Applied Thermal Engineering: Research Paper

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 17

Applied Thermal Engineering 110 (2017) 1483–1499

Contents lists available at ScienceDirect

Applied Thermal Engineering


journal homepage: www.elsevier.com/locate/apthermeng

Research Paper

Numerical comparison of PCCI combustion and emission of diesel


and biodiesel fuels at low load conditions using 3D-CFD models
coupled with chemical kinetics
Alborz Zehni, Rahim Khoshbakhti Saray ⇑, Kamran Poorghasemi
Faculty of Mechanical Engineering, Sahand University of Technology, Sahand New Town, Tabriz, Iran

h i g h l i g h t s

 Comparison of PCCI combustion of B100 and B0 fuels was performed.


 30 CAD ATDC is the optimum SOI timing for the lowest NOx + THC and CO emissions.
 NOx, CO, THC for B100 were lowered by 13%, 27%, 12% respectively compared to B0.
 ISFC for B100 case is increased by 6% compared to B0 case.

a r t i c l e i n f o a b s t r a c t

Article history: LTC (Low Temperature Combustion) has demonstrated abilities to lower diesel engine emissions while
Received 24 January 2016 achievement of better fuel economy in comparison with conventional combustion. One of the major
Revised 4 July 2016 LTC methods is PCCI (Premixed Charge Compression Ignition) combustion strategy in which high levels
Accepted 11 September 2016
of EGR (exhaust gas recirculation) is utilized to reduce overall combustion temperatures and to lengthen
Available online 12 September 2016
the ignition delay. Increase in ignition delay provides time for fuel evaporation and reduces in-
homogeneities in the reactant mixture, thus NOx formation from local temperature spikes and Soot
Keywords:
formation from locally rich mixtures can be reduced. However, as dilution is increased to the limits,
PCCI
Biodiesel
Soot, HC and CO can significantly increase. One of the ways to control these emissions in PCCI combustion
Diesel is utilizing biodiesel fuel due to its different physical and chemical properties compared to petroleum-
CO based diesel fuel. Hence, in the present work, a numerical study is performed to compare the combustion,
THC performance and emission characteristics of neat biodiesel and diesel fuel surrogates under the condition
of PCCI combustion mode in a light-duty, single-cylinder diesel engine by KIVA-CHEMKIN code. The oper-
ating conditions are at the engine speed of 2000 rpm and load of 5.5 bar IMEP. Simulations are capable of
capturing, not only the bulk combustion characteristics, as well as the first and second-stage combustion
processes, but also the details of the important reactive species and emission formation. For both diesel
and biodiesel fuels, it has been tried to find optimum cases which are only based on sweep of injection
timing and do the comparisons at those optimum states. The results indicate that for both diesel and bio-
diesel fuels, 30 CAD ATDC SOI timing can be regarded as an optimum SOI timing for the lowest NOx
+ THC and CO emissions. For the optimum case, NOx, CO, THC, thermal efficiency and ISFC for the biodie-
sel case are lowered by 13%, 27%, 12%, 9.5% and increased by 6%, respectively compared to the diesel case.
Ó 2016 Elsevier Ltd. All rights reserved.

1. Introduction including HCCI/RCCI/PCCI (Homogeneous/Reactivity controlled


and Premixed Charge Compression Ignition) are receiving
Since emission regulations become more stringent, new com- increased attention due to their potential for simultaneously
bustion concepts should be developed to meet these emission reducing Soot and NOx emissions from DI (Direct Injection) diesel
standards. Nowadays, LTC (Low Temperature Combustion) modes engines [1].
For the HCCI mode, in-cylinder homogeneity may cause rapid
combustion and pressure rise by simultaneous ignition throughout
⇑ Corresponding author. the cylinder space and the mixture can be susceptible to great
E-mail address: khoshbakhti@sut.ac.ir (R. Khoshbakhti Saray). combustion noise. It is also very tough to control the combustion

http://dx.doi.org/10.1016/j.applthermaleng.2016.09.056
1359-4311/Ó 2016 Elsevier Ltd. All rights reserved.
1484 A. Zehni et al. / Applied Thermal Engineering 110 (2017) 1483–1499

Nomenclature

A0 fluid flow constant q_ cm density change rate due to chemistry (kg/m3s)


a dimensionless number used for low Mach number flows q_ s density change rate owing to Spray (kg/m3s)
D diffusion coefficient (m2/s) e dissipation rate of turbulence (J/kgs)
F rate of momentum gain per unit volume x_ m molar production rate (kmol/m3s)
g specific body force, assumed constant (m/s2)
I specific internal energy (J/kg) Abbreviations
J heat flux (W/m2) B0/B20/B100 neat diesel/20% (vol) biodiesel/neat biodiesel
k turbulent kinetic energy (J) ATDC After Top Dead Center
m_ mass flow rate (kg/s) BTDC Before Top Dead Center
M number of reactions in chemical kinetics mechanism CAD crank angle degree
P pressure (Pa) CFD Computational Fluid Dynamics
Q heat content (kJ/kg) CI Compression Ignition
Q_ c source term due to chemistry (W/m3) CO Carbon Monoxide
Q_ s source term owing to spray (W/m3) DI Direct Injection
t time (s) DVODE double-precision variable-coefficient ordinary
u velocity vector (m/s) differential equation solver
W molecular weight (kg/kmol) EOI End of Injection
Y mass fraction EGR exhaust gas recirculation
EVO Exhaust Valve Open
Subscripts HCCI Homogeneous Charge Compression Ignition
HV heating value HRR Heat Release Rate (J/deg)
f fuel IMEP Indicated Mean Effective Pressure (Pa)
i indicated IVC Inlet Valve Closure
m number of species in chemical kinetics mechanism LTC Low Temperature Combustion
t thermal LTHR/HTHR low/high temperature heat release
MD/MD9D methyl decanoate/methyl-9-decanoate
Superscripts NOx Nitrogen Oxides
c chemistry source PCCI Premixed Charge Compression Ignition
s spray contribution RCCI Reactivity Controlled Compression Ignition
PM Particulate Matter
SMD sauter mean diameter (m)
Greek symbols
SOI/ASOI Start of Injection/After Start of Injection
dml dirac delta function
THC/HC Total Hydrocarbon/Hydrocarbon
Dhf heat of formation of species m at absolute zero (kJ/kmol) W_ power (kW)
q density (kg/m3)
g efficiency

phases. PCCI and RCCI concepts have been evolved from the HCCI In the PCCI combustion process, fuel can be injected into the
combustion mode for the sake of better control over the start of combustion chamber in three ways. These are, advanced direct
combustion (SOC) and controlling burn duration. injection, port fuel injection and late direct injection [7]. Advanced
RCCI combustion concept, has arisen as a combustion mode in direct injection and port fuel injection are more similar to the HCCI
which combustion is controlled mainly by fuel reactivity [2,3]. This strategy and suffer from fuel spray impingement on the cylinder
combustion mode symbolizes an evolution from the PCCI concept, walls and incomplete fuel evaporation. Consequently, contributing
because fuel and air are mixed before combustion, but in this case to a dramatic increase in HC and CO emissions [8–10]. Late direct
fuel reactivity varies across the cylinder. High-octane fuel (low
reactivity) with early direct injected high-cetane fuel (high reactiv-
ity) are blended in order to handle auto-ignition and tune the reac-
tivity of the combustion mixture [4]. PCCI is not fully homogeneous
like HCCI and achieves desired ignition delay through extensive
use of exhaust gas recirculation (EGR). Adoption of high EGR per-
mits longer ignition delay. Hence, it permits better premixing of
air–fuel compared to the conventional diesel combustion, results
in less fuel-rich pockets followed by a low temperature combus-
tion, which simultaneously reduces NOx and Soot levels [5]. The
operation of PCCI combustion decouples injection and heat release
events. Since combustion starts after end of injection, the process is
primarily governed by chemical kinetics and not by diffusive
mixing anymore like in Compression Ignition (CI) engines. Fig. 1
compares the general trend of ignition delay, heat release rate,
in-cylinder temperature and pressure for PCCI and conventional
diesel combustion. More ignition delay, two stage combustion
and lower cylinder temperature for PCCI mode is apparent com- Fig. 1. Comparison of general combustion characteristics of PCCI and conventional
pared to conventional diesel combustion [6]. CI combustion [6].
A. Zehni et al. / Applied Thermal Engineering 110 (2017) 1483–1499 1485

injection avoids the fuel-wall impingement and gives a way to heavy-duty diesel engine. They concluded that despite the occur-
switch the combustion style to the conventional diesel engines. rence of NOx/HC trade-off, ethanol-diesel-biodiesel is effective in
The results of studies imply that although PCCI combustion reducing NOx and smoke simultaneously in the PCCI combustion.
reduces NOx significantly but, like HCCI strategy, the problem of Su et al. [23] experimentally showed that in the PCCI combustion,
HC and CO emissions are not solved completely [11–14]. higher fuel oxygen content, lower stoichiometric air–fuel ratio and
One way to solve this problem is the application of biodiesel reduced aromatic content of biodiesel relative to diesel causes the
fuels as an alternative fuel for the PCCI combustion [15,16]. As more reduction of Soot formation. Tompkins et al. [24] experimen-
can be seen in Table 1, biodiesel has different physical and tally determined the behavior of low temperature heat release for
chemical properties compared to petroleum-based diesel fuel diesel and biodiesel fuels in late injection low temperature com-
[17]. Biodiesel has approximately 11% oxygen (by weight) which bustion in a medium duty diesel engine and the influence of such
results in reduced Soot, CO, and HC emissions compared to its behavior on LTC torque and emissions. The results showed that
diesel counterpart. In addition, biodiesel contains low sulfur, and Diesel fuel displayed a longer and more intense low temperature
therefore produces low sulfur oxides or sulfates [18]. However, heat release phase. Lower amounts of low temperature heat
biodiesel has a higher viscosity, density, speed of sound, and sur- release in the biodiesel causes less sensitivity to EGR, less instabil-
face tension that may cause differences in injection and combus- ity, and produces better torque and emission characteristics. Jung
tion processes. Also, due to high oxygen content, it may cause to et al. [25] compared PM of diesel and biodiesel fuels under conven-
the slight increase of NOx [19,20]. Thus, there is a need for more tional and PCCI combustions by some experimental methods. The
comprehensive research work to conclusively determine the bene- results revealed that the amount of desorbed volatile hydrocarbons
fits and drawbacks of biodiesel for utilizing in the LTC combustion from biodiesel PM is larger than that of diesel PM, and with either
especially in the PCCI combustion branch. fuel it is larger for the LTC mode than for conventional compression
By reference to research on modeling of diesel and biodiesel LTC ignition combustion mode. Between the conventional CI combus-
combustion, the modeling of HCCI combustion by various models tion modes, the population of PM with larger portion of diffusion
such as multi-zone and multi-dimensional models are abundant combustion phase biased toward larger size.
which Komninos and Rakopoulos [21] have pointed the important Regarding multi-zone modeling, it can be pointed to the works
works in a review paper. On the other hand, although there are of Rakopoulos et al. [26,27] in which the conventional combustion
some publications regarding combustion, emission and perfor- and emissions of diesel and biodiesel fuels were compared. But,
mance of diesel fuel under the PCCI combustion mode, the PCCI they did not expand their comparison for the PCCI combustion
combustion of biodiesel fuel is still under development. Moreover, mode. On the contrary, Kuleshov [28] studied the PCCI combustion
most biodiesel PCCI combustion works are conducted experimen- and emissions of diesel fuel with multi-zone models. But, the bio-
tally and the need of comprehensive multi-zone or 3D- diesel fuel has not been taken into account in his simulations.
computational fluid dynamic (CFD) simulation works still exists The main CFD studies related to the PCCI combustion of diesel
in order to identify the uncertainties and results which cannot be and biodiesel fuels are such as work by Lee et al. [29] that simu-
obtained by experimental works easily. lated the PCCI combustion with soybean biodiesel blends of neat
The recent significant experimental studies regarding PCCI diesel and B20 (20% biodiesel, 80% diesel) for a heavy duty test
combustion of diesel and biodiesel fuels include the work of Fang
et al. [22] in which they studied the effects of ethanol-diesel-
biodiesel blends on PCCI combustion and its emissions for a Table 2
Chemical formula and molecular structure of the soybean biodiesel components [39].

Ester component Chemical formula Molecular structure


Table 1
Comparison of physical and chemical properties of diesel and biodiesel fuels [17]. Methyl palmitate C17H34O2

Property Diesel fuel Biodiesel fuel (soybean)


Methyl stearate C19H38O2
Density @ 15 °C (g/cm3) 0.825 0.89
Lower heating value (MJ/kg) 42.6 37.528
Cetane number 50.74 53.80 Methyl oleate C19H36O2
Kinematic viscosity @ 40 (mm2/s) 3.196 5.249
Flash point (°C) 74 148
Methyl linoleate C19H34O2
Cold filter plugging point (°C) 15 +13
Distilation (°C)
Initial boiling point (IBP) 162.5 332 Methyl linolenate C19H32O2
Finial boiling point (FBP) 359.7 353.6
Sulfur content (ppm) 50 2.5

Fig. 2. Schematic diagram for coupling between KIVA code and Chemkin chemistry solver.
1486 A. Zehni et al. / Applied Thermal Engineering 110 (2017) 1483–1499

Fig. 3. Evolution of the main thermo physical properties of methyl oleate, methyl palmitate and tetradecane versus temperature [40,42].

engine by KIVA. A surrogate mechanism was constructed by


Table 3 combining the skeletal mechanisms of methyl butanoate (MB)
Engine characteristics of GM 1.9 L [32].
and n-heptane. The results indicated that there is no significant
Bore/stroke 82.0/90.4 [mm] difference between diesel and B20 for ignition delays in the test
Displacement per cylinder 477 [cm3] conditions. In addition, CO is suppressed for B20 in LTC, while Soot
Aspiration Turbocharged
Compression ratio 16.7
decreases in LTC with minor effects of oxygenated fuel. It should be
Connecting rod length 145.5 [mm] notified that because of the insufficient length of the carbon chain
Squish height 0.617 [mm] in MB, this mechanism does not adequately capture the ignition
IVC (CA°ATDC) 132 delay. Moreover, cool flame regime cannot be observed during its
EVO (CA°ATDC) 112°
oxidation processes [30,31]. Brakora and Reitz [32] developed a
Injector characteristics: bosch common rail CRIP2.2 reduced biodiesel mechanism including n-heptane, methyl decan-
Type/sac volume Mini Sac/0.23 [mm3]
oate (MD) and methyl-9-decanoate (MD9D) species for use in 3-D
Holes num./nozzle diam. 7/d = 0.14 [mm]
Included angle 155° simulation of biodiesel PCCI combustion. The predicted in-cylinder
Rail pressure 860 [bar] pressure, heat release histories and emissions obtained from the
KIVA3V-Chemkin coupling indicated good agreement with the

Table 4
Engine initial conditions and operating conditions.

Case group 1 [46] Case group 2 [32]


IMEP [bar] 3 5.5
Engine speed [rpm] 1500 2000
Injected fuel [mg/inj] 8.8 B0:13.5 B100:16.07
SOI/EOI [°CA ATDC] Case 1: 31.1/25.9 B0: Case 4: 38/26.77, Case 5: 34/23.27, Case 6: 30/20.42, Case 7: 26/16.5
Case 2: 26.6/21.4 Case 8: 22/12.78
Case 3:15.8/10.6 B100: Case 9: 38/28.06, Case 10: 34/24.28, Case 11: 30/19.63, Case 12: 26/15.7
Case 13: 22/11.63
EGR [%] 67 67
Fuel type B0 B0, B100
Initial O2/N2/CO2/H2O mass 0.09571/0.74012/0.11185/ B0: 0.09571/0.74012/0.11185/0.05232
fraction 0.05232 B100: 0.09538/ 0.73182/0.12261/0.05019
Intake temperature [K] 363 338
Intake pressure [bar] 1.5 1.62
Swirl ratio 2.2 2.2
Cylinder wall temperature 403 403
Cylinder head temperature 513 513
Piston temperature 535 535
A. Zehni et al. / Applied Thermal Engineering 110 (2017) 1483–1499 1487

corresponding experimental data. However, Brakora did not inves- momentum, energy and species transport equations are formu-
tigate the biodiesel PCCI combustion performance and emission lated in KIVA as given in equations (1–4), respectively.
formation in detail.   
@ qm q
The main goal of the present work is to perform a detail com- þ r  ðqm uÞ ¼ r  qDr m þ q_ cm þ q_ s dml ð1Þ
parison of PCCI combustion, performance and emission character-
@t q
istics of pure biodiesel (B100) and diesel (B0) fuel surrogates by  
@ qu 1 2
multi-dimensional KIVA-3V code coupled with Chemkin. The þ r  ðquuÞ ¼  2 rp  A0 r qk þ r  r þ Fs þ qg ð2Þ
model predictions are validated using measured in-cylinder pres- @t a 3
sure histories and exhaust emissions. The developed model is used
to compare the PCCI combustion and its emission characteristics of @
ðqIÞ þ r  ðquIÞ ¼ Pr  u þ ð1  A0 Þr : ru  r  J
the diesel and biodiesel fuels at the optimum state of emissions. @t
The conclusions can open a new prospect for the future researches. þ A0 qe þ Q_ c þ Q_ s ð3Þ

@
2. Model description ðqYm Þ þ r  ðquYkm Þ ¼ r  ðDym rYm Þ þ q_ cm ð4Þ
@t

2.1. Flow field, spray and combustion treatment where q_ cm in Eqs. (1) and (4) and Q_ c in Eq. (3) are the source terms
that need to be calculated by combustion model. Mathematical
Flow simulation is carried out using a version of the Los Alamos descriptions of these terms are as follows:
National Laboratory (LANL) CFD code, KIVA 3V, with improved tur-
q_ cm ¼ Wm x_ m ð5Þ
bulence and spray/wall interaction models [33]. The KIVA code
solves the turbulent, three dimensional transient conservation
X
M
equations of mass, momentum, energy and species using a finite- Q_ c ¼  x_ m ðDhf Þm ð6Þ
volume, temporal-differencing scheme. This solution procedure, m¼1
namely the Arbitrary Lagrangian-Eulerian (ALE) method, decouples
By reference to the above equations, it can be declared that the
calculations of the diffusion and convection terms from chemical
ultimate goal of a sample combustion model is to determine the
source terms. Hence, each computational cell can be treated as a
chemical species net production rates, x _ m . To calculate the molar
homogeneously mixed reactor at each time-step. Continuity,
production rate of chemical species participated in chemical kinet-
ics mechanism, the gas phase kinetics library of CHEMKIN-II [34] is
integrated into KIVA code. In this procedure, the KIVA chemistry
subroutine chem has been replaced by a new subroutine developed
to perform chemistry solutions by iterative calling of DVODE
[35]. This new subroutine acts as an interface between KIVA
and CHEMKIN and updates the combustion source terms. The

Fig. 4. Normalized injection rate shape [47]. Fig. 6. Mesh independency based on the in-cylinder pressure history.

(a) Entire cylinder mesh-front view (b) Entire cylinder mesh-Top view (c) Mesh of bowl section
Fig. 5. Computational mesh of the GM 1.9 Lit Engine geometry at different views.
1488 A. Zehni et al. / Applied Thermal Engineering 110 (2017) 1483–1499

-21 CAD ATDC -20 CAD ATDC -19 CAD ATDC -18 CAD ATDC

B100

B0

-15 CAD ATDC -14 CAD ATDC -13 CAD ATDC -12 CAD ATDC

B100

B0

Fig. 7. Comparison of B100 and B0 spray tip penetrations at the various crank angle degrees, top plane (SOI = 22 CAD ATDC, rpm = 2000).

Fig. 8. History of spray tip penetration and SMD for B100 and B20 fuels (SOI = 22 CAD ATDC, rpm = 2000).
A. Zehni et al. / Applied Thermal Engineering 110 (2017) 1483–1499 1489

multi-dimensional KIVA code provides the initial species concen- volume reactor. The CHEMKIN subroutines construct M-set of stiff
tration and thermodynamic information of each individual cell at ordinary differential equations and DVODE subroutine is then
every time step to pass to CHEMKIN solver. During combustion successively called to compute the species production rate at the
calculations, each computational cell is treated as a constant end of each time step. Using the procedure discussed above, the

(a) B0-Case 1 (b) B0-Case 2

(c) B0-Case 3
(d) B0-Case 6

(e) B0-Case 7 (f) B0-Case 8


Fig. 9. Comparison of measured and predicted in-cylinder pressure and HRR histories for various B0 cases.
1490 A. Zehni et al. / Applied Thermal Engineering 110 (2017) 1483–1499

flow and chemistry solutions are fully coupled. Fig. 2 summarizes and unsaturated biodiesel surrogates was used in this study. The
the above explanations by a schematic diagram. applied multi-component biodiesel mechanism contains 71 spe-
The turbulence, modeled by RNG k-e model [36], affects the com- cies and 192 reactions. For pure biodiesel (B100), a fuel mixture
bustion by convection, property transport and heat flux. Heat trans- of 25% MD, 25% MD9D and 50% n-heptane in mole basis was used
fer subroutine is also modified in the current model to calculate wall [41]. The detailed thermo-physical properties such as normal
heat flux based on Han and Reitz formulation [37]. Droplet breakup boiling point, critical properties, vapor pressure, latent heat of
was modeled using a hybrid Kelvin Helmholtz (KH) – Rayleigh vaporization, liquid density, liquid viscosity, liquid thermal
Taylor (RT) model [38]. For both diesel and biodiesel fuels, constants conductivity, gas diffusion coefficients and surface tension of
for KH and RT were set to 60 and 0.2, respectively. methyl palmitate and methyl oleate were assigned to saturated
In the present study, the fuel library of the KIVA code was mod- methyl ester MD and unsaturated methyl ester MD9D, respectively
ified to include the thermodynamic and physical properties of bio- [41].
diesel fuel blends mainly based on soybean-derived biodiesel fuel. For the pure diesel fuel (B0), a reduced single component reac-
Soybean biodiesel is mainly composed of five methyl esters: tion mechanism for n-heptane fuel with 29 species and 52 reac-
methyl palmitate, methyl stearate, methyl oleate, methyl linoleate tions, which was developed by Patel et al. [42], was used to
and methyl linolenate. As seen in Table 2, methyl palmitate and calculate the detailed chemical kinetics of fuel/air mixture burning.
methyl stearate are saturated, while methyl oleate, methyl linole- While then-heptane mechanism is used for reaction calculations in
ate, and methyl linolenate contain one, two, and three double- each cell, the thermo-physical properties of the fuel were repre-
bonds, respectively [39]. sented with those of tetra-decane (C14H30) due to the diesel and
For biodiesel reaction mechanism, multi-chemistry mechanism C14H30 similar thermo-physical specifications [43]. Fig. 3 shows
developed by Brakora at the ERC [40] using methyl decanoate (MD) the evolution of the main thermo physical properties of methyl
and methyl-9-decenoate (MD9D) as representatives of saturated oleate, methyl palmitate and tetradecane versus temperature.

(a) B100-Case 11 (b) B100-Case 12

(c) B100-Case 13
Fig. 10. Comparison of measured and predicted in-cylinder pressure and HRR histories for various B100 cases.
A. Zehni et al. / Applied Thermal Engineering 110 (2017) 1483–1499 1491

(a) B0 cases (b) B100 cases


Fig. 11. Numerical and measured exhaust NOx, CO and THC emissions for B0 and B100 cases over sweep of the SOI timings.
1492 A. Zehni et al. / Applied Thermal Engineering 110 (2017) 1483–1499

C2H2 (acetylene) is as an inception specie for Soot formation and 3.2. Simulation conditions
can be defined as an appropriate Soot surrogate [43,44]. Hence,
C2H2 is the representative of Soot behavior in the present work. All the simulations are carried out from IVC to EVO and they are
For the calculation of NOx formation, a 4 species (N, NO, N2O and based on the two operating conditions. For the case group 1 includ-
NO2) and 9 reaction NOx mechanism was used that has been ing case 1–3, fuel is diesel type and IMEP = 3 bar at the engine
reduced from the GRI-Mech 3.0 NOx mechanism [45] and added speed of 1500 rpm. The only goal of involving this case group is
to the diesel and biodiesel reaction mechanisms. more comprehensive validation of the PCCI combustion model at
the other engine loads and rpms. But, for the case group 2, the PCCI
3. Experimental and simulation procedure combustion modeling is considered for both the diesel (case 4–8)
and biodiesel (case 9–13) fuels at IMEP = 5.5 bar and engine speed
3.1. Engine specification of 2000 rpm and the comparison between the two fuels will be
performed in this condition. Engine operating conditions are given
The model is validated by experimental data of a GM 1.9L, 4- in Table 4. EGR percent is 67 for the both case groups which is
cylinder light duty diesel engine which was conducted at the uni- equivalent to 9.5% inlet O2concentration and it is calculated as
versity of Wisconsin-Madison Engine Research Center [32]. Table 3 the ratio of intake to exhaust CO2 concentration, as shown in
lists the engine and injector specifications. Eq. (7).
The measurement of pressure data was accomplished via a ½CO2 int
piezoelectric KISTLER 6125B pressure transducer. The resolution EGR% ¼  100 ð7Þ
½CO2 exh
of data was 0.1 CAD, corresponding to 3600 counts per revolution.
Exhaust emissions were measured using a California Analytical Injection profile is the same for all the cases of case group 1 and
Instruments bench capable of recording CO2, CO, O2 (Model 300), 2. Fig. 4 the injection profile which is modeled by extrapolation
HC (Model 300S-HFID) and NOx (Model 400-HCLD). from the measured injection rates [47].

Fig. 12. Spray penetration of B0 fuel with CO and fuel vapor contours at 4 CAD ASOI for the sweep of SOI timings.
A. Zehni et al. / Applied Thermal Engineering 110 (2017) 1483–1499 1493

3.3. Computational mesh and mesh independency the figures, SMD and spray tip penetration for B100 fuel are higher
compared to the B0 fuel which is due to the higher density, surface
Fig. 5 shows the computational mesh at different views used for tension and viscosity of B100 compared to the B0 fuel [46].
the present calculations. The full circle computational mesh gener- Consequently, as can be seen from Fig. 8, less spray break-up and
ation was carried out in the ANSYS ICEM CFD software. The average poor evaporation of the B100 case compared to theB0 case is
cell dimensions were 0.7–1.8 mm, 0.29–0.5 mm and4–5 mm in occurred. As expected, long liquid spray tip Penetrations produce
radial, vertical and azimuth directions, respectively. To validate a significant amount of wall wetting for the B100 cases.
the mesh independence, a finer mesh with the same engine geom- Figs. 9 and 10 reveal the numerical and measured in-cylinder
etry was created. The cylinder pressure curves were compared for pressure and heat release rate history for various B0 and B100
the condition of case 4 (see Fig. 6). It can be seen that with further cases. The experimental heat release rate is evaluated by means
refinements on the mesh, no significant difference was observed of a single zone diagnostic code on the basis of the in-cylinder
on the predicted in-cylinder pressure. By considering the computa- pressure distribution [48]. It can be observed that for all the cases,
tional time, the medium mesh was used for all the simulations. the model predictions of ignition delay, peak pressure and the
Using this mesh, for an Intel corei7 single processor with 12 GB maximum heat release rate locations and magnitudes are in
RAM, run times of simulation took about 12 and 7 h for B100 relatively good agreement with the corresponding experimental
and B0 cases, respectively. data. Slight discrepancies of calculated pressure during the
combustion and expansion phases which include maximum 3%
4. Results and discussion error relative to the measured data may be due to heat transfer
and crevice flow modeling [49].
Figs. 7 and 8 compare the spray tip penetration traces as well as Fig. 11 indicates the numerical and measured exhaust NOx, CO
evolution of spray tip penetration and sauter mean diameter and THC emissions as well as ISFC for the B0 and B100 cases over
(SMD) versus time for theB100 and B0 fuels under the conditions sweep of the SOI timings. Good agreements between the predicted
of SOI = 22 CAD ATDC in the case group 2. As can be seen from and experimental results are obtained. Low discrepancies of

Fig. 13. Spray penetration of B100 fuel with CO and fuel vapor contours at 4 CAD ASOI for the sweep of SOI timings.
1494 A. Zehni et al. / Applied Thermal Engineering 110 (2017) 1483–1499

measured and calculated results for THC and CO are likely due to fuels. Regarding ISFC, no considerable trend can be found from
not considering crevices volume. It can be seen that at 30 CAD the results and the amounts are relatively at the same level.
ATDC SOI timing, low levels of CO and THC are observed for both Fig. 12 shows the reasons for choosing 30 CAD ATDC SOI as the
B0 and B100 cases. Although, for this SOI timing NOx is relatively optimum injection timing case. In this figure, B0 spray penetration
higher compared to the other SOI timings, but, because total with CO and fuel vapor contours are shown at 4 CAD ASOI for the
exhaust NOx for all the cases is on the safe side (<0.3 g/kg-fuel), various SOI timings. Rectangle shape is used for marking the
30 CAD ATDC SOI timing can be regarded as an optimum SOI tim- area of the in-cylinder CO and liquid fuel distributions. Also,
ing for the lowest NOx + THC and CO emissions for B100 and B0 elliptical shape is utilized for defining the relatively high local CO

Fig. 14. Comparison of the numerical exhaust NOx, CO and THC emissions for the B0 and B100 cases over sweep of the SOI timings.

(a) B0 (b) B100


Fig. 15. Trends of numerical exhaust NOx and Soot results for the B0 and B100 cases over sweep of the SOI timing.
A. Zehni et al. / Applied Thermal Engineering 110 (2017) 1483–1499 1495

concentration regions. As can be observed, for 22 CAD ATDC case, bustion process in the proximity of the cold cylinder walls is likely
the SOI timing is so that fuel is over rich under the region of the the other reason for high CO and HC emissions especially at later
spray penetration trajectory which causes the formation of high crank angle degrees. Among the cases, 30 and 26 CAD ATDC
local CO concentrations (large elliptical shapes) as well as more SOI cases have better conditions related to the lowest levels of
spray bowl rim and squish wall impingement. Furthermore, for CO and HC. For the 30 CAD ATDC case, more oxygen availability
the 34 and 38 CAD ATDC SOI cases, in spite of small region of causes the more vaporization of the fuel parcels before impinging
high local CO concentration, in-cylinder cold wall impingement is to the bowl rim. Hence, number of fuel parcels in the squish region
occurred which causes the wall film formation and increase of are lower for the 30 CAD ATDC case compared to the 26 CAD
CO and HC levels (large rectangle shapes). Quenching of the com- ATDC case. Furthermore, the remaining parcels for the 30 CAD

Fig. 16. Comparison of calculated in-cylinder pressure, HRR, temperature, CH2O, OH, O2, O, N2O, NOx, THC, CO and Soot history for the B100 and B0 fuels, SOI = 30 CAD
ATDC.
1496 A. Zehni et al. / Applied Thermal Engineering 110 (2017) 1483–1499

ATDC case are oxidized in the squish and bowl regions at the later mechanism as well as the higher in-cylinder temperature in the
crank angle degrees due to the relatively higher in-cylinder tem- expansion stroke. It should be notified that in the LTC combustion,
perature. As shown in Fig. 13, this finding is consistent for the the role of the thermal NOx mechanism at temperatures below
B100 cases, too. approximately 1800 K is lowered compared to the N2O-
Fig. 14 compares the numerical exhaust NOx, CO and THC emis- intermidiate mechanism [54].
sions as well as ISFC for the B0 and B100 cases over sweep of the Fig. 16i shows that in spite of the higher THC formation stage
SOI timings. As can be seen, for the various SOI timings, NOx, CO for the B100 case which is the consequence of higher spray pene-
and THC emissions for the B100 cases are lower compared to the tration and wall impingement for the B100 case, the oxidation pro-
B0 cases. But, ISFC of B100 cases is higher compared to the B0 cases cess compensates for the higher THC concentration compared to
which is due to the lower heating value of biodiesel fuel compared the B0 case in the expansion stroke. Furthermore, Fig. 16j and k
to the diesel fuel [18]. In other words, with the same indicated reveal that for the B100 case, the processes of CO and Soot forma-
work or IMEP, more B100 fuel should be injected compared to tion is lower and their oxidation is higher compared to the B0 case.
the B0 fuel and consequently fuel consumption is increased. For The cause stems from the higher cylinder temperature in the HTHR
the optimum case, NOx, CO, THC and ISFC for the B100 case are combustion phase as well as higher in-cylinder oxygen for the
lowered by 13%, 27% and 12%, and increased by 6%, respectively B100 case compared to the B0 case.
compared to the B0 case. Fig. 16l shows the indicated work for both B100 and B0 cases. As
Fig. 15 plots the trends of numerical exhaust NOx and soot can be seen, indicated work for the B100 case is lower compared to
results for the B0 and B100 cases over sweep of the SOI timing. B0 case. Thermal efficiency is a parameter which is the ratio of the
NOx and Soot trade-off can be seen for some cases by tracing the indicated work to the heat supplied to it. Table 5 shows the param-
trends of these two emissions versus SOI timing. Hence, it can be eters which are utilized for obtaining a qualitative comparison
concluded that for some cases PCCI combustion cannot defeat between the thermal efficiency of different B100 and B0 cases.
the NOx/Soot trade-off completely. The experimental results of The calculations show that for all the cases, thermal efficiency for
Refs. [50,51] confirm the explanations. It is also worth noting that B100 cases is lower compared to B0 cases. Also, for the optimum
at 30 CAD ATDC, minimum soot is observed for the both cases cases 6 and 11, B100 and B0 contain 48.06% and 43.47% (9.5% dif-
which is due to the similar behavior of soot and CO and this con- ference) of thermal efficiency, respectively.
clusion will be explained with more details in this paper. Fig. 17 demonstrates the liquid spray penetration and Contour
Fig. 16 indicates the history of combustion parameters, impor- plots of temperature, oxygen, CH2O, OH, O, N2O, NO, Fuel vapor,
tant reactive radicals and emissions of B100 and B0 fuels for the CO and Soot species at 10 CAD ATDC (time of LTHR), 0 CAD ATDC
optimum case. Fig. 16a shows that HRR has two peaks for both (time of HTHR) and 10 CAD ATDC for B100 and B0 cases at
B100 and B0 fuels andB100 fuel experiences identical combustion SOI = 30 CAD ATDC.
stages with the B0 fuel. The first peak is related to the low temper- At 10 CAD ATDC, it can be observed that both fuels appear to
ature heat release (LTHR) or cool flame which is similar to the HCCI have parcels caught on the bowl rim. The least volatile fuel, B100,
type and corresponds to 830 K in-cylinder temperature for both has the most parcels remaining. Another point is that in-cylinder
B100 and B0 cases. Also, the second peak of HRR is due to the high temperature is relatively homogenous for the both cases. Also,
temperature heat release (HTHR) which is harmonized with the in- CH2O with high concentration as well as OH radical with low con-
cylinder temperature of1330 K and 1347 K for B100 and B0 cases, centration are formed corresponding to the oxygen consumed
respectively. The ignition delay of the low temperature heat regions. The dependence of NO formation region on N2O and O rad-
release which is the duration between the start of injection and ical is more vivid for the B0 case compared to the B100 case. As
onset of LTHR [52] is 12 CAD and 12.2 CAD for B0 and B100 cases, expected, CO is formed in the regions where the in-cylinder oxygen
respectively. Moreover, as can be seen, pressure and temperature is low. Also, there is an amazing similarity between CO and Soot
rise for B100 case are advanced compared to the B0 case which formation and oxidation zones. The reason is due to the C2H2
can be due to the formaldehyde (CH2O) and hydroxyl (OH) radicals oxidation mechanisms i.e. C2H2 + OH = CH3 + CO and C2H2 + O =
formation. Formaldehyde and hydroxyl radicals are the symptoms CH2 + CO in the fuel rich region which indicates that in the
of LTHR and HTHR onset, respectively [53]. The slight decrease of
Table 5
ignition delay for B0 case is related to the advancing of CH2O for- Comparison of thermal efficiency between different B100 and B0 cases.
mation. On the other hand, since formation of OH radical for
B100 case is advanced compared to the B0 case, peak of HTHR is
advanced as well. The time of second peak of HRR for B100 and
B0 cases corresponds to 1.4 CAD ATDC and 1.4 CAD ATDC for
B100 and B0 cases, respectively. In the current work, SOC is defined
as the crank angle degree where 10 percent of cumulative heat is
released. SOC for B100 and B0 cases are occurred at 10.61 and
6.74 CAD ATDC, respectively. In other words, SOC for B100 case
is advanced nearly 4 CAD compared to B0 case. As shown in
Fig. 16e, advance of HTHR for the B100 case leads to accelerate
the in-cylinder oxygen consumption rate from 18 CAD ATDC up
to 5.2 CAD ATDC which is consistent with peak of OH concentra-
tion. However, after 5.2 CAD ATDC, in-cylinder oxygen concentra-
tion for the B100 case remains higher in comparison with the B0
case which is due to the oxygenated structure of the B100 fuel.
Fig. 16h shows that NOX formation history for the B100 case is
higher compared to the B0 case. Moreover, NOx formation reac-
tions are frozen at 44 CAD and 26 CAD for B0 and B100 cases,
respectively. The main reason is due to the relatively higher O rad-
ical and N2O specie concentrations (Fig. 16f and g) of B0 case for
the forward reaction of N2O + O = 2NO in the N2O-intermidiate
A. Zehni et al. / Applied Thermal Engineering 110 (2017) 1483–1499 1497

-10 CAD ATDC 0 CAD ATDC 10 CAD ATDC


B0 B100 B0 B100 B0 B100

Temp
O2
CH2O
OH
O
N2O
NO
Fuelvapor
CO
Soot

Fig. 17. Spray tip penetration and Contour plots of Temperature, O2, NO, CO and fuel vapor (HC) for B100 and B0 at various crank angle degrees, SOI = 30 CAD ATDC.

presence of O and OH radicals the conversion of Soot to CO is pos- tion CH2O + C2H3 = C3H4 + OH which causes the similarity between
sible. Also, the similarity of CH2O contour with fuel vapor contour CH2O and OH during HTHR period [57].
is in agreement with the work of Lachaux and Musculus [55] which Total energy, which is released during oxidation for the B0 case,
showed that formaldehyde tracks well with the mixture of hydro- the more dependency of NO on N2O and O radical is remained like
carbons formed during the early stage reactions. LTHR condition. At this point, NO and fuel vapor pervade into the
At 0 CAD ATDC, high temperature region with high rate of oxy- squish region. Also, for the B0 case, the entire bowl of the cylinder
gen consumption in the bowl region is obvious for the both cases. is filled with CO whilst, for the B100 case it is partially filled.
In other words, areas of low oxygen correspond to the region of the Dominant factor for the lower CO and Soot concentration of B100
spray where fuel oxidation occurs and temperature is increased. case seems to be from the side of higher local temperature
For the both cases, a trade-off is seen for the CH2O and OH concen- compared to the B0 case.
tration regions between HTHR and LTHR times,as each specie At 10 CAD ATDC, local in-cylinder temperature of the B0 case
remains on opposite sides of the combustion chamber at 10 tends to outpace the temperature of the B100 case. Higher oxygen
CAD ATDC and TDC, respectively. The reason is that formaldehyde availability for the B100 case which is the consequence of higher
formed during low-temperature reactions, is consumed as the oxygen content of B100 fuel assists fuel vapor, CO and Soot oxida-
combustion process transitions to the second stage ignition [56]. tion up to 40 CAD ATDC (time of freezing oxidation mechanisms).
On the other hand, CH2O is converted to OH via the following reac- As can be observed, CH2O, OH and O radicals are settled in the
1498 A. Zehni et al. / Applied Thermal Engineering 110 (2017) 1483–1499

squish and bottom of the piston bowl following the remained oxy- [2] Z. Jia, I. Denbratt, Experimental Investigation of Natural Gas-Diesel Dual-Fuel
RCCI in a Heavy-Duty Engine, SAE Paper 2015-01-0838, 2015.
gen regions.
[3] J. Benajes, S. Molina, A. García, E. Belarte, M. Vanvolsem, An investigation on
RCCI combustion in a heavy duty diesel engine using in-cylinder blending of
5. Conclusions diesel and gasoline fuels, Appl. Therm. Eng. 63 (2014) 66–76.
[4] Y. Li, M. Jia, Y. Chang, Y. Liu, M. Xie, T. Wang, L. Zhou, Parametric study and
optimization of a RCCI (reactivity controlled compression ignition) engine
The aim of the present work was to perform a detail comparison fueled with methanol and diesel, Energy 65 (2014) 319–332.
of PCCI combustion, performance and emission characteristics of [5] R. Kiplimo, E. Tomita, N. Kawahara, S. Yokobe, Effects of spray impingement,
injection parameters, and EGR on the combustion and emission characteristics
pure biodiesel (B100) and diesel (B0) fuel surrogates by of a PCCI diesel engine, Appl. Therm. Eng. 37 (2012) 165–175.
multi-dimensional KIVA-3V code coupled with Chemkin. Reduced [6] G. Di Blasio, Premixed Combustion in a Light Duty Compression Ignition
mechanisms for n-heptane (diesel surrogate) and mixture of 50% Engine Through Fuel and Injection System Design: An Experimental Approach
PhD Dissertation, University of Napoli federico II, 2011.
n-heptane, 25% MD and 25% MD9d (biodiesel surrogates) were [7] M. Han, D.N. Assanis, S.V. Bohac, Sources of hydrocarbon emissions from low-
implemented to study the combustion and emission characteris- temperature premixed compression ignition combustion from a common rail
tics for the both fuels. The main conclusions were made as follows: direct injection diesel engine, Combust. Sci. Technol. 181 (2009) 496–517.
[8] T. Kanda, T. Hakozaki, T. Uchimoto, J. Hatano, N. Kitayama, H. Sono, PCCI
Operation with Early Injection of Conventional Diesel Fuel, SAE Paper 2005-01-
1. SMD and spray tip penetration for B100 fuel are higher com- 0378, 2005.
pared to the B0 fuel which is due to the higher density, surface [9] L. Cao, H. Su, S. Mosbach, M. Kraft, M. Bhave, S. Kook, C. Bae, Studying the
Influence of Direct Injection on PCCI Combustion and Emissions at Engine Idle
tension and viscosity of B100 compared to the B0 fuel.
Condition using Two Dimensional CFD and Stochastic Reactor Model, SAE
2. -30 CAD ATDC SOI timing can be regarded as an optimum SOI Paper 2008-01-0021, 2008.
timing for the lowest NOx + THC and CO emissions for B100 [10] T. Fang, R.E. Coverdill, C.-fF. Lee, R.A. White, Effects of injection angles on
and B0 cases. For the optimum case, NOx, CO, THC, thermal effi- combustion processes using multiple injection strategies in an HSDI diesel
engine, Fuel 87 (2008) 3232–3239.
ciency and ISFC for the B100 case were lowered by 13%, 27%, [11] M.P.B. Musculus, T. Lachaux, L.M. Pickett, C.A. Idicheria, End-of-Injection
12%, 9.5% and increased by 6%, respectively compared to the Overmixing and Unburned Hydrocarbon Emissions in Low-Temperature
B0 case. Combustion Diesel Engines, SAE Paper 2007-01-0907, 2007.
[12] I.W. Ekoto, W.F. Colban, P.C. Miles, S.W. Park, D.E. Foster, R.D. Reitz, UHC and
3. For some cases PCCI combustion cannot defeat the NOx/Soot CO Emissions Sources from a Light-Duty Diesel Engine Undergoing Dilution
trade-off completely. Controlled Low-Temperature Combustion, SAE Paper 2009-24-0043, 2009.
4. The first peak of the heat release rate is related to the low tem- [13] J.E. Parks, V. Prikhodko, J.M.E. Storey, T.L. Barone, S.A. Lewis, M.D. Kass, S.P.
Huff, Emissions from premixed charge compression ignition (PCCI)
perature heat release (LTHR) or cool flame which is similar to combustion and effect on emission control devices, Catal. Today 151 (2010)
the HCCI type and corresponds to 830 K in-cylinder tempera- 278–284.
ture for the both B100 and B0 cases. Also, the second peak of [14] S. Imtenan, M. Varman, H.H. Masjuki, M.A. Kalam, H. Sajjad, M.I. Arbab, I.M.
Rizwanul Fattah, Impact of low temperature combustion attaining strategies
HRR is due to the high temperature heat release (HTHR) which on diesel engine emissions for diesel and biodiesels: a review, Energy Convers.
is harmonized with the in-cylinder temperature of 1330 K and Manage. 80 (2014) 329–356.
1347 K for B100 and B0 cases, respectively. [15] M. Zheng, M.C. Mulenga, G.T. Reader, M. Wang, D.S.K. Ting, J. Tjong, Biodiesel
engine performance and emissions in low temperature combustion, Fuel 87
5. Advance of second peak of HRR as well as in-cylinder pressure
(2008) 714–722.
and temperature rise for the B100 case is due to advance of [16] S. Lee, J. Jang, S. Oh, Y. Lee, J. Kim, K. Lee, Comparative Study on Effect of Intake
OH radical formation for the B100 case compared to the B0 case. Pressure on Diesel and Biodiesel Low Temperature Combustion Characteristics
6. Higher O and N2O specie concentrations of B0 case for the for- in a Compression Ignition Engine, SAE Paper 2013-01-2533, 2013.
[17] Ö. Can, E. Öztürka, H. Solmazb, F. Aksoyc, C. Çinarb, H.S. Yücesu, Combined
ward reaction of N2O + O = 2NO reaction in the N2O- effects of soybean biodiesel fuel addition and EGR application on the
intermidiate mechanism explains for the higher concentration combustion and exhaust emissions in a diesel engine, Appl. Therm. Eng. 95
of NOx for the B0 case compared to the B100 case. (2016) 115–124.
[18] S.M. Palash, H.H. Masjuki, M.A. Kalam, B.M. Masum, A. Sanjid, M.J. Abedin,
7. Despite of the higher THC formation stage for the B100 case State of the art of NOx mitigation technologies and their effect on the
which is the consequence of higher spray penetration and wall performance and emission characteristics of biodiesel-fueled compression
impingement for the B100 case, the oxidation process compen- ignition engines, Fuel 76 (2013) 400–420.
[19] E. Sadeghinezhad, S.N. Kazi, A. Badarudin, C.S. Oon, M.N.M. Zubir, M. Mehrali, A
sates for the higher THC concentration compared to the B0 case comprehensive review of bio-diesel as alternative fuel for compression
in the expansion stroke. Besides, for the B100 case, the process ignition engines, Renew. Sust. Energy Rev. 28 (2013) 410–424.
of CO and Soot formation is lower and their oxidation is higher [20] F.J. Jiménez-Espadafor, M. Torres, J.A. Velez, E. Carvajal, J.A. Becerra,
Experimental analysis of low temperature combustion mode with diesel and
compared to the B0 case. biodiesel fuels: a method for reducing NOx and Soot emissions, Fuel Process.
8. There is an amazing similarity between CO and Soot formation Technol. 103 (2012) 57–63.
and oxidation zones. The reason is due to the C2H2 oxidation [21] N.P. Komninos, C.D. Rakopoulos, Modeling HCCI combustion of biofuels: a
review, Renew. Sust. Energy Rev. 16 (2012) 1588–1610.
mechanisms i.e. C2H2 + OH = CH3 + CO and C2H2 + O = CH2 + CO
[22] Q. Fang, J. Fang, J. Zhuang, Z. Huang, Effects of ethanol diesel biodiesel blends
in the fuel rich region which indicates that in the presence of on combustion and emissions in premixed low temperature combustion, Appl.
O and OH radicals the conversion of Soot to CO is possible. Therm. Eng. 54 (2013) 541–548.
9. For B100 and B0 cases, a trade-off is seen for the CH2O and OH [23] J. Su, H. Zhu, S.V. Bohac, Particulate matter emission comparison from
conventional and premixed low temperature combustion with diesel,
concentration regions between HTHR and LTHR times, as each biodiesel and biodiesel–ethanol fuels, Fuel 113 (2013) 221–227.
specie remains on opposite sides of the combustion chamber [24] B.T. Tompkins, H. Song, T.J. Jacobs, Low Temperature Heat Release of Palm and
at 10 CAD ATDC and TDC, respectively. Soy Biodiesel in Late Injection Low Temperature Combustion. SAE Paper 2014-
01-1381, 2014.
[25] Y. Jung, J. Hwang, C. Bae, Assessment of particulate matter in exhaust gas for
biodiesel and diesel under conventional and low temperature combustion in a
Acknowledgement compression ignition engine, Fuel 165 (2016) 413–424.
[26] C.D. Rakopoulos, K.A. Antonopoulos, D.C. Rakopoulos, Multi-zone modeling of
The financial support for this research study by National Iranian diesel engine fuel spray development with vegetable oil, bio-diesel or diesel
fuels, Energy Convers. Manage. 47 (2006) 1550–1573.
Oil, Refinery and Distribution Company (NIORDC) is acknowledged.
[27] C.D. Rakopoulos, K.A. Antonopoulos, D.C. Rakopoulos, Development and
application of multi-zone model for combustion and pollutants formation in
References direct injection diesel engine running with vegetable oil or its bio-diesel,
Energy Convers. Manage. 48 (2007) 1881–1901.
[28] A.S. Kuleshov, Multi-Zone Diesel Spray Combustion Model for Thermodynamic
[1] Q. Fang, J. Fang, J. Zhuang, Z. Huang, Influences of pilot injection and exhaust
Simulation of Engine with PCCI and High EGR Level, SAE Paper 2009-01-1956,
gas recirculation (EGR) on combustion and emissions in a HCCI-DI combustion
2009.
engine, Appl. Therm. Eng. 48 (2012) 97–104.
A. Zehni et al. / Applied Thermal Engineering 110 (2017) 1483–1499 1499

[29] Y. Lee, K. Jang, K. Han, K.Y. Huh, S. Oh, Simulation of a Heavy Duty Diesel [44] R. Solsjö, Large Eddy Simulation of Turbulent Combustion in PPC and Diesel
Engine Fueled with Soybean Biodiesel Blends in Low Temperature Engines PhD Dissertation, University of Lund, 2014.
Combustion, SAE Paper 2013-01-1100, 2013. [45] http://combustion.berkeley.edu/gri-mech/version30/text30.html.
[30] Z. Luo, M. Plomer, T. Lu, S. Som, D.E. Longman, S.M. Sarathy, W.J. Pitz, A [46] I.W. Ekoto, W.F. Colban, P.C. Miles, S. Park, D.E. Foster, R.D. Reitz, Sources of
reduced mechanism for biodiesel surrogates for compression ignition engine UHC Emissions from a Light-Duty Diesel Engine Operating in a Partially
applications, Fuel 92 (2012) 143–153. Premixed Combustion Regime, SAE Paper 2009-01-1446, 2009.
[31] Y. Chang, M. Jia, Y. Li, Y. Zhang, M. Xie, H. Wang, R.D. Reitz, Development of a [47] F. Perini, D. Sahoo, P.C. Miles, R.D. Reitz, Modeling the Ignitability of a Pilot
skeletal oxidation mechanism for biodiesel surrogate, Proc. Combust. Inst. 35 Injection for a Diesel Primary Reference Fuel: Impact of Injection Pressure,
(2015) 3037–3044. Ambient Temperature and Injected Mass, SAE Paper 2014–01-1258, 2014.
[32] J. Brakora, R. Reitz, A Comprehensive Combustion Model for Biodiesel-Fueled [48] M. Fathi, R. Khoshbakhti Saray, M.D. Checkel, Detailed approach for apparent
Engine Simulations, SAE Paper 2013-01-1099, 2013. heat release analysis in HCCI engines, Fuel 89 (2010) 2323–2330.
[33] A.A. Amsden, KIVA-3V: A Block-Structured KIVA Program for Engines with [49] C.D. Rakopoulos, G.M. Kosmadakis, A.M. Dimaratos, E.G. Pariotis, Investigating
Vertical or Canted Valves, LA-13313-MS; 1997. the effect of crevice flow on internal combustion engines using a new simple
[34] R.J. Kee, F.M. Rupley, J.A. Miller, CHEMKIN-II: A FORTRAN Chemical Kinetics crevice model implemented in a CFD code, Appl. Energy 88 (2011) 111–126.
Package for the Analyses of Gas Phase Chemical Kinetics, Sandia Report 1989; [50] H. Zhu, S.V. Bohac, K. Nakashima, L.M. Hagen, Z. Huang, D.N. Assanis, Effect of
SAND 89-8009. fuel oxygen on the trade-offs between soot, NOx and combustion efficiency in
[35] P.N. Brown, G.D. Byrne, A.C. Hindmarsh, VODE: a variable coefficient ODE premixed low-temperature diesel engine combustion, Fuel 112 (2013) 459–
solver, SIAM J. Sci. Stat. Comput. 10 (1989) 1038–1051. 465.
[36] Z. Han, R.D. Reitz, Turbulence modeling of internal combustion engines using [51] B.T. Tompkins, T.J. Jacobs, Low Temperature Combustion with Biodiesel: The
RNG k-e models, Combust. Sci. Technol. 106 (1995) 267–295. Role of Oxygenation in Improving Efficiency and Emissions Spring Technical
[37] Z. Han, R.D. Reitz, A temperature wall function formulation for variable- Meeting of the Central States Section of the Combustion Institute April 22–24,
density turbulent flows with application to engine convective heat transfer 2012.
modeling, Int. J. Heat Mass Transf. 40 (1997) 613–625. [52] S. Cong, G.P. McTaggart-Cowan, C.P. Garner, Wahab, M. Peckham,
[38] J.C. Beale, R.D. Reitz, Modeling spray atomization with the Kelvin-Helmholtz/ Experimental investigation of low temperature diesel combustion processes,
Rayleigh-Taylor hybrid model, Atom. Sprays 9 (1999) 623–650. Combust. Sci. Technol. 183 (2011) 1376–1400.
[39] V.I. Golovitchev, J. Yang, Construction of combustion models for rapeseed [53] R. Collin, J. Nygren, M. Richter, M. Aldén, Simultaneous OH- and
methyl ester bio-diesel fuel for internal combustion engine applications, Formaldehyde-LIF Measurements in an HCCI Engine, SAE Paper 2003-01-
Biotechnol. Adv. 27 (2009) 641–655. 3218, 2003.
[40] J.L. Brakora, A Comprehensive Combustion Model for Biodiesel-Fueled Engine [54] S.R. Turns, An Introduction to Combustion: Concepts and Applications,
Simulations PhD Dissertation, University of Wisconsin-Madison, 2012. McGraw-Hill, New York, 2012, pp. 149–178.
[41] H. An, W.M. Yang, A. Maghbouli, S.K. Chou, K.J. Chua, Detailed physical [55] T. Lachaux, M.P.B. Musculus, In-cylinder unburned hydrocarbon visualization
properties prediction of pure methyl esters for biodiesel combustion during low-temperature compression-ignition engine combustion using
modeling, Appl. Energy 102 (2013) 647–656. formaldehyde PLIF, Proc. Combust. Inst. 31 (2) (2007) 2921–2929.
[42] A. Patel, S.C. Kong, R.D. Reitz, Development and Validation of a Reduced [56] S.L. Kokjohn, D.A. Splitter, R.M. Hanson, R.D. Reitz, Modeling charge
Reaction Mechanism for HCCI Engine Simulations, SAE Paper 2004-01-0558, preparation and combustion in diesel fuel, ethanol, and dual-fuel PCCI
2004. engines, in: ILASS-Americas 22nd Annual Conference on Liquid Atomization
[43] R. Opat, Y. Ra, M.A. Gonzalez, R. Krieger, R.D. Reitz, D.E. Foster, R.P. Durrett, R. and Spray Systems, Cincinnati, OH, May 2010.
M. Siewert, Investigation of Mixing and Temperature Effects on HC/CO [57] M.P.B. Musculus, T. Lachaux, Visualization of UHC Emissions for Low-
Emissions for Highly Dilute Low Temperature Combustion in a Light Duty Temperature Diesel Engine Combustion, Diesel Engine Emissions Reduction
Diesel Engine, SAE Technical Paper 2007-01-0193; 2007. Conference Marriott Renaissance Center, Detroit, MI, 2006.

You might also like