You are on page 1of 39

Accepted Manuscript

Effect of compression ratio on performance, combustion and emissions


characteristics of compression ignition engine fuelled with jojoba methyl ester

Meshack Hawi, Ahmed Elwardany, Shinichi Ookawara, Mahmoud Ahmed

PII: S0960-1481(19)30525-7
DOI: https://doi.org/10.1016/j.renene.2019.04.041
Reference: RENE 11465

To appear in: Renewable Energy

Received Date: 4 September 2018


Revised Date: 3 March 2019
Accepted Date: 9 April 2019

Please cite this article as: Hawi M, Elwardany A, Ookawara S, Ahmed M, Effect of compression ratio
on performance, combustion and emissions characteristics of compression ignition engine fuelled with
jojoba methyl ester, Renewable Energy (2019), doi: https://doi.org/10.1016/j.renene.2019.04.041.

This is a PDF file of an unedited manuscript that has been accepted for publication. As a service to
our customers we are providing this early version of the manuscript. The manuscript will undergo
copyediting, typesetting, and review of the resulting proof before it is published in its final form. Please
note that during the production process errors may be discovered which could affect the content, and all
legal disclaimers that apply to the journal pertain.
ACCEPTED MANUSCRIPT
Effect of Compression Ratio on Performance, Combustion and Emissions Characteristics of
Compression Ignition Engine Fuelled with Jojoba Methyl Ester

Meshack Hawi1, Ahmed Elwardany1,2 and Shinichi Ookawara1,3, Mahmoud Ahmed1,4,*


1
Energy Resources Engineering Department, Egypt- Japan University of Science and
Technology (E-JUST), P.O. Box 179-21934, New Borg El-Arab City, Alexandria,
Egypt,
meshack.ochieng@ejust.edu.eg
2
Mechanical Engineering Department, Faculty of Engineering, Alexandria University,

PT
Alexandria 21544, Egypt
ahmed.elwardany@alexu.edu.eg
3
Department of Chemical Science and Engineering, School of Materials and Chemical

RI
Technology, Tokyo Institute of Technology, S1-7, 2-12-1 Ookayama,
Meguro-ku, Tokyo 152-8552 Japan,
sokawara@chemeng.titech.ac.jp

SC
4
Mechanical Engineering Department, Faculty of Engineering, Assiut University, Assiut
71516, Egypt
*
Corresponding author mahmoud.ahmed@ejust.edu.eg and/ or aminism@aun.edu.eg
(M.Ahmed)

U
AN
Abstract
1 The major challenges facing the energy sector are the cost of fossil fuel and its depletion.
M

2 Therefore, the current work presents an experimental investigation on the effect of different
3 blends of jojoba methyl ester (JME) in diesel engine performance, combustion and emissions
D

4 characteristics. A numerical investigation on the effect of compression ratio (CR) on using


TE

5 neat JME is also presented. Thermophysical properties of JME and raw jojoba oil are
6 measured and characterized by FT-IR and GC-MS analysis. Engine performance parameters,
7 combustion characteristics and emission characteristics are measured for single cylinder,
EP

8 four-stroke, direct injection diesel engine fueled with diesel and different blends of JME in
9 diesel (5%, 10%, and 20% by volume). A comprehensive numerical setup using ANSYS
C

10 FORTE code is developed and validated against new measurements. Results illustrated that
AC

11 increasing CR from 18 to 23 peak incylinder pressure is increased. Additionally, increasing


12 CR leads to higher NOx and CO, UHC concentrations for diesel and JME100. The reduction
13 in peak incylinder pressure results from using JME100 instead of diesel could be recovered
14 by increasing CR from 21.5 to 23. This highlights the possibility of using neat JME in
15 compression ignition engines with less emissions and minimal losses in the output power.

16 Keywords: diesel engine; compression ratio; jojoba methyl ester; engine simulation;
17 combustion characteristics.

Page 1 of 37
ACCEPTED MANUSCRIPT
18 1. Introduction

19 Compression ignition (CI) engines are widely used in transport, construction, mining, power
20 generation and agricultural sectors due to their durability and higher thermal efficiency
21 compared to spark ignition (SI) engines [1]. However, CI engines are currently facing two
22 major challenges including rise in the cost of diesel fuel and increasingly stringent legislation
23 on emission of oxides of nitrogen (NOx) and particulate matter (PM). To address these two

PT
24 issues, there is considerable research effort by different researchers on production and
25 utilization of renewable alternative fuels in CI engines [2,3]. Among the leading renewable

RI
26 alternative fuels for CI engines is biodiesel. Biodiesel obtained from different vegetable oils
27 including soybean oil, castor oil, rapeseed oil, jatropha curcas oil, sunflower oil, karanja oil,

SC
28 rubber seed oil, jojoba oil, among others, could be produced locally, and at relatively low
29 cost. Additionally, it generates less pollutants to the environment compared to those from

U
30 mineral diesel [4].
AN
31 Biodiesel could be used directly in the diesel engine with minimal modifications in the
32 engine design [5,6]. However, some recent studies have reported that for better performance
M

33 of diesel engine when operated with alternative fuels, the engine design parameters should be
34 investigated [1,4]. A number of studies [7,8] have also proposed development of fully
dedicated biodiesel engines with optimized design parameters before large-scale substitution
D

35
36 of petroleum diesel with biodiesel can be implemented globally. In an experimental study on
TE

37 performance of a diesel engine fueled with neat rapeseed oil and its blends with mineral
38 diesel, Buyukkaya [9] noted impressive performance with neat biodiesel, though with a slight
EP

39 increase of about 12 % in NOx emission. It was suggested that further investigations on the
40 effect of engine design parameters such as injection timing and duration is necessary for
41 better combustion of biodiesel in diesel engines. Hamid et al. [10] demonstrated through an
C

42 extensive numerical investigation of in-cylinder air flow, that modification of the combustion
AC

43 chamber could significantly improve CI engine performance with emulsified biodiesel.

44 Given the variation in chemical and physical properties of biodiesel from standard diesel, the
45 combustion characteristics may also vary significantly [11]. Therefore, the design parameters
46 of the engine need to be adjusted according to the physicochemical properties and
47 combustion characteristics of the fuel to obtain the optimum performance of the engine [12].
48 Compression ratio is one of the design parameters which greatly affects the engine
49 performance [13]. Laguitton et al. [14] showed through experimental investigation that soot

Page 2 of 37
ACCEPTED MANUSCRIPT
50 and NOx emissions could be reduced in CI engines by lowering the CR. Raheman and
51 Ghadge [15] investigated the effect of CR on CI engine fueled with Mahua oil and its blends
52 with mineral diesel. It was reported that an increase in CR from 18 to 20 led to a significant
53 increase in brake thermal efficiency (BTE) by up to 33 %. A decrease in brake specific fuel
54 consumption (BSFC) by up to 10.7 % and 19.3 % was also recorded for diesel and biodiesel,
55 respectively, when the CR was increased from 18 to 19.

PT
56 Jindal et al. [16] experimentally studied the effect of CR on performance of a direct injection
57 (DI) diesel engine fueled with jatropha methyl ester. Their study indicated that increasing the

RI
58 CR from the standard setting of 17.5 to 18 and the injection pressure from 210 to 250 bar
59 resulted in an increase in the BTE by 8.2 %. In an extensive experimental investigation by

SC
60 Sayin and Gumus [17] on the effect of CR on performance and emission characteristics of a
61 DI diesel engine fueled with biodiesel-diesel blends, engine CR of 19, 18 and 17 were

U
62 considered. The study reported improved BSFC, BTE and brake specific energy consumption
63 (BSEC) with increase in CR from the original setting of 18 to 19. A reduction in carbon
AN
64 monoxide (CO) and unburned hydrocarbon (UHC) emissions with increase in CR was also
65 reported. In an experimental study by Muralidharan and Vasudevan [18] on the effect of CR
M

66 on the performance, combustion and emission characteristics of a variable compression ratio


67 (VCR) engine operated on waste cooking oil and its blends with mineral diesel, a superior
D

68 performance by B40 (40% biodiesel) blend as compared to mineral diesel at CR of 21 was


TE

69 reported.

70 EL_Kassaby and Nemit_allah [19] studied the effect of compression ratio on a diesel engine
EP

71 fueled with blends of biodiesel from waste cooking oil and standard diesel at 10 %, 20 %, 30
72 % and 50 % biodiesel (B10, B20, B30 and B50, respectively), and at three conditions of
73 compression ratio (14, 16 and 18). It was reported that a change of CR from 14 to 18 resulted
C

74 in 18.39 %, 27.48 %, 18.5 %, and 19.82 % increase in brake thermal efficiency in case of
AC

75 B10, B20, B30, and B50, respectively. A reduction in UHC and CO emissions by up to 52 %
76 and 37.2 %, respectively, was also reported. However, NOx emission increased by 36.84 %.
77 The study concluded that increasing the CR had more benefits for biodiesel fueled engine
78 than diesel fueled engine. From an elaborate literature review by Ghazali et al. [20] on effects
79 of biodiesel on CI engine performance and emissions, many authors reported an increase in
80 NOx emissions (5.6 – 52 %) when using biodiesel while several others showed that the
81 blended fuels lead to a reduction in NOx (5-35 %). Generally, most studies reported an
82 improvement in combustion and performance of CI engine with increase in compression

Page 3 of 37
ACCEPTED MANUSCRIPT
83 ratio, as well as considerable decrease in UHC, CO and smoke emissions [21,22]. However,
84 in most of the research investigations, NOx is reported to increase with the increase in CR
85 because of higher peak combustion temperatures [23].

86 The present study is motivated by the need to find alternative sources of energy which are
87 sustainable and environmentally friendly, in order to overcome the challenges of growing
88 global energy demand and environmental pollution resulting from the use of fossil fuels. The

PT
89 global energy demand is constantly increasing with the increase in world population and the
90 total proven oil reserves in the world, estimated by British Petroleum (BP) statistical review

RI
91 as 1707 billion barrels by 2016 could meet only approximately 50 years of global energy
92 demand [24]. Furthermore, it is predicted that by 2050 the global population will exceed 9

SC
93 billion and the demand for fuel is expected to rise significantly [25]. With increased use of
94 fossil fuels, a corresponding rise in the negative impacts of the pollutant emissions is

U
95 expected. Renewable energy presents a viable and sustainable solution to these challenges.
96 Biodiesel, as a renewable fuel, is considered a suitable alternative fuel for CI engine.
AN
97 Currently, several countries worldwide are encouraging production of biodiesel fuel,
98 especially from non-edible vegetable oils to avoid food security concerns [26,27].
M

99 Simmondsia chinensis, commonly known as Jojoba is one of the crops currently promoted by
100 governments for plantation for its renewable source of high quality oil [28]. Its seeds store
D

101 about 50 % of a light yellow, odorless liquid wax ester commonly referred to as Jojoba oil
TE

102 (JO) [28]. Cultivation of Jojoba requires less water, and its productivity in marginal lands is
103 high, hence it is considered a sustainable source of biodiesel. Therefore, Jojoba biodiesel has
104 been explored as a cheap, sustainable fuel that can serve as a substitute for petroleum diesel,
EP

105 hence significant research effort is currently ongoing to optimize its performance in CI
106 engines. However, to the best of our knowledge, the effect of compression ratio on
C

107 combustion, performance and emission characteristics of a direct injection CI engine has not
AC

108 been clearly investigated when using Jojoba oil methyl ester.

109 Most of research efforts to implement neat biodiesel in CI engines have been focused on
110 biodiesel from other different sources including waste cooking oil, jatropha biodiesel,
111 soybean methyl ester and rapeseed methyl ester. Therefore, in the present study, neat jojoba
112 methyl ester was considered as a potential alternative fuel for CI engine because it is
113 produced from non-edible oil. The main aim of this study is to investigate numerically the
114 effect of CR on combustion and emission characteristics of a CI engine fueled with neat
115 jojoba methyl ester compared to those of mineral diesel.

Page 4 of 37
ACCEPTED MANUSCRIPT
116 To carry out this investigation computational fluid dynamics (CFD) simulations are
117 conducted on a 3-D numerical model of a single cylinder four-stroke CI engine. The model is
118 validated using experimental data from tests performed on a single cylinder four-stroke diesel
119 engine using mineral diesel and its blends with JME produced from raw jojoba oil.
120 Experiments are conducted at a CR of 21.5 (on a constant CR engine), while in the
121 simulations four other CR settings are considered, including 18, 20, 22 and 23.

PT
122 2. Experimental setup and procedures

RI
123 2.1 System description

124 The experimental data for validation of the numerical model was obtained from the

SC
125 experimental setup shown in Fig. 1. The engine test rig (Model GUNT-CT100.22) consists of
126 a CI engine of the technical specifications presented in Table 1, an asynchronous motor

U
127 Model TFCP 132SB-2 connected to the engine for cranking and for torque measurement, air
128 flow and fuel flow measuring systems for measuring the air and fuel flowrates, respectively.
AN
129 More details about the engine test rig could be found in [29].
M
D
TE
C EP
AC

Page 5 of 37
ACCEPTED MANUSCRIPT
Exhaust gas
analyzer
GUNT Control Unit (CT 100.22)
Amplifier CO, UHC,
NOx
Speed Torque Temperature
4
7 Exhaust gases

12
3 5 Intake air
1 2
11

PT
Asynchronous motor 6
CI Engine
USB – DAQ Force sensor (Torque)
8
System
9 10

RI
PC with GUNT
Software

SC
13 14

U
measurement
Air flow
AN
Fuel flow

Fuel tank Air filter


Air box
M

Air flow (Quietening


vessel/ settling
Fuel pump with strainer tank)
D
TE

1 Electric motor switch, 2 Ignition switch, 3 Coolant pump switch, 4 Air consumption gauge, 5 Intake (negative)
EP

pressure gauge, 6 Fuel pump switch, 7 Fuel measuring tube, 8 Fuel consumption meter, 9 Speed sensor, 10 TDC
sensor, 11 Kistler pressure transducer, 12 Exhaust gas temperature sensor, 13 Emergency switch, 14 Main switch.
130
131 Fig. 1. Schematic of the engine experimental setup.
C

132
AC

133

134

135

136

137

138

139

Page 6 of 37
ACCEPTED MANUSCRIPT
140

141 Table 1. Specifications of the test engine


Engine parameter Specification
Engine model HATZ-1B30-2
Engine type Single cylinder 4-stroke direct injection CI
Bore (mm) 80
Stroke (mm) 69

PT
Crank radius (mm) 34.5
Connecting rod length (mm) 114.5
Displacement volume (cm3) 347
Compression ratio 21.5:1

RI
Rated power (kW/rpm) 5.5/3500
Idle speed (rpm) 1000
Type of cooling Air cooling

SC
Start-up Electrical
142

U
143 2.2 Experimental procedure AN
144 Experiments were conducted at a constant engine speed of 2000 rpm and a torque of 13.5
145 N.m, with varying fuel conditions of pure mineral diesel (D100) and biodiesel (Jojoba Methyl
146 Ester: JME) blends of 5 %, 10 %, and 20 % (JME5, JME10 and JME20, respectively). Air
M

147 flow rate into the engine was measured using orifice meter (diameter of 20.6 mm) while fuel
148 consumption was measured using flow meter sensor (Huba control type 680-out signal 0-10
D

149 VDC, Accuracy ±0.25 % FS). Engine brake torque was measured by a force sensor of Model
TE

150 FLINTEC ZLB-200Kg-C3. Cylinder pressure and engine speed were measured by Kistler
151 piezoelectric pressure sensor of Model 6052C (linked with a charge amplifier of Model
GUNT CT100.13) and a proximity sensor (WACHENDORFF of type PNP-N.O, Sn 4 mm,
EP

152
153 10-30VDC, and 200 mA), respectively. Rotation of the crank shaft was recorded by an
154 optical encoder while a proximity switch of model WACHENDORFF PNP-N.O with a
C

155 detecting distance of 4 mm was used in determining the position of the top dead center
AC

156 (TDC).

157 A high-speed data acquisition (DAQ) system (Model USB-AD16f) is used to collect data
158 from different sensors of the system. Three K-type thermocouples located at the intake port,
159 exhaust port and fuel line were used to measure the ambient air temperature, exhaust gas
160 temperature and fuel temperature, respectively. A desktop computer (PC) with LabVIEW
161 software (GUNT software) was used to analyze the output data. Engine emissions were
162 measured using exhaust gas emissions analyser, Model ECA 450. Emission measurements

Page 7 of 37
ACCEPTED MANUSCRIPT
163 were recorded in triplicate and the mean values evaluated and reported. The technical
164 specifications of the emission analyser are given in Table 2.

165 In-cylinder pressure data was recorded for fifty cycles. The average pressure was evaluated in
166 each case and used in calculation of the experimental heat release rate. Performance
167 parameters including torque, engine speed, fuel flowrate and air flowrate were also recorded
168 for fifty cycles and the mean values are calculated for determination of performance

PT
169 characteristics such as BSFC and BTE.

170 2.3 Uncertainty analysis

RI
171 Uncertainty can be defined as the magnitude of error in results or measurements [30]. In the

SC
172 present study, uncertainties in the calculated parameters were obtained from uncertainties in
173 measured quantities using Eq. (1) [31].

U
௡ ଶ ଶ
ܷ௒ 1 ߲ܻ
(1)
= ෍ ቈ൬ ܷ ൰ ቉
ܻ ܻ ܻ߲ܺ௜ ௑೔
AN
௜ୀଵ

174 where Y is the dependent parameter, calculated from measured parameters, ܺ௜ , ܷ௒ denotes
the uncertainty in Y and ܷ௑೔ denotes uncertainty in ܺ௜ . Table 3 gives the uncertainties of the
M

175

176 various instruments used in the present study, as well as percentage uncertainties of
D

177 calculated parameters like brake power, BSFC and BTE.


TE

178 Table 2. Specifications of the exhaust gas analyzer


Gas Measuring range Resolution Accuracy
EP

CO 0 – 4000 ppm 1 ppm ±5 % of reading or ±10 ppm


CO2 0 – 20 % vol. 0.1 % vol. ±0.5 % of reading
UHC 0 – 10 % vol. 0.01 % vol. ±0.3 % of reading
O2 0 – 20.9 % vol. 0.01 % vol. ±0.3 % of reading
C

NOx 0 – 4000 ppm 1 ppm ±5 % of reading or ±5 ppm


-20 to 1315oC 1oC ±2oC
AC

Stack temperature
Probe tip temperature 800oC max - -

179 Table 3. Experimental uncertainties


Instrument Range Accuracy Uncertainty
Torque indicator, N.m 0 - 200 ±1% of reading ±1
Fuel burette, cc 153 ±0.2 ±1
Speed sensor, rpm 0 - 10000 ±5 rpm ±0.1
Exhaust gas analyzer
CO, ppm 0 – 4000 ±10 ppm ±1
UHC, %vol 0 – 10 ±0.3 % of reading ±0.1

Page 8 of 37
ACCEPTED MANUSCRIPT
NOx, ppm 0 – 4000 ±5 ppm ±1
Pressure transducer, bar 250 ±1% of reading ±1
Crank angle encoder, degree 0 - 720 ±0.5 ±0.3
Brake power - - ±1
BSFC - - ±2
BTE - - ±3.2
180

181 2.4 Experimental heat release rate

PT
182 Heat release rate (HRR) is the rate at which energy stored in the chemical bonds of reactants
183 (fuel and oxidizer) is converted into sensible heat during combustion [13]. It is expressed in

RI
184 joules per crank angle degree (J/o). The net value (Net HRR) is first obtained from the
185 measured cylinder pressure data and the calculated cylinder volume. Losses are then

SC
186 approximated and added to the net HRR to obtain the gross HRR.

187 In this study, the experimental heat release rate is calculated using the first law-single zone
188 model equation as follows [13]:
U
AN
݀ܳ௚௥௢௦௦ ߛ(ܶ) ܸ݀ 1 ݀‫ܳ݀ ݌‬௪௔௟௟
= ×‫݌‬ + ×ܸ +
(2)
݀ߠ ߛ(ܶ) − 1 ݀ߠ ߛ(ܶ) − 1 ݀ߠ ݀ߠ
M

189

190 where the specific heat ratio ߛ(ܶ), is given by


D

ߛ(ܶ) = 1.35 − 6 × 10ିହ × ܶ + 10ି଼ × ܶ ଶ (3)


TE

191

192 With ܶ being the mean gas temperature.


EP

193 The mean gas temperature is calculated from the measured cylinder pressure and the
194 calculated cylinder volume using the equation of state as given by
C

ܶ௥ ‫ܸ݌‬
ܶ=
(4)
AC

‫݌‬௥ ܸ௥
195

196 The thermodynamic properties (pressure ‫݌‬௥ , temperature ܶ௥ , and volume ܸ௥ ) are estimated at
197 the reference condition of intake valve closing - IVC (ܲூ௏஼ , ܶூ௏஼ , ܸூ௏஼ ). ܶூ௏஼ and ܲூ௏஼ are the
198 temperature and the pressure at IVC, taken as 360 K and 1.013 ×105 Pa, respectively. The
199 contents of the cylinder are assumed to behave as an ideal gas (air) with the specific heat

Page 9 of 37
ACCEPTED MANUSCRIPT
200 varying with temperature [32]. The temperature and pressure of the combustion products are
201 also assumed to be uniform at any time during the combustion process.

202 Heat loss through cylinder wall is calculated using convection heat transfer equation, as
203 follows:

݀ܳ௪௔௟௟ 1
= ℎ௖ ‫(ܣ‬ఏ) (ܶ − ܶ௪௔௟௟ ) ൬ ൰
(5)
݀ߠ 6ܰ

PT
204

where the wall temperature, ܶ௪௔௟௟ is assumed as 420 K and the convection heat transfer

RI
205
206 coefficient, ℎ௖ is estimated according to Eq. (6) [33].

ℎ௖ = ‫ܥ‬ଵ × ܸ ି଴.଴଺ × ‫݌‬଴.଼ × ܶ ଴.ସ × (‫ܥ‬ଶ + ܸ௠ )଴.଼

SC
(6)
207 where ‫ ݌‬is the instantaneous pressure in bar, while ‫ܥ‬ଵ and ‫ܥ‬ଶ are constants (‫ܥ‬ଵ = 130 and
‫ܥ‬ଶ = 1.4).

U
208
209 2.5 Biodiesel fuel preparation
AN
210 Jojoba methyl ester (JME) was produced from commercially available jojoba oil through
211 transesterification process. The choice of jojoba oil was due to its availability and potential as
M

212 a source of biodiesel from non-edible oil. The properties of the raw jojoba oil [34] were
213 provided by the supplier. The methyl ester produced from jojoba oil (wax ester) was
D

214 characterized by Fourier transform infrared (FT-IR) and gas chromatography mass
TE

215 spectrometry (GC-MS) analysis. A transesterification procedure suggested by [35] was


216 followed to produce the methyl ester.
EP

217 Kinematic viscosity ν, of the resulting biodiesel was determined at 40oC in accordance with
218 American Society of Testing and Materials standards (ASTM D 445). All measurements
219 were taken in triplicate and average values calculated and recorded. The kinematic viscosity
C

220 of the biodiesel obtained from the transesterification process is approximately half that of the
AC

221 raw oil but still higher than the recommended limit by ASTM standards, for engine
222 application. The presence of long-chain fatty alcohols and residual jojoba oil in the produced
223 biodiesel could be the major contributing factor to the high kinematic viscosity of the Jojoba
224 methyl ester (JME) produced [36]. The biodiesel was blended with mineral diesel to reduce
225 the kinematic viscosity further to meet the limits recommended by ASTM standards [37]. The
226 measured kinematic viscosities, densities, flash points and fire points of the blends (JME5,
227 JME10, JME20, JME50, and JME80), neat JME, as well as mineral diesel are shown in Table
228 4.

Page 10 of 37
ACCEPTED MANUSCRIPT
229 Table 4 shows the measured properties of the raw oil, the produced biodiesel and biodiesel-
230 diesel blends of varying proportions.

231 Table 4. Properties of diesel, raw oil, biodiesel and biodiesel-diesel blends
D100 Raw oil JME100 JME5 JME10 JME20 JME50 JME80
Kinematic 3.5899 24.4394 12.3928 4.2574 4.4914 5.2532 7.8238 11.0871
viscosity

PT
at 40 oC
(mm2/s)

RI
Density at 0.8423 0.878 0.869 0.844 0.845 0.848 0.856 0.864
21.5 oC
(g/cm3)

SC
Flash 74.00 295.0 - - - - - -
point ( oC)

U
Fire point 106.00 324.0 - - - - - -
( oC)
AN
232

233 3.0 Theoretical analysis


M

234 3.1 Physical and numerical models

235 ANSYS Forte [38] is used to perform CFD simulations in this work. The sub-models
D

236 implemented in the computations, including physical and chemistry sub-models are listed in
TE

237 Table 5.
238 Table 5. Physical and chemistry sub-models used in the computations
Sub-models
EP

Turbulence model RNG k-ε model [39]


Breakup model KH-RT model [40]
Collision model Radius of influence (ROI) model [41]
C

Vaporization model Discrete multi-component (DMC) model [42]


Fuel chemistry Skeletal n-heptane mechanism [43], reduced
AC

methyl decanoate mechanism [44]


NOx mechanism Extended Zeldovich mechanism [45]
239
240 Renormalization group (RNG) k-ε turbulence model, Kelvin-Helmholtz/ Rayleigh-Taylor
241 (KH-RT) breakup model and Radius of Influence (ROI) collision model were used. The KH-
242 RT hybrid breakup model controls the spray atomization and droplet breakup of solid-cone
243 sprays. It combines the Kelvin-Helmholtz (KH) model which predicts the jet’s primary
244 breakup region (breakup length), and the Rayleigh-Taylor (RT) model which predicts
245 secondary breakup of spray droplets beyond the breakup length from the nozzle exit [40]. To

Page 11 of 37
ACCEPTED MANUSCRIPT
246 represent the vaporization of spray droplets, the Discrete Multi-component (DMC) fuel-
247 vaporization model [42] was used. The DMC vaporization model tracks the individual fuel
248 droplets during the evaporation process and allows coupling with the reaction kinetics of the
249 individual fuel components [42]. The model considers a spherical liquid fuel droplet
250 consisting of a finite number of components vaporizing without chemical reactions in a liquid
251 environment.

PT
252
253 In-cylinder turbulence was modelled using RNG k-ε turbulence model. In-cylinder flow

RI
254 processes are compressible and highly turbulent due to compression of the charge, swirl and
255 combustion processes. The RNG k-ε turbulence model was selected due to its high accuracy

SC
256 in predicting swirling flows [39,46]. The collision of fuel particles was predicted using the
257 ROI collision model [41]. In this model, one particle collides with another only if that other
258 particle resides within its radius of influence. The model is preferred as it eliminates mesh-
259

U
size dependency and time-step dependency for the droplet collision process [41]. A skeletal n-
AN
260 heptane mechanism consisting of 68 species and 283 reactions [43] is used to represent
261 mineral diesel combustion, while a reduced methyl decanoate (MD) mechanism [44]
M

262 consisting of 123 species and 394 reactions is used to model biodiesel combustion. Sazhin et
263 al. [47] demonstrated through a detailed study on modelling of biodiesel fuel droplet heating
D

264 and evaporation that multi-component as well as single component surrogates such as MD
265 can sufficiently be used to represent biodiesel combustion. NOx formation in CI combustion
TE

266 process involves mainly thermal and prompt NOx pathways. Thermal NOx is formed
267 downstream of the flame front and it highly depends on the gas temperature. Thermal NOx
EP

268 formation is commonly modelled using the extended Zeldovich mechanism. In this study, the
269 extended Zeldovich mechanism is integrated into the fuel combustion chemistries for
C

270 prediction of in-cylinder NOx evolution.


271
AC

272 The simulation study was conducted for a single cylinder four-stroke CI engine operated at a
273 constant speed and torque of 2000 rpm and 13.5 N.m, respectively. Due to the symmetry
274 imposed by the four equally-spaced nozzle holes of the fuel injector, a 90o sector mesh is
275 used to represent the full 360o engine geometry. Mesh independence study is conducted to
276 ensure that the generated mesh gives grid-independent results. The simulation covers only the
277 portion of engine cycle from intake valve closure (IVC) at 120o bTDC to exhaust valve
278 opening (EVO) at 125o aTDC. Since only one injection event is considered, engine misfire is
279 detected by observing the heat release rate and in-cylinder pressure profiles. In-cylinder

Page 12 of 37
ACCEPTED MANUSCRIPT
280 pressure profile which coincides with the motoring curve as well as lack of heat release in an
281 engine cycle with fuel injection indicates misfire.
282 The intake pressure and temperature are maintained at 1.013 bar and 360 K, respectively. The
283 wall temperature is fixed as 420 K while the piston surface temperature is taken as 500 K.
284 Five different conditions of CR are considered i.e 18, 20, 21.5, 22 and 23. The CR is varied to
285 study its effect on combustion and emission characteristics of a CI engine fueled with neat

PT
286 JME. Fuel temperature at injection is set as 368 K and the injector nozzle diameter is 0.15
287 mm. The spray axis is set at 75o from the vertical axis and a spray cone angle of 15o is used.

RI
288 The mass of fuel injected is 15 mg/cycle, calculated from the experimental fuel consumption
289 measurements. All simulation cases are numerically solved using parallel processing of Dell

SC
290 Precision T7500 workstation with Intel Xeon® processor of 3.75GH, 12-Processors, and 64-
291 GB installed memory. Figure 2 shows the computational grid at the start of compression
292 stroke.

U
AN
M
D
TE
C EP
AC

293

Page 13 of 37
ACCEPTED MANUSCRIPT
294 Fig. 2. Computational grid
295 The combustion process is treated as a turbulent reacting flow where the basic fluid dynamics
296 are governed by the Navier-Stokes equations. The governing equations for the flow field
297 include species conservation, fluid continuity, momentum and energy conservation equations,
298 as well as the gas-phase mixture equation of state.

299 Spatial differencing of the governing equations was conducted by implementing the Arbitrary

PT
300 Lagrangian-Euler method [48] on hexahedron mesh. The technique allows mesh vertices to
301 move in an arbitrarily prescribed manner, enabling piston motion. The algebraic finite-

RI
302 volume equations resulting from discretization of the governing equations were solved by
303 implicit methods. The semi-implicit method for pressure linked equations (SIMPLE method)

SC
304 [49] was applied for the coupling of pressure and velocities to solve the flow field variables
305 in a two-step iterative procedure. The method extrapolates the pressure, iteratively solves for

U
306 velocities, then temperature, and finally the pressure [49]. The quasi-second-order upwind
307 (QSOU) method was used to solve the convection terms in the momentum and energy
AN
308 conservation equations.

3.2 Mesh independence study


M

309
310 Eight different mesh resolutions are evaluated to determine the effect of element size on the
311 accuracy of the obtained results. Data from experimental firing cases of the test engine are
D

312 used as the benchmark for comparison with the numerical results. A mesh with 12,180 cells is
TE

313 found to give satisfactory results with respect to accuracy and computational runtime. Finer
314 mesh resolutions result in increased computational time without significant improvement in
EP

315 the results. Details of the mesh independence study are summarized in Table 6.
316
317 Table 6. Mesh independence study cases: using n-heptane mechanism (68 species, 283
C

318 reactions) [43]


Case Number Element Runtime Peak Peak Peak HRR Peak HRR
AC

of cells size Pressure Pressure (J/ degree): (J/ degree):


(mm) (bar): (bar): Numerical Experiment
Numerical Expt.
1 1,170 13 26 min, 12.3 80.81 72.14 69.08 31.25
sec
2 2,192 8 32 min, 1.5 76.39 " 46.74 "
sec
3 4,120 6 41 min, 18.6 73.34 " 39.95 "
sec
4 12,180 1 1 hr, 24 min, 71.73 " 31.30 "
0.6 sec
5 32,460 0.9 2 hrs, 37 min, 71.13 " 31.41 "

Page 14 of 37
ACCEPTED MANUSCRIPT
28.4 sec
6 81,120 0.75 6 hrs, 54 min, 70.05 " 26.70 "
18.7 sec
7 228,640 0.50 1 d, 5 hrs, 10 71.56 " 38.60 "
min, 51.6 sec
8 296,320 0.25 1 d, 7 hrs, 16 71.31 " 36.15 "
min, 25.4 sec.
319

PT
320 Figures 3 and 4 show the effect of mesh resolution on in-cylinder pressure and heat release
321 rates, respectively, for the investigated mesh resolutions. Four cases only are presented for
322 comparison. The selected cases sufficiently show the effect of mesh resolution on

RI
323 computational results. It is observed that the accuracy of the output parameters improve with
324 reduction in element size. Coarse mesh resolution results in overprediction of both the peak

SC
325 in-cylinder pressure and HRR by 12 % and 120 % times, respectively. It is also observed that
326 the mesh of 12,180 cells yielded satisfactory results at a lower runtime compared to finer

U
327 mesh resolutions, hence it was selected for the present numerical study. Finer mesh resolution
AN
328 would lead to longer computational time without any significant improvement in the accuracy
329 of the results.
M

100
Cylinder Pressure (bar)

1170 Cells
2192 Cells
D

80 12180 Cells
228640 Cells
Measured
TE

60

40
EP

20
-20 -10 0 10 20
C

330
Crank Angle (degree ATDC)
AC

331 Fig. 3. Effect of mesh resolution on cylinder pressure.

Page 15 of 37
ACCEPTED MANUSCRIPT
80
1170 Cells
2192 Cells

HRR (J/ degree)


60 12180 Cells
228 640 Cells
Measured
40

20

PT
0
-10 -5 0 5 10

RI
332
Crank Angle (degree ATDC)

333 Fig. 4. Effect of mesh resolution on heat release rate.

SC
334

335 3.3 Model validation

336
U
The motoring and firing in-cylinder pressure profiles are compared against the measured
AN
337 results to determine the accuracy of the numerical model in predicting the performance of the
338 test engine. The peak in-cylinder pressure magnitude and location, and the general shape of
M

339 pressure profiles are used as the validation criteria. Figures 5 and 6 show comparisons of the
340 numerical and experimental pressure profiles for the motoring and firing (diesel) cases,
D

341 respectively, while Fig. 7 shows a comparison for JME20 firing case.
TE

60
Cylinder Pressure (bar)

Measured
Numerical
EP

40
C

20
AC

0
-80 -40 0 40 80
342
Crank Angle (degree ATDC)

343 Fig. 5. Comparison between the numerical and measured cylinder pressure for motoring case.

Page 16 of 37
ACCEPTED MANUSCRIPT
80 100

Cylinder Pressure (bar)


Measured
Numerical
80

HRR (J/degree)
60 Measured
Numerical
60
40
40
20 20

PT
0 0
-80 -40 0 40 80

RI
344
Crank Angle (degree ATDC)

345 Fig. 6. Comparison between the numerical and measured cylinder pressure for diesel.

SC
80
Cylinder Pressure (bar)

Measured (JME20)
Numerical (JME20)

U
60
AN
40
M

20

0
D

-80 -40 0 40 80
Crank Angle (degree ATDC)
TE

346
347 Fig. 7. Comparison between the numerical and measured cylinder pressure for biodiesel-
348 diesel blend (JME20).
EP

349 Comparison of the in-cylinder pressure profiles and HRR curves show satisfactory agreement
350 between the numerical results and experimental data. This approves the validity of the
C

351 numerical setup to be used for different CR and different diesel/biodiesel blends.
AC

352 4.0 Results and Discussion

353 This section is divided into three main subsections. The first subsection presents measured
354 results on characterization of the biodiesel fuel used in this study while the second part
355 presents the experimental results of the effect of jojoba methyl ester on combustion and
356 emission characteristics of the CI engine. The final subsection discusses the numerical results
357 of the effect of CR on CI engine combustion and emission characteristics.

358 4.1 Biodiesel fuel analysis

Page 17 of 37
ACCEPTED MANUSCRIPT
359 Properties of the jojoba methyl ester produced from transesterification were investigated
360 through FT-IR and GC-MS analysis. Fuel properties such as kinematic viscosity and density
361 were also measured and were presented in Table 4. Figures 8(a) and (b) show the FT-IR
362 spectra of raw JO and the produced biodiesel, respectively.
100
90

PT
80
Transmittance [%]

RI
60 70

SC
50 40

U
30

AN
3458.58

2925.47
2857.44

2679.61

2318.35

2034.28

1738.95
1654.34

1456.74
1365.71

1243.35
1173.75

864.83

720.18

520.05
593.11
M

3500 3000 2500 2000 1500 1000 500


Wavenumber cm-1
363
D

364 (a)
100

TE
90
80

EP
Transmittance [%]
60 70

C
50

AC
40 30
20

3403.69

2924.65
2856.07

2680.40

2035.20

1740.00
1653.96
1568.20

1456.87
1366.54

1247.77

1057.51
1174.15

867.19

721.38

592.78
519.01

3500 3000 2500 2000 1500 1000 500


Wavenumber cm-1
365
366 (b)

Page 18 of 37
ACCEPTED MANUSCRIPT
367 Fig.8. FT-IR spectrum of (a) raw jojoba oil and (b) the biodiesel produced.
368 Though they look quite similar, a few noticeable differences exist in the FT-IR spectra of the
369 oil and the biodiesel, indicating the conversion of the oil to biodiesel.

370 The retention time and fragmentation pattern data of the GC-MS analysis revealed the main
371 components of the produced biodiesel as Octadecadienoic acid methyl ester, cis-13-
372 Eicosenoic acid methyl ester, Cyclopropaneoctanoic acid, 8-Octadecenoic acid, 9-Octadecen-

PT
373 1-ol and 13-Docosen-1-ol. The identified fatty acid methyl esters were verified with available
374 data from previous studies [50].

RI
375 4.2 Effect of biodiesel on combustion characteristics: Experimental investigation

SC
376 In-cylinder pressure variation with crank angle was recorded for each test. Figure 9 shows the
377 effect of jojoba biodiesel on measured in-cylinder pressure for D100, JME5, JME10 and

U
378 JME20. Peak cylinder pressure is observed to decrease with increase in biodiesel quantity in
379 the fuel blend owing to the lower calorific value and higher viscosity of biodiesel. The effect
AN
380 of JME substitution on HRR is shown in Fig. 10.

381 The lower in-cylinder peak pressure and HHR of JME blends could be attributed to the lower
M

382 calorific value of JME compared to diesel, leading to lower power output [51]. The higher
383 viscosity and density also affect fuel vaporization and spray pattern, and consequently the
D

384 peak pressure.


TE

80
Cylinder Pressure (bar)
EP

60

40
C

D100 (Measured)
JME5 (Measured)
AC

20 JME10 (Measured)
JME20 (Measured)

0
-20 -10 0 10 20 30 40
385
Crank Angle (degree ATDC)

386 Fig. 9. Effect of jojoba biodiesel on cylinder pressure at 2000 rpm and 13.5 N.m load.

Page 19 of 37
ACCEPTED MANUSCRIPT
40
D100 (Measured)
JME5 (Measured)

HRR (J/degree)
30 JME10 (Measured)
JME20 (Measured)

20

10

PT
0
-60 -40 -20 0 20 40

RI
387
Crank Angle (degree ATDC)

388 Fig.10. Effect of jojoba biodiesel on heat release rate at 2000 rpm and 13.5 N.m load.

SC
389 4.3 Effect of biodiesel on measured NOx and CO emission: Experimental investigation

Figures 11(a) and (b) show the effect of quantity of biodiesel in fuel blend on measured

U
390
391 emissions of CO and NOx, respectively. The CO emission is higher at low engine load but
AN
392 decreases with increase in engine load for all fuel conditions. It is also observed that higher
393 quantities of biodiesel in the fuel blend leads to increase in CO emission. The increase in CO
M

394 emission could be attributed to higher density and viscosity of biodiesel which impedes the
395 fuel atomization process. NOx emission generally increases with increase in engine load, for
D

396 all the fuel types. Low substitution level of JME5 shows the highest NOx levels under all
397 engine loads. The NOx levels then decrease significantly with higher quantities of biodiesel in
TE

398 the blend. This could be attributed to the effect of biodiesel on spray atomization and
399 consequently on combustion and incylinder pressure as seen from Fig. 9. The highest peak
EP

400 incylinder pressure was found to be for JME5 while any further increase in JME percentage
401 leads to reduction in incylinder pressure.
C
AC

Page 20 of 37
ACCEPTED MANUSCRIPT
120
(a) D100

CO Emission (ppm)
JME5
JME10
80 JME20

40

PT
0
0 3 6 9 12 13.5

RI
402
Engine load (N.m)

200

SC
(b)
NOx Emission (ppm)

D100
150 JME5

U
JME10
JME20
100
AN
50
M

0
0 3 6 9 12 13.5
D

403
Engine load (N. m)
TE

404 Fig. 11. Effect of biodiesel blend on measured (a) CO emission and (b) NOx emission of the
405 CI engine.
406 4.4 Effect of compression ratio on combustion characteristics: Numerical investigation
EP

407 Figures 12(a) and (b) show the effect of CR on in-cylinder pressure of the CI engine fueled
408 with mineral diesel and neat jojoba methyl ester (JME100), respectively. It could be noticed
C

409 for original CR of engine, 21.5, that the peak pressure is reduced by 10% when using
AC

410 JME100 compared to neat diesel and the ignition is delayed by approximately 5 CA. It could
411 be also observed that this reduction in peak pressure and increase in ignition delay could be
412 recovered by increasing CR from 21.5 to 23 (approximately 7%). Hence, it could be noted
413 that the neat diesel fuel could be replaced by JME100 fuel without loss in power by
414 increasing CR by 7%. This could be attributed to higher cetane number of JME [53] which
415 leads to low accumulation of fuel during the ignition delay period, and subsequently to lower
416 in-cylinder peak pressure. Lowering the CR of the engine from the baseline value of 21.5 to
417 18 leads to a reduction in peak in-cylinder pressure from 72.68 to 69.12 bar (4.9 %) for

Page 21 of 37
ACCEPTED MANUSCRIPT
418 diesel, and from 66.45 to 52.24 bar (21.38 %) for biodiesel. On the other hand, raising the CR
419 to 23 leads to 2.4 % and 14.5 % increase in peak pressure for diesel and JME, respectively
420 The results indicate that in-cylinder peak pressure improves with the increase in CR for both
421 fuels and the impact of CR is higher in biodiesel case than that in diesel. The peak pressure
422 shifts away from top dead centre, indicating increase in ignition delay. Lower CR leads to
423 lower gas temperature, hence longer ignition delay. Hence, the peak pressure shifts away

PT
424 from top dead centre, indicating increase in ignition delay. This is in agreement with the
425 results of Selvan et al. [54] for a CI engine fueled with jatropha biodiesel blends with diesel.

RI
80
Cylinder Pressure (bar)

(a)

SC
60 Fuel: D100

40

U
CR 23
CR 22
AN
20 CR 21.5 (Baseline)
CR 20
CR 18
0
M

-20 -10 0 10 20 30
426
Crank Angle (degree ATDC)
D

80
Cylinder Pressure (bar)

(b)
TE

60 Fuel: JME100
EP

40
CR 23
CR 22
20 CR 21.5 (Baseline)
C

CR 20
CR 18
0
AC

-20 -10 0 10 20 30
427
Crank Angle (degree ATDC)

428 Fig. 12. Effect of CR on in-cylinder pressure of (a) diesel fueled CI engine and (b) biodiesel
429 (JME100) fueled CI engine.
430 The effect of CR on heat release rates (HRR) for diesel and JME combustion is shown in
431 Figs. 13(a) and (b), respectively. A slight delay in peak HRR and start of combustion with
432 decrease in CR was observed for both fueling conditions. Increasing the CR from 21.5 to 23
433 results in 1.8% (30.34 to 30.89) and 14.5% (40.85 to 45.95) increase in peak HHR for diesel

Page 22 of 37
ACCEPTED MANUSCRIPT
434 and JME, respectively. An increase in peak HRR, from 30.34 to 34.07 J/ oCA (12.3%) with
435 decrease in CR from 21.5 to 18 was also observed in the case of diesel operation. This could
436 be attributed to accumulation of more fuel in the combustion chamber due to longer ignition
437 delay caused by lower in-cylinder gas temperature at start of injection. A similar trend of
438 increasing HRR with decrease in CR (at high load) was observed by Hariram et al. [22] in an
439 experimental investigation on the influence of CR on combustion and performance

PT
440 characteristics of a direct injection CI engine. In the case of engine operation with neat
441 biodiesel in the present study, higher HRR compared to diesel was observed at corresponding

RI
442 compression ratios, especially at higher CR settings. This indicates faster combustion rate,
443 possibly due to the oxygen atom present in biodiesel molecules. However, unlike in engine

SC
444 operation with diesel, biodiesel reveals a decrease in peak HRR with reduction in CR.
445 Reducing the CR from 21.5 to 18 led to a decrease in the peak HRR by 29.8%, from 40.85 to
446 28.66 J/ oCA. This also could be as a result of poor atmization in neat biodiesel case which
447 results from high viscosity
U
AN
50
CR 23 (a)
CR 22
40 Fuel: D100
M
HRR (J/ degree)

CR 21.5 (Baseline)
CR 20
30 CR 18
D

20
TE

10

0
EP

-20 -10 0 10 20
448
Crank Angle (degree ATDC)
C

CR 23 (b)
AC

60 CR 22 Fuel: JME100
HRR (J/ degree)

CR 21.5 (Baseline)
CR 20
40 CR 18

20

0
-40 -20 0 20 40
449
Crank Angle (degree ATDC)

Page 23 of 37
ACCEPTED MANUSCRIPT
450 Fig. 13. Effect of CR on HRR of (a) diesel fueled CI engine and (b) biodiesel fueled CI
451 engine.
452

453 4.5 Effect of compression ratio on emission characteristics

454 Figures 14(a) and (b) show the effect of compression ratio on NO emission formation during
455 the combustion process, for engine operation with diesel and biodiesel, with reference to

PT
456 engine crank angle. Figure 14(c) on the other hand presents a comparison of the NO emission
457 for diesel and biodiesel operation at different CR. The results show that formation of NO

RI
458 emission is more sensitive to CR in the case of engine operation with JME than with standard
459 diesel. By reducing the compression ratio from 21.5 to 18, NO emission reduced by

SC
460 approximately 1.79 % , from 0.4473 g/kg-fuel to 0.4393 g/kg-fuel for diesel and by 40.7 % ,
461 from 0.3144 g/kg-fuel to 0.1865 g/kg-fuel for JME. This shows a significant reduction in

U
462 emission of NO when standard diesel is fully replaced with neat JME. However, the in-
463 cylinder peak pressure reduces remarkably, as was discussed in Fig. 12(b). On the other hand,
AN
464 increasing the CR to 23 leads to an increase in NO emission by 2.2 % and 4.6 % for diesel
465 and JME, respectively.
M

0.8
NO Emission (g/kg-fuel)

(a)
D

Fuel: D100
0.6
TE

0.4
CR 23
CR 22
0.2
EP

CR 21.5 (Baseline)
CR 20
CR 18
0.0
C

-40 0 40 80 120
Crank Angle (degree ATDC)
AC

466

Page 24 of 37
ACCEPTED MANUSCRIPT
0.8

NO Emission (g/kg-fuel)
CR 23 (b)
CR 22 Fuel: JME100
0.6 CR 21.5 (Baseline)
CR 20
CR 18
0.4

0.2

PT
0.0
-20 0 20 40 60 80 100 120

RI
467
Crank Angle (degree ATDC)

0.8

SC
NO Emission (g/kg-fuel)

D100 (c)
JME100
0.6

0.4
U
AN
0.2
M

0.0
17 18 19 20 21 22 23 24
D

468
Compression Ratio (CR)
TE

469 Fig. 14. Effect of CR on NO emission of (a) diesel fueled CI engine, (b) biodiesel fueled CI
470 engine and (c) diesel and biodiesel NO comparison.
471 Many researchers [51,55–57] have reported higher NOx production for biodiesel fueled CI
EP

472 engines, while others [58–60] have reported lower NOx emissions. Generally, NOx formation
473 is influenced by equivalence ratio, oxygen concentration, combustion temperature and
C

474 reaction time [13]. In the present study, NOx emission is lower for biodiesel combustion.
AC

475 Further investigation is therefore necessary to ascertain the effect of biodiesel on NOx
476 emission.

477 The effect of compression ratio on CO emission is shown in Fig. 15. CO emission highly
478 depends on the combustion temperature and availability of oxygen. A reduction in CR from
479 21.5 to 18 leads to a decrease in CO emission by 14.29 % and 14.75 % for diesel and neat
480 JME, respectively, while an increase in CR to 23 increases the CO emission by 1.4% and
481 27.2% for diesel and JME, respectively. Lower CO emission was recorded under engine
482 operation with JME as compared to diesel, at all CR settings. Biodiesel recorded lower CO

Page 25 of 37
ACCEPTED MANUSCRIPT
483 emission at all compression ratios owing to the additional oxygen present in the biodiesel
484 chemical structure. Similar results have been reported from previous experimental studies
485 using biodiesel from different sources [18].

800

CO Emission (g/kg-fuel)
(a)
600 Fuel: D100

PT
400
CR 23
200 CR 22

RI
CR 21.5 (Baseline)
CR 20
0 CR 18

SC
-40 -20 0 20 40 60 80 100 120
486
Crank Angle (degree ATDC)

1000
U
AN
CO Emission (g/kg-fuel)

(b) Fuel: JME100


800 CR 23
CR 22
600 CR 21.5 (Baseline)
M

CR 20
CR 18
400
D

200
TE

0
-20 0 20 40 60 80 100 120
Crank Angle (degree ATDC)
EP

487
600
CO Emission (g/kg-fuel)

(c)
C

D100
500 JME100
AC

400

300

200

100
17 18 19 20 21 22 23 24
488
Compression Ratio (CR)

Page 26 of 37
ACCEPTED MANUSCRIPT
489 Fig. 15. Effect of CR on CO emission of (a) diesel fueled CI engine, (b) biodiesel fueled CI
490 engine and (c) diesel and biodiesel comparison.
491 The effect of CR on UHC emission is shown in Fig. 16. Compression ratio variation is
492 observed to have marginal effect on UHC emission. Lower UHC emission levels are also
493 observed under engine operation with biodiesel, owing to higher combustion temperatures
494 and availability of more oxygen from biodiesel molecular structure. The results are in good

PT
495 agreement with the findings of Muralidharan et al. [18].
UHC Emission (g/kg-fuel)

800 (a)

RI
Fuel: D100
600 CR 23

SC
CR 22
CR 21.5 (Baseline)
400 CR 20
CR 18

U
200 AN
0
-40 -20 0 20 40 60 80 100 120
496
Crank Angle (degree ATDC)
M

400
UHC Emission (g/kg-fuel)

(b) Fuel: JME100


D

300 CR 23
TE

CR 22
CR 21.5 (Baseline)
200 CR 20
CR 18
EP

100

0
C

-40 -20 0 20 40 60 80 100 120


AC

497
Crank Angle (degree ATDC)

Page 27 of 37
ACCEPTED MANUSCRIPT
200

UHC Emission (g/kg-fuel)


(c)
D100
160 JME100

120

80

40

PT
0
17 18 19 20 21 22 23 24

RI
498
Compression Ratio (CR)

499 Fig. 16. Effect of CR on UHC emission of (a) diesel fueled CI engine, (b) biodiesel fueled CI

SC
500 engine and (c) diesel and biodiesel comparison.
501 Table 7 summarizes the effect of CR on combustion and emission characteristics of the CI

U
502 engine fueled with D100 and JME100. Comparison of the combustion and emission
characteristics of the CI engine fueled with standard diesel and JME shows that reduction of
AN
503
504 CR leads to decrease in in-cylinder peak pressure, pressure rise rate, heat release rate as well
505 as CO, UHC and NOx emissions, irrespective of the fuel type. While reduction in all the
M

506 emissions with lower CR is beneficial, reduction in the in-cylinder peak pressure is
507 undesirable as it leads to lower power output of the engine.
D

508 Table 7. Effect of CR on combustion and emission characteristics of CI engine fuelled with
TE

509 mineral diesel (D100) and pure jojoba methyl ester (JME)
Compression ratio (CR)
EP

23 22 21.5 20 18
D JME D JME D JME D JME D JM
100 100 100 100 100 100 100 100 100 E
C

100
Peak
AC

Cylinder 74.41 76.0 73.50 71.6 72.68 66.45 71.72 56.91 69.1 52.2
Pressure 87 856 34 195 193 4
(bar)
Peak 6.028 7.81 5.590 6.47
Pressure 2 87 5 92 5.49 5.53 5.31 3.34 5.20 2.25
Rise Rate
(bar/oCA)
Peak 30.89 45.9 30.57 44.3 30.34 40.85 31.99 34.88 34.0 28.6
HRR 00 516 71 677 7 6
(J/degree
)
CO 323.3 269. 328.6 231. 327.8 211.8 313.3 182.9 280. 180.

Page 28 of 37
ACCEPTED MANUSCRIPT
[g/kg- 693 3017 133 4928 1 0 7 5 99 57
fuel]
HC 106.7 30.6 110.3 26.5 108.7 25.21 105.5 19.58 94.7 18.0
[g/kg- 853 343 129 457 4 0 8 9
fuel]
NOx 0.457 0.32 0.452 0.32 0.447 0.314 0.445 0.244 0.43 0.18
[g/kg- 1 88 5 54 3 4 7 6 93 65
fuel]
510

PT
511
Change in peak pressure (%)

RI
20
(a)
D100
10 JME

SC
0

-10

U
AN
-20
Peak Pressure: compared to CR 21.5
-30
17 18 19 20 21 22 23
M

512
Compression ratio (CR)

20
Change in Peak HRR (%)

(b) Peak HRR: compared to CR 21.5


10
TE

0
-10
EP

-20
D100
-30 JME100
C

-40
AC

17 18 19 20 21 22 23
513
Compression ratio (CR)

514

Page 29 of 37
ACCEPTED MANUSCRIPT

Increase in NOx Emission (%)


20
(c) NOx: compared to CR 21.5
10
0
-10
-20
-30 D100

PT
JME100
-40
-50

RI
17 18 19 20 21 22 23
515
Compression ratio (CR)

SC
516

517

U
518
Reduction in CO Emission (%)

AN
30
(d) CO: compared to CR 21.5
20
D100
M

JME100
10
D

0
TE

-10

-20
17 18 19 20 21 22 23
EP

519
Compression Ratio (CR)
Reduction in UHC Emission (%)
C

30
(e) UHC: compared to CR 21.5
AC

20
10
0
-10
-20 D100
JME100
-30
-40
17 18 19 20 21 22 23
520
Compression ratio (CR)

Page 30 of 37
ACCEPTED MANUSCRIPT
521 Fig. 17. Percentage change in (a) in-cylinder peak pressure, (b) HRR, (c) NOx, (d) CO and
522 (e) UHC emission, as a result of CR change relative to the original setting of 21.5.

523 Increasing the compression ratio results in higher in-cylinder peak pressure, HRR. NO, CO
524 and UHC emissions. Conversely, lowering the CR leads to a decrease in the in-cylinder peak
525 pressure, HRR. NO, CO and UHC emissions. JME also shows greater sensitivity to change in
526 CR compared to standard diesel.

PT
527 5.0 Conclusion

RI
528 The effect of CR on combustion and emission characteristics of a CI engine fuelled with neat
529 jojoba methyl ester have been investigated and compared with mineral diesel. Numerical

SC
530 simulations were conducted on a 3-D CFD simulation model of a single cylinder four-stroke
531 CI engine, operated at a constant speed and load of 2000 rpm and 13.5 N.m, respectively. The
532 model is validated using experimental data obtained at a compression ratio of 21.5, followed
533

U
by application of the model to investigate engine performance at CR of 18, 20, 22 and 23.
AN
534 The following conclusions are drawn from the investigations.

535 1. Increasing the compression ratio from 21.5 to 23 leads to increase in peak pressure by 2.4
M

536 % and 14.5 % for D100 and JME100, respectively, indicating greater improvement for
537 biodiesel combustion at higher CR. Conversely, decreasing the CR from 21.5 to 18 lowers
D

538 the peak pressure by 4.9 % and 21.9 % for D100 and JME, respectively. Therefore, based on
TE

539 the recorded peak pressure, higher CR should be considered for optimum performance of CI
540 engine with neat biodiesel. This would enhance biodiesel atomization, vaporization and
541 mixing with air, to improve combustion.
EP

542 2. Regarding emissions, increasing the compression ratio from 21.5 to 23 leads to increase in
543 CO by 1.4 % and 27.2 % for D100 and JME100, respectively, and increase in UHC by 1.8 %
C

544 and 21.5 % for D100 and JME, respectively. In addition, NOx emission increases by 2.2 %
AC

545 and 4.6 % for D100 and JME100, respectively, for the same change in CR. Higher
546 compression ratio leads to higher temperature at start of combustion and lower temperature
547 after expansion, which could hinder complete oxidation of CO and UHC, leading to increase
548 in CO and UHC emissions. However, comprehensive experimental investigations on a
549 variable compression ratio (VCR) engine are recommended in future studies to establish the
550 findings on combustion and emission characteristics of CI engine fuelled with neat JME.
551 Application of acoustic signatures to ascertain engine stability is also recommended.

Page 31 of 37
ACCEPTED MANUSCRIPT
552 Acknowledgement
553 The authors gratefully acknowledge the financial support of Japan International Corporation
554 Agency (JICA) in acquiring the simulation software used in this study. This work was also
555 supported by Egypt-Japan University of Science and Technology (E-JUST) through provision
556 of the equipment and facilities used in the experimental and numerical study.

References

PT
557

558 [1] Wamankar AK, Satapathy AK, Murugan S. Experimental investigation of the effect of
559 compression ratio, injection timing & pressure in a DI (direct injection) diesel engine

RI
560 running on carbon black-water-diesel emulsion. Energy 2015;93:511–20.
561 doi:10.1016/j.energy.2015.09.068.

SC
562 [2] Özener O, Yüksek L, Ergenç AT, Özkan M. Effects of soybean biodiesel on a DI
563 diesel engine performance, emission and combustion characteristics. Fuel
564 2014;115:875–83. doi:10.1016/j.fuel.2012.10.081.

U
565 [3] Abu-Qudais M, Haddad O, Qudaisat M. The effect of alcohol fumigation on diesel
566 engine performance and emissions. Energy Convers Manag 2000;41:389–99.
AN
567 doi:http://dx.doi.org/10.1016/S0196-8904(99)00099-0.
568 [4] Xue J, Grift TE, Hansen AC. Effect of biodiesel on engine performances and
569 emissions. Renew Sustain Energy Rev 2011;15:1098–116.
M

570 doi:10.1016/j.rser.2010.11.016.
571 [5] Lin CY, Lin HA. Diesel engine performance and emission characteristics of biodiesel
572 produced by the peroxidation process. Fuel 2006;85:298–305.
D

573 doi:10.1016/j.fuel.2005.05.018.
TE

574 [6] Muralidharan K, Vasudevan D, Sheeba KN. Performance, emission and combustion
575 characteristics of biodiesel fuelled variable compression ratio engine. Energy
576 2011;36:5385–93. doi:10.1016/j.energy.2011.06.050.
EP

577 [7] Agarwal AK, Gupta JG, Dhar A. Potential and challenges for large-scale application of
578 biodiesel in automotive sector. Prog Energy Combust Sci 2017;61:113–49.
579 doi:10.1016/j.pecs.2017.03.002.
C

580 [8] Lapuerta M, Rodríguez-Fernández J, Agudelo JR. Diesel particulate emissions from
581 used cooking oil biodiesel. Bioresour Technol 2008;99:731–40.
AC

582 doi:10.1016/j.biortech.2007.01.033.
583 [9] Buyukkaya E. Effects of biodiesel on a DI diesel engine performance, emission and
584 combustion characteristics. Fuel 2010;89:3099–105. doi:10.1016/j.fuel.2010.05.034.
585 [10] Hamid MF, Idroas MY, Sa’ad S, Saiful Bahri AJ, Sharzali CM, Abdullah MK, et al.
586 Numerical investigation of in-cylinder air flow characteristic improvement for
587 Emulsified biofuel (EB) application. Renew Energy 2018;127:84–93.
588 doi:10.1016/j.renene.2018.04.006.
589 [11] Imtenan S, Masjuki HH, Varman M, Kalam MA, Arbab MI, Sajjad H, et al. Impact of
590 oxygenated additives to palm and jatropha biodiesel blends in the context of
591 performance and emissions characteristics of a light-duty diesel engine. Energy

Page 32 of 37
ACCEPTED MANUSCRIPT
592 Convers Manag 2014;83:149–58. doi:10.1016/j.enconman.2014.03.052.
593 [12] Bora BJ, Saha UK, Chatterjee S, Veer V. Effect of compression ratio on performance,
594 combustion and emission characteristics of a dual fuel diesel engine run on raw biogas.
595 Energy Convers Manag 2014;87:1000–9. doi:10.1016/j.enconman.2014.07.080.
596 [13] Heywood JB. Internal Combustion Engine Fundamentals. 1st ed. New York: McGraw-
597 Hill, Inc.; 1988.
598 [14] Laguitton O, Crua C, Cowell T, M RH, Gold M. The effect of compression ratio on

PT
599 exhaust emissions from a PCCI diesel engine. Energy Convers Manag Manag
600 2007;48:2918–24.
601 [15] Raheman H, Ghadge S. Performance of diesel engine with biodiesel at varying

RI
602 compression ratio and ignition timing. Fuel 2008;87:2659–66.
603 [16] Jindal S, Nandwanaa BP, Rathore NS, Vashistha V. Experimental investigation of the

SC
604 effect of compression ratio and injection pressure in a direct injection diesel engine
605 running on Jatropha methyl ester. Appl Therm Eng 2010;30:442–8.
606 [17] Sayin C, Gumus M. Impact of compression ratio and injection parameters on the

U
607 performance and emissions of a DI diesel engine fueled with biodiesel blended diesel
608 fuel. Appl Therm Eng 2011;31:3182–8.
AN
609 [18] Muralidharan K, Vasudevan D. Performance, emission and combustion characteristics
610 of a variable compression ratio engine using methyl esters of waste cooking oil and
611 diesel blends. Appl Energy 2011;88:3959–68.
M

612 [19] El_Kassaby M, Nemit_allah MA. Studying the effect of compression ratio on an
613 engine fueled with waste oil produced biodiesel/diesel fuel. Alexandria Eng J
614 2013;52:1–11. doi:10.1016/j.aej.2012.11.007.
D

615 [20] Wan Ghazali WNM, Mamat R, Masjuki HH, Najafi G. Effects of biodiesel from
TE

616 different feedstocks on engine performance and emissions: A review. Renew Sustain
617 Energy Rev 2015;51:585–602. doi:10.1016/j.rser.2015.06.031.
618 [21] Gnanamoorthi V, Devaradjane G. Effect of compression ratio on the performance,
EP

619 combustion and emission of di diesel engine fueled with ethanol e Diesel blend. J
620 Energy Inst 2015;88:19–26. doi:10.1016/j.joei.2014.06.001.
621 [22] Hariram V, Vagesh Shangar R. Influence of compression ratio on combustion and
C

622 performance characteristics of direct injection compression ignition engine. Alexandria


623 Eng J 2015;54:807–14. doi:10.1016/j.aej.2015.06.007.
AC

624 [23] Hirkude J, Padalkar AS. Experimental investigation of the effect of compression ratio
625 on performance and emissions of CI engine operated with waste fried oil methyl ester
626 blend. Fuel Process Technol 2014;128:367–75. doi:10.1016/j.fuproc.2014.07.026.
627 [24] BP Statistical Review of World Energy 2017. Br Pet 2017:1–52.
628 [25] Graham-rowe D. Beyond food versus fuel. Nature 2011;474:3–5.
629 [26] Demirbas A. Biofuels Sources, Biofuel Policy, Biofuel Economy and Global Biofuel
630 Projections. Energy Convers Manag 2008;49:2106–16.
631 [27] Ahmed W, Nazar MF, Ali SD, Rana UA, Khan SUD. Detailed investigation of
632 optimized alkali catalyzed transesterification of Jatropha oil for biodiesel production. J

Page 33 of 37
ACCEPTED MANUSCRIPT
633 Energy Chem 2015;24:331–6. doi:10.1016/S2095-4956(15)60319-9.
634 [28] Canoira L, Alca’ntara R, Garcı’a-Martı’nez J, Carrasco J. Biodiesel from Jojoba oil-
635 wax : Transesterification with methanol and properties as a fuel. Biomass and
636 Bioenergy 2006;30:76–81. doi:10.1016/j.biombioe.2005.07.002.
637 [29] El-seesy AI, Abdel-rahman AK, Hassan H, Ookawara S, Hawi M. Biodiesel-diesel
638 fuel mixture with addition of nanoparticles. Proc. ASME 2017 Power Conf. Jt. With
639 ICOPE-17, Charlotte: ASME; 2017, p. 1–9.

PT
640 [30] Robinson PRB and DK. Data Reduction and Error Analysis for the Physical Sciences.
641 third edit. New York: Mc Graw Hill; 2003.
642 [31] H.W C, J.W.G S. Experimentation and uncertainty analysis for engineers. second edi.

RI
643 New York: John Willey & Sons; 1989.
644 [32] Gumus M. A comprehensive experimental investigation of combustion and heat

SC
645 release characteristics of a biodiesel (hazelnut kernel oil methyl ester) fueled direct
646 injection compression ignition engine. Fuel 2010;89:2802–14.
647 doi:10.1016/j.fuel.2010.01.035.

U
648 [33] G. F. Hohenberg. Advance approaches for heat transfer calculation. Soc Automot Eng
649 SAE Tech Pap NO 790825 19792788–98 1979.
AN
650 [34] N. EM. Jojoba: the green gold hope for the Egyptian desert development - United
651 Nations: Economic and Social commission for Western Asia,Report of the Experts
652 Group Meeting Manama (Bahrain) 2002. 2002.
M

653 [35] Alptekin E, Canakci M. Determination of the density and the viscosities of biodiesel-
654 diesel fuel blends. Renew Energy 2008;33:2623–30.
655 doi:10.1016/j.renene.2008.02.020.
D

656 [36] Shah SN, Sharma BK, Moser BR, Erhan SZ. Preparation and evaluation of Jojoba oil
TE

657 methyl esters as biodiesel and as a blend component in ultra-low sulfur diesel fuel.
658 Bioenergy Res 2010;3:214–23. doi:10.1007/s12155-009-9053-y.
659 [37] Gumus M, Sayin C, Canakci M. Effect of fuel injection timing on the injection,
EP

660 combustion, and performance characteristics of a direct-injection ( DI ) diesel engine


661 fueled with canola oil methyl ester-diesel fuel blends. Energy & Fuels 2010;24:3199–
662 213. doi:10.1021/ef9014247.
C

663 [38] ANSYS Forte 17.2, ANSYS: San Diego, 2016. n.d.
AC

664 [39] Han Z, Reitz RD. Turbulence Modeling of Internal Combustion Engines Using RNG κ
665 - ε Models. Combust Sci Technol 1995;106:267–95.
666 [40] Beale JC, Reitz RD. Modeling spray atomization with the Kelvin- Helmholtz /
667 Rayleigh-Taylor hybrid model. At Sprays 1999;145.
668 [41] Munnannur A, Reitz RD. A Comprehensive collision model for multi-dimensional
669 engine spray computation. At Sprays 2009;19:597–619.
670 doi:10.1615/AtomizSpr.v19.i7.
671 [42] Ra Y, Reitz RD. A vaporization model for discrete multi-component fuel sprays. Int J
672 Multiph Flow 2009;35:101–17. doi:10.1016/j.ijmultiphaseflow.2008.10.006.
673 [43] Lu T, Law CK, Yoo CS, Chen JH. Dynamic stiffness removal for direct numerical

Page 34 of 37
ACCEPTED MANUSCRIPT
674 simulations. Combust Flame 2009;156:1542–51.
675 doi:10.1016/j.combustflame.2009.02.013.
676 [44] Luo Z, Plomer M, Lu T, Som S, Longman DE. A reduced mechanism for biodiesel
677 surrogates with low temperature chemistry for compression ignition engine
678 applications. Combust Theory Model 2012;16:369–85.
679 doi:10.1080/13647830.2011.631034.
680 [45] H.N.Gupta. Fundamentals of internal combustion engine. 2nd ed. New Delhi: Asoke
681 K. Ghosh, PHI Learning Private Limited; 2006.

PT
682 [46] Han Z, Uludogan A, Hampson GJ, Reitz RD. Mechanism of soot and NOx emission
683 reduction using multiple-injection in a diesel engine. SAE Pap 960633 1996.
684 doi:10.4271/960633.

RI
685 [47] Sazhin SS, Al Qubeissi M, Kolodnytska R, Elwardany AE, Nasiri R, Heikal MR.
686 Modelling of biodiesel fuel droplet heating and evaporation: Effects of fuel

SC
687 composition. Fuel 2015;115:308–18. doi:10.1016/j.fuel.2015.03.051.
688 [48] Pracht WE. Calculating three-dimensional fluid flows at all speeds with an Eulerian-
689 Lagrangian computing mesh. J Comput Phys 1975;17:132–59. doi:10.1016/0021-

U
690 9991(75)90033-9.
AN
691 [49] Patankar VS V. Numerical heat transfer and fluid flow. New York: Hemisphere
692 Publishing Corporation: Washington D.C.; 1980.
693 [50] Shah M, Tariq M, Ali S, Guo Q. Transesterification of jojoba oil, sunflower oil, neem
M

694 oil, rocket seed oil and linseed oil by tin catalysts. Biomass and Bioenergy 2014.
695 doi:10.1016/j.biombioe.2014.08.029.
696 [51] Canakci M. Combustion characteristics of a turbocharged DI compression ignition
D

697 engine fueled with petroleum diesel fuels and biodiesel. Bioresour Technol
698 2007;98:1167–75. doi:10.1016/j.biortech.2006.05.024.
TE

699 [52] Ng J, Kiat H, Gan S. Engine-out characterisation using speed-load mapping and
700 reduced test cycle for a light-duty diesel engine fuelled with biodiesel blends. Fuel
701 2011;90:2700–9. doi:10.1016/j.fuel.2011.03.034.
EP

702 [53] Radwan MS, Abu-Elyazeed OSM, Attai YA, Morsy ME. On the ignition delay of
703 jojoba bio-diesel and its blends with gas oil. SAE Tech Pap 2014;2014–Octob.
704 doi:10.4271/2014-01-2654.
C

705 [54] Selvan VAM, Anand RB, Udayakumar M. Combustion characteristics of diesohol
AC

706 using biodiesel as an additive in a direct injection compression ignition engine under
707 various compression ratios. Energy and Fuels 2009;23:5413–22.
708 doi:10.1021/ef900587h.
709 [55] Graboski MS, McCormick RL. Combustion of fat and vegetable oil derived fuels in
710 diesel engines. Prog Energy Combust Sci 1998;24:125–64.
711 [56] Gonzalez Gomez ME, Howard-Hildige R, Leahy JJ, O’Reilly T, Supple B, Malone M.
712 Emission and performance characteristics of a 2 litre Toyota diesel van operating on
713 esterified waste cooking oil and mineral diesel fuel. Environ Monit Assess
714 2000;65:13–20. doi:10.1023/A:1006446326210.
715 [57] Agarwal AK, Das LM. Biodiesel development and characterization for use as a fuel in

Page 35 of 37
ACCEPTED MANUSCRIPT
716 compression ignition engines. J Eng Gas Turbines Power 2001;123:440–7.
717 doi:10.1115/1.1364522.
718 [58] Dorado MP, Ballesteros E, Arnal JM, Gómez J, López FJ. Exhaust emissions from a
719 diesel engine fueled with transesterified waste olive oil. Fuel 2003;82:1311–5.
720 doi:10.1016/S0016-2361(03)00034-6.
721 [59] Kalam M., Masjuki H. Biodiesel from palmoil-an analysis of its properties and
722 potential. Biomass and Bioenergy 2002;23:471–9. doi:10.1016/S0961-9534(02)00085-
723 5.

PT
724 [60] Kalam MA, Husnawan M, Masjuki HH. Exhaust emission and combustion evaluation
725 of coconut oil-powered indirect injection diesel engine. Renew Energy 2003;28:2405–
726 15. doi:10.1016/S0960-1481(03)00136-8.

RI
727

SC
Nomenclature

Abbreviations
SOI start of injection IVC intake valve closing

U
CI compression ignition EVO exhaust valve opening
AN
CAD bTDC crank angle degree before top HRR heat release rate
dead centre
CAD aTDC crank angle degree after top EGR exhaust gas recirculation
dead centre
M

JME Jojoba oil methyl ester SOC start of combustion


FT-IR Fourier transform infrared
NOx oxides of nitrogen GC-MS Gas chromatography mass
D

spectroscopy
UHC unburned hydrocarbon PMC premixed combustion
TE

IC internal combustion
SI spark ignition MD methyl decanoate
DI direct injection
PM particulate matter ROPR rate of pressure rise
EP

ID ignition delay LHV lower heating value


BSFC brake specific fuel consumption CFD computational fluid dynamics
BTE brake thermal efficiency CR compression ratio
C

SOC start of combustion BTDC before top dead centre


JO Jojoba oil ATDC after top dead centre
AC

DICI direct injection compression


ignition

Symbols Subscripts

γ specific heat ratio T temperature


C1, C2, constants r reference condition
hc convection heat transfer wall cylinder wall
coefficient
p instantaneous cylinder pressure
(bar) gross gross value

Page 36 of 37
ACCEPTED MANUSCRIPT
T mean gas temperature (K) c convection
V cylinder volume
Vr, Tr Pr volume, temperature and
pressure at any reference
condition
θ crank angle
A surface area
ν kinematic viscosity

PT
728

RI
U SC
AN
M
D
TE
C EP
AC

Page 37 of 37
ACCEPTED MANUSCRIPT

Highlights
1. Neat JME gives lower emissions of NOx, CO and UHC as compared to diesel
2. Increasing CR improves peak pressure of JME fueled CI engine by up to 14.5%
3. Neat JME fueled CI engine requires high CR for improved output power
4. Higher CR results in higher NOx emission for both mineral diesel and JME
5. Fully dedicated biodiesel CI engine would enhance engine performance with JME

PT
RI
U SC
AN
M
D
TE
C EP
AC

You might also like