Sie sind auf Seite 1von 16

SAE TECHNICAL PAPER SERIES

2000011081

A Review of Investigations Using the Second Law of Thermodynamics to Study InternalCombustion Engines
Jerald A. Caton
Texas A&M University

Reprinted From: SI Combustion (SP1517)

SAE 2000 World Congress Detroit, Michigan March 69, 2000


400 Commonwealth Drive, Warrendale, PA 15096-0001 U.S.A. Tel: (724) 776-4841 Fax: (724) 776-5760

2000011081

A Review of Investigations Using the Second Law of Thermodynamics to Study Internal-Combustion Engines
Jerald A. Caton
Texas A&M University
Copyright 2000 Society of Automotive Engineers, Inc.

ABSTRACT
Investigations that have used the second law of thermodynamics to study internal-combustion engines in a detailed manner date back to the late 1950s. Over two dozen previous investigations which have used the second law of thermodynamics or availability analyses were identified. About two-thirds of these have been completed for diesel engines, and the other one-third have been completed for spark-ignition engines. The majority of these investigations have been completed since the 1980s. A brief description of each of these investigations is provided. In addition, representative results are presented for both compression-ignition (diesel) and spark-ignition engines to illustrate the type of information obtained by the use of second law analyses. Both instantaneous values for the engine availability, and the overall values for energy and availability are described.

significant findings. The next subsections will review the concept of availability, and provide basic analytical results. This will be followed by major sections on previous work, example results, and summary. AVAILABILITY Related to the analysis based on the second law of thermodynamics is the concept of availability which is also known as essergy (essence of energy) and exergy [14]. Availability, a thermodynamic property of a system and its surroundings, is a measure of the maximum useful work that a given system may attain as the system is allowed to reversibly transition to a thermodynamic state which is in equilibrium with its environment. One key aspect of availability is the fact that a portion of a given amount of energy is available to produce useful work, while the remaining portion of the original energy is unavailable for producing useful work. In general, the processes of interest are the thermal, mechanical and chemical processes. An example of the thermal aspect of availability is a case where the system temperature is above the environmental temperature. By utilizing an ideal heat engine (such as a Carnot engine), the availability from the system could be converted to work until the system temperature equaled the environmental temperature (the remaining energy is, therefore, the unavailable portion of the energy). An example of the mechanical aspect of availability is a system which is at a pressure above the environment. By utilizing an ideal expansion device (such as an ideal turbine), the energy of the system could be converted to work until the system pressure equaled the environmental pressure. A final consideration is the chemical aspect1 of availability. This aspect considers the potential to complete work by exploiting the concentration differences of the various species relative to the related concentrations in the environment. The consideration of
1. The chemical aspect of availability by convention refers to the concentration differences between the species in the system and in the environment [810]. In contrast, the (chemical) fuel energy is included in the availability terms since the total (chemical and sensible) energy is used for the internal energy and for the enthalpy.

INTRODUCTION
Reports on the detailed use of the second law of thermodynamics to study internal combustion engines have been published for over 40 years. While the use of a second law analysis is not necessary for general performance computations, the insight provided by a second law analysis is invaluable in understanding the details of the overall thermodynamics of engine operation. The second law of thermodynamics is a rich and powerful statement of related physical observations that has a wide range of implications with respect to engineering design and operation of thermal systems. For example, the second law can be used to determine the direction of processes, to establish the conditions of equilibrium, to specify the maximum possible performance of thermal systems, and to identify those aspects of processes that are detrimental to overall performance. The objective of the current work was to provide a comprehensive listing and description of all the known work in this area, and to compare and contrast the more 1

the species concentration component of availability is often neglected (particularly when considering mobile engine applications) due to the practical difficulties of implementing such a system and the relatively small amounts of work produced [4, 5]. DETERMINATION OF AVAILABILITY The determination of availability is based on the values of other thermodynamic properties. In this development, the kinetic and potential energies are neglected (and can be shown to be negligible). Since the overall engine operation includes both closed system and open system portions, two forms of availability are needed. At all times, for the complete system:
a = (u u o ) ( p o (v v o )) To ( s so )

A = Aend Astart A = Ain Aout + AQ AW Adest

(4)

(1)

where a is the specific availability (or exergy), u is the specific internal energy, uo, vo and so are the specific internal energy, specific volume and specific entropy for the dead state1, respectively, po and To are the pressure and temperature of the dead state, respectively, v is the specific volume, and s is the specific entropy. The dead state is defined as the conditions of the environment at a temperature of To and a pressure of po. The term, po(vvo), represents the work completed against the atmosphere at po and hence is not useful. For the flow periods (open system), the flow availability (or exergy for flows), af , is given by:
a f = ( h ho ) To ( s s o )

where A is the change of the total system availability for a process, Aend is the total availability at the end of the period, Astart is the total availability at the start of the period, Ain is the total availability transferred into the system accompanying flow into the system, Aout is the availability transferred out of the system accompanying flow out of the system, AQ is the availability transferred accompanying the heat transfer, AW is the availability transfer due to work, and Adest is the availability which is destroyed by irreversible processes. This relation may be used to ascertain the destruction of availability by solving eq. 4 to find Adest. That is,

Adest = Astart Aend + Ain Aout + AQ AW

(5)

For work interactions, the availability is equal to the amount of the work:
AW = W

(6)

For heat transfer, the availability which is transferred out of the system is equal to the available portion of the heat transfer:

T AQ = 1 o Q T

(7)

(2)

where h is the specific enthalpy, ho and so are the specific enthalpy and specific entropy of the dead state, respectively, and s is the specific entropy of the flowing matter. For flows out of the system, the flowing matter is the cylinder contents, and for flows into the system, the flowing matter must be specified. With the above relations (eqs. 1 and 2) and a specification of the reference (or dead) state, the availability of the cylinder contents may be determined throughout the cycle. The total availability is obtained by
A = ma

where AQ is the available portion of the heat transfer, Q is the differential heat transfer which is transferred at a system (boundary) temperature of T. The availability that transfers into the system (Ain) and out of the system (Aout) due to flows are given as follows:

Ai =

 (m a )dt
i f ,i

(8)

where the subscript i refers to each individual flow (for this study, intake or exhaust). The fuel availability of the fuel is needed. this is found from standard relationships, and is often only a few percent higher than the lower heating value of the fuel [4]. ANALYTICAL RESULTS In this sub-section, results are presented which illustrate the general characteristics of availability. These results are general, and not necessarily related only to engines. First, the portion of energy that is available to do work is described. Second, the amount of available energy that is destroyed due to a heat transfer process from a high temperature to a lower temperature is presented. Finally, the amount of available energy which is lost due to combustion processes is described. Only a certain portion of a given amount of energy can be converted to work. The maximum portion would be obtained if the original energy was used in a Carnot heat engine: 2

(3)

where m is the system mass and a is the specific availability. Availability is not a conserved property, but in fact, may be destroyed by irreversible processes such as heat transfer through a finite temperature difference, combustion, friction, and mixing processes. Between any end states, therefore, the change in the availability may be related to the relevant processes:

1. This is actually a restricted dead state since composition equilibrium with the environment is not considered [1].

Wmax

T = Qtotal 1 o T gas

(9)

where Wmax is the maximum work that could be obtained from the original energy, Qtotal is the total original energy, To is the environment temperature, and Tgas is the constant temperature of the original energy. Therefore, the percentage which is available is

above. Another irreversible process is combustion. During combustion, the fuel availability is converted from a chemical form to a thermal form, and in the process the potential to do work is reduced. In the following few paragraphs, the destruction of availability by heat transfer and by combustion are examined. For heat transfer processes across finite temperature differences, a portion of the availability is destroyed. This is due to the fact that at the higher gas temperature a greater portion of the energy is available to produce work. Once the energy is deposited at the wall at a lower temperature, the energys capability to produce work is diminished. This may be determined from the following

AQ Wmax T = = 1 o Qtotal Qtotal Tgas

(10)

Figure 1 shows the percentage of the energy which is available (or unavailable) as a function of gas temperature for an environment temperature of 300 K. As shown, the percentage of available energy increases as the gas temperature increases. Conversely, for low temperatures, a smaller percentage of the original energy is available to produce work. Energy at the ambient temperature (300 K for these results) has no potential to do work, and hence, the available energy is 0% of the original energy. On the other hand, energy at 3500 K is over 91% available.
100 90 80 70 60 50 40 30 20 10
To = 300 K AVAILABLE ENERGY UNAVAILABLE ENERGY

T Agas = Qtotal 1 o Tgas


T Awall = Qtotal 1 o T wall
110
Tw = 300 K

(11)

(12)

100 90 DESTROYED AVAILABILITY (%) 80 70


Tw = 450 K

AVAILABLE ENERGY (%)

60 50
Tw = 600 K

40 30 20 10 0 1000

To = 300 K

1500

2000

2500

3000

GAS TEMPERATURE (K)

0 500 1000 1500 2000 2500 3000 3500 GAS TEMPERATURE (K)

Figure 2.

Figure 1.

Percentage of the energy which is available (and unavailable) as a function of gas temperature for an environment temperature of 300 K.

Percentage of the availability which is destroyed during the heat transfer process from the gas temperature to the wall temperature for an environment temperature of 300 K.

As mentioned above, available energy may be used to produce useful work, may be transferred via heat transfer or mass flows, and may be destroyed by irreversible processes. One such irreversible process is the heat transfer across finite temperature differences described 3

where Agas and Awall are the available energy of the gas and wall, respectively, and Tgas and Twall are the temperatures of the gas and wall, respectively. The percentage of the availability destroyed due to the heat transfer process is

A Adest Agas Awall = = 1 wall Agas Agas Agas


Combining eqs. 11, 12 and 13 yields

(13)

Adest Agas

T 1 o Twall = 1 T 1 o Tgas

(14)

Figure 2 shows the percentage of the availability which is destroyed due to the heat transfer process from the gas temperature to the wall temperature for an environment temperature of 300 K. First, for the case with a wall temperature of 300 K, 100% of the availability is destroyed since all the energy is at the environment temperature and can not produce work. For the other two cases (wall temperatures of 450 and 600 K), the percentage destroyed increases with gas temperature. This is because the availability at the wall temperature remains the same, but the initial availability of the gas is higher for the higher temperatures. Therefore, more availability is destroyed for higher gas temperatures. In other words, the larger the temperature difference, the larger the destruction of availability. Finally, for the higher wall temperatures, the percentage destroyed decreases since the higher wall temperatures retain more of the original availability.
Constant Volume, Adiabatic Combustion pR = 500 kPa = 1.0 Octane-air

Finally, this sub-section will end with comments on another process which destroys available energy, combustion. The chemical energy of the fuel represents the potential to yield a maximum amount of work (available energy). As this energy is converted to thermal energy at a specific temperature, some portion of that available energy is destroyed. Caton [6] has presented results from an analytical study which examined combustion processes in a constant volume, adiabatic system. This system was selected to isolate the combustion destruction of available energy from the other processes. Results were obtained for a variety of conditions for octane and air mixtures. As an example, figure 3 shows the destroyed availability as a percentage of the original available energy as a function of the maximum (adiabatic flame) temperature for an equivalence ratio of 1.0 for an initial reactant pressure of 500 kPa. This pressure, 500 kPa, is representative of the pressure at the start of combustion for a range of internal combustion engines [4]. This range of maximum temperatures was obtained by varying the initial temperature from 500 to 2500 K. The results in figure 3 show that the percentage of the total reactant availability destroyed by combustion decreases monotonically from about 20 to 7.5% as the maximum temperature increases from about 2800 to 3400 K for these conditions. In other words, the destruction of the original availability decreases as the temperature of the combustion process increases. This result was shown to be a direct consequence of the characteristics of the specific availability as a function of temperature [6]. In general, these results suggest that as the combustion temperature increases, the destruction of availability decreases. For the assumptions of this study, however, the destruction of availability does not attain zero even for unrealistically high temperatures. In any case, these high temperatures and pressures are beyond the practical limits of todays designs and materials for combustion devices. Although higher gas temperatures may minimize the destruction of available energy by combustion [5, 6], these higher temperatures may lead to other losses of available energy in practical (actual) engineering systems. In particular, the higher temperatures may result in higher heat transfer which will remove the available energy. Also, if not utilized, the higher availability will be expelled with the exhaust gases. Another consideration would be the potential for higher nitric oxide (NO) formation rates at these higher temperatures.

20

18 DESTROYED AVAILABILITY (%)

16

14

12

10

8 2800 3000 3200 3400

MAXIMUM TEMPERATURE (K)

Figure 3.

Percentage of the availability destroyed by the combustion process for a constant volume, adiabatic system (adapted from Caton [6]). 4

In summary, the above analytical results are intended to illustrate the general characteristics of availability analyses. Specifically, the above discussion has shown that energy has an available and an unavailable portion. Also, two modes of availability destruction, due to heat transfer, and combustion, were described.

PREVIOUS STUDIES
Over two dozen previous studies employing the second law of thermodynamics or availability analyses with respect to internal combustion engines were identified. The majority of these have been completed for diesel engines. The following is a chronological presentation of descriptions of these studies. This presentation is divided into two subsections: (1) early work (19571989), and (2) recent work (19902000). EARLY WORK: 1957 to 1989 One of the earliest documented studies was a brief report presented by Traupel [7] in 1957. Although there were few details, he apparently completed calculations to determine the availability values based on measurements of the principal energy terms. He compared a naturally aspirated diesel engine and a turbocharged diesel engine. He stated that the combustion process accounted for a destruction of about 22.5% and 21.9% of the fuels availability for natural aspirated and turbocharged diesel engines, respectively. He also reported on the losses related to cooling, exhaust, mechanical, and aerodynamic processes. A pioneering work on this topic was reported by Patterson and van Wylen [8] in 1964. They described an early version of a thermodynamic cycle simulation for spark-ignition engines in which they included determination of entropy values. With the entropy values, they then determined availability for the compression and expansion strokes. They isolated the availability destruction associated with the heat transfer and combustion processes. Some of the simplifications of this early work included (1) idealized induction and exhaust processes with instantaneous valve events occurring at top dead and bottom dead center, (2) the induction, compression, and exhaust processes were assumed adiabatic, and (3) the cylinder pressure during the induction and exhaust processes was assumed constant and specified. They summarized their findings by stating that of the availability at the beginning of the compression process, 1/3 was delivered as work, 1/3 was lost due to the combustion and heat transfer processes, and 1/3 was expelled. Clarke [9] examined the Otto, Joule and Atkinson airstandard cycles from the perspective of availability and the associated availability destruction. He described the possibilities of achieving higher thermal efficiencies by recognizing the fundamental availability loss mechanisms for internal combustion engines. Clarke stated that to achieve minimum destruction of availability, the combustion process should be under conditions of near chemical equilibrium. He suggested strategies to achieve minimum destruction of availability. Edo and Foster [10] in 1984 reported on an availability analysis for an engine which utilized dissociated methanol. The use of dissociated methanol was motivated by the potential to capture exhaust energy by dissociating liquid methanol into more readily used 5

gaseous species such as carbon monoxide (CO) and hydrogen (H2). The dissociated products then have the potential to be used as a much leaner reactant mixture thus improving fuel efficiency and reducing emissions. In the course of this study, they completed an availability analysis which used a simple adiabatic, air-standard analysis with an instantaneous heat release for the combustion process, but with equilibrium products. They reported availability as a function of equivalence ratio, and showed the various transfers and destruction of the fuel availability. Beginning in the mid1980s, a number of more detailed investigations were reported on the use of availability. Perhaps the most notable contributions were from a series of investigations by researchers at the Cummins Engine Company. In 1984, the first of these was reported by Flynn et al. [5]. They used a second law analysis to study a turbocharged, intercooled diesel engine. The engine for this study was a 14-liter, in-line six-cylinder, diesel engine operating at 300 kW at 2100 rpm. In particular, they used the second law analysis to evaluate low heat rejection (LHR) engine concepts and secondary heat recovery devices. Essentially they used a standard thermodynamic cycle simulation to obtain the thermodynamic states for a particular engine cycle. They then determined entropy and availability values for these state points, and completed availability balances for the given engine cycle. They showed (for the engine cylinder) that of the original fuel availability about 46% was delivered as useful indicated work, 26% was destroyed, 10% was transferred as heat, and 18% was exhausted. They showed that, as expected, the work output per unit of fuel increased as the equivalence ratio became leaner. Also, as the equivalence ratio becomes leaner, the destruction of availability becomes greater. The reason that the work output increases anyway is that the availability transfers due to heat transfer and exhaust flow decrease much faster as equivalence ratio decreases. The net result, therefore, is an increase in the work output per unit of fuel for the leaner mixtures. This observation has also been reported by others [4]. Further details from this study [5] are presented in a subsequent section of this paper. Primus [11] reported on a second law analysis of exhaust systems for a turbocharged, intercooled diesel engine. This was a companion study to the one reported by Flynn et al. [5]. Primus reported on the influence of the exhaust manifold cross-sectional area upon a number of characteristics such as frictional losses for a 14-liter diesel engine operating at 1900 rpm with an air-fuel ratio of 34.4. He was able to determine an optimum exhaust manifold diameter which minimized the overall loses. Primus et al. [12] described another study which was a continuation of their earlier work (Flynn et al. [5] ). In this study, they used the second law analysis to assess the benefits of turbocharging, charge air cooling, turbocompounding, the implementation of a bottoming cycle, and the use of insulating techniques. The baseline

engine for this study was a 14-liter direct-injection, natural aspirated, diesel engine rated at 185 kW at 2100 rpm. They showed that as the combustion becomes leaner (excess air), the availability destruction increases due to increased mixing and lower bulk gas temperatures. This happens because the high temperature products of combustion are mixed with the excess air. For mixtures closer to or at stoichiometric, this effect is minimized (less excess air, higher temperatures), and hence, the conversion of chemical potential to work is more effective. This finding explained the relative merits of the various options they investigated since each option would have a unique stoichiometry (amount of excess air). For this reason, they found that for turbocharging (with a higher AF ratio) relative to natural aspiration (with a lower AF ratio) the combustion destruction of availability was higher. In 1985, Primus and Flynn [13] reported on a continuation of their earlier work. The engine they used in this study was an inline 10-liter, six-cylinder, turbocharged and aftercooled, direct-injected diesel engine. They conducted a detailed parametric study which examined the effects of a number of engine parameters on the various thermodynamic processes of the engine operation. The parameters examined were engine speed, load, peak cylinder pressure limit, compression ratio, intake air temperature, injection timing and apparent heat release rate shape. They presented their results for the distribution of availability uses and transfers in three forms: tables of the numerical values, graphs of the absolute availability for each mode of availability use or transfer, and graphs of the percentage of the fuel availability for each mode of availability use or transfer. As an example of their results, they demonstrated that as the combustion duration is shortened the combustion destruction of availability decreases due to the increase in the cylinder pressures and temperatures. Also, they showed that as the injection timing is retarded, the combustion destruction of availability increases due to the decrease in the cylinder pressures and temperatures. They listed the percentage of the availability destroyed by combustion as increasing from 21.8 to 32.5% as load (equivalence ratio) decreased. Primus and Flynn [14] in 1986 reported on a further study which continued their previous work. They focused on itemizing the various loss mechanisms associated with a 10-liter, six-cylinder, turbocharged and aftercooled, direct-injected diesel engine. They demonstrated how the second law enhanced their understanding of the thermodynamic processes. They studied in-cylinder and out-of-cylinder processes: in-cylinder heat transfer, combustion, exhaust, friction, turbine, exhaust valve, compressor, aftercooler, intake valve, and exhaust manifold heat transfer. They provided examples of parametric variations of key engine parameters such as intake manifold temperature, injection timing, and exhaust manifold size. 6

van Gerpen and Shapiro [15] also used a second law analysis with a standard cycle simulation for a diesel engine. In contrast to the previous investigations, this work included the chemical component 1 of the availability. Some simplifications of this work were (1) the initial cylinder conditions at bottom dead center (BDC) were assumed to be the ambient conditions with no residual gases, and (2) only compression and expansion strokes were considered (no flows were included). This study [15] was based on a diesel engine and used a dead state based on standard saturated air (with trace CO2, H2O, and Ar) at 298.15 K and 101.35 kPa. They found that the chemical contribution to the availability is highly dependent on the equivalence ratio. For the case they studied, they reported that for lean and stoichiometric equivalence ratios the chemical availability was about 15% of the total availability. For rich cases, the availability was shown to be as high as 90% of the total availability for an especially rich equivalence ratio of 2.0. The large contribution of the chemical component to the availability for the rich cases was a direct result of the relatively high concentrations of species such as H2 and CO which possess significant amounts of chemical (fuel) energy. In other words, for the rich cases, the presence of CO and H2 and other such species have unused fuel energy which means that the availability would be dominated by the chemical component. At least to some extent, the results for the rich cases are not unexpected. From an energy perspective, for these rich cases, the combustion inefficiency2 would be high. The quantification of these losses by the availability analysis is an alternative way to view these inefficiencies. Alkidas [16, 17], in 1988 and 1989, reported on a study which examined the application of a second law analysis for a diesel engine. The engine he used was a 2.0-liter single-cylinder, direct-injection, open-chamber, diesel engine operated at 1200 and 1800 rpm with various loads. This work was different than many of the other investigations in two major ways. First, he defined the thermodynamic system as outside the engine cylinder. Second, he used experimental measurements of the energy rejected to the coolant and lubricating oil, of the brake work, and of the air and fuel flow rates. He then calculated availability values from the thermodynamic states based on the measured values. Alkidas [16, 17] showed that the heat transfer was responsible for the greatest availability transfer, and that the combustion destruction of availability was the next most important mechanism of availability removal. For the cases he studied, the combustion destruction was
1. The chemical component, as mentioned earlier, refers to the potential to do work due to the species concentrations relative to the concentrations in the surroundings. This does not refer to the chemical energy of the fuel. 2. Combustion inefficiency is defined as the ratio of the chemical energy carried out of the engine (due primarily to the presence of combustible species) and the chemical energy of the fuel. [4].

between 25 and 43% of the original fuel availability. Alkidas stated that preheating the intake air decreases the combustion irreversibilities due to the fact that the combustion temperatures increased. In the second paper [17], Alkidas also studied a low-heatrejection diesel engine. This engine used air-gapinsulated piston, liner, fire deck and exhaust port. The engine was tested for 10 operating conditions at 1200 and 1800 rpm. Alkidas showed that the low-heatrejection engine more effectively utilized the fuels availability largely due to the reduced heat losses and the higher combustion temperatures. McKinley and Primus [18] described an assessment of a number of turbocharging systems from both a first law and a second law perspective. They studied a 10-liter, inline six-cylinder, diesel engine operating at 224 kW at 2100 rpm. They examined variable geometry turbocharging, wastegating, and resonant intake systems. The baseline turbocharging system used fixed geometry with no wastegate. Air-to-air aftercooling was employed for all systems. In general, the results of the second law analysis were dominated by the associated changes in the air-fuel ratio used with each of the turbocharging systems. Kumar et al. [19] reported on a second law analysis of a single-cylinder, direct-injected, diesel engine using a comprehensive simulation. This report included only preliminary results for an operating condition of 2000 rpm with an equivalence ratio of 0.7. For the one condition examined, they reported that 16.1% of the fuel availability was destroyed during the combustion process. Lipkea and DeJoode [20] reported on the use of both experimental and simulation results to assess the performance of two direct-injection, 7.6-liter, six-cylinder, heavy-duty, turbocharged, intercooled diesel engines from a second law perspective. Details concerning this engine are provided by Whiting et al. [21]. They included chemical availability in their analysis. For the dead state, they selected standard air at 101.34 kPa and 298.15 K with trace amounts of H2O, CO2, and other species. One objective of their work was to determine the effect of major engine parameters on the fuel consumption. They used an availability analysis to identify the sources of irreversibilities and availability losses during the engine cycle. Lipkea and DeJoode [20] completed an availability analysis for each of the various engine components (such as the turbocharger, intercooler, ports/manifolds, and cylinder). They showed that the exhaust and the heat transfer accounted for about 60% of the fuel energy, but only about 20% of this energy could be used potentially to produce additional work. About 40% of the fuel availability was lost due to internal irreversibilities such as combustion, friction, mixing and heat transfer. Shapiro and van Gerpen [22] extended their earlier work [15] to include a two-zone combustion model and applied this model to both a compression-ignition and a spark7

ignition engine. As before, this study included chemical availability considerations. Their work considered only the compression and expansion strokes, and included no consideration of intake or exhaust flows. They presented the time-resolved values of the availability for cases with different equivalence ratios, residual fractions, and burn durations. They showed, for example, that the combustion irreversibility increases with increasing burn duration. RECENT WORK: 1990 to 2000 In 1991, Bozza et al. [23] described a second law analysis of an indirectinjected, four-cylinder turbocharged, diesel engine. They used experimental measurements to obtain information for the heat release and flow expressions in their simulation. As an example, one operating condition studied was at 4500 rpm and an equivalence ratio of 0.56. They found that for steady-state operation the percentage of the fuel availability destroyed by combustion ranged between about 22 and 26% depending on the values used for the ignition delay, aspiration, turbocharger speed, and other parameters. They also examined transient operation with particular emphasis to the turbocharger performance. Gallo and Milanez [24] reported in 1992 on the use of a cycle simulation to determine the instantaneous irreversibilities, and other second law considerations for a spark-ignition engine using ethanol and gasoline. They focused on the combustion process and valve timings. They examined the effects of ignition timing, duration of combustion, combustion shape factor, and equivalence ratio on second law efficiencies. They found that the use of ethanol (at a compression ratio of 12 compared to a compression ratio of 8 for gasoline) relative to gasoline provided a more effective use of the fuel energy. Further, the combustion irreversibilities were less with ethanol than for gasoline. Al-Najem and Diab [25] presented a short technical note which described brief results for turbocharged diesel engine operated at 243 kW with an air-fuel ratio of 20. They stated that about 50% of the fuel availability is destroyed due to unaccounted factors such as combustion, 15% is removed via exhaust and cooling water, and about 1% is destroyed in the turbocharger. Rakopoulos [26] in 1993 described a first and second law analysis of a spark ignition engine using a cycle simulation and experiment. The engine studied was a variable compression Ricardo E-6 spark ignition engine. The major parameters studied were the compression ratio, fuel-air ratio, and ignition advance. The authors model included the development of a spherical flame front. Only the valve closed period was studied. The author discusses possible ways for improving cycle performance by reducing availability losses due to combustion through improvements in combustion chamber design, fuel-air mixing, and ignition processes. Rakopoulos and Andritsakis [27] in 1993 presented results for the irreversibility rates of two four-stroke cycle

diesel engines. The first engine was a high-speed, direct injection (DI), naturally aspirated, single-cylinder, diesel engine, and the second engine was a medium-speed, indirect-injection (IDI), turbocharged six-cylinder diesel engine. They used experimental information to determine the fuel burning rate, and then used the second law of thermodynamics to deduce the irreversibility rates for each engine. They showed that the accumulated irreversibility was proportional to the fuel burned fraction for a wide range of engine loads, speeds, and injection timings. For the DI engine, the destroyed availability was between about 21 and 31% of the original fuel availability. For the IDI engine, the destroyed availability (for the case of both combustion chambers) was between about 24 and 29% of the original fuel availability. Also, for the IDI engine, the irreversibility of the flow between the prechamber and main chamber was identified. Rakopoulos et al. [28] in 1993 reported on the availability accumulation and destruction in a high-speed, directinjection, naturally aspirated diesel engine. They completed experiments to determine the fuel reaction rates, and then computed the associated second law quantities including the irreversibility production rate. They limited their considerations to the valve closed period. They completed this work for a range of speeds and loads. They also studied limited cooling conditions to determine the implications from a second law perspective on improving efficiency. They considered the use of exhaust heat recovery devices to utilize the extra availability present in the exhaust gases for the limited cooling cases. Table 1. Date 1957 Summary of Previous Investigations (Part 1) (1957 1989) Investigators Engine * Traupel [7] CI Comments Values based on measurements; few details Compression and expansion strokes; simple treatment of intake and exhaust Otto, Joule and Atkinson air-standard cycles Simple Otto airstandard cycle model with equilibrium products (no flows) Comprehensive model of all processes

Table 1. 1984

Summary of Previous Investigations (Part 1) (1957 1989) Primus [11] CI Comprehensive model of all processes; focused on exhaust system optimization Comprehensive model of all processes Comprehensive model of all processes Comprehensive model of all processes Compression and expansion strokes; no intake or exhaust strokes; included chemical availability Experimental measurements of energy terms; calculated availability Comprehensive model of all processes; evaluation of turbocharging systems Comprehensive model of all processes including manifold flow dynamics; included chemical availability; only preliminary results Comprehensive model of all processes; included chemical availability; included experimental measurements

1984 1985 1986 1987 (& 1990)

Primus et al. [12] Primus and Flynn [13] Primus and Flynn [14] van Gerpen and Shapiro [15]

CI CI CI CI

1988 (& 1989) 1988

Alkidas [16, 17]

CI

McKinley and Primus [18] Kumar et al. [19]

CI

1989

CI

1989

Lipkea and DeJoode [20]

CI

1964

Patterson and van Wylen [8] Clarke [9]

SI

1989

Shapiro and SI & CI Compression and van Gerpen expansion strokes; no [22] intake or exhaust strokes; included chemical availability

1976

SI/CI

1984

Edo and Foster [10]

SI=

*SI: spark ignition engine; CI: compression ignition (diesel) engine. =SI engine using dissociated methanol. Rakopoulos et al. [28] stated that their results indicated that the irreversibilities decrease and the availability of the exhaust gas increases with increasing fuel-air ratio (or increasing equivalence ratio). On the other hand, they reported that both the irreversibility and the availability of the exhaust gas increased with engine speed, and slightly decreased with increasing injection timing. 8

1984

Flynn et al. [5]

CI

Table 2. Table I. Summary of Previous Investigations (Part 2) (1990 2000) Date 1991 Investigators Engine * Bozza et al. [23] CI Comments Comprehensive model of all processes; included experimental measurements Comprehensive model of all processes Brief results for a turbocharged diesel engine Compression and expansion strokes; no intake or exhaust strokes; included transient operation Calculated availability; experimental measurements of energy terms; considered only valve closed period; related combustion irreversibility to fuel reacted fraction Experimental measurements of energy terms; calculated availability; considered only valve closed period Comprehensive model of all processes; included experimental measurements Comprehensive model of all processes; included experimental measurements; included transient operation Experimental measurements of energy terms; calculated availability Comprehensive model of all processes; included experimental measurements;

Table 2. Table I. Summary of Previous Investigations (Part 2) (1990 2000) 1998 Anderson et al. [33] Caton [34 37] SI Comprehensive quasidimensional model of all processes Comprehensive model of all processes

1999, 2000

SI

*SI: spark ignition engine; CI: compression ignition (diesel) engine. =Used gasoline and a 30% butanol-gasoline blend. IUsed ethanol and gasoline. Completed for both a conventional cycle and a Miller cycle. Rakopoulos and Giakoumis [29] in 1997 reported on the use of a computer analysis to assess the performance of a turbo-charged, aftercooled, indirect-injected, sixcylinder marine-duty, diesel engine operated over a range of engine speeds, loads and compression ratios. A number of the engine sub-assemblies were studied. These included the compressor, turbine, inlet and exhaust systems, and in-cylinder processes. They showed that the combustion irreversibilities decreased with increasing compression ratio. This observation was due to the fact that the equivalence ratio increased as compression ratio increased due to the corresponding decrease in the compressor pressure ratio. Rakopoulos and Giakoumis [30] in 1997 reported on the use of a computer analysis to study the energy and exergy performance of an indirect-injection, naturallyaspirated diesel engine operating under steady-state and transient conditions. The engine was a Ricardo E-6 research diesel engine with about a 21:1 compression ratio. As an example of their transient results, they considered an acceleration which started at 15% load at 1500 rpm and accelerated to 100% of full load at 1500 rpm in 0.2 seconds. They presented the engine response to the imposed acceleration for speed, injected fuel, engine and load torques, and maximum cylinder pressures as a function of time (or engine cycles). They reported that the combustion irreversibility decreased during acceleration due to slightly higher fueling rates associated with this transient event. Alasfour [31] in 1997 described the results of an availability analysis completed for a single cylinder, spark-ignition fuel-injected Hydra engine using both gasoline and a 30% butanol-gasoline blend. The majority of this work was an experimental study during which he obtained general engine performance results as a function of equivalence ratio. Once he had obtained engine performance, he was able to report the results in terms of energy quantities: brake work, friction work, heat transfer to the coolant, energy out the exhaust, and unaccounted energy losses. He then used these results to determine the related second law quantities. He found 9

1992 1992

Gallo and Milanez [24] Al-Najem and Diab [25] Rakopoulos [26]

SII CI

1993

SI

1993

Rakopoulos and Andritsakis [27]

CI

1993

Rakopoulos et al. [28]

CI

1997

Rakopoulos and Giakoumis [29] Rakopoulos and Giakoumis [30]

CI

1997

CI

1997

Alasfour [31]

SI=

1997

Rakopoulos and Giakoumis [32]

CI

for an equivalence ratio of 0.9 for the butanol-gasoline blend that 49.4% of the fuels availability was not used to produce useful work. In addition, he found that both the first and second law efficiencies increased for lean operation. Rakopoulos and Giakoumis [32] in 1997 described their use of a computer analysis to assess the cumulative and availability rate balances of a multi-cylinder diesel engine. They studied a six-cylinder, turbocharged and aftercooled, indirect-injection diesel engine at full load and 1500 rpm. They neglected chemical dissociation. They included all individual components from compressor through the cylinder to the turbine. They included all processes for both the closed valve and open valve portion of the cycle. They showed that 21.4% of the fuels availability left the cylinder with the exhaust, but after the turbine, the exhaust only contained 13.5%. The combustion irreversibility was responsible for destroying 21.9% of the fuels availability. Anderson et al. [33] in 1998 reported on an investigation of a naturally-aspirated, Miller cycle spark-ignition engine using late intake valve closure. Using a comprehensive quasi-dimensional engine cycle simulation, they compared the Miller cycle strategy with a conventional spark-ignition engine. The advantage of the Miller cycle is that it can use late intake valve closure to control load down to 35% of full load with the use of a throttle. Below this load, the Miller cycle would use supplemental throttling. First law considerations showed that the Miller cycle increased the indicated thermal efficiency at light loads by as much as 6.8%. The second law analysis showed that the conventional throttle destroys up to 3% of the availability. Caton [3437] reported on the use of the second law of thermodynamics to study a spark-ignition engine. This work was based on the use of a comprehensive thermodynamic cycle simulation. In one portion of this study, he examined the effects of engine load and speed on a number of performance, energy and availability terms [37]. A commercial, V-8, spark-ignition engine was selected for this study. Engine loads corresponding to brake mean effective pressures (bmep) of 163, 325, and 655 kPa, and engine speeds of 700, 1400, and 2800 rpm were examined. For these conditions, the availability displaced to the cylinder wall via heat transfer (as a percentage of the fuel availability) ranged between 15.9 and 31.5%. The net availability expelled with the exhaust gases ranged between 21.0 and 28.1%, and the availability destroyed by the combustion process ranged between 20.3 and 21.4%. In addition, this study showed that the mixing of the inlet charge with existing cylinder gases was an additional (but small) mechanism for the destruction of available energy [3437]. Table I is a chronological summary of these previous investigations which used a second law analysis for engine evaluations. The table lists the year of the report, 10

the investigator(s), the type of engine (SI or CI), and comments. In addition, appendix A contains two tables which provide some additional details about the engines used in the above studies. In summary, over two dozen previous studies have been identified which have used the second law of thermodynamics and the concept of availability to examine engine operation in some detail. Most of these studies have use some type of engine simulation, although several based their results on measurements of the principal energy terms. The majority of this previous work has been completed for diesel engines. Although most of the previous work has considered conventional engines, at least four (4) studies included some nonconventional characteristics. These non-conventional characteristics included the use of alternative fuels (such as butanol, ethanol and methanol), and a Miller cycle engine.

EXAMPLE RESULTS
In this section, representative results derived from use of the second law are presented for both a compressionignition and a spark-ignition engine. The presentation of these results is intended to provide a general overview of second law analyses. For detailed results, the referenced works should be consulted. Direct comparisons between the results from the two engines are not possible since the two engines are operating at different conditions and with different outputs. Furthermore, each study has chosen a different thermodynamic system. The following results, therefore, are meant to be illustrative of the nature of second law analyses. COMPRESSION-IGNITION ENGINE For representative results for a compression-ignition (diesel) engine, sample results from the work of Flynn et al. [5] will be presented. In one of their investigations, they studied a 14-liter (with a bore of 140 mm and a stroke of 152mm) turbocharged and intercooled, direct injection diesel engine. The engine was operated at 2100 rpm producing 300 kW of brake power with a 16:1 compression ratio. Figure 4 shows the percentages of the fuels energy and availability for the indicated work, heat transfer, and net exhaust flow. In addition, for the availability, this figure shows the percentage destroyed by combustion and valve throttling irreversibilities. As shown, the indicated work is 47.6% of the energy and 45.8% of the availability (the slight difference is because the availability of the fuel is 1.0317 times the fuels heating value). The heat transfer accounts for 12.6% of the fuel energy, but only 9.7% of the fuel availability. This is because not all the energy of the heat transfer is available to do work. Also, the net exhaust flow consists of 41.4% of the fuel energy, but only 18.3% of the fuel availability. Finally, the combustion irreversibilities are significant: 21.0% of the available energy is destroyed. The throttling losses due to

the flows past the intake and exhaust valves accounted for about 5.3% of the fuels availability.
Compression Work Net Useful Indicated Work Out

this simulation included the computation of entropy, availability, irreversibilities, and the related entropy and availability balances. From the balances, destruction of availability was determined. This simulation was used to complete first and second law analyses for a commercial, spark-ignition engine operating at a part load condition. The selected engine was a V8 configuration with a compression ratio of 8.1:1, and with a bore and stroke of 101.6 and 88.4 mm, respectively. A part load operating condition at 1400 rpm with an equivalence ratio of 1.0 was selected. The instantaneous availability of the system was the net result of the transfer of availability through heat transfer, flows and work, and the destruction of availability due to combustion. Figure 5 shows the system availability and the availability transfers as a function of crank angle. The availability transfers are exhibited as accumulative values which lead to the final system availability. First, the useful work is shown as the top (dashed) curve. During compression, availability is transferred into the system due to the compression work. After top dead center (0CA), the availability transfer is out as the system delivers work. The net indicated useful work is equal to the final value (0.286 kJ) at the end of the cycle (at 584aTDC). The next curve down is for the availability destroyed during the combustion process. Once combustion ends, the difference between the two top curves remains constant. The next curve down (in fig. 5) represents the transfer of availability due to heat transfer, and the final curve accounts for the availability transfer due to flows. When the exhaust valve opens (EVO), the availability decreases sharply. This decrease due to exhaust flow continues until fresh charge enters and availability is then transferred into the system. Near the end of the intake process, the flow reverses and flows out of the system into the intake manifold. Eventually, at the end of the cycle, the system availability of the system has returned to the original value. The final (darkest) curve, therefore, is the instantaneous total system availability. In addition to the instantaneous values of availability, the distribution of the total energy and availability values for the cycle is of interest. Figure 6 shows the percentage of the total fuel energy and total fuel availability that each of the major processes uses. The left-hand bar is for the first law (energy) analyses, and the right-hand bar is for the second law (availability) analyses. First, with respect to the net indicated work, the values using energy units are the same since the indicated work is 100% available energy. The percentages are slightly different due to the slightly higher availability of the fuel relative to its energy value [36].

0.0

USEFUL SYSTEM AVAILABILITY (kJ) AVAILABILITY TRANSFERS (kJ)

-0.2

Combustion Starts

-0.4
Combustion Ends

Availability Destruction due to Combustion

-0.6

Availability Transfer due to Heat Loss EVO

-0.8

Availability Transfer due to Flows Fresh Charge Enters

-1.0

Final System Availability

-180

180 CRANK ANGLE

360

540

Figure 4.

Useful system availability and availability transfers as a function of crank angle for the spark ignition engine [36].
Valve Losses (5.3%) (41.4%)* Destruction due to Combustion (21.0%) Net Transfer Out Due to Flows (12.6%)* (47.6%)* Total Indicated Work Heat Transfer (18.3%)

ENERGY or AVAILABILITY (%)

100

80

60

(9.7%) (45.8%)

40

20

0 1 2

Energy

Availability

Figure 5.

Percentage of the fuels energy and availability for a compression-ignition engine. (*Note: energy values do not add to 100% because of a reported imbalance of 1.6%). Adapted from Flynn et al. [5].

SPARK-IGNITION ENGINE The representative results for a spark-ignition engine are based on recent work by Caton [36]. A thermodynamic engine cycle simulation was extended to include an analysis based on the second law of thermodynamics and the associated computation of availability. The major augmentations to 11

Unused Fuel (0.7%)

ENERGY or AVAILABILITY (%)

100

Destruction due to Inlet Mixing (1.3%) Destruction due to Combustion (20.6%)

REFERENCES
1. Moran, M. J., Availability Analysis A Guide to Efficient Energy Use (Corrected Edition), The American Society of Mechanical Engineers, New York, NY, 1989. 2. Moran, M. J., and Shapiro, H. N., Fundamentals of Engineering Thermodynamics, John Wiley & Sons, Inc., New York, New York, third edition, 1995. 3. Wark, K., Jr., and Richards, D. E., Thermodynamics, sixth edition, McGraw-Hill Company, New York, NY, 1999. 4. Heywood, J. B., Internal Combustion Engine Fundamentals, McGraw-Hill Book Company, New York, New York, 1988. 5. Flynn, P F., Hoag, K. L., Kamel, M. M., and Primus, . R. J., A New Perspective on Diesel Engine Evaluation Based on Second Law Analysis, Society of Automotive Engineers, SAE Paper no. 840032, 1984. 6. Caton, J. A., On the Destruction of Availability (Exergy) Due to Combustion Processes with Specific Application to Internal-Combustion Engines, submitted to Energy, 04 August 1999. 7. Traupel, W., Reciprocating Engine and Turbine in Internal Combustion Engineering, in proceedings of the International Congress of Combustion Engines (CIMAC), Zurich, Switzerland, 1957. 8. Patterson, D. J. and van Wylen, G., A Digital Computer Simulation for Spark-Ignited Engine Cycles, in SAE Progress in Technology, Digital Calculations of Engine Cycles, vol. 7, 1964. 9. Clarke, J. M., The Thermodynamic Cycle requirements for Very High Rational Efficiencies, proceedings of the Sixth Thermodynamics and Fluid Convention, University of Durham, paper no. C53/76, Institute of Mechanical Engineers, London, England, 68 April 1976. 10. Edo, T., and Foster, D., A Computer Simulation of a dissociated Methanol Engine, proceedings of the IV International Symposium on Alcohol Fuel Technology, Ottawa, Canada, May 1984. 11. Primus, R. J., A Second Law Approach to Exhaust System Optimization, Society of Automotive Engineers, SAE Paper no. 840033, 1984. 12. Primus, R. J., Hoag, K. L., Flynn, P F., and Brands, . M. C., An Appraisal of Advanced Engine Concepts Using Second Law Analysis Techniques, Society of Automotive Engineers, SAE Paper no. 841287, also, International Conference on Fuel Efficient Power Trains and Vehicles, the Institution of Mechanical Engineers, paper no. C440/84, pp. 7387, 1984. 13. Primus, R. J., and Flynn, P F., Diagnosing the Real . Performance Impact of Diesel Engine Design Parameter Variation (a Primer in the Use of Second Law Analysis), in International Symposium on Diagnostics and Modelling of Combustion in Reciprocating Engines, pp. 529538, 1985. 14. Primus, R. J., and Flynn, P F., The Assessment of . Losses in Diesel Engines Using Second Law Analysis, in Computer-Aided Engineering of Energy 12

Unused Fuel (0.7%) Net Transfer Out Due to Flows

80

60

(40.0%)

(24.7%) 40 (28.7%) 20 (30.6%) 0 1 2 Total Indicated Work (29.7%) Heat Transfer (23.0%)

Energy

Availability

Figure 6.

Percentage of the fuels energy and availability for a spark-ignition engine [36].

For the heat loss, although 0.268 kJ of energy is transferred out of the system, only 0.221 kJ of this is available energy. Hence the percentages are different. Similarly, for the net flow out, only 24.7% of the available energy is expelled, but 40.0% of the actual energy is expelled. The next two categories apply only to the availability accounting, and not to the first law (energy) aspect. These two categories quantify the availability destruction due to combustion and inlet mixing. The availability destroyed was 20.6 and 1.3%, respectively, for these two processes. Finally, for the parameters selected [36], about 0.7% of the availability and energy of the fuel were not used.

SUMMARY
This paper has reviewed investigations that have used the second law of thermodynamics in studying internalcombustion engines. Over 40 years of efforts and over 28 technical papers have been identified. About two-thirds of these have been completed for diesel engines, and the other one-third has been completed for spark-ignition engines. Almost all of these investigations have been completed since the 1980s. The second law of thermodynamics was shown to provide a framework which leads to a more thorough understanding of the energy conversion process, provides a quantitative measure of the capability to produce useful work, and identifies those processes that are destructive to the goals of high performance and high efficiency engines. Representative results were presented for both compression-ignition (diesel) and spark-ignition engines to illustrate the type of information obtained by the use of second law analyses. Both instantaneous values for the engine availability, and the overall values for energy and availability were presented.

15.

16.

17.

18.

19.

20.

21.

22.

23.

24.

25.

26.

27.

Systems, ed. by R. A. Gaggioli, the American Society of Mechanical Engineers, Advanced Energy Systems Division, AES-Vol. 23, December 1986. van Gerpen, J. H., and Shapiro, H. N., Second-Law Analysis of Diesel Engine Combustion, Journal of Engineering for Gas Turbines and Power, Vol. 112, pp. 129137, 1990, also in Analysis and Design of Advanced Energy Systems: Computer-Aided Analysis and Design, ed. by M. J. Moran, R. A. Bajura, and G. Tsatsaronis, the American Society of Mechanical Engineers, Advanced Energy Systems Division, AES-Vol. 33, December 1987. Alkidas, A. C., The Application of Availability and Energy Balances to a Diesel Engine, Transactions of the ASME, Journal of Engineering for Gas Turbines and Power, vol. 110, pp. 462469, July 1988. Alkidas, A. C., The Use of Availability and Energy Balances in Diesel Engines, Society of Automotive Engineers, SAE paper no. 890822, 1989. McKinley, T. L., and Primus, R. J., An Assessment of Turbocharging Systems for Diesel Engines from First and Second Law Perspectives, Society of Automotive Engineers, SAE Paper no. 880598, 1988. Kumar, S. V., Minkowycz, W. J., and Patel, K. S., Thermodynamic Cycle Simulation of the Diesel Cycle: Exergy as a Second Law Analysis Parameter, International Communications in Heat and Mass Transfer, vol. 16, pp. 335346, 1989. Lipkea, W. H., and DeJoode, A. D., A Comparison of the Performance of Two Direct Injection Diesel Engines from a Second Law Perspective, Society of Automotive Engineers, SAE paper no. 890824, 1989. Whiting, T. M., Hewlitt, R. W., and Shea, M. H., New Deere 7.6L Engine, Society of Automotive Engineers, SAE paper no. 881284, 1988. Shapiro, H. N., and van Gerpen, J. H., Two Zone Combustion Models for Second Law Analysis of Internal Combustion Engines, Society of Automotive Engineers, SAE paper no. 890823, 1989. Bozza, F., Nocera, R., Senatore, A., and Tuccillo, R., Second Law Analysis of Turbocharged Engine Operation, Society of Automotive Engineers, SAE paper no. 910418, 1991. Gallo, W. L. R., and Milanez, L., F., Exergetic Analysis of Ethanol and Gasoline Fueled Engines, Society of Automotive Engineers, SAE paper no. 920809, 1992. Al-Najem, N. M., and Diab, J. M., Energy-Exergy Analysis of a Diesel Engine, Heat Recovery Systems & CHP, vol. 12, No. 6, pp. 525529, 1992. Rakopoulos, C. D., Evaluation of a Spark Ignition Engine Cycle Using First and Second Law Analysis Techniques, Energy Conversion and Management, vol. 34, no. 12, pp. 12991314, 1993. Rakopoulos, C. D., and Andritsakis, E. C., DI and IDI Combustion Irreversibility Analysis, in Thermodynamics and the Design, Analysis and Improvement of Energy Systems, edited by H. J. Richter, Proceedings of ASME-WAM, AES-vol. 30, (also, HTD-vol. 266), pp. 1732, New Orleans, LA, 1993. 13

28. Rakopoulos, C. D., Andritsakis, E. C., and Kyritsis, D. K., Availability Accumulation and Destruction in a DI Diesel Engine with Special Reference to the Limited Cooled Case, Heat Recovery Systems & CHP, vol. 13, pp. 261276, 1993. 29. Rakopoulos, C. D., and Giakoumis, E. G., Speed and Load Effects on the Availability Balance and Irreversibilities Production in a Multi-Cylinder Turbocharged Diesel Engine, Applied Thermal Engineering, vol. 17, no. 3, pp. 299313, 1997. 30. Rakopoulos, C. D., and Giakoumis, E. G., Simulation and Exergy Analysis of Transient Diesel-Engine Operation, Energy, vol. 22, no. 9, pp. 875885, 1997. 31. Alasfour, F. N., Butanol A Single-Cylinder Engine Study: Availability Analysis, Applied Thermal Engineering, vol. 17, no. 6, pp. 537549, 1997. 32. Rakopoulos, C. D., and Giakoumis, E. G., Development of Cumulative and Availability Rate Balances in a Multi-Cylinder, Turbocharged, Indirect Injection Diesel Engine, Energy Conversion and Management, vol. 38, no. 4, pp. 347369, 1997. 33. Anderson, M. K., Assanis, D. N., and Filipi, Z. S., First and Second Law Analyses of a NaturallyAspirated, Miller Cycle, SI Engine with Late Intake Valve Closure, Society of Automotive Engineers, SAE Paper No. 980889, 1998. 34. Caton, J. A., Incorporation and Use of an Analysis Based on the Second Law of Thermodynamics for Spark-Ignition Engines (Using a Comprehensive Cycle Simulation), Report No. ERL9901, Engine Research Laboratory, Texas A&M University, Department of Mechanical Engineering, Version 2.0, 15 February 1999. 35. Caton, J. A., Performance, Energy and Availability Characteristics as Functions of Speed and Load for a Spark-Ignition Engine Using a Thermodynamic Cycle Simulation, Report No. ERL9902, Engine Research Laboratory, Texas A&M University, Department of Mechanical Engineering, Version 1.0, 12 April 1999. 36. Caton, J. A., Results From the Second-Law of Thermodynamics For a Spark-Ignition Engine Using a Cycle Simulation, proceedings of the 1999 Fall Technical Conference, the American Society of Mechanical Engineers, Internal Combustion Engine Division, Ann Arbor, MI, October 1999. 37. Caton, J. A., Operation Characteristics of a SparkIgnition Engine Using the Second Law of Thermodynamics: Effects of Speed and Load, the Society of Automotive Engineers, 2000 SAE International Congress & Exposition, Cobo Center, Detroit, MI, 69 March 2000.

CONTACT INFORMATION
Dr. Jerald A. Caton is a professor in the Department of Mechanical Engineering at Texas A&M University, College Station, Texas, 778433123. He has been working on topics associated with internal combustion

engines since 1972. He also has worked in the areas of gas turbines, selective noncatalytic removal (SNCR) of nitric oxides, alternative fuels, cogeneration, fundamental

combustion topics, and boiler combustion. He may be contacted at: jcaton@mengr.tamu.edu

APPENDIX A ENGINE CHARACTERISTICS


The following two tables are lists of the major engine characteristics for the reviewed investigations for compression-ignition and spark-ignition engines, respectively. The brake power and speed listed are for the stated design point or base operating condition of the Table 01List of CI Engines Reviewed First Author Ref. No. Vd (L) ----Flynn Primus Primus 5 11 12 14 14 14 NA TC TC/AC TC/AC NA TC/AC Primus Primus van Gerpen Alkidas McKinley Kumar Lipkea Shapiro Bozza Al-Najem Rakopoulos 13 14 15 16, 17 18 19 20 22 23 25 27 10 10 1.17 2.0 10 0.78 7.6 1.17 1.37 --0.48 16.6 Rakopoulos Rakopoulos Rakopoulos Rakopoulos 28 29 30 32 0.48 16.6 0.51 16.6 TC/AC TC/AC NA TC/AC NA --TC/AC NA TC TC NA TC/AC NA TC/AC NA TC/AC 6 6 1 1 6 1 6 1 4 --1 6 1 6 1 6 Type No. of Cyls ----6 6 6 Brake Power (kW) Trauple 7 ----300 268 185 220 224 224 --333 224 --~170 --55.8 243 4.0 235 4.0 235 4.0 235 2100 2100 --1200 & 1800 2100 2000 ~2200 --4500 --2000 1500 2000 1500 2000 1500 ----2100 1900 2100 Speed (rpm) study. Where no information was available, dashes are entered.

NA: naturally aspirated; TC: turbo-charged; AC: aftercooled

14

Table 02List of SI Engines Reviewed First Author Ref. No. Patterson Edo Shapiro Gallo Rakopoulos Alsafour Anderson Caton 8 10 22 24 26 31 33 3437 Vd (L) ----1.17 0.4 0.51 0.45 2.0 5.7 NA --NA NA NA NA NA NA Type No. of Cyls 1 --1 1 1 1 4 8 Brake Power (kW) 14.2 --------5.7 6.67 21.9 2800 ----5200 2500 1700 2000 1400 Speed (rpm)

NA: naturally aspirated; TC: turbo-charged; AC: aftercooled

15

Das könnte Ihnen auch gefallen