Вы находитесь на странице: 1из 9

CHALMERS TEKNISKA HOGSKOLA

Institutionen for Termo- och Fluiddynamik

CHALMERS UNIVERSITY OF TECHNOLOGY


Department of Thermo and Fluid Dynamics

Analysis of Assumptions Commanding Detailed Chemistry


EDC-Based Model for Diesel Spray Combustion
by
V. I. Golovitchev

Paper submitted to the 6th International Symposium on Diagnostics


and Modeling of Combustion in Internal Combustion Engines
Yokohama, Japan, August 2-5, 2004

Horsalsvagen,
7
412 96, Goteborg - Sweden
http://www.tfd.chalmers.se/ valeri
phone: +46-31-772-1396 e-mail: valeri@tdf.chalmers.se fax: +46-31-18-09-76

Analysis of Assumptions Commanding Detailed Chemistry


EDC-Based Model for Diesel Spray Combustion
V.I. Golovitchev

Chalmers University of Technology


412 96, Goteborg - Sweden
Keywords: Modeling of diesel sprays, detailed chemistry, MK combustion

ABSTRACT
The new eddy dissipation concept (EDC) model based on the operator-splitting procedure applied
to the mass conservation equations for species participating in reversible chemical reactions representing combustion in a PaSR partially stirred reactor (PaSR) volume is analyzed. The model
has been implemented in the KIVA-3V code together with a new droplet collision procedure, and
its application to spray combustion simulation is illustrated by the results of the 3-D modeling of
the DI Diesel Volvo D12C engine. The chemical mechanism of diesel oil surrogate consisted of 68
species (including soot aromatic precursors) participating in 285 reversible reactions has been applied to simulation of MK (Modulated Kinetics) diesel spray combustion regime.Mean features of
MK combustion under conditions of delayed injection when auto-ignition has much in common with
the HCCI process are predicted in accordance with experimental data.
INTRODUCTION
The MK combustion process based on a lowtemperature premixed mode was suggested for simultaneous reduction of NOx and particulate matter. To combat emissions formation, the injection
timing is retarded extending the ignition delay and
premixed combustion mode contribution. Optimization of the process requires modeling based on the
detailed chemistry. Since the time scales for kinetic
and turbulent micro-mixing are comparable during
the long ignition period characteristic of the MK
mode, turbulence/chemistry interaction must be accounted for. The Eddy Dissipation Concept (EDC)
is a promising technique to facilitate such combustion analysis. However, the EDC is the model which
was not yet analyzed in details, albeit the model is
said to be implemented in practically all commercial
codes available. For example, in the latest review by
Veynante and Vervisch published recently in [1], the
model has been referred as belonging to the family
of infinitely fast chemistry models and summarized
in only 15 half-lines of the following text: The eddy
dissipation model (EDC) is a direct extension to nonpremixed flame of the eddy break up (EBU) closure
initially devoted to turbulent premixed combustion...
The fuel burning rate is calculated according to:
Yo
Yp

),
wf =
min(Yf , ,
k
s
1+s

(1)

where and are adjustable parameters of the closure. In the above equation, ... the reaction rate is
limited by a deficient species. To account for the existence of burnt gases bringing the energy to ignite
the fresh reactants, this species may be the reaction
product (when is non-zero)... Actually, for large ,
the model is difficult to justify.
In fact, the above review attributes the EDC model
only to its first, Magnussen and Hjertager, M-H, formulation [2]. In this model, the reaction product
term (when is non-zero) is introduced into the rate
expression to account for finite-rate chemistry effects (ignition/quenching) that makes it incorrect to
place the EDC approach in the family of infinitely
fast chemistry models. Some other versions of so
called EDC model, e.g. [3], were formulated without a proper connection with the original EDC approach. Also, Fluent [4] recently claimed the implementation of a new version of the EDC model ...for
the modeling of turbulent finite-rate chemistry. It is
completely different from the eddy dissipation model
that has long been available. This paper is aimed to
fill a gap in the EDC based models analysis starting
with an outline of the classic EDC model features.
Since the retarded injection was simulating on a
relatively coarse mesh generated during the expansion stroke, it required a modification of the collision
model because some default collision models introduce computational artifacts into the spray simula-

tion especially pronounced on coarse meshes.

rewritten as
dc1
c1 c o
c c1
=
= fr (c) =
d

mix

NEW PHYSICAL MODELS


where

Turbulence/Chemistry Interaction
The new model formulation is based on the
operator-splitting procedure applied to the reactive
species conservation equations. In terms of this approach, the time differencing was performed in three
steps: the first step was assumed to be the convection
contribution, the second one - diffusion effect without a contribution of micro-mixing, and the third step
was the chemical kinetics effect coupled with micromixing. The last step of such a mass balance can be
interpreted as representing combustion in a constant
volume partially stirred reactor (PaSR) of a computational cell size, where reactions occur in a fraction of
its volume. Since the reaction zone parameters, as a
rule, can not be resolved on a computational grid, the
sub-grid diffusion term due to micro-mixing was approximated with the help of introduction of a micromixing time as suggested in the interaction with the
mean approach [5]. Then, assuming that chemical
processes proceed in such a way that the shortest
chemical time associated with the particular (reference) species was constrained by the micro-mixing
time, a system of PaSR equations consisting of two
sets can obtained
dc1
c
c1 c o
c c c1
= ,
=
= ,
dt

c
mix
c

(2)

where is the time increment, c is the chemical


reaction time, and mix is the micro-mixing time.
The model distinguishes (see [6] for more details) between the concentration (in mean molar density) at
the reactor exit, c1 , the concentrations in the reaction zone, c, and in the feed, c0 . When time proceeds,
c1 trades place for c0 .
After algebraic manipulations, one can yield the
analytical solutions of the problem (2) which, finally,
can be represented as
c1 c o
c1
1
c1 c1
= ( ) = H( ,
),

c
2
c mix

(4)

(3)

where = c /(mix + c ), and H is a harmonic mean.


The solution (3), if the process is expressed in
terms of the reactor output parameters, shows that
the turbulent combustion time is a sum of the mixing and reaction times, as the main consequence of
the PaSR model.
For the detailed chemical mechanism, the species
production rate of the particular r-chemical reaction
(separated into creation and destruction rates) contrary to the first equation of the system (2) can be

fr (c) = (r00 r0 ) r (c)

where r00 and r0 are stoichiometric coefficients of the


backward and forward stages, r is a progress variable of the r-reaction. Species indices above are omitted for simplification.
The reaction progress variable r (c) is calculated
from the mass-action law
r (c) = kfr (T )

Ns
Y

(cs )

0
sr

s=1

kbr (T )

Ns
Y

(cs )

00
sr

s=1

where kfr and kbr are the rate coefficients for the forward and backward stages of the r-reaction, T is the
temperature, and Ns is the total number of species
in the mixture.
There is a substantial difference between Eqs (2)
and (4) - the first model has the analytical solution,
while the application of the second one requires numerical integration. Combination of the terms in the
model (4) leads to a definition of the reactive volume
as introduced in [7]:
c1 = c + co (1 ),

(5)

where the multiplier = /( + mix ) is wedged between 0 and 1, while is the integration step.
Rewriting the first relation of (4) in terms of the
reactor exit parameters, one can get
c1 c o

= fr (c) ' fr (c1 ) + (fr /c)|c=c1 (c c1 )


= fr (c1 )

c c1
c

The rhs of the above equation is constructed using


Taylors expansion at the state c1 , assuming that the
reaction times can be estimated as reciprocal values of the Jacobian matrix elements evaluated at
unknown values c=c1 , i.e., c |fr /c|1 and that
(fr /c)|c=c1 < 0. Elimination of c-values using c1 values taken from Eq. (5)
c1 c o
c1
co (1 )
= fr (c1 ) [
c1 ]/c ,

(6)

if the reaction time is defined as


c 1

= fr o /(cs o + term
r ),

immediately puts the reaction rate into the form


c1s cos

c
c + mix
cs o f r o
,

cs o + term
r + termr mix

= fr (c1 )
=

(7)

containing no reaction zone parameters c which cannot be resolved on a grid, but replacing their effect
with the rate multiplier . The superscript o and the
subscript s denotes the value at a start of the integration step and the index of the reference species
defined below, respectively.
This expression looks different from a finite-rate
formulation of the original EDC model [8] known as
ws =

M
c c1s
s
,
1 M
mix

(8)

where M is the fine structure volume, and denotes


the fine-structure quantities, but a generic proximity
of the approximations can be established using the
PaSR formulation.
The chemical reaction times are formally defined
as characteristic times of the destruction rates, if the
species chemical production rates are presented as a
sum of creation and destruction terms. In this way,
for each reaction of the mechanism, chemical times
can be attributed to each species participating in the
depletion stages, e.g., -kr ci ck ci /c,i ck /c,k ,
where i, k are species indices. In terms of the EDC
model, the shortest time is restricted from below by
the micro-mixing time that explains the usage of deficient (limiting) species in the rate expression (1).
Application of the rate expression (7) to a linear
approximation of the Arrhenius reaction:
fro = cos /c
leads to the formula
cos +

cos fro

termr + term
r mix
c1s
=

c + mix

cos
+ c + mix

1
c1 c1
= H( s , s )
2
c mix

which is perfectly consistent with the PaSR analytical solution (3). In a limiting case of chemical equilibrium, the model approaches the M-H model.

Chemical Time Definition


The reference species is similar to the limiting
species used in the M-H model [2] reaction rate definition. Let us consider the balance equation of sspecies participating in the r-reaction:
c1s

cos

= ( 00 sr 0 sr )[kfr (T )1 kbr (T )2 ] =

[ 0 sr kbr 2 + 00 sr kfr 1 ] [ 0 sr kfr 1 + 00 sr kbr 2 ],


{z
} |
{z
}
|
term
r

term
r

where

1 =

Ns
Y

n=1

(c1n )

0
nr

, 2 =

Ns
Y

n=1

(c1n )

00
nr

For further analysis, the above equation written in


a form separating the net chemical production rate

into creation, term
r , and destruction, termr , terms
is assumed to be linearized as
c1s cos

1
cs | =0

= (term
r

c1s
c
)
c c + mix

(9)

= cos ,

This linearized expression is mostly accurate for the


species which concentration is less than those of the
other reaction partners. Thus, such a species (called
reference species) is playing a special role in the
analysis. The factor before this species has a dimensional representation of an inverse time. This time
is the shortest time for the particular reaction. By
algebraic manipulation with Eq. (9), one can get the
unknown terms in Eq. (7), viz., fr (c1 ) and c :
f r (c1s )
c 1

cos for
,
cos + term
r

= (f r /c)|c=c1 fr o /(cs o + term


r )

As shown in [7], introduction of reference species assures the equilibrium conditions implementation at
for each reaction in the mechanism.

Micro-mixing Time Definition


The correct definition of the micro-mixing time is a
matter of importance for any EDC turbulent combustion model. The micro-mixing time definition taken
in this study is applicable to the system with natural initial mixture non-uniformities and based on
the -model 1 proposed by Frisch [9, p.135-140]. In
terms of this approach, at each stage of the Richardson cascade, the number of eddies formed from a
given parent eddy is chosen such that the fraction
of volume occupied by active eddies is decreased by
a factor < 1. In this way, the intermittency is introduced in a geometric form. The factor is an adjustable parameter of the model. If lo is the initial
eddy size, and l is the size of the eddy formed on the
n-th step of fragmentation, the fraction pl of the active space can be calculated as
l
pl = n = ( )3D ,
lo

(10)

where the exponent D can be interpreted as a fractal


dimension of the turbulent field. The expression (10)
can be derived by considering the analytical representation of the Richardson cascade in the form:
f (x)f (y/x) = f (y)f (1),

(11)

1 This has nothing in common with that mentioned in the


review paper [1].

where f (1) is a number of eddies at the initial stage,


f (x) = f (lo /) is a number of -fragments at the particular stage of fragmentation, x = lo /, y = lo /,
and , < are fragment sizes. The above equation
has a general solution- the fragmentation law
N f (x) = CxD ,

(12)

where C and D are the solution parameters, N is the


number of fragments. The fraction of the active volume can be calculated as
V olN

= f (x) 3 = C(lo /)D 3


= CloD 3D = Clo3 (/lo )3D

This relation is identical to the expression (10). Since


the active eddies of size l fill only a fraction pl of a
total volume, the specific energy associated with this
scale is
l
El = vl 2 pl = vl 2 ( )3D
lo

(13)

assuming the energy flux from scales l to smaller


scales is El /l , where l = l/vl is the eddy turnover
time. In the inertial range of scales, the energy flux
is independent of l, i.e.,
 = vo3 /lo ,
where  is a dissipation rate of turbulent kinetic energy. After some algebra, one can get the expression
for the eddy turnover time
l =

lo l 2 + 3D
( )3 3
v o lo

(14)

The viscous cutoff scale of the -model is obtained


by equating the eddy turnover time and the viscous
diffusion time d = d2 / that gives an expression for
the dissipation scale
3

d = lo Re 1+D ,

(15)

where Re = lo vo /, and is molecular viscosity.


When D = 3, the classic Kolmogorov expression is recovered; his assumption neglects the intermittency.
If lo is defined in terms of the k  model, from Eq.
(15), the viscous dissipation time can be derived as
d = mix

= ( ) [( )1/2 ](1)


1
k
= ( ) (c /Ret ) 2 ,


(16)

where = 3(D3)
1+D , and k and  are the turbulent
kinetic energy and energy dissipation rate, respectively. From experiments reported in [10] follows
that the fractal dimension for turbulent dissipation
is about D=2.7, i.e
k
mix = ( ) (c /Ret )0.621


In particular, the eddy break-up time k/ corresponds to D=5, and Kolmogorov micro-scale time
k = (/)1/2 - to D=3.
Another approach developed is based on the usage
of Kolmogorov expression for the micro-mixing time
with the molecular viscosity replaced by a fraction of
the effective viscosity assumed to be responsible for
micro-mixing. If, for example, the RNG k  model
is employed, the turbulent viscosity related to k, the
turbulent kinetic energy, and , the dissipation rate
of k, is given by the general expression.
s
c k 2 / 2
t = l (1 +
)
(17)
l

= (l + s t + 2 l s t ),
where s t is the standard k- kinematic viscosity.
Provided that the first two terms in the expression
(17) contribute to conventional diffusion transport,
the geometrical mean term can be assumed to determine the characteristic time scale for micro-mixing
in the turbulence/chemistry interaction model, if
written in a form similar to the Kolmogorov time definition, k = (l /)1/2 replacing the molecular viscosity by its turbulent counterpart, i.e.,
mix = (2

p
k
l ts /)1/2 = (2 c k )1/2 ,


(18)

where c =0.09 is the constant of the k- model.


The above expression formally corresponds to
D=3.8 in the general formula (16) giving only a provisional value for mix which must be refined in comparisons with experiments. The model is assumed to
be valid across a full range of flow conditions from
low to high Reynolds numbers, if k and  are determined from the transport equations of the compressible RNG k-  model.

Diesel Fuel Surrogate Model


Practical diesel fuels consist of a great number of
aliphatic and aromatic compounds, and their combustion is too complex to be modeled using a comprehensive chemical mechanism. Despite the existence
of the accurate multi-component evaporation model
such as the distillation curve (DC) model, [11], the
combustion processes in practical numerical simulations are feasible only for the fuel surrogates presenting a bland of very few constituent components.
Aliphatic components can be represented by long
chain hydrocarbons such as n-heptane or n-dodecane
to their cetane number of approximately 56, which
is similar to the cetane number of conventional diesel
fuel. Aromatic components significantly contribute
to soot formation. The diesel fuel surrogate, which
can sufficiently represent the cetane number as well

as the other properties of real fuel is assumed to be


a 70/30 % mixture of n-heptane, C7 H16 , and toluene,
C7 H8 . The alternative model consisting of the mixture of n-dodecane and -methylnaphthalene is also
conceivable. Since the physical properties of the fuel
surrogate are represented by real diesel oil properties, the difference between seems not critical.
Features of Fuel Surrogate Model
(Time history of surrogate and its components)
0.04

Mole fraction

0.02

0.01

0.0002

0.0004
0.0006
Time, s

0.0008

The simulation of a spray liquid core fragmentation starts by injecting droplets of the size of the
nozzle orifice. The followed droplet breakup is modeled by providing an initial drop size distribution in
a form of the power law as proposed in [16]
 n+1 r n
if 0 r ro
ro ( ro )
f (r) =
,
(19)
0
otherwise
where n=0.5, and r and ro are the drop and injector
nozzle radii, respectively.
Since in the KIVA code the droplets are injected as
parcels of the identical mass, the distribution r3 f(r)
normalized to unit total mass is used to describe the
parcel size distribution
 n+4 r n+3
if 0 r ro
ro ( ro )
(20)
g(r) =
0
otherwise

C14H28
C7H16
C7H8

0.03

Modified Spray Models

0.001

Figure 1: Calculated fuel surrogate and its component time histories in the course of combustion development in the premixed mixture (=0.5) at po =41.0
bar, To =900 K.
Detailed reaction mechanisms for n-heptane and
toluene oxidation have been developed, validated using shock-tube and constant volume auto-ignition experimental data, reduced (using SENKIN code and
reaction path analysis software of the Chemkin-2
[12] package) to the mechanism consisting of 68
species participating in 285 reactions and implemented into the KIVA-3V code [13] to simulate diesel
spray combustion. This mechanism was proven (see
[14]) to predict the fuel auto-ignition, combustion development and formation of polycyclic aromatic hydrocarbons (PAHs) till gaseous soot precursors based
on acenaphthylene, A2R5, then converted directly
into soot, the substance assumed to have graphite
properties. The mechanism for both aliphatic and
aromatic components includes principal fragments
of C1 -C7 and O/H chemistry taken from different
sources, but mainly from LLNL and MIT databases
posted in the Internet. The fuel surrogate decomposition into constituent components is treated as
two global stages of oxidative pyrolysis. The reaction
mechanism and species thermodynamic properties
are provided by [15]. Fig. 1 illustrates the dynamics of the diesel fuel surrogate decomposition into
constituent components in chemical kinetics simulations. The physical properties of the fuel surrogate
were taken as properties of real diesel fuel compiled
in the DI model of the KIVA-3V fuel library with the
original fuel chemical formula replaced by C14 H28 .

In fact, the spray behavior is not very sensitive to the


value of the exponent n.
Optimization of retarded as well as multiple (pilot and post) injections can be realized on a relatively coarse computational grid. Since the standard
droplet collision model implemented in the KIVA
code is inherently grid dependent (see [17]) due to
the collision frequency depends on computational
cell sizes, the collision model was modified based on
the approach of [18] reducing this grid dependence.
There are few other algorithms of such a type, mostly
based on approaches used in molecular dynamics.
The standard collision procedure is as follows. The
probability Pn that a droplet undergoes n collisions
with other droplets is calculated according to the
Poisson distribution law
Pn =

n
exp(),
n!

(21)

where = t is the mean expected number of binary droplet collisions during a computational time
increment, t, and the collision frequency between
the droplets reads
=

n2
(R1 + R2 )2 |~u12 |
V ol

Above, n2 is a number of drops in a largest parcel 2,


R1 , R2 and ~u12 = ~u1 ~u2 are the radii and approach
relative velocity of colliding droplets, respectively.
The computational grid artifact is introduced in
the modeling due to the presence of the cell volume
V ol where the droplets can collide. In the new model
implemented, the droplet collision is impending if
and only if the geometric requirements are satisfied
~u12 ~r12
t |~r1 ~r2 | (R1 + R2 )
|r12 |
12 crit ,

where ~r1 and ~r2 define droplet positions, 12 is the


angle between the vectors ~v12 and ~r12 , crit defines
an angle of the collision cone in which the colliding
parcels are distributed.
Finally, the parcel will not collide if the random
number sampled is less than Po , the initial probability in the Poisson law, defined as
Po = min(1, C1

R1 + R2 C2
t ,
)e
12

(22)

where is defined as the residence time of colliding


droplets in the vicinity (characterized by the minimum distance 12 ) of the apparent intersection of
parcel assumed straight-line trajectories; C1 and C2
are the model constants to be determined. Description of the outcome of collisions, coalescence or separation, was left unchanged.
The modified spray models were validated using
experimental data obtained in constant volume vessels. In particular, the collision model was proven to
be grid independent, but requires more validation.
MODEL APPLICATIONS
To complete the EDC model analysis, let us consider the finite-rate chemistry EDC based on the rate
expression (8) implemented in the Fluent-6 code [4].
The Fluent model can be deduced from the twoset PaSR model (2) by assuming that combustion
occurs in the sub-grid fine structures submerged in
a cell-size reactor with initial conditions following
from the fact of identity of both sets of the system
(2) at = mix . This allows replacing c1 co in
the first set by c c1 , and to integrate equations of
the first set numerically with a stiff ODE solver 2
until Kolmogorov micro-mixing time giving the reaction zone parameters. This integration procedure,
in fact, replaces the reaction zone parameters in the
rate expressions with the non-identical reactor values. Thus, the system of equations becomes overdetermined, and the second set of system (2) can be
used to define the reaction source term as a new solution variable. However, the general mathematical
structure of the PaSR model was not much violated,
and, as a consequence, the solution when applied to
the linear one-step Arrhenius reaction, appears to be
similar to that of the sub-grid PaSR model. To establish a connection with the PaSR reaction rate, let us
calculate the EDC source term (8) for this case:
cs
cs c1s
= f (c) =

c
Solving the above equation, one can get
c
cs = c1s
+ c
2 Usage of a PSR solver of the Chemkin-2 (or Cantera) package
with the residence time equal to mix seems more reasonable.

Since in Eq. (8), M /(1 M ) = mix / , and keeping


in mind that the integration must proceed over mix
time, one can yield
M cs c1s
c1
= s =
1 M mix
c +
1
c1s c1s
c1s
= H( ,
),

c + mix
2
c mix

f (c) =

that is once again consistent with the PaSR solution.


Thus, the EDC combustion model is characterized by
a unique set of postulates.

Diesel Engine Modeling


The computational model developed has been applied to simulations of the axisymmetric bowl-inpiston Volvo D12C DI Diesel heavy-duty engine geometry with a peak in the center of the bowl. Details
of the engine combustion chamber and injector specifications are given in Tab. 1.
Table 1: Volvo D12C engine specification.
Bore
Stroke
Squish
Connecting rod
Injector nozzle dia ()
Engine speed
Injection timing
Injection period
Injection mode
Injected mass/stroke
Injection vel.
Initial pressure
Initial gas temp
Inclination angle of spray
Spray cone angle
Initial droplet temp

131 103 m
150 103 m
1.0 103 m
260 103 m
0.235 103 m
1500-1700 rpm
0-5.8 ATDC
8.4 CA
velocity table
0.062 g
300 - 500 m/s
1.31 bar
340-350 K
153.0 deg
9.5 - 12.5 deg
350 K

In the simulations, the 60-degree sector mesh


(76x30x30 cells in the squish region + 45x30x20 cells
in the bowl region) has been employed. Injection of
diesel oil has been started at 0-5.8 deg ATDC with
the duration of 8.4 deg. The injection velocity has
been specified using the tabular data taken from experiments. The injection umbrella made an angle
153 deg, the total mass injected amounted for 62.2
mg per stroke, 13.3 mg per injector (6 injectors).
Qualitatively, the earlier predicted features of diesel
spray combustion (see e.g., [14]) were found in accordance with in-cylinder pressure measurements,
RoHR (rate of heat release) estimations and endoscopic video data.

From experimental observation follows that the


heat release rate in the course of spray combustion
development (as calculated from the rate of pressure
rise) has two distinct phases. The first one is characterized by a high rate of heat release (rapid pressure growth) while the second phase is more gradual.
The first stage is called premixed (although it is not
close to be well mixed) and is the result of combined
effect of auto-ignition and flame propagation. The
Volvo D12C Engine Modeling
(Retarded injection)
100

Incylinder pressure, bar

80

Figure 3: Predicted temperature distributions on a


slice of the a sector mesh at 7.5 Ca; SOI takes place
at TDC.

60

40

Calculation, SOI=0. Ca
Experimental data

20

0
30

20

10

10
20
Ca ATDC

30

40

50

60

(a)
Volvo D12C Engine Modeling
(Retarded injection)
300

250

SOI=0.0 Ca
SOI=3.0 Ca
SOI=6.0 Ca

RoHR, J/Ca

200

150

100

50

0
10

10

20

30

such conditions, the ignition can occur at several regions simultaneously and then rapidly spreads over
the chamber volume similar to the HCCI (homogeneous charge compression ignition) or MK (modulated kinetics) regime. In this case the entire heat release occurs in the premixed mode, although again
it is apparent that the mixture is not well premixed.
This regime of diesel spray combustion is reproduced
in the modeling as illustrated by Figs 2 a-b) by predicted in-cylinder pressure and RoHR vs Ca histories
and temperature distributions in Fig. 3. The temperature plot illustrates that the combustion zone is
more distributed compared with the classic diffusion
flame structure. From the plot one can also conclude
that with the help of the new collision model a realistic spray pattern has been predicted. The corresponding in-cylinder pressure vs time histories for

40

Volvo D12C Engine Modeling

Ca ATDC

(Retarded injection)
100

(b) ca=7.5 ATDC

Figure 2: a) Predicted and measured (courtesy of


T.Rente) in-cylinder pressure and b) predicted RoHR
vs Ca histories at different SOI illustrating MK combustion under retarded injections.

Pressure, atm

80

60

40
SOI=0.0 Ca
SOI=3.0 Ca
SOI=6.0 Ca

20

later stage is referred to as the mixing-controlled


one since it develops after the fuel spray is fully surrounded by flame and combustion is the transportcontrolled process. If fuel injection takes place after
the TDC, the ignition starts after an extended period of cold flow development under the condition of
dropping pressure and temperature. As observed at

0
10

10

20

30

40

Ca ATDC

Figure 4: Calculated n-cylinder pressure vs time histories at different SOI illustrating MK combustion.

different SOIs (start of injection) are presented in


Fig. 4. At the engine speed 1700 rpm, the agreement between predictions and experimental data
was found quite reasonable. However, the onset of
the feeble combustion regime has been predicted
more close to TDC than it took place in the current
experiments.
CONCLUSIONS
The paper summarizes the results of the development and validation of a new integrated computer
model (based on the KIVA-3V code) for compression ignited combustion simulation illustrated by the
model application to diesel engine analysis based on
a fuel surrogate model as a substitute of practical
hydrocarbon fuel kinetics. The improved approach
includes new and modified models for spray atomization, droplet collision, and turbulence/chemistry
interaction.
Using the 3-D sector mesh, more realistic predictions of the diesel spray combustion has been
achieved for the Volvo D12C DI engine. Thus, the
transition of the main diesel combustion mode to the
MK mode in the case of retarded fuel injection has
been reproduced in a reasonable agreement with experimental data.
ACKNOWLEDGMENTS
This work was supported by the Chalmers Combustion Engine Research Center, CERC.

References
[1] Veynante, D., Vervisch, L., Turbulence Combustion Modeling, Progress in Energy and Combustion Science, 28: 193-266 (2002)
[2] Magnussen, B.F., Hjertager, B.H., On the Numerical Modeling of Turbulent Combustion
with Special Emphasis on Soot Formation and
Combustion, Sixteenth Symposium (International) on Combustion, Pittsburgh: The Combustion Institute, p. 719-729 (1976)
[3] Kong, S.F., Marriot, C.D.,Reitz, R.D., and Christensen, M., Modeling and Experiments of HCCI
Engine Combustion Using Detailed Chemical
Kinetics with Multidimensional CFD, SAE Paper 2001-01-1026 (2001)
[4] Fluent-6:http://www.fluent.com/software/fluent/
focus/reacting.htm (2002)
[5] Aubry, C., and Villermaux, J., J. Chemical Engineering Science, 30, p. 457 (1975)

[6] Golovitchev, V.I., Nordin,N., Jarnicki,R., and


Chomiak, J., 3-D Diesel Spray Simulations Using a New Detailed Chemistry Turbulent Combustion Model, SAE Paper 2000-01-1891 (2000)
[7] Golovitchev, V.I., Nordin, N., Detailed Chemistry Sub-Grid Scale Model of Turbulent Spray
Combustion for the KIVA Code, Paper 99-ICE237, ICE-Vol. 33-3: p.17-25 (1999)
[8] M AGNUSSEN, B.F., Modeling of Reaction Processes in Turbulent Flames with Special Emphasis on Soot Formation and Combustion, in
Particulate Carbon Formation during Combustion, ed. by D.C. Siegla and G.W. Smith, pp.
321- 341, Plenum Publishing Co. (1981)
[9] Frisch, U., Turbulence, the Legacy of A.N. Kolmogorov, Cambridge University Press (1995)
[10] Sreenivasan, K.R., and Meneveau, C., The
fractal facet of turbulence, J. Fluid Mech.: vol.
173, pp. 357-386 (1986)
[11] Chin, J.S., Calculation Method for Multi- Component Stagnant Droplet Evaporation with Finite Diffusion, ASME-94-GT-440 (1994)
[12] Lutz, A.E., Kee, R.J., and Miller, J.A., SENKIN:
A Fortran Chemical Program for Predicting Homogeneous Gas Phase Chemical Kinetics with
Sensitivity Analysis, Sandia Report SAND878248, UC-401 (September 1994)
[13] Amsden A.A., KIVA-3v: A Block-structured
KIVA Program for Engines with Vertical or
Canted Valves, LA-13313-MS, July (1997)
[14] Golovitchev, V.I., Atarashiya, K., Tanaka, K.
and Yamada, S., Towards Universal EDCBased Combustion Model for Compression Ignited Engine Simulations, SAE Paper 2003-011849 (2003)
[15] Golovitchev, V.I., http://www.tfd.chalmers.se
/ valeri/MECH (1998)
[16] Tanner, F.X., and Weissler, G., Simulation of Liquid Jet Atomization for Fuel Sprays by Means
of a Cascade Drop Breakup Model, SAE Paper
980808 (1998)
[17] Schmidt, D.P., and Rutland, C.J., A New
Droplet Collision Algorithm. Journal of Computational Physics, 164: 62-80 (2000)
[18] Nordin, N., Complex Chemistry Modeling of
Diesel Spray Combustion, Ph.D. Thesis, Department of Thermal and Fluid Dynamics,
Chalmers University of Technology (2001)

Вам также может понравиться