Sunteți pe pagina 1din 24

Downloaded from SAE International by Indian Institute of Technology - Chennai, Monday, July 17, 2017

SAE TECHNICAL
PAPER SERIES 2006-01-0197

The Effect of Swirl Ratio and Fuel Injection


Parameters on CO Emission and Fuel Conversion
Efficiency for High-Dilution, Low-Temperature
Combustion in an Automotive Diesel Engine
Sanghoon Kook and Choongsik Bae
Korea Advanced Institute of Science and Technology

Paul C. Miles and Dae Choi


Sandia National Laboratories

Michael Bergin and Rolf D. Reitz


University of Wisconsin-Madison

Reprinted From: Compression Ignition Combustion Process 2006


(SP-2012)

2006 SAE World Congress


Detroit, Michigan
April 3-6, 2006

400 Commonwealth Drive, Warrendale, PA 15096-0001 U.S.A. Tel: (724) 776-4841 Fax: (724) 776-5760 Web: www.sae.org
Downloaded from SAE International by Indian Institute of Technology - Chennai, Monday, July 17, 2017

The Engineering Meetings Board has approved this paper for publication. It has successfully completed
SAE's peer review process under the supervision of the session organizer. This process requires a
minimum of three (3) reviews by industry experts.

All rights reserved. No part of this publication may be reproduced, stored in a retrieval system, or
transmitted, in any form or by any means, electronic, mechanical, photocopying, recording, or otherwise,
without the prior written permission of SAE.

For permission and licensing requests contact:

SAE Permissions
400 Commonwealth Drive
Warrendale, PA 15096-0001-USA
Email: permissions@sae.org
Tel: 724-772-4028
Fax: 724-776-3036

For multiple print copies contact:

SAE Customer Service


Tel: 877-606-7323 (inside USA and Canada)
Tel: 724-776-4970 (outside USA)
Fax: 724-776-0790
Email: CustomerService@sae.org

ISSN 0148-7191
Copyright 2006 SAE International
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.

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

Printed in USA
Downloaded from SAE International by Indian Institute of Technology - Chennai, Monday, July 17, 2017

2006-01-0197

The Effect of Swirl Ratio and Fuel Injection Parameters


on CO Emission and Fuel Conversion Efficiency for
High-Dilution, Low-Temperature Combustion in an
Automotive Diesel Engine
Sanghoon Kook and Choongsik Bae
Korea Advanced Institute of Science and Technology

Paul C. Miles and Dae Choi


Sandia National Laboratories

Michael Bergin and Rolf D. Reitz


University of Wisconsin-Madison

Copyright 2006 SAE International

ABSTRACT (often termed the thermal efficiency), which can exceed


the fuel conversion efficiency of modern SI engines by
Engine-out CO emission and fuel conversion efficiency as much as 40% [1]. However, for conventional diesel
were measured in a highly-dilute, low-temperature diesel combustion regimes, emissions of NOx and particulates
combustion regime over a swirl ratio range of 1.447.12 from these engines are excessive, and costly exhaust-
and a wide range of injection timing. At fixed injection gas after-treatment systems are required to meet current
timing, an optimal swirl ratio for minimum CO emission emission regulations.
and fuel consumption was found. At fixed swirl ratio, CO
emission and fuel consumption generally decreased as Recently, diesel combustion systems employing low
injection timing was advanced. Moreover, a sudden combustion temperatures and enhanced fuel-air pre-
decrease in CO emission was observed at early injection mixing have been proposed to drastically reduce NOx
timings. Multi-dimensional numerical simulations, and particulate emissions (e.g. [2]-[6]). In these systems,
pressure-based measurements of ignition delay and low combustion temperatures are achieved by
apparent heat release, estimates of peak flame employing high levels of exhaust gas recirculation (EGR)
temperature, imaging of natural combustion luminosity or retarded injection timing, while pre-mixing is
and spray/wall interactions, and Laser Doppler enhanced by the use of split injections, high-injection
Velocimeter (LDV) measurements of in-cylinder pressures, or measures that increase the ignition delay.
turbulence levels are employed to clarify the sources of This work focuses on those systems that employ high
the observed behavior. Mixing processes occurring after EGR rates to maintain low combustion temperatures,
the pre-mixed burn are found to be the likely source of with no special measures taken to increase pre-mixing
the optimal swirl ratio, while enhanced pre-combustion other than the increased ignition delay associated with
mixing dominates the reduction in CO with earlier dilute mixtures.
injection. Liquid fuel films formed on the bowl lip are not
found to significantly impact CO emissions, and Traditionally, low combustion temperatures are thought
increased injection pressure typically reduces CO to reduce NOx emissions but increase particulate
emissions at this high dilution rate. Fuel conversion emissions due to a reduction in particulate oxidation
efficiency is examined in terms of component rates. However, as EGR rates are increased to very high
efficiencies related to combustion phasing, heat loss, levels (approximately 60% or greater) simultaneous soot
and combustion efficiency. The influence of swirl and and NOx reduction is observed [3, 5]. At these high
injection timing on each of these efficiencies is dilution levels, characterized by O2 concentrations of
discussed. about 11%, the flame temperatures are limited to levels
at which the soot and NOx formation rates are low.
Unfortunately, the region of applicability of these low
INTRODUCTION
soot and NOx combustion regimes within the load-speed
Diesel engines are an attractive option for automotive operating map is often limited by high CO and UHC
powerplants due to their high fuel conversion efficiency emissions, with an accompanying fuel economy penalty.
Downloaded from SAE International by Indian Institute of Technology - Chennai, Monday, July 17, 2017

2000 300
Skip-fired CO Emission [ppm]

290

1500 CO 280

ISFC [g/kW-hr]
270

1000 260

250

500 240

230
' ISFC | 4%
0 220
-30 -25 -20 -15 -10
SOI [CAD]

Figure 1 The influence of injection timing on CO


emissions and fuel economy for a highly-dilute
(O2=10.1%) intake charge measured in the same
engine of this study. 1500 RPM, 3 bar IMEP,
Pinj=800bar, Rs=3.77. Reproduced from Ref. [9] Figure 2 Optical engine geometry and experimental set-
up for velocity measurements. High-speed video
imaging of combustion luminosity and spray penetration
was also performed, observing the combustion chamber
CO emission generally stems from two primary sources. from below (via the 45 mirror) and from the side (via the
The first source, under-mixing of fuel, is the CO formed liner windows)
in the products of rich pre-mixed combustion. If mixing
rates are too low, this CO will fail to mix with sufficient
additional O2 to complete the oxidation process before
expansion quenches chemical reactions. The second In this paper, we focus on clarifying the causes of the
source, over-mixing, is associated with fuel that is CO emission and fuel economy trends observed in
mixed to very lean equivalence ratios during the ignition Fig. 1, through application of a combined approach
delay period. After ignition, these overly lean mixtures do employing conventional pressure-based diagnostics,
not reach sufficiently high temperatures to complete the optical diagnostics, and numerical simulation. In addition
oxidation process on engine time scales. to injection timing, we have also varied both the injection
pressure and the swirl ratio, in an effort to further clarify
For thoroughly pre-mixed, very lean combustion systems, both the dominant source of CO emission as well as the
the over-mixed fuel and subsequent low peak particular aspects of the mixing processes that influence
combustion temperature is the source of engine CO CO emission and fuel economy. Increased injection
emission [7]. Similarly, the results of numerical pressure can be expected to primarily influence mixing
simulations point to over-mixed fuel as the principal processes during the ignition delay period, while
source of CO emission from conventional diesel increased swirl is anticipated to influence the later
combustion systems [8]. However, as dilution levels mixing processes to a greater extent than the injection-
increase, the correlation between CO emission and dominated early mixing processes. Additionally, swirl is
ignition delay observed in dilute, low-temperature diesel a parameter which has been previously cited as
combustion systems becomes negative [9]that is, important for reducing emissions [2] in low-temperature
increased ignition delay is found to correspond to combustion systems, and is generally found to be
reduced CO emissions for highly-dilute mixtures. beneficial for soot emission and fuel economy in
Additionally, the average equivalence ratio at ignition is conventional diesel combustion systems (e.g. [10]-[12]).
estimated to exceed 2, and to increase with increasing
dilution level. Jointly, these observations indicate that for The dilution level employed is characteristic of the
highly-dilute systems the dominant CO emission source Toyota smokeless combustion regime [3] or the AVL
more likely stems from under-mixed fuel. HCLI regime [5], and represents a very low-NOx, low-
soot operating regime where CO emissions and fuel
More importantly, significant changes in CO emission economy have already begun to suffer [9]. In clarifying
and in fuel economy can be obtained by changing the the source of CO emission and changes in fuel economy,
injection timing, as is shown in Fig. 1. With early our objective is to identify potential pathways for
injection, i.e. SOI=-26 CAD ATDC, CO emission is improvement of the performance of these combustion
decreased 3-fold over emission levels measured with systems.
injection timing at -10 CAD ATDC. Clearly, mixing
processes can be significantly influenced by injection
timing, and potentially by other factors as well.
Downloaded from SAE International by Indian Institute of Technology - Chennai, Monday, July 17, 2017

EXPERIMENT Table 1 Engine specifications

RESEARCH ENGINE Basic Geometry


Bore: 79.5 [mm] Stroke: 85.0 [mm]
A single-cylinder optically-accessible diesel research Disp. Vol.: 422 [cm3] Comp. Ratio: 18.7
engine was employed to study the effect of swirl and
injection timing on CO emission and fuel conversion Valve Events
efficiency. The engine has characteristics typical of
small-bore engines intended for automotive applications, IVO: -360 CAD ATC EVO: 145 CAD ATC
i.e.: four valves; a central, vertical injector; a 6-hole; IVC: -140 CAD ATC EVC: -350 CAD ATC
minisac nozzle; and a concentric, re-entrant bowl. A
schematic of the engine is shown in Fig. 2, and its main Fuel Injection Equipment
specifications are shown in Table 1. The floor of the Bosch Flow No.: 320 Included Angle: 145
piston bowl was comprised of a quartz insert that Number of Holes: 6 Hydro-erosion: 10 %
permitted in-cylinder fluid velocity measurements to be Nozzle Style: Cylindrical Minisac (Bosch DLLA)
obtained from below via laser Doppler velocimetry.

The engine is equipped with a common-rail fuel injection


system capable of a maximum injection pressure of
1350 bar. This system allows the injection pressure,
injected mass (duration of injection) and start-of-injection mean effective pressure (IMEP) and the apparent heat
(SOI) to be set independently. The injector is release rate. Care was exercised to ensure that
instrumented with a needle-lift sensor that permits the pressure-record filtering did not introduce artificial side-
actual start of injection to be accurately determined. lobes in the calculated apparent heat release rates.
Additional details of the calculation methodology are
The approximate fueling rate was determined by presented in our earlier work [9]. In-cylinder pressure
measuring the mass injected into a small chamber which data, and the velocity data described below, were
seals against the engine head. The calibration was acquired with a resolution of 0.25 crank angle degrees
performed with the head/injector at operating (CAD). Unless otherwise specified, all crank angles
temperature. Fuel was injected into the chamber for 575 presented in this paper are reported as CAD after TDC
injection events, and the average injected quantity compression.
determined by dividing the total mass by the number of
injection events. This averaged value was used to Engine-out CO emission was measured using a
calculate the fuel conversion efficiency. Note that this California Analytical Instruments NDIR (Non-Dispersive
procedure accounts for thermal changes in the injector Infra-Red) CO analyzer (model 300). Samples were
performance, but does not account for the higher taken from the exhaust plenum and transferred to the
ambient pressures encountered when the engine is analyzer through a heated sample line. Moisture and
operating. Accordingly, the average injected mass is condensable hydrocarbons were removed from the
likely to be overestimated, resulting in low values for the sample by a condenser prior to the analyzer.
fuel conversion efficiency. Neglect of the influence of
cylinder pressure will also introduce a bias into the OPTICAL DIAGNOSTICS
estimated fueling rate, wherein the injected fuel amount
at later injection timing will be less due to the higher
Spray and Luminosity Imaging
cylinder pressures.
To provide information on spray-targeting, an expanded
The magnitude of the overestimation/bias in the injected
514.5 nm Ar+ ion laser beam was directed into the
fuel mass can be estimated by assuming that, at the
baseline injection pressure of 800 bar, a mean effective combustion chamber through one of the side liner
injection pressure of 400 bar is achieved. At the earliest windows shown in Fig. 2. A high speed digital video
injection timing employed, the average cylinder pressure camera (Integrated Design Tools Inc., X-Stream Vision)
during injection is 22 bar, resulting in an overestimation was employed to view the spray penetration and wall-
of the injected fuel by 3%. At the latest injection timing, film formation from below, via the extended piston and
the 47 bar cylinder pressure results in a 6% 45 mirror. The frame rate was set to 21,593 frames per
overestimation. As will be seen below, the magnitude of second, corresponding to 0.4168 CAD between images
this bias is small as compared to the magnitude of the at 1,500 rpm. The image resolution is 100 x 120 pixels
changes in fuel conversion efficiency observed. and an exposure time of 30 s was employed.

In-cylinder pressure measurements were acquired using High speed videos were also obtained of the natural
a water-cooled (KISTLER 6043A60) piezoelectric luminosity from soot emissions, viewing the combustion
pressure transducer. Measurements obtained over 50 chamber through a side window. In this case, the
engine cycles were averaged to calculate the indicated framing rate employed was 9,000 frames per second,
corresponding to 1 image per crank degree.
Downloaded from SAE International by Indian Institute of Technology - Chennai, Monday, July 17, 2017

Table 2 Operating Parameters Table 3 Variation in fueling rate with swirl ratio

Basic Geometry Swirl Ratio Inj. Duration Fueling Rate


Speed [rpm] 1500 [Psec] [mg/stroke]
Load 3 bar gross IMEP 1.44 409 8.325
2.59 409 8.325
TCoolant [C] 88
3.77 421 8.908
Simulated EGR [Mass %] 65
4.94 425 9.108
10.1% O2, 80.6% N2,
Intake charge composition 7.12 456 10.258
9.3% CO2
Typical In-Cylinder I 0.57
PInjection [bar] 600, 800, 1000, 1200
Swirl Ratio (Rs) 1.44, 2.59, 3.77, 4.94, 7.12 fuel injection occurring on only one of every four engine
PIntake [bar] 1.2 cycles. This skip-firing period allowed the species
concentrations in the exhaust system and the emission
TIntake [C] 90
analyzers to stabilize. Finally, data were acquired over
PExhaust [bar] 1.3 50 additional skip-fired cycles. For velocity
Start-of-Injection (SOI) measurements, data were acquired for 120 s after the
-30.25 to -7.75
[CAD ATDC] commencement of skip-fired operation. After data
acquisition, a fixed cool-down/cleaning period was
employed to ensure that the gross thermal state of the
engine was the same for all measurements.

Flow Velocity The swirl ratio is varied by throttling the low-swirl intake
ports. To achieve the large range of swirl ratios reported
Measurements of the tangential component of velocity here, the intake valve in the high-swirl port was fitted
were acquired with a custom-designed, fiber-coupled with a 180 shroud, oriented to enhance the component
LDV. A laser power of approximately 260 mW per beam of the intake flow tangential to the cylinder wall. The
was used at a wavelength of 514.5 nm. A 5.0 MHz swirl ratios reported in Table 2 are the numeric values
differential frequency shift removed directional ambiguity. resulting from flow-bench measurements performed by
ZrO2 seed particles were introduced with a fluidized bed Ricardo, Inc. Radial profiles of the tangential velocity at
seeder. The seed particles have a mean diameter of three axial locations, measured with the LDV through the
0.48 m and a standard deviation of 0.15 m. Additional liner windows, provide a more direct measure of the in-
details can be found in Ref. 21.
cylinder angular momentum. At -55 CAD, integration of
these profiles resulted in swirl ratios of 1.10, 2.75, 3.86,
ENGINE OPERATION
5.05, and 6.53. With the exception of the lowest swirl
ratio, these measurements are within 8% of the flow-
The engine operating parameters are described in
bench swirl ratios. The swirl ratios reported throughout
Table 2. All data were acquired at a fixed engine speed
the remainder of this document are those determined via
of 1500 rpm.
flow-bench.
Recirculated exhaust gas was simulated by diluting the
The fueling rate was controlled to provide a constant 3
intake air stream with N2 and CO2. The relative
bar gross IMEP (indicated mean effective pressure) for
proportions of CO2 and N2 were chosen to match the each swirl ratio, at the injection timing providing best fuel
mixture molar specific heat of real engine exhaust gas at economy. Thereafter, as SOI was varied, the fueling rate
the selected load and O2 concentration. This matching was held fixed. The most advanced injection timing
was performed with gas properties evaluated at 600 K. considered at each swirl ratio was determined by the
The 10.1% O2 concentration selected corresponded to a achievable IMEP. When the IMEP dropped significantly
65% exhaust gas recirculation (EGR) rate. At Rs = 3.77 from the desired load no earlier timings were considered.
the typical in-cylinder equivalence ratio ( I ), was 0.57 As discussed below, excessive spray penetration due to
(O= 1.75), corresponding to an equivalence ratio based low in-cylinder gas density is thought to limit the earliest
on intake air of 0.79 (O= 1.27). practical injection timing. Injection timing was retarded to
the crank angle at which impaired fuel economy was
Prior to obtaining in-cylinder pressure, exhaust observed at the lower swirl ratios, typically -10.25 CAD.
emissions and flow velocity, the engine was motored for The approximate fueling rates employed for each swirl
a minimum of 90 s in order to pre-heat the combustion ratio are listed in Table 3. The fuel employed was a 2007
chamber walls and to allow the intake plenum pressure emission certification diesel with a cetane number of
to stabilize. For in-cylinder pressure and exhaust 47.1 and a lower heating value of 42.98 MJ/kg.
emission data, the engine was skip-fired for 70 s, with
Downloaded from SAE International by Indian Institute of Technology - Chennai, Monday, July 17, 2017

NUMERICAL SIMULATION 7 4 Skip-Fired


CO Emission 3500
In previous studies, the three-dimensional intake flow
field was calculated with STAR-CD. At IVC, those results
6 3000

Swirl Ratio
were mapped to a sector grid and used to initialize the
KIVA-3V [14] simulations of the combustion process. 5 2500
However, a comparison of the temporal history of the
tangential velocity and turbulent kinetic energy for 4 2000
various initialization schemes revealed that a simpler
solid-body-rotation swirl initialization performed well 3 1500
compared to the mapped results [15]. Due to the 1000
computational expense of computing the intake flow for 2
the range of swirl velocities considered here (Rs = 1.44 [ppm]
7.12), solid-body-rotation was used for initialization of -30 -25 -20 -15 -10
the tangential velocity in this study. The usual practice SOI [CAD ATDC]
of initializing the axial velocities by linearly interpolating
between the piston surface and the head, and initializing
the radial velocities to zero was followed. Figure 3 Effect of swirl ratio and injection timing on CO
emission
The spray and combustion sub-models employed are
described more fully elsewhere ([15],[16]). The
characteristic time combustion sub-model [17], is of observed optimal swirl ratio is somewhat unexpected.
greatest relevance to the present study. In this model, Higher swirl ratios are typically thought to enhance
the instantaneous production and oxidation rates of the turbulence production and mixing processes, resulting in
important combustion species are not calculated directly more rapid, complete combustion [18]. Consequently, if
via either simplified or detailed chemical mechanisms. under-mixed fuel is the dominant source of CO
Rather, the rate of change of each species is calculated emissions, we would anticipate a continuous
as proportional to the difference between the actual improvement in CO emissions as the swirl level is
species mass fraction and its instantaneous thermo- increased.
dynamic equilibrium value. The constant of propor-
tionality (the inverse of the characteristic time) is a The existence of an optimal swirl level suggests that a
weighted combination of a chemical timescale and a mechanism by which the highest swirl levels impede
turbulent timescale, with progressively greater weighting mixing and combustion is present. However, the
toward the turbulent scale as the combustion proceeds. traditional mechanism by which excessive swirl is
thought to impede mixing, by limiting jet penetration
RESULTS AND DISCUSSION (e.g. [19],[20]) is not believed to be significant for the
small bore diameters and high injection pressures that
The results of this study are presented in four main parts. characterize modern automotive diesel engines. This
In the first part, the variation in CO emissions with swirl belief is supported by Fig. 4, which shows the fuel mass
ratio and injection timing is reported. Second, the source distributions predicted for various swirl ratios at -17 CAD
of the influence of swirl ratio on CO emissions is clarified ATDC. In these simulations, SOI was fixed at -22.25
through examination of the heat release characteristics, CAD and the crank angle shown occurs shortly after the
numerical simulations and in-cylinder imaging of end of injection. A high fuel mass fraction is observed
combustion luminosity. Third, the factors potentially near the bowl lip at all swirl ratios, indicating penetration
responsible for the observed dependency of CO of the fuel jet to the maximum possible extent. At the
emissions on injection timing are examined. Finally, we higher swirl ratios, more fuel is observed in the cut-plane
examine the influence of both swirl ratio and injection shown (22 down-swirl of the fuel jet) due to greater
timing on fuel conversion efficiency. convection by the swirl velocity.

CO EMISSION TRENDS The second dominant feature of Fig. 3 is the rapid


decrease in CO emission observed at fixed swirl ratio
Figure 3 shows the engine-out CO emissions as a when the injection timing is advanced. This behavior is
function of both swirl ratio and injection timing (SOI). also apparent in Fig. 1. The decrease is larger at the
Two features are particularly notable: higher swirl ratios. For a swirl ratio of 7.12, CO emission
was reduced by over 83% as the injection timing was
First, at fixed injection timing, there is generally an advanced from -22.25 to -30.25 CAD ATDC.
optimal swirl ratio that gives the lowest CO emission.
The minimum CO emission is typically measured near In the following sections we investigate the causes of
the swirl ratio of 2.59. As noted in the Introduction, at these two behaviors.
this high dilution rate the dominant source of CO
emission is thought to be under-mixed fuel. Thus, the OPTIMAL SWIRL RATIO
Downloaded from SAE International by Indian Institute of Technology - Chennai, Monday, July 17, 2017

80

App. Heat Release Rate [J/CAD]


R
70 s

1.44
60 2.59
50 3.77
4.94
40 7.12
30
20
10
0
0.6

Cumulative AHRR / mf QLHV


0.5

0.4

0.3

0.2 Increasing Swirl

0.1

0
-15 -10 -5 0 5 10 15 20
Crank Angle [CAD]
Figure 5 Variation in the apparent heat release rate and
the cumulative heat release rate as the swirl ratio is
varied. SOI = -22.25 CAD

Beyond TDC the cumulative heat release curves


continue to diverge, but by this time a clear ordering with
swirl ratio in the amount of heat released has been
established. This ordering persists until EVO. The
Figure 4 Predicted fuel mass distribution for various behavior of the cumulative heat release curves suggests
swirl ratios at -17 CAD ATDC: SOI = -22.25 CAD, that the mixing processes responsible for this ordering
Pinj = 800 bar. The r-z plane velocity vectors are also commence between -5 CAD and TDC, and endure for
shown superimposed several crank degrees thereafter.

The ordering in the cumulative heat release is also found


As a first step toward examining the cause of the optimal to correlate inversely with the engine-out CO emissions.
swirl ratio exhibited in Fig. 3, the variation in the That is, lower CO emissions are observed for swirl ratios
measured apparent heat release rate and in the providing the highest cumulative heat release.
integrated, cumulative heat release rate with flow swirl Accordingly, insight into the mechanisms responsible for
level are shown in Fig. 5 for fixed SOI = -22.25 CAD. the CO emission behavior are likely to be found by
Note that, as observed previously for conventional diesel examining the mixing processes occurring during the
combustion over a more limited range of swirl ratios [12], latter portion of the combustion process, beyond -5 CAD.
increased swirl is found to reduce the ignition delay and
to advance the phasing of the early portion of the heat From Fig. 5, it is also apparent that, in addition to high
release. CO emissions, the high swirl cases will suffer from
reduced fuel economya topic that is discussed in
A second, more important observation can be made greater detail below.
from the cumulative heat release behavior shown in the
lower part of Fig. 5. By about -4 CAD, the heat release To help clarify the cause of the behavior seen in Figs. 3
rate observed for Rs = 7.12 slows considerably, followed and 5, the flow structures and spatial distributions of CO
by slowing of the heat release observed at Rs = 4.94 and (or partially-burned fuel) and O2 computed in the
Rs = 3.77.
Downloaded from SAE International by Indian Institute of Technology - Chennai, Monday, July 17, 2017

Figure 6 The mean r z plane flow structure and the Figure 7 Changes in the spatial distribution of O2 at
spatial distribution of CO at 3 CAD for various swirl 3 CAD for various swirl ratios, as predicted by numerical
ratios, as predicted by numerical simulation. simulation. SOI = -22.25 CAD, cut-plane 22 down-swirl
SOI = -22.25 CAD, cut-plane 22 down-swirl from the from the fuel jet
fuel jet

numerical simulation were examined. In our past work, crank angle is during the mixing-controlled portion of
we have shown that the mean (bulk) flow structures combustion, just after the secondary peak in the
predicted by the simulation capture the measured mean apparent heat release rates seen in Fig. 5. At this time,
flow structures remarkably well [21]. However, this the cumulative heat release rates seen in the lower
comparison was performed for a single, moderate swirl portion of Fig. 5 continue to diverge.
ratio. To extend this validation to the higher swirl ratios
employed here, we have measured axial profiles of the At the lowest swirl ratio, Rs = 1.44, the bulk flow
tangential velocity in the central region of the bowl at structure is dominated by a single clockwise-rotating
each swirl ratio, and compare these to the simulation structure that fills the outer portion of the bowl. In the
results in the Appendix. Although some differences exist, central region of the bowl, above the pip, a smaller
overall the comparison shows that the axial distribution secondary vortex rotating in the opposite direction is
of the tangential velocity is similar to the computed seen. From Fig. 7, which shows the distribution of the
results, and evolves in a like manner. remaining O2 in the combustion chamber, it is seen that
this secondary vortex transports O2 to the same region
Figure 6 shows the variation in the r z plane flow where the dominant vortex transports CO. In this
structures and in the spatial distribution of CO mass interfacial region between the two vortices, steep
fraction predicted for various swirl ratios at 3 CAD. This gradients in swirl velocity (Fig. 8), as well as high rates
Downloaded from SAE International by Indian Institute of Technology - Chennai, Monday, July 17, 2017

remaining O2. The mixing system created is thus


anticipated to be more effective than is seen for
Rs = 1.44, an expectation which is supported by the heat
release characteristics seen in Fig. 5. The dual-vortex
structures formed at these swirl ratios are long-lived; the
simulation results suggest that they continue to provide
effective mixing for at least another 5 CAD beyond the
crank angle illustrated in Figs. 68.

The formation of these structures correlates well with the


secondary peak in the apparent heat release rate, seen
near TDC in Fig. 5. Similar dual-vortex structures, whose
formation coincides with increased heat release rates,
have also been observed under late-injection operating
conditions [23].

Secondary peaks in the experimental heat release rate


near TDC are also seen in Fig. 5 for the higher swirl
ratios. Although small, these peaks are both
reproducible and of a magnitude that easily exceeds the
precision of the measured heat release during this
portion of the cycle, approximately 1 J/CAD [12].
Corresponding to the appearance of these peaks, the
numerical simulations also predict the formation of
counter-rotating vortex structures in the higher swirl
cases, though at slightly earlier crank angles. The
important distinction appears to be the lifetime of these
structures, and the manner in which they evolve and
transport the unburned fuel and oxidant later in the cycle.
Already by 3 CAD, Figures 6 and 7 indicate that for
Rs = 3.77 a third vortex structure is forming in the central
portion of the cylinder. The original upper vortex is
displaced outward, and rather than acting to bring the
partially-burned fuel (CO) and O2 to the same interface,
it now predominantly acts to transport the partially-
burned fuel away from the O2 in the center of the bowl.
This effect becomes even more pronounced as the swirl
level increases further. By Rs = 7.12 yet a fourth vortical
structure is apparent, and the original bowl vortex
dominant at Rs = 1.44is now barely visible in the
reentrant portion of the bowl.
Figure 8 Changes in the spatial distribution of the swirl
velocity at 3 CAD for various swirl ratios, as predicted by Comparison of the CO spatial distributions in Fig. 6 with
numerical simulation. SOI = -22.25 CAD, cut-plane 22 the swirl velocity distributions in Fig. 8 reinforces the
down-swirl from the fuel jet above discussion. In general, the partially-burned fuel is
spatially co-located with the high-swirl fluid. As the swirl
ratio increases, this co-location results in greater
of mean flow deformation in the r z plane ([22],[23]) average radial locations of the CO, due to the return of
lead to enhanced turbulence production. Thus, the the high-swirl fluid to the combustion chamber periphery.
observed pair of counter-rotating flow structures forms a At moderate swirl ratios (e.g. Rs = 2.59) this is advanta-
mixing system, wherein large-scale structures transport geous: not only is the dual-vortex structure formed, but
the partially-burned fuel and O2 to a common interface, the partially-burned fuel is positioned to exit the bowl into
while simultaneously providing mechanisms to produce the squish volume, where additional turbulence
the turbulence needed to achieve rapid small-scale generated by the exiting flow is anticipated to facilitate
mixing. mixing with the remaining O2 in the squish volume.

As swirl ratio is further increased to 2.59, the same dual- In contrast, at the highest swirl ratios, co-location with
vortex flow structure is again apparent. In this case, the high-swirl fluid both impedes mixing with the O2 in the
centrifugal forces acting on the high angular momentum center of the bowl and interferes with the exit of the
fluid transported into the inner regions of the bowl are partially-burned fuel from the bowl with the squish flow.
larger, and the lower vortex is smaller due to the Fundamentally, the centrifugal forces acting on the high-
tendency of this high angular momentum fluid to return swirl fluid prevent the inward displacement required to
to the bowl periphery. Consequently, the upper vortex is
larger and encompasses a greater fraction of the
Downloaded from SAE International by Indian Institute of Technology - Chennai, Monday, July 17, 2017

Figure 9 Simulation results illustrating the co-location of CO (partially-burned fuel) and high-swirl fluid during the expan-
sion stroke. Rs = 7.12

Figure 10 Images of natural luminosity from soot emissions at three swirl ratios. At higher swirl, the visible soot is lower
and more uniformly fills the bowl. Clockwise from the upper-left corner of each figure are superimposed the transmissivity
of the neutral density filter employed, the exposure time, the crank angle relative to SOI, and the true crank angle at which
the image was obtained

move the fluid over the bowl lip. The co-location of the digital video camera. Figure 10 compares the side-view
CO and the high-swirl fluid therefore persists until well images of the natural soot luminosity observed at
into the expansion stroke, effectively trapping the CO in Rs = 1.44 with that observed at higher swirl ratios,
the re-entrant portion of the bowl. This is illustrated in Rs = 3.77 and Rs = 7.12. Although the intensity of the
Fig. 9 for the highest swirl case, Rs = 7.12. soot luminosity is less, images obtained at the higher
swirl ratios clearly show that the luminous soot (which is
To provide additional experimental support for this anticipated to be co-located with partially-burned fuel) is
picture, the spatial distribution of soot luminosity, and its found deeper in the bowl and more uniformly fills the
temporal evolution, was captured using the high speed bowl than at the lower swirl ratios. This condition
Downloaded from SAE International by Indian Institute of Technology - Chennai, Monday, July 17, 2017

1.5 100

Rich Mass Fraction ( o > 1)


Apparant Heat Release Rate [J/CAD]
CO Mass [mg] SOI: -22.25 CAD ATDC
0.3
In-Cylinder

SOI: -30.25 CAD ATDC


1.0 80 SOI: -22.25 CAD ATDC

0.5 60
In-Cylinder CO mass 0.2
0.0
40 SOI: -30.25 CAD ATDC

Apparent Heat Release Rate


20 0.1

0
-30 -20 -10 0 10 20 0
Crank Angle [CAD ATDC] -40 0 40 80 120
Crank Angle [CAD ATDC]
Figure 11 Spatially-integrated in-cylinder CO mass from
the numerical simulations, and corresponding exper- Figure 12 Numerical simulation of the mass fraction of
imental apparent heat release for different injection the cylinder contents characterized by a (pre-
timings. Rs=7.12 combustion) equivalence ratio greater than 1. Rs=7.12

persists over all later crank angles during which the bowl experimentally-determined apparent heat release rates
is visible. for each case. There is a noticeable difference in the
amount of CO present in the cylinder, in agreement with
In summary, the numerically predicted bulk flow the experimental results shown in Fig. 3. For both cases,
structures provide a mechanism that explains the CO formation started in the middle of the premixed burn
observed optimal swirl ratio. At the lowest swirl ratios, a phase and the peak CO mass is found near the start of
dual-vortex structure forms that enhances the mixing of the mixing-controlled burn phase. Examining the
partially-burned fuel and air. For moderately increased evolution of the CO mass, it is apparent that the latter
swirl ratios, this dual-vortex system appears to be more oxidation rates are similar and that the lower CO mass
effective in enhancing the mixing processes. However, observed for the earlier injection timing is principally due
as swirl is increased further, the bulk structures created to lower initial CO formation or more rapid early
actually impede the late-cycle mixing, even though early oxidation of CO.
mixing and heat release appears to be enhanced.
Ultimately, too high a swirl ratio leads to trapping of The nature of the characteristic-time combustion sub-
partially-burned fuel within the bowl. model is such that CO must be formed from rich
mixtures. For lean mixtures, instantaneous equilibrium
RAPID DECREASE IN CO EMISSION WITH CO concentrations are small and the model will not
ADVANCED SOI predict large CO production rates; i.e., CO emissions
stemming from incomplete combustion of over-mixed
The potential causes of the second major observation fuel will not be captured. Thus, the higher CO mass
made from Fig. 3, a decrease in CO emissions as SOI is formed at the later injection timing must be caused by
advanced, are investigated in this section from several the presence of a greater amount of rich mixture. The
perspectives. Once again we rely heavily on the results fraction of rich mixture found in the cylinder for both
of numerical simulations, which suggest that the initial injection timings, shown in Fig. 12, supports this
degree of fuel-air mixing achieved is a critical factor. To conclusion. At the earlier injection timing, the peak
support this finding, we examine the measured influence fraction of the in-cylinder mass characterized by fuel-rich
on CO emission of injection pressure and ignition delay. conditions is smaller, and the fuel-rich mass fraction is
The correlation of CO emissions with peak flame already decreasing at the onset of combustion near -12
temperature is also examined for evidence that early CO CAD. For the later injection timing, however, the peak
oxidation rates could also be an important factor mass of rich in-cylinder mixture is not reached until after
influencing CO emission. Finally, we scrutinize other TDCwell into the mixing-controlled portion of the
processes that influence the early mixture preparation combustion process.
processincluding wall-wetting, spray-targeting, and
flow turbulence. The better mixture preparation characteristics observed
at the earlier injection timing in Fig. 12 are not the result
Simulation Results of an increased ignition delay. While increased ignition
delay is undoubtedly beneficial, the mixing process itself
Figure 11 shows simulation results for the total, in- is enhanced with the earlier injection timing, such that for
cylinder mass of CO for two different injection timings at equal times after the start of injection, a smaller fraction
the highest swirl ratio, Rs = 7.12. Also shown are of the in-cylinder mass is fuel-rich for the earlier injection
Downloaded from SAE International by Indian Institute of Technology - Chennai, Monday, July 17, 2017

4500 Rs=3.77
P : 600 bar 1800

CO Emission [ppm]
inj 800 bar
CO Emission [ppm]
1000 bar
3500

4 Skip-Fired
1200 bar SOI=-22.25 CAD ATDC
4 Skip-Fired

1400
2500
1000
SOI=-28.25 CAD ATDC
1500
SOI=-26.25 CAD ATDC

600
500
1.4 1.6 1.8 2 2.2
-30 -25 -20 -15 -10 -5
[ms]
SOI [CAD ATDC] id

Figure 13 Effect of injection pressure and injection timing Figure 14 Correlation of CO emission with the ignition
on CO emission at swirl ratio of 3.77 delay at swirl ratio of 3.77

timing. This enhanced mixture preparation is reflected in with the numerical predictions. At the earliest injection
the higher peak heat release rates seen in Fig. 11 for timings, however, increased injection pressure may be
SOI=-30.25, behavior which is well-captured in the detrimental.
simulation results, as are the ignition delay periods
observed for both injection timings. Ignition delay is well-known to increase as injection
timing is advanced, due to the lower ambient
Examination of the spatial distributions of fuel predicted temperature and pressure at earlier SOI [24]. Here, the
numerically suggest that the enhanced mixing observed ignition delay is defined as the time from SOI until the
with earlier injection is not due to a significant difference time at which 10% of the total heat release has occurred.
in the flow structures existing during the ignition delay The 10% heat release timing correlates well with the
period, but rather to initial deposition of a significant beginning of rapid, high-temperature heat release. As
fraction of fuel vapor within the squish volume. As the noted in the introduction, we have previously observed
piston rises, this fuel is returned to the bowl with the an inverse correlation at high dilution rates between CO
squish flow and results in a better mixed, broader spatial emission and ignition delay. In general provided spray
distribution of fuel vapor than is seen with later injection. targeting, wall impingement, or spray/flow interactions
Peak fuel vapor concentrations (and subsequent CO do not vary significantly with SOI increased ignition
concentrations), for both injection timings, are found delay should correlate well with a higher degree of fuel-
deep in the bowl, as can be seen in the lower portion of air premixing. In turn, more premixing will correlate
Fig. 6. In both cases, the CO remains trapped in the inversely with CO emission, provided CO emission is
lower regions of the bowl during expansion. For the dominated by the combustion of rich mixtures.
earlier injection timing, however, there is far less of it. Conversely, CO emissions resulting from over-mixed
fuel would be expected to increase with increasing
Thus, the simulation results suggest that the dominant ignition delay.
cause of the reduced CO emissions with advanced
injection timing is associated with reduced peak CO The correlation between ignition delay and CO emission
mass during the early combustion process. Early mixing measured for Rs=3.77 is shown in Fig. 14. This
processes, occurring prior to or during the first portion of intermediate swirl ratio was selected for detailed
the premixed burn, are at least partly responsible for the examination because a minimum in CO emission is
reduced CO mass. observed near SOI =-26 CAD in both Figs. 1 and 3. In
keeping with the predictions of the numerical simulations,
The Influence of Injection Pressure and Ignition Delay the variable injection pressure tests, and our previous
results, Fig. 14 shows a strong negative correlation of
The numerical simulations of the previous section CO emission with increasing ignition delay, indicating
suggest that more rapid initial mixing is an effective the benefits of increased early mixing for reducing CO
method of reducing CO emissionprovided that CO emission.
emission stems predominantly from under-mixed fuel. A
straight-forward test of this prediction can be made At the earliest injection timing shown in Fig. 14,
through increasing injection pressure, a measure that increased CO emission is observed despite an increase
will enhance initial mixing rates but which is less likely to

significantly impact the late-cycle mixing. As This view must be adopted with caution. As shown in
demonstrated by Fig. 13, increased injection pressure is the previous section, under some circumstances mixing-
found to generally reduce CO emissions, in agreement rates can be affected significantly by changes in SOI.
Downloaded from SAE International by Indian Institute of Technology - Chennai, Monday, July 17, 2017

Rs=3.77 temperature increases monotonically as injection is


2000
CO Emission [ppm] advanced. Notably, as SOI is advanced to -22.25 CAD,
SOI=-22.25 CAD ATDC
there is little change in CO emission despite a near 30 K
change in peak flame temperature. As SOI is further
4 Skip-Fired

advanced to -26.25 CAD, however, a significant re-


duction in CO emission is observed, although the flame
temperature only increases by 5 K. Overall, the
1000
900 observed correlation between peak flame temperature
SOI=-28.25 CAD ATDC
800 and CO emission is not consistent with the proposition
SOI=-26.25 CAD ATDC
that early CO oxidation significantly influences the
700 emission level. The subsequent increase in CO emission
600 as SOI is further advanced to -28.25 CAD, while the
2000 2010 2020 2030 2040 peak flame temperature continues to increase,
reinforces this view.
Peak Flame Temperature [K]
Spray Penetration and Targeting
Figure 15 Correlation of CO emission with peak flame
temperature at a swirl ratio of 3.77 An additional important factor related to early mixture
formation processes in small-bore diesels is spray
penetration and targeting. Excessive penetration will
in ignition delay. This behavior is both repeatable and clearly lead to piston wall-wetting, and in extreme cases
reflected in the CO emission behavior seen at other swirl can lead to liquid fuel impacting the cylinder liner (e.g.
ratios (see Fig. 3). Furthermore, it is consistent with the [25]). Spray targeting is expected to influence the initial
behavior seen at the earliest injection timings of Fig. 13, mixture formation process by affecting the initial
where increased injection pressure was no longer splitting of the spray upon striking the bowl lip, and to
beneficial. It would appear that at very early injection further influence the size and location of liquid films that
timing, factors other than ignition delay are influencing may be formed on the piston surface.
pre-mixing and subsequent CO emissionbehavior that
will be examined more closely below. Figures 16 and 17 illustrate the liquid spray penetration
and wall-film formation, as determined from high-speed
Flame Temperature imaging of the liquid sprays. Also shown, below the
experimental images, is the evolution of the gas-phase
The evidence presented thus far suggests that fuel distribution predicted in the numerical simulations.
decreased peak CO mass early in the combustion The computed fuel distributions are shown for a vertical
process is primarily responsible for the decrease in CO cut-plane 22 downstream of the spray axis. The center
emissions observed with advanced injection timing. of the piston and location of the bowl lip are scaled to
Simulation results suggest that the decreased CO mass the spray image and are designated with dashed lines.
is due to enhanced mixing. However, earlier injection The time after the start of injection and the current crank
results in advanced combustion phasing which results in angle are shown at the right-hand side of each gas-
higher peak cylinder pressures and higher peak phase fuel mass image.
combustion temperatures. Accordingly, it seems
plausible that the reduction in peak CO mass may be For the injection timing of -26.25 CAD seen in Fig. 16,
partly associated with the more rapid early oxidation of where the lowest CO emission was observed at the
CO due to these higher temperatures. Rs = 3.77 swirl ratio, the spray tip clearly impacts the
bowl lip and creates a liquid film. In contrast, with
We have selected the peak adiabatic flame temperature SOI = -22.25 CAD, Fig. 17 shows that the spray tip
to characterize the CO oxidation rates early in the reaches to the bowl lip, but little or no liquid film can be
combustion process (i.e., during the premixed burn seen. Formation of liquid films is generally thought to
period). The peak flame temperature is the adiabatic increase UHC emissions (e.g. [26]), and would likely
flame temperature achieved by combustion of a increase CO emissions as well due to an overall
stoichiometric mixture at the peak core charge retarded fuel preparation process. The lower CO
temperature, which is estimated assuming adiabatic emission observed at the earlier injection timing
compression of the in-cylinder charge. Although CO suggests, therefore, that formation of wall films is not a
oxidation will always occur at temperatures lower than dominant process influencing CO emission here.
this, the relative magnitude of the peak flame
temperature is likely a good indicator of the relative A second interesting observation can be made from
magnitude of CO oxidation temperatures. Figs. 16 and 17. The predicted gas-phase fuel mass
distributions show that for the earlier, -26.25 CAD
The correlation of CO emission with peak flame injection timing, some of the fuel penetrates into the
temperature is shown in Fig. 15, where the CO squish area. However, the squish flow induced by the
emissions measured at Rs = 3.77 are plotted on a rising piston pushes this fuel back into the bowl. By the
logarithmic axis. As anticipated, the peak flame end of the sequence shown, the fuel distribution is
Downloaded from SAE International by Indian Institute of Technology - Chennai, Monday, July 17, 2017

Figure 16 Visualized spray impingement on the bowl lip Figure 17 Visualized spray impingement on the bowl lip
and predicted fuel mass distribution: SOI = -26.25 CAD, and predicted fuel mass distribution: SOI = -22.25 CAD,
Rs = 3.77 Rs = 3.77

almost identical for both injection timings. Accordingly, At these more advanced injection timings, liquid films
the initial distribution of the vapor phase fuel does not can be expected to form not only on the bowl rim, but
appear to be a dominant factor in determining the CO also on the piston top. Figure 18 shows an idealized
formation and ultimately the CO emission level. This view of the spray targeting on the piston for various
stands in contrast to our previous discussion of the fuel crank angles. With a start of injection at -26 CAD and
vapor distributions observed for Rs = 7.12. Recall that, in later, we predominantly expect wetting of the bowl rim,
that case, significantly more advanced injection timing as was seen in Fig. 16. Minor wetting of the piston top
was considered. may occur with SOI = -26 CAD, but only from the earliest
portion of the injection event. For earlier injection timings,
Downloaded from SAE International by Indian Institute of Technology - Chennai, Monday, July 17, 2017

0
-5 CAD TDC
D
-5

z [mm]
-10
-15 CAD
-15

-20 -25 CA
AD ATDC
-22 CAD 0
r = 13.6
3 mm

Figure 19 Velocity measurement location and piston


position at -25 CAD ATDC: r = 13.6 and various axial (z)
locations
-26 CAD
greater period of time, and could significantly enhance
the early mixture preparation process.

To examine this potential source of enhanced mixing,


tangential velocity data were acquired at a radius of
-30 CAD r = 13.6 mm and axial locations of z = 2 20 mm below
the cylinder head. The data discussed below were
obtained at the locations z = 4 10 mm, which are
shown in Fig. 19, along with superimposed bowl outlines
Figure 18 Idealized depiction of the un-deflected spray which show the piston position relative to these
targeting for various crank angles measurement locations at various crank angles. At the
early injection timings considered here, these
wetting of the piston top is expected to become measurement locations were approximately centered
increasingly important, due to both increased along the path of a fuel spray during the injection event
penetration and the spray targeting. At high swirl ratios, (cf. Fig. 18). Later, these locations are characteristic of
we anticipate that vaporization of these films will be locations within the central region of the bowlwhere
more rapid, and their potential influence on emissions much of the later mixing occurs. All measurements were
diminished. Piston top wall films are thus a possible obtained in a plane 22 down-swirl of a fuel jet.
explanation for the increased CO emissions seen in Fig.
3 for the earliest injection timings and lower swirl ratios. The measured temporal evolution of the mean flow
velocity and RMS fluctuations are shown in Figs. 20 and
Turbulence Enhancement 21, respectively. Velocities measured under both
motored and fired operation are presented in each figure.
A final factor that may influence the early mixture The swirl ratio shown is 7.12. Recall that, as the injection
preparation process and increase mixing rates during timing is advanced, the maximum reduction in CO
the ignition delay period is enhanced flow turbulence. In emission was observed at Rs = 7.12 (see Fig. 3).
earlier work we have shown that, over the course of the Measurements obtained at injection timings of -22.25
latter compression stroke, production of turbulence by and -30.25 CAD are presented.
bulk compression acting on the isotropic normal stresses
can often dominate [16]. For a uniform density flow, this The mean flow evolution, shown in Fig. 20, is consistent
production can be expressed as: with both physical expectations and with the results of
numerical simulations. For motored operation, Fig. 20
shows that the tangential velocity increases in the latter
1 dV
Piso 2 k (1) part of the compression stroke as high angular
3 V dt
momentum is transferred into the bowl by the squish
flow. With fuel injection, little influence of the injection
Here, the production of turbulence is directly proportional event is observed until after injection has ended. After
to the level of pre-existing turbulent kinetic energy k, and the fuel injection event, a distinct perturbation in the
the negative of the logarithmic rate-of-change of cylinder tangential velocity is observed at all injection timings and
volume - 1 V dV dt a quantity which is large between measurement locations, associated with a displacement
about -30 and -5 CAD. Thus, turbulent energy injected of the high angular momentum fluid near the bowl rim by
earlier is amplified by the compression process over a the fuel spray. This perturbation rapidly subsides,
however, and beyond about -10 CAD there is little
Downloaded from SAE International by Indian Institute of Technology - Chennai, Monday, July 17, 2017

6
Rs = 7.12 Rs = 7.12
10 Pinj = 800 bar Pinj = 800 bar
r = 13.6 mm 5 r = 13.6 mm
8 z = 4 mm
4
6 3 Motored
Fired, -22.25 SOI
z = 4 mm Fired, -30.25 SOI
4 2
Fuel Injection Event
Motored
2 Fired, -22.25 SOI 1
Fired, -30.25 SOI
Fuel Injection Event
0
10
5
8 z = 6 mm
4
Motored
6 Fired, -22.25 SOI
3 Fuel Injection Event
Fired, -30.25 SOI

Tangential RMS Velocity [Sp]


z = 6 mm
4 2
Tangential Velocity [Sp]

Fuel Injection Event Motored


2 Fired, -22.25 SOI 1
Fired, -30.25 SOI

0
10
5
8 z = 8 mm
4
Fuel Injection Event Motored
6
3 Fired, -22.25 SOI
Fired, -30.25 SOI
z = 8 mm
4 2
Motored
2 -22.25 SOI 1
Fuel Injection Event
-30.25 SOI

0
10
5
8 z = 10 mm
4 Motored
Fired, -22.25 SOI
6 Fired, -30.25 SOI
3
Fuel Injection Event
z = 10 mm
4
2
Motored
2 Fuel Injection Event Fired, -22.25 SOI 1
Fired, -30.25 SOI

0 0
-30 -20 -10 0 10 -30 -20 -10 0 10
Crank Angle [CAD ATDC] Crank Angle [CAD ATDC]
Figure 20 Measured mean velocity in tangential direction Figure 21 Measured tangential velocity RMS fluctuations
at r = 13.6 and various axial (z) locations for Rs =7.12 at r = 13.6 and various axial (z) locations for Rs =7.12
Downloaded from SAE International by Indian Institute of Technology - Chennai, Monday, July 17, 2017

evidence that the mean flow evolves in a significantly


different manner than the motored flow. For SOI = - 7 Fuel Conversion Efficiency 34
30.25 CAD combustion begins at about -11 CAD, while
for SOI = -22.25 CAD combustion begins near -8 CAD 6 32

Swirl Ratio
(see Fig. 11). Hence, at these locations, we see little
effect of combustion on the mean flow evolution. 5 30
The corresponding RMS fluctuations are shown in
Figure 21. Changes in RMS velocity fluctuations could
4 28
be due to cycle-to-cycle fluctuations in the mean flow
structures or to variability of the fuel injection event, 3 26
rather than additional turbulence. However, previous
cycle-resolved analysis has shown that enhanced RMS 2 24
fluctuations observed after fuel injection correlate well
with enhanced small-scale, high-frequency fluctuations -30 -25 -20 -15 -10 [%]
characteristic of turbulent motions [27]. Consequently,
we believe that the increased RMS fluctuations seen in SOI [CAD ATDC]
Fig. 21 represent, at least in part, increased flow
turbulence. Figure 22 Effect of swirl ratio and injection timing on fuel
conversion efficiency
In general, increased RMS fluctuations are observed
coincident with the period in which the largest
perturbation in the mean flow structure is seen. At the 7
highest axial location, z = 4 mm, these fluctuations are Combustion Efficiency
particularly large, and are both larger in magnitude and
94
of a longer duration for the earlier injection timing. Later Swirl Ratio 6 92
in the cycle, a more moderate increase in fluctuations
over the motored fluctuations is seen, which typically 5 90
persists for at least 10 CAD. Considering the -11 and -8
CAD start-of-combustion for the -30.25 and -22.25 CAD 4 88
injection timings, respectively, we see that this more
moderate increased turbulence, which appears earlier 86
for the earlier injection timing, may also increase pre- 3
combustion mixing. Overall, the measurements indicate 84
that with earlier injection, turbulence levels are higher 2 82
and act on the pre-combustion mixture for a longer
period. Increased turbulent mixing may therefore assist -30 -25 -20 -15 -10 [%]
in the early mixture formation process, complementing
the longer ignition delay. However, it must be recognized
SOI [CAD ATDC]
that the spatial locations sampled with LDV are sparse,
and increased turbulence will not increase mixing rates if Figure 23 Effect of swirl ratio and injection timing on
it occurs in regions with a homogeneous mixture combustion efficiency
composition.

FUEL CONVERSION EFFICIENCY The gross indicated fuel conversion efficiency is the ratio
of the gross indicated work W to the chemical energy in
In the previous sections we have examined how swirl the injected fuel, calculated from the product of the fuel
ratio and combustion phasing (injection timing) influence mass and the lower-heating value QLHV of the fuel
CO emission, which is an indication of complete,
efficient combustion. Complete combustion is a pre- W
K fc (2)
requisite for obtaining a high fuel conversion efficiency m f Q LHV
Kfc . However, other factors also influence Kfc . In this
section our goal is to clarify the broader impact of As an aid to understanding the influence of various
variations in swirl ratio and injection timing on Kfc . In factors on Kfc (including CO emission and combustion
pursuing this goal, we consider only the gross indicated efficiency), we have previously proposed [9] that Kfc be
fuel conversion efficiency. The net fuel conversion decomposed as a product of three separate efficiencies
efficiency may be influenced through other factors (e.g.
increased pumping work associated with high-swirl port K
designs) that are not addressed here. Kwc
 K hl
 c 

W Qchem  Qhl Qchem
K fc (3)
Q  Q Q m Q
chem hl chem f LHV
Downloaded from SAE International by Indian Institute of Technology - Chennai, Monday, July 17, 2017

The fuel conversion efficiency, calculated from Eq.(2), is


7 Work Conversion Efficiency shown in Fig. 22. For comparison, an estimate of the
40 combustion efficiency is calculated from the CO
6 emissions (see Fig. 3) using the expression
Swirl Ratio

5 38 'h 0f ,CO  'h 0f ,CO2


Kc 1  f (4)
'H R
4 36

In Eq.(4), f represents the fraction of fuel carbon found in


3 34 the CO emissions, 'H R is the negative of the fuel lower
heating value, and 'h 0f is the heat of formation of the
2 32 sub-scripted species. Additional details can be found in
[%] Ref. 9. Combustion efficiencies estimated from Eq.(4)
-30 -25 -20 -15 -10 are shown in Fig. 23.
SOI [CAD ATDC]
Although many similarities exist between Figs. 22 and
Figure 24 Effect of swirl ratio and injection timing on 23, there are also distinct differences that bear noting:
work conversion efficiency
1) While the combustion efficiency remains high in the
upper-left quadrant of Fig. 23, K fc is decreasing at
the higher swirl ratios.

7 Heat Loss Efficiency 95 2) The best K fc is observed at somewhat lower swirl


ratios and more retarded injection timing than the
6 90 best Kc
Swirl Ratio

3) In the lower-right quadrant of Fig. 22, a rapid


5 85 decrease in K fc is observed. No similar decrease
can be seen in Kc .
4 80
4) In the lower-left quadrant, a local minimum in K fc is
3 75 seen that is not present for Kc

2 70 To assist in the explanation of these differences, the


work conversion efficiency is presented in Fig. 24. All of
-30 -25 -20 -15 -10 [%] the differences between K fc and Kc identified above can
be seen to correlate well with the variation of K wc with
SOI [CAD ATDC] SOI and with swirl ratio. First, the work conversion
efficiency can be seen to decrease gradually in the
Figure 25 Effect of swirl ratio and injection timing on upper-left quadrant of Fig. 24. Thus, decreased K fc at
heat-loss efficiency early SOI and high swirl ratio can be, at least partially,
explained by decreased K wc . Second, peak work
conversion efficiencies are observed at lower swirl ratios
Because the difference between the chemical heat and at more retarded injection timing than peak Kc , an
release Q chem and the energy losses due to heat transfer observation that helps explain the second difference
and crevice flows Q hl is the apparent net heat release, between Figs. 22 and 23 enumerated above. Third, the
the first term on the right-hand-side (RHS) represents behavior of K fc in both the lower-left and the lower-right
the fraction of the apparent heat release that is quadrants is mirrored in K wc .
converted into useful workthe work conversion
efficiency K wc . Changes in this efficiency are primarily In addition to explaining the differences between K fc and
due to changes in the combustion phasing. The center Kc , the trends in K wc seen in Fig. 24 also serve to
term on the RHS of Eq.(3) is a straight-forward illustrate the mechanisms by which changes in injection
representation of the fraction of the chemical heat timing and swirl affect K fc . For example, a general
release that is manifested as apparent heat releasea characteristic of Fig. 24 is a decrease in K wc for both
heat-loss efficiency Khl , and the last term on the RHS is early and late injection timings. In the former case, early
the combustion efficiency Kc . heat release results in negative work as the combustion-
induced pressure rise occurs before compression has
ended. Although some of this work is recovered during
the subsequent expansion, due to heat transfer and
Downloaded from SAE International by Indian Institute of Technology - Chennai, Monday, July 17, 2017

blow-by some of this work is inevitably lost. In the latter SUMMARY AND CONCLUSION
case, heat release is sufficiently retarded that much of
the volume expansion has already occurred. CO emission and fuel conversion efficiency are
examined in a highly-dilute, low-temperature diesel
A second observation is the low K wc in the upper-central combustion regime as the injection timing and swirl ratio
region of Fig. 24, which coincides with a minimum in the are varied. To clarify the dominant factors influencing the
combustion efficiency. As discussed in the section emission and efficiency behavior, the cylinder pressure,
describing the optimal swirl ratio, the cause of this engine-out emissions, and flow velocities are measured.
minimum in Kc is thought to be trapping of CO in the High speed video imaging and multi-dimensional numer-
bowl. Apparently, this trapping also retards the heat ical simulations of the combustion process are also
release thereby influencing K wc . performed.

Finally, K wc also exhibits a general decreasing trend as The CO emission behavior exhibits two interesting
the swirl level increases. As discussed elsewhere [28], trends:
the work conversion efficiency is influenced by heat
transfer losses that occur in a non-combusting cycle. First, an optimal swirl ratio exists at which the lowest CO
emission and best fuel conversion efficiency are
The higher heat transfer losses incurred in high swirl
observed. Heat release analysis, the results of the
flows result in a decreased K wc . Consequently, the
numerical simulations, and imaging of combustion
influence of heat loss is not fully-confined to the heat-
luminosity indicate that this behavior is likely due to
loss efficiency. mixing processes occurring after the premixed burn
period. At the lowest swirl ratios, a dual-vortex vertical
Although Q chem and Q hl have not been measured plane structure forms that enhances the mixing of
directly, the heat-loss efficiency K hl can be estimated partially-burned fuel and air. This structure is most
from the remaining efficiencies via Eq.(3). The results effective at a moderate swirl ratio (| 2.5). At higher swirl
must be employed cautiously, however, as the ratios, the flow structures formed impede mixing, and
combustion efficiencies calculated from Eq.(4) will likely can lead to trapping of partially-burned fuel within the
be overestimated, due to the neglect of the energy bowl.
content in the unburned hydrocarbons and other
unmeasured species. Provided that these species follow Second, CO emission generally exhibits a rapid
the same trends as the CO emissions, their neglect decrease from the maximum as SOI is advanced,
implies that the changes observed in Kc are particularly at the highest swirl ratios. The numerical
underestimated. simulations indicate that, at a fixed swirl ratio, earlier
injection timing leads to enhanced pre-combustion
Fig. 25 illustrates the resulting estimates of K hl . Note mixing and hence lower peak in-cylinder CO mass. The
that the overall variation in K hl is larger than is seen in enhanced mixing is due not only to increased ignition
either K wc or Kc . However, because the variation in Kc is delay, but also to increased mixing rates under high-
underestimated, the variation seen in K hl is likely swirl conditions. A reduction in CO emission with
excessive. Nevertheless, the magnitude of the true increased pre-mixing CO implies that CO emission
variation is likely to be significant, and the trends stems predominantly from under-mixed fuel (rich
observed are physically plausible and bear consideration. mixtures), a finding which is supported experimentally by
First, a general trend of decreasing heat-loss efficiency a strong tendency toward reduced CO emission with
is seen as swirl is increaseda result that is not increased ignition delay and with increased injection
pressure. Correlation of CO emission with peak
unexpected. More interestingly, the heat loss efficiency
adiabatic flame temperature suggests that temperature
shows a greater variation with injection timing. This
related changes to the early CO oxidation rates are not a
variation might be expected based on the simple
significant factor influencing peak in-cylinder CO mass
observation that there is more time available for heat and subsequent engine-out emission. Spray penetration
transfer at the earlier injection timing. At least three other and spray targeting, with concomitant liquid film
factors also contribute to this behavior, though. First, formation was also examined as a potential source of
with earlier combustion phasing the average surface modified premixing. Significant liquid film formation
area-to-volume ratio of the hot, in-cylinder gases is along the bowl lip was observed at early injection timings.
increased. Second, mixture that burns first and then is However, low CO emission was observed under these
compressed reaches a higher temperature than mixture conditions, signifying that wall-film formation in this
that burns after compression. Third, convective heat region is not a dominant factor influencing CO emission.
transfer coefficients are expected to be larger at the At the earliest injection timings, the likely existence of
higher peak cylinder pressures observed with early wall films on the top of the piston is identified as a
injection. Each of these factors is expected to increase potential source of increasing CO. Finally, the possibility
heat transfer losses as injection is advanced. that enhanced pre-combustion turbulence was
promoting mixing under early-injection conditions was
examined. Increased turbulence was observed with
Downloaded from SAE International by Indian Institute of Technology - Chennai, Monday, July 17, 2017

earlier injection timing, and likely assists in the initial 8. Adomeit, P., Pischinger, S., Becker, M., Rohs, H.,
mixing taking place in the ignition delay period. and Greis, A., Laser Optical Diagnostics and
Numerical Analysis of HSDI Combustion, THIESEL
The impact of swirl level and injection timing on the fuel 2004: Conference on Thermo- and Fluid Dynamic
conversion efficiency is also assessed. Qualitative Processes in Diesel Engines, Sept. 8-10, Valencia,
differences are observed between the fuel conversion Spain, 2004.
efficiency and the combustion efficiency derived from the 9. Kook, S., Bae, C., Miles, P. C., Choi, D, and Pickett,
CO emissions. These differences can be accounted for L. M., The Influence of Charge Dilution and
by changes in the work conversion efficiency Injection Timing on Low-Temperature Diesel
(associated primarily with combustion phasing) and by Combustion and Emissions, SAE Paper 2005-01-
heat transfer losses.
3837, 2005.
10. Timoney, D.L., Smoke and Fuel Consumption
ACKNOWLEDGMENTS Measurements in a Direct Injection Diesel Engine
with Variable Swirl, SAE Paper No. 851542, 1985.
Support for this research was provided by the U.S.
11. Van Gerpen, J.H., Hwang, C.-W., and Borman, G.L.,
Department of Energy, Office of FreedomCAR and
The Effects of Swirl and Injection Parameters on
Vehicle Technologies. The research was performed at
Diesel Combustion and Heat Transfer, SAE Paper
the Combustion Research Facility, Sandia National
Laboratories, Livermore, California. Sandia is a No. 850265, 1985.
multiprogram laboratory operated by Sandia Corporation, 12. Miles, P. C., The Influence of Swirl on HSDI Diesel
a Lockheed Martin Company, for the United States Combustion at Moderate Speed and Load, SAE
Department of Energys National Nuclear Security Paper 2000-01-1829, 2000.
Administration under contract DE-AC04-94AL85000. The 13. Miles, P.C., Choi, D., Pickett, L.M., Singh, I.P.,
BK21 and Future Vehicle Technology Development Henein, N., RempelEwert, B.H., Yun, H., and Reitz,
Corps. of Korea supported Sanghoon Kook's visiting R. D., Rate-Limiting Processes in Late-Injection,
research. The authors express their appreciation to Mark Low-Temperature Diesel Combustion Regimes,
Musculus and Lyle Pickett for providing the high speed THIESEL 2004: Conference on Thermo- and Fluid
camera and the Matlab source code to calculate the Dynamic Processes in Diesel Engines, Sept. 8-10,
adiabatic flame temperature. Valencia, Spain, 2004.
14. Amsden, A. A., KIVA-3V: A Block Structured KIVA
REFERENCES Program for Engines with Vertical or Canted Valves,
Los Alamos National Laboratory Report No. LA-
1. Cowland, C., Gutmann, P. and Herzog, P.L., 13313-MS, 1997.
Passenger Vehicle Diesel Engines for the US, SAE 15. RempelEwert, Bret H., The Influence of Swirl on
Paper No. 2004-01-1452, 2004. Flow, Fuel Injection, and Emissions in an HSDI
2. Kimura, S., Aoki, O., Kitahara, Y. and Aiyoshizawa Diesel Engine, Masters Thesis, Dept. of
E., Ultra-Clean Combustion Technology Combining Mechanical Engineering, University of Wisconsin
a Low-Temperature and Premixed Combustion Madison, 2004.
Concept for Meeting Future Emission Standards, 16. Miles, P., Choi, D., Megerle, M., RempelEwert, B.,
SAE Paper No. 2001-01-0200, 2001. Reitz, R. D., Lai, M. D., Sick, V., The Influence of
3. Akihama, K., Takatori, Y., Inagaki, K., Sasaki, S., Swirl Ratio on Turbulent Flow Structure in a Motored
and Dean, A.M., Mechanism of the Smokeless Rich HSDI Diesel Engine A Combined Experimental
Diesel Combustion by Reducing Temperature, SAE and Numerical Study, SAE Paper 2004-01-1678,
Paper No. 2001-01-0655, 2001. 2004.
4. Walter, B. and Gatellier, B. Development of the 17. Kong, S.-C., Han, Z., and Reitz, R.D., The
High Power NADI Concept Using Dual Mode Diesel Development and Application of a Diesel Ignition
Combustion to Achieve Zero NOx and Particulate and Combustion Model for Multidimensional Engine
Emissions, SAE Paper No. 2002-01-1744, 2002. Simulation, SAE Paper 950278, 1995.
5. Weibck, M., Csat, J., Glensvig, M., Sams, T. and 18. Ikegami, M., Role of Flows and Turbulent Mixing in
Herzog, P., Alternative Brennverfahren ein Ansatz Combustion and Pollutant Formation in Diesel
fr den Zuknftigen Pkw-Dieselmotor, MTZ: vol 64: Engines, COMODIA 90: Proceedings of the
pp 718-727, 2003. International Symposium on Diagnostics and
6. Hasegawa R. and Yanagihara, H., HCCI Modeling of Combustion in Internal Combustion
Combustion in DI Diesel Engine, SAE Paper No. Engines, 35 September, Kyoto, Japan, 1990.
2003-01-0745, 2003. 19. Khan, I.M., Wang, C.H.T., and Langridge, B.E.,
7. Sjberg, M., and Dec, J. E., An Investigation into Effect of Air Swirl on Smoke and Gaseous
the Lowest Acceptable Combustion Temperatures Emissions from Direct-Injection Diesel Engines,
for Hydrocarbon Fuels in HCCI Engines, SAE Paper No. 720102, 1972.
Proceedings of the Combustion Institute, Vol. 30, 20. Binder, K. and Hilburger, W., Influence of the
pp.2719-2726, 2005. Relative Motions of Air and Fuel Vapor on the
Downloaded from SAE International by Indian Institute of Technology - Chennai, Monday, July 17, 2017

Mixture Formation Processes of the Direct Injection


Diesel Engine, SAE Paper No. 810831, 1981.
21. Miles, P., Megerle, M., Sick, V., Richards, K., Nagel,
Z., and Reitz. R. D., The Evolution of Flow Struc-
tures and Turbulence in a Fired HSDI Diesel Engine,
SAE Paper 2001-01-3501, 2001.
22. Miles, P.C., RempelEwert, B.H., and Reitz, R.D.,
Squish-Swirl and Injection-Swirl Interaction in Direct
Injection Diesel Engines, ICE 2003: 6th International
Conference on Engines for Automobiles, Capri,
Naples, Italy, Sept. 14-19, 2003
23. Miles, P.C., In-Cylinder Turbulent Flow Structure in
Direct-Injection, Swirl-Supported Diesel Engines,
Ch. 1 in Flow and Comb. in Automotive Engines, C.
Arcoumanis, ed. Springer-Verlag, in press.
24. Heywood, J. B., Internal Combustion Engine
Fundamentals, International Edition, McGraw-Hill,
Inc., 1988.
25. Akagawa, H., Miyamoto, T., Harada, A., Sasaki, S.,
Shimazaki, N., Hashizume, T., and Tsujimura, K.,
Approaches to Solve Problems of the Premixed
Lean Diesel Combustion, SAE Paper 1999-01-0183,
1999.
26. Henein, N. Analysis of Polllutant Formation and
Control and Fuel Economy in Diesel Engines, Prog.
Energy Combust. Sci., Vol. 1, pp.165-207, 1978.
27. Miles, P., Megerle, M., Hammer, J., Nagel, Z., Reitz,
R. D., Sick, V., Late-Cycle Turbulence Generation
in Swirl-Supported, Direct-Injection Diesel Engines,
SAE Paper 2002-01-0891, 2002.
28. Choi, D. and Miles, P.C., A Parametric Study of
Low-Temperature, Late-Injection Combustion in an
HSDI Diesel Engine, COMODIA 2004: 6th Intl.
Symp. On Diagnostics and Modeling of Combustion
in IC Engines, Aug 2-5, Yokohama, 2004.

NOMENCLATURE

ATDC After Top Dead Center

CAD Crank Angle Degrees

EGR Exhaust Gas Recirculation

EVC Exhaust Valve Closing

EVO Exhaust Valve Opening

IMEP Indicated Mean Effective Pressure

IVC Intake Valve Closure

IVO Intake Valve Opening

SOI Start of Injection

Sp Mean Piston Speed


Downloaded from SAE International by Indian Institute of Technology - Chennai, Monday, July 17, 2017

Rs = 2.59, SOI = -22.25 CAD ATDC


0 0
APPENDIX: COMPARISON OF MEASURED AND
-5 -5
SIMULATED TANGENTIAL VELOCITIES
-10 -10

Figures A1 ~ A5 compare the evolution of the measured


-15 -15
mean tangential velocities with the numerical predictions Experiment
pe
Computation
m
for Rs = 1.44 ~ 7.12. The velocity data are normalized -20 -25 CAD -20 -5 CAD
with the piston speed Sp: 4.25 m/s at 1500 rpm. At each 0 0

crank angle in the figures, axial profiles of the tangential -5 -5


mean velocities are depicted on a schematic showing
the relative position of the piston. The velocities were -10 -10

measured at a fixed radial location of r = 13.6 mm, which -15 -15


is represented by the vertical, dashed lines. The crank
angle range investigated commences well before -20 -20 TDC

z [mm]
-20 CAD

z [mm]
0 0
injection occurs at -22 CAD ATDC and continues to the
end of the premixed combustion at around 10 CAD -5 -5
ATDC.
-10 -10

The overall development of the measured and predicted -15 -15


mean velocity fields is very similar, though several
-20 -15 CAD -20 5 CAD
discrepancies in the detailed shape of the profiles exist.
0 0
For example, the measured velocities are lower than the
computation at many locations lower in the bowl. These -5 -5
data are currently being re-assessed for noise
-10 -10
contributions arising from scattering from the bowl floor.
-15 -15

A second discrepancy is under-predicted velocities at -


-20 -10 CAD -20 10 CAD
15 CAD ATDC for Rs = 3.77 ~ 7.12. At this crank angle, Measurement
e Location Measurement
e Location
0 2 4 6 8 10 12 0 2 4 6 8 10 12
the calculations show a marked influence of the fuel
Tangential Velocity [Sp] Tangential Velocity [Sp]
injection event (SOI=-22.25 CAD ATDC) near z = -6 mm,
resulting in a minimum in the tangential velocities. This
Figure A2 Comparison of the axial profile of the
measured mean tangential velocities with the results of
Rs = 1.44, SOI = -22.25 CAD ATDC the numerical simulation. Rs = 2.59.
0 0
Rs = 3.77, SOI = -22.25 CAD ATDC
-5 -5 0 0

-10 -10 -5 -5

-15 -15 -10 -10


Experiment
pe
Computation
m
-20 -20 -15 -15
-25 CAD -5 CAD Experiment
pe
0 0 Computation
m
-20 -25 CAD -20 -5 CAD
-5 -5 0 0

-10 -10 -5 -5

-15 -15 -10 -10

-20 -20 -15 -15


TDC
z [mm]

-20 CAD
z [mm]

0 0
-20 -20 TDC
z [mm]

-20 CAD
z [mm]

-5 0 0
-5

-10 -10 -5 -5

-15 -10 -10


-15

-20 -20 -15 -15


-15 CAD 5 CAD
0 0
-20 -15 CAD -20 5 CAD
0 0
-5 -5

-5 -5
-10 -10

-10 -10
-15 -15

-15 -15
-20 -10 CAD -20 10 CAD
Measurement
e Location Measurement
e Location
0 2 4 6 8 10 12 0 2 4 6 8 10 12 -20 -20
-10 CAD 10 CAD
Measurement
e Location Measurement
e Location
Tangential Velocity [Sp] Tangential Velocity [Sp]
0 2 4 6 8 10 12 0 2 4 6 8 10 12

Figure A1 Comparison of the axial profile of the Tangential Velocity [Sp] Tangential Velocity [Sp]
measured mean tangential velocities with the results of
the numerical simulation Rs = 1 44 Figure A3 Comparison of the axial profile of the
measured mean tangential velocities with the results of
the numerical simulation. Rs = 3.77.
Downloaded from SAE International by Indian Institute of Technology - Chennai, Monday, July 17, 2017

Rs = 4.94, SOI = -22.25 CAD ATDC Rs = 7.12, SOI = -22.25 CAD ATDC
0 0 0 0

-5 -5 -5 -5

-10 -10 -10 -10

-15 -15 -15 -15


Experiment
pe Experiment
pe
Computation
m Computation
m
-20 -20 -20 -20
-25 CAD -5 CAD -25 CAD -5 CAD
0 0 0 0

-5 -5 -5 -5

-10 -10 -10 -10

-15 -15 -15 -15

-20 -20 TDC -20 -20 TDC


z [mm]

-20 CAD

z [mm]
-20 CAD
z [mm]

z [mm]
0 0 0 0

-5 -5 -5 -5

-10 -10 -10 -10

-15 -15 -15 -15

-20 -20 -20 -20


-15 CAD 5 CAD -15 CAD 5 CAD
0 0 0 0

-5 -5 -5 -5

-10 -10 -10 -10

-15 -15 -15 -15

-20 -20 -20 -20


Measurement
e Location
-10 CAD 10 CAD Measurement
e Location
-10 CAD 10 CAD
Measurement
e Location Measurement
e Location

0 2 4 6 8 10 12 0 2 4 6 8 10 12 0 2 4 6 8 10 12 0 2 4 6 8 10 12
Tangential Velocity [Sp] Tangential Velocity [Sp] Tangential Velocity [Sp] Tangential Velocity [Sp]

Figure A4 Comparison of the axial profile of the Figure A5 Comparison of the axial profile of the
measured mean tangential velocities with the results of measured mean tangential velocities with the results of
the numerical simulation. Rs = 4.94 the numerical simulation. Rs = 7.12.

mm, resulting in a minimum in the tangential velocities. Despite the discrepancies mentioned above, the
This behavior is not seen in the measurements. Finally, important point to note is that, overall, the measured
the overall trends in the axial velocity profile prior to velocities and their evolution correlate well with the
injection differs for Rs = 7.12the measured tangential numerical simulations.
velocity has a maximum near z = -12 mm, while the
simulation predicts a maximum near the cylinder head. CONTACT
This behavior was repeatable. Additional experiment
and analysis are needed to clarify this discrepancy. Sanghoon Kook, kooks@kaist.edu

S-ar putea să vă placă și