Download PDF
Review  |  Open Access  |  14 Mar 2024

Recent advances in 2D structured materials with defect-exploiting design strategies for electrocatalysis of nitrate to ammonia

Views: 593 |  Downloads: 57 |  Cited:   0
Energy Mater 2024;4:400020.
10.20517/energymater.2023.67 |  © The Author(s) 2024.
Author Information
Article Notes
Cite This Article

Abstract

Ammonia has been used in a wide variety of applications, and with the recent interest in hydrogen energy as a green energy source, it is emerging as a cost-effective, high-density hydrogen carrier due to its three hydrogen atoms. Currently, ammonia is produced by the Haber-Bosch method at high temperatures and pressure, which is energy-intensive and emits large amounts of carbon dioxide. As a viable alternative, the electrochemical conversion of nitrate to ammonia has emerged as an efficient and eco-friendly synthesis method. To encourage further exploration in this field, this review offers insights into utilizing two-dimensional materials as electrochemical catalysts, focusing on designs that exploit defects for nitrate reduction to ammonia.

Keywords

Nitrate reduction reaction, ammonia synthesis, 2D electrocatalyst, defect engineering

INTRODUCTION

As the world transitions to a renewable energy and hydrogen (H2) economy to achieve a carbon-neutral society, ammonia is gaining traction as a hydrogen carrier[1,2] and carbon-free fuel[3,4] for hard-to-decarbonize sectors such as power generation[5,6] and transportation[7,8]. Unlike hydrogen, which liquefies at around -253 °C, ammonia liquefies at -34 °C, making it easy to store and transport [Table 1][9,10]. Globally, about 180 million tons of ammonia are produced yearly[11], of which 80% is used as a fertilizer for agriculture, and the remaining 20% is used in explosives, pharmaceuticals, and more[12]. Ammonia is produced under high temperature and pressure conditions (350-550 °C, 150-250 bar) via the Haber-Bosch process, which is an energy-intensive procedure that requires 1%-2% of the total energy produced by humans on the earth[13] [Figure 1A]. In addition, the amount of methane consumed in the production process is about 3.5 times higher, and more than 400 million tons of CO2 is emitted annually to produce ammonia, which can cause environmental problems[14].

Recent advances in 2D structured materials with defect-exploiting design strategies for electrocatalysis of nitrate to ammonia

Figure 1. Schematic diagrams of ammonia synthesis. (A) The Haber-Bosch process (Reprinted with permission from Ref.[27]. Copyright 2019 Royal Society of Chemistry)[27] and (B) The electrochemical synthesis of ammonia process.

Table 1

Comparison of liquid hydrogen and liquid ammonia

PropertiesUnitLiquid hydrogenLiquid ammoniaRef.
Densitykg/m370.6682[13]
Transported temperature, pressure°C, bar-253, 1.2 bar-33, 1 bar[13]
The gravimetric energy density (LHV)MJ/kg12018.6[14]
Volumetric energy density (LHV)MJ/L8.4912.7[14]
Gravimetric hydrogen contentwt%10017.8[14]
Volumetric hydrogen contentkg H2/m370.8121[14]
Explosive limit in airvol%18.3-5916-25-
Storage methodLiquefactionLiquefaction-

Electrochemical ammonia production has been proposed as an alternative to the Haber-Bosch process and is gaining traction due to several advantages [Figure 1B]. First, it is energy efficient, as ammonia can be produced under ambient conditions, as opposed to the high temperature and pressure conditions of the Haber-Bosch method[15]; Second, it is environmentally friendly as it can be combined with renewable energy sources (solar, tidal, and wind) with low carbon emissions[16]; Third, it uses nitrogen and water as reactants, eliminating the use of fossil fuels as a source of H2, with the required protons (H+) generated on-site through the oxidation of water[17]. Finally, the process has advantages such as flexible reaction control[18], scalability, and on-demand on-site NH3 production[19]. Electrochemical ammonia synthesis is mainly based on nitrogen reduction reactions (NRR)[20,21], and although many preliminary results have been obtained, the low efficiency has limited its commercialization[22]. To overcome this limitation, electrochemical ammonia synthesis using the ionic states of nitrogen oxides has recently been investigated[23-27].

From the nitrogen cycle, nitrogen oxides are stable forms under normal aerobic conditions[28]; however, as the accumulation of nitrogen oxides increases due to industrial activities, it causes health problems such as liver damage[29] and blue baby[30]. To restore the nitrogen cycle, the World Health Organization (WHO) stated that the maximum tolerable concentration of NO3- in water and wastewater is 50 mg/L[31]. For the removal of nitrogen oxides, physical removal methods are still used, such as reverse osmosis[32] or ion exchange[33]; however, these costs are high, about $1-2 million per year per city (based on a population of 500,000 in the US)[34].

Recently, nitrate reduction reaction (NitRR), also known as NO3RR, has been studied due to its ability to synthesize ammonia and remove pollutants simultaneously[35,36]. In addition, following advantages over nitrogen, the solubility of NO2- or NO3- is about 40,000 times that of N2 gas, and the dissociation of N-O (204 kJ/mol) compared to N≡N (945 kJ/mol) is much easier, making NitRR more economical than NRR[37,38]. As a result, the total conversion and energy efficiency of NitRR is much higher than NRR, and its performance can be further improved by strategic catalyst design. NitRR mechanisms involve nine protons, eight electrons, and reaction pathways depending on the reaction conditions[39]. Due to their complex pathways, NitRR to ammonia has a variety of by-products, such as N2, NO, and NO2-, which increases the difficulty of analyzing the mechanism[40]. Various papers have revealed the mechanism, selectivity, and activity of NitRR. In previous studies, the rate-determining step (RDS) of NitRR is generally known to be NO3- to NO2- on transition metal (TM) surfaces and under acidic conditions[41]. The RDS can be seen in Figure 2. To increase the efficiency of the overall NitRR, it is necessary to develop a catalyst that facilitates the adsorption/desorption of NO3-. The metal-based catalysts, such as Pt[42], Pd[43], Ag[44], Sn[45], and Cu-based catalysts[46-48], are considered universal selections in the electrocatalytic process to lower the NitRR overpotential due to their high activity and appropriate adsorption energy of nitrate[49]. These are verified theoretically by the online differential electrochemical mass spectrometry (DEMS) and density functional theory (DFT) calculation[43,46,48].

Recent advances in 2D structured materials with defect-exploiting design strategies for electrocatalysis of nitrate to ammonia

Figure 2. Schematic illustrations of the NitRR mechanisms (Reprinted with permission from Ref.[49]. Copyright 2022 Wiley-VCH GmbH)[49].

Over the past decade, there has been increasing interest in atomistic, low-cost two-dimensional (2D) nanosheet materials with unique physicochemical properties, such as high specific surface area, desirable morphology with myriad binding sites, and excellent reaction rates due to fast ion diffusion pathways[50]. In particular, 2D nanosheets are promising materials for catalysis due to their high surface efficiency[51], which can be exposed to active sites. Many studies have been conducted on layered double hydroxides (LDH)[52,53], chalcogenides[54,55], MXenes[56,57], metal-organic frameworks (MOFs)[58,59], etc.[60-62]. These 2D materials have found applications as catalysts or substrates, leveraging their intrinsic properties such as conductivity and activity for NitRR. Certain 2D materials, such as graphene, show limited electrocatalytic properties for NitRR; however, they have high conductivity, rendering them more suitable as substrates than active materials. This enhances conductivity and creates a synergy effect when used with functional materials.

To synthesize efficient electrocatalysts, porosity engineering is critical in carrying abundant active sites and shortened channels through meso-/macropores for mass/charge transport[63]. Also, the oxygen vacancy (OV)[64,65] in oxides is used as a powerful strategy to regulate the crystalline structure, electronic structure, and surface properties. In addition, doping heteroatoms efficiently modifies the electron structure of electrocatalysts to decrease the value of GH*∆ of the electrocatalysts and to improve electrochemical activity[66]. It can make synergistic effects to boost mass transfer on 2D nanosheets and increase active sites through the multiscale defect. Moreover, the adjustment of interlayer spacing can not only tune the atomic distance and electron interaction but also manipulate the electronic structure of host materials, promising higher electrical conductivity[54]. Therefore, design strategies for electrocatalysts make it very possible to modulate the binding strength of intermediates, such as NO2-, and the competitive hydrogen evolution reaction (HER) process in NitRR.

In this review, we provide an overview of 2D materials as electrochemical catalysts and report the latest results of applying defects, such as doping, OVs, and topological defects, to enhance their activity as electrocatalysts in the electrochemical synthesis of ammonia. We also present our views on future research directions for ultrathin 2D materials to encourage further active research.

2D STRUCTURE ELECTROCATALYSTS: DEFECT ENGINEERING

Layered double hydroxides

The earth-abundant TM catalysts have been focused on overcoming the disadvantages of noble metal catalysts. However, the TMs quickly foam M-H bonds, leading to low Faradaic efficiency (FE) and poor selectivity of NH3[67,68]. High HER performance makes NitRR performance soft because HER is a competitive reaction with NitRR. Moreover, another problem exists in that NO3- has an innately lower binding affinity with TMs in the aqueous solution because of strong hydrogen bonding and its symmetrical resonant structure[47]. Recently, TM-based LDHs (TM LDHs, [M1-x2+x3+(OH)2]x+(An-)x/n·mH2O), sometimes called Hydrotalcite-like compounds (HTLCs), have been focused on NitRR fields. The TM LDH consists of divalent (M2+: e.g., Ni2+, Mg2+, or Zn2+) and trivalent (M´3+, e.g., Mn3+, Fe3+, and Al3+) metal cations with interlayer exchangeable inorganic or organic anions (An-: e.g., NO3-, SO42-, CO32-, and Cl-)[69]. TM LDH catalysts have many advantages, such as large surface area, active site, low cost, easily tunable electric structure, and chemical composition[70]. The various synthesis methods of TM LDHs have been researched because of their unique 2D structure. The exfoliation method belongs to the top-down method and is usually used to make 2D materials, including TM LDH nanosheets. Other top-down synthesis methods have been studied, such as plasma etching and liquid exfoliation. Otherwise, bottom-up methods also have attention owing to their relatively simple process. Microemulsion, Coprecipitation, and hydrothermal processes were representative bottom-up methods to synthesize TM LDH nanosheets[71-73]. TM LDH catalysts could be easily synthesized through a one-step or two-step hydrothermal process[70].

TM LDH catalysts can easily apply reforming methods to improve performance due to their intrinsic 2D structure. Distortion is a frequently used method in 2D materials, leading to a distorted surface, which can cause more exposure of the edge site to enhance electrocatalytic performance. OVs are one of the representative defect engineering methods used to improve the electrochemical properties in the TM LDH catalysts by boosting the charge transfer/separation and optimizing the reaction kinetics. These vacancies also make TM LDH catalysts higher electrical conductivity and increase adsorption for oxygen-based intermediates[74]. Recently, the hybrid TM LDH-based catalysts for NitRR, including two kinds of metal, have been studied in most research. These mixed TM LDH catalysts show good reactive performance due to synergistic effects. Du et al. efficiently synthesized the CoFe LDH catalysts with the hydrothermal method in Figure 3A, and the catalyst performed excellent FE, NH3 selectivity, high activity, and durability in alkaline conditions[52]. The DFT calculations indicated that CoFe LDH has a downhill reaction step (NO3- → *NO3- → *NO2-) free energy and a moderate desorption step (*NH3 → NH3) free energy compared to that of Co LDH and Fe LDH [Figure 3B]. NH3 selectivity and FE of CoFe LDH shows 98.93% and 97.68% at -0.45 V (vs. reversible hydrogen electrode (RHE)) in 1 M KOH and 1,400 ppm NO3- electrolytes [Figure 3C]. The operation durability of performance of the CoFe LDH in NitRR maintained for 12 cycles with 36 h in Figure 3D. To increase the conductivity of TM LDH catalysts, TM LDH-decorated Cu foam electrodes have been studied by Kim et al.[53]. The Ni3Fe-CO3 LDH/Cu achieved the NH3 selectivity of 95.8%, the NH3 productivity of 1.261 mg h-1 cm-2, and FE of 96.8% at -0.2 V (vs. RHE) in 1 M KOH electrolytes with 5 mM KNO3 [Figure 3E and F]. Various multi-TM LDH/Cu foams were used as electrocatalysts to compare the electrocatalyst performance. Among them, Ni3Fe-CO3 LDH/Cu foam shows the highest electrocatalytic performance [Figure 3G]. The catalyst suppresses H-H bond formation by hindering the Heyrovsky step (Hads* + H2O + e- → H2 + OH) relating to the Volmer step (H2O + e- → Hads* + OH-). Ternary TM LDH catalysts have been studied to apply complementary effects. Li et al. investigated Cu2CoAl LDH catalyst for NitRR [Figure 3H][75]. The Cu/Al LDHs show lower structural stability than Co/Al LDH because of the Jahn-Teller effect. However, the Cu-based catalyst has a higher electrocatalytic activity for NitRR than the Co-based catalyst. Cu2CoAl LDH was synthesized to utilize each advantage, and the catalyst shows an NH3 conversion of 83% at -0.74 V (vs. Hg/HgO) in 1 M NaOH electrolytes with 0.1 M NaNO3 in Figure 3I. Therefore, research into TM LDHs as electrocatalysts has significantly advanced owing to their large active area and cost-effective manufacturing. They can be applied concurrently with defect engineering during hydrothermal processes. Thus, due to their low manufacturing cost, a simple synthesis system holds high potential for commercializing the TM LDH catalyst in the industry.

Recent advances in 2D structured materials with defect-exploiting design strategies for electrocatalysis of nitrate to ammonia

Figure 3. (A) Schematic image of the synthesis process for CoFe LDH and its cell composition for NitRR. (B) Reaction-free energies for *NO3, *NO2, and *NH3 on the Fe LH, Co LH, and CoFe LDH surfaces. (C) FE and NH3 selectivity of CoFe LDH according to potential. (D) The consecutive recycling test of CoFe LDH at -0.45 V (vs. RHE). (A-D) Figures are reprinted with permission from Ref.[52]. Copyright 2022 Elsevier[52]. (E) Comparison of NH3 yield rate and FE of Ni3Fe-CO3 LDH with reference materials. (F) The recycling test of Ni3Fe-CO3 LDH/Cu foam electrode at -0.2 V (vs. RHE). (G) NH3 yield rate and FE using multi-metallic TM LDH/Cu foam electrodes. (E-G) figures are reprinted with permission from Ref.[53]. Copyright 2023 The Royal Society of Chemistry[53]. (H) XRD patterns of Cu/Co/Al-HTLCs. (I) The transformation of NO3 and total nitrogen (TN) on Cu/Co/Al-HTLCs according to time at -0.74 V (vs. RHE). (H and I) Figures are reprinted with permission from Ref.[75]. Copyright 2020 MDPI[75].

Chalcogenides

The low-cost TM chalcogenide materials have been considered to replace the high-cost precious metal catalysts. A TM chalcogenide is a chemical compound of at least one TM cation and at least one more chalcogen anion (S, Se, and Te). Among the TM chalcogenides, TM dichalcogenides (TMDs), such as metal disulfides (MS2) and diselenides (MSe2), with M4+ (Mo4+, Ti4+, and V4+) have been widely used in various electrochemical applications due to the unique physicochemical and structural features[76]. TMDs have unique layered structures, and one of them, TiS2, was investigated as a cathode material for secondary batteries because of its layered structure that can be intercalated by lithium in the 1970s[77]. In addition to these structure features, TMDs can be easily doped with various TMs to tune the electric structure and chemical composition[78].

More recently, to improve the electrocatalytic performance, various nanostructures of TMDs have been studied. Partially, 2D TMDs exhibit the potential for NitRR owing to their large active area and tunable electronic structure. These 2D TMDs can be categorized into trigonal prismatic (hexagonal, H), octahedral (tetragonal, T), and distorted phase (T´)[79]. Various defect engineering is also studied, and sulfur vacancies (SVs) are one of the representatives in the field of TMD catalysts[80].

Doping single metal atoms into 2D materials is considered one of the effective strategies to enhance the electrochemical performance of a catalyst[81,82]. Such strategies can significantly improve the electrochemical NitRR performance of a catalyst by incorporating metal dopants into 2D materials. Li et al. investigated the iron single atomic catalyst (SAC) supported on 2D MoS2 (Fe-MoS2) nanosheets for NitRR [Figure 4A][54]. Additionally, Cu-, Ni-, and Co-MoS2 nanosheets were also studied to compare with Fe-MoS2 catalysts, and Fe-MoS2 has higher electrochemical performance than that of other catalysts [Figure 4B and C]. DFT studies show that Fe-MoS2 presents the deoxidation of the *NO to *N intermediate, decreasing the energy barrier of the reaction pathway to synthesize ammonia. Ding et al. also studied Fe-MoS2 catalysts through DFT calculations for NitRR[83]. Since Fe on the MoS2 improves the activating NitRR and suppressing HER, Fe-MoS2 achieved a high FE of 90% and a yield rate of 9.75 mg h-1 cm-2 for NH3 production at -0.85 V (vs. RHE) [Figure 4D]. Figure 4E and F show the charge density difference results, where the surface of Fe-MoS2 has a lower NO3- adsorption energy (-1.92 eV) than that of the surface of MoS2 (-0.11 eV). The DFT results were calculated on the different pathways of NitRR, unlike the study above by Li et al., Still, the Fe-MoS2 reduces the energy barrier of the path, which is the protonation of *NO to *NOH intermediates in Figure 4G[54]. The 2D Cu-SnS2-x nanosheet arrays were investigated for NitRR by Li et al.[55]. Cu doping makes the abundant SVs on the SnS2 nanosheet, which increases the electrocatalytic activity of the catalyst. The 2D Cu-SnS2-x nanosheet arrays showed two times higher NitRR performance, which is a yield rate of 0.63 mmol h-1 mgcat.-1 and FE of 93.8% at -0.7 V (vs. RHE), than those of pristine SnS2 in 0.1 M KOH with 0.1 M KNO3 solution. The adsorption energy of Cu-SnS2-x (-1.6158 eV) is higher than that of SnS2-x(-2.0755 eV); however, the too-high adsorption energy may bring about difficulty in the desorption of the intermediaries according to Sabatier principle. DFT studies indicate that doped Cu stabilizes intermediates on the NitRR pathway and tunes the electron distribution. In addition, SVs enhance the adsorption capacity due to charge transfer to Sn. Thus, the TMDs have shown substantial promise for NitRR due to their advantages, including a large active area and tunable electronic structure. However, research in the NitRR field has primarily focused on a few TMD catalysts, such as MoS2, MoSe2, SnS2, and SeSe2, leaving various combinations of TMDs, including TaS2, NbS2, VSe2, and MoTe2, largely unexplored. These catalysts offer unlimited potential for NitRR. Exploring other related materials is warranted to gain a more comprehensive understanding of 2D TMDs.

Recent advances in 2D structured materials with defect-exploiting design strategies for electrocatalysis of nitrate to ammonia

Figure 4. (A) TEM image of Fe-MoS2 nanosheets on the carbon support. (B) FEs for NH3 on Cu-MoS2, Ni-MoS2, Co-MoS2, and Fe-MoS2 nanosheets at different potentials. (C) Reaction Gibbs free energies of intermediates on M-MoS2 nanosheets. (A-C) figures are reprinted with permission from Ref.[54]. Copyright 2021 Wiley-VCH GmbH[54]. (D) Potential dependent NH3 yield rate and FE of Fe-MoS2. Charge density difference results of NO3- on (E) MoS2 and (F) Fe-MoS2. (G) Reaction Gibbs free energies for NitRR on Fe-MoS2 and MoS2. (D-G) Figures are reprinted with permission from Ref.[83]. Copyright 2023 Elsevier[83].

Transition metal oxides

In the electrocatalytic field, low-cost TM oxide (TMO) catalysts have attracted attention for commercialization by replacing precious metal catalysts. TMO catalysts have many advantages, such as ease of synthesis, thermal and chemical stability, low cost, and high earth abundance to be used as catalysts[84]. However, these catalysts cannot be directly used for NitRR because TMOs have some disadvantages when used as catalysts because of their low electronic conductivity, relatively low active site, and low selectivity[85]. Optimization strategies, such as nanostructure and defect engineering, have been considered to overcome these challenges and improve electrocatalytic performance. Nanostructure engineering is a process method for making catalysts have nanoscale structures, including nanoparticles and 2D structures. These structures increase the higher active area, narrow the band gap, and promote electronic conductivity[86]. Synthesizing 2D TMOs through the top-down method has challenges, leading to multiple attempts to explore the bottom-up method. For instance, Sun et al. and Sun et al. synthesize ultrafine 2D nanosheets of TMOs[87,88]. Huang et al. demonstrated the fabrication of graphene-like TMO nanosheets by utilizing graphene oxide (GO) as a 2D template, wherein metal precursors were deposited on the GO surface and then calcined[89]. Various TMs could synthesize 2D TMOs using either a 2D template or surfactant molecules. Defect engineering, such as OVs, is commonly used to improve NitRR performance with changes in electronic structure and surface properties. Standard modification methods, such as hydrogen heat treatment, thermal annealing, wet chemical reduction, and electroreduction, can quickly generate OVs in TMOs[65].

Wang et al. studied Cu/CuOx in-plane heterostructured nanosheet arrays with OVs (Cu/CuOx/CF) for NitRR [Figure 5A][60]. This 2D nanosheet structure has rich exposing active sites and high mass/charge transfer. Moreover, OVs and in-plane heterojunctions tune electronic structures and improve adsorption capabilities and properties. The test results show that Cu/CuOx/CF achieves higher electrocatalytic performance, FE of 93.58%, and an NH3 production of 0.23 mmol h-1 cm-2 at -1.3 V (vs. saturated calomel electrode, SCE) than that of CuOx/CF, Cu/CF, and Cu Foam [Figure 5B and C]. For five cycles, the retention of the yield rate and FE of Cu/CuOx/CF was maintained at a high level in Figure 5D. Zhao et al. studied the modulation of multi-scale defects, such as OVs and nanoholes, on the Co3O4 catalyst for NitRR[61]. The defective Co3O4/Co with an interlaced nanosheet was synthesized through electrodeposition and calcination. The OVs improved the adsorption of NO3- and increased active sites. The rich nanoholes displayed in Figure 5E may make additional channels for convenient mass transfer. As shown in Figure 5F, the adsorption energy of NO3- on Co3O4-2Ov is lower than others at -1.81 eV. In addition, the energy barrier of HER on Co3O4-2Ov is the highest value of 1.15 eV [Figure 5G]. These results show that the OVs could enhance NO3- adsorption and suppress HER. Figure 5H shows that Co3O4/Co-I has a higher yield rate of 4.43 mg h-1 cm-2 and FE of 88.7% than that of Cu3O4/Co-h in 0.1 M Na2SO4 with 1 mg mL-1 of KNO3 solution. Co3O4/Co exhibits a high performance in a flow electrolyzer system, with a yield rate of 8.96 mg h-1 cm-2 at 0.1 A cm-2. Research on 2D TMOs for NitRR is widespread due to their cost-effectiveness and stability. However, their inherent drawbacks, notably low conductivity and electrocatalyst activity, pose significant challenges. Addressing this, introducing defects in TMOs compensates for their intrinsic low conductivity. These defects overcome the conductivity issue and augment the electrochemical performance of catalysts by increasing both active area and activity. Consequently, defect engineering emerges as a promising solution to tackle the intrinsic problems of using oxides as electrocatalysts.

Recent advances in 2D structured materials with defect-exploiting design strategies for electrocatalysis of nitrate to ammonia

Figure 5. (A) Schematic image of the synthesis process for Cu/CuOx/CF. (B) NH3 yield rates and FEs of Cu/CuOx/CF at different potentials. (C) NH3 yield rates and FEs of Cu/CuOx/CF, CuOx/CF, Cu/CF, and CF. (D) The Cu/CuOx/CF recycling test was at -1.3 V (vs. SCE). (A-D) Figures are reprinted with permission from Ref.[60]. Copyright 2022 American Chemical Society[60]. (E) SEM images of CO3O4/Co-h. (F) Calculated adsorption energies of NO3- on various Co3O4 models. (G) Computed Gibbs free energies of H2 production. (H) The NH3 yield rate of Co3O4/Co-l and Co3O4/Co-h at different potentials. (E-H) Figures are reprinted with permission from Ref.[61]. Copyright 2023 Elsevier[61].

Metal-organic frameworks

MOFs are new crystalline porous materials that can be formed by coupling metal ions/clusters with organic ligands[86]. They have been widely used in batteries[90], supercapacitors[91], catalysis[92], etc.[93,94] due to their advantages such as high specific surface area[95], tunable pore size, and high porosity[96]. The 2D nanosheets of MOFs have recently emerged as a promising material that makes them valuable in widespread electrocatalytic fields due to their atomic-level thickness, abundant active sites, and large surface area. Additionally, 2D MOF/COF nanosheet synthesis methods can be divided into top-down and bottom-up[97]. In the former, bulk crystals are synthesized first and then split into multi-layer or even single-layer nanosheets. The latter involves directly synthesizing 2D nanosheets by controlling the growth of bulk crystals in a specific direction. Bulk crystals connected between layers by hydrogen bonds or van der Waals forces are successfully exfoliated using ultrasound, mechanical force, etc. Due to the electrocatalytic reactions being related to the surface properties of the materials, 2D MOF materials provide an opportunity to construct an efficient electrode with the advantages of flexible surface and active sites for electrocatalysis. In addition, the crystal structure of 2D MOF nanosheets provides highly exposed surface area and active sites at the atomic level. These consequences significantly shorten the diffusion lengths of the products and reactants to improve the performance of an electrocatalyst and have shown excellent performance in many fields. However, some traditional MOFs still suffer from low conductivity and stability.

Lv et al. reported 2D-In MOFs as an electrocatalyst with a uniformly dispersed single-atom site and enzyme mimic system for efficient NitRR[98]. The 2D-In MOFs were synthesized through a wet chemical method, and they showed a high ammonia yield rate of 278.8 μg h-1 cm-2 at -0.7 V (vs. RHE) and pH 2 with an optimal FE of 92%. These high performances are due to their stable microenvironment space and reversible ligand dissociation, as shown in Figure 6A. Figure 6B shows the solid-state cyclic voltammograms of 2D-In MOFs, showing two peaks of 0.22 and 0.55 V (vs. FC/FC+). Both peaks are attributed to redox-active tetrathiafulvalene (TTF) ligand and quasi-reversible electron transfer process. The reversible ligand dissociation at various potentials and pH is studied by in situ Raman spectroscopy in Figure 6C and D. The two peaks of 363 and 505 cm-1 correspond to the In(III), and after the applied negative potential, two weak peaks of 438 and 806 cm-1 appeared, which are due to In(II) and NO3- adsorption at pH 1~3, respectively. On the other hand, apparent differences in peaks are not found in Figure 6D, which means the ligand dissociation mechanism does not work at the high pH. Lastly, the DFT method verifies the ligands’ dynamic dissociation catalytic mechanism theoretically [Figure 6E].

Recent advances in 2D structured materials with defect-exploiting design strategies for electrocatalysis of nitrate to ammonia

Figure 6. (A) Schematic illustration of the multifunction motifs assembled in 2D-In-MOF and their functional characteristics and dynamic ligand dissociation mechanism. (B) Solid-state cyclic voltammograms of 2D-In-MOF conducted in 0.1 M LiBF4 in CH3CN electrolyte. (C) Ex situ and in situ Raman spectra of 2D-In-MOF before (black) and after (red) electrocatalysis in H2SO4 (pH = 1) with 0.5 g L-1 KNO3 and (D) in 50 mM H2SO4 (pH = 5) with 0.5 g L-1 KNO3. (E) Free energy diagram of electrochemical conversion of NO3- to NH3 for 2D-In-MOF and the optimized geometric structures of intermediates corresponding to the N-end pathway (pH = 2/3). (A-E) Figures are reprinted with permission from Ref.[98]. Copyright 2022 American Chemical Society[98]. (F) HR-TEM image of Cu1Co1HHTP, (G) The electrocatalytic activities of various CuxCoyHHTP catalysts at -0.6 V (vs. RHE). (H) d-orbit PDOS of Cu site in Cu1Co1HHTP and CuHHTP slabs. (I) Free energy diagrams of the NitRR process on Cu1Co1HHTP and CuHHTP slabs. (F-J) Figures are reprinted with permission from Ref.[99]. Copyright 2023 Elsevier[99].

Liu et al. reported bimetallic conductive copper-cobalt hexahydroxytriphenylene MOF (CuxCoyHHTP), which has a synergy effect between different single-metal sites and shows a great yield rate and a FE of 299.9 μmol h-1 cm-2 and 96.4%, respectively[99]. CuxCoyHHTP was synthesized via the hydrothermal method, and the morphology was analyzed by High-Resolution Transmission Electron Microscopy (HR-TEM) [Figure 6F]. To investigate the synergistic effect of Co doping to Cu active sites, electrochemical NitRR with various Co/Cu ratios is performed. The Cu1Co1HHTP showed an optimal ratio of Cu:Co (1:1) with the highest ammonia yield rate and FE [Figure 6G]. After the doping of Co, εd of the Cu site in Cu1Co1HHTP d = -3.22 eV) is closer to the Fermi level than CuHHTP. This result explains that the doping of heteroatoms affects the electronic structure of Cu sites, which can adsorb the intermediate properly [Figure 6H]. Lastly, Cu1Co1HHTP is verified more favorable for NitRR from the Gibbs free-energy diagram, the ΔG of NO3- adsorption and potential determining step (PDS) (NO* → NOH*) on the Cu site of Cu1Co1HHTP are -0.82 and 0.54 eV, respectively, which are lower than CuHHTP (-0.55 and 1.56 eV) [Figure 6I and J].

Fang et al. reported 2D-Fe cyano NSs (nanosheets) as multi-functional electrocatalysts that have abundant Fe0 active sites for NitRR and oxygen evolution reaction (OER) via topotactic conversion and electro-reduction [Figure 7A][100]. As shown in Figure 7B, Fe cyano NSs have a large active surface area, which is 1.54 times the Fe-cyano bulk and super hydrophilic properties from electrical double-layer capacitance (EDLC) and contact angle data. These properties are due to 2D engineering and hydrophilic functional groups and are favorable for electrode contact with electrolytes and NO3- adsorption. In addition, the X-ray Photoelectron Spectroscopy (XPS) data shows the generation of Fe0 in Fe cyano NSs after the electro-reduction process, and the Fe0 sites are verified as active sites for NitRR by poisoning Fe sites using potassium thiocyanide (KSCN) [Figure 7C and D]. DFT simulation shows negative ΔG of NO3- adsorption on Fe0 sites that supports the suppression of HER and NitRR selectivity. In the electrochemical NitRR test, the Fe cyano NSs exhibit a high ammonia yield rate of 42.1 mg h-1 mgcat.-1 and an FE of over 90% at -0.5 V (vs. RHE) [Figure 7E]. Furthermore, Fe cyano NS-based material as a multifunctional catalyst showed excellent durability without degradation of cell voltage at 60 mA cm-2 for an electrolyzer of OER and NitRR [Figure 7F]. Therefore, recent research on 2D-MOFs as electrocatalysts has progressed due to their high surface area and shortened diffusion lengths. In particular, 2D-MOFs, which have uniformly dispersed single atom sites and doped single heteroatom sites, show a synergistic effect on electrochemical activity and can improve conductivity and stability with reversible reactions.

Recent advances in 2D structured materials with defect-exploiting design strategies for electrocatalysis of nitrate to ammonia

Figure 7. (A) Synthesis of ultra-large Fe-cyano NSs via directional freezing. (B) EDLC of blank CFP electrode, Fe-cyano bulk, and NS electrocatalysts. Inset of (B) CAs of (i) blank CFP electrode, (ii) Fe-cyano bulk, and (iii) NS electrocatalysts. (C) Poisoning experiment on Fe-cyano-R NSs. (D) H2O and NO3- adsorption on Fe0. (E) Ammonia yield rate and Faradaic efficiency (FE) at various potentials on Fe-cyano NSs. (F) Stability of electrolyzer. Reprinted with permission from Ref.[100]. Copyright 2022 American Chemical Society[100].

MXenes

MXenes are generally defined as 2D TM carbides or nitrides, with Ti3C2Tx synthesized from the Mn+1AXn (MAX) phase of Ti3AlC2 first reported in 2011[101]. The thickness of MXenes is typically in the 1 nm range and depends on the n value (Mn+1XnTx) of the MXenes[102]. MXenes are primarily synthesized by selectively etching the A layer from the precursor MAX phase, which is a precursor of MXenes and is typically a series of ternary layer compounds, where M is a TM, A is an element of group IIIA or IVA in the periodic table, and X is carbon or nitrogen [Figure 8A][103]. Because the M-X covalent bonds and M-A metal bonds on MAX are powerful, for breaking the M-A bonds, potent etching agents such as hydrofluoric acid (HF) and lithium fluoride-hydrochloric acid mixtures (LiF-HCl) are required. As a result, the surface of MXenes is terminated with abundant functional groups such as -OH and -O, which vary in type and amount depending on the etching route [Figure 8B][104]. Various stoichiometries of MXenes have been studied, such as monometallic MXenes, bimetallic MXenes, and multi-metal MXenes [Figure 8C][105].

Recent advances in 2D structured materials with defect-exploiting design strategies for electrocatalysis of nitrate to ammonia

Figure 8. (A) A schematic representation of MAX structures from n = 1 to n = 4 and their etched MXene structures. Reprinted with permission from Ref.[103]. Copyright 2021 Wiley-VCH)[103]. (B) A schematic of different surface functional groups depending on the etching route. Reprinted with permission from Ref.[104]. Copyright 2018 Wiley-VCH[104]. (C) Schematic depicting different MXene stoichiometries. Reprinted with permission from Ref.[105]. Copyright 2023 Wiley-VCH[105].

Due to their numerous properties, including unique mechanical properties[106], metallic conductivity[107], in-plane anisotropic structure[108], and tunable band gap[109], MXenes have been arising as multipurpose 2D materials for promising applications[110-112]. Compared to other 2D materials, they have abundant terminal functional groups[113], enabling reversible redox reactions[114]. Unlike other 2D materials, they do not require processing to increase the exposure and density of active sites. In general, MXene basal planes are considered catalytically inactive[115]. However, these can become active by tuning their terminal groups. Hu et al. reveal the mechanism of NO3- to ammonia on M3C2 MXenes, and their catalytic activity depends on the terminated group, OVs, and environment of different pH via DFT calculations[116]. Generally, most M3C2 MXenes prefer HER selective; on the other hand, O-terminated Ti3C2 and Ti3C2O2 with OV groups have NitRR selectivity due to the electronic state of N intermediate on Ti3C2O2 with OVs showed closer than Ti3C2 to the Fermi level. These results are due to the increased oxidation state of surface Ti, which weakens *NHx adsorption and facilitates the hydrogenation process.

Cai et al. reported that surface hydroxyl on oxygen-functionalized Ti3C2 MXene (Of-TiMX) plays a vital role in the NitRR process[115]. The Of-TiMX was synthesized via three steps of etching, exfoliation, and calcination to make an oxygen group on the surface of TiMX [Figure 9A]. The Of-TiMX shows not only selective NitRR but also a high FE and an ammonia yield rate of 90.4% and 0.99 mg h-1 cm-2 at -1.7 V (vs. SCE), which indicates HER is effectively suppressed on its surface [Figure 9B].

Recent advances in 2D structured materials with defect-exploiting design strategies for electrocatalysis of nitrate to ammonia

Figure 9. (A) Schematic illustration of synthesizing the Of-TiMX for NitRR. (B) FEs of ammonia and nitrite of Of-TiMX at different potentials. (C) Pourbaix diagrams for Of-TiMX at pH 7. (D) Free-energy diagrams for NitRR on the surface of full OH-terminated Of-TiMX at U = 0.0 V and U = -1.2 V (vs. SHE). Reprinted with permission from Ref.[115]. Copyright 2023 Elsevier[115].

From the surface Pourbaix diagram, the trend of the amount of surface hydroxyl grows as the potential increases [Figure 9C]. This trend is related to the experimental results that the Of-TiMX has the highest NitRR performance at -1.7 V (vs. SCE) because the hydroxyl groups allow interfacial H-bonding with NO3- and other intermediates for NH3 production. In addition, the free energy diagram for NitRR on the hydroxyl-covered surface of Of-TiMX suggests the NitRR pathway to synthesize ammonia and support high optimized potential and selectivity [Figure 9D]. Hence, recent research on MXenes and advances in computational science have been extensive. Significantly, the surface modification of MXenes, which have abundant and tuneable functional groups on their surface, is determined to improve electrocatalytic activity and selectivity with a high ammonia yield rate.

Graphene-based materials

To reduce the use of precious metals, carbon nanomaterials are commonly used in electrocatalytic applications due to their high electrical conductivity, low cost, and chemical stability[117]. Graphene, one of the carbon nanomaterials, has mainly been utilized as a catalyst or support to improve the electrocatalytic performance of the NitRR owing to its 2D structure. However, when used alone, graphene has relatively poor electrocatalytic properties, such as low activity and selectivity. Various engineering approaches have been investigated to overcome these problems, including a combination of TMs and defect engineering.

Graphene has been widely used as a support for single-atom catalysts (SACs) due to its suitability as an excellent structural support. SACs are dispersed or coordinated atomically on the surface of 2D materials such as graphene, and surface defects are often utilized as their active sites. These catalysts have many advantages, including high stability, activity, selectivity, efficient active sites, and drastic cost reduction. Additionally, they allow for the tuning of the selectivity and activity of electrocatalytic reactions[118]. To synthesize the SACs, lots of synthesis methods, such as wet-chemical, MOF-derived, atomic layer deposition (ALD), co-precipitation, and impregnation, have been investigated[119].

Wu et al. synthesized Fe SAC anchored on graphene through pyrolysis and etching, as shown in Figure 10A[62]. The Fe SAC on graphene exhibits higher electrocatalytic performance with a yield rate of ~20,000 μg h-1 mgcat.-1 at -0.85 V (vs. RHE) and a FE of ~75% at - 0.66 V (vs. RHE) than those of iron nanoparticles supported on nitrogen-doped carbon (FeNP/NC) and NC [Figure 10B and C]. The N-N coupling pathway for N2 production, one of the competitive reactions, can be inhibited due to the isolated active Fe atoms, as the *N on neighboring active sites cannot exist side by side. Figure 10D displays the minimum energy pathway (MEP) of NitRR on the Fe single atom site. DFT calculations show that the MEP converts to a downhill conformation when U = -0.3 V (vs. RHE) (black line) in Figure 10E; the results support that the experimental onset potential [~-0.4 V (vs. RHE)] is reasonable.

Recent advances in 2D structured materials with defect-exploiting design strategies for electrocatalysis of nitrate to ammonia

Figure 10. (A) Schematic image of the synthesis process for Fe SAC. (B) NH3 partial current density and yield rate of Fe SAC, FeNP/NC, NC. (C) NH3 FE of Fe SAC at different potentials. (D) Calculated minimum energy pathway on Fe SAC for NitRR. (E) Calculated accessible energy pathway at U = 0 and -0.3 V (vs. RHE). (A-E) Figures are reprinted with permission from Ref.[62]. Copyright 2021 Springer Nature Limited[62]. (F) The metal atoms and substrates considered for constructing the DACs on N-graphenes. (G) Calculated free energy of NitRR on Cu2/N3-6 and the involved intermediates. (H) Potential-dependent NH3 yield rate and FE of Cu2/N3-6. (F-H) Figures are reprinted with permission from Ref.[121]. Copyright 2023 Elsevier[121]. (I) HR-TEM image of ox-LIG. (J) NH3 FE and (K) ox-LIG, LIG, and rGO yield rates at different potentials. (I-K) Figures are reprinted with permission from Ref.[122]. Copyright 2023 Wiley-VCH GmbH[122].

Beyond the study of SACs, dual-atom catalysts (DACs) have been investigated because of their unique structure. Their structure could suppress HER on their surface and increase the selectivity of NitRR due to an ensemble effect[120]. Zhao et al. calculated the DFT of DACs embedded in N-graphene for NitRR and demonstrated the results with subsequent experiments[121]. The geometric structure of DACs is determined by the d-band type of the constituent metals [Figure 10F and G]. The DFT indicated the synergistic effects of two active sites improve the activation of catalysts, which is related to intrinsic d-band centers of active sites. Then, it is revealed that Cu2@N3-6 catalysts have high electrocatalytic performance for NitRR in DFT results, and Cu2@N3-6 achieved a high FE of 97.4% and yield rate of 18.2 mg h-1 cm-2 at -0.8 V (vs. RHE) in the experiment [Figure 10H].

Defect engineering can increase the electrochemical activity of the graphene for the NitRR due to the change in electronic structures. Huang et al. show a skyscraping FE of nearly 100% with amorphous graphene for NitRR[122]. The amorphous graphene was synthesized via laser induction, which resulted in numerous defects, disordered pentagons, hexagons, and heptagons in graphene, and Figure 10I shows these defects in graphene. These defects increase the selectivity of the graphene for NitRR. The amorphous graphene has high electrocatalytic performances, which are FE of about 100% and NH3 production of 2,859 μg h-1 cm-2 at -0.93 V (vs. RHE) [Figure 10J and K]. As mentioned above, graphene-based materials have gained attention as catalysts and support for NitRR due to their high conductivity, cost-efficiency, and expansive surface area, reducing the reliance on precious catalysts while maintaining electrochemical performance. Recently, studies have explored the potential of graphene as a standalone catalyst for NitRR, particularly when coupled with defect engineering. However, some defect engineering, such as N-S or N-P co-doping on graphene, remains unapplied for NitRR. Additionally, other carbon-based materials, such as carbon nanotubes (CNTs), with defect engineering have not received extensive investigation, suggesting that carbon-based materials hold further potential for NitRR.

CONCLUSIONS AND PERSPECTIVES

Electrochemical synthesis of green ammonia has been focused on as an alternative way to the energy-intensive Harber-Bosch process due to its advantages, such as zero-carbon emission, scalability, and production under ambient conditions. However, commercialization of NitRR undergoes inevitable drawbacks, such as low selectivity of ammonia and high cost of electrocatalysts, due to immature technology and competitive HER. To increase the electrochemical performance of the NitRR, the design strategies of electrocatalysts have been studied over the past decade. Furthermore, 2D structured materials have been investigated as promising candidates for efficient electrocatalysts due to their advantages, such as a particular surface area with an abundant active site, ease of modification, and low cost. In addition, various defect engineering, such as oxygen defect, doping, and topological defects, have been considered to improve the electrocatalytic performance because these defects in 2D electrocatalysts not only increase the mass/charge transport, active site, and electrical conductivity but can also tune the electronic structure.

This review summarized the recent advances and development of 2D catalysts with defect engineering for NitRR. While some 2D materials are employed as catalysts for NitRR, certain drawbacks become apparent. For example, TMOs exhibit low conductivity, and graphene demonstrates low activity. However, defects emerge as a solution, addressing these intrinsic issues by enhancing catalyst conductivity and activity. Unexplored 2D materials, such as TaS2, MoTe2, and defect-rich CNTs, remain untapped for converting from nitrate to ammonia. Additionally, various defect engineering techniques, such as N-P or N-S co-doping and topological defects, await NitRR exploration. Consequently, the remaining candidates can potentially increase ammonia yield and FE.

Recent studies demonstrated the high electrocatalytic performance of NitRR, as shown in Table 2. Understanding the NitRR mechanism on the catalyst is essential to improve electrocatalytic performance. In previous studies, the NitRR mechanism of defects-containing 2D catalysts has also been investigated by the DFT method. However, most DFT calculations have focused on point defects, such as vacancies and doping. There are still unresolved issues related to NitRR regarding complex 2D and 3D defects, such as grain boundaries (GBs), twins, distortion, voids, and cracks. In realistic situations, catalysts have numerous GBs, and the unsaturated coordination of GB atoms can enhance catalytic activity. DFT calculations can utilize two GB bulk models built based on the coincidence site lattice (CSL) theory to address this. Moreover, appropriate conditions are also necessary for understanding the precise properties of each type of defect.

Table 2

Comparison of different 2D-structured materials as electrocatalysts for NitRR

CategoryElectrocatalystsSynthesis methodSubstrateElectrolyteNH3 yield rateFE (%)Selectivity (%)Ref.
Layered double hydroxide (LDH)CoFe LDHHydrothermalNi foam1 M KOH + 1,400 ppm KNO30.93 mmol h-1 cm-2
@-0.45 V (vs. RHE)
97.6898.93[52]
Ni3Fe-CO3 LDHCo-precipitationCu foam1 M KOH + 5 mM KNO31.261 mg h-1 cm-2
@-0.2 V (vs. RHE)
96.8095.80[53]
Chalcogenide Fe-MoS2HydrothermalCarbon cloth0.1 M Na2SO4 + 0.1 M NaOH + 0.1 M NaNO30.03 mmol h-1 cm-2
@-0.5 V (vs. RHE)
98[54]
Fe-MoS2HydrothermalCarbon cloth1 M LiCl + 0.25 M LiNO39.75 mg h-1 cm-2
@-0.85 V (vs. RHE)
90[83]
Cu-SnS2-xSolvothermalCarbon cloth0.1 M KOH + 0.1 M KNO30.63 mmol h-1 mgcat.-1
@-0.7 V (vs. RHE)
93.80[55]
B-MoS2HydrothermalCarbon cloth0.5 M Na2SO4 + 0.1 M NaNO310.8 mg h-1 cm-2
@-0.7 V (vs. RHE)
92.30[123]
Transition-metal oxide
(TMO)
Co3O4/Co-lElectrodeposition
& calcination
Co foam1.000 ppm NO3- + 0.1 M Na2SO44.43 mg h-1 cm-2
@-0.8 V (vs. RHE)
88.7083.30[61]
Cu/CuOx/CFOxidation treatmentCu foam0.5 M K2SO4 + 200 mg L-1 NO3-0.23 mmol h-1 cm-2
@-1.3 V (vs. SCE)
93.5895.00[60]
CuO NWAs@Fe3O4Hydrolysis
& pyrolysis
Cu foam1 M KOH + 1,400 ppm NO3-1.691 mmol h-1 cm-2
@-0.27 V (vs. RHE)
99.4688.68[124]
Metal-organic framework (MOF) Ni-MOFSolvothermalNi foam70 mg·L-1 NaNO3, 0.1 M Na2SO4110.13 ug h-1 cm-2
@-1.4 V (vs. SCE)
96.4080[58]
cMOF(CuxCoyHHTP)HydrothermalCarbon cloth0.5 M Na2SO4 + 0.1 M NaNO3299.9 μmol h-1 cm- 2
@-0.6 V (vs. RHE)
96.40[99]
Fe-cyano NSsSol-gelCarbon fiber
paper
0.1 M KNO3 + 1 M KOH42.1 mg h-1 mgcat.-1
@-0.5 V (vs. RHE)
90.20[100]
In-MOF In8SolvothermalCarbon paper5 mM H2SO4 + 15 g·L-1 KNO3256.9 μg h-1 mgcat.-1
@-0.7 V (vs. RHE)
90.1099.30[98]
MXene Ti3C2Etching &
ultrasonication
Carbon paper0.5 M K2SO4 + 200 ppm KNO30.99 mg h-1 cm-2
@-1.7 V (vs. SCE)
90[115]
BP/Nb2CSelf-assemblyCarbon cloth0.1 M K2SO4 + 0.05 M KNO31,967.0 μg h-1 cm-2
@-0.6 V (vs. RHE)
90.4
@-0.5 V vs. RHE
[125]
Bi2O3/MXeneSolvothermalCarbon paper1,000 ppm NO3- + 0.5 M Na2SO47.00 mg h-1 cm-2
@-1.8 V (vs. SCE)
91.10[126]
Fe1Cu2@MXeneWet-chemicalCarbon cloth100 mg L-1 KNO3 + 0.1 M Na2SO41.2 mg h-1 cm-2
@-0.95 V (vs. RHE)
5299.60[127]
FeNPs@MXeneWet-chemicalCarbon cloth100 mg L-1 NO3 + 0.5 M Na2SO4~0.8 mg h-1 cm-2
@-0.95 V (vs. RHE)
20.0076.80[128]
CuCl_BEFWet-chemicalCarbon cloth100 mg L-1 NO3 + 0.5 M Na2SO41.82 mg h-1 cm-2
@-1.0 V (vs. RHE)
44.7098.60[129]
Mo2C NSsCarbonizationCarbon paper1 M NaOH + 0.1 M NaNO325.2 mg h-1 mgcat.-1
@-0.4 V (vs. RHE)
81.4
@-0.3 V vs. RHE
[130]
Graphene-basedmaterial Fe SACWet-chemicalGlassy carbon0.1 M K2SO4 + 0.5 M KNO320,000 ug h-1 mgcat.-1
@-0.85 V (vs. RHE)
75
@-0.66 V vs. RHE
69[62]
Cu2@N3-6Pyrolysis methodCarbon cloth0.5 M Na2SO4 + 0.1 M NaNO318.2 mg h-1 cm-2
@-0.8 V (vs. RHE)
97.40[121]
ox-LIGLaser inductionCellulose paper/Au layer1 M NaNO32,859 μg h-1 cm-2
@-0.93 V (vs. RHE)
96.40> 70[122]

Understanding fluid dynamics within the flow electrolyzer is a critical concern in the scale-up process. These considerations become particularly vital as cell designs progress from H-type cells to flow cells for NitRR. Understanding NitRR in the flow state is paramount. Various approaches can be employed to achieve a comprehensive understanding of NitRR in a flow state, including computational simulations such as computational fluid dynamics (CFD) and molecular dynamics (MD). Additionally, in situ characterization techniques, such as in situ Fourier-transform infrared spectroscopy (FT-IR), in situ Raman spectroscopy, in situ X-ray absorption spectroscopy (XAS), in situ scanning electrochemical microscopy (SECM), and online DEMS, need further exploration to accurately decipher the conversion of NO3- to NH3 accurately and to gain insights into the kinetics of the reaction. These in situ measurements will provide clues to suppress the production of by-products such as H2, N2, and NO2-. Although some challenges remain, progress is highly expected through continuous research. This review will help bridge the gap between theory and experiment to understand 2D materials and defect engineering for NitRR.

DECLARATIONS

Author’s contributions

Conceptualization, investigation, visualization, validation, writing - original draft, writing - review & editing: Choi J, Im S, Kim JK, Sim U

Investigation, visualization, writing - review & editing: Choi J

Validation, writing - review & editing: Surendran S

Writing - review & editing: Moon DJ, Kim JY

Availability of data and materials

Not applicable.

Financial support and sponsorship

This work was supported by the Korea Institute of Energy Technology Evaluation and Planning (KETEP) grant funded by the Korean government (MOTIE) (No. RS-2023-00242227 (Clean Hydrogen and Ammonia Innovation Research Center) and No. 20224000000320). This work was supported from the National Research Foundation of Korea (NRF) grant funded by the Korean government (MSIT) (No. 2022R1A2C1012419). This work was supported by the New & Renewable Energy Core Technology Program of the Korea Institute of Energy Technology Evaluation and Planning (KETEP) granted financial resource from the Ministry of Trade, Industry & Energy, Republic of Korea (No. 20213030040590). This research was supported by Nano·Material Technology Development Program through the National Research Foundation of Korea (NRF) funded by Ministry of Science and ICT (2020M3H4A3106313). Additionally, this work was backed by the KENTECH Research Grant funded by the Korea Institute of Energy Technology, Republic of Korea (KRG2022-01-016). This research also benefited from the "Regional Innovation Strategy (RIS)" through the National Research Foundation of Korea (NRF), funded by the Ministry of Education (MOE) (2021RIS-002).

Conflicts of interest

All authors declared that there are no conflicts of interest.

Ethical approval and consent to participate

Not applicable.

Consent for publication

Not applicable.

Copyright

© The Author(s) 2024.

REFERENCES

1. Klerke A, Christensen CH, Nørskov JK, Vegge T. Ammonia for hydrogen storage: challenges and opportunities. J Mater Chem 2008;18:2304-10.

2. Wang Y, Zhou X, Liu L. Theoretical investigation of the combustion performance of ammonia/hydrogen mixtures on a marine diesel engine. Int J Hydrog Energy 2021;46:14805-12.

3. Chehade G, Dincer I. Progress in green ammonia production as potential carbon-free fuel. Fuel 2021;299:120845.

4. Guo Y, Pan Z, An L. Carbon-free sustainable energy technology: direct ammonia fuel cells. J Power Sources 2020;476:228454.

5. Valera-Medina A, Xiao H, Owen-Jones M, David W, Bowen P. Ammonia for power. Prog Energy Combust Sci 2018;69:63-102.

6. Cheng S, Logan BE. Ammonia treatment of carbon cloth anodes to enhance power generation of microbial fuel cells. Electrochem Commun 2007;9:492-6.

7. Giddey S, Badwal SPS, Munnings C, Dolan M. Ammonia as a renewable energy transportation media. ACS Sustain Chem Eng 2017;5:10231-9.

8. Wang W, Herreros JM, Tsolakis A, York AP. Ammonia as hydrogen carrier for transportation; investigation of the ammonia exhaust gas fuel reforming. Int J Hydrog Energy 2013;38:9907-17.

9. Fu E, Gong F, Wang S, Xiao R. Chemical looping technology in mild-condition ammonia production: a comprehensive review and analysis. Small 2024;20:e2305095.

10. Erisman JW, Sutton MA, Galloway J, Klimont Z, Winiwarter W. How a century of ammonia synthesis changed the world. Nature Geosci 2008;1:636-9.

11. Gao Y, Wang R, Li Y, et al. Regulating dynamic equilibrium of active hydrogen for super-efficient nitrate electroreduction to ammonia. Chem Eng J 2023;474:145546.

12. Yang B, Ding W, Zhang H, Zhang S. Recent progress in electrochemical synthesis of ammonia from nitrogen: strategies to improve the catalytic activity and selectivity. Energy Environ Sci 2021;14:672-87.

13. Noh H, Kang K, Seo Y. Environmental and energy efficiency assessments of offshore hydrogen supply chains utilizing compressed gaseous hydrogen, liquefied hydrogen, liquid organic hydrogen carriers and ammonia. Int J Hydrog Energy 2023;48:7515-32.

14. Asif M, Sidra Bibi S, Ahmed S, et al. Recent advances in green hydrogen production, storage and commercial-scale use via catalytic ammonia cracking. Chem Eng J 2023;473:145381.

15. Soloveichik G. Electrochemical synthesis of ammonia as a potential alternative to the Haber-Bosch process. Nat Catal 2019;2:377-80.

16. El-shafie M, Kambara S. Recent advances in ammonia synthesis technologies: toward future zero carbon emissions. Int J Hydrog Energy 2023;48:11237-73.

17. Lazouski N, Chung M, Williams K, Gala ML, Manthiram K. Non-aqueous gas diffusion electrodes for rapid ammonia synthesis from nitrogen and water-splitting-derived hydrogen. Nat Catal 2020;3:463-9.

18. Verleysen K, Parente A, Contino F. How sensitive is a dynamic ammonia synthesis process? Global sensitivity analysis of a dynamic Haber-Bosch process (for flexible seasonal energy storage). Energy 2021;232:121016.

19. Guo J, Chen P. Catalyst: NH3 as an energy carrier. Chem 2017;3:709-12.

20. Zhang L, Ji X, Ren X, et al. Electrochemical ammonia synthesis via nitrogen reduction reaction on a MoS2 catalyst: theoretical and experimental studies. Adv Mater 2018;30:e1800191.

21. Montoya JH, Tsai C, Vojvodic A, Nørskov JK. The challenge of electrochemical ammonia synthesis: a new perspective on the role of nitrogen scaling relations. ChemSusChem 2015;8:2180-6.

22. Wu D, Peng X, Li L, et al. Commercial biogas plants: review on operational parameters and guide for performance optimization. Fuel 2021;303:121282.

23. Giddey S, Badwal S, Kulkarni A. Review of electrochemical ammonia production technologies and materials. Int J Hydrog Energy 2013;38:14576-94.

24. Shen H, Choi C, Masa J, et al. Electrochemical ammonia synthesis: mechanistic understanding and catalyst design. Chem 2021;7:1708-54.

25. Long J, Chen S, Zhang Y, et al. Direct electrochemical ammonia synthesis from nitric oxide. Angew Chem Int Ed 2020;59:9711-8.

26. Jiang H, Chen GF, Hai G, et al. A nitrogen battery electrode involving eight-electron transfer per nitrogen for energy storage. Angew Chem Int Ed 2023;62:e202305695.

27. Smith C, Hill AK, Torrente-murciano L. Current and future role of Haber-Bosch ammonia in a carbon-free energy landscape. Energy Environ Sci 2020;13:331-44.

28. Thakur IS, Medhi K. Nitrification and denitrification processes for mitigation of nitrous oxide from waste water treatment plants for biovalorization: challenges and opportunities. Bioresour Technol 2019;282:502-13.

29. Hon WM, Lee KH, Khoo HE. Nitric oxide in liver diseases: friend, foe, or just passerby? Ann N Y Acad Sci 2002;962:275-95.

30. Knobeloch L, Salna B, Hogan A, Postle J, Anderson H. Blue babies and nitrate-contaminated well water. Environ Health Perspect 2000;108:675-8.

31. WHO. Nitrate and nitrite in drinking-water. 2016. Avaliable from: https://www.who.int/docs/default-source/wash-documents/wash-chemicals/nitrate-nitrite-background-document.pdf [Last accessed on 20 Feb 2024].

32. Schoeman J, Steyn A. Nitrate removal with reverse osmosis in a rural area in South Africa. Desalination 2003;155:15-26.

33. Cliford D, Liu X. Ion exchange for nitrate removal. J Am Water Works Assoc 1993;85:135-43.

34. Wei L, Liu D, Rosales BA, Evans JW, Vela J. Mild and selective hydrogenation of nitrate to ammonia in the absence of noble metals. ACS Catal 2020;10:3618-28.

35. 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.

36. Xu H, Ma Y, Chen J, Zhang WX, Yang J. Electrocatalytic reduction of nitrate - a step towards a sustainable nitrogen cycle. Chem Soc Rev 2022;51:2710-58.

37. Wang Y, Shu S, Peng M, et al. Dual-site electrocatalytic nitrate reduction to ammonia on oxygen vacancy-enriched and Pd-decorated MnO2 nanosheets. Nanoscale 2021;13:17504-11.

38. Cui Y, Yang H, Dai C, Ren P, Song C, Ma X. Coupling of LaFeO3 - plasma catalysis and Cu+/Cu0 electrocatalysis for direct ammonia synthesis from air. Ind Eng Chem Res 2022;61:4816-23.

39. Deng X, Yang Y, Wang L, Fu XZ, Luo JL. Metallic Co nanoarray catalyzes selective NH3 production from electrochemical nitrate reduction at current densities exceeding 2 A cm-2. Adv Sci 2021;8:2004523.

40. Rosca V, Duca M, de Groot MT, Koper MT. Nitrogen cycle electrocatalysis. Chem Rev 2009;109:2209-44.

41. Jung W, Hwang YJ. Material strategies in the electrochemical nitrate reduction reaction to ammonia production. Mater Chem Front 2021;5:6803-23.

42. Wang Q, Zhao X, Zhang J, Zhang X. Investigation of nitrate reduction on polycrystalline Pt nanoparticles with controlled crystal plane. J Electroanal Chem 2015;755:210-4.

43. Lim J, Liu C, Park J, et al. Structure sensitivity of Pd facets for enhanced electrochemical nitrate reduction to ammonia. ACS Catal 2021;11:7568-77.

44. Liu H, Park J, Chen Y, et al. Electrocatalytic nitrate reduction on oxide-derived silver with tunable selectivity to nitrite and ammonia. ACS Catal 2021;11:8431-42.

45. Siriwatcharapiboon W, Kwon Y, Yang J, et al. Promotion effects of Sn on the electrocatalytic reduction of nitrate at Rh nanoparticles. ChemElectroChem 2014;1:172-9.

46. Wang Y, Zhou W, Jia R, Yu Y, Zhang B. Unveiling the activity origin of a copper-based electrocatalyst for selective nitrate reduction to ammonia. Angew Chem Int Ed 2020;59:5350-4.

47. He W, Zhang J, Dieckhöfer S, et al. Splicing the active phases of copper/cobalt-based catalysts achieves high-rate tandem electroreduction of nitrate to ammonia. Nat Commun 2022;13:1129.

48. Wang C, Ye F, Shen J, Xue KH, Zhu Y, Li C. In situ loading of Cu2O active sites on island-like copper for efficient electrochemical reduction of nitrate to ammonia. ACS Appl Mater Interfaces 2022;14:6680-8.

49. Anastasiadou D, van Beek Y, Hensen EJM, Costa Figueiredo M. Ammonia electrocatalytic synthesis from nitrate. Electrochem Sci Adv 2023;3:e2100220.

50. Chia X, Pumera M. Characteristics and performance of two-dimensional materials for electrocatalysis. Nat Catal 2018;1:909-21.

51. Tao H, Fan Q, Ma T, et al. Two-dimensional materials for energy conversion and storage. Prog Mater Sci 2020;111:100637.

52. Du F, Li J, Wang C, et al. Active sites-rich layered double hydroxide for nitrate-to-ammonia production with high selectivity and stability. Chem Eng J 2022;434:134641.

53. Kim K, Lee H, Huang X, et al. Energy-efficient electrochemical ammonia production from dilute nitrate solution. Energy Environ Sci 2023;16:663-72.

54. Li J, Zhang Y, Liu C, et al. 3.4% Solar-to-ammonia efficiency from nitrate using Fe single atomic catalyst supported on MoS2 nanosheets. Adv Funct Mater 2022;32:2108316.

55. Li H, Xu X, Lin X, et al. Cooperative interaction between Cu and sulfur vacancies in SnS2 nanoflowers for highly efficient nitrate electroreduction to ammonia. J Mater Chem A 2023;11:2014-22.

56. Li Y, Ma J, Waite TD, Hoffmann MR, Wang Z. Development of a mechanically flexible 2D-MXene membrane cathode for selective electrochemical reduction of nitrate to N2: mechanisms and implications. Environ Sci Technol 2021;55:10695-703.

57. Karthikeyan P, Elanchezhiyan S, Preethi J, Talukdar K, Meenakshi S, Park CM. Two-dimensional (2D) Ti3C2Tx MXene nanosheets with superior adsorption behavior for phosphate and nitrate ions from the aqueous environment. Ceram Int 2021;47:732-9.

58. Pan F, Zhou J, Wang T, et al. 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.

59. Sivan SE, Kang KH, Han SJ, et al. Facile MOF-derived one-pot synthetic approach toward Ru single atoms, nanoclusters, and nanoparticles dispersed on CeO2 supports for enhanced ammonia synthesis. J Catal 2022;408:316-28.

60. Wang H, Guo Y, Li C, et al. Cu/CuOx in-plane heterostructured nanosheet arrays with rich oxygen vacancies enhance nitrate electroreduction to ammonia. ACS Appl Mater Interfaces 2022;14:34761-9.

61. Zhao F, Hai G, Li X, Jiang Z, Wang H. Enhanced electrocatalytic nitrate reduction to ammonia on cobalt oxide nanosheets via multiscale defect modulation. Chem Eng J 2023;461:141960.

62. Wu ZY, Karamad M, Yong X, et al. Electrochemical ammonia synthesis via nitrate reduction on Fe single atom catalyst. Nat Commun 2021;12:2870.

63. Wang H, Liu X, Niu P, Wang S, Shi J, Li L. Porous Two-dimensional materials for photocatalytic and electrocatalytic applications. Matter 2020;2:1377-413.

64. Jia R, Wang Y, Wang C, Ling Y, Yu Y, Zhang B. Boosting selective nitrate electroreduction to ammonium by constructing oxygen vacancies in TiO2. ACS Catal 2020;10:3533-40.

65. Liu Y, Deng P, Wu R, Zhang X, Sun C, Li H. Oxygen vacancies for promoting the electrochemical nitrogen reduction reaction. J Mater Chem A 2021;9:6694-709.

66. Wang C, Liu Z, Hu T, et al. Metasequoia-like nanocrystal of iron-doped copper for efficient electrocatalytic nitrate reduction into ammonia in neutral media. ChemSusChem 2021;14:1825-9.

67. Cui X, Tang C, Zhang Q. A review of electrocatalytic reduction of dinitrogen to ammonia under ambient conditions. Adv Energy Mater 2018;8:1800369.

68. Chen G, Ren S, Zhang L, et al. Advances in electrocatalytic N2 reduction - strategies to tackle the selectivity challenge. Small Methods 2019;3:1800337.

69. Wang Q, O'Hare D. Recent advances in the synthesis and application of layered double hydroxide (LDH) nanosheets. Chem Rev 2012;112:4124-55.

70. Yi H, Liu S, Lai C, et al. Recent advance of transition-metal-based layered double hydroxide nanosheets: synthesis, properties, modification, and electrocatalytic applications. Adv Energy Mater 2021;11:2002863.

71. Zhang S, Zhao Y, Shi R, et al. Efficient photocatalytic nitrogen fixation over Cuδ+ -modified defective ZnAl-layered double hydroxide nanosheets. Adv Energy Mater 2020;10:1901973.

72. Li T, Li R, Luo H. Facile in situ growth of Ni/Co-LDH arrays by hypothermal chemical coprecipitation for all-solid-state asymmetric supercapacitors. J Mater Chem A 2016;4:18922-30.

73. Xu J, Deng H, Song J, Zhao J, Zhang L, Hou W. Synthesis of hierarchical flower-like Mg2Al-Cl layered double hydroxide in a surfactant-free reverse microemulsion. J Colloid Interface Sci 2017;505:816-23.

74. Li J, Lian R, Wang J, He S, Jiang SP, Rui Z. Oxygen vacancy defects modulated electrocatalytic activity of iron-nickel layered double hydroxide on Ni foam as highly active electrodes for oxygen evolution reaction. Electrochim Acta 2020;331:135395.

75. Li L, Yang J, Yun Y, Hu S, Huang Y. Characterization and electrochemical behaviour of nanoscale hydrotalcite-like compounds toward the reduction of nitrate. Nanomaterials 2020;10:1926.

76. Yu MS, Jesudass SC, Surendran S, Kim JY, Sim U, Han MK. Synergistic interaction of MoS2 nanoflakes on La2Zr2O7 nanofibers for improving photoelectrochemical nitrogen reduction. ACS Appl Mater Interfaces 2022;14:31889-99.

77. Whittingham MS. Lithium batteries and cathode materials. Chem Rev 2004;104:4271-301.

78. Choi W, Choudhary N, Han GH, Park J, Akinwande D, Lee YH. Recent development of two-dimensional transition metal dichalcogenides and their applications. Mater Today 2017;20:116-30.

79. Sukanya R, da Silva Alves DC, Breslin CB. Review - recent developments in the applications of 2D transition metal dichalcogenides as electrocatalysts in the generation of hydrogen for renewable energy conversion. J Electrochem Soc 2022;169:064504.

80. Wang J, Sun Z, Li Y, et al. Sulfur vacancy MoS2 for electrocatalytic reduction of nitrate to ammonia with enhanced selectivity. J Alloys Compd 2023;955:170199.

81. Mohanty B, Jena BK, Basu S. Single atom on the 2D matrix: an emerging electrocatalyst for energy applications. ACS Omega 2020;5:1287-95.

82. Fan M, Cui J, Wu J, Vajtai R, Sun D, Ajayan PM. Improving the catalytic activity of carbon-supported single atom catalysts by polynary metal or heteroatom doping. Small 2020;16:e1906782.

83. Ding J, Hou X, Qiu Y, et al. Iron-doping strategy promotes electroreduction of nitrate to ammonia on MoS2 nanosheets. Inorg Chem Commun 2023;151:110621.

84. Xu T, Liang J, Li S, et al. Recent advances in nonprecious metal oxide electrocatalysts and photocatalysts for N2 reduction reaction under ambient condition. Small Sci 2021;1:2000069.

85. Cui X, Tang C, Liu XM, Wang C, Ma W, Zhang Q. Highly selective electrochemical reduction of dinitrogen to ammonia at ambient temperature and pressure over iron oxide catalysts. Chemistry 2018;24:18494-501.

86. Xiong W, Yin H, Wu T, Li H. Challenges and opportunities of transition metal oxides as electrocatalysts. Chemistry 2023;29:e202202872.

87. Sun Z, Liao T, Dou Y, et al. Generalized self-assembly of scalable two-dimensional transition metal oxide nanosheets. Nat Commun 2014;5:3813.

88. Sun Y, Liu Q, Gao S, et al. Pits confined in ultrathin cerium(IV) oxide for studying catalytic centers in carbon monoxide oxidation. Nat Commun 2013;4:2899.

89. Huang Z, Zhou A, Wu J, et al. Bottom-up preparation of ultrathin 2D aluminum oxide nanosheets by duplicating graphene oxide. Adv Mater 2016;28:1703-8.

90. Li C, Liu L, Kang J, et al. Pristine MOF and COF materials for advanced batteries. Energy Stor Mater 2020;31:115-34.

91. Sheberla D, Bachman JC, Elias JS, Sun CJ, Shao-Horn Y, Dincă M. Conductive MOF electrodes for stable supercapacitors with high areal capacitance. Nat Mater 2017;16:220-4.

92. Wang Q, Astruc D. State of the art and prospects in metal-organic framework (MOF)-based and MOF-derived nanocatalysis. Chem Rev 2020;120:1438-511.

93. Wu T, Liu X, Liu Y, et al. Application of QD-MOF composites for photocatalysis: energy production and environmental remediation. Coord Chem Rev 2020;403:213097.

94. Ma T, Li H, Ma JG, Cheng P. Application of MOF-based materials in electrochemical sensing. Dalton Trans 2020;49:17121-9.

95. Chae HK, Siberio-Pérez DY, Kim J, et al. A route to high surface area, porosity and inclusion of large molecules in crystals. Nature 2004;427:523-7.

96. Freund R, Zaremba O, Arnauts G, et al. The current status of MOF and COF applications. Angew Chem Int Ed 2021;60:23975-4001.

97. Wang J, Li N, Xu Y, Pang H. Two-dimensional MOF and COF nanosheets: synthesis and applications in electrochemistry. Chemistry 2020;26:6402-22.

98. Lv Y, Su J, Gu Y, et al. Atomically precise integration of multiple functional motifs in catalytic metal-organic frameworks for highly efficient nitrate electroreduction. JACS Au 2022;2:2765-77.

99. Liu P, Yan J, Huang H, Song W. Cu/Co bimetallic conductive MOFs: electronic modulation for enhanced nitrate reduction to ammonia. Chem Eng J 2023;466:143134.

100. Fang Z, Jin Z, Tang S, Li P, Wu P, Yu G. Porous two-dimensional iron-cyano nanosheets for high-rate electrochemical nitrate reduction. ACS Nano 2022;16:1072-81.

101. Zhang T, Pan L, Tang H, et al. Synthesis of two-dimensional Ti3C2Tx MXene using HCl+LiF etchant: enhanced exfoliation and delamination. J Alloys Compd 2017;695:818-26.

102. Ihsanullah I. MXenes (two-dimensional metal carbides) as emerging nanomaterials for water purification: progress, challenges and prospects. Chem Eng J 2020;388:124340.

103. Wyatt BC, Rosenkranz A, Anasori B. 2D MXenes: tunable mechanical and tribological properties. Adv Mater 2021;33:e2007973.

104. Qian A, Seo JY, Shi H, Lee JY, Chung CH. Surface functional groups and electrochemical behavior in dimethyl sulfoxide-delaminated Ti3C2Tx MXene. ChemSusChem 2018;11:3719-23.

105. Tsounis C, Kumar PV, Masood H, et al. Advancing MXene electrocatalysts for energy conversion reactions: surface, stoichiometry, and stability. Angew Chem Int Ed 2023;62:e202210828.

106. Yorulmaz U, Özden A, Perkgöz NK, Ay F, Sevik C. Vibrational and mechanical properties of single layer MXene structures: a first-principles investigation. Nanotechnology 2016;27:335702.

107. Wu Y, Nie P, Wu L, Dou H, Zhang X. 2D MXene/SnS2 composites as high-performance anodes for sodium ion batteries. Chem Eng J 2018;334:932-8.

108. Wu B, Cai X, Shui L, Gao E, Liu Z. Extraordinary electromechanical actuation of Ti2C MXene. J Phys Chem C 2021;125:1060-8.

109. Hu C, Du Z, Wei Z, Li L, Shen G. Functionalized Ti3C2Tx MXene with layer-dependent band gap for flexible NIR photodetectors. Appl Phys Rev 2023;10:021402.

110. Er D, Li J, Naguib M, Gogotsi Y, Shenoy VB. Ti3C2 MXene as a high capacity electrode material for metal (Li, Na, K, Ca) ion batteries. ACS Appl Mater Interfaces 2014;6:11173-9.

111. Yang Q, Huang Z, Li X, et al. A wholly degradable, rechargeable Zn-Ti3C2 MXene capacitor with superior anti-self-discharge function. ACS Nano 2019;13:8275-83.

112. Le TA, Tran NQ, Hong Y, Kim M, Lee H. Porosity-engineering of MXene as a support material for a highly efficient electrocatalyst toward overall water splitting. ChemSusChem 2020;13:945-55.

113. Zhou Z, Yu F, You Y, Zhan J, Zhang L. Tunable functional groups on MXene regulating the catalytic property of anchored cobalt phthalocyanine for electrochemical CO2 reduction. Inorg Chem Front 2023;10:5371-8.

114. Wang X, Bak S, Han M, et al. Surface redox pseudocapacitance of partially oxidized titanium carbide mxene in water-in-salt electrolyte. ACS Energy Lett 2022;7:30-5.

115. Cai J, Huang J, Cao A, et al. Interfacial hydrogen bonding-involved electrocatalytic ammonia synthesis on OH-terminated MXene. Appl Catal B 2023;328:122473.

116. Hu T, Wang M, Guo C, Li CM. Functionalized MXenes for efficient electrocatalytic nitrate reduction to ammonia. J Mater Chem A 2022;10:8923-31.

117. Brownson DA, Kampouris DK, Banks CE. An overview of graphene in energy production and storage applications. J Power Sources 2011;196:4873-85.

118. Zhuo HY, Zhang X, Liang JX, Yu Q, Xiao H, Li J. Theoretical understandings of graphene-based metal single-atom catalysts: stability and catalytic performance. Chem Rev 2020;120:12315-41.

119. Cheng N, Zhang L, Doyle-davis K, Sun X. Single-atom catalysts: from design to application. Electrochem Energ Rev 2019;2:539-73.

120. Rehman F, Kwon S, Musgrave CB, Tamtaji M, Goddard WA, Luo Z. High-throughput screening to predict highly active dual-atom catalysts for electrocatalytic reduction of nitrate to ammonia. Nano Energy 2022;103:107866.

121. Zhao T, Chen K, Xu X, et al. Homonuclear dual-atom catalysts embedded on N-doped graphene for highly efficient nitrate reduction to ammonia: from theoretical prediction to experimental validation. Appl Catal B 2023;339:123156.

122. Huang L, Cheng L, Ma T, et al. Direct synthesis of ammonia from nitrate on amorphous graphene with near 100% efficiency. Adv Mater 2023;35:e2211856.

123. Luo Y, Chen K, Shen P, et al. B-doped MoS2 for nitrate electroreduction to ammonia. J Colloid Interface Sci 2023;629:950-7.

124. Li J, Liu H, Du F, et al. Microenvironmental corrosion and hydrolysis induced two-dimensional heterojunction of copper oxide@ferriferrous oxide for efficient electrochemical nitrate reduction to ammonia. Chem Eng J 2023;471:144488.

125. Wang S, Song C, Cai Y, et al. Interfacial polarization triggered by covalent-bonded MXene and black phosphorus for enhanced electrochemical nitrate to ammonia conversion. Adv Energy Mater 2023;13:2301136.

126. Zhang H, Li L, Sun W, He J, Xu Q, Lu J. Highly dispersed in-situ grown Bi2O3 nanosheets on Ti3C2Tx MXene for selective electroreduction of nitrate to ammonia. ChemElectroChem 2023;10:e202201001.

127. Wang Y, Rahimnejad S, Sun WJ, et al. Bimetallic Cu-Fe catalysts on MXene for synergistically electrocatalytic conversion of nitrate to ammonia. J Colloid Interface Sci 2023;648:595-603.

128. Sun W, Li L, Zhang H, He J, Lu J. A bioinspired iron-centered electrocatalyst for selective catalytic reduction of nitrate to ammonia. ACS Sustain Chem Eng 2022;10:5958-65.

129. Sun WJ, Ji HQ, Li LX, et al. Built-in electric field triggered interfacial accumulation effect for efficient nitrate removal at ultra-low concentration and electroreduction to ammonia. Angew Chem Int Ed Engl 2021;60:22933-9.

130. Zhu D, Li G, Yan X, Geng C, Gao L. Electrochemical nitrate reduction to high-value ammonia on two-dimensional molybdenum carbide nanosheets for nitrate-containing wastewater upcycling. Sci Total Environ 2023;878:163145.

Cite This Article

Export citation file: BibTeX | RIS

OAE Style

Choi J, Im S, Choi J, Surendran S, Moon DJ, Kim JY, Kim JK, Sim U. Recent advances in 2D structured materials with defect-exploiting design strategies for electrocatalysis of nitrate to ammonia. Energy Mater 2024;4:400020. http://dx.doi.org/10.20517/energymater.2023.67

AMA Style

Choi J, Im S, Choi J, Surendran S, Moon DJ, Kim JY, Kim JK, Sim U. Recent advances in 2D structured materials with defect-exploiting design strategies for electrocatalysis of nitrate to ammonia. Energy Materials. 2024; 4(2): 400020. http://dx.doi.org/10.20517/energymater.2023.67

Chicago/Turabian Style

Choi, Jinuk, Sejin Im, Jihyun Choi, Subramani Surendran, Dae Jun Moon, Joon Young Kim, Jung Kyu Kim, Uk Sim. 2024. "Recent advances in 2D structured materials with defect-exploiting design strategies for electrocatalysis of nitrate to ammonia" Energy Materials. 4, no.2: 400020. http://dx.doi.org/10.20517/energymater.2023.67

ACS Style

Choi, J.; Im S.; Choi J.; Surendran S.; Moon DJ.; Kim JY.; Kim JK.; Sim U. Recent advances in 2D structured materials with defect-exploiting design strategies for electrocatalysis of nitrate to ammonia. Energy Mater. 2024, 4, 400020. http://dx.doi.org/10.20517/energymater.2023.67

About This Article

Special Issue

This article belongs to the Special Issue 2D Materials: From Research to Application
© The Author(s) 2024. Open Access This article is licensed under a Creative Commons Attribution 4.0 International License (https://creativecommons.org/licenses/by/4.0/), which permits unrestricted use, sharing, adaptation, distribution and reproduction in any medium or format, for any purpose, even commercially, as long as you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons license, and indicate if changes were made.

Data & Comments

Data

Views
593
Downloads
57
Citations
0
Comments
0
5

Comments

Comments must be written in English. Spam, offensive content, impersonation, and private information will not be permitted. If any comment is reported and identified as inappropriate content by OAE staff, the comment will be removed without notice. If you have any queries or need any help, please contact us at support@oaepublish.com.

0
Download PDF
Cite This Article 13 clicks
Like This Article 5 likes
Share This Article
Scan the QR code for reading!
See Updates
Contents
Figures
Related
Energy Materials
ISSN 2770-5900 (Online)
Follow Us

Portico

All published articles are preserved here permanently:

https://www.portico.org/publishers/oae/

Portico

All published articles are preserved here permanently:

https://www.portico.org/publishers/oae/