Sei sulla pagina 1di 7

Energy Conversion and Management 79 (2014) 5965

Contents lists available at ScienceDirect

Energy Conversion and Management


journal homepage: www.elsevier.com/locate/enconman

The Miller cycle effects on improvement of fuel economy in a highly


boosted, high compression ratio, direct-injection gasoline engine: EIVC
vs. LIVC
Tie Li , Yi Gao, Jiasheng Wang, Ziqian Chen
School of Mechanical Engineering, Shanghai Jiao Tong University, PR China

a r t i c l e

i n f o

Article history:
Received 18 October 2013
Accepted 5 December 2013
Available online 29 December 2013
Keywords:
Fuel economy
Gasoline engines
Intake boosting
Miller cycle
Knock
Pumping loss

a b s t r a c t
A combination of downsizing, highly boosting and direct injection (DI) is an effective way to improve fuel
economy of gasoline engines without the penalties of reduced torque or power output. At high loads,
however, knock problem becomes severer when increasing the intake boosting. As a compromise, geometric compression ratio (CR) is usually reduced to mitigate knock, and the improvement of fuel economy is discounted. Application of Miller cycle, which can be realized by either early or late intake
valve closing (EIVC or LIVC), has the potential to reduce the effective CR and suppress knock. In this paper,
the effects of EIVC and LIVC on the fuel economy of a boosted DI gasoline production engine reformed
with a geometric CR of 12.0 are experimentally compared at low and high loads. Compared to the original
production engine with CR 9.3, at the high load operation, the brake specic fuel consumption (BSFC) is
improved by 4.7% with CR12.0 and LIVC, while the effect of EIVC on improving BSFC is negligibly small. At
the low load operation, combined with CR12.0, LIVC and EIVC improve the fuel economy by 6.8% and
7.4%, respectively, compared to the production engine. The mechanism behind the effects of LIVC and
EIVC on improving the fuel economy is discussed. These results will be a valuable reference for engine
designers and researchers.
2013 Elsevier Ltd. All rights reserved.

1. Introduction
Requirement for better fuel economy and reduction in exhaust
gas emissions in automobiles are driving the introduction of a wide
range of new gasoline engine technologies. One of the most promising technologies is the downsized, highly boosted, direct-injection spark-ignition (DISI) engine [17]. In automotive gasoline
engines, the most frequent operations are at low loads and the load
control is usually implemented by using a throttle valve to restrict
airow into cylinders, which results in pumping loss during gas exchange strokes. With keeping the engine torque output constant, a
decrease in the engine displacement will lead to a shift in the typical engine operating range to higher loads, which will help reduce
the pumping loss and thus the fuel consumption. To meet vehicle
requirements for the maximum power output, intake boost is generally necessary for downsized engines at high loads, causing
knock a more severe problem compared with their naturally
aspirated (NA) counterparts. A reduction in geometric compression
ratio (CR) is the most typical solution to mitigate the knock prob Corresponding author. Address: School of Mechanical Engineering, Shanghai
Jiao Tong University, 800 Dong Chuan Rd., Shanghai 200240, PR China. Tel./fax: +86
21 3420 3749.
E-mail address: litie@sjtu.edu.cn (T. Li).
0196-8904/$ - see front matter 2013 Elsevier Ltd. All rights reserved.
http://dx.doi.org/10.1016/j.enconman.2013.12.022

lem in downsized gasoline engines at high loads, but this strategy


is at the expanses of reduced fuel economy benet from engine
downsizing.
A strategy of shortening compression stroke relative to expansion stroke, called Miller cycle, is effective to reduce the temperature and pressure at the end of compression stroke, and therefore
suppress knock. Miller cycle can be realized by either early or late
intake valve closing (EIVC or LIVC). The original Miller engine employed a compression control valve on the cylinder head to release
part of the charge to the exhaust port during the compression
stroke to reduce the effective CR [8,9]. With the thermodynamic
analysis of ideal Miller cycle, Wu et al. [10] concluded that the
Miller cycle has no inherent efciency advantage over the Otto cycle because of the shorten compression stroke. However, they
emphasized that the engine knock problem of the Otto engine
could be suppressed by the lowered temperature and pressure of
the airfuel mixture before combustion with LIVC or EIVC, which
should be benecial for supercharging. Analysis of air-standard
Miller cycle with considerations of heat transfer loss, friction and
variable specic heats of working uid demonstrated that Miller
cycle is more efcient than Otto cycle and the effects of heat
transfer loss, friction and specic heat ratio on fuel efciency are
signicant and should be considered in practical cycle analysis
[1114].

60

T. Li et al. / Energy Conversion and Management 79 (2014) 5965

Nomenclature
A
a
b
Cv
E
m
ne
P
Q
R
Rc
T
U
V
Vc
X
e

c
h

constant ()
constant ()
constant ()
isochoric specic heat (J/(mol K))
activity energy (J)
molar number (mol)
engine speed (rpm)
pressure (bar)
apparent heat release (J)
universal gas constant (J/(mol K))
the ratio of connecting rod length to crank radius ()
temperature (K)
internal energy (J)
volume (cm3)
the clearance volume (cm3)
parameter related to fuel or mixture properties ()
compression ratio ()
ratio of specic heat ()
crank angle (CAD ATDC)

In practical spark-ignition engines, as aforementioned, pumping


loss signicantly limits the engine efciency at low loads, while
knock restricts the compression ratio and spark advances that result in low thermal efciency at high loads [1517]. Cleary et al.
[18] investigated the part-load fuel consumption potential of
unthrottled operation of a single cylinder engine using the EIVC
variable valve actuation. They reported that a 7% fuel consumption
improvement is achieved through the optimization of the intake
valve lift, duration and timing while maintaining the conventional
exhaust valve event. Not as they expected, however, the enhanced
in-cylinder charge motion owing to the unthrottled operation with
EIVC did not lead to signicant combustion performance improvement. Miklanek et al. [19] pointed out that due to the shortened
compression stroke with Miller cycle, the lowered in-cylinder
charge temperature at the beginning of the combustion may lead
to a slower combustion speed compared to a standard Otto cycle,
which could negatively inuence the engine efciency. They proposed a strategy combining Miller cycle and mixture heating, and
with one dimensional simulation they demonstrated a signicant
improvement in the fuel economy. Anderson et al. [20] conducted
the rst and second law analysis and compared a NA, Miller cycle,
SI engine with LIVC to a conventionally-throttled Otto cycle counterpart at light load conditions. The rst law analysis revealed that
as much as 6.3% improvements in the indicated thermal efciency
are obtained with the Miller cycle, and these improvements are
primarily due to reductions in pumping and compression work
while heat transfer losses are comparable. The second law analysis
showed that the throttling process in the conventional engine destroys up to 3% of the fuel availability and that the higher pressure
in the LIVC intake manifold leads to a notable thermo-mechanical
advantage over the conventional throttled engine.
In the original design of Miller engine, a supercharger was attempted to remedy the inadequacy of power output due to the
shortened compression stroke [8,9]. From another viewpoint, the
shortened compression stroke can be actively utilized to mitigate
knock in boosted or high CR engines. In the development of Mazda
V6 Miller cycle engine equipped with a LIVC mechanism and a Lysholm supercharger, the ram-pulsation effects of the Miller cycle
were well utilized to enhance the anti-knocking performance
[2123]. Wang et al. [24] examined the effects of the Miller cycle
on a high CR SI engine, and concluded that LIVC with a high valve
lift should be implemented at high loads to reduce the effective CR

ignition delay of end gas (ms)

Acronyms
ATDC
after top dead center
BMEP
brake mean effective pressure (MPa)
BSFC
brake specic fuel consumption (kg/kW h)
CA50
combustion phasing dened by the crank angle of 50%
accumulative heat release (CA)
CAD
crank angle degree
CR
compression ratio ()
DISI
spark-ignited direction-injection
EIVC
early intake valve closing
LIVC
late intake valve closing
MBT
maximum brake torque
NA
naturally aspirated
PMEP
pumping mean effective pressure (MPa)
ROHR
rate of heat release
TDC
top dead center

and knock tendency while EIVC with a low valve lift should be
adopted at low loads to improve the thermal efciency. Gottschalk
et al. [25] investigated the effect of the Miller cycle with EIVC on
knock control in a turbocharged, downsized DISI engine at the middle loads. They demonstrated that the Miller cycle strategies offer
high potential in cyclic variations of combustion, fuel economy and
exhaust gas temperature.
From the above paragraphs, it can be seen that there have been
extensive researches on Miller cycle including both theoretical cycle analysis and experimental studies on practical engines. A combination of Miller cycle and modern highly boosted downsized
engine should have great potentials in further improvement of fuel
economy in SI engines. Either EIVC or LIVC is the most practical approach to realize the Miller cycle, thanks to technical advances in
variable valve actuations and electronic controls. However, a systematic comparison of the performances of EIVC and LIVC on
improving the fuel economy of a modern highly-boosted, downsized DI gasoline engine with high CR at low and high loads is
rarely reported, although such information is very valuable for engine designers.
In this study, the Miller cycle effects of EIVC and LIVC on the fuel
economy of a boosted DI gasoline production engine modied with
a geometric CR of 12.0 are experimentally compared at low and
high loads, meanwhile the mechanism of these effects is discussed.
The details are reported as follows.

2. Experimental approaches and procedure


A 2.0 liter 4-cylinder DISI engine was operated at low and high
load conditions: 2000 rpm 0.4 MPa brake mean effective pressure
(BMEP) and 1000 rpm 1.32 MPa BMEP. Here the condition of
2000 rpm 0.4 MPa BMEP is one of the frequently operating conditions of the engine, representative of typical partial load in this
study, while 1.32 MPa BMEP is wide open throttle (WOT) condition
at 1000 rpm for the engine, and it was used as a representative of
low-speed high-load operation in this study. Engine performance
was compared at the geometric compression ratios (CR) of 9.3
and 12.0 congured with the base camshaft and CR12.0 with the
EIVC and LIVC camshafts. A summary of the engine specications
are provided in Table 1, where the rated power and torque are taken from the original production engine. The lift proles of the in-

T. Li et al. / Energy Conversion and Management 79 (2014) 5965

Fig. 2 shows a schematic of the experiment0al setup. An AVL AC


dynamometer system was operated in speed-control mode to
maintain the desired engine speed. The engine coolant and oil temperatures were maintained at 90 1.5 C using the conditioning
systems. An inlet air conditioning system along with an intercooler
conditioning system was used to maintain the temperature,
humidity, and pressure of the inlet air at desired values. Because
the temperature at the compressor outlet varied in a range of
4452 C at the low load and 73112 C at the high load for the different strategies, an in-house made intercooler was employed to
keep the intake temperature at 34 1.0 C for all the tests in this
study. Accelerometer-type knock sensors were used to monitor
knocking combustion. A fuel ow meter (AVL733S) with a precision of 0.12% and a high-precision (0.01%) torque sensor (HBM)
_ f and the torque, T
were utilized to measure the fuel ow rate, m
and then the brake specic fuel consumption (BSFC) can be calculated as

Table 1
Engine specications, fuel and test conditions.
Cylinder number
Bore  stroke
Displacement
Connecting rod length
Crank radius
Compression ratio
Intake and exhaust systems
Valve number
Fuel delivery system
Fuel
Rated powera
Rated torquea
Intake boosting
Test conditions

In-line 4
86  86 mm
1998 cm3
145.5 mm
43.0 mm
9.3, 12.0
Duel VVT
16
Direct injection
#97 gasoline (RON97)
190 kW @ 5300 rpm
353 Nm @ 2000  5250 rpm
Turbocharger + supercharger
1000 rpm 1.32 MPa BMEP
2000 rpm 0.4 MPa BMEP
0.9 at 1000 rpm 1.32 MPa BMEP
1.0 at 2000 rpm 0.4 MPa BMEP

Excess air ratio


a

For the original production engine.

BSFC
Base Cam
LIVC Cam

Lift

EIVC Cam

TDC

BDC

Crank Angle
Fig. 1. A schematic of valve lift proles for the different camshafts.

take valve for the original, EIVC and LIVC camshafts are schematically shown in Fig. 1. LIVC is obtained by extending the duration of
the maximum lift of the base camshaft, while EIVC is realized with
a reduced lift compared to the original prole. The high-pressure
injectors were located below the two intake ports (side injection)
with an operating pressure up to 15.0 MPa. Both a supercharger
and a turbocharger were used to provide sufcient intake air quantity to achieve the desired torque levels when using different camshafts. ETAS hardware and INCA software were used to
communicate with the engine controller to optimize the engine
performance at each operating condition. The commercial #97 gasoline (research octane number 97) in Chinese market was used as
fuels for all the tests.

_f
_f
m
30m

T  2pne =60 T  pne

h
i
1=2
1
2
V=V c 1 e  1 Rc 1  cos h  Rc 2  sin h
2

where Vc is the clearance volume, e the compression ratio and Rc the


ratio of connecting rod length to crank radius.
Based on the rst law of thermodynamics, the released heat at
the crank angle h is

Fresh air

Inter
cooler

ExhaustTurbine

where ne is the engine speed in revolution per minute.


Before each test, at rst, the engine was well conditioned by
warming-up operation from low to high loads for 30 min to ensure
the data repeatability. Then, after the engine was set at the desired
operating conditions, the engine was optimized in terms of fuel
economy by tuning the fuel injection timing and pressure, the excess air ratio as well as the phasing of intake and exhaust valve
events. Finally, spark sweeps were carried out, and the points of
knock limit at the high load operation and maximum brake torque
(MBT) at the low load operation were identied and the data were
recorded by the sampling system.
Piezoelectric in-cylinder pressure transducers (Kistler 6125A)
with ame arrestors were ush-mounted through the cylinder
head between the intake and exhaust valves to collect the in-cylinder pressure. The pressure signals were processed using the charge
ampliers (Kistler 5064B) and a high-speed data acquisition system at a resolution of 0.2 crank angle degrees (CAD) triggered by
a crankshaft encoder. For an internal combustion engine with a
mechanism of connecting rod and crank shaft, the cylinder volume
V at any crank position hcan be written as

Throttle valve

Dyno

61

compressor

Air
cleaner

Supercharger

Fig. 2. A schematic of experimental setup for the engine test bench.

62

T. Li et al. / Energy Conversion and Management 79 (2014) 5965

CA50 (CAD ATDC)

BSFC (g/kwh)

35

4.5%
2.2%
5

CR 9.3

CR 12.0

1000 rpm
1.32 MPa BMEP

CR 9.3

30
25
20
15
10
5
0

CR 9.3

CR 12.0

2000 rpm
0.4 MPa BMEP

CR 12.0

Fig. 4. Effect of the higher compression ratio on the combustion phasing (CA50) at
the knock limited point (ne = 1000 rpm, BMEP = 1.32 MPa).

Fig. 3. Effect of the higher compression ratio (CR) on the brake specic fuel
consumption (BSFC) at the high and low load operating conditions.

where U is the internal energy, m the molar number, T the average


temperature of the working gas, and Cv the isochoric specic heat
given by

Cv

c1

where R is gas constant and c the specic heat ratio that can be calculated using the data in the JANAF table [26,27]. Combining the Eq.
(4) and the equation of state of ideal gas PV = mRT into Eq. (3), the
rate of heat release is obtained as



dQ
1
dP
dV
PV
dc


V
P
dh c  1
dh
dh
c  12 dh

s AXa Pb exp

E
RT


6

where A, a and b are constants, X is parameters related to fuel or


mixture properties, and E is related to activity energy. It is clear
from Eq. (6) that an increase in the intake pressure leads to higher
P of the end gas, increasing the knock propensity, and a higher CR
deteriorates the knock problem by increasing both P and T. Therefore, despite the potential of fuel economy improvement at light
loads, a relatively low CR is usually employed for boosted engine.
3.2. Effects of EIVC and LIVC at high loads

3. Results and discussion


3.1. Effect of higher compression ratio on fuel economy
Tests were conducted at geometric compression ratios of 9.3
and 12.0 with the base camshaft to compare the effect of implementing higher CR piston on the fuel economy. An operating condition of 2000 rpm 0.4 MPa bar in brake mean effective pressure
(BMEP) was chosen to represent the part load and 1000 rpm
1.32 MPa BMEP for the high load.
Fig. 3 shows the fuel consumption results for the geometric CRs
of 9.3 and 12.0 at the MBT point for the low load condition and the
knock limited point for the high load condition. A 4.5% improvement in the brake specic fuel consumption (BSFC) is observed
when increasing the geometric CR from 9.3 to 12.0 for the low load
condition.
Although the higher CR improves the part-load fuel consumption, knock becomes severer at the high load condition. Here
CA50, the crank angle where 50% of the inducted fuel energy has
been released, was used to characterize the engines knock performance. When increasing CR from 9.3 to 12.0, the engine becomes
more knock limited, where CA50 is retarded by about 11.0 CAD
at 1000 rpm 0.4 MPa BMEP engine operating condition, as shown
in Fig. 4. As a result, as shown in Fig. 3, compared to CR of 9.3,
implementation of the CR 12.0 increases BSFC by 2.2% at the knock
limited spark timing (KLSA) due to the CA50 penalty.
Knock is one of the major constraints to improve the thermal
efciency of spark-ignition (SI) engines. Moderate knock, usually
dened by the clanging or pinging sound, increases the engine
vibration and noise, while heavy knock may directly lead to the engine failure [28]. It is generally accepted that knock is attributed to
the oscillation of pressure wave resulted from the rapid heat

Shown in Fig. 5 are the effects of EIVC and LIVC on improving


BSFC at the knock limited point under the high load condition.
For comparison purpose, the data at the same operating condition
in Fig. 3 are also plotted. In comparison to the case of base camshaft and CR 12.0, both EIVC and LIVC result in signicant improvements of BSFC. Particularly, BSFC is improved by 6.9% by using LIVC
compared to the case of base camshaft and CR 12.0, and it is even
better than the case of base camshaft and CR 9.3. However, no
improvement in BSFC is observed by using the EIVC strategy, compared to the original production engine of CR 9.3. Here it should be
mentioned that the intake valve closure was swept in a range of
161191 and 231251 CAD ATDC during gas exchange strokes
for EIVC and LIVC, respectively, and the data in Fig. 5 are the optimized results in terms of the intake camshaft phasing for the lowest fuel consumption.
Fig. 6 shows the effects of EIVC and LIVC on the combustion
phasing (CA50) at the knock limited point under the high load condition. Compare to the case of base camshaft and CR 12.0, the combustion phasing is advanced by 3 crank angle degrees (CAD) for

BSFC (g/kwh)

dQ dU
dV
d
dV

P
mC v T P
dh
dh
dh dh
dh

release of auto-ignited unburned mixture (so called end gas) ahead


of the propagating ame front. The ignition delay s of end gas can
be expressed as an Arrhenius-type correlation [2931].

2.2%

2.0%
6.9%

CR 9.3

CR 12.0

CR 12.0

Base Cam Base Cam LIVC

CR 12.0

EIVC

Fig. 5. Effects of EIVC and LIVC on improving BSFC at the knock limited point under
the high load condition (ne = 1000 rpm, BMEP = 1.32 MPa).

63

T. Li et al. / Energy Conversion and Management 79 (2014) 5965

30
25
20
15
10
5
0
CR 9.3
CR 12.0 CR 12.0
Base Cam Base Cam
LIVC

CR 12.0
EIVC

Cylinder pressure (MPa)

Fig. 6. Effects of EIVC and LIVC on the combustion phasing (CA50) at the knock
limited point under the high load condition (ne = 1000 rpm, BMEP = 1.32 MPa).

Base Cam CR12.0


LIVC Cam CR12.0
EIVC Cam CR12.0

64
6.4
32
3.2
1.6
16
8
0.8
4
0.4
0.2
2
0.1
1
0.5
0.05
40

80

160

320

640

Volume (cm3)

Cylinder pressure (MPa)

Fig. 7. PV diagrams of the whole engine cycle with EIVC and LIVC at the knock
limited point under the high load condition (ne = 1000 rpm, BMEP = 1.32 MPa).

0.14

downward. As a result, there would be signicant heat transfer


from the hot cylinder wall to the charge, which can be deduced
from the pressure trace after BDC, as shown in Fig. 8. For LIVC,
the pressure at the closure of intake valve is higher than for EIVC.
However, for LIVC, since the charge is continually inducted until
the valve closure after BDC, both the temperature and heat capacity of the charge around BDC could be higher than for EIVC. Therefore, there could be less heat transfer from the cylinder to the
charge and thus the smaller temperature rise for LIVC than for
EIVC. In addition, after BDC some charge would be pushed out
the cylinder by the piston upward motion owing to the late valve
closure, taking away some heat from the cylinder. As a consequence, as shown in Figs. 7 and 9, both the pressure and temperature during the compression stroke are lower for LIVC than for
EIVC, resulting in the better anti-knock performance for LIVC than
for EIVC as have been shown in Fig. 6.
Note that in Fig. 8 while the pressure during the intake stroke
before the valve closure is similar for both EIVC and LIVC, the pressure during the exhaust stroke for LIVC is signicantly lower than
for EIVC. This may be attributed to the fact that LIVC advanced the
combustion phasing as has been shown in Fig. 6, reducing the pressures during the expansion stroke, after the blow down as well as
during the exhaust stroke as can be seen in Fig. 7. As a result, LIVC
leads to a larger pumping work than EIVC.
Fig. 10 shows the comparison of pumping work for the different
cases at the knock limited point under the high load condition. The
data in the case of base camshaft and CRs of 9.3 and 12.0 are also
plotted in the gure for reference. To maintain the load of 1.32 MPa
BEMP at 1000 rpm, the intake charge was aggressively boosted so
that the intake pressure is higher than the exhausting pressure.
Therefore, here the pumping work plays a positive role in the fuel
economy. In comparison to the base camshaft, while EIVC reduces
the pumping work, LIVC signicantly increases the pumping work.
Therefore, here the better performance of LIVC in improvement
of fuel consumption over EIVC can be attributed to the advanced
combustion phasing as well as the increased pumping work.

Intake valve closure

3.3. Effects of EIVC and LIVC at low loads


LIVC Cam CR12.0
EIVC Cam CR12.0
0.07

40

80

160

320

640

Volume (cm3)
Fig. 8. PV diagrams of the exhaust and intake strokes of the engine with EIVC and
LIVC at the knock limited point under the high load condition (ne = 1000 rpm,
BMEP = 1.32 MPa).

EIVC, while it is advanced by up to 10 CAD for LIVC to a value comparable to the case of CR 9.3 and base camshaft. While both strategies suppress the engine knock, LIVC exhibits higher potentials in
the suppression of knock than EIVC, despite the higher effective CR
for LIVC than for EIVC as will be shown below.
Figs. 7 and 8, respectively, show the PV diagrams of the whole
cycle and during the intake and compression strokes of the engine
with EIVC and LIVC at the knock limited point under the operating
condition of 1000 rpm 1.32 MPa BMEP. In this study, the effective
CR is dened as the volume ratio of the in-cylinder charge at the
moment of intake valve closure and at the top dead center (TDC).
As shown in Fig. 7, despite the higher effective CR of 9.9 for LIVC
than the CR 9.3 for EIVC, the pressure at the end of compression
stroke for LIVC is lower than for EIVC. This may be attributed
mostly to the heat transfer from the cylinder to the charge. For
EIVC, since the intake valve closes early before the bottom dead
center (BDC), the charge is actually expanded as the piston moves

Fig. 11 shows the effects of EIVC and LIVC on improving BSFC at


MBT point under the operating condition of 2000 rpm 0.4 MPa bar
BMEP. Both EIVC and LIVC exhibit good performance: a 2.9% reduction in BSFC for EIVC and a 2.3% reduction for LIVC are obtained
compared to the case of base camshaft and CR 12. Compared to
the case of base camshaft and CR 9.3, the improvements are 7.2%
and 6.7% for EIVC and LIVC, respectively. Here the better fuel economy performance of EIVC than LIVC is consistent with the earlier
studies [24,32].

2500

Cylinder temperature (K)

CA50 (CAD ATDC)

35

2000
1500
1000

Base Cam CR12.0


LIVC Cam CR12.0
EIVC Cam CR12.0

500
0
-30

30

60

90

Crank Angle (ATDC)


Fig. 9. History of in-cylinder averaged temperature for various camshafts at the
knock limited point under the high load condition (ne = 1000 rpm,
BMEP = 1.32 MPa).

64

T. Li et al. / Energy Conversion and Management 79 (2014) 5965

0.07

-0.025
-0.03

PEMP (MPa)

PEMP (MPa)

0.06
0.05
0.04
0.03
0.02

-0.035
-0.04
-0.045

0.01
-0.05

0
CR 9.3
CR 12.0 CR 12.0
Base Cam Base Cam LIVC

CR 9.3 CR 12.0 CR 12.0


Base Cam Base Cam
LIVC

CR 12.0
EIVC

Fig. 10. Effects of EIVC and LIVC on the pumping work at the knock limited point
under the high load condition (ne = 1000 rpm, BMEP = 1.32 MPa).

CR 12.0
EIVC

Fig. 13. Effects of EIVC and LIVC on pumping work at the MBT point under the low
load condition (ne = 2000 rpm, BMEP = 0.4 MPa).

7.4%

CR 9.3
Base Cam

CR 12.0 CR 12.0
Base Cam
LIVC

CR 12.0
EIVC

Fig. 11. Effects of EIVC and LIVC on improving BSFC at the MBT point under the low
load condition (ne = 2000 rpm, BMEP = 0.4 MPa).

Rate of heat release (J/deg)

BSFC (g/kwh)

6.8%

35
30
25

100

CR9.3 Base Cam


CR12 Base Cam
CR12 LIVC
CR12 EIVC

90
80
70
60

20

50

15

40
30

10

20

5
0
-60

10
0
-40

-20

20

40

60

Normalized cumulative heat release (%)

40

4.5%

Crank angle degree (CAD ATDC)

Cylinder pressure (MPa)

0.16

Fig. 14. Effects of EIVC and LIVC on rate of heat release at the MBT point under the
low load condition (ne = 2000 rpm, BMEP = 0.4 MPa).

Base Cam CR12.0


LIVC Cam CR12.0
EIVC Cam CR12.0

0.08

Intake valve closure


0.04
40

80

160

320

640

Volume (cm3 )
Fig. 12. PV diagrams of the exhaust and intake strokes of the engine with EIVC and
LIVC at the MBT point under the low load condition (ne = 2000 rpm,
BMEP = 0.4 MPa).

Fig. 12 shows the PV diagrams during the intake and compression strokes of the engine with EIVC and LIVC at the MBT point under the conditions of 2000 rpm 0.4 MPa BMEP. Here the effective
CRs are 9.7 and 11.4 for EIVC and LIVC, respectively. It should be
pointed out that the variable valve timing mechanisms were used
for both the intake and exhaust systems, and here the valve timings have been optimized for the best fuel economy for the different camshafts. At the low load operation, without the intake boost
and with the throttle, the in-cylinder pressure during the intake
stroke is lower than that during compression stroke. Therefore,
the pumping work plays a negative role in the fuel economy.
Fig. 13 shows the effects of EIVC and LIVC on the pumping effective mean pressure (PMEP) at 2000 rpm 4.0 bar BMEP. The EIVC
strategy results in the smallest absolute PMEP, followed by LIVC,
and the base camshafts for the two CRs shows similar performance.
An explanation can be found from Fig. 12. With EIVC, a relatively
large open angle of the intake throttle valve could be implemented,

resulting in relatively higher intake pressure. With the similar


pressure during the exhaust stroke, the higher intake pressure
would lead to a less pumping loss. Therefore, the reduction of
pumping loss for EIVC should be one of the most important contributors to the improved fuel economy. The reason for the superior performance of EIVC over LIVC is not yet fully understood,
and one explanation goes to the higher lift for LIVC that could require more throttled operation to restrict the air ow and thus result in more pumping loss than for EIVC.
With EIVC, besides the reduced pumping loss, improved fuel
air mixing and enhanced in-cylinder turbulence strength were reported, which could contribute to combustion [18,32]. As shown in
Fig. 14, however, EIVC shows the lowest rate of heat release, followed by LIVC, despite the almost identical CA50 for all the cases.
As mentioned in reference [19] and shown in Fig. 12 in this study,
the lowered pressure and temperature at the end of compression
stroke could be the reason for the slower rate of heat release in
the case of EIVC. However, also shown in Fig. 12, LIVC leads to a
higher pressure during the compression stroke than the cased of
base camshaft and CR12.0. Here the reason is not clear at present
and further study is needed to address this issue.
4. Conclusions
Part- and high-load evaluations have been conducted with respect to the performance of EIVC and LIVC with a high CR of 12.0
for the boosted DISI engine. The results are summarized as follows:
1. At the part-load operation, a fuel consumption improvement is
obtained with increasing the geometric CR from 9.3 to 12.0.
However, at the high-load condition, the higher CR increases
knock tendency and the fuel consumption deteriorates.

T. Li et al. / Energy Conversion and Management 79 (2014) 5965

2. Improved anti-knock performance is achieved when implementing either EIVC or LIVC strategies for the CR 12.0 at the
high load condition. A fuel economy even better than the case
of base camshaft and CR9.3 is obtained with a combination of
CR12 and LIVC.
3. LIVC is advantageous over EIVC in the fuel consumption
improvement at the high load, primarily owing to the better
knock resistance and greater pumping work of LIVC than EIVC.
4. At the low load, the fuel economy can be improved with either
EIVC or LIVC, primarily owing to the decreased pumping loss.
5. The reduction in the pumping loss with EIVC is greater than
with LIVC. However, the rate of heat release for EIVC is slower
for LIVC, probably owing to the lowered pressure and temperature at the end of compression stroke.

Acknowledgments
The Supports by Doctoral Fund of Ministry of Education of China (20120073120059), Shanghai Pujiang Program (13PJ1404300),
and the Recruitment Program of Global Youth Experts of Chinese
Government are gratefully acknowledged.
References
[1] Lecointe B. Monnier G. Downsizing a gasoline engine using turbo-charging
with direct injection. SAE Paper; 2003 [2003-01-0542].
[2] Pallotti P, Torella E. New application of an electric boosting system to a small,
four-cylinder S.I. engine. SAE Paper; 2003 [2003-32-0039].
[3] Lake T, Stokes J, Murphy R, Osborne R, Schamel A. Turbo-charging concepts for
downsized DI gasoline engines. SAE Paper; 2004 [2004-01-0036].
[4] Knigstein A, Hock C, Frensch M. Comparison of advanced turbocharging
technologies under steady-state and transient conditions. SAE Paper; 2006
[2006-05-0364].
[5] Zaccardi JM, Pagot A, Vangraefschepe F. Optimal design for a highly downsized
gasoline engine. SAE Paper; 2009 [2009-01-1794].
[6] Alger T, Gingrich J, Roberts C, Mangold B. Cooled exhaust-gas recirculation for
fuel economy and emissions improvement in gasoline engines. Int J Engine Res
2011;12:25264.
[7] Li T, Wu D, Xu M. Thermodynamic analysis of EGR effects on the rst and
second law efciencies of a boosted spark-ignited direction-injection gasoline
engine. Energy Convers Manage 2013;70:1308.
[8] Miller R. Supercharge and internally cooling for high output. ASME Trans
1947;69:45364.
[9] Miller R. Supercharged engine. Patent US2817322 A; 1957.
[10] Wu C, Puzinauskas PV, Tsai JS. Performance analysis and optimization of a
supercharged Miller cycle Otto engine. Appl Therm Eng 2003;23:51121.
[11] Ge Y, Chen L, Sun F, Wu C. Effects of heat transfer and friction on the
performance of an irreversible air-standard miller cycle. Int Commun Heat &
Mass Trans 2005;32(8):104556.

65

[12] Hou SS. Comparison of performances of air standard Atkinson and Otto cycles
with
heat
transfer
considerations.
Energy
Convers
Manag
2007;48(5):168390.
[13] Lin JC, Hou SS. Performance analysis of an air-standard Miller cycle with
considerations of heat loss as a percentage of fuels energy, friction and
variable specic heats of working uid. Int J Therm Sci 2008;47(2):18291.
[14] Al-Sarkhi A, Jaber JO, Probert SD. Efciency of a Miller engine. Appl Energy
2006;83:34351.
[15] Kutlar OA, Arslan H, Calik AT. Methods to improve efciency of four stroke,
spark ignition engines at part load. Energy Convers Manage
2005;46(20):320220.
[16] Alkidas AC. Combustion advancements in gasoline engines. Energy Convers
Manage 2007;48:275161.
[17] Caton JA. The thermodynamic characteristics of high efciency, internalcombustion engines. Energy Convers Manage 2012;58:8493.
[18] Cleary D, Silvas G. Unthrottled engine operation with variable intake valve lift,
duration, and timing. SAE Paper; 2007 [2007-01-1282].
[19] Miklanek L, Vitek O, Gotfryd O, Klir V. Study of unconventional cycles
(Atkinson and Miller) with mixture heating as a means for the fuel economy
improvement of a throttled SI engine at part load. SAE Int J Engines
2012;5(4):162436.
[20] Anderson M, Assanis D, Filipi Z. First and second law analyses of a naturallyaspirated, Miller cycle, SI engine with late intake valve closure. SAE Paper;
1998. p. 980889.
[21] Choshi M, Asanomi K, Abe H, Okamoto S, Shoji M. Development of V6 Miller
cycle engine. JSAE Rev 1994;15(3):195200.
[22] Hatamura K, Hayakawa M, Goto T, Hitomi M. A study of the improvement
effect of Miller-cycle on mean effective pressure limit for high-pressure
supercharged gasoline engines. JSAE Rev 1997;18(2):1016.
[23] Iwata N, Miyagoshi K, Nakatani S, Hitomi M. Improvement of anti-knocking
performance by supercharged Miller-cycle engine ram-pulsation effects on
anti-knocking performance by Miller-cycle. JSAE Rev 1995;16(3):3138.
[24] Wan Y, Du A. Reducing part load pumping loss and improving thermal
efciency through high compression ratio over-expanded cycle. SAE Paper;
2013 [2013-01-1744].
[25] Gottschalk W, Lezius U, Mathusall L. Investigations on the potential of a
variable Miller cycle for SI knock control. SAE Paper; 2013 [2013-01-1122].
[26] Lemmon E, McLinden M, Friend D. Thermo Physical Properties of Fluid
Systems. NIST, Gaithersburg, MD. <Http://webbook.nist.gov.entro> [accessed
on 18/10/2012].
[27] Kroenlein K, Muzny C, Kazakov A, Diky V, Chirico R, Magee J, Abdulagatov I,
Frenkel M. NIST/TRC Web Thermo-Tables (WTT), Technical report version 22011-4-Pro, NIST; 2011. <Http://wtt-pro.nist.gov/wtt-pro> [accessed on 18/
10/2012].
[28] Heywood JB. Internal combustion engine fundamentals. McGraw-Hill; 1988.
[29] Livengood JC, Wu PC. Correlation of autoignition phenomena in internal
combustion engines. In: 5th Symp (Int) on Combust; 1955. p. 34756.
[30] Soylu S. Prediction of knock limited operating conditions of a natural gas
engine. Energy Convers Manage 2005;46(1):12138.
[31] Verhelst S, Sheppard CGW. Multi-zone thermodynamic modeling of sparkignition engine combustion an overview. Energy Convers Manage
2009;50(5):132635.
[32] Diana S, Iorio B, Giglio V, Police G. The Effect of valve lift shape and timing on
air motion and mixture formation of DISI engines adopting different VVA
actuators. SAE Paper; 2001 [2001-01-3553].

Potrebbero piacerti anche