Sie sind auf Seite 1von 13

Downloaded from SAE International by Univ of Calif-Los Angeles, Saturday, September 03, 2016

SAE TECHNICAL
PAPER SERIES 980789

Modelling the Influence of Fuel


Injection Parameters on Diesel Engine
Emissions
D. K. Mather, R. D. Reitz
University of Wisconsin-Madison

International Congress and Exposition


Detroit, Michigan
February 23-26, 1998

400 Commonwealth Drive, Warrendale, PA 15096-0001 U.S.A. Tel: (724) 776-4841 Fax: (724) 776-5760
Downloaded from SAE International by Univ of Calif-Los Angeles, Saturday, September 03, 2016

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

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

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

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

All SAE papers, standards, and selected


books are abstracted and indexed in the
Global Mobility Database

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

ISSN 0148-7191
Copyright 1998 Society of Automotive Engineers, Inc.

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

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

Printed in USA
Downloaded from SAE International by Univ of Calif-Los Angeles, Saturday, September 03, 2016
980789
Modeling the In uence of Fuel
Injection Parameters on Diesel Engine
Emissions
D. K. Mather and R. D. Reitz
Engine Research Center,

University of Wisconsin-Madison

Abstract to improve engine performance. Zolver et al.[1] 1 success-


fully used 3-dimensional modeling to help assess the e ects
Rate shaping of the fuel injection process is known to signif- of piston bowl shape for a turbo-charged direct-injected
icantly impact emissions production in diesel engines. To diesel engine. Some other recent examples of the applica-
demonstrate the ability of multidimensional engine model- tion of multidimensional modeling to the understanding of
ing to quantify and explain the e ect of rate shaping and in- direct injected diesel engines include the study of Han et
jection duration, three injection pro les typical of common al. [2] where multidimensional modeling was employed to
diesel fuel injection systems were investigated for three in- identify the mechanism by which pulsed fuel injection is
jection durations and injection timings. The present study able to reduce soot while maintaining low NOx . Senecal
uses an improved version of the KIVA-II engine simulation et al. [3] explained the bene ts on combustion and emis-
code employing the characteristic time combustion model, sions of a novel engine design utilizing many fuel injectors
the Kelvin-Helmholtz and the Rayleigh-Taylor spray atom- directed tangentially to the piston bowl in order to induce
ization mechanisms, the extended Zeldovich thermal NOx swirling ow. Another aspect of the study of Senecal et al.
production model, and a single species soot model. To- [3] was to use the detailed information about the location of
gether, these models satisfactorily reproduce the known soot from their simulation to propose modi cations to the
trends in engine behavior with respect to injection rate combustion chamber geometry. The bene ts of the modi-
shaping and injection duration for a representative high- ed engine geometries were demonstratead by simulation.
load high-speed operating condition at a xed fueling rate; In the present work multidimensional modeling is used to
namely, increased NOx with decreased injection duration; explore the in uence of injection characteristics on diesel
decreased soot production with decreased injection dura- engine performance and emissions. The results could be
tion. At high injection pressures, the falling injection pro- useful to suggest injection parameters for improved engine
le produces more NOx and less soot than either the square performance.
or rising injection pro le. The rising injection pro le pro-
duces more soot and less NOx than the square pro le. The
computational results are also used to explain the origin Model Description
of anomalous soot-NOx tradeo trends where soot minima
are seen as the injection timing is varied. The numerical calculations were performed using a mod-
i ed version of the KIVA-II code [4] with improvements
Introduction for the turbulence, gas-to-wall heat transfer, ignition, and
combustion models as described in Han et al. [2] A recently
In a compression ignition engine the injection velocity, du- improved soot model and spray model were also included.
ration, and timing in uence the atomization, vaporization, The models used in the present study are summarized be-
mixing, distribution, and combustion of the injected fuel. low.
These injection characteristics dominate the ignition be-
havior, emission levels, and power output of a particular Ignition Model
engine. As pollutant emission regulations have become
more stringent it has become increasingly more dicult In the study of Kong and Reitz [5], the `Shell' auto-
to meet emission targets with an economical and ecient ignition model [6] was adapted for diesel ignition and imple-
engine. One of the tools that is becoming more useful mented into KIVA-II. This model introduces three generic
for investigating engine phenomena is multidimensional en- species with reactions to represent hydrocarbon ignition.
gine modeling. These models are able to reproduce the The eight reactions comprising this model are given below:
global behavior of the engine (e.g., pressure, heat release, Initiation:
engine-out emissions) and provide insight into the under- RH + O2 ! 2R (1)
lying causes and mechanisms responsible for the trends to
increase our understanding of diesel engines. 1 Numbers in brackets denote References listed at the end of the

There has been much progress in using CFD modeling paper.

1
Downloaded from SAE International by Univ of Calif-Los Angeles, Saturday, September 03, 2016

Propagation: accurate predictions of combustion heat release. A sim-


R ! R + P (2) pli cation that had been previously made is that c;i was
R ! R + B (3) assumed to be constant for all of the six reactive species
considered: fuel, O2 , H2 O, H2 , CO , CO2 . The basic
R ! R + Q (4) assumption in the above formulation is that chemical re-
R + Q ! R + B (5) actions and mixing processes are identical for each of the
species. In a more general case for the six species and three
Branching: atom-balance system, it can be shown that each species
B ! 2R (6) approaches its respective equilibrium density at a rate de-
Linear Termination: termined by one of three distinct characteristic times [10].
According to the present characteristic time combustion
R ! non-reactive species (7) model, the rates of change for the densities of fuel, CO,
and H2 are given by
Quadratic Termination:
@fuel = , fuel , fuel (12)
2R ! non-reactive species (8) @t c;fuel
where RH and P are the hydrocarbon and the products of @CO = , CO , CO (13)
combustion, respectively. The generic ignition species are @t c;CO
R , B and Q which represent the radical formed from the
fuel, the branching agent, and a labile intermediate species, @H2 = , H2 , H2 (14)
respectively. In the present implementation, the heats of @t c;H2
formation of the generic species are neglected. Thus, the Then, the rates of change of the species CO2 , H2 O , and
only reactions that consume or produce thermal energy are O2 can be determined from the atom conservation equa-
Eqs. 1 and 2. tions,
It has been found that the formation rate of the labile
species, Q, of Eq. 4 is the rate limiting step in the kinetic @CO2 = , @CO , n @fuel (15)
path. The formation rate of Q is given by @t @t @t
rateQ;formation = f4 Kp (9) @H2 O = , @H2 , (n + 1) @fuel (16)
@t @t @t
where  ,E 
f4 = Af 04 exp f4 x4 y4  
<T [O2 ] [R] (10) @O2 = 1 @CO + 1 @H2 + 3n + 1 @fuel (17)
@t 2 @t 2 @t 2 @t
and Ef 4 = 3:0  104 , x4 = ,1 , y4 = 0:35 , and the
rate Kp is given by Theobald and Cheng [7] and Kong where n indicates the carbon atom number of a straight
et al. [8]. The sensitivity of the ignition delay to the Q chained hydrocarbon fuel molecule.
formation rate was investigated by Theobald and Cheng In the standard implementation of the characteristic
[7] who found that the total delay is sensitive to the pre- time combustion model, the characteristic time, c;i is the
exponential factor Af 04 . The value used in this present same for all species and is the sum of a turbulent-mixing
study is 6:5  105 which is similar to that used by Kong et time-scale and a laminar chemical-kinetics time-scale given
al [8]. by
Combustion Model c = c;i = laminar + fturbulent (18)
Combustion was simulated using the characteristic time In the multiple time-scale combustion model a similar
combustion model, initially developed for spark ignited en- formulation is used where the time-scales for CO and H2
gines by Abraham et al. [9] and adapted to diesel combus- conversion are assumed to be 10% of the fuel time-scale
tion by Kong et al. [8], [10].
In Eq. 18, the longer of the two time-scales has more
@i = , i , i (11) of an in uence on the conversion rate. The laminar time-
@t c;i scale for diesel combustion is based on a correlation for the
where i is the local instantaneous thermodynamic equi- laminar ame speed of a single burning fuel droplet given
librium value of the species' partial density [10] assuming by
that the local temperature remains constant, and c;i is the  
characteristic time for the ith species to reach that equilib- laminar = A [Cn H2n+2 ] 4 [O2 ], 2 exp <ET
3 3
(19)
rium density. The computations were limited to only the
major chemical species, whose inclusion is necessary for

2
Downloaded from SAE International by Univ of Calif-Los Angeles, Saturday, September 03, 2016

where A = 1:625  10,11 is a model constant, < is the Emissions Models


universal gas constant, and E = 30; 000cal=mole is the
activation energy [8]. THERMAL NOx : The current approach to modeling NOx
Kuo and Reitz [11] studied the e ect of residual gas emissions from diesel engines is to use the extended Zel-
concentration on the laminar characteristic time in com- dovich thermal NO mechanism and neglect other sources
putations of 2-stroke engine combustion with high levels of NOx formation. When comparing to experimental NOx
of residual gases and high EGR rates. They found that data, a scaling factor equal to the ratio of molecular weights
the laminar characteristic time increases with residual gas of NO and NO2 is applied to the NO predictions. The
concentration. This e ect was accounted for by dividing extended Zeldovich mechanism consists of the following
the pre-exponential constant, A in Eq. 19 by a correction mechanisms, as described by Bowman [13]:
factor, CF , given by O + N2 $ NO + N (24)
CF = 1:27 (1 , 2:1 ) (20) N + O2 $ NO + O (25)
where is the residual mass fraction (i.e., = massresidual N + OH $ NO + H (26)
/ masstotal .) This modi cation was chosen to be consis- with the concentrations of H , OH , and O determined from
tent with Metghalchi and Keck's correlation [12] for the
retarding e ect of residual mass fraction on laminar ame O + OH $ O2 + H (27)
speed and is adopted in the present study.
The turbulent time-scale, turbulent , is assumed to be and assuming equilibrium. The concentration of N is as-
linearly proportional to the turn-over time of the charac- sumed to be at steady state (i.e., @N=@t = 0) in the rate
teristic turbulent eddy. For the k ,  type of turbulence equations resulting from the Zeldovich mechanism. This
model, the eddy turn-over time is the quotient of the tur- mechanism can be rewritten as an explicit expression for
bulent kinetic energy, k, and the turbulent dissipation rate, the rate of change of the concentration of NO [14]:
, and the turbulent time-scale becomes 0 1
@ [NO] = 2k [O][N ] @ 1 , K12[NO ] 2
[O2 ][N2] A (28)
turbulent = C2 k @t 1f 2 b [NO ]
(21) 1 + k2f [Ok21]+ k3f [OH ]
where C2 = 0:1 is a model constant [8]. where K12 = k1f =k1b  k2f =k2b , and the subscripts 1,2 and
Finally the function f that appears in Eq. 18 is a delay 3 refer to Eqs. 24, 25, and 26 ,respectively. The recom-
coecient, that was found essential for adapting the char- mended rate constants are given by Bowman [13]:
acteristic time combustion model to diesel engines, given ,
k1f = 7:6  1013exp ,38T;000

by (29)
k1b = 1:6  1013
f = 11 , e, (22)
, 
k2f = 6:4  109Texp , ,3T;150 
, e,1 k2b = 1:5  109 Texp ,19T;500 (30)
where is a progress variable for the chemical reactions,
equal to the ratio of the mole fraction of combustion prod- k3f = 1:0  1014 ,  (31)
ucts to the total mole fraction of reacting species, k3b = 2:0  1014exp ,23T;650
= 1Y,COY2 ;R + Y+HY2 O;R + +YCO;R + YH2 ;R (23) SOOT FORMATION : Hiroyasu et al. [15] developed a
CO2 ;r H2 O;r YCO;r + YH2 ;r simple soot model which was later applied to multidimen-
where the subscripts R and r indicate whether the source of sional diesel combustion by Belardini et al. [16]. The
the chemical species is either chemical reaction or residual model predicts the soot mass production rate, @Ms =@t, by
gas, respectively. a single-step reaction composed of a soot formation rate,
Physically, f accounts for uid that is already mixed Msf , and a soot oxidation rate, Mso , according to
on a large scale comparable to the turbulent scales, and @Ms = M , M
mixing on the molecular level that is responsible for the sf so (32)
laminar ame speed. Early in the combustion, the chemi- @t
cal time-scale is dominated by the laminar time-scale. As The Arrhenius formation rate is proportional to the fuel
the reactions progress, more products of combustion are vapor mass, Mfv , given by
present and the chemical time-scale becomes dominated Msf = Kf Mfv (33)
by the turbulent time-scale. Without f , the characteristic
time combustion model is not able to capture the premixed where the formation coecient is a function of both pres-
portion of diesel combustion adequately. sure and temperature
 ,E 
Kf = Asf P exp
1
2
sf (34)
<T
3
Downloaded from SAE International by Univ of Calif-Los Angeles, Saturday, September 03, 2016

with Esf = 12500, and Asf is a model constant. We is the Weber number for the gas (We = g u2r r0 =) with
ur the relative velocity between the gas and the drop, and
SOOT OXIDATION : The Hiroyasu soot model has been p is the surface tension. Z is the Ohnesorge number (Z =
modi ed in the present study by employing the Nagle Strick- Wel =Rel), where Wel is the Weber number based
p on the
land - Constable (NSC) soot oxidation model. In this liquid density. T is the Taylor number (T = Z We.)
model, soot oxidation occurs by two surface reaction mech- The breakup time,  , is tted to two regimes: bag
anisms: at more reactive `A' sites and at less reactive `B' breakup at low We, and stripping breakup at high We.
sites. Three reaction equations are formulated to describe The breakup time is expressed as
the oxidation process with the net rate given as
 =
3:788B1ro (38)
M = 6 (MWs ) M R
so s Ds s total (35) K ,H K ,H
where MWs is the molecular weight of soot (carbon 12 with B1 = 60 a constant chosen to match experimental
g/mole), s is the soot density (2 g/cc), Ds is the soot di- data, which re ects initial disturbances to the drop surface
ameter (3.0e-6 cm), and Ms is the soot mass. The details caused by ow within the nozzle [19]. The drop size shed
of the net reaction rate Rtotal are given by Patterson et al. during breakup, rshed , is assumed to be proportional to the
[17]. Other soot oxidation models are available that are wavelength of Eq. 36
based on mixing rates (e.g., Hampson and Reitz [18]). The rshed = B0 K ,H (39)
NSC soot oxidation model used in the present study has
the characteristic of predicting lower soot levels in diesel where B0 is an empirical constant taken to be 0.61. Finally,
engines at retarded injection timings than are observed ex- the rate equation governing the change in the parent drop
perimentally. This characteristic has been traced to the size, r, is expressed as
model's sensitivity to the given value of the soot diame- @r = r , rshed
ter, an ostensibly unknown quantity for diesel combustion. (40)
The soot diameter a ects soot oxidation by determining @t 
the surface area of the soot mass available for oxidation. The R-T breakup mechanism [20] is employed to ac-
Computational results employing the NSC model with a count for the e ect of rapid deceleration of the drops on
smaller soot diameter (0.6e-6 cm) predicts higher soot lev- the atomization process. This mechanism is cast in the
els at retarded injection timings than the NSC model in same form as the K-H mechanism. The fastest growing
the present study, but still predicts lower soot levels than wavelength is given by
are observed experimentally. r 3
Spray Model  = 2 R,T al (41)
Spray atomization is a complicated phenomenon to model. and the fastest growing frequency is given by
It is very sensitive to fuel properties and ambient conditions r r

R,T = 2a al
as well as to the details of the fuel injector nozzle design and
operating conditions. As the spray is the dominant feature 3 3
4
(42)
of the ow within a diesel engine combustion chamber, ac-
curately resolving the spray atomization is a prerequisite where the drop acceleration is given by
for diesel engine simulation. The approach to modeling at- 2
omization in this study is to assume that the mechanisms a = 83 Cd g urr . (43)
responsible for the breakup of the spray are classical uid l
dynamic instabilities which act at the interface between The breakup time is found by taking the reciprocal of the
two uids of di erent densities: the Kelvin-Helmholtz (K- frequency of Eq. 42, and the drop is assumed to break into
H) inertia instability and the Rayleigh-Taylor (R-T) accel- two drops of equal size. The K-H and R-T mechanisms
erative instability. For both of the instabilities, the time compete with each other to break up the spray. More de-
that the spray droplet breaks up is determined from the tails of the breakup mechanisms are described in Su et al.
growth rate of the fastest growing wavelength predicted by [20], and Patterson and Reitz [21].
the classical instabilities. For the K-H mechanism the wave Another aspect of modeling the spray atomization pro-
length, K ,H of the fastest growing wave is given by Reitz cess is the modeling of the intact core found close to the
[19] as nozzle. In the present study, the intact core is accounted
p for by delaying the onset of breakup by the R-T mecha-
K ,H = 9:02r0(1 (1 + 0:45 Z )(1 + 0:4T 0:7) (36) nism until the spray droplet has moved beyond the intact
+ 0:865We1:67)0:6 core length, L. An expression for L is found by applying
and its frequency,
K ,H , is given by the theory of Levich [22] as
1:5 r  r
0 : 34 + 0 : 385 We

K ,H = (1 + Z )(1 + 1:4T 0:6)  r3 (37) L = C l do (44)
l 0 g

4
Downloaded from SAE International by Univ of Calif-Los Angeles, Saturday, September 03, 2016

where C is a model constant which accounts for the details


of the nozzle design and operating condition, l and g 1.0e+05

are the liquid and gas densities, respectively, and do is the


baseline
long
initial diameter of the intact core taken to be the nozzle 8.0e+04 medium
hole diameter. More details of this model are given by
short

injection velocity (cm/s)


Ricart et al. [23]. 6.0e+04

Combustion Modeling of Injection 4.0e+04

Rate Shaping 2.0e+04


Fuel injection characteristics have a signi cant e ect on the
performance and emissions of diesel engines. To demon-
strate the ability of multidimensional engine modeling to
0.0e+00
-5 0 5 10 15 20 25
quantify and explain the e ects of rate-shaping and injec- crank angle (degrees ATDC @ 1600 rpm)
tion duration on the emissions from a heavy duty diesel en-
gine, three injection pro les characteristic of typical diesel (a) falling
fuel injectors were investigated with three injection du-
rations. Figure 1(a) shows a falling injection rate, char- 1.0e+05
acteristic of accumulator-type common-rail fuel injectors. baseline
Figure 1(b) shows a rising injection rate, characteristic of long
medium
mechanical or electric jerk-type fuel injectors. Figure 1(c) 8.0e+04
short

injection velocity (cm/s)


shows a square injection rate, characteristic of common rail
fuel injectors [24]. (The solid line on Figures 1(a) - 1(c) is 6.0e+04
the measured injection pro le for a common rail fuel injec-
tor installed on the engine for which the models have been 4.0e+04
calibrated [2].)
A range of fuel injection timings and injection durations
were simulated in the present study to generate soot-NOx 2.0e+04
tradeo curves for each modeled fuel injection system oper-
ating in the baseline engine. These injection pro les were 0.0e+00
simulated for a mid-range diesel truck engine at a xed -5 0 5 10 15 20 25
equivalence ratio (i.e., a xed fueling rate.) The speci -
crank angle (degrees ATDC @ 1600 rpm)

cations of the engine are given in Table 1. The multidi- (b) rising
mensional model used has the demonstrated capability of
reproducing the experimental data for the engine with a
common rail injector. This is shown in Figs. 2(a), 2(b), 1.0e+05
and 2(c) for the cylinder pressure, heat release and soot- baseline
long
NOx tradeo , respectively. These data is included here for 8.0e+04 medium
comparison to the other injection pro les. short
injection velocity (cm/s)

Figure 2(c) also shows the in uence of the choice of as-


sumed soot particle size in the NSC soot oxidation model, 6.0e+04

Eq. (35). The agreement between the measured and com-


puted soot emissions is good at advanced timings, but de- 4.0e+04
teriorates at retarded injection timings, as also discussed
by Hampson and Reitz [18]. 2.0e+04
The computational grid used in the present study is
shown along with a typical distribution of fuel droplets in
the combustion chamber in Figure 3 for the baseline injec- 0.0e+00
tion case at 6 degrees ATDC where the injection started
-5 0 5 10 15 20 25
crank angle (degrees ATDC @ 1600 rpm)
at -9 degrees ATDC. These calculations take advantage of
the six-fold symmetry from the six-hole centrally located (c) square
fuel injector by simulating only a one-sixth sector of the
combustion chamber. Figure 1: Three injection pro les with three injection dura-
The predicted tradeo curves of the engine-out soot and tions: long (21.5 CA), medium (17.55 CA), and short (15.2
NOx for the various injection pro les are shown in Fig. 4 CA). The solid line is the baseline experimental injection
with three injection timings: -10 CA ATDC, -5 CA ATDC, pro le included for comparison.
5
Downloaded from SAE International by Univ of Calif-Los Angeles, Saturday, September 03, 2016

12
measured
10 computed
pressure (MPa)

2
Figure 3: Perspective view of computational grid and fuel
0 droplet distribution at 6 degrees ATDC (-9 CA ATDC
-100 -50 0 50 100 SOI). Computations are on a one-sixth sector of the com-
crank angle (degrees ATDC) bustion chamber.
(a) pressure
and TDC. The falling injection pro les, Fig.4(a), were the
most sensitive to changes in the injection duration. As can
700 been seen, the long duration falling injection produced the
600
measured
computed
highest soot levels, while the short duration falling injec-
tion produced the lowest soot levels of the three injection
heat release rate (J/deg)

500 pro les considered. The tradeo for the rising injection
pro les, Fig. 4(b), shows the general trends of increas-
ing the injection duration decreasing the NOx levels and
400

300 increasing the soot levels. The tradeo for the square injec-
tion pro les, Fig. 4(c), show the same trends as the rising
200
injection pro les with respect to changes in the injection
100 duration.
The measured data for the experimental engine con g-
0
-20 -10 0 10 20 30 40 50 60
ured with the common rail injector are repeated on the
crank angle (degrees ATDC) tradeo curve for the square injection in Fig. 4(c) showing
that the actual injection is similar in performance to the
(b) heat release middle duration square injection case.
An interesting feature of the tradeo curve for the long
injection duration falling pro le is that there is a minimum
2.2
in the soot levels. In fact, there is a portion of that tradeo
2 measured
computed (3.0e-6 cm) curve where soot level decreases with retarding fuel injec-
1.8 computed (0.6e-6 cm) tion timing that will be explained next.
1.6
The combustion mechanisms responsible for the ob-
soot (g/kg-fuel)

1.4
served trends are re ected in the NOx and soot evolution
1.2
curves for the di erent injection rate shapes. Figure 5
1
presents emission results for the three di erent injection
0.8
pro les with the short injection duration of 15.2 CA and
0.6
-5 CA SOI. The results show that the predicted total in-
0.4
cylinder NOx evolution time (Fig. 5(a)) for the falling
0.2
injection pro le rises about 50% higher than the curves for
0
15 20 25 30 35 40 45 50 55 60 65 either the square or the rising injection pro les. Figure
NOx (g/kg-fuel) 5(b) shows that the total incylinder soot evolution curve
for the falling injection pro le reaches a maximum of only
(c) soot-NOx tradeo 60% of the maximum for the square and rising injection
pro les.
Figure 2: Comparison between the computed and mea- These emissions evolution results can be explained by
sured (a) pressure, (b) heat release and (c) emissions. (a) considering the local instantaneous equivalence ratios and
and (b) are for an injection timing of -9 degrees ATDC. temperatures within the combustion chamber. Figure 6(a)
(c) shows engine out emissions for injection timings of -12, shows that much less of the combustion chamber gas mass
-9, -6, -3, and 0 degrees ATDC for the NSC soot oxidation for the falling injection pro le is rich (i.e., above 2.5 equiv-
model with soot diameters of 3.0e-6 cm, and 0.6e-6 cm. alence ratio) compared to either the rich mass of the rising

6
Downloaded from SAE International by Univ of Calif-Los Angeles, Saturday, September 03, 2016

1.2
long
1 medium
short
soot (g/kg-fuel)

0.8

0.6

0.4
70

0.2 60

0 50

NOx (g/kg-fuel)
0 10 20 30 40 50 60 70 80 90
NOx (g/kg-fuel) 40

(a) falling rate-of-injection 30 falling


rising
square
20
1.2
long 10
1 medium
short 0
-20 0 20 40 60 80 100
soot (g/kg-fuel)

0.8 crank angle (degrees ATDC)

0.6 (a) NOx evolution


0.4
2.5
falling
0.2 rising
2 square
0
0 10 20 30 40 50 60 70 80 90
soot (g/kg-fuel)

NOx (g/kg-fuel) 1.5

(b) rising rate-of-injection 1

1.2 0.5
long
1 medium
short 0
baseline computed -20 0 20 40 60 80 100
soot (g/kg-fuel)

0.8 crank angle (degrees ATDC)

0.6 (b) soot evolution


0.4
Figure 5: NOx and soot evolution with varying injection
0.2
pro les for 15.2 CA duration injection at -5 CA ATDC SOI

0
0 10 20 30 40 50 60 70 80 90
NOx (g/kg-fuel)

(c) square rate-of-injection

Figure 4: Predicted soot-NOx tradeo curves for the three


injection rate shapes at three injection durations.
7
Downloaded from SAE International by Univ of Calif-Los Angeles, Saturday, September 03, 2016

or square injection pro les. Less rich gas mass within the
combustion chamber translates into less soot producing re-
gions within the combustion chamber.
To explain the NOx results, Fig. 6(b) shows the mass
of the cylinder gas which has temperature above 2800K for
the three di erent injection schemes with the short injec-
tion duration and -5 SOI. The results indicate that a higher
percent of the cylinder mass is above 2800K for the falling
pro le than for either the square or the falling pro les, and
since NOx formation is highly temperature dependent, this

percent cylinder mass at equivalence ration > 2.5


9
explains the higher NOx of the falling pro le. 8
falling
rising
square

Reduced Soot at Retarded Injection


7

Timings
6
5

Finally, the mechanism responsible for the anomalous trend 4


of decreasing soot with retarding injection timing for the 3
falling rate seen in Fig. 4(a) for the long injection duration
can be explained by showing the di erences in the spatial 2
variation of the soot elds. Figure 7 shows contours de n- 1
ing the location of the soot cloud at 39 CA ATDC which
is a representative time during the soot oxidation period 0
-10 -5 0 5 10 15 20 25 30
for the long duration falling pro le for the three injection crank angle (degrees ATDC)
timings, -10 CA, -5 CA, and TDC, respectively. At both
the -10 CA and the TDC injection timings, the soot cloud (a) Percent cylinder mass above equivalence ratio of 2.5
is located in the bottom of the piston bowl. In this case
the soot is located in a region devoid of oxygen and soot
oxidation is inhibited. For the -5 CA injection timing, the
1.6

soot cloud is has been swept though the combustion cham- 1.4
ber and is located near the head where there is available
percent cylinder mass > 2800 K

oxygen which promotes soot oxidation. The di erence in 1.2

location of the soot is the reason for the anomalous de- 1


crease in soot levels as the injection timing was retarded
from -10 CA to -5 CA. For early injection with a long in- 0.8
jection duration the lower spray momentum con nes the 0.6
combustion to the bowl as seen in Fig. 7(a). As the in-
jection timing is retarded, the spray interacts more with 0.4
the reverse squish ow, moving the soot cloud nearer to
the head, as seen in Fig. 7(b). Finally, at very retarded 0.2

timings, the spray momentum is again directed into the 0


piston bowl and the soot remains away from the head, as -10 -5 0 5 10 15 20 25 30
seen in Fig. 7(c). crank angle (degrees ATDC)

Interestingly, the short injection duration falling pro- (b) Percent cylinder mass above 2800K
les do not show a decrease in soot levels with retarding
injection timing. This can be explained by the location of
the soot cloud in comparison to its location for the long Figure 6: Evolution of rich and high temperature regions
injection duration case. For the short duration injection, with varying injection pro les for 15.2 CA duration injec-
the soot cloud is located near the head of the combustion tion at -5 CA ATDC SOI
chamber for all of the injection timings, as can be seen in
Fig. 8. Evidently, the higher spray momentum of the short
duration injections is sucient for the spray to penetrate
farther into the combustion chamber such that more of the
soot cloud reaches the head region in all cases.
This mode change from the soot being located on the
cylinder head to the soot being located in the piston bowl
is shown conceptually in Fig. 9. By changing the injec-

8
Downloaded from SAE International by Univ of Calif-Los Angeles, Saturday, September 03, 2016

(a) -10 CA ATDC SOI (a) -10 CA ATDC SOI

(b) -5 CA ATDC SOI (b) -5 CA ATDC SOI

(c) TDC SOI (c) TDC SOI

Figure 7: Soot contours for the long duration falling injec- Figure 8: Soot contours for the short duration falling injec-
tion pro le at 39 CA ATDC for di erent start-of-injection tion pro le at 39 CA ATDC for di erent start-of-injection
timings (SOI). timings (SOI).
9
Downloaded from SAE International by Univ of Calif-Los Angeles, Saturday, September 03, 2016

tion timing and changing the location of the soot during


its oxidation period, the engine is e ectively moved to a
di erent soot-NOx tradeo curve, as indicated schemati- injector
cally in Fig. 9. In this case the tradeo curve is higher for
the soot located near the piston than for the soot located
on the head.

Conclusions
soot cloud

The injection pro les from three typical classes of diesel


fuel injector were simulated for a range of fuel injection
timings and injection durations. Overall, the predicted ef-
fects on emissions of decreasing the injection duration is to (a) retarded injection
increase NOx and to decrease soot, consistent with mea-
sured data. The falling injection pro les show the greatest
sensitivity to decreasing the injection duration. The falling injector
injection pro le results in higher soot levels than either the
square or rising injection pro les at the same injection du-
ration. However, at the shortest injection duration inves- soot cloud
tigated (i.e., highest injection pressure) the falling pro le
produced the lower soot levels than either the square or
the rising injection pro les. The rising injection pro les all
have higher soot and lower NOx than the square pro les
at the same injection duration. Similarly, the trend of in-
creasing soot and decreasing NOx can be seen in many of
the tradeo curves. Finally, the phenomenon of decreasing
soot with retarding fuel injection timing has been identi- (b) early injection
ed and the underlying mechanism proposed. It is shown
to be due to a regime transition where the soot cloud is
located in di erent regions in the combustion chamber.
Table 1 Caterpillar Engine Speci cations
bore 13.719 cm
stroke 16.510 cm
connecting rod length 26.162 cm (1)
displacement 2.44 liters
soot

compression ratio 15:1


piston crown Mexican hat
engine speed 1600 rpm
intake pressure 1.84 bar mode change
injection duration 22.5 degrees (2)
fuel injected 0.1625 grams/cycle

Acknowledgments
This work was supported by Caterpillar Inc., with addi- NO x
tional funding from the Army Research Oce. The au-
thors thank Drs. Simon Chen, Sudhakar Das and Andre (c) Mode shift from one soot-NOx tradeo to another
Kazakov for helpful discussions and comments.
Figure 9: Mode shift with the soot cloud located near the
References cylinder head to soot located in the piston bowl. The com-
1. Zolver, M., and Griard, C., and Henriot, S., \3D bustion jumps from one soot-NOx tradeo curve at (1 -
Modeling Applied to the Development of a DI Diesel piston bowl) to a lower one at (2 - combustion chamber
Engine: E ect of Piston Bowl Shape," SAE 971599, head)
1997.

10
Downloaded from SAE International by Univ of Calif-Los Angeles, Saturday, September 03, 2016

2. Han, Z., Uludogan, A., Hampson, G. J., and Reitz, 15. Hiroyasu, H., and Kadota, T., \Models for Combus-
R. D., \Mechanisms of Soot and NOx Emission Re- tion and Formation of Nitric Oxide and Soot in DI
duction Using Multiple-Injection in a Diesel Engine," Diesel Engines," SAE 760129, 1976.
SAE 960633, 1996.
16. Belardini, P., Bertoli, C., Cameretti, M. C., and Gi-
3. Senecal, P. K., Uludogan, A., and Reitz, R. D., \De- acomo, N. D., \A Coupled Diesel Combustion and
velopment of Novel Direct-Injection Diesel Engine Soot Formation Model for KIVA-II Code: Charac-
Combustion Chamber Designs Using Computational teristics and Experimental Validation," International
Fluid Dynamics," SAE 971594, 1997. Symposium COMODIA 94, 315-323, 1994.
4. Amsden, A. A., O'Rourke, P. J., and Butler, T. D., 17. Patterson, M. A., Kong, S.-C., Hampson, G. J., and
\KIVA-II: A Computer Program for Chemically Re- Reitz, R. D., \Modeling the E ects of Fuel Injec-
active Flows with Sprays," Los Alamos National Lab- tion Characteristics on Diesel Engine Soot and NOx
oratory Report No. LA-11560-MS, 1989. Emissions," SAE 940523, 1994.
5. Kong, S.-C., and Reitz, R. D., \Multidimensional 18. Hampson, G. J., and Reitz, R. D., \Two-Color Imag-
Modeling of Diesel Ignition and Combustion Using ing of In-Cylinder Soot Concentration and Temper-
a Multistep Kinetics Model," Journal of Engineering ature in a Heavy-Duty DI Diesel Engine with Com-
for Gas Turbines and Power, 115, 781-789, 1993. parison to Multidimensional Modeling for Single and
6. Halstead, M. P., Kirsh, L. J., and Quinn, C. P., \The Split Injections," SAE 980524, 1998.
Autoignition of Hydrocarbon Fuels at High Tempera- 19. Reitz, R. D., \Modeling Atomization Processes in
ture and Pressures { Fitting of a Mathematical Model," High-Pressure Vaporizing Sprays," Atomization and
Combustion and Flame, 30, 45-60, 1977. Spray Technology, 3, 309-337, 1987.
7. Theobald, M. A., and Cheng, W. K., \A Numeri- 20. Su, T. F., Patterson, M., Reitz, R. D., and Farrel,
cal Study of Diesel Ignition," ASME Energy-Source P. V., \Experimental and Numerical Studies of High
Technology Conference and Exhibition, Dallas, Texas, Pressure Multiple Injection Sprays," SAE 960861,
February, 1987. 1996.
8. Kong, S.-C., Han, Z., and Reitz, R. D., \The De- 21. Patterson, M., and Reitz, R. D., \Modeling the Ef-
velopment and Application of a Diesel Ignition and fects of Fuel Spray Characteristics on Diesel Engine
Combustion Model for Multidimensional Engine Sim- Combustion Emissions," SAE 980131, 1998.
ulation," SAE 950278, 1995.
22. Levich, V. G., Physiochemical Hydrodynamics, Prentice-
9. Abraham, J., Bracco, F. V., and Reitz, R. D., \Com- Hall Inc., New Jersey, 1962.
parison of Computed and Measured Premixed Charge
Engine Combustion," Combustion and Flame, 60, 23. Ricart, L. M., Xin, J., Bower, G. R., and Reitz, R.
309-322, 1985. D., \In-cylinder Measurement and Modeling of Liq-
uid Fuel Spray Penetration in a Heavy-Duty Diesel
10. Xin, J., Montgomery, D., and Reitz, R. D., \Multi- Engine," SAE 97159, 1997.
dimensional Modeling of Combustion for a Six-Mode
Emissions Test Cycle on a DI Diesel Engine," ASME 24. Reitz, R. D., Chen, S. K., and Mather, D. K., \Next
Journal of Engineering for Gas Turbines and Power, Generation of High Power Density Diesels for Mid-
119, 683-691, 1997. Range Truck Applications { Part II: The Develop-
11. Kuo, T. W., and Reitz, R. D., \Three{Dimensional ment of Advanced Design Process for Combustion
Computations of Combustion in Premixed{Charge System Optimization," SAE 972689, 1997.
and Direct{Injected Two{Stroke Engines," SAE 920425,
1992.
12. Metghalchi, M. and Keck, J., \Burning Velocities of
Mixtures of Air with Methanol, Isooctane, and Indo-
lene at High Pressures and Temperatures," Combus-
tion and Flame, 48, 191-210, 1982.
13. Bowman, C. T., \Kinetics of Pollutant Formation
and Destruction in Combustion," Prog. Energy Com-
bust. Sci., 1, 33-45, 1975.
14. Heywood, J. B., Internal Combustion Engines, McGraw-
Hill, 1988.

11

Das könnte Ihnen auch gefallen