Open Access Article
This Open Access Article is licensed under a Creative Commons Attribution-Non Commercial 3.0 Unported Licence

Accelerating N2H4(ads) formation by frustrated Lewis pairs in an oxyhydroxide for electrocatalytic ammonia oxidation into N2

Meng-Ying Yinab, Xing-Yuan Xiaab, Ting Daiab, Xia Chenab, Qiu-Ju Xingab, Lei Tian*abc and Jian-Ping Zou*ab
aKey Laboratory of Jiangxi Province for Persistent Pollutants Prevention Control and Resource Reuse, Nanchang Hangkong University, Nanchang 330063, P. R. China
bNational-Local Joint Engineering Research Center of Heavy Metals Pollutants Control and Resource Utilization, Nanchang Hangkong University, Nanchang 330063, P. R. China. E-mail: zjp_112@126.com
cSchool of Chemistry and Chemical Engineering, Institute of Clean Energy and Materials, Guangzhou Key Laboratory for Clean Energy and Materials, Key Laboratory for Water Quality and Conservation of the Pearl River Delta, Ministry of Education, Guangzhou University, Guangzhou 510006, P. R. China

Received 26th August 2025 , Accepted 4th October 2025

First published on 6th October 2025


Abstract

Conventional catalysts based on the individual Oswin and Salomon (O–S) or Gerischer and Mauerer (G–M) mechanism cannot achieve direct electrocatalytic ammonia (NH3) oxidation into nitrogen (N2) with high activity and selectivity. Herein, a bimetallic nickel–cobalt oxyhydroxide (Ni0.5–Co0.5-OOH) with frustrated Lewis pairs was developed through an elaborate analysis of the binding types of NH3 with the metal-oxide anode, efficiently integrating O–S and G–M mechanisms for converting NH3 into N2 with high activity (94%) and selectivity (63%), which is much superior to the anodes in the previous reports. The evidence of batch experiments, in situ characterization, and theoretical calculations confirms that two NH3 molecules bind to Co3+ sites (Lewis acid) in CoOOH and hydroxy sites (Lewis base) in NiOOH, respectively. Then, the NH2(ads) generated on the Lewis acid sites can quickly recombine with the NH2(ads) desorbed from the Lewis base sites, accelerating the formation of N2H4(ads) and preventing the peroxidation of NH3. The electrocatalytic system assembled with the Ni0.5–Co0.5-OOH anode shows excellent performance for NH3 elimination in the secondary aerobic process effluent. Our work provides precious guidance for the design of novel anodes and sheds light on further promoting the performance of ammonia conversion.


Introduction

The electrocatalytic ammonia oxidation reaction (EAOR) is viewed as a promising conversion measure due to its ability to address the growing problem of ammonia-containing wastewater pollution.1–4 Nitrate and nitrite are the common products during the EAOR process, which could exacerbate secondary pollution.5–7 As a comparison, nitrogen (N2) is the most desirable product of the EAOR.8–11 Therefore, it is urgent to exploit effective strategies to efficiently convert ammonia into N2.

Direct electrocatalytic oxidation, based on anodic electron transfer, has been widely applied to ammonia pollution purification.12–14 It is more environmentally friendly than indirect oxidation, which is related to the electrochlorination process because there are no toxic secondary products (chlorine gas and chlorine-containing intermediates) produced.15–17 Until now, two possible conversion paths of the EAOR by direct electrocatalytic oxidation have been proposed. One of them is the Oswin and Salomon (O–S) mechanism,18 in which the adsorbed NH3 dehydrogenates to N(ads) step by step and then converts into N2 after the recombination of N(ads) (eqn (1)–(5)). During this process, the active sites in the anode have high Lewis acidity and mainly bind to the N atom in NH3, ensuring strong binding energy for subsequent dehydrogenation. Unfortunately, excessive binding prevents the generated N(ads) from recombining, making it highly susceptible to peroxidation to generate NO2 and NO3.19–21 The O–S mechanism demonstrates strong activity for the EAOR but extremely low selectivity for N2 generation. The other reaction path for the EAOR is the Gerischer and Mauerer (G–M) mechanism,22,23 in which the formation of N2 is derived from the recombination of NxHy(ads) intermediates (eqn (1)–(3) and (6)). The G–M mechanism shows a higher selectivity for N2 formation than the O–S mechanism because the anodic active sites have high Lewis basicity and mainly bind to the H atom in NH3, which is easily desorbed from the anodic surface after dehydrogenation, further efficiently converting into N2. However, the binding energy is relatively weak when the active site binds to the H atom in NH3, resulting in low efficiency of NH3 dehydrogenation to NH2(ads) and sluggish EAOR kinetics.24,25 The anode materials reported so far do not exhibit simultaneously high selectivity and high activity of the EAOR. Therefore, it is necessary to develop novel anode materials to realize the highly efficient and selective conversion of NH3 into N2 through direct oxidation.

 
NH3(aq) → NH3(ads) (1)
 
NH3(ads) → NH2(ads) + H+ + e (2)
 
NH2(ads) → NH(ads) + H+ + e (3)
 
NH(ads) → N(ads) + H+ + e (4)
 
N(ads) + N(ads) → N2 (5)
 
NHx(ads) + NHy(ads) → N2 + (x + y)H+ + (x + y)e, x, y = 1 or 2 (6)

Inspired by the above mechanisms, it is found that the dehydrogenation of NH3 into NH2(ads) is the rate-limiting step of the EAOR, while the recombination of NHx(ads) and NHy(ads) is the determining step for the selectivity of N2 generation.26 Theoretically, if the dehydrogenation in the O–S mechanism and the recombination of NHx(ads) and NHy(ads) in the G–M mechanism can be integrated together, the oxidation of NH3 into N2 can be effectively achieved with high activity and selectivity. Among the numerous anode materials,27–30 metal oxyhydroxides (MOOH) are the desired anode materials for the efficient and selective conversion of NH3 into N2 due to the presence of frustrated Lewis pairs.31,32 In detail, the metal atoms as Lewis acid sites can preferentially bind to the N atom due to the strong interaction between empty d-orbitals and lone pair electrons in the N atom, while the oxygen-containing functional groups as Lewis base sites can bind to the H atom (O–H–N) through hydrogen bonding. NHx(ads) generated on the Lewis acid sites will quickly recombine with NHy(ads) desorbed from the Lewis base sites, which can prevent the peroxidation of NH3, thereby promoting EAOR efficiency and N2 generation selectivity. More importantly, the presence of surface hydroxy groups enables oxyhydroxides to effectively resist the erosion of Cl, thereby demonstrating high reliability in actual wastewater treatment. Although a small number of hydroxyl oxide electrodes (NiOOH or FeOOH) have been reported for use in the EAOR process, the activity of the EAOR and the selectivity for N2 generation are poor (below 30%).33,34 Therefore, it is urgent to clarify the conversion law of NH3 on different oxyhydroxides and develop highly active and selective oxyhydroxide anodes to achieve the EAOR.

Herein, a series of mono-metallic oxyhydroxide electrodes (MOOH, M: Mn, Co, Fe, Cu, Ni, and Zn) were synthesized, and their activity and selectivity of N2 formation in the EAOR were investigated. Combined with the electronic structure of different transition metal ions, the transformation law between NH3 and the active sites with different Lewis acidity/basicity was summarized. Based on this, a bimetallic oxyhydroxide was designed and synthesized, which can bind to the N and H atoms in NH3, respectively, integrating the O–S and the G–M mechanisms together, further enhancing the EAOR efficiency and the selectivity of N2 formation. The synergistic effect of the bimetallic oxyhydroxide was explored by batch performance evaluation and structural characterization. Moreover, through XPS, in situ diffuse reflectance infrared Fourier transform spectroscopy (DRIFTS), and theoretical calculations, the conversion mechanism of NH3 mediated by the bimetallic oxyhydroxide was further demonstrated. This work provides a new avenue for the design of anodic materials for the EAOR, which is in favor of the technological development of ammonia pollution purification.

Materials and methods

Materials and characterization

All chemicals were commercially obtained and used without further purification and the detailed information of materials is shown in Text S1 in the SI. The results of electrochemical analysis and characterization for catalyst structure and morphology are displayed in Text S2 in the SI. The evolution of conversion performance is displayed in Text S3 in the SI. The analytical methods for electrochemical characterization are displayed in Text S4 in the SI.

Preparation of the anodes

Nickel foam (NF) was first cleaned through sonicating consecutively in acetone, 3 M HCl, ethanol, and deionized water for 15 min to remove the oxide layer and degrease the surface of the NF. Then, the treated NF was directly used as the substrate to deposit Ni–Co precursors. In a typical procedure of Ni–Co precursor fabrication, 2.0 mmol (582.06 mg) of Co(NO3)2·6H2O, 2.0 mmol (497.68 mg) of Ni(CH3COO)2·4H2O, 3.75 mmol (138.8 mg) of NH4F and 15 mmol (900.7 mg) of CO(NH2)2 were well dissolved in 50 mL of deionized water to form a mixture solution. The obtained solution was transferred into a 100 mL Teflon-lined stainless-steel autoclave. After immersing the washed NF (1.5 × 4 cm2) substrate in the homogeneous solution, the autoclave was sealed and heated at 90 °C for 7 h. After the autoclave cooled down to room temperature, the Ni–Co precursor loaded NF was taken out from the autoclave and rinsed with ultrapure water several times. The black electrode of Ni–Co oxyhydroxide was obtained by immersing the precursors in 5% NaClO solution as an oxidant for 30 min. The chemical oxidation of the precursors was carried out at pH 3–4 adjusted with H2SO4 or NaOH.

Catalytic experiments

The initial concentration of ammonia was 20 mg N L−1 and 50 mM Na2SO4 electrolyte was added into a 100 mL undivided cylindrical glass electrolytic cell. The different anodes and the Pt cathode were fixed at a gap of 1 cm in the reactor and connected to a potentiostat (CHI 660E electrochemical workstation (Shanghai, Chenhua, China)). Ag/AgCl was used as the reference electrode. The immersed area of the electrode was 4.5 cm2. During the reaction process, the samples were taken with syringes at designated time intervals and immediately filtered through 0.45 μm polytetrafluoroethylene (PTFE) membranes. All experiments were conducted three times.

Analysis methods

The concentration of ammonia was measured by Nessler's reagent colorimetric method. The samples were distilled first and the absorbance was obtained from a UV-vis spectrophotometer (UV-9000S, Metash Instruments Ltd, China) at the wavelength of 420 nm.35 The concentrations of nitrate (NO3) and nitrite (NO2) were determined by the ultraviolet spectrophotometry and N-(1-naphthyl)-ethylenediamine dihydrochloride spectrophotometric method, respectively.36 The concentration of total nitrogen (TN) was measured by the traditional alkaline persulfate oxidation method.37 The pH value was measured with a pH meter (Mettler Toledo-FE28, Swit). The ammonia removal (R) and selectivity of nitrogen conversion (SN) were calculated using equations below:
 
image file: d5sc06524k-t1.tif(7)
 
image file: d5sc06524k-t2.tif(8)
where [N] denotes the concentration of nitrogen products, including [NO2–N], [NO3–N] and [N2–N].

Current efficiency is expressed as eqn (9):

 
image file: d5sc06524k-t3.tif(9)
Here, C0 is the initial concentration of NH3–N, Ct is the concentration of NH3–N at degradation time t (s), V is the volume of the electrolyte (0.1 L), M is the molar mass of NH3–N (14 g mol−1), I is the current density (A m−2), A is the effective area of the electrodes (4.5 m−2), F is the Faraday constant (96[thin space (1/6-em)]485.3 C mol−1) and n is the number of electrons needed for the oxidation of one mole of ammonia (n = 3).

The energy consumption (EC, kWh m3) was calculated through eqn (10).

 
image file: d5sc06524k-t4.tif(10)
where U is the applied voltage (V), I is the average current (A), t is the time (h), V is the solution volume (L), and log(C0/Ct) is the logarithm of the initial and instant concentrations of the ammonia.

Theoretical calculation

The present first principle DFT calculations are performed with the Vienna Ab initio Simulation Package (VASP) with the projector augmented wave (PAW) method. Detailed information about the calculation procedure is provided in Text S5 in the SI.

Results and discussion

Lewis acidity/basicity on NH3 conversion

The structure characterization studies of different MOOH are shown in Fig. S1 and S2. The prototypical MOOH oxyhydroxide has a typical cubic crystal structure. Metal atoms are located at the centre of MO6 octahedra and the layers are linked by hydrogen bonds (Fig. 1a). When there are many unoccupied d orbitals, metal atoms in MOOH can bind with N atoms in NH3 during the adsorption process.38 In contrast, the O atoms in MOOH may bind with H in NH3 through a hydrogen bond (Fig. 1b).39 With the outer 3d orbitals being gradually filled by electrons, the Lewis acidity of metal sites decreases and the Lewis basicity of hydroxy sites increases (Fig. 1c, e and S3). The EAOR performance of MOOH (from Mn3+ to Zn3+) increases first and then decreases (CoOOH > MnOOH > NiOOH > FeOOH > ZnOOH > CuOOH). When the d orbitals of the metal atom in MOOH are before half-filled, the adsorption energies of NH3 at the metal sites are higher than those at the hydroxy sites (Fig. 1d), which can strongly bind with the N atom in NH3, thus giving it strong EAOR performance. However, the overly strong force leads to the peroxidation of NHx(ads) intermediates into NO2/NO3, which follows the typical O–S mechanism and exhibits low N2 formation selectivity. As a comparison, after the d orbital is half-filled, NH3 adsorbs more strongly at the hydroxy sites than at the metal sites. Though the EAOR performance of MOOH is relatively sluggish, the N2 formation selectivity is dramatically increased, which can be attributed to the change of the catalytic mechanism from O–S to G–M. In the two stages before and after a half-filled electronic orbital, CoOOH and NiOOH exhibited optimal Lewis acidity/basicity and superior activity and selectivity for the EAOR compared to other MOOH or bimetallic systems (Fig. S4), respectively. Hence, Co and Ni elements were used to assemble the subsequent bimetallic oxyhydroxide catalysts to integrate the O–S and G–M mechanisms on a catalyst.
image file: d5sc06524k-f1.tif
Fig. 1 (a) Crystal structure of MOOH; (b) different adsorption types with NH3; (c) EAOR activity and selectivity of MOOH with different transition metal ions (E = 1.8 V, pH = 11, 0.05 M Na2SO4); (d) calculated adsorption energies of ammonia molecules at metal and hydroxyl sites in mono-metallic oxyhydroxides; (e) diagram of d orbital electron configuration of different transition metal ions.

Synthesis and characterization of Ni–Co-OOH

The Ni–Co oxyhydroxide (Ni–Co-OOH) was synthesized following the schematic diagram in Fig. 2a. The XRD pattern of Ni–Co-OOH shows the presence of characteristic peaks attributed to NiOOH and CoOOH, respectively (Fig. S5). As shown in Fig. 2b, Raman spectra further demonstrate that the peaks located at 474 and 557 cm−1 are attributed to the Eg bending and A1g stretching vibrations of Ni–O of NiOOH, respectively, whereas the peaks observed at 506 and 680 cm−1 are assigned to the CoOOH in Ni–Co-OOH.40,41 The Raman spectra show a slight shift in the vibrational frequencies of Ni–O and Co–O bonds compared to NiOOH and CoOOH, indicating there is electronic interaction between NiOOH and CoOOH. Besides, X-ray photoelectron spectroscopy (XPS) shows that two peaks in Co 2p3/2 at 780.8 and 781.9 eV are assigned to Co3+ and Co2+,42–44 respectively. For Ni 2p3/2 spectra, the peaks at 855.36 and 856.66 eV are attributable to the Ni2+ and Ni3+,45–47 respectively (Fig. 2c and d). Ni/Co–O bonds are also identified by the characteristic peak at 529.7 eV from O 1s spectra (Fig. S6). The peaks of Ni3+ and Co3+ are observed to shift by 0.3 eV and 0.5 eV, respectively, towards a lower binding energy compared to their mono-metallic oxyhydroxide, proving the electron transfer from Co3+ to Ni3+. The coupling of Ni3+ and Co3+ via the bridging O2− facilitates π-donation from O2− to Ni3+, resulting in the overall electron transfer from Co3+ to Ni3+, which is in agreement with the Raman spectra results. In addition, scanning electron microscopy (SEM) images show that the morphology of Ni–Co-OOH is a nanoplate structure covered with nano-flower spheres, where NiOOH and CoOOH are nanoplates and nano-flower spheres, respectively, confirming that Ni–Co-OOH is a composite and CoOOH is loaded on the surface of NiOOH (Fig. 2e, S7 and S8). A further inspection of these nanosheet structures with high-resolution TEM and spherical aberration-corrected high angle annular dark field scanning transmission electron microscopy (HAADF-STEM) was conducted. As shown in Fig. 2f and g, the images demonstrate the heterojunction region of Ni–Co-OOH, indicating that there are two different phases on both sides of the interface. The lattice distances are 1.6, 2.1 and 2.3 Å, which are consistent with the (2 2 0) facet of NiOOH, (0 0 6) facet of CoOOH and (1 0 2) facet of NiOOH, respectively. Therefore, the aforementioned characterization evidently confirms the successful preparation of the Ni–Co bimetallic oxyhydroxide catalyst by a two-stage method of hydrothermal and chemical oxidation and it is a biphasic composite with a heterojunction structure.
image file: d5sc06524k-f2.tif
Fig. 2 (a) Synthesis diagram of Ni–Co-OOH; (b) Raman spectra of Ni–Co-OOH, CoOOH and NiOOH and nickel foam (NF); (c) Ni 2p XPS of Ni–Co-OOH and NiOOH; (d) Co 2p XPS of Ni–Co-OOH and CoOOH; (e) SEM images of Ni–Co-OOH; (f) TEM images of Ni–CoOOH; (g) HAADF-STEM image of Ni–Co-OOH.

EAOR activity, N2 selectivity and practical application

The EAOR activity of CoOOH and NiOOH is 93.8% and 24.1%, whereas the selectivity of N2 is 1.3% and 35.7% and the selectivity of NO3 is 74% and 24%, respectively, which typically follows the O–S and G–M mechanisms, respectively (Fig. 3a). Nix–Co1−x-OOH catalysts with different Ni/Co ratios were obtained by varying the precursor dosages based on the similar solubility products of Ni(OH)2 and Co(OH)2 (ksp(Ni(OH)2) = 2 × 10−15, ksp(Co(OH)2) = 1.6 × 10−15) (Table S1). Ni0.5–Co0.5-OOH (Ni[thin space (1/6-em)]:[thin space (1/6-em)]Co = 0.5[thin space (1/6-em)]:[thin space (1/6-em)]0.5) exhibits higher EAOR activity (94.0%) and N2 formation selectivity (63.0%) than Ni0.75–Co0.25-OOH, Ni0.25–Co0.75-OOH and the physically mixed CoOOH/NiOOH, which is superior to the electrocatalysts reported in the previous literature (Fig. S9–S11 and Table S2).48–54 Since the strong interaction between the Co-based Lewis acid sites and the N atom can cause difficulty in the desorption of N-containing intermediates from the anodic surface, the introduction of excessive CoOOH into NiOOH inhibits the formation of N2, while generating more NO3 during the EAOR process. In addition, as shown in Fig. 3b, the NiOOH anode shows low faradaic efficiency (7.5%) and high energy consumption (150 kWh kg N−1) for NH3 conversion. Conversely, the current efficiency of Co-based oxyhydroxides significantly improves to 20–40%, while the low N2 selectivity makes the NO2 and NO3 predominate the energy consumption. The Ni0.5–Co0.5-OOH can reach the highest faradaic efficiency (42%) and the energy consumption reaches a minimum value of 26.3 kWh kg N−1, which further verifies the superior activity of Ni0.5–Co0.5-OOH for the EAOR.
image file: d5sc06524k-f3.tif
Fig. 3 (a) EAOR activity and product selectivity with different anodes (E = 1.8 V, pH = 12, 0.05 M Na2SO4); (b) Faradaic efficiency and energy consumption of the EAOR with different anodes (E = 1.8 V, pH = 12, 0.05 M Na2SO4); (c) EAOR activity and product selectivity of Ni0.5–Co0.5-OOH under different conditions; (d) ten cycles of Ni0.5–Co0.5-OOH (E = 1.8 V, pH = 12, 0.05 M Na2SO4); (e) EAOR activity of Ni0.5–Co0.5-OOH for practical wastewater (E = 2.0 V, 0.05 M Na2SO4).

With the increase of applied voltage from 1.6 V to 2.1 V vs. RHE, the EAOR activity of Ni0.5–Co0.5-OOH increases, while the N2 formation selectivity is optimal (63.0%) at 1.8 V (Fig. 3c and S12). When excessive voltage is applied, the anodic oxygen evolution reaction (OER) may be severe, and NH3 can be easily peroxided into NO3 and NO2, resulting in a low N2 formation selectivity. Besides, the removal of NH3 and the formation selectivity of N2 is significantly inhibited with increasing NH3 concentrations (Fig. 3c and S13), which may be due to the saturation of the active sites on the electrode surface. In addition, the oxidation efficiency of NH3 increases from 3.9% to 99.3%, whereas the N2 formation selectivity declines from 100% to 38.4% with the increase in pH from 10 to 13 (Fig. 3c and S14). At a relatively low pH, NH3 can be protonated into NH4+ (pKa(NH4+/NH3) = 9.25), leading to the sluggish anodic oxidation of NH4+ caused by electrostatic repulsion. In contrast, at a high pH, N2 production selectivity is significantly inhibited due to the serious OER and NH3 peroxidation (eqn (S4)–(S10)) at the anode, leading to the generation of NO2 and NO3. These results indicate that the best performance of the Ni0.5–Co0.5-OOH anode can be obtained when the voltage is 1.8 V, the NH3 concentration is 20 mg L−1 and the pH is 12.

To evaluate the practical application of the Ni0.5–Co0.5-OOH anode, ten cycles of experiments were carried out. As shown in Fig. 3d, the NH3 conversion is around 90%, while the N2 formation selectivity is still over 60% within 10 cycles, indicating that Ni0.5–Co0.5-OOH shows remarkable stability. Besides, there are no leaching concentrations of cobalt/nickel ions detected even in chloride-containing solutions and no change is observed in the metal value state and structure during ten cycles (Fig. S15–S18), which further confirms the stability of Ni0.5–Co0.5-OOH. Unlike the lattice oxygen mechanism involved in the OER processes (Fig. S19), surface reconstruction does not occur during the EAOR process. In addition, the electrocatalytic system assembled with the Ni0.5–Co0.5-OOH anode was applied to treat the practical ammonia-containing wastewater, which was the secondary aerobic process effluent. As shown in Fig. 3e and S20, the NH3 conversion is around 99%, while the N2 formation selectivity (55%) is much greater than that of nitrate (31%) and nitrite (11%). In conclusion, the Ni0.5–Co0.5-OOH electrode has excellent stability and shows good application potential in the treatment of practical ammonia-containing wastewater.

Electrochemical characterization

Diverse electrochemical characterization studies were performed to explore the EAOR activity of Ni0.5–Co0.5-OOH. As shown in the linear sweep cyclic voltammetry (LSV) and Tafel curves (Fig. 4a and S21–S23), a lower onset potential (1.28 V) and a lower Tafel slope (128 mV dec−1) are observed in the presence of NH3, demonstrating that Ni0.5–Co0.5-OOH is more susceptible to the EAOR than the OER (1.45 V, 244 mV dec−1). Besides, the polarization current of Ni0.5–Co0.5-OOH in the presence of NH3 is beyond that of CoOOH and the corresponding Tafel slope is the lowest derived from these catalysts, which further indicates exceptional activity of Ni0.5–Co0.5-OOH for the EAOR (Fig. 4b, S24 and S25). Moreover, the electrochemical surface area is calculated from the values of double layer capacitance (Cdl) to be 35.5, 31.75, 50.25, 38 and 69.5 cm2 for NiOOH, CoOOH, Ni0.75–Co0.25-OOH, Ni0.25–Co0.75-OOH and Ni0.5–Co0.5-OOH, respectively (Fig. 4c and S26). This result confirms that there are more catalytically active sites on the Ni0.5–Co0.5-OOH than other oxyhydroxides because of the synergistic effect of CoOOH and NiOOH. Furthermore, Ni0.5–Co0.5-OOH displays the smallest Nyquist semicircle diameter, suggesting the lowest charge transfer resistances (1.2 Ω) during the reaction process (Fig. 4d). Meanwhile, the in situ electrochemical impedance spectroscopy (EIS) tests were conducted to uncover the dynamic interfacial changes of the catalyst (Fig. 4e and f). It is observed that the phase angle of the HF region in the NH3 solution sharply decreased from 1.26 V, indicating the fast charge-transfer kinetics of the EAOR. Thus, the Ni0.5–Co0.5-OOH/NF electrocatalyst exhibits extraordinary activity for the EAOR.
image file: d5sc06524k-f4.tif
Fig. 4 (a) LSV curves of Ni0.5–Co0.5-OOH with or without NH3; (b) LSV curves of different oxyhydroxides in the presence of NH3; (c) electrochemical double-layer capacitance (Cdl) of different oxyhydroxides; (d) electrochemical impedance spectroscopy of different oxyhydroxides, the inset is the equivalent circuit for EIS fitting; (e and f) Bode phase plots of Ni0.5–Co0.5-OOH for the EAOR and OER.

Direct NH3 oxidation mechanism

XPS and in situ diffuse reflectance infrared Fourier transform spectroscopy (DRIFTS) were used to detect the variation of Ni0.5–Co0.5-OOH/NF and NH3 molecules during the EAOR process. As shown in Fig. 5a and b, the proportions of Ni3+/Ni2+ and Co3+/Co2+ first decrease and then increase close to the initial value during the EAOR process (Tables S3 and S4). Regardless of O–S and G–M mechanisms, trivalent metal ions with high oxidative potential in oxyhydroxides act as the electron acceptors to induce the dehydrogenation of NH3(ads) into NH2(ads). As a result, the decrease of Ni3+ and Co3+ in Ni0.5–Co0.5-OOH occurs at the initial stage of the EAOR. Subsequently, the formed Ni2+ and Co2+ in Ni0.5–Co0.5-OOH can be further oxidized into Ni3+ and Co3+ through anodic electron transfer. Besides, it can be seen from the potential-dependent in situ DRIFTS that the peak around 1102 cm−1 is ascribable to the symmetric N–H bending vibration, which stems from the NH3(ads) or NHx(ads) intermediates (Fig. 5c and S27). Notably, there is a new peak at 1244 cm−1 with increasing potential, which is assigned to the stretching of H2N–NH2(ads).55,56 This phenomenon manifests that the formation of N2 during the EAOR process is derived from the further conversion of N2H4(ads). A negligible peak around 1605 cm−1 in DRIFTS, corresponding to the N–O vibration, is also observed, resulting from a few NO2 or NO3 generated from the peroxidation of NH3. In addition, as shown in Fig. 5d, in situ Raman spectra of CoOOH and NiOOH show an increasing intensity of Ni–OH and Co–N peaks, respectively, with the potential increasing from 1.2 V to 1.9 V vs. RHE. It is noted that the stretching vibrations of Co–N and Ni–OH bonds are observed in Ni–Co-OOH throughout the reaction, which serves to further verify the synergistic effect of the frustrated Lewis pairs. The intensity of the Ni–OH peak, which arises from strong hydrogen bonding between NiOOH and NH3,57–59 initially increases and then decreases as the potential is raised from 1.2 V to 1.9 V. This trend suggests that NiOOH undergoes reversible transformation to Ni(OH)2 and eventually returns to its initial state, accompanied by the dehydrogenation of NH3, formation of N2H4(ads), and generation of N2 on frustrated Lewis pairs. A similar reversible behavior is observed for CoOOH in the Ni–Co-OOH system. The change in Co–N strength also confirms that Co3+ is the site for NH3 oxidation. The frustrated Lewis acid–base pair composed of Co3+ and hydroxy facilitates N2H4(ads) formation, thus promoting the selective ammonia oxidation into N2.
image file: d5sc06524k-f5.tif
Fig. 5 (a) Ni 2p XPS of the Ni0.5–Co0.5-OOH system within 4 h; (b) Co 2p XPS of the Ni0.5–Co0.5-OOH system; (c) in situ DRIFTS spectra of the Ni0.5–Co0.5-OOH system with applied potential positive scanning from 0.91 to 2.0 V vs. RHE; (d) in situ Raman test of CoOOH (left), NiOOH (middle) and Ni0.5–Co0.5-OOH (right) systems; (e) the modeled N2H4(ads) formation pathway of the Ni0.5–Co0.5-OOH and NiOOH system, respectively; (f) the pivotal potential gaps of RDS and the formation step of N2H4(ads) in Ni–Co-OOH, CoOOH and NiOOH systems; (g) Gibbs free energy diagrams of the EAOR by G–M and O–S mechanisms of the Ni0.5–Co0.5-OOH and CoOOH system, respectively.

DFT calculations were performed to explore the conversion path of NH3 mediated by Ni0.5–Co0.5-OOH. According to the structural characterization, the Ni0.5–Co0.5-OOH model with the heterojunction is simulated (Fig. S28). It is known that NiOOH following the G–M mechanism exhibits high selectivity for N2 generation, which is primarily derived from the conversion of N2H4(ads) intermediates. The N2H4(ads) generation energy barrier is compared to determine the N2 generation selectivity of Ni–Co-OOH. As shown in Fig. 5e and f, for Ni0.5–Co0.5-OOH, two NH3 molecules bind to Co3+ sites and hydroxy sites (the adsorption model as displayed in Fig. S29), which is the integration of O–S and G–M mechanisms. Then, the NH2(ads) generated on the Co3+ sites can quickly recombine with the NH2(ads) desorbed from the hydroxy sites into N2H4(ads) (eqn (11)–(16)). As a comparison, for NiOOH, the formation of N2H4(ads) is derived from the recombination of two NH2(ads) at hydroxy sites. The lower free energy of the N2H4(ads) formation confirms that the Ni0.5–Co0.5-OOH anode exhibits a higher N2 selectivity for the EAOR than NiOOH. In addition, the NH3 conversion path is further analysed to explore the EAOR activity of the Ni0.5–Co0.5-OOH anode. As shown in Fig. 5f and g, after the formation of the N2H4(ads) intermediate, its continuous dehydrogenation into N2 occurs at Co3+ sites. The rate-limiting step of the EAOR with the Ni0.5–Co0.5-OOH anode is the formation of N2H2(ads) (1.56 eV), which is the lowest in comparison with the O–S mechanism of CoOOH (2.02 eV) and G–M mechanism of NiOOH (3.25 eV). This result further verifies that the EAOR activity of the Ni0.5–Co0.5-OOH anode is superior to that of CoOOH. In summary, due to the synergetic effect of the frustrated Lewis pairs in Ni0.5–Co0.5-OOH, the O–S and G–M mechanisms are integrated, and the recombination of NH2(ads) intermediates into N2H4(ads) intermediates is facilitated, thereby immensely promoting the EAOR activity and N2 formation selectivity.

 
CoIIIOOH + NH3 → CoIIIOOH–NH3(ads) (11)
 
CoIIIOOH–NH3(ads) → CoII(OH)2–NH2(ads) (12)
 
NiIIIOOH + NH3 → NiIIIOOH–NH3(ads) (13)
 
NiIIIOOH–NH3(ads) → NiII(OH)2–NH2(ads) (14)
 
CoII(OH)2–NH2(ads) + NiII(OH)2–NH2(ads) → CoII(OH)2–N2H4(ads) + NiII(OH)2 (15)
 
CoII(OH)2–N2H4(ads) + 5OH → CoIIIOOH + N2 + 4H2O + 5e (16)

Conclusions

In conclusion, this study developed a bimetallic nickel–cobalt oxyhydroxide (Ni–Co-OOH) to integrate O–S and G–M mechanisms together. The synergy of the frustrated Lewis acid–base pairs is utilized for the selective conversion of NH3 into N2. Noteworthily, two NH3 molecules bind to Co3+ sites (Lewis acid) in CoOOH and hydroxy sites (Lewis base) in NiOOH. Then, the NH2(ads) desorbed from the hydroxy sites can rapidly recombine with the NH2(ads) generated in Co3+ into N2H4(ads), leading to the formation of N2. The Ni–Co-OOH anode exhibits excellent activity (94%) and selectivity (63%) for the EAOR beyond other catalysts. This work provides a new idea for designing and preparing novel catalysts to improve the performance of ammonia conversion technology.

Author contributions

Meng-Ying Yin: writing – original draft, formal analysis, data curation, conceptualization. Xing-Yuan Xia: formal analysis, data curation. Ting Dai: formal analysis, data curation, software. Xia Chen: formal analysis, conceptualization, software. Qiu-Ju Xing: resources, supervision, visualization. Lei Tian: project administration, writing – review & editing, methodology. Jian-Ping Zou: writing – review & editing, methodology, resources, project administration, funding acquisition.

Conflicts of interest

There are no conflicts to declare.

Data availability

Data will be made available on request.

The general information, methods and supplementary data associated with this article are provided in the supplementary information (SI). See DOI: https://doi.org/10.1039/d5sc06524k.

Acknowledgements

We gratefully acknowledge the financial support of the National Natural Science Foundation of China (Grant No. 52470079, 52170082, 52300081, 51938007, and 52100186), Natural Science Foundation of Jiangxi Province (Grant No. 20212ACB203008), Science and Technology Department Project of Jiangxi Province (No. 20223AEI91001), and Key Laboratory of Jiangxi Province for Persistent Pollutants Prevention Control and Resource Reuse (No. 2023SSY02061).

Notes and references

  1. M. E. Ahmed, M. R. Boroujeni, P. Ghosh, C. Greene, S. Kundu, J. A. Bertke and T. H. Warren, Electrocatalytic Ammonia Oxidation by a Low-Coordinate Copper Complex, J. Am. Chem. Soc., 2022, 144, 21136–21145 CrossRef CAS PubMed.
  2. T. Gupta, Sanyam, S. Saraswat, A. Mondal and B. Mondal, Electrochemical Ammonia Oxidation with a Homogeneous Molecular Redox Mediator, Chem. Sci., 2025, 16, 14377–14383 RSC.
  3. Y. Chen, G. Zhang, Q. Ji, H. Lan, H. Liu and J. Qu, Visualization of Electrochemically Accessible Sites in Flow-through Mode for Maximizing Available Active Area toward Superior Electrocatalytic Ammonia Oxidation, Environ. Sci. Technol., 2022, 56, 9722–9731 CrossRef CAS PubMed.
  4. M. E. Ahmed, R. J. Staples, T. R. Cundari and T. H. Warren, Electrocatalytic Ammonia Oxidation by Pyridyl-Substituted Ferrocenes, J. Am. Chem. Soc., 2025, 144, 6514–6522 CrossRef PubMed.
  5. S. Bose, Y. Xia and R. N. Zare, Understanding the formation of nitrate from nitrogen at the interface of gas-water microbubbles, Chem. Sci., 2024, 15, 19764–19769 RSC.
  6. H. Y. Liu, H. M. C. Lant, J. L. Troiano, G. Hu, B. Q. Mercado, R. H. Crabtree and G. W. Brudvig, Electrocatalytic, Homogeneous Ammonia Oxidation in Water to Nitrate and Nitrite with a Copper Complex, J. Am. Chem. Soc., 2022, 144, 8449–8453 CrossRef CAS PubMed.
  7. S. He, Y. Chen, M. Wang, H. Nuomin, P. Novello, X. Li, S. Zhu and J. Liu, Metal nitride nanosheets enable highly efficient electrochemical oxidation of ammonia, Nano Energy, 2021, 80, 105528 CrossRef CAS.
  8. L. Tian, P. Chen, X. H. Jiang, L. S. Chen, L. L. Tong, H. Y. Yang, J. P. Fan, D. S. Wu, J. P. Zou and S. L. Luo, Mineralization of cyanides via a novel Electro-Fenton system generating •OH and •O2-, Water Res., 2022, 209, 117890 CrossRef CAS PubMed.
  9. Q. Jiang, D. M. Xia, X. F. Li, H. Zhang, R. J. Yin, H. J. Xie, H. B. Xie, J. Jiang, J. W. Chen and J. S. Francisco, Rapid N2O formation from N2 on water droplet surfaces, Angew. Chem., Int. Ed., 2025, 64, e202421002 CrossRef CAS PubMed.
  10. F. Habibzadeh, S. L. Miller, T. W. Hamann and M. R. Smith, Homogeneous electrocatalytic oxidation of ammonia to N2 under mild conditions, Proc. Natl. Acad. Sci. U. S. A., 2019, 116, 2849–2853 CrossRef CAS PubMed.
  11. L. Tian, M. Y. Yin, L. L. Zheng, Y. Chen, W. Liu, J. P. Fan, D. S. Wu, J. P. Zou and S. L. Luo, Extremely efficient mineralizing CN- into N2 via a newly developed system of generating sufficient ClO•/Cl2•- and self-decreasing pH, Sep. Purif. Technol., 2023, 309, 123021 CrossRef CAS.
  12. S. I. Venturini, D. R. Martins De Godoi and J. Perez, Challenges in Electrocatalysis of Ammonia Oxidation on Platinum Surfaces: Discovering Reaction Pathways, ACS Catal., 2023, 13, 10835–10845 CrossRef CAS.
  13. Y. Li, H. S. Pillai, T. Wang, S. Hwang, Y. Zhao, Z. Qiao, Q. Mu, S. Karakalos, M. Chen, J. Yang, D. Su, H. Xin, Y. Yan and G. Wu, High-performance ammonia oxidation catalysts for anion-exchange membrane direct ammonia fuel cells, Energy Environ. Sci., 2021, 14, 1449–1460 RSC.
  14. F. Fang, Q. Y. Cheng, M. F. Wang, Y. Z. He, Y. F. Huan, S. S. Liu, T. Qian, C. L. Yan and J. M. Lu, Identifying Upper d-Band Edge as Activity Descriptor for Ammonia Oxidation on PtCo Alloys in Low-Temperature Direct Ammonia Fuel Cells, ACS Nano, 2025, 19, 1260–1270 CrossRef CAS PubMed.
  15. L. Tian, X. Y. Xia, L. J. Zhou, L. L. Zheng, Q. J. Xing, L. S. Zhang, J. P. Zou and Z. Q. Liu, Enhanced chlorine enrichment via electron-deficient centers of Co(III) for efficient electrochlorination and ammonia removal, Appl. Catal., B, 2024, 340, 123260 CrossRef CAS.
  16. Z. P. Yu, G. J. Xia, V. M. Diaconescu, L. Simonelli, A. P. LaGrow, Z. X. Tai, X. Y. Xiang, D. H. Xiong and L. F. Liu, Atomically dispersed dinuclear iridium active sites for efficient and stable electrocatalytic chlorine evolution reaction, Chem. Sci., 2024, 15, 9216–9223 RSC.
  17. J. Z. He, C. Zhang, Y. Yang, J. W. Kang, C. Y. Zhang, D. He and J. X. Ma, Chlorine-Mediated Ammonia and Organics Transformation during Electrochemical Ammonia Recovery from Human Urine, Environ. Sci. Technol., 2025, 59, 13096–13107 CrossRef CAS PubMed.
  18. H. G. Oswin and M. Salomon, The anodic oxidation of ammonia at platinum black electrodes in aqueous KOH electrolyte, Can. J. Chem., 1963, 41, 1686–1694 CrossRef CAS.
  19. H. Y. Liu, H. M. C. Lant, C. Decavoli, R. H. Crabtree and G. W. Brudvig, pH-Dependent Electrocatalytic Aqueous Ammonia Oxidation to Nitrite and Nitrate by a Copper(II) Complex with an Oxidation-Resistant Ligand, J. Am. Chem. Soc., 2025, 147, 1624–1630 CrossRef CAS PubMed.
  20. G. Chen, X. L. Ding, P. He, T. Cheng, Y. Chen, J. Lin, X. Zhang, S. Zhao, N. Qiao and X. Y. Yi, Understanding the factors governing the ammonia oxidation reaction by a mononuclear ruthenium complex, Chem. Sci., 2025, 16, 7573–7578 RSC.
  21. M. Anand, C. S. Abraham and J. K. Nørskov, Electrochemical oxidation of molecular nitrogen to nitric acid -towards a molecular level understanding of the challenges, Chem. Sci., 2021, 12, 6442–6448 RSC.
  22. H. Gerischer and A. Mauerer, Untersuchungen Zur anodischen Oxidation von Ammoniak an Platin-Elektroden, J. Electroanal. Chem. Interfacial Electrochem., 1970, 25, 421–433 CrossRef CAS.
  23. K. Yang, J. Liu and B. Yang, Mechanism and Active Species in NH3 Dehydrogenation under an Electrochemical Environment: An Ab initio Molecular Dynamics Study, ACS Catal., 2021, 11, 4310–4318 CrossRef CAS.
  24. X. Chen, Z. Y. Ke, X. Wang, H. Q. Jin, Y. W. Cheng, Y. K. Xiao, R. Jiang, Y. M. Da, L. Fan, H. X. Li, D. M. Liu, S. F. Yang and W. Chen, Fullerene Network-Buffered Platinum Nanoparticles Toward Efficient and Stable Electrochemical Ammonia Oxidation Reaction for Hydrogen Production, Angew. Chem., Int. Ed., 2025, 137, e202505180 CrossRef.
  25. G. L. Li, Y. H. Shi, R. Y. Huang, T. G. Ma, Q. Zhuo, L. Xu, F. Jiang, Y. K. Zhai, Y. H. Xiao, Y. Yan and Y. Zhao, An ingenious structural design of NiCu-MOF bar-like nanosheet array: harnessing synergistic effects to attain exceptional ammonia oxidation performance, Chem. Eng. J., 2025, 521, 166548 CrossRef CAS.
  26. X. Yang, L. Sun, X. Liu, Z. Yang, H. Sun, W. Liu and H. Chen, Vacancy-driven ammonia electrooxidation reaction on the nanosized CeOx electrode in nonaqueous electrolyte, ACS Catal., 2024, 14, 6236–6246 CrossRef CAS.
  27. H. Y. Fu, F. C. Gu, Y. Z. Niu, S. X. Liao, Z. Y. Bu, H. N. Wang, D. Yang, X. S. Wang and Q. Li, Spatially confined transition metals boost high initial coulombic efficiency in alloy anodes, Chem. Sci., 2025, 16, 418–424 RSC.
  28. D. Sun, X. Hong, K. Wu, K. S. Hui, Y. Du and K. N. Hui, Simultaneous removal of ammonia and phosphate by electro-oxidation and electrocoagulation using RuO2-IrO2/Ti and microscale zero-valent iron composite electrode, Water Res., 2020, 169, 115239 CrossRef CAS PubMed.
  29. C. X. Liu, Z. F. Teng, X. Liu, R. Zhang, J. Q. Chi, J. W. Zhu, J. F. Qin, X. B. Liu, Z. X. Wu and L. Wang, Stabilizing NiFe active sites using a high-valent Lewis acid for selective seawater oxidation, Chem. Sci., 2025, 16, 16321–16330 RSC.
  30. F. Chang, W. Gao, J. Guo and P. Chen, Emerging Materials and Methods toward Ammonia-Based Energy Storage and Conversion, Adv. Mater., 2021, 33, 2005721 CrossRef CAS PubMed.
  31. Z. Hu, S. Lu, F. Tang, D. Yang, C. Zhang, Q. Xiao and P. Ming, High-performance precious metal-free direct ammonia fuel cells endowed by Co-doped Ni4Cu1 anode catalysts, Appl. Catal., B, 2023, 334, 122856 CrossRef CAS.
  32. H. Zhong, Q. Zhang, J. Yu, X. Zhang, C. Wu, Y. Ma, H. An, H. Wang, J. Zhang, X. Wang and J. Xue, Fundamental Understanding of Structural Reconstruction Behaviors in Oxygen Evolution Reaction Electrocatalysts, Adv. Energy Mater., 2023, 13, 2301391 CrossRef CAS.
  33. X. Jiang, D. Ying, X. Liu, M. Liu, S. Zhou, C. Guo, G. Zhao, Y. Wang and J. Jia, Identification of the role of Cu site in Ni-Cu hydroxide for robust and high selective electrochemical ammonia oxidation to nitrite, Electrochim. Acta, 2020, 345, 136157 CrossRef CAS.
  34. J. J. Medvedev, Y. Tobolovskaya, X. V. Medvedeva, S. W. Tatarchuk, F. Li and A. Klinkova, Pathways of ammonia electrooxidation on nickel hydroxide anodes and an alternative route towards recycled fertilizers, Green Chem., 2022, 24, 1578–1589 RSC.
  35. APHA, AWWA and WEF, Standard Methods for the Examination of Water and Wastewater, APHA, Washington, DC, 22th edn, 2012 Search PubMed.
  36. M. Pandurangappa and Y. Venkataramanappa, Aminophenyl Benzimidazole as a New Reagent for the Estimation of NO2/Nitrite/Nitrate at Trace Level: Application to Environmental Samples, Anal. Lett., 2007, 40, 2974–2991 CrossRef CAS.
  37. W. J. Duan, G. Li, Z. C. Lei, T. G. Zhu, Y. Z. Xue, C. H. Wei and C. H. Feng, Highly active and durable carbon electrocatalyst for nitrate reduction reaction, Water Res., 2019, 161, 126–135 CrossRef CAS PubMed.
  38. Y. H. Wang, F. Y. Gao, X. L. Zhang, Y. Yang, J. Liao, Z. Z. Niu, S. Qin, P. P. Yang, P. C. Yu, M. Sun and M. R. Gao, Efficient NH3-Tolerant Nickel-Based Hydrogen Oxidation Catalyst for Anion Exchange Membrane Fuel Cells, J. Am. Chem. Soc., 2023, 145, 17485–17494 CrossRef CAS PubMed.
  39. M. Zhu, Y. Yang, S. Xi, C. Diao, Z. Yu, W. S. V. Lee and J. Xue, Deciphering NH3 Adsorption Kinetics in Ternary Ni-Cu-Fe Oxyhydroxide toward Efficient Ammonia Oxidation Reaction, Small, 2021, 17, 2005616 CrossRef CAS PubMed.
  40. N. Yao, G. Wang, H. Jia, J. Yin, H. Cong, S. Chen and W. Luo, Intermolecular Energy Gap-Induced Formation of High-Valent Cobalt Species in CoOOH Surface Layer on Cobalt Sulfides for Efficient Water Oxidation, Angew. Chem., Int. Ed., 2022, 61, e202117178 CrossRef CAS PubMed.
  41. Z. Zhou, Y. Xie, L. Sun, Z. Wang, W. Wang, L. Jiang, X. Tao, L. Li, X. H. Li and G. Zhao, Strain-induced in situ formation of NiOOH species on Co-Co bond for selective electrooxidation of 5-hydroxymethylfurfural and efficient hydrogen production, Appl. Catal., B, 2022, 305, 121072 CrossRef CAS.
  42. L. Y. Hao, Z. J. Tang, C. Y. Cai, Y. C. Zhao, L. Tian, N. Li and Z. Q. Liu, Electron-delocalized Cu2+ activates spin channels in spinel oxides to selectively produce 1O2 for wastewater treatment, Angew. Chem., Int. Ed., 2025, 64, e202504426 CrossRef CAS PubMed.
  43. L. Tian, L. Zhang, L. Zheng, Y. Chen, L. Ding, J. Fan, D. Wu, J. Zou and S. Luo, Overcoming Electrostatic Interaction via Strong Complexation for Highly Selective Reduction of CN- into N2, Angew. Chem., Int. Ed., 2022, 61, e202214145 CrossRef CAS PubMed.
  44. J. J. Liang, C. Y. Du, Y. J. Xian, C. Y. Cai, L. Tian, Y. H. Ao, P. F. Wang and Z. Q. Liu, Accelerating *OH desorption via electron-delocalized CuTd2+-O-CoOh3+ for water purification, Adv. Funct. Mater., 2025, 35, 2503514 CrossRef CAS.
  45. L. Tian, Z. J. Tang, L. Y. Hao, T. Dai, J. P. Zou and Z. Q. Liu, Efficient homolytic cleavage of H2O2 on hydroxyl-enriched spinel CuFe2O4 with dual Lewis acid sites, Angew. Chem., Int. Ed., 2024, 63, e202401434 CrossRef CAS PubMed.
  46. V. T. T. Phan, Q. P. Nguyen, B. Wang and I. J. Burgess, Oxygen Vacancies Alter Methanol Oxidation Pathways on NiOOH, J. Am. Chem. Soc., 2024, 146, 4830–4841 CrossRef CAS PubMed.
  47. Z. Tian, M. Wang, G. Chen, J. Chen, Y. Da, H. Zhang, R. Jiang, Y. Xiao, B. Cui, C. Jiang, Y. Ding, J. Yang, Z. Sun, C. Han and W. Chen, n-ZrS3/p-ZrOS Photoanodes with NiOOH/FeOOH Oxygen Evolution Catalysts for Photoelectrochemical Water Oxidation, Angew. Chem., Int. Ed., 2024, e202414209 Search PubMed.
  48. M. Gonzalez-Reyna, M. S. Luna-Martínez and J. F. Perez-Robles, Nickel supported on carbon nanotubes and carbon nanospheres for ammonia oxidation reaction, Nanotechnology, 2020, 31, 235706 CrossRef CAS PubMed.
  49. J. Yao, M. Zhou, D. Wen, Q. Xue and J. Wang, Electrochemical conversion of ammonia to nitrogen in non-chlorinated aqueous solution by controlling pH value, J. Electroanal. Chem., 2016, 776, 53–58 CrossRef CAS.
  50. J. Song, Y. Yinhai, Y. Jia, T. Wang, J. Wei, M. Wang, S. Zhou, Z. Li, Y. Hou, L. Lei and B. Yang, Improved NH3-N conversion efficiency to N2 activated by BDD substrate on NiCu electrocatalysis process, Sep. Purif. Technol., 2021, 276, 119350 CrossRef CAS.
  51. Y. J. Shih and C. H. Hsu, Kinetics and highly selective N2 conversion of direct electrochemical ammonia oxidation in an undivided cell using NiCo oxide nanoparticle as the anode and metallic Cu/Ni foam as the cathode, Chem. Eng. J., 2021, 409, 128024 CrossRef CAS.
  52. R. Wang, H. Liu, K. Zhang, G. Zhang, H. Lan and J. Qu, Ni(II)/Ni(III) redox couple endows Ni foam-supported Ni2P with excellent capability for direct ammonia oxidation, Chem. Eng. J., 2021, 404, 126795 CrossRef CAS.
  53. S. S. P. Rahardjo and Y. J. Shih, Electrochemical characteristics of silver/nickel oxide (Ag/Ni) for direct ammonia oxidation and nitrogen selectivity in paired electrode system, Chem. Eng. J., 2023, 452, 139370 CrossRef CAS.
  54. Y. J. Shih, Y. H. Huang and C. P. Huang, Electrocatalytic ammonia oxidation over a nickel foam electrode: role of Ni(OH)2(s)-NiOOH(s) nanocatalysts, Electrochim. Acta, 2018, 263, 261–271 CrossRef CAS.
  55. P. Zhou, S. Zhang, Z. Ren, X. Tang, K. Zhang, R. Zhou, D. Wu, J. Liao, Y. Zhang and C. Huang, In Situ Cutting of Ammonium Perchlorate Particles by Co-Bipy “scalpel” for High Efficiency Thermal Decomposition, Adv. Sci., 2022, 9, 2204109 CrossRef CAS PubMed.
  56. Y. Peng, W. Si, X. Li, J. Chen, J. Li, J. Crittenden and J. Hao, Investigation of the Poisoning Mechanism of Lead on the CeO2-WO3 Catalyst for the NH3-SCR Reaction via in Situ IR and Raman Spectroscopy Measurement, Environ. Sci. Technol., 2016, 50, 9576–9582 CrossRef CAS PubMed.
  57. R. M. S. Yoo, D. Yesudoss, D. Johnson and A. Djire, A Review on the Application of In-situ Raman Spectroelectrochemistry to Understand the Mechanisms of Hydrogen Evolution Reaction, ACS Catal., 2023, 13, 10570–10601 CrossRef CAS.
  58. L. Bai, F. Franco, J. Timoshenko, C. Rettenmaier, F. Scholten, H. S. Jeon, A. Yoon, M. Rüscher, A. Herzog, F. T. Haase, S. Kühl, S. W. Chee, A. Bergmann and R. C. Beatriz, Electrocatalytic Nitrate and Nitrite Reduction toward Ammonia Using Cu2O Nanocubes: Active Species and Reaction Mechanisms, J. Am. Chem. Soc., 2024, 146, 9665–9678 CrossRef CAS PubMed.
  59. H. Chen, H. Ze, M. Yue, D. Wei, Y. A, Y. Wu, J. Dong, Y. Zhang, H. Zhang, Z. Tian and J. Li, Unmasking the Critical Role of the Ordering Degree of Bimetallic Nanocatalysts on Oxygen Reduction Reaction by In situ Raman Spectroscopy, Angew. Chem., Int. Ed., 2022, 134, e202117834 CrossRef.

This journal is © The Royal Society of Chemistry 2025
Click here to see how this site uses Cookies. View our privacy policy here.