Sie sind auf Seite 1von 6

Applied Energy 87 (2010) 14271432

Contents lists available at ScienceDirect

Applied Energy
journal homepage: www.elsevier.com/locate/apenergy

Application of spray combustion simulation in DI diesel engine


I. Dhuchakallaya *, A.P. Watkins
Energy, Environment and Climate Change Research Group, School of Mechanical, Aerospace and Civil Engineering, University of Manchester, United Kingdom

a r t i c l e

i n f o

Article history:
Received 2 June 2009
Received in revised form 29 July 2009
Accepted 22 August 2009
Available online 15 September 2009
Keywords:
Combustion
Chemical equilibrium
Diesel
Modelling
Spray
Simulation

a b s t r a c t
This work presents the development and implementation of combustion model for DI diesel engines by
using the PDF-Chemical Equilibrium combustion model. The key concept of this approach is to predict the
thermochemical variables (e.g., temperature, species mass fractions) and then the average scalars of these
variables are evaluated by a probability density function (PDF) averaging approach. To realize ame propagation, the reaction time scale is employed to relax the innitely fast chemistry of chemical equilibrium.
The PDF-Eddy Break Up ignition model is adopted in the auto-ignition calculation. With regard to the
comparison results, the simulation results are in good agreement with the experimental results in both
ignition and combustion modes. In addition, the predicted lift-off length also corresponds to a power-law
scaling of Siebers et al.
2009 Elsevier Ltd. All rights reserved.

1. Introduction
Nowadays, the direct fuel injection technology has become
more and more popular both in compression ignition and spark
ignition engines. The direct injection (DI) engines are characterised
by a superior fuel economy compared to indirect injection pre- and
swirl-chamber engines, but the vibration and noise control in these
engines are necessary to consider. In the basic concept for spray
combustion, the high-pressure liquid fuel is injected into the combustion chamber occupied by high-pressure and high-temperature
quiescent air. The high pressure of injection disturbs the liquid jet
surface and leads to disintegration of isolated liquid fragments or
ligaments which consequently breakup further into smaller droplets. The disintegrated liquid droplets are then evaporated due to
heat transfer between liquid droplets and ambient air, and subsequently mixed with the hot air. A rich mixture is generally located
at core centre region and the entrained mixing air gradually increases towards the spray periphery. In the suitable ignition conditions including primarily high temperature and concentration of
mixture within the ammability limits, the combustible mixture
then burns rapidly leading to a sharp increase in heat transfer
due to the chemical reaction. In the experimental results, the ignition generally occurs somewhere in the downstream portion of the
fuel vaporised cloud. This is the onset of combustion.
Recently, the development and application of mathematical
models for spray combustion in engines have been increasingly
* Corresponding author.
E-mail address: dhuchakallaya@yahoo.com (I. Dhuchakallaya).
0306-2619/$ - see front matter 2009 Elsevier Ltd. All rights reserved.
doi:10.1016/j.apenergy.2009.08.029

attractive to researchers [13] because these useful techniques offer fairly cheap and reliable results to describe and predict the phenomena of spray combustion. They can provide the output every
single variable of a problem at any position in physical space and
at any time during the process, while the experiments cannot give
the complete set of information due to the following reasons.
Firstly, it is extremely difcult and expensive to apply the sophisticated optical measurement techniques to a rapidly oscillating
combustion engine without affecting the boundary conditions of
spray development and combustion, for example, in dense spray
closed to injector where the breakup and atomization of the spray
mostly occur, in spray head where most soot is located. Furthermore, the experiments generally provide information limited to
only two dimensions and hardly yield three-dimensional results.
Another benet of mathematical models is the articial parameters
that are intentionally designed for investigating the effects of simulated factors on spray combustion processes. These features are
important in order to obtain more detailed information about
spray combustion phenomena.
The principle objective of the present work is to develop a simulation model for a spray combustion based on the spray number
size distribution moments introduced by Beck [4]. These moments
employed in governing equations for spray are based on the moment-average quantities. In the crucial purpose of the present
work, the mathematical models adopted for auto-ignition is the
PDF-Eddy Break Up (EBU) ignition model [5]. The key concept of
this ignition approach is to combine the chemical reaction rate
dealing with low-temperature mode, and the turbulence reaction
rate governing the high-temperature part by a reaction progress

1428

I. Dhuchakallaya, A.P. Watkins / Applied Energy 87 (2010) 14271432

variable coupling function which represents the level of reaction.


The average reaction rate here is evaluated by a probability density
function (PDF) averaging approach. The PDF-EBU ignition model
yields the ignition delay time in good agreement with the Shell
ignition model prediction. However, the ignition kernel location
predicted by the Shell model is slightly nearer injector than that
by the PDF-EBU model leading to shorter lift-off length. For combustion mode, the PDF-Chemical Equilibrium (EQ) combustion
model is introduced. The details of this model will be discussed
later on.

The global single-step reaction describing the combustion can be


expressed as

2. Mathematical modelling

Applying the CHNO atom balances, thus

Based on work of Burke and Schumann [6], the fuel and oxidant
species never coexist in space and chemical reaction also convert
completely to nal products with no reaction rate or chemical
equilibrium information required. Hence, it cannot predict intermediate species formation or dissociation effects and ame temperature results are often over-predicted. The multi-step reaction
models generally reduced from the detailed mechanism which involves thousands of chemical reactions and hundreds of species
seem to the reasonable solution. These models have been employed successfully in engine applications, for examples, the MIT
model [7], Schreibers model [8], and Jones and Lindsted [9]. However, these schemes have a limitation to describe the combustion
mechanisms of higher hydrocarbons such as diesel due to the complexity of their mechanisms and expensive computational cost. In
order to depict the combustion species, the chemical equilibrium
model is introduced due to its inexpensiveness and effectiveness.
In this approach, the chemistry is assumed to be rapid enough
for chemical equilibrium. The equilibrium mass fraction of each
species related to mixture fraction is calculated. Then it can predict
information of intermediate species without knowledge of detailed
chemical kinetics.
Hence, the equilibrium state here refers to the chemical equilibrium condition with consideration of the effect of dissociation of
major species. To simplify, the basic species including
CO2 ; CO;H2 O;H2 ; H; OH;O2 ; O; NO; N; and N2 are assumed to occur
in products. Thus, the relevant chemical equilibrium reactions are

Cn Hm aO2 3:76N2 ! x1 O2 x2 N2 x3 CO2 x4 H2 O


x5 H x6 H2 x7 N x8 NO x9 O
x10 OH x11 CO
where a is related to the equivalence ratio U as

n m=4

C; n x3 x11
H; m 2x4 x5 2x6 x10

N; 7:52a 2x2 x7 x8
O; 2a 2x1 2x3 x4 x8 x9 x10 x11

The reaction equilibrium constants Kp of each reaction are obtained


from;

p
p
p
y5 p 0
y9 p0
y7 p0
y y2 p0
Kp2 p
Kp3 p
Kp4 6 21
Kp1 p

y6
y1
y2
y4
p
pp
y10 y6 p0
y4 y11
y6 y8 p0
Kp5
Kp6
Kp7
p
y4 y2
y6 y3
y4

where p0 P=P0 is the normalised pressure inside the combustion


chamber and yi is the mole fraction of species i which can be calculated from

xi
yi P11

i1 xi

For given values of pressure, temperature, and equivalence ratio,


Eqs. (1)(6) are solved by using the NewtonRaphson iteration
method yielding the mole fractions of eleven species consequently.
The equilibrium ame temperature corresponding to given
pressure and equivalence ratio can be determined simultaneously
as follows;
NR
h
i X
h
i
NP
0
m m00i hf ;i hs;i T eq  hs;i T 0
m0i h0f ;i hs;i T R  hs;i T 0

i1

Kp1
Kp5
1
1
H2
H H2 O
OH H2
2
2
Kp2
Kp6
1
O2
O H2 CO2
H2 O CO
2
Kp3
Kp7
1
1
N2
N H2 O N2
H2 NO
2
2

i1

where subscripts eq; 0; R are the equilibrium, reference (298 K), and
reactants states, respectively.
In order to reduce the computational cost, the database of average chemical equilibrium properties (e.g., species mass fraction,
temperature) for different mixture fraction variances is generated
by using the PDF approach as

Kp4

2H2 O
2H2 O2
where Kpi are the reaction equilibrium constants i 1; 2; . . . ; 7.
These constants can be calculated from Gibbs function depending
on temperature as





1
Kp1 exp
Ru T
g H2 T  g H T
2



1
Ru T
g O2 T  g O T
Kp2 exp
2



1
Kp3 exp
Ru T
g N2 T  g N T
2
Kp4 exp 2g H2 O T  2g H2 T  g O2 T=Ru T



1
Ru T
Kp5 exp
g H2 O T  g OH T  g H2 T
2

e
/

/eq ZPZdZ

According to Pires Da Cruz et al. [1], the Beta PDF is particularly well
suited in describing the turbulent mixing between two species in a
non-reactive ow comparing to the DNS measurements. Thus, the
presumed PDF of mixture fraction for all of this work is assumed
to have a Beta distribution.
Originally, the chemical equilibrium model is based on innitely fast chemistry assumption. Due to unsteady state of ame
development, the characteristic time scale is required in order to
evaluate how fast a process is departing from equilibrium state.
According to Abraham et al. [10], the consuming/producing rates
of species mass fraction can be rewritten as
eq

dY i Y i  Y i

dt
s

Kp6 exp g H2 T g CO2 T  g H2 O T  g CO T =Ru T





1
Ru T
Kp7 exp
g H2 O T g N2 T  g H2 T  g NO T
2
1

where Y i and Y eq
i are the local and chemical equilibrium mass fraction of species i. The characteristic time scale s assumed to be the
same for all species is given as

I. Dhuchakallaya, A.P. Watkins / Applied Energy 87 (2010) 14271432

s slam f stur

10

where f is a delay coefcient given as

f 1  er =0:632

11

and

Y CO2 Y H2 O Y CO Y H2 Y OH Y NO Y N Y O Y H
1  Y N2

12

The laminar time scale employed here based on work of Westbrook


and Dryer[11] is given as

slam A1 Y F0:25 Y 1:50


exp15100=T
O2

13

and the turbulent time scale is expressed as

stur Bk=e

14

where A and B are the determining constants.


3. Results and discussion
The diesel spray combustion measurements of Akiyama et al.
[12] are adopted here to validate the combustion application of
present model. High pressure diesel fuel is injected into a high
temperature environment in a rapid compression machine. The
ignition delay time is evaluated by the combustion chamber pressure history. Spray combustion ame photographs are taken by
ICCD camera and analysed by the two colour method to estimate
the ame temperature. The experimental details are given in Table
1. A computational grid is used throughout this section as presented in Fig. 1 with inlet SMR of 10 lm. Based on a RANS approach, the solution of the transport equation system is carried
out in an EulerianEulerian framework. The further details of
transport equations for gas and liquid phases including the source
terms for droplet breakup, collision and evaporation can be found
in Beck [4].
The auto-ignition in this experiment, dened as the time which
the pressure rises, is reported approximately at 1.4 ms. However,
the rst appearance of the signicant luminous ame in the experiment is presented later at around 2.8 ms. The reaction rate

Table 1
Spray combustion test case details of Akiyama et al. [12].
Time step ls
The smallest grid width (mm)
Cell grid (cells)
Trap pressure (MPa)
Injection pressure (MPa)
Trap temperature (K)
Liquid temperature (K)
Injector radius (mm)
Injection duration (ms)

0.5
0.5
109  73
2.7
80.0
830
298
0.09
3.8

1429

obtained by calculating the pressure difference in combustion


chamber reaches a peak at 2.5 ms after injection. In the simulation,
the auto-ignition here is dened as the moment at which temperature is over 1000 K. Hence, the predicted ignition results here
employing the PDF-EBU ignition model [5] is established at
2.13 ms and located at around 32.6 mm away from nozzle and
4.8 mm above centreline. This simulation corresponds to the
experimental results of [1316] that the ignition site of higher
injection pressure, more than 100 MPa, appears in the leading vapour zone, while the ignition occurs in the periphery of liquid core
for the lower pressure. Conspicuously, the auto-ignition occurs in
the lean mixture region at mixture fraction of 0.015, corresponding
to the mixture fraction at stoichiometric condition of 0.069. This is
also evidently exhibited in the simulation results of Chomiak and
Karlsson [17] and Tap and Veynante [18]. Then the ignition kernel
expands quickly downstream along the vapour region inwards the
main body of the fuel jet to consume the mixture formed, and the
hot air entrained by the liquid fuel jet provides a premixed region
which stabilised the ame where the high reaction rate is located.
The reaction zone is prevented propagating upstream towards the
nozzle exit by the combined effect of high injection speed and
evaporative cooling of liquid jet. The increased gas temperature enhances the evaporation rate of liquid fuel resulting in a large mixture fraction, especially near the centreline where there is a plenty
of liquid fuel. Due to a large of liquid fuel to evaporate, the temperature level at centreline is the lowest in the cross-section of spray.
However, in the outer regions of the spray, fuel droplets are actively involved in the diffusion combustion process. During the fuel
vapour consumed, the ame becomes shorter and wider, and then
it appears like a reball nally. The simulated combustion process
is quite similar to the simulation results of [1720].
Fig. 2 shows the comparison between the direct photographs of
luminous ames and the predicted ame temperature contours. At
injection period of 3.0 ms, the predicted gas temperature contours
are rather higher than the experimental result. This is due to late
appearance of luminous ame at 2.8 ms in the experiment. Due
to no certain threshold temperature to dene the lift-off length
in simulation, here the threshold temperature of 2200 K as in
[18,21] is chosen to determine the lift-off length. Hence, the liftoff length predicted is approximately 47 mm. Unfortunately, the
luminous ames are not clear enough to identify the exact lift-off
length. However, the lift-off length approximated by a power-law
scaling of Siebers et al. [22,23] for this operating condition gives
value in range of 4055 mm. The higher lift-off length causes the
increase of air entrainment in the premixed combustion zone
resulting in high reaction rate [22,23]. Considering the spray shape,
there is a contracted section of spray before reaching spray head
which is called neck. This is due to the recirculation of gas velocity from spray tip to spray neck. Thus, at the neck of spray the
ame growth is certainly disturbed by this gas circulation. The ux
of backward gas velocity is relatively high in the spray neck

Fig. 1. Grid used for all test cases in the sprays (The domain is 0.2 m long and 0.03 m in radius with 109  73 cells).

1430

I. Dhuchakallaya, A.P. Watkins / Applied Energy 87 (2010) 14271432

Fig. 2. Comparison between ame luminosity from the experiment of Akiyama et al. [12] (left) and the predicted ame temperature distributions (right).

resulting in high turbulence and mixture fraction variance. This


substantially obstructs the ame propagation due to high uctuation of combustible mixture. The ame shapes at 4.2 and 5.2 ms
are quite comparable to the luminous ames, except the appearance of spray tail at 5.2 ms in the experimental data.
The chemical reaction is dominant in the rst period of combustion, and the turbulent reaction is carried on in the main combustion subsequently [5]. This effect of dominant reaction is also
presented evidently in Fig. 3. In the early period, the chemical reaction time scale is relatively higher than the turbulent reaction time
scale leading to the chemical reaction governing the burning rate.
Accordingly, all reaction time scales decrease rapidly due to the increased temperature, especially the chemical reaction time scale.
Beyond 2.4 ms, the turbulent reaction time scale provides the higher time scale resulting in controlling the reaction rate until the end
of combustion process where all reaction time scales move suddenly up again.
Fig. 4 shows schematically the spray diffusion ame with the
relative location of spray boundary. The gas temperature, reactant
and product species distributions predicted are taken at time of
4.0 ms after injection and axial position of 74 mm away from injector. Beyond the lift-off length where is the location of the maximum reaction layer, the smaller droplets evaporate rapidly to
provide fuel vapour which is consumed mostly near the outer portion of the turbulent diffusion ame, and then the larger droplets
can travel further before evaporation and combustion processes

Reaction time scale (s)

1E-4

Chemical reaction time scale


Turbulent reaction time scale
Reaction time scale

1E-5
1E-6
1E-7
1E-8
1E-9
1E-10
1E-11
2

tig

4
5
6
Time after injection (ms)

Fig. 3. The development of reaction time scales during combustion predicted by the
PDF-Chemical Equilibrium combustion model.

begin due to the relatively high inertia. As seen, the reactant species predicted by both models present the overlapping region of
fuel vapour and oxidizer. This overlapping region occurs possibly
due to the reversible reaction and/or a limited time of reactants
to react. The Chemical Equilibrium model is based on reversible
fast equilibrium chemistry, hence both fuel and oxidant certainly
present around stoichiometric ratio. In addition, the reaction time
scale scheme can relax the innitely fast chemistry. Remarkably,
the overlapping region being the location of stoichiometric point
also provides a maximum temperature. Additionally, the PDF
shapes along radius employed for evaluating the mean scalar values are also presented. At the centreline, the shape of PDF looks
like nearly homogeneous mixture behaviour. With increasing distance from centreline, the peak point of PDF shape gradually leans
to y-axis referring to the decrease of mixture fraction, and nally,
at the outer spray boundary, its shape transforms from K to L
shapes which exhibits nearly pure air behaviour. In additional,
the cross-sectional distributions of reactants and products are also
illustrated. The major products including CO2 and H2 O species
reach peak at around stoichiometric condition where there is the
overlap of fuel vapour and oxidant, and the peaks of CO and H2
are located at centreline of spray.
The ame distribution is also compared in term of ame area as
shown in Fig. 5. Occurrence of the predicted ame is obviously earlier than the experimental result, and then the predicted maximum
ame area becomes relatively comparable to the experimental
ame area but its peak point is located at different time. Nevertheless, the ame area of the experimental data develops in the same
rate as the simulation result. Fig. 6 shows how accurately present
model can capture heat release rate. In the early stage, heat release
rate slightly decreases below zero due to heating up the liquid
droplets. It then rapidly increases after ignition occurs resulting
in high reaction rate. Later on, the diffusion ame becomes dominant to control the reaction. The ame approaches combustion
chamber wall at time of 5.2 ms in the experiment. Hence the results after time of 5.2 ms are not considered. Comparing with the
experimental results, the peak of reaction rate obtained from the
experiment is lower than the peak of the simulation result. In main
combustion, combustion models present higher reaction rate than
the experimental data until time of around 4.5 ms. This is possible
that the evaporation model employed here performs imperfectly.
At low temperature, it can produce less fuel vapour than the experimental result leading to a longer time for heating up and

1431

I. Dhuchakallaya, A.P. Watkins / Applied Energy 87 (2010) 14271432

evaporating later on in early period. On the other hand, the fuel vapour generated is larger than the experimental results at high temperature. Thus, there is a lot of fuel vapour to consume in main
combustion leading to higher heat release rate predicted by this
model. This is also evident that the combustion period of the
experiment is slightly longer than that of the simulation result.
The details of ame temperature distribution represented in
term of temperature histogram are also illustrated in Fig. 7. As
seen, the simulation results present good agreement with the
experimental results at time of 4.0 ms after start of injection. The
predicted mean temperature corresponds to the experimental data
at around 1950 K. At time of 5.0 ms, model can capture well only
the average value of histogram matching to temperature of
2100 K whilst its amplitude is far different from the experimental
data. This is because the remaining fuel vapour used for future
burning in simulation at present time is considerably less than that

Heat release rate (J/ms)

Fig. 4. Schematic representation of spray diffusion ame predicted by the PDF-Equilibrium model.

900

Experiment [12]

700

Simulation

500
300
100
-100
0

Time after injection (ms)


Fig. 6. Time history of heat release rates for the experiment of Akiyama et al. [12].

2000
Exp [12] at 4.0 ms

Experiment [12]

2000

Flame area (mm2)

Flame area (mm2)

2500

Simulation

1500
1000
500
0
1.0

2.0

3.0

4.0

5.0

Exp [12] at 5.0 ms

1600

Sim at 4.0 ms
Sim at 5.0 ms

1200
800
400
0
1400

1900

2400

2900

Flame temperature (K)

Time after injection (ms)


Fig. 5. Development of ame areas for the experiment of Akiyama et al. [12].

Fig. 7. Comparing ame temperature histogram for the experiment of Akiyama


et al. [12].

1432

I. Dhuchakallaya, A.P. Watkins / Applied Energy 87 (2010) 14271432

in the experimental data. Although the ame area is quite similar,


large ame area appearing in this simulation presents in low-temperature range, especially in spray tail. Principally, the PDF-EQ
combustion model is capable to capture this spray combustion
experiment fairly well.
4. Conclusion
The combustion model employed in this work is applied to diesel spray combustion applications, especially in the commercial
diesel engines. The PDF-Chemical Equilibrium combustion model
is developed here in order to effectively perform in combustion
mode. In this combustion model, the average scalars of the chemical equilibrium mass fractions of reactants and products and adiabatic ame temperature based on the PDF approach are required
to calculate and store in the database. Thus, the dissociation of major species is allowed leading to further investigation of the combustion products and pollution. The reaction time scale is
employed to relax the innitely fast chemistry in the chemical
equilibrium model. In the auto-ignition section, the PDF-Eddy
Break Up ignition model[5] is adopted which performs successfully
similar to the Shell ignition model. With regard to the comparison
results, the simulation results are satisfactory in predicting the
ignition delay time and ame development with the experimental
results. In addition, the predicted lift-off length also corresponds to
a power-law scaling of Siebers et al. [22,23].
References
[1] Pires da Cruz A, Baritaud TA, Poinsot TJ. Self-ignition and combustion modeling
of initially nonpremixed turbulent systems. Combust Flame 2001;124:6581.
[2] Tao F, Golovitchev VI, Chomiak J. Application of complex chemistry to
investigate the combustion zone structure of DI diesel sprays under enginelike conditions. In: The fth international symposium on diagnostics and
modeling of combustion in internal combustion engines (COMODIA 2001),
Nagoya; 2001.
[3] Tang Q, Zhao W, Bockelie M, Fox RO. Multi-environment probability density
function method for modelling turbulent combustion using realistic chemical
kinetics. Combust Theory Model 2007;11:889907.
[4] Beck JC. Computational modelling of polydisperse sprays without segregation
into droplet size classes. PhD thesis, Manchester, UMIST; 2000.
[5] Dhuchakallaya I. Development and application of the drop number size
moment modelling sprays to engine simulations and application of
combustion models. Manchester: The University of Manchester; 2009.

[6] Burke SP, Schumann TW. Diffusion ames. In: Proceedings of the rst
symposium on combustion, Pittsburgh, Swampscott, MA; 1928 (reprint of
proceedings published by The Combustion institute in 1965).
[7] Cowart JS, Keck JC, Heywood JB, Westbrook CK, Pitz WJ. Engine knock
predictions using fully-detailed and a reduced chemical kinetic mechanism. In:
The 23rd international symposium on combustion, Pittsburgh; 1990.
[8] Schreiber M, Sakak AS, Lingens A, Grifths JF. A reduced thermokinetic model
for the autoignition of fuels with variable octane ratings. In: The 25th
international symposium on combustion, Irvine; 1994.
[9] Jones WP, Lindstedt RP. Global reaction schemes for hydrocarbon combustion.
Combust Flame 1988;73:23349.
[10] Abraham J, Bracco FV, Reitz RD. Comparisons of computed and measured
premixed charge engine combustion. Combust Flame 1985;60:30922.
[11] Westbrook CK, Dryer FL. Chemical kinetic modelling of hydrocarbon
combustion. Prog Energy Combust 1984;10:157.
[12] Akiyama H, Nishimura H, Ibaraki Y, Iida N. Study of diesel spray combustion
and ignition using high-pressure fuel injection and a micro-hole nozzle with a
rapid compression machine: improvement of combustion using low cetane
number fuel, vol. 19. Society of Automotive Engineers of Japan; 1998. p. 319
27.
[13] Crua C. Combustion processes in a diesel engine. PhD thesis, Brighton,
University of Brighton; 2002.
[14] Bruneaux G, Aug M, Lemenand C. A study of combustion structure in high
pressure single hole common rail direct diesel injection using laser induced
uorescence of radicals. In: The sixth international symposium on diagnostics
and modeling of combustion in internal combustion engines, COMODIA 2004,
Yokohama, Japan; 2004.
[15] Ganippa LC, Andersson S, Chomiak J. Combustion characteristics of diesel
sprays from equivalent nozzles with sharp and rounded inlet geometries.
Combust Sci Technol 2003;175:101532.
[16] Larsson A. Optical studies in a DI diesel engine. SAE technical paper series no.
1999-01-3650; 1999.
[17] Chomiak J, Karlsson A. Flame lift-off in diesel sprays. In: The 26th international
symposium on combustion, Pittsburgh; 1996.
[18] Tap FA, Veynante D. Simulation of ame lift-off on a diesel jet using a
generalized ame surface density modeling approach. In: Proceedings of the
Combustion Institute, vol. 30; 2005 p. 91926.
[19] Golovitchev VI, Nordin N, Chomiak J. Modeling of spray formation, ignition and
combustion in internal combustion engines, publication nr. 98/1. Thermo &
uid dynamics. Chalmers University of Technology; 1998.
[20] Lehtiniemi H, Mauss F, Balthasar M, Magnusson I. Modeling diesel spray
ignition using detailed chemistry with a progress variable approach.
Combustion Science and Technology 2006;178:197797.
[21] Senecal PK, Pomraning E, Richards KJ, Briggs TE, Choi CY, McDavid RM,
Patterson MA. Multi-dimensional modeling of direct-injection diesel spray
liquid length and ame lift-off using CFD and parallel detailed chemistry. SAE
technical paper series no. 2003-01-1043; 2003.
[22] Siebers DL, Higgins BS. Flame lift-off on direct-injection diesel sprays under
quiescent conditions. SAE technical paper series no. 2001-01-0530; 2001.
[23] Siebers DL, Higgins BS, Pickett LM. Flame lift-off on direct injection diesel fuel
jets: oxygen concentration effects. SAE technical paper series no. 2002-010890; 2002.

Das könnte Ihnen auch gefallen