You are on page 1of 10

Energy Conversion and Management 84 (2014) 217226

Contents lists available at ScienceDirect

Energy Conversion and Management


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

Simultaneous reduction of NOx and smoke in a dual fuel DI diesel engine


Debabrata Barik , S. Murugan
Internal Combustion Engines Laboratory, Department of Mechanical Engineering, National Institute of Technology, Rourkela 769008, India

a r t i c l e

i n f o

Article history:
Received 14 December 2013
Accepted 13 April 2014
Available online 4 May 2014
Keywords:
Biodiesel
Biogas
Diesel engine
Dual fuel
Energy share

a b s t r a c t
This paper presents the results of an experimental investigation conducted on a compression ignition (CI)
engine, modied to run on dual fuel mode, using biogas as a primary fuel and KME (Karanja methyl ester)
as a pilot fuel. The biogas was produced by anaerobic digestion of Pongamia pinnata (Karanja) seed cakes.
In dual fuel mode, the biogas was inducted at four different ow rates, viz. 0.3 kg/h, 0.6 kg/h, 0.9 kg/h and
1.2 kg/h through the intake manifold of the engine. The biogas ow rate of 0.9 kg/h gave a better performance and lower emissions, than those of the other ow rates. The NO and smoke emissions were found
to be lower by about 34% and 14%, than those of KME operation, at full load. The ignition delay was longer
by about 12 CA in the dual fuel operation, than that of KME at full load. The part load performance was
found to be better in dual fuel operation, with reduced emissions of NO and smoke, in comparison with
KME. The ignition delay at part load in dual fuel operation was also lower than that of KME operation.
2014 Elsevier Ltd. All rights reserved.

1. Introduction
An increase in the greenhouse gas (GHG) emission increases the
global warming potential (GWP) and ozone depletion potential
(ODP). Due to these, different problem arises which include, rise
in global temperature, melting of glaciers, severe drought, reduction in the fertility of plants, etc. The main sources of the GHG
emission are automotive vehicles, power plants, and refrigeration
and air conditioning plants [1]. In 1997, the Kyoto protocol was
signed under the United Nations Framework Convention on Climate Change (UNFCCC). According to this protocol, many countries
agreed to reduce their emissions of CO2 and ve other GHG emissions, by implementing three mechanisms namely, Joint Implementation (JI), International Emissions Trading (IET) and Clean
Development Mechanism (CDM). Now, the Kyoto protocol covers
more than 160 countries in the world. The CDM is a means for
the developed countries to achieve part of their Kyoto target by
purchasing certied emission reductions from the GHG reducing
projects in developing countries. The production and utilization
of biofuels is one of the methods to implement the CDM in the
developing countries. It is reported that, during the period 2008
2012, about 55.7% of the CDM projects were hosted by China, followed by 30% by India and the remaining was by the other Asian
countries to implement the CDM [2]. Fuel switching such as
conventional petroleum fuels to renewable fuels in transportation
Corresponding author.
E-mail addresses: debabrata93@gmail.com (D. Barik), s.murugan@nitrkl.ac.in
(S. Murugan).
http://dx.doi.org/10.1016/j.enconman.2014.04.042
0196-8904/ 2014 Elsevier Ltd. All rights reserved.

sector is one of the technologies to be adopted in the CDM. The


Indian biofuel program has already been initiated by the government of India, to use 5% bioethanol with gasoline in automotive
vehicles. The government of India has also introduced the biodiesel
policy to support the CDM. As India is an agrarian country and has
vast agricultural lands, the production and utilization of biofuels
will greatly support the CDM.
Biodiesel is a good renewable fuel for CI engines. It is produced
by the transesterication of edible and nonedible vegetable oils,
animal fats, algae, etc. [3]. In India, nonedible oil seeds, such as
Jatropha, Karanja and Mahua are considered to be potential feed
stocks for biodiesel production. But the availability of such feed
stocks for biodiesel production is yet to be realistic. Biodiesel and
its diesel blends can be used directly in CI engines. Many researchers have reported that, biodiesel fueled diesel engines gave lower
carbon monoxide (CO), unburned hydrocarbons (UHC), particulate
matter (PM), and sulphur dioxide (SO2), but higher NOx than those
of diesel fueled engines [46], while some researchers reported the
opposite [7,8]. The reasons for the higher NO emission in biodiesel
engines are, the advance of injection timing as a result of higher
bulk modulus, higher adiabatic ame temperature, heat release
rate, more stoichiometric burning, presence of oxygen in the fuel
and less radiative heat transfer, and properties such as the boiling
point, viscosity and surface tension [916]. However, the exact
cause of NOx is yet to be unpredictable. The NOx emissions from
biodiesel fueled engines can be reduced by fuel modications such
as the emulsication of water with biodiesel, and the blending of
the high latent heat of vaporization fuel with the low latent
heat of vaporization fuel [1720]. Engine modications such as

218

D. Barik, S. Murugan / Energy Conversion and Management 84 (2014) 217226

retarding the injection timing, adopting exhaust gas recirculation


(EGR), and water injection are the viable options for reducing the
NOx emission [2125]. All the fuel or engine modications reportedly reduce the NOx emission, but the trade-off was not compromised in many techniques. It is reported that, the biogas dual
fuel operation has the potential to reduce both the NOx and
particulate emissions in CI engine, without affecting its thermal
efciency [2628].
Biogas is a potentially renewable fuel, cheaply producible and
environment friendly. It can be obtained from organic substances
through the anaerobic digestion process. The main components
in biogas are methane and carbon dioxide. The composition of biogas depends on the type of feed stocks, and production processes.
Different types of feed stocks have been explored for the production of biogas, such as cow dung, non-edible seed cakes, animal
waste, food waste, agricultural waste, municipal waste, and sewage sludge [29]. Biogas is a low cetane fuel, and it cannot be
directly used in a conventional diesel engine. It requires an ignition
source as the self-ignition temperature is high [30]. A pilot fuel
with a high cetane number is used to ignite the biogas, when it
is used in CI engines. The CO2 present in the biogas acts as an agent
for the reduction in NOx, when it is used in dual fuel engine [31,32].
Many research works have been documented on the utilization
of biogas in conventional diesel engines on dual fuel mode. It was
reported that, the thermal efciency of the engine dropped at part
loads [6,26,33], while it was unaffected, and increased at full load
[27,3438]. The NO and particulate matter were found to be lower
[26,27,31,39,40]. The HC and CO were reported to be higher in
some cases [31,40], and lower in others [26,27,35]. The volumetric
efciency of the biogas fueled engine was found to be lower than
that of diesel engine [30]. The ignition delay was reported to be
longer in dual fuel operation than that of diesel operation
[30,31]. The maximum cylinder pressure and rate of pressure rise
were found to be higher [31].
The purpose of the present study is to meet two objectives; (i)
to replace biodiesel by biogas which is also a renewable fuel, and
(ii) to evaluate the combustion, performance and emission of the
dual fuel engine. For this purpose, biogas at four different ow
rates was inducted in suction, while biodiesel was injected as a
pilot fuel. The engine behavior in terms of combustion, performance and the exhaust emission characteristics were assessed, in
comparison with KME operation in the same engine and presented
in this paper.
2. Materials and method
2.1. Test fuels
The important physical properties of KME are given in Table 1.
The primary fuel used in this study was biogas, produced by the

anaerobic digestion of Karanja seed cake (KSC). Table 2 gives the


properties of biogas obtained from the KSC, in comparison with
biogas obtained from cow dung (CD). The compositions of biogas
produced from the KSC were measured with the help of a nondispersive infrared biogas analyzer and given in Table 3.

2.2. Experimental setup


The detailed schematic diagram of the experimental setup is
shown in Fig. 1. The engine used in this research is a single cylinder, four stroke, naturally aspirated, air cooled, direct injection (DI)
diesel engine, with a displacement volume of 662 cm3 (stroke
110 mm, bore 87.5 mm) and compression ratio of 17.5. Table 4
gives the detailed specications of the test engine.
The power developed by the engine was measured by an electrical dynamometer (Kirloskar, WHD10075, ACG). The pilot fuel
injection system was composed of a mechanically actuated plunger type unit injection (MAPUI) system (MICO BOSCH) with a three
hole nozzle. The in-cylinder pressure at each crank angle (CA) was
measured with a quartz piezo-electric pressure transducer (Make:
KISTLER, Model: 5395A) mounted on the engine cylinder head, and
connected to a charge amplier. In each test, the combustion pressure was obtained every 0.6 CA interval by the data acquisition
system. The cylinder pressure vs. CA data was analyzed, applying
the rst law of thermodynamics to calculate the HRR (heat release
rate). The detailed discussion for the HRR calculation is provided in
Section 3. A twenty two channel signal analyzer was used for data
acquisition, and the acquired data was transferred through the
Ethernet cable, and stored in a PC for ofine analysis. A non-contact
type sensor was mounted near the ywheel, to measure the engine
speed. The exhaust gas temperature was measured with the help of
a K-type thermocouple tted in the exhaust manifold.
The biogas generated in the biogas plant (oating dome type)
was stored in a gas holder and supplied to the engine by a hose
pipe. A multi hole gas mixing kit was attached to the intake
manifold, for ensuring proper air and biogas mixture supply to

Table 2
Properties of biogas.
Properties

Test method,
ASTM

Biogas from
KSC

Biogas from
CD

Density, kg/m3
Caloric value, MJ/kg
Auto-ignition
temperature, C
Flame speed, m/s
A/F ratio, kg of air/kg of
fuel
Flammability limit, vol.%
in air
Octane number

D 3588
D 1945
E 659

1.2
27.53
600650

1.31
17.2
640670

D 7424
D 4891

25
17.23

21
15.3

D 6793

7.514

7.511.7

D 2699

130

110

Table 1
Properties of KME.
Properties

Test method, ASTM

Diesel

KME

Density, kg/m3
Caloric value, MJ/kg
Auto-ignition temperature, C
Flash point, C
Fire point, C
Pour point, C
Carbon residue, %
Cetane number
Carbon, wt.%
Hydrogen, wt.%
Sulphur, wt.%
Oxygen, wt.%

D 4052
D 4809
E 659
D 93
D 93
D 97
D 4530
D 613
D 3178
D 3178
D 3177
E 385

830
43.8
210350
50
56
6
0.1
50
85.3
13.19
0.3

880
40.96
170320
230
258
3
0.71
57.6
65.74
10.04
0.001
24.01

Table 3
Composition of biogas obtained from KSC.
Compounds

Molecular
formula

Biogas from KSC,


vol.%

Biogas from CD
vol.%

Carbon dioxide
Oxygen
Carbon
monoxide
Hydrogen
Methane
Nitrogen
Hydrogen
sulphide

CO2
O2
CO

17.37
1.5

2530
03

H2
CH4
N2
H2S

1.4
73
6.5
0.23

01
5070
010
03

D. Barik, S. Murugan / Energy Conversion and Management 84 (2014) 217226

219

continuous gas sampling combined with a heated and temperature


controlled smoke chamber compensates for changes in pressure
and test conditions to give the most accurate value of the smoke
opacity.
2.3. Experimental procedure
Initially, the reference data on the combustion, performance
and emissions of the engine were collected for the KME operation.
Then the engine was run on dual fuel mode. The biogas induction
rate was varied from 0.3 kg/h to 1.2 kg/h in steps of 0.3 kg/h. The
fuel injection pressure and engine speed were held constant at
200 bar and 1500 rpm. In both the KME and dual fuel operation,
the KME injection timing was kept constant at 23 CA bTDC. The
engine was loaded at 0%, 25%, 50%, 75% and 100% in both the
KME and dual fuel operation. The fuel injection quantities in both
the operations were controlled by the governor, with variations in
the engine load. The test engine was operated at 28 2 C of ambient temperature under all modes of operations.
1 Engine
2 Flywheel
3 Speed sensor
4 AC dynamometer
5 Resistive load cell
6 EGT sensor
7 Exhaust manifold
8 Fuel injector
9 Pressure transducer
10 Air filter

11 Air flow sensor


12 Air box
13 Mixing kit
14 Gas flow meter
15 Pressure gauge
16 Biogas regulator
17 Hose pipe
18 Biogas holder
19 Digester
20 Feedstock feeding funnel

21 Fuel tank
22 High fuel level optical sensor
23 Low fuel level optical sensor
24 Fuel pump
25 Control panel
26 Exhaust gas analyzer
27 Smoke meter
28 Data acquisition system
29 Monitor
30 Engine base

Fig. 1. Schematic diagram of engine experimental setup.

2.4. Error analysis


The measurements obtained in the experimentation might have
errors and uncertainties, which could arise from the instruments
used, operating condition, calibration, observation, test planning
and environment. To examine the accuracy of the measurement,
an uncertainty analysis was performed, using the method
described by Coleman and Steele [41].

U R B2R P 2R
Table 4
Specications of the test engine.
Make/model
Brake power, kW
Rated speed, rpm
Cooling system
Burning clearance, mm
Injection nozzle
Nozzle opening pressure, bar
Injection timing, CA bTDC
Inlet valve open, CA bTDC
Inlet valve close, CA aBDC
Exhaust valve open, CA bBDC
Exhaust valve close, CA aTDC

Kirloskar TAF 1
4.4
1500
Air
1.11.2
3-Hole
200
23
4.5
35.5
35.5
4.5

the engine. The quantity of biogas inducted into the engine in each
operating module was controlled and measured with a biogas ow
meter (Make: Siya, Model: SI10), attached before the mixing kit
between the intake manifold and biogas holder. The air consumption by the engine was measured with a differential pressure sensor tted in the air box. The pilot fuel consumption was measured
by a vertical burette tted with two optical fuel level sensors.
The exhaust emission measurements without catalytic treatment were done with a gas analyzer. The HC, CO and CO2 emissions
were measured on the NDIR (non-dispersive infrared) principle,
and the NO emission was measured by an electro chemical sensor.
The smoke opacity of the engine exhaust was measured with a diesel smoke meter. This works on the principle of the Hartridge
smoke meter, and measures the smoke opacity. The Hartridge
smoke meter consists of an optical unit mounted inside a measuring head and a separate electronic control unit. The measurement
principle is based on light extinction detection. The collimated
beam from the light-source is absorbed and scattered by the particulate exhaust emissions. A photodiode determines the light
intensity of the attenuated beam and the corresponding opacity
value is transmitted to a separate remote display. Partial ow

1=2

UR is the uncertainty of the physical parameters, using the rootsum-square (RSS) method, at 5% signicance level [42]. BR and PR
are the systematic and random uncertainties respectively.

" 
2 #1=2
n
X
BR
1 @R

Bi
R @X i
R
i1

" 
2 #1=2
n
X
PR
1 @R

Pi
R @X i
R
i1

In the equation cited, R is the physical parameter dependent on


the variable Xi. The symbol PR and Pi denote the uncertainty in R
and the measurement level, respectively. As a result, the maximum
uncertainty of the experiment obtained was 2.201%. Table 5 lists
out the instruments used in the present investigation and their
uncertainties.
2.5. Biogas energy share and air excess ratio
In dual fuel operation, the energy share of the primary gaseous
fuel is an important parameter for analyzing the premixed lean
combustion. In order to generate a certain amount of power, both
the primary fuel and pilot fuel contributes energy. During combustion, the consumption of the primary fuel remains unchanged with
the change in load, while the pilot fuel consumption varies with
the load. The energy share by a fuel strongly depends on the
caloric value and rate of fuel consumption. The energy share in
the dual fuel operation is dened as the ratio of energy supplied
by the primary fuel, to the sum of the energy supplied by the
primary fuel and the pilot fuel. The biogas energy share is quantied by Eq. (4).

Biogas energy share

Energy equivalent of biogas


 100
Energy equivalent of KME biogas
4

220

D. Barik, S. Murugan / Energy Conversion and Management 84 (2014) 217226

Table 5
Range, accuracy, and uncertainty of the instruments used.
Instruments used

Parameter measured with units

Range

Accuracy

Uncertainty, %

Load indicator
Temperature indicator
Burette
Air ow meter
Biogas ow meter
Speed sensor
Charge amplier
Pressure transducer
Crank angle encoder
Data acquisition system
Smoke meter
Exhaust gas analyzer

Engine load, W
Temperature, C
Fuel volume, cm3
Air consumption, m3/min
Biogas consumption, m3/min
Engine speed, rpm
Amplies input voltage, V
In-cylinder pressure, bar
Crank position, CA
Converts signal to digital values, bit
Smoke opacity, %
Exhaust emission
NO, ppm
HC, ppm
CO, %

2506000
0900
130
0.550
0.125
010,000

0110
0720
64
0100
NO: 05000
HC: 020,000
CO: 010

10
1
0.2
0.1
0.1
10
1%
0.1
0.6
0.1
1
50
10
0.03

0.2
0.15
0.5
0.5
0.02
1
0.1
0.15
0.01
0.001
1
1
0.5
0.03
2.201

Over all uncertainty

Table 6
Energy share of KME and biogas at 1.2 kg/h.
Load, Mass
%
of
KME,
kg/h

Biogas
KME
Air
Energy
Mass of Energy
energy
equivalent excess energy
biogas, equivalent
share, % share, %
of KME, kW of biogas, ratio
kg/h
kW

0
25
50
75
100

1.2
1.2
1.2
1.2
1.2

0.406
0.537
0.844
1.101
1.353

4.62
6.11
9.60
12.58
15.39

9.17
9.17
9.17
9.17
9.17

0.88
0.81
0.67
0.59
0.52

33.4
39.9
51.1
57.7
62.6

66.6
60.1
48.9
42.3
37.4

where,

Energy equivalent of KME

_ KME  CVKME
m
3600

and,

Energy equivalent of biogas

_ biogas  CVbiogas
m
3600

The k for the dual fuel operation was in the range of 1.980.52, from
no load to full load. The effects of biogas energy share, on the k at no
load and full load for the KME and dual fuel operation, are given in
Table 7.
The variation of the k with biogas energy share is depicted in
Fig. 2. It can be observed from the gure that, an increase in the
biogas fraction in the fuel air mixture, results in a reduction in
the k, as biogas replaces the air in the intake mixture in the dual
fuel operation. The gaseous fuel would result in a more air displacement, and hence lower the k.
The variation of biogas induction quantity with biogas energy
share is depicted in Fig. 3.
It can be observed from the gure that, the biogas energy share
is low at full load, while the energy share is high at no load, for all
the ow rates of biogas in dual fuel operation. This is due to more
KME consumption at relatively high load than that of no load. The
biogas ow rate of 1.2 kg/h gives the maximum energy share in
comparison with the other ow rates throughout the load spectrum. At full load, the biogas ow rates of 0.3 kg/h, 0.6 kg/h,

_ KME and m
_ biogas is the mass ow rate of KME and biogas,
where m
and CV is the caloric value of the fuel used. The energy share of
biogas at a ow rate of 1.2 kg/h is given in Table 6.
For in-depth analysis of dual fuel operation, the air excess ratio
was calculated from the stoichiometric airfuel ratio. The air
excess ratio, k equals the airfuel ratio over the stoichiometric
ratio. The expression for the air excess ratio in dual fuel operation
was calculated from the following correlation.

kdual h 
A

F biogas
stoic

where

h 
A

F biogas
stoic

_ air
m
_ biogas
m
and

  A

A

F KME stoic

F KME stoic

_ KME
m

Load 0%
Load 25%
Load 50%
Load 75%
Load 100%

Biogas: 0.3kg/h, 0.6kg/h,


0.9kg/h, 1.2kg/h

are the stoichiometric airfuel

0
0

ratios of biogas and KME respectively. The stoichiometric airfuel


ratio for KME was calculated with the molecular composition of
KME (C19.4H38O2) based on the average fatty acid composition.

15

30

45

60

75

90

Biogas energy share, %


Fig. 2. Effect of biogas energy share on the air excess ratio.

Table 7
Effects of biogas energy share on the k for KME and dual fuel operation.
Mode of operation

KME
KME + biogas
KME + biogas
KME + biogas
KME + biogas

Biogas ow rate, kg/h

0.3
0.6
0.9
1.2

Biogas energy share, %

k
No load

Full load

Load 0%

Load 25%

Load 50%

Load 75%

Load 100%

4.41
1.98
1.37
1.07
0.88

1.41
0.78
0.67
0.59
0.52

29.8
46.9
58.5
66.6

25.5
41.5
52.6
60.1

18.3
31.4
41.5
48.9

14.6
25.7
34.7
42.3

11.8
21.8
30.2
37.4

221

D. Barik, S. Murugan / Energy Conversion and Management 84 (2014) 217226

Biogas induction, kg/h

Load 0%
Load 25%
Load 50%
Load 75%
Load 100%

1.6
1.2

pressure in the premixed combustion phase [39]. It is observed


that the presence of CO2 in biogas did not signicantly affect the
maximum cylinder pressure, but the ignition and peak cylinder
pressure occurred little later and shifts some degrees toward the
expansion process [38,40].
The HRR, Q_ was calculated using the rst law analysis, and the
expression is given below.

Biogas: 0.3kg/h, 0.6kg/h,


0.9kg/h, 1.2kg/h

0.8

c dV c dP
Q_ 1 P

V
qw
c
dh c1 dh

0.4
0
0

15

30

45

60

75

90

Biogas energy share, %


Fig. 3. Variation of biogas induction quantity with energy share.

0.9 kg/h and 1.2 kg/h gives energy shares of about 11.8%, 21.8%,
30.2% and 37.4% respectively.
3. Results and discussion
3.1. Combustion analysis
The peak cylinder pressure and HRR at full load for the KME and
dual fuel operation with different biogas energy shares are
depicted in Fig. 4. The peak cylinder pressure mainly depends on
the rate of combustion in the premixed combustion stage [43].
The peak cylinder pressure for KME is about 71.33 bar, which
occurs at 12.41 CA aTDC. In dual fuel operation the peak cylinder
pressures of 73.71 bar, 73.57 bar, 72.97 bar and 74.29 bar occur at
12.95 CA aTDC, 12.76 CA aTDC, 13.18 CA aTDC and 12.78 CA
aTDC, for the biogas ow rates of 0.3 kg/h, 0.6 kg/h, 0.9 kg/h and
1.2 kg/h respectively, at full load. In the dual fuel operation, the
increase in the biogas ow rate results in an increase in the cylinder pressure. The reason for the higher peak cylinder pressure in
the dual fuel operation than that of KME operation is due to the
induction of biogas with the intake-air charge brings about a
decrease and dilution of oxygen concentration, which may cause
ignition delay to extend, leading to a higher rate of increase in

75

KME
KME+biogas 0.6 kg/h
KME+biogas 1.2 kg/h

90

KME+biogas 0.3 kg/h


KME+biogas 0.9 kg/h

75

Heat release rate, J/ o CA

Cylinder pressure, bar

90

60
60
45
45

30

30

15

15
0
-50

0
-15
-25

25

50

75

Crank angle, o CA
Fig. 4. Variation of cylinder pressure and HRR with CA at full load.

where P is the cylinder pressure, h is the crank angle and c is the


ratio of specic heats, Cp/Cv. The appropriate range for c for the diesel heat release analysis is 1.31.35. V is the instantaneous cylinder
volume (m3). The instantaneous cylinder volume was obtained from
the engine geometry and crank angle values. The wall heat transfer
qw, and blow by losses are not considered in this study to nd the
heat released inside the cylinder. The caloric value of KMEbiogas
at full load with different biogas ow rates are given in Table 8.
The HRR in the premixed combustion phase depends on the
ignition delay, mixture formation and the combustion rate in the
initial stages [38]. The dual fuel operation gives a higher HRR than
that of KME operation. This is due to the combined effect of the
combustion of pilot fuel and gaseous fuel in the immediate vicinity
of the ignition and combustion centers of the pilot fuel [40]. The
dual fuel operation with the biogas energy share of 3058% gives
the maximum heat release of 55.2 J/CA at full load, which is higher
than that of other energy shares. Maximum amount of energy is
found to release immediately following the commencement of
auto-ignition of the pilot fuel. This is due to increase in concentration of biogas in the air and corresponding k which modies and
extends signicantly the ammability zone around the pilot fuel
[44,45]. The occurrence of HRR is found to be earlier in KME operation, than that of dual fuel operation is due to its higher cetane
number and the presence of oxygen molecules in the fuel itself.
The maximum HRR for KME is observed at 11.83 CA aTDC,
whereas, in dual fuel operation, the maximum HRR is observed
with biogas ow rate of 0.9 kg/h, at 13.18 CA aTDC, and with a
maximum biogas energy share of 58%, at full load.
Fig. 5 depicts the effects of the biogas energy share on the ignition delay. The ignition delay is a time period measured in CA
between the start of injection and start of combustion. The start
of combustion may be taken as the minimum point on the HRRCA curve (where HRR becomes positive, Fig. 4). The ignition delay
is mainly inuenced by parameters, such as the fuel type, fuel quality, airfuel ratio, engine speed, fuel atomization quality, inducted
air temperature, compression ratio and pressure [43].
In dual fuel operation, the biogas energy share of 11.866.6%
exhibits a longer ignition delay than that of KME operation
throughout the load spectrum. The ignition delay in the dual fuel
operation increases by about 12 CA than that of KME, with the
increase in the biogas energy share. This is due to the presence
of biogas in fuelair mixture, which alters the pre-ignition property of the charge, and reduces the oxygen concentration [31]. A
shorter ignition delay is noticed for the fuels at full load, than that
of no load, are due to the higher cylinder temperature at full load.
The ignition delay period usually increases with an increase in the
amount of gaseous fuel in the charge for most gaseous fuels under
normal operating conditions [40].

Table 8
Caloric value for dual fuel operation with different biogas ow rates at full load.
Properties

KME + biogas 0.3 kg/h

KME + biogas 0.6 kg/h

KME + biogas 0.9 kg/h

KME + biogas 1.2 kg/h

Caloric value, MJ/kg

43.98

46.74

49.21

51.14

222

D. Barik, S. Murugan / Energy Conversion and Management 84 (2014) 217226

25

19

Fuel: KME and KME+Biogas, with


biogas flow rate of 0.3kg/h, 0.6kg/h,
0.9kg/h, 1.2kg/h

Maximum cylinder pressure, bar

Ignition delay, o CA

22

90

Load 0%
Load 25%
Load 50%
Load 75%
Load 100%
KME

16
13

80
KME

70

60

50

10
-15

15

30

45

60

75

-15

90

Combustion duration, o CA

Fuel: KME and KME+Biogas, with


biogas flow rate of 0.3kg/h, 0.6kg/h,
0.9kg/h, 1.2kg/h

KME

40

30

20
-15

120

Mass fracion burned, %

The effect of biogas energy share on combustion duration is


depicted in Fig. 6. The combustion duration for KME is 38.3 CA,
at full load. The combustion duration for the dual fuel operation
increases with increase in biogas energy share and vice versa.
The biogas energy share of 11.8%, 21.8%, 30.2% and 37.4% gives a
marginal increase in combustion duration of 39.2 CA, 40 CA,
40.8 CA and 41.8 CA respectively, at full load. This is due to induction of more quantity of biogas. The inducted biogas alters the
physical properties of the charge being compressed, reduces the
oxygen concentration in the charge mixture causes slower diffusion combustion, which results prolonged combustion duration.
In addition, biogas causes a slower rate of burning of the pilot fuel,
as it contains about 17% CO2. In both KME and dual fuel operation,
longer combustion duration is noticed at full load than that of no
load, is due to the consumption of more fuel at relatively high load.
Fig. 7 portrays the variation of the maximum cylinder pressure
with different energy shares. The cylinder pressure decreases with
the increase in the biogas energy share. The maximum cylinder
pressure for KME is about 71.3 bar. In dual fuel operation, the maximum cylinder pressure is 73.7 bar, 73.5 bar, 72.9 bar, and 74.2 bar
with the biogas energy shares of 11.8%, 21.8%, 30.2% and 37.4%
respectively, at full load. The increase in the maximum cylinder
pressure in the dual fuel operation is due to the induction of biogas
with the intake-air charge brings about a decrease and dilution of
oxygen concentration, which may cause ignition delay to extend,
leading to a higher rate of increase in pressure in the premixed
combustion phase [39].
The variation of the mass fraction burned (MFB) with the CA, for
both the KME and dual fuel operation, at full load is depicted in
Fig. 8. The energy conversion during a specic combustion cycle
can be expressed by the MFB at a specic CA. The MFB at each

50

30

45

60

75

90

Fig. 7. Variation of the maximum cylinder pressure with biogas energy share.

Fig. 5. Effect of the biogas energy share on ignition delay.

Load 0%
Load 25%
Load 50%
Load 75%
Load 100%

15

Biogas energy share, %

Biogas energy share, %

60

Load 0%
Load 25%
Load 50%
Load 75%
Load 100%

Fuel: KME and KME+Biogas, with


biogas flow rate of 0.3kg/h, 0.6kg/h,
0.9kg/h, 1.2kg/h

KME
KME+biogas 0.3 kg/h
KME+biogas 0.6 kg/h
KME+biogas 0.9 kg/h
KME+biogas 1.2 kg/h

100
80
60
40
20
0
-20

-10

10

20

30

40

50

Crank angle, o CA
Fig. 8. Variation of mass fraction burned with CA at full load.

individual combustion cycle is a normalized quantity with a scale


01, describing the process for the conversion of chemical energy
into work, as a function of CA. This is also indicative of the start
and end of combustion. The correlation for estimating the MFB
from the cylinder pressure and volume data is given as:

MFB

Pi
mb i
DP c
PN0
mb total
0 DP c

where 0 denotes the start of combustion, N is the total number of


CA intervals, DPc is the pressure rise due to combustion and i is
the CA locator.
The mass fraction burned for KME is marginally higher than
that of dual fuel operation, at full load. The 90% MFB for KME is
occurred at about 30 CA aTDC. In dual fuel operation, the occurrence of 90% mass fraction burned is noticed at 31 CA aTDC, 32
CA aTDC, 32.8 CA aTDC and 33.4 CA aTDC, at the biogas ow
rates of 0.3 kg/h, 0.6 kg/h, 0.9 kg/h and 1.2 kg/h respectively. The
advanced MFB for KME is due to its higher cetane number and
availability of oxygen molecules in the fuel itself. The delay in
MFB for dual fuel operation than that of KME is due to the drop
in k and dilution of charge due to the available CO2 in biogas.
The values of ignition delay, cylinder peak pressure, maximum
HRR, combustion duration and MFB at full load for the fuels tested
are given in Table 9.
3.2. Performance analysis

15

30

45

60

75

90

Biogas energy share, %


Fig. 6. Variation of combustion duration with biogas energy share.

The variations in the brake specic energy consumption (BSEC)


with the biogas energy share, for the fuels tested, are depicted in
Fig. 9. The BSEC increases with the increase in the biogas energy

223

D. Barik, S. Murugan / Energy Conversion and Management 84 (2014) 217226


Table 9
Combustion parameters of KME and KMEbiogas operation at full load.
Parameters

KME

KME + biogas 0.3 kg/h

KME + biogas 0.6 kg/h

KME + biogas 0.9 kg/h

KME + biogas 1.2 kg/h

Ignition delay, CA
Maximum cylinder pressure, bar
Maximum HRR, J/CA
Combustion duration, CA
Mass fraction burned (for 90%), CA aTDC

11.5
71.33
52.4
38.31
30

12
73.71
53.1
39.27
31

12.2
73.57
54.1
40.05
32

12.4
72.97
55.2
40.84
32.8

12.6
74.29
52.6
41.89
33.4

Load 25%
Load 50%
Load 75%
Load 100%

60

40

Fuel: KME and KME+Biogas,


with biogas flow rate of 0.3kg/h,
0.6kg/h, 0.9kg/h, 1.2kg/h

KME

20

0
-15

15

30

45

60

75

90

Biogas energy share, %


Fig. 9. Variation of BSEC with biogas energy share.

Exhaust gas temperature, o C

share, for all the fuels tested in the dual fuel operation. A higher
BSEC is noticed in the dual fuel operation than that of KME at part
loads. This is due to the lower energy density of biogas, lower cylinder temperature, and the presence of CO2 in biogas which prevents faster burning. The difference in BSEC between KME and
dual fuel operation are not signicantly different at high operating
loads. At full load the dual fuel operation has similar fuel-energy
conversion efciency to that of KME [46,47]. Because less energy
from the fuel is required at full load compared to no load, due to
the increased cylinder temperature at full load [40]. The biogas
ow rate of 1.2 kg/h shows a higher BSEC of 51.8 MJ/kWh at a
higher energy share of 60%, at 25% load. The BSEC for KME is
11.8 MJ/kWh at full load, and for the dual fuel operation with the
biogas ow rates of 0.3 kg/h, 0.6 kg/h, 0.9 kg/h and 1.2 kg/h, with
biogas energy share of 11.8%, 21.8%, 30.2% and 37.4%, the BSEC
are 15.9 MJ/kWh, 17.3 MJ/kWh, 19.2 MJ/kWh and 21.5 MJ/kWh
respectively, at full load.
Fig. 10 illustrates the effect of the biogas energy share on the
exhaust gas temperature. The KME operation gives the highest
exhaust gas temperature throughout the load spectrum. The
oxygen present in the ester molecule enhances the combustion

550

Load 0%
Load 25%
Load 50%
Load 75%
Load 100%

Fuel: KME and KME+Biogas,


with biogas flow rate of 0.3kg/h,
0.6kg/h, 0.9kg/h, 1.2kg/h

450

KME

350
250
150
50
-15

process, and results in a higher exhaust gas temperature. In the


dual fuel operation, the exhaust gas temperature goes down with
the increase in the biogas energy share. This is because, the biogas
in the inducted air gets hotter to auto ignite during the combustion
process, and absorbs the heat energy, which decreases the local
adiabatic ame temperature, leading to a reduction in the exhaust
gas temperature. On the other hand the lower exhaust gas temperature is due the increase in diluent resulting from the 17% CO2 in
the biogas.
Fig. 11 depicts the variation of volumetric efciency with the
biogas energy share. The volumetric efciency of KME is found to
be higher than that of the dual fuel operation. The volumetric efciency of KME is found to be about 84.5% at full load, whereas, the
volumetric efciencies of dual fuel operation at the biogas ow
rates of 0.3 kg/h, 0.6 kg/h, 0.9 kg/h and 1.2 kg/h, are about
83.77%, 82.92%, 82% and 79.9% respectively, at full load. The reduction in the volumetric efciency in the dual fuel operation is due to
the induction of biogas with the air through the intake manifold,
replaces some of fresh air. The k for the dual fuel operation
decreases with the increase in the biogas energy share (given in
Table 7), which directly inuences the volumetric efciency. The
drop in volumetric efciency in full load in comparison to part load
may be due to the increase in the inducted air temperature,
because of the hot cylinder wall, which results in the reduction
of air density, causing a reduction in the volumetric efciency at
full load [48].

3.3. Emission analysis


The concentration of the BSCO emission with respect to the biogas energy share is depicted in Fig. 12. In general, the BSCO emission is formed as a result of incomplete combustion of the fuel. The
BSCO emission in the dual fuel operation with the biogas energy
share of 11.837.4% is considerably higher than that of KME under
all test conditions. This may be due to the dilution of the charge by
CO2 present in the biogas gives a higher CO emission. The biogas
energy shares of 11.8%, 21.8%, 30.2% and 37.4% shows higher BSCO
emissions of 9%, 16.6%, 28.5% and 37.5% respectively, than that of

93

Volumetric efficiency, %

BSEC, MJ/kWh

80

Load 0%
Load 25%
Load 50%
Load 75%
Load 100%

Fuel: KME and KME+Biogas, with


biogas flow rate of 0.3kg/h, 0.6kg/h,
0.9kg/h, 1.2kg/h

90
87
84

KME

81
78

15

30

45

60

75

Biogas energy share, %


Fig. 10. Variation of exhaust gas temperature with biogas energy share.

-15

15

30

45

60

75

90

Biogas energy share, %


Fig. 11. Variation in the volumetric efciency with biogas energy share.

224

D. Barik, S. Murugan / Energy Conversion and Management 84 (2014) 217226

0.075

BSCO, g/kWh

0.06

0.045

Fuel: KME and KME+Biogas,


with biogas flow rate of 0.3kg/h,
0.6kg/h, 0.9kg/h, 1.2kg/h

KME

0.03

Load 25%

Fuel: KME and KME+Biogas,


with biogas flow rate of 0.3kg/h,
0.6kg/h, 0.9kg/h, 1.2kg/h

BSNO, g/kWh

Load 25%
Load 50%
Load 75%
Load 100%

Load 50%
Load 75%

KME

Load 100%

4
3
2

0.015

1
-15

0
-15

15

30

45

60

75

90

15

30

45

60

75

90

Biogas energy share, %

Biogas energy share, %


Fig. 14. Variation in BSNO emission with biogas energy share.
Fig. 12. Variation of BSCO with biogas energy share.

Load 25%
Load 50%
Load 75%
Load 100%

BSHC, g/kWh

0.3

KME

0.09

0.02
15

30

45

60

75

Biogas energy share, %


Fig. 13. Variation of BSHC emissions with biogas energy share.

Load 50%
Load 75%
Load 100%

KME

0
-15

15

30

45

60

75

90

Biogas energy share, %


Fig. 15. Variation of BSCO2 emission with biogas energy share.

concentration of the charge followed by overall decrease in cycle


temperature. Hence, the NO formation is suppressed with the combined effect of these phenomena [5254]. The higher BSNO emission for KME operation is due to the higher oxygen concentration
[42]. A reduction of 24%, 29%, 34% and 39.5% in the BSNO emission
is obtained in the dual fuel operation, with the biogas energy share
of 11.8%, 21.8%, 30.2% and 37.4% respectively, in comparison with
KME, at full load.
The variation of BSCO2 with biogas energy share is depicted in
Fig. 15.
KME gives a maximum BSCO2 of 1.18 g/kW h at full load. This is
due to the additional oxygen molecules present in the KME attribute to complete combustion [27]. In dual fuel operation the BSCO2

65

0.16

Load 25%

Fuel: KME and KME+Biogas,


with biogas flow rate of 0.3kg/h,
0.6kg/h, 0.9kg/h, 1.2kg/h

0.23

-15

Fuel: KME and KME+Biogas,


with biogas flow rate of 0.3kg/h,
0.6kg/h, 0.9kg/h, 1.2kg/h

Smoke opacity, %

0.37

BSCO2, g/kWh

KME operation at full load. The lower BSCO emission from the KME
operation is attributed to the additional oxygen present in the fuel.
Also, the increased cetane number of KME lowers the probability of
forming a fuel rich zone, and advanced ignition [49]. It is also
apparent from the gure that, the BSCO emissions are high at
low loads due to poor mixture formation.
The variation of the BSHC emission with biogas energy share is
depicted in Fig. 13. The dual fuel operation shows a higher BSHC
emission than that of KME operation. The higher BSHC emission
in the dual fuel operation is due to the induction of excess amount
of biogas through the intake manifold, which reduces the volume
of inducted air forming a richer mixture, and an increase in the
partial burning with less oxygen [50,51]. Thus, it is evident that
the BSHC level goes up, whenever the biogas energy share
increases and vice-versa. The BSHC emission for KME is lower than
that of dual fuel operation. Probably, the oxygen content in the
KME gives complete combustion and reduces the level of BSHC
emission [31]. About 12%, 18%, 23% and 42% increment in the BSHC
emission is obtained in the dual fuel operation, at the biogas ow
rates of 0.3 kg/h, 0.6 kg/h, 0.9 kg/h and 1.2 kg/h respectively than
that of KME, at full load.
The variation of the BSNO emission with biogas energy share is
shown in Fig. 14. The concentration of BSNO for the dual fuel operation is considerably lower than that of KME operation, with
increase in the biogas energy share. The BSNO formation is highly
dependent on the combustion temperature, availability of oxygen,
compression ratio and the retention time for the reaction. In the
dual fuel operation, the presence of CO2 having a high molar
specic heat, dilutes the charge and lowers the cycle temperature
signicantly. In addition, CO2 in biogas lowers the oxygen

90

50

Load 0%
Load 25%
Load 50%
Load 75%
Load 100%

Fuel: KME and KME+Biogas,


with biogas flow rate of 0.3kg/h,
0.6kg/h, 0.9kg/h, 1.2kg/h

KME

35

20

5
-15

15

30

45

60

75

90

Biogas energy share, %


Fig. 16. Variation of smoke emission with biogas energy share.

225

D. Barik, S. Murugan / Energy Conversion and Management 84 (2014) 217226


Table 10
Performance and emission parameters of KME and KMEbiogas operation at full load.
Parameters

KME

KME + biogas 0.3 kg/h

KME + biogas 0.6 kg/h

KME + biogas 0.9 kg/h

KME + biogas 1.2 kg/h

BSEC, MJ/kWh
Volumetric efciency, %
BSCO, g/kWh
BSHC, g/kWh
BSNO, g/kWh
BSCO2, g/kWh
Smoke opacity, %

11.80
84.5
0.01
0.036
2.61
1.18
26.2

15.98
83.7
0.011
0.041
1.96
0.88
25

17.37
82.9
0.012
0.044
1.85
0.78
22.8

19.25
82
0.014
0.047
1.72
0.56
22.4

21.5
79.9
0.016
0.063
1.58
0.46
20.3

decreases with the increase in the biogas energy share. The biogas
energy share of 11.8%, 21.7%, 30.2% and 37.4% gives a lower BSCO2
of 25%, 33%, 52% and 60% respectively, than that of KME at full load.
This reduction in the BSCO2 emission in the dual fuel operation is
attributed to the lower volumetric efciency, lower k and higher
CO2 in biogas.
Fig. 16 depicts the variation in the concentration of smoke opacity with biogas energy share.
It is observed that the dual fuel operation is a potential way of
reducing smoke emission. Specically, at low engine loads in dual
fuel operation, the energy share by biogas is more and the percentage of KME substitution increases, hence smoke opacity decreases.
The dual fuel operation with the biogas energy share of 11.8%,
21.8%, 30.2%, and 37.4% give lower smoke emission of 4.5%,
12.9%, 14.5% and 22.5% than that of KME, at full load, respectively.
It is apparent that, in dual fuel operation, using biogas is a very
effective method to reduce smoke emission at almost all engine
operating points. This is due to the presence of methane in biogas,
as the main constituent, that is the lower member in the parafn
family, possesses very small tendency to produce soot [55]. In general the reduction of smoke is attributed to ame temperature
reduction and increased oxidation of soot precursors in the soot
forming region by the enhanced concentration of O and OH around
the ame (resulting in high oxidation) produced from the CO2 in
biogas [40]. The CO2 concentration in the fuel causes a decrease
in overall cycle temperature, which has a positive (i.e. reduction)
impact on smoke formation and, at the same time, does not seem
to have an adverse (i.e. increase) effect on the NO emissions, as
generally happens in normal diesel engines.
The performance and emission parameters for KME and dual
fuel operation at full load are summarized in Table 10.
4. Conclusions
The conclusions of the results obtained in the investigation are
as follows;
 In dual fuel operation, the biogas ow rate of 0.9 kg/h with the
biogas energy share of 30.158.4% shows the optimum results,
in terms of the performance, combustion and emission, than
that of other ow rates, tested in this study.
 In dual fuel operation, the BSEC was found to be higher by about
38% than that of KME operation, at the biogas ow rate of
0.9 kg/h, at full load.
 The k for the dual fuel operation dropped from 0.78 to 0.52, with
a change in the biogas energy share of 11.8% to 37.4%.
 In dual fuel operation, about 30% of KME replacement was possible at the biogas ow rate of 0.9 kg/h, at full load.
 The CO and HC emissions were found to be higher by about 28%
and 23% at the biogas ow rate of 0.9 kg/h, in comparison with
KME at full load.
 The NO, smoke and CO2 emissions at full load were found to be
lower by about 34%, 14% and 52% respectively, in dual fuel operation at a biogas ow rate of 0.9 kg/h, in comparison with KME.

The adoption of this KMEbiogas dual fuel technology will


boost the proportion of the farms renewable energy usage and
can reduce the KME production cost. It is also a promising technique for on farm electricity generation. This biogas dual fuel
engine, not only consume a wide range of gaseous fuel resources
effectively, but also has the potential to avoid much of the current
and future problems facing to diesel engine including very signicant reduction in their exhaust emissions tradeoffs.
References
[1] Heywood JB. Internal combustion engine fundamentals. New York: McGrawHill; 1988.
[2] Azhar Khan Muhammad, Zahir Khan Muhammad, Zaman Khalid, Naz Lubna.
Global estimates of energy consumption and greenhouse gas emissions. Renew
Sustain Energy Rev 2014;29:33644.
[3] Rakopoulos CD, Antonopoulos KA, Rakopoulos DC. Development and
application of multi-zone model for combustion and pollutants formation in
direct injection diesel engine running with vegetable oil or its bio-diesel.
Energy Convers Manage 2007;48:1881901.
[4] Yoon SH, Suh HK, Lee CS. Effect of spray and EGR rate on the combustion and
emission characteristics of biodiesel fuel in a compression ignition engine.
Energy Fuels 2009;23:148693.
[5] Kivevele Thomas T, Kristf Lukcs, Bereczky kos, Mbarawa Makame M.
Engine performance, exhaust emissions and combustion characteristics of a CI
engine fuelled with croton megalocarpus methyl ester with antioxidant. Fuel
2011;90:27829.
[6] Luijten CCM, Kerkhof E. Jatropha oil and biogas in a dual fuel CI engine for rural
electrication. Energy Convers Manage 2011;52:142638.
[7] Dorado MP, Ballesteros E, Arnal JM, Gmez J, Lpez F. Exhaust emissions form a
diesel engine fueled with transesteried waste olive oil. Fuel 2003;82:13115.
[8] Lapuerta M, Armas O, Ballesteros R, Fernndez J. Diesel emissions from biofuels
derived from Spanish potential vegetable oils. Fuel 2005;84:77380.
[9] Bittle JA, Knight BM, Jacobs TJ. Interesting behavior of biodiesel ignition delay
and combustion duration. Energy Fuels 2010;24:416677.
[10] Agarwal Avinash Kumar, Srivastava Dhananjay Kumar, Dhar Atul, Maurya
Rakesh Kumar, Shukla Pravesh Chandra, Singh Akhilendra Pratap. Effect of fuel
injection timing and pressure on combustion, emissions and performance
characteristics of a single cylinder diesel engine. Fuel 2013;111:37483.
[11] Roy Murari Mohon, Wang Wilson, Bujold Justin. Biodiesel production and
comparison of emissions of a DI diesel engine fueled by biodieseldiesel and
canola oildiesel blends at high idling operations. Appl Energy
2013;106:198208.
[12] Ng JH, Ng HK, Gan SY. Characterisation of engine-out responses from a light
duty diesel engine fuelled with palm methyl ester (PME). Appl Energy
2012;90:5867.
[13] Xue JL, Grift TE, Hansen AC. Effect of biodiesel on engine performances and
emissions. Renew Sustain Energy Rev 2011;15:1098116.
[14] Sun JF, Caton JA, Jacobs TJ. Oxides of nitrogen emissions from biodieselfuelled
diesel engines. Progess Energy Combust Sci 2010;36:67795.
[15] Buyukkaya E. Effects of biodiesel on a DI diesel engine performance, emission
and combustion characteristics. Fuel 2010;89:3099105.
[16] Di Y, Cheung CS, Huang ZH. Experimental investigation on regulated and
unregulated emissions of a diesel engine fueled with ultra-low sulfur diesel
fuel blended with biodiesel from waste cooking oil. Sci Total Environ
2009;407:83546.
[17] Subramanian KA. A comparison of waterdiesel emulsion and timed injection
of water into the intake manifold of a diesel engine for simultaneous control of
NO and smoke emissions. Energy Convers Manage 2011;52:84957.
[18] Yilmaz Nadir, Vigil Francisco M, Burl Donaldson A, Darabseh Tariq.
Investigation of CI engine emissions in biodieselethanoldiesel blends as a
function of ethanol concentration. Fuel 2014;115:7903.
[19] Fang Qiang, Fang Junhua, Zhuang Jian, Huang Zhen. Effects of ethanoldiesel
biodiesel blends on combustion and emissions in premixed low temperature
combustion. Appl Therm Eng 2013;54:5418.
[20] Saravanana S, Nagarajan G, Sampath S. Combined effect of injection timing,
EGR and injection pressure in NOx control of a stationary diesel engine fuelled
with crude rice bran oil methyl ester. Fuel 2013;104:40916.

226

D. Barik, S. Murugan / Energy Conversion and Management 84 (2014) 217226

[21] Zhao Junfeng, Wang Junmin. Effect of exhaust gas recirculation on biodiesel
blend level estimation in diesel engines. J Dynam Syst Meas Control
2013;135:0110107. http://dx.doi.org/10.1115/1.4006884.
[22] Agarwal Deepak, Singh Shrawan Kumar, Agarwal Avinash Kumar. Effect of
Exhaust Gas Recirculation (EGR) on performance, emissions, deposits and
durability of a constant speed compression ignition engine. Appl Energy
2011;88:29007.
[23] Tesfa B, Mishra R, Gu F, Ball AD. Water injection effects on the performance
and emission characteristics of a CI engine operating with biodiesel. Renew
Energy 2012;37:33344.
[24] Tauzia Xavier, Maiboom Alain, Shah Samiur Rahman. Experimental study of
inlet manifold water injection on combustion and emissions of an automotive
direct injection diesel engine. Energy 2010;35:362839.
[25] Sahin Zehra, Tuti Mustafa, Durgun Orhan. Experimental investigation of the
effects of water adding to the intake air on the engine performance and
exhaust emissions in a DI automotive diesel engine. Fuel 2014;115:88495.
[26] Roubaud Anne, Favrat Daniel. Improving performances of a lean burn
cogeneration biogas engine equipped with combustion prechambers. Fuel
2005;84:20017.
[27] Makareviciene Violeta, Sendzikiene Egle, Pukalskas Saugirdas, Rimkus
Alfredas, Vegneris Ricardas. Performance and emission characteristics of
biogas used in diesel engine operation. Energy Convers Manage
2013;75:22433.
[28] Yamasaki Yudai, Kanno Masanobu, Suzuki Yoshitaka, Kaneko Shigehiko.
Development of an engine control system using city gas and biogas fuel
mixture. Appl Energy 2013;101:46574.
[29] Jingura RM, Matengaifa R. Optimization of biogas production by anaerobic
digestion for sustainable energy development in Zimbabwe. Renew Sustain
Energy Rev 2009;13:111620.
[30] Debabrata Barik, Murugan S. Performance and emission characteristics of a
biogas fueled DI diesel engine. In: SAE Paper 2013-01-2507; 2013.
[31] Yoon SH, Lee CK. Experimental investigation on the combustion and exhaust
emission characteristics of biogasbiodiesel dual-fuel combustion in a CI
engine. Fuel Process Technol 2011;92:9921000.
[32] Debabrata Barik, Murugan S. Investigation on performance and exhaust
emissions characteristics of a DI diesel engine fueled with Karanja methyl
ester and biogas in dual fuel mode. In: SAE Technical Paper 2014-01-1311;
2014.
[33] Maizonnasse Mark, Plante Jean-Sbastien, David Oh, Laamme Claude B.
Investigation of the degradation of a low-cost untreated biogas engine using
preheated biogas with phase separation for electric power generation. Renew
Energy 2013;55:50113.
[34] Chandra R, Vijay VK, Subbarao PMV, Khura TK. Performance evaluation of a
constant speed IC engine on CNG methane enriched biogas and biogas. Appl
Energy 2011;88:396977.
[35] Bedoya ID, Arrieta AA, Cadavid FJ. Effects of mixing system and pilot fuel
quality on dieselbiogas dual fuel engine performance. Bioresour Technol
2009;100:66249.
[36] Tippayawong N, Promwungkwa A, Rerkkriangkrai P. Long-term operation of a
small biogas/diesel dual-fuel engine for on-farm electricity generation. Biosyst
Eng 2007;98:2632.
[37] Bari S. Effect of carbon dioxide on the performance of biogas/diesel dual fuel
engine. Renew Energy 1996;9(14):100710.

[38] Duc Phan Minh, Wattanavichien Kanit. Study on biogas premixed charge
diesel dual fuelled engine. Energy Convers Manage 2007;48:2286308.
[39] Tonkunya Nutthapong, Wongwuttanasatian Tanakorn. Utilization of biogas
diesel mixture as fuel in a fertilizer pelletising machine for reduction of
greenhouse gas emission in small farms. Energy Sustain Develop
2013;17:2404.
[40] Musta Nirendra N, Raine Robert R, Verhelst Sebastian. Combustion and
emissions characteristics of a dual fuel engine operated on alternative gaseous
fuels. Fuel 2013;109:66978.
[41] Coleman HW, Steele Jr WG. Experimentation and uncertainty analysis for
engineers. New York: John Wiley & Sons; 1989.
[42] Ryu Kyunghyun. Effects of pilot injection timing on the combustion and
emissions characteristics in a diesel engine using biodieselCNG dual fuel.
Appl Energy 2013;111:72130.
[43] Satyanarayana M, Muraleedharan C. A comparative study of vegetable oil
methyl esters (biodiesels). Energy 2011;36:212937.
[44] Karim GA. Combustion in gas fueled compression ignition engines of the dual
fuel type. ASME J Eng Gas Turb Power 2003;125:82736.
[45] Patterson J, Clarke A, Chen R. Experimental study of the performance and
emissions characteristics of a small diesel Genset operating in dual-fuel mode
with three different primary fuels. In: SAE Paper No. 2006-01-0050; 2006.
[46] Papagiannakis RG, Hountalas DT. Combustion and exhaust emission
characteristics of a dual fuel compression ignition engine operated with
pilot diesel fuel and natural gas. Energy Convers Manage 2004;45:297187.
[47] Selim MYE. Sensitivity of dual fuel engine combustion and knocking limits to
gaseous fuel composition. Energy Convers Manage 2004;45:41125.
[48] Lakshmanan T, Nagarajan G. Experimental investigation of timed manifold
injection of acetylene in direct injection diesel engine in dual fuel mode.
Energy 2010;35:31728.
[49] Muralidharan K, Vasudevan D, Sheeba KN. Performance, emission and
combustion characteristics of biodiesel fuelled variable compression ratio
engine. Energy 2011;36:538593.
[50] Papagiannakis RG, Hountalas DT, Rakopoulos CD. Theoretical study of the
effects of pilot fuel quantity and its injection advance on the performance and
emissions of a dual fuel diesel engine. Energy Convers Manage
2007;48:295161.
[51] Ryu Kyunghyun. Effects of pilot injection pressure on the combustion and
emissions characteristics in a diesel engine using biodieselCNG dual fuel.
Energy Convers Manage 2013;76:50616.
[52] Bedoya ID, Saxena S, Cadavid FJ, Dibble RW, Wissink M. Experimental study of
biogas combustion in a HCCI engine for power generation with high indicated
efciency and ultra-low NOx emissions. Energy Convers Manage
2012;53:15462.
[53] Nathan SS, Mallikarjuna JM, Ramesh A. An experimental study of the biogas
diesel HCCI mode of engine operation. Energy Convers Manage
2010;51:134753.
[54] Al-Dawody MF, Bhatti SK. Optimization strategies to reduce the biodiesel NOx
effect in diesel engine with experimental verication. Energy Convers Manage
2013;68:96104.
[55] Papagiannakis RG, Rakopoulos CD, Hountalas DT, Rakopoulos DC. Emission
characteristics of high speed, dual fuel, compression ignition engine operating
in a wide range of natural gas/diesel fuel proportions. Fuel 2010;89:1397406.

You might also like