Download as pdf or txt
Download as pdf or txt
You are on page 1of 18

Fuel 360 (2024) 130269

Contents lists available at ScienceDirect

Fuel
journal homepage: www.elsevier.com/locate/fuel

Full Length Article

Combustion of liquid ammonia and diesel in a compression ignition engine


operated in high-pressure dual fuel mode
Karl Oskar Pires Bjørgen a, *, David Robert Emberson b, Terese Løvås a
a
Department of Energy and Process Engineering, NTNU, Kolbjørn Hejes vei 1, Trondheim 7034, Norway
b
Department of Marine Technology, NTNU, Jonsvannsveien 82, Trondheim 7050, Norway

A R T I C L E I N F O A B S T R A C T

Keywords: This study investigates the characteristics of a compression ignition engine fuelled by ammonia and diesel in
Ammonia high-pressure dual fuel mode. The study focuses on the effect of injection strategies on engine performance and
Internal combustion engine important emissions like NH3, NO/NO2 and N2O. The engine was operated with fixed intake air conditions and at
Dual fuel combustion
an engine speed of 1500 rpm for all operating points, and liquid ammonia was directly injected at 180 bar using a
Compression ignition
GDI injector. The ammonia energy shares tested were 40 %, 50 % and 60 %. Various combustion modes were
Ammonia spray
achieved by fixing the diesel injection timing at − 15 CAD aTDC and varying the ammonia injection timing
between -80 and 2.5 CAD aTDC. Injecting ammonia early (-80/-60 CAD aTDC) led to premixed-type combustion
with very high ammonia slip (65 %). Delaying the ammonia injection to -30 CAD aTDC, i.e. 15 CAD before the
diesel injection, yielded a very long ignition delay due to the strong influence of the ammonia injection on the
diesel injection, resulting in premixed combustion of ammonia and diesel. Injecting ammonia simultaneously or
after diesel did not affect the diesel ignition, yielding stable combustion where the ammonia burned during the
mixing-controlled phase. Temporally overlapping injections of ammonia and diesel resulted in the highest
combustion efficiency, lowest ammonia slip, highest NOx emissions, and lowest N2O emissions, i.e., the optimal
cases found. For temporally overlapping ammonia and diesel injections, extending the ammonia injection to after
the end of diesel injection was found to increase ammonia slip compared to injecting ammonia before diesel, due
to the limited time for liquid ammonia to vaporise, mix and combust in the expansion stroke. The NOx and N2O
emissions were found to have opposite trends, where the highest NOx and lowest N2O concentrations were
achieved for the operating points with the highest combustion efficiency. Delaying the combustion phasing
improved the combustion efficiency for the 40 % and 50 % ammonia energy share cases, while for the 60 %
cases, it deteriorated, which can be explained by the injection of ammonia after the end of diesel injection for the
60 % ammonia energy share cases.

World War. The fuel system on six buses was adapted to carry anhydrous
ammonia and coal gas. The engines were reported to operate success­
1. Introduction fully with ammonia in spark ignition (SI) mode and gave an early
indication of the potential of ammonia as engine fuel. Following this
Introducing ammonia as fuel in internal combustion engines (ICE) is work, several studies were conducted by the United States Army during
regarded as a promising step in replacing fossil fuels used in ships, the 1960s aiming to evaluate ammonia as a substitute fuel in SI engines
heavy-duty vehicles, and stationary engines for electricity production in [3,4,5,6,7], all concluding that ammonia can provide output power
the coming years [1]. The main driving factors of this are the achieved equal to the fossil alternatives in SI engines. In 1967, Pearsall and
reduction in carbon dioxide emissions and ammonia’s potential to be Garabedian [8] performed a study on using ammonia in SI, compression
produced from renewable energy sources via water electrolysis and the ignition (CI) and reactivity-controlled compression ignition (RCCI) en­
Haber-Bosch process, effectively serving as a chemical energy storage. gine mode, concluding that using a diesel pilot with ammonia mixed
The concept of ammonia as a transportation fuel is not new. The first into the port of the engine was possible. However, issues related to
occurrence of ammonia as fuel in ICEs took place in Belgium in 1943 [2] unreliable diesel injections resulted in inconsistent ignition, indicating
as a result of limited access to gasoline/diesel fuel during the Second

* Corresponding author.
E-mail address: karl.o.bjorgen@ntnu.no (K.O.P. Bjørgen).

https://doi.org/10.1016/j.fuel.2023.130269
Received 31 August 2023; Received in revised form 30 October 2023; Accepted 30 October 2023
Available online 6 December 2023
0016-2361/© 2023 The Author(s). Published by Elsevier Ltd. This is an open access article under the CC BY license (http://creativecommons.org/licenses/by/4.0/).
K.O.P. Bjørgen et al. Fuel 360 (2024) 130269

Nomenclature X Molar fraction [-]


MW Molar weight [kmol/kg]
HRRaccum Accumulated heat release [J] dQn/dt Net heat release rate [J/CAD]
AES Ammonia energy share [-] γ Specific heat ratio [-]
rNH3,slip Ammonia slip ratio [-]
w Characteristic in-cylinder gas velocity [m/s] Abbreviations
ηc Combustion efficiency [-] aTDC After top dead centre
V Cylinder volume [m3] CA Crank angle
ρ Density [kg/m3] CAD Crank angle degree
rρ Density ratio of ammonia and nitrogen [-] CI Compression ignition
B Engine bore [m] CR Compression ratio
dQ/dt Heat release rate [J/CAD] DME Dimethyl ether
h Heat transfer coefficient [W/m2K] FTIR Fourier Transform Infrared
p In-cylinder pressure [kPa] GDI Gasoline direct injection
Tcyl In-cylinder temperature [K] HPDF High pressure dual fuel
ηth Indicated thermal efficiency [-] HRR Heat release rate
Wi Indicated work [J] ICE Internal combustion engine
Efuel Injected fuel energy per cycle [J] IDT Ignition delay time
m Injected fuel mass per cycle [kg] NTNU Norges Teknisk-Naturvitenskaplige Universitet
LHV Lower heating value [J/kg] RCCI Reactivity-controlled compression ignition
ṁNH3,exhaust Mass flow rate of ammonia in exhaust [kg/s] rpm Revolutions per minute
ṁNH3,input Mass flow rate of ammonia into engine [kg/s] SI Spark ignition
ṁdiesel,input Mass flow rate of diesel into engine [kg/s] TDC Top dead centre
ṁinput Mass flow rate of fuel and air to engine [kg/s] THC Total hydrocarbon
Y Mass fraction [-]

that SI mode should be the preferred choice. An attempt to achieve found that the ammonia energy share (AES)1 should not exceed 60 %
compression ignition of directly injected ammonia in an engine with a since IDTs become too long, deteriorating engine operation signifi­
compression ratio of 30:1 was, in their case, unsuccessful. cantly. Gill et al. [16] investigated the performance of pure ammonia
Ammonia was thereafter not systematically investigated for ICEs for and dissociated ammonia (i.e. a mixture of hydrogen and nitrogen) as
several decades until it regained interest during the early 2000s, moti­ the main fuels and diesel as the pilot fuel in RCCI mode. The RCCI mode
vated by the need to reduce CO2 emissions. Grannell et al. [9] studied cases achieved the same engine load as the diesel-only case using 10-15
the combustion of ammonia in SI engines operating with gasoline, % less diesel by mass. The engine performance with ammonia and
finding that ammonia can replace most of the energy supplied by gas­ dissociated ammonia were both acceptable, while ammonia emissions
oline, but not entirely, due to operational problems such as high coef­ were very high for the neat ammonia case, indicating that unburned
ficient of variance for indicated mean effective pressure and low ammonia is a concern for such a configuration. Niki et al. [17] also
indicated thermal efficiency. Mørch et al. [10] tested ammonia with performed an experimental campaign with ammonia and diesel com­
hydrogen as a fuel in an SI engine, and found that a 10 % addition of bustion in RCCI mode. The maximum share of ammonia heat release
hydrogen resulted in the highest efficiency and mean effective pressure. contribution with respect to diesel was 18 %. For the tested AESs, the
Later, the same group investigated nitric oxide (NO) emissions from an concentration of unburned ammonia in the exhaust was very low.
SI engine almost identical to the former setup fuelled with 20 vol% Yousefi et al. [18] investigated the effect of AES (up to 40 %) and diesel
hydrogen and 80 vol% ammonia [11] and concluded that NO formation injection timing on the combustion performance of ammonia and diesel
is lower for stoichiometric combustion compared to gasoline operation, in a CI engine in RCCI mode. They found that increasing the AES caused
but for leaner mixtures, an increase of NO emissions was observed which overall poorer combustion characteristics due to the high auto-ignition
was attributed to fuel-bound nitrogen in ammonia combusting under temperature and low flame speed of ammonia, causing an increase in
lower temperatures. N2O emissions are also important to consider since ammonia and N2O emissions. The combustion efficiency achieved for
they have a global warming potential of 265 times that of CO2 over a AES at 40 % was approximately 65 %. Førby et al. [19] injected
100-year period [12]. N2O emissions were also measured in this study ammonia using a GDI injector and used a pilot injection of heptane to
and were found to increase for leaner mixtures, which was explained by ignite the mixture. The tested AES range was 80-98.5 %. Due to pre­
an increase in NH2 radicals and NO2 for the leaner conditions, which in mixed ammonia being trapped in crevices, the measured ammonia
turn promotes the production of N2O [13]. concentration in the exhaust was high, up to 27,000 ppm, but with
Gross and Kong [14] performed tests of direct injection of neat increasing AES, the ammonia emissions decreased due to increased
dimethyl ether (DME) and 80/20 % and 60/40 % DME-ammonia mix­ combustion efficiency. Increased combustion efficiency also led to lower
tures in a CI engine. The DME-ammonia was injected using a gasoline N2O and higher NO emissions. Similar findings were made by Nadimi
direct injection (GDI) injector, where both single and double injections et al. [20], where for increasing AES the unburned ammonia concen­
were tested. The tests showed that increasing the ammonia content in tration in the exhaust increased significantly, and simultaneously
the fuel mixture caused longer ignition delay times (IDT) and limited the decreased the N2O emissions.
engine load conditions. Increasing the injection pressure improved Combusting ammonia in SI or RCCI mode has shown to be chal­
combustion and emissions for the highest ammonia-content mixture. lenging in terms of keeping unburned ammonia emissions low and not
Reiter and Kong [15] studied ammonia and diesel in RCCI mode and having the possibility of altering the air-ammonia mixture inside the

1
see Equation 2.5.

2
K.O.P. Bjørgen et al. Fuel 360 (2024) 130269

cylinder. Direct injection of ammonia in a CI engine has shown to be ammonia mixture fraction distribution and combustion for injection of
equally challenging with poor autoignition properties of ammonia. A ammonia at -60, -40 and -20 CAD aTDC and fixed diesel injection at -15
promising alternative is to inject ammonia directly using a high-pressure CAD aTDC is discussed. The engine and injector configuration in the
fuel injector together with a separate direct injection of diesel, often simulation is similar to the engine used in the current study. The early
referred to as high-pressure dual fuel (HPDF) mode. The individual in­ injections of ammonia at -60 CAD aTDC are found to be flash boiling,
jection timings, in-cylinder flow conditions, and spray patterns can be leading to an increased vaporisation rate and a significant decrease in
used to improve engine performance and emissions. There are only a few the in-cylinder gas temperature and pressure. The early injection led to
experimental studies performed with ammonia in HPDF mode. the highest degree of premixing, resulting in combustion with high HRR
Scharl and Sattelmayer [21] and Scharl et al. [22] performed an and a high level of unburned ammonia. For ammonia injections at -40
experimental campaign in a rapid compression expansion machine fitted CAD aTDC, the ammonia was completely vaporised at the start of diesel
with two single-hole injectors for ammonia and diesel, injecting injection but did not achieve complete mixing with air. Despite the lack
ammonia at 480-530 bar. They measured the heat release rate (HRR) of premixing, a significant reduction in unburned ammonia was found
and optically observed spray and combustion characteristics, giving compared to -60 CAD aTDC. Injecting ammonia late at -20 CAD aTDC
insight into the details of the combustion. In [21], they concluded that a resulted in the highest combustion efficiency, along with the lowest
high ammonia burnout rate is achieved when there is a strong interac­ emissions of unburned ammonia and N2O, and the highest emission of
tion of the diesel and ammonia sprays combined with the diesel injection NO.
starting before the ammonia injection, causing the ammonia to burn as a As seen from previous work on HPDF mode with ammonia, it offers
diffusion flame. However, when ammonia is injected before or simul­ better utilisation of the inducted air, where the ammonia does not need
taneously with diesel for overlapping sprays, a longer diesel IDT was to be premixed with air in the intake, which also leads to significantly
observed. Also, wall quenching of the flame was observed to influence less unburned ammonia emissions related to trapped gas in crevices.
the HRR. The optimal condition and spray configuration were optically Additionally, more degrees of freedom are introduced related to the
investigated in [22]. The flame lift-off was found to not stabilise for the injection strategy, which enables increased control of the combustion
conditions tested, increasing linearly over the duration of the injection. mode and more options for optimising the engine performance and
Based on a validated spray model, they concluded that the ammonia emissions. The presented research above does not include measurement
spray quickly becomes lean downstream of the nozzle, even upstream of of the unburned ammonia and N2O, leaving a knowledge gap in the
the liquid length, with low temperatures close to the nozzle exit. The literature. These particular emissions are very important for ammonia
lack of flame stabilisation was explained by ammonia’s high auto­ combustion. Unburned ammonia, also referred to as ammonia slip, is a
ignition temperature and re-entrainment of product gases with rela­ measure of ammonia combustion efficiency as well as being an unde­
tively low temperatures caused by ammonia’s high latent heat of sirable emission due to its toxicity and environmental impact [25]. N2O
vaporisation and low adiabatic flame temperature. emissions have a very high global warming potential and are hard to
Zhang et al. [23] performed an experimental study on the combus­ reduce in after-treatment systems [26]. Based on this, the objective of
tion of ammonia and diesel in a low-speed two-stroke single-cylinder CI this study is to experimentally investigate the effect of ammonia energy
engine operating in HPDF mode relevant for marine applications. They share and injection strategy on the combustion of ammonia and diesel in
used two injectors for injecting ammonia and diesel separately, where a CI engine, where the engine performance and important emissions like
the ammonia injection pressure was set to 650 bar. They found that by ammonia, N2O and NOx are measured and reported.
varying the relative injection timing of ammonia and diesel, different
combustion regimes could be achieved, i.e. premixed, diffusion and 2. Methods
premixed-diffusion combustion. By keeping the diesel and ammonia
injection timing fixed at -8 crank angle degrees after top dead centre 2.1. Research engine rig
(CAD aTDC) and changing the injection duration of ammonia, they
observed that the ignition of diesel was undisturbed for the different The research engine rig at NTNU Department of Energy and Process
injected masses of ammonia. For emissions, the specific emissions in g/ Engineering is based on a single-cylinder 4-stroke CI engine fitted to a
kWh of total hydrocarbons (THC), CO and soot decrease with increased 7.5 kW electric AC drive. Details about the engine are listed in Table 1.
AES, while NOx increases. Keeping AES at 50 % and the diesel injection The engine is equipped with an air heater and a compressor capable of
timing fixed at -8 CAD aTDC, but varying the ammonia injection timing conditioning the intake air temperature and pressure. The diesel fuel
between -16, -8 and 0 CAD aTDC did not affect the initial diesel ignition. system consists of a high-pressure diesel pump driven by an AC motor.
Simultaneously injecting ammonia and diesel yielded the shortest The diesel fuel tank is placed on a scale logging the fuel mass con­
combustion duration, highest thermal efficiency, highest NOx, and THC/ sumption rate during engine operation and is further used for calcu­
CO/soot-specific emissions. Advancing the ammonia injection was lating the injected diesel fuel per cycle. The ammonia fuel system
argued to result in more ammonia-air premixing, generating regions consists of manual stainless steel ball valves and 6 mm Swagelok con­
with super lean mixtures of ammonia, while imposing diffusion com­ nections, including a 0.5-litre high-pressure vessel, where the liquid
bustion, shortening the combustion duration and increasing the indi­ ammonia is stored. The liquid ammonia is pressurised by nitrogen to a
cated thermal efficiency. The influence of diesel injection timing was maximum of 200 bar and fed to the engine. The high-pressure vessel
also investigated, where ammonia injection timing was fixed at -8 CAD
aTDC with 50 % AES, while the diesel injection was varied between -12 Table 1
to -6 CAD aTDC. The initial ignition coincided well with the diesel in­ Engine specifications. *10 % valve lift.
jection timing and was unaffected, even when ammonia was injected 2
Engine model Hatz 1B30
CAD (0.88 ms) before diesel. Comparing diesel-only with ammonia- Number of cylinders 1 –
diesel for the same total fuel energy input, the initial stages of com­ Compression ratio 17.5 –
bustion are similar, while the combustion duration is significantly Bore/Stroke 80.0/69.0 mm
shortened for the ammonia-diesel cases. NOx concentrations are overall Connecting rod length 105.5 mm
Crank radius 34.5 mm
higher for the ammonia-diesel cases. Displaced volume 0.3468 L
The current research group has performed a numerical study of the Intake valve opening -405 (-357*) CAD aTDC
injection of ammonia piloted with diesel using the computational fluid Intake valve closing -105 (-153*) CAD aTDC
dynamics software CONVERGE by Convergent Science Inc., WI, USA Exhaust valve opening 107 (153*) CAD aTDC
Exhaust valve closing 405 (357*) CAD aTDC
[24]. The effect of ammonia vaporisation, vaporisation cooling,

3
K.O.P. Bjørgen et al. Fuel 360 (2024) 130269

containing liquid ammonia is freely mounted in the vertical direction rotational orientation is such that the cone centre axis points away from
and sits on a scale. The liquid ammonia bottle is kept at room the engine head, towards the piston. The GDI injector is controlled using
temperature. a GDI injection driver from Life Racing Ltd, UK, while the diesel injec­
During the injection of ammonia, the mass of ammonia and com­ tion is controlled using a Drivven CompactRIO module from National
pressed nitrogen inside the ammonia fuel tank is measured by the scale. Instruments Corp, TX, USA. The injectors’ orientations are shown in
The decrease in mass is due to ammonia flowing out of the tank. Fig. 1.
Meanwhile, nitrogen flows into the tank to compensate for the pressure
drop, maintaining the high injection pressure. By assuming that the 2.3. Fuels
volume of ammonia flowing out of the cylinder is replaced by com­
pressed nitrogen flowing in, the mass flow rate of ammonia out can be The anhydrous ammonia used in the experiments was acquired from
calculated from ṁNH3 = (ṁtankrρ)/(rρ - 1), where ṁNH3 is the mass flow Linde Plc in a 26-litre bottle with a dip tube containing 13.8 kg of
rate of ammonia and ṁtank is the change in mass per time unit of the ammonia. The ammonia is refrigerant-grade with purity of 99.98 %,
ammonia fuel tank, and rρ = ρNH3 /ρN2 is the ratio of the ammonia and moisture levels below 200 ppm and oil impurities below 3 ppm.
nitrogen densities at the tank pressure and temperature. Ammonia has a very high latent heat of vaporisation, 1371.2 kJ/kg,
The in-cylinder pressure and the absolute intake pressure are logged compared to 270 kJ/kg for diesel fuel, meaning that ammonia requires
using crank angle resolved Kistler sensors, Kistler 6052C and 4011A, much more heat from the surroundings while vaporising, which has
respectively. The crankshaft position is logged using a shaft encoder been shown to affect combustion [22]. Ammonia has a lower heating
with a crank angle resolution of 4096 readings per rotation, i.e. 0.0879 value of 18.6 MJ/kg, approximately half that of diesel (43.3 MJ/kg).
degrees between each reading (9.76 µs between each reading at 1500 However, ammonia can combust completely with approximately half
rpm). The engine rig is controlled remotely from an external control the mass of oxygen than diesel, i.e. a stoichiometric air-fuel ratio of 6.07
room via National Instruments equipment for automation and data compared to 14.55 for diesel. This means more ammonia can be com­
acquisition. This is an important safety measure due to the very high busted with the same air mass flow rate, resulting in nearly the same
toxicity of ammonia. In addition, other measures have been imple­ amount of fuel energy as diesel. The diesel fuel used in the experiments
mented, such as improved ventilation for achieving negative pressure in was a reference diesel containing no biodiesel produced by Coryton
the engine room, an ammonia gas detection system, personal safety Advanced Fuels Ltd. Table 2 lists important properties of the reference
equipment and detailed routines for handling ammonia. diesel and ammonia.
Exhaust gas concentrations are measured using a Horiba MEXA ONE-
RS Gas Analyzer for the dry concentrations of O2 and CO2, having a
2.4. Operating conditions
sampling frequency of 1.0 Hz and a reported uncertainty of 0.1 % of the
measuring range. The O2 is measured by the magneto-pneumatic oxygen
The engine operating conditions are listed in Table 3. The intake air
principle, and CO2 by the non-dispersive infrared technique. The ana­
condition was kept fixed at temperature and pressure of 40◦ C and 1.2
lysers were calibrated with span and zero gases on a daily basis. For the
bar, respectively. The diesel injection energising time was kept fixed at
measurement of the wet concentrations of NO, NO2, NH3, N2O, CO, CO2
1.00 ms, resulting in an injected mass of diesel fuel per cycle of 115.34
and H2O, a CX4000 Fourier transform infrared (FTIR) gas analyser from
mg. The ammonia energising time was varied between 1.06 ms, 1.45 ms,
Gasmet Technologies Oy, Finland, was used. The FTIR sampling unit was
and 2.05 ms, resulting in an averaged injected mass of ammonia of
fitted to the exhaust stream via a heated probe, all holding a temperature
14.19 mg, 20.79 mg and 31.51 mg, corresponding to 40 %, 50 % and 60
of 180 ◦ C. The FTIR sampling time was set to 5 s and was continuously
% AES. The three operating points are referred to as AES40, AES50 and
operated during each engine operating point. The reported measure­
AES60, respectively. The maximum AES of 60 % was chosen since higher
ment uncertainty is <2 % of the measuring range by the manufacturer.
AES yielded very poor combustion of ammonia for the current operating
The exhaust concentrations presented in the Results section are aver­
conditions and engine configuration.
aged values over the last 500 engine cycles for each operating point. All
The injection timing scheme is shown in Fig. 2. For AES40, the first
exhaust gas concentrations are presented in wet molar fractions. The
parameter study was performed while keeping the diesel injection
CO2 and O2 wet molar fractions are calculated based on the water molar
timing fixed at -15 CAD aTDC, and varying the ammonia timing between
fraction measured by the FTIR.
-80 CAD aTDC and 2.5 CAD aTDC. The second parameter study was
2.2. Injectors

The engine head has been modified to fit a common rail (CR) diesel
injector and a GDI injector, where the GDI injector is used for injecting
ammonia. The CR diesel injector has a modified nozzle (DLLA 150P
1437) where four out of six holes have been blocked by laser welding,
leaving two holes with hole diameters of 0.12 mm. The drill hole angle
with respect to the axial direction is 75◦ and 120◦ separates the two
holes when observing the nozzle along its axial direction. The GDI
injector (Bosch HDEV-5-2LS) is a solenoid direct-acting injector with a
six-hole nozzle. The advantage of using a GDI injector compared to a
diesel CR injector for injecting ammonia is that there is no return fuel
flow and that the injector can operate at lower injector pressures given
that the ammonia injection system is limited to a maximum of 200 bar.
The GDI nozzle geometry was characterised using the silicone mould
technique [27] and CT scan, where the hole has a counterbore design
with a large diameter of 0.46556 mm (standard deviation 0.00368 mm)
and a small diameter of 0.23505 mm (standard deviation 0.00600 mm). Fig. 1. Diesel and ammonia injectors’ orientation with respect to the piston and
The angles with respect to the axial direction of the six holes vary be­ engine head. In the illustration, the top of the piston is positioned 3.00 mm from
tween 22.22◦ and 34.41◦ , meaning that the centre axis of the total spray the engine head, corresponding to a crank angle position of -20 or 20
cone is approximately 6.1◦ off from the injector axis. The GDI injector’s CAD aTDC.

4
K.O.P. Bjørgen et al. Fuel 360 (2024) 130269

Table 2 injection timings later than -15 CAD aTDC due to ammonia being
Fuel properties of diesel and liquid ammonia. aassumed based on [28]. bLiquid injected into the high-pressure gas in the cylinder created by diesel
ammonia, 25◦ C, 10 bar [29,30]. cDiesel and air, ϕ = 1.0, 470 K, 1 bar [31]. combustion, thereby resulting in a lower differential pressure over the
d
Ammonia and air, ϕ = 1.0, 473 K, 1 bar [32]. e90 % volume fraction distilled. GDI injector nozzle and consequently a lower injected mass of ammonia.
Diesel Ammonia Unit For example, for ammonia injection timings at -15 CAD aTDC, the AES is
Cetane number 60.4 0a
– 41 %, while for 2.5 CAD aTDC, the AES is 34 %, i.e. a gradual decrease of
Density 823.0 603.1b kg/m3 7 % points.
Kinematic viscosity 2.534 0.237b mm2/s
Lower heating value 43.33 18.6 MJ/kg
Latent heat of vaporisation 270 1371.2 kJ/kg 2.5. Calculations
Laminar burning velocity 82c 16.08d cm/s
Total aromatics 18.1 0 w% The HRR (dQ/dt) was calculated based on the measured in-cylinder
Boiling point at 1 bar 304.0e -33.6 ◦
C pressure and an estimated heat transfer to the walls [33]. The
Hydrogen content 13.3 17.6 w%
apparent HRR (dQn/dt, also known as net HRR) is the result of pressure
work on the piston and the change in sensible internal energy of the
performed by varying the combustion phasing by keeping the relative trapped gas in the cylinder and is calculated from the in-cylinder pres­
injection timing fixed. Thereby, this was repeated for AES50 and AES60, sure. The specific heat ratio of air (γ) was determined for the calculated
but only including the injection timing combinations that yielded low thermodynamic conditions at each CAD. The current calculation does
ammonia slip. Additionally, the effect of changing the AES for fixed not consider the effects of blow-by and fuel vaporisation on the HRR.
diesel injection timing at -15 CAD aTDC for ammonia injection timing at The wall heat transfer is calculated based on Woschni et al. [34], using
-30, -25 and -20 CAD aTDC was also studied. Some of the injection the parameter values C1 = 2.28, C2 = 0.00324, h =
timing combinations lead to more or less temporally overlapping in­ 3.26B́ -0.2p0.8T-0.53w0.8, where C1 and C2 are constants suitable for diesel
jections of ammonia and diesel, which is defined as the time at which engines, h is the heat transfer coefficient in W/m2K, B is the engine bore
ammonia and diesel injections are simultaneously taking place. in m, p is the in-cylinder pressure in kPa, Tcyl is the in-cylinder tem­
Considerations should be taken for AES and global equivalence ratio perature in K, w is a characteristic in-cylinder gas velocity in m/s. The
(including ammonia and diesel) for the fixed diesel injection timing engine wall temperature was assumed 200◦ C. Additionally, a correction
study, see Fig. 3. The AES and equivalence ratio decreases for ammonia to the calculated HRR is done by subtracting the apparent HRR calcu­
lated for the motored cycle. This eliminates the HRR offset caused by

Table 3
The average values for important operating parameters are calculated based on quantities measured during the current experimental campaign. The standard deviation
of each operating parameter is shown in the parenthesis.
AES40 AES50 AES60 unit

Engine speed 1503.8 (4.1) rpm


Intake pressure 1.2003 (0.0010) bar
Intake temperature 40.38 (0.50) ◦
C
Injection pressure diesel 931.84 (11.85) bar
Injection pressure ammonia 182.07 (6.62) bar
Air mass flow rate 5.14794 (0.02216) g/s
Diesel mass flow rate 0.11534 (0.00100) g/s
Ammonia mass flow rate 0.17754 (0.02216) 0.26082 (0.01597) 0.39735 (0.00721) g/s
Injection energising time diesel 1.00 (-) ms
Injection energising time ammonia 1.06 (-) 1.45 (-) 2.05 (-) ms
Global equivalence ratio 0.5350 (0.0245) 0.6352 (0.0176) 0.8004 (0.0047) –
AES 0.3959 (0.0316) 0.4927 (0.0178) 0.5969 (0.0044) –

Fig. 2. The injection timing combinations of ammonia and diesel investigated in the current study. The markers with different shades of blue within the circles
represent the AESs tested for the indicated injection timing combination. The dotted line represents the simultaneous injection of ammonia and diesel. The study is
organised in two parts, where the diesel injection timing is kept fixed (horizontal sweep), and where the combustion phasing is kept fixed (sweeps parallel to the
dotted line).

5
K.O.P. Bjørgen et al. Fuel 360 (2024) 130269

Fig. 3. Ammonia energy share and global equivalence ratio (including ammonia and diesel) of the tested conditions for fixed diesel injection timing. The equivalence
ratio for the diesel-only case was 0.312. The dotted line represents the simultaneous injection of ammonia and diesel.

geometrical errors based on the compression ratio or piston position. weight of the exhaust gas mixture. The ammonia slip ratio rNH3,slip is
The HRR curves presented in the Results section (Section 3) are the thereby defined as
ensemble-averaged curves of the last 500 engine cycles from each
ṁNH3 ,exhaust
operating point. rNH3 ,slip = , (4)
ṁNH3 ,input
The accumulated HRR curve is calculated by numerical integration
of the HRR curve from the position -100 CAD aTDC until 125 CAD aTDC, The ammonia energy share is defined as
without correcting for the cooling enthalpy from the vaporisation of
ṁNH3 ,input LHV NH3
ammonia and diesel. The crank angle positions for 10 %, 50 % and 90 % AES = (5)
ṁNH3 ,input LHV NH3 + ṁdiesel,input LHV diesel
accumulated heat release, i.e. CA10, CA50 and CA90, are based on the
accumulated HRR curve. The IDT was defined as the CAD position where
the HRR first reached 5 J/CAD. The accumulated heat release for an 3. Results
engine cycle is denominated as HRRaccum.
The indicated work per engine cycle Wi was calculated from the line 3.1. Effect of ammonia injection timing
integral of the in-cylinder pressure p and cylinder volume V as Wi =
∮ 3.1.1. Combustion characteristics
pdV. The indicated thermal efficiency ηth was thereby calculated by
The ensemble-averaged in-cylinder pressures for the motored curve,
dividing the indicated work by the injected fuel energy Efuel, calculated
diesel-only and ammonia-diesel operation are shown for the various
based on the measured injected mass of ammonia and diesel per engine
ammonia injection timing and fixed diesel injection timing at -15 CAD
cycle (mNH3 and mdiesel) and their respective lower heating values (LHV),
aTDC for all AESs (40 %, 50 % and 60 %) in Fig. 4. Likewise, the
i.e. Efuel = mNH3LHVNH3 + mdieselLHVdiesel.
ensemble-averaged HRR curves for diesel-only and ammonia-diesel
Wi cases are shown for fixed diesel timing at -15 CAD aTDC and varying
ηth = (1)
Efuel ammonia timings in Fig. 5, Fig. 6 and Fig. 7 for all AESs (40 %, 50 % and
60 %), respectively.
The combustion efficiency was calculated by dividing the accumu­
The diesel-only HRR curve shows a classical diesel combustion case
lated HRR (HRRaccum) by the injected fuel energy.
where the premixed combustion phase is apparent, followed by a
HRRaccum mixing-controlled combustion phase into the expansion stroke. The
ηc = (2)
Efuel premixed phase is affected by the IDT of diesel, allowing the fuel and air
to mix before ignition, where shorter ignition delay times normally
The ammonia slip ratio is calculated to provide a metric on how
result in less combustion as premixed and vice versa. The fluctuations
much of the input ammonia is emitted through the exhaust, i.e. not
seen on the HRR curve during the expansion stroke could be due to
combusted in the engine. It is defined as the ratio between the ammonia
acoustic oscillations caused by travelling pressure waves inside the
mass flow rate in the exhaust (ṁNH3,exhaust) and the input ammonia mass
channel where the pressure transducer is installed [35]. These should
flow rate (ṁNH3,input). The mass flow rate of ammonia in the exhaust is
therefore be considered an experimental uncertainty.
calculated by
In the case of ammonia and diesel burning in dual fuel mode, the
MWNH3 conventional diesel definitions of the premixed phase and the mixing-
ṁNH3 ,exhaust = ṁinput YNH3 = ṁinput XNH3 , (3)
MWexhaust controlled phase become invalid since ammonia is mixed in with the
air to different degrees, or ammonia is directly injected into the diesel
where the ṁinput is the total mass flow rate into the engine (including flame. However, the terms premixed combustion phase and mixing-
ammonia, diesel and air), YNH3 is the mass fraction of ammonia in the controlled combustion phase are still used in the discussion below to
exhaust, XNH3 is the measured molar fraction of ammonia in the exhaust, differentiate between the initial premixed phase and the subsequent
MWNH3 is the molar weight of ammonia, and MWexhaust is the molar mixing-controlled phase during the engine cycle. In the case of ammonia

6
K.O.P. Bjørgen et al. Fuel 360 (2024) 130269

Fig. 4. In-cylinder pressure for varying ammonia injection timing, fixed diesel injection timing at -15 CAD aTDC and for AES of 40 %, 50 % and 60 %. The diesel-only
and motored cases are shown as the dashed red and grey lines, respectively. The red horizontal line indicates the diesel energizing time.

being injected before the diesel pilot, the initial premixed phase is ini­ reduce the IDT because less ammonia is vaporised before the diesel in­
tialised by the diesel pilot, resulting in partially premixed combustion of jection starts, causing higher local temperatures. As the ammonia in­
diesel and ammonia, where the level of ammonia-air premixing depends jection timing is delayed to -20 CAD aTDC, the diesel premixed peak
on the injection timing of ammonia. The initial premixed phase thereby starts to appear again, and for ammonia injection timing at -17.5 aTDC
transitions into the mixing-controlled phase, where the diesel pilot CAD, the initial premixed diesel peak is at a similar CAD position as for
burns as a diffusion flame while any surrounding partially premixed or the diesel-only case.
premixed ammonia burns simultaneously. After the end of diesel in­ For ammonia injections simultaneous or later than diesel, i.e. for -15
jection, the mixing-controlled phase continues where ammonia and to 2.5 CAD aTDC (third and fourth row in Fig. 5), the initial ignition
diesel burn out in the expansion stroke. In the case where ammonia is occurs at the same CAD position as the diesel-only case and is not
injected late, i.e. such that it does not interfere with the initial diesel influenced by the ammonia injection. The premixed combustion phase
injection, the ammonia burns as a diffusion flame during the mixing- for the ammonia-diesel case is similar to the diesel-only case. However,
controlled phase. the IDT is slightly shorter, causing the premixed HRR peak to be lower
Fig. 5 shows the HRR for AES40. The first row shows ammonia in­ due to the shorter time for air and diesel to mix before ignition. The
jection timings of -80, -60, -40 and -30 CAD aTDC, where ammonia is shortened IDT is likely caused by higher in-cylinder gas temperatures
injected early enough to not overlap with the diesel injection. An initial due to higher cylinder wall temperatures and internal exhaust gas
premixed combustion peak varying in size is observed. By injecting recirculation due to the higher loads of AES40 compared to diesel-only.
ammonia at -80 CAD aTDC, it has a long time (65 CAD/7.22 ms) to During the mixing-controlled combustion phase, the additional HRR
vaporise and mix with air before the injected diesel ignites the mixture, from the ammonia injection is observed and is gradually shifted as the
resulting in no clear transition from the premixed to the mixing- ammonia injection timing is delayed.
controlled phase in the HRR curves, suggesting that premixed The HRR curves for AES50 are shown in Fig. 6. A similar behaviour is
ammonia burns simultaneously with diesel during the premixed and the observed for AES50, as for AES40, when delaying the ammonia injec­
mixing-controlled phase. A similar observation is made when injecting tion. For the early injections, with the longer IDT, i.e. -25 CAD aTDC
ammonia at -60 CAD aTDC. For both cases, the HRR is overall low, ammonia injection timing, premixed combustion is observed, where
suggesting that the ammonia is not combusting properly. As ammonia is diesel and ammonia burn fast with a high HRR due to the long IDT.
injected at -40 CAD aTDC, a more distinct premixed combustion phase is For AES60, only ammonia injections at -30, -25 and -20 CAD aTDC
observed, coinciding well with the premixed peak of the diesel-only were performed and are shown in Fig. 7. The IDTs for -30 and -25 CAD
case, followed by a high HRR during the mixing-controlled phase. The aTDC are long due to the injection of ammonia prior to the diesel pilot.
injection timing at -40 CAD aTDC leads to less premixed ammonia at the The long IDT results in fast combustion with high HRR and high
time of diesel ignition, causing mostly diesel to initially ignite followed fluctuations.
by combustion of the more inhomogeneous ammonia-air mixture. For Fig. 8 compares ammonia injections at -30, -25 and -20 CAD aTDC
-30 CAD aTDC, the longer ignition delay and the slow development of for AES40, AES50 and AES60, which gives a clearer picture of the effect
the premixed combustion phase indicate that the ignition of diesel is of increasing AES. The ignition delay times for these ammonia injection
affected by the ammonia injection. timings are similar for all AESs, as shown in Fig. 9. This indicates that the
In the second row in Fig. 5, the ammonia injection timing varies with IDT is unaffected by the amount of ammonia being injected but is mainly
a step of 2.5 CAD between each operating point of -25, -22.5, -20 and driven by the injection timing of ammonia. The HRR curves for all cases
-17.5 CAD aTDC. For ammonia timing of -25 CAD aTDC, compared to in this figure, except the -25 CAD aTDC for AES40, have a two-stage
the earlier injections, the IDT is longer, indicating that diesel ignition combustion process. For the AES40 case, for ammonia injection timing
has been significantly hampered. The long IDT gives diesel and at -30 CAD aTDC, the initial HRR peak is higher but less defined
ammonia time to mix before ignition, resulting in the highest HRR compared to the other AESs. When injecting ammonia for longer, the
achieved for all ammonia injection timings tested for AES40. The long initial HRR peak is lower, possibly due to more liquid ammonia being
ignition delay is due to ammonia vaporising and cooling the air in the present during this phase, requiring heat to vaporise. For AES60, the
region where diesel is injected, resulting from the high latent heat of HRR curve is wider and slightly lower than that of AES40 and AES50,
ammonia and the orientation of sprays relative to each other. This effect meaning that the longer injection of ammonia causes the mixing-
appears to be highest for the -25 CAD aTDC case, suggesting that the controlled combustion phase to burn with a lower HRR, most likely
earlier injection timing results in a larger reduction in local temperature also due to more liquid ammonia being present, requiring heat from the
(in the diesel spray region) due to the entire ammonia spray vaporising combustion for vaporisation and resulting in lower oxygen availability.
before diesel is injected. Ammonia injections later than -25 CAD aTDC The -25 CAD aTDC case shows that the longer injections (AES50 and

7
K.O.P. Bjørgen et al. Fuel 360 (2024) 130269

Fig. 5. Ensemble-averaged heat release rate profiles for varying ammonia injection timing, fixed diesel injection timing at -15 CAD aTDC and AES of 40 %. The blue
curves represent the ammonia/diesel cases, while the red dashed lines represent the diesel-only cases. The shaded area indicates the standard deviation of the
measured 500 engine cycles. The horizontal red and blue lines indicate the diesel and ammonia energizing times, respectively.

AES60) of ammonia affect the ignition of ammonia, causing a two- becomes less homogeneous the later the ammonia is injected. When
staged combustion process. In contrast, the AES40 case results in delaying the ammonia injection from -50 to -25 CAD aTDC, an increase
single-staged combustion of ammonia and diesel. For the latest injection in IDT is observed, showing that the mechanism is now not driven by the
of ammonia at -20 CAD aTDc, the ignition process is equal for all AESs, premixed ammonia-air gas temperature but rather the interaction be­
since the ammonia injections start simultaneously and all overlap with tween the liquid ammonia and diesel spray. As discussed above, strong
the time of ignition. In the expansion stroke, the HRR is proportional to interaction between the sprays causes the IDT to reach a maximum of
the AES. 0.85–0.89 ms at -25 CAD aTDC for all AESs. The relative injection timing
The ignition delay times are shown for varying ammonia injection between diesel and ammonia causing the highest IDT is thought to be
timings and fixed diesel timing at -15 CAD aTDC in Fig. 9. The ignition very dependent on the spray orientation in the engine cylinder, based on
delay reaches a local minimum by delaying the ammonia injection from the findings in the previous study by Scharl et al. [21]. For the ammonia
-80 to -60/-50 CAD aTDC. The shorter IDT for later injections could be injection at -30 CAD aTDC, AES60 has longer IDT than AES40 and
caused by liquid ammonia having less time to vaporise (before the diesel AES50 which could be due to it having the largest injected mass of
is injected) and mix with hot air before ignition, leading to a higher ammonia before ignition, i.e. the ammonia injection ends close to the
temperature of the ammonia-fuel mixture. The ammonia-air mixture start of ignition. By delaying the ammonia injection from -25 to -15 CAD

8
K.O.P. Bjørgen et al. Fuel 360 (2024) 130269

Fig. 6. Ensemble-averaged heat release rate profiles for varying ammonia injection timing, fixed diesel injection timing at -15 CAD aTDC and AES of 50 %. The blue
curves represent the ammonia/diesel cases, while the red dashed lines represent the diesel-only cases. The shaded area indicates the standard deviation of the
measured 500 engine cycles. The horizontal red and blue lines indicate the diesel and ammonia energizing times, respectively.

aTDC, i.e. from injecting ammonia 10 CAD before diesel, to simulta­ at -40 CAD aTDC. This means the initial combustion phase has a lower
neous injections, the IDT decreases by approximately 0.35 ms (40 % HRR for -40 CAD aTDC, as discussed in the section above. For injections
decrease). All AESs follow this trend quite closely. The minimum IDT for after -15 CAD aTDC, IDT-CA10 is stable due to the HRR only being
all ammonia injection timings is reached at -12.5 CAD aTDC for AES40 influenced by the diesel injection. For AES50, IDT-CA10 is stable for
and -10 CAD aTDC for AES50 and stabilises close to 0.5 ms for AES40 injections later than -30 CAD aTDC and is slightly longer than IDT-CA10
and AES50 for later injections, where the diesel injection timing controls of AES40. When advancing the ammonia injection from -25 to -30 CAD
the IDT. aTDC, IDT-CA10 for AES50 increases, which is also observed for AES60.
The time between IDT and the occurrence of 10 %, 50 % and 90 % For IDT-CA50, a minimum is found for ammonia injection timings of
released heat is shown in Fig. 10, referred to as IDT-CA10, IDT-CA50 and -25 and -22.5 CAD aTDC for all AESs. The minimum is due to the overall
IDT-CA90 durations, respectively. IDT-CA90 indicates how long the high HRR peak achieved for these injection timings due to long IDTs
combustion lasts, while IDT-CA10 and IDT-CA50 indicate the duration causing the fuel to burn in premixed mode, as seen above in the HRR
of the initial combustion phase (premixed phase) and the middle phase, curves. By delaying the ammonia injection toward -15 CAD aTDC, IDT-
respectively. Overall, IDT-CA10 is relatively stable for all AES40 CA50 becomes longer due to the gradual transition to two-staged com­
ammonia injection timings but reaches a maximum for early injections bustion with the premixed and mixing-controlled phases. Delaying the

9
K.O.P. Bjørgen et al. Fuel 360 (2024) 130269

Fig. 7. Ensemble-averaged heat release rate profiles for varying ammonia injection timing, fixed diesel injection timing at -15 CAD aTDC and AES of 60 %. The blue
curves represent the ammonia/diesel cases, while the red dashed lines represent the diesel-only cases. The shaded area indicates the standard deviation of the
measured 500 engine cycles. The horizontal red and blue lines indicate the diesel and ammonia energizing times, respectively.

Fig. 8. Ensemble-averaged heat release rate profiles for varying ammonia injection timing and fixed diesel injection timing at -15 CAD aTDC, for AES of 40 %, 50 %
and 60 %. The horizontal red and blue lines indicate the diesel and ammonia energizing times, respectively.

injection after -15 CAD aTDC for AES40 and AES50, IDT-CA50 is
observed to level off, i.e. approximately between -12.5 and -7.5 CAD
aTDC, but increases again for later injections.
IDT-CA90 for AES40 is relatively stable from -80 CAD aTDC to -17.5
CAD aTDC, having only a slight decrease during the period of long IDTs
and high HRR peaks for -25 CAD aTDC. Between -25 and -17.5 CAD
aTDC, a transition from long IDTs resulting in premixed combustion to
shorter IDTs controlled by the diesel injection occurs, causing IDT-CA90
to increase for AES50 and AES60 while only affecting AES40 to a small
degree. For AES40 and AES50, a sudden increase occurs between -17.5
and -15 CAD aTDC and levels off after -12.5 CAD aTDC, which is the
transition where ammonia is injected after diesel. For the latest injection
timings of AES40, ammonia is injected into the expansion stroke, which
extends the combustion period since liquid ammonia is vaporising and
combusting late in the expansion stroke, where the diesel HRR is low.
Fig. 9. Ignition delay time as a function of ammonia injection timing for fixed Generally, a longer combustion duration leads to higher heat losses,
diesel injection timing at -15 CAD aTDC for AES of 0 %, 40 %, 50 % and 60 %.
which leads to lower thermal efficiencies. The longer combustion du­
The vertical dashed line indicates the start of diesel injection.
rations for late injection of ammonia, for all AESs, will likely contribute
to a deterioration of the thermal efficiency.
The combustion and thermal efficiencies are shown in Fig. 11. The
combustion efficiency (left axis) for AES40 is lowest for early injections

10
K.O.P. Bjørgen et al. Fuel 360 (2024) 130269

For later injections of ammonia, the combustion efficiency decreases


slightly and increases for very late injections. It is important to note that
the actual AES for AES40 decreases after -15 CAD aTDC (see Fig. 3 in
Section 2, Methods). The influence of AES on combustion efficiency is
high, as observed when comparing AES40, AES50 and AES60. There­
fore, the increase in combustion efficiency for late injections is likely due
to the decrease in AES for the AES40 and AES50 operating points.
Although the AES gradually decreases from -15 CAD aTDC to 2.5 CAD
aTDC, the combustion efficiency decreases between -15 and -10 CAD
aTDC, meaning that the combustion is likely strongly deteriorated by the
delayed ammonia injection timing.
For AES50, the combustion efficiency is the same as AES40 for -30
CAD aTDC and decreases when delaying the injections. A local
maximum is observed for injections at -20 CAD aTDC and decreases
similarly to AES40 for injections after the diesel injection timing. AES60
shows a decreasing trend when delaying the ammonia injection. The
combustion efficiency trends are discussed together with the emissions
measurements in the following section.
The indicated thermal efficiency (right axis) follows a similar
behaviour to the combustion efficiency, although it can be optimised by
changing the combustion phasing, which is investigated later.

3.2. Emission characteristics

Fig. 10. Combustion durations for 10 %, 50% and 90 % of the released heat As discussed in the Introduction section (Section 1), the level of
(IDT-CA10 (○), IDT-CA50 (□) and IDT-CA90 (⋄)), for AES of 0 %, 40 %, 50 % unburnt fuel in the exhaust is of concern when dealing with ammonia
and 60 %, with fixed diesel injection timing at -15 CAD aTDC, as a function of
combustion. The ammonia emissions are shown in molar fraction in
ammonia injection timing. The combustion duration is calculated as the time
Fig. 12a and as the ammonia slip ratio in Fig. 12b, calculated as the ratio
between IDT and the respective fraction of heat released. The vertical dashed
line indicates the start of diesel injection.
of ammonia mass flow rate in the exhaust stream to the input ammonia
mass flow rate. For AES40, the concentration of ammonia emission is
close to 4 % (40,000 ppm) for the early injections at -80 CAD aTDC, and
falls below 1 % (10,000 ppm) for ammonia injections at -40 CAD aTDC
and later. The early injections of ammonia cause a very high ammonia
slip ratio, and between 20 % and 65 % of the ammonia is not combusted,
indicating that the combustion is poor for these operating points for the
current configuration. The high ammonia slip for early injections of
ammonia is likely due to ammonia mixing and distributing to areas in
the cylinder where the diesel flame is not reached. Additionally, the
ammonia flame is known to easily quench at the engine walls, which
occurs for early injections of ammonia due to premixing.
For ammonia injections at -30 CAD aTDC, the ammonia slip ratio for
all AESs falls within the range of 8–11 %. As the ammonia injection is
delayed further, the ammonia slip for AES40 and AES50 continues to
decline, while it increases for AES60. The increase in ammonia slip is
supported by the discussion above regarding the increased combustion
duration and drop in combustion efficiency, where late injected
ammonia burns poorly during the expansion stroke, causing a higher
probability of flame quenching. Likewise, for AES50, the ammonia slip
increases when injecting close to the diesel injection timing at -15 CAD
aTDC, also supported by the above discussion. For AES40, a sudden
increase, as seen for AES50 and AES60, is not observed, instead, a small
increase is observed between -10 and -7.5 CAD aTDC, also correlating
with the increase in the combustion duration in Fig. 10. The increase in
Fig. 11. Combustion efficiency (○) and indicated thermal efficiency (□), for ammonia slip for all the AESs is therefore driven by late ammonia in­
AES of 0 %, 40 %, 50 % and 60 %, with fixed diesel injection timing at -15 CAD jection timing relative to the diesel injection. It seems like continuing
aTDC, as a function of ammonia injection timing. The vertical dashed line in­ ammonia injection after the diesel injection is ended, causes increased
dicates the start of diesel injection.
ammonia slip. Based on this, for combusting ammonia properly, there
has to be an overlap of diesel flame and ammonia injection. Otherwise,
of ammonia, i.e. 72.5 % for -80 CAD aTDC injection timing, and grad­ the ammonia injection quenches the diesel post-flame. This is consistent
ually increases as the injection is delayed. After -30 CAD aTDC, the with previous research on the topic [21] and [24], where overlapping
combustion efficiency levels off and slightly increases when approach­ injection timings and interacting sprays yielded the highest burn rates.
ing simultaneous injections at -15 CAD aTDc. Around -15 CAD aTDC, a Interestingly, the ammonia injection timing at -30 CAD aTDC led to
local maximum is reached. This shows that the early injections of similar ammonia slips, which could indicate that if the injection dura­
ammonia, where ammonia is allowed to mix and vaporise before the tion was kept constant at 1.06 ms (i.e. same as for AES40) for all AESs by
diesel injection, are not good for a proper burnout of ammonia in this increasing the mass flow rate, a similar ammonia slip ratio curve could
configuration. possibly be achieved for AES50 and AES60.

11
K.O.P. Bjørgen et al. Fuel 360 (2024) 130269

Fig. 12. The concentration of ammonia in the exhaust stream and the ammonia slip ratio for AES of 0 %, 40 %, 50 % and 60 %, with fixed diesel injection timing at
-15 CAD aTDC, as a function of ammonia injection timing. The ammonia slip ratio indicates the share of ammonia which is unburned and emitted via the exhaust.
The vertical dashed line indicates the start of diesel injection.

Comparing the results from ammonia and diesel combustion in HPDF


mode with previous studies using RCCI mode, the emission trends are
different when considering unburned ammonia. Previous studies on
RCCI mode [18,19,20] all found that for increasing AESs the unburned
ammonia monotonically increased while keeping the diesel injection
timing fixed. For the current study, the ammonia emissions did not in­
crease significantly when increasing the AES from 40 % to 50 %, for
ammonia injections between -30 and -20 CAD aTDC and for diesel in­
jections at -15 CAD aTDC. However, when increasing to AES of 60 %, the
ammonia emissions increased significantly. For RCCI mode, the increase
in ammonia emissions can partly be explained by trapped ammonia in
crevices, which is avoided for HPDF mode. However, other mechanisms
such as spray-to-spray interaction and the relative injection timing of
ammonia and diesel play important roles in HPDF mode.
The concentrations of NOx, NO and N2O in the engine exhaust stream Fig. 14. The concentration of N2O in the exhaust stream, for AES of 0 %, 40 %,
are shown in Fig. 13 and 14, respectively. The NOx concentration is the 50 % and 60 %, with fixed diesel injection timing at -15 CAD aTDC, as a
sum of the NO and NO2 concentrations, where the NO concentration is function of ammonia injection timing. The vertical dashed line indicates the
also shown independently in the figure to indicate its fraction of the total start of diesel injection.
NOx. It is apparent that the NOx and N2O concentrations have a some­
what opposite trend. Generally, where NOx is high, N2O is low. For the
AES40 case, the NOx concentration is highest for ammonia injections at
-20 and -17.5 CAD aTDC, i.e. 5.0 and 2.5 CAD before the diesel injection.
For AES50 and AES60, the peak is reached slightly before, i.e., for -25
CAD aTDC, showing similar behaviour to AES40, although AES60 has
too few operating points to conclude this.

Fig. 15. The share of NO2 in NOx, for AES of 0 %, 40 %, 50 % and 60 %, with
fixed diesel injection timing at -15 CAD aTDC, as a function of ammonia in­
jection timing. The vertical dashed line indicates the start of diesel injection.

Based on the NOx and NO concentrations, the share of NO2 is


consistently low for the ammonia injection timings with high NOx for
AES40 and AES50, i.e. 6–9 % NO2, as shown in Fig. 15. Delaying the
Fig. 13. The NOx (○) and NO (□) concentrations measured in the engine ammonia injection, the NO2 share gradually increases for AES40 and
exhaust stream, for AES of 0 %, 40 %, 50 % and 60 %, with fixed diesel injection AES50. While for AES60, the NO2 share is 22 % for -30 and -25 CAD
timing at -15 CAD aTDC, as a function of ammonia injection timing. The ver­ aTDC and decreases to 13 % for -20 CAD aTDC.
tical dashed line indicates the start of diesel injection.

12
K.O.P. Bjørgen et al. Fuel 360 (2024) 130269

The formation of thermal NO from N2 is very temperature dependent


and is well described by the Zeldovich mechanism [36]. For ammonia
combustion, the fuel-bound nitrogen also contributes to the NO forma­
tion when ammonia is oxidised. The high NOx emissions follow the
minimum ammonia slip ratio points of the different AESs, which can be
explained by the high conversion rate of ammonia and the corre­
sponding higher temperatures achieved for these points. As NO is
formed, some are oxidised to NO2 in the flame zone but generally con­
verted back to NO in a post-flame region with low oxygen concentration
via NO2 + O → NO + O2 [37]. If the temperature in the post-flame region
is low, the reduction of NO2 is “frozen”, leading to higher NO2 emissions
[11]. From Fig. 15 it can therefore be assumed that the higher NO2 share
for the late injections is caused by lower temperatures, likely due to the
poorer combustion of ammonia. Fig. 16. The concentration of CO in the exhaust stream, for AES of 0 %, 40 %,
The formation of N2O is also affected by quenching effects in the 50 % and 60 %, with fixed diesel injection timing at -15 CAD aTDC, as a
expansion stroke or at the cold engine walls. N2O is mainly formed via function of ammonia injection timing. The vertical dashed line indicates the
two reactions involving the combination of NH and NO at high tem­ start of diesel injection.
peratures and NH2 and NO2 at low temperatures [13]. At high temper­
atures, N2O will be reduced either from thermal dissociation or reacting increasing AES, the CO concentration is increased, indicating that the
with atomic H, forming N2. In this case, the low-temperature route is combustion of diesel is poorer for higher AESs. For AES40, the CO
likely the main cause of N2O emissions. The presence of NO2 and NH2 concentration is lower for ammonia injection timings 2.5 CAD before the
from unburned ammonia in the expansion stroke with low temperatures diesel injection compared to the other timings, indicating that this
are prerequisites for the engine out emission of N2O, leading to a operating point yields more complete combustion, as seen for the other
competition between NO2 and N2O. For ammonia injections at the peak measured quantities for this point. For AES50 and AES60, the CO con­
NOx point for AES40 and AES50, the N2O concentration is at its mini­ centration does not change much when changing the ammonia injection
mum, which can be explained by the high temperature achieved due to timing.
the high combustion efficiency. When delaying the ammonia injection
to -12.5 CAD aTDC, i.e. 2.5 CAD after the diesel injection, the N2O
concentration reaches a local maximum for AES40 and AES50. Further 3.3. Effect of combustion phasing
delaying the ammonia injection causes a slightly lower N2O concen­
tration before it again increases for late injections of ammonia. Based on 3.3.1. Combustion characteristics
the discussion above, the local maximum of N2O at -12.5 CAD aTDC is The in-cylinder pressures for varying combustion phasing are shown
likely due to a decrease in temperature in the expansion stroke due to in Fig. 17. For AES40, the injection of ammonia starts simultaneously
poor burnout of ammonia. This is also observed in combustion efficiency with diesel, while for AES50 and AES60, ammonia starts 5 CAD (0.56
and ammonia slip ratio Figs. 11 and 12b, respectively. Since this occurs ms) before the diesel injection. A comparison between AES40 and AES50
at the same ammonia injection timings for AES40 and AES50, it is provides insight into the effect of injecting ammonia 5 CAD before diesel
hypothesised to be caused by the lack of overlapping ammonia and (AES50). While comparing AES50 and AES60 gives insight into the ef­
diesel injections. fect of injecting ammonia 0.5 ms after the end of diesel injection. The
The maximum NOx and minimum N2O emissions both occur for change in the total mass of injected ammonia for these comparisons
ammonia injection timings slightly before the diesel injection, as seen in must naturally be taken into account.
Figs. 13 and 13. This could be explained by the physical distance of In Fig. 18, the effect of combustion phasing on the HRR is shown for
approximately 15 mm between the ammonia and diesel injector in the AES40, AES50 and AES60. For AES40, the HRR curve does not consid­
engine (see Fig. 1). For the case of simultaneous ammonia and diesel erably change its shape, but the first HRR peak during the premixed
injections at -15 CAD aTDC for AES40, where the injection durations are phase decreases slightly when delaying the injections. This is due to a
nearly the same, the diesel spray gets a head start since the diesel shortening in IDT since the fuel is injected into hotter air closer to the top
injector is located downstream of the ammonia spray. In preliminary dead centre (TDC), as seen in Fig. 19. Consequently, more fuel burns
ammonia spray measurements for similar ambient gas densities, the during the mixing-controlled phase for the delayed combustion phasing.
ammonia gas penetration length reaches 15 mm in approximately 0.25 For AES50, the shape of the HRR curve is more affected by delayed in­
ms, corresponding to 2.25 CAD in the engine. Likewise, the ammonia jection timings. This is due to ammonia being injected 5 CAD (0.56 ms)
injection will have a 15 mm non-overlapping part at the end of injection before diesel, where the cooling effect from ammonia vaporisation in­
since the diesel injection is located 15 mm downstream. By injecting fluences the start of combustion for AES50 and AES60, causing a 0.10 ms
ammonia 2.5 CAD aTDC before the diesel injection, the spatial overlap longer IDT compared to AES40. As the injections are delayed, the pre­
between the two sprays increases, potentially leading to better com­ mixed phase of the HRR curve becomes more pronounced, and the HRR
bustion, and thereby higher NOx and lower N2O concentrations. peak becomes lower. Since the IDT decreases for delayed combustion
Lewandowski et al. [24] also observed an increase in NO concen­ phasing, the latest injections have less time for ammonia to vaporise and
tration when delaying the ammonia injection towards the diesel injec­ mix with air before ignition. This means that the distinct premixed
tion while simultaneously observing a decrease in N2O and ammonia. combustion phase is mainly caused by the combustion of diesel, while
Also, comparing the NOx emissions with the study of Zhang et al. [23], for the earlier combustion phasing, the initial combustion phase is a
they kept the diesel injection timing fixed at -8 CAD aTDC while varying combination of combustion of diesel and ammonia in premixed mode.
the ammonia injection timing between -16, -8 and 0 CAD aTDC. They The same is observed for AES60.
observed a similar trend as in Fig. 13, where the NOx was the highest for The effect of combustion phasing on the combustion durations is
simultaneous injections of ammonia and diesel. They argue that late shown in Fig. 20. For AES40, the IDT-CA50 duration is near constant for
injection of ammonia could lead to deNOx reactions, reducing NOx all injection timings, while the IDT-CA50 duration for AES50 and AES60
during the expansion stroke, which could also be the case in the current increases when delaying the combustion phasing. The increased com­
study. bustion duration for IDT-CA50 is caused by the transition from the
The CO concentrations are shown in Fig. 16, showing that for premixed combustion during the initial phase of ammonia and diesel

13
K.O.P. Bjørgen et al. Fuel 360 (2024) 130269

Fig. 17. The ensemble-averaged in-cylinder pressures for varying combustion phasing for AESs of 40 %, 50 % and 60 %. The horizontal red and blue lines indicate
the diesel and ammonia energizing times, respectively. Note that the start of ammonia injection occurs simultaneously as diesel for AES40 and 5 CAD before diesel for
AES50 and AES60.

Fig. 18. The ensemble-averaged heat release rate curves for varying combustion phasing for AESs of 40 %, 50 % and 60 %. The horizontal red and blue lines indicate
the diesel and ammonia energizing times, respectively. Note that the start of ammonia injection occurs simultaneously as diesel for AES40 and 5 CAD before diesel for
AES50 and AES60.

due to longer ignition delay times, to premixed combustion of mainly


diesel for shorter ignition delay times, as discussed above. The IDT-CA90
duration shows a clear difference between AES40/AES50 and AES60,
where the combustion duration increases linearly for AES60 when
delaying the combustion phasing. This suggests that the injected
ammonia after the end of diesel injection prolongs the overall combus­
tion duration, indicating that the late injected ammonia combusts at a
slower rate due to the resulting lower temperatures in the expansion
stroke. As mentioned, AES40 and AES50 do not have ammonia injected
after the end of diesel injection, confirming that the late injected
ammonia is the cause of the longer IDT-CA90 duration.
For AES50, the combustion efficiency, shown in Fig. 21, increases
from 80 % to 86 % when delaying the diesel timing from -20 to -15 CAD
aTDC, but levels off when further delaying it. The benefit of injecting
closer to TDC is that more of the spray injection period, ignition and
Fig. 19. The ignition delay time for varying combustion phasing for AESs of combustion of the fuel mixture occur under higher ambient gas tem­
0 %, 40 %, 50 % and 60 %. The combustion phasing is indicated as the diesel perature conditions, enhancing the combustion process. For AES60, the
injection timing along the x-axis. Note that the start of ammonia injection oc­ combustion efficiency does not change significantly between -20 and
curs simultaneously as diesel for AES40 and 5 CAD before diesel for AES50 -17.5 CAD aTDC diesel injection timing, while for -15 CAD aTDC, it is
and AES60. reduced by 2 % points, which could be explained by the

14
K.O.P. Bjørgen et al. Fuel 360 (2024) 130269

Fig. 22. The concentration of ammonia in the exhaust stream for varying
combustion phasing for AESs of 0 %, 40 %, 50 % and 60 %. The combustion
phasing is indicated as the diesel injection timing along the x-axis. Note that the
start of ammonia injection occurs simultaneously as diesel for AES40 and 5 CAD
before diesel for AES50 and AES60.

Fig. 20. The combustion durations for 10 %, 50% and 90 % heat released (IDT-
CA10 (○), IDT-CA50 (□) and IDT-CA90 (⋄)) for varying combustion phasing for
AESs of 0 %, 40 %, 50 % and 60 %. The combustion phasing is indicated as the
diesel injection timing along the x-axis. Note that the start of ammonia injection
occurs simultaneously as diesel for AES40 and 5 CAD before diesel for AES50
and AES60.

Fig. 23. The concentration of NOx (○) and NO (□) in the exhaust stream for
varying combustion phasing for AESs of 0 %, 40 %, 50 % and 60 %. The
combustion phasing is indicated as the diesel injection timing along the x-axis.
Note that the start of ammonia injection occurs simultaneously as diesel for
AES40 and 5 CAD before diesel for AES50 and AES60.

Fig. 24. The concentration of N2O in the exhaust stream for varying combus­
Fig. 21. The combustion (○) and thermal (□) efficiencies for varying com­ tion phasing for AESs of 0 %, 40 %, 50 % and 60 %. The combustion phasing is
bustion phasing for AESs of 0 %, 40 %, 50 % and 60 %. The combustion phasing indicated as the diesel injection timing along the x-axis. Note that the start of
is indicated as the diesel injection timing along the x-axis. Note that the start of ammonia injection occurs simultaneously as diesel for AES40 and 5 CAD before
ammonia injection occurs simultaneously as diesel for AES40 and 5 CAD before diesel for AES50 and AES60.
diesel for AES50 and AES60.
3.4. Emission characteristics
abovementioned late injections of ammonia causing the combustion to
extend far into the expansion stroke. The indicated thermal efficiency The effect of combustion phasing on the emission concentrations of
increases when delaying the combustion phasing since more of the heat ammonia, NOx and N2O in the exhaust stream is shown in Fig. 22, 23 and
release occurs after TDC. 24, respectively. The remaining emissions gases did not significantly
change when varying the combustion phasing. The ammonia concen­
trations for AES40 and AES50 marginally increase when advancing the
combustion phasing. For AES60, the ammonia concentration increases

15
K.O.P. Bjørgen et al. Fuel 360 (2024) 130269

linearly from 10,000 ppm to 22,000 ppm when delaying the combustion reported a monotonically decreasing trend of GHG specific emissions for
phasing by 2.5 and 5 CAD from diesel injection timing at -20 CAD aTDC, increasing AES for combustion of ammonia and diesel in RCCI mode.
which suggests that the ammonia injected after the end of diesel injec­ The minimum GHG specific emissions achieved within each AES in the
tion causes a poor burn out of ammonia, also supported by the above current study do not have a monotonically decreasing trend with
discussion on the increased combustion duration. The NOx concentra­ increasing AES. This could be due to more degrees of freedom for the
tion decreases, and the N2O concentration increases when delaying the HPDF mode compared to the RCCI mode, which therefore required more
combustion phasing. optimisation. The GHG specific emissions are clearly affected by the
The effect of combustion phasing shows that further optimisation is AES, where higher AES reduces the CO2-related GHG emissions. Oper­
possible, where the engine performance can be increased while reducing ating the engine with higher AES is planned for future experiments,
emissions. The combustion and indicated thermal efficiencies are where the engine load will be kept constant while increasing the AES.
increased by delaying the combustion phasing. However, the N2O The current experimental campaign is an initial step in exploring
emissions are simultaneously increased, which is the hardest emission to ammonia/diesel in HPDF mode for the current configuration, requiring
aftertreat. Therefore, in this configuration, the ideal would be to further optimisation.
advance the combustion phasing, leading to lower N2O emissions at the
expense of lower engine performance and higher NOx emissions. 4. Conclusions

3.5. Greenhouse gas emissions This study has systematically investigated the combustion and
emissions characteristics of direct injection of ammonia and diesel in a
The main objective of using ammonia as fuel in IC engines is to CI engine. The focus was to study the effect of the injection timing of
reduce GHG emissions compared to carbon-based fuels. The sum of the ammonia and diesel, where the ammonia injection timing was varied
specific emissions of CO2 and 265 times N2O, representing the GHG between -80 CAD aTDC and 2.5 CAD aTDC while holding the diesel
specific emissions in g/kWh, are presented in Fig. 25. For comparison, injection timing fixed at -15 CAD aTDC. The effect of combustion
the diesel-only case is also presented. The overall lowest specific GHG phasing with fixed relative injection timings for each ammonia energy
emissions are achieved for AES60 with injections of ammonia at -30 CAD share was also studied. The following conclusions can be made based on
aTDC. For AES40 and AES50, ammonia injections at -17.5 and -20 CAD the results presented:
aTDC result in the minimum GHG specific emissions. N2O related GHG
specific emissions are a major part of the total GHG specific emissions, • The combustion mode varies significantly when varying the
where the highest share of 49 % is found for the very early injection of ammonia injection timing for fixed diesel timing at -15 CAD aTDC.
ammonia at -80 CAD aTDC for AES40, while the lowest share of 20 % is For very early ammonia injections between -80 and -60 CAD aTDC,
found for ammonia injections at -17.5 CAD aTDC for AES40. This result the ammonia premixes with the air, leading to premixed-type com­
emphasises the importance of minimising N2O emissions from ammonia bustion with low HRR, where ammonia is likely distributed to re­
combustion. N2O should be accurately measured in experimental work gions in the cylinder/piston where the diesel flame does not reach,
and in industrial applications, where the aim is to replace carbon-based leading to poor overall combustion, very high ammonia slip (65 %
fuels with ammonia in order to reduce GHG emissions. The largest for -80 CAD aTDC) and high N2O emissions.
reduction in GHG specific emissions in this study is 23 % for AES60 • For later ammonia injections approaching -25 CAD aTDC (i.e. 10
compared to the diesel-only case, which is far from the potential CAD before diesel), the ignition delay time reaches its maximum for
reduction of 55 % if the emissions of N2O are fully abated. More work is all ammonia energy shares tested. The long ignition delay time is
needed to understand the mechanisms behind N2O production and explained by a large amount of cold ammonia vapour in the region
consumption in an engine operating in HPDF mode. Comparing the where diesel is injected. Thereby, premixed-type combustion was
current GHG specific emissions results with previous work on ammonia/ achieved for this operating point.
diesel in RCCI mode shows many similarities. Yousefi et al. [18] found • Injecting ammonia simultaneously or after diesel does not affect the
that the minimum GHG specific emissions for 40 % AES were found for ignition delay time since the initial diesel spray remains undisturbed.
earlier injections of diesel, resulting in 606.2 g/kWh. Nadimi et al. [20] Injecting ammonia into the diesel flame thereby results in the com­
bustion of ammonia in a mixing-controlled mode, leading to a more
stable combustion.
• Temporally overlapping injections of ammonia yield the highest
combustion efficiency, lowest ammonia slip, highest NOx emissions,
and lowest N2O emissions. For temporally overlapping injections,
the ammonia slip ratio was found to be higher for ammonia in­
jections continuing after the end of diesel injection. For delayed
combustion phasing, injecting ammonia after the end of diesel in­
jection led to an increase in the ammonia slip ratio, which is
explained by ammonia requiring sufficient time to vaporise, mix and
combust the ammonia in the expansion stroke. On the contrary,
delaying the combustion phasing without injecting ammonia after
the end of diesel injection did not affect the ammonia slip ratio.
• The ammonia slip ratio was found to be similar when injecting
ammonia at -30 CAD aTDC (i.e. 15 CAD before the diesel injection)
for all tested ammonia energy shares and similar for AES40 and
AES50 between -30 and -17.5 CAD aTDC. A sudden increase in the
ammonia slip ratio was observed for later injections of ammonia.
• The N2O concentration reaches a minimum for ammonia injections
Fig. 25. The CO2-equivalent GHG specific emissions based on the sum of the 2.5 and 5.0 CAD before the diesel injection for AES40 and AES50.
CO2 and N2O (265 times) specific emissions, for AES of 0 %, 40 %, 50 % and 60 However, an increase was observed for ammonia injections 2.5 CAD
% and all tested ammonia injection timings with diesel injection timing at -15 after the diesel injection, correlating inversely with the trend of
CAD aTDC. The horizontal dashed line indicates the diesel-only level. combustion efficiency.

16
K.O.P. Bjørgen et al. Fuel 360 (2024) 130269

• Combustion phasing affected the indicated thermal and combustion system. Fuel 2011;90:854–64. http://www.sciencedirect.com/science/article/pii/
S0016236110005132. https://doi.org/10.1016/j.fuel.2010.09.042.
efficiencies and NOx/N2O emissions, where delaying the start of
[11] Westlye FR, Ivarsson A, Schramm J. Experimental investigation of nitrogen based
combustion yielded higher engine performance but simultaneously emissions from an ammonia fueled SI-engine. Fuel 2013;111:239–47. https://doi.
increased the N2O emissions due to late injection of ammonia. org/10.1016/j.fuel.2013.03.055. http://www.sciencedirect.com/science/article/
pii/S0016236113002433.
[12] Myhre G, Shindell D, Bŕeon F-M, Collins W, Fuglestvedt J, Huang J, et al.
The optimal injection strategy for this configuration is found to be Anthropogenic and Natural Radiative Forcing, Technical Report. Cambridge,
injecting ammonia and diesel temporally close to each other, and to United Kingdom and New York, NY, USA: Cambridge University Press; 2013.
avoid injecting ammonia after the diesel flame. The conclusions found in [13] Glarborg P. The NH3/NO2/O2 system: Constraining key steps in ammonia ignition
and N2O formation, Combustion and Flame (2022) 112311. https://www.scienc
this study are highly dependent on the spray orientation, and further edirect.com/science/article/pii/S0010218022003261. https://doi.org/10.1016/j.
investigation into the effect of ammonia and diesel spray interaction is combustflame.2022.112311.
planned for future work. [14] Gross CW, Kong S-C. Performance characteristics of a compression-ignition engine
using direct-injection ammonia–DME mixtures. Fuel 2013;103:1069–79. https://
doi.org/10.1016/j.fuel.2012.08.026. http://www.sciencedirect.com/science/artic
CRediT authorship contribution statement le/pii/S001623611200662X.
[15] Reiter AJ, Kong S-C. Combustion and emissions characteristics of compression-
ignition engine using dual ammonia-diesel fuel. Fuel 2011;90:87–97. http://www.
Karl Oskar Pires Bjørgen: . David Robert Emberson: Writing – sciencedirect.com/science/article/pii/S001623611000414X. https://doi.org/
review & editing, Methodology, Funding acquisition, Conceptualization. 10.1016/j.fuel.2010.07.055.
Terese Løvås: Writing – review & editing, Resources, Project adminis­ [16] Gill SS, Chatha GS, Tsolakis A, Golunski SE, York APE. Assessing the effects of
partially decarbonising a diesel engine by co-fuelling with dissociated ammonia.
tration, Funding acquisition, Conceptualization. Int J Hydrogen Energy 2012;37:6074–83. https://doi.org/10.1016/j.
ijhydene.2011.12.137. http://www.sciencedirect.com/science/article/pii/
S0360319911028710.
Declaration of competing interest
[17] Niki Y, Nitta Y, Sekiguchi H, Hirata K. Emission and Combustion Characteristics of
Diesel Engine Fumigated With Ammonia. In: ASME 2018 Internal Combustion
The authors declare that they have no known competing financial Engine Division Fall Technical Conference. American Society of Mechanical
interests or personal relationships that could have appeared to influence Engineers Digital Collection; 2019. https://doi.org/10.1115/ICEF2018-9634.
[18] Yousefi A, Guo H, Dev S, Liko B, Lafrance S. Effects of ammonia energy fraction and
the work reported in this paper. diesel injection timing on combustion and emissions of an ammonia/diesel dual-
fuel engine. Fuel 2022;314:122723. https://doi.org/10.1016/j.fuel.2021.122723.
Data availability https://www.sciencedirect.com/science/article/pii/S0016236121025886.
[19] Førby N, Thomsen TB, Cordtz RF, Bræstrup F, Schramm J. Ignition and combustion
study of premixed ammonia using GDI pilot injection in CI engine. Fuel 2023;331:
Data will be made available on request. 125768. https://www.sciencedirect.com/science/article/pii/S0016236
122025947. https://doi.org/10.1016/j.fuel.2022.125768.
[20] Nadimi E, Przybyla G, Lewandowski MT, Adamczyk W. Effects of ammonia on
Acknowledgement combustion, emissions, and performance of the ammonia/diesel dual-fuel
compression ignition engine. J Energy Inst 2023;107:101158. https://doi.org/
This work was funded by the EEA and Norway Grants via the Polish- 10.1016/j.joei.2022.101158. https://www.sciencedirect.com/science/article/pii/
S1743967122002069.
Norwegian research project ACTIVATE [NOR/POLNOR/ACTIVATE/ [21] Scharl V, Sattelmayer T. Ignition and combustion characteristics of diesel piloted
0046/2019-00]. ammonia injections, Fuel. Communications 2022;11:100068. https://www.scienc
edirect.com/science/article/pii/S2666052022000188. https://doi.org/10.1016/j.
jfueco.2022.100068.
Appendix A. Supplementary data [22] Scharl V, Lackovic T, Sattelmayer T. Characterization of ammonia spray
combustion and mixture formation under high-pressure, direct injection
Supplementary data to this article can be found online at https://doi. conditions. Fuel 2023;333:126454. https://doi.org/10.1016/j.fuel.2022.126454.
https://www.sciencedirect.com/science/article/pii/S0016236122032781.
org/10.1016/j.fuel.2023.130269. [23] Zhang K, Shen Y, Palulli R, Ghobadian A, Nouri J, Duwig C. Combustion
characteristics of steam-diluted decomposed ammonia in multiple-nozzle direct
References injection burner. Int J Hydrogen Energy 2023. https://doi.org/10.1016/j.
ijhydene.2023.01.091. https://www.sciencedirect.com/science/article/pii/
S0360319923000927.
[1] Valera-Medina A, Amer-Hatem F, Azad AK, Dedoussi IC, de Joan-non M,
[24] Lewandowski MT, Pasternak M, Haugsvær M, Løvås T. Simulations of ammonia
Fernandes RX, et al. Review on Ammonia as a Potential Fuel: From Synthesis to
spray evaporation, cooling, mixture formation and combustion in a direct injection
Economics. In: Energy Fuels. 35. publisher: American Chemical Society; 2021.
compression ignition engine. Int J Hydrogen Energy 2023. https://doi.org/
p. 6964–7029.
10.1016/j.ijhydene.2023.06.143. https://www.sciencedirect.com/science/article/
[2] Kroch E. Ammonia - A fuel for motor buses. J Inst Pet 1945;31:213–23.
pii/S0360319923030227.
[3] Cornelius W, Huellmantel LW, Mitchell HR. Ammonia as an Engine Fuel, 1965, p.
[25] Wyer KE, Kelleghan DB, Blanes-Vidal V, Schauberger G, Curran TP. Ammonia
650052. https://www.sae.org/content/650052/. doi:10.4271/650052.
emissions from agriculture and their contribu- tion to fine particulate matter: A
[4] Garabedian CG, Johnson JH. The theory of operation of an ammonia burning
review of implications for human health. J Environ Manage 2022;323:116285.
internal combustion engine, Technical Report, army tank-automotive center
https://www.sciencedirect.com/science/article/pii/S0301479722018588.
warren MI, 1966. https://apps.dtic.mil/sti/citations/AD0634681, section:
https://doi.org/10.1016/j.jenvman.2022.116285.
Technical Reports.
[26] Park Y-K, Kim B-S. Catalytic removal of nitrogen oxides (NO, NO2, N2O) from
[5] Starkman ES, Newhall HK, Sutton R, Maguire T, Farbar L. Ammonia as a Spark
ammonia-fueled combustion exhaust: A review of applicable technologies. Chem
Ignition Engine Fuel. Theory Appl 1966:660155. https://www.sae.org/conten
Eng J 2023;461:141958. https://doi.org/10.1016/j.cej.2023.141958. https://
t/660155/. https://doi.org/10.4271/660155.
www.sciencedirect.com/science/article/pii/S1385894723006897.
[6] Gray JT, Dimitroff E, Meckel NT, Quillian RD. Ammonia Fuel - Engine
[27] Macian V, Bermudez V, Payri R, Gimeno J. New Technique for Determination of
Compatibility and Combustion, 1966, p. 660156. https://www.sae.org/conten
Internal Geometry of a Diesel Nozzle with the Use of Silicone Methodology,
t/660156/. doi:10.4271/660156.
Experimental Techniques 27 (2003) 39–43. http://doi.wiley.com/10.1111/j.1
[7] Sawyer RF, Starkman ES, Muzio L, Schmidt WL. Oxides of Nitrogen in the
747-1567.2003.tb00107.x. https://doi.org/10.1111/j.1747-1567.2003.tb00107.x.
Combustion Products of an Ammonia Fueled Reciprocating Engine, 1968, p.
[28] Mounaim-Rousselle C, Mercier A, Brequigny P, Dumand C, Bouriot J, Houillé S.
680401. https://www.sae.org/content/680401/. doi:10.4271/680401.
Performance of ammonia fuel in a spark assisted compression Ignition engine. In:
[8] Pearsall TJ, Garabedian CG. Combustion of Anhydrous Ammonia in Diesel Engines
Int J Engine Res. 14680874211038726. publisher: SAGE Publications; 2021.
1967;670947. https://doi.org/10.4271/670947. https://www.sae.org/conten
https://doi.org/10.1177/14680874211038726.
t/670947/.
[29] Haar L, Gallagher JS. Thermodynamic properties of ammonia. URL: J Phys Chem
[9] Grannell SM, Assanis DN, Bohac SV, Gillespie DE. The Fuel Mix Limits and
Ref Data 1978;7:635–792. https://doi.org/10.1063/1.555579. https://pubs.aip.
Efficiency of a Stoichiometric, Ammonia, and Gasoline Dual Fueled Spark Ignition
org/aip/jpr/article/7/3/635-792/242263.
Engine. J Eng Gas Turbines Power 130 (2008). https://asmedigitalcollection.asme.
[30] Fenghour A, Wakeham WA, Vesovic V, Watson JTR, Millat J, Vogel E. The
org/gasturbinespower/article/130/4/042802/476304/. https://doi.org/10.1115/
Viscosity of Ammonia. J Phys Chem Ref Data 1995;24:1649–67. https://doi.org/
1.2898837, publisher: American Society of Mechanical Engineers Digital
10.1063/1.555961. URL: https://pubs.aip.org/aip/jpr/article/24/5/1649-1667/
Collection.
241712.
[10] Mørch CS, Bjerre A, Gøttrup MP, Sorenson SC, Schramm J. Ammonia/hydrogen
mixtures in an SI-engine: Engine performance and analysis of a proposed fuel

17
K.O.P. Bjørgen et al. Fuel 360 (2024) 130269

[31] Chong CT, Hochgreb S. Measurements of laminar flame speeds of liquid fuels: Jet- [35] Maurya RK. Reciprocating Engine Combustion Diagnostics: In-Cylinder Pressure
A1, diesel, palm methyl esters and blends using particle imaging velocimetry (PIV). Measurement and Analysis, Mechanical Engineering Series. Cham: Springer
Proc Combust Inst 2011;33:979–86. https://doi.org/10.1016/j.proci.2010.05.106. International Publishing; 2019. https://doi.org/10.1007/978-3-030-11954-6.
https://www.sciencedirect.com/science/article/pii/S1540748910002695. http://link.springer.com/10.1007/978-3-030-11954-6.
[32] Lhuillier C, Brequigny P, Lamoureux N, Contino F, C. Mounaim-Rousselle,. [36] Zeldovich. The Oxidation of Nitrogen in Combustion and Explosions. Acta
Experimental investigation on laminar burning velocities of ammonia/hydrogen/ Physicochimica 1946;21:577–628.
air mixtures at elevated temperatures. URL: Fuel 2020;263:116653. https://doi. [37] Merryman EL, Levy A. Nitrogen oxide formation in flames: The roles of NO2 and
org/10.1016/j.fuel.2019.116653. https://www.sciencedirect.com/science/artic fuel nitrogen. Symp (Int) Combust 1975;15:1073–83. https://www.sciencedirect.
le/pii/S0016236119320071. com/science/article/pii/S0082078475803729. https://doi.org/10.1016/S0082-
[33] Heywood JB. Internal Combustion Engine Fundamentals. McGraw-Hill; 1988. 0784(75)80372-9.
[34] Woschni G. A Universally Applicable Equation for the Instantaneous Heat Transfer
Coefficient in the Internal Combustion Engine, in: SAE Technical Paper Series,
volume No. 670931, 1967.

18

You might also like