Download as pdf or txt
Download as pdf or txt
You are on page 1of 20

Fuel 367 (2024) 131450

Contents lists available at ScienceDirect

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

Full Length Article

Effect of liquid ammonia HPDI strategies on combustion characteristics and


emission formation of ammonia-diesel dual-fuel heavy-duty engines
Zheng Zhang a, Liming Di a, b, *, Lei Shi c, Xiyu Yang a, Tengfei Cheng a, Cheng Shi a, *
a
School of Vehicle and Energy, Yanshan University, Qinhuangdao 066004, China
b
Hebei Key Laboratory of Special Carrier Equipment, Qinhuangdao 066004, China
c
Weichai Power Co., Ltd., Weifang 261061, China

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

Keywords: Reducing carbon emissions from internal combustion engines has become a crucial topic due to the pressure
Ammonia brought about by global warming. Ammonia, a carbon-free fuel, has significant potential for heavy-duty diesel
Diesel engine applications. However, the laminar flame speed of ammonia fuel is relatively slow, resulting in subop­
Injection timing
timal combustion performance. Therefore, the present study focuses on ammonia-diesel dual-fuel heavy-duty
Injection direction
engines. A novel liquid ammonia high-pressure direct injection (HPDI) strategy has been developed to address
Heavy-duty engine
the challenges associated with poor combustion performance and increased unburned ammonia emissions in
ammonia-diesel dual-fuel internal combustion engines. Firstly, a three-dimensional numerical simulation method
was utilized to establish and verify a model of a heavy-duty internal combustion engine that incorporates HPDI of
liquid ammonia. Afterward, numerical studies comparing the engine’s combustion and emission characteristics
are conducted with varied ammonia energy fractions, liquid ammonia injection timings, and liquid ammonia
injection directions. The results indicate that, compared to the pure diesel mode, the novel liquid ammonia HPDI
strategy can alter the combustion mode, significantly improving fuel–air mixture efficiency. This leads to more
thorough combustion. After replacing 80% of the fuel energy input with ammonia, there is still an increase of
8.9% in indicated mean effective pressure and a 10.6% improvement in indicated thermal efficiency. Moreover,
modifying the HPDI strategy reduces greenhouse gas emissions effectively, and due to the thermal de-NOx re­
action, there is a notable decrease in NOx emissions. Furthermore, by altering the liquid ammonia injection
timing and direction, finer control of fuel combustion and pollutant generation can be achieved, thereby
reducing unburned ammonia emissions in conditions of high ammonia energy fractions.

them, ammonia stands out, as its complete combustion results only in


nitrogen and water [6]. Ammonia is convenient to store, with a liquid
1. Introduction form energy density 1.5 times higher than that of hydrogen [7].
Furthermore, heavy-duty internal combustion engines can adapt to
In response to the demand for carbon neutrality following the ammonia fuel without significant adjustments [8]. Therefore, ammonia
enactment of the Paris Climate Agreement., countries worldwide are is considered one of the most promising carbon-free fuels for successful
actively pursuing ways to decrease carbon emissions [1]. As a major application in heavy-duty internal combustion engines in this century.
component of power output devices, heavy-duty Compression Ignition In recent decades, there has been significant research on the use of
(CI) diesel engines generate significant amounts of soot and NOX during ammonia fuel in CI diesel engines. Due to the slow flame propagation
operation. In recent years, numerous advanced combustion control speed of gaseous ammonia [9], achieving a compression ratio of 35:1 is
methods have been researched, which have effectively lowered soot and required for the operation of a CI diesel engine with pure ammonia fuel
NOX emissions from diesel engines [2,3]. However, due to the utilization [10]. Therefore, when using ammonia as an alternative fuel in CI diesel
of fossil fuels, diesel engines still emit significant quantities of CO2 and engines, diesel is commonly employed as a co-reactant to ignite the
other greenhouse gases [4], significantly impacting the present and ammonia [11]. In 1966, Gray et al. [12] conducted a study on the
future applications of diesel engines. One crucial method to reduce application of ammonia fuel in diesel engines. They demonstrated that
greenhouse gas emissions is to explore alternative fuels [5]. Among

* Corresponding authors.
E-mail addresses: diliming@ysu.edu.cn (L. Di), shicheng@ysu.edu.cn (C. Shi).

https://doi.org/10.1016/j.fuel.2024.131450
Received 25 December 2023; Received in revised form 17 February 2024; Accepted 10 March 2024
Available online 12 March 2024
0016-2361/© 2024 Elsevier Ltd. All rights reserved.
Z. Zhang et al. Fuel 367 (2024) 131450

engines. This results in an ITE of 39.7 %, exceeding that of pure diesel


Nomenclature operation. At the same time, it reduces unburned ammonia emissions by
85.3 % and greenhouse gas emissions by 30.6 %.
AMR Adaptive mesh refinement It can be seen that in the ammonia-diesel dual-fuel internal com­
ATDC After top dead center bustion engines in the above studies, ammonia enters the combustion
BMEP Brake mean effective pressure chamber through the intake manifold in a premixed manner, which is
CA Crank angle referred to as Low-Pressure Injection (LPI). Although this LPI method
EAIT End of ammonia injection timing requires minimal modification of the existing engine, the low injection
EDIT End of diesel injection timing pressure and gaseous nature of ammonia present notable challenges.
IMEP Indicated mean effective pressure This results in a decrease in both engine power and efficiency compared
HPDI High-pressure direct injection with the traditional pure diesel mode [20], accompanied by an increase
ITE Indicated thermal efficiency in NOx and unburned emissions [21]. Compared to the LPI method, the
LPI Low-pressure injection HPDI method involves the direct injection of liquid ammonia into the
TDC Top dead center combustion chamber, requiring significantly higher injection pressures,
SAIT Start of ammonia injection timing typically above 80 MPa. Therefore, the HPDI has a higher ammonia
SDIT Start of diesel injection timing delivery efficiency and a more flexible injection strategy, which allows
3D Three-dimensional better control of combustion and ensures combustion efficiency under
conditions of high ammonia energy substitution. Li et al. [22] compared
the LPI mode and HPDI mode of ammonia fuel in an ammonia-diesel
dual-fuel two-stroke low-speed marine engine. They found that the
in CI diesel engines with a relatively low compression ratio (15.2:1), maximum energy substitution of ammonia under the LPI mode was
diesel could be used as a co-reactant to achieve stable combustion of approximately 80 %, whereas the HPDI mode achieved up to 97 %. At
ammonia. For ammonia-diesel dual-fuel models that use diesel as a the same time, the LPI mode showed a higher ITE, while the HPDI mode
combustion agent, ammonia is predominantly introduced into the resulted in lower emissions of NH3, NOX, and greenhouse gases. In their
combustion chamber after pre-mixing in the intake manifold. Reiter subsequent study, Zhou et al. [23] found that by optimizing the
et al. [13] conducted tests on a multi-cylinder turbocharged diesel en­ ammonia-diesel fuel blending process, the ITE could reach more than 50
gine test bench, investigating various ammonia-diesel ratios. Their study % in both the intake manifold LPI of ammonia and the HPDI of liquid
revealed that for the same engine torque, CO2 emissions gradually ammonia scenarios, but at this time, the NOx emissions under the LPI
decreased with increasing ammonia substitution in the fuel. Further­ strategy increased by 140.6 %. In the optimal emission condition of the
more, despite the presence of nitrogen in ammonia fuel, the combustion in-cylinder HPDI of the liquid ammonia scheme, compared to the pure
of ammonia in the engine does not necessarily result in increased NOx diesel mode, NOX emissions were reduced by approximately 47 %,
emissions. As long as the energy substitution of ammonia does not greenhouse gas emissions were reduced by approximately 97 %, and
exceed 60 %, the engine’s NOx emissions can be maintained at lower unburnt NH3 emissions were negligible. Subsequently, Zhang et al. [24]
levels. They then conducted further studies on the combustion of conducted a study on the ammonia-diesel dual direct injection mode in a
ammonia-diesel dual-fuel [14]. They found that under constant engine low-speed two-stroke engine. The results indicate that varying the
power conditions, the preferred diesel energy fraction for optimal timing of diesel injection can alter the ignition timing of ammonia in­
combustion efficiency is between 40 % and 60 %. Subsequently, Niki jection, allowing precise adjustment of the engine combustion process
et al. [15] conducted a similar investigation into the effects of ammonia- and improving ITE. Moreover, optimizing the injection timing of both
diesel blending in CI engines. The results indicate that under constant diesel and ammonia can reduce the fuel NOx emissions generated from
engine power conditions, increasing ammonia supply leads to a reduc­ ammonia combustion.
tion in engine cylinder pressure curve and peak pressure, accompanied In summary, current research on the liquid ammonia HPDI method
by an increase in ignition delay. Simultaneously, Nadimi et al. [16] has primarily focused on the impact of injection strategies on the com­
observed that in ammonia-diesel dual-fuel heavy-duty engines, the bustion process and emission formation in two-stroke marine diesel
indicated thermal efficiency (ITE) decreases with increasing ammonia engines. Due to the capability of in-cylinder direct injection to control
energy fraction, accompanied by a significant increase in unburned fuel combustion directly by modifying the injection process, employing
ammonia emissions. In order to optimize engine performance and HPDI strategies in four-stroke engines similarly exerts a significant in­
mitigate unburnt ammonia emissions, researchers have begun to explore fluence on the combustion process and emission formation. However,
novel fuel injection strategies. due to differences in intake organization and combustion modes, the
Among these, the dual-stage injection strategy for diesel is widely operational cycles of four-stroke engines differ significantly from those
employed. Mi et al. [17] investigated the strategy of diesel pilot injection of two-stroke engines. In contrast, there is a paucity of research reports
in an ammonia-diesel dual-fuel internal combustion engine using on the application of HPDI strategies of liquid ammonia in four-stroke
ammonia energy fractions of 40 %, 50 %, 60 %, and 70 %. The results engines. Notably, the larger combustion chambers of heavy-duty four-
indicate that the unburned NH3 emissions decrease from approximately stroke diesel engines also provide sufficient space for the arrangement of
8000 ppm to 1000 ppm, and the reduction in unburned NH3 becomes liquid ammonia injectors. Moreover, diesel injectors can be utilized for
more pronounced with an increase in the quantity of pilot-injected ammonia fuel without requiring modification, thus obviating the need
diesel. However, this injection strategy leads to higher emissions of for additional costs associated with developing ammonia injectors. In
NOX, CO, and HC. Yousefi et al. [18] similarly investigated the dual- comparison to the LPI scheme, the incremental cost incurred under the
stage diesel injection strategy. They initially examined the influence of HPDI scheme only encompasses the additional two injectors and the
different ammonia energy fractions and diesel injection timing on fuel conversion of the original gaseous ammonia supply system to a liquid
combustion. They found that the use of ammonia fuel led to a reduction ammonia supply system. Therefore, this study investigates the applica­
of 58.8 % in NOx emissions, primarily attributed to the presence of tion of a liquid ammonia HPDI mode in ammonia-diesel dual-fuel heavy-
thermal de-NOx reactions during NH3 combustion. However, the rela­ duty engines. Additionally, a novel HPDI strategy is proposed, effec­
tively slow flame propagation speed of ammonia resulted in a decrease tively enhancing engine performance while reducing the emissions of
in the ITE from 38.1 % to 27.2 %. In their subsequent study, Yousefi et al. unreacted fuel. The outcomes of this research provide a theoretical basis
[19] found that the use of a segmented diesel injection strategy can for the development of dual-fuel internal combustion engines using
promote a more thorough combustion of ammonia-diesel dual-fuel ammonia-diesel, thereby contributing to advancements in this field.

2
Z. Zhang et al. Fuel 367 (2024) 131450

2. Methodology

2.1. Model establishment

The ammonia-diesel dual-fuel heavy-duty engine was selected as the


research subject. The specific specifications of the engine are presented
in Table 1. The 3D numerical simulation was created using CONVERGE,
a computational fluid dynamics (CFD) software that offers great ad­
vantages in dynamic mesh handling and allows for adaptive mesh
refinement (AMR) to shorten the computation time. Since the current
study focuses on fuel mixture, fuel combustion, and emissions genera­
tion, and considering computational costs, each computational scheme
undergoes a single-cycle simulation. The simulation duration ranges
from intake valve opening (-385◦ after top dead center (ATDC)) to
exhaust valve opening (150◦ ATDC), and in all simulated schemes, intake
pressure and temperature are fixed at 1.35 bar and 313.15 K, respec­
tively. Grid independence validation should be performed before the
simulation. In the pure diesel scheme, Three base grid sizes (5 mm, 4
mm, 3 mm) are selected and identical refinement is applied to different
base grids. As shown in Fig. 1, there is no significant difference between
Fig. 1. Grid independence validation for in-cylinder mean pressure.
the results obtained under 4 mm and 3 mm mesh. While the cylinder
pressure curve decreases too much under 5 mm mesh, a 4 mm mesh size
is selected due to the balance of accuracy and efficiency of calculation.
Fig. 2 shows the mesh generated in the cross-section over the intake
valve during the opening of the intake valve.
To solve issues such as the reduction of unburned ammonia emis­
sions, a novel liquid ammonia HPDI strategy is established. Therefore, it
is necessary to define the ammonia injector in the 3D model. Fig. 3
shows the schematic defining the liquid ammonia injectors. the liquid
ammonia nozzles are arranged symmetrically, with the central axis of
symmetry being the Z-axis. Here, “l” represents the distance between the
liquid ammonia nozzle and the center of the cylinder head, “α” and “β”
respectively denote the angles between the liquid ammonia injection
direction and the X-axis and Z-axis, and the diameter of the liquid
ammonia nozzle is 0.22 mm.

2.2. Model description

In the 3D simulation, the turbulence model employed the RNG k-ε


model [25], which can reduce the amount of computation while
ensuring the accuracy of the computation. In order to reflect the spray of
diesel and liquid ammonia within the combustion chamber, the frag­
Fig. 2. Schematic diagram of mesh generation during intake valve opening.
mentation model adopted was the KH-RT model [26], which combines
the KH and RT models to calculate the fuel injection process jointly [27].
The NTC model [28] and Kuhnke Film Splash model [29] were chosen to by Xu et al. [31]. This mechanism consists of 69 species and 389
predict fuel collisions and wall impact. To accurately capture the com­ elementary reactions. In order to predict engine NOx emissions effec­
bustion process of ammonia-diesel within the combustion chamber, the tively, the Extended Zel’dovich model was chosen in this study [32].
SAGE combustion model was employed in this study[30], which allows Table 2 summarizes the computational models employed in the 3D
the utilization of user-defined combustion mechanism files. In this simulation process in this research.
study, the chemical reaction mechanism employed the chemical kinetic
skeleton mechanism for ammonia/n-heptane combustion, as proposed 2.3. Simulation conditions

Table 1 The boundary conditions and initial conditions have a significant


Engine specifications. impact on the accuracy of CFD simulations. The initial conditions for the
The main parameters Parameter description CFD simulations are consistent with the experimental parameters, where
Engine model Caterpillar 3401
engine speed is configured at the commonly employed rate of 910 rpm,
Number of cylinders 1 and the brake mean effective pressure (BMEP) is set to 8.1 bar, corre­
Displacement (L) 2.44 sponding to the half-load condition of the engine. Table 3 shows other
Piston stroke (mm) 165.1 initial conditions used during the simulation process. Because of the
Cylinder diameter (mm) 137.2
introduction of ammonia into the combustion chamber through in-
Compression ratio 16.25
Connecting rod length (mm) 261.62 cylinder direct injection, it is necessary to define the quantity of
Intake valve opening (◦ ATDC) − 385 ammonia. In this study, the usage of ammonia is defined using %NH3,
Intake valve closing (◦ ATDC) − 169 which represents the energy fraction of ammonia. It signifies the per­
Exhaust valve opening (◦ ATDC) 145 centage of energy contributed by the utilized ammonia to the total fuel
Exhaust valve closing (◦ ATDC) 348
energy. The calculation formula is expressed as follows:

3
Z. Zhang et al. Fuel 367 (2024) 131450

Fig. 3. Schematic diagram of the definition of the liquid ammonia injector.

combustion, and emissions formation. Firstly, a comparative study of


Table 2
different %NH3 was conducted to discuss the effects on the engine under
Summary of the key computational model.
the condition of constant injection pressure, with the liquid ammonia
The main parameters Parameter description injection pressure set at above 110 MPa. Secondly, with the %NH3 fixed
Turbulence RNG κ-ε at 80 % and the injection direction set at α = 60◦ , β = 75◦ , the effects of
Evaporation Frossling different ammonia injection timings were investigated. Furthermore, to
Spray breakup KH-RT evaluate the influence of ammonia injection directions, with the %NH3
Drop turbulent dispersion Wall Film-O’Rourke
Combustion SAGE
fixed at 80 and the ammonia injection timing fixed at − 12◦ ATDC, an
NOX formation Extended Zeldovich exploration was performed within the α range of 30◦ to 75◦ .

2.4. Model validation


Table 3
Summary of boundary and initial conditions. To validate the accuracy of the 3D numerical simulation model in
terms of combustion and emissions, experimental data from Yousefi
Region Type Temperature Pressure
et al. [19] on an ammonia-diesel dual-fuel engine were selected for
Air intake Inflow 313 K 1.35 bar verification. The engine model and initial conditions in this numerical
Inlet port Fixed wall 420 K NA
model are consistent with their experiments, and some key experimental
Exhaust outlet Outflow 800 K 1.50 bar
Outlet port Fixed wall 500 K NA operating conditions are presented in Table 5. Numerical simulation
Piston surface Moving wall 553 K NA results of in-cylinder average pressure curves and heat release rates were
compared with experimental results for different ammonia energy
fractions, as shown in Fig. 4. It can be observed that, under three
ṁNH3 × LHVNH3
%NH3 = × 100% (1) different ammonia energy fractions, the in-cylinder average pressure
ṁD × LHVD + ṁNH3 × LHVNH3 curves of the model align well with the experimental values, accurately
Here, ṁD represents the mass flow rate of diesel. ṁNH3 represents the predicting the combustion process within the engine chamber.
mass flow rate of ammonia. LHVD denotes the lower heating value of Furthermore, to ensure accurate predictions of emission generation by
diesel fuel. LHVNH3 represents the lower heating value of ammonia. the model, the production of several major emissions was also
After model validation, three specific simulation cases were estab­ compared. The results, illustrated in Fig. 5, demonstrate that the trends
lished, as illustrated in Table 4. The simulations primarily investigated of various emissions in the simulation results align consistently with
the effects of different strategies on the ammonia-diesel mixture, fuel experimental values as the ammonia energy fraction changes. Addi­
tionally, the predicted emission quantities are within a 5 % error range
compared to experimental values, indicating the model’s reasonable
Table 4 representation of emission generation aspects. Moreover, Fig. 6 com­
CFD simulation operating conditions. pares the emission quantities of major unburned products during two
Case1 Case2 Case3 combustion processes. The simulation results show close agreement
with experimental values for unburned ammonia and HC emissions,
Ammonia energy 0, 40, 60, 80 60 60
fraction/% affirming the model’s capability to accurately predict unburned product
Diesel injection timing/ − 14.2 − 14.2 − 14.2 emissions.

ATDC Experimental data from Zhou et al. [23] were employed to validate
Diesel mass flow rate/kg/ 3.38, 2.73, 2.08, 1.37 1.37
h 1.37, 0.69
Ammonia injection − 12 − 14, − 12, − 12 Table 5
timing/◦ ATDC − 10, − 8 The experimental operating conditions were conducted by Yousefi et al. [18].
Ammonia mass flow rate/ 0, 1.56, 3.18, 4.84, 4.84 4.84
Ammonia energy fraction /% 0 20 40
kg/h 6.45
α/◦ 60 60 30, 45, 60, Diesel injection timing /◦ ATDC − 14.2 − 14.2 − 14.2
75 Diesel mass flow rate/ kg/h 3.38 2.73 2.08
β/◦ 75 75 75 Air mass flow rate/ kg/h 83.26 80.81 78.11
l/mm 42 42 42 Ammonia mass flow rate/ kg/h 0 1.56 3.18

4
Z. Zhang et al. Fuel 367 (2024) 131450

Fig. 4. Validation of in-cylinder pressure curves and heat release rate curves for varying %NH3 schemes.

the liquid ammonia injection process. The environmental conditions set 3. Results and discussions
for the spray model validation align with those of their experiments.
They conducted liquid ammonia spray experiments using a constant- 3.1. Effect of different ammonia energy fractions on combustion and
volume combustion apparatus, achieving atomization through a emission performance
single-hole injector with a diameter of 0.22 mm. Wherein the ambient
temperature is set to 900 K, liquid ammonia injection pressure is set to In this section, a comprehensive analysis of the influence of different
60 MPa, and the liquid ammonia temperature is 350 K. Real-time cap­ ammonia energy fractions on combustion in an ammonia-diesel dual-
ture of spray images was accomplished using a high-speed camera and a fuel internal combustion engine is presented. Fig. 9 shows the average
Schlieren testing system. Regarding the model computation parameters in-cylinder pressure curves and a comparison of ammonia-diesel injec­
associated with spray phenomena, in order to accurately characterize tion timing at different %NH3. When ammonia is used to replace a
the fragmentation process of liquid ammonia droplets, the KH model portion of the diesel fuel, the cylinder pressure increases compared to
breakup time constant is set to 32, the KH model size constant is 0.6, the the pure diesel fuel scheme, with the 60 % NH3 scheme having the
RT model size constant is 0.68, and the RT model breakup length con­ highest peak pressure, followed by the 80 %NH3 and 40 %NH3 schemes.
stant is 18. Fig. 7 illustrates the experimental and simulated results of The most crucial factor is the change in combustion pattern intro­
the liquid ammonia spray morphology over time after the onset of in­ duced by the new injection strategies. Fig. 10 compares the flow field
jection. The experimental results of the spray front shape and width and temperature distribution during the fuel injection and combustion
closely matched the numerical simulation. Additionally, a comparison processes between the pure diesel scheme and the 80 %NH3 scheme. It
of the simulated and experimental penetration distances is presented in can be seen that at − 10.5◦ ATDC, flames begin to develop in the pure
Fig. 8. The simulation results for the liquid ammonia spray penetration diesel scenario, while in the 80 %NH3 scheme, the diesel flames come
distance align well with the experimental data, validating macroscopic into contact with the liquid ammonia jets at this point, igniting the
characteristic parameters. This indicates that the chosen spray model liquid ammonia. Although liquid ammonia possesses a significant latent
accurately describes the injection process of liquid ammonia spray heat of vaporization, the high injection pressure and the start of
within the combustion chamber. ammonia injection timing (SAIT) configuration of ammonia liquid in­
jection lead to its ignition in liquid form by the diesel flame upon entry
into the combustion chamber. Consequently, there is minimal evapo­
ration of liquid ammonia. In the subsequent flame development, it can
be observed that in the pure diesel scheme, the flames are predomi­
nantly distributed around the diesel jet, spreading outward after
impacting the cylinder wall. In contrast, within the 80 %NH3 scheme,

5
Z. Zhang et al. Fuel 367 (2024) 131450

Fig. 5. Validation of major emissions for varying %NH3 schemes.

Fig. 6. Validation of partially combusted emissions for varying %NH3 schemes.

the flames distribute throughout the entire combustion chamber chaotic turbulent flow field, augmenting turbulence that further am­
following the motion generated by the liquid ammonia jets. Observing plifies the subsequent fuel–air mixing efficiency during injection. As the
the flow field distribution under these two schemes reveals that, in the injection process progresses, the liquid ammonia jets induce a relatively
case of pure diesel, the flow field in the cylinder first follows the diesel regular clockwise rotating vortex flow field within the cylinder. The
jet from the center of the piston to the cylinder wall, and a part of it generated two forms of flow field enhance the comprehensive move­
bounces back after hitting the cylinder wall, and a part of it spreads ment of fuel within the combustion chamber, effectively increasing the
around with the cylinder wall. When incorporating ammonia into the efficiency of fuel–air mixing. This indicates that the in-cylinder liquid
cylinder through HPDI, the injected liquid ammonia moves along the ammonia direct injection method proposed in this study can alter the
piston bowl wall. High-pressure, substantial volume direct injection of combustion mode, resulting in multi-point rapid combustion compared
liquid-phase ammonia significantly enhances in-cylinder turbulence and to the pure diesel mode. This enhances fuel combustion efficiency,
vorticity. Initially, upon contact with the diesel jet, this results in a more ensuring the uniform filling of the combustion chamber with high-

6
Z. Zhang et al. Fuel 367 (2024) 131450

temperature regions.
Another factor contributing to the observed increase in cylinder
pressure in the dual-fuel scheme is the reduction in fuel injection
duration. From Fig. 11, it can be observed that by ensuring constant
injection pressure and total input energy, the diesel injection duration
gradually decreases as the ammonia energy substitution increases. At
the same time, the liquid ammonia injection duration increases. How­
ever, due to the constant values of the start of diesel injection timing
(SDIT) and SAIT, the time required for fuel injection in the 40 %NH3 and
60 %NH3 schemes is shorter among the three ammonia-diesel dual-fuel
schemes. This results in a more concentrated fuel injection and com­
bustion, leading to a higher peak pressure of the average in-cylinder
pressure curve.
By examining the heat release rate curves in Fig. 12 for different
schemes, it is observed that, under 40 %NH3 conditions, since the diesel
injection process entirely overlaps with the liquid ammonia injection
Fig. 7. Comparison of spray morphologies between the experiment and process, the heat release rate curve is similar to that of the pure diesel
simulations. mode, exhibiting two peaks indicative of a typical diffusion combustion
heat release rate curve. In contrast, for the 60 %NH3 and 80 %NH3
scenarios, the heat release rate curves display three peaks, demon­
strating characteristics of segmented combustion. Across the four
schemes, the first peak for each occurs approximately around
− 11◦ ATDC, indicating that the ignition timing for this engine is in the
vicinity of − 11◦ ATDC. Due to the combined accumulation of diesel and
liquid ammonia prior to ignition timing in the ammonia-blended
schemes, there is a greater amount of fuel participating in combustion
at ignition. Consequently, the peak value of the first peak in the heat
release rate curves shows an increase in the three ammonia-diesel dual-
fuel schemes. Furthermore, it can be seen from Fig. 11 that all three
schemes cumulatively inject the same amount of fuel before the moment
of ignition, so the peak size of the first peak is also the same. On the other
hand, in the pure diesel scheme, there is only diesel injection and due to
the longer injection time, the second peak is lower and appears later. In
the other three schemes, the second peak occurs between − 10◦ ATDC
and − 4◦ ATDC, indicating that the liquid ammonia injected at
− 10◦ ATDC is continuously ignited by the flame in the combustion
chamber. At the same time, it can be observed from Fig. 11 that during
the period from − 10◦ ATDC to − 4◦ ATDC, the liquid ammonia injection
quantity remains relatively constant in the three ammonia-diesel dual
Fig. 8. Comparison of liquid ammonia spray penetrations between the exper­
fuel schemes, while the diesel injection quantity gradually decreases.
iments and simulations.
Consequently, the total amount of fuel involved in combustion de­
creases, resulting in a gradual reduction of the second peak of the heat
release rate in these three schemes with increasing %NH3.
Adopting the strategy of direct liquid ammonia injection into the
cylinder not only enhances combustion efficiency but also improves
engine performance. As depicted in Fig. 13, employing this novel liquid
ammonia HPDI method with ammonia-blended fuel results in increased
combustion efficiency, leading to elevated indicated mean effective
pressure (IMEP) values for all three %NH3 schemes. Specifically, the
IMEP differences between the 40 %NH3 and 60 %NH3 are minimal,
measuring 1.13 and 1.14 MPa, respectively. However, the 80 %NH3
exhibits a slight reduction in IMEP, reaching 1.10 MPa, due to a longer
injection duration of fuel, causing less concentrated combustion.
Nevertheless, this value represents an improvement compared to the
pure diesel scheme. Observing the energy balance comparison charts for
each NH3 in Fig. 14(a), it can be noted that the improvement in fuel–air
mixture efficiency contributes to an increase in engine thermal effi­
ciency. However, it is noteworthy that unburned losses gradually in­
crease. This is primarily due to the inevitable production of more
ammonia vapor during the fuel combustion process with increasing
liquid ammonia usage and prolonged injection duration. Therefore, in
schemes with higher %NH3, the peak ammonia gas content within the
Fig. 9. Comparison of cylinder pressure curves at different %NH3 schemes.
cylinder is elevated. As ammonia combustion in gaseous form occurs at a
slower rate, the limited consumption of ammonia gas during the com­
bustion process ultimately results in unburned ammonia emissions.
Fig. 14(c) illustrates that unburned ammonia linearly increases with the

7
Z. Zhang et al. Fuel 367 (2024) 131450

Fig. 10. Comparison of temperature and velocity field distribution between pure diesel and 80 %NH3 schemes.

Fig. 11. Comparison of injection timing at different %NH3 schemes.

rise in %NH3. The unburned ammonia emissions for the 40 %NH3, 60 % Fig. 12. Comparison of heat release rate curves under the conditions of
different %NH3 schemes.
NH3, and 80 %NH3 scenarios are 0.06, 0.17, and 0.41 g/kWh, respec­
tively. Notably, the 40 %NH3 and 60 %NH3 exhibit lower ammonia
emissions, while the 80 %NH3 shows higher emissions. However, the combustion efficiency of the fuel but also by the blending ratio. As the
optimization of injection strategies can effectively mitigate unburned blending ratio increases, the amount of evaporated liquid ammonia and
ammonia emissions under a high %NH3. the absorbed heat also increase. Therefore, among the three %NH3, the
The alterations in the combustion characteristics not only impact peak in-cylinder average temperature decreases with an increase in %
engine performance but also influence the formation of emissions. NH3. By the 80 %NH3, the heat loss from ammonia blending offsets the
Fig. 15 compares the in-cylinder average temperature curves and NOx beneficial effects of the new injection method on combustion, resulting
emissions at different %NH3. It is evident that, compared to the pure in minimal differences in peak in-cylinder average temperature
diesel scheme, the in-cylinder average temperature increases after compared to the pure diesel scheme.
adopting this novel liquid ammonia HPDI method. This is primarily Typically, NOx emissions are positively correlated with in-cylinder
attributed to the improved mixing of fuel and air in the cylinder, average temperature. However, it is observed that only the NOx emis­
resulting in an accelerated combustion rate. However, after dual-fuel sions in the 40 %NH3 scenario are higher than the pure diesel scenario,
blending, the cylinder temperature is influenced not only by the while the NOx emissions in the remaining two %NH3 schemes decrease.
This is primarily due to the different pathways for NOx formation in

8
Z. Zhang et al. Fuel 367 (2024) 131450

ammonia fuel compared to pure diesel. Pure diesel combustion involves


only thermal NOx generation, whereas ammonia fuel combustion in­
troduces three factors influencing NOx emissions: thermal NOx gener­
ation, fuel NOx generation, and thermal de-NOx processes. The process
of NOx generation is illustrated by equations (2–3) [33].

H + NH3⇔H2 + NH2(2)

NO2 + NH2⇔NO + H2NO(3)

The process of thermal de-NOx is illustrated by equations (4–7)


[34,35].

NH3 + OH⇔NH2 + H2O(4)

NH3 + O⇔NH2 + OH(5)

NH2 + NO⇔N2 + H + OH(6)

NH2 + NO⇔N2 + H2O(7)

Fig. 16 compares the in-cylinder NOx generation curves between the


pure diesel and 80 %NH3. From the graph, it is evident that both sce­
Fig. 13. Comparison of IMEP under the conditions of different %NH3 schemes.
narios exhibit a rapid increase in NOx concentration. However, in the 80
%NH3, in addition to the generation of thermal NOx, it is apparent that

Fig. 14. Comparison of energy balance and in-cylinder ammonia gas content and unburned ammonia emission at different %NH3 schemes.

9
Z. Zhang et al. Fuel 367 (2024) 131450

Fig. 15. Comparison of In-cylinder average temperature and NOx emission under the conditions of different %NH3 schemes.

The primary objective of substituting ammonia fuel for diesel is to


mitigate greenhouse gas emissions. However, the combustion of
ammonia gives rise to another greenhouse gas, N2O. According to data
from the United States Environmental Protection Agency in 2022, over a
100-year time scale, the greenhouse gas effect of N2O is 273 times that of
CO2 [38]. To compare the total greenhouse gas emissions among various
injection schemes, this study defines the total greenhouse gas emissions
(EGHG), as shown in Equation (8), signifying the equivalent CO2 emis­
sions of the internal combustion engine’s greenhouse gas emissions.

EGHG = ECO2 + 273EN2O(8)

In the equation, EGHG represents the total greenhouse gas emissions,


ECO2 denotes CO2 emissions, and EN2O signifies N2O emissions.
N2O primarily functions as an intermediate product in the combus­
tion of ammonia, participating in several reactions as outlined by
Equations (9–13) [39,40].

NH3 + OH⇔NH2 + H2O(9)

NH2 + NO2⇔N2O + H2O(10)


Fig. 16. Comparison of in-cylinder NOx concentration curves between pure
diesel and 80% NH3 schemes. NH + NO⇔N2O + H(11)

O + N2O⇔N2 + O2(12)
during the liquid ammonia injection period, ammonia combustion
contributes to a portion of fuel NOx. The NOx generation rate is OH + N2O⇔N2 + HO2(13)
noticeably faster, and the NOx concentration at the same crank angle is
higher. Subsequently, as the combustion progresses, the in-cylinder NOx The generation of N2O involves a complex, multifactorial process,
concentration reaches its peak. In the pure diesel scheme, the concen­ with its principal pathways illustrated in Fig. 17. During the combustion
tration of NOx remained nearly constant, whereas in the 80 %NH3 process, the high-temperature region within the combustion chamber
scheme, there was a discernible decrease in NOx concentration. The initially gives rise to OH radicals within the flame. Subsequently, these
reduction in NOx concentration is attributed to the occurrence of the radicals undergo oxidation reactions with nearby ammonia, leading to
thermal de-NOx reaction. the formation of N2O [41]. As N2O serves as an intermediate product in
This reaction primarily occurs within the temperature range of 1100 ammonia combustion, its predominant generation occurs at the fore­
K to 1400 K. Below 1100 K, the reaction is almost negligible. At tem­ front of the ammonia flame. This is evident in the substantial overlap
peratures above 1400 K, NH3 is primarily oxidized into NO [36,37]. among the 2100 K temperature iso-surface, the 0.07 % molar fraction
Therefore, in the 60 %NH3 and 80 %NH3 schemes, where the in-cylinder OH radical iso-surface, and the 0.0001 M fraction N2O iso-surface.
average temperature is lower and decreases rapidly, the final NOx In the fuel injection scheme of liquid ammonia HPDI, liquid
emissions decrease due to thermal de-NOx. In contrast, the 40 %NH3 ammonia rapidly ignites upon entering the combustion chamber,
scheme experiences an excessive rise in-cylinder average temperature, resulting in swift flame propagation and near-complete combustion.
leading to an early-stage increase in NOx generation. Additionally, due Only a small fraction of unburned ammonia, originating from ammonia
to the complete coverage of liquid ammonia injection time by diesel vaporization, leads to incomplete combustion, consequently resulting in
injection time in this scenario, coupled with a shorter fuel injection minimal N2O production. As depicted in Fig. 18, even when considering
duration, ammonia combustion is more and more complete. There is the greenhouse effect of N2O being 273 times that of CO2 and multi­
almost no residual ammonia participating in subsequent thermal de- plying the N2O emission by 273, the relative N2O emissions compared to
NOx reactions after the in-cylinder average temperature decreases. As CO2 emissions in each scheme can be virtually neglected. Therefore, the
a result, the final NOx emissions increase to 40 %NH3. ultimate total greenhouse gas emissions are primarily influenced by CO2

10
Z. Zhang et al. Fuel 367 (2024) 131450

Fig. 17. Evolution of N2O generation in the 80 %NH3 scheme.

Fig. 19. Comparison of cylinder pressure curves at different SAIT schemes.


Fig. 18. Comparison of Greenhouse gas emission under the conditions of
different %NH3 schemes. impose restrictions on the maximum pressure rise rate within the cyl­
inder. In this engine, the maximum pressure rise rate should not exceed
emissions, which, in turn, are predominantly influenced by the %NH3. 1.1 MPa/deg. However, in the SAIT = -14◦ ATDC scenario, the maximum
The greater the substitution of diesel by ammonia, the lesser the CO2 pressure rise rate is calculated to be 3.61 MPa/deg, which is detrimental
emissions. Thus, at an 80 % ammonia energy fraction, the total green­ to the stable operation of the engine. The maximum pressure rise rate
house gas emissions are minimized, measuring only 176.1 g/kW⋅h, typically manifests when the combustion of fuel initiates within the
representing a 74.3 % reduction compared to the pure diesel scheme. combustion chamber and is associated with the first peak in the heat
release rate curve [42]. Upon comparing the heat release rate curves in
Fig. 20, it is observed that the magnitude of the first peak decreases as
3.2. Effects of different ammonia injection timing on combustion and the SAIT is retarded. Notably, the first peak values under the SAIT = -10
emission performance and − 8◦ ATDC schemes are identical. This phenomenon is primarily
attributed to the varying total fuel mass within the combustion chamber
In this section, with the 80 % ammonia energy fraction fixed, the at different ignition timings. As the end of diesel injection timing (EDIT)
impact of liquid ammonia injection timing on combustion is primarily in all schemes was set at − 10.4◦ ATDC, and the ignition timings are
investigated. Diesel injection in all four schemes commences at approximate − 11◦ ATDC for all schemes, the diesel mass within the
− 14.2◦ ATDC and concludes at − 10.4◦ ATDC. Liquid ammonia injection combustion chamber is consistent just before ignition. However, due to
initiation times are varied at − 14, − 12, − 10, and − 8◦ ATDC, each lasting the variation in SAIT, the accumulated amount of liquid ammonia
for 16◦ CA. Fig. 19 presents a comparison of in-cylinder average pressure within the combustion chamber differs before ignition, thereby resulting
curves for different SAIT. It is observed that with the retardation of the in a higher total fuel mass at the ignition timing in scenarios with more
SAIT, the peak cylinder pressure systematically decreases and is advanced SAIT. Conversely, in scenarios with SAIT values of − 10 and
delayed. This phenomenon is mainly attributed to the fact that with an − 8◦ ATDC, where SAIT is already after the ignition timing, the total fuel
earlier injection timing, more fuel is injected into the combustion mass at the ignition timing corresponds to the injected diesel quantity.
chamber before the engine reaches the top dead center, and since the Consequently, the ranking of the total fuel mass within the combustion
combustion chamber volume is at its minimum at the top dead center, chamber at the ignition timing is as follows: − 14◦ ATDC scheme >
more fuel participation before this point contributes more to the cylin­ -12◦ ATDC schemes > -10◦ ATDC schemes = -8◦ ATDC schemes. More­
der pressure elevation. On the other hand, an earlier liquid ammonia over, the amount of fuel accumulated in the combustion chamber before
injection timing can increase the interaction time between diesel and ignition influences the magnitude of the instantaneous heat release
liquid ammonia, resulting in higher intensities of early-stage turbulence during the ignition moment. Therefore, the ranking of the first peak heat
and late-stage vortices. This promotes thorough contact between fuel release rate corresponds to the ranking of the accumulated fuel mass in
and air in the combustion chamber, enhances combustion speed, and the combustion chamber just before ignition. Additionally, due to the
increases in-cylinder pressure. longer interval between SAIT and EDIT in the SAIT = -8◦ ATDC
In order to ensure the stability of the engine, it is common practice to

11
Z. Zhang et al. Fuel 367 (2024) 131450

Fig. 20. Comparison of heat release rate curve and maximum pressure rise rate under the conditions of different SAIT schemes.

compared to the SAIT = -10◦ ATDC, it can be observed that the crank schemes of SAIT = -14 and − 12◦ ATDC, the shared accumulation of
angle corresponding to the second rise in the heat release rate curve is diesel and liquid ammonia in the combustion chamber occurs prior to
more delayed in the SAIT = -8◦ ATDC. ignition due to the SAIT occurring before the ignition moment. Conse­
Fig. 21 compares the distribution of 1800 K temperature contours quently, the initial flame in both schemes comprises a coexistence of
under different liquid ammonia injection timings. It can be observed diesel and liquid ammonia flames. Furthermore, in the SAIT =
that the stages involved in the flame development process are generally -14◦ ATDC scheme, the injection of liquid ammonia occurs earlier than in
consistent across different schemes. Initially, flames appear in the the SAIT = -12◦ ATDC scheme, leading to a greater accumulation of
combustion chamber around − 10.5◦ ATDC. Subsequently, with the in­ liquid ammonia in the combustion chamber and a longer distance
jection of liquid ammonia, a circular flame develops around the piston traveled by the liquid ammonia. Consequently, the initial flame in the
bowl, and after the EDIT, the diesel flame gradually diminishes, giving SAIT = -14◦ ATDC scheme is the largest among the four schemes. In the
way to the dominance of the liquid ammonia flame. However, the details schemes of SAIT = -10 and − 8◦ ATDC, the appearance of the liquid
of flame development under different SAIT schemes vary slightly. In the ammonia flame is observed at − 8.5◦ ATDC and − 7◦ ATDC, respectively.

Fig. 21. The distribution of the 1800 K temperature contour under various SAIT schemes.

12
Z. Zhang et al. Fuel 367 (2024) 131450

Different SAIT schemes also slightly influence the flame development


speed within the combustion chamber. An earlier SAIT implies an earlier
interaction between the liquid ammonia jet and diesel jet, leading to an
earlier formation of turbulence and vortices, consequently accelerating
flame development. At − 5◦ ATDC, it is evident that the flame area de­
creases with the delay of SAIT.
As evident from the combustion phase comparative chart in Fig. 22
(a), the CA10, CA50, and CA90, are delayed with the postponement of
SAIT. The temporal distribution of heat release during different crank
angles influences the work done by the fuel on the piston. An excessively
early release of heat relative to the top dead center (TDC) may result in a
higher negative work done by the fuel on the piston, while a delayed
release may lead to fuel burning expansion in a larger combustion
chamber volume, thereby reducing thermal efficiency. Contrasting
Fig. 22(b), it is apparent that the fuel combustion timing in the SAIT =
-10◦ ATDC scheme is more conducive to work done. CA50 occurs just
after TDC, indicating superior fuel work efficiency. Consequently, this
scheme achieves the maximum IMEP of 1.14 MPa, representing a 12.9 %
increase compared to the pure diesel scheme with an IMEP of 1.01 MPa.
Fig. 23 presents an energy balance comparison under different
ammonia injection timings. It is observed that the unburned losses
decrease with the delay of SAIT. In the SAIT = -8◦ ATDC scheme, the Fig. 23. Comparison of energy balance at different SAIT schemes.
engine exhibits the highest thermal efficiency; however, due to the late
combustion in this scheme, the IMEP is lower compared to the SAIT = development is in its initial stage, limiting the amount of ammonia gas
-10◦ ATDC scheme. The 0.1 % higher thermal efficiency compared to the that can be ignited by the flame, thus resulting in a slower decline in
SAIT = -10◦ ATDC scheme is primarily compensated by the reduction in ammonia content. Nonetheless, due to the early completion of liquid
unburned losses. Examination of Fig. 24 reveals that unburned ammonia ammonia injection in the SAIT = -14◦ ATDC scheme, the decline in
emissions are higher in schemes with advanced SAIT, with an evident ammonia content also occurs relatively early, resulting in only slightly
decrease in emissions occurring between SAIT = -12◦ and − 10◦ ATDC, higher unburned ammonia emissions compared to the SAIT = -12◦ ATDC
with ignition timing at 11◦ ATDC serving as a dividing point. This is strategy. In contrast, in the SAIT = -12◦ and − 10◦ ATDC strategies, flame
primarily because liquid ammonia may undergo evaporation during the development during the liquid ammonia injection process is more
injection process, leading to the generation of gaseous ammonia. The complete, leading to a faster decrease in ammonia gas content. There­
combustion of ammonia in the gaseous phase is more challenging fore, in the two schemes with SAIT after ignition timing, unburned
compared to liquid ammonia. If liquid ammonia is injected before the ammonia emissions are lower. Among these, the SAIT = -8◦ ATDC
ignition timing, there is a greater likelihood of its evaporation into gas scheme exhibits the minimum unburned ammonia emission, measuring
during the time interval from SAIT to ignition. Conversely, if liquid 0.312 g/kW⋅h. This represents a 23.1 % reduction compared to the
ammonia is injected after the ignition timing, it is rapidly ignited upon original injection scheme with an 80 % ammonia energy fraction.
entry into the combustion chamber, significantly reducing the time Fig. 25 compares the in-cylinder average temperature curves and
available for evaporation. Therefore, the peak in-cylinder ammonia gas NOx emissions under different liquid ammonia injection timings. The
content evaporated within the cylinder increases with advancing SAIT. delayed SAIT results in a postponed combustion process of the fuel in the
Particularly notable is the SAIT = -14◦ ATDC scheme, where the furthest combustion chamber. Simultaneously, the delayed SAIT weakens its
distance from ignition leads to the highest ammonia gas evaporation. As interaction with diesel, leading to a reduction in the generation of tur­
the combustion process progresses, the evaporated ammonia within the bulence and vortex intensity. This decrease in turbulence and vortex
cylinder is ignited. However, due to the early injection of liquid intensity lowers the efficiency of fuel–air mixing, resulting in a slower
ammonia in the SAIT = -14◦ and − 12◦ ATDC schemes, flame flame combustion speed. Consequently, the observed maximum in-

Fig. 22. Combustion phase and IMEP under the conditions of different SAIT schemes.

13
Z. Zhang et al. Fuel 367 (2024) 131450

Fig. 24. Comparison of the in-cylinder ammonia gas content and unburned ammonia emission at different SAIT schemes.

Fig. 25. Comparison of in-cylinder average temperature and NOx emission under the conditions of different SAIT schemes.

cylinder average temperature decreases with the delayed SAIT. Therefore, the NOx emissions under these four SAIT schemes are cate­
Furthermore, in the SAIT = -14◦ ATDC scheme, where liquid ammonia gorized based on the ignition timing, and within each category, NOx
injection occurs earlier, and the ignition timing is around − 11◦ ATDC, emissions show a positive correlation with cylinder temperature.
there is a portion of liquid ammonia evaporating and absorbing heat In the comparison of greenhouse gas emissions, as depicted in
during the short interval between liquid ammonia injection and ignition. Fig. 26, the disparity in CO2 emissions is minimal. This is attributed to
This phenomenon is reflected in the in-cylinder average temperature the consistent ammonia energy fraction of 80 % across various SAIT
curve of the SAIT = -14◦ ATDC scheme, showing a temporary decrease in schemes, which replace an equivalent amount of diesel consumption.
temperature during the interval from SAIT to ignition timing. Normally, Notably, there is variation in the emissions of N2O. In the SAIT =
the emission of NOx is directly proportional to the in-cylinder average -14◦ ATDC scheme, the earlier injection of liquid ammonia enhances
temperature. However, an interesting observation in Fig. 25(b) is that, turbulence and vortex intensity within the combustion chamber. This
with the ignition timing as a reference point, NOx emissions increase in results in improved fuel–air mixing, rapid combustion, and lower
the schemes where SAIT occurs after ignition timing. This might be emissions of N2O, which consequently serves as an intermediate product
attributed to the influence of thermal de-NOx reactions, which require in ammonia combustion. Although there is a 67.2 % difference between
ammonia participation. In the two schemes with SAIT before ignition the highest and lowest N2O emissions among the four schemes, the
timing, liquid ammonia is involved throughout the combustion process, actual emissions of N2O in these schemes are relatively low compared to
facilitating continuous participation in thermal de-NOx reactions. In CO2 emissions. Even when scaled by a factor of 273, the difference re­
contrast, the schemes with SAIT after ignition timing allow for only a mains on the order of two magnitudes, indicating that the overall
relatively shorter duration of participation in thermal de-NOx reactions. greenhouse gas emissions closely approximate those of CO2 emissions.
Additionally, it is noted that, after the engine operates to 20◦ ATDC, the
schemes with SAIT = -10 and − 8◦ ATDC exhibit slower declines in in- 3.3. Effects of different ammonia injection directions on combustion and
cylinder average temperature. Consequently, these two schemes have emission performance
more extensive high-temperature regions in the combustion chamber,
which is unfavorable for subsequent thermal de-NOx reactions. Fig. 27 illustrates the comparison of in-cylinder average pressure

14
Z. Zhang et al. Fuel 367 (2024) 131450

α, the location where the liquid ammonia jets impinge on the piston
bowl gradually moves away from the ammonia injection nozzle.
Simultaneously, this alteration in α also results in a change in the angle
at which the liquid ammonia jets impinge on the edge of the piston bowl,
significantly influencing the subsequent flame development
morphology. It is evident that when α is greater than 30◦ , the liquid
ammonia flame within the combustion chamber moves and develops
along the edge of the piston bowl, forming a relatively uniform annular
flame. As α decreases, the flame development within the combustion
chamber accelerates, leading to a more concentrated combustion, and a
gradual increase in the peak cylinder pressure. However, as α continues
to decrease, the impingement angle of the liquid ammonia jets increases.
At this point, the velocity direction of the liquid ammonia jet can be
decomposed into components along the tangential direction (V1) to the
edge of the piston bowl and the vertical direction(V2). A noticeable
observation is that the magnitude of V2 gradually increases as α de­
creases, causing the liquid ammonia jet flame to impact the outer region
of the piston bowl rather than following the periphery of the piston
bowl. Consequently, with the reduction of α, the proportion of flame
outside the piston bowl gradually increases. When α decreases to 30◦ ,
the magnitude of V2 is significantly greater than that of V1. At this point,
Fig. 26. Comparison of greenhouse gas emissions under the conditions of the liquid ammonia flame essentially ceases to move along the edge of
different SAIT schemes. the piston bowl and instead surges out of the piston bowl in the direction
of injection. Thus, in the α = 30◦ scheme, the fuel does not exhibit vortex
curves and combustion phases under different liquid ammonia injection flow, leading to a reduction in the efficiency of the fuel–air mixture, an
directions. It can be observed that the rise rate of the in-cylinder pres­ uneven distribution of flames within the combustion chamber, and ul­
sure curve gradually increases with the decrease in α. This phenomenon timately a decrease in the peak average cylinder pressure.
is primarily associated with combustion speed [43]. Upon examining the Fig. 29 compares the in-cylinder heat release rate curves and IMEP
comparative combustion phase chart, it is evident that the combustion under different liquid ammonia injection directions. Since the SAIT is
duration progressively shortens with decreasing α. Therefore, changes in consistent across the four liquid ammonia injection direction schemes,
the injection direction affect the fuel combustion speed, with smaller α the accumulated fuel mass inside the combustion chamber before the
values corresponding to faster combustion speeds. However, it is note­ ignition moment is identical for all four schemes, resulting in a consis­
worthy that in the α = 30◦ scheme, although the combustion speed is the tent first peak in the heat release rate curves. However, with the
fastest among the four schemes, the peak cylinder pressure under this decrease in α, the occurrence position of the second peak gradually
scheme does not follow the pattern of decreasing α, leading to larger advances, and the rate of ascent of the heat release rate curves also in­
peak values. In comparison to the α = 45◦ scheme, the peak cylinder creases. From Fig. 28, it can be observed that in the schemes with larger
pressure under the α = 30◦ scheme instead experiences a slight α, the liquid ammonia jets impact the edge of the piston bowl shortly
reduction. after entering the combustion chamber. Due to compression, atomiza­
Fig. 28 contrasts the 1800 K temperature contour plots under four tion is restricted, leading to delayed combustion. However, as α de­
different liquid ammonia injection directions, with the yellow dashed creases, the movement distance of the liquid ammonia jets before
lines representing the directions of liquid ammonia injection. It is impacting the piston bowl increases. Consequently, atomization and
evident that the direction of liquid ammonia injection has a significant combustion with air occur before impacting the piston bowl in schemes
impact on the subsequent development of the flame. With a decrease in with smaller α, resulting in an earlier occurrence of the second peak in

Fig. 27. Comparison of in-cylinder average pressure curve and combustion phase under the conditions of different α schemes.

15
Z. Zhang et al. Fuel 367 (2024) 131450

Fig. 28. The distribution of the 1800 K temperature contour under various α scheme conditions.

Fig. 29. Comparison of heat release rate and IMEP under the conditions of different α schemes.

the heat release rate. Nevertheless, the examination of IMEP for the four combustion performance can only be achieved by varying the value of α
schemes reveals a non-linear correlation with the advancement of the within a certain range. Beyond a certain angle of α, the vortex flame
second heat release rate peak. In the α = 30◦ scheme, there is a decrease morphology does not occur. Among the four schemes, α = 45◦ exhibits
in IMEP. This phenomenon is attributed to a change in flame develop­ the maximum IMEP at 1.11 MPa. This represents a 10.1 % increase
ment morphology in this scheme. In the α = 30◦ scheme, the liquid compared to the pure diesel scheme with an IMEP of 1.01 MPa.
ammonia flame, after impacting the edge of the piston bowl, does not Fig. 30 compares the energy balance and in-cylinder ammonia gas
propagate along the edge as observed in other schemes. Consequently, content and unburned ammonia emissions under different liquid
the efficiency of fuel–air mixing decreases, leading to poorer combustion ammonia injection directions. In the α = 60◦ and 45◦ schemes, the
performance and a reduction in IMEP. Therefore, improvement in combustion chamber exhibits higher fuel–air mixing efficiency, with the

16
Z. Zhang et al. Fuel 367 (2024) 131450

Fig. 30. Comparison of energy balance and in-cylinder ammonia gas content and unburned ammonia emission under the conditions of different α schemes.

flame forming a swirling vortex that expands outward. This leads to NOx emissions under different liquid ammonia injection directions. It
more thorough fuel combustion, increased work output for the engine, can be observed that, as α decreases, the heat release from ammonia
and improved thermal efficiency. However, it is noteworthy that these combustion gradually advances. Consequently, the rate of rise in the in-
two schemes also exhibit higher unburned losses. This is primarily due cylinder average temperature profile increases with decreasing α, and,
to the longer distance traveled by liquid ammonia from injection to wall except for the α = 30◦ scheme, the peak of the in-cylinder average
impact in the α = 60◦ and 45◦ schemes, resulting in a portion of liquid temperature profile also increases with decreasing α. In the α = 30◦
ammonia evaporating and absorbing heat. Conversely, in the α = 75◦ scheme, influenced by the flow field within the combustion chamber,
scheme, liquid ammonia impacts the piston shortly after being injected the efficiency of fuel–air mixing decreases, resulting in a reduction in
from the nozzle. This compression effect may reduce the atomization fuel combustion efficiency compared to the α = 45◦ scheme. As a result,
efficiency, resulting in less ammonia evaporation and lowering the po­ the peak of the in-cylinder average temperature is slightly lower. Typi­
tential for unburned ammonia emissions. It is worth noting that in the α cally, in diesel single-fuel internal combustion engines, NOx emissions
= 30◦ scheme, unburned ammonia is also minimal. This is attributed to are positively correlated with temperature. However, it is observed that
the rapid combustion of fuel immediately after injection, with the fuel NOx emissions in the four schemes are lower in cases with higher tem­
atomization-to-combustion process being very short and the flame peratures. This is primarily due to the presence of the thermal de-NOx
spreading smoothly. Examining the final unburned ammonia emissions process within the reaction temperature range of 1100 K-1400 K dur­
reveals that, in the α = 60◦ and 45◦ schemes, there is a higher level of ing ammonia-diesel dual-fuel combustion. Consequently, during the
unburned ammonia emissions. In contrast, the α = 75◦ scheme exhibits initial rise in cylinder average temperature, the in-cylinder NOx content
the lowest unburned ammonia emissions at 0.291 g/KW⋅h. This repre­ is predominantly governed by the thermal NOx generation process,
sents a 28.3 % reduction compared to the original injection scheme with while after the cylinder average temperature begins to decline, the
an 80 % ammonia energy fraction. thermal NOx generation process essentially ceases. Simultaneously, an
Fig. 31 compares the in-cylinder average temperature profiles and increase in the occurrence of lower temperature regions within the

17
Z. Zhang et al. Fuel 367 (2024) 131450

Fig. 31. Comparison of in-cylinder average temperature and NOx concentration curve of different α schemes.

combustion chamber is observed, marking the initiation of the domi­


nance of the thermal de-NOx process over the NOx content.
It can be observed that the peak in-cylinder NOX concentration is
determined by both the rate of increase and the duration of the ascent
during the NOX concentration rising phase. Analyzing Fig. 31 and Fig. 27
reveals that the rate of increase in NOX concentration rises with the
increasing rate of in-cylinder average temperature elevation. The
duration of the NOX concentration rise is related to the combustion
phase, it can be noted that as the combustion duration extends, the
duration of the NOX generation process also prolongs. Among the three
schemes other than α = 75◦ , although the duration of thermal NOX
generation decreases with the decrease in α, it is considered a secondary
factor compared to the rate of increase during the NOX concentration
rising phase in influencing the peak NOX content. Therefore, the
magnitude of the peak NOX content is primarily influenced by the rate of
increase during the in-cylinder average temperature rising phase, the
ultimate peak NOX concentration increases with the decrease in α.
However, in the α = 75◦ scheme, although the rate of increase in in-
cylinder average temperature is close to that of the α = 60◦ scheme,
due to the longer combustion duration, the peak NOX content in the α =
75◦ scheme is only lower than that in the α = 30◦ scheme. In the sub­
sequent phase dominated by thermal de-NOX, the duration of the NOX
Fig. 32. Comparison of greenhouse gas emission under the conditions of
concentration reduction phase remains roughly the same, with the pri­
different α schemes.
mary factor influencing the decrease in NOX concentration being the
rate of decline in in-cylinder average temperature. The α = 45◦ scheme
exhibits the highest rate of decline in in-cylinder average temperature impact, instead, it rushes out of the piston bowl. Consequently, the flame
and consequently a lower peak NOX content, thus achieving the lowest front exhibits more folds compared to other schemes, resulting in a more
NOX emission level ultimately. The rates of decline in in-cylinder pronounced quenching phenomenon. As a result, N2O emissions relative
average temperature for the α = 30◦ and α = 60◦ schemes are similar, to α = 45◦ increase. However, due to the relatively small quantity of N2O
resulting in comparable rates of NOX concentration reduction; however, emissions, the ultimate total greenhouse gas emissions are determined
due to the higher peak NOX content in the α = 30◦ scheme, the NOX by CO2 emissions. The lowest CO2 emissions occur in the α = 45◦
emission is higher for the α = 30◦ scheme compared to the α = 60◦ scheme, measuring 168.1 g/KW⋅h.
scheme. In the α = 75◦ scheme, where the temperature declines the
slowest, less NOX is consumed within the combustion chamber, resulting 4. Conclusions
in the highest NOX emissions among the four schemes.
Fig. 32 presents a comparative analysis of greenhouse gas emissions This paper presents a novel liquid ammonia direct injection strategy
under various ammonia injection directions. It is evident that, in and conducts a comprehensive computational investigation based on
schemes with higher α values, the combustion of the flame is slower, this fuel injection method, comparing the influences of different
leading to an extended combustion time for the ammonia flame. The ammonia energy fractions, liquid ammonia injection timing, and liquid
probability of incomplete combustion product N2O generation during ammonia injection directions on the combustion and emission charac­
ammonia combustion is higher in schemes with larger α values. There­ teristics of an ammonia-diesel dual-fuel heavy-duty engine. The main
fore, when α exceeds 30◦ , the emission of N2O decreases with a reduc­ conclusions are as follows:
tion in α. (1) Through investigation of the combustion process within an
In the case of α = 30◦ , although the combustion time is shorter, the ammonia-diesel dual-fuel heavy-duty engine, it was observed that under
flame in this scheme does not follow the edge of the piston bowl after constant total fuel energy input, the novel liquid ammonia HPDI strategy

18
Z. Zhang et al. Fuel 367 (2024) 131450

developed in this study could alter the in-cylinder flow field during the References
combustion process, generating an early turbulent flow field followed by
a later vortex flow field. This modification enhances the efficiency of [1] Zhao Y, Su Q, Li B, Zhang Y, Wang X, Zhao H, et al. Have those countries declaring
“zero carbon” or “carbon neutral” climate goals achieved carbon emissions-
fuel and air mixing, facilitating the formation of multi-point combustion economic growth decoupling? J Clean Prod 2022;363:132450.
within the combustion chamber and subsequently improving combus­ [2] Ayhan V, Ece YM. New application to reduce NOx emissions of diesel engines:
tion efficiency. electronically controlled direct water injection at compression stroke. Appl Energy
2020;260:114328.
(2) Compared to the pure diesel mode, dual-fuel in-cylinder direct [3] Sachuthananthan B, Vinoth R, Satya R, Sandeep B, Sudheer N, Deekshith C. Study
injection results in an enhancement of IMEP and thermal efficiency. on the use of selective catalytic reduction technique for NOx emission reduction in
Simultaneously, the use of ammonia fuel significantly reduces green­ an diesel engine fuelled with methyl ester of water hyacinth. Mater Today: Proc
2022;68:1415–21.
house gas emissions, and due to the presence of the thermal de-NOx [4] Fayyazbakhsh A, Bell ML, Zhu X, Mei X, Koutný M, Hajinajaf N, et al. Engine
reaction, NOx emissions also decrease at high ammonia energy frac­ emissions with air pollutants and greenhouse gases and their control technologies.
tions. In comparison to the pure diesel mode, the greenhouse gas and J Clean Prod 2022;376:134260.
[5] Bao J, Qu P, Wang H, Zhou C, Zhang L, Shi C. Implementation of various bowl
NOx emissions under the 80 %NH3 scheme are reduced by 74.3 % and
designs in an HPDI natural gas engine focused on performance and pollutant
59.9 %, respectively. However, unburned losses also increase, leading to emissions. Chemosphere 2022;303:135275.
elevated unburned ammonia emissions. [6] Li J, Lai S, Chen D, Wu R, Kobayashi N, Deng L, et al. A review on combustion
(3) Different timings of liquid ammonia injection can alter the characteristics of ammonia as a carbon-free fuel. Front Energy Res 2021;9:760356.
[7] Olabi A, Abdelkareem MA, Al-Murisi M, Shehata N, Alami AH, Radwan A, et al.
interaction time of the ammonia-diesel jet, affecting the turbulence and Recent progress in green ammonia: production, applications, assessment; barriers,
vortex intensity within the combustion chamber. Simultaneously, it and its role in achieving the sustainable development goals. Energ Conver Manage
modifies the fuel distribution at the ignition moment, influencing the 2023;277:116594.
[8] Kurien C, Mittal M. Review on the production and utilization of green ammonia as
development of the flame within the combustion chamber. Advancing an alternate fuel in dual-fuel compression ignition engines. Energ Conver Manage
the timing of liquid ammonia injection can enhance the flame devel­ 2022;251:114990.
opment speed; however, it adversely affects engine stability, leading to a [9] Sekhar SJ, Al-Shahri ASA, Glivin G, Le T, Mathimani T. A critical review of the
state-of-the-art green ammonia production technologies-mechanism, advancement,
reduction in overall engine performance. In addition, delaying the challenges, and future potential. Fuel 2024;358:130307.
timing of liquid ammonia injection can effectively reduce unburned [10] Qi Y, Liu W, Liu S, Wang W, Peng Y, Wang Z. A review on ammonia-hydrogen
ammonia emissions. fueled internal combustion engines. eTransportation 2023;18:100288.
[11] Comotti M, Frigo S. Hydrogen generation system for ammonia–hydrogen fuelled
(4) Changes in the direction of liquid ammonia injection can influ­ internal combustion engines. Int J Hydrogen Energy 2015;40(33):10673–86.
ence the subsequent flame development morphology. When α = 30◦ , the [12] Gray JT, Dimitroff E, Meckel NT, Quillian Jr R. Ammonia fuel—engine
angle at which the liquid ammonia jets impact the edge of the piston compatibility and combustion. SAE Trans 1967:785–807.
[13] Reiter AJ, Kong S-C. Demonstration of compression-ignition engine combustion
bowl is excessively large. In this case, the liquid ammonia flame does not
using ammonia in reducing greenhouse gas emissions. Energy Fuel 2008;22(5):
move along the piston bowl edge as in other schemes. This alteration in 2963–71.
flame development patterns leads to a decrease in IMEP, thermal effi­ [14] Reiter AJ, Kong S-C. Combustion and emissions characteristics of compression-
ciency, and in-cylinder average temperature. However, for α > 30◦ , ignition engine using dual ammonia-diesel fuel. Fuel 2011;90(1):87–97.
[15] Niki Y, Yoo DH, Hirata K, Sekiguchi H. Effects of ammonia gas mixed into intake
gradually reducing α accelerates the development of the liquid ammonia air on combustion and emissions characteristics in diesel engine. ASME 2016
flame in the combustion chamber, enhances vortex intensity, and im­ Internal Combustion Engine Division Fall Technical Conference. 2016.
proves engine performance. Nevertheless, unburned ammonia emissions [16] Nadimi E, Przybyła G, Lewandowski MT, Adamczyk W. Effects of ammonia on
combustion, emissions, and performance of the ammonia/diesel dual-fuel
increase as α decreases, with the lowest unburned ammonia emissions compression ignition engine. J Energy Inst 2023;107:101158.
occurring in the α = 75◦ scheme, representing a 28.3 % reduction [17] Mi S, Wu H, Pei X, Liu C, Zheng L, Zhao W, et al. Potential of ammonia energy
compared to the original 80 %NH3 injection scheme. fraction and diesel pilot-injection strategy on improving combustion and emission
performance in an ammonia-diesel dual fuel engine. Fuel 2023;343:127889.
[18] Yousefi A, Guo H, Dev S, Liko B, Lafrance S. Effects of ammonia energy fraction and
CRediT authorship contribution statement diesel injection timing on combustion and emissions of an ammonia/diesel dual-
fuel engine. Fuel 2022;314:122723.
[19] Yousefi A, Guo H, Dev S, Lafrance S, Liko B. A study on split diesel injection on
Zheng Zhang: Writing – original draft, Validation, Investigation, thermal efficiency and emissions of an ammonia/diesel dual-fuel engine. Fuel
Conceptualization. Liming Di: Supervision, Funding acquisition. Lei 2022;316:123412.
Shi: Resources, Methodology. Xiyu Yang: Writing – review & editing, [20] Manigandan S, Ryu JI, Praveen Kumar TR, Elgendi M. Hydrogen and ammonia as a
primary fuel – a critical review of production technologies, diesel engine
Software. Tengfei Cheng: Software, Data curation. Cheng Shi: Writing
applications, and challenges. Fuel 2023;352:129100.
– review & editing, Supervision, Resources, Project administration. [21] Wu B, Wang Y, Wang D, Feng Y, Jin S. Generation mechanism and emission
characteristics of N2O and NOx in ammonia-diesel dual-fuel engine. Energy 2023;
Declaration of competing interest 284:129291.
[22] Li T, Zhou X, Wang N, Wang X, Chen R, Li S, et al. A comparison between low-and
high-pressure injection dual-fuel modes of diesel-pilot-ignition ammonia
The authors declare that they have no known competing financial combustion engines. J Energy Inst 2022;102:362–73.
interests or personal relationships that could have appeared to influence [23] Zhou X, Li T, Wang N, Wang X, Chen R, Li S. Pilot diesel-ignited ammonia dual fuel
low-speed marine engines: a comparative analysis of ammonia premixed and high-
the work reported in this paper. pressure spray combustion modes with CFD simulation. Renew Sustain Energy Rev
2023;173:113108.
Data availability [24] Zhang Z, Long W, Dong P, Tian H, Tian J, Li B, et al. Performance characteristics of
a two-stroke low speed engine applying ammonia/diesel dual direct injection
strategy. Fuel 2023;332:126086.
Data will be made available on request. [25] Raj AGS, Mishra CS. Simulation and experimental data resemblance of Darmstadt
spark ignition engine with different turbulence models – a computational fluid
dynamics cold flow data. Data Brief 2022;43:108340.
Acknowledgments [26] Jia M, Pan H, Bian Y, Zhang Z, Chang Y, Liu H. Calibration of the constants in the
kelvin-helmholtz rayleigh-Taylor (KH-RT) breakup model for diesel spray under
The authors would like to acknowledge the financial support pro­ wide conditions based on advanced data analysis techniques. Atom Sprays 2022;32
(6).
vided by Chunhui Project Foundation of the Education Department of
[27] Bao J, Wang H, Wang R, Wang Q, Di L, Shi C. Comparative experimental study on
China (Grant No. HZKY20220239) and Science Research Project of macroscopic spray characteristics of various oxygenated diesel fuels. Energy Sci
Hebei Education Department (Grant No. QN2023224). Eng 2023;11(5):1579–88.
[28] Schmidt DP, Rutland CJ. A new droplet collision algorithm. J Comput Phys 2000;
164(1):62–80.
[29] Han Z, Xu Z, Trigui N. Spray/wall interaction models for multidimensional engine
simulation. Int J Engine Res 2000;1(1):127–46.

19
Z. Zhang et al. Fuel 367 (2024) 131450

[30] Shi C, Chai S, Di L, Ji C, Ge Y, Wang H. Combined experimental-numerical analysis [38] Shin J, Park S. Numerical analysis for optimizing combustion strategy in an
of hydrogen as a combustion enhancer applied to wankel engine. Energy 2023;263: ammonia-diesel dual-fuel engine. Energ Conver Manage 2023;284:116980.
125896. [39] Mathieu O, Petersen EL. Experimental and modeling study on the high-temperature
[31] Xu L, Chang Y, Treacy M, Zhou Y, Jia M, Bai X-S. A skeletal chemical kinetic oxidation of ammonia and related NOx chemistry. Combust Flame 2015;162(3):
mechanism for ammonia/n-heptane combustion. Fuel 2023;331:125830. 554–70.
[32] Shi C, Chai S, Wang H, Ji C, Ge Y, Di L. An insight into direct water injection [40] Duynslaegher C, Jeanmart H, Vandooren J. Flame structure studies of premixed
applied on the hydrogen-enriched rotary engine. Fuel 2023;339:127352. ammonia/hydrogen/oxygen/argon flames: experimental and numerical
[33] Kobayashi H, Hayakawa A, Somarathne KKA, Okafor EC. Science and technology investigation. Proc Combust Inst 2009;32(1):1277–84.
of ammonia combustion. Proc Combust Inst 2019;37(1):109–33. [41] Nakamura H, Hasegawa S, Tezuka T. Kinetic modeling of ammonia/air weak
[34] Lyon RK. The NH3-NO-O2 reaction. Int J Chem Kinet 1976;8(2):315–8. flames in a micro flow reactor with a controlled temperature profile. Combust
[35] Lee G-W, Shon B-H, Yoo J-G, Jung J-H, Oh K-J. The influence of mixing between Flame 2017;185:16–27.
NH3 and NO for a De-NOx reaction in the SNCR process. J Ind Eng Chem 2008;14 [42] Gong C, Li Z, Sun J, Liu F. Evaluation on combustion and lean-burn limitof a
(4):457–67. medium compression ratio hydrogen/methanol dual-injection spark-ignition
[36] Glarborg P, Miller JA, Ruscic B, Klippenstein SJ. Modeling nitrogen chemistry in engine under methanol late-injection. Appl Energy 2020;277:115622.
combustion. Prog Energy Combust Sci 2018;67:31–68. [43] Gong C, Yi L, Zhang Z, Sun J, Liu F. Assessment of ultra-lean burn characteristics
[37] Joo J, Lee S, Kwon O. Effects of ammonia substitution on combustion stability for a stratified-charge direct-injection spark-ignition methanol engine under
limits and NOx emissions of premixed hydrogen–air flames. Int J Hydrogen Energy different high compression ratios. Appl Energy 2020;261:114478.
2012;37(8):6933–41.

20

You might also like