Next Article in Journal
Ethylene Dimerization Performance of NiBTCs Synthesized Using Different Solvents
Next Article in Special Issue
Pt-Based Electrocatalyst Modified by CsH2PO4/SiP2O7 for Electrochemical Oxidation of NH3 to H2 in Solid Acid Electrolysis Cell
Previous Article in Journal
Green Oxidative Catalytic Processes for the Preparation of APIs and Precursors
Previous Article in Special Issue
Photodeposition of Fe-Based Cocatalysts Capable of Effectively Promoting the Oxygen Evolution Activity of BaTaO2N
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

A Review of Transition Metal Nitride-Based Catalysts for Electrochemical Nitrogen Reduction to Ammonia

1
Department of Chemical Engineering, Department of Integrative Engineering for Hydrogen Safety, Kangwon National University, Chuncheon 24341, Republic of Korea
2
Department of Chemical Engineering, Hanyang University, Seoul 04763, Republic of Korea
*
Authors to whom correspondence should be addressed.
Catalysts 2023, 13(3), 639; https://doi.org/10.3390/catal13030639
Submission received: 20 February 2023 / Revised: 14 March 2023 / Accepted: 20 March 2023 / Published: 22 March 2023
(This article belongs to the Special Issue Theme Issue in Honor of Prof. Dr. Jae Sung Lee)

Abstract

:
Electrochemical nitrogen reduction (NRR) has attracted much attention as a promising technique to produce ammonia at ambient conditions in an environmentally benign and less energy-consuming manner compared to the current Haber–Bosch process. However, even though much research on the NRR catalysts has been conducted, their low selectivity and reaction rate still hinder the practical application of the NRR process. Among various catalysts, transition metal nitride (TMN)-based catalysts are expected to be promising catalysts for NRR. This is because the NRR process can proceed via the unique Mars–Van Krevelen (MvK) mechanism with a compressed competing hydrogen evolution reaction. However, a controversial issue exists regarding the origin of ammonia produced on TMN-based catalysts. The instability of the TMN-based catalysts can lead to ammonia generation from lattice nitrogen instead of supplied N2 gas. Thus, this review summarizes the recent progress of TMN-based catalysts for NRR, encompassing the NRR mechanism, synthetic routes, characterizations, and controversial opinions. Furthermore, future perspectives on producing ammonia electrochemically using TMN-based catalysts are provided.

1. Introduction

Next to sulfuric acid, ammonia (NH3) is the second most produced chemical in the world and is one of the essential commodities in our lives [1]. NH3 is employed in various fields, encompassing fertilizers, plastics, textiles, explosives, pharmaceuticals, dyes, and refrigerants [2,3,4]. In addition, it is considered an attractive hydrogen energy carrier due to its high gravimetric hydrogen density (17.8 wt%) and high volumetric energy density (4.32 kWh/L), similar to methanol [1,4,5,6]. Furthermore, the liquefaction point of NH3 (−33.4 °C) is much more favorable for storage and transportation compared to that of hydrogen (−253 °C), facilitating a CO2-neutral energy system [7,8].
N2(g) + 3H2(g) ↔ 2NH3(aq)     (ΔH° = −91.8 kJ/mol)
The Haber–Bosch process (HB process), for which Pritz Haber and Carl Bosch won the Nobel Prize, has been the dominant way to produce industrial ammonia since its invention in 1909 (Equation (1)) [9,10]. Most of the NH3 produced from this process (over 80%) is converted to fertilizers, enabling the growth of food production and, subsequently, of the world population [10,11]. However, there are several drawbacks to this process. The NH3 production via the HB process requires high temperatures (400–500 °C) and high pressures (150–300 atm) using Fe-based catalysts [12,13], consuming more than 1% of the total global energy [14]. Additionally, the hydrogen molecules used in the HB process are mainly produced from the steam methane reforming (SMR) of fossil fuels, emitting approximately 1.67 tons of CO2 per ton of NH3 [15,16]. Considering that the annual production of NH3 by the process was 235 million tons in 2021, the amount of CO2 emission from this process is equivalent to 390 million tons [1]. Furthermore, due to the energy- and capital-intensive characteristics of the HB process, NH3 production sectors are localized in China, Russia, the USA, and India [17]. Thus, an NH3 transportation step is required using a ship, tank truck, or pipeline, accompanied by additional cost, energy, and CO2 emissions.
Recently, much attention has been paid to electrochemical nitrogen reduction reactions (NRR) to produce ammonia to replace the current HB process. NRR can be carried out at ambient conditions over heterogeneous catalysts with lower energy consumption compared to the HB process. During the NRR process, hydrogen can be obtained from water splitting powered by renewable energy sources such as solar or wind energy. Thus, CO2 emissions can be removed at the NH3 production step [18,19,20]. Moreover, NRR enables on-site production of NH3 [21,22]; thereby, additional reduction of CO2 emissions may be achieved through the reduced need for NH3 transportation (Figure 1).
Although the NRR process is environmentally benign with zero carbon emissions and reduced energy consumption, challenging issues, such as insufficient faradaic efficiency (FE) and low NH3 yield for practical application, still exist. The FE and NH3 yields represent the selectivity for NH3 production and the NRR reaction rate (per unit time and unit catalyst mass or unit area), respectively. The equilibrium potential of the hydrogen evolution reaction (HER) is similar to that of NRR. Thus, HER becomes a competing reaction, lowering FE (i.e., selectivity) [23]. Additionally, a large overpotential is required to activate the nitrogen-nitrogen triple bond in an inert N2 molecule, resulting in a low NH3 yield (i.e., reaction rate). Until now, various catalysts, including precious metal-based materials [24,25,26,27,28,29,30,31,32], transition metal compounds (transition metal carbide [33,34,35]/nitride [36,37,38]/sulfide [39,40,41]/oxide [42,43,44]/phosphide [45,46,47]), and non-metallic series [48,49,50], have been employed for NRR to improve the FE and NH3 yields. Despite many efforts, the NRR performances of the reported catalysts are still far below practical utilization goals (current density of 300 mA cm−2, FE of 90%, and energy efficiency of 60%) according to the REFUEL program of the US Department of Energy [51]. Thus, the investigation of highly active, durable, and cost-effective catalysts for NRR is strongly required.
Currently, transition metal nitride (TMN)-based catalysts are extensively studied as a new class of non-precious metal catalysts for NRR. TMN is an interstitial alloy that incorporates nitrogen atoms into the interstitial sites of the crystal lattice of metals [52]. TMN has unique physical properties, including hardness, wear resistance, and superconductivity, which renders them suitable for use as coating agents for cutting tools and refractory materials [53,54,55]. More importantly, TMN has been employed as a catalyst in various traditional catalytic reactions to replace Pt-group metals, including hydrogenation and hydrodesulfurization, because they exhibit similar electronic structures to those of noble metals [55,56,57]. Additionally, TMNs have gained great attention in the field of energy, such as fuel cells [58,59], photocatalysts [60,61], solar cells [62,63], and water splitting [64,65]. Especially in NRR, many TMN-based catalysts have been reported as potential catalysts inspired by density functional theory (DFT) results [66,67,68,69]. The most promising candidates turned out to be VN, CrN, NbN, and ZrN because they might be more selective toward NRR than the competing HER [70].
In this review, we present an overview of the state-of-the-art TMN-based catalysts for NRR, focusing on the reaction mechanism (such as the Mars–Van Krevelen mechanism), synthetic routes, and characterization methods. In addition, conflicting opinions pointing out the inability of TMN to contribute to NRR due to their instability will also be considered. Finally, challenges and future perspectives will be discussed.

2. Mechanisms of NRR

The following equations represent the reactions for NRR in acidic (Equation (2)) and basic solutions (Equation (3)), with the equilibrium potentials referenced to the reversible hydrogen electrode (RHE). Equations (4) and (5) are for HER under acidic and alkaline media, respectively. The NRR involves six electrons and is thus kinetically unfavorable compared to the two-electron-involved HER. In addition, the equilibrium potentials of NRR are similar to those of HER in both acidic and alkaline conditions, suggesting that HER is a competitive reaction to NRR.
N2(g) + 6H+(aq) + 6e ↔ 2NH3(aq)     (E° = 0.092 V vs. RHE)
N2(g) + 6H2O(l) + 6e ↔ 2NH3 + 6OH(aq)    (E° = 0.092 V vs. RHE)
2H+(aq) + 2e ↔ H2(g)     (E° = 0.000 V vs. RHE)
2H2O(l) + 2e ↔ H2(g) + 2OH(aq)     (E° = 0.000 V vs. RHE)
Consisting of multiple protonation steps, NRR generates several intermediates during the process, which may act as a blockage, making it difficult to react thermodynamically. The initial addition of the proton and electron to N2 needs a highly negative equilibrium potential (i.e., a very high energy barrier), revealing that the formation of N2H is difficult (Equation (6)). Furthermore, the N2H can be further reduced to two possible intermediates, diazene (N2H2) and hydrazine (N2H4). The equilibrium potentials for diazene and hydrazine (Equations (7) and (8)) are larger compared to NH3 production (Equations (2) and (3)), indicating a thermodynamic difficulty in the electroreduction of N2.
N2 + H+ + e ↔ N2H     (E° = −3.20 V vs. RHE)
N2 + 2H+ + 2e ↔ N2H2(g)     (E° = −1.10 V vs. RHE)
N2 + 4H+ + 4e ↔ N2H4(g)     (E° = −0.36 V vs. RHE)
In general, mechanisms of NRR are classified as the dissociative pathway, the associative pathway, and the Mars–Van Krevelen (MvK) pathway (Figure 2). In the dissociative pathway (Figure 2a), by breaking the N≡N bond, two N atoms are first adsorbed solely on the catalyst surface, and subsequent hydrogenation occurs to produce NH3. The cleavage of the triple bond of N2 requires a large quantity of energy (945 kJ/mol) [71,72]. Given that the HB process follows this pathway, it is not surprising that the HB process needs such a high temperature, high pressure, and substantial energies. In the associative pathway, N2 molecules are first adsorbed on the catalyst surface without the N≡N bond breaking, followed by the addition of hydrogen atoms. Thus, the N–N bond is broken simultaneously with the generation of the first NH3 molecule. This mechanism can be subclassified as the distal pathway, the alternative pathway, and the enzymatic pathway. In the associative distal pathway (Figure 2b), the N2 molecule reaches the catalyst surface and is adsorbed in end-on mode. The remote N atom (i.e., the distal N atom) is hydrogenated first and released as NH3. Another N atom remaining on the catalyst surface is continuously reduced to produce the second NH3. In the associative alternative pathway (Figure 2c), which is different from the distal pathway, the H atoms are added to both N atoms, and two molecules of NH3 are released consecutively. In the enzymatic pathway (Figure 2d), the two N atoms of the N2 molecule are both adsorbed on the surface of the catalyst (side-on mode). Then two N atoms are reduced by H atoms alternatively, as demonstrated for the associative alternative pathway.
The MvK mechanism has been suggested to explain the NRR mechanism for the TMN catalysts (Figure 2e) and is in contrast to the other mechanisms. Firstly, H atoms are attached to the surface N atom of the TMNs. The N atom is electrochemically reduced, and the first NH3 is released, leaving an N vacancy on the catalyst surface. Secondly, the N-vacancies are replenished by the supplied N2 molecules, and a subsequent hydrogenation process occurs, producing the second NH3. Finally, the catalyst’s surface is regenerated, and this cycle is repeated during the NRR. Notably, the MvK mechanism possesses unique merits. The first NH3 production does not require the adsorption and cleavage of N2 on the surface of the catalyst, reducing the energy barrier for NH3 production. Furthermore, the surface energy of the catalysts with N-vacancies is higher, facilitating favorable adsorption of N2 and cleavage of the N≡N bond. For the TMN catalyst, Abghoui et al. calculated the important factors in each step, such as the initial NH3 formation, N-vacancy replenishment, catalytic cycle endurance, stability of the N vacancies, and poisoning. They claimed that the most promising catalysts were VN, CrN, ZrN, and NbN, and their rocksalt (100) facet was the most likely to be active among the various facets [70]. Because candidates have stable N vacancies and surface vacancies free from poisoning by −H or −O, higher NRR activities than those of HER can be expected.

3. TMN-Based Catalysts

3.1. TMN-Based Catalysts with Catalytic Activity

3.1.1. Vanadium Nitride-Based Catalysts

VN nanowire array on carbon cloth (VN/CC) was synthesized by annealing V2O5 precursor under NH3 flow and utilized as a catalyst for NRR [73]. The nanowire structure can be confirmed through scanning electron microscope (SEM) images (Figure 3a). The VN/CC showed an FE of 3.58% with an NH3 yield of 2.48 × 10−10 mol s−1 cm−2 at −0.3 V vs. RHE in a 0.1 M HCl solution. At potentials lower than −0.3 V, the FE and NH3 yields decreased remarkably because HER became the dominant reaction. Additionally, no hydrazine was detected through the Watt and Chrisp method, which indicates a great selectivity of the VN/CC. The FE and NH3 yields were maintained without significant performance decay over 10 times of consecutive recycling tests at −0.3 V. The stability test showed no significant changes in the catalysts’ morphology, crystallinity, or surface states, suggesting the activity and stability of the VN/CC for NRR in acidic media.
A VN nanosheet on Ti mesh (VN/TM) was fabricated through the nitridation of a VO2 nanosheet array on Ti mesh (VO2/TM) under NH3 flow for 3 h at 700 °C [74]. The SEM and transmission electron microscope (TEM) showed that the catalyst is composed of VN nanosheets, and the energy dispersive X-ray spectroscopy (EDX) mapping results confirmed the uniform distribution of the V and N elements on Ti. In the linear sweep voltammetry (LSV) results, the current density under N2 flow was higher than that under Ar flow, suggesting that VN/TM is active for NRR in 0.1 M HCl. It achieved an NH3 yield of 8.40 × 10−11 mol s−1 cm−2 at −0.5 V with an FE of 2.25% (Figure 3b). The TM itself did not show NRR activity, further suggesting that VN is an active site in VN/TM. In the EDX results after the NRR test under Ar flow for 3 h, the N in VN disappeared, indicating the MvK mechanism of VN/TM. In the NMR spectra using the 15N2 isotope, both doublet peaks of 15NH4+ and triplet peaks of 14NH4+ were detected, further implying that the NRR proceeded through the MvK mechanism on VN/TM. In addition, FE and NH3 yield values were maintained for 10 recycling tests, indicative of good stability of VN/TM toward NRR.
Yang et al. employed a membrane electrode assembly (MEA) configuration to study the NRR for VN nanoparticles due to the low solubility of N2 in an aqueous solution (Figure 3c) [37]. The VN and Pt/C catalysts were loaded on carbon paper and used as cathodes and anodes, respectively. In this setup, the Pt anode acted as a reversible hydrogen electrode. At −0.1 V, the VN catalyst exhibited an FE of 6.0% with an NH3 yield of 3.3 × 10−10 mol s−1 cm−2. A steady-state NH3 yield of 1.1 × 10−10 mol s−1 cm−2 and a 1.6% FE were maintained for 116 h. In contrast, at −0.2 V, an FE of 6.5% and an NH3 yield of 5.0 × 10−10 mol s−1 cm−2 were recorded within the first hour. However, the NH3 yield at −0.2 V plunged more than 95% to 1.1 × 10−11 mol s−1 cm−2 after 2 h. NRR tests with a 15N2 isotope proved that the reaction was conducted via the MvK mechanism. Triplet (14NH4+) and doublet (15NH4+) peaks were detected. The active phase and deactivation mechanism were identified by characterizing fresh and spent VN samples. VN0.7O0.45 is proposed as an active phase from the X-ray photoelectron spectroscopy (XPS) results because the VN0.7O0.45/VN ratios correlate well with the reactivity results (Figure 3d). Furthermore, in the operando X-ray absorption near edge structure (XANES) of the V K-edge results, the intensity of the pre-edge peak at 5468.4 eV (oxynitride species) decreased with increasing NRR reaction time, and more negative potentials accelerated the intensity decreasing rate (Figure 3e). These results indicate that the conversion of VN0.7O0.45 to VN causes the deactivation of NRR. The deactivation mechanism was hypothesized such that since the active phase, VN0.7O0.45, is unstable under reduction conditions, O atoms nearby N atoms are removed faster at a more negative potential. Hence, the generated vacancies are filled with N2, finally producing the inactive phase VN (Figure 3f).

3.1.2. Chromium Nitride-Based Catalysts

For the Cr2N and CrN catalysts, the active species and deactivation pathway were investigated using an MEA configuration, as shown in Figure 3c [75]. The commercial Cr2N (CN-P) and CrN (CN-S) synthesized by the urea-glass route have been used for NRR. The CN-P showed an NH3 yield of 1.4 × 10−11 mol s−1 cm−2 and an FE of 0.58% at −0.2 V for 24 h under 80 °C conditions. From −0.2 V to −0.8 V, the NH3 yield increased, but the FE decreased because HER became the dominant reaction as the potential increased. Additionally, the catalyst deactivation of the CN-P is more severe at a more negative potential. Over 24 h, the NH3 yield decreased by 20% at −0.2 V, while an 87% decrease at −0.8 V was observed. In the NMR spectra using the 15N2 isotope, both 15NH4+ and 14NH4+ peaks were detected, confirming the MvK mechanism on the CN-P catalyst. The N/Cr ratio measured from the XPS results was pointed out as an important parameter. After the reaction at −0.2 and −0.4 V, the N/Cr ratio was around 0.5, similar to the fresh sample ratio. However, the ratio abruptly decreased after operation at −0.6 and −0.8 V. Especially at −0.8 V, as observed by depth-profiling XPS, the N/Cr ratio was 0.2 (Figure 4a). These results indicate that N was leached out on the catalyst surface during NRR, which in turn caused the deactivation of CN-P. The NRR performance of CN-S was compared with CN-P. The FE and NH3 yields of CN-S were 4 and 5 orders of magnitude times lower than those of CN-P. According to XPS analysis, all species in CN-P were present except for the Cr2N, which provides strong evidence that Cr2N is the active phase for NRR.
Ma et al. reported that CrN is an active catalyst for NRR [38]. Here, CrN was synthesized by a glucose-mediated hydrothermal calcination route and ammonolysis. Glucose is used for creating a hollow structure of metal oxide, which results from the collapse of the carbon body during calcination. A metal precursor prepared by hydrothermal treatment was converted to Cr2O3 after calcination and subsequent nitridation under an NH3 and H2 atmosphere, and CrN nanocubes (NCs) were generated. In the LSV curves, the current density under N2 was higher than that under Ar, indicating that CrN NCs are active for NRR. After a chronoamperometry (CA) test for 2 h, CrN NCs recorded an NH3 yield of 31.11 μg mgcat−1 h−1 with an FE of 16.64% at −0.5 V. In the control experiments (at open circuit potential under N2 flow and −0.5 V under Ar flow), the produced amounts of ammonia were negligible. Furthermore, there were only small variations of FE and NH3 yields in the cycling tests that were repeated 5 times, showing the stability of CrN NCs for NRR. To understand the reaction based on the MvK mechanism, STEM-EDX analysis was conducted after the NRR tests of CrN NCs. The chemical composition showed an N-rich shell and an O-rich core. This is in contrast to what was expected to be an O-rich shell and N-rich core owing to the formation of oxide anions on the surface to replace nitride ions (leached from the lattice by the MvK mechanism). The authors explained that this resulted from the leaching of N atoms and consecutive rapid re-nitridation only at the surfaces, not the core.
Yao et al. synthesized CrOxNy via the urea-glass route [76]. The stoichiometry was estimated to be CrO0.66N0.56 from STEM-EDS. For NRR tests, a proton exchange membrane electrolyzer (PEMEL) configuration was employed, using CrO0.66N0.56 as a cathode and IrO2 as an anode. For the CrO0.66N0.56 catalyst, the highest NH3 yield was found to be 8.94 × 10−11 mol s−1 cm−2, and the highest FE of 6.7% was achieved at 2.0 V and 1.8 V, respectively (Figure 4b,c). Note that the measurements were conducted using an electrolytic cell. Meanwhile, the Cr2O3 exhibited 2.78 × 10−11 mol s−1 cm−2 at 2.0 V and 2.8% at 1.8 V, implying that N-doping on chromium oxides can lead to better NRR activity. In addition, an NH3 yield of 6.2 × 10−11 mol s−1 cm−2 at 2.0 V was obtained on pure CrN; it was assumed to come from the synergy effects between N and O atoms. In the XPS Cr3/2 spectra, the binding energies of Cr3+ and Cr6+ species in CrO0.66N0.56 shifted to higher energy levels compared to those in Cr2O3 because of the electron transfer from Cr to N. Additionally, the binding energy of CrN (575.6 eV) is 1.5 eV lower than that of CrO0.66N0.56 in the Cr3/2 spectra, indicating a charge transfer from Cr to O. On the other hand, in the N-1s spectra, the binding energy of CrO0.66N0.56 (397.9 eV) is higher than that of CrN (396.6 eV). The above results suggest that the enhanced charge transfers from Cr to O with respect to N. Thus, the N on CrN0.66N0.56 can be reduced more easily than CrN, generating ammonia and N-vacancy. The crystallinity and surface states were not changed after a 3 h electrochemical test, verifying the stability of the CrO0.66N0.56 catalysts for NRR.

3.1.3. Molybdenum Nitride-Based Catalysts

Zhang et al. prepared MoN nanosheets on carbon cloth (MoN NA/CC) as a catalyst for NRR [36]. A mixture of Na2MoO4 and thiourea was hydrothermally treated with the CC substrate, and subsequent nitridation at 800 °C for 3 h generated the MoN NA/CC. The nanosheet structure of the catalyst was observed by TEM. The NRR performance of the MoN NA/CC in 0.1 M HCl solution recorded an FE of 1.15% and an NH3 yield of 3.01 × 10−10 mol s−1 cm−2 at −0.3 V with no hydrazine generation. Furthermore, the MoN NA/CC showed good stability in 7 successive cycling tests and long-term durability tests over 27 h at −0.3 V. In addition, there was no correlation between N2 flow rates and NRR performance, indicating that NRR is not dependent on the gas-solid interface (Figure 4d). After a long-term reaction of 62 h, the absence of the N atoms in MoN under the Ar atmosphere was identified by the XPS and EDX spectra, while the maintenance of N in MoN was determined under the N2 atmosphere. Additionally, XRD patterns indicated metallic Mo formation after 62 h under Ar flow. These results suggest that the MoN NA/CC catalyzes NRR via the MvK mechanism. The detection of triplet 14NH4+ and doublet 15NH4+ signals in NMR using 15N2 results further confirmed that NRR followed the MvK mechanism.
A Mo2N nanorod was prepared by nitridation of MoO2 at 700 °C for 3 h and used as an NRR catalyst in a 0.1 M HCl electrolyte [77]. The catalyst was loaded on the glassy carbon electrode (GCE) and acted as a working electrode. At −0.3 V, Mo2N/GCE showed an NH3 yield of 78.4 μg mgcat−1 h−1 and FE of 4.5%, while MoO2/GCE exhibited an NH3 yield of 16.8 μg mgcat−1 h−1 with an FE of 1.3%. These results indicate that N plays a critical role in NRR. Indeed, DFT calculations reveal that the free energy barrier of the potential determining step (PDS) of NRR on MoO2 decreases dramatically after nitrogenization. For the MoO2 catalyst, reductive protonation of the adsorbed N2 (*N2) to *NNH is the PDS requiring 1.26 eV. However, for Mo2N, reductive protonation of *NH in the second NH3 formation is suggested as the PDS with a free energy of 0.66 eV. The FE and yield decreased below −0.3 V due to the competitive HER, which was proven by gas chromatography (GC). The produced amount of H2 at −0.1 V and −0.2 V was much less than 5 μmol, and the corresponding FE was about 30%. However, at more negative than −0.3 V, H2 production increased dramatically with an FE of 80%. The Mo2N/GCE showed no obvious performance decay during the recycling tests for 10 times and long-term reactions for 20 h, proving good stability for the NRR.
Figure 4. (a) The ratio of N to Cr in spent catalysts (Cr2N) through XPS-depth profiling. (b) J-t curves for 3 h at different potentials on CrO0.66N0.56. (c) NH3 formation rate with faradaic efficiency using CrO0.66N0.56. (d) NH3 yields and FEs with different N2 flow rates on MoN/CC at −0.3 V. (e) Schematic illustration of MV-MoN@NC synthesis through the sacrificial-hard-template route. (f) EPR spectra of MV-MoN@NC and MoN@NC based on g = 2.0018. Reproduced with permission from the American Chemical Society [36,75], Wiley [76], and Elsevier [78].
Figure 4. (a) The ratio of N to Cr in spent catalysts (Cr2N) through XPS-depth profiling. (b) J-t curves for 3 h at different potentials on CrO0.66N0.56. (c) NH3 formation rate with faradaic efficiency using CrO0.66N0.56. (d) NH3 yields and FEs with different N2 flow rates on MoN/CC at −0.3 V. (e) Schematic illustration of MV-MoN@NC synthesis through the sacrificial-hard-template route. (f) EPR spectra of MV-MoN@NC and MoN@NC based on g = 2.0018. Reproduced with permission from the American Chemical Society [36,75], Wiley [76], and Elsevier [78].
Catalysts 13 00639 g004
Yang et al. studied the role of cation vacancy for NRR by fabricating MoN nanocrystals with abundant Mo vacancy [78]. Mo-vacancy rich MoN nanocrystal on hierarchical porous carbon frame (MV-MoN@NC) is synthesized by annealing the mixture of SiO2 spheres, molybdic acid, and dicyandiamide (DCDA) under NH3 and Ar flow, and then MoN@NC connected with SiO2 is formed. In the chemical etching process of SiO2 by HF, the Si–O–Mo connection is easily removed, and the Mo vacancies are made on the MoN surface (Figure 4e). For comparison, MoN@NC is also synthesized with the same procedure, except that a polystyrene (PS) sphere was used instead of the SiO2 sphere. The MV-MoN@NC showed a 3D honeycomb structure with ~500 nm cavities, and abundant defects in the MV-MoN@NC are observed in TEM images. The electron paramagnetic resonance (EPR) results further verify the defects in the catalyst. In the EPR results, the intensity of the MV-MoN@NC peak is larger and wider than that of MoN@NC, indicating that it has more defects due to the unpaired electrons in the nitrogen atoms of MV-MoN@NC (Figure 4f). The MV-MoN@NC achieved the FE of 6.9% with an NH3 yield of 76.9 μg mgcat−1 h−1 at −0.2 V in a 0.1 M HCl solution. The MoN@NC showed lower NRR performance with an NH3 yield of 31.1 μg mg−1 h−1. After the 15N2 isotope test, the MvK mechanism on MV-MoN@NC was identified by detecting both 15NH4+ and 14NH4+ signals in NMR spectra. The enhanced NRR performance on MV-MoN@NC was investigated by DFT calculations. For the MoN (200) surface, the PDS is the reduction step of *NH to *NH2, with a relatively large energy barrier of 1.40 eV. In contrast, for the MoN (200) surface with Mo vacancies, the reductive protonation of *NH2 to *NH3 is the PDS, and its energy barrier is 0.61 eV. The reaction barrier, ΔE, is inversely proportional to the reaction rate, k, by the Arrhenius equation. Therefore, the decrease of the energy barrier from 1.40 eV to 0.61 eV by the introduction of a Mo vacancy leads to an enhancement of the reaction rate, which results in better NRR performance.

3.1.4. Titanium Nitride-Based Catalysts

Commercial TiN was treated by ball-milling (TiN-BM) and subsequent plasma-etching (TiN-PE), and the generated titanium oxynitride (TiOxNy) on TiN was pointed out as an NRR active phase [79]. The particle size of commercial TiN was 1~3 μm, and the size decreased to 100~500 nm for TiN-BM. After the plasma-etching of TiN-BM, the surface morphology and particle size of TiN-PE were similar compared to TiN-BM. In comparison with TiN, the (200) peak in the XRD patterns of the TiN-PE moved to a lower angle due to the introduction of the larger O atoms relative to the N atoms. The introduction of O atoms was also observed in the Raman spectra. The peak intensity at 130 cm−1 (vibration of Ti–O) increased in TiN-PE compared to TiN-BM and TiN samples, suggesting plasma-etching treatment facilitates the formation of TiOxNy through surface oxidation of TiN. Likewise, in the XPS results, the nitrogen content of TiN-PE was lower, and the oxygen content of TiN-PE was higher than that of TiN-BN and TiN, which indicates that plasma-etching is effective for the introduction of high O content into the TiN structure, resulting in the formation of TiOxNy. With the TiN-PE, the electrochemical test was conducted in a 0.1 M Na2SO4 solution, exhibiting an FE of 9.1% and an NH3 yield of 3.32 × 10−10 mol s−1 cm−2 at −0.6 V (Figure 5a). Furthermore, the durability test conducted at −0.6 V showed good stability for 12 h, and there was no significant change in NH3 yield and FE during 10 consecutive recycling tests. The time-dependent experiment of NH3 yield under an Ar-saturated electrolyte (3 h) and sequential N2-saturated electrolyte (3 h) supported the MvK mechanism on the TiN-PE. In the Ar-saturated electrolyte, the NH3 yield initially slowly increased and stabilized at 6.61 × 10−11 mol s−1 cm−2 within 1 h. After the electrolyte was changed to an N2-saturated solution, the NH3 yield increased rapidly, reaching 3.95 × 10−10 mol s−1 cm−2 after 3 h. The produced NH3 in an Ar-saturated solution originated from TiOxNy in TiN-PE, i.e., the surface N atoms of TiOxNy were reduced to NH3, leaving N-vacancies. In an N2-saturated solution, subsequent adsorption and activation of N2 on N-vacancies proceeded via the MvK mechanism, resulting in increased NH3 yield. NMR results with the 15N2 isotope further confirmed that the NRR proceeded with the MvK mechanism, where doublet 15NH4+ and triplet 14NH4+ signals appeared together (Figure 5b).
Two-dimensional Ti2N MXene was utilized for NRR with the expectation of a high surface area-to-volume ratio and easy access to N sites [80]. The Ti2AlN (MAX phase) was treated with a molten salt at 550 °C under Ar flow to form Ti2AlN-MST, and subsequent acid treatment for fluoride salt removal generates multi-layer Ti2N MXnene (Ti2N-ML). Further exfoliation of Ti2N-ML by sonication in water produces only single layer Ti2N MXene (Ti2N-FL) (Figure 5c). The interlayer spacing of Ti2N-FL is three times higher than that of the MAX phase based on (002) peak position in XRD patterns, meaning that the clear separation of the layers from the MAX phase, in turn, increases the surface area by exposing more N sites, leading to enhanced NRR performance. The widened peaks because of the distortion of the structure caused by the electron density of the surface terminal group in the Raman analysis confirmed the transformation from the MAX phase to the MXene phase. The NRR activity of Ti2N-FL was compared with that of Ti3CN (carbonitride) and Ti3C2 (carbide). In cyclic voltammetry (CV) measurements, Ti2N-FL and Ti3CN showed larger areas under N2 than under Ar, indicating that more NH3 could be generated because of the larger charge storage capacity. In the LSV curves, the current density of Ti2N-FL was about 10 times higher than that of Ti3CN due to a greater amount of N sites and improved reactivity. Considering that the Ti3C2 showed limited NRR performance, the ratio of N atoms to Ti atoms within the lattice structure might affect the NRR performance. Ti2N-FL showed an FE of 19.85% and an NH3 yield of 11.33 μg cm−2 h−1 at −0.25 V. In contrast, Ti3CN exhibited an FE of 0.03% with an NH3 yield of 3.352 μg cm−2 h−1 at −0.55 V. NH3 was not detected when a bare glassy carbon electrode without catalyst was used after a 4-hour CA test at −0.25 V. On the other hand, an NH3 yield of 5.24 μg cm−2 h−1 was recorded at −0.25 V under Ar flow, suggesting that NRR proceeds through the MvK mechanism.

3.1.5. Other TMN-Based Catalysts

Jin et al. used N-vacancy to introduce 2D-layered W2N3 (NV-W2N3) in NRR [81]. The W2N3 was prepared by nitridation of Na2W4O13 under an NH3 atmosphere, and additional annealing of W2N3 under a 5% H2/Ar atmosphere generates N-vacancies on the W2N3 surface (NV-W2N3). In the XPS N-1s spectra, the peak intensity at 400.2 eV (assigned to N-vacancy) of NV-W2N3 increased compared to W2N3, suggesting a higher amount of N-vacancies on NV-W2N3. Furthermore, the concentrations of the N-vacancies were calculated to be 6.6% for NV-W2N3 and 4.3% for W2N3 based on the W/N atom ratio through XPS data. Additionally, in the W-L3 edge in the extended X-ray absorption fine structure (EXFAS) spectra, the intensity of the peak for NV-W2N3 at 1.8 Å (W-N bonding) was reduced relative to W2N3 (Figure 5d), implying a decreased coordination number of the W-N and thus, a formation of N-vacancies. The NV-W2N3 achieved an NH3 yield of 3.80 ± 0.32 × 10−11 mol s−1 cm−2 with an FE of 11.67 ± 0.93% at −0.2 V in 0.1 M KOH. The W2N3 showed a lower NH3 yield compared to the NV-W2N3, revealing that N-vacancies affected the NRR activities (Figure 5e). During NRR measurements, hydrazine was not detected, and in the NMR test with the 15N2 isotope, only 15NH4+ doublet signals were observed. These results suggest that the NRR on NV-W2N3 proceeded with the alternative distal mechanism, not the MvK mechanism. In DFT calculations, the release of the second NH3 is pointed out as the PDS for NV-W2N3, pointing thermodynamically uphill with a free energy change of 0.97 eV. In comparison, the corresponding free energy change on W2N is above 2 eV, verifying that the electron-deficient area induced by the N-vacancies can reduce the thermodynamic limiting potential and thus promote the NRR performance.
Cu3N/CF was prepared by a simple one-step synthesis as an NRR catalyst in a 0.1 M Na2SO4 solution [82]. The Cu foam (downstream side) and urea (upstream side) were located separately on the alumina boat and annealed at 300 °C under Ar flow. During the annealing process, urea can release ammonia, resulting in the growth of Cu3N on the Cu foam. The Cu3N/CF showed an NH3 yield of 1.12 × 10−10 mol s−1 cm−2, and an FE of 1.5% at −0.2 V. Bare Cu foam recorded an NH3 yield of 1.0 × 10−11 mol s−1 cm−2. The Cdl values from CV measurements were 12.4 mF cm−2 for Cu3N/CF and 3.8 mF cm−2 for Cu foam, suggesting that sheet-structured Cu3N/CF had a wider electrochemically active surface area compared to Cu foam with a flat structure. The NRR activities of Cu3N/CF were maintained after electrolysis for 14 h and 5 cycling tests. After an NRR test under Ar for 4 h, the N atoms on the surface of Cu3N/CF disappeared from the EDX spectra, indicating the MvK mechanism (Figure 5f).
TMN-based composite catalysts were investigated in NRR to improve the performance of NRR. γ-Mo2N nanoparticles on a 2D hexagonal boron nitride (h-BN) were prepared by one-step nitridation with molybdenum oxide, boric acid, and urea [83]. The initial amount of molybdenum oxide affected the size of the Mo2N nanoparticle. Additionally, the larger the particle size, the more distorted the BN surface, creating B and N vacancies. The Mo atoms in γ-Mo2N hybridize with N2 by receiving the electron lone pair for which the electron-deficient nitrogen vacancies could draw electrons. In compensation, Mo atoms donate the electron to adsorbed N2 molecules through π-back donation for activation. Therefore, vacancy-driven electron transfer between γ-Mo2N/h-BN and N2 weakens the triple bond in N2, lowering the energy barrier and enhancing the faradaic efficiency. The composite γ-Mo2N on 2D-h-BN catalysts showed an NH3 yield of 35.9 μg mg−1 h−1 and FE of 61.5% at −0.3 V in a 0.1 M Na2SO4 solution.
Single Bi atom-incorporated hollow TiN nanorods encapsulated in nitrogen-doped carbon (NC) layer supported on carbon cloth (NC/Bi SAs/TiN/CC) were prepared by a 4-step process composed of BiOI electrodeposition on TiN/CC, ligand exchange, electrodeposition of aniline film, and pyrolysis with dicyandiamide at 900 °C in N2 flow [84]. The NC layers with 2 nm thickness were incorporated with single Bi atoms, which are coated on the surface of TiN. The composite NC/Bi SAs/TiN/CC catalysts showed an NH3 yield of 76.15 μg mgcat−1 h−1 at −0.8 V and FE of 24.6% at −0.5 V in a 0.1 M Na2SO4 solution. In the NMR results, doublet 15NH4+ and triplet 14NH4+ signals were observed when using 15N2 and 14N2 as the feed gases, confirming that the NH3 is produced from NRR. The DFT study suggests that there exists a strong synergistic effect between Bi SAs and TiN substrate, which could simultaneously promote the hydrogenation of N2 molecules into NH3 * on the TiN substrate and the desorption of NH3 * from single atomic Bi sites and thus boost NRR. The NRR performances and experimental conditions of the TMN-based catalysts are summarized in Table 1.

3.2. TMN-Based Catalysts with Non-Catalytic Activity (Leaching or Decomposition)

3.2.1. Vanadium Nitride-Based Catalysts

Manjunatha reported chemically and electrochemically unstable VN for NRR [85]. The NRR measurements were conducted in acidic (0.1 M HCl), neutral (0.1 M Na2SO4), and alkaline (0.1 M KOH) media with a commercial VN catalyst. All experimental instruments were cleaned with acid, and a control experiment was conducted to prevent contamination from N-source. Significant leaching of V and N elements was observed in the chemical and electrochemical experiments using VN. Notably, the concentration of V in an alkaline solution after chemical leaching was 12 times higher compared to other solutions (Figure 6a). Again, after the electrochemical reaction, V-leaching was the most severe in alkaline media, followed by acid and neutral media. After a chemical and electrochemical test in an alkaline solution, O elements, which were not present in pristine VN, were generated on VN, as evidenced by EDX results. This might be due to the replenishment of the N-vacancies with OH- ions in an alkaline solution. In acid and neutral media, O elements were not detected on the VN catalyst by EDX. Thus, N-vacancies remained unoccupied, resulting in a decreased N content compared to pristine VN (Figure 6b). Chemical leaching showed a higher NH3 yield in acidic solutions compared to neutral solutions. In contrast, electrochemical leaching showed a higher NH3 yield in neutral media compared to an acidic solution. The HER generally occurs more easily under acidic conditions and generated H2 bubbles can prevent contact between the surface of VN with the electrolyte, thereby protecting V and N from dissolution. Therefore, the authors claimed that the chemical and electrochemical instability of VN limits the application of NRR due to the influence of NH3 produced by N-leaching.
Du et al. also reported VN as an inactive catalyst for NRR [86]. The V(OH)x precursor was prepared by microwave radiation of VCl3 and urea, which was then annealed at 600 °C under NH3 flow to obtain the VN catalyst. Then, the VN was loaded onto carbon paper, which was used as electrode material. In CV tests under 0.05 M H2SO4, the current density with N2 flow was higher compared to Ar flow. However, the amount of NH3 produced after CV was similar to the amount of NH3 produced after the CA test for 2 h at −0.6 V under an N2-saturated solution. Furthermore, a similar amount of NH3 was detected under the Ar-saturated solution after the CA test, suggesting that the produced NH3 was irrelevant to N2 gas and possibly originated from the lattice N atoms. In addition, adsorbed NH3 (which was absorbed on the catalyst surface after the electrochemical reaction) can be detected by soaking the electrode. The adsorbed NH3 was detected even without an applied potential under an Ar-saturated solution, provided that the produced NH3 is not related to the electrochemical reaction but mainly related to the unaided decomposition of the catalyst (i.e., the stability of the catalyst). For further confirmation, consecutive tests were conducted. In the 1st test with N2, NH3 was generated (17 mmol molcat−1) after the CV test. However, during the subsequent CA test for 2 h, no NH3 was produced. Just a few “adsorbed” NH3 molecules were released (5 mmol molcat−1) by soaking the electrode. In the 2nd test with Ar and the 3rd test with N2, the trend is identical to the 1st test, meaning that NH3 was detected only after the CV test and no NH3 was produced after the subsequent CA test. Notably, the produced amount of NH3 gradually decreased as the tests proceeded from the 1st to the 3rd, suggesting the losses of surface N sites of VN. After these tests, a new peak at 402 eV corresponding to adsorbed ammonia or another form of protonated nitrogen was detected in the XPS N-1s spectra, indicating that N-vacancies cannot accept N2 continuously from the electrolyte and thus, VN cannot catalyze the NRR.
Figure 6. (a) Comparison of leached vanadium in VN after chemical and electrochemical reactions in acidic, neutral, and alkaline media. (b) The proposed NH3 production mechanism of VN in each condition. (c) CV curves of CrN NPs under N2- and Ar-saturated solutions. (d) Cr3+ concentration and dissolution ratio at −0.4 V and −0.6 V for 1 h electrolysis. (e) XPS N-1s spectra after electrolysis for 6 h at −0.6 V. Reproduced with permission from Wiley [85] and the Royal Society of Chemistry [87].
Figure 6. (a) Comparison of leached vanadium in VN after chemical and electrochemical reactions in acidic, neutral, and alkaline media. (b) The proposed NH3 production mechanism of VN in each condition. (c) CV curves of CrN NPs under N2- and Ar-saturated solutions. (d) Cr3+ concentration and dissolution ratio at −0.4 V and −0.6 V for 1 h electrolysis. (e) XPS N-1s spectra after electrolysis for 6 h at −0.6 V. Reproduced with permission from Wiley [85] and the Royal Society of Chemistry [87].
Catalysts 13 00639 g006

3.2.2. Chromium Nitride-Based Catalysts

The stability of lattice N and the deactivation path of CrN nanoparticles (CrN NP) were investigated [87]. Through the nitridation of the Cr-urea xerogel precursor at 600 °C in the NH3 atmosphere, the CrN NP was prepared. In the CV results of CrN NP, the current density in N2 was higher than that in Ar with negligible HER activity, which indicates that CrN is active for NRR (Figure 6c). After conducting CA tests for 1 h, an FE of 15.12% with an NH3 yield of 7.47 μg h−1 mg−1 was achieved at −0.5 V. At potentials below −0.6 V, the FE and NH3 yield decreased because of the increased HER activity and N atom loss from the CrN lattice. Notably, produced NH3 was detected at all potentials under Ar flow, suggesting unavoidable lattice N leakage from CrN. To identify the stability of lattice N, cycling tests were carried out at −0.4 V and −0.6 V. During 6 consecutive tests for 1 h, there were continuous decreases in activity at both potentials, and the tendency of activity decay was alleviated from the third cycle with steady NH3 production. At the end of the cycle, the NH3 yields at both potentials were similar, which was insufficient to explain the potential-induced changes. To find out other reasons for the activity decay according to the potentials, inductively coupled plasma-mass spectroscopy (ICP-MS) was employed, which revealed that the overall collapse of the structure is an important reason for the decay. The detected amount of Cr3+ at −0.6 V was 20 times higher than that measured at −0.4 V (Figure 6d), indicating that a more negative potential accelerates the collapse of the structure. In the XPS results conducted after the cycling test, the N/Cr ratios were determined to be 4.35 at −0.4 V and 3.34 at −0.6 V, which were much higher than the 0.57 of fresh CrN. The N/Cr ratio incensement was due to the adsorbed NH3 on the catalyst surface, as identified by the peak at 400 eV in the XPS N-1s spectra (Figure 6e). In the NH3-TPD experiments, adsorbed NH3 on CrN was desorbed at 658 °C, revealing that the bond between CrN and NH3 is strong, and thus the strong bond between the CrN surface and NH3 resulted in catalyst poisoning. Hence, two possible deactivation mechanisms were suggested in this research: one is a potential-induced overall collapse of the structure, and the other is NH3 poisoning of the CrN.

3.2.3. Molybdenum Nitride-Based Catalysts

Hu et al. synthesized a Mo2N catalyst via the urea-glass route as a catalyst for NRR, which showed a pure tetragonal Mo2N crystalline structure [88]. Titanium (Ti) was selected as a substrate due to its inactivity for HER. Electrochemical tests were conducted in N2- and Ar-saturated 0.1 M HCl solutions. At potentials below −0.22 V, the current density in the N2-saturated solution was slightly higher compared to the Ar-saturated solution in the LSV curves. The Mo2N-loaded Ti (Mo2N@Ti) catalyst exhibited an FE of 42.3% with an NH3 yield of 1.99 μg mg−1 h−1 at −0.05 V and an FE of 28.4% with an NH3 yield of 2.73 μg mg−1 h−1 at 0.05 V. At −0.15 V, the FE substantially decreased to 1.93%. However, in the NMR tests using 15N2 gas, only 14NH4+ triplet signals were observed without the 15NH4+ doublet signals, which implies that the produced NH3 originated not from the catalytic reaction but from the decomposition of Mo2N (Figure 7a). Even after incubating the Mo2N@Ti electrode for 1 h without bias, a similar amount of NH3 was produced in all electrolytes (0.1 M HCl, 0.5 M NaCl, and 0.1 M NaOH) compared to the NH3 amount measured after the reaction at 0.05 V (Figure 7b). These results indicate the decomposition of Mo2N in aqueous media. Moreover, it had a more severe instability in the alkaline solution and was dissolved entirely after 1 week of incubation in 0.1 M NaOH.

3.2.4. Niobium Nitride-Based Catalysts

Du et al. investigated the non-catalytic activity of NbN and Nb4N5 for NRR [86]. The NbN catalyst was prepared by the urea-glass route, and the particles are clustered due to the high annealing temperature (1000 °C). On the surface of NbN, oxide species, as well as nitrides, were detected by XPS resulting from the unavoidable surface oxidation. The surface oxide species of the NbN were removed by H3PO4 treatment and tested for NRR. CV and CA tests were conducted in acidic, neutral, and alkaline media. Although the specific surface area of the NbN was low, the NbN loaded on the electrode was sufficient to produce NH3. However, NH3 was just marginally produced, indicating that the NbN has no catalytic activity for NRR. Because of the high synthetic temperature of 1000 °C, NbN possesses a highly stable structure that cannot generate surface N-vacancies. Meanwhile, Nb4N5 on carbon cloth (Nb4N5/CC) was synthesized by microwave irradiation at 200 °C and subsequent annealing at 700 °C under an NH3 atmosphere. For the Nb4N5/CC, the produced amount of NH3 after the CV test in a 0.05 M H2SO4 solution was similar under N2 and Ar flows, suggesting that NH3 might be produced from the release of N atoms in the crystal lattice. This can be explained by the MvK mechanism without a catalytic reaction. The generated NH3 in the first step originates from the surface nitrogen, leaving an N vacancy. However, the N-vacancy is blocked by other species and not replenished by N2. Hence, it does not proceed as a catalytic reaction. In addition, adsorbed ammonia was produced not only after the electrochemical reaction but also when no potential was applied, suggesting that it was also affected by the decomposition of the catalyst. Over 3 consecutive tests (each test includes the amount of NH3 after CV and adsorbed NH3), the NH3 yield decreased from the first to the third test (Figure 7c,d). In the first test, the amount of adsorbed NH3 (90 mmol molcat−1) was higher than the amount after CV (60 mmol molcat−1), and it repeated during the tests, suggesting an instability of the Nb4N5. After the 3 consecutive tests, the physicochemical properties of the Nb4N5/CC have changed compared to the fresh catalyst. The morphology was unevenly agglomerated in the SEM images, and the crystalline peaks weakened in XRD. In the XPS results, the intensity of the peak around 204~205 eV for Nb4N5 in the Nb 3d spectra decreased, and the peak intensities regarding oxygen species (207~208 eV) increased (Figure 7e). In the N 1s spectra, a peak at 402 eV corresponding to adsorbed ammonia was generated. The characterization results for Nb4N5/CC indicate weak structural stability during the electrochemical reaction, implying that Nb4N5/CC is not suitable for NRR.

4. Summary and Perspective

NH3 is an indispensable chemical in various industrial fields. The most commonly employed method for NH3 production is the HB process, which consumes a lot of energy with heavy CO2 emissions. The NRR is a promising alternative to produce NH3 to overcome the shortcomings of the HB process. Despite the merits of NRR, including environmental friendliness with carbon neutrality and cost-effectiveness with reduced energy consumption, insufficient FE and low NH3 yields still limit its practical application. Obviously, the development of highly active and durable catalysts is of the utmost importance to realizing NRR at a practical level. In this review, we have summarized the recent research progress of TMN-based catalysts for NRR and provided information on their mechanism, synthetic routes, and characterization methods. Furthermore, contrasting views were considered, which indicated the inadequacy of the TMN-based catalysts for NRR due to chemical and electrochemical instability.
To produce NH3 electrochemically using TMN-based catalysts, the following points should be considered: First, the rigorous procedure to detect NH3 suggested by Chorkendorff and Simonov should be employed [89,90]. As we have introduced, there exists a controversial issue in the NRR activity of TMN-based catalysts (MvK mechanism or leaching). Furthermore, during the synthetic procedure of the TMN-based catalysts, N-containing chemicals are vitally used and can be a cause for false positives, influencing the NH3 yield. Accordingly, careful control experiments should be conducted, especially for the TMN-based catalysts, to identify the origin of the NH3 and quantify the amount of NH3 produced. The 15N2 isotope test can provide a distinct clue as to whether NH3 is produced from the supplied N2 gas. Indeed, 15N2 experiments are becoming imperative to ensure the reliability of the NRR results since 2019. On the other hand, to verify that the produced NH3 originated from the leaching of the TMN-based catalysts, additional experiments, such as ICP-OES, are suggested to measure the concentration of metal ions and correlate them with the produced amount of NH3.
Second, the NRR mechanism should be further investigated in detail using advanced in situ characterization techniques. The NRR is a complex process involving multiple proton-coupled electron transfer steps, generating various intermediates on the catalyst surface. Thereby, the NRR mechanism has not been deeply understood so far, and currently, proposed mechanisms are mainly dependent on the computational simulation. Advanced in situ techniques, such as in situ Raman spectroscopy or in situ XAS, could provide useful information regarding the reaction intermediates, active sites, and, thus, NRR mechanism. By combining in situ characterization techniques with computational simulations, the development of highly efficient catalysts for NRR might be more achievable.
The above first and second points could help to clarify the debate regarding the origin of produced NH3 on TMN-based catalysts in the NRR. The third point is designing TMN-based catalysts with high activity and stability for NRR. To facilitate the NRR (via the MvK mechanism) while suppressing the inevitable competing HER reaction, several strategies can be considered. Defect engineering, including heteroatom doping and N-vacancy formation, can modify the electronic structures (i.e., intrinsic activity), regulating the binding energies of N2 and intermediates [91,92,93]. Additionally, morphology engineering techniques such as 2D materials, single-atom catalysts, facet-controlled catalysts, and composite catalysts (with catalyst supports) can help to increase the NRR activity by exposing more active sites and improving the contact area with the electrolyte.
Advanced system engineering is necessary in addition to catalyst design to enhance the NRR performance. The low solubility of N2 (0.0126 mg∙g−1) in H2O results in a low concentration of N2 molecules near the catalyst surface in an aqueous electrolyte system, resulting in low FE and NH3 yields. One effective solution is the use of a gas diffusion electrode (GDE) to increase the local N2 concentration and improve the NRR performance. Indeed, the design and implementation of GDE in NRR significantly affected its performances [23,94]. Moreover, the use of GDE can decrease the uncertainty by transferring low-current density aqueous electrolyte systems to high-current density practical GDE-based devices. Additionally, the GDE implementation can lead to a reduction in the cost of isotope experiments compared to the typical aqueous cell system by decreasing the required gas flow rate [95,96].
NRR is a strong option to produce NH3, substituting the HB process. To realize the NRR at a practical level, highly active and durable catalysts for the NRR are crucial. Although the TMN-based catalysts are considered promising candidates due to their unique MvK mechanism, a careful approach is required because of the controversial views on the origin of the produced NH3. By combining a rigorous procedure for NH3 detection and in situ characterizations, researchers can gain clear insight into the origin of NH3 and NRR mechanisms on the TMN-based catalysts. In turn, designing efficient TMN-based catalysts might be possible.

Author Contributions

Conceptualization, S.Y.P. and D.H.Y.; methodology, S.Y.P. and D.H.Y.; software, S.Y.P., Y.J.J. and D.H.Y.; validation, S.Y.P., Y.J.J. and D.H.Y.; investigation, S.Y.P., Y.J.J. and D.H.Y.; resources, Y.J.J. and D.H.Y.; data curation, S.Y.P. and D.H.Y.; writing—original draft preparation, S.Y.P., Y.J.J. and D.H.Y.; writing—review and editing, S.Y.P., Y.J.J. and D.H.Y.; visualization, S.Y.P., Y.J.J. and D.H.Y.; supervision, Y.J.J. and D.H.Y.; project administration, Y.J.J. and D.H.Y.; funding acquisition, D.H.Y. All authors have read and agreed to the published version of the manuscript.

Funding

This study was supported by the National Research Foundation of Korea (NRF) and grant-funded by the Korean government (Ministry of Education) (2019R1I1A3A01052741). This work was also supported by the Korean Institute of Energy Technology Evaluation and Planning (KETEP) and the Ministry of Trade, Industry, and Energy (MOTIE) of the Republic of Korea (No. 20224000000080).

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Ghavam, S.; Vahdati, M.; Wilson, I.; Styring, P. Sustainable ammonia production processes. Front. Energy Res. 2021, 9, 34. [Google Scholar] [CrossRef]
  2. Afif, A.; Radenahmad, N.; Cheok, Q.; Shams, S.; Kim, J.H.; Azad, A.K. Ammonia-fed fuel cells: A comprehensive review. Renew. Sustain. Energy Rev. 2016, 60, 822–835. [Google Scholar] [CrossRef]
  3. Erisman, J.W.; Sutton, M.A.; Galloway, J.; Klimont, Z.; Winiwarter, W. How a century of ammonia synthesis changed the world. Nat. Geosci. 2008, 1, 636–639. [Google Scholar] [CrossRef]
  4. Bird, F.; Clarke, A.; Davies, P.; Surkovic, E. Ammonia: Zero-Carbon Fertiliser, Fuel and Energy Store. In The Royal Society; Policy Briefing: London, UK, 2020. [Google Scholar]
  5. Wang, X.; Yang, J.; Salla, M.; Xi, S.; Yang, Y.; Li, M.; Zhang, F.; Zhu, M.K.; Huang, S.; Huang, S. Redox-Mediated Ambient Electrolytic Nitrogen Reduction for Hydrazine and Ammonia Generation. Angew. Chem. Int. Ed. 2021, 60, 18721–18727. [Google Scholar] [CrossRef]
  6. Guo, X.; Zhu, Y.; Ma, T. Lowering reaction temperature: Electrochemical ammonia synthesis by coupling various electrolytes and catalysts. J. Energy Chem. 2017, 26, 1107–1116. [Google Scholar] [CrossRef]
  7. Chatterjee, S.; Parsapur, R.K.; Huang, K.-W. Limitations of Ammonia as a Hydrogen Energy Carrier for the Transportation Sector. ACS Energy Lett. 2021, 6, 4390–4394. [Google Scholar] [CrossRef]
  8. 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, 672–687. [Google Scholar] [CrossRef]
  9. Humphreys, J.; Lan, R.; Tao, S. Development and recent progress on ammonia synthesis catalysts for Haber–Bosch process. Adv. Energy Sustain. Res. 2021, 2, 2000043. [Google Scholar] [CrossRef]
  10. Qing, G.; Ghazfar, R.; Jackowski, S.T.; Habibzadeh, F.; Ashtiani, M.M.; Chen, C.-P.; Smith III, M.R.; Hamann, T.W. Recent advances and challenges of electrocatalytic N2 reduction to ammonia. Chem. Rev. 2020, 120, 5437–5516. [Google Scholar] [CrossRef]
  11. Faria, J.A. Renaissance of ammonia synthesis for sustainable production of energy and fertilizers. Curr. Opin. Green Sustain. Chem. 2021, 29, 100466. [Google Scholar] [CrossRef]
  12. Cao, N.; Zheng, G. Aqueous electrocatalytic N2 reduction under ambient conditions. Nano Res. 2018, 11, 2992–3008. [Google Scholar] [CrossRef]
  13. Guo, X.; Du, H.; Qu, F.; Li, J. Recent progress in electrocatalytic nitrogen reduction. J. Mater. Chem. A 2019, 7, 3531–3543. [Google Scholar] [CrossRef]
  14. Chen, J.G.; Crooks, R.M.; Seefeldt, L.C.; Bren, K.L.; Bullock, R.M.; Darensbourg, M.Y.; Holland, P.L.; Hoffman, B.; Janik, M.J.; Jones, A.K. Beyond fossil fuel–driven nitrogen transformations. Science 2018, 360, eaar6611. [Google Scholar] [CrossRef] [Green Version]
  15. Smith, C.; Hill, A.K.; Torrente-Murciano, L. Current and future role of Haber–Bosch ammonia in a carbon-free energy landscape. Energy Environ. Sci. 2020, 13, 331–344. [Google Scholar] [CrossRef]
  16. Bicer, Y.; Dincer, I.; Zamfirescu, C.; Vezina, G.; Raso, F. Comparative life cycle assessment of various ammonia production methods. J. Clean. Prod. 2016, 135, 1379–1395. [Google Scholar] [CrossRef]
  17. Egenhofer, C.; Schrefler, L.; Rizos, V.; Marcu, A.; Genoese, F.; Renda, A.; Wieczorkiewicz, J.; Roth, S.; Infelise, F.; Luchetta, G. The Composition and Drivers of Energy Prices and Costs in Energy-Intensive Industries: The Case of Ceramics, Glass and Chemicals; Centre for European Policy Studies: Brussels, Belgium, 2014. [Google Scholar]
  18. Service, R.F. Liquid sunshine. Science 2018, 361, 120–123. [Google Scholar] [CrossRef]
  19. Wang, L.; Xia, M.; Wang, H.; Huang, K.; Qian, C.; Maravelias, C.T.; Ozin, G.A. Greening Ammonia toward the Solar Ammonia Refinery. Joule 2018, 2, 1055–1074. [Google Scholar] [CrossRef] [Green Version]
  20. Bouckaert, S.; Pales, A.F.; McGlade, C.; Remme, U.; Wanner, B.; Varro, L.; D’Ambrosio, D.; Spencer, T. Net Zero by 2050: A Roadmap for the Global Energy Sector; International Energy Agency: Paris, France, 2021. [Google Scholar]
  21. Tang, C.; Qiao, S.-Z. How to explore ambient electrocatalytic nitrogen reduction reliably and insightfully. Chem. Soc. Rev. 2019, 48, 3166–3180. [Google Scholar] [CrossRef]
  22. Foster, S.L.; Bakovic, S.I.P.; Duda, R.D.; Maheshwari, S.; Milton, R.D.; Minteer, S.D.; Janik, M.J.; Renner, J.N.; Greenlee, L.F. Catalysts for nitrogen reduction to ammonia. Nat. Catal. 2018, 1, 490–500. [Google Scholar] [CrossRef]
  23. Hu, L.; Xing, Z.; Feng, X. Understanding the electrocatalytic interface for ambient ammonia synthesis. ACS Energy Lett. 2020, 5, 430–436. [Google Scholar] [CrossRef]
  24. Li, S.-J.; Bao, D.; Shi, M.-M.; Wulan, B.-R.; Yan, J.-M.; Jiang, Q. Amorphizing of Au Nanoparticles by CeOx–RGO Hybrid Support towards Highly Efficient Electrocatalyst for N2 Reduction under Ambient Conditions. Adv. Mater. 2017, 29, 1700001. [Google Scholar] [CrossRef] [PubMed]
  25. Bao, D.; Zhang, Q.; Meng, F.-L.; Zhong, H.-X.; Shi, M.-M.; Zhang, Y.; Yan, J.-M.; Jiang, Q.; Zhang, X.-B. Electrochemical Reduction of N2 under Ambient Conditions for Artificial N2 Fixation and Renewable Energy Storage Using N2/NH3 Cycle. Adv. Mater. 2017, 29, 1604799. [Google Scholar] [CrossRef] [PubMed]
  26. Wang, D.; Azofra, L.M.; Harb, M.; Cavallo, L.; Zhang, X.; Suryanto, B.H.; MacFarlane, D.R. Energy-efficient nitrogen reduction to ammonia at low overpotential in aqueous electrolyte under ambient conditions. ChemSusChem 2018, 11, 3416–3422. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  27. Ruiyi, L.; Keyang, H.; Pengwu, X.; Wendong, W.; Nana, L.; Haiyan, Z.; Zaijun, L.; Xiaohao, L. Synthesis of a ruthenium–graphene quantum dot–graphene hybrid as a promising single-atom catalyst for electrochemical nitrogen reduction with ultrahigh yield rate and selectivity. J. Mater. Chem. A 2021, 9, 24582–24589. [Google Scholar] [CrossRef]
  28. Huang, H.; Xia, L.; Shi, X.; Asiri, A.M.; Sun, X. Ag nanosheets for efficient electrocatalytic N2 fixation to NH3 under ambient conditions. ChemComm 2018, 54, 11427–11430. [Google Scholar] [CrossRef]
  29. Li, W.; Li, K.; Ye, Y.; Zhang, S.; Liu, Y.; Wang, G.; Liang, C.; Zhang, H.; Zhao, H. Efficient electrocatalytic nitrogen reduction to ammonia with aqueous silver nanodots. Commun. Chem. 2021, 4, 10. [Google Scholar] [CrossRef]
  30. Wang, X.; Luo, M.; Lan, J.; Peng, M.; Tan, Y. Nanoporous Intermetallic Pd3Bi for Efficient Electrochemical Nitrogen Reduction. Adv. Mater. 2021, 33, 2007733. [Google Scholar] [CrossRef]
  31. Xie, H.; Geng, Q.; Zhu, X.; Luo, Y.; Chang, L.; Niu, X.; Shi, X.; Asiri, A.M.; Gao, S.; Wang, Z. PdP2 nanoparticles–reduced graphene oxide for electrocatalytic N2 conversion to NH3 under ambient conditions. J. Mater. Chem. A 2019, 7, 24760–24764. [Google Scholar] [CrossRef]
  32. Liu, H.-M.; Han, S.-H.; Zhao, Y.; Zhu, Y.-Y.; Tian, X.-L.; Zeng, J.-H.; Jiang, J.-X.; Xia, B.Y.; Chen, Y. Surfactant-free atomically ultrathin rhodium nanosheet nanoassemblies for efficient nitrogen electroreduction. J. Mater. Chem. A 2018, 6, 3211–3217. [Google Scholar] [CrossRef]
  33. Liu, Y.; Zhu, X.; Zhang, Q.; Tang, T.; Zhang, Y.; Gu, L.; Li, Y.; Bao, J.; Dai, Z.; Hu, J.-S. Engineering Mo/Mo2C/MoC hetero-interfaces for enhanced electrocatalytic nitrogen reduction. J. Mater. Chem. A 2020, 8, 8920–8926. [Google Scholar] [CrossRef]
  34. Feng, J.; Zhu, X.; Chen, Q.; Xiong, W.; Chen, X.; Luo, Y.; Alshehri, A.A.; Alzahrani, K.A.; Jiang, Z.; Li, W. Ultrasmall V8C7 nanoparticles embedded in conductive carbon for efficient electrocatalytic N2 reduction toward ambient NH3 production. J. Mater. Chem. A 2019, 7, 26227–26230. [Google Scholar] [CrossRef]
  35. Ren, X.; Zhao, J.; Wei, Q.; Ma, Y.; Guo, H.; Liu, Q.; Wang, Y.; Cui, G.; Asiri, A.M.; Li, B. High-performance N2-to-NH3 conversion electrocatalyzed by Mo2C nanorod. ACS Cent. Sci. 2018, 5, 116–121. [Google Scholar] [CrossRef] [Green Version]
  36. Zhang, L.; Ji, X.; Ren, X.; Luo, Y.; Shi, X.; Asiri, A.M.; Zheng, B.; Sun, X. Efficient electrochemical N2 reduction to NH3 on MoN nanosheets array under ambient conditions. ACS Sustain. Chem. Eng. 2018, 6, 9550–9554. [Google Scholar] [CrossRef]
  37. Yang, X.; Nash, J.; Anibal, J.; Dunwell, M.; Kattel, S.; Stavitski, E.; Attenkofer, K.; Chen, J.G.; Yan, Y.; Xu, B. Mechanistic insights into electrochemical nitrogen reduction reaction on vanadium nitride nanoparticles. J. Am. Chem. Soc. 2018, 140, 13387–13391. [Google Scholar] [CrossRef]
  38. Ma, Z.; Chen, J.; Luo, D.; Thersleff, T.; Dronskowski, R.; Slabon, A. Structural evolution of CrN nanocube electrocatalysts during nitrogen reduction reaction. Nanoscale 2020, 12, 19276–19283. [Google Scholar] [CrossRef]
  39. Zhang, L.; Ji, X.; Ren, X.; Ma, Y.; Shi, X.; Tian, Z.; Asiri, A.M.; Chen, L.; Tang, B.; Sun, X. Electrochemical ammonia synthesis via nitrogen reduction reaction on a MoS2 catalyst: Theoretical and experimental studies. Adv. Mater. 2018, 30, 1800191. [Google Scholar] [CrossRef]
  40. Kong, J.; Kim, M.-S.; Akbar, R.; Park, H.Y.; Jang, J.H.; Kim, H.; Hur, K.; Park, H.S. Electrochemical nitrogen reduction kinetics on a copper sulfide catalyst for NH3 synthesis at low temperature and atmospheric pressure. ACS Appl. Mater. Interfaces 2021, 13, 24593–24603. [Google Scholar] [CrossRef]
  41. Zhao, X.; Lan, X.; Yu, D.; Fu, H.; Liu, Z.; Mu, T. Deep eutectic-solvothermal synthesis of nanostructured Fe3S4 for electrochemical N2 fixation under ambient conditions. ChemComm 2018, 54, 13010–13013. [Google Scholar] [CrossRef]
  42. Han, J.; Ji, X.; Ren, X.; Cui, G.; Li, L.; Xie, F.; Wang, H.; Li, B.; Sun, X. MoO3 nanosheets for efficient electrocatalytic N2 fixation to NH3. J. Mater. Chem. A 2018, 6, 12974–12977. [Google Scholar] [CrossRef]
  43. Liu, Q.; Zhang, X.; Zhang, B.; Luo, Y.; Cui, G.; Xie, F.; Sun, X. Ambient N2 fixation to NH3 electrocatalyzed by a spinel Fe3O4 nanorod. Nanoscale 2018, 10, 14386–14389. [Google Scholar] [CrossRef]
  44. Kong, J.; Lim, A.; Yoon, C.; Jang, J.H.; Ham, H.C.; Han, J.; Nam, S.; Kim, D.; Sung, Y.-E.; Choi, J.; et al. Electrochemical Synthesis of NH3 at Low Temperature and Atmospheric Pressure Using a γ-Fe2O3 Catalyst. ACS Sustain. Chem. Eng. 2017, 5, 10986–10995. [Google Scholar] [CrossRef]
  45. Mou, T.; Liang, J.; Ma, Z.; Zhang, L.; Lin, Y.; Li, T.; Liu, Q.; Luo, Y.; Liu, Y.; Gao, S.; et al. High-efficiency electrohydrogenation of nitric oxide to ammonia on a Ni2P nanoarray under ambient conditions. J. Mater. Chem. A 2021, 9, 24268–24275. [Google Scholar] [CrossRef]
  46. Guo, W.; Liang, Z.; Zhao, J.; Zhu, B.; Cai, K.; Zou, R.; Xu, Q. Hierarchical cobalt phosphide hollow nanocages toward electrocatalytic ammonia synthesis under ambient pressure and room temperature. Small Methods 2018, 2, 1800204. [Google Scholar] [CrossRef]
  47. Zhu, X.; Wu, T.; Ji, L.; Liu, Q.; Luo, Y.; Cui, G.; Xiang, Y.; Zhang, Y.; Zheng, B.; Sun, X. Unusual electrochemical N2 reduction activity in an earth-abundant iron catalyst via phosphorous modulation. ChemComm 2020, 56, 731–734. [Google Scholar] [CrossRef]
  48. Lan, J.; Luo, M.; Han, J.; Peng, M.; Duan, H.; Tan, Y. Nanoporous B13C2 towards highly efficient electrochemical nitrogen fixation. Small 2021, 17, 2102814. [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. Song, P.; Wang, H.; Kang, L.; Ran, B.; Song, H.; Wang, R. Electrochemical nitrogen reduction to ammonia at ambient conditions on nitrogen and phosphorus co-doped porous carbon. ChemComm 2019, 55, 687–690. [Google Scholar] [CrossRef]
  51. MacFarlane, D.R.; Cherepanov, P.V.; Choi, J.; Suryanto, B.H.R.; Hodgetts, R.Y.; Bakker, J.M.; Ferrero Vallana, F.M.; Simonov, A.N. A Roadmap to the Ammonia Economy. Joule 2020, 4, 1186–1205. [Google Scholar] [CrossRef]
  52. Chen, J.G. Carbide and Nitride Overlayers on Early Transition Metal Surfaces:  Preparation, Characterization, and Reactivities. Chem. Rev. 1996, 96, 1477–1498. [Google Scholar] [CrossRef]
  53. Yang, M.; Allen, A.J.; Nguyen, M.T.; Ralston, W.T.; MacLeod, M.J.; DiSalvo, F.J. Corrosion behavior of mesoporous transition metal nitrides. J. Solid State Chem. 2013, 205, 49–56. [Google Scholar] [CrossRef]
  54. Marchand, R.; Tessier, F.; DiSalvo, F.J. New routes to transition metal nitrides: And characterization of new phases. J. Mater. Chem. 1999, 9, 297–304. [Google Scholar] [CrossRef]
  55. Youn, D.H.; Bae, G.; Han, S.; Kim, J.Y.; Jang, J.-W.; Park, H.; Choi, S.H.; Lee, J.S. A highly efficient transition metal nitride-based electrocatalyst for oxygen reduction reaction: TiN on a CNT–graphene hybrid support. J. Mater. Chem. A 2013, 1, 8007–8015. [Google Scholar] [CrossRef]
  56. Xie, J.; Xie, Y. Transition Metal Nitrides for Electrocatalytic Energy Conversion: Opportunities and Challenges. Chem. Eur. J. 2016, 22, 3588–3598. [Google Scholar] [CrossRef]
  57. Dongil, A.B. Recent Progress on Transition Metal Nitrides Nanoparticles as Heterogeneous Catalysts. Nanomaterials 2019, 9, 1111. [Google Scholar] [CrossRef] [Green Version]
  58. Dong, Y.; Deng, Y.; Zeng, J.; Song, H.; Liao, S. A high-performance composite ORR catalyst based on the synergy between binary transition metal nitride and nitrogen-doped reduced graphene oxide. J. Mater. Chem. A 2017, 5, 5829–5837. [Google Scholar] [CrossRef]
  59. Luo, J.; Tian, X.; Zeng, J.; Li, Y.; Song, H.; Liao, S. Limitations and Improvement Strategies for Early-Transition-Metal Nitrides as Competitive Catalysts toward the Oxygen Reduction Reaction. ACS Catal. 2016, 6, 6165–6174. [Google Scholar] [CrossRef]
  60. Go, H.; Akio, I.; Tsuyoshi, T.; Kondo, J.N.; Michikazu, H.; Kazunari, D. Ta3N5 as a Novel Visible Light-Driven Photocatalyst (λ < 600 nm). Chem. Lett. 2002, 31, 736–737. [Google Scholar]
  61. Kim, J.Y.; Lee, M.H.; Kim, J.-H.; Kim, C.W.; Youn, D.H. Facile nanocrystalline Ta3N5 synthesis for photocatalytic dye degradation under visible light. Chem. Phys. Lett. 2020, 738, 136900. [Google Scholar] [CrossRef]
  62. Seol, M.; Youn, D.H.; Kim, J.Y.; Jang, J.-W.; Choi, M.; Lee, J.S.; Yong, K. Mo-Compound/CNT-Graphene Composites as Efficient Catalytic Electrodes for Quantum-Dot-Sensitized Solar Cells. Adv. Energy Mater. 2014, 4, 1300775. [Google Scholar] [CrossRef]
  63. Youn, D.; Seol, M.; Kim, J.; Jang, J.-W.; Choi, Y.; Yong, K.; Lee, J.S. TiN Nanoparticles on CNT-Graphene Hybrid Support as Noble-Metal-Free Counter Electrode for Quantum-Dot-Sensitized Solar Cells. ChemSusChem 2013, 6, 261–267. [Google Scholar] [CrossRef]
  64. Park, S.H.; Jo, T.H.; Lee, M.H.; Kawashima, K.; Mullins, C.B.; Lim, H.-K.; Youn, D.H. Highly active and stable nickel–molybdenum nitride (Ni2Mo3N) electrocatalyst for hydrogen evolution. J. Mater. Chem. A 2021, 9, 4945–4951. [Google Scholar] [CrossRef]
  65. Park, S.H.; Kang, S.H.; Youn, D.H. Direct One-Step Growth of Bimetallic Ni2Mo3N on Ni Foam as an Efficient Oxygen Evolution Electrocatalyst. Materials 2021, 14, 4768. [Google Scholar] [CrossRef]
  66. Abghoui, Y.; Garden, A.L.; Hlynsson, V.F.; Björgvinsdóttir, S.; Ólafsdóttir, H.; Skúlason, E. Enabling electrochemical reduction of nitrogen to ammonia at ambient conditions through rational catalyst design. Phys. Chem. Chem. Phys. 2015, 17, 4909–4918. [Google Scholar] [CrossRef] [PubMed]
  67. 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]
  68. Abghoui, Y.; Skúlasson, E. Transition Metal Nitride Catalysts for Electrochemical Reduction of Nitrogen to Ammonia at Ambient Conditions. Procedia Comput. Sci. 2015, 51, 1897–1906. [Google Scholar] [CrossRef] [Green Version]
  69. Abghoui, Y.; Skúlason, E. Electrochemical synthesis of ammonia via Mars-van Krevelen mechanism on the (111) facets of group III–VII transition metal mononitrides. Catal. Today 2017, 286, 78–84. [Google Scholar] [CrossRef]
  70. Abghoui, Y.; Garden, A.L.; Howalt, J.G.; Vegge, T.; Skúlason, E. Electroreduction of N2 to Ammonia at Ambient Conditions on Mononitrides of Zr, Nb, Cr, and V: A DFT Guide for Experiments. ACS Catal. 2016, 6, 635–646. [Google Scholar] [CrossRef] [Green Version]
  71. Gambarotta, S.; Scott, J. Multimetallic Cooperative Activation of N2. Angew. Chem. Int. Ed. 2004, 43, 5298–5308. [Google Scholar] [CrossRef]
  72. Kitano, M.; Inoue, Y.; Yamazaki, Y.; Hayashi, F.; Kanbara, S.; Matsuishi, S.; Yokoyama, T.; Kim, S.-W.; Hara, M.; Hosono, H. Ammonia synthesis using a stable electride as an electron donor and reversible hydrogen store. Nat. Chem. 2012, 4, 934–940. [Google Scholar] [CrossRef]
  73. Zhang, X.; Kong, R.-M.; Du, H.; Xia, L.; Qu, F. Highly efficient electrochemical ammonia synthesis via nitrogen reduction reactions on a VN nanowire array under ambient conditions. ChemComm 2018, 54, 5323–5325. [Google Scholar] [CrossRef]
  74. Zhang, R.; Zhang, Y.; Ren, X.; Cui, G.; Asiri, A.M.; Zheng, B.; Sun, X. High-Efficiency Electrosynthesis of Ammonia with High Selectivity under Ambient Conditions Enabled by VN Nanosheet Array. ACS Sustain. Chem. Eng. 2018, 6, 9545–9549. [Google Scholar] [CrossRef]
  75. Nash, J.; Yang, X.; Anibal, J.; Dunwell, M.; Yao, S.; Attenkofer, K.; Chen, J.G.; Yan, Y.; Xu, B. Elucidation of the Active Phase and Deactivation Mechanisms of Chromium Nitride in the Electrochemical Nitrogen Reduction Reaction. J. Phys. Chem. C 2019, 123, 23967–23975. [Google Scholar] [CrossRef]
  76. Yao, Y.; Feng, Q.; Zhu, S.; Li, J.; Yao, Y.; Wang, Y.; Wang, Q.; Gu, M.; Wang, H.; Li, H.; et al. Chromium Oxynitride Electrocatalysts for Electrochemical Synthesis of Ammonia Under Ambient Conditions. Small Methods 2019, 3, 1800324. [Google Scholar] [CrossRef]
  77. Ren, X.; Cui, G.; Chen, L.; Xie, F.; Wei, Q.; Tian, Z.; Sun, X. Electrochemical N2 fixation to NH3 under ambient conditions: Mo2N nanorod as a highly efficient and selective catalyst. ChemComm 2018, 54, 8474–8477. [Google Scholar]
  78. Yang, X.; Ling, F.; Su, J.; Zi, X.; Zhang, H.; Zhang, H.; Li, J.; Zhou, M.; Wang, Y. Insights into the role of cation vacancy for significantly enhanced electrochemical nitrogen reduction. Appl. Catal. B 2020, 264, 118477. [Google Scholar] [CrossRef]
  79. Kang, S.; Wang, J.; Zhang, S.; Zhao, C.; Wang, G.; Cai, W.; Zhang, H. Plasma-etching enhanced titanium oxynitride active phase with high oxygen content for ambient electrosynthesis of ammonia. Electrochem. Commun. 2019, 100, 90–95. [Google Scholar] [CrossRef]
  80. Johnson, D.; Hunter, B.; Christie, J.; King, C.; Kelley, E.; Djire, A. Ti2N nitride MXene evokes the Mars-van Krevelen mechanism to achieve high selectivity for nitrogen reduction reaction. Sci. Rep. 2022, 12, 657. [Google Scholar] [CrossRef] [PubMed]
  81. Jin, H.; Li, L.; Liu, X.; Tang, C.; Xu, W.; Chen, S.; Song, L.; Zheng, Y.; Qiao, S.-Z. Nitrogen Vacancies on 2D Layered W2N3: A Stable and Efficient Active Site for Nitrogen Reduction Reaction. Adv. Mater. 2019, 31, 1902709. [Google Scholar] [CrossRef]
  82. He, H.-y.; Wang, S.; Ji, L.-L. Fabrication of self-supported Cu3N electrode for electrocatalytic nitrogen reduction reaction. J. Fuel Chem. Technol. 2022, 50, 484–493. [Google Scholar] [CrossRef]
  83. Yesudoss, D.K.; Lee, G.; Shanmugam, S. Strong catalyst support interactions in defect-rich γ-Mo2N nanoparticles loaded 2D-h-BN hybrid for highly selective nitrogen reduction reaction. Appl. Catal. B Environ. 2021, 287, 119952. [Google Scholar] [CrossRef]
  84. Xi, Z.; Shi, K.; Xu, X.; Jing, P.; Liu, B.; Gao, R.; Zhang, J. Boosting Nitrogen Reduction Reaction via Electronic Coupling of Atomically Dispersed Bismuth with Titanium Nitride Nanorods. Adv. Sci. 2022, 9, 2104245. [Google Scholar] [CrossRef]
  85. Manjunatha, R.; Karajić, A.; Teller, H.; Nicoara, K.; Schechter, A. Electrochemical and Chemical Instability of Vanadium Nitride in the Synthesis of Ammonia Directly from Nitrogen. ChemCatChem 2020, 12, 438–443. [Google Scholar] [CrossRef]
  86. Du, H.-L.; Gengenbach, T.R.; Hodgetts, R.; MacFarlane, D.R.; Simonov, A.N. Critical Assessment of the Electrocatalytic Activity of Vanadium and Niobium Nitrides toward Dinitrogen Reduction to Ammonia. ACS Sustain. Chem. Eng. 2019, 7, 6839–6850. [Google Scholar] [CrossRef]
  87. Guo, W.; Liang, Z.; Tang, Y.; Cai, K.; Qiu, T.; Wu, Y.; Zhang, K.; Gao, S.; Zou, R. Understanding the lattice nitrogen stability and deactivation pathways of cubic CrN nanoparticles in the electrochemical nitrogen reduction reaction. J. Mater. Chem. A 2021, 9, 8568–8575. [Google Scholar] [CrossRef]
  88. Hu, B.; Hu, M.; Seefeldt, L.; Liu, T.L. Electrochemical Dinitrogen Reduction to Ammonia by Mo2N: Catalysis or Decomposition? ACS Energy Lett. 2019, 4, 1053–1054. [Google Scholar] [CrossRef]
  89. Andersen, S.Z.; Čolić, V.; Yang, S.; Schwalbe, J.A.; Nielander, A.C.; McEnaney, J.M.; Enemark-Rasmussen, K.; Baker, J.G.; Singh, A.R.; Rohr, B.A.; et al. A rigorous electrochemical ammonia synthesis protocol with quantitative isotope measurements. Nature 2019, 570, 504–508. [Google Scholar] [CrossRef] [Green Version]
  90. Choi, J.; Suryanto, B.H.R.; Wang, D.; Du, H.-L.; Hodgetts, R.Y.; Ferrero Vallana, F.M.; MacFarlane, D.R.; Simonov, A.N. Identification and elimination of false positives in electrochemical nitrogen reduction studies. Nat. Commun. 2020, 11, 5546. [Google Scholar] [CrossRef]
  91. Wei, J.; Jing, Y.; Zhao, Z.; Fan, Z.; Liang, Z.; Huang, J.; Wu, H.; Xie, Z.; Liu, D.; Qu, D.; et al. Catalyst-Support interactions enhanced electrochemical nitrogen reduction on Au/ZrO2. Electrochim. Acta 2021, 381, 138222. [Google Scholar] [CrossRef]
  92. Wang, J.; Wei, J.; An, C.; Tang, H.; Deng, Q.; Li, J. Electrocatalyst design for the conversion of energy molecules: Electronic state modulation and mass transport regulation. Chem. Commun. 2022, 58, 10907–10924. [Google Scholar] [CrossRef]
  93. Ologunagba, D.; Kattel, S. A Density Functional Theory Study of Electrochemical Nitrogen Reduction to Ammonia on the (100) Surface of Transition-Metal Oxynitrides. J. Phys. Chem. C 2022, 126, 17045–17055. [Google Scholar] [CrossRef]
  94. Wei, X.; Vogel, D.; Keller, L.; Kriescher, S.; Wessling, M. Microtubular Gas Diffusion Electrode Based on Ruthenium-Carbon Nanotubes for Ambient Electrochemical Nitrogen Reduction to Ammonia. ChemElectroChem 2020, 7, 4679–4684. [Google Scholar] [CrossRef]
  95. Kolen, M.; Antoniadis, G.; Schreuders, H.; Boshuizen, B.; van Noordenne, D.D.; Ripepi, D.; Smith, W.A.; Mulder, F.M. Combinatorial Screening of Bimetallic Electrocatalysts for Nitrogen Reduction to Ammonia Using a High-Throughput Gas Diffusion Electrode Cell Design. J. Electrochem. Soc. 2022, 169, 124506. [Google Scholar] [CrossRef]
  96. Kolen, M.; Ripepi, D.; Smith, W.A.; Burdyny, T.; Mulder, F.M. Overcoming Nitrogen Reduction to Ammonia Detection Challenges: The Case for Leapfrogging to Gas Diffusion Electrode Platforms. ACS Catal. 2022, 12, 5726–5735. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Schematic illustration for comparing the HB and NRR processes.
Figure 1. Schematic illustration for comparing the HB and NRR processes.
Catalysts 13 00639 g001
Figure 2. The mechanisms of NRR. (a) Dissociative pathway; (b) associative distal pathway; (c) associative alternative pathway; (d) associative enzymatic pathway; and (e) Mars–Van Krevelen (MvK) pathway.
Figure 2. The mechanisms of NRR. (a) Dissociative pathway; (b) associative distal pathway; (c) associative alternative pathway; (d) associative enzymatic pathway; and (e) Mars–Van Krevelen (MvK) pathway.
Catalysts 13 00639 g002
Figure 3. (a) SEM image of VN/CC. (b) NH3 yield and FE at different potentials on VN/TM. (c) Schematic illustration of MEA. (d) VNxOy/VN ratio after 1 h of reaction at various potentials. (e) Operando XAS results of K-edge XANES of VN at −0.2 V. (f) Proposed MvK mechanism of NRR and deactivation pathway on VN0.7O0.45. Reproduced with permission from the Royal Society of Chemistry [73] and the American Chemical Society [37,74,75].
Figure 3. (a) SEM image of VN/CC. (b) NH3 yield and FE at different potentials on VN/TM. (c) Schematic illustration of MEA. (d) VNxOy/VN ratio after 1 h of reaction at various potentials. (e) Operando XAS results of K-edge XANES of VN at −0.2 V. (f) Proposed MvK mechanism of NRR and deactivation pathway on VN0.7O0.45. Reproduced with permission from the Royal Society of Chemistry [73] and the American Chemical Society [37,74,75].
Catalysts 13 00639 g003
Figure 5. (a) NH3 yield and FEs on TiN-PE. (b) 1H NMR spectra for 14NH4+, 15NH4+ from standard (14NH4)2SO4, (15NH4)2SO4, and 14N2, 15N2 bubbled in electrolyte using TiN-PE. (c) A schematic illustration of the formation of Ti2N MXene from the Ti2AlN MAX phase. (d) FT-EXFAS plot of NV-W2N3 and W2N3. (e) The difference between NV-W2N3 and W2N3 on NH3 yield rate and FEs. (f) EDX analysis after electrolysis of Cu3N/CF in Ar-saturated electrolyte. Reproduced with permission from Springer Nature [80], Wiley [81], and Elsevier [79,82].
Figure 5. (a) NH3 yield and FEs on TiN-PE. (b) 1H NMR spectra for 14NH4+, 15NH4+ from standard (14NH4)2SO4, (15NH4)2SO4, and 14N2, 15N2 bubbled in electrolyte using TiN-PE. (c) A schematic illustration of the formation of Ti2N MXene from the Ti2AlN MAX phase. (d) FT-EXFAS plot of NV-W2N3 and W2N3. (e) The difference between NV-W2N3 and W2N3 on NH3 yield rate and FEs. (f) EDX analysis after electrolysis of Cu3N/CF in Ar-saturated electrolyte. Reproduced with permission from Springer Nature [80], Wiley [81], and Elsevier [79,82].
Catalysts 13 00639 g005
Figure 7. (a) 1H NMR spectra of 14N2 and 15N2 on Mo2N. (b) Quantification of ammonia after incubating Mo2N in acidic, neutral, and alkaline solutions without and with a bias for 1 h. (c) J-V curves on Nb4N5/CC in consecutive tests (1st—N2, 2nd—N2, 3rd—Ar). (d) Changes of produced NH3 for 3 times of consecutive tests. (e) XPS spectra in the Nb 3d region. Reproduced with permission from the American Chemistry Society [86,88].
Figure 7. (a) 1H NMR spectra of 14N2 and 15N2 on Mo2N. (b) Quantification of ammonia after incubating Mo2N in acidic, neutral, and alkaline solutions without and with a bias for 1 h. (c) J-V curves on Nb4N5/CC in consecutive tests (1st—N2, 2nd—N2, 3rd—Ar). (d) Changes of produced NH3 for 3 times of consecutive tests. (e) XPS spectra in the Nb 3d region. Reproduced with permission from the American Chemistry Society [86,88].
Catalysts 13 00639 g007
Table 1. The NRR performances of TMN-based catalysts.
Table 1. The NRR performances of TMN-based catalysts.
TMNs CatalystElectrolytePotential
/V vs. RHE
Production Rate
/mol s−1 cm−2
FE
/%
IsotopeTest Cell
Condition
Ref
VN/CC0.1 M HCl−0.32.48 × 10−103.58XH-cell[73]
VN/TM0.1 M HCl−0.58.40 × 10−112.25OH-cell[74]
VN nanoparticles-−0.13.3 × 10−106.0OMEA[37]
Cr2N-−0.21.4 × 10−110.58OMEA[75]
CrN0.1 M HCl−0.56.1 × 10−1116.6XH-cell[38]
CrO0.66N0.56-2.08.94 × 10−116.7 (at 1.8 V)XPEMEL[76]
MoN NA/CC0.1 M HCl−0.33.01 × 10−101.15OH-cell[36]
Mo2N/GCE0.1 M HCl−0.378.4 μg mgcat−1 h−14.5XH-cell[77]
MV-MoN@NC0.1 M HCl−0.25.02 × 10−106.9OH-cell[78]
TiN-PE0.1 M Na2SO4−0.63.32 × 10−109.1OTwo-compartment[79]
Ti2N MXene0.1 M HCl−0.251.85 × 10−1019.85XH-cell[80]
2D layered W2N30.1 M KOH−0.23.8 ± 0.32 × 10−1111.67 ± 0.93OH-cell[81]
Cu3N/CF0.1 M Na2SO4−0.21.12 × 10−101.5XH-cell[82]
γ-Mo2N on 2D-h-BN0.1 M Na2SO4−0.335.9 μg mg−1 h−161.5XH-cell[83]
NC/Bi SAs/TiN/CC0.1 M Na2SO4−0.875.15 μg mgcat−1 h−124.6OH-cell[84]
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

Park, S.Y.; Jang, Y.J.; Youn, D.H. A Review of Transition Metal Nitride-Based Catalysts for Electrochemical Nitrogen Reduction to Ammonia. Catalysts 2023, 13, 639. https://doi.org/10.3390/catal13030639

AMA Style

Park SY, Jang YJ, Youn DH. A Review of Transition Metal Nitride-Based Catalysts for Electrochemical Nitrogen Reduction to Ammonia. Catalysts. 2023; 13(3):639. https://doi.org/10.3390/catal13030639

Chicago/Turabian Style

Park, So Young, Youn Jeong Jang, and Duck Hyun Youn. 2023. "A Review of Transition Metal Nitride-Based Catalysts for Electrochemical Nitrogen Reduction to Ammonia" Catalysts 13, no. 3: 639. https://doi.org/10.3390/catal13030639

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