Previous Article in Journal
Preparation of g-C3N4/Co3O4/NS-CQDs Composite Materials and Their Application in the Detection of Hydrogen Peroxide and Glucose
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Enhanced Nitrate Reduction Performance of Cu-Doped Nanoporous Co2P Electrocatalyst

1
School of Materials Science and Engineering, Tianjin University, Tianjin 300350, China
2
State Key Laboratory of Precious Metal Functional Materials, Tianjin 300350, China
3
Tianjin Key Laboratory of Composite and Functional Materials, Tianjin 300350, China
*
Authors to whom correspondence should be addressed.
Nanomaterials 2025, 15(10), 753; https://doi.org/10.3390/nano15100753 (registering DOI)
Submission received: 21 April 2025 / Revised: 13 May 2025 / Accepted: 15 May 2025 / Published: 17 May 2025
(This article belongs to the Section Energy and Catalysis)

Abstract

:
Electrocatalytic nitrate reduction to ammonia (NO3RR) is a promising approach to recycle nitrogen from nitrate pollutants, yet it remains challenged by low Faradaic efficiency and insufficient NH3 production. Herein, Cu-doped nanoporous Co2P (np-Co2−xCuxP) is reported as electrocatalyst for NO3RR, achieving an ammonia yield rate of 30.6 mg h−1 cm−2 with a Faradaic efficiency of 93.4% at −0.3 V vs. RHE. In-situ spectroscopic analyses indicate that Cu incorporation modifies the surface electronic structure, resulting in the promotion of *H adsorption and *NO2 hydrogenation, thereby facilitating efficient ammonia generation.

1. Introduction

Ammonia is a widely used industrial raw product and considered an environmentally friendly fuel option for low-carbon energy technologies [1,2,3]. Electrochemical ammonia synthesis offers a promising alternative method for ambient NH3 production compared to the traditional energy-intensive Haber-Bosch process [4,5]. Electrochemical nitrate reduction offers a more practical pathway for ammonia synthesis, benefiting from nitrate’s high solubility and the relatively low N=O bond dissociation energy (204 kJ mol−1) [6,7,8]. Meanwhile, nitrate, commonly found in wastewater due to human activities, presents an appealing opportunity for converting pollutants into valuable ammonia from an environmental standpoint [9,10,11,12]. This highlights the importance of developing efficient electrocatalysts that can facilitate the multi-electron transfer process and enhance NH3 selectivity under ambient conditions. However, the NO3 conversion process is generally regarded as the rate-determining step, which has a low probability of occurrence due to ubiquitous energy-scaling relations [13,14]. Consequently, high overpotentials are required to overcome sluggish kinetics, which may trigger the competing hydrogen evolution reaction (HER), which undesirably consumes electrons and reduces nitrate reduction efficiency, thereby, reducing NH3 selectivity. Therefore, the development of catalysts that accelerate nitrate activation while suppressing HER at low overpotentials is critical for improving selectivity and energy efficiency.
Recent studies have witnessed breakthroughs in transition metal-based NO3RR catalysts, along with strategies aimed at enhancing activity and selectivity [15,16]. Among these, cobalt phosphides are highly effective for generating active hydrogen species and demonstrate exceptional catalytic performance towards NO3RR [17,18,19]. The cobalt centers with partial positive charges can effectively adsorb nitrate anions, while phosphorus centers with partial negative charges can serve as proton-acceptor sites. Tailoring the local electronic structure of cobalt phosphides by doping with other metals such as copper is a promising approach to optimize the adsorption strength of reaction intermediates and modulate hydrogen adsorption behavior. Unlike heterostructure engineering, metal doping can maintain the phase integrity of the host material while enabling fine-tuning of its surface reactivity. For instance, Cu-doped Fe3O4 catalysts have demonstrated nearly 97% Faradaic efficiency and high ammonia yield in NO3RR, attributed to the favorable electronic modulation introduced by Cu doping [20]. Similarly, Co-doped Fe@Fe2O3 catalysts derived from metal–organic frameworks exhibit enhanced NO3RR performance, where Co doping adjusts the Fe d-orbitals, improving intermediate adsorption and suppressing competing hydrogen evolution reactions [21]. In the context of cobalt phosphide systems, doping with transition metals like Fe, Ni has been explored to modulate electronic properties and catalytic activity. Studies indicate that such doping can influence the crystal structure, morphology, and electronic characteristics of Co2P-based catalysts, thereby affecting their performance in reactions [22,23]. However, conventional phosphide synthesis methods that often require toxic reagents or high-temperature treatments [24]. In addition, studies on Cu-doped Co2P for nitrate reduction remain limited, the appropriate modification of electronic structure by Cu doping for the metal phosphides deserves further investigation.
Herein, we report green and facile approach to prepare Cu-doped nanoporous Co2P (np-Co2−xCuxP) via a simple dealloying method for efficient electrochemical nitrate reduction. The catalyst achieves a notable ammonia production rate of 30.6 mg h−1 cm−2 and maintains a high Faradaic efficiency of 93.4% at −0.3 V versus RHE. Notably, it delivers a high current density exceeding 600 mA cm−2 at −0.4 V versus RHE, indicating excellent catalytic kinetics and conductivity. X-ray photoelectron spectroscopy (XPS) indicates that Cu incorporation modifies the electronic environments of Co and P atoms, which in turn facilitates the adsorption and transformation of crucial reaction intermediates. In-situ attenuated total reflection surface-enhanced infrared absorption spectroscopy (ATR-SEIRAS) results indicate a reduced onset potential for intermediate formation, which facilitates *H adsorption and *NO2 hydrogenation. These results demonstrate that the np-Co2−xCuxP catalyst, prepared through a facile and scalable method, offers excellent activity and selectivity toward ammonia production. By tuning the Cu content, the electronic structure and catalytic behavior of cobalt phosphide can be effectively modulated, showing great promise for practical nitrate-to-ammonia conversion applications.

2. Materials and Methods

2.1. Materials Fabrication

The precursor alloys were synthesized using a vacuum high-frequency induction melting furnace. Co, Cu, and Co2P were weighed according to atomic ratios of Co:Cu:P = 80:0:20, 75:5:20, 70:10:20, and 65:15:20, respectively, and then melted to form the desired alloys (Figure S1). The corresponding mass of each component is listed in Table S1. According to the actual Cu ratios, the four samples with Co:Cu:P = 80:0:20, 75:5:20, 70:10:20, and 65:15:20 were named as np-Co2P, np-Co1.94Cu0.06P, np-Co1.34Cu0.66P, and np-Co1.04Cu0.96P, respectively. The resulting melt was then remelted in a quartz cone tube with one end opening and quickly ejected onto a copper wheel rotating at 3500 rpm. This rapid cooling process produced the precursor alloy ribbons. The dealloying process was performed electrochemically in 1 M HCl at −0.05 V, employing a standard three-electrode setup [25]. Following dealloying, the resulting ribbons were thoroughly washed in deionized water and ethanol.

2.2. Materials Characterization

The phase composition was analyzed using X-ray diffraction (XRD model: Bruker D8 GmbH, Karlsruhe, Germany) with Cu Kα radiation. Morphological, microstructural, and compositional characterizations were carried out using a transmission electron microscope (TEM: JEOL JEM-2100F Akishima, Tokyo, Japan) coupled with an energy-dispersive X-ray spectrometer (EDS) and a scanning electron microscope (SEM: S-4800 Hitachi Tokyo, Japan). The elemental valence states were determined by X-ray photoelectron spectroscopy (XPS; Axis Supra Kratos Analytical Ltd., Manchester, UK).

2.3. Electrochemical Measurements

The electrocatalytic reactions were performed in an H-type electrochemical cell, with the anode and cathode compartments separated by a Nafion membrane. Before assembly, the Nafion membrane was handled with a 5 wt.% hydrogen peroxide solution at 80 °C for 60 min, followed by rinsing with deionized water for 30 min. It was then boiled in a 5 wt.% sulfuric acid solution at 80 °C for 60 min and washed again in deionized water for 30 min.
In the cathode compartment, the np-Co2P or np-Co2−xCuxP catalyst was applied as the working electrode. A platinum mesh served as the counter electrode in the anode compartment, and an Ag/AgCl electrode was used as the reference electrode. All electrochemical tests were conducted at room temperature. The potentials were converted to the reversible hydrogen electrode (RHE) using the following Equation (1):
E R H E = E A g / A g C l + 0.0591 × p H + 0.198

2.4. Determination of NH3

Ammonia (NH3) levels in the electrolyte were determined by the indophenol blue method. Extract the reaction solution from the reaction chamber and dilute it to 2 mL at a specific ratio. Then, 1 M NaOH solution with salicylic acid and sodium citrate, 0.05 M NaClO and C5FeN6Na2O (1 wt.%) was added to the diluted 2 mL of electrolyte. After two hours, the absorbance at 655 nm was measured using a UV fluorescence photometer. For comparison, a series of standard absorbance curves at different concentrations were prepared using NH4Cl. The ammonia production rate and Faraday efficiency were determined using the following equations:
The NH3 yield was calculated using Equation (2):
Y i e l d   NH 3 = C NH 3 V t S
where CNH3 refers to the NH3 concentration, which was measured using a UV-Vis spectrophotometer (UV-2700 Shimadzu Corporation, Kyoto, Japan). and calculated based on the standard curve. V denotes the electrolyte volume, t corresponds to the electrolysis duration, and S represents the surface area of the electrode.
Faradaic efficiency (FE) was calculated using Equation (3):
FE = 8 F C NH 3 V Q
where 8 represents the total electrons transferred during the reduction of nitrate to ammonia, F refers to the Faraday constant (96,485 C/mol), CNH3 denotes the concentration of NH3, V indicates the total volume of electrolyte used in the reaction, and Q represents the actual charge used during the test. The corresponding ammonia standard curves for different concentrations are shown in Figure S2.

2.5. In-Situ ATR-SEIRAS Measurements

In-situ ATR-SEIRAS was conducted using a Thermo Scientific IS50 spectrometer (Thermo Fisher Scientific Inc., Waltham, MA, USA). The standard three-electrode system was employed, with a platinum mesh as the counter electrode and an Ag/AgCl electrode as the reference electrode. The working electrode was prepared by sonicating the catalyst to form an ink, which was drop-cast onto a gold-sputtered silicon column. A Polaris long-life air-cooled mid/far-infrared source was used as the infrared light source. Infrared spectra were collected at room temperature and atmospheric pressure, with a constant applied potential. The spectra were acquired over a collection time of 90 s.

3. Results and Discussion

The np-Co2−xCuxP catalyst was prepared by selectively etching the Co phase from the Co-Cu-P precursor alloy (Figure S1). The Co-Cu-P precursor alloys were prepared by melt-spinning method. The precursor alloys with designed compositions were melted into liquid state in quartz tubes and then ejected on a rotating copper roller to realize rapid cooling. During the solidification process, the Cu atoms trends to alloy with metallic Co and dope into the Co2P in the nucleation and growth process, forming the metallic Co phase, Co0.52Cu0.48 alloy phase and Co2P phase in the precursor alloy. The phase compositions of these precursor alloys were examined by X-ray diffraction (XRD). As shown in Figure S3, the diffraction peaks located at 44.4°, 51.5°, and 75.9° are attributed to metallic Co, while peaks at 41.1°, 43.2°, and 52.3° correspond to the Co2P phase [25,26,27]. With increasing Cu content, additional peaks at 50.6° and 74.3° emerge, which can be assigned to the Co0.52Cu0.48 alloy phase [28,29]. This indicates that Cu atoms gradually occupy the lattice sites of Co, forming a solid solution alloy phase when the Cu content becomes sufficiently high. Compared with the binary Co80P20 precursor, the Cu-doped samples show no significant phase difference at low Cu contents; however, a noticeable increase in the intensity of Co0.52Cu0.48-related peaks is observed at higher Cu loadings, suggesting enhanced alloying behavior for Cu. After electrochemical dealloying, the Co and Co0.52Cu0.48 alloy phases are selectively dissolved, forming a porous Co2P structure (Figure 1). The complete disappearance of metallic Co and Co0.52Cu0.48 peaks confirms their full removal during dealloying. This process is critical for generating high surface area and accessible active sites [25]. Figure S4 shows the enlarged XRD patterns focusing on the (201) plane of Co2P for all samples. A clear shift of the diffraction peak towards lower angles with increasing Cu content is observed, indicating a lattice expansion due to the incorporation of Cu atoms, which possess a larger atomic radius than Co. This confirms the successful substitutional doping of Cu into the Co2P lattice, leading to the modulation of its crystal structure. Therefore, Co not only facilitates the formation of a dealloyable structure but also indirectly influences the final composition and morphology of the catalyst.
To further investigate the influence of Cu content on the morphology and composition of np-Co2−xCuxP, SEM and EDX analyses were performed. EDX results (Table S2) confirm that the Cu content can be effectively tuned from 1.8% to 24.05% and the corresponding ICP analysis (Table S3) further verifies the actual elemental composition is consistent with the designed values. As shown in Figure 2, all samples exhibit interconnected nanoporous structures, but with distinct features depending on Cu content. The np-Co2P sample displays uniform ~200 nm pores, resulting from selective removal of the Co phase and retention of the Co2P framework. With increased Cu content, smaller pores (~50 nm) emerge alongside the larger ones due to the involvement of Co0.52Cu0.48 alloy in the dealloying process, which enhances surface area and active site density. However, further increasing Cu content leads to coarsened porous structures (~300 nm), likely due to excessive corrosion of Co–Cu phases, resulting in reduced surface area.
To gain deeper insights into the morphology, phase composition, and elemental distribution of the np-Co1.34Cu0.66P, transmission electron microscopy (TEM) and energy-dispersive X-ray spectroscopy (EDS) analyses were conducted. As shown in Figure 3a, the TEM image reveals a hierarchical porous structure, consistent with the SEM observations. The high-resolution TEM (HRTEM) image and FFT patterns in Figure 3b displays well-defined lattice fringes with interplanar spacings of 2.203 Å and 1.856 Å, corresponding to the (201) and (031) planes of Co2P, respectively. The FFT pattern shows two pairs of centrosymmetric diffraction spots, corresponding to the (201) and (031) planes of Co2P, which further confirms the good crystallinity and specific orientation of the sample [30,31]. The measured angle between these two planes is approximately 81°, which aligns well with theoretical values, further confirming the Co2P phase of the material. Elemental mapping via EDS (Figure 3c) shows a homogeneous distribution of Co, P, and Cu throughout the porous framework, indicating that Cu atoms are uniformly doped into the Co2P lattice.
To gain insight into the electronic structure modulation induced by Cu incorporation, X-ray photoelectron spectroscopy (XPS) was conducted on the np-Co2−xCuxP samples, as displayed in Figure 3. The Co 2p spectra (Figure 4a,d,g) show two main peaks around 778.1 eV and 793.1 eV, corresponding to Co 2p3/2 and Co 2p1/2 of Co0. Additional peaks at 781.1 eV and 797.5 eV are attributed to Co2+ and Co3+ species, likely due to surface oxidation during measurement [18,32]. In the P 2p spectra (Figure 4b,e,h), the peaks at 129.3 eV and 130.2 eV are assigned to the 2p3/2 and 2p1/2 of metal phosphides, while the peak at 133.2 eV is associated with oxidized phosphorus [33,34]. For Cu (Figure 4c,f,i), the characteristic Cu-P peak appears at 932.2 eV (Cu 2p3/2) and 952.1 eV (Cu 2p1/2), confirming the successful incorporation of Cu. As the Cu content increases, a positive shift of Co 2p and a negative shift of P 2p binding energies are observed, indicating electron redistribution due to Cu doping-where Cu lowers the electron density around Co and promotes electron transfer from Co to P. Notably, the Cu 2p3/2 peak of np-Co1.34Cu0.66P appears at the highest binding energy (932.5 eV), suggesting the strongest interaction between Cu and the Co2P matrix, which is beneficial for enhancing electron transfer during electrocatalytic nitrate reduction [35,36].
To further evaluate the electrocatalytic nitrate reduction performance of np-Co2−xCuxP catalysts, a series of electrochemical measurements were conducted. As shown in Figure 5a,b, np-Co1.34Cu0.66P exhibits the highest NO3RR current density among all tested samples, especially at low overpotentials, significantly outperforming np-Co2P. This indicates that appropriate Cu doping effectively enhances the catalytic activity of Co2P. The catalytic activity was further evaluated by the potentiostatic method. The corresponding i–t curve is shown in Figure S5. Figure 5c,d display the NH3 yield and Faradaic efficiency (FE) of np-Co1.34Cu0.66P under various potentials. As the potential becomes more negative, NH3 yield increases, while FE peaks at −0.3 V with a maximum of 93.7%, which can be attributed to the competitive hydrogen evolution at high overpotential. Figure 5e demonstrates the NH3 yield and FE of different samples at −0.3 V. The np-Co1.34Cu0.66P catalyst achieves the highest NH3 yield and FE among the Co2P-based catalysts. To assess the long-term durability, a 10-h electrolysis test was conducted at −0.3 V, during which both NH3 yield and FE remained stable, indicating excellent catalytic stability (Figure S6). In addition, no obvious difference can be detected before and after stability test, confirming the high stability of the nanoporous Cu doped Co2P. To verify that the detected ammonia originates from the electrochemical reaction, control experiments without applied potential and without nitrate were conducted. In both cases, no significant ammonia was detected, as shown in Figure S7. To further confirm the suppression of the hydrogen evolution reaction by Cu incorporation, gas chromatography (GC) measurements were conducted. As shown in Figure S8, the np-Co1.34Cu0.66P exhibited a H2 Faradaic efficiency of 2.59% at −0.3 V vs. RHE, which is lower than that of np-Co2P (6.08%), indicating the suppression of HER by Cu doping. This trend can be attributed to the fact that pure Co2P has limited water dissociation ability, resulting in insufficient *H for NO3- hydrogenation at low potentials. Moderate Cu doping promotes H2O dissociation and provides more reactive hydrogen, improving NH3 selectivity and FE. However, excessive Cu content leads to weakened NO3 and intermediate adsorption. A comparison of np-Co1.34Cu0.66P with other state-of-the-art catalysts is presented in Figure 5f and Table S4, indicating that the np-Co1.34Cu0.66P exhibits superior catalytic performance compared with other NO3RR electrocatalysts [21,37,38,39,40,41,42,43,44,45,46].
To further explore the enhanced mechanism caused by Cu doping, in situ ATR-SEIRAS was conducted to monitor reaction intermediates. As shown in Figure 6, the spectra of np-Co2P and np-Co1.34Cu0.66P display characteristic peaks at 1209, 1438, and 1632 cm−1, corresponding to *NO2, *NHx, and H2O species, respectively [11,47,48,49]. For np-Co2P, the *NHx signal emerges at −0.2 V and strengthens as the potential becomes more negative, indicating progressive reduction of nitrate. The presence of *NO2 at 0.2 V implies that Co2P enables nitrate activation at low overpotential, but insufficient hydrogen availability results in incomplete reduction and byproduct formation. In contrast, the *NHx signal for np-Co1.34Cu0.66P appears at 0 V, suggesting improved conversion of *NO2 to *NHx due to Cu incorporation [50,51]. Moreover, a blue shift in the O-H stretching vibration (from 3300 to 3400 cm−1) indicates enhanced water dissociation on the Cu-doped surface, leading to more active hydrogen species that facilitate nitrate hydrogenation and promote catalytic activity [52,53,54].

4. Conclusions

In summary, nanoporous Cu-doped Co2P (np-Co2−xCuxP) electrocatalysts were successfully fabricated via chemical dealloying method and used as the electrocatalyst for nitrate reduction reaction to ammonia. Structural characterizations demonstrate that moderate Cu incorporation not only retained the nanoporous framework but also modulated the electronic structure of Co2P. At −0.3 V vs. RHE, np-Co1.34Cu0.66P delivers an impressive ammonia production rate of 30.6 mg h−1 cm−2, along with a notable Faradaic efficiency of 93.4%, outperforming both pristine Co2P and other Cu-doped variants. In-situ ATR-SEIRAS experiments reveal that Cu doping facilitates water dissociation and provides more active hydrogen species for efficient nitrate hydrogenation, thereby promoting the transformation of *NO2 to *NHx intermediates. This study highlights the crucial role of electronic structure engineering in optimizing NO3RR performance and offers a promising strategy for designing efficient non-noble metal catalysts for sustainable ammonia synthesis.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/nano15100753/s1.

Author Contributions

Y.H.: Conceptualization, Methodology, Data curation, Writing—original draft. Writing—review & editing. X.Z.: Data curation, Writing—original draft. Writing—review & editing Y.L.: Conceptualization, Writing—original draft. H.J.: Supervision, Writing-review & editing. Z.L.: Methodology. S.W.: Formal analysis. Z.G.: Conceptualization, Writing—review & editing. S.Z.: Supervision, Project administration, Writing—review & editing, Funding acquisition. Z.C.: Supervision, Writing—review & editing, Funding acquisition. W.X.: Conceptualization, Methodology, Project administration, Supervision, Writing—review & editing, Funding acquisition. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by National Natural Science Foundation of China (No. 52371161, No. 52371232, No. 52271152).

Data Availability Statement

The original contributions presented in this study are included in the article/Supplementary Materials. Further inquiries can be directed to the corresponding author.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Gunbatar, S.A.; Kiran, B.E.; Boz, Y.; Oztay, E.S. A systematic review of green and sustainable chemistry training research with pedagogical content knowledge framework: Current trends and future directions. Chem. Educ. Res. Pract. 2025, 26, 34–52. [Google Scholar] [CrossRef]
  2. Cheng, Q.; Muhammad, A.; Kaario, O.; Ahmad, Z.; Martti, L. Ammonia as a sustainable fuel: Review and novel strategies. Renew. Sustain. Energy Rev. 2025, 207, 114995. [Google Scholar] [CrossRef]
  3. Moosazadeh, M.; Mansourimarand, A.; Ajori, S.; Taghikhani, V.; Yoo, C. Waste-to-Ammonia: A sustainable pathway for energy transition. Renew. Sustain. Energy Rev. 2025, 208, 115012. [Google Scholar] [CrossRef]
  4. Li, W.; Liu, W.; Lin, Y.; Liu, C.; Wang, X.; Xu, J. Optimal dispatching of integrated energy system with hydrogen-to-ammonia and ammonia-mixed/oxygen-enriched thermal power. Energy 2025, 316, 134514. [Google Scholar] [CrossRef]
  5. Fang, Y.; Li, M.; Gao, Y.; Wen, Y.; Shan, B. Static Organic p-n Junctions in Photoelectrodes for Solar Ammonia Production with 86 % Internal Quantum Efficiency. Angew. Chem. Int. Ed. Engl. 2025, 64, e202415729. [Google Scholar] [CrossRef]
  6. Ahsan, A.; Hussain, M.; Rashad, M.A.; Akhter, P.; Jamil, F.; Cho, K.; Park, Y.-K. Critical review on electrocatalytic reduction of nitrogen and nitrate to ammonia. J. Ind. Eng. Chem. 2025, 02, 29. [Google Scholar] [CrossRef]
  7. Jiang, H.; Li, T.; Gao, Y.; Fan, J.; Gan, D.; Yuan, S.; Hong, L.; Feng, Y.; Sun, J.; Song, Q.; et al. Sustainable ammonia synthesis: Opportunities for electrocatalytic nitrate reduction. J. Energy Chem. 2025, 105, 630–638. [Google Scholar] [CrossRef]
  8. Kang, B.; Xu, B.; Chen, Z.; Li, F.; Wang, Y. Promoting active hydrogen supply for kinetically matched tandem electrocatalytic nitrate reduction to ammonia. Appl. Catal. B-Environ. 2025, 360, 124528. [Google Scholar] [CrossRef]
  9. Bhatti, U.A.; Bhatti, M.A.; Tang, H.; Syam, M.S.; Awwad, E.M.; Sharaf, M.; Ghadi, Y.Y. Global production patterns: Understanding the relationship between greenhouse gas emissions, agriculture greening and climate variability. Environ. Res. 2024, 245, 118049. [Google Scholar] [CrossRef]
  10. Zhang, R.; Li, C.; Cui, H.; Wang, Y.; Zhang, S.; Li, P.; Hou, Y.; Guo, Y.; Liang, G.; Huang, Z.; et al. Electrochemical nitrate reduction in acid enables high-efficiency ammonia synthesis and high-voltage pollutes-based fuel cells. Nat. Commun. 2023, 14, 8036. [Google Scholar] [CrossRef]
  11. Wang, M.; Khan, M.A.; Mohsin, I.; Wicks, J.; Ip, A.H.; Sumon, K.Z.; Dinh, C.-T.; Sargent, E.H.; Gates, I.D.; Kibria, M.G. Can sustainable ammonia synthesis pathways compete with fossil-fuel based Haber–Bosch processes? Energy Environ. Sci. 2021, 14, 2535–2548. [Google Scholar] [CrossRef]
  12. Zhang, H.; Fang, K.; Yang, J.; Chen, H.; Ning, J.; Wang, H.; Hu, Y. Strategies and applications of electrocatalytic nitrate reduction towards ammonia. Coord. Chem. Rev. 2024, 506, 215723. [Google Scholar] [CrossRef]
  13. Han, S.; Li, H.; Li, T.; Chen, F.; Yang, R.; Yu, Y.; Zhang, B. Ultralow overpotential nitrate reduction to ammonia via a three-step relay mechanism. Nat. Catal. 2023, 6, 402–414. [Google Scholar] [CrossRef]
  14. Xu, H.; Ma, Y.; Chen, J.; Zhang, W.X.; Yang, J. Electrocatalytic reduction of nitrate—a step towards a sustainable nitrogen cycle. Chem. Soc. Rev. 2022, 51, 2710–2758. [Google Scholar] [CrossRef] [PubMed]
  15. Wang, Y.; Liu, C.; Zhang, B.; Yu, Y. Self-template synthesis of hierarchically structured Co3O4@NiO bifunctional electrodes for selective nitrate reduction and tetrahydroisoquinolines semi-dehydrogenation. Sci. China Mater. 2020, 63, 2530–2538. [Google Scholar] [CrossRef]
  16. Hu, Q.; Yang, K.; Peng, O.; Li, M.; Ma, L.; Huang, S.; Du, Y.; Xu, Z.X.; Wang, Q.; Chen, Z.; et al. Ammonia Electrosynthesis from Nitrate Using a Ruthenium-Copper Cocatalyst System: A Full Concentration Range Study. J. Am. Chem. Soc. 2024, 146, 668–676. [Google Scholar] [CrossRef] [PubMed]
  17. Jia, Y.; Ji, Y.G.; Xue, Q.; Li, F.M.; Zhao, G.T.; Jin, P.J.; Li, S.N.; Chen, Y. Efficient Nitrate-to-Ammonia Electroreduction at Cobalt Phosphide Nanoshuttles. ACS Appl. Mater. Interfaces 2021, 13, 45521–45527. [Google Scholar] [CrossRef]
  18. Ye, S.; Chen, Z.; Zhang, G.; Chen, W.; Peng, C.; Yang, X.; Zheng, L.; Li, Y.; Ren, X.; Cao, H.; et al. Elucidating the activity, mechanism and application of selective electrosynthesis of ammonia from nitrate on cobalt phosphide. Energy Environ. Sci. 2022, 15, 760–770. [Google Scholar] [CrossRef]
  19. Qi, R.; Wang, Z.; Zhong, M.; Wang, C.; Bai, F.; Lu, X. Synergistic Integration of Amorphous Cobalt Phosphide with a Conductive Channel for Highly Efficient Electrocatalytic Nitrate Reduction to Ammonia. Small 2024, 20, 2308311. [Google Scholar] [CrossRef]
  20. Cheng, X.; Zhao, H.; Liu, P.; Han, R.; Wang, Y.-Q. A Cu/Fe3O4@CN tandem catalyst for efficient ammonia electrosynthesis from nitrate reduction. J. Colloid Interface Sci. 2025, 682, 703–714. [Google Scholar] [CrossRef]
  21. Zhang, S.; Li, M.; Li, J.; Song, Q.; Liu, X. High-ammonia selective metal organic framework derived Co-doped Fe/Fe2O3 catalysts for electrochemical nitrate reduction. Proc. Natl. Acad. Sci. USA 2022, 119, e2115504119. [Google Scholar] [CrossRef]
  22. Zhang, L.; Dong, W.; Zhang, Y.; Song, X.; Jiang, H. Precisely anchoring Ni-doped cobalt phosphide nanoparticles on phosphatized carbon nitride for efficient photocatalytic water splitting. Chem. Eng. J. 2023, 472, 144898. [Google Scholar] [CrossRef]
  23. Li, F.; Bu, Y.; Lv, Z.; Mahmood, J.; Han, G.-F.; Ahmad, I.; Kim, G.; Zhong, Q.; Baek, J.-B. Porous Cobalt Phosphide Polyhedrons with Iron Doping as an Efficient Bifunctional Electrocatalyst. Small 2017, 13, 1701167. [Google Scholar] [CrossRef] [PubMed]
  24. Li, Y.; Lu, Z.; Zheng, L.; Yan, X.; Xie, J.; Yu, Z.; Zhang, S.; Jiang, F.; Chen, H. The synergistic catalysis effect on electrochemical nitrate reduction at the dual-function active sites of the heterostructure. Energy Environ. Sci. 2024, 17, 4582–4593. [Google Scholar] [CrossRef]
  25. Tan, Y.; Wang, H.; Liu, P.; Cheng, C.; Zhu, F.; Hirata, A.; Chen, M. 3D Nanoporous Metal Phosphides toward High-Efficiency Electrochemical Hydrogen Production. Adv. Mater. 2016, 28, 2951–2955. [Google Scholar] [CrossRef]
  26. Callejas, J.F.; Read, C.G.; Popczun, E.J.; McEnaney, J.M.; Schaak, R.E. Nanostructured Co2P Electrocatalyst for the Hydrogen Evolution Reaction and Direct Comparison with Morphologically Equivalent CoP. Chem. Mater. 2015, 27, 3769–3774. [Google Scholar] [CrossRef]
  27. Xu, W.; Fan, G.; Zhu, S.; Liang, Y.; Cui, Z.; Li, Z.; Jiang, H.; Wu, S.; Cheng, F. Electronic Structure Modulation of Nanoporous Cobalt Phosphide by Carbon Doping for Alkaline Hydrogen Evolution Reaction. Adv. Funct. Mater. 2021, 31, 2107333. [Google Scholar] [CrossRef]
  28. Guo, R.; Zhang, Y.; Zhang, X.; Yang, H.; Hu, T. Synthesis of Co0·52Cu0·48/Cu@S–C arrays for the enhanced performance of hydrazine-assisted hydrogen production. Int. J. Hydrogen Energy 2024, 68, 472–480. [Google Scholar] [CrossRef]
  29. Sang, W.; Wang, C.; Zhang, X.; Yu, X.; Yu, C.; Zhao, J.; Wang, X.; Yang, X.; Li, L. Dendritic Co0.52Cu0.48 and Ni0.19Cu0.81 alloys as hydrogen generation catalysts via hydrolysis of ammonia borane. Int. J. Hydrogen Energy 2017, 42, 30691–30703. [Google Scholar] [CrossRef]
  30. Liu, G.; Wang, M.; Xu, Y.; Wang, X.; Li, X.; Liu, J.; Cui, X.; Jiang, L. Porous CoP/Co2P heterostructure for efficient hydrogen evolution and application in magnesium/seawater battery. J. Power Sources 2021, 486, 229351. [Google Scholar] [CrossRef]
  31. Chen, L.; Song, X.-L.; Ren, J.-T.; Yuan, Z.-Y. Precisely modifying Co2P/black TiO2 S-scheme heterojunction by in situ formed P and C dopants for enhanced photocatalytic H2 production. Appl. Catal. B Environ. 2022, 315, 121546. [Google Scholar] [CrossRef]
  32. Liu, W.; Gao, H.; Zhang, Z.; Zheng, Y.; Wu, Y.; Fu, X.; Su, J.; Gao, Y. CoP/Cu3P heterostructured nanoplates for high-rate supercapacitor electrodes. Chem. Eng. J. 2022, 437, 135352. [Google Scholar] [CrossRef]
  33. Gao, Y.; Wang, K.; Xu, C.; Fang, H.; Yu, H.; Zhang, H.; Li, S.; Li, C.; Huang, F. Enhanced electrocatalytic nitrate reduction through phosphorus-vacancy-mediated kinetics in heterogeneous bimetallic phosphide hollow nanotube array. Appl. Catal. B-Environ. 2023, 330, 122627. [Google Scholar] [CrossRef]
  34. Li, J.; Li, M.; An, N.; Zhang, S.; Song, Q.; Yang, Y.; Li, J.; Liu, X. Boosted ammonium production by single cobalt atom catalysts with high Faradic efficiencies. Proc. Natl. Acad. Sci. USA 2022, 119, e2123450119. [Google Scholar] [CrossRef] [PubMed]
  35. Liu, W.; Zhang, Z.; Zhang, Y.; Zheng, Y.; Liu, N.; Su, J.; Gao, Y. Interior and Exterior Decoration of Transition Metal Oxide Through Cu0/Cu+ Co-Doping Strategy for High-Performance Supercapacitor. Nano-Micro Lett. 2021, 13, 61. [Google Scholar] [CrossRef] [PubMed]
  36. Lee, Y.S.; Lim, K.Y.; Chung, Y.D.; Whang, C.N.; Jeon, Y. XPS core-level shifts and XANES studies of Cu–Pt and Co–Pt alloys. Surf. Interface Anal. 2000, 30, 475–478. [Google Scholar] [CrossRef]
  37. Wang, Y.; Zhou, W.; Jia, R.; Yu, Y.; Zhang, B. Unveiling the activity origin of a copper-based electrocatalyst for selective nitrate reduction to ammonia. Angew. Chem. Int. Ed. Engl. 2020, 59, 5350–5354. [Google Scholar] [CrossRef]
  38. Yang, M.; Li, B.; Li, S.; Dong, Q.; Huang, Z.; Zheng, S.; Fang, Y.; Zhou, G.; Chen, X.; Zhu, X.; et al. Highly Selective Electrochemical Nitrate to Ammonia Conversion by Dispersed Ru in a Multielement Alloy Catalyst. Nano Lett. 2023, 23, 7733–7742. [Google Scholar] [CrossRef]
  39. Li, Y.; Ma, J.; Wu, Z.; Wang, Z. Direct Electron Transfer Coordinated by Oxygen Vacancies Boosts Selective Nitrate Reduction to N2 on a Co-CuOx Electroactive Filter. Environ. Sci. Technol. 2022, 56, 8673–8681. [Google Scholar] [CrossRef]
  40. Yu, Y.; Wang, C.; Yu, Y.; Wang, Y.; Zhang, B. Promoting selective electroreduction of nitrates to ammonia over electron-deficient Co modulated by rectifying Schottky contacts. Sci. China Chem. 2020, 63, 1469–1476. [Google Scholar] [CrossRef]
  41. Jiang, M.; Su, J.; Song, X.; Zhang, P.; Zhu, M.; Qin, L.; Tie, Z.; Zuo, J.-L.; Jin, Z. Interfacial Reduction Nucleation of Noble Metal Nanodots on Redox-Active Metal–Organic Frameworks for High-Efficiency Electrocatalytic Conversion of Nitrate to Ammonia. Nano Lett. 2022, 22, 2529–2537. [Google Scholar] [CrossRef] [PubMed]
  42. Jia, R.; Wang, Y.; Wang, C.; Ling, Y.; Yu, Y.; Zhang, B. Boosting Selective Nitrate Electroreduction to Ammonium by Constructing Oxygen Vacancies in TiO2. ACS Catal. 2020, 10, 3533–3540. [Google Scholar] [CrossRef]
  43. Yang, J.; Qi, H.; Li, A.; Liu, X.; Yang, X.; Zhang, S.; Zhao, Q.; Jiang, Q.; Su, Y.; Zhang, L. Potential-driven restructuring of Cu single atoms to nanoparticles for boosting the electrochemical reduction of nitrate to ammonia. J. Am. Chem. Soc. 2022, 144, 12062–12071. [Google Scholar] [CrossRef]
  44. Lv, Y.; Ke, S.W.; Gu, Y.; Tian, B.; Tang, L.; Ran, P.; Zhao, Y.; Ma, J.; Zuo, J.L.; Ding, M. Highly Efficient Electrochemical Nitrate Reduction to Ammonia in Strong Acid Conditions with Fe2M-Trinuclear-Cluster Metal-Organic Frameworks. Angew. Chem. Int. Ed. Engl. 2023, 62, e202305246. [Google Scholar] [CrossRef]
  45. Bu, Y.; Wang, C.; Zhang, W.; Yang, X.; Ding, J.; Gao, G. Electrical pulse-driven periodic self-repair of Cu-Ni tandem catalyst for efficient ammonia synthesis from nitrate. Angew. Chem. Int. Ed. Engl. 2023, 62, e202217337. [Google Scholar] [CrossRef] [PubMed]
  46. Song, Z.; Liu, Y.; Zhong, Y.; Guo, Q.; Zeng, J.; Geng, Z. Efficient electroreduction of nitrate into ammonia at ultralow concentrations via an enrichment effect. Adv. Mater. 2022, 34, 2204306. [Google Scholar] [CrossRef] [PubMed]
  47. Zhao, C.X.; Liu, J.N.; Li, B.Q.; Ren, D.; Chen, X.; Yu, J.; Zhang, Q. Multiscale Construction of Bifunctional Electrocatalysts for Long-Lifespan Rechargeable Zinc-Air Batteries. Adv. Funct. Mater. 2020, 30, e202003619. [Google Scholar] [CrossRef]
  48. Mao, X.; Bai, X.; Wu, G.; Qin, Q.; O’Mullane, A.P.; Jiao, Y.; Du, A. Electrochemical Reduction of N2 to Ammonia Promoted by Hydrated Cation Ions: Mechanistic Insights from a Combined Computational and Experimental Study. J. Am. Chem. Soc. 2024, 146, 18743–18752. [Google Scholar] [CrossRef]
  49. Yu, S.; Yamauchi, H.; Wang, S.; Aggarwal, A.; Kim, J.; Gordiz, K.; Huang, B.; Xu, H.; Zheng, D.J.; Wang, X.; et al. CO2-to-methanol electroconversion on a molecular cobalt catalyst facilitated by acidic cations. Nat. Catal. 2024, 7, 1000–1009. [Google Scholar] [CrossRef]
  50. Wang, Y.H.; Zheng, S.; Yang, W.M.; Zhou, R.Y.; He, Q.F.; Radjenovic, P.; Dong, J.C.; Li, S.; Zheng, J.; Yang, Z.L.; et al. In situ Raman spectroscopy reveals the structure and dissociation of interfacial water. Nature 2021, 600, 81–85. [Google Scholar] [CrossRef]
  51. Zhu, S.; Qin, X.; Yao, Y.; Shao, M. pH-Dependent Hydrogen and Water Binding Energies on Platinum Surfaces as Directly Probed through Surface-Enhanced Infrared Absorption Spectroscopy. J. Am. Chem. Soc. 2020, 142, 8748–8754. [Google Scholar] [CrossRef] [PubMed]
  52. Concha, B.M.; Chatenet, M.; Coutanceau, C.; Hahn, F. In situ infrared (FTIR) study of the borohydride oxidation reaction. Electrochem. Commun. 2009, 11, 223–226. [Google Scholar] [CrossRef]
  53. Fang, J.-Y.; Zheng, Q.-Z.; Lou, Y.-Y.; Zhao, K.-M.; Hu, S.-N.; Li, G.; Akdim, O.; Huang, X.-Y.; Sun, S.-G. Ampere-level current density ammonia electrochemical synthesis using CuCo nanosheets simulating nitrite reductase bifunctional nature. Nat. Commun. 2022, 13, 7899. [Google Scholar] [CrossRef]
  54. Zhang, J.; Quast, T.; Eid, B.; Chen, Y.-T.; Zerdoumi, R.; Dieckhöfer, S.; Junqueira, J.R.C.; Seisel, S.; Schuhmann, W. In-situ electrochemical reconstruction and modulation of adsorbed hydrogen coverage in cobalt/ruthenium-based catalyst boost electroreduction of nitrate to ammonia. Nat. Commun. 2024, 15, 8583. [Google Scholar] [CrossRef] [PubMed]
Figure 1. XRD pattern of dealloyed alloys with different Cu contents.
Figure 1. XRD pattern of dealloyed alloys with different Cu contents.
Nanomaterials 15 00753 g001
Figure 2. SEM image of (a) np-Co2P (b) np-Co1.94Cu0.06P (c) np-Co1.34Cu0.66P (d) np-Co104Cu0.96P alloy.
Figure 2. SEM image of (a) np-Co2P (b) np-Co1.94Cu0.06P (c) np-Co1.34Cu0.66P (d) np-Co104Cu0.96P alloy.
Nanomaterials 15 00753 g002
Figure 3. TEM image (a), HRTEM and FFT patterns for HRTEM images (b), and EDS elemental mapping (c) of dealloyed np-Co1.34Cu0.66P.
Figure 3. TEM image (a), HRTEM and FFT patterns for HRTEM images (b), and EDS elemental mapping (c) of dealloyed np-Co1.34Cu0.66P.
Nanomaterials 15 00753 g003
Figure 4. XPS spectra of (ac) Co 2p, P 2p, and Cu 2p for np-Co1.94Cu0.06P, (df) Co 2p, P 2p, and Cu 2p for np-Co1.34Cu0.66P, and (gi) Co 2p, P 2p, and Cu 2p for np-Co1.04Cu0.96P.
Figure 4. XPS spectra of (ac) Co 2p, P 2p, and Cu 2p for np-Co1.94Cu0.06P, (df) Co 2p, P 2p, and Cu 2p for np-Co1.34Cu0.66P, and (gi) Co 2p, P 2p, and Cu 2p for np-Co1.04Cu0.96P.
Nanomaterials 15 00753 g004
Figure 5. (a) LSV curves of np-Co2−xCuxP catalysts in 1 M KOH; (b) LSV curves in 1 M KOH + 0.1 M NaNO3; (c) NH3 yield and (d) Faradaic efficiency of np-Co1.34Cu0.66P at different applied potentials; (e) NH3 yield and Faradaic efficiency of np-Co2−xCuxP at −0.3 V; (f) Comparison of Faradaic efficiency with recently reported high-performance NO3RR catalysts.
Figure 5. (a) LSV curves of np-Co2−xCuxP catalysts in 1 M KOH; (b) LSV curves in 1 M KOH + 0.1 M NaNO3; (c) NH3 yield and (d) Faradaic efficiency of np-Co1.34Cu0.66P at different applied potentials; (e) NH3 yield and Faradaic efficiency of np-Co2−xCuxP at −0.3 V; (f) Comparison of Faradaic efficiency with recently reported high-performance NO3RR catalysts.
Nanomaterials 15 00753 g005
Figure 6. In-situ ATR-SEIRAS spectra of (a) np-Co2P and (b) np-Co1.34Cu0.66P.
Figure 6. In-situ ATR-SEIRAS spectra of (a) np-Co2P and (b) np-Co1.34Cu0.66P.
Nanomaterials 15 00753 g006
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

Huang, Y.; Zhang, X.; Liang, Y.; Jiang, H.; Wu, S.; Li, Z.; Cui, Z.; Zhu, S.; Gao, Z.; Xu, W. Enhanced Nitrate Reduction Performance of Cu-Doped Nanoporous Co2P Electrocatalyst. Nanomaterials 2025, 15, 753. https://doi.org/10.3390/nano15100753

AMA Style

Huang Y, Zhang X, Liang Y, Jiang H, Wu S, Li Z, Cui Z, Zhu S, Gao Z, Xu W. Enhanced Nitrate Reduction Performance of Cu-Doped Nanoporous Co2P Electrocatalyst. Nanomaterials. 2025; 15(10):753. https://doi.org/10.3390/nano15100753

Chicago/Turabian Style

Huang, Yunduo, Xiechen Zhang, Yanqin Liang, Hui Jiang, Shuilin Wu, Zhaoyang Li, Zhenduo Cui, Shengli Zhu, Zhonghui Gao, and Wence Xu. 2025. "Enhanced Nitrate Reduction Performance of Cu-Doped Nanoporous Co2P Electrocatalyst" Nanomaterials 15, no. 10: 753. https://doi.org/10.3390/nano15100753

APA Style

Huang, Y., Zhang, X., Liang, Y., Jiang, H., Wu, S., Li, Z., Cui, Z., Zhu, S., Gao, Z., & Xu, W. (2025). Enhanced Nitrate Reduction Performance of Cu-Doped Nanoporous Co2P Electrocatalyst. Nanomaterials, 15(10), 753. https://doi.org/10.3390/nano15100753

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Article metric data becomes available approximately 24 hours after publication online.
Back to TopTop