1 s2.0 S0926669023000961 Main

You might also like

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

Industrial Crops & Products 194 (2023) 116332

Contents lists available at ScienceDirect

Industrial Crops & Products


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

Impact of fuel injection pressure on the common rail direct fuel injection
engine powered by microalgae, kapok oil, and soybean biodiesel blend
Paras Gupta a, Upendra Rajak b, Tikendra Nath Verma a, *, Manoj Arya a,
Thokchom Subhaschandra Singh c
a
Department of Mechanical Engineering, Maulana Azad National Institute of Technology, Bhopal 462003, India
b
Department of Mechanical Engineering, Rajeev Gandhi Memorial College of Engineering and Technology, Nandyal, Andhra Pradesh, India
c
Department of Mechanical Engineering, National Institute of Technology, Manipur 795004, India

A R T I C L E I N F O A B S T R A C T

Keywords: In this paper, the viability of microalgae as a viable replacement feedstock for compression ignition engines is
Edible biodiesel investigated. The characteristics of diesel fuel alternatives, including microalgae (MAME100), kapok oil
Non-edible biodiesel (KA100), and soybean (SME100), as well as their blends with diesel, were studied. The results were used to
Fuel injection pressure
compare the engine performance and emissions of D100, MAME100, MAME20, KA100, KA20D80, SME100, and
Particulate matter (PM) emission
SME20D80. The findings demonstrate that the addition of MAME20, KA20D80, and SME20D80 have been
NOx emission
employed to assess engine output and emission. At 200 bar, the brake thermal efficiency (BTE) decreases by 2.0
% for MAME20D80, 1.5 % for KA20D80, and 2.0 % for SME20D80 as compared to D100. Specific fuel con­
sumption (SFC) was shown to rise as KA20D80 levels rose by 3.8 %. The use of MAME20D80 and SME20D80
resulted in reductions of NOx emissions of 10.0 % and 14.2 % as well as soot emissions of 10 % and 4.0 % as
compared to D100. However, employing SME20D80, the blends produce greater carbon dioxide (CO2) emissions
of 11.4 %. In conclusion, a diesel engine can employ MAME20D80 and KA20D80.

Asokan et al. (2016) conducted experimental research on the com­


bustion and performance characteristics of CI (compression ignition)
1. Introduction engines using kapok oil methyl ester as the propellant in 2016. It was
discovered that as the load changes, SFC (specific fuel consumption) and
The world population is expanding quickly, which is driving up en­ EGT (exhaust gas temperature) rise while BTE (brake thermal effi­
ergy demand significantly. The majority of the energy used in the world ciency), CP (cylinder pressure), and HRR fall. The BTE, HC (unburned
today is produced from fossil fuels like coal, gasoline, etc. Developing hydrocarbons), and SE increased by 9.2 %, 31.25 %, and 20 %,
nations that heavily rely on non-renewable energy sources to provide all respectively, but the HRR (heat release rate), MRPR(maximum rate of
of their home and industrial energy needs. Our current energy needs are pressure rise), EGT, and NOx decreased by 2.58 %, 12.9 %, 1.5 %, and
largely satisfied by non-renewable resources. Using fossil fuels despite 4.85 %, respectively, at 240 b TDC and full load in comparison to an
the fact that biofuels have a higher conversion efficiency for a very long engine fuelled with diesel, according to Gnanasekaran et al. (2016)a
period. Some of the sources from which biofuels can be obtained include study. In order to predict the engine characteristics, Khandal et al.
forest debris, non-edible feedstock, and edible feedstock. Biofuels are a (2017) conducted an experimental examination of the effects of EGR
renewable energy source that is derived from natural sources, therefore (exhaust gas recirculation), FIP (fuel injection pressure), and IT (injec­
it may be replenished in a matter of years rather than millions of years. tion timing) on variable compression ratio engines filled with honge
The majority of people view biofuels as a cheap, efficient, and very biodiesel. FIP shows maximum BTE and minimal carbon monoxide
useful source of energy. Using a range of edible and non-edible oils, such (CO), nitrogen oxides (NOx), and hydrocarbon emission at 100 b TDC
as Kapok, Karanja, Microalgae, soybean, and Jatropha, to produce bio­ and 900 bar.
fuels. The physical and chemical characteristics of these alternative fuels The engine emission and performance characteristics of biodiesels
roughly contrast with diesel fuel (Arunkumar et al., 2019; Martin et al., produced by the peroxidation process were experimentally studied by
2017; Cherng-Yuan et al., 2006).

* Corresponding author.
E-mail addresses: tnverma@manit.ac.in, verma.tikks@gmail.com (T.N. Verma).

https://doi.org/10.1016/j.indcrop.2023.116332
Received 30 March 2022; Received in revised form 15 December 2022; Accepted 17 January 2023
Available online 3 February 2023
0926-6690/© 2023 Elsevier B.V. All rights reserved.
P. Gupta et al. Industrial Crops & Products 194 (2023) 116332

Nomenclature BTE Brake thermal efficiency.


HRR Heat release rate.
D100 100% Diesel. CO2 Carbon dioxide.
MAME100 100% Microalgae methyl ester biodiesel. CO Carbon monoxide.
SME100 100% Soybean methyl ester biodiesel. NOx Oxides of nitrogen.
KA100 100% Kapok oil biodiesel. HC Hydrocarbon.
KA20D80 20% Kapok oil biodiesel and 80% diesel. To Temperature (K).
MAME20D80 20% Microalgae methyl ester biodiesel and 80% Po In- cylinder pressure.
diesel. Eactivation Activation energy in the auto-ignition process.
SME20D80 20% Soybean methyl ester biodiesel and 80% diesel. dX
dτ o Heat release rate, 1/sec.
FIP Fuel injection pressure. Ø0 ≈ ØI ≈ ØII ≈ ØIII Function characterising completeness the fuel
IT Injection timing. vapour combustion in the zone.
a TDC After top dead centre. MF Cycle fuel mass.
a BDC After bottom dead centre. A0, AI, AIII Empirical factor controlled by intensity of swirl and speed
b TDC Before top dead centre. of engine.
b BDC Before bottom dead centre. σUD Percentage of fuel evaporated in the present instant of
BSFC Brake specific fuel consumption. time.
MRPR Maximum rate of pressure rise. σU Percentage of fuel evaporated in the course of ignition
EGT Exhaust gas temperature. delay period.
CP Cylinder pressure. α Equivalence fraction of A/F in the cylinder.
CT Cylinder temperature. X Burnt Fuel fraction.
ID Ignition delay. τo Period of auto-ignition delay, sec.
ITE Indicated thermal efficiency.

Lin in et al. (2006). BTE and EGT increased while SFC, CO2, CO, and NOx demonstrated that the NOx emission, particulate matter discharge, and
decreased. Behçet et al. (2014) conducted an experimental investigation SE scale-up with increasing injection timing, swirl ratio, and nozzle
to compare the performance and emission characteristics of waste diameter. Through the KIVA 4 CHEMKIN platform, Zhao et al. (2018)
cooking oil and fish oil in diesel engines. They found that the perfor­ report a computational investigation of the effect of fuel structure
mance and emission characteristics of the fish oil were superior to those (quad-component skeletal mechanism) on soot particle size and mass in
of the waste cooking oil. Another experimental investigation on diesel engines. The findings indicate that soot particle mass and number will
engines running on biodiesel n-butanol blends (KBB5, KBB10, KBB15 increase as unsaturated fatty acid methyl ester (FAME) concentration in
and KBB20) was conducted by Killol in et al. (2019) and discovered that biodiesel fuel increases.
KBB15 exhibits good performance and lower emissions. Another Using edible (soybean oil) and non-edible (kapok oil, microalgae
researcher, Perumal et al. (2017), used pongamia biodiesel as the pro­ methyl ester) oils as a propellant in CI engines, the current work focuses
pellant in CI engines to conduct an experimental examination of engine on the numerical analysis of engine characteristics. According to the
characteristics. It was observed that when diesel was utilised as the fuel, literature study, earlier research on engine characteristics has relied on
CO, HC, and thermal efficiency decreased by 8.2 %, 8.9 %, and 2.4 %, kapok oil, microalgae methyl ester, and soybean oil, however, to the
respectively, but SFC increased by 4.2 %. author’s knowledge, neither effort has taken into account changes in
In an experimental study utilising castor biodiesel as the fuel, both load and fuel injection pressure nor simultaneous application of the
Arunkumar et al. (2019) found that SFC was 4 % higher, CO, HC, and same. Therefore, author has made an effort to close the information gap
thermal efficiency were 9 %, 8.8 %, and 2.2 % higher, respectively, than by conducting a well-organised evaluation on the effects of different
with diesel. Islam et al. (2015), experiment on a CRDI (Common rail loading conditions (25 %, 50 %, 75 %, and 100 %) and fuel injection
direct injection) engine using microalgae methyl ester as the propellant; pressure (180, 200 and 220 bar) on diesel engine performance. Addi­
instantaneous cylinder pressure and engine output torque decreased tionally, the author performed two experiment validations.
while NOx emission increased for a 50 % blend in comparison to diesel.
The performance and exhaust emissions of a diesel engine running on 2. Material and methods
Ceiba pentandra biodiesel were experimentally examined by Silitonga
et al. (2013). It was found that a 10 % mix produced the best perfor­ 2.1. Alternative and alternative properties
mance at 1900 rpm and full throttle. Shrivastava and Verma (2020a)
conducted a study, according to research on the effects of FIP and load The substitute was created by transesterification with methanol
on the characteristics of CI engines charged with roselle oil, ignition using edible soybean oil, non-edible kapok oil, and microalgae species.
delay, summary of emission (SE), brake thermal efficiency, and indi­ According to earlier studies, the soybean oil, microalgae, and kapok oils
cated thermal efficiency scale down as FIP increases, whilst fuel con­ were transformed into alternatives through the chemical reaction pro­
sumption, exhaust gas temperature, maximum pressure rise rate, CO2, cedure of transesterification, in which the oils were shared with ethanol
and NOx scale up. or methanol in the attendance of a compound to produce fatty-acid-
The use of computational fluid dynamics (CDF) has shown to save esters and glycerine. Methanol is typically utilised in place of ethanol
research costs and time while still producing reliable and accurate re­ because it lowers the cost of production for alternatives, increasing their
sults from experimental models. Numerous solver packages, such as ability to compete in the marketplace for petroleum based diesel fuel.
Diesel-RK, Converge CFD, AVL-fire, Kiva code, and ANSYS ICE, are Alternative energy that can be used to reduce reliance on fossil fuels.
capable of simulating CI engines. By adjusting injection timing, nozzle Since the market for edible oils is becoming more competitive and more
diameter, and swirl ratio, Rajak et al. (2020) performed a numerical expensive to produce alternative, there has been a lot of interest in the
analysis on the emission characteristics of diesel engines fueled with use of non-edible oil (Geng et al., 2020; Mofijur M et al., 2013a; b).
diesel, spirulina microalgae, and ethanol mixes. The results The fuels employed in this analysis were diesel (D100), microalgae

2
P. Gupta et al. Industrial Crops & Products 194 (2023) 116332

methyl ester (MAME100), kapok oil (KA100) and soybean methyl ester computer. The eddy current dynamometer was set to Zero(0) position to
(SME100). First, 80 % of diesel fuel was blended with 20 % of micro­ signify the initial zero loading condition. The inbuilt fuel consumption
algae methyl ester, 20 % kapok oil, and 20 % soybean methyl ester by measurement system was initiated for taking the readings of fuel con­
volume basis is called to MAME20D80, KA20D80 and SME20D80, sumption at different loading conditions. The DAQ system collected the
respectively as per investigations by Rajak et al. (2020). The alternative analog signals from the sensors and fed the same to the computer and the
fuel properties as shown in Table 1. As a result, Eq. (1) was applied to same being translated to the GUI. The data is then exported from the GUI
determine the density of the mix (ρ), and Eq. (2) was used to get the during the experiment and the raw data is collected from the computer.
viscosity of the blend (v). The cetane number and carbon percentage of The experimental engine was tested in an air-fuel ratio of 15:1.
blends were calculated using Eqs. (3) and (4). All of the propellant’s Although diesel engines in general operates between 14.7 and 18 and
characteristics come from earlier studies. Table 1 lists the characteristics above, a more rich-mixture based mixture is implemented in the present
of the fuel and its mixtures (Silitonga et al. 2013, Islam et al. 2015, study. This is to facilitate the engine to combust the fuel properly.
Kuleshov and Mahkamov, 2008; Rajak et al., 2019b). Numerically, the tool will exhibit results at any specified air: fuel ratio,
yet the authors have kept similar air: fuel ratio between experimental

n
ρB = X r ρr (1) and numerical works. The richer side of the ratio is to accelerate the
r=1 combustion phenomenon of the different fuels.


n
lnνB = Xr lnυr (2) 2.3. Uncertainty assessment
r=1

An uncertainty analysis has been undertaken to establish the



n
CNB = Xr CNr (3) repeatability of the investigations with the specific goal of identifying
r=1 and enumerating mistakes originating in the experimentations due to
estimating accuracy, instruments used and their calibration, and tech­

n
nique used in an ambient state. In this study, the combined uncertainty
CB = X r Cr (4)
r=1 for engine performance and exhaust fumes has been measured using the
root mean square method, where U is the overall uncertainty of a certain
quantity R, has been listed by employing Eq. (5). In this, R is defined as f
2.2. Investigational system [y1, y2,., yn], where y1, y2,., yn are the independent variables of R and
each of them has its own individual error. Table 3 displays the sensors
A naturally aspirated, single cylinder, four-stroke, direct injection, and the measurement devices uncertainty (Bhowmik et al., 2018; Rav­
compression-ignition diesel-engine is used in the trials, along with an ikumar et al., 2022, 2022). The Eq. (5) has been used to calculate the
eddy current dynamometer. Fig. 1 shows a schematic representation of total uncertainty for the engine and exhaust emissions based on the
the diesel engine test arrangement. These kinds of engines are frequently average values of five consecutive readings collected over a 2-minute
used for research. The type of fuel injector used for the present study is a sample period. Additionally, a further standard deviation index has
diesel solenoid injector (Make: BOSCH) with 3 holes and 120◦ spray been calculated in order to increase the reliability of the uncertainty
angle with injection pressure upto 1100 bar. Table 2 provides technical analysis.
information about the diesel engine. To measure engine performance [
exhaust emissions under various operating situations, simple in­ ΔU = √ (dynamometer)2 + (smoke sensor)2 +
struments are used. A computer system with ignition programming al­ (pressure sensor)2 + (speed senor)2 +
(temperature sensor)2 + (crank angle encoder)2 + (5)
lows for the study of tested samples engine presentation and ignition
(load indicator)2 + (NOx sensor)2 + ]
data. With the help of this programme, estimates of variables like cyl­
(CO2 sensor)2 + (pressure experiment)2
inder pressure, heat release rate, fuel flow rate, air measurement, calo­
rimeter water flow rate, and others can be made. The TESTO-350 gas [
ΔU = √ (0.15)2 + (1.0)2 + (0.5)2 +
analyser measures carbon-monoxide, carbon-dioxide, nitrogen-oxide,
and other gases at the engine tail pipe. For determining smoke opacity (1.0) + (0.15)2 + (0.2)]2 + (0.2)2 +
2

(1.0)2 + (0.5)2 + (3.5)2 ΔU = 3.98%


concentration, an AVL smoke metre is used. Given the unpredictability
of all parameters, the equipment used in this experiment are periodically
tuned. The readings were taken after running the engine in diesel mode 2.4. Numerical model description
for 30 min to acquire stabilised figures. The exhaust gas analyzer was
calibrated without gas before the experiment. In the current study, Diesel-RK software has been utilised to
The experimental set-up was checked for any possible errors and numerically analyse the performance, combustion, and emission char­
later the engine was set to run on diesel fuel for 10 h to stabilise before acteristics of a diesel engine using neat microalgae (MAME100), kapok
commencement of the experiments. The throttle was set to fully open oil (KA100), and soybean (SME100), its 20 % blend with diesel. We
position during the initial stage of operation. The engine was manually looked into the connections between crank angle and cylinder pressure,
cranked and the graphical user interface (GUI) was opened in the temperature, heat release rate, pressure rise rate, and other important

Table 1
Characteristics of fuel.
Characteristics D100 MAME100 KA100 SME100 MAME20D80 KA20D80 SME20D80

C (%) 86.14 78.41 78 77.3 84.27 84.51 84.07


H (%) 13.86 11.12 12.5 11.8 13.31 13.58 13.44
O (%) 0.4 10.47 9.5 10.8 2.414 2.22 2.48
Flash point 64 95 156.5 120 70.2 82.5 75.2
Density (kg/m3) 830 912 876.9 885 846.24 839.38 841
Viscosity (mm2/s) 3.61 5.06 4.61 5.23 3.9 3.81 3.934
Heating value (MJ/kg) 42.5 37.42 40.825 36.2 41.484 42.165 41.24
Cetane Number 48 46.5 57.2 51.3 47.7 49.84 48.66

3
P. Gupta et al. Industrial Crops & Products 194 (2023) 116332

Fig. 1. (a) Experiment setup. (b). Picture of experimental system.

Table 2 Table 3
Specification of engine. Uncertainty assessment.
System specification Standards Tool Percentage of indecision

Engine type 1-cylinder, 4-stroke, CI engine Eddy current dynamometer ±0.15


Bore (mm)×Stroke (mm) 80 × 110 Smoke meter ±1.0
Compression ratio 17.5 Pressure sensor ±0.5
Rated power (KW) 3.7 Speed sensor ±1.0
Speed (rpm) 1500 Temperature sensor ±0.15
Method of cooling Water Encoder ±0.2
Fuel injection timing 230 b TDC Load indicator ±0.2
NOX ±0.5
CO2 ±1

4
P. Gupta et al. Industrial Crops & Products 194 (2023) 116332

parameters. From the diesel engine experiment, simulated data and


experimental data were contrasted. The Diesel-RK tool, which makes use
of the multi-zone concept, was used to simulate the combustion of an
artificial compression ignition engine. This model takes into account the
governing equations that Fiveland and Assanis specified using the
Diesel-RK programme. The Zeldovich Mechanism is utilised to produce
nitrogen oxide emissions, while the Alkidas Mechanism, and Mechanism
are used to predict ignition delay, soot, and particle matter. Diesel spray
zone characteristics are shown in Fig. 2, and vaporisation is shown in
Fig. 3.
Spray model:

(a) Earlier jet impingement


I. Firm conical core Fig. 3. Vaporisation of spray of CI engine (Kuleshov et al., 2010).
II. Dense progressive front
III. Attenuated outer sleeve d(mYi ) ∑ j
(b) Latterly spray impingement = mjY + Sg (8)
dt j
i

I. Conical axial core at near wall flow ∑(mj )( j ) Ωi Wmw


II. Dense core of near wall flow on piston bowl surface Yi = Y i − Y cyl
i + (9)
m ρ
III. Latterly front of near wall flow j

IV. Attenuated outer zone of near wall flow Conservation of energy.


According to Fiveland and Assanis, the energy equation for an open
Penetration spray tip thermodynamic system can be expressed as follows:
( )1.5 ( )
V1 L1 d(mu) dν dQht ∑
V0
= 1−
L2 =− p + + mj hj (10)
dt dt dt j

Where;L1 = prevailing distance between location of EFM & nozzle The rate at which energy changes within the system is indicated on
injectorL2 =penetration distance of EFMV0 = EFM initial velocity at the left hand side of the equation above. The displacement work rate,
nozzle of the injectorV1 = ddLτk ; EFM current velocity. heat transfer rate, and enthalpy flow are each represented by the first,
Heat release rate and ignition delay. second, and third terms on the right side of the equation.
The Diesel-RK software multi-zone combustion model divides the Modified Tolstov’s equation.
heat release process into four essential parts. Although they have Although their physical and chemical characteristics differ, each has
different physical and chemical properties, each one affects how quickly an impact on the pace at which heat is released. The first stage of the
heat is released. The ignition delay period, the initial phase of heat heat release is the ignition delay period, which is calculated using the
release, is computed using a modified form of Tolstov’s equation. modified Tolstov’s equation as follows:
The following equations describe the species conservation equation, The calculating of auto ignition delay period by:
which takes into account the evaluation and extinction of each species (
√̅̅̅̅̅ (
) To
)
Eactivation 70
on a mass fraction basis. τo = 3.8 × 10− 6 1 − 1.6 × 10− 4 × rpm exp −
Po 8.312To CN + 25
dm ∑ (11)
= mj (6)
dt j Heat release rate.
During premixed combustion phase
mi
Yi = (7) [ ( ) ] ( )
m dX MF dσUD
= ∅o A0 (σUD − x0 )(0.1σUD + Xo ) + ∅I (12)
dτo vI dτ

Fig. 2. Several of zones of fuel (diesel) spray (Kuleshov et al., 2010).

5
P. Gupta et al. Industrial Crops & Products 194 (2023) 116332

During the mixing-controlled combustion phase Table 4


( ) ( ) Validation conditions.
dX dσU MF
= ∅I + ∅II AII (σU − X)(α − X) (13) System specification Validation
dτo dτ vC
Engine type 4-stroke, one cylinder, CI engine
During the late combustion phase Bore (mm)×Stroke (mm) 80 × 110
Compression ratio 17.5
dX
= ØIII × AIII × Kt × {1 − X} × {ξB α − X} (14) Rated power (kW) 3.7
d τo Speed (rpm) 1500
Method of cooling Water
It is assumed in those equations that Φ0 = ΦI = ΦII = ΦIII which Fuel injection timing 23◦ b TDC
specifies the degree of fuel vapour combustion in the zones, is true. The
cylinder heat transfer is considered, and Woschni’s relationship is used
to get the corresponding heat transfer coefficients for each of its many
Table 5
zones.
Boundary condition for validation.
Specific fuel consumption.
Brake specific fuel consumption is the amount of gasoline consumed Parameter Validation 2 (Rajak and Verma, 2018b)

to produce one unit of braking power from an engine. Injection timing 230 b TDC
Fuel injection pressure 180–220 bar
MF Inlet valve opening 4.50 b TDC
SFC = (15)
PBrake Inlet valve closing 35.50 a BDC
Exhaust valve opening 35.50 b BDC
Brake thermal efficiency Exhaust valve closing 4.50 a TDC

PBrake
ηBrake = (16)
MF × CV analysed simultaneously, it was discovered that the peak cylinder
Equation calculates the thermal NO using the chain Zeldovich pressure at full load varied by 3.5 % of validation I and 4.64 % of
method and offers the following details: (17− 20). The full set of equa­ validation II, respectively. In the experiment, a diesel engine with a
tions consists of fourteen equilibrium equations, three equations for the 3.7 kW rated output was used. The analysis for the second validation-II
material balance, and the Dalton equation for partial pressure. Each of the experimental and numerical results is presented in Table 5. Cyl­
stage requires computing the equilibrium concentrations of eighteen inder pressure was examined for Rajak and Verma (2018b) as shown in
species. Fig. 5. On the basis of historical data for internal combustion engines,
reputable researchers are implementing the Diesel-RK programme.
O2 ↔ 2O Several researchers, including Kuleshov et al. (2010) and Kuleshov and
Mahkamov (2008), validated the DIESEL-RK SOFTWARE by comparing
N2 + O ↔ NO + O
actual and model results. Using engine parameters from Kuleshov et al.,
2008, the multi zone model was utilised to investigate the combustion
N + O2 ↔ NO + O (17)
characteristics of a compression ignition machine running in HCCI
38020
− 7 − T
{ ( / )2 } mode.
d[NO] P × 2.333 × 10 .e b [N2 ]e .[O]e . 1 − [NO] [NO]e 1
= ( ) . (18)
dθ R.Tb . 1 + 2365.e
2365
Tb [NO]
. ω 3. Result and discussions
Tb [NO]
e

The NO concentration in a cylinder is given by: 3.1. Combustion characteristics

rNOc = rNO (19)


3.1.1. Cylinder pressure
And the specific NO in g/kWh is expressed as Cylinder pressure determines a fuels capacity to mix uniformly with
air and burn effectively. According to Nautiyal et al. (2020), the com­
30 × rNO × Mbg
eNO = (20) bustion properties of DI engines are significantly influenced by the
L C × ηM
chemical and physical qualities of the fuel used. The variance in cylinder
Soot formation. pressure for the D100, MAME 100, SME 100, KA100, KA20D80, MAME
The Hatridge smoke level, which is represented as given in equation, 20D80, and SME 20D80 engines is shown in Fig. 6 as a function of crank
can be used to calculate the level of soot generation (21− 22). The angle. The cylinder pressure for the D100 was 90.986 bar at full load,
calculation determines how much soot will form in the burning zone. compared to 81.885, 80.694, 86.03, 90.105, 88.77, and 88.506 bar for
( ) the MAME 100, SME 100, KA100, KA20D80, MAME20D80, and SME
d[C] q dx
= 0.004 c (21) 20D80, respectively. The MAME100 may have had the lowest cylinder
dt K V dt pressure of all the biodiesels tested due to its high viscosity and low
calorific value. Diesel has a somewhat higher cylinder pressure than
Hartridge = 100{1 − 0.9545exp( − 2.4226[C])} (22)
both pure biodiesel and its mixes. A shorter ignition delay is also asso­
ciated with lower cylinder pressure. As more fuel is charged into the
combustion chamber, cylinder pressure rises. From Fig. 6, it was
2.5. Diesel-RK authentication
observed that the in-cylinder pressure of all fuels combusted inside the
cylinder was similar in-nature. This is not the case and hence the peak
The Diesel-RK simulation programme has been validated using the
value at the corresponding crank angle has been taken out for all the
various engine appearances of a diesel-powered CI engine. The bound­
fuels under study and plotted as “peak cylinder pressure” variation at
ary conditions that were first used in numerical simulation and experi­
different loading and fuel injection pressure in Fig. 7. In Fig. 7, the
mental results are shown in Tables 4 and 5. The study of the
variation in peak cylinder pressure corresponding to the load and fuel
experimental and numerical data for cylinder-pressure is shown in Fig. 4
injection pressure is shown. The D100 cylinder pressure decreases by
(validation-I for present investigation). Fig. 5 illustrates a further vali­
4.66 % and 9.74 %, respectively, when comparing 220 bar fuel injection
dation for the cylinder pressure using the same operating conditions and
pressure to 200 bar and 180 bar fuel injection pressure. Similar to this,
prior data. When the numerical and experimental findings were

6
P. Gupta et al. Industrial Crops & Products 194 (2023) 116332

Fig. 4. Validation of cylinder pressure.

Fig. 5. Validation of a) cylinder pressure.

the reductions for MAME 100, SME 100, KA 100, KA20D80, MAME 3.1.2. Heat release rate
20D80, and SME 20D80 at 200 bar fuel injection pressure are 4.40 %, The chart clearly shows that the rate of heat release increases as the
4.46 %, 5.07 %, 4.55 %, 4.58 %, and 4.58 %, respectively, while at percentage of biodiesel in the mixed fuel increases and happens further
180 bar fuel injection pressure, the reductions are 9.74 %, 9.83 %, 10.1 from top dead centre, which is also a sign of ignition delay. The greater
%, 11.1 %, 10.9 %, and 9.6 %, as shown in figure (Shrivastava and ignition delay and delayed start of combustion may be to blame for the
Verma, 2020a; Dunga et al., 2022). higher rate of heat release. This type of behaviour is also caused by the
Due to dissimilar nature of the fuel characteristics under test (diesel, greater cetane number and latent heat of evaporation of biodiesel. More
biodiesel blends), they exhibit different combustion characteristics. One fuel is burnt during the premixed phase of combustion as a result of the
such factor is the variation of calorific value of the different fuel blends. prolonged ignition delay and fuel accumulation. Additionally, the bio­
The peculiar significance of Fig. 7 is to show the difference of the peak diesel molecular structure contains oxygen, which enhances the diffu­
cylinder pressure values upon combustion in the cylinder. In most of the sion stage of combustion. A higher instantaneous heat release rate
cases, diesel was observed to have higher values compared to other fuel. results from the higher volatility, which also causes a bigger volume of
This is due to the fact that higher calorific fuel exhibit better combus­ fuel to accumulate during the premixed phase combustion. Additionally,
tibility and hence higher cylinder pressure. the findings demonstrate that the biodiesel blended fuel exhibits a larger

7
P. Gupta et al. Industrial Crops & Products 194 (2023) 116332

Fig. 6. Cylinder pressure with crank angle.

Fig. 7. Peak cylinder pressure with load, FIP and biodiesel blends.

neat heat release rate compared to diesel fuel. This is because biodiesel 2007). Fig. 9 illustrates the divergence of the heat release rate as a
has a larger mass proportion of oxygen and a higher latent heat of function of load and fuel injection pressure. It demonstrated that the
evaporation than diesel fuel (Datta et al., 2016; Goodrum, 2002). heat release rate for D100 decreases by 11.12 % and 26.12 %, respec­
The quantity of heat emitted during a combustion process at a tively, at 200 and 180 bar fuel injection pressures compared to 220 bar
particular crank angle is referred to as the heat release rate. A CI engine fuel injection pressure. For MAME 100, SME 100, KA 100, KA 20D80,
goes through multiple stages of combustion, such as premixed, rapid, MAME 20D80, and SME 20D80, the fuel injection pressure is lowered by
regulated combustion, and the afterburning phase (Rajak et al., 2019a). 13.7 %, 15.1 %, 11.9 %, 12.77 %, 11.4 % and 12.2 %, respectively, at
At full load, Fig. 8 displays the change in heat release rate versus crank 200 bar, while it is reduced by 22.5 %, 30.3 %, 24.1 %, 28.8 %, 27.12 %,
angle for the D100, MAME 100, SME 100, KA100, KA20D80, MAME and 27.8 %, respectively, at 180 bar.
20D80, and SME 20D80. At full load, the heat release rates were 44.668,
35.37, 31.42, 32.92, 42.99, 42.33, and 41.7 J/deg for D100, MAME 100, 3.1.3. Ignition delay
SME 100, KA20D80, MAME20D80, and SME20D80, respectively. The The time interval between the beginning of fuel injection and the
law of thermodynamics is used to calculate the rate of heat emission. It beginning of fuel combustion is known as the ignition delay. Ignition
was shown that the heat release rate of biodiesel and its blends is lower delay includes both a physical delay as well as a chemical delay (thanks
than that of pure diesel. The rate of fuel-air mixing during the premixed to the pre-combustion process) (accredit of disintegration, vaporisation
combustion phase, the fuel’s calorific value, and the ignition delay all and blending of air-fuel occur) Lakshminarayanan et al. (2022). The
affect how much heat is generated (Rajak and Verma, 2018; Agarwal, physical and chemical delays manifest at the same time. It is a key

8
P. Gupta et al. Industrial Crops & Products 194 (2023) 116332

Fig. 8. Heat release rate with crank angle.

Fig. 9. Heat release rate with load, fuel injection pressure.

consideration when analysing the likelihood of knocking as well as the by 0.766 %, 1.182 %, 0.0613 %, and 10.8 %, respectively, while falling
fuel’s ability to ignite. Fig. 10 depicts how the ignition delay varies with for SME100 and KA20D80 at 200 bar in comparison to its value at
load; it was found that for D100, it was 12.903. While it decreases for 220 bar fuel injection pressure at full load. Again, at 180 bar for
SME100, KA100, KA20D80, and SME20D80, respectively, it increases MAME100, SME100, KA100, KA20D80, MAME20D80, and SME20D80,
for MAME100 and MAME20D80, respectively, by 7.564 % and 1.108 %. it was seen that it increases by 1.35 %, 0.208 %, 1.714 %, 0.565 %,
This may be the result of MAME100 and MAME20D80’s biodiesels 0.5135 % and 10.82 % accordingly as it compares at 220 bar fuel in­
having a lower cetane number than diesel whereas other biodiesels and jection pressure. With an improved delay period, charges that accu­
blends have a greater cetane number. The engine speed, oxygen content, mulate in the combustion chamber burn quickly, cutting the time it
and cetane number all affect how quickly the ignition starts (Shrivastava takes for combustion to occur. The environment’s pressure and tem­
and Verma 2020b). The cylinder surface and continued gas temperature perature have a significant impact on the ignition delay.
cause the ignition delay to rise as load decreases (Rajak et al., 2019b).
For diesel, biodiesel, and their blends, the variation of ignition delay 3.1.4. Cylinder temperature
versus load versus fuel injection pressure is shown in Fig. 10. In com­ The findings show that for all of the tested fuels, the in-cylinder
parison to ignition delay at 220 bar fuel injection pressure, it was found temperature rises as the load situation increases. The temperature in­
that for D100, ignition delay increased by 0.1395 % & 0.658 % for 200 & side the cylinder rises as the load is larger because more fuel is
180 bar. It rises for MAME100, KA100, MAME20D80, and SME20D80 consumed. The in-cylinder temperature for biodiesel is greater during

9
P. Gupta et al. Industrial Crops & Products 194 (2023) 116332

Fig. 10. Variation of ignition delay with FIP.

the entire test condition. The primary cause of the decrease in maximum and SME20D80 were observed to be 1939.7, 1755.3, 1735.6, 1730.9,
in-cylinder temperature is the higher latent heat of evaporation of bio­ 1790.7, 1896.7, and 1893 K. With the increase in load, cylinder tem­
diesel compared to the biodiesel blends. The equivalent temperatures of perature increased. The temperature of the combustion cylinder has a
evaporation for neat SME100 biodiesel, SME20D80 blend, and diesel significant impact on engine knocking and pollutants. The amount of
fuel are 325 kJ/kg, 265 kJ/kg, and 250 kJ/kg, respectively. The varia­ NOx released is significantly increased by higher cylinder temperature.
tion of cylinder temperature vs load at 220 bar fuel injection pressure is The researcher noticed the same pattern as well (Rajak et al., 2018a).
shown in Fig. 11. As observed from the figure, the lowest values of For diesel, biodiesel, and their blends, it was shown that as fuel injection
cylinder temperature are for the fuel- MAME100 & SME100, followed by pressure is increased, cylinder temperature rises. At full load, the fuel
KA 100. This is due to the fact that the fuels under test were not blended injection pressure for D100, MAME100, SME100, KA100, KA20D80,
with diesel at any proportion and they are subjected as it is inside the MAME20D80, and SME20D80 decreased by 3.07 %, 3.23 %, 3.29 %,
cylinder. Since calorific value of the fuels are low, they will naturally 2.91 %, 3.079 %, and 3.074 % at 200 bar, respectively, while it
exhibit lower heat dissipation during combustion. This is observed as decreased by 6.61 %, 7.54 %, 7.58 %, 7.314 %, 7.09 %, 6.606 %, and
“cylinder temperature”. Yet, upon blending with diesel at 20:80 bio­ 6.62 % at 180 bar.
diesel: diesel ratio, the calorific values of the blend fuels significantly
increased and therefore exhibit higher cylinder temperature upon
combustion. The result may also vary depending upon the operating 3.2. Performance characteristics
conditions of the engine. At full loading condition, cylinder temperature
values for D100, MAME 100, SME100, KA100, KA20D80, MAME20D80, 3.2.1. Specific fuel consumption
One kilowatt of output power is produced in one hour using a specific

Fig. 11. Cylinder temperature variation with load, FIP and biodiesel blends.

10
P. Gupta et al. Industrial Crops & Products 194 (2023) 116332

amount of fuel, which is the fuel consumption. In particular, it measures %, 32.6 %, 32.4 %, and 32.3 % at 220 bar with full load conditions. The
how well an engine can convert the chemical energy of fuel into useful findings show that for all of the tested fuels, the brake thermal efficiency
work output. Flaming the fuel and producing the necessary work output rises as engine load or injection pressure rises.
is a crucial factor in determining the CI engine efficiency. The specific
fuel fluctuation with load at 220 bar fuel injection pressure is shown in 3.2.3. Exhaust gas temperature
Fig. 12. For D100, MAME100, SME100, KA100, KA20D80, Fig. 14 shows how the tested fuels exhaust gas temperature changes
MAME20D80, and SME20D80 biodiesel and blends, the BSFC values as a function of engine load and injection pressure. The highest exhaust
under full load condition are notable to be 0.257, 0.298, 0.311, 0.270, gas temperature is depicted, the values of 926, 889.2, 868.1, 925.7,
0.259, 0.264, and 0.266 kg/kWh, respectively. The lowest BSFC value 923.2, 920, and 916.2 K for D100, MAME100, SME100, KA100,
for diesel is 0.257 kg/kWh, and the second-lowest figure for KA100 is KA20D80, MAME20D80, and SME20D80, respectively. In comparison to
0.259 kg/kWh. The variations in the energy content (calorific value) of diesel, the temperature of the exhaust gas is lower for biodiesel and its
the blended fuels are the primary cause of this type of behaviour. Lower blend. Due to the premixed and diffusion combustion processes, it relies
fuel energy content means higher fuel consumption to produce one unit on the combustion process and the calorific value of the fuel. Due to
of braking power (Datta and Mandal, 2017). This chart makes it evident improved combustion, exhaust gas temperature improves when engine
that, in comparison to diesel under operational conditions, its value rises load increases Karthickeyan (2019). The amount of oxygen in the pro­
when the blend ratio is increased. The calorific value, density, air fuel pellant also plays a role. Increased compression ratio also results in a
ratio, and viscosity all play a role. Higher viscosity and density cause a decrease in exhaust gas temperature (Rajak et al., 2019c). According to
higher specific fuel consumption despite a lower calorific value. Blended research, fuel injection pressure decreases and load increases are
fuel has a higher density and a lower heating value than diesel since accompanied by an increase in exhaust gas temperature. Another
biodiesel is an oxygenated fuel, which aids in improved combustion important factor in determining the temperature of the exhaust gas is the
(Rajak et al., 2019b, 2020). latent heat of propellant vaporisation. The NOx emission can be mini­
mised by lowering the exhaust gas temperature (Nanthagopal et al.,
3.2.2. Brake thermal efficiency 2018).
Brake thermal efficiency is the proportion of produced braking
power to heat input. Fig. 13 shows how the tested fuels brake thermal
3.3. Discharge characteristics
efficiency changes as a function of engine load and injection pressure.
The graphic shows that with increased load, brake thermal efficiency
3.3.1. CO2 discharge
rises due to enhanced combustion. The fuel calorific value, viscosity,
Fuel combustion produces carbon dioxide (CO2), which builds up
lower heating value, and density affect the brake thermal efficiency
and has a negative impact on the environment. Today, CO2 emissions are
according to Muralidharan and Vasudevan (2011). It was found that
a big concern for the entire world, and several international agreements
when the 20 % blend ratio uses, the value of brake thermal efficiency
make it a top priority to reduce them. Fig. 15 shows how the tested fuels
decreases as per investigation. Low calorific value, high viscosity, and
carbon dioxide changes as a function of engine load and injection
density result in the fuel spray qualities, which in turn impact how the
pressure. With increasing load, CO2 emissions are dropping. At 220 bar
fuel and air combine and burn, as well as the fuel, low thermal efficiency
fuel injection pressure and full load operating conditions, it was revealed
when used as a brake. It was discovered that the brake thermal efficiency
that 828.81, 866.4, 892.5, 841.2, 831, 830, and 830.94 g/KWh for
increases as the fuel injection pressure increased. The biodiesel high
D100, MAME 100, SME100, KA100, KA20D80, MAME20D80, and
oxygen content, which causes apodictic combustion of the fuel inside the
SME20D80. Due to its higher oxygen content than diesel, biodiesel has
combustion chamber. For D100, MAME 100, SME 100, KA100,
an excessively high CO2 value (Karthickeyan et al., 2019). Complete
KA20D80, MAME 20D80, and SME 20D80, the value of brake thermal
combustion occurred with a high oxygen concentration, resulting in
efficiency waste was determined to be 32.72 %, 31.76 %, 31.48 %, 32.2
higher emissions than diesel. Additionally, as the compression ratio

Fig. 12. Specific fuel consumption variation with load, FIP and biodiesel blends.

11
P. Gupta et al. Industrial Crops & Products 194 (2023) 116332

Fig. 13. BTE variation with load, FIP and biodiesel blends.

Fig. 14. EGT variation with load, FIP and biodiesel blends.

rises, carbon dioxide emissions climb as well because the fuel combi­ burning of a hydrocarbon propellant should only produce CO2 and water
nation is more combustible (Rajak et al., 2019c). It was observed that, (H2O), which depends on the carbon hydrogen ratio in the propellant.
with the exception of KA100 and KA20D80, for which it declines, the This is also affected by the carbon-hydrogen ratio in the fuel. However,
value increased with a drop in fuel injection pressure for D100, MAME some research has found that CO2 emissions from biodiesel are lower
100, SME100, MAME20D80, and SME20D80. For D100, MAME 100, than those from diesel (Shrivastava and Verma, 2020a; Labeckas and
SME100, MAME20D80, and SME20D80, the value increased by 1.87 %, Slavinskas, 2006), despite the fact that it has a higher CO2 content than
4.4 %, 4.14 %, 1.848 %, and 2.148 % at 180 bar compared to 220 bar diesel.
fuel injection pressure at full load, and by 0.6925 %, 1.518 %, 1.319 %,
0.4578 %, and 0.876 % at 200 bar. At 180 bar and 200 bar fuel injection 3.3.2. SO2 discharge
pressure as compared to 220 bar pressure, CO2 emission drops by 4.6 %, Fig. 16 shows how the tested fuels fluctuation of SO2 emission
6.3 % for KA100 and 2.29 %, 3.89 % for KA20D80. This KA100 and changes as a function of engine load and injection pressure. It shows that
KA20D80 anomaly could be caused by a greater equivalency ratio. The under operating conditions, SO2 emission decreases as load increases.

12
P. Gupta et al. Industrial Crops & Products 194 (2023) 116332

Fig. 15. CO2 variation with load, FIP and biodiesel blends.

Fig. 16. SO2 variation with load, FIP and biodiesel blends.

For D100, MAME100, SME100, KA100, KA20D80, MAME20D80, and 2.26 %, and 2.15 %, while at 200 bar fuel injection pressure, it decreases
SME20D80, it was discovered that 0.01029, 0.00597, 0.00624, 0.00541, by 1.85 %, 1.50 %, 1.28 %, 0.92 %, 0.64 %, 0.94 %, and 0.83 % as
0.00935, 0.00953, and 0.00961 g/KWh. The sulphur content of the opposed to 220 bar fuel injection pressure, as. The SO2 emission is
gasoline is what drives SO2 emissions. When compared to biodiesel, SO2 declining in the rich fuel combination because less additional air is
emissions are high. It was discovered that as fuel injection pressure is required for SO2 generation. In addition, compared to conventional
reduced, SO2 emissions rise. At 180 bar fuel injection pressure, it de­ fuels, the biodiesels have a relatively low sulphur concentration.
creases for D100, MAME 100, SME 100, KA100, KA20D80, MAME
20D80, and SME 20D80 by 1.846 %, 4.61 %, 4.17 %, 2.52 %, 2.21 %,

13
P. Gupta et al. Industrial Crops & Products 194 (2023) 116332

Fig. 17. NOX variation with load, FIP and biodiesel blends.

3.3.3. NOX discharge ignition delay, and fuel properties like oxygen content, viscosity, and
Fig. 17 shows how the tested fuels fluctuation of nitrogen oxides calorific value. The temperature of the exhaust gases has an impact as
(NOx) emission changes as a function of engine load and injection well. Heat release rate and cylinder temperature in an engine also
pressure. NOx emission was estimated to be 2361.2, 2487, 2126, 2967, control NOx generation (Singh et al., 2018). It was found that NOx
2504.4, 2120.1, and 2047.6 ppm for D100, MAME100, SME100, KA100, emission for KA100, KA20D80, MAME100, and MAME20D80 are higher
KA20D80, MAME20D80, and SME20D80, respectively, at full load and than D100. A researcher also made a similar finding (Silitonga et al.,
220 bar fuel injection pressure. The amount of nitrogen oxide that is 2013; Islam et al., 2015, 2019). The higher iodine value of the charge is
released depends on the engine’s torque, load, cylinder pressure, the cause of the increased NOx emission. NOx emissions with long-chain

Fig. 18. PM variation with load, FIP and biodiesel blends.

14
P. Gupta et al. Industrial Crops & Products 194 (2023) 116332

unsaturated fatty acids side by side (Rahman et al., 2014). The pro­ appearances in diesel engines.
duction of NOx emissions is also governed by the air-fuel ratio. Ac­
cording to figure, the NOx emission rises as the fuel injection pressure 4. Conclusion
rises and load.
Using both edible and non-edible oil, the impact of fuel injection
3.3.4. PM discharge pressure and load on the engine parameters of a CI engine was investi­
PM release occurs when propellant burns in inadequate air, which is gated. The following findings are drawn from the current investigation.
a prominent source of pollution from CI engines. The fluctuation in PM
discharge with load and fuel injection pressure is shown in Fig. 18. The • As fuel injection pressure rises, in-cylinder pressure rises as well. For
graph clearly shows that fuel injection pressure has a negative impact on the KA20D80, the peak cylinder pressure at 220 bar is 90.01 bar as
PM discharge, which means that as fuel injection pressure rises, PM opposed to 85.9 bar at 200 bar. Diesel has slightly greater cylinder
discharge falls. At full load, the values are 0.905, 1.220, 1.289, 0.989, pressure than KA20D80 mix at 180 bar. Fuel injection pressure led to
0.936, 0.980, and 0.990 g/kWh at 180 bar for D100, MAME100, an increase in HRR as well.
SME100, KA100, KA20D80, MAME20D80, and SME20D80. At 200 bar, • The value of SFC drops when fuel injection pressure rises from
these values are: 0.758, 0.985, 1.047, 0.812, 0.756, 0.831, and 0.840 g/ 180 bar to 220 bar; however, the KA20D80, MAME20D80, and
kWh; at 220 bar, they are: 0.638, 0.819, 0.874, 0.665, 0.641, 0.700, and SME20D80 mix exhibits somewhat greater BSFC when compared to
0.709 g/kWh. When there is not enough air, the propellant molecules diesel at 220 bar.
thermally crack, causing oxygen to split in the presence of acetylene and • In comparison to neat diesel, the BTE of the KA20D80, MAME20D80,
then polymerise into carbon-rich macromolecules that clump together and SME20D80 blends is lower at 220 bar. For all of the tested fuel,
to form particles (Lakshminarayanan and Aghav, 2010). It relies on the BTE rises as fuel injection pressure rises, reaching a maximum at
oxygen content, aromatic content, and heat release rate during the 220 bar. When compared to its BTE at 200 bar, the BTE of the
premixed combustion phase. The length of the biodiesel’s carbon chain KA20D80 blend at 220 bar injection pressure rises by 1.6 % and 0.73
also affects PM discharge (Rajak et al., 2019c; Wei, 2017). % at low and high load conditions, respectively.
Table 6 displays a revision of the evaluation. The updated evaluation
document explains how biodiesel-diesel test fuels affect emission

Table 6
An overview of the test fuels’ performance and emission characteristics with biodiesel-diesel blends.

15
P. Gupta et al. Industrial Crops & Products 194 (2023) 116332

• The CO2 and SO2 emission reduces with increase in injection pres­ on diesel engine emissions for diesel and biodiesels: a review. Energy Convers.
Manag. 80, 329–356. https://doi.org/10.1016/j.enconman.2014.01.020.
sure. KA100 has lowest SO2 emission than another biodiesel and
Islam, Muhammad Aminul, Rahman, M.M., Heimann, Kirsten, Nabi, Md. Nurun,
diesel. Ristovski, Z.D., Dowell, Ashley, Thomas, George, Feng, Bo, Alvensleben, Nicolas von,
• As injection pressure and load increases, NOx discharge also rises. Brown, Richard J., 2015. Combustion analysis of microalgae methyl ester in a
When the load is 180, 200, or 220 bar for the KA100, the NOx common rail direct injection diesel engine. Fuel 143, 351–360. https://doi.org/
10.1016/j.fuel.2014.11.063.
discharge rises by 57.4 %, 63.76 %, and 66.98 %, respectively. The Karthickeyan, V., 2019a. Effect of cetane enhancer on Moringa oleifera biodiesel in a
minimal NOx discharge for SME20D80 gasoline. thermal coated direct injection diesel engine. Fuel 235, 538–550. https://doi.org/
• As injection pressure rises, PM discharge falls. As load increases, PM 10.1016/J.FUEL.2018.08.030.
Karthickeyan, V., Ashok, B., Nanthagopal, K., Thiyagarajan, S., Edwin Geo, V., 2019b.
discharge decreases by 47.2 %, 46.1 %, and 48.2 % for SME100 Investigation of novel Pistacia Khinjuk biodiesel in DI diesel engine with post
gasoline at 180, 200, and 220 bar injection pressure, respectively. combustion capture system. Appl. Therm. Eng. 159, 113969 https://doi.org/
10.1016/j.applthermaleng.2019.113969.
Khandal, S.V., Banapurmath, N.R., Gaitonde, V.N., 2017. Effect of exhaust gas
CRediT authorship contribution statement recirculation, fuel injection pressure and injection timing on the performance of
common rail direct injection engine powered with honge biodiesel (BHO. Energy
Paras Gupta: Methodology, Resources, Writing – original draft. 139, 828–841. https://doi.org/10.1016/j.energy.2017.08.035.
Killol, Abhijeet, Niklesh Reddy, Santosh Paruvada, Murugan, S., 2019. Experimental
Upendra Rajak: Resources, Writing - Review & Editing. Tikendra Nath studies of a diesel engine run on biodiesel n-butanol blends. Renew. Energy 135,
Verma: Supervision. Manoj Arya: Resources, Writing - Review & 687–700. https://doi.org/10.1016/j.renene.2018.12.011.
Editing. Thokchom Subhaschandra Singh: Resources, Writing - Re­ Kuleshov, A., Mahkamov, Khamid, 2008. Multi-Zone diesel fuel spray combustion model
for the simulation of a diesel engine running on biofuel. Proc. Inst. Mech. Eng. Part A
view & Editing.
J. Power Energy 222 (3), 309–321. https://doi.org/10.1243/09576509JPE530.
Kuleshov, Andrey, Kozlov, A. v, Mahkamov, Khamid, 2010. Self-Ignition delay prediction
Declaration of Competing Interest in PCCI direct injection diesel engines using multi-zone spray combustion model and
detailed chemistry (2010-01-1960). SAE Tech. Pap. https://doi.org/10.4271/2010-
01-1960.
The authors declare that they have no known competing financial Labeckas, Gvidonas, Slavinskas, Stasys, 2006. The effect of rapeseed oil methyl ester on
interests or personal relationships that could have appeared to influence direct injection diesel engine performance and exhaust emissions. Energy Convers.
the work reported in this paper. Manag. 47 (13–14), 1954–1967. https://doi.org/10.1016/j.enconman.2005.09.003.
Lakshminarayanan, P.A., Aghav, Y.V., 2010. Particulate matter from direct injection
diesel engines. In: Modelling Diesel Combustion. In: Mechanical Engineering Series.
Data availability Springer, Dordrecht. 〈https://doi.org/10.1007/978-90-481-3885-2_14〉.
Lakshminarayanan, P.A., Yoghesh v Aghav, 2022. Modelling Diesel Combustion,
Springer Nature, Singapore, ISBN: 9811667411, 9789811667411.
Data will be made available on request. Lin, Cherng Yuan, Hsiu An Lin, 2006a. Diesel engine performance and emission
characteristics of biodiesel produced by the peroxidation process. Fuel 85 (3),
References 298–305. https://doi.org/10.1016/j.fuel.2005.05.018.
Martin, M., Leenus Jesu, V., Edwin, Geo, Nagalingam, B., 2017. Effect of fuel inlet
temperature on cottonseed oil–diesel mixture composition and performance in a di
Agarwal, Avinash Kumar, 2007. Biofuels (alcohols and biodiesel) applications as fuels for
diesel engine. J. Energy Inst. 90 (4), 563–573. https://doi.org/10.1016/j.
internal combustion engines. Prog. Energy Combust. Sci. 33 (3), 233–271. https://
joei.2016.05.005.
doi.org/10.1016/j.pecs.2006.08.003.
Mofijur, H.H., Masjuki, M.A., Kalam, A.E.Atabani, 2013b. Evaluation of biodiesel
Arunkumar, M., Kannan, M., Murali, G., 2019. Experimental studies on engine
blending, engine performance and emissions characteristics of Jatropha curcas
performance and emission characteristics using Castor biodiesel as fuel in CI engine.
methyl ester: Malaysian perspective. Energy 55, 879–887. https://doi.org/10.1016/
Renew. Energy 131, 737–744. https://doi.org/10.1016/j.renene.2018.07.096.
j.energy.2013.02.059.
Asokan, M.A., Vijayan, R., Senthur Prabu, S., Venkatesan, N., 2016. Experimental studies
Mofijur M, A.E., Atabani, H.H., Masjuki, M.A., Kalam, B.M.Masum, 2013a. A study on the
on the combustion characteristics and performance of a DI diesel engine. Int. J. Oil
effects of promising edible and non-edible biodiesel feedstocks on engine
Gas Coal Technol. 12, 105–119. https://doi.org/10.1504/IJOGCT.2016.075843.
performance and emissions production: a comparative evaluation. Renew. Sustain.
Behçet, Rasim, Yumrutaş, Recep, Oktay, Hasan, 2014. Effects of fuels produced from fish
Energy Rev. 23, 391–404. https://doi.org/10.1016/j.rser.2013.03.009.
and cooking oils on performance and emissions of a diesel engine. Energy 71,
Muralidharan, K., Vasudevan, D., 2011. Performance, emission and combustion
645–655. https://doi.org/10.1016/j.energy.2014.05.003.
characteristics of a variable compression ratio engine using methyl esters of waste
Bellappu, Venkat Jayanth, Victor Depoures, Melvin, Kaliyaperumal, Gopal,
cooking oil and diesel blends. Appl. Energy 88 (11), 3959–3968. https://doi.org/
Dillikannan, Damodharan, Jawahar, Dilipsingh, Palani, Kumaran, Meravanigee
10.1016/j.apenergy.2011.04.014.
Shivappa, Ganesha Prasad, 2021. A comprehensive study on the effects of multiple
Nanthagopal, K., Ashok, B., Saravanan, B., Korah, Shane Mathew, Chandra, Snehith,
injection strategies and exhaust gas recirculation on diesel engine characteristics that
2018. “Effect of next generation higher alcohols and Calophyllum inophyllum
utilize waste high density polyethylene oil. Energy Sources Part A Recovery, Util.
methyl ester blends in diesel engine. J. Clean. Prod. 180, 50–63. https://doi.org/
Environ. Eff. https://doi.org/10.1080/15567036.2021.1924313.
10.1016/j.jclepro.2018.01.167.
Bhowmik, S., Paul, A., Panua, S., Ghosh, S.K., Debroy, D., 2018. Performance-exhaust
Nautiyal, Piyushi, K. A. Subramanian, M.G. Dastidar, Kumar, Ashok, 2020. Experimental
emission prediction of diesosenol fueled diesel engine: an ANN coupled MORSM
assessment of performance, combustion and emissions of a compression ignition
based optimization. Energy 153, 212–222. https://doi.org/10.1016/j.
engine fuelled with spirulina platensis biodiesel. Energy 193. https://doi.org/
energy.2018.04.053.
10.1016/j.energy.2019.116861.
Cherng-Yuan, Lin, Hsiu-An, Lin, 2006. Diesel engine performance and emission
Perumal, Varatharaju, Ilangkumaran, M., 2017. Experimental analysis of engine
characteristics of biodiesel produced by the peroxidation process. Fuel 85 (3),
performance, combustion and emission using pongamia biodiesel as fuel in CI
298–305. https://doi.org/10.1016/j.fuel.2005.05.018.
engine. Energy 129, 228–236. https://doi.org/10.1016/j.energy.2017.04.120.
Datta, A., Mandal, B.K., 2017. Engine performance, combustion and emission
Rahman, M.M., Pourkhesalian, A.M., Jahirul, M.I., Stevanovic, S., Pham, P.X., Wang, H.,
characteristics of a compression ignition engine operating on different biodiesel-
Masri, A.R., Brown, R.J., Ristovski, Z.D., 2014. Particle emissions from biodiesels
alcohol blends. Energy 125, 470–483.
with different physical properties and chemical composition. Fuel 134, 201–208.
Dunga, Simhana Devi, Kumar, Ravinder, Rajak, Upendra, 2022. Experimental
https://doi.org/10.1016/j.fuel.2014.05.053.
investigation of performance, emission and combustion characteristics of a CI engine
Rajak, U., Verma, T.N., 2018b. Spirulina Microalgae Biodiesel – A novel renewable
fuelled by blends of waste plastic oil with diesel. Energy Sources, Part A Recovery
alternative energy source for compression ignition engine. J. Clean. Prod. 201,
Util. Environ. Eff. 44 (3), 7693–7708. https://doi.org/10.1080/
343–357. https://doi.org/10.1016/j.jclepro.2018.08.057.
15567036.2022.2115582.
Rajak, U., Prerana, N., Verma, T.N., Pugazhendhi, A., 2019b. Alternating the
Geng, L., Yanjuan, Wang, Yueying, Wang, Huimei, Li, 2020. Effect of the injection
environmental benefits of aegle-diesel blends used in compression ignition. Fuel 256.
pressure and orifice diameter on the spray characteristics of biodiesel. J. Traffic
https://doi.org/10.1016/j.fuel.2019.115835.
Transp. Eng. (Engl. Ed. ) 7 (3), 331–339. 〈https://doi.org/10.1016/j.jtte.2018
Rajak, Upendra, Nashine, Prerana, Singh, Thokchom Subhaschandra, Verma, Tikendra
.12.004〉.
Nath, 2018. Numerical investigation of performance, combustion and emission
Gnanasekaran, Sakthivel, Saravanan, N., Ilangkumaran, M., 2016. Influence of injection
characteristics of various biofuels. Energy Convers. Manag. 156, 235–252. https://
timing on performance, emission and combustion characteristics of a DI diesel
doi.org/10.1016/j.enconman.2017.11.017.
engine running on fish oil biodiesel. Energy 116, 1218–1229. https://doi.org/
Rajak, Upendra, Nashine, Prerana, Verma, Tikendra Nath, 2019. Assessment of diesel
10.1016/j.energy.2016.10.039.
engine performance using spirulina microalgae biodiesel. Energy 166, 1025–1036.
Goodrum, J.W., 2002. Volatility and boiling points of biodiesel from vegetable oils and
https://doi.org/10.1016/j.energy.2018.10.098.
tallow. Biomass Bioenergy 22 (3), 205–211. 〈https://doi.org/10.1016/S0961-9534
Rajak, Upendra, Prerana, Nashine, Verma, Tikendra Nath, Pugazhendhi, Arivalagan,
(01)00074-5〉.
2019a. Performance, combustion and emission analysis of microalgae spirulina in a
Imtenan, S.M., Varman, H.H., Masjuki, M.A., Kalam, H., Sajjad, M.I., Arbab, I.M.
Rizwanul Fattah, 2014. Impact of low temperature combustion attaining strategies

16
P. Gupta et al. Industrial Crops & Products 194 (2023) 116332

common rail direct injection diesel engine. Fuel 255, 115855. https://doi.org/ blends of two higher alcohols with lemongrass oil biodiesel and diesel fuel. Energy
10.1016/j.fuel.2019.115855. Environ. 33 (6), 1134–1159 https://doi: 10.1177/0958305×211051323.
Rajak, Upendra, Prerana, Nashine, Verma, Tikendra Nath, 2020. Numerical study on Shrivastava, Pankaj, Verma, Tikendra Nath, 2020a. Effect of fuel injection pressure on
emission characteristics of a diesel engine fuelled with diesel-spirulina microalgae- the characteristics of ci engine fuelled with biodiesel from Roselle oil. Fuel 265,
ethanol blends at various operating conditions. Fuel 262. https://doi.org/10.1016/j. 117005. https://doi.org/10.1016/j.fuel.2019.117005.
fuel.2019.116519. Shrivastava, Pankaj, Verma, Tikendra Nath, 2020b. Effect of fuel injection pressure on
Rajak, Upendra, Verma, Tikendra Nath, 2018a. Effect of emission from ethylic biodiesel the characteristics of CI engine fuelled with biodiesel from Roselle oil. Fuel 265,
of edible and non-edible vegetable oil, animal fats, waste oil and alcohol in ci engine. 117005. https://doi.org/10.1016/j.fuel.2019.117005.
Energy Convers. Manag. 166, 704–718. https://doi.org/10.1016/j. Silitonga, A.S., Masjuki, H.H., Mahlia, T.M.I., Hwai Chyuan, Ong, Chong, W.T., 2013.
enconman.2018.04.070. Experimental study on performance and exhaust emissions of a diesel engine fuelled
Rajasekaran, Shanmugam, Dillikannan, Damodharan, Kaliyaperumal, Gopal, with Ceiba pentandra biodiesel blends. Energy Convers. Manag. 76, 828–836.
Poures, Melvin Victor De, Babu, Rajesh Kumar, 2021. A comprehensive study on the https://doi.org/10.1016/j.enconman.2013.08.032.
effects of 1-decanol, compression ratio and exhaust gas recirculation on diesel engine Singh, T.S., Verma, T.N., Nashine, P., Shijagurumayum, C., 2018. BS-III diesel vehicles in
characteristics powered with low density polyethylene oil. Energy Sources, Part A: Imphal, India: An emission perspective. In: Sharma, N., Agarwal, A., Eastwood, P.,
Recovery, Util., Environ. Eff. 43 (23), 3064–3081. https://doi.org/10.1080/ Gupta, T., Singh, A. (Eds.), Air Pollution and Control, Energy, Environment, and
15567036.2020.1833112. Sustainability. Springer, Singapore. https://doi.org/10.1007/978-981-10-7185-0_5.
Rajesh, Adhinarayanan, Ramakrishnan, Aravindh, Kaliyaperumal, Gopal, De Sun, Xiuxiu, Xingyu, Liang, Gequn, Shu, Yuesen, Wang, Yong, Chen, 2019. Effect of
Poures, MelvinV.íctor, Babu, Rajesh Kumar, Dillikannan, Damodharan, 2020. toluene content on the combustion and emissions of large two-stroke marine diesel
Comparative analysis on the effect of 1-decanol and di-n-butyl ether as additive with engine. Appl. Therm. Eng. 159, 113909 https://doi.org/10.1016/j.
diesel/LDPE blends in compression ignition engine. Energy Sources, Part A: applthermaleng.2019.113909.
Recovery, Util., Environ. Eff. https://doi.org/10.1080/15567036.2020.1773967. Wei, L., Cheung, C.S., Ning, Z., 2017. Influence of waste cooking oil biodiesel on
Ravikumar, Jayabal, Subramani, Sekar, Dillikannan, Damodharan, combustion, unregulated gaseous emissions and particulate emissions of a direct-
Devarajan, Yuvarajan, Thangavelu, Lakshmanan, Nedunchezhiyan, Mukilarasan, injection diesel engine. Energy 127, 175–185. https://doi.org/10.1016/j.
Kaliyaperumal, Gopal, Poures, Melvin Victor De, 2022. Multi-objective optimization energy.2017.03.117.
of performance and emission characteristics of a CRDI diesel engine fueled with Zhao, Feiyang, Wenming Yang, Wenbin Yu, Han Li, Yu Yun Sim, Teng Liu, and Kun Lin
sapota methyl ester/diesel blends. Energy 250, 123709. https://doi.org/10.1016/j. Tay. (2018). Numerical study of soot particles from low temperature combustion of
energy.2022.123709. engine fueled with diesel fuel and unsaturation biodiesel fuels. Applied Energy. 211:
Seeniappan, K., Venkatesan, B., Krishnan, N.N., et al., 2022. A comparative assessment of 187–193. https://doi.org/10.1016/j.apenergy.2017.11.056.
performance and emission characteristics of a DI diesel engine fuelled with ternary

17

You might also like