Next Article in Journal
An In Vivo Assessment of the Effect of Hexane Extract from Endlicheria paniculata Branches and Its Main Compound, Methyldehydrodieugenol B, on Murine Sponge-Induced Inflammation
Next Article in Special Issue
Electrode Materials, Structural Design, and Storage Mechanisms in Hybrid Supercapacitors
Previous Article in Journal
Comparison of Salvianolic Acid A Adsorption by Phenylboronic-Acid-Functionalized Montmorillonites with Different Intercalators
Previous Article in Special Issue
Controllable Synthesis of N2-Intercalated WO3 Nanorod Photoanode Harvesting a Wide Range of Visible Light for Photoelectrochemical Water Oxidation
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Recent Progress on Hydrogen Production from Ammonia Decomposition: Technical Roadmap and Catalytic Mechanism

1
School of Energy and Environment, Anhui University of Technology, Ma’anshan 243002, China
2
School of Electrical and Energy Power Engineering, Yangzhou University, Yangzhou 225002, China
*
Authors to whom correspondence should be addressed.
Molecules 2023, 28(13), 5245; https://doi.org/10.3390/molecules28135245
Submission received: 9 June 2023 / Revised: 30 June 2023 / Accepted: 2 July 2023 / Published: 6 July 2023
(This article belongs to the Special Issue Development of Electrochemical Energy Storage Materials)

Abstract

:
Ammonia decomposition has attracted significant attention in recent years due to its ability to produce hydrogen without emitting carbon dioxide and the ease of ammonia storage. This paper reviews the recent developments in ammonia decomposition technologies for hydrogen production, focusing on the latest advances in catalytic materials and catalyst design, as well as the research progress in the catalytic reaction mechanism. Additionally, the paper discusses the advantages and disadvantages of each method and the importance of finding non-precious metals to reduce costs and improve efficiency. Overall, this paper provides a valuable reference for further research on ammonia decomposition for hydrogen production.

1. Introduction

Reducing carbon emissions is a pressing global challenge. The development of clean and renewable sources of energy, such as wind or solar, is an effective way. At the same time, efficient and reliable energy storage and carriers are indispensable due to the intermittency and variability of such natural energy sources. Hydrogen is considered an excellent energy carrier due to its carbon-free, renewable, and environmentally friendly features [1,2,3,4]. However, low volume energy density and serious safety issues concerning storage and transportation hinder its large-scale applications. A feasible approach is reversibly storing hydrogen in liquid or solid hydrides (hydrogen carriers) and then generating hydrogen through catalytic processes as required. Among potential hydrogen storage intermediates, ammonia (NH3) is the most promising candidate due to its competitive advantages in terms of hydrogen storage capacity, zero carbon emissions, availability, cost, and safety [5,6,7,8].
Ammonias’s hydrogen content is as high as 17.6 wt%, and the hydrogen volume density is about 123 kg-H2/m3, much higher than other modern hydrogen storage systems [5]. Moreover, ammonia is readily liquefied at room temperature (25 °C) and a light pressure of ca.10atm, resulting in a higher volumetric energy density than pressurized H2 [6,7]. In addition, ammonia is widely used as an important chemical material in modern industries, such as nitrogen fertilizers, refrigerants, and NOx-reducing agents. Due to the strong, long-term demand, the infrastructure for its production, storage, and supply has been well developed. As such, compared to other zero-carbon fuels, ammonia is a widely available, practical, and affordable choice [8]. Furthermore, the extensive use of ammonia over the years has contributed to developing safety codes and standards. Thus, ammonia safety has been ensured.
Ammonia decomposition plays a crucial role in the renewable energy sources to hydrogen user’s process, as outlined in Figure 1. The development of these technologies is not synchronized. The production capacity of hydrogen and ammonia is not a significant obstacle to meeting the high demand, while the current ammonia decomposition system does not allow mass hydrogen production with adequate purity. Therefore, ammonia decomposition may become the bottleneck of the hydrogen chain, which is why it is getting dramatically increased attention [9,10,11].
In the past ten years, and especially in the past five, pioneering research and follow-up studies have resulted in new conceptions, methods, and materials for ammonia decomposition. These approaches and underlying mechanisms urgently need to be summarized and understood in detail and are reviewed herein. This paper first analyzes the benefits and drawbacks of various technologies and then describes their characteristics, principles, and mechanisms. Afterward, the composition, performance, and synthesis methods of a series of catalysts are covered.

2. Categories of NH3 Decomposition

The utilization of NH3 as a carrier for H2 has become more widespread due to its beneficial properties and importance in industrial settings, as well as its ability to mitigate air pollution [2,12]. Compared to H2, NH3 is more easily liquefied and stored, making it a more feasible alternative.
N H 3 g 1 2 N 2 g + 3 2 H 2 g H = + 46   k J / m o l
To produce clean H2, various technologies have been developed, such as thermocatalytic NH3 decomposition, photocatalytic NH3 decomposition, plasma-catalytic NH3 decomposition, and electrocatalytic NH3 decomposition. Figure 2 shows the advantages and disadvantages of each method [10,11,13,14,15,16,17,18]. Understanding these methods is key to determining the current capabilities of the NH3-to-H2 conversion technologies. Figure 3 compares the performance of various technical methods for ammonia decomposition. The results were obtained under different conditions, enabling a thorough evaluation of the effectiveness of each approach [10,19,20,21,22,23,24,25]. Further discussion has been explored regarding the benefits and drawbacks of each approach below.

2.1. Thermocatalytic NH3 Decomposition

It should be noted that without a catalyst, the temperature needed for ammonia decomposition is quite high. This is due to the strong hydrogen bonds present within ammonia molecules, which demand a high level of energy to break apart. As a result, ammonia molecules break down into hydrogen and nitrogen [10,13,26,27]. Therefore, ammonia decomposition without a catalyst is not practical, as it necessitates high temperature and energy consumption as well as having a low reaction rate. Consequently, catalysts have been created to hasten the sluggish kinetics of NH3 decomposition and stimulate H2 production, as shown in Figure 4.
The catalytic dissociation of ammonia can be explained as a process where ammonia molecules are absorbed into the active site of the catalyst, leading to a sequence of reactions involving dehydrogenation and recombinative desorption [7,9,28]. The reaction is generally accepted to occur in several steps, as shown in Scheme 1. Firstly, there is a molecular adsorption of NH3 that generates surface NH3* at the active sites of metal catalysts. Then, NH3* undergoes incremental dehydrogenation, resulting in the production of N* and H* atoms. Finally, N* and H* undergo recombinative desorption, leading to the release of N2 and H2 into the gas phase. These reactions culminate in the creation of di-nitrogen and di-hydrogen on the surface of the catalyst.
Regarding the ammonia synthesis reaction, dissociative nitrogen adsorption is typically viewed as the rate-determining step. Nevertheless, there’s a possibility that recombinative nitrogen desorption could also impact the rate of the decomposition reaction. This matter requires further investigation and research to determine its actual impact on the reaction kinetics. Lucentini et al. conducted a survey and analysis of various catalytic systems. The technological status and challenges of ammonia decomposition were first presented in their paper. Then, state-of-the-art catalytic systems and related reaction mechanisms were described in detail. Finally, the structured reactors for the decomposition reaction of ammonia were explored [10]. In summary, the nitrogen desorption is the dominant slow step, but the complexity of the reaction increases with the catalyst composition, active phase, and reaction conditions.
Regarding the catalysts, there has been a growing interest in using transition metal-based catalysts for NH3-to-H2 conversion. Although Ru-based catalysts have demonstrated excellent performance, they are less practical for large-scale applications due to their high cost [30,31,32]. Therefore, Fe and Ni-based catalysts have emerged as promising alternatives due to their superior performance, but efforts are still underway to further reduce costs, optimize activity, and increase their operational lifespan [33,34,35,36,37,38]. One approach to improving the activity of these catalyst systems is to use rare earth element-based heteroatoms, such as La and Ce, for doping. This results in improving the surface properties, increasing the number of active sites, and enhancing the catalytic activity of Fe and Ni-based catalysts [39,40,41,42,43]. Another approach is to create multi-metal alloy-based structures, which can further enhance the catalytic activity by providing better dispersion of the metal atoms, improving the stability of the catalyst, and increasing the resistance to poisoning [28,44,45,46,47,48,49].
The conventional process of H2 production through thermocatalytic NH3 decomposition is slow. It requires external energy input to heat the catalyst to at least 400 ℃ for a sufficient time, resulting in time delay and low energy efficiency. Therefore, a more suitable approach for the H2 production process is needed that does not require external energy input. One such approach involves exposing a pretreated catalyst to a mixture of NH3 and O2 gases at a specific ratio at room temperature. This approach is known as low-temperature ammonia oxidation, and it removes the need for the high-temperature decomposition of NH3. The reaction proceeds as follows:
N H 3 + 1 / 4 O 2 H 2 + 1 / 2 N 2 + 1 / 2 H 2 O   ( H = 75   k J / m o l )
Katsutoshi Nagaoka et al. studied hydrogen production using an acidic RuO2/γ-Al2O3 catalyst by exposing ammonia and O2 at room temperature, as shown in Figure 5 [50]. The exothermic adsorption of ammonia on the catalyst causes the catalyst bed to rapidly heat up to the auto-ignition temperature of catalytic ammonia, resulting in the oxidative decomposition of ammonia to generate hydrogen. A differential calorimeter and a volumetric gas sorption analyzer were used in the study to measure the heat released by the physisorption of multiple ammonia molecules and the chemisorption of ammonia onto both the RuO2 and acidic sites on the γ-Al2O3. The results showed a significant amount of heat evolved during both processes.

2.2. Non-Thermal Plasma-Catalytic NH3 Decomposition

Thermocatalytic decomposition of NH3 generally occurs at a temperature above 873 K, while traditional thermal reforming or decomposition of NH3 typically requires a high temperature of about 1300 K. Despite the presence of a catalyst, the reaction temperature is still relatively high, restricting the reaction’s useful applications. To start chemical reactions at lower temperatures, researchers have investigated the potential of alternate strategies involving electrical discharges or non-thermal plasmas [18,21,51]. Researchers have explored the potential of plasma-catalytic NH3 decomposition. The results show that non-thermal plasma can effectively transform NH3 into H2 with less influence from the gas temperature [52,53,54]. The results of studies conducted on plasma-catalytic NH3 decomposition suggest that it can be produced with lower energy consumption and higher efficiency [22,55,56,57].
Generally, the decomposition of ammonia induced through the direct application of plasma is restrained due to the following reasons: (1) Energy efficiency: Direct application of plasma does not now have any significant advantages in terms of energy use in contrast to catalytic decomposition [58,59,60]. This is because, beneath equilibrium conditions, ammonia can decompose at relatively low temperatures. (2) Plasma stabilization: When ammonia undergoes direct-discharge conditions, it decomposes to produce nitrogen and high concentrations of hydrogen [61]. This makes it challenging to stabilize the plasma, which similarly complicates the process of ammonia decomposition. However, one can still take advantage of the benefits of plasma, such as the speedy introduction of high-temperature reaction conditions. In the catalytic reaction system, plasma can function solely in the preliminary start-up stage segment till the catalyst bed is heated up to the reaction temperature [14]. In this case, hydrogen can be produced immediately at the beginning of the plasma system via the capability of a decomposition reaction. At the same time, the catalyst can be heated quickly via the warmth supplied via the plasma.
The mechanism for the plasma-assisted conversion of NH3 can be divided into two parts. (1) Plasma mechanisms: These involve electron-induced chemistry, such as electron impact reactions, quenching of excited species, and ion reactions. The energetic electrons in the plasma interact with the NH3 molecules, leading to various chemical reactions. (2) Thermal mechanisms: These involve conventional thermal mechanisms for thermally induced chemistry. The heat generated by the plasma contributes to the overall conversion of NH3. Thermal energy helps break and form chemical bonds, facilitating the conversion process. Bang et al. conducted an experimental and numerical study of the plasma-assisted conversion of NH3 (Figure 6). They used a mixture of NH3 (1 mol%) diluted in N2 at atmospheric pressure. This allowed them to investigate the effects of plasma on NH3 conversion under specific conditions [62]. They observed that the conversion of NH3 took place even at room temperature during their study. They found a local peak in conversion at around 600 K. This finding suggests that the plasma-assisted cracking of NH3 has the potential to be cost-effective by reducing the temperature range required for the cracking reaction. Lowering the temperature window for the cracking reaction can lead to energy savings and improved efficiency in the conversion process.

2.3. Electrocatalytic NH3 Decomposition

The electrochemical process is also a promising alternative for onboard use, as hydrogen and nitrogen can be gained from the decomposition of ammonia at a moderate temperature [17,63,64]. Liquid ammonia has a theoretical electrolysis voltage of 0.077 V, significantly lower than water’s electrolysis voltage of 1.23 V [16,63]. When ammonia is electrolyzed, hydrogen molecules are produced at the cathode, and amide ions are released. Amide ions are oxidized at the anode to form nitrogen molecules. The electrolysis reactor must be designed as a highly closed electrolytic cell under stringent experimental conditions to prevent the oxidation and hydration of metal amides [63]. However, the current efficiency is only 85% at a high cell voltage of 2 V due to the inevitable reverse reaction in liquid ammonia. To make ammonia electrolysis practical, it is necessary to reduce the cell voltage as much as possible at a high current.
The electrocatalytic NH3 decomposition can occur in an aqueous electrolyte under certain pH conditions, either through the NH3 oxidation in an alkaline environment caused by OH ions after the adsorption of NH3 onto the electrode or through NH4+ oxidation by the oxidants such as hypochlorous acid in low pH environments [65,66,67]. However, electro-corrosion and sluggish kinetics in acidic electrolytes can lead to low efficiency in the electrochemical process. Alternative approaches to address this issue involve using alkaline electrolytes for electrocatalytic NH3 decomposition, which have been extensively studied and offer the potential to overcome undesirable problems related to acidic environments towards electrode materials. Ideally, in an NH3 decomposition system, NH3 is oxidized to N2 at the anode, while H2O is reduced to H2 at the cathode, producing the gaseous H2 and N2 from an ammonia solution. However, it is well known that the side reaction of OER may be a competing process with the Electrocatalytic NH3 decomposition at the anode.
Anode reaction:
2 N H 3 + 6 O H N 2 + 6 H 2 O + 6 e   E θ = 0.77   V   v s .   S H E
Cathode reaction:
2 H 2 O + 2 e H 2 + 2 O H   E θ = 0.829 V     v s .   S H E
Overall reaction:
2 N H 3 3 H 2 + N 2   E θ = 0.059   V
The use of electrocatalytic NH3 decomposition to produce H2 has many benefits, including simplicity and cost-effectiveness. However, problems remain, such as slow kinetics and poor selectivity. To increase the electrocatalytic NH3 decomposition efficiency, various types of electrocatalysts have been studied, but their performance still cannot satisfy the requirements of industrial applicability. Therefore, it is important to develop new high-efficiency ammonia electrolysis electrodes for hydrogen production, as shown in Figure 7. Although platinum and other precious metal-based catalysts are effective for the process of electrocatalytic NH3 decomposition (alkaline water electrolysis), their high cost and limited availability are significant drawbacks [16,17,68,69,70]. Recently, catalysts based on transition metals have shown promise for electrocatalytic NH3 decomposition through various structural and morphological engineering techniques, including alloyed/core-shell formation, shape control, heteroatom doping, and self-supporting materials [40,66,71,72,73]. However, the current density of oxidation and selectivity achieved by these catalysts is still lower than what is required for practical application. Therefore, further research and development in this area is necessary.

2.4. Photocatalytic NH3 Decomposition

The photocatalytic method of splitting NH3 into N2 and H2 is also a promising approach, as it can be carried out using recyclable catalysts at room temperature and can easily control light exposure with a switch. Using sunlight to decompose ammonia through photocatalysis is an artificial photosynthetic reaction that occurs in alkaline conditions [15]. As shown in Figure 8, when the energy of the irradiating light exceeds the band gap of the photocatalyst it creates electron-hole pairs. The resulting holes act as strong oxidants, while the electrons in the conduction band can reduce O2 to form hydroxyl radicals. To achieve the photocatalytic degradation of ammonia, the electrons and holes generated on the surface of a semiconductor must have appropriate reduction and oxidation abilities to interact with the adsorbed species on the catalyst surface and produce free radicals or other products.
P h o t o c a t a l y s t + h v e + h +
2 N H 3 + 6 h + N 2 + 6 H +
2 H + + 2 e H 2
So far, only a limited number of photocatalysts have been found to be effective in the decomposition of an aqueous ammonia solution. These include common photocatalysts, such as TiO2, ZnO, ZnS, Mo2N, and graphene, as well as their hybrid materials that are loaded with metals [15,74,75,76,77,78,79,80,81,82,83]. Utsunomiya et al. studied the photocatalytic activity of TiO2 loaded with different metals on the decomposition of ammonia and explored the NH3 decomposition mechanism by proposing three reaction pathways [83]. As shown in Figure 9, route 1 involves the formation of NH radicals by removing one hydrogen atom from each of the two NH2 radicals, while route 2 involves the coupling of adjacent NH2 radicals to form NH2-NH2. Route 2′ involves the formation of NH2-NH2 through the formation of H2N-NH3. The activation energies for routes 1 and 2 calculated by density functional theory (DFT) are 236 kcal/mol and 74.8 kcal/mol, respectively. Of these pathways, route 2 is considered energetically more favorable than route 1. Possible reaction pathways for the formation of N2 and H2 through NH2-NH2 coupling were further divided into two routes: route 2, which involves the coupling of NH2 radicals to form H2N-NH2, and route 2′, which involves the interaction of NH2 with one NH3 molecule in the gas phase. The activation energies of the two routes were estimated to be 74.4 kcal/mol and 59.2 kcal/mol, respectively. Therefore, it is plausible that NH3 decomposition occurs via the formation of NH2-NH2 through routes 2 and 2′. Razak et al. reported the linear generation of H2 from NH3 over Pd/TiO2 catalyst. They suggest that the production of H2 from NH3 was initiated by the interaction of N atoms on the Pd surface. This interaction led to the dissociation of N-H bonds under the catalysis of photogenerated electrons [84].
Recently, Yuan et al. have discovered that Cu-Fe-AR, which is not an effective thermocatalyst, can function as an excellent photocatalyst for NH3 decomposition when exposed to short-pulse laser illumination [85]. The formation of adsorbate–metal excited states by hot carriers leads to the lowering of activation barriers, active site cleaning, and product desorption, resulting in enhanced reactivity and stability of Cu-Fe-AR compared to traditional thermocatalysts. Under continuous-wave LED illumination, Cu-Fe-AR manifests a high efficiency comparable to Cu-Ru-AR. Although Cu-Ru-AR exhibits slightly greater reactivity due to photothermal heating, the study demonstrates that photocatalysis can be efficiently performed with inexpensive LED photon sources, as shown in Figure 10. The results suggested that abundantly available metals might serve as productive and cost-effective catalyst substrates for plasmonic antenna-reactor photocatalysis.

3. Catalysts for NH3 Decomposition

The choice of catalyst is critical for the efficient decomposition of ammonia to produce hydrogen at low temperatures, and different technologies require different catalysts. For example, the Ru group of catalysts is commonly used in thermocatalytic NH3 decomposition, while the Pt group is suitable for both electrocatalytic and photocatalytic NH3 decomposition. In addition to co-catalysts, such as Ru and Pt, photocatalysts are also required in photocatalytic NH3 decomposition [9,26,72,73]. In non-thermal plasma-catalytic NH3 decomposition, Fe-based catalysts have been the most studied [48,51,54]. Furthermore, for the industrialization of hydrogen production from ammonia decomposition, catalyst support is vital as it has a synergistic effect on the catalyst’s performance. Therefore, this review will focus on precious metal catalysts, non-precious metal catalysts, multi-alloy catalysts, and other aspects, emphasizing the research emphasis on various catalysts in different NH3 decomposition techniques. By understanding the unique roles catalysts play in different technologies, researchers can optimize performance and improve the overall efficiency of the ammonia decomposition process.

3.1. Precious Metal Catalysts

Precious metals are a significant type of catalyst material, with Ru-based materials being particularly noteworthy for their exceptional performance in NH3 decomposition [7,9,10,31]. As a result, they have been extensively studied. The catalysis of NH3 decomposition on Ru-based catalysts is primarily attributed to B5 active sites [31,86,87,88]. The B5-type site comprises a configuration of three Ru atoms in one layer and two additional Ru atoms in the layer directly above it at a monoatomic step on a Ru(0001) terrace [89], as shown in Figure 11. Thus, the number of B5 active sites is frequently linked to the size and shape of Ru nanoparticles on the catalyst surface [90,91].
Kim et al. investigated the influence of the crystalline phase of alumina on the decomposition of NH3 over Ru catalysts supported on alumina [87]. The results show that the dispersion and morphology of Ru are key factors for H2 production from NH3 decomposition over Ru/Al2O3 catalysts and can be regulated by the support and calcination temperature. Among the Ru/Al2O3 catalysts calcined at various temperatures and reduced at 573 K, Ru/γ-Al2O3 with Ru particle sizes of 7~8 nm displayed the highest rate of ammonia decomposition. Feng et al. have demonstrated that Yttrium oxide (Y2O3) can effectively disperse and stabilize Ru nanoparticles as a functional support for catalysts [92]. Highly dispersed Ru nanoparticles on Y2O3 (Ru/Y2O3) were prepared, and the catalytic activity of ammonia decomposition was evaluated. The results showed that at a relatively high weight gas hourly space velocity of 30,000 mLg−1h−1, the 5 wt.% Ru/Y2O3 catalyst exhibited excellent catalytic activity, achieving near-complete NH3 conversion at 475 °C, superior to many Ru catalysts supported by typical oxides. Cha et al. reported the successful loading of highly monodisperse Ru particles onto alkali-exchanged zeolite Y support using an ion-exchange method coupled with a vacuum calcination treatment [93]. As shown in Figure 12, the resulting Ru/M-Y catalysts exhibit an average particle size of approximately 1 nm and contain both nanometer- and sub-nanometer-sized Ru particles. Hu et al. reported the successful synthesis of stable ruthenium (Ru) clusters with a size of about 1.5 nm by reduction of single Ru atoms under an ammonia atmosphere at 550 °C [94]. These Ru clusters are uniformly dispersed on the surface of the ceria. The obtained supported ruthenium cluster catalysts exhibit exceptional activity for ammonia decomposition with an extremely high hydrogen production rate. Ru-based materials have potential applications in Photocatalytic NH3 decomposition and Plasma-catalytic NH3 decomposition. Iwase et al. have reported that a Ru-loaded ZnS photocatalyst demonstrated remarkable activity in decomposing an aqueous ammonia solution into H2 and N2 with a stoichiometric amount under simulated solar radiation [79]. In this reaction, the Ru co-catalyst was crucial in promoting ammonia oxidation to N2.
On the other hand, the support used also strongly affects the catalytic activity of Ru-based materials. Firstly, a support with a high specific surface area can enhance the dispersal of Ru nanoparticles, leading to improved NH3 adsorption capacity [92,93,94]. Secondly, the characteristics of the support itself can have a synergistic effect with Ru nanoparticles, thereby enhancing catalytic activity [23,95,96,97,98,99]. For example, Jeon et al. synthesized a Y-doped BaCeO3 perovskite and constructed a strong metal support interaction (SMSI) interface with Ru particles, as shown in Figure 13 [95]. The catalyst achieved 100% NH3 decomposition efficiency at 500 °C and had a higher H2 generation rate than most Ru-based catalysts. The improvement in catalytic performance can be attributed to the promotion of N2 desorption by the SMSI interface. Yamazaki et al. studied Ru catalysts supported on CeO2−PrOx composites (Ru/CP) prepared through coprecipitation [99]. The study found that the activity of the Ru/CP catalyst is higher than that of the Ru/CeO2 and Ru/PrOx catalysts, and the Pr content (=Pr/(Ce + Pr)) reached the target value of 33–67%. It was also revealed that increasing the Pr content in the catalysts enhanced the Ru dispersion but suppressed Ru metalation due to a strong “metal–support interaction”.
Apart from Ru-based materials, Pt-based materials are also extensively studied in the field of NH3 decomposition. While Ru-based materials are mainly used for thermocatalytic NH3 decomposition, Pt-based materials are typically employed for photocatalytic and electrocatalytic NH3 decomposition [16,17,68,69,75,76,100]. In a study by Kominami et al., the photocatalytic NH3 decomposition performance of metal-supported TiO2 catalysts was investigated [76]. Pt/TiO2 exhibited the best performance, with the amount of H2 and N2 reaching 90 and 28 umol, respectively, after 4 h of light irradiation, as shown in Figure 14. The H2/N2 ratio was 3.2, consistent with the stoichiometric ratio of NH3 decomposition. Dong et al. successfully electrolyzed liquid ammonia with metal amide as the supporting electrolyte using Pt electrodes [16]. In a subsequent study, they replaced the Pt electrode with a Pt/Rh/Ir alloy electrode, which tripled the current density [69].

3.2. Non-Precious Metal Catalysts

Although precious metal materials have excellent NH3 decomposition ability, their high cost as precious metals limits their potential for large-scale applications. Therefore, researchers are exploring the development of non-precious metal catalysts. Currently, non-precious metals such as Ni, Co, and Fe have demonstrated good NH3 decomposition activity and are being investigated as potential alternatives [6,9,10,101].
Ni-based catalysts have received significant attention among numerous non-precious metal catalysts due to their lower cost and significant activity [7,102]. The activity of Ni-based catalysts is primarily influenced by the size of Ni nanoparticles present on the catalyst surface [103,104,105]. Deng et al. obtained Ni nanoparticles with varying particle sizes through direct thermal reduction of nickel-containing Ca-Al layered double hydroxides (LDH) and tested them for NH3 decomposition, as shown in Figure 15 [104]. Among these catalysts, the Ni/CaAlOx catalyst with Ni(NO3)2 as the precursor exhibited the best NH3 decomposition performance, with an NH3 conversion rate of 99% at 550 °C. This is because the average particle size of the Ni nanoparticles, reduced with Ni (NO3)2 as the precursor, is the smallest (4.7 nm), allowing for better dispersion on the surface of CaAlOx and, thus, more effective adsorption of NH3.
Besides the size of Ni nanoparticles, the support used in Ni-based catalysts also plays a crucial role in their catalytic activity. Currently, metal oxides such as CeO2 [43,106,107,108,109], Y2O3 [110,111], Al2O3 [112,113,114], and La2O3 [115] are commonly used as supports for Ni-based catalysts used in NH3 decomposition. The high activity of these metal oxide-supported Ni-based catalysts is typically due to the high dispersity of Ni nanoparticles, the strong interaction between the support and Ni nanoparticles, and the presence of oxygen vacancies on the support surface. Li et al. successfully used a Ni single-atom supported CeO2 (SA Ni/CeO2) catalyst for NH3 decomposition into H2 under solar heating conditions [109]. The experimental results showed that at 300 °C, the rate of NH3 decomposition to H2 reached 3.544 mmol g−1 min−1, which was superior to all non-precious metal catalysts and most precious metal catalysts. Based on this, they combined their self-made solar heating system, and the H2 generation rate of the catalyst under 1 sun reached 1.58 mmol g−1 min−1, which was 100 times higher than the previous record. Do et al. synthesized a series of Ni/AlCeOx composites using the cation-anion double hydrolysis method and used them as catalysts for NH3 decomposition to H2 production [114]. The experimental results showed that the catalyst exhibited excellent catalytic performance and could completely decompose NH3 at around 550 °C. The improved catalytic performance can be attributed to the highly dispersed and reducible Ni atoms, the increased amount of surface oxygen vacancies, and the significantly enhanced NH3 adsorption affinity.
Co-based materials have also been extensively studied as catalysts for NH3 decomposition [116,117,118]. Many properties of Co-based catalysts are similar to those of Ni-based catalysts, and their catalytic activity is also influenced by the size of Co nanoparticles and the support structure [119,120,121,122,123,124,125]. Lei et al. successfully prepared small Co nanoparticles dispersed on N-doped carbon carriers (Co/NC-X) by pyrolyzing ZIF-67 in an N2 atmosphere at different temperatures (X = 500, 600, 700, 800 °C). They used them as catalysts for NH3 decomposition to H2, as shown in Figure 16 [125]. The Co NPs were evenly distributed and highly dispersed on NC, resulting in high catalytic activity. Among these catalysts, the NH3 conversion rate of Co/NC-600 at 500 °C was 80%, and the H2 production rate was 26.8 mmol g−1 min−1. Yu et al. investigated the effect of BaNH, CaNH, and Mg2N3 on the catalytic activity of Co in the NH3 decomposition reaction [124]. In the reaction temperature range of 300–550 °C, the formation rate of H2 was in the order of Co-BaNH > Co-CaNH > Co-Mg2N3. Notably, the H2 generation rate of Co-BaNH at 500 °C reached 20 mmol g−1 min−1, comparable to the active Ru/Al2O3. In-depth studies revealed that a [Co-N-Ba]-like intermediate species was formed at the interface of Co metal and BaNH, which led to the higher catalytic activity of Co-BaNH.
The NH3 decomposition activity of Fe-based materials is generally lower than that of Ni-based materials and Co-based materials due to the higher bond energy of Fe-N compared to Ni-N and Co-N [9,101]. However, despite this, researchers still have great interest in Fe-based materials due to their low price [36,37,38,52,55,57,126,127]. Lu et al. found a hidden active phase in Fe-based catalysts for NH3 decomposition, with the highest dehydrogenation rate corresponding to an evanescent Fe/Fe4N mixing phase [38]. However, the deposition of excess nitrogen atoms on the Fe-based catalyst can cause deactivation and a decrease in catalytic efficiency. Chen et al. observed a “particle size effect” in the activity of Fe3O4/Al2O3 catalyst in NH3 decomposition, with TOFs increasing as the size of Fe3O4 nanoparticles increased [57]. As shown in Figure 17, this effect was also observed in plasma-catalyzed NH3 decomposition.

3.3. Multi-Metallic Catalysts

To overcome the limitations of single noble metal catalysts and single non-noble metal catalysts in the large-scale application of NH3 decomposition, researchers are exploring the use of multi-metallic catalysts. These catalysts offer a balance between cost and performance, with low cost, good catalytic performance, and high stability. Additionally, the composition and morphology of different metals in the catalysts can be regulated, providing more possibilities for optimization.
At present, bimetallic catalysts are the mainstream of multi-metallic catalysts [28]. Common bimetallic catalysts include Ru-Ni [128,129,130], Ni-Co [42,46,49,131,132], Ni-Fe [47,48,133], etc. Hansgen et al. developed a computational framework that utilizes nitrogen binding energies to identify potential monolayer bimetallic catalysts [134]. Through this framework, they were able to predict that Ni-Pt-Pt(111) would be an effective bimetallic catalyst for NH3 decomposition, even exceeding the activity of Ru. Their calculations proved to be successful in identifying this high-performing catalyst. Tabassum et al. synthesized CoNi alloy nanoparticles that were well dispersed on mixed oxide support of MgO-CeO2-SrO, with potassium promotion, resulting in the K–CoNialloy–MgO–CeO2–SrO catalyst, as shown in Figure 18 [49]. This catalyst exhibited high efficiency in converting NH3, with a conversion rate of 97.7% under reaction conditions of 450 °C and 6000 mLh−1gcat−1. The study suggests that the active sites located on the metal/oxide interface promote the recombination of adsorbed N atoms, leading to N2 desorption and a significant reduction in activation energy barriers. Shiraishi et al. developed a bimetallic alloy nanoparticle-supported TiO2 photocatalyst (Pt0.9Au0.1/TiO2) with 90 mol% Pt and 10 mol% Au [20]. This photocatalyst demonstrated effective decomposition of NH3 into H2 and N2, with significantly higher catalytic activity than Pt/TiO2. The enhanced performance can be ascribed to the presence of Au in the alloy, which reduces the Schottky barrier height at the metal/TiO2 interface, leading to a more efficient transfer of conduction band electrons on TiO2 to the metal particles. Yi et al. conducted a study where they prepared various bimetallic catalysts (Fe-Co, Mo-Co, Fe-Ni, and Mo-Ni) and utilized them for plasma-catalyzed NH3 decomposition [22]. The results revealed that the Fe-Ni catalyst displayed the highest catalytic activity compared to the other catalysts. Further analysis indicated that the Fe-Ni catalyst had the highest NH3 adsorption capacity, likely the primary reason for its superior catalytic activity. Jiang et al. investigated various morphologies of NiCo2N compounds. They specifically focused on the structure of nanoneedles grown on 3D nickel foam. It was found that this particular structure exhibited several advantages, including a significantly increased surface area, more exposed active sites, enhanced charge transfer, and improved gas diffusion. Furthermore, the NiCo2N composite demonstrated the most optimal catalytic activity in both the hydrogen evolution reaction (HER) and ammonia electrolysis [135].
Developing multi-metal catalysts containing three or more metal elements is a more challenging task, but it also holds greater potential. Xie et al. successfully utilized a novel high-entropy alloy (HEA) CoMoFeNiCu nanoparticles for highly efficient NH3 decomposition [44]. As illustrated in Figure 19, they overcame the miscibility limit of bimetallic CoMo alloys by stably tuning the Co/Mo elemental ratio in CoMoFeNiCu HEA nanoparticles. The HEA catalyst has a lower cost than Ru-based catalysts and exhibits better catalytic performance than Co-Mo catalysts. At a reaction temperature of 500 °C, the optimal conversion efficiency of NH3 using this catalyst can reach 100%. Furthermore, the alloy composition and surface adsorption properties of the HEA catalyst can be well-tuned, demonstrating its immense potential for practical applications.

3.4. Other Catalysts

In addition to metal nanoparticles, some metal nitrides can also be used as NH3 decomposition catalysts, such as Mo2N [136,137,138], MnN [139], Co3Mo3N [140,141,142], etc. Huo et al. designed Mo2N nanocrystals anchored on two-dimensional (2D) mesoporous silica/reduced graphene oxide (rGO) hybrid nanosheets (Mo2N/SBA-15/rGO) as an NH3 decomposition catalyst [136]. Benefiting from the highly dispersed Mo2N nanocrystals and modest Mo-N band strength, the catalyst exhibits excellent catalytic performance with an ammonia decomposition rate as high as 30.58 mmol g−1 min−1. Srifa et al. synthesized Mo2N-based catalysts with Fe, Ni, and Co additives, as shown in Figure 20 [141]. The NH3 decomposition activities of Co3Mo3N, Ni3Mo3N, and Fe3Mo3N catalysts were higher than Mo2N catalysts. Among them, Co3Mo3N has the best catalytic performance, and the NH3 conversion efficiency reaches 94% at 550 °C. The increase in catalytic activity can be attributed to adding Co, Ni, and Fe to increase the particle size and specific surface area of the catalyst and promote the recombination and desorption of N atoms.
Researchers have also developed some NH3 decomposition catalysts of sulfides and carbides, but the performance is average. Kraupner et al. synthesized mesoporous Fe3C with high crystallinity [143]. The material has transmission pores with a diameter of 20 nm and a high specific surface area (about 415 m2 g−1). At 700 °C, the conversion rate of NH3 is over 95%. Krishnan et al. successfully synthesized MoS2-supported acid, and alkaline functionalized (Al, Ti, and Zr)-Laponite catalysts [144]. The results show that the presence of heteroatoms (Al, Ti, and Zr) in the Laponite framework plays a key role in improving the reducibility, acid-base functional groups, and textural properties of MoS2. Among them, the MoS2/Zr-Laponite catalyst has the best NH3 decomposition performance, and its NH3 conversion rate at 600 °C degrees is 94%.

4. Conclusions

Ammonia decomposition is a highly efficient, environmentally friendly, and sustainable technology for hydrogen production that has attracted significant attention. This article explores the technical path and catalysts used for hydrogen production through ammonia decomposition. The article discusses various technologies and their catalytic mechanisms, highlighting their advantages and disadvantages and their potential for industrialization. The review also covers recent progress in catalyst research for ammonia decomposition, emphasizing the importance of finding non-precious metals to reduce costs and improve efficiency. Overall, this article provides valuable insights into the current state of NH3 decomposition for H2 production and the potential for future advancements in this field.

Author Contributions

Conceptualization, X.H. and W.F.; investigation, K.L. and X.L.; writing—original draft preparation, X.H. and Y.M.; writing—review and editing, W.F.; project administration, X.H.; funding acquisition, X.H. All authors have read and agreed to the published version of the manuscript.

Funding

The authors gratefully acknowledge the financial support from The University Synergy Innovation Program of Anhui Province (grant No. GXXT-2020-006 and grant No. GXXT-2022-025).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

All necessary data of this review has been included in the paper and no further data is needed.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Awad, O.I.; Zhou, B.; Harrath, K.; Kadirgama, K. Characteristics of NH3/H2 blend as carbon-free fuels: A review. Int. J. Hydrogen Energy, 2022; in press. [Google Scholar] [CrossRef]
  2. Wan, Z.; Tao, Y.; Shao, J.; Zhang, Y.; You, H. Ammonia as an effective hydrogen carrier and a clean fuel for solid oxide fuel cells. Energy Convers. Manag. 2021, 228, 113729. [Google Scholar] [CrossRef]
  3. Klerke, A.; Christensen, C.H.; Nørskov, J.K.; Vegge, T. Ammonia for hydrogen storage: Challenges and opportunities. J. Mater. Chem. 2008, 18, 2304–2310. [Google Scholar] [CrossRef]
  4. Chang, F.; Gao, W.; Guo, J.; Chen, P. Emerging Materials and Methods toward Ammonia-Based Energy Storage and Conversion. Adv. Mater. 2021, 33, e2005721. [Google Scholar] [CrossRef] [PubMed]
  5. Makhloufi, C.; Kezibri, N. Large-scale decomposition of green ammonia for pure hydrogen production. Int. J. Hydrogen Energy 2021, 46, 34777–34787. [Google Scholar] [CrossRef]
  6. Sun, S.; Jiang, Q.; Zhao, D.; Cao, T.; Sha, H.; Zhang, C.; Song, H.; Da, Z. Ammonia as hydrogen carrier: Advances in ammonia decomposition catalysts for promising hydrogen production. Renew. Sustain. Energy Rev. 2022, 169, 112918. [Google Scholar] [CrossRef]
  7. Le, T.A.; Do, Q.C.; Kim, Y.; Kim, T.-W.; Chae, H.-J. A review on the recent developments of ruthenium and nickel catalysts for COx-free H2 generation by ammonia decomposition. Korean J. Chem. Eng. 2021, 38, 1087–1103. [Google Scholar] [CrossRef]
  8. Lim, D.; Kim, A.; Cheon, S.; Byun, M.; Lim, H. Life cycle techno-economic and carbon footprint analysis of H2 production via NH3 decomposition: A Case study for the Republic of Korea. Energy Convers. Manag. 2021, 250, 114881. [Google Scholar] [CrossRef]
  9. Mukherjee, S.; Devaguptapu, S.V.; Sviripa, A.; Lund, C.R.F.; Wu, G. Low-temperature ammonia decomposition catalysts for hydrogen generation. Appl. Catal. B Environ. 2018, 226, 162–181. [Google Scholar] [CrossRef]
  10. Lucentini, I.; Garcia, X.; Vendrell, X.; Llorca, J. Review of the Decomposition of Ammonia to Generate Hydrogen. Ind. Eng. Chem. Res. 2021, 60, 18560–18611. [Google Scholar] [CrossRef]
  11. Tran, D.T.; Nguyen, T.H.; Jeong, H.; Tran, P.K.L.; Malhotra, D.; Jeong, K.U.; Kim, N.H.; Lee, J.H. Recent engineering advances in nanocatalysts for NH3-to-H2 conversion technologies. Nano Energy 2022, 94, 106929. [Google Scholar] [CrossRef]
  12. Juangsa, F.B.; Irhamna, A.R.; Aziz, M. Production of ammonia as potential hydrogen carrier: Review on thermochemical and electrochemical processes. Int. J. Hydrogen Energy 2021, 46, 14455–14477. [Google Scholar] [CrossRef]
  13. Peters, S.; Abdel-Mageed, A.M.; Wohlrab, S. Thermocatalytic Ammonia Decomposition—Status and Current Research Demands for a Carbon-Free Hydrogen Fuel Technology. ChemCatChem 2022, 15, e202201185. [Google Scholar] [CrossRef]
  14. Li, S.; Miao, P.; Zhang, Y.; Wu, J.; Zhang, B.; Du, Y.; Han, X.; Sun, J.; Xu, P. Recent Advances in Plasmonic Nanostructures for Enhanced Photocatalysis and Electrocatalysis. Adv. Mater. 2021, 33, e2000086. [Google Scholar] [CrossRef] [PubMed]
  15. Zhang, S.; He, Z.; Li, X.; Zhang, J.; Zang, Q.; Wang, S. Building heterogeneous nanostructures for photocatalytic ammonia decomposition. Nanoscale Adv. 2020, 2, 3610–3623. [Google Scholar] [CrossRef] [PubMed]
  16. Dong, B.X.; Ichikawa, T.; Hanada, N.; Hino, S.; Kojima, Y. Liquid ammonia electrolysis by platinum electrodes. J. Alloys Compd. 2011, 509, S891–S894. [Google Scholar] [CrossRef]
  17. Little, D.J.; Smith, I.I.I.M.R.; Hamann, T.W. Electrolysis of liquid ammonia for hydrogen generation. Energy Environ. Sci. 2015, 8, 2775–2781. [Google Scholar] [CrossRef]
  18. Lee, D.H.; Kang, H.; Kim, Y.; Song, H.; Lee, H.; Choi, J.; Kim, K.-T.; Song, Y.-H. Plasma-assisted hydrogen generation: A mechanistic review. Fuel Process. Technol. 2023, 247, 107761. [Google Scholar] [CrossRef]
  19. Mohammadi, Z.; Sharifnia, S.; Shavisi, Y. Photocatalytic degradation of aqueous ammonia by using TiO2ZnO/LECA hybrid photocatalyst. Mater. Chem. Phys. 2016, 184, 110–117. [Google Scholar] [CrossRef]
  20. Shiraishi, Y.; Toi, S.; Ichikawa, S.; Hirai, T. Photocatalytic NH3 Splitting on TiO2 Particles Decorated with Pt–Au Bimetallic Alloy Nanoparticles. ACS Appl. Nano Mater. 2020, 3, 1612–1620. [Google Scholar] [CrossRef]
  21. Gao, Y.; Hu, E.; Yi, Y.; Yin, G.; Huang, Z. Plasma-assisted low temperature ammonia decomposition on 3d transition metal (Fe, Co and Ni) doped CeO2 catalysts: Synergetic effect of morphology and co-doping. Fuel Process. Technol. 2023, 244, 107695. [Google Scholar] [CrossRef]
  22. Yi, Y.; Wang, L.; Guo, Y.; Sun, S.; Guo, H. Plasma-Assisted ammonia decomposition over Fe–Ni alloy catalysts for COx-Free hydrogen. AIChE J. 2018, 65, 691–701. [Google Scholar]
  23. Le, T.A.; Kim, Y.; Kim, H.W.; Lee, S.-U.; Kim, J.-R.; Kim, T.-W.; Lee, Y.-J.; Chae, H.-J. Ru-supported lanthania-ceria composite as an efficient catalyst for COx-free H2 production from ammonia decomposition. Appl. Catal. B Environ. 2021, 285, 119831. [Google Scholar] [CrossRef]
  24. Shavisi, Y.; Sharifnia, S.; Mohamadi, Z. Solar-light-harvesting degradation of aqueous ammonia by CuO/ZnO immobilized on pottery plate: Linear kinetic modeling for adsorption and photocatalysis process. J. Environ. Chem. Eng. 2016, 4, 2736–2744. [Google Scholar] [CrossRef]
  25. Gu, Y.-Q.; Jin, Z.; Zhang, H.; Xu, R.-J.; Zheng, M.-J.; Guo, Y.-M.; Song, Q.-S.; Jia, C.-J. Transition metal nanoparticles dispersed in an alumina matrix as active and stable catalysts for COx-free hydrogen production from ammonia. J. Mater. Chem. A 2015, 3, 17172–17180. [Google Scholar] [CrossRef]
  26. Alturaifi, S.A.; Mathieu, O.; Petersen, E.L. An experimental and modeling study of ammonia pyrolysis. Combust. Flame 2022, 235, 111694. [Google Scholar] [CrossRef]
  27. Hayashi, F.; Toda, Y.; Kanie, Y.; Kitano, M.; Inoue, Y.; Yokoyama, T.; Hara, M.; Hosono, H. Ammonia decomposition by ruthenium nanoparticles loaded on inorganic electride C12A7:e−. Chem. Sci. 2013, 4, 3124–3130. [Google Scholar] [CrossRef]
  28. Khan, W.U.; Alasiri, H.S.; Ali, S.A.; Hossain, M.M. Recent Advances in Bimetallic Catalysts for Hydrogen Production from Ammonia. Chem. Rec. 2022, 22, e202200030. [Google Scholar] [CrossRef]
  29. Ao, R.; Lu, R.; Leng, G.; Zhu, Y.; Yan, F.; Yu, Q. A Review on Numerical Simulation of Hydrogen Production from Ammonia Decomposition. Energies 2023, 16, 921. [Google Scholar] [CrossRef]
  30. Wang, Z.; Cai, Z.; Wei, Z. Highly Active Ruthenium Catalyst Supported on Barium Hexaaluminate for Ammonia Decomposition to COx-Free Hydrogen. ACS Sustain. Chem. Eng. 2019, 7, 8226–8235. [Google Scholar] [CrossRef]
  31. Chen, C.; Wu, K.; Ren, H.; Zhou, C.; Luo, Y.; Lin, L.; Au, C.; Jiang, L. Ru-Based Catalysts for Ammonia Decomposition: A Mini-Review. Energy Fuels 2021, 35, 11693–11706. [Google Scholar] [CrossRef]
  32. Zhiqiang, F.; Ziqing, W.; Dexing, L.; Jianxin, L.; Lingzhi, Y.; Qin, W.; Zhong, W. Catalytic ammonia decomposition to COx-free hydrogen over ruthenium catalyst supported on alkali silicates. Fuel 2022, 326, 125094. [Google Scholar] [CrossRef]
  33. Wang, Y.; Mao, X.; Yang, J.; Wang, J.; Guan, W.; Chen, J.; Han, B.; Tian, Z. One-step synthesis of Ni/yttrium-doped barium zirconates catalyst for on-site hydrogen production from NH3 decomposition. Int. J. Hydrogen Energy 2022, 47, 2608–2621. [Google Scholar] [CrossRef]
  34. Qiu, Y.; Fu, E.; Gong, F.; Xiao, R. Catalyst support effect on ammonia decomposition over Ni/MgAl2O4 towards hydrogen production. Int. J. Hydrogen Energy 2022, 47, 5044–5052. [Google Scholar] [CrossRef]
  35. Kocer, T.; Saraç-Öztuna, F.E.; Kurtoğlu-Öztulum, S.F.; Unal, U.; Uzun, A. Effect of Nickel Precursor on the Catalytic Performance of Graphene Aerogel-Supported Nickel Nanoparticles for the Production of COx-free Hydrogen by Ammonia Decomposition. Energy Technol. 2022, 10, 2100794. [Google Scholar] [CrossRef]
  36. Cui, H.-Z.; Gu, Y.-Q.; He, X.-X.; Wei, S.; Jin, Z.; Jia, C.-J.; Song, Q.-S. Iron-based composite nanostructure catalysts used to produce COx-free hydrogen from ammonia. Sci. Bull. 2016, 61, 220–226. [Google Scholar] [CrossRef] [Green Version]
  37. Martirez, J.M.P.; Carter, E.A. First-Principles Insights into the Thermocatalytic Cracking of Ammonia-Hydrogen Blends on Fe(110): 1. Thermodynamics. J. Phys. Chem. C 2022, 126, 19733–19744. [Google Scholar] [CrossRef]
  38. Lu, B.; Li, L.; Ren, M.; Liu, Y.; Zhang, Y.; Xu, X.; Wang, X.; Qiu, H. Ammonia decomposition over iron-based catalyst: Exploring the hidden active phase. Appl. Catal. B Environ. 2022, 314, 121475. [Google Scholar] [CrossRef]
  39. Hu, X.-C.; Wang, W.-W.; Jin, Z.; Wang, X.; Si, R.; Jia, C.-J. Transition metal nanoparticles supported La-promoted MgO as catalysts for hydrogen production via catalytic decomposition of ammonia. J. Energy Chem. 2019, 38, 41–49. [Google Scholar] [CrossRef] [Green Version]
  40. Maslova, V.; Fourré, E.; Veryasov, G.; Nesterenko, N.; Grishin, A.; Louste, C.; Nassar, M.; Guignard, N.; Arrii, S.; Batiot-Dupeyrat, C. Ammonia Decomposition in Electric Field over Ce-based Materials. ChemCatChem 2023, 15, e202201626. [Google Scholar] [CrossRef]
  41. Podila, S.; Driss, H.; Ali, A.M.; Al-Zahrani, A.A.; Daous, M.A. Influence of Ce substitution in LaMO3 (M = Co/Ni) perovskites for COx-free hydrogen production from ammonia decomposition. Arab. J. Chem. 2022, 15, 103547. [Google Scholar] [CrossRef]
  42. Huang, C.; Li, H.; Yang, J.; Wang, C.; Hu, F.; Wang, X.; Lu, Z.-H.; Feng, G.; Zhang, R. Ce0.6Zr0.3Y0.1O2 solid solutions-supported Ni-Co bimetal nanocatalysts for NH3 decomposition. Appl. Surf. Sci. 2019, 478, 708–716. [Google Scholar] [CrossRef]
  43. Sima, D.; Wu, H.; Tian, K.; Xie, S.; Foo, J.J.; Li, S.; Wang, D.; Ye, Y.; Zheng, Z.; Liu, Y.-Q. Enhanced low temperature catalytic activity of Ni/Al–Ce0.8Zr0.2O2 for hydrogen production from ammonia decomposition. Int. J. Hydrogen Energy 2020, 45, 9342–9352. [Google Scholar] [CrossRef]
  44. Xie, P.; Yao, Y.; Huang, Z.; Liu, Z.; Zhang, J.; Li, T.; Wang, G.; Shahbazian-Yassar, R.; Hu, L.; Wang, C. Highly efficient decomposition of ammonia using high-entropy alloy catalysts. Nat. Commun. 2019, 10, 4011. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  45. Sayas, S.; Morlanés, N.; Katikaneni, S.P.; Harale, A.; Solami, B.; Gascon, J. High pressure ammonia decomposition on Ru–K/CaO catalysts. Catal. Sci. Technol. 2020, 10, 5027–5035. [Google Scholar] [CrossRef]
  46. Wu, Z.-W.; Li, X.; Qin, Y.-H.; Deng, L.; Wang, C.-W.; Jiang, X. Ammonia decomposition over SiO2-supported Ni–Co bimetallic catalyst for COx-free hydrogen generation. Int. J. Hydrogen Energy 2020, 45, 15263–15269. [Google Scholar] [CrossRef]
  47. Simonsen, S.B.; Chakraborty, D.; Chorkendorff, I.; Dahl, S. Alloyed Ni-Fe nanoparticles as catalysts for NH3 decomposition. Appl. Catal. A Gen. 2012, 447–448, 22–31. [Google Scholar] [CrossRef] [Green Version]
  48. Fu, E.; Qiu, Y.; Lu, H.; Wang, S.; Liu, L.; Feng, H.; Yang, Y.; Wu, Z.; Xie, Y.; Gong, F.; et al. Enhanced NH3 decomposition for H2 production over bimetallic M(M=Co, Fe, Cu)Ni/Al2O3. Fuel Process. Technol. 2021, 221, 106945. [Google Scholar] [CrossRef]
  49. Tabassum, H.; Mukherjee, S.; Chen, J.; Holiharimanana, D.; Karakalos, S.; Yang, X.; Hwang, S.; Zhang, T.; Lu, B.; Chen, M.; et al. Hydrogen generation via ammonia decomposition on highly efficient and stable Ru-free catalysts: Approaching complete conversion at 450 °C. Energy Environ. Sci. 2022, 15, 4190–4200. [Google Scholar] [CrossRef]
  50. Nagaoka, K.; Eboshi, T.; Takeishi, Y.; Tasaki, R.; Honda, K.; Imamura, K.; Sato, K. Carbon-free H2 production from ammonia triggered at room temperature with an acidic RuO2/gamma-Al2O3 catalyst. Sci. Adv. 2017, 3, e1602747. [Google Scholar] [CrossRef] [Green Version]
  51. Navascués, P.; Obrero-Pérez, J.M.; Cotrino, J.; González-Elipe, A.R.; Gómez-Ramírez, A. Unraveling Discharge and Surface Mechanisms in Plasma-Assisted Ammonia Reactions. ACS Sustain. Chem. Eng. 2020, 8, 14855–14866. [Google Scholar] [CrossRef]
  52. Wang, L.; Yi, Y.; Guo, Y.; Zhao, Y.; Zhang, J.; Guo, H. Synergy of DBD plasma and Fe-based catalyst in NH3 decomposition: Plasma enhancing adsorption step. Plasma Process. Polym. 2017, 14, 1600111. [Google Scholar] [CrossRef]
  53. Akiyama, M.; Aihara, K.; Sawaguchi, T.; Matsukata, M.; Iwamoto, M. Ammonia decomposition to clean hydrogen using non-thermal atmospheric-pressure plasma. Int. J. Hydrogen Energy 2018, 43, 14493–14497. [Google Scholar] [CrossRef]
  54. El-Shafie, M.; Kambara, S.; Hayakawa, Y. Alumina particle size effect on H2 production from ammonia decomposition by DBD plasma. Energy Rep. 2020, 6, 25–30. [Google Scholar] [CrossRef]
  55. Wang, L.; Zhao, Y.; Liu, C.; Gong, W.; Guo, H. Plasma driven ammonia decomposition on a Fe-catalyst: Eliminating surface nitrogen poisoning. Chem. Commun. 2013, 49, 3787–3789. [Google Scholar] [CrossRef] [PubMed]
  56. Wang, L.; Yi, Y.; Zhao, Y.; Zhang, R.; Zhang, J.; Guo, H. NH3 Decomposition for H2 Generation: Effects of Cheap Metals and Supports on Plasma–Catalyst Synergy. ACS Catal. 2015, 5, 4167–4174. [Google Scholar] [CrossRef]
  57. Chen, G.; Qu, J.; Cheah, P.; Cao, D.; Zhao, Y.; Xiang, Y. Size-Dependent Activity of Iron Nanoparticles in Both Thermal and Plasma Driven Catalytic Ammonia Decomposition. Ind. Eng. Chem. Res. 2022, 61, 11436–11443. [Google Scholar] [CrossRef]
  58. El-Shafie, M.; Kambara, S.; Hayakawa, Y. Energy and exergy analysis of hydrogen production from ammonia decomposition systems using non-thermal plasma. Int. J. Hydrogen Energy 2021, 46, 29361–29375. [Google Scholar] [CrossRef]
  59. Lin, Q.F.; Jiang, Y.M.; Liu, C.Z.; Chen, L.W.; Zhang, W.J.; Ding, J.; Li, J.G. Instantaneous hydrogen production from ammonia by non-thermal arc plasma combining with catalyst. Energy Rep. 2021, 7, 4064–4070. [Google Scholar] [CrossRef]
  60. El-Shafie, M.; Kambara, S.; Hayakawa, Y. Plasma-enhanced catalytic ammonia decomposition over ruthenium (Ru/Al2O3) and soda glass (SiO2) materials. J. Energy Inst. 2021, 99, 145–153. [Google Scholar] [CrossRef]
  61. Andersen, J.A.; Christensen, J.M.; Østberg, M.; Bogaerts, A.; Jensen, A.D. Plasma-catalytic ammonia decomposition using a packed-bed dielectric barrier discharge reactor. Int. J. Hydrogen Energy 2022, 47, 32081–32091. [Google Scholar] [CrossRef]
  62. Bang, S.; Snoeckx, R.; Cha, M.S. Kinetic Study for Plasma Assisted Cracking of NH3: Approaches and Challenges. J. Phys. Chem. A 2023, 127, 1271–1282. [Google Scholar] [CrossRef] [PubMed]
  63. Hanada, N.; Hino, S.; Ichikawa, T.; Suzuki, H.; Takai, K.; Kojima, Y. Hydrogen generation by electrolysis of liquid ammonia. Chem. Commun. (Camb) 2010, 46, 7775–7777. [Google Scholar] [CrossRef] [PubMed]
  64. Yang, Y.; Kim, J.; Jo, H.; Seong, A.; Lee, M.; Min, H.-K.; Seo, M.-g.; Choi, Y.; Kim, G. A rigorous electrochemical ammonia electrolysis protocol with in operando quantitative analysis. J. Mater. Chem. A 2021, 9, 11571–11579. [Google Scholar] [CrossRef]
  65. Goshome, K.; Yamada, T.; Miyaoka, H.; Ichikawa, T.; Kojima, Y. High compressed hydrogen production via direct electrolysis of liquid ammonia. Int. J. Hydrogen Energy 2016, 41, 14529–14534. [Google Scholar] [CrossRef] [Green Version]
  66. Akagi, N.; Hori, K.; Sugime, H.; Noda, S.; Hanada, N. Systematic investigation of anode catalysts for liquid ammonia electrolysis. J. Catal. 2022, 406, 222–230. [Google Scholar] [CrossRef]
  67. Ponikvar, Ž.; Likozar, B.; Gyergyek, S. Electrification of Catalytic Ammonia Production and Decomposition Reactions: From Resistance, Induction, and Dielectric Reactor Heating to Electrolysis. ACS Appl. Energy Mater. 2022, 5, 5457–5472. [Google Scholar] [CrossRef]
  68. Jiang, M.; Zhu, D.; Zhao, X. Electrolysis of ammonia for hydrogen production catalyzed by Pt and Pt-Ir deposited on nickel foam. J. Energy Chem. 2014, 23, 1–8. [Google Scholar] [CrossRef]
  69. Dong, B.-X.; Tian, H.; Wu, Y.-C.; Bu, F.-Y.; Liu, W.-L.; Teng, Y.-L.; Diao, G.-W. Improved electrolysis of liquid ammonia for hydrogen generation via ammonium salt electrolyte and Pt/Rh/Ir electrocatalysts. Int. J. Hydrogen Energy 2016, 41, 14507–14518. [Google Scholar] [CrossRef]
  70. Little, D.J.; Edwards, D.O.; Smith, M.R., 3rd; Hamann, T.W. As Precious as Platinum: Iron Nitride for Electrocatalytic Oxidation of Liquid Ammonia. ACS Appl. Mater. Interfaces 2017, 9, 16228–16235. [Google Scholar] [CrossRef]
  71. Wu, M.; Du, J.; Tao, C.; Liu, Z.; Li, Y. A tri-functionalised PtSnOx-based electrocatalyst for hydrogen generation via ammonia decomposition under native pH conditions. J. Colloid Interface Sci. 2019, 542, 451–459. [Google Scholar] [CrossRef]
  72. Zhang, S.; Zhao, Y.; Yan, L.; Jiang, H.; Yang, X.; Wang, Y.; Song, H.; Zhao, X. Electrochemical ammonia oxidation reaction on defect-rich TiO nanofibers: Experimental and theoretical studies. Int. J. Hydrogen Energy 2021, 46, 39208–39215. [Google Scholar] [CrossRef]
  73. Zhang, S.; Yan, L.; Jiang, H.; Yang, L.; Zhao, Y.; Yang, X.; Wang, Y.; Shen, J.; Zhao, X. Facile Fabrication of a Foamed Ag3CuS2 Film as an Efficient Self-Supporting Electrocatalyst for Ammonia Electrolysis Producing Hydrogen. ACS Appl. Mater. Interfaces 2022, 14, 9036–9045. [Google Scholar] [CrossRef] [PubMed]
  74. Dong, Y.; Bai, Z.; Liu, R.; Zhu, T. Preparation of fibrous TiO2 photocatalyst and its optimization towards the decomposition of indoor ammonia under illumination. Catal. Today 2007, 126, 320–327. [Google Scholar] [CrossRef]
  75. Fuku, K.; Kamegawa, T.; Mori, K.; Yamashita, H. Highly dispersed platinum nanoparticles on TiO2 prepared by using the microwave-assisted deposition method: An efficient photocatalyst for the formation of H2 and N2 from aqueous NH3. Chem. Asian J. 2012, 7, 1366–1371. [Google Scholar] [CrossRef]
  76. Kominami, H.; Nishimune, H.; Ohta, Y.; Arakawa, Y.; Inaba, T. Photocatalytic hydrogen formation from ammonia and methyl amine in an aqueous suspension of metal-loaded titanium(IV) oxide particles. Appl. Catal. B Environ. 2012, 111–112, 297–302. [Google Scholar] [CrossRef]
  77. Yuzawa, H.; Mori, T.; Itoh, H.; Yoshida, H. Reaction Mechanism of Ammonia Decomposition to Nitrogen and Hydrogen over Metal Loaded Titanium Oxide Photocatalyst. J. Phys. Chem. C 2012, 116, 4126–4136. [Google Scholar] [CrossRef]
  78. Reli, M.; Edelmannová, M.; Šihor, M.; Praus, P.; Svoboda, L.; Mamulová, K.K.; Otoupalíková, H.; Čapek, L.; Hospodková, A.; Obalová, L.; et al. Photocatalytic H2 generation from aqueous ammonia solution using ZnO photocatalysts prepared by different methods. Int. J. Hydrogen Energy 2015, 40, 8530–8538. [Google Scholar] [CrossRef]
  79. Iwase, A.; Ii, K.; Kudo, A. Decomposition of an aqueous ammonia solution as a photon energy conversion reaction using a Ru-loaded ZnS photocatalyst. Chem. Commun. 2018, 54, 6117–6119. [Google Scholar] [CrossRef]
  80. Jung, S.-C.; Chung, K.-H.; Choi, J.; Park, Y.-K.; Kim, S.-J.; Kim, B.-J.; Lee, H. Photocatalytic hydrogen production using liquid phase plasma from ammonia water over metal ion-doped TiO2 photocatalysts. Catal. Today 2022, 397–399, 165–172. [Google Scholar] [CrossRef]
  81. Wu, Z.; Ambrožová, N.; Eftekhari, E.; Aravindakshan, N.; Wang, W.; Wang, Q.; Zhang, S.; Kočí, K.; Li, Q. Photocatalytic H2 generation from aqueous ammonia solution using TiO2 nanowires-intercalated reduced graphene oxide composite membrane under low power UV light. Emergent Mater. 2019, 2, 303–311. [Google Scholar] [CrossRef]
  82. Zaman, S.F.; Jolaoso, L.A.; Podila, S.; Al-Zahrani, A.A.; Alhamed, Y.A.; Driss, H.; Daous, M.M.; Petrov, L. Ammonia decomposition over citric acid chelated γ-Mo2N and Ni2Mo3N catalysts. Int. J. Hydrogen Energy 2018, 43, 17252–17258. [Google Scholar] [CrossRef]
  83. Utsunomiya, A.; Okemoto, A.; Nishino, Y.; Kitagawa, K.; Kobayashi, H.; Taniya, K.; Ichihashi, Y.; Nishiyama, S. Mechanistic study of reaction mechanism on ammonia photodecomposition over Ni/TiO2 photocatalysts. Appl. Catal. B Environ. 2017, 206, 378–383. [Google Scholar] [CrossRef]
  84. Abdul Razak, S.; Mahadi, A.H.; Thotagamuge, R.; Prasetyoko, D.; Bahruji, H. Photocatalytic Hydrogen Gas Production from NH3 and Alkylamine: Route to Zero Carbon Emission Energy. Catal. Lett. 2022, 153, 1013–1023. [Google Scholar] [CrossRef]
  85. Yuan, Y.; Zhou, L.; Robatjazi, H.; Bao, J.L.; Zhou, J.; Bayles, A.; Yuan, L.; Lou, M.; Lou, M.; Khatiwada, S.; et al. Earth-abundant photocatalyst for H2 generation from NH3 with light-emitting diode illumination. Science 2022, 378, 889–893. [Google Scholar] [CrossRef]
  86. Leung, K.C.; Tan, E.; Li, G.; Ng, B.K.Y.; Ho, P.-L.; Lebedev, K.; Tsang, S.C.E. Metal-loaded zeolites in ammonia decomposition catalysis. Faraday Discuss, 2023; Advance Article. [Google Scholar] [CrossRef]
  87. Kim, H.B.; Park, E.D. Ammonia decomposition over Ru catalysts supported on alumina with different crystalline phases. Catal. Today 2023, 411–412, 113817. [Google Scholar] [CrossRef]
  88. Pinzón, M.; Avilés-García, O.; de la Osa, A.R.; de Lucas-Consuegra, A.; Sánchez, P.; Romero, A. New catalysts based on reduced graphene oxide for hydrogen production from ammonia decomposition. Sustain. Chem. Pharm. 2022, 25, 100615. [Google Scholar] [CrossRef]
  89. Lu, X.; Zhang, J.; Chen, W.K.; Roldan, A. Kinetic and mechanistic analysis of NH3 decomposition on Ru(0001), Ru(111) and Ir(111) surfaces. Nanoscale Adv. 2021, 3, 1624–1632. [Google Scholar] [CrossRef]
  90. Armenise, S.; Cazaña, F.; Monzón, A.; García-Bordejé, E. In situ generation of COx-free H2 by catalytic ammonia decomposition over Ru-Al-monoliths. Fuel 2018, 233, 851–859. [Google Scholar] [CrossRef] [Green Version]
  91. García-García, F.R.; Guerrero-Ruiz, A.; Rodríguez-Ramos, I. Role of B5-Type Sites in Ru Catalysts used for the NH3 Decomposition Reaction. Top. Catal. 2009, 52, 758–764. [Google Scholar] [CrossRef]
  92. Feng, J.; Liu, L.; Ju, X.; Wang, J.; Zhang, X.; He, T.; Chen, P. Highly Dispersed Ruthenium Nanoparticles on Y2O3 as Superior Catalyst for Ammonia Decomposition. ChemCatChem 2021, 13, 1552–1558. [Google Scholar] [CrossRef]
  93. Cha, J.; Lee, T.; Lee, Y.-J.; Jeong, H.; Jo, Y.S.; Kim, Y.; Nam, S.W.; Han, J.; Lee, K.B.; Yoon, C.W.; et al. Highly monodisperse sub-nanometer and nanometer Ru particles confined in alkali-exchanged zeolite Y for ammonia decomposition. Appl. Catal. B Environ. 2021, 283, 119627. [Google Scholar] [CrossRef]
  94. Hu, X.-C.; Fu, X.-P.; Wang, W.-W.; Wang, X.; Wu, K.; Si, R.; Ma, C.; Jia, C.-J.; Yan, C.-H. Ceria-supported ruthenium clusters transforming from isolated single atoms for hydrogen production via decomposition of ammonia. Appl. Catal. B Environ. 2020, 268, 118424. [Google Scholar] [CrossRef]
  95. Jeon, N.; Kim, S.; Tayal, A.; Oh, J.; Yoon, W.; Kim, W.B.; Yun, Y. Y-Doped BaCeO3 Perovskite-Supported Ru Catalysts for COx-Free Hydrogen Production from Ammonia: Effect of Strong Metal–Support Interactions. ACS Sustain. Chem. Eng. 2022, 10, 15564–15573. [Google Scholar] [CrossRef]
  96. Im, Y.; Muroyama, H.; Matsui, T.; Eguchi, K. Investigation on catalytic performance and desorption behaviors of ruthenium catalysts supported on rare-earth oxides for NH3 decomposition. Int. J. Hydrogen Energy 2022, 47, 32543–32551. [Google Scholar] [CrossRef]
  97. Cao, C.-F.; Wu, K.; Zhou, C.; Yao, Y.-H.; Luo, Y.; Chen, C.-Q.; Lin, L.; Jiang, L. Electronic metal-support interaction enhanced ammonia decomposition efficiency of perovskite oxide supported ruthenium. Chem. Eng. Sci. 2022, 257, 117719. [Google Scholar] [CrossRef]
  98. Zhang, X.; Liu, L.; Feng, J.; Ju, X.; Wang, J.; He, T.; Chen, P. Metal–support interaction-modulated catalytic activity of Ru nanoparticles on Sm2O3 for efficient ammonia decomposition. Catal. Sci. Technol. 2021, 11, 2915–2923. [Google Scholar] [CrossRef]
  99. Yamazaki, K.; Matsumoto, M.; Ishikawa, M.; Sato, A. NH3 decomposition over Ru/CeO2-PrOx catalyst under high space velocity conditions for an on-site H2 fueling station. Appl. Catal. B Environ. 2023, 325, 122352. [Google Scholar] [CrossRef]
  100. Obata, K.; Kishishita, K.; Okemoto, A.; Taniya, K.; Ichihashi, Y.; Nishiyama, S. Photocatalytic decomposition of NH3 over TiO2 catalysts doped with Fe. Appl. Catal. B Environ. 2014, 160–161, 200–203. [Google Scholar] [CrossRef]
  101. Bell, T.E.; Torrente-Murciano, L. H2 Production via Ammonia Decomposition Using Non-Noble Metal Catalysts: A Review. Top. Catal. 2016, 59, 1438–1457. [Google Scholar] [CrossRef] [Green Version]
  102. Zhang, J.; Xu, H.; Li, W. Kinetic study of NH3 decomposition over Ni nanoparticles: The role of La promoter, structure sensitivity and compensation effect. Appl. Catal. A Gen. 2005, 296, 257–267. [Google Scholar] [CrossRef]
  103. Hu, Z.-P.; Weng, C.-C.; Yuan, G.-G.; Lv, X.-W.; Yuan, Z.-Y. Ni nanoparticles supported on mica for efficient decomposition of ammonia to COx-free hydrogen. Int. J. Hydrogen Energy 2018, 43, 9663–9676. [Google Scholar] [CrossRef]
  104. Deng, L.; Lin, H.; Liu, X.; Xu, J.; Zhou, Z.; Xu, M. Nickel nanoparticles derived from the direct thermal reduction of Ni-containing Ca–Al layered double hydroxides for hydrogen generation via ammonia decomposition. Int. J. Hydrogen Energy 2021, 46, 38351–38362. [Google Scholar] [CrossRef]
  105. Zhang, H.; Alhamed, Y.A.; Kojima, Y.; Al-Zahrani, A.A.; Miyaoka, H.; Petrov, L.A. Structure and catalytic properties of Ni/MWCNTs and Ni/AC catalysts for hydrogen production via ammonia decomposition. Int. J. Hydrogen Energy 2014, 39, 277–287. [Google Scholar] [CrossRef]
  106. Wu, K.; Cao, C.-F.; Zhou, C.; Luo, Y.; Chen, C.-Q.; Lin, L.; Au, C.; Jiang, L. Engineering of Ce3+-O-Ni structures enriched with oxygen vacancies via Zr doping for effective generation of hydrogen from ammonia. Chem. Eng. Sci. 2021, 245, 116818. [Google Scholar] [CrossRef]
  107. Zhou, C.; Wu, K.; Huang, H.; Cao, C.-F.; Luo, Y.; Chen, C.-Q.; Lin, L.; Au, C.; Jiang, L. Spatial Confinement of Electron-Rich Ni Nanoparticles for Efficient Ammonia Decomposition to Hydrogen Production. ACS Catal. 2021, 11, 10345–10350. [Google Scholar] [CrossRef]
  108. Chen, C.; Fan, X.; Zhou, C.; Lin, L.; Luo, Y.; Au, C.; Cai, G.; Wang, X.; Jiang, L. Hydrogen production from ammonia decomposition over Ni/CeO2 catalyst: Effect of CeO2 morphology. J. Rare Earths 2023, 41, 1014–1021. [Google Scholar] [CrossRef]
  109. Li, Y.; Guan, Q.; Huang, G.; Yuan, D.; Xie, F.; Li, K.; Zhang, Z.; San, X.; Ye, J. Low Temperature Thermal and Solar Heating Carbon-Free Hydrogen Production from Ammonia Using Nickel Single Atom Catalysts. Adv. Energy Mater. 2022, 12, 2202459. [Google Scholar] [CrossRef]
  110. Okura, K.; Okanishi, T.; Muroyama, H.; Matsui, T.; Eguchi, K. Ammonia Decomposition over Nickel Catalysts Supported on Rare-Earth Oxides for the On-Site Generation of Hydrogen. ChemCatChem 2016, 8, 2988–2995. [Google Scholar] [CrossRef]
  111. Okura, K.; Okanishi, T.; Muroyama, H.; Matsui, T.; Eguchi, K. Additive effect of alkaline earth metals on ammonia decomposition reaction over Ni/Y2O3 catalysts. RSC Adv. 2016, 6, 85142–85148. [Google Scholar] [CrossRef]
  112. Gu, Y.; Ma, Y.; Long, Z.; Zhao, S.; Wang, Y.; Zhang, W. One-pot synthesis of supported Ni@Al2O3 catalysts with uniform small-sized Ni for hydrogen generation via ammonia decomposition. Int. J. Hydrogen Energy 2021, 46, 4045–4054. [Google Scholar] [CrossRef]
  113. Okura, K.; Okanishi, T.; Muroyama, H.; Matsui, T.; Eguchi, K. Promotion effect of rare-earth elements on the catalytic decomposition of ammonia over Ni/Al2O3 catalyst. Appl. Catal. A Gen. 2015, 505, 77–85. [Google Scholar] [CrossRef] [Green Version]
  114. Do, Q.C.; Kim, Y.; Le, T.A.; Kim, G.J.; Kim, J.-R.; Kim, T.-W.; Lee, Y.-J.; Chae, H.-J. Facile one-pot synthesis of Ni-based catalysts by cation-anion double hydrolysis method as highly active Ru-free catalysts for green H2 production via NH3 decomposition. Appl. Catal. B Environ. 2022, 307, 121167. [Google Scholar] [CrossRef]
  115. Yu, Y.; Gan, Y.-M.; Huang, C.; Lu, Z.-H.; Wang, X.; Zhang, R.; Feng, G. Ni/La2O3 and Ni/MgO–La2O3 catalysts for the decomposition of NH3 into hydrogen. Int. J. Hydrogen Energy 2020, 45, 16528–16539. [Google Scholar] [CrossRef]
  116. Varisli, D.; Kaykac, N.G. COx free hydrogen production over cobalt incorporated silicate structured mesoporous catalysts. Appl. Catal. B Environ. 2012, 127, 389–398. [Google Scholar] [CrossRef]
  117. Lendzion-Bielun, Z.; Narkiewicz, U.; Arabczyk, W. Cobalt-based Catalysts for Ammonia Decomposition. Materials 2013, 6, 2400–2409. [Google Scholar] [CrossRef] [Green Version]
  118. Pinzón, M.; Romero, A.; de Lucas-Consuegra, A.; de la Osa, A.R.; Sánchez, P. COx-free hydrogen production from ammonia at low temperature using Co/SiC catalyst: Effect of promoter. Catal. Today 2022, 390–391, 34–47. [Google Scholar] [CrossRef]
  119. Zhang, H.; Alhamed, Y.A.; Chu, W.; Ye, Z.; AlZahrani, A.; Petrov, L. Controlling Co-support interaction in Co/MWCNTs catalysts and catalytic performance for hydrogen production via NH3 decomposition. Appl. Catal. A Gen. 2013, 464–465, 156–164. [Google Scholar] [CrossRef]
  120. Duan, X.; Ji, J.; Qian, G.; Fan, C.; Zhu, Y.; Zhou, X.; Chen, D.; Yuan, W. Ammonia decomposition on Fe(1 1 0), Co(1 1 1) and Ni(1 1 1) surfaces: A density functional theory study. J. Mol. Catal. A Chem. 2012, 357, 81–86. [Google Scholar] [CrossRef]
  121. Czekajło, Ł.; Lendzion-Bieluń, Z. Effect of preparation conditions and promoters on the structure and activity of the ammonia decomposition reaction catalyst based on nanocrystalline cobalt. Chem. Eng. J. 2016, 289, 254–260. [Google Scholar] [CrossRef]
  122. Varisli, D.; Kaykac, N.G. Hydrogen from ammonia over cobalt incorporated silicate structured catalysts prepared using different cobalt salts. Int. J. Hydrogen Energy 2016, 41, 5955–5968. [Google Scholar] [CrossRef]
  123. Huang, C.; Yu, Y.; Tang, X.; Liu, Z.; Zhang, J.; Ye, C.; Ye, Y.; Zhang, R. Hydrogen generation by ammonia decomposition over Co/CeO2 catalyst: Influence of support morphologies. Appl. Surf. Sci. 2020, 532, 147335. [Google Scholar] [CrossRef]
  124. Yu, P.; Wu, H.; Guo, J.; Wang, P.; Chang, F.; Gao, W.; Zhang, W.; Liu, L.; Chen, P. Effect of BaNH, CaNH, Mg3N2 on the activity of Co in NH3 decomposition catalysis. J. Energy Chem. 2020, 46, 16–21. [Google Scholar] [CrossRef]
  125. Li, G.; Zhang, H.; Yu, X.; Lei, Z.; Yin, F.; He, X. Highly efficient Co/NC catalyst derived from ZIF-67 for hydrogen generation through ammonia decomposition. Int. J. Hydrogen Energy 2022, 47, 12882–12892. [Google Scholar] [CrossRef]
  126. Lanzani, G.; Laasonen, K. NH3 adsorption and dissociation on a nanosized iron cluster. Int. J. Hydrogen Energy 2010, 35, 6571–6577. [Google Scholar] [CrossRef]
  127. Othman, N.E.F.; Salleh, H.M.; Purwanto, H. Utilization of Low-grade Iron Ore in Ammonia Decomposition. Procedia Chem. 2016, 19, 119–124. [Google Scholar] [CrossRef] [Green Version]
  128. Lucentini, I.; García Colli, G.; Luzi, C.; Serrano, I.; Soler, L.; Divins, N.J.; Martínez, O.M.; Llorca, J. Modelling and simulation of catalytic ammonia decomposition over Ni-Ru deposited on 3D-printed CeO2. Chem. Eng. J. 2022, 427, 131756. [Google Scholar] [CrossRef]
  129. Lucentini, I.; García Colli, G.; Luzi, C.D.; Serrano, I.; Martínez, O.M.; Llorca, J. Catalytic ammonia decomposition over Ni-Ru supported on CeO2 for hydrogen production: Effect of metal loading and kinetic analysis. Appl. Catal. B Environ. 2021, 286, 119896. [Google Scholar] [CrossRef]
  130. Lucentini, I.; Casanovas, A.; Llorca, J. Catalytic ammonia decomposition for hydrogen production on Ni, Ru and Ni-Ru supported on CeO2. Int. J. Hydrogen Energy 2019, 44, 12693–12707. [Google Scholar] [CrossRef]
  131. Li, H.; Guo, L.; Qu, J.; Fang, X.; Fu, Y.; Duan, J.; Wang, W.; Li, C. Co-Ni supported yttrium oxide material as a catalyst for ammonia decomposition to COx-free hydrogen. Int. J. Hydrogen Energy 2023, 48, 8985–8996. [Google Scholar] [CrossRef]
  132. He, H.; Jiang, H.; Yang, F.; Liu, J.; Zhang, W.; Jin, M.; Li, Z. Bimetallic NixCo10−x/CeO2 as highly active catalysts to enhance mid-temperature ammonia decomposition: Kinetics and synergies. Int. J. Hydrogen Energy 2023, 48, 5030–5041. [Google Scholar] [CrossRef]
  133. Silva, H.; Nielsen, M.G.; Fiordaliso, E.M.; Damsgaard, C.D.; Gundlach, C.; Kasama, T.; Chorkendorff, I.B.; Chakraborty, D. Synthesis and characterization of Fe–Ni/ɣ-Al2O3 egg-shell catalyst for H2 generation by ammonia decomposition. Appl. Catal. A Gen. 2015, 505, 548–556. [Google Scholar] [CrossRef] [Green Version]
  134. Hansgen, D.A.; Vlachos, D.G.; Chen, J.G. Using first principles to predict bimetallic catalysts for the ammonia decomposition reaction. Nat. Chem. 2010, 2, 484–489. [Google Scholar] [CrossRef] [PubMed]
  135. Jiang, K.; Li, K.; Liu, Y.-Q.; Lin, S.; Wang, Z.; Wang, D.; Ye, Y. Nickel-cobalt nitride nanoneedle supported on nickel foam as an efficient electrocatalyst for hydrogen generation from ammonia electrolysis. Electrochim. Acta 2022, 403, 139700. [Google Scholar] [CrossRef]
  136. Huo, L.; Han, X.; Zhang, L.; Liu, B.; Gao, R.; Cao, B.; Wang, W.-W.; Jia, C.-J.; Liu, K.; Liu, J.; et al. Spatial confinement and electron transfer moderating Mo-N bond strength for superior ammonia decomposition catalysis. Appl. Catal. B Environ. 2021, 294, 120254. [Google Scholar] [CrossRef]
  137. Xu, J.; Yan, H.; Jin, Z.; Jia, C.J. Facile Synthesis of Stable MO2N Nanobelts with High Catalytic Activity for Ammonia Decomposition. Chin. J. Chem. 2019, 37, 364–372. [Google Scholar] [CrossRef]
  138. Podila, S.; Zaman, S.F.; Driss, H.; Alhamed, Y.A.; Al-Zahrani, A.A.; Petrov, L.A. Hydrogen production by ammonia decomposition using high surface area Mo2N and Co3Mo3N catalysts. Catal. Sci. Technol. 2016, 6, 1496–1506. [Google Scholar] [CrossRef]
  139. Guo, J.; Wang, P.; Wu, G.; Wu, A.; Hu, D.; Xiong, Z.; Wang, J.; Yu, P.; Chang, F.; Chen, Z.; et al. Lithium imide synergy with 3d transition-metal nitrides leading to unprecedented catalytic activities for ammonia decomposition. Angew. Chem. Int. Ed. Engl. 2015, 54, 2950–2954. [Google Scholar] [CrossRef]
  140. Jolaoso, L.A.; Zaman, S.F.; Podila, S.; Driss, H.; Al-Zahrani, A.A.; Daous, M.A.; Petrov, L. Ammonia decomposition over citric acid induced γ-Mo2N and Co3Mo3N catalysts. Int. J. Hydrogen Energy 2018, 43, 4839–4844. [Google Scholar] [CrossRef]
  141. Srifa, A.; Okura, K.; Okanishi, T.; Muroyama, H.; Matsui, T.; Eguchi, K. COx-free hydrogen production via ammonia decomposition over molybdenum nitride-based catalysts. Catal. Sci. Technol. 2016, 6, 7495–7504. [Google Scholar] [CrossRef]
  142. Srifa, A.; Okura, K.; Okanishi, T.; Muroyama, H.; Matsui, T.; Eguchi, K. Hydrogen production by ammonia decomposition over Cs-modified Co3Mo3N catalysts. Appl. Catal. B Environ. 2017, 218, 1–8. [Google Scholar] [CrossRef]
  143. Kraupner, A.; Markus, A.; Palkovits, R.; Schlicht, K.; Giordano, C. Mesoporous Fe3C sponges as magnetic supports and as heterogeneous catalyst. J. Mater. Chem. 2010, 20, 6019–6022. [Google Scholar] [CrossRef]
  144. Santhana Krishnan, P.; Neelaveni, M.; Tamizhdurai, P.; Mythily, M.; Krishna Mohan, S.; Mangesh, V.L.; Shanthi, K. COx-free hydrogen generation via decomposition of ammonia over al, Ti and Zr−Laponite supported MoS2 catalysts. Int. J. Hydrogen Energy 2020, 45, 8568–8583. [Google Scholar] [CrossRef]
Figure 1. The status of hydrogen production from ammonia decomposition in ammonia economy.
Figure 1. The status of hydrogen production from ammonia decomposition in ammonia economy.
Molecules 28 05245 g001
Figure 2. Advantages and disadvantages of different NH3 decomposition methods.
Figure 2. Advantages and disadvantages of different NH3 decomposition methods.
Molecules 28 05245 g002
Figure 3. A comparison of various technical methods for ammonia decomposition. (As the evaluation of electrocatalytic NH3 decomposition for H2 production is mainly based on current density, no comparison is made here with other technical methods) [10,19,20,21,22,23,24,25].
Figure 3. A comparison of various technical methods for ammonia decomposition. (As the evaluation of electrocatalytic NH3 decomposition for H2 production is mainly based on current density, no comparison is made here with other technical methods) [10,19,20,21,22,23,24,25].
Molecules 28 05245 g003
Figure 4. Energy profiles for ammonia decomposition without catalyst and with catalyst.
Figure 4. Energy profiles for ammonia decomposition without catalyst and with catalyst.
Molecules 28 05245 g004
Scheme 1. Elementary kinetic steps for the ammonia decomposition reaction; (*) refers to an unoccupied adsorption site on the catalyst [29].
Scheme 1. Elementary kinetic steps for the ammonia decomposition reaction; (*) refers to an unoccupied adsorption site on the catalyst [29].
Molecules 28 05245 sch001
Figure 5. Schematic of the catalytic cycle developed for oxidative decomposition of ammonia (reproduced with permission from [50], Copyright Science, 2017).
Figure 5. Schematic of the catalytic cycle developed for oxidative decomposition of ammonia (reproduced with permission from [50], Copyright Science, 2017).
Molecules 28 05245 g005
Figure 6. Schematic of the primary chain reactions for (a) PCMech−1 and (b) PCMech−2 for the NH3/N2 combination (1/99 mol%) with Tg = 500 K, E/N = 150 Td, and Pdis = 10 W. (reproduced with permission from [62], Copyright American Chemical Society, 2023).
Figure 6. Schematic of the primary chain reactions for (a) PCMech−1 and (b) PCMech−2 for the NH3/N2 combination (1/99 mol%) with Tg = 500 K, E/N = 150 Td, and Pdis = 10 W. (reproduced with permission from [62], Copyright American Chemical Society, 2023).
Molecules 28 05245 g006
Figure 7. Pt and Fe catalysts decompose ammonia (reproduced with permission from [70], Copyright American Chemical Society, 2017).
Figure 7. Pt and Fe catalysts decompose ammonia (reproduced with permission from [70], Copyright American Chemical Society, 2017).
Molecules 28 05245 g007
Figure 8. Illustration of the process of photocatalytic ammonia decomposition (reproduced with permission from [15], Copyright Royal Soc Chemistry, 2020).
Figure 8. Illustration of the process of photocatalytic ammonia decomposition (reproduced with permission from [15], Copyright Royal Soc Chemistry, 2020).
Molecules 28 05245 g008
Figure 9. Reaction mechanism of NH3 decomposition to N2 and H2 over TiO2 photocatalyst (reproduced with permission from [83], Copyright Elsevier, 2017).
Figure 9. Reaction mechanism of NH3 decomposition to N2 and H2 over TiO2 photocatalyst (reproduced with permission from [83], Copyright Elsevier, 2017).
Molecules 28 05245 g009
Figure 10. Schematic comparison of the energy diagram for photocatalysis (excited states) and thermocatalysis (ground state) on Cu-Fe- and Cu-Ru-ARs. Red arrows refer to electronic excitation on Fe-N, and blue to Ru-N. Only the two possible RDSs are included for simplicity; migrations of nitrogen species are not considered for the same reason (reproduced with permission from [85], Copyright Science, 2022).
Figure 10. Schematic comparison of the energy diagram for photocatalysis (excited states) and thermocatalysis (ground state) on Cu-Fe- and Cu-Ru-ARs. Red arrows refer to electronic excitation on Fe-N, and blue to Ru-N. Only the two possible RDSs are included for simplicity; migrations of nitrogen species are not considered for the same reason (reproduced with permission from [85], Copyright Science, 2022).
Molecules 28 05245 g010
Figure 11. Scheme of associative adsorption of the two N atoms to form an N2 molecule in a B5-type site: (1) N atoms diffusion to the B5-type site, (2) association of N atoms to form an N2 molecule, and (3) desorption of the N2 molecule (‘three stages of the growth of an Ru particle’ reproduced with permission from [91], Copyright Springer, 2009).
Figure 11. Scheme of associative adsorption of the two N atoms to form an N2 molecule in a B5-type site: (1) N atoms diffusion to the B5-type site, (2) association of N atoms to form an N2 molecule, and (3) desorption of the N2 molecule (‘three stages of the growth of an Ru particle’ reproduced with permission from [91], Copyright Springer, 2009).
Molecules 28 05245 g011
Figure 12. (a) HAADF-STEM images, (b) particle size distribution, and (c) EDS elemental mapping images of Ru/M-Y catalysts (M = H, Na, K, and Rb) (reproduced with permission from [93], Copyright Elsevier, 2021).
Figure 12. (a) HAADF-STEM images, (b) particle size distribution, and (c) EDS elemental mapping images of Ru/M-Y catalysts (M = H, Na, K, and Rb) (reproduced with permission from [93], Copyright Elsevier, 2021).
Molecules 28 05245 g012
Figure 13. Schematic diagram of NH3 decomposition of Ru/Y-dosed BaCeO3 catalyst (reproduced with permission from [95], Copyright American Chemical Society, 2022).
Figure 13. Schematic diagram of NH3 decomposition of Ru/Y-dosed BaCeO3 catalyst (reproduced with permission from [95], Copyright American Chemical Society, 2022).
Molecules 28 05245 g013
Figure 14. Schematic diagram of Pt/TiO2 photocatalytic NH3 decomposition performance and reaction mechanism (reproduced with permission from [76], Copyright Elsevier, 2012).
Figure 14. Schematic diagram of Pt/TiO2 photocatalytic NH3 decomposition performance and reaction mechanism (reproduced with permission from [76], Copyright Elsevier, 2012).
Molecules 28 05245 g014
Figure 15. Schematic illustration of the synthesis of Ni nanoparticles derived from the thermal reduction of Ni-containing Ca-Al LDH (reproduced with permission from [104], Copyright Elsevier, 2021).
Figure 15. Schematic illustration of the synthesis of Ni nanoparticles derived from the thermal reduction of Ni-containing Ca-Al LDH (reproduced with permission from [104], Copyright Elsevier, 2021).
Molecules 28 05245 g015
Figure 16. Schematic illustration of the synthesis of Co/NC-X (reproduced with permission from [125], Copyright Elsevier, 2022).
Figure 16. Schematic illustration of the synthesis of Co/NC-X (reproduced with permission from [125], Copyright Elsevier, 2022).
Molecules 28 05245 g016
Figure 17. Schematic diagram of NH3 decomposition ability of Fe/Al2O3 catalysts driven by plasma and thermo as the size of Fe3O4 nanoparticles increases (reproduced with permission from [57], Copyright American Chemical Society, 2022).
Figure 17. Schematic diagram of NH3 decomposition ability of Fe/Al2O3 catalysts driven by plasma and thermo as the size of Fe3O4 nanoparticles increases (reproduced with permission from [57], Copyright American Chemical Society, 2022).
Molecules 28 05245 g017
Figure 18. Schematic illustration for the decorated CoNi alloy on the oxide support of MgO–CeO2–SrO (reproduced with permission from [49], Copyright Royal Society of Chemistry, 2022).
Figure 18. Schematic illustration for the decorated CoNi alloy on the oxide support of MgO–CeO2–SrO (reproduced with permission from [49], Copyright Royal Society of Chemistry, 2022).
Molecules 28 05245 g018
Figure 19. HEA catalysts breaking the miscibility limitation of conventional binary alloys (reproduced with permission from [44], Copyright Nature, 2019).
Figure 19. HEA catalysts breaking the miscibility limitation of conventional binary alloys (reproduced with permission from [44], Copyright Nature, 2019).
Molecules 28 05245 g019
Figure 20. Schematic diagram of Mo2N catalysts with the addition of Co, Ni, and Fe for NH3 decomposition (reproduced with permission from [141], Copyright Royal Society of Chemistry, 2016).
Figure 20. Schematic diagram of Mo2N catalysts with the addition of Co, Ni, and Fe for NH3 decomposition (reproduced with permission from [141], Copyright Royal Society of Chemistry, 2016).
Molecules 28 05245 g020
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

Huang, X.; Lei, K.; Mi, Y.; Fang, W.; Li, X. Recent Progress on Hydrogen Production from Ammonia Decomposition: Technical Roadmap and Catalytic Mechanism. Molecules 2023, 28, 5245. https://doi.org/10.3390/molecules28135245

AMA Style

Huang X, Lei K, Mi Y, Fang W, Li X. Recent Progress on Hydrogen Production from Ammonia Decomposition: Technical Roadmap and Catalytic Mechanism. Molecules. 2023; 28(13):5245. https://doi.org/10.3390/molecules28135245

Chicago/Turabian Style

Huang, Xiangyong, Ke Lei, Yan Mi, Wenjian Fang, and Xiaochuan Li. 2023. "Recent Progress on Hydrogen Production from Ammonia Decomposition: Technical Roadmap and Catalytic Mechanism" Molecules 28, no. 13: 5245. https://doi.org/10.3390/molecules28135245

Article Metrics

Back to TopTop