Sie sind auf Seite 1von 12

SAE TECHNICAL

PAPER SERIES 2000-01-0662

3-D CFD Analysis of the Combustion Process in a


DI Diesel Engine using a Flamelet Model
H. Bensler, F. Bühren and E. Samson
Volkswagen AG

L. Vervisch
INSA de Rouen & LMFN – CORIA / CNRS

Reprinted From: Multi-Dimensional Engine Modeling


(SP–1512)

SAE 2000 World Congress


Detroit, Michigan
March 6–9, 2000

400 Commonwealth Drive, Warrendale, PA 15096-0001 U.S.A. Tel: (724) 776-4841 Fax: (724) 776-5760
The appearance of this ISSN code at the bottom of this page indicates SAE’s consent that copies of the
paper may be made for personal or internal use of specific clients. This consent is given on the condition,
however, that the copier pay a $7.00 per article copy fee through the Copyright Clearance Center, Inc.
Operations Center, 222 Rosewood Drive, Danvers, MA 01923 for copying beyond that permitted by Sec-
tions 107 or 108 of the U.S. Copyright Law. This consent does not extend to other kinds of copying such as
copying for general distribution, for advertising or promotional purposes, for creating new collective works,
or for resale.

SAE routinely stocks printed papers for a period of three years following date of publication. Direct your
orders to SAE Customer Sales and Satisfaction Department.

Quantity reprint rates can be obtained from the Customer Sales and Satisfaction Department.

To request permission to reprint a technical paper or permission to use copyrighted SAE publications in
other works, contact the SAE Publications Group.

All SAE papers, standards, and selected


books are abstracted and indexed in the
Global Mobility Database

No part of this publication may be reproduced in any form, in an electronic retrieval system or otherwise, without the prior written
permission of the publisher.

ISSN 0148-7191
Copyright © 2000 Society of Automotive Engineers, Inc.

Positions and opinions advanced in this paper are those of the author(s) and not necessarily those of SAE. The author is solely
responsible for the content of the paper. A process is available by which discussions will be printed with the paper if it is published in
SAE Transactions. For permission to publish this paper in full or in part, contact the SAE Publications Group.

Persons wishing to submit papers to be considered for presentation or publication through SAE should send the manuscript or a 300
word abstract of a proposed manuscript to: Secretary, Engineering Meetings Board, SAE.

Printed in USA
2000-01-0662

3-D CFD Analysis of the Combustion Process in a DI Diesel


Engine using a Flamelet Model
H. Bensler, F. Bühren and E. Samson
Volkswagen AG

L. Vervisch
INSA de Rouen & LMFN – CORIA / CNRS

Copyright © 2000 Society of Automotive Engineers, Inc.

ABSTRACT combustion process itself. In direct-injection (DI) Diesel


engines the complexity of the CFD simulation is even
A 3-dimensional numerical study has been conducted higher through the presence of partially mixed
investigating the combustion process in a VW 1.9l TDI combustion.
Diesel engine. Simulations were performed modeling the In this study, the Diesel combustion process was
spray injection of a 5-hole Diesel injector in a pressure investigated both numerically and experimentally in a VW
chamber. A graphical methodology was utilized to match 1,9l TDI engine. The following sections will describe the
the spray resulting from the widely used Discrete Droplet experimental and numerical setups utilized. A brief
Spray model to pressure chamber spray images. review of turbulent combustion modeling will be
Satisfactory agreement has been obtained regarding the presented along with the theory behind the new RTZF
simulated and experimental spray penetration and cone combustion model used for the simulations. Finally, the
angles. Thereafter, the combustion process in the engine results will be discussed and improvements to the
was simulated. Using engine measurements to initialize combustion model will be proposed.
the combustion chamber conditions, the compression
stroke, the spray injection and the combustion simulation
COMBUSTION MODELING THEORY
was performed. The novel RTZF two-zone flamelet
combustion model was used for the combustion
Although non-premixed combustion is not the most
simulation and was tested for partial load operating
efficient way to burn, it is the most common regime found
conditions. An objective analysis of the model is
in internal combustion engines. The modeling of non-
presented including the results of a numerical parameter
premixed (and partially premixed) turbulent combustion
study. A comparison of numerical and experimental
requires expressing the mean reaction rate, as a function
engine results was found to be encouraging and show
of chemistry and flow. As it is impossible to directly
potential for the further development and implementation
average the reaction rate due to its highly non-linearity, it
of the combustion model.
is necessary to propose other approaches to express the
mean reaction rate ω& .
INTRODUCTION
The different tools used to model turbulent combustion
Due to increasing market demands, there is a great need can be listed as follows:
to design and develop cleaner and optimized Diesel
• the mean values method, where the mean reaction
engines in ever shorter time periods. Through the use of
rate is calculated using the mean values of different
3-dimensional Computational Fluid Dynamics (CFD)
properties [1]
simulations, the design process can be greatly speeded
up with minimal costs. The amount of manufacturing and • the geometric description method, where the flame is
testing of prototypes can be reduced through the pre- described as a geometrical entity [2]
optimization of engine processes via numerical • the scalar dissipation method, where the reaction
simulations. Of these engine processes, the 3-D rate is calculated using the dissipation rate of a
simulation of the combustion process has one of the scalar field [3]
highest levels of complexity. Here all the various aspects • the statistical description method, which uses a
and parameters of the combustion process need to be stochastic approach to determine ω& [4,5,6].
modeled including the proper flow field (main flow and it’s
associated turbulence), the spray injection and the

1
Some assumptions need to be made to simplify concentrations and thermal properties are calculated for
chemistry and transport. Depending on the conditions, a given mixture composition, pressure and temperature.
several hypotheses can be made: The mixture compositions consist of the mass fractions of
air, unburned fuel, burned air and burned fuel. There are
• infinitely fast chemistry
11 chemical species making up the combustion products:
• finite rate chemistry, where diffusion and reaction are CO , CO 2 , H , H 2 , H 2 O , N 2 , NO , O , O 2 , OH and N .
coupled as in laminar flames (the laminar flamelet These represent the major chemical species in both the
assumption) complete and incomplete combustion products. These
• finite rate chemistry, where diffusion and reaction are chemical species concentrations are then used for the
treated separately, as in probability density function determination of the emissions, which are calculated
(pdf) methods. separately. Their formation reactions have much slower
rates and make a small contribution to the total energy
Modeling combustion in DI diesel engines requires to
balance.
take into account the possibility to burn under different
regimes: premixed, non-premixed and partially premixed,
SPECIES SCHEME – The computational cells are
due to the inhomogeneous mixing of the reactants.
notionally divided by the flame front into two zones: an
Models have to represent those phenomena accurately.
unburned zone and a burned zone. Air, fuel vapor and
This approach is the one of pdf-generator modeling [5].
residual gases are found in the unburned zone and
Other ways can be followed, using a more empirical
combustion products are in the burned zone. The
formulation such as in the mean values and geometrical
unburned region is further divided into two regions: a
description methods.
segregated and a fully mixed region. In the segregated
region, the air and fuel are not mixed at a molecular level
RTZF COMBUSTION MODEL and are therefore not ready to react chemically. In the
other region, the air, fuel and residuals are fully mixed.
A combination of these two methods is the novel Ricardo
Two Zone Flamelet (RTZF) combustion model [7] The segregated region increases in mass from newly
investigated by this study. In this model, a semi-empirical evaporated fuel from a fuel spray or from the flow from an
approach is adopted where the turbulent burning velocity inlet boundary. The fluid flow and molecular diffusion
and the burning rate are estimated based on fractal convert segregated reactants to fully mixed reactants;
geometry and basic dimensional analysis of turbulent combustion then consumes the fully mixed reactants and
flames. In this way, difficulties such as the description of converts them into combustion products. The mixture
the flame front area can be avoided. In addition, there is a components in the unburned zone have effectively three
simple and clear relationship between the turbulence and mass fractions: a segregated, a fully mixed and an overall
burning velocity, which can be easily compared with mass fraction; only two of these are however
measurements. A fractal-based model provides limited independent. The segregated and overall mass fractions
physical insight, but it quickly provides the burning rate, are solved for and the fully mixed mass fraction is
which is a major requirement for combustion modeling. obtained from the other two.
The proper description of the effects of chemistry is an The species concentrations of the combustion products
essential of combustion modeling. A large amount of are determined from the initial air/fuel ratio before
computing resources is required for the simulation of full combustion. Thus the mass fractions of both the burned
chemical reaction schemes; however, as an engineering air and fuel are used rather than the total mass fraction of
tool, this approach is too costly and complex. On the the combustible products in the burned zone. In a similar
other hand, simplified reduced reaction schemes have manner, the separate mass fractions of the residual air
been developed for a few particular fuel types. These are, and fuel are also used.
however, lacking in generality and can be over simplified.
An example of this is the previously mentioned
Magnussen model [1], which uses an over-simplified
scheme, based on a one-step reaction of burning at a
stoichiometric air/fuel ratio. Important effects such as the
incomplete combustion of fuel-rich mixture burning are
ignored with this method. Moreover, information
concerning the chemical composition of the combustion
products is also not provided thus emissions formation
cannot be determined.
The current model uses a chemical equilibrium approach
for the chemistry modeling. Since the reactions, which
are responsible for the heat release, are always
performed at fast rates, they can therefore be treated at Figure 1. Schematic of the RTZF Combustion model.
equilibrium conditions. The chemical species

2
MIXING AND FLAME PROPAGATION – The different where c is the regress variable and is defined as the
components in the segregated region are converted to sum of mass fractions of unburned air, unburned fuel and
the fully mixed mass fraction by turbulent mixing. The residuals ( c = Y A + YF + YRA + YRF ). Before the start of
turbulent mixing is based on the large-eddy scale and
combustion, the regress variable for a given cell is equal
has a mixing rate of:
to 1; it has a value of 0 after the completion of
k combustion. The characteristic cell length is evaluated by
ω αmix = c mix ρYαseg
ε (1) the cubic root of the cell volume ( Lc = 3 Vc ) and the
coefficient g has an approximate value of 4.
where c mix is the mixing constant ranging from a value of
2 to 32 (a value of 20 is the recommended default value For the case of non-premixed flames, the ratio of the fully
mixed reactant volume to the total volume ( rvol ) is
for c mix ). estimated for the degree of segregation. The segregated
Reactants, which are fully mixed, are consumed by the air/fuel mixture causes holes in the flame front. Thus the
combustion. Low temperature reactions occur for auto- non-premixed burning rate is slower than a pre-mixed
ignition and high-temperature reactions occur for the rate.
flame propagation. High-temperature reactions always
follow after auto-ignition. Thereafter, the combustibles AUTO-IGNITION MODEL – Since low temperature
can be consumed by either more auto-ignition or by reactions are responsible for auto-ignition and high-
normal flame propagation. For the normal flame temperature reactions for the normal flame, these two
propagation, the burning rate is determined by: reactions are treated separately. Low-temperature
reactions for the auto-ignition are considered as a
AT lumped one-step reaction of a generic intermediate
ω f = ρr SL I Σ rvol
A ignition species Yig , which is inversely proportional to the
(2)
where the effective unstretched laminar burning velocity ignition time ( Yig = 1 / τ ig ). The ignition probability, p ig , is
S L is derived from measured maximum laminar burning determined via the integral:
velocity where the effects of the reference condition, fuel-
dt
air equivalence ratio and residuals are taken into p ig = ∫
account. τ ig
(6)
The effect of strain on flame burning velocity is
represented by the flame stretch factor I . With The ignition delay is a function of temperature and
increasing strain, the propagating flame will be partially pressure given by:
quenched and can even be extinguished. For the
turbulent flame, the ratio of flame surface area to the τ ig = AP − n e (T A / T )
(7)
projected transverse area ( AT / A ) is determined from
fractal geometry: where the coefficients A , n and T A (activation
temperature) are determined from experimental ignition
D −2 delay measurements. A transport equation for the ignition
AT l 
=  I 
 probability is solved and auto-ignition occurs when p ig
A  lk  has reached a value of unity. The contribution of heat
(3)
from the low-temperature reactions is considered
Here l I is the integral length scale and l k is the negligible.
Kolmogorov scale. When considering the transition from
laminar to turbulent regime, the fractal dimension D is COMMENTS – The mixing rate is built in the same way
correlated as a function of the ratio of turbulence intensity as the Eddy Break-Up model [8], using a linear relaxation
to laminar burning velocity: model. Only the effect of turbulence is taken into account.
Reaction takes place in a thin zone between burned and
  ′ 
 − 0,5 / ln  1+ u  unburned gas, assimilated to a set of laminar flames. Its
  S 
  
D = 2 + 0,75e L formulation seems to be a mix between the BML model
(4)
modified by Bray et al. [9] and the approach of Gouldin et
The flame projected transverse area per unit volume al. [10]. The BML model supposes that a point of the flow
( Σ = A / V ) is a shape function, and is defined as is either in the burned or in the unburned zones,
separated by the reaction area. The reaction rate is then
gc (1 − c ) proportional to the pass frequency of the flame front.
Σ=
Lc
(5)

3
Those models are formulated in terms of flame density. TDI ENGINE – The experimental engine used was a VW
The first one, with reference of equation (5) for the TDI 1.9l direct injection Diesel engine. The engine
formulation of the flame density, makes the flame density characteristics are found in Table 1 below.
to be proportional to the crossing frequency of the flame
front. In the other one, the flame density is expressed as Table 1. Engine Characteristics.
a fractal surface and gives an expression quite analog to
(3). TDI 1,9 liter
The RTZF model uses the approach of Bray et al. [9] for Type 4 in-line-cylinder
the reaction rate, the fractal formulation being used for Injection Direct
the calculation of the surface area of the flame. The
model enables to simulate both non-premixed and Arrangement Row
premixed combustion. For non-premixed combustion, Timing 1 overhead camshaft (belt)
reactants are first mixed locally, in each cell, and are then
burned. For pre-mixed combustion, only the second step Valves 2 per cylinder
is required. For more RTZF model details see [7]. Capacity 1896cm 3

EXPERIMENTATION Bore x stroke 79 ,5 mm × 95 ,5 mm

Compression ratio 17,8


There were two experiments involved for the validation of
the numerical results: a pressure chamber for the spray Connecting Rod 144 mm
modeling and a TDI Diesel engine. Their respective
Turbocharger Garrett VNT15
setups will be discussed in the sections below.
Engine speed 1000 rpm
PRESSURE CHAMBER – The pressure chamber is a
cylindrical chamber equipped with small windows, which
enable to take pictures of the inside-phenomena. Fig. 2 The engine was mounted on a standard VW test bed.
shows a schematic of the pressure chamber. A light Important engine parameters were measured including
source is used to illuminate the spray and a digital video the injected mass of the fuel, the cylinder pressure, and
camera then records the images. A high-pressure emissions such as NOx, Soot, Hydrocarbons, CO , CO 2
compressor and a heat exchanger allow the pressure and O 2 . Air and exhaust temperatures were measured
and temperature in the chamber to be adjusted. In this using thermal resistance sensors at 8 different locations
study, the chamber pressure was set to 20 bar with a ranging from the air filter to the exhaust pipe. For the
constant temperature of 293K. The Diesel fuel was numerical simulations conducted in this study, the
injected with a 5-hole injector and had a temperature of measured intake manifold temperature was of prime
303K and an injection pressure of 400 bar. Pictures were importance.
taken through the bottom window of the chamber each
50 µs for a duration of 2 ms. In order to determine the time-dependent fuel injection
rate for a given injector, measurements were made
F uel-
separately on a hydraulic injector test bed. Here the fuel
H eat - H igh-P r es s ure-
pres s ure
ex chan ger com pres s or is injected into a tube filled with the same fuel. Time
dependent pressure waves are measured, whose integral
P ow er U nit DU T
L ight L ight D at a
is proportional to the injected mass of fuel. In this
S our ce 60°

40°
60°

40°
S our ce S av in g manner, an accurate injection law can be determined for
20° 20°
any type of injector. This methodology provides an
K-EFAM
VOLKSWAGEN
E lect rical
accurate description of the time dependent injected fuel
10° 10°
D at a mass flow rate law.
A quis it ion
Calculat ion
Engine measurements for this study were made for part
O pt rical
S ignal-
con t roller
K am era
D at a
load operating conditions with an engine speed of 1000
A quis it ion
RPM, a Pmep of 1 bar and with 6 mg of injected fuel. The
fuel was injected with a pressure of 400 bar in two
Figure 2. Schematic of pressure chamber.
phases: a pre- and a main-injection. The mass of the
injected fuel equaled 1.6 mg and 4.4 mg, respectively.
The corresponding injection law can be seen in Fig. 3.

4
DISCUSSION AND RESULTS

PRESSURE CHAMBER – The two main parameters


considered for the correlation of the simulated and
photographed spray images were the spray cone angle
and the penetration. Several calculations were needed to
get an acceptable match with the pressure chamber
images. The procedure was to begin with an estimated
value of the cone angle of 8 degrees resulting from
correlations existing in the literature [14]. The simulation
was conducted and the width of the spray angle and the
penetration were visually compared to the experimental
images. If a poor correlation was obtained, an iterative
process was used either increasing or decreasing the old
value of the cone angle. Once a suitable cone angle was
obtained, the other parameters such as droplet size
Figure 3. Measured injection rate. distribution and droplet interaction models were tested.
Results showed that a cone angle of 10 degrees
NUMERICAL ANALYSIS
corresponded well with the experimental images. An
example of the computational mesh and the
The 3-dimensional simulations were performed using the
corresponding spray are seen in Fig. 4. Each of the 5
finite-volume CFD code VECTIS [11]. The CFD solver
holes was modeled identically which resulted in 5 very
provides 3-D time-dependent, compressible or
similar jets. Note that the mesh in the area near the
incompressible solution of the continuity, Navier-Stokes,
injector is twice as fine as in the area of the spray jets.
and energy equations. The turbulence model used is the
This was done to increase the numerical accuracy of the
standard k − ε model [12] based on the assumption of
spray modeling. The mesh density used for the pressure
local isotropic turbulence at high Reynolds numbers.
chamber simulations was also used later for modeling the
Equations are solved fully-implicitly with coupling
spray and combustion in the TDI engine. It should be
between variables and non-linear effects using iterative
further noted that the only the main injection described in
or predictor-corrector methods. An orthogonal structured
Fig. 3 was modeled in the pressure chamber since no
Cartesian mesh is used with wall cell volume and face
photographs were available for the pre-injection.
area adaptation. The computational meshes are
generated using a fully automatic mesh generator. Local
mesh refinement is utilized for the resolution of small-
scale flow structures near boundaries and in regions of
high gradients. The standard logarithmic “law of the wall”
is used for the evaluation of the wall boundary layer.

SPRAY MODELLING – The fuel injection simulation is


done using a Lagrangian stochastic spray model of
discrete droplets [7]. The spray is modelled as a
dispersed liquid phase which interacts and penetrates
the surrounding continuous gas phase. The spray is
described by an ensemble of discrete droplets parcels,
where each parcel contains a quantity of droplets with the
identical velocity, temperature and size. Both the droplet
parcels and the gas phase interact with each other
through drag forces and heat and mass transfer.
Differential equations for the droplet mass, momentum
and temperature are solved. In addition, other spray
phenomena such as droplet-coalescence, -turbulence
interaction, -wall interaction and –breakup are also
modelled. The droplet-breakup model used for this study
is that of Patterson and Reitz [13].
Figure 4. Pressure chamber computational mesh and
spray simulation.

5
Figure 5. Comparison of simulated and pressure chamber spray penetration: Pchamber = 20 bar, Tchamber = 293 K,
Prail=400bar and Trail=303K.

A comparison of the simulated and photographed spray A view of the computational mesh used for the engine
results can be seen in Fig. 5. A total of six pairs of images simulations can be seen in Fig. 6. Note that in the area of
are presented showing the cone angle and penetration the injector, local mesh refinement is used to improve the
for one representative jet of the spray. Both the simulated accuracy of the spray modeling. The cell size in the bowl
and photographic images were directly comparable was 1mm square that resulted in an average mesh size
through the use of a 10-mm square reference grid. A of 100000 internal cells. The calculation time on a single-
spray image is shown every 250 microseconds. The lines processor workstation which included the compression,
at the tip and sides of the sprays had identical spray and combustion simulation equaled about 40
dimensions for both the simulated and photographic hours.
images. One can see that although the slightly non-
symmetric nature of the photographed spray was not
found in simulated spray, general good agreement was
found with regards to both the spray cone angle and
penetration.

TDI ENGINE – The engine numerical simulations were


done from –40° CA BTDC until 40° ATDC. The Figure 6. Cross-sectional view of the computational
calculations were initialized with measured swirl ratio of mesh at 10°CA ATDC.
2.2 from another study done on a transparent VW TDI
engine [15]. The initial conditions in the combustion were A comparison of the simulated and measured in-cylinder
taken from engine measurements where the pressure pressure can be seen in Fig. 7 based on the default
and temperature equaled 9.2 bar and 735K, respectively. values of the combustion model constants. The
The species in the chamber were initialized to take into measured pressure curve has a maximum value of about
account the 30% EGR set up on the experimental 43 bar. Thereafter, two pressure peaks are found at 5 and
engine. The injection parameters used for the simulations 17° ATDC, which indicate combustion occurring after the
were the same as those for the engine measurements pre- and main injections. When considering the simulated
mentioned in the Experimentation section. pressure curve, very good agreement is observed from –

6
40 to –5° CA BTDC. The simulations also show two EGR level of 30%, the chemical composition of the
peaks smaller in magnitude, however both beginning very chamber gases would be in effect different compared to
close to the measured start of combustion. the unfired engine.
The simulation results based on the default values of the
combustion model constants already showed an
encouraging correlation compared to the measured
pressure curve. However, in order to further test the
model, a parameter study was conducted by changing
the mixing coefficient (Cmix), the Reaction Scaling Factor
(RSF) and the Ignition Scaling Factor (ISF). The mixing
coefficient is a real constant in the model and directly
controls the rate of mixing of the species, i.e. their
conversion from a segregated to a mixed condition. In
addition, Cmix has a strong influence on the overall
combustion reaction rate. The other two parameters
investigated are used for fine-tuning of the model. The
RSF controls the burning velocity and thus the overall
reaction rate. Finally, the ISF tunes the auto-ignition pdf,
effecting the rate of the low-temperature reactions that
lead to auto ignition. The default values for the
Figure 7. Comparison of simulated and measured in- parameters are Cmix=20, RSF=1, and ISF=1.
cylinder pressure using the default values of
the combustion model constants In order to investigate the effects of varying Cmix, three
(Cmix=20, RSF=1, ISF=1). simulations were performed with Cmix set at 10, 20 and
32; the values of RSF and ISF were left at their default
values of 1. Fig. 9 shows the percentage difference
between the simulated and the measured cylinder
pressure curves as a function of °CA. It can be seen that
combustion begins at about 3° CA ATDC for all the three
Cmix values. For Cmix=10, the pressure is always more
under-predicted than for the case of the other two Cmix
values and has a mean difference of –3.0 %. When
Cmix=32, the mean difference equals –2.2% and has a
higher value of the pressure over the entire 25°CA. The
default value of Cmix=20 is very similar to Cmix=32 with
a mean difference of -2.5%. Thus the higher the value of
Cmix, the smaller was the mean difference between the
simulated and measured pressure.

Figure 8. Measured and simulated in-cylinder pressure


with and without combustion with default
values of the combustion model constants
(zoom of Fig.7).

The area of combustion in Fig. 7 from 0 to 20 °CA is


shown in more detail in Fig. 8. Here it can be seen that
there is a general under-prediction of the calculated
pressure curve except in a region between 10° and 12°
CA ATDC. Moreover, one can see that there is a
difference of about 0.4 bar (approximately 1%) between
the measured pressure curves with and without
combustion where the unfired pressure curve is lower in
magnitude. An explanation for this discrepancy is that the Figure 9. Percent difference between simulated and
combustion chamber gas temperature and pressure measured combustion chamber pressure for
could be influenced by wall heating effects since a unfired varying mixing coefficient as a function of
engine would have lower wall temperatures than a fired crank angle (RSF=1, ISF=1).
engine. In addition, since the fired engine was run with an

7
The variation of RSF is shown in Fig. 10. As the RSF In Fig. 12, it is shown that the percent difference in the
value decreased, the percent difference increased from a pressure for the different ISF values becomes relatively
mean difference of –2.5% at RSF=1 to –3.1% at similar with increasing °CA where the mean difference
RSF=0.5. Once again the point of combustion begin at ranges from –2.5 to -2.7%.
5°CA ATDC was not affected. Thus the default value of
RSF=1 showed the better fit to the experimental pressure
curve.

Figure 12. Percent difference between simulated and


measured combustion chamber pressure for
varying ignition scaling factor as a function of
Figure 10. Percent difference between simulated and crank angle (CMIX=20, RSF=1).
measured combustion chamber pressure for
varying reaction scaling factor as a function of In order to better understand the RTZF model and to
crank angle (CMIX=20, ISF=1). provide ways in which the model could be improved, data
from the literature were analyzed. Direct Numerical
As expected, by changing the ISF, the point of ignition Simulation (DNS) studies [16,17,18] have reported that
begin was influenced. The ISF was simulated at three the time evolution of the mixture during auto-ignition
different values: 0.5, 0.75 and 1.0. In Fig. 11, it can be strongly depends on the time evolution of the mixture
clearly seen that the time of ignition begin occurs fraction dissipation rate itself. The mixture fraction
progressively later as ISF is decreased; moreover, the dissipation rate is one of the key quantities of non-
magnitude of the initial pressure peak also becomes premixed turbulent combustion, where the extent of
progressively smaller. After the initial point of pre- mixing between evaporated fuel and oxidiser is usually
injection combustion, the effect of changing ISF becomes measured using a mixture fraction [19]. The control
smaller with increasing °CA. parameters of a reaction-diffusion layer are: a diffusive
time measured from the diffusive flux of mixture fraction,
used to define the inverse of the scalar dissipation rate,
and a chemical time. Both are retained to build a
Damkohler number useful for characterising various
combustion regimes [20].
DNS results suggest that one important point to
reproduce auto-ignition is to capture the time history of
the mean mixture fraction dissipation rate evaluated
under stoichiometric conditions. Practically speaking,
when all the local characteristic micro-mixing times
(inverse of the scalar dissipation rate) are much smaller
than the local ignition times of the stoichiometric mixture
(or of the most reactive mixture), ignition is more or less
uniformly distributed. Accordingly, when local turbulent
micro-mixing leads to possibilities of non-uniform
ignition, edge-flame combustion involving partially
premixed flames will control the development of the heat
Figure 11. Measured and simulated in-cylinder pressure release rate.
with and without combustion as a function of
crank angle for varying ISF values
(CMIX=20, RSF=1).

8
In the light of these DNS observations, one possibility to REFERENCES
improve the combustion model, and to mimic these
important features of auto-ignition, would consist of 1. Magnussen B. and Mjertager B., “On mathematical
modifying the calculation of the probability of finding modelling of turbulent combustion”, 16th International
ignition. Following the work of Ravet and Vervisch [5], Symposium on Combustion, pp. 719-727, The
based on an idea first proposed by Borghi [4], a table Combustion Institute, 1976.
look up of ignition time delay could be constructed. The 2. Marble F. and Broadwell J., “The coherent flame
resulting ignition times would be a function of the mean model of non-premixed turbulent combustion”,
temperature, mean species mass fraction and position in Project Squid TRW-9-PU, Project Squid
mixture fraction space. Introducing the probability density Headquarters, Chaffee Hall, Purdue University, 1977.
distribution of micro-mixing time, the probability of finding 3. Bilger R.W., “The structure of diffusion flames”,
ignition for each possible instantaneous value of the Combustion Science Technology, Vol. 13, pp. 155-
mixture fraction is then measured from the probability of 170, 1976.
having micro-mixing times greater than the ignition time. 4. Borghi R., “Turbulent combustion modelling”, Prog.
A presumed beta-pdf for the mixture fraction could be Energy Combust. Sci., Vol. 14, pp. 245-292, 1988.
used to average over the range of variation of mixture 5. Ravet F. and Vervisch L., “Modelling non-premixed
fraction. The existence of characteristic mixing times turbulent combustion in aeronautical engines using
smaller than ignition delay would prevent auto-ignition pdf-generator”, 36th Aerospace Sciences Meeting
and then combustion. However, when ignition prevails and Exhibit AIAA, paper 98-1027, Reno, Nevada,
over micro-mixing, combustion would occur. This USA, 1998.
modification of the model would lead to a more accurate 6. Pope, S.B., “Pdf method for turbulent reacting flows”,
description of the first stage of combustion in the engine. Prog. Energy Combust. Sci., Vol. 11, pp. 119-195,
Finally, the ignition table look-up could be implemented 1985.
using a detailed chemical mechanism; this would allow 7. VECTIS Theory Manual, Version 3.4, Ricardo
for better accuracy in the determination of the burning Consulting Engineers Ltd, December, 1999.
rate. 8. Spalding, D., “Mixing and chemical reaction in steady
confined turbulent flames”, 13th Symposium
CONCLUSION (International) on Combustion, Pittsburgh, pp. 649-
657, The Combustion Institute, Pittsburgh, 1971.
A numerical and experimental study was conducted 9. Bray, K., Champion, M. and Libby P., “The interaction
investigating the combustion process in a 1.9L TDI Diesel between turbulence and chemistry in premixed
engine. Simulations and measurements were conducted turbulence flames”, Turbulent Reacting Flows,
in both for the case of a pressure chamber and the Volume 40 of Lecture Notes in Engineering, R.
engine. A graphical methodology was used which Borghi and S. Murphy (Eds.), pp. 541-563, Springer,
matched the simulated spray well with the spray images 1989.
taken in the pressure chamber. The novel RTZF flamelet 10. Gouldin, F., Bray, K. and Chen, J., “Chemical closure
model was used for the combustion simulations. A model for fractal flamelets”, Combust. Flame, Vol. 77,
comparison of the calculated and measured in-cylinder pp. 241-259, 1989.
pressure indicated good agreement at partial load 11. VECTIS Users Guide, Computational Fluid
operating conditions with a mean difference between the Dynamics, Release 3.3, Ricardo Consulting
simulation and measurement pressure equalling 3%. Engineers Ltd, January, 1999.
Furthermore, suggestions based on Direct Numerical 12. Launder, B. E. and Spalding D.B, Mathematical
Simulations from the literature have been proposed how Models of Turbulence, Academic Press, New York,
to improve the auto-ignition model utilized. In conclusion, 1972.
this study has shown that the RTZF flamelet model tested 13. Patterson, M.A. and Reitz, R.D, “Modeling the effects
shows encouraging results and thus merits further of fuel spray charactreristics on Diesel eingine
development and investigation at other engine operating combustion and emission”, SAE 980131, 1998.
conditions. 14. Reitz, R.D. and Bracco, F.V., “Mechanisms of
atomization of a liquid jet”, Phys. Fluids, Vol. 25,
ACKNOWLEDGMENTS pp.1730-1742, 1982.
15. Arcoumanis, C., Whitelaw, J.H., Hentschel, W. and
The authors would like to thank Arno Homburg and Schindler, K.P., “Flow and combustion in a
Reinhard Schulz for their help concerning the pressure transparent 1.9 liter direct injection diesel engine”,
chamber measurements and diagrams. Finally, thanks Proce. Inst. Mech. Engrs., Vol. 208, pp.191-205,
are given to Dr. H.-J. Oberg for his useful discussions 1994.
during the course of this study. 16. Vervisch, L. and Poinsot, T., “Direct numerical
simulation of non-premixed turbulent flame”, Annu.
Rev. Fluid Mech., Vol. 30, pp. 655-692, 1998.

9
17. Domingo, P. and Vervisch, L., “Triple flames and
partially premixed combustion in autoignition of non-
premixed turbulent mixtures”, 26th International
Symposium on Combustion, Naples, pp. 233-240,
The Combustion Institute, Pittsburgh, 1996.
18. Mastorakos, E., Baritaud, T. and Poinsot, T.J.,
“Numerical simulations of autoignition in turbulent
mixing flows”, Direct numerical simulation for
turbulent reacting flows, T. Baritaud, T. Poinsot, and
M. Baum (Eds), pp. 242-276, Editions Technip, 1996.
19. Peters, N., “Laminar flamelet concepts in turbulent
combustion”, 21st International Symposium on
Combustion, Irvine, pp. 1231-1250, The Combustion
Institute, Pittsburgh, 1986.
20. Linán, A., “The asymptotic structure of counter flow
diffusion flames for large activation energies”, Acta
Astronautica 1007 (1), 1974.

10

Das könnte Ihnen auch gefallen