Ni-loaded Co-NC catalysts for promoting electrocatalytic nitrate reduction to ammonia

Fang Zhaoa, Yidi Liua, Chengjie Lib, Zhen Yuana, Qianqian Huaa, Liguo Gaoa, Xuefeng Ren*a, Peixia Yang*c and Anmin Liu*a
aSchool of Chemical Engineering, Ocean and Life Sciences, Dalian University of Technology, Panjin 124221, China. E-mail: renxuefeng@dlut.edu.cn; liuanmin@dlut.edu.cn
bShandong Engineering Research Center of Green and High-value Marine Fine Chemical, Weifang University of Science and Technology, Weifang 262700, China
cMIIT Key Laboratory of Critical Materials Technology for New Energy Conversion and Storage, State Key Laboratory of Urban Water Resource and Environment, School of Chemistry and Chemical Engineering, Harbin Institute of Technology, Harbin 150001, China. E-mail: yangpeixia@hit.edu.cn

Received 30th July 2024 , Accepted 20th August 2024

First published on 20th August 2024


Abstract

The Haber–Bosch process, the traditional method of ammonia synthesis, uses hydrogen derived from steam reforming of hydrocarbons; and involves harsh operating conditions of high temperatures (300–600 °C) and pressures (200–400 atm), expending a vast amount of energy each year. Recently, there has been a lot of interest in the electrochemical nitrogen reduction process (NRR) to NH3, which is inspired by natural microbial nitrogen fixation. However, the stable N[triple bond, length as m-dash]N violent hydrogen evolution reaction hindered the development of the NRR. In comparison, the electrocatalytic nitrate reduction reaction (NO3RR) has significant advantages. The much lower dissociation energy of N[double bond, length as m-dash]O (204 kJ mol−1) is required; nitrate is widespread in surface water. Herein, an electrocatalyst loaded with Ni onto CoZn@ZIF by a simple impregnation method is reported, which possesses a nitrogen-doped graphitic carbon structure after pyrolytic carbonization. Detailed experiments showed that the NiCo-NC catalyst significantly accelerated the NO3RR compared to Co-NC. NiCo-NC exhibited remarkable NO3RR activity. At −0.6 V and −1.1 V, ammonia yields of 5.01 mg cm−2 h−1 and 10.12 mg cm−2 h−1 were obtained, with FEs reaching 92.75% and 96.65%, respectively. The catalyst showed excellent electrochemical stability in 24-hour electrolysis experiments and five-cycle stability tests. Meanwhile, 15N isotope labeling experiments further verified the source of N in NH4+ from NO3.


1. Introduction

Ammonia is an essential component of artificial fertilizers and a perfect hydrogen-rich, carbon-free energy source.1 Currently, ammonia synthesis relies heavily on the Haber–Bosch process, which involves hydrogen derived from steam reforming of hydrocarbons, carried out at high temperatures and pressures.2–4 Over 300 million tons of CO2 emissions are produced annually from energy consumption, which makes up 2% of the world's total energy consumption. In recent years, inspired by natural microbial nitrogen fixation, the electrochemical nitrogen reduction reaction (NRR) to NH3 has attracted great interest.5,6 However, nitrogen's industrial utilization is hindered by its minimal water solubility, the stable N[triple bond, length as m-dash]N bond (dissociation energy 941 kJ mol−1) and the intense competition for hydrogen evolution reactions.7,8 These factors lead to low NH3 and low current densities. Compared with the NRR, the interesting method for synthesizing NH3 under ambient conditions is the electrochemical nitrate reduction reaction (NO3RR), which is thermodynamically more viable than the NRR because of the lower N[double bond, length as m-dash]O bond dissociation energy of 204 kJ mol−1.9 In addition, residual NO3 in wastewater removed from wastewater treatment plants is an essential source of nitrogen.

Nonetheless, ammonia yield and selectivity are still constrained by competition from a variety of by-products (NO2, NO2, NO, N2O, N2, NH2OH, NH2NH2).10 The NO3RR is an eight-electron, nine-proton reaction mechanism.11 More seriously, nitrites are a prevalent by-product of the NO3RR and are more toxic and carcinogenic than nitrates.12 Among the reported NO3RR catalysts, several types of catalysts are included, such as non-metallic, noble metal, and transition metal.9 Although catalysts based on precious metals have demonstrated potent activity in the NO3RR, their broad applicability is constrained by the high cost. Transition metals are widely used in electrocatalysis due to their abundant reserves, metallic nature, and environmental friendliness.

Metal–organic framework (MOF) derived materials combine the advantages of organic and inorganic materials and have received increasing attention for their higher porosity, good electrical conductivity, and more metal–ligand sites.13,14 Metal nanomaterials produced by pyrolysis of MOFs can retain the precursor's characteristics. Metal cations are linked to ligand molecules in MOFs, and the arrangement and localization of ligands are determined by the coordination configuration.15,16 A zeolitic imidazolate framework (ZIF) is an ideal carrier for the construction of catalyst metal sites by pyrolytically converting other elements, such as nitrogen, into graphitic nitrogen and pyridine nitrogen.17,18 However, the NO3RR electrocatalytic process involves complex 8e and 9H* transfer mechanisms.19 Bimetallic composite materials have more advantages over monometallic compounds, such as stronger metal-to-metal interactions, enhanced charge transfer capabilities, and optimized electronic structures, showing more excellent catalytic activity and stability.20,21 Co-based catalysts have been reported for the NO3RR,22–24 such as CuCo2O4,25 PP–Co,26 Co3O4/CC,27 Cu–Co3O4.28 Metal Co has excellent stability at the electrochemical reduction potential, which facilitates the dissociation of H2O and provides the required electrons. Nonetheless, electron rich metal Co will hinder the electrostatic adsorption of NO3 and the formation of *NHx substances.29 Chu et al.30 dispersed the Co single atoms on the C3N4 substrate, and the monatomic Co site can effectively activate NO2 and optimize the formation energy of key NOH intermediates. In additions, Li et al.31 prepared copper–nickel nanoparticles (CuNi/TM) on a titanium mesh via a simple and tunable electrodeposition method. The CuNi/TM catalyst showed excellent NO3RR performance. The metal Ni on the catalyst surface provides H adsorption sites where hydrogen molecules are adsorbed to form H-Ni bonds.32,33 Reduces the adsorption energies of the critical intermediates *NO3, *NO2, and *NH2.34 Consequently, the Ni–Co bimetallic catalysts can further improve the catalytic activity by coordinating the adsorption strength of the dual active sites with the intermediates and overcoming the restriction of the single active site.35,36

Herein, we synthesized CoZn@ZIF precursors in aqueous solution under ambient conditions, loaded metal Ni onto the surface of the precursors by a simple impregnation method, and subsequently successfully prepared bimetallic NiCo-NC catalysts by pyrolytic carbonization. Compared to Co-NC catalysts, they have the ability to accelerate the NO3RR eight-electron reaction and efficiently promote the synthesis of NH3. NiCo-NC has a Faraday efficiency of 92.75% and 96.65% with yields of 5.01 mg cm−2 h−1 and 10.12 mg cm−2 h−1 at −0.6 V and −1.1 V, respectively. Meanwhile, 15N isotope labeling experiments further verified the source of N in NH3 from NO3.

2. Results and discussion

2.1. Structure and morphology

The precursor CoZn@ZIF was synthesized at ambient temperature, impregnated under ultrasonic conditions to adsorb Ni2+, and pyrolytically carbonized to prepare Ni-loaded Co-NC materials with the dodecahedral structure of ZIF-8. The morphology of the catalysts during the synthesis process was investigated by scanning electron microscopy (SEM). As shown in Fig. 1, CoZn@ZIF impregnated with Ni2+ showed a slightly rough surface, and after pyrolysis at high temperatures, exhibited a slight collapse of the carbon skeleton, and had metal particles on the surface. The size of the Ni-coated Co-NC was around 1.5 μm. For comparison, we synthesized Ni-coated Co-NC materials with different Ni loadings under equivalent conditions (Fig. S1). The morphology of bimetallic NiCo-NC was further investigated by transmission electron microscopy (TEM) and energy-dispersive X-ray spectroscopy (EDS), as shown in Fig. 2. The catalyst possesses a dodecahedral framework of ZIF-8, and the surface nanoparticles are in the range of 50 nm in diameter. The high-resolution transmission electron microscope (HRTEM) image of NiCo-NC shows a lattice spacing of 0.345 nm, which corresponds to the (002) plane of C. EDS mapping (Fig. 2c) analysis showed that the elements Ni, Co, N, and C were uniformly distributed throughout the amorphous porous structure, indicating that Ni and Co are highly dispersed coexisting substances.
image file: d4cy00942h-f1.tif
Fig. 1 (a) Synthesis diagram of the Ni-coated Co-NC composite material; SEM image of (b) CoZn@ZIF, (c) Ni loaded CoZn@ZIF and (d) pyrolyzed Ni-coated Co-NC.

image file: d4cy00942h-f2.tif
Fig. 2 (a) TEM image of bimetallic NiCo-NC; (b) HRTEM image of NiCo-NC; (c) TEM images of Co, Ni, N, and C and corresponding EDS elemental mappings.

The synthesized catalysts were analyzed for the crystal structure and chemical state by X-ray polycrystalline diffraction (XRD) and X-ray photoelectron spectroscopy (XPS). As shown in Fig. 3a, the metallic peaks of Zn are negligible. The diffraction peaks at 2θ = 44.32°, 51.48° and 75.78° correspond to the (111), (200), and (220) planes of NiCo metal, respectively.37–39 The diffraction peak near 2θ = 25° corresponds to the low crystallization (002) plane of graphitic carbon, which corresponds to the HRTEM image. In addition, the XRD patterns of precursors NiCoZn@ZIF and CoZn@ZIF are nearly similar to the pattern of ZIF-8, indicating that the absorption of Ni2+ ions did not destroy the ZIF crystal structure (Fig. S3a). XPS was performed for the study of the surface composition and valence states of Co-NC and NiCo-NC. As shown in Fig. 3b, the elements determined in NiCo-NC are C, N, Co, and Ni. The C 1s orbital spectra of NiCo-NC catalysts are shown in Fig. 3c; the obvious characteristic peaks were found at binding energies of 284.8 eV, 285.5 eV, 286.8 eV, 288.65 eV, and 291.0 eV, which can be attributed to C–C, C[double bond, length as m-dash]N, and C[double bond, length as m-dash]C, and very low-content C–O and C[double bond, length as m-dash]O peaks, respectively, indicating high graphitization. The C 1s spectra of Co-NC are given in the ESI (Fig. S3c). As shown in Fig. 3d, the N 1s peaks at binding energies of 398.8 eV, 399.65 eV, and 401.2 eV were attributed to pyridine N, M–N, graphitic N and some oxygenated moieties (oxidized N and pyridine-N-oxide), respectively, which indicated that ZIF imidazole during pyrolysis was successfully converted to graphitic carbon. Furthermore, the catalytic efficiency is enhanced by pyridine N and graphite N, which work together to synergize the electron distribution of the active sites off-domain, providing enough anchoring for the monometallic sites of the M–N–C catalysts.40 The Co 2p peaks of the NiCo-NC catalysts at binding energies of 778.55 eV and 795.0 eV correspond to Co3+ species in the 2p3/2 and 2p1/2 orbitals, respectively, and the peaks at 780.35 eV and 796.7 eV correspond to Co2+ species in the 2p3/2 and 2p1/2 orbitals. Similar results have also been observed from Co-NC. The Ni 2p spectrum in Fig. 3f demonstrates Ni2+ 2p3/2 and 2p1/2 peaks at 854.4 eV and 871.5 eV, respectively.41


image file: d4cy00942h-f3.tif
Fig. 3 (a) XRD spectra of NiCo-NC and Co-NC; (b) XPS survey spectra of NiCo-NC; XPS spectra of (c) C 1s for NiCo-NC; (d) N 1s for NiCo-NC and Co-NC; (e) Co 2p for NiCo-NC and Co-NC; (f) Ni 2p for NiCo-NC.

2.2. Electrocatalytic performance of catalysts in the NO3RR

Initially, the NO3RR catalytic activity and onset potentials of NiCo-NC catalysts were investigated by linear sweep voltammetry (LSV), which indicates a notable variation in current densities. As shown in Fig. 4, the current density gradually increased with the increase of applied potential. When NO3 was added to the solution, the current density changed significantly. At an applied potential of −1.0 V, the current density was as high as 117 mA cm−2, indicating the occurrence of the NO3RR on the NiCo-NC catalyst. LSV comparisons of Co-NC catalysts with different Ni loadings are shown in Fig. S3. With the addition of Ni, the current density gradually increased. When the addition of Ni was 0.4 mmol, the current density decreased to 105 mA cm−2. The maximum current density was achieved when the Ni loading during the synthesis of the material was 0.3 mmol. Therefore, all subsequent tests were based on NiCo-NC catalysts with a Ni loading of 0.3 mmol. Subsequently, we measured the current density variations of NiCo-NC catalysts in an electrolyte of 0.1 M KOH & KNO3 at different potentials and performed electrochemical tests for 2 h under an Ar atmosphere. As shown in Fig. 4b, the current density increases gradually with the increase of applied potential, and the current density curves remain stable at different potentials without large fluctuations, indicating that NiCo-NC has excellent electrochemical stability in the catalytic NO3RR process. The content of NH4+ in the electrolyte was determined using indophenol blue colorimetry (Fig. 4c). NiCo-NC obtained 92.75% FE and 5.01 mg cm−2 h−1 ammonia yield at −0.6 V; 96.65% FE and 10.12 mg cm−2 h−1 ammonia yield at −1.1 V. The ammonia yield and FE were maintained at a high level in the −0.6–−1.1 V interval of the applied potential. The FE of the by-product NO2 showed completely opposite properties (Fig. 4d), with both the by-product nitrite yield and the FE as a whole remaining at a low level. The NiCo-NC catalyst could effectively inhibit the accumulation of by-products and promote the reduction of NO3 to NH3.
image file: d4cy00942h-f4.tif
Fig. 4 (a) Linear scanning voltammetry curve of NiCo-NC; (b) it test with different applied potentials; (c) NiCo-NC ammonia yield and FE at different potentials; (d) comparison of the FE of NiCo-NC – nitrite and ammonia.

The long-term stability and cycling stability of the catalyst are important factors in evaluating the performance of the catalyst. The NiCo-NC catalyst was subjected to five cycle stability tests at −0.6 V and −1.1 V and 24 h cycle stability test. As shown in Fig. 5, NiCo-NC performed remarkable cycling stability at a potential of −0.6 V, where both NH4+ and FE were well maintained at a potential of −1.1 V. There was no discernible drop in current density throughout the 24 h long term current test (Fig. 5d), suggesting that the NiCo-NC catalyst has outstanding durability.


image file: d4cy00942h-f5.tif
Fig. 5 (a) Five cycle stability it tests at applied potentials of −0.6 V and −1.1 V; (b) (c) ammonia yield and FE corresponding to cycle stability tests at −0.6 V and −1.1 V, respectively; (d) 24 h long term stability i–t tests at applied potentials of −0.6 V and −1.1 V; (e) comparison of the ammonia yield and FE of Co-NC and NiCo-NC at −0.6 V; (f) comparison of the ammonia yield and FE of Co-NC and NiCo-NC at −1.1 V.

Based on the above analysis, we further explored the performance differences between NiCo-NC and Co-NC. As shown in Fig. 5e, the ammonia yields of Co-NC at applied potentials of −0.6 V and −1.1 V were 3.85 mg cm−2 h−1 and 8.32 mg cm−2 h−1, and the FEs were 82.4% and 86.4%; which were much lower than the ammonia yields and the FEs of NiCo-NC. The superiority of NiCo-NC over Co-NC for the NO3RR was further confirmed. To demonstrate that the N in NH4+ is derived from NO3, we designed and performed a comparative test in a blank 0.1 M KOH solution (Fig. 6). Negligible NH3 eliminates contamination from catalysts, electrolytes, or other external environments that might contain N. Furthermore, 15N and 14N isotopic labeling experiments were performed to demonstrate the N source of the product NH3 detected by 1H NMR spectroscopy. After electrocatalytic K15NO3 reduction, 1H NMR of the electrolyte showed typical 15NH4+ double peaks and no typical 14NH4+ triple peaks, confirming that the NH4+ formed was exclusively from the electroreduction of NO3 (Fig. 6c).


image file: d4cy00942h-f6.tif
Fig. 6 (a) UV-vis absorption spectra of NH4+ under different conditions of Ni-coated Co-NC; (b) NH4+ yields under different conditions; (c) 1H NMR spectra of the electrolyte after the NO3RR with 15NO3 and 14NO3 as nitrogen sources.

3. Conclusion

In summary, Ni-coated Co-NC catalysts were successfully prepared by pyrolytic carbonization through a simple impregnation method by loading metal Ni onto CoZn@ZIF precursors. Ni-coated Co-NC was experimentally demonstrated to be an efficient electrocatalyst for ammonia synthesis. Ammonia yields of 5.01 mg cm−2 h−1 and 10.12 mg cm−2 h−1 were achieved at applied potentials of −0.6 V and −1.1 V, with FEs of 92.75% and 96.65%, respectively. 15N isotope labeling experiments further verified that the source of N in NH4+ is derived from NO3.

Data availability

The data used to support the paper are included within the article.

Conflicts of interest

There are no conflicts to declare.

Acknowledgements

Support from the National Natural Science Foundation of China (21902021, 21908017), the Fundamental Research Funds for the Central Universities (DUT22LK09), the Open Foundation of Key Laboratory of Industrial Ecology and Environmental Engineering, MOE (KLIEEE-20-01, KLIEEE-21-02), the Foundation of State Key Laboratory of High-efficiency Utilization of Coal and Green Chemical Engineering (2022-K70), and the Hefei Advanced Computing Center for this work is gratefully acknowledged.

References

  1. L. Ouyang, J. Liang, Y. Luo, D. Zheng, S. Sun, Q. Liu, M. S. Hamdy, X. Sun and B. Ying, Recent advances in electrocatalytic ammonia synthesis, Chin. J. Catal., 2023, 50, 6–44 CrossRef CAS .
  2. H. Liu, J. Qin, J. Mu and B. Liu, In situ interface engineered Co/NC derived from ZIF-67 as an efficient electrocatalyst for nitrate reduction to ammonia, J. Colloid Interface Sci., 2023, 636, 134–140 CrossRef CAS PubMed .
  3. F. Zhao, G. Li, Q. Hua, J. Cao, J. Song, L. Gao, T. Ma, X. Ren and A. Liu, A two-dimensional MXene-supported CuRu catalyst for efficient electrochemical nitrate reduction to ammonia, Catal. Sci. Technol., 2023, 13(19), 5543–5548 RSC .
  4. W. Guo, Z. Liang, J. Zhao, B. Zhu, K. Cai, R. Zou and Q. Xu, Hierarchical Cobalt Phosphide Hollow Nanocages toward Electrocatalytic Ammonia Synthesis under Ambient Pressure and Room Temperature, Small Methods, 2018, 2(12), 1800204 CrossRef .
  5. Y. Liu, Y. Su, X. Quan, X. Fan, S. Chen, H. Yu, H. Zhao, Y. Zhang and J. Zhao, Facile Ammonia Synthesis from Electrocatalytic N2 Reduction under Ambient Conditions on N-Doped Porous Carbon, ACS Catal., 2018, 8(2), 1186–1191 CrossRef CAS .
  6. U. K. Ghorai, S. Paul, B. Ghorai, A. Adalder, S. Kapse, R. Thapa, A. Nagendra and A. Gain, Scalable Production of Cobalt Phthalocyanine Nanotubes: Efficient and Robust Hollow Electrocatalyst for Ammonia Synthesis at Room Temperature, ACS Nano, 2021, 15(3), 5230–5239 CrossRef CAS PubMed .
  7. Z.-Y. Niu, L. Jiao, T. Zhang, X.-M. Zhao, X.-F. Wang, Z. Tan, L.-Z. Liu, S. Chen and X.-Z. Song, Boosting Electrocatalytic Ammonia Synthesis of Bio-Inspired Porous Mo-Doped Hematite via Nitrogen Activation, ACS Appl. Mater. Interfaces, 2022, 14(50), 55559–55567 CrossRef CAS PubMed .
  8. W. Tao, P. Wang, H. Li, R. Huang and G. Zhou, Engineering sulfur vacancies optimization in Ni3Co6S8 nanospheres toward extraordinarily efficient nitrate electroreduction to ammonia, Appl. Catal., B, 2023, 324, 122193 CrossRef CAS .
  9. Y. Li, S. Xiao, X. Li, C. Chang, M. Xie, J. Xu and Z. Yang, A robust metal-free electrocatalyst for nitrate reduction reaction to synthesize ammonia, Mater. Today Phys., 2021, 19, 100431 CrossRef CAS .
  10. H. Chen, C. Zhang, L. Sheng, M. Wang, W. Fu, S. Gao, Z. Zhang, S. Chen, R. Si, L. Wang and B. Yang, Copper single-atom catalyst as a high-performance electrocatalyst for nitrate-ammonium conversion, J. Hazard. Mater., 2022, 434, 128892 CrossRef CAS PubMed .
  11. J. Liang, Z. Li, L. Zhang, X. He, Y. Luo, D. Zheng, Y. Wang, T. Li, H. Yan, B. Ying, S. Sun, Q. Liu, M. S. Hamdy, B. Tang and X. Sun, Advances in ammonia electrosynthesis from ambient nitrate/nitrite reduction, Chem, 2023, 9(7), 1768–1827 CAS .
  12. Z. Zhang, N. Zhang, J. Zhang, B. Deng, Z. Cao, Z. Wang, G. Wei, Q. Zhang, R. Jia, P. Xiang and S. Xia, Critical review in electrocatalytic nitrate reduction to ammonia towards a sustainable nitrogen utilization, Chem. Eng. J., 2024, 483, 148952 CrossRef CAS .
  13. J. Dong, X. Han, Y. Liu, H. Li and Y. Cui, Metal–Covalent Organic Frameworks (MCOFs): A Bridge Between Metal–Organic Frameworks and Covalent Organic Frameworks, Angew. Chem., Int. Ed., 2020, 59(33), 13722–13733 CrossRef CAS PubMed .
  14. R. Zhao, Z. Liang, R. Zou and Q. Xu, Metal-Organic Frameworks for Batteries, Joule, 2018, 2(11), 2235–2259 CrossRef CAS .
  15. S. Xu, A. Dong, Y. Hu, Z. Yang, S. Huang and J. Qian, Multidimensional MOF-derived carbon nanomaterials for multifunctional applications, J. Mater. Chem. A, 2023, 11(18), 9721–9747 RSC .
  16. X. Wang, A. Dong, Y. Hu, J. Qian and S. Huang, A review of recent work on using metal–organic frameworks to grow carbon nanotubes, Chem. Commun., 2020, 56(74), 10809–10823 RSC .
  17. M. Zhang, H. Mao, Y. Liang and X. Yu, Recent progress in zeolitic imidazolate frameworks (ZIFs)-derived nanomaterials for effective lithium polysulfide management in lithium–sulfur batteries, J. Mater. Chem. A, 2023, 11(34), 17892–17919 RSC .
  18. V. K. Velisoju, J. L. Cerrillo, R. Ahmad, H. O. Mohamed, Y. Attada, Q. Cheng, X. Yao, L. Zheng, O. Shekhah, S. Telalovic, J. Narciso, L. Cavallo, Y. Han, M. Eddaoudi, E. V. Ramos-Fernández and P. Castaño, Copper nanoparticles encapsulated in zeolitic imidazolate framework-8 as a stable and selective CO2 hydrogenation catalyst, Nat. Commun., 2024, 15(1), 2045 CrossRef CAS PubMed .
  19. H. Zhang, H. Wang, X. Cao, M. Chen, Y. Liu, Y. Zhou, M. Huang, L. Xia, Y. Wang, T. Li, D. Zheng, Y. Luo, S. Sun, X. Zhao and X. Sun, Unveiling Cutting-Edge Developments in Electrocatalytic Nitrate-to-Ammonia Conversion, Adv. Mater., 2024, 36(16), 2312746 CrossRef CAS PubMed .
  20. R. Meng, Q. Du, N. Zhong, X. Zhou, S. Liu, S. Yin and X. Liang, A Tandem Electrocatalysis of Sulfur Reduction by Bimetal 2D MOFs, Adv. Energy Mater., 2021, 11(47), 2102819 CrossRef CAS .
  21. Q. Huo, J. Li, X. Qi, G. Liu, X. Zhang, B. Zhang, Y. Ning, Y. Fu, J. Liu and S. Liu, Cu, Zn-embedded MOF-derived bimetallic porous carbon for adsorption desulfurization, Chem. Eng. J., 2019, 378, 122106 CrossRef CAS .
  22. J. Ni, J. Yan, F. Li, H. Qi, Q. Xu, C. Su, L. Sun, H. Sun, J. Ding and B. Liu, Atomic Co–P Catalytic Pair Drives Efficient Electrochemical Nitrate Reduction to Ammonia, Adv. Energy Mater., 2024, 2400065 CrossRef CAS .
  23. T. Xie, X. He, L. He, K. Dong, Y. Yao, Z. Cai, X. Liu, X. Fan, T. Li, D. Zheng, S. Sun, L. Li, W. Chu, A. Farouk, M. S. Hamdy, C. Xu, Q. Kong and X. Sun, CoSe2 nanowire array enabled highly efficient electrocatalytic reduction of nitrate for ammonia synthesis, Chin. Chem. Lett., 2024, 35(11), 110005 CrossRef CAS .
  24. C. Ma, L. Bao, X. Fan, X. He, X. Liu, W. Chu, A. Farouk, M. S. Hamdy, S. Sun, Q. Li, M. Wu and X. Sun, Co nanoparticle-decorated radix cynanchi daniculati-derived carbon for efficient electrocatalytic nitrite reduction to ammonia, Catal. Sci. Technol., 2024, 14(11), 3007–3011 RSC .
  25. J. Wang, S. Zhang, C. Wang, K. Li, Y. Zha, M. Liu, H. Zhang and T. Shi, Ambient ammonia production via electrocatalytic nitrate reduction catalyzed by a flower-like CuCo2O4 electrocatalyst, Inorg. Chem. Front., 2022, 9(10), 2374–2378 RSC .
  26. Q. Chen, J. Liang, Q. Liu, K. Dong, L. Yue, P. Wei, Y. Luo, Q. Liu, N. Li, B. Tang, A. A. Alshehri, M. S. Hamdy, Z. Jiang and X. Sun, Co nanoparticle-decorated pomelo-peel-derived carbon enabled high-efficiency electrocatalytic nitrate reduction to ammonia, Chem. Commun., 2022, 58(26), 4259–4262 RSC .
  27. X. Xu, L. Hu, Z. Li, L. Xie, S. Sun, L. Zhang, J. Li, Y. Luo, X. Yan, M. S. Hamdy, Q. Kong, X. Sun and Q. Liu, Oxygen vacancies in Co3O4nanoarrays promote nitrate electroreduction for ammonia synthesis, Sustainable Energy Fuels, 2022, 6(18), 4130–4136 RSC .
  28. Z. Niu, S. Fan, X. Li, Z. Liu, J. Wang, J. Duan, M. O. Tadé and S. Liu, Facile Tailoring of the Electronic Structure and the d-Band Center of Copper-Doped Cobaltate for Efficient Nitrate Electrochemical Hydrogenation, ACS Appl. Mater. Interfaces, 2022, 14(31), 35477–35484 CrossRef CAS PubMed .
  29. F. Zhao, G. Hai, X. Li, Z. Jiang and H. Wang, Enhanced electrocatalytic nitrate reduction to ammonia on cobalt oxide nanosheets via multiscale defect modulation, Chem. Eng. J., 2023, 141960 CrossRef CAS .
  30. H. Zhao, J. Xiang, Z. Sun, S. Shang and K. Chu, Electroreduction of Nitrite to Ammonia over a Cobalt Single-Atom Catalyst, ACS Sustainable Chem. Eng., 2024, 12(7), 2783–2789 CrossRef CAS .
  31. R. Li, D. Zhao, L. Zhang, K. Dong, Q. Li and G. Fan, Electrodeposited copper–nickel nanoparticles as highly efficient electrocatalysts for nitrate reduction to ammonia, Sustainable Energy Fuels, 2023, 7(18), 4417–4422 RSC .
  32. Y. Men, X. Su, P. Li, Y. Tan, C. Ge, S. Jia, L. Li, J. Wang, G. Cheng, L. Zhuang, S. Chen and W. Luo, Oxygen-Inserted Top-Surface Layers of Ni for Boosting Alkaline Hydrogen Oxidation Electrocatalysis, J. Am. Chem. Soc., 2022, 144(28), 12661–12672 CrossRef CAS PubMed .
  33. Y. Luo, Z. Zhang, F. Yang, J. Li, Z. Liu, W. Ren, S. Zhang and B. Liu, Stabilized hydroxide-mediated nickel-based electrocatalysts for high-current-density hydrogen evolution in alkaline media, Energy Environ. Sci., 2021, 14(8), 4610–4619 RSC .
  34. M. Gong, W. Zhou, M.-C. Tsai, J. Zhou, M. Guan, M.-C. Lin, B. Zhang, Y. Hu, D.-Y. Wang, J. Yang, S. J. Pennycook, B.-J. Hwang and H. Dai, Nanoscale nickel oxide/nickel heterostructures for active hydrogen evolution electrocatalysis, Nat. Commun., 2014, 5(1), 4695 CrossRef CAS PubMed .
  35. J. Ma, Y. Zhang, B. Wang, Z. Jiang, Q. Zhang and S. Zhuo, Interfacial Engineering of Bimetallic Ni/Co-MOFs with H-Substituted Graphdiyne for Ammonia Electrosynthesis from Nitrate, ACS Nano, 2023, 17(7), 6687–6697 CrossRef CAS PubMed .
  36. D. Chen, J. Luo, Q. Sun, C. Han, Q. Li and J. Qian, Bimetallic phosphide nanoparticles embedded in carbon nanostrips for electrocatalytic water oxidation, Int. J. Hydrogen Energy, 2022, 47(43), 18700–18707 CrossRef CAS .
  37. Z. Cai, I. Yamada and S. Yagi, ZIF-Derived Co9–xNixS8 Nanoparticles Immobilized on N-Doped Carbons as Efficient Catalysts for High-Performance Zinc–Air Batteries, ACS Appl. Mater. Interfaces, 2020, 12(5), 5847–5856 CrossRef CAS PubMed .
  38. Y. Zhang, J. Zhao, G. Fan, L. Yang and F. Li, Robust MOF-derived carbon-supported bimetallic Ni–Co catalysts for aqueous phase hydrodeoxygenation of vanillin, Dalton Trans., 2022, 51(6), 2238–2249 RSC .
  39. J. Chen, M. R. Ahasan, J.-S. Oh, J. A. Tan, S. Hennessey, M. M. Kaid, H. M. El-Kaderi, L. Zhou, K. U. Lao, R. Wang and W.-N. Wang, Highly efficient CO2 electrochemical reduction on dual metal (Co–Ni)–nitrogen sites, J. Mater. Chem. A, 2024, 12(8), 4601–4609 RSC .
  40. X. An, Q. Tang, H. Lan, H. Liu, X. Yu, J. Qu, H. Lin and J. Ye, Facilitating Molecular Activation and Proton Feeding by Dual Active Sites on Polymeric Carbon Nitride for Efficient CO2Photoreduction, Angew. Chem., Int. Ed., 2022, 61(46), e202212706 CrossRef CAS PubMed .
  41. L. He, F. Yao, Y. Zhong, C. Tan, K. Hou, Z. Pi, S. Chen, X. Li and Q. Yang, Achieving high-performance electrocatalytic reduction of nitrate by N-rich carbon-encapsulated Ni-Cu bimetallic nanoparticles supported nickel foam electrode, J. Hazard. Mater., 2022, 436, 129253 CrossRef CAS PubMed .

Footnote

Electronic supplementary information (ESI) available. See DOI: https://doi.org/10.1039/d4cy00942h

This journal is © The Royal Society of Chemistry 2024