Sunteți pe pagina 1din 9

Fuel 175 (2016) 240–248

Contents lists available at ScienceDirect

Fuel
journal homepage: www.elsevier.com/locate/fuel

Effect of biodiesel saturation on soot formation in diesel engines


Zhi Wang a,⇑, Li Li a, Jianxin Wang a, Rolf D. Reitz b
a
State Key Laboratory of Automotive Safety and Energy, Tsinghua University, Beijing, China
b
Engine Research Center, University of Wisconsin-Madison, Madison, WI 53706-1687, USA

h i g h l i g h t s

 Numerical study on soot formation of biodiesel from different feedstocks.


 The acetylene production is proportional to the content of unsaturated FAME.
 The biodiesel with less C@C result in the lower formation of soot emissions.
 Reduced soot formation and improved soot oxidation were achieved for biodiesel.

a r t i c l e i n f o a b s t r a c t

Article history: To understand soot formation in a diesel engine fueled with different biodiesels, a numerical study was
Received 4 August 2015 performed using the KIVA-3V code, combined with a multi-step phenomenological soot model. The sim-
Received in revised form 10 February 2016 ulations were used to predict differences in soot formation for three various biodiesel feedstock types.
Accepted 13 February 2016
Good agreements on soot emissions were achieved in comparisons of engine experiments and simula-
Available online 19 February 2016
tions at various engine operating conditions. The experimental data and simulated results showed that
the degree of saturation and the oxygen content of biodiesel fuels are the major factors responsible for
Keywords:
biodiesel soot production. The reduction of soot mass concentration for biodiesel is achieved due to
Biodiesel
Soot
the suppressed soot formation process and improved oxidation rate compared with diesel. It is observed
Fuel saturation that the acetylene generated in the pyrolysis of biodiesel is proportional to the content of unsaturated
Diesel engine fatty acid methyl ester (the number of CAC double bonds). Among the three different biodiesel fuels,
the lowest soot tendency was found for the Jatropha Methyl Ester because of its lowest amount of unsat-
urated alkyl esters through both numerical modeling and diesel engine experiments.
Ó 2016 Elsevier Ltd. All rights reserved.

1. Introduction weight, and therefore reduce exhaust emissions. Most researchers


have reported that a decrease in particular matter (PM) was
Nowadays, conventional diesel engines and gasoline engines observed when using neat biodiesel or the biodiesel/diesel blends
face challenge of soot emissions and super-knock [1,2]. Increas- [7–10]. As well, the fuel composition has an important effect on
ingly stringent environmental regulations and energy consump- the formation of soot emissions in diesel combustion. Recently
tion have put forward a strong demand on seeking alternative numerical modeling has shown that oxygen in the fuel is beneficial
solutions to reduce dependence on conventional fossil fuels. for suppressing the soot precursor formation in fuel-rich regions,
Biodiesel, as an attractive renewable fuel, can be used directly in thus reducing soot production [11–14].
compression ignition engines which need almost no modification Biodiesel comprised of a variety of long-chain fatty acid alkyl
[3–5]. Usually, biodiesel refers to a mixture of alkyl esters that esters including saturated and unsaturated compounds. The pres-
are produced from various biological feedstocks. Methyl biodiesel ence of carbon–carbon double bonds in unsaturated mono-alkyl
is comprised primarily of methyl palmitate (C17H34O2), methyl esters appears to influence sooting tendency significantly. Sarathy
stearate (C19H38O2), methyl oleate (C19H36O2), methyl linoleate et al. [15] and Gaïl et al. [16] conducted experimental and compu-
(C19H34O2), and methyl linoleate (C19H32O2) [6]. Biodiesel is an tational studies of methyl butanoate (MB) and methyl crotonate
alternative fuel with oxygen content of approximately 10% by (MC) combustion using a jet stirred reactor (JSR) and an
opposed-flow diffusion flame. They found that the unsaturated C4
FAME (i.e., MC) shows higher sooting tendency than the saturated
⇑ Corresponding author. C4 FAME (i.e., MB) due to the presence of C@C bond in MC. Garner
E-mail address: wangzhi@tsinghua.edu.cn (Z. Wang).

http://dx.doi.org/10.1016/j.fuel.2016.02.048
0016-2361/Ó 2016 Elsevier Ltd. All rights reserved.
Z. Wang et al. / Fuel 175 (2016) 240–248 241

et al. [17] experimentally investigated the combustion characteris- phenomenological soot model. The combustion process and spatial
tic of methyl octenoate and methyl octanoate in shock tube. The distributions of soot-relevant intermediate species for different
positive correlation was found between the increased tendency of biodiesel fuels were studied in detail.
acetylene formation and the number of C@C bonds in the ester
structure or the degree of unsaturation. Feng et al. [18] also 2. Methodology
observed that the higher amount of the CAC double bond in
methyl esters results in greater soot propensity. 2.1. Experimental approach
Still much work needs to be conducted to gain a deep insight
into the effect of fuel composition on the soot emissions both The experiments were conducted using a Cummins ISDe4 diesel
experimentally and numerically. Especially, it is essential to under- engine equipped with a common rail fuel injection system. Table 1
stand how the saturation of biodiesel affects soot formation so as lists the specifications of the engine, and the detailed operating
to find the difference in soot emissions for different biodiesels. conditions are shown in Table 2. An AVLSPC-472 partial flow dilu-
Kinetic modeling provides an effective method to obtain thorough tion smart sampler was used to collect soot samples on filters.
understanding of the combustion features of fuel in internal com- Before and after sampling, the filters are weighed using an elec-
bustion engines [19]. Considering the complexity of real biodiesel, tronic scale with 10 lg sensitivity, which were kept under constant
simpler esters were usually used as surrogates of large methyl temperature (22 ± 3 °C) and relative humidity (45% ± 8%) condi-
esters during the combustion simulation. A detailed chemical tions. The total PM was separated into the dry soot and non-soot
kinetic model for the methyl butanoate (C5H10O2) was proposed fractions by extraction in dichloromethane (CH2Cl2) and then in
by Fisher et al. [20] to represent the characteristic of biodiesel. de-ionized water.
However, the simpler species do not adequately represent the The three different biodiesels studied were: Jatropha Methyl
combustion features of biodiesel due to the insufficient length of Ester (JME), Cottonseed Methyl Ester (CME), and Rubber seed oil
the carbon chain of the surrogate molecule. Therefore, Herbinet Methyl Ester (RME). Table 3 gives the properties of the reference
et al. [21] developed a combined oxidation model using a three diesel fuel and the biodiesels. Compared with the diesel fuel, the
component blends of methyl decanoate (MD), methyl-9- biodiesels have lower heating values, higher densities and viscosi-
decenoate (MD9D), and n-heptane, which is capable of matching ties. Moreover, the cetane number of JME is the highest while RME
the C/H/O ratio for the real fuels. Moreover, this detailed chemical has the lowest.
kinetic mechanism was validated against the jet stirred reactor
(JSR) results by Dagaut and coworkers [22], and well reproduced
2.2. Numerical methodology
the reactivity properties of rapeseed methyl ester. Later, Brakora
and Reitz [23] reduced the MD/MD9D detailed mechanism of Her-
2.2.1. Biodiesel reaction mechanism
binet, and applied it in multi-dimensional engine simulations. The
The multi-dimensional computations were performed using the
reduced model was validated against experimental results of spray
KIVA-3V code coupled with CHEMKIN II. The original biodiesel
and combustion at conventional diesel operating conditions.
combustion mechanism was developed by combining the mecha-
A series of complicated physical and chemical processes are
nisms of MD (C11H22O2) and MD9D (C11H20O2) from LLNL [6,21].
involved with the soot formation in compression ignition engines.
The biodiesel mechanisms used in this study was developed by
Thus, a good soot model is essential to better understand soot for-
Brakora et al. [35], which was reduced from a combined mecha-
mation in diesel engines fueled with biodiesel. The most widely
nism of MD, MD9D and n-heptane. Various biofuel compositions
used two-step soot model includes the soot formation expression
in biodiesels were represented by adjusting the mole fractions of
by Hiroyasu and Kadota [24] and the oxidation expression by Nagel
MD, MD9D, and n-heptane. As discussed above, biodiesel fuels
and Strickland Constable [25]. Since the formation process of the
derived from different feedstocks mainly contain five long-chain
empirical soot model is simplified and couldn’t reflect the effect
methyl esters: methyl palmitate (MP; C16:0), methyl stearate
of fuel composition on soot production. Further, it provides no
(MS; C18:0), methyl oleate (MO; C18:1), methyl linoleate (MLO;
knowledge about soot number density and agglomeration. Conse-
C18:2), and methyl linolenate (MLN; C18:3). The first two species
quently some more detailed multi-step phenomenological soot
(MP and MS) are saturated methyl esters, while the last three com-
models have been proposed [26–28]. Tao et al. [28] developed a
ponents contain one (MO), two (MLO), and three (MLN) double
phenomenological soot model. In this model, acetylene is involved
bonds, respectively. The physical properties of real biodiesel fuels
during precursor formation after fuel pyrolysis. Then precursor
are represented by those of the five methyl esters. While in terms
species formation, particle inception and coagulation, surface
of chemistry, the two saturated species was represented by MD
growth, and oxidation are considered as subsequent steps. Valida-
and the other unsaturated species was represented by MD9D.
tions of this model over wide ranges of engine conditions were pre-
The fractions of these five components in JME, CME and RME are
sented in [29–31]. Moreover, this soot model can be implemented
listed in Fig. 1. The five components are listed (left to right) in
in CFD codes for three-dimensional simulations where detailed
the order of decreasing saturation. This composition for the differ-
soot mechanisms may be too computationally expensive.
ent biodiesel fuels was also applied in the simulations. For exam-
McEnally and Pfefferle [32] found that the structure of biodiesel
fuel affects sooting tendencies in laboratory-scale flames. Previous
experimental investigations have proved that different FAME fuels Table 1
produce different levels of soot emissions in engine tests [33,34]. Engine specifications.
Few modeling studies capture biodiesel composition effects on Engine type Four cylinder DI
soot formation. Therefore, it is necessary to study the combustion Bore  stroke 107  124 mm
of different FAME fuels to account for the effect of fuel structure on Connecting rod 192 mm
soot formation. Displacement 4.5 L
Compression ratio 17.5
The purpose of this work was to investigate the influence of bio- Injection pressure 160 MPa
diesel composition on soot formation and oxidation. Three FAMEs Nozzle diameter 0.168 m
with different levels of saturation were tested using a diesel engine Nozzle number 8 holes
and the results were compared with CFD modeling results using a Intake valve close timing 155 °CA BTDC
Exhaust valve open timing 120 °CA ATDC
multi-component biodiesel mechanism coupled with a multi-step
242 Z. Wang et al. / Fuel 175 (2016) 240–248

Table 2 Table 5
Details of the operating conditions. Chemical processes in the current soot model.

Load Engine BMEP Intake Intake Injection Injection Chemical processes Expression
ratio speed (MPa) pressure temperature times quantity
1. Acetylene formation R_ 1
(%) (r/min) (MPa) (°C) per cycle Fuel! m
2 C2 H2
(mg) 2. Precursor formation R_ 2
C2 H2 ! 2Z R þ H2
25 1416 0.50 0.118 82 3 30 3. Particle inception R_ 3
100 1416 1.56 0.19 94 3 100 R!P
4. Particle coagulation R_ 4
xP!P
5. Surface growth R_ 5
P + C2 H2 !P + H2
Table 3 6. O2 oxidation R_ 6
P + O2!P + 2CO
Fuel properties.
7. OH oxidation R_ 7
P + OH!P + CO + 12 H2
Methods Diesel JME CME RME
8. Acetylene oxidation R_ 8
C2H2 + O2!2CO + H2
Density (kg/m3) GB/T2540 829 873 875 878
9. Precursor oxidation R_ 9
Cetane number GB/T386 59 57.2 54 51.7 R + OH!CO + 12 H2
Heating value (MJ/kg) GB/T384 42.7 39.8 39.6 39.4
Viscosity (40 °C mm2/s) GB/T265 – 6.71 6.37 6.13
Viscosity (20 °C mm2/s) GB/T265 3.954 – – –
Sulfur content (ppm) SH/T0689 26 2.4 6.3 1.5 diesel and diesel fuel. While in the diesel combustion modeling,
Oxygen content (wt.%) Element – 11.37 11.61 11.28 the physical properties of C14H30 were used to model the spray
analyser characteristics and the oxidation chemistry of n-heptane was
Hydrogen content (wt. SH/T0656 13.34 11.99 11.73 11.76
applied to simulate the combustion process of diesel fuel. Acety-
%)
Carbon content (wt.%) SH/T0656 85.92 76.64 76.36 76.96 lene was considered as the main product of fuel pyrolysis in the
multi-step soot model, which contributes precursor formation.
Furthermore, Acetylene formation for biodiesel is dependent on
80
the fuel structure and is predicted by the oxidation kinetics and
not by the soot model.
Fuel Composition [mole frction]

70
JME
60 CME 2.2.3. Engine modeling and simulating conditions
RME Taking symmetrical advantage of the centrally mounted eight-
50 hole equally-spaced injector hole nozzle, only a 45° sector of the
40
cylinder was simulated, and the simulations were performed using
sector meshes of the engine. Each computation was conducted
30 from the intake valve closing to the exhaust valve opening. The
mesh consisted of about 28,000 cells at top dead center (TDC), as
20
shown in Fig. 2, and was constructed based on the bowl geometry
10 of the Cummins engine. The simulation conditions were chosen
based on the experimental engine operation conditions.
0
C16:0 C18:0 C18:1 C18:2 C18:3
3. Results and discussions
Fig. 1. Composition of biodiesel fuels from three different feedstocks.

3.1. Validation of the soot model


ple, JME contains five components of 15% MP, 12% MS, 71% MO, 2%
MLO and 0% MLN. So for chemistry model, 100% JME would repre- The reduced multi-component biodiesel surrogate mechanism
sented as 13.5% MD, 36.5% MD9D and 50% n-heptane. coupled with the multi-step phenomenological soot model was
applied to predict the sooting behavior of the different biodiesel
fuels. Fig. 4 compares the in-cylinder pressure and heat release rate
2.2.2. Numerical models (HRR) curves between the experimental and simulation results for
The sub-models used in this study are listed in Table 4. The diesel and JME at 25% and 100% engine load. The injection timing
characteristics of turbulent flows were simulated using the RNG and injection pulse width for each injection are indicated using
k–e turbulence model. Spray atomization and droplet breakup the dotted line. As shown in Fig. 3, the model predictions for both
were modeled by Kelvin–Helmholtz and Rayleigh–Taylor model.
The nine-step phenomenological soot model, which was developed
by Tao et al. [28], was used to simulate the soot formation and oxi-
dation processes as listed in Table 5. The MD/MD9D/heptane
mechanism was used to simulate the combustion process for bio-

Table 4
The computational sub-models applied in the simulations.

Phenomenon Model Reference


Turbulence RNG k–e model [40]
Spray breakup Hybrid KH-RT model [41,42]
Evaporation Discrete multicomponent [43]
Combustion SpeedChem [44]
Droplet collision ROI collision model [35]
Soot formation Multi-step phenomenological [29]
Fig. 2. Computational grid of the combustion chamber for the 45° sector.
Z. Wang et al. / Fuel 175 (2016) 240–248 243

Fig. 3. Comparison of the in-cylinder pressure and heat release rate curves between the experimental results and simulation results for diesel and biodiesel under 25% and
100% load.

biodiesel and diesel are in good agreement with the experimental acteristics for all the fuels. Very little amount of soot is generated
results for all test cases, indicating that the MD/MD9D/heptane during the pilot-injection combustion phase because of its pre-
mechanism is sufficient to represent the biodiesel combustion for mixed lean combustion. Subsequently, the soot mass increases
these conditions. The fuel injection quantity and the injec- sharply during the main diffusion combustion, and decreases after
tion strategy for all fuels were kept the same, which means lower the main combustion stage due to soot oxidation. Similar phenom-
total energy for the biodiesel case because of its lower heating ena were observed at the post-injection combustion stage. Lower
value compared with diesel. As a consequence, the peak cylinder acetylene, precursor and soot peaks were observed for the JME fuel
pressure was reduced by 6.1% and 3.5% for JME under the 100% compared with diesel under both engine conditions. The eventual
and 25% load conditions, respectively. The heat release rate curves engine-out soot emission depends on the balance between soot
of all fuels have similar profiles, with a pilot-injection combustion formation and soot oxidation. The oxygen in the biodiesel sup-
phase followed by a main-injection and post-injection combustion presses soot emissions because of the reduced soot precursor for-
phase. mation and improved soot oxidation process, which is consistent
Fig. 4 compares the evolution of in-cylinder normalized acety- with the results found in Ref. [36]. Compared with diesel, JME exhi-
lene, precursor and soot mass between diesel and JME under the bits lower soot emissions by 25% at the 25% load condition and by
25% and 100% loads. Fig. 5 compares the equivalence ratio at the 80% at the 100% load condition, indicating that biodiesel reduces
timing of 20 °CA ATDC between the diesel and biodiesel under soot emission more effectively at high engine load. On the one
100% load. Fig. 6 compares the measured and simulated engine- hand, more fuels were injected into the cylinder at the higher
out soot emissions. As shown in Fig. 4, the predicted histories of engine load, leading to a broader rich mixture area during the com-
acetylene, precursor and soot mass show three-peak-shaped char- bustion. As a result, more carbon is available for soot production
244 Z. Wang et al. / Fuel 175 (2016) 240–248

8 4

25% load Diesel Diesel


Biodiesel
Acetylene (g/kg fuel) 6 3 100% load Biodiesel

Acetylene (g/kg fuel)


4 2

2 1

0
0
-30 -20 -10 0 10 20 30 40 50 60 -30 -20 -10 0 10 20 30 40 50 60
Crank Angle (( ° ) CA ATDC ) Crank Angle ((°)CA ATDC )

3 2.0
Diesel Diesel
25% load Biodiesel 100% load Biodiesel
Precursor (g/kg fuel)

1.5

Precursor (g/kg fuel)


2

1.0

1
0.5

0 0.0

-30 -20 -10 0 10 20 30 40 50 60 -30 -20 -10 0 10 20 30 40 50 60


Crank Angle (( ° ) CA ATDC ) Crank Angle ((°)CA ATDC )

120 50

100 40 Diesel
Diesel 100%load
25%load Biodiesel
Soot (g/kg fuel)

Biodiesel
Soot (g/kgfuel)

80
30
60
20
40
10
20
0
0
-30 -20 -10 0 10 20 30 40 50 60 -30 -20 -10 0 10 20 30 40 50 60
Crank Angle (( ° ) CA ATDC ) Crank Angle ((°)CA ATDC )

Fig. 4. Comparison of predicted acetylene, precursor and soot histories between the diesel and biodiesel under 25% and 100% load.

20°CA ATDC
0.5
Soot formation rate (g/cm3.s)

Diesel formation rate


B100 formation rate
0.4 100%load B100 oxygenation rate
Diesel oxygenation rate
Phi 0.3
3.0 (a) diesel
1.5
0.2

0.0 0.1

0.0
(b) biodiesel
-30 -20 -10 0 10 20 30 40 50 60
Crank Angle ((°)CA ATDC)

Fig. 5. Comparison of predicted equivalence ratio and soot formation rate between the diesel and biodiesel under 100% load.

because of the deficiency of oxygen. The presence of oxygen in the decrease soot formation. This also can be observed from the equiv-
biodiesel plays a more significant role in reducing soot formation. alence ratio distribution during the main combustion period
In addition, the molecular structure of oxygenated fuels is found to between the diesel and biodiesel under 100% load as shown in
Z. Wang et al. / Fuel 175 (2016) 240–248 245

0.04 the experimental results in this study. Similar results can also be
Diesel Simulation
found in Refs. [34,39]
Diesel Experiment Fig. 7 presents the net soot production of the three biodiesel
0.03 Biodiesel Simulation fuels, comparing the results as measured in the experiments to
Biodiesel Experiment those predicted by the simulations. Although the simulations
Soot [[g/kW.h]

under-predict the soot emissions for all the fuels, they correctly
0.02 predict the trend of soot with FAME unsaturation level. In Fig. 7,
the RME exhibits higher soot emission due to its higher unsatura-
tion level, while JME results in lower soot production because of its
0.01 lower unsaturation level. The results confirm that the simulations
are able to reproduce differences in soot emissions due to differ-
ences in biofuel saturation level.
0.00 Fig. 8(a) shows the effect of biodiesel composition on the mass
25% load 100% load concentration of acetylene, soot precursors, and soot predicted by
the multi-step phenomenological model. To identify the difference
Fig. 6. Comparison of measured and simulated soot mass for the diesel and of the final production for the three species, magnified profiles
biodiesel fuel under test conditions.
from 60 to 100 °CA ATDC were presented in Fig. 8(b). It can be seen
that the histories of C2H2, soot precursor, and soot showed the
Fig. 5. On the other hand, although not considered in the present same increasing trends with the increase of fuel unsaturation.
models, the radiation from soot can be an important source for From JME, and CME to RME, the peak mass concentrations of acet-
heat transfer in diesel engine flames, some researchers have pro- ylene, soot precursors and soot increased with the increasing level
posed that the radiative heat losses for biodiesel are lower in com- of unsaturation. Similar trends of net soot production were also
parison with diesel, which produce a higher flame temperature observed in the experiments for three biodiesels. As a result, an
[37]. Others have suggested that biodiesel has a slightly higher adi- increase in net soot production for RME was observed due to its
abatic flame temperature than diesel [38]. Overall, a higher flame higher fraction of unsaturated methyl esters compared to those
temperature is attained for biodiesel, which improved soot oxida- of JME and CME, as shown in Fig. 7. Both the experiments and sim-
tion at high engine load. ulations indicate that the soot tendency of biodiesel correlates
As shown in Fig. 6, the net soot emissions predicted by the strongly to the fuel composition and increases in the following
multi-step phenomenological soot model were in good agreement order: JME, CME and RME.
with the experimental data for each fuel. Therefore, it is reasonable
to believe that the present MD/MD9D/heptane surrogate mecha- 3.3.2. Spatial distributions of intermediate species for soot
nism, coupled with the multi-step phenomenological soot model Fig. 9 shows the spatial distributions of mass concentration of
captures the soot emission trends with acceptable accuracy. acetylene, soot, OH radicals, average temperature, and equivalence
ratio at the timing of 10 °CA ATDC (during main injection). It is
3.2. Combustion characteristics of different biodiesels clear that the acetylene and soot are detected in relatively high
temperature (1600–1800 K) and fuel rich (above equivalence ratio
Fig. 3c shows the calculated pressure and HRR traces for the dif- (phi) = 2) regions for all fuels. In Fig. 9(a) and (b), the three differ-
ferent biodiesel fuels at 100% load conditions. It is obvious that the ent biodiesels all produce lower amount of acetylene and soot
pressure and HRR for the three biodiesel fuels does not show much compared to the diesel fuel. Moreover, it can be seen that RME
difference at the same test conditions, and the calculated results shows higher acetylene and soot mass concentrations than the
are in excellent agreement with the measured engine data, as other two biodiesel fuels, which illustrates that more soot is
shown in Fig. 3. To further investigate the effect of the different formed for the fuel with more carbon double bonds. These results
compositions on soot emissions, predictions were made for soot are consistent with the experimental results.
mass evolution and the mass concentrations of soot-relevant Fig. 10(d) shows the distribution for the OH radical during the
species. main combustion phase, which represents the reactant involved
in soot oxidation reactions. The pattern of OH distribution is very
similar to that of the temperature, indicating that OH is generated
3.3. Soot emissions
in high temperature regions. The net soot formation is a competi-
3.3.1. Predicted evolution of soot-relevant species
As discussed above, biodiesel fuel is mainly composed of five 0.008
methyl esters. Here the three biodiesel fuels derived from three
JME Simulation JME Experiment
different feedstocks contain different proportions of methyl esters.
CME Simulation CME Experiment
Fig. 2 illustrates the mole fractions of the five major compositions 0.006 RME Simulation RME Experiment
Soot [[g/kW.h]

of the various biodiesels. JME has the lowest amount of unsatu-


rated methyl esters (i.e., fewer carbon double bonds), while RME
has the highest amount of unsaturated methyl esters (i.e., more 0.004
carbon double bonds) among the three biodiesel fuels. The carbon
double bond has an important influence on sooting tendency. Sar-
athy et al. [15] and Gaïl et al. [16] found that the acetylene (C2H2) 0.002
yield during biodiesel pyrolysis is proportional to the number of
C@C double bonds in the methyl esters. The acetylene species are
intermediates to soot precursor formation, a higher level of C2H2 0.000
JME CME RME
leads to more soot production. This suggests that an unsaturated
methyl ester would have a greater tendency of soot formation Fig. 7. Comparison of measured and simulated soot mass for the three different
compared to a saturated methyl ester, which is consistent with biodiesel fuel under test conditions.
246 Z. Wang et al. / Fuel 175 (2016) 240–248

3
0.0009
100% load JME JME
CME 100% load CME
Acetylene (g/kg fuel) 2 RME

Acetylene (g/kg fuel)


0.0006 RME

1 0.0003

0.0000
0

-0.0003
-30 -20 -10 0 10 20 30 40 50 60 60 70 80 90 100
Crank Angle ((°)CA ATDC) Crank Angle ((°)CA ATDC)

2.0 0.06

JME 0.05 JME


1.5 100% load CME 100% load CME
RME RME

P r e c u r s o r ( g /k g f u e l)
Precursor (g/kg fuel)

0.04

1.0 0.03

0.02
0.5

0.01
0.0
0.00
-30 -20 -10 0 10 20 30 40 50 60 60 70 80 90 100
Crank Angle ((°)CA ATDC) Crank Angle ((°)CA ATDC)

30 0.06

JME 0.05 JME


100% load CME 100% load CME
Soot (g/kg fuel)

20
Soot (g/kg fuel)

RME 0.04 RME

0.03

10
0.02

0.01
0
0.00
-30 -20 -10 0 10 20 30 40 50 60 60 70 80 90 100
Crank Angle ((°)CA ATDC) Crank Angle ((°)CA ATDC)
(a) Global figures (b) local magnified figures
Fig. 8. Mass concentration of acetylene, soot precursors and soot for different biodiesel.

tion between the rate of soot formation and the rate of precursor lower soot than diesel fuel. Furthermore, the mass concentration of
and soot oxidation. Bartok and Sarofim [39] stated that soot oxida- acetylene increases with increasing number of carbon double
tion is most likely to be dominated by OH under stoichiometric and bonds in biodiesel fuels. At 25 °CA ATDC, the post injected fuel
fuel-rich conditions, while soot is oxidized by both OH and O2 burns after the main combustion phase, which oxidizes the soot
under lean conditions. The simulation results in Fig. 9(d) show that formed during the main combustion. Also, the post injection
the OH concentration for diesel fuel is lower than that of the bio- improves fuel air mixing, which further accelerates soot oxidation.
diesel fuels, resulting in a decreasing soot oxidation by OH for It is seen in Fig. 10(c) that a small amount of soot number density is
the diesel fuel. Moreover, more oxygen is available for reacting left for all fuels.
with carbon in the biodiesel. As a result, biodiesel fuels are effec-
tive for soot reduction mainly due to the suppressed soot precursor
formation and improved soot oxidation. 4. Concluding remarks
Fig. 10 presents the spatial distributions of mass concentration
of acetylene, soot, OH radical, temperature, and equivalence ratio Experiments were conducted in a multi-cylinder diesel engine
at the timing of 25 °CA ATDC (during post-injection). As can been to study the soot emission characteristics of three different biodie-
seen in Fig. 8, two peaks of in-cylinder soot concentration were sels. To better understand the test results, a multi-component bio-
observed near the timing of 10 °CA and 25 °CA ATDC. Similar diesel mechanism, coupled with a multi-step soot model was
results are also been found in Fig. 10. All three biodiesels produce validated and used to predict soot formation of different biodiesel
Z. Wang et al. / Fuel 175 (2016) 240–248 247

JME CME RME Diesel

(a) Acetylene

0e-03 3.0e-03
0e-03 3.0e-03 6.0e-03
6.0e-03

(b) Soot mass

10 3 5 8 10
(ppm)

(c) Soot number


desity

0e-10 5.0e-10
0e-10 5.0e-10 1.0e-09
1.0e-09

(d) OH

0e-03 1.0e-03 2.0e-03

(e) Temp

800K 1700K 2600K

(f)Equivalence
ratio

0.0 1.5 3.0

Fig. 10. Spatial distributions of mass concentrations of acetylene, soot, OH radicals, and soot number density, in-cylinder temperature, equivalence ratio at 25 °CA ATDC.

fuels. The proposed model well reproduce the soot trends of differ- Acknowledgments
ent fuels obtained in the engine experiments and reveal the effect
of fuel saturation on soot formation. This work was sponsored by the Ministry of Science and Tech-
nology of China through the China–Singapore Project
(1) Similar to the experimental results, the simulation results 2012DFG61960 and the Project of the National Key Basic Research
show that biodiesels produce lower soot compared to diesel Plan (Chinese ‘‘973’’ Plan) under Grant No. 2013CB228404. The
fuel because of the oxygen atoms in the fuel. The oxygenated authors also appreciate the useful discussion with Dr. Xin He at
characteristics of biodiesel both reduce soot formation and Tsinghua University on biodiesel combustion.
enhance soot oxidation.
(2) During the combustion, higher saturated biodiesel fuel pro-
duces less acetylene and soot precursors than un-saturated References
biodiesel. Furthermore, acetylene production is proportional
[1] Wang Z, Qi Y, He X, et al. Analysis of pre-ignition to super-knock: hotspot-
to the number of carbon–carbon double bonds in the FAME induced deflagration to detonation. Fuel 2015;144:222–7.
structure, which is observed both in the experiments and [2] Wang Z, Liu H, Song T, He X. Relationship between super-knock and pre-
simulations. Similar trends can also be seen by studying ignition. Int J Engine Res 2015;16(2):166–80.
[3] Buyukkaya E. Effects of biodiesel on a DI diesel engine performance, emission
the in-cylinder spatial distributions of mass concentrations and combustion characteristics. Fuel 2010;89(10):3099–105.
of acetylene and precursor species. [4] Di Y, Cheung CS, Huang Z. Experimental investigation on regulated and
(3) The net soot production is the result of the combined effect unregulated emissions of a diesel engine fueled with ultra-low sulfur diesel
fuel blended with biodiesel from waste cooking oil. Sci Total Environ 2009;407
of acetylene and precursor species formation. Biodiesel fuels (2):835–46.
with a lower fraction of unsaturated FAME (less C@C) result [5] Li L, Wang J, Wang Z, Xiao J. Combustion and emission characteristics of diesel
in lower soot emissions. RME exhibits the highest level of engine fueled with diesel/biodiesel/pentanol fuel blends. Fuel
2015;156:211–8.
soot production, while the JME has the lowest level of soot
[6] Herbinet O, Pitz WJ, Westbrook CK. Detailed chemical kinetic oxidation
production. mechanism for a biodiesel surrogate. Combust Flame 2008;154(3):507–28.
248 Z. Wang et al. / Fuel 175 (2016) 240–248

[7] Agarwal AK, Gupta T, Kothari A. Particulate emissions from biodiesel vs diesel [25] Nagel J, Strickland-Constable. Oxidation of carbon between 1000 and 2000C.
fuelled compression ignition engine. Renew Sustain Energy Rev 2011;15 In: Proceedings of the fifth conference on carbon; 1962. p154.
(6):3278–300. [26] Kazakov A, Foster DE. Modeling of soot formation during DI diesel combustion
[8] Zhang J, He K, Shi X, Zhao Y. Comparison of particle emissions from an using a multi-step phenomenological model. SAE paper 982463; 1998.
engine operating on biodiesel and petroleum diesel. Fuel 2011;90(6): [27] Fusco A, Knox-Kelecy AL, Foster DE. Application of a phenomenological soot
2089–97. model to diesel engine combustion. In: Conference on modeling and
[9] Giakoumis EG, Rakopoulos CD, Dimaratos AM, Rakopoulos DC. Exhaust diagnostics for advanced engine systems (COMODIA), vol. 94; 1994. p. 571–6.
emissions of diesel engines operating under transient conditions with [28] Tao F, Golovitchev VI, Chomiak J. A phenomenological model for the prediction
biodiesel fuel blends. Prog Energy Combust 2012;38(5):691–715. of soot formation in diesel spray combustion. Combust Flame 2004;136
[10] Di Y, Cheung CS, Huang Z. Comparison of the effect of biodiesel–diesel and (3):270–82.
ethanol–diesel on the gaseous emission of a direct-injection diesel engine. [29] Tao F, Reitz RD, Foster DE, Liu Y. Nine-step phenomenological diesel soot
Atmos Environ 2009;43(17):2721–30. model validated over a wide range of engine conditions. Int J Therm Sci
[11] Cheng AS, Dibble RW, Buchholz BA. The effect of oxygenates on diesel engine 2009;48(6):1223–34.
particulate matter. SAE paper 2002-01-1705; 2002. [30] Tao F, Liu Y, RempelEwert BH, Foster DE, Reitz RD, et al. Modeling the effects of
[12] Chen G, Yu W, Fu J, Mo J, Huang Z, Yang J, et al. Experimental and modeling EGR and injection pressure on soot formation in a high-speed direct-injection
study of the effects of adding oxygenated fuels to premixed n-heptane flames. (HSDI) diesel engine using a multi-step phenomenological soot model. SAE
Combust Flame 2012;159(7):2324–35. technical paper 2005-01-0121; 2005.
[13] Yu W, Chen G, Huang Z, Chen Z, Gong J, Yang J, et al. Experimental and kinetic [31] Liu Y, Tao F, Foster DE, Reitz RD. Application of a multiple-step
modeling study of methyl butanoate and methyl butanoate/methanol flames phenomenological soot model to HSDI diesel multiple injection modeling.
at different equivalence ratios and C/O ratios. Combust Flame 2012;159 SAE technical paper 2005-01-0924; 2005.
(1):44–54. [32] McEnally CS, Pfefferle LD. Sooting tendencies of oxygenated hydrocarbons in
[14] Li L, Wang J, Wang Z, Liu H. Combustion and emissions of compression ignition laboratory-scale flames. Environ Sci Technol 2011;45:2498–503.
in a direct injection diesel engine fueled with pentanol. Energy [33] Wu F, Wang J, Chen W, Shuai S. A study on emission performance of a diesel
2015;80:575–81. engine fueled with five typical methyl ester biodiesels. Atmos Environ 2009;43
[15] Sarathy SM, Gaïl S, Syed SA, Thomson MJ, Dagaut P. A comparison of saturated (7):1481–5.
and unsaturated C4 fatty acid methyl esters in an opposed flow diffusion flame [34] Johansson M, Yang J, Ochoterena R, Gjirja S, Denbratt I. NOx and soot emissions
and a jet stirred reactor. Proc Combust Inst 2007;31(1):1015–22. trends for RME, SME and PME fuels using engine and spray experiments in
[16] Gaïl S, Thomson MJ, Sarathy SM, Syed SA, Dagaut P, Diévart P, et al. A wide- combination with simulations. Fuel 2013;106:293–302.
ranging kinetic modeling study of methyl butanoate combustion. Proc [35] Brakora JL, Ra Y, Reitz RD. Combustion model for biodiesel-fueled engine
Combust Inst 2007;31(1):305–11. simulations using realistic chemistry and physical properties. SAE technical
[17] Garner S, Sivaramakrishnan R, Brezinsky K. The high-pressure pyrolysis of paper 2011-01-0831; 2011.
saturated and unsaturated C7 hydrocarbons. Proc Combust Inst 2009;32 [36] Westbrook CK, Pitz WJ, Curran HJ. Chemical kinetic modeling study of the
(1):461–7. effects of oxygenated hydrocarbons on soot emissions from diesel engines. J
[18] Feng Q, Jalali A, Fincham AM, Wang YL, Tsotsis TT, Egolfopoulos FN. Soot Phys Chem 2006;110:6912–22.
formation in flames of model biodiesel fuels. Combust Flame 2012;159 [37] Musculus M. Measurements of the influence of soot radiation on in-cylinder
(5):1876–93. temperatures and exhaust NOx in a heavy-duty DI diesel engine. SAE technical
[19] Wang Z, Wang Y, Reitz RD. Pressure oscillation and chemical kinetics coupling paper 2005-01-0925; 2005.
during knock processes in gasoline engine combustion. Energy Fuels [38] Ban-Weiss GA, Chen JY, Buchholz BA, et al. Numerical investigation into the
2012;26:7107–19. anomalous slight NOx increase when burning biodiesel; a new (old) theory.
[20] Fisher EM, Pitz WJ, Curran HJ, Westbrook CK. Detailed chemical kinetic Fuel Process Technol 2007;88(7):659–67.
mechanisms for combustion of oxygenated fuels. Proc Combust Inst 2000;28 [39] Bartok W, Sarofim AF. Fossil fuel combustion: a source book. University of
(2):1579–86. California; 1991.
[21] Herbinet O, Pitz WJ, Westbrook CK. Detailed chemical kinetic mechanism for [40] Yakhot V, Orszag SA. Renormalization-group analysis of turbulence. Phys Rev
the oxidation of biodiesel fuels blend surrogate. Combust Flame 2010;157 Lett 1986;57(14):1722–4.
(5):893–908. [41] Patterson MA, Reitz RD. Modeling the Effects of fuel spray characteristics on
[22] Dagaut P, Gaı LS, Sahasrabudhe M. Rapeseed oil methyl ester oxidation over diesel engine combustion and emission. SAE technical paper 980131; 1998.
extended ranges of pressure, temperature, and equivalence ratio: [42] Beale JC, Reitz RD. Modeling spray atomization with the Kelvin-Helmholtz/
experimental and modeling kinetic study. Proc Combust Inst 2007;31 Rayleigh-Taylor hybrid model. Atomization and Sprays 1999;9(6):623–50.
(2):2955–61. [43] Ra Y, Reitz RD. A vaporization model for discrete multi-component fuel sprays.
[23] Brakora JL, Reitz RD. A comprehensive combustion model for biodiesel-fueled Int J Multiphas Flow 2009;35(2):101–17.
engine simulations. SAE technical paper 2013-01-1099; 2013. [44] Kong SC, Reitz RD. Use of detailed chemical kinetics to study HCCI engine
[24] Hiroyasu H, Kadota T. models for combustion and formation of nitric oxide and combustion with consideration of turbulent mixing effects. J Eng Gas Turb
soot in DI diesel engines.SAE technical paper 760129; 1976. Power 2002;124(3):702–7.

S-ar putea să vă placă și