Sunteți pe pagina 1din 36

energies

Article
Experimental Study of DI Diesel Engine Operational
and Environmental Behavior Using Blends of City
Diesel with Glycol Ethers and RME
Theodoros C. Zannis 1, * , Roussos G. Papagiannakis 2 , Efthimios G. Pariotis 1 and
Marios I. Kourampas 1
1 Naval Architecture and Marine Engineering Section, Hellenic Naval Academy, 18539 Piraeus, Greece;
pariotis@hna.gr (E.G.P.) marioskourampas@gmail.com (M.I.K.)
2 Thermodynamic & Propulsion Systems Section, Aeronautical Sciences Department,
Hellenic Air Force Academy, Dekelia Air Force Base, 1010 Dekelia, Attiki, Greece;
r.papagiannakis@gmail.com
* Correspondence: thzannis@hna.gr; Tel.: +30-210-458-1663

Received: 29 March 2019; Accepted: 22 April 2019; Published: 24 April 2019 

Abstract: An experimental investigation is performed in a single-cylinder direct-injection (DI) diesel


engine using city diesel oil called DI1 and two blends of DI1 with a mixture of glycol ethers. The
addition of glycol ethers to fuel DI1 produced oxygenated fuels GLY10 (10.2 mass-% glycol ethers)
and GLY30 (31.3 mass-% glycol ethers) with 3% and 9% oxygen content, respectively. The addition of
biofuel rapeseed methyl ester (RME) to fuel DI1 produced oxygenated blend RME30 (31.2 mass-%
RME) with 3% oxygen content. Engine tests were performed with the four fuels in the DI diesel
engine at 2500 RPM and at 20%, 40%, 60%, and 80% of full load. The experimental diesel engine
was equipped with devices for recording cylinder pressure, injection pressure, and top dead center
(TDC) position and also it was equipped with exhaust gas analyzers for measuring soot, NO, CO,
and HC emissions. A MATLAB 2014 code was developed for analyzing recorded cylinder pressure,
injection pressure, and TDC position data for all obtained engine cycles and for calculating the main
engine performance parameters. The assessment of the experimental results showed that glycol
ethers have more beneficial impact on soot and NO emissions compared to RME, whereas RME have
less detrimental impact on engine performance parameters compared to glycol ethers.

Keywords: diesel; performance; emissions; oxygenated fuels; glycol ethers; rapeseed methyl ester

1. Introduction
During recent years, internal combustion engine manufacturers in cooperation with research
institutes and universities from all over the world have managed to develop advanced technology
four-stroke and two-stroke diesel engines for various industrial sectors with considerably improved
thermal efficiency and environmental behavior compared to the past decades [1]. It is characteristic
that nowadays the brake-specific fuel consumption (BSFC) of medium-speed marine diesel engines has
reached the value of 168 g/kWh [2] and the corresponding BSFC of large two-stroke slow-speed diesel
engines is close to 156 g/kWh [3], which corresponds to the highest thermal efficiency and to the lowest
specific CO2 emissions of all fossil fuel-powered propulsion and electric power generation thermal
engines. However, diesel engines still remain a major source of gaseous and particulate emissions,
which have serious detrimental effects on human health and on the environment [1]. For this reason and
also for compliance with strict environmental regulations currently issued in many industrial sectors
such as automotive, maritime, and the electric power generation sector, engine manufacturers and
research organizations have strived to find solutions for further reducing the gaseous and particulate

Energies 2019, 12, 1547; doi:10.3390/en12081547 www.mdpi.com/journal/energies


Energies 2019, 12, 1547 2 of 36

diesel-emitted pollutants and also to further reduce the carbon footprint of diesel engines by improving
their efficiency [4].
One direct mean for improving mainly the environmental behavior not only of future but also
of existing compression ignition engines is the use of alternative gaseous fuels such as natural gas,
methane, and hydrogen. Natural gas combustion in compression ignition engines under dual-fuel
mode has proven to be quite effective in reducing both soot and NOx emissions with positive effects
also in brake-specific energy consumption [5]. However, compression ignition (CI) combustion with
natural gas either as a diesel oil supplement or using pilot diesel oil quantity for combustion initiation
and flame development resulted in significant deterioration of CO and HC emissions compared to
conventional diesel operation [5]. The influence of biogas and waste fats methyl esters on NOx,
CO, and PM emissions of dual fuel engines has been examined in the past with quite encouraging
results regarding the improvement of the environmental performance of these engines [6]. Also,
the combustion of alternative gaseous fuels such as a mixture of methane with hydrogen have been
examined in spark-ignition (SI) engines with quite positive results regarding the environmental
repercussions of this type of engine [7].
Another more cost-effective method than natural gas thought to also drastically reduce mainly
diesel-emitted particulate matter (PM) is the use of other alternative gaseous and liquid biofuels such
as straight vegetable oils, methyl esters of vegetable oils, alcohols, and ethers. Vegetable oils and
vegetable oil esters can be used in CI engines either as neat fuels or as blends with conventional diesel
oils without any serious engine modifications. Experimental and theoretical studies conducted in the
past have examined the use of vegetable oils and their esters in CI engines under both steady-state
and transient conditions. Specifically, a heat release rate and performance analysis and also an
environmental behavior assessment of a six-cylinder turbocharged (T/C) direct-injection (DI) diesel
engine using blends of diesel oil with esters of cottonseed oil or sunflower oil had been performed in
the past [7]. The experimental results of this study [8] showed significant reduction of diesel-emitted
soot, which increase with increasing biofuel content in fuel blend. In another experimental study [9],
the combustion of cottonseed oil, cottonseed oil methyl ester, ethanol, and butanol on the performance
characteristics and the exhaust emissions of a T/C multi-cylinder CI engine was investigated, and it
was found that the use of either ethanol or butanol results in the defeat of the well-known soot/NOx
trade-off of conventional CI engines. In a previous experimental investigation [10], the beneficial
effects of blends of butanol and diethyl ether with either vegetable oils or biodiesels on diesel engine
environmental behavior have been demonstrated. Also, in another important study [11], the individual
impact of ethanol, n-butanol, and diethyl ether as biofuel supplements on CI engine performance
characteristics, cyclic variability, and emissive behavior has thoroughly been examined.
Chauhan et al. [12] reviewed the effect of various blends of conventional diesel oil with various
types of ethers and vegetable oils on CI engine performance and emissions and concluded that the
use of 10% to 20% biodiesel addition to diesel fuel can be favorable for long-term use in CI engines.
Also, in a recent study [13], the impact of various additives, including oxygenated ones, to diesel
engine-out regulated and non-regulated emissions was examined. Apart from the aforementioned
reviews, Giakoumis et al. [14] examined the influence of various biofuel blends on combustion-induced
noise radiation using internal and published data. Also, the relative impact of various biodiesels on
exhaust emissions under diesel engine transient operation has been thoroughly been reviewed [15].
In a recent review [16], very informative correlations for predicting biodiesel cetane number using
various fatty acid compositions have also been developed. Apart from the experimental investigations
performed in the past using various liquid and gaseous biofuels in CI engines, theoretical studies
have also been performed using either multi-dimensional computational fluid dynamics (CFD) or
phenomenological simulations models of the operational performance and the polluting potentiality
of CI engines burning alternative oxygenated fuels. Detailed two-zone phenomenological models
initially developed for conventional CI engine performance and emission analysis [17] and also for
EGR analysis [18] have been properly modified and used for investigating the effect of oxygenated
Energies 2019, 12, 1547 3 of 36

fuels on CI engine operational and environmental behavior [19]. Also, multi-zone models have been
used for examining the influence of diesel/biofuel blending ratio and molecular structure on CI engine
performance and emissions [20]. Finally, comprehensive CFD models have been used for assessing the
impact of oxygenated fuel properties on diesel spray combustion and soot formation [21]. The influence
of fuels chemical composition and properties on CI engine transient operational and environmental
performance has been examined in the past using detailed simulations models [22].
The present study emphasizes the experimental examination of the use of blends of diesel oil with
either a mixture of diethylene glycol dimethyl ether and diethylene glycol dibutyl ether called glycol
ethers or rapeseed methyl ester (RME). Studies conducted in the past investigated the impact of either
pure RME or blends of RME with conventional diesel oil on diesel engine performance characteristics
and exhaust emissions showing that the use of RME can mainly result in reduction of diesel-emitted
soot concentration and particle size and number. Specifically, Labecki et al. [23] examined the influence
of diesel and RME on high-speed DI diesel engine-emitted soot particle number–size distribution
considering various injection parameters and EGR rates. Klyus [24] investigated the impact of various
biofuels on CI engine technology. Buyukkaya [25] analyzed the impact of biodiesel content and type on
DI diesel engine combustion characteristics and pollutant emissions and Allocca et al. [26] performed
measurements in a quiescent combustion chamber and in a EURO5 CI engine using conventional diesel
oils and RME. Imran et al. [27] thoroughly examined the impact of RME on CI engine performance
and emissions under various engine speeds and loads and demonstrated the positive influence of
this biofuel on CI engine soot reduction. Previously published experimental studies have compared
the relative impact of RME on CI engine performance and emissions compared to other biodiesels
such as soybean oil methyl ester (SME) and palm oil methyl ester (PME) [28], gas-to-liquid (GTL) [29],
and bioethanol [30]. As in many cases, RME combustion in CI engines results in deterioration of NOx
emissions compared to conventional diesel operation; the use of exhaust gas recirculation (EGR) has
also been examined in CI engine burning blends of conventional diesel oil and RME as an effective
method for NOx emissions curtailment [31]. Experimental studies performed in the past have also
investigated the use of glycol ethers and RME either individually or on a comparative basis for
assessing their relative influence on CI engine combustion characteristics, performance parameters,
and pollutant emissions. Specifically, Gómez-Cuenca et al. [32] and Song et al. [33] performed engine
tests with blends of conventional diesel oil and glycol ethers and they demonstrated the positive
influence of ethers on CI engine environmental behavior. In addition, the relative impact of glycol
ethers and RME as additives to diesel oil has been experimentally assessed in a Ricardo/Hydra CI
engine and showed that glycol ethers can be more beneficial on soot reduction compared to RME [34].
Also, the addition of ethers to conventional diesel fuels with increasing aromatic content has been
investigated showing that the combination of low aromatic content in the parental fuel with increased
glycol ether content can substantially reduce diesel-emitted soot [35]. Of particular importance is a
previous review study [36] where various alternative oxygenated fuels were compared with intake-air
oxygen enrichment as two in-cylinder mixture oxygen-enhancement techniques, which showed that
fuel-bound oxygen and oxygenated fuel properties have a direct effect on local soot formation and
especially on local soot oxidation rate inside fuel spray. Another important experimental study is the
one of Song et al. [33], which examined the effect of blends of a Finnish summer grade diesel oil with
glycol ethers at two different proportions (10 mass-% and 30 mass-%) and RME (30 mass-%) and they
derived informative results regarding the impact of ethers and RME on combustion evolution through
photographic studies, on performance characteristics, and on exhaust emissions.
Having thoroughly examined existing literature in the field of liquid oxygenated biofuels,
one question is raised; whether alternative oxygenated fuels such as glycol ethers and RME can be used
in old technology diesel engines for substantially improving their environmental performance without
considerable detrimental effects in their operational availability and their performance characteristics.
This question is essential since most of the related published studies refer to modern diesel engines
with electronically controlled fuel injection systems. However, worldwide, there is a large fleet of
Energies 2019, 12, 1547 4 of 36

surpassed technology diesel engines, which require serious improvement of their environmental
behavior. Another question raised is the feasibility of the development of an experimental apparatus
based on a DI diesel engine with very low budget, which can be used for fuels testing and also for
educational purposes since it has been proved that the training of young engineers on diesel engine
technology can be quite favorable on the development of their engineering skills [37].
Also, engine development [38] and engine measuring hardware and software companies [39]
have developed commercial instrumentation and software for analyzing cylinder pressure sensor
data and performance CI engine performance analysis without details about the methods used for
analyzing the cylinder pressure signals and for performing performance and combustion analysis.
In addition, a relevant experimental study [33] conducted in the past with glycol ethers and RME
as diesel oil additives in single-cylinder DI diesel engine have not described in detail the cylinder
pressure processing and performance analysis methodology.
Hence, the main objective of the present study is to experimentally examine the effect of
conventional diesel oil and its blends with either glycol ethers or RME for assessing their individual
impact on the combustion characteristics and soot, NO, CO, and total unburned hydrocarbons (HC)
emissions of a relatively old technology DI diesel engine. Besides that, another objective of the present
analysis is to demonstrate the development and use of a DI diesel engine experimental installation with
limited funding, which can be used not only for fuels testing, but also for other applications and for
young engineers training. Another objective of the present study is to describe, in detail, a theoretical
methodology for processing cylinder pressure, injection pressure, and top dead center (TDC) position
data obtained over many engine cycles for calculating the engine performance parameters and
combustion characteristics either on a cycle-to-cycle basis or on average cycle basis. Unlike the past,
the detailed description of this methodology can be adopted by scientists and engineers working on
CI engine development and it can also be used for educational purposes. Consequently, tests were
performed in single-cylinder naturally aspirated DI diesel engine (“Lister LV1”) using one conventional
diesel oil, two of its blends with glycol ethers, and a blend of diesel oil with RME at 2500 rpm and at 20%,
40%, 60%, and 80% load. Also, a computational model was developed in MATLAB [40] for processing
measured cylinder pressure, injection pressure, and TDC position data from all engine operating
points and for calculating engine performance and combustion characteristics. The evaluation of
the experimental results showed that for the same blending proportion in conventional diesel oil,
a biodiesel-agent (RME) appears to have less detrimental impact on engine performance parameters
and specific fuel consumption and more positive impact on HC emissions reduction compared to a
mixture of synthetic oxygenates (glycol ethers). Oppositely, the same percentage of glycol ethers in a
diesel blend appears to have more beneficial impact on soot and NO emissions compared to RME.

2. Test Fuels Description


The test fuels examined in the present study were prepared under a combined theoretical and
experimental investigation, which was conducted in the past and aimed at the computational and
experimental examination of various conventional and alternative blends that could be used in both
existing (at the time) and future fleets of diesel-powered vehicles [20,34,36]. The main purpose of
this extensive and detailed investigation was the determination of the optimum diesel fuel chemical
synthesis and physical and chemical properties for attaining reduction of diesel emitted pollutants
without deteriorating or, if possible, further improving the specific fuel consumption of existing and
future diesel engines. Hence, under this elaborate research investigation, a series of non-oxygenated
and oxygenated diesel fuels were prepared by an oil refinery in Finland for examining, among other
fuel parameters, the effect of fuel oxygen content and oxygenate molecular structure on diesel engine
performance characteristics and gaseous and particulate emissions [20,34,36]. Initially, Finnish summer
grade city diesel oil was chosen as the reference fuel or, in other words, the “base” fuel (DI1) of
this investigation. For the development of “base” fuel DI1, three oxygenated diesel blends were
prepared with oxygen content ranging from 3% to 9%. In particular, it was decided to use diethylene
Energies 2019, 12, 1547 5 of 36

glycol dimethyl ether (Diglyme), diethylene glycol dibutyl ether (Butyl Diglyme), and rapeseed
methyl ester (RME) as oxygenated additives to “base” fuel DI1. Both diglyme and butyl diglyme
are called glymes [20,34,36]. The blending of Finnish summer grade city diesel (DI1) with RME at
31.2 mass-% provided an oxygenated diesel blend with 3% oxygen. Rapeseed methyl ester is a biofuel
with high viscosity and density, which can be prepared from biological sources such as vegetable
oils (rapeseed oil). Esters of vegetable oils such as rapeseed oil, soybean oil, and cottonseed oil are
considered as renewable alternative solutions for diesel oils [26–29]. Methyl esters are produced from
corresponding vegetable oils through esterification process [9,16,26–29]. The addition of a mixture
of diglyme (C6 H14 O3 ) and butyl diglyme (C12 H26 O3 ) to “base” fuel DI1 at different blending ratios
produced two oxygenated diesel fuels, namely GLY10 (89.8 mass-% DI110.2 mass-% glymes) and
GLY30 (68.7 mass-% DI1-1.3 mass-% glymes) with 3% and 9% oxygen content, respectively. Glycol
ethers are synthetically produced oxygenated substances that indicate high cetane number and low
value of heat of combustion. The chemical composition and the physical and chemical properties of
the test fuels DI1, RME30, GLY10, and GLY30 are presented in Table 1 as delivered from an oil refinery
in Finland. It is worth mentioning here that the measured results for the composition of fuels GLY10
(89.8 mass-% DI1-0.2 mass-% glymes) and GLY30 (68.7 mass-% DI1–31.3 mass-% glymes) in glymes
and in base fuel DI1 are slightly different compared to the corresponding scheduled percentages of
the initial fuel recipes as evidenced from the data of Table 1. The symbol “mass-%” corresponds to
the blending mass composition of a test fuel e.g., test fuel GLY30 contains 16.6 mass-% diglyme and
14.7 mass-% butyl diglyme, which means that 1 kg of test fuel GLY10 contains 166 g diglyme and 147 g
butyl diglyme; thus, 313 g glymes.

Table 1. Chemical composition, physical properties, and chemical properties of test fuels DI1, RME30,
GLY10, and GLY30 examined in the present study [20,34,36].

Fuel Method DI1 RME30 GLY10 GLY30


DI1 + RME DI1 + Glymes DI1 + Glymes
Description “Base” fuel
3% oxygen 3% oxygen 9% oxygen
90.7 mass-% 71.4 mass-%
DI1 DI1
Finnish 70 mass-% DI1
5.0 mass-% 15.0 mass-%
Fuel Recipe summer grade 30 mass-%
Diglyme Diglyme
city diesel RME
4.3 mass- % 13.6 mass-%
Butyl Diglyme Butyl Diglyme
Paraffines, mass-% ASTM D2425 40.1 28.1 36.4 28.6
Naphtenics, mass-% ASTM D2425 40.1 28.1 36.4 28.6
Aromatics, mass-% ASTM D2549 19.8 13.9 18.0 14.1
Density, 15 ◦ C, kg/m3 ISO 3675 833.7 848.2 840.5 854.9
Distillation, ◦ C, 5% ISO 3405 208.0 221.5 192.9 175.2
Distillation, ◦ C, 50% ISO 3405 270.2 307.3 265.3 254.6
Distillation, ◦ C, 95% ISO 3405 347.9 350.6 348.8 345.2
Cetane number ISO 5165 53.3 54.2 59.5 72.7
Viscosity +40 ◦ C, mm2 /s ISO 3104 2.94 3.27 2.54 1.99
LHV, MJ/kg ISO 1928 43.03 41.22 41.48 38.75
Compressibility, 60 bar,
FEV 6.81 6.56 6.87 6.98
20 ◦ C, [× 10−5 bar−1 ]
ISO 8754 /
Sulphur, mg/kg 31 23 28 23
D3210
Diglyme, mass-% GC-AED 0 0 5.5 16.6
Butyl Diglyme, mass-% GC-AED 0 0 4.7 14.7
Glymes, mass-% Calc. 0 0 10.2 31.3
RME Content GC-AED 0 31.2 0 0
Carbon, % ASTM D5291 86.1 83.0 83.2 78.1
Hydrogen, % ASTM D5291 14.0 13.3 13.6 13.2
C/H ratio ASTM D5291 6.1 6.20 6.10 5.94
Oxygen content Calc. 0 3 3 9
Energies 2019, 12, 1547 6 of 36

From the examination of the specifications provided in Table 1, it can be concluded that the
blending of RME and glymes with “base” fuel DI1 provide the opportunity to investigate the effect
of oxygenated additive type on diesel engine performance and combustion characteristics and also
on diesel-emitted pollutants. This examination is facilitated using test fuels RME30 and GLY30,
which have almost the same percentage (31 mass-%) of two different oxygenated additives, namely
RME and glymes. Also, test fuels DI1 (0% oxygen), GLY10 (3% oxygen), and GLY30 (9% oxygen)
formulate a series of fuels with progressively increasing oxygen content using the same oxygenated
additives (mixture of glymes) at different blending ratios. As a result, the experimental results obtained
for the examined diesel engine performance parameters and exhaust emissions can be used for the
examination of the influence of increasing oxygen content on HSDI diesel engine operational and
environmental behavior.

3. Description of the Diesel Engine Experimental Installation, the Fuel Testing Procedure, and the
Processing Methodology of the Measured Cylinder Pressure, Injection Pressure, and TDC
Position Data

3.1. Description of the Diesel Engine Experimental Installation


An experimental apparatus was installed in the past at the Internal Combustion Engines Laboratory
of National Technical University of Athens, Greece based on a single-cylinder DI diesel engine coupled
with a hydraulic dynamometer. In the specific experimental installation, devices for controlling diesel
engine operation and monitoring its operational parameters were installed. The diesel engine used in
the experiments is “Lister LV1” engine, which was developed by Lister Petter Company. The main
technical data of the “Lister LV1” engine are given in Table 2. In Table 3, the main operating parameters
of the “Lister LV1” engine, which were used during the engine tests, are given. The main operating
parameters shown in Table 3 are the measured barometric pressure in the laboratory, the measured
relative humidity, the dry air temperature of the laboratory, and the hydraulic dynamometer load in
kg. Also, an effort was made that during engine tests, the pressure of the lubricant oil system to was
have limited variations in the range between 0.7 and 1.0 bar. “Lister LV1” engine is coupled with a
Heenan and Froude hydraulic dynamometer [20,34,36].

Table 2. Main technical data of the diesel engine “Lister LV1” used in the experiments [20,34,36].

Technical Specification Definition


Number of cylinders 1
Engine operation type Four-stroke
Air breathing type Naturally aspirated
Cooling method Air cooled
Combustion chamber Direct injection—“Bowl-in-piston”
Cylinder bore 85.73 mm
Piston stroke 82.55 mm
Connecting rod length 188.5 mm
Compression ratio 17:1
Nominal speed range 1000–3000 RPM
Maximum power 6.7 kW @ 3000 RPM
Maximum torque 25 Nm @ 2000 RPM
Number of fuel injector holes 3
Nozzle orifice diameter 0.250 mm
Injector opening pressure 180 bar
Number of valves per cylinder 2
Intake valve opening 15 degCA BTDC
Intake valve closing 41 degCA ABDC
Exhaust valve opening 41 degCA BBDC
Exhaust valve closing 15 degCA ATDC
Energies 2019, 12, 1547 7 of 36

Table 3. Main operating parameters of the diesel engine “Lister LV1” used in the experiments.

Laboratory Laboratory Laboratory Lubricant


Engine Hydraulic Engine
Barometric Relative Dry Air Engine Oil System
Speed Dynamometer Power
Pressure Humidity Temperature Load (%) Pressure
(RPM) Load (kg) (kW)
(mmHg) (%) (o C) (bar)
771.8 58.5 26.2 1 1.23 20 0.7–1
770.6 57 27.5 2 2.45 40 0.7–1
2500
769.5 56.5 27.8 3 3.68 60 0.7–1
768 56 28 4 4.90 80 0.7–1

The main measuring equipment comprised of an Alcock air flow meter (viscous type), fuel
tank, and flow-meter for measuring engine fuel consumption, thermocouples for recording exhaust
gas temperature, lubricant oil temperature and engine coolant temperature, a magnetic pickup for
recording TDC position, a crankshaft rotational speed indicator, and a piezoelectric transducer for
measuring in-cylinder pressure [35,41]. The piezoelectric transducer used for recording cylinder
pressure was developed by Kistler and it was a Kistler 6001 type. The operational temperature range
of the Kistler 6001 piezoelectric transducer was from −196 to 350 ◦ C, its sensitivity was 15 pCb/bar,
and its linearity was lower than ±0.8% at full scale indication [35,41].
A similar piezoelectric transducer was fitted to the high-pressure fuel line between pump and
injector close to the injector for monitoring fuel injection pressure. This piezoelectric transducer was
also developed by Kistler and it was a Kistler 6005 type. The operational temperature range of the
Kistler 6005 piezoelectric transducer was from −196 to 240 ◦ C, its sensitivity was 7.91pCb/bar and its
linearity was lower than ±0.1% at full scale indication [35,41]. An oscilloscope was used to monitor
cylinder pressure or injection pressure signal.
It was also used a fast data acquisition system for recording cylinder pressure, injection pressure,
and TDC position measurements and storing them in a PC [35,41]. The data acquisition system was
based on an analog-to-digital (ADC) converter DAS-1801ST of Keithley, which was installed in a
personal computer. The specific ADC converter can record 8 differential or 16 single-phase signals with
12-bit resolution in each one. In the engine tests described in this study, the ADC converter recorded 3
analog signals from two piezoelectric transducers and the TDC position magnetic pickup through a
STP-50/C terminal board [35,41].
In addition, exhaust gas analyzers were used for the measurement of tailpipe soot, NO, CO,
and HC values. A Bosch RTT100 smoke meter was used for the measurement of diesel-emitted soot
concentration in mg/L at the engine exhaust. NO emissions were measured in ppm using a Signal
4000 type chemiluminescent analyzer and HC emissions (equivalent propane) were measured also in
ppm using a flame ionization analyzer manufactured by Ratfisch. Both NO and HC analyzers were
equipped with thermostatically controlled heated lines. Finally, CO emissions were measured in ppm
using a Signal 3001 type non-dispersive infrared analyzer [35,41].
Figure 1 shows a schematic view of the experimental apparatus used to perform engine tests in
“Lister LV1” DI diesel engine using test fuels DI1, RME30, GLY10, and GLY30.
Figure 2 shows a photograph of the Alcock viscous type air flow meter used during engine tests to
measure intake air flow rate. Alcock air flow meter is specifically designed in a way that the measuring
errors from air flow pulsations in engine intake duct due to the reciprocating piston movement are
eliminated. In the examined experimental apparatus, the Alcock flow meter is placed after the air
filter. The measuring element of the Alcock flow meter is a cell with triangular cross section passages,
which each one of them 76 mm long and 0.43 mm high. The pressure drop of the air is attributed
to the cell resistance, which is analogous to the air flow velocity. Hence, the air volume flow rate
through the measuring element of the Alcock flow meter varies linearly with the pressure drop, thus
minimizing the measuring error. The inclined tube of the Alcock flow meter contains a mixture of
water and alcohol with specific weight equal to 0.82 at ambient temperature. The tube has a variable
inclination corresponding to three different positions, which each inclined position corresponding to
Energies 2019, 12, 1547 8 of 36

different range of measured flow rates. In the present study, the middle inclined position of the tube
was used
Energies [35,41].
2018, 11, x FOR PEER REVIEW 8 of 36
Energies 2018, 11, x FOR PEER REVIEW 8 of 36

GasGas
Data Acquisition Card
NO CO HC
Data Acquisition Card

Exhaust
NO CO HC

Exhaust
TDC Magnetic
Pickup
TDC Signal
Magnetic Filter
Pickup Signal
Amplifier Filter
Alcock Air
Flowmeter Amplifier
Alcock Air Injection Amplifier Smokemeter
Flowmeter Pressure Signal Amplifier
Injection Smokemeter
Pressure Signal
Cylinder Pressure Signal
Cylinder Pressure Signal
FUEL TANK F T LISTER Exhaust Gas
FUEL TANK F T LV1
LISTER T Exhaust Gas
ENGINE
LV1
T
ENGINE T Thermometer
Heenan & T Thermometer
Froude&
Speed F Flowmeter
Heenan Indicator
Dynamometer
Froude
Speed F Flowmeter
Indicator
Dynamometer
1. Schematic
Figure 1. Schematicview
viewofofthe
theexperimental
experimentalinstallation used
installation to perform
used engine
to perform teststests
engine in “Lister LV1”
in “Lister
Figure
DI
LV1” DI1.diesel
diesel Schematic
engine usingview
engine oftest
theDI1,
test fuels
using experimental
RME30,
fuels installation
GLY10,
DI1, RME30, used
and GLY30.
GLY10, and to perform engine tests in “Lister
GLY30.
LV1” DI diesel engine using test fuels DI1, RME30, GLY10, and GLY30.

Figure 2. Photographic view of the Alcock air flow meter installed at the intake of the diesel engine
Figure 2.
“Lister Photographic
LV1” [35,41]. view of the Alcock air flow meter installed at the intake of the diesel engine
“Lister LV1”
“Lister LV1”[35,41].
[35,41].
It is of particular importance to discuss the installation and the use of the magnetic pickup on
It is of particular
of flywheel importance
particularforimportance to discuss the
theinstallation and thethe useuse
of the magnetic pickup on the
the engine recording to discuss
TDC position installation
signal. For andthis reason, of
in the magnetic
Figure pickup
3a is shown on
the
engine
the flywheel
engine for recording
flywheel for TDC position
recording TDC signal. For
position this For
signal. reason,
this inreason,
Figure in
3a is shown
Figure 3a the
is installation
shown the
installation of the magnetic pickup on “Lister LV1” engine flywheel and in Figure 3b is shown a
of the magnetic
installation pickup
of the on “Lister LV1”onengine flywheel and in flywheel
Figure 3bandis shown a schematic view ofa
schematic view ofmagnetic pickup
the complete TDC “Lister
position LV1” engine
signal measuring and in Figure
recording 3b is shown
apparatus. As
the complete
schematic TDC
view ofposition
the signal
complete measuring
TDC and
position recording
signal apparatus.
measuring As
and evidenced
recording from Figure
apparatus. 3a,
As
evidenced from Figure 3a, a plastic bar is placed on the flywheel following the rotation of the engine
aevidenced
plastic barfrom
is placed
Figure on3a,
thea flywheel
plastic following the rotation of the following
engine crankshaft. When the engine
plastic
crankshaft. When the plastic bar is bar is placed
aligned with onthethe flywheel
magnetic pickup, thenthe therotation
circuit isof closed
the and
bar is aligned with the magnetic pickup, then the circuit is closed and the TDC position signal becomes
the TDC position signal becomes equal to zero. Though, during recent decades, a shaft encoder and
crankshaft. When the plastic bar is aligned with the magnetic pickup, then the circuit is closed has
equal
the TDC to zero. Though,
signalduring recent decades, a shaft encoder hasrecent
been used for monitoring cylinder
been usedposition
for monitoring becomes
cylinderequal to zero.
pressure, whichThough, during
is capable of recording decades, a shaft
pressures withencoder has
very small
been used for monitoring cylinder pressure, which is capable of recording pressures
crank angle steps, e.g., 0.1 or 0.2 of crank angle degree, the use of magnetic pickup is a cost-effective with very small
crank
and angle steps,
reliable method e.g.,
for0.1 or 0.2 of crank
measuring TDC angle
positiondegree,
signalthe useitof
and hasmagnetic
been usedpickup is a cost-effective
in many engine tests
and reliable method for measuring TDC
with conventional and alternative diesel oils [42]. position signal and it has been used in many engine tests
with conventional and alternative diesel oils [42].
Energies 2019, 12, 1547 9 of 36

pressure, which is capable of recording pressures with very small crank angle steps, e.g., 0.1 or 0.2 of
crank angle degree, the use of magnetic pickup is a cost-effective and reliable method for measuring
TDC position signal and it has been used in many engine tests with conventional and alternative diesel
oils [42].
Energies 2018, 11, x FOR PEER REVIEW 9 of 36

Analog to Digital
Recording System on PC

Plastic
bar Signal
Amplifier Oscilloscope

“LISTER LV1” HYDRAULIC


DIESEL ENGINE BRAKE

TDC Position Recording


Magnetic Pickup

(a) (b)
3. (a) Photographic view of the magnetic pickup installation on “Lister LV1” engine flywheel
Figure 3.
and (b) schematic
schematic view of the
the complete
complete installation
installation of the top
top dead
dead center
center (TDC)
(TDC) position
position magnetic
magnetic
pickup
pickup with
withthe
thesignal amplifier,
signal the the
amplifier, oscilloscope, and the
oscilloscope, andPCtheanalog-to-digital recording
PC analog-to-digital system [35,41].
recording system
[35,41].
3.2. Fuels Testing Procedure
3.2. Fuels Testing
Engine testsProcedure
were performed with test fuels DI1, RME30, GLY10, and GLY30 at 2500 rpm and
at four engine loads namely 20%, 40%, 60%, and 80% of full load. All tests were carried out using
Engine tests were performed with test fuels DI1, RME30, GLY10, and GLY30 at 2500 rpm and at
constant static injection timing (26 degCA BTDC) of the fuel injection system to avoid variations in fuel
four engine loads namely 20%, 40%, 60%, and 80% of full load. All tests were carried out using
injection commencement inside the combustion chamber due to static injection timing variations. Also,
constant static injection timing (26 degCA BTDC) of the fuel injection system to avoid variations in
a serious effort was made all engine tests to be performed without noticeable variations of intake air
fuel injection commencement inside the combustion chamber due to static injection timing
temperature and lubricant oil temperature as a method for avoiding engine operation fluctuations and
variations. Also, a serious effort was made all engine tests to be performed without noticeable
most importantly,
variations of intakeengine loading variations.
air temperature Engineoil
and lubricant testing procedure
temperature ascomprised
a method of
forthe following
avoiding two
engine
steps [35,41]:
operation fluctuations and most importantly, engine loading variations. Engine testing procedure
comprised
1. of engine
Initially, the following twocarried
tests were steps [35,41]:
out at 2500 rpm and at all engine loads (20%, 40%, 60%, and 80%
of full load) using only “base” fuel
1. Initially, engine tests were carried out DI1. At each
at 2500 rpmoperating
and at all point,
enginevarious engine
loads (20%, 40%,operational
60%, and
parameters were recorded such as fuel consumption, exhaust gas temperature,
80% of full load) using only “base” fuel DI1. At each operating point, various engine intake temperature,
flow mass rate,
operational cylinder pressure,
parameters and injection
were recorded such as pressure. Hence, usingexhaust
fuel consumption, this testing
gas methodology,
temperature,
the engine
intake baseline operation
temperature, flow massfor thecylinder
rate, reference fuel DI1and
pressure, wasinjection
constituted.
pressure. Hence, using this
2. testing
The previous testing procedure
methodology, the enginewas repeated
baseline for thefor
operation same
the engine operating
reference fuel DI1conditions for each
was constituted.
one previous
2. The of the oxygenated test fuels was
testing procedure RME30, GLY10,
repeated forand
the GLY30.
same engine operating conditions for each
one of the oxygenated test fuels RME30, GLY10, and GLY30.
It should be clarified that all engine tests at 2500 rpm and at each engine load were performed by
It should
keeping be clarified load
the dynamometer that all
(i.e.,engine testsbrake
hydraulic at 2500 rpm and
weight) at each
constant engine
when load were
changing from performed
one fuel to
by keeping
another. the dynamometer
Hence, since variationsload (i.e., of
in heat hydraulic
combustionbrake weight)
were constantbetween
experienced when changing
fuels DI1,from one
RME30,
fuel to another.
GLY10, Hence,
and GLY30, the since variations in was
fuel consumption heatadjusted
of combustion
at eachwere
loadexperienced
when moving between fuels
from one DI1,
fuel to
RME30, in
another GLY10,
order and GLY30,
to keep the fuel consumption
the dynamometer was adjusted at each load when moving from one
weight constant.
fuel to another in order to keep the dynamometer weight constant.
As already mentioned, the experimental single-cylinder DI diesel engine considered in the
present study (“Lister LV1”) is directly connected to a proper hydraulic dynamometer. The
hydraulic brake manufacturer provided the following equation for the calculation of engine brake
power [35,41]:
W ⋅ RPM
Pe = (1)
1500
where Pe is the engine brake power in CV, W is the hydraulic dynamometer load indication in kg,
and RPM is the rotational speed of the engine shaft.
Energies 2019, 12, 1547 10 of 36

As already mentioned, the experimental single-cylinder DI diesel engine considered in the present
study (“Lister LV1”) is directly connected to a proper hydraulic dynamometer. The hydraulic brake
manufacturer provided the following equation for the calculation of engine brake power [35,41]:

W·RPM
Pe = (1)
1500
where Pe is the engine brake power in CV, W is the hydraulic dynamometer load indication in kg,
and RPM is the rotational speed of the engine shaft.
Engine intake air flow rate was measured using an Alcock measuring device, which measures
intake pressure variation and it was calibrated for air temperature of 20 ◦ C. Hence, the intake air
volume flow rate is calculated using the following relation, which has been derived according to a
proper diagram accompanying the Alcock measuring device [35,41]:
.
VA = 1.698·(Alcock value) (2)
.
where VA is the intake air volume flow rate in m3 /h and Alcock value is measured in cm. During
experiments, the laboratory dry air temperature Troom and the barometric pressure proom were recorded
and the results are given above in Table 3 at each operating point. The measurements for laboratory air
temperature and barometric pressure were used for calculating the engine intake air mass flow rate
using measured intake air volume flow rate and the ideal gas equation of state [35,41]:

. proom .
mA = VA (3)
RTroom

A constant volume tube of 50 mL is used for measuring engine fuel consumption. The measurement
of fuel consumption is based on the measurement of time required for the evacuation of the 50 mL
diesel fuel tube from the engine. Consequently, the fuel consumption in kg/h is calculated using the
following relation [35,41]:
 kg 
−6 m3
!     
s 
. kg  50(mL)·ρf m3 ·10 mL ·3600 h 
mf =  (4)
∆t(s)

h  

where ρf is fuel density and ∆t in s is the time required for the evacuation of the 50 mL diesel fuel tube
during engine operation. Results for the estimated intake air mass flow rate in kg/h and the measured
fuel flow rate in kg/h of the engine at each operating point are given in Table 4.

Table 4. Results for the estimated intake air mass flow rate and the measured fuel flow rate of the
engine for each one of the test fuels DI1, RME30, GLY10, and GLY30 at 2500 RPM and at 80% of full load.

Intake Air
Intake Air Intake Air Measured Time Fuel
Alcock Volume
Fuel Density Mass Flow of 50 mL Fuel Consumption
Value (cm) Flow Rate
(kg/m3 ) Rate (kg/h) Consumption (s) (kg/h)
(m3 /h)
DI1 16.8 28.53 1.197 34.159 105 1.429
RME30 16.7 28.36 1.190 33.756 99 1.542
GLY10 16.7 28.36 1.188 33.675 101 1.498
GLY30 16.9 28.70 1.184 33.989 103 1.494

3.3. Description of the Processing Procedure of the Cylinder Pressure, Injection Pressure, and TDC Position
Signals Initially Obtained from the Diesel Engine
A computational model was developed in MATLAB [40] for processing initial experimental signals
for cylinder pressure, injection pressure, and TDC position previously obtained during an engine
testing procedure in “Lister LV1” using test fuels DI1, RME30, GLY10, and GLY30. Engine tests were
Energies 2019, 12, 1547 11 of 36

performed at 2500 rpm and at four engine loads namely 20%, 40%, 60%, and 80% of full engine load.
The developed model was used to process the aforementioned experimental signals for generating
the average cylinder pressure–crank angle profile and the average fuel injection pressure–crank
angle profile and then to use these profiles for calculating the main performance and combustion
characteristics of the “Lister LV1” for all test fuels examined. Of particular importance is the description
of the mathematical process adopted from the developed model to process the initial cylinder pressure,
injection pressure, and TDC position signals over all obtained complete engine cycles for generating
the average cylinder pressure and the average injection pressure profiles for each examined engine
operating point and each examined test fuel.
Hence, during the engine testing procedure and at each engine operating point after setting the
rack position
Energies 2018, 11,in the PEER
x FOR fuel pump
REVIEW and after securing constant operating conditions, the following signals 11 of 36
obtained from the engine were recorded in a personal PC through a fast data acquisition system:
system:
The The cylinder
cylinder pressurepressure
signal, thesignal, the injection
injection pressure pressure
signal, and signal, and the
the TDC TDC signal.
position positionInsignal.
FigureIn4,
Figure
the 4, the characteristic
characteristic signals of signals
cylinder ofpressure
cylinder (Figure
pressure (Figure
4a), 4a), pressure
injection injection (Figure
pressure (Figure
4b), and TDC4b),
and TDC
position position
(Figure (Figure
4c) are shown,4c)as are shown,
obtained fromasthe
obtained from during
diesel engine the diesel engine
testing during
procedure testing
of fuel DI1
procedure
at 2500 rpmof fuel
and 80%DI1ofat 2500
full rpm
load. and 80%
It should of full load.that
be mentioned It should be mentioned
in the specific that ininvestigation,
experimental the specific
experimental
100 consecutive investigation,
engine cycles 100 consecutive
were recorded engine
at eachcycles
enginewere recorded
operating at as
point each engine operating
evidenced also from
point as
Figure evidenced
4a–c. In Figure also from
4a,b, Figurepressure
cylinder 4a–c. InandFigure 4a,b, pressure
injection cylinder signals
pressure forand injection pressure
10 consecutive cycles
signals
are shown.for 10 consecutive cycles are shown.

(a) (b)

0.6
"LISTER LV1" ENGINE
2500 RPM - 60% LOAD
0.4 DI1

0.2

-0.2

-0.4
0 1200 2400 3600 4800 6000 7200
NUMBER OF MEASUREMENTS (-)
(c)
Figure 4. (a) Cylinder
Cylinder pressure
pressuresignal,
signal,(b)
(b)fuel
fuelinjection
injectionpressure
pressuresignal,
signal,andand(c)(c)
TDCTDCposition signal
position as
signal
obtained from
as obtained thethe
from “Lister LV1”
“Lister engine
LV1” during
engine testing
during procedure
testing procedure forfor
testtest
fuel DI1
fuel at 2500
DI1 rpm
at 2500 rpmand at
and
60% of full engine load.
at 60% of full engine load.

As evidenced from
As evidenced from Figure
Figure5,5,TDC
TDCposition
positionsignal
signal undergoes,
undergoes, at at a specific
a specific acquisition
acquisition point,
point, an
an abrupt rise and immediately afterwards a steep reduction around zero. The cross-section
abrupt rise and immediately afterwards a steep reduction around zero. The cross-section point of point
the
of
TDCthe position
TDC position
curve,curve,
whichwhich connects
connects the the
locallocal maximum
maximum with
with thethe localminimum
local minimumofof the
the TDC
TDC
position signal with the zero horizontal line, corresponds to the piston immobilization position at
TDC (Figure 6). The number of measurements between two consecutive positions at which the TDC
position signal becomes equal to zero provides the number of measurements, which were actually
received during a complete engine cycle. The point at which the TDC position signal becomes equal
to zero is numerically tracked using the bisection method.
Energies 2019, 12, 1547 12 of 36

position signal with the zero horizontal line, corresponds to the piston immobilization position at
TDC (Figure 6). The number of measurements between two consecutive positions at which the TDC
position signal becomes equal to zero provides the number of measurements, which were actually
received during a complete engine cycle. The point at which the TDC position signal becomes equal to
zero is 2018,
Energies
Energies numerically
2018, 11,
11, xx FOR tracked
FOR PEER
PEER using the bisection method.
REVIEW
REVIEW 12
12 of
of 36
36

Figure
Figure 5. 5. Graphical
Graphical explanation
Graphical explanation of
explanation of the
of the procedure
the procedure followed
followed for
followed for the
the determination
determination of
determination of the
the number
number of of
complete engine
complete engine cycles.
engine cycles. Experimental
cycles. Experimental data
Experimentaldata are
dataare given
aregiven for
givenfor “Lister
for“Lister LV1”
“ListerLV1” engine
LV1”engine
engine at
atat 2500
2500
2500 rpm
rpm
rpm and
and
and at
at at
60%60%
60%
of
of
fullfull
of full engine
engine
engine load
load
load using
using
using fuel
fuel
fuel DI1.
DI1.
DI1.

"LISTER
"LISTER LV1"
LV1" ENGINE
ENGINE
2500
2500 RPM
RPM -- 60%
60% LOAD
LOAD -- FUEL
FUEL DI1
DI1
Injection
Injection Pressure Signal
Pressure Signal
Cylinder Pressure Signal
Cylinder Pressure Signal
TDC
TDC Position
Position Signal
Signal
5
5 Injection
Injection Pressure
Pressure
4 Measurement
Measurement
4
corresponding
corresponding
3
3 to
to "Hot
"Hot TDC"
TDC"
2 i.e.
2 i.e. 180
180 degCA
degCA
1
1
0
0
-1 Cylinder
-1 Cylinder Pressure
Pressure
4
4 Measurement
Measurement
corresponding
corresponding
3
3 to
to "Hot
"Hot TDC"
TDC"
i.e.
i.e. 180
180 degCA
degCA
2
2

1
1

0
0
0.6
0.6
0.5
0.5 "Hot
"Hot TDC"
TDC" --
0.4 Local
Local Positive
Positive Bisection
0.4
0.3 Extremum Bisection Method
Method
0.3
0.2
Extremum
0.2
0.1
0.1
0
0
-0.1
-0.1
-0.2
-0.2 Local
Local Negative
Negative
-0.3
-0.3 Extremum
-0.4 Extremum
-0.4
1480
1480 1490
1490 1500
1500 1510
1510 1520
1520 1530
1530 1540
1540 1550
1550 1560
1560
NUMBER
NUMBER OF
OF MEASUREMENTS
MEASUREMENTS (-)
(-)

Figure 6. Graphical
Figure 6. Graphical explanation
explanation of
explanation of “hot
“hot TDC”
TDC” position determination per
position determination per engine cycle. Also,
engine cycle. Also, in
Also, in the
the
figure, the graphical explanation of the cylinder pressure and the injection pressure measurements
figure, the graphical explanation of the cylinder pressure and the injection pressure measurements
tracking corresponding
tracking correspondingtoto
corresponding tothe “hot
the
the TDC”
“hot
“hot TDC”
TDC” is shown.
is
is shown.Experimental
shown. datadata
Experimental
Experimental are given
data are for “Lister
are given
given for LV1” engine
for “Lister
“Lister LV1”
LV1”
at 2500 at
engine
engine rpm
at and
2500
2500 at 60%
rpm
rpm and of
and at full
at 60%engine
60% of full load
of full engine
engineusing
loadfuel
load DI1.
using
using fuel
fuel DI1.
DI1.

At the point
At point that the
the TDC position
position signal becomes
becomes equal to to zero, it
it can be
be matched with
with a value
At the
the point that
that the TDC
TDC position signal
signal becomes equal
equal to zero,
zero, it can
can be matched
matched with aa value
value
of
of crank angle (CA) equal to 180 degCA or 540 degCA (it is assumed that the crank angle isequal
equaltoto
of crank angle (CA) equal to 180 degCA or 540 degCA (it is assumed that the crank angle is equal to 00
crank angle (CA) equal to 180 degCA or 540 degCA (it is assumed that the crank angle is
0 degCA whenthe
degCA the pistonisisatatBDC).
BDC). Thisprocedure
procedure providesus us withthe
the opportunity to to correlate the
the
degCA when
when the piston
piston is at BDC). ThisThis procedure provides
provides us with with the opportunity
opportunity to correlate
correlate the
measured
measured values
values of of cylinder
cylinder pressure
pressure and
and injection
injection pressure
pressure withwith specific
specific crank
crank angles
angles for
for aa
complete engine cycle. Having specified the points at which the TDC position signal
complete engine cycle. Having specified the points at which the TDC position signal becomes equal becomes equal
to
to zero,
zero, we
we can
can then
then determine
determine the the points
points that
that correspond
correspond to to TDC
TDC positions
positions during
during combustion
combustion (180
(180
degCA) i.e., “hot TDCs” since all other TDC position zero points correspond
degCA) i.e., “hot TDCs” since all other TDC position zero points correspond to the gas exchange to the gas exchange
period of an engine cycle (540 degCA). The TDC position zero points that are assigned to 180 degCA
Energies 2019, 12, 1547 13 of 36

measured values of cylinder pressure and injection pressure with specific crank angles for a complete
engine cycle. Having specified the points at which the TDC position signal becomes equal to zero,
we can then determine the points that correspond to TDC positions during combustion (180 degCA)
i.e., “hot TDCs” since all other TDC position zero points correspond to the gas exchange period of
an engine cycle (540 degCA). The TDC position zero points that are assigned to 180 degCA are those
that correspond to the peak values of the cylinder pressure signal. As shown in Figure 6, the tracking
of the “hot TDC” can be used for also identifying the cylinder pressure measurement corresponding
to “hot TDC” or, in other words, to 180 degCA using the measured cylinder pressure signal and the
TDC position signal. Also, as evidenced, the measured injection pressure signal and TDC position
signal can be used for correlating the “hot TDC” measurement with the measured injection pressure
corresponding to “hot TDC” i.e., to 180 degCA.
Having determined the TDC position points and knowing the number of measurements at
each obtained complete engine cycle, the actual acquisition step can be calculated, which is equal to
the ratio of the theoretical number of measurements, which are supposed to be obtained during a
complete four-stroke engine cycle in the case where the acquisition step was set equal to 1 degCA
(720 measurements) to the actual number of measurements obtained during engine tests at a complete
engine cycle (e.g., 722 measurements). Small deviations between the theoretical and the actual total
number of measurements can be ascribed to the fact the engine crankshaft speed indicates small
deviations compared to the specified value during engine tests. Depending on the shaft rotational
speed small variation, the actual acquisition step can be lower than 1 degCA, e.g., if the theoretical
measurements during a complete engine cycle are 720 and the actual measurements are 722, then the
actual acquisition step of measurements is 0.997 degCA. Hence, the distance in crank angle degrees
between the cylinder pressure value at 180 degCA i.e., at “hot TDC” and the previous cylinder pressure
measurement will be equal to the actual acquisition step in crank angle degrees. The same distance in
crank angle degrees will be the cylinder pressure value at 180 degCA and the next cylinder pressure
measurement. For example, if the actual acquisition step is 0.997 degCA and starts from the cylinder
pressure measurement at 180 degCA, the previous cylinder pressure measurement will be assigned
to 180 − 0.997 = 179.003 degCA and the next cylinder pressure measurement will be assigned to
180 + 0.997 = 180.997 degCA. Consequently, using the actual acquisition step in crank angle degrees,
we can start from the cylinder pressure values at 180 degCA and move backward and forward assigning
cylinder pressure measurements to crank angles for all the complete engine cycles. It should be
underlined here that not all measured engine cycles are complete, meaning that the recording of
the pressure and the TDC position signals does not start from the beginning of the first cycle and
it does not end at the end of the last cycle. Hence, usually the first and the last engine cycles are
excluded from the analysis since they do not contain measurements close to pre-described theoretical
total number of measurements per cycle i.e., 720. Having assigned the measured cylinder pressure
values to non-integer crank angle degrees for each complete cycle a linear interpolation can be used
for calculating cylinder pressure values corresponding to integer crank angle degrees i.e., from 1 to
720 degCA and for each complete engine cycle. The cylinder pressure piezoelectric transducer constant
in bar/Volt is then used for converting cylinder pressure measurements from Volt to bars. This way,
the cylinder pressure crank angle degree profile for each complete engine cycle is generated. The same
procedure is followed for tracking the injection pressure value corresponding to “hot TDC” and thus
to 180 degCA, and then to derive the injection pressure–crank angle degree profile for each complete
engine cycle. It should be noticed that in the case of fuel injection–crank angle degree profiles an
adjustment of the fuel injection pressure was made at each operating case for taking into account the
residual pressure of the fuel injection system through piezoelectric transducer drift.
Also, it should be mentioned that the cylinder pressure and injection pressure data analysis
is performed considering constant shaft speed, which is equal to the one derived from the actual
acquisition step. As known, the engine rotational speed varies during an engine operation cycle mainly
due to the inertia of the moving masses of the main kinematic mechanism. However, the effect of
Energies 2019, 12, 1547 14 of 36

reciprocating and rotating masses of “Lister LV1” engine on its angular velocity cycle variation is
partially counterbalanced by the flywheel. Also, “Lister LV1” is a small lightweight engine and thus,
high inertia forces are not expected. For this reason, the present “Lister LV1” engine performance
analysis is performed considering constant angular velocity i.e., engine speed during an engine
operation cycle.
At this point, it should be underlined that all three signals of TDC position, cylinder pressure,
and injection pressure are not received at exactly the same time instant from the fast acquisition card.
In other words, there is a time phase between the three recorded signals. Specifically, during an
acquisition step, all three signals are recorded. Hence, the time phase between two consecutive signals
is equal with 1/3 of the acquisition step e.g., if the acquisition step is set to 1 degCA, then at the first
1/3 of the degree the cylinder pressure signal is recorded, at the second 1/3 of the degree the injection
pressure signal is recorded, and at the final 1/3 of the degree the TDC position signal is recorded.
The values of cylinder pressure and injection pressure for each complete engine cycle are then used
to calculate the average cylinder pressure–crank angle and the average injection pressure–crank angle
profiles. The determination of the average cylinder pressure–crank angle and the average injection
pressure–crank angle profiles provides us with the opportunity to calculate the engine performance
parameters of the examined diesel engine such as indicated work and power, IMEP, ISFC, gross and
net heat release rate, ignition delay, and combustion duration. The derivation of individual cylinder
pressure and injection pressure profiles for all recorded complete engine cycles and their processing
for performance and heat release rate analysis has proven quite useful in engine cases where there is
cyclic variation on performance and combustion characteristics between consecutive engine cycles.
Having estimated the cylinder pressure–crank angle degree profile for each recorded complete
engine cycle and also for the average cycle, the indicated work per cylinder can be calculated and
through this the cylinder indicated power. Hence, the measured cylinder pressure–crank angle
degree profiles are transformed to cylinder pressure–instantaneous cylinder volume profiles and
then, the trapezoidal method is used for integrating each cylinder pressure–volume profile and for
calculating indicated power and IMEP [43,44]. This way, the covariance (COV) of the IMEP can be
estimated for each obtained complete cycle providing the opportunity to examine the effect of test fuels
on the cyclic variability of the cylinder pressure profile. According to Byttner et al. [45], the COV(IMEP)
is calculated as follows:
σ(IMEP)
COV(IMEP) = (5)
µ(IMEP)
where σ and µ are the standard deviation and the mean value of IMEP over all recorded complete
engine cycles. A similar mathematical relation with the aforementioned one was also used to calculate
the covariance of the mean injection pressure for assessing the effect of test fuels on the cyclic variation
of mean fuel line pressure.
The calculation of the heat release rate is of high importance in diesel engines since, from its
processing, very important information can be extracted, such as the initiation and the completion of
the combustion, the quality of combustion, the fuel burning rate, and the duration and the intensity of
the premixed and the diffusion-controlled in-cylinder combustion. Hence, the following mathematical
formula, which has been thoroughly analyzed by Heywood [46], is used to calculate the instantaneous
net heat release rate:

dQnet dQgross dQloss c dV cV dp


= − = (1 + V )p + V . (6)
dϕ dϕ dϕ R dϕ R dϕ
In the previous relation, the term dQgross /dϕ is called total or gross heat release rate, whereas
the difference dQnet /dϕ is called net heat release rate [46,47]. The term dQloss /dϕ corresponds to the
instantaneous in-cylinder gas heat transfer losses, which are transferred to the cylinder walls mainly
through heat convection and secondarily, after combustion initiation, mainly through additional heat
radiation due to flame development and combustion-released burning particles. In the present study,
Energies 2019, 12, 1547 15 of 36

the well-known Annand model [46,48] is used to calculate the instantaneous heat losses since this
model has already been used successfully for heat release rate calculations [34–36,41] and also in
experimental transfer studies [49]. It is worth mentioning that Annand model was calibrated in the
present analysis at each operating point using the following relation:
.
. Qgross,tot
mf = (7)
LHV
. .
where mf the measured fuel injected quantity per engine cycle, Qgross,tot is the total gross heat released
during an engine cycle, and LHV is the lower heating value of each examined fuel.
The accurate calculation of the combustion duration during an engine cycle is not easy and simple
since the precise determination of the end of combustion is quite difficult. This is due to the fact the
heat release rate curve indicates fluctuations towards the end of combustion and thus the precise end
of combustion is quite difficult to be spotted. For this reason, in many diesel combustion investigations,
various combustion durations are defined, which correspond to different proportions of fuel injected
mass such as CA5, CA25, CA50, and CA90 or CA95, which correspond to the combustion duration in
crank angle degrees of the 5%, 25%, 50%, and 90% or 95% of fuel injected mass per cycle, respectively.
The computational model can be used for experimental data processing of either four-stroke or
two-stroke diesel engines where measurements of cylinder pressure, injection pressure, and TDC
position (three sensors) were obtained, or it can be used for processing only measurements of cylinder
pressure and TDC position (two sensors). Also, it should be underlined that most of the input
data of the computational model can be found in any case of examined diesel engine from the
engine manufacturer manual, the manufacturers of piezoelectric transducers for cylinder pressure and
injection pressure transducers, and the fuel preparation refineries. Hence, the developed experimental
data processing model can provide reliable results for the following diesel engine performance and
combustion characteristics:

• Cylinder pressure–crank angle and injection pressure–crank angle profiles for each measured
complete engine cycle at a certain engine operating point;
• Average cylinder pressure–crank angle and injection pressure–crank angle profiles over all
measured complete engine cycles at a certain engine operating point;
• Indicated work and power;
• Indicated mean effective pressure (IMEP) per engine cycle and per average cylinder pressure
profile at a certain engine operating point;
• Indicated specific fuel consumption (ISFC);
• Engine mechanical efficiency;
• Brake-specific fuel consumption (BSFC);
• Actual acquisition step in crank angle degrees;
• Actual engine speed;
• Instantaneous gross and net heat release rates for each measured cycle cylinder pressure profile
and for the average cylinder pressure profile at a certain engine operating point;
• Instantaneous heat transfer loss rate for each measured cycle cylinder pressure profile and for the
average cylinder pressure profile at a certain engine operating point;
• In-cylinder gas temperature–crank angle profile for each measured cycle cylinder pressure profile
and for the average cylinder pressure profile at a certain engine operating point;
• Cumulative gross and net heat release rates for each measured cycle cylinder pressure profile and
for the average cylinder pressure profile at a certain engine operating point;
• Cumulative heat loss rate for each measured cycle cylinder pressure profile and for the average
cylinder pressure profile at a certain engine operating point;
Energies 2019, 12, 1547 16 of 36

• Instantaneous and cumulative fuel burning mass rates for the average cylinder pressure profile at
a certain engine operating point;
• Ignition angle and ignition delay for the average cylinder pressure profile at a certain engine
operating point;
• Peak cylinder pressure and corresponding crank angle for each measured cycle cylinder pressure
profile and for the average cylinder pressure profile at a certain engine operating point;
• Combustion durations of 5%, 25%, 50%, 90%, and 100% of the total injected fuel mass for each
measured cycle cylinder pressure profile and for the average cylinder pressure profile at a certain
engine operating point;
• Average injection pressure, dynamic injection timing (i.e., start of injection, SOI), peak injection
pressure, and fuel injection duration for the average cylinder pressure profile at a certain engine
operating point.

4. Results and Discussion

4.1. Effect of RME30 and GLY30 on HSDI Diesel Engine-Measured Cylinder Pressure and Injection Pressure
Cycle-to-Cycle Variation and Average Cycle Heat Release Rate Analysis Results
The oxygenated fuels RME30 and GLY30 have almost the same mass percentage (31 mass-%) of two
different oxygenated additives (RME and mixture of glymes). Hence, the evaluation of experimental
results for cylinder pressure, injection pressure, and heat release rates for these two fuels at 2500 rpm
and at 80% of full load facilitate the extraction of conclusions about their effect on diesel injection and
combustion mechanisms.
In Figure 7a–c, experimental results for fuel injection pressure profiles of RME30 and GLY30 of
the first (Figure 7a), the fifth (Figure 7b), and the ninth (Figure 7c) engine recorded cycle at 2500 rpm
and at 80% of full load are shown. All experimental results shown in Figure 7a–c are given for the
high-speed single-cylinder DI diesel engine “Lister LV1”. It was decided that the effect of RME30 and
GLY30 on the fuel injection pressure profiles of the first, the fifth, and the ninth recorded complete
engine cycle would be examined to assess the impact of these oxygenated fuels on the cycle variation
of the fuel line pressure, since the physical properties of these fuels may influence the behavior of the
fuel injection system.
From the observation of Figure 7a–c, it is concluded that the use of fuel RME30 in “Lister LV1”
engine resulted in an earlier initiation of fuel injection pressure rise and also in a steeper injection
pressure rise compared to fuel GLY30. These observations can be attributed to the lower compressibility
factor of RME30 compared to GLY30. Also as evidenced from Figure 7a–c, the individual impact of
fuels RME30 and GLY30 on fuel injection pressure profile do not vary noticeably from one engine cycle
to another, indicating that the fuel injection pressure system performs almost the same when using
either RM30 or GLY30.
In Figure 8a–c, experimental results for cylinder pressure profiles of RME30 and GLY30 of the first
(Figure 8a), the fifth (Figure 8b), and the ninth (Figure 8c) complete engine cycle at 2500 rpm and at 80%
of full load of “Lister LV1” engine are shown. We also decided to assess the effect of RME30 and GLY30
on the cylinder pressure profiles of the first, the fifth, and the ninth obtained complete engine cycles to
evaluate the impact of these oxygenated fuels on the cycle variation of the measured cylinder pressure.
From the observation of Figure 7a–c, it is concluded that the use of fuel RME30 in “Lister LV1”
engine resulted in an earlier initiation of fuel injection pressure rise and also in a steeper injection
pressure rise compared to fuel GLY30. These observations can be attributed to the lower
compressibility factor of RME30 compared to GLY30. Also as evidenced from Figure 7a–c, the
individual impact of fuels RME30 and GLY30 on fuel injection pressure profile do not vary
Energies 2019, 12, 1547from one engine cycle to another, indicating that the fuel injection pressure system 17 of 36
noticeably
performs almost the same when using either RM30 or GLY30.

INJECTION PRESSURE (bar)

INJECTION PRESSURE (bar)


Energies 2018, 11, x FOR PEER REVIEW 17 of 36
(a) (b)
In Figure 8a–c, experimental results for cylinder pressure profiles of RME30 and GLY30 of the
first (Figure 8a), the fifth (Figure 8b), and the ninth (Figure 8c) complete engine cycle at 2500 rpm
INJECTION PRESSURE (bar)

and at 80% of full load of “Lister LV1” engine are shown. We also decided to assess the effect of
RME30 and GLY30 on the cylinder pressure profiles of the first, the fifth, and the ninth obtained
complete engine cycles to evaluate the impact of these oxygenated fuels on the cycle variation of the
measured cylinder pressure.
From the examination of Figure 8a–c, it results that fuel RM30 indicates higher cylinder
pressures around TDC and slightly higher cylinder pressures during expansion stroke compared to
fuel GLY30. Hence, slightly higher amounts of combustion energy are released during the first
stages of combustion in the case of RME30 compared to GLY30. Also, slightly higher combustion
energy is released during diffusion-controlled combustion, thus leading to slightly higher
in-cylinder pressures during expansion stroke (c) for fuel RME30 compared to GLY30. These
observations can be possibly ascribed to the earlier injection pressure rise and, thus, earlier initiation
Figure 7.Figure 7. Comparison
Comparison of experimentalresults
of experimental results for
for fuel
fuelinjection
injectionpressure profile
pressure of (a) of
profile the(a)
first
theengine
first engine
of in-cylinder fuel injection for fuel RME30 compared to GLY30 led to the preparation of slightly
operationoperation
cycle, cycle,
(b) (b)fifth
the the fifth engine
engine operation cycle,
operation cycle, and
and (c) the
(c) ninth
the engine
ninth operation
engine cycle for
operation test for test
cycle
higher fuel quantity in the case of RME30. Also, according to Figure 8a–c, no noticeable variations
fuels RME30 and GLY30 at 2500 rpm and at 80% of full load. All experimental results are given for
fuels RME30
from and GLY30
onehigh-speed
cycle at 2500
to another rpm and
in cylinder at 80% profile
pressure of full differences
load. All experimental
between fuels results
RME30 areand
given
GLY30for the
the single-cylinder DI diesel engine “Lister LV1”.
high-speed single-cylinder
are evidenced, leading to DIthediesel engine
conclusion “Lister
that LV1”. oxygenated fuels do not cause significant
the specific
cyclic variation regarding cylinder pressure.
CYLINDER PRESSURE (bar)

CYLINDER PRESSURE (bar)

(a) (b)
CYLINDER PRESSURE (bar)

(c)
Figure
Figure 8. 8. Comparison
Comparison of experimentalresults
of experimental results for
for cylinder
cylinder pressure
pressureprofile of (a)of
profile the(a)
first
theengine
first engine
operation cycle, (b) the fifth engine operation cycle, and (c) the ninth engine operation cycle for test
operation cycle, (b) the fifth engine operation cycle, and (c) the ninth engine operation cycle for test
fuels RME30 and GLY30 at 2500 rpm and at 80% of full load. All experimental results are given for
fuels RME30 and GLY30 at 2500 rpm and at 80% of full load. All experimental results are given for the
the high-speed single-cylinder DI diesel engine “Lister LV1”.
high-speed single-cylinder DI diesel engine “Lister LV1”.
In Figure 9a, comparative experimental results for instantaneous gross heat release rate for test
fuels RME30 and GLY30 at 2500 rpm and at 80% of full load are given. Initial predictions for
instantaneous gross heat release rate were smoothed using an intrinsic smoothing spline function of
MATLAB [40] and the smoothed gross heat release rate results are given in Figure 9a. In addition, in
Energies 2018, 11, x FOR PEER REVIEW 18 of 36
Energies 2019, 12, 1547 18 of 36

Figure 9b, comparative results for the instantaneous in-cylinder heat loss rate for RME30 and GLY30
at 2500 rpm
From theand at 80% of of
examination full load 8a–c,
Figure are presented. Gross
it results that fuelheat
RM30 release and higher
indicates heat loss rate results
cylinder for
pressures
fuels RME30 and GLY30, provided in Figure 9a,b, were calculated
around TDC and slightly higher cylinder pressures during expansion stroke compared to fuel GLY30. using the corresponding
experimental
Hence, slightlyresults
higherof cylinder
amounts of pressure
combustion of the average
energy cycle from
are released allthe
during recorded onesofatcombustion
first stages 2500 rpm
and at 80% load.
in the case of RME30 compared to GLY30. Also, slightly higher combustion energy is released during
As observed from
diffusion-controlled Figure 9a,
combustion, both
thus fuels to
leading RME30
slightlyand GLY30
higher indicatepressures
in-cylinder almost the same
during ignition
expansion
angles, thus
stroke for meaning
fuel RME30combustion
compared toisGLY30.
initiatedThesefor both fuels at almost
observations can bethe same angle.
possibly Also,
ascribed to according
the earlier
to Figure 9a, the intensity of premixed and diffusion-controlled combustion phases
injection pressure rise and, thus, earlier initiation of in-cylinder fuel injection for fuel RME30 compared are relatively
higher for test fuel RME30 compared to fuel GLY30. This observation, as already
to GLY30 led to the preparation of slightly higher fuel quantity in the case of RME30. Also, according explained, can be
to
attributed
Figure 8a–c, to no
thenoticeable
earlier initiation
variationsof injection
from oneprocesscycle tofor test fuel
another in RME30
cylindercompared to GLY30,
pressure profile which
differences
results
between infuels
the premixed
RME30 and combustion
GLY30 areof a relatively
evidenced, highertofuel
leading the mass. The spontaneous
conclusion that the specificcombustion
oxygenated of
afuels
relatively
do nothigher fuel mass incyclic
cause significant the case of RME30
variation with slightly
regarding cylinderhigher heat of combustion compared
pressure.
to GLY30 results in higher premixed combustion-released
In Figure 9a, comparative experimental results for instantaneous energy and thus,
grossmore intense
heat release premixed
rate for
combustion. Also, the use of RME30 results in slightly higher amounts
test fuels RME30 and GLY30 at 2500 rpm and at 80% of full load are given. Initial predictions for of combustion-released
energy during gross
instantaneous diffusion-controlled
heat release rate combustion
were smoothed compared
usingtoanthe use of GLY30.
intrinsic smoothing spline function of
MATLAB [40] and the smoothed gross heat release rate results arethe
From the observation of Figure 9b, it can be concluded that fuelinRME30
given Figureindicates higher
9a. In addition,
instantaneous heat loss rates during all stages of combustion (i.e.,
in Figure 9b, comparative results for the instantaneous in-cylinder heat loss rate for RME30 and GLY30 both premixed and
diffusion-controlled
at 2500 rpm and at 80% combustion)
of full loadaround TDC and
are presented. during
Gross expansion
heat release and stroke
heat loss compared
rate resultsto for
GLY30.
fuels
This observation can be attributed to the higher in-cylinder bulk gas temperatures
RME30 and GLY30, provided in Figure 9a,b, were calculated using the corresponding experimental experienced for
test fuelofRME30
results cylindercompared
pressure of to the
GLY30
averageduringcycle combustion as a result
from all recorded onesofat the
2500aforementioned
rpm and at 80%higher load.
in-cylinder pressures.
HEAT RELEASE RATE (J/deg)
INSTANTANEOUS GROSS

HEAT LOSS RATE (J/deg)


INSTANTANEOUS

(a) (b)
Comparison of
Figure 9. Comparison of experimental
experimental results
results for
for (a)
(a) instantaneous
instantaneous in-cylinder
in-cylinder gross
gross heat
heat release
release rate
and (b) instantaneous
and instantaneous in-cylinder
in-cylinder heal
heal loss
loss rate
rate for
for test
test fuels
fuels RME30
RME30 and
and GLY30
GLY30 at
at 2500
2500 rpm
rpm and
and at
80% of full load. All experimental results are given for the high-speed single-cylinder DI diesel
of full load. All experimental results are given for the high-speed single-cylinder DI diesel engine
“Lister
engine LV1”.
“Lister LV1”.

As Figure
In observed 10a,from Figure 9a,experimental
comparative both fuels RME30 resultsandforGLY30 indicate almost
the cumulative gross the
heatsame ignition
release rate
angles, thus
between meaning
fuels RME30combustion
and GLY30isare initiated
given. for both
Also, in fuels
Figure at almost the same angle.
10b, comparative Also, according
experimental results
to Figure
for 9a, the intensity
the cumulative net heat of premixed
release and diffusion-controlled
rate between fuels RME30 and GLY30 combustion phases are
are presented, andrelatively
finally,
higher
in for10c,
Figure testcomparative
fuel RME30results
compared to fuel
for the GLY30. heat
cumulative Thisloss
observation,
rate betweenas already explained,
fuels RME30 can be
and GLY10
are given. All results shown in Figure 10a–c correspond to “Lister LV1” engine operation atwhich
attributed to the earlier initiation of injection process for test fuel RME30 compared to GLY30, 2500
results
rpm in the
and premixed
at 80% of fullcombustion
load. From of atherelatively higheroffuel
observation mass.10a,b,
Figure The spontaneous combustion
it results that fuel RME30 of a
relativelyhigher
indicates highervalues
fuel mass in the case gross
of cumulative of RME30
and netwith slightly
heat releasehigher heat of combustion
rate compared compared
to fuel GLY30 after
to GLY30 results in higher premixed combustion-released energy and thus,
TDC during expansion stroke. This observation follows the observations previously made for the more intense premixed
combustion.inAlso,
differences the use of RME30
instantaneous gross andresults
netinheat
slightly higher
release amounts
rates between of combustion-released
fuels RME30 and GLY30. energy
during diffusion-controlled
Finally, the higher instantaneous combustion
heat loss compared to the use
rate observed of GLY30.
in Figure 9a for test fuel RME30 compared
From the observation of Figure 9b, it can be
to fuel GLY30 also results in higher cumulative heat loss rate for RME30 concluded that the fuel RME30
compared indicates
to GLY30 after
higherasinstantaneous
TDC, evidenced from heat loss10a.
Figure rates during all stages of combustion (i.e., both premixed and
diffusion-controlled
Hence, from the combustion)
observationaround of theTDC andofduring
effects RME30 expansion
and GLY30 strokeoncompared to GLY30.
fuel injection and
in-cylinder combustion characteristics, it can be concluded that the combustion mechanism for
This observation can be attributed to the higher in-cylinder bulk gas temperatures experienced is
Energies 2019, 12, 1547 19 of 36

test fuel RME30 compared to GLY30 during combustion as a result of the aforementioned higher
in-cylinder pressures.
In Figure 10a, comparative experimental results for the cumulative gross heat release rate
between fuels RME30 and GLY30 are given. Also, in Figure 10b, comparative experimental results
for the cumulative net heat release rate between fuels RME30 and GLY30 are presented, and finally,
in Figure 10c, comparative results for the cumulative heat loss rate between fuels RME30 and GLY10
are given. All results shown in Figure 10a–c correspond to “Lister LV1” engine operation at 2500 rpm
Energies 2018, 11, x FOR PEER REVIEW 19 of 36
and at 80% of full load. From the observation of Figure 10a,b, it results that fuel RME30 indicates
higher values of
primarily cumulative
affected by thegross
fueland net heat
injection releasewhich,
process rate compared to fuel is
as evidenced, GLY30 after TDC
controlled during
by oxygenated
expansion stroke. This observation follows the observations previously made for
fuels physical properties and mainly by the lower compressibility factor of RME30 compared the differences in to
instantaneous
GLY30. This gross and net heat
observation release
should be rates
kept between
in mind fuels RME30with
in contrast and GLY30.
the higherFinally, the higher
oxygen content of
instantaneous heat loss rate observed in Figure 9a for test fuel RME30 compared
GLY30 compared to RME30, which it would be expected to enhance local fuel combustion to fuel GLY30 also
rates.
results in higher cumulative heat loss rate for RME30 compared to GLY30 after TDC,
Hence, the lower compressibility factor of RME30 appears to overwhelm the higher oxygen content as evidenced
fromofFigure
GLY3010a. regarding their impact on in-cylinder combustion mechanism and its characteristics.

1000 1000
"LISTER LV1" ENGINE
800 800 2500 RPM - 80% LOAD
GLY30
RME30
600 600
"LISTER LV1" ENGINE
2500 RPM - 80% LOAD
400 400
GLY30
RME30
200 200

0 0

-200 -200
160 180 200 220 240 260 280 160 180 200 220 240 260 280
CRANK ANGLE (deg) CRANK ANGLE (deg)
(a) (b)

(c)
Figure 10. Comparison
Figure of experimental
10. Comparison results
of experimental for (a)
results forin-cylinder cumulative
(a) in-cylinder gross
cumulative heatheat
gross release rate,rate,
release
(b) in-cylinder cumulative
(b) in-cylinder net heat
cumulative netrelease rate, and
heat release rate,(c)and
the (c)
in-cylinder cumulative
the in-cylinder heal lossheal
cumulative rateloss
for test
rate for
fuelstest
RME30 and GLY30 at 2500 rpm and at 80% of full load. All experimental results are given
fuels RME30 and GLY30 at 2500 rpm and at 80% of full load. All experimental results are given for the
high-speed single-cylinder
for the high-speed DI diesel engine
single-cylinder “Lister
DI diesel LV1”.
engine “Lister LV1”.

Hence,
4.2. Effectfrom the observation
of GLY10 and GLY30of onthe effects
HSDI of RME30
Diesel and GLY30
Engine-Measured on fuelPressure
Cylinder injectionand
and in-cylinder
Injection Pressure
combustion characteristics,
Cycle-to-Cycle it can
Variation and be concluded
Average that
Cycle Heat the combustion
Release mechanism
Rate Analysis Results is primarily affected
by the fuel injection process which, as evidenced, is controlled by oxygenated fuels physical properties
and mainlyTheby oxygenated
the lower fuels GLY10 and
compressibility GLY30
factor have almost
of RME30 the same
compared oxygenated
to GLY30. additive type
This observation
(glymes)
should be keptat different
in mind inproportions 10.2the
contrast with mass-%
higher and 31.3 mass-%,
oxygen content respectively. The different
of GLY30 compared blending
to RME30,
which it would be expected to enhance local fuel combustion rates. Hence, the lower compressibility(3%
ratios of oxygenated agents result in different fuel oxygen contents between fuels GLY10
oxygen) and GLY30 (9% oxygen). Hence, the assessment of experimental results for cylinder
pressure, injection pressure, and heat release rates for these two fuels at 2500 rpm and at 80% of full
load will facilitate the extraction of conclusions about their effect on diesel injection and combustion
mechanisms. Initially, we decided to examine the effect of GLY10 and GLY30 on the cyclic variation
of measured fuel injection and cylinder pressure profiles for assessing the impact of these two
oxygenated additives on the behavior of the fuel injection system and on engine operation and
Energies 2019, 12, 1547 20 of 36

factor of RME30 appears to overwhelm the higher oxygen content of GLY30 regarding their impact on
in-cylinder combustion mechanism and its characteristics.

4.2. Effect of GLY10 and GLY30 on HSDI Diesel Engine-Measured Cylinder Pressure and Injection Pressure
Cycle-to-Cycle Variation and Average Cycle Heat Release Rate Analysis Results
The oxygenated fuels GLY10 and GLY30 have almost the same oxygenated additive type (glymes)
at different proportions 10.2 mass-% and 31.3 mass-%, respectively. The different blending ratios of
oxygenated agents result in different fuel oxygen contents between fuels GLY10 (3% oxygen) and GLY30
(9% oxygen). Hence, the assessment of experimental results for cylinder pressure, injection pressure,
and heat release rates for these two fuels at 2500 rpm and at 80% of full load will facilitate the extraction
of conclusions about their effect on diesel injection and combustion mechanisms. Initially, we decided
to examine the effect of GLY10 and GLY30 on the cyclic variation of measured fuel injection and cylinder
pressure profiles for assessing the impact of these two oxygenated additives on the behavior of the
fuel injection system and on engine operation and performance. In Figure 11a–c, experimental results
for fuel injection pressure profiles of GLY10 and GLY30 of the first (Figure 11a), the fifth (Figure 11b),
and the ninth (Figure 11c) engine cycle at 2500 rpm and at 80% of full load are shown. All experimental
results shown in Figure 11a–c are given for the high-speed single-cylinder DI diesel engine “Lister
LV1”.Energies
From2018,
the11,
observation of Figure 11a–c, it is concluded that the use of fuel GLY10 in “Lister
x FOR PEER REVIEW 20 of 36 LV1”
engine results in slightly earlier initiation and steeper injection pressure rise compared to fuel GLY30
at all and steeper
cycles injectionThis
examined. pressure rise compared
observation to fuel
can also beGLY30 at all
ascribed tocycles examined.
the lower This observation
compressibility factor of
can also be ascribed to the lower compressibility factor of fuel GLY10
fuel GLY10 compared to GLY30. In fact, since the difference in fuel compressibilitycompared to GLY30. In fact,
factor between
since the difference in fuel compressibility factor between fuels GLY10 and GLY30 is lower
fuels GLY10 and GLY30 is lower compared to the corresponding difference between fuels RME30 and
compared to the corresponding difference between fuels RME30 and GLY30, the pertinent effects on
GLY30, the pertinent effects on fuel injection pressure are more pronounced in the case of fuel pair
fuel injection pressure are more pronounced in the case of fuel pair RME30-GLY30 compared to the
RME30-GLY30 compared to the fuel pair GLY10-GLY30.
fuel pair GLY10-GLY30.
INJECTION PRESSURE (bar)

INJECTION PRESSURE (bar)

(a) (b)
INJECTION PRESSURE (bar)

(c)
Figure 11. Comparison
Figure 11. Comparison of experimental
of experimentalresults
results for fuelinjection
for fuel injectionpressure
pressure profile
profile of the
of (a) (a) first
the first
engineengine
operation
operation cycle,
cycle, (b) (b)
thethe fifth
fifth engineoperation
engine operation cycle,
cycle, and
and(c) (c)the
theninth
ninthengine
engineoperation
operationcyclecycle
for test
for test
fuels fuels
GLY10 GLY10 and GLY30
and GLY30 at 2500
at 2500 rpmrpmandand at 80%
at 80% of of
fullfull load.All
load. Allexperimental
experimental results
results are
aregiven
givenfor for the
the high-speed single-cylinder DI diesel engine
high-speed single-cylinder DI diesel engine “Lister LV1”. “Lister LV1”.

In Figure 12a–c, experimental results for cylinder pressure profiles of GLY10 and GLY30 of the
first (Figure 12a), the fifth (Figure 12b), and the ninth (Figure 12c) complete engine cycle at 2500 rpm
and at 80% of full load of “Lister LV1” engine are shown. From the examination of Figures 12a–c, it
results that fuel GLY10 indicates slightly higher cylinder pressures around TDC and during
expansion stroke compared to fuel GLY30. Hence, though both fuels appear to ignite almost
Energies 2019, 12, 1547 21 of 36

In Figure 12a–c, experimental results for cylinder pressure profiles of GLY10 and GLY30 of the first
(Figure 12a), the fifth (Figure 12b), and the ninth (Figure 12c) complete engine cycle at 2500 rpm and at
80% of full load of “Lister LV1” engine are shown. From the examination of Figure 12a–c, it results
that fuel GLY10 indicates slightly higher cylinder pressures around TDC and during expansion stroke
compared to fuel GLY30. Hence, though both fuels appear to ignite almost simultaneously, higher
amounts of combustion energy are released during combustion, thus leading to higher in-cylinder
pressures during expansion stroke for fuel GLY10 compared to GLY30. This difference in measured
cylinder pressure between fuels GLY10 and GLY30 is associated with the differences evidenced in fuel
injection pressure between these two fuels due to lower compressibility factor of fuel GLY10 compared
to fuel GLY30. Differences in measured cylinder pressure are less pronounced in the case of fuel pair
GLY10–GLY30 compared to fuel pair RME30–GLY30. Also, no noticeable cyclic variation is observed
when the “Lister LV1” engine burned either GLY10 or GLY30.
Energies 2018, 11, x FOR PEER REVIEW 21 of 36
CYLINDER PRESSURE (bar)

CYLINDER PRESSURE (bar)

(a) (b)
CYLINDER PRESSURE (bar)

(c)
Figure
Figure 12. 12. Comparison
Comparison of experimentalresults
of experimental results for
for fuel
fuelinjection
injection pressure profile
pressure of (a) of
profile the(a)
first
theengine
first engine
operation
operation cycle, cycle,
(b) the(b)fifth
the fifth engine
engine operationcycle,
operation cycle, and
and(c)(c)the ninth
the engine
ninth operation
engine cycle for
operation test for test
cycle
fuels GLY10 and GLY30 at 2500 rpm and at 80% of full load. All experimental results are given for
fuels GLY10 and GLY30 at 2500 rpm and at 80% of full load. All experimental results are given for the
the high-speed single-cylinder DI diesel engine “Lister LV1”.
high-speed single-cylinder DI diesel engine “Lister LV1”.
In Figure 13a, comparative experimental results for instantaneous gross heat release rate for test
In Figure 13a, and
fuels GLY10 comparative experimental
GLY30 at 2500 rpm and at 80%results forload
of full instantaneous grossshown
are given. Results heat in
release
Figurerate
12a for test
fuels GLY10 and GLY30
are smoothed usingatan
2500 rpm and
intrinsic MATLABat 80% of full [40].
function loadInare given. in
addition, Results
Figureshown in Figure 12a are
13b, comparative
smoothed results
usingfor the
an instantaneous
intrinsic MATLABin-cylinder heat loss
function rate In
[40]. for addition,
GLY10 andin GLY30
Figureat 2500
13b,rpm and at 80% results
comparative
of full load are presented.
for the instantaneous in-cylinderAs observed
heat lossfrom
rateFigure 13a,b, both
for GLY10 andfuels
GLY30GLY10 and GLY30
at 2500 rpm andindicate no of full
at 80%
serious differences in ignition angle, thus meaning combustion is initiated for both fuels at almost
load are presented. As observed from Figure 13a,b, both fuels GLY10 and GLY30 indicate no serious
the same angle. Also, according to Figure 13a,b, the intensity of premixed and diffusion-controlled
differences in ignition angle, thus meaning combustion is initiated for both fuels at almost the same
combustion phases are relatively higher for test fuel GLY10 compared to fuel GLY30. This
angle. Also, according
observation, to Figure
as already 13a,b,can
explained, thebeintensity
ascribed toofthepremixed and diffusion-controlled
earlier initiation of injection process forcombustion
test
phases are
fuel relatively higher for
GLY10 compared to test fuel The
GLY30. GLY10 compared
slightly higher to fuel GLY30.
cylinder pressuresThisandobservation,
correspondingas already
in-cylinder
explained, can be bulk gas temperatures
ascribed to the earlier experienced
initiationfor fuel GLY10process
of injection comparedfor to fuel
test GLY30
fuel GLY10 results in
compared to
slightly higher instantaneous heat loss rates for fuel GLY10 relative to GLY30
GLY30. The slightly higher cylinder pressures and corresponding in-cylinder bulk gas temperatures as witnessed from the
observation of Figure 13b.
experienced for fuel GLY10 compared to fuel GLY30 results in slightly higher instantaneous heat loss
In Figure 14a–c, comparative results for the cumulative gross heat release rate (Figure 14a),
rates forcumulative
fuel GLY10 netrelative to GLY30
heat release as witnessed
rate (Figure 14b), and from the observation
cumulative heat loss rateof(Figure
Figure14c)
13b.between
fuels GLY10 and GLY30 at 2500 rpm and at 80% of full load are given. From the observation of
Figure 14a,b, it results that fuel GLY10 demonstrates slightly higher values of cumulative gross and
net heat release rate compared to fuel GLY30 after TDC during expansion stroke. These observations
are in accordance with the previously made observations for the differences in instantaneous gross
and net heat release rates between fuels GLY10 and GLY30. Finally, the slightly higher instantaneous
heat loss rate observed in Figure 13b for test fuel GLY10 compared to fuel GLY30 also results in
Energies 2018, 11, x FOR PEER REVIEW 22 of 36

Consequently, the use of GLY30 results in a delay of fuel injection process and, thus, results in
lower cylinder pressures and in-cylinder gas temperatures after TDC during expansion stroke
compared to fuels RME30 and GLY10. Hence, RME30 is the most influential oxygenated fuel on fuel
Energies 2019, 12, 1547 22 of 36
injection pressure and injection process and on in-cylinder combustion mechanism, followed by
GLY10 and GLY30.

HEAT LOSS RATE (J/deg)


INSTANTANEOUS
Energies 2018, 11, x FOR PEER REVIEW 22 of 36

Consequently, the use of GLY30 results in a delay of fuel injection process and, thus, results in
lower cylinder pressures and in-cylinder gas temperatures after TDC during expansion stroke
compared to fuels RME30 and GLY10. Hence, RME30 is the most influential oxygenated fuel on fuel
injection pressure and injection process and on in-cylinder combustion mechanism, followed by
(a) (b)
GLY10 and GLY30.
Figure
Figure 13. 13. Comparison
Comparison of experimental
of experimental results
results forfor
(a)(a) instantaneous in-cylinder
instantaneous in-cylinder gross
gross heat release
heat release rate
rate and (b) instantaneous in-cylinder heal loss rate for test fuels GLY10 and GLY30
and (b) instantaneous in-cylinder heal loss rate for test fuels GLY10 and GLY30 at 2500 rpm at 2500 rpm and and at

HEAT LOSS RATE (J/deg)


at 80% of full load. All experimental results are given for the high-speed single-cylinder DI diesel
80% of full load. All experimental results are given for the high-speed single-cylinder DI diesel engine

INSTANTANEOUS
engine “Lister LV1”.
“Lister LV1”.
1000 1000
"LISTER LV1" ENGINE
In Figure
800
14a–c, comparative results for the cumulative 800
gross heat release
2500 RPM rate (Figure 14a),
- 80% LOAD

cumulative net heat release rate (Figure 14b), and cumulative heat loss rate GLY30 (Figure
GLY10 14c) between fuels
600 600
GLY10 and GLY30 at 2500 rpm"LISTER and atLV1"80% of full load are given. From the observation of Figure 14a,b,
ENGINE
400 2500 RPM - 80% LOAD 400
it results that fuel GLY10 demonstrates GLY30 slightly higher values of cumulative gross and net heat release
GLY10
rate compared 200 to fuel GLY30 after TDC during expansion 200 stroke. These observations are in accordance
(a)
with the previously
0 made observations for the differences 0 in instantaneous(b) gross and net heat release
rates between fuelsFigure 13. Comparison
GLY10 and GLY30. of experimental results
Finally, the for (a)higher
slightly instantaneous in-cylinder gross
instantaneous heat heat
loss release
rate observed
-200 -200
rate and (b) instantaneous in-cylinder heal loss rate for test fuels GLY10 and GLY30 at 2500 rpm and
in Figure 13b 160 for test
180 fuel
200 GLY10
220 compared
240 260 to fuel GLY30
280 160 also
180 results
200 220in slightly
240 260 higher
280 cumulative
at 80% of CRANK ANGLE
full load. All (deg)
experimental results are given for theCRANK ANGLEsingle-cylinder
high-speed (deg) DI diesel
heat loss rate for GLY10 compared
engine “Lister(a)
LV1”.
to GLY30 after TDC, as evidenced (b) from Figure 14c.

1000 1000
CUMULATIVE HEAT LOSS RATE (J)

"LISTER LV1" ENGINE


800 800 2500 RPM - 80% LOAD
GLY30
GLY10
600 600
"LISTER LV1" ENGINE
400 2500 RPM - 80% LOAD 400
GLY30
GLY10
200 200

0 0

-200 -200
160 180 200 220 240 260 280 160 180 200 220 240 260 280
CRANK ANGLE (deg) CRANK ANGLE (deg)
(c)
(a) (b)
Figure 14. Comparison of experimental results for (a) in-cylinder cumulative gross heat release rate,
CUMULATIVE HEAT LOSS RATE (J)

(b) in-cylinder cumulative net heat release rate, and (c) the in-cylinder cumulative heal loss rate for
test fuels GLY10 and GLY30 at 2500 rpm and at 80% of full load. All experimental results are given
for the high-speed single-cylinder DI diesel engine “Lister LV1”.

4.3. Effect of Test Fuels DI1, RME30, GLY10, and GLY30 on the Cyclic Variation of Mean Injection Pressure
and Indicated Mean Effective Pressure (IMEP)

(c)
Figure 14. Comparison
Figure 14. Comparison of experimental
of experimental results
results for in-cylinder
for (a) (a) in-cylindercumulative
cumulative gross
grossheat
heatrelease rate,rate,
release
(b) in-cylinder cumulative net heat release rate, and (c) the in-cylinder cumulative heal
(b) in-cylinder cumulative net heat release rate, and (c) the in-cylinder cumulative heal loss rate forloss rate for test
test fuels GLY10 and GLY30 at 2500 rpm and at 80% of full load. All experimental results are given
fuels GLY10 and GLY30 at 2500 rpm and at 80% of full load. All experimental results are given for the
for the high-speed single-cylinder DI diesel engine “Lister LV1”.
high-speed single-cylinder DI diesel engine “Lister LV1”.
4.3. Effect of Test Fuels DI1, RME30, GLY10, and GLY30 on the Cyclic Variation of Mean Injection Pressure
and Indicated Mean Effective Pressure (IMEP)
Energies 2019, 12, 1547 23 of 36

Consequently, the use of GLY30 results in a delay of fuel injection process and, thus, results
in lower cylinder pressures and in-cylinder gas temperatures after TDC during expansion stroke
compared to fuels RME30 and GLY10. Hence, RME30 is the most influential oxygenated fuel on fuel
injection pressure and injection process and on in-cylinder combustion mechanism, followed by GLY10
and GLY30.

4.3. Effect of Test Fuels DI1, RME30, GLY10, and GLY30 on the Cyclic Variation of Mean Injection Pressure
Energies
and 2018, 11,Mean
Indicated x FOREffective
PEER REVIEW
Pressure (IMEP) 23 of 36

A statistical analysis has also been performed to examine examine the effect of test fuels DI1, RME30,
GLY10, and GLY30 GLY30 on on the
thecyclic
cyclicvariation
variationofofmeasured
measuredmean mean injection
injection pressure
pressure andand IMEP.
IMEP. Hence,
Hence, in
in Figure
Figure 15,15, results
results forfor
thethe covarianceofofmeasured
covariance measuredmean meaninjection
injectionpressure
pressureandandIMEP
IMEPof oftest
test fuels
fuels DI1,
DI1,
RME30, GLY10, and GLY30 GLY30 at at 2500
2500 rpm
rpm and and atat 80%
80% of of full
full load
load areare given.
given. These results have been
derived from the statistical analysis of 100 consecutive
derived from the statistical analysis of 100 consecutive cycles obtained cycles obtained fromfrom
“Lister LV1” LV1”
“Lister engineengine
using
the aforementioned
using the aforementioned test fuels.
test As evidenced
fuels. As evidencedfrom Figure 15, the15,
from Figure covariance of IMEP
the covariance of is similar
IMEP for all
is similar
examined fuels and, in all cases, does not exceed 3%, thus demonstrating
for all examined fuels and, in all cases, does not exceed 3%, thus demonstrating the addition of RME the addition of RME and
glycol ethers
and glycol to base
ethers fuel DI1
to base fuel did
DI1not
didseriously
not seriouslyaffectaffect
the operation
the operation of theofexamined
the examined“Lister LV1” LV1”
“Lister from
one
fromcycle to another.
one cycle A similar
to another. behavior
A similar for the
behavior foreffect of theofexamined
the effect the examined test fuels on COV(IMEP)
test fuels on COV(IMEP) was
also evidenced at 20%, 40%, and 60% of full load. Hence, the addition of
was also evidenced at 20%, 40%, and 60% of full load. Hence, the addition of RME and glycol ethers RME and glycol ethers to base
fuel DI1fuel
to base did DI1
not affect
did notnoticeably “Lister LV1”
affect noticeably engine
“Lister LV1”steady-state operation. operation.
engine steady-state Moreover, Moreover,
as evidenced as
from Figure
evidenced 15, Figure
from the addition
15, the of RME to
addition offuel
RMEDI1 resulted
to fuel in reduction
DI1 resulted of the covariance
in reduction of mean
of the covariance of
injection pressure, whereas the addition of glymes in the case of
mean injection pressure, whereas the addition of glymes in the case of fuels GLY10 and GLY30fuels GLY10 and GLY30 resulted in
the increase
resulted of increase
in the the covariance of mean injection
of the covariance of mean pressure
injectioncompared
pressure to conventional
compared diesel operation.
to conventional diesel
Hence, fuel injection system operation indicated slightly lower cyclic
operation. Hence, fuel injection system operation indicated slightly lower cyclic variation with thevariation with the addition of
RME compared
addition of RMEtocompared
conventional diesel operation,
to conventional dieselwhereas
operation, the whereas
additionthe of glymes
addition toof
base fuel DI1
glymes led
to base
to a more
fuel DI1 led unsteady
to a more cyclic behavior
unsteady of the
cyclic fuel injection
behavior system,
of the fuel which,system,
injection however, did not
which, exceeddid
however, the not
5%
covariance, which is within the limits of steady-state operation.
exceed the 5% covariance, which is within the limits of steady-state operation.

7
"LISTER LV1" ENGINE
2500 RPM - 80% LOAD
6
MEAN INJECTION PRESSURE (bar)
IMEP (bar)
5

0
DI1 RME30 GLY10 GLY30
Figure 15. Effect of test fuels DI1, RME30, GLY10, and GLY30 on on the
the covariance
covariance of measured
measured mean
injection pressure and indicated mean effective pressure
pressure (IMEP). Results are given for the high-speed
DI diesel engine “Lister LV1”
LV1” at
at 2500
2500 rpm
rpm and
and at
at 80%
80%of
offull
fullload.
load.

4.4. Effect of Test Fuels DI1, RME30, GLY10, and GLY30 on Examined HSDI Diesel Engine Main Performance
4.4. Effect of Test Fuels DI1, RME30, GLY10, and GLY30 on Examined HSDI Diesel Engine Main
Parameters and Combustion Characteristics
Performance Parameters and Combustion Characteristics
In this section, the effect of conventional fuel DI1 and oxygenated fuels RME30, GLY10, and GLY30
In this section, the effect of conventional fuel DI1 and oxygenated fuels RME30, GLY10, and
on measured in-cylinder pressure and injection pressure and on instantaneous net heat release rate
GLY30 on measured in-cylinder pressure and injection pressure and on instantaneous net heat
at 2500 rpm and at 80% of full load is analyzed. We also examined the influence of oxygenated
release rate at 2500 rpm and at 80% of full load is analyzed. We also examined the influence of
fuels RME30, GLY10, and GLY30 on the percentage change of the indicated power, ISFC, injection
oxygenated fuels RME30, GLY10, and GLY30 on the percentage change of the indicated power,
duration, SOI, ignition angle, ignition delay, and combustion durations of 5% (CA5), 25% (CA25), 50%
ISFC, injection duration, SOI, ignition angle, ignition delay, and combustion durations of 5% (CA5),
(CA50), and 90% (CA90) of fuel injected mass per cycle. The percentage change of each one of the
25% (CA25), 50% (CA50), and 90% (CA90) of fuel injected mass per cycle. The percentage change of
each one of the aforementioned performance and combustion characteristics for each oxygenated
fuel is expressed with reference to the pertinent value of “base” fuel DI1. The main objective of this
investigation is to examine the relative deviation of the main performance and combustion
parameters in the case of oxygenated fuels combustion compared to conventional “base” fuel
Energies 2019, 12, 1547 24 of 36

aforementioned performance and combustion characteristics for each oxygenated fuel is expressed
with reference to the pertinent value of “base” fuel DI1. The main objective of this investigation is to
examine the relative deviation of the main performance and combustion parameters in the case of
oxygenated fuels combustion compared to conventional “base” fuel combustion. Another objective of
this particular examination is the evaluation, on a comparative basis, of the individual effect of each
one of the three oxygenated fuels examined in this study not only compared to “base” fuel DI1, but also
among them. In other words, this investigation will facilitate the assessment of the effect of increasing
fuel oxygen content for the same oxygenated additive by examining engine performance results for test
fuels GLY10 (3% oxygen) and GLY30 (9% oxygen). Also, this investigation will facilitate the evaluation
of the influence a biofuel additive (RME) in contrast with a synthetic oxygenated additive (Glymes) by
comparing performance results of fuels RME30 and GLY30, which have almost the same blending
ratio (31mass-%) of two different oxygenated agents. Overall, this particular analysis will reveal not
only the operational behavior of oxygenated fuels compared to conventional diesel oil, but also will
support the selection of the best oxygenated fuel in terms of engine performance and combustion
characteristics among the three examined in this study.
In Figure 16a, measured injection pressure profiles for test fuels DI1, RME30, GLY10, and GLY30 at
2500 rpm and at 80% of full load are compared. Results in Figure 16a are given for the average engine
cycle. In Figure 16b, the relative changes of injection duration for oxygenated blends RME30, GLY10,
and GLY30 with reference to the corresponding value of conventional diesel oil DI1 at 2500 rpm and at
80% of full load are shown. In Figure 16c, the relative changes of dynamic injection timing (i.e., start of
injection, SOI) for oxygenated blends RME30, GLY10, and GLY30 with reference to the corresponding
value of conventional diesel oil DI1 at 2500 rpm and at 80% of full load are presented. As evidenced from
Figure 16a, with reference to crank angle when the injector needle is lifted (i.e., 180 bar), all oxygenated
fuels indicate a delay in injection process initiation inside the cylinder compared to fuel DI1. The
delay in injection process is less pronounced in the case of oxygenated fuel RME30, in which it has
almost the same SOI with “base” fuel DI1, and is more pronounced in the case of oxygenated fuel
GLY30. Conventional fuel DI1 and oxygenated fuels RME30 and GLY10 indicate almost the same end
of fuel injection process as evidenced from Figure 16a, whereas the fuel injection process completion
for oxygenated fuel GLY30 is delayed compared to all other fuels. The small delay in injection process
initiation (i.e., small retardation of injection timing) observed in the case of oxygenated fuel GLY10
and especially in the case of oxygenated fuel GLY30 is associated with the progressive increase of
fuel compressibility factor observed in the cases of GLY10 and GLY30 compared to conventional fuel
DI1. The injection process appears to close almost at the same angle in the case of fuels DI1, RME30,
and GLY10, revealing a reduction of injection duration mainly due to similar values of fuel viscosity in
the case of fuel pair DI1-RME30 and due to lower fuel viscosity in the case of fuel GLY10 compared to
fuel DI1. In the case of oxygenated fuel GLY30, the injection process end is slightly delayed compared
to all other fuels. However, the difference between injection process end and injection process initiation
in the case of oxygenated fuel GLY30 is lower compared to fuels DI1 and RME30, thus leading to a
reduction of injection duration mainly due to significantly lower viscosity of fuel GLY30 compared to
the viscosities of fuels DI1 and RME30. Variations in fuel injection process initiation and completion are
reflected in the relative changes of injection duration of oxygenated fuels RME30, GLY10, and GLY30
compared to conventional fuel DI1 shown in Figure 16b. As evidenced from Figure 16b, the highest
reductions of injection duration are evidenced in the case of oxygenated fuels GLY10 and GLY30
compared to conventional fuel DI1 reaching up 5.5%. The relative reduction of injection duration in
the case of biofuel-added blend RME30 is limited compared to conventional oil DI1 not exceeding
1.5%. Also, according to Figure 16c, the progressive transition from diesel oil DI1 to RME30 and then
to fuels GLY10 and GLY30 results in a relative reduction of SOI (i.e., distance in crank angles from
TDC) and thus to a progressive delay in fuel injection process initiation inside the engine cylinder. The
effect on SOI reduction is more pronounced in the case of oxygenated fuel GLY30 with the highest
Energies 2019, 12, 1547 25 of 36

compressibility factor compared to diesel oil DI1 and is less pronounced in the case of oxygenated
fuel RME30.
Energies 2018, 11, x FOR PEER REVIEW 25 of 36

INJECTION PRESSURE (bar)

(a)
0

-1

-2

-3

-4

-5

-6
"LISTER LV1" ENGINE
-7 2500 RPM - 80% LOAD

-8
RME30 GLY10 GLY30
DI1 DI1 DI1
(b) (c)
16. Comparative
Figure 16. Comparativeassessment
assessment of of
experimental
experimentalresults of (a)ofinjection
results pressure
(a) injection profiles,profiles,
pressure (b) relative
(b)
changes changes
relative of injection duration,duration,
of injection and (c) relative change
and (c) of start
relative changeof injection
of start of(SOI) for test(SOI)
injection fuelsfor
DI1, RME30,
test fuels
GLY10,
DI1, and GLY30.
RME30, GLY10,Alland relative changes
GLY30. are expressed
All relative changes witharereference to the
expressed corresponding
with reference tovalue the
of fuel DI1. Experimental
corresponding value of fuelresults are given at 2500
DI1. Experimental resultsrpmareand
givenat 80% of full
at 2500 rpmload
and for the high-speed
at 80% of full load
single-cylinder
for the high-speedDI diesel engine “Lister
single-cylinder LV1”.
DI diesel engine “Lister LV1”.

Overall,
In Figureit 17a,
can be concluded
measured that thepressure
cylinder biodiesel-added
profiles blend
for test RME30
fuels indicates
DI1, RME30, an almost
GLY10, similar
and
behavior with conventional diesel oil DI1 regarding its effect on fuel
GLY30 at 2500 rpm and at 80% of full load are compared. Results in Figure 17a are given for the injection process whereas the
diesel/glycol
average engineethers fuelInGLY30
cycle. Figuredemonstrated
17b, the relativethe highest
changesimpact on fuel injection
of indicated power for process leadingblends
oxygenated to the
highest reduction in injection duration and the highest delay in injection
RME30, GLY10, and GLY30 with reference to the corresponding value of conventional diesel oil DI1 process initiation (i.e., highest
reduction
at 2500 rpm of SOI)
and at compared
80% of fullto conventional diesel oilInDI1.
load are presented. Figure 17c, the relative changes of ISFC for
In Figure 17a, measured cylinder pressure
oxygenated blends RME30, GLY10, and GLY30 with reference profiles for test fuels to DI1, RME30,
the GLY10, and
corresponding GLY30
value of
at 2500 rpm and at 80% of full load are compared. Results in Figure 17a
conventional diesel oil DI1 at 2500 rpm and at 80% of full load. As evidenced from Figure 17a, the are given for the average
engine cycle.
transition from In fuel
Figure DI117b,
to the relative changes
oxygenated fuel RME30of indicated
results power for oxygenated
in slightly higher cylinderblends RME30,
pressures
around TDC mainly due to the higher fuel heating power (fuel consumption in kg/s x LHV inatkJ/kg)
GLY10, and GLY30 with reference to the corresponding value of conventional diesel oil DI1 2500
rpm and at 80% of full load are presented. In Figure 17c, the relative
of RME30 compared to fuel DI1. The slightly higher cylinder pressures around TDC observed for changes of ISFC for oxygenated
blends RME30,
oxygenated fuelGLY10,
RME30and GLY30 to
compared with reference
fuel to the
DI1 result, corresponding
as evidenced morevalue
clearlyof from
conventional
Figure 17b,diesel
in
oil DI1 at 2500 rpm and at 80% of full load. As evidenced from Figure
the small increase of engine indicated power by almost 2%. The transition from fuel DI1 to fuel 17a, the transition from fuel
DI1 to oxygenated
GLY10 results in almostfuel RME30similarresults in slightly
cylinder higher
pressures, cylinder
which are pressures aroundcompared
slightly lower, TDC mainly due
to fuel
to the higher fuel heating power (fuel consumption in kg/s x LHV in
RME30 and to fuel DI1, mainly due to the almost same fuel heating power values between fuels DI1kJ/kg) of RME30 compared to
fuel DI1. The slightly higher cylinder pressures around TDC observed
and GLY10. The slightly lower cylinder pressures around TDC observed for oxygenated fuel GLY10 for oxygenated fuel RME30
compared to
compared to fuel
fuelDI1 DI1result,
result,asasevidenced
evidenced more
moreclearly from
clearly Figure
from 17b,17b,
Figure in the
in small increase
the small of engine
reduction of
indicated power by almost
engine indicated power by almost 1%. 2%. The transition from fuel DI1 to fuel GLY10 results in almost similar
cylinder
The pressures,
transition whichfrom fuelare DI1
slightly lower,
to fuel compared
GLY30 resultstoinfuel
lower RME30 and pressures,
cylinder to fuel DI1,which
mainlyare due to
also
the almost
lower same fuel
compared heating
to other twopower values fuels,
oxygenated between fuelsdue
mainly DI1 toand theGLY10.
lower Thefuel slightly
heating lower
powercylinder
of fuel
GLY30 compared to all other fuels. The lower cylinder pressures around TDC and during the
expansion stroke observed for oxygenated fuel GLY30 compared to fuel DI1 result, as evidenced
more clearly from Figure 17b, in reduction of engine indicated power by almost 5.5%, which is the
highest indicated power reduction observed in the present analysis compared to conventional diesel
to the LHV of fuel DI1.
Overall, it can be concluded from the comparison of the effects of fuels RME30 and GLY30,
which have the same blending ratio of two different oxygenated additives (RME vs. glycol ethers),
on indicated power and ISFC that the synthetic oxygenated additive (glycol ethers) have more
detrimental impact on indicated power and ISFC compared to the same proportion of a RME mainly
Energies 2019, 12, 1547 26 of 36
due to lower LHV. The comparison of the effects of fuels GLY10 and GLY30, which have the same
oxygenated additive (glycol ethers) at different proportions, also showed that GLY30 has more
pressures around
detrimental TDCon
influence observed
indicatedforpower
oxygenated fuelcompared
and ISFC GLY10 compared
to GLY10to mainly
fuel DI1due
result, as evidenced
to lower LHV as
more clearly from Figure 17b, in the small reduction of
a result of the higher oxygen content of GLY30 compared to GLY10.engine indicated power by almost 1%.

CYLINDER PRESSURE (bar)

(a)
11
10
"LISTER LV1" ENGINE
9
2500 RPM - 80% LOAD
8
7
6
5
4
3
2
1
0
RME30 GLY10 GLY30
DI1 DI1 DI1
(b) (c)
17. Comparative
Figure 17. Comparativeassessment
assessment of of
experimental
experimentalresults of (a)ofcylinder
results pressure
(a) cylinder profiles,profiles,
pressure (b) relative
(b)
changeschanges
relative of indicated power, and
of indicated (c) relative
power, and (c)change
relativeofchange
indicated specific fuel
of indicated consumption
specific (ISFC) for
fuel consumption
test fuels
(ISFC) forDI1,
testRME30, GLY10,
fuels DI1, and GLY30.
RME30, GLY10, All
andrelative
GLY30.changes are expressed
All relative changes with reference towith
are expressed the
reference to the corresponding value of fuel DI1. Experimental results are given at 2500 rpm andfor
corresponding value of fuel DI1. Experimental results are given at 2500 rpm and at 80% of full load at
the high-speed
80% of full loadsingle-cylinder DI diesel
for the high-speed engine “Lister
single-cylinder LV1”.engine “Lister LV1”.
DI diesel

TheFigure
In transition
18a,from fuel DI1 toresults
experimental fuel GLY30 results in lower
for instantaneous cylinder
net pressures,
heat release rateswhich are fuels
for test also lower
DI1,
compared to other two oxygenated fuels, mainly due to the lower fuel heating
RME30, GLY10, and GLY30 at 2500 rpm and at 80% of full load are compared. Results in Figure 18a power of fuel GLY30
compared
are given forto all
theother fuels.engine
average The lower
cycle, cylinder pressures
and all heat releasearound TDC and
rate profiles areduring the expansion
smoothed using an
stroke observed for oxygenated fuel GLY30 compared to fuel DI1
intrinsic function of MATLAB [40]. In Figure 18b, the relative changes of ignition result, as evidenced more
angleclearly
for
from Figure 17b, in reduction of engine indicated power by almost 5.5%,
oxygenated blends RME30, GLY10, and GLY30 with reference to the corresponding value of which is the highest indicated
power reduction
conventional observed
diesel oil DI1inat
the2500
present
rpmanalysis
and at compared
80% of fulltoload
conventional diesel operation
are presented. Values ofwith fuel
ignition
DI1. According to Figure 17a, the transition from diesel oil DI1 to oxygenated
angles as distances in crank angle degrees from TDC, which were derived from the processing of thefuel RME30, to GLY10,
and finally to fuel
instantaneous heatGLY30
releaseresults in a progressive
rate profiles, increase
are used for of ISFC, which
the calculation is associated
of results shown inwith the18b.
Figure lower
In
values of LHV of oxygenated fuels RME30, GLY10, and GLY30 compared to
Figure 18c, the relative changes of ignition delay for oxygenated blends RME30, GLY10, and GLY30the LHV of fuel DI1.
with Overall,
referenceit to
canthe
becorresponding
concluded fromvalue the comparison of thediesel
of conventional effectsoil
of DI1
fuelsatRME30
2500 rpm and and
GLY30, which
at 80% of
have the same blending ratio of two different oxygenated additives (RME vs. glycol ethers), on indicated
power and ISFC that the synthetic oxygenated additive (glycol ethers) have more detrimental impact
on indicated power and ISFC compared to the same proportion of a RME mainly due to lower LHV.
The comparison of the effects of fuels GLY10 and GLY30, which have the same oxygenated additive
(glycol ethers) at different proportions, also showed that GLY30 has more detrimental influence on
indicated power and ISFC compared to GLY10 mainly due to lower LHV as a result of the higher
oxygen content of GLY30 compared to GLY10.
In Figure 18a, experimental results for instantaneous net heat release rates for test fuels DI1,
RME30, GLY10, and GLY30 at 2500 rpm and at 80% of full load are compared. Results in Figure 18a are
initiation of combustion observed when changing from fuel DI1 to fuel GLY10 and also to fuel
GLY30 can be attributed to the higher cetane number of these two oxygenated fuels compared to fuel
DI1. Also, the transition from fuel DI1 to fuel GLY10 and also to fuel GLY30 results in less intense
premixed combustion phases, which is associated with the relative reductions of ignition delay
when
Energieschanging from fuel DI1 to fuel GLY10 and to fuel GLY30, as observed from Figure 18c. 27 of 36
2019, 12, 1547
According to Figure 18c, the highest relative reduction of ignition delay when changing from
conventional diesel oil to oxygenated fuels is observed in the case of the blend containing 31.3
given forglycol
mass-% the average
ethers engine
(GLY30)cycle, and all heat
as evidenced fromrelease
Figurerate profiles
18c, whichare is smoothed
associatedusing
with thean intrinsic
highest
function of MATLAB [40]. In Figure 18b, the relative changes of ignition
reduction of SOI (i.e., highest retardation of injection timing) as observed from Figure 16c. angle for oxygenated blends
RME30, GLY10, andthe
Consequently, GLY30
use ofwith referencefuels
oxygenated to theincorresponding
“Lister LV1” engine value resulted
of conventional diesel oil DI1
in the acceleration of
at 2500 rpm and at 80% of full load are presented. Values of ignition
combustion initiation compared to conventional diesel oil operation with the effects more angles as distances in crank
angle degrees
pronounced infrom TDC,
the case ofwhich were derived
biodiesel-added fuelfrom
RME30.the processing
Also, the use of of
theoxygenated
instantaneousblendsheatresulted
release
rate
in theprofiles,
reductionare used for thedelay
of ignition calculation
compared of results shown in Figure
to conventional diesel18b. In Figure with
oil operation 18c, the therelative
effects
changes of ignition delay for oxygenated blends RME30, GLY10,
more pronounced in the case of glycol ethers-added fuel GLY30, which had the highest cetane and GLY30 with reference to the
corresponding
number from value of conventional
all fuels examined. Finally,diesel oilthe DI1useat 2500 rpm and at additives
of oxygenated 80% of fullresulted
load areingiven. the
As evidenced from Figure 18a, the transition from fuel DI1 to oxygenated
downplaying of premixed combustion phase intense compared to conventional diesel oil operation fuel RME30 results in an
earlier initiation of combustion, thus leading to the relative increase of
with the effects to be more pronounced in the case of glymes-added fuel GLY30, which indicated the ignition angle (i.e., distance
from TDC)
highest as evidenced
reductions moredelay
in ignition clearlyandfrom Figure
thus, 18b.
to fuel The earlier
physical initiation preparation
and chemical of combustion observed
time.
whenItchanging from fuel DI1 to fuel RME30 can be attributed to the slightly
should also be mentioned that the experimental results shown in this section for cylinder higher cetane number of
fuel RME30 compared to fuel DI1. Also, the transition from fuel DI1 to
pressure (Figure 17a) and for ignition delay relative reduction (Figure 18c) are in accordance on afuel RME30 results in a less
intense premixed
qualitative basis withcombustion phase, experimental
corresponding which is associated
resultswith the relative
derived reduction
in a Ricardo Hydra of optical
ignitionengine
delay
when changing from fuel DI1 to fuel RME30,
[33] with the same oxygenated fuels used in the present study. as observed from Figure 18c.

30
"LISTER LV1" ENGINE
25 2500 RPM - 80% LOAD
AVERAGE CYCLE
GLY30
20
GLY10
RME30
15
DI1
10

-5
160 165 170 175 180 185 190 195 200 205 210 215 220
CRANK ANGLE (deg)
Energies 2018, 11, x FOR PEER REVIEW 28 of 36
(a)
RELATIVE CHANGE OF
IGNITION DELAY (%)

(a) (b)
Figure 18. Comparative assessment of experimental results of (a) instantaneous net heat release rate
profiles,
profiles, (b)
(b) relative
relativechange
changeofofignition
ignitionangle,
angle,and (c)(c)
and relative change
relative of ignition
change delay
of ignition for test
delay for fuels DI1,
test fuels
RME30, GLY10, and GLY30. All relative changes are expressed with reference
DI1, RME30, GLY10, and GLY30. All relative changes are expressed with reference to the to the corresponding
value of fuel DI1.
corresponding Experimental
value results
of fuel DI1. are given results
Experimental at 2500 are
rpmgiven
and at
at80%
2500ofrpm
full and
loadatfor80%
the of
high-speed
full load
single-cylinder
for the high-speed DI diesel engine “Lister
single-cylinder LV1”.
DI diesel engine “Lister LV1”.

As Figure
In observed fromthe
19a–d, Figure 18a, changes
relative the transition from fuel DI1
of combustion to oxygenated
durations in crankfuel GLY10
angle of 5%and also
(CA5)
to oxygenated fuel GLY30 results in an earlier initiation of combustion, thus leading
(Figure 19a), 25% (CA25) (Figure 19b), 50% (CA50) (Figure 19c), and 90% (CA90) (Figure 19d) of to the relative
increase of ignition
oxygenated angle inGLY10,
fuels RME30, both cases,
andasGLY30
evidenced
withmore clearly
reference tofrom Figure
“base” fuel18b.
DI1 The earlier
at 2500 rpminitiation
and at
of combustion observed when changing from fuel DI1 to fuel GLY10 and also to fuel
80% of full load are shown. As evidenced from Figure 19a–d, the diesel engine operation with all GLY30 can be
attributed to the higher cetane number of these two oxygenated fuels compared to fuel
examined oxygenated fuels results in the relative increase of all combustion durations compared to DI1. Also,
conventional diesel operation with fuel DI1. The highest impact on the elongation of all combustion
durations compared to conventional diesel oil operation is evidenced in the case of the oxygenated
fuel GLY30, which indicates the lowest LHV compared to all other fuels, whereas the lowest impact
on the relative increase of all combustion durations is witnessed in the case of biofuel-added blend
RME30. The highest values of relative increase of combustion duration for all examined oxygenated
Energies 2018, 11, x FOR PEER REVIEW 28 of 36

Energies 2019, 12, 1547 28 of 36

RELATIVE CHANGE OF
IGNITION DELAY (%)
the transition from fuel DI1 to fuel GLY10 and also to fuel GLY30 results in less intense premixed
combustion phases, which is associated with the relative reductions of ignition delay when changing
from fuel DI1 to fuel GLY10 and to fuel GLY30, as observed from Figure 18c.
According to Figure 18c, the highest relative reduction of ignition delay when changing from
conventional diesel oil to oxygenated fuels is observed in the case of the blend containing 31.3 mass-%
glycol ethers (GLY30) as evidenced from Figure 18c, which is associated with the highest reduction of
SOI (i.e., highest retardation of injection timing) as observed from Figure 16c.
Consequently, the use(a) of oxygenated fuels in “Lister LV1” engine resulted (b) in the acceleration of
combustion initiation compared to conventional diesel oil operation with the effects more pronounced
Figure 18. Comparative assessment of experimental results of (a) instantaneous net heat release rate
in the case of biodiesel-added fuel RME30. Also, the use of oxygenated blends resulted in the reduction
profiles, (b) relative change of ignition angle, and (c) relative change of ignition delay for test fuels
of ignition delay compared to conventional diesel oil operation with the effects more pronounced in the
DI1, RME30, GLY10, and GLY30. All relative changes are expressed with reference to the
case of glycol
correspondingethers-added fuelDI1.
value of fuel GLY30, which had
Experimental the highest
results are givencetane
at 2500 number
rpm and from all fuels
at 80% of fullexamined.
load
Finally, the use of oxygenated additives resulted in
for the high-speed single-cylinder DI diesel engine “Lister LV1”. the downplaying of premixed combustion phase
intense compared to conventional diesel oil operation with the effects to be more pronounced in the
case In
of glymes-added
Figure 19a–d, fuel the GLY30,
relativewhich
changes indicated the highest
of combustion reductions
durations in ignition
in crank angledelay
of 5% and(CA5)
thus,
to fuel physical
(Figure 19a), 25% and(CA25)
chemical preparation
(Figure 19b), 50% time.(CA50) (Figure 19c), and 90% (CA90) (Figure 19d) of
It should
oxygenated also
fuels be mentioned
RME30, GLY10, and that GLY30
the experimental
with reference results shownfuel
to “base” in this
DI1section
at 2500 forrpm cylinder
and at
80% of full load are shown. As evidenced from Figure 19a–d, the diesel engine operation withonalla
pressure (Figure 17a) and for ignition delay relative reduction (Figure 18c) are in accordance
qualitativeoxygenated
examined basis with corresponding
fuels results in experimental
the relativeresults
increase derived
of allincombustion
a Ricardo Hydra optical
durations engine [33]
compared to
with the same oxygenated fuels used in the present study.
conventional diesel operation with fuel DI1. The highest impact on the elongation of all combustion
In Figure
durations compared19a–d,tothe relative changes
conventional diesel oil of operation
combustion durations in
is evidenced inthecrank
caseangle
of theofoxygenated
5% (CA5)
(Figure
fuel 19a),which
GLY30, 25% (CA25)
indicates (Figure 19b),LHV
the lowest 50% compared
(CA50) (Figure 19c), and
to all other fuels,90% (CA90)
whereas the(Figure
lowest 19d)impactof
oxygenated fuels RME30, GLY10, and GLY30 with reference to “base” fuel
on the relative increase of all combustion durations is witnessed in the case of biofuel-added blend DI1 at 2500 rpm and at 80%
of full load
RME30. Theare shown.
highest As evidenced
values of relativefrom Figureof19a–d,
increase the diesel
combustion enginefor
duration operation with all
all examined examined
oxygenated
oxygenated
fuels fuels results
is evidenced in thein the
caserelative
of CA5 increase
(Figure of all
19),combustion
whereas the durations
effectscompared to conventional
of oxygenated fuels on
diesel operation with fuel DI1. The highest impact on the elongation
combustion duration are reduced, progressively moving to the initial stages of combustion (CA5) to of all combustion durations
compared
the to conventional
final stages of combustion diesel(CA90).
oil operation
Hence,isthe evidenced
effects of inoxygenated
the case of the oxygenated
fuels fuel GLY30,
on the elongation of
which indicates the lowest LHV compared to all other fuels, whereas
combustion compared to baseline diesel operation are more pronounced in the case of initial the lowest impact on the relative
increase of combustion
premixed all combustion durations
(CA5, Figureis 19a)
witnessed
and theyin the case
are of biofuel-added
progressively blend RME30.
downplayed The highest
as moving from
values of relative
premixed increase of combustion
to diffusion-controlled duration for all examined oxygenated fuels is evidenced in the
combustion.
case of CA5 (Figure
Overall, it can be 19), whereas the
concluded thateffects of oxygenated
the glycol ethers-added fuelsblend
on combustion
GLY30 with duration are reduced,
the highest oxygen
progressively
content resultsmoving to the initial
in the highest stages
increase of of combustion
both premixed(CA5) to thecombustion
and total final stagesduration
of combustion
compared (CA90).
to
Hence, the
baseline effects
diesel of oxygenated
operation mainly due fuelstoonitsthe elongation
highest of combustion
fuel consumption compared
compared to other
to all baseline diesel
fuels. On
operation
the are more
other hand, the pronounced
lowest relative in the case of initial
deterioration of premixed
both premixed combustion
and total (CA5, Figure 19a)
combustion and they
duration is
are progressively
evidenced in the casedownplayed as movingblend
of biodiesel-added from premixed
RME30 comparedto diffusion-controlled combustion.
to diesel-only operation.

60 60
55 55
"LISTER LV1" ENGINE "LISTER LV1" ENGINE
50 2500 RPM - 80% LOAD 50 2500 RPM - 80% LOAD
45 45
40 40
35 35
30 30
25 25
20 20
15 15
10 10
5 5
0 0
RME30 GLY10 GLY30 RME30 GLY10 GLY30
DI1 DI1 DI1 DI1 DI1 DI1
(a) (b)
Figure 19. Cont.
Energies 2019, 12, 1547 29 of 36
Energies 2018, 11, x FOR PEER REVIEW 29 of 36

60 60
55 55
"LISTER LV1" ENGINE "LISTER LV1" ENGINE
50 2500 RPM - 80% LOAD 50 2500 RPM - 80% LOAD
45 45
40 40
35 35
30 30
25 25
20 20
15 15
10 10
5 5
0 0
RME30 GLY10 GLY30 RME30 GLY10 GLY30
DI1 DI1 DI1 DI1 DI1 DI1
(c) (d)
Comparative assessment
Figure 19. Comparative assessment of experimental
experimental results
results for the relative changes of combustion
durations of (a) 5% of fuel injected mass (CA5), (b) 25% of fuel injected mass (CA25), (c) 50% of fuel
injected mass
mass (CA50),
(CA50),andand(d)
(d)90%
90%ofof fuel
fuel injected
injected mass
mass (CA90)
(CA90) for for
testtest
fuelsfuels
DI1,DI1, RME30,
RME30, GLY10,
GLY10, and
and GLY30.
GLY30. All relative
All relative values
values are are expressed
expressed withwith reference
reference to to thecorresponding
the correspondingvalue
valueofoffuel
fuel DI1.
DI1.
Experimental
Experimental results
resultsare
aregiven
givenatat2500
2500rpm
rpmand
andatat80%
80% ofof
full
fullload
loadforfor
thethe
high-speed
high-speedsingle-cylinder DI
single-cylinder
diesel engine “Lister LV1”.
DI diesel engine “Lister LV1”.

Overall,
4.5. Effect of Testit can
Fuels beRME30,
concluded that and
GLY10, the glycol
GLY30ethers-added
on the Percentage blend GLY30
Change with the
of HSDI Dieselhighest
Engine oxygen
BSFC
content results in the highest increase of both premixed and
and Exhaust Soot, NO, CO, and HC Emissions Compared to Diesel Operation with “Base” Fuel DI1 total combustion duration compared
to baseline diesel operation mainly due to its highest fuel consumption compared to all other fuels.
On theFigure
other 20ahand, depicts the variation
the lowest of the percentage
relative deterioration change ofand
of both premixed measured BSFC ofduration
total combustion the three is
oxygenated fuels RME30, GLY10, and GLY30 with reference
evidenced in the case of biodiesel-added blend RME30 compared to diesel-only operation. to “base” fuel DI1 as function of engine
load at 2500 rpm. As evidenced from Figure 20a, the use of oxygenated fuels RME30, GLY10, and
4.5.
GLY30Effect of TestinFuels
results RME30, GLY10,
a deterioration of BSFCand GLY30
compared on theto Percentage
baseline dieselChange of HSDI Diesel
operation Engine loads.
at all engine BSFC
and
The Exhaust
highestSoot, NO, CO, and
deterioration HC Emissions
of measured BSFCCompared to DieselinOperation
is evidenced the casewith “Base” Fuel
of GLY30 at allDI1 examined
engine loads,
Figure 20awhich
depicts ranged from 7%
the variation ofatthe20% load to 15%
percentage changeat 80% load. The
of measured lowest
BSFC worsening
of the of BSFC
three oxygenated
compared
fuels RME30, to GLY10,
baseline anddiesel
GLY30 operation
with referenceis witnessed
to “base” infuel
theDI1case as of GLY10ofat
function all examined
engine load at 2500 engine
rpm.
loads, which ranged from 2% at 20% load to 6% at 80%
As evidenced from Figure 20a, the use of oxygenated fuels RME30, GLY10, and GLY30 resultsload. The aforementioned relative BSFC in
variations
a deteriorationbetween of BSFC GLY30 and DI1 to
compared and GLY10diesel
baseline and DI1 are associated
operation at all enginewith theloads.fact The
that highest
GLY30
indicates the of
deterioration highest
measured LHVBSFCreduction (9.95%)in
is evidenced compared
the case of toGLY30
DI1, whereas GLY10 indicates
at all examined engine loads, the lowest
which
LHV reduction (3.6%) compared to DI1.
ranged from 7% at 20% load to 15% at 80% load. The lowest worsening of BSFC compared to baseline
dieselFigure
operation20b shows the variation
is witnessed of theofpercentage
in the case GLY10 at all change
examinedof measured soot (in
engine loads, mg/l)ranged
which of the three
from
oxygenated fuels RME30, GLY10, and GLY30 with reference
2% at 20% load to 6% at 80% load. The aforementioned relative BSFC variations between GLY30 and to baseline fuel DI1 as function of
engine load at 2500 rpm. At all engine loads, the replacement of
DI1 and GLY10 and DI1 are associated with the fact that GLY30 indicates the highest LHV reduction the non-oxygenated fuel DI1 from
oxygenated
(9.95%) fuels RME30
compared and GLY10
to DI1, whereas GLY10 results in an the
indicates increase
lowest ofLHV
measuredreductiontailpipe
(3.6%) soot concentration,
compared to DI1.
which Figure 20b shows the variation of the percentage change of measured soot (in mg/L)engine
is more pronounced in the case of RME30 compared to GLY10 at all examined load
of the three
with the exception of 60% load where the effects of both RME30
oxygenated fuels RME30, GLY10, and GLY30 with reference to baseline fuel DI1 as function of engineand GLY10 on soot relative increase
are almost
load at 2500similar.
rpm. AtOppositely, the transition
all engine loads, from baseline
the replacement non-oxygenated
of the non-oxygenated fuel
fuel DI1 DI1
fromto oxygenated
oxygenated
fuel GLY30 results in reduction of measured exhaust soot
fuels RME30 and GLY10 results in an increase of measured tailpipe soot concentration, at all examined loads with the exception
which of is
40% load, where a relative increase is observed at almost 15%
more pronounced in the case of RME30 compared to GLY10 at all examined engine load with the compared to baseline operation. Soot
reductionsofexperienced
exception 60% load where withtheGLY30 range
effects from
of both 16% atand
RME30 80% load on
GLY10 to 21.5% at 60%increase
soot relative load compared
are almost to
DI1. Hence, though that the oxygenated fuel GLY30 indicated
similar. Oppositely, the transition from baseline non-oxygenated fuel DI1 to oxygenated fuel GLY30the highest fuel consumption at all
loads due
results to its lowest
in reduction LHV compared
of measured exhaust all soot other
at allfuels, it exhibits
examined loads the withlowest soot exhaust
the exception of 40%values
load,
where a relative increase is observed at almost 15% compared to baseline operation. Sootobservation
compared to both non-oxygenated DI1 and oxygenated fuels RME30 and GLY10. This reductions
can be attributed
experienced with GLY30possibly to the
range from fact16%thatat the
80%highload oxygen
to 21.5%content
at 60% (9%) of fuel GLY30,
load compared to DI1. which
Hence, is
readily inside
though that the fuel jet zones during
oxygenated fuel GLY30 combustion,
indicatedpromotes,
the highest as fuel
evidenced
consumptionsignificantly, soot oxidation
at all loads due to its
rate despite the increased average fuel/air equivalence
lowest LHV compared all other fuels, it exhibits the lowest soot exhaust values comparedratio due to increased fuel consumption,
to both
which is expected to enhance in-cylinder fuel formation rate.
non-oxygenated DI1 and oxygenated fuels RME30 and GLY10. This observation can be attributed
In Figure
possibly to the20c, factthe variation
that the high of oxygen
the percentage
contentchange
(9%) ofoffuelmeasured
GLY30,NO emissions
which is readily(in ppm)
insideoffuel
the
three oxygenated fuels RME30, GLY10, and GLY30 with reference to baseline fuel DI1 as function of
engine load at 2500 rpm. The transition from baseline fuel DI1 to oxygenated fuel RME30 results in a
Energies 2019, 12, 1547 30 of 36

jet zones during combustion, promotes, as evidenced significantly, soot oxidation rate despite the
increased average fuel/air equivalence ratio due to increased fuel consumption, which is expected to
enhance in-cylinder fuel formation rate.
In Figure 20c, the variation of the percentage change of measured NO emissions (in ppm) of the
three oxygenated fuels RME30, GLY10, and GLY30 with reference to baseline fuel DI1 as function of
engine load at 2500 rpm. The transition from baseline fuel DI1 to oxygenated fuel RME30 results in a
relative increase of NO emissions at all examined loads except from 40% load, where the relative change
of NO emissions between DI1 and RME30 is imperceptible. The percentage increase of NO emissions in
the case of RME30 ranges from 16% at 20% load to 3% at 60% load. The transition from “base” fuel DI1
to glymes-added fuel GLY10 results in an increase of NO emissions at 20% load (15.5% relative increase)
and at 80% load (7% relative increase), whereas at 40% and at 60%, NO emissions are decreased by
9% and 2%, respectively, compared to baseline diesel operation. The same behavior is evidenced
also in the case of transition from non-oxygenated fuel DI1 to oxygen-added fuel GLY30. Specifically,
a relative increase of NO emissions is witnessed at low and at high load, which spans from 10.6% at
20% load to 4.9% at 80% load when changing from fuel DI1 to GLY30. Oppositely, as evidenced from
Figure 20c, transition from fuel DI1 to fuel GLY30 results in reduction of NO emissions by 5.8% at 40%
of full load and by 4.4% at 60% of full load. Higher NO emissions observed in the case of fuel RME30
compared to baseline fuel DI1 at most of the examined engine loads are possibly associated with the
higher cylinder pressures and bulk-gas temperatures observed in the case of RME30 compared to fuel
DI1, which in combination with the slightly higher local fuel-bound oxygen availability, promoted
in-cylinder NO formation. In the case of oxygenated fuels GLY10 and GLY30, the lower fuel supplied
thermal energy compared to “base” fuel DI1 resulted in lower cylinder pressures and in-cylinder gas
temperatures resulted in lower negative effects to in-cylinder NO formation at 20% load compared to
RME30 and to exhaust NO reductions at 40% and at 60% load compared to baseline diesel operation.
A quite encouraging outcome is that, as evidenced from Figure 20b,c, in the case of oxygenated fuel
GLY30 and at 60% load, the well-known soot/NOx trade-off is reversed since both diesel-emitted soot
and NO are reduced compared to baseline diesel operation. As previous studies have shown, NO
emissions are correlated with various engine operating parameters such as intake temperature, fuel
consumption, intake air moisture, or in-cylinder water content [50]. For this reason, it should be
noticed here that further experimental studies are scheduled to be performed using different injection
nozzles for the test fuels DI1, RME30, GLY10, and GLY30. The different injection nozzles will help to
control the fuel injection process by keeping the fuel injection rate almost the same for the examined
test fuels and thus, for controlling the heat released energy per engine cycle and through this, the
diesel-generated flame temperatures. This experimental campaign will be conducted to potentially
isolate the effect of different LHV and physical properties of test fuels on in-cylinder gas temperature
and thus, to potentially clarify the impact of oxygenated additive structure and oxygen content not
only on NOx, but also on soot, CO, and HC emissions.
Figure 20d demonstrates the variation of the percentage change of measured CO emissions
(in ppm) of the three oxygenated fuels RME30, GLY10, and GLY30 with reference to baseline fuel DI1
as function of engine load at 2500 rpm. As observed from Figure 20d, with the exception of 20% load,
the transition from “base” fuel DI1 to oxygenated fuels RME30, GLY10, and GLY30 at all engine loads
resulted in a considerable deterioration of CO emissions compared to baseline diesel operation. The
more characteristic relative deterioration of tailpipe CO values is evidenced at 80% of full load, where
the relative increase of CO emissions ranges from 143% for GLY30 to 196% for GLY10 compared to
“base” fuel DI1. The substantial worsening of CO emissions when burning oxygenated fuels at high
engine load (i.e., 80%) can be ascribed to the increase of fuel consumption in the case of oxygenated
fuels due to LHV reduction for sustaining constant engine load, which lead to the increase of fuel/air
equivalence ratio and to the substantial promotion of CO emissions.
Figure 20e depicts the variation of the relative change of measured HC emissions (in ppm) of
the three oxygenated fuels RME30, GLY10, and GLY30 with reference to baseline fuel DI1 as function
Energies 2019, 12, 1547 31 of 36

of engine load at 2500 rpm. As evidenced from Figure 20e, the transition from baseline fuel DI1 to
oxygenated fuel RME30 results in a relative reduction of HC emissions at all examined loads except
from 80% load where the relative increase of NO emissions between DI1 and RME30 is rather limited.
The percentage reduction of HC emissions in the case of RME30 ranges from 41% at 20% load to 20%
at 60% load. Oppositely, the transition from baseline fuel DI1 to oxygenated fuel GLY10 results in
a relative increase of HC emissions ranging from 1.3% at 80% load to 7.85% at 40% load. At 60%
Energies 2018, 11, x FOR PEER REVIEW 31 of 36
load, the replacement of fuel DI1 by fuel GLY10 results in a small decrease of HC emissions of 2%.
According
2%. According to Figure 20e,20e,
to Figure the the
combustion
combustion of of
oxygenated
oxygenated fuel GLY30
fuel GLY30results
resultsininaareduction
reductionof
of HC
HC
emissions compared to conventional diesel oil DI1 combustion at all examined
emissions compared to conventional diesel oil DI1 combustion at all examined engine loads. Thisengine loads. This
reduction
reduction spans
spans from
from 0.4%
0.4% at
at 80%
80% load
load to
to 4%
4%atat40%
40%load.
load.
160
"LISTER LV1" ENGINE
140 2500 RPM
RELATIVE CHANGE OF

120 RME30/DI1
GLY10/DI1
100 GLY30/DI1
BSFC (%)

80
60
40
20
0
-20
-40
20 40 60 80
ENGINE LOAD (%)

(a) (b)
200
"LISTER LV1" ENGINE
175 2500 RPM
RELATIVE CHANGE OF

RME30/DI1
NO EMISSIONS (%)

150 GLY10/DI1
125 GLY30/DI1

100

75

50

25

-25
20 40 60 80
ENGINE LOAD (%)
(c) (d)

(e)
Figure 20.Comparative
Figure 20. Comparative assessment of experimental
assessment results for
of experimental the relative
results for the changes of (a)
relative brake-specific
changes of (a)
fuel consumption
brake-specific fuelBSFC, (b) measured
consumption BSFC, exhaust soot, (c) measured
(b) measured exhaust
exhaust soot, NO, (d) measured
(c) measured exhaust NO,exhaust
(d)
CO, and (e)exhaust
measured measuredCO,exhaust
and (e)total unburned
measured hydrocarbons
exhaust (HC) for
total unburned test fuels DI1,
hydrocarbons RME30,
(HC) GLY10,
for test fuels
and
DI1, GLY30.
RME30,Experimental results Experimental
GLY10, and GLY30. are given at 2500 rpm
results areand at 80%
given of full
at 2500 rpmload
andfor the high-speed
at 80% of full load
single-cylinder DI diesel engine “Lister LV1”.
for the high-speed single-cylinder DI diesel engine “Lister LV1”.

5. Conclusions
In the present study, an experimental investigation was performed in a single-cylinder
naturally-aspirated high-speed DI diesel engine “Lister LV1” at 2500 rpm and at four engine loads
(20%, 40%, 60%, and 80%) using one non-oxygenated diesel oil DI1, one oxygenated blend RME30
containing 68.8 mass-% DI1 and 31.2 mass-% RME, and two oxygenated blends GLY10 and GLY30
Energies 2019, 12, 1547 32 of 36

5. Conclusions
In the present study, an experimental investigation was performed in a single-cylinder
naturally-aspirated high-speed DI diesel engine “Lister LV1” at 2500 rpm and at four engine loads
(20%, 40%, 60%, and 80%) using one non-oxygenated diesel oil DI1, one oxygenated blend RME30
containing 68.8 mass-% DI1 and 31.2 mass-% RME, and two oxygenated blends GLY10 and GLY30
containing 10.2 mass-% and 31.3 mass-% glycol ethers respectively. During engine tests, experimental
data were received for cylinder pressure, injection pressure, TDC position, fuel consumption, and soot,
NO, CO, and HC emissions at all examined loads. An experimental data model was developed in
MATLAB [40] for processing initial signals for cylinder pressure, injection pressure, and TDC position,
and for generating engine performance characteristics and heat release rate results.
From the comparative evaluation of the effects of oxygenated fuels RME30, GLY10, and GLY30
compared to baseline diesel operation with fuel DI1 on engine performance characteristics and
measured exhaust emissions it can be concluded that:

• The oxygenated fuel GLY30 indicated the highest impact on fuel injection process leading to
the highest reduction in injection duration and the highest delay in injection process initiation
compared to all other oxygenated fuels.
• The combustion of the most highly oxygenated fuel GLY30 resulted in the highest reductions
in indicated power, ISFC, BSFC, and ignition delay compared to other oxygenated fuels. Also,
oxygenated fuel GLY30 showed the highest combustion durations compared to all other fuels
both during premixed and diffusion-controlled combustion phases.
• The combustion of oxygen-added fuel GLY30 indicated the lowest soot exhaust values compared
to both non-oxygenated DI1 and oxygenated fuels RME30 and GLY10.
• Combustion of all examined oxygenated fuels resulted in the deterioration of NO emissions
compared to baseline diesel operation at most of the examined operating points. However, there is
an engine operating point (60% load) where the combustion of the highly-oxygenated fuel GLY30
can reverse the well-known soot/NO trade-off of conventional diesel engines leading to reduction
of both soot and NO emissions.
• Combustion of all examined oxygenated fuels resulted in substantial deterioration of CO emissions
compared to baseline diesel operation at all examined engine loads.
• The use of the oxygenated fuel with the highest oxygen content (GLY30, 9% oxygen) resulted in
reduction of HC emissions compared to conventional diesel oil operation.

Overall, it can be concluded that for almost the same blending proportion (31 mass-%) in
conventional diesel oil, a biodiesel-agent (RME) appears to be have less detrimental impact on engine
performance parameters and specific fuel consumption and more positive impact on HC emissions
reduction compared to a mixture of synthetic oxygenates (glycol ethers). Oppositely, the same
percentage of glycol ethers in a diesel blend appears to have more beneficial impact on soot and NO
emissions compared to rapeseed methyl ester.

Author Contributions: Conceptualization, T.C.Z.; software, T.C.Z. and M.I.K.; validation, R.G.P.; data curation,
E.G.P. and M.I.K.; writing—original draft preparation, T.C.Z. and R.G.P.; writing—review and editing, E.G.P.
and M.I.K.
Funding: This research was funded by European Commission under “NeDeNeF–New Diesel Engines and New
Diesel Fuels” research program with contract number G3RD-CT 1999-00021.
Acknowledgments: Authors wish to express their gratitude to the European Commission for funding the research
program “NeDeNeF” under which the experimental investigation was conducted and to the Institut Francais du
Petrole (IFP) for coordinating it. Authors wish to thank also Fortum Oil & Gas Oy for preparing all the test fuels
and supplying us with invaluable raw data.
Conflicts of Interest: The authors declare no conflict of interest.
Energies 2019, 12, 1547 33 of 36

Nomenclature
ABDC After bottom dead center
ATDC After top dead center
BBDC Before bottom dead center
BDC Bottom dead center
BSFC Brake specific fuel consumption
BTDC Before top dead center
CA Crank angle
CA25 Combustion duration of 25% of fuel injected mass per cycle
CA5 Combustion duration of 5% of fuel injected mass per cycle
CA50 Combustion duration of 50% of fuel injected mass per cycle
CA90 Combustion duration of 90% of fuel injected mass per cycle
CO Carbon monoxide
deg Degrees
DI Direct injection
EGR Exhaust gas recirculation
EVO Exhaust valve opening
GLY Glycol ethers—glymes
HC Total unburned hydrocarbons
IMEP Indicated mean effective pressure
ISFC Indicated specific fuel consumption
IVC Inlet valve closing
NO Nitrogen monoxide
ppm Part per million
RME Rapeseed methyl ester
RPM Rotations per minute
SOI Start of injection
TDC Top dead center

References
1. Mollenhauer, K.; Tschoeke, H. Handbook of Diesel Engines; Springer–Verlag: Berlin/Heidelberg, Germany, 2010.
2. Wartsila 31, World’s Most Efficient 4-Stroke Diesel Engine. Available online: https://www.wartsila.com/docs/
default-source/product-files/engines/ms-engine/brochure-o-e-w31.pdf (accessed on 21 April 2019).
3. MAN Diesel & Turbo, Marine Engine IMO Tier II and Tier III Programme 2018. Available online: https:
//www.man-es.com/docs/default-source/marine/4510_0017_02web.pdf (accessed on 21 April 2019).
4. DieselNet: Engine & Emission Technology Online. Available online: https://www.dieselnet.com (accessed
on 21 April 2019).
5. Papagiannakis, R.G.; Krishnan, S.R.; Rakopoulos, D.C.; Srinivasan, K.K.; Rakopoulos, C.D. A combined
experimental and theoretical study of diesel fuel injection timing and gaseous fuel/diesel mass ratio effects
on the performance and emissions of natural gas-diesel HDDI engine operating at various loads. Fuel 2017,
202, 675–687. [CrossRef]
6. Golimowski, W.; Krzaczek, P.; Marcinkowski, D.; Gracz, W.; Walowski, G. Impact of biogas and waste fats
methyl esters on NO, NO2 , CO and PM emission by dual fuel diesel engine. Sustainability 2019, 11, 1799.
[CrossRef]
7. Kosmadakis, G.M.; Rakopoulos, D.C.; Rakopoulos, C.D. Methane/hydrogen fueling a spark-ignition engine
for studying NO, CO and HC emissions with a research CFD code. Fuel 2016, 185, 903–915. [CrossRef]
8. Rakopoulos, D.C. Heat release analysis of combustion in heavy-duty turbocharged diesel engine operating on
blends of diesel fuel with cottonseed or sunflower oils and their bio-diesel. Fuel 2012, 96, 524–534. [CrossRef]
9. Rakopoulos, D.C.; Rakopoulos, C.D.; Giakoumis, E.G. Impact of properties of vegetable oil, bio-diesel,
ethanol and n-butanol on the combustion and emissions of turbocharged HDDI diesel engine operating
under steady and transient conditions. Fuel 2015, 156, 1–19. [CrossRef]
Energies 2019, 12, 1547 34 of 36

10. Rakopoulos, D.C.; Rakopoulos, C.D.; Kyritsis, D.C. Butanol or DEE blends with either straight vegetable
oil or biodiesel excluding fossil fuel: Comparative effects on diesel engine combustion attributes, cyclic
variability and regulated emissions trade-off. Energy 2016, 115, 314–325. [CrossRef]
11. Rakopoulos, D.C.; Rakopoulos, C.D.; Giakoumis, E.G.; Komninos, N.P.; Kosmadakis, G.M.;
Papagiannakis, R.G. Comparative evaluation of ethanol, n-butanol, and diethyl ether effects as biofuel
supplements on combustion characteristics, cyclic variations, and emissions balance in light-duty diesel
engine. ASCE J. Energy Eng. 2017, 143, 04016044. [CrossRef]
12. Chauhan, B.-S.; Singh, R.-K.; Cho, H.M.; Lim, H.C. Practice of diesel fuel blends using alternative fuels: A
review. Renew. Sust. Energy Rev. 2016, 59, 1358–1368. [CrossRef]
13. Hosseinzadeh–Bandbafha, H.; Tabatabaei, M.; Aghbashlo, M.; Khanali, M.; Demirbas, A. A comprehensive
review on the environmental impacts of diesel/biodiesel additives. Energy Convers. Manag. 2018, 174,
579–614. [CrossRef]
14. Giakoumis, E.G.; Rakopoulos, D.C.; Rakopoulos, C.D. Combustion noise radiation during dynamic diesel
engine operation including effects of various biofuel blends: A review. Renew. Sust. Energy Rev. 2016, 54,
1099–1113. [CrossRef]
15. Giakoumis, E.G.; Rakopoulos, C.D.; Dimaratos, A.M.; Rakopoulos, D.C. Exhaust emissions of diesel engines
operating under transient conditions with biodiesel fuel blends. Prog. Energy Combust. Sci. 2012, 38, 691–715.
[CrossRef]
16. Giakoumis, E.G.; Sarakatsanis, C.K. A comparative assessment of biodiesel cetane number predictive
correlations based on fatty acid composition. Energies 2019, 12, 422. [CrossRef]
17. Rakopoulos, C.D.; Rakopoulos, D.C.; Kyritsis, D.C. Development and validation of a comprehensive two-zone
model for combustion and emissions formation in a DI diesel engine. Int. J. Energy Res. 2003, 27, 1221–1249.
[CrossRef]
18. Rakopoulos, C.D.; Rakopoulos, D.C.; Mavropoulos, G.C.; Kosmadakis, G.M. Investigating the EGR rate and
temperature impact on diesel engine combustion and emissions under various injection timings and loads
by comprehensive two-zone modeling. Energy 2018, 157, 990–1014. [CrossRef]
19. Rakopoulos, D.C.; Rakopoulos, C.D.; Giakoumis, E.G.; Papagiannakis, R.G. Evaluating oxygenated fuel’s
influence on combustion and emissions in diesel engines using a two-zone combustion model. ASCE J.
Energy Eng. 2018, 144, 04018046. [CrossRef]
20. Rakopoulos, C.D.; Hountalas, D.T.; Zannis, T.C. Theoretical Study Concerning the Effect of Oxygenated Fuels on
DI Diesel Engine Performance and Emissions; SAE Technical Paper 2004-01-1838; SAE International: New York,
NY, USA, 2004. [CrossRef]
21. Park, W.; Park, S.; Reitz, R.D.; Kurtz, E. The effect of oxygenated fuel properties on diesel spray combustion
and soot formation. Combust. Flame 2017, 180, 276–283. [CrossRef]
22. Rakopoulos, C.D.; Giakoumis, E.G.; Hountalas, D.T.; Rakopoulos, D.C. The Effect of Various Dynamic,
Thermodynamic and Design Parameters on the Performance of a Turbocharged Diesel Engine Operating under
Transient Load Conditions; SAE Technical Paper 2004-01-0926; SAE International: New York, NY, USA, 2004.
[CrossRef]
23. Labecki, L.; Lindner, A.; Winklmayr, W.; Uitz, R.; Cracknell, R.; Ganippa, L. Effects of injection parameters
and EGR on exhaust soot particle number-size distribution for diesel and RME fuels in HSDI engines. Fuel
2013, 112, 224–235. [CrossRef]
24. Klyus, O. Biofuels for self-ignition engines. Sci. J. Marit. Univ. Szczecin 2005, 7, 153–156.
25. Buyukkaya, E. Effects of biodiesel on a DI diesel engine performance, emission and combustion characteristics.
Fuel 2010, 89, 3099–3105. [CrossRef]
26. Allocca, L.; Mancaruso, E.; Montanaro, A.; Sequino, L.; Vaglieco, B.-M. Evaluation of RME (rapeseed methyl
ester) and mineral diesel fuels behaviour in quiescent vessel and EURO 5 engine. Energy 2014, 77, 783–790.
[CrossRef]
27. Imran, S.; Emberson, D.R.; Wen, D.S.; Diez, A.; Crookes, R.J.; Korakianitis, T. Performance and specific
emissions contours of a diesel and RME fueled compression-ignition engine throughout its operating speed
and power range. Appl. Energy 2013, 111, 771–777. [CrossRef]
28. Johansson, M.; Yang, J.; Ochoterena, R.; Gjirja, S.; Denbratt, I. NOx and soot emissions trends for RME, SME
and PME fuels using engine and spray experiments in combination with simulations. Fuel 2013, 106, 293–302.
[CrossRef]
Energies 2019, 12, 1547 35 of 36

29. Mancaruso, E.; Sequino, L.; Vaglieco, B.-M. GTL (Gas To Liquid) and RME (Rapeseed Methyl Ester)
combustion analysis in a transparent CI (compression ignition) engine by means of IR (infrared) digital
imaging. Energy 2013, 58, 185–191. [CrossRef]
30. Guido, C.; Beatrice, C.; Napolitano, P. Application of bioethanol/RME/diesel blend in a Euro5 automotive
diesel engine: Potentiality of closed loop combustion control technology. Appl. Energy 2013, 102, 13–23.
[CrossRef]
31. Tsolakis, A.; Megaritis, A.; Wyszynski, M.L.; Theinnoi, K. Engine performance and emissions of a diesel
engine operating on diesel-RME (rapeseed methyl ester) blends with EGR (exhaust gas recirculation). Energy
2007, 32, 2072–2080. [CrossRef]
32. Gómez-Cuenca, F.; Gómez-Marín, M.; Folgueras-Díaz, M.B. Effects of ethylene glycol ethers on diesel fuel
properties and emissions in a diesel engine. Energy Convers. Manag. 2011, 52, 3027–3033. [CrossRef]
33. Song, H.; Quinton, K.-S.; Peng, Z.; Zhao, H.; Ladommatos, N. Effects of oxygen content of fuels on combustion
and emissions of diesel engines. Energies 2016, 9, 28. [CrossRef]
34. Zannis, T.C.; Hountalas, D.T.; Kouremenos, D.A. Experimental investigation to specify the effect of oxygenated
additive content and type on di diesel engine performance and emissions. Trans. SAE J. Fuels Lubr. 2004, 113,
1723–1743. [CrossRef]
35. Zannis, T.C.; Hountalas, D.T. DI diesel engine performance and emissions from the oxygen enrichment of
fuels with various aromatic content. Energy Fuels 2004, 18, 659–666. [CrossRef]
36. Rakopoulos, C.D.; Hountalas, D.T.; Zannis, T.C.; Levendis, Y.A. Operational and environmental evaluation
of diesel engines burning oxygen-enriched air or oxygen-enriched fuels: A. review. Trans. SAE J. Fuels Lubr.
2004, 113, 1723–1743. [CrossRef]
37. Laskowski, R.; Chybowski, L.; Gawdzinska, K. An engine room simulator as a tool for environmental
education of marine engineers. In Advances in Intelligent Systems and Computing; Springer: Berlin/Heidelberg,
Germany, 2015.
38. AVL GCA—Gas Exchange and Combustion Analysis Software. Available online: https://www.avl.com/-/gca-
gas-exchange-and-combustion-analysis-software (accessed on 21 April 2019).
39. Kongsberg, K-Chief Vessel Performance System. Available online: https://www.kongsberg.com (accessed on
21 April 2019).
40. MATLAB and Statistics Toolbox Release 2014a, The MathWorks, Inc., Natick, Massachusetts, United States.
Available online: https://www.mathworks.com/products/matlab.html (accessed on 21 April 2019).
41. Zannis, T.C.; Hountalas, D.T.; Papagiannakis, R.G.; Levendis, Y.A. Effect of fuel chemical structure and
properties on diesel engine performance and pollutant emissions: Review of the results of four European
research programs. Trans. SAE J. Fuels Lubr. 2008, 117. [CrossRef]
42. Bueno, A.V.; Velásquez, J.A.; Milanez, L.F. Internal Combustion Engine Indicating Measurements. In Applied
Measuring Systems; Md. Zahurul Haq; IntechOpen: London, UK, 2012; Volume 2, ISBN 978-953-51-0103-1.
43. Chapra, S.C.; Canale, R.P. Numerical Methods for Engineers, 6th ed.; McGraw-Hill: New York, NY, USA, 2010;
ISBN 978-0-07-340106-5.
44. Hahn, B.; Valentine, D.T. Essential MATLAB for Scientists and Engineers, 3rd ed.; Butterworth-Heinemann:
Burlington, VT, USA, 2007; ISBN 13: 9-78-0-75-068417-0.
45. Byttner, S.; Rongvaldsson, T.; Wickstrom, N. Estimation of Combustion Variability Using In-cylinder Ionization
Measurements; SAE Technical Paper 2001-01-3485; SAE International: New York, NY, USA, 2001. [CrossRef]
46. Heywood, J.B. Internal Combustion Engine Fundamentals; McGraw-Hill: New York, NY, USA, 1988.
47. Ferguson, C.R. Internal Combustion Engines—Applied Thermosciences; John Wiley: New York, NY, USA, 1986.
48. Annand, W.J.D. Heat transfer in the cylinders of reciprocating internal combustion engines. Proc. Inst.
Mech. Eng. 1963, 177, 973–990.
49. Rakopoulos, C.D.; Mavropoulos, G.C. Experimental instantaneous heat fluxes in the cylinder head and
exhaust manifold of an air-cooled diesel engine. Energy Convers. Manag. 2000, 41, 1265–1281. [CrossRef]
Energies 2019, 12, 1547 36 of 36

50. Chybowski, L.; Laskowski, R.; Gawdzinska, K. An overview of systems supplying water into the combustion
chamber of diesel engines to decrease the amount of nitrogen oxides in exhaust gas. J. Mar. Sci. Technol.
2015, 20, 393–405. [CrossRef]

© 2019 by the authors. Licensee MDPI, Basel, Switzerland. This article is an open access
article distributed under the terms and conditions of the Creative Commons Attribution
(CC BY) license (http://creativecommons.org/licenses/by/4.0/).

S-ar putea să vă placă și