Download as pdf or txt
Download as pdf or txt
You are on page 1of 19

Energy 296 (2024) 131171

Contents lists available at ScienceDirect

Energy
journal homepage: www.elsevier.com/locate/energy

Parametric analysis and optimization of the combustion process and


pollutant performance for ammonia-diesel dual-fuel engines
Cheng Shi a, Zheng Zhang a, Huaiyu Wang b, Jingyi Wang c, Tengfei Cheng a, Liang Zhang a, *
a
School of Vehicle and Energy, Yanshan University, Qinhuangdao 066004, China
b
School of Mechanical Engineering, Beijing Institute of Technology, Beijing, 100081, China
c
AECC Beijing Institute of Aeronautical Materials, Beijing 100095, China

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

Handling editor: X Ou As a low-cost, carbon-free, and widely available hydrogen carrier fuel, ammonia has now regained peoples’
attention in the future development of the carbon-neutral engine. To evaluate the feasibility of ammonia as a
Keywords: carbon-reducing fuel for a high-power diesel engine, a computational fluid dynamics simulation model of a
Diesel engine Caterpillar 3401 diesel engine has been established and validated. Initially, an analysis was conducted to
Combustion characteristics
examine the effects of different ammonia energy fractions on combustion and the generation of emissions.
Ammonia-diesel
Subsequently, to address issues related to incomplete ammonia combustion and a decline in engine power
Pollutants formation
Diesel injection strategies performance, a novel fuel injection strategy was developed. This strategy optimizes the distribution of diesel
within the combustion chamber and the combustion process of ammonia-diesel dual-fuel by coordinating the
injection timing of diesel fuel with the injection angle. The results indicate that an ammonia energy fraction of
40% is deemed to be an optimal mixing ratio for ammonia addition. Compared with pure diesel mode, green­
house gas emissions are reduced by 34.3%, and the indicated mean effective pressure is only reduced by 4%
when the ammonia energy fraction is 40%. Collaborative optimization of fuel injection timing and angle can
achieve multi-point pilot ignition, shorten the combustion duration, and thus approach the effect of premixed
combustion. This leads to a substantial decrease in unburned ammonia emissions, with the optimal unburned
ammonia emission injection scheme registering only 0.279 g/kW⋅h, a mere 1.63% of the emissions under the
original injection scheme with a 40% ammonia energy fraction. Additionally, the rapid combustion contributes
to an increase in in-cylinder temperature, resulting in a significant reduction in N2O and CO emissions.

1. Introduction Ammonia is convenient to store, with a liquid form energy density 1.5
times higher than that of hydrogen [11]. Furthermore, heavy-duty diesel
In response to the demand for carbon neutrality following the engines can adapt to ammonia fuel without significant adjustments [12].
enactment of the Paris Climate Agreement, countries worldwide are Moreover, heavy-duty diesel engines commonly utilized in industries
actively pursuing ways to decrease carbon emissions [1]. As a major such as heavy manufacturing, heavy-duty vehicles, and shipping are not
component of power output devices, heavy-duty compression ignition suitable for electrification [13]. Therefore, ammonia is considered one
(CI) diesel engines generate significant amounts of soot and NOX during of the most promising carbon-free fuels for successful application in
operation [2]. In recent years, numerous advanced combustion control heavy-duty diesel engines in this century.
methods have been researched [3,4], which have effectively lowered In recent decades, there has been significant research on the use of
soot and NOX emissions from diesel engines [5,6]. However, due to the ammonia fuel in CI diesel engines. The flame propagation speed of
utilization of fossil fuels, diesel engines still emit significant quantities of ammonia is relatively slow [14], and improving the compression ratio
CO2 and other greenhouse gases, which severely impacts their sustain­ can effectively enhance ammonia’s combustion performance [15].
ability. One crucial method to reduce carbon emissions is to explore Achieving a compression ratio of 35:1 is required for the operation of a
alternative fuels [7]. Among them [8], ammonia stands out, as its CI diesel engine with pure ammonia fuel [16]. Therefore, when using
complete combustion results only in nitrogen and water [9,10]. ammonia as an alternative fuel in CI diesel engines, diesel is commonly

* Corresponding author.
E-mail addresses: shicheng@ysu.edu.cn (C. Shi), zhangliang402@ysu.edu.cn (L. Zhang).

https://doi.org/10.1016/j.energy.2024.131171
Received 17 February 2024; Received in revised form 25 March 2024; Accepted 30 March 2024
Available online 4 April 2024
0360-5442/© 2024 Elsevier Ltd. All rights reserved.
C. Shi et al. Energy 296 (2024) 131171

employed as a co-reactant to ignite the ammonia [17]. In 1966, Gray influences the distribution of diesel fuel within the combustion chamber.
et al. [18] conducted a study on the application of ammonia fuel in This objective can be more readily achieved by modifying the injection
diesel engines. They demonstrated that in CI diesel engines with a angle of the diesel nozzle. Modifying the injection angle significantly
relatively low compression ratio (15.2:1), diesel could be used as a impacts both the distribution of fuel within the combustion chamber and
co-reactant to achieve stable combustion of ammonia. For the combustion process. Moreover, once the injector is installed,
ammonia-diesel dual-fuel models that use diesel as a combustion agent, changing the injection angle becomes challenging. After a change in the
ammonia is predominantly introduced into the combustion chamber fuel used by the engine, the original injection angle may no longer be
after pre-mixing in the intake manifold. Reiter et al. [19] conducted tests suitable. Therefore, a novel injection strategy is developed in this study,
on a multi-cylinder turbocharged diesel engine test bench, investigating which optimizes the distribution of diesel fuel within the combustion
various ammonia-diesel ratios. Their study revealed that for the same chamber by coordinating the injection timing of diesel with the injection
engine torque, CO2 emissions gradually decreased with increasing angle, ensuring a uniform mixture of diesel and ammonia. This opti­
ammonia substitution in the fuel. Furthermore, despite the presence of mization aims to enhance the combustion process in ammonia-diesel
nitrogen in ammonia fuel, the combustion of ammonia in the engine dual-fuel engines. This research provides theoretical support for the
does not necessarily result in increased NOx emissions. As long as the development of injection strategies in ammonia-diesel dual-fuel engines.
energy substitution of ammonia does not exceed 60%, the engine’s NOx
emissions can be maintained at lower levels. They then conducted 2. Methodology
further studies on the combustion of ammonia-diesel dual-fuel [20].
They found that under constant engine power conditions, the preferred 2.1. Model establishment
diesel energy fraction for optimal combustion efficiency is between 40%
and 60%. Subsequently, Niki et al. [21] conducted a similar investiga­ In this study, a one-dimensional (1D) numerical simulation model of
tion into the effects of ammonia-diesel blending in CI engines. The re­ the engine was constructed to obtain boundary conditions that were not
sults indicate that under constant engine power conditions, increasing achievable in experimental settings for three-dimensional (3D) simula­
ammonia supply leads to a reduction in engine cylinder pressure curve tions. Additionally, a 3D numerical simulation model was developed to
and peak pressure, accompanied by an increase in ignition delay. validate and predict the combustion performance of ammonia-diesel
Simultaneously, Nadimi et al. [22] observed that in ammonia-diesel dual-fuel within the combustion chamber. All model parameters were
dual-fuel heavy-duty engines, the indicated thermal efficiency (ITE) consistent with the experimental study conducted by Yousefi et al. [25],
decreases with increasing ammonia energy fraction, accompanied by a and the specific parameters are presented in Table 1. The construction of
significant increase in unburned ammonia emissions. Sun et al. [23] also the 1D numerical simulation model utilized the GT-Power software. This
found that even with the utilization of post-treatment methods, the re­ software is capable of simulating the engine operating conditions
sidual ammonia emitted from the engine remains challenging to elimi­ through multi-cycle transient real-time simulations, and the overall
nate when utilizing ammonia-diesel dual fuel. In order to optimize model is illustrated in Fig. 1. According to the GT-Power user manual,
engine performance and mitigate unburnt ammonia emissions, re­ the intake manifold grid length was set to 0.4 times the cylinder diam­
searchers have begun to explore novel fuel injection strategies. eter. In comparison, the exhaust manifold grid length was set to 0.55
Among these, the dual-stage injection strategy for diesel is widely times the cylinder diameter. Fig. 2 illustrates the pressure curve during
employed. Mi et al. [24] investigated the strategy of diesel pilot injection the intake process obtained from the 1D simulation results.
in an ammonia-diesel dual-fuel engine using ammonia energy fractions The 3D numerical simulation was created using CONVERGE, a
of 40%, 50%, 60%, and 70%. The results indicate that the unburned NH3 computational fluid dynamics (CFD) software that offers great advan­
emissions decrease from approximately 8000 ppm–1000 ppm, and the tages in dynamic mesh handling and allows adaptive mesh refinement
reduction in unburned NH3 becomes more pronounced with an increase (AMR) to shorten the computation time. Since the current study focuses
in the quantity of pilot-injected diesel. However, this injection strategy on fuel mixture, fuel combustion, and emissions generation, the simu­
leads to higher emissions of NOX, CO, and HC. Yousefi et al. [25] simi­ lation duration for each computational scheme ranges from intake valve
larly investigated the dual-stage diesel injection strategy. They initially opening (− 385◦ after top dead center (ATDC)) to exhaust valve opening
examined the influence of different ammonia energy fractions and diesel (150◦ ATDC). Grid independence validation should be performed before
injection timing on fuel combustion. They found that the use of the simulation. In the pure diesel scheme, three base grid sizes (5 mm, 4
ammonia fuel led to a reduction of 58.8% in NOx emissions, primarily mm, 3 mm) are selected and identical AMR is applied to different base
attributed to the presence of thermal de-NOx reactions during NH3 grids. As shown in Fig. 3, there is no significant difference between the
combustion. However, the relatively slow flame propagation speed of results obtained under 4 mm and 3 mm mesh. While the cylinder pres­
ammonia resulted in a decrease in the ITE from 38.1% to 27.2%. In their sure curve decreases too much under 5 mm mesh, a 4 mm mesh size is
subsequent study, Yousefi et al. [26] found that the use of a segmented selected due to the balance of accuracy and efficiency of calculation.
diesel injection strategy can promote a more thorough combustion of
ammonia-diesel dual-fuel engines. This results in an ITE of 39.7%, 2.2. Model description
exceeding that of pure diesel operation. At the same time, it reduces
unburned ammonia emissions by 85.3% and GHG emissions by 30.6%. In the 1D simulation, a heat transfer model based on the Flow heat
In summary, in ammonia-diesel dual-fuel engines, the diesel injec­ transfer model was employed. This model is capable of calculating heat
tion strategy has a direct impact on combustion and emissions. An
appropriate injection strategy not only enhances the emission perfor­
Table 1
mance of the engine but also ensures the combustion efficiency of the Engine specifications.
fuel in the combustion chamber. Therefore, exploration of the ammonia-
The main parameters Parameter description
diesel injection strategy is essential. However, current research mostly
focuses on diesel injection timing, with limited exploration of diesel Engine model Caterpillar 3401
injection angles. In these studies, attempts have been made to influence Displacement (L) 2.44
Bore × stroke (mm × mm) 137.2 × 165.1
the combustion process within the combustion chamber by employing Compression ratio 16.25
strategies such as early injection or dual-stage injection. However, Intake valve opening (◦ ATDC) − 385
achieving lower residual ammonia emissions while ensuring engine Intake valve closing (◦ ATDC) − 169
performance remains challenging. A critical factor affecting the com­ Exhaust valve opening (◦ ATDC) 145
Exhaust valve closing (◦ ATDC) 348
bustion process is the alteration of injection timing, which directly

2
C. Shi et al. Energy 296 (2024) 131171

Fig. 1. 1D performance simulation model of the original engine.

Fig. 2. Intake pressure curve output by 1D simulation calculation.


Fig. 3. Grid independence validation for in-cylinder mean pressure.
transfer between combustion in the cylinder and the cylinder wall by
considering detailed flow characteristics within the cylinder. The com­ study, the chemical reaction mechanism employed the chemical kinetic
bustion model utilized was EngCylCombDiWiebo. When employing this skeleton mechanism for ammonia/n-heptane combustion, as proposed
combustion model, the engine’s geometric structure does not influence by Xu et al. [33]. This mechanism consists of 69 species and 389
the combustion rate; rather, it is solely affected by the relevant param­ elementary reactions. In order to predict engine NOx emissions effec­
eters within the combustion model. This characteristic significantly re­ tively, the Extended Zel’dovich model was chosen in this study [34].
duces the computational cost of simulations. The fuel properties of diesel Table 2 summarizes the computational models employed in the 3D
and ammonia were modeled using the diesel2-combust and nh3-vap fuel simulation process in this research.
models provided in the GT-Power fuel library.
In the 3D simulation, the turbulence model employed the RNG k-ε
model [27], 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 Table 2
mentation model adopted was the KH-RT model [28], which combines Summary of the key computational model.
the KH and RT models to calculate the fuel injection process jointly [29]. The main parameters Parameter description
The NTC model [30] and Kuhnke Film Splash model [31] were chosen to Turbulence RNG κ-ε
predict fuel collisions and wall impact. To accurately capture the com­ Evaporation Frossling
bustion process of ammonia-diesel within the combustion chamber, the Spray breakup KH-RT
SAGE combustion model was employed in this study [32], which allows Drop turbulent dispersion Wall Film-O’Rourke
Combustion SAGE
the utilization of user-defined combustion mechanism files. In this
NOX formation Extended Zeldovich

3
C. Shi et al. Energy 296 (2024) 131171

2.3. Initial and boundary conditions Table 4


Summary of boundary and initial conditions.
The boundary conditions and initial conditions have a significant Region Type Temperature Pressure
impact on the accuracy of CFD simulations. The initial conditions for the
Air intake Inflow Based on 1D result Based on 1D result
1D and 3D simulations are consistent with the experimental parameters, Inlet port Fixed wall 420 K NA
where engine speed is configured at the commonly employed rate of Exhaust outlet Outflow Based on 1D result Based on 1D result
910 rpm, and the brake mean effective pressure (BMEP) is set to 8.1 bar, Outlet port Fixed wall 500 K NA
corresponding to the half-load condition of the engine. These conditions Piston surface Moving wall 553 K NA

are in accordance with the experimental conditions reported by Yousif


et al. [25], Table 3 presents CFD simulation operating conditions.
Validation was conducted with ammonia energy fractions of 0, 20, and Table 5
40, assuming uniform mixing of ammonia and air entering the com­ CFD simulation operating conditions.
bustion chamber through the intake port. Table 4 shows the initial NH3 energy fraction/% 40
conditions used during the simulation process, those initial conditions Diesel injection timing/ ATDC

− 24,-30,-46,-62
presented in the table are average values. In the actual 3D simulation Diesel injection angle/◦ 85,80,65,50
process, to enhance simulation accuracy, the pressure and temperature Diesel mass flow rate/kg/h 2.08
in the intake and exhaust were controlled by curves derived from the Air mass flow rate/kg/h 78.11
Ammonia mass flow rate/kg/h 3.18
results of the 1D simulation. Due to the utilization of ammonia-diesel
dual fuel in the simulation model, it is necessary to define the quan­
tity of ammonia. In this study, the usage of ammonia is defined using %
NH3, which represents the energy fraction of ammonia. It signifies the
percentage of energy contributed by the utilized ammonia to the total
fuel energy. The calculation formula is expressed as follows:
ṁNH3 × LHVNH3
%NH3 = × 100% (1)
ṁD × LHVD + ṁNH3 × LHVNH3

Here, ṁD represents the mass flow rate of diesel. ṁNH3 represents the
mass flow rate of ammonia. LHVD denotes the lower heating value of
diesel fuel. LHVNH3 represents the lower heating value of ammonia.
After completing the validation of a 1D and 3D model and analyzing
the impact of different %NH3 schemes on the combustion of ammonia-
diesel dual-fuel engines, it was observed that a 40%NH3 is relatively
suitable for this operating condition. Therefore, under the fixed condi­
tion of %NH3 at 40, optimization was pursued by injection timing and
altering injection angles of diesel to enhance the distribution of diesel
within the combustion chamber, four different diesel injection timings
were selected, followed by an exploration of four different injection
angles under each injection timing. This was undertaken to address is­
sues related to incomplete ammonia combustion and a decline in engine
power performance. Due to the inherent diesel injection timing in the
Fig. 4. Schematic diagram of the definition of the diesel injection angle.
engine, achieving a uniform distribution of diesel within the combustion
chamber is challenging. The original injection timing is no longer suit­
able for engines utilizing ammonia-diesel dual fuels. Therefore, the capability of the 1D and 3D simulation models in capturing overall en­
diesel injection timing investigated in this study is progressively gine performance and in-cylinder combustion conditions.
advanced from − 24◦ ATDC. The specific parameters and boundary For 1D simulations, the primary objective was to provide boundary
conditions are detailed in Table 5. The definition of the diesel injection conditions for 3D simulations, thus focusing primarily on the validation
angle, denoted as α and illustrated in Fig. 4, signifies the angle between of the engine operating process. Since the fuel flow rate in the 1D
the injection direction and the negative half-axis of the Z-axis. simulation is automatically controlled by the control mode program
after the simulation starts, while in the 3D simulation, it is set directly
before computations, it is essential to validate the fuel flow rate to
2.4. Model validation ensure the accuracy of ammonia energy fractions within the 1D simu­
lation. Fig. 6 illustrates the comparison of fuel flow rates, revealing that
Fig. 5 presents a comparative analysis between the 1D and 3D nu­ the errors between experimental and simulated values are within 5%,
merical simulation results of in-cylinder average pressure curves and indicating the assurance of ammonia mass fraction in 1D simulation.
heat release rates, under varying %NH3 schemes, against experimental The above data comparisons demonstrate that the constructed 1D model
values. The relative errors between the 1D and 3D simulation data of the in this study can effectively simulate the engine operating process,
cylinder pressure and the experimental data are both found to be less providing a valuable reference for boundary conditions in 3D numerical
than 5% upon calculation. This affirms the accurate predictive simulations.
For 3D simulations, the primary focus lies in investigating the pro­
Table 3 cesses of emission generation and in-cylinder combustion performance
CFD simulation operating conditions. of the engine. Therefore, it is imperative to validate several key emis­
NH3 energy fraction/% 0 20 40 60 sions for a comprehensive analysis. In particular, due to the consistently
Diesel injection timing/◦ ATDC − 14.2 − 14.2 − 14.2 − 14.2 low levels of NO2 emissions in both experimental and simulated con­
Diesel mass flow rate/kg/h 3.38 2.73 2.08 1.37 ditions, its contribution is negligible compared to NO emissions.
Air mass flow rate/kg/h 83.26 80.81 78.11 77.46 Therefore, the NOX emission is predominantly represented by NO. The
Ammonia mass flow rate/kg/h 0 1.56 3.18 4.84
results, illustrated in Fig. 7, demonstrate that the trends of various

4
C. Shi et al. Energy 296 (2024) 131171

Fig. 5. Validation of in-cylinder pressure curves and heat release rate curves for varying %NH3 schemes.

Fig. 6. Validation of fuel flow rates for different %NH3 schemes under 1D simulation.

emissions in the simulation results align consistently with experimental 3. Results and discussion
values as the ammonia energy fraction changes. Additionally, the pre­
dicted emission quantities are within a 5% error range compared to 3.1. Effect of different ammonia energy fractions on combustion
experimental values, indicating the model’s reasonable representation
of emission generation aspects. Moreover, Fig. 8 compares the emission In this section, a comprehensive analysis is presented regarding the
quantities of major unburned products under different %NH3 schemes. impact of varying ammonia fractions on the combustion process in
The simulation results show close agreement with experimental values ammonia-diesel dual-fuel engines. Fig. 9 illustrates the distribution of
for unburned ammonia and HC emissions, affirming the model’s capa­ air velocities during diesel injection for different %NH3 schemes. During
bility to accurately predict unburned product emissions. the diesel injection process, the air velocities near the diesel jet decrease
proportionally with the increase in %NH3 at different time points.
Moreover, as the %NH3 concentration rises, the reduction in air velocity

5
C. Shi et al. Energy 296 (2024) 131171

Fig. 7. Validation of major emissions for different %NH3 schemes under 3D simulation.

Fig. 8. Validation of partially combusted emissions for different %NH3 schemes under 3D simulation.

becomes more pronounced. This phenomenon is attributed to the diesel and ammonia deteriorates.
increased concentration of ammonia, coupled with the decrease in the Fig. 10 illustrates the distribution of ammonia concentration during
mass of injected diesel, leading to an augmented hindrance to the mo­ the combustion process at different %NH3 schemes. During combustion,
tion during the diesel injection process. When %NH3 is less than 40%, ammonia is initially ignited by the diesel jet. As the ammonia content
the diesel jet can impact the piston bowl earlier, creating a vortex along increases and diesel decreases, the ammonia directly ignited by the
the piston bowl wall, as depicted in the schematic direction. This en­ diesel jet gradually diminishes, and diesel is rapidly consumed. The
hances the contact area between diesel and ammonia, promoting better remaining ammonia is predominantly concentrated above the diesel jet
mixing of the two fuels. However, as %NH3 increases, the intensity of the and outside the piston bowl, indicating a transition from premixed
vortex significantly diminishes. The discussion above indicates that the combustion to diffusion combustion. Due to the relatively slow flame
velocity field and vortex strength within the combustion chamber are propagation speed of ammonia, incomplete combustion and an increase
influenced by %NH3. With an increase in %NH3, the mixing efficiency of in unburned ammonia emissions are likely to occur. This phenomenon

6
C. Shi et al. Energy 296 (2024) 131171

Fig. 9. Comparison of the airflow velocity distribution during diesel injection at various %NH3 schemes.

Fig. 10. Comparison of ammonia mass during the combustion process at various %NH3 schemes.

becomes particularly pronounced when %NH3 exceeds 40. Therefore, in quantity increases, leading to a decrease in flame propagation speed and
scenarios with higher %NH3, adjustments to the diesel injection strategy incomplete combustion of the remaining ammonia. Under the three
are recommended to increase the portion of ammonia directly ignited by ammonia blending schemes, the emissions of unburned ammonia are
diesel. Fig. 11 presents the emissions of unburned ammonia at various % 5.4 g/kW⋅h, 16.5 g/kW⋅h, and 26.8 g/kW⋅h, respectively. Compared to a
NH3 schemes. It can be observed that, with an increase in %NH3, the 40% ammonia fraction, the unburned ammonia emissions at a 60%
emission of unburned ammonia gradually rises. This is primarily ammonia fraction increase by 62.4%.
attributed to the shift towards diffusion combustion as the ammonia Fig. 12 compares the distribution of the 1800 K temperature iso-
surface during the combustion process at different %NH3 schemes. In
comparison to the pure diesel scheme, the introduction of ammonia did
not significantly affect the ignition timing, but an increase in %NH3
notably influenced the flame propagation speed. During the initial
stages of combustion, the flame is primarily initiated by the diesel jet
and expands outward. In schemes with %NH3 less than 40, where there
is a higher proportion of diesel and a better ammonia-diesel mixture, the
flame propagates faster. However, when %NH3 exceeds 40%, diesel is
rapidly consumed, resulting in a decrease in ammonia available for
direct ignition by diesel. Subsequent combustion relies mainly on
ammonia diffusion combustion. Due to the slower flame propagation
speed of ammonia, incomplete combustion is more likely to occur within
the cylinder, and since the flame is initiated by diesel, which is primarily
present in the piston bowl, most of the flame is concentrated within the
piston bowl. The external flame diffusion is slow, resulting in a notice­
able circular flame surface expanding outward. Fig. 13 depicts the
variation of the combustion phase during the in-cylinder combustion
process at different %NH3 schemes. The addition of ammonia signifi­
cantly increases the combustion duration, and with an increase in %
NH3, both CA10-50 and CA50-90 gradually lengthen. When %NH3 in­
creases from 0 to 40, CA10-50 and CA50-90 slightly increase, indicating
Fig. 11. Comparison of unburned ammonia emission at different % a relatively small impact on combustion speed. However, when the %
NH3 schemes. NH3 increases from 40 to 60, CA10-50 and CA50-90 extend by 2.67◦ CA

7
C. Shi et al. Energy 296 (2024) 131171

Fig. 12. The distribution of the 1800K temperature contour under different %NH3 schemes.

and 33.6%, respectively. It is evident that an excessive mixing of


ammonia adversely affects engine performance. However, subsequent
optimization of injection strategies can be employed to compensate for
these drawbacks.
The introduction of ammonia not only influences combustion on a
macroscopic scale but also exerts an impact on the microscale genera­
tion of reactive species during the combustion process. Fig. 15 illustrates
the distinctions in the in-cylinder NO generation process between the
pure diesel and 40% NH3 schemes. During pure diesel combustion, the
combustion occurs around the diesel spray, forming a high-temperature
diffusive flame, leading to the formation of thermal NOx in the vicinity
of the flame. In the subsequent combustion process, the fuel diffuses
within the combustion chamber, rapidly increasing the temperature in
the high-temperature regions, resulting in the swift production of ther­
mal NOx in areas above 1800 K. In contrast, in the 40%NH3 scheme, not
only is thermal NOx generated, but also a portion of NOx formed by the
fuel. This is attributed to the presence of nitrogen atoms in ammonia
itself, inevitably leading to the formation of fuel-NOx during the com­
bustion process. This process is represented by Eqs. (2) and (3) [35].

Fig. 13. Comparison of combustion phase at different %NH3 schemes. H + NH3⇔H2+NH2 (2)

NO2+NH2⇔NO + H2NO (3)


and 4.90◦ CA, representing increases of 29.9% and 19.9%, respectively.
This suggests that as %NH3 increases, the phenomenon of uneven fuel Therefore, as depicted in Fig. 15(b), in the 40% NH3 scheme, the
mixture becomes more pronounced. Due to the slower combustion speed simultaneous formation of thermal NOx generated at high temperatures
of ammonia, the flame propagation slows down during this period. and fuel-NOx from ammonia combustion results in a higher concentra­
Fig. 14 presents a comparison of in-cylinder pressure curves and tion of NOx during the combustion phase in the 40% NH3 scheme.
indicated mean effective pressure (IMEP) under different %NH3 In addition, there is a thermal de-NOx process in the ammonia-diesel
schemes. With the increase in %NH3, there is an augmentation in the dual-fuel engines, which also influences the ultimate NOx emission. The
crank angle corresponding to the peak in-cylinder average pressure, a reaction for the thermal de-NOx process is depicted by Eqs. (4)–(7) [36].
delay in the timing of the peak in-cylinder average pressure, and a
gradual decrease in IMEP. This is primarily attributed to the slow NH3+OH ⇔ NH2+ H2O (4)
combustion rate of ammonia, resulting in an excessively prolonged NH3+O ⇔ NH2+ OH (5)
combustion time. During the expansion phase of combustion, when the
volume of the engine combustion chamber is relatively large, the NH2+NO ⇔ N2+H + OH (6)
contribution of fuel combustion expansion to the rise in cylinder pres­
NH2+NO ⇔ N2+H2O (7)
sure is limited. This leads to a reduction in the efficiency of fuel com­
bustion work on the piston. At %NH3 of 40 and 60, the IMEP values are This reaction predominantly takes place within the temperature
8.01 MPa and 6.71 MPa, respectively. In comparison to the IMEP of 1.01 range of 1100 K–1400 K. Below 1100 K, the reaction is nearly negligible,
MPa under the pure diesel mode, these represent reductions of 20.7% while above 1400 K, NH3 is primarily oxidized into NO [37]. Fig. 16

8
C. Shi et al. Energy 296 (2024) 131171

Fig. 14. Comparison of cylinder pressure curves and IMEP at different %NH3 schemes.

Fig. 15. Comparison of NOx generation process between pure diesel and 40%NH3 schemes.

compares the NO content curve and temperature curve in the combus­ Furthermore, following the decrease in temperature in the combustion
tion chamber between the pure diesel and 40% NH3 schemes. It is chamber, the thermal de-NOx process continues, resulting in a further
evident that in the pure diesel scheme, as the temperature in the com­ reduction in cylinder NO content, ultimately leading to a significant
bustion chamber starts to decrease, the NO content reaches its maximum decrease in NO emissions.
and remains relatively stable. In contrast, the 40% NH3 scheme, during Fig. 17 illustrates the generation process of another GHG, N2O, in the
the temperature increase, rapidly generates thermal NOx and fuel-NOx. 40%NH3 scheme. The greenhouse effect of this particular GHG is 273
Subsequently, as the temperature in the combustion chamber rises to times that of CO2 over a 100-year time scale [38]. The reaction equation
around 1100 K, the presence of NH3 triggers the thermal de-NOx pro­ governing the production of this GHG is depicted in Eqs. 8–10 [39].
cess. Comparatively, the peak NO content within the cylinder decreases
in the 40%NH3 scheme compared to the pure diesel scheme. NH3+OH ⇔ NH2+H2O (8)

9
C. Shi et al. Energy 296 (2024) 131171

Fig. 16. Comparison of NO content curves in the combustion chamber between pure diesel and 40%NH3 schemes.

Fig. 17. The process of N2O generation in the 40%NH3 scheme.

NH2+NO2⇔N2O + H2O (9)

NH + NO ⇔ N2O + H (10)

From the figure, it can be observed that during the combustion


process, OH radicals are initially generated in the high-temperature
region of the combustion chamber flame. Subsequently, they undergo
oxidative reactions with nearby ammonia, producing N2O, which,
however, undergoes decomposition within the temperature range of
1073–1273 K [40,41]. Therefore, the emissions of N2O are influenced by
both NH3 combustion and the temperature within the combustion
chamber, constituting a multifactorial and intricate process. Notably, in
the process of N2O generation, it consistently occurs at the ammonia
flame front. As an incomplete byproduct of ammonia combustion, the
likelihood of N2O generation increases in schemes with longer propa­
Fig. 18. The cylinder temperature curve and the in-cylinder concentration
gation distances and durations of the ammonia flame front. This is
curves of NH3 and N2O in the 40%NH3 scheme.
because, during such conditions, there is a higher probability of flame
wrinkling, which raises the likelihood of flame quenching, thereby
a significant impact on N2O emissions, such as alterations in injection
facilitating incomplete combustion of ammonia. In Fig. 18, the cylinder
timing and spray angles.
concentration curves and temperature curves of NH3 and N2O are
compared. During the initial stages of combustion, the concentration of
N2O rapidly increases with the combustion of ammonia. However, as the
3.2. Effects of different injection timing and injection angle on combustion
temperature in the combustion chamber gradually decreases, the con­
and emission performance
tent of N2O decreases, signifying that the rate of N2O decomposition
exceeds the rate of NH3 combustion oxidation to produce N2O. Given the
3.2.1. Combustion performance
complexity of N2O generation, different combustion strategies may have
In this section, the impact of optimizing injection timing and spray

10
C. Shi et al. Energy 296 (2024) 131171

angles based on a 40%NH3 was explored on diesel distribution, outside the piston bowl, resulting in slower combustion of ammonia
ammonia-diesel dual-fuel engine combustion, and emissions. Fig. 19 inside the piston bowl. Notably, in the four injection schemes at
illustrates the spatial distribution of diesel during the injection process − 62◦ ATDC injection timing, due to increased mixing time between
in various diesel injection schemes within the combustion chamber. It diesel and ammonia before combustion, the ammonia-diesel mixture is
can be observed that a later injection timing leads to the ignition of a uniform. The flame loses a distinct shape, approaching homogeneous
portion of diesel during the injection process, causing non-uniform combustion. In the scheme in the upper right part of the figure, the
mixing with ammonia before combustion initiation. Additionally, initial flame is located inside the piston bowl, making it challenging for
different diesel injection schemes alter the location of diesel jet impact the flame to propagate outside the bowl.
and its spatial distribution within the combustion chamber; smaller Fig. 21 presents a comparison of the combustion phase and ignition
diesel spray angles and earlier injection timings shift the impact position delay under various injection schemes. Ignition delay is defined as the
away from the cylinder head. Among the four injection timings, each time elapsed from diesel injection to CA10. It can be observed that
exhibits an injection angle causing the diesel jet to collide with the edge advancing the injection timing leads to an earlier start of combustion
of the piston bowl, depicted in orange in the figure. Under these and a shorter combustion duration. However, the ignition delay in­
schemes, diesel distributes more uniformly within and outside the piston creases, primarily due to the lower temperature and pressure within the
bowl after impact on the bowl surface. However, under the injection combustion chamber during the early stages of the compression stroke.
timings of − 46◦ ATDC and − 62◦ ATDC, even though diesel is primarily This prolongs the time from injection to diesel ignition, resulting in an
directed toward the cylinder wall, the early injection timing results in an extended ignition delay. The shortened combustion time is mainly
increased residence time of diesel within the combustion chamber. In attributed to the increased time for diesel movement within the com­
comparison to the − 30◦ ATDC injection timing, under these two injec­ bustion chamber with the advancement of injection timing. The more
tion timings, diesel eventually attains a more uniform distribution uniform distribution of diesel in the cylinder allows most ammonia to be
within the combustion chamber. On the other hand, under the same ignited by diesel, reducing the need for self-diffusion combustion and,
injection timing, when the injection angle is smaller than that of the consequently, minimizing the combustion time. Notably, the scheme
orange scheme, as depicted in the upper right portion of the figure, the with − 24◦ ATDC injection timing and 85◦ injection angle exhibits uni­
diesel spray primarily impacts the piston bowl. Except for the scheme form diesel distribution but slower combustion speed. This is attributed
with an injection timing of − 30◦ ATDC and an angle of 50◦ , after diesel to the delayed injection timing, causing a portion of diesel to ignite
jets impact the piston bowl, they move along the edge of the piston bowl, during the injection process, with the majority of diesel not yet reaching
collide with the cylinder head, and subsequently disperse outside the the piston bowl. As a result, ammonia within the piston bowl needs to
piston bowl. Ultimately, this leads to a uniform distribution within the self-diffuse for combustion, leading to higher values of CA50-90. Be­
combustion chamber. In contrast, in most other schemes, the majority of sides, smaller combustion duration corresponds to schemes with more
diesel is distributed within the central region of the combustion uniform diesel distribution. Specifically, the scheme with the shortest
chamber. combustion duration is the one with − 62◦ ATDC injection timing and 80◦
Fig. 20 compares the distribution of temperature during the com­ injection angle, measuring 9.32◦ CA.
bustion process under different injection schemes. It can be observed Fig. 22 presents a comparison of cylinder pressure curves under
that the flame initially expands outward from the diesel injection, and different injection schemes. It is evident that advancing the injection
among the four orange schemes where diesel jets precisely target the timing increases the maximum cylinder pressure. This is attributed to
edge of the piston bowl, the diesel distribution is more uniform. In these the comprehensive mixing of diesel with ammonia and air when the
schemes, the initial flame exists simultaneously inside and outside the injection timing is advanced, with premixed combustion dominating the
piston bowl, allowing ammonia within the combustion chamber to combustion process, leading to rapid combustion and higher cylinder
ignite from multiple locations, promoting more thorough combustion. In pressures. Furthermore, except for the injection timing schemes at
contrast, in the lower left part of the figure, diesel is primarily injected − 24◦ ATDC and − 30◦ ATDC under an injection angle of 85◦ , where the

Fig. 19. The diesel distribution during the diesel injection process under various injection schemes.

11
C. Shi et al. Energy 296 (2024) 131171

Fig. 20. The temperature distribution during the combustion process under different injection schemes.

piston’s upward motion [42]. This unfavorable condition adversely af­


fects the operation of the engine piston, resulting in a decrease in IMEP.
In contrast, with a relatively delayed injection timing, the timing of the
engine combustion process becomes more appropriate. The CA50-90
mostly occurs after the TDC and exhibits relative concentration. The
IMEP of the engine can be assured. In the case of the injection timing at
− 24◦ ATDC and an injection angle of 65◦ , the maximum IMEP is ach­
ieved at 1.01 MPa, which is comparable to the IMEP under the pure
diesel mode.

3.2.2. Emission performance


Fig. 24 illustrates the contour plots of the 1% ammonia mass fraction
during the combustion process at various injection schemes, character­
izing the flame front propagation of ammonia. Analysis of the figure
indicates that the combustion of ammonia is influenced by the distri­
bution of diesel. Under the same injection timing, in schemes with a
uniform distribution of diesel, ammonia can be ignited from multiple
points within the combustion chamber. However, in schemes where the
diesel jet impacts the piston bowl, as shown in the upper right scheme of
the figure, the consumption of both the flame and ammonia initiates
from within the piston bowl, making it challenging for the flame to
propagate toward the cylinder wall in the vicinity of the piston bowl
during the later stages of combustion. A distinct circularly spreading
flame front during the late combustion stages is observable. In contrast,
in scenarios where diesel is predominantly distributed outside the piston
bowl, as depicted in the lower-left scheme of the figure, ammonia starts
Fig. 21. Comparison of combustion phase and ignition delay under different
injection schemes. to be consumed near the cylinder wall, resulting in a portion of ammonia
near the cylinder center that is challenging to combust. Furthermore, the
rate of ammonia consumption is directly correlated with the combustion
average cylinder pressure was lower compared to other schemes with
duration. Under the injection timing of − 62◦ ATDC, due to the homo­
the same injection timing, varying injection angles at different injection
geneous mixing of diesel and ammonia, combustion manifests as ho­
timings did not significantly impact the cylinder pressure. Fig. 23 il­
mogeneous burning, proceeding rapidly. For the four schemes at this
lustrates the IMEP under different injection schemes. The introduction
injection timing, it can be observed that, within a mere 4◦ CA timeframe,
of ammonia into the engine leads to a reduction in IMEP. This decline is
ammonia is essentially completely consumed.
primarily attributed to the slow combustion of ammonia, resulting in
Fig. 25 compares the ammonia concentration curves within the
lower combustion efficiency compared to the pure diesel mode. How­
combustion chamber under different injection schemes. With the
ever, it is possible to mitigate the impact of combustion efficiency on
advancement of injection timing, the ignition timing of ammonia com­
IMEP by adjusting the injection timing and injection angle. When the
bustion also advances. Moreover, the combustion rate of ammonia
diesel injection timing is advanced, the combustion of diesel primarily
significantly increases. This is primarily attributed to the earlier
occurs before the top dead center (TDC), leading to negative work on the

12
C. Shi et al. Energy 296 (2024) 131171

Fig. 22. Comparison of cylinder pressure curves under different injection schemes.

are expressed by Eq. (11).

EGHG = ECO2+273EN2O (11)

In the equation, EGHG represents the total GHG emissions, ECO2 denotes
CO2 emissions, and EN2O signifies N2O emissions. From the graph, it can
be observed that, compared to the CO2 emission level of 685 g/kW⋅h in
the pure diesel mode, the CO2 emissions for all schemes exhibit a sig­
nificant reduction, with CO2 emissions increasing with advancing in­
jection timing. However, under the same injection timing, the
differences in CO2 emissions among various injection angle schemes are
minor. This is attributed to an equivalent substitution of diesel with
ammonia, both accounting for 40% of the total fuel energy. Since N2O is
Fig. 23. Comparison of IMEP under different injection schemes. an intermediate product during ammonia combustion, it tends to be
generated during ammonia flame quenching or incomplete combustion.
injection timing, which prolongs the mixing time between diesel and With advanced injection timing, the combustion flame inside the engine
ammonia. It allows for the ignition of ammonia from multiple points, combustion chamber initiates from multiple points, leading to rapid
accelerating flame propagation. Under the same diesel injection timing, flame development and fast propagation. Consequently, the probability
the ignition timing of ammonia undergoes minimal changes, with only of N2O generation is significantly reduced. In contrast, in schemes with
variations in ammonia consumption rates. Appropriate fuel injection slower flame propagation, where fuel combustion time and flame
angles facilitate a more uniform distribution of diesel in the combustion propagation distance are extended, the likelihood of flame quenching at
chamber, enabling ammonia to combust from multiple points and the flame front increases. This raises the possibility of incomplete
reducing the flame propagation distance. Consequently, it is evident that combustion at the flame front, resulting in higher N2O emissions.
in schemes where diesel distribution is uniform, the ammonia con­ Therefore, under the same injection timing, schemes with unreasonable
sumption rate is higher. fuel injection angles exhibit uneven diesel distribution, slower flame
Fig. 26 contrasts the emissions of unburned ammonia under different development, and higher N2O emissions. The final comparison of total
injection schemes. It is evident from the graph that advancing the in­ GHG emissions in Fig. 27(c) reveals that, compared to the pure diesel
jection timing reduces the emissions of unburned ammonia signifi­ mode, all schemes exhibit reduced GHG emissions. The lowest emissions
cantly. Additionally, at a constant injection timing, the fuel injection occur in the − 30◦ ATDC injection timing with a 50◦ injection angle
angle also has a substantial impact on unburned ammonia emissions. In scheme. Although this scheme does not yield the lowest N2O emissions,
schemes where diesel distribution is uneven, the slow development of its overall GHG emissions are minimized to 450 g/kW⋅h due to lower
flames results in a substantial increase in unburned ammonia emissions. CO2 emissions. This represents a 34.3% reduction compared to the pure
For instance, under two injection timings of − 24◦ and − 30◦ ATDC, the diesel mode.
maximum emissions of unburned ammonia occur at an 85◦ injection Fig. 28 presents a comparison of in-cylinder temperature and NO
angle scheme, with values of 89.5 g/kW⋅h and 78.8 g/kW⋅h, respec­ emissions under different injection schemes. With %NH3 held constant,
tively. Conversely, in schemes with faster combustion rates, the emis­ the formation of NOX is primarily influenced by thermo-NOX formation,
sions of unburned ammonia are lower. When the injection timings of where temperature serves as the dominant factor. Lower in-cylinder
− 24◦ , − 30◦ , − 46◦ , and − 62◦ ATDC, the schemes with the lowest un­ temperatures correspond to reduced NOX formation. Additionally,
burned ammonia emissions correspond to injection angles of 50◦ , 50◦ , schemes with lower in-cylinder temperatures are conducive to the
65◦ , and 80◦ , with emission rates of 12.8 g/kW⋅h, 7.0 g/kW⋅h, 3.6 g/ progress of thermal de-NOX reduction reactions. As observed in the
kW⋅h, and 0.28 g/kW⋅h, respectively. These values represent 77.5%, graph, NO emissions increase with advancing injection timing. This is
42.4%, 21.8%, and 1.63% of the original injection scheme with a 40% attributed to the increased motion time of diesel within the combustion
NH3. chamber, allowing ammonia to ignite from multiple locations. In com­
Fig. 27 compares GHG emissions under different injection schemes. parison to schemes with delayed injection timing, the combustion pro­
In the ammonia-diesel dual-fuel engine, due to the presence of both CO2 cess is faster, resulting in higher in-cylinder temperatures and
and N2O GHG emissions, total GHG emissions are defined in this study to substantial thermo-NOX generation. Conversely, in schemes with un­
facilitate comparison with pure diesel scenarios. Total GHG emissions even diesel distribution and slow combustion, in-cylinder temperatures

13
C. Shi et al. Energy 296 (2024) 131171

Fig. 24. The distribution of iso-surfaces with a 1% mass fraction of NH3 under various injection schemes.

Fig. 25. Comparison of in-cylinder ammonia concentration under different injection schemes.

are lower, leading to lower NO emissions. The scheme with the lowest readily oxidized, leading to lower CO emissions with higher in-cylinder
NO emissions is the − 24◦ ATDC injection timing with an 85◦ injection temperatures [43]. As depicted in Fig. 29(b), an earlier injection timing
angle, registering at 4.59 g/kW⋅h. generates more OH radicals. Moreover, under the same injection timing,
Fig. 29 illustrates a comparison of CO emissions and in-cylinder OH schemes with higher in-cylinder temperatures exhibit higher OH radical
radical peak concentrations under different injection schemes and in the concentrations. This implies that in injection schemes with uneven
pure diesel mode. The generation of CO is primarily influenced by the diesel distribution and lower in-cylinder temperatures, the oxidation of
concentrations of OH radicals and temperature. Higher OH radical CO is reduced, resulting in higher final CO emissions. Conversely, in
concentrations lead to increased oxidation of CO, resulting in lower CO injection schemes with more complete combustion, CO emissions are
emissions. Similarly, in elevated temperature conditions, CO is more minimized. The lowest CO emissions are observed in the − 62◦ ATDC

14
C. Shi et al. Energy 296 (2024) 131171

Fig. 26. Comparison of unburned ammonia under different injection schemes.

Fig. 28. Comparison of in-cylinder temperature and NO emissions under


different injection schemes.

Fig. 27. Comparison of GHG emissions under different injection schemes. Fig. 29. Comparison of CO emissions and in-cylinder OH radical peak con­
centrations under different injection schemes.
injection timing with an 85◦ injection angle scheme, measuring at 0.062
g/kW⋅h, which is 6.67% of the pure diesel scheme. combustion times and a decrease in combustion efficiency. Under
optimal GHG emission conditions, the engine achieves an IMEP of 0.97
3.3. Optimized combustion strategy for the ammonia-diesel dual-fuel MPa, accounting for 96% of the IMEP in the pure diesel scenario.
engine Considering the reduction in GHG emissions, a 4% loss in IMEP is
deemed acceptable. Fig. 31 depicts contour plots of unburned ammonia
Fig. 30 presents contour plots of GHG emissions and IMEP under the and NO emissions based on diesel injection timing and diesel injection
influence of diesel injection angle and diesel injection timing. Under the angle. With the advancement of injection timing, unburned ammonia
precondition of a 40%NH3, higher GHG emissions are concentrated in emissions significantly decrease. In the scenario of − 62◦ ATDC injection
the central and upper-right regions of the plots. A reduction in GHG timing and 80◦ injection angle, unburned ammonia emissions are merely
emissions is observed with smaller injection angles at relatively delayed 0.28 g/kW⋅h. This is mainly attributed to the homogeneous mixing of
injection timings. Additionally, at an injection angle of 85◦ , the engine diesel and ammonia, resulting in combustion approaching homogeneous
exhibits inferior combustion performance. This is primarily attributed to combustion and rapid consumption of ammonia. In Fig. 31(b), NO
the uneven mixing of diesel and ammonia, leading to prolonged emissions primarily increase with the advancement of injection timing,

15
C. Shi et al. Energy 296 (2024) 131171

Fig. 30. GHG emissions and IMEP for various diesel injection timings and injection angles.

Fig. 31. Unburned ammonia and NO emissions for various diesel injection timings and injection angles.

with minimal impact from injection angle variations. This is primarily In order to comprehend the energy losses during the combustion
attributed to the fixed %NH3 at 40, where the generation of NO is pre­ process of an ammonia-diesel dual-fuel engine, Fig. 32 provides a
dominantly influenced by temperature. Advancing the injection timing comparative analysis of the energy distribution in the fuel. The un­
increases the in-cylinder temperature unavoidably, leading to a conse­ burned loss refers to the energy loss caused by the portion of the fuel that
quent rise in NOx emissions, which can be effectively reduced through does not undergo combustion and thermal efficiency denotes the ratio of
post-processing methods. the work done by the fuel on the piston to the total energy released

Fig. 32. Comparison of energy balance under different injection schemes.

16
C. Shi et al. Energy 296 (2024) 131171

during fuel combustion. The aggregate of the remaining energy losses is Table 6
represented as other losses. It can be observed that, under the same The optimal injection schemes for the four emissions.
injection timing, in schemes with uneven diesel distribution, the ability Scheme The lowest emissions of Parameters
of diesel to ignite ammonia is limited. Additionally, due to the slow pollutants
combustion of ammonia, there is an increase in unburned ammonia, S1 Unburned NH3 − 62◦ ATDC injection timing with an 80◦
leading to a significant proportion of unburned losses. Simultaneously, injection angle
advancing the injection timing can mitigate the emission of unburned S2 GHG − 30◦ ATDC injection timing with a 50◦
ammonia. At an injection timing of − 62◦ ATDC, there is almost negli­ injection angle
S3 NO − 24◦ ATDC injection timing with an 85◦
gible unburned loss. However, at this timing advance, combustion ini­ injection angle
tiates earlier, and as the combustion approaches homogeneous burning, S4 CO − 62◦ ATDC injection timing with an 85◦
the combustion rate significantly accelerates. Most of the fuel combusts injection angle
before reaching the top dead center, resulting in a reduction of the work
done by the fuel combustion on the piston and a subsequent decrease in
4. Conclusions
thermal efficiency.
Fig. 33 provides a comparative analysis of the emissions of unburned
This study employed CFD simulation software of GT-power and
NH3, GHG, NOX, and CO for four injection strategies with the lowest
CONVERGE to construct a 1D and 3D simulation model for an ammonia-
emissions of each pollutant. The emissions of these four pollutants and
diesel dual-fuel engine, respectively. The study primarily introduces a
the corresponding injection strategies are detailed in Table 6. There is a
novel fuel injection strategy that coordinates the timing and angle of
marginal difference in the total GHG emissions among the four strate­
diesel fuel injection, computational investigating its impact on flame
gies. In the S3 strategy, the NO emissions are the lowest, but the emis­
development and pollutant generation in ammonia-diesel dual-fuel en­
sions of unburned NH3 and CO are the highest among the four strategies.
gine. The main conclusions are summarized as follows:
In contrast, the S1 and S4 strategies exhibit the lowest emissions of
unburned NH3 and CO; however, the NO emissions in these two stra­
(1) The introduction of ammonia significantly reduces CO2 emissions
tegies are comparatively elevated compared to the other strategies. The
and, due to the presence of thermal de-NOx reactions, also
S2 strategy demonstrates relatively balanced emissions for all four pol­
markedly decreases NO emissions. However, an increase in the
lutants, with the lowest GHG emissions. In practical engineering appli­
percentage of NH3 leads to uneven mixing of ammonia-diesel
cations, the choice of injection strategy can be tailored according to
fuel, reduces flame propagation speed, prolongs flame combus­
specific requirements.
tion duration, decreases IMEP, and increases unburned NH3

Fig. 33. Comparative analysis of emissions among the four schemes yielding minimal generation of pollutants.

17
C. Shi et al. Energy 296 (2024) 131171

emissions. This situation can be substantially improved by opti­ [2] Wang X, Gao J, Chen H, Chen Z, Zhang P, Chen Z. Diesel/methanol dual-fuel
combustion: an assessment of soot nanostructure and oxidation reactivity. Fuel
mizing the diesel injection strategy.
Process Technol 2022;237:107464.
(2) Through a comparative analysis of engine combustion and [3] Chen Z, Li K, Liu J, Wang X, Jiang S, Zhang C. Optimal design of glucose solution
emissions under different injection schemes, it was observed that emulsified diesel and its effects on the performance and emissions of a diesel
changes in injection strategy alter the spatial distribution of engine. Fuel 2015;157:9–15.
[4] Chen Z, Wang L, Wei Z, Wang Y, Deng J. Effect of components on the
diesel within the combustion chamber. Advancing the injection emulsification characteristic of glucose solution emulsified heavy fuel oil. Energy
timing increases the mixing time of diesel and ammonia, while 2022;244:123147.
changes in the injection angle alter the trajectory of diesel spray. [5] Ayhan V, Ece YM. New application to reduce NOx emissions of diesel engines:
electronically controlled direct water injection at compression stroke. Appl Energy
Proper coordination of injection timing and angle facilitates 2020;260:114328.
uniform combustion of ammonia in the combustion chamber, [6] Sachuthananthan B, Vinoth R, Satya R, Sandeep B, Sudheer N, Deekshith C. Study
achieving near-premixed combustion effects and reducing com­ on the use of selective catalytic reduction technique for NOx emission reduction in
an diesel engine fuelled with Methyl ester of Water Hyacinth. Mater Today Proc
bustion time. 2022;68:1415–21.
(3) Following adjustments to the diesel injection strategy, the mixing [7] Fayyazbakhsh A, Bell ML, Zhu X, Mei X, Koutný M, Hajinajaf N, et al. Engine
efficiency between diesel and ammonia increased, resulting in emissions with air pollutants and greenhouse gases and their control technologies.
J Clean Prod 2022;376:134260.
more uniform ammonia combustion within the combustion [8] Bao J, Qu P, Wang H, Zhou C, Zhang L, Shi C. Implementation of various bowl
chamber and significantly impacting unburned ammonia emis­ designs in an HPDI natural gas engine focused on performance and pollutant
sions. Ultimately, the lowest unburned ammonia emissions were emissions. Chemosphere 2022;303:135275.
[9] Wang H, Ji C, Wang D, Wang Z, Yang J, Meng H, et al. Investigation on the
achieved in the injection scheme with a − 62◦ ATDC injection
potential of using carbon-free ammonia and hydrogen in small-scaled Wankel
timing and 80◦ injection angle, measuring at 0.279 g/kW⋅h. rotary engines. Energy 2023;283:129166.
Merely 1.63% of the original injection scheme. However, in [10] Li J, Lai S, Chen D, Wu R, Kobayashi N, Deng L, et al. A review on combustion
schemes with lower unburned ammonia emissions, flame devel­ characteristics of ammonia as a carbon-free fuel. Front Energy Res 2021;9:760356.
[11] Wang H, Ji C, Shi C, Yang J, Wang S, Ge Y, et al. Multi-objective optimization of a
opment was faster, leading to an increase in cylinder temperature hydrogen-fueled Wankel rotary engine based on machine learning and genetic
and subsequent elevated NOx emissions, necessitating the algorithm. Energy 2023;263:125961.
implementation of effective post-treatment methods to mitigate [12] Kurien C, Mittal M. Review on the production and utilization of green ammonia as
an alternate fuel in dual-fuel compression ignition engines. Energy Convers Manag
NOx emissions. 2022;251:114990.
(4) Under different injection schemes, GHG emissions all showed a [13] Liu H, Ampah JD, Zhao Y, Sun X, Xu L, Jiang X, et al. A perspective on the
decrease. Given the same %NH3, GHG emissions were principally overarching role of hydrogen, ammonia, and methanol carbon-neutral fuels
towards net zero emission in the next three decades. Energies 2023;16(1):280.
influenced by N2O emissions. The lowest GHG emissions were [14] Sekhar SJ, Al-Shahri ASA, Glivin G, Le T, Mathimani T. A critical review of the
observed at − 30◦ ATDC injection timing and 50◦ injection angle, state-of-the-art green ammonia production technologies-mechanism, advancement,
amounting to 450 g/kW⋅h. This represented a 34.3% reduction challenges, and future potential. Fuel 2024;358:130307.
[15] Wen M, Liu H, Cui Y, Ming Z, Feng L, Wang G, et al. Study on combustion stability
compared to the pure diesel mode. At this point, the engine’s and flame development of ammonia/n-heptane dual fuel using multiple optical
IMEP was 0.97 MPa, indicating only a 4% loss compared to the diagnostics and chemical kinetic analyses. J Clean Prod 2023;428:139412.
pure diesel scheme. [16] Qi Y, Liu W, Liu S, Wang W, Peng Y, Wang Z. A review on ammonia-hydrogen
fueled internal combustion engines. eTransportation 2023;18:100288.
[17] Comotti M, Frigo S. Hydrogen generation system for ammonia–hydrogen fuelled
CRediT authorship contribution statement internal combustion engines. Int J Hydrogen Energy 2015;40(33):10673–86.
[18] Gray JT, Dimitroff E, Meckel NT, Quillian Jr R. Ammonia fuel—engine
compatibility and combustion. SAE Trans 1967:785–807.
Cheng Shi: Writing – original draft, Project administration, Meth­ [19] Reiter AJ, Kong S-C. Demonstration of compression-ignition engine combustion
odology, Funding acquisition, Conceptualization. Zheng Zhang: using ammonia in reducing greenhouse gas emissions. Energy & Fuels 2008;22(5):
Writing – original draft, Validation, Software, Data curation. Huaiyu 2963–71.
[20] Reiter AJ, Kong S-C. Combustion and emissions characteristics of compression-
Wang: Validation, Investigation, Formal analysis. Jingyi Wang: ignition engine using dual ammonia-diesel fuel. Fuel 2011;90(1):87–97.
Writing – original draft, Formal analysis. Tengfei Cheng: Software, [21] Niki Y, Yoo DH, Hirata K, Sekiguchi H. Effects of ammonia gas mixed into intake
Investigation. Liang Zhang: Supervision, Resources, Investigation. air on combustion and emissions characteristics in diesel engine. ASME 2016
Internal Combustion Engine Division Fall Technical Conference 2016.
[22] Nadimi E, Przybyła G, Lewandowski MT, Adamczyk W. Effects of ammonia on
combustion, emissions, and performance of the ammonia/diesel dual-fuel
Declaration of competing interest compression ignition engine. J Energy Inst 2023;107:101158.
[23] Sun X, Li M, Li J, Duan X, Wang C, Luo W, Liu H, Liu J. Nitrogen oxides and
ammonia removal analysis based on three-dimensional ammonia-diesel dual fuel
The authors declare that they have no known competing financial engine coupled with one-dimensional SCR model. Energies 2023;16(2):908.
interests or personal relationships that could have appeared to influence [24] Mi S, Wu H, Pei X, Liu C, Zheng L, Zhao W, et al. Potential of ammonia energy
fraction and diesel pilot-injection strategy on improving combustion and emission
the work reported in this paper. performance in an ammonia-diesel dual fuel engine. Fuel 2023;343:127889.
[25] Yousefi A, Guo H, Dev S, Liko B, Lafrance S. Effects of ammonia energy fraction and
Data availability diesel injection timing on combustion and emissions of an ammonia/diesel dual-
fuel engine. Fuel 2022;314:122723.
[26] Yousefi A, Guo H, Dev S, Lafrance S, Liko B. A study on split diesel injection on
Data will be made available on request. thermal efficiency and emissions of an ammonia/diesel dual-fuel engine. Fuel
2022;316:123412.
[27] Raj AGS, Mishra CS. Simulation and experimental data resemblance of darmstadt
Acknowledgments spark ignition engine with different turbulence models – a computational fluid
dynamics cold flow data. Data Brief 2022;43:108340.
The present work was funded by Chunhui Project Foundation of the [28] 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
Education Department of China (Grant No. HZKY20220239) and Sci­ wide conditions based on advanced data analysis techniques. Atomization Sprays
ence Research Project of Hebei Education Department (Grant No. 2022;32(6).
QN2023224). [29] Bao J, Wang H, Wang R, Wang Q, Di L, Shi C. Comparative experimental study on
macroscopic spray characteristics of various oxygenated diesel fuels. Energy Sci
Eng 2023;11(5):1579–88.
References [30] Schmidt DP, Rutland CJ. A new droplet collision algorithm. J Comput Phys 2000;
164(1):62–80.
[31] Han Z, Xu Z, Trigui N. Spray/wall interaction models for multidimensional engine
[1] Zhao Y, Su Q, Li B, Zhang Y, Wang X, Zhao H, et al. Have those countries declaring
simulation. Int J Engine Res 2000;1(1):127–46.
“zero carbon” or “carbon neutral” climate goals achieved carbon emissions-
economic growth decoupling? J Clean Prod 2022;363:132450.

18
C. Shi et al. Energy 296 (2024) 131171

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

19

You might also like