You are on page 1of 11

Environ. Sci. Technol.

1999, 33, 3957-3967

Just as rich and extremely lean air/fuel ratios have been


Mechanisms of Particulate Matter observed to cause high HC emissions (8-10), so have they
Formation in Spark-Ignition Engines. been found to cause high PM emissions (7, 11-13). Moreover,
cold engines temperatures and cold intake air have been
1. Effect of Engine Operating found to produce more PM than warm ones during transient
driving schedules (5, 7, 14), presumably because of the same
Conditions reason that they emit more HCs (15): they inhibit liquid fuel
vaporization, thus increasing the amount of liquid fuel in
DAVID KAYES* AND SIMONE HOCHGREB the combustion chamber, which cannot burn completely
before being emitted to the exhaust (16, 17). However, the
Massachusetts Institute of Technology, 77 Massachusetts
correlation between HC and PM emissions is not perfect; for
Avenue, Room 31-167, Cambridge, Massachusetts 02139
example, Quader (using a Bosch smoke meter) found that
increased engine load increases smoke emissions but de-
creases HC emissions (7).
A combined experimental and modeling effort was Even if there were a correlation between HC and PM
emissions, it would not necessarily imply similarity of
performed in order to understand how particulate matter
causation. For instance, HC emissions arise from such sources
(PM) is formed in spark-ignition (SI) internal combustion as leaky exhaust valves and crevices inside of which HCs
engines. Parameters that affect global and local air/fuel escape the chemical oxidation reaction. However, PM is
ratios strongly affect PM. Minimum PM number and mass formed by chemical reaction; therefore, these mechanisms
concentrations are emitted at a global air/fuel ratio may not contribute directly to PM formation. Rather, HC
within 10% of stoichiometric, and concentrations increase and PM emissions need to be examined over a broad range
by as many as 3 orders of magnitude when the air/fuel of engine and fuel conditions to establish whether they are
ratio is either increased or decreased 30% from stoichiometric. correlated and, if so, why: specifically, what are the origins
Burning liquid fuel is a significant source of PM, as of PM in SI engines?
evidenced by the fact that open valve fuel injection increases While research in the past has focused largely on single
characteristics of particlessnumber emissions, mass emis-
PM concentrations by up to 3 orders of magnitude
sions, or size distribution characteristicssin a limited set of
relative to closed valve injection. Coolant and oil temperatures, steady-state operating conditions or has examined the
spark timing, and exhaust gas recirculation (EGR) affect integrated result of some of these characteristics over a
PM through their effect on intake port and cylinder transient driving schedule, past research has not systemati-
temperatures as well as through the effect on the availability cally investigated the effect of a variety of steady-state
of liquid fuel in the cylinder. Particle sizes as a function conditions on all three characteristics. Therefore, the objec-
of engine operating conditions are discussed. tives of the present worksin this paper and the companion
paper (3)sare to examine the effect of engine operating
parameters (air/fuel ratio, speed, load, spark timing, EGR,
Introduction and operating temperature), fuel parameters (fuel chemistry,
injection timing, and control strategy), lubricating oil pa-
Recent studies suggest that atmospheric particulate matter
rameters (composition and viscosity), and catalytic converter
(PM) is an important factor associated with mortality and
parameters on PM number and mass emissions as well as
morbidity in urban areas (1). While diesel engines emit
characteristic particle sizes.
significantly higher concentrations of particles than gasoline
engines, the total number of vehicle miles traveled (VMT) by
gasoline-powered vehicles in urban centers greatly exceeds
Experimental Apparatus
that of diesels, such that gasoline vehicles may contribute A scanning mobility particle sizer (SMPS) manufactured by
up to 68% of the vehicle engine-generated particulate mass TSI Inc. is used to measure PM number- and volume-
and 90% of the particulate number (2). Therefore, it is weighted concentrations as well as size distribution char-
important to understand how PM is formed in SI engines acteristics in the diluted exhaust from a gasoline engine;
and how it might be mitigated. Moreover, it is important to SMPS operation is described in ref 2. Due to the strong effect
understand not only the number- or mass-weighted emis- of atmospheric dilution on exhaust temperature and hence
sions but also the size distribution characteristics insofar as particulate mass (2, 18), the U.S. EPA mandates a maximum
they influence respirability. To this end, a combined ex- sample temperature of 52 °C (125 °F), effectively requiring
perimental and modeling effort was undertaken: the current dilution by a factor of at least about 10. In the present
paper details the experimental results as a function of steady- experimental setup, engine exhaust flows through a muffler
state engine operating conditions, companion paper 2 details and is introduced into the centerline of the dilution tunnel,
experimental results as a function of fuel and oil charac- where it is mixed with filtered air. Diluent air for the dilution
teristics as well as the effect of the catalytic converter (3), tunnel is drawn through a carbon filter to remove organics
and companion paper 3 describes a predictive model (4). and a high efficiency particulate air (HEPA) filter, with a
The results of previous investigations of PM formation measured particle trapping efficiency of 99.8% on a mass
will be examined so as to elucidate some of the possible basis. The dilution tunnel has an inner diameter of 15 cm
mechanisms by which SI engines form PM. Correlations and consists entirely of stainless steel and Teflon since other
between HC and PM emissions observed in previous research materials may retain and release organic vapors (19).
(5-7) as well as the necessity of HC for PM formation suggest Exhaust is supplied by either a Ford Zetec or a Saturn
that the mechanisms of HC and PM emission may be similar. DOHC 4-cylinder spark-ignition engine, each of which is a
modern production engine with four valves per cylinder,
* Corresponding author phone (617)253-3358; fax (617)253-9453; pent-roof geometries, and port fuel injection (PFI). The Ford
e-mail: kayes@mit.edu. has 2.0 L displacement and a compression ratio of 9.6; while
10.1021/es9810991 CCC: $18.00  1999 American Chemical Society VOL. 33, NO. 22, 1999 / ENVIRONMENTAL SCIENCE & TECHNOLOGY 9 3957
Published on Web 10/12/1999
the Saturn has 1.9 L and a compression ratio of 9.5. Both approximately 20%; for mass concentration, it is 25%.
engines are coupled to dynamometers. Exhaust is taken Uncertainty in measurements of mean particle sizes made
directly from the engines (no catalytic converter) unless via the engine/tunnel/SMPS system is approximately 3%;
specified. The fuel injection of the Ford Zetec engine can be for mode particle sizes, it is 10%. These numbers represent
controlled by either the standard production engine control- maximum deviations between any measurement and the
ler or by a manual computerized proportional/integral (PI) average measured value in repeated experiments at the
controller, the latter of which maintains the air/fuel ratio baseline operating conditions. At engine operating conditions
within a more narrow range than the former and is used for where the engine frequently misfired (for example, when
all tests unless specified. Spark timing for the Ford engine operating at a lean air/fuel ratio of φ ) 0.7), differences
is controlled by a manual controller and is set for MBTs between any two measurements could be significantly larger
maximum brake torque or minimum spark advance for best than these quoted uncertainties; however, in these cases,
torquesunless specified. A manufacturer-supplied computer, the deviations are presumed to be actual variations in the
which overrides the set points in the standard production engine exhaust concentrations as opposed to artifacts of the
engine controller, controls the fuel injection and spark timing measurement process. Moreover, at engine operating condi-
for the Saturn engine. Again, spark timing is MBT unless tions leading to large diameter particles, the SMPS may
specified. The lubricating oils used in both engines are SAE accurately measure the number-weighted mean and mode
viscosity 30; neither engine has significant oil consumption. sizes, but the volume-weighted mean and mode (i.e.,
The SMPS outputs the particle size distribution as well as weighted by diameter to the third power and equivalent to
integrated results: concentrations, mean particle sizes, and the mass-weighted mean and mode, respectively) may fall
mode particle sizes, all of which are given on number, surface above the SMPS measurement range. The SMPS measure-
area, and volume-weighted bases (20). (Because of a pre- ment range (9-378 nm) was optimized for number-weighted
sumed error in the SMPS software, volume-weighted mean particle concentrations; according to measurements Rickeard
size is calculated manually; the software calculates all other et al. made of particle concentrations using two types of
parameters correctly.) The mean particle size is the average particle analyzers, less than 1% of the particles on a number-
size of all particles, as detailed in ref 21, and the mode is the weighted basis should fall above the SMPS measurement
most common size, i.e., the size at which the highest range (31).
concentration of particles is measured. Volume-weighted SMPS particulate measurements, taken in the dilution
sizes are equivalent to mass-weighted sizes for uniform tunnel and shown in the figures that follow, are corrected for
density particles. The volume-weighted concentration (or dilution ratio using the following equation so that they reflect
volume fraction) is directly related to the mass concentration the particle number or mass per unit volume of exhaust,
by the particulate density. Most researchers assume a density [PM]e, cooled to standard temperature and pressure:
of 1000 (22-24) to 2000 kg/m3 (25-28) to the 2260 kg/m3
density of solid carbon (29); an assumed density of 2000 [PM]e ) Λ[PM]d - (Λ - 1)[PM]a (1)
kg/m3sthe median and the average of the above values (using
the available 1 digit accuracy)sis used in the calculation of
all mass concentrations in the present work. A density of where [PM]d is the particulate number or mass concentration
1000 kg/m3, which is closer to the density of the organic measured in the diluted exhaust, Λ is the dilution ratio, and
carbon condensate/absorbate that may make up 38-75% of [PM]a is the particulate concentration measured in the diluent
the particle mass (30), could just as well have been assumed, air.
in which case all PM mass concentrations would be one-half
the presently quoted values. Quoted particles sizes are the Experimental Conditions
electrical mobility diameter, which is the size of a sphere An experimental matrix was developed to test the effect of
that would have the same amount of drag as the particle steady-state engine conditions and fuel delivery parameters
when being drawn through air by an electric charge. on particulate emissions. (The effect of transient conditions
In conjunction with PM measurements, HCs concentra- on PM will be explored in an upcoming work.) The matrix
tions are measured with a Rosemount analytical model 402 centers about the “baseline operating conditions” listed in
hydrocarbon analyzer, which is a flame ionization detector Table 1. Some tests in the matrix were performed on the
(FID) with a heated inlet line. Water is not purged from Ford Zetec engine, and some were performed on the Saturn
exhaust before entering the FID, so the measurements engine, with two sets of tests performed on both engines in
represent “wet” HC concentrations. O2 concentrations are order to compare their respective emissions. The comparison
measured with a Beckman model OM-11 EA meter and are will be discussed in the Results section.
shown as wet values. NOx concentrations are measured with
a Thermo Environmental Instruments model 10 chemilu- Experimental Methodology
minescence analyzer, which requires desiccation, so mea- The following is a description of testing procedures. Each
surements represent “dry” NOx concentrations. Dilution ratio test, in which one parameter is swept about the baseline
(Λ)sdefined as the ratio of mass flow rates of diluted exhaust value, is performed on a single day, thereby ensuring that
to undiluted exhaust, with Λ ) 1 corresponding to undiluted day-to-day engine variations are not an issue and that diluent
exhaustsis calculated by comparing dry CO2 concentrations air characteristics remain consistent. At the beginning of a
measured with a Rosemount Analytical model 880A NDIR test day, the SMPS is warmed and used to measure the
analyzer in the diluted exhaust, undiluted exhaust, and atmospheric particulate concentration. Then the SMPS is
diluent air as per ref 2. Dilution ratios within 10% of 15:1 are used to measure the particle concentration in the diluent
used because they minimized the sensitivity of measured air, i.e., the air (without any engine exhaust) drawn into the
PM concentrations to deviations in dilution ratio (2). The dilution tunnel through the HEPA and charcoal filters.
flow rate of diluted exhaust in the present experiments is Experiments begin only after the particulate mass concen-
3600 L/min (Re ) 33 000), the average of the flow rates tration in the diluent air drops below 0.4% of the atmospheric
examined in ref 2. concentration, which corresponds to less 0.1% of the
Measurement uncertainties were described and quantified measured particle concentration at the baseline engine
in a previous investigation (2), but the results will be operating conditions and approximately 0.4% of the mini-
summarized here. Uncertainty in measurements of number mum measured PM concentration (corrected for dilution
concentration made via the engine/tunnel/SMPS system is ratio) using indolene fuel.

3958 9 ENVIRONMENTAL SCIENCE & TECHNOLOGY / VOL. 33, NO. 22, 1999
TABLE 1. Baseline Engine Operating Conditionsa
parameter value
engine speed 2000 (( 10) rpm
IMAP 0.4 (( 0.04) bar, absolute
φ 1.00 (( 0.02)
fuel injection timing CVI
fuel injection control PI control using UEGO sensor for
strategy Ford engine; manual override of
set points for standard production
controller for Saturn engine,
monitored by heated oxygen
sensor
spark timing MBT (( 3 °CA)
coolant and oil temps 87 (( 2) °C
EGR none
fuel type indolene
lubricating oil type conventional 30 weight mineral oil
catalyst none (measurements are
engine-out emissions)
exhaust dilution ratio 15 (( 2):1
a IMAP, intake manifold air pressure. φ, fuel/air equivalence ratio.

CVI, closed valve injection, as opposed to open valve injection (OVI). FIGURE 1. PM as a function of fuel/air equivalence ratio. Baseline
PI, proportional/integral control. UEGO sensor, universal exhaust gas engine operating conditions except equivalence ratio and dilution
O2 sensor. MBT, minimum spark advance for best torque.°CA, crank ratio, as explained in text. Data taken on both Ford and Saturn
angle degrees. EGR, exhaust gas recirculation. Indolene, a research engines. Data corrected for dilution ratio to represent PM
gasoline with low sulfur content. concentration in the cooled engine exhaust.

Engine conditioning is handled as follows: before any


day of experiments, the engine is warmed at the baseline
engine speed and load until the temperatures of the coolant
and oil are within 2° of the 87 °C set point. Once the engine
is fully warmed, the dilution ratio is adjusted until it is within
10% of 15:1. Tests are then performed using a randomized
protocol. Between any two engine operating conditions in
an experiment, the engine is conditioned for a minimum of
10 min at the new operating condition. The 10-min condi-
tioning period is deemed more than sufficient based upon
measurements that 5 min after a step change in the operating
conditions the PM concentration are within 1% of the steady-
state value. By comparison, Graskow et al., who used a similar
SMPS apparatus on a SI engine connected to a dilution tunnel,
found that conditioning takes only 5-6 min (32).

Results and Discussion


The following is a description of results from individual
experiments in which individual operating conditions were
FIGURE 2. HC emissions (top right scale) and oxygen emissions
varied from the baseline conditions. This section also
(left scale) as a function of fuel/air equivalence ratio. Emissions
describes the mechanisms responsible for the experimentally are corrected in order to represent mole fraction in the “wet” exhaust
observed PM concentrations. gas. Baseline engine operating conditions except equivalence ratio.
PM as a Function of Equivalence Ratio. In experiments Also, flame temperature (bottom right scale) as a function of fuel/
where the fuel/air equivalence ratio (φ) is varied by incre- air equivalence ratio, calculated using GM Engine-Simulation
ments of 0.1 φ units from 0.7 to 1.3 while MBT spark timing Program (32). Baseline engine operating conditions except equiva-
is maintained; PM number and mass concentrations are a lence ratio and using isooctane fuel.
minimum at approximately stoichiometric (φ ) 1.0) and
increase as φ is made either lean or rich (Figure 1). As opposed dilution ratio, these data will not be discussed and are only
to HC emissions (Figure 2), which vary by less than an order plotted in Figure 1 to emphasize the consistency of trends
of magnitude over the range φ ) 0.7-1.3, PM number and observed for PM emissions as a function of φ. Specifically,
mass concentrations vary by 1 and 2 orders of magnitude, the behaviorsminimum measured emissions at, or within
respectively. (The difference between the increases in number 10% of, stoichiometric and increased emissions at lean or
and mass concentrations is due to the change in particle rich φsis consistent from engine to engine even at different
sizes with respect to φ, as will be discussed.) Experiments by dilution ratios, as is the magnitude of the corrected PM
other researchers have shown that PM emissions from SI concentration.) In all cases, data in Figure 1 are corrected for
engines reach a minimum at approximately 10-20% lean of dilution ratio to represent PM concentration per unit of
the stoichiometric air/fuel ratio and increase at rich or exhaust gas cooled to STP, according to eq 1.
extremely lean φ (11, 13). A combination of mechanisms contribute to the behavior
Experimental measurement of the variation in PM emis- of PM with respect to equivalence ratio: nucleation, oxida-
sion with respect to φ was performed on the Saturn engine tion, and growth. Particle nucleation is affected by temper-
using the baseline dilution ratio of 15. (The same experiment ature and availability of particle precursors, such as liquid
was also performed on the Ford engine, although using a fuel droplets or pools, fuel-rich regions, and products of
dilution ratio between 8 and 9. Because of the nonstandard incomplete fuel oxidation. Similarly, oxidation is affected by

VOL. 33, NO. 22, 1999 / ENVIRONMENTAL SCIENCE & TECHNOLOGY 9 3959
temperature and oxygen availability. Growth is affected by
the amount of HCs available for adsorption/absorption, the
amount of surface area available on which HCs can adhere,
and the number of particles with which any one particle can
coagulate. The following discussion details the effects of air/
fuel ratio on these various mechanisms and their controlling
factors.
Formation of particles in gas-phase reactions depends
on soot precursor availability and temperature. As the fuel/
air ratio is increased, greater HC concentrations in the flame
lead to greater concentrations of soot precursors. Moreover,
as the fuel/air ratio is increased from φ ) 0.7 to 1.1, cylinder
temperatures increase: for example, peak flame temperature
[estimated using the GM Engine-Simulation Program (33)]
increases from about 2350 to 2600 K (Figure 2). [Although
the GM Engine-Simulation Program stimulations were
performed based on isooctane fuel, peak flame temperature
varies by no more than 3% or 100 K at any given φ between
indolene and isooctane (34). Thus, engine simulation results
FIGURE 3. Number-weighted mean and mode particle sizes as a
using isooctane fuel should be generalizable to indolene.] function of equivalence ratio, measured on both Ford and Saturn
But as φ is increased from 1.1 to 1.3, peak cylinder temper- engines. Baseline conditions except equivalence ratio and dilution
atures decrease by 70 K (Figure 2). The combined result of ratio, as explained in text.
the soot precursor availability and temperature (as discussed
in ref 4) is that peak particle formation via gas-phase reactions to PM mass when the exhaust is cooled and diluted. But
occurs at φ ) 1.2 and monotonically decreases on either side regardless of φ, the exhaust is diluted to approximately 30 °C
of that value (4). While these gas-phase reactions explain the in the dilution tunnel, so temperature should not affect the
general increase in PM as φ increases up to 1.2, the model relative amount of adsorption/absorption as a function of φ.
presented in ref 4 suggests that liquid fuel combustion is the Rather, HC availability in the dilution tunnel (which, with
cause of increasing PM as φ is decreased below 1.0. fixed dilution ratio, is a minimum at φ ) 0.8 and is
Formation of particles in heterogeneous phase reactionss proportional to mole fractions shown in Figure 2) and the
i.e., generation of particles from burning droplets or pools amount of surface area on which adsorption/absorption can
of liquid fuelsdepends on the availability of the liquid fuel occur (which is proportional to the amount of particles
and oxygen necessary for combustion as well as on the nucleated minus the amount oxidized) govern the amount
temperature necessary to ignite the mixture. Witze and Green of mass accreted. Growth via coagulation of particles scales
used optical techniques to show that pool fires can occur but with the number of particles squared: thus, like mass
do so only after flame passage when temperatures are high accretion via adsorption/absorption, coagulation depends
but oxygen concentrations are significantly lower than in strongly on the number of particles nucleated and results in
the unburned charge (15). According to the model presented ever increasing sizes of aggregate particles. Therefore,
in ref 4, the probability that a liquid fuel droplet or pool will coagulation should cause the most growth at the leanest and
ignite in the post-flame environment is so strongly dependent richest air/fuel ratios. In summary, cases with the highest
upon post-flame oxygen concentration that ignition is several concentrations of particles and HC before the growth
orders of magnitude more likely at φ ) 0.7 where post-flame processes should yield the highest PM mass concentrations
oxygen mole fraction is approximately 7% than at φ ) 1.0 after the growth processes; the end product from the
where oxygen mole fraction is 0.7%, despite approximately formation, adsorption, absorption, and coagulation processes
30% less liquid fuel and the 230 K lower flame temperature is the final PM concentration shown in Figure 1.
(Figure 2) (33). Since formation of PM from liquid fuel requires Comparison of mean and mode particle sizes at the
that the liquid ignite (rather than escape the combustion various air/fuel ratios corroborates the above explanation of
chamber completely unburned) and since the concentration PM emissions (for example, number-weighted sizes are
of post-flame oxygen required for ignition decreases mono- shown in Figure 3). On both number- and mass-weighted
tonically with increased φ, PM concentrations formed by bases, the minimum particle sizes occur within 10% of φ )
heterogeneous-phase reactions decrease monotonically with 1.0, while particle size increases by as much as a factor of 3
increased φ. at both rich and lean equivalence ratios, presumably because
Particle oxidation follows an Arrhenius dependence on of coagulation and adsorption/absorption as discussed
oxygen concentration and temperature. Although in-cylinder above.
temperatures reach a maximum at φ ) 1.1, oxygen availability The occurrence of minimum PM emissions stoichiometric
is a maximum at the minimum φ and decreases monotoni- air/fuel ratio suggests that PM emissions can be minimized
cally with increased φ (Figure 2). The result of this competition by tightening the control of air/fuel ratio around φ ) 1. A test
between temperature and oxygen availability, based on was performed to confirm this: two engine control strategies
calculations following the model presented in ref 4, is that were employed, one allowing larger φ fluctuations than the
the fraction of PM oxidized is a maximum at the leanest φ other but with all other conditions (including spark timing)
and decreases monotonically as φ is increased (4). fixed at the steady-state baseline (Table 1). In one case, the
Growth of particles via adsorption/absorption depends standard production Ford engine controller (the EEC IV) was
primarily on the availability of HCs and PM surface area. used with the standard oxygen sensor and, in the other, a
Adsorption/absorption, as modeled for vapors on plane computerized proportional/integral (PI) controller was used
surfaces by Langmuir and adapted for HCs on diesel particles with a universal exhaust gas oxygen (UEGO) sensor. The PI
by Plee and MacDonald (35, 36), is enhanced by low controller with the UEGO sensor allowed φ to fluctuate within
temperatures and high concentrations of both unburned HCs about (1% of stoichiometric, while the standard production
and particle surface area. While adsorption/absorption begins controller/sensor system allowed (2-3% fluctuation, as
as soon as particles and vapors are brought in contact inside measured with a separate UEGO sensor. PM number
the combustion chamber, it contributes most substantially concentration is a factor of 7 lower (as measured in repeated

3960 9 ENVIRONMENTAL SCIENCE & TECHNOLOGY / VOL. 33, NO. 22, 1999
Both HC and PM concentrations have a trough at injection
timings about the middle of the intake stroke.
The following is a reasonable explanation for the measured
trends above. Fuel injected onto the closed valves sufficiently
advanced of IVO resides in the intake port long enough to
allow vaporization. Thus, for these injection timings, a
relatively small amount of liquid fuel enters the combustion
chamber, which in turn causes low HC emissions. The fact
that HC emissions are insensitive to fuel injection timing
during the CVI period at steady-state operation has been
observed by Arcoumanis et al., Alkidas, and Quader (7, 38,
39). Analogously, fuel injection timings that allow significant
evaporation time likely result in low PM emissions, presum-
ably because of the absence of liquid fuel needed for sooty
droplet or pool fires.
By contrast, fuel injected through the open valves has a
short-circuit by which it can arrive in the combustion
chamber without undergoing vaporization. Meyer and Hey-
wood estimate that as much as seven times more fuel survives
in liquid phase until spark ignition if it entered the cylinder
via OVI as compared with CVI (40). A number of studies have
shown that OVI leads to higher steady-state HC emissions
than CVI, precisely because of the added liquid fuel present
FIGURE 4. HC and PM emissions as a function of the time in Crank (7, 38, 39, 41). While OVI leads to high HC emissions by
Angle Degrees (CAD) After Intake Top Dead Center (AITDC) at which allowing some of the fuel to escape full oxidation in the flame,
injected fuel hits the intake valves. Baseline engine operating it generates PM in two ways. Primarily, burning droplets or
conditions except fuel injection timing. pools of liquid fuel nucleate particles, as observed qualita-
tively by Witze and Green (15) and as discussed above in
tests) with the PI controller than with the production reference to the fuel-lean PM production. Of lesser impor-
controller; mass concentration is a factor 10 lower; HC mole tance is the fact that liquid fuel does not burn fully and thus
fraction is 250 ppmC1 (about 10% of the total HC mole leads to HC emissions, which may then adsorb/absorb onto
fraction) lower. The degree to which the PI control reduced existing particles, thereby increasing their mass. The reason
PM emissions is greater than that predicted by the model in that this mechanism is believed to be of lesser importance
ref 4 but suggests the strong effect of nonstoichiometric is that HC concentration increases by only 67% while PM
conditions on PM emissions. Number- and mass-weighted concentrations increase by orders of magnitude; if such a
mean particle sizes are respectively 14% and 12% smaller for small change in HC concentration could create such a large
PI control, presumably due to less HC adsorption, absorption, change in PM concentration, then the correlation between
and coagulation. However, mode sizes for the different the PM and HC emission rates would be stronger than that
control strategies are all within 10% of each other, which is observed both in past work (5-7) and in companion paper
within measurement uncertainty. Hence, finer control of φ 2 (3). [The R 2 correlation coefficient between PM and HC
results in smaller, less concentrated particles as well as less concentrations for experiments described in this paper and
HCs. a companion paper (3) is approximately 0.2.] As for other
PM as a Function of Fuel Injection Timing. The effect PM formation mechanisms, fuel injection timing has a
of fuel injection timing (particularly, closed versus open valve negligible effect on cylinder temperatures (as demonstrated
injection) was investigated by varying the start of fuel injection by the NOx emissions, which varied by no more than the
at 25 °CA intervals throughout the OVI period and 100 °CA measurement uncertainty) and thus should not significantly
intervals throughout the CVI period. The experiments were affect the amount of particulate generated through gas-phase
performed on the Ford engine which has intake valve open nucleation reactions, which are Arrhenius in nature (42).
(IVO) at 10 °CA BTC and intake valve closing (IVC) at 230 °CA Moreover, the small effect of injection timing on in-cylinder
ATC. Fuel injection timing given in Figure 4 is the time when temperature and the lack of significant effect on in-cylinder
fuel hits the intake valves: calculated as the crank angle at oxygen concentration should lead to negligible variation in
which the fuel leaves the fuel injector plus the estimated, the PM oxidation rate with injection timing.
average time necessary for fuel droplets to travel from the The trough in HC and PM emissions at 50-70 °CA appears
injector to the intake valve. On the basis of a force balance to be the result of injecting fuel into the fast-moving intake
analysis following Ladommatos and Rose (37), the fuel travel airstream. High relative velocities result in a high heat transfer
time is assumed approximately equal to the transit time for rate from intake air to the droplets and hence a relatively
intake air from the injectors to the intake valves. Thus, it is high evaporation rate (37, 41). Moreover, injection into fast
essentially the distance of travel divided by the average air moving airstreams may facilitate droplet breakup (37, 43),
speed during the intake process, which is calculated using which increases the surface/volume ratio and thus further
the GM Engine-Simulation Program (33). enhances the evaporation rate.
HC and PM emissions (Figure 4) are insensitive to fuel Particle sizes vary by no more than the measured scatter
injection timing during most of the CVI period. However, during the closed valve injection period (Figure 5); number-
the injection timing immediately before IVOsshown as 23 weighted mean diameters are approximately 65 nm, and
°CA before IVO in Figure 4sresults in 10% higher HC mass-weighted mean diameters are 100 nm. Although
emissions that than the CVI average, while PM number- and minimum particle sizes are observed during OVI (38 and 69
mass-weighted concentrations for this injection timing are nm on number- and mass-weighted bases, respectively), not
both five times the CVI level. During the OVI period, HC all OVI timings result in smaller particles than CVI timings.
concentrations increase by as much as 67% over the CVI The observed particle sizes corroborate the aforementioned
level, while PM concentrations on either number- or mass- assertion that the difference between emission rates at the
weighted bases increase by nearly 3 orders of magnitude. different injection timings is primarily due to the number of

VOL. 33, NO. 22, 1999 / ENVIRONMENTAL SCIENCE & TECHNOLOGY 9 3961
FIGURE 6. HC, NOx, and mass-weighted PM emissions as a function
of coolant and oil temperature. Baseline operating conditions except
for coolant and oil temperatures.
FIGURE 5. Number- and mass-weighted mean particle sizes as a
function of the time in Crank Angle Degrees (CAD) After Intake Top
Dead Center (AITDC) at which injected fuel hits the intake valves. Similarly, the added presence of liquid fuel in the cylinder
Baseline engine operating conditions except fuel injection timing. due to cooler component temperatures should produce
higher PM emissions, not only because the liquid fuel in the
particles nucleated by burning liquid fuel. That is, if a cylinder can burn and produce soot but also because the
combination of nucleation and growth mechanisms led to increased HC emissions can adsorb/absorb on the particles,
the measured particle concentrations, then the particles causing growth.
generated with OVI (those having the most HC available for Experimental measurements show that the 53 °C decrease
growth) would be the largest. By contrast, the fact that small in coolant and oil temperature leads to a 20 °C decrease in
particles and high HC concentrations are observed concur- intake port and manifold temperature (as measured with a
rently suggest that growth cannot dominate the observed thermocouple inserted in the intake manifold approximately
differences in particle concentrations as a function of in- 11 cm upstream of the port). Figure 6 shows that a 53 °C
jection timing, the alternative being that nucleation must. decrease in coolant and oil temperature results in a measured
The implications of this test are that direct injection spark- 19% increase in HCs, presumably due to the added liquid
ignition enginessfor which the amount of liquid fuel inside fuel presence. NOx emission levelssa surrogate for in-cylinder
the cylinder greatly exceeds that in standard PFI enginess temperaturessdecrease by only 2% from the warm condition
should have much higher PM emissions than PFI engines at to the cold, which is a statistically insignificant decrease
similar conditions. compared to the approximate 10% uncertainty. The de-
PM as a Function of Engine Oil/Coolant Temperature. creased coolant and oil temperatures result in a 60% increase
In light of research by Williams et al. and Maricq et al. showing in mass-weighted PM emissions. (However, number-
that PM emissions during bag 1 of the Federal Test Procedure weighted PM emissions differ by less than the measured
(cold test phase) are higher than those during bag 3 (warm scatter.) Mass-weighted particle sizes increase by 26% from
test phase) (5, 44), a test was performed to measure the effect the warm conditions to the cold ones, in itself a large enough
of steady-state coolant and oil temperature on steady-state increase to account for the increased PM mass concentration.
PM emissions. An additional heat exchanger was added to Larger particles during cold operation than warm means
the cooling circuits for the coolant and lubricating oil, either that enhanced nucleation increases the rate of
allowing the coolant and oil to be maintained at either the coagulation or that absorption/adsoprtion are enhanced, or
baseline conditions or at a temperature of 34 ( 6 °C during both.
steady-state operation. The heat exchanger circuits consisted PM as a Function of Spark Timing. A test was performed
of copper pipes submerged (a) in ice baths to maintain the where the spark timing was varied in 5 °CA increments from
temperature of 34 °C or (b) in ambient air to maintain the 5 to 55 °CA BTC (20 °CA before MBT timing to 30 °CA after),
baseline temperature of 87 °C. Besides the temperature with all other conditions held fixed. The effect of advanced
reservoir in which the heat exchanger circuits were posi- spark timing is an increase in number- and mass-weighted
tioned, no changes were made to the engine between tests PM emissions from spark timings of 5-40 °CA BTC (the latter
at the two temperatures to minimize measurement artifacts. of which is about 5 °CA advanced of MBT timing), after which
Decreased coolant and oil temperature leads to decreased further advance results in decreased emissions (Figure 7).
intake port and manifold temperature by virtue of the greater Similarly, advancing spark timing up to 40 °CA BTC results
heat transfer to the coolant and oil. However, there is not a in increased HC emissions; but as spark timing is advanced
one-to-one correlation between decreased coolant/oil tem- beyond 40 °CA BTC, HC emissions remain constant to within
perature and decreased port/manifold temperature, in part the measurement uncertainty (Figure 8).
because of compression and combustion heating of the intake The effect of spark timing on PM is not as simple as that
valves. Nevertheless, decreased coolant and oil temperatures on HCs: the behavior of HCs with respect to spark timing
result in lower rates of fuel evaporation in the intake port, is consistent with an increased rate of oxidation at the higher
thus higher concentrations of in-cylinder liquid fuel and post-flame temperatures associated with more retarded spark
greater concentrations of in-cylinder liquid fuel have, in the timings (for example, exhaust temperatures shown in Figure
past, been observed to lead to higher HC emissions (16, 45). 8). On the other hand, the variation in PM concentrations

3962 9 ENVIRONMENTAL SCIENCE & TECHNOLOGY / VOL. 33, NO. 22, 1999
FIGURE 7. PM emissions as a function of spark timing. Baseline
operating conditions except spark timing. Tests repeated on 2 FIGURE 9. Number- and mass-weighted mean and mode particle
days: data shown as filled and open symbols. Outlier shown in sizes as a function of spark timing. Baseline engine operating
parentheses. conditions except spark timing.

of HCs available for surface growth remains fixed; hence, the


fact that PM concentrations decrease suggests that the
increase in the particle oxidation rate is greater than that for
nucleation. Mean and mode particle sizes show no monotonic
trend as spark timing is changed (Figure 9), reflecting the
competing effects of the different particle dynamics mech-
anisms.
PM as a Function of EGR. Experiments were performed
in which the level of exhaust gas recirculated to the intake
was varied with all other conditions (including IMAP and
engine speed) fixed. EGR was varied from the baseline level
(no EGR) to the level such that the brake torque produced
by the engine was approximately zero. (Even when the engine
produces no brake torque, the dynamometer maintains the
baseline engine speed of 2000 rpm.) EGR levels are referenced
as the percentage of the EGR rate necessary to stall the engine,
i.e., produce zero brake torque. Values tested range from 0
to 100% in 25% increments.
Experimental results show that, as the level of EGR
increases, the PM number and mass emissions decrease
FIGURE 8. HC emissions, peak flame temperature, and exhaust port exponentially (Figure 10); number- and mass-weighted PM
temperature as a function of spark timing. Baseline operating emissions are 2 orders of magnitude lower at the maximum
conditions except spark timing. Symbols represent data; curves EGR case than at the baseline conditions. The number- and
accompanying HC mole fractions and exhaust temperatures mass-weighted emissions have R 2 correlation coefficients of
represent fit to data. Peak flame temperature calculated using GM 0.78 and 0.71 for their respective exponential fits, showing
Engine-Simulation Program (32). that the data has a fair amount of scatter, despite the fact
that as many as five measurements were made at each EGR
with respect to spark timing reflects the competing effects level. The scatter in the data may not be an artifact of the
of both flame and post-flame temperatures on nucleation PM measurement process but may reflect real variability in
and oxidation as well as the effect of HC concentrations on PM emissions; this conclusion is consistent with the fact
adsorption/absorption. Advancing spark timing results in that, at the high EGR levels used in this experiment, misfires
increased peak temperature (Figure 8) and thus raises the and partial burn cycles were frequent although not necessarily
rate of Arrhenius-dependent PM nucleation rates (4, 33) while periodic. (Scatter is also observed in particle sizes, which are
at the same time reducing PM concentrations by enhancing discussed below.)
post-flame oxidation, which is also Arrhenius-dependent (4, The decrease in PM emissions with increased EGR is
46, 47). Furthermore, as spark timing is advanced, exhaust consistent with the effect of EGR on characteristic temper-
temperatures decrease (Figure 8), thus slowing post-flame atures and species concentrations in the engine. Increased
HC oxidation and leaving more HCs available for adsorption/ EGR results in increased intake port/manifold temperature
absorption. However, as spark timing is advanced beyond but decreased flame and post-flame temperatures, as indi-
40 °CA BTC, exhaust temperatures decrease no more than cated by both direct and indirect measurements. In particular,
the 1-2% measurement uncertainty; consequently, the measured temperatures in the intake manifold increase
Arrhenius-dependent rate of HC oxidation does not decrease monotonically from 50 to 88 °C when EGR is increased from
significantly, resulting in the plateau in HC emissions. So, as 0 to 100% of that needed to stall (Figure 11) since the hot
spark timing is advanced beyond 40 °CA BTC, nucleation exhaust gas is recirculated directly into the intake manifold;
rates increase, in-cylinder oxidation rates increase, exhaust thus, increased EGR enhances liquid fuel vaporization rates.
oxidation rates remain approximately fixed, and the amount On the other hand, peak temperatures decrease, as evidenced

VOL. 33, NO. 22, 1999 / ENVIRONMENTAL SCIENCE & TECHNOLOGY 9 3963
FIGURE 12. Particle sizes as a function of EGR: mean sizes on both
number- and mass-weighted bases plus mode sizes on number-
FIGURE 10. Number- and mass-weighted PM emissions as a function weighted basis. Baseline engine operating conditions except EGR
of EGR rate. Baseline conditions except EGR rate. Lines represent rate. Mode sizes on mass-weighted basis are too scattered and
fits to data, described in text. thus are not plotted.

in more rapid evaporation of liquid fuel. Plus, the recirculated


exhaust gas displaces fresh charge in order to maintain fixed
intake manifold pressure. [Estimates of the burned gas
fraction in the cylinder prior to combustion made using the
GM Engine-Simulation Codesmade by varying burned gas
fraction until spark timing and brake torque matched
measured valuessare that burned gas fraction increases from
approximately 10% in the zero EGR case to approximately
70% in the maximum EGR case (33).] The combination of
increased vaporization rate and decreased amount of charge
supplied to the engine is that the in-cylinder gas-phase fuel
concentration decreases by an estimated 2/3, while the in-
cylinder liquid fuel concentration decreases by an estimated
80-90% (4). PM nucleation mechanisms depend strongly
on the gas- and liquid-phase fuel concentrations as well as
temperature (4, 42), so that decreased concentration of both
fuel phases combined with decreased in-cylinder temper-
atures (discussed above) results in a sharp decrease in the
amount of PM nucleated as EGR increases.
Nonetheless, as EGR increases, oxidation rates should
decrease and growth rates should increaseseffects that
FIGURE 11. Intake manifold temperature, exhaust HC concentration, compete with the decreased PM nucleation. Specifically, the
and exhaust NOx concentration as a function of EGR. Baseline engine decreased peak/exhaust temperatures and oxygen concen-
operating conditions except EGR rate. tration result in decreased PM oxidation, since the oxidation
rate depends on oxygen and exponentially on temperature
by NOx emissions, which decrease from 2400 to 25 ppm as (4, 46, 47). Furthermore, as EGR increases, exhaust HC
EGR is increased (Figure 11). While EGR does displace concentration increases, so the amount of vapor available
nitrogen and oxygen, the main reason for the decreased NOx for adsorption/absorption increases. However, as EGR
formation is that EGR acts as a diluent and therefore holds increases, the increase in vapor available for growth is
the peak cylinder temperatures down (48). Similarly, as EGR potentially offset by the decrease in available surface area on
increases, post-flame temperatures decrease. For example, particle nuclei. The combined result, as demonstrated by
as EGR is increased from 0 to 100% of the level to stall, experimental results (Figure 10), is that the decreased
measured exhaust port temperature decreases from ap- nucleation overwhelms the decreased oxidation and poten-
proximately 560 to approximately 490 °C. Moreover, HC tially increased growth.
emissions increase from 2400 to 13 200 ppmC1 (Figure 11); Both number- and mass-weighted mean particle sizes
the reason for this is twofold: increased frequency of misfires decrease as EGR is increased from 0% to 50 or 75% of the
and partial burn cycles wherein HCs escape oxidation in the amount needed to stall, after which point the sizes increase
flame plus decreased post-flame and exhaust temperatures (Figure 12). However, measured particle sizes have significant
during nonmisfire cycles, thus leading to decreased post- scatter as discussed above. Number-weighted mean particle
flame oxidation. sizes are approximately 40 nm at the zero EGR case and
Given these measurements, the theoretical expectation decrease to 20 nm at 75% EGR but increase to 25 nm at 100%
why increased EGR results in an exponential decrease in PM EGR. Number-weighted mode sizes decrease monotonically
is the following. As more hot exhaust gas is introduced into from 36 to 11 nm as EGR is increased. Mass-weighted mean
the intake manifold, the intake temperature rises and results particles size are 50 nm at zero EGR, approximately 35 nm

3964 9 ENVIRONMENTAL SCIENCE & TECHNOLOGY / VOL. 33, NO. 22, 1999
FIGURE 13. Number-weighted PM concentrations as a function of FIGURE 14. Mass-weighted PM concentrations as a function of
intake pressure (engine load). Three engine speeds, all other intake pressure (engine load). Three engine speeds, all other
conditions same as baseline. conditions same as baseline.

for EGR between 25 and 75%, and 48 nm at 100% EGR. Mass- Measured exhaust temperatures increase monotonically by
weighted mode sizes show a high degree of scatter. The 80-120 °C as load is increased at all three speeds, and
simultaneous decrease in number-weighted mode sizes and (perhaps as a result of increasing post-flame oxidation)
increase in mean sizes is due to a general though non- exhaust HC emissions decrease monotonically by 390-450
monotonic spreading of the size distributionsparticularly ppmC1 (a decrease of 21-27% of the average HC mole
toward large particles, although also toward small onessas fractions) at the three speeds.
EGR is increased: the standard deviation of the size Increased load may cause increased PM for the following
distribution is a minimum between 25% and 50% EGR and reasons. As load is increased, the mass of fresh charge drawn
increases as EGR is either increased or decreased from those into the cylinder increases, because intake air and fuel flow
levels. rates increase and because the residual gas fraction in the
The spreading of the size distribution is consistent with cylinder decreases (49). But increased intake flow rates as
the following physical phenomena. As EGR is increased, the well as decreased blow-back into the intake ports result in
average size of particle nuclei decreases, presumably for the decreased intake temperatures. The result is that not only
same reasons that the concentration of nuclei decreases, as does the concentration of gas phase fuel increase but so
discussed above. However, as EGR is increased, the increased does the amount of liquid fuel, since the liquid fuel is less
HC concentration encourages growth. The particles that are likely to evaporate due to lower port/manifold temperatures
most likely to grow are those with the most surface area: (50). Moreover, as load increases, peak cylinder temperatures
thus, the largest particles grow faster than the smaller and pressures increase, as calculated using the GM Engine-
particles. The combined result of the competing nucleation Simulation Code (33). Consequently, the amount of par-
and growth effects is that nucleation-mode particles are ticulate nucleatedswhich depends strongly on gas and liquid
smaller at high levels of EGR than at the low levels, while the phase fuel availability as well as temperaturesshould increase
agglomeration-mode particles are larger at high levels of EGR as load increases. Furthermore, the decreased residual gas
than at the low levels. The reason number- and mass- fraction and increased cylinder pressure result in higher
weighted mean particle sizes are smaller at 25% EGR than oxygen concentrations at higher loads; since the probability
at zero EGR may be that the decreased size of nuclei (due of liquid fuel ignition and subsequent soot production is
to decreased liquid- and gas-phase fuel concentrations as strongly dependent on oxygen concentration (as discussed
well as decreased temperatures) overwhelms the slight above), the increased oxygen concentration with increased
increase in HCs available for growth (2400 ppmC1 at zero load should further increase the amount of PM nucleated
EGR versus 2550 ppmC1 at 25% EGR). through liquid fuel burning.
PM as a Function of Engine Load. The effect of engine
load on PM number (Figure 13) and mass (Figure 14) Conversely, as load increases, in-cylinder and exhaust
emissions was measured at three engine speeds. All condi- temperatures increase as does the partial pressure of oxygen,
tions were the baseline conditions, except IMAP (load) and thus facilitating PM oxidation. Increased load also results in
speed, which were the nine permutations of 0.4, 0.7, and 1.0 decreased mole fractions of exhaust HCs, thus resulting in
bar IMAP and 1500, 2000, and 2500 rpm, respectively. (Even a decreased concentration of HC vapor available for adsorp-
though the torque meter measured MBT spark timing at 7-10 tion/absorption. Nevertheless, the general increase in PM
°CA BTC for the three wide open throttle conditions, light with increased load (Figures 13 and 14) demonstrates that
and erratic knock occurred at those conditions.) the increased nucleation rate overwhelms the increased
As load is increased, PM concentrations increase mono- oxidation and decreased growth.
tonically. Number concentrations increase relative to con- Characteristic PM sizes show no universal trends as a
centrations at the baseline engine load by a factor of 7 at function of engine load: increasing with IMAP at some speeds
2500 rpm, a factor of 90 at 2000 rpm, and a factor of 430 at and loads, decreasing at others. The fact that particle
1500 rpm. Mass concentrations increase by 30% at 2500 rpm, concentrations monotonically increase with increased load
by a factor of 80 at 2000 rpm, and by a factor of 370 at 1500 while particle sizes do not have any consistent trend
rpm. Measured intake manifold temperature decreases by corroborates the inference that higher loads induce nucle-
4 °C as IMAP is increased from 0.4 to 1.0 bar at 2000 rpm. ation of greater numbers of particles.

VOL. 33, NO. 22, 1999 / ENVIRONMENTAL SCIENCE & TECHNOLOGY 9 3965
increase with speed at some speeds and loads but decrease
at others. As with particle concentrations, the lack of
monotonic trends may reflect competing formation, growth,
and oxidation mechanisms.

Acknowledgments
This work was sponsored by the EPA Center On Airborne
Organics at MIT, the MIT Engine-Fuels Interaction Consor-
tium (DaimlerChrysler, Ford, General Motors, Mobil, Peugeot,
Renault, Shell, and Volvo), and the National Science Foun-
dation. Invaluable assistance was lent by undergraduate
assistant Scott Whitehead and by Scott Carpenter at GM
Saturn.

Literature Cited
(1) Breath-Taking: Premature Mortality Due to Particulate Air
Pollution in 239 American Cities; Natural Resources Defense
Council: New York, 1996.
(2) Kayes, D.; Hochgreb, S. SAE Tech. Pap. Ser. 1998, No. 982601.
(3) Kayes, D.; Hochgreb, S. Environ. Sci. Technol. 1999, 33, 3968-
FIGURE 15. Number-weighted PM concentrations as a function of 3977.
engine speed. Three intake pressures, all other conditions same (4) Kayes, D.; Hochgreb, S. Environ. Sci. Technol. 1999, 33, 3978-
as baseline. 3992.
(5) Williams, D. J.; Milne, J. W.; Roberts, D. B.; Kimberlee, M. C.
Atmos. Environ. 1989, 23, 2639.
(6) Kosowski, M. G. SAE Tech. Pap. Ser. 1985, No. 850294.
(7) Quader, A. A. SAE Tech. Pap. Ser. 1989, No. 890623.
(8) Kayes, D.; Hochgreb, S. Combust. Sci. Technol. 1997, 127, 333
(correction: 1998, 37, 391).
(9) Thompson, N. D.; Wallace, J. S. SAE Tech. Pap. Ser. 1994, No.
940480.
(10) Min, K.; Cheng, W. K.; Heywood, J. B. SAE Tech. Pap. Ser. 1994,
No. 940306.
(11) Pedersen, P. S.; Ingwersen, J.; Nielsen, T.; Larsen, E. Environ.
Sci. Technol. 1980, 14, 71.
(12) Fanick, E. R.; Whitney, K. A.; Bailey, B. K. SAE Tech. Pap. Ser.
1996, No. 961089.
(13) Kittelson, D. B.; Dolan, D. F.; Beaver, M. C. Presented at Central
States Section of the Combustion Institute, 1979; Paper CSS/
CI-79-09.
(14) Maricq, M. M.; Dearth, M. A.; Chase, R. E.; Ball, J. C. Presented
at Health Effects Institute Workshop on Particle Characteriza-
tion, Cambridge, MA, December 1996.
(15) Witze, P. O.; Green, R. M. SAE Tech. Pap. Ser. 1997, No. 970866.
(16) Alkidas, A. C.; Drews, R. J. SAE Tech. Pap. Ser. 1996, No. 961958.
(17) Fulcher, S. K.; Gajdeczko, B. F.; Felton, P. G.;, Bracco, F. V. SAE
Tech. Pap. Ser. 1995, No. 952482.
(18) Amann, C. A.; Stivender, D. L.; Plee, S. L.; MacDonald, J. S. SAE
FIGURE 16. Mass-weighted PM concentrations as a function of Tech. Pap. Ser. 1980 No. 800251.
(19) Hildemann, L. M.; Cass, G. R.; Markowski, G. R. Aerosol Sci.
engine speed. Three intake pressures, all other conditions same Technol. 1989, 10, 193.
as baseline. (20) Model 3934 SMPS (Scanning Mobility Particle Sizer) Instruction
Manual; TSI Incorporated: St. Paul, MN, 1996; pp A-1, B-2-5,
PM as a Function of Engine Speed. In conjunction with B-17-19.
measurements of PM at different load conditions, number- (21) Seinfeld, J. H. Atmospheric Chemistry and Physics of Air Pollu-
and mass-weighted PM concentrations were measured as a tion, 1st ed.; John Wiley and Sons Publishing: New York, 1986;
function of engine speed at three load conditions (Figures p 281.
15 and 16). Neither the number nor the mass concentration (22) Hildemann, L. M.; Markowski, G. R.; Jones, M. C.; Cass, G. R.
showed a universal, monotonic trend with respect to engine Aerosol Sci. Technol. 1991, 14, 138.
(23) Kittelson, D. B. J. Aerosol Sci. 1998, 29 (5/6), 575.
speed. (24) Hofeldt, D. L. SAE Tech. Pap. Ser. 1993, No. 930079.
Increased engine speed results in higher gas temperatures (25) Roessler, D. M.; Faxvog, F. R.; Stevenson, R.; Smith, G. W. In
(both in-cylinder and exhaust) (9, 33) and in lower residence Particulate Carbon: Formation During Combustion, 1st ed.;
times. Higher peak temperatures increase both PM formation Siegla, D. C., Smith, G. W., Eds.; Plenum: New York, 1981; p 57.
and oxidation rates, whereas the reduced residence time (26) Dyer, T. M.; Flower, W. L. In Particulate Carbon: Formation
limits the duration over which the rates extend. The During Combustion, 1st ed.; Siegla, D. C., Smith, G. W., Eds.;
Plenum: New York, 1981; p 363.
competing factors are reflected in the rise and fall of PM (27) Kadota, T.; Henein, N. A. In Particulate Carbon: Formation
number concentrations with speed and in the load-depend- During Combustion, 1st ed.; Siegla, D. C., Smith, G. W., Eds.;
ent behavior of mass-weighted PM concentrations as a Plenum: New York, 1981; p 391.
function of speed. HC mole fractions vary by no more than (28) Lindstedt, P. R. In Soot Formation in Combustion, 1st ed.;
14% as a function of engine speed and do not vary Bockhorn, H., Ed.; Springer-Verlag: New York, 1994; Chapter
monotonically with speed. Therefore, HCs should not 27.
(29) Melton, L. A. Appl. Opt. 1983, 23 (13), 2201.
contribute to a significant, monotonic trend in growth via (30) Hildemann, L. M.; Markowski, G. R.; Cass, G. R. Environ. Sci.
adsorption/absorption as a function of speed. Technol. 1991, 25 (4), 744.
Characteristic PM sizes show no universal trends as a (31) Rickeard, D. J.; Bateman, J. R.; Kwon, Y. K.; McAughey, J. J.;
function of engine speed. Number- and mass-weighted sizes Dickens, C. J. SAE Tech. Pap. Ser. 1996, No. 961980.

3966 9 ENVIRONMENTAL SCIENCE & TECHNOLOGY / VOL. 33, NO. 22, 1999
(32) Graskow, B. R.; Kittelson, D. B.; Ahmadi, M. R.; Morris, J. E. SAE (43) Gelfand, B. E. Prog. Energy Combust. Sci. 1996, 22, 201.
Tech. Pap. Ser. 1998, No. 980528. (44) Maricq, M. M.; Podsiadlik, D. H.; Chase, R. E. Environ. Sci.
(33) Meintjes, K. General Motors Engine-Simulation Program (CSC), Technol., 1999, 33, 1618.
version 1.0; General Motors: Warren, MI, 1987. (45) Cheng, W. K.; Hamrin, D.; Heywood, J. B.; Hochgreb, S.; Min,
(34) Reynolds, W. C. STANJAN; Stanford University: Palo Alto, CA, K.; Norris, M. SAE Tech. Pap. Ser. 1993, No. 932708.
1986. (46) Nagle, J.; Strickland-Constable, R. F. In Proceedings of the
(35) Langmuir, I. J. Am. Chem. Soc. 1918, 40, 1361. Fifth Carbon Conference, 1st ed.; Pergamon: New York, 1962;
(36) Plee, S. L.; MacDonald, J. S. SAE Tech. Pap. Ser. 1980, No. 800186. p 154.
(37) Ladommatos, N.; Rose, D. W. SAE Tech. Pap. Ser. 1996, No. (47) Appleton, J. P.; Park, C. Combust. Flame 1973, 20, 369.
960467. (48) Quader, A. A. SAE Tech. Pap. Ser. 1971, No. 710009.
(38) Arcoumanis, C.; Gold, M. R.; Whitelaw, J. H.; Xu, H. M.; Gaade,
(49) Fox, J. W.; Cheng, W. K.; Heywood, J. B. SAE Tech. Pap. Ser.
J. E.; Wallace, S. SAE Tech. Pap. Ser. 1998, No. 981186.
1993, No. 931025.
(39) Alkidas, A. C. SAE Tech. Pap. Ser. 1994, No. 941959.
(40) Meyer, R.; Heywood, J. B. SAE Tech. Pap. Ser. 1999, No. 1999- (50) Posylkin, M.; Taylor, A. M. K. P.; Vannobel, F.; Whitelaw, J. H
01-0567. SAE Tech. Pap. Ser. 1994, No. 941989.
(41) Ladommatos, N.; Rose, D. W. SAE Tech. Pap. Ser. 1996, No.
960468. Received for review October 23, 1998. Revised manuscript
(42) Kowalik, R. M.; Ruth, L. A.; Edelman, R. B.; Farmer, R. C.; Wong, received July 12, 1999. Accepted September 2, 1999.
E.; Wang, T. S. Technical Report DOE/ET/11313-11; U.S. Depart-
ment of Energy: Washington, DC, 1982. ES9810991

VOL. 33, NO. 22, 1999 / ENVIRONMENTAL SCIENCE & TECHNOLOGY 9 3967

You might also like