You are on page 1of 16

Downloaded from SAE International by University of Minnesota, Tuesday, July 31, 2018

SAE TECHNICAL
PAPER SERIES 2004-01-2973

An Experimental Study on the Effect of Intake


Primary Runner Blockages on Combustion
and Emissions in SI Engines under
Part-Load Conditions
A. Selamet, S. Rupal and Y. He
The Ohio State University

P. S. Keller
DaimlerChrysler Corporation

Reprinted From: SI Engine Experiment and Modeling


(SP-1901)

Powertrain & Fluid Systems


Conference and Exhibition
Tampa, Florida USA
October 25-28, 2004

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 University of Minnesota, Tuesday, July 31, 2018

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
Fax: 724-772-4891
Tel: 724-772-4028

For multiple print copies contact:

SAE Customer Service


Tel: 877-606-7323 (inside USA and Canada)
Tel: 724-776-4970 (outside USA)
Fax: 724-776-1615
Email: CustomerService@sae.org

ISBN 0-7680-1523-5
Copyright © 2004 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 of a proposed manuscript to: Secretary, Engineering Meetings Board, SAE.

Printed in USA
Downloaded from SAE International by University of Minnesota, Tuesday, July 31, 2018

2004-01-2973

An Experimental Study on the Effect of Intake Primary


Runner Blockages on Combustion and Emissions
in SI Engines under Part-Load Conditions
A. Selamet, S. Rupal and Y. He
The Ohio State University

P. S. Keller
DaimlerChrysler Corporation

Copyright © 2004 SAE International

ABSTRACT relationship, ST ~ u’, that renders the turbulent scale


insignificant under normal operating conditions, has been
Charge motion is known to accelerate and stabilize combustion supported by combustion bomb and engine experiments for the
through its influence on turbulence intensity and flame bulk of the burn process (Figs. 34 and 35 in Tabaczynski [3],
propagation. The present work investigates the effect of charge Figs. 15 and 16 in Lancaster et al. [4], and Fig. 9-30 in
motion generated by intake runner blockages on combustion Heywood [5].) As the turbulence intensity diminishes, the
characteristics and emissions under part-load conditions in SI turbulent flame speed approaches the kinetically controlled
engines. Firing experiments have been conducted on a laminar flame speed SL. At large values of u’, ST may deviate
DaimlerChrysler (DC) 2.4L 4-valve I4 engine, with spark from the linear relationship (due presumably to extinction [3],
range extending around the Maximum Brake Torque (MBT) Fig. 47 in [2], Figs. 15 and 16 in Khalighi et al. [6].)
timing. Three blockages with 20% open area are compared to Regardless of the exact nature of relationship between the two,
the fully open baseline case under two operating conditions: the strong influence of u’ in increasing ST and therefore
2.41 bar brake mean effective pressure (bmep) at 1600 rpm, enhancing the burn rate has now been well established.
and 0.78 bar bmep at 1200 rpm. The blocked areas are shaped
to create different levels of swirl, tumble, and cross-tumble. The foregoing enhancement is particularly needed in engines
Crank-angle resolved pressures have been acquired, including operating under part-load conditions where the throttling
cylinders 1 and 4, intake runners 1 and 4 upstream and reduces the mass flow rate as well as the intake charge density,
downstream of the blockage, and exhaust runners 1 and 4. which in turn reduces the in-cylinder peak pressures.
Exhaust runner temperature and emissions, including NOx, Furthermore, the throttling increases the burned gas residual
HC, CO, CO2, and O2 are also measured. Combustion is fraction within the cylinder. If the engine is operated in a lean
quantified by calculating Coefficient of Variation (COV), burn mode or with Exhaust Gas Recirculation (EGR), these
Lowest Normalized Value (LNV), and burn duration. The diluents (excess air or burned products) will also decrease the
impact of blockages on the increase of exhaust gas temperature peak flame temperature. All of these factors reduce the laminar
by means of spark retardation is also analyzed. flame speed predominantly through the Boltzmann factor,
hence the burn rate. Thus, the mechanisms to counter the
INTRODUCTION reduction in flame speed by increasing the turbulence intensity
in fresh mixture near spark are highly desirable. Increased
A number of approaches have been proposed in literature to burn rate makes the combustion less sensitive to cyclic
represent the turbulent flame speed ST, which defines the burn variations in pressure, temperature and mixture composition,
rate when combined with the unburned mixture density and leading to improved stability [5, Chap. 9]. In general, a
flame area. One of the concepts dates back to Schelkin [1] considerable amount of turbulence is generated during the
who considered the distortion (or wrinkling) of the flame intake process by the incoming jet, with integral length scales
surface by turbulent eddies larger than the flame thickness on the order of the valve lift. Tabaczynski [3] provided useful
(see, for example, Glassman [2]). Though somewhat heuristic, estimates of both the turbulent dissipation rate and the eddy
the approach captures through effective area increase, the decay time. Estimated at about 1 ms, the latter is typically an
functional relationship between the turbulent flame speed and order of magnitude smaller than engine time scales. Thus, the
turbulence intensity u’, the former varying nearly linearly with anisotropic intake process decays rapidly by turbulent
the latter at highly turbulent flows. This near-linear dissipation (see also Lumley [7]). Aside from shearing in the
1
Downloaded from SAE International by University of Minnesota, Tuesday, July 31, 2018

jet, a superimposed organized intake charge motion, on the known to cut down the HC emissions under normal operating
other hand, stores some of the initial momentum through its conditions and increase the port and runner oxidation due to
mean rotation and breaks down into small scale turbulent higher exhaust temperatures. Improved combustion stability
eddies during compression. due to charge motion combined with lean AFR is also useful to
reduce cold start HC, since typically the engine needs to be
Swirl and tumble are the two types of charge motion that have started with an enriched mixture. Thus, some of the recent
been studied extensively. The generation mechanisms have work has concentrated on the effect of aggressive spark
included, for example, port designs [8,9], valve shrouding retardation on HC emissions, with charge motion retaining the
[6,10] and masking [3], chamber shrouding [9], various valve combustion stability [18-22]. Reference [22] provides a useful
lift [7] and dual port strategies [11], charge motion control summary of additional applications of charge motion that
valves (CMCV) [10,12], and their immobile counterparts: flow improve the combustion stability at cold start with lean
blockages [13,14] and sleeves [15]. One of the earlier works mixture and retarded spark.
with firing engines by Kent et al. [16] involved the
relationship between swirl and burn rate with different intake The objective of the current work is to gain a comprehensive
ports and valve lifts. With a single-cylinder engine running understanding of the effect of intake runner blockages through
part load at 1500 rpm, Hadded and Denbratt [9] considered the firing experiments under part-load conditions and
effect of tumble on burn duration and combustion stability stoichiometric mixture strength. A spark sweep is performed
with varying EGR and Air/Fuel Ratio (AFR). They provided for unblocked as well blocked runners to establish the
not only the measurements of mean velocity and turbulence minimum of Brake Specific Fuel Consumption (BSFC), which
intensity, but also the estimates of integral length scales in low corresponds to the MBT timing. The impact of charge motion
and high tumble-port cases. Endres et al. [17] reported the on combustion parameters, fuel consumption, and emissions is
impact of tumble and swirl towards improving part-load then studied in detail. Intake (and exhaust) runner pressures
operation and extending lean burn limit and EGR tolerance. are also analyzed to understand the flow dynamics.
Urushihara et al. [10] illustrated the effect of tumble on the
enhancement of air-fuel mixing. Floch et al. [13] examined the EXPERIMENTAL SETUP
impact of increased turbulence intensity caused by swirl and
tumble generated through flow blockages on COV and burn Experiments have been performed on a DC 2.4L 4-valve I4
duration using propane in a single-cylinder engine under a port-fuel-injection engine with pentroof combustion chamber
part-load condition at 2000 rpm. Urushihara et al. [12] (Table 1). Figure 1 shows the main components of the engine
discussed a similar work on an SI engine running on set up and the position of different transducers. Gasoline with
homogeneous charge. In an I4 engine, Kang et al. [8] reported H/C = 1.842 (thus stoichiometric AFR = 14.56) is used as the
the effect of tumble on lean burn characteristics, in terms of fuel. Emission analysis is carried out using Horiba’s Mexa
combustion stability and burn duration. The intake port design 7100 analyzer with gas samples from exhaust runners at
with stronger tumble led to a more stable combustion locations 1” away from the cylinder head face. Chrysler
combined with the reduced burn delay. Arcoumanis et al. [15] Motors Instrumentation’s Interrogator Monitor is used to
conducted experiments at MBT spark timing under idling and manually control the fuel injection for stoichiometric AFR and
part-load operating conditions on a single cylinder engine with the spark timing. Exhaust runner temperature is measured by
an intake port sleeve that generates tumble, to understand the Omega Type K thermocouples placed 1.5” downstream of the
effects on combustion and emissions such as NOx and HC head face.
under lean and stoichiometric equivalence ratios. They also
provided a comparison of flame images with and without
Table 1. Engine specifications.
sleeves. Many of the foregoing engine studies have used 4-
valve pentroof combustion chambers, and also incorporated Bore (mm) 87.5
steady flow measurements and optical diagnostic techniques
Stroke (mm) 101
such as Laser Doppler Velocimetry [8-10,12,13,15] to quantify
the flow field within the cylinder. Pertinent literature on other Connecting rod length (mm) 151
diagnostic techniques, including Particle Tracking Velocimetry Compression ratio 9.47
and Molecular Tagging Velocimetry may be found in Schock 3
et al. [14]. Insightful flame growth images obtained by high- Clearance volume (cm ) 71.43
speed CCD cameras have also been reported [8,12,15]. Firing Order 1-3-4-2

Intentional spark retard from MBT results in an increase in the In-cylinder pressure is measured using Kistler type 6125B
exhaust gas temperature, due to lower work extraction during piezoelectric transducers (250 bar) for 256 firing cycles. The
expansion, promotes an early catalyst light-off, and therefore blockages (Fig. 2) of 3 mm thickness are introduced 0.875”
reduces cold start HC emissions. The retarded spark increases upstream of the cylinder head face. Intake pressure upstream
cycle-to-cycle variations since the combustion energy is and downstream of the blockage is acquired using Kistler type
released into an expanding volume as opposed to an 4045A2 piezoresistive transducers (2 bar) located 0.375”
approximately constant volume with slowly moving piston upstream and downstream of the blockage and averaged for 64
near TDC at the MBT timing. This increased cyclic variation cycles. Same type of piezoresistive transducers are used in the
can be countered by the charge motion. Spark retard is also
2
Downloaded from SAE International by University of Minnesota, Tuesday, July 31, 2018

Dyno

11
9
13 6 5 Blockage
15
16 Cylinders

17 c4 2

c3
Exhaust Muffler Intake
ACB
Outlet
c2

Catalyst
c1 1

Exhaust Intake
manifold manifold
14
12 4 3
10

8 7

c1-c4: cylinders; 1, 2: in-cylinder pressure transducers; 3-6: intake runner pressure transducers; 7: theta sensor; 8: TDC sensor; 9, 10: exhaust
pressure transducers; 11, 12: exhaust thermocouples; 13, 14: emission taps; 15: ECU O2 sensor; 16: monitored O2 sensor; 17: catalyst thermocouple.
Figure 1. Engine setup and location of transducers

exhaust runners to measure pressure 1” downstream of the where W = v³ PdV is the work done per cylinder/cycle over
cylinder head face. The resolution of pressure traces is 1 Crank compression and expansion strokes, Vd is the displacement
Angle Degree (CAD) dictated by the optical encoder volume, and Nc is the number of cylinders. Being a measure of
connected to the crankshaft. All measurements have been combustion stability, COV in IMEP is then determined by
made in cylinders and runners 1 and 4. In-cylinder
piezoelectric transducers output charge proportional to
V imep
pressure differences and it is necessary to reference (peg) the COVimep = u 100% , (2)
relative pressure magnitudes to an absolute value known at IMEP
some CAD. In the current experiments, 3 CAD average of the
intake runner pressure downstream of the blockage around the where IMEP and V imep are the mean and the standard
Intake Bottom Dead Center (IBDC) is used as the reference.
deviation for 256 engine cycles, respectively.

Another useful parameter to quantify the combustion stability


is the LNV in IMEP defined as

IMEPlowest
LNV u100% , (3)
(a) tumble (b) swirl (c) swumble IMEP
Figure 2. Schematic representation of intake runner blockages.
where IMEPlowest is the lowest IMEP in 256 cycles. Lower
LNV would suggest less stable combustion.

DATA ANALYSIS Burn duration is the period starting from ignition until End of
Combustion (EOC). The first step in the calculation of burn
COV in indicated mean effective pressure (IMEP), LNV, and duration is determining the thermodynamic property n,
burn duration are determined from in-cylinder pressure to assuming polytropic compression and expansion, given by
characterize combustion. Knowing in-cylinder pressure (P)
and volume (V), IMEP is defined as PV n Const . ,

W Nc
IMEP , (1)
Vd
3
Downloaded from SAE International by University of Minnesota, Tuesday, July 31, 2018

A linear regression fit to N number of CAD (index j) on a log transfer during expansion. The heat losses become significant,
P-log V plot gives [23], particularly under operating conditions that lead to slow
combustion such as light loads and retarded ignition.
N N N
N ¦ log V j log Pj  ¦ log V j ¦ log Pj
j 1 j 1 j 1 RESULTS AND DISCUSSION
n N N
, (4)
N ¦ (log V j ) 2  ( ¦ log V j ) 2
j 1 j 1 Experiments have been conducted at two part-load operating
conditions: 2.41 bar (35 psi) bmep at 1600 rpm engine speed
The intervals for compression and expansion have to be (DC World-wide Point = WP) and 0.78 bar bmep (11 ft-lb
selected to account for the deviation of the P-V diagram from a brake torque) at 1200 rpm (DC high speed Idle Point = IP).
straight line due to combustion and flow through the valves. Comparisons are provided between three 20% open area
The definition of the polytropic exponent for compression nc is blockages- tumble, swirl, and cross-tumble (or, equivalently,
based on a CAD range starting at 40° before spark timing and swumble, which is a combination of tumble and swirl) vs. fully
ending at the ignition point. In order to determine the interval open (baseline) case (Fig. 2). Two sets of data have been
for polytropic expansion ne, an iterative method of slope included to illustrate the repeatability. The MBT timing that
comparison is followed to estimate an approximate value for corresponds to minimum BSFC is identified for each load-
EOC (AEOC). n is calculated over a short interval, initially blockage combination and experiments are run sweeping
assumed between Top Dead Center (TDC) and 7° after TDC. through a spark range from retarded to advanced ignition
This 7° interval is shifted by one CAD after TDC and the (Table 2). The variation of BSFC with Spark Advance (SA) is
value of n is compared to that of nc until the difference falls shown in Fig. 3 with the corresponding MBT timings being
below 0.1, when the CAD corresponding to the starting of the circled (same identification is also used in the following
interval for n calculation is assumed as AEOC. Knowing figures). The minimum in the BSFC variation is well
AEOC, the polytropic exponent for expansion, ne is calculated established for WP, whereas the curve lacks the distinct U-
over an interval extending from AEOC until 10° before shape for the lighter load at IP presenting therefore a degree of
Exhaust Valve Opening (EVO). For the combustion period challenge to the determination of MBT. The dramatic effect of
between compression and expansion, a weighted average of nc the port blockages on combustion is reflected by the shift in
and ne is calculated to yield n as a function of CAD. MBT spark timing, with tumble producing the maximum effect
(SA of 14°, 24°, and 22° for tumble, swirl, and swumble,
respectively vs. 34° for fully open at WP and 22°, 32°, and 30°
Knowing n, EOC is determined from the combustion pressure
for tumble, swirl, and swumble, respectively vs. 40° for fully
rise method. The total pressure rise per CAD is considered to
open at IP). The reduced SA (particularly with tumble) is
be the sum of the increase due to change in the volume
helpful in retaining combustion stability at retarded ignition.
produced by the piston motion and the increment due to
The introduction of the blockage increases fuel consumption at
combustion at constant volume [24,25]. The increase in
the WP over the entire spark range. The BSFC remains
pressure due to combustion is computed as:
comparable near MBT at the IP between fully open and
blocked runners. These trends in BSFC will be elaborated
n
ª § V j 1 · º Vj upon later in connection with the indicated in-cylinder
'Pj « Pj  ¨ P
¸ j 1 » V . (5) quantities, including pumping loop observations.
«¬ © Vj ¹ »¼ spark
For combustion parameters, results have been presented for 5
spark timings chosen such that 3 of them cover the MBT
Combustion is assumed to end when the sum of the ratio of timing, while the remaining 2 are representative of the most
instantaneous to total combustion pressure rise for 3 retarded and advanced spark timings. Figure 4 compares
successive CAD is less than 0.005. Mass Fraction Burned COVimep between fully open and blocked runner cases. The
(MFB) is determined by assuming proportionality to the effect of charge motion on COV is not dramatic under either
fractional pressure rise due to the combustion from the time of operating condition at MBT. In view of the fact that the
spark based on a modified form of Rassweiler-Withrow stoichiometric mixture strength has been maintained
method given by: throughout the present study, this observation is consistent
with that of Arcoumanis et al. [15], who also showed
i
significant COV variation with leaner mixtures. Here, a larger
¦ 'P impact of restriction on COV is observed in the most retarded
spark spark timings for IP, where the introduction of tumble results
MFB (CAD) EOC
. (6)
in COV reduction of about 5%, swirl 4% and swumble 3%
¦
spark
'P (based on SA of 20°, 14°, 24°, and 18° for fully open, tumble,
swirl, and swumble, respectively), with important implications
for cold start HC reduction. Figures 5a-5d show the in-cylinder
Burn duration calculation of the present work is known to be pressure for 256 consecutive cycles corresponding to fully
reliable under a wide variety of operating conditions. The open, tumble, swirl, and swumble cases at MBT of IP. The
objective of the iterative slope comparison approach is to make reference (zero) for CAD is chosen as TDC of compression.
a realistic prediction of AEOC that accounts for the heat Note the reduction in the spread of peak pressure over the
4
Downloaded from SAE International by University of Minnesota, Tuesday, July 31, 2018

individual cycles due to combustion variation with the 3.5


blockages relative to the fully open case, although the a) WP
corresponding COVs at MBT are not substantially different. 3

The LNV in IMEP for fully open and blocked runner cases is
2.5
compared in Fig. 6. The introduction of charge motion

COV (%)
increases LNV at MBT for both operating conditions, most 2
noticeably with tumble at IP. Thus, LNV may be used as an
additional indicator of combustion stability improvements with 1.5
blockages, particularly for stoichiometric mixtures. Although it
1
is not the case here as evidenced from individual cycle
pressures in Fig. 5, the LNV approach may, in general, be 0.5
susceptible to low pressures of an isolated cycle. 4 8 12 16 20 24 28 32 36 40 44 48
400 SA
a) WP
390 10
b) IP
BSFC (g/kW-hr)

380
8
370

360 6

COV (%)
350
4
340
2
330
4 8 12 16 20 24 28 32 36 40 44 48
SA 0
12 16 20 24 28 32 36 40 44 48
Set 1 Set 2
Fully open: SA
Tumble: Figure 4. Effect of blockages on COV in imep (cylinder #1)
[The legend is same as Fig. 3.]
Swirl:
Swumble:
780 The variation of 0-100% burn duration with SA is depicted in
b) IP Fig. 7. The strong reduction of burn duration with tumble
blockage is clear at MBT, with the decrease being about 24° at
740
WP and 26° at IP. Swirl does not appear to reduce the burn
BSFC (g/kW-hr)

duration at WP, in contrast to a decrease of about 8° at IP.


700 Swumble shows a decrease of 6° at WP and 14° at IP. The total
burn duration is split next into conventional 0-10% (burn
delay) and 10-90% durations. The impact of blockages on the
660
0-10% burn duration is shown in Fig. 8, with the trends often
resembling those of 0-100% duration. The reduction with
620 tumble blockage at MBT reaches 14° for WP and 15° for IP
12 16 20 24 28 32 36 40 44 48 compared to the fully open runners. Swirl and swumble also
SA show a decrease of 5° and 8° for WP, respectively, and 9° for
Figure 3. Effects of blockages on BSFC. IP in the 0-10% burn duration. Figure 9 compares the variation
of 10-90% burn duration between the fully open and partially
blocked runners. By comparison with the 0-10% burn
Table 2. MBT spark timing. duration, charge motion does not exert strong influence on the
10-90% burn duration, with the exception of tumble showing a
Operating Spark range MBT point decrease of 7° compared to the fully open case at both
Blockage operating conditions. Such disparity observed here associated
condition (BTDC) (CAD)
with the effect of tumble on burn delay and duration is
fully open 26-46 34 consistent with, for example, Hadded and Denbratt [8]. MFB
tumble 6-26 14 variation at MBT is compared in Fig. 10 with the origin for
WP swirl 14-34 24 time chosen as the ignition point for all configurations. This
swumble 10-30 22 selection of origin for Fig. 10 is the only exception among the
fully open 20-40 40 results presented here using TDC of compression as the
tumble 14-34 22 reference for CAD. For both operating conditions, blockages
IP swirl 24-44 32 exhibit a steeper rise in MFB over the fully open case
swumble 18-38 30 consistent

5
Downloaded from SAE International by University of Minnesota, Tuesday, July 31, 2018

14 100
a) fully open a) WP
12
98

10
96
Pressure (bar)

LNV (%)
8
94
6
92
4

90
2

88
0
-90 -60 -30 0 30 60 90 4 8 12 16 20 24 28 32 36 40 44 48
CAD SA
100
b) IP
14

b) tumble 90
12

80

LNV (%)
10
Pressure (bar)

8
70
6

60
4

2 50
12 16 20 24 28 32 36 40 44 48
0 SA
-90 -60 -30 0 30 60 90
CAD Figure 6. Effect of blockages on LNV in IMEP (Cylinder #1)
[The legend is same as Fig. 3.]
14
c) swirl
12 80
a) WP
0-100% Burn duration (CAD)

10
70
Pressure (bar)

60
6

4 50

2
40
0
-90 -60 -30 0 30 60 90
CAD 30
4 8 12 16 20 24 28 32 36 40 44 48
SA
14

d) swumble 105
b) IP
0-100% Burn duration (CAD)

12
95
10

85
Pressure (bar)

75
6

4
65

2 55

0
-90 -60 -30 0 30 60 90 45
CAD 12 16 20 24 28 32 36 40 44 48
SA
Figure 5. Cyclic variation of in-cylinder pressure at MBT of IP.
Figure 7. Effect of blockages on 0-100% burn duration (cylinder #1)
[The legend is same as Fig. 3.]

6
Downloaded from SAE International by University of Minnesota, Tuesday, July 31, 2018

consistent with the preceding figures, illustrating faster burn, 35

10-90% Burn duration (CAD)


with tumble producing the largest effect. The corresponding a) WP
30
burn delay (0-10% MFB), flame propagation (10-90% MFB),
and total burn duration (0-100% MFB) periods at MBT are 25
listed in Table 3. The burn delay period is thus affected by all
blockages, whereas only tumble appears to reduce the flame 20
propagation period noticeably. Tumble, being a vertically
15
oriented vortex with its axis transverse to the cylinder, starts
breaking up during the early compression due to piston 10
motion, resulting in higher turbulence intensity during ignition,
thereby initiating stable flame growth and accelerating the 5

burn rate. The turbulence intensity combined with the mean 4 8 12 16 20 24 28 32 36 40 44 48


SA
motion results in tumble’s strong influence on the reduction of
the burn delay, consistent with other engine experiments 40

10-90% Burn duration (CAD)


[11,12; particularly the latter, which provided the details of in- b) IP
cylinder velocity field and turbulence late in the compression 35
stroke]. As the breakdown of the tumble vortex progresses
30
through the combustion period, most of the mean momentum
is converted into small scale turbulent eddies and hence, 10- 25
90% burn duration seems to be dictated by the turbulence
intensity alone. Swirl, on the other hand, maintains its 20

momentum until late compression and hence may not be as 15


effective in reducing the burn duration. The differences in
local mixture motion due to tumble and swirl near spark plug 10
will also influence the flame kernel convection. These intrinsic 12 16 20 24 28 32 36 40 44 48
mechanisms can therefore explain the relative trends observed SA
in Figs. 7-10. Figure 9. Effect of blockages on 10-90% burn duration (cylinder #1)
[The legend is same as Fig. 3.]
35
0-10% Burn duration (CAD)

a) WP 1

30 a) WP
0.9

0.8
25
0.7
Mass Fraction Burned

20
0.6

15 0.5

0.4
10
0.3

5 0.2
4 8 12 16 20 24 28 32 36 40 44 48
0.1
SA
0
0 10 20 30 40 50 60
40 CAD
0-10% Burn duration (CAD)

b) IP Ignition

35 fully open tumble swirl swumble


1

30 b) IP
0.9

0.8
25
0.7
Mass Fraction Burned

20 0.6

0.5

15 0.4
12 16 20 24 28 32 36 40 44 48
0.3
SA
0.2
Figure 8. Effect of blockages on 0-10% burn duration (cylinder #1)
[The legend is same as Fig. 3.] 0.1

0
0 10 20 30 40 50 60 70 80 90
CAD
Ignition

Figure 10. Effect of blockages on MFB at MBT (cylinder #1).


7
Downloaded from SAE International by University of Minnesota, Tuesday, July 31, 2018

Table 3. Comparison of MFB at MBT (set 1). 0.55


a) fully open vs. tumble at WP
Operating 0-10% MFB 10-90% MFB 0-100% MFB 0.5
Blockage
condition (CAD) (CAD) (CAD)
0.45

Pressure (bar)
fully open 24 22 60
tumble 10 15 36
WP 0.4
swirl 19 26 59
swumble 16 20 54 fully open: upstream
0.35
fully open 35 30 88 fully open: downstream
tumble 20 23 62 0.3 tumble: upstream
IP
swirl 26 32 80 tumble: downstream
Swumble 26 29 74 0.25
0 90 180 270 360 450 540 630 720

Intake runner pressure upstream and downstream of tumble CAD


IVO IVC
and swirl blockages are compared with the fully open case at 0.55
MBT for WP in Figs.11a and 11b, respectively, and for IP in b) fully open vs. swirl at WP
Figs. 11c and 11d. The blockage introduces changes in both 0.5
magnitude and phasing of the upstream and downstream
runner pressures. The pressure drop due to flow separation 0.45

Pressure (bar)
across the blockage is reflected in the downstream runner
0.4
pressure, showing a maximum difference from upstream
pressure midway through the intake stroke. The increase of the
0.35
downstream pressure close to TDC of intake with the fully open: upstream
fully open: downstream
introduction of the blockage can be attributed to the backflow 0.3 swirl: upstream
of exhaust gases into the intake port. The mean of fully-open
swirl: downstream
runner pressures is observed to be lower than those of the 0.25
blocked runner particularly upstream of the blockage location. 0 90 180 270 360 450 540 630 720
This trend is more visible at WP with significantly higher flow CAD
rates than IP. The mean pressure difference may be attributed IVO IVC
to the changes in the throttle position, with the larger throttle 0.4
fully open: upstream
opening for the blocked runners leading to higher mean runner c) fully open vs. tumble at IP
fully open: downstream
pressures on the manifold side of blockage location. Measured tumble: upstream
0.35
mean intake manifold pressures depicted in Fig. 12 support tumble: downstream
Pressure (bar)

this argument further, by revealing higher pressures for


blocked runners relative to the fully-open case. 0.3

By providing the relative indicated (gross and net) quantities,


the in-cylinder P-V diagrams help explain the impact of 0.25
blockages further. Figure 13a is a log P-log V plot at the IP for
illustrative purposes, with the pumping loop expanded in Fig.
13b. The pumping mean effective pressure (PMEP) along with 0.2
0 90 180 270 360 450 540 630 720
its ratio to IMEP determined from such P-V diagrams is
CAD
compared in Table 4 among four configurations at both
operating conditions. Note that PMEP (hence pumping work) IVO IVC
0.4
has always increased with blockages, more so at IP (lighter d) fully open vs. swirl at IP fully open: upstream
load) than WP. The PMEP increase with blockages at WP fully open: downstream
correlates well with the BSFC trends in Fig. 3a, while such 0.35
swirl: upstream
correlation does not exist at IP, suggesting that the other swirl: downstream
Pressure (bar)

factors such as in-cylinder heat transfer may play a larger


relative role for the latter. Also included in the table is the 0.3
indicated specific fuel consumption (ISFC), whose reciprocal
represents the fuel conversion efficiency within lower heating
value of the fuel. At both operating points, ISFC improves 0.25

with tumble (indicating higher fuel conversion efficiency) and


increases with swirl. Swumble improves ISFC at WP, while
0.2
manifesting no change at IP. The results of an approximate 0 90 180 270 360 450 540 630 720
analysis of in-cylinder heat transfer based on Gatowski et al. CAD
[26] are also included in Table 4. The net heat release ratio Ș
IVO IVC
designates the sum of piston work and sensible internal energy
change relative to total chemical energy input via fuel. 1 – Ș Figure 11. Intake runner pressure (runner #1).
8
Downloaded from SAE International by University of Minnesota, Tuesday, July 31, 2018

represents the fraction due to heat transfer to cylinder walls


10
plus enthalpy leakage through crevices. With the assumption
a) P-V diagram
the former mechanism dominates, the heat transfer is estimated
5
to increase with all blockages, presumably due to higher
convective heat transfer coefficient with increased mean flow.
Estimated exhaust enthalpy flow is also included in Table 4,

Pressure (bar)
with blockages showing a slight increase at WP and decrease
at IP. 1

0.55
0.5
a) WP
Intake Manifold Pressure (bar)

0.5
0.1
0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1
Volume/ Max Volume

0.45 fully open tumble swirl swumble

1.2

1
0.9
0.8
0.4 0.7
4 8 12 16 20 24 28 32 36 40 44 48 0.6
b) Pumping loop
SA 0.5

Pressure (bar)
0.4
0.34
b) IP 0.3
Intake Manifold Pressure (bar)

0.2

0.32

0.1
0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1
0.3 Volume/ Max Volume

Figure 13. In-cylinder pressure vs. volume at IP (cylinder #1).

Figure 14 shows the variation of exhaust runner mean gas


0.28
temperature with SA. Retarded ignition results in higher
12 16 20 24 28 32 36 40 44 48
exhaust temperature for all cases. The exhaust temperature
SA
shows an 8°C reduction with tumble and an increase of 7°C
Figure 12. Intake manifold pressure and 11°C with swirl and swumble, respectively, compared to
[The legend is same as Fig. 3.]
the fully-open case at the MBT timing of WP. At the IP, a
similar comparison gives a reduction of about 70°C, 30°C, and
50°C with tumble, swirl, and swumble, respectively. In view
Table 4. Comparison of PMEP, ISFC, exhaust enthalpy flow, and of comparable in-cylinder peak pressures in Fig. A1b among
net heat release ratio at MBT (set 1).
the fully open and blocked runner cases at IP, similar, though
not identical, in-cylinder peak temperatures would be expected
PMEP/ Exhaust Net heat
Operating Runner
SA
PMEP
IMEP
ISFC
enthalpy release
in the absence of other factors. The exhaust runner gas
condition blockage (kPa) (g/kW-hr) temperatures downstream of the cylinder with blockages at IP
(%) flow (kW) ratio Ș
may be further influenced by the charge motion: improved
fully open 34 50.49 16.63 270.05 9.03 0.69
work extraction from the gases during the expansion as
inferred from P-V diagrams (with the exception of swumble
WP tumble 14 59.16 18.31 262.69 9.22 0.66 yielding nearly identical expansion work to the fully open
swirl 24 58.59 18.98 275.32 9.43 0.67 runner) and an increase in the heat losses to the cylinder walls,
swumble 22 61.70 19.29 264.87 9.56 0.67 as discussed earlier.
fully open 40 65.56 40.80 315.91 3.40 0.64
Figures 15-17 show the variation of specific emissions of nitric
IP tumble 22 72.85 42.89 299.21 2.80 0.60 oxide (NOx), hydrocarbon (HC), and carbon monoxide (CO)
swirl 32 73.64 46.21 323.98 3.12 0.58 with SA. The well-established NOx reduction with spark
swumble 30 74.57 47.58 315.16 3.16 0.61 retard, as illustrated here in Fig. 15, correlates well with the
lower in-cylinder peak pressures (Refer to Fig. A2) due to the
9
Downloaded from SAE International by University of Minnesota, Tuesday, July 31, 2018

significance of temperature in the Zeldovich mechanism. For 720


near-stoichiometric, steady-state engine operation, the a) WP

Exhaust temperature (oC)


700
dominant single source of unburned hydrocarbons is known to
be the trapping of unburned mixture in combustion chamber 680
crevices [5, Chap. 11;27;28]. Thus, the lower in-cylinder
pressures may also inhibit such storage in crevices, therefore 660
reducing the HC emissions with retarded ignition, as observed
in Fig. 16. HC emissions have also been depicted in Fig. 18 as 640
a function of exhaust enthalpy flow with the MBT values
620
included in Table 4 as discussed earlier. With the moderate
levels of ignition retard in the present study, recall the 600
corresponding increase in exhaust gas temperatures from Fig. 4 8 12 16 20 24 28 32 36 40 44 48
14, a trend which helps better oxidation of HC particularly SA
with aggressive retardation. Blockages also exhibit CO 600
reduction in Fig. 17. Figures 19 and 20 depict the specific b) IP

Exhaust temperature (oC)


oxygen (O2) and carbon dioxide (CO2), respectively. For the
two operating conditions considered, swumble appears to yield 550
the least emissions. Somewhat comparable in-cylinder
pressures at MBT for both operating conditions (Recall Fig.
500
A1) do not allow an extrapolation for the behavior in the
resulting temperature versus NOx emissions. At the IP, with
comparable in-cylinder pressure, volume and mass flow rates 450
of air and fuel, leading possibly to similar average
temperatures, the unfavorable effect of blockages on NOx may
400
be attributed to higher local temperatures during and soon after
12 16 20 24 28 32 36 40 44 48
combustion due to higher burn rates, particularly for tumble.
SA
CO reduction with charge motion appears somewhat related to
mixing, which possibly decreases the local fuel-rich zones that Figure 14. Effect of blockages on exhaust temperature (Cylinder #4)
are the potential CO formation regions. At the WP, charge [The legend is same as Fig. 3.]
motion shows a reduction in O2 (Fig. 19), pointing to an
improved charge mixing and oxidation. At the IP, with the 8
exception of swumble exhibiting similar O2 levels, both tumble a) WP
7
and swirl again show reduction in O2. Finally, the CO2
emissions depicted in Fig. 20 closely resemble the behavior of
BSNOx (mg/kW-s)

6
BSFC, as expected. The emission comparisons here are based
5
on samples collected from exhaust primary runners. While this
approach provides a relative assessment of the blockages, the 4
absolute values of certain species (particularly HC and CO
3
reacting in ports and runners, unlike frozen NOx) are expected
to vary with location, for examples, from runner to manifold 2
collector (upstream of the converter brick).
1
4 8 12 16 20 24 28 32 36 40 44 48
CONCLUDING REMARKS SA

5
The effect of organized intake charge motion on combustion b) IP
and emission characteristics has been investigated by
4
comparing three blockages of 20% open area (swirl, tumble,
BSNOx (mg/kW-s)

and swumble) with an unrestricted intake runner. In summary:


3

x The blockages exhibit substantial and varying impact


2
on combustion, re-emphasizing the importance of
charge motion and turbulence intensity on burn rate.
1
The two part-load operating conditions exhibit, at
times, significant differences in trends, with more
dramatic improvements observed in combustion 0
12 16 20 24 28 32 36 40 44 48
parameters at the IP, which closely resembles engine
SA
idling.
Figure 15. Effect of blockages on NOx emission (cylinder #4)
[The legend is same as Fig. 3.]

10
Downloaded from SAE International by University of Minnesota, Tuesday, July 31, 2018

3.5 3.5
a) WP a) WP
3 3

BSHC (mg/kW-s)
BSHC (mg/kW-s)

2.5 2.5

2 2

1.5 1.5

1 1

0.5 0.5
4 8 12 16 20 24 28 32 36 40 44 48 8 9 10 11 12 13
SA Enthalpy flow (kW)
12 12
b) IP b) IP
10 10

BSHC (mg/kW-s)
BSHC (mg/kW-s)

8 8

6 6

4 4

2 2

0 0
12 16 20 24 28 32 36 40 44 48 2 2.5 3 3.5 4 4.5 5
SA
Enthalpy flow (kW)
Figure 16. Effect of blockages on HC emission (cylinder #4) Figure 18. HC emissions vs. exhaust gas enthalpy flow comparison (set 1)
[The legend is same as Fig. 3.] [The legend is same as Fig. 3.]

2.6 44
a) WP a) WP
2.4 40
BSO2 (mg/kW-s)
BSCO (mg/kW-s)

36
2.2
32
2
28
1.8
24
1.6
20

1.4 16
4 8 12 16 20 24 28 32 36 40 44 48 4 8 12 16 20 24 28 32 36 40 44 48
SA SA
7 96
b) IP b) IP
88
6.5
BSCO (mg/kW-s)

80
BSO2 (mg/kW-s)

6
72

5.5 64

56
5
48

4.5 40
12 16 20 24 28 32 36 40 44 48 12 16 20 24 28 32 36 40 44 48
SA SA
Figure 17. Effect of blockages on CO emission (cylinder #4) Figure 19. Effect of blockages on O2 emission (cylinder #4)
[The legend is same as Fig. 3.] [The legend is same as Fig. 3.]

11
Downloaded from SAE International by University of Minnesota, Tuesday, July 31, 2018

250 x Considerable changes are observed in the magnitude


a) WP
245
and phasing of the intake runner pressures
downstream of the blockage due to flow separation
BSCO2 (mg/kW-s)

240 and back flow during the intake stroke, leading to


235 higher pumping losses due to the restrictions, which
is reflected by the increase in BSFC at WP. The
230
exhaust runner pressures did not reveal a pattern
225 worthy of further discussion here.
220
x Retarded ignition results in increased exhaust runner
215 mean gas temperatures. At the WP, the effect of
4 8 12 16 20 24 28 32 36 40 44 48 charge motion on the exhaust runner gas temperatures
SA is nearly negligible, whereas at the IP, the gas
temperatures are reduced with the blockages
500
b) IP considerably; due possibly to the improved work
480 extraction during the expansion and increased heat
losses.
BSCO2 (mg/kW-s)

460

440
x In-cylinder peak pressures are reduced with the spark
retard due to combustion delay and expansion of
420 burned gases into a larger volume due to the piston
motion, leading to reductions in NOx and HC
400 emissions (due to lower in-cylinder temperatures and
presumably less trapped mixture in crevices,
380
12 16 20 24 28 32 36 40 44 48
respectively) with retarded ignition for all cases. For
SA both operating conditions at MBT, tumble and swirl
reduce CO emissions, while increasing either NOx or
Figure 20. Effect of blockages on CO2 emission (cylinder #4) HC at times. Swumble reduces emissions of NOx,
[The legend is same as Fig. 3.] HC, and CO at both operating conditions (with the
exception of a slight increase in NOx at the IP),
indicating that the emission results cannot simply be
x Combustion is considerably enhanced by tumble for averaged between tumble and swirl. From HC
the IP, with a reduction of 26q in the total burn emissions viewpoint, swumble appears to be most
desirable. The reductions in CO due to blockages
duration, in contrast to 8q of swirl and 14q of
often correlate with the reductions in O2, suggesting
swumble. The corresponding values for the WP are
improved combustion due to better mixing.
24q, 1q, and 6q respectively. In agreement with
Hadded and Denbratt [9], the larger part of the
improvement under both operating conditions takes It is emphasized that the distinct burn rate improvements due
place during burn delay. In-cylinder pressure remains to charge motion are demonstrated in the present study using a
comparable with the blockage vs. fully-open case at stoichiometric fuel-air mixture near MBT. Such a mixture
MBT for the IP, and with tumble vs. fully-open case strength leads to relatively limited benefits in combustion
for the WP, suggesting that the peak pressure stability due to charge motion and, at times, trade-offs for fuel
magnitude is an indicator of the improved combustion consumption. Further benefits are expected with more
phasing resulting from the charge motion. The aggressive spark retard for catalyst light-off and enleanment.
determination of the BSFC minimum, hence the MBT The relative merits of the three blockages used in this study
timing, is clear at the WP, while the lighter load at IP will thus be further explored under such inherently less stable,
presents a challenge. yet practically significant, combustion conditions.

x Consistent with the findings of Arcoumanis et al.


[15] for stoichiometric mixtures, the blockages do not
produce marked differences in COV at either
operating condition using MBT timing. For the
retarded ignition, however, the tumble shows a
noticeable reduction of 5% and swirl 4% at the IP.
The trends in LNV appear to provide additional
insight into the combustion stability with charge
motion under stoichiometric conditions.

12
Downloaded from SAE International by University of Minnesota, Tuesday, July 31, 2018

REFERENCES [14] Schock, H., Shen, Y., Timm, E., Stuecken, T., Fedewa,
A. and Keller, P., 2003, The measurement and control of
[1] Schelkin, K.I., 1943, On combustion in a turbulent flow, cyclic variation of flow in a piston cylinder assembly,
Journal of Technical Physics (USSR), Vol. XIII, Nos. 9- SAE paper 2003-01-1357.
10. (English translation, NACA TM 1110, 1947).
[15] Arcoumanis, C., Godwin, S.N. and Kim, J.W., 1998,
[2] Glassman, I., 1996, Combustion, Third edition, Chapter Effect of tumble strength on exhaust emissions in a
4, Academic Press, New York. single cylinder 4-valve Spark-Ignition engine, SAE paper
981044.
[3] Tabaczynski, R.J., 1976, Turbulence and turbulent
combustion in spark-ignition engines, Prog. Energy [16] Kent, J.C., Haghgooie, M., Mikulec, A., Davis, G.C. and
Comb. & Sci., Vol. 2, pp. 143-165. Tabaczynski, R.J., 1987, Effects of intake port design
and valve lift on in-cylinder flow and burn rate, SAE
[4] Lancaster, D.R., Krieger, R.B., Sorenson, S.C. and Hull, paper 872153.
W.L., 1976, Effects of turbulence on spark-ignition
engine combustion, SAE paper 760160. [17] Endres, H., NeuBer, H-J., and Wurms, R., 1992,
Influence of swirl and tumble on the economy and
[5] Heywood, J.B., 1988, Internal Combustion Engine emissions of multi-valve SI engines, SAE paper 920516.
Fundamentals, McGraw Hill, New York.
[18] Russ, S., Lavoie, G. and Dai, W., 1999 (1), SI engine
[6] Khalighi, B., Tahry, S.H.E., Haworth, D.C. and Huebler, operation with retarded ignition: Part 1 – Cyclic
M.S., 1995, Computation and measurement of flow and variations, SAE paper 1999-01-3506.
combustion in a four-valve engine with intake variations,
SAE paper 950287. [19] Russ, S., Thiel, M. and Lavoie, G., 1999 (2), SI engine
operation with retarded ignition: Part 2 – HC Emissions
[7] Lumley, J.L., 1999, Engines – An Introduction, pp. 134- and Oxidation, SAE paper 1999-01-3507.
181, Cambridge University Press.
[20] Nishizawa, K., Momoshima, S., Koga, M. and Tsuchida,
[8] Kang, K-Y., Oh, S-M., Lee, J-W., Lee, K-H. and Bae, C- H., 2000, Development of new technologies targeting
S., 1997, The effects of tumble flow on lean burn zero emission for gasoline engines, SAE paper 2000-01-
characteristics in a 4-valve SI engine, SAE paper 0890.
970791.
[21] Hallgren, B.E. and Heywood, J.B., 2003, Effects of
[9] Hadded, O. and Denbratt, I., 1991, Turbulence substantial spark retard on SI engine combustion and
characteristics of tumbling air motion in 4-valve S.I. hydrocarbon emissions, SAE paper 2003-01-3237.
engines and their correlation with combustion
parameters, SAE paper 910478. [22] Russ, S., 2004, Spark Retardation for Improving Catalyst
Light-Off Performance, [Chap. 6 in Advanced
[10] Urushihara, T., Nakada, T., Kakuhou, A. and Takagi, Y., Developments in Ultra-Clean Gasoline-Powered
1996, Effects of swirl/tumble motion on in-cylinder Vehicles, Edited by F. Zhao], SAE, Pennsylvania.
mixture formation in a lean-burn engine, SAE paper
961994. [23] Kothamasu, V., 1998, Effect of intake and exhaust
elements on sound attenuation and engine performance:
[11] Stockhausen, W.F., Wiemero, T.A., Ives, D.C. and An experimental and computational investigation,
Kronik, A.Y., 1996, Development and application of the Master’s Thesis, The Ohio State University.
Ford split port induction concept, SAE paper 961151.
[24] Shayler, P.J., Wiseman, M.W. and Ma, T., 1990,
[12] Urushihara, T., Murayama, T., Takagi, Y. and Lee, K.H., Improving the determination of mass fraction burnt, SAE
1995, Turbulence and cycle-by-cycle variation of mean paper 900351.
velocity generated by swirl and tumble flow and their
effects on combustion, SAE paper 950813. [25] Grimm, B.M. and Johnson, R.T., 1990, Review of simple
heat release computations, SAE paper 900445.
[13] Floch, A., Frank, J.V. and Ahmed, A., 1995, Comparison
of effects of intake-generated swirl and tumble on [26] Gatowski, J.A., Balles, E.N., Chun, K.M., Nelson, F.E.,
turbulence characteristics in a 4-valve engine, SAE paper Ekchian, J.A. and Heywood, J.B., 1984, Heat release
952457. analysis of engine pressure data, SAE paper 841359.

[27] Cheng, W.K., Hamrin, D., Heywood, J.B., Hochgreb, S.,


Min, K. and Norris, M., 1993, An overview of
13
Downloaded from SAE International by University of Minnesota, Tuesday, July 31, 2018

hydrocarbon emissions mechanism in spark-ignition and product gas expansion into a larger volume due to moving
engines, SAE paper 932708. piston, which is demonstrated in Figs. A2a and A2b for fully
open and tumble cases at the IP, respectively.
[28] Eng, J.A., Leppard, W.R., Najt, P.M. and Dryer, F.L.,
1998, The effect of fuel composition on hydrocarbon 20
emissions from a spark ignition engine: iso- a) WP
octane/toluene and n-octane/toluene fuel mixtures, SAE 16
paper 982557.

Pressure (bar)
12

ABBREVIATIONS
8

AFR Air Fuel Ratio


4
BMEP Brake Mean Effective Pressure
0
BSFC Brake Specific Fuel Consumption -90 -60 -30 0 30 60 90
CAD
ISFC Indicated Specific Fuel Consumption
fully open tumble swirl swumble
LNV Lowest Normalized Value 12
CAD Crank Angle Degree b) IP

COV Coefficient of Variation 9

Pressure (bar)
EGR Exhaust Gas Recirculation
6
EOC End of Combustion
EVO Exhaust Valve Opening 3
IBDC Intake Bottom Dead Center
IMEP Indicated Mean Effective Pressure 0
-90 -60 -30 0 30 60 90
PMEP Pumping Mean Effective Pressure CAD

IP high speed Idle Point Figure A1. Effect of blockages on in-cylinder pressure.

MBT Maximum Brake Toque 15


SA a) fully open
MFB Mass Fraction Burnt
12 20
SA Spark Advance 28
Pressure (bar)

9
TDC Top Dead Center 30

38
WP World-wide Point 6
40
CMCV Charge Motion Control Valve
3

0
-90 -60 -30 0 30 60 90
APPENDIX A: SAMPLE IN-CYLINDER CAD
PRESSURE COMPARISONS 15
b) tumble
SA
At MBT, Fig. A1a shows that in-cylinder peak pressures for 12
14
swirl and swumble are slightly lower than the fully open case
Pressure (bar)

at the WP, while tumble remains comparable. The peak 9


22

magnitudes are also comparable between partially blocked and 24


fully open runners for the IP (Fig. A1b), demonstrating the 6 26
retention of pressure magnitudes as the impact of improved
34
combustion with charge motion. In-cylinder peak pressure is 3
typically reduced with spark retard due to delayed combustion
0
-90 -60 -30 0 30 60 90
CAD

Figure A2. Variation of in-cylinder pressure with SA at IP.

14

You might also like