You are on page 1of 9

ARTICLE IN PRESS

Energy 32 (2007) 20722080


www.elsevier.com/locate/energy

Engine performance and emissions of a diesel engine operating on


diesel-RME (rapeseed methyl ester) blends with EGR
(exhaust gas recirculation)
A. Tsolakisa,, A. Megaritisb, M.L. Wyszynskia, K. Theinnoia
a

Mechanical and Manufacturing Engineering, School of Engineering, University of Birmingham, Birmingham B15 2TT, UK
b
Mechanical Engineering, School of Engineering and Design, Brunel University, West London, Uxbridge UB8 3PH, UK
Received 25 January 2007

Abstract
The effects of biodiesel (rapeseed methyl ester, RME) and different diesel/RME blends on the diesel engine NOx emissions, smoke, fuel
consumption, engine efciency, cylinder pressure and net heat release rate are analysed and presented. The combustion of RME as pure
fuel or blended with diesel in an unmodied engine results in advanced combustion, reduced ignition delay and increased heat release rate
in the initial uncontrolled premixed combustion phase. The increased in-cylinder pressure and temperature lead to increased NOx
emissions while the more advanced combustion assists in the reduction of smoke compared to pure diesel combustion. The lower caloric
value of RME results in increased fuel consumption but the engine thermal efciency is not affected signicantly. When similar
percentages (% by volume) of exhaust gas recirculation (EGR) are used in the cases of diesel and RME, NOx emissions are reduced to
similar values, but the smoke emissions are signicantly lower in the case of RME. The retardation of the injection timing in the case of
pure RME and 50/50 (by volume) blend with diesel results in further reduction of NOx at a cost of small increases of smoke and fuel
consumption.
r 2007 Elsevier Ltd. All rights reserved.
Keywords: Biodiesel; Emissions; Combustion; EGR; Retarded injection

1. Introduction
Research on alternative renewable fuels has become very
important worldwide due to concerns about the effects of
fossil fuel usage on global warming. They can be made
from renewable raw material and can offer reduction of
fossil fuel consumption. Modern diesel engines require a
clean burning, stable fuel that performs well under a
variety of operating conditions.
Partial or complete replacement of petroleum-based
fuels for diesel engines has been seriously studied in
various parts of the world [13]. Using chemically
unaltered vegetable oil directly can cause performance
problems because of their high viscosity and low volatility
[2,3]. A feasible solution is to trans-esterify the oils with
Corresponding author. Tel.: +44 121 4144170; fax: +44 121 4143958.

E-mail address: a.tsolakis@bham.ac.uk (A. Tsolakis).


0360-5442/$ - see front matter r 2007 Elsevier Ltd. All rights reserved.
doi:10.1016/j.energy.2007.05.016

methanol (or ethanol) to form esters. Biodiesel is an


alternative fuel that can be used directly in any existing,
unmodied diesel engine as pure or blended with diesel.
Because it has similar properties to petroleum diesel fuel,
biodiesel can be blended in any ratio with diesel [4,5].
The combustion of biodiesel fuel in compression ignition
(CI) engines in general results in lower smoke, particulate
matter, carbon monoxide and hydrocarbon emissions
compared to standard diesel fuel combustion while the
engine efciency is either unaffected or improved [16].
Recent engine testing studies continue to present NOx
emission results that vary widely and appear to depend
upon engine manufacturer or engine design [7]. In general,
the combustion of biodiesel in unmodied CI engines
equipped with pumplinenozzle fuel systems results in
advancement of the combustion process compared to diesel
fuel. This is due to differences in chemical and physical
properties of the fuels [816]. For engines equipped with

ARTICLE IN PRESS
A. Tsolakis et al. / Energy 32 (2007) 20722080

unit injector or common rail injection systems, this


relationship is not so clear [7]. Although higher overall
cylinder temperature is an indicator of higher NOx, it must
be stated that the temperature distribution in the cylinder is
of more importance, thus causing in some cases the NOx to
increase and others to decrease [14,15]. Another feature of
biodiesel is thatas recent exhaust-gas fuel reforming
studies have shownit can signicantly improve the
hydrogen production and reforming process efciency
compared to diesel [17].
In this study, the combined effects of rapeseed methyl
ester (RME) biodiesel combustion with the incorporation
of exhaust gas recirculation (EGR) on the engine
performance and emissions are analysed and compared
with the results obtained from the engine operating on
ultra-low sulphur diesel (ULSD). The diesel engine used in
this work is equipped with pumplinenozzle type fuel
injection system; as such the effects may not be fully
relevant to common rail or unit injection equipped engines.
2. Experimental setup
The experiments were carried out on a Lister-Petter TR1
engine. The engine is a 773-cm3, naturally aspirated, aircooled, single-cylinder direct injection diesel engine. The
main engine specications are: bore 98.4 mm, stroke
101.6 mm, conrod length 165.0 mm, compression ratio
15.5, maximum power 8.6 kW at 2500 rpm and maximum
torque 39.2 N m at 1800 rpm. The fuel injection system
used in this study is as given by the manufacturer and has a
three-hole nozzle with hole diameter of 0.25 mm, located
near the combustion chamber centre with an opening
pressure of 180 bar. The engine piston is a bowl-in-piston
design. An electric dynamometer with a motor and a load
cell was coupled to the engine and used to load and motor
the engine. The full engine test rig has been described in
detail in previous publications (e.g. Ref. [4]).
The EGR ow was controlled manually by a valve and
the EGR level was determined volumetrically as the
percentage reduction in volume ow rate of air at a xed
engine operating point. The inlet charge was kept as much
as possible at the same temperature (in the range of
2530 1C) when using EGR, so that the effects of the inlet
charge temperature on the ignition delay and combustion
process could be eliminated.
A KISTLER 6125B pressure transducer (1% measurements accuracy), mounted ush at the cylinder head and
connected via a KISTLER 5011 charge amplier to a
National Instruments data acquisition board, was used to
record the cylinder pressure. The crankshaft position was
measured using a digital shaft encoder. The test rig
included other standard engine instrumentation such as
thermocouples to measure oil, air, inlet manifold and
exhaust temperatures and pressure gauges mounted at
relevant points. Normal engine test bed safety features
were also included. Atmospheric conditions (humidity,
temperature, pressure) were monitored during the tests.

2073

The maximum fuel injection pressure was measured using


another pressure transducer that was tted to the highpressure fuel pipe between the pump and the injector.
Data acquisition and combustion analysis were carried
out using in-house developed LabVIEW-based software.
Output from the analysis of 200 consecutive engine cycles
included peak engine cylinder pressure, values of indicated
mean effective pressure (IMEP), percentage coefcient of
variation (% COV) of IMEP, average values and
percentage COV of peak cylinder pressures, average crank
angle for ignition delay, burn duration, 50% burn point,
etc. The COVs of IMEP and peak cylinder pressure were
used as criteria for combustion stability (cyclic variability).
In all tests presented here, the COVs were below 2%.
An AVL Digas4000 analyser was used to measure NOx,
CO, and CO2, by NDIR (non-dispersive infrared gas
analysis), and oxygen (O2) concentrations in the exhaust
(electrochemical method). Additional measurements of
NONOx were carried out using chemiluminescence
analyzer. Hydrocarbons were measured with heated ame
ionization detector (FID). Smoke was measured using an
EFAW 68A Bosch smoke meter. Table 1 shows the
accuracy of the measurements and the uncertainty of the
computed results of the various parameters.
The combustion of different ULSDRME mixtures was
examined at two engine operating conditions: 1500 rpm
engine speed, 20 N m torque; and 1500 rpm engine speed,
30 N m torque. As shown in Table 2, the corresponding
IMEPs were 4.5 and 6.1 bar. For both engine conditions,
pure biodiesel (RME) and two ULSD/RME blends (B20
and B50) were examined. The ULSD and RME specications are shown in Table 3. Table 4 shows the percentages
of ULSD and RME by volume and by mass in the different
fuel blends.
The results regarding emissions and engine performance
obtained with B20, B50 and pure RME (B100) were
compared with the results obtained with ULSD at the same
engine operating condition. Two EGR percentages of
Table 1
Accuracy of measurements and uncertainly of computed results
Measurements

Accuracy

NOx
CO
CO2
HC
Smoke (BSN)
Time
Speed
Torque

1 ppm (Digas 4000) and


5 ppm (chemiluminescence)
0.01%
0.01%
1 ppm
0.2
0.5%
5 rpm
0.2 N m

Computed results

Uncertainty (%)

Fuel volumetric rate


Power
Brake specic fuel consumption
Efciency

1
1
1.5
1.5

ARTICLE IN PRESS
A. Tsolakis et al. / Energy 32 (2007) 20722080

2074
Table 2
Engine conditions and fuel mixtures tested
Engine
condition

Speed
(rpm)

IMEP
(bar)

Torque
(N m)

Fuel

1500

4.5

20

ULSD
B20
B50
RME

1500

6.1

30

ULSD
B20
B50
RME

Table 3
Fuel properties
Fuel analysis

Method

Ultra-low
sulphur diesel
(ULSD)

Rapeseed
methyl ester
(RME)

Cetane number
Density at 15 1C
(kg/m3)
Viscosity at 40 1C
(cSt)
50% distillation
(1C)
90% distillation
(1C)
LCV (MJ/kg)
Sulphur (mg/kg)

ASTM D613
ASTM D4052

53.9
827.1

54.7
883.7

ASTM D445

2.467

4.478

ASTM D86

264

335

ASTM D86

329

342

ASTM D2622

42.7
46

39
5

Table 4
ULSD and RME volume and mass percentages of the tested fuel mixtures
Fuel

ULSD vol.
(%)

ULSD mass
(%)

RME vol.
(%)

RME mass
(%)

ULSD
B20
B50
RME

100
80
50
0

100
78.9
48.3
0

0
20
50
100

0
21.1
51.7
100

1070.6% and 2070.6% were used. The exhaust gas added


to the engine inlet as EGR was cooler than the engine
exhaust gas but it was at a higher temperature than the
inlet air charge. The effects of retarding the injection timing
by 31 crank angle (CA) on the NOx, smoke emissions and
engine performance for the engine operation on B50 and
RME were also examined. The standard injection timing
was 221 CA before top dead centre (BTDC).
3. Results
3.1. Fuel blends combustion
Figs. 1 and 2 show the NOx, Bosch Smoke Number
(BSN), HC and CO emissions, obtained with ULSD, B20,

B50 and RME with different EGR percentages (0%, 10%


and 20%) for the two engine operating conditions.
Similarly, the fuel and EGR effects on the engine brake
specic fuel consumption (BSFC) and engine efciency are
shown in Fig. 3. The effects of EGR are discussed in
Section 3.2.
The increase of the RME percentage in the fuel blend
reduces the smoke, HC and CO but increases the NOx for
both engine conditions. The increased fuel consumption
(BSFC) with the increase of the biodiesel content of the fuel
blend is mainly due to the lower caloric value of RME
(37 MJ/kg) compared to ULSD (42.6 MJ/kg), which
resulted in the increased fuel ow rates. The increased fuel
ow in the case of B20, B50 and RME in conjunction with
the higher density of RME (883.7 kg/m3) compared to
ULSD (827.1 kg/m3) resulted in the increased fuel mass
used and hence higher BSFC. However, the engine thermal
efciency n (%) was not affected signicantly from the
combustion of fuel blends.
Fig. 4 shows the cylinder pressure and net heat release
rate (NHRR) patterns from the combustion of the four
fuels (ULSD, B20, B50 and RME) for both engineoperating conditions. Although the NHRR does not take
into account the heat loss during the combustion process, it
is generally used (numerous publications) to provide an
acceptable indication of the combustion history of fuels
(assuming that the heat losses are not much different). The
increase of the RME percentage in the fuel blend
(vol. 20%, 50% and 100% pure RME) appears to reduce
the ignition delay, increase the rate of fuel burnt in the
premixed phase and shift the start of combustion to an
earlier stage and hence increase the in-cylinder pressure
compared to ULSD combustion. The overall combustion
duration CA was 301, 301, 281 and 281 for the engine
operation at 4.5 bar IMEP for ULSD, B20, B50 and RME,
respectively. For the engine operation at 6.1 bar IMEP the
combustion duration was 341 CA for all the fuel blends.
Biodiesel such as RME is less compressible than diesel
fuel, so the pressure in the pumplinenozzle type fuel
injection system can develop faster, and pressure waves can
propagate faster in biodiesel than diesel even at the same
nominal pump timing [913]. As a result, the injection of
biodiesel fuel starts earlier with higher pressure and rate
and at the same CA degree, the mass of biodiesel injected is
higher than the corresponding mass of diesel. Furthermore,
the increased RME viscosity leads to reduced fuel losses
during the injection process, compared to lower-viscosity
fuels such as ULSD. The reduction in fuel losses during the
injection process leads to a faster evolution of pressure and,
thus, advances the injection timing [8]. It has to be claried
here that the actual injection timing associated with the
RME and ULSD/RME blends was not quantied and the
ignition delays mentioned in this paper are dened as
the interval between 221 CA BTDC (standard injection
timing) and fuel ignition.
The combination of the increased injection pressure and
similar cetane number of RME compared to ULSD (cetane

ARTICLE IN PRESS
A. Tsolakis et al. / Energy 32 (2007) 20722080

400

1000
0% EGR

10% EGR

20% EGR

350

800

300
CO (ppm)

NOx (ppm)

2075

600
400

250
200
150
100

200

50
0

0
ULSD

B20

B50

ULSD

B100
500

2.0

400
HC (ppm)

2.5

BSN

1.5
1.0

B20

B50

B100

B50

B100

300
200
100

0.5
0.0

0
ULSD

B20

B50

ULSD

B100

B20

Fig. 1. Effects of fuel blend composition and EGR on the engine exhaust emissions. IMEP 4.5 bar.

1400
1200

500
0% EGR

10% EGR

20% EGR
400
CO (ppm)

NOx (ppm)

1000
800
600

300
200

400
100

200
0

0
ULSD

B20

B50

B100
500

4.0

400
HC (ppm)

5.0

3.0
BSN

ULSD

2.0
1.0

B20

B50

B100

300
200
100

0.0

0
ULSD

B20

B50

B100

ULSD

B20

B50

Fig. 2. Effects of fuel blend composition and EGR on the engine exhaust emissions. IMEP 6.1 bar.

B100

ARTICLE IN PRESS
A. Tsolakis et al. / Energy 32 (2007) 20722080

2076

320

300
0% EGR

IMEP = 6.1 bar

20% EGR
280

IMEP = 4.5 bar

BSFC (g/kWh)

BSFC (g/kWh)

300

10% EGR

280

260
240

260

220
200

240
ULSD

B20

B50

ULSD

B100

B20

B50

B100

B50

B100

36.0

34.0
IMEP = 4.5 bar

35.5

IMEP = 6.1 bar

Efficiency n (%)

Efficiency n (%)

33.0
32.0
31.0
30.0

35.0
34.5
34.0
33.5
33.0
32.5

29.0
ULSD

B20

B50

B100

ULSD

B20

70
60
50

90

60

80

50
40
30

40

20

30

10

20

0
IMEP 4.5 bar

10
0
20

70

10
20

10

0
10
20
Crank Angle (deg)

30

40

Cylinder Pressure (bar)

ULSD
B20
B50
B100

NHRR (J/degCA)

Cylinder Pressure (bar)

80

ULSD
B20
B50
B100

70
60
50
40
30
20
IMEP 6.1 bar

10
0
20

10

0
10
20
Crank Angle (deg)

30

80
70
60
50
40
30
20
10
0
10
20

NHRR (J/degCA)

Fig. 3. Effects of fuel blend composition and EGR on the engine BSFC and efciency. IMEP 4.5 and 6.1 bar.

40

Fig. 4. Effects of fuel blend composition on the cylinder pressure and NHRR. IMEP 4.5 and 6.1 bar.

number of both fuels is approximately 54) results in an


increased amount of fuel undergoing premixed combustion
at an earlier stage as indicated by the NHRR patterns
presented in Fig. 5. For the 4.5 bar IMEP engine operation
the injection pressures were 251, 254, 257 and 262 bar for
ULSD, B20, B50 and RME, respectively. Similarly, for the
engine operation at 6.1 bar IMEP the fuel injection
pressures were 253, 255, 260 and 265 bar for ULSD, B20,
B50 and RME, respectively.
The shorter ignition delay and the increased amount of
fuel undergoing premixed combustion results in higher
cylinder pressure and hence temperature. The rate of

formation of NOx emissions in diesel engines is primarily a


function of ame temperature, which is closely related to
the peak cylinder pressure and hence temperature. Szybist
et al. [1] have shown that when the impact of fuels and
injection timing are considered, the timing of the maximum
cylinder temperature correlates with the trends of the NOx
emissions. In addition, the higher density of RME
compared to ULSD in conjunction with the increased
injection pressure, results in the delivery of a higher
amount of fuel at the same injection setting conditions.
Combustion, therefore, takes place over a shorter period of
time, and this possibly allows less time for cooling by heat

ARTICLE IN PRESS
A. Tsolakis et al. / Energy 32 (2007) 20722080

30

30

IMEP 4.5 bar

IMEP 6.1 bar

25
NHRR (J/degCA)

NHRR (J/degCA)

25
B100
20
15
B50

2077

B20

10
ULSD
5

B100

20
B50
15
B20

10

ULSD

0
9

Crank Angle (deg)

7
6
5
Crank Angle (deg)

Fig. 5. Premixed combustion NHRR patterns with different fuel blends. IMEP 4.5 and 6.1 bar.

transfer and dilution. The above-described trends can


explain the higher NOx formation associated with the
combustion of RME.
Furthermore, the oxygen content of the biodiesel may
contribute to improved fuel oxidation even in locally rich
fuel combustion zones, thus resulting in the reduction of
smoke. The increased oxygen concentration in the sootforming region at the centre line of the fuel jet and the
reduced residence time of a fuel element in the sootforming jet may enhance the soot reduction. The increased
availability of oxygen from the combustion of biodiesel in
the rich premixed reaction zones results in lower production of soot precursor species and hence in reduced rates of
the soot producing reactions. The amount of soot
precursor species available to produce soot strongly
depends on the amount of oxygen available in the mixture.
When sufcient oxygen is available, soot precursors species
react with molecular oxygen or oxygen-containing radicals
(like OH, O) and eventually produce CO rather than
aromatics and soot [6]. The reduction of smoke with RME
compared to ULSD can also be attributed up to a certain
extent to the signicantly lower sulphur content of RME
(5 mg/kg) compared to ULSD (46 mg/kg).
Several reasons have been proposed to explain the
decrease in HC and CO emissions when substituting
conventional diesel by biodiesel. An obvious reason for
the reduction of both HC and CO emissions in the case of
using pumplinenozzle injection system is the advanced
injection timing with the use of biodiesel. Rakopoulos et al.
[18] have concluded in their review work that both HC and
CO emissions decrease as the oxygen in the combustion
chamber increases, either with oxygenated fuels such as
biodiesel or oxygen-enriched air.
3.2. Use of EGR
In all fuelling cases, EGR addition resulted in increased
smoke, HC, CO and reduced NOx emissions as illustrated
earlier in Figs. 1 and 2. The use of EGR was more effective
(higher reduction of NOx with lower increase of smoke) in

the case of B20, B50 and RME combustion compared to


ULSD. It is clear from Figs. 1 and 2 that with 20% EGR,
the NOx emissions were reduced to values similar to those
of ULSD with the same EGR percentage while the smoke
levels were kept at considerably lower values. The use of
EGR resulted in increased fuel consumption and reduced
efciency as expected (Fig. 3). The effects of EGR on the
emission trends and engine performance were more
pronounced in the case of the high load (6.1 bar IMEP)
engine operating condition.
The use of EGR seems to suit better the B20, B50 and
RME combustion, as apart from resulting in the higher
NOx reduction, it maintained the smoke (soot, particulate
matter) at relatively low levels. One reason for the higher
NOx reduction is probably the different composition of the
engine exhaust gas obtained from the combustion of RME
and ULSD. The increased fuel consumption with biodiesel
resulted in increased H2O and CO2 in the engine exhaust
gas. For example, the CO2 contents of the engine exhaust
gas from the combustion of ULSD, B20, B50 and RME
were 4.86%, 4.90%, 5.00% and 5.10%, respectively, for
engine condition 1 (IMEP 4.5 bar) and increased to
6.60%, 6.90%, 7.00% and 6.90%, respectively, for engine
condition 2 (IMEP 6.1 bar). Furthermore, when a
constant amount of EGR is added, the relative air/fuel
ratio l (lambda) moves in the direction of stoichiometry
(l 1) as the biodiesel quantity in the fuel blend is
increased. For example, for the engine operating condition
of 4.5 bar IMEP with 20% EGR, the measured l values
were 2.43, 2.29, 2.17 and 2.00 in the cases of ULSD, B20,
B50 and RME fuelling, respectively. Similarly for the
6.1 bar IMEP with 20% EGR, the l values were 1.72, 1.68,
1.65 and 1.64 for ULSD, B20, B50 and RME fuelled
engine, respectively.
Fig. 6 shows the heat release rate (NHRR) patterns of
the premixed combustion phase from the engine operating
with ULSD and RME with 0%, 10% and 20% EGR. In
the case of the ULSD fuelled engine, the fuel injection
timing was not optimised for operation with EGR. In this
case, advanced injection timing will be required to enhance

ARTICLE IN PRESS
A. Tsolakis et al. / Energy 32 (2007) 20722080

2078

NHRR (J/degCA)

the effects of the residual gas on the combustion process


and hence achieve NOx emission reductions while keeping
the smoke and fuel consumption levels low.
The use of EGR in the case of biodiesel mixtures
combustion was benecial in terms of NOx emissions
because it resulted in the retardation of the combustion of
RME, which is advanced, compared to ULSD. The use of
EGR in the case of the biodiesel-fuelled engine resulted in

the increase of the ignition delay and shifted the start and
end of combustion to later stages in the compression stroke
and expansion stroke, respectively. For both fuels, the start
of combustion in the case of 10% EGR was the same as in
the case where 20% EGR was added. This effect is
attributed to the increased inlet charge temperature by
5 1C due to the higher temperature of EGR compared to
the inlet air.

35

3.3. Retarded injection timing

30

Figs. 7 and 8 show the effects of the use of retarded


injection timing by 31 CA on emissions (NOx, BSN, HC
and CO) and performance (BSFC and engine efciency)
with B50 and RME fuelling for the two tested engine
operating conditions. For comparison, the data for ULSD
at standard injection timing (221 CA BTDC) are also
shown.
The retardation of the injection timing resulted in
reduced NOx emissions and increased smoke, CO and
HC emissions. In the case of 4.5 bar IMEP engineoperating condition, the reduction of NOx emissions was
10% and 30% while the BSN increased by about 25% and
55% for B50 and pure RME, respectively. In the case of
6.1 bar IMEP engine-operating condition, the reduction of
NOx emissions was around 20% and the increase of BSN
was around 10% for both B50 and RME.
The use of retarded injection timing did not affect
signicantly the fuel consumption and engine efciency.

25
20
15
10
5
0
10

7
6
5
Crank Angle Degree

RME0% EGR
RME10% EGR
RME20% EGR

ULSD0% EGR
ULSD10% EGR
ULSD20% EGR

Fig. 6. Effects of EGR addition on the premixed combustion NHRR


patterns for the cases of ULSD and RME fuelling. IMEP 6.1 bar.

1400
1200

250

IMEP 4.5 bar


IMEP 6.1 bar

200
CO (ppm)

NOx (ppm)

1000
800
600

150
100

400
50

200
0
ULSD

B50

B50 Ret.

B100

B100 Ret.

3.0
2.5

B50

B50 Ret.

B100

B100 Ret.

ULSD

B50

B50 Ret.

B100

B100 Ret.

400
HC (ppm)

2.0
BSN

ULSD
500

1.5
1.0

300
200
100

0.5
0.0

0
ULSD

B50

B50 Ret.

B100

B100 Ret.

Fig. 7. Effects of retarded injection timing by 31 CA on the engine exhaust emissions for the cases of B50 and RME fuelling. IMEP 4.5 and 6.1 bar.

ARTICLE IN PRESS
A. Tsolakis et al. / Energy 32 (2007) 20722080

In the cases of both RME and B50 combustion, the


retarded injection timing resulted in a reduction of the
engine efciency by 0.10.2% for the two engine operating
conditions. Correspondingly, the BSFC was increased for
both B50 and pure RME.
Figs. 9 and 10 show the cylinder pressure and heat
release rate for B50 and RME fuelling, respectively, with
and without retarded injection timing for the two tested

engine operating conditions. For comparison, the combustion of ULSD at standard injection timing (221 CA BTDC)
is also shown.
The retarded injection timing resulted in the reduction of
the cylinder pressure at levels lower or similar to those of
ULSD combustion. The start of combustion was shifted to
a later stage in the compression stroke and the end of
combustion occurred later in the expansion stroke.

36.0

400
IMEP 4.5 bar
IMEP 6.1 bar

350

35.0
Efficiency n (%)

300
BSFC (g/kWh)

2079

250
200
150
100

34.0
33.0
32.0
31.0
30.0

50

29.0

0
ULSD

B50

B50 Ret.

B100

ULSD

B100 Ret.

B50

B50 Ret.

B100

B100 Ret.

Fig. 8. Effects of retarded injection timing by 31 CA on the engine BSFC and efciency for the cases of B50 and RME fuelling. IMEP 4.5 and 6.1 bar.

ULSD
B50
B50 retard

60

50

50
30
40
30

10

20
10
20

10
10

0
10
20
Crank Angle (deg)

30

70

70

60

50

50
30

40
30

10

20
10
20

40

ULSD
B50
B50 retard

IMEP 6.1 bar

NHRR (J/degCA)

IMEP 4.5 bar

Cylinder Pressure (bar)

70

80

70

NHRR (J/degCA)

Cylinder Pressure (bar)

80

10
10

10

20

30

40

Crank Angle (deg)

Fig. 9. Effects of retarded injection timing by 31 CA on the cylinder pressure and NHRR for B50 combustion. IMEP 4.5 and 6.1 bar.

RME retard

60

50

50

40

40

30

30

20

20

10
0

10
0
20

10
10

0
10
20
Crank Angle (deg)

30

70

60

RME

40

70

ULSD

IMEP 6.1 bar

60

RME
RME retard

60

50

50

40

40

30

30

20

20

10

10

0
20

NHRR (J/degCA)

ULSD

Cylinder Pressure (bar)

IMEP 4.5 bar

NHRR (J/degCA)

Cylinder Pressure (bar)

70

80

70

80

10
10

0
10
20
Crank Angle (deg)

30

40

Fig. 10. Effects of retarded injection timing by 31 CA on the cylinder pressure and NHRR for RME combustion. IMEP 4.5 and 6.1 bar.

ARTICLE IN PRESS
2080

A. Tsolakis et al. / Energy 32 (2007) 20722080

The heat release rate patterns in the premixed combustion


phase and the peak values of the heat release rate in the
diffusion combustion phase were not affected signicantly
but they were shifted to later stages in the engine cycle.
With the retarded injection timing, the overall combustion
duration was shorter possibly due to early ame quenching
resulting in incomplete combustion (normally not likely in
diesel engines) and increased smoke.
4. Conclusions
The combustion of RME pure or blended with ULSD at
20% and 50% by volume (B20 and B50) was investigated
and the main ndings are summarised below.
The combustion of RME, B20 and B50 in the
unmodied engine with pumplinenozzle injection system
resulted in advanced combustion compared to ULSD, as
indicated by the presented heat release rate patterns. The
ignition delay was reduced while the initial uncontrolled
premixed combustion phase (uncontrolled heat release
phase) was increased. This resulted in increased cylinder
pressure and temperature and hence early fuel ignition.
The advanced RME combustion resulted in the reduction of smoke, HC and CO while both NOx emissions and
fuel consumption were increased. The combustion of
different fuel blends did not affect signicantly the engine
efciency. The increased amount of oxygen in the RME
molecule and hence in the locally fuel-rich combustion
zones is believed to be an additional reason for the reduced
smoke. The increase of the fuel consumption is mainly due
to the lower caloric value (LCV) of RME compared to
ULSD.
The use of EGR was more effective in the case of
biodiesel blends combustion compared to ULSD combustion. The NOx emissions were reduced at levels similar to
those of ULSD with the use of similar volumetric
percentages of EGR while the smoke was kept low. The
main reasons for the higher NOx reduction with the use of
EGR in the case of biodiesel fuelling are: (i) the increased
CO2 dilution as more CO2 enters the combustion as part of
the EGR compared to ULSD, (ii) the lower relative air/fuel
ratio l with biodiesel compared to operation with diesel
(for the same EGR level), and (iii) the retardation of the
already advanced combustion from the use of biodiesel.
The retardation of the injection timing in the case of B50
and pure RME combustion resulted in small increases of
smoke and fuel consumption and a signicant reduction of
NOx emissions.

References
[1] Szybist JP, Kirby SR, Boehman AL. NOx emissions of alternative
diesel fuels: a comparative analysis of biodiesel and FT diesel. Energy
Fuels 2005;19:148492.
[2] Graboski MS, McCormick LR. Combustion of fat vegetable oil
derived fuels in diesel engines. Prog Energy Sci 1998;24:12564.
[3] Graboski MS, Ross JD, McCormick RL. Transient emissions from
No. 2 diesel and biodiesel blends in a DDC series 60 engine. SAE
paper no. 961166, 1996.
[4] Tsolakis A, Megaritis A, Wyszynski ML. Applications of exhaust gas
fuel reforming in compression ignition engines fuelled by diesel and
biodiesel fuel mixtures. Energy Fuels 2003;17:146473.
[5] Peterson CL, Reece DL. Emissions testing with blends of esters of
rapeseed oil fuel with and without a catalytic converter. SAE paper
no. 961114, 1996.
[6] Mueller CJ, Picket LM, Siebers DL, Pitz WJ, Westbrook CK, Martin
GC. Effects of oxygenates on soot processes in DI diesel engines:
experimental and numerical simulations. SAE paper no. 2003-011791, 2003.
[7] McCormick RL, Williams A, Ireland J, Brimhall M, Hayes RR.
Effects of biodiesel blends on vehicle emissions, National Renewable
Energy Laboratory-NREL/MP-540-40544, located /www.nrel.gov/
vehiclesandfuels/npbf/pdfs/40554.pdfS; October 2006.
[8] Choi CY, Bower GR, Reitz RD. Effects of biodiesel blended fuels
and multiple injections on D.I. diesel engines. SAE paper no. 970218,
1997.
[9] Rakopoulos CD, Hountalas DT. A simulation analysis of a DI diesel
engine fuel injection system tted with a constant pressure valve.
Energy Convers Manage 1996;37:13550.
[10] Arcoumanis C, Gavaises M, Yamanishi M, Oiwa J. Applications of
FIE computer model to an in-line pump-based injection system for
diesel engines. SAE paper no. 970348, 1997.
[11] Tat ME, Van Gerpen JH. Physical properties and composition
detection of biodiesel-diesel fuel blends. ASAE paper no. 026084,
2002.
[12] Szybist JP, Boehman AL. Behavior of a diesel injection system with
biodiesel fuel. SAE paper no. 2003-01-1039, 2003.
[13] Boehman AL, Morris D, Szybist J, Esen E. The impact of the bulk
modulus of diesel fuels on fuel injection timing. Energy Fuels
2004;18:187782.
[14] Scholl KW, Sorenson SC. Combustion of soybean oil methyl ester in
a direct injection diesel engine. SAE paper no. 930934, 1993.
[15] Rakopoulos CD, Antonopoulos KA, Rakopoulos DC, Hountalas
DT, Giakoumis EG. 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 2006;47:327287.
[16] Demirbas A. Progress and recent trends in bio-fuels. Prog Energy
Combust Sci 2007;37:118.
[17] Tsolakis A, Megaritis A. Exhaust gas assisted reforming of rapeseed
methyl aster for reduced exhaust emissions of CI engines. Biomass
Bioenergy 2004;27:493505.
[18] Rakopoulos CD, Hountalas DT, Zannis TC, Levendis YA. Operational and environmental evaluation of diesel engines burning
oxygen-enriched intake air or oxygen-enriched fuels: a review. SAE
paper no. 2004-01-2924, 2004.

You might also like