Está en la página 1de 30

SPECIAL ISSUE PAPER 97

Validation of engine combustion models against


detailed in-cylinder optical diagnostics data for
a heavy-duty compression-ignition engine
S Singh1*, R D Reitz1, M P B Musculus2, and T Lachaux2
1 Department of Mechanical Engineering, The University of Wisconsin, Madison, Wisconsin, USA
2 Sandia National Laboratories, Livermore, California, USA

The manuscript was accepted after revision for publication on 11 October 2006.

DOI: 10.1243/14680874JER02406

Abstract: Three different approaches for modelling in-cylinder compression-ignition engine


processes for partially premixed combustion modes are compared with experimentally
observed cylinder pressure and in-cylinder images of liquid- and vapour-fuel penetration,
ignition, combustion, and soot formation in an optically accessible heavy-duty direct injection
engine. A multi-dimensional computational fluid dynamics model for engine combustion,
KIVA-3V, served as a common platform into which three different combustion submodels
were integrated: (1) characteristic time combustion (KIVA-CTC); (2) representative interactive
flamelet (KIVA-RIF); and (3) direct integration using detailed chemistry (KIVA-CHEMKIN).
Three different engine operating strategies with significant premixing of fuel and air prior to
ignition were investigated: low-temperature combustion achieved by charge dilution, with fuel
injection either (1) early, or (2) late in the engine cycle, and (3) long ignition delay, high-
temperature combustion (i.e. no charge dilution) with fuel injection near top dead centre of
the piston stroke.
Comparison of simulated cylinder pressure and heat-release rates with the experimental
results shows that all the combustion submodels predict the cylinder pressures and heat-release
rates reasonably well, but predictions of in-cylinder phenomena were significantly different
among the submodels. The KIVA-CHEMKIN submodel predictions agree best with experimental
observations of the location of ignition sites and the spatial distribution of soot and OH. The
KIVA-RIF model, which uses global quantities to account for turbulence–chemistry interactions,
under-predicts the flame lift-off, while ignition sites and species distributions are broader than
observed experimentally. The KIVA-CTC submodel greatly over-predicts the spatial extent and
total amount of in-cylinder soot.

Keywords: OH planar laser-induced fluorescence, soot laser-induced incandescence, laser


diagnostics, representative interactive flamelet (RIF), CHEMKIN, characteristic time
combustion (CTC), combustion model

1 INTRODUCTION but also because they can complement experimental


data to yield a deeper understanding of the under-
Computational models are widely used for prediction lying chemical and physical processes responsible
and analysis of internal combustion engine perform- for experimental observations. Spatial resolution of
ance and emissions. Predictive models are useful models can range from relatively simple and fast
not only because they can replace time-consuming zero-dimensional simulations, e.g. reference [1], in
and costly research and development experiments, which the entire cylinder charge is assumed to be
ideally mixed, to a more complex multi-dimensional
* Corresponding author: Department of Mechanical Engineering, computational fluid dynamics (CFD) codes, e.g. ref-
The University of Wisconsin, 1500 Engineering Drive, Room 1011, erence [2], which solves for transient fluid flow, spray
Madison, WI 53706, USA. email: singh002@erc.wisc.edu dynamics, and combustion processes on a spatial

JER02406 © IMechE 2007 Int. J. Engine Res. Vol. 8

Downloaded from jer.sagepub.com at University of Birmingham on June 2, 2015


98 S Singh, R D Reitz, M P B Musculus, and T Lachaux

grid. With recent advances in computing technol- are implemented: (1) characteristic-time combustion
ogy, multi-dimensional models are becoming more (KIVA-CTC); (2) CFD integrated with detailed chem-
widely used, owing to their ability to accommodate istry (KIVA-CHEMKIN); and (3) representative inter-
complex engine geometries at high resolution and active flamelet (KIVA-RIF). All other submodels
with reasonable computational cost. For example, (including spray dynamics, turbulence, heat transfer,
the multi-dimensional computer code KIVA-3V etc.) are the same for each model formulation, so
coupled with a genetic algorithm was recently used that differences in the model predictions should be
by Wickman et al. [3] to optimize the combustion entirely owing to differences in combustion model-
chamber bowl geometry in a small-bore diesel ling approaches. The range of applicability of these
engine for reduced emissions and improved fuel con- models was validated against experimental data from
version efficiency. three low-load engine operating strategies having
Typically, model predictions are reliable only over significant premixing of fuel and charge gases before
a limited range of conditions, which depends upon the onset of combustion: low-temperature combus-
the formulation of the model and on the available tion (LTC) achieved by charge dilution, with fuel
experimental data against which it is validated. Over injection either (1) early, or (2) late in the engine
the last two decades, multi-dimensional models have cycle, and (3) long ignition delay, high-temperature
been extensively validated and used to predict com- combustion (HTC; i.e. no charge dilution) with fuel
pression-ignited (CI) and spark-ignited (SI) engine injection near top dead centre (TDC) of the piston
performance and emissions [4–6]. However, much of stroke.
this validation was limited to conventional engine In a previous publication by the current authors
combustion strategies. Recently, to improve fuel [8], the same three combustion submodels were
economy and reduce pollutant emissions, engine validated against cylinder pressure, apparent heat-
developers have been exploring alternative engine release rate (AHRR), exhaust NO emissions, two-
x
operational strategies that are very unlike conven- colour soot thermometry data, and high-speed soot
tional operating conditions, and computational luminosity imaging. In this paper, the model predic-
models are being challenged to predict over a wider tions are further compared with in-cylinder planar
range of engine operating conditions. laser diagnostic images of: (1) laser light elastically
The predictive capability of multi-dimensional scattering off the liquid-fuel spray; laser-induced
engine combustion models is limited by many fac- fluorescence of both (2) OH and (3) vapour-phase
tors, including the accuracy of their submodels. A fuel molecules; (4) laser-induced incandescence of
few of the more important submodels are for turbu- soot; and line-of-sight imaging of (5) ignition chemi-
lence, spray dynamics, and combustion. Full multi- luminescence and (6) soot luminosity.
dimensional models are usually validated using
global engine performance data (e.g. cylinder press-
ure and apparent chemical heat-release rate) and
2 EXPERIMENTAL SET-UP
exhaust emissions. Agreement of model-predicted
cylinder pressure and exhaust emissions with corre-
2.1 Engine and operating conditions
sponding experimental data does not, however,
guarantee that the model-predicted in-cylinder pro- An optically accessible, single-cylinder, direct-
cesses accurately reflect the true in-cylinder pro- injection (DI), four-stroke diesel engine based on a
cesses. To validate in-cylinder predictions of spray Cummins N-series production engine was used in
dynamics, combustion, and pollutant formation, a this investigation. Figure 1 shows a schematic of the
more comprehensive comparison with in-cylinder optical engine, and Table 1 summarizes the specifi-
experimental observations is required. This paper cations of this engine. A complete description of the
presents a comparison of multi-dimensional com- engine and its specifications is available in references
puter model predictions against in-cylinder obser- [9] to [11].
vations of liquid- and vapour-fuel penetration, The engine cylinder bore is 139.7 mm and the
ignition, combustion, and pollutant formation ob- piston stroke is 152.4 mm, yielding a displacement
tained from multiple imaging diagnostics in an of 2.34 l per cylinder. To provide optical access, the
optically accessible engine. The KIVA-3V release 2 engine is equipped with an extended piston and flat
code developed at the Los Alamos National Lab- piston-crown window. As shown in Fig. 1, a station-
oratory [7] and modified at the University of ary mirror placed below the piston window provides
Wisconsin provides a multi-dimensional model a view of the combustion chamber to the camera
framework into which three combustion submodels imaging system. Another window replaces one of the

Int. J. Engine Res. Vol. 8 JER02406 © IMechE 2007

Downloaded from jer.sagepub.com at University of Birmingham on June 2, 2015


Validation of engine combustion models 99

Table 1 Engine and injector specifications


Engine base type Cummins N-14, DI diesel
Number of cylinders 1
Number of intake valves 2
Number of exhaust valves 1*
Combustion chamber Quiescent, DI
Swirl ratio 0.5
Bore×stroke (cm) 13.97×15.24
Bowl width, depth (cm) 9.78, 1.55
Displacement (l) 2.34
Connecting rod length (cm) 30.48
Geometric compression ratio 11.2:1
Simulated compression ratio 16:1
Fuel injector type Common-rail, pilot-valve
actuated
Cup (tip) type Mini sac
Number of holes 8, equally spaced
Spray pattern included angle 152°
Rail pressure (bar) 1200 or 1600
Nozzle orifice diameter (mm) 0.196
Nozzle orifice L/D 5

* In this optically accessible diesel engine, one of the two


exhaust valves of the production cylinder head was replaced
by a window and periscope.

different combustion environments owing to the


extent of dilution and the relative differences be-
tween duration of injection (DOI) and ignition delay
(ID) [8]. The first condition in Table 2 is characterized
as HTC because the charge gas is undiluted air (21
per cent oxygen by volume), and the resulting peak
adiabatic–stoichiometric flame temperature is rela-
tively high (2700 K). For this condition, fuel injection
was initiated near TDC and a relatively long ID was
achieved by using cool engine intake temperatures.
This condition represents a typical example of low-
load conventional diesel combustion with a large
premixed burn and relatively little mixing-controlled
combustion (MCC).
Fig. 1 Engine schematic with laser/imaging diagnostic The two remaining conditions are characterized as
set-up (top), and top view of the optical piston LTC because the oxygen volume fraction in the
and piston bowl, showing orientation of the charge gas was reduced to 12.7 per cent by dilution,
eight diesel jets, the laser sheet, and the window so that the peak adiabatic–stoichiometric flame tem-
in the piston bowl wall peratures were reduced to approximately 2200 K (see
Table 2). In a production engine, exhaust gas recircu-
two exhaust valves in the cylinder head to provide lation (EGR) typically provides dilution to reduce the
imaging access to the squish region above the piston, intake oxygen volume fraction. In the optical engine
but this perspective was not used in the current of the current study, however, the fuel injector was
study. Additional windows located around the top of skip-fired, with fuel injected every tenth engine cycle
the cylinder allow optical access for laser imaging to avoid thermal loading problems, so EGR was not
diagnostics. As shown in the bottom of Fig. 1, the an option (engine speed was maintained between
piston bowl wall was fitted with an optical window fired cycles by an electric dynamometer connected
that matched the bowl contours to allow laser sheet to the crankshaft). Rather, the intake stream was
access into the combustion bowl of the piston along diluted with a metered flow of nitrogen gas to achieve
the axis of one of the eight fuel jets. 12.7 per cent oxygen by volume.
Three different operating conditions were selected Two different fuel injection strategies were selected
as representative of the three modes of long ignition for the LTC operating conditions: early injection,
delay partially premixed combustion, as described in starting at −22° ATDC, and late injection, starting at
Table 2. The three operating conditions produce TDC (see Table 2). Note that for both of these LTC

JER02406 © IMechE 2007 Int. J. Engine Res. Vol. 8

Downloaded from jer.sagepub.com at University of Birmingham on June 2, 2015


100 S Singh, R D Reitz, M P B Musculus, and T Lachaux

Table 2 Test matrix


High temperature, Low temperature, Low temperature,
long ID early injection late injection

Speed (r/min) 1200 1200 1200


IMEP (bar) 4.5 3.9 4.1
Injection pressure (bar) 1200 1600 1600
Intake temperature (°C) 47 90 70
BDC temperature (°C) 62 92 78
Intake pressure (kPa) 192 214 202
TDC motored temperature (K) 800 870 840
TDC motored density (kg/m3) 22.3 22.9 22.5
Peak adiabatic flame temperature (K)* 2700 2256 2164
SOI (°ATDC) −5 −22 0
Injection quantity (mg) 61 56 56
DOI (CAD) 10 7 7
O concentration (vol. %) 21 12.7 12.7
2
IMEP, indicated mean effective pressure; BDC, before dead centre; SOI, start of injection; DOI, duration of
injection; ATDC, after top dead centre; CAD crank angle degree. The intake temperatures and pressures
shown are significantly higher than typical values for production engines because of the low compression
ratio of the optical engine. Owing to compromises necessary to implement optical access in the engine, the
geometric compression ratio of the optical engine is only 11.2:1, compared with 16:1 in the production
engine. As a result, elevated intake-air temperatures and pressures are required to yield charge conditions
near TDC that are typical of those in the production engine.
* The peak adiabatic flame temperatures for each condition were calculated for a stoichiometric mixture of
fuel and charge gas adiabatically compressed to the peak cylinder pressure, as described in reference [11].

conditions, the fuel injection pressure was increased angle of 152° (14° down-angle from the firedeck).
to 1600 bar, compared with 1200 bar for the HTC The eight fuel orifices are equally spaced and have
condition. The higher injection pressure was selected nominal diameters of 0.196 mm. The rate of injection
to promote faster pre-combustion mixing, and to was measured using a momentum-based meter, as
increase the rate of fuel delivery. With faster fuel described in reference [11]. As described above, the
delivery, the DOI was shortened such that the ID fuel was injected in only one out of ten cycles to
became comparable with, or longer than, the DOI. minimize thermal loading on the optical engine. The
Note that the late-injection condition is somewhat properties of the diesel fuel used in the current study
similar to the modulated by kinetics (MK) combus- are given in reference [14].
tion operating condition suggested by Kimura et al.
[12, 13], but with two significant differences: first, 2.3 Laser sheets and camera set-up
the swirl ratio (0.5) is not as high as in typical MK
Two different laser systems were used in this study.
combustion conditions (3 or more); second, the
First, an optical parametric oscillator, pumped by the
period between the end of injection (EOI) and start
third harmonic of an Nd:YAG laser generated 10 Hz,
of combustion (SOC), termed ‘ignition dwell’, is
10 ns pulsed laser radiation near 568 nm, which was
smaller than that of the previous MK combustion
frequency doubled to near 284 nm at 17 mJ per pulse
implementations. In spite of these differences in
for the fluorescence diagnostics. Second, a 10 Hz
engine controlling parameters, a major part of the
frequency doubled Nd:YAG laser generating 10 ns
heat release is still expected to be controlled by kin-
pulses at 532 nm with 35 mJ per pulse. Both beams
etics (not mixing), similar to the MK combustion
were converted into thin (less than 1 mm thick)
operating condition of Kimura et al. [12, 13].
sheets using a combination of cylindrical and spheri-
cal lenses, as shown in Fig. 1. The two sheets were
2.2 Fuel and fuel injector
overlapped and directed through a cylinder-wall
The engine is equipped with a non-production, high- window and into the combustion chamber at a slope
pressure, electronically controlled, common-rail fuel of 14° from the horizontal and parallel to the axis of
injector. Specifications for the fuel injector are the fuel jet that was inline with the cylinder-wall and
included in Table 1. This injector uses a solenoid- piston-bowl-wall windows. After being clipped by the
actuated pilot valve and a pressure-balanced needle window aperture, the sheets were about 30 mm wide
to control fuel delivery. It is capable of multiple injec- inside the combustion chamber. For a portion of the
tions at up to 2000 bar fuel rail pressure. For the piston stroke (crank angle range of 22°–30° from
experiments presented here, an eight-hole, mini-sac TDC) the top surface of the bowl-wall window inter-
injector cup (tip) was employed, having an included sected the nominal jet axis, along which the laser

Int. J. Engine Res. Vol. 8 JER02406 © IMechE 2007

Downloaded from jer.sagepub.com at University of Birmingham on June 2, 2015


Validation of engine combustion models 101

sheets were normally aligned. To avoid interference, comparison with model predictions in the Results
from 22° to 26° from TDC, the laser sheets were section.
moved vertically down by 3 mm from the nominal
fuel-jet axis, and from 27° to 30° from TDC, the sheets 2.4 Optical diagnostics
were moved vertically up by 3 mm from the nominal
jet axis. After the top of the piston-bowl-wall window The in-cylinder spray, combustion, and pollutant-
had cleared the path of the laser sheets, the sheet formation processes were visualized using six
elevation was returned to the elevation of the nom- optical/imaging diagnostics. The diagnostics are
inal jet axis. briefly listed in Table 3, and more detailed (and
Six different optical/imaging diagnostics were lengthy) descriptions of each diagnostic can be
combined in various simultaneous pairings. Either found in the references cited in Table 3. Although
naturally occurring luminescence or laser-induced detailed descriptions of the optical diagnostics are
emission for each diagnostic was recorded by a not provided here, guidelines to aid the reader to
pair of cameras viewing the combustion chamber interpret the diagnostic images are provided in the
through the piston-crown window. After reflecting off Results section.
a stationary mirror, the emitted light was chromati-
cally separated and directed to the two gated, inten-
sified, ultraviolet (UV)-sensitive cameras by a 450 nm 3 COMPUTATIONAL MODELS
cut-off dichroic beamsplitter, which reflected UV
light and transmitted visible light (see Fig. 1). The three engine operating conditions were simu-
Camera 1, which received long-wavelength (visible) lated using the KIVA-3V release 2 [7], implemented
light, was equipped with a Nikkor 105 mm f/2.5, glass with three different combustion submodels. All
camera lens, while camera 2, which received short- other submodels, which included the hybrid Kelvin
wavelength (UV) light, was equipped with a UV- Helmholtz–Rayleigh Taylor (KH–RT) spray break-up
Nikkor 105 mm f/4.2 camera lens. In addition, a filter model [23] and the re-normalized group (RNG) k–e
pack appropriate for each diagnostic was placed in turbulence model developed by Han and Reitz [24],
front of the lenses for each camera. Owing to speed were identical for all three implementations.
limitations of the pulsed lasers (10 Hz), only one
image could be acquired for each engine cycle. 3.1 KIVA-CTC submodel
Accordingly, to generate a statistical data set, images This widely used submodel is attractive owing to its
were acquired at each crank angle of interest in sets relative simplicity and its validated accuracy for HTC
of 12, from 12 different engine cycles. Although the conditions [25]. It has the shortest computational
IMEP for these operating conditions was relatively time of the three combustion submodels of the cur-
stable, cycle-to-cycle variation of many of the imaged rent study.
in-cylinder phenomena was significant, so that com-
parison of instantaneous images with model predic-
3.1.1 Auto-ignition
tions was difficult. Ensemble averaging of the images
was deemed inappropriate, because it blurred the Ignition is predicted by the Shell model [26], which
observations and the resulting averaged image was uses a simplified reaction mechanism to simulate the
not representative of the features observed in instan- auto-ignition of hydrocarbon fuels. The mechanism
taneous images. Instead, the full experimental image consists of five generic species (fuel, oxygen, radicals,
set for each diagnostic was carefully surveyed, and intermediate species, and branching agents) and
the image most representative of the most frequently eight generic reactions with initiation, propagation,
observed in-cylinder features was selected for branching, and termination steps. Details of the

Table 3 Optical diagnostics


Diagnostic Measurement

Laser-elastic (Mie) scattering (LMS) [15] Planar liquid-fuel spray imaging


Planar laser-induced fluorescence of OH (OH-PLIF) [16] Planar imaging of OH radicals
Broadband planar laser-induced fluorescence (BB-PLIF) [17] Planar imaging of unburned fuel and combustion
intermediates
Planar laser-induced incandescence of soot (soot-LII) [18–20] Planar soot distribution
Chemiluminescence imaging [19, 21] Line-of-sight ignition and combustion locations
Soot luminosity [19, 22] Line-of sight distribution of hot, radiating soot

JER02406 © IMechE 2007 Int. J. Engine Res. Vol. 8

Downloaded from jer.sagepub.com at University of Birmingham on June 2, 2015


102 S Singh, R D Reitz, M P B Musculus, and T Lachaux

model implementation can be found in Kong and experiments [31]. While this reduced n-heptane
Reitz [27]. The standard ignition model constants [4] mechanism is a suitable model for the combustion
were tuned slightly to better predict the ID, but were reactions of diesel fuel, the physical properties of
thereafter kept the same for all three operating con- heptane (e.g. density, distillation) are unlike those of
ditions. The ignition model was used only for the diesel fuel. The physical properties of heavier hydro-
low-temperature chemistry. After ignition (i.e. when carbons (dodecane, tetradecane, hexadecane, etc.)
the local gas temperature became greater than are more similar to diesel fuel, and any of these could
1100 K), the main combustion submodel described be used as a surrogate for diesel fuel. In this investi-
below simulated the subsequent chemistry. gation, tetradecane was chosen as a surrogate to
model the non-chemical-kinetic processes of fuel
3.1.2 Main combustion injection, spray break-up, and evaporation.
The CHEMKIN chemistry solver [33] was inte-
The KIVA-CTC model calculates the equilibrium con-
grated into the KIVA-3V code to solve the n-heptane
centration of each species and the corresponding
reaction mechanism. The convective and diffusive
laminar and turbulent characteristic times to deter-
transport is modelled using the RNG k–e turbulence
mine species conversion rates [27]. Seven major
model, and sub-grid-scale turbulence–chemistry
combustion species are considered: fuel, O , N , CO ,
2 2 2 interaction is not considered (i.e. the mixture is
H , H O, and CO. The time rate of change of the
2 2 assumed to be homogeneous within each compu-
concentration of species k is given as [4]
tational cell). Further details of model formulation
dY Y −Y* and implementation can be found in references [34]
k =− k k (1)
dt t and [35].
c
where Y is the concentration of species k, Y* is the 3.2.2 NO and soot
k k x
local and instantaneous thermodynamic equilibrium
concentration, and t is the characteristic time (1/e) In the KIVA-CHEMKIN submodel, NO emissions
c x
are simulated by a reduced NO mechanism that
to reach equilibrium. The characteristic time is cal- x
culated as the sum of a kinetic laminar time-scale was derived from the Gas Research Institute (GRI)
and a turbulent mixing time-scale, as described in NO mechanism [36]. Consequently, four additional
species (N, NO, NO , N O) and nine reactions are
reference [4]. 2 2
added to the n-heptane chemistry mechanism [37].
The reactions account only for thermal NO forma-
3.1.3 NO and soot
x x
tion; prompt NO chemistry is not modelled. Three
x
In the KIVA-CTC submodel, NO emissions are pre- of the nine reactions are the extended Zeldovich
x
dicted using the extended Zeldovich mechanism reactions [28], and the other six account for thermal
[28], and soot emissions are predicted by a phenom- NO formation through NO and N O reactions path-
x 2 2
enological soot model that uses the Hiroyasu soot ways. The NO and N O reactions were included
2 2
formation model [29] and the Nagle–Strickland- because LTC conditions favour formation of these
Constable (NSC) soot oxidation model [30]. In the two species more than HTC conditions do [28].
formation model, the parent fuel molecule is the As in the KIVA-CTC submodel, soot emissions are
species from which soot forms. predicted by the Hiroyasu soot formation model [29]
and the NSC soot oxidation model [30]. However, in
3.2 KIVA-CHEMKIN submodel the KIVA-CHEMKIN submodel, acetylene (C H ) is
2 2
used as the soot formation species rather than the
This is the most chemically detailed combustion
parent fuel molecule because acetylene is a more
model implemented in the current study, and re-
appropriate soot precursor species, and is part of
quired the longest computational times.
the reduced n-heptane mechanism (acetylene is not
present in the KIVA-CTC chemistry). Other than its
3.2.1 Ignition and main combustion
use of acetylene as a precursor species, the soot
A skeletal reaction mechanism for n-heptane with 30 model is not part of the n-heptane main combustion
species and 65 reactions [31] simulated the diesel mechanism. That is, the soot formation and oxi-
fuel chemistry. The skeletal mechanism retains the dation reactions are calculated separately from the
main features of a larger, more detailed mechanism main combustion and NO chemistry calculations.
[32] and includes reactions of polycyclic aromatic The present soot model has been validated under
hydrocarbons. The mechanism was validated using constant volume conditions and predicts the sooting
constant-volume ID data and engine combustion tendency of diesel sprays well [35].

Int. J. Engine Res. Vol. 8 JER02406 © IMechE 2007

Downloaded from jer.sagepub.com at University of Birmingham on June 2, 2015


Validation of engine combustion models 103

3.3 KIVA-RIF submodel study adopts the model proposed by Demoulin and
Borghi [40]. The detailed model equations are de-
This model is intermediate in terms of computa-
scribed in references [40] and [41]. The fuel chemis-
tional cost between the KIVA-CTC and KIVA-
try was modelled using the same reduced n-heptane
CHEMKIN models. The flamelet equations provide a
mechanism as described above for the KIVA-
provision for considering sub-grid-scale turbulence,
CHEMKIN submodel.
although this feature is not resolved locally because
sub-grid-scale effects are simulated using domain-
3.3.2 NO and soot
averaged turbulence quantities. x
The NO and soot models used were the same as for
x
3.3.1 Ignition and main combustion the KIVA-CHEMKIN sub-model.
The flamelet concept is based on the assumption
3.4 Computational grid and initial conditions
that the smallest turbulent time and length scales are
much larger than the corresponding chemical scales, The piston geometry and computational grid are
and there exists a locally undisturbed thin sheet shown in Fig. 2. The diesel injector has eight equally
where chemical reaction occurs [5]. This sheet is spaced nozzle holes, so the combustion chamber was
treated as an ensemble of counter-flow diffusion represented by a 45° sector mesh with periodic
flames, called flamelets. The advantage of the above boundary conditions. All major geometric dimen-
treatment is that all reacting scalars can be uniquely sions of the combustion chamber – including cylin-
related to the mixture fraction variable, Z, which der bore, stroke, bowl dimensions, and squish height
describes the local fuel-to-air ratio for non-premixed – were exactly replicated in the computational grid.
combustion, such as in conventional diesel engines. An additional artificial grid in the crevice region
The local flamelet structure is described by the accounted for the additional volume occupied by the
flamelet equations for species and enthalpy assum- ring land, valve, and optical window crevices. In this
ing unity Lewis number [38] as way, the geometric compression ratio in the model
was the same as for the actual engine. The mesh was
A B
∂Y x ∂2Y v̇ W
k= k + k k (2) created using the standard KIVA-3V pre-processor.
∂t 2 ∂Z2 r
The mesh was composed of about 80 000 compu-

A B
∂T x ∂2h N x ∂2Y v̇ W ∂P tational cells at BDC with 1.2×1.2×1.2 mm cell size
c = −∑h k+ k k − (3)
p ∂t 2 ∂Z2 k 2 ∂Z2 r ∂t near the piston bowl wall. The cell size is smallest
k=1 near the cylinder axis (Fig. 2) and largest at the cylin-
where x is the scalar dissipation rate defined as der wall (1.2×1.5×1.2 mm).
The computations were started from intake valve
A B
∂Z 2
x=2D (4) closure (IVC) with a uniform mixture distribution in
∂x
the cylinder. The swirl was initialized based on wheel
where D is the diffusion coefficient of the mixture flow velocity profile [2]. For low-swirl heavy-duty DI
fraction. engines, the turbulent kinetic energy near TDC is not
The CFD code provides the necessary parameters very sensitive to initialization of the turbulent quan-
to the flamelet code for solving the flamelet equa- tities at IVC [42]. A low value of turbulent kinetic
tions, and the species mass fractions obtained from
the flamelet code are passed on to the CFD code.
Appropriate modeling of the scalar dissipation rate
accounts for the transient effects of the turbulent
flow on chemistry. The turbulent mean values of the
reactive scalars can be evaluated once the con-
ditional scalar dissipation rate and the probability
density function (PDF) of the mixture fraction, P̃(Z),
are known, as described in references [38] and [39].
Turbulent combustion models originally devel-
oped for gaseous flames have been applied to simu-
late spray combustion processes without considering
the effects of vaporization on small-scale turbulent
mixing and turbulent combustion. In order to ac- Fig. 2 A 45° sector computational grid at 0° ATDC
count for the effects of vaporization, the present (80 000 cells at BDC)

JER02406 © IMechE 2007 Int. J. Engine Res. Vol. 8

Downloaded from jer.sagepub.com at University of Birmingham on June 2, 2015


104 S Singh, R D Reitz, M P B Musculus, and T Lachaux

energy (<1 per cent of mean piston speed) and a 4.1.2 Liquid- and vapour-fuel distributions
small length scale (10 per cent of cylinder bore) were
Figure 4 shows a comparison of experimentally
assumed at IVC to initialize the turbulent flow field.
observed and model-predicted liquid-fuel and
The pressure at IVC was initialized based on the
vapour-fuel distributions. The left column shows the
experimental pressure trace, and the IVC tempera-
experimental images, and the remaining three col-
ture was calculated assuming isentropic com-
umns show the model predictions, as labelled for
pression from BDC to IVC. The injection rate shape
each of the three combustion submodels. The per-
was taken from experimental data, acquired using a
spective for the images is an upward view through
momentum-based rate-of-injection meter [11].
the large piston-crown window. The field of view
does not encompass the entire combustion chamber,
but rather it focuses on one of the eight jets, at the
4 RESULTS AND DISCUSSION
3 o’clock position from the camera’s perspective. The
white curve on the right of each experimental image
4.1 HTC condition
represents the inner surface of the piston bowl,
4.1.1 AHRR and cylinder pressure which is about 49 mm away from the injector tip,
For the HTC condition, the 8.75 CAD ID is relatively indicated by a white dot on the left side of each
long, and rapid combustion of a significant quantity image. The number in the bottom right corner of
of premixed fuel causes a high peak AHRR with each experimental image is the crank angle at which
mostly premixed combustion, as shown in Fig. 3. the image was taken. For the model, recall that only
Note that prior to calculating the experimental a single fuel jet is simulated within a 45° sector grid.
AHRR, the pressure data were smoothed using a The model images thus show a 45° slice of the full
Fourier series low-pass filter to remove significant 360° view of the bowl region of the combustion
acoustic ringing in the pressure data [11]. Although chamber. The curved edges on the right side of the
this filtering technique does not affect the integrated model images are the inner surface of the piston
heat release (total area under the AHRR curve), the bowl, and the apex of the triangular domain at the
magnitude of the peak AHRR is reduced somewhat left of the model images is the location of the injec-
and the sharp features in the AHRR data smoothed. tor tip.
Therefore, the model-predicted peak AHHR cannot Representative, instantaneous images from two
be directly compared with the experimental peak simultaneous optical diagnostics are shown in the
AHRR. However, the KIVA-CHEMKIN and the KIVA- left column of Fig. 4, with laser-light Mie scattering
RIF models clearly predict a shorter ID than the off liquid fuel droplets (LMS) false-coloured blue and
experiment, while the KIVA-CTC model predicts a planar BB-PLIF of fuel molecules false-coloured
longer ID. Furthermore, the KIVA-RIF model pre- green. The relative gains of the camera image intensi-
dicted a higher peak AHRR compared with KIVA-CTC fiers are also noted at the top of each image. Model
and the KIVA-CHEMKIN models. predictions of liquid fuel are shown in Fig. 4 in two
ways. First, the modelled liquid-fuel parcels, each of
which represents a group of droplets, are represented
by circles with equal diameters, as shown in Fig. 4
(small red circles). This rendering technique is some-
what lacking, however, because the circle distri-
bution is not clearly quantitative of the local liquid-
fuel mass. Second, as a complementary visualization,
the local liquid-fuel ‘density’ in each cell (mass of
liquid fuel divided by the cell volume) along the
plane of the nominal jet axis is also presented in the
model predictions of Fig. 4, using a blue colourmap.
Because of the finite grid scale, the post-processing
scheme described above blurs the blue-colour
liquid-fuel density distribution somewhat, so that it
is slightly broader than the distribution of red par-
Fig. 3 Experimentally measured and modelled cylin- cels. Model predictions of vapour-fuel density in the
der pressure and AHRR, and measured mass plane of the nominal jet axis are also shown in the
rate of injection profile for the HTC condition three model columns of Fig. 4, false-coloured green.

Int. J. Engine Res. Vol. 8 JER02406 © IMechE 2007

Downloaded from jer.sagepub.com at University of Birmingham on June 2, 2015


Validation of engine combustion models 105

Fig. 4 Experimentally observed LMS (blue) and BB-PLIF (green) and model-predicted liquid-
fuel (red circles and blue colour) and vapour-fuel mass fraction (green) for the HTC
condition

The liquid fuel (LMS, blue) first appears in the this BB-PLIF diagnostic is only capable of detecting
experimental image at −5° ATDC, coincident with fuel near the leading edge of the jet, where attenu-
the actual start of injection. Up to −1° ATDC, the ation is minimal, as described in reference [17].
leading edges of the LMS and the BB-PLIF images Therefore, at 1° ATDC and later, the BB-PLIF signal
remain coincident, indicating that liquid and vapour (green) represents a visualization of the leading edge
phases of the fuel penetrate almost equally. At 1° of the vapour-fuel jet only.
ATDC, the leading edge of the BB-PLIF distribution The liquid- and vapour-fuel penetration distances
(green) in the experimental image penetrates farther predicted by all three models agree reasonably well
than the leading edge of the LMS signal (blue). In with the strongest LMS signal and BB-PLIF per-
this image, the downstream BB-PLIF signal arises imeter. In many of the experimental LMS images,
primarily from vapour-phase fuel, since the LMS however, some weak elastic-scattering signal is
image shows that significant liquid fuel does not exist present downstream of the bright leading edge of the
so far downstream. Also note that the intensity of the liquid-fuel-spray tip apparent in Fig. 4. The authors
BB-PLIF in the experimental image at 1° ATDC is believe that this weak signal is from pockets of un-
very weak in the upstream direction (towards the vaporized fuel, probably the heavy ends of the multi-
injector). This occurs because the vapour fuel component fuel, and these pockets of liquid fuel are
strongly attenuates the UV laser light, so that up- also observed for other operating conditions. Only
stream regions of the jet are not illuminated as well, KIVA-CTC and KIVA-CHEMKIN predict small pockets
and therefore yield a smaller fluorescence signal. of liquid fuel (blue) in the regions downstream of the
Vapour fuel is certainly present farther upstream, but strong LMS perimeter. As will be discussed in the

JER02406 © IMechE 2007 Int. J. Engine Res. Vol. 8

Downloaded from jer.sagepub.com at University of Birmingham on June 2, 2015


106 S Singh, R D Reitz, M P B Musculus, and T Lachaux

next section, these differences in the predicted extent are readily apparent in the model-predicted fuel dis-
of fuel vaporization are caused by differences in tributions in Fig. 4, especially in the 1° and 3° ATDC
low-temperature chemical heat release predicted by images for the KIVA-CTC and KIVA-CHEMKIN
the various models. At 3° ATDC, both the KIVA- models. When these isolated pockets eventually
CHEMKIN and the KIVA-RIF models predict de- vaporize, the LL suddenly snaps back to the furthest
pletion of vapour fuel at the periphery of the jet downstream liquid fuel remaining, which is often sig-
owing to the start of second-stage combustion (SSC). nificantly further upstream because of the ‘break’ in
The penetration of liquid and vapour fuel for both the liquid-fuel distribution. The sudden decreases in
the experiments and models is plotted in Fig. 5. the model-predicted LL plotted in Fig. 5 correspond
For the experiments, the liquid- and vapour-fuel to these isolated liquid pocket vaporization events,
penetration are quantified as the farthest down- which occur multiple times for the KIVA-CTC and
stream distance that either the elastic-scattering or especially the KIVA-CHEMKIN model, causing the
fluorescence signal is detected, using an appropriate apparent oscillation in the LL late in the injection
intensity threshold [15, 43]. The model-predicted event. Finally, isolated pockets of liquid fuel are also
liquid-fuel penetration is defined as the distance apparent in many of the experimental LMS images.
from the injector to the furthest liquid particle from Recall that the single, instantaneous LMS images are
the injector, and the vapour penetration is measured acquired from different cycles, so it is not possible to
from the model images using a threshold of 10 per observe evolution and evaporation of isolated liquid-
cent of the maximum vapour-fuel concentration. fuel pockets in the experimental data. However, such
Initially, the model-predicted liquid-fuel penetration, pocket-vaporization events would be manifested by
or liquid length (LL), matches very well with the increased cycle-to-cycle variation in the experimen-
measurements (Fig. 5). Later in the injection event, tal data (black dots in Fig. 5), which clearly increases
all three model solutions frequently predict pockets at about the same time that the models show liquid-
of liquid fuel at the tip of the spray that are isolated pocket vaporization events.
from the rest of the more continuous upstream The model-predicted LL is also affected by the
liquid-fuel spray. These isolated liquid-fuel pockets amount and location of low-temperature ignition
sites. First, regarding the KIVA-CTC model, the pre-
dicted ID for SSC is too long, and the Shell ignition
model predicts little heat release during low-
temperature ignition, so that temperatures in the jet
remain relatively low. The liquid fuel, including
isolated pockets at the tip of the jet, can therefore
survive much longer, so that the KIVA-CTC model
greatly over-predicts the LL. Note that around 4°
ATDC, an isolated pocket of liquid fuel at the tip of
the KIVA-CTC jet finally vaporizes, resulting in a
sudden and drastic drop in LL (Fig. 5). Regarding the
KIVA-CHEMKIN and KIVA-RIF models, the reduced
n-heptane chemistry mechanism predicts more
significant low-temperature heat release early in
ignition than does the Shell model. The increase in
temperature resulting from the low-temperature heat
release aids vaporization of liquid fuel, especially at
the tip of the jet, resulting in a more accurate predic-
tion of the LL. For the KIVA-RIF model, however, a
large portion of the liquid-fuel spray vaporizes sud-
denly at 1° ATDC, leading to sudden drop in LL as
shown in both Figs 4 and 5. Furthermore, the LL
predicted by KIVA-RIF remains short, and does not
oscillate, as it does for the KIVA-CHEMKIN and KIVA-
Fig. 5 Experimentally observed and model-predicted CTC models. This is because at 1° ATDC, KIVA-RIF
liquid length (LL), vapour penetration, mass predicts cool-flame (CF) ignition throughout the jet
rate of injection profile, and in-cylinder liquid- cross-section, which helps to vaporize the isolated
fuel fraction for the HTC condition liquid pockets in the mid- and downstream region

Int. J. Engine Res. Vol. 8 JER02406 © IMechE 2007

Downloaded from jer.sagepub.com at University of Birmingham on June 2, 2015


Validation of engine combustion models 107

of the jet. On the other hand, for the KIVA-CHEMKIN [22]. (3) Finally, soot may form shortly after the start
mode, these isolated liquid pockets survive for a of HTC, and the resulting soot luminosity is typically
longer time. The penetration and subsequent evap- another several orders of magnitude stronger than
oration of the liquid pockets result in oscillation in the high-temperature chemiluminescence [21, 22].
LL. Clearly, for this long ID operating condition, it is In the current study, camera 1 in Fig. 1 imaged all
important to predict the temperature rise and the three components of natural emission, together.
resulting effects on liquid-fuel vaporization owing to The emission during ignition reactions is dominated
ignition chemistry. Such a significant ignition chem- by chemiluminescence, but can be very weak, requir-
istry influence on liquid-fuel vaporization is in con- ing a very high camera gain. As other, brighter com-
trast to conventional diesel combustion, for which ponents of natural emission arose, such as soot
the liquid-fuel vaporization is primarily driven by luminosity, a much lower effective intensifier gain
entrainment of hot ambient gases [43]. Finally, the was used, so that the weaker emission from chemi-
liquid fuel in the experimental image was last luminescence was very dim. For all natural lumin-
observed at 4° ATDC, but the models predict a small osity images, the camera gate width was set to 70 ms,
amount of residual liquid fuel near the nozzle and which is equivalent to about one-half CAD at
along the jet axis for a short time after the end of 1200 r/min (the recorded natural emission is inte-
injection. At the end of injection near 5° ATDC, the grated over one-half CAD).
models predict that about 4 per cent of the total fuel The accuracy of model predictions of the onset
injected remains as liquid, which subsequently and progression of low- and high-temperature
vaporizes almost completely by 8° ATDC, as shown ignition sites is validated by comparison of chemi-
in Fig. 5. This late-vaporizing liquid fuel, if present luminescence images with the model-predicted
in the experiments, surely would have been detected cylinder gas temperature along the plane of the
by LMS. nominal fuel-jet axis. This comparison is justified
In the experiments, prior to the start of SSC when because the reactions that produce chemilumin-
the AHRR increases rapidly (near 4° ATDC), the only escence emission accompany exothermic ignition
likely sources of broadband fluorescence from UV and combustion reactions that increase the local gas
laser excitation are fuel molecules. Therefore, fluo- temperature. While the KIVA-CHEMKIN and KIVA-
rescence in the BB-PLIF images up to 4° ATDC can RIF submodels predict significant increases in gas
be interpreted entirely as fuel fluorescence, and its temperature during LTC, the Shell ignition model in
penetration is compared with the model predictions the KIVA-CTC submodel yields little change in gas
in Fig. 5. All of the models predict penetration of the temperature until the very end of LTC, when the gas
leading edge of vapour fuel quite well. Of course, temperature increases rapidly. Consequently, tem-
after the start of SSC, the KIVA-CHEMKIN and KIVA- perature is a poor indicator of the progress of LTC
RIF models show a decreasing vapour fuel penetra- for KIVA-CTC. However, the model does predict a
tion owing to consumption of fuel near the bowl gradual accumulation of generic (representative of
wall. This occurs after the tip of the vapour-fuel jet alkyl radical ‘R’) ignition radical species, which do
impinges on the piston bowl wall in the experimental indicate the locations of low-temperature ignition
images, so that reliable vapour-fuel distribution sites. Accordingly, for KIVA-CTC, the ignition radical
cannot be discerned from the BB-PLIF images. The species of the Shell model, rather than local gas tem-
KIVA-CTC model predicts longer SSC ID, so it pre- perature, is used as an indicator of the location of
dicts vapour fuel persisting near the piston bowl wall ignition sites. Finally, it is important to recognize that
until about 8° ATDC. the recorded emission in the chemiluminescence
images is integrated along the camera’s line of sight,
so that it does not directly correspond to the model-
4.1.3 Onset and development of auto-ignition sites
predicted planar distributions of either temperature
Naturally occurring luminous emission from engine or ignition radical species.
combustion can be divided into three sources. The first column on the left in Fig. 6 shows simul-
(1) During early ignition stages, species that include taneous experimental images of LMS from liquid-
CH, C , and formaldehyde (CH O) emit weak, rela- fuel droplets (blue) and ignition chemiluminescence
2 2
tively broadband chemiluminescence emission from (green). The second column shows liquid-fuel (red
about 360 to about 560 nm [19, 21]. (2) Later, high- particles and blue colour) and mass fraction of
temperature reactions produce greater amounts of radicals (green) for the KIVA-CTC model, and the
OH and CH, yielding an increase in the chemilumi- third and fourth columns show liquid-fuel and local
nescence intensity by several orders of magnitude gas temperature (green) for the KIVA-CHEMKIN and

JER02406 © IMechE 2007 Int. J. Engine Res. Vol. 8

Downloaded from jer.sagepub.com at University of Birmingham on June 2, 2015


108 S Singh, R D Reitz, M P B Musculus, and T Lachaux

Fig. 6 Left column: experimentally observed simultaneous LMS (blue) and natural emission
(green). Right columns: model-predicted liquid-fuel (red circles and blue colour) and
either ignition radical concentration or local gas temperature (green) in the plane of the
nominal jet axis for the HTC condition

the KIVA-RIF models respectively, in a plane along that the chemiluminescence is slightly stronger on
the jet axis. The temperature scale for the model the ‘sides’ of the jet, away from the nominal jet
images is from 850 to 1100 K. The sequence of images centre-line. The KIVA-CTC and KIVA-CHEMKIN
begins at 0° ATDC, just before the chemilumin- models first indicate ignition near 1° ATDC, in similar
escence is first observed, and ends at 3° ATDC. For regions of the jet as for the experimental obser-
this operating condition, the models predict sig- vations. The KIVA-RIF model predicts the earliest
nificantly different ignition timings relative to each indications of ignition 1 CAD earlier, near 0° ATDC,
other and to the experimental observations, so the on the ‘sides’ of the jet but also along the liquid-fuel
experimental and model-predicted images at a par- spray, very close to the injector (greenish tint on the
ticular crank angle do not correspond to the same left side of the top-right image in Fig. 6). After about
location on respective AHRR curves. As a result, the 1 CAD (near 1° ATDC for the KIVA-RIF and near
model predictions and the experimental images 2° ATDC for the KIVA-CTC and KIVA-CHEMIN
must be compared with care, with consideration for models), the KIVA-RIF model suddenly predicts
differences in the predicted ID. ignition throughout the jet cross-section, while the
Weak chemiluminescence from ignition reactions KIVA-CTC and KIVA-CHEMKIN models still predict
is first observed at 1° ATDC, even as the experimental ignition only on the ‘sides’ of the jet. This sudden,
AHRR is negative (Fig. 3). By 2° ATDC, the natural broad occurrence of ignition leads to a larger and
emission is slightly stronger and more downstream more distinct CF heat release prediction by the KIVA-
of the initial ignition sites. The shape of the brightest RIF model, which was neither observed in the experi-
regions varies somewhat from cycle to cycle, but they ments nor predicted by the KIVA-CTC and KIVA-
are always located downstream of the LMS signal, in CHEMKIN models. The large CF heat release also
the mid-stream region of the jet. In addition, note contributes to the rapid fuel vaporization predicted

Int. J. Engine Res. Vol. 8 JER02406 © IMechE 2007

Downloaded from jer.sagepub.com at University of Birmingham on June 2, 2015


Validation of engine combustion models 109

by KIVA-RIF, as discussed in the previous section. In magnitude, so the chemiluminescence recorded by


addition, note that although the CHEMKIN and RIF the camera is most probably from high-temperature
chemistry mechanisms are the same, differences in reactions (low-temperature chemiluminescence
the combustion model formulations result in dif- would be too weak to be recorded by the camera).
ferent ID predictions. The RIF model accounts for The location of the bright chemiluminescence spot
turbulence–chemistry interactions through scalar varies slightly from cycle to cycle but it is mostly
dissipation rate, which affects the ID predictions. present near the tip and on one side of the liquid-
At the start of SSC, defined as the time when the fuel jet. The KIVA-CTC model predicts SSC at
AHRR becomes more than about 50 J/deg (near 3° 6° ATDC leeward side of the jet with respect to the
ATDC in Fig. 3), the chemiluminescence intensity in counter-clockwise swirl flow. That is, the swirl is
the experimental images becomes stronger by nearly counter-clockwise in the images, and ignition occurs
an order of magnitude from the initially detected on the leeward side of the swirl flow with respect to
emission, as indicated by reduced camera gain in the the jet. The KIVA-CHEMKIN model predicts SSC on
3° ATDC image compared with the 1° ATDC image. both ‘sides’ of the jet at 3° ATDC. The KIVA-RIF model
The KIVA-RIF and the KIVA-CHEMKIN models predicts the start of SSC throughout the jet cross-
predict the start of SSC at 1.7° and 2.5° ATDC respect- section, and extends far upstream, near the injector
ively. For the KIVA-RIF model, significant tempera- nozzle.
ture rise owing to SSC is predicted throughout the
jet cross-section, similar to the CF temperature dis-
4.1.4 In-cylinder OH and soot distributions
tribution. The sudden occurrence of SSC throughout
the jet leads to high peak AHRR prediction by the PLIF of OH was excited by pumping a narrow OH
KIVA-RIF model. For large premixed-burn conditions absorption line with the 284 nm laser sheet.
such as this one, the diffusion-flamelet assumptions Camera 2 in Fig. 1 imaged the fluorescence emission,
of the RIF equations are less appropriate, probably using a set of spectral filters to isolate the OH
contributing to inaccurate predictions of premixed fluorescence in the 308–320 nm range and to reject
combustion. Figure 7 shows the experimental and elastic scatter [16]. Unfortunately, other species, such
model images at the start of SSC, using a temperature as fuel (as in the BB-PLIF measurement), polycyclic
scale from 1100 to 2500 K for the model images. To aromatic hydrocarbons (PAHs), and soot may also
account for differences in the start of SSC timing emit light within the filter passband after excitation
predicted by the models, images were selected from near 284 nm. Fortunately, OH fluorescence is excited
similar positions at the start of the SSC AHRR (see only for a very specific laser wavelength (narrow-line
Fig. 3) for the experiment and model predictions, absorber), while the other sources are excited over a
which have different crank positions. Strong chemi- wide range of laser wavelengths (broadband absor-
luminescence is typically observed near the tip of the bers). Therefore, OH fluorescence may be identified
LMS signal at 4° ATDC. To avoid image saturation, by comparing the fluorescence signal with the
the camera gain was reduced by over two orders of laser wavelength tuned to overlap with yhe OH

Fig. 7 Left column: experimentally observed simultaneous LMS (blue) and natural emission
(green). Right columns: model-predicted liquid-fuel (red circles and blue colour) and local
gas temperature (green) in the plane of the nominal jet axis for the HTC condition

JER02406 © IMechE 2007 Int. J. Engine Res. Vol. 8

Downloaded from jer.sagepub.com at University of Birmingham on June 2, 2015


110 S Singh, R D Reitz, M P B Musculus, and T Lachaux

absorption line (‘online’, near 284 nm) to the variation in both the soot luminosity and OH-PLIF
fluorescence signal with the laser wavelength tuned distributions, the fluorescence is often strongest in a
so that it does not overlap an OH absorption line ring-like structure surrounding the soot-producing
(‘offline’, near 283.9 nm). A strong contrast between central region of the jet, similar to the quasi-steady
online and offline fluorescence intensity indicates diffusion-flame-like structure of the diesel concep-
that the fluorescence is dominated by OH, while tual model proposed by Dec [10]. In Dec’s investi-
weaker contrast indicates that other sources contrib- gation of a hotter intake-temperature condition with
ute to the signal. With a single laser, simultaneous a shorter ID, the temporal overlap between injection
images of both online and offline fluorescence could and combustion events was more substantial, such
not be acquired during the same engine cycle, so that a quasi-steady diffusion flame was sustained for
representative online and offline images from a relatively long time. For the operating condition
different cycles at each crank angle were compared investigated here, the injection ends at 5° ATDC and
to assess the general spatial contribution of OH-PLIF a major part of the AHRR occurs after the end of
to the total fluorescence signal. After comparing the injection. Therefore, OH and soot structures like
online and offline images, regions of strong online/ those at 6° ATDC are observed only for a short dur-
offline OH-PLIF contrast have been marked on the ation. Indeed, by 8° ATDC, the OH distribution
experimental images (white boundary). The fluores- becomes broader and both OH and soot are found
cence inside the boundary should be interpreted as only in the downstream regions of the jet. By 10°
OH-PLIF, while fluorescence outside the boundary ATDC, the OH distribution becomes even more uni-
should be interpreted as BB-PLIF (possibly from fuel, form and there is no indication of any thin, ring-like
PAHs, soot, etc.). OH structures. Also, the AHRR becomes very low
The in-cylinder distribution of soot for the HTC after 8° ATDC (see Fig. 3) and the OH and soot are
condition was observed by line-of-sight soot- concentrated primarily near the bowl wall. Finally,
luminosity imaging. A comparison of the model- the soot luminosity and OH-PLIF gradually decrease
predicted soot with the high-speed soot luminosity until only a weak soot luminosity and OH-PLIF are
imaging has been presented in a previous publi- observed in the downstream regions of the jet by
cation [8]. In this paper, simultaneous images of both 22° ATDC.
soot luminosity and OH-PLIF for the HTC condition The KIVA-CTC model predicted a longer ID com-
are presented and compared with model predictions. pared with the experiments, with SSC starting near
The left column in Fig. 8 shows simultaneous 6° ATDC. Consequently, soot formation is also
images of online OH-PLIF (green) and line-of-sight delayed. At the start of SSC, the KIVA-CTC model
soot luminosity (red). The other three columns show predicts soot on the leeward side of the jet with
model-predicted mass fractions of OH (green) and respect to the counter-clockwise swirl flow, where the
soot (red) in a plane along the nominal fuel-jet axis. high-temperature SSC was first predicted (cf. Fig. 7).
The yellow colour in the images is a result of the Soot immediately follows HTC with the KIVA-CTC
overlapping of red and green colours. The KIVA-CTC model because the soot model is activated when the
model does not predict OH radicals; therefore only local cell temperature reaches 1100 K. Recall that the
the soot distribution is presented. For the model- model uses the ‘fuel’ as a soot precursor, so the soot
predicted soot and OH, the colour-maps are set to is predicted in all regions of the jet that contain some
the same lower and upper levels for all the models. amount of fuel and are hotter than 1100 K. The local
In this way, the difference in colour brightness from rate of soot formation is directly proportional to the
one model to another indicates a difference in the amount of fuel, and the soot formation and oxidation
concentration of soot or OH. rates can be controlled by tuning the soot model con-
At 5° ATDC, the crank angle of peak experimental stants for the two-step NSC model. Generally, these
AHRR, weak soot luminosity is first observed in the model constants are tuned to match the exhaust soot
downstream region of the jet. At this crank angle, the measurements, but since such measurements were
intensity of fluorescence (green) is completely insen- not available, the standard soot model constants
sitive to laser wavelength, and therefore interpreted were chosen [4]. In addition, for the type of compari-
as entirely BB-PLIF (no OH). Starting at 6° ATDC, the son presented in this paper, only the qualitative dis-
contrast between online and offline in the down- tribution of soot, and not its absolute concentration,
stream region of the OH-PLIF images increases can be compared with the experimental measure-
dramatically, suggesting an abundance of OH rad- ments. Change in soot model constants does not
icals in this area, as indicated by the white-dashed greatly affect the soot distribution in the jet and only
boundary. Although there is significant cycle-to-cycle changes in absolute amount of soot in the cylinder

Int. J. Engine Res. Vol. 8 JER02406 © IMechE 2007

Downloaded from jer.sagepub.com at University of Birmingham on June 2, 2015


Validation of engine combustion models 111

Fig. 8 Left column: experimentally observed simultaneous OH-PLIF (green) and soot luminosity
(red) images. Right columns: model-predicted OH (green) and soot (red) mass fractions
in the plane of the nominal jet axis for the HTC condition

at any given time. At 8° ATDC, much of the volume dict a high concentration of OH radicals on the per-
of the jet is at high temperature, for the KIVA-CTC iphery of the jet at 6° ATDC surrounding the soot-
model, so that soot is predicted throughout the jet. producing central region, similar to the experimental
Compared with the experimental observations of distributions of OH and soot. Details of the predicted
soot luminosity, the KIVA-CTC model predicts soot flame structures are somewhat different for the two
in a larger portion of the jet. The model also predicts models, however. The KIVA-RIF model predicts a
soot in the upstream regions of the jet long after the broader OH distribution, and the flame (as indicated
end of combustion (e.g. at 22° ATDC in Fig. 8). This by OH radicals) extends fully upstream to the injector
is clearly in disagreement with the experimental nozzle. By contrast, the KIVA-CHEMKIN model pre-
observations of soot luminosity. dicts a shell-like OH distribution, and the flame does
Both KIVA-CHEMKIN and KIVA-RIF generally pre- not extend as far upstream. In the experimental

JER02406 © IMechE 2007 Int. J. Engine Res. Vol. 8

Downloaded from jer.sagepub.com at University of Birmingham on June 2, 2015


112 S Singh, R D Reitz, M P B Musculus, and T Lachaux

image at 6° ATDC, the OH-PLIF from the 3 o’clock 4.2 Early-injection LTC condition
jet does not extend as far upstream as the model-
4.2.1 AHRR and cylinder pressure
predicted OH distributions, but attenuation of the
284 nm sheet probably reduces the fluorescence For this case, the start of fuel injection is early, at
excitation in the upstream regions, such that the −22° ATDC, as shown in Fig. 9. Heat release from
fluorescence signal is weaker upstream. That is, the the CF reactions is first observed in the experiments
284 nm laser light is absorbed by fuel molecules, at −15° ATDC after an ID of 7 CAD. The CF ID is
combustion intermediates, and/or soot in the down- under-predicted and the rate of heat release is over-
stream region of the jet, such that OH-PLIF nearer predicted by both the KIVA-CHEMKIN and the KIVA-
the injector becomes weaker. Indeed, BB-PLIF is RIF models. This is most probably owing to the
clearly present in the upstream regions of both of the characteristics of the n-heptane chemistry mechan-
jets adjacent to the 3 o’clock jet (see 6° and 8° ATDC ism. It is likely that the n-heptane mechanism over-
images in Fig. 8), which probably experience much predicts fuel reactivity in the CF regime. The KIVA-
less attenuation upstream because they are illumi- CTC model, which uses the simplified Shell model
nated partially from the sides. The absence of for auto-ignition, predicts very low and non-
BB-PLIF in the upstream region of the 3 o’clock jet distinctive heat release during the CF period. In the
is therefore clear evidence of attenuation of the experimental AHRR, after a period of very low heat
284 nm light in the upstream region of the 3 o’clock release, the SSC starts at −12° ATDC. The SSC ID,
jet. Accordingly, the OH-PLIF may actually extend defined as the time in CAD from the start of fuel
further upstream than is apparent in the 3 o’clock injection to the time when the second-stage AHHR
jet, but conclusive evidence of upstream OH is not first exceeds the CF AHHR, is fairly well predicted by
discernible from the current data. The online/offline all the models.
contrast in upstream regions of the adjacent jets is
very low, which indicates that the fluorescence from 4.2.2 Liquid- and vapour-fuel distributions
the near-nozzle area is mainly from broadband
Figure 10 shows a comparison of the experimental
sources such as fuel (not OH). Thus, there is no evi-
LMS from liquid-fuel droplets (blue) and BB-PLIF
dence that OH is present in detectable levels very
(green) images (left column) and model-predicted
near the injector, as predicted by the KIVA-RIF
liquid- and vapour-fuel distributions (three columns
model. This is probably owing to a loss in spatial
on right). The liquid fuel first appears in the exper-
resolution caused by assuming a single flamelet and
imental images at −22° ATDC. At −18° ATDC, the
one global scalar dissipation rate for the entire com-
penetration of BB-PLIF slightly exceeds the penetra-
putational domain. That is, although the scalar dissi-
tion of the tip of LMS while the models predict equal
pation rate is calculated locally in each computation
penetration of liquid and vapour fuel. For this
cell, in the single flamelet approach, a domain-
operating condition, the ID is longer than for the
averaged scalar dissipation is used in the flamelet
HTC condition and therefore, the jet penetrates
equations to obtain instantaneous values of species
mass fraction and temperature. The instantaneous
values are then converted into mean values using
the PDF approach in each computational cell. As a
result, the RIF can predict combustion very near the
injector.
Similar to the experimental observations, both
KIVA-CHEMKIN and KIVA-RIF predict similar devel-
opment of the OH and soot regions in Fig. 8. Both
models predict increased broadness of the OH distri-
bution and decreased area of the soot cloud. During
later stages of the combustion, the OH and soot are
concentrated primarily near the bowl wall. For the
KIVA-RIF model, however, a large soot-producing
region overlaps with the OH regions as indicated by
a light yellow colour of the image. In contrast, the Fig. 9 Experimentally measured and modelled cylin-
KIVA-CHEMKIN model predicts soot and OH regions der pressure and AHRR, and measured mass
that are more spatially displaced from each other rate of injection profile for the early-injection
(less yellow colour), consistent with the experiments. LTC condition

Int. J. Engine Res. Vol. 8 JER02406 © IMechE 2007

Downloaded from jer.sagepub.com at University of Birmingham on June 2, 2015


Validation of engine combustion models 113

Fig. 10 Experimentally observed LMS (blue) and BB-PLIF (green) and model-predicted liquid-
fuel (red circles and blue colour) and vapour-fuel mass fraction (green) in the plane of
the nominal jet axis for the early-injection LTC condition

further into the combustion chamber before the start tion of vapour-fuel penetration, although the arrival
of SSC. The models predict the penetration and of vapour fuel at the piston bowl wall is predicted 1
width of vapour fuel very well, especially the spread- CAD earlier than observed in the experiments.
ing of the jet near the bowl wall at −14° ATDC.
However, near −14° ATDC, the models predict liquid
4.2.3 Onset and development of auto-ignition sites
fuel near the injector while no LMS is observed in
the corresponding experimental image. The first column on the left in Fig. 12 shows simul-
The penetration of liquid and vapour fuel is plotted taneous images of LMS (blue) and ignition chemi-
in Fig. 11. All three models predict the LL well for the luminescence (green). The second column shows
first half of the injection event. As observed for the KIVA-CTC-predicted LL and ignition radical species,
HTC condition, a small disconnected pocket of liquid and the third and fourth columns show KIVA-
fuel at the tip of the jet leads to over-prediction of LL CHEMKIN- and KIVA-RIF-predicted LL and gas tem-
from −19° to −18° ATDC. Near −17° ATDC, evapor- perature in a plane along the jet axis. The images are
ation of this pocket causes a sudden decrease of LL presented only during the CF period from −16° to
in all three of the models, similar to the HTC con- −12° ATDC, and the temperature scale for the model
dition. Again, some liquid fuel persists until very late images is 850–1100 K.
in the cycle, but only 5 per cent of total fuel injected In the experiments, weak chemiluminescence is
remains at −14° ATDC, and essentially zero remains first observed at −15° ATDC, which is just after a
by 2° ATDC. The models give reasonably good predic- positive CF AHRR is observed (Fig. 9). This typical

JER02406 © IMechE 2007 Int. J. Engine Res. Vol. 8

Downloaded from jer.sagepub.com at University of Birmingham on June 2, 2015


114 S Singh, R D Reitz, M P B Musculus, and T Lachaux

larger region of the jet, and the models predict


ignition throughout the jet cross-section, especially
in the downstream regions. At −13° ATDC, the
chemiluminescence increases by an order of magni-
tude as the CF AHRR reaches its maximum value
(Fig. 9), and is strongest from the mid-stream to the
downstream regions of the jet. At the same time, the
models slightly over-predict the vapour-fuel penetra-
tion (Fig. 10), causing a greater spread of vapour fuel
along the piston bowl wall. This greater spread is
manifested as a wider ignition zone near the piston
bowl wall. At −12° ATDC, near the end of the CF
stage, as the AHRR begins to decrease (Fig. 9), the
chemiluminescence overall becomes weaker, but
brighter emission is observed extending upstream to
the injector. The models do not predict a concomi-
tant temperature rise near the injector, and liquid-
fuel evaporation near the injector is under-predicted.
In the experiments, no liquid fuel (blue) is detected
in the −12° ATDC image, even with the camera gain
increased by an order of magnitude compared with
the earlier LMS images. In all three models, however,
the fuel near the injector is slower to vaporize and
Fig. 11 Experimentally observed and model-predicted some liquid fuel remains, as indicated by the red
LL, vapour penetration, mass rate of injection
circles and blue colour in the upstream regions of
profile, and in-cylinder liquid-fuel fraction for
the early-injection LTC condition the model images at −12° ATDC. Under-prediction
of fuel vaporization in all of the models may be
indicative of poor prediction of the CF chemistry that
image shows strong chemiluminescence in the mid- raises local temperatures and drives vaporization of
stream region and on the ‘sides’ of the jet, although the liquid fuel.
the actual location of the brightest chemilumi- The transition to the rapid, exothermic SSC reac-
nescence varies from cycle to cycle. All of the models tions occurs near −11° ATDC, as shown in Fig. 13.
predict ignition on the periphery of the jet near The left column in Fig. 13 shows simultaneous
−16° ATDC, but the KIVA-CTC and KIVA-CHEMKIN images of BB-PLIF (blue) and line-of-sight chemi-
models predict ignition only in the mid-stream luminescence (green). The other three columns show
region, while the KIVA-RIF model predicts ignition model predictions of OH mass fraction (blue) and
along the entire length of the jet. Similar to the exper- the local gas temperature (green) in a plane along
imental observations in this study, some numerical the jet axis. For the model images, the temperature
[44] and other experimental [45] investigations of scale is from 1100 to 2000 K, which spans the range
ignition in a diesel jet have shown that ignition typi- of flame temperatures observed during SSC for this
cally first occurs in the mid-stream region of the jet. diluted condition. For the KIVA-CTC model, only gas
Thus, the KIVA-RIF predictions disagree with the temperature is shown, since OH is not modelled. The
experimental chemiluminescence observations of AHRR from SSC first becomes positive at −11° ATDC,
this study and observation from other studies as well. but strong chemiluminescence is not observed until
By −15° ATDC, the KIVA-CTC and KIVA-CHEMKIN −10° ATDC. Note that both chemiluminescence
models still predict ignition in a shell on the periph- images in Fig. 13 are about 80 times as bright as at
ery of the jet, while the KIVA-RIF model predicts −12° ATDC in the last row in Fig. 12, as indicated by
ignition throughout the jet cross-section. Recall that the reduction in camera gain. The KIVA-CHEMKIN
a similar, sudden appearance of ignition throughout model predicts the start of SSC at −11° ATDC, and
the jet cross-section was also predicted by the KIVA- by −10° ATDC the OH mass fraction rises signifi-
RIF model for the HTC condition. cantly on the periphery of the high-temperature
Just one-half CAD later, at −14.5° ATDC, the zone. The KIVA-RIF and the KIVA-CTC models pre-
chemiluminescence emission in the experimental dict the start of SSC later, at −10° ATDC. Similar to
images becomes stronger and is recorded over a the experiments, all the models predict ignition close

Int. J. Engine Res. Vol. 8 JER02406 © IMechE 2007

Downloaded from jer.sagepub.com at University of Birmingham on June 2, 2015


Validation of engine combustion models 115

Fig. 12 Left column: experimentally observed simultaneous LMS (blue) and natural emission
(green). Right columns: model-predicted liquid-fuel (red circles and blue colour) and
either ignition radical concentration or local gas temperature (green) in the plane of the
nominal jet axis for the early-injection LTC condition

to the piston bowl wall, but temperature distri- 532 nm laser sheet used for LMS. Camera 1 in Fig. 1
butions are somewhat different for each model. captured the resulting LII emission, using a 450 nm
short-wave-pass filter to reject much of the longer-
4.2.4 In-cylinder OH and soot
wavelength naturally occurring soot luminosity, and
For this operating condition, relatively low levels of a 532 nm notch filter to reject elastically scattered
in-cylinder soot are produced owing to extensive pre- laser-light. This LII filtering scheme is essentially the
combustion mixing [14]. For this operating con- same as that described in detail in previous studies
dition, soot is present in tiny isolated pockets in on the same engine [18, 20, 46].
somewhat random locations. The most representa- The image sequence in Fig. 14 starts at −9° ATDC,
tive image of the ensemble-averaged behaviour at near the peak experimental AHRR, and ends at 4°
each crank angle has been chosen from a set of 20 ATDC. Prior to −7° ATDC, the fluorescence is insen-
images from different engine cycles. These images sitive to laser wavelength (i.e. it is BB-PLIF) and is
are presented in Fig. 14, which shows simultaneous therefore attributed to mostly broadband sources
images of OH-PLIF (green) and soot LII (red) from (fuel, formaldehyde, PAH, etc.). At −7° ATDC, strong
the experiments (left column), along with the model- fluorescence is observed throughout the length of
predicted OH (green) and soot (red) mass fractions the jet, even in the upstream region. Recall that
(right columns), all in a plane along the jet axis. In unburned fuel and intermediates in the downstream
the current study, soot LII was excited with the same regions of the jet (when present) severely attenuate

JER02406 © IMechE 2007 Int. J. Engine Res. Vol. 8

Downloaded from jer.sagepub.com at University of Birmingham on June 2, 2015


116 S Singh, R D Reitz, M P B Musculus, and T Lachaux

Fig. 13 Left column: experimentally observed simultaneous OH-PLIF (blue) and natural emis-
sion (green). Right columns: model-predicted OH mass fraction (blue) and local gas
temperature (green) in the plane of the nominal jet axis for the early-injection LTC
condition

the laser sheet owing to absorption, so that species of the soot pockets is significant, but on average, soot
in the upstream region are not well illuminated. The is located in the downstream region of the jet, usually
sudden appearance of fluorescence in the upstream near the head vortex, but sometimes it is also found
regions suggests a dramatic reduction in the amount at other random downstream locations. At −5°
of unburned fuel in the downstream region, probably ATDC, for the same LII camera gain, increased inten-
owing to consumption of fuel by rapid exothermic sity of LII indicates a higher amount of soot in the
SSC reactions. Indeed, comparison of online and head vortex; but again, the LII intensity varies from
offline OH-PLIF images shows strong contrast only cycle to cycle. It is important to mention here that
in the downstream regions near the piston bowl wall, the LII diagnostic illuminates only the soot that
indicating the presence of the highly oxidizing radical resides in the plane of the laser sheet, along the jet
OH, and thus significant SSC chemistry. In the axis. Additional soot could be present out of this
upstream regions, however, the online and offline plane. Indeed, the line-of-sight soot luminosity
OH-PLIF signals are of similar strength, so other images presented in a previous publication [8] show
broadband fluorescence sources contribute to the that soot is present in a relatively larger region in the
majority of the emission (i.e. BB-PLIF). That is, fuel three-dimensional jet, near the piston bowl wall. The
or other combustion intermediates are still present strength of the LII signal decreases at −3° ATDC. At
in the upstream region, so the transition to SSC, 0° and 4° ATDC, the brightest red spots are most
which consumes fuel and produces OH, has not yet probably from LII of soot while the weak background
occurred in the upstream region. In the mid- to red colour is most probably soot luminosity that
downstream regions, the OH-PLIF appears in a broad leaks through the LII camera filters. The luminosity
distribution, filling the jet cross-section. A similar is from soot that is out of the plane of the laser sheet.
broad OH-PLIF distribution persists through −5° Starting at the beginning of the sequence in Fig. 14,
ATDC. Starting at −3° ATDC, the BB-PLIF signal in at −9° ATDC, both the KIVA-CHEMKIN and the
the upstream regions of the jet begins to decrease, KIVA-RIF models predict soot near the piston bowl
and by 0° and 4° ATDC, the fluorescence emission wall, surrounded by OH radicals. The KIVA-
(from OH-PLIF) is strong only near the piston bowl CHEMKIN model predicts a ribbon-like OH struc-
wall. ture, while the KIVA-RIF model predicts a broad
Soot first appears in the experimental images (red spatial distribution of OH. At −7° ATDC, the KIVA-
colour in the left column of Fig. 14) at −9° ATDC in CHEMKIN and the KIVA-RIF soot and OH distri-
small randomly located pockets near the piston bowl butions are similar, but the KIVA-CHEMKIN model
wall. Cycle-to-cycle variation in the actual location still predicts OH and soot spatially displaced from

Int. J. Engine Res. Vol. 8 JER02406 © IMechE 2007

Downloaded from jer.sagepub.com at University of Birmingham on June 2, 2015


Validation of engine combustion models 117

Fig. 14 Left column: experimentally observed simultaneous OH-PLIF (green) and soot LII (red).
Right columns: model predicted OH (green) and soot (red) mass fractions in the plane
of the nominal jet axis for the early-injection LTC condition

each other, while the KIVA-RIF model predicts some experimental observations, with OH forming in the
overlapping of OH and soot (yellow colour) which mid- to downstream regions of the jet. At −5° ATDC
was not seen in the experiments. At −7° ATDC, the and later, the KIVA-CHEMIKIN and KIVA-RIF models
first unambiguous evidence of OH-PLIF appears in predict soot mainly in the head vortex region. The
the downstream region of the experimental images, KIVA-RIF model continues to predict more overlap
while the KIVA-CHEMKIN and KIVA-RIF models pre- of soot and OH (yellow colour) compared with the
dict significant OH as early as −11° ATDC. OH may KIVA-CHEMKIN model predictions.
have existed in the experiment before −7° ATDC, For the KIVA-CTC model, soot appears immedi-
but it certainly was not detectable with confidence ately at the start of SSC at −10° ATDC, and soot
(strong online/offline contrast) until after −7° ATDC. remains in the first image shown in Fig. 14, at −9°
Also, all three models under-predict either the CF or ATDC. Similar to the HTC condition, the soot forma-
the SSC ID, so that the onset of OH is somewhat tion region overlaps with the high-temperature
advanced. However, when the OH-PLIF is first region. Clearly, the size of the soot cloud is much
detected in the experiments, the OH distributions larger than that observed in the experiments or as
predicted by the models are somewhat similar to the predicted by either the KIVA-CHEMKIN or the KIVA-

JER02406 © IMechE 2007 Int. J. Engine Res. Vol. 8

Downloaded from jer.sagepub.com at University of Birmingham on June 2, 2015


118 S Singh, R D Reitz, M P B Musculus, and T Lachaux

RIF models. In addition, the KIVA-CTC model pre- model. The KIVA-CTC ID prediction matches with
dicts a large amount of soot surviving even after the the experiments, but the rate of pressure rise, and
end of combustion. Late in the cycle, at 4° ATDC, hence the AHRR, is higher. The simple time-scale
the local soot concentration decreases somewhat formulation of the KIVA-CTC combustion model
(reduced intensity of red colour), but it is still present seems to be inadequate for predicting the AHRR for
over a large area from mid-stream to the downstream this LTC condition, for which chemical kinetic time-
regions of the jet, which is very unlike the experimen- scales may be more important than fluid dynamic
tal observations. The rate of oxidation is greatly time-scales [12, 13].
under-predicted by the KIVA-CTC model, as was also
discussed in Singh et al. [8]. This indicates that the 4.3.2 Liquid- and vapour-fuel distributions
NSC-type oxidation model that was originally pro-
Figure 16 shows the experimentally observed LMS
posed to model soot oxidation for high-sooting diesel
(blue) and BB-PLIF (green) images (left column) and
engine conditions, needs improvements before it can
model-predicted liquid- (red circles and blue colour)
be used for low-sooting LTC conditions.
and vapour-fuel (green) distributions for the late-
injection LTC condition. Liquid fuel first appears in
the experimental image at 0° ATDC. Up to 2° ATDC,
4.3 Late-injection LTC condition
the LMS and BB-PLIF distributions penetrate almost
4.3.1 AHRR and cylinder pressure equally, and all three models predict the penetration
reasonably well during this period. At 4° ATDC, both
Figure 15 shows the cylinder pressure and AHRR for
the experimentally observed and model-predicted
the late-injection LTC condition. The fuel injection
vapour-fuel penetration exceeds the penetration of
begins at TDC. Significant heat release from CF reac-
the liquid fuel. Initially, the liquid penetration is
tions is observed after an ID of about 6.25 CAD.
slightly under-predicted by the models, as indicated
Unlike the early-injection LTC condition, for which
in Fig. 16. As observed for the other two operating
there is a distinct AHRR peak attributable to the CF
conditions, after the end of injection at 8° ATDC, the
chemistry, in this case the experimental CF heat
experimental image does not show significant LMS,
release is not as distinct. The CF AHRR attains a local
but the models still predict considerable liquid fuel
peak value and stays relatively constant until the
near the injector. This is also evident in Fig. 17, as
start of SSC. As for the early-injection LTC condition,
over-prediction of LL after the end of injection. The
the CF ID is under-predicted by both the KIVA-
amount of this liquid fuel is, however, only 2 per cent
CHEMKIN and the KIVA-RIF models. The SSC ID is
of the total liquid fuel injected, and decreases to less
slightly under-predicted by the KIVA-CHEMKIN
than 0.01 per cent by 11° ATDC.
model and is over-predicted by the KIVA-RIF model.
Furthermore, the KIVA-RIF model predicts a very
4.3.3 Onset and development of auto-ignition sites
high peak AHRR compared with the KIVA-CHEMKIN
The leftmost column in Fig. 18 shows simultaneous
experimental images of LMS (blue) and ignition
chemiluminescence (green). The other three col-
umns show the model-predicted liquid-fuel (red par-
ticles and blue colour) and either ignition radical
concentration or local gas temperature (green) in a
plane along the jet axis. Similar to the early-injection
LTC condition, the images are presented during the
CF AHRR, and the temperature scale for the model
images is set from 850 to 1100 K. The onset and
development of auto-ignition sites for this condition
is very similar to the early-injection LTC condition
already described.
Natural emission from ignition chemilumin-
escence in the experimental images is first observed
Fig. 15 Experimentally measured and modelled cylin- at 5.5° ATDC, downstream of the tip of the LMS, in
der pressure and AHRR, and measured mass the mid-stream region of the jet. Note that the
rate of injection profile for the late-injection chemiluminescence is once again stronger on the
LTC condition ‘sides’ of the jet. Similar to the early-injection LTC

Int. J. Engine Res. Vol. 8 JER02406 © IMechE 2007

Downloaded from jer.sagepub.com at University of Birmingham on June 2, 2015


Validation of engine combustion models 119

Fig. 16 Experimentally observed LMS (blue) and BB-PLIF (green) and model-predicted liquid-
fuel (red circles and blue colour) and vapour-fuel mass fraction (green) in the plane of
the nominal jet axis for the late-injection LTC condition

condition, the CF ID is under-predicted by both the zone. At 8° ATDC, as the vapour fuel approaches the
KIVA-CHEMKIN and the KIVA-RIF models. As a piston bowl wall (cf. Fig. 16), the chemiluminescence
result, the models predict an earlier start of CF is strongest in the mid-stream to downstream regions
ignition at 4° ATDC. The KIVA-CHEMKIN model pre- of the jet. The KIVA-CHEMKIN and the KIVA-RIF
dicts ignition at a few locations on the periphery of models also predict progression of the ignition
the jet, while the KIVA-RIF model predicts ignition towards the piston bowl wall, and the structures of
along the entire length of the jet. At 5.5° ATDC, both ignition zones are similar for the two models. At 9°
models predict an increase in the size of the ignition ATDC, for the same camera gain, the intensity of
zone, although the KIVA-CHEMKIN model predicts chemiluminescence becomes slightly stronger, but
ignition mostly on the periphery of the jet, while the the source is still probably CF reactions. At the start
KIVA-RIF model predicts ignition throughout the jet of SSC, the strength of chemiluminescence suddenly
cross-section. One-half CAD later, at 6° ATDC, the increases, requiring a large reduction in camera gain
chemiluminescence in the experimental image to avoid image saturation [22]. For the experimental
becomes stronger and is observed over a larger area image at 9° ATDC, CF chemiluminescence is strong
of the jet. The models predict an earlier start of CF near the bowl wall and on the upper edge of the jet.
ignition, and they also over-predict spreading of The locations of bright spots vary somewhat from
vapour fuel near the piston bowl wall, so the size of cycle to cycle. The KIVA-CHEMKIN and the KIVA-RIF
ignition zones predicted by the models is larger than models predict widening of the ignition zone near
the experimentally observed chemiluminescence the bowl wall at 9° ATDC.

JER02406 © IMechE 2007 Int. J. Engine Res. Vol. 8

Downloaded from jer.sagepub.com at University of Birmingham on June 2, 2015


120 S Singh, R D Reitz, M P B Musculus, and T Lachaux

vation, while the KIVA-RIF model predicts a longer


SSC ID. Thus, the model images are phase shifted
such that the first image for each model is presented
at the start of SSC. At the start of SSC at 10° ATDC,
KIVA-CHEMKIN predicts high temperatures near the
piston bowl wall, in a region similar to the region of
strong natural emission observed in the experimen-
tal image. The model also predicts a slight increase
in the size of the high-temperature region at 11°
ATDC, along with the appearance of OH radicals at
the periphery of this zone. The KIVA-CTC model pre-
dicts the start of SSC at 11° ATDC. The size of the
predicted high-temperature zone is quite large com-
pared with the experimental observations and the
KIVA-CHEMKIN predictions. The large size of the
ignition zone also agrees with a high rate of AHRR
for the KIVA-CTC model (Fig. 15). The KIVA-RIF
model predicts the start of SSC later, at 14° ATDC.
Initially, the model predicts SSC in a narrow region
near the piston bowl wall, which then spreads to a
bigger region at 14° ATDC. Note that unlike the KIVA-
Fig. 17 Experimentally observed and model-predicted
LL, vapour penetration, mass rate of injection
CHEMKIN model, the KIVA-RIF model predicts OH
profile, and in-cylinder liquid-fuel fraction for radicals (blue) in the entire high-temperature region
the late-injection LTC condition of the jet.
In Fig. 19, the BB-PLIF distribution was not very
visible because the natural emission image obscured
Figure 19 shows a comparison of the simul- the BB-PLIF image. In Fig. 20, only BB-PLIF images
taneously acquired images of BB-PLIF (blue) and are compared with the model-predicted OH distri-
naturally occurring luminous emission (green) with butions. The image sequence continues from Fig. 19.
the model-predicted OH radical concentration (blue) Since KIVA-CTC does not predict OH radicals, only
and the local gas temperature (green) in a plane KIVA-CHEMKIN and KIVA-RIF predictions are com-
along the jet axis. For the model images, the tempera- pared with the experiments. Again, the images are
ture scale is 1100–2000 K. At 10° ATDC, strong natural phase shifted such that the first image is presented
emission is observed in downstream regions of the at the crank angle of first appearance of OH radicals
jet. A large reduction in camera gain (factor of 22) in the model images. The online/offline fluorescence
from the previous image in Fig. 18 indicates the start contrast improves at 13° ATDC indicating the appear-
of brighter chemiluminescence that accompanies ance of OH radicals in the jet. At 12° ATDC, the KIVA-
SSC. At 11° ATDC, the intensity of chemilumi- CHEMKIN model predicts an arc-like, thin structure
nescence becomes even stronger and moves slightly of OH radicals. The experimental image does not
upstream, indicating SSC occurring in a larger area. show a similar distinct arc of BB-PLIF, which may
The BB-PLIF is stronger in the regions of strong contain some OH-PLIF. On the contrary, the exper-
chemiluminescence, indicating consumption of fuel, imental image shows an arc that is turned backwards.
and thus greater penetration of the laser sheet into The authors believe that this is owing to the attenu-
the jet interior. Near the start of SSC, the contrast ation of the laser sheet. Initially, the presence of fuel
between online and offline fluorescence is very poor and other intermediate species near the piston bowl
indicating that the fluorescence (blue) is primarily wall attenuates the laser sheet and prevents it from
from broadband sources, although some OH radicals reaching the upstream regions of the jet. Only the
could also be present. Considering the possibility of OH radicals, fuel, and intermediates at the leading
the presence of OH radicals, the model-predicted OH edge of the jet are excited by the laser sheet. At 13°
distribution was compared with the experimentally ATDC, as the laser sheet penetrates further into the
observed BB-PLIF. In doing so, the different IDs for jet, fluorescence is now observed in a larger region
SSC predicted by each of the models must be near the piston bowl wall. There is some indication
accounted for. The KIVA-CHEMKIN and the KIVA- of ribbon-like structures in some images, but the
CTC IDs are shorter than the experimental obser- shape of these structures varies greatly from cycle to

Int. J. Engine Res. Vol. 8 JER02406 © IMechE 2007

Downloaded from jer.sagepub.com at University of Birmingham on June 2, 2015


Validation of engine combustion models 121

Fig. 18 Left column: experimentally observed simultaneous LMS (blue) and natural emission
(green). Right columns: model-predicted liquid-fuel (red circles and blue colour) and
either ignition radical concentration or local gas temperature (green) in the plane of the
nominal jet axis for the late-injection LTC condition

cycle, and they never appeared as a well-defined arc cence is observed only near the piston bowl wall.
as predicted by the KIVA-CHEMKIN model. The By 16° ATDC, fluorescence emission is observed
KIVA-RIF model predicts a broad distribution of OH throughout the length of the jet. However, the fluo-
at the start of SSC at 15° ATDC, and the distribution rescence from OH radicals, as indicated by strong
continues to be broad at 16° ATDC. online/offline contrast, dominates only in the down-
stream region, as marked by the white boundary in
4.3.4 In-cylinder OH and soot distributions
the experimental images. The fluorescence in the
Figure 21 shows a comparison of the model- upstream region is insensitive to laser wavelength
predicted OH (green) and soot (red) mass fractions and is therefore mostly from broadband sources,
with simultaneous experimental images of OH-PLIF probably unburned fuel and combustion intermedi-
(green) and soot-LII (red), in a plane along the jet ates. From 16° to 22° ATDC, the distribution and
axis. The experimentally observed OH-PLIF distri- intensity of fluorescence remains nearly unchanged
bution for this late-injection LTC condition is similar in the downstream region but it continues to
to the previously discussed early-injection LTC con- brighten in the upstream region. By 30° ATDC, only
dition. The OH-PLIF first appears in the downstream BB-PLIF remains, in the upstream region of the jet,
region of the jet. Again, as for the early-injection LTC near the injector. Note that some of this signal very
condition, initially, the laser does not penetrate into near the injector tip is from reflection of 284 nm laser
upstream regions of the jet, and as a result, fluores- light off the cylinder head, which appears as a thick

JER02406 © IMechE 2007 Int. J. Engine Res. Vol. 8

Downloaded from jer.sagepub.com at University of Birmingham on June 2, 2015


122 S Singh, R D Reitz, M P B Musculus, and T Lachaux

Fig. 19 Experimentally observed simultaneous OH-PLIF (blue) and natural emission (green) and
model-predicted OH mass fraction (blue) and local gas temperature (green) in the plane
of the nominal jet axis for the late-injection LTC condition

of Fig. 21. At 22° ATDC, the intensity of LII decreases,


which indicates that either soot oxidizes or some of
it moves out of the laser sheet. The latter is more
likely because the line-of-sight soot luminosity
images presented in a previous publication [14] indi-
cate that soot survives somewhere in the jet as late
as 40° ATDC.
All three models predict similar distributions of
soot and OH as for the early-injection LTC condition.
As usual, the KIVA-CTC model predicts soot in the
entire high-temperature region. Clearly, there is a dis-
crepancy between the area of the experimentally
observed soot-LII and the KIVA-CTC-predicted soot.
The model also predicts slower soot oxidation, as
was also the case for the early-injection LTC con-
dition. Even long after the end of combustion at 30°
ATDC, the KIVA-CTC model still predicts soot in a
large portion of the jet. Both the KIVA-CHEMKIN-
Fig. 20 Experimentally observed BB-PLIF (green) and
OH mass fraction (green) predicted by KIVA- and the KIVA-RIF-predicted OH and soot distri-
CHEMKIN and KIVA-RIF models in the plane butions are very similar to the early-injection LTC
of the nominal jet axis for the late-injection condition. The KIVA-CHEMKIN model predicts a
LTC condition high-concentration ribbon-like OH structure sur-
rounding the soot formation region. Unlike exper-
bright band (marked under the solid white rectangle) imental observations of soot, the model predicts
on the left-hand side of the image. soot throughout the jet cross-section, however only
The experimentally observed soot-LII distributions in a thin region near the piston bowl wall. Later, the
for this late-injection LTC condition are also some- soot is convected from the middle of the jet to the
what similar to the early-injection LTC condition. head vortex region. The KIVA-RIF model predicts
The cycle-to-cycle variations of the location of bright soot mainly in the head vortex region similar to the
spots of LII are lower, however, and the soot is still experimental image. As discussed earlier, the KIVA-
found mostly in the head vortex region, as evident RIF model predicts a longer SSC ID. During this
from the 16° and 18° ATDC images in the left column additional time available for jet development, the

Int. J. Engine Res. Vol. 8 JER02406 © IMechE 2007

Downloaded from jer.sagepub.com at University of Birmingham on June 2, 2015


Validation of engine combustion models 123

Fig. 21 Experimentally observed simultaneous OH-PLIF (green) and soot LII (red) and model-
predicted OH (green) and soot (red) mass fractions in the plane of the nominal jet axis
for the late-injection LTC condition

fuel in the middle region of the jet is convected to lowing conclusions can be drawn based on the
the head vortex region due to the momentum of the observations.
jet. This results in a less-rich mixture in the middle
region of the jet such that at the start of SSC, soot 1. All three combustion models give reasonable pre-
is formed only in the head-vortex region. Beyond dictions of the cylinder pressure and AHRR trends
18° ATDC the KIVA-CHEMKIN and the KIVA-RIF for the HTC and early- and late-injection LTC con-
images of soot become similar to each other and to ditions. KIVA-CTC and KIVA-RIF models predicted
the experiments. higher rates of combustion and higher peak AHRR
compared with the KIVA-CHEMKIN model.
2. The models predict liquid- and vapour-fuel pene-
5 SUMMARY AND CONCLUSIONS tration well for the first half of the injection event.
For the second half of the injection event, the LL
Three different approaches to modelling ignition and is slightly over-predicted, and vaporization of
combustion in diesel engines were compared against liquid fuel near the nozzle near the end of injec-
cylinder pressure, AHRR, and in-cylinder images of tion is under-predicted by the spray and ignition
spray, combustion, and emissions for a heavy-duty models used in this study.
DI diesel engine. The models were implemented into 3. The KIVA-CHEMKIN model best predicts the
the same version of the KIVA-3V code, which pro- onset and development of CF ignition sites in the
vides a common platform for comparing previously mid-stream region on the ‘sides’ of the jet. The
proposed ignition and combustion models. The fol- KIVA-RIF model predicts blurred-out ignition

JER02406 © IMechE 2007 Int. J. Engine Res. Vol. 8

Downloaded from jer.sagepub.com at University of Birmingham on June 2, 2015


124 S Singh, R D Reitz, M P B Musculus, and T Lachaux

sites. The KIVA-RIF model also predicts faster 8. Finally, the KIVA-CHEMKIN approach gives good
appearance of ignition throughout the jet as com- results without the need to account for sub-grid-
pared with the experimental observations and the scale turbulence–chemistry interactions. The
KIVA-CHEMKIN model. The KIVA-CTC model present KIVA-RIF model implementation that uses
does not predict considerable heat release during global quantities to account for turbulence-
the CF period. However, it still predicts ignition chemistry interactions does not perform as well.
reactions, as indicated by an accumulation of alkyl It is possible that the performance of RIF-type
radicals. models could be improved if turbulence–chemis-
4. Although the timing of SSC is not always well pre- try interactions were considered locally in each
dicted, the spatial locations of SSC are well pre- computational cell, instead of using a limited
dicted by the models. For the LTC conditions, both number of flamelets for the entire computational
experimentally observed and model-predicted domain. However, the good agreement with
SSC starts near the piston bowl wall. However, the the experiments demonstrated by the KIVA-
actual size and shape of the SSC zone is different CHEMKIN model suggests that this level of detail
for all the models. In general, at the start of SSC, is not necessary for accurate modelling of the
KIVA-CTC predicts a larger combustion zone com- present diesel engine combustion regimes. The
pared with the experiment and with the other very simplified representation of chemistry in the
two models. KIVA-CTC model does not seem appropriate for
5. The KIVA-CHEMKIN model always predicts a rela- modelling LTC conditions, especially during the
tively thin diffusion flame (as indicated by OH CF period.
radicals) surrounding a soot-producing region,
especially at the start of SSC. Such thin, ribbon-
like OH distributions were not observed in the ACKNOWLEDEGMENTS
experiments. However, similar to the experimental
observations, the KIVA-CHEMKIN predicted OH The experiments were performed at the Combustion
and soot producing regions that are spatially dis- Research Facility, Sandia National Laboratories,
placed from each other. On the other hand, the Livermore, California. Sandia is a multi-programme
KIVA-RIF model predicts a broader OH distri- laboratory operated by Sandia Corporation, a Lock-
bution and predicts overlapping of soot and OH heed Martin Company, for the United States
in certain regions of the jet that was not seen in Department of Energy’s (DOE) National Nuclear
the experiments. For the LTC conditions, the Security Administration under contract DE-AC04-
KIVA-CHEMKIN- and KIVA-RIF-predicted OH and 94AL85000. Financial support was provided by the
soot distributions are different at the start of SSC DOE’s Office of FreedomCAR and Vehicle Technol-
but become similar during the latter part of the ogies, managed by Kevin Stork and Gurpreet Singh.
combustion event. The KIVA-CTC model always Additional financial support for the modelling activi-
predicts soot in a larger region compared with the ties at the University of Wisconsin was provided by
experiments and the other two models. The model Caterpillar Inc. The authors also thank David Cicone
also predicts slower soot oxidation. of Sandia National Laboratories for his assistance in
6. The KIVA-RIF model does not predict a lifted maintaining the optical-access research engine used
flame for the HTC condition. Unlike the exper- in these experiments.
imental observations, the predicted flame (as indi-
cated by OH radicals) extends fully upstream to
the injector. This is owing to a loss in spatial reso- REFERENCES
lution caused by assuming a single flamelet and
one global scalar dissipation rate for the entire 1 Nguyen, H. L., Addy, H. E., Bond, T. H., Lee, C. M.,
computational domain. and Chun, K. S. Performance and efficiency evalu-
7. The KIVA-CTC model predicts soot formation over ation and heat release study of a direct-injection
a greater region of the jet and slower soot oxi- stratified-charge rotary engine. SAE paper 870445,
dation. The soot in the jet is observed even long 1987.
2 Amsden, A. A., Butler, T. D., O’Rourke, P. J., and
after the end of combustion. This suggests that
Ramshaw, J. D. Kiva – a comprehensive model for
the two-step NSC model (based on ‘fuel’ as the 2-D and 3-D simulations. SAE paper 850554, 1985.
soot inception species) probably needs improve- 3 Wickman, D. D., Senecal, P. K., and Reitz, R. D.
ment to predict soot oxidation, especially for the Diesel engine combustion chamber geometry
LTC, low-sooting conditions. optimization using genetic algorithms and multi-

Int. J. Engine Res. Vol. 8 JER02406 © IMechE 2007

Downloaded from jer.sagepub.com at University of Birmingham on June 2, 2015


Validation of engine combustion models 125

dimensional spray and combustion modeling. SAE 18 Dec, J. E. Soot distribution in a D.I. diesel engine
paper 2001-01-0574, 2001. using 2-D imaging of laser-induced incandescence,
4 Kong, S. C., Han, Z. W., and Reitz, R. D. The devel- elastic scattering, and flame luminosity. SAE paper
opment and application of a diesel ignition and 920115. SAE Trans., 1992, 101(4), 101–112.
combustion model for multidimensional engine 19 Dec, J. E. and Espey, C. Ignition and early soot for-
simulations. SAE paper 950278. SAE Trans., 1995, mation in a D.I. diesel engine using multiple 2-D
104(3), 502–518. imaging diagnostics. SAE paper 950456. SAE Trans.,
5 Pitsch, H., Wan, Y. P., and Peters, N. Numerical 1995, 104(3), 853–875.
investigation of soot formation and oxidation under 20 Dec, J. E. and Zion, P. L. K. The effects of ignition
diesel engine conditions. SAE paper 952357. SAE timing and diluent addition on late-combustion
Trans., 1995, 104(4), 938–949. soot burnout in a DI diesel engine based on simul-
6 Tan, Z. and Reitz, R. D. Modeling ignition and com- taneous 2-D imaging of OH and soot. SAE paper
bustion in spark-ignition engines using a level set 2000-01-0238, 2000.
method. SAE paper 2003-01-0722, 2003. 21 Dec, J. E. and Espey, C. Chemiluminescence imag-
7 Amsden, A. A. KIVA-3V release 2, improvements to ing of autoignition in a DI diesel engine. SAE paper
KIVA-3V. Los Alamos National Laboratory report 982685. SAE Trans., 1998, 107.
LA-UR-99-915, 1999. 22 Pickett, L. M., Siebers, D. L., and Idicheria, C. A.
8 Singh, S., Reitz, R. D., and Musculus, M. P. B. Relationship between ignition processes and the lift-
Comparison of the characteristic time (CTC), rep- off length of diesel fuel jets. SAE paper 2005-01-
resentative interactive flamelet (RIF), and direct 3843, 2005.
integration with detailed chemistry models against 23 Patterson, M. A. and Reitz, R. D. Modeling the
optical diagnostic data for multi-mode combustion effects of fuel spray characteristics on diesel engine
in a heavy-duty DI diesel engine. SAE paper 2006- combustion and emissions. SAE paper 980131. SAE
01-0055, 2006. Trans., 1998, 107(3), 27–43.
9 Espey, C. and Dec, J. E. Diesel engine combustion 24 Han, Z. Y. and Reitz, R. D. Turbulence modeling of
studies in a newly designed optical-access engine internal combustion engines using RNG k–e models.
using high-speed visualization and 2-D laser imag- Combust. Sci. Technol., 1995, 106, 267–295.
ing. SAE paper 930971. SAE Trans., 1993, 102(4), 25 Reitz, R. D. Computational fluid dynamics modeling
703–723. of diesel engine spray combustion and emissions.
10 Dec, J. E. A conceptual model of D.I. diesel combus- In Diesel engine reference book (Eds B. Challen and
tion based on lased sheet imaging. SAE paper R. Baranescu), 2nd Edition, 1998 (Butterworth-
970873. SAE Trans., 1997, 106(3), 1319–1348. Heinemann, Oxford).
11 Musculus, M. P. B. On the correlation between NO 26 Halstead, M., Kirsh, L., and Quinn, C. The autoigni-
x
emissions and the diesel premixed burn. SAE paper tion of hydrocarbon fuels at high temperatures
2004-01-1401. SAE Trans., 2004, 113(4), 631–651. and pressures – fitting of a mathematical model.
12 Kimura, S., Aoki, O., Ogawa, H., Muranaka, S., and Combust. Flame, 1977, 30, 45–60.
Enomoto, Y. New combustion concept for ultra- 27 Kong, S. C. and Reitz, R. D. Multidimensional
clean and high-efficiency small DI diesel engines. modeling of diesel ignition and combustion using a
SAE paper 1999-01-3681. SAE Trans., 1999, 108(3), multistep kinetics model. J. Engng Gas Turbines
2128-2137. Power, 1993, 115(4), 781–789.
13 Kimura, S., Aoki, O., Kitahara, Y., and Aiyoshizawa, 28 Turns, S. R. An introduction to combustion, 2000
E. Ultra-clean combustion technology combining a (McGraw Hill, New York).
low-temperature and premixed combustion concept 29 Hiroyasu, H. and Kadota, T. Models for combustion
for meeting future emission standards. SAE paper and formation of nitric oxide and soot in DI diesel
2001-01-0200. SAE Trans., 2001, 110(4), 239–248. engines. SAE paper 760129. SAE Trans., 1976, 85.
14 Singh, S., Reitz, R. D., and Musculus, M. P. B. 30 Nagle, J. and Strickland-Constable, R. F. Oxidation
2-color thermometry experiments and high-speed of carbon between 1000 and 2000 °C. In Proceedings
imaging of multi-mode diesel engine combustion. of the Fifth Carbon Conference, Vol. 1, 1962, p. 154
SAE paper 2005-01-3842, 2005. (Pergamon Press, London).
15 Espey, C. and Dec, J. E. The effect of TDC tempera- 31 Patel, A., Kong, S. C., and Reitz, R. D. Development
ture and density on the liquid-phase penetration in and validation of a reduced reaction mechanism for
a D.I. diesel engine. SAE paper 952456. SAE Trans., HCCI engine simulations. SAE paper 2004-01-0558,
1995, 104(4), 1400–1414. 2004.
16 Dec, J. E. and Coy, E. B. OH radical imaging in a DI 32 Govovitchev, V. I. Chalmers University of
diesel engine and the structure of early diffusion Technology, Gothenburg, Sweden, 2000. http:
flame. SAE paper 960831. SAE Trans., 1996, 105(3), //www.tfd.chalmers.se/~valeri/MECH.html.
1127–1148. 33 Kee, R. J., Rupley, F. M., and Miller, J. A.
17 Musculus, M. P. B. Multiple simultaneous optical CHEMKIN-II: a FORTRAN chemical kinetics pack-
diagnostic imaging of early-injection low- age for the analyses of gas phase chemical kinetics.
temperature combustion in a heavy-duty diesel Sandia Report, SAND 89-8009, 1989.
engine. SAE paper 2006-01-0079, 2006. 34 Kong, S. C. and Reitz, R. D. Use of detailed chemical

JER02406 © IMechE 2007 Int. J. Engine Res. Vol. 8

Downloaded from jer.sagepub.com at University of Birmingham on June 2, 2015


126 S Singh, R D Reitz, M P B Musculus, and T Lachaux

kinetics to study HCCI engine combustion with con- studies in a newly designed optical-access engine
sideration of turbulent mixing effects. J. Engng Gas using high-speed visualization and 2-D laser imag-
Turbines Power, 2002, 124(3), 702–707. ing. SAE paper 930971, 1993.
35 Kong, S. C. and Reitz, R. D. Application of detailed
chemistry and CFD for predicting direct injection
HCCI Engine combustion and emissions. In Pro- APPENDIX
ceedings of the 29th International Symposium on
Combustion, 21–26 July 2003, pp. 663–669.
36 Smith, G. P., Golden, D. M., Frenklach, M., Notation
Moriarty, N. W., Eiteneer, B., Goldenberg, M., AHRR apparent heat-release rate
Bowman, C. T., Hanson, R. K., Song, S., Gardiner,
ATDC after top dead centre
W. C., Lissianski, V. V., and Qin, Z. 2000. http:
//www.me.berkeley.edu/gri_mech/. BB broadband
37 Kong, S. C., Sun, Y., and Reitz, R. D. Modeling diesel BDC bottom dead centre
spray flame lift-off, sooting tendency and NO emis- CAD crank angle degree
x
sions using detailed chemistry with phenomeno- CF cool flame
logical soot model. ASME ICES2005-1009, 2005. CFD computational fluid dynamics
38 Pitsch, H., Barths, H., and Peters, N. Three dimen- CI compression ignited
sional modeling of NOx and soot formation in CTC characteristic time combustion
DI-diesel engines using detailed chemistry based
DI direct injection
on the interactive flamelet approach. SAE paper
962057. SAE Trans., 1996, 105(4), 2010–2025. DOI duration of injection
39 Barths, H., Antoni, C., and Peters, N. Three- EGR exhaust gas recirculation
dimensional simulation of pollutant formation in a EOI end of injection
DI diesel engine using multiple interactive flamelets. HTC high-temperature combustion
SAE paper 982459. SAE Trans., 1998, 107(4), 984– ID ignition delay
994. IMEP indicated mean effective pressure
40 Demoulin, F. X. and Borghi, R. Presumed PDF
IVC intake valve closure
modeling of turbulent spray combustion. Combust.
Sci. Techol., 2000, 158, 249–271. KH–RT Kelvin Helmholtz–Rayleigh Taylor
41 Kong, S. C., Kim, H., Reitz, R. D., and Kim, Y. LII laser-induced incandescence
Comparisons of combustion simulations using a LMS laser Mie scattering
representative interactive flamelet model and direct LTC low-temperature combustion
integration of CFD with detailed chemistry. ASME LL liquid length
ICES2005-1010, 2005. MCC mixing-controlled combustion
42 El Tahri S. H. A numerical study on the effects of MK modulated by kinetics
fluid motion at inlet-valve closure on subsequent
NO nitric oxides
fluid motion in a motored engine. SAE paper x
820035, 1982. NSC Nagle–Strickland-Constable
43 Siebers, D. L. Liquid-phase fuel penetration in PAH polycyclic aromatic hydrocarbons
diesel sprays. SAE paper 980809. SAE Trans., 1998, PDF probability density function
107(3), 1205–1227. PLIF planar laser-induced fluorescence
44 Tao, F. and Chomiak, J. Numerical investigation of RIF representative interactive flamelet
reaction zone structure and flame liftoff of DI diesel RNG re-normalized group
sprays with complex chemistry. SAE paper
SI spark ignited
2002-01-1114, 2002.
45 Higgins, B., Siebers, D., and Aradi, A. Diesel-spray SOC start of combustion
ignition and premixed-burn behavior. SAE paper SOI start of injection
2000-01-0940, 2000. SSC second-stage combustion
46 Espey, C. and Dec, J. E. Diesel engine combustion TDC top dead centre

Int. J. Engine Res. Vol. 8 JER02406 © IMechE 2007

Downloaded from jer.sagepub.com at University of Birmingham on June 2, 2015

También podría gustarte