Download as pdf or txt
Download as pdf or txt
You are on page 1of 8

Properties and use of Moringa oleifera biodiesel and diesel fuel blends

in a multi-cylinder diesel engine


M. Mojur
a,
, H.H. Masjuki
a
, M.A. Kalam
a
, A.E. Atabani
a,
, M.I. Arbab
a
, S.F. Cheng
b
, S.W. Gouk
b
a
Department of Mechanical Engineering, University of Malaya, 50603 Kuala Lumpur, Malaysia
b
Unit of Research on Lipids (URL), Department of Chemistry, Faculty of Science, University of Malaya, 50603 Kuala Lumpur, Malaysia
a r t i c l e i n f o
Article history:
Received 4 September 2013
Accepted 24 February 2014
Available online 28 March 2014
Keywords:
Biodiesel production
Characterization
Blending
Engine performance
Emissions
a b s t r a c t
Researchers have recently attempted to discover alternative energy sources that are accessible, techni-
cally viable, economically feasible, and environmentally acceptable. This study aims to evaluate the phys-
ico-chemical properties of Moringa oleifera biodiesel and its 10% and 20% by-volume blends (B10 and B20)
in comparison with diesel fuel (B0). The performance and emission of M. oleifera biodiesel and its blends
in a multi-cylinder diesel engine were determined at various speeds and full load conditions. The prop-
erties of M. oleifera biodiesel and its blends complied with ASTM D6751 standards. Over the entire range
of speeds, B10 and B20 fuels reduced brake power and increased brake specic fuel consumption com-
pared with B0. In engine emissions, B10 and B20 fuels reduced carbon monoxide emission by 10.60%
and 22.93% as well as hydrocarbon emission by 9.21% and 23.68%, but slightly increased nitric oxide
emission by 8.46% and 18.56%, respectively, compared with B0. Therefore, M. oleifera is a potential feed-
stock for biodiesel production, and its blends B10 and B20 can be used as diesel fuel substitutes.
2014 Elsevier Ltd. All rights reserved.
1. Introduction
The reserves of petroleum-derived fuels are diminishing with
their increasing demand every day. Moreover, the combustion
products that result from burning these fuels are considered harm-
ful to the environment. Several factors such as depletion of petro-
leum derived fuel, climate change, and increase in the price of
petroleum products have generated interest in discovering alterna-
tive energy sources among researchers [13]. In the last decades,
many researchers worldwide have searched for new alternative
energy sources that are available, technically feasible, economi-
cally viable, and environmentally acceptable [4]. Biodiesel is con-
sidered one of the best alternative energy sources because of its
potential to reduce dependency on fossil diesel fuel, capacity to de-
crease environmental pollutant output, and application in com-
pression ignition (CI) engines with no modication [5,6].
Biodiesel is nonexplosive, biodegradable, nonammable, renew-
able, nontoxic, environment friendly, and similar to diesel fuel
[7,8]. The main advantages of biodiesel include the following: it
can be blended with diesel fuel at any proportion; it can be used
in a CI engine with no modication; it does not contain any
harmful substances; and it produces less harmful emissions to
the environment than diesel fuel [9,10].
Biodiesel can be obtained through transesterication of vegeta-
ble oils, animal fats, waste cooking oil, and waste restaurant
greases [11]. It originates from edible and nonedible sources. The
most common edible oils of biodiesel include palm oil, rapeseed
oil, sunower oil, coconut oil, and peanut oil, whereas the noned-
ible oil sources of biodiesel are Jatropha, neem, cotton, jojoba, rub-
ber, Moringa, Mahua, castor, and animal tallow [12,13]. The
present study aims to evaluate the potential of biodiesel produc-
tion from Moringa oleifera oil as a promising feedstock that is easily
accessible worldwide. This study characterizes the physico-chem-
ical properties of M. oleifera biodiesel and its 10% and 20% by-vol-
ume blends. The properties that were investigated include
kinematic viscosity, density, ash point, cloud point, pour point,
and cold lter plugging point, viscosity index, and oxidation stabil-
ity. Then, the performance of the 10% and 20% by-volume blends of
M. oleifera biodiesel was assessed in a diesel engine. The relevant
fuel properties of M. oleifera biodiesel, such as engine performance
and emission characteristics, were fully investigated and compared
with those of diesel fuel.
2. Literature review
M. oleifera is a member of the Moringaceae family, which
mainly grows in tropical countries [14]. This drought-tolerant
http://dx.doi.org/10.1016/j.enconman.2014.02.073
0196-8904/ 2014 Elsevier Ltd. All rights reserved.

Corresponding authors. Tel./fax: +60 03 79674448 (M. Mojur). Tel.: +60


122314659 (A.E. Atabani).
E-mail addresses: mojduetme@yahoo.com (M. Mojur), a_atabani2@msn.com
(A.E. Atabani).
Energy Conversion and Management 82 (2014) 169176
Contents lists available at ScienceDirect
Energy Conversion and Management
j our nal homepage: www. el sevi er. com/ l ocat e/ enconman
pioneer species is locally known in Malaysia as kachang kelur, but
commercially it called ben oil or behen oil. M. oleifera contains
behenic (docosanoic) acid and signicantly prevents oxidative deg-
radation. M. oleifera can be used for medicinal and clinical pur-
poses, and contains a substantial amount of nutrition value. The
species is locally distributed in northwest India, Southeast Asia,
Africa, South America, and Arabia [15,16]. However, it is also cur-
rently available in the Central America, Philippines, North America,
Malaysia, and Cambodia. M. oleifera grows fast, can withstand a
wide range of rainfall (25 cm to 300+ cm per year), and sustain life
in poor soil (pH 59) [17,18]. The height of M. oleifera tree can
range from 5 m to 10 m [19]. The seeds of M. oleifera are triangular
and contain approximately 40% of oil by weight. The oil produced
from the seed kernel of M. oleifera is golden yellow. M. oleifera oil
reportedly contains elevated amounts of oleic acid, which com-
prises approximately 74.41% of its entire fatty acid prole [20].
Recent studies [21,22] have investigated the potential of biodie-
sel production from edible oil and nonedible oil sources, and their
utilization in a diesel engine. Only a few studies [14,16,2326]
have reported on the potential of biodiesel production from M.
oleifera, a nonedible oil source, and evaluated the blends of M. oleif-
era in a multi-cylinder diesel engine. Only Rajaraman et al. [27]
have reported on the performance and emission characteristics of
Moringa oil methyl ester and its blends (B20 to B100) in a direct
injection diesel engine at various load conditions. They reported
that M. oleifera blend exhibits lower brake thermal efciency
(BTE) than diesel fuel because of the formers lower heating value
and higher viscosity and density than the latter. In engine emis-
sions, M. oleifera blend produces lower HC, CO and PM emission
but NO
x
emission than diesel fuel. The properties of M. oleifera bio-
diesel and its blends meet the ASTM D6751 specications, which
are the standard specication for biodiesel (B100) fuels and indi-
cate the products suitability to be used in diesel engines.
3. Materials and methods
3.1. Materials
Crude M. oleifera oil (CMOO) was collected from University
Sains Malaysia. All other chemicals, reagents, and accessories were
purchased from LGC Scientic Sdn Bhd (Malaysia). The experimen-
tal investigation was performed using diesel fuel (B0), B10 (90%
diesel and 10% M. oleifera biodiesel), and B20 (20% M. oleifera bio-
diesel and 80% diesel).
3.2. Equipment list
Table 1 highlights the equipment used to measure the physico-
chemical properties of M. oleifera biodiesel and its blends.
3.3. Biodiesel production
The high acid value of CMOO causes a problem during the sep-
aration process. Therefore, a two-step process (acidbase catalyst)
was suggested to convert M. oleifera oil into biodiesel (methyl es-
ter). The production of biodiesel was conducted at the Energy
Lab of University Malaya using a 1 L batch reactor with a reux
condenser, a magnetic stirrer, a thermometer, and a sampling out-
let. The summary of the biodiesel production process is given in
Table 2. Furthermore, a comprehensive view of the biodiesel pro-
duction processes is furnished in the following section.
3.3.1. Acid-catalyzed process
For biodiesel production, an acid-catalyzed process was used
before transesterication to reduce the high acid value of crude
oils. In this process, a molar ratio of 12:1 methanol to CMOO and
1% (v/v oil) of sulfuric acid (H
2
SO
4
) were added to the preheated
oil at 60 C for 3 h at 600 rpm stirring speed. After the reaction,
the product was transferred to a separating funnel to separate
the esteried oil (lower layer) from the upper layer, which includes
excess alcohol, sulfuric acid, and impurities. The lower layer was
then loaded into a control rotary evaporator (IKA) and heated at
60 C under vacuum conditions for 1 h to remove methanol and
water from the esteried oil. After esterication, the acid value
was reduced to less than 4.
3.3.2. Alkaline-catalyzed process
In the alkaline-catalyzed process, a molar ratio of 6:1 of meth-
anol and 1% (w/w oil) of potassium hydroxide (KOH) were added
to the preheated esteried M. oleifera oil at 60 C for 2 h at
600 rpm stirring speed. After the reaction, the produced methyl es-
ter was deposited in a separation funnel for 16 h to separate glyc-
erol from methyl ester. The lower layer, which contains glycerol
and impurities, was drained.
3.3.3. Post-treatment process
The methyl ester was washed with warm distilled water to re-
move the impurities and glycerol. In this process, 50% (v/v oil) of
distilled water at 60 C was sprayed over the surface of the ester
and stirred gently. This washing process was repeated several
times until the pH of the biodiesel became neutral. The lower layer
was discarded, and the upper layer was poured into a control ro-
tary evaporator (IKA) to remove water and excess methanol from
methyl ester. The methyl ester was poured into a ask, dried using
anhydrous sodium sulfate (Na
2
SO
4
), and then further dried using
the control rotary evaporator. Finally, the produced biodiesel was
ltered using a qualitative lter paper (150 mm diameter, No. 1)
to obtain the nal product. The percentage of yield of the produced
biodiesel was more than 90%.
Nomenclature
ASTM American Society for Testing and Materials
BP brake power
B0 diesel fuel
B10 10% biodiesel + 90% diesel
B20 20% biodiesel + 80% diesel
BSFC brake specic fuel consumption
BTE brake thermal efciency
CMOO crude Moringa oleifera oil
CO carbon monoxide
CO
2
carbon dioxide
FAC fatty acid composition
FT-IR Fourier transform-Infra red
HC hydrocarbon
NO nitric oxide
NO
x
oxides of nitrogen
PM particulate matter
170 M. Mojur et al. / Energy Conversion and Management 82 (2014) 169176
3.4. Fatty acid composition (FAC)
The FAC of M. oleifera biodiesel was analyzed using gas chroma-
tography (Shidmadzu, Japan) equipped with a ame ionization
detector. Moreover, the results of FAC of M. oleifera biodiesel are
shown in Table 3.
3.5. Properties analysis
The physico-chemical properties of the produced biodiesel were
characterized according to the ASTM D6751 standard. Cetane num-
ber (CN), iodine value (IV), and saponication value (SV) were
determined using the following equations:
CN 46:3 5458=SV 0:225 IV 1
SV R560 A
i
=M
Wi
2
IV R254 A
i
D=M
Wi
3
where A
i
is the percentage of each component, D is the number of
double bond, and MW
i
is the molecular mass of each component.
M. oleifera biodiesel was also characterized by FT-IR using a Per-
kin Elmer biodiesel FAME analyzer equipped with an MIR TGS
detector in the range of 4000400 cm
1
and processed with the
computer software program spectrum. The resolution was 4 cm
1
and 8 scans.
3.6. Blending of biodiesel
The test fuel was blended with diesel using a homogenizer de-
vice at a speed of 2000 rpm. The homogenizer was xed with a
clamp on a vertical stand, which allows for changing the homoge-
nizers height. To mix the fuels by using the homogenizer, the plug
was turned on and the appropriate speed was selected by using the
selector located at the top of the homogenizer.
3.7. Engine tests
The test engine used was a Mitsubishi Pajero (model 4D56T)
multi-cylinder diesel engine. The test rig of the engine is shown
in Fig. 1. The detailed specications of the engine are shown in Ta-
ble 4. Before running the engine with the biodiesel-blended fuels,
the engine was rst operated with diesel fuel for a few minutes
to warm up the engine. Diesel fuel was also used before engine
shutdown. The same procedure was maintained for each fuel. To
perform engine performance and emission tests, the engine was
run at various speeds (10004000 rpm) at full load conditions.
The engine test conditions were monitored by a REO-DCA control-
ler connected through a desktop to the engine test bed (Fig. 1). A
BOSCH exhaust gas analyzer (model BEA-350) was used to mea-
Table 1
Summary of the equipment used to measure the properties of M. oleifera biodiesel and its blend.
Property Equipment Test method Accuracy
Kinematic viscosity SVM 3000 (Anton Paar, UK) ASTM D445 0.1%
Flash point Pensky-martens ash point automatic NPM 440 (Normalab, France) ASTM D93 0.1 C
Oxidation stability 873 Rancimat (Metrohm, Switzerland) EN ISO 14112 0.01 h
Cloud and pour point Cloud and pour point tester automatic NTE 450 (Normalab, France) ASTM D2500 0.1 C
ASTM D97
Cold lter plugging point Cold lter plugging point tester automatic NTL 450 (Normalab, France) ASTM D6371 0.1 C
Density DM 40 (Mettler Toledo, USA) ASTM D1298 0.1 kg/m
3
Caloric value C2000 basic calorimeter (IKA, UK) ASTM D240 0.001 MJ/kg
Acid value Automation titration rondo 20 (Mettler Toledo, Switzerland) ASTM D664 0.001 mgKOH/g
Table 2
Summary of biodiesel production process from M. oleifera oil.
S/n Process parameter Process specication
01 Process selected Acidbase catalyst process
02 Reaction temperature 60 C
03 Catalyst used 98% Pure sulfuric acid (1% v/v) and 99%
pure potassium hydroxides (1% w/w)
04 Alcohol used Methanol
05 Molar ratio 12:1 For esterication and 6:1 for
transesterication
06 Reaction time 3 h For esterication and 2 h
for transesterication
07 Setting time 15 h
08 Stirring speed 600 rpm
Table 3
Fatty acid composition of M. oleifera biodiesel.
Sl. no. Fatty acid (as methyl ester) Molecular weight Shorthand designation Systematic name Formula MOME MOME
a
1 Lauric 200 12:0 Dodecanoic C
12
H
24
O
2
0 0
2 Myristic 228 14:0 Tetradecanoic C
14
H
28
O
2
0.1 0
3 Palmitic 256 16:0 Hexadecanoic C
16
H
32
O
2
7.9 6.5
4 Palmitoleic 254 16.1 Hexadec-9-enoic C
16
H
30
O
2
1.7 0
5 Stearic 284 18:0 Octadecanoic C
18
H
36
O
2
5.5 6.0
6 Oleic 282 18:1 Cis-9-Octadecenoic C
18
H
34
O
2
74.1 72.2
7 Linoleic 280 18:2 Cis-9-cis-12 Octadecadienoic C
18
H
32
O
2
4.1 1.0
8 Linolenic 278 18:3 Cis-9-cis-12 C
18
H
30
O
2
0.2 0
9 Arachidic 312 20:0 Eicosanoic C
20
H
40
O
2
2.3 4.0
10 Eicosanoic 310 20:1 Cis-11-eicosenoic C
20
H
38
O
2
1.3 2.0
11 Behenic 340 22:0 Docosanoic C
22
H
44
O
2
2.8 7.1
12 Other 0 1
Saturated 18.6 23.6
Monounsaturated 77.1 74.2
Polyunsaturated 4.3 1
Total 100 99.8
a
Data obtained from Ref. [16].
M. Mojur et al. / Energy Conversion and Management 82 (2014) 169176 171
sure the NO, HC, and CO contents of the exhaust emission gases.
The details of gas analyser are shown in Table 5. All tests were rep-
licated three times, and the average was obtained.
3.8. Error analysis
Errors and uncertainties in the experiments can arise from
instrument selection, condition, calibration, environment, observa-
tion, reading, and test planning. Uncertainty analysis was required
to prove the accuracy of the experiments. The accuracy of the
speed, fuel measurement, brake power (BP), and time was
10 rpm, 1% of the reading, 0.07 kW, and 0.1 s respectively.
The relative uncertainty of brake-specic fuel consumption (BSFC)
was determined using the linearized approximation method of
uncertainty. Table 6 shows the summary of the values of measure-
ment accuracy and the relative uncertainty of various parameters,
including BP, BSFC, CO, HC, and NO emission.
4. Results and discussion
4.1. Characterization of M. oleifera biodiesel and blends
To characterize pure M. oleifera biodiesel (B100), several proper-
ties such as density, ash point, viscosity, viscosity index, caloric
value, cloud and pour points, and oxidation stability were exam-
ined and compared based on ASTM D6751 standards. Table 7
shows the detailed physico-chemical characteristics of M. oleifera
biodiesel (B100) and its blends (B10 and B20). All the physico-
chemical properties of M. oleifera biodiesel were found to meet
the ASTM D6751 standards. Thus, M. oleifera biodiesel can be used
in a diesel engine as diesel fuel substitutes. Moreover, Table 7
emphasizes that one of the main features of M. oleifera biodiesel
is its high oxidation stability (26.2 h). This result agrees with
[16], which indicated that M. oleifera possesses signicant resis-
tance to oxidative degradation.
4.2. Engine performance
In this study, engine performance in terms of BP and BSFC was
evaluated. The inuence of M. oleifera biodiesel on engine perfor-
mance depends on the relationship between the fuel injection sys-
tem and the fuel properties, oxygenation nature of the biodiesel,
higher viscosity, and lower caloric value. These effects have a ma-
jor inuence on spray formation and combustion. The following
section discusses the obtained results of these parameters. Table 8
shows the results of statistical analysis for the test fuel at full load
conditions.
4.2.1. BP
Fig. 2 shows the trend line of the engine BP output of M. oleifera
biodiesel at different engine speeds. The BP increased steadily as
the engine speed increased up to 3500 rpm and then decreased be-
cause of high frictional force. Over the entire range of speed, the
average BP values were 28.72, 27.51, and 26.41 kW for fuel sam-
ples B0, B10, and B20, respectively. The biodiesel blend fuels pro-
duced lower BP than diesel fuel. In addition, biodiesel blend fuels
B10 and B20 decreased the BP by 4.22% and 8.03% compared with
diesel fuel. The decrease in BP can be attributed to the biodiesel
blends lower caloric values and higher viscosities than diesel fuel
(Table 7). These properties of the biodiesel blends inuenced the
combustion system. The prolonged ignition delays due to the
blends resulted in incomplete combustion because of their higher
viscosity compared with diesel fuel. Uneven combustion character-
istic of biodiesel fuel reduces the engine BP [28,29].
4.2.1. BSFC
The different BSFC values of diesel and biodiesel blend fuels
with their trend line are shown in Fig. 3. The BSFC values were
higher when biodiesel blend fuel was used. This result is supported
by previous ndings [3032]. Compared with diesel fuel, the BSFC
slightly increased with increasing biodiesel blend ratio. The BSFC of
diesel engine depends on the relationship among volumetric fuel
injection system, fuel density, viscosity, and lower heating value
[33]. Over the entire range of speed, the average BP values were
386, 406 and 418 g/kW h for fuel samples B0, B10, and B20, respec-
tively. The biodiesel blend fuels consumed a higher amount of fuel
to produce a unit kW of power than diesel fuel. In addition, biodie-
sel blend fuel samples B10 and B20 increased the BSFC by 5.13%
and 8.39%, respectively, compared with diesel fuel. The higher BSFC
of biodiesels can be attributed to the effects of the lower heating
value (Table 7) of the blends [28]. Biodiesel fuel is delivered into
the engine on a volumetric basis per stroke; thus, larger quantities
of biodiesel are fed into the engine.
Fig. 1. Engine test bed set-up.
172 M. Mojur et al. / Energy Conversion and Management 82 (2014) 169176
Therefore, to produce the same power, more biodiesel fuel is
needed because biodiesel has a lower caloric value compared
with diesel fuel [34].
4.2.2. BTE
The trend line of engine BTE of M. oleifera biodiesel at different
engine speeds is shown in Fig. 4. With increasing engine speed, the
BTE of the Moringa biodiesel blends increased together with diesel.
However, the brake thermal efciency of the Moringa biodiesel
blends was lower than that of diesel fuel throughout the entire
range. The possible reasons for this reduction are the lower
caloric value and increased fuel consumption of Moringa biodie-
sel blends as compared with diesel fuel.
4.3. Emission analysis
4.3.1. CO emission
CO is produced though incomplete combustion of fuels without
any oxygen molecules, such as petroleum fuels. Several factors
such as airfuel ratio, engine speed, injection timing, injection
pressure, and type of fuels affect CO emission [35]. The different
CO emission values of diesel and biodiesel blend fuel with their
trend line are shown in Fig. 5. Over the entire speed range, fuel
samples B10 and B20 reduced CO emission by 10.60% and 22.93%
compared with B0 fuel, respectively. This result agrees with previ-
ous reports [27,3638]. This result can be attributed to the higher
oxygen content and higher cetane number of the biodiesel fuel
than diesel fuel. Biodiesel fuel contains 12% higher oxygen than
diesel fuel [14]. As the percentage of biodiesel increased in the
blend, the higher oxygen content in biodiesel allows more carbon
molecules to burn and facilitates the completion of combustion.
Thus, CO emission can be reduced by using the biodiesel blends
in a diesel engine.
4.3.2. HC emission
Unburned HC results from the incomplete combustion of fuel
and ame quenching [28]. The different HC emission values of die-
sel and biodiesel blend fuels with their trend line are shown in
Fig. 6. The unburned HC emissions for B10 and B20 were lower
than that for diesel fuel. Over the entire range of speed, B10 and
B20 reduced HC emission by 9.21% and 23.68%, respectively. In
addition, HC emission reduced as the percentage of biodiesel frac-
tion in the blends increased. This result can be attributed to the
high oxygen contents of biodiesel fuel. Biodiesel contains higher
oxygen and lower carbon and hydrogen than diesel fuel, all of
Table 4
Details specication of the engine.
Engine model 4D56T (Mitsubishi Pajero)
Engine type 4 Cylinder inline
Displacement L 2.5
Cylinder bore stroke mm
2
91.1 95
Compression ratio 21:1
Maximum engine speed rpm 4200
Maximum power kW 78
Fuel system Distribution type jet pump (indirect injection)
Lubrication system Pressure feed
Combustion chamber Swirl type
Cooling system Radiator cooling
Table 5
Details of the exhaust gas analyzer.
Equipment Method Measurement Upper
limit
Accuracy
BOSCH gas
analyser
Non-dispersive
infrared
CO 10.00 vol% 0.001 vol%
Non-dispersive
infrared
CO
2
18.00 vol% 0.01 vol%
Flame ionization
detector (FID)
HC 9999 ppm l ppm
Electro-chemical
transmitter
NO 5000 ppm 1 ppm
Table 6
Summary of measurements uncertainty.
Measurements Accuracy Relative uncertainty
BP 0.07 kW 0.243
BSFC 5 g/kW h 0.013
CO 0.001 vol% 0.003
NO 1 ppm 0.005
HC 1 ppm 0.090
CO
2
0.01 vol% 0.001
Table 7
Physico-chemical characteristics of M. oleifera biodiesel and its blends.
Properties Units Standards B0 B10 B20 B100 MOME
a
ASTM D6751
Kinematic viscosity at 40 C mm
2
/s ASTM D445 3.23 3.54 3.67 5.05 4.83 1.96
Viscosity index 90 101.1 111.6 184.6
Density kg/m
3
ASTM D1298 827.2 830.6 833.6 859.6
Flash point C ASTM D93 68.5 79.5 82.5 150.5 130 min
Cloud point C ASTM D2500 8 7 8 19 18
Pour point C ASTM D97 0 3 6 19 17
Cold lter plugging point C ASTM D6371 5 6 6 18
Caloric value MJ/kg ASTM D240 45.30 44.74 43.30 40.05
Iodine value g I/100 g EN 14111 77.5
Saponication value 199
Oxidation stability h EN ISO 14112 26.2 3.61 3
Cetane number ASTM D613 48 56.30 67.07 47 min
Acid value mg KOH/g ASTM D664 0.22 0.80
Total glycerol ASTM D6584 0.11 0.24% Mass
a
Data collected from Ref. [16].
M. Mojur et al. / Energy Conversion and Management 82 (2014) 169176 173
which trigger an improved and complete combustion process
[39,40].
4.3.3. NO emission
The different NO emission values of diesel and biodiesel blend
fuel with their trend line are shown in Fig. 7. The NO values were
higher when biodiesel blend fuel was used. Similar ndings were
reported in previous studies [41]. B10 and B20 emitted 8.46% and
18.56% higher NO than diesel fuel over the entire speed range. This
result can be attributed to the lean air/fuel ratio. Biodiesel is an
oxygenated fuel and contains 12% more oxygen in its molecular
structure, which causes higher chamber temperature by improving
combustion at its warmed-up condition [42]. Thus, NO emission is
increased with the use of the biodiesel blend compared with diesel
fuel. Moreover, increased NO can be explained in terms of adiabatic
ame temperature. Biodiesel fuel contains higher percentages of
unsaturated fatty acids that have higher adiabatic ame tempera-
ture, which causes higher NO emission [41].
Table 8
Statistical analysis of the test fuel at full load.
BP (Fig. 2) BSFC (Fig. 3) BTE (Fig. 4) CO (Fig. 5) HC (Fig. 6) NO (Fig. 7)
B0 B10 B20 B0 B10 B20 B0 B10 B20 B0 B10 B20 B0 B10 B20 B0 B10 B20
Mean 28.72 27.51 26.42 386 405.51 418.09 20.93 20.16 19.72 0.37 0.33 0.29 10.86 9.86 8.29 210.86 228.71 250
Median 32.49 31.45 29.96 380 406 401 20.91 20.93 20.47 0.31 0.26 0.23 10 9 7 229 250 264
Std 8.82 8.3 7.99 54.56 53.27 55.83 2.77 2.64 2.53 0.16 0.16 0.17 3.72 3.18 3.45 50.06 49.15 52.64
8
13
18
23
28
33
38
500 1000 1500 2000 2500 3000 3500 4000
B
r
a
k
e

p
o
w
e
r

[
K
W
]
Engine speed [rpm]
B0
B20
B10
Fig. 2. Variation of brake power with respect to the speed at full load condition.
300
350
400
450
500
550
600
500 1000 1500 2000 2500 3000 3500 4000
B
S
F
C

[
g
/
K
w
.
h
]
Engine speed [rpm]
B10
B20
B0
Fig. 3. Variation of brake specic fuel consumption with respect to the speed at full
load condition.
5
10
15
20
25
500 1000 1500 2000 2500 3000 3500 4000
B
T
E

(
%
)
Engine speed [rpm]
B0
B10
B20
Fig. 4. Variation of brake thermal efciency with respect to the speed at full load
condition.
0
0.1
0.2
0.3
0.4
0.5
0.6
0.7
0.8
500 1000 1500 2000 2500 3000 3500 4000
C
O

[
v
o
l

%
]
Engine speed [rpm]
B10
B20
B0
Fig. 5. Variation of CO emissions with respect to the speed at full load condition.
0
2
4
6
8
10
12
14
16
18
20
500 1000 1500 2000 2500 3000 3500 4000
H
C

[
p
p
m
]
Engine speed [rpm]
B10
B20
B0
Fig. 6. Variation of HC emissions with respect to the speed at full load condition.
174 M. Mojur et al. / Energy Conversion and Management 82 (2014) 169176
5. Conclusions
Biodiesel is one of the best alternative sources of fuel because of
its potential to reduce dependency on fossil diesel fuel, capacity to
decrease environmental pollutant output, and application in CI en-
gines without further modication. This study aims to produce bio-
diesel from CMOO and to evaluate its 10% and 20% by-volume
blends in a diesel engine. The following conclusions were drawn
from the experiments. After esterication process acid value re-
duced signicantly and after transesterication process the per-
centage of yield was found more than 90%. The properties of M.
oleifera biodiesel and its blends conform to ASTM D6751 standards.
Over the entire speed range, B10 and B20 yielded an average
reduction of BP compared with diesel fuel. Meanwhile, the average
BSFC of B10 and B20 was slightly higher than that of diesel fuel.
B10 and B20 reduced CO emission and slightly increased NO emis-
sion than diesel fuel. Therefore, M. oleifera is a potential feedstock
for biodiesel production, and B10 and B20 blends can be used as di-
rect diesel fuel substitutes.
Acknowledgment
The authors would like to acknowledge University of Malaya for
nancial support through High Impact Research Grant UM.C/HIR/
MOHE/ENG/07.
References
[1] Palash SM, Kalam MA, Masjuki HH, Masum BM, Rizwanul Fattah IM, Mojur M.
Impacts of biodiesel combustion on NO
x
emissions and their reduction
approaches. Renew Sust Energy Rev 2013;23:47390.
[2] Mojur M, Atabani AE, Masjuki HH, Kalam MA, Masum BM. A study on the
effects of promising edible and non-edible biodiesel feedstocks on engine
performance and emissions production: a comparative evaluation. Renew Sust
Energy Rev 2013;23:391404.
[3] Shahabuddin M, Liaquat AM, Masjuki HH, Kalam MA, Mojur M. Ignition delay,
combustion and emission characteristics of diesel engine fueled with biodiesel.
Renew Sust Energy Rev 2013;21:62332.
[4] Liaquat AM, Kalam MA, Masjuki HH, Jayed MH. Potential emissions reduction
in road transport sector using biofuel in developing countries. Atmos Environ
2010;44:386977.
[5] Tan P-q, Hu Z-y, Lou D-m, Li Z-j. Exhaust emissions from a light-duty diesel
engine with Jatropha biodiesel fuel. Energy 2012;39:35662.
[6] Tesfa B, Mishra R, Zhang C, Gu F, Ball AD. Combustion and performance
characteristics of CI (compression ignition) engine running with biodiesel.
Energy 2013;51:10115.
[7] Lee HV, Tauq-Yap YH, Hussein MZ, Yunus R. Transesterication of jatropha oil
with methanol over MgZn mixed metal oxide catalysts. Energy
2013;49:128.
[8] Mojur M, Masjuki HH, Kalam MA, Atabani AE, Shahabuddin M, Palash SM,
et al. Effect of biodiesel from various feedstocks on combustion characteristics,
engine durability and materials compatibility: a review. Renew Sust Energy
Rev 2013;28(2013):44155.
[9] Shahabuddin M, Kalam MA, Masjuki HH, Bhuiya MMK, Mojur M. An
experimental investigation into biodiesel stability by means of oxidation and
property determination. Energy 2012;44:61622.
[10] Ushakov S, Valland H, sy V. Combustion and emissions characteristics of
sh oil fuel in a heavy-duty diesel engine. Energy Convers Manage
2013;65:22838.
[11] Costa JF, Almeida MF, Alvim-Ferraz MCM, Dias JM. Biodiesel production using
oil from sh canning industry wastes. Energy Convers Manage 2013;74:1723.
[12] Atabani AE, Silitonga AS, Badruddina Irfan Anjum, Mahlia TMI, Masjuki HH,
Mekhilef S. A comprehensive review on biodiesel as an alternative energy
resource and its characteristics. Renew Sust Energy Rev 2012;16:207093.
[13] Atabani AE, Mahlia TMI, Anjum Badruddin I, Masjuki HH, Chong WT, Lee KT.
Investigation of physical and chemical properties of potential edible and non-
edible feedstocks for biodiesel production, a comparative analysis. Renew Sust
Energy Rev 2013;21:74955.
[14] Santana CR, Pereira DF, de Araujo NA, Cavalcanti EB, da Silva GF. Physical
chemical characterization of the Moringa Lam Revista. Brasileira de Produtos
Agroindustriais, Campina Grande 2010;12:5560.
[15] Atabani AE, Silitonga AS, Ong HC, Mahlia TMI, Masjuki HH, Badruddin IA, et al.
Non-edible vegetable oils: a critical evaluation of oil extraction, fatty acid
compositions, biodiesel production, characteristics, engine performance and
emissions production. Renew Sust Energy Rev 2013;18:21145.
[16] Rashid U, Anwar F, Moser BR, Knothe G. Moringa oleifera oil: a possible source
of biodiesel. Bioresour Technol 2008;99:81759.
[17] Orwa. Agroforestry Database 4.0. [cited 2013 14th April]; Available from
<http://wwwworldagroforestryorg/treedb2/AFTPDFS/Moringa_oleiferapdf>.
[18] Borugadda VB, Goud VV. Biodiesel production from renewable feedstocks:
status and opportunities. Renew Sust Energy Rev 2012;16:476384.
[19] Kibazohi O, Sangwan RS. Vegetable oil production potential from Jatropha
curcas, Croton megalocarpus, Aleurites moluccana, Moringa oleifera and
Pachira glabra: assessment of renewable energy resources for bio-energy
production in Africa. Biomass Bioenergy 2011;35:13526.
[20] Rahman IMM, Barua S, Nazimuddin M, Begum ZA, Rahman MA, Hasegawa H.
Physicochemical properties of Moringa Oleifera Lam. seed oil of the
Indigenous-cultivar of Bangladesh. J Food Lipids 2009;16:54053.
[21] Ndayishimiye P, Tazerout M. Use of palm oil-based biofuel in the internal
combustion engines: performance and emissions characteristics. Energy
2011;36:17906.
[22] Wander PR, Altani CR, Colombo AL, Perera SC. Durability studies of mono-
cylinder compression ignition engines operating with diesel, soy and castor oil
methyl esters. Energy 2011;36:391723.
[23] da Silva JPV, Serra TM, Gossmann M, Wolf CR, Meneghetti MR, Meneghetti
SMP. Moringa oleifera oil: studies of characterization and biodiesel production.
Biomass Bioenergy 2010;34:152730.
[24] Kafuku G, Lam MK, Kansedo J, Lee KT, Mbarawa M. Heterogeneous catalyzed
biodiesel production from Moringa oleifera oil. Fuel Process Technol
2010;91:15259.
[25] Kafuku G, Mbarawa M. Alkaline catalyzed biodiesel production from Moringa
oleifera oil with optimized production parameters. Appl Energy
2010;87:25615.
[26] Rashid U, Anwar F, Ashraf M, Saleem M, Yusup S. Application of response
surface methodology for optimizing transesterication of Moringa oleifera oil:
biodiesel production. Energy Convers Manage 2011;52:303442.
[27] Rajaraman S, Yashwanth GK, Rajan T, Kumaran RS, Raghu P. Experimental
investigations of performance and emission characteristics of Moringa oil
methyl ester and its diesel blends in a single cylinder direct injection diesel
engine. In: Proceedings of the ASME 2009 international mechanical
engineering congress & exposition, November 1319, Lake Buena Vista,
Florida, USA; 2009.
[28] Mojur M, Masjuki HH, Kalam MA, Atabani AE. Evaluation of biodiesel
blending, engine performance and emissions characteristics of Jatropha curcas
methyl ester: Malaysian perspective. Energy 2013;55:87987.
[29] Muralidharan K, Vasudevan D, Sheeba KN. Performance, emission and
combustion characteristics of biodiesel fuelled variable compression ratio
engine. Energy 2011;36:538593.
[30] Wang X, Ge Y, Yu L, Feng X. Comparison of combustion characteristics and
brake thermal efciency of a heavy-duty diesel engine fueled with diesel and
biodiesel at high altitude. Fuel 2013;107:8528.
[31] Chauhan BS, Kumar N, Cho HM. A study on the performance and emission of a
diesel engine fueled with Jatropha biodiesel oil and its blends. Energy
2012;37:61622.
[32] Shahabuddin M, Masjuki HH, Kalam MA, Mojur M, Hazrat MA, Liaquat AM.
Effect of additive on performance of C.I. engine fuelled with bio diesel. Energy
Proc 2012;14:16249.
[33] Qi DH, Chen H, Geng LM, Bian YZ. Experimental studies on the combustion
characteristics and performance of a direct injection engine fueled with
biodiesel/diesel blends. Energy Convers Manage 2010;51:298592.
[34] Tsolakis A, Megaritis A, Wyszynski ML, Theinnoi K. Engine performance and
emissions of a diesel engine operating on diesel-RME (rapeseed methyl ester)
blends with EGR (exhaust gas recirculation). Energy 2007;32:207280.
[35] Gumus M, Sayin C, Canakci M. The impact of fuel injection pressure on the
exhaust emissions of a direct injection diesel engine fueled with biodiesel
diesel fuel blends. Fuel 2012;95:48694.
[36] Hirkude JB, Padalkar AS. Performance and emission analysis of a compression
ignition: engine operated on waste fried oil methyl esters. Appl Energy
2012;90:6872.
100
125
150
175
200
225
250
275
300
500 1000 1500 2000 2500 3000 3500 4000
N
O

[

p
p
m
]
Engine speed [rpm]
B10
B20
B0
Fig. 7. Variation of NO emissions with respect to the speed at full load condition.
M. Mojur et al. / Energy Conversion and Management 82 (2014) 169176 175
[37] Lapuerta M, Armas O, Rodrguez-Fernndez J. Effect of biodiesel fuels on diesel
engine emissions. Prog Energy Combust Sci 2008;34:198223.
[38] Kim H, Choi B. The effect of biodiesel and bioethanol blended diesel fuel on
nanoparticles and exhaust emissions from CRDI diesel engine. Renew Energy
2010;35:15763.
[39] Lin B-F, Huang J-H, Huang D-Y. Experimental study of the effects of vegetable
oil methyl ester on DI diesel engine performance characteristics and pollutant
emissions. Fuel 2009;88:177985.
[40] Qi DH, Chen H, Geng LM, Bian YZ, Ren XC. Performance and combustion
characteristics of biodieseldieselmethanol blend fuelled engine. Appl
Energy 2010;87:167986.
[41] El-Kasaby M, Nemit-allah MA. Experimental investigations of ignition delay
period and performance of a diesel engine operated with Jatropha oil biodiesel.
Alexandria Eng J 2013;52:1419.
[42] Devan PK, Mahalakshmi NV. A study of the performance, emission and
combustion characteristics of a compression ignition engine using methyl
ester of paradise oileucalyptus oil blends. Appl Energy 2009;86:67580.
176 M. Mojur et al. / Energy Conversion and Management 82 (2014) 169176

You might also like