Você está na página 1de 17

Applied Energy 165 (2016) 166–182

Contents lists available at ScienceDirect

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

Experimental analysis of ethanol dual-fuel combustion in a heavy-duty


diesel engine: An optimisation at low load
Vinícius B. Pedrozo ⇑, Ian May, Macklini Dalla Nora, Alasdair Cairns, 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

 Dual-fuel combustion offers promising results on a stock heavy-duty diesel engine.


 The use of split diesel injections extends the benefits of the dual-fuel mode.
 Ethanol–diesel dual-fuel combustion results in high indicated efficiencies.
 NOx and soot emissions are significantly reduced.
 Combustion efficiency reaches 98% with an ethanol energy ratio of 53%.

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

Article history: Conventional diesel combustion produces harmful exhaust emissions which adversely affect the air qual-
Received 10 August 2015 ity if not controlled by in-cylinder measures and exhaust aftertreatment systems. Dual-fuel combustion
Received in revised form 7 November 2015 can potentially reduce the formation of nitrogen oxides (NOx) and soot which are characteristic of diesel
Accepted 14 December 2015
diffusion flame. The in-cylinder blending of different fuels to control the charge reactivity allows for
lower local equivalence ratios and temperatures. The use of ethanol, an oxygenated biofuel with high
knock resistance and high latent heat of vaporisation, increases the reactivity gradient. In addition,
Keywords:
renewable biofuels can provide a sustainable alternative to petroleum-based fuels as well as reduce
Dual-fuel combustion
Ethanol
greenhouse gas emissions. However, ethanol–diesel dual-fuel combustion suffers from poor engine effi-
Split diesel injections ciency at low load due to incomplete combustion. Therefore, experimental studies were carried out at
Engine-out emissions 1200 rpm and 0.615 MPa indicated mean effective pressure on a heavy-duty diesel engine. Fuel delivery
Combustion losses was in the form of port fuel injection of ethanol and common rail direct injection of diesel. The objective
Low load was to improve combustion efficiency, maximise ethanol substitution, and minimise NOx and soot emis-
sions. Ethanol energy fractions up to 69% were explored in conjunction with the effect of different diesel
injection strategies on combustion, emissions, and efficiency. Optimisation tests were performed for the
optimum fuelling and diesel injection strategy. The resulting effects of exhaust gas recirculation, intake
air pressure, and rail pressure were investigated. The optimised combustion of ethanol ignited by split
diesel injections resulted in higher net indicated efficiency when compared to diesel-only operation.
For the best emissions case, NOx and soot emissions were reduced by 65% and 29%, respectively.
Aftertreatment requirements that are generally associated with cost and fuel economy penalties can
be minimised. Combustion efficiency of 98% was achieved at the expense of higher NOx emissions.
Ó 2015 Elsevier Ltd. All rights reserved.

1. Introduction capability, reliability, as well as superior fuel conversion efficiency


[1]. However, conventional diesel combustion incurs a wide range
Heavy-duty (HD) diesel engines have been widely utilised in on of local in-cylinder equivalence ratios and temperatures which can
and off-road transportation sectors due to their high torque result in NOx and soot formation [2]. These emissions can
adversely affect the air quality if not controlled by exhaust
⇑ Corresponding author at: Brunel University London, Centre for Advanced aftertreatment technologies and in-cylinder measures. Several
Powertrain and Fuels Research (CAPF), Kingston Lane, Uxbridge, Middlesex UB8 combustion concepts have been developed, arising from costly
3PH, United Kingdom. Tel.: +44 7473361328; fax: +44 1895266698. aftertreatment systems and strict fuel efficiency and emissions
E-mail addresses: vinicius.pedrozo@brunel.ac.uk, pedrozo.vinicius@gmail.com regulations [3]. In addition, according to economic growth
(V.B. Pedrozo).

http://dx.doi.org/10.1016/j.apenergy.2015.12.052
0306-2619/Ó 2015 Elsevier Ltd. All rights reserved.
V.B. Pedrozo et al. / Applied Energy 165 (2016) 166–182 167

projections, it is predicted an increase in the demand for petroleum elevated knock resistance and latent heat of vaporisation of the
and other energy sources by more than 30% from 2010 to 2040 [4]. ethanol allow the use in high compression ratio and highly boosted
This may result in elevated prices for liquid fuels as well as engines [30]. Moreover, early dual-fuel results obtained from an
compromise their cost competitiveness, opening opportunities optical engine showed that ethanol can suppress soot formation
for improved sustainability and greenhouse gas (GHG) emissions in high temperature regions of the conventional diesel combustion
reduction via biofuels [5]. chamber [31]. Recent experimental analyses with ethanol–diesel
The alternative combustion technologies are generally centred DF combustion demonstrated noticeable NOx reductions at engine
on improved fuel atomisation and mixture preparation, lower local loads above 0.8 or 1.0 MPa net indicated mean effective pressure
equivalence ratios, reduced peak in-cylinder temperatures, and (IMEP) [32–36].
faster burn rates. This is usually referred to Low Temperature Com- Asad et al. [37] investigated the load range of ethanol–diesel
bustion (LTC) [3]. Among the combustion strategies proposed is low temperature combustion using a single cylinder light-duty
Homogeneous Charge Compression Ignition (HCCI). This is charac- engine. The concept, named Premixed Pilot Assisted Combustion
terised by early fuel injections promoting a fully premixed charge, (PPAC), utilised high EGR levels and single diesel injections near
long ignition delays, and short combustion durations. However, the firing top dead centre (TDC). The main challenge encountered
lack of direct control of ignition timing and combustion phasing, was the elevated levels of unburnt HC and CO emissions at low
particularly under transient conditions, is still the major drawback. loads, as observed by other ethanol dual-fuel combustion studies
It also exhibits elevated combustion losses, combustion noise, and [38,39]. Asad et al. [37] attributed these losses to the resistance
sensitivity to temperature [6–8]. In comparison, some slightly of ethanol to auto-ignition and proposed an alternative combus-
more heterogeneous combustion concepts have been developed. tion strategy to enable clean combustion and higher efficiencies
Premixed Charge Compression Ignition (PCCI) [9–12], Partially Pre- at these specific conditions. From idle to low loads, the engine
mixed Charge Compression Ignition (PPCI) [13], Modulated Kinet- would operate under conventional diesel combustion, utilising
ics (MK) [14], and Uniform Bulky Combustion System (UNIBUS) high levels of EGR and boost combined with retarded injections
[15] name a few. These allow a higher degree of combustion phas- and elevated rail pressures. After the engine reaches a certain load,
ing control at low and medium loads while maintaining low soot the combustion would switch to ethanol dual-fuel combustion.
and NOx emissions. However, these less premixed combustion However, high levels of EGR might not be feasible and would place
modes tend to suffer from lower indicated efficiency, increased greater demand on the boosting system to maintain the required
unburnt hydrocarbons (HC) and carbon monoxide (CO) emissions, equivalence ratio.
and limited load range due to high exhaust gas recirculation (EGR) Sarjovaara et al. [40] studied the effect of diesel injection
and boost requirements. parameters on ethanol dual-fuel combustion using a modified
Gasoline Direct Injection Compression Ignition (GDCI) [16,17] six-cylinder diesel engine with a compression ratio of 14.2:1. The
and Partially Premixed Combustion (PPC) [18–20] are some alter- use of small pilot injections corresponding to approximately 10–
natives to diesel LTC. They expand the high efficiency window 20% of the total diesel fuel injected helped maintaining acceptable
and achieve very low NOx emissions operating up to full load with PRR levels while running with high ethanol percentages. However,
moderate-high EGR rates. As these concepts utilise gasoline, they the maximum ethanol substitution ratio was limited to only 39% at
do not reduce the dependence on liquid fossil fuels. They also 25% load, and exhaust gas emissions were neglected.
require engine hardware modifications such as the piston and Considering the previous works, an experimental study was
injection system, and ignition or lubricant improvers, depending carried out on a single cylinder HD engine. The combustion system
on the fuel selected. Some drawbacks regarding soot levels at remained stock, with a re-entrant piston bowl design and a geo-
higher loads, due to low air–fuel ratio, accompanied with metric compression ratio of 16.8:1. Ethanol was port fuel injected
significant CO and HC emissions at low loads are also reported. while diesel fuel was directly injected into the cylinder. The objec-
Recent PPC studies with renewable fuels, including ethanol, have tive was to reduce combustion losses and maximise the use of
demonstrated high thermal efficiency and further soot reductions ethanol while maintaining low levels of NOx and soot emissions.
[21–23]. However, high acoustic noise and elevated peak heat The diesel injection strategy tested uses a pre-injection to adjust
release rates have been experienced due to a fast premixed com- mixture flammability and reduce PRR, and one injection around
bustion, requiring lower intake air pressures and larger amounts TDC to maintain combustion control. This concept is slightly differ-
of EGR, which reduce combustion efficiency [24]. ent from conventional dual-fuel using a single diesel injection near
Dual-fuel (DF) combustion, such as Reactivity Controlled Com- TDC for ignition [33–37,41]. The strategy also differs from the early
pression Ignition (RCCI) [25–27], have been developed to overcome diesel injections utilised in RCCI combustion [25,26,28].
the majority of the previously mentioned issues. The concept uses The experiments were performed at 1200 rpm and 0.615 MPa
different fuels to control the in-cylinder reactivity gradient while IMEP, with varying ethanol energy fractions up to 69%. The impact
achieving a wide operating range with near zero levels of NOx of different diesel injection strategies on combustion, emissions,
and soot, acceptable pressure rise rate (PRR), and very high indi- and efficiency were explored. Pre-injections corresponding to up
cated efficiency [28]. The primary method of fuel delivery is the to 60% of the total diesel fuel injected were evaluated without
port fuel injection of a low reactivity fuel (i.e. gasoline, alcohol, EGR. Subsequent investigation of the pre-injection timing and
propane, natural gas, etc.) to create a well-mixed charge of fuel– quantity was performed for the optimum fuelling and injection
air-EGR. The high reactivity fuel (i.e. diesel) serves as the ignition strategy using an EGR rate of 25%. Finally, the effect of higher
source and is directly injected into the combustion chamber. How- intake air pressure and diesel injection pressure were explored.
ever, RCCI is sensitive to variations in the intake air temperature The best dual-fuel results were compared against diesel-only
and pressure. This is expected as the combustion is sufficiently operation.
premixed and governed by chemical kinetics [25]. Furthermore,
the combustion phasing is generally controlled by varying the fuel
reactivity (i.e. substitution ratio), which might not be the optimum 2. Experimental setup
at certain engine loads.
Ethanol is attractive as a low reactivity fuel because it can be The experiments were carried out on a single cylinder HD diesel
produced from biomass and can offset the demand for engine equipped with a high pressure common rail diesel injection
petroleum-based fuels in internal combustion engines [29]. The system, representing the engine of a modern heavy goods vehicle
168 V.B. Pedrozo et al. / Applied Energy 165 (2016) 166–182

(i.e. truck). The test cell layout and main engine specifications are Table 1
depicted in Fig. 1 and Table 1, respectively. Measurement device Single cylinder HD diesel engine specifications.

specifications are shown in the Appendix A. Two large-volume Parameter Value


surge tanks were installed to damp out pressure fluctuations in Bore 129 mm
the intake and exhaust manifolds. Fresh intake air was supplied Stroke 155 mm
to the engine via an external compressor system with closed loop Swept volume 2026 cm3
control for the boost pressure. An intake throttle provided fine con- Geometric compression ratio 16.8:1
Maximum in-cylinder pressure 18 MPa
trol over the intake manifold air pressure. The fresh air flow rate Piston type Re-entrant bowl
was measured with a thermal mass flow meter. High-pressure loop Number of valves 4
cooled external EGR was supplied to the engine via an EGR valve Diesel injection system Common rail, inj. pressure of 50–220 MPa,
and an electronically controlled exhaust back pressure valve centrally mounted diesel injector, 8 holes
Ethanol injection system PFI peak-and-hold Marelli IWP069,
located downstream of the exhaust surge tank.
included spray angle of 15°
Auxiliary equipment such as the high-pressure diesel pump
(HPP) and the engine coolant and oil pumps are not coupled to
the engine but driven by separate electric motors. Coolant and
oil temperatures were kept within 353 ± 5 K. Oil pressure was set Table 2
Fuel properties.
to 350 ± 10 kPa throughout the experiments. An independent
low-pressure system supplied diesel to the common rail injection Characteristic Diesel Ethanol
system. Two Coriolis flow meters were used to measure the diesel Product Gasoil (Ultra Low Sulphur) Ethyl alcohol
flow rate (m _ diesel ) by considering the total fuel supplied to and from Density at 293 K 827 kg/m3 789 kg/m3
the injector and HPP. Ethanol was injected into the intake port Cetane number 45 –
RON [42] – 107
through a high flow-rate peak-and-hold port fuel injector (PFI). It
Alcohol content NA 99.1–99.5% (v/v)
was mounted into the intake air manifold so that the spray was Water content <0.2 g/kg <1.14 (w/w)
directed towards the back of the intake valves, located approxi- Boiling point/range 443–643 K 351 K
mately 0.3 m downstream. Heat of vaporisation [42] 270 kJ/kg 840 kJ/kg
An in-house injector driver controlled the PFI pulse width, Carbon content 86.6% 52.1%
Hydrogen content 13.2% 13.1%
adjusted according to the desired ethanol substitution ratio. The Oxygen content 0.2% 34.8%
ethanol start of injection (SOI) was set to the firing TDC to max- LHV 42.9 MJ/kg 26.9 MJ/kg
imise the time for air–fuel mixture preparation before the intake
valve opening event. The ethanol mass flow rate (m _ ethanol ) was
obtained from the injector calibration curve. Ethanol injection emissions [43] and confirmed by the air and fuel flow rates. The
pressure was continuously monitored by a pressure transducer, lower heating value of the DF combustion mode (LHVDF) was calcu-
so that a constant delta pressure of 300 ± 10 kPa could be main- lated using Eq. (1).
tained across the injector. A heat exchanger held the fuel temper- _ ethanol LHVethanol þ m
m _ diesel LHVdiesel
ature constant at 293 ± 5 K. The relevant properties of the fuel used LHVDF ¼ ð1Þ
_ ethanol þ m
ðm _ diesel Þ
in this work are listed in Table 2.
The stoichiometric air/fuel ratio was determined by the conser- The ethanol substitution ratio (EF) was calculated on an energy
vation of mass of each chemical element in the reactants [42]. The input basis, defined as the ratio of the energy content of the etha-
global equivalence ratio was calculated based on the engine-out nol to the total fuel energy (Eq. (2)).

Fig. 1. Schematic diagram of the engine experimental setup.


V.B. Pedrozo et al. / Applied Energy 165 (2016) 166–182 169

_ ethanol LHVethanol
m engine operation condition is close to operation points #3 and
EF ¼ ð2Þ
_ ethanol LHVethanol þ m
m _ diesel LHVdiesel #8 of the World Harmonized Stationary Cycle (WHSC) and #7 of
the former European Stationary Cycle (ESC13) for HD engines.
The in-cylinder pressure was measured by a piezoelectric pres-
Fig. 2 shows where the test point (white circle with black dots pat-
sure sensor. Intake and exhaust pressures were measured by two
tern) is located over an estimated speed and load map of a HD die-
water cooled piezoresistive absolute pressure sensors. The intake
sel engine. The WHSC and ESC13 test cycle points are also
valve lift profile was obtained by measuring the displacement of
displayed. The bigger the circle, the higher is the relative weight
the valve spring retainer with a displacement sensor. Tempera-
over the test cycle. The selected test point is located in a high res-
tures and pressures at relevant locations were measured by K-
idency area of a typical HD diesel engine drive cycle. It also repre-
type thermocouples and pressure gauges, respectively. Two
sents an engine load where dual-fuel combustion is adversely
National Instruments data acquisition (DAQ) cards were used to
affected by elevated CO and unburnt HC emissions. The main goal
acquire the signals from the measurement device. While a high
of this study was to mitigate combustion losses and improve effi-
speed DAQ card received the crank angle resolved data synchro-
ciency while achieving NOx and soot levels close to Euro VI legis-
nised with an optical encoder of 0.25 crank angle degrees (CAD)
lation emissions limits (0.40 and 0.01 g/kW h, respectively)
resolution, a lower speed DAQ card acquired the low frequency
utilising low levels of intake air pressure and EGR.
engine operation conditions. These data were collected and dis-
The engine is equipped with a prototype variable valve actua-
played live by an in-house DAQ program. The pumping work was
tion system (VVA). Variable intake valve closing timing (IVC) and
included in all the calculations, such as IMEP and net indicated effi-
the resulting effective compression ratio (ECR) can be selected dur-
ciency (g_ind). The apparent net heat release rate (HRR), denoted
ing the engine operation. The intake valve opening and closing tim-
by (dQn/dt), was calculated using the following well-known Eq. (3):
ings were set at 367 CAD and 150 CAD after firing top dead centre
dQ n c dV 1 dp (ATDC), respectively, at 0.5 mm valve lift. This valve timing pro-
¼ p þ V ð3Þ
dt c  1 dt c  1 dt vided an ECR of approximately 16.0:1. The average PRR and COV_-
IMEP limits were set to 2 MPa/CAD and 5%, respectively.
where c is the ratio of specific heats, t is time, and V and p stand for
in-cylinder volume and pressure, respectively. Since the absolute
value of heat released is not as important to this study as the bulk 4. Results and discussion
shape of the curve with respect to crank angle, a c of 1.33 was
assumed. Combustion phasing (CA50) was determined by the crank 4.1. Conventional diesel combustion baseline
angle position of 50% mass fraction burnt (MFB). Ignition delay was
defined as the period of time between the diesel start of injection The primary objective of these tests was to obtain the best indi-
and start of combustion (SOC), set to 0.3% MFB point of the average cated specific fuel consumption (ISFC)/NOx/soot trade-off in
cycle. Cycle-to-cycle variability was measured by the coefficient of diesel-only operation by means of moderate amounts of EGR, ele-
variation of the IMEP (COV_IMEP), defined as the ratio of the stan- vated injection pressures, and optimised SOI. The conventional die-
dard deviation in IMEP and the mean IMEP over the sampled cycles. sel baseline was performed at four different EGR rates of 0%, 10%,
For the sake of simplification, the average in-cylinder gas tempera- 21%, and 25%. EGR temperature varied from 296 to 354 K as its
ture was calculated by applying the ideal gas model, considering ratio increased. Intake manifold air temperature also rose from
each species in the mixture. 292 K with no EGR to 306 K at 25% EGR. Two intake air pressures
Exhaust emissions were measured by a Horiba MEXA-7170 of 103 and 125 ± 1 kPa were included as a reference to the DF
DEGR emission analyser equipped with a heated line and a high experiments. A 10 kPa difference between intake air and exhaust
pressure module, allowing high-pressure sampling. The EGR rate gas back pressure was applied throughout the tests to maintain
was calculated by the ratio of intake and exhaust CO2 concentra- consistent pumping losses. The maximum COV_IMEP observed
tions measured by the same analyser. According to [44–46], deter- during these tests was 1.8%.
mining the actual hydrocarbons emissions measured by the flame Rail pressure (RP) was increased from 70 to 125 MPa as larger
ionisation detector (FID) needs to be calibrated for oxygenated percentages of EGR were added. A single injection strategy could
compounds such as ethanol. This is due to relative insensitivity only be applied in some cases in order to keep the PRR under the
of the equipment towards alcohols and aldehydes. Therefore, the limit of 2 MPa/CAD. A small pilot injection of approximately
FID response to ethanol was corrected by the method developed 3 mm3, with a dwell timing of 1 ms (7.2 CAD) and a split ratio of
in [45] with an updated factor of 0.68 [46]. This correction uses a
second order polynomial and the volumetric ethanol content as
an input. Smoke was measured by an AVL 415SE Smoke Meter.
The results were converted from FSN to mg/m3, according to
[47]. The calculation of specific exhaust gas emissions was based
on the UN Regulation number 49, with NOx and CO emissions cor-
rected to a wet basis [48]. Finally, combustion efficiency (gc) was
calculated using the following Eq. (4):
ðISCO LHV þ ISHC LHV Þ P
gc ¼ 1  _ CO DF i
_ diesel LHVdiesel Þ
ð4Þ
ðmethanol LHVethanol þ m
where ISCO and ISHC represent the net indicated specific emissions
of CO and unburnt HC in g/kW h, respectively, and Pi represents the
net indicated power.

3. Test conditions

The test point selected for this study was an engine speed of Fig. 2. The selected test point, and the WHSC and ESC13 test cycle points over an
1200 rpm and a load of 25%, equivalent to 0.615 MPa IMEP. This estimated HD diesel engine speed-load map.
170 V.B. Pedrozo et al. / Applied Energy 165 (2016) 166–182

28/72, in average, was used to decrease the pressure rise rates. The 4.2. Determination of the optimum fuelling and diesel injection
split ratio calculation was based on the ratio of the energising time strategy in DF combustion mode
(ET) of each injection to the total energising time. The pilot injec-
tion resulted in shorter ignition delay (ID) periods (i.e. SOI_2 to In this section, the optimum ethanol substitution ratio and die-
SOC), reducing the rate of premixed combustion represented by sel injection strategy to ignite and efficiently burn the well-mixed
the first peak heat release in Fig. 3. The reduced combustion noise charge is identified. No external EGR was used in this initial phase
(PRR) was achieved at the expense of higher soot emissions, linked to reduce the complexity of the test. Intake air and exhaust back
to the formation of high equivalence ratio zones at intermediate pressures were held constant at 103 ± 1 kPa and 113 ± 1 kPa,
temperatures [1]. respectively. Intake manifold air temperature was maintained at
The sensitivity of the PRR, ISFC, ISNOx and ISSoot to different 295 ± 3 K throughout this set of experiments. Upon using a single
intake air pressures, injection strategies, EGR rates, and combus- diesel injection close to TDC, dual-fuel combustion had limited
tion phasing (CA50) are shown in Fig. 4. The legend makes refer- operating range due to low fuel conversion efficiency and high
ence to the intake pressure, EGR rate, RP, and split ratio. A higher PRR [40,41], caused by fast heat release of the diesel fuel. Split die-
intake air pressure of 125 kPa improved fuel efficiency as leaner sel injections were then adopted to improve the mixture reactivity
mixtures reduced in-cylinder temperatures and subsequent heat stratification and enhance the ignition process.
transfer losses [49]. Higher oxygen availability and higher injection In this test, the start of the second diesel injection (SOI_2) was a
pressures reduced soot emissions by minimising fuel-rich combus- result of the stipulated start of the first injection (SOI_1) and the
tion and enhancing diesel atomisation and mixing. NOx production dwell timing between injections. The resulting SOI_2 remained
decreased as more EGR was added due to lower combustion tem- within 12 to 2 CAD ATDC. This differs from RCCI operation, where
peratures, a result of the higher heat capacity of the charge and the second injection is generally delivered near 35 CAD ATDC
lower oxygen concentration [50]. These are typical trade-offs of a [27]. The required energising time for the first injection (ET_1)
conventional diesel combustion system and directly related to was set using the ECU’s application program. The energising time
combustion temperature and local equivalence ratio. of the second injection (ET_2) was automatically adjusted by the
Higher diesel fuel injection pressure and the ‘correct’ EGR level engine speed governor. The estimate of the quantity injected dur-
resulted in an optimum NOx/soot trade-off without penalising fuel ing ET_1 and ET_2 was based on the ratio of the energising time of
consumption. This operation point will be used for future compar- each injection to the total injection time (i.e. split ratio), as the die-
isons to the best DF combustion results. The selected diesel-only sel fuel mass injected at each energising time cannot be easily
calibration is circled on the curve denoted with ‘‘x” markers. It determined.
was achieved at an intake air pressure of 125 kPa, a rail pressure The ethanol mass flow rates tested in this experimental analysis
of 125 MPa, and 25% EGR, resulting in ISSoot and ISNOx emissions were 34.5, 53.6, and 72.0 mg/inj. The lowest injection quantity was
of 0.018 and 2.01 g/kW h, respectively. A detailed summary of this limited by the minimum pulse width provided by the injector
operating condition is given in Table 6 and will be discussed later drive. Considering this initial value represented approximately
in the paper. 40% on a mass input basis, the next two pulse widths were set at
around 60% and 80% of the total fuel mass injected. These fractions
changed as the indicated efficiency increased during the injection
strategy sweeps. A more detailed summary of the resultant ethanol
fractions for each fuelling and diesel injection strategy is depicted
in Table 3. The data were placed beside a similar diesel-only oper-
ating condition (i.e. not the optimum), used as a baseline in this
first section. The diesel injection pressure was held constant at
70 MPa.
The longest ET_1 of 0.90 ms could not be tested at the lowest
ethanol mass flow rate of 34.5 mg/inj because of rapid heat release.
At the highest ethanol substitution ratio (i.e. 72.0 mg/inj), an ET_1
of 0.43 ms was not applied as a short pre-injection did not allow
enough time for mixture preparation prior to the SOC, causing
knock. A longer dwell timing of 2 ms (⁄) was required with an
ET_1 of 0.59 ms to avoid elevated PRR. This particular injection
strategy was plotted with a label of 1 ms with the aim of simplify-
ing the analysis of the results. An ET_1 of 0.90 ms was removed by
the ECU at 72.0 mg/inj because ET_2 was too short to be
maintained.
Fig. 5 depicts the ignition delay from SOI_1 and SOI_2 to SOC for
those fuelling and diesel injection strategies showed in Table 3. The
legend makes reference to the ET_1 and dwell timing, while the
colours differentiate the fuelling strategy. It is observed that the
ignition delay from SOI_1 to SOC steadily rises for longer and more
advanced pre-injections. The increased amount of diesel injected
earlier in the cycle resulted in better mixture preparation, reducing
high reactivity zones within the piston bowl. The same behaviour
is seen when employing elevated ethanol fractions, due to lower
charge reactivity (i.e. higher RON) and lower bulk gas temperatures
introduced by the higher heat of vaporisation of the ethanol
[37,41]. While maintaining the same diesel injection strategy of
Fig. 3. In-cylinder pressure, diesel injection, and HRR curves of conventional diesel the diesel-only mode (ET_1 of 0.43 ms, dwell timing of 1 ms), the
combustion running without a pilot injection and with a split ratio of 26/74. utilisation of 34.5 and 53.6 mg of ethanol resulted in shorter
V.B. Pedrozo et al. / Applied Energy 165 (2016) 166–182 171

3.0 202
Higher intake air
No pilot injection pressure,
EGR rate, and RP
198
PRR [MPa/CAD]

ISFC [g/kWh]
2.0

194

1.0
190

Higher RP

0.0 186
0 3 6 9 12 15 0 3 6 9 12
CA50 [CAD ADTC] ISNOx [g/kWh]

0.075

103 kPa, 0% EGR, RP 70 MPa, No Pilot


Higher intake air 103 kPa, 0% EGR, RP 70 MPa, 25/75
ISSoot [g/kWh]

0.050
pressure,
EGR rate, and RP 125 kPa, 0% EGR, RP 70 MPa, No Pilot

125 kPa, 0% EGR, RP 70 MPa, 26/74

0.025 125 kPa, 10% EGR, RP 90 MPa, 28/72

125 kPa, 21% EGR, RP 110 MPa, 28/72

125 kPa, 25% EGR, RP 125 MPa, 30/70


0.000
0 3 6 9 12
ISNOx [g/kWh]

Fig. 4. Conventional diesel combustion sensitivities of PRR, ISFC, ISNOx and ISSoot to different intake air pressures, injection strategies, EGR rates, and CA50.

Table 3
Fuelling and diesel injection strategies.

Parameter Unit Value


Ethanol flow rate mg/inj 0.0 34.5 53.6 72.0
Ethanol mass fraction – 0.00 0.41 0.43 0.44 0.58 0.61 0.63 0.65 0.77 0.78
Ethanol energy fraction – 0.00 0.31 0.32 0.33 0.47 0.50 0.52 0.54 0.67 0.69
LHVDF MJ/kg 42.9 36.3 36.1 35.9 33.6 33.1 32.8 32.5 30.6 30.4
ET_1 ms 0.43 0.43 0.59 0.76 0.43 0.59 0.76 0.90 0.59 0.76
Diesel inj. split ratio – 25/75 28/72 37/63 45/55 31/69 40/60 49/51 60/40 42/58 52/48
Dwell timing ms 1 1 1 3 1 1 3 5 2(⁄) 3

ignition delays than diesel-only operation. The reduced mixing the diesel injections were advanced as they allowed more mixing
times might be a result of high reactivity caused by the greater time prior to combustion. For the same diesel injection strategy,
amount of diesel injected to compensate the combustion ineffi- the combustion duration was shortened as more ethanol was
ciencies and keep the load constant. When the ignition delay injected. This is a result of the faster combustion promoted by its
between the SOI_2 and the SOC was plotted, negative values were homogeneous distribution and flame propagation, generally lead-
observed for the more advanced and longer pre-injections (i.e. tri- ing to higher PRR. Combustion remained stable with COV_IMEP
angles and squares). It depicts the fact that the SOI_2 took place in the range of 1.1–2.3% throughout the tests.
after combustion had already started. This was attributed to the The results in Fig. 7 show that a lower NOx/soot emissions
progressive auto-ignition of diesel fuel injected at SOI_1 and the trade-off can be achieved using DF combustion when compared
entrained air and ethanol in the piston bowl [26]. to equivalent conventional diesel combustion. Elevated levels of
Fig. 6 shows that the combustion phasing was shifted linearly soot with low ethanol fractions were associated with shorter igni-
towards TDC for a pre-injection taking place prior to 20 CAD tion delays and rapid rate of premixed combustion of the diesel
ATDC. However, no clear correlation between CA50 and SOI_1 fuel. Higher ethanol fractions and earlier pre-injections helped
could be observed as the pre-injections were advanced and the lowering soot emissions due to lower overall carbon to hydrogen
energising times were lengthened. In some cases, a higher degree mass ratio and reduced diesel diffusion combustion [51]. Addition-
of stratification in the piston bowl promoted by a retarded SOI_1 ally, the use of advanced and longer SOI_1 improved the mixture
advanced the CA50 position. The auto-ignition of the premixed die- preparation by creating multiple ignition sites, reducing charge
sel was hindered by the lower reactivity of the ethanol-air charge stratification and NOx emissions [52].
in the cylinder during the first injection and was more prone to Fig. 8 depicts the combustion and net indicated efficiencies.
cyclic variations of flow and mixture motion. Despite the different Indicated efficiency decreased for shorter and retarded pre-
trend in CA50, the combustion duration (CA10–CA90) decreased as injections due to reduced in-cylinder reactivity gradient and
172 V.B. Pedrozo et al. / Applied Energy 165 (2016) 166–182

Fig. 5. Ignition delay from SOI_1 and SOI_2 to SOC.

delayed combustion process towards the expansion stroke. Com-


bustion efficiency was associated with the amount of ethanol
injected. The reduced amount of diesel fuel (i.e. reduced in-
cylinder charge reactivity) inhibits the combustion progression,
resulting in increased unburnt HC and CO emissions levels [53].
Combustion modelling also showed that the majority of late cycle
unburnt HC is located in the piston-to-liner crevices, and that the
bulk of the CO resides near the liner and the crevice regions [54].
This phenomenon is exacerbated at elevated ethanol fractions
due to higher amount of fuel trapped in the crevices and squish
volumes of the conventional diesel combustion system, not Fig. 6. CA50, CA10–CA90 and PRR.
designed to operate on a premixed fuel [28].
The optimum DF operating mode was achieved using an etha-
nol mass flow rate of 53.6 mg/inj, with a SOI_1 between 40 and 0.24
0.43 ms, Dwell 1 ms Diesel
35 CAD ATDC and a split ratio of approximately 60/40. This fuel-
0.59 ms, Dwell 1 ms* 34.5 mg/inj
ling strategy is equivalent to an ethanol energy fraction of 54%. The 0.76 ms, Dwell 3 ms 53.6 mg/inj
diesel injection strategy selected allowed for appropriate reactivity 0.18
ISSoot [g/kWh]

0.90 ms, Dwell 5 ms 72.0 mg/inj


gradient. Combustion efficiency reached levels of a lower ethanol
substitution ratio. It is believed that higher intake air temperatures
0.12
could improve the ethanol evaporation process and thus combus-
tion efficiency at higher ethanol fractions [55]. A lower substitution
ratio (34.5 mg/inj) did not show advantages, worsening exhaust
0.06
emissions and net indicated efficiency.
Fig. 9 shows a comparison between the optimum DF
combustion strategy and an equivalent diesel-only case. Lower
0.00
in-cylinder pressure and bulk gas temperature were observed dur- 0 3 6 9 12
ing the compression stroke, prior to the SOC. This is mainly attrib- ISNOx [g/kWh]
uted to the evaporation of the ethanol, generally called cooling
effect [32,56]. The fuelling and injection strategy also resulted in Fig. 7. NOx/soot emissions trade-off.
a single stage heat release, differing from the typical double-
hump of conventional diesel combustion. Finally, the faster and
more advanced combustion process of the dual-fuel mode led to of the burnt gas temperature would require appropriate modelling
higher peak pressure (Pmax) and higher average combustion of the fuel distribution and heat release of the premixed and
temperatures than the diesel-only operation. The determination non-premixed regions which is out of the scope of this work.
V.B. Pedrozo et al. / Applied Energy 165 (2016) 166–182 173

Fig. 8. Combustion and net indicated efficiencies.

Combustion phasing could still be controlled by the injection


timing without the need for large amounts of EGR.
Table 4 compares the resulting performance and emissions of
the two combustion modes. In dual-fuel mode, the longer ignition
delay and the reduced amount of high reactivity fuel minimised
fuel-rich and high temperatures zones [26,28]. As a result, NOx
and soot emissions were reduced by more than 50% when com-
pared to the conventional diesel combustion case. The faster and
more advanced combustion process also lowered heat transfer
losses [28] and allowed for higher indicated efficiency. Combustion
efficiency dropped as a result of reduced fuel reactivity and more Fig. 9. In-cylinder pressure, diesel injection, HRR, and average in-cylinder gas
premixed fuel trapped in the crevice and squish volumes [53]. As temperature curves. Optimum DF combustion strategy vs. equivalent diesel-only
the Euro VI emissions standards for NOx have not been fulfilled, case.

investigations with EGR were performed in the next section.

Table 4
4.3. The effect of EGR in DF combustion Comparison between the optimum DF combustion strategy and an equivalent diesel-
only case running with an intake air pressure of 103 kPa, an RP of 70 MPa, and no EGR.
This section presents the results of the experiments conducted Parameter Unit Diesel DF
to evaluate the effect of EGR in the optimum DF combustion
EF – 0.00 0.54
strategy. The boundary conditions were kept constant with the SOI_1 CAD ATDC 14.5 36.1
exception of intake manifold air temperature, which was elevated ET_1 ms 0.45 0.90
to 306 ± 1 K due to the addition of 25% EGR at 343 ± 5 K. The etha- SOI_2 CAD ATDC 7.3 0.2
nol mass flow rate was kept at 53.6 mg/inj. As the control of com- Split ratio % 25/75 60/40
Ignition delay – SOI_1 to SOC ms 1.67 3.20
bustion phasing in the DF mode relies on the diesel injection COV_IMEP % 1.2 1.4
strategy, SOI_1 was fixed at around 36.5 CAD ATDC with a con- Pmax MPa 7.12 8.67
stant energising timing ET_1 of 0.90 ms, providing a split ratio of PRR MPa/CAD 1.07 0.94
approximately 60/40. Experiments were carried out first with con- CA50 CAD ATDC 7.0 2.1
CA10–CA90 CAD 28.5 13.5
stant SOI_2 and then at similar CA50, by advancing SOI_2. Fig. 10
Uglobal – 0.48 0.44
shows the in-cylinder pressure, diesel injection, and HRR curves ISSoot g/kW h 0.031 0.011
of the optimum DF strategy at 0% and 25% EGR. Table 5 ISNOx g/kW h 7.85 3.56
summarises the resulting performance and emissions of the two ISCO g/kW h 0.80 4.56
experiments with and without EGR. ISHC g/kW h 0.42 4.90
gc % 99.7 97.4
In the case of constant SOI_2, adding EGR delayed combustion g_ind % 43.6 45.5
into the expansion stroke and increased the combustion duration.
174 V.B. Pedrozo et al. / Applied Energy 165 (2016) 166–182

peak heat release. Net indicated efficiency with EGR was slightly
lower than that without EGR, possibly as a consequence of non-
optimised combustion phasing and higher global equivalence ratio.
As the Euro VI NOx and soot emissions targets have not been
reached, further experiments were carried out by varying the first
diesel injection timing and split ratio at a constant SOI_2, as
described in the following section.

4.4. Optimisation of the SOI_1 and split ratio at a constant SOI_2

Sensitivity studies of the first diesel injection timing and split


ratio on combustion, emissions, and efficiency of the optimum
DF strategy were carried out in this section. This is prior to testing
a higher rail pressure and a lower equivalence ratio, which are
addressed in the subsequent section. The operating conditions for
this series of experiments were the same as Section 4.3 at an
EGR rate of 25%. Unlike the constant dwell timing experiments per-
formed in Section 4.2, SOI_1 and ET_1 were varied while the SOI_2
was held at approximately 7.5 CAD ATDC. This provided a safety
margin to PRR and CA50. The injection pressure was set at 70 MPa.
The ethanol mass flow rate was kept at 53.6 mg/inj, resulting in an
energy fraction varying from 50% to 53%.
Fig. 11 depicts the SOI_1 sweeps at a split ratio of approxi-
mately 57/43. A maximum PRR of 1.71 MPa/CAD was found with
an SOI_1 occurring at 40 CAD ATDC. It was also the calibration
with the shortest combustion duration of 7.6 CAD. Earlier pre-
injections reduced fuel-rich zones, which was supported by a drop
in soot emissions and increase in COV_IMEP. The result was a
Fig. 10. In-cylinder pressure, diesel injection, and HRR curves of the optimum DF slower and retarded combustion process, increasing combustion
strategy at 0% and 25% EGR.
losses and hindering the indicated efficiency. Pre-injections later
than 40 CAD ATDC reduced the time available for mixing and
CA50 was retarded from 2.1 to 10.6 CAD ATDC and CA10–CA90 was regions of higher local equivalence ratio prevailed. This stratifica-
extended from 13.5 to 17 CAD, drastically reducing NOx tion advanced the CA50 and increased the peak in-cylinder pres-
emissions from 3.56 to 0.69 g/kW h. Soot, CO, and unburnt HC sure and temperatures, resulting in higher NOx formation.
emissions increased to a certain extent due to late and poor Another consequence of over advanced combustion phasing was
combustion development. Diesel fuel consumption also increased the increase in the compression work followed by a reduction in
with a more diluted charge in order to keep the engine speed net indicated efficiency. Accordingly, the optimum emissions/effi-
and IMEP constants, reducing the ethanol energy fraction from ciency trade-off was obtained by an SOI_1 taking place around
54% to 51%. 36.5 CAD ATDC.
In the second case, the SOI_2 was advanced to phase the MFB Fig. 12 shows the results of the split ratio sweep performed
profile closer to the CA50 of the 0% EGR case. The introduction of with a constant SOI_1 at 36.5 CAD ATDC. The ET_1 was varied
EGR allowed for a longer ignition delay. Once the combustion from 0.80 to 1.01 ms, representing 49–64% of the total energising
started, a more readily ignitable charge burnt in almost half the time. Increasing the fuel quantity injected in the first injection
time. NOx emissions were curbed by EGR. CO and unburnt HC yielded similar effects to retarding SOI_1. At the largest first injec-
emissions decreased as a result of longer mixing period and higher tion amount (i.e. split ratio of 64/36), the more reactive mixture led

Table 5
The effect of EGR on combustion, emissions, and efficiency of the optimum DF strategy running at an intake air pressure of 103 kPa and an RP of 70 MPa.

Parameter Unit 0% EGR 25% EGR, similar SOI_1 and SOI_2 25% EGR, similar SOI_1 and CA50
EF – 0.54 0.51 0.54
SOI_1 CAD ATDC 36.1 36.6 36.6
ET_1 ms 0.90 0.90 0.90
SOI_2 CAD ATDC 0.2 0.2 9.8
Split ratio % 60/40 57/43 57/43
Ignition delay – SOI_2 to SOC ms 1.83 0.75 0.29
COV_IMEP % 1.4 1.5 1.1
Pmax MPa 8.67 6.45 9.13
PRR MPa/CAD 0.94 0.56 1.82
CA50 CAD ATDC 2.1 10.6 1.9
CA10–CA90 CAD 13.5 17.0 6.9
Uglobal – 0.44 0.63 0.60
ISSoot g/kW h 0.011 0.018 0.013
ISNOx g/kW h 3.56 0.69 2.73
ISCO g/kW h 4.56 5.86 3.26
ISHC g/kW h 4.90 5.41 3.62
gc % 97.4 97.1 98.1
g_ind % 45.5 44.2 45.1
V.B. Pedrozo et al. / Applied Energy 165 (2016) 166–182 175

0.46 8

ISNOx, ISCO, ISHC [g/kWh]


6
0.44
_ind [-]

0.42
Net indicated eff.
2
CA10-CA90
CA50 ISNOx
PRR ISCO
0.40 18 0 ISHC 0.04
ISSoot
COV

CA50 [CAD ATDC]

COV_IMEP [%/100]
CA10-CA90 [CAD]
0.03

PRR [MPa/CAD]

ISSoot [g/kWh]
10

0.02

2
0.01

-6 0.00
-50 -45 -40 -35 -30 -25 -50 -45 -40 -35 -30 -25
SOI_1 [CAD ATDC] SOI_1 [CAD ATDC]

Fig. 11. Effect of the SOI_1 on combustion, emissions, and efficiency.

0.46 7.5
ISNOx, ISCO, ISHC [g/kWh]

0.44 5.0
_ind [-]

Net indicated eff.


CA10-CA90
0.42 2.5
CA50
PRR
ISNOx
ISCO
0.40 12 0.0 0.03
ISHC
ISSoot
COV
COV_IMEP [%/100]
CA50 [CAD ATDC]
CA10-CA90 [CAD]

PRR [MPa/CAD]

ISSoot [g/kWh]

6 0.02

0 0.01

-6 0.00
45% 50% 55% 60% 65% 70% 45% 50% 55% 60% 65% 70%
ET_1/total energising time [%] ET_1/total energising time [%]

Fig. 12. Effect of split ratio on combustion, emissions and efficiency.

to a rapid and early combustion, reducing unburnt HC and CO pre-injection (i.e. split ratio of 49/51) reduced the mixture flamma-
emissions at the expense of higher NOx and PRR. Soot also bility and delayed the combustion process, degrading fuel conver-
increased as the mixing time available to the main injection was sion efficiency. The optimum split ratio represents an ET_1
reduced. The opposite was true for decreasing the ET_1, however equivalent to 53–55% of the total energising time. The COV_IMEP
a trade-off in indicated efficiency was observed. A short remained below 1.4%.
176 V.B. Pedrozo et al. / Applied Energy 165 (2016) 166–182

4.5. The effect of higher intake air pressure and rail pressure in DF
combustion

In this section, dual-fuel experiments were carried with higher


intake air pressure and diesel injection pressure. The objective was
to improve combustion and indicated efficiencies while reducing
soot and NOx emissions. Plots of two different rail pressures (70
and 90 MPa) and intake manifold air pressures (103 and 125 kPa)
were compared on a CA50 basis. The ethanol energy fraction varied
from 51% to 54%, with an average of 53%, as the amounts of diesel
fuel were automatically adjusted by the ECU to maintain a con-
stant engine speed and load. The SOI_1 was set at around 36.5
CAD ATDC with an ET_1 equivalent to a split ratio of approximately
54/46. The SOI_2 was altered within the range 9.5 to 1.0 CAD
ATDC to delay or advance the CA50. The EGR was introduced into
the system at 353 ± 10 K and maintained at 25.0 ± 0.5%. The result-
ing intake air charge temperature was 308 ± 2 K. The pumping
losses were the same, as the delta between the intake air pressure
and the exhaust back pressure was held at 10 kPa. Combustion sta-
bility was considered acceptable, with COV_IMEP between 0.9%
and 2.2%.
Fig. 13 depicts the in-cylinder pressure, diesel injection, and
HRR curves for a sweep of SOI_2 at 125 kPa intake air pressure
and 90 MPa injection pressure. Although the SOC’s took place
almost at the same crank angle, a retarded SOI_2 shifted the com-
bustion process towards the expansion stroke, decreasing the peak
in-cylinder pressure and peak HRR.
To expand on this trend, SOI_2, combustion duration, ignition
delay, and PRR with respect to CA50 are shown in Fig. 14. The dif-
fusion combustion of the second diesel injection allowed control
over the combustion phasing and burn rate. A higher intake air
pressure required a retarded SOI_2 due earlier auto-ignition of
the diesel fuel injected at the 36.5 CAD ATDC, suggested by the
shorter ignition delays. Earlier main injections generally allowed
for more time available between the SOI_2 and the SOC. However,

Fig. 14. SOI_2, ignition delay, CA10–CA90, and PRR vs. CA50.

the most advanced cases at 125 kPa intake air pressure presented
short ignition delays (i.e. highlighted by the circled region). The
higher in-cylinder pressures and temperatures advanced the onset
of auto-ignition of the premixed diesel (i.e. ET_1). Despite the
earlier SOC, an intake air pressure of 125 kPa possesses the longest
CA10–CA90 and the lowest PRR. This is a result of the dilution with
excess air, which diminishes the reaction rates [57]. An injection
Fig. 13. In-cylinder pressure, diesel injection, and HRR curves for a sweep of SOI_2. pressure of 90 MPa resulted in better atomisation and more
V.B. Pedrozo et al. / Applied Energy 165 (2016) 166–182 177

Fig. 15. Effect of intake air pressure and diesel injection pressure in DF combustion.

homogeneous charge prior to the SOI_2, increasing the PRR and


shortening the combustion process. The PRR limit of 2.0 MPa/
CAD was exceeded during the most advanced cases at 103 kPa
intake air pressure.
Fig. 15 shows the effect of higher boost and diesel injection
pressures on in-cylinder pressure and the resulting heat release
at a CA50 of approximately 5.8 CAD ATDC. A lower rail pressure
required a retarded SOI_2 to obtain the same combustion phasing
because of the formation of high equivalence ratio regions within
the piston bowl. Another important observation concerns the
HRR profile of the DF combustion at 103 kPa intake air pressure
and 90 MPa injection pressure. The combination of lower in-
cylinder pressure and improved atomisation resulted in longer
ignition delay and elevated PRR. It is believed that the two distinct
high temperature heat release events are a result of the fast burn-
ing of the second diesel injection, followed by the ignition of the
entrained premixed charge. Higher intake air pressure and lower
RP allowed for a more progressive HRR.
Fig. 16 depicts the effect of intake air pressure and injection
pressure on emissions and efficiencies of DF combustion. Despite
the higher in-cylinder pressures prior to the SOC at 125 kPa intake
air pressure, the dilution with excess air constrained the charge
reactivity and heat release peaks, mitigating NOx formation. The
opposite occurs if rail pressure is increased as a result of better
atomisation (i.e. presence of close-to-stoichiometric regions) and
Fig. 16. The effect of intake air pressure and injection pressure on emissions and
faster combustion process. The improved mixture preparation also
efficiencies of DF combustion.
helps with fuel efficiency gains and smoke reduction. When com-
paring the emissions on a CA50 basis, an advanced SOI_2 resulted
in higher NOx emissions due to shorter CA10–CA90 and elevated Combustion efficiency in DF mode remains lower than conven-
in-cylinder peak pressures and temperatures. In addition, an ear- tional diesel combustion (i.e. generally above 99%). It is attributed
lier SOI_2 at an intake air pressure of 103 kPa led to higher soot to the reduced reactivity of the ethanol and the fundamental differ-
emissions after poor mixing and rapid combustion. The operation ences between a turbulent diffusion flame in diesel-only operation
at 125 kPa exhibited the contrary behaviour, with soot emissions and a progressive and sequential combustion from the high to low
increasing as the SOI_2 was retarded. This is a consequence of a reactivity regions in DF mode [1]. The diesel consumption does not
poor combustion process and the formation of low temperature progress to the crevice regions, not releasing enough energy for the
fuel-rich regions. combustion of all the premixed fuel, which increases combustion
178 V.B. Pedrozo et al. / Applied Energy 165 (2016) 166–182

an injection pressure of 125 kPa and 90 MPa, respectively. Fig. 17


compares the in-cylinder pressure, injector signal, and MFB curves
of the optimum emissions trade-off in diesel-only (see Section 4.1)
and DF combustion against the most fuel efficient case attained
during this study. As showed in Section 4.2, dual-fuel mode is char-
acterised by advanced combustion phasing, shorter combustion
durations, and higher peak in-cylinder pressures. This is due to
the early first diesel injection which increases the flammability of
the in-cylinder charge and promotes a more reactive mixture prior
to the second injection. The main diesel injection close to TDC
determines the combustion phasing. It is believed that advanced
SOI_2 will be required at higher ethanol substitution ratios to
increase the degree of stratification and avoid misfire.
Table 6 compares the resulting performance and emissions of
the optimum conventional diesel combustion baseline (deter-
mined in Section 4.1) against the two selected dual-fuel operating
points. The first DF strategy represents the operating point with
optimum g_ind/ISNOx/ISSoot trade-off, while the second is the
most fuel efficient calibration obtained at this particular load. It
can be observed that a premixed charge of ethanol resulted in sev-
eral benefits. These included higher indicated efficiency, and lower
soot and NOx emissions than diesel-only operation. The use of
optimised split diesel injections, intake air pressure, and rail pres-
sure also kept ISCO and ISHC below 10 g/kW h, demonstrating a
considerable improvement in comparison to previous studies
[37–39]. The higher indicated efficiency associated with dual-fuel
combustion led to lower exhaust gas temperatures (EGT) than con-
ventional diesel combustion [58]. Elevated injection pressure of
Fig. 17. In-cylinder pressure, diesel injection, and MFB curves of the optimum 125 MPa was required to curb soot emissions in diesel-only mode.
emissions trade-off in diesel-only and DF combustion modes compared against the The radar chart showed in Fig. 18 summarises the main trends
most fuel efficient case.
observed. Net indicated efficiency, peak in-cylinder pressure, soot
and NOx emissions, and combustion efficiency results were
losses [53]. The use of an intake air pressure of 103 kPa and earlier normalised to obtain a clear distinction of the benefits. Euro VI
SOI_2’s improved the flammability of the in-cylinder charge and legislation emissions limits were not fully met under conventional
resulted in combustion efficiencies up to 98%, at the expense of diesel combustion or dual-fuel mode. However, the best results
higher NOx emissions. A higher intake air pressure and a retarded were obtained with utilisation of ethanol. The combustion of a
SOI_2 reduced NOx emissions and combustion efficiencies due to well-mixed charge ignited by split diesel injections considerably
overly lean regions and lower local in-cylinder temperatures. reduced NOx and soot emissions, which are the main limiting
factors of HD diesel engines. Net indicated efficiency increased
by 3.2% in the most efficient case as a result of lower local
4.6. Comparison between DF and conventional diesel combustion in-cylinder temperatures and faster combustion process. The
majority of the unburnt HC and CO emissions produced by DF
The optimum dual-fuel operating points were determined by operation can be removed by an oxidation catalyst, assuming an
the best indicated efficiency/emissions trade-off at an intake and EGT above 570 K [58].

Table 6
Comparison between the optimum diesel-only calibration and the selected dual-fuel operating points, running with 25% EGR and 125 kPa intake air pressure.

Parameter Unit Diesel, trade-off DF, trade-off DF, most efficient


EF – 0.00 0.53 0.54
SOI_1 CAD ATDC 9.3 36.7 36.7
ET_1 ms 0.38 0.76 0.76
SOI_2 CAD ATDC 2.0 0.8 2.7
Split ratio % 30/70 53/47 55/45
Rail pressure MPa 125 90 90
Ignition delay – SOI_2 to SOC ms 0.49 1.12 0.77
COV_IMEP % 1.8 1.6 1.4
Pmax MPa 7.56 7.85 9.09
PRR MPa/CAD 1.42 0.67 1.26
CA50 CAD ATDC 9.2 7.8 4.1
CA10–CA90 CAD 22.3 18.6 14.6
EGT K 596 578 567
Uglobal – 0.53 0.51 0.50
ISSoot g/kW h 0.0175 0.0125 0.0118
ISNOx g/kW h 2.01 0.71 1.32
ISCO g/kW h 0.67 9.55 6.94
ISHC g/kW h 0.14 6.69 5.34
gc % 99.8 96.0 96.8
g_ind % 44.1 44.4 45.5
V.B. Pedrozo et al. / Applied Energy 165 (2016) 166–182 179

Pmax Diesel, trade-off


+20%
EF 0.53, trade-off
+3.8%
EF 0.54, most fuel efficient

+0.7%
_ind ISSoot
-33%
+3.2% -29%

-4% -65%
-3%
-34%
c ISNOx

Fig. 18. Normalised net indicated efficiency, peak in-cylinder pressure, ISSoot and ISNOx emissions, and combustion efficiency results of the optimum trade-offs in diesel-
only and DF combustion modes compared against the most fuel efficient case.

5. Conclusions use of a second diesel injection around TDC rather than early
injections is a differentiator from RCCI.
In this paper, ethanol dual-fuel combustion was experimentally  Reducing rail pressure in dual-fuel mode created a higher
investigated using an HD diesel engine operating at 1200 rpm and degree of stratification, advancing the SOC. The shorter mix-
0.615 MPa. The impact of three ethanol fractions and different die- ing period resulted in longer diffusion combustion and higher
sel injection strategies on combustion, emissions, and efficiency soot emissions, but lower NOx levels. A higher injection
were analysed and discussed. Two separate engine calibrations pressure improved diesel atomisation and resulted in a more
with the highest efficiency and lowest emissions were quantified, homogeneous charge, delaying the SOC. This contributed to
complete with effects from EGR, intake air pressure, and injection shorter combustion process, higher PRR, and higher NOx
pressure. The main findings can be summarised as follows: emissions.
 A leaner operation (higher intake air pressure) decreased local
 Diesel-only combustion requires a combination of very high in-cylinder temperatures and heat release peaks. This resulted
injection pressures and EGR rates to achieve low engine-out in lower NOx formation, but higher unburnt HC and CO
emissions of soot and NOx. Intake air pressure also needs to emissions.
be increased to avoid a fuel economy penalty.
 Ethanol dual-fuel combustion with a single diesel injection Ethanol dual-fuel combustion offers the potential to minimise
close to TDC has a limited operating range due to high PRR exhaust aftertreatment requirements and improve the efficiency
and low indicated efficiency. A split injection strategy adjusted of heavy-duty engines. Furthermore, the use of ethanol allows for
the mixture flammability and promoted in-cylinder reactivity less dependence on petroleum-based fuels and a reduction in
gradients, increasing the fuel conversion efficiency and reduc- greenhouse gas emissions [5]. While Euro VI emissions standards
ing CO and unburnt HC emissions. still remain a challenge, the calibration for the best dual-fuel emis-
 In the majority of the cases tested, the homogenous charge of sions case reduced NOx emissions by 65% and soot emissions by
ethanol reduced local in-cylinder temperatures and fuel-rich 29% when compared against the optimum conventional diesel
zones, resulting in lower NOx and soot emissions compared to combustion mode. The most efficient dual-fuel calibration allowed
conventional diesel combustion. for a 3.2% gain in indicated efficiency while retaining at least 33%
 High ethanol substitution ratios resulted in poor combustion reduction in NOx and soot emissions. Combustion losses were mit-
efficiency at this particular load due low in-cylinder charge igated by maintaining CO and unburnt HC under 10 g/kW h, but a
temperature and high amount of fuel trapped in the crevice low temperature catalyst may be required [25]. Dual-fuel opera-
and squish volumes. Low substitution ratios did not demon- tion at higher loads might be limited by the maximum cylinder
strate benefits in terms of net indicated efficiency and emis- pressure of the engine. Future work will be carried out at different
sions reduction. engine speeds and loads to maximise the substitution of ethanol
 The optimum DF strategy was an ethanol substitution ratio of while minimising emissions.
54% on an energy basis, with a diesel pre-injection timing
around 36.5 CAD ATDC and a split ratio of corresponding to
53–55% of the total energising time. Acknowledgments
 The addition of EGR reduced NOx emissions and promoted
longer ignition delays, consequently allowing more time for The authors would like to acknowledge the Brazilian Federal
the charge to mix prior to the SOC. Agency for Support and Evaluation of Postgraduate Education
 The split diesel injection strategy used in this work allowed for (CAPES) and the National Council for Scientific and Technological
control over combustion phasing without the need for varying Development (CNPq) for supporting the PhD studies of Mr. Pedrozo
the overall fuel reactivity (i.e. ethanol substitution ratio). The and Mr. Dalla Nora at Brunel University London.
180 V.B. Pedrozo et al. / Applied Energy 165 (2016) 166–182

Appendix A. Measurement device specifications

Measured variable Device Manufacturer Dynamic Linearity/ Repeatability


range accuracy
CO (low content) AIA-721A Horiba (MEXA 7170 0–2.5k ppm 6 ±1.0% FS or Within ±0.5% of
CO (mid-high content) AIA-722 DEGR) 0–12 vol% ±2.0% of FS
CO2 AIA-722 0–20 vol% readings
NOx CLA-720MA 0–500 ppm or
0–10k ppm
O2 MPA-720 0–25 vol%
Unburnt HC FIA-725A 0–500 ppm or
0–50k ppm
Diesel injector current signal Current Probe PR30 LEM 0–20 A ±1% of reading
±2 mA
Diesel flow rate (return) PROline promass 83A Endress + Hauser 0–100 kg/h ±0.10% of ±0.05% of
DN01 reading reading
Diesel flow rate (supply) PROline promass 83A 0–20 kg/h ±0.10% of ±0.05% of
DN02 reading reading
Intake and exhaust pressures Piezoresistive Kistler 0–1 MPa 6 ±0.50% of FS
pressure sensor Type within 0–353 K
4049A
Amplifier Type
4622A
In-cylinder pressure Piezoelectric Kistler 0–30 MPa 6 ±0.40% of FS
pressure sensor Type
6125C
Amplifier FI Piezo AVL 6 ±0.01% of FS
Intake valve lift S-DVRT-24 LORD MicroStrain 0–24 mm ±1% of reading ±1.0 lm
displacement sensor using straight
DEMOD-DVRT-TC line
conditioner
Intake air mass flow rate Proline t-mass 65F Endress + Hauser 0–910 kg/h ±1.5% of reading ±0.5% of reading
(10–100% of FS)
Oil and ethanol pressure Pressure transducer GE 0–1 MPa < ±0.20% of FS
UNIK 5000
Smoke number 415SE AVL 0–10 FSN – Within ±0.005
FSN +3% of
reading
Speed AG150 Froude Hofmann 0–8000 rpm ±1 rpm
Torque Dynamometer 0–500 Nm ±0.25% of FS
Temperature Thermocouple K RS 233–1473 K 6± 2.5 K or
Type (Class 2) ±0.75% of
readings
⁄ FS = full scale.

References
Renew Sustain Energy Rev 2014;38. http://dx.doi.org/10.1016/j.
[1] Reitz RD. Directions in internal combustion engine research. Combust Flame rser.2014.07.019.
2013;160:1–8. http://dx.doi.org/10.1016/j.combustflame.2012.11.002. [8] Yao M, Zheng Z, Liu H. Progress and recent trends in homogeneous charge
[2] Dec JE. A conceptual model of DI diesel combustion based on laser sheet compression ignition (HCCI) engines. Prog Energy Combust Sci
imaging. SAE Technical Paper 1997. http://dx.doi.org/10.4271/970873. 2009;35:398–437. http://dx.doi.org/10.1016/j.pecs.2009.05.001.
[3] Imtenan S, Varman M, Masjuki HH, Kalam Ma, Sajjad H, Arbab MI, et al. Impact [9] Kook S, Bae C. Combustion control using two-stage diesel fuel injection in a
of low temperature combustion attaining strategies on diesel engine emissions single-cylinder PCCI engine. SAE Technical Paper 2004. http://dx.doi.org/10.
for diesel and biodiesels: a review. Energy Convers Manage 2014;80. http://dx. 4271/2004-01-0938.
doi.org/10.1016/j.enconman.2014.01.020. [10] Neely GD, Sasaki S, Huang Y, Leet JA, Stewart DW. New diesel emission control
[4] U.S. Energy Information Adminstration. International Energy Outlook 2014. strategy to meet US tier 2 emissions regulations. SAE Technical Paper 2005.
Washington; 2014. http://dx.doi.org/10.4271/2005-01-1091.
[5] United Nations Environment Programme. Towards sustainable production and [11] Kook S, Bae C, Miles PC, Choi D, Pickett LM. The influence of charge dilution
use of resources: assessing biofuels; 2009. and injection timing on low-temperature diesel combustion and emissions.
[6] Asad U, Zheng M, Ting DS-K, Tjong J. Implementation challenges and solutions SAE Technical Paper 2005. http://dx.doi.org/10.4271/2005-01-3837.
for homogeneous charge compression ignition combustion in diesel engines. J [12] Opat R, Ra Y, Gonzalez D. MA, Krieger R, Reitz RD, Foster DE, et al. Investigation
Eng Gas Turbines Power 2015;137. http://dx.doi.org/10.1115/1.4030091. of mixing and temperature effects on HC/CO emissions for highly dilute low
[7] Bendu H, Murugan S. Homogeneous charge compression ignition (HCCI) temperature combustion in a light duty diesel engine. SAE Technical Paper
combustion: mixture preparation and control strategies in diesel engines. 2007. http://dx.doi.org/10.4271/2007-01-0193.
V.B. Pedrozo et al. / Applied Energy 165 (2016) 166–182 181

[13] Musculus MPB, Miles PC, Pickett LM. Conceptual models for partially premixed [40] Sarjovaara T, Alantie J, Larmi M. Ethanol dual-fuel combustion concept on
low-temperature diesel combustion. Prog Energy Combust Sci 2013;39. http:// heavy duty engine. Energy 2013;63:76–85. http://dx.doi.org/10.1016/j.
dx.doi.org/10.1016/j.pecs.2012.09.001. energy.2013.10.053.
[14] Kimura S, Aoki O, Ogawa H, Muranaka S, Enomoto Y. New combustion concept [41] Padala S, Woo C, Kook S, Hawkes ER. Ethanol utilisation in a diesel engine
for ultra-clean and high-efficiency small DI diesel engines. SAE Technical Paper using dual-fuelling technology. Fuel 2013;109:597–607. http://dx.doi.org/
1999. http://dx.doi.org/10.4271/1999-01-3681. 10.1016/j.fuel.2013.03.049.
[15] Hasegawa R, Yanagihara H. HCCI combustion in DI diesel engine. SAE Technical [42] Heywood JB. Internal combustion engine fundamentals. 1st ed. McGraw-Hill
Paper 2003. http://dx.doi.org/10.4271/2003-01-0745. Inc.; 1988.
[16] Sellnau M, Foster M, Hoyer K, Moore W, Sinnamon J, Husted H. Development of [43] Silvis WM. An algorithm for calculating the air/fuel ratio from exhaust
a gasoline direct injection compression ignition (GDCI) engine. SAE Int J Eng emissions. SAE Technical Paper 1997. http://dx.doi.org/10.4271/970514.
2014;7:835–51. http://dx.doi.org/10.4271/2014-01-1300. [44] Zhao H, Ladammatos N. Engine combustion instrumentation and diagnostics.
[17] Sellnau M, Moore W, Sinnamon J, Hoyer K, Foster M, Husted H. GDCI multi- Warrendale (PA, USA); 2001.
cylinder engine for high fuel efficiency and low emissions. SAE Int J Eng [45] Kar K, Cheng WK. Speciated engine-out organic gas emissions from a PFI-SI
2015;8. http://dx.doi.org/10.4271/2015-01-0834. engine operating on ethanol/gasoline mixtures. SAE Int J Fuels Lubr 2009;2.
[18] Kalghatgi GT, Risberg P, Ångström H-E. Partially pre-mixed auto-ignition of http://dx.doi.org/10.4271/2009-01-2673.
gasoline to attain low smoke and low NOx at high load in a compression [46] Wallner T. Correlation between speciated hydrocarbon emissions and flame
ignition engine and comparison with a diesel fuel. SAE Technical Paper 2007. ionization detector response for gasoline/alcohol blends. J Eng Gas Turbines
http://dx.doi.org/10.4271/2007-01-0006. Power 2011;133. http://dx.doi.org/10.1115/1.4002893. 082801–082801 – 8.
[19] Manente V, Zander C, Johansson B, Tunestal P, Cannella W. An advanced [47] AVL. AVL 415SE smoke meter – product guide. Graz, Austria; 2013.
internal combustion engine concept for low emissions and high efficiency [48] Economic Commission for Europe of the United Nations (UN/ECE). Regulation
from idle to max load using gasoline partially premixed combustion. SAE No 49 – uniform provisions concerning the measures to be taken against the
Technical Paper 2010. http://dx.doi.org/10.4271/2010-01-2198. emission of gaseous and particulate pollutants from compression-ignition
[20] Manente V, Johansson B, Cannella W. Gasoline partially premixed combustion, engines and positive ignition engines for use in vehicles. Off J Eur Union 2013.
the future of internal combustion engines? Int J Engine Res 2011;12:194–208. [49] Splitter DA, Reitz RD. Fuel reactivity effects on the efficiency and operational
http://dx.doi.org/10.1177/1468087411402441. window of dual-fuel compression ignition engines. Fuel 2014;118. http://dx.
[21] Manente V, Johansson B, Tunestal P. Characterization of partially premixed doi.org/10.1016/j.fuel.2013.10.045.
combustion with ethanol: EGR sweeps, low and maximum loads. J Eng Gas [50] Zhao H. HCCI and CAI engines for automotive industry. Cambridge: Woodhead
Turbines Power 2010;132. http://dx.doi.org/10.1115/1.4000291. Publishing Limited; 2007.
[22] Shen M, Tuner M, Johansson B, Cannella W. Effects of EGR and intake pressure [51] Tutak W, Lukács K, Szwaja S, Bereczky Á. Alcohol–diesel fuel combustion in the
on PPC of conventional diesel, gasoline and ethanol in a heavy duty diesel compression ignition engine. Fuel 2015;154:196–206. http://dx.doi.org/
engine. SAE Technical Paper 2013;01. http://dx.doi.org/10.4271/2013-01- 10.1016/j.fuel.2015.03.071.
2702. [52] May I, Cairns A, Zhao H, Pedrozo V, Wong HC, Whelan S, et al. Reduction of
[23] Shen M, Tuner M, Johansson B. Close to stoichiometric partially premixed methane slip using premixed micro pilot combustion in a heavy-duty natural
combustion – the benefit of ethanol in comparison to conventional fuels. SAE gas-diesel engine. SAE Technical Paper 2015. http://dx.doi.org/10.4271/2015-
Technical Paper 2013. http://dx.doi.org/10.4271/2013-01-0277. 01-1798.
[24] Kaiadi M, Johansson B, Lundgren M, Gaynor JA. Experimental investigation on [53] Desantes JM, Benajes J, García A, Monsalve-Serrano J. The role of the in-
different injection strategies for ethanol partially premixed combustion. SAE cylinder gas temperature and oxygen concentration over low load reactivity
Technical Paper 2013. http://dx.doi.org/10.4271/2013-01-0281. controlled compression ignition combustion efficiency. Energy
[25] Reitz RD, Duraisamy G. Review of high efficiency and clean reactivity 2014;78:854–68. http://dx.doi.org/10.1016/j.energy.2014.10.080.
controlled compression ignition (RCCI) combustion in internal combustion [54] Kokjohn SL, Hanson RM, Splitter DA, Reitz RD. Experiments and modeling of
engines. Prog Energy Combust Sci 2015;46:12–71. http://dx.doi.org/10.1016/j. dual-fuel HCCI and PCCI combustion using in-cylinder fuel blending. SAE Int J
pecs.2014.05.003. Engines 2009;2. http://dx.doi.org/10.4271/2009-01-2647.
[26] Benajes J, Molina S, García A, Belarte E, Vanvolsem M. An investigation on RCCI [55] Sarjovaara T, Larmi M, Vuorinen V. Effect of charge air temperature on E85
combustion in a heavy duty diesel engine using in-cylinder blending of diesel dual-fuel diesel combustion. Fuel 2015;153:6–12. http://dx.doi.org/10.1016/
and gasoline fuels. Appl Therm Eng 2014;63:66–76. http://dx.doi.org/10.1016/ j.fuel.2015.02.096.
j.applthermaleng.2013.10.052. [56] Tutak W. Bioethanol E85 as a fuel for dual fuel diesel engine. Energy Convers
[27] Kokjohn SL, Hanson R, Splitter D, Kaddatz J, Reitz RD. Fuel reactivity controlled Manage 2014;86:39–48. http://dx.doi.org/10.1016/j.enconman.2014.05.016.
compression ignition (RCCI) combustion in light- and heavy-duty engines. SAE [57] Eichmeier J, Wagner U, Spicher U. Controlling gasoline low temperature
Int J Fuels Lubr 2011;4. http://dx.doi.org/10.4271/2011-01-0357. combustion by diesel micro pilot injection. J Eng Gas Turbines Power
[28] Kokjohn SL, Hanson RM, Splitter Da, Reitz RD. Fuel reactivity controlled 2012;134. http://dx.doi.org/10.1115/1.4005997.
compression ignition (RCCI): a pathway to controlled high-efficiency clean [58] Prikhodko VY, Curran SJ, Parks JE, Wagner RM. Effectiveness of diesel oxidation
combustion. Int J Engine Res 2011;12:209–26. http://dx.doi.org/10.1177/ catalyst in reducing HC and CO emissions from reactivity controlled
1468087411401548. compression ignition. SAE Int J Fuels Lubr 2013;6. http://dx.doi.org/10.4271/
[29] Sarathy SM, Oßwald P, Hansen N, Kohse-Höinghaus K. Alcohol combustion 2013-01-0515. 2013–01 – 0515.
chemistry. Prog Energy Combust Sci 2014;44:40–102. http://dx.doi.org/
10.1016/j.pecs.2014.04.003.
[30] Martins MES, Lanzanova TDM. Full-load Miller cycle with ethanol and EGR:
Glossary
potential benefits and challenges. Appl Therm Eng 2015;90:274–85. http://dx.
doi.org/10.1016/j.applthermaleng.2015.06.086. ATDC: after firing top dead centre
[31] He B, Wang J, Shuai S, Yan X. Homogeneous charge combustion and emissions CA10–CA90: combustion duration (10–90% cumulative heat release)
of ethanol ignited by pilot diesel on diesel engines. SAE Technical Paper 2004. CA50: crank angle of 50% cumulative heat release
http://dx.doi.org/10.4271/2004-01-0094. CAD: crank angle degree
[32] Gao T, Divekar P, Asad U, Han X, Reader GT, Wang M, et al. An enabling study of CO: carbon monoxide
low temperature combustion with ethanol in a diesel engine. J Energy Res CO2: carbon dioxide
Technol 2013;135. http://dx.doi.org/10.1115/1.4024027. COV_IMEP: coefficient of variation of the IMEP
[33] Han X, Xie K, Tjong J, Zheng M. Empirical study of simultaneously low NOx and DAQ: data acquisition
soot combustion with diesel and ethanol fuels in diesel engine. J Eng Gas
DF: dual-fuel
Turbines Power 2012;134. http://dx.doi.org/10.1115/1.4007163.
ECR: effective compression ratio
[34] Divekar P, Yang Z, Ting D, Chen X, Zheng M, Tjong J. Efficiency and emission
trade-off in diesel-ethanol low temperature combustion cycles. SAE Technical ECU: engine control unit
Paper 2015. http://dx.doi.org/10.4271/2015-01-0845. EGR: exhaust gas recirculation
[35] Han X, Zheng M, Tjong J. Clean combustion enabling with ethanol on a dual- EGT: exhaust gas temperature
fuel compression ignition engine. Int J Engine Res 2015:1–13. http://dx.doi. ESC13: European Stationary Cycle
org/10.1177/1468087415575646. ET: energising time
[36] Han X, Divekar P, Reader G, Zheng M, Tjong J. Active injection control for ET_1: energising time of the first diesel injection
enabling clean combustion in ethanol-diesel dual-fuel mode. SAE Int J Engines ET_2: energising time of the second diesel injection
2015;8:890–902. http://dx.doi.org/10.4271/2015-01-0858. FID: flame ionisation detector
[37] Asad U, Kumar R, Zheng M, Tjong J. Ethanol-fueled low temperature FSN: filter smoke number
combustion: a pathway to clean and efficient diesel engine cycles. Appl GDCI: Gasoline Direct Injection Compression Ignition
Energy 2015. http://dx.doi.org/10.1016/j.apenergy.2015.01.057. GHG: greenhouse gas
[38] Júnior RFB, Martins CA. Emissions analysis of a diesel engine operating in
HC: hydrocarbons
Diesel–Ethanol Dual-Fuel mode. Fuel 2014;148:191–201. http://dx.doi.org/
HCCI: Homogeneous Charge Compression Ignition
10.1016/j.fuel.2014.05.010.
HD: heavy-duty
[39] Ogawa H, Shibata G, Kato T, Zhao P. Dual fuel diesel combustion with
premixed ethanol as the main fuel. SAE Technical Paper 2014. http://dx.doi. HPP: high-pressure diesel pump
org/10.4271/2014-01-2687. HRR: apparent net heat release rate
182 V.B. Pedrozo et al. / Applied Energy 165 (2016) 166–182

ID: ignition delay PFI: port fuel injector


IMEP: net indicated mean effective pressure PPC: partially premixed combustion
ISFC: net indicated specific fuel consumption PPCI: partially premixed charge compression ignition
ISCO: net indicated specific emissions of CO RCCI: Reactivity Controlled Compression Ignition
ISHC: net indicated specific emissions of unburnt HC RP: rail pressure
ISNOx: net indicated specific emissions of NOx SOC: start of combustion
ISSoot: net indicated specific emissions of soot SOI: start of injection
IVC: intake valve closing SOI_1: start of the first diesel injection
LTC: low temperature combustion SOI_2: start of the second diesel injection
MFB: mass fraction burnt TDC: firing top dead centre
MK: Modulated Kinetics UNIBUS: Uniform Bulky Combustion System
NOx: nitrogen oxides VVA: variable valve actuation
O2: oxygen WHSC: World Harmonized Stationary Cycle
Pmax: peak in-cylinder pressure c: ratio of specific heats
PCCI: Premixed Charge Compression Ignition Uglobal: Global Equivalence Ratio

Você também pode gostar