You are on page 1of 24

Int. J. Exergy, Vol. 14, No.

1, 2014 101

Effects of rapeseed oil addition to a diesel fuel


on thermodynamic efficiencies

E. Buyukkaya and H.S. Soyhan


Engineering Faculty,
Department of Mechanical Engineering,
Sakarya University,
Esentepe Campus, 54187, Sakarya, Turkey
E-mail: ebkaya@sakarya.edu.tr
E-mail: hsoyhan@hotmail.com

B. Gokalp*
Engineering Faculty,
Department of Mechanical Engineering,
Kocaeli University, 41040, Kocaeli, Turkey
E-mail: burak.gokalp1@kocaeli.edu.tr
*Corresponding author

Abstract: This paper presents a comparative energy and exergy analyses


of a six-cylinder, four-stroke, direct injection diesel engine using diesel fuel
and its mixture with rapeseed oil. Engine was used to obtain the experimental
data at full-load conditions for constant 2100 rpm speed. It was found that the
first- and second-law efficiencies were increased with biodiesel addition.
On the other hand, exhaust exergy was found to be inversely proportional to the
biodiesel addition. It is also shown with extensive emission analysis that the
addition of biodiesel fuels to the standard diesel fuel enhances the emission
characteristics of our diesel engine.

Keywords: energy; exergy; emission; diesel fuel; rapeseed oil.

Reference to this paper should be made as follows: Buyukkaya, E.,


Soyhan, H.S. and Gokalp, B. (2014) ‘Effects of rapeseed oil addition to
a diesel fuel on thermodynamic efficiencies’, Int. J. Exergy, Vol. 14, No. 1,
pp.101–124.

Biographical notes: Ekrem Buyukkaya is a member of the Department of


Mechanical Engineering, Sakarya University since 1992. He got his BEng
(1991), MSc (1994) and PhD (1997) from Istanbul Technical University.
Currently, he is been working on Energy Technologies, Air Pollution and Its
Control Automotive as an Associate Professor in the Engineering Faculty at
Sakarya University.

Hakan Serhad Soyhan is a member of the Department of Mechanical


Engineering, Sakarya University since 1992. He got his BEng (1992), MSc
(1995) and PhD (2000) from Istanbul Technical University, did post-doctoral
researchs in chemical kinetics at the Combustion Physics Division at Lund
University in Sweden and on HCCI engines and chemical kinetics at Shell
Global Solutions in Chester, UK. Currently, he is been working on combustion
modelling studies in relation to emission reduction and control of emissions

Copyright © 2014 Inderscience Enterprises Ltd.


102 E. Buyukkaya et al.

from transportation as an Associate Professor in the Engineering Faculty at


Sakarya University. He is Head of Local Energy Research Association and a
member of associates in the Turkish Society of Mechanical Engineers as well
as in the ASME ICE section.

Burak Gokalp received his BS (1997) from the Department of Mechanical


Engineering, Naval, Academy İstanbul, Turkey, and gained MS (2005) and
PhD (2009) from the Department of Mechanical Engineering, Kocaeli
University, Kocaeli, Turkey. His research interests include internal combustion
engines, renewable energy sources, clean energy, marine engines.

1 Introduction

Biodiesel has received much attention in the past decade owing to its ability to replace
fossil fuels, which are more likely to run out within a century. Especially, the
environmental issues concerned with the exhaust gases emission by the usage of fossil
fuels also encourage the usage of biodiesel, which has proved to be ecofriendly far more
than fossil fuels. Biodiesel is known as a carbon neutral fuel because carbon present in
the exhaust was originally fixed from the atmosphere (Srivathsan et al., 2008). One of the
most common methods in producing biodiesel is transesterification of vegetable oil,
waste animal fats and waste restaurant greases (yellow grease) with a short-chain alcohol.
These oils are identified as one of the future contenders to fulfil the demand gap
produced by the depletion of fossil diesel fuels (Ma and Hanna, 1999; Zheng et al., 2006,
2007). High-purity methyl ester can be achieved by transesterification of fresh vegetable
oils with methanol in the presence of an alkaline catalyst (Dmytryshyn et al., 2004).
Transesterification of rapeseed oil produces ester whose properties are comparable
with those of the conventional diesel fuels (Lang et al., 2001a). Rapeseed oil is a low
erucic acid, which is now the second largest oilseed crop after soybean and the third
largest vegetable oil after soybean oil and palm oil. Rapeseed contains approximately
40% oil (Ackman, 1990; Lang et al., 2001b).
In comparison with the conventional diesel fuels, biodiesel is 100% renewable.
However, this proportion is reduced to around 90% (if the balance is made in mass) or
95% (if the balance is made in carbon mass) when fossil alcohol (usually methanol) is
used. The life-cycle analyses of CO2 emissions should be accounted for to evaluate the
impact of biodiesel on the global greenhouse effect (Beer et al., 2001; Toyota Motor
Corporation, 2006; Sheehan et al., 1998). Biodiesel fuel can effectively reduce engine-out
emissions of particulate matter, carbon monoxide (CO) and unburned hydrocarbons in
modern four-stroke compression–ignition engines (Beer et al., 2001; Lapuerta et al.,
2008a; Graboski and McCormick, 1998).
Nevertheless, low-temperature combustion strategies, such as homogeneous charge
compression–ignition enabling technologies and smokeless diesel combustion, offer a
promising solution to simultaneously reduce the formation of NOx and particulate matter.
Many investigators have studied the performance and exhaust emissions of diesel
engines using various biodiesel (Usta, 2005; Almeida et al., 2002; Kalligeros et al., 2003;
Canakcı and Van Gerpen, 2001; Altıparmak et al., 2007; Nwafor, 2004). Rakopoulos
et al. (2008a) conducted an evaluation of the use of sunflower and cottonseed oil methyl
esters (biodiesels) of Greek origin as supplements in the diesel fuel at blend ratios
Effects of rapeseed oil addition to a diesel fuel 103

of 10:90 and 20:80, in a fully instrumented, six-cylinder, turbocharged, after-cooled,


direct-injection (DI), Mercedes-Benz, mini-bus diesel engine. They showed that all of the
tested biodiesel blends can be used safely and advantageously in the present bus diesel
engine, at least in these small blending ratios, with the cottonseed biodiesel showing
emission-wise a small superiority over its sunflower biodiesel counterpart concerning
soot and CO emissions (Rakopoulos et al., 2008a; Rakopoulos and Giakoumis, 2006a).
In the other experimental study, Rakopoulos et al. (2006) evaluated and compared the use
of various diesel fuel supplements at blend ratios of 10:90 and 20:80. Theoretical aspects
of the diesel engine combustion, combined with the widely differing physical and
chemical properties of these diesel fuel supplements against the normal diesel fuel, are
used to aid the correct interpretation of the observed engine behaviour. Van Gerpen and
Shapiro (1990) investigated the second-law analysis of diesel engine combustion, defined
a thermodynamic system as outside the engine cylinder and discussed the chemical
availability. Primus and Flynn (1986) showed the benefits of using the second law in
determining various energy losses in a diesel engine. Rakopoluulos et al. investigated the
accumulation and destruction of energy in a direct-injection diesel engine based on
experimental data (Lipkea and DeJoode, 1998; Rakapoulos et al., 1993). Bozza and
Rakopoulos suggested that this portion of chemical availability should not be taken
into account when studying internal combustion engine applications (Rakopoulos and
Andritsakis, 1993; Bozza et al., 1991).
Analysis of thermodynamic systems and processes are usually based on the first law
of thermodynamics. This type of analysis permits the designer to evaluate the internal
energy variation as a function of the energy transfers across the boundaries as heat or
work and the enthalpies associated with the mass flow crossing these boundaries.
However, the first law is deficient for evaluating some features of energy resource
use (Cengel and Boles, 1998; Moran, 1982; Benson and Whitehouse, 1979). The
thermodynamic details of thermal system operations can be better understood by
performing exergy analysis and not only energy analysis of the system. The exergy
analysis reveals the locations, sources and magnitudes of energy resources wasted in the
system (Cengel and Boles, 1998; Rakopoulos and Giakoumis, 2006a).
Several researchers have studied the exergy analysis on internal combustion engines
(Cengel and Boles, 1998; Moran, 1982; Benson and Whitehouse, 1979; Moran and
Shapiro, 2000; Ferguson, 1986; Rakopoulos et al., 2008b). Kumar et al. (1989) used the
definition of total available flow and a simulation model of a diesel engine cycle to obtain
the distribution of losses and overall engine irreversibility. Rakopoulos and Giakoumis
(2006b) analysed a turbocharged diesel engine operating under transient load conditions
to simulate the transient operation on a crank angle degree basis, using a detailed analysis
of mechanical friction. Flynn et al. (1984) theoretically studied effects of insulating a
cylinder wall on the availability destruction and showed that the lower availability
destruction of the heat release in the in-cylinder gases could be obtained by insulation.
In this study, neat rapeseed oil was considered as a potential alternative fuel for an
unmodified diesel engine because it has high oil content (around 40%). Experimental and
computational analyses on first- and second-law efficiencies have been performed for a
diesel engine fuelled with rapeseed oil and its diesel blends compared with those of
standard diesel. Emission analyses are also performed to investigate effects of biodiesel
fuels mixed with standard diesel fuel to enhance emission characteristics of diesel
engines using Standard diesel fuel.
104 E. Buyukkaya et al.

2 Energy balances

In this study, five species were taken into account (CO, CO2, O2, H2O and N2) in the
calculation of internal energy and specific heats for the first- and second-law analyses.
All species were assumed to behave as perfect gases. The fuels used in comparison
are diesel (C14,09H24,78) with a lower heating value (LHV) of 42,640 kJ/kg, B100
(C18,74H34,43O2) with an LHV of 37,388 kJ/kg, B50 with an LHV of 40,014 kJ/kg,
B20 with an LHV of 41,589 kJ/kg and B5 with an LHV of 42,377 kJ/kg. Ambient
temperature, Te, was assumed to be equal to 25°C. Table 1 shows the standard
environment considered in the second-law analysis.

Table 1 Standard environment considered in the exergy analysis (T0 = 298.15 K,


P0 = 1.01325 bar)

Reference species Mole fraction (%)


N2 75.670
O2 20.350
H2O 3.0300
CO2 0.0345
CO 0.0007

The thermodynamic system is shown in Figure 1. In this system, it is assumed that air and
fuel enter the combustion chamber and the combustion products leave the system as
exhaust gases to the atmosphere. Cooling water enters and exits the system and
lubricating oil re-circulates in the control volume. It is assumed that heat exchange occurs
only between the engine and environment.

Figure 1 The engine as a thermodynamic open system (see online version for colours)

Energy equation for the system shown in Figure 1 is written as

Ws = m f LHV + Q w + Q env − W f − E ex (1)


Effects of rapeseed oil addition to a diesel fuel 105

where W s is the shaft power, m f LHV is the fuel input power, Q w is the rate of heat
transfer to the coolant, W f is the friction power, E ex is the exhaust power and Q env is the
rate of heat transfer to the environment. Q env was calculated by using energy balance
equation (equation (1)) since it is the only unknown in equation.

2.1 Friction power


The friction power for the test engine was calculated via an empirical formula
(Rakopoulos and Giakoumis, 2006) as follows:
p n
W f = pfmeVd ni = fme (2)
2570
where Vd is the displacement volume (m3), n is the engine speed (rpm) and i is the
number of power stroke per each revaluation (0.5 for four stroke engines) and Pfme is the
friction means effective pressure (bar), which was calculated by using the following
equation:
pfme = 0.123ε + 0.0004774n. (3)

2.2 Exhaust energy


Exhaust energy was calculated by using enthalpies of the products. In rich combustion,
an equilibrium among the species of H 2 O, CO2 , CO, H2 and O 2 is often assumed to
determine the burned gas compositions. In that case, burned gas composition of a fuel can
be calculated via the water–gas reaction. In the case of incomplete reaction, a mole-based
reaction equation can be written as
Cc H h Oo + λ O min (O 2 + 3.762N 2 ) → nCO2 CO 2 + nCO CO + nH2 O H 2 O + nH2 H 2
(4)
+ nN2 N 2 + nO2 O 2

The exhaust power was calculated as follows:


E ex = n f ∆H pr
o
(5)
o
where ∆H pr is the enthalpy of the products and n f is the fuel molar rate.

3 Exergy balance

Maximum extractable shaft power, Ws ,max , was formulated for the equivalent
thermodynamic open system as shown in Figure 1:

 T   T 
W s ,max = m a ba + m f b f − W f − Q ex + Q w 1 − e  + Q env 1 − e  (6)
 Tw   Tw 

where Tw is the mean coolant temperature and calculated as


106 E. Buyukkaya et al.

Tw =
(Tw,in + Tw,out )
. (7)
2
In this system, two heat exchanges occur: one with the cooling circuit at constant
temperature, Tw , and the other one with the environment at constant temperature, Te.
In equations (6) and (7), it is assumed that surface temperature of the engine is equal to
the average coolant temperature.
The flow availability associated with fuel was given (Moran and Shapiro, 2000)
for hydrocarbon liquid fuels of general type CmHn as
 n 0.042 
b f = LHV 1.04224 + 0.011925 − . (8)
 m m 
The exergy of the exhaust gas flow can be divided into two main components as follows:

bex = etm + ekim (9)

where etm is the thermo-mechanical exergy and ekim is the chemical exergy. The total
exergy for species contained in one mole of gas mixture is
n  n xip 
bex = ∑ xi hi (T ) − hi (T0 ) − T0 [ si (T ) − si (T0 ) ] + ∑ RT0 ln  (10)
i =1  i =1 xi0 p0 
where n is the number of species. The maximum amount of power that can be extracted
from an exhaust stream at a given condition can be defined as

Wex = nex bex (11)

The maximum extractable shaft power W s ,max is written in the following form

Ws ,max = Ws + Te S p (12)

where the second term in the right-hand side states the total irreversibility (availability
destruction) of the system. S p is the rate of entropy production owing to irreversibility
within the control volume. It can be easily seen from equation (12) that the maximum
extractable shaft power W is achieved when the processes within the control volume
s ,max
are reversible. It also shows the reversible limit of the system. The second-law efficiency
of the engine was defined as
W W
η II =  s = 1 −  lost . (13)
Ws ,max Ws ,max

The lost power, which occurs owing to irreversibility of the system, is formulated as

Wlost = Ws ,max − Ws = Te S p (14)

It is clear from equation (12) that magnitude of the power loss depends on the choice
of the standard environment temperature.
Effects of rapeseed oil addition to a diesel fuel 107

The first-law efficiency was defined as


W s
ηI = (15)
m f LHV

where m f is the mass flow rate of the fuel.

4 Irreversibility sources in the engine

Entropy production in an open system can be given in a general form as follows


(Figure 2):
Q
S p = ∑ m out sout −∑ m in sin + ∑ i (16)
Ti

It must be noted that the sign of Qi must be taken as negative since heat transfer occurs
from the system to the environment. Irreversibility of the system, I, was calculated from
the following equation (11)
T T
I = Te S p = Te (m ex sex − m a sa − m f s f ) + W f + Q w e + Q env e (17)
Tw Tw

The first term at the right side of equation (17) accounts for the internal irreversibility
of the system, the second term represents irreversibility arising from total friction, the
third and the last terms represent irreversibility owing to the heat transfer processes
across the system boundary.

Figure 2 The engine for the equivalent thermodynamic open system


108 E. Buyukkaya et al.

5 Experimental set-up

The experiments were conducted on a six-cylinder, four-stroke, turbocharged direct


injection diesel engine. The experimental set-up is shown schematically in Figure 3. The
accuracies of the measured parameters and the uncertainties in the calculated parameters
are given in Table 2. The engine specifications are given in Table 3. The test engine was
coupled to a hydraulic dynamometer providing a maximum engine power of 350 kW
with a ±0.1 kW of uncertainty to control engine speed and load. Full-load characteristics
of the DI diesel engine fuelled with neat rapeseed, its blends and diesel fuels were at the
constant engine speeds, ranging from 1000 rpm to 2100 rpm with an interval of 200 rpm.
At each speed, the maximum engine torque was reached for each fuel. In other words, the
test engine was operated at different torques when different fuels were tested. Engine
speed was measured by a digital tachometer with a resolution of 1 rpm.

Figure 3 The test setup

Table 2 The accuracies of the measurements and the uncertainties in the calculated results

Temperatures ±1°C
CO ±0.5%
HC ±10 ppm
CO2 ±0.5%
NOx 300 ppm
Smoke meter ±5%
Calculated results Uncertainty
BSFC max. ±2%
BTE max. ±2%
Pressure ±1 bar
Crank angle encoder ±1°
Engine speed ±1 rpm
Effects of rapeseed oil addition to a diesel fuel 109

The fuel consumption was measured by gravimetric method with an AVL 733 S model
fuel meter, for which a closed graduated cylinder having a total volume of 250 cm3 was
previously filled with fuel and, while the engine was operated in the prescribed
conditions, graduated cylinder was opened and a record of the time required for the given
amount of fuel to be consumed was obtained. The volumetric flow rate of the intake air
was measured using a rotary-type flow meter. A surge tank located between the air flow
meter and intake manifold was used for damping out the pulsations produced by the
engine, thus obtaining a steady air flow. The exhaust gas temperature was measured
using a thermocouple connected to the exhaust pipe just downstream of the exhaust
manifold. The cooling water temperatures at the inlet and outlet of the engine block were
measured using ‘Pt 100’ thermocouples.

Table 3 Engine specifications

Type of engine MAN, direct injection, turbocharged


Cylinder number 6
Cylinder diameter 150 mm
Stroke 121 mm
Compression ratio 17/1
Maximum engine power 164 kW at 2100 rpm
Maximum engine torque 819 Nm at 1600 rpm
Injection pump 22° BTDC
Injector opening pressure Mechanically controlled in-line type
Type of fuel injection 220 bar
Nozzle hole diameter Pump-line-nozzle injection system
Nozzle type 0.3 mm
Nozzle hole number Multi hole

A MOBYDIC (5000 + 5100 combined) portable combustion analyser was used for direct
volumetric analysis of the O2, CO, NOx and CO2 contents in the exhaust gas and of the
exhaust gas and ambient temperatures. A Bosch GMBH-ETD02050A model smoke
meter was used to measure the opacity in the exhaust gas by drawing 1.80 l of exhaust
gas through a paper filter. The biodiesel used in this study was provided from a local
producer that uses a transesterification process, which was catalysed by potassium
hydroxide (1.0% by weight of oil). Then, biodiesel was analysed by an established
research institution following the ASTM D6751 standard, a test specification for neat
rapeseed oil. Additionally, ASTM D5291 was applied to obtain the C/H/O ratios. The
important properties of neat rapeseed oil are presented in Table 4.
As concerned with the engine performance, a measurement methodology was
employed such that for each type of fuel utilised, the engine speed was initially set at
1000 rpm. At the full load, the power output and torque were measured on the monitoring
equipment available at the test bench. The engine speed was then increased to 1200 rpm
and this procedure was repeated for other engine speeds. The determination of the
power and torque curves was conducted following this technique by varying speed
from 200 rpm to 2000 rpm. First, diesel fuel was used as fuel at the diesel engine.
110 E. Buyukkaya et al.

Then, the mixtures of diesel and biodiesel having 5, 20 and 70% volumetric proportions
of biodiesel were named as B5, B20 and B70, respectively. Finally, pure biodiesel fuel
was used as fuel at the diesel engine.

Table 4 Properties diesel fuel and neat rapeseed oil

Unit Method Commercial diesel Neat rapeseed oil


3
Density (at 15°C) kg/m ISO 3675 837 920
2
Viscosity (at 40°C) mm /s ISO 3104 2.6 3.5
Calorific value MJ/kg ASTM D 4809 43.35 37.1
Flash point °C ISO 2719 72 240
Pour point °C EN 116 –20 –5
Cetane number ISO 5165 49 39
C mass fraction % kg/kg ASTM D 5291 85.4 78
H mass fraction % kg/kg ASTM D 5291 15 13.5
O mass fraction % kg/kg – 0 8.9

6 Experimental results

All tests were performed under steady-state conditions. The brake-specific fuel
consumption (BSFC), brake thermal efficiency (BTE), mechanical efficiency (ME),
exhaust gas temperature and exhaust emissions, such as CO, CO2 and NOx, have been
investigated.
The BSFC, the fuel consumption rate ( m f ) per unit effective power ( Pe ), was
calculated as
3600 m f
BSFC = . (18)
Pe

Figure 4(a) shows the variations in the BSFC of the test engine with engine speed at
full-load condition. At 1400 rpm, the BSFCs diesel and B100 fuels were 232 g/kWh and
251 g/kWh, respectively. BSFCs of the B5, B20, B70 and B100 fuels were observed to
be higher by 2.5, 3, 5.5 and 7.5% than that of the diesel fuel, respectively. The higher fuel
consumption of the B100 and their blends could be primarily related to LHV of the B100.
A large majority of authors reported that they found increases in biodiesel fuel
consumption in proportion to the biodiesel content in the blends and to the loss of heating
value (Lapuerta et al., 2008a; Ming et al., 2008; Ramadhas et al., 2004a, 2005; Raheman
and Phadatare, 2004a). Labeckas and Slavinskas (2006) claimed that the higher fuel
consumption of the blends could be related to the lower, on average, by 12.5% of net
heating value of rapeseed oil methyl ester. Another reason for the increase in BSFC with
biodiesel would be a change in the combustion timing caused by the biodiesel’s higher
cetane number as well as injection timing changes.
Effects of rapeseed oil addition to a diesel fuel 111

Figure 4 Variations of (a) the BSFC; (b) the BTE; (c) the power output and (d) the engine torque
with engine speed at full load

(a) (b)

(c) (d)

It is also seen that the BSFC increases slightly with the increase in the percentage of the
rapeseed oil in blend fuels. On the other hand, starting at the minimum BSFC point,
increasing or decreasing speed at a constant load increases BSFC as a result of mainly the
reduced engine volumetric efficiency. It is generally accepted that fuel consumption is
proportional to volumetric energy density of the fuel based on the lower and net heating
value (Graboski and McCormick, 1998).
The BTE indicates the energetic performance of the engine and is defined as the ratio
of the power output to the fuel energy input as
Pe
BTE = (19)
m f LHV

where LHV is the Lower Heating Value of the fuel. BTE of the engine fuelled
with diesel, B5, B20, B70 and B100 fuels is shown in Figure 4(b). It is observed
that soybean methly ester (SME) yields maximum BTE values. It is noticed that the
use of oxygen-rich biodiesel promotes a better combustion, thus improving the thermal
efficiency, compared with D2, where LHV is the lower heating value of the fuel. The
BTE is defined as the actual brake power per cycle divided by the amount of chemical
energy as indicated by the LHV of the fuel (Canakcı and Van Gerpen, 2001). It is noticed
112 E. Buyukkaya et al.

that the use of oxygen-rich biodiesel promotes a better combustion, thus improving the
thermal efficiency, compared with the other blends. At full load, BTE increases with
increasing engine speed up to 2000 rpm. At 2000 rpm, the BTE values are 0.427, 0.425,
0.425, 0.424 and 0.423 for B20, B5, B70, B100 and diesel fuels, respectively. At
2100 rpm, the BTE decreased for all fuels. The reason is insufficient air causing
incomplete combustion of the fuel. The BTE obtained with B100 and their blends close
to that of the one obtained with the diesel fuel. Labeckas and Slavinskas (2006) tested a
4750 cc engine under different steady modes using 5, 10, 20, 35% blends and pure
rapeseed-oil biodiesel. They obtained higher BTEs with 5–10% blends compared with the
others. On the other hand, Ramadhas et al. (2005b) obtained higher BTEs with 10% and
20% blends. This improved efficiency was explained by authors with increased lubricity
of these blends when compared with diesel fuel.
Figure 4(c) shows the variation in engine power at full load for different fuels.
The results show that there are no noticeable differences in the measured engine power
output between diesel and B5 fuels. However, the measured engine power for other
blends is lower than that of the diesel fuel. Maximum reduction in engine power for B20,
B70 and B100 fuels is 6, 8 and 10 kW, respectively. LHV of the rapeseed oil is
responsible for this reduction. Similar results were reported by Kaplan et al. (2006),
who compared sunflower-oil biodiesel and diesel fuels at full and partial loads and at
different engine speeds in diesel engine. Their results showed a power reduction between
5% and 10%. They also explained this power reduction with LHV of the biodiesel.
Figure 4(d) shows the variation in engine torque at full load for different fuels.
Maximum torque was obtained at 1600 rpm for each kind of fuel. At 1600 rpm, power
and torque of the diesel and B5 fuels were almost imperceptible. At higher speeds, the
torque delivered with B5 fuel was higher approximately 2 Nm, on average, than the
torque delivered by diesel fuel. But, a more pronounced torque drop was observed for
B20, B70 and B100 fuels, and the average torque drop between diesel and B20, B70 and
B100 fuels is 19.7, 32 and 38.7 Nm, respectively. At the speed of 1800 rpm, the
maximum difference of the measured torques between diesel and B20, B70 and B100
fuels was found to be 2.2, 4 and 5%, respectively.
Figure 5(b) shows the exhaust gas temperature traces for different fuels. The exhaust
gas temperature depends on the energy density of fuel and the combustion efficiency. The
exhaust gas temperature from burning SME and D2/SME blends was increased because
of the O2 content of SME. Exhaust temperatures are affected by ignition delay.
A biodiesel that has a slightly lower cetane rating results in a longer ignition delay
and slower burning rate (Nwafor, 2004). Higher ignition delay results in a delayed
combustion and higher exhaust temperature. The gas temperature also presented
variations when the injection system was inspected because of the improved performance
of the diesel engine (Almeida et al., 2002). It can be seen from the figure that biodiesel
fuel and its blends give higher exhaust gas temperatures than diesel fuel for all of the
engine speeds. The exhaust gas temperature reached its maximum level at B100 since
rapeseed oil content increased.
The NOx emissions of the engine for diesel, B5, B20, B70 and B100 fuels are shown
in Figure 5(c). It is known that NOx formation is dependent on volumetric efficiency,
combustion duration and especially temperature arising from high activation energy
needed for the reactions involved. Although the exhaust gas temperatures increased,
Effects of rapeseed oil addition to a diesel fuel 113

the NOx emissions were observed to decrease with the increase in engine speed
(Figure 5(b) and (c)). The reaction time of each engine cycle was thereafter reduced so
that the residence time of the gas temperature within the cylinder was shortened. This led
to lower NOx emissions under higher engine speeds. The increase in NOx emissions was
proportional to the amount of biodiesel. In the case of pure biodiesel, the increase in NOx
emission was 12% compared with the diesel fuel. There were also 6 and 9% increase in
NOx emissions for B20 and B70, respectively. Similar conclusions were drawn by other
authors in the literature (Lapuerta et al., 2008b; de Almeida Silvio et al., 2002; Marshall
et al., 1995; Nwafor et al., 2000). Data from the soybean diesel study of Graboski et al.
have been re-analysed to gain a greater understanding of the increase in NOx emissions.
The NOx emission is clearly dependent on speed and load. Analysis of the test results
shows that the maximum NOx emissions increase proportionally to the mass percent
of oxygen in the SME (Graboski and McCormick, 1998; Kalligeros et al., 2003).
On the other hand, fuel density is known to affect emissions of NOx from the diesel
engine. The NOx increase occurred because the fuel injectors inject a constant volume or
larger mass for the more dense fuel, a given speed and load (Graboski and McCormick,
1998). However, Dorado et al. (2003) indicated a decrease in NOx emissions using waste
olive oil methyl ester instead of diesel fuel.
Figure 5(d) shows the CO emission of the engine as a function of engine speed at
full-load traces for different fuels. This emission resulted from incomplete oxidation of
the fuel. Biodiesel fuels containing around 11% oxygen by weight enhance conditions to
get nearly complete combustion, and thus cause more CO2 formation rather than CO
(Ramadhas et al., 2005c). It was observed that the CO emission decreased with the
increase in engine speed. At low engine speeds, the CO emissions of the B5, B20, B70
and B100 are 12, 25, 31 and 35% lower than that of diesel fuel, respectively. This
decrease may be due to the oxygen content of the blends and pure biodiesel. Poor
atomisation and uneven distribution of small portions of fuel across the combustion
chamber, along with a low gas temperature, may cause local oxygen deficiency and
incomplete combustion (Roskilly et al., 2008). The highest CO emission of 900 ppm was
measured for diesel fuel at 1000 rpm. The CO emissions are shown to decrease more
rapidly for all fuels from 1000 rpm to1400 rpm. Reduced CO emissions were maintained,
probably, thanks to oxygen inherently present in the biofuel, which makes it easier to be
burnt at higher temperature in the cylinder. Similar results can be found in other studies
(Roskilly et al., 2008; Raheman and Phdatare, 2004b).
The variations of HC emissions for diesel and biofuels are shown in Figure 5(e).
The emission of unburned HC is negligibly small for all the fuels. The HC emissions of
B20, B70 and B100 fuels were lower than that of diesel fuel. The increased gas
temperature and the higher cetane number as responsible for this decrease may be
explained. The higher temperature of the burned gases prevented condensation of the
heaviest hydrocarbons in the sampling line, suggesting proper conditions for HC
emission analysis. The higher cetane number of biodiesel causes a decrease in HC
emissions owing to the decrease in combustion delay (Monyem et al., 2001).
Figure 6(a) shows the variation of cylinder pressure with crank angle for diesel, neat
rapeseed oil and its blends at 2000 rpm and full-load conditions. From this figure, it is
clear that the peak cylinder pressure is decreased with the increase in rapeseed oil
addition in the blends. However, the combustion process of the test fuels is similar,
114 E. Buyukkaya et al.

consisting of a phase of premixed combustion following by a phase of diffusion


combustion. Premixed combustion phase is controlled by the ignition delay period and
spray envelope of the injected fuel (Senthil et al., 2005; Ozsezen et al., 2009a; Devan and
Mahalakshmi, 2009). Therefore, the viscosity and volatility of the fuel have a very
important role to increase atomisation rate and to improve air fuel mixing formation. The
cylinder peak pressure because of the high viscosity and low volatility of rapeseed oil and
its blends is lower than that of standard diesel. It is observed that the peak pressures of
154.7, 150, 149, 148.5 and 147 bar were recorded for standard diesel, B5, B20, B70 and
B100, respectively. Similar conclusions were drawn by other authors in the literature
(Senthil et al., 2005; Ozsezen et al., 2009a; Devan and Mahalakshmi, 2009). Similar
results were reported by Devan and Mahalakshmi (2009), who compared poon oil
biodiesel and diesel fuels at full load in single-cylinder diesel engine. They observed that
cylinder pressures of 67.5, 63 and 60 bar were recorded for standard diesel, B20 and poon
oil, respectively. They explained pressure reduction with the expected effects of poon oil
viscosity on fuel spray, and reduction of air entrainment and fuel/air mixing rates.
However, the cylinder peak pressure of biodiesel fuels with methyl ester was lower than
that of the pure biodiesel or was worth close to diesel fuel owing to the improvement in
the preparation of air fuel mixture as a result of low fuel viscosity (Sezer and Bilgin,
2008; Ozsezen et al., 2009b).
The heat release rate is used to identify the start of combustion, the fraction of fuel
burned in the premixed mode and differences in combustion rates of fuels (Ozsezen et al.,
2009a). Analyses of cylinder pressure data to obtain the heat release rate for neat
rapeseed oil and its blends were conducted. Heat release rate shown in Figure 6(b)
indicates that the ignition delay for B100 and the blends are lower than that for diesel.
The maximum heat release rate of standard diesel, B5, B20, B70 and B100 is 84, 79.7,
77.50, 74.9 and 72.2 J/° CA, respectively. This is because, as a consequence of the
shorter ignition delay, such as seen in Table 4, the premix combustion phase for neat
rapeseed oil and its blends is less intense. On the other hand, increased accumulation of
fuel during the relatively longer delay period resulted in higher rate of heat release while
running with diesel. Because of the shorter delay, peak heat release rate occurs earlier for
neat rapeseed oil and its blends in comparison with diesel. For B5, B20 and B70 blends,
the heat release peak was higher than that of B100 owing to reduced viscosity and better
spray formation. The less intense premixed combustion phase was due to the shorter
ignition delay of neat rapeseed oil compared with that of diesel. This was probably the
result of the chemical reactions during the injection of vegetable oil at high temperature.
Although similar conclusions were drawn by other authors in the literature, they were at
different conclusions. Ozsezen et al. (2009a) studied on combustion characteristics of
crude sunflower oil and diesel fuels at 3000 rpm and full load. They observed that the
start of combustion of diesel fuel was earlier than those of crude sunflower oil and diesel
fuel had less ignition delay period. They explained that the crude sunflower oil exhibited,
on average, 2.08° CA longer ignition delay owing to its lower cetane number when
compared with diesel fuel. One of the most important parameters in the combustion
phenomenon is the ignition delay. The increase in fuel viscosity, particularly for
petroleum-derived fuels, results in poor atomisation, slower mixing, increased penetration
and reduced cone angle. These result in longer ignition delay. But, biodiesel is not
derived from crude petroleum, and the opposite trend is seen in the case of biodiesel
Effects of rapeseed oil addition to a diesel fuel 115

and their blends (Kanoglu et al., 2008). In this study, the ignition delay was calculated in
terms of the crank angle between the start of fuel injection and the start of combustion.
The ignition delays for B0, B5, B20, B70 and B100 fuels were found as 8.5, 7.75, 7.25,
6.50 and 5.75° CA, respectively. The ignition delay slightly decreased with the use of
biodiesels. Biodiesel usually includes a small percentage of diglycerides having higher
boiling points than diesel. However, the chemical reactions during the injection of
biodiesel at high temperature resulted in the breakdown of the high molecular weight
esters. These complex chemical reactions led to the formation of gases of low molecular
weight. Rapid gasification of this lighter oil in the fringe of the spray spreads out the jet,
and thus volatile combustion compounds ignited earlier and reduced the delay period.

Figure 5 Variations of (a) the smoke opacity; (b) the exhaust gas temperature; (c) the NOx
emissions; (d) the CO emissions (e) and the HC emissions with engine speed at full load

(a) (b)

(c) (d)

(e)
116 E. Buyukkaya et al.

Figure 6 Variations of (a) the cylinder gas pressure and (b) the heat release rate with respect
to CA and fuel type

(a) (b)

7 Computational results of exergy

The experimental measurements shown in Figures 7–14 show the exergy analysis of the
fuels in the engine.

Figure 7 Fuel exergy

Figure 8 Exhaust exergy


Effects of rapeseed oil addition to a diesel fuel 117

Figure 9 Exergy transfer by heat transfer

Figure 10 Exergy destruction by combustion

Figure 11 Total entropy


118 E. Buyukkaya et al.

Figure 12 Engine exergy efficiency

Figure 13 Total heat loss

Figure 14 Effective power exergy

Figures 7–14 show the exergy analysis of the fuels in the engine. Figure 7 shows specific
chemical exergies of the fuels. Because B100 and B50 fuels have lower heat capacity,
their chemical exergy gives the minimum levels. In most reported studies, fuel
consumption is calculated from the CO2 emissions and an analysis of the fuel carbon
Effects of rapeseed oil addition to a diesel fuel 119

content. A more accurate approach is due to a combination of CO2 emission


measurements with a gravimetric measurement of fuel consumption (Graboski and
McCormick, 1998).
Fuel exergy is calculated by multiplying the specific exergies with their flow rate as
seen in Figure 7. It is seen that the diesel fuel with a low addition of biodiesel gives
enhancement in engine parameters compared with the diesel fuel since biodiesel fuel
increases homogeneity of the mixture in the combustion chamber and thus increases the
combustion and thermal efficiencies.
As seen in Figure 8, exhaust exergy decreases with increasing biodiesel addition to
the standard diesel fuel because the cetane number of the biodiesel is lower compared
with the standard diesel fuel. This causes longer ignition delay, slower combustion, and
thus, a higher exhaust gas temperature. Existence of oxygen in fuel-rich regions also
increases the combustion efficiency and thus high exhaust gas temperature. This causes
higher exhaust exergy with increasing biodiesel content in the standard diesel fuel and
higher exhaust gas temperature and reaches a maximum level for B100. Reducing the
exhaust gas temperature and exhaust gas concentration can reduce the loss with a higher
exhaust gas temperature.
Figure 9 shows the exergy transfer by heat transfer from the engine to the
environment. As seen here, B5 fuel has the maximum level. Entropy produced in the
engine is a function of exergy destruction. As seen in Figure 10, exergy destruction
decreases with increasing biodiesel content in the fuel. It is because of the better
combustion by the homogeneous mixture in the cylinder. Figure 12 clearly shows that
addition of biodiesel fuel to the standard diesel fuel brings better exergy efficiency. Total
entropy, effective power exergy and total heat loss are described in Figures 11, 13 and 14,
respectively.

8 Conclusion

In this work, an experimental and computational first- and second-law analysis is


performed for a six-cylinder diesel engine at 2100 rpm by using several fuels. Engine
tests were done for rapeseed oil and its 5, 20 and 70% blends with diesel fuel. The test
results indicated that the only low-concentration blends in terms of performance
efficiency and environmentally friendly emissions (particularly for B20 and lower
blends) could be recognised as the potential candidates to be certificated for full-scale
usage in unmodified diesel engines. B20 gives the best BTE of engine. The maximum
thermal efficiency is observed with B100 and the minimum with diesel fuel. It is shown
that biodiesel addition increases the engine performance and enhances the exhaust
emissions compared with the standard diesel fuel. On the other hand, the use of rapeseed
oil as a blend has no positive effect on BSFC of the engine. At the same time, it should be
noted that higher NOx formation occurred when biodiesel is used. Therefore, researches
should be concentrated to propose NOx reduction strategies for biodiesel combustion.
From the combustion analysis, a shorter ignition delay was observed for all the test fuels.
However, further research and development on the additional fuel property measures,
long-term run and wear analysis of biodiesel fuelled engine are also necessary along with
injection timing and duration for better combustion of biodiesel in diesel engines.
Biodiesel addition to the standard diesel fuel in several ratios increases the exergy
efficiency as well. As known, exergy destruction occurs by irreversibilities in the system
120 E. Buyukkaya et al.

and decreases the system efficiency. One of the most important sources of this decrease is
the exergy destruction by exhaust gas and heat transfer. For avoiding these losses, EGR
systems can be used to reduce the in-cylinder temperatures after combustion process and
thus to reduce heat transfer.

References
Ackman, R.G. (1990) Canola and Rapeseed Production, Chemistry, Nutrition and Processing
Technology, Avi Book van Nostrand Reinhold, New York.
Almeida, S.C.A., Belchior, C.R., Nascimento, M.V.G., Vieira Leonardo dos, S.R. and
Fleury, G. (2002) ‘Performance of a diesel generator fuelled with palm oil’, Fuel, Vol. 81,
pp.2097–2102.
Altıparmak, D., Keskin, A., Koca, A. and Gürü, M. (2007) ‘Alternative fuel properties of tall oil
fatty acid methyl ester-diesel fuel blends’, Biosour. Technol., Vol. 98, pp.241–246.
Beer, T., Grant, T., Morgan, G., Lapszewicz, J., Anyon, P., Edwards, J., Nelson, P., Watson, H. and
Williams, D. (2001) Comparison of Transport Fuels Final Report (Ev45a/2/F3c) to the
Australian Greenhouse Office on the Stage 2 Study of Life-Cycle Emissions Analysis of
Alternative Fuels for Heavy Vehicles, Australia.
Benson, R.S. and Whitehouse, N.D. (1979) Internal Combustion Engines, Pergamon Press, Oxford.
Bozza, F., Nocera, R., Senatore, A. and Tuccilo, R. (1991) ‘Second law analysis of turbocharged
engine operation’, Trans. SAE J. Engines, Vol. 100, pp.547–560.
Canakcı, M. and Van Gerpen, J.H. (2001) The Performance and Emissions of a Diesel Engine
Fueled with Biodiesel from Yellow Grease and Soybean Oil, ASAE Paper 01-6050.
Cengel, Y.A. and Boles, M.A. (1998) Thermodynamics: An Engineering Approach, McGraw-Hill,
New York.
de Almeida Silvio, C.A., Rodrigues Belchior, C., Nascimento Marcos, V.G.,
Vieira Leonardo dos, S.R. and Fleury, G. (2002) ‘Performance of a diesel generator fuelled
with palm oil’, Fuel, Vol. 81, pp.2097–2102.
Devan, P.K. and Mahalakshmi, N.V. (2009) ‘Study of the performance, emission and combustion
characteristics of a diesel engine using poon oil-based fuels’, Fuel Processing Technology,
Vol. 90, pp.513–519.
Dmytryshyn, S.N., Dalai, A.K., Chaudhari, S.T., Misha, H.K. and Reaney, M.J. (2004) ‘Synthesis
and characterization of vegetable oil derived esters: evaluation for their diesel additive
properties’, Bioresource Technology, Vol. 92, pp.55–64.
Dorado, M.P., Ballesteros, E., Arnal, J.M., Gomez, J. and Lopez, F.J. (2003) ‘Exhaust emissions
from a diesel engine fuelled with transesterified waste olive oil’, Fuel, Vol. 82, pp.1311–1315.
Ferguson, C.R. (1986) Internal Combustion Engines, Wiley, New York.
Flynn, P.F., Hoag, K.L., Kamel, M.M. and Primus, R.J. (1984) A New Perspective on Diesel
Engine Evaluation Based on Second Law Analysis, SAE Tech. Pap. 840032.
Graboski, M.S. and McCormick, R.L. (2008) ‘Combustion of fat and vegetable oil derived fuels
in diesel engines’, Progr. Energy Combust. Sci., Vol. 24, pp.125–164.
Kalligeros, S., Zannikos, F., Stournas, S., Lois, E., Anastopoulos, G., Teas, Ch. and
Sakellaropoulos, F. (2003) ‘An investigation of using biodiesel/marine diesel blends on the
performance of a stationary diesel engine’, Biomass Bioenergy, Vol. 24, pp.141–149.
Kanoglu, M., Dincer, I. and Rosen, M.A. (2008) ‘Exergetic performance investigation of a
turbocharged stationary diesel engine’, Int. J. of Exergy, Vol. 5, No. 2, pp.193–203.
Kaplan, C., Arslan, R. and Surmen, A. (2006) ‘Performance characteristics of sunflower methyl
esters as biodiesel’, Energy Sources, Part A, Vol. 28, pp.751–755.
Effects of rapeseed oil addition to a diesel fuel 121

Kumar, S., Minkowycz, W.J. and Patel, K.S. (1989) ‘Thermodynamic cycle simulation of the diesel
engine: availability as second law analysis parameter’, Intern. Combust. Heat Mass Transfer,
Vol. 16, pp.335–346.
Labeckas, G. and Slavinskas, S. (2006) ‘The effect of rapeseed oil methly ester on direct injection
diesel engine performance and exhaust emissions’, Energy Conversion & Management,
Vol. 7, pp.1954–1967.
Lang, X., Dalai, A.K., Bakhshi, N.N., Reaney, M.J. and Hertz, P.B. (2001a) ‘Preparation and
characterization of bio-diesels from various bio-oils’, Bioresource Technology, Vol. 80,
pp.53–62.
Lang, X., Dalai, A.K., Reaney, M.J. and Hertz, P.B. (2001b) ‘Biodiesel esters as lubricity additives:
effects of process variables and evaluation of low-temperature properties’, Fuels Int.,
pp.207–227.
Lapuerta, M., Armas, O. and Fermandez, J.R. (2008a) ‘Effect of biodiesel fuels on diesel engine
emissions’, Progress in Energy and Combustion Sciences, Vol. 34, pp.198–223.
Lapuerta, M., Armas, O. and Fernandez, J.R. (2008b) ‘Biodiesel engine performance and emissions
in low temperature combustion’, Progress in Energy and Combustion Science, Vol. 34,
pp.198–223.
Lipkea, W.H. and DeJoode, A.D. (1998) A Comparison of the Performance of Two Direck
İnjection Diesel Engines from a Second Law Perspective, SAE Tech. Pap. 890824.
Ma, F. and Hanna, M.F. (1999) ‘Biodiesel production: a review’, Biosource Technology, Vol. 70,
pp.1–15.
Marshall, W., Schumacher, L. and Howell, S. (1995) Engine Exhaust Emissions Evaluation of a
Cummins L10E When Fueled with a Biodiesel Blend, SAE Technical Paper 952363,
doi:10.4271/952363.
Ming, Z., Mwila, C., Mulenga Graham, T.R., Meiping, W., David, S.K. and Ting Jimi, T. (2008)
‘Biodiesel engine performance and emissions in low temperature combustion’, Fuel, Vol. 87,
pp.714–722.
Monyem, A., Van Gerpen, J.H. and Canakci, M. (2001) ‘The effect of timing and oxidation on
emissions from biodiesel-fueled engines’, ASAE, Vol. 44, pp.35–42.
Moran, M.J. (1982) Availability Analysis: A Guide to Efficient Energy Use, Prentice-Hall,
Englewood Cliffs, NJ.
Moran, M.J. and Shapiro, H.N. (2000) Fundamentals of Thermodynamics, Wiley, New York.
Nwafor, O.M.I. (2004) ‘Emission characteristics of diesel engine operating on rapeseed methyl
ester’, Renewable Energy, Vol. 29, pp.119–129.
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’, Renew. Energy, Vol. 21, pp.433–444.
Ozsezen, A.N., Canakci, M., Turkcan, A. and Sayin, C. (2009b) ‘Performance and combustion
characteristics of a DI diesel engine fueled with waste palm oil and canola oil methyl esters’,
Fuel, Vol. 88, pp.629–636.
Ozsezen, A.N., Turkcan, A. and Canakcı, M. (2009a) ‘Combustion analysis of preheated
crude sunflower oil in an IDI diesel engine’, Biomass & Bioenergy, Vol. 33, No. 5, May,
pp.760–767.
Primus, R.J. and Flynn, P.F. (1986) ‘The assessment of losses in a diesel engines using second law
analysis’, in Gupta, G. (Ed.): Computer-Aided Engineering of Energy Systems, Advanced
Energy Systems, American Society of Mechanical Engineers (ASME), New York, Vol. 3,
pp.61–68.
Raheman, H. and Phadatare, A.G. (2004a) ‘Diesel engine emissions and performance from blends
of karanja methyl ester and diesel’, Biomass Bioenergy, Vol. 27, pp.393–400.
Raheman, H. and Phdatare, A.G. (2004b) ‘Diesel engine emissions and performance from blends
of karanja methly ester and diesel’, Biomass and Bioenergy, Vol. 27, pp.393–397.
122 E. Buyukkaya et al.

Rakapoulos, C.D., Andritsakis, E.C. and Kiyritsis, D.K. (1993) ‘Availability accumulation and
destruction in a Dl diesel engine with special reference to the limited cooling case’, Heat
Recovery Syst. CHP, Vol. 13, pp.261–276.
Rakopoulos, C.D. and Andritsakis, E.C. (1993) ‘DI and IDI diesel engines combustion
irreversibility analysis’, Proceedings of Audio Engineering Society (AES), ASMEWA Meeting,
New Orleans, LA, Vol. 30, pp.17–32.
Rakopoulos, C.D. and Giakoumis, E.G. (2006a) ‘Second-law analyses applied to internal
combustion engines operation’, Prog. Energy Combust. Sci., Vol. 32, pp.2–47.
Rakopoulos, C.D. and Giakoumis, E.G. (2006b) ‘Comparative first and secondlaw parametric study
of transient diesel engine operation’, Energy, Vol. 31, pp.1591–1606.
Rakopoulos, C.D., Antonopoulos, K.A., Rakopoulos, D.C., Hountalas, D.T. and Giakoumis, E.G.
(2006) ‘Comparative performance and emissions study of a direct injection diesel engine using
blends of diesel fuel with vegetable oils or bio-diesels of various origins’, Energy Convers.
Manage., Vol. 47, pp.3272–3287.
Rakopoulos, C.D., Rakopoulos, D.C., Hountalas, D.T., Giakoumis, E.G. and Andritsakis, E.C.
(2008a) ‘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, pp.147–157.
Rakopoulos, C.D., Scott, M.A., Kyritsis, D.C. and Giakoumis, E.G. (2008b) ‘Availability analysis
of hydrogen/natural gas blends combustion in internal combustion engines’, Energy, Vol. 33,
pp.248–255.
Ramadhas, A.S., Jayaraj, S. and Muraleedharan, C. (2005a) ‘Characterization and effect of using
rubber seed oil as fuel in compression ignition engines’, Renew. Energy, Vol. 30, pp.795–803.
Ramadhas, A.S., Jayaraj, S. and Muraleedharan, C. (2005b) ‘Performance and emission evaluation
of a diesel engine fueled with methyl esters of rubber seed oil’, Renew. Energy, Vol. 30,
pp.1789–1800.
Ramadhas, A.S., Muraleedharan, S. and Jayaraj, S. (2005c) ‘Performance and emission evaluation
of a diesel engine fueled with methly esters of rubber seed oil’, Renewable Energy, Vol. 30,
pp.1789–1800.
Ramadhas, S., Jayaraj, S. and Muraleedharan, C. (2004) ‘Use of vegetable oils as I.C. engine
fuels-a review’, Renew. Energ., Vol. 29, pp.727–742.
Roskilly, A.P., Nanda, S.K., Wang, Y.D. and Chirkowski, J. (2008) ‘The performance and the
gaseous of two small marine craft diesel engines fuelled with biodiesel’, Applied Thermal
Engineering, Vol. 28, pp.872–880.
Senthil, K.M., Kerihuel, A., Bellettre, J. and Tazerout, M. (2005) ‘Experimental investigations on
the use of preheated animal fat as fuel in a compression ignition engine’, Renewable Energy
Vol. 30, pp.2314–2323.
Sezer, I. and Bilgin, A. (2008) ‘Exergy analysis of SI engines’, Int. J. Exergy, Vol. 5, No. 2,
pp.204–217.
Sheehan, J., Camobreco, V., Duffield, J., Graboski, M. and Shapouri H. (1998) Life Cycle
Inventory of Biodiesel and Petroleum Diesel for Use in an Urban Bus, National Renewable
Energy Laboratory, http://www.nrel.gov/docs/Legosti/fy98/24089.pdf
Srivathsan, V.R., Srinivasan, L.N. and Karuppan, M. (2008) ‘An overview of enzymatic production
of biodiesel’, Bioresource Technology, Vol. 99, No. 10, pp.3975–3981.
Toyota Motor Corporation (2006) Well-to-Wheel Analysis of Greenhouse Gas Emissions of
Automotive Fuels in the Japanese Context, Available on line: /http://www.mizuhoir.co.jp/
english/knowledge/documents/wtwghg041130.pdfS
Usta, N. (2005) ‘An experimental study on performance and exhaust emissions of a diesel engine
fuelled with tobacco seed oil methyl ester’, Energy Convers. Manage., Vol. 46, pp.2373–2386.
Van Gerpen, J.H. and Shapiro, H.N. (1990) ‘Second law analysis of diesel engine combustion’,
J. Eng. Gas Turbines Power, Vol. 112, No. 1, pp.129–137.
Effects of rapeseed oil addition to a diesel fuel 123

Zheng, M., Mulenga, M.C., Reader, G.T., Tan, Y., Wang, M. and Tjong, J. (2007) Neat Biodiesel
Fuel Engine Tests and Preliminary Modeling, SAE 2007 010616.
Zheng, M., Mulenga, M.C., Reader, G.T., Wang, M. and Ting, D.S.K. (2006) Influence of
Biodiesel Fuel on Diesel Engine Performance and Emissions in Low Temperature
Combustion, SAE 2006-01-3281.

Bibliography
Alasfour, F.N. (1997) ‘Butanols – a single cylinder engine study: availability analysis’, Appl.
Therm. Eng., Vol. 17, No. 6, pp.537–549.
Alkidas, A.C. (1988) ‘The application of availability and energy balances to a diesel engine’,
J. Eng. Gas Turbines Power, Vol. 110, pp.462–468.
Caton, J.A. (2000) A Review of Investigations using the Second Law of Thermodynamics to Study
Internal-Combustion Engines, SAE Tech. Pap. 2000-01-1081, 2000.
Ferguson, C.R. and Kirkpatrick, A.T. (2001) Internal Combustion Engines, 2nd ed., Wiley,
Singapore, pp.285–286.
Hovelius, K. and Hansson, P-A. (1999) ‘Energy and exergy analysis of rape seed oil rape seed
methyl (RME) production under Swedish conditions’, Biomass Bioenergy, Vol. 17,
pp.279–290.
Puhan, S. and Vedaraman, N., Sankaranarayanan, G. and Bharat Ram, B.V. (2005) ‘Performance
and emission study of Mahua oil (Madhuca indica oil) ethyl ester in a 4-stroke natural
aspirated direct injection diesel engine’, Renew. Energy, Vol. 30, pp.1269–1278.
Rakopoulos, C.D. (1993) ‘Evaluation of a spark ignition engine cycle using first and second-law
analysis techniques’, Energy Convers. Manage., Vol. 34, pp.1299–1314.

Nomenclature
B Flow availability (kJ/kg)
E ex Exhaust power (kW)
etm Thermo-mechanical exergy (kJ/kmol)
ekim Chemical exergy (kJ/kmol)
H Specific enthalpy (kJ/kg)
I Total irreversibility (kW)
LHV Lower heating value
m Mass flow rate (kg/s)
MEXP Maximum extractable power
N Engine speed (rpm)
n f Fuel molar rate (kg/s)

S p Rate of entropy production (kW/K)


Pfme Friction mean effective pressure (bar)
R Universal gas constant (kJ/kg K)
W f Friction power (kW)
124 E. Buyukkaya et al.

Q Rate of heat transfer (kW)


S Specific entropy (kJ/kgK)
T Temperature (K)
T Mean temperature (K)
Q ex ,max Maximum extractable (available) exhaust power (kW)
Wlost Lost power (kW)
W s Shaft power (kW)
W s ,max Maximum extractable shaft power (kW)
ηI First-law efficiency

η II Second-law efficiency

ε Compression ratio
o
∆H pr Enthalpy of the products (kJ/kg)

λ Excess air ratio


Subscripts
a Air
ch Chemical
com Combustion
e Standard atmosphere
env Environment
Ex Exhaust
F Fuel
I Species
in Inlet
out Outlet

You might also like