You are on page 1of 13

Abou Al-Sood et al.

International Journal of Energy and Environmental Engineering 2012, 3:13


http://www.journal-ijeee.com/content/3/1/13

ORIGINAL RESEARCH Open Access

Rapid thermodynamic simulation model for


optimum performance of a four-stroke,
direct-injection, and variable-compression-ratio
diesel engine
Maher M Abou Al-Sood*, Mahmoud Ahmed and Yousef M Abdel-Rahim

Abstract
A thermodynamic simulation model for the performance of a four-stroke, direct-injection diesel engine is
developed. The simulation model includes detailed sub-models for fuel burning rate, combustion products,
thermodynamic properties of working fluid, heat transfer, fluid flow, and both soot and oxides of nitrogen (NOx)
formation mechanisms. To validate the model, comparisons between experimental and predicted results for
different engines, operating under different conditions, were conducted. The comparisons show that there is a
good concurrence between measured and predicted values. An optimization analysis is conducted for seeking an
optimum variation of compression ratio to achieve pre-set objective targets such as constant minimum
brake-specific fuel consumption and constant maximum torque. The optimization analysis is performed under the
constraint that the maximum pressure and temperature inside the cylinder do not exceed the maximum allowable
pressure and temperature of the conventional engine (constant compression ratio).The varying compression ratio is
optimized with each of the previous conditions separately. Results indicated that varying the compression ratio to
achieve previous targets leads to saving fuel consumption, higher brake efficiency and power, and reduction in
soot emission from the engine. Also, an increase in NOx is noticed at low speed. This drawback is considerable and
can be overcome by reducing the operation speed range.
Keywords: Thermodynamic, Combustion, Compression ratio, Optimum performance, Diesel engine

Background limited set of experimental data, are preferred due to


Thermodynamic models (zero-dimensional) and turbu- their simplicity, of being less time-consuming, and their
lent models (multi-dimensional) are the two types of relatively accurate results [7-10].
models that have been used in internal combustion en- Requirements to provide environmentally friendly
gine simulation modeling. Nowadays, trends in combus- vehicles (i.e., low fuel consumption with less emission)
tion engine simulations are towards the development of have resulted in using the most complicated and intel-
comprehensive multi-dimensional models that accur- lectual means of internal combustion engine develop-
ately describe the performance of engines at a very high ment. Nowadays, new ideas, which could not be
level of details. However, these models need a precise ex- discussed two decades ago, are being considered by
perimental input and substantial computational power, automotive manufacturers. In particular, many leading
which make the process significantly complicated and automotive companies have approached practically the
time-consuming [1-6]. On the other hand, zero- very complicated design ideas with different aspects of
dimensional models, which are mainly based on energy diesel/petrol engine design. These aspects have been
conservation (first law of thermodynamics) with a under extensive theoretical and experimental investiga-
tions. The most important aspect of design is aimed to
vary the engine compression ratio (rc) depending on
* Correspondence: m_aboualsood@hotmail.com
Department of Mechanical Engineering, Assiut University, Assiut 71516, Egypt
load, speed, or both. Several trials have been done in

© 2012 Abou Al-Sood et al.; licensee Spinger. This is an Open Access article distributed under the terms of the Creative
Commons Attribution License (http://creativecommons.org/licenses/by/2.0), which permits unrestricted use, distribution, and
reproduction in any medium, provided the original work is properly cited.
Abou Al-Sood et al. International Journal of Energy and Environmental Engineering 2012, 3:13 Page 2 of 13
http://www.journal-ijeee.com/content/3/1/13

that respect with extensive design, experimentation, and optimized for a full range of driving conditions, such as
measurements [11-19]. All attempts to change the rc are acceleration, speed, and load. Such optimized perform-
achieved by one or more of the following concepts [11]: ance was obtained by incorporating some means to sup-
(1) moving the cylinder head, (2) variation of combus- press the increase in emission rate to be applicable in
tion chamber volume, (3) variation of piston deck the automotive field due to the strict anti-pollution laws.
height, (4) modification of connecting rod geometry The effect of varying compression ratio and other
(usually by means of some intermediate member), (5) parameters such as injection timing, engine load, and
moving the crankpin within the crankshaft (effectively fuel on the performance of diesel engine was recently
varying the stroke), and (6) moving the crankshaft axis. studied experimentally [17-20]. Raheman and Ghadge
Benefits and challenges of variable-compression-ratio [17] studied the performance of a Ricardo EG engine
(VCR) spark-ignition engines had been illustrated, exam- using biofuel and its blend with high-speed diesel fuel at
ined, and critically reviewed by Roberts [11]. Also, the varying compression ratio, injection timing, and engine
implications for volume manufacture and the strategy load. They concluded that the performance of the engine
for VCR implementation in order to produce the max- reduced with the increase in the percentage of biofuel in
imum benefit had been discussed. the blend above 20% compared to that of conventional
Ladommatos and Balian [12] studied the effects of an engine. This reduction in engine performance can be
increased clearance volume on the performance of a overcome by increasing the compression ratio and injec-
direct-injection diesel engine. This study was conducted tion timing. Jindal et al. [18] investigated the effect of
to overcome problems with combustion at standard rc compression ratio and injection pressure on the per-
for high-speed, direct-injection (DI) diesel engine. Modi- formance of DI diesel engine running on Jatropha me-
fications were done in the injection system and operat- thyl ester. They demonstrated that an increase in
ing conditions to improve the power output and fuel compression ratio associated with an increase in injec-
consumption. Sobotowski et al. [13] modified a proto- tion pressure improves the performance of the engine in
type DI diesel engine to incorporate BP Oil's VCR en- terms of brake-specific fuel consumption, brake thermal
gine concept. Their objective was achieved by altering efficiency, and emissions (i.e., hydrocarbon, NOx, and
the phase relation between two pistons linked to separ- smoke opacity). They concluded that for fuelling the en-
ate crankshafts and operating in two cylinders arranged gine with bio-diesel, one should go for higher compres-
in an opposed engine or interconnected through a trans- sion ratio associated with higher injection pressure.
fer port. According to their concept, a novel crankshaft Muralidharan et al. [19] estimated the performance,
phasing mechanism was employed to achieve rc vari- emission, and combustion characteristics of a single-cy-
ation with a number of significant advantages over other linder, four-stroke, VCR, multi-fuel engine fuelled with
VCR engine designs. These advantages include (a) mech- waste cooking oil methyl ester and its blends with stand-
anical simplicity and compact design, (b) capability to ard diesel. The performance parameters include brake
externally control rc during engine operation, (c) auto- thermal efficiency, brake-specific fuel consumption,
matic variation of valve timing, and (d) compatibility brake power, indicated mean effective pressure, mechan-
with current production technology. ical efficiency, and exhaust gas temperature. They com-
Rychter et al. [14] evaluated the concept of varying rc pared and analyzed their results with standard diesel,
to give different expansion and compression ratios by and a considerable improvement in the performance
means of a theoretical simulation of a turbocharged parameters as well as exhaust emissions was observed
diesel engine. They varied the ratio of connecting rod especially at full load. Gandure and Ketlogetswe [20]
length to crank throw to study the effect of rc on engine carried out their work to compare the performance of
performance at a fixed load. The principal benefits, at variable compression engine fuelled with native mar-
3/4 load and rc of 20, were a reduction in fuel consump- ula seed oil and petrodiesel fuels. The results indi-
tion by about 2% and a reduction in ignition delay that cated that engine performance when powered with
leads to an estimated 6-dB reduction in combustion marula oil as fuel was very close to that when pow-
noise. Variable-stroke engine mechanism patented by ered with petrodiesel and at an 80% load, the com-
Freudenstein and Maki [15] was used by Yamin and pression ratio of 16:1 yields optimum performance in
Dado [16] as it gives availability to change the stroke terms of engine torque and brake power for both pet-
length and the corresponding rc in the range of 6.82 to rodiesel and marula oil fuels.
10. Their results showed a significant improvement in The purpose of this study is to develop a simple, rapid,
the engine's power and reduction in the fuel consump- and accurate simulation model based on thermodynamic
tion, in addition to an increase in carbon monoxide and analysis without the need for a great deal of computa-
nitric oxides compared with the actual engine. Finally, tional power or knowledge of precise engine geometrical
they concluded that the engine performance can be data to predict the range of compression ratio
Abou Al-Sood et al. International Journal of Energy and Environmental Engineering 2012, 3:13 Page 3 of 13
http://www.journal-ijeee.com/content/3/1/13

corresponding to the optimum engine performance In the present work, the detailed sub-models are for-
(minimum brake-specific fuel consumption (bsfc) and mulated as follows:
maximum brake torque) without an increase in the max-
imum cylinder pressure and temperature of the conven- Engine geometry
tional engine. The developed model is based on mass The rate of change of the volume with respect to time is
and energy conservation laws, equation of state, heat given as
transfer between the working fluid and cylinder walls,  
rc  1 1 þ cosθ
fuel energy release rate, and engine friction losses. dot; V ¼ Vc sinθ pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
ffi dot; θ : ð1Þ
Single-zone combustion is assumed, and thermodynamic 2 λ2  sin2 θ
properties are calculated from chemical equilibrium
principles for six and/or eleven species. A stepwise solu- Combustion product model
tion is carried out for the varying compression ratio and For temperature <1,600 K, the equilibrium composition
the engine speed over its entire range (i.e., 1,500 to of gases produced by the combustion of a general hydro-
2,800 rpm). In calculations, the temperature and pres- carbon fuel with air (having a general fuel-air molecule
sure cannot exceed the maximum allowable temperature in the form CwHxOyNz) is calculated based on the as-
and pressure of the conventional engine, i.e., constant sumption that only gaseous phases are considered. The
compression ratio, which are 2,100 K and 10.7 MPa, general combustion equation of the fuel in air can be
respectively. written as

Methods Cw Hx Oy Nz ! n1 CO2 þ n2 H2 O þ n3 CO þ n4 H2
Mathematical model þ n5 O2 þ n6 N2 : ð2Þ
A thermodynamic analysis of a four-stroke, direct-
For temperature ≥1,600 K, 11 combustion products
injection diesel engine is conducted. It is based on a
are assumed. The chemical reaction can be written as
single-zone combustion modeling with a rapid premixed
burning phase followed by a slower mixing-controlled Cw Hx Oy Nz ! n1 CO2 þ n2 H2 O þ n3 CO þ n4 H2
burning phase. Although a single-zone model is an old þ n5 O2 þ n6 N2 þ n7 OH þ n8 H
one, it was selected in this study for the sake of its sim- þ n9 NO þ n10 N þ n11 O: ð3Þ
plicity and of being less time-consuming for the pro-
gram execution. Works with two- and multi-zone Obtaining the moles of the 11 combustion products,
models are in progress. The analysis includes detailed 11 equations are needed (i.e., four equations from the
sub-models representing different engine aspects inte- mass balance of C, H, O, and N in Equation 3, and seven
grated in the main engine simulation model. These sub- equations from the chemical equilibrium). These equa-
models are engine geometry, combustion products, tions have been solved by the procedure mentioned in
thermodynamic properties, engine friction and heat [21].
transfer, mass exchange, combustion rate, temperature-
pressure, and emission mechanisms. The analysis Thermodynamic properties of cylinder content model
presented in this article is based on the following During each thermal cycle, the cylinder contains, in gen-
assumptions: eral, air, fuel vapor, and combustion products with differ-
ent proportions depending on the engine conditions
1. The content of the cylinder is a homogeneous gas under consideration. The specific heat under constant
mixture of air (79% N2 and 21% O2 by volume), pressure, enthalpy, and entropy of the cylinder content
where fuel vapor, combustion products, and its are given by
thermodynamic properties are calculated using ideal
X
zz X
zz
gas law with temperature-dependent specific heats. cp ¼ Xi C ¼
p ; h i ;
Xi h ð4Þ
i
2. Pressure and temperature inside the cylinder are i¼1 i¼1
uniform and vary with the crank angle.
Xzz
3. The gas motion inside the cylinder (created by where Xi ¼ ni = ni , and zz (number of different gases)
pressure gradients due to piston motion) is i¼1
neglected. is equal to 6 and 11 for T < 1,600 K, and T ≥ 1,600 K,
4. A single-zone combustion process starts with a rapid respectively.
premixed burning phase, followed by a slower
mixing-controlled burning phase. Engine friction model
5. Temperatures of the cylinder head, cylinder walls, Empirical correlations of losses are in terms of compres-
and piston crown are assigned constant values. sion ratio, engine speed, bore, stroke, intake and exhaust
Abou Al-Sood et al. International Journal of Energy and Environmental Engineering 2012, 3:13 Page 4 of 13
http://www.journal-ijeee.com/content/3/1/13

manifold pressures, and valve geometry [22] are used in used with assumed specified values for piston crown, cy-
this work. These correlations were concluded to the fol- linder head, and cylinder wall temperatures.
lowing friction categories as follows:
1. Piston losses: Mass exchange model
  Air intake and exhaust exchange. The mass flow rates
ðfmepÞp ¼ 6:2nr þ 0:606 rc þ 1:254 rc 1:370:0238Sp
þ 9:3Sp1:03 through the valves are usually described by the equation
ð5Þ for compressible, one-dimensional, isentropic flow. The
real flow effects are included by means of an experimen-
2. Blowby losses:
tally determined discharge coefficient. In the present simu-
ðfmepÞb ¼ 11:86rc0:4  ð3:38 þ 0:103rc Þ þ 6:89½N=1; 0001:185 lation, the discharge coefficient is taken as reported by
 1=3
ð6Þ Campbell [29] and is equal to sin π ðθ  θvo Þ=Δθopen .
The gas flow rate is related to the upstream stagna-
3. Exhaust and inlet system throttling losses: tion pressure, stagnation temperature, and static pres-
ðfmepÞei ¼ Pim þ Pem =2:75 ð7Þ sure just downstream of the flow restriction. For the
flow into the cylinder through intake valve, the up-
4. Crankcase mechanical losses: stream stagnation pressure and temperature are the
intake manifold pressure and temperature (pim, Tim),
 
dN respectively, and the downstream static pressure is
ðfmepÞc ¼ 12:122 the cylinder pressure (p). For the flow out of the cy-
 000L
1;    
4N 1:75
niv Div linder through exhaust valve, the upstream stagnation
þ 0:07 30  pressure and temperature are the cylinder pressure
1; 000 d2 L
 1:5 and temperature (p, T), respectively, and the down-
N
þ 2:688 ð8Þ stream static pressure is the exhaust manifold pres-
1; 000
sure (pem) [29,32]. The mass flow rates entering and
5. Combustion chamber and valve pumping losses: leaving the cylinder are defined as follows [33]:

k1 k

 1:298 For subsonic inflow where ppin < kþ1 2


;
ðfmepÞcc ¼ 8:96ðN=1; 000Þ1:7  2:98 Vd =niv nc D2iv
vffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
!
ð9Þ u  k2  k1
u 2k p p k
dot; min ¼ Ain pim t 1 :
The intake and exhaust manifold gage pressures ðk  1ÞRTim pim pim
(pim, pem) in Equation 7 are taken as reported by
ð12Þ
Eichelberg [23] as pim ¼ 0; pem ¼ 0:062ðN=1; 000Þ2 .
The overall engine frictional power losses are obtained
k1
k
p
For sonic and supersonic inflow where ≥ 2
;
by pin kþ1
vffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
!
u
X u k   2 kþ1
5
t
k1
ðfmepÞoverall ¼ Vd ðN=120Þnc ðfmepÞi : ð10Þ dot; min ¼ Ain pim : ð13Þ
i¼1 RTim kþ1


k1
k
p
Engine heat transfer model For subsonic outflow where pem < 2
kþ1 ;
Eichelberg [23] derived a simple equation for calcula- sffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
ting the instantaneous heat flux out of the engine   k2  em k1 
2k pem k
dot; qw (in Watts per square meter) in terms of cylinder dot; mex ¼ Aex p 1  pem p Þ :
ðk  1ÞRT p
temperature, pressure, and piston mean velocity. Rako-
poulos and Hountalas [24] modified this equation ð14Þ
into the form
k1
k
pem
For sonic and supersonic outflow where p ≥ 2
kþ1 ;
dot; qw ¼ 1:99ðSpÞ1=3 ðpT Þ1=2 ðT  T wÞ: ð11Þ vffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
!ffi
u   kþ1
u k 2
dot; mex ¼ Aex pt
k1
: ð15Þ
Radiation heat transfer was neglected in some studies RT kþ1
(e.g., [25-28]), but it had been considered in others (e.g.,
[29-31]). In the present simulation, the Eichelberg model Mass of injected fuel. The mass of fuel injected into
[23] as modified by Rakopoulos and Hountalas [24] is the cylinder mf (relative to total injected mass mft) is
Abou Al-Sood et al. International Journal of Energy and Environmental Engineering 2012, 3:13 Page 5 of 13
http://www.journal-ijeee.com/content/3/1/13

written (according to Zelenik [34] and Gu and Ball [35]) Equation 18 as follows:
in terms of crank angle (θ) as follows:
    Temperature-pressure relationships
mf 1 π ðθ  θo Þ The changes of pressure and temperature with respect
¼ 1  cos : ð16Þ
mft 2 Δθinjection to time are defined as follows:
 
ρ dot; V dot; T @ρ dot; φ @ρ dot; m
Differentiation of the above equation with respect to dot; P ¼    þ ;
@ρ=@P V ρ @T ρ @φ m
time leads to the following equation:
      ð21Þ
dot; mf π π ðθ  θo Þ
¼ sin dot; θ: (   !)
mft 2Δθinjection Δθinjection dot; T ¼
B∘

dot; V h dot; m C ∘
þ 1 ∘  ∘ dot; φ þ
1 X
dot; mj hj þ dot; Qw ;
A∘ V B m B mB ∘

ð17Þ j

ð22Þ
Combustion rate model @h 

where A∘ ¼ @T þ ðð@ρ=@T Þ 1 @h ∘ 1ρð@h=@P Þ
@ρ=@PÞ ρ  @P ; B ¼ ð@ρ=@P Þ ;
Watson et al. [36] developed equations for fuel energy

release appropriate for diesel engine simulations. In their C ∘ ¼ @φ


@h
þ ðð@ρ=@P
@ρ=@φÞ 1 @h
Þ ρ  @P :
development, the combustion process starts from a rapid
premixed burning phase (represented by function f1),
Soot mechanism
followed by a slower mixing-controlled burning phase
Hiroyasu and Kadota [39] developed a simple soot
(represented by function f2), with both functions empir-
model to predict the production rate of soot mass (
ically linked to the duration of ignition delay ( ~τ id ) and
dot; ms ) using a single-step competition between the
the duration of combustion (Δθcomb). According to their
soot mass formation rate dot; msf and the soot mass oxi-
model, the burned mass of fuel at any angle (θ) related
dation rate dot; mso according to dot; ms ¼ dot; msf 
to total injected mass (mft) is

dot; mso . Morel and Keribar [40] and Wahiduzzaman
mfb θ^ =mft ¼ βf1 þ ð1  βÞf2 ; ð18Þ et al. [41] suggested models for soot formation and their
oxidation in diesel engines. In their models, the time

k k  rates of soot formation and oxidation, respectively, are


where f1 ¼ 1  1  θ^ , θ^ ¼ θ  θig =Δθcomb , and given by
0:26
β ¼ 1  0:92φ0:37 τ id :  
3; 000
The ignition delay (~τ id ) is calculated in degrees of crank dot; msf ¼ 0:3 mfb exp  dot; θ; ð23Þ
Taf
angle by empirical correlations mentioned by Heywood
[32] and Mehta et al. [37]. These correlations are in terms pffiffiffiffiffiffiffi!
0:4ms exp  10;000 pO2
of intake manifold temperature (Tim, in Kelvin) and pres- dot; mso ¼ T
dot; θ; ð24Þ
6ρs ds N
sure (pim, in bars), compression ratio (rc), fuel cetane num-
ber (CN), and mean piston velocity (Sp, in meters per where mfb is the mass of fuel burned, Taf is the adiabatic
second) as follows: flame temperature at an equivalence ratio of 1.1, ds (the
  soot particle diameter) = 0.022 × 10−6 m, ρs (soot density) =
 618; 840
~τ id ¼ 0:36 þ 0:622Sp exp 1,800 kg/m3 [42].
CN þ 25
   0:63 
1 1 21:2 NOx mechanism
   :
RTim rck1 17; 190 Pim rck  12:4 The extended Zeldovich mechanism for NOx formation
ð19Þ rate is formed as follows [29]:
The ignition delay period can be expressed in millise- !
d 1  ½NO2 =K12 ½O2 e ½N2 e
conds according to τ id ¼ ½~τ id =0:006N . Values of the em- ½NO ¼ 2k1f ½Oe ½N2 e  ;
dt 1 þ k1b ½NO= k2f ½O2 e þ k3f ½OH  e
pirical coefficients k1, k2, k3, and k4 are reported by
Heywood [32] and Rossini et al. [38] as follows: k1 ¼ ð25Þ
2 þ 1:25  108 ðτ id =N Þ2:4 , k2 ¼ 5000, k3 ¼ ð14:2=φ0:644 Þ,
and k4 ¼ 0:79k30:25 . The fuel burning rate with respect to where K12 ¼ ðk1f =k1b Þ=ðk2f =k2b Þ . The constants above

the crank angle (θ) can be obtained by differentiating are taken as follows: k1f ¼ 7:6  1013 exp  38;000
T , k2f ¼

dmfb 

k 1 ðk 2 1Þ ðk 1 1Þ ðk 4 1Þ


¼ mft βk 1 k 2 1  θ^ θ^ þ k 3 k 4 ð1  βÞθ^ exp k 3 θ^


k4

: ð20Þ
Δθcomb
Abou Al-Sood et al. International Journal of Energy and Environmental Engineering 2012, 3:13 Page 6 of 13
http://www.journal-ijeee.com/content/3/1/13


6:4  109 T exp  3;150
T ,k3f ¼ 4:1  1013 , k1b ¼ 1:6  1013, Results and discussions
 Simulation results and its validation
and k2b ¼ 1:5  109 T exp  19;500
T .
The numerical simulation results are performed for a con-
Solution procedure
ventional DI diesel engine with constant compression ratio
The governing equations (Equations 1 to 25) of the com- (specifications and operating conditions are given in
prehensive model subjected to the assumptions previ- Table 1). Samples of these results represent the engine
ously mentioned in the mathematical model are solved thermodynamic cycle, the temperature and pressure vari-
for different processes of the engine cycle. The fourth- ation throughout the cycle, the rates of mass entering and
order Runge-Kutta method is used to simulate the com- leaving the cylinder, and the in-cylinder NOx and soot his-
prehensive model equations at the prescribed initial con- tory along the cycle which are displayed in Figures 1, 2,
ditions. The engine cycle starts from the moment the and 3, respectively.
exhaust valve closes. The solution is carried out with In order to validate the present developed model,
end results of each process taken as the starting condi- comparisons between the numerical results and the
tions for the following process and end values of the available measurements are conducted. Two sets of ex-
completed thermal cycle taken as the starting values for perimental data reported by Han and Reitz [3] and
the subsequent cycle. The solution is iterated and con- Wahiduzzaman et al. [41] are used to validate the devel-
sidered satisfactory when the end values of the oped model as shown in Figure 4. Figure 4 shows the
temperature and pressure of the cycle are within comparison between the predicted and measured pres-
±0.001% from the starting values of the temperature and sures and heat release versus the crank angle. Based on
pressure of the same cycle. Inputs to the code are the this figure, one can observe that there is a good concur-
engine specifications, operating conditions, and fuel data rence between measured and predicted results for differ-
listed in Table 1. The numerical output results are the ent engines under different operating conditions. More
instantaneous pressure, temperature, volume, heat trans- details about the simulation results and validation can
fer, work, mass exchange, fuel heat release, and engine be found in a previous work of the authors [43].
performance variables such as bsfc, indicated mean ef-
Variation of the compression ratio
fective pressure, brake mean effective pressure, indicated
power, brake power, and volumetric, indicated, brake, The numerical simulation results are used to determine
and mechanical efficiencies. the range of rc corresponding to the optimum engine
performance for two cases: (1) minimum bsfc and (2)
maximum brake torque. This was computed using
optimization techniques under the restriction of max-
imum allowable pressure and temperature. The numer-
Table 1 Engine specifications and operating conditions
ical results of the optimization process, for engine
Specifications/operating conditions Inputs/values
Engine type HELWAAN M114
Crank Angle (Degree)
Cylinder bore 112 mm 200 267 334 401 468 535 602
10 2000
Stroke 115 mm
Number of cylinders 4
Intake valve open angle 9° BTDC 8 1600
Intake valve close angle 43° ABDC P-V
Temperature
Temperature (K)
Pressure (MPa)

Exhaust valve open angle 40° BBDC


6 1200
Exhaust valve close angle 12° ATDC
Start of injection 22° BTDC
4 800
Number of injector nozzles 1
Diameter of injector nozzles 0.45 mm
Pressure
Conventional compression ratio 16.4 2 400
Engine speed range 1,500 to 2,800 rpm
Overall equivalence ratio 0.47
0 0
Fuel type C10.8 H18.7 0 0.0002 0.0004 0.0006 0.0008 0.001 0.0012
Volume (m 3)
Lower latent heat of fuel 42 MJ/kg
Figure 1 Pressure-volume diagram along engine cycle; cylinder
BTDC before top dead center, ABDC after bottom dead center, BBDC before pressure and temperature versus crank angle at 2,100 rpm.
bottom dead center, ATDC after top dead center.
Abou Al-Sood et al. International Journal of Energy and Environmental Engineering 2012, 3:13 Page 7 of 13
http://www.journal-ijeee.com/content/3/1/13

0.035 0.07 Crank angle for pressure (degree)


Mass in
Mass out 200 275 350 425 500
400 16
0.03 0.06
Predicted HR
Measured HR [3]
0.025 0.05

Mass out (gm/degree)


Predicted Pressure
Mass in (gm/degree)

Measured pressure [41]

Heat release rate (J/degree)


300 12
0.02 0.04

Pressure (MPa)
0.015 0.03
200 8
0.01 0.02

0.005 0.01
100 4

0 0
0 180 360 540 720
Crank angle (degree)

Figure 2 Instantaneous mass flow rates through intake and 0 0


340 360 380 400 420
exhaust valves versus crank angle at 2,100 rpm. Crank angle for heat release (degree)

Figure 4 Predicted and measured pressures and heat release


specifications previously mentioned as shown in Table 1, versus crank angle.
are presented in Figures 5,6,7,8,9, Figures 10,11,12,13,14.
It is worth mentioning that for all figures, the solid lines 238.5 and 261 g/kW.h passing through a minimum value
represent the performance of optimum variable rc, while of 230 g/kW.h at an engine speed of about 1,900 rpm. In
dashed lines represent the performance of conventional addition, for the variable-compression-ratio engine, the
constant rc engine. bsfc is kept constant at its minimum value of 230 g/kW.
h by varying the compression ratio between 16.4 and
Case 1: minimum brake-specific fuel consumption 17.8. The shown variation trend in compression ratio
Figures 5, 6, 7, 8, and 9 show the results at minimum emphasizes that the variation in compression ratio
bsfc. Figure 5 shows the comparison between the oper- should be considered to decrease the bsfc to its mini-
ation under constant compression ratio (rc = 16.4) mum possible value. Since the increase in compression
and the hypothetical operation under varying compres- ratio is accompanied by an increase in both cylinder
sion ratio to achieve working with minimum bsfc (230 pressure and brake power, consequently, a decrease in
g/kW.h) as possible. The figure shows that the bsfc for bsfc is achieved. The percentage reduction in bsfc ranges
the constant-compression-ratio engine changes between

16 4
Soot
NOx

12 3
Mass out (gm/degree)
Mass in (gm/degree)

8 2

4 1

0 0
340 380 420 460 500
Crank angle (degree)

Figure 3 Soot and NOx inside the cylinder of the engine versus Figure 5 bsfc and compression ratio versus speed for constant-
crank angle at 2,100 rpm. and variable-compression-ratio engines.
Abou Al-Sood et al. International Journal of Energy and Environmental Engineering 2012, 3:13 Page 8 of 13
http://www.journal-ijeee.com/content/3/1/13

6 90

vol
5 89

Delay Period (degree), RGF (%)

Vol. Efficiency (%)


4 88
Delay Period

RGF
3 87

2 86

1 85
1500 1760 2020 2280 2540 2800
Figure 6 Maximum pressure and temperature versus speed for Engine Speed (rpm)
constant- and variable-compression-ratio engines.
Figure 8 Delay period, RGF, and volumetric efficiency versus
speed for constant- and variable-rc engines for minimum bsfc.
between 3.7% at 1,500 rpm and 4.4% at 2,460 rpm pas-
sing through 0% at 1,900 and 2,800 rpm. Changing the rpm and 40 K at 2,460 rpm passing through 0 K at both
compression ratio for an operating engine is desirable 1,900 and 2,800 rpm. These increases still do not exceed
and could be attained. Its effect on maximum cycle pres- the engine design's maximum pressure and temperature
sure and temperature must be studied as they represent of the conventional engine. Changes in brake power and
important parameters in the design of engine's different torque with speed for the variable compression ratio at
parts. constant minimum bsfc are shown in Figure 7. It shows
Figure 6 shows the increase in maximum pressure and an increase in brake power by about 1.67 kW at 1,500
temperature when operating under variable compression rpm and 1.6 kW at 2,460 rpm. Moreover, the brake
ratio compared to the operation under constant com- torque increases by 9 N.m at 1,500 rpm and 8 N.m at
pression ratio. Compared to the case of a constant com- 2,460 rpm.
pression ratio, the increase in maximum cycle pressure Figure 8 shows that varying the compression ratio
is about 0.53 MPa at 1,500 rpm and 0.55 MPa at 2,460 results in a decrease in the delay period, volumetric effi-
rpm passing through 0 MPa at both 1,900 and 2,800 ciency, and residual gas fraction (RGF) over the engine
rpm with variable compression ratio. The noticeable in- speed range (1,500 to 2,460 rpm). It is seen that this
crease in maximum temperature is about 72 K at 1,500

Figure 7 Brake power and torque versus speed for constant- Figure 9 Soot and NOx emission versus speed for constant- and
and variable-compression-ratio engines. variable-rc engines for minimum bsfc.
Abou Al-Sood et al. International Journal of Energy and Environmental Engineering 2012, 3:13 Page 9 of 13
http://www.journal-ijeee.com/content/3/1/13

Figure 10 Brake torque and power versus speed for constant- Figure 12 bsfc and thermal efficiency versus speed for
and variable-compression-ratio engines. constant- and variable-rc engines.

extreme conditions, when auto-ignition of most of the


decrease changes between 0.2° of crank angle at 1,500 injected fuel occurs, an audible knocking sound is
rpm and 0.2° of crank angle at 2,460 rpm. The decrease noticed, often referred to as a ‘diesel knock.’ Also, if long
occurs due to the fact that the increase in compression ignition delay occurs sufficiently late and reaches the ex-
ratio results in an increase of the air pressure and pansion process, the burning process may be quenched.
temperature at the end of the compression process This leads to incomplete combustion, reduced power
which are inversely proportional to ignition delay. The output, and poor fuel-power conversion efficiency. Also,
decrease in the ignition delay is desirable as it causes ig- the figure shows that the volumetric efficiency and RGF
nition to occur before greater amounts of fuel is injected are reduced by about 4.3% and 0.3%, respectively, at a
in the cylinder. Consequently, the rates of heat release speed of 1,500 rpm. These reductions continue to de-
and pressure rise are then controlled primarily by the crease until reaching 0% at 1,900 rpm. These reductions,
rate of injection and air-fuel mixing, resulting in a then, increase again until reaching about 4.3% and 0.1%
smoother engine operation. Long ignition delay allows at 2,460 rpm. This variation was attributed to the in-
for more fuel to be injected before auto-ignition occurs, crease in compression ratio. It means that a decrease in
which results in a very rapid pressure rise as the com- clearance volume leads to a decrease in charging-up the
bustion starts and reaches high-peak pressures. Under
6 90

vol
5 89
Delay Period (degree), RGF (%)

Vol. Efficiency (%)

4 88
Delay Period

RGF
3 87

2 86

1 85
1500 1760 2020 2280 2540 2800
Engine Speed (rpm)

Figure 13 Delay period, RGF, and volumetric efficiency versus


Figure 11 Maximum cylinder pressure and temperature versus speed for constant- and variable-rc engines for maximum
speed for constant- and variable-compression-ratio engines. torque.
Abou Al-Sood et al. International Journal of Energy and Environmental Engineering 2012, 3:13 Page 10 of 13
http://www.journal-ijeee.com/content/3/1/13

of fuel (rc = constant) at 1,500 rpm. Beyond 1,900 rpm,


the NOx emission from the engine with variable com-
pression is also greater than that of conventional engine
and is about 0.97 g/kg of fuel at 2,460 rpm. The increase
of NOx is a drawback to the advantages mentioned when
using variable-compression-ratio engines to achieve
minimum bsfc (maximum efficiency). This drawback
can be overcome by reducing the engine speed range
from 1,500 to 2,800 rpm to 1,665 to 2,800 rpm as illu-
strated on the figure.
To conclude based on Figures 6, 7, 8, and 9, varying
the compression ratio, within the entire range of engine
speed of 1,500 to 2,800 rpm, to attain a minimum bsfc
results in significant variations in the predicted perform-
ance variables compared to those predicted at constant
compression ratio (rc = 16.4). These variations can be
Figure 14 Soot and Nox emission versus speed for constant- summarized as follows: (1) at an engine speed of 2,460
and variable-compression-ratio engines.
and 2,800 rpm, there is no variation in all investigated
performance variables; (2) for the other values of engine
cylinder. The general decrease in residual gas fraction is speed, the maximum pressure, maximum temperature,
desirable because it decreases the dilution of the charge brake torque, brake power, and unfavorable NOx are in-
in the cylinder and increases the scavenging efficiency of creasing. On the other hand, the delay period, volumet-
the engine. However, these reductions are small and can ric efficiency, RGF, and soot formation are decreasing.
be neglected. Table 2 summarizes the predicted results of performance
The effect of the optimal variation of the compression variables at different values of engine speed.
ratio corresponding to the minimum bsfc on soot and
NOx emissions is shown in Figure 9. For all reported Case 2: maximum brake torque
speed range, the soot emission from the engine with The predicted performance variables corresponding to
variable compression is less than that of the conven- maximum brake torque compared to those predicted at
tional engine. The decrease in soot emission is found to a conventional constant compression ratio (rc = 16.4) are
be about 21% at 1,500 rpm and decreases until it reaches displayed in Figures 10, 11, 12, 13, and 14. Figure 10
0% at 1,900 rpm. Then, it increases again to 9% at 2,460 shows that the brake torque, in the case of a constant
rpm and returns to 0% at 2,800 rpm. On the other hand, compression ratio, changes between 281 and 252 N.m
the formed NOx inside the variable-compression-ratio passing through a maximum of 295 N.m at about 1,900
engine is much higher compared to that inside the rpm. For the variable compression ratio, the brake
constant-compression-ratio engine. The maximum in- torque is kept constant at its maximum value of 295 N.
crease reaches about 11.2 g/kg of fuel instead of 6.4 g/kg m over a speed ranging from 1,500 to 2,460 rpm by

Table 2 Results of varying engine rc compared to constant rc (rc = 16.3) to attain the minimum bsfc
Engine performance N = 1,500 rpm N = 1,900 rpm N = 2,460 rpm
variables
Increase (%) Decrease (%) Increase (%) Decrease (%) Increase (%) Decrease (%)
bsfc - 3.7 0 0 - 4.4
Maximum pressure 6 - 0 0 5.5 -
Maximum temperature 4.3 - 0 0 1.9 -
Brake torque 3.3 - 0 0 2.8 -
Brake power 3.8 - 0 0 2.2 -
Delay period - 6.9 0 0 - 4.5
Volumetric efficiency - 0.3 0 0 - 0.1
RGF - 4.3 0 0 - 4.3
Soot - 21 0 0 - 8.5
NOx 75 - 0 0 76 -
bsfc brake-specific fuel consumption, RGF residual gas fraction.
Abou Al-Sood et al. International Journal of Energy and Environmental Engineering 2012, 3:13 Page 11 of 13
http://www.journal-ijeee.com/content/3/1/13

Table 3 Results of varying engine rc compared to constant rc (rc = 16.4) to attain maximum brake torque
Engine performance N = 1,500 rpm N = 1,900 rpm N = 2,460 rpm
variables
Increase (%) Decrease (%) Increase (%) Decrease (%) Increase (%) Decrease (%)
bsfc - 4.6% 0 0 - 2.7%
Maximum pressure 10.3 - 0 0 6% -
Maximum temperature 3.6 - 0 0 2 -
Brake torque 5.4 - 0 0 4.6 -
Thermal efficiency 4.4 - 0 0 2.8 -
Delay period - 15.3 0 0 - 6.8
Volumetric efficiency - 0.48 0 0 - 0.25
RGF - 6.5 0 0 - 5
Soot - 9.7 0 0 - 12
NOx 21.7 - 0 0 90.7 -
bsfc brake-specific fuel consumption, RGF residual gas fraction.

varying the compression ratio between 16.4 and 19.3. figure shows that the volumetric efficiency is reduced
This means that the brake torque increases by 5.4% at by about 0.48% at a speed of 1,500 rpm and 0.25% at
1,500 rpm and 4.6% at 2,460 rpm. The shown variation 2,460 rpm passing through 0% both at 1,900 and
trend in compression ratio urges that actual variation in 2,800 rpm. Also, this figure shows that variation of
compression ratio should be considered to increase the the compression ratio for optimum brake torque
brake torque to its constant maximum value. results in a general decrease in RGF by about 6.5% at
Figure 11 shows the increase in pmax and Tmax 1,500 rpm and 5% at 2,460 rpm passing through 0%
when operating under variable compression ratio both at 1,900 and 2,800 rpm.
compared to the operation under constant compres- The effect of the optimal variation of compression
sion ratio. The increase in pmax is about 0.9 MPa at ratio for maximum brake torque on soot and NOx emis-
1,500 rpm and 0.6 MPa at 2,460 rpm passing through sion is shown in Figure 14. The soot emission from the
0 MPa both at about 1,900 and 2,800 rpm. The no- engine with variable compression ratio is less than that
ticeable increase in Tmax is about 60 K at 1,500 rpm of the conventional engine by about 9.7% at 1,500 rpm,
and 40 K at 2,460 rpm passing through 0 K both at 0% both at 1,900 and 2,800 rpm, and 12% at 2,460 rpm.
1,900 and 2,800 rpm. These increases in pmax and The formed NOx inside the variable-compression-ratio
Tmax are tolerable, as they are still below the allow- engine is very high compared to that inside the
able pressure and temperature of the conventional constant-compression-ratio engine by about 21.7% (2.65
engine. Similar observations for the changes of brake g/kg of fuel) at 1,500 rpm, 90.7% (0.461 g/kg of fuel) at
efficiency and bsfc versus engine speed for operation 2,460 rpm, and 0% at both 1,800 and 2,800 rpm. This in-
under constant and variable compression ratios are crease is due to the increase of cylinder temperature
shown in Figure 12. If the compression ratio is con- with increasing compression ratio. The increase in NOx
stant, the brake thermal efficiency varies between 36% at low speed, i.e., 1,500 rpm, can be avoided by avoiding
and 32.9% passing through a maximum value of the engine to run in the speed range of 1,500 to 1,620
37.3% at an engine speed of 1,900 rpm. In case of rpm as illustrated in Figure 14.
varying compression ratio (between 16.4 and 19.3 for The conclusion based on Figures 10, 11, 12, 13, and 14
maximum torque), the efficiency changes between is that varying the compression ratio of the diesel en-
37.6% at 1,500 rpm and 36.5% at 2,460 rpm. In gine, within the entire range of engine speed of 1,500 to
addition to gains in brake torque and efficiency, the 2,800 rpm, to achieve a maximum engine torque leads
figure shows a decrease in the bsfc over an engine to variations in the predicted performance variables
speed ranging from 4.6% at 1,500 rpm to 2.7% at compared to those predicted at constant compression
2,460 rpm passing through 0% both at 1,900 and ratio (rc = 16.4). These variations can be summarized as
2,800 rpm. follows: (1) at an engine speed of 1,500 and 2,460 rpm,
Figure 13 shows a significant decrease in delay pe- there is no variation in all studied variables; (2) for the
riod, volumetric efficiency, and RGF over an engine other values of engine speed, the maximum pressure,
speed ranging from 1,500 to 2,460 rpm with values maximum temperature, brake torque, thermal efficiency,
between 15.3% (0.45° of crank angle) at 1,500 rpm and unfavorable NOx are increasing. On the contrary,
and 6.8% (0.3° of crank angle) at 2,460 rpm. Also, the the delay period, volumetric efficiency, residual gas
Abou Al-Sood et al. International Journal of Energy and Environmental Engineering 2012, 3:13 Page 12 of 13
http://www.journal-ijeee.com/content/3/1/13

fraction, and soot formation are decreasing. Table 3 professor in the Department of Mechanical Engineering at the Assiut
summarizes the predicted results of performance vari- University, Egypt. His current research interests focus on sprays and liquid
atomization, fuel cells, solar hydrogen generation, and two-phase flows. YM
ables at different values of engine speed. received his Ph.D. degree in Mechanical Engineering from Kansas State
University, USA, in 1984. Now, he is an emeritus professor in the Department
Conclusions of Mechanical Engineering at the Assiut University, Egypt. His research
interests focus on combustion and management and optimization of
The present rapid thermodynamic simulation for the thermal systems.
direct-injection diesel engine is based on a single-zone
combustion modeling with a rapid premixed burning Acknowledgment
This work has been fully supported by Assiut University and the Mechanical
phase followed by a slower mixing-controlled burning Engineering Department.
phase. Comparison with other experimental results
(under constant compression ratio) verifies very well the Received: 13 March 2012 Accepted: 19 June 2012
Published: 7 August 2012
simulation predictions of cycle performance. The main
conclusions of this study indicate that varying the com- References
pression ratio to achieve minimum bsfc or maximum 1. Kaario, O, Lendormy, E, Sarjovaara, T, Larmi, M, Rantanen, P: In-cylinder flow
field of a diesel engine. SAE Paper No. 2007-01-4046., (2007). doi:10.4271/
brake torque for moderate diesel engine under the re- 2007-01-4046
striction that the maximum cylinder pressure and 2. Lee, C-FF, Zhao, J: Modeling of blowby in a small-bore high-speed direct
temperature do not exceed those of the conventional en- injection optically accessible diesel engine. SAE Paper No. 2006-01-0649.,
(2006). doi:10.4271/2006-01-0649
gine results in saving fuel consumption, higher brake ef- 3. Han, Z, Reitz, RD: Turbulence modeling of internal combustion engines
ficiency and power, and reduction in soot emission from using RNG k-E models. Combust. Sci. Technol. 106, 267–295 (1995)
the engine. Also, an increase in NOx is noticed at low 4. Hessel, RP, Rutland, CJ: Intake flow modeling in a four-stroke diesel engine
using KIVA-3. J. Propulsion Power 11, 378–384 (1995)
speed. This drawback is considerable and can be over- 5. Reitz, RD, Rutland, CJ: Development and testing of diesel engine CFD
come by reducing the operation speed range from 1,500 models. Prog. Energy Sci. 21, 173–196 (1995)
to 2,800 rpm to 1,665 to 2,800 rpm in the case of mini- 6. Kopa, RD, Hollander, BR, Hollander, FH, Kimura, H: Combustion temperature,
pressure and products at chemical equilibrium. Report No. 64–1 Dept. of
mum bsfc and from 1,500 to 1,800 rpm to 1,620 to Mech. Eng., Univ. of California, (1964)
2,800 rpm in the case of maximum brake torque. 7. Arrègle, J, Lòpez, JJ, Garcia, JM, Fenollosa, C: Development of of a zero-
dimensional diesel combustion model. Part 1: analysis of the quasi-steady
Abbreviations diffusion combustion phases. Appl. Therm. Eng. 23, 1301–1317 (2003)
CN: fuel cetane number; cp, cv: constant pressure and volume specific heats; 8. Kouremenous, DA, Rakopoulos, CD, Hountalas, D: Thermodynamic analysis
Div: intake valve head diameter; D: cylinder bore; H, h: enthalpy and specific of indirect injection diesel engine by using two-zone modeling of
enthalpy; K: specific heat ratio; M: mass; N: crank speed; N: number of moles; combustion. Trans. ASME, J. Eng. Gas Turbine Power 112, 138–149 (1990)
nc: number of cylinders per engine; niv, nr: number of intake valves and rings 9. Chmela, FG, Pirker, GH, Wimmer, A: Zero-dimensional ROHR simulation for
per cylinder; p: pressure; Q, q: heat transfer and heat flux; rc: compression DI diesel engines–a generic approach. Energy Convers. Manage. 48, 2942–
ratio; R: gas constant; Sp: mean piston velocity; T: cylinder temperature; V, Vc, 2950 (2007)
Vd: swept, clearance, and maximum displacement volumes; X: mole fraction; 10. Basha, SA, Gopal, KR: In-cylinder fluid flow, turbulence and spray models—a
Θ: crank position angle; Λ: connecting rod length to crank radius ratio; θevo, review. Renew. Sustain. Energy Rev. 13, 1620–1627 (2009)
θivo: exhaust and intake valve opening angles; Δθeopen: opening duration 11. Roberts, M: Benefits and challenges of variable compression ratio (VCR). SAE
angle of exhaust valve; Δθiopen: opening duration angle of intake valve; Paper No. 2003-01-0398., (2003). doi:10.4271/2003-01-0398
ρ: density; ϕ: equivalence ratio; τ, τ: ignition delay period and ignition delay 12. Ladommatos, N, Balian, RA: Combustion in a direct injection diesel engine with
angle; B: backward; comb: combustion; e: exit, equilibrium; em: exhaust increased clearance volume. Proc. Instn. Mech. Engr. 204, 187–197 (1990)
manifold; ex: exhaust; f: fuel, forward; fb: fuel burning; ft: total fuel; 13. Sobotowski, RA, Potrer, BC, Pilley, AD: The development of novel variable
i: indicated, inlet, injection start; ig: ignition; im: intake manifold; in: intake; compression ratio, direct injection diesel engine. SAE Paper No. 910484.,
o: ambient condition; p: piston; s: stoichiometry, soot; sf: soot formation; (1991). doi:10.4271/910484
so: soot oxidation; w: wall, atoms of carbon in fuel-air molecule; x: hydrogen 14. Rychter, TJ, Teodorczyk, A, Stone, CR, Leonard, HJ, Ladommatos, N, Charlton,
atoms in fuel-air molecule; y: oxygen atoms in fuel-air molecule; z: nitrogen SJA: Theoretical study of a variable compression ratio turbocharged diesel
atoms in fuel-air molecule. engine. Proc. Instn. Mech. Engr. 206, 227–238 (1992)
15. Freudenstein, F, Maki, ER: Variable-displacement piston engine. US Patent
Competing interests 4,270,495, 2 June 1981
The authors declare that they have no competing interests. 16. Yamin, JA, Dado, MH: Performance simulation of a four-stroke engine with
variable stroke-length and compression ratio. Appl. Energy 77, 447–463 (2004)
Authors’ contributions 17. Raheman, H, Ghadge, SV: Performance of diesel engine with biodiesel at
MM conceived the concept and procedures of the present work, developed varying compression ratio and ignition timing. Fuel 87, 2659–2666 (2008)
the model, and carried out the computations and analysis. MA checked the 18. Jindal, S, Nandwana, BP, Rathore, NS, Vashistha, V: Experimental
equations and analysis, and reviewed the revised manuscript. YM reviewed investigation of the effect of compression ratio and injection pressure in a
the graphs and results. All authors read and approved the final manuscript. direct injection diesel engine running on Jatropha methyl ester. Appl.
Therm. Eng. 30, 442–448 (2010)
Authors’ information 19. Muralidharan, K, Vasudevan, D, Sheeba, KN: Performance, emission and
MM earned his Ph.D. degree in Mechanical Engineering from the University combustion characteristics of biodiesel fuelled variable compression ratio
of Manitoba, Canada, in 2007. He is currently an assistant professor in the engine. Energy 36, 5385–5393 (2011)
Department of Mechanical Engineering at the Assiut University, Egypt. He 20. Gandure, J, Ketlogetswe, C: Comparative performance analysis of marula oil
has been interested in heat and mass transfer relating to droplets and spray and petrodiesel fuels on a variable compression ratio engine, pp. 13–15.
modeling, turbulence, combustion, and modeling and simulation of internal AFRICON, 2011, Victoria Falls, Livingstone (2011)
combustion engines. MA received his Ph.D. degree in Mechanical 21. Bishop, IN: Effect of design variables on friction and economy. SAE Trans. 3,
Engineering from Mississippi State University, USA, in 1994. He is currently a 334–358 (1965)
Abou Al-Sood et al. International Journal of Energy and Environmental Engineering 2012, 3:13 Page 13 of 13
http://www.journal-ijeee.com/content/3/1/13

22. Heywood, JB: Modeling combustion and performance characteristics of


internal combustion engines. In: Proceedings of the 1976 Heat Transfer and
Fluid Mechanics Institute, pp. 180–195. Stanford University Press, Palo Alto
(1976)
23. Eichelberg, G: Some new investigations on old combustion-engine
problems. Engineering 149, 463–547 (1939)
24. Rakopoulos, CD, Hountalas, DT: Net and gross heat release rate calculation
in DI diesel engine using various heat transfer models, pp. 13–19. Paper
presented at the 1994 American Society of Mechanical Engineers-winter
annual meeting, Chicago (1994)
25. Annand, WJD: Heat transfer in cylinders of reciprocating internal
combustion engines. Proc. Instn. Mech. Engr. 177, 973–990 (1963)
26. Annand, WJD, Ma, TH: Instantaneous heat transfer rates to cylinder head
surface of small compression ignition engine. Proc. Instn. Mech. Engr. 185,
967–987 (1970)
27. Woschni, GA: Universally applicable equation for the instantaneous heat
transfer coefficient in internal combustion engine. SAE Trans. 76, 3065–3083
(1967)
28. Blawton, B: Effect of compression and expansion on instantaneous heat
transfer in reciprocating internal combustion engines. Proc. Instn. Mech.
Engr. 201, 175–186 (1987)
29. Campbell, AS: Thermodynamic Analysis of Combustion Engines. John Wiley
& Sons Press, New York (1979)
30. Ogurt, T, Takeda, A, Torii, K, Inaba, S: Radiation heat transfer of combustion
flames in a diesel engine. Bull. JSME 28, 638–646 (1985)
31. Cheung, CG, Leung, CW, Leung, TP: Modelling spatial radiative heat flux
distribution in a direct injection diesel engine. Proc. Instn. Mech. Engr. 208,
275–283 (1994)
32. Heywood, JB: Internal Combustion Engine Fundamentals, 1st edn. McGraw-
Hill Science/Engineering/Math, New York (1988)
33. Ramos, JL: Internal Combustion Engine Modeling, 1st edn. Taylor & Francis,
London (1989)
34. Zelenik, FJ: Combustion modeling in internal combustion engines.
Combust. Sci. Technol. 12, 159–164 (1976)
35. Gu, F, Ball, AD: Diesel injector dynamic modelling and estimation of
injection parameters from impact response. Part 1: modelling and analysis
of injector impacts. Proc. Instn. Mech. Engr 210, 293–302 (1996)
36. Watson, N, Pilley, AD, Marzouk, MA: Combustion correlation for diesel
engine simulation. SAE Paper No. 800029., (1980). doi:10.4271/800029
37. Mehta, PS, Singal, SK, Pundir, BPA: Comprehensive simulation model for
mixing and combustion characteristics of small direct injection diesel
engine. Proc. Instn. Mech. Engr. 209, 117–126 (1995)
38. Rossini, FD, Pitzer, KS, Arnett, RL, Braun, RM, Primentel, GC: Selected Values
of Physical and Thermodynamic Properties of Hydrocarbons and Related
Compounds. Carnegie Press, Pittsburgh (1953)
39. Hiroyasu, H, Kadota, T: Models of combustion and formation of nitric oxide and
soot in DI diesel engine. SAE Paper No. 760129., (1976). doi:10.4271/760129
40. Morel, T, Keribar, R: Heat radiation in D.I. diesel engines. SAE Paper No.
860445., (1986). doi:10.4271/860445
41. Wahiduzzaman, S, Morel, T, Timar, J, DeWitt, DP: Experimental and analytical
study of heat radiation in a diesel engine. SAE Paper No. 870571., (1987).
doi:10.4271/870571
42. Han, Z, Uludogan, A, Hampson, GJ, Reitz, RD: Mechanism of soot and NOx
emission reduction using multiple-injection in a diesel engine. SAE Paper
No. 960633., (1996). doi:10.4271/960633
43. Al-Sood, M, Abdel-Latif, A, Abdel-Rahim, Y, Ibrahim, A: Optimum compression
ratio variation of a 4-stroke, direct injection diesel engine for minimum bsfc.
SAE Paper No. 1999-01-2519. (1999). doi:10.4271/1999-01-2519
Submit your manuscript to a
doi:10.1186/2251-6832-3-13 journal and benefit from:
Cite this article as: Abou Al-Sood et al.: Rapid thermodynamic
simulation model for optimum performance of a four-stroke, 7 Convenient online submission
direct-injection, and variable-compression-ratio diesel engine.
International Journal of Energy and Environmental Engineering 2012 3:13.
7 Rigorous peer review
7 Immediate publication on acceptance
7 Open access: articles freely available online
7 High visibility within the field
7 Retaining the copyright to your article

Submit your next manuscript at 7 springeropen.com

You might also like