You are on page 1of 23

Downloaded from rspa.royalsocietypublishing.

org on August 11, 2013

Mathematical analysis of spark ignition engine operation via the combination of the first and second laws of thermodynamics
Ismet Sezer and Atilla Bilgin Proc. R. Soc. A 2008 464, doi: 10.1098/rspa.2008.0190, published 8 December 2008

References Subject collections Email alerting service

This article cites 16 articles

http://rspa.royalsocietypublishing.org/content/464/2100/3107.full. html#ref-list-1 Articles on similar topics can be found in the following collections mechanical engineering (166 articles)
Receive free email alerts when new articles cite this article - sign up in the box at the top right-hand corner of the article or click here

To subscribe to Proc. R. Soc. A go to: http://rspa.royalsocietypublishing.org/subscriptions

Downloaded from rspa.royalsocietypublishing.org on August 11, 2013

Proc. R. Soc. A (2008) 464, 31073128 doi:10.1098/rspa.2008.0190 Published online 22 July 2008

Mathematical analysis of spark ignition engine operation via the combination of the rst and second laws of thermodynamics
SMET S EZER 1, * BY I
1

AND

A TILLA B ILGIN 2

Besikduzu Vocational School, and 2Mechanical Engineering Department, Karadeniz Technical University, 61080 Trabzon, Turkey

This study aims at the theoretical exergetic evaluation of spark ignition engine operation. For this purpose, a two-zone quasi-dimensional cycle model was installed, not considering the complex calculation of uid motions. The cycle simulation consists of compression, combustion and expansion processes. The combustion phase is simulated as a turbulent ame propagation process. Intake and exhaust processes are also computed by a simple approximation method. The results of the model were compared with experimental data to demonstrate the validation of the model. Principles of the second law are applied to the model to perform the exergy (or availability) analysis. In the exergy analysis, the effects of various operational parameters, i.e. fuelair equivalence ratio, engine speed and spark timing on exergetic terms have been investigated. The results of exergy analysis show that variations of operational parameters examined have considerably affected the exergy transfers, irreversibilities and efciencies. For instance, an increase in equivalence ratio causes an increase in irreversibilities, while it decreases the rst and also the second law efciencies. The irreversibilities have minimum values for the specied engine speed and optimum spark timing, while the rst and second law efciencies reach a maximum at the same engine speed and optimum spark timing.
Keywords: spark ignition engine; quasi-dimensional model; availability; irreversibilities; exergy analysis

1. Introduction The cycle models of spark ignition (SI) engines are one of the most effective tools for the analysis of engine performance, parametric examinations and assistance to new developments. These models are mainly classied as thermodynamics-based models and uid dynamics-based models, namely computational uid dynamics (CFD) models. Thermodynamic-based models consist of zero- and quasidimensional (QD) models. Zero-dimensional (ZD) models do not have any spatial resolutions and the cylinder composition is regarded as mean values. Geometric features of the uid motion cannot be predicted in these models due to the nonexistence of uid ow modelling. However, they are fast and cheap in terms of computation. On the other hand, uid dynamics-based models are multidimensional (MD) models that are a solution of the full conservation differential
* Author for correspondence (isezer@ktu.edu.tr).
Received 7 May 2008 Accepted 27 June 2008

3107

This journal is q 2008 The Royal Society

Downloaded from rspa.royalsocietypublishing.org on August 11, 2013

3108

. Sezer and A. Bilgin I


Table 1. Nomenclature.

A Aexh Af Af,ch AQ Atm AW Atot bmep bsfc Cp d iv D E f h I lT Lcr Liv Ls m MBT n p R Rs SA Q LVH Qw T r rf S U v Vc Vf xb XsZR/R s Greek letters Fed f hI hII hv q qs r

availability or exergy, J exergy transfer with exhaust gases, J area of ame front, m2 fuel chemical exergy, J exergy transfer with heat, J thermomechanical exergy, J exergy transfer with work, J total exergy, J brake mean effective pressure, bar brake-specic fuel consumption, kg (kW h)K1 specic heat at constant pressure, J kgK1 K intake valve diameter, m cylinder diameter, m energy, J mass residual gas fraction, dimensionless specic enthalpy, J kgK1 irreversibilities, J characteristic length scale of turbulent ame, m connecting rod length, m maximum intake valve lift, m stroke length, m mass, kg maximum brake torque engine speed, rpm pressure, bar cylinder radius, m radius of spark plug location from cylinder axis, m spark advance, CAD lower heating value of fuel, kJ kgK1 total heat transfer to cylinder walls, J absolute temperature, K compression ratio, dimensionless ame radius, m entropy, J internal energy, J specic volume, m3 kgK1 volume of combustion chamber, m3 enamed volume, m3 mass fraction burned, dimensionless spark plug location from cylinder axis, dimensionless charge-up efciency, dimensionless fuelair equivalence ratio, dimensionless rst law efciency, % second law efciency, % volumetric efciency, % crank angle, CAD spark timing crank angle, CAD density, kg mK3
(Continued.)

Proc. R. Soc. A (2008)

Downloaded from rspa.royalsocietypublishing.org on August 11, 2013

Exergy analysis of SI engine operation


Table 1. (Continued.) tb u subscripts 0 a b cr comb dest e g kin max pot r s tot T u w characteristic burning time of eddy at size of l T, s angular speed, sK1 reference or dead-state conditions intake conditions burned connecting rod combustion destruction entrainment gas kinetic maximum potential exhaust conditions spark total turbulent unburned wall

3109

(NavierStokes) equations in both time and spatial dimensions. CFD models can predict chamber geometry and uid motion in detail. Many sub-models for turbulence, combustion and chemical reactions are required in CFD models. Moreover, a computational mesh and detailed boundary and initial conditions are also needed for the solution. For these reasons, they are very slow and expensive in terms of computation. Furthermore, CFD models suffer from some uncertainties during numerical modelling due to the complexity of engine processes. Therefore, CFD models are mostly used to determine the turbulent uid ow eld and optimum combustion chamber geometry in the case of no combustion conditions (Wengang 1995; Tallio 1998; Sezer 2008). QD models, on the other hand, are fundamentally thermodynamics-based models, as cited before, and they need empirical inputs for turbulent intensities and mass burn rate. QD models have the advantages of both ZD and MD models due to some inuence of chamber geometry and uid motion, are fast in terms of computation time and are easy to use (Dai et al. 1996; Caton 2001; Georgios 2005; Sezer 2008). For these reasons, QD models have been widely used for predicting the effects of ow parameters, mixture composition and combustion chamber geometry in SI engine combustion and heat transfer, and also overall engine performance and emissions (Agarwal et al. 1998; table 1). Both ZD and QD models have been commonly used over the years for research and design studies, but most such simulation studies have been used only for the rst law of thermodynamics. Furthermore, in the few past decades, it has been clearly understood that the rst law of thermodynamics is not capable of providing a suitable insight into engine operations. For this reason, the use of the second law of thermodynamics has been intensied in internal combustion engines (Caton 2000a; Rakopoulos & Giakoumis 2006). The analysis of a process or a system with the second law of thermodynamics is termed availability or exergy analysis.
Proc. R. Soc. A (2008)

Downloaded from rspa.royalsocietypublishing.org on August 11, 2013

3110

. Sezer and A. Bilgin I

The application of exergy analysis in engineering systems is very useful because it provides quantitative information about irreversibilities and exergy losses in the system. In this way, the thermodynamic efciency can be quantied and poor efciency areas identied, so that systems can be designed and operated more efciently (Cengel & Boles 1994; Moran & Shapiro 2000; Rezac & Metghalchi 2004). A series of papers was published on second law or exergy analysis applied to internal combustion engines in the last few decades. A review study was published by Caton (2000a) and was extended by Rakopoulos & Giakoumis (2006). It can be seen from these review papers that numerous studies have been performed on the application of exergy analysis to SI engines (Shapiro & Van Gerpen 1989; Gallo & Milanez 1992; Rakopoulos 1993; Alasfour 1997; Caton 1999a, 2000b, 2002; Kopac & Kokturk 2005) and compression ignition engines (Flynn et al. 1984; Alkidas 1988; Kumar et al. 1989; Van Gerpen & Shapiro 1990; Velasquez & Milanez 1994; Rakopoulos & Giakoumis 1997; Kanoglu et al. 2005; Parlak et al. 2005). In most SI engine studies, the combustion process has been generally simulated by using simple empirical equations (Shapiro & Van Gerpen 1989; Gallo & Milanez 1992; Caton 1999a, 2000b), and studies performed with detailed combustion models are limited (Rakopoulos 1993; Caton 2002). However, the combustion process is the most important stage during SI engine operation and modelling of combustion in a realistic way is very important for exergetic computations. Therefore, this study has been devoted to the investigation of SI engine operation from the second law perspective by using a QD two-zone cycle model, and the combustion period is simulated as a turbulent entrainment process. The details of the cycle model and exergy analysis are presented in the next sections. 2. Mathematical model (a ) Governing equations of the cycle model A QD thermodynamic cycle model, which is mainly based on the model of Ferguson (1985), has been used in this study. Equations of the original model have been rearranged for the model used here. The governing equations of the cycle model were derived from the rst law of thermodynamics (the energy equation) by assuming that the cylinder content obeys the ideal gas law. The energy equation in crank angle basis is written as du dm dQ dV Cu Z Kp ; 2:1 dq dq dq dq where m is the mass of cylinder content; u is the specic internal energy; Q is the heat transfer; p is the pressure; V is the volume; and q is the crank angle. Equation (2.1) shows the variations of the thermodynamic properties with respect to the crank angle. Moreover, these thermodynamic properties are also functions of temperature and pressure, which will be determined in the model. To determine instantaneous cylinder volume, pressure, and burned and unburned gas temperatures, the following governing equations have been used: & i!' r K1 1 h 2 2 0:5 V q Z Vc 1 C 1 Kcos q C 1 K1 Kr cr sin q ; 2:2 2 rcr DU Z Q KW 0 m
Proc. R. Soc. A (2008)

Downloaded from rspa.royalsocietypublishing.org on August 11, 2013

Exergy analysis of SI engine operation

3111

where rcr is half of the ratio of stroke length Ls to connecting rod length Lcr, rcrZLs/2Lcr. dp A C B C C Z ; 2:3 dq D CE where     ! h g vb v ln v b 1 dV Tw vu v ln vu Tw ; BZ 1K 1K Ab C Au ; AZ m dq um Cpb v ln Tb Tb Cpu v ln Tu Tu !   ! dx b v ln v b h b K h u v2 v ln vb 2 v b v ln v b b ; D Zx b C v b K vu C v b C ZK dq v ln Tb Cpb Tb Cpb Tb v ln Tb p v ln p and  ! 2  vu v ln vu 2 vu v ln vu E Z 1K x b C : Cpu Tu v ln Tu p v ln p dTb Kh g Ab Tb K Tw v v ln v b dp h u K hb dxb C b C Z umCpb x b dq Cpb v ln Tb dq x b Cpb dq and dTu Khg Au Tu K Tw dp v v ln vu : C u Z umCpu 1 K x b dq Cpu v ln Tu dq 2:5

2:4

The extra governing equations are also used to compute work output and heat loss, d W d V Zp dq dq and hg d Qw Z Ab Tb K Tw C Au Tu K Tw : dq u 2:7 2:6

In the above equations, heat transfer coefcients h g and the cylinder wall temperature Tw are taken as 500 W mK2 KK1 and 400 K, respectively, as suggested by Ferguson (1985). Ab and Au are also the wetted areas by burned and unburned gases, respectively, which are computed by a geometrical sub-model given below. Further details of the cycle model can be found in Ferguson (1985) and Sezer (2008). (b ) Combustion model After the initiation of combustion at a specied crank angle, two zones, namely burned and unburned, exist in the combustion chamber. Each zone is assumed to be uniform in temperature and homogeneous in composition. It is also assumed that a uniform pressure distribution exists in the cylinder. The combustion process is simulated as turbulent ame propagation and it is supposed that the ame
Proc. R. Soc. A (2008)

Downloaded from rspa.royalsocietypublishing.org on August 11, 2013

3112

. Sezer and A. Bilgin I

front progresses spherically in the unburned gases. Under these assumptions, the following equations, rst presented by Blizard & Keck (1974) and Keck (1982), developed by Tabaczynski et al. (1977, 1980) and also by Bayraktar & Durgun (2003), have been used for the determination of burnt mass fraction: dm e Z ru Af Ue ; 2:8 dq dm b 2:9 Z ru Af SL C me K m b =tb dq and tb Z l T =SL ; 2:10 where Af is the area of the ame front and tb is the characteristic burning time of the eddy of size l T. The rate of mass burned is proportional to the ame front area and the ame speed. The Af has been computed instantaneously depending on the enamed volume Vf with the help of the geometrical sub-model. Moreover, the following equations were used for determining the turbulent entrainment speed Ue, the turbulent speed UT and the characteristic length scale of the turbulent ame l T: Ue Z UT C SL ;  i ri =re 1=2 ; UT Z 0:08U  i Z hv Apc =Aiv Ls n U 30 and l T Z 0:8L iv re =ri 3=4 ; 2:14 2:11 2:12 2:13

 i is the speed of gases entering the cylinder during induction; Apc is the where U area of the piston crown; Aiv is the maximum opening area of the intake valve; and L iv is the maximum intake valve lift. Laminar ame speed SL is also calculated by the equations developed by Gu lder (1984), SL f; T ; p Z SL;0 Tu =T0 a p=p0 b 1 Kjf ; 2:15

where SL,0 is the laminar ame speed at standard conditions; T0Z298 K; and p0Z1 bar and is calculated as SL;0 f Z ZW fh expKxf K 1:0752 : 2:16

The values of a, b, Z, W, h and x are given in Gu lder (1984) and j is selected as 2.5 for 0%f R0.3. (c ) Geometric sub-model In the geometric sub-model, it is supposed that the spherical ame propagates from the spark plug through the combustion chamber, as shown in gure 1. Under this assumption, the enamed volume Vf, ame surface area Af and heat
Proc. R. Soc. A (2008)

Downloaded from rspa.royalsocietypublishing.org on August 11, 2013

Exergy analysis of SI engine operation


spark plug cylinder axis

3113

Figure 1. Flame front geometry.

transfer surface area Aw have been computed, depending on the instantaneous enamed volume Vf and chamber height h, by using an iterative method. The values of Vf and h have been determined from the thermodynamic cycle model as follows: Vf q Z Vb q C m e q K m b q=ru and hq Z Ls 1 Kr K1 C 0:51 Kcosq: 2:18 2:17

The geometric features of the ame front in gure 1 have been determined for any ame radius from the following mathematical relations (Blizard & Keck 1974; Keck 1982; Bilgin 2002): zZy Af rf Z 2rf az dz 2:19
zZ0

and Vf rf Z 2rf

zZy
zZ0

az r 2 z C bz R2 K Xs R sin bz dz ;

2:20

where cos a Z Xs2 C r 2 z KR2 =2Xs r z ; cos b Z Xs2 C r 2 z KR2 =2Xs r z ; 2 and r 2 z Z r 2 f Kz . The total combustion chamber surface area wetted by the burned gases is the sum of the areas of the cylinder head, cylinder wall and piston crown wetted by
Proc. R. Soc. A (2008)

Downloaded from rspa.royalsocietypublishing.org on August 11, 2013

3114 the burned gases as below,

. Sezer and A. Bilgin I

Aw rf Z Aw;ch rf C Aw;cw rf C Aw;pc rf ; Aw;ch rf Z a0r 2 0 C b0R 2 K Xs R sin b0; zZy Aw;cw rf Z 2 Rbz dz
zZ0

2:21 2:22 2:23

and Aw;pc rf Z ay r 2 y C by R2 K Xs R sin by: 2:24 The area wetted by the unburned gases is determined by assuming the total chamber area is the sum of the areas in contact with the burned and unburned zones as follows: Atot Z Ab C Au : 2:25 (d ) Computation of the cycle As known, SI engine cycles consist of four consecutive processes, namely intake, compression, expansion and exhaust. Intake and exhaust processes are computed by using the approximation method given by Bayraktar & Durgun (2003). In this method, pressure loss during the intake process is calculated from the Bernoulli equation for one-dimensional uncompressible ow, and intake pressure and temperature are determined as pa Z p0 KDpa and Ta Z T0 C DT C gr Tr =1 C gr Tr ; 2:27 2:26

where p0 and T0 are the ambient pressure and temperature, respectively, and Dpa is the pressure loss. The volumetric efciency is determined as hv Z Fed r pa T0 ; r K 1 p 0 T0 C DT C g r Tr 2:28

where Fed is the charge-up efciency; r is the compression ratio; and gr is the molar residual gas fraction. Compression, combustion and expansion processes have been computed by arranging the governing equations (2.3)(2.7) for each process in a suitable manner. Exhaust pressure pr and exhaust temperature Tr are also specied in terms of ambient pressure p0 and burned gas temperature Tb, respectively (Bayraktar & Durgun 2003), pr Z 1:05 K 1:25p0 and Tr Z Tb =pb =pr 1=3 :
Proc. R. Soc. A (2008)

2:29

2:30

Downloaded from rspa.royalsocietypublishing.org on August 11, 2013

Exergy analysis of SI engine operation

3115

Once the computation of cycle is completed, engine performance parameters, i.e. brake mean effective pressure (bmep) and brake-specic fuel consumption (bsfc), can be determined by using well-known equations (Ferguson 1985; Bayraktar & Durgun 2003; Sezer 2008). (e ) The second law (exergy ) concept The second law is analogous to the statement of entropy balance as follows (Cengel & Boles 1994; Moran & Shapiro 2000): DS Z Q=T boundary C s; 2:31 where s is the total entropy generated due to the internal irreversibilities. Considering the combination of the rst and second laws of thermodynamics, the exergy or availability equation can be stated for a closed system (Cengel & Boles 1994; Moran & Shapiro 2000), A Z E C p0 V K T0 S ; 2:32 where E is the total energy, which is the sum of the internal, kinetic and potential energies (E totZUCE kinCEpot); V and S are the volume and entropy of the system, respectively; and p0 and T0 are the xed pressure and temperature of the dead state, respectively. Availability (or exergy) is dened as the maximum theoretical work that can be obtained from a combined system (combination of a system and its reference environment) when the system comes into equilibrium (as thermally, mechanically and chemically) with the environment (Cengel & Boles 1994; Caton 2000a; Moran & Shapiro 2000; Rezac & Metghalchi 2004; Rakopoulos & Giakoumis 2006). The maximum available work from a system emerges as the sum of two contributions: thermomechanical exergy Atm and chemical exergy Ach. Thermomechanical exergy is dened as the maximum extractable work from the combined system as the system comes into thermal and mechanical equilibria with the environment, and it is determined as (Van Gerpen & Shapiro 1990; Moran & Shapiro 2000; Rakopoulos & Giakoumis 2006) X Atm Z E C p0 V K T0 S K m0;i m i ; 2:33 where m i and m 0,i are the mass and chemical potential of species i, respectively, calculated at restricted dead-state conditions. At the restricted dead-state conditions, the system is in thermal and mechanical equilibria with the environment and no work potential exists between the system and the environment due to temperature and pressure differences. But the system does not reach chemical equilibrium with the environment, as the contents of the system are not permitted to mix with the environment or enter the chemical reaction by environmental components (Van Gerpen & Shapiro 1990). In principle, the difference between the compositions of the system at the restricted dead-state conditions and the environment can be used to obtain additional work to reach chemical equilibrium. The maximum work obtained in this way is called chemical exergy and is determined as (Van Gerpen & Shapiro 1990; Moran & Shapiro 2000; Rakopoulos & Giakoumis 2006) X Ach Z m i m 0;i Km0 2:34 i ;
Proc. R. Soc. A (2008)

Downloaded from rspa.royalsocietypublishing.org on August 11, 2013

3116

. Sezer and A. Bilgin I

where m0 i is the chemical potential of species i calculated at true deadstate conditions. The availability balance for a closed system for any process can also be written as (Caton 1999a, 2000a,b) DA Z A2 K A1 Z AQ K AW K Adest ; 2:35 where DA is the variation of the total system availability; A2 is the total availability at the end of the process; A1 is the total availability at the start of the process; AQ is the availability transfer due to heat transfer interactions; AW is the availability transfer due to work interactions; and Adest is the destroyed availability by irreversible processes. Considering fuel chemical exergy, the exergy balance equation for the engine cylinder is written as (Zhang 2002; Sezer 2008)     dA T dQ dW dV m dx b dI Z 1K 0 K K p0 a K comb : 2:36 C f dq dq dq T dq m dq f ;ch dq The left-hand side of equation (2.36) is the rate of change in the total exergy of the cylinder contents. The rst and second terms on the right-hand side represent exergy transfers with heat and work, respectively. The third term on the right-hand side corresponds to burned fuel exergy; here, m f and m tot are the masses of fuel and total cylinder contents and af,ch is the fuel chemical exergy. a f,ch is calculated by using the following equation that is developed for liquid fuels by Kotas (1995): h0 o0 Af ;ch Z a f ;ch m f Z QLHV 1:0401 C 0:01728 0 C 0:0432 0 c c 0  0 ! s h C0:2196 0 1 K2:0628 0 ; c c

2:37

where QLHV is the lower heating value of the fuel that is calculated by using the Mendeleyev formula QLHV Z 33:91c 0 C 125:6h 0 K 10:89o 0 Ks 0 K2:519h 0 Kw 0 : 2:38

The quantities h 0 , c 0 , o 0 , s 0 and w 0 in equations (2.37) and (2.38) represent the mass fractions of the elements carbon, hydrogen, oxygen, sulphur and water content in the fuel, respectively. The last term on the right-hand side of equation (2.36) illustrates exergy destruction in the cylinder due to combustion. It is calculated as _ comb ; I_comb Z T0 S 2:39 _ comb is the rate of entropy generation due to combustion irreversibilities. where S It is calculated from the two-zone combustion model depending on entropy balance as (Zhang 2002; Sezer 2008) _ comb Z dm b s b C dm u su ; S dq dq 2:40

where m b and m u are the masses and sb and su are the specic entropy values of the burned and unburned gases of the cylinder contents, respectively.
Proc. R. Soc. A (2008)

Downloaded from rspa.royalsocietypublishing.org on August 11, 2013

Exergy analysis of SI engine operation

3117

Moreover, the total exergy destruction considered here consists of combustion and heat transfer irreversibilities as follows (Sezer 2008): I_tot Z I_comb C I_Q : 2:41 Entropy production sourced from the heat transfer process is as in equation (2.42), and exergy destruction due to heat transfer has already been determined in equation (2.36), _ _ _Q Z Qb CQu ; S 2:42 Tb Tu _ u are the rates of heat loss from the burned and unburned gas _ b and Q where Q zones at temperatures Tb and Tu, respectively. The efciency is dened to be able to compare different engine size applications or evaluate various improvement effects from the perspective of either the rst or second laws (Rakopoulos & Giakoumis 2006). The rst law (or energy-based) efciency is dened as energy out as work W hI Z Z ; 2:43 energy in m f QLHV where W is the indicated work output. Various second law efciencies (exergetic or availability efciency, or effectiveness) have been dened in the literature (Rakopoulos & Giakoumis 2006). In this study, the following denition is used for the second law efciency: hII Z exergy out as work AW : Z exergy in m f a f ;ch 2:44

3. Numerical applications (a ) Computer program and solution procedure A computer code has been written for the presented SI engine cycle model. Input parameters in the software were r, n, f, Xs, qs, properties of the fuel and ambient pressure and temperature. Once determining the intake conditions, the thermodynamic state of the cylinder charge is predicted by solving the governing differential equations. To integrate these differential equations, the DVERK subroutine is used. The composition and thermodynamic properties of the cylinder content in the simulation are computed using the FORTRAN subroutines fuelairresidual gas and equilibrium-combustion-products, which were originally developed by Ferguson (1985). Exergetic calculations are performed simultaneously, depending on the thermodynamic state of the cylinder content. Finally, the results obtained have been corrected using error analysis as follows (Ferguson 1985; Sezer 2008): 31 Z 1 Kvm =V and 32 Z 1 C W =Dmu C Qw :
Proc. R. Soc. A (2008)

3:1 3:2

Downloaded from rspa.royalsocietypublishing.org on August 11, 2013

3118
50

. Sezer and A. Bilgin I

40

pressure (bar)

30

20

10

Rakopoulos (1993)

presented model 0 120 80 40 0 40 crank angle (CAD) 80 120

Figure 2. Comparisons of computed and experimental cylinder pressures. rcZ7; fZ1.0; fZ0.1; XsZ0.1; qsZK30 CAD; nZ2500 rpm; fuel, gasoline. Table 2. Specications of the engines. specication experimental; Rakopoulos (1993) parametric; Ferguson (1985) r 7.5 10 Xs 0.1 0 D (mm) Ls (mm) Lcr (mm) d iv (mm) Liv (mm) 76.2 100 110 80 241.3 160 35 40 4.8 5

Selecting the values of 31 and 32 at the 10K4 level, the condence of the analysis and the computer program is fullled. (b ) Validation of the model To demonstrate the reliability of the presented cycle model, the predicted values are compared with the experimental data in gure 2 for the conditions specied in the gure and the engine specications given in table 2. Cylinder pressure is selected as the comparison parameter, and predictions for pressure are in good agreement with the experimental data, as shown in the gure. Therefore, it can be said that the presented model has a sufcient level of condence for the analysis of engine performance and parametric investigation. 4. Results and discussion Figure 3 shows typical variations of the exergetic terms during the investigated part of the cycle for the conditions given in the gure. As shown in the gure, the thermomechanical exergy (Atm) increases gradually during the compression
Proc. R. Soc. A (2008)

Downloaded from rspa.royalsocietypublishing.org on August 11, 2013

Exergy analysis of SI engine operation


3000

3119

2500

2000 exergetic terms (J)

1500

1000

500

500 180 150 120 90 60 30 0 30 60 crank angle (CAD) 90 120 150 180

Figure 3. Typical variations of exergetic terms as a function of crank angle. fZ1.05; fZ0.1; rZ10; XsZ0; qsZK30 CAD; nZ3000 rpm; DqbZ64 CAD.

stroke up to the beginning of combustion, while fuel chemical exergy (Af,ch) remains constant. During compression, increase in the thermomechanical exergy directly depends on the exergy transfer with work (AW) and AW shows a symmetrical variation by thermomechanical exergy with a negative sign. There is no remarkable variation in irreversibilities (I ) due to a negligible exergy transfer with heat (AQ) in this period. The variation in the total exergy (Atot) reects the behaviour of the thermomechanical exergy variation. With the start of combustion, at the crank angle of 30 CAD before top dead centre, the chemical exergy of the fuel decreases rapidly due to conversion to heat. As a result of this conversion, i.e. combustion of fuel, the temperature and pressure increase in the cylinder, which results in a steep increase in thermomechanical exergy and an increase in exergy transfer from the cylinder contents to the walls by heat transfer. Both heat transfer and, especially, combustion give a rapid increase in the irreversibilities due to entropy generation. Combustion ends at the crank angle of 34 CAD after top dead centre and the expansion process continues until the piston reaches the bottom dead centre. During this part of the cycle, decreases in Atm and Atot continue due to the exergy transfers (both work and heat) from the system, while irreversibilities stand at almost a constant level. The remaining exergy in the cylinder at the end of the expansion emits with exhaust gases, which is called exergy transfer with exhaust (Aexh). The distributions of the exergetic terms in the fuel exergy, i.e. AQ, AW, Aexh and I are approximately 8, 36.5, 36.9 and 18 per cent, respectively, for the conditions given in gure 3.
Proc. R. Soc. A (2008)

Downloaded from rspa.royalsocietypublishing.org on August 11, 2013

3120
(a) exergy transfer with heat (J) 50 0 50

. Sezer and A. Bilgin I


(b) 1200 exergy transfer with work (J) 1.1 800 1.0 400 = 0.9

= 0.9 100 150 200 250 1.1 1.0

400 1.1 1.0 = 0.9 (d ) 2500 thermomechanical exergy (J) 2000 1500 1000 500 0 ( f ) 3000 1.1 1.0 = 0.9 total exergy (J) 2500 2000 1500 1000 500 180 120 60 0 60 120 180 crank angle (CAD) 1.1 1.0 = 0.9 = 0.9 1.0 1.1

(c)

500 400

irreversibilities (J) fuel exergy (J)

300 200 100 0

(e) 3000 2500 2000 1500 1000 500 0 180 120 60 0 60 120 180 crank angle (CAD)

Figure 4. Effects of equivalence ratio on exergetic terms. (af ) f Z0; rZ10; XsZ0; qsZMBT; nZ3000 rpm.

Figure 4 shows the effects of equivalence ratio (f) on exergetic terms during the periods of compression, combustion and expansion. As shown in the gure, the effects of f appear clearly after the start of combustion in AQ and AW, the irreversibilities and also the thermomechanical exergy. Af,ch and Atot variations,
Proc. R. Soc. A (2008)

Downloaded from rspa.royalsocietypublishing.org on August 11, 2013

Exergy analysis of SI engine operation


Table 3. The values obtained from a parametric study for different conditions. parameter f 0.9 1.0 1.1 n (rpm) 1500 3000 4500 SA (CAD) 15 30 45 4.36!10K2 4.86!10K2 5.38!10K2 4.86!10K2 4.86!10K2 4.87!10K2 4.85!10K2 4.86!10K2 4.80!10K2 1957.72 2180.55 2194.49 2180.82 2180.55 2187.43 2179.22 2180.55 2155.90 2112.32 2352.84 2604.59 2353.06 2352.84 2359.99 2351.41 2352.84 2331.53 2796.91 2875.05 2877.81 2892.88 2875.05 2858.49 2790.36 2875.05 2939.42 12.79 13.86 13.98 13.33 13.86 13.85 13.50 13.86 13.06 70 65 66 36 65 110 77 65 67 m f (g) Qf ( J) Af,ch ( J) Tb,max (K)

3121

bmep (bar) Dqb (CAD)

however, are affected from the beginning of the compression. The values of AQ in gure 4a have the maximum values for a stoichiometric mixture (fZ1.0), while they decrease to lower values for both the lean mixture (fZ0.9) and rich mixture (fZ1.1). This can be attributed to the temperature values given in table 3; the stoichiometric and rich mixtures give higher combustion temperatures in comparison with the lean mixture due to more fuel (m f) and fuel energy (Qf). The values of AQ for rich and lean mixtures are less than the value for a stoichiometric mixture, approximately 6.9 and 9.2 per cent, respectively. However, the AW values in gure 4b increase with increasing equivalence ratio; thus the maximum values have been obtained with a rich mixture, but it is also noted that the rich mixture gives much closer AW values to those of a stoichiometric mixture. These variations can be attributed to the bmep values given in table 3, which are also a direct result of the xed maximum brake torque timing. The values of AW for fZ0.9 and 1.0 are less than the value of fZ1.1, approximately 7.6 and 0.8 per cent, respectively. The irreversibilities in gure 4c have their maximum values for fZ1.1 and decrease to lower values with decreasing equivalence ratios. The variations in the irreversibilities can be explained by combustion efciency. Decreases in the equivalence ratio beyond that of a stoichiometric mixture result in the leaning of the mixture, and combustion becomes more efcient due to the abundant oxygen in the fuelair mixture, which results in a decrease in the combustion irreversibilities. Furthermore, increases in the irreversibilities with increasing equivalence ratio can be explained by the composition and the amounts of combustion products, as cited by Caton (1999b). He stated that the destroyed value of availability, on a per mass of mixture basis, increased with increasing equivalence ratio. The values of the irreversibilities for fZ0.9 and 1.0 are less than the value of fZ1.1, approximately 6.2 and 5.7 per cent, respectively. There are closer variations in gure 4d between the curves of Atm for fZ1.0 and 1.1, while a lean mixture gives considerably lower values, especially during expansion. These variations can be explained by the fact that the increased fuel exergy contribution for the rich mixture by excess fuel cannot be converted to work due to insufcient oxygen,
Proc. R. Soc. A (2008)

Downloaded from rspa.royalsocietypublishing.org on August 11, 2013

3122
(a) exergy transfer with heat (J) 100 0 100 200 300

. Sezer and A. Bilgin I


(b) 1200 exergy transfer with work (J)
0 rp m

4500 rpm
30

800

300

00

rp

400

rpm 4500 rpm 1500

1500 rpm 400

400 (d ) 2500 thermomechanical exergy (J) 2000 1500 1000 1500 rpm 500 0 ( f ) 3000 1500 rpm total exergy (J) 2500 2000 1500 1000 1500 rpm 500 0 180 120 60 0 60 120 180 crank angle (CAD) 4500 rpm 3000 rpm 4500 rpm 3000 rpm

(c)

500 400

irreversibilities (J)

300 200 100 0

rpm 1500 4500 rpm

3000 rpm

(e) 2500 2000 fuel exergy (J) 1500 3000 rpm 1000 500 4500 rpm

0 180 120 60 0 60 120 crank angle (CAD)

180

Figure 5. Effects of engine speed on exergetic terms. (a f ) fZ1.0; f Z0; rZ10; XsZ0; qsZMBT.

which presents incomplete combustion. Furthermore, the reduced cylinder pressures and temperatures come out for the lean mixture due to fuel deciency. Thus, the stoichiometric mixture gives the maximum Atm values due to the lower irreversibilities than the rich mixture. Enriching of the fuelair mixture increases Af,ch in gure 4e, but, as cited above, excess fuel, more than required, cannot be
Proc. R. Soc. A (2008)

Downloaded from rspa.royalsocietypublishing.org on August 11, 2013

Exergy analysis of SI engine operation

3123

converted efciently to useful work and the remaining fuel is wasted with exhaust, as seen in the gure. The main contribution to variations in Atot given in gure 4f is supplied by the fuel exergy during compression, while both thermomechanical and fuel exergies donate to it in the rest of the cycle. Aexh increases by increasing the equivalence ratio so rich mixtures give the maximum exhausted exergy. The values of Aexh for fZ1.0 and 1.1 are higher than that of fZ0.9, approximately 24.5 and 42.7 per cent, respectively. Figure 5 shows the effects of engine speed (n) on exergetic terms during the examined part of the cycle. As shown in gure 5a, AQ values decrease with increasing engine speed. This can be explained by the variation of cycle duration, such that the quantity of heat transferred to the cylinder wall increases as the engine runs slowly due to the enlarging of the cycle duration. The combustion temperatures in table 3 have also affected the AQ values. The values of AQ for 3000 rpm and 4500 rpm are less than the value of 1500 rpm, approximately 42.5 and 58.7 per cent, respectively. However, AW in gure 5b has the maximum values for 3000 rpm during expansion, while 4500 rpm has the maximum AW during compression. The variations in AW can be explained by the bmep values in table 3; the useful work and AW also increase as the bmep increases. The values of AW for 1500 rpm and 4500 rpm are less than that of 3000 rpm, approximately 3 and 1 per cent, respectively. Irreversibilities in gure 5c also reach to the minimum for 3000 rpm. The values of the irreversibilities for 1500 rpm and 4500 rpm are greater than the value for 3000 rpm, approximately 3.5 and 0.3 per cent, respectively. The results for AW and the irreversibilities show that the engine speed of 3000 rpm is an optimum speed among examined speeds. Atm in gure 5d is signicantly affected by the variation of engine speed, especially during expansion. The increase in engine speed has not varied the peak values of Atm, but it causes an increase in Aexh at the end of expansion. The Af,ch values in gure 5e do not vary with engine speed for the same amount of fuel supplied to the cylinder, as shown in table 3. The steep variation takes place in Af,ch during combustion and the slopes of the lines decrease with increasing engine speed due to the variation of combustion duration. The variations in Atot in gure 5f reveal the combination of thermomechanical and fuel exergies. The Aexh increases by increasing the engine speed due to the enlarging of the combustion duration. The values of Aexh for 1500 rpm and 4500 rpm are greater than the value for 3000 rpm, approximately 35 and 47.3 per cent, respectively. Figure 6 shows the effects of the spark timing (qs) on exergetic terms during the compression, combustion and expansion periods. AQ values in gure 6a decrease to lower values with the retarding of the spark timing, thus the minimum AQ values are obtained by K15 CAD. The advances in the beginning of combustion increase AQ due to the extension of the contact duration of hot gases with the cylinder walls. The combustion temperatures in table 3 are the other factor in the variations in AQ; increasing combustion temperature increases heat transfer naturally. The values of AQ for K15 and K30 CAD are less than that of K45 CAD, approximately 3.2 and 6.96 per cent, respectively. In gure 6b, AW has the maximum values for K30 CAD during expansion, while increasing spark timing causes increases in the compression work. Thus, additional compression work is required during compression for the advanced spark timing, but the advanced spark timing cannot contribute to the exergy output as useful work during expansion. The values of AW for K15 and K45 CAD are less than
Proc. R. Soc. A (2008)

Downloaded from rspa.royalsocietypublishing.org on August 11, 2013

3124
(a) exergy transfer with heat (J) 50 0 50 100 150 200 250 (c) 500 400 irreversibilities (J) 300 200 100 0 (e) 2500 2000

. Sezer and A. Bilgin I


(b) 1200 exergy transfer with work (J) 30 CAD 800 45 CAD

400

30 CAD

45 CAD

400 (d ) 3000 thermomechanical exergy (J) 2500 45 CAD 2000 1500 1000 500 0 ( f ) 3000 2500 total exergy (J) 2000 1500 1000 30 CAD 500 0 0 60 120 180 180 120 60 crank angle (CAD) 45 CAD 30 CAD

= 1 s

5 CA

CAD 45 CAD 30

fuel exergy (J)

1500 1000 500 45 CAD 0 180 120 60 0 60 120 180 crank angle (CAD) 30 CAD

Figure 6. Effects of spark timing on exergetic terms. (af ) fZ1.0; f Z0; rZ10; XsZ0; nZ3000 rpm.

the value for K30 CAD in magnitude, approximately 2.5 and 5.4 per cent, respectively. The irreversibilities in gure 6c also attain the minimum values for K30 CAD. The values of the irreversibilities for both K15 and K45 CAD are greater by 0.6 per cent than that of K30 CAD. The variations indicate that the spark timing of K30 CAD is an optimal condition in the tested spark timings. It is also noted that there is no signicant effect of varying spark timing on the
Proc. R. Soc. A (2008)

Downloaded from rspa.royalsocietypublishing.org on August 11, 2013

Exergy analysis of SI engine operation


(a) first and second law efficiencies (%) 50 40 30 20 10 0 0.9 1.0 1.1 equivalance ratio (c) first and second law efficiencies (%) 50 40 30 20 10 0 15 240 210 180 150 120 90 60 30 0 (b)

3125
240 210 180 150 120 90 60 30 0

1500

3000 4500 engine speed ( ) 240 bsfc (kg kW 1 h 1) 210 180 150 120 90 60 30 0

45 30 spark advance (CAD) Figure 7. Effects of (a) equivalence ratio, (b) engine speed and (c) spark timing on the engine performance parameters. Right-hatched bars, hI; left-hatched bars, hII; horizontal bars, bsfc.

irreversibilities. Therefore, ignition timing has to be optimized to give the least AQ and to extract maximum exergy as useful work. When the spark timing is xed earlier, Atm rises in advance, as shown in gure 6d. Conversely, if ignition is delayed, Atm diminishes due to the decrease in pressure and temperature, as shown in table 3. There is almost no effect on the fuel exergy in gure 6e, but Af,ch varied during combustion due to variations in the combustion duration shown in table 3. Atot variations in gure 6f reect the combination of Atm and Af,ch, and Aexh reach the minimum for K30 CAD at the end of the expansion. The values of Aexh for K15 and K45 CAD are greater than the value of K30 CAD, approximately 4.5 and 4.7 per cent, respectively. Figure 7 shows the variations of hI, hII and bsfc with equivalence ratio, engine speed and spark advance, respectively. hI and hII decrease, while bsfc increases with increasing equivalence ratio, as shown in gure 7a. These variations are suitable for the denitions of the efciencies. The leaning of the fuelair mixture results in increases in hI and hII and decreases in bsfc due to the improvement of combustion efciency. However, the enriching of the charge has negative effects on the efciencies and bsfc due to the increasing fuel supplied to the cylinder, as shown in table 3. The variations in the efciencies are as cited by Ferguson (1985). The increments obtained with fZ0.9 and 1.0 are approximately 4.9 and 3.6 per cent in hI and 4.6 and 3.4 per cent in hII, respectively, when compared with fZ1.1. The increments in the bsfc are approximately 4.7 per cent for
Proc. R. Soc. A (2008)

bsfc (kg kW 1 h 1)

Downloaded from rspa.royalsocietypublishing.org on August 11, 2013

3126

. Sezer and A. Bilgin I

fZ1.0 and 21.6 per cent for fZ1.1 in comparison with fZ0.9. The variations of performance parameter with engine speed are given in gure 7b. hI and hII have the maximum values for 3000 rpm and the minimum bsfc obtained at this engine speed. The decrements for 1500 rpm and 4500 rpm are approximately 1.7 and 1.3 per cent in hI and 1.5 and 0.5 per cent in hII, when compared with 3000 rpm. The increments in the bsfc are also approximately 1.4 per cent for 1500 rpm and 10.4 per cent for 4500 rpm, in comparison with 3000 rpm. The effects of spark timing on the engine performance parameters are shown in gure 7c. hI and hII have the maximum values for K30 CAD, and the minimum bsfc values are obtained at that spark timing. The decrements for K15 and K45 CAD are approximately 1 and 1.9 per cent in hI and 0.9 and 1.7 per cent in hII, respectively, when compared with K30 CAD. The increments in the bsfc are also approximately 5 per cent for K15 CAD and 9 per cent for K45 CAD, in comparison with K30 CAD. 5. Conclusions To evaluate SI engine operation from a second law perspective, a QD cycle model was used in this study. The effects of fuelair equivalence ratio, engine speed and spark timing on the exergy transfers, irreversibilities and efciencies have been investigated theoretically. The following general conclusions can be drawn. (i) A parametric exergetic analysis provides a better understanding of interactions between operating conditions, energy conversions and transfer processes, which permits the revelation of the magnitude of work potential lost during the cycle in a more realistic way than the rst law analysis, and points to several possible ways for improving engine performance. (ii) The effects of equivalence ratio, engine speed and spark timing show different trends in manner and magnitude. Exergy transfer with heat is a maximum for the stoichiometric mixture (fZ1.0), while irreversibilities on a per mass of mixture basis increase as the equivalence ratio increases. The exergy transfer with exhaust and work also increases by increasing the equivalence ratio due to increasing fuel exergy in the cylinder. However, the increment in exergy transfer with work for a rich mixture (fZ1.1) is very slight. The peak values of the rst and second law efciencies were obtained with the lean mixture (fZ0.9) and this equivalence ratio gives the minimum bsfc. (iii) Exergy transfer with heat and exhaust increases by increasing the engine speed. However, exergy transfer with work reaches a maximum at the engine speed of 3000 rpm and irreversibilities reach a minimum at this engine speed. The maximum rst and second law efciencies and the minimum bsfc were obtained for 3000 rpm. (iv) Exergy transfer with heat increases as the spark timing increases. However, exergy transfer with work reaches a maximum, while the irreversibilities and exhausted exergy have minimum values at an optimum spark timing of K30 CAD. The best rst and second law efciencies and the minimum bsfc value were obtained at that spark timing.
Proc. R. Soc. A (2008)

Downloaded from rspa.royalsocietypublishing.org on August 11, 2013

Exergy analysis of SI engine operation

3127

References
Agarwal, A., Filipi, Z. S., Assanis, D. N. & Baker, D. M. 1998 Assessment of single- and two-zone turbulence formulations for quasi-dimensional modeling of spark-ignition engine combustion. Combust. Sci. Technol. 136, 1339. (doi:10.1080/00102209808924163) Alasfour, F. N. 1997 Butanola single-cylinder engine study: availability analysis. Appl. Therm. Eng. 17, 537549. (doi:10.1016/S1359-4311(96)00069-5) Alkidas, A. C. 1988 The application of availability and energy balances to a diesel engine. Trans. ASME J. Eng. Gas Turbines Power 110, 462469. Bayraktar, H. & Durgun, O. 2003 Mathematical modeling of spark-ignition engine cycles. Energ. Sources 25, 651 666. (doi:10.1080/00908310303404) Bilgin, A. 2002 Geometric features of the ame propagation process for an SI engine having dual-ignition system. Int. J. Energy Res. 26, 9871000. (doi:10.1002/er.832) Blizard, N. C. & Keck, J. C. 1974 Experimental and theoretical investigation of turbulent burning model for internal combustion engines. SAE paper no. 740191, pp. 846864. Caton, J. A. 1999a Results from the second-law of thermodynamics for a spark-ignition engine using a cycle simulation. In Proc. ASME-ICED Fall Technical Conf., 1016 October, Ann Arbor, MI, pp. 3549. Caton, J. A. 1999b On the destruction of availability (exergy) due to the combustion process-with specic application to the internal combustion engine. Energy 25, 10971117. (doi:10.1016/ S0360-5442(00)00034-7) Caton, J. A. 2000a A review of investigations using the second law of thermodynamics to study internal-combustion engines. In SAE World Congress, 69 March, Detroit, MI, pp. 115. Caton, J. A. 2000b Operation characteristics of a spark-ignition engine using the second law of thermodynamics: effects of speed and load. In SAE World Congress, 69 March, Detroit, MI, pp. 117. Caton, J. A. 2001 A multiple-zone cycle simulation for spark-ignition engines: thermodynamic details. In Proc. ASME-ICED Fall Technical Conf., 2326 September, Argonne, IL, pp. 118. Caton, J. A. 2002 A cycle simulation including the second law of thermodynamics for a sparkignition engine: implications of the use of multiple-zones for combustion. In SAE World Congress, 47 March Detroit, MI, pp. 119. Cengel, Y. A. & Boles, M. A. 1994 Thermodynamics, an engineering approach, 2nd edn. New York, NY: McGraw-Hill. Dai, W., Newman, C. E. & Davis, G. C. 1996 Predictions of in-cylinder tumble ow and combustion in SI engines with a quasi-dimensional model. SAE paper no. 961962, pp. 157168. Ferguson, C. R. 1985 Internal combustion engine, applied thermosciences. New York, NY: Wiley. Flynn, P. F., Hoag, K. L., Kamel, M. M. & Primus, R. J. 1984 A new perspective on diesel engine evaluation based on second law analysis. SAE paper no. 840032, pp. 198211. Gallo, W. L. R. & Milanez, L. F. 1992 Exergetic analysis of ethanol and gasoline fueled engines. SAE paper no. 920809, pp. 907915. Georgios, Z. 2005 Mathematical and numerical modeling of ow and combustion processes in a spark ignition engine. PhD thesis, Department of Applied Mathematics, University of Crete, Heraklion, Greece. . 1984 Correlations of laminar combustion data for alternative S.I. engine fuels. SAE Gu lder, O paper no. 841000, pp. 123. Kanoglu, M., Isik, S. K. & Abusoglu, A. 2005 Performance characteristics of a diesel engine power plant. Energy Convers. Manage. 46, 16921702. (doi:10.1016/j.enconman.2004.10.005) Keck, J. C. 1982 Turbulent ame structure and speed in spark-ignition engines. In 19th Int. Symp. Combustion, pp. 14511466. Haifa, Israel: The Combustion Institute. Kopac, M. & Kokturk, L. 2005 Determination of optimum speed of an internal combustion engine by exergy analysis. Int. J. Exergy 2, 4054. (doi:10.1504/IJEX.2005.006432) Kotas, T. J. 1995 The exergy method of thermal plant analysis. Malabar, FL: Krieger Publishing.
Proc. R. Soc. A (2008)

Downloaded from rspa.royalsocietypublishing.org on August 11, 2013

3128

. Sezer and A. Bilgin I

Kumar, S. V., Minkowycz, W. J. & Patel, K. S. 1989 Thermodynamic cycle simulation of the diesel cycle: exergy as a second law analysis parameter. Int. Commun. Heat Mass Transf. 16, 335 346. (doi:10.1016/0735-1933(89)90082-1) Moran, M. J. & Shapiro, H. N. 2000 Fundamentals of engineering thermodynamic, pp. 265315, 3rd edn. New York, NY: Wiley. Parlak, A., Yasar, H. & Eldogan, O. 2005 The effect of thermal barrier coating on a turbo-charged diesel engine performance and exergy potential of the exhaust gas. Energy Convers. Manage. 46, 489499. (doi:10.1016/j.enconman.2004.03.006) Rakopoulos, C. D. 1993 Evaluation of a spark ignition engine cycle using rst and second law analysis techniques. Energy Convers. Manage. 34, 1299 1314. (doi:10.1016/0196-8904(93) 90126-U) Rakopoulos, C. D. & Giakoumis, E. G. 1997 Simulation and exergy analysis of transient dieselengine operation. Energy 22, 875885. (doi:10.1016/S0360-5442(97)00017-0) Rakopoulos, C. D. & Giakoumis, E. G. 2006 Second law analyses applied to internal combustion engines operation. Prog. Energy Combust. Sci. 32, 247. (doi:10.1016/j.pecs.2005.10.001) Rezac, P. & Metghalchi, H. 2004 A brief note on the historical evolution and present state of exergy analysis. Int. J. Exergy 1, 426437. (doi:10.1504/IJEX.2004.005787) Sezer, I. 2008 Application of exergy analysis to spark ignition engine cycle. PhD dissertation, Karadeniz Technical University, Trabzon, Turkey. Shapiro, H. N. & Van Gerpen, J. H. 1989 Two zone combustion models for second law analysis of internal combustion engines. SAE paper no. 890823, pp. 14081422. Tabaczynski, R. J., Ferguson, C. R. & Radhakrishnan, K. 1977 A turbulent entrainment model for spark-ignition combustion. SAE paper no. 770647, pp. 24142432. Tabaczynski, R. J., Trinker, F. H. & Sahnnon, B. A. S. 1980 Further renement of a turbulent ame propagation model for spark-ignition engines. Combust. Flame 39, 111121. (doi:10.1016/ 0010-2180(80)90011-5) Tallio, K. V. 1998 A multiuid turbulent entrainment combustion model. PhD dissertation, Drexel University, Philadelphia, USA. Van Gerpen, J. H. & Shapiro, H. N. 1990 Second law analysis of diesel engine combustion. Trans. ASME J. Eng. Gas Turbines Power 112, 129 137. (doi:10.1115/1.2906467) Velasquez, J. A. & Milanez, L. F. 1994 Analysis of the irreversibilities in diesel engines. SAE paper no. 40673, pp. 10601068. Wengang, D. 1995 Quasi-dimensional modeling of spark ignition engine. PhD dissertation, University of Texas, Austin, USA. Zhang, S. 2002 The second law analysis of a spark ignition engine fueled with compressed natural gas. MS dissertation, University of Windsor, Ontario, Canada.

Proc. R. Soc. A (2008)

You might also like