Sie sind auf Seite 1von 170

Our updated Terms of Use will become effective on May 25, 2012. Find out more.

Category:Thermodynamic cycles
From Wikipedia, the free encyclopedia

Jump to: navigation, search The main article for this category is Thermodynamic cycle. Wikimedia Commons has media related to: Thermodynamics cycles

Subcategories
This category has only the following subcategory.

Hot air engines (1 C, 9 P)

Pages in category "Thermodynamic cycles"


The following 40 pages are in this category, out of 40 total. This list may not reflect recent changes (learn more).

Thermodynamic cycle

H cont.

Homogeneous charge compression ignition Humphrey cycle

Rankine cycle Regenerative cooling

Absorption refrigerator Atkinson cycle

S I

Integrated gasification combined cycle

Barton evaporation engine K Bordwell thermodynamic cycle Brayton cycle

Kalina cycle Kleemenko cycle

Scuderi cycle Sea surface temperature Siemens cycle Solarhydrogen energy cycle Staged combustion cycle (rocket) Stirling cycle Stoddard engine Flash-gas

L
Carnot cycle Combined cycle

Subcooling

Lenoir cycle

M D

Vapor-compression refrigeration

Diesel cycle

Ericsson cycle

Magnetic refrigeration Mercury vapour turbine Miller cycle Mixed/dual cycle

O
Gas-generator cycle (rocket)

Organic Rankine cycle Otto cycle

P
HampsonLinde cycle Heat pump and refrigeration cycle High-efficiency hybrid cycle

Pseudo Stirling cycle Pulse tube refrigerator

Retrieved from "http://en.wikipedia.org/w/index.php?title=Category:Thermodynamic_cycles&oldid=485 126556" Categories: Thermodynamic processes Engine technology


Personal tools

Log in / create account

Namespaces

Category Talk

Variants

Views Actions Search


Search Special:Search

Read Edit View history

Navigation

Main page Contents Featured content Current events Random article Donate to Wikipedia

Interaction Toolbox

Help About Wikipedia Community portal Recent changes Contact Wikipedia

What links here Related changes Upload file Special pages Permanent link

Print/export

Create a book Download as PDF Printable version

Languages

Bosanski

Catal Dansk Deutsch Espaol Esperanto Franais Galego Bahasa Indonesia Italiano Nederlands Polski Slovenina / Srpski Svenska Trke Ting Vit This page was last modified on 2 April 2012 at 08:28. Text is available under the Creative Commons Attribution-ShareAlike License; additional terms may apply. See Terms of use for details. Wikipedia is a registered trademark of the Wikimedia Foundation, Inc., a nonprofit organization. Contact us Privacy policy About Wikipedia Disclaimers Mobile view

Our updated Terms of Use will become effective on May 25, 2012. Find out more.

Atkinson cycle
From Wikipedia, the free encyclopedia

Jump to: navigation, search

Thermodynamics

Branches[show]

Laws[show]

Systems[show]

System properties[show]

Material properties[show]

Specific heat capacity Compressibility Thermal expansion

Equations[show]

Potentials[show]

Internal energy Enthalpy Helmholtz free energy Gibbs free energy


History and culture[show]

Scientists[show]

v t e

The Atkinson cycle engine is a type of internal combustion engine invented by James Atkinson in 1882. The Atkinson cycle is designed to provide efficiency at the expense of power density, and is used in some modern hybrid electric applications.

Contents
[hide]

1 Design 2 Ideal thermodynamic cycle 3 Modern Atkinson cycle engines 4 Rotary Atkinson cycle engine 5 Vehicles using Atkinson cycle engines 6 See also 7 References 8 External links

[edit] Design
The original Atkinson cycle piston engine allowed the intake, compression, power, and exhaust strokes of the four-stroke cycle to occur in a single turn of the crankshaft and was designed to avoid infringing certain patents covering Otto cycle engines.[1] Due to the unique crankshaft design of the Atkinson, its expansion ratio can differ from its compression ratio and, with a power stroke longer than its compression stroke, the engine can achieve greater thermal efficiency than a traditional piston engine. While Atkinson's original design is no more than a historical curiosity, many modern engines use unconventional valve timing to produce the effect of a shorter compression stroke/longer power stroke, thus realizing the fuel economy improvements the Atkinson cycle can provide.[2]

[edit] Ideal thermodynamic cycle

Figure 1: Atkinson Gas Cycle The ideal Atkinson cycle consists of following operations:

1-2 Isentropic or reversible adiabatic compression 2-3 Isochoric heating (Qp) 3-4 Isobaric heating (Qp') 4-5 Isentropic expansion 5-6 Isochoric cooling (Qo) 6-1 Isobaric cooling (Qo')

[edit] Modern Atkinson cycle engines

A small engine with Atkinson-style linkages between the piston and flywheel. Modern Atkinson cycle engines do away with this complex energy path. Recently Atkinson cycle has been used to describe a modified Otto cycle engine in which the intake valve is held open longer than normal to allow a reverse flow of intake air into the intake manifold. The effective compression ratio is reduced (for a time the air is escaping the cylinder freely rather than being compressed) but the expansion ratio is unchanged. This means the compression ratio is smaller than the expansion ratio. Heat gained from burning fuel increases the pressure, thereby forcing the piston to move, expanding the air volume beyond the volume when compression began. The goal of the modern Atkinson cycle is to allow the pressure in the combustion chamber at the end of the power stroke to be equal to atmospheric pressure; when this occurs, all the available energy has been obtained from the combustion process. For any given portion of air, the greater expansion ratio allows more energy to be converted from heat to useful mechanical energy meaning the engine is more efficient. The disadvantage of the four-stroke Atkinson cycle engine versus the more common Otto cycle engine is reduced power density. Due to a smaller portion of the compression stroke being devoted to compressing the intake air, an Atkinson cycle engine does not take in as much air as would a similarly designed and sized Otto cycle engine. Four-stroke engines of this type with this same type of intake valve motion but with a supercharger to make up for the loss of power density are known as Miller cycle engines.

[edit] Rotary Atkinson cycle engine

Rotary Atkinson cycle engine The Atkinson cycle can be used in a rotary engine. In this configuration an increase in both power and efficiency can be achieved when compared to the Otto cycle. This type of engine retains the one power phase per revolution, together with the different compression and expansion volumes of the original Atkinson cycle. Exhaust gases are expelled from the engine by compressed-air scavenging. This modification of the Atkinson cycle allows the use of alternative fuels like diesel and hydrogen. Disadvantages of this design include the requirement that rotor tips seal very tightly on the outer housing wall and the mechanical losses suffered through friction between rapidly oscillating parts of irregular shape. See External Links for more information.

[edit] Vehicles using Atkinson cycle engines

2004 Toyota Prius hybrid

2010 Ford Fusion Hybrid (North America) While a modified Otto cycle engine using the Atkinson cycle provides good fuel economy, it is at the expense of a lower power-per-displacement as compared to a traditional four-stroke engine.[3] If demand for more power is intermittent, the power of the engine can be supplemented by an electric motor during times when more power is needed. This forms the basis of an Atkinson cycle-based hybrid electric drivetrain. These electric motors can be used independently of, or in combination with, the Atkinson cycle engine, to provide the most efficient means of producing the desired power. This drive train first entered production in late 1997 in the Japanese-market Toyota Prius. At this writing, most production full hybrid-electric vehicles use Atkinson cycle engines:

Chevrolet Tahoe Hybrid electric (four-wheel drive) with a compression ratio of 10.8:1 Ford Escape/Mercury Mariner/Mazda Tribute electric (front- and four-wheel drive) with a compression ratio of 12.4:1 Ford Fusion Hybrid/Mercury Milan Hybrid/Lincoln MKZ Hybrid electric (frontwheel drive) with a compression ratio of 12.3:1 Hyundai Sonata Hybrid (front-wheel drive) Infiniti M35h Hybrid (rear-wheel drive) Kia Optima Hybrid (front-wheel drive) Lexus CT200h (front-wheel drive) Lexus HS250h (front-wheel drive) Lexus RX 450h hybrid electric (rear-wheel drive) Mercedes ML450 Hybrid (four-wheel drive) electric Mercedes S400 Blue Hybrid (rear-wheel drive) electric Toyota Highlander Hybrid (2011 and newer)[4] Toyota Prius hybrid electric (front-wheel drive) with a (purely geometric) compression ratio of 13.0:1 Toyota Camry Hybrid electric (front-wheel drive) with a compression ratio of 12.5:1 Lexus GS450h hybrid electric (Rear-Wheel drive) with a compression ratio of 13.0:1

[edit] See also

History of the internal combustion engine

[edit] References
1. ^ U.S. Patent 367,496 2. ^ http://www.canadiandriver.com/2010/07/14/auto-tech-atkinson-cycle-engines-andhybrids.htm 3. ^ Heywood, John B. Internal Combustion Engine Fundamentals, pp. 184-186.

4. ^ http://www.edmunds.com/toyota/highlander-hybrid/2011/road-test.html

[edit] External links

Animation of Atkinson Cycle Engine Note that this animation shows the true Atkinson engine, which uses a complex linkage that allows different stroke lengths for intake/compression and power/exhaust. However, the illustration shows the engine with the linkage laid out to generate 4 equal strokes. To alter the ratio of the strokes, the rightmost pivot point (the one that is attaching the horizontal green link to the frame) should be moved downwards along the frame. This will allow more angular movement as the link rotates up, giving a longer piston stroke for power and exhaust, and less angular movement as the link rotates down, giving a shorter piston stroke for intake and compression. In fact, a sliding pivot point at that location would allow the engine to dynamically change the stroke ratios. Modified Atkinson Cycle Engine: Alternative variable valve timing strategy increases low speed torque obtainable from Atkinson Cycle Engine. Comparison of Prime Movers Suitable for USMC Expeditionary Power Sources, Oak Ridge National Laboratory Libralato Engines - developing a rotary Atkinson cycle engine Rotary Atkinson cycle engine - gives details of this engine as well as comparisons with conventional and Wankel engines The Prius's Not So Secret Gas-Mileage Secrets - how the Prius uses the Atkinson cycle to get better results than an Otto cycle engine [hide]

v t e

Thermodynamic cycles
Without phase change (hot air engines)

External combustion cycles

Bell Coleman Brayton/Joule Carnot Ericsson Stirling Stirling (Pseudo / Adiabatic) Stoddard

With phase change

Kalina

Rankine (Organic Rankine) Regenerative Atkinson Brayton/Joule Diesel Expander Gas-generator Homogeneous Charge Compression Ignition Lenoir Miller Otto Pressure-fed Staged combustion Combined HEHC Mixed/Dual Hampson-Linde Kleemenko Pulse tube Regenerative cooling Transcritical Vapor absorption Vapor-compression Siemens Vuilleumier Barton Humphrey Scuderi

Internal combustion cycles

Mixed cycles

Refrigeration cycles

Uncategorized

Retrieved from "http://en.wikipedia.org/w/index.php?title=Atkinson_cycle&oldid=485958953" View page ratings

Rate this page


What's this? Trustworthy Objective Complete

Well-written I am highly knowledgeable about this topic (optional) Submit ratings Saved successfully Your ratings have not been submitted yet Categories: Thermodynamic cycles Hybrid vehicles
Personal tools

Log in / create account

Namespaces

Article Talk

Variants Views Actions Search


Search Special:Search

Read Edit View history

Navigation

Main page Contents Featured content Current events Random article Donate to Wikipedia

Interaction

Help

Toolbox

About Wikipedia Community portal Recent changes Contact Wikipedia

What links here Related changes Upload file Special pages Permanent link Cite this page Rate this page

Print/export

Create a book Download as PDF Printable version

Languages

Deutsch Espaol Franais Galego Italiano Magyar Polski Slovenina This page was last modified on 6 April 2012 at 19:43. Text is available under the Creative Commons Attribution-ShareAlike License; additional terms may apply. See Terms of use for details. Wikipedia is a registered trademark of the Wikimedia Foundation, Inc., a nonprofit organization. Contact us Privacy policy About Wikipedia Disclaimers Mobile view

Our updated Terms of Use will become effective on May 25, 2012. Find out more.

Barton evaporation engine


From Wikipedia, the free encyclopedia

Jump to: navigation, search The topic of this article may not meet Wikipedia's notability guidelines for products and services. Please help to establish notability by adding reliable, secondary sources about the topic. If notability cannot be established, the article is likely to be merged, redirected, or deleted. (June 2011) This article needs additional citations for verification. Please help improve this article by adding citations to reliable sources. Unsourced material may be challenged and removed. (June 2011)

Thermodynamics

Branches[show]

Laws[show]

Systems[show]

System properties[show]

Material properties[show]

Specific heat capacity

Compressibility Thermal expansion

Equations[show]

Potentials[show]

Internal energy Enthalpy Helmholtz free energy Gibbs free energy


History and culture[show]

Scientists[show]

v t e

The Barton evaporation engine (BEE) is a heat engine developed by Sunoba Pty Ltd.

Contents
[hide]

1 Principle 2 Performance

3 See also 4 References 5 External links

[edit] Principle
The principle of the BEE is based on: (1) adiabatic expansion of unsaturated air; (2) evaporative cooling at reduced pressure; and (3) re-compression back to atmospheric pressure with further evaporation. By this means, the engine produces power and cooled moist air from water and hot dry air: hot dry air + water power + cooled moist air With a modest amount of passive solar pre-heating, the engine theoretically is able to produce power in hot arid climates. As well as being a heat engine, the BEE can also be used as an evaporative cooler. Moreover, the BEE has broadly comparable theoretical efficiency to simple Rankine steam turbines, without need for high-pressure boiler or condenser. The BEE can function well on industrial waste heat, particularly the exhaust gas of open cycle gas turbines.

[edit] Performance
It is claimed [1] that under suitable weather conditions, and assuming evaporation to saturation at constant volume in the low-pressure section and no further evaporation, a BEE engine would convert approximately 4-5% of energy collected at 30-40C above ambient into electrical or mechanical power (further evaporation during the recompression step (3) would more than double this estimate). Whilst this efficiency is not high, the inputs will be inexpensive since the requisite pre-heating can be accomplished by passive solar methods, and hot air is both the heat transfer medium and the working gas. The engine does not require heat exchangers or condensers as with Rankine cycle engines. In general, the efficiency of the BEE increases with the inlet temperature and the expansion ratio. A full thermodynamic analysis is available [2]. If the inlet air is sourced from an Open-Cycle Gas Turbine exhaust at 500 degrees Celsius, the theoretical work output is approximately 150 kJ per kg of dry air throughput. The thermodynamic cycles for the BEE and the Brayton gas engine can be combined in an engine consisting of a single device which would have remarkably high efficiency.

[edit] See also

Drinking bird - a toy that works on similar principles.

[edit] References
1. ^ N.G. Barton, Thermodynamic Model for a Two-Stroke Evaporation Engine, in Proc Australian and New Zealand Solar Energy Society Conference, Canberra, 1315 September 2006. 2. ^ N.G. Barton, "An Evaporation Heat Engine and Condensation Heat Pump", The ANZIAM Journal, 49 (2008), to appear.

[edit] External links

About Sunoba Pty Ltd [hide]


v t e

Thermodynamic cycles
Without phase change (hot air engines)

External combustion cycles

Bell Coleman Brayton/Joule Carnot Ericsson Stirling Stirling (Pseudo / Adiabatic) Stoddard

With phase change


Kalina Rankine (Organic Rankine) Regenerative Atkinson Brayton/Joule Diesel Expander Gas-generator

Internal combustion cycles

Homogeneous Charge Compression Ignition Lenoir Miller Otto Pressure-fed Staged combustion Combined HEHC Mixed/Dual Hampson-Linde Kleemenko Pulse tube Regenerative cooling Transcritical Vapor absorption Vapor-compression Siemens Vuilleumier Barton Humphrey Scuderi

Mixed cycles

Refrigeration cycles

Uncategorized

Retrieved from "http://en.wikipedia.org/w/index.php?title=Barton_evaporation_engine&oldid=43552666 4" View page ratings

Rate this page


What's this? Trustworthy Objective Complete Well-written I am highly knowledgeable about this topic (optional) Submit ratings Saved successfully Your ratings have not been submitted yet

Categories: Thermodynamic cycles Engines Hidden categories: Articles with topics of unclear notability from June 2011 All articles with topics of unclear notability Articles needing additional references from June 2011 All articles needing additional references
Personal tools

Log in / create account

Namespaces

Article Talk

Variants Views Actions Search


Search Special:Search

Read Edit View history

Navigation

Main page Contents Featured content Current events Random article Donate to Wikipedia

Interaction

Help About Wikipedia Community portal

Toolbox

Recent changes Contact Wikipedia

What links here Related changes Upload file Special pages Permanent link Cite this page Rate this page

Print/export

Create a book Download as PDF Printable version This page was last modified on 21 June 2011 at 20:53. Text is available under the Creative Commons Attribution-ShareAlike License; additional terms may apply. See Terms of use for details. Wikipedia is a registered trademark of the Wikimedia Foundation, Inc., a nonprofit organization. Contact us Privacy policy About Wikipedia Disclaimers Mobile view

Our updated Terms of Use will become effective on May 25, 2012. Find out more.

Bordwell thermodynamic cycle


From Wikipedia, the free encyclopedia

Jump to: navigation, search A Bordwell thermodynamic cycle use experimentally determined and reasonable estimates of Gibbs free energy (G) values to determine unknown and experimentally inaccessible values.[1][2]

[edit] Overview
Analogous to Hess's Law which deal with the summation of enthalpy (H) values, Bordwell thermodynamic cycles deal with the summation of Gibbs free energy (G) values. Free energies used in these systems are most often determined from equilibriums and redox potentials, both of which correlate with free energy. This is with the caveat that redox scales are not absolute and thus it is important that all electrons are evaluated in redox pairs. This removes the offset of a given reference potential, otherwise the values are reported as potentials (V) against that reference. Its also worth recognizing that the values of the pKa system are just moderately transformed Keq values. When working with equilibrium energy values such as G and E1/2 values it common to employ a naught () symbol. The naught has a two component definition. The first more common component is that it refers to the physical conditions being at standard state. The second more significant component is that energy refers to an equilibrium energy even if there is a conditionally defined standard state. Just as activation energy with the double dagger G refers the energy difference between reactants and the transition state, G refers to the energy difference between reactants and products. The nought is assumed when working with equilibrium values such as Keq and pKa. The example below contains four reactions that can be related through their associated free energies. Given any three values and the fourth can be calculated. Its important to note that the fourth reaction in the series is an inverted homolytic bond cleavage stated in terms of free energy. The chemical transformation for the associated -G is the same it would be for a bond dissociation energy (BDE). However, the -G is not a BDE, since BDE are by definition stated in terms of enthalpy (H). The two values are of course related by G = H - TS and as a result educated comparisons can be made between G and H. R- e- + R. (Reaction 1) Go rxn 1 = -nFE1/2 H+ + e- H. (Reaction 2)

Go = -nFE1/2 RH H+ + R- (Reaction 3) Go rxn 3 = RT(2.303)pKa R. + H. RH (Reaction 4) Go rxn 4 = -RTln(Keq)


rxn 2

[edit] Conversions
Relationships between Keq, pKeq, E1/2, and G.[3] G = -RTln(Keq) G = (2.303)RT(pKeq) G = -nFE1/2 Useful conversion factors: -23.06 (kcal/mol)(e-)-1(V)-1 1.37(pKeq) kcal/mol 1.37[-log(Keq)] kcal/mol

[edit] References
1. ^ Bordwell, Frederick G. (1988-12-01). "Equilibrium acidities in dimethyl sulfoxide solution". Accounts of Chemical Research 21 (12): 456463. doi:10.1021/ar00156a004. 2. ^ Gardner, Kimberly A.; Linda L. Kuehnert, James M. Mayer (1997-05-01). "Hydrogen Atom Abstraction by Permanganate: Oxidations of Arylalkanes in Organic Solvents". Inorganic Chemistry 36 (10): 20692078. doi:10.1021/ic961297y. PMID 11669825. 3. ^ Silberberg, Martin (2004). Chemistry: The Molecular Nature of Matter and Change (4 ed.). McGraw-Hill Science/Engineering/Math. ISBN 0073101699.

Retrieved from "http://en.wikipedia.org/w/index.php?title=Bordwell_thermodynamic_cycle&oldid=4566 44122" View page ratings

Rate this page


What's this? Trustworthy Objective Complete

Well-written I am highly knowledgeable about this topic (optional) Submit ratings Saved successfully Your ratings have not been submitted yet Categories: Thermodynamic cycles
Personal tools

Log in / create account

Namespaces

Article Talk

Variants Views Actions Search


Search Special:Search

Read Edit View history

Navigation

Main page Contents Featured content Current events Random article Donate to Wikipedia

Interaction

Help About Wikipedia

Toolbox

Community portal Recent changes Contact Wikipedia

What links here Related changes Upload file Special pages Permanent link Cite this page Rate this page

Print/export

Create a book Download as PDF Printable version This page was last modified on 21 October 2011 at 07:31. Text is available under the Creative Commons Attribution-ShareAlike License; additional terms may apply. See Terms of use for details. Wikipedia is a registered trademark of the Wikimedia Foundation, Inc., a nonprofit organization. Contact us Privacy policy About Wikipedia Disclaimers Mobile view

Our updated Terms of Use will become effective on May 25, 2012. Find out more.

Brayton cycle
From Wikipedia, the free encyclopedia

Jump to: navigation, search

Thermodynamics

Branches[show]

Laws[show]

Systems[show]

System properties[show]

Material properties[show]

Specific heat capacity Compressibility Thermal expansion

Equations[show]

Potentials[show]

Internal energy Enthalpy Helmholtz free energy Gibbs free energy


History and culture[show]

Scientists[show]

v t e

The Brayton cycle is a thermodynamic cycle that describes the workings of the gas turbine engine, basis of the airbreathing jet engine and others. It is named after George Brayton (18301892), the American engineer who developed it, although it was originally proposed and patented by Englishman John Barber in 1791.[1] It is also sometimes known as the Joule cycle. The Ericsson cycle is similar but uses external heat and incorporates the use of a regenerator.

Contents
[hide]

1 History 2 Model o 2.1 Solar Brayton cycle 3 Methods to increase power 4 Methods to improve efficiency 5 Reverse Brayton cycle

6 See also 7 References 8 External links

[edit] History
This unreferenced section requires citations to ensure verifiability. In 1872, George Brayton applied for a patent for his Ready Motor. The engine used a separate piston compressor and expander. The compressed air was heated by internal fire as it entered the expander cylinder. Brayton produced and sold "Ready Motors" to perform a variety of tasks like water pumping, mill operation, even marine propulsion. Critics of the day claimed the engines ran smoothly and had an efficiency of about 17%. Today the term Brayton cycle is generally associated with the gas turbine, even though Brayton only built piston engines. The Brayton cycle is a cycle which can be used in both internal combustion engines (such as jet engines) and for external combustion engines. Although the Brayton cycle is usually run as an open system (and indeed must be run as such if internal combustion is used), it is conventionally assumed for the purposes of thermodynamic analysis that the exhaust gases are reused in the intake, enabling analysis as a closed system. Another interesting piece of Brayton cycle history was its use in the Selden patent. In the early days of the automobile, a creative attorney "Selden" claimed to have a patent for the internal combustion powered version. The patent drawings showed the use of Brayton cycle engine. Instead of paying royalties, Henry Ford fought the Selden patent. Ford argued his cars used the four-stroke Otto cycle and not the Brayton engine shown used in the Selden auto. Ford won the appeal of the original case.

[edit] Model
A Brayton-type engine consists of three components:

A gas compressor A mixing chamber An expander

In the original 19th-century Brayton engine, ambient air is drawn into a piston compressor, where it is compressed; ideally an isentropic process. The compressed air then runs through a mixing chamber where fuel is added, an isobaric process. The heated (by compression), pressurized air and fuel mixture is then ignited in an expansion cylinder and energy is released, causing the heated air and combustion products to expand through a piston/cylinder; another ideally isentropic process. Some of the work extracted by the piston/cylinder is used to drive the compressor through a crankshaft arrangement. The term Brayton cycle has more recently been given to the gas turbine engine. This also has three components:

A gas compressor A burner (or combustion chamber) An expansion turbine

Ideal Brayton cycle:


isentropic process - Ambient air is drawn into the compressor, where it is pressurized. isobaric process - The compressed air then runs through a combustion chamber, where fuel is burned, heating that aira constant-pressure process, since the chamber is open to flow in and out. isentropic process - The heated, pressurized air then gives up its energy, expanding through a turbine (or series of turbines). Some of the work extracted by the turbine is used to drive the compressor. isobaric process - Heat rejection (in the atmosphere).

Actual Brayton cycle:


adiabatic process - Compression. isobaric process - Heat addition. adiabatic process - Expansion. isobaric process - Heat rejection.

Idealized Brayton cycle Since neither the compression nor the expansion can be truly isentropic, losses through the compressor and the expander represent sources of inescapable working inefficiencies. In general, increasing the compression ratio is the most direct way to increase the overall power output of a Brayton system.[2]

The efficiency of the ideal Brayton cycle is , where is the heat capacity ratio.[3] Figure 1 indicates how the cycle efficiency changes with an increase in pressure ratio. Figure 2 indicates how the specific power output changes with an increase in the gas turbine inlet temperature for two different pressure ratio values.

Figure 1: Brayton cycle efficiency

Figure 2: Brayton cycle specific power output

[edit] Solar Brayton cycle


In 2002 a hybrid open solar Brayton cycle was operated for the first time consistently and effectively with relevant papers published, in the frame of the EU SOLGATE program.[4] The air was heated from 570 K to over 1000 K into the combustor chamber. Further hybridization was achieved during the EU Solhyco project running a hybridized Brayton cycle with solar energy and Biodiesel only.[5]

[edit] Methods to increase power


The power output of a Brayton engine can be improved in the following manners:

Reheat, wherein the working fluidin most cases airexpands through a series of turbines, then is passed through a second combustion chamber before expanding to ambient pressure through a final set of turbines. This has the advantage of increasing the power output possible for a given compression ratio without exceeding any metallurgical constraints (typically about 1000 C). The use of an afterburner for jet aircraft engines can also be referred to as "reheat"; it is a different process in that the reheated air is expanded through a thrust nozzle rather than a turbine. The metallurgical constraints are somewhat alleviated, enabling much higher reheat temperatures (about 2000 C). Reheat is most often used to improve the specific power (per throughput of air), and is usually associated with a drop in efficiency, this effect is especially pronounced in afterburners due to the extreme amounts of extra fuel used. Overspray, wherein, after a first compressor stage, water is injected into the compressor, thus increasing the mass-flow inside the compressor, increasing the turbine output power significantly and reducing compressor outlet temperatures.[6] In a second compressor stage the water is completely converted to a gas form, offering some intercooling via its latent heat of vaporization.

[edit] Methods to improve efficiency


The efficiency of a Brayton engine can be improved in the following manners:

Increasing pressure ratio - As Figure 1 above shows, increasing the pressure ratio increases the efficiency of the Brayton cycle. This is analogous to the increase of efficiency seen in the Otto cycle when the compression ratio is increased. However, there are practical limits when it comes to increasing the pressure ratio. First of all, increasing the pressure ratio increases the compressor discharge temperature. This can cause the temperature of the gasses leaving the combustor to exceed the metallurgical limits of the turbine. Also, the diameter of the compressor blades becomes progressively smaller in higher pressure stages of the compressor. Because the gap between the blades and the engine casing increases in size as a percentage of the compressor blade height as the blades get smaller in diameter, a greater percentage of the compressed air can leak back past the blades in higher pressure stages. This causes a drop in compressor efficiency, and is most likely to occur in smaller gas turbines (since blades are inherently smaller to begin with). Finally, as can be seen in Figure 1, the efficiency levels off as pressure ratio increases. Hence, there is little to gain by increasing the pressure ratio further if it is already at a high level.

Regeneration, wherein the still-warm post-turbine fluid is passed through a heat exchanger to preheat the fluid just entering the combustion chamber. This directly offsets fuel consumption for the same operating conditions, improving efficiency; it also results in less power lost as waste heat. However, at higher pressure ratios, the compressor discharge temperature can exceed the exhaust temperature. Under these conditions, regeneration would be counterproductive. Therefore, regeneration is only an option when the pressure ratio is sufficiently low that the exhaust temperature is higher than the compressor discharge temperature.

This feature is only available if the exhaust heat is not used otherwise, as in cogeneration or combined cycle applications.

A Brayton engine also forms half of the combined cycle system, which combines with a Rankine engine to further increase overall efficiency. However, although this increases overall efficiency, it does not actually increase the efficiency of the Brayton cycle itself. Cogeneration systems make use of the waste heat from Brayton engines, typically for hot water production or space heating.

[edit] Reverse Brayton cycle


A Brayton cycle that is driven in reverse, via net work input, and when air is the working fluid, is the air refrigeration cycle or Bell Coleman cycle. Its purpose is to move heat, rather than produce work. This air cooling technique is used widely in jet aircraft.

[edit] See also


Wikimedia Commons has media related to: Brayton cycle

Gas turbine Jet engine Heat engines Thermodynamics Power HVAC Engineering Gerotor Closed-cycle gas turbine

[edit] References
1. ^ according to Gas Turbine History

2. ^ Lester C. Lichty, Combustion Engine Processes, 1967, McGraw-Hill, Inc., Lib.of Congress 67-10876 3. ^ http://web.mit.edu/16.unified/www/SPRING/propulsion/notes/node27.html Ideal cycle equations, MIT lecture notes 4. ^ Research 5. ^ Solhyco.com Retrieved 2012-01-09 6. ^ http://www.max-boost.co.uk/max-boost/resources/docs/SwirlFlash_WI.pdf

[edit] External links


Today in Science article on Brayton Engine http://scitation.aip.org/getabs/servlet/GetabsServlet?prog=normal&id=JSEEDO00 0126000003000872000001&idtype=cvips&gifs=yes http://elib.dlr.de/46328/ Test and evaluation of a solar powered gas turbine system [hide]

v t e

Thermodynamic cycles
Without phase change (hot air engines)

External combustion cycles

Bell Coleman Brayton/Joule Carnot Ericsson Stirling Stirling (Pseudo / Adiabatic) Stoddard

With phase change


Kalina Rankine (Organic Rankine) Regenerative Atkinson Brayton/Joule Diesel Expander Gas-generator Homogeneous Charge Compression Ignition

Internal combustion cycles

Lenoir Miller Otto Pressure-fed Staged combustion Combined HEHC Mixed/Dual Hampson-Linde Kleemenko Pulse tube Regenerative cooling Transcritical Vapor absorption Vapor-compression Siemens Vuilleumier Barton Humphrey Scuderi

Mixed cycles

Refrigeration cycles

Uncategorized

Retrieved from "http://en.wikipedia.org/w/index.php?title=Brayton_cycle&oldid=488663651" View page ratings

Rate this page


What's this? Trustworthy Objective Complete Well-written I am highly knowledgeable about this topic (optional) Submit ratings Saved successfully Your ratings have not been submitted yet Categories:

Thermodynamic cycles Hidden categories: Articles needing additional references from December 2011 All articles needing additional references
Personal tools

Log in / create account

Namespaces

Article Talk

Variants Views Actions Search


Search Special:Search

Read Edit View history

Navigation

Main page Contents Featured content Current events Random article Donate to Wikipedia

Interaction Toolbox

Help About Wikipedia Community portal Recent changes Contact Wikipedia

What links here Related changes Upload file Special pages Permanent link Cite this page Rate this page

Print/export

Create a book Download as PDF Printable version

Languages

Catal Deutsch Espaol Franais Galego Italiano Magyar Polski Portugus Slovenina / Srpski Svenska Trke Ting Vit This page was last modified on 22 April 2012 at 15:42. Text is available under the Creative Commons Attribution-ShareAlike License; additional terms may apply. See Terms of use for details. Wikipedia is a registered trademark of the Wikimedia Foundation, Inc., a nonprofit organization. Contact us Privacy policy About Wikipedia

Disclaimers Mobile view

Our updated Terms of Use will become effective on May 25, 2012. Find out more.

Carnot cycle
From Wikipedia, the free encyclopedia

Jump to: navigation, search

Thermodynamics

Branches[show]

Laws[show]

Systems[show]

System properties[show]

Material properties[show]

Specific heat capacity Compressibility Thermal expansion

Equations[show]

Potentials[show]

Internal energy Enthalpy Helmholtz free energy Gibbs free energy


History and culture[show]

Scientists[show]

v t e

The Carnot cycle is a theoretical thermodynamic cycle proposed by Nicolas Lonard Sadi Carnot in 1824 and expanded by Benoit Paul mile Clapeyron in the 1830s and 40s. It can be shown that it is the most efficient cycle for converting a given amount of thermal energy into work, or conversely, creating a temperature difference (e.g. refrigeration) by doing a given amount of work. Every thermodynamic system exists in a particular thermodynamic state. When a system is taken through a series of different states and finally returned to its initial state, a thermodynamic cycle is said to have occurred. In the process of going through this cycle, the system may perform work on its surroundings, thereby acting as a heat engine. A system undergoing a Carnot cycle is called a Carnot heat engine, although such a 'perfect' engine is only a theoretical limit and cannot be built in practice. The Carnot cycle when acting as a heat engine consists of the following steps: 1. Reversible isothermal expansion of the gas at the "hot" temperature, TH (isothermal heat addition or absorption). During this step (A to B on Figure 1, 1 to 2 in Figure 2) the expanding gas makes the piston work on the surroundings.

The gas expansion is propelled by absorption of quantity Q1 of heat from the high temperature reservoir. 2. Isentropic (reversible adiabatic) expansion of the gas (isentropic work output). For this step (B to C on Figure 1, 2 to 3 in Figure 2) the piston and cylinder are assumed to be thermally insulated, thus they neither gain nor lose heat. The gas continues to expand, working on the surroundings. The gas expansion causes it to cool to the "cold" temperature, TC. 3. Reversible isothermal compression of the gas at the "cold" temperature, TC. (isothermal heat rejection) (C to D on Figure 1, 3 to 4 on Figure 2) Now the surroundings do work on the gas, causing quantity Q2 of heat to flow out of the gas to the low temperature reservoir. 4. Isentropic compression of the gas (isentropic work input). (D to A on Figure 1, 4 to 1 in Figure 2) Once again the piston and cylinder are assumed to be thermally insulated. During this step, the surroundings do work on the gas, compressing it and causing the temperature to rise to TH. At this point the gas is in the same state as at the start of step 1.

Figure 1: A Carnot cycle acting as a heat engine, illustrated on a temperature-entropy diagram. The cycle takes place between a Figure 2: A Carnot cycle acting as a heat hot reservoir at temperature TH and a cold engine, illustrated on a pressure-volume reservoir at temperature TC. The vertical diagram to illustrate the work done. axis is temperature, the horizontal axis is entropy.

Contents
[hide]

1 Properties and significance o 1.1 The temperature-entropy diagram o 1.2 The Carnot cycle o 1.3 Carnot's theorem o 1.4 Efficiency of real heat engines 2 See also 3 References 4 External links

[edit] Properties and significance


[edit] The temperature-entropy diagram

A generalized thermodynamic cycle taking place between a hot reservoir at temperature TH and a cold reservoir at temperature TC. By the second law of thermodynamics, the cycle cannot extend outside the temperature band from TC to TH. The area in red QC is the amount of energy exchanged between the system and the cold reservoir. The area in white W is the amount of work energy exchanged by the system with its surroundings. The amount of heat exchanged with the hot reservoir is the sum of the two. If the system is behaving as an engine, the process moves clockwise around the loop, and moves counter-clockwise if it is behaving as a refrigerator. The efficiency of the cycle is the ratio of the white area (work) divided by the sum of the white and red areas (heat absorbed from the hot reservoir). The behaviour of a Carnot engine or refrigerator is best understood by using a temperature-entropy (TS) diagram, in which the thermodynamic state is specified by a point on a graph with entropy (S) as the horizontal axis and temperature (T) as the vertical axis. For a simple system with a fixed number of particles, any point on the graph will represent a particular state of the system. A thermodynamic process will consist of a

curve connecting an initial state (A) and a final state (B). The area under the curve will be:

which is the amount of thermal energy transferred in the process. If the process moves to greater entropy, the area under the curve will be the amount of heat absorbed by the system in that process. If the process moves towards lesser entropy, it will be the amount of heat removed. For any cyclic process, there will be an upper portion of the cycle and a lower portion. For a clockwise cycle, the area under the upper portion will be the thermal energy absorbed during the cycle, while the area under the lower portion will be the thermal energy removed during the cycle. The area inside the cycle will then be the difference between the two, but since the internal energy of the system must have returned to its initial value, this difference must be the amount of work done by the system over the cycle. Referring to figure 2, mathematically, for a reversible process we may write the amount of work done over a cyclic process as:

Since dU is an exact differential, its integral over any closed loop is zero and it follows that the area inside the loop on a T-S diagram is equal to the total work performed if the loop is traversed in a clockwise direction, and is equal to the total work done on the system as the loop is traversed in a counterclockwise direction.

A Carnot cycle taking place between a hot reservoir at temperature TH and a cold reservoir at temperature TC.

[edit] The Carnot cycle


Evaluation of the above integral is particularly simple for the Carnot cycle. The amount of energy transferred as work is

The total amount of thermal energy transferred between the hot reservoir and the system will be

and the total amount of thermal energy transferred between the system and the cold reservoir will be

The efficiency is defined to be:

where is the work done by the system (energy exiting the system as work), is the heat put into the system (heat energy entering the system), is the absolute temperature of the cold reservoir, and is the absolute temperature of the hot reservoir. is the maximum system entropy is the minimum system entropy This efficiency makes sense for a heat engine, since it is the fraction of the heat energy extracted from the hot reservoir and converted to mechanical work. A Rankine cycle is usually the practical approximation.

[edit] Carnot's theorem


Main article: Carnot's theorem (thermodynamics) It can be seen from the above diagram, that for any cycle operating between temperatures T_H and T_C, none can exceed the efficiency of a Carnot cycle.

A real engine (left) compared to the Carnot cycle (right). The entropy of a real material changes with temperature. This change is indicated by the curve on a T-S diagram. For this figure, the curve indicates a vapor-liquid equilibrium (See Rankine cycle). Irreversible systems and losses of heat (for example, due to friction) prevent the ideal from taking place at every step. Carnot's theorem is a formal statement of this fact: No engine operating between two heat reservoirs can be more efficient than a Carnot engine operating between those same reservoirs. Thus, Equation 3 gives the maximum efficiency possible for any engine using the corresponding temperatures. A corollary to Carnot's theorem states that: All reversible engines operating between the same heat reservoirs are equally efficient. Rearranging the right side of the equation gives what may be a more easily understood form of the equation. Namely that the theoretical maximum efficiency of a heat engine equals the difference in temperature between the hot and cold reservoir divided by the absolute temperature of the hot reservoir. To find the absolute temperature in kelvin, add 273.15 degrees to the Celsius temperature. Looking at this formula an interesting fact becomes apparent. Lowering the temperature of the cold reservoir will have more effect on the ceiling efficiency of a heat engine than raising the temperature of the hot reservoir by the same amount. In the real world, this may be difficult to achieve since the cold reservoir is often an existing ambient temperature. In other words, maximum efficiency is achieved if and only if no new entropy is created in the cycle. Otherwise, since entropy is a state function, the required dumping of heat into the environment to dispose of excess entropy leads to a reduction in efficiency. So Equation 3 gives the efficiency of any reversible heat engine. In mesoscopic heat engines, work per cycle of operation fluctuates due to thermal noise. For the case when work and heat fluctuations are counted, there is exact equality that relates average of exponents of work performed by any heat engine and the heat transfer from the hotter heat bath N. A. Sinitsyn (2011). "Fluctuation Relation for Heat Engines". J. Phys. A: Math. Theor. 44: 405001. . This relation transforms the Carnot's inequality into exact equality that is applied to an arbitrary heat engine coupled to two heat reservoirs and operating at arbitrary rate.

[edit] Efficiency of real heat engines

Carnot realized that in reality it is not possible to build a thermodynamically reversible engine, so real heat engines are less efficient than indicated by Equation 3. In addition, real engines that operate along this cycle are rare. Nevertheless, Equation 3 is extremely useful for determining the maximum efficiency that could ever be expected for a given set of thermal reservoirs. Although Carnot's cycle is an idealisation, the expression of Carnot efficiency is still useful. Consider the average temperatures,

at which heat is input and output, respectively. Replace TH and TC in Equation (3) by <TH> and <TC> respectively. For the Carnot cycle, or its equivalent, <TH> is the highest temperature available and <TC> the lowest. For other less efficient cycles, <TH> will be lower than TH , and <TC> will be higher than TC. This can help illustrate, for example, why a reheater or a regenerator can improve the thermal efficiency of steam power plants -- and why the thermal efficiency of combined-cycle power plants (which incorporate gas turbines operating at even higher temperatures) exceeds that of conventional steam plants. See also: Heat Engine efficiency and other performance criteria

[edit] See also


Carnot heat engine Reversible process (thermodynamics) Carnot cycle graphs (above) should not be confused with Karnaugh maps in boolean logic and digital electronics.

[edit] References

Carnot, Sadi (1824). Rflexions sur la puissance motrice du feu et sur les machines propres dvelopper cette puissance. Paris: Bachelier. http://books.google.com/books?id=YcY9AAAAMAAJ. (French) Carnot, Sadi; Thurston, Robert Henry (editor and translator) (1890). Reflections on the Motive Power of Heat and on Machines Fitted to Develop That Power. New York: J. Wiley & Sons. (full text of 1897 ed.)) (html) Feynman, Richard P.; Leighton, Robert B.; Sands, Matthew (1963). The Feynman Lectures on Physics. Addison-Wesley Publishing Company. pp. 444f. ISBN 0201021161.

Halliday, David; Resnick, Robert (1978). Physics (3rd ed. ed.). John Wiley & Sons. pp. 541548. ISBN 0471024562. Kittel, Charles; Kroemer, Herbert (1980). Thermal Physics (2nd ed. ed.). W. H. Freeman Company. ISBN 0-7167-1088-9.

[edit] External links


Hyperphysics article on the Carnot cycle. Interactive Java applet showing behavior of a Carnot engine. [hide]

v t e

Thermodynamic cycles
Without phase change (hot air engines)

External combustion cycles

Bell Coleman Brayton/Joule Carnot Ericsson Stirling Stirling (Pseudo / Adiabatic) Stoddard

With phase change


Kalina Rankine (Organic Rankine) Regenerative Atkinson Brayton/Joule Diesel Expander Gas-generator Homogeneous Charge Compression Ignition Lenoir Miller Otto Pressure-fed Staged combustion

Internal combustion cycles

Mixed cycles

Combined HEHC Mixed/Dual Hampson-Linde Kleemenko Pulse tube Regenerative cooling Transcritical Vapor absorption Vapor-compression Siemens Vuilleumier Barton Humphrey Scuderi

Refrigeration cycles

Uncategorized

Retrieved from "http://en.wikipedia.org/w/index.php?title=Carnot_cycle&oldid=487201885" View page ratings

Rate this page


What's this? Trustworthy Objective Complete Well-written I am highly knowledgeable about this topic (optional) Submit ratings Saved successfully Your ratings have not been submitted yet Categories: Thermodynamic cycles Fundamental physics concepts Atmospheric thermodynamics Hidden categories: Articles with French language external links

Personal tools

Log in / create account

Namespaces

Article Talk

Variants Views Actions Search


Search Special:Search

Read Edit View history

Navigation

Main page Contents Featured content Current events Random article Donate to Wikipedia

Interaction Toolbox

Help About Wikipedia Community portal Recent changes Contact Wikipedia

What links here Related changes Upload file Special pages Permanent link

Cite this page Rate this page

Print/export

Create a book Download as PDF Printable version

Languages

Bosanski Catal esky Deutsch Eesti Espaol Esperanto Euskara Franais Galego Hrvatski Italiano Lietuvi Magyar Nederlands Polski Portugus Slovenina Slovenina / Srpski Suomi Trke

This page was last modified on 13 April 2012 at 17:08. Text is available under the Creative Commons Attribution-ShareAlike License; additional terms may apply. See Terms of use for details. Wikipedia is a registered trademark of the Wikimedia Foundation, Inc., a nonprofit organization. Contact us Privacy policy About Wikipedia Disclaimers Mobile view

Our updated Terms of Use will become effective on May 25, 2012. Find out more.

Combined cycle
From Wikipedia, the free encyclopedia

Jump to: navigation, search

Thermodynamics

Branches[show]

Laws[show]

Systems[show]

System properties[show]

Material properties[show]

Specific heat capacity Compressibility Thermal expansion

Equations[show]

Potentials[show]

Internal energy Enthalpy Helmholtz free energy Gibbs free energy


History and culture[show]

Scientists[show]

v t e

"CCGT" redirects here. It is not to be confused with Closed-cycle gas turbine. In electric power generation a combined cycle is an assembly of heat engines that work in tandem off the same source of heat, converting it into mechanical energy, which in turn usually drives electrical generators. The principle is that the exhaust of one heat engine is used as the heat source for another, thus extracting more useful energy from the heat, increasing the system's overall efficiency. This works because heat engines are only able to use a portion of the energy their fuel generates (usually less than 50%). The remaining heat (e.g., hot exhaust fumes) from combustion is generally wasted. Combining two or more thermodynamic cycles results in improved overall efficiency, reducing fuel costs. In stationary power plants, a successful, common combination is the Brayton cycle (in the form of a turbine burning natural gas or synthesis gas from coal) and the Rankine cycle (in the form of a steam power plant). Multiple stage turbine or steam cylinders are also common. Historically successful combined cycles have used hot cycles with mercury vapor turbines, magnetohydrodynamic generators or molten carbonate fuel cells, with steam plants for the low temperature bottoming cycle. Bottoming cycles operating from a steam

condenser's heat are theoretically possible, but uneconomical because of the very large, expensive equipment needed to extract energy from the small temperature differences between condensing steam and outside air or water. However, it is common in cold climates (such as Finland) to drive community heating systems from a power plant's condenser heat. Such cogeneration systems can yield theoretical efficiencies above 95%. In automotive and aeronautical engines, turbines have been driven from the exhausts of Otto and Diesel cycles. These are called turbo-compound engines. Aside from turbochargers, they have failed commercially because their mechanical complexity and weight are less economical than multistage turbines. Stirling engines are also a good theoretical fit for this application. In a combined cycle power plant (CCPP), or combined cycle gas turbine (CCGT) plant, a gas turbine generator generates electricity, and the heat of its exhaust is used to make steam, which in turn drives a steam turbine to generate additional electricity. This last step enhances the efficiency of electricity generation, and combined-cycle plants can achieve efficiencies of 60%. Many new gas power plants in North America and Europe are of this type. Such an arrangement used for marine propulsion is called combined gas (turbine) and steam (turbine) (COGAS).

Contents
[hide]

1 Design principle o 1.1 Typical size of CCGT plants o 1.2 Efficiency of CCGT plants o 1.3 Boosting Efficiency o 1.4 Supplementary firing and blade cooling 2 Fuel for combined cycle power plants o 2.1 Configuration of CCGT plants 3 Integrated gasification combined cycle (IGCC) 4 Integrated solar combined cycle (ISCC) 5 Automotive use 6 Aeromotive use 7 See also 8 References 9 External links

[edit] Design principle

Working principle of a combined cycle power plant (Legend: 1-Electric generators, 2Steam turbine, 3-Condenser, 4-Pump, 5-Boiler/heat exchanger, 6-Gas turbine) In a thermal power station, water is the working medium. High pressure steam requires strong, bulky components. High temperatures require expensive alloys made from nickel or cobalt, rather than inexpensive steel. These alloys limit practical steam temperatures to 655 C while the lower temperature of a steam plant is fixed by the boiling point of water. With these limits, a steam plant has a fixed upper efficiency of 35 to 42%. An open circuit gas turbine cycle has a compressor, a combustor and a turbine. For gas turbines the amount of metal that must withstand the high temperatures and pressures is small, and lower quantities of expensive materials can be used. In this type of cycle, the input temperature to the turbine (the firing temperature), is relatively high (900 to 1,400 C). The output temperature of the flue gas is also high (450 to 650 C). This is therefore high enough to provide heat for a second cycle which uses steam as the working fluid; (a Rankine cycle). In a combined cycle power plant, the heat of the gas turbine's exhaust is used to generate steam by passing it through a heat recovery steam generator (HRSG) with a live steam temperature between 420 and 580 C. The condenser of the Rankine cycle is usually cooled by water from a lake, river, sea or cooling towers. This temperature can be as low as 15 C In an automotive powerplant, an Otto, Diesel, Atkinson or similar engine would provide one part of the cycle and the waste heat would power a Rankine cycle steam or Stirling engine, which could either power ancillaries (such as the alternator) or be connected to the crankshaft by a turbo compounding system.[citation needed]

[edit] Typical size of CCGT plants


For large scale power generation, a typical set would be a 270 MW gas turbine coupled to a 130 MW steam turbine giving 400 MW. A typical power station might consist of between 1 and 6 such sets. Plant size is important in the cost of the plant. The larger plant sizes benefit from economies of scale (lower initial cost per kilowatt) and improved efficiency.

A single shaft combined cycle plant comprises a gas turbine and a steam turbine driving a common generator. In a multi-shaft combined cycle plant, each gas turbine and each steam turbine has its own generator. The single shaft design provides slightly less initial cost and slightly better efficiency than if the gas and steam turbines had their own generators. The multi-shaft design enables 2 or more gas turbines to operate in conjunction with a single steam turbine, which can be more economical than a number of single shaft units. The primary disadvantage of multiple stage combined cycle power plants is that the number of steam turbines, condensers and condensate systems - and perhaps the number of cooling towers and circulating water systems - increases to match the number of gas turbines. For a multiple shaft combined cycle power plant there is only one steam turbine, condenser and the rest of the heat sink for up to three gas turbines; only their size increases. Having only one large steam turbine and heat sink results in low cost because of economies of scale. A larger steam turbine also allows the use of higher pressures and results in a more efficient steam cycle. Thus the overall plant size and the associated number of gas turbines required have a major impact on whether a single shaft combined cycle power plant or a multiple shaft combined cycle power plant is more economical. Gas turbines of about 150 MW size are already in operation manufactured by at least four separate groups - General Electric and its licensees, Alstom, Siemens, and Westinghouse/Mitsubishi. These groups are also developing, testing and/or marketing gas turbine sizes of about 200 MW. Combined cycle units are made up of one or more such gas turbines, each with a waste heat steam generator arranged to supply steam to a single steam turbine, thus forming a combined cycle block or unit. Typical Combined cycle block sizes offered by three major manufacturers (Alstom, General Electric and Siemens) are roughly in the range of 50 MW to 500 MW and costs are about $600/kW.

[edit] Efficiency of CCGT plants


To avoid confusion, the efficiency of heat engines and power stations should be stated HHV (aka Gross Heating Value) or LCV (aka Net Heating value), and whether Gross output at the generator terminals or Net Output at the power station fence are being considered. In general in service Combined Cycle efficiencies are over 50 percent on a lower heating value and Gross Output basis. Most combined cycle units, especially the larger units, have peak, steady state efficiencies of 55 to 59%. Research aimed at 1370C (2500F) turbine inlet temperature has led to even more efficient combined cycles and 60 percent efficiency has been reached in the combined cycle unit of Baglan Bay, a GE Htechnology gas turbine with a NEM 3 pressure reheat boiler, utilising steam from the HRSG to cool the turbine blades. Siemens AG announced in May 2011 to have achieved a 60.75% net efficiency with a 578 megawatts SGT5-8000H gas turbine at the Irsching Power Station.[1]

By combining both gas and steam cycles, high input temperatures and low output temperatures can be achieved. The efficiency of the cycles add, because they are powered by the same fuel source. So, a combined cycle plant has a thermodynamic cycle that operates between the gas-turbine's high firing temperature and the waste heat temperature from the condensers of the steam cycle. This large range means that the Carnot efficiency of the cycle is high. The actual efficiency, while lower than this, is still higher than that of either plant on its own.[2] The actual efficiency achievable is a complex area.[3] The electric efficiency of a combined cycle power station, calculated as electric energy produced as a percent of the lower heating value of the fuel consumed, may be as high as 58 percent when operating new, i.e. unaged, and at continuous output which are ideal conditions. As with single cycle thermal units, combined cycle units may also deliver low temperature heat energy for industrial processes, district heating and other uses. This is called cogeneration and such power plants are often referred to as a Combined Heat and Power (CHP) plant.

[edit] Boosting Efficiency


The efficiency of CCGT and GT can be boosted by pre-cooling combustion air. This is practised in hot climates and also has the effect of increasing power output. This is achieved by evaporative cooling of water using a moist matrix placed in front of the turbine, or by using Ice storage air conditioning. The latter has the advantage of greater improvements due to the lower temperatures available. Furthermore, ice storage can be used as a means of load control or load shifting since ice can be made during periods of low power demand and, potentially in the future the anticipated high availability of other resources such as renewables during certain periods.

[edit] Supplementary firing and blade cooling


Supplementary firing may be used in combined cycles (in the HRSG) raising exhaust temperatures from 600C (GT exhaust) to 800 or even 1000C. Using supplemental firing will however not raise the combined cycle efficiency for most combined cycles. For single boilers it may raise the efficiency if fired to 700- 750C - for multiple boilers however, supplemental firing is often used to improve peak power production of the unit, or to enable higher steam production to compensate for failure of a second unit. Maximum supplementary firing refers to the maximum fuel that can be fired with the oxygen available in the gas turbine exhaust. The steam cycle is conventional with reheat and regeneration. Hot gas turbine exhaust is used as the combustion air. Regenerative air preheater is not required. A fresh air fan which makes it possible to operate the steam plant even when the gas turbine is not in operation,increases the availability of the unit. The use of large supplementary firing in Combined Cycle Systems with high gas turbine inlet temperatures causes the efficiency to drop. For this reason the Combined Cycle Plants with maximum supplementary firing are only of minimal importance today, in comparison to simple Combined Cycle installations. However, they have two advantages

that is a) coal can be burned in the steam generator as the supplementary fuel, b) has very good part load efficiency. The HRSG can be designed with supplementary firing of fuel after the gas turbine in order to increase the quantity or temperature of the steam generated. Without supplementary firing, the efficiency of the combined cycle power plant is higher, but supplementary firing lets the plant respond to fluctuations of electrical load. Supplementary burners are also called duct burners. More fuel is sometimes added to the turbine's exhaust. This is possible because the turbine exhaust gas (flue gas) still contains some oxygen. Temperature limits at the gas turbine inlet force the turbine to use excess air, above the optimal stoichiometric ratio to burn the fuel. Often in gas turbine designs part of the compressed air flow bypasses the burner and is used to cool the turbine blades. Supplementary firing raises the temperature of the exhaust gas from 800 to 900 degree Celsius. Relatively high flue gas temperature raises the condition of steam (84 bar, 525 degree Celsius) thereby improving the efficiency of steam cycle.

[edit] Fuel for combined cycle power plants


The turbines used in Combined Cycle Plants are commonly fueled with natural gas. However, global natural gas reserves are expected to be fully consumed by 2070.[4] Despite this fact, it is becoming the fuel of choice for an increasing amount of private investors and consumers because it is more versatile than coal or oil and can be used in 90% of energy applications. Chile which once depended on hydropower for 70% of its electricity supply, is now boosting its gas supplies to reduce reliance on its drought afflicted hydro dams. Similarly China is tapping its gas reserves to reduce reliance on coal, which is currently burned to generate 80% of the countrys electric supply. Where the extension of a gas pipeline is impractical or cannot be economically justified, electricity needs in remote areas can be met with small scale Combined Cycle Plants, using renewable fuels. Instead of natural gas, Combined Cycle Plants can be filled with biogas derived from agricultural and forestry waste, which is often readily available in rural areas. Combined cycle plants are usually powered by natural gas, although fuel oil, synthesis gas or other fuels can be used. The supplementary fuel may be natural gas, fuel oil, or coal. Biofuels can also be used. Integrated solar combined cycle power stations combine the energy harvested from solar radiation with another fuel to cut fuel costs and environmental impact. The first such system to come online is Yazd power plant, Iran[5][6] and more are under construction at Hassi R'mel, Algeria and Ain Beni Mathar, Morocco. Next generation nuclear power plants are also on the drawing board which will take advantage of the higher temperature range made available by the Brayton top cycle, as well as the increase in thermal efficiency offered by a Rankine bottoming cycle.

Low-Grade Fuel for Turbines: Gas turbines burn mainly natural gas and light oil. Crude oil, residual, and some distillates contain corrosive components and as such require fuel treatment equipment. In addition, ash deposits from these fuels result in gas turbine deratings of up to 15 percent. They may still be economically attractive fuels however, particularly in combined-cycle plants. Sodium and potassium are removed from residual, crude and heavy distillates by a water washing procedure. A simpler and less expensive purification system will do the same job for light crude and light distillates. A magnesium additive system may also be needed to reduce the corrosive effects if vanadium is present. Fuels requiring such treatment must have a separate fuel-treatment plant and a system of accurate fuel monitoring to assure reliable, low-maintenance operation of gas turbines.

[edit] Configuration of CCGT plants


The combined-cycle system includes single-shaft and multi-shaft configurations. The single-shaft system consists of one gas turbine, one steam turbine, one generator and one Heat Recovery Steam Generator (HRSG), with the gas turbine and steam turbine coupled to the single generator in a tandem arrangement on a single shaft. Key advantages of the single-shaft arrangement are operating simplicity, smaller footprint, and lower startup cost. Single-shaft arrangements, however, will tend to have less flexibility and equivalent reliability than multi-shaft blocks. Additional operational flexibility is provided with a steam turbine which can be disconnected, using a synchro-self-shifting (SSS) Clutch,[7] for start up or for simple cycle operation of the gas turbine. Multi-shaft systems have one or more gas turbine-generators and HRSGs that supply steam through a common header to a separate single steam turbine-generator. In terms of overall investment a multi-shaft system is about 5% higher in costs. Single- and multiple-pressure non-reheat steam cycles are applied to combined-cycle systems equipped with gas turbines having rating point exhaust gas temperatures of approximately 540 C or less. Selection of a single- or multiple-pressure steam cycle for a specific application is determined by economic evaluation which considers plant installed cost, fuel cost and quality, plant duty cycle, and operating and maintenance cost. Multiple-pressure reheat steam cycles are applied to combined-cycle systems with gas turbines having rating point exhaust gas temperatures of approximately 600 C. The most efficient power generation cycles are those with unfired HRSGs with modular pre-engineered components. These unfired steam cycles are also the lowest in cost. Supplementary-fired combined-cycle systems are provided for specific application. The primary regions of interest for cogeneration combined-cycle systems are those with unfired and supplementary fired steam cycles. These systems provide a wide range of thermal energy to electric power ratio and represent the range of thermal energy

capability and power generation covered by the product line for thermal energy and power systems.

[edit] Integrated gasification combined cycle (IGCC)


An integrated gasification combined cycle, or IGCC, is a power plant using synthesis gas (syngas). Syngas can be produced from a number of sources, including coal and biomass. The system utilizes gas and steam turbines, the steam turbine operating off of the heat leftover from the gas turbine. This process can raise electricity generation efficiency to around 50%.

[edit] Integrated solar combined cycle (ISCC)


An integrated solar combined cycle, or ISCC, is a power plant using solar thermal collectors. This is typically in the form of parabolic troughs.

[edit] Automotive use


Any turbocharged engine is effectively a combined cycle with the turbo charger extracting extra energy from the exhaust gases. Theoretically, this extracted energy could be used to drive the wheels, but it is more practical to use it to force air into the engine which reduces the suction loss and thereby improves the efficiency overall. On large marine diesels turbo-compounding has been employed where the turbocharger physically pushes the engine around via some sort of gearing arrangement. Combined cycles have traditionally only been used in large power plants. BMW, however, has proposed that automobiles use exhaust heat to drive steam turbines.[8] This can even be connected to the car or truck's cooling system to save space and weight, but also to provide a condenser in the same location as the radiator and preheating of the water using heat from the engine block. However, stirling engines can also be used if light weight is a priority (as in a sports car or racing application), because they use a gas such as air rather than water as the working fluid. It may be possible to use the pistons in a reciprocating engine for both combustion and steam expansion like in the Crower six stroke.[9]

[edit] Aeromotive use


Some versions of the Wright R-3350 were produced as turbo-compound engines. Three turbines driven by exhaust gases, known as power recovery turbines, provided nearly 600 hp at takeoff. These turbines added power to the engine crankshaft through bevel gears and fluid couplings.[10]

[edit] See also

Heat recovery steam generator Hydrogen-cooled turbogenerator Mercury vapour turbine Cost of electricity by source COGAS

[edit] References
1. ^ "Siemens pushes world record in efficiency to over 60 percent while achieving maximum operating flexibility". Siemens AG. 19 May 2011. http://www.siemens.com/press/en/pressrelease/?press=/en/pressrelease/2011/fossil_powe r_generation/efp201105064.htm. 2. ^ "Efficiency by the Numbers" by Lee S. Langston 3. ^ "The difference between LCV and HCV (or Lower and Higher Heating Value, or Net and Gross) is clearly understood by all energy engineers. There is no right or wrong definition.". Claverton Energy Research Group. http://www.claverton-energy.com/thedifference-between-lcv-and-hcv-or-lower-and-higher-heating-value-or-net-and-gross-isclearly-understood-by-all-energy-engineers-there-is-no-right-or-wrong-definition.html. 4. ^ "Natural Gas reserves". BP. http://www.bp.com/sectiongenericarticle800.do?categoryId=9037178&contentId=706862 4. Retrieved 19 September 2011. 5. ^ "Yazd Solar Energy Power Plant 1st in its kind in world". Payvand Iran news. 13 April 2007. http://payvand.com/news/07/apr/1132.html. 6. ^ "CCGT Plants in Iran - other provinces". Power Plants Around the World Photo Gallery. http://www.industcards.com/cc-iran.htm. 7. ^ "SSS Clutch Operating Principle". SSS Gears Limited. http://www.sssclutch.com/howitworks/100-2SSSPrinciples.pdf. 8. ^ "BMW Turbosteamer gets hot and goes" by John Neff, AutoBlog, December 9, 2005 9. ^ "Inside Bruce Crowers Six-Stroke Engine" By Pete Lyons, AutoWeek, February 23, 2006 10. ^ Goleta Air and Space Museum: 2002 Camarillo EAA Fly-in

[edit] External links


Hunstown: Ireland's most efficient power plant @ Siemens Power Generation website ABB Power Generation website @ ABB_Group Natural Gas Combined-cycle Gas Turbine Power Plants Northwest Power Planning Council, New Resource Characterization for the Fifth Power Plan, August 2002 Combined cycle solar power [show]

v t

Thermodynamic cycles

[show]

v t e

Electricity delivery

Nonrenewable

Coal Fossil-fuel power station Natural gas Petroleum Nuclear Oil shale Biomass Geothermal Hydro Marine Solar Wind

Renewable

Retrieved from "http://en.wikipedia.org/w/index.php?title=Combined_cycle&oldid=489167875" View page ratings

Rate this page


What's this? Trustworthy Objective Complete Well-written I am highly knowledgeable about this topic (optional) Submit ratings Saved successfully Your ratings have not been submitted yet Categories: Thermodynamic cycles Mechanical engineering Power station technology Energy conversion Hidden categories: All articles with unsourced statements Articles with unsourced statements from March 2011
Personal tools

Log in / create account

Namespaces

Article Talk

Variants Views

Actions Search
Search

Read Edit View history

Special:Search

Navigation

Main page Contents Featured content Current events Random article Donate to Wikipedia

Interaction Toolbox

Help About Wikipedia Community portal Recent changes Contact Wikipedia

What links here Related changes Upload file Special pages Permanent link Cite this page Rate this page

Print/export

Create a book Download as PDF Printable version

Languages

Deutsch Espaol

Franais Hrvatski Nederlands Polski Slovenina Svenska This page was last modified on 25 April 2012 at 15:21. Text is available under the Creative Commons Attribution-ShareAlike License; additional terms may apply. See Terms of use for details. Wikipedia is a registered trademark of the Wikimedia Foundation, Inc., a nonprofit organization. Contact us Privacy policy About Wikipedia Disclaimers Mobile view

Our updated Terms of Use will become effective on May 25, 2012. Find out more.

Diesel cycle
From Wikipedia, the free encyclopedia

Jump to: navigation, search

Thermodynamics

Branches[show]

Laws[show]

Systems[show]

System properties[show]

Material properties[show]

Specific heat capacity Compressibility Thermal expansion

Equations[show]

Potentials[show]

Internal energy Enthalpy Helmholtz free energy Gibbs free energy


History and culture[show]

Scientists[show]

v t e

The Diesel cycle is the thermodynamic cycle which approximates the pressure and volume of the combustion chamber of the Diesel engine, invented by Rudolph Diesel in 1897. It is assumed to have constant pressure during the first part of the "combustion" phase ( to in the diagram, below). This is an idealized mathematical model: real physical Diesels do have an increase in pressure during this period, but it is less pronounced than in the Otto cycle. The idealized Otto cycle of a gasoline engine approximates constant volume during that phase, generating more of a spike in a p-V diagram.

Contents
[hide]

1 The Idealized Diesel Cycle o 1.1 Maximum thermal efficiency 2 Applications o 2.1 Diesel engines

o 2.2 Other internal combustion engines without spark plugs 3 References 4 See also

[edit] The Idealized Diesel Cycle

p-V Diagram for the Ideal Diesel cycle. The cycle follows the numbers 1-4 in clockwise direction. The image on the left shows a p-V diagram for the ideal Diesel cycle; where is pressure and is specific volume. The ideal Diesel cycle follows the following four distinct processes (The color references refer to the color of the line on the diagram.):

Process 1 to 2 is isentropic compression of the fluid (blue colour) Process 2 to 3 is reversible constant pressure heating (red) Process 3 to 4 is isentropic expansion (yellow) Process 4 to 1 is reversible constant volume cooling (green)[1]

The Diesel is a heat engine: it converts heat into work. The isentropic processes are impermeable to heat: heat flows into the loop through the left expanding isobaric process and some of it flows back out through the right depressurizing process, and the heat that remains does the work.

Work in ( ) is done by the piston compressing the working fluid Heat in ( ) is done by the combustion of the fuel Work out ( ) is done by the working fluid expanding on to the piston (this produces usable torque)

Heat out (

) is done by venting the air

[edit] Maximum thermal efficiency


The maximum thermal efficiency of a Diesel cycle is dependent on the compression ratio and the cut-off ratio. It has the following formula under cold air standard analysis:

where is thermal efficiency is the cut-off ratio combustion phase) (ratio between the end and start volume for the

r is the compression ratio


is ratio of specific heats (Cp/Cv)[2] The cut-off ratio can be expressed in terms of temperature as shown below:

can be approximated to the flame temperature of the fuel used. The flame temperature can be approximated to the adiabatic flame temperature of the fuel with corresponding air-to-fuel ratio and compression pressure, . can be approximated to the inlet air temperature. This formula only gives the ideal thermal efficiency. The actual thermal efficiency will be significantly lower due to heat and friction losses. The formula is more complex than the Otto cycle (petrol/gasoline engine) relation that has the following formula;

The additional complexity for the Diesel formula comes around since the heat addition is at constant pressure and the heat rejection is at constant volume. The Otto cycle by comparison has both the heat addition and rejection at constant volume. Comparing the two formulae it can be seen that for a given compression ratio (r), the ideal Otto cycle will be more efficient. However, a Diesel engine will be more efficient overall since it will have the ability to operate at higher compression ratios. If a petrol engine were to have the same compression ratio, then knocking (self-ignition) would occur and this would severely reduce the efficiency, whereas in a Diesel engine, the self ignition is the desired behavior. Additionally, both of these cycles are only idealizations, and the actual behavior does not divide as clearly or sharply. And the ideal Otto cycle formula stated above does not include throttling losses, which do not apply to Diesel engines. The Diesel cycle is a combustion process of a reciprocating internal combustion engine. In it, fuel is ignited by heat generated by compressing air in the combustion chamber, into which fuel is injected. This is in contrast to igniting it with a spark plug as in the Otto cycle (four-stroke/petrol) engine. Diesel engines (heat engines using the Diesel cycle) are used in automobiles, power generation, Diesel-electric locomotives, and submarines.

[edit] Applications
[edit] Diesel engines
Main article: Diesel engine The Diesel engine has the lowest specific fuel consumption of any large internal combustion engine, 0.26 lb/hp.h (0.16 kg/kWh) for very large marine engines. Twostroke Diesels with high pressure forced induction, particularly turbocharging, make up a large percentage of the very largest Diesel engines. In North America, Diesel engines are primarily used in large trucks, where the low-stress, high-efficiency cycle leads to much longer engine life and lower operational costs. These advantages also make the Diesel engine ideal for use in the heavy-haul railroad environment.

[edit] Other internal combustion engines without spark plugs


Many model airplanes use very simple "glow" and "Diesel" engines. Glow engines use glow plugs. "Diesel" model airplane engines have variable compression ratios. Both types depend on special fuels (easily obtainable in such limited quantities) for their ignition timing. Some 19th century or earlier experimental engines used external flames, exposed by valves, for ignition, but this becomes less attractive with increasing compression. (It was

the research of Nicolas Lonard Sadi Carnot that established the thermodynamic value of compression.) A historical implication of this is that the Diesel engine would eventually have been invented without the aid of electricity. See the development of the hot bulb engine and indirect injection for historical significance.

[edit] References
1. ^ Eastop & McConkey 1993, Applied Thermodynamics for Engineering Technologists, Pearson Education Limited, Fifth Edition, p.137 2. ^ The Diesel Engine

[edit] See also


Diesel engine Hot bulb engine Mixed/dual cycle [hide]


v t e

Thermodynamic cycles
Without phase change (hot air engines)

External combustion cycles

Bell Coleman Brayton/Joule Carnot Ericsson Stirling Stirling (Pseudo / Adiabatic) Stoddard

With phase change


Kalina Rankine (Organic Rankine) Regenerative Atkinson Brayton/Joule Diesel

Internal combustion cycles

Expander Gas-generator Homogeneous Charge Compression Ignition Lenoir Miller Otto Pressure-fed Staged combustion Combined HEHC Mixed/Dual Hampson-Linde Kleemenko Pulse tube Regenerative cooling Transcritical Vapor absorption Vapor-compression Siemens Vuilleumier Barton Humphrey Scuderi

Mixed cycles

Refrigeration cycles

Uncategorized

Retrieved from "http://en.wikipedia.org/w/index.php?title=Diesel_cycle&oldid=479999058" View page ratings

Rate this page


What's this? Trustworthy Objective Complete Well-written I am highly knowledgeable about this topic (optional) Submit ratings Saved successfully

Your ratings have not been submitted yet Categories: Thermodynamic cycles
Personal tools

Log in / create account

Namespaces

Article Talk

Variants Views Actions Search


Search Special:Search

Read Edit View history

Navigation

Main page Contents Featured content Current events Random article Donate to Wikipedia

Interaction Toolbox

Help About Wikipedia Community portal Recent changes Contact Wikipedia

What links here

Related changes Upload file Special pages Permanent link Cite this page Rate this page

Print/export

Create a book Download as PDF Printable version

Languages

Deutsch Espaol Esperanto Galego Hrvatski Italiano Magyar Nederlands Norsk (nynorsk) Polski Slovenina This page was last modified on 3 March 2012 at 16:37. Text is available under the Creative Commons Attribution-ShareAlike License; additional terms may apply. See Terms of use for details. Wikipedia is a registered trademark of the Wikimedia Foundation, Inc., a nonprofit organization. Contact us Privacy policy About Wikipedia Disclaimers Mobile view

Our updated Terms of Use will become effective on May 25, 2012. Find out more.

Ericsson cycle
From Wikipedia, the free encyclopedia

Jump to: navigation, search

Thermodynamics

Branches[show]

Laws[show]

Systems[show]

System properties[show]

Material properties[show]

Specific heat capacity Compressibility Thermal expansion

Equations[show]

Potentials[show]

Internal energy Enthalpy Helmholtz free energy Gibbs free energy


History and culture[show]

Scientists[show]

v t e

Rendering of an Ericsson engine. A cold gaseous working fluid, such as atmospheric air (shown in blue), enters the cylinder via a non-return valve at the top-right. The air is compressed by the piston (black) as the piston moves upward. The compressed air is stored in the pneumatic tank (at left). A two-way valve (gray) moves downward to allow pressurized air to pass through the regenerator where it is preheated. The air then enters the space below the piston, which is an externally-heated expansion-chamber. The air

expands and does work on the piston as it moves upward. After the expansion stroke, the two-way valve moves upward, thus closing off the tank and opening the exhaust port. As the piston moves back downward in the exhaust stroke, hot air is pushed back through the regenerator, which reclaims most of the heat, before passing out the exhaust port (left) as cool air. The Ericsson cycle is named after inventor John Ericsson, who designed and built many unique heat engines based on various thermodynamic cycles. He is credited with inventing two unique heat engine cycles and developing practical engines based on these cycles. His first cycle is very similar to what is now called the "Brayton cycle", with the exception that it uses external combustion. His second cycle is now called the Ericsson cycle.

Contents
[hide]

1 Ideal Ericsson cycle o 1.1 Comparison with Stirling and Carnot cycles o 1.2 Comparison with the Brayton cycle 2 Ericsson engine 3 Regenerator 4 History 5 Today's potential 6 References 7 External links

[edit] Ideal Ericsson cycle


The following is a list of the four processes that occur between the four stages of the ideal Ericsson cycle:

Process 1 -> 2: Isothermal compression. The compression space is assumed to be intercooled, so the gas undergoes isothermal compression. The compressed air flows into a storage tank at constant pressure. In the ideal cycle, there is no heat transfer across the tank walls. Process 2 -> 3: Isobaric heat addition. From the tank, the compressed air flows through the regenerator and picks up heat at a high constant-pressure on the way to the heated power-cylinder. Process 3 -> 4: Isothermal expansion. The power-cylinder expansion-space is heated externally, and the gas undergoes isothermal expansion. Process 4 -> 1: Isobaric heat removal. Before the air is released as exhaust, it is passed back through the regenerator, thus cooling the gas at a low constant pressure, and heating the regenerator for the next cycle.

[edit] Comparison with Stirling and Carnot cycles


The Ericsson cycle is often compared to the Stirling cycle, since the engine designs based on these respective cycles are both external combustion engines with regenerators. The Ericsson is perhaps most similar to the so called "double-acting" type of Stirling engine, in which the displacer piston also acts as the power piston. Theoretically, both of these cycles have so called ideal efficiency, which is the highest allowed by the second law of thermodynamics. The most well known ideal cycle is the Carnot cycle, although a real Carnot engine is not known to have been invented. The theoretical efficiencies for both, Ericsson and Stirling cycles acting in the same limits are equal to the Carnot Efficiency for same limits.

[edit] Comparison with the Brayton cycle


Main article: Brayton Cycle The first cycle Ericsson developed is now called the "Brayton cycle", commonly applied to the rotary jet engines for airplanes. The second Ericsson cycle is the cycle most commonly referred to as simply the "Ericsson cycle". The (second) Ericsson cycle is also the limit of an ideal gas-turbine Brayton cycle, operating with multistage intercooled compression, and multistage expansion with reheat and regeneration. Compared to the Brayton cycle which uses adiabatic compression and expansion, the second Ericsson cycle uses isothermal compression and expansion, thus producing more net work per stroke. Also the use of regeneration in the Ericsson cycle increases efficiency by reducing the required heat input. For further comparisons of thermodynamic cycles, see heat engine. Cycle/Process Compression Heat addition Expansion Heat rejection isobaric adiabatic isobaric Ericsson (First, 1833) adiabatic isobaric isothermal isobaric Ericsson (Second, 1853) isothermal isobaric adiabatic isobaric Brayton (Turbine) adiabatic

[edit] Ericsson engine


The Ericsson engine is based on the Ericsson cycle, and is known as an "external combustion engine", because it is externally heated. To improve efficiency, the engine has a regenerator or recuperator between the compressor and the expander. The engine can be run open- or closed-cycle. Expansion occurs simultaneously with compression, on opposite sides of the piston.

[edit] Regenerator
Main article: Regenerative heat exchanger

Ericsson coined the term "regenerator" for his independent invention of the mixed-flow counter-current heat exchanger. However, Rev. Robert Stirling had invented the same device, prior to Ericsson, so the invention is credited to Stirling. Stirling called it an "economiser" or "economizer", because it increased the fuel economy of various types of heat processes. The invention was found to be useful, in many other devices and systems, where it became more widely used, since other types of engines became favored over the Stirling engine. The term "regenerator" is now the name given to the component in the Stirling engine. The term "recuperator" refers to a separated-flow, counter-current heat exchanger. As if this weren't confusing enough, a mixed-flow regenerator is sometimes used as a quasiseparated-flow recuperator. This can be done through the use of moving valves, or by a rotating regenerator with fixed baffles, or by the use of other moving parts. When heat is recovered from exhaust gases and used to preheat combustion air, typically the term recuperator is used, because the two flows are separate.

[edit] History
In 1791, before Ericsson, Barber proposed a similar engine. The Barber engine used a bellows compressor and a turbine expander, but it lacked a regenerator/recuperator. There are no records of a working Barber engine. Ericsson invented and patented his first engine using an external version of the Brayton cycle in 1833 (number 6409/1833 British). This was 18 years before Joule and 43 years before Brayton. Brayton engines were all piston engines and for the most part, internal combustion versions of the unrecuperated Ericsson engine. The "Brayton Cycle" is now known as the gas turbine cycle, which differs from the original "Brayton Cycle" in the use of a turbine compressor and expander. The gas turbine cycle is used for all modern gas turbine and turbojet engines, however simple cycle turbines are often recuperated to improve efficiency and these recuperated turbines more closely resemble Ericsson's work. Ericsson eventually abandoned the open cycle in favor of the traditional closed Stirling cycle. Ericsson's engine can easily be modified to operate in a closed-cycle mode, using a second, lower-pressure, cooled container between the original exhaust and intake. In closed cycle, the lower pressure can be significantly above ambient pressure, and He or H2 working gas can be used. Because of the higher pressure difference between the upward and downward movement of the work-piston, specific output can be greater than of a valveless Stirling engine. The added cost is the valve. Ericsson's engine also minimizes mechanical losses: the power necessary for compression does not go through crank-bearing frictional losses, but is applied directly from the expansion force. The piston-type Ericsson engine can potentially be the highest efficiency heat engine arrangement ever constructed. Admittedly, this has yet to be proven in practical applications.[citation needed]

The Ericsson-cycle engine (the second of the two discussed here) was used to power a 2,000-ton ship, the caloric ship Ericsson, and ran flawlessly for 73 hours. The combination engine produced about 300 horsepower (220 kW). It had a combination of four dual-piston engines; the larger expansion piston/cylinder, at 14 feet (4.3 m) in diameter, was perhaps the largest piston ever built. Rumor has it that tables were placed on top of those pistons (obviously in the cool compression chamber, not the hot power chamber) and dinner was served and eaten, while the engine was running at full power.[citation needed] At 6.5 RPM the pressure was limited to 8 psi (55 kPa). According to the official report it only consumed 4200 kg coal per 24 hours (original target was 8000 kg, which is still better than contemporary steam engines). The one sea trial proved that even though the engine ran well, the ship was underpowered. Sometime after the trials, the Ericsson sank. When it was raised, the Ericsson-cycle engine was removed and a steam engine took its place. Ericsson designed and built a very great number of engines running on various cycles including steam, Stirling, Brayton, externally heated diesel air fluid cycle. He ran his engines on a great variety of fuels including coal and solar heat. Ericsson was also responsible for an early use of the screw propeller for ship propulsion, in the USS Princeton, built in 184243.

[edit] Today's potential


The Ericsson cycle (and the similar Brayton cycle) receives renewed interest today to extract power from the exhaust heat of gas (and producer gas) engines and solar concentrators. An important advantage of the Ericsson cycle over the widely known Stirling engine is often not recognized:[by whom?] the volume of the heat exchanger does not adversely affect the efficiency. For medium and large engines the cost of valves can be small compared to this advantage. Turbocompressor plus turbine implementations seem favorable in the MWe range, positive displacement compressor plus turbine for Nx100 kWe power, and positive displacement compressor+expander below 100 kW. With high temperature hydraulic fluid, both the compressor and the expander can be liquid ring pump even up to 400C, with rotating casing for best efficiency.

[edit] References

Ericsson's patents. 1833 British and 1851 USA The evolution of the heat engine, by: Ivo Kolin Published Moriya Press, 1972 by Longman Hot Air Caloric and Stirling Engines, by: Robert Sier. Published 1999, by L A Mair. New York Times 1853-03-01 The Caloric Ship Ericsson - Official Report and Correspondence

[edit] External links

1979 RAND report on a new "Ericsson Cycle Gas Turbine Powerplant" design [1] [hide]

v t e

Thermodynamic cycles
Without phase change (hot air engines)

External combustion cycles

Bell Coleman Brayton/Joule Carnot Ericsson Stirling Stirling (Pseudo / Adiabatic) Stoddard

With phase change


Kalina Rankine (Organic Rankine) Regenerative Atkinson Brayton/Joule Diesel Expander Gas-generator Homogeneous Charge Compression Ignition Lenoir Miller Otto Pressure-fed Staged combustion Combined HEHC Mixed/Dual Hampson-Linde Kleemenko Pulse tube

Internal combustion cycles

Mixed cycles

Refrigeration cycles

Regenerative cooling Transcritical Vapor absorption Vapor-compression Siemens Vuilleumier Barton Humphrey Scuderi

Uncategorized

Retrieved from "http://en.wikipedia.org/w/index.php?title=Ericsson_cycle&oldid=480656663" View page ratings

Rate this page


What's this? Trustworthy Objective Complete Well-written I am highly knowledgeable about this topic (optional) Submit ratings Saved successfully Your ratings have not been submitted yet Categories: Thermodynamic cycles Piston engines Hidden categories: All articles with unsourced statements Articles with unsourced statements from February 2011 Articles with specifically marked weasel-worded phrases from February 2011
Personal tools

Log in / create account

Namespaces

Article Talk

Variants Views Actions Search


Search Special:Search

Read Edit View history

Navigation

Main page Contents Featured content Current events Random article Donate to Wikipedia

Interaction Toolbox

Help About Wikipedia Community portal Recent changes Contact Wikipedia

What links here Related changes Upload file Special pages Permanent link Cite this page Rate this page

Print/export

Create a book

Download as PDF Printable version

Languages

Catal Deutsch Franais Slovenina Trke This page was last modified on 7 March 2012 at 13:04. Text is available under the Creative Commons Attribution-ShareAlike License; additional terms may apply. See Terms of use for details. Wikipedia is a registered trademark of the Wikimedia Foundation, Inc., a nonprofit organization. Contact us Privacy policy About Wikipedia Disclaimers Mobile view

Our updated Terms of Use will become effective on May 25, 2012. Find out more.

Gas-generator cycle (rocket)


From Wikipedia, the free encyclopedia

Jump to: navigation, search

Gas generator rocket cycle. Some of the fuel and oxidizer is burned separately to power the pumps and then discarded. Most Gas-generator engines use the fuel for nozzle cooling. The gas generator cycle is a power cycle of a bipropellant rocket engine. Some of the propellant is burned in a gas-generator and the resulting hot gas is used to power the engine's pumps. The gas is then exhausted. Because something is "thrown away" this type of engine is also known as open cycle. There are several advantages to the gas generator cycle over its counterpart, the staged combustion cycle. The gas generator turbine does not need to deal with the counter pressure of injecting the exhaust into the combustion chamber. This simplifies plumbing and turbine design, and results in a less expensive and lighter engine. The main disadvantage is lost efficiency due to discarded propellant. Gas generator cycles tend to have lower specific impulse than staged combustion cycles.[citation needed] As in most cryogenic rocket engines, some of the fuel in a gas-generator cycle is used to cool the nozzle and combustion chamber. Current construction materials cannot withstand extreme temperatures of rocket combustion processes by themselves. Cooling permits the use of rocket engines for relatively longer periods of time with todays

material technology. Without rocket combustion chamber and nozzle cooling, the engine would fail catastrophically.[1]

Contents
[hide]

1 Usage 2 See also 3 References 4 External links

[edit] Usage
Gas generator combustion engines include the following:

Vulcain[2] Merlin rocket engine[3] RS-68 J-2X F-1 (rocket engine) RD-107 CE-20

Gas generator combustion engines rockets:


Ariane 5 Falcon 9 Delta IV Space Launch System Saturn V Soyuz (rocket family) Geosynchronous Satellite Launch Vehicle III

[edit] See also


Expander cycle Pressure-fed cycle Rocket engine Staged combustion cycle

[edit] References

1. ^ Liquid Hydrogen as a Propulsion Fuel, 1945-1959 2. ^ "Vulcain 2 Rocket Engine". EADS. http://cs.astrium.eads.net/sp/LauncherPropulsion/Vulcain-2-Rocket-Engine.html. 3. ^ "SpaceX Merlin Engine". SpaceX. http://www.spacex.com/falcon1.php#merlin_engine.

[edit] External links


Rocket power cycles [1] [show]


v t e

Spacecraft propulsion
State Liquid-fuel rocket Solid-fuel rocket Hybrid rocket Liquid propellants (Cryogenic Hypergolic) Monopropellant Propellants Bipropellant (Staged combustion cycle Expander cycle Gasgenerator cycle Pressure-fed cycle) Tripropellant Electrostatic Electromagnetic Electrothermal Other Closed system Colloid thruster Ion thruster (Electrostatic ion thruster Hall effect thruster Field Emission Electric Propulsion) Pulsed inductive thruster Magnetoplasmadynamic thruster Electrodeless plasma thruster VASIMR Pulsed plasma thruster Helicon double-layer thruster Arcjet rocket MagBeam Resistojet rocket High-power electric propulsion Mass driver

Nuclear electric rocket Nuclear thermal rocket (Radioisotope Salt-water Gas core "Lightbulb")

Nuclear pulse propulsion (Antimatter) Fusion rocket (Bussard Open system ramjet) Fission-fragment rocket (Fission sail) Nuclear photonic rocket

[show]

v t e

Thermodynamic cycles

Retrieved from "http://en.wikipedia.org/w/index.php?title=Gasgenerator_cycle_(rocket)&oldid=487970419" View page ratings

Rate this page


What's this? Trustworthy Objective Complete Well-written I am highly knowledgeable about this topic (optional) Submit ratings Saved successfully Your ratings have not been submitted yet Categories: Combustion Rocket engines Spacecraft propulsion Thermodynamic cycles Hidden categories: All articles with unsourced statements Articles with unsourced statements from January 2010
Personal tools

Log in / create account

Namespaces

Article Talk

Variants Views Actions Search


Search Special:Search

Read Edit View history

Navigation

Main page Contents Featured content Current events Random article Donate to Wikipedia

Interaction Toolbox

Help About Wikipedia Community portal Recent changes Contact Wikipedia

What links here Related changes Upload file Special pages Permanent link Cite this page Rate this page

Print/export

Create a book Download as PDF Printable version

Languages

Catal Espaol This page was last modified on 18 April 2012 at 08:31. Text is available under the Creative Commons Attribution-ShareAlike License; additional terms may apply. See Terms of use for details. Wikipedia is a registered trademark of the Wikimedia Foundation, Inc., a nonprofit organization. Contact us Privacy policy About Wikipedia Disclaimers Mobile view

Our updated Terms of Use will become effective on May 25, 2012. Find out more.

Heat pump and refrigeration cycle


From Wikipedia, the free encyclopedia

Jump to: navigation, search For details of practical heat pumps, see Heat pump.

Thermodynamics

Branches[show]

Laws[show]

Systems[hide] State: Equation of state Ideal gas Real gas Phase of matter Equilibrium Control volume Instruments

Processes: Isobaric Isochoric Isothermal Adiabatic Isentropic Isenthalpic Quasistatic Polytropic Free expansion Reversibility Irreversibility Endoreversibility

Cycles: Heat engines Heat pumps Thermal efficiency System properties[show]

Material properties[show]

Specific heat capacity Compressibility Thermal expansion

Equations[show]

Potentials[show]

Internal energy Enthalpy Helmholtz free energy Gibbs free energy


History and culture[show]

Scientists[show]

v t e

Thermodynamic heat pump cycles or refrigeration cycles are the conceptual and mathematical models for heat pumps and refrigerators. A heat pump is a machine or device that moves heat from one location (the 'source') at a lower temperature to another location (the 'sink' or 'heat sink') at a higher temperature using mechanical work or a high-temperature heat source.[1] Thus a heat pump may be thought of a "heater" if the objective is to warm the heat sink (as when warming the inside of a home on a cold day), or a "refrigerator" if the objective is to cool the heat source (as in the normal operation of a freezer). In either case, the operating principles are identical.[2] Heat is moved from a cold place to a warm place.

Contents
[hide]

1 Thermodynamic cycles o 1.1 Vapor-compression cycle o 1.2 Vapor absorption cycle o 1.3 Gas cycle o 1.4 Stirling cycle o 1.5 Comparison with Combined Heat and Power, CHP. 2 References 3 External links

[edit] Thermodynamic cycles


According to the second law of thermodynamics heat cannot spontaneously flow from a colder location to a hotter area; work is required to achieve this.[3] An air conditioner requires work to cool a living space, moving heat from the cooler interior (the heat source) to the warmer outdoors (the heat sink). Similarly, a refrigerator moves heat from inside the cold icebox (the heat source) to the warmer room-temperature air of the kitchen (the heat sink). The operating principle of the refrigeration cycle was described mathematically by Sadi Carnot in 1824 as a heat engine. A heat pump can be thought of as heat engine which is operating in reverse. Heat pump and refrigeration cycles can be classified as vapor compression, vapor absorption, gas cycle, or Stirling cycle types.

[edit] Vapor-compression cycle


Main article: Vapor-compression refrigeration The vapor-compression cycle is used in most household refrigerators as well as in many large commercial and industrial refrigeration systems. Figure 1 provides a schematic diagram of the components of a typical vapour-compression refrigeration system.

Figure 1: Vapor compression refrigeration The thermodynamics of the cycle can be analyzed on a diagram[4][5] as shown in Figure 2. In this cycle, a circulating refrigerant such as Freon enters the compressor as a vapor. The vapor is compressed at constant entropy and exits the compressor superheated. The superheated vapor travels through the condenser which first cools and removes the superheat and then condenses the vapor into a liquid by removing additional heat at constant pressure and temperature. The liquid refrigerant goes through the expansion valve (also called a throttle valve) where its pressure abruptly decreases, causing flash evaporation and auto-refrigeration of, typically, less than half of the liquid.

Figure 2:TemperatureEntropy diagram That results in a mixture of liquid and vapor at a lower temperature and pressure. The cold liquid-vapor mixture then travels through the evaporator coil or tubes and is completely vaporized by cooling the warm air (from the space being refrigerated) being blown by a fan across the evaporator coil or tubes. The resulting refrigerant vapor returns to the compressor inlet to complete the thermodynamic cycle. The above discussion is based on the ideal vapor-compression refrigeration cycle, and does not take into account real-world effects like frictional pressure drop in the system, slight thermodynamic irreversibility during the compression of the refrigerant vapor, or non-ideal gas behavior (if any).

[edit] Vapor absorption cycle


Main article: Absorption refrigerator In the early years of the twentieth century, the vapor absorption cycle using waterammonia systems was popular and widely used but, after the development of the vapor compression cycle, it lost much of its importance because of its low coefficient of performance (about one fifth of that of the vapor compression cycle). Nowadays, the vapor absorption cycle is used only where waste heat is available or where heat is derived from solar collectors. The absorption cycle is similar to the compression cycle, except for the method of raising the pressure of the refrigerant vapor. In the absorption system, the compressor is replaced by an absorber which dissolves the refrigerant in a suitable liquid, a liquid pump which raises the pressure and a generator which, on heat addition, drives off the refrigerant vapor from the high-pressure liquid. Some work is required by the liquid pump but, for a

given quantity of refrigerant, it is much smaller than needed by the compressor in the vapor compression cycle. In an absorption refrigerator, a suitable combination of refrigerant and absorbent is used. The most common combinations are ammonia (refrigerant) and water (absorbent), and water (refrigerant) and lithium bromide (absorbent).

[edit] Gas cycle


When the working fluid is a gas that is compressed and expanded but does not change phase, the refrigeration cycle is called a gas cycle. Air is most often this working fluid. As there is no condensation and evaporation intended in a gas cycle, components corresponding to the condenser and evaporator in a vapor compression cycle are the hot and cold gas-to-gas heat exchangers in gas cycles. The gas cycle is less efficient than the vapor compression cycle because the gas cycle works on the reverse Brayton cycle instead of the reverse Rankine cycle. As such the working fluid does not receive and reject heat at constant temperature. In the gas cycle, the refrigeration effect is equal to the product of the specific heat of the gas and the rise in temperature of the gas in the low temperature side. Therefore, for the same cooling load, a gas refrigeration cycle will require a large mass flow rate and would be bulky. Because of their lower efficiency and larger bulk, air cycle coolers are not often applied in terrestrial refrigeration. The air cycle machine is very common, however, on gas turbine-powered jet airliners since compressed air is readily available from the engines' compressor sections. These jet aircraft's cooling and ventilation units also serve the purpose of pressurizing the aircraft cabin.

[edit] Stirling cycle


Main article: Stirling cycle The Stirling cycle heat engine can be driven in reverse, using a mechanical energy input to drive heat transfer in a reversed direction (i.e. a heat pump, or refrigerator). There are several design configurations for such devices that can be built. Several such setups require rotary or sliding seals, which can introduce difficult tradeoffs between frictional losses and refrigerant leakage. The Free Piston Stirling Cooler (FPSC) is an elegant, completely sealed heat transfer system that has only two moving parts (a piston and a displacer), and uses helium as the working fluid. The piston is typically driven by an oscillating magnetic field that is the source of the power needed to drive the refrigeration cycle. The magnetic drive allows the piston to be driven without requiring any seals, gaskets, O-rings, or other compromises to the hermetically sealed system. [6] Claimed advantages for the system include environmental friendliness, cooling capacity, light weight, compact size, precise controllability, and high efficiency.[7]

The FPSC was invented in 1964 by William Beale, a professor of Mechanical Engineering at Ohio University in Athens, Ohio. He founded and continues to be associated with Sunpower Inc., [8] which specializes primarily in researching and developing FPSC systems for a wide variety of military, aerospace, industrial, and commercial applications. Sunpower also makes cryocoolers and special pulse tube coolers capable of reaching below 40K (around 390F, or 230C). A FPSC cooler made by Sunpower was used by NASA to cool instrumentation in satellites.[9] Since 2002. another leading supplier of FPSC technology has been the Twinbird Company [10] in Japan, which also markets a broad line of household appliances. Both Sunpower and Twinbird appear to work in collaboration with Global Cooling NV, which is located in the Netherlands, but has a research center in Athens, Ohio.[11]

[edit] Comparison with Combined Heat and Power, CHP.


A heat pump may be compared with a chp unit, in that for a condensing steam plant, as it switches to produced heat, then electrical power is lost or becomes unavailable, just as the power used in a heat pump becomes unavailable. Typically for every unit of power lost, then about 6 units of heat are made available at about 90oC. Thus CHP has an effective COP compared to a heat pump of 6[12]. It is noteworthy that the unit for the CHP is lost at the high voltage network and therefore incurs no losses, whereas the heat pump unit is lost at the low voltage part of the network and incurs on average a 6% loss. Because the losses are proportional to the square of the current, during peak periods losses are much higher than this and it is likely that widespread ie city wide application of heat pumps would cause overloading of the distribution and transmission grids unless they are substantially reinforced.

[edit] References
1. ^ The Systems and Equipment volume of the ASHRAE Handbook, ASHRAE, Inc., Atlanta, GA, 2004 2. ^ Thermodynamics an engineering approach Yunus A. engel, Michael A. Boles. Published 1989 by McGraw-Hill in New York, ISBN 007121688X 3. ^ Fundamentals of Engineering Thermodynamics, by Howell and Buckius, McGrawHill, New York. 4. ^ The Ideal Vapor-Compression Cycle 5. ^ Scroll down to "The Basic Vapor Compression Cycle and Components" 6. ^ Twinbird Company. "Welcome to the Dr. Cool's FAQ Room !". Twinbird Company. http://fpsc.twinbird.jp/en/faq_e.html. Retrieved 2011-04-06. 7. ^ Twinbird Company. "About FPSC". Twinbird Company. http://fpsc.twinbird.jp/en/about_fpsc_e.html. Retrieved 2011-04-06. 8. ^ Sunpower Inc.. "[Homepage"]. Sunpower Inc.. http://www.sunpower.com/. Retrieved 2011-04-06. 9. ^ Sunpower Inc.. "Cryocoolers". Sunpower Inc.. http://www.sunpower.com/index.php?pg=79. Retrieved 2011-04-06.

10. ^ Twinbird Company. "Twinbird Company Profile". Twinbird Company. http://www.twinbird.jp/english/profile.html. Retrieved 2011-04-06. 11. ^ Global Cooling NV. "About". Global Cooling NV. http://www.globalcooling.nl/about/. Retrieved 2011-04-06. 12. ^ http://www.sciencedirect.com/science/article/pii/S030142151100379X

Notes

Turns, Stephen (2006). Thermodynamics: Concepts and Applications. Cambridge University Press. pp. 756. ISBN 0521850428. http://books.google.com/books?id=fy5hs04OeMQC&dq=Heat+pump+and+refrig eration+cycle. Dincer, Ibrahim (2003). Refrigeration Systems and Applications. John Wiley and Sons. pp. 598. ISBN 0471623512. Whitman, Bill (2008). Refrigeration and Air conditioning Technology. Delmar.

[edit] External links


The Refrigeration Cycle: How Air Conditioners Work "The Basic Refrigeration Cycle" [hide]

v t e

Thermodynamic cycles
Without phase change (hot air engines)

External combustion cycles

Bell Coleman Brayton/Joule Carnot Ericsson Stirling Stirling (Pseudo / Adiabatic) Stoddard

With phase change


Kalina Rankine (Organic Rankine) Regenerative Atkinson

Internal combustion cycles

Brayton/Joule Diesel Expander Gas-generator Homogeneous Charge Compression Ignition Lenoir Miller Otto Pressure-fed Staged combustion Combined HEHC Mixed/Dual Hampson-Linde Kleemenko Pulse tube Regenerative cooling Transcritical Vapor absorption Vapor-compression Siemens Vuilleumier Barton Humphrey Scuderi

Mixed cycles

Refrigeration cycles

Uncategorized

Retrieved from "http://en.wikipedia.org/w/index.php?title=Heat_pump_and_refrigeration_cycle&oldid=4 85915469" View page ratings

Rate this page


What's this? Trustworthy Objective Complete Well-written

I am highly knowledgeable about this topic (optional) Submit ratings Saved successfully Your ratings have not been submitted yet Categories: Thermodynamic cycles Heat pumps
Personal tools

Log in / create account

Namespaces

Article Talk

Variants Views Actions Search


Search Special:Search

Read Edit View history

Navigation

Main page Contents Featured content Current events Random article Donate to Wikipedia

Interaction

Help About Wikipedia Community portal Recent changes

Toolbox

Contact Wikipedia

What links here Related changes Upload file Special pages Permanent link Cite this page Rate this page

Print/export

Create a book Download as PDF Printable version

Languages

Italiano Portugus This page was last modified on 6 April 2012 at 15:19. Text is available under the Creative Commons Attribution-ShareAlike License; additional terms may apply. See Terms of use for details. Wikipedia is a registered trademark of the Wikimedia Foundation, Inc., a nonprofit organization. Contact us Privacy policy About Wikipedia Disclaimers Mobile view

Our updated Terms of Use will become effective on May 25, 2012. Find out more.

Integrated gasification combined cycle


From Wikipedia, the free encyclopedia

Jump to: navigation, search An integrated gasification combined cycle (IGCC) is a technology that turns coal into gassynthesis gas (syngas). It then removes impurities from the coal gas before it is combusted and attempts to turn any pollutants into re-usable byproducts. This results in lower emissions of sulfur dioxide, particulates, and mercury. Excess heat from the primary combustion and generation is then passed to a steam cycle, similarly to a combined cycle gas turbine. This then also results in improved efficiency compared to conventional pulverized coal.

Contents
[hide]

1 Significance 2 Operations 3 Installations 4 Cost and reliability 5 IGCC Emission Controversy 6 See also 7 References 8 External links

[edit] Significance
Coal can be found in abundance in America and many other countries and its price has remained relatively constant in recent years. Consequently it is used for about 50 percent of U.S. electricity needs.[1] Thus the lower emissions that IGCC technology allows may be important in the future as emission regulations tighten due to growing concern for the impacts of pollutants on the environment and the globe.[1]

[edit] Operations
Below is a schematic flow diagram of an IGCC plant:

Block diagram of IGCC power plant, which utilizes the HRSG The gasification process can produce syngas from high-sulfur coal, heavy petroleum residues and biomass. The plant is called integrated because its syngas is produced in a gasification unit in the plant which has been optimized for the plant's combined cycle. In this example the syngas produced is used as fuel in a gas turbine which produces electrical power. To improve the overall process efficiency heat is recovered from both the gasification process and also the gas turbine exhaust in 'Waste Heat Boilers' producing steam. This steam is then used in steam turbines to produce additional electrical power.

[edit] Installations
In 2007 there were only two IGCC plants generating power in the U.S.;[citation needed] however, several new IGCC plants are expected to come online in the U.S. in the 20122020 time frame. The DOE Clean Coal Demonstration Project helped construct 3 IGCC plants: Wabash River Power Station in West Terre Haute, Indiana, Polk Power Station in Tampa, Florida (online 1996), and Pinon Pine in Reno, Nevada. In the Reno demonstration project, researchers found that then-current IGCC technology would not work more than 300 feet (100m) above sea level.[2] The DOE report in reference 3

however makes no mention of any altitude effect, and most of the problems were associated with the solid waste extraction system. The plant failed.[3] Poland's Kdzierzyn will soon host a Zero-Emission Power & Chemical Plant that combines coal gasification technology with Carbon Capture & Storage (CCS). The supplement of up to 10% biomass in the combustion process will make this plant even more environmentally-friendly. The first generation of IGCC plants polluted less than contemporary coal-based technology, but also polluted water; for example, the Wabash River Plant was out of compliance with its water permit during 19982001[4] because it emitted arsenic, selenium and cyanide. The Wabash River Generating Station is now wholly owned and operated by the Wabash River Power Association. IGCC is now touted as capture ready and could potentially capture and store carbon dioxide.[5] (See FutureGen) There are several advantages and disadvantages when compared to conventional post combustion carbon capture and various variations and these are fully discussed at reference 6.[6]

[edit] Cost and reliability


The main problem for IGCC is its extremely high capital cost, upwards of $3,593/kW.[7] Official US government figures give more optimistic estimates [8] of $1,491/kW installed capacity (2005 dollars) v. $1,290 for a conventional clean coal facility, but in light of current applications, these cost estimates have been demonstrated to be incorrect.[citation
needed]

Outdated per megawatt-hour cost of an IGCC plant vs a pulverized coal plant coming online in 2010 would be $56 vs $52, and it is claimed that IGCC becomes even more attractive when you include the costs of carbon capture and sequestration, IGCC becoming $79 per megawatt-hour vs. $95 per megawatt-hour for pulverized coal.[9] Recent testimony in regulatory proceedings show the cost of IGCC to be twice that predicted by Goddell, from $96 to 104/MWhr.[10][11] That's before addition of carbon capture and sequestration (sequestration has been a mature technology at both Weyburn in Canada (for enhanced oil recovery) and Sleipner in the North Sea at a commercial scale for the past ten years)capture at a 90% rate is expected to have a $30/MWh additional cost.[12] Wabash River was down repeatedly for long stretches due to gasifier problems. The gasifier problems have not been remediedsubsequent projects, such as Excelsior's Mesaba Project, have a third gasifier and train built in. However, the past year has seen Wabash River running reliably, with availability comparable to or better than other technologies.

The Polk County IGCC has design problems. First, the project was initially shut down because of corrosion in the slurry pipeline that fed slurried coal from the rail cars into the gasifier. A new coating for the pipe was developed. Second, the thermocoupler was replaced in less than two years; an indication that the gasifier had problems with a variety of feedstocks; from bituminous to sub-bituminous coal. The gasifier was designed to also handle lower rank lignites. Third, unplanned down time on the gasifier because of refractory liner problems, and those problems were expensive to repair. The gasifier was originally designed in Italy to be half the size of what was built at Polk. Newer ceramic materials may assist in improving gasifier performance and longevity. Understanding the operating problems of the current IGCC plant is necessary to improve the design for the IGCC plant of the future. (Polk IGCC Power Plant, http://www.cleanenergy.us/projects/polk_florida.html.) Keim, K., 2009, IGCC A Project on Sustainability Management Systmes for Plant Re-Design and Re-Image. This is an unpublished paper from Harvard University) General Electric is currently designing an IGCC model plant that should introduce greater reliability. GE's model features advanced turbines optimized for the coal syngas. Eastman's industrial gasification plant in Kingsport, TN uses a GE Energy solid-fed gasifier. Eastman, a fortune 500 company, built the facility in 1983 without any state or federal subsidies and turns a profit.[13][14] There are several refinery-based IGCC plants in Europe that have demonstrated good availability (90-95%) after initial shakedown periods. Several factors help this performance: 1. None of these facilities use advanced technology (F type) gas turbines. 2. All refinery-based plants use refinery residues, rather than coal, as the feedstock. This eliminates coal handling and coal preparation equipment and its problems. Also, there is a much lower level of ash produced in the gasifier, which reduces cleanup and downtime in its gas cooling and cleaning stages. 3. These non-utility plants have recognized the need to treat the gasification system as an up-front chemical processing plant, and have reorganized their operating staff accordingly. Another IGCC success story has been the 250 MW Buggenum plant in The Netherlands. It also has good availability. This coal-based IGCC plant currently uses about 30% biomass as a supplemental feedstock. The owner, NUON, is paid an incentive fee by the government to use the biomass. NUON is constructing a 1,300 MW IGCC plant in the Netherlands. The Nuon Magnum IGCC power plant will be commissioned in 2011. Mitsubishi Heavy Industries has been awarded to construct the power plant.[15] A new generation of IGCC-based coal-fired power plants has been proposed, although none is yet under construction. Projects are being developed by AEP, Duke Energy, and Southern Company in the US, and in Europe by ZAK/PKE, Centrica (UK), E.ON and RWE (both Germany) and NUON (Netherlands). In Minnesota, the state's Dept. of Commerce analysis found IGCC to have the highest cost, with an emissions profile not

significantly better than pulverized coal. In Delaware, the Delmarva and state consultant analysis had essentially the same results. The high cost of IGCC is the biggest obstacle to its integration in the power market; however, most energy executives recognize that carbon regulation is coming soon. Bills requiring carbon reduction are being proposed again both the House and the Senate, and with the Democratic majority it seems likely that with the next President there will be a greater push for carbon regulation. The Supreme Court decision requiring the EPA to regulate carbon (Commonwealth of Massachusetts et al. v. Environmental Protection Agency et al.)[16] also speaks to the likelihood of future carbon regulations coming sooner, rather than later. With carbon capture, the cost of electricity from an IGCC plant would increase approximately 30%. For a natural gas CC, the increase is approximately 33%. For a pulverized coal plant, the increase is approximately 68%. This potential for less expensive carbon capture makes IGCC an attractive choice for keeping low cost coal an available fuel source in a carbon constrained world. In Japan, electric power companies, in conjunction with Mitsubishi Heavy Industries has been operating a 200 t/d IGCC pilot plant since the early '90s. In September 2007, they started up a 250 MW demo plant in Nakoso. It runs on air-blown (not oxygen) dry feed coal only. It burns PRB coal with an unburned carbon content ratio of <0.1% and no detected leaching of trace elements. It employs not only F type turbines but G type as well. (see gasification.org link below) Next generation IGCC plants with CO2 capture technology will be expected to have higher thermal efficiency and to hold the cost down because of simplified systems compared to conventional IGCC. The main feature is that instead of using oxygen and nitrogen to gasify coal, they use oxygen and CO2. The main advantage is that it is possible to improve the performance of cold gas efficiency and to reduce the unburned carbon (char). With a 1300 degrees C class gas turbine it is possible to achieve 42% net thermal efficiency, rising to 45% with a 1500 degree class gas turbine, with CO2 capture. In case of conventional IGCC systems, it is only possible to achieve just over 30% efficiency with a 1300 degree gas turbine.[citation needed] The CO2 extracted from gas turbine exhaust gas is utilized in this system. Using a closed gas turbine system capable of capturing the CO2 by direct compression and liquefication obviates the need for a separation and capture system.[17]

[edit] IGCC Emission Controversy


In 2007, the New York State Attorney General's office demanded full disclosure of "financial risks from greenhouse gases" to the shareholders of electric power companies proposing the development of IGCC coal-fired power plants. "Any one of the several new or likely regulatory initiatives for CO2 emissions from power plants - including state carbon controls, EPA's regulations under the Clean Air Act, or the enactment of federal

global warming legislation - would add a significant cost to carbon-intensive coal generation";[18] U.S. Senator Hillary Clinton from New York has proposed that this full risk disclosure be required of all publicly-traded power companies nationwide.[19] This honest disclosure has begun to reduce investor interest in all types of existing-technology coal-fired power plant development, including IGCC. Senator Harry Reid (Majority Leader of the 2007/2008 U.S. Senate) told the 2007 Clean Energy Summit that he will do everything he can to stop construction of proposed new IGCC coal-fired electric power plants in Nevada. Reid wants Nevada utility companies to invest in solar energy, wind energy and geothermal energy instead of coal technologies. Reid stated that global warming is a reality, and just one proposed coal-fired plant would contribute to it by burning seven million tons of coal a year. The long-term healthcare costs would be far too high, he claimed (no source attributed). "I'm going to do everything I can to stop these plants.", he said. "There is no clean coal technology. There is cleaner coal technology, but there is no clean coal technology."[20] One of the most efficient ways to treat the H2S gas from a IGCC plant, is by converting it into sulphuric acid in a wet gas sulphuric acid process wsa process However, the majority of the H2S treating plants utilize the modified Claus process, as the sulphur market infrastructure and the transportation costs of sulphuric acid versus sulphur are in favour of sulphur production.

[edit] See also


Relative cost of electricity generated by different sources Environmental impact of the coal industry

[edit] References
1. ^ a b Schon, Samuel C., and Arthur A. Small III. "Climate change and the potential of coal gasification." Geotimes 51.9 (Sept 2006): 20(4). Expanded Academic ASAP. Gale. University of Washington. 28 Oct. 2008 |date=October 29, 2008 2. ^ Source: Joe Lucas, Executive Director of Americans for Balanced Energy Choices, as interviewed on NPR's Science Friday, Friday May 12, 2006 3. ^ Information Bridge: DOE Scientific and Technical Information - Sponsored by OSTI 4. ^ Wabash River Energy Ltd. (August 2000). "Wabash River Coal Gasification Repowering Project Final Technical Report" (PDF). Work performed under Cooperative Agreement DE-FC21-92MC29310. The U.S. Department of Energy / Office of Fossil Energy / National Energy Technology Laboratory / Morgantown, West Virginia. http://www.osti.gov/bridge/servlets/purl/787567-a64JvB/native/787567.pdf. Retrieved 2008-06-30. "As a result, process waste water arising from use of the current feedstock, remains out of permit compliance due to elevated levels of arsenic, selenium and cyanide. To rectify these concerns, plant personnel have been working on several potential equipment modifications and treatment alternatives to bring the discharge back into compliance. Wabash River is currently obligated to resolve this issue by September 2001. [p. ES-6] Elevated levels of selenium, cyanide and arsenic in the waste water have caused the process waste water to be out of permit compliance. Daily maximum values, though

5. 6. 7. 8. 9. 10. 11. 12.

13. 14. 15. 16. 17. 18. 19. 20.

not indicated in the table above, were routinely exceeded for selenium and cyanide, and only occasionally for arsenic. [p. 6-14, Table 6.1L]" ^ http://www.gepower.com/prod_serv/products/gasification/en/app_power.htm ^ http://www.claverton-energy.com/integrated-gasification-combined-cycle-for-carboncapture-storage.html Claverton Energy Group conference 24th October Bath. ^ Excelsior's Mesaba Project ^ http://www.eia.doe.gov/oiaf/aeo/assumption/pdf/electricity.pdf#page=3 ^ Goodell, Jeff. "Big Coal." pg. 214. New York, Houghton Mifflin. 2006 ^ Testimony of Dr. Elion Amit, Minnesota Dept. of Commerce. ^ http://www.mncoalgasplant.com/puc/05-1993%20pub%20rebuttal.pdf. ^ http://www.newenergymatters.com/?p=list&t=insightnote&sp=1&f=sectors,ccs/publish, Yes/&sort=2&id=1766 ^ Goodell, Jeff. "Big Coal." New York, Houghton Mifflin. 2006 ^ Eastman - Eastman Chemical Company - Home Page ^ http://www.nuon.com/about-nuon/Innovative-projects/magnum.jsp ^ Massachusetts, et al. v. Environmental Protection Agency, 05-1120 - FindLaw US Supreme Court Center ^ Inumaru,Jun - senior research scientist, Central Research Institute of Electric Power Industry (CRIEPI)(Japan) G8 Energy Ministerial Meeting Symposium, Nikkei Weekly. ^ http://www.marketwire.com/mw/rel_us_print.jsp?id=776699 ^ http://www.hillaryclinton.com/files/pdf/poweringamericasfuture.pdf ^ http://publicutilities.utah.gov/news/cleanenergysummitreidopposescoal.pdf

[edit] External links


Hunstown: Ireland's most efficient power plant @ Siemens Power Generation website Natural Gas Combined-cycle Gas Turbine Power Plants Northwest Power Planning Council, New Resource Characterization for the Fifth Power Plan, August 2002 Combined cycle solar power

Retrieved from "http://en.wikipedia.org/w/index.php?title=Integrated_gasification_combined_cycle&oldi d=473563693" View page ratings

Rate this page


What's this? Trustworthy Objective Complete Well-written

I am highly knowledgeable about this topic (optional) Submit ratings Saved successfully Your ratings have not been submitted yet Categories: Thermodynamic cycles Chemical engineering Power station technology Energy conversion Hidden categories: All articles with unsourced statements Articles with unsourced statements from September 2008 Articles with unsourced statements from March 2011 Articles with unsourced statements from November 2008
Personal tools

Log in / create account

Namespaces

Article Talk

Variants Views Actions Search


Search Special:Search

Read Edit View history

Navigation

Main page Contents Featured content Current events

Random article Donate to Wikipedia

Interaction Toolbox

Help About Wikipedia Community portal Recent changes Contact Wikipedia

What links here Related changes Upload file Special pages Permanent link Cite this page Rate this page

Print/export

Create a book Download as PDF Printable version

Languages

Deutsch Espaol Franais Polski This page was last modified on 27 January 2012 at 19:37. Text is available under the Creative Commons Attribution-ShareAlike License; additional terms may apply. See Terms of use for details. Wikipedia is a registered trademark of the Wikimedia Foundation, Inc., a nonprofit organization. Contact us Privacy policy About Wikipedia Disclaimers Mobile view

Our updated Terms of Use will become effective on May 25, 2012. Find out more.

Miller cycle
From Wikipedia, the free encyclopedia

Jump to: navigation, search This article does not cite any references or sources. Please help improve this article by adding citations to reliable sources. Unsourced material may be challenged and removed. (November 2010)

Thermodynamics

Branches[show]

Laws[show]

Systems[show]

System properties[show]

Material properties[show]

Specific heat capacity Compressibility Thermal expansion

Equations[show]

Potentials[show]

Internal energy Enthalpy Helmholtz free energy Gibbs free energy


History and culture[show]

Scientists[show]

v t e

In engineering, the Miller cycle is a combustion process used in a type of four-stroke internal combustion engine. The Miller cycle was patented by Ralph Miller, an American engineer, in the 1940s. This type of engine was first used in ships and stationary power-generating plants, and is now used for some railway locomotives such as the GE PowerHaul. It was adapted by Mazda for their KJ-ZEM V6, used in the Millenia sedan, and in their Eunos 800 sedan (Australia) luxury cars. More recently, Subaru has combined a Miller cycle flat-4 with a hybrid driveline for their concept "Turbo Parallel Hybrid" car, known as the Subaru B5TPH.

[edit] Overview

A traditional Otto cycle engine uses four "strokes", of which two can be considered "high power" the compression stroke (high power consumption) and power stroke (high power production). Much of the internal power loss of an engine is due to the energy needed to compress the charge during the compression stroke, so systems that reduce this power consumption can lead to greater efficiency. In the Miller cycle, the intake valve is left open longer than it would be in an Otto cycle engine. In effect, the compression stroke is two discrete cycles: the initial portion when the intake valve is open and final portion when the intake valve is closed. This two-stage intake stroke creates the so called "fifth" stroke that the Miller cycle introduces. As the piston initially moves upwards in what is traditionally the compression stroke, the charge is partially expelled back out the still-open intake valve. Typically this loss of charge air would result in a loss of power. However, in the Miller cycle, this is compensated for by the use of a supercharger. The supercharger typically will need to be of the positive displacement (Roots or Screw) type due to its ability to produce boost at relatively low engine speeds. Otherwise, low-rpm torque will suffer. A key aspect of the Miller cycle is that the compression stroke actually starts only after the piston has pushed out this "extra" charge and the intake valve closes. This happens at around 20% to 30% into the compression stroke. In other words, the actual compression occurs in the latter 70% to 80% of the compression stroke. In a typical spark ignition engine, the Miller cycle yields an additional benefit. The intake air is first compressed by the supercharger and then cooled by an intercooler. This lower intake charge temperature, combined with the lower compression of the intake stroke, yields a lower final charge temperature than would be obtained by simply increasing the compression of the piston. This allows ignition timing to be advanced beyond what is normally allowed before the onset of detonation, thus increasing the overall efficiency still further. An additional advantage of the lower final charge temperature is that the emission of NOx in diesel engines is decreased, which is an important design parameter in large diesel engines on board ships and power plants. Efficiency is increased by raising the compression ratio. In a typical gasoline engine, the compression ratio is limited due to self-ignition (detonation) of the compressed, and therefore hot, air/fuel mixture. Due to the reduced compression stroke of a Miller cycle engine, a higher overall cylinder pressure (supercharger pressure plus mechanical compression) is possible, and therefore a Miller cycle engine has better efficiency. The benefits of utilizing positive displacement superchargers come with a cost. 15% to 20% of the power generated by a supercharged engine is usually required to do the work of driving the supercharger, which compresses the intake charge (also known as boost). A similar delayed-valve closing method is used in some modern versions of Atkinson cycle engines, but without the supercharging. These engines are generally found on

hybrid electric vehicles, where efficiency is the goal, and the power lost compared to the Miller cycle is made up through the use of electric motors.

[edit] Source

Miller cycle [hide]


v t e

Thermodynamic cycles
Without phase change (hot air engines)

External combustion cycles

Bell Coleman Brayton/Joule Carnot Ericsson Stirling Stirling (Pseudo / Adiabatic) Stoddard

With phase change


Kalina Rankine (Organic Rankine) Regenerative Atkinson Brayton/Joule Diesel Expander Gas-generator Homogeneous Charge Compression Ignition Lenoir Miller Otto Pressure-fed Staged combustion Combined HEHC

Internal combustion cycles

Mixed cycles

Mixed/Dual Hampson-Linde Kleemenko Pulse tube Regenerative cooling Transcritical Vapor absorption Vapor-compression Siemens Vuilleumier Barton Humphrey Scuderi

Refrigeration cycles

Uncategorized

Retrieved from "http://en.wikipedia.org/w/index.php?title=Miller_cycle&oldid=484871857" View page ratings

Rate this page


What's this? Trustworthy Objective Complete Well-written I am highly knowledgeable about this topic (optional) Submit ratings Saved successfully Your ratings have not been submitted yet Categories: Thermodynamic cycles Hidden categories: Articles lacking sources from November 2010 All articles lacking sources
Personal tools

Log in / create account

Namespaces

Article Talk

Variants Views Actions Search


Search Special:Search

Read Edit View history

Navigation

Main page Contents Featured content Current events Random article Donate to Wikipedia

Interaction Toolbox

Help About Wikipedia Community portal Recent changes Contact Wikipedia

What links here Related changes Upload file Special pages Permanent link Cite this page Rate this page

Print/export

Create a book Download as PDF Printable version

Languages

Catal Deutsch Espaol Franais Galego Italiano Magyar Polski Portugus Slovenina Trke This page was last modified on 31 March 2012 at 17:23. Text is available under the Creative Commons Attribution-ShareAlike License; additional terms may apply. See Terms of use for details. Wikipedia is a registered trademark of the Wikimedia Foundation, Inc., a nonprofit organization. Contact us Privacy policy About Wikipedia Disclaimers Mobile view

Our updated Terms of Use will become effective on May 25, 2012. Find out more.

Otto cycle
From Wikipedia, the free encyclopedia

Jump to: navigation, search See also: Otto engine and Four-stroke engine

Thermodynamics

Branches[show]

Laws[show]

Systems[show]

System properties[show]

Material properties[show]

Specific heat capacity Compressibility Thermal expansion

Equations[show]

Potentials[show]

Internal energy Enthalpy Helmholtz free energy Gibbs free energy


History and culture[show]

Scientists[show]

v t e

An Otto cycle is an idealized thermodynamic cycle which describes the functioning of a typical spark ignition reciprocating piston engine,[1] the thermodynamic cycle most commonly found in automobile engines.[clarification needed]

Pressure-Volume diagram

Temperature-Entropy diagram The idealized diagrams of a four-stroke Otto cycle Both diagrams: the intake (A) stroke is performed by an isobaric expansion, followed by an adiabatic compression (B) stroke. Through the combustion of fuel, heat is added in an isochoric process, followed by an adiabatic expansion process, characterizing the power (C) stroke. The cycle is closed by the exhaust (D) stroke, characterized by isochoric cooling and isobaric compression processes. The Otto cycle is constructed out of: TOP and BOTTOM of the loop: a pair of quasi-parallel adiabatic processes LEFT and RIGHT sides of the loop: a pair of parallel isochoric processes

The adiabatic processes are impermeable to heat: heat flows into the loop through the left pressurizing process and some of it flows back out through the right depressurizing process, and the heat which remains does the work. The processes are described by:[2]

Process 1-2 is an isentropic compression of the air as the piston moves from bottom dead centre (BDC) to top dead centre (TDC). Process 2-3 is a constant-volume heat transfer to the air from an external source while the piston is at top dead centre. This process is intended to represent the ignition of the fuel-air mixture and the subsequent rapid burning. Process 3-4 is an isentropic expansion (power stroke). Process 4-1 completes the cycle by a constant-volume process in which heat is rejected from the air while the piston is a bottom dead centre.

The Otto cycle consists of adiabatic compression, heat addition at constant volume, adiabatic expansion, and rejection of heat at constant volume. In the case of a four-stroke Otto cycle, technically there are two additional processes: one for the exhaust of waste heat and combustion products (by isobaric compression), and one for the intake of cool oxygen-rich air (by isobaric expansion); however, these are often omitted in a simplified analysis. Even though these two processes are critical to the functioning of a real engine, wherein the details of heat transfer and combustion chemistry are relevant, for the simplified analysis of the thermodynamic cycle, it is simpler and more convenient to assume that all of the waste-heat is removed during a single volume change. A P-V animation of the Otto cycle is very useful in the analysis of the entire process.[3]

Contents
[hide]

1 History 2 Processes o 2.1 Process 1-2 (B on diagrams) o 2.2 Process 2-3 (C on diagrams) o 2.3 Process 3-4 (D on diagrams) o 2.4 Process 4-1 (A on diagrams) o 2.5 Exhaust and intake strokes 3 Cycle Analysis 4 References

[edit] History

The four-stroke engine was first patented by Alphonse Beau de Rochas in 1861. Before, in about 185457, two Italians (Eugenio Barsanti and Felice Matteucci) invented an engine that was rumored to be very similar, but the patent was lost.
"The request bears the no. 700 of Volume VII of the Patent Office of the Reign of Piedmont. We do not have the text of the patent request, only a photo of the table which contains a drawing of the engine. We do not even know if it was a new patent or an extension of the patent granted three days earlier, on December 30, 1857, at Turin." f. Eugenio Barsanti and Felice Matteucci, June 4, 1853 [4]

The first person to build a working four stroke engine, a stationary engine using a coal gas-air mixture for fuel (a gas engine), was German engineer Nicolaus Otto.[5] This is why the four-stroke principle today is commonly known as the Otto cycle and four-stroke engines using spark plugs often are called Otto engines.

[edit] Processes
[edit] Process 1-2 (B on diagrams)
Piston moves from crank end (bottom dead centre) to cover end (top dead centre) and an ideal gas with initial state 1 is compressed isentropically to state point 2, through compression ratio . Mechanically this is the adiabatic compression of the air/fuel mixture in the cylinder, also known as the compression stroke. Generally the compression ratio is around 9-10:1 (V1:V2) for a typical engine.[6]

[edit] Process 2-3 (C on diagrams)


The piston is momentarily at rest at TDC and heat is added to the working fluid at constant volume from an external heat source which is brought into contact with the cylinder head. The pressure rises and the ratio is called the "explosion ratio". At this instant the air/fuel mixture is compressed at the top of the compression stroke with the volume essentially held constant, also known as ignition phase.

[edit] Process 3-4 (D on diagrams)


The increased high pressure exerts a greater amount of force on the piston and pushes it towards the BDC. Expansion of working fluid takes place isentropically and work is done by the system. The volume ratio is called "isentropic expansion ratio". Mechanically this is the adiabatic expansion of the hot gaseous mixture in the cylinder head, also known as expansion (power) stroke.

[edit] Process 4-1 (A on diagrams)

The piston is momentarily at rest at BDC and heat is rejected to the external sink by bringing it in contact with the cylinder head. The process is so controlled that ultimately the working fluid comes to its initial state 1 and the cycle is completed.

[edit] Exhaust and intake strokes


Exhaust stroke-ejection of the gaseous mixture via an exhaust valve through the cylinder head. Induction stroke-intake of the next air charge into the cylinder. The volume of the exhaust gasses is the same as the air charge.[6]

[edit] Cycle Analysis


Processes 1-2 and 3-4 do work on the system but no heat transfer occurs during adiabatic expansion and compression. Processes 2-3 and 4-1 are isochoric; therefore, heat transfer occurs but no work is done. No work is done during a isochoric (constant volume) because work requires movement; when the piston volume does not change no shaft work is produced by the system. Four different equations can be derived by neglecting kinetic and potential energy and considering the first law of thermodynamics (energy conservation). Assuming these conditions the first law is rewritten as:[2]

Applying this to the Otto cycle the four process equations can be derived:

Since the first law is expressed as heat added to the system and work expelled from the system then ( ) and ( ) will always produce positive values. However, since work always involves movement, processes 2-3 and 4-1 will be omitted because they occur at a constant volume. The net work can be expressed as:

The net work can also be found by evaluating the heat added minus the heat leaving or expelled.

Thermal efficiency is the quotient of the net work to the heat addition into system. Upon rearrangement the thermal efficiency can be obtained (Net Work/Heat added): Equation 1:

Alternatively, thermal efficiency can be derived by strictly heat added and heat rejected.

In the Otto cycle, there is no heat transfer during the process 1-2 and 3-4 as they are reversible adiabatic processes. Heat is supplied only during the constant volume processes 2-3 and heat is rejected only during the constant volume processes 4-1.[7] Equation 1 can now be related to the specific heat equation for constant volume. The specific heats are particularly useful for thermodynamic calculations involving the ideal gas model.

Rearranging yields:

Inserting the specific heat equation into the thermal efficiency equation (Equation 1) yields.

Upon rearrangement:

Next, noting from the diagrams The equation then reduces to: Equation 2:

, thus both of these can be omitted.

Since the Otto cycle is an isentropic process the isentropic equations of ideal gases and the constant pressure/volume relations can be used to yield Equations 3 & 4. Equation 3:

Equation 4:

imagine The derivation of the previous equations are found by solving these four equations respectively (where is the gas constant):

Further simplifying Equation 4, where is the compression ratio Equation 5:

Also, note that

where is the specific heat ratio From inverting Equation 4 and inserting it into Equation 2 the final thermal efficiency can be expressed as:[7] Equation 6:

From analyzing equation 6 it is evident that the Otto cycle depends directly upon the compression ratio . Since the for air is 1.4, an increase in will produce an increase in . However, the for the combustion products of the fuel/air mixture is taken at approximately 1.3. The foregoing discussion implies that it is more efficient to have a high compression ratio. The standard ratio is approximately 10:1 for typical automobiles. Usually this does not increase much because of the possibility of autoignition, or "knock", which places an upper limit on the compression ratio.[2] During the compression process 1-2 the temperature rises, therefore an increase in the compression ratio causes an increase in temperature. Autoignition occurs when the temperature of the fuel/air mixture becomes too high before it is ignited by the flame front. The compression stroke is intended to compress the products before the flame ignites the mixture. Therefore if the compression ratio was increased, the mixture could be compressed before ignition leading to "engine knocking". This can damage engine components and will decrease the original horsepower of the engine.

[edit] References
1. ^ Wu, Chih. Thermodynamic Cycles: Computer-aided Design and Optimization. New York: M. Dekker, 2004. Print. 2. ^ a b c Moran, Michael J., and Howard N. Shapiro. Fundamentals of Engineering Thermodynamics. 6th ed. Hoboken, N.J. : Chichester: Wiley ; John Wiley, 2008. Print.

3. ^ "Animated Diagram". Leipzig. 2006. http://techni.tachemie.unileipzig.de/otto/index_e.html. Retrieved 2010-09-22. 4. ^ "Documenti Storici". Barsantiematteucci.it. http://www.barsantiematteucci.it/inglese/documentiStorici.html. Retrieved 2010-09-22. 5. ^ Gunston, Bill (1999). Development of Piston Aero Engines (2 ed.). Sparkford, UK: Patrick Stephens Ltd. p. 21. ISBN 0-7509-4478-1. 6. ^ a b "Heat Cycles - Electropeaedia". Woodbank Communications Ltd. http://www.mpoweruk.com/heat_engines.htm. Retrieved 2011-04-11. 7. ^ a b Gupta, H. N. Fundamentals of Internal Combustion. New Delhi: Prentice-Hall, 2006. Print.

Retrieved from "http://en.wikipedia.org/w/index.php?title=Otto_cycle&oldid=488628678" View page ratings

Rate this page


What's this? Trustworthy Objective Complete Well-written I am highly knowledgeable about this topic (optional) Submit ratings Saved successfully Your ratings have not been submitted yet Categories: Piston engines Thermodynamic cycles Hidden categories: Wikipedia articles needing clarification from December 2011
Personal tools

Log in / create account

Namespaces

Article Talk

Variants

Views Actions Search


Search Special:Search

Read Edit View history

Navigation

Main page Contents Featured content Current events Random article Donate to Wikipedia

Interaction Toolbox

Help About Wikipedia Community portal Recent changes Contact Wikipedia

What links here Related changes Upload file Special pages Permanent link Cite this page Rate this page

Print/export

Create a book Download as PDF Printable version

Languages

Catal Dansk Deutsch Espaol Franais Galego Italiano Magyar Nederlands Norsk (nynorsk) Portugus This page was last modified on 22 April 2012 at 10:14. Text is available under the Creative Commons Attribution-ShareAlike License; additional terms may apply. See Terms of use for details. Wikipedia is a registered trademark of the Wikimedia Foundation, Inc., a nonprofit organization. Contact us Privacy policy About Wikipedia Disclaimers Mobile view

Our updated Terms of Use will become effective on May 25, 2012. Find out more.

Rankine cycle
From Wikipedia, the free encyclopedia

Jump to: navigation, search

Thermodynamics

Branches[show]

Laws[show]

Systems[show]

System properties[show]

Material properties[show]

Specific heat capacity Compressibility Thermal expansion

Equations[show]

Potentials[show]

Internal energy Enthalpy Helmholtz free energy Gibbs free energy


History and culture[show]

Scientists[show]

v t e

The Rankine cycle is a cycle that converts heat into work. The heat is supplied externally to a closed loop, which usually uses water. This cycle generates about 90% of all electric power used throughout the world,[1] including virtually all solar thermal, biomass, coal and nuclear power plants. It is named after William John Macquorn Rankine, a Scottish polymath and Glasgow University professor. The Rankine cycle is the fundamental thermodynamic underpinning of the steam engine.

Contents
[hide]

1 Description 2 The four processes in the Rankine cycle 3 Variables 4 Equations 5 Real Rankine cycle (non-ideal) 6 Variations of the basic Rankine cycle

6.1 Rankine cycle with reheat 6.2 Regenerative Rankine cycle 7 Organic Rankine cycle 8 References

o o

[edit] Description

Physical layout of the four main devices used in the Rankine cycle The Rankine cycle most closely describes the process by which steam-operated heat engines most commonly found in power generation plants generate power. The two most common heating processes used in these power plants are nuclear fission and the combustion of fossil fuels such as coal, natural gas, and oil. The Rankine cycle is sometimes referred to as a practical Carnot cycle because, when an efficient turbine is used, the TS diagram begins to resemble the Carnot cycle. The main difference is that heat addition (in the boiler) and rejection (in the condenser) are isobaric in the Rankine cycle and isothermal in the theoretical Carnot cycle. A pump is used to pressurize the working fluid received from the condenser as a liquid instead of as a gas. All of the energy in pumping the working fluid through the complete cycle is lost, as is most of the energy of vaporization of the working fluid in the boiler. This energy is lost to the cycle because the condensation that can take place in the turbine is limited to about 10% in order to minimize blade erosion; the vaporization energy is rejected from the cycle through the condenser. But pumping the working fluid through the cycle as a liquid requires a very small fraction of the energy needed to transport it as compared to compressing the working fluid as a gas in a compressor (as in the Carnot cycle). The efficiency of a Rankine cycle is usually limited by the working fluid. Without the pressure reaching super critical levels for the working fluid, the temperature range the cycle can operate over is quite small: turbine entry temperatures are typically 565C (the creep limit of stainless steel) and condenser temperatures are around 30C. This gives a theoretical Carnot efficiency of about 63% compared with an actual efficiency of 42% for

a modern coal-fired power station. This low turbine entry temperature (compared with a gas turbine) is why the Rankine cycle is often used as a bottoming cycle in combinedcycle gas turbine power stations. The working fluid in a Rankine cycle follows a closed loop and is reused constantly. The water vapor with entrained droplets often seen billowing from power stations is generated by the cooling systems (not from the closed-loop Rankine power cycle) and represents the waste heat energy (pumping and condensing) that could not be converted to useful work in the turbine. Note that cooling towers operate using the latent heat of vaporization of the cooling fluid. While many substances could be used in the Rankine cycle, water is usually the fluid of choice due to its favorable properties, such as nontoxic and nonreactive chemistry, abundance, and low cost, as well as its thermodynamic properties. One of the principal advantages the Rankine cycle holds over others is that during the compression stage relatively little work is required to drive the pump, the working fluid being in its liquid phase at this point. By condensing the fluid, the work required by the pump consumes only 1% to 3% of the turbine power and contributes to a much higher efficiency for a real cycle. The benefit of this is lost somewhat due to the lower heat addition temperature. Gas turbines, for instance, have turbine entry temperatures approaching 1500C. Nonetheless, the efficiencies of actual large steam cycles and large modern gas turbines are fairly well matched.

[edit] The four processes in the Rankine cycle

Ts diagram of a typical Rankine cycle operating between pressures of 0.06bar and 50bar

There are four processes in the Rankine cycle. These states are identified by numbers (in brown) in the above Ts diagram.

Process 1-2: The working fluid is pumped from low to high pressure. As the fluid is a liquid at this stage the pump requires little input energy. Process 2-3: The high pressure liquid enters a boiler where it is heated at constant pressure by an external heat source to become a dry saturated vapor. The input energy required can be easily calculated using mollier diagram or h-s chart or enthalpy-entropy chart also known as steam tables. Process 3-4: The dry saturated vapor expands through a turbine, generating power. This decreases the temperature and pressure of the vapor, and some condensation may occur. The output in this process can be easily calculated using the Enthalpy-entropy chart or the steam tables. Process 4-1: The wet vapor then enters a condenser where it is condensed at a constant temperature to become a saturated liquid.

In an ideal Rankine cycle the pump and turbine would be [[isentropic], i.e., the pump and turbine would generate no entropy and hence maximize the net work output. Processes 12 and 3-4 would be represented by vertical lines on the T-S diagram and more closely resemble that of the Carnot cycle. The Rankine cycle shown here prevents the vapor ending up in the superheat region after the expansion in the turbine, [1] which reduces the energy removed by the condensers.

[edit] Variables
Heat flow rate to or from the system (energy per unit time) Mass flow rate (mass per unit time) Mechanical power consumed by or provided to the system (energy per unit time) Thermodynamic efficiency of the process (net power output per heat input, dimensionless) Isentropic efficiency of the compression (feed pump) and expansion (turbine) processes, dimensionless The "specific enthalpies" at indicated points on the T-S diagram The final "specific enthalpy" of the fluid if the turbine were isentropic The pressures before and after the compression process

[edit] Equations
In general, the efficiency of a simple Rankine cycle can be defined as:

Each of the next four equations[1] is easily derived from the energy and mass balance for a control volume. defines the thermodynamic efficiency of the cycle as the ratio of net power output to heat input. As the work required by the pump is often around 1% of the turbine work output, it can be simplified.

When dealing with the efficiencies of the turbines and pumps, an adjustment to the work terms must be made. = - p/pump ( = - ( - )*turbine turbine/
pump/

)/pump

[edit] Real Rankine cycle (non-ideal)

Rankine cycle with superheat

In a real Rankine cycle, the compression by the pump and the expansion in the turbine are not isentropic. In other words, these processes are non-reversible and entropy is increased during the two processes. This somewhat increases the power required by the pump and decreases the power generated by the turbine. In particular the efficiency of the steam turbine will be limited by water droplet formation. As the water condenses, water droplets hit the turbine blades at high speed causing pitting and erosion, gradually decreasing the life of turbine blades and efficiency of the turbine. The easiest way to overcome this problem is by superheating the steam. On the Ts diagram above, state 3 is above a two phase region of steam and water so after expansion the steam will be very wet. By superheating, state 3 will move to the right of the diagram and hence produce a drier steam after expansion.

[edit] Variations of the basic Rankine cycle

Rankine cycle with reheat The overall thermodynamic efficiency (of almost any cycle) can be increased by raising

the average heat input temperature of that cycle. Increasing the temperature of the steam into the superheat region is a simple way of doing this. There are also variations of the basic Rankine cycle which are designed to raise the thermal efficiency of the cycle in this way; two of these are described below.

[edit] Rankine cycle with reheat


In this variation, two turbines work in series. The first accepts vapor from the boiler at high pressure. After the vapor has passed through the first turbine, it re-enters the boiler

and is reheated before passing through a second, lower pressure turbine. Among other advantages, this prevents the vapor from condensing during its expansion which can seriously damage the turbine blades, and improves the efficiency of the cycle, as more of the heat flow into the cycle occurs at higher temperature.

[edit] Regenerative Rankine cycle


The regenerative Rankine cycle is so named because after emerging from the condenser (possibly as a subcooled liquid) the working fluid is heated by steam tapped from the hot portion of the cycle. On the diagram shown, the fluid at 2 is mixed with the fluid at 4 (both at the same pressure) to end up with the saturated liquid at 7. This is called "direct contact heating". The Regenerative Rankine cycle (with minor variants) is commonly used in real power stations. Another variation is where bleed steam from between turbine stages is sent to feedwater heaters to preheat the water on its way from the condenser to the boiler. These heaters do not mix the input steam and condensate, function as an ordinary tubular heat exchanger, and are named "closed feedwater heaters". The regenerative features here effectively raise the nominal cycle heat input temperature, by reducing the addition of heat from the boiler/fuel source at the relatively low feedwater temperatures that would exist without regenerative feedwater heating. This improves the efficiency of the cycle, as more of the heat flow into the cycle occurs at higher temperature.

[edit] Organic Rankine cycle


Main article: Organic Rankine cycle The organic Rankine cycle (ORC) uses an organic fluid such as n-pentane[2] or toluene[3] in place of water and steam. This allows use of lower-temperature heat sources, such as solar ponds, which typically operate at around 7090 C.[4] The efficiency of the cycle is much lower as a result of the lower temperature range, but this can be worthwhile because of the lower cost involved in gathering heat at this lower temperature. Alternatively, fluids can be used that have boiling points above water, and this may have thermodynamic benefits. See, for example, mercury vapour turbine. The Rankine cycle does not restrict the working fluid in its definition, so the inclusion of an organic cycle is simply a marketing concept that should not be regarded as a separate thermodynamic cycle.

[edit] References
1. ^ Wiser, Wendell H. (2000). Energy resources: occurrence, production, conversion, use. Birkhuser. p. 190. ISBN 9780387987446.

http://books.google.com/books?id=UmMx9ixu90kC&pg=PA190&dq=electrical+power+ generators+steam+percent&hl=en&ei=JppoTpVexNmBB4C72MkM&sa=X&oi=book_re sult&ct=result&resnum=2&ved=0CDgQ6AEwATgK#v=onepage&q=steam&f=false. 2. ^ Canada, Scott; G. Cohen, R. Cable, D. Brosseau, and H. Price (2004-10-25). "Parabolic Trough Organic Rankine Cycle Solar Power Plant". 2004 DOE Solar Energy Technologies (Denver, Colorado: US Department of Energy NREL). http://www.nrel.gov/csp/troughnet/pdfs/37077.pdf. Retrieved 2009-03-17. 3. ^ Batton, Bill (2000-06-18). "Organic Rankine Cycle Engines for Solar Power". Solar 2000 conference. Barber-Nichols, Inc.. http://www.nrel.gov/csp/troughnet/pdfs/batton_orc.pdf. Retrieved 2009-03-18. 4. ^ Nielsen et al., 2005, Proc. Int. Solar Energy Soc.

^Van Wyllen 'Fundamentals of thermodynamics' (ISBN 85-212-0327-6) Moran & Shapiro 'Fundamentals of Engineering Thermodynamics' (ISBN 0-47127471-2) Wikibooks Engineering Thermodynamics

[show]

v t e

Thermodynamic cycles

[show]

v t e

Electricity delivery

Nonrenewable

Coal Fossil-fuel power station Natural gas Petroleum Nuclear Oil shale Biomass Geothermal Hydro Marine Solar Wind

Renewable

Retrieved from "http://en.wikipedia.org/w/index.php?title=Rankine_cycle&oldid=488869861" View page ratings

Rate this page


What's this? Trustworthy Objective Complete Well-written I am highly knowledgeable about this topic (optional) Submit ratings Saved successfully Your ratings have not been submitted yet

Categories: Thermodynamic cycles Scottish inventions


Personal tools

Log in / create account

Namespaces

Article Talk

Variants Views Actions Search


Search Special:Search

Read Edit View history

Navigation

Main page Contents Featured content Current events Random article Donate to Wikipedia

Interaction Toolbox

Help About Wikipedia Community portal Recent changes Contact Wikipedia

What links here

Related changes Upload file Special pages Permanent link Cite this page Rate this page

Print/export

Create a book Download as PDF Printable version

Languages

Catal Deutsch Espaol Franais Galego Hrvatski Bahasa Indonesia Italiano Lietuvi Magyar Polski Portugus Romn Svenska Trke This page was last modified on 23 April 2012 at 19:59. Text is available under the Creative Commons Attribution-ShareAlike License; additional terms may apply. See Terms of use for details. Wikipedia is a registered trademark of the Wikimedia Foundation, Inc., a nonprofit organization. Contact us

Privacy policy About Wikipedia Disclaimers Mobile view

Our updated Terms of Use will become effective on May 25, 2012. Find out more.

Siemens cycle
From Wikipedia, the free encyclopedia

Jump to: navigation, search

Thermodynamics

Branches[show]

Laws[show]

Systems[show]

System properties[show]

Material properties[show]

Specific heat capacity Compressibility Thermal expansion

Equations[show]

Potentials[show]

Internal energy Enthalpy Helmholtz free energy Gibbs free energy


History and culture[show]

Scientists[show]

v t e

The Siemens cycle is a technique used to cool or liquefy gases.[1]. A gas is compressed, leading to an increase in its temperature (by Gay-Lussac's law relating pressure and temperature). The compressed gas is then cooled by a heat exchanger, then the cool, compressed gas is allowed to decompress, further cooling it (again by Gay-Lussac's law). This results in a gas (or liquefied gas) that is colder than the original and at the same pressure. Carl Wilhelm Siemens patented the Siemens cycle in 1857[2].

[edit] See also


Adiabatic process Gas compressor Hampson-Linde cycle Regenerative cooling Timeline of low-temperature technology

[edit] References
1. ^ Adiabatic Expansion Cooling of Gases 2. ^ The Siemens cycle

[hide]

v t e

Thermodynamic cycles
Without phase change (hot air engines)

External combustion cycles

Bell Coleman Brayton/Joule Carnot Ericsson Stirling Stirling (Pseudo / Adiabatic) Stoddard

With phase change


Kalina Rankine (Organic Rankine) Regenerative Atkinson Brayton/Joule Diesel Expander Gas-generator Homogeneous Charge Compression Ignition Lenoir Miller Otto Pressure-fed Staged combustion Combined HEHC Mixed/Dual

Internal combustion cycles

Mixed cycles

Refrigeration cycles

Hampson-Linde Kleemenko Pulse tube Regenerative cooling Transcritical Vapor absorption Vapor-compression Siemens Vuilleumier Barton Humphrey Scuderi

Uncategorized

This physics-related article is a stub. You can help Wikipedia by expanding it.

Retrieved from "http://en.wikipedia.org/w/index.php?title=Siemens_cycle&oldid=380126820" View page ratings

Rate this page


What's this? Trustworthy Objective Complete Well-written I am highly knowledgeable about this topic (optional) Submit ratings Saved successfully Your ratings have not been submitted yet Categories: Thermodynamic cycles Physics stubs
Personal tools

Log in / create account

Namespaces

Article Talk

Variants Views Actions Search


Search Special:Search

Read Edit View history

Navigation

Main page Contents Featured content Current events Random article Donate to Wikipedia

Interaction Toolbox

Help About Wikipedia Community portal Recent changes Contact Wikipedia

What links here Related changes Upload file Special pages Permanent link Cite this page Rate this page

Print/export

Create a book Download as PDF Printable version

Languages

Deutsch Franais Nederlands This page was last modified on 21 August 2010 at 10:23. Text is available under the Creative Commons Attribution-ShareAlike License; additional terms may apply. See Terms of use for details. Wikipedia is a registered trademark of the Wikimedia Foundation, Inc., a nonprofit organization. Contact us Privacy policy About Wikipedia Disclaimers Mobile view

Our updated Terms of Use will become effective on May 25, 2012. Find out more.

Stirling cycle
From Wikipedia, the free encyclopedia

Jump to: navigation, search

Thermodynamics

Branches[show]

Laws[show]

Systems[show]

System properties[show]

Material properties[show]

Specific heat capacity Compressibility Thermal expansion

Equations[show]

Potentials[show]

Internal energy Enthalpy Helmholtz free energy Gibbs free energy


History and culture[show]

Scientists[show]

v t e

This article is about the "adiabatic" Stirling cycle. For the "idealized" Stirling cycle , see the Stirling engine article. The Stirling cycle is a thermodynamic cycle that describes the general class of Stirling devices. This includes the original Stirling engine that was invented, developed and patented in 1816 by Reverend Dr. Robert Stirling with help from his brother, an engineer.[1] The cycle is reversible, meaning that if supplied with mechanical power, it can function as a heat pump for heating or refrigeration cooling, and even for cryogenic cooling. The cycle is defined as a closed-cycle regenerative cycle with a gaseous working fluid. "Closed-cycle" means the working fluid is permanently contained within the thermodynamic system. This also categorizes the engine device as an external heat engine. "Regenerative" refers to the use of an internal heat exchanger called a regenerator which increases the device's thermal efficiency. The cycle is the same as most other heat cycles in that there are four main processes: 1.Compression, 2. heat-addition, 3. expansion and 4. heat removal. However, these processes are not discrete, but rather the transitions overlap.

Contents
[hide]

1 Idealized Stirling cycle thermodynamics 2 Technical complexity of topic 3 Piston motion variations 4 Volume variations 5 Pressure-versus-volume graph 6 Particle/mass motion 7 Heat-exchanger pressure-drop 8 Pressure versus crank-angle 9 Temperature versus crank-angle 10 Cumulative heat and work energy 11 See also 12 References 13 External links

[edit] Idealized Stirling cycle thermodynamics

A pressure/volume graph of the idealized Stirling cycle. In real applications of the Stirling cycles (e.g. Stirling engines) this cycle is quasi-elliptical. The idealized Stirling cycle consists of four thermodynamic processes acting on the working fluid ( See diagram to right): 1. Isothermal Expansion. The expansion-space is heated externally, and the gas undergoes near-isothermal expansion. 2. Constant-Volume (known as isovolumetric or isochoric) heat-removal. The gas is passed through the regenerator, thus cooling the gas, and transferring heat to the regenerator for use in the next cycle.

3. Isothermal Compression. The compression space is intercooled, so the gas undergoes near-isothermal compression. 4. Constant-Volume (known as isovolumetric or isochoric) heat-addition. The compressed air flows back through the regenerator and picks-up heat on the way to the heated expansion space.

[edit] Technical complexity of topic


The Stirling cycle is a highly advanced subject that has defied analysis by many experts for over 190 years. Highly advanced thermodynamics are required to describe the cycle. Professor Israel Urieli writes: "...the various 'ideal' cycles (such as the Schmidt cycle) are neither physically realizable nor representative of the Stirling cycle" [2] The analytical problem of the regenerator (the central heat exchanger in the Stirling cycle) is judged by Jakob to rank 'among the most difficult and involved that are encountered in engineering '.[3][4]

[edit] Piston motion variations

A model of a four-phase Stirling cycle Most thermodynamic textbooks use a highly-simplified form of a Stirling cycle consisting of 4-processes. This is known as an "ideal Stirling cycle", because it is an "idealized" model, and not necessarily an optimized cycle. Theoretically, the "ideal cycle" does have high net work output per cycle. However, it is rarely used for practical reasons, in part because other cycles are simpler or reduce peak stresses on bearings and/or other components. For convenience, the designer may elect to use piston motions dictated by system dynamics, such as the mechanical linkage mechanisms. At any rate, the efficiency and cycle power are nearly as good as an actual implementation of the idealized case. A typical piston-crank or linkage in a so named "kinematic" design, often results in a near-sinusoidal piston motion. Some designs will cause the piston to "dwell" at either extreme of travel. Many kinematic linkages, such as the well known "Ross yoke", will exhibit nearsinusoidal motion. However, other linkages, such as the "rhombic drive", will exhibit more non-sinusoidal motion. To a lesser extent, the ideal cycle introduces complications,

since to implement the cycle in a real engine would require somewhat higher accelerations of the pistons and higher viscous pumping-losses of the working fluid, although the material stresses and pumping-losses in an optimized engine, would only be intolerable when approaching the "ideal cycle" and/or at high cycle rates. Other issues include the time required for heat transfer, particularly for the isothermal processes. In an engine with a cycle approaching the "ideal cycle", the cycle rate might have to be slowed down to address these issues. In the most basic model of a free piston device, the kinematics will result in simple harmonic motion.

[edit] Volume variations


In beta and gamma engines, generally the phase angle difference between the piston motions is not the same as the phase angle of the volume variations. However, in the alpha Stirling, they are the same.[5] The rest of the article assumes sinusoidal volume variations, as in an alpha Stirling with co-linear pistons, so named an "opposed piston" alpha device.

[edit] Pressure-versus-volume graph


This type of plot is used to characterize almost all thermodynamic cycles. The result of sinusoidal volume variations is the quasi-elliptical shaped cycle shown in Figure 1. Compared to the idealized cycle, this cycle is a more realistic representation of most real Stirling engines. The four points in the graph, label the crank-angle in degrees.[6]

The adiabatic Stirling cycle is similar to the idealized Stirling cycle; however, the four thermodynamic processes are slightly different (see graph above):

180 to 270, pseudo-Isothermal Expansion. The expansion-space is heated externally, and the gas undergoes near-isothermal expansion. 270 to 0, near-constant-Volume (or near-isometric or isochoric) heat-removal. The gas is passed through the regenerator, thus cooling the gas, and transferring heat to the regenerator for use in the next cycle. 0 to 90, pseudo-Isothermal Compression. The compression space is intercooled, so the gas undergoes near-isothermal compression. 90 to 180, near-constant-Volume (near-isometric or isochoric) heat-addition. The compressed air flows back through the regenerator and picks-up heat on the way to the heated expansion space.

With the exception of a Stirling thermoacoustic engine, none of the gas particles actually flows through the complete cycle. So this approach is not amenable to further analysis of the cycle. However, it provides an overview and indicates the cycle work.

[edit] Particle/mass motion


Figure 2, shows the streaklines which indicate how gas flows through a real Stirling engine. The vertical colored lines, delineate the volume spaces of the engine. From leftto-right they are: the volume swept by the expansion (power) piston, the clearance volume (which prevents the piston from contacting the hot heat-exchanger), the heater, the regenerator, the cooler, the cooler clearance volume, and the compression volume swept by the compression piston.

Alpha type Stirling. Animated version.

[edit] Heat-exchanger pressure-drop


Also referred to as "pumping losses", the pressure drops shown in Figure 3, are caused by viscous flow through the heat exchangers. The red line represents the heater, green is the regenerator, and blue is the cooler. To properly design the heat exchangers, multivariate optimization is required to obtain sufficient heat transfer with acceptable flow losses.[5] The flow losses shown here are relatively low, and they are barely visible in the following image, which will show the overall pressure variations in the cycle.

[edit] Pressure versus crank-angle


Figure 4 shows results from an "adiabatic simulation" with non-ideal heat exchangers. Note that the pressure-drop across the regenerator is very low compared to the overall pressure variation in the cycle.

[edit] Temperature versus crank-angle


Figure 5 illustrates the adiabatic properties of a real heat exchanger. The straight lines represent the temperatures of the solid portion of the heat exchanger, and the curves are the gas temperatures of the respective spaces. The gas temperature fluctuations are caused by the effects of compression and expansion in the engine, together with non-ideal heat exchangers which have a limited rate of heat transfer. When the gas temperature deviates above and below the heat exchanger temperature, it causes thermodynamic losses known as "heat transfer losses" or "hysteresis losses". However, the heat exchangers still work well enough to allow the real cycle to be effective, even if the actual thermal efficiency of the overall system is only about half of the theoretical limit.

[edit] Cumulative heat and work energy


Figure 6 shows a graph of the alpha-type Stirling engine data, where 'Q' denotes heat energy, and 'W' denotes work energy. The blue dotted-line shows the work output of the compression space. As the trace dips down, and work is done on the gas as it is compressed. During the expansion process of the cycle, some work is actually done on the compression piston, as reflected by the upward movement of the trace. At the end of the cycle, this value is negative, indicating that compression piston requires a net input of work. The blue solid line shows the heat flowing out of the cooler heat-exchanger. Notice that the heat from the cooler, and the work from the compression piston both have the same cycle energy! This is consistent with the zero-net heat transfer of the regenerator (solid green line). As would be expected, the heater and the expansion space both have positive energy flow. The black dotted-line shows the net-work output of the cycle. On this trace, the cycle ends higher that it started, indicating that the heat engine converts energy from heat into work.

[edit] See also


Pseudo Stirling cycle Stirling engine [hide]


v t e

Thermodynamic cycles
Without phase change (hot air engines) External combustion cycles

Bell Coleman

Brayton/Joule Carnot Ericsson Stirling Stirling (Pseudo / Adiabatic) Stoddard

With phase change


Kalina Rankine (Organic Rankine) Regenerative Atkinson Brayton/Joule Diesel Expander Gas-generator Homogeneous Charge Compression Ignition Lenoir Miller Otto Pressure-fed Staged combustion Combined HEHC Mixed/Dual Hampson-Linde Kleemenko Pulse tube Regenerative cooling Transcritical Vapor absorption Vapor-compression Siemens Vuilleumier Barton Humphrey Scuderi

Internal combustion cycles

Mixed cycles

Refrigeration cycles

Uncategorized

[edit] References
1. ^ Robert Sier (1999). Hot air caloric and stirling engines. Vol.1, A history (1st Edition (Revised) ed.). L.A. Mair. ISBN 0-9526417-0-4. 2. ^ Organ, "The Regenerator and the Stirling Engine", p.xxii, Forward by Urieli 3. ^ Organ, "The Regenerator and the Stirling Engine", p.7 4. ^ Jakob, M. (1957) Heat Transfer II John Wiley, New York, USA and Chapman and Hall, London, UK 5. ^ a b Organ, "The Regenerator and the Stirling Engine" 6. ^ Israel Urieli (Dr. Iz), Associate Professor Mechanical Engineering: Stirling Cycle Machine Analysis

[edit] External links


Retrieved from "http://en.wikipedia.org/w/index.php?title=Stirling_cycle&oldid=488546765" View page ratings

Rate this page


What's this? Trustworthy Objective Complete Well-written I am highly knowledgeable about this topic (optional) Submit ratings Saved successfully Your ratings have not been submitted yet Categories: Thermodynamic cycles Stirling engines
Personal tools

Log in / create account

Namespaces

Article Talk

Variants Views Actions Search


Search Special:Search

Read Edit View history

Navigation

Main page Contents Featured content Current events Random article Donate to Wikipedia

Interaction Toolbox

Help About Wikipedia Community portal Recent changes Contact Wikipedia

What links here Related changes Upload file Special pages Permanent link Cite this page Rate this page

Print/export

Create a book Download as PDF Printable version

Languages

Deutsch Espaol Franais Italiano This page was last modified on 21 April 2012 at 20:36. Text is available under the Creative Commons Attribution-ShareAlike License; additional terms may apply. See Terms of use for details. Wikipedia is a registered trademark of the Wikimedia Foundation, Inc., a nonprofit organization. Contact us Privacy policy About Wikipedia Disclaimers Mobile view

Our updated Terms of Use will become effective on May 25, 2012. Find out more.

Solarhydrogen energy cycle


From Wikipedia, the free encyclopedia

Jump to: navigation, search SolarHydrogen energy cycle is an energy cycle where a solar powered electrolyzer is used to convert water to hydrogen and oxygen. Hydrogen and oxygen produced thus are stored to be used by a fuel cell to produce electricity when no sunlight is available.[1]

Contents
[hide]

1 Working 2 Features 3 Use of hydrogen iodide 4 Advantages 5 References

[edit] Working
Photovoltaic panels convert sunlight to electricity. In this cycle, the excess electricity produced after consumption by devices connected to the system, is used to power an electrolyzer. The electrolyzer converts water into hydrogen and oxygen, which is stored. This hydrogen is used up by a fuel cell to produce electricity, which can power the devices when sunlight is unavailable.[1]

[edit] Features
The SolarHydrogen energy cycle can be incorporated using organic thin film solar cells[2] and microcrystalline silicon thin film solar cells[3] This cycle can also be incorporated using photoelectrochemical solar cells. These solar have been incorporated since 1972[4] for hydrogen production[5] and is capable of directly converting sunlight into chemical energy.[4]

[edit] Use of hydrogen iodide


An aqueous solution of hydrogen iodide has been proposed as an alternative to water as a fuel that can be used in this cycle. Splitting of hydrogen iodide is easier than splitting water as its Gibbs energy change for decomposition is lesser. Hence silicon

photoelectrodes can decompose hydrogen iodide into hydrogen and iodine without any external bias.[4]

[edit] Advantages

This cycle is pollution free as the only effluent from this cycle is pure water.[1] Since a hydrogen powered energy economy is more stable than conventional energy economies, this cycle can be incorporated in politically unstable countries.[6]

[edit] References
1. ^ a b c "Schatz Solar Hydrogen Project". schatzlab.org. http://www.schatzlab.org/projects/real_world/schatz_solar.html. Retrieved 2011-06-18. 2. ^ Nakato, Y.; Jia, G.; Ishida, M.; Morisawa, K.; Fujitani, M.; Hinogami, R.; Yae, S. (10 June 1998). "Efficient Solar-to-Chemical Conversion by One Chip of n-Type Silicon with Surface Asymmetry". Electrochem. Solid-State Lett. (Osaka, Japan: Electrochemical Society) 1 (2): 7173. doi:10.1149/1.1390640. http://scitation.aip.org/getabs/servlet/GetabsServlet?prog=normal&id=ESLEF600000100 0002000071000001&idtype=cvips&gifs=yes&ref=no. Retrieved 2011-07-20. 3. ^ Yae, Shinji; Kobayashi, Tsutomu; Abe, Makoto; Nasu, Noriaki; Fukumuro, Naoki; Ogawa, Shunsuke; Yoshida, Norimitsu; Nonomura, Shuichi et al (15 February 2007). "Solar to chemical conversion using metal nanoparticle modified microcrystalline silicon thin film photoelectrode". Solar Energy Materials and Solar Cells (Japan: ScienceDirect) 91 (4): 224229. doi:10.1016/j.solmat.2006.08.010. http://www.sciencedirect.com/science/article/pii/S092702480600362X. Retrieved 201107-20. 4. ^ a b c "Water splitting to produce solar hydrogen using silicon thin film". spie.org. http://spie.org/x8413.xml?ArticleID=x8413. Retrieved 2011-08-30. 5. ^ Fujishima, Akira; Honda, Kenichi (7 July 1972). "Electrochemical Photolysis of Water at a Semiconductor Electrode". Nature (Japan: Nature Publishing Group) 238 (1): 3738. doi:10.1038/238037a0. PMID 12635268. http://www.nature.com/nature/journal/v238/n5358/abs/238037a0.html. Retrieved 201107-20. 6. ^ Turner, John A; Williams, Mark C; Rajeshwar, Rajeshwar (2004). "Hydrogen Economy based on Renewable Energy Sources". CSA Illumina (http://md1.csa.com)&#32;13 (3): 2430. http://md1.csa.com/partners/viewrecord.php?requester=gs&collection=TRD&recid=A06 4421610AH&q=%22Hydrogen+economy+based+on+renewable+energy+sources%22&u id=790995690&setcookie=yes. Retrieved 2011-08-30.

Retrieved from "http://en.wikipedia.org/w/index.php?title=Solar%E2%80%93hydrogen_energy_cycle&o ldid=455853796" View page ratings

Rate this page


What's this?

Trustworthy Objective Complete Well-written I am highly knowledgeable about this topic (optional) Submit ratings Saved successfully Your ratings have not been submitted yet Categories: Thermodynamic cycles
Personal tools

Log in / create account

Namespaces

Article Talk

Variants Views Actions Search


Search Special:Search

Read Edit View history

Navigation

Main page Contents Featured content Current events Random article

Donate to Wikipedia

Interaction Toolbox

Help About Wikipedia Community portal Recent changes Contact Wikipedia

What links here Related changes Upload file Special pages Permanent link Cite this page Rate this page

Print/export

Create a book Download as PDF Printable version This page was last modified on 16 October 2011 at 15:10. Text is available under the Creative Commons Attribution-ShareAlike License; additional terms may apply. See Terms of use for details. Wikipedia is a registered trademark of the Wikimedia Foundation, Inc., a nonprofit organization. Contact us Privacy policy About Wikipedia Disclaimers Mobile view

Das könnte Ihnen auch gefallen