Sie sind auf Seite 1von 24

Int. J. Oil, Gas and Coal Technology, Vol. 12, No.

1, 2016 81

Experimental studies on variable compression ratio


engine fuelled with cottonseed oil methyl ester
biodiesel

M. Santhosh*
Department of Mechanical Engineering,
P.S.N.A College of Engineering and Technology,
Dindigul, Tamilnadu, India
Email: srinithims78@gmail.com
*Corresponding author

K.P. Padmanaban
SBM College of Engineering and Technology,
Dindigul, Tamilnadu, India
Email: padmarubhan@yahoo.co.in
Abstract: This research work investigates the effects of compression ratio on
engine performances, combustion characteristics and emission of exhaust gases
of variable compression ratio engine fuelled with biodiesel. The cottonseed oil
methyl ester is blended with standard diesel and used as fuel in the variable
compression ratio single cylinder four stroke multi fuel engine at 1,440 rpm of
80% load with different compression ratios i.e., 18:1, 19:1, 20:1, 21:1 and 22:1.
The analyses of engine performances, combustion characteristics and emission
of exhaust gases were carried out for different proportions of cottonseed oil
methyl ester blends (b15%, b30%, b45%, b60%, b75%, b90% and b100%).
The investigation reveals that at higher compression ratios, heat release rate,
combustion pressure and brake thermal efficiency is higher. The investigation
also reveals that the brake power and mechanical efficiency are found to have
increased with increase in compression ratios. It is also found that the higher
blend proportions with standard diesel reduce the emission of HC and CO2 but
increase the emission of NOX. [Received: November 5 2013; Accepted:
October 4 2014]
Keywords: variable compression ratio engine; brake thermal efficiency; BTE;
heat release rate; HRR; emission; cottonseed oil methyl ester; COME.
Reference to this paper should be made as follows: Santhosh, M. and
Padmanaban, K.P. (2016) ‘Experimental studies on variable compression ratio
engine fuelled with cottonseed oil methyl ester biodiesel’, Int. J. Oil, Gas and
Coal Technology, Vol. 12, No. 1, pp.81–104.
Biographical notes: M. Santhosh is working as an Associate Professor in the
Department of Mechanical Engineering, P.S.N.A College of Engineering and
Technology, Dindigul, Tamilnadu, India. He is doing research in the area of
biodiesel and its application in internal combustion engines.
K.P. Padmanapan is working as a Principal in SBM College of Engineering and
Technology, Dindigul, Tamilnadu, India. His research interests are fixture
design, finite element analysis, evolutionary techniques and internal
combustion engines.

Copyright © 2016 Inderscience Enterprises Ltd.


82 M. Santhosh and K.P. Padmanaban

1 Introduction

Fossil fuels are source of energy for industrialisation and their increased use causes
depletion and environmental effects all over the world. The use of alternative fuels such
as vegetable oils like sunflower oil, palm oil, corn oil, cottonseed oil, canola oil, rapeseed
oil, alcohols and petroleum gases, etc is comparatively higher than recent years (Donghui
et al., 2011; Basha et al., 2009). The process like transesterification is used to derive
biodiesel from vegetable oils. Biodiesel is free from sulphur and used directly in
compression ignition (CI) engines (Rao et al., 2010; Ozsezen and Canakci, 2011). The
presence of more amount of oxygen in biodiesel leads to higher brake thermal efficiency
(BTE) and reduced emission of exhaust gases (Ndayishimiye and Tazerout, 2011). Lin
et al. (2009) studied the engine performance and emission characteristics of exhaust gases
in diesel engine fuelled with vegetable oil methyl ester. Higher heat release rate (HRR)
was observed due to oxygen content of biodiesel and enhanced ignition quality. Ilkilic
(2009) tested a single cylinder diesel engine using biodiesel and standard diesel and
compared the results. Kalam et al. (2011) concluded that the emission of exhaust gases
hydrocarbon (HC), carbon monoxide (CO) and smoke decreased when using biodiesel in
diesel engine. The palm and coconut oil were used separately as fuel in diesel engine and
it was found that the engine power decreased for b5 compared to standard diesel.
Biodiesel derived from the cottonseed oil is used in diesel engine without any
modification in engine and produces better engine performance and emission
characteristics than conventional fuels. Change in compression ratio (CR) gives better
results in CI engine performance, combustion and emission characteristics. Karabektas
et al. (2008) and Keskin et al. (2008) studied the performance and emission
characteristics of diesel engine fuelled with cottonseed oil methyl ester (COME) as fuel
at full load conditions. The COME was preheated at different temperatures and higher
BTE with lower CO emission was obtained up to 90°C (Karabektas et al., 2008). When
the diesel engine fuelled with cotton oil soap-stock biodiesel-diesel blends the reduction
in engine power and torque was 6.2% and 5.8% respectively than diesel. Nabi et al.
(2009), He and Bao (2005) and Aydin and Bayindir (2010) investigated COME as a fuel
in a diesel engine. When the blend proportion of COME is increased, there is reduction in
emission of exhaust gases (CO, PM and smoke) with an increase in NOX emission.
Higher thermal efficiency was obtained for 30% blend of COME with standard diesel.
Raheman and Ghadge (2008) conducted an experiment on CI engine (IDI with the
pre-combustion chamber) by varying CR from 18 to 20 fuelled with mahua oil. It was
found that better results were obtained when there is an increase in CR. Jindal et al.
(2010) concluded that at CR 18:1, the performance is optimum for diesel engine fuelled
with jatropha methyl ester. Sayin and Gumus (2011) inferred the engine performance and
emission of exhaust characteristics of diesel engine by varying CR (17:1, 18:1 and 19:1).
It is evident that from the result the enhanced BTE was attained when increasing CR and
blend proportions of biodiesel with standard diesel. The literature reveals that the effects
of variation in CR on the BTE indicate that at higher CRs improve the engine efficiency.
Higher CR increased the NOx emissions and reduced the HC and CO emissions when
compared to the original CR.
Experimental studies on variable compression ratio engine fuelled 83

However, the above literature reveals that the effect of higher CRs (18:1 to 22:1) on
the engine performances, combustion characteristics and emission of exhaust gases of
variable compression ratio (VCR) engine using COME blended with diesel has not been
intensively studied. Hence, in this research work, an attempt is made to use the COME
blended with diesel in VCR multi fuel engine. The analysis is carried out to investigate
the effects of different CRs on the VCR engine using different proportions of COME
blends and compared with standard diesel.

2 Production of biodiesel

Transesterification is a most common and well recognised chemical reaction in which


alcohol reacts with triglycerides of fatty acids in existence of catalyst to form glycerol
and esters. One litre of refined cotton seed oil is taken in the beaker and heated to 70°C.
Sodium Hydroxide is used as the catalyst and methanol as solvent. Then the catalyst was
mixed with methyl alcohol to dissolve and added to heated cottonseed oil. After the
mixture was stirred, it was transferred to another container. Then the glycerine settles
down and methyl ester formed at the top. Then the distillation process is carried out at
about 110°C to remove the water contained in the esterified cottonseed oil. Allow the
COME to reach the room temperature and store it in a container. The various properties
of COME, standard diesel and COME blend with diesel at different proportions are
shown in Table 1. The elemental analysis of diesel and COME is shown in Table 2. The
typical values of the free fatty acid composition are shown on Table 3.
Table 1 Properties of Diesel and blends of COME with diesel

Test
Properties method Diesel B15 B30 B45 B60 B75 B90 B100
(ASTM)
Density (kg/l) D4052 0.85 0.854 0.855 0.855 0.86 0.86 0.865 0.870
Kinematic D445 4.0 4.11 4.19 4.23 4.27 4.31 4.36 4.39
viscosity at 40°C
(mm2/s)
Calculated cetane D4737 47 40 43 46 51 52 54 55
number (CCN)
Calorific value D240 42,500 42,417 42,364 42,190 41,548 40,905 39,523 38,523
(kJ/kg)
Flash point (°C) D93 60 62 64 66 68 70 120 185
Fire point (°C) - 69 70 72 74 76 78 138 205
84 M. Santhosh and K.P. Padmanaban

Table 2 Elemental analyses of diesel and COME

Elemental analysis (%W/W) Diesel Cottonseed oil methyl ester


C 83.03 82.11
H 13.02 11.83
O 0.190 05.98
N 1.760 0.042
S 0.025 0.004

Table 3 Free fatty acid composition of COME

Fatty acid composition Weight %


Palmitic 16:0 28
Stearic 18:0 01
Oleic 18:1 13
Linoleic 18:2 58

Table 4 Engine specifications

Make Legion brothers


BHP 3 to 5 Hp
Speed 1,400 to 1,600 rpm
Number of cylinders 1
Compression ratio 5:l to 22:1
Bore 80 mm
Stroke 110 mm
Type of ignition Compression/spark ignition
Method of loading Eddy current dynamometer
Method of starting Manual crank start
Method of cooling Water cooled

3 Experimental apparatus and procedure

3.1 Experimental set up


The experiments were conducted on a four-stroke, naturally aspirated, water cooled,
single cylinder VCR multi fuel engine. Details of the engine specifications were shown in
Table 4 and test bench set up is shown in Figures 1(a) and (b). An air cooled eddy current
dynamometer has been directly coupled to the engine output shaft using a tyre coupling.
The engine and air cooled eddy current dynamometer are coupled using a tyre coupling
while the output shaft of the dynamometer is fixed to a strain gauge type load cell for
measuring the load applied to the engine. The fuels from the tank is connected through
solenoid valve to the glass burette and then to the engine through manual ball valve. The
levels of the fuel (low, high) in the burette are sensed by using optical lot sensor and the
timing of the events are recorded in the computer. The fuel consumption is measured
Experimental studies on variable compression ratio engine fuelled 85

along the test run. The 11 bit 2050 step crank angle (CA) encoder is mounted on the
camshaft to measure engine CA. A shell tube gas to liquid heat exchanger is used as
calorimeter for conducting the heat balance. The emission of exhaust gas is measured by
flue gas analyser. The CR is varied by using raising the bore and head of the engine. As
the bore and the head of the engine are raised and lowered, the clearance volume is
changed which effects variation in CR. The exhaust gas emissions were measured from
VCR engine by MARS flue gas analyser.

Figure 1 (a) Schematic diagram of experimental setup (b) VCR engine experimental setup
(see online version for colours)

(a)

(b)

The VCR engine consists of combustion analyser kit developed by Legion brothers,
includes software, combustion pressure sensor and CA encoder. The combustion pressure
sensor is mounted on the cylinder head. Optimised piezoelectric sensor with sensitivity of
±0.5% is used for continuous monitoring of combustion pressure with average for
86 M. Santhosh and K.P. Padmanaban

20 working cycles. The CA encoder consists of 14 analogue inputs (12–16 bits depending
on speed) with rotational speed of maximum 12,000 min–1 and connector type of 9416-12
pin. Windows based software is used for data measurement, auto zoom graphs and digital
display of data in computer, stored indefinite number of graphs with facilities export data
to Microsoft excel. These system capabilities are to calculate P-θ diagram, mass fraction
burned angle, HRR and combustion pressure rise with stored data in excel file. The
calorific value and viscosity are measured by bomb calorimeter and redwood viscometer
respectively. Calculated cetane number is determined by measuring the aniline point. The
acrylic body rota meter is used for the measurement of water flow measurement with
range 40–400 LPH for engine cooling and 10–100 LPH for calorimeter cooling. K type
sensors with a range of 0–300°C are used to measure the inlet and outlet gas temperature
from calorimeter, inlet and outlet water temperature from the cylinder. Another, K type
sensor with a range of 0–1,500°C is used to measure the temperature of exhaust gases at
inlet and outlet of calorimeter. For the measurement of ambient temperature also
standalone K type sensor with range air 0–300°C is used. All the measured parameters
from the sensor are connected to the computer. A MN-05 model, MARS portable gas
analyser is used for measuring the exhaust gas emissions. This analyser is able to
measure the emission of NOX, CO, CO2, O2 and HC. The probe of the analyser is inserted
into the exhaust pipe of the engine before taking the measurements. In this analyser the
zero setting function sets the sensor to zero using span gas. On a timed basis the analyser
first being turned ON a zero will be requested automatically at 7, 15 and then 25 minute
time intervals. Subsequent request will be for every 25 minutes. The zero air port allows
the gas analyser to zero without removing the sampling probe from the exhaust tail. The
ambient air from the zero port drawn through a charcoal filter helps in setting HC to zero.
The zero key on the keyboard enables to set zero for CO, CO2 and NOX values to zero
with ambient air and is calibrated to 20.9% oxygen by volume. The computerised data
acquisition system is used to collect the data which can be analysed for engine
performance, combustion characteristics and emission of exhaust gases.
Table 5 Percentage uncertainties of measurements

Measurements Accuracy
Engine speed ±3 rpm
Specific fuel consumption ±2%
Power ±2%
Temperatures ±1°C
Time ±0.6%
Crank angle encoder ±0.5°CA
Combustion pressure ±0.5 bar
Heat release rate ±0.05J/CA
Mass fraction burned ±0.5%
Piezoelectric sensor ±0.5%
Experimental studies on variable compression ratio engine fuelled 87

3.2 Experimental procedures


The experiments were conducted at four different CRs i.e., 18:1, 19:1, 20:1, 21:1 and
22:1. All tests were conducted at 80%load of the engine. The initial tests were conducted
with standard diesel for different CRs to obtain the base data of the engine. Then, by
using different blends of COME (b15, b30, b45, b60, b75, b90 and b100), engine
performances, combustion characteristics and emission of exhaust gases are
experimentally obtained and the results are compared with that of standard diesel. The
emission of exhaust gases is measured by MN-05/MARS flue gas analyser. The exhaust
gas emissions HC, CO and NOX have been measured. The accuracies of the
measurements for different parameters are shown in Table 5. The specification of exhaust
gas analyser is given in Table 6.
Table 6 Specification of exhaust gas analyser

Gas measured Measuring range Resolution


CO 0–10% Vol 0.01%
CO2 0–20% Vol 0.1%
HC 0–15,000 ppm 1 ppm
O2 0–25% Vol 0.01%
NOx 0–10,000 ppm 1 ppm

4 Results and discussions on engine performance

4.1 Theoretical aspects of blends combustion in the engine


In the atomisation process, the mean droplet size is larger when bio-diesel is used against
the diesel fuel case. This is mainly due to the bio-diesel having a higher kinematic
viscosity than that of the neat diesel fuel. The evaporation process will be slower for the
bio-diesel and thus could affect the combustion process. Taking into account that the
stoichiometric air–fuel ratio is on the order of 12.5 with the neat bio-diesel and of 15 with
the neat diesel fuel, this is translated into that the air/bio-diesel oil mixtures can reach the
stoichiometric conditions nearly 15% faster than the air/diesel fuel mixture. The higher
cetane number of the bio-diesel, as compared to the neat diesel fuel case, can decrease the
NOx emissions due to the relatively lower ignition delay and thus shorter premixed
combustion during which NOx is mainly formed. However, all these may be obscured by
the delicate distribution of the fuel–air ‘packets’ inside the sprays, given that NOx
production is favoured by high local temperatures and near to stoichiometric local
conditions. The lower calorific value of the bio-diesel compared to that for the neat diesel
fuel, at first sight, requires higher fuel rates against the respective neat diesel fuel case,
given that the corresponding comparison must be effected at the same load, i.e., brake
mean effective pressure.
88 M. Santhosh and K.P. Padmanaban

4.2 Brake thermal efficiency


The BTE for varied CRs while using COME blends and standard diesel are found and
shown in Figure 2. The BTE values dependency on CR and blend ratios are given in
Table 7. The effect of CR indicates that BTE increases with an increase in CR for
standard diesel and blends of COME with standard diesel. The increase in BTE for b30 at
CR 19:1, 20:1, 21:1 and 22:1 are 2.51%, 3.12%, 7.45% and 9.61% respectively. This is
due to the lower volatility, reduced ignition delay and higher temperature at higher CR
which enhances the combustion. The maximum BTE obtained for the blend b15 at CR
22:1 is 0.4218. By increasing the temperature of the fluid, the intermolecular attraction
between different layers of the fluid decreases, thus viscosity decreases. The reduction in
viscosity leads to improved atomisation, fuel vaporisation and combustion. It may also be
due to better utilisation of heat energy and better air entrainment. For blend b15 and b75
better results are observed among the other blends and diesel. For blend b15 the
percentage increase in BTE for CR 19:1, 20:1, 21:1 and 22:1 are 1.35%, 4.13%, 5.96%
and 12.24% respectively as compared to CR 18:1. For blend b75, the percentage increase
in BTE for CR 22:1 is 6.14% as compared to CR 18:1. This can be attributed to better
combustion at higher CR and lubricity of biodiesel (Sayin and Gumus, 2011). The
reduction in the viscosity enhances atomisation and fuel vaporisation (Devan and
Mahalakshmi, 2009). By increasing the temperature of the fluid, the intermolecular
attraction between different layers of the fluid decreases, thus viscosity decreases. For
proper functioning of the engine it is necessary to reduce the viscosity of a fuel. Fuel with
relatively higher viscosity will not break into fine particles when sprayed. Large particles
will burn slowly resulting in poor engine performance. The improved BTE is obtained by
increasing the blend proportion of biodiesel at all CRs as the lubricity of the biodiesel is
higher (Buyukkaya, 2010; Yilmaz and Morton, 2011; Rakopoulos et al., 2008). The
increment in BTE for b15, b30, b45, b60, b75, b90 and b100 at CR 18:1 are 3.87%,
1.65%, 1.66%, 1.85%, 3.57%, 1.11% and 1.79% respectively compared to standard
diesel. This is due to oxygen content in the biodiesel which leads to the improved
combustion. The enhancement in BTE is obtained at CR 21:1 with the increase of blend
proportions due to the better combustion. It is evident from the figure that the enhanced
BTE is obtained with the increase of CR and blend proportions.

Figure 2 Comparison of BTE for different CRs


Experimental studies on variable compression ratio engine fuelled 89

Table 7 BTE results dependency on CR and blend conditions

Diesel
CR b15 (%) b30 (%) b45 (%) b60 (%) b75 (%) b90 (%) b100 (%)
(%)
18 36.17421 37.57651 36.23193 36.77240 36.84544 37.46853 36.57273 36.82350
19 37.15784 38.07974 37.13675 37.30563 36.71120 37.43978 36.65077 37.15953
20 37.90760 39.11611 37.36452 36.94092 36.84312 37.26929 36.98301 38.11236
21 37.45104 39.81099 38.92850 37.49010 37.47015 37.00943 37.17156 38.65506
22 39.78810 42.17942 39.70643 37.88535 37.87245 39.76231 37.87178 38.69348

4.3 Brake specific fuel consumption

The BSFC for different CRs and different proportions of blends and for the standard
diesel is shown in Figure 3. The BSFC is found to be higher with increasing blend
proportions of biodiesel. The increments in BSFC at CR 21:1 for b15, b30, b45, b60, b75,
b90 and b100 are 7.83%, 11.66%, 14.81%, 15.25%, 18.49%, 20.98% and 20.64%
respectively. The lower heating value of the biodiesel blends which leads the more
amount of fuel injected into the cylinder to obtain the same power output (Bueno et al.,
2011; Wander et al., 2011). The increase in CR decreases the BSFC for b60 by 0.82%,
3.29%, 4.94% and 6.17% for CR 19:1, 20:1, 21:1 and 22:1 respectively. The reason is
that when the CR increased cylinder pressure increases which lead to higher power and
thereby decreased BSFC is obtained (Ulusoy et al., 2004). For blend b75 the higher
BSFC is obtained at CR 21:1 as compared to 22:1. The possible reason for this trend
could be that decrease in maximum cylinder pressure attained in cylinder (Sayin and
Gumus, 2011). That is for b75 at CR 21:1 the decrease in maximum cylinder pressure is
obtained when compared to other blend proportions. Therefore, fuel consumption per
output power will increase.

Figure 3 Comparison of brake specific fuel consumption for different CRs


90 M. Santhosh and K.P. Padmanaban

4.4 Brake power and mechanical efficiency


The BP for different CRs and different proportions of blends and standard diesel is found
and shown in Figure 4. The BP is found to have increased when CR and blend
proportions of biodiesel increases. This is due to higher oxygen content of the biodiesel
(Song and Zhang, 2008). The increase in BP at CR 22:1 for b15, b30, b45, b60, b75, b90
and b100 are 0.13%, 0.36%, 0.46%, 0.60%, 0.71%, 0.78% and 0.83% respectively. The
mechanical efficiency for different CRs and different proportions of blends and for
standard diesel is found and shown in Figure 5. From the figure, it is evident that
mechanical efficiency increases with increasing blend proportions at all CRs. The
increase in mechanical efficiency at CR 20:1 for b15, b30, b45, b60, b75, b90 and b100
are 1.42%, 1.46%, 2.23%, 2.38%, 2.96%, 3.25% and 3.43% respectively (Muralidharan
and Vasudevan, 2011).

Figure 4 Comparison of brake power for different CRs

Figure 5 Comparison of mechanical efficiency for different CRs


Experimental studies on variable compression ratio engine fuelled 91

4.5 Discussion
The BTE was increased when increasing CR 18:1 to 22:1. Similar result was obtained
when increasing CR from 18:1 to 20:1 the BTE of the diesel engine (IDI with pre
combustion chamber) was increased by 33% (Raheman et al., 2008). Higher BTE is
obtained by Buyukkaya (2010) for blends b20, b50, b70, b100 at 2,000 rpm are 0.427,
0.425, 0.424 and 0.423 respectively. The BSFC is found to be higher with increasing
blend proportions of biodiesel. Similar trend was obtained by Buyukkaya (2010) at 1,400
rpm. BSFC for b5, b20, b70 and b100 fuels were observed to be higher by 2.5%, 3%,
5.5% and 7.5% respectively than that of diesel fuel. Also the result obtained by Raheman
and Ghadge (2008) reveals that the mean BSFCs of high speed diesel, b20, b40, b60, b80
and b100 were found to be 0.421, 0.437, 0.478, 0.503, 0.539 and 0.583 kg/kWh
respectively. The increase in CR decreases the BSFC, similar trend was obtained by
Gogoi and Baruah (2011) i.e., decrease in BSFC at full load condition is 0.36, 0.343,
0.337, 0.326, 0.323 and 0.317 kg/kWh for CR from 13:1 to 18:1 respectively.

5 Combustion analysis

5.1 Maximum combustion pressure


The combustion pressure with respect to CA for different CRs and different proportions
of blends and for standard diesel is found and shown in Figures 6–10. It is evident that
the combustion pressure, obtained using COME for different blends and standard diesel
is almost equal. The maximum combustion pressure is found to be comparatively higher
for the b75 at CR 18:1. This is due to longer ignition delay which leads to more amount
of fuel admitted in cylinder (Puhan et al., 2009). That is the maximum combustion
pressure obtained at CR 18:1 for diesel, b15,b30,b45,b60,b75,b90 and b100 are 63.18,
62.85, 61.55, 64.11, 63.06, 65.27, 62.87 and 62.11 bar respectively. For diesel the
pre-combustion phase is very intense compared to COME and its blends, leading to a
high pressure rise, as more fuel is accumulated during the delay period. The high
premixed combustion rate can be attributed to the high value of cylinder pressure at high
loads. The increase in temperature reduces the viscosity of the oil by breaking down the
intermolecular bonds and increases the possibility of ignition and hence ignition delay is
reduced (Lakshmi Narayana Rao et al., 2007). It is also evident that combustion pressure
decreases for b15 at CR 18:1 due to the reduced ignition delay and lower calorific value
of biodiesel (Puhan et al., 2009; Lakshmi Narayana Rao et al., 2007). The increase in
combustion pressure obtained for b45, b75 are 1.47% and 3.31% respectively when
compared to diesel. Higher combustion pressure is obtained for blend b60 at CR 21:1 (as
shown in Figure 9), that is the percentage increase in combustion pressure for blend b60
is 5.91% as compared to diesel. This may be due to the increase in the flash and fire point
of the blend this may lead to the shorter ignition delay that increased the combustion
pressure. The combustion pressure increases for all blends and even for standard diesel as
there is an increase in CR due to complete combustion of fuel inside combustion
chamber.
92 M. Santhosh and K.P. Padmanaban

Figure 6 Comparison of combustion pressure and CA data at CR 18:1

Figure 7 Comparison of combustion pressure and CA data at CR 19:1

Figure 8 Comparison of combustion pressure and CA data at CR 20:1


Experimental studies on variable compression ratio engine fuelled 93

Figure 9 Comparison of combustion pressure and CA data at CR 21:1

Figure 10 Comparison of combustion pressure and CA data at CR 22:1

5.2 Heat release rate


The HRR has been calculated and verified through the repeatability analysis using the
factorial design in the design of experiments. The five test runs are repeated at the same
test conditions and the observations are averaged. It has been noticed that as the CR
increases, the HRR for the diesel decreases. This may be due to the air entrainment and
lower air/fuel mixing rate and effect of viscosity of the blends. The HRR with respect to
CA for different CRs and for different proportions of blends and for the standard diesel is
shown in Figures 11–15. The maximum HRR obtained from graph for diesel at CR 18:1,
19:1, 20:1, 21:1 and 22:1 are 17.71, 15.41, 14.79, 13.73 and 12.53 J/deg respectively. The
similar trend is obtained for b100 at CR 18:1, 19:1, 20:1, 21:1 and 22:1 are 14.27, 13.78,
13.27, 12.46 and 12.09 J/deg respectively. That is HRR decreases slightly when the CR is
increased due to the air entrainment and lesser air-fuel mixing in combustion chamber
(Muralidharan and Vasudevan, 2011). It is also observed that the HRR obtained for
COME is lower than that of standard diesel due to shorter ignition delay (Lakshmi
94 M. Santhosh and K.P. Padmanaban

Narayana Rao et al., 2007). The figures imply that HRR obtained for b90 is lower among
the blends. That is the HRR attained for diesel, b15, b30, b45, b60, b75, b90 and b100 are
17.71, 17.18, 16.98, 13.71, 16.48, 15.51, 13.24 and 14.27 J/deg respectively at CR 18:1.
This is due to reduced viscosity and lower heating value of the fuel (Shehata and Abdel
Razek, 2011). It is also found that the HRR which is obtained for b15 at CR 18:1 is
slightly higher than the HRR of standard diesel due to the oxygen in COME and longer
delay period which results complete combustion (Haik et al., 2011).

Figure 11 Comparison of HRR and CA data at CR 18:1

Figure 12 Comparison of HRR and CA data at CR 19:1


Experimental studies on variable compression ratio engine fuelled 95

Figure 13 Comparison of HRR and CA data at CR 20:1

Figure 14 Comparison of HRR and CA data at CR 21:1

Figure 15 Comparison of HRR and CA data at CR 22:1


96 M. Santhosh and K.P. Padmanaban

5.3 Mass fraction burned


The mass fraction burned with respect to CA for different CRs, different proportions of
blends and for the standard diesel is found and shown in Figures 16–20. It is found that
the mass fraction burned increases with increase in blend proportions at CR 18:1. The
higher mass fraction burned is obtained at CR 20:1 due to supplementary oxygen content
of blends and thereby sustains combustion. It is evident from the graph that the dip in the
mass fraction at about 360°CA might be due to the combustion lagging of the mass
fraction. It is also found that mass fraction burned for all proportions of blends is lower
than that of standard diesel at CR 22:1 (Arul Mozhi Selvan et al., 2009). From the figure,
it is evident that mass fraction burned is slightly higher for b75 when compared with
other blends and standard diesel at CRs 18:1, 19:1 and 20:1.

Figure 16 Comparison of mass fraction burned and CA data for CR 18:1

Figure 17 Comparison of mass fraction burned and CA data for CR 19:1


Experimental studies on variable compression ratio engine fuelled 97

Figure 18 Comparison of mass fraction burned and CA data for CR 20:1

Figure 19 Comparison of mass fraction burned and CA data for CR 21:1

Figure 20 Comparison of mass fraction burned and CA data for CR 22:1


98 M. Santhosh and K.P. Padmanaban

5.4 Exhaust gas temperature


The EGT for different CRs, different proportions of blends and for standard diesel is
found and shown in Figure 21. It is found that EGT decreases with increase in blend
proportions at higher CRs. At CR 21:1 EGT reduced by 8.17%, 10.49%, 11.48%,
11.43%, 11.64%, 12.19% and 12.52% for b15, b30, b45, b60, b75, b90 and b100
respectively. It is also found that at CRs 18:1 and 19:1, EGT is higher for COME than
standard diesel due to its lower cetane number. That is longer ignition delay of biodiesel
due to the lower cetane number which contributes as slower burn rate and higher exhaust
gas temperature (Nwafor et al., 2000). For b15 the EGT decreased by 0.75%, 1.39%,
2.07% and 4.59% for CR 19:1, 20:1, 21:1 and 22:1 due to reduction of ignition lag which
leads to better and complete combustion due to the increase in cylinder air temperature.
The difference between the diesel and blend b15 is due to the lower calorific value of
blend b15 when compared to diesel, which reduces the total energy released and hence
lowering the exhaust temperature.

Figure 21 Variation of exhausts gas temperature for different CRs

Figure 22 Variation of NOX emission for different CRs


Experimental studies on variable compression ratio engine fuelled 99

6 Emission analysis

6.1 Nitrogen oxide emissions


The emission of NOx for different CRs, different proportions of blends and for standard
diesel are measured and shown in Figure 22. It is found that the emission of NOx
increases with an increase in CR for all blend proportions. The higher values for diesel at
higher CR are due to higher in-cylinder temperature attained during combustion. The
higher NOx emission was observed for diesel at CR 20:1 during the testing of the engine
and this is due to the calorific value of fuel. The increase in NOX emission for b100 at CR
19:1, 20:1, 21:1 and 22:1 are 2.28%, 3.84%, 5.84% and 12.1% respectively. This is due
to the increase in combustion temperature which results in better combustion of fuel
(Sureshkumar et al., 2008). The increase in combustion temperature is due to the oxygen
content of the biodiesel. That is NOX emission increase with the mass percent of oxygen
in fuel (Park et al., 2012; Baiju et al., 2009; Raslavicius and Bazaras, 2010a). The figure
also implies that the emission of NOx with increase in blend proportions of COME is
higher than that of standard diesel (Karabektas et al., 2008). The decrement in emission
of NOX at CR 22:1 with increase in blend proportions are 12.91%, 11.07%, 13.65%,
9.59%, 6.64%, 7.01% and 0.48% for b15, b30, b45, b60, b75, b90 and b100 respectively
when compared to diesel. This is due to reduction of ignition lag which leads to better
and complete combustion. This higher temperature is the cause for increase in the
emission of NOX (Kwanchareon et al., 2007).

Figure 23 Variation of HC emission for different CRs

6.2 HC emission
The emission of HC for different CRs and different proportions of blends and for the
standard diesel is found and shown in Figure 23. It is found that the emission of HC
decreases with an increase in blend proportions and CRs (Lapuerta et al., 2008). The
decrease in emission of HC is obtained for b15 at CR 19:1, 20:1, 21:1 and 22:1 are
21.62%, 32.43%, 40.54% and 48.65% respectively when CR is increased. The higher
emission of HC is obtained at lower CR due to ignition delay and flame quenching in
100 M. Santhosh and K.P. Padmanaban

cylinder wall which leads to incomplete combustion (Karabektas et al., 2008; Sayin and
Gumus, 2011). From the figure, it is evident that at CR 22:1, the decrease in emission of
HC is obtained for b15, b30, b45, b60, b75, b90 and b100 are 26.08%, 30.43%, 52.17%,
56.52%, 61.30%, 65.65% and 69.13% respectively, when compared with standard diesel.
This is due to the over rich mixture during biodiesel addition (Sayin and Gumus, 2011).

6.3 CO emissions
The emission of CO for different CRs and different proportions of blends and for
standard diesel is predicted and shown in Figure 24. It is found that the emission of CO
decreases with an increase in CR. The CO emission is decreased for diesel fuel at CR
19:1, 20:1, 21:1, 22:1 are 7.8%, 15.10%, 17.19% and 38.02% respectively. Similar result
is obtained for b100 at CR 19:1, 20:1, 21:1, 22:1 are 4.13%, 19.01%, 57.02% and 52.07%
respectively. This is due to higher temperature in combustion chamber reduces ignition
lag and complete combustion is achieved in cylinder (Park et al., 2012). When compared
to standard diesel, the emission of CO decreases with an increase in blend proportions at
all CRs (Ghorbani et al., 2011). The emission of CO decreased by 18.48%, 26.05%,
30.25% and 35.29%, 36.13%, 49.57% and 51.26% at CR 22:1 for b15, b30, b45, b60,
b75, b90 and b100 respectively. This is due to the oxygen content of the biodiesel which
promotes the combustion process. It is evident from Figure 24 that complete combustion
occurs when CO has a complete oxidation (Godiganur et al., 2010; Raslavicius et al.,
2010b).

Figure 24 Variation of co emission for different CRs

7 Conclusions

The engine performance, combustion characteristics and emission of exhaust gases of a


VCR engine fuelled with different proportions of COME blends and standard diesel are
experimentally investigated for different CRs i.e., 18:1, 19:1, 20:1, 21:1 and 22:1. The
conclusions are:
Experimental studies on variable compression ratio engine fuelled 101

1 influences on engine performance


• the BTE is found to be increased with an increase in CR. At CR 22:1 the
maximum BTE obtained is 0.4218 for blend b15. The percentage increment in
BTE obtained for b15 is 12.24% at CR 22:1 when compared to CR 18:1. The
BTE is found to have increased with higher blend proportions as the lower
heating value of COME blends results in the consumption of more fuel. The
BSFC is found to be decreased with an increase in CR. The BP and mechanical
efficiency are found to be increased with an increase in CR.
2 Influences on combustion characteristics
• The enhancement in maximum combustion pressure is obtained as CR increases
for all fuels. The increase in maximum combustion pressure for diesel at CR
19:1, 20:1, 21:1, 22:1 are 1.64%, 4.72%, 7.96% and 14.71% respectively when
compared with CR 18:1. The HRR increases with an increase in CR which
enhances the density of air in the cylinder during combustion which enhances
the combustion pressure and thereby absolute combustion is achieved. The
HRR, which is slightly higher than that of standard diesel, is obtained for the
b15 at CR 22:1, as the delay period of COME blend increases when compared to
that of standard diesel. The enhancement in mass fraction burned is obtained as
the blend proportion increases for all fuels. The mass fraction burned at CR
18:1, 19:1, 20:1, 21:1, 22:1 are 20.36%, 17.32%, 13.58%, 12.47%, 51.42%
respectively for diesel.
• The EGT for diesel, b15, b30, b45, b60, b75, b90, b100 are 270.48°C, 260.34°C,
258.47°C, 231.78°C, 251.69°C, 251.75°C, 253.12°C, 259.09°C respectively at
CR 20:1. The reason for the reduction in exhaust gas temperature at CR 20:1 is
due to the lower calorific value of blended fuel as compared to the standard
diesel and lower temperature, at the end of compression.
3 Influences on emission of exhaust gases
• The NOX emission increases with the increase of CR. At higher CR the
decrement in NOX emission is obtained for blends when compared to diesel but
the reverse trend is obtained at lower CR. At CR 22:1 the decrement in NOX
emission obtained for b15 is 12.91% when compared to diesel.
• The emission of HC decreases with an increase in blend proportion and it is
lesser than that of standard diesel at CRs 18:1 and 19:1. At CR 22:1 the decrease
in HC emission obtained for b15 is 48.65% when compared to CR 18:1.
• The emission of CO decreases with an increase in blend proportion and it is
lesser than that of standard diesel at CRs 19:1 and 21:1. At CR 22:1 the decrease
in CO emission obtained for b15 is 18.48% when compared to CR 18:1.
From the above observations, it has been found that the performance of the b15 blend is
better when compared with the conventional standard diesel at CR 22:1. The
experimental result reveals that lower and medium percentages of COME can be
substituted as diesel fuel.
102 M. Santhosh and K.P. Padmanaban

Acknowledgements

The authors thank the All India Council for Technical Education (AICTE), Government
of India for providing grant (NO.8024/RID/MOD/70/08/09) under Modernisation and
Removal of Obsolescence (MODROB) scheme for providing matching grand for the
purchase of variable CR multi fuel engine test rig.

Nomenclature
b100 100% biodiesel
b15 15% biodiesel + diesel
b30 30% biodiesel + diesel
b45 45% biodiesel + diesel
b60 60% biodiesel + diesel
b75 75% biodiesel + diesel
b90 90% biodiesel + diesel
BP Brake power (kW)
BSFC Brake specific fuel consumption (kg/kWhr)
BTE Brake thermal efficiency (%)
CA Crank angle (deg)
CO Carbon monoxide (g/kWhr)
COME Cottonseed oil methyl ester
CR Compression ratio
EGT Exhaust gas temperature (oC)
HC Hydro carbon (g/kWhr)
HRR Heat release rate (J/CA)
NOX Nitrogen oxides (g/kWhr)
VCR Variable compression ratio

References
AliKeskin, M.G., Altiparmak, D. and Aydin, K. (2008) ‘Using of cotton oil soap-stock
biodiesel-diesel fuel blends as an alternative diesel fuel’, Renewable Energy, Vol. 33, No. 4,
pp.553–557.
Arul Mozhi Selvan, V., Anand, R.B. and Udayakumar, M. (2009) ‘Combustion characteristics of
diesohol using bio diesel as an additive in a direct injection ignition engine under various
compression ratios’, Energy and Fuels, Vol. 23, No. 11, pp.5413–5422.
Aydin, H. and Bayindir, H. (2010) ‘Performance and emission analysis of cottonseed oil methyl
ester in a diesel engine’, Renewable Energy, Vol. 35, No. 3, pp.588–592.
Baiju, B., Naik, M.K. and Das, L.M. (2009) ‘A comparative evaluation of compression ignition
engine characteristics using methyl and ethyl esters of Karanja oil’, Renewable Energy,
Vol. 34, No. 6, pp.1616–1621.
Experimental studies on variable compression ratio engine fuelled 103

Basha, S.A., Gopal, K.R. and Jebaraj, S. (2009) ‘A review on biodiesel production, combustion,
emissions and performance’, Renewable & Sustainable Energy Reviews, Vol. 13, Nos. 6–7,
pp.1628–1634.
Bueno, A.V., Velasquez. J.A. and Milanez, L.F. (2011) ‘Heat release and engine performance
effects of soybean oil ethyl ester blending into diesel fuel’, Energy, Vol. 36, No. 6,
pp.3907–3916.
Buyukkaya, E. (2010) ‘Effects of biodiesel on a DI diesel engine performance, emission and
combustion characteristics’, Fuel, Vol. 89, No. 10, pp.3099–3105.
Devan, P.K. and Mahalakshmi, N.V. (2009) ‘A study of performance, emission and combustion
characteristics of a compression ignition engine using methyl ester of paradise oil – eucalyptus
oil blends’, Applied Energy, Vol. 86, No. 5, pp.675–680.
Donghui, Q., Leick, M., Liu, Y. and Lee, C-F.F. (2011) ‘Effect of EGR and injection timing on
combustion and emission characteristics of split injection strategy DI-diesel engine fuelled
with biodiesel’, Fuel, Vol. 90, No. 5, pp.1884–1891.
Ghorbani, A., Bazooyar, B., Shariati, A., Jokar, S.M., Ajami, H. and Naderi, A. (2011)
‘A comparative study of combustion performance and emission of biodiesel blends and diesel
in an experimental boiler’, Applied Energy, Vol. 88, No. 12, pp.4725–4732.
Godiganur, S., Murthy, C.S. and Reddy, R.P. (2010) ‘Performance and emission characteristics of a
Kirloskar HA394 diesel engine operated on fish oil methyl esters’, Renewable Energy,
Vol. 35, No. 2, pp.355–359.
Gogoi, T.K. and Baruah, D.C. (2011) ‘The use of Koroch seed oil methyl ester blends as fuel in a
diesel engine’, Applied Energy, Vol. 88, No. 8, pp.2713–2725.
Haik, Y., Selim, M.Y.E. and Abdulrehman, T. (2011) ‘Combustion of algae oil methyl ester in an
indirect injection diesel engine’, Energy, Vol. 36, No. 3, pp.1827–1835.
He, Y. and Bao, Y.D. (2005) ‘Study on cottonseed oil as a partial substitute for diesel oil in fuel for
single-cylinder diesel engine’, Renewable Energy, Vol. 30, No. 5, pp.805–813.
Ilkilic, C. (2009) ‘Emission characteristics of a diesel engine fuelled by 25% sunflower oil methyl
ester and 75% diesel fuel blend’, Energy Sources, Part A: Recovery, Utilization, and
Environmental Effects, Vol. 31, No. 6, pp.480–491.
Jindal, S., Nandwana, B.P., Rathore, N.S. and Vashistha, V. (2010) ‘Experimental investigation of
the effect of compression ratio and injection pressure in a DI diesel engine running on
Jatropha methyl ester’, Applied Thermal Engineering, Vol. 30, No. 5, pp.442–448.
Kalam, M.A., Masjuki, H.H., Jayed, M.H. and Liaquat, A.M. (2011) ‘Emission and performance
characteristics of an indirect ignition diesel engine fuelled with waste cooking oil’, Energy,
Vol. 36, No. 1, pp.397–402.
Karabektas, M., Ergen, G. and Hosoz, M. (2008) ‘The effects of preheated cottonseed oil methyl
ester on the performance and exhaust emissions of a diesel engine’, Applied Thermal
Engineering, Vol. 28, Nos. 17–18, pp.2136–2143.
Kwanchareon, P., Luengnaruemitchai, A. and Jai-In, S. (2007) ‘Solubility of a diesel-bio-diesel-
ethanol blend, its fuel properties, and its emission characteristics from diesel engine’, Fuel,
Vol. 86, Nos. 7–8, pp.1053–1061.
Lakshmi Narayana Rao, G., Durga Prasad, B., Sampath, S. and Rajagopal, C.K. (2007)
‘Combustion analysis of diesel engine fuelled with Jatropha oil methyl ester – diesel blends’,
International Journal of Green Energy, Vol. 4, No. 6, pp.645–658.
Lapuerta, M., Armas, O. and Rodriguez-Fernandez, J. (2008) ‘Effect of biodiesel fuels on diesel
engine emissions’, Progress in Energy and Combustion Science, Vol. 34, No. 2, pp.198–223.
Lin, B-F., Huang, J-H. and Huang, D-Y. (2009) ‘Experimental study of the effects of vegetable oil
methyl ester on DI diesel engine performance characteristics and pollutant emissions’, Fuel,
Vol. 88, No. 9, pp.1779–1785.
Muralidharan, K. and Vasudevan, D. (2011) ‘Performance, emission and combustion
characteristics of a variable compression ratio engine using methyl esters of waste cooking oil
and diesel blends’, Applied Energy, Vol. 88, No. 11, pp.3959–3968.
104 M. Santhosh and K.P. Padmanaban

Nabi, M.N., Rahman, M.M. and Akhter, M.S. (2009) ‘Biodiesel from cotton seed oil and its effect
on engine performance and exhaust emissions’, Applied Thermal Engineering, Vol. 29,
Nos. 11–12, pp.2265–2270.
Ndayishimiye, P. and Tazerout, M. (2011) ‘Use of palm oil-based biofuel in the internal
combustion engines: performance and emissions characteristics’, Energy, Vol. 36, No. 3,
pp.1790–1796.
Nwafor, O.M.I., Rice, G. and Ogbonna, A.I. (2000) ‘Effect of advanced injection timing
on the performance of rapeseed oil in diesel engines’, Renewable Energy, Vol. 21, Nos. 3–4,
pp.433–444.
Ozsezen, A.N. and Canakci, M. (2011) ‘Determination of performance and combustion
characteristics of a diesel engine fuelled with canola and waste palm oil methyl esters’, Energy
Conversion and Management, Vol. 52, No. 1, pp.108–116.
Park, S.H., Cha, J. and Lee, C.S. (2012) ‘Impact of biodiesel in bioethanol blended diesel on the
engine performance and emissions characteristics in compression ignition engine’, Applied
Energy, Vol. 99, pp.334–343.
Puhan, S., Jegan, R., Balasubbramanian, K. and Nagarajan, G. (2009) ‘Effect of injection pressure
on performance, emission and combustion characteristics of high linolenic linseed oil methyl
ester in a DI diesel engine’, Renewable Energy, Vol. 34, No. 5, pp.1227–1233.
Raheman, H. and Ghadge, S.V. (2008) ‘Performance of diesel engine with biodiesel at varying
compression ratio and ignition timing’, Fuel, Vol. 87, No. 12, pp.2659–2666.
Rakopoulos, C.D., Rakopoulos, D.C., Hountalas, D.T., Giakoumis, E.G. and Andritsakis, E.C.
(2008) ‘Performance and emissions of bus engine using blends of diesel fuel with bio-diesel of
sunflower or cottonseed oils derived from Greek feedstock’, Fuel, Vol. 87, No. 2, pp.147–157.
Rao, Y.V.H., Voleti, R.S., Raju, A.V.S. and Reddy, P.N. (2010) ‘The effect of cottonseed oil
methyl ester on the performance and exhaust emissions of a diesel engine’, International
Journal of Ambient Energy, Vol. 31, No. 4, pp.203–210.
Raslavicius, L. and Bazaras, Z. (2010a) ‘Ecological assessment and economic feasibility to utilize
first generation biofuels in cogeneration output cycle – the case of Lithuania’, Energy, Vol. 35,
No. 9, pp.3666–3673.
Raslavicius, L. and Bazaras, Z. (2010b) ‘Variations in oxygenated blend composition to meet
energy and combustion characteristics very similar to the diesel fuel’, Fuel Processing
Technology, Vol. 91, No. 9, pp.1049–1054.
Sayin, C. and Gumus, M. (2011) ‘Impact of compression ratio and injection parameters on the
performance and emissions of a DI diesel engine fuelled with biodiesel-blended diesel fuel’,
Applied Thermal Engineering, Vol. 31, No. 16, pp.3182–3188.
Shehata, M.S. and Abdel Razek, S.M. (2011) ‘Experimental investigation of diesel engine
performance and emission characteristics using jojoba/diesel blend and sunflower oil’, Fuel,
Vol. 90, No. 2, pp.886–897.
Song, J-T. and Zhang, C-H. (2008) ‘An experimental study on the performance and exhaust
emissions of a diesel engine fuelled with soybean oil methyl ester’, P I MechEng D-J Aut,
Vol. 222, No. 12, pp.2487–2496.
Sureshkumar, K., Velraj, R. and Ganesan, R. (2008) ‘Performance and exhaust emission
characteristics of a CI engine fuelled with pongamiapinnata methyl ester (PPME) and its
blends with diesel’, Renewable Energy, Vol. 33, No. 10, pp.2294–2302.
Ulusoy, Y., Tekin, Y., Cetinkaya, M. and Karaosmanoglu, F. (2004) ‘The engine tests of biodiesel
from used frying oils’, Energy Sources, Vol. 26, No. 10, pp.927–932.
Wander, P.R., Altafini, C.R., Moresco, A.L., Colombo, A.L. and Lusa, D. (2011) ‘Performance
analysis of a mono-cylinder diesel engine using soy straight vegetable oil as fuel with varying
temperature and injection angle’, Biomass and bio-energy, Vol. 35, No. 9, pp.3995–4000.
Yilmaz, N. and Morton, B. (2011) ‘Comparative characteristics of compression ignited engines
operating on biodiesel produced from waste vegetable oil’, Biomass and Bio-Energy, Vol. 35,
No. 5, pp.2194–2199.

Das könnte Ihnen auch gefallen