Next Article in Journal
The Effect of Ageing on Phase Transformations and Mechanical Behaviour in Ni-Rich NiTi Alloys
Previous Article in Journal
Reinforcement Effects on Tensile Behavior of Ultra-High-Performance Concrete (UHPC) with Low Steel Fiber Volume Fractions
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Progress Made in Non-Metallic-Doped Materials for Electrocatalytic Reduction in Ammonia Production

by
Gerald D. S. Quoie Jr
1,2,
Mingshuo Jiao
1,2,
Krisztina Lászlód
3 and
Ying Wang
1,2,*
1
State Key Laboratory of Pollution Control and Resources Reuse, College of Environmental Science and Engineering, Tongji University, Shanghai 200092, China
2
Shanghai Institute of Pollution Control and Ecological Security, Shanghai 200092, China
3
Department of Physical Chemistry and Materials Science, Budapest University of Technology and Economics, H-1521 Budapest, Hungary
*
Author to whom correspondence should be addressed.
Materials 2024, 17(10), 2419; https://doi.org/10.3390/ma17102419
Submission received: 8 April 2024 / Revised: 2 May 2024 / Accepted: 15 May 2024 / Published: 17 May 2024
(This article belongs to the Special Issue Featured Reviews in Catalytic Materials)

Abstract

:
The electrocatalytic production of ammonia has garnered considerable interest as a potentially sustainable technology for ammonia synthesis. Recently, non-metallic-doped materials have emerged as promising electrochemical catalysts for this purpose. This paper presents a comprehensive review of the latest research on non-metallic-doped materials for electrocatalytic ammonia production. Researchers have engineered a variety of materials, doped with non-metals such as nitrogen (N), boron (B), phosphorus (P), and sulfur (S), into different forms and structures to enhance their electrocatalytic activity and selectivity. A comparison among different non-metallic dopants reveals their distinct effects on the electrocatalytic performance for ammonia production. For instance, N-doping has shown enhanced activity owing to the introduction of nitrogen vacancies (NVs) and improved charge transfer kinetics. B-doping has demonstrated improved selectivity and stability, which is attributed to the formation of active sites and the suppression of competing reactions. P-doping has exhibited increased ammonia generation rates and Faradaic efficiencies, likely due to the modification of the electronic structure and surface properties. S-doping has shown potential for enhancing electrocatalytic performance, although further investigations are needed to elucidate the underlying mechanisms. These comparisons provide valuable insights for researchers to conduct in-depth studies focusing on specific non-metallic dopants, exploring their unique properties, and optimizing their performance for electrocatalytic ammonia production. However, we consider it a priority to provide insight into the recent progress made in non-metal-doped materials and their potential for enabling long-term and efficient electrochemical ammonia production. Additionally, this paper discusses the synthetic procedures used to produce non-metal-doped materials and highlights the advantages and disadvantages of each method. It also provides an in-depth analysis of the electrochemical performance of these materials, including their Faradaic efficiencies, ammonia yield rate, and selectivity. It examines the challenges and prospects of developing non-metallic-doped materials for electrocatalytic ammonia production and suggests future research directions.

1. Introduction

Ammonia (NH3) is a highly significant chemical compound in contemporary life and is extensively used in agriculture, pharmaceuticals, textiles, and various other industries [1,2,3,4,5]. Due to its high energy density (4.32 kW h kg−1), high hydrogen content, and zero carbon emissions, it is an excellent energy carrier conducive to carbon neutrality [6,7]. The Haber–Bosch ammonia synthesis process, developed in 1905, remains the primary method for ammonia production. This process requires harsh conditions, such as reaching temperatures of 400–500 °C and pressures of 150–300 atm [8,9]. It is associated with significant energy consumption, accounting for 1–2% of global energy usage, and contributes to over 1% of the total CO2 emission [10,11]. The exploration of non-metallic-doped materials for electrocatalytic reduction in ammonia production has witnessed significant advancements in recent years, reflecting the quest for sustainable and efficient nitrogen fixation processes [12,13,14]. Early studies, such as those by Humphrey Davy [15], Martín et al. [16], Fitchter and Suter [17], Tamelen et al. [18], Ren et al. [19], Tsumeto et al. [20], Rod et al. [21], Kordali et al. [22], Köleli et al. [23], Skulason et al. [24], Azofra et al. [25], Abghoui and Skúlason [26], Zhao et al. [27], and Du et al. [28], laid the groundwork by demonstrating the electrocatalytic capabilities of nitrogen reduction reactions (NRRs) for NH3 synthesis.
Therefore, electrochemistry is an effective method for studying redox processes. The electrochemical hydrogenation reduction of N2 to synthesize NH3 is feasible at room temperature, circumventing the need for hydrogen with high energy density. Furthermore, the electrochemical reduction approach is capable of efficiently treating nitrate-laden wastewater, facilitating an eight-electron transfer to yield NH3 [29,30]. Owing to its reduced energy demand and enhanced Faradaic efficiency, this methodology has received increasing scholarly attention [31,32,33]. The bond enthalpy of N=O in nitrate is substantially lower (204 kJ/mol) than that of the N≡N triple bond (941 kJ/mol), indicating that ammonia production via nitrate reduction is more energy-efficient than traditional ammonia synthesis via nitrogen hydrogenation [34,35,36].
Additionally, this technique provides a novel remedial strategy for aquatic nitrate pollution, offering concurrent advantages for environmental conservation and economic progress [37]. Under high overpotential conditions, electrocatalytic reactions are concurrent with a pronounced hydrogen evolution reaction (HER), which affects the Faradaic efficiency (FE) [38]. Moreover, given the diverse oxidation states of nitrogen, the reduction pathway may produce multiple side products, thereby attenuating the yield [39]. Consequently, the utilization of high-performance catalysts to amplify reaction kinetics and selectivity is of paramount importance [40,41]. Non-metallic doping is a viable approach to bolster the activity of catalysts.
Despite significant advancements in the field of electrocatalytic reduction for ammonia production, there is still a need for a comprehensive understanding of how non-metallic-doped materials enhance catalytic efficiency. Although there have been many investigations into the use of non-metallic catalysts, there is a relative lack of exploration of the potential advantages and mechanisms of non-metallic doping. This study gap emphasizes the need for additional investigations to clarify the fundamental principles governing the electrocatalytic activity of non-metallic-doped materials and their influence on the efficiency of NH3 production. Therefore, we present a thorough examination of advancements in the utilization of non-metallic-doped materials for electrocatalytic reduction in the production of NH3. Through our investigation of this research void, our objective was to provide insight into the fundamental mechanisms, improvements in performance, and potential future development of non-metallic-doped catalysts for the advancement of sustainable NH3 synthesis methods.

2. Construction Principle of Non-Metallic-Element-Doped Catalyst

Various concepts guide the construction of non-metallic-doped catalysts, with the aim of enhancing the electrocatalytic activity by modifying its structure and characteristics. Figure 1 depicts the doping of non-metallic elements like boron (B), nitrogen (N), phosphorus (P), and sulfur (S) into metal-based or carbon-based electrodes.
Multiple proton reduction steps characterize the process of ammonia production through reduction. In the pathway of nitrate reduction, the rate-determining step is the reduction of nitrate to nitrite [42]. In the case of nitrogen reduction, there are two main pathways: one involves breaking the nitrogen-to-nitrogen bond and, in the final step, concurrently producing ammonia as the end product (Figure 2a), while the other consists of breaking the nitrogen-to-nitrogen bond in the first step (Figure 2b) [43].

2.1. Defect Engineering

Defect engineering has been widely used to improve electrocatalytic performance [44]. This is fundamental for the development of non-metallic catalysts for NH3 synthesis. Non-metallic materials are incorporated to alter the structure and surface characteristics of the catalyst and introduce imperfections [45]. Defect engineering pertains to the deliberate introduction of imperfections or aberrations into the catalyst structure. These flaws may contain edges, voids, or dislocations that can act as catalytic reaction sites [46]. Adding these flaws increases the reactivity and surface area, improving the electrocatalytic effectiveness of reactions such as the reduction of nitrate (NO3) to NH3. The most commonly employed non-metallic elements for catalyst doping in NH3 production include B, N, P, and S [47]. These elements are preferred because of their ability to adjust the catalyst surface characteristics and electronic structure, thereby enhancing the catalytic activity for NH3 production. Despite the challenges and limitations related to defect control and stability, defective non-metallic catalysts hold significant promise for sustainable and efficient catalytic applications.
Huang et al. made gC3N4, a porous polymeric carbon nitride with a controllable number of nitrogen vacancies (NV), by calcining the material in two easy steps. This study found that introducing NV can significantly enhance the electrocatalytic reduction performance of gC3N4 for nitrate, achieving a nitrate removal rate of 89.1% and high Faraday efficiency while also exhibiting high NH3 selectivity [48]. Researchers have constructed reduced graphene oxides via defect engineering, where carbon vacancies facilitate strong binding with N2 while suppressing H binding. This approach has achieved high Faradaic efficiencies (up to 22.0%) across a wide pH range [49].

2.2. Post-Synthesis Doping Strategies

Post-synthetic doping approaches involve direct interactions between a carbon material that has already been synthesized and sources containing nitrogen, such as urea, NH3, dicyandiamide (C2H4N4), dimethylformamide (C3H7NO), and melamine (C3H6N6) For example, the incorporation of nitrogen into the carbon network can be facilitated by high temperatures, prolonging the reaction periods, adding transition metals, or preparing a carbon surface [50]. Although these methods have some disadvantages, they are widely used because of their simplicity in producing N-doped carbons. The nitrogen content and arrangement were affected by the temperature and time of the process. [51]. While some studies indicate that NH3, as a precursor, shows potential at 500–600 °C, other studies propose temperatures exceeding 900 °C for peak nitrogen content with NH3 [52].
Post-synthetic doping methods involve a range of techniques, such as microwave-assisted carbonization, which leads to the production of N-doped reduced graphene oxide (N-RGO) containing pyridinic, pyrrolic, and graphitic nitrogen groups [53,54]. Urea is another nitrogen-containing precursor that is known for its low pyrolysis temperature and low cost. For instance, N-doped carbon with 4% nitrogen content was later treated with urea, increasing its nitrogen content to 11.08% and enhancing its adsorption characteristics [55]. Therefore, the modification of carbon compounds by the deposition and subsequent carbonization of N-containing precursors is a noteworthy approach.
Several approaches have demonstrated efficacy. Chen et al. [56] employed hydrazine (N2H4) and NH3 to decrease the size of graphene oxide (GO) at temperatures lower than 200 °C, resulting in the production of nitrogen-doped graphene sheets (NG) with a nitrogen content of 5%. The polymerization and carbonization of aniline produce N-doped carbon fibers [57]. Innovative methods, such as the tour technique employing aryldiazonium salts, enable the tuning of N-functionalities in carbon nanotubes (CNTs) or graphene [58]. Electrochemical functionalization with aminobenzene acids yields N-modified CNTs with various nitrogen and oxygen groups [59].
Nitric acid treatment, either as a pretreatment or for chemical reactions with nitrogen-containing compounds, has been effective in diverse synthesis approaches [60,61]. The utilization of nitric acid and amine functionalization has been demonstrated to be successful in achieving the solubilization of single-walled CNTs. This technique involves the hydrolysis of N-doped carbon at elevated temperatures [62]. Lu et al. [63] demonstrated the adaptability of nitric acid as a pretreatment for various N-doped carbon compounds for the oxygen reduction reaction (ORR). They specifically employed nitric acid and melamine to create N-doped-ordered mesoporous carbon.
Table 1 summarizes several post-synthesis methods, showing the enhanced characteristics of doped materials employing non-metals, such as N, B, carbon (C), and silicon (Si) from various sources across different materials.
Chemical vapor deposition (CVD) has been employed to create co-doped microporous carbons with Y-zeolite infiltration, revealing predominant pyridinic, pyrrolic, and graphitic nitrogen [70]. CVD with C2H4/NH3 combinations also successfully produced nitrogen-doped carbon nanofibers (N-CNFs) with enhanced pyridinic group content. Carbonization uses a variety of nitrogen-containing sources, including conducting polymers, ionic liquids, and biomass, to produce N-doped carbons with various topologies and nitrogen concentrations [71,72].
For instance, Gavrilov et al. [73] produced N-doped carbon materials by carbonizing polyaniline (PANI) at a temperature of 800 °C. This process resulted in a nitrogen content of approximately 9%, mostly comprising pyridinic and graphitic groups. The N-doped carbon compounds obtained from PANI exhibited varied proportions of graphitic and pyridinic groups depending on the experimental conditions. This highlights the impact of factors such as carbonization temperature, monomer type, and polymerization length [74]. Imidazolium and pyridinium ionic liquids have also been used to produce mesoporous N-doped carbon with a nitrogen content exceeding 15% by weight [75]. Moreover, the distinctive characteristics of Metal–Organic Frameworks (MOFs) containing N-organic ligands have prompted investigations into their potential as precursors for N-doped carbon materials [76].
Furthermore, in situ techniques involving biomass, particularly hydrothermal carbonization (HTC), offer an eco-friendly route. Nevertheless, hydrochars typically require supplementary procedures, such as heat treatment or chemical activation, to improve their surface area, which may result in a reduction in the amount of nitrogen present [71]. The traditional decomposition of biomass without chemical activation effectively generates N-doped carbon nanosheets with a high surface area from chitosan [77]. A novel method utilizing KCl/ZnCl2 to combine biomass and melamine resulted in the production of mesoporous N-doped carbons with a significant nitrogen concentration of 11.9% [78]. Despite the obstacles in regulating nitrogen speciation and content, the synthesis of N-doped carbon materials by diverse in situ processes has been significant.

2.3. In Situ Doping Methods

As seen in Table 2, in the context of in situ doping synthesis, simultaneous carbon material synthesis and nitrogen doping are achieved, with CVD being a widely utilized technique, particularly for graphene, CNTs, CNFs, and microporous carbon production [79]. Several researchers have devised novel techniques for manufacturing co-doped microporous carbons with large surface areas. The process involves impregnating the Y zeolite with 1-ethyl-3-methylimidazolium tetracyanoborate, an ionic liquid containing nitrogen, and chemical vapor deposition (CVD). This method results in the formation of microporous carbon co-doped with nitrogen in the form of pyridinic, pyrrolic, and graphitic nitrogen [70].
A wide range of N-containing sources, including conducting polymers, ionic liquids, and biomass, have been utilized for carbonization to produce nitrogen-doped carbons with different structures and N levels. Lu et al. [82] used electrospinning to create hollow-structured N-doped carbon nanofibers (N-CNFs) with significant graphitic characteristics by combining PAN and ZIF-8. These nanofibers achieved a nitrogen content of 8% after carbonization at 900 °C. The N-containing groups in the final N-CNFs primarily consisted of pyridinic and pyrrolic groups, with the latter changing to a graphitic form at elevated carbonization temperatures [77].
Hydrothermal carbonization (HTC) is an eco-friendly and cost-efficient method for producing N-doped carbon compounds [83]. Although in situ techniques involving biomass, mainly through hydrothermal carbonization (HTC), offer an eco-friendly route, the resulting hydrochars often require additional steps for heat treatment or chemical activation to enhance the surface area and deplete some nitrogen content [71]. Chitosan was successfully converted into N-doped carbon nanosheets with large surface areas using traditional biomass pyrolysis without chemical activation [84].

3. Regulation of Nitrogen Reduction by Non-Metallic-Doped Catalysts

The foundational element in crafting catalysts doped with non-metallic materials, such as N-CNTs, for NH3 synthesis is the integration of non-metallic elements, such as B and N, to create active sites. For instance, incorporating B-N prevents undesired HER while providing ample active sites on carbon nanosheets. Utilizing non-metallic materials, mainly two-dimensional nano-non-metallic materials with doped heteroatoms, has significantly advanced eco-friendly and sustainable synthesis [85].

3.1. Boron-Doped (B-Doped) Catalysts

B-doped catalysts, particularly B-N-doped materials, exhibit notable potential as active sites for the electrocatalytic and photocatalytic reduction of N2 to NH3 [86,87]. These materials offer sufficient active sites for NH3 generation while effectively suppressing competing hydrogen evolution processes, making them advantageous for targeted electrocatalytic nitrogen reduction [88]. However, NH3 synthesis using B-doped catalysts involves the electrocatalytic reduction of NO3 to NH3. The B-doped catalyst facilitates redox reactions, converting nitrate to nitrite by transferring electrons from the catalyst to nitrate [89,90]. The adsorption of nitrite on the catalyst surface leads to the adsorption of N species, which are highly reactive and easily reduced to NH3 [91,92].
Further reduction of the adsorbed N species on the B-doped catalyst produced ammonia. This process involves the gain of electrons from nitrogen species, which interact with hydroxyl groups on the catalyst surface to form NH3. The essential role of the B-doped catalysts lies in facilitating these redox processes, lowering the energy barrier for nitrogen reduction, and enhancing the overall reaction rate. The synthesis concluded that the generated ammonia was desorbed from the catalyst surface. The unique structure and electrical characteristics of the B-doped catalysts contribute to efficient NH3 desorption, yielding significant amounts of NH3 [93].
Ongoing research has explored the impact of boron concentration on B-doped catalysts in NH3 synthesis, revealing significant effects on electrocatalytic and photocatalytic nitrogen reduction [47]. The boron content of B-doped carbon nanosheets influences the efficiency and selectivity of the electrocatalytic nitrogen reduction to ammonia [94]. In electrocatalytic ammonia generation, graphene with a boron doping concentration of 6.2% demonstrated superior faradaic efficiencies and NH3 yields [95].
The dispersion of distinct boron structures in boron-doped graphene impacts electrocatalytic nitrogen reduction, with a higher ratio indicating an increased catalytic activity. The type of catalyst material and the boron concentration collectively influence the specific catalytic activity [96]. The lower boron concentration in the B-doped Ni/SBA-15 catalysts reduces their catalytic activity [97]. Boron clusters (BC3) in B-doped graphene serve as significant electrocatalysts for the nitrogen reduction reaction (NRR) sites, enhancing graphene reactivity and affecting surface oxidation and catalytic activity [98]. A typical synthesis method for B-doped materials is shown in Figure 3.

3.2. Nitrogen-Doped Catalysts

N-doped catalysts, including B-doped variants, are now increasingly employed in the electrocatalytic reduction of NO3 to NH3 because most undoped electrocatalysts show low efficiencies and NH3 yields, primarily because of the non-polar nature and high binding energy (940.95 kJ/mol) of the triple bond of N molecules, making it difficult to break and slow down the NRR [89]. These doped catalysts generate active sites that cleave the N-N bond, enabling N2 adsorption and the subsequent activation of nitrogen atoms for reduction. The resulting adsorbed nitrogen species are highly reactive and readily reduced to ammonia [99]. N-doped catalysts facilitate electron gain from N species and interactions with hydroxyl groups, thereby accelerating the reduction process. The synthesis ended with NH3 desorption from the catalyst surface, yielding a high output. N-doped phosphorene has shown exceptional performance in this process [100].
Table 3 provides a concise summary of the different synthetic techniques used to prepare the metal catalysts supported on N-doped carbon materials. These methods include the use of capping agents to regulate the formation of nanoparticles, nitrogen-doped carbon to ensure a uniform dispersion of nanoparticles, and MOFs as sacrificial templates for creating metal/N-carbon compositions with distinct properties. For example, N-doped carbon has emerged as a compelling alternative for supporting nanoparticles (NPs) and promoting small and well-dispersed NPs, thereby enhancing catalytic performance [101]. Although capping agents have historically regulated NP growth, their impact on catalytic activity is debatable. N-doped carbon supports are promising alternatives for overcoming the limitations associated with capping agents [102]. Nitrogen functionalities in N-doped carbon act as “tethers”, aiding NP dispersion and conferring resistance to agglomeration and coarsening, thus enhancing catalytic performance [103].
Existing research demonstrates that nitrogen functional groups on carbon supports have a significant impact on the deposition of NPs, the interaction between the metal and support, the electrical properties of the NPs, and the acid–base properties of the resulting materials. Several synthetic methods have been investigated to control the nucleation, growth, and stabilization of N-containing carbons in nanoparticles [104]. The synthesis of NPs supported on N-doped carbon materials can be divided into three main approaches: (i) deposition of NPs onto pre-existing N-doped carbon materials, (ii) simultaneous integration of nitrogen and metal phases onto pre-existing carbon materials, and (iii) on-site formation of NPs on N-doped carbon supports [106].
Typically, methods such as impregnation, deposition–precipitation, and sol immobilization are used to load nanoparticles onto pre-synthesized N-doped carbon materials. These methods control the properties of the support material; however, they require numerous processes [105]. The process of synthesizing NPs on carbon supports doped with nitrogen, undertaken through a pyrolysis stage, enables the simultaneous integration of metal and nitrogen functionalizations. This incorporation occurs under certain conditions and affects the features of the NPs [107]. One notable method involves the use of MOFs as sacrificial templates, in which metal joints and organic linkers undergo precision pyrolysis to transform them into carbon metal–carbon composites with nitrogen. This approach produces catalysts with fascinating compositions and structures, such as gold nanoparticles embedded in N-doped nanoporous carbon, cobalt nanoparticles in N-doped porous carbon, and various other configurations [110].
N-phosphorene was produced by combining N-methyl-2-pyrrolidone (NMP) solvent with ball milling and microwave-assisted exfoliation. Bulk BP crystals were initially ground to increase the surface reactivity [111]. Subsequently, the materials were ball-milled in NMP with or without NH4OH, resulting in the production of N-doped or O-phosphorene. Ball milling weakens the van der Waals contacts and produces reactive species at nanosheet boundaries [108]. Ultrathin phosphorene nanosheets were produced by microwave-assisted liquid-phase exfoliation. AFM measurements revealed the thicknesses of non-doped phosphorene (~4.9 nm), O-phosphorene (~5.5 nm), and N-phosphorene (~8.6 nm) [112]. The Raman spectra showed no post-treatment structural changes. The UV-vis spectra displayed characteristic peaks and blueshifts in the doped samples. XPS analysis revealed P–P and P–O bonds in all samples, with variations in the PxOy peak intensity, indicating successful N-doping and reduced oxygen groups in N-phosphorene, further supported by the HR N 1s spectrum and XPS survey scan [111].
Table 4 offers a comprehensive summary of multiple electrocatalysts utilized in the conversion process of NO3 into NH3. The catalysts listed comprise Pd-Cu/TinO2n−1, Pd-In/γ-Al2O3, 30%Cu–70%Pd, Fe3Ni-N-C, a single Fe atom, Co3O4 nanorod arrays, Rh/C, Cu@Th-BPYDC, RuNi-MOF, and Pd-NDs/Zr-MOF. The table provides information on the NH3 yield, FE, and electrolytes utilized. The properties and characteristics of each catalyst differentiate them, showcasing their capacity to facilitate the conversion of NO3 into NH3 across diverse applications.
The enhanced number of active sites and electronic structure of N-doping likely contribute to the superior catalytic activity of N-phosphorene. Control investigations verified that NH3 gas originated from the injected N2 gas. No NH3 was found in electrolytes saturated with Ar when N-phosphorene was used in copper foam (CF) [111]. Furthermore, a comparison of N-, O-, and phosphorene at various potentials revealed consistent NH3 yield rates (~15 μg h−1 mgcat−1), with N-phosphorene demonstrating superior NRR activity and HER suppression. N-phosphorene also exhibited a lower charge transfer resistance (Rct) (24.2 Ω) than phosphorene (56.3 Ω) and O-phosphorene (34.1 Ω), indicating favorable charge transfer mechanisms due to heteroatom presence [126]. Additionally, durability assessment over six consecutive electrolysis cycles at 0 V showed minimal declines in the NH3 yield rate and FE, confirming the stability of N-phosphorene for continuous NRR electrocatalysis. Long-term testing over 15 h confirmed the constant current density of N-phosphorene, and structural and compositional analyses after six cycles revealed negligible surface composition changes due to N2 adsorption during catalysis, demonstrating the remarkable stability of N-phosphorene [127]. Figure 4 shows the method for the in situ synthesis of nitrogen-doped metal materials.

3.3. Phosphorous-Doped Catalysts

Phosphorus doping often changes the crystal structure of the original catalyst, as shown in Figure 5. However, phosphorus-doped catalysts, such as phosphorus-doped potassium peroxyniobate (P-KNO), have been utilized for the electrocatalytic reduction of nitrogen to ammonia, effectively improving N2-to-NH3 conversion. P-KNO, with enriched oxygen vacancies, demonstrated an NH3 production rate of 23.01 μg h−1 mgcat−1 and an FE of 39.77% at −0.4 V vs. RHE (reversible hydrogen evolution) in a 0.1 M solution [129].
Phosphorus-doped carbon nitride, with abundant nitrogen vacancies, exhibited an outstanding NRR catalytic performance in 0.1 M Na2SO4 electrolyte, with an FE of 22.5% and an NH3 production of 28.67 μg h−1mgcat −1 [131]. In addition, phosphorus in the catalyst structure enhances N2 fixation activity and accelerates N2 adsorption and activation during NRRs, leading to a higher NH3 generation and improved selectivity [132].
The combination of phosphorus and potassium in P-KNO efficiently increased N2-to-NH3 conversion [133]. Nitrogen-doped phosphorene (N-phosphorene) with high crystallinity was synthesized using ball milling and microwave techniques. This synthesis method leads to excellent electrocatalytic performance, with an NH3 yield rate and FE of up to 18.79 μg h−1 mgcat−1 and 21.51%, respectively, at a low overpotential (0 V) compared to RHE [134].

3.4. Sulfur-Doped Catalysts

Sulfur-doped catalysts show promise for the electrocatalytic production of NH3 by reducing nitrogen. The incorporation of sulfur into diverse carbon-based materials, including graphene and porous carbon, enhances their electrocatalytic performance in ammonia synthesis via N2 reduction [135]. Sulfur-doped graphene (S-G) and sulfur-doped porous carbon (S-NPC) have emerged as effective and stable electrocatalysts that promote the reduction of N2 to NH3. Specifically, S-G exhibits efficacy and steadfastness in catalyzing the NRR under normal circumstances, leading to sufficient N2-to-NH3 conversion [136].
The proposed mechanism for S-doped graphene (S-G) involves the development of catalytically active sites, improved adsorption of nitrogen species, facilitated electron transfer, high NH3 yield, and enhanced Faradaic efficiency [135,137]. Density functional theory simulations indicate that carbon atoms adjacent to sulfur substitutions act as catalytically active sites for the NRR on S-G [135]. Incorporating sulfur increases the adsorption of nitrogen species, which is a crucial factor for the electrocatalytic NRR. By facilitating electron transmission between the reactants and electrolytes, sulfur-doped graphene reduces the energy barrier for the NRR, resulting in an accelerated reaction rate. In 0.1 M HCl, S-G exhibits exceptional performance with a substantial NH3 production of 27.3 μg h−1 mgcat−1 and a high Faradaic efficiency of 11.5% at −0.6 and −0.5 V vs. RHE, respectively, surpassing undoped graphene values (6.25 μg h−1 mgcat−1; 0.52%) [137]. The remarkable electrocatalytic performance of sulfur-doped graphene makes it a promising and sustainable electrocatalyst for N2-to-NH3 fixation [138].

4. Nitrogen Synthesis of Ammonia Using Non-Metal-Doped Catalysts

The utilization of non-metallic materials as catalysts for the N2 synthesis of NH3 represents a promising avenue for sustainable and environmentally friendly NH3 production methods. Therefore, as shown in Table 5, NH3 production via electrocatalytic nitrogen reduction presents several challenges. An example of such an issue is the rivalry between the HER and NRR. Enhancing the inhibition of HER improves the effectiveness of the NRR [139]. However, because of their sensitive catalytic structure, precious metal catalysts such as Ru and Rh show increased activity, with Ru on titanium felt enhancing NRRs and NH3 production [140]. Furthermore, Guo et al. found that fluorine-doped graphene demonstrates excellent performance in the oxygen reduction reaction [141], and it can be synthesized via a hydrothermal reflux method [142]. This material holds significant implications for electrocatalytic ammonia production.

5. Synthesizing and Activating Biomass-Derived Carbon Materials for NH3 Production

Biomass-derived carbon materials (BCMs) are a new type of material that possesses various catalytic characteristics, including a significant surface area and porosity. These qualities make them highly suitable for use as heterogeneous catalysts and catalytic supports. The materials exhibit compatibility with a diverse array of processes, encompassing organic transformations, electrocatalytic reactions, and photocatalytic reactions [148,149,150,151]. Biomass carbon, mostly sourced from plants, possesses the qualities of being inexpensive, abundant, eco-friendly, and sustainable, rendering it a very promising substance for electrochemical applications [152,153,154,155]. Recently created materials generated from biomass, such as porous carbons and carbon nanotubes, have superior characteristics, including a larger surface area, improved thermal stability, increased adsorption capacity, and higher electrical conductivity. These advantages are due to their small size and increased porosity.
Nevertheless, they are still very inexperienced in heterogeneous catalysis and require more refinement in order to supplant their counterparts obtained from synthetic carbon precursors [148,152,156,157,158]. Hence, the mutually beneficial association between carbon compounds produced from biomass and non-metallic-doping agents is a crucial pathway in the pursuit of environmentally friendly and high-performing materials for various uses. Biomass-derived carbons may be efficiently modified by using non-metallic-dopants such as N, P, S, and B. These dopants enhance the renewable and structurally adaptable properties of biomass-derived carbons [159]. Nitrogen-doped carbon materials, for instance, have shown exceptional promise in improving the electrochemical properties crucial for energy storage devices like supercapacitors [50,160,161].
Similarly, S-doped carbons have demonstrated heightened catalytic prowess in environmental remediation applications [135,162,163]. The comprehensive characterization of these materials, facilitated by advanced techniques such as X-ray photoelectron spectroscopy (XPS) and Raman spectroscopy, has been instrumental in elucidating the intricate structural and chemical alterations induced by doping strategies [164,165]. Li et al. made N-doped porous carbon from biomass using a basic hydrothermal–pyrolysis method. This material was very good at catalyzing reactions. The material demonstrated a significant ammonia production and FE of 16.3 μgh−1mgcat−1 and 27.5%, respectively [166].
Also, an electrocatalyst made from alfalfa and made up of porous carbon showed a lot of activity in the NRR, reaching an FE of 9.98% [52]. However, the carbon material that was made from bamboo shoots and used to make NH3 through electrochemistry only worked 27.5% of the time under normal temperature and pressure conditions in an acidic solution [166]. Therefore, Giddey et al. state that the faradaic efficiencies mentioned are insufficient for practical use. They suggest that practical needs require a minimum FE of 50% [167]. Carbon-based materials are abundant and have the unique capacity to endure acidic conditions, which makes them well suited for certain catalytic purposes [168,169,170,171,172]. This makes them very suitable for developing carbon-based catalysts without any metal content, specifically for NRRs [173,174]. Huang et al. conducted experiments that show that amorphous-oxygen-doped carbon nanosheets (O-CNs) made from biomass may be very useful as electrocatalysts for NRRs, showing great selectivity with an FE of 4.97% [175].
Similarly, in a study conducted by Wu et al., it was found that oxygen-doped hollow carbon microtubes (O-KFCNTs) derived from natural kapok fibers are a highly efficient electrocatalyst for converting N2 to NH3 without the use of metals. These microtubes exhibit excellent selectivity and achieve a high FE of 9.1% [153]. Therefore, future studies must prioritize enhancing synthesis techniques to ensure scalability and investigate the synergistic impacts of combining doping with complementary modifications to materials. Carbon materials obtained from biomass can improve performance in several multifunctional material platforms.

6. Conclusions and Prospects

Extensive research conducted over the years on non-metallic-doped materials for electrocatalytic NH3 synthesis has made considerable advancements and has the potential to solve urgent challenges in the sectors of sustainability, emission reduction, and energy transition. The incorporation of non-metallic dopants has shown encouraging outcomes in improving the electrocatalytic performance and specificity of materials. By altering various crystal structures, these dopants have demonstrated the capacity to enhance the rate of NH3 production, faradaic efficiency, and stability. More importantly, there has been progress in understanding the basic principles that control the electrocatalytic performance of materials doped with non-metallic elements. These basic principles have provided insights into the effects of doping on the electronic structure, surface characteristics, and reaction kinetics. This understanding serves as a strong basis for the development and enhancement in future catalysts for achieving efficient NH3 generation. The creation of non-metallic-doped materials for electrocatalytic NH3 generation is a significant advancement in sustainable and ecologically friendly synthesis processes.
Nevertheless, the electrochemical method presents the possibility of achieving carbon-neutral NH3 synthesis using sustainable energy sources and effective nitrate reduction. The shift towards electrocatalytic NH3 generation is in line with worldwide initiatives to reduce greenhouse gas emissions and move towards a sustainable energy future. In electrocatalytic NH3 synthesis, there are significant prospects and difficulties related to non-metallic-doped materials that need to be considered in the future. Additional investigations are required to examine other substances that can be added and their combined effects, as well as to refine the procedures used to create the substance to improve its performance and capacity to be produced on a larger scale. Thorough research is necessary to ensure the long-term stability and durability of the catalysts under actual working conditions.
Furthermore, the incorporation of catalysts doped with non-metallic elements into electrocatalytic systems that are applicable in real-world settings and the expansion of these systems to an industrial scale continue to be a crucial stage in the process of making them commercially viable. Effective collaboration among academics, industry, and policymakers is essential in addressing the technological, economic, and regulatory factors necessary for the effective deployment of electrocatalytic NH3 synthesis. The advancements achieved in non-metallic-doped materials for electrocatalytic NH3 generation represent a promising approach for achieving sustainable NH3 synthesis. These catalysts have the potential to transform the NH3 manufacturing business and contribute to a more environmentally friendly and sustainable future by addressing important issues, such as sustainability, emission reduction, and energy transition.

Author Contributions

Conceptualization, writing—original draft preparation, writing—review and editing, G.D.S.Q.J.; conceptualization, writing—review and editing, M.J.; supervision and funding acquisition, K.L. and Y.W. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by the Shanghai Municipal Committee of Science and Technology (grant number 22230712800).

Acknowledgments

The authors wish to extend their profound gratitude to the Shanghai Municipal Committee of Science and Technology for their financial support.

Conflicts of Interest

The authors declare no conflicts of interest.

Nomenclature

NVNitrogen vacancies
BBoron
PPhosphorus
SSulfur
NH3 Ammonia
NO3Nitrate
MOFs Metal–organic frameworks
HER Hydrogen evolution reaction
CVD Chemical vapor deposition
HTC Hydrothermal carbonization
ORR Oxygen reduction reaction
GO Graphene oxide
NG Graphene sheets
N-RGO N-doped reduced graphene oxide
CNTs Carbon nanotubes
N-CNFs N-doped carbon nanofibers
NRRs Nitrogen reduction reactions
HER Hydrogen evolution reaction
FE Faradaic efficiency
O-KFCNTsHollow carbon microtubes
NMPN-methyl-2-pyrrolidone
gC3N4 porous polymeric carbon nitride
C2H4N4 Dicyandiamide
C3H7NODimethylformamide
C3H6N6Melamine
BCMsBiomass-derived carbon materials

References

  1. Tao, W.; Wang, P.; Hu, B.; Wang, X.; Zhou, G. Accelerating the reaction kinetics from nitrate to ammonia by anion substitution in NiCo-based catalysts. J. Environ. Chem. Eng. 2023, 11, 109117. [Google Scholar] [CrossRef]
  2. Zhai, G.-Y.; Li, Q.-Y.; Zhang, S.-N.; Xu, D.; Xia, S.-Y.; Gao, P.; Lin, X.; Lin, Y.-X.; Cheng, J.-H.; Hu, W.-Y.; et al. Accelerating the Activation of NOx−on Ru Nanoparticles for Ammonia Production by Tuning Their Electron Deficiency. CCS Chem. 2022, 4, 3455–3462. [Google Scholar] [CrossRef]
  3. Wyer, K.E.; Kelleghan, D.B.; Blanes-Vidal, V.; Schauberger, G.; Curran, T.P. Ammonia emissions from agriculture and their contribution to fine particulate matter: A review of implications for human health. J. Environ. Manag. 2022, 323, 116285. [Google Scholar] [CrossRef] [PubMed]
  4. Liu, L.; Xu, W.; Lu, X.; Zhong, B.; Guo, Y.; Lu, X.; Zhao, Y.; He, W.; Wang, S.; Zhang, X.; et al. Exploring global changes in agricultural ammonia emissions and their contribution to nitrogen deposition since 1980. Proc. Natl. Acad. Sci. USA 2022, 119, e2121998119. [Google Scholar] [CrossRef] [PubMed]
  5. Mehr, H.G.; Torabiardekani, N.; Rahimpour, M.R. Chapter eleven—Ammonia application in fabric, textile, and leather products. In Progresses in Ammonia: Science, Technology and Membranes; Basile, A., Rahimpour, M.R., Eds.; Elsevier: Amsterdam, The Netherlands, 2024; pp. 223–239. [Google Scholar]
  6. Fan, K.; Xie, W.; Li, J.; Sun, Y.; Xu, P.; Tang, Y.; Li, Z.; Shao, M. Active hydrogen boosts electrochemical nitrate reduction to ammonia. Nat. Commun. 2022, 13, 7958. [Google Scholar] [CrossRef] [PubMed]
  7. Yüzbaşıoğlu, A.E.; Avşar, C.; Gezerman, A.O. The current situation in the use of ammonia as a sustainable energy source and its industrial potential. Curr. Res. Green Sustain. Chem. 2022, 5, 100307. [Google Scholar] [CrossRef]
  8. Shahed Gharahshiran, V.; Zheng, Y. Sustainable ammonia synthesis: An in-depth review of non-thermal plasma technologies. J. Energy Chem. 2024, 96, 1–38. [Google Scholar] [CrossRef]
  9. Wu, T.; Fan, W.; Zhang, Y.; Zhang, F. Electrochemical synthesis of ammonia: Progress and challenges. Mater. Today Phys. 2021, 16, 100310. [Google Scholar] [CrossRef]
  10. Zhou, Y.; Lv, J.; Peng, Y.; Lin, X.; Lv, X.; Ye, Q.; Liu, S.; Wu, A.; Li, X. Ar-plasma enhanced copper-nickel alloy catalysis for ammonia synthesis. Waste Dispos. Sustain. Energy 2022, 4, 149–155. [Google Scholar]
  11. Shan, Y.; Huang, Q.; Guan, D.; Hubacek, K. China CO2 emission accounts 2016–2017. Sci. Data 2020, 7, 54. [Google Scholar] [CrossRef]
  12. Zhang, X.; Wang, Y.; Liu, C.; Yu, Y.; Lu, S.; Zhang, B. Recent advances in non-noble metal electrocatalysts for nitrate reduction. Chem. Eng. J. 2021, 403, 126269. [Google Scholar] [CrossRef]
  13. Yuan, S.; Xue, Y.; Ma, R.; Ma, Q.; Chen, Y.; Fan, J. Advances in iron-based electrocatalysts for nitrate reduction. Sci. Total Environ. 2023, 866, 161444. [Google Scholar] [CrossRef] [PubMed]
  14. Huang, Z.; Rafiq, M.; Woldu, A.R.; Tong, Q.-X.; Astruc, D.; Hu, L. Recent progress in electrocatalytic nitrogen reduction to ammonia (NRR). Coord. Chem. Rev. 2023, 478, 214981. [Google Scholar] [CrossRef]
  15. Davy, H.S.I. The Bakerian Lecture, on some chemical agencies of electricity. Philos. Trans. R. Soc. Lond. 1807, 31, 1–56. [Google Scholar]
  16. Martín, A.J.; Shinagawa, T.; Pérez-Ramírez, J. Electrocatalytic Reduction of Nitrogen: From Haber-Bosch to Ammonia Artificial Leaf. Chem 2019, 5, 263–283. [Google Scholar] [CrossRef]
  17. Fichter, F.; Suter, R. Zur Frage der kathodischen Reduktion des elementaren Stickstoffs. Helv. Chim. Acta 2004, 5, 246–255. [Google Scholar] [CrossRef]
  18. Tamelen, E.E.v.; Åkermark, B. Electrolytic reduction of molecular nitrogen. J. Am. Chem. Soc. 1968, 90, 4492–4493. [Google Scholar] [CrossRef]
  19. Ren, Y.; Yu, C.; Tan, X.; Huang, H.; Wei, Q.; Qiu, J. Strategies to suppress hydrogen evolution for highly selective electrocatalytic nitrogen reduction: Challenges and perspectives. Energy Environ. Sci. 2021, 14, 1176–1193. [Google Scholar] [CrossRef]
  20. Tsuneto, A.; Kudo, A.; Sakata, T. Lithium-mediated electrochemical reduction of high pressure N2 to NH3. J. Electroanal. Chem. 1994, 367, 183–188. [Google Scholar] [CrossRef]
  21. Rod, T.H.; Logadottir, A.; Nørskov, J.K. Ammonia synthesis at low temperatures. J. Chem. Phys. 2000, 112, 5343–5347. [Google Scholar] [CrossRef]
  22. Kordali, V.; Kyriacou, G.; Lambrou, C. Electrochemical synthesis of ammonia at atmospheric pressure and low temperature in a solid polymer electrolyte cell. Chem. Commun. 2000, 6, 1673–1674. [Google Scholar] [CrossRef]
  23. Köleli, F.; Röpke, T. Electrochemical hydrogenation of dinitrogen to ammonia on a polyaniline electrode. Appl. Catal. B Environ. 2006, 62, 306–310. [Google Scholar] [CrossRef]
  24. Skúlason, E.; Bligaard, T.; Gudmundsdóttir, S.; Studt, F.; Rossmeisl, J.; Abild-Pedersen, F.; Vegge, T.; Jónsson, H.; Nørskov, J.K. A theoretical evaluation of possible transition metal electro-catalysts for N2 reduction. Phys. Chem. Chem. Phys. 2012, 14, 1235–1245. [Google Scholar] [CrossRef]
  25. Azofra, L.M.; Li, N.; MacFarlane, D.R.; Sun, C. Promising prospects for 2D d2–d4 M3C2 transition metal carbides (MXenes) in N2 capture and conversion into ammonia. Energy Environ. Sci. 2016, 9, 2545–2549. [Google Scholar] [CrossRef]
  26. Abghoui, Y.; Skúlason, E. Onset potentials for different reaction mechanisms of nitrogen activation to ammonia on transition metal nitride electro-catalysts. Catal. Today 2017, 286, 69–77. [Google Scholar] [CrossRef]
  27. Zhao, J.; Zhang, L.; Xie, X.-Y.; Li, X.; Ma, Y.; Liu, Q.; Fang, W.-H.; Shi, X.; Cui, G.; Sun, X. Ti3C2Tx (T = F, OH) MXene nanosheets: Conductive 2D catalysts for ambient electrohydrogenation of N2 to NH3. J. Mater. Chem. A 2018, 6, 24031–24035. [Google Scholar] [CrossRef]
  28. Du, H.L.; Chatti, M.; Hodgetts, R.Y.; Cherepanov, P.V.; Nguyen, C.K.; Matuszek, K.; MacFarlane, D.R.; Simonov, A.N. Electroreduction of nitrogen with almost 100% current-to-ammonia efficiency. Nature 2022, 609, 722–727. [Google Scholar] [CrossRef]
  29. Fan, X.; Zhao, D.; Deng, Z.; Zhang, L.; Li, J.; Li, Z.; Sun, S.; Luo, Y.; Zheng, D.; Wang, Y.; et al. Constructing Co@TiO2 Nanoarray Heterostructure with Schottky Contact for Selective Electrocatalytic Nitrate Reduction to Ammonia. Small 2023, 19, e2208036. [Google Scholar] [CrossRef]
  30. Li, H.; Wang, Y.; Chen, S.; Peng, F.; Gao, F. Boosting Electrochemical Reduction of Nitrate to Ammonia by Constructing Nitrate-Favored Active Cu–B Sites on SnS2. Small 2024, 20, 2308182. [Google Scholar] [CrossRef]
  31. Zhang, N.; Zhang, G.; Tian, Y.; Tang, Y.; Chu, K. FeS2 nanoparticles on reduced graphene oxide: An efficient electrocatalyst for nitrate electroreduction to ammonia. Dalton Trans. 2022, 51, 16805–16810. [Google Scholar] [CrossRef]
  32. Zhang, N.; Zhang, G.-k.; Shen, P.; Zhang, H.; Ma, D.; Chu, K. Lewis Acid Fe-V Pairs Promote Nitrate Electroreduction to Ammonia. Adv. Funct. Mater. 2023, 33, 2211537. [Google Scholar] [CrossRef]
  33. Zhang, S.; Li, M.; Li, J.; Song, Q.; Liu, X. N-doped carbon–iron heterointerfaces for boosted electrocatalytic active and selective ammonia production. Proc. Natl. Acad. Sci. USA 2023, 120, e2207080119. [Google Scholar] [CrossRef] [PubMed]
  34. Gao, W.; Xie, K.; Xie, J.; Wang, X.; Zhang, H.; Chen, S.-P.; Wang, H.; Li, Z.; Li, C. Alloying of Cu with Ru Enabling the Relay Catalysis for Reduction of Nitrate to Ammonia. Adv. Mater. 2023, 35, e2202952. [Google Scholar] [CrossRef] [PubMed]
  35. Guo, H.; Li, M.; Yang, Y.; Luo, R.; Liu, W.; Zhang, F.; Tang, C.; Yang, G.; Zhou, Y. Self-Supported Pd Nanorod Arrays for High-Efficient Nitrate Electroreduction to Ammonia. Small 2023, 19, e2207743. [Google Scholar] [CrossRef] [PubMed]
  36. Chen, K.; Ma, Z.; Li, X.; Kang, J.; Ma, D.; Chu, K. Single-Atom Bi Alloyed Pd Metallene for Nitrate Electroreduction to Ammonia. Adv. Funct. Mater. 2023, 33, 2209890. [Google Scholar] [CrossRef]
  37. Pan, F.; Zhou, J.; Wang, T.; Zhu, Y.; Ma, H.; Niu, J.; Wang, C. Revealing the activity origin of ultrathin nickel metal–organic framework nanosheet catalysts for selective electrochemical nitrate reduction to ammonia: Experimental and density functional theory investigations. J. Colloid Interface Sci. 2023, 638, 26–38. [Google Scholar] [CrossRef] [PubMed]
  38. Kim, K.; Zagalskaya, A.; Ng, J.L.; Hong, J.; Alexandrov, V.; Pham, T.A.; Su, X. Coupling nitrate capture with ammonia production through bifunctional redox-electrodes. Nat. Commun. 2023, 14, 823. [Google Scholar] [CrossRef]
  39. Xu, B.; Chen, Z.; Zhang, G.; Wang, Y. On-Demand Atomic Hydrogen Provision by Exposing Electron-Rich Cobalt Sites in an Open-Framework Structure toward Superior Electrocatalytic Nitrate Conversion to Dinitrogen. Environ. Sci. Technol. 2022, 56, 614–623. [Google Scholar] [CrossRef]
  40. Xiong, R.; Xu, B.; Chen, Z.; Cheng, C.; Wang, Y. Self-supporting electrode incorporating active Co sites for ultrafast ammonia production from nitrate reduction. Colloids Surf. A Physicochem. Eng. Asp. 2024, 687, 133557. [Google Scholar] [CrossRef]
  41. Feng, T.; Wang, J.; Wang, Y.; Yu, C.; Zhou, X.; Xu, B.; László, K.; Li, F.; Zhang, W. Selective electrocatalytic reduction of nitrate to dinitrogen by Cu2O nanowires with mixed oxidation-state. Chem. Eng. J. 2022, 433, 133495. [Google Scholar] [CrossRef]
  42. Hu, Q.; Qin, Y.J.; Wang, X.D.; Zheng, H.J.; Gao, K.R.; Yang, H.P.; Zhang, P.X.; Shao, M.H.; He, C.X. Grain Boundaries Engineering of Hollow Copper Nanoparticles Enables Highly Efficient Ammonia Electrosynthesis from Nitrate. Ccs Chem. 2022, 4, 2053–2064. [Google Scholar] [CrossRef]
  43. Chanda, D.; Xing, R.M.; Xu, T.; Liu, Q.; Luo, Y.L.; Liu, S.H.; Tufa, R.A.; Dolla, T.H.; Montini, T.; Sun, X.P. Electrochemical nitrogen reduction: Recent progress and prospects. Chem. Commun. 2021, 57, 7335–7349. [Google Scholar] [CrossRef]
  44. Lin, J.; Xu, T.; Qi, J. Chapter 6—Defect engineering for modifying transition metal oxides. In Metal Oxides and Related Solids for Electrocatalytic Water Splitting; Qi, J., Korotcenkov, G., Eds.; Elsevier: Amsterdam, The Netherlands, 2022; pp. 161–190. [Google Scholar] [CrossRef]
  45. Qiu, W.; Liu, Y.; Xie, M.; Jin, Z.; Li, P.; Yu, G. Structural engineering of catalysts for ammonia electrosynthesis from nitrate: Recent advances and challenges. EES Catal. 2024, 2, 202–219. [Google Scholar] [CrossRef]
  46. Yu, S.; Xiang, T.; Alharbi, N.S.; Al-aidaroos, B.A.; Chen, C. Recent development of catalytic strategies for sustainable ammonia production. Chin. J. Chem. Eng. 2023, 62, 65–113. [Google Scholar] [CrossRef]
  47. Liang, X.; Zhu, H.; Yang, X.; Xue, S.; Liang, Z.; Ren, X.; Liu, A.; Wu, G. Recent Advances in Designing Efficient Electrocatalysts for Electrochemical Nitrate Reduction to Ammonia. Small Struct. 2022, 4, 2200202. [Google Scholar] [CrossRef]
  48. Huang, Y.; Long, J.; Wang, Y.; Meng, N.; Yu, Y.; Lu, S.; Xiao, J.; Zhang, B. Engineering Nitrogen Vacancy in Polymeric Carbon Nitride for Nitrate Electroreduction to Ammonia. ACS Appl. Mater. Interfaces 2021, 13, 54967–54973. [Google Scholar] [CrossRef]
  49. Zhang, M.; Choi, C.; Huo, R.; Gu, G.H.; Hong, S.; Yan, C.; Xu, S.; Robertson, A.W.; Qiu, J.; Jung, Y.; et al. Reduced graphene oxides with engineered defects enable efficient electrochemical reduction of dinitrogen to ammonia in wide pH range. Nano Energy 2020, 68, 104323. [Google Scholar] [CrossRef]
  50. Salinas-Torres, D.; Navlani-García, M.; Mori, K.; Kuwahara, Y.; Yamashita, H. Nitrogen-doped carbon materials as a promising platform toward the efficient catalysis for hydrogen generation. Appl. Catal. A Gen. 2019, 571, 25–41. [Google Scholar] [CrossRef]
  51. Al-Naggar, A.H.; Shinde, N.M.; Kim, J.-S.; Mane, R.S. Water splitting performance of metal and non-metal-doped transition metal oxide electrocatalysts. Coord. Chem. Rev. 2023, 474, 214864. [Google Scholar] [CrossRef]
  52. Zhao, C.; Zhang, S.; Han, M.; Zhang, X.; Liu, Y.; Li, W.; Chen, C.; Wang, G.; Zhang, H.; Zhao, H. Ambient Electrosynthesis of Ammonia on a Biomass-Derived Nitrogen-Doped Porous Carbon Electrocatalyst: Contribution of Pyridinic Nitrogen. ACS Energy Lett. 2019, 4, 377–383. [Google Scholar] [CrossRef]
  53. Li, X.; Fang, Y.; Zhao, S.; Wu, J.; Li, F.; Tian, M.; Long, X.; Jin, J.; Ma, J. Nitrogen-doped mesoporous carbon nanosheet/carbon nanotube hybrids as metal-free bi-functional electrocatalysts for water oxidation and oxygen reduction. J. Mater. Chem. A 2016, 4, 13133–13141. [Google Scholar] [CrossRef]
  54. Mandal, B.; Saha, S.; Das, D.; Panda, J.; Das, S.; Sarkar, R.; Tudu, B. Supercapacitor performance of nitrogen doped graphene synthesized via DMF assisted single-step solvothermal method. FlatChem 2022, 34, 100400. [Google Scholar] [CrossRef]
  55. Chen, W.; Yang, H.; Chen, Y.; Chen, X.; Fang, Y.; Chen, H. Biomass pyrolysis for nitrogen-containing liquid chemicals and nitrogen-doped carbon materials. J. Anal. Appl. Pyrolysis 2016, 120, 186–193. [Google Scholar] [CrossRef]
  56. Chen, H.; Ding, L.; Zhang, K.; Chen, Z.; Lei, Y.; Zhou, Z.; Hou, R. Preparation of chemically reduced graphene using hydrazine hydrate as the reduction agent and its NO2 sensitivity at room temperature. Int. J. Electrochem. Sci. 2020, 15, 10231–10242. [Google Scholar] [CrossRef]
  57. Moorthy, M.; Kumar, V.B.; Porat, Z.e.; Gedanken, A. Novel polymerization of aniline and pyrrole by carbon dots. New J. Chem. 2018, 42, 535–540. [Google Scholar] [CrossRef]
  58. Guo, H.-L.; Su, P.; Kang, X.; Ning, S.-K. Synthesis and characterization of nitrogen-doped graphene hydrogels by hydrothermal route with urea as reducing-doping agents. J. Mater. Chem. A 2013, 1, 2248–2255. [Google Scholar] [CrossRef]
  59. González-Gaitán, C.; Ruiz-Rosas, R.; Morallón, E.; Cazorla-Amorós, D. Functionalization of carbon nanotubes using aminobenzene acids and electrochemical methods. Electroactivity for the oxygen reduction reaction. Int. J. Hydrogen Energy 2015, 40, 11242–11253. [Google Scholar] [CrossRef]
  60. Wong, W.Y.; Daud, W.R.W.; Mohamad, A.B.; Kadhum, A.A.H.; Loh, K.S.; Majlan, E.H. Recent progress in nitrogen-doped carbon and its composites as electrocatalysts for fuel cell applications. Int. J. Hydrogen Energy 2013, 38, 9370–9386. [Google Scholar] [CrossRef]
  61. Mohan, T.V.R.; Nallagangula, M.; Kala, K.; Hernandez-Tamargo, C.E.; De Leeuw, N.H.; Namitharan, K.; Bhat, V.T.; Sasidharan, M.; Selvam, P. Pyridinic-nitrogen on ordered mesoporous carbon: A versatile NAD(P)H mimic for borrowing-hydrogen reactions. J. Catal. 2023, 419, 80–98. [Google Scholar] [CrossRef]
  62. Mallakpour, S.; Soltanian, S. Surface functionalization of carbon nanotubes: Fabrication and applications. RSC Adv. 2016, 6, 109916–109935. [Google Scholar] [CrossRef]
  63. Lu, J.; Bo, X.; Wang, H.; Guo, L. Nitrogen-doped ordered mesoporous carbons synthesized from honey as metal-free catalyst for oxygen reduction reaction. Electrochim. Acta 2013, 108, 10–16. [Google Scholar] [CrossRef]
  64. Wei, X.; Wang, M.-S.; Bando, Y.; Golberg, D. Post-Synthesis Carbon Doping of Individual Multiwalled Boron Nitride Nanotubes via Electron-Beam Irradiation. J. Am. Chem. Soc. 2010, 132, 13592–13593. [Google Scholar] [CrossRef] [PubMed]
  65. Jiang, Z.; Kato, Y.; Al-zubaidi, A.; Yamamoto, K.; Kawasaki, S. Effect of post-synthesis nitrogen doping in nanocarbons on cathode reaction of metal-air cells. Mater. Express 2014, 4, 337–342. [Google Scholar] [CrossRef]
  66. Milliken, S.; Cheong, T.; Cui, K.; Veinot, J.G.C. Post-Synthesis Boron Doping of Silicon Quantum Dots via Hydrosilsesquioxane-Capped Thermal Diffusion. ChemRxiv 2022. [Google Scholar] [CrossRef]
  67. Kim, Y.; Ryu, J.; Park, M.; Kim, E.S.; Yoo, J.M.; Park, J.; Kang, J.H.; Hong, B.H. Vapor-phase molecular doping of graphene for high-performance transparent electrodes. ACS Nano 2014, 8, 868–874. [Google Scholar] [CrossRef] [PubMed]
  68. Ma, Y.; Lu, Z.; Shan, G.; Chen, Y. Solvothermal Fabrication of Nitrogen-Doped Carbon Nanoparticles as Efficient Catalyst for Oxygen Reduction in KOH Electrolyte. ChemistrySelect 2017, 2, 5390–5393. [Google Scholar] [CrossRef]
  69. Dlamini, M.W.; Phaahlamohlaka, T.N.; Kumi, D.O.; Forbes, R.; Jewell, L.L.; Coville, N.J. Post doped nitrogen-decorated hollow carbon spheres as a support for Co Fischer-Tropsch catalysts. Catal. Today 2020, 342, 99–110. [Google Scholar] [CrossRef]
  70. Han, Q.; Wu, C.; Jiao, H.; Xu, R.; Wang, Y.; Xie, J.; Guo, Q.; Tang, J. Rational Design of High-Concentration Ti3+ in Porous Carbon-Doped TiO2 Nanosheets for Efficient Photocatalytic Ammonia Synthesis. Adv. Mater. 2021, 33, 2008180. [Google Scholar] [CrossRef] [PubMed]
  71. Matsagar, B.M.; Yang, R.-X.; Dutta, S.; Ok, Y.S.; Wu, K.C.W. Recent progress in the development of biomass-derived nitrogen-doped porous carbon. J. Mater. Chem. A 2021, 9, 3703–3728. [Google Scholar] [CrossRef]
  72. Mainali, K.; Mood, S.H.; Pelaez-Samaniego, M.R.; Sierra-Jimenez, V.; Garcia-Perez, M. Production and applications of N-doped carbons from bioresources: A review. Catal. Today 2023, 423, 114248. [Google Scholar] [CrossRef]
  73. Gavrilov, N.; Pašti, I.A.; Mitrić, M.; Travas-Sejdić, J.; Ćirić-Marjanović, G.; Mentus, S.V. Electrocatalysis of oxygen reduction reaction on polyaniline-derived nitrogen-doped carbon nanoparticle surfaces in alkaline media. J. Power Sources 2012, 220, 306–316. [Google Scholar] [CrossRef]
  74. Saha, D.; Kienbaum, M.J. Role of oxygen, nitrogen and sulfur functionalities on the surface of nanoporous carbons in CO2 adsorption: A critical review. Microporous Mesoporous Mater. 2019, 287, 29–55. [Google Scholar] [CrossRef]
  75. Wang, Y.; Shao, Y.; Wang, H.; Yuan, J. Advanced Heteroatom-Doped Porous Carbon Membranes Assisted by Poly(ionic liquid) Design and Engineering. Acc. Mater. Res. 2020, 1, 16–29. [Google Scholar] [CrossRef] [PubMed]
  76. Payattikul, L.; Chen, C.-Y.; Chen, Y.-S.; Raja Pugalenthi, M.; Punyawudho, K. Recent Advances and Synergistic Effects of Non-Precious Carbon-Based Nanomaterials as ORR Electrocatalysts: A Review. Molecules 2023, 28, 7751. [Google Scholar] [CrossRef] [PubMed]
  77. Liu, K.; Cao, P.; Chen, W.; Ezeh, C.I.; Chen, Z.; Luo, Y.; Liu, Q.; Zhao, H.; Rui, Z.; Gao, S.; et al. Electrocatalysis enabled transformation of earth-abundant water, nitrogen and carbon dioxide for a sustainable future. Mater. Adv. 2022, 3, 1359–1400. [Google Scholar] [CrossRef]
  78. Zhang, S.; Tian, K.; Cheng, B.-H.; Jiang, H. Preparation of N-Doped Supercapacitor Materials by Integrated Salt Templating and Silicon Hard Templating by Pyrolysis of Biomass Wastes. ACS Sustain. Chem. Eng. 2017, 5, 6682–6691. [Google Scholar] [CrossRef]
  79. Song, Y.; Yang, J.; Wang, K.; Haller, S.; Wang, Y.; Wang, C.; Xia, Y. In-situ synthesis of graphene/nitrogen-doped ordered mesoporous carbon nanosheet for supercapacitor application. Carbon 2016, 96, 955–964. [Google Scholar] [CrossRef]
  80. Suboch, A.N.; Podyacheva, O.Y. Pd Catalysts Supported on Bamboo-Like Nitrogen-Doped Carbon Nanotubes for Hydrogen Production. Energies 2021, 14, 1501. [Google Scholar] [CrossRef]
  81. Wang, B.; Liu, B.; Dai, L. Non-N-Doped Carbons as Metal-Free Electrocatalysts. Adv. Sustain. Syst. 2020, 5, 2000134. [Google Scholar] [CrossRef]
  82. Lu, L.; Cao, X.; Shen, Z.; Li, L.; Huo, J.; Chen, W.; Liu, C.; Liu, H. Electrospun nitrogen-doped carbon nanofibers for electrocatalysis. Sustain. Mater. Technol. 2020, 26, e00221. [Google Scholar] [CrossRef]
  83. Zhou, Q.; Yu, K.; Pan, Z.; Huang, Z.; Xu, Y.; Hu, G.; Wu, S.; Chen, C.; Lin, L.; Lin, Y. Research on a novel Ni-doped TiN modified N-doped CNTs supported Pt catalysts and their synergistic effect for methanol electrooxidation. Int. J. Hydrogen Energy 2018, 43, 22519–22528. [Google Scholar] [CrossRef]
  84. Tong, X.; Chen, Z.; Zhuo, H.; Hu, Y.; Jing, S.; Liu, J.; Zhong, L. Tailoring the physicochemical properties of chitosan-derived N-doped carbon by controlling hydrothermal carbonization time for high-performance supercapacitor application. Carbohydr. Polym. 2019, 207, 764–774. [Google Scholar] [CrossRef] [PubMed]
  85. Hamsa, A.P.; Arulprakasam, M.; Unni, S.M. Electrochemical nitrogen fixation on single metal atom catalysts. Chem. Commun. 2023, 59, 10689–10710. [Google Scholar] [CrossRef] [PubMed]
  86. Zhao, R.; Li, Q.; Chen, Z.; Jose, V.; Jiang, X.; Fu, G.; Lee, J.-M.; Huang, S. B, N-doped ultrathin carbon nanosheet superstructure for high-performance oxygen reduction reaction in rechargeable zinc-air battery. Carbon 2020, 164, 398–406. [Google Scholar] [CrossRef]
  87. Wang, J.; Shi, M.; Yi, G.; Meng, J.; Li, Q. Computational identification of B substitutional doped C9N4 monolayer for electrocatalytic N2 reduction. Mol. Catal. 2021, 511, 111726. [Google Scholar] [CrossRef]
  88. Zhang, G.; Li, Y.; He, C.; Ren, X.; Zhang, P.; Mi, H. Recent Progress in 2D Catalysts for Photocatalytic and Electrocatalytic Artificial Nitrogen Reduction to Ammonia. Adv. Energy Mater. 2021, 11, 2003294. [Google Scholar] [CrossRef]
  89. Lu, X.; Yu, J.; Cai, J.; Zhang, Q.; Yang, S.; Gu, L.; Waterhouse, G.I.N.; Zang, S.-Q.; Yang, B.; Lu, S. Exclusive nitrate to ammonia conversion via boron-doped carbon dots induced surface Lewis acid sites. Cell Rep. Phys. Sci. 2022, 3, 100961. [Google Scholar] [CrossRef]
  90. Zheng, M.; Xu, Z.-N.; Fu, M.-L. Boron-doped activated carbon catalyzed reduction of dilute nitrite acid in oxidative esterification reaction in the coal to ethylene glycol process. J. Environ. Chem. Eng. 2022, 10, 107932. [Google Scholar] [CrossRef]
  91. Gou, F.; Wang, H.; Fu, M.; Jiang, Y.; Shen, W.; He, R.; Li, M. Boron-induced electron localization in Cu nanowires promotes efficient nitrate reduction to ammonia in neutral media. Appl. Surf. Sci. 2023, 612, 155872. [Google Scholar] [CrossRef]
  92. Gou, Q.; Mao, Y.; Lv, S.; Gou, F.; Jiang, Y.; Shen, W.; Li, M.; Wang, Y.; He, R. Enhancing nitrate reduction to ammonia by synergistic and interface coupling effects of binary metal sites. Appl. Catal. B Environ. Energy 2024, 348, 123810. [Google Scholar] [CrossRef]
  93. Kuang, P.; Natsui, K.; Feng, C.; Einaga, Y. Electrochemical reduction of nitrate on boron-doped diamond electrodes: Effects of surface termination and boron-doping level. Chemosphere 2020, 251, 126364. [Google Scholar] [CrossRef]
  94. Ren, C.; Zhang, Y.; Li, Y.; Zhang, Y.; Huang, S.; Lin, W.; Ding, K. Whether Corrugated or Planar Vacancy Graphene-like Carbon Nitride (g-C3N4) Is More Effective for Nitrogen Reduction Reaction? J. Phys. Chem. C 2019, 123, 17296–17305. [Google Scholar] [CrossRef]
  95. Majumder, M.; Saini, H.; Dědek, I.; Schneemann, A.; Chodankar, N.R.; Ramarao, V.; Santosh, M.S.; Nanjundan, A.K.; Kment, Š.; Dubal, D.; et al. Rational Design of Graphene Derivatives for Electrochemical Reduction of Nitrogen to Ammonia. ACS Nano 2021, 15, 17275–17298. [Google Scholar] [CrossRef] [PubMed]
  96. Ouyang, W.; Zhi, Q.; Gong, L.; Sun, H.; Liu, M.; Zhang, J.; Han, X.; Xia, Z.; Zhang, L. Rational design of boron-containing co-doped graphene as highly efficient electro-catalysts for the nitrogen reduction reaction. J. Mater. Chem. A 2021, 9, 24590–24599. [Google Scholar] [CrossRef]
  97. Singh, S.; Nguyen, T.D.; Siang, T.J.; Phuong, P.T.T.; Huy Phuc, N.H.; Truong, Q.D.; Lam, S.S.; Vo, D.-V.N. Boron-doped Ni/SBA-15 catalysts with enhanced coke resistance and catalytic performance for dry reforming of methane. J. Energy Inst. 2020, 93, 31–42. [Google Scholar] [CrossRef]
  98. Wang, F.; Xia, L.; Li, X.; Yang, W.; Zhao, Y.; Mao, J. Nano-Ferric Oxide Embedded in Graphene Oxide: High-performance Electrocatalyst for Nitrogen Reduction at Ambient Condition. Energy Environ. Mater. 2020, 4, 88–94. [Google Scholar] [CrossRef]
  99. Yan, X.; Liu, D.; Cao, H.; Hou, F.; Liang, J.; Dou, S.X. Nitrogen Reduction to Ammonia on Atomic-Scale Active Sites under Mild Conditions. Small Methods 2019, 3, 1800501. [Google Scholar] [CrossRef]
  100. Peng, X.; Wang, C.; Tan, Z.; Ni, J.; Lin, B.; Lin, J.; Wang, X.; Zheng, L.; Au, C.-t.; Jiang, L. N-Induced Electron Transfer Effect on Low-Temperature Activation of Nitrogen for Ammonia Synthesis over Co-Based Catalysts. ACS Sustain. Chem. Eng. 2021, 9, 1529–1539. [Google Scholar] [CrossRef]
  101. Nichols, F.; Liu, Q.; Sandhu, J.; Azhar, Z.; Cazares, R.; Mercado, R.; Bridges, F.; Chen, S. Platinum-complexed phosphorous-doped carbon nitride for electrocatalytic hydrogen evolution. J. Mater. Chem. A 2022, 10, 5962–5970. [Google Scholar] [CrossRef]
  102. Rafeeq, H.; Hussain, A.; Ambreen, A.; Zill e, H.; Waqas, M.; Bilal, M.; Iqbal, H.M.N. Functionalized nanoparticles and their environmental remediation potential: A review. J. Nanostructure Chem. 2022, 12, 1007–1031. [Google Scholar] [CrossRef]
  103. Sun, J.; Lu, J.; Huang, C.; Wu, Q.; Xia, L.; Xu, Q.; Yao, W. Modification of Ni3N with a Cobalt-Doped Carbon Shell for High-Performance Hydrogen Evolution in Alkaline Media. ACS Sustain. Chem. Eng. 2021, 9, 1994–2002. [Google Scholar] [CrossRef]
  104. Fiorio, J.L.; Garcia, M.A.S.; Gothe, M.L.; Galvan, D.; Troise, P.C.; Conte-Junior, C.A.; Vidinha, P.; Camargo, P.H.C.; Rossi, L.M. Recent advances in the use of nitrogen-doped carbon materials for the design of noble metal catalysts. Coord. Chem. Rev. 2023, 481, 215053. [Google Scholar] [CrossRef]
  105. Rossi, L.M.; Fiorio, J.L.; Garcia, M.A.S.; Ferraz, C.P. The role and fate of capping ligands in colloidally prepared metal nanoparticle catalysts. Dalton Trans. 2018, 47, 5889–5915. [Google Scholar] [CrossRef] [PubMed]
  106. Feng, X.; Bai, Y.; Liu, M.; Li, Y.; Yang, H.; Wang, X.; Wu, C. Untangling the respective effects of heteroatom-doped carbon materials in batteries, supercapacitors and the ORR to design high performance materials. Energy Environ. Sci. 2021, 14, 2036–2089. [Google Scholar] [CrossRef]
  107. Zhang, P.; Zhu, H.; Dai, S. ChemInform Abstract: Porous Carbon Supports: Recent Advances with Various Morphologies and Compositions. ChemCatChem 2015, 7, 2788–2805. [Google Scholar] [CrossRef]
  108. Yang, N.; Li, L.; Li, J.; Wei, Z. Modifying the sensibility of nonmetal-doped phosphorene by local or global properties. Phys. Chem. Chem. Phys. 2019, 21, 4899–4906. [Google Scholar] [CrossRef] [PubMed]
  109. Cheng, X.; Shen, Z.; Jiao, L.; Yang, L.; Wang, X.; Wu, Q.; Hu, Z. Tuning metal catalysts via nitrogen-doped nanocarbons for energy chemistry: From metal nanoparticles to single metal sites. EnergyChem 2021, 3, 100066. [Google Scholar] [CrossRef]
  110. Shen, K.; Chen, X.; Chen, J.; Li, Y. Development of MOF-Derived Carbon-Based Nanomaterials for Efficient Catalysis. ACS Catal. 2016, 6, 5887–5903. [Google Scholar] [CrossRef]
  111. Xu, G.; Li, H.; Bati, A.S.R.; Bat-Erdene, M.; Nine, M.J.; Losic, D.; Chen, Y.; Shapter, J.G.; Batmunkh, M.; Ma, T. Nitrogen-doped phosphorene for electrocatalytic ammonia synthesis. J. Mater. Chem. A 2020, 8, 15875–15883. [Google Scholar] [CrossRef]
  112. Batmunkh, M.; Myekhlai, M.; Bati, A.S.R.; Sahlos, S.; Slattery, A.D.; Benedetti, T.M.; Gonçales, V.R.; Gibson, C.T.; Gooding, J.J.; Tilley, R.D.; et al. Microwave-assisted synthesis of black phosphorus quantum dots: Efficient electrocatalyst for oxygen evolution reaction. J. Mater. Chem. A 2019, 7, 12974–12978. [Google Scholar] [CrossRef]
  113. Gayen, P.; Spataro, J.; Avasarala, S.; Ali, A.-M.; Cerrato, J.M.; Chaplin, B.P. Electrocatalytic Reduction of Nitrate Using Magnéli Phase TiO2 Reactive Electrochemical Membranes Doped with Pd-Based Catalysts. Environ. Sci. Technol. 2018, 52, 9370–9379. [Google Scholar] [CrossRef] [PubMed]
  114. Chaplin, B.P.; Shapley, J.R.; Werth, C.J. The Selectivity and Sustainability of a Pd–In/γ-Al2O3 Catalyst in a Packed-Bed Reactor: The Effect of Solution Composition. Catal. Lett. 2009, 130, 56–62. [Google Scholar] [CrossRef]
  115. Machida, M.; Sato, K.; Ishibashi, I.; Hasnat, M.A.; Ikeue, K. Electrocatalytic nitrate hydrogenation over an H+-conducting solid polymer electrolyte membrane–modified cathode assembly. Chem. Commun. 2006, 7, 732–734. [Google Scholar] [CrossRef] [PubMed]
  116. Hasnat, M.A.; Karim, M.R.; Machida, M. Electrocatalytic ammonia synthesis: Role of cathode materials and reactor configuration. Catal. Commun. 2009, 10, 1975–1979. [Google Scholar] [CrossRef]
  117. Gao, Z.; Lai, Y.; Tao, Y.; Xiao, L.; Zhang, L.; Luo, F. Constructing Well-Defined and Robust Th-MOF-Supported Single-Site Copper for Production and Storage of Ammonia from Electroreduction of Nitrate. ACS Cent. Sci. 2021, 7, 1066–1072. [Google Scholar] [CrossRef] [PubMed]
  118. Wu, Z.-Y.; Karamad, M.; Yong, X.; Huang, Q.; Cullen, D.A.; Zhu, P.; Xia, C.; Xiao, Q.; Shakouri, M.; Chen, F.-Y.; et al. Electrochemical ammonia synthesis via nitrate reduction on Fe single atom catalyst. Nat. Commun. 2021, 12, 2870. [Google Scholar] [CrossRef] [PubMed]
  119. Wang, Y.; Yu, Y.; Jia, R.; Zhang, C.; Zhang, B. Electrochemical synthesis of nitric acid from air and ammonia through waste utilization. Natl. Sci. Rev. 2019, 6, 730–738. [Google Scholar] [CrossRef] [PubMed]
  120. Guo, Y.; Cai, X.; Shen, S.; Wang, G.; Zhang, J. Computational prediction and experimental evaluation of nitrate reduction to ammonia on rhodium. J. Catal. 2021, 402, 1–9. [Google Scholar] [CrossRef]
  121. Sun, J.; Gao, W.; Fei, H.; Zhao, G. Efficient and selective electrochemical reduction of nitrate to N2 by relay catalytic effects of Fe-Ni bimetallic sites on MOF-derived structure. Appl. Catal. B Environ. 2022, 301, 120829. [Google Scholar] [CrossRef]
  122. Qin, J.; Wu, K.; Chen, L.; Wang, X.; Zhao, Q.; Liu, B.; Ye, Z. Achieving high selectivity for nitrate electrochemical reduction to ammonia over MOF-supported RuxOy clusters. J. Mater. Chem. A 2022, 10, 3963–3969. [Google Scholar] [CrossRef]
  123. Jiang, M.; Su, J.; Song, X.; Zhang, P.; Zhu, M.; Qin, L.; Tie, Z.; Zuo, J.-L.; Jin, Z. Interfacial Reduction Nucleation of Noble Metal Nanodots on Redox-Active Metal–Organic Frameworks for High-Efficiency Electrocatalytic Conversion of Nitrate to Ammonia. Nano Lett. 2022, 22, 2529–2537. [Google Scholar] [CrossRef] [PubMed]
  124. Lim, J.; Liu, C.-Y.; Park, J.; Liu, Y.-H.; Senftle, T.P.; Lee, S.W.; Hatzell, M.C. Structure Sensitivity of Pd Facets for Enhanced Electrochemical Nitrate Reduction to Ammonia. ACS Catal. 2021, 11, 7568–7577. [Google Scholar] [CrossRef]
  125. Chen, G.-F.; Yuan, Y.; Jiang, H.; Ren, S.-Y.; Ding, L.-X.; Ma, L.; Wu, T.; Lu, J.; Wang, H. Electrochemical reduction of nitrate to ammonia via direct eight-electron transfer using a copper–molecular solid catalyst. Nat. Energy 2020, 5, 605–613. [Google Scholar] [CrossRef]
  126. Luo, Y.; Jiang, H.; Ding, L.-X.; Chen, S.; Zou, Y.; Chen, G.-F.; Wang, H. Selective Synthesis of Either Nitric Acid or Ammonia from Air by Electrolyte Regulation in a Plasma Electrolytic System. ACS Sustain. Chem. Eng. 2023, 11, 11737–11744. [Google Scholar] [CrossRef]
  127. Kong, J.; Kim, H.; Park, H.S. Electrochemical NH3 production: In-situ evaluation of the activity and durability of nitrogen-reduction catalysis using scanning electrochemical microscopy (SECM). Appl. Catal. B Environ. 2023, 338, 123019. [Google Scholar] [CrossRef]
  128. Wang, J.; Ling, L.; Deng, Z.; Zhang, W.-x. Nitrogen-doped iron for selective catalytic reduction of nitrate to dinitrogen. Sci. Bull. 2020, 65, 926–933. [Google Scholar] [CrossRef] [PubMed]
  129. Fan, S.; Zhao, F.; Wang, X.; Wang, Q.; Zhao, Q.; Li, J.; Liu, G. A phosphorus-doped potassium peroxyniobate electrocatalyst with enriched oxygen vacancies boosts electrocatalytic nitrogen reduction to ammonia. Dalton Trans. 2022, 51, 11163–11168. [Google Scholar] [CrossRef] [PubMed]
  130. Gao, J.; Jiang, B.; Ni, C.; Qi, Y.; Bi, X. Enhanced reduction of nitrate by noble metal-free electrocatalysis on P doped three-dimensional Co3O4 cathode: Mechanism exploration from both experimental and DFT studies. Chem. Eng. J. 2020, 382, 123034. [Google Scholar] [CrossRef]
  131. Zhang, N.; Zhang, G.; Tian, Y.; Guo, Y.; Chu, K. Boron phosphide as an efficient metal-free catalyst for nitrate electroreduction to ammonia. Dalton Trans. 2023, 52, 4290–4295. [Google Scholar] [CrossRef]
  132. Shi, L.; Yin, Y.; Wang, S.; Sun, H. Rational Catalyst Design for N2 Reduction under Ambient Conditions: Strategies toward Enhanced Conversion Efficiency. ACS Catal. 2020, 10, 6870–6899. [Google Scholar] [CrossRef]
  133. Mushtaq, M.A.; Arif, M.; Yasin, G.; Tabish, M.; Kumar, A.; Ibraheem, S.; Ye, W.; Ajmal, S.; Zhao, J.; Li, P.; et al. Recent developments in heterogeneous electrocatalysts for ambient nitrogen reduction to ammonia: Activity, challenges, and future perspectives. Renew. Sustain. Energy Rev. 2023, 176, 113197. [Google Scholar] [CrossRef]
  134. Rehman, F.; Delowar Hossain, M.; Tyagi, A.; Lu, D.; Yuan, B.; Luo, Z. Engineering electrocatalyst for low-temperature N2 reduction to ammonia. Mater. Today 2021, 44, 136–167. [Google Scholar] [CrossRef]
  135. Xia, L.; Yang, J.; Wang, H.; Zhao, R.; Chen, H.; Fang, W.; Asiri, A.M.; Xie, F.; Cui, G.; Sun, X. Sulfur-doped graphene for efficient electrocatalytic N2-to-NH3 fixation. Chem. Commun. 2019, 55, 3371–3374. [Google Scholar] [CrossRef] [PubMed]
  136. Qin, Y.; Zhang, S.; Gao, G.; Ding, S.; Su, Y. A theoretical study on molybdenum and sulfur co-doped graphene for electrocatalytic nitrogen reduction. Mol. Catal. 2022, 517, 112048. [Google Scholar] [CrossRef]
  137. Ruan, Y.; He, Z.-H.; Liu, Z.-T.; Wang, W.; Hao, L.; Xu, L.; Robertson, A.W.; Sun, Z. Emerging two-dimensional materials for the electrocatalytic nitrogen reduction reaction to yield ammonia. J. Mater. Chem. A 2023, 11, 22590–22607. [Google Scholar] [CrossRef]
  138. Braun, A.; Bora, D.K.; Lauterbach, L.; Lettau, E.; Wang, H.; Cramer, S.P.; Yang, F.; Guo, J. From inert gas to fertilizer, fuel and fine chemicals: N2 reduction and fixation. Catal. Today 2022, 387, 186–196. [Google Scholar] [CrossRef] [PubMed]
  139. Zhao, X.; Hu, G.; Chen, G.F.; Zhang, H.; Zhang, S.; Wang, H. Comprehensive Understanding of the Thriving Ambient Electrochemical Nitrogen Reduction Reaction. Adv. Mater. 2021, 33, 2007650. [Google Scholar] [CrossRef] [PubMed]
  140. Zhang, L.; Chen, G.F.; Ding, L.X.; Wang, H. Advanced Non-metallic Catalysts for Electrochemical Nitrogen Reduction under Ambient Conditions. Chem.–A Eur. J. 2019, 25, 12464–12485. [Google Scholar] [CrossRef] [PubMed]
  141. Guo, J.H.; Zhang, J.G.; Zhao, H.Q.; Fang, Y.S.; Ming, K.; Huang, H.; Chen, J.M.; Wang, X.C. Fluorine-doped graphene with an outstanding electrocatalytic performance for efficient oxygen reduction reaction in alkaline solution. R. Soc. Open Sci. 2018, 5, 180925. [Google Scholar] [CrossRef]
  142. Gu, J.; Guo, X.; Xia, W.; Peng, P.; Du, F.; Li, L. Facile synthesis of fluorinated graphene for surface self-assembly of aluminum hydride. FirePhysChem 2023, 3, 201–207. [Google Scholar] [CrossRef]
  143. Insausti, M.; Timmis, R.; Kinnersley, R.; Rufino, M.C. Advances in sensing ammonia from agricultural sources. Sci. Total Environ. 2020, 706, 135124. [Google Scholar] [CrossRef] [PubMed]
  144. Sanetuntikul, J.; Hyun, S.; Ganesan, P.; Shanmugam, S. Cobalt and nitrogen co-doped hierarchically porous carbon nanostructure: A bifunctional electrocatalyst for oxygen reduction and evolution reactions. J. Mater. Chem. A 2018, 6, 24078–24085. [Google Scholar] [CrossRef]
  145. Singh, K.; Razmjooei, F.; Yu, J.-S. Active sites and factors influencing them for efficient oxygen reduction reaction in metal-N coordinated pyrolyzed and non-pyrolyzed catalysts: A review. J. Mater. Chem. A 2017, 5, 20095–20119. [Google Scholar] [CrossRef]
  146. Palma, V.; Cortese, M.; Renda, S.; Ruocco, C.; Martino, M.; Meloni, E. A Review about the Recent Advances in Selected NonThermal Plasma Assisted Solid–Gas Phase Chemical Processes. Nanomaterials 2020, 10, 1596. [Google Scholar] [CrossRef] [PubMed]
  147. Ngorot Kembo, J.P.; Wang, J.; Luo, N.; Gao, F.; Yi, H.; Zhao, S.; Zhou, Y.; Tang, X. A review of catalytic oxidation of carbon monoxide over different catalysts emphasis on Hopcalite Catalysts. New J. Chem. 2023, 47, 20222–20247. [Google Scholar] [CrossRef]
  148. Shetty, A.; Molahalli, V.; Sharma, A.; Hegde, G. Biomass-Derived Carbon Materials in Heterogeneous Catalysis: A Step towards Sustainable Future. Catalysts 2022, 13, 20. [Google Scholar] [CrossRef]
  149. Tang, Z.-E.; Lim, S.; Pang, Y.-L.; Ong, H.-C.; Lee, K.-T. Synthesis of biomass as heterogeneous catalyst for application in biodiesel production: State of the art and fundamental review. Renew. Sustain. Energy Rev. 2018, 92, 235–253. [Google Scholar] [CrossRef]
  150. Wang, Y.; Zhang, M.; Shen, X.; Wang, H.; Wang, H.; Xia, K.; Yin, Z.; Zhang, Y. Biomass-Derived Carbon Materials: Controllable Preparation and Versatile Applications. Small 2021, 17, 2008079. [Google Scholar] [CrossRef] [PubMed]
  151. Kaur, P.; Verma, G.; Sekhon, S.S. Biomass derived hierarchical porous carbon materials as oxygen reduction reaction electrocatalysts in fuel cells. Prog. Mater. Sci. 2019, 102, 1–71. [Google Scholar] [CrossRef]
  152. Zhou, L.; Cao, H.; Zhu, S.; Hou, L.; Yuan, C. Hierarchical micro-/mesoporous N- and O-enriched carbon derived from disposable cashmere: A competitive cost-effective material for high-performance electrochemical capacitors. Green Chem. 2015, 17, 2373–2382. [Google Scholar] [CrossRef]
  153. Wu, T.; Li, P.; Wang, H.; Zhao, R.; Zhou, Q.; Kong, W.; Liu, M.; Zhang, Y.; Sun, X.; Gong, F. Biomass-derived oxygen-doped hollow carbon microtubes for electrocatalytic N2-to-NH3 fixation under ambient conditions. Chem. Commun. 2019, 55, 2684–2687. [Google Scholar] [CrossRef] [PubMed]
  154. He, H.; Zhang, R.; Zhang, P.; Wang, P.; Chen, N.; Qian, B.; Zhang, L.; Yu, J.; Dai, B. Functional Carbon from Nature: Biomass-Derived Carbon Materials and the Recent Progress of Their Applications. Adv. Sci. 2023, 10, 2205557. [Google Scholar] [CrossRef] [PubMed]
  155. Hu, H.; Yan, M.; Jiang, J.; Huang, A.; Cai, S.; Lan, L.; Ye, K.; Chen, D.; Tang, K.; Zuo, Q.; et al. A state-of-the-art review on biomass-derived carbon materials for supercapacitor applications: From precursor selection to design optimization. Sci. Total Environ. 2024, 912, 169141. [Google Scholar] [CrossRef]
  156. Lesbayev, B.; Auyelkhankyzy, M.; Ustayeva, G.; Yeleuov, M.; Rakhymzhan, N.; Maltay, A.; Maral, Y. Recent advances: Biomass-derived porous carbon materials. S. Afr. J. Chem. Eng. 2023, 43, 327–336. [Google Scholar] [CrossRef]
  157. Wang, L.; Wang, T.; Hao, R.; Wang, Y. Synthesis and applications of biomass-derived porous carbon materials in energy utilization and environmental remediation. Chemosphere 2023, 339, 139635. [Google Scholar] [CrossRef] [PubMed]
  158. Priya, D.S.; Kennedy, L.J.; Anand, G.T. Emerging trends in biomass-derived porous carbon materials for energy storage application: A critical review. Mater. Today Sustain. 2023, 21, 100320. [Google Scholar] [CrossRef]
  159. Paramasivam, N.; Sambandam, A.; Nastesan, B. Metalloids (B, Si) and non-metal (N, P, S) doped graphene nanosheet as a supercapacitor electrode: A density functional theory study. Mater. Today Commun. 2023, 35, 105905. [Google Scholar] [CrossRef]
  160. Inagaki, M.; Toyoda, M.; Soneda, Y.; Morishita, T. Nitrogen-doped carbon materials. Carbon 2018, 132, 104–140. [Google Scholar] [CrossRef]
  161. Huang, Z.; Liao, Z.; Yang, W.; Zhou, H.; Fu, C.; Gong, Y.; Chen, L.; Kuang, Y. Different types of nitrogen species in nitrogen-doped carbon material: The formation mechanism and catalytic role on oxygen reduction reaction. Electrochim. Acta 2017, 245, 957–966. [Google Scholar] [CrossRef]
  162. Ma, G.; Ning, G.; Wei, Q. S-doped carbon materials: Synthesis, properties and applications. Carbon 2022, 195, 328–340. [Google Scholar] [CrossRef]
  163. Wang, M.J.; Wang, L.; Li, Q.; Bai, Y.; Wang, D.; Feng, Y. Controllable Synthesis of N-/S-Doped and N,S-doped hollow carbon spheres for the oxygen reduction Reaction: A universal Mono-Micelle Self-Assembly strategy. Mater. Lett. 2022, 309, 131315. [Google Scholar] [CrossRef]
  164. Sudhanshu, R. X-ray Photoelectron Spectroscopy (XPS) Technology; Springer: Berlin/Heidelberg, Germany, 2020. [Google Scholar]
  165. Krishna, D.N.G.; Philip, J. Review on surface-characterization applications of X-ray photoelectron spectroscopy (XPS): Recent developments and challenges. Appl. Surf. Sci. Adv. 2022, 12, 100332. [Google Scholar] [CrossRef]
  166. Li, Q.; Chen, X.; Yang, Y. Biomass-Derived Nitrogen-Doped Porous Carbon for Highly Efficient Ambient Electro-Synthesis of NH3. Catalysts 2020, 10, 353. [Google Scholar] [CrossRef]
  167. Giddey, S.; Badwal, S.P.S.; Kulkarni, A. Review of electrochemical ammonia production technologies and materials. Int. J. Hydrogen Energy 2013, 38, 14576–14594. [Google Scholar] [CrossRef]
  168. Jia, Y.; Chen, J.; Yao, X. Defect electrocatalytic mechanism: Concept, topological structure and perspective. Mater. Chem. Front. 2018, 2, 1250–1268. [Google Scholar] [CrossRef]
  169. Liu, X.; Dai, L. Carbon-based metal-free catalysts. Nat. Rev. Mater. 2016, 1, 16064. [Google Scholar] [CrossRef]
  170. Zheng, Y.; Jiao, Y.; Li, L.H.; Xing, T.; Chen, Y.; Jaroniec, M.; Qiao, S.Z. Toward Design of Synergistically Active Carbon-Based Catalysts for Electrocatalytic Hydrogen Evolution. ACS Nano 2014, 8, 5290–5296. [Google Scholar] [CrossRef] [PubMed]
  171. Hu, C.; Dai, L. Carbon-Based Metal-Free Catalysts for Electrocatalysis beyond the ORR. Angew. Chem. Int. Ed. 2016, 55, 11736–11758. [Google Scholar] [CrossRef] [PubMed]
  172. Zheng, Y.; Jiao, Y.; Zhu, Y.; Li, L.H.; Han, Y.; Chen, Y.; Du, A.; Jaroniec, M.; Qiao, S.Z. Hydrogen evolution by a metal-free electrocatalyst. Nat Commun 2014, 5, 3783. [Google Scholar] [CrossRef]
  173. Yu, X.; Han, P.; Wei, Z.; Huang, L.; Gu, Z.; Peng, S.; Ma, J.; Zheng, G. Boron-Doped Graphene for Electrocatalytic N2 Reduction. Joule 2018, 2, 1610–1622. [Google Scholar] [CrossRef]
  174. Liu, Y.; Su, Y.; Quan, X.; Fan, X.; Chen, S.; Yu, H.; Zhao, H.; Zhang, Y.; Zhao, J. Facile Ammonia Synthesis from Electrocatalytic N2 Reduction under Ambient Conditions on N-Doped Porous Carbon. ACS Catal. 2018, 8, 1186–1191. [Google Scholar] [CrossRef]
  175. Huang, H.; Xia, L.; Cao, R.; Niu, Z.; Chen, H.; Liu, Q.; Li, T.; Shi, X.; Asiri, A.M.; Sun, X. A Biomass-Derived Carbon-Based Electrocatalyst for Efficient N(2) Fixation to NH(3) under Ambient Conditions. Chemistry 2019, 25, 1914–1917. [Google Scholar] [CrossRef]
Figure 1. Non-metallic-element-doped material schematic.
Figure 1. Non-metallic-element-doped material schematic.
Materials 17 02419 g001
Figure 2. (a). The first possible electrocatalytic pathway for nitrogen to produce ammonia. (b) The second possible electrocatalytic pathway for nitrogen to produce ammonia.
Figure 2. (a). The first possible electrocatalytic pathway for nitrogen to produce ammonia. (b) The second possible electrocatalytic pathway for nitrogen to produce ammonia.
Materials 17 02419 g002
Figure 3. Schematic illustration of BCDs/NiCo2O4/CC nanowire array preparation. Adapted with permission from Lu et al. [89]. Copyright 2022, ScienceDirect.
Figure 3. Schematic illustration of BCDs/NiCo2O4/CC nanowire array preparation. Adapted with permission from Lu et al. [89]. Copyright 2022, ScienceDirect.
Materials 17 02419 g003
Figure 4. Synthetic route of nitride–carbon-encapsulated N-doped iron (FeN-NC). Adapted with permission from Wang et al. [128]. Copyright 2020, Elsevier.
Figure 4. Synthetic route of nitride–carbon-encapsulated N-doped iron (FeN-NC). Adapted with permission from Wang et al. [128]. Copyright 2020, Elsevier.
Materials 17 02419 g004
Figure 5. A typical phosphorus-doped catalyst. Adapted with permission from Gao et al. [130]. Copyright 2020, ScienceDirect.
Figure 5. A typical phosphorus-doped catalyst. Adapted with permission from Gao et al. [130]. Copyright 2020, ScienceDirect.
Materials 17 02419 g005
Table 1. Summary of post-synthesis doping methods.
Table 1. Summary of post-synthesis doping methods.
Post-Synthesis Doping MethodSource of Non-MetalMaterial DopedTemperature Range (°C)TimeResultRef.
In situ electron-beam irradiationParaffin waxBoron nitride nanotubes--A well-controlled, little-damaging, simple strategy[64]
Direct nitrogen reactionNitrogen and cyanamideSingle-walled carbon nanotubes/graphene sample500–800 °C1 hImproved catalytic properties[65]
Thermal-diffusion-based dopingBoron from boric acidSilicon quantum dots400–600 °C1 hNetwork-interconnected Si particles.[66]
Ethylene amine vapor dopingEthylene amineGraphene300 °C1 hHigh-performance transparent electrodes[67]
Solvothermal methodNitrogen from melamineCarbon nanoparticles from CCl4--Low-cost alternative to commercial Pt/C catalysts[68]
Solid template approachNitrogen from melaminecore–shell SiO2@RF composites600 and 900 °C1 hHigh specific surface areas and thermal stabilities[69]
Table 2. Synthesis methods for in situ-N-doped carbon compounds.
Table 2. Synthesis methods for in situ-N-doped carbon compounds.
Synthesis TechniqueDescriptionSource/ProductRef.
CVDSimultaneous synthesis of carbon material and nitrogen doping.N-doped carbons[79]
Infiltration and subsequent CVDA novel approach involving Y zeolite infiltration with an ionic liquid and subsequent CVD. Co-doped microporous carbon exhibited pyridinic, pyrrolic, and graphitic nitrogen.Co-doped microporous carbons[70]
CVDUtilization of C2H4/NH3 combinations in CVD for creating N-doped carbon nanofibers (N-CNFs)N-CNFs[80]
CarbonizationN-doped carbon material preparation from polyaniline (PANI) through carbonization at 800 °C using various salt precursorsSynthesis from PANI[73]
CarbonizationEmploying imidazolium/pyridinium ionic liquids (ILs) for N-doped carbon synthesisN-doped carbons[81]
Electrospinning, carbonizationCreation of hollow-particle-based N-doped carbon nanofibers (N-CNFs) using PAN and ZIF-8, followed by carbonization.Hollow-particle-based N-CNFs[82]
Hydrothermal carbonization (HTC)Hydrothermal carbonization (HTC) of biomass for creating N-doped carbon compounds.Biomass-derived N-doped carbon[83]
Heat treatment, chemical activationClosed porosity and low surface area in resulting hydrochars necessitating additional treatment.Biomass-derived N-doped carbon[71]
PyrolysisConventional pyrolysis of biomass for producing N-doped carbon nanosheets without chemical activation.Advanced pyrolysis for N-doped carbon Nanosheets[84]
Table 3. Summary of synthesis methods for metal catalysts supported on N-doped carbon materials.
Table 3. Summary of synthesis methods for metal catalysts supported on N-doped carbon materials.
MethodologyDescriptionReference
Capping-agent-protected NPsUses capping agents for NP growth regulation; concerns about the negative impact on catalytic activity.[101,102]
N-doped carbon supportsAchieves small, evenly distributed NPs. Nitrogen functionalities enhance catalytic performance.[103,104]
Synthetic techniques for NPs on N-doped carbonThree categories: loading on pre-synthesized N-doped carbon, simultaneous incorporation, and in situ synthesis.[83,105]
Conventional techniques for loading NPsIncludes impregnation, deposition–precipitation, and sol immobilization with subsequent reduction.[106,107]
In situ synthesis of NPs on N-doped carbonThe pyrolysis step allows the simultaneous incorporation of metal and nitrogen.[107,108]
Use of MOF as a sacrificial templateMOFs are used as templates for intriguing metal/N–carbon compositions.[109,110]
Table 4. Summary of electrocatalysts for the reduction of NO3 to NH3.
Table 4. Summary of electrocatalysts for the reduction of NO3 to NH3.
CatalystsNH3 YieldFE (%)ElectrolyteYear PublishedReference
Pd-Cu/TinO2n−1-221 mM NaNO32018[113]
Pd–In/γ-Al2O3-71.51.4 mM NO32009[114]
Pt-493000 mg·L−1 NO32006[115]
30%Cu-70%Pd-580.05 M KNO32009[116]
Fe3Ni-N-CNitrate-to-N299.3100 ppm NO32021[117]
Single Fe atom0.46 mmol·h−1·cm−2750.50 M KNO32021[118]
Co3O4 nanorod arrays0.854 mmol·h−1·cm233.6100 g L−1 NO32019[119]
Rh/C34.4 μg h−1cm−220.80.1 M KNO32021[120]
Cu@Th-BPYDC225.3 μmolh−1cm−292.5100 mM KNO32022[121]
RuNi-MOF274 mgh−1mgcat−17350 mg L−1 NO32022[122]
Pd-NDs/Zr-MOF287.31 mmol·h−1·gcat−1.58.1500 ppm NO32022[123]
Pd cuboctahedrons306.8 μgh−1mgcat−13520 mM NO32021[124]
PTCDA/O-Cu436 μg h−1cm−285.9500 ppm NO32020[125]
Table 5. Types and characteristics of electrocatalyst for N2 synthesis of NH3.
Table 5. Types and characteristics of electrocatalyst for N2 synthesis of NH3.
Catalyst Type/ProcessCharacteristics Challenges and LimitationsRef.
Metallic catalysts (Ru, Rh)Ru electroplated on titanium felt enhances NH3 production; Ru on porous graphite improves NRRCatalytic structure sensitivity; cost and scarcity of precious metals[139,140]
Monatomic catalystsStrong chemical interaction with nitrogen; outstanding activityHigh cost and scarcity for large-scale application[143,144]
Non-precious metal catalystsLow Faraday efficiency; activation via sp hybrid orbitalsLimited Faraday efficiency despite low cost[145]
Metal-free catalysts (N, P)Carbon-based compounds doped with N and P; PCN-NVs increase ammonia yield and Faraday efficiencyCost and scarcity challenges for large-scale use[78]
Perovskite catalystsLaFeO3 effective in oxidizing N2; NTP enhances current density; Cu/CuO catalyst reduces NOx−Gibbs free energy spectrum investigation; challenges in N2 activation[146,147]
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

Quoie Jr, G.D.S.; Jiao, M.; Lászlód, K.; Wang, Y. Progress Made in Non-Metallic-Doped Materials for Electrocatalytic Reduction in Ammonia Production. Materials 2024, 17, 2419. https://doi.org/10.3390/ma17102419

AMA Style

Quoie Jr GDS, Jiao M, Lászlód K, Wang Y. Progress Made in Non-Metallic-Doped Materials for Electrocatalytic Reduction in Ammonia Production. Materials. 2024; 17(10):2419. https://doi.org/10.3390/ma17102419

Chicago/Turabian Style

Quoie Jr, Gerald D. S., Mingshuo Jiao, Krisztina Lászlód, and Ying Wang. 2024. "Progress Made in Non-Metallic-Doped Materials for Electrocatalytic Reduction in Ammonia Production" Materials 17, no. 10: 2419. https://doi.org/10.3390/ma17102419

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