24.07.2013 Views

Numerical Modeling of Diesel Spray Formation and Combustion C ...

Numerical Modeling of Diesel Spray Formation and Combustion C ...

Numerical Modeling of Diesel Spray Formation and Combustion C ...

SHOW MORE
SHOW LESS

You also want an ePaper? Increase the reach of your titles

YUMPU automatically turns print PDFs into web optimized ePapers that Google loves.

<strong>Numerical</strong> <strong>Modeling</strong> <strong>of</strong> <strong>Diesel</strong> <strong>Spray</strong> <strong>Formation</strong> <strong>and</strong> <strong>Combustion</strong><br />

C. Bekdemir ∗ , L.M.T. Somers, L.P.H. de Goey<br />

Mechanical Engineering, Eindhoven University <strong>of</strong> Technology, The Netherl<strong>and</strong>s<br />

www.combustion.tue.nl<br />

Abstract<br />

A study is presented on the modeling <strong>of</strong> fuel sprays in diesel engines. The objective <strong>of</strong> this study is in the first place to<br />

accurately <strong>and</strong> efficiently model non-reacting diesel spray formation, <strong>and</strong> secondly to include ignition <strong>and</strong> combustion.<br />

For that an efficient 1D Euler-Euler spray model [21] is implemented <strong>and</strong> applied in 3D CFD simulations. Concerning<br />

combustion, a detailed chemistry tabulation approach, called FGM (Flamelet Generated Manifold), is adopted.<br />

Results are compared with EHPC (Eindhoven High Pressure Cell) experiments, data from S<strong>and</strong>ia <strong>and</strong> IFP. The newly<br />

created combination <strong>of</strong> the 1D spray model with 3D CFD gives a good overall performance in terms <strong>of</strong> spray length<br />

<strong>and</strong> shape prediction, <strong>and</strong> also numerically it has advantages above Euler-Lagrange type models. Together with the<br />

FGM, also auto-ignition <strong>and</strong> a flame lift-<strong>of</strong>f length is achieved.<br />

Introduction<br />

Due to ever increasing dem<strong>and</strong>s from emission legislation<br />

(NOx <strong>and</strong> soot), fuel economy (CO2) <strong>and</strong> fuel<br />

flexibility (bio-fuels) diesel engines become more <strong>and</strong><br />

more complex. Therefore, conventional engine design approaches<br />

that rely on prototype development become too<br />

time-consuming <strong>and</strong> expensive. The development <strong>of</strong> predictive<br />

<strong>and</strong> efficient computational tools would represent<br />

a significant step forward in the ability to rapidly design<br />

high efficiency, low emission engines [8].<br />

Modern diesel engine technology unequivocally applies<br />

liquid fuel injection with high pressure, that forms<br />

a non-homogenous mixture leading to relatively high levels<br />

<strong>of</strong> soot. This spray formation process may seem<br />

straightforward, but in reality it is dauntingly complex.<br />

Furthermore, combustion presents especially great challenges<br />

[14]. For that reason, accurate <strong>and</strong> fast CFD is<br />

needed. The objective <strong>of</strong> this study is to accurately <strong>and</strong> efficiently<br />

model diesel spray formation, <strong>and</strong> capture autoignition<br />

<strong>and</strong> flame lift-<strong>of</strong>f by means <strong>of</strong> a tabulated chemistry<br />

method called FGM (Flamelet Generated Manifold).<br />

<strong>Spray</strong> formation <strong>and</strong> combustion modeling are described<br />

in the following sections, respectively. Subsequently<br />

some reacting spray results are presented. And finally<br />

some conclusions <strong>and</strong> recommendations are given.<br />

<strong>Spray</strong> <strong>Modeling</strong><br />

Fluent’s DPM model (Euler-Lagrange method) is extensively<br />

used to model evaporating, but inert heptane<br />

sprays. This is done with special attention for temperature<br />

dependent material properties <strong>and</strong> for many different<br />

setups, including various meshes, solver timesteps <strong>and</strong><br />

amount <strong>of</strong> parcels (groups <strong>of</strong> identical droplets). The results<br />

are compared with a measurement on the EHPC <strong>and</strong><br />

experimental data from S<strong>and</strong>ia. From these comparisons<br />

it is found that the DPM model gives unsatisfactory results<br />

concerning spray <strong>and</strong> liquid lengths. From a numerical<br />

point <strong>of</strong> view there are also major disadvantages. The<br />

∗ Corresponding author: c.bekdemir@tue.nl<br />

Proceedings <strong>of</strong> the European <strong>Combustion</strong> Meeting 2009<br />

results are highly mesh <strong>and</strong> timestep dependent <strong>and</strong> <strong>of</strong>ten<br />

convergence problems occur. Also the statistical approach<br />

with parcels is a source <strong>of</strong> problems due to large<br />

computing times when parcels accumulate in the domain.<br />

Many authors tried to overcome these problems by fine<br />

tuning the submodels for specific cases, but this is obviously<br />

not the way to go due to the fundamental discrepancy<br />

between, on one h<strong>and</strong> the limitation to cell sizes <strong>and</strong><br />

lack <strong>of</strong> parallelization possibilities, <strong>and</strong> on the other h<strong>and</strong><br />

solving in-cylinder velocity fields <strong>and</strong> turbulence with increasingly<br />

finer meshes.<br />

However, others also developed (semi) Euler-Euler<br />

methods to model spray formation. One <strong>of</strong> them is the socalled<br />

ICAS (Interactive Cross section Averaged <strong>Spray</strong>)<br />

model [17, 22] that combines a 1D (Eulerian) spray model<br />

with the existing Langrangian CFD model. The region<br />

near the nozzle exit is then covered with the 1D spray<br />

model <strong>and</strong> somewhere further downstream droplet source<br />

terms are introduced to incorporate 3D spray formation<br />

in the CFD environment. By introducing parcels anyway,<br />

the influence <strong>of</strong> in-cylinder flow to spray formation would<br />

still be captured [18], but to do so the Euler-Euler region<br />

should be not more than a few millimeters from the nozzle<br />

exit, because further downstream (behind the liquid<br />

length) there are no droplets anymore. This makes complex,<br />

adaptive meshes necessary, <strong>and</strong> the DPM model <strong>of</strong><br />

Fluent would still be needed. However, ideally the CFD<br />

code should be used for gas phase calculations only in order<br />

to circumvent the complex discrete phase interaction.<br />

Therefore a 1D model that covers the complete spray region<br />

is presented in this chapter. This model is coupled<br />

to Fluent with appropriate source terms for mass (fuel vapor),<br />

momentum <strong>and</strong> energy.<br />

In the following, first the phenomenological spray<br />

model proposed by Versaevel et al. [21] is introduced.<br />

The model is then implemented in Matlab <strong>and</strong> validated<br />

with measurements <strong>of</strong> IFP [20] <strong>and</strong> S<strong>and</strong>ia [7][15]. Last<br />

but not least, source terms are extracted from the 1D<br />

model <strong>and</strong> put into Fluent via UDFs (User-Defined Function),<br />

<strong>and</strong> the resulting 3D solutions are also compared<br />

with measurements.


1D Euler-Euler <strong>Spray</strong> Model<br />

The 1D quasi steady spray model <strong>of</strong> Versaevel et al.<br />

[21] is an extension <strong>of</strong> the earlier efforts <strong>of</strong> Naber et al.<br />

[11] <strong>and</strong> Siebers [15]. Naber <strong>and</strong> Siebers developed a 1D<br />

model for non-vaporizing spray penetration first, <strong>and</strong> later<br />

Siebers added some thermodynamics to distinguish liquid<br />

penetration from vapor penetration. But Siebers’ contribution<br />

is based on the assumption that only at the steady<br />

liquid length position thermodynamic equilibrium exists.<br />

This approach implies that no temperature information is<br />

available, except at the liquid length position. Also the<br />

composition <strong>of</strong> the spray volume between the nozzle exit<br />

<strong>and</strong> liquid length is unknown. Versaevel et al. overcame<br />

this shortcoming by introducing a void fraction m that<br />

couples the mass, momentum <strong>and</strong> energy equations.<br />

The model is based on five basic assumptions from<br />

which the first four are the same as in the work <strong>of</strong> Naber<br />

et al. [11] <strong>and</strong> Siebers [15].<br />

⇒ no velocity slip between gas <strong>and</strong> liquid phases<br />

⇒ whole system at constant pressure<br />

⇒ uniform velocity, density <strong>and</strong> temperature pr<strong>of</strong>iles<br />

⇒ constant spray angle<br />

⇒ whole system at thermodynamic equilibrium<br />

The last assumption is necessary to gain information<br />

about the temperature <strong>and</strong> composition <strong>of</strong> the spray in<br />

the entire 1D domain. The mass, momentum <strong>and</strong> energy<br />

equations are derived by considering a control volume as<br />

defined in Figure 1. The spray is described in one direction<br />

due to the constant angle <strong>and</strong> the axisymmetry. The<br />

figure shows that from fuel injection ( ˙mfl,0) into the xdirection<br />

the spray diverges due to air entrainment ( ˙ma)<br />

into the spray volume. Air entrainment is controlled by<br />

the prescribed spray angle ( θ<br />

2 ). For this purpose an experimental<br />

dispersion relation is chosen. At the liquid length<br />

just enough hot air is entrained into the spray to evaporate<br />

all liquid fuel, so from that point on the fuel penetrates the<br />

surrounding gas as a vapor.<br />

The phenomenological spray model is implemented in<br />

Matlab. The st<strong>and</strong>ard non-linear solver <strong>of</strong> Matlab is used<br />

for this purpose. Material properties, except the liquid fuel<br />

.<br />

mfl,0<br />

.<br />

ma<br />

.<br />

ma<br />

control volume<br />

/2<br />

d eff<br />

gas <strong>and</strong> liquid gas only<br />

Figure 1: Definition <strong>of</strong> the control volume used in the derivation<br />

<strong>of</strong> the 1D model [21]<br />

df<br />

x<br />

2<br />

density, are temperature dependent, <strong>and</strong> are obtained from<br />

the thermophysical database <strong>of</strong> DIPPR [6]. The calculated<br />

spray length compares good with IFP measurements [20]<br />

as shown in Figure 2, indicated with the solid en dotted<br />

lines, respectively.<br />

SL [mm]<br />

50<br />

45<br />

40<br />

35<br />

30<br />

25<br />

20<br />

15<br />

case: IFP heptane<br />

10<br />

5<br />

3D model<br />

Fluent DPM<br />

1D model<br />

fit through IFP measurements<br />

0<br />

0 0.1 0.2 0.3 0.4 0.5<br />

Time [ms]<br />

0.6 0.7 0.8 0.9 1<br />

Figure 2: <strong>Spray</strong> length as function <strong>of</strong> time with the Euler-Euler<br />

3D model, compared to Fluent DPM, Euler-Euler 1D model <strong>and</strong><br />

IFP measurement<br />

3D <strong>Spray</strong> Simulation<br />

The 1D phenomenological spray model discussed in<br />

the previous section is, in contrary to the earlier model <strong>of</strong><br />

Naber <strong>and</strong> Siebers, suitable to apply in combination with a<br />

3D CFD code. To accomplish such an interaction, source<br />

terms are extracted from the 1D model <strong>and</strong> are assigned<br />

to the corresponding transport equations in Fluent (see [3]<br />

for the details). Subsequently the combined model is validated<br />

through spray length comparison with experimental<br />

data. IFP [20] <strong>and</strong> S<strong>and</strong>ia [15][7] measurements are used<br />

for validation purposes. These are all for single component<br />

fuels that are well documented, so thermophysical<br />

data needed for the numerical model is found in literature.<br />

The implemented 3D model (circles) predicts the<br />

spray length better than Fluent’s DPM model (stars), as is<br />

shown in Figure 2. The correctness <strong>of</strong> the 3D model prediction<br />

is best visualized with the contours <strong>of</strong> fuel mass<br />

fraction at the spray cross-section shown in Figure 3. The<br />

upper spray is a DPM simulation result <strong>and</strong> the other one<br />

is gained with the 3D model, both at 1 ms. Apart from the<br />

obvious spray length difference, the shape/width <strong>of</strong> the<br />

sprays are also dissimilar. DPM gives too wide sprays,<br />

since relatively large cells have to be used to meet the requirements<br />

<strong>of</strong> the Lagrangian approach. In the 3D Euler-<br />

Euler case one can refine the grid until the spray is resolved<br />

sufficiently, without having discrete phase related<br />

problems.<br />

Summarizing, the better overall performance (spray<br />

length <strong>and</strong> shape) <strong>and</strong> the proper mesh resolution behavior<br />

(higher resolution gives better solutions) <strong>of</strong> the 3D<br />

Euler-Euler model, together with the ability to parallelize<br />

is a major advantage compared with Fluent’s DPM model.<br />

Since the DPM model uses only one CPU to do all discrete


Figure 3: Case: IFP heptane. Contours <strong>of</strong> fuel mass fraction<br />

gained with Fluent’s DPM model <strong>and</strong> the implemented 3D spray<br />

model. Note the minimum/maximum values between the brackets<br />

at the right h<strong>and</strong> side, the colorbar is scaled for each separately<br />

phase calculations no matter how many CPUs are available.<br />

The consequence is large computation times despite<br />

the relatively small amount <strong>of</strong> cells.<br />

Possible improvements to the 1D/3D spray model for<br />

the future may be the following:<br />

⇒ Make 3D spray model pressure dependent in order to<br />

simulate spray formation in a variabele volume combustion<br />

chamber.<br />

⇒ Sound spray angle prediction. Not really an improvement<br />

to the model itself, but spray angle prediction<br />

certainly has a large effect on the final results. Therefore<br />

the quest for proper <strong>and</strong> generic angle prediction<br />

methods/relations should be promoted.<br />

<strong>Combustion</strong> <strong>Modeling</strong><br />

In the preceding chapters a 3D Euler-Euler spray<br />

model is implemented <strong>and</strong> validated for evaporating inert<br />

fuel sprays. This model is mesh <strong>and</strong> solver timestep<br />

independent, <strong>and</strong> is suitable for parallel simulations. So,<br />

regarding the status <strong>of</strong> modeling the mixing process, additional<br />

modeling features, which may require fine spatial<br />

<strong>and</strong> time resolutions, can be included. In this chapter an<br />

attempt is made to add combustion, more specific, the emphasis<br />

is on the application <strong>of</strong> FGMs (Flamelet Generated<br />

Manifolds) in modeling <strong>of</strong> the turbulent combustion <strong>of</strong> a<br />

transient igniting spray.<br />

In the following, first the principle <strong>of</strong> flamelets <strong>and</strong> its<br />

use for modeling combustion with tabulated chemistry is<br />

shortly mentioned. Then, the procedure <strong>of</strong> a FGM generation<br />

is shown. Subsequently, its implementation into<br />

Fluent is described.<br />

FGM Approach<br />

Detailed models can be very accurate, but unfortunately<br />

also computationally very expensive. To overcome<br />

3<br />

the impractical computing times, while solving the combustion<br />

process still with high detail (depends on used reaction<br />

scheme), an approach with tabulated chemistry is<br />

applied. This so-called FGM approach is developed by<br />

van Oijen [19] for laminar premixed flames, <strong>and</strong> makes<br />

use <strong>of</strong> 1D laminar flamelet data to tabulate composition,<br />

density, temperature etc. as function <strong>of</strong> local control<br />

variables. However, the flamelet concept views a turbulent<br />

flame as an ensemble <strong>of</strong> thin, laminar, locally 1D<br />

flames, called flamelets, embedded within the turbulent<br />

flow field. Furthermore, the concept is based on the assumption<br />

that the smallest turbulent time <strong>and</strong> length scales<br />

are much larger than the chemical ones, <strong>and</strong> there exists a<br />

locally undisturbed sheet where chemical reactions occur<br />

[16]. So, later Ramaekers [13] extended the application to<br />

turbulent partially-premixed combustion by choosing one<br />

control variable describing non-premixed (mixture fraction<br />

Z) <strong>and</strong> one describing premixed (reaction progress<br />

variable P V ) combustion, <strong>and</strong> by PDF (Probability Density<br />

Function) integration to account for turbulence.<br />

In this study non-premixed flamelets for a counterflow<br />

setup are solved with CHEM1D [1], which is a specialized<br />

one-dimensional laminar flame code developed at the<br />

Eindhoven University <strong>of</strong> Technology. A heptane flamelet<br />

database at constant pressure is calculated, making use <strong>of</strong><br />

a reduced n-heptane mechanism [12].<br />

In non-premixed combustion it is common practice to<br />

introduce the mixture fraction Z, here the definition <strong>of</strong><br />

Bilger [4] is adopted:<br />

Z =<br />

2 YC−YC,2<br />

MC<br />

2 YC,1−YC,2<br />

MC<br />

1 +<br />

YH −YH,2 (YO−YO,2)<br />

2 − MH<br />

MO<br />

+ 1 YH,1−YH,2<br />

2 − MH<br />

YO,1−YO,2<br />

MO<br />

, (1)<br />

where Y st<strong>and</strong>s for mass fraction, M is the molar mass<br />

<strong>and</strong> the subscripts C, H <strong>and</strong> O indicate the quantities<br />

for the elements carbon, hydrogen <strong>and</strong> oxygen, respectively.<br />

The subscripts 1 <strong>and</strong> 2 refer to the constant mass<br />

fraction in the original fuel <strong>and</strong> oxidizer streams, respectively.<br />

In the fuel stream the mixture fraction is equal to<br />

unity <strong>and</strong> monotonically decreases to zero at the oxidizer<br />

stream. An additional control variable, called the reaction<br />

progress variable P V , is introduced to parameterize the<br />

progress <strong>of</strong> the irreversible combustion process. In this<br />

study a combination <strong>of</strong> CO2, CO <strong>and</strong> CH2O mass fractions<br />

is chosen as a reaction progress variable:<br />

P V = YCO2<br />

MCO2<br />

+ YCO<br />

+<br />

MCO<br />

YCH2O<br />

. (2)<br />

MCH2O<br />

The succes <strong>of</strong> this concept is related to the fact that<br />

all occurring compositions tend to have a common,<br />

low-dimensional, attractor in composition space, a socalled<br />

intrinsic low-dimensional manifold (ILDM) [9].<br />

Hence, the complex chemistry is reduced <strong>and</strong> completely<br />

described by the mixture fraction Z <strong>and</strong> the reaction<br />

progress variable P V .<br />

FGM Construction <strong>and</strong> Implementation<br />

FGMs can be generated in many ways. For stationary<br />

flames, there is a classical way with steady flamelets only,


PV [−]<br />

Flamelet database generation<br />

igniting flamelet<br />

extinguishing flamelet<br />

igniting PSRs<br />

0 0.1 0.2 0.3 0.4 0.5<br />

Z [−]<br />

0.6 0.7 0.8 0.9 1<br />

Lowest strainrate<br />

Steady solutions region<br />

Highest non−quenching strainrate<br />

Timedependently extinguishing or igniting flamelet<br />

Perfectly Stirred Reactors (PSRs) before ignition<br />

Figure 4: Ways to generate a ’full’ flamelet database<br />

where a sequence <strong>of</strong> steady flames with strainrates varying<br />

from a low value to the quenching value is computed.<br />

An illustrative example is shown in Figure 4, see the gray<br />

area between the solution for the lowest strainrate <strong>and</strong> the<br />

solution at which the strainrate reached its maximum before<br />

extinction.<br />

But a spray event is unsteady <strong>and</strong> initially nonreacting,<br />

so to cover the ignition process the table should<br />

also contain information in the area beneath the quenching<br />

strainrate solution. Several ways exist to fill this gap<br />

in the Z-P V plane. One way is to solve a timedependent<br />

flamelet with a higher strainrate than the highest possible<br />

non-quenching strainrate, in this way forcing the flame to<br />

extinguish <strong>and</strong> in the mean time sampling data to fill the<br />

gap. Another approach, that is more appropriate for this<br />

study, is solving timedependent flamelets from a mixed,<br />

but non-reacting initial state. The ignition behavior is followed<br />

in time until a steady flame is reached. The third<br />

possibility is to reproduce ignition <strong>of</strong> mixtures covering<br />

the entire Z-space with PSR (Perfectly Stirred Reactor)<br />

auto-ignition calculations [10]. All three methods to fill<br />

the Z-P V gap are depicted schematically in Figure 4.<br />

The way(s) a FGM is constructed in this study is depicted<br />

schematically in Figure 5.<br />

Due to the unsteady nature <strong>of</strong> a diesel injection event,<br />

ignition modeling is at least as important as combustion<br />

modeling. Following the FGM approach, besides combustion,<br />

ignition should be covered inherently. But not<br />

surprisingly this depends on the way the FGM is generated.<br />

The extinguishing flamelet approach is applied <strong>and</strong><br />

does not lead to ignition <strong>of</strong> the spray. Instead, only local<br />

temperatures slightly above the initial ambient temperature<br />

are found, <strong>and</strong> the source <strong>of</strong> the reaction progress<br />

variable is not big enough to end in total ignition within<br />

a few milliseconds. However, a FGM constructed with<br />

4<br />

an igniting flamelet database does result is auto-ignition<br />

<strong>of</strong> the whole spray in short time. Therefore, in this paper<br />

only the results <strong>of</strong> the igniting flamelet approach are<br />

presented.<br />

The turbulence-chemistry interaction is accounted for<br />

by integrating the quantities ϕ in the 2D table with a β-<br />

PDF function as follows:<br />

˜ϕ =<br />

1 1<br />

0<br />

0<br />

ϕ(Z, P V ) ˜ P (Z) ˜ P (P V ) dZdP V. (3)<br />

Note that this explicit formulation assumes that Z <strong>and</strong><br />

P V are statistically independent. The overtilde st<strong>and</strong>s for<br />

Favre (mass) averaged. Both control variables are from<br />

then on described with a mean value ( ˜ Z, P V ) <strong>and</strong> a variance<br />

(Z” 2 , P V ” 2 ), so a quantity is defined by the probability<br />

<strong>of</strong> occurrence for several states instead <strong>of</strong> one fixed<br />

state. The chemistry is in this way extended to a 4D lookup<br />

table with the means <strong>and</strong> variances <strong>of</strong> the two control<br />

variables as the parameters (look-up indices).<br />

The 4D FGM combustion model is implemented in<br />

Fluent, in order to do turbulent spray combustion simulations<br />

in 3D space. The four scalars ( ˜ Z, P V , Z” 2 , P V ” 2 )<br />

are solved with user-defined scalar transport equations, in<br />

addition to the st<strong>and</strong>ard continuity, momentum <strong>and</strong> turbulence<br />

equations. All species concentrations <strong>and</strong> corre-<br />

quenching flame<br />

Z-PV domain filled with<br />

stationary loop over strainrates ,<br />

<strong>and</strong> with timedependent<br />

quenching flame<br />

TRF mechanism<br />

48 species<br />

248 reactions<br />

CHEM1D<br />

solves:<br />

flamelet equations<br />

(constant pressure )<br />

2D FGM<br />

Z, PV – table<br />

(laminar)<br />

interpolated laminar flamelet data ρ,<br />

sPV, Yi <strong>and</strong> T as function <strong>of</strong><br />

Z <strong>and</strong> PV<br />

4D FGM<br />

"2 "2<br />

Z, Z , PV ,PV – table<br />

(turbulence included )<br />

2D data integrated with PDF<br />

functions. ρ, sPVm, <br />

"2<br />

sPV ,sPV sPVv, Yi <strong>and</strong> T<br />

as function <strong>of</strong> mmmmmmmm "2<br />

, , "2<br />

Z Z PV ,PV<br />

Figure 5: FGM construction scheme<br />

igniting flame<br />

Z-PV domain filled with , from<br />

initial pure mixing solution ,<br />

igniting flame , using the<br />

timedependent solver


sponding temperatures are in principle known from the<br />

flamelet database for any mixture fraction <strong>and</strong> progress<br />

variable combination.<br />

Results <strong>and</strong> Discussion<br />

The evolution from the early stage <strong>of</strong> ignition to further<br />

combustion <strong>of</strong> the spray, injected from the left, is<br />

shown at six moments in time in Figure 6. The upper half<br />

<strong>of</strong> the plots represent the values <strong>of</strong> the progress variable<br />

<strong>and</strong> the lower parts are contours <strong>of</strong> temperature. Several<br />

interesting observations are done from this figure. First, at<br />

the outer edge <strong>of</strong> the spray activity begins, here shown by<br />

means <strong>of</strong> an increased (<strong>and</strong> still increasing) progress variable<br />

<strong>and</strong> a corresponding increase in temperature; from<br />

an initial 800 K ambient to around 1000 K locally. This<br />

activity is particularly present close to the place the flame<br />

lift-<strong>of</strong>f will settle. Further in time the outer contour <strong>of</strong> the<br />

igniting spray is becoming clearer due to high values <strong>of</strong><br />

P V <strong>and</strong> T . Finally, the full outer edge will be reacting<br />

<strong>and</strong> the combustion region exp<strong>and</strong>s to the inner volume<br />

<strong>and</strong> the maximum temperature continues to rise.<br />

Also a much simpler combustion model is used that<br />

is available in Fluent, called eddy-dissipation model [2].<br />

This model is, like the flamelet approach, a mixinglimited<br />

combustion model, with the differences that only<br />

the global reaction from fuel <strong>and</strong> O2 to CO2 <strong>and</strong> H2O is<br />

considered <strong>and</strong> immediate reaction is assumed. So, it is<br />

not expected that this simple model can be as accurate as<br />

the much more detailed FGM method, but it gives nice<br />

material to compare regarding for example temperature<br />

pr<strong>of</strong>iles. Such a comparative picture is given in Figure 7,<br />

taken at 1 ms. In contrast to the eddy-dissipation model,<br />

apart from the inherent auto-ignition, also a flame lift-<strong>of</strong>f<br />

settles automatically using tabulated chemistry.<br />

The diffusion flame at the spray edge stays the hottest<br />

region, as can be expected on the basis <strong>of</strong> the used nonpremixed<br />

flamelet database. But according to the conceptual<br />

diesel combustion model <strong>of</strong> Dec [5] it is stated that<br />

in DI diesel injection, regions with premixed <strong>and</strong> nonpremixed<br />

combustion can be distinguished. The next step<br />

Tmax = 807 K Tmax = 853 K<br />

Figure 6: Case: IFP heptane. Temporal sequence <strong>of</strong> progress<br />

variable <strong>and</strong> temperature contours showing the auto-ignition<br />

process resulting in total combustion<br />

5<br />

Figure 7: Case: IFP heptane. Contours <strong>of</strong> temperature gained<br />

with the eddy-dissipation model <strong>and</strong> the implemented FGM<br />

model. Note the minimum/maximum values between the brackets<br />

at the right h<strong>and</strong> side, the colorbar is scaled for each separately<br />

may be the creation <strong>of</strong> a partially-premixed database in<br />

order to capture the presence <strong>of</strong> both regimes at the same<br />

time.<br />

Unfortunately, there are no experimental ignition delay<br />

<strong>and</strong> lift-<strong>of</strong>f length data published for this case in the<br />

paper <strong>of</strong> Verhoeven et al. [20]. However, the numerically<br />

found delay time <strong>and</strong> lift-<strong>of</strong>f length are in the order<br />

<strong>of</strong> magnitude <strong>of</strong> comparable fuel sprays. Also two important<br />

observations that are reported by Verhoeven et al.<br />

are the auto-ignition position <strong>and</strong> the subsequent burning<br />

behavior. They state that the first visible emission corresponds<br />

roughly to the flame lift-<strong>of</strong>f distance during quasisteady<br />

state combustion phase observed later. And that<br />

this auto-ignited kernel progresses along the spray until<br />

it reaches the tip, after which no other development occurs,<br />

except for further penetration. The above presented<br />

numerical results <strong>of</strong> the Euler-Euler spray model together<br />

with the FGM combustion model is completely coherent<br />

with these experimental observations.<br />

Conclusions<br />

A 1D Euler-Euler spray model is implemented into 3D<br />

CFD (Fluent). This 3D spray model is validated with inert<br />

fuel spray penetration measurements <strong>and</strong> is able to predict<br />

spray lengths <strong>and</strong> shapes quantitatively well. It also<br />

<strong>of</strong>fers the advantage <strong>of</strong> a proper mesh resolution behavior<br />

(higher resolution gives better solutions), <strong>and</strong> is suitable<br />

for parallel computing.<br />

<strong>Combustion</strong> <strong>of</strong> the fuel spray is modeled with a tabulated<br />

chemistry approach (FGM). The manifold is created<br />

with igniting diffusion flame solutions. Important characteristics<br />

like auto-ignition <strong>and</strong> flame lift-<strong>of</strong>f are captured<br />

without explicitly modeling them, showing the generic nature<br />

<strong>and</strong> therefore the potential <strong>of</strong> the applied method.<br />

A first study with heptane as a surrogate for diesel


fuel shows promising results concerning spray formation,<br />

<strong>and</strong> subsequently auto-ignition <strong>and</strong> the settling <strong>of</strong> a lift-<strong>of</strong>f<br />

length.<br />

Outlook - Future Research<br />

In the future, direct validation with ignition delay<br />

times <strong>and</strong> flame lift-<strong>of</strong>f lengths will be done. And at the<br />

same time the influence <strong>of</strong> the preprocessing phase on the<br />

combustion behavior will be investigated. One can think<br />

about the choice <strong>of</strong> a progress variable <strong>and</strong> the applied<br />

FGM generation method.<br />

References<br />

[1] CHEM1D, A one-dimensional laminar flame<br />

code, Eindhoven University <strong>of</strong> Technology,<br />

http://www.combustion.tue.nl/chem1d.<br />

[2] Fluent 6.3 User’s Guide, September 2006.<br />

[3] C. Bekdemir. <strong>Numerical</strong> modeling <strong>of</strong> diesel spray<br />

formation <strong>and</strong> combustion. Master’s thesis, Eindhoven<br />

University <strong>of</strong> Technology, <strong>Combustion</strong> Technology,<br />

2008.<br />

[4] R.W. Bilger, S.H. Starner, <strong>and</strong> R.J. Kee. On reduced<br />

mechanisms for methane-air combustion in nonpremixed<br />

flames. <strong>Combustion</strong> <strong>and</strong> Flame, 80:135–<br />

149, 1990.<br />

[5] John E. Dec. A conceptual model <strong>of</strong> di diesel combustion<br />

based on laser-sheet imaging. SAE paper,<br />

(SAE 970873), February 1997.<br />

[6] DIPPR Design Institute for Physical Properties.<br />

http://dippr.byu.edu/.<br />

[7] ECN Engine <strong>Combustion</strong> Network.<br />

http://public.ca.s<strong>and</strong>ia.gov/ecn/index.php.<br />

[8] J.T. Farrell, N.P. Cernansky, F.L. Dryer, D.G. Friend,<br />

C.A. Hergart, C.K. Law, R.M. McDavid, C.J.<br />

Mueller, A.K. Patel, <strong>and</strong> H. Pitsch. Development<br />

<strong>of</strong> an experimental database <strong>and</strong> kinetic models for<br />

surrogate diesel fuels. SAE paper, (SAE 2007-01-<br />

0201), 2007.<br />

[9] U. Maas <strong>and</strong> S.B. Pope. Simplifying chemical kinetics:<br />

Intrinsic low-dimensional manifolds in composition<br />

space. <strong>Combustion</strong> <strong>and</strong> Flame, 88(1992):239–<br />

264, 1992.<br />

[10] Jean-Baptiste Michel, Olivier Colin, <strong>and</strong> Denis Veynante.<br />

<strong>Modeling</strong> ignition <strong>and</strong> chemical structure <strong>of</strong><br />

partially premixed turbulent flames using tabulated<br />

chemistry. <strong>Combustion</strong> <strong>and</strong> Flame, 152(2008):80–<br />

99, September 2007.<br />

[11] Jeffrey D. Naber <strong>and</strong> Dennis L. Siebers. Effects <strong>of</strong><br />

gas density <strong>and</strong> vaporization on penetration <strong>and</strong> dispersion<br />

<strong>of</strong> diesel sprays. SAE paper, (SAE 960034),<br />

February 1996.<br />

[12] N. Peters, G. Paczko, R. Seiser, <strong>and</strong> K. Seshadri.<br />

Temperature cross-over <strong>and</strong> non-thermal runaway at<br />

two-stage ignition <strong>of</strong> n-heptane. <strong>Combustion</strong> <strong>and</strong><br />

Flame, 128:38–59, 2002.<br />

[13] W.J.S. Ramaekers. The application <strong>of</strong> flamelet generated<br />

manifolds in modelling <strong>of</strong> turbulent partiallypremixed<br />

flames. Master’s thesis, Eindhoven Uni-<br />

6<br />

versity <strong>of</strong> Technology, <strong>Combustion</strong> Technology,<br />

2005.<br />

[14] R.D. Reitz <strong>and</strong> C.J. Rutl<strong>and</strong>. Development <strong>and</strong> testing<br />

<strong>of</strong> diesel engine cfd models. Prog. Energy Combust.<br />

Sci., 21:173–196, 1995.<br />

[15] Dennis L. Siebers. Scaling liquid-phase fuel penetration<br />

in diesel sprays based on mixing-limited vaporization.<br />

SAE paper, (SAE 1999-01-0528), March<br />

1999.<br />

[16] Satbir Singh, Rolf D. Reitz, <strong>and</strong> Mark P.B. Musculus.<br />

Comparison <strong>of</strong> the characteristic time (ctc), representative<br />

interactive flamelet (rif), <strong>and</strong> direct integration<br />

with detailed chemistry combustion models<br />

against optical diagnostic data for multi-mode<br />

di diesel engine. SAE paper, (SAE 2006-01-0055),<br />

April 2006.<br />

[17] R. Steiner, C. Bauer, C. Kruger, F. Otto, <strong>and</strong><br />

U. Maas. 3d-simulation <strong>of</strong> di-diesel combustion applying<br />

a progress variable approach accounting for<br />

complex chemistry. SAE paper, (SAE 2004-01-<br />

0106), March 2004.<br />

[18] G. Stiesch. <strong>Modeling</strong> engine spray <strong>and</strong> combustion<br />

processes. Springer, Berlin, 2003.<br />

[19] J.A. van Oijen. Flamelet-Generated Manifolds: Development<br />

<strong>and</strong> Application to Premixed Laminar<br />

Flames. PhD thesis, Eindhoven University <strong>of</strong> Technology,<br />

<strong>Combustion</strong> Technology, 2002.<br />

[20] Dean Verhoeven, Jean-Luc Vanhemelryck, <strong>and</strong><br />

Thierry Baritaud. Macroscopic <strong>and</strong> ignition<br />

characteristics <strong>of</strong> high-pressure sprays <strong>of</strong> singlecomponent<br />

fuels. SAE paper, (SAE 981069), February<br />

1998.<br />

[21] Philippe Versaevel, Paul Motte, <strong>and</strong> Karl Wieser. A<br />

new 3d model for vaporizing diesel sprays based<br />

on mixing-limited vaporization. SAE paper, (SAE<br />

2000-01-0949), March 2000.<br />

[22] Y.P. Wan <strong>and</strong> N. Peters. Application <strong>of</strong> the crosssectional<br />

average method to calculations <strong>of</strong> the<br />

dense spray region in a diesel engine. SAE paper,<br />

(SAE 972866), 1997.

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!