Next Article in Journal
Advancements in Wind Farm Control: Modelling and Multi-Objective Optimization Through Yaw-Based Wake Steering
Previous Article in Journal
CO2 Geothermal Power Generation: Laboratory Experiment on the Interaction Between Carbonated Water and Rishiri Island Basalt in the Vicinity of Injection Wells
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Review of Ammonia Oxy-Combustion Technologies: Fundamental Research and Its Various Applications

1
Graduate Program, Department of Energy & Mechanical Engineering, Gyeongsang National University, Tongyeong 53064, Republic of Korea
2
Department of Smart Energy & Mechanical Engineering, Gyeongsang National University, Tongyeong 53064, Republic of Korea
*
Author to whom correspondence should be addressed.
Energies 2025, 18(9), 2252; https://doi.org/10.3390/en18092252
Submission received: 26 March 2025 / Revised: 23 April 2025 / Accepted: 26 April 2025 / Published: 28 April 2025
(This article belongs to the Section B: Energy and Environment)

Abstract

:
The combustion of ammonia with oxygen presents a promising pathway for global energy transformation using carbon dioxide-neutral energy solutions and carbon capture. Ammonia, a carbon-free fuel, offers several benefits, owing to its non-explosive nature, high octane rating, and ease of storage and distribution. However, challenges such as low flammability and excessive nitrogen oxide (NOx) emissions must be addressed. This paper explores the recent advances in ammonia oxy-combustion and highlights recent experimental and numerical research on NOx emission traits, combustion, and flame propagation across various applications, including gas furnaces, internal combustion engines, and boilers. Furthermore, this review discusses the diverse approaches to overcoming the challenges of ammonia combustion, including oxygen enrichment, fuel blending, plasma assistance, preheating, multiple injections, and burner design modifications. By summarizing the advancements in ammonia oxy-combustion investigation, this paper aims to provide valuable insights that can serve as reference information for prospective ammonia oxy-combustion research and applications toward the transition to sustainable energy.

1. Introduction

Excessive reliance on fossil fuels, including propane, natural gas, coal, and petroleum, in the combustion processes of power plants, industrial processes, and transportation is the primary source of CO2 emissions that significantly contribute to the current level of greenhouse gas (GHG) emissions [1,2]. Accelerating the transition to a reasonable, reliable, and sustainable energy framework aimed at achieving CO2-neutral energy processes is crucial, as emphasized by the sustainable development goals outlined by the United Nations [3]. The energy industry will face significant challenges and opportunities as global awareness of climate change increases and zero-carbon targets are established. Transitions to eliminate fossil fuel dependence are essential for achieving these goals [4]. The Paris Agreement calls for a collective focus on reducing GHG emissions and climate change impacts by setting carbon neutrality targets [5]. Strategies for mitigating the effects of climate change include implementing an emission trading system, applying carbon credits, and establishing carbon taxes to encourage sustainable practices through innovation and investment in green technologies [6]. A tax of USD 40 per ton of CO2, which covers 30% of total emissions, is estimated to reduce cumulative emissions by 4–6% [7], as depicted in Figure 1. Additionally, environmental taxes significantly affect green development goals and technological inventions in high- and middle-income countries. By learning from developed countries, developing countries can accelerate their transitions to advanced technologies [8].
The most important strategy to reduce carbon emissions involves utilizing renewable energy. Solar, wind, and hydroelectric energy are the most used renewable sources. However, due to the lack of continuity in terms of reliability, the performance of these renewable energy sources depends on natural conditions. It may not be easy to use directly for power generation because it will require large-scale batteries and challenging long-distance transportation [9]. Therefore, an energy transition is necessary to reduce GHG emissions and even to prevent the damaging effects of reliance on traditional fuels. By embracing green energy production and innovative technologies, the community can rapidly develop alternative energy sources, such as low-carbon fuels, hydrogen, and ammonia [10,11,12]. Carbon-free fuels promise to reduce fossil energy consumption and to contribute significantly to energy security by driving a well-established renewable energy infrastructure that encloses industrial decarbonization, supply chains, and regional development [13].
Among the three examples of carbon-free fuels mentioned above, ammonia has potential applications in the energy sector as a decarbonizing fuel, hydrogen vector, and energy storage medium. It has several desirable properties, such as a hydrogen capacity of 17.8% by mass and a high energy density per unit volume, establishing ammonia as an ideal hydrogen carrier and ensuring safety, ease of storage and transport, and zero carbon emissions by the end of the process [14,15,16]. The production process is relatively straightforward and efficient, and it utilizes the Haber–Bosch (HB) method for ammonia synthesis [17]. Ammonia is used globally as an energy source in gas turbines, industrial furnaces in thermal power plants, and internal combustion engines (ICEs) in marine transportation [18]. Compared with other fuels (hydrogen, methanol, dimethyl ether, gasoline, and diesel), ammonia has an extensive operating range and high power output [19]. Ammonia is the primary carbon-free fuel for decarbonizing in marine transportation. Ammonia and hydrogen contribute over 60% of generated energy, enabling the 2050 Net Zero Emissions scenario [20]. Ammonia can be mixed with other fuels to assist coal-fired power plants in reducing carbon pollution through co-combustion with up to 60% ammonia content in the fuel. The evolution to 50–60% ammonia co-firing in the 2030s was predicted to achieve full ammonia firing by the 2040s [18].
Although ammonia can be used directly or substituted with fossil fuels in combustion systems, several fundamental issues exist because of its low flammability, low laminar burning velocity (LBV), elevated nitrogen oxide (NOx) emissions, and weak radiation intensity [14]. Furthermore, the presence of water can corrode fuel systems [21]. Research has been conducted to address these issues [22]. The corrosion problems can be addressed using advanced coatings, such as composite silane coatings, or materials resistant to ammonia corrosion at high temperatures, such as 15CrMoG steel [23]. Considerable efforts have focused on comprehending the mechanism of NOx generation and advances in NOx emission reduction technologies, such as staged combustion, oscillating combustion, selective catalytic reduction, moderate or intense low-oxygen dilution (MILD), and preheating [2,24].
A widely recognized approach to enhancing the combustion rate of ammonia involves reactivity stratification using a high-reactivity fuel such as hydrogen [25]. Oxygen-enriched combustion is used in combustion systems to tackle issues in the combustion characteristics of ammonia by increasing the LBV of ammonia and the adiabatic temperature due to rising oxygen levels in the combustion process [26]. This process is referred to as oxy-combustion, where pure O2 is used as a substitute for air through CO2 recycling [27]. The addition of oxygen affects the combustion system by reducing the mass flow of exhaust gases and increasing the adiabatic temperature, which significantly increases the efficiency [28,29]. Although ammonia oxy-combustion offers considerable advantages, challenges remain, including the boost in NOx emissions due to the elevated maximum temperature in ammonia combustion [30,31].
The previous paragraph discussed the urgency of emission mitigation, which is a priority to be addressed along with the development of the industry. Along with this development, the solutions offered to suppress GHG emissions are becoming increasingly progressive. As is known, the involvement of ammonia in a combustion system still has several obstacles that need to be addressed. Most recent literature reviews have focused on ammonia combustion characteristics and applications [32,33,34,35,36,37,38,39,40,41,42] or a combination of ammonia and other fuel blends, especially with hydrogen, while often neglecting ammonia oxy-fuel combustion. This paper contributes significantly to recent research developments and provides crucial insights for future ammonia oxygen combustion studies. Currently, a full assessment of energy concerns focuses primarily on ammonia as a carbon-free fuel, as discussed in Section 1. Section 2 briefly summarizes the properties, advantages, and downsides. Section 3 examines the progress in research on ammonia oxy-fuel combustion. Section 4 highlights the characteristics of ammonia oxy-fuel combustion across various applications, and Section 5 highlights the progress and forecasts the prospective research and development patterns of ammonia oxy-fuel combustion.

2. Ammonia Fuel Combustion Characteristics

2.1. Advantages of Ammonia Fuel Combustion

Ammonia has been predicted as an ammonia energy storage system for hydrogen and other effective energy carriers, owing to its high gravimetric hydrogen density of 17.8 wt% [43,44]. Methanol is a strong competitor to ammonia because of its capacity to store significant amounts of energy and ability to facilitate long-distance transmission. Nonetheless, utilizing methanol for hydrogen storage has environmental consequences because it releases CO2 when used directly or after degradation [45,46]. Meanwhile, ammonia is usually transported in liquid form because of its high density. It can be stored at −33 °C below ambient pressure or at room temperature at approximately 10 bar. Using refrigerants is effective for large-volume storage [15]. The infrastructure for handling and transporting ammonia is well established, enabling efficient global transmission using trucks, trains, barges, vessels, and pipelines [47]. Table 1 provides a comparative overview, along with the chemical kinetics investigation results for the explosive properties of NH3. Ammonia has a narrow explosive range and the lowest explosive reactivity among fuels such as hydrogen and methane [48]. It can be directly utilized in combustion without generating CO2 emissions, leading to carbon-neutral energy systems. The reaction of an ammonia–air mixture under stoichiometric conditions can be expressed as follows [49]:
N H 3 + 0.75 ( O 2 + 3.76 N 2 ) 1.5 H 2 O + 3.32 N 2 .
Table 1 shows that ammonia has an octane number of 130, higher than those of the other fuels. An engine fueled by ammonia has the potential to reduce the chances of knocking, operate effectively at higher compression ratios, and enhance thermal efficiency [54]. Investigations were conducted in a rapid compression engine to understand the effects of ammonia’s addition on isooctane knock combustion. The observed data imply that the intensity of knocking decreases as the mole fraction of ammonia increases to 0% (A0), 40% (A40), and 80% (A80) while maintaining consistent thermodynamic conditions, as illustrated in Figure 2. The effect of the initial pressure on the knock intensity was noted [55].
More than 98% of the ammonia produced globally in significant amounts is generated using the HB method, which involves the catalytic interaction of nitrogen and hydrogen at high pressures and temperatures [56]. Ammonia can be classified into several types based on its production process. “Brown ammonia” refers to the coal gasification method for producing hydrogen. “Gray ammonia” is ammonia delivered from hydrogen extracted from methane. “Blue ammonia” resembles gray ammonia, but it integrates carbon capture, utilization, and storage technology into its production process. These three production processes use fossil fuels that generate GHG emissions. However, more sustainable ammonia production can be achieved by using renewable energy as an electricity source, namely “green ammonia” [57,58,59]. The HB process is used to synthesize green ammonia by combining H2 produced through the electrolysis of H2O with N2 extracted from the air in an air separation unit [60].

2.2. Challenges in Ammonia Fuel Combustion

Despite its advantages, ammonia utilization presents numerous challenges because of its unique characteristics. Ammonia toxicity is a crucial safety concern because ammonia can be hazardous if leaked. Ammonia poisoning represents the most significant health risk in the event of a leak [61].
Fuel-bond NOx emissions and the amount of unburned ammonia in the exhaust gas are other critical challenges in ammonia combustion. NOx has adverse effects on human health by contributing to the formation of acid rain and has substantial GHG consequences [24,62]. NOx generation during ammonia combustion is a convoluted process controlled by several parameters, including the temperature, pressure, and oxygen ratio in the combustion zone [63]. During the burning process, the nitrogen atoms in the ammonia fuel react with oxygen to assemble NO molecules, which indicates that NO formation can be minimized by limiting the oxygen level available to participate in the reaction in Equation (2). The oxygen-to-fuel ratio decreases as the equivalence ratio increases [64]:
N H i + O X N O + H i X
The low combustion efficiency of ammonia as a fuel presents an immediate challenge due to its lower heating value and higher heating value. Moreover, its LBV is significantly lower than those of other hydrocarbons, suggesting that the energy generated from burning ammonia is significantly less intense [65]. Additionally, the high auto-ignition temperature of ammonia (650 °C), which is higher than those of hydrogen (520 °C), propane (450 °C), and methane (630 °C), makes it challenging to ignite and results in low flammability. Nevertheless, the LBV of ammonia can be improved by integrating it with enriched oxygen, owing to the increased reactivity rates of OH, H, O, and NH2 molecules in the reaction zone at a higher O2 content [26]. Figure 3 plots the LBV at different O2 concentrations.

3. Fundamental Research on Ammonia Fuel Combustion

This section is divided into two subsections. Section 3.1 reviews recent trends in experimental techniques, and Section 3.2 discusses modeling using chemical kinetic analysis.

3.1. Experimental Studies

Several studies on flame propagation in ammonia combustion have been conducted to investigate the fuel characteristics. Ichimura et al. [66] investigated the dynamics of turbulent flame propagation during ammonia combustion. The fan-stirred chamber was tested using a mixture of ammonia and air at diverse equivalence ratios (Φ) ranging from 0.6 to 1.3. Figure 4 presents Schlieren images of ammonia–air mixture burning at Φ values of 0.8, 1.0, and 1.2. Figure 5 indicates that a fuel-lean mixture at Φ = 0.9 can maintain propagation at the most remarkable turbulence intensity caused by Lewis’s influence, despite the LBV peaking at a Φ value of approximately 1.1, as shown in Figure 6 [66].
Karan et al. [67] reported experimental data on oxy-ammonia combustion at elevated temperatures and pressures in a constant-volume spherical chamber to evaluate the LBV under controlled conditions. Spherical flames spreading outward in a constant-volume chamber provide a method for measuring the LBV. Tests were performed using ammonia–oxygen mixtures with equivalence ratios of 0.8, 1.1, and 1.3. These ratios represent the lean, stoichiometric, and rich combustion states at an initial temperature of 300 K and pressure of 1–4 bar. The analysis revealed that the LBV obtained by Nakamura and Stagni closely matched the experimental results. The oxygen level in the mixture determines the LBV. Lean mixtures tend to have more efficient combustion processes, leading to LBVs higher than those of rich mixtures, with the maximum LBV observed at Φ = 1.1. Mei et al. [68] also analyzed the LBV of an ammonia/oxygen/nitrogen combustion mixture under oxygen enrichment and pressure consequences. They remarked that oxygen enrichment enhanced flame propagation and reduced buoyancy by increasing the adiabatic flame temperature and radical concentrations of H, OH, and NH2. Additionally, they found that increased pressure could suppress the LBV, owing to reduced mixture reactivity.
Zhang et al. [69] experimentally and numerically investigated the LBV and emission characteristics of a Bunsen burner under oxy-ammonia combustion conditions (100% oxygen). Figure 7 illustrates the experimental set-up. They aimed to estimate the LBV in the surroundings of a constant-volume combustion bomb by utilizing a Bunsen burner with an output diameter of 6 mm and an area contraction coefficient of 178. This methodology assumes that the LBV remains uniform across the entire surface area of the flame. The LBV can be determined by applying the following mass conservation balance:
S L = ρ b A b S b ρ u A b = Q A b ,
where Q is the total volumetric flow rate, Ab is the intended flame surface area, and ρb and ρu represent the densities of burned and unburned reactants, respectively. Ab was calculated from the NH* distribution image captured by an intensified high-speed CMOS camera system, which is a better method than the Schlieren technique. The impact of the preheating method was compared across a temperature ( T u , the temperature at 50 mm upstream of the outlet) range of 298–520 K and a Φ range of 0.7–1.6. Figure 8 shows that at ambient temperature and pressure, the highest possible LBV was 1.25 m/s, which is approximately 18 times greater than that of NH3/air combustion, attributed to the accelerated reaction dynamics of OH, H, O, and NH2 radicals within the reaction zone.
Co-firing ammonia with other fuels, particularly using oxy-combustion technology, is an innovative area of research and application. Okafor et al. [70] reported the LBVs of ammonia and methane in premixed air flames using experimental and computational methodologies. The tests were performed in a cylindrical combustion chamber with a constant volume. The chamber had an inner diameter of 270 mm and a length of 410 mm. Two quartz windows 60 mm in size were installed at opposite sides to provide optical access to the chamber. Schlieren photography was employed to observe and record the flame propagation over the quartz chamber windows. A high-speed digital camera was used while the equivalency ratio was changed from 0.8 to 1.3, and the ammonia concentration varied from 0 to 0.3 based on the heat fraction of the fuel. These findings demonstrate that the unstretched LBV diminished as the ammonia concentration increased. This highlights the significant effect of ammonia on the combustion characteristics of the combination.
Liu et al. [71] presented experimental and modeling findings on the laminar flame speed of a blending fuel of methane and ammonia under oxy-combustion conditions. They explored varying the NH3/CH4 ratio from 0.1 to 0.2, the CO2 mole fraction from 45% to 65%, and the O2 mole fraction from 35% to 40%, all within a two-burner counter flow arrangement. The bottom burner released a premixed combustible mixture of CH4, NH3, O2, and CO2, whereas the top burner emitted N2. These experimental results were compared with the computational results from three chemical kinetic models: the HUST, Okafor, and Mendiara mechanisms. Notably, the Mendiara mechanism consistently yielded results that exceeded the experimental findings. In contrast, the Okafor mechanism tended to underpredict the laminar flame speeds at an NH3/CH4 (II) ratio of 0.1 when analyzing O2/CO2 mixtures with a 35:65 ratio. Regardless, the HUST mechanism demonstrated its effectiveness as a dependable predictor of all experimental outcomes.
Li et al. [72] inspected the LBV and flame propagation characteristics of two clean-energy blends, namely ammonia and hydrogen, during oxygen combustion under varying conditions. Their findings emphasized the importance of Φ and the initial pressure in defining combustion performance. The analysis demonstrated that the LBVs of the ammonia/hydrogen/oxygen mixtures increased with an increasing fuel ratio. The LBV increased significantly with the hydrogen content in the blend, emphasizing the role of hydrogen in enhancing the LBV. Increasing the hydrogen-to-ammonia ratio improved the flame stability and LBV. The highest LBV was approximately 0.8, which deviated slightly from the stoichiometric ratio, owing to the complex interactions between the flame temperature and diffusion. Li et al. [73] highlighted the flame propagation and inherent instability mechanisms in hydrogen–ammonia–oxygen combustion. When the fuel ratio was gradually increased, the Lewis number (diffusional–thermal instability) shifted from one, and the hydrodynamic instability was limited. In addition, an increase in pressure affected this instability. Moreover, the flame thickness of the hydrogen–ammonia combustion in air was higher than that of the oxygen combustion under the same operating conditions, indicating that the oxygen atmosphere significantly increased the hydrodynamic instability of the hydrogen–ammonia-blended flame.
Shi et al. [74] explored the LBV and combustion characteristics of co-firing reactive hydrocarbons and oxygenated fuels, namely ammonia with dimethyl ether (DME) (CH3OCH3), in an O2/CO2 atmosphere. The observed LBV of NH3/O2/CO2 increased significantly, exceeding 65 cm/s when γ reached 70%. In contrast, higher oxygen levels significantly increased the LBV propagation rate of NH3/DME, as shown in Figure 9. Furthermore, the LBV of (NH3/DME)/(O2/CO2) typically increased with the DME concentration, despite its comparatively low diffusivity (Figure 10). Although chemical effects primarily enhance the LBV of NH3 in an O2/CO2 environment with DME co-combustion, thermal and transport effects have a minimal impact.
Barbas et al. [75] conducted experimental and computational investigations on the formation and emission of CO and NO by NH3 doping during oxy-methane combustion. Under the control conditions, the excess oxygen coefficient was evaluated to be 1.05, 1.1, 1.15, 1.2, and 1.25. The mass flow rate of CO2 was adjusted accordingly for each excess oxygen coefficient. The tested oxidizer arrangements retained blends of 21% O2 with 79% CO2, 23% O2 with 77% CO2, 25% O2 with 75% CO2, 27% O2 with 73% CO2, and 29% O2 with 71% CO2, in addition to air. The research revealed that NH3 significantly contributes to NO production through HNO intermediates, such as H N O + H H 2 + N O , which are essential to the process. During oxy-fuel combustion, NH3 transforms the nitrogen bound in the fuel into NO. The study also revealed that NO generation is higher for oxidizers with a higher oxygen content in oxy-fuel applications. Nevertheless, it decreases slightly as the excess oxygen coefficient increases. The addition of ammonia to methane combustion under oxy-fuel conditions considerably impacts NO emissions while affecting CO build pathways, owing to altered radical dynamics and competition. A particularly elevated oxygen coefficient increases CO emissions while decreasing NO emissions, as shown in Figure 11 and Figure 12.
Wu et al. [76] provided further insights into the formation of NO from the co-firing of hydroxylated fuels, namely methanol (CH3OH) and ethanol (C2H5OH), with NH3 oxidation in diverse oxy-fuel atmospheres by simulating the formation and reduction of NO during the oxidation of CH3OH/NH3 and C2H5OH/NH3 in O2/CO2, O2/H2O, and O2/CO2/H2O atmospheres under oxy-fuel conditions. The authors found that the presence of H2O and CO2 affected NO generation. In particular, the addition of H2O enhanced NO reduction more effectively than the addition of CO2. At higher oxygen-to-fuel ratios, the reduction effects of H2O and CO2 were more pronounced and stronger for C2H5OH than for CH3OH.

3.2. Numerical Study

Kinetic modeling was implemented to determine the unstretched LBV, species concentrations, and species production rates. Additionally, it contributed to the sensitivity analysis and emission generation during the combustion process. This method requires advanced process details and optimization. The experimental LBV results were validated to understand the chemical kinetics mechanisms. These mechanisms were utilized in Chemkin-Pro to model the LBV and to compare the modeling results with the experimental data. Various simulations were performed under isobaric and isothermal conditions. The results of these simulations verified that the mechanisms were driven by temperature rather than pressure because the change in LBV increased with the temperature but showed limited sensitivity to pressure shifts beyond 5 bar [67].
The experimental results were further analyzed using the PREMIX and OPPOSED code packages. Figure 13 and Figure 14 illustrate that the LBV exhibits a linear increase with an increasing O2 concentration or decreasing CO2 level under specified conditions [71].
In addition to analyzing the propagating flame sub-model, the nitrogen conversion pathway was examined using Chemkin-Pro. The analysis revealed that NO emissions from NH3/O2 combustion under stoichiometric conditions were approximately 7216 ppmv at 15% O2. This level was approximately 1.5 times higher than that of NH3/air combustion and 13 times greater than that of CH4/air combustion. As illustrated in Figure 15, the mole fraction and N-normalized mole fraction of NO increased with Φ, and preheating also contributed to an increase in NO emissions [69].
Okafor et al. [70] conducted simulations to replicate experimental results and to analyze the LBV and species concentration. Various detailed combustion mechanisms were utilized for the numerical simulations, including GRI Mech 3.0, Mendiara Mech, and Tian Mech. The study revealed that the Tian Mech mechanism significantly underestimated the unstretched LBV, particularly at lower ammonia concentrations. In contrast, GRI Mech 3.0 provided predictions closer to the experimental results but lacked some crucial ammonia oxidation steps for high ammonia content flames. A new detailed chemical kinetic model was developed to address these limitations by combining elements from GRI Mech 3.0 and Tian Mech. This new model includes 59 species and 356 elementary reactions. The model was validated against experimental data and demonstrated improved accuracy in predicting the LBV and NO concentration in flames. Li et al. [72] highlighted the use of ammonia–hydrogen blends, stating that although the LBV increased with a higher hydrogen ratio, NOx emissions, primarily NO emissions, decreased under rich combustion conditions. Figure 16 illustrates the mole fractions of reactants and products in a hydrogen/ammonia/oxygen flame under different Φ values, providing significant insights. H2, NH3, and O2 were the primary reactants, whereas H2O, N2, and NO are the primary products. NH3 was consumed the fastest among the reactants, and the concentrations of H2 and O2 stabilized after a certain amount was consumed. The most significant NOx-producing substance in the mixed fuel was NO. The amount of NO produced decreased gradually with increasing Φ values. A higher Φ value reduced NO emissions, achieving maximum NO reduction under fuel-rich conditions. This characteristic makes ammonia–hydrogen blends promising carbon-free fuel alternatives for industrial and energy applications.
Table 2 summarizes the research developments through experimental methods and chemical kinetics simulations to describe the ammonia oxy-fuel combustion process in the above research.
Table 2. The recent literature on experimental and kinetic modeling studies of ammonia oxy-fuel combustion.
Table 2. The recent literature on experimental and kinetic modeling studies of ammonia oxy-fuel combustion.
AuthorsResearch MethodCombustion ConditionsApproach (Injector Equipment, Methods, etc.)Key Findings
Ichimura et al. [66]Experiment
  • Initial pressure and temperature of 1 atm and 298 K, respectively
  • Φ = 0.6–1.3
  • Ammonia air combustion
  • Constant-volume chamber
  • Schlieren technique
  • The maximum intensity of turbulence occurred at Φ = 0.9.
  • The highest LBV was achieved at Φ = 1.1.
Karan et al. [67]Experiment and simulation
  • Initial pressure and temperature of 1–4 bar and 300 K, respectively
  • Φ = 0.8, 1.1, and 1.3
  • Ammonia oxy-combustion
  • Constant-volume spherical chamber
  • CMOS technique
  • The maximum LBV occurred at Φ = 1.1.
  • The variation in LBV increased significantly with temperature.
Mei et al. [68]Experiment and simulation
  • Initial pressure and temperature of 1–5 atm and 298 K, respectively
  • Φ = 0.6–1.6
  • Ammonia/O2/N2 (oxygen concentrations of 25–45%)
  • Premixed combustion
  • A high-pressure, constant-volume cylindrical combustion vessel
  • Schlieren technique
  • Increased oxygen content led to a higher LBV while minimizing buoyancy effects.
  • The LBV decreased with increasing pressure.
Zhang et al. [69]Experiment and simulation
  • Preheating temperature (298–520 K)
  • Φ = 0.7–1.6
  • Ammonia oxy-combustion
  • Bunsen burner
  • Constant-volume chamber
  • ICMOS technique
  • The LBV of a flame was 1.25 m/s, which is approximately 18 times that of NH3/air combustion.
  • NO emissions increased with Φ, and preheating increased.
Okafor et al. [70]Experiment and simulation
  • Initial pressure and temperature of 0.10 MPa and 298 K, respectively
  • Φ = 0.8, 1.1, and 1.3
  • NH3, CH4, air combustion
  • Constant-volume chamber
  • Schlieren technique
  • The LBV decreased as the ammonia concentration increased.
  • GRI Mech 3.0 and the Tian Mech model improved the accuracy in predicting the LBV and NO concentration in flames.
Liu et al. [71]Experiment and simulation
  • Initial pressure and temperature of 1 atm and 300 ± 2 K, respectively
  • NH3/CH4 ratio: 0.1–0.2; CO2 mole fraction: 45–65%; O2 mole fraction: 35–40%
  • CH4/NH3/O2/CO2 combustion
  • Two-burner counterflow
  • DPIV technique
  • The HUST mechanism was considered practical for estimating the LBV.
  • The LBV increased significantly with higher O2 or decreased CO2 levels.
Li et al. [72]Experiment and simulation
  • Initial pressure of 0.1–1.0 atm
  • Φ = 0.5–1.5
  • Fuel ratio: 0.5–2
  • NH3/H2 oxy-combustion
  • Cylindrical constant-volume chamber
  • Z-arranged Schlieren technique
  • The maximum flame was achieved near Φ = 0.8.
  • Maximum NO reduction was reached under fuel-rich conditions.
Li et al. [73]Experiment
  • Initial pressure of 0.1–1.0 atm
  • Φ = 0.5–1.5
  • Fuel ratio: 0.5–2
  • NH3/H2 oxy-combustion
  • Cylindrical constant-volume chamber
  • Z-arranged Schlieren technique
  • The combustion in oxygen significantly enhanced the hydrodynamic instability.
Shi et al. [74]Experiment and simulation
  • Initial pressure and temperature of up to 5 atm and 373 K, respectively
  • Φ = 0.7–1.5
  • NH3/DME in O2/CO2 combustion
  • Constant-volume chamber
  • Schlieren technique
  • Oxygen enrichment accelerated the LBV propagation.
  • The LBV of the (NH3/DME)/(O2/CO2) mixture demonstrated a notable increase as the proportion of DME rose.
Barbas et al. [75]Experiment and simulation
  • Operating at atmospheric pressure
  • Excess oxygen coefficients (1.05, 1.1, 1.15, 1.2, and 1.25)
  • NH3 doped in CH4 oxy-combustion
  • Water-cooled burner
  • NH3 significantly contributed to producing NO via HNO intermediates.
  • A specific excess oxygen coefficient led to higher CO emissions and lower NO emissions.
Wu et al. [76]Simulation
  • Oxygen concentration: 30%
  • Oxygen-to-fuel ratio α: 0.8 and 1.2
  • CH3OH/C2H5OH/NH3 in oxy-combustion
  • Plug flow reactor
  • The NO decrease was more significant at high oxygen-fuel levels and with C2H5OH instead of CH3OH.
Both the experimental and simulation research methods reviewed provide important insights into the combustion characteristics of ammonia oxy-combustion, namely the flame stability, LBV, and NOx formation characteristics. The experimental approach offers valuable insights into real flame observation, but this method is limited because it is only usable at small scales and has high costs. On the other hand, modeling approaches such as chemical kinetic methods or computational fluid dynamics (CFD) enable broader and more detailed parametric studies on a larger scale. Both methods have their respective advantages. Experiments have advantages in flame validation and chemical reactivity processes, whereas numerical analysis can predict and optimize combustion processes such as flame characteristics and emission production. Okafor et al. [70] and Li et al. [72] combined experimental and modeling methods and showed significant improvements in the accuracy of kinetic mechanisms and NO formation.

4. Ammonia Oxy-Fuel Combustion in Various Applications

Numerous studies, including experimental and numerical analyses like CFD investigations, have assessed the feasibility of employing ammonia energy in oxygen combustion. The numerical CFD method is widely used because it enables more complex analyses and can be applied to larger or even industrial-scale systems, unlike experiments or other numerical methods, which are typically limited to simpler scales and analyses. CFD enables tasks like optimizing or redesigning combustion processes on a much larger scale. CFD can produce high-precision results by employing a range of numerical equations. This section is divided into three subsections. Section 4.1, Section 4.2 and Section 4.3 review the trends in applying ammonia oxy-fuel combustion in gas furnaces, ICEs, and boilers, respectively, with approaches demonstrated according to the actual conditions. The end of each section summarizes the reviewed literature studies, presented in Table 3, Table 4 and Table 5. This paper aims to evaluate the research progress and assess the impacts of various methods on the characteristics of ammonia oxy-fuel combustion.

4.1. Ammonia Oxy-Fuel Combustion in Gas-Fired Furnaces

This subsection focuses on the research on combustion performance and emissions associated with ammonia oxy-combustion within gas furnaces to evaluate the research progress and analyze the advantages and disadvantages of the various methods employed in recent research, the summary is provided in Table 3. El-Adawy and Nemitallah [77] conducted a relevant study on gas furnace conditions and investigated the flame combustion performance and NOx emissions against the inlet mixture temperature and velocity of ammonia oxy-combustion in a gas turbine combustor. They used a flamelet-generated manifold model to deliver an efficient and accurate simulation of complex combustion phenomena, enabling analysis of the temperature distribution, flame dynamics, and emissions. The initial temperature increased with the inlet mixture temperature under stoichiometric conditions because it determined the flame position. In addition, the turbulent flames of NH3 and O2 increased with the inlet mixture velocity, indicating that a higher inlet velocity can lead to enhanced mixing and combustion efficiencies. Meanwhile, the impact of the inlet velocity (Vin) on the NOx emissions was minimal.
Cai et al. [78] systematically analyzed the thermal performance and NO emissions during ammonia and oxygen burning in a microplanar combustor. They examined the effects of various parameters, including Φ, the inlet pressure (Pin), and the inlet temperature (Tin), which affect the outer wall mean temperature (OWMT). The results indicated that NO production is susceptible to Φ. Specifically, decreasing Φ increased the NO emissions due to higher flame temperatures and oxygen atom concentrations. Optimizing Φ can result in drastically lower NO concentrations. Tin was essential in influencing the production of OWMT and NO. Increasing Tin decreased the concentrations of OWMT and NO. By contrast, an increase in Pin suggests that the OWMT also increases, owing to the intensified mixing process that drives the flame downstream at high temperatures. However, less NO was produced during combustion with increasing Pin values.
Zhang et al. [79] presented a concept for non-premixed ammonia/oxygen combustion in a dual-inlet microfluidic system. In their study, a 2D numerical model was used to compare premixed combustion using one channel and non-premixed combustion using two channels, in which oxygen and ammonia entered through different channels into the combustion chamber. Figure 17 shows that the non-premixed flame prevents flashback, particularly when NH3 and O2 are introduced through separate channels. This strategy improved the mixing and combustion stability, optimized the thermal performance, and reduced emissions. The aforementioned paper also discussed the impact of Φ on the combustion performance. The authors concluded that Φ = 0.9 produced the most optimal thermal performance and emission characteristics. Figure 18 and Figure 19 show that higher inlet velocities increased the wall temperature and reduced NO emissions, indicating that enhanced convective heat transfer improved the thermal performance of the combustor.
Cafiero et al. [80] examined the properties and NOx emissions of ammonia and oxygen combustion within a 20 kW semi-industrial furnace operating under MILD conditions. The compliance measure in the study involved varying the O2 enrichment, increasing the operational range from 21% to 50%, and controlling two oxidizer injectors to achieve average inlet velocities of 150 m/s for ID16 and 80 m/s for ID20. The authors observed enhanced reactivity in NH3 combustion due to O2 enrichment under conventional and MILD conditions. Nonetheless, higher O2 levels augmented NOx emissions. Optimizing O2 enrichment with MILD combustion could minimize NH3 destruction while controlling NOx. O2 enrichment of 37% was identified to be effective for MILD combustion, significantly reducing the ammonia drop while maintaining the NOx emissions at an allowable level of 146 ppmvd. Moreover, the effect of the oxidizer velocity for two injector sizes (ID16 and ID20) was tested to determine its impact on the inlet combustion. An elevated inlet oxidizer velocity stabilized the combustion and affected the temperature distribution. NOx emissions were lower with ID16 than with ID20 across all O2 levels at the optimal Φ value of 0.95, which achieved the maximum temperature as the optimal compromise between NOx emissions and ammonia slip. At Φ = 1.01, the minimum NOx emissions were achieved, whereas the NH3 slip was higher than the minimum limit that could be assessed during the experiment, reaching 4000 ppmvd before any corrections. These findings align closely with those obtained by Sato et al. [81], who found that unburnt ammonia became significant at Φ ≥ 1.05.
Plasma assistance is a strategy for improving ammonia ignition and flame propagation under challenging combustion conditions. This research group showed how plasma-enhanced systems can reduce ignition delay and stabilize combustion, which is critical for practical applications. Matveev et al. [82] conducted research on the plasma-assisted burning of ammonia in air and oxygen atmospheres. They utilized ANSYS Chemkin to determine the kinetic temperature and concentration of the combustion product components with pressure variations (1–10 bar) and an excess oxidizer coefficient. This technique uses plasma from a radio frequency plasma torch to stabilize a mixture of ammonia and oxygen, thus efficiently enhancing the combustion process. This stability is achieved by the formation of O, H, and OH radicals while reducing the ignition delay. Extending the combustion pressure reduces NOx emissions during the ammonia combustion process in an air–oxygen setting. Matveev and Serbin [83] conducted follow-up research on plasma-assisted ammonia and oxy-combustion, using 3D modeling to analyze the combustion chamber. Their examination focused on the combustion properties of ammonia. The model employed the eddy dissipation concept, which integrates detailed chemical reactions from the Arrhenius equation with the turbulence in the flame. These findings indicate that the plasma torch is an essential starting point for ammonia ignition. The combustion remained constant even when the oxygen flow rate increased. Predominant NO formation occurred in the central region of the combustion chamber, where ammonia combustion was more intense and the highest temperature was achieved. In contrast, NO2 was produced in areas with lower temperatures.
Advanced burner redesigns, such as bluff body designs, have been explored to assess their effectiveness in stabilizing the flame, modifying the temperature profile, and affecting pollutant formation. This deconstruction provides insights into how burner design affects ammonia combustion performance under oxygen fuel conditions. Cai et al. [84] investigated the incorporation of a bluff body within a microplanar combustor by utilizing premixed ammonia and oxygen. They focused on the effects of thermal efficiency and NOx emissions, as depicted in Figure 20. They employed a 3D model integrating flow dynamics, combustion processes, thermodynamics, and heat transfer. The SIMPLE mode couples the pressure and velocity to evaluate the impact of a bluff body on the combustor’s performance. This analysis included an exploration of the dimensionless (h) influence and the bluff body dimensionless location (l) effect. The findings showed that the presence of a bluff body enhanced the outer wall temperature (OWT) around it, unlike in the absence of a bluff body. Regardless, the temperature at the combustor outlet tended to be lower in the presence of the bluff body. Bluff body application can also reduce NO formation under certain conditions because the ammonia flow rate is adjusted to three distinct values, as depicted in Figure 21. NO production had minimal variation at low fuel flow rates of ≤1000 mL/min. Regardless, at a flow rate of 1200 mL/min, the bluff body caused a notable reduction in the NO concentration. The presence of the bluff body significantly affects the temperature field distribution. The flow field and temperature profile show differences without the bluff body, especially around the bluff body region. Figure 22 implies that, compared to the case without the bluff body, the flame temperature downstream of the bluff body along the flow direction (axial) is substantially reduced. This decrease effectively reduced the rate of NO formation. Increasing h increases the OWT but can also inhibit NO formation in certain instances. However, a higher bluff body height increases the pressure loss. Increasing l can increase the OWT and reduce low NO concentrations, probably because a larger l value intensifies the heat transfer between the blend and the lower surface of the combustor.
The examination of NOx emission performance in ammonia–oxygen premixed combustion with a new design in a microplanar burner that adds a perforated plate was introduced by Cai et al. [85], with the design shown in Figure 23. They used a 3D model validated with experimental data incorporating 22 species and 67 reactions to explore the flow area, impacts of preferential diffusion, and conjugate heat transfer. Several critical findings emerged from this study. Perforated plates expanded the downstream recirculation zone, lowered the temperature and flame rate, and decreased NO formation in the recirculation zone. The primary factors affecting NOx emissions included the flow dynamics, conjugate heat transfer, and preferential diffusion. Conjugate heat transfer analysis considers the interaction between the combustion gases and the burner wall. This is essential because it affects the location of the flame and the rate of chemical reactions, thereby affecting NO formation. The quality of the burner wall material significantly affects the heat transfer, flame containment, and NOx output. Three types of materials were selected: quartz, steel, and nickel. Materials exhibiting high thermal conductivities are particularly effective in facilitating heat transfer among solid products during combustion.
Bektaş et al. [86] dissected the capacities of methane and ammonia under air–oxygen combustion conditions against the influence of different equivalence ratios in a two-inlet cyclone combustion chamber. They revealed that ammonia–oxygen combustion performed better at 1828 K, compared with air–methane at 1538 K or air–ammonia at 994 K. Employing oxygen as an oxidizer enhanced NOx emissions, although fuel-rich conditions had the potential to mitigate NO emissions. Cyclone combustors have prominent characteristics for assembling a strong swirling flow, which can improve the mixing stability and long residence times within the combustion chamber, ultimately leading to more efficient combustion. As shown in Figure 24, during the air–ammonia combustion process, the maximum axial velocity was achieved while maintaining consistent thermal power across all instances. This finding implies that a higher flow rate is introduced into the combustor to correspond to an identical thermal power because ammonia has a lower calorific value than methane. The ideal LBV for ammonia is achieved under fuel-rich conditions, specifically at an equivalence ratio of Φ = 1.1 when using oxy-ammonia. However, combustion that can create a more stable flame condition occurs under stoichiometric conditions. According to a report by Davies et al. [87], unburned NH3 emissions contribute to flame instability, where Φ > 1.1 causes the amount of unburned NH3 to increase so much that the combustion efficiency cannot be calculated. Therefore, combustion under fuel-rich conditions should be reconsidered because it increases the risk of incomplete combustion, leading to hydrocarbon emissions from ammonia slip.
The studies reviewed in this subsection investigated various fuel-blending and MILD combustion techniques to overcome the low laminar combustion velocity of ammonia and its high potential for NOx emissions. This approach is critical to developing optimal and efficient ammonia combustion systems. Zhao et al. [88] studied CH4/NH3 combustion with MILD oxy-combustion technology to improve the thermal efficiency, carbon technology, and NOx reduction in a jet flame in a hot co-flow (JHC) combustor. Their initial study used 2D modeling to examine how CO2 and H2O affect NO formation. The simulation utilized an in situ adaptive tabulation algorithm to calculate the efficiency and stability of the complex heterogeneous chemical kinetics. This study revealed four major NH3-to-NO conversion routes with HNO as the central intermediate and indicated that the significant addition of CO2 facilitated NO formation through reactions involving NH and NH2 radicals. H2O also contributed to NO formation by increasing the number of OH radicals. Compared with the addition of CO2, the addition of H2O led to a smaller increase in the N conversion rate. As illustrated in Figure 25, the emission trends of the three dilutions differed, with MILD-N2 exhibiting the highest emissions followed by disused MILD-CO2, whereas MILD-H2O achieved the lowest emissions.
In a subsequent study, Zhao et al. [89] investigated the effects of O2 co-flow (XO2) and temperature (Tcof) on the combustion system and fuel NO formation in a CH4/NH3 jet diffusion flame utilizing hot O2/CO2 co-flow. The differences in XO2 ranged from 3% to 30%, whereas Tcof ranged from 1300 to 2100 K. The findings revealed that the MILD strategy was attainable when XO2 ≤ 21% for all Tcof levels but required a higher Tcof for higher XO2 values (e.g., Tcof ≥ 1500 K for XO2 = 24%). However, increasing the maximum temperature above the auto-ignition temperature led to a transition of the combustion regime to high-temperature combustion. Increasing XO2 and Tcof changed the radicals (H, O, and OH), reducing the O radicals, which shifted the NO formation route. Increasing the availability of radicals enabled efficient NH3 oxidation and NO reduction through recombustion, thereby decreasing the NO emission index.
Zhao et al. [90] continued their research, focusing on investigating the mechanism of fuel and NO formation during MILD oxy-combustion of CH4/NH3 under different diluent environments: N2, CO2, and water vapor. The analysis revealed that MILD-N2 combustion had the highest temperature and most intense reaction zone that caused NO emission. In contrast, in MILD-CO2 combustion, the flame temperature decreased, and the flame zone expanded because the CO2 absorbed heat. MILD-H2O combustion produced a slightly higher flame temperature than CO2 but yielded a minor reaction zone because H2O enhanced the oxidation process by generating OH radicals.
Blending ammonia oxy-combustion with other fuels is highly desirable in gas combustors to achieve high fuel activity. Cai and Zhao [91] analyzed the micro-power system by investigating the impact of changing the fuel composition ratio (ε) and the hydrogen molar fraction in the ammonia–hydrogen mixture. In addition, they examined the influence of the wall thermal conductivity (WTC) on the thermal efficiency and NOx pollution. The examination revealed that adding hydrogen reduced the NO concentration and changed the flame position. Figure 26 depicts the change in radiation efficiency (RE) with and without H2 inclusion, clarifying that a higher ε ratio led to a reduced RE value. Nevertheless, an increased fuel flow rate resulted in an upward trend in the RE. However, beyond a certain point, any further increase caused a decline in RE. Hydrogen addition also impacted the thermal performance, reducing the OWT, which correlated with decreased NOx emissions. Further analysis showed that a high WTC could improve the OWT at low flow rates. A high WTC also produced a more consistent temperature distribution across the outer wall.
Luo et al. [92] explored ammonia, hydrogen, and oxygen in premixed combustion within a segmented nozzle micro-combustor in a thermophotovoltaic system, as shown in Figure 27. This technique aims to enhance the stability of ammonia combustion and control NOx emissions. The numerical model uses the pressure-based COUPLE algorithm because it accurately represents the interaction between the flame and wall, enabling more efficient convergence. Figure 28 shows that under identical inlet velocity conditions, the average wall temperature of the segmented structure exceeds that of the non-segmented structure. In addition, the segmented channel produces lower NO emissions than the non-segmented channel. As the hydrogen concentration increased from 10% to 30%, the efficiency of the burner radiation improved by 13.5%. However, this increase in hydrogen led to higher NO emissions. This study also revealed specific combustion conditions that minimize NO emissions while maintaining high RE. For instance, under Φ H2 = 1.0 and Φ NH3 = 1.2, the NO emissions could be diminished by up to 13.3% while maintaining RE above 60%.
Sun et al. [93] performed a 3D numerical analysis of a rotating detonation engine fueled by a blend of ammonia/hydrogen and oxygen-enriched air. The detonation wave structure, propagation mode, burner performance, and emission formation were analyzed under different hydrogen concentrations and equivalence ratios. Figure 29 depicts the temperature and pressure fields of the combustion chamber as the hydrogen concentration increased, demonstrating that an increase in the hydrogen concentration increased the temperature, pressure, and velocity. Meanwhile, the elevation of the detonation pulse was expected to decline, but an increase in the detonation velocity would increase the fuel consumption. The cycle efficiency is calculated as follows:
η = L E / Q i n ,
Q i n = m h 2 H h 2 + m n h 3 H n h 3
where mh2 is the hydrogen mass flow rate and mnh3 denotes the ammonia mass flow rate. Hydrogen and ammonia have low calorific values: Hh2 and Hnh3. The addition of hydrogen generates a constant detonation wave, thereby enhancing the cycle efficiency. However, it does not necessarily improve the effectiveness. As observed in Figure 30, the optimal cycle efficiency was attained when the hydrogen level was maintained at 0.3, with the Φ value remaining below 1. Furthermore, the NOx emission was primarily influenced by the equivalency ratio of the fuel mixture instead of the hydrogen concentration, and in the range of Φ = 0.7–1.4, the NO emissions significantly decreased.
Ilbas et al. [94] reviewed the combustion characteristics and NOx emissions of ammonia/kerosene fuel under air and pure oxygen combustion conditions in a swirl gas turbine combustor. The ammonia composition was determined for a blend of 70% kerosene and 30% ammonia in increments of 10% based on the heat fraction. Figure 31 shows no significant change in the maximum temperature of the kerosene fuel when ammonia was added to the combustor. However, when pure oxygen was used instead of air, the temperature of the combustion chamber significantly improved. Owing to the thermal NOx mechanism, the high-temperature zone drove NOx formation during oxygen combustion. However, the predicted NOx emission levels at the combustion chamber’s exit were not significantly elevated.
Table 3. Recent research applications of ammonia oxy-combustion in a gas furnace.
Table 3. Recent research applications of ammonia oxy-combustion in a gas furnace.
AuthorsResearch MethodCombustion ConditionsApproach
Equipment
Key Findings
El-Adawy and Nemitallah [77]Computational fluid dynamics (CFD) simulation
  • Tin = 298–498 K
  • Vin = 5.2–9.0
  • Φ = 0.6–1.2
  • Ammonia oxy-combustion
Swirl gas turbine combustor
  • Flamelet-generated manifold models for complex turbulent combustion.
  • Higher Tin and peak combustion temperatures.
  • Higher Vin resulted in adequate mixing and homogenous temperature profiles.
Cai et al. [78]CFD simulation
  • Initial pressure of 1–2 atm
  • Initial temperature of 300–500 K
  • Φ = 0.8–1.2
  • Ammonia oxy-combustion
Micro-planar combustor
  • Φ = 0.9 yielded the optimal heat transfer performance.
  • Increasing Tin decreased the outer wall mean temperature (OWMT) and NO.
  • Increasing Pin increased the OWMT but reduced NO formation.
Zhang et al. [79]CFD simulation
  • Initial pressure and temperature of 1.0 atm and 300 K, respectively
  • Φ = 0.6–1.1
  • Velocity: 0.1–1.1 m/s
  • Non-premixed combustion
Micro-planar combustor
  • Non-premixed combustion prevented flashback and provided improved heat and emission efficiency.
  • Φ = 0.9 yielded the optimal heat transfer performance.
  • Increasing the velocity increased the temperature.
Cafiero et al. [80]Experiment
  • Inlet velocity: 80–50 m/s
  • Φ = 0.9–1.01
  • Ammonia oxy-combustion
  • Oxygen concentration: 21–50%
20-kW semi-industrial-scale furnace
  • O2 enrichment of 37% is optimal for moderate or intense low-oxygen diffusion (MILD) combustion.
  • A higher inlet oxidizer velocity stabilized the combustion and temperature distribution.
  • At Φ = 0.95 and the maximum temperature, the optimal compromise between nitrogen oxide (NOx) emissions and ammonia slip can be achieved.
Matveevet al. [82]Chemkin simulation
  • Average temperature at the plasma torch outlet: 4950 K
  • RF plasma torch frequency: 3–7 MHz
  • Pressure: 1–10 bar
  • Fuel and oxidizer temperatures: 288 K and 409 K, respectively
Perfectly stirred reactor
  • Plasma assistance significantly stabilized the combustion.
  • Increasing the pressure in combustion decreased the NOx emissions.
Matveevet al. [83]CFD simulation
  • Pressure: 0.3 MPa
  • Initial flow rate of ammonia: 53.76 g/s; temperature: 288 K
  • Plasma oxygen flow rate: 3 g/s; average temperature: 4990 K
  • Oxidizer flowrate: 400–720 g/s; temperature: 409 K
Radial swirl combustion chamber
  • The eddy dissipation concept model was implemented.
  • Plasma assistance trigger factor in ammonia ignition.
  • NO formed in a central zone, whereas NO2 was produced at reduced temperatures.
Cai et al. [84]CFD simulation
  • Initial temperature: 300 K
  • Flow rate: 600–1200 mL/min
  • h = 1/5–4/5
  • l = 1/5–4/5
  • Ammonia oxy-premixed combustion
Micro-planar combustor with a bluff-body
  • A bluff body can increase the outer wall temperature (OWT) and reduce NO generation.
  • The ideal dimensions of the bluff body are h = 4/5 and l = 4/5, which optimize the maximum OWT while minimizing the NOx emissions.
Cai et al. [85]CFD simulation
  • Initial pressure and temperature: 1.0 atm and 300 K, respectively
  • Φ = 0.8–1
  • Ammonia oxy-premixed combustion
Micro-planar combustor with perforated plates
  • Perforated plates reduced the LBV and NO formation.
  • High-thermal-conductivity materials were effective in conjugate heat transfer.
Bektas et al. [86]CFD simulation
  • Φ = 1.1–1.68
  • Ammonia oxy- and air-premixed combustion
  • Methane air-premixed combustion
Two-inlet cyclone combustor
  • The performance in oxy-ammonia combustion was outstanding compared with that in air–methane and air–ammonia combustion.
  • The maximum flame occurred at Φ = 1.1 for oxy-ammonia.
Zhao et al. [88]CFD simulation
  • Fuel inlet: CH4/NH3 at 39.9 m/s with a temperature of 305 K
  • Co-flow inlet: 3.2 m/s with a temperature of 1700 K
  • XCO2 = 0–41%
  • XH2O = 0–50%
  • MILD-oxy condition
Jet flame in hot co-flow (JHC) burner
  • Increasing the CO2 and H2O contents increased the formation of NO.
Zhao et al. [89]CFD simulation
  • Fuel inlet: CH4/NH3 Re 9482 (39.9 m/s) and temperature 305 K
  • Co-flow inlet: 3.2 m/s and temperature 1300–2100 K
  • XO2 = 3–30%
  • XCO2 = 97–70%
  • MILD-oxy condition
JHC burner
  • MILD combustion occurred at lower XO2 levels or higher Tcof levels.
  • The NO emission index decreased as XO2 and Tcof increased.
Zhao et al. [90] CFD simulation
  • Fuel inlet: (CH4/NH3) Re 39.9 m/s and temperature 305 K
  • N2, CO2, and H2O dilution inlet 39.9 m/s and temperature 305 K
  • MILD-oxy condition
JHC burner
  • MILD-H2O was most suitable for achieving lower NOx emissions while maintaining combustion stability.
Cai and Zhao [91]CFD simulation
  • ε = 0.1–0.5
  • The wall thermal conductivity (WTC) was increased from 1 to 100 W/mK
  • Ammonia/hydrogen–oxygen combustion
Micro-planar combustor
  • Adding hydrogen reduced the NO concentration and changed the flame position.
  • High WTC improved the temperature uniformity and reduced the NO emissions.
  • The RE increased with the WTC.
Luo et al. [92] CFD simulation
  • Φ H2 and NH3 = 1–1.2
  • Hydrogen content: 10–30%
  • Ammonia/hydrogen–oxygen combustion
  • Premixed combustion
Segmented nozzle micro-combustor in thermophotovoltaic system
  • The segmented constructs enhanced the thermal efficiency and minimized NO emissions.
  • Increasing the H2 content resulted in enhanced NO emissions.
  • ΦNH3 = 1.2 and ΦH2 = 1.0 were the effective conditions.
Sun et al. [93]CFD simulation
  • Initial pressure and temperature: 1.0 atm and 300 K, respectively
  • Φ = 0.6–1.4
  • Hydrogen concentration: 0.1–0.5
  • Ammonia/hydrogen-oxygen-air combustion
  • Premixed combustion
Rotating detonation engine
  • A sustained detonation wave requires a hydrogen concentration of at least 0.2.
  • Φ must be maintained between 0.9 and 1.3.
  • Φ is an effective feature for managing emissions.
Ilbas et al. [94]CFD simulation
  • Ammonia content: 10–30%
  • Ammonia/kerosene–oxygen combustion
  • Non-premixed combustion
Swirl gas turbine combustor
  • An ammonia addition did not significantly alter the maximum temperature of kerosene combustion.
Ammonia oxy-combustion in gas furnaces can be improved without a drastic redesign but requires system modification in the ignition process. Ammonia in oxygen combustion can improve combustion efficiency, although flame stability and NOx emissions remain significant challenges. NOx formation is significantly influenced by flame temperature and oxygen availability factors that can be overcome by MILD combustion. Most researchers use fuel blending to overcome such ammonia combustion problems, incorporating fuels such as methane, hydrogen, and kerosene, which positively impacts the flame speed and flame stability. Although hydrogen improves combustion efficiency and significantly reduces the amount of unburned ammonia, a high proportion of blended hydrogen increases NOx emissions; we observed NOx emissions under fuel-rich conditions. This will provide an opportunity for a practical strategy, and blending ammonia and kerosene (up to 30% ammonia by heat fraction) does not significantly affect the combustion performance. The main changes occur in the maximum temperature level, which increases slightly, and the combustion distribution. In addition, integrating fuel mixing with advanced burner configurations, such as by performing a redesign with non-premixed combustion, incorporating a bluff body, and conducting micro-segmentation, is also effective in improving the combustion performance and reducing NOx. However, it must be studied thoroughly, especially on an industrial scale.

4.2. Ammonia Oxy-Fuel Combustion in Internal Combustion Engines

The prior studies related to ammonia oxy-fuel combustion in internal combustion engines are further discussed and summarized in Table 4. Park et al. [95] examined the potential of ammonia energy in an ICE and determined the characteristics of ammonia-enriched oxygen combustion. They observed that adding oxygen reduced the ignition delay under low-speed, low-load conditions, as evidenced by a reduction in the coefficient of variation for the indicated mean effective pressure. Figure 32 shows that increasing the oxygen concentration can improve stability and thermal efficiency, which increases significantly with the addition of 25% oxygen. Nonetheless, if the oxygen addition was greater than this value, then the increase in efficiency was no longer significant. Oxygen enrichment significantly decreased the amount of unburned NH3 and N2O emissions but increased NOx emissions.
Nonavinakere et al. [96] explored how diverse oxygen levels influence the ammonia combustion sparked in a constant-volume combustion chamber. Figure 33 shows a series of high-speed photographs capturing the varying oxygen levels observed at various times following ignition. These findings reveal that increasing the oxygen concentration renders the flame velocity sensitive to the air–fuel ratio. According to Figure 34, the highest flame velocity recorded was 112.7 cm/s at 40% oxygen with a Φ value of 1.1 under 10 bar of pressure. Lower oxygen levels also impacted the flame velocity by causing flame instability, which resulted in asymmetric flame development triggered by buoyancy effects within the burner. The study also highlighted other practical effects of increasing the oxygen concentration, such as improved chemical reactivity and combustion completeness. As shown in Figure 35, these variables increased the peak heat release rate by approximately 2.5 times. Furthermore, they decreased the ignition delay and prolonged the flames.
Wang et al. [97] dissected the ignition and combustion enhancement processes of an NH3/water mixture with auxiliary oxygen using a turbulent jet ignition (TJI) system. They analyzed the effects of various parameters, including the pre-chamber equivalence ratio (ΦPC), primary chamber equivalence ratio (ΦMC), and orifice diameter (d), on the combustion performance. The study revealed that auxiliary oxygen injection significantly improved the ignition performance, resulting in better flame propagation and stability. ΦPC had no significant effect on the jet power, but increasing ΦPC could decrease the ignition delay due to the increased availability of O2. Although ΦMC controls the ignition delay and combustion duration, combustion was most efficient at ΦMC values between 1.2 and 1.3, balancing the mixture reactivity and oxygen availability. This study also highlighted that decreasing the orifice diameter increased the jet velocity, as shown in Figure 36, thereby advancing the flow turbulence. However, the turbulent flow caused a lack of ignition performance because NH3/air cannot be ignited under high-turbulence velocity conditions.
Hong et al. [98] examined the effect of oxygen additions on the performance of an ammonia–hydrogen dual-fuel engine by applying a high compression ratio using hydrogen direct injection. They explored the impacts of single- and dual-hydrogen injection modes on the flame effect, combustion stability, and thermal efficiency. In addition, the effects of oxygen enrichment levels on burning and flame propagation were analyzed. Some injection strategies decreased the primary combustion duration, cyclic variation (CoVPmax), BMEP, and brake thermal efficiency (BTE). These improvements were due to the development of a hydrogen jet and mixture distribution for the optimal combustion process. In addition, increasing the oxygen concentration could enhance the heat release, shorten the ignition delay, and improve the combustion stability. However, higher oxygen levels reduced the BTE because of the increased adiabatic flame temperature, which caused more significant heat transfer losses. The dual hydrogen injection mode was more suitable when combined with oxygen addition strategies, resulting in improved engine performance. Figure 37 shows a hybrid power system design utilizing an ammonia–hydrogen dual-fuel engine as a future technology development based on the results of this and previous studies.
Park et al. [99] obtained significant insights through experiments on spark ignition engines using ammonia as preliminary fuel and the role of selective catalytic reduction (SCR) in reducing NOx and unburned ammonia emissions. The effect of increasing the oxygen concentration in the intake air was also evaluated. The authors found that increasing the amount of O2 in the intake air enhanced the combustion rate of ammonia and consequently improved the stability and efficiency of combustion. The increase in combustion rate with the addition of O2 also affected the concentrations of the components in the exhaust gas, and as the O2 concentration increased, NOx increased, whereas the amount of unburned NH3 decreased. The SCR post-treatment system showed that the O2 concentration in the intake air could maximize the conversion efficiency of low NOx and unburned NH3 at 25.5% under high-load conditions. In contrast, under low-load conditions, the ideal O2 concentration was 23% for NOx reduction and 24% for unburned NH3 reduction. Therefore, the operating conditions could be utilized to maximize the conversion efficiency in the SCR post-treatment system by adjusting the ratio of NOx and unburned NH3 via changing the O2 concentration.
Table 4. Recent studies on the application of ammonia oxy-combustion in an internal combustion engine.
Table 4. Recent studies on the application of ammonia oxy-combustion in an internal combustion engine.
AuthorsResearch MethodCombustion ConditionsApproach EquipmentKey Findings
Park et al. [95]Experiment
  • Operating conditions: 1500 rpm low load and 2000 rpm high load
  • Oxygen content: 20–30%
  • Ammonia oxy-combustion
  • SI engine
  • Applying oxygen to the intake air improved the spark timing, burning speed, and thermal efficiency.
Nonavinakere et al. [96]Experiment and Chemkin simulation
  • Initial pressure and temperature: 10 bar and 373 K, respectively
  • Φ = 0.9–1.15
  • Ammonia oxy-combustion (oxygen concentrations of 15–40%)
  • Constant-volume chamber
  • Schlieren technique
  • Increasing the oxygen concentration positively impacted the ammonia combustion, flame velocity, ignition delay time, and HHR.
Wang et al. [97]Experiment
  • ΦPC = 0.6–1.2
  • ΦMC = 1.1–1.4
  • d = 3–5 mm
  • Ammonia oxy-combustion
  • TJI
  • Constant-volume combustion bomb
  • O2 boosted the ignition capability of the unburnt fuel in the primary chamber.
  • The orifice diameter controlled the jet velocity and turbulent intensity.
Hong et al. [98]Experiment
  • Hydrogen direct injection
  • For multiple injections, a secondary injection ratio of 0.5
  • Oxygen content: 21–34.29%
  • Ammonia–hydrogen oxy-combustion
  • Dual-fuel engine
  • The engine performance metrics were enhanced using multiple injection techniques.
  • Oxygen enrichment improved the combustion, but the thermal efficiency was compromised.
Park et al. [99]Experiment
  • Operating conditions: low and high loads of 1000 and 1500 rpm, respectively
  • Oxygen content: 20–30%
  • Ammonia oxy-combustion
  • SI engine
  • Adjusting the O2 concentration in the selective catalytic reduction post-treatment system can enhance the conversion efficiency by optimizing the ratio of NOx to unburned ammonia.
Research on ammonia oxy-combustion in ICEs has shown that NOx is highly sensitive to the combustion phase, ignition timing, and oxygen concentration. An adaptive ignition control strategy is needed to achieve optimal performance without lowering emission standards. In addition, hydrogen injection can stabilize ammonia combustion. However, analyzing other additives such as methanol, ethanol, and propane is worthwhile, as they can offer better performance improvement between the initial ignition and flame speed [100,101]. This research is mainly conducted experimentally at a lab scale. In practical research, studies using a machine-scale approach with real load conditions are needed.

4.3. Ammonia Oxy-Fuel Combustion in Coal-Fired Furnaces

Another application of ammonia oxy-fuel combustion that is in coal-fired furnaces is discussed in this subsection and summarized in Table 5. Ghadi et al. [102] conducted a simulation analysis of a pilot-scale furnace encompassing ammonia–coal oxy-co-firing with variations from pure coal to 50% ammonia in a swirl burner. They investigated the consequences of ammonia oxy-co-firing on the fluid motion and heat transfer, species concentration, volatile combustion char burnout, and flue gas emissions. The authors presented a simulation using Ansys Fluent 22.0 for gas and particle modeling using the Lagrangian method. Additionally, the turbulent flow in combustion was analyzed using the realizable k-ε modeling method, which is particularly effective at predicting swirling flows. In addition, the effect of radiation on the heat transfer process was considered using the weighted sum of gray gases and discrete ordinate models. The authors found that the swirling motion improved the integration of the fuel and oxidizer, thus affecting the combustion efficiency. Increasing the ammonia content intensified the formation of internal recirculation zones. The ammonia ignited sooner and burned faster than the coal, leading to increased oxygen consumption near the burner. CO2 emissions were reduced with increasing additions of ammonia in the fuel mixture compared with pure coal. In contrast, the NO emissions in Figure 38 and Figure 39 show a sharp increase when 10% ammonia is added. However, the NO formation rate stabilizes when the amount of ammonia is increased to 50%, where the NO level for the 50% coal-50% ammonia mixture (shown by the red line) is about 30 ppm higher than the NO levels for 100% coal combustion (indicated by the purple line).
Pan et al. [103] utilized an experimental approach and CFD-DEM numerical modeling to analyze the interaction between ammonia and oxygen bubbles in a fluidized bed combustion process. The mechanics of diffusion combustion between ammonia and oxygen bubbles in fluidized beds have been studied, emphasizing the effects of mass transfer on combustion reactions. Moreover, the combustion process occurs in three stages, as shown in Figure 40: the bubble injection process, bubble approach process, and merging or bursting of bubbles. Mass transfer occurs during the bubble approach process. The oxygen is repositioned from the bottom to the top of the bubble. At the same time, NH3 from the upper bubble is absorbed by the lower bubble through the emulsion, which promotes combustion. The interval affects the bubble coalescence and combustion efficiency, with a shorter interval time increasing the oxygen transfer and combustion.
The role of oxygen and char nitrogen oxidation mechanisms in ammonia and coal combustion on the formation of NOx emissions was investigated by Zheng et al. [104] using density functional theory. This examination revealed three primary reaction pathways involved in ammonia–coal combustion: the formation of N2, NO2, and NO. These findings indicate that oxygen promotes the oxidation of NO to NO2, which effectively reduces the activation energy of this process. As the temperature increased from 1000 K to 2000 K, the NO reduction rate by ammonia consistently exceeded the NO production rate.
Chen et al. [105] studied the dynamics of NO production during the combustion of ammonia and coal using experimental studies and quantum chemistry computations under O2/CO2 atmospheric conditions. The results revealed that elevated temperatures and oxygen concentrations promoted NO formation, which peaked earlier. Meanwhile, CO2 significantly inhibited NO formation when the temperature increased. Using quantum calculations, the authors discovered that O2 reduced the energy barrier for ammonia–nitrogen oxidation and promoted its conversion into nitrogen oxide while increasing the energy barrier for coal–nitrogen oxidation. The combustion of oxygen fuel along with flue gas recirculation is an effective method for suppressing NO emissions in combustion processes involving ammonia and coal. Lei et al. [106] investigated the NO pollution characteristics of coal and ammonia co-firing under O2/N2 and O2/CO2 atmospheres using experiments and reactive force field molecular dynamics simulations (ReaxFF MD). The findings revealed that burning oxygen, especially in an O2/CO2 environment, suppressed NO emission generation compared with combustion with conventional air (O2/N2). Nevertheless, NO formation increased with the temperature and NH3 blending ratio, whereas the CO2 concentration inhibited the transformation of NH3 into NO at low temperatures. In contrast, elevated temperatures accelerated the reaction, increasing the number of OH radicals in the burning process and encouraging the conversion of NH3 into NO.
Xia et al. [107] conducted an experiment in a constant-volume chamber stirred by a fan to examine the flame propagation during the combustion of ammonia and coal under varying NH3/O2/N2 Φ values. They observed that the flame propagation rate during the combustion of ammonia and coal exceeded that during pure ammonia combustion under lean conditions. Regardless, under ammonia-rich conditions, the speed of flame propagation diminished for the coal and ammonia mixture due to heat being absorbed by coal particles and an elevated fuel Φ value, which reduced the LBV under stoichiometric conditions. The flame propagation velocity was determined to be similar for co-combustion and pure ammonia combustion.
Table 5. Recent studies on the application of ammonia oxy-combustion in a boiler.
Table 5. Recent studies on the application of ammonia oxy-combustion in a boiler.
AuthorsResearch MethodCombustion ConditionsApproach EquipmentKey Findings
Zare Ghadi et al. [102]CFD simulation
  • LHV of ammonia: 18.6 MJ/kg; LHV of coal: 20.9 MJ/kg
  • Ammonia content: 0–50%
  • Ammonia/coal oxy-combustion
  • Swirl burner
  • Increasing the ammonia content improved combustion efficiency and reduced CO2 emissions, whereas the NO emissions rose initially but stabilized as the ammonia content increased further.
Pan et al. [103]Experiment and CFD-DEM simulation
  • Initial temperature: 200 °C
  • Pressure: 3 bar
  • Particle diameter: 0.5 mm
  • Ammonia–oxygen
  • Fluidized bed furnace
  • The mass transfer process of bubbles significantly impacted the combustion efficiency.
  • Shorter interval times enhanced oxygen transport and combustion.
Zheng et al. [104]Quantum chemistry calculations
  • NH3 content: 0–100%
  • NH3/coal/O2 combustion
  • A single-layer graphene structure
  • Oxygen promoted the conversion of NO into NO2 during combustion.
Chen et al. [105]Experiment and quantum chemistry calculations
  • Ammonia content: 15%
  • Pulverized coal: 119 mg
  • Oxygen content: 21–40%
  • Ammonia/coal oxy-combustion
  • High-temperature tube furnace
  • Constant-temperature fixed bed
  • Oxy-fuel combustion with flue gas recirculation significantly prevented NO emissions.
Lei et al. [106]Experiment and ReaxFF MD simulation
  • Bituminous coal
  • Blending ratio of NH3: 0–60%
  • Coal/NH3/O2/CO2/N2 combustion
  • Drop tube furnace
  • The NO in an O2/CO2 atmosphere exceeded that in an O2/N2 atmosphere.
Xia et al. [107]Experiment
  • Φ ammonia/oxygen/nitrogen in co-combustion: 0.4–1.4
  • Turbulence intensity: 0–1.29 m/s
  • Coal/ammonia/oxygen/nitrogen combustion
  • A fan-stirred constant volume chamber
  • Schlieren images
  • The flame propagation rate in co-combustion rose under ammonia-lean conditions.
  • Stoichiometry indicated the same rate for both combustion types.
Integrating oxy-fuel ammonia combustion into coal-fired furnaces is an essential and fascinating topic to research in the effort to decarbonize legacy thermal power systems, which are widely used for electricity generation. The combustion strategy uses ammonia with coal in oxygen combustion to decrease CO2 significantly. However, controlling NOx under this combustion condition remains a concern because of the fuel-bound nitrogen in both ammonia and coal.

5. Conclusions

Ammonia oxy-combustion has emerged as a technological development that offers renewable energy and carbon-neutral fuels. It addresses the global challenges in energy supply and climate change issues generated by carbon emissions from traditional fuels. Using ammonia as fuel has several advantages, such as not generating pollution during combustion, not being easily explosive, and being easy to store in a mature supply chain infrastructure. On the production side, ammonia can be produced on a large scale by the HB method. However, ammonia also poses challenges, particularly due to its toxicity, making stringent safety measures to prevent leakage a critical concern. Another challenge with ammonia is the LBV, which makes ignition difficult and reduces its efficiency compared with conventional hydrocarbons. Additionally, its combustion generates fuel-bound NOx and unburned ammonia, but oxygen enrichment in ammonia combustion can promote combustion efficiency.
Research on ammonia oxy-combustion has been conducted comprehensively, particularly focusing on turbulent flame propagation and LBV. The effects of Φ as well as the oxygen content, initial temperature or pressure, and dual and trinary fuel modes have been analyzed. In addition to chemical kinetic analysis, examinations have been conducted to validate experimental discoveries, establish the performance of sensitivity analyses, and identify the emission pathways. Previous research on ammonia–oxygen combustion revealed that the maximum LBV occurs at a Φ value of approximately 1.1 when the combustion is fuel-rich. However, this condition must be considered because it significantly increases the ammonia slip in the exhaust gas. This gain raises concerns about environmental pollution due to hydrocarbon emissions from unburned ammonia, resulting in decreased combustion efficiency. Moreover, other factors also influence the combustion process, including an elevated ammonia concentration and a higher starting pressure, which can decrease the LBV. Meanwhile, regarding NO emission formation, scholars have found that preheating enlargement caused an increase in NO. However, under conditions of higher pressure and fuel-rich combustion, the potential exists for NO emissions to be reduced.
This review examined the advancements in ammonia oxy-combustion for diverse applications in gas furnaces, ICEs, and boilers. Experimental and computational studies, such as CFD investigations, have been used to observe combustion phenomena. Miscellaneous technical innovations have also been employed to enhance combustion stability and decrease NO emissions. In gas furnaces, approaches such as MILD combustion, plasma assistance, non-premixed combustion, bluff body integration, WTC, and segmented nozzle design have been implemented. The dual-fuel modes include ammonia–methane, ammonia–hydrogen, and ammonia–kerosene. For ICE applications, the advancements focus on the multiple-injection method, redesigning the orifice diameter, and utilizing dual-fuel mode ammonia–hydrogen. Similarly, in boiler systems, these strategies use flue gas recirculation and the dual-fuel ammonia–coal mode.
In summary, ammonia oxy-combustion offers a promising route to improving combustion efficiency. Despite the observed improvements, this approach provokes issues primarily because it increases NO emissions. However, this review shows that the development of ammonia oxygen combustion is still focused on the laboratory and experimental scales, with most researchers concentrating on parameters such as the equivalence ratio and oxygen enrichment. Therefore, a comprehensive study on applying various methods in the ammonia oxy-combustion process is necessary to ensure stable combustion and control NOx generation.
The general future research directions that could accelerate the transition to ammonia oxy-combustion systems, such as a staged combustion method and the application of a staged combustion system for ammonia oxy-combustion in gas furnaces, gas turbines, and boilers, are less explored. Future research is needed to analyze the effectiveness of the combustion stages, including the fuel and air stages, on the combustion performance and NOx formation. Furthermore, mixing ammonia oxy-combustion with other fuels, such as hydrogen, methane, alcohol, and DME, requires further analysis under various oxygen combustion conditions in all applications, such as gas furnaces, ICEs, and boilers. Advanced burners with plasma assistance are important to analyze further because they can reduce ignition delay in ammonia combustion. More widely, research should be conducted on different combinations of pressures, temperatures, and burner designs. Ammonia slip must also be mitigated, especially under fuel-rich conditions such as those present in SCR applications. Overall, persistent research, invention, and partnerships are necessary to develop practical applications of this technology and support the global energy transition.

Author Contributions

Data curation, formal analysis, investigation, and writing—original draft, N.D.; software, validation, visualization, and writing—review and editing, K.I.; funding acquisition, methodology, project administration, resources, supervision, and writing—review and editing, Y.S. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by a Korea Evaluation Institute of Industrial Technology (KEIT) grant funded by the Korean government (MOTIE) (No. RS-2022-00155548).

Conflicts of Interest

The authors declare that they have no known competing financial interests or personal relationships that could have appeared to influence the work reported in this paper.

Abbreviations

BTEBrake thermal efficiency
CFDComputational fluid dynamics
DMEDimethyl ether
GHGGreenhouse gas
HBHaber–Bosch
ICEInternal combustion engine
JHCJet-flame in hot co-flow
LBVLaminar burning velocity
MILDModerate or intense low-oxygen dilution
NOxNitrogen oxide
OWMTOuter wall mean temperature
OWTOuter wall temperature
RERadiation efficiency
SCRSelective catalytic reduction
SISpark ignition
WTCWall thermal conductivity

References

  1. National Academies of Sciences, Engineering, and Medicine. Greenhouse Gas Emissions Information for Decision Making: A Framework Going Forward; The National Academies Press: Washington, DC, USA, 2022. [Google Scholar] [CrossRef]
  2. Jang, B.; Oh, C.; Ahn, S.; Kim, Y.; Park, J.; Choi, M.; Sung, Y. Nitric oxide emission reduction and thermal characteristics of fuel-pulsed oscillating combustion in an industrial burner system. Energy 2021, 216, 119263. [Google Scholar] [CrossRef]
  3. United Nations. UN Sustainable Development Goals. 2024. Available online: https://www.un.org/sustainabledevelopment/sustainable-development-goals/ (accessed on 21 October 2024).
  4. Otomo, J.; Koshi, M.; Mitsumori, T.; Iwasaki, H.; Yamada, K. Chemical kinetic modeling of ammonia oxidation with improved reaction mechanism for ammonia/air and ammonia/hydrogen/air combustion. Int. J. Hydrogen Energy 2018, 43, 3004–3014. [Google Scholar] [CrossRef]
  5. United Nations. The Paris Agreement. 2015. Available online: https://unfccc.int/documents/184656 (accessed on 21 October 2024).
  6. World Bank. State and Trends of Carbon Pricing 2023. 2023. Available online: http://hdl.handle.net/10986/39796 (accessed on 21 October 2024).
  7. Metcalf, G.E.; Stock, J.H. The macroeconomic impact of Europe’s carbon taxes. Am. Econ. J. Macroecon. 2023, 15, 265–286. [Google Scholar] [CrossRef]
  8. Karmaker, S.C.; Hosan, S.; Chapman, A.J.; Saha, B.B. The role of environmental taxes on technological innovation. Energy 2021, 232, 121052. [Google Scholar] [CrossRef]
  9. Liu, J.; Huang, Z.; Fan, M.; Yang, J.; Xiao, J.; Wang, Y. Future energy infrastructure, energy platform and energy storage. Nano Energy 2022, 104, 107915. [Google Scholar] [CrossRef]
  10. Yu, Z.; Jianquan, L.; Jie, W.; Wenlong, Z. 9E heavy-duty gas turbine ammonia-hydrogen mixture combustion characteristics and NOx emission characteristics research. Int. J. Hydrogen Energy 2024, 87, 997–1010. [Google Scholar] [CrossRef]
  11. Feng, Z.; Pan, J.; Li, P.; Fan, B.; Nauman, M.; Yang, W. Numeral study of the mixture formation and combustion process of an ammonia/hydrogen rotary engine. Int. J. Hydrogen Energy 2024, 69, 1469–1480. [Google Scholar] [CrossRef]
  12. Jamrozik, A.; Tutak, W. The impact of ammonia and hydrogen additives on the combustion characteristics, performance, stability and emissions of an industrial DF diesel engine. Appl. Therm. Eng. 2024, 257, 124189. [Google Scholar] [CrossRef]
  13. Zhao, H. The role of green ammonia in meeting challenges towards a sustainable development in China. Energy 2024, 310, 133283. [Google Scholar] [CrossRef]
  14. Kobayashi, H.; Hayakawa, A.; Somarathne, K.D.; Okafor, E.C. Science and technology of ammonia combustion. Proc. Combust. Inst. 2019, 37, 109–133. [Google Scholar] [CrossRef]
  15. Ong, C.W.; Chang, N.; Tsai, M.L.; Chen, C.L. Decarbonizing the energy supply chain: Ammonia as an energy carrier for renewable power systems. Fuel 2024, 360, 130627. [Google Scholar] [CrossRef]
  16. Miura, D.; Tezuka, T. A comparative study of ammonia energy systems as a future energy carrier, with particular reference to vehicle use in Japan. Energy 2014, 68, 428–436. [Google Scholar] [CrossRef]
  17. Gu, T.; Araya, S.S.; Yin, C.; Liso, V. Exploring decentralized ammonia synthesis for hydrogen storage and transport: A comprehensive CFD investigation with experimental validation and parametric study. Energy Convers. Manag. 2023, 295, 117604. [Google Scholar] [CrossRef]
  18. IRENA; AEA. Innovation Outlook: Renewable Ammonia; International Renewable Energy Agency: Abu Dhabi, United Arab Emirates; Ammonia Energy Association: Brooklyn, NY, USA, 2022; ISBN 978-92-9260-423-3. Available online: https://www.irena.org/publications/2022/May/Innovation-Outlook-Renewable-Ammonia (accessed on 21 October 2024).
  19. Zhao, Z.; Zhang, M.; Wu, Y.; Song, W.; Yan, J.; Qi, X.; Yang, J.; Zhang, H. Ammonia energy: Synthesis and utilization. Ind. Eng. Chem. Res. 2024, 63, 8003–8024. [Google Scholar] [CrossRef]
  20. IEA. Net Zero Roadmap: A Global Pathway to Keep the 1.5 °C Goal in Reach. 2023. Available online: http://www.iea.org (accessed on 25 October 2024).
  21. Subramani, A.K.; Duraisamy, G.; Govindan, N.; Hossain, A.K. An innovative method of ammonia use in a light-duty automotive diesel engine to enhance diesel combustion, performance, and emissions. Int. J. Hydrogen Energy 2024, 49, 38–58. [Google Scholar] [CrossRef]
  22. Xu, J.; Zhao, R.; Sun, S.; Wu, Z.; Lu, Z.; Xiao, L.; Hou, L. Enhancement of ammonia corrosion resistance of coatings through improving dispersion and binding site of graphene oxide in silane coatings. J. Alloys Compd. 2024, 989, 174174. [Google Scholar] [CrossRef]
  23. An, T.; Wei, B.; Ma, R.; Chen, L.; Wang, S.; Xu, M.; Liu, K. Study on the high-temperature corrosion mechanism of boiler steel 15CrMoG in ammonia-coal co-firing environment. Fuel 2024, 378, 132892. [Google Scholar] [CrossRef]
  24. Tian, J.; Wang, L.; Xiong, Y.; Wang, Y.; Yin, W.; Tian, G.; Wang, Z.; Cheng, Y.; Ji, S. Enhancing combustion efficiency and reducing nitrogen oxide emissions from ammonia combustion: A comprehensive review. Process Saf. Environ. Prot. 2024, 183, 514–543. [Google Scholar] [CrossRef]
  25. Liang, W.; Law, C.K. Enhancing ammonia combustion using reactivity stratification with hydrogen addition. Proc. Combust. Inst. 2023, 39, 4419–4426. [Google Scholar] [CrossRef]
  26. Li, J.; Huang, H.; Kobayashi, N.; He, Z.; Osaka, Y.; Zeng, T. Numerical study on effect of oxygen content in combustion air on ammonia combustion. Energy 2015, 93, 2053–2068. [Google Scholar] [CrossRef]
  27. Fan, G.; Chen, M.; Wang, C.; Feng, Q.; Sun, Y.; Xu, J.; Du, Y.; Che, D. Numerical study on oxy-fuel combustion characteristics of industrial furnace firing coking dry gas. Energy 2024, 286, 129643. [Google Scholar] [CrossRef]
  28. Zhang, Y.; Xin, Y.; Si, M.; Liu, F.; Lou, C. Experimental study on soot formation and primary particle size in oxy-combustion ethylene diffusion flames under CO2 substitution for N2. Case Stud. Therm. Eng. 2023, 48, 103060. [Google Scholar] [CrossRef]
  29. Wachter, P.; Gaber, C.; Demuth, M.; Hochenauer, C. Towards a recuperative, stationary operated thermochemical reformer: Experimental investigations on the methane conversion and waste heat recovery. Appl. Therm. Eng. 2021, 183, 116121. [Google Scholar] [CrossRef]
  30. Gan, Z.; Yang, S.; Wang, H. Investigation of the heating characteristics of turbulent non-premixed gas combustion in the industrial-scale walking beam type reheating furnace. Appl. Therm. Eng. 2024, 257, 124212. [Google Scholar] [CrossRef]
  31. Li, Z.; Zhang, Y.; Zhang, H. Kinetics modeling of NO emission of oxygen-enriched and rich-lean-staged ammonia combustion under gas turbine conditions. Fuel 2023, 355, 129509. [Google Scholar] [CrossRef]
  32. Sun, J.; Zhao, N.; Zheng, H. A comprehensive review of ammonia combustion: Fundamental characteristics, chemical kinetics, and applications in energy systems. Fuel 2025, 394, 135135. [Google Scholar] [CrossRef]
  33. Valera-Medina, A.; Vigueras-Zuniga, M.; Shi, H.; Mashruk, S.; Alnajideen, M.; Alnasif, A.; Davies, J.; Wang, Y.; Zhu, X.; Yang, W.; et al. Ammonia combustion in furnaces: A review. Int. J. Hydrogen Energy 2023, 49, 1597–1618. [Google Scholar] [CrossRef]
  34. Yao, N.; Chen, Y.; Wei, L.; Xu, Q.; Pan, W. Progress in CFD simulation for ammonia-fueled internal combustion engines and gas turbines. J. Energy Inst. 2025, 119, 101951. [Google Scholar] [CrossRef]
  35. Elbaz, A.M.; Wang, S.; Guiberti, T.F.; Roberts, W.L. Review on the recent advances on ammonia combustion from the fundamentals to the applications. Fuel Commun. 2022, 10, 100053. [Google Scholar] [CrossRef]
  36. Wang, Y.; Wang, X.; Zeng, W.; Wang, W.; Song, Z. Advancements in turbulent combustion of ammonia-based fuels: A review. Int. J. Hydrogen Energy 2024, 88, 1332–1355. [Google Scholar] [CrossRef]
  37. Kang, L.; Pan, W.; Zhang, J.; Wang, W.; Tang, C. A review on ammonia blends combustion for industrial applications. Fuel 2022, 332, 126150. [Google Scholar] [CrossRef]
  38. Hua, Y.; Zhang, Y.; Gao, D. Ammonia composite combustion with alcohol/ether: A systematic review from engine applications to combustion enhancement and emission mechanisms. Energy Convers. Manag. 2024, 322, 119160. [Google Scholar] [CrossRef]
  39. Moses, J.K.; Tivfa, T.A.; Qasem, N.A.; Alquaity, A.B. Combustion characteristics of hydrogen, ammonia, and their blends: A review. Fuel 2025, 388, 134404. [Google Scholar] [CrossRef]
  40. Biebl, M.; Leicher, J.; Giese, A.; Wieland, C. A comprehensive study of non-premixed combustion of ammonia and its blends: Flame stability and emission reduction. Fuel 2025, 386, 134501. [Google Scholar] [CrossRef]
  41. Zhang, X.; Zhao, S.; Zhang, Q.; Wang, Y.; Zhang, J. A Review of Ammonia Combustion Reaction Mechanism and Emission Reduction Strategies. Energies 2025, 18, 1707. [Google Scholar] [CrossRef]
  42. Zhou, L.; Zhong, L.; Liu, Z.; Wei, H. Toward highly-efficient combustion of ammonia–hydrogen engine: Prechamber turbulent jet ignition. Fuel 2023, 352, 129009. [Google Scholar] [CrossRef]
  43. Kojima, Y.; Yamaguchi, M. Ammonia as a hydrogen energy carrier. Int. J. Hydrogen Energy 2022, 47, 22832–22839. [Google Scholar] [CrossRef]
  44. Tawalbeh, M.; Murtaza, S.Z.M.; Al-Othman, A.; Alami, A.H.; Singh, K.; Olabi, A.G. Ammonia: A versatile candidate for the use in energy storage systems. Renew. Energy 2022, 194, 955–977. [Google Scholar] [CrossRef]
  45. Tsiklios, C.; Schneider, S.; Hermesmann, M.; Müller, T.E. Efficiency and optimal load capacity of E-fuel-based energy storage systems. Adv. Appl. Energy 2023, 10, 100140. [Google Scholar] [CrossRef]
  46. Li, L.; Wang, B.; Jiao, K.; Ni, M.; Du, Q.; Liu, Y.; Li, B.; Ling, G.; Wang, C. Comparative techno-economic analysis of large-scale renewable energy storage technologies. Energy AI 2023, 14, 100282. [Google Scholar] [CrossRef]
  47. Bicer, Y.; Dincer, I. Environmental impact categories of hydrogen and ammonia driven transoceanic maritime vehicles: A comparative evaluation. Int. J. Hydrogen Energy 2018, 43, 4583–4596. [Google Scholar] [CrossRef]
  48. Guo, Q.; Liu, J.; Zhang, L.; Feng, X.; Wang, H. Explosion limits of binary mixtures of ammonia and additives (hydrogen, natural gas, and diethyl ether). Fuel 2025, 381, 133519. [Google Scholar] [CrossRef]
  49. Reiter, A.J.; Kong, S.-C. Demonstration of compression-ignition engine combustion using ammonia in reducing greenhouse gas emissions. Energy Fuels 2008, 22, 2963–2971. [Google Scholar] [CrossRef]
  50. Liu, J.; Liu, J. Experimental investigation of the effect of ammonia substitution ratio on an ammonia-diesel dual-fuel engine performance. J. Clean. Prod. 2024, 434, 140274. [Google Scholar] [CrossRef]
  51. Chatterjee, S.; Parsapur, R.K.; Huang, K.-W. Limitations of ammonia as a hydrogen energy carrier for the transportation sector. ACS Energy Lett. 2021, 6, 4390–4394. [Google Scholar] [CrossRef]
  52. Wakasugi, T.; Tsuru, D.; Hatta, N.; Hattori, N.; Haruna, S.; Tashima, H.; Watanabe, H. Characteristics of high-pressure spray combustion of liquid ammonia blended with methanol/propane in a marine engine. Fuel 2025, 391, 134808. [Google Scholar] [CrossRef]
  53. National Institute of Standards and Technology, NIST. Chemistry WebBook, Thermophysical Properties of Fluid Systems, SRD 69. Available online: https://webbook.nist.gov/chemistry/fluid/ (accessed on 11 November 2024).
  54. Valera-Medina, A.; Xiao, H.; Owen-Jones, M.; David, W.I.F.; Bowen, P.J. Ammonia for power. Prog. Energy Combust. Sci. 2018, 69, 63–102. [Google Scholar] [CrossRef]
  55. Zhang, R.; Liu, W.; Zhang, Q.; Qi, Y.; Wang, Z. Auto-ignition and knocking combustion characteristics of iso-octane-ammonia fuel blends in a rapid compression machine. Fuel 2023, 352, 129088. [Google Scholar] [CrossRef]
  56. Trotta, L.; Bernardini, V.; Ermini, M.V.; Campana, L.G.; Verna, M.; Raimondi, G.; Spazzafumo, G. Climate finance-driven feasibility study of a green ammonia and urea production plant in Italy. Int. J. Hydrogen Energy, 2024; in press. [Google Scholar] [CrossRef]
  57. Hatzell, M.C. The colors of ammonia. ACS Energy Lett. 2024, 9, 2920–2921. [Google Scholar] [CrossRef]
  58. Mayer, P.; Ramirez, A.; Pezzella, G.; Winter, B.; Sarathy, S.M.; Gascon, J.; Bardow, A. Blue and green ammonia production: A techno-economic and life cycle assessment perspective. iScience 2023, 26, 107389. [Google Scholar] [CrossRef]
  59. Zincir, B. Environmental and economic evaluation of ammonia as a fuel for short-sea shipping: A case study. Int. J. Hydrogen Energy 2022, 47, 18148–18168. [Google Scholar] [CrossRef]
  60. Campion, N.; Gutiérrez-Alvarez, R.; Bruce, J.T.F.; Münster, M. The potential role of concentrated solar power for off-grid green hydrogen and ammonia production. Renew. Energy 2024, 236, 121410. [Google Scholar] [CrossRef]
  61. Hu, L.; Niu, J.; Wu, W.; Zhao, Z.; Du, M.; Chen, L.; Zheng, Q.; Cao, H. Quantitative analysis of toxicity risks in the operation of ammonia-fueled tugboats. Ocean. Eng. 2024, 310, 118759. [Google Scholar] [CrossRef]
  62. Mashruk, S.; Shi, H.; Mazzotta, L.; Ustun, C.E.; Aravind, B.; Meloni, R.; Alnasif, A.; Boulet, E.; Jankowski, R.; Yu, C.; et al. Perspectives on NOX emissions and impacts from ammonia combustion processes. Energy Fuels 2024, 38, 19253–19292. [Google Scholar] [CrossRef]
  63. Hill, S.C.; Douglas Smoot, L. Modeling of nitrogen oxides formation and destruction in combustion systems. Prog. Energy Combust. Sci. 2000, 26, 417–458. [Google Scholar] [CrossRef]
  64. Nozari, H.; Karabeyoğlu, A. Numerical study of combustion characteristics of ammonia as a renewable fuel and establishment of reduced reaction mechanisms. Fuel 2015, 159, 223–233. [Google Scholar] [CrossRef]
  65. Wang, B.; Wang, H.; Yang, C.; Hu, D.; Duan, B.; Wang, Y. Effect of different ammonia/methanol ratios on engine combustion and emission performance. Appl. Therm. Eng. 2024, 236, 121519. [Google Scholar] [CrossRef]
  66. Ichimura, R.; Hadi, K.; Hashimoto, N.; Hayakawa, A.; Kobayashi, H.; Fujita, O. Extinction limits of an ammonia/air flame propagating in a turbulent field. Fuel 2019, 246, 178–186. [Google Scholar] [CrossRef]
  67. Karan, A.; Dayma, G.; Chauveau, C.; Halter, F. Experimental study and numerical validation of oxy-ammonia combustion at elevated temperatures and pressures. Combust. Flame 2022, 236, 111819. [Google Scholar] [CrossRef]
  68. Mei, B.; Zhang, X.; Ma, S.; Cui, M.; Guo, H.; Cao, Z.; Li, Y. Experimental and kinetic modeling investigation on the laminar flame propagation of ammonia under oxygen enrichment and elevated pressure conditions. Combust. Flame 2019, 210, 236–246. [Google Scholar] [CrossRef]
  69. Zhang, Y.; Zhang, W.; Yu, B.; Li, X.; Zhang, L.; Zhao, Y.; Sun, S. Experimental and kinetic modeling study on laminar flame speeds and emission characteristics of oxy-ammonia premixed flames. Int. J. Hydrogen Energy 2024, 63, 857–870. [Google Scholar] [CrossRef]
  70. Okafor, E.C.; Naito, Y.; Colson, S.; Ichikawa, A.; Kudo, T.; Hayakawa, A.; Kobayashi, H. Experimental and numerical study of the laminar burning velocity of CH4−NH3–air premixed flames. Combust. Flame 2018, 187, 185–198. [Google Scholar] [CrossRef]
  71. Liu, S.; Zou, C.; Song, Y.; Cheng, S.; Lin, Q. Experimental and numerical study of laminar flame speeds of CH4/NH3 mixtures under oxy-fuel combustion. Energy 2019, 175, 250–258. [Google Scholar] [CrossRef]
  72. Li, R.; Zhang, Y.; Chen, X.; Wang, D.; Liu, Q.; Liu, X. Experimental study on the combustion characteristics of hydrogen/ammonia blends in oxygen. Case Stud. Therm. Eng. 2024, 60, 104732. [Google Scholar] [CrossRef]
  73. Li, R.; Liu, X.; Chen, X.; Zhang, Y.; Zhang, Q. Flame propagation characteristics and inherent instability in premixed carbon-free hydrogen/ammonia/oxygen combustibles. J. Chem. Thermodyn. 2024, 197, 107326. [Google Scholar] [CrossRef]
  74. Shi, X.; Li, W.; Zhang, J.; Fang, Q.; Zhang, Y.; Xi, Z.; Li, Y. Exploration of NH3 and NH3/DME laminar flame propagation in O2/CO2 atmosphere: Insights into NH3/CO2 interactions. Combust. Flame 2024, 260, 113245. [Google Scholar] [CrossRef]
  75. Barbas, M.; Costa, M.; Vranckx, S.; Fernandes, R.X. Experimental and chemical kinetic study of CO and NO formation in oxy-methane premixed laminar flames doped with NH3. Combust. Flame 2015, 162, 1294–1303. [Google Scholar] [CrossRef]
  76. Wu, Q.; Sun, Z.; Liu, Y.; Li, H.; Zhang, A. The characteristics and mechanism NO formation during hydroxylated fuels/NH3 oxidation in oxy-fuel atmospheres. Energy. 2024, 307, 132558. [Google Scholar] [CrossRef]
  77. El-Adawy, M.; Nemitallah, M. Unveiling Numerically the Impact of Inlet Mixture Composition, Temperature, and Velocity on Ammonia-Oxy Combustion for Gas Turbines. Available online: https://ssrn.com/abstract=4974910 (accessed on 20 November 2024).
  78. Cai, T.; Zhao, D.; Wang, B.; Li, J.; Guan, Y. NO emission and thermal performances studies on premixed ammonia-oxygen combustion in a CO2-free micro-planar combustor. Fuel 2020, 280, 118554. [Google Scholar] [CrossRef]
  79. Zhang, Y.; Lu, Q.; Fan, B.; Quaye, E.K.; Weng, J.; Xiao, S.; Pan, J. Effect of intake method on ammonia/oxygen non-premixed combustion in the micro combustor with dual-inlet. Fuel 2022, 317, 123504. [Google Scholar] [CrossRef]
  80. Cafiero, M.; Sharma, S.; Kamal, M.M.; Lubrano Lavadera, M.; Iavarone, S.; Coussement, A.; Parente, A. Enhancing pure NH3 combustion: Impacts of O2 enrichment under MILD conditions in a 20-kW semi-industrial scale furnace. Proc. Combust. Inst. 2024, 40, 105336. [Google Scholar] [CrossRef]
  81. Sato, D.; Davies, J.; Mazzotta, L.; Mashruk, S.; Valera-Medina, A.; Kurose, R. Effects of Reynolds number and ammonia fraction on combustion characteristics of premixed ammonia-hydrogen-air swirling flames. Proc. Combust. Inst. 2024, 40, 105283. [Google Scholar] [CrossRef]
  82. Matveev, I.B.; Serbin, S.I.; Wolf, K. Plasma-assisted ammonia combustion—Part 1: Possibilities of plasma combustion of ammonia in air and oxygen. IEEE Trans. Plasma Sci. 2023, 51, 1446–1450. [Google Scholar] [CrossRef]
  83. Matveev, I.B.; Serbin, S.I. Plasma-assisted ammonia combustion—Part 2: Combustion of ammonia in oxygen. IEEE Trans. Plasma Sci. 2023, 51, 2290–2294. [Google Scholar] [CrossRef]
  84. Cai, T.; Zhao, D.; Jiaqiang, E. Bluff-body effect on thermal and NOx emission characteristics in a micro-planar combustor fueled with premixed ammonia-oxygen. Chem. Eng. Process. Process Intensif. 2020, 153, 107979. [Google Scholar] [CrossRef]
  85. Cai, T.; Becker, S.M.; Cao, F.; Wang, B.; Tang, A.; Fu, J.; Han, L.; Sun, Y.; Zhao, D. NO emission performance assessment on a perforated plate-implemented premixed ammonia-oxygen micro-combustion system. Chem. Eng. J. 2021, 417, 128033. [Google Scholar] [CrossRef]
  86. Bektaş, A. Oxy- and air-ammonia premixed combustion in a two-inlet cyclone combustor. Int. J. Energy Res. 2022, 46, 13874–13888. [Google Scholar] [CrossRef]
  87. Davies, J.; Mashruk, S.; Sato, D.; Mazzotta, L.; Pugh, D.; Valera-Medina, A. Emissions analyses of humidified cracked ammonia swirling flames. Combust. Flame 2025, 274, 113984. [Google Scholar] [CrossRef]
  88. Zhao, Z.; Zhang, Z.; Zha, X.; Gao, G.; Mao, W.; Wu, F.; Luo, C.; Zhang, L. Fuel-NO formation mechanism in MILD-oxy combustion of CH4/NH3 fuel blend. Fuel 2023, 331, 125817. [Google Scholar] [CrossRef]
  89. Zhao, Z.; Li, X.; Zhang, Z.; Zha, X.; Chen, Y.; Gao, G.; Wu, F.; Luo, C.; Zhang, L. Combustion regimes and fuel-NO mechanism of CH4/NH3 jet diffusion flames in hot O2/CO2 co-flow. Fuel Process Technol. 2022, 229, 107173. [Google Scholar] [CrossRef]
  90. Zhao, Z.; Zhang, T.; Li, X.; Zhang, L.; Zhang, Z.; Chen, Y.; Wu, F.; Luo, C.; Zheng, C. NO formation mechanism of CH4/NH3 jet flames in hot co-flow under MILD-oxy condition: Effects of co-flow CO2 and H2O. Fuel 2022, 313, 123030. [Google Scholar] [CrossRef]
  91. Cai, T.; Zhao, D. Effects of fuel composition and wall thermal conductivity on thermal and NOx emission performances of an ammonia/hydrogen-oxygen micro-power system. Fuel Process Technol. 2020, 209, 106527. [Google Scholar] [CrossRef]
  92. Luo, B.; Jiaqiang, E.; Chen, J.; Zhang, F.; Ding, J. Effect of NH3/H2/O2 premixed combustion on energy conversion enhancement and NOx emission reduction of the segmented nozzle micro-combustor in thermophotovoltaic system. Renew. Energy 2024, 228, 120682. [Google Scholar] [CrossRef]
  93. Sun, Z.; Huang, Y.; Luan, Z.; Gao, S.; You, Y. Three-dimensional simulation of a rotating detonation engine in ammonia/hydrogen mixtures and oxygen-enriched air. Int. J. Hydrogen Energy 2023, 48, 4891–4905. [Google Scholar] [CrossRef]
  94. Ilbas, M.; Kumuk, O.; Karyeyen, S. Numerical study of a swirl gas turbine combustor for turbulent air and oxy-combustion of ammonia/kerosene fuels. Fuel 2021, 304, 121359. [Google Scholar] [CrossRef]
  95. Park, C.; Jang, I.; Park, G.; Min, C.; Kim, M.; Kim, Y.; Choi, Y. Effect of oxygen concentrations in intake air on combustion characteristics of ammonia direct injection SI engine. Fuel 2024, 376, 132643. [Google Scholar] [CrossRef]
  96. Nonavinakere Vinod, K.; Fang, T. Experimental characterization of spark ignited ammonia combustion under elevated oxygen concentrations. Proc. Combust. Inst. 2023, 39, 4319–4326. [Google Scholar] [CrossRef]
  97. Wang, Z.; Ji, C.; Wang, D.; Zhang, T.; Wang, S.; Yang, H.; Zhai, Y.; Wang, H. Experimental investigation on combustion characteristics of ammonia/air using turbulent jet ignition with auxiliary oxygen in pre-chamber. Appl. Therm. Eng. 2024, 243, 122622. [Google Scholar] [CrossRef]
  98. Hong, C.; Ji, C.; Wang, S.; Xin, G.; Qiang, Y.; Yang, J. Evaluation of hydrogen injection and oxygen enrichment strategies in an ammonia-hydrogen dual-fuel engine under high compression ratio. Fuel 2023, 354, 129244. [Google Scholar] [CrossRef]
  99. Park, C.; Choi, Y.; Park, G.; Jang, I.; Kim, M.; Kim, Y.; Choi, Y. Investigation on the reduction in unburned ammonia and nitrogen oxide emissions from ammonia direct injection SI engine by using SCR after-treatment system. Heliyon 2024, 10, e37684. [Google Scholar] [CrossRef]
  100. Uddeen, K.; Tang, Q.; Shi, H.; Turner, J.W. Ammonia-methanol and ammonia-ethanol dual-fuel combustion in an optical spark-ignition engine: A multiple flame generation approach. Appl. Therm. Eng. 2025, 265, 125544. [Google Scholar] [CrossRef]
  101. Zhang, Q.; Zhang, R.; Qi, Y.; Wang, Z. Ignition characteristics of ammonia–methanol blended fuel in a rapid compression machine. Fuel 2024, 368, 131636. [Google Scholar] [CrossRef]
  102. Zare Ghadi, A.; Gu, B.; Lim, H. CFD analysis on the effect of ammonia addition on oxyfuel coal combustion in a swirl burner. Energy Fuels 2024, 38, 19798–19813. [Google Scholar] [CrossRef]
  103. Pan, S.; Ma, J.; Chen, X.; Liang, C. Combustion of an ammonia bubble with an oxygen bubble in a fluidized bed. Appl. Therm. Eng. 2024, 244, 122591. [Google Scholar] [CrossRef]
  104. Zheng, Z.; Shen, J.; Han, Z.; Chen, Y.; Bai, Y. Unveiling the role of oxygen in ammonia coal combustion: A DFT study on NOx emission mechanism. Comput. Theor. Chem. 2024, 1241, 114893. [Google Scholar] [CrossRef]
  105. Chen, P.; Wang, H.; Qiao, L.; Gu, M.; Luo, K.; Fan, J. Study of the NO formation characteristics and ammonia-N/coal-N transformation mechanism of ammonia-coal co-combustion in O2/CO2 atmosphere. Combust. Flame 2024, 270, 113756. [Google Scholar] [CrossRef]
  106. Lei, M.; Zhao, Z.; Hong, D.; Zeng, Y.; Tian, X.; Zhang, L.; Zhang, Q. Experimental study and molecular dynamics simulation on co-combustion of pulverized coal and NH3 at O2/N2 and O2/CO2 atmospheres. Fuel 2024, 364, 131043. [Google Scholar] [CrossRef]
  107. Xia, Y.; Hadi, K.; Hashimoto, G.; Hashimoto, N.; Fujita, O. Effect of ammonia/oxygen/nitrogen equivalence ratio on spherical turbulent flame propagation of pulverized coal/ammonia co-combustion. Proc. Combust. Inst. 2021, 38, 4043–4052. [Google Scholar] [CrossRef]
Figure 1. Potential innovation “leapfrog” benefits for developing nations [8].
Figure 1. Potential innovation “leapfrog” benefits for developing nations [8].
Energies 18 02252 g001
Figure 2. Knock intensity versus initial pressure [55].
Figure 2. Knock intensity versus initial pressure [55].
Energies 18 02252 g002
Figure 3. Mechanism of NH3–air laminar burning velocity (LBV) with equivalence ratios at different O2 levels in combustion air [26].
Figure 3. Mechanism of NH3–air laminar burning velocity (LBV) with equivalence ratios at different O2 levels in combustion air [26].
Energies 18 02252 g003
Figure 4. Schlieren imaging of LBV generated from an ammonia–air mixture at various Φ values [66].
Figure 4. Schlieren imaging of LBV generated from an ammonia–air mixture at various Φ values [66].
Energies 18 02252 g004
Figure 5. Flame propagation chances of ammonia–air flames at various turbulence intensities (u′). The dashed line represents the separation between flame propagation and extinction regions [66].
Figure 5. Flame propagation chances of ammonia–air flames at various turbulence intensities (u′). The dashed line represents the separation between flame propagation and extinction regions [66].
Energies 18 02252 g005
Figure 6. Experimental results for the changes in the unstretched LBV (Sl) as a function of Φ [66].
Figure 6. Experimental results for the changes in the unstretched LBV (Sl) as a function of Φ [66].
Energies 18 02252 g006
Figure 7. Schematic of the Bunsen burner system testing configuration [69].
Figure 7. Schematic of the Bunsen burner system testing configuration [69].
Energies 18 02252 g007
Figure 8. Experimental and modeled values of SL at various (a) Φ and (b) Tu as well as fitted values (c) [69].
Figure 8. Experimental and modeled values of SL at various (a) Φ and (b) Tu as well as fitted values (c) [69].
Energies 18 02252 g008
Figure 9. LBV of NH3/(O2/CO2) mixtures with γ = 50% and (NH3/dimethyl ether (DME))/(O2/CO2) mixtures with α = 50% and γ = 32.5% at Φ = 1.0 for different γ values [74].
Figure 9. LBV of NH3/(O2/CO2) mixtures with γ = 50% and (NH3/dimethyl ether (DME))/(O2/CO2) mixtures with α = 50% and γ = 32.5% at Φ = 1.0 for different γ values [74].
Energies 18 02252 g009
Figure 10. (a) LBV for (NH3/DME)/(O2/CO2) for multiple DME concentrations and γ = 32.5%, Pu = 1 atm, Tu = 373 K, and Φ = 1.0. The square symbols represent the LBVs observed in this study, whereas the curves show the simulation outcomes obtained using the current and prior models. (b) Calculated LBV given by the current model as an indicator of α using the revised fictional diluent gas approach [74].
Figure 10. (a) LBV for (NH3/DME)/(O2/CO2) for multiple DME concentrations and γ = 32.5%, Pu = 1 atm, Tu = 373 K, and Φ = 1.0. The square symbols represent the LBVs observed in this study, whereas the curves show the simulation outcomes obtained using the current and prior models. (b) Calculated LBV given by the current model as an indicator of α using the revised fictional diluent gas approach [74].
Energies 18 02252 g010
Figure 11. Measured and predicted CO emissions based on the excess oxygen coefficient for three distinct oxidizer set-ups [75].
Figure 11. Measured and predicted CO emissions based on the excess oxygen coefficient for three distinct oxidizer set-ups [75].
Energies 18 02252 g011
Figure 12. NO emissions measured and predicted based on the excess oxygen coefficient of the fuel for air with and without ammonia [75].
Figure 12. NO emissions measured and predicted based on the excess oxygen coefficient of the fuel for air with and without ammonia [75].
Energies 18 02252 g012
Figure 13. Impact of CO2 mole fraction of the oxidizer on the laminar flame speed of CH4/NH3/O2/CO2/N2 [71].
Figure 13. Impact of CO2 mole fraction of the oxidizer on the laminar flame speed of CH4/NH3/O2/CO2/N2 [71].
Energies 18 02252 g013
Figure 14. Laminar flame speeds of CH4/NH3/O2/CO2 mixtures analyzed based on the O2 mole fraction in the oxidizer [71].
Figure 14. Laminar flame speeds of CH4/NH3/O2/CO2 mixtures analyzed based on the O2 mole fraction in the oxidizer [71].
Energies 18 02252 g014
Figure 15. Mole fraction range (a,b) and the rate of generation evaluation (c,d) of NO [69].
Figure 15. Mole fraction range (a,b) and the rate of generation evaluation (c,d) of NO [69].
Energies 18 02252 g015
Figure 16. Mole fractions of principal compounds in an ammonia/hydrogen/oxygen flame across different Φ values [72].
Figure 16. Mole fractions of principal compounds in an ammonia/hydrogen/oxygen flame across different Φ values [72].
Energies 18 02252 g016
Figure 17. Combustor design overview. (a) Premixed combustor and (b) non-premixed combustion chamber featuring a singular entry point for the oxidizer [79].
Figure 17. Combustor design overview. (a) Premixed combustor and (b) non-premixed combustion chamber featuring a singular entry point for the oxidizer [79].
Energies 18 02252 g017
Figure 18. Temperature profiles for A9 and A13–A17 at Φ = 0.9 and Lin = 1 mm [79].
Figure 18. Temperature profiles for A9 and A13–A17 at Φ = 0.9 and Lin = 1 mm [79].
Energies 18 02252 g018
Figure 19. Temperature distributions along the X of outer wall 3 with different velocities (u) in meters per second [79].
Figure 19. Temperature distributions along the X of outer wall 3 with different velocities (u) in meters per second [79].
Energies 18 02252 g019
Figure 20. Geometric specifications for the basic micro-planar combustor incorporating a bluff body design [84].
Figure 20. Geometric specifications for the basic micro-planar combustor incorporating a bluff body design [84].
Energies 18 02252 g020
Figure 21. Comparison of the effects of the presence and absence of a bluff body on the mole fraction of NO at varying NH3 flow rates [84].
Figure 21. Comparison of the effects of the presence and absence of a bluff body on the mole fraction of NO at varying NH3 flow rates [84].
Energies 18 02252 g021
Figure 22. Temperature profiles and streamlines in combustors with and without a bluff body [84].
Figure 22. Temperature profiles and streamlines in combustors with and without a bluff body [84].
Energies 18 02252 g022
Figure 23. Schematic of a micro-planar burner with an additional perforation plate [85].
Figure 23. Schematic of a micro-planar burner with an additional perforation plate [85].
Energies 18 02252 g023
Figure 24. Velocity streamlines associated with methane and air/oxy-ammonia [87].
Figure 24. Velocity streamlines associated with methane and air/oxy-ammonia [87].
Energies 18 02252 g024
Figure 25. Temperature distribution and reaction zone contours for N2, CO2, and H2O at O2 levels of 9%, 21%, and 40%, respectively [88].
Figure 25. Temperature distribution and reaction zone contours for N2, CO2, and H2O at O2 levels of 9%, 21%, and 40%, respectively [88].
Energies 18 02252 g025
Figure 26. Radiation efficiencies (REs) at diverse NH3 flow rates (mL/min) across different ε values [91].
Figure 26. Radiation efficiencies (REs) at diverse NH3 flow rates (mL/min) across different ε values [91].
Energies 18 02252 g026
Figure 27. Design of a micro-combustor incorporating a segmented channel [92].
Figure 27. Design of a micro-combustor incorporating a segmented channel [92].
Energies 18 02252 g027
Figure 28. (a) Average temperature of the wall. (b) NO mass flow rate at an outlet for a segmented micro-combustor, as opposed to one without, under different vin conditions [92].
Figure 28. (a) Average temperature of the wall. (b) NO mass flow rate at an outlet for a segmented micro-combustor, as opposed to one without, under different vin conditions [92].
Energies 18 02252 g028
Figure 29. Simulation outputs for cases 0.3–1.1, 0.4–1.1, and 0.5–1.1 (0.3, 0.4, and 0.5 amounts of hydrogen at Φ = 1.1), displayed from left to right [93].
Figure 29. Simulation outputs for cases 0.3–1.1, 0.4–1.1, and 0.5–1.1 (0.3, 0.4, and 0.5 amounts of hydrogen at Φ = 1.1), displayed from left to right [93].
Energies 18 02252 g029
Figure 30. Cycle efficiency curves for various hydrogen concentrations of (red) 0.0, (orange) 0.1, (light green) 0.2, (green) 0.3, (light blue) 0.4, and (blue) 0.5 with different equivalence ratios [93].
Figure 30. Cycle efficiency curves for various hydrogen concentrations of (red) 0.0, (orange) 0.1, (light green) 0.2, (green) 0.3, (light blue) 0.4, and (blue) 0.5 with different equivalence ratios [93].
Energies 18 02252 g030
Figure 31. Temperature distributions observed in various sections for various fuel blends during air/oxygen ignition [94].
Figure 31. Temperature distributions observed in various sections for various fuel blends during air/oxygen ignition [94].
Energies 18 02252 g031
Figure 32. Oxygen levels in intake air at 1500 rpm, with a brake mean effective pressure (BMEP) of 0.4 MPa, affecting the spark ignition (SI) timing and NOx emissions [95].
Figure 32. Oxygen levels in intake air at 1500 rpm, with a brake mean effective pressure (BMEP) of 0.4 MPa, affecting the spark ignition (SI) timing and NOx emissions [95].
Energies 18 02252 g032
Figure 33. Images illustrating the progression of flames at an equivalency ratio of 1.1 captured under various oxygen levels [96].
Figure 33. Images illustrating the progression of flames at an equivalency ratio of 1.1 captured under various oxygen levels [96].
Energies 18 02252 g033
Figure 34. Effects of oxygen levels and equivalence ratio on flame velocity [96].
Figure 34. Effects of oxygen levels and equivalence ratio on flame velocity [96].
Energies 18 02252 g034
Figure 35. Heat release rate variation with Φ and oxygen content [96].
Figure 35. Heat release rate variation with Φ and oxygen content [96].
Energies 18 02252 g035
Figure 36. Velocity profiles across distances for various orifice sizes in pre-chamber combustion [97].
Figure 36. Velocity profiles across distances for various orifice sizes in pre-chamber combustion [97].
Energies 18 02252 g036
Figure 37. Hybrid power system using an ammonia–hydrogen dual-fuel engine. The red arrows indicate exhaust gas flow to the selective catalytic reduction (SCR). The purple dashed lines show the electrical connection between the inverter controller and generator. The orange and red chambers explain the H2, NH3, and O2 supplied for combustion [98].
Figure 37. Hybrid power system using an ammonia–hydrogen dual-fuel engine. The red arrows indicate exhaust gas flow to the selective catalytic reduction (SCR). The purple dashed lines show the electrical connection between the inverter controller and generator. The orange and red chambers explain the H2, NH3, and O2 supplied for combustion [98].
Energies 18 02252 g037
Figure 38. NO distribution occurring within the furnace for (a) 100% coal and (b) a blend of 50% NH3 and 50% coal [102].
Figure 38. NO distribution occurring within the furnace for (a) 100% coal and (b) a blend of 50% NH3 and 50% coal [102].
Energies 18 02252 g038
Figure 39. Association between NO emissions and co-firing ratio of coal to ammonia [102].
Figure 39. Association between NO emissions and co-firing ratio of coal to ammonia [102].
Energies 18 02252 g039
Figure 40. Temperature, ammonia, and oxygen fraction distributions during the ascent of two bubbles, analyzed at a time interval (t2inj) of 0.3 s [103].
Figure 40. Temperature, ammonia, and oxygen fraction distributions during the ascent of two bubbles, analyzed at a time interval (t2inj) of 0.3 s [103].
Energies 18 02252 g040
Table 1. Comparison of ammonia with hydrogen, propane, and methane in terms of critical properties [14,50,51,52,53].
Table 1. Comparison of ammonia with hydrogen, propane, and methane in terms of critical properties [14,50,51,52,53].
PropertyAmmoniaHydrogenPropaneMethane
Chemical formulaNH3 H2C3H8CH4
Octane number130111120
Density (kg/m3)0.730.092.010.66
Molar mass (g/mol)17.032.01644.09616.04
Boiling point (°C)−33.4−253−42.1−161
Adiabatic flame temperature (°C) 1800211020001950
Flashpoint (°C) 132−253−104−188
Auto-ignition temperature (°C) 650520450630
Critical pressure (MPa)11.31.294.254.6
Critical temperature (°C) 132.4−240.2196.9−82.59
Lower heating value (LHV) (MJ/kg)18.61205046.4
Higher heating value (HHV) (MJ/kg)22.5141.950.455.5
Explosive limits in air (%)15–284–752.37–9.54.4–17
Minimum ignition energy (mJ)80.0110.260.28
Maximum laminar burning velocity (m/s)0.072.910.430.37
Gravimetric hydrogen density (wt%)17.810018.225
Flammability limit (equivalence ratio)0.63–1.40 0.10–7.10.51–2.50.50–1.7
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Dwi, N.; Ischia, K.; Sung, Y. Review of Ammonia Oxy-Combustion Technologies: Fundamental Research and Its Various Applications. Energies 2025, 18, 2252. https://doi.org/10.3390/en18092252

AMA Style

Dwi N, Ischia K, Sung Y. Review of Ammonia Oxy-Combustion Technologies: Fundamental Research and Its Various Applications. Energies. 2025; 18(9):2252. https://doi.org/10.3390/en18092252

Chicago/Turabian Style

Dwi, Novianti, Kurniawati Ischia, and Yonmo Sung. 2025. "Review of Ammonia Oxy-Combustion Technologies: Fundamental Research and Its Various Applications" Energies 18, no. 9: 2252. https://doi.org/10.3390/en18092252

APA Style

Dwi, N., Ischia, K., & Sung, Y. (2025). Review of Ammonia Oxy-Combustion Technologies: Fundamental Research and Its Various Applications. Energies, 18(9), 2252. https://doi.org/10.3390/en18092252

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop