You are on page 1of 26

Accepted Manuscript

Combustion and emissions of a Euro VI heavy-duty natural gas engine using EGR
and TWC

Qiang Zhang, Zishun Xu, Menghan Li, Sidong Shao

PII: S1875-5100(15)30308-5
DOI: 10.1016/j.jngse.2015.12.015
Reference: JNGSE 1164

To appear in: Journal of Natural Gas Science and Engineering

Received Date: 3 November 2015


Revised Date: 10 December 2015
Accepted Date: 10 December 2015

Please cite this article as: Zhang, Q., Xu, Z., Li, M., Shao, S., Combustion and emissions of a Euro VI
heavy-duty natural gas engine using EGR and TWC, Journal of Natural Gas Science & Engineering
(2016), doi: 10.1016/j.jngse.2015.12.015.

This is a PDF file of an unedited manuscript that has been accepted for publication. As a service to
our customers we are providing this early version of the manuscript. The manuscript will undergo
copyediting, typesetting, and review of the resulting proof before it is published in its final form. Please
note that during the production process errors may be discovered which could affect the content, and all
legal disclaimers that apply to the journal pertain.
ACCEPTED MANUSCRIPT

Combustion and emissions of a Euro VI heavy-duty natural gas engine using EGR and TWC
Qiang Zhanga,*, Zishun Xua, Menghan Lia, Sidong Shaob
a Department of Energy and Power Engineering, Shandong University, Jinan, 250061, China
b State Key Laboratory for Reliability of Internal Combustion Engines at Weichai Power, 26 Minsheng East Street, Weifang, Shandong, 261000,
China

Abstract: The application of cooled exhaust gas recycle (EGR) and three-way catalyst (TWC) on natural gas engines has

PT
been considered to be a cost effective solution for meeting Euro VI emission standards. In the present work, systematic
experiments have been performed on a six-cylinder commercial heavy truck spark ignition (SI) natural gas engine with a
cooled EGR and TWC after-treatment system. The combustion and emission characteristics of the engine were fully

RI
analyzed. The results show that by combining stoichiometric combustion with EGR and TWC, EURO VI emission
standards for NOx, CO and NMHC, as well as smoke emission standards, could be comfortably met. Although with the

SC
current configuration, the EURO VI limits for CH4 and NH3 emissions could not be satisfied, this could possibly be settled
by the recalibration of the excess air ratio and the addition of an ammonia slip catalyst (ASC) device.
Key word: exhaust gas recirculation; natural gas engine; three way catalyst; combustion; emissions;

U
AN
1 Introduction
Natural gas engines using methane as the principal fuel are capable of yielding lower greenhouse gas emissions

compared to their equivalent diesel engines[1-3]. Recently, heavy-duty natural gas engines have been more widely applied
M

due to the demand of reducing petroleum fuel consumption and toxic emissions. However, the advent of more stringent
D

emission regulations, such as EPA 10 and EURO VI, has presented challenges to the development of natural gas engines[4].

For diesel engines, the EURO VI emission standards could be met with the adoption of SCR and EGR for NOx reduction,
TE

DOC for HC and CO removal and DPF for PM control[5,6]. In natural gas engines, gaseous natural gas combusts in a

premixed manner with extremely low PM emissions. Therefore, in order to comply with the EURO VI emission standards,
EP

stoichiometric combustion should be adopted to enable the effective use of a three-way catalyst[7].
C

As commonly accepted, the thermal load of premixed natural gas engines is relatively higher than those of their

equivalent diesel engines as a result of the slower flame propagation speed [8-12]. Meanwhile, the improvement in the
AC

thermal efficiency and emission levels is limited by their knocking tendency[13-15]. Although the reduction of HC and CO

emissions could be achieved by the combination of lean-burn combustion technology and DOC, an additional SCR device

(which is expensive) should also be equipped to meet the requirements of EURO VI emission standards[16]. With the

Abbreviations: ATDC, after top dead center; ASC, BTDC, before top dead center; CA, crank angle; CO, carbon monoxide; EAR, excess air ratio;
EGR, exhaust gas recirculation; HC, hydrocarbon; HPL, high pressure loop; HLPL, high/low-pressure EGR loop; FS, full scale; H.R.R, heat
release rate; LPL, low EGR pressure loop; MFB, mass fraction burnt; MFB0-10%, duration between the spark timing and phase angle of 10%
mass fraction burnt; MFB50%, phase angle of 50% mass fraction burnt; MFB10-90%, duration of 10-90% mass fraction burnt; NMHC,
non-methane hydrocarbons; Pmax, maximum cylinder pressure; SI, spark ignition; Tmax, maximum in-cylinder temperature; TWC, three-way
catalyst;
*
Corresponding author Tel.: +86 13791033095
E-mail address: sduzqtg@163.com(Qiang Zhang)
1
ACCEPTED MANUSCRIPT

adoption of stoichiometric combustion with TWC, the EURO VI emission standards could possibly be met[17]. However,

the durability of the key parts (piston, exhaust valves, etc.) of the engine and the durability of the TWC after treatment

system will decrease due to an increase in the thermal stress caused by higher combustion temperature and the subsequent

higher exhaust gas temperature. This problem is more prominent in the heavy-duty natural gas engines with relatively larger

bore diameters and corresponding longer distance for flame propagation[18]. Adding cooled EGR is considered to be a

PT
promising method to solve the problem of thermal load because natural gas/air mixture is diluted with inert gases and the

combustion temperature and the exhaust temperature are reduced accordingly. Additionally, the knocking tendency could

RI
also be reduced. All of the above factors contribute to the more reliable application of stoichiometric combustion with TWC

SC
on heavy duty natural gas engines[19,20].

Several studies have been focused on stoichiometric natural gas engines with EGR. Corbo et al.[21] conducted contract

U
analysis of the lean-burn and stoichiometric operations on a 9.5 L 6-cylinder SI natural gas engine under the 13 modes of

ECE-R49 procedure. According to their results, lean burn operation with an excess air ratio of 1.4 to 1.6 has a power output
AN
and efficiency similar to the stoichiometric operation with an EGR rate in the range of 10% to 20%. However, the emissions

of stoichiometric operation are much lower. Einewall et al.[22] compared the emission characteristics of stoichiometric SI
M

natural gas engine with TWC and cooled EGR (excess air ratio from 0.98 to 1.003 and EGR rate from 0% to 35%) with that
D

of lean-burn operation (excess air ratio from 1.0 to 1.6) on a 9.6 L 6-cylinder engine under engine load in the range of 0

MPa to 1.6 MPa with an engine speed of 1200 rpm. They concluded that NOx and HC emissions are significantly lower at
TE

stoichiometric operation with penalties in CO emissions. Ibarhim and Bari[23] evaluated the use of EGR on a stoichiometric

0.5 L single-cylinder SI natural gas engine with TWC under engine loads of 1.0 MPa and 1.1 MPa and an engine speed of
EP

1500 rpm. They reported that increasing the EGR rate from 0% to 12% helps to reduce the burning rate, the combustion

temperature and the possibility of the occurrence of abnormal combustion, which in turn improves the stability of engine
C

operation and reduces NOx emissions. Schöffler et al.[24] confirmed that adding EGR could result in a reduced heat release
AC

rate by changing the EGR rate from 0% to 30% on a 1.3 L single-cylinder SI stoichiometric natural gas engine with an

engine load in the range of 1.1 MPa to 2.0 MPa and an engine speed in the range of 1200 rpm to 2200 rpm. Moreover, they

also reported the benefits of cooled EGR in controlling engine thermal load. The impacts of EGR components were further

explored by Li et al.[25] on a 6.6 L 6-cylinder SI natural gas engine under an engine load of 0.7 MPa and an engine speed of

1450 rpm. They summarized the diluent and thermal effects on NOx emissions and compared the effects of different

components. They also noted that although the NOx emissions could be reduced by adding dilution gas, the decrease in the

thermal efficiency caused by adding N2 and CO2 is inevitable.


2
ACCEPTED MANUSCRIPT

Currently, three configurations are available for the introduction of EGR, i.e., high-pressure EGR loop (HPL),

low-pressure EGR loop (LPL), and high/low-pressure EGR loop (HLPL)[26]. In the HPL configuration, EGR is taken from

the exhaust manifold and introduced downstream of the compressor, which permits a quicker response of the EGR loop.

However, the instant exhaust pressure needs to be higher than the intake pressure to ensure the flow of EGR, which induces

reductions in the efficiency. In LPL configuration, EGR is taken from the after-turbine exhaust gas flow and inducted into

PT
the intake air before the compressor. Thus, the intake pressure can be higher than the exhaust pressure and higher efficiency

can be obtained. Because all of the exhaust gases go through the turbocharger, additional devices for intake pressure control

RI
should be added. The HLPL configuration, in which EGR is taken from the exhaust manifold and introduced before the

SC
compressor, acts as a compromise between HPL and LPL[27]. This configuration has a faster response than the LPL

configuration, and the applied intake pressure can be higher than the exhaust pressure. It should also be mentioned that in

U
LPL and HLPL configurations, the components of EGR gases may have some detrimental effects on the durability of the

turbocharger[28]. Consequently, for commercial heavy duty trucks, where fast response and good reliability are in great
AN
demand, HPL is more commonly used, especially in commercial applications.

This paper aims to investigate the general performance of a stoichiometric SI natural gas engine with an HPL EGR
M

system and TWC. A full set of experiments was conducted in order to thoroughly study the combustion and emissions of the
D

test engine. All of the experimental results are presented by contour maps and the relationship between different parameters

was fully analyzed. This will provide some valuable information and give some instructions for the optimization for SI
TE

stoichiometric natural gas engine operating with cooled EGR and TWC. Systematic research work on this type of engines,

which concerns such a wide range of operation conditions and combines combustion analysis with emission analysis, has
EP

not been performed before.


C

2 Experimental apparatus and test conditions


AC

2.1 Experimental apparatus


The test engine is a turbocharged SI natural gas engine for heavy-duty trucks utilizing the strategy of stoichiometric

combustion with TWC. The specifications of the test engine are listed in Table 1.

Table 1 Specifications of the test engine


Engine Parameters Specifications
Engine type Spark-ignition, 6-cylinder, in line, 4-stroke
Charged method Turbocharged and intercooled
Bore/Stroke 126 mm/155 mm
Displacement 11.6 L

3
ACCEPTED MANUSCRIPT

Rated speed 1800 rpm


Rated power 260 kW
Idle speed 700 rpm
Valve number of each cylinder 4
Compression ratio 11.5:1
Piston cavity shape Re-entrant type
After-treatment TWC

PT
Table 2 Emission test method and accuracies

RI
Parameter Measuring method Accuracy
Flame ionization detector
HC emissions ±1% FS
(FID)

SC
Nondispersive infrared
CO emissions ±1% FS
(NDIR)
Chemiluminescent detector

U
NOx emissions ±1% FS
(CLD)
AN
Nondispersive infrared
CO2 emissions ±1% FS
(NDIR)
M
D
TE
C EP
AC

Fig. 1 layout of the test bed

The intake charge is pressurized by a single-stage turbocharger with an electrically controlled waste gate; thereby, the

boost ratio could be flexibly adjusted. To mitigate the knocking tendency and reduce the thermal load of heavy-duty natural

gas engine with a large bore, cooled EGR is applied with an HPL configuration, where exhaust gas is taken from one side of

the exhaust manifold and the ratio of EGR is electronically controlled by an EGR valve. Natural gas is supplied into a

multi-hole air/fuel mixer after the compressor and intercooler mix with air, forming a natural gas/air mixture. The natural

gas/air mixture then comes into a secondary mixer to mix with the cooled EGR gas before being induced into the cylinder.
4
ACCEPTED MANUSCRIPT

The layout of the test bed is shown in Fig. 1. The test engine is coupled to an electric dynamometer and the operating

parameters, including engine speed, torque, power, exhaust gas temperature, temperatures before and after intercooler,

temperatures before and after EGR cooler and temperatures of lubricating oil and coolant, are collected by the Power-link

engine system (Type FC2000). The natural gas mass flow rate and intake air mass flow rate are measured by a Coriolis type

mass flow meter (Micromotion CMF025) with an accuracy of ±0.01 kg/h and a laminar flowmeter (ToCeiL-LFE400) with

PT
an accuracy of ±0.5%. The in-cylinder pressure is recorded by a combustion analyzer (AVL 641) and a piezoelectric

pressure transducer (Kistler 6052C) with a resolution of 0.2°CA. The HC, CO and NOx emissions before and after TWC

RI
and the concentrations of CO2 in the intake charge and exhaust gases are measured by the HORIBA MEXA-7100DEGR

SC
exhaust analyzers. The emission test method and accuracies are given in Table 2, and all of the emission results are based on

the average value of three duplicate tests.

U
2.2 Calculation of combustion parameters
AN
The average cylinder pressure of 100 consecutive cycles was acquired to eliminate the impact of cyclic variation on heat

release calculations. The corresponding net heat release rate was calculated using the following equation [29]:
M

dQnet γ dV 1 dp (1)
= p + V
dθ γ − 1 dθ γ − 1 dθ
D

Cp (2)
γ =
CV
TE

pV (3)
T=
mR
dQnet
where dθ denotes the net heat release rate, θ is the crank angle, p is the transient cylinder pressure, V is the transient
EP

working volume, Cp
is the specific heat at constant pressure, CV is the specific heat at constant volume, γ is the polytropic

coefficient and T is the transient in-cylinder temperature. In this paper, the combustion parameters, such 50% and 90%
C

combustion phase angle and the maximum in-cylinder temperature, are calculated from the net heat release rate.
AC

2.3 Test conditions


In the test process, the engine speed was varied from 1000 to 1800 rpm at intervals of 200 rpm, and BMEP was varied

from 0.15 to 1.5 MPa. All of the engine performance parameters were collected at every test point. Fig. 2 provides the spark

timing at different conditions. It can be seen that the spark timings are earlier at higher engine speeds because the

combustion duration in the crank angle is prolonged at higher engine speeds. Hence, the combustion process should be

advanced to achieve higher thermal efficiency and lower exhaust temperature. In addition, spark timing first advances and

then delays with an increase in BMEP. This is because at low loads, the duration for fuel combustion is undoubtedly shorter

5
ACCEPTED MANUSCRIPT

owing to the smaller quantity of induced fuel. However, as the engine load increases to a certain extent, the advance of

spark timing will be constrained by the maximum pressure and the NOx emissions, leading to a relatively retarded spark

timing used at high loads. Moreover, spark timing is retarded in regions with low engine loads and low engine speeds in

order to maintain exhaust temperature high enough for the effective operation of TWC.

Engine spark timing

PT
1.5
°ATDC
-31.0
1.2 -27.5

RI
-24.0
-20.5
-17.0
BMEP (MPa)

0.9 -13.5
-10.0

SC
-6.5
-3.0
0.6

U
0.3
AN
1000 1200 1400 1600 1800
Engine speed(rpm)

Fig. 2 Contour map of spark timing


M

Fig. 3 depicts the control strategy of the EGR rate. As illustrated, the EGR rate shows a generally increasing trend with

an increase in the engine load. This is because at lower engine loads, the combustion temperature and the tendency of NOx
D

formation are relatively lower, so adopting lower EGR rates is beneficial for the thermal efficiency and emission levels. At
TE

higher loads, the tendency of knocking and the maximum combustion pressure should be controlled by the application of

higher EGR rates. In the present work, the emission results are compared with EURO VI standards for transient cycles
EP

because this study is aimed at the optimization of parameters for on-road applications, where operation condition varies

frequently. However, some transient conditions, e.g., acceleration and deceleration, could not be taken into consideration.
C
AC

6
ACCEPTED MANUSCRIPT

EGR rate
1.5
%
0.5
3.0
1.2
5.5
7.9

BMEP (MPa)
10.4
12.9
0.9
15.4
17.8
20.3

PT
0.6

RI
0.3

1000 1200 1400 1600 1800


Engine speed(rpm)

SC
Fig. 3 Contour map of EGR rate

3 Test results and analysis

U
3.1 Effects of EGR and excess air ratio on engine performance
AN
As shown by Fig. 4, a milder increase in the cylinder pressure can be observed when more EGR is added. This could

be a result of the slower flame propagation speed caused by the dilution effects of inertial gases in EGR, which could also
M

lead to prolonged flame development, extended heat release duration and reduced heat release in each crank angle. It is also

presented that although the spark timing is advanced at higher EGR rates to gain better thermal efficiency, the peak value of
D

heat release is still lower and the end of heat release is still retarded as EGR increases.
TE

8
Speed 1300rpm,BMEP1.5MPa, EAR=1.01 Speed 1300rpm,BMEP1.5MPa,EAR=1.01
120
EGR0.5%
EGR0.5% EGR5.0%
100 EGR5.0% 6
EGR10.0%
EP

EGR10.0% EGR15.0%
Cylinder pressure(bar)

EGR15.0%
H.R.R(%/°CA)

EGR20.0%
80 EGR20.0%
4

60
C

40 2
AC

20
0
0
-30 -15 0 15 30 45 60 -30 -15 0 15 30 45 60

crank angle (°CA) crank angle (°CA)

(a) (b)
Fig. 4 Cylinder pressure (a) and heat release rate (b) at different EGR rates

Fig. 5 demonstrates the effects of EGR on maximum in-cylinder temperature and engine-out NOx emissions. At both

BMEPs of 1.5 MPa and 1.05 MPa, the possibility of oxidation of nitrogen is evidently reduced owing to the reduced flame

temperature resulting from the dilution effects of EGR. Further, as indicated by the data in Fig. 5, engine-out NOx

7
ACCEPTED MANUSCRIPT

emissions are higher at the higher engine load of 1.5 MPa when the EGR rate is higher than 5% as a result of the higher

in-cylinder temperature. However, when the EGR rate is reduced to 0.5%, NOx emissions are lower at a BMEP of 1.5 MPa.

This is most likely because at higher BMEP values, the back pressure is relatively higher, leading to poorer scavenging and,

subsequently, larger proportion of residual gas and lower engine-out NOx emissions.

PT
2700
Speed 1300rpm, BMEP1.5MPa, EAR1.01 25
Maximum in-cylinder temperature(K)

2600 Tmax1.5
Tmax1.05 20

RI
NOx1.5
2500

NOx(g/kW·h)
NOx1.05
15

2400

SC
10
2300

5
2200

U
0
0 5 10 15 20
AN
EGR rate (%)

Fig. 5 Effects of EGR rate on maximum in-cylinder temperature


6.0 2300
Speed 1300rpm,BMEP 1.5MPa,Spark timing -16°ATDC
M

5.5 2250 Tmax(K)


NOx
Pmax
5.0 Tmax 2200
D
NOx(g/kW·h)

4.5
2150
TE

4.0
Pmax (bar)

105
3.5

100
3.0
EP

EGR rate 17.4%,


2.5 95
0.98 1.00 1.02 1.04 1.06
Excess air ratio
C

Fig. 6 Effects of excess air ratio on NOx emissions, maximum cylinder temperature and maximum cylinder pressure

Fig. 6 offers the effects of excess air ratio on engine-out NOx emissions. It can be seen that the maximum cylinder
AC

pressure increases while the maximum cylinder temperature decreases with an increasing excess air ratio. Although the

amplitudes of variation for maximum pressure and maximum temperature appear to be small, engine-out NOx emissions

increase significantly when the induced mixture becomes leaner. This is because when excess air ratio changes from 0.98 to

1.05, although the maximum in-cylinder temperature decreases, the NOx emissions increase rapidly due to an increase in

the oxygen concentration. It can also be noticed that in this region, NOx emissions tend to be quite sensitive to the excess

air ratio[30]. Thus, the accurate control of both the excess air ratio and EGR rate are crucial for NOx reduction.

3.2 Combustion process analysis


8
ACCEPTED MANUSCRIPT

Fig. 7 shows the maximum pressure and the corresponding angle of occurrence. As manifested by Fig. 7a, the

maximum cylinder pressure increases with increased engine load owing to the increased fuel quantity and the increase in a

comparably small magnitude in the engine speed because the effects of decreasing heat release in every crank angle could

be offset by the adjustment of spark timing. Although the EGR rate first increases and then decreases with an increase in the

engine load, as shown by Fig. 7b, delayed occurrence of the maximum pressure appears in regions with low speeds and low

PT
loads, and the distribution of the phase angle of maximum pressure is opposite to that of spark timing while distinctive from

that of the EGR rate, implying that spark timing is the more dominant influencing factor for the phase angle of maximum

RI
cylinder pressure.

SC
Max cylinder pressure Phase angle of maximum cylinder pressure
1.5 1.5
bar °ATDC
14.0 7.4
29.8

U
1.2 1.2 9.5
45.5 11.7
61.3 13.8
BMEP (MPa)

77.0
AN
BMEP (MPa)

16.0
92.8 18.1
0.9 0.9
108.5 20.2
124.3 22.4
140.0 24.5
0.6
M

0.6

0.3 0.3
D

1000 1200 1400 1600 1800 1000 1200 1400 1600 1800
TE

Engine speed(rpm) Engine speed(rpm)

(a) (b)
Fig. 7 Contour map of maximum cylinder pressure (a) and the corresponding phase angle (b)
EP

Fig. 8 shows the maximum pressure rise and the corresponding phase angle. It can be observed in Fig. 8a that the

variation in the maximum pressure rise rate with the engine speed is consistent with that of the maximum cylinder pressure
C

at lower engine loads, while it exhibits a different distribution at higher engine loads. This is because at higher engine loads,
AC

the EGR rate, which can deter the increase in the pressure rise rate, reaches a higher value and plays a more important role,

which leads to a decreasing trend of maximum pressure rise rate with increasing speed. However, for the phase angle of

maximum pressure rise rate (Fig. 8b), spark timing remains the principal influencing factor.

9
ACCEPTED MANUSCRIPT

Max pressure rise rate Phase angle of max pressure rise rate
1.5 1.5
bar/°CA °ATDC
0.4 0.0
1.2 2.3
1.2 1.2 4.5
1.9
2.7 6.8

BMEP (MPa)
BMEP (MPa)

3.4 9.0
4.2 11.3
0.9 0.9
5.0 13.5
5.7 15.8

PT
6.5 18.0
0.6 0.6

RI
0.3 0.3

1000 1200 1400 1600 1800 1000 1200 1400 1600 1800

SC
Engine speed(rpm) Engine speed(rpm)

(a) (b)
Fig. 8 Contour map of maximum pressure rise rate (a) and the corresponding phase angle (b)

U
Fig. 9a shows the distribution of MFB 0-10% (duration between the spark timing and the 10% mass fraction burnt, i.e.,
AN
the flame development duration) at different operating points. As is known, the flame development duration is contingent

upon the flow movement adjacent to the spark plug, the ignition energy and the temperature and pressure at the time of
M

ignition. At engine loads lower than 0.3 MPa, although the lower in-cylinder temperature may have some negative effects

on the propagation of the initial flame, the flame development duration has relatively smaller values. This is because in
D

these regions, the in-cylinder pressure and EGR rate are also at a very low level, leading to relatively smaller energies
TE

necessary for ignition and enhanced ignition quality. At engine speeds lower than 1100 rpm, the flame development duration

is also very short due to the smaller energy required for ignition and the weaker flow movement near the spark plug. It
EP

should be observed that the longest flame development duration appears at high engine speeds with medium engine loads.

This is because in these regions, the spark timings are the most advanced, causing lower pressure and temperature at the
C

ignition period, which are unfavorable for the flame propagation at initial stages.
AC

10
ACCEPTED MANUSCRIPT

MFB0-10%
1.5
°CA
12.2
1.2 14.1
16.0
17.9
19.8

BMEP (MPa)
0.9 21.7
23.6
25.5
27.5

PT
0.6

0.3

RI
1000 1200 1400 1600 1800
Engine speed(rpm)

SC
(a)
MFB50% MFB10-90%
1.5 1.5
°ATDC °CA
3.0 14.4

U
1.2 5.0 1.2 17.0
7.0 19.5
9.0 22.1
AN
11.0 24.6
BMEP (MPa)
BMEP (MPa)

0.9 13.0 0.9 27.2


15.0 29.8
17.0 32.3
19.0 34.9
M

0.6 0.6

0.3
D

0.3

1000 1200 1400 1600 1800 1000 1200 1400 1600 1800
TE

Engine speed(rpm) Engine speed(rpm)


(b) (c)
Fig. 9 Contour map of MFB 0-10% (a), MFB-50% (b) and MFB 10-90% (c)
EP

Fig. 9b and Fig. 9c offer the phase angle of 50% mass fraction burnt and duration of 10-90% mass fraction burnt. MFB

-50% is excessively delayed in regions with low engine speeds and low engine loads, which is in accordance with the
C

sparking timing and the phase angle of maximum pressure rise rate. In regions with relatively higher engine loads and
AC

engine speeds, the phase angle of MFB 50% is within the ideal range of 3 to 10°ATDC; thus, considerable efficiency can be

guaranteed. It is worth noting that MFB 50% and the phase angle of maximum pressure rise rate exhibit similar distribution

because both parameters are strongly affected by the sparking timing. As can be observed in Fig. 9c, MBF 10-90% increases

with the engine speed as the actual time for every crank angle decreases at higher engine speeds. Additionally, MBF 10-90%

is apparently longer in regions with higher engine speeds and medium engine loads, which is attributed to the combined

effects of the EGR rate and ignition reliability.

The maximum heat release rate and the corresponding phase angle are shown in Fig. 10. As shown in Fig. 10a, the

11
ACCEPTED MANUSCRIPT

maximum heat release rate seems to be strongly influenced by the engine load. This can be explained by the remarkably

increased fuel quantity and the subsequent increased heat release at higher engine loads. The phase angle of heat release rate,

however, is predominantly determined by the spark timing, which is why its distribution and changing trends are in

accordance with those of the spark timing.

Max H.R.R Phase angle of max H.R.R


1.5

PT
1.5
J/°CA °ATDC
62 0.5
1.2 116 1.2 2.9
171

RI
5.3
225 7.6
BMEP (MPa)

279 10.0

BMEP (MPa)
0.9 333 0.9 12.4
388 14.8

SC
442 17.1
496 19.5
0.6 0.6

U
0.3 0.3
AN
1000 1200 1400 1600 1800 1000 1200 1400 1600 1800
Engine speed(rpm) Engine speed(rpm)
(a) (b)
M

Fig. 10 Contour map of maximum heat release rate (a) and the corresponding phase angle (b)
IMEP COV
1.5
%
D

0.44
1.2 0.81
TE

1.18
1.55
1.92
BMEP (MPa)

0.9 2.30
2.67
3.04
EP

3.41
0.6
C

0.3
AC

1000 1200 1400 1600 1800


Engine speed(rpm)

Fig. 11 Contour map of cyclic variation for IMEP


The COV of IMEP is considered as a key parameter for spark-ignition natural gas engine. As previously investigated,

the COV of IMEP is an important indicator for the evaluation of engine performance because it is closely related to

emissions, fuel economy and vehicle drivability. Fig. 11 provides the COV of IMEP for the test engine. As shown, the COV

of IMEP is lower than 2% in all regions, suggesting that high combustion stability can be ensured in most operating

conditions[31]. The COV of IMEP is higher when operating at lower engine loads as a result of higher fraction of residual

12
ACCEPTED MANUSCRIPT

exhaust gas, lower mass flow rate and larger diversities in the cycle-to-cycle mixture quality. At medium and high engine

loads, the COV for IMEP is generally smaller than 1%, implying better combustion stability; further, at engine speeds

higher than 1300 rpm, the COV of IMEP at high engine loads is slightly higher than that at medium loads due to an increase

in the EGR rate. It should also be noticed that the largest cyclic variation exists in regions with an engine speed of

approximately 1800 rpm and an engine load in the range of 0.3 to 0.6 MPa, caused by the unstable combustion resulting

PT
from earlier spark timing.

3.2 Emission characteristics

RI
3.2.1 NOx emissions
Fig. 12 shows the engine-out and tailpipe NOx emissions. It can be noted that the engine-out NOx emissions are higher

SC
at engine loads lower than 0.45 MPa, where engine-out NOx emissions exhibit a decreasing trend with an increase in the

engine load. As widely accepted, NOx emissions increase with an increase in the combustion temperature and the prolonged

U
residence time in high-temperature regions[32]. Furthermore, by adding EGR, the heat capacity of the intake charge and the
AN
tendency for thermal dissociation of CO2 are increased[33,34], both leading to decreased flame temperature and,

subsequently, decreased NOx generation. Additionally, the oxygen concentration is reduced when more EGR is introduced,
M

resulting in less possibility of reactions between N2 and O2, namely, lower NOx emissions. According to the test results, it

can be clearly noticed that the distribution of engine-out NOx emissions is more dependent on the EGR rate than the other
D

two influencing factors. Moreover, as shown by Fig. 12b, tailpipe NOx emissions are lower than 0.14 g/kW·h in the whole
TE

range of the engine speed. Therefore, for a stoichiometric natural gas engine with EGR and TWC, the Euro VI emission

standard for NOx emissions (0.46 g/kW·h) can be achieved without difficulties.
EP

Engine-out BSNOx Tailpipe BSNOx


1.5 1.5
g/kW·h g/kW·h
C

1.9 0.01
1.2 4.9 1.2 0.03
7.9 0.04
AC

10.9 0.06
BMEP (MPa)
BMEP (MPa)

13.9 0.07
0.9 16.8 0.9 0.09
19.8 0.11
22.8 0.12
25.8 0.14
0.6 0.6

0.3 0.3

1000 1200 1400 1600 1800 1000 1200 1400 1600 1800
Engine speed(rpm) Engine speed(rpm)

(a) (b)
Fig. 12 Contour map of engine-out NOx emissions (a) and tailpipe NOx emissions (b)
13
ACCEPTED MANUSCRIPT

Fig. 13 depicts the conversion efficiency of NOx emissions. In the whole range of operation, the conversion efficiency

of NOx emissions is higher than 99%, indicating that the CO and HC concentration in the exhaust gas is sufficient for NOx

reduction. Furthermore, as shown by Fig. 13b, the inlet temperature of TWC is in the range of 386 to 603°C (Fig. 13b),

which is enough for the effective operation of TWC.

NOx conversion efficiency TWC inlet temperature

PT
1.5 1.5
% °C
97.8 386
1.2 98.0 1.2 413
98.3 440

RI
98.5 467
BMEP (MPa)

98.8 495

BMEP (MPa)
0.9 99.1 0.9 522
99.3 549

SC
99.6 576
99.8 603
0.6 0.6

U
0.3 0.3
AN
1000 1200 1400 1600 1800 1000 1200 1400 1600 1800
Engine speed(rpm) Engine speed(rpm)

(a) (b)
M

Fig. 13 Contour map of NOx conversion efficiency (a) and TWC inlet temperature (b)

3.2.2 HC emissions
D

Fig. 14 demonstrates the engine-out THC emissions and the CH4 to THC ratio. As widely accepted, engine-out THC
TE

emissions are mainly associated with unburned fuel and partial combustion[35,36]. When operating at lower engine loads

with lower in-cylinder temperature, the possibility of bulk quenching in the late combustion stages increases, which could

result in increased THC emissions [37]. Moreover, at higher engine speeds, the phenomena of local extinction caused by
EP

higher intensity of turbulent movement and less time available for the combustion event could further increase HC
C

emissions. All of the above-mentioned reasons help to explain the distribution of THC emissions in Fig. 14a, where

engine-out THC emissions are relatively higher at regions with an engine speed higher than 1600 rpm and an engine load
AC

lower than 0.3 MPa.

With an increase in the engine load, the quantity of intake charge is significantly increased, providing more fuel for the

combustion process. Thereby, the combustion temperature, together with the temperatures of piston and cylinder, increases

as the engine load increases, thus reducing the occurrence of bulk quenching in near wall areas and enhancing the late

combustion and post-oxidation events, all of which contribute to the reduction of HC emissions. Moreover, although the

EGR rate, which has the potential to cause an increase in the HC emissions[38], is evidently higher at higher engine loads,

the HC emissions are more affected by the effects of crevices and bulk quenching, thus exhibiting a decreasing trend with
14
ACCEPTED MANUSCRIPT

an increase in the engine load. In addition, with earlier spark timing, the occurrence of retarded combustion is reduced. As a

consequence, a slight reduction with engine speed can be observed in regions with higher engine speeds and higher engine

loads.

As shown in Fig. 14 b, CH4 is the dominant component of HC emissions, accounting for over 80% of the whole THC

emissions in most test conditions. CH4 escaped during the valve overlapping period and the unburned fuel resulting from

PT
the crevice effects, together with the insufficient in-cylinder combustion, are all sources of CH4 emissions[39]. The regions

with low CH4-to-THC ratios mainly distribute in regions with low engine loads, which is attributed to increased occurrences

RI
of partial combustion. In the meantime, when operating at lower engine loads, the boost ratio is relatively lower with

SC
smaller opening of the throttle valve, suppressing the charge exchanging process and reducing the unburned fuel going

directly into the exhaust system, which is another possible reason for this phenomena. It can also be noticed that high a

U
CH4-to-THC ratio distributes in regions with engine speeds of approximately 1400 rpm and medium and high engine loads,

suggesting that less partial combustion occurs at these conditions.


AN
Engine-out BSTHC CH4/THC(%)
1.5 1.5
g/kW·h %
M

1.8 79.2
1.2 2.9 1.2 80.3
4.0 81.3
5.2 82.4
D

6.3 83.5
BMEP (MPa)
BMEP (MPa)

0.9 7.4 0.9 84.5


8.6 85.6
TE

9.7 86.6
10.9 87.7
0.6 0.6
EP

0.3 0.3

1000 1200 1400 1600 1800 1000 1200 1400 1600 1800
Engine speed(rpm) Engine speed(rpm)
C

(a) (b)
AC

Fig. 14 Contour map of engine out THC emissions (a) and CH4-to-THC ratio (b)

Methane, which is a predominant component of THC emissions, has a high activation energy and is difficult to be

catalyzed[40]. As manifested by Fig. 15a, tailpipe CH4 emissions are highest in regions with high engine speeds and low

engine loads, which is in accordance with the distribution of high engine-out THC emissions. With an increase in the engine

load, tailpipe CH4 emissions show a trend of first decreasing and then increasing, while with an increase in the engine speed,

a generally increasing trend is illustrated. As seen in Fig. 15b, tailpipe NMHC emissions are lower in regions with low

engine speeds and higher in regions with higher engine speeds due to the effects of engine-out THC emissions and the

corresponding conversion efficiency. Generally, tailpipe NMHC emissions are lower than the EURO VI emission standard
15
ACCEPTED MANUSCRIPT

of 0.16 g/kW·h in the whole operating range.

Tailpipe BSCH4 Tailpipe BSNMHC


15 1.5
g/kW·h g/kW·h
0.05 0.00
12 0.37 1.2 0.01
0.69 0.02
1.00 0.03
BMEP (MPa)

1.32 0.04

BMEP (MPa)

PT
9 1.64 0.9 0.05
1.96 0.06
2.27 0.07
2.59 0.08
6 0.6

RI
3 0.3

SC
1000 1200 1400 1600 1800 1000 1200 1400 1600 1800
Engine speed(rpm) Engine speed(rpm)

(a) (b)

U
Fig. 15 Contour map of tailpipe CH4 emissions (a) and tailpipe NMHC emissions (b)
AN
As shown in Fig. 16a, the conversion efficiency of THC emissions is higher (approximately 80%) in regions with

engine loads in the range of 0.3 to 0.9 MPa, while in some regions with high engine loads, the conversion efficiency of THC
M

can be lower than 60%. As observed in Fig. 16a and Fig. 16b, when the engine is operating at high engine loads, the

engine-out oxygen concentration is obviously lower. Hence, the conversion efficiency of THC, which acts as the reducer in
D

TWC, is correspondingly lower due to the lack of oxidation. When the engine is operating at medium engine loads, the
TE

engine-out O2 has the highest value, leading to high conversion efficiency of THC emissions. Consequently, it can be

deduced that the conversion efficiency of THC emissions, which act as the reducer in TWC, is determined not only by the
EP

characteristics of catalyst but also by the oxygen concentration in the exhaust gas, indicating that the problem of high CH4

emissions could be mitigated by raising excess air ratio.


C
AC

16
ACCEPTED MANUSCRIPT

THC conversion efficiency Engine-out O2


1.5 1.5
% %
39.4 0.302
1.2 45.8 1.2 0.355
52.2 0.408

BMEP (MPa)
58.6 0.461
65.0 0.514
BMEP (MPa)

0.9 71.3 0.9 0.567


77.7 0.620
84.1 0.673

PT
90.5 0.726
0.6 0.6

RI
0.3 0.3

1000 1200 1400 1600 1800 1000 1200 1400 1600 1800
Engine speed(rpm) Engine speed(rpm)

SC
(a) (b)
Fig. 16 Contour map of THC conversion efficiency (a) and engine-out O2 (b)

U
3.2.3 CO emissions
Fig. 17 shows the engine-out and tailpipe CO emissions. For stoichiometric combustion, engine-out CO emissions,
AN
resulting from the lack of oxygen, are considerably high due to the low oxygen concentration and the formation of a local

over-rich mixture. In regions with low engine loads, increased engine-out CO emissions are incurred by the lower
M

combustion temperature and the subsequent suppressed oxidation of CO, while at medium and high engine loads,
D

engine-out CO emissions are relatively lower and change slightly with engine load. Moreover, when engine speed increases,
TE

the time available for CO oxidation decreases. Hence, increase in CO emissions is induced due to incomplete combustion. It

is worth mentioning that with an increase in the engine speed, the flow movement of natural gas and cooled EGR is

intensified, enhancing the mixing of natural gas and intake air and reducing the formation of local over-rich areas. In
EP

addition, the distribution of CO emissions is related to the EGR rate. In high-speed, high-load regions with high EGR rates,
C

increases in engine-out CO emissions can be clearly noticed.

As presented in Fig. 17b, at the whole test conditions, tailpipe CO emissions are lower than the EURO VI emissions
AC

standard of 4.0 g/kW·h, and under most conditions, CO emissions are lower than 1.5 g/kW·h. High CO emissions distribute

in regions with high engine loads as well as in regions with high engine speeds and low engine loads. The distribution of

CO emissions is primarily determined by the conversion efficiency. Fig. 17 c presents the conversion efficiency of CO

emissions. CO emissions are easily oxidized in TWC and the conversion efficiency is strongly influenced by the oxygen

concentration. When operating in regions with high engine loads, the engine-out oxygen concertation is relatively lower,

providing insufficient oxidant for the conversion of CO. However, because the conversion of CO is much easier than that of

CH4, the overall conversion efficiency is higher than that of THC emissions. When operating in regions with low engine
17
ACCEPTED MANUSCRIPT

loads, conversion efficiency appears to be higher owing to the correspondingly higher engine out oxygen concertation.

Engine-out BSCO Tailpipe BSCO


1.5 1.5
g/kW·h g/kW·h
10.5 0.4
1.2 13.4 1.2 0.9
16.3 1.3
19.2 1.8
22.1 2.2
BMEP (MPa)

BMEP (MPa)

PT
0.9 25.0 0.9 2.6
27.9 3.1
30.8 3.5
33.7 4.0
0.6 0.6

RI
0.3 0.3

SC
1000 1200 1400 1600 1800 1000 1200 1400 1600 1800
Engine speed(rpm) Engine speed(rpm)
(a) (b)

U
CO conversion efficiency
1.5
%
AN
69.0
1.2 72.6
76.2
79.8
M

83.3
BMEP (MPa)

0.9 86.9
90.5
94.1
97.7
D

0.6
TE

0.3

1000 1200 1400 1600 1800


EP

Engine speed(rpm)
(c)
Fig. 17 Contour map of engine-out CO emissions (a), tailpipe CO emissions (b), and CO conversion efficiency (c)
C

3.2.4 NH3 and smoke emissions


AC

18
ACCEPTED MANUSCRIPT

Tailpipe NH3
1.5
ppm
50
1.2 110
170
230
290

BMEP (MPa)
0.9 350
410
470

PT
530
0.6

RI
0.3

1000 1200 1400 1600 1800


Engine speed(rpm)

SC
Fig. 18 Contour map of tailpipe NH3 emissions

In the TWC, NO will react with H2, producing NH3 and H2O. Therefore, the emission of NH3 is unavoidable. As

U
shown by Fig. 18, NH3 emissions are highest in regions with low engine speeds and high engine loads and also relatively
AN
higher in regions with low engine loads and engine speeds higher than 1200 rpm. This is because in regions with low engine

speeds and high engine loads, engine-out NOx emissions are relatively higher with low CO and THC emissions, increasing
M

the possibility of the reaction with H2 and resulting in high NH3 emissions; in low load regions with engine speeds higher

than 1200 rpm, NOx emissions have the highest value, which will undoubtedly contribute to the generation of NH3. It
D

should be noted that tailpipe NH3 emissions exceed the EURO VI emission limit (10 ppm). Adoption of an Ammonia Slip
TE

Catalyst (ASC) may be a solution for this.

For natural gas fueled engines, methane, which has no carbon-carbon bonds and possesses much lower tendency for
EP

soot generation than diesel, is the primary constituent of the fuel burnt in the cylinder[41,42]. Thus, the formation of soot is

mainly incurred by the pyrolysis of the evaporated lubricating oil, which contains more carbon and is consequently more
C

likely to produce soot precursor. In this case, when more EGR is added, more soot may be brought into the cylinder, causing
AC

negative effects on the smoke emissions. Hence, it can be concluded that higher smoke opacity appears at medium load

conditions, which could be attributed to the combined effects of EGR rate and spark timing.

19
ACCEPTED MANUSCRIPT

Smoke
1.5
FSN
0.000
1.2 0.003
0.005

BMEP (MPa)
0.008
0.010
0.9
0.013
0.015

PT
0.018
0.6

RI
0.3

1000 1200 1400 1600 1800


Engine speed(rpm)

SC
Fig. 19 Contour map of smoke opacity
3.3 Fuel economy

U
Fig. 20 illustrates the distribution of thermal efficiency and CO2 emissions. It can be seen that brake thermal efficiency
AN
increases, while the CO2 emissions decrease with increasing engine load attributed to the improved combustion process and

the increase in mechanical efficiency. With an increase in the engine speed, however, an opposite trend can be clearly seen.
M

This can be explained by the shortened time for combustion and increased mechanical losses.

Brake thermal efficiency(%) Tailpipe BSCO2(g/kW·h)


1.5 1.5
D

14.5 546
17.4 643
20.3 740
TE

1.2 23.2 1.2 836


26.1 933
BMEP (MPa)

BMEP (MPa)

29.0 1030
31.9 1127
0.9 34.8 0.9 1223
EP

37.7 1320

0.6 0.6
C

0.3 0.3
AC

1000 1200 1400 1600 1800 1000 1200 1400 1600 1800
Engine speed(rpm) Engine speed(rpm)

(a) (b)
Fig. 20 Contour map of brake thermal efficiency (a) and tailpipe CO2 emissions (b)

4. Conclusions
Aiming at achieving Euro VI emission standards, the combustion process and the emissions of a natural gas engine

using cooled EGR and TWC were investigated using contour maps. The following main conclusions have been drawn:

1. Adding EGR could significantly reduce maximum cylinder pressure, in-cylinder temperature and thus engine-out NOx

20
ACCEPTED MANUSCRIPT

emissions. However, the excess air ratio should be very accurately controlled. This is because under near stoichiometric

conditions, only a slight increase in the excess air ratio could induce an obvious increase in the engine-out NOx emissions.

2. Maximum cylinder pressure and maximum heat release increase with an increased engine load due to the increased fuel

quantity and occur later in regions with low speeds and low loads. Although the COV for IMEP deteriorates in regions with

more EGR or relatively advanced spark timing, it is lower than 2% in most conditions, indicating that good combustion

PT
stability could be achieved.

3. Engine-out NOx emissions are higher at engine loads lower than 0.45 MPa, and the high-emission regions of tailpipe

RI
NOx emissions are in accordance with engine-out NOx emissions. Meanwhile, NOx emissions after TWC are generally

SC
lower than 0.14 g/kW·h, which is much lower than the EURO VI emission standard.

4. Tailpipe CH4 emissions are higher in regions with high speeds and low loads, and tailpipe NMHC emissions are higher in

U
regions with higher engine speeds. Under all operating conditions, tailpipe NMHC emissions can comply with the EURO

VI emissions standard, whereas tailpipe CH4 emissions should be further reduced by improving the excess air ratio
AN
calibration.

5. In regions with low engine loads, higher CO emissions are induced by the lower combustion temperature, while at
M

medium and high engine loads, CO emissions are relatively lower and change slightly with engine load. In most regions, the
D

conversion efficiency for CO emissions is higher than 90%; thus, the tailpipe CO emissions can meet the demand of the

EURO VI emission standard.


TE

6. NH3 emissions are highest in regions with low engine speeds and high engine loads and also relatively higher in regions

with low engine loads and engine speeds higher than 1200 rpm, while smoke opacity seems to be excessively low under all
EP

conditions. It is worth mentioning that NH3 emissions tend to be higher than the EURO VI emission standard, suggesting

that an additional ASC device is essential.


C

7. Brake thermal efficiency is higher at higher engine loads and lower at high engine speeds. The changing trend of CO2
AC

emissions is opposite to that of brake thermal efficiency.

Acknowledgments
This work was supported by the National High-tech R&D Program of China (863 Program 2014AA052802).

References:
[1] Weaver CS. Natural gas vehicles-a review of the state of the art. SAE technical Paper 892133; 1989.
[2] Maclean HL, Lave LB. Evaluating automobile fuel/propulsion system technologies. Progr Eneryg Combust Sci 2003;
29:1-69.
[3] Shiga S, Ozone S, Machacon HTC, Karasawa T, Nakamura H. A study of the combustion and emission Characteristics
of Compressed natural gas direct injection stratified combustion using a rapid compression machine. Combustion and flame
2002; 129:1-10.
21
ACCEPTED MANUSCRIPT

[4] Dimopoulos P., Bach C., Solitic P, Boulouchos K. Hydrogen-Natural gas blends fuelling passenger car engines:
Combustion, emissions and well-to-wheels assessment. International journal of hydrogen energy 2008; 33(23): 7224-7236,
2008. doi:10.1016/j.ijhydene. 2008.07.012.
[5] Vressner A, Gabrielsson P, Gekas I, Senar E. Meeting the EURO VI NOx Emission Legislation using a EURO IV base
engine and a SCR/ASC/DOC/DPF configuration in the world harmonized transient cycle. SAE technical papers
2010-01-1216; 2010.
[6] Squaiella LLF, Martins CA, Lacava PT. Strategies for emission control in diesel engine to meet Euro VI. Fuel 2013; 104:
183-193.
[7] Bielaczyc P, Woodburn J, Szczotka A. An assessment of regulated emissions and CO2 emissions from a European

PT
light-duty CNG-fueled vehicle in the context of Euro 6 emissions regulations. Applied Energy 2014; 117: 134–141.
[8] Rousseau S, Lemoult B, Tazerout M. Combustion characteristics of natural gas in a lean burn spark-ignition engine.
Proc Inst Mech Eng Part D J Automob Eng 1999; 213:481-489.
[9] Chiodi M, Berner HJ, Bargende M. Investigation on mixture formation and combustion process in a CNG-engine by

RI
using a fast response 3D-CFD-simulation. SAE Paper 2004; 2004-01-3004.
[10] Ehsan M. Effect of spark advance on a gas run automotive spark ignition engine. Journal of Chemical Engineering
2006, 24(1):42-49.

SC
[11] Davis SG, Law CK. Determination of and Fuel Structure Effects on Laminar Flame Speeds of C1 to C8 Hydrocarbons. )
Determination of and fuel structure effects on laminar flame speeds of C1 to C8 hydrocarbons. Combustion Science and
Technology 1998, 140:1-6, 427-449. DOI: 10.1080/00102209808915781
[12] Patricia D, Herve LG, Roda B, Olvier H, Pierrealexandre G, Alexander K and Frederique BL. Measurements of

U
laminar flame velocity for components of natural gas. Energy & fuels 2011; 25: 3875-3884.
[13] Cho HM, He BQ. Spark ignition natural gas engines—A review. Energy Conversion and Management 2007; 48: 608–
618.
AN
[14] Gersen S, Essena MV, Dijka GV, Levinskya H. Physicochemical effects of varying fuel composition on knock
characteristics of natural gas mixtures 2014; 161(10): 2729-2737.
[15] Saikalya K, Correb OL, Rahmounia C, Truffetc L. Preventive knock protection technique for stationary SI engines
fueled by natural gas. Fuel Processing Technology 2010; 91(6): 641-652.
M

[16] Kaiadi M. Diluted operation of a heavy-duty natural gas engine. Doctoral Thesis 2011. Lund University, lund Sweden.
[17] Yan BW, Yao MF, Mao B, Li YZ, Qin YF. A Comparative Study on the Fuel Economy Improvement of a Natural Gas
SI Engine at the Lean Burn and the Stoichiometric Operation both with EGR under the Premise of Meeting EU6 Emission
D

Legislation. SAE paper 2015-01-1958; 2015.


[18] Cheenkachorn K, Poompipatpong C, Choi GH. Performance and emissions of a heavy-duty diesel engine fueled with
diesel and LNG (liquid natural gas). Energy 2013; 55. 52-57.
TE

[19] Ibrahim A, Bari S. A comparison between EGR and lean burn strategies employed in a natural gas SI engine using a
two-zone combustion model. Energy Conversion and Management 2009; 50(12):3129-3139.
[20] Li, W., Liu, Z., Wang, Z., Li, C. et al., "Effect of CO2, N2, and Ar on Combustion and Exhaust Emissions Performance
in a Stoichiometric Natural Gas Engine," SAE Technical Paper 2014-01-2693; 2014. doi:10.4271/2014-01-2693.
EP

[21] Corbo P, Gambino M, Lannaccone S, Unich A. Comparison between lean-burn and stoichiometric technologies for
CNG heavy duty engines. SAE paper 950057; 1995.
[22] Einewall P, Tunestal P, Johansson B. Lean Burn Natural Gas Operation vs. Stoichiometric Operation with EGR and a
Three Way Catalyst. SAE paper 2005-01-0250; 2005.
C

[23] Ibrahim A, Bari S. An experimental investigation on the use of EGR in a supercharged natural gas SI engine. Fuel 2010;
89:1721-1730.
AC

[24] Schöffler T, Hoffmann K, Daimler AG. Stoichiometric natural gas combustion in a single cylinder SI engine and
impact of charge dilution by means of EGR. SAE technical papers 2013-24-0113; 2013.
[25] Li WF, Liu GC, Wang ZS, Xu Y. Experimental investigation of the thermal and diluent effects of EGR components on
combustion and NOx emissions of a turbocharged natural gas SI engine. Energy Conversion and Management 2014;
88:1041-1050.
[26] Heavy-duty waste hauler with chemically correct natural gas engine diluted with EGR and using a Three-Way catalyst.
T. Reppert, J. Chiu. Subcontract report NREL/SR-540-38222.
[27] Simio LD, Gambino M, Lannaccone S. A study of different EGR routes on a heavy duty stoichiometric natural gas
engine. SAE technical paper 2009-24-0096; 2009.
[28] Reifarth S, Angstrom HE. Transient EGR in a high-speed DI diesel engine for a set of different EGR-routings. SAE
Technical paper 2010-01-1271; 2010.
[29] Rakopoulos CD, Rakopoulos DC, Giakoumis EG, Dimaratos AM. Investigation of the combustion of neat cottonseed
oil or its neat bio-diesel in a HSDI diesel engine by experimental heat release and statistical analyses. Fuel 2010; 89(12),
3814–3826.
22
ACCEPTED MANUSCRIPT

[30] Lean-burn or rich-burn? http://www.gallois.be/ggmagazine_2013/gg_04_07_2013_148.pdf.


[31] Schöffler T, Hoffmann K, Koch T. Stoichiometric Natural Gas Combustion in a Single Cylinder SI Engine and Impact
of Charge Dilution by Means of EGR. SAE Technical paper 2013-24-0113; 2013.
[32] Hill, S. C., and Smoot, L. D. (2000). “Modeling of nitrogen oxides formation and destruction in combustion systems.”
Prog. Energy Combust.Sci., 26(4–6), 417–458.
[33] Ladommatos N, Abdelhalim S, Zhao H, Hu Z. The dilution, chemical and thermal effects of exhaust gas recirculation
on diesel emissions – part 2: effects of carbon dioxide. Society of Automotive Engineers. SAE paper 961167; 1996.
[34] Ladommatos N, Abdelhalim S, Zhao H, Hu Z. The dilution, chemical, and thermal effects of exhaust gas recirculation
on diesel engine emissions – part 3: effects of water vapour. Society of Automotive Engineers. SAE paper 971659; 1997.

PT
[35] Heywood JB. Internal combustion engine fundamentals McGraw-Hill Book Co, New York (1988) Ramos JI. Internal
combustion engine modeling. New York: Hemisphere; 1989.
[36] Mendez S, Kashdan JT, Bruneaux G, Thirouard B, Vangraefschepe F. Formation of unburned hydrocarbons in low
temperature diesel combustion. SAE paper 2009-01-2729; 2009.

RI
[37] Jung GS, Sung YH, Choi BS, Lee CW, M. T. Lim MT. Major sources of hydrocarbon emissions in a premixed charge
compression ignition engine. International Journal of Automotive Technology 2012; 13(3): 347−353.
[38] Hussain J, Palaniradja K, Alagumurthi N, Manimaran R. Effect of Exhaust Gas Recirculation (EGR) on Performance

SC
and Emission characteristics of a Three Cylinder Direct Injection Compression Ignition Engine. Alexandria Engineering
Journal 2012; 51(4): 241–247.
[39] Lampert JK, Kazi MS, Farrauto RJ. Palladium catalyst performance for methane emissions abatement from lean burn
natural gas vehicles. Applied Catalysis B: Environmental 1997; 14(3-4):211–223.

U
[40] Thierry Leprince, Joe Aleixo, Kamal Chowdhury, Mojghan Naseri, Shazam Williams. Development of pre-turbo
catalyst for natural gas engines. Proceedings of ICES03 2003 spring technical conference of the ASME internal combustion
engine division ICE2003-568; Salzburg, Austria, May 11-14, 2003.
AN
[41] Papagiannakis RG, Rakopoulos CD, Hountalas DT, Rakopoulos DC. Emission characteristics of high speed, dual fuel,
compression ignition engine operating in a wide range of natural gas/diesel fuel proportions. Fuel 2010; 89:1397–406.
[42] Barik D, Murugan S. Simultaneous reduction of NOx and smoke in a dual fuel DI diesel engine. Energy Conversion
and Management 2014; 84:217–226.
M
D
TE
C EP
AC

23
ACCEPTED MANUSCRIPT

Figure Captures
Fig.1 layout of the test bed
Fig.2 Contour map of spark timing
Fig.3 Contour map of EGR rate
Fig.4 Cylinder pressure(a) and heat release rate(b) at different EGR rates
Fig.5 Effects of EGR rate on maximum in-cylinder temperature

PT
Fig.6 Effects of excess air ratio on NOx emissions, maximum cylinder temperature and maximum cylinder pressure
Fig.7 Contour map of maximum cylinder pressure(a) and the corresponding phase angle(b)

RI
Fig.8 Contour map of maximum pressure rise rate(a) and the corresponding phase angle(b)
Fig.9 Contour map of MFB0-10%(a), MFB50%(b) and MFB10-90%(c)

SC
Fig.10 Contour map of maximum heat release rate(a) and the corresponding phase angle(b)
Fig.11 Contour map of cyclic variation for IMEP
Fig.12 Contour map of engine-out NOx emissions (a) and tailpipe NOx emissions(b)

U
Fig.13 Contour map of NOx conversion efficiency(a) and TWC inlet temperature(b)
AN
Fig.14 Contour map of engine out THC emissions(a) and CH4 to THC ratio(b)
Fig.15 Contour map of tailpipe CH4 emissions(a) and tailpipe NMHC emissions(b)
Fig.16 Contour map of THC conversion efficiency(a) and engine-out O2(b)
M

Fig.17 Contour map of engine-out CO emissions(a), tailpipe CO emissions(b) and CO conversion efficiency(c)
Fig.18 Contour map of tailpipe NH3 emissions
D

Fig.19 Contour map of smoke opacity


Fig. 20 Contour map of brake thermal efficiency(a) and tailpipe CO2 emissions(b)
TE
C EP
AC

24
ACCEPTED MANUSCRIPT
 The test engine is equipped with cooled EGR and TWC.
 The sensitivity of EGR rate and excess air ratio to NOx emissions has been
analyzed.
 The smoke emissions are extremely low and can easily meet the EURO VI
emission standard.
 The necessity of the addition of ASC device has been pointed out.

PT
RI
U SC
AN
M
D
TE
C EP
AC

You might also like