You are on page 1of 15

Applied Energy 159 (2015) 117131

Contents lists available at ScienceDirect

Applied Energy
journal homepage: www.elsevier.com/locate/apenergy

High load performance and combustion analysis of a four-valve direct


injection gasoline engine running in the two-stroke cycle
Macklini Dalla Nora , Hua Zhao
Brunel University London, Centre for Advanced Powertrain and Fuels Research (CAPF), Kingston Lane, Uxbridge, Middlesex UB8 3PH, United Kingdom

h i g h l i g h t s

 The two-stroke cycle was achieved in a four-valve highly boosted gasoline engine.
 High charging efficiencies were realised by means of valve timing optimisation.
 The engine achieved 2.4 MPa equivalent IMEP with 7 MPa in-cylinder pressure.
 Scavenging inefficiencies and poor airfuel mixing limited the high speed operation.

a r t i c l e i n f o a b s t r a c t

Article history: With the introduction of CO2 emissions legislation or fuel economy standards in Europe and many coun-
Received 25 March 2015 tries, significant effort is being made to improve spark ignition gasoline engines because of their domi-
Received in revised form 23 August 2015 nant market share in passenger cars and potential for better fuel economy. Amongst several
Accepted 27 August 2015
approaches, the engine downsizing technology has been adopted by the automotive companies as one
of the most effective methods to reduce fuel consumption of gasoline engines. However, aggressive
engine downsizing is constrained by excessive thermal and mechanical loads as well as knocking com-
Keywords:
bustion and low speed pre-ignition (also known as super-knock). In order to overcome such difficulties,
Two-stroke poppet valve engine
Gasoline direct injection
a gasoline direct injection single cylinder engine was modified to run under the two-stroke cycle by oper-
Engine downsizing ating the intake and exhaust valves around bottom dead centre (BDC) at every crankshaft revolution. The
Controlled auto-ignition combustion products were scavenged by means of a reversed tumble flow of compressed air during the
positive valve overlap period at BDC. The engine output was determined by the charging and trapping
efficiencies, which were directly influenced by the intake and exhaust valve timings and boost pressures.
In this research a valve timing optimisation study was performed using a fully flexible valve train unit,
where the intake and exhaust valve timings were advanced and retarded independently at several speeds
and loads. A supercharger was used to vary the load by increasing the intake pressure. The effects of valve
timing and boost pressure in this two-stroke poppet valve engine were investigated by a detailed analysis
of the gas exchange process and combustion heat release. Gaseous and smoke emissions were measured
and analysed. The results confirmed that the two-stroke cycle operation enabled the indicated mean
effective pressure to reach 1.2 MPa (equivalent to 2.4 MPa in a four-stroke cycle) with an in-cylinder
pressure below 7 MPa at an engine speed as low as 800 rpm. The engine operation was limited by scav-
enging inefficiencies and short time available for proper airfuel mixing at high speeds using the current
fuel injector. The large amounts of hot residual gas trapped induced controlled auto-ignition combustion
at high speeds, and thus the abrupt heat release limited higher loads.
2015 Elsevier Ltd. All rights reserved.

Abbreviations: ATDC, after top dead centre; BSFC, brake specific fuel consumption; CA, crank angle; CAI, controlled auto-ignition; CO, carbon monoxide; COVIMEP,
covariance of the indicated mean effective pressure; dP/dh, rate of pressure rise; ECR, effective compression ratio; EER, effective expansion ratio; EGR, exhaust gas recycling;
EVC, exhaust valve closing; EVO, exhaust valve opening; GDI, gasoline direct injection; UHC, unburned hydrocarbon; IMEP, indicated mean effective pressure; ISCO, indicated
specific carbon monoxide; ISFC, indicated specific fuel consumption; ISUHC, indicated specific unburned hydrocarbon; ISNOx, indicated specific oxides of nitrogen; IVC,
intake valve closing; IVO, intake valve opening; KLS, knock limited spark advance; LHV, lower heating value; MBT, minimum spark advance for maximum break torque; NOx,
oxides of nitrogen; rpm, revolutions per minute; SI, spark ignition.
Corresponding author. Tel.: +44 (0)74630 95392.
E-mail address: macklini.dallanora@brunel.ac.uk (M. Dalla Nora).

http://dx.doi.org/10.1016/j.apenergy.2015.08.122
0306-2619/ 2015 Elsevier Ltd. All rights reserved.
118 M. Dalla Nora, H. Zhao / Applied Energy 159 (2015) 117131

1. Introduction mechanical inefficiencies [10]. Nevertheless, significant advances


have been made in high pressure fuel direct injection systems
Two-stroke engines are well known for their superior power and high efficiency boosting devices (superchargers, turbochargers
density and reduced weight compared to equivalent four-stroke and e-boosters). In addition, flexible variable valve actuation sys-
units and are employed to power handheld tools to large marine tems, particularly fast variable cam devices, have been developed
engines [1,2]. Their use for high performance purposes is widely to enhance air management for production engines [11]. Such
spread for motorbikes, snowmobiles and outboard vehicles, with technological improvements have prompted renewed interest in
claimed power densities above 220 kW/L [3]. However, these developing two-stroke poppet valve gasoline engines [12,13] and
advantages, mainly related to crank-case scavenged two-stroke diesel engines [10,14], considering their potential for engine down-
engines, are often offset by drawbacks regarding gaseous emis- sizing with lower in-cylinder pressures and less structural stresses
sions, thermal efficiency and engine components durability [4]. than downsized four-stroke engines [15,16]. Furthermore, the two-
On the subject of emissions, the fuel short-circuiting in mixture stroke poppet valve engine shares nearly all components from the
scavenged two-stroke engines results in significant unburned contemporary four-stroke engine architecture and hence can be
hydrocarbon (UHC) emissions. The lubricant added to the fuel produced from the same manufacturing process.
has much less effect on emissions from crank-case scavenged The majority of studies with two-stroke poppet valve engines
two-stroke engines according to [3], as modern units use propor- has taken advantage of the inherent residual gas trapped at
tions as low as 1% of oil in the fuel. Regarding the thermal effi- reduced charging efficiencies to achieve high in-cylinder thermal
ciency, conventional two-stroke engines usually lose expansion conditions, and hence induce auto-ignition combustion in both
work in favour of enhanced scavenging through early exhaust port spark-ignition [12,13,17] and compression-ignition engines
opening. This procedure uses the exhaust blow-down phase to [18,19]. In the auto-ignition combustion high thermal efficiency
reduce the levels of residual gas trapped prior to the intake pro- and near zero oxides of nitrogen (NOx) emissions can be achieved
cess, ensuring higher degrees of charge purity [5]. Lastly, the across the part load engine operation range. However, the low
reduced components durability (piston, rings and liner) of ported charging efficiency, usually 70% at most, results in large amounts
two-stroke engines can be attributed to uneven thermal loads of hot residual gas trapped and increased combustion noise at high
and reduced lubricant oil film when UHC emissions is a concern loads. The violent combustion in this case limits the engine load to
[6]. It is important to keep in mind that all these disadvantages values of indicated mean effective pressure (IMEP) from 0.5 to
are related to cross-scavenged and loop-scavenged two-stroke 0.7 MPa using commercial gasoline.
engines with intake and exhaust ports, where the crank-case is Considering this, the present study aims at investigating the
employed as a pump for the air or air/fuel mixture and therefore two-stroke poppet valve engine behaviour under charging efficien-
lubricant oil needs to be added to the air stream. Such problems cies above 90% at low speeds, showing the full potential of this
can be avoided by the uniflow two-stroke engine concept, in which type of engine for aggressive downsizing and downspeeding com-
externally compressed air is supplied through ports at bottom dead pared to four-stroke engines. Such level of charging efficiency
centre (BDC) and the exhaust gas is forced out through conven- could be only achieved with high values of boost pressure and a
tional poppet valves in the cylinder head. Greater charging efficien- valve timing optimisation study, where the intake and exhaust
cies can be achieved with such designs [1], but production valves were operated independently aiming at reducing the levels
complexity and packaging restrictions have limited its application of residual gas trapped to a minimum. The effect of the scavenging
to large marine diesel engines so far, though some attempts have process on the engine performance was evaluated at several
been made to adopt such an engine design for vehicular applica- speeds and loads, whilst measurements of gaseous and smoke
tions [7]. emissions, as well as a heat release analysis, were performed to
In the beginning of 1990 a new concept of two-stroke operation study the airfuel mixing and combustion process. High values of
was proposed as a possible solution to overcome the problems IMEP with low in-cylinder pressures are expected in this two-
related to conventional ported two-stroke engines. Based on the stroke direct injection gasoline poppet valve engine.
design of modern four-stroke engines, the two-stroke scavenging
process was achieved through the overlap period of overhead
intake and exhaust valves around BDC at every engine revolution 2. Experiments
[2,8]. Because of the use of poppet valves higher power outputs
could be achieved with the same engine durability as four-stroke 2.1. Experimental facilities
engines. The high levels of UHC emissions due to fuel short-
circuiting had been addressed by direct fuel injection and air- All the experiments were conducted on a single cylinder
assisted fuel injection [5]. The lubricant oil consumption, charac- research engine mounted on a transient test bed. The engine was
teristic of crank-case scavenged engines, had been eliminated by equipped with an electro-hydraulic fully variable valve train unit
using wet sump and external scavenge pump, mostly roots blower capable of independent control over the timings and lifts of each
superchargers. When applying this concept to Diesel engines, 40% of the four valves, enabling both two-stroke and four-stroke cycles
higher torque and reduced combustion noise compared to an operation [20]. The valve control unit operated under closed loop
equivalent four-stroke model was demonstrated by Toyota [6]. control over the oil pressure, oil temperature and valve timing/lift,
A reported problem of gasoline direct injection (GDI) two- ensuring precise valve operation up to 3000 rpm in the two-stroke
stroke poppet valve engines was the insufficient mixing between cycle. A Ricardo rCube engine control unit was used to manage the
fuel and air in the cylinder, mainly attributed to the shorter time throttle position, injection pulse width, spark timing and valve
available and the relatively lower injection pressures used at the parameters. An AC dynamometer enabled both motored and fired
time [8]. This poor airfuel mixing results in stratified charging operations whilst an external cooling system provided fully auto-
and incomplete combustion, with large emissions of carbon mated control over the engine oil and coolant temperatures. Com-
monoxide (CO), UHC and soot [9]. The boosted air supplier, essen- mercial gasoline (95 RON) was directly injected into the
tial for this type of two-stroke engine, has been also reported as combustion chamber through a Denso solenoid double-slit type
responsible for hindering fuel consumption due to isentropic and injector [21]. The instantaneous fuel mass flow rate was measured
M. Dalla Nora, H. Zhao / Applied Energy 159 (2015) 117131 119

by an Endress + Hauser Promass 83A Coriolis flow meter, with a


maximum error of 0.2% for the flow range studied. The boosted
air was supplied by an AVL 515 sliding vanes compressor unit with
closed loop control over the pressure. The air mass flow rate was
measured by a Hasting HFM-200 laminar flow meter with a max-
imum error of 1%. The intake and exhaust pressures were mea-
sured by two Kistler piezo-resistive transducers installed in the
intake plenum (4007BA20F) and in the exhaust port (4007BA5F),
with a maximum error of 0.1%. The in-cylinder pressure was mea-
sured by a Kistler 6061B piezo-electric sensor, with a maximum
measurement error of 0.5%. To record the crank angle position a
LeineLinde incremental encoder with a resolution of 720 pulses
per revolution was employed. Averaged temperatures were mea-
sured at the intake plenum, exhaust runner, oil gallery, coolant
jacket and fuel rail by using K-type thermocouples with an accu-
racy of 1%. An AVL 415SE smoke meter was used to measure
the smoke levels, with repeatability better than 3% of the measured
Fig. 2. Cylinder head details.
values. Gaseous emissions were analysed by a Horiba MEXA
7170DEGR using the non-dispersive infrared principle for CO, a
heated flame ionization detector for UHC, a paramagnetic detector intake ports, as shown by a small recession in the engine geometry.
for oxygen (O2) and a heated chemiluminescence detector for NOx. The unconventional intake ports arrangement aimed at improving
The overall error attained to each gas measurement was smaller the reverse tumble flow necessary to scavenge the burned gases,
than 2%. The location of all instruments described above can be while keeping acceptable values of discharge coefficients.
found in Fig. 1, as well as the temperature and pressure measure- The reverse tumble flow was improved by masking the cylinder
ment points labelled as T and P, respectively. head region around both intake valves, so the intake charge was
A National Instruments 6353 USB X card was used for data prevented from going straight into the exhaust system in the so
acquisition (DAQ) and an in-house software was employed for called air short-circuiting. The majority of the flow was then direc-
combustion analysis. ted downwards in the cylinder, as seen in Fig. 3, and the scaveng-
ing process could be improved.
2.2. Engine details
2.3. Test procedures
The prototype engine had an 81.6 mm bore and 66.9 mm stroke,
which resulted in a swept volume of 350 cm3 and an oversquared The two-stroke cycle was achieved by opening both the intake
bore-to-stroke ratio of 1.22. A geometrical compression ratio of and exhaust valves around BDC as presented in Fig. 4. The long
11.8:1 was achieved with a dome-in-piston and a pent roof com- valve overlap period allowed the inlet boosted air to scavenge
bustion chamber. The intake valves were 28 mm in diameter while the combustion products. The start of fuel injection (SOI) occurred
the exhaust valves were 30 mm, which is peculiar compared to after all the valves were closed to avoid fuel short-circuiting to the
conventional four-stroke engines where the intake valves are usu- exhaust. In addition, the SOI after intake valve closing (IVC) pre-
ally bigger. In the two-stroke cycle, although, the scavenging pro- vented fuel from entering into the intake ports through backflow,
cess at high engine speeds is more dependent on the exhaust which may occur if the in-cylinder pressure becomes higher than
flow characteristics and hence the exhaust flow area should be the intake port pressure. The fuel entrained in the intake port could
greater as seen in ported two-stroke engines. The engine had two then be carried back into the cylinder and pass directly to the
conventional side exhaust ports and two individual upright- exhaust port in the following cycle, contributing to increased
straight intake ports shown in blue and green, respectively, in UHC emissions.
Fig. 2. The spark plug was located in the centre of the combustion At each of the engine speeds studied, i.e. 800, 1500, 2200 and
chamber while the fuel injector was side mounted between the 3000 5 rpm, five intake pressure levels were applied (where

Fig. 1. Research engine and test cell facilities.


120 M. Dalla Nora, H. Zhao / Applied Energy 159 (2015) 117131

4 IVO 130 IVO 135


IVO 140 IVO 145
IVO 150 EVO 130
3

Valve lift (mm)


2

0
110 150 190 230 270
Crank angle ()

Fig. 5. Intake valve timing optimisation.

Fig. 3. Cross section of cylinder head at the valves plane.

EVO 120 EVO 125


4
possible) as a way to control the engine load. By increasing the EVO 130 EVO 135
boost pressures from 120 2 kPa to 280 3 kPa the delivery ratio EVO 140 IVO 140
increased and less residual gas was trapped, resulting in greater 3

Valve lift (mm)


air mass in the cylinder and higher engine power output. At some
operation points stable combustion was not achieved as the
2
covariance of the indicated mean effective pressure (COVIMEP)
reached a limit of 10%. This value seems high for four-stroke
engines where a value around 5% is usually considered [22]. 1
However, bearing in mind the doubled firing frequency of
two-stroke engines the torque variation is reduced and the levels
0
of vibration and harshness are attenuated. In a previous study, 110 150 190 230 270
stable operation in a two-stroke poppet valve engine was claimed
Crank angle ()
at COVIMEP values as high as 35% [5].
At each engine speed, nine different combinations of intake and Fig. 6. Exhaust valve timing optimisation.
exhaust opening/closing timings were tested as shown in Figs. 5
and 6. The intake and exhaust valve durations were kept constant
at 100 CA (crank angle) and 120 CA, respectively. At each engine from 295 K to 305 K, except for the maximum intake pressure of
speed and a given boost pressure, the exhaust valve timing was 280 kPa at 800 rpm when it reached 325 K due to insufficient air
kept fixed and firstly the intake valve opening (IVO) was varied cooling.
from 130 CA to 150 CA after top dead centre (ATDC), in steps of The ignition timing was set to minimum spark advance for
5 CA. Then, the intake valve timing was fixed and the exhaust maximum brake torque (MBT) or knock limited spark (KLS) at con-
valve opening (EVO) was varied from 120 CA to 140 CA ATDC, ditions when knocking combustion occurred. A knocking combus-
also in steps of 5 CA. tion threshold of 1 MPa/CA was set for the maximum rate of
In the previous research at part-load conditions [23] the pressure rise (dP/dh).
exhaust valve closing (EVC) took place before the IVC to increase To ensure the same airfuel mixing conditions for all the valve
the residual gas trapped for controlled auto-ignition (CAI) combus- timings studied the SOI was set to 260 CA ATDC, which was the
tion. In this study, however, the EVC was delayed to obtain higher latest EVC timing tested. The fuel injection pressure was set to
charging efficiencies. In addition, the exhaust valve opened earlier 15.0 0.5 MPa and its temperature kept at 293 5 K.
during the expansion process to increase the exhaust blow-down At each intake pressure and engine speed tested, the fuelling
period. During the experiments the maximum advance of IVO rate was increased until a fuel rich in-cylinder charge was
was set to the EVO. obtained. This fuelling rate was then kept constant as the intake
The engine coolant and oil temperatures were kept at 353 3 K and exhaust valve timings were varied, so the effect of the gas
for all cases studied. The intake air temperature was in the range exchanging process could be solely evaluated. Furthermore, as
the engine speed changed, the fuel flow rate was also varied
4 Intake valves Exhaust valves accordingly to ensure a fuel rich in-cylinder charge as seen in
Table 1. The reason why lean mixtures could not be employed
relies on the method used in this research to calculate the gas
Valve lift (mm)

3
exchanging parameters, particularly the air trapping efficiency.
2 By evaluating the exhaust gas composition, the technique reported
SOI
by [4,24] considers that all free oxygen in the exhaust stream is
1
resulted from inefficiencies during the scavenging process. In this
case a sufficiently rich mixture was required during the combus-
tion process to ensure the minimum possible oxygen remaining
0
0 90 180 270 360 within the cylinder at EVO, otherwise the air trapping efficiency
Crank angle ( ) would be underestimated. The equation used to calculate the trap-
ping efficiency based on this presumption is presented in the fol-
Fig. 4. Two-stroke cycle operation principle. lowing section.
M. Dalla Nora, H. Zhao / Applied Energy 159 (2015) 117131 121

Table 1 0:5yCOCO CO2 


Fuel flow rates at different engine speeds and intake pressures. H2  6
CO 3CO2 
Fuel flow rate (mg/cycle) Intake pressure (kPa)
where [H2] is the exhaust concentration of hydrogen, y is the hydro-
120 160 200 240 280 gen to carbon ratio of the fuel (considered 1.87), [CO] is the exhaust
Engine speed (rpm) 800 17.1 20.4 24.8 27.1 30.8 concentration of carbon monoxide and [CO2] is the exhaust concen-
1500 12.3 20.5 25.5 34.3 tration of carbon dioxide. All exhaust gas concentrations in ppm
2200 8.86 16.8
(volumetric basis).
3000 6.01
The air trapping efficiency, defined as the ratio of in-cylinder
trapped air mass to the total intake air mass, was calculated based
on the analytical method developed for fuel rich combustion in
2.4. Data analysis two-stroke engines [24]:

Trapeff A
Based on the in-cylinder pressure and crank-angle measure-  
yKCO2 
ments, the mass fraction burnt was calculated by integrating the 0:5CO CO2  0:25 COKCO 
CO CO 2  0:5NOx
 2

net heat release rate as presented in [25]: yKCO2 
0:5CO CO2  O2  0:25 COKCO CO CO2  0:5NOx
2
dQ net c dV 1 dp 7
p V 1
dh c  1 dh c  1 dh
where Trapeff(A) is the air trapping efficiency, [CO] is the exhaust
where Qnet is the net heat release in J/CA, h is the crank angle in CA, concentration of carbon monoxide, [CO2] is the exhaust concentra-
c is the ratio of specific heats (considered constant and equal to tion of carbon dioxide, y is hydrogen to carbon ratio of the fuel (con-
1.33), p is the in-cylinder pressure in Pa and V is the in-cylinder vol- sidered 1.87), K is the watergas reaction equilibrium constant
ume in m3. (considered 3.5), [NOx] is the exhaust concentration of oxides of
The effective compression and expansion ratios were calculated nitrogen and [O2] is the exhaust concentration of oxygen. All
from the instantaneous in-cylinder volumes at EVC and EVO, exhaust gas concentrations in ppm (volumetric basis).
respectively. Due to the air short-circuiting from the intake to the exhaust
V EVO V clr during the valve overlap period, the measured exhaust lambda
EER 2 value differed from the in-cylinder lambda. The in-cylinder lambda
V clr
was then calculated based on the air trapping efficiency and fuel
V EVC V clr trapping efficiency [24]:
ECR 3
V clr Trapeff A
kin-cylinder kexhaust 8
where EER is the effective expansion ratio, ECR is the effective com- Trapeff F
pression ratio, VEVO is the instantaneous in-cylinder volume at EVO,
where kin-cylinder is the in-cylinder lambda, kexhaust is the exhaust
VEVC is the instantaneous in-cylinder volume at EVC and Vclr is the
lambda, Trapeff(A) is the air trapping efficiency and Trapeff(F) is the
clearance volume (32.4 cm3).
fuel trapping efficiency.
Exhaust emissions were converted from parts per million (ppm)
The fuel trapping efficiency (defined as the ratio of in-cylinder
to g/kW h based on the UN Regulation number 49 [26]:
trapped fuel mass to the total injected fuel mass) in a GDI engine
ugas cgas kh q_ exh is expected to be 100%, where no fuel short-circuiting is supposed
ISgas 4 to happen. However, due to the high levels of fuel stratification
Pi
resulted from the short time available for airfuel mixing at high
where ISgas is the indicated specific gas emission (ISCO, ISUHC or speeds and loads, some of the fuel could not take part in the
ISNOx) in g/kW h, ugas is the dimensionless specific gas constant combustion process and left the cylinder unburned. Thus, the fuel
(CO = 0.000966, UHC = 0.000499 and NOx = 0.001587) for gasoline trapping efficiency was introduced to take into account the
fuelled engines, cgas is the gas concentration in the exhaust stream short-circuited fuel from the previous cycle, similar to that used
in ppm (volumetric basis), kh is the dimensionless correction factor in conventional ported two-stroke engines [24]:
to convert gaseous measurements of CO and NOx from a dry basis to
wet basis (see Ref. [26] for more details), q_ exh is the exhaust mas CO CO2 
Trapeff F 9
flow rate (g/h) and Pi is the indicated power output (kW). CO CO2  UHC
The combustion efficiency was calculated based on the emis- where Trapeff(F) is the fuel trapping efficiency, [CO] is the exhaust
sions products not fully oxidized during the combustion. The CO, concentration of carbon monoxide, [CO2] is the exhaust concentra-
UHC and H2 mass flow rates were acquired by multiplying their tion of carbon dioxide and [UHC] is the exhaust concentration of
respective indicated specific emissions by the indicated power, so unburned hydrocarbons. All exhaust gas concentrations in ppm
the values in g/h could be obtained. (volumetric basis).
m _ UHC LHVUHC m
_ CO LHVCO m _ H2 LHVH2 The charging efficiency, described as the ratio of delivered air
gc 1  _ fuel LHVfuel
5 mass retained in the cylinder charge to the total in-cylinder charge
m
at intake conditions, was used to indicate how efficiently the
where gc is the combustion efficiency, m _ CO is the mass flow rate of burned gases were displaced during the scavenging process. It
CO, LHVCO is the lower heating value (LHV) of CO (10.1 MJ/kg), m _ UHC can be calculated based on the air trapping efficiency and delivery
is the mass flow rate of UHC, LHVUHC is the LHV of UHC (44 MJ/kg), ratio, as follows:
m_ H2 is the mass flow rate of H2, LHVH2 is the LHV of H2 (120 MJ/kg),  
mair
m_ fuel is the fuel mass flow rate and LHVfuel is the LHV of the fuel Charg eff Trapeff A 10
Swv ol qair
(44 MJ/kg).
Emissions of hydrogen (H2) were estimated based on the mea- where Chargeff is the charging efficiency, Trapeff(A) is the air trapping
surements of CO and CO2 according to [27]: efficiency, mair is the delivered air mass per cycle in kg, Swvol is the
122 M. Dalla Nora, H. Zhao / Applied Energy 159 (2015) 117131

engine swept volume in m3 and qair is the air density at intake con- When the fuelling rate was reduced to avoid excessive heat release
ditions in kg/m3. The term between brackets in Eq. (10) is the deliv- rate at higher boost pressures, unstable combustion occurred as
ery ratio, which compares the current delivered air mass per cycle measured by the higher COVIMEP values. On the other hand, when
to the reference air mass in an ideal charging process. the fuelling rate was increased to avoid combustion instabilities,
the dP/dh rose above the knock limit. The occurrence of violent
combustion or unstable combustion was likely related to the large
3. Results and discussion
amount of hot residual gas trapped, resulted from insufficient time
available for scavenging at higher engine speeds [28]. The presence
The results presented here are averaged over 100 consecutive
of hot residual gas raised the charge temperature and accelerated
cycles and plotted as a function of valve timings at given engine
the occurrence of auto-ignition combustion in the unburned mix-
speeds and intake pressures. The nomenclature of the different
ture, resulting in rapid and violent heat release rate. In addition,
valve timings studied consists of the IVO and the EVO timings in
since the SOI took place as late as 260 CA ATDC (similar to that
CA ATDC. The Y-axis is further divided into four parts according
used in stratified charge four-stroke GDI engines), significant fuel
to the engine speed.
stratification was present resulting in increased mixture reactivity.
At 800 rpm all the boosting levels could be tested throughout
3.1. Performance and combustion analysis the valve timings studied except for the latest IVO (150 CA) and
earliest EVO (120 CA), when combustion became excessively
Fig. 7 shows the indicated mean effective pressure (IMEP) unstable. From the left to the middle point along the x-axis the
results at different engine speeds and boost pressures. It is noted IVO was retarded from 130 to 150 CA ATDC at a constant EVO
that highly boosted operation was not possible at higher speeds of 130 CA ATDC. At the lowest boost pressure of 120 kPa the IMEP
(2200 rpm and 3000 rpm) due to violent and unstable combustion. values varied little with IVO. When the boost pressure was higher
than 160 kPa, the IMEP increased with the retarded IVO and
reached its peak at IVO 150 CA ATDC. It is noted that the higher
0.26 3000 RPM the boost pressure the more pronounced is the change in IMEP
with IVO. This can be explained by an increase in the charging effi-
ciency as presented in Fig. 8, resulted from higher pressure differ-
0.23
ence between the intake and exhaust ports. When the IVO was
retarded, a more effective blow-down event without intake air
0.20 contamination was allowed. Such effect would be even more pro-
nounced at higher boost pressures. At 1500 rpm the IVO and EVO
2200 RPM
0.59 sweeps had similar effects on the IMEP, but no stable combustion
could be achieved at the boost pressure of 280 kPa.
0.47 From the right to the middle along the x-axis in Figs. 7 and 8,
when the EVO was advanced from 140 to 120 CA ATDC and the
0.35 IVO kept at 140 CA ATDC, the charging efficiency (and therefore
the IMEP) changed little at lower boost pressures but rose steadily
0.23 to reach its peak at the middle of the graph. This behaviour mir-
rored the left part of the curve and can be explained by the
1500 RPM increased blow-down period and higher pressure ratio across the
IMEP (MPa)

1.20
exhaust valves at an earlier EVO. In addition, the difference
between the intake air pressure and the in-cylinder burned gases
1.00
was greater at the same IVO as the in-cylinder pressure had
dropped to a lower value due to extended exhaust blow-down.
0.80
At 800 rpm the peak IMEP of 1.2 MPa was achieved at an intake
pressure of 280 kPa, producing a specific torque of 195 N m/L with
0.60
the in-cylinder peak pressure as low as 6.8 MPa. To produce the
same torque at the same speed in a four-stroke engine of the same
0.40
displacement, the engine would need to be operated at 2.4 MPa
800 RPM IMEP. The in-cylinder pressure in this case would be expected to
1.23
be nearly twice as high (13.6 MPa), resulting in great structural
1.03
stresses and thermal load issues. Such operation condition could
only be achieved in a highly downsized engine, assuming the
0.83 engine would not be limited by knocking combustion and/or low
speed pre-ignition (LSPI) inducing super-knock [29]. This high
0.63 value of torque at low speeds is comparable to extremely boosted
modern diesel engines under the concept of downspeeding, where
0.43 120 kPa 160 kPa 200 kPa the engine operation region is shifted towards lower speeds with
240 kPa 280 kPa reduced friction and gas exchange losses [30]. As the engine speed
0.23 increased, although, the charging efficiency dropped and therefore
IVO 130, EVO 130

IVO 135, EVO 130

IVO 140, EVO 130

IVO 145, EVO 130

IVO 150, EVO 130

IVO 140, EVO 120

IVO 140, EVO 125

IVO 140, EVO 135

IVO 140, EVO 140

lower loads could be realised at reduced values of charge purity. At


2200 rpm, for instance, the maximum specific torque dropped to
about 88 N m/L and resulted in a modest specific power output
of about 20 kW/L. This result is still less than half of that found
in an equivalent downsized four-stroke engine operating at the
same speed, where specific power values around 54 kW/L are
Fig. 7. Indicated mean effective pressure. reported [31]. The linear trend of specific torque and in-cylinder
M. Dalla Nora, H. Zhao / Applied Energy 159 (2015) 117131 123

20 250 7.5
3000 RPM

In-cylinder pressure (MPa)


18 200 6.0

Specific torque (Nm/litre)


16
150 4.5

44
2200 RPM
100 3.0
36

28 50 1.5
Specific torque

20
In-cylinder pressure
0 0.0
1500 RPM 500 1250 2000 2750 3500
90
Engine speed (rpm)
efficiency (%)
Charging

75 Fig. 9. Maximum specific torque and corresponding in-cylinder pressure achieved


at the speeds tested.

60

It can be seen from the PV diagram that the largest amount of


45 useful work was achieved with the earliest EVO (IVO 140, EVO 120)
and the latest IVO (IVO 150, EVO 130), when the scavenging pro-
30 cess was optimised and less residual gas was trapped. As the valve
timing was moved towards IVO 130, EVO 130, the greater charge
95 800 RPM dilution promoted by the internal EGR reduced the heat release
rate and hence the peak pressure. It can be seen that in this case
80
as the intake and exhaust valves opened at the same time, part
of the burned gases mixed with the intake charge and thus com-
65
promised the in-cylinder charge purity during the next cycle
50 [32]. The valve timing IVO 140, EVO 140 was characterised with
even lower in-cylinder peak pressure as a result of greater amounts
35 120 kPa 160 kPa 200 kPa of residual gas trapped, as shown by the lower charging efficiency
240 kPa 280 kPa (Fig. 8). As presented by the zoomed part of the PV diagram in
20 Fig. 7, in this case the EVO was the most retarded and the
IVO 135, EVO 130

IVO 140, EVO 130

IVO 145, EVO 130

IVO 150, EVO 130

IVO 140, EVO 120

IVO 140, EVO 125

IVO 140, EVO 135

IVO 140, EVO 140


IVO 130, EVO 130

expansion loop was the longest amongst those shown. These


two extreme valve timings also showed the highest in-cylinder
pressures around BDC, which caused the poor scavenging as the
pressure drop across the intake valves decreased. Moreover, the
in-cylinder pressure at the end of the compression phase for these
two cases was about 50% lower than that for IVO 150, EVO 130
Fig. 8. Charging efficiency. and IVO 140, EVO 120, resulted from less trapped fresh air mass
and higher levels of residual gas with larger heat capacity.
The two valve timings with the highest in-cylinder pressures in
pressure with the engine speed is presented in Fig. 9, for the Fig. 10, i.e.: IVO 150, EVO 130 and IVO 140, EVO 120, presented
selected valve timing IVO 145, EVO 130. similar peak pressures (less than 4% of difference), although the
The results in Fig. 8 illustrate that the maximum IMEP was a early EVO case had reduced useful work and hence 2% lower IMEP.
direct result of the most completed scavenging process achieved At this speed it is possible to confirm that the exhaust blown-down
at the latest IVO (IVO 150, EVO 130) and earliest EVO (IVO 140, phase can be partially replaced by a later EVO (130) with
EVO 120). Because the fuelling rate was kept constant at a given improved expansion work, without compromising the purity of
intake pressure, it would have been possible to achieve even higher the charge. For this two valve timings the difference in charging
engine power outputs by increasing the fuelling rate at these valve efficiency was less than 0.5% (Fig. 8), whilst the IMEP increased
timings at 800 rpm. However, it would have been at the expense of by 2% with later EVO (Fig. 7).
poorer combustion efficiency and higher fuel consumption. At any The gas exchange process in this two-stroke poppet valve
given IVO and EVO timings the charging efficiency dropped with engine was also affected by the actuation speed of the hydraulic
the increased engine speed because of the reduced time available valve train. As shown in Fig. 11, the valve opening and closing
for gas exchanging. For instance, at 2200 rpm and 120 kPa the slopes became less steep as the engine speed increased, resulting
residual gas level was found around 75%, whilst at 3000 rpm it in reduced effective flow area. Such limitation of the camless sys-
reached 82%. Furthermore, at each engine speed the charging effi- tem can be overcome by using a conventional camshaft of higher
ciency decreased from the middle to the both sides of the x-axis, lift driven by the crankshaft at the same speed.
reaching a minimum when the valves opened at the same time, i. Whilst the charging efficiency measured the effectiveness of the
e. IVO 130, EVO 130 and IVO 140, EVO 140. In order to better removal of burned gases, the air trapping efficiency was calculated
understand the scavenging results, the pressurevolume (PV) dia- to determine the air short-circuiting rate. As shown in Fig. 12, the
grams of four valve timings at 800 rpm and 200 kPa are plotted in trapping efficiency rose steadily with the engine speed as a result
Fig. 10. of shorter time available for gas exchanging and consequent lower
124 M. Dalla Nora, H. Zhao / Applied Energy 159 (2015) 117131

6.4 IVO 130, EVO 130 0.6


IVO 150, EVO 130
5.6 IVO 140, EVO 120 0.5

In-cylinder pressure (MPa)


IVO 140, EVO 140
4.8 0.4

4.0 0.3

3.2 0.2

2.4 0.1
270 290 310 330 350
1.6

0.8

0.0
25 50 75 100 125 150 175 200 225 250 275 300 325 350
In-cylinder volume (cm3)

Fig. 10. Pressurevolume diagram for selected valve timings at 800 rpm and 200 kPa intake pressure.

800 rpm 1500 rpm


4
2200 rpm 3000 rpm 3000 RPM
72
3
Valve lift (mm)

66

2
60

1
2200 RPM
68

0 64
110 140 170 200 230 260
Crank angle () 60

Fig. 11. Effect of engine speed on valve opening and closing durations. 56

70 1500 RPM
air short-circuiting rate. Higher trapping efficiencies were found
efficiency (%)
Air trapping

for earlier EVO and hence earlier EVC, particularly at 2200 rpm 55
and 3000 rpm, when the overlap period was reduced.
It is noted that when the intake air pressure was set at 120 kPa 40
the air trapping efficiency at 800 rpm and 1500 rpm exhibited dif-
ferent trends from the other pressures. This different pattern may
25
be attributed to a transition from a displacement dominated scav-
enging process to a mixing dominated scavenging process, as ide-
10
alised by the Benson-Brandham two-part model for gas
exchanging in two-stroke engines [33]. According to this theory 120 kPa 160 kPa 800 RPM
50
the scavenging was firstly dominated by a displacement process 200 kPa 240 kPa
until it reached a certain value of delivery ratio, which in this case 280 kPa
42
was around 1.5 at 800 rpm and 0.6 at 1500 rpm. After this point
the fresh air and the burned gases were more prone to mix until 34
the end of the scavenging process.
The combustion duration, calculated from 10% to 90% of the 26
mass fraction burned (MFB), is presented in two parts according
to the intake pressures: the first part for 120/160 kPa (Fig. 13) 18
and the second part for 200/240/280 kPa (Fig. 14).
At 800 rpm it is noted that the combustion durations decreased 10
IVO 135, EVO 130

IVO 145, EVO 130

IVO 150, EVO 130

IVO 140, EVO 125

IVO 140, EVO 135

IVO 140, EVO 140


IVO 130, EVO 130

IVO 140, EVO 130

IVO 140, EVO 120

slightly as the boost pressure and load increased because of the


higher charge temperatures and pressures. In addition, it can be
seen from Figs. 13 and 14 that the combustion duration was
between 13 CA and 19 CA at 800 rpm, which is much shorter than
that of spark ignition (SI) combustion in four-stroke engines. This
suggests that the heat release process might have taken place in
the form of a spark ignited flame around the spark plug and Fig. 12. Air trapping efficiency.
M. Dalla Nora, H. Zhao / Applied Energy 159 (2015) 117131 125

3000 RPM 50% greater than that at 800 rpm as a result of poorer charging effi-
24 ciencies at higher speeds. In addition, it can be seen that the most
retarded KLS occurred at the earliest EVO because of the minimum
12 residual gas concentration as evidenced by the highest charging
efficiency (Fig. 8). For the same reason, the KLS timing became
0 more retarded when the IVO was moved from 130 to 150 CA ATDC
and less residual gas was trapped. When the boost pressure was set
2200 RPM to 120 kPa, MBT could be achieved for all the valve timings and
35
more advanced MBT timings were realised near the middle of
25 the x-axis, when both the charging efficiency and trapping effi-
ciency were maximized.
At 800 rpm and 1500 rpm the valve timing IVO 150, EVO 130
15
presented the overall best indicated specific fuel consumption
(ISFC) seen in Fig. 17. At 800 rpm the ISFC of 252 g/kW h was found
5
at 1.1 MPa IMEP, while 264 g/kW h was achieved at 0.67 MPa. At
31 1500 RPM 1500 rpm the ISFC increased at all loads and the fuel consumption
deteriorated to about 316 g/kW h at the same 1.1 MPa IMEP. At
duration ( CA)
Combustion

2200 rpm and 3000 rpm the engine operation benefited from ear-
28
lier intake and exhaust valve opening as given by the valve timing
IVO 140, EVO 120. At these speeds the burning duration short-
25 ened due to the occurrence of CAI combustion and hence the
expansion work was not hindered by early EVO. The ISFC was
22 found at 264 g/kW h and 246 g/kW h in the range of loads 0.24
0.33 MPa IMEP at 2200 rpm and 3000 rpm, respectively.
19 Compared to downsized four-stroke engines operating at simi-
lar loads and speeds, the ISFC values were found in the same order
800 RPM of magnitude as the brake specific fuel consumption (BSFC) values
21
presented by [16] for a 850 cm3 two-cylinder GDI engine. Although
19 the two-stroke poppet valve engine does not experience pumping
losses, the power consumed by a supercharger should be consid-
17 ered alongside the friction losses when making direct comparisons
to BSFC results. In another study conducted in a highly boosted sin-
15
gle cylinder 400 cm3 engine running at stoichiometric airfuel
13 ratio [36], the ISFC of 253 g/kW h was registered at 1.6 MPa IMEP
120 kPa 160 kPa and 2000 rpm. For the sake of comparison, the poppet valve
11
IVO 130, EVO 130

IVO 135, EVO 130

IVO 140, EVO 130

IVO 145, EVO 130

IVO 150, EVO 130

IVO 140, EVO 120

IVO 140, EVO 125

IVO 140, EVO 135

IVO 140, EVO 140

31 1500 RPM

28

25
Fig. 13. Combustion duration for 120 kPa and 160 kPa intake pressures.
22
duration (CA)
Combustion

auto-ignition combustion of some premixed charge in the end-gas


19
[34,35]. As the engine speed went up to 1500 rpm the combustion
duration increased in terms of crank angles, but decreased slightly 800 RPM
20
in absolute time as the flame speed was accelerated by the higher
flow turbulence and the auto-ignition combustion was favoured by
18
the hotter residual gas.
At 2200 and 3000 rpm stable engine operation was mainly lim- 16
ited to the boost pressure of 120 kPa. During such operation it was
found that the spark timing had little effect on the combustion 14
phasing and auto-ignition combustion became the dominant heat
release process as evidenced by the very short combustion dura- 12
tions. The combustion duration remained nearly independent of 200 kPa 240 kPa 280 kPa
10
the IVO variation when the EVO was set to 130 CA ATDC. In com-
IVO 135, EVO 130

IVO 140, EVO 130

IVO 145, EVO 130

IVO 150, EVO 130

IVO 140, EVO 120

IVO 140, EVO 125

IVO 140, EVO 135

IVO 140, EVO 140


IVO 130, EVO 130

parison, the EVO had a more pronounced effect on the combustion


duration as shown by the earliest EVO (120 CA ATDC) producing
the shortest burning duration.
Figs. 15 and 16 show the spark timings set for MBT (coloured
symbols) or KLS (grey symbols) at 800 rpm and 1500 rpm, above
which CAI combustion took place and the spark timing had little
effect. It is noted that the presence of KLS at 1500 rpm was about Fig. 14. Combustion duration for 200 kPa, 240 kPa and 280 kPa intake pressures.
126 M. Dalla Nora, H. Zhao / Applied Energy 159 (2015) 117131

-10 340
1500 RPM 3000 RPM
300
-13
260
-16
220

-19 2200 RPM


420
Spark timing
(CA ATDC)

-22 360

800 RPM
-2 300

-6 240

-10 1500 RPM

ISFC (g/kWh)
450
-14
400
-18
120 kPa 160 kPa 350
-22
IVO 130, EVO 130

IVO 140, EVO 130

IVO 140, EVO 125

IVO 140, EVO 135


IVO 135, EVO 130

IVO 145, EVO 130

IVO 150, EVO 130

IVO 140, EVO 120

IVO 140, EVO 140

300

250

120 kPa 160 kPa 800 RPM


440
200 kPa 240 kPa
Fig. 15. Spark timings set for MBT (coloured symbols) or KLS (grey symbols) for 400 280 kPa
120 kPa and 160 kPa intake pressures. (For interpretation of the references to colour
in this figure legend, the reader is referred to the web version of this article.) 360

320

280

1500 RPM 240


-2
IVO 130, EVO 130

IVO 135, EVO 130

IVO 140, EVO 130

IVO 145, EVO 130

IVO 150, EVO 130

IVO 140, EVO 120

IVO 140, EVO 125

IVO 140, EVO 135

IVO 140, EVO 140


-7

-12

-17
Fig. 17. Indicated specific fuel consumption.
Spark timing
(CA ATDC)

-22

800 RPM
-2
11
-6
Effective compression/
expansion ratio

-10 10

-14
9
-18
200 kPa 240 kPa 280 kPa
8 Effective compression ratio
-22
IVO 130, EVO 130

IVO 135, EVO 130

IVO 140, EVO 130

IVO 145, EVO 130

IVO 150, EVO 130

IVO 140, EVO 120

IVO 140, EVO 125

IVO 140, EVO 135

IVO 140, EVO 140

Effective expansion ratio


7
IVO 130, EVO 130

IVO 135, EVO 130

IVO 140, EVO 130

IVO 145, EVO 130

IVO 150, EVO 130

IVO 140, EVO 120

IVO 140, EVO 125

IVO 140, EVO 135

IVO 140, EVO 140

Fig. 16. Spark timings set for MBT (coloured symbols) or KLS (grey symbols) for
200 kPa, 240 kPa and 280 kPa intake pressures. (For interpretation of the references
to colour in this figure legend, the reader is referred to the web version of this
article.) Fig. 18. Effective compression and expansion ratios.
M. Dalla Nora, H. Zhao / Applied Energy 159 (2015) 117131 127

two-stroke engine achieved 276 g/kW h at 1500 rpm and 0.8 MPa work by the higher EER did not result in improved ISFC as a result
IMEP (1.6 MPa equivalent in the four-stroke cycle). This repre- of the lowest ECR amongst all points. Similarly, the case IVO 140,
sented 9% higher fuel consumption for the particularly optimised EVO 120, which presented the highest ECR, could not achieve the
valve timing IVO 150, EVO 130. highest efficiency at 800 and 1500 rpm due to retarded spark tim-
It could be observed that the ISFC was intrinsically linked to the ing and combustion phasing. At higher speeds, when CAI combus-
expansion work, charging efficiency and in-cylinder mixture com- tion prevailed, this increment in effective compression ratio
position and combustion. Therefore, there was a trade-off between ensured the lowest ISFC of 246 g/kW h.
higher scavenging rates through exhaust blow-down with early The most significant cause for the change in ISFC as a function of
EVO and higher expansion works achieved with late EVO. This valve timings was found from the combustion efficiency plots in
effect was demonstrated in the PV diagram in Fig. 10, though it Fig. 19. It can be seen that the combustion efficiency results mir-
is clearly seen in Fig. 18 where the effective compression ratio rored those of ISFC presented in Fig. 17. The highest combustion
(ECR) and effective expansion ratio (EER) were plotted as a func- efficiencies and lowest ISFCs occurred in the middle of the graphs
tion of the valve timings. around IVO 150, EVO 130 and IVO 140, EVO 120 at 800 rpm.
Fig. 18 shows that for a given exhaust valve timing both the The combustion efficiency decreased with higher engine speeds
effective compression and expansion ratios were constant and at the same boost pressure, and it dropped with higher boost pres-
the EER was higher than the ECR by about one unit. The effective sures at each particular engine speed. Besides the fuel rich mixture
compression and expansion ratios matched each other at IVO at all operation points, which already reduced the overall combus-
140, EVO 120, and when the EVO was retarded from 120 to tion efficiency, the main reason attributed to the low values
140 CA ATDC the EER increased and the ECR was reduced. The achieved at higher speeds and loads relied on the time available
highest EER and hence highest expansion work was achieved with for mixture formation with a SOI as late as 260 CA ATDC. For
the most retarded EVO. However, such an increase in the useful instance, at 120 kPa intake pressure the combustion efficiency for

100 0.98
3000 rpm 3000 rpm

85 0.94

70 0.90

2200 rpm 2200 rpm


95 0.93

80 0.87

65 0.81

50 0.75
1500 rpm
In - cylinder lambda (-)

94
1500 rpm
0.96
efficiency (%)
Combustion

78
0.90

62 0.84

46 0.78

30 0.72

800 rpm 800 rpm


95
0.99

83 0.93

71 0.87

59 0.81
120 kPa 160 kPa 200 kPa 120 kPa 160 kPa 200 kPa
240 kPa 280 kPa 240 kPa 280 kPa
47 0.75
IVO 130, EVO 130

IVO 140, EVO 130

IVO 150, EVO 130

IVO 140, EVO 125

IVO 140, EVO 140


IVO 135, EVO 130

IVO 140, EVO 130

IVO 145, EVO 130

IVO 150, EVO 130

IVO 140, EVO 120

IVO 140, EVO 125

IVO 140, EVO 135

IVO 140, EVO 140

IVO 130, EVO 130

IVO 135, EVO 130

IVO 145, EVO 130

IVO 140, EVO 120

IVO 140, EVO 135

Fig. 19. Combustion efficiency. Fig. 20. In-cylinder lambda.


128 M. Dalla Nora, H. Zhao / Applied Energy 159 (2015) 117131

the best valve timing remained around 90% at all speeds. However, have been hindered. In this case the gas exchanging analysis
at 160 kPa it continually decreased from 94% at 800 rpm to only required the adoption of fuel rich mixtures to evaluate the air trap-
70% at 1500 rpm even at in-cylinder lambda values around 0.94, ping efficiency, so lean mixtures would have underestimated it due
as seen in Fig. 20 at IVO 140, EVO 120. to the presence of free oxygen in the exhaust.
As shown by the in-cylinder lambda values in Fig. 20, the
change in combustion efficiency with valve timings could be 3.2. Emission analysis
attributed to the variation of in-cylinder air/fuel mixture with
the gas exchanging process. The higher the relative air/fuel ratio As shown in Fig. 21, CO emission increased significantly as the
(lambda) the more complete the combustion became. The leanest mixture became richer with more advanced IVO or retarded EVO
mixture of near stoichiometric air/fuel ratio was reached at at each engine speed. Fig. 21 shows that negligible CO emission
800 rpm and resulted in a combustion efficiency of 94%. As the was produced at 800 rpm when the charging efficiency and lambda
engine speed was increased from 800 rpm to 2200 rpm, the were maximised. Based on the estimated in-cylinder lambda
decreased charging efficiencies led to richer air/fuel mixtures and results in Fig. 20, some noticeable CO emission was expected by
lower combustion efficiencies. At the lowest boost pressure of combustion of overall fuel rich mixture. The lower than expected
120 kPa, the combustion efficiency became higher at 3000 rpm CO emission could be caused by the oxidation of CO to CO2 by
(90%) than 2200 rpm (88%) mainly because of the leaner mixture the short-circuited air mixed with the burned gas during the scav-
and faster heat release rate (Fig. 13). The extremely low results enging process. As the engine speed was increased from 800 rpm
for combustion efficiency at both ends of the valve timings studied to 2200 rpm, the poorer scavenging and combustion of richer mix-
were then justified by the low in-cylinder lambda values. These tures resulted in significant increase in CO and UHC emissions at
results could have been greatly improved by reducing the fuelling higher engine speeds. In addition, the mixture was less homoge-
rate, although the solely effect of valve timing alterations would neous at higher engine speed because of the reduced time available

100 3000 RPM 3000 RPM


40

65 20

30 0

2200 RPM 2200 RPM


90
290

60
170
30

50 0

1500 RPM 1500 RPM


ISCO (g/kWh)

ISUHC (g/kWh)

270
200

180 150

100
90
50

0 0
120 kPa 160 kPa 800 RPM 120 kPa 160 kPa 800 RPM
240 200 kPa 240 kPa 100 200 kPa 240 kPa
280 kPa 280 kPa
180 75

120 50

60 25

0 0
IVO 130, EVO 130
IVO 130, EVO 130

IVO 135, EVO 130

IVO 140, EVO 130

IVO 145, EVO 130

IVO 150, EVO 130

IVO 140, EVO 120

IVO 140, EVO 125

IVO 140, EVO 135

IVO 140, EVO 140


IVO 135, EVO 130

IVO 140, EVO 130

IVO 145, EVO 130

IVO 150, EVO 130

IVO 140, EVO 120

IVO 140, EVO 125

IVO 140, EVO 135

IVO 140, EVO 140

Fig. 21. Indicated specific carbon monoxide emissions. Fig. 22. Indicated specific unburned hydrocarbon emissions.
M. Dalla Nora, H. Zhao / Applied Energy 159 (2015) 117131 129

between the end of injection and the beginning of combustion. smoke emission was noticeably more affected by the engine load
This could have contributed to the very rapid rise in CO emissions and speed than by the valve timing itself, as the fuel impingement
when the engine speed was changed from 800 rpm to 1500 rpm, as increased with longer injection durations at higher loads.
in the case of the valve timing IVO 150, EVO 130 when this The overall excessive smoke emission presented in Fig. 23
emission increased from 4 g/kW h to 43 g/kW h in the intake pres- reflected the level of charge stratification resulted from late fuel
sure range 120200 kPa. injections, besides the global fuel rich mixtures and the relatively
As shown in Fig. 22, UHC emission showed less dependency on low injection pressure used (15 MPa). Even at 800 rpm and
valve timings and lower correlation with the charging efficiency 0.67 MPa IMEP the smoke level remained around 0.55 FSN at the
and in-cylinder lambda. As late injections were employed, most best valve timing tested, while it increased by about four-times
UHC emissions were likely produced by the fuel rich combustion when the engine speed reached 1500 rpm. This was certainly a
as well as fuel impingement due to retarded injection. The UHC drawback of the two-stroke poppet valve engine, particularly when
emissions are not only dependent on the overall air/fuel ratio but comparing it to the values achieved by four-stroke GDI engines.
also on its homogeneity. As injection took place after 260 CA Tests performed by [16] in a two-cylinder downsized engine
ATDC, there was limited time available for a homogeneous mixture demonstrated FSN values around 0.15 when running at 1200 rpm
to form and very rich mixtures could be present in some regions and loads between 0.7 and 1.8 MPa IMEP. Nevertheless, such study
producing UHC emissions. In addition, at higher loads and boost employed a multi-hole centrally mounted injector working at
pressures, the end of injection could be as late as 290 CA ATDC, 20 MPa, and the time available for airfuel mixing was about three
when the piston was only at about 25 mm from the cylinder head. times longer than those realised in the present study.
Thus, the fan shaped spray impinged onto the piston and formed From the NOx emission presented in Fig. 24, more residual gas
pool fires on its top. For the same reasons, high smoke emission was trapped (lower charging efficiencies) when moving along the
was observed as seen in Fig. 23. Compared to UHC emissions, the

3000 RPM 1.0


3000 RPM
3.0
0.5
1.5

0.0 0.0

2200 RPM 2200 RPM


7.5
4.2

5.0 2.8

2.5 1.4

0.0 0.0

1500 RPM 1500 RPM


Smoke (FSN)

ISNOx (g/kWh)

5.0 11.0

4.0 8.0

3.0 5.0

2.0 2.0
120 kPa 160 kPa 800 RPM 120 kPa 160 kPa 800 RPM
2.6 200 kPa 240 kPa 200 kPa 240 kPa
280 kPa 37
280 kPa
29
1.8

21
1.0
13

0.2 5
IVO 145, EVO 130
IVO 130, EVO 130

IVO 135, EVO 130

IVO 140, EVO 130

IVO 150, EVO 130

IVO 140, EVO 120

IVO 140, EVO 125

IVO 140, EVO 135

IVO 140, EVO 140

IVO 135, EVO 130

IVO 140, EVO 130

IVO 145, EVO 130

IVO 150, EVO 130

IVO 140, EVO 120

IVO 140, EVO 125

IVO 140, EVO 135

IVO 140, EVO 140


IVO 130, EVO 130

Fig. 23. Smoke emissions. Fig. 24. Indicated specific oxides of nitrogen emissions.
130 M. Dalla Nora, H. Zhao / Applied Energy 159 (2015) 117131

x-axis from the middle to the both sides. Because of the increased  The ISFC was primarily determined by the combustion effi-
heat capacity of CO2 and reduced oxygen availability by the pres- ciency, which was directly related to the in-cylinder air/fuel
ence of recycled burned gases, the combustion temperature and ratio. The relative air/fuel ratios of the in-cylinder mixture could
hence NOx formation were significantly reduced [37]. be increased by optimisation of the valve timings for maximum
At 800 rpm the early EVO raised the charge oxygen availability charging efficiency. Furthermore, by reducing the fuelling rate
increasing NOx emissions to levels of downsized four-stroke engi- better ISFC would be expected, although it would not be possi-
nes operating at similar conditions [22]. As the speed increased ble to calculate the gas exchanging parameters using the
from 800 to 3000 rpm, the combustion mode progressed from selected approach.
SI towards CAI as a result of higher levels of hot residual gas  Compared to four-stroke downsized engines operating at simi-
trapped. Consequently, the NOx emissions progressively lar loads and speeds, the overall two-stroke poppet valve engine
decreased thanks to the higher charge dilution and lower com- efficiency was found below values normally achieved by those
bustion temperature. engines. Improved and competitive results are possible to be
From Fig. 24 it is also noted that the NOx emissions were more achieved by leaner combustion and engine operation at those
sensitive to the valve timings studied than to the load itself, espe- valve timings which demonstrated better charging efficiencies
cially at 800 rpm. At this speed the emission of oxides of nitrogen i.e. IVO 150, EVO 130 and IVO 140, EVO 120.
increased by 20% as the boost pressure was changed from 120 to  The CO emissions were directly affected by the in-cylinder
280 kPa (0.661.22 MPa IMEP). In comparison, by retarding the lambda. At 800 rpm, negligible CO emission was measured with
IVO in 10 CA from 130 to 140 CA ATDC the NOx emissions nearly optimised valve timings, although at higher speeds these values
doubled. The spark timing also played an important role in NOx increased due to the lower temperature combustion.
emissions, as shown by the point IVO 140, EVO 120 at 200 kPa  Besides the overall fuel rich combustion, uHC and smoke emis-
boost. The ignition timing in this case had to be retarded to avoid sions were found also affected by fuel impingement considering
knocking combustion (Fig. 16), reducing the in-cylinder peak tem- the late SOI adopted. At 1500 rpm and 120/160/200 kPa, these
perature and NOx production. emissions had a lighter variation than that found for CO under
At 2200 and 3000 rpm and intake pressure of 120 kPa, pure CAI constant fuelling rate and regardless the valve timing used.
combustion took place. At 2200 rpm the NOx emissions rose  NOx emission was found very low at higher engine speeds
rapidly as the boost pressure was increased from 120 kPa to when there was high residual gas concentration and CAI com-
160 kPa, as a result of both reduced residual gas concentration bustion. At high loads and low speeds, although, these values
and presence of high temperature flame in the spark assisted CAI increased due to the greater charge oxygen concentration and
combustion. higher combustion temperatures.

4. Conclusions The above results have demonstrated that the scavenging pro-
cess and fuel preparation are the two most important issues affect-
In this study, a four-valve direct injection gasoline engine was ing the two-stroke poppet valve engines performance and
operated in the two-stroke cycle mode by opening both the intake emissions. The scavenging process can be further optimised by
and exhaust valves around BDC. The exhaust gas was scavenged by additional experiments with different valve opening durations,
compressed air during the valve overlap period. At each engine particularly at higher engine speeds.
speed and boost pressure, the engine output was measured as a After the acquisition of all gas exchanging parameters under
function of intake and exhaust valve timings. The results can be fuel rich conditions, additional tests can be performed with lean
summarised as follows: combustion so that improved ISFC results can be achieved. In this
case the delivery ratio needs to be kept constant so the in-cylinder
 At 800 rpm the peak IMEP of 1.2 MPa was achieved at an intake lambda can be evaluated at the same air and fuel trapping efficien-
pressure of 280 kPa, producing a specific torque of 195 N m/L cies. At fuel lean conditions the mixture preparation process can be
with the in-cylinder peak pressure as low as 6.8 MPa. At each further improved by higher injection pressures and/or split injec-
engine speed, the maximum IMEP was obtained with the high- tions. A more robust stratified charge combustion system design,
est charging efficiency. As the engine speed was increased, the such as a centrally mounted fast DI injector, would be also
maximum output was limited by the scavenging process and desirable.
violent heat release rate. At 2200 rpm and 3000 rpm the maxi-
mum loads were limited to 0.55 and 0.25 MPa IMEP, respec- Acknowledgement
tively, which is equivalent to 1.1 and 0.5 MPa IMEP in the
four-stroke cycle. The first author would like to acknowledge the Brazilian council
 For the given valve opening duration and valve lift, the maxi- for scientific and technological development (CNPq Brasil) for
mum charging efficiency of 95% could be achieved at 800 rpm. supporting his PhD at Brunel University London.
At any given IVO and EVO timings the charging efficiency
dropped with the increased engine speed due to the reduced
References
time available for gas exchanging. At 2200 rpm the charging
efficiency was limited to 45%, while at 3000 rpm it dropped to [1] Hooper PR, Al-Shemmeri T, Goodwin MJ. Advanced modern low-emission two-
about 20%. stroke cycle engines. Proc Instit Mech Eng, Part D: J Automob Eng
 At 120 kPa intake pressure, the trapping efficiency increased 2011;225:153143. http://dx.doi.org/10.1177/0954407011408649.
[2] Hundleby G. Development of a poppet-valved two-stroke engine the flagship
from about 35% to 70% with higher engine speeds as the air concept. SAE Technical Paper, 1990; 900802. http://dx.doi.org/10.4271/
short-circuiting rate was reduced. At low speeds and long valve 900802.
overlaps, although, the trapping efficiency was reduced to val- [3] Kenny RG. Developments in two-stroke cycle engine exhaust emissions. Proc
Instit Mech Eng, Part D: J Automob Eng 1992;206:93106. http://dx.doi.org/
ues as low as 15%. 10.1243/PIME_PROC_1992_206_165_02.
 At 800 rpm and 1500 rpm the heat release process was domi- [4] Blair G. Design and simulation of two-stroke engines. Warrendale: Society of
nated by spark ignited flame propagation combustion. At higher Automotive Engineers; 1996, ISBN 978-1-56091-685-7.
[5] Stokes J, Hundleby G, Lake T, Christie M. Development experience of a poppet-
engine speeds, CAI combustion took place and the spark timing valved two-stroke flagship engine. SAE Technical Paper, 1992; 920778. http://
had little effect. dx.doi.org/10.4271/920778.
M. Dalla Nora, H. Zhao / Applied Energy 159 (2015) 117131 131

[6] Nomura K, Nakamura N. Development of a new two-stroke engine with [22] Zhao H. Advanced direct injection combustion engine technologies and
poppet-valves: Toyota S-2 engine, a new generation of two-stroke engines for development, vol. 1. Cambridge: Woodhead Publishing; 2010, ISBN
the future? Proceedings of the IFP international seminar, Paris, France, 1993. p. 9781845693893.
5362. [23] Zhang Y, Ojapah M, Cairns A, Zhao H. 2-Stroke CAI Combustion Operation in a
[7] Knoll R. AVL Two-Stroke Diesel Engine. SAE Technical Paper, 1998; 981038. GDI Engine with Poppet Valves. SAE Technical Paper, 2012; 2012-01-1118.
http://dx.doi.org/10.4271/981038. http://dx.doi.org/10.4271/2012-01-1118.
[8] Nakano M, Sato K, Ukawa H. A two-stroke cycle gasoline engine with poppet [24] Douglas R. AFR and emissions calculations for two-stroke cycle engines. SAE
valves on the cylinder head. SAE Technical Paper 1990; 901664, http://dx.doi. Technical Paper, 1990; 901599. http://dx.doi.org/10.4271/901599.
org/10.4271/901664. [25] Heywood JB. Internal combustion engines fundamentals. New York: McGraw-
[9] Sementa P, Vaglieco BM, Catapano F. Thermodynamic and optical Hill; 1988, ISBN 978-0071004992.
characterizations of a high performance GDI engine operating in [26] United Nations Regulation 49. Uniform provisions concerning the measures to
homogeneous and stratified charge mixture conditions fuelled with gasoline be taken against the emission of gaseous and particulate pollutants from
and bio-ethanol. Fuel 2012;96:20419. http://dx.doi.org/10.1016/ compression-ignition engines and positive ignition engines for use in vehicles.
j.fuel.2011.12.068. Off J Eur Union, 2013. <http://eur-lex.europa.eu/legal-content/EN/TXT/?uri=
[10] Pohorelsky L, Brynych P, Macek J, Vallaude P, et al. Air system conception for a CELEX:42013X0624(01)>.
downsized two-stroke diesel engine. SAE Technical Paper, 2012; 2012-01- [27] Xu R. A Convenient technique for determining two-stroke emissions
0831. http://dx.doi.org/10.4271/2012-01-0831. measurements accuracy and A/F ratio. SAE Technical Paper, 1996; 961804.
[11] Cairns A, Zhao H, Todd A, Aleiferis P. A study of mechanical variable valve http://dx.doi.org/10.4271/961804.
operation with gasolinealcohol fuels in a spark ignition engine. Fuel [28] Andwari AM, Aziz AA, Said MFM, Latiff ZA. Experimental investigation of the
2013;106:80213. http://dx.doi.org/10.1016/j.fuel.2012.10.041. influence of internal and external EGR on the combustion characteristics of a
[12] Osborne R, Li G, Sapsford S, Stokes J. et al. Evaluation of HCCI for future controlled auto-ignition two-stroke cycle engine. Appl Energy 2014;134:110.
gasoline powertrains. SAE Technical Paper, 2003; 2003-01-0750. http://dx.doi. http://dx.doi.org/10.1016/j.apenergy.2014.08.006.
org/10.4271/2003-01-0750. [29] Wang Z, Qi Y, He X, Wang J, Shuai S, Law CK. Analysis of pre-ignition to super-
[13] Zhang Y, Zhao H. Investigation of combustion, performance and emission knock: Hotspot-induced deflagration to detonation. Fuel 2015;144:2227.
characteristics of 2-stroke and 4-stroke spark ignition and CAI/HCCI operations http://dx.doi.org/10.1016/j.fuel.2014.12.061.
in a DI gasoline. Appl Energy 2014;130:24455. http://dx.doi.org/10.1016/j. [30] Wetzel P. Downspeeding a light duty diesel passenger car with a combined
apenergy.2014.05.036. supercharger and turbocharger boosting system to improve vehicle drive cycle
[14] Tribotte P, Ravet F, Dugue V, Obernesser P, et al. Two strokes diesel engine fuel economy. SAE Technical Paper, 2013; 2013-01-0932. http://dx.doi.org/10.
promising solution to reduce CO2 emissions. Procedia Soc. Behav. Sci. 4271/2013-01-0932.
2012;48:2295314. http://dx.doi.org/10.1016/j.sbspro.2012.06.1202. [31] Lumsden G, OudeNijeweme D, Fraser N, Blaxill H. Development of a
[15] Martin S, Beidl C, Mueller R. Responsiveness of a 30 bar BMEP 3-cylinder turbocharged direct injection downsizing demonstrator engine. SAE Int J
engine: opportunities and limits of turbocharged downsizing. SAE Technical Engine 2009;2(1):142032. http://dx.doi.org/10.4271/2009-01-1503.
Paper, 2014; 2014-01-1646. http://dx.doi.org/10.4271/2014-01-1646. [32] Li N, Xie H, Chen T, Li L, Zhao H. The effects of intake backflow on in-cylinder
[16] Eichhorn A, Lejsek D, Hettinger A, Kufferath A. Challenge determining a situation and auto ignition in a gasoline controlled auto ignition engine. Appl
combustion system concept for downsized SI-engines comparison and Energy 2013;101:75664. http://dx.doi.org/10.1016/j.apenergy.2012.07.050.
evaluation of several options for a boosted 2-cylinder SI-engine. SAE Technical [33] Benson RS, Brandham PT. A method for obtaining a quantitative assessment of
Paper, 2013; 201301-1730. http://dx.doi.org/10.4271/2013-01-1730. the influence of charging efficiency on two-stroke engine performance. Int J
[17] Zhang Y, Zhao H, Ojapah M, Cairns A. CAI combustion of gasoline and its Mech Sci 1969;11:30312. http://dx.doi.org/10.1016/0020-7403(69)90048-4.
mixture with ethanol in a 2-stroke poppet valve DI gasoline engine. Fuel [34] Xie H, Li L, Chen T, Yu W, Wang X, Zhao H. Study on spark assisted compression
2013;109:6618. http://dx.doi.org/10.1016/j.fuel.2013.03.002. ignition (SACI) combustion with positive valve overlap at mediumhigh load.
[18] Benajes J, Molina S, Novella R, De Lima D. Implementation of the partially Appl Energy 2013;101:62233. http://dx.doi.org/10.1016/j.apenergy.2012.
premixed combustion concept in a 2-stroke HSDI diesel engine fuelled with 07.015.
gasoline. Appl Energy 2014;122:94111. http://dx.doi.org/10.1016/j. [35] Ortiz-Soto EA, Lavoie GA, Martz JB, Wooldridge MS, Assanis DN. Enhanced heat
apenergy.2014.02.013. release analysis for advanced multi-mode combustion engine experiments.
[19] Benajes J, Novella R, De Lima D, Tribott P. Analysis of combustion concepts in Appl Energy 2014;136:46579. http://dx.doi.org/10.1016/j.apenergy.2014.
a newly designed two-stroke high-speed direct injection compression ignition 09.038.
engine. Int J Engine Res 2015;16:5267. http://dx.doi.org/10.1177/ [36] Li Y, Zhao H, Stansfield P, Freeland, P. Synergy between boost and valve
1468087414562867. timings in a highly boosted direct injection gasoline engine operating with
[20] Osborne R, Stokes J, Lake T, Carden P, et al. Development of a two-stroke/four- miller cycle. SAE Technical Paper, 2015; 2015-01-1262. http://dx.doi.org/10.
stroke switching gasoline engine the 2/4SIGHT Concept. SAE Technical Paper 4271/2015-01-1262.
2005; 2005-01-1137. http://dx.doi.org/10.4271/2005-01-1137. [37] Zhao H. HCCI and CAI engines for the automotive industry. Cambridge:
[21] Sadakane S, Sugiyama M, Kishi H, Abe S, et al. Development of a New V-6 high Woodhead Publishing; 2007, ISBN 978-1-84569-128-8.
performance stoichiometric gasoline direct injection engine. SAE Technical
Paper, 2005; 2005-01-1152. http://dx.doi.org/10.4271/2005-01-1152.

You might also like