You are on page 1of 17

See discussions, stats, and author profiles for this publication at: https://www.researchgate.

net/publication/278670426

Effects of Fuel Physical Properties on Diesel Engine Combustion Using Diesel


and BioDiesel Fuels

Article · April 2008


DOI: 10.4271/2008-01-1379

CITATIONS READS
28 2,736

4 authors:

Youngchul Ra Rolf Reitz


Mountain Top University University of Wisconsin–Madison
52 PUBLICATIONS   1,616 CITATIONS    617 PUBLICATIONS   20,251 CITATIONS   

SEE PROFILE SEE PROFILE

Joanna Mcfarlane Stuart Daw


Oak Ridge National Laboratory Oak Ridge National Laboratory
66 PUBLICATIONS   750 CITATIONS    230 PUBLICATIONS   3,375 CITATIONS   

SEE PROFILE SEE PROFILE

Some of the authors of this publication are also working on these related projects:

SANS for Geochemistry and Geophysics View project

Computational Pyrolysis Consortium View project

All content following this page was uploaded by Joanna Mcfarlane on 23 February 2016.

The user has requested enhancement of the downloaded file.


08 PFL-944

Effects of Fuel Physical Properties on Diesel Engine Combustion


using Diesel and Bio-diesel Fuels

Youngchul Ra and Rolf D. Reitz


University of Wisconsin, Madison

Joanna McFarlane and Stuart C. Daw


Oak Ridge National Laboratory

Copyright © 2008 SAE International

ABSTRACT and the chemical oxidation mechanisms between the


two fuels, biodiesel combustion is not as well understood
A computational study using multi-diemsional CFD as the combustion of regular diesel.
modeling was performed to investigate the effects of
physical properties on diesel engine combustion While the effects of the physical and chemical properties
characteristics with bio-diesel fuels. Properties of typical of the fuel on combustion can be investigated
bio-diesel fuels that were either calculated or measured experimentally, a computational study is an effective
are used in the study and the simulation results are approach for investigating parametric effects of physical
compared with those of conventional diesel fuels. The properties on mixture preparation, ignition and
sensitivity of the computational results to individual combustion, and subsequent pollutant formation. In
physical properties is also investigated, and the results recent evolution of computational methods of diesel
provide information about the desirable characteristics of engine operation, more effort has been made in
the blended fuels. The properties considered in the modeling the in-cylinder processes. As spray injection,
study include liquid density, vapor pressure, surface which is a critical physical process in mixture
tension, liquid viscosity, liquid thermal conductivity, liquid preparation, affects the subsequent combustion process,
specific heat, latent heat, vapor specific heat, vapor the accuracy of physical models is of significant
diffusion coefficient, vapor viscosity and vapor thermal importance. The performance of physical models also
conductivity. The results show significant effects of the strongly depends on the accurate estimation of physical
fuel physical properties on ignition delay and burning and chemical properties of the fuel.
rates at various engine operating conditions. It is seen
that there is no single physical property that dominates Following a review of the literature concerning the
differences of ignition delay between diesel and bio- physical properties of alternative fuels, Chakravarthy et
diesel fuels. However, among the 11 properties al. [8] compiled and presented a complete set of bio-
considered in the study, the simulation results were diesel properties required for detailed engine
found to be most sensitive to the liquid fuel density, simulations. Based on that research, in this paper, a
vapor pressure and surface tension through their effects study of the effects of physical property differences
on the mixture preparation processes. between biodiesel and diesel fuels on CI engine
operation is presented. The sensitivity of computations
INTRODUCTION of engine combustion to the individual properties is also
investigated. In order to decouple the effect of
Biodiesel fuel is seen as a promising alternative to differences in reaction chemistry between biodiesel and
conventional diesel for use in compression-ignition (CI) diesel, the simulations were performed by replacing the
engines. As a plant-based fuel, it can be produced physical properties of diesel with those of biodiesel,
domestically, is a renewable energy source, and while keeping the same reaction mechanism as that
essentially has a closed carbon cycle over its lifetime [1]. used for diesel engine simulations [9].
In terms of emissions, it has been reported that biodiesel
tends to reduce unburned hydrocarbons, carbon BIODIESEL PROPERTIES
monoxide, and particulate matter compared to its
petroleum counterpart, while the NOx level is likely to The physical properties of mixtures of hydrocarbons and
increase with increasing biodiesel fraction in blended biodiesel can be determined by calculating the physical
fuels [2,3]. Although it has been suggested through properties of each individual component, and then
experimental research [2-7] that the difference of imposing mixing rules to derive the properties of the
combustion between biodiesel and regular diesel can be mixtures. This method is feasible for biodiesel, which
attributed to differences in both the physical properties consists of a relatively simple formulation of fatty acid
esters. Properties of biodiesel and their methods of Table 2. Assumed composition of biodiesel.
estimation have been described by Chakravarthy et al.
[8]. Table 1 lists various physical properties that are component amount [mol %]
required for spray and combustion simulations and how Hexadecanoic acid, methyl ester 17.0
these are commonly derived. In the present study, the Octadecanoic acid, methyl ester 9.0
same strategy as that used by Chakravarthy et al. was 9-Octadecenoic acid, methyl ester 30.0
employed to obtain the necessary properties of biodiesel 9,12-Octadecadienoic acid, methyl
surrogate fuel. 11 physical properties (p1 to p11) were 44.0
ester
considered in the study of biodiesel engine operation
and its sensitivity to physical properties.

The assumed composition of biodiesel for the calculation THERMODYNAMIC PROPERTIES


of physical properties, which is based on Gas-
Chromatography analysis of Soy-biodiesel, is listed in The critical properties of biodiesel are needed for
Table 2. thermodynamic calculations. Although it is difficult to
obtain critical properties of prototypical biodiesel
experimentally because the compounds decompose at
Table 1. Methods of estimation of physical properties. temperatures below the critical point, they were
estimated using group additivity and empirical
No. Physical Method of analysis/references formulations taken from Yuan et al. [10] following Reid et
property al. [11]. Mixture properties were determined using Lee-
Kesler mixing rules. An equation of state model was
p1 Liquid density Orrick & Ebar Group contribution used as suggested by Araújo and Meireles [40]. In the
[10], Rackett equation [11], Araújo present study, the molecular weight and critical
and Meireles [40] temperature of the considered biodiesel-surrogate were
p2 Vapor Pitzer method[10,11], Araújo and set to 292 g/mole and 775 K, respectively.
pressure Meireles [40]
The density of the biodiesel-surrogate was obtained
p3 Surface Weighted average of individual from data predicted by the Rackett equation [11], which
tension surface tensions [10] matches the available experimental value, ρref = 0.8976
p4 Liquid Group contribution method [10], g/cm3, at Tref of 273.15K [24]. The densities are plotted
viscosity empirical correlations [12], in Fig. 1. Liquid density variation is considered in the
Logarithmic equation [13], present spray models, as well as in the mass and
weighted average of individual momentum balance equations in the gas and liquid
components [14] phase interactions. Note that, the density of the diesel-
surrogate is somewhat underestimated since diesel is
p5 Liquid thermal GC TCD experimental analysis represented by C14H30 in this study.
conductivity [15], numerical analysis [16],
empirical correlations [17]
The vapor pressure of the biodiesel-surrogate was
p6 Heat of Pitzer method[10,11], Araújo and computed using a Peng-Robinson equation of state (PR-
vaporization Meireles [40] EOS) based on lumped parameter critical properties,
which closely matches the literature formulation [25].
p7 Vapor heat GC TCD experimental analysis
Figure 2 shows the vapor pressure variation of the two
capacity [15], numerical analysis [16],
fuels as a function of liquid temperature. Compared to
empirical correlations [17]
that of the diesel-surrogate, the vapor pressure of the
p8 Vapor Chapman-Enskog kinetic theory biodiesel-surrogate is much lower, which is expected to
diffusivity (Skelland [18]) influence the spray evaporation process significantly.
p9 Vapor Correlation from Chung and
The heat of vaporization of the biodiesel-surrogate was
viscosity coworkers [19,20]
also calculated up to the critical point, and is plotted
p10 Vapor thermal Correlation from Chung and along with that of the diesel-surrogate in Fig. 3.
conductivity coworkers [19,20] Compared to the diesel-surrogate, the heat of
vaporization of the biodiesel-surrogate is lower at low
p11 Liquid Heat Interpolation of Lagrange
temperatures, while it becomes higher at high
capacity polynomials [21], group
temperatures. Since spray droplets are heated up
contribution [22]
quickly during the vaporization process in the
p12 Critical Group contribution [10], correlation combustion chamber, it is expected that the evaporation
properties based on 400 fuels [11] of the biodiesel-surrogate will be more affected by the
high temperature characteristics of the heat of
p13 Mixing rules Lee-Kesler equation [10], Lumped
vaporization.
for critical parameter continuous
properties of thermodynamics [23]
mixtures
1.0 0.03
Biodiesel
0.8 0.025 Diesel

surface tension [N/m]


density [kg/L]

0.02
0.6
0.015
0.4
0.01
0.2 Biodiesel
0.005
Diesel
0.0
0
200 300 400 500 600 700 800
300 400 500 600 700
temperature [K] temperature [K]
Figure 2. Liquid density profiles for biodiesel and diesel Figure 4. Surface tension profiles of biodiesel and diesel
surrogate fuels. The density of biodiesel-surrogate was surrogate fuels.
calculated as in Ref. [8].

1.E+07 4.5

liquid heat capacity [J/kg-K]


Biodiesel
4.0
vapor pressure [Pa]

1.E+05 Diesel
3.5
1.E+03
3.0

1.E+01 2.5
2.0
1.E-01 Biodiesel
C14H30 1.5
1.E-03 1.0
300 400 500 600 700 200 300 400 500 600 700
temperature [K] temperature [K]
Figure 2. Vapor pressure profiles for biodiesel and diesel Figure 5. Profiles of liquid heat capacity for biodiesel and
surrogate fuels. diesel surrogate fuels.

500 0.01
heat of vaporization [kJ/kg]

Biodiesel
liquid viscosity [N-s/m^2]

400 Diesel Biodiesel


0.008
Diesel
300
0.006

200
0.004
100
0.002
0
200 300 400 500 600 700 800 0

temperature [K] 300 400 500 600 700


temperature [K]
Figure 3. Profiles of heat of vaporization for biodiesel Figure 6. Profiles of liquid viscosity for biodiesel and
and diesel surrogate fuels. diesel surrogate fuels.
0.16 1.8E-04
Biodiesel
liquid thermal conductivity
0.14 1.6E-04

vapor viscosity [Pa-s]


Diesel 1.4E-04
0.12
1.2E-04
0.10
[J/m-s-K]

1.0E-04
0.08
8.0E-05
0.06
6.0E-05
0.04 4.0E-05 Biodiesel
0.02 2.0E-05 Diesel
0.00 0.0E+00
300 400 500 600 700 300 500 700 900 1100
temperature [K] temperature [K]
Figure 7. Profiles of liquid thermal conductivity for
Figure 10. Profiles of vapor viscosity for biodiesel and
biodiesel and diesel surrogate fuels.
diesel surrogate fuels.

0.1
0.005

vapor thermal conductivity


Biodiesel
vapor heat capacity [J/kg-K]

0.08 Diesel
0.004

0.06

[W/m-K]
0.003

0.04
0.002

Biodiesel 0.02
0.001
Diesel
0
0
300 400 500 600 700 800 900 1000
0 500 1000 1500 2000 2500 3000
temperature [K]
temperature [K]

Figure 8. Profiles of vapor heat capacity for biodiesel Figure 11. Profiles of vapor thermal conductivity for
and diesel surrogate fuels. biodiesel and diesel surrogate fuels.

Surface tension is one of the key properties considered


0.1 in spray breakup and collision/coalescence models.
Biodiesel Empirical correlations were used to obtain surface
0.08 Diesel tension of the biodiesel-surrogate. The surface tension
diffusivity [cm^2/s]

is a linear function of temperature, with little change with


the weighting factors of the different methyl esters. The
0.06 surface tension of the biodiesel-surrogate is about the
same as that for the diesel-surrogate at room
0.04 temperature, but it decreases slower with temperature.
In Fig. 4, the biodiesel-surrogate surface tension is
compared with that of the diesel-surrogate as a function
0.02 of temperature.

0 The heat capacity for methyl esters can be correlated as


300 400 500 600 700 800 900 1000 a function of the number of carbon atoms in the parent
carboxylic acid moiety [26]. Compared to diesel,
temperature [K] biodiesel’s heat capacity per unit mole is nearly 50%
higher at temperatures near boiling point. However,
Figure 9. Profiles of diffusivity in air for biodiesel and
when it is compared per unit mass, as used in the
diesel surrogate fuels at ambient pressure.
energy equation, the heat capacity of the biodiesel-
surrogate is lower than that of the diesel surrogate. This
implies that the biodiesel-surrogate droplets may be estimate the sensitivity to each property, only one
heated faster than the diesel-surrogate droplets. The property was varied from the value of the base fuel to
liquid heat capacities of the two fuels are compared in that of the other fuel while all other properties were fixed
Fig. 5. at those of the base fuel. Two ways of sensitivity
checking were considered; one considered the diesel-
Liquid viscosity is considered in modeling drop internal surrogate as the reference and the properties were
flow, drop breakup and wall film motion. The liquid changed individually from the diesel-surrogate (base) to
viscosity of the biodiesel-surrogate is higher than that of the biodiesel-surrogate and the other is from the
the diesel-surrogate, especially at low temperatures biodiesel-surrogate (base fuel) to the diesel-surrogate.
where most of the breakup processes take place in the
cylinder. Therefore, it is expected that the breakup The energy conservation equation between the liquid
process is most affected by the viscosity difference. and vapor phases in the evaporation process is given as
Comparison of the viscosities of the two fuels is shown
in Fig. 6. eliq = evap + RT / W − Pv / ρ l − hvap
Liquid thermal conductivity is used in the present drop
interior model to calculate the heat transfer between the where eliq is the liquid internal energy, evap is the vapor
drop interior and the surface where a transient internal energy, R is the gas constant, T is the
temperature distribution is assumed. The thermal temperature, W is the molecular weight, Pv is the vapor
conductivity of the biodiesel-surrogate is slightly lower pressure at temperature T, ρl is the liquid density, and
than that of the diesel-surrogate, as shown in Fig. 7. hvap is the heat of vaporization.

The vapor heat capacity of the fuel is important in However, it should be noted that the vapor heat capacity
calculating the internal energy and temperature (involved through the vapor internal energy), heat of
distribution of gas mixtures surrounding the spray drops, vaporization, vapor pressure and liquid heat capacity
which substantially affects the transient heat transfer (involved through the liquid internal energy) are also
from the surrounding gas mixture to the drop surface. coupled through the Clausius-Clapeyron equation, i.e.,
This is especially true where the fuel drops vaporize
rapidly so that the fuel/air mixtures become rich. The ⎛ dPvap ⎞ hvap
vapor heat capacity of biodiesel was surrogated with that ⎜⎜ ⎟⎟ =
for Methyl Oleate, which is available in the literature [27]. ⎝ dT ⎠ sat T (1 / ρ g − 1 / ρ l )
Its vapor heat capacity is slightly lower than that of the
diesel-surrogate over the temperature range of interest,
When any one of the properties considered in the
as shown in Fig. 8.
Clausius-Clapeyron equation is changed, the equality of
the energy equation is inevitably violated. Therefore,
TRANSPORT PROPERTIES adjustment must be made to one of the other properties
to maintain equality of the equation, as listed in Table 3.
The transport properties of the vapor phase, i.e.,
diffusivity, viscosity, and thermal conductivity, were Table 3. Properties adjusted with sensitivity variable.
estimated for the soy biodiesel mixture at a number of
elevated temperatures and pressures, including above Sensitivity property Adjusted property
the critical point [8]. Liquid heat capacity Heat of vaporization
Vapor heat capacity Heat of vaporization
The diffusivity for biodiesel vapor is much lower than that Heat of vaporization Liquid heat capacity
for diesel. Note that, in the present study, the pressure Vapor pressure Liquid heat capacity
dependency of vapor diffusivity is taken into account
through the density variation in the mass diffusivity. The
viscosity of biodiesel vapor is slightly higher than that of
diesel while its thermal conductivity is lower than that of PHYSICAL SUB-MODELS
diesel. Viscosity and thermal conductivity show little
dependency on pressure, thus only the temperature For simulating the spray processes and the mixing and
dependency was considered in the present study. combustion of fuel/air mixtures in the cylinder, various
Figures 9 to 11 show comparisons of the diffusivity, physical sub-models were employed in the KIVA-ERC-
viscosity and thermal conductivity of the two surrogate CHEMKIN code which is based on KIVA3V Release 2
fuels, respectively. [28] coupled with the CHEMKIN II library [29]. The sub
models include ones related drop breakup, collision and
coalescence, drop deformation, drop evaporation, wall
SENSITIVITY TO PHYSICAL PROPERTIES impingement and vaporization, etc.

As mentioned above, 11 physical properties were A hybrid primary break-up model [30] that is
considered to study the sensitivity of numerical computationally effective as well as comprehensive
simulations of engine operation to the individual physical enough to account for the effects of aerodynamics, liquid
properties of the biodiesel-surrogate. In order to properties and nozzle flows was employed. In this
model, the injected fuel “blobs” are tracked by a COMBUSTION MODELING
Lagrangian method while the break-up of each blob is
calculated from considerations of jet stability from Detailed chemistry was employed by coupling the
Kelvin-Helmholtz (KH) instability theory. For the KIVA3V (Release 2) with the Chemkin II library [29]. A
secondary and further break-up processes, a Kelvin skeletal reaction mechanism for n-heptane fuel with 34
Helmholtz (KH) – Rayleigh Taylor (RT) hybrid model [30] species and 77 reactions [35], which was modified for
was used. hydrogen oxidation reactions and some key reactions of
the low temperature path of n-heptane to intermediate
A droplet collision model based on the stochastic particle hydrocarbon species from the mechanism of Patel et al.
method [28] was used, in which the collision frequency is [36], was used to calculate the detailed chemical kinetics
used to calculate the probability that a drop in one parcel of combustion. Various n-heptane mechanisms have
will undergo a collision with a drop in another parcel, been widely employed for diesel combustion calculations
assuming all drops in each parcel behave in the same by many researchers because of its similar ignition
manner. The probability of coalescence is determined characteristics to those of diesel fuel. In the present
considering the Weber number that includes the effects study, the n-heptane mechanism was also used for
of density and surface tension of liquid droplets. reaction calculations in each cell, but the physical
properties of the fuel were represented with those of the
Droplet deformation in terms of its distortion from surrogate fuels in the calculations of spray processes,
sphericity is modeled using a forced, damped harmonic and in gaseous species mixing. For the calculation of
oscillator model, where the surface tension and viscosity NOx formation, a 4 species (N, NO, N2O and NO2) and
of the droplet are the major properties used in the 9 reaction NOx mechanism was used that has been
restoring force and damping terms, respectively [31]. reduced from the GRI NOx mechanism [37] and added
Distortion of droplets affects the momentum change to the n-heptane reaction mechanism.
between droplets and the ambient gas, and
subsequently drop velocities (or relative velocity The same reaction chemistry mechanism was used for
between the drop and the gas) that are the governing the combustion calculations of both biodiesel and diesel
parameters in the breakup and evaporation processes. surrogate fuels. Therefore, the present work only
addresses differences in combustion that result from
The droplet vaporization model [32, 33] considers the differences in the fuel physical properties. Inclusion of
droplet temperature range from flash-boiling conditions the effect of chemistry has been considered by Brakora
to normal evaporation. The improved model accounts for et al. [9].
variable internal droplet temperatures, and considers an
unsteady internal heat flux with internal circulation, and a A phenomenological soot model [38] modified from the
new model for the determination of the droplet surface Hiroyasu soot model [39] was employed to predict soot
temperature. The model uses an effective heat transfer emissions. In the modified model, acetylene (C2H2) is
coefficient model for the heat flux from the surrounding used as an inception species for soot formation, which
gas to the droplet surface. Also, the variable density of not only enables the soot model to be coupled with the
the diesel-surrogate fuel as a function of temperature is detailed chemistry calculation, but also improves the
considered in the governing equations and the relevant soot emission predictions.
sub-models. The effective heat transfer coefficient
calculated in the model is also used to determine the COMPUTATIONAL CONDITIONS
amount of fuel to be treated as vapor when the drop
surface temperature reaches the critical temperature Simulations of combustion in heavy and light-duty diesel
while the drop interior is still in sub-critical condition. The engines with various injection timings were performed
detailed formulation is shown in Appendix A. for biodiesel and diesel surrogate fuels to study the
effects of differences of physical properties on
Effects associated with spray/wall interactions, including combustion. The physical properties of the biodiesel-
droplet splash, film spreading due to impingement surrogate mentioned in the previous section were used
forces, and motion due to film inertia were considered in as the fuel properties in the simulations. The properties
a wall film submodel [34], in addition to calculations of of diesel surrogate fuel were represented by those of
film transport on complex surfaces with heating and tetra-decane (C14H30).
vaporization of the film, and separation and re-
entrainment of films at sharp corners. In order to study the sensitivity of mixture preparation
and combustion to the individual physical properties of
In the 2-phase transport equations, droplets are treated the fuel, 11 physical properties (refer to the previous
as point sources and the wall film fuel flow is not section) were varied between those of the biodiesel and
resolved on the computational grid. Therefore, it is diesel surrogate fuels. In addition, simulations of
assumed that the vaporized fuel in a computational cell evaporation of single drops were performed under two
where droplets or wall film parcels exist mixes non-reacting flow conditions; stagnant and convective
completely with the gaseous mixture within the cell. flow. Assuming constant ambient conditions, the
Thus, stratification of gaseous species within a single effects of the transport properties of the fuel vapor were
cell is not resolved. not considered in the single drop cases. Spray
Table 4. Computational conditions.
1.2
Engine: 1.9 L light-duty engine
Bore x Stroke [mm] 82 x 90.4 1

normalized injection rate


Compression ratio 16.5
0.8
Engine speed [rev/min] 2000
7-hole,
injector 0.6
included angle=155 deg
Nozzle hole diameter [µm] 141 0.4
Injection timings [deg atdc] -20.0 ~ -6
Injection duration [degCA] 8.62 0.2
Injection amount [mg] 14.6
IVC temperature [K] 357, 373 0
IVC pressure [bar] 1.93, 2.93 0 0.2 0.4 0.6 0.8 1

Oxygen fraction @ IVC [%] 21, 15 normalized time


Computation crank angle
-132 ~ 112 Figure 13. Injection rate shape used for the small bore
[deg atdc]
light-duty engine simulations. Injection duration was 8.62
Engine: 2.4 L heavy-duty engine
degCA for an engine speed of 2000 rev/min. Rates are
Bore x Stroke [mm] 137.16 x 165.1
normalized by the maximum value during the injection.
Compression ratio 16.2
Engine speed [rev/min] 1737
6-hole, simulations, injection timings were varied from -20 ~ -6
injector
included angle=130 deg deg atdc that ranges from PCCI to conventional DI
Nozzle hole diameter [µm] 192 modes. In the injection sweep, the same initial
Injection timings [deg atdc] -20.0 ~ -6 conditions were used for all simulations. The sensitivity
Injection duration [degCA] 16.5 study was performed for the case with injection timing of
Injection amount [mg] 109 -6 deg atdc.
IVC temperature [K] 367
IVC pressure [bar] 2.21 A representative computational grid for the small bore
Oxygen fraction @ IVC [%] 21 light-duty engine is shown in Fig. 12. Sector grids (1/7th
Computation crank angle
-143 ~ 110 and1/6th sectors for the small and large bore engines,
[deg atdc] respectively) were employed to save computation time.
A total of 32656 cells at BTDC were considered. 20 cells
were used for the 360/7 degree sector in the azimuthal
direction and 1.078 x 0.87 mm resolution was used for
the cells in the bowl region in radial and vertical
directions, respectively. At the time of injection, the
squish region has a resolution of 1.078 x 1.216 mm in
radial and vertical directions, respectively. The resolution
of the grid turned out to be within the grid-independence
range from extensive grid-resolution dependence
investigation.

It was concluded that a mesh resolution of ~ 1mm is


satisfactory. A total of 32656 cells at BTDC were
Figure 12. Vertical cross-section view of the considered in the mesh shown in Fig. 12. 20 cells were
computational grid for the small bore light-duty engine. used for the 360/7 degree sector in the azimuthal
direction and 1.078 x 0.87 mm resolution was used for
combustion was also simulated in the sensitivity study the cells in the bowl region in the radial and vertical
with all of the 11 physical species considered. directions, respectively. At the time of injection, the
squish region has a resolution of 1.078 x 1.216 mm in
A small bore light-duty (1.9L) diesel engine with a the radial and vertical directions, respectively.
compression ratio of 16.5 [35] was used for the
simulations. In order to study the effects of boost The fuel injection profiles were based on
pressure and engine geometry, the same engine with a measurements of diesel spray injections [35], as shown
boost pressure of 2.93 bar and a large bore heavy-duty in Fig. 13, and the same profile was used for the
engine [33] were also simulated for the two fuels. different injection timings and different fuels.

The computational conditions are listed in Table. 4. For Due to the difference of physical properties (especially,
single drop simulations, the ambient temperature and density, vapor pressure and bulk modulus) of biodiesel
pressure were fixed at 1000 K and 40 bar, respectively. from those of diesel, the biodiesel flow upstream of
The initial drop size was 100 µm. For the engine nozzle exit may differ from that of diesel, and
subsequently, the injection rates of biodiesel sprays may
be altered. In the present study, however, the same
injection rate shape was used for the simulation of the drop size, biodiesel
biodiesel-surrogate spray injections in order to facilitate 1.6 1.6
drop size, diesel
the analysis of spray behavior differences between 1.4 surface fuel mass ftaction, biodiesel 1.4

fuel mass fraction @ surface


biodiesel and diesel surrogate fuels. The effects of

nomalized drop diameter


surface fuel mass ftaction, diesel
biodiesel physical properties on spray injection are left 1.2 1.2
for further investigation. 1 1

0.8 0.8
RESULTS AND DISCUSSION
0.6 0.6
SINGLE DROP SIMULATION 0.4 0.4

For initial assessment of the effect of fuel properties, the 0.2 0.2
evaporation of a single droplet was simulated under 0 0
stagnant and convective conditions with the ambient 0.00 0.02 0.04 0.06 0.08 0.10
pressure and temperature fixed at 1 bar, 800 K,
respectively. The initial drop size and temperature were time [sec]
set to 100 µm and 340 K, respectively. For the (a)
convective condition, the initial drop velocity was 5 m/s.

Figure 14 shows the variation of normalized drop 650


diameter, fuel mass fraction at the drop surface and drop
interior and surface temperatures for biodiesel and 600

drop temperature [K]


diesel surrogate fuel droplets. The biodiesel-surrogate 550
drop has a much longer lifetime than the diesel-
surrogate drop. Although the surface temperature of the 500
biodiesel-surrogate drop is higher than that of the diesel-
surrogate drop, the corresponding surface fuel mass 450
interior, biodiesel
fraction is much lower than that of the diesel-surrogate
400 surface, biodiesel
because the vapor pressure of the biodiesel-surrogate is
much lower than that of the diesel-surrogate at the same 350 interior, diesel
temperature (refer to Fig. 2). Note that, during the surface, diesel
vaporization of a stagnant drop, the drop temperature 300
increases rapidly due to the heat transfer from the 0.00 0.02 0.04 0.06 0.08 0.10
ambient gases. The drop size increases as the drop time [sec]
density decreases while the drop mass decreases. The
heating process lasts until the drop surface (b)
temperatures become equal to the boiling temperature,
624.9 and 525.6 K for biodiesel and diesel surrogate Figure 14. Predicted temporal profiles of drop conditions.
fuels, respectively. While the biodiesel-surrogate drop is (a) drop diameter normalized by the initial drop diameter
predicted to reach boiling evaporation, the diesel- (=100 µm) and fuel mass fraction at the drop surface, (b)
surrogate drop vaporizes in the normal evaporation drop interior and surface temperatures. Ambient gas
mode during the entire drop lifetime due to its much pressure and temperature are 1 bar and 800 K,
faster evaporation rate than the biodiesel-surrogate. It is respectively.
notable that, under the typical in-cylinder gas conditions
of CI engines, the ambient gas pressure is likely to be
super-critical. In this case, the spray droplets do not As can be seen in Figs. 15 (a) and (b), it turns out that
experience the boiling regime, but transcritical drop vaporization is not sensitive to variations of surface
vaporization is likely to occur. tension (p3), liquid viscosity (p4), and heat of
vaporization (p6). It is expected that the surface tension
The sensitivity of the drop vaporization to the individual does not affect the vaporization of a single drop because
physical properties was calculated for stagnant and no break-up process is assumed. Differences in the
convective conditions, and the normalized drop size liquid viscosity and heat of vaporization between the
profiles are plotted in Fig. 15 for some of the parameter biodiesel and diesel surrogate fuels are relatively small,
cases. Since the ambient conditions are assumed to be thus the sensitivity to these properties is minimal.
constant and stationary (no effect of the entrainment
velocity of the ambient gases was assumed, although In terms of the drop lifetime, the most sensitive property
the decrease of drop momentum by the drag force was was the vapor pressure (p2) in both cases. It is
taken into account for the convective cases) during the interesting that the vapor pressure variation results in
drop lifetime, the effects of gas phase transport almost the same drop lifetime as that of the diesel-
properties are not considered in the sensitivity surrogate under the convective evaporation condition
simulations for single drops. without drop break-up. Vapor specific heat (p7) and the
biodiesel, cainj= -20
1.2 p1 16 biodiesel, cainj= -15
p6
biodiesel, cainj= -10
biodiesel biodiesel, cainj= -6
1 14
nomalized drop diameter

diesel, cainj= -20


p2 p5 diesel, cainj= -15

pressure [MPa]
0.8 12
-20 diesel, cainj= -10
diesel, cainj= -6
p7 10
0.6
p11
0.4 8
diesel -6
0.2 6
p3, p4
0 4
0.00 0.02 0.04 0.06 0.08 0.10 -20 -10 0 10 20 30
crank angle [deg atdc]
time [sec]

(a) Figure 16. Effects of physical property differences on


combustion in a large bore heavy-duty engine. Injection
1.2 timings were varied from -20 to -6 deg atdc. Solid lines
p1
p6
are for the biodiesel-surrogate and dotted lines are for
1 the diesel-surrogate.
nomalized drop diameter

biodiesel

0.8
p11
p2 surface. Therefore, variation of vapor specific heat alters
0.6 p7 the heat transfer from the surrounding gas mixtures.
p5 Liquid density (p1) change mostly affects the swelling of
0.4 the drop, which implies that the sensitivity to this
diesel property may be significantly coupled to the breakup,
0.2 p3, p4 collision/coalescence, and drop deformation processes
in spray injection cases.
0
0.00 0.01 0.02 0.03 0.04 0.05 0.06 SPRAY SIMULATIONS
time [sec]
Simulations of biodiesel and diesel surrogate sprays
(b) injected into an engine cylinder were performed for three
different operating conditions. Two different piston bowl
Figure 15. Sensitivity of single drop vaporization to geometries and two boost pressures were considered so
individual physical properties listed in Table 1. (a) as to compare the effects of spray impingement and
stagnant drop vaporization, (b) convective drop ambient pressures. Due to the sufficiently large bore of
vaporization with initial drop velocity of 5 m/s. Baseline the heavy-duty engine, spray impingement on the
properties were set to those of the biodiesel-surrogate cylinder walls is minimized. By comparing the two cases
and changed individually to the corresponding the with different boost pressures for the same bowl
diesel-surrogate property. geometry of the 1.9 L light-duty engine, the effects of
physical property differences due to boost pressure
differences can be examined.
liquid density (p1) were predicted to be the second most
sensitive properties. In Fig. 16, pressure profiles of combustion of biodiesel
and diesel surrogate fuels in the large bore engine are
It is notable that use of the liquid specific heat of the compared for four different injection timings. Earlier
diesel-surrogate for that of the biodiesel-surrogate (see injection timing allows more time for vaporization and
Fig. 5) slows down evaporation such that the drop life mixing before ignition. This results in smaller differences
time becomes longer than that for the biodiesel- of ignition timing than for cases with later injection
surrogate case. Vapor pressure, which is the most timings. Note that, since the same combustion model
different property between the biodiesel and diesel was used for both fuels, the differences of ignition
surrogate fuels (refer to Fig. 2), governs the fuel mass timings are only due to the effects of the physical
fraction at the drop surface, and thus the evaporation property differences of the fuels on the mixture
until the boiling temperature (or critical temperature preparation process. Fuel sprays that are injected into
under supercritical ambient pressure conditions) is an already ignited gas mixture meet higher temperature
reached. The fuel vapor specific heat (p7) becomes an gases, and thus their vaporization is further enhanced.
important factor when the fuel drop vaporizes rapidly Therefore, it is expected that the effects of the physical
near the boiling regime because the temperature property differences become less important in terms of
distribution near the drop surface is significantly affected their effect on the peak cylinder gas pressure as
by the specific heat of the gas mixture that includes a injection timings are advanced.
large fraction of fuel vapor supplied from the drop
biodiesel diesel

biodiesel diesel
(a)
(a)

biodiesel diesel

biodiesel diesel
(b) (b)

biodiesel diesel

Figure 17. Distributions of temperature and drop (c)


locations at 20 deg atdc in the plane of the spray for
injection timings of (a) –20 deg atdc and (b) –6 deg atdc.
Simulated engine is the large bore heavy-duty engine.
biodiesel diesel
biodiesel, cainj= -20
13 biodiesel, cainj= -15
12 biodiesel, cainj= -10
(d)
biodiesel, cainj= -6
11
diesel, cainj= -20
pressure [MPa]

10 diesel, cainj= -15


9
-20 diesel, cainj= -10
diesel, cainj= -6
8
7
Figure 19. Distributions of temperature and drop
6 locations in the plane of the spray for two different
5 -6 injection timings. (a) at crank angle -4 deg atdc with
4
injection at –20 deg atdc, (b) at crank angle +20 deg
-20 -10 0 10 20 30
atdc with injection at –20 deg atdc, (c) at crank angle +4
deg atdc with injection at –6 deg atdc, and (d) at crank
crank angle [deg atdc]
angle +20 deg atdc with injection at –6 deg atdc. The
Figure 18. Effects of physical property difference on simulated engine is the small bore light-duty engine.
combustion in a small bore light-duty engine. Injection
timings were varied from -20 to -6 deg atdc. Solid lines to move to the outer region of the piston bowl and the
are for the biodiesel-surrogate and dotted lines are for squish region in the biodiesel-surrogate case.
the diesel-surrogate.
When the small bore light duty diesel engine with a
similar compression ratio (16.5) was simulated, the
The coupled effects of injection timing and physical effects of injection timing variation were found to be
property differences can be seen in the distributions of complicated by spray-wall interactions. Thus, the
temperature and drop locations shown in Fig. 17. For the operating conditions for the engine were chosen to
injection timing of –20 deg atdc, the spray drops highlight the effects of spray breakup and spray-wall
vaporize well and fuel vapor mixes well with air in both interactions. Compared to the large bore case, the
cases. It is seen that the air in the entire cylinder region sprays are expected to have poorer breakup
is utilized in the combustion even though the distribution characteristics (the ambient gas density at the time of
of high temperature regions is a bit different due to the injection is lower) and more wall impingement (maximum
different vaporization history of the sprays in the two fuel available distance for penetration before impingement is
cases. much shorter). Figure 18 shows the pressure profiles of
combustion of biodiesel and diesel surrogate fuels in the
However, for the injection timing of –6 deg atdc, more light-duty engine with a IVC gas pressure of 1.93 bar.
droplets are seen, and the high temperature zones tend
For the injection timings considered in this study, the
trend of ignition timing differences with injection timing
variation are seen to be similar to those of the large bore
biodiesel, cainj= -20
18 biodiesel, cainj= -15
biodiesel, cainj= -10
16 9.5
biodiesel, cainj= -6
diesel
14 -20 diesel, cainj= -20
9.0 p1
diesel, cainj= -15
pressure [MPa]

diesel, cainj= -10 p2 biodiesel


12 8.5

pressure [MPa]
diesel, cainj= -6
p3
10 8.0
p7
7.5
8 -6 p4
p11
6 7.0
p6
4 6.5
p5, p8,p9,p10
-20 -10 0 10 20 30
6.0
crank angle [deg atdc] 0 5 10 15 20

Figure 20. Effects of physical property difference on crank angle [deg atdc]
combustion in a small bore light-duty engine operated
(a)
under boost pressure (Pivc=2.926 bar). Injection timings
were varied from -20 to -6 deg atdc. Solid lines are for 1
the biodiesel-surrogate and dotted lines are for the

CA50 difference [deg CA]


diesel-surrogate. 0

-1
engine, but the differences of ignition timings and peak
cylinder pressures are predicted to be larger. Therefore, -2
it can be stated that the effects of physical property
differences become more significant at operating -3
conditions where more spray impingement is expected.
In order to confirm the argument, in-cylinder spray -4
distributions for the two fuel cases are shown in Fig. 19, te re

va eat
liq ion
su pr sity

nt d

liq d
p. ff
p. c
va Cp
liq isc

od p

l
po en l

se
se

va Vis
la Con

n
va . Di
together with the corresponding temperature fields for

Bi C
ce su

Co
ns

h
.V

ie
ie

p.

.
rfa es

p
D
d

.
two injection timings.

te
r
va

In the case with earlier injection timing (-20 deg atdc), it


is expected that more spray droplets impinge on the (b)
walls (note that the ambient gas density at the time of
injection is lower than in later injection cases, thus the 100
liquid length is longer and the drop sizes after breakup
vaporized fuel @ EOI [%]

are bigger). However, earlier injection also allows a 90


longer time for fuel vaporization and fuel/air mixing in the
high temperature environment before ignition occurs,
which compensates for increased wall impingement. 80
Investigation of cases with various vapor pressures
indicated that the competition between the two effects 70
depends on the vapor pressures of the fuel (results not
shown). For the vapor pressure of the biodiesel- 60
surrogate used in the present study, the overall spray
impingement on the walls was predicted not to vary
va e at
te re
liq ion
rfa es y

liq d
nt d

p. ff
p . sc
va Cp
liq isc

od p

l
d el

se
su pr sit

la Con

n
va . Di

Bi C
ce su
s

va Vi
Co
ns

h
.V

ie
or e n
ie

p.

much for the injection timings considered in this study,


p
D

.
te

as seen in Fig. 19..


p
va

When the ambient gas density at the time of injection is (c)


increased to promote spray breakup by elevating the
boost level, it is evident from Fig. 20 that the reduced Figure 21. Sensitivity of combustion to individual
wall impingement helps form a better-mixed charge and physical properties. (a) Profiles of in-cylinder gas
the effects of physical property differences become less pressure near the time of ignition, (b) crank angle of
dependent on the injection timings. Figure 20 shows 50% heat release point compared to the baseline, and
very similar trends to those of the large bore engine (Fig. (c) vaporized mass at the time of end of injection
16) in terms of the pressure profiles for the combustion normalized by the total injected fuel. Base fuel is the
of the biodiesel and diesel surrogate fuels for the various biodiesel-surrogate. Small bore light-duty engine
injection timings. operated at the lower boost pressure. Injection timing
was –6 deg atdc.
9.5 Soot, biodiesel
diesel 0.14 140
Soot, diesel
9
0.12 NOx, biodiesel 120
p3, p4, p5, p6, p11
8.5 NOx, diesel
pressure [MPa]

0.1 100

Soot [g/kg-f]

NOx [g/kg-f]
8
0.08 80
7.5 p7
0.06 60
7
p1
0.04 40
6.5 p2 biodiesel
0.02 20
6
0 5 10 15 20 0 0
crank angle [deg ATDC] -25 -20 -15 -10 -5 0
start of injection command [deg atdc]
(a)
4.0 Figure 23. Predicted emissions of the biodiesel and
diesel surrogates for various injection timings. The small
CA50 difference [dgeCA]

3.5
3.0 bore light-duty engine with the lower boost pressure was
2.5 simulated
2.0
1.5
The sensitivity of the engine simulations to the individual
1.0 physical properties was investigated by replacing one
0.5 property of the base fuel with that of the other. Two
0.0 sensitivity runs with each of the two fuels set to the
properties of the base fuel were performed. All of the 11
te r e

va a t
liq ion
su pr ity

liq d
nt d

p. ff
p. c
va Cp
liq isc

od p

l
de e l

se
va Vis

physical properties were considered in the study.


l a on

va Di

Bi C
he
ce su
or n s
s

Co
ns
.V

ie
ie

p.
.C

.
p.
rfa es
D

te

Figure 21 shows the results of the sensitivity simulations


p
va

in terms of the pressure profiles near the time ignition,


(b) the crank angle of 50% mass burnt, and the vaporized
100
fuel mass at the end of injection. In this case, the base
fuel was the biodiesel-surrogate. The small bore light-
vaporized fuel @ EOI [%]

duty engine operated at the lower boost pressure was


90 simulated with injection timing of –6 deg atdc.

80 From the comparison of the results of the biodiesel and


diesel surrogate combustion, it is seen that no single
70
property has a dominant effect on the combustion
characteristics. However, the effects of the individual
properties are predicted to be significant.
60

As in the single drop vaporization cases (refer to Fig.


va at
te re
liq ion
su r pr sity

liq d
nt d

p ff
p . sc
va Cp
liq isc

od p

l
po den l

se
se

l a on

n
va Di

Bi C
he
ce su

va . Vi
Co
ns
.V

ie
ie

p.
.C

15), the liquid density (p1) and vapor heat capacity (p7)
p.
rfa es
D

te

are predicted to be very influential in the engine


va

combustion cases, as well. It is interesting that the vapor


(c) pressure does not seem to have as dominant an effect
on the combustion as in the single drop cases. It is
Figure 22. Sensitivity of combustion to individual notable that the surface tension change (p3) advanced
physical properties. (a) Profiles of in-cylinder gas the ignition timing significantly. As can be seen from the
pressure near the time of ignition, (b) crank angle of vaporized fuel mass at the end of injection shown in Fig.
50% heat release point compared the baseline, and (c) 21 (c), the reduction of surface tension from that of the
vaporized mass at the time of end of injection biodiesel-surrogate to that of the diesel-surrogate affects
normalized by the total injected fuel. Base fuel is the the breakup process, resulting in a smaller average drop
diesel-surrogate. Small bore light-duty engine with size, and this substantially enhances vaporization.
cr=16.5. Injection timing was –6 deg atdc. Although the vaporized mass in the case of density (p1)
variation is increased up to 92% of the diesel-surrogate
case by the end of injection, its ignition timing and 50%
mass burning point are still much more retarded than simulations. The same reaction mechanism was used
that of the diesel-surrogate, This indicates that the for the two fuels in order to separate chemistry effects
overall combustion behavior is described not by a single from physical property effects. Three different operating
property effect but by the coupled effects of all of the conditions were considered to study the coupled effects
properties. of the physical property differences and spray injection
conditions. Also, the effects of injection timing on
It is also interesting that the variation of individual
combustion phasing for the two fuels were investigated.
properties considered here does not always seem to
The sensitivity of combustion calculations to the
affect the combustion in the direction toward the diesel-
individual properties was also examined for non-reacting
surrogate case. For example, it was predicted that the
single drop evaporation and engine spray combustion.
change of latent heat (p6) and liquid heat capacity (p11)
Based on the results, the following conclusions were
slowed down the evaporation process and retarded the
drawn.
ignition timing and the 50% mass burning point most.
With the properties of the diesel-surrogate as the 1. Differences of the physical properties between the
baseline, the results of the sensitivity simulations that biodiesel and diesel surrogate fuels significantly affect
were obtained for the same operating conditions as used in-cylinder mixture conditions before ignition, as well as
for the biodiesel-based case are shown in Fig. 22. Vapor the subsequent combustion process. Biodiesel
pressure (p2) was again the most influential property in properties tend to slow down vaporization, increase fuel
terms of its effect on ignition timing, while the 50% impingement on the walls, delay the start of ignition, and
burning point was most sensitive to the density change lower the peak in-cylinder gas pressure.
(p1). It is interesting that the 50% burning points were
later than those of the diesel-surrogate for all the 2. The effects of physical property differences on mixture
sensitivity runs considered in this case. Overall, the preparation tend to be reduced as the injection timing is
sensitivity of the engine combustion calculations to the advanced. However, the injection timing effects are
individual physical properties was much less than that in strongly coupled with the extent of fuel impingement on
the biodiesel-based case. Therefore, it does not seem to the combustion chamber walls.
be appropriate to use the properties of the diesel-
surrogate for spray combustion simulations of the 3. From the single drop evaporation results, the liquid
biodiesel-surrogates with only some of the properties density and vapor pressure were found to be the most
replaced with those of the biodiesel-surrogate. Accurate influential properties among the 8 liquid phase fuel
estimation of as many individual physical properties of properties considered in that study.
biodiesel as possible is desirable for reliable predictions
of biodiesel fueled engine operation. 4. The results from the engine combustion study indicate
that vapor pressure differences do not affect the
The soot and NOx emissions of the two fuel cases are simulations as much as would be expected from single
compared in Fig. 23. As the injection timing is retarded, drop studies due to the competing effects of other fuel
ignition timings are delayed and the peak values of the properties. When the physical properties of the diesel-
local maximum and average temperatures decrease. surrogate fuel were chosen as the baseline values,
This results in deceased NOx emissions and a slight calculations of spray combustion were found to be much
increase in soot emissions. However, the predictions of less sensitive to changes in individual property values
the NOx and soot magnitudes do not agree with corresponding to those in the biodiesel-based case. This
experimental observations that NOx increases and soot indicates that it is important to accurately specify all of
decreases with increasing biodiesel fraction [2,3,7]. the physical properties for biodiesel fuel, since they have
These observations probably result from differences in a coupled effect.
the chemical oxidation mechanisms of diesel and
biodiesel, which were not modeled in the present study. 5. No single physical property explained the differences
The effect of reaction chemistry differences between the seen in the ignition delay between the biodiesel and
two fuels is under investigation using a reduced kinetics diesel surrogate fuels. The overall differences in
mechanism for a biodiesel surrogate fuel [9]. combustion behavior are due to the coupled effects of
Nonetheless, the usefulness of the present differences of multiple properties rather than a single
computational predictions for explaining the effects of dominant property.
the physical property differences is well demonstrated in
the present study. 6. Replacement of the physical properties alone may
not capture the emission characteristics associated with
biodiesel-fueled engine operation reported in literature.
SUMMARY AND CONCLUSIONS Therefore, an appropriate chemistry mechanism such as
that of Brakora et al. [9] needs to be incorporated for
Using the physical properties of the biodiesel-surrogate accurate prediction of emissions.
fuels obtained from accurate compilations of the
individual components, numerical simulations of ACKNOWLEDGMENTS
combustion in small and large bore compression ignition
engines were performed and compared with This research was partially sponsored by the Laboratory
corresponding results from diesel fueled engine Directed Research and Development Program of Oak
Ridge National Laboratory (ORNL), managed by UT- characterization of biodiesel low sulfur diesel fuel
Battelle, LLC for the U. S. Department of Energy under blends,” National Biodiesel Board, U. Missouri, Dec. 20,
Contract No. DE-AC05-00OR22725. The authors thank 1995.
ORNL for funding support. 15. Herb, S.F., Magidman, P., Reimansc, R.W.,
“Observations on response factors for thermal
REFERENCES conductivity detectors in GLC analysis of fatty acid
methyl esters,” J. Am. Oil Chemists Soc. 44(1), p. 32,
1. Fuel Fact Sheets, National Biodiesel Board, 1967.
www.biodiesel.org. 16. Tseng, C.C., Viskanta, R., ”Effect of radiation
2. Akasaka, Y., Suzuki, T., and Sakurai, Y. 1997. absorption on fuel droplet evaporation,” Combustion
”Exhaust Emissions of a DI Diesel Engine Fueled with Science and Technology 177(8), 1511-1542, 2005.
Blends of Biodiesel and Low Sulfur Diesel Fuel,” SAE 17. Huber, M.L., Perkins, R.A., “Thermal conductivity
972998, 1997. correlations for minor constituent fluids in natural gas: n-
3. McCormick, R. L., Alvarez, J. R., Graboski, M. S., octane, n-nonane, and n-decane,” Fluid Phase Equilibria
Tyson, K. S., and Vertin, K., ”Fuel Additive and Blending 227, 47-55, 2005.
Approaches to Reducing NOx Emissions from 18. Skelland, A.H.P., Diffusional Mass Transfer, Krieger
Biodiesel,” SAE 2002-01-1658, 2002. Publishing Company, Malabar FL.
4. Szybist, J. P., and Boehman, A. L., “Behavior of a 19. Chung, T.-H., Ajlan, M., Lee, L.L., Starling, K.E.,
Diesel Injection System with Biodiesel Fuel,” SAE 2003- ”Generalized multiparameter correlation for nonpolar
01-1039, 2003. and polar fluid transport properties,” Ind. Eng. Chem.
5. Tat, M. E., Van Geroen, J. H., and Wang, P. S., Res. 27, 671-679, 1988.
”Fuel Property Effects on Injection Timing, Ignition 20. Chung, T.H., Lee, L.L., Starling, K.E., “Applications
Timing and Oxides of Nitrogen Emissions from of kinetic gas theory and multiparameter correlation for
Biodiesel-Fueled Engines,” 2004 ASAE/CSAE Annual prediction of dilute gas viscosity and thermal
International Meeting, Ottowa, Ontario, Canada, August conductivity,” Ind. Eng. Chem. Fundam. 23, 8-13, 1984.
1-4, 2004. 21. Duran, A., Carmona, M., Monteagudo, J.M.,
6. Cheng, A.S., Upatnieks, A., and Mueller, C.J., “Modeling soot and SOF emissions from a diesel
“Investigation of the impact of biodiesel fuelling on NOx engine”, Chemosphere 56(3), 209-225, 2004.
emissions using an optical direct injection diesel engine,” 22. Fisher, E.M., Pitz, W.J., Curran, H.J., Westbrook,
Int. J. Engine. Research, 7, 297-318, 2006. C.K., “Detailed chemical kinetics mechanisms for the
7. Marchese, A. J., Vaughn, T. L., Hammill, M., and combustion of oxygenated fuels,” Proceedings of the
Harris, M., “Ignition Delay of Bio-Ester Fuel Droplets,” Combustion Institute 28, 1579-1586, 2000.
SAE 2006-01-3302, 2006. 23. Tamim, J., Hallett, W.L.H., ”A continuous
8. Chakravarthy, K., McFarlane J., Daw, S.C. Ra, Y. thermodynamics model for multicomponent droplet
and Reitz, R.D., “Physical Properties of Soy Bio-diesel & vaporization,” Chem. Eng. Science 50, 2933-2942,
Implications for Use of Bio-diesel in Diesel Engines,” 1995.
SAE 2007-01-4030, 2007. 24. Tat, M.E., van Gerpen, J.H., ”The specific gravity of
9. Brakora, J.L., Ra, Y., Reitz, R.D., McFarlane J., biodiesel and its blends with diesel fuel,” The Journal of
and Daw, S.C., “Development and Validation of a the American Oil Chemists’ Society 77(2), 115-119,
Reduced Reaction Mechanism for Biodiesel-Fueled 2000.
Engine Simulations,” submitted to SAE 2008 World 25. Yaw, C. L., CRC Handbook of Chemicals and
Congress, 2007. Physics, various editions; and Chemical Properties
10. Yuan, W., Hansen, A.C., Zhang, Q., ”Predicting the Handbook, McGraw-Hill Publishing Company, 1999.
physical properties of biodiesel for combustion 26. van Bommel, M.J., Oonk, H.A.J., van Miltenberg,
modeling,” Transactions of the ASAE 46(6), 1487-1493, J.C., ”Heat capacity measurements of 13 methyl esters
2003. of n-carboxylic acids from methyl octanoate to methyl
11. Reid, R.C., Prausnitz, J.M., Sherwood, T.K., eicosanoate between 5K and 350K,” J. Chem. Eng. Data
“Properties of Gases and Liquids, 4th Ed., New York, 49, 1036-1042, 2004.
N.Y., McGraw-Hill, 1987. 27. Gallant, R.W., and Yaws, C.L., Physical Properties
12. Yamane, K., Ueta, A., Shimamoto, Y., “Influence of of Hydrocarbons, 3rd ed., Houston, Gulf Pub. Co., 1995.
physical and chemical properties of biodiesel fuel on 28. Amsden, A.A., KIVA-3V, Release 2, Improvements
injection, combustion, and exhaust emission to KIVA-3V, LA-UR-99-915, 1999.
characteristics in a DI-CI engine,” In Proceedings of the 29. Kee, R.J., Rupley, F.M., Miller, J.A., “CHEMKIN-II: A
5th International Symposium on Diagnostics and FORTRAN Chemical Kinetics Package for the Analysis
Modeling of Combustion in Internal Combustion of Gas Phase Chemical Kinetics,” Sandia Report SAND
Engines, COMODIA 2001, July 1-4, 2001, Nagoya. 89-8009, 1989.
13. Allan, C.A.W., Watts, K.C., Ackman, R.G., Pegg, 30. Beale, J.C., and Reitz, R.D., “Modeling Spray
M.J., Fuel 78, 1319, 1999. Atomization with the Kelvin-Helmholtz/Rayleigh-Taylor
14. Schumacher, L., Chellappa, A., Wetherell, W., Hybrid Model,” Atomization and Sprays, 9, 623-650,
Russell, M.D., “The physical and chemical 1999.
31. Liu, A.B., Mather, D., Reitz, R.D., “Modeling the Combustion in a Light Duty Diesel Engine,” SAE 2007-
Effects of Drop Drag and Breakup on Fuel Sprays,” SAE 01-0193, 2007.
930072, 1993. 36. Patel, A., Kong, S.C., and Reitz, R.D., “Development
32. Ra, Y., and Reitz, R.D., “A Model for Droplet and Validation of a Reduced Reaction Mechanism for
Vaporization for Use in Gasoline and HCCI Engine HCCI Engine Simulations,” SAE 2004-01-0558, 2004.
Applications,” Journal of Engineering for Gas Turbines & 37. http://www.me.berkeley.edu/gri_mech/
Power, 126(2), 422-428, 2004. 38. Kong, S.C., Sun, Y., and Reitz, R.D., “Modeling
33. Ra, Y. and Reitz, R.D., “The Application of a Multi- Diesel Spray Flame Lift-Off, Sooting Tendency and NOx
Component Vaporization Model to Gasoline Direct Emissions Using Detailed Chemistry with
Injection Engines,” Int. J. of Engine Research, 4(3), 193- Phenomenological Soot Model,” Journal of Gas Turbines
218, 2003. and Power, 129, 245-251, 2007.
34. O'Rourke, P.J., and Amsden, A. A., “A spray/wall 39. Hiroyasu, H. and Kadota, T., “Models for
interaction submodel for the KIVA-3 wall film model,” Combustion and Formation of Nitric Oxide and Soot in
SAE 2000-01-0271, 2000. DI Diesel Engines,” SAE 760129, 1979.
35. Opat, R., Ra, Y., Gonzalez, M.A., Krieger, D., R., 40. Araújo, M.E., Meireles, M.A.A., “Improving phase
Reitz, R.D., Foster, D. E., Durrett, R.P., and Siewert, equilibrium calculation with the Peng-Robinson EOS for
R.M., “Investigation of Mixing and Temperature Effects fats and oils related compounds/supercritical CO2
on HC/CO Emissions for Highly Dilute Low Temperature systems.” Fluid Phase Equilibria 169, 49-64, 2000.

APPENDIX A. A model of transcritical droplet be determined from the thermal diffusivity, α (= k / ρcv ) ,
vaporization. and a characteristic heating time, τc, i.e.,
During the lifetime of drop evaporation at supercritical
ambient conditions, the vapor pressure of a liquid droplet k
l c ~ ατ c = τc A-(1)
at a subcritical liquid temperature cannot exceed the ρc v
ambient pressure, thus it cannot be subjected to boiling
evaporation. However, when droplets are heated by heat
transfer from the surrounding gases at supercritical where k and ρ are the thermal conductivity and density
conditions, which is possible under diesel condition, the of the liquid droplet, respectively, evaluated at the
surface temperature of the droplet can reach the critical interior drop temperature.
mixing temperature, while the temperature of the drop
interior remains sub-critical. In this case, the phase The characteristic heating time can be obtained from the
change near the critical mixing line on a phase elapsed time to heat up the entire drop mass from the
equilibrium plot is very complex and involved due to high initial temperature to final temperature assuming no
pressure gas absorption and non-ideal gas behavior. In critical phase change from the drop surface and given
order to model the phase change of liquid fuel drops the heat flux from the drop surface to the drop interior.
under transcritical conditions, a simple model that The energy balance equation for this heating process is
employs the concept of effective radius of phase change given as
was used to estimate the amount of fuel to be treated as
vaporized during a given time step during the (Td , f − Td )
evaporation calculation. ρVcv ~ Ah f ,eff (Ts − Td )
τc
The assumptions made in the model include; 1) the
critical mixing line is independent of ambient pressure, or
i.e., the critical mixing temperature is equal to the critical
temperature. 2) no absorption of gases into the liquid rd (Td , f − Td )
drop is assumed. 3) the droplet and the surrounding gas τ c ~ ρcv A-(2)
mixtures are bounded by a sharp phase transition 3 h f ,eff (Ts − Td )
boundary. 4) the density of the liquid drop interior is
uniform and is evaluated at the average drop interior where V is the volume of the drop, cv is the specific heat
temperature. 5) only the temperature gradient within the of the liquid drop, Td,f is the final drop temperature after
drop-gas interface is considered in evaluation of heat heating, Td is the initial drop temperature, A is the drop
transfer into the drop interior. 6) a quasi-steady state surface area, hf,eff is the effective heat transfer coefficient
vaporization process is assumed. that takes the effects of internal circulation of the liquid
drop into account, rd is the drop radius, and Ts is the
The effective radius of a droplet beyond which the liquid drop surface temperature, which is equal to the critical
is regarded vaporized is calculated from the temperature. Therefore, the effective radius of phase
characteristic length of heat transfer at the drop surface change is determined using a model constant, χ, as
that is at the critical temperature. From dimensional
analysis, the characteristic length of heat transfer, lc, can
k r (Td , f − Td )
l c ~ ατ c = χ ρcv d
ρcv 3 h f ,eff (Ts − Td )

rd k (Td , f − Td )
=χ A-(3)
3h f ,eff (Ts − Td )

A constant value of χ=0.05 was used, in the present


study. The transcritical vaporization model was
combined with the evaporation model reported in Ref.
[32].

View publication stats

You might also like