DOI:
10.1039/D5TA02053K
(Paper)
J. Mater. Chem. A, 2025,
13, 22383-22391
Electrodeposited CoP2 on CO2-laser-modified graphite felt: a robust electrocatalyst for nitrite reduction to ammonia†
Received
13th March 2025
, Accepted 17th April 2025
First published on 18th April 2025
Abstract
The conversion of nitrite-based pollutants to value-added ammonia (NH3) via sustainable electrocatalysis represents a remarkable advancement in waste management research. Herein, a two-step strategy was developed to synthesize well-dispersed cobalt phosphide (CoP2) on graphene oxide (GO)–graphite felt (GF), termed CoP2/GO–GF. The electrodeposited CoP2 exhibited exceptional performance in the electrocatalytic NO2− to NH3 reduction reaction (NO2RR), achieving a maximum NH3 yield rate of 10.6 mg h−1 cm−2 with a faradaic efficiency of 80% at −0.4 V vs. the reversible hydrogen electrode (RHE). The high efficiency of CoP2/GO–GF is attributed to its improved surface-active site density, enhanced electrochemical double-layer capacitance (3.37 mF cm−2), and optimized electron transfer resistance (13.31 Ω). Furthermore, a turnover frequency analysis of the NO2RR indicated the abundance of active sites, facilitating smooth charge tunneling from CoP2 to CO2 laser-developed GO on GF in CoP2/GO–GF. In situ FTIR analysis confirmed the sequential reduction pathway from NO2− to NH3, identifying NO as a key intermediate. Additionally, density functional theory (DFT) calculations revealed a moderate free energy barrier (0.26 eV) for the rate-limiting step, thus validating the thermodynamic feasibility of the reaction. Furthermore, durability tests demonstrated stable performance over 10 reuse cycles, confirming the efficiency and robustness of CoP2/GO–GF as an electrocatalyst in the NO2RR.
1. Introduction
Nitrogenous waste generated by modern agricultural practices, sewage discharge, and industrial processes has become a major concern for groundwater quality. Reports indicate that elevated levels of nitrate/nitrite (NO3−/NO2−) contribute to various health-related issues, including reproductive problems and methemoglobinemia, while also damaging aquatic life by creating oxygen (O)-deficient “dead zones” in aquatic ecosystems.1 To facilitate NO2− removal from water, methods such as catalytic reduction,2–4 adsorption,5 membrane filtration,6 and ion exchange7 are increasingly being used. Compared with conventional methods, NO2−-to-ammonia (NH3) conversion is a more efficient and sustainable approach for NO2− removal.8 Thus, the electrocatalytic conversion of NO2− to NH3 is a promising strategy. Consequently, optimizing this process could substantially enhance agricultural efficiency and environmental sustainability. In NO2− reduction, the considerably low N
O bond strength makes NO2− an effective precursor for NH3 production through N–H bond formation.9–11 However, efficiency and selectivity are major issues in the case of the NO2− to NH3 reduction reaction (NO2RR), as competing reactions forming N2 and H2 hinder the NH3 yield.12 Thus, developing materials capable of simultaneously mitigating pollution and producing value-added chemicals remains a critical challenge.
Metal phosphides, particularly transition metal phosphides (TMPs), have recently attracted attention owing to their unique electronic and chemical structure. These properties render them promising candidates for diverse electrocatalytic applications, such as water splitting reactions,13 carbon dioxide (CO2) reduction reactions (CO2RR),14,15 and NO2−/NO3− reduction reactions (NO2RR/NO3RR).16,17 The combination of metals such as cobalt (Co), nickel, molybdate, and iron with phosphorus (P) promotes a delocalized electron system, resulting in high electrical conductivity, which is crucial for electrochemical reactions.18–20 Among TMPs, cobalt phosphide (CoP2) is a particularly attractive material owing to its excellent electrical conductivity, strong metal–support interactions, and robust catalytic activity.21,22 Recently, Hou et al. demonstrated that ultrafine deposition of CoP2 on carbon nanosheets markedly enhanced water oxidation reaction performance by improving mass and electron transport properties of CoP2.23 Furthermore, Song et al. revealed that CoP2 facilitates the hydrogenation of carbon monoxide (CO), with P incorporation altering the electronic and adsorption properties of metallic Co. Notably, Coδ+ sites were identified as active centers for the formation of higher oxygenates responsible for CO hydrogenation.24
However, enhancing the overall catalyst reactivity in catalytic reactions requires tuning the material for highly active surfaces, which is of utmost importance. In this case, graphite felt (GF) largely fulfills these requirements by providing large surfaces with high chemical stability. However, the low wettability of GF restricts charge transfer during electrochemical processes.25 The surface engineering of GF with graphene oxide (GO) or the introduction of surface functionalities, such as oxidative modification, has proven to be an effective strategy for increasing the electrode–electrolyte interaction.26–28 Furthermore, to optimize the graphitic surface and enhance its surface characteristics, CO2 laser treatment has gained traction due to its rapid, controllable, and localized energy input. This technique effectively introduces surface micro-defects and increases surface roughness, thereby enhancing wettability, which is advantageous for electrodeposition and electrochemical performance.29–31
Thus, this study aims to establish strong interactions between GO–GF and metal phosphides via coupled CO2-laser-based thermal treatment and electrodeposition to boost NO2RR performance. Furthermore, defect-modulated CoP2 minimizes the charge transfer resistance, enhances the surface redox activity, and boosts NH3 yield at −0.4 V vs. RHE. The efficient NO2RR results further demonstrate the consistency between density functional theory predictions and experimental findings.
2. Experimental section
2.1. Preparation of CoP2-loaded GF
CoP2 was decorated on GF using a two-step method. First, the GF surface was converted to GO by treating it with a CO2 laser at 26.5 W (95% power) for 15 min. Thereafter, electrodeposition was performed on the treated GF using a platinum wire as the counter electrode and Ag/AgCl as the reference electrode at −1.3 V (vs. Ag/AgCl) for 25 min in a 40 mL electrolyte solution containing 0.001 M CoCl2·6H2O, 0.5 M NaH2PO2·H2O, and 0.1 M NaCl. Following electrodeposition, the resulting CoP2-loaded GF was rinsed with deionized water and ethanol, followed by drying at 60 °C. For simplicity, the untreated GF, CO2-treated GF, and CoP2-decorated CO2-treated GF were termed bare GF, GO–GF, and CoP2/GO–GF, respectively.
3. Results and discussion
3.1. Surface transformation and morphological evolution of the graphite felt
The electrodeposition of CoP2 onto GO-engineered GF was performed in a two-step process, as illustrated in Fig. 1a. Owing to the poor wettability of bare GF hindering the electrolytic process, a two-step modification was employed to electrodeposit CoP2 onto the GF carbon (C) surface. First, the GF was thermally etched using a CO2 laser, thereby introducing a GO structure to enhance the wettability and interaction between the GF and CoP2. This tailored GF surface provided a highly receptive platform for the subsequent deposition of CoP2. In the second step, CoP2 was electrochemically deposited at −1.3 V (vs. Ag/AgCl) onto the GO-engineered GF, forming CoP2/GO–GF. This surface modification induced noticeable morphological and compositional changes, altering the elemental ratio of the material (Fig. 1b and c). The initial C content of the bare GF (Fig. 1b) was 99.68 wt%. This decreased to 97.68% after CO2 laser irradiation (Fig. 1c), with an increase in the O content to 2.32%. Finally, electrodeposition introduced Co (2.3%) and P (5.7%), thereby reducing the C content to 82.5%. This confirmed successful electrodeposition, as shown in Fig. 1d. In addition, energy dispersive spectroscopy (EDS) spectra (Fig. S1a and b†) confirmed the formation of CoP2, with compositional analysis of Co and P revealing a stoichiometry consistent with CoP2. This result is further supported by the TEM studies, as depicted in Fig. S2.† The TEM image shows the nanoscale characteristics of the synthesized CoP2/GO–GF, demonstrating a uniform distribution of Co, P, O, and C.
 |
| Fig. 1 (a) Schematic of the CoP2/GO–GF synthesis process and FESEM-EDS analysis of (b) bare GF, (c) GO–GF, and (d and e) CoP2/GO–GF. | |
3.2. Structural and chemical characterization
The crystalline, molecular, and chemical characteristics of the prepared electrocatalytic materials were analyzed through X-ray diffraction (XRD), Raman spectroscopy, and X-ray photoelectron spectroscopy (XPS), as presented in Fig. 2a–e. The Raman spectra of the GF, GO–GF, and CoP2/GO–GF samples (Fig. 2a) exhibited characteristic D and G bands at 1347 and 1595 cm−1, respectively, indicating the presence of C-based structures. The ratio of D to G band intensities (ID/IG) provided insights into the structural order of the C surface, with bare GF exhibiting a high ID/IG ratio (1.22), indicating a highly disordered C structure. However, the functionalization of GF with GO reduced the ID/IG ratio to 1.18, suggesting an increased structural order owing to the layered structure of GO. In contrast, the electrodeposition of CoP2 onto the GO-modified surface increased the ID/IG ratio to 1.31, implying a high defect density. This increase in the ID/IG ratio likely resulted from CoP2-induced disruptions in the C framework, altering the electronic structure of the CoP2/GO–GF and introducing localized disorder within the matrix catalyst. The diffraction pattern presented in Fig. 2b revealed a prominent peak at 10.68° upon CO2 laser irradiation, thereby confirming the presence of GO on GF, along with C reflections at 26.6°, 44°, and 54.7° (JCPDS No. 00-026-1080).32 However, the low concentration of CoP2 yielded weak diffraction signals. To address this, Fig. 2c presents an expanded diffraction view of CoP2/GO–GF, showing two additional peaks at 2θ values of 33° and 35°, which are attributed to electrodeposited CoP2 (JCPDS No. 00-026-0481).
 |
| Fig. 2 Structural and compositional analysis of bare GF, GO–GF, and CoP2/GO–GF: (a) Raman spectra, (b) XRD patterns, (c) expanded XRD view, and XPS spectra of (d) C 1s, (e) O 1s, (f) Co 2p, and (g) P 2p. | |
The XPS survey spectra of GF, GO–GF, and CoP2/GO–GF showed the elemental composition view of C, O, Co, and P, as presented in Fig. S3.† As expected, the survey spectra exhibited an increase in O content from bare GF to GO–GF, providing clear evidence of GF surface transformation. Fig. 2f presents the core C 1s spectra, which display multiple deconvoluted peaks at binding energies (BEs) of 283.8, 284.6, 285.1, 286.1, and 286.6 eV, corresponding to various C bonding states. The peaks at 283.8 and 284.6 eV were attributed to sp2 hybridized C
C and sp3 hybridized C–C bonds, respectively, originating from the graphitic framework of GF. Meanwhile, the peaks at 285.1, 286.1, and 286.6 eV were attributed to O-functionalized C species, specifically the –C–OH, –C–O, and O
C–O groups, respectively.33 In addition, a peak at 291.5 eV observed in GO–GF corresponded to the π–π* transition, which is characteristic of the conjugated electronic structure in GO.34 Furthermore, the deconvoluted O 1s core-level spectra (Fig. 2g) revealed three distinct peaks corresponding to the C
O, C–O, and H–O–H functional groups, with BEs of 531.78, 532.81, and 534.5 eV, respectively, in GF, GO–GF, and CoP2/GO–GF.35 Interestingly, an additional peak at 532.41 eV, observed in the GO–GF and CoP2/GO–GF spectra, corresponded to C–O–H,33 indicating the surface functionalization of GF via GO incorporation. Thereafter, the high-resolution Co spectrum (Co 2p3/2 and Co 2p1/2), depicted in Fig. 2f, exhibited distinct Co states upon deconvolution. The Co 2p signals at BEs of 782.5 and 798.7 eV were attributed to Co–P species, whereas those centered at 779.3 and 794.3 eV corresponded to the Co2+ species, accompanied by satellite peaks at 789.5 and 805.72 eV.36,37 Furthermore, the P 2p signals at low BEs of 129.1 eV (P 2p3/2) and 130 eV (P 2p1/2), were attributed to Co–P linkage.38 However, the peak at 133.2 eV indicated the oxidized P species, a common feature in metal phosphide-based materials.39,40
3.3. Electrocatalytic NO2− reduction reaction (NO2RR)
The electrocatalytic conversion of NO2− to NH3 was investigated using the as-synthesized GF, GO–GF, and CoP2/GO–GF electrocatalysts in an H-type electrochemical cell setup (Fig. S3†), with the results presented in Fig. 3a–h. Following 200 activation cycles (Fig. S4†), linear sweep voltammetry (LSV) was conducted in the presence and absence of NO2−. It revealed notable changes in surface reactions across all studied electrocatalysts in the NO2−-containing environment (Fig. 3a). Furthermore, a substantial enhancement in the current density was observed for CoP2/GO–GF compared to that for bare GF and GO–GF (Fig. 3b), which is attributed to the strong interaction between the electroactive CoP2 and GO–GF.41 This observation was further supported by the Tafel slope analysis (Fig. 3c), which indicated the trend CoP2/GO–GF (148.8 mV dec−1) < GO–GF (220.6 mV dec−1) < GF (335.8 mV dec−1). The small Tafel slope of the CoP2/GO–GF catalyst demonstrated fast charge transfer kinetics, highlighting its superior intrinsic catalytic performance. Similarly, the enhanced charge transfer capability of the catalyst was analyzed using the Nyquist plot measured at −0.13 V vs. RHE (Fig. S5a†). The charge transfer resistance (Rct) was calculated from the Nyquist plot. It was 13.31 Ω for CoP2/GO–GF, which was lower than the values of 49.9 and 49.04 Ω for GF and GO–GF, respectively (Fig. S5b†). The remarkably lower Rct value of CoP2/GO–GF indicates improved electron transport kinetics compared to GO–GF and GF, thereby highlighting its superior conductivity, which facilitates the efficient NO2RR. In addition, notable phase shifts at higher frequencies, as shown in Fig. S6,† reflect the optimized charge transfer dynamics of CoP2/GO–GF, enhancing the rate of the electrochemical NO2RR. Furthermore, the CV curves recorded at different scan rates (10–100 mV s−1, Fig. S7†) provided insights into the surface capacitance and active surface area of the electrocatalysts.
 |
| Fig. 3 Electrocatalytic NO2RR measurements: (a) LSV measurements; (b) comparative current density (mA cm−2) vs. potential (V vs. RHE) trend; (c) Tafel slopes; (d) NH3 partial current densities of GF, GO–GF, and CoP2/GO–GF; (e) NH3 yield via UV-vis spectroscopy; (f) FE of NH3 production; (g) NH4+ yield via NMR spectroscopy; (h) reaction stability over 10 h for CoP2/GO–GF; and (i) comparison of ammonia yield with reported studies. | |
Thus, the considerably higher electrochemical double-layer capacitance (Cdl) of CoP2/GO–GF (3.37 mF cm−2) compared to that of GO–GF (2.84 mF cm−2) and bare GF (1.38 mF cm−2), as depicted in Fig. S8a†, suggests that CoP2/GO–GF possesses a greater density of accessible active sites. A high Cdl indicates the increased ability of the electrocatalyst to store and release charge, thereby increasing its interaction with the electrolyte and yielding an enhanced catalytic performance.42 Electrochemically active surface area calculations were conducted (Fig. S8b†) and revealed a higher density of catalytically active sites in CoP2/GO–GF (0.038 cm2), outperforming GF (0.015 cm2) and GO–GF (0.032 cm2). This resulted in the enhancement of surface–reactant interactions and consequently boosted the electrocatalytic efficiency.
The partial current density is a key metric for evaluating the efficiency of the electrocatalysts in producing the reaction products. In Fig. 3d, the highest partial current density for NH3 production was achieved with CoP2/GO–GF at −0.4 V vs. RHE, surpassing both GO–GF and bare GF. This NO2RR trend (CoP2/GO–GF > GO–GF > GF) was also reflected in the ultraviolet-visible (UV-vis) absorbance (Fig. S9a–c†), which exhibited a higher absorption peak for CoP2/GO–GF compared to that for GO–GF and GF. A similar trend was observed in the proton nuclear magnetic resonance (1H NMR) spectra (Fig. S10†), showing an increased peak area for CoP2/GO–GF. To quantify the NH3 yield, a standard calibration curve was plotted using the indophenol blue method with known NH3 concentrations (Fig. S11a and b†). Bulk electrolysis on CoP2/GO–GF (Fig. S12†) was performed at 0 to −0.4 V vs. RHE, with NH3 yield rates determined from the UV-vis absorption spectra shown in Fig. 3e. A linear increase in NH3 production was observed, reaching a maximum production capacity of 10.6 mg h−1 cm−2 at an optimal potential of −0.4 V vs. RHE. Faradaic efficiency (FE) is a crucial parameter for determining the effectiveness of electrochemical reactions toward target products. However, CoP2/GO–GF demonstrated an exceptional FE of 80% for NH3 measured at −0.4 V vs. RHE, thereby confirming its outstanding NO2RR performance (Fig. 3f). The accuracy of the NH3 yield rate for CoP4/GO–GF was evaluated through three repetitive cycles at each tested potential (−0.1 to −0.4 V vs. RHE), as shown in Fig. S13–S16,† revealing high reproducibility with negligible variations. Meanwhile, a gradual drop in NH3 production and FE at elevated potentials was observed in all the studied materials (Fig. S17a and b†). This suggests that increased competing hydrogen evolution reaction (HER) becomes predominant at high negative potentials.43
Furthermore, 1H NMR was employed to understand the accuracy of the NH3 yield, with known NH3 concentrations (Fig. S18a†) used to construct a calibration curve for comparison with NH3 produced during the NO2RR. Thereafter, the 1H NMR spectra (Fig. S18b†) of the electrolyte confirmed an NH3 yield of approximately 10.35 mg h−1 cm−2 over CoP2/GO–GF at −0.4 V vs. RHE (Fig. 3g), with an FE of ∼78%. This validated the precision and consistency of the reaction, as corroborated by the indophenol blue method (Fig. S19†). The improved performance of CoP2/GO–GF is largely attributed to its high active site density, which enhances its NO2− adsorption and activation capabilities. Active site density calculations for CoP2/GO–GF, GO–GF, and GF yielded values of 17.32, 16.02, and 10.91 μmol cm−2, respectively (Fig. S20†). Consistent with the results reported by Pan et al.,44 who demonstrated that strong active site interactions enhanced the electrochemical performance in Co–P-based HER catalysts, our turnover frequency analysis (Fig. S21†) revealed that CoP2/GO–GF possessed higher intrinsic catalytic activity than GF and GO–GF, thus aligning with its remarkable performance in the NO2RR.
Extended studies on the reusability of the CoP2/GO–GF electrocatalyst were conducted over 10 cycles, with the electrolyte replaced every 1 h. The material exhibited minimal variations in the current density, FE, and NH3 yield, thereby confirming its stable catalytic performance (Fig. 3h and S22a–c†). Furthermore, the comparative analysis presented in Fig. 3i and Table S1† illustrates the superior performance of CoP2/GO–GF relative to recently reported catalysts for NO2− reduction. Following 10 repetitive cycles, the compositional stability of the material was assessed using SEM-EDS color mapping (Fig. S23†) on CoP2/GO–GF. Apart from slight variations, no profound changes in elemental distribution were observed. This finding was further supported by XPS analysis of CoP2/GO–GF (Fig. S24†), exhibiting consistent peak splitting in the C 1s, O 1s, Co 2p, and P 2p spectra before and after the reaction. Notably, the Co 2p and P 2p peaks at BEs of approximately 782 eV (2p3/2) and 130/129 eV (2p1/2/2p3/2), respectively, indicated no structural deformation in CoP2. The material characterization of CoP2/GO–GF following the stability study aligned with the results from the multiple cycle analysis (Fig. 3h), as demonstrated earlier. Additionally, an extended 50 h chronoamperometric stability test was conducted (see Fig. S25†), during which the catalyst retained ∼80% of its initial current density (from −125.5 to −100.97 mA cm−2), thereby underscoring the prolonged durability of CoP2/GO–GF.
3.4.
In situ NO2RR and DFT studies
The ex situ FTIR analysis was conducted to investigate the detailed NO2− reduction reaction (NO2RR) pathway under varying applied potentials, ranging from open circuit potential (OCP) to −0.5 V vs. RHE. The results, presented in Fig. 4a–c, provide insights into the sequential transformation of nitrite into ammonium. At OCP, the IR spectra exhibit a characteristic peak corresponding to NO2− vibrations at 1226 cm−1, along with a broad peak at 1637 cm−1 attributed to H2O. However, due to the overlap between the N–O stretching vibration (∼1600–1700 cm−1) and the H–O–H bending mode of water, distinguishing NO as an intermediate is challenging.45 Despite this spectral overlap, as the applied potential becomes more negative, the progressive evolution of the spectra reveals the formation of NH4+ (∼1400 cm−1), confirming a stepwise electrochemical reduction process. This spectral trend suggests that NO appears as a key intermediate in the NO2RR pathway, undergoing further reduction to NH → NH2 before its ultimate conversion to NH4+. Fig. 4b and c provide a colormap representation of potential (V vs. RHE) versus wavenumber, enabling a clearer visualization of intermediate species and product formation. The color gradient represents IR absorbance intensity, where red-shifted regions at more negative potentials indicate an increased presence of NH4+ alongside residual NO2−. A noticeable shift in peak intensity in Fig. 4c, particularly at −0.3 V and −0.4 V, correlates with enhanced NH4+ production, suggesting that these potentials favor the highest ammonium yield. These results confirm that NO acts as an active intermediate, undergoing further reduction (NO → NH → NH2 → NH3) in the NO2RR pathway.46,47
 |
| Fig. 4
Ex situ FTIR spectra (a)–(c) and DFT studies of the CoP2/GO structure: (d) optimized structure, (e) band and DOS plot, (f) PDOS plot of the CoP2/GO structure, and (g) reaction profile of the NO2RR pathway over CoP2/GO–GF. | |
To gain deeper insights into the structural and electronic interactions governing NO2RR efficiency, density functional theory (DFT) calculations were performed. The CoP2/GO–GF structure was modeled using the monoclinic CoP2 crystal system (P21/c) with the (200) plane, as observed in the XRD pattern. Graphene oxide (GO) was incorporated into the crystal system with CoP2 as the active surface, and the structure was subsequently optimized. The optimized structure is shown in Fig. 4d, where the C–C bond lengths are approximately 1.49 Å. This length results from the heterostructure formation and strain induced in GO by the monoclinic crystal structure, aligning with previous studies.48–50 The Co–P, P–P, C–O, and O–H bond lengths are measured to be ∼2.33 Å, 2.27 Å, 1.44 Å, and 0.99 Å, respectively, closely matching previously reported values of 2.32 Å, 2.26 Å, 1.46 Å, and 0.99 Å.51–54 To evaluate the structural stability of the system, the cohesive energy (EC) was calculated, yielding a value of −1.85 eV per atom, which confirms the structural integrity of the CoP2/GO heterostructure. Additionally, the formation energy (EF) was computed to evaluate the thermodynamic feasibility of the structure, revealing a negative EF value of approximately −1.23 eV, indicating that the formation is energetically favorable. To examine the electronic properties of the system, the band structure and density of states (DOS) were analyzed, with the corresponding plots displayed in Fig. 4e. The band structure of the CoP2/GO heterostructure shows metallic characteristics, featuring overlapping energy states near the Fermi level, which is consistent with the DOS analysis. The DOS plots reveal symmetric energy states for both spin-up and spin-down channels, indicating a non-magnetic nature with no net spin polarization. The atomic contributions to the energy states were further investigated using the projected DOS plot in Fig. 4f. The Co atom shows significant contributions to energy states in the valence band region and near the Fermi level, extending into the conduction bands. Notably, there is a strong overlap of C atoms with CoP2 in the valence band region. At the Fermi level, the overlap between C and Co atoms suggests strong interactions between GO and Co atoms, facilitating charge transfer and enhancing electronic coupling.
To evaluate NO2RR performance, the NO2 molecule was initially adsorbed onto the CoP2/GO structure. The calculated adsorption energy of approximately −1.55 eV indicates that NO2 adsorption is highly favorable. The catalytic activity of the system primarily depends on the adsorption of reaction intermediates and the reaction energies of elementary steps.55 The intermediates were placed at an initial distance of 3 Å, optimized, and the corresponding structures are presented in Fig. S26.† The free energy changes (ΔG) for each reaction step were calculated, and the reaction energy profile is shown in Fig. 4g. The free energy of the initial
adsorption was determined to be −1.55 eV, suggesting that the reaction proceeds spontaneously. In the fifth step of the NO2RR pathway, two possible reaction mechanisms can occur: (i) NOH* → N*, involving the elimination of H2O, and (ii) NOH* → NHOH*, where an H* atom interacts with the N atom before the H2O elimination. Based on the calculated ΔG values, the NOH* → N* pathway exhibits a lower reaction free energy, indicating that it is more thermodynamically favorable than the NOH* → NHOH* pathway. This suggests that the direct elimination of H2O from NOH* to form N* is the preferred route in the NO2RR process. Finally, for NH4+ formation, the ΔG value for the
step must be sufficiently low to facilitate the reaction. The observed ΔG of −0.13 eV confirms the feasibility of NH4+ formation. The highest reaction energy among all elementary steps is considered the rate-limiting step (RLS), which determines the catalytic performance of the material.56,57 In this study, the third step (NO2H* → NO*) is identified as the RLS, with a reaction energy barrier of 0.26 eV, suggesting that the material must overcome a moderate energy barrier. However, this energy barrier remains within a feasible range for the reaction to proceed. Furthermore, the faradaic efficiency (FE) for the NO2RR process was calculated, yielding a value of ∼100% under ambient conditions for the RLS of 0.26 eV. This result indicates that the material exhibits outstanding catalytic performance.58 These findings suggest that the CoP2/GO heterostructure demonstrates remarkable energetic stability and high selectivity toward NO2, with an RLS of 0.26 eV and an FE of 100%, making it a highly promising candidate for NO2 reduction applications.
4. Conclusion
This study demonstrated the fabrication of a highly active CoP2/GO–GF electrocatalyst, developed through the stepwise modification of GF with a CO2 laser followed by electrodeposition of CoP2. Comprehensive structural and compositional analysis revealed prominent surface transformations, indicating the importance of enhanced active site density in the NO2RR. The CoP2/GO–GF electrocatalyst exhibited remarkable NO2RR performance, achieving a high NH3 yield rate of 10.6 mg h−1 cm−2 and an FE of 80% at −0.4 V vs. RHE. Interestingly, enhanced charge transfer kinetics, lower resistance, and a high density of accessible active sites were key to its superior catalytic activity. Ex situ FTIR and DFT calculations confirmed NO* as a key intermediate in the NO2RR pathway, with the NO2H* → NO* step identified as the rate-limiting step (0.26 eV), ensuring a feasible and selective conversion to NH3. Furthermore, the material demonstrated excellent durability and reusability in NH3 formation, retaining consistent performance over 10 extended cycles. Its high stability and selectivity highlight its potential for practical applications in sustainable nitrogen conversion. This study showcases the rapid design of CoP2/GO–GF and its promising performance for sustainable NO2− to NH3 conversion, thereby demonstrating its applicability in waste-to-value-added product processes.
Data availability
Data are available within the article and its ESI.†
Conflicts of interest
The authors declare no competing financial interests.
Acknowledgements
This research was supported by the Korea Basic Science Institute (National Research Facilities and Equipment Center) grant funded by the Ministry of Education (RS-2024-00434932). The authors express their gratitude for the financial assistance of the National Research Foundation of Korea (NRF) (2022R1A2C2010686 and RS-2024-00405324). This work was supported by the Glocal University 30 Project Fund of Gyeongsang National University in 2024.
References
- W. de Vries, Current Opinion in Environmental Science & Health, 2021, 21, 100249 Search PubMed.
- H. Liu, X. Liu, Y. Yu, W. Yang, J. Li, M. Feng and H. Li, J. Mater. Chem. A, 2018, 6, 4611–4616 RSC.
- S. Lee, Y. Lee and W. Choi, Appl. Catal., B, 2023, 327, 122432 CrossRef CAS.
- Y. Lee, J. Theerthagiri, N. Yodsin, A. Min, C. J. Moon, S. Jungsuttiwong and M. Y. Choi, Angew. Chem., Int. Ed., 2024, 63, e202413774 CrossRef CAS PubMed.
- S. Lu, Q. Zhu and R. Li, J. Colloid Interface Sci., 2023, 652, 1481–1493 CrossRef CAS PubMed.
- Y. Fan, X. Wang, C. Butler, A. Kankam, A. Belgada, J. Simon, Y. Gao, E. Chen and L. R. Winter, Nat. Water, 2024, 2, 684–696 CrossRef CAS.
- A. M. Bergquist, J. K. Choe, T. J. Strathmann and C. J. Werth, Water Res., 2016, 96, 177–187 CrossRef CAS.
- J. Theerthagiri, J. Park, H. T. Das, N. Rahamathulla, E. S. F. Cardoso, A. P. Murthy, G. Maia, D. V. N. Vo and M. Y. Choi, Environ. Chem. Lett., 2022, 20, 2929–2949 CrossRef CAS.
- F. Wang, H. Zhao, G. Zhang, H. Zhang, X. Han and K. Chu, Adv. Funct. Mater., 2024, 34, 2308072 CrossRef CAS.
- H. Zhao, J. Xiang, Z. Sun, S. Shang and K. Chu, ACS Sustain. Chem. Eng., 2024, 12, 2783–2789 CrossRef CAS.
- C. Yang, T. Wei, C. Wang, F. Yue, X. Li, H. Pang, X. Zheng, Y. Zhang and F. Fu, Mater. Horiz., 2025, 12, 877–885 RSC.
- H. Liu, J. Park, Y. Chen, Y. Qiu, Y. Cheng, K. Srivastava, S. Gu, B. H. Shanks, L. T. Roling and W. Li, ACS Catal., 2021, 11, 8431–8442 CrossRef CAS.
- J. Theerthagiri, A. P. Murthy, S. J. Lee, K. Karuppasamy, S. R. Arumugam, Y. Yu, M. M. Hanafiah, H.-S. Kim, V. Mittal and M. Y. Choi, Ceram. Int., 2021, 47, 4404–4425 CrossRef CAS.
- K. U. D. Calvinho, A. W. Alherz, K. M. K. Yap, A. B. Laursen, S. Hwang, Z. J. L. Bare, Z. Clifford, C. B. Musgrave and G. C. Dismukes, J. Am. Chem. Soc., 2021, 143, 21275–21285 CrossRef CAS PubMed.
- M. Esmaeilirad, A. Kondori, N. Shan, M. T. Saray, S. Sarkar, A. M. Harzandi, C. M. Megaridis, R. Shahbazian-Yassar, L. A. Curtiss, C. U. Segre and M. Asadi, Appl. Catal., B, 2022, 317, 121681 CrossRef CAS.
- A. Zhang, Y. Liang, X. He, X. Fan, C. Yang, L. Ouyang, D. Zheng, S. Sun, Z. Cai, Y. Luo, Q. Liu, S. Alfaifi, L. Cai, H. Wang and X. Sun, Inorg. Chem., 2023, 62, 12644–12649 CrossRef CAS PubMed.
- Z. Jiang, Y. Wang, Z. Lin, Y. Yuan, X. Zhang, Y. Tang, H. Wang, H. Li, C. Jin and Y. Liang, Energy Environ. Sci., 2023, 16, 2239–2246 RSC.
- J. Zou, S. Wu, Y. Lin, X. Li, Q. Niu, S. He and C. Yang, Environ. Sci. Technol., 2024, 58, 14895–14905 CrossRef CAS PubMed.
- S. K. T. Aziz, S. Sultana, A. Kumar, S. Riyajuddin, M. Pal and A. Dutta, Cell Rep. Phys. Sci., 2023, 4, 101747 CrossRef CAS.
- G. Li, H. Chen, B. Zhang, H. Guo, S. Chen, X. Chang, X. Wu, J. Zheng and X. Li, Appl. Surf. Sci., 2022, 582, 152404 CrossRef CAS.
- L. Shen, Y. Qian, D.-H. Kim and D. J. Kang, Int. J. Hydrogen Energy, 2024, 61, 996–1003 CrossRef CAS.
- J. Wang, Z. Liu, Y. Zheng, L. Cui, W. Yang and J. Liu, J. Mater. Chem. A, 2017, 5, 22913–22932 RSC.
- C.-C. Hou, L. Zou, Y. Wang and Q. Xu, Angew. Chem., Int. Ed., 2020, 59, 21360–21366 CrossRef CAS PubMed.
- X. Song, Y. Ding, W. Chen, W. Dong, Y. Pei, J. Zang, L. Yan and Y. Lu, Energy Fuels, 2012, 26, 6559–6566 CrossRef CAS.
- W. Zhou, X. Meng, J. Gao, F. Sun and G. Zhao, Chemosphere, 2021, 278, 130382 CrossRef CAS.
- S. S. Barton, M. J. B. Evans, E. Halliop and J. A. F. MacDonald, Langmuir, 1997, 13, 1332–1336 CrossRef CAS.
- Y. Wang, W. Zhou, J. Gao, Y. Ding and K. Kou, J. Electroanal. Chem., 2019, 833, 258–268 CrossRef CAS.
- F. Tang, D. He, H. Jiang, R. Wang, Z. Li, W. Xue and R. Zhao, Carbon, 2022, 197, 163–170 CrossRef CAS.
- Y. Zhao, S. Wang, X. Wang, L. Zhang, J. Sha, X. Zhang, R. Mu and Z. Dong, Fuel, 2024, 373, 132318 CrossRef CAS.
-
D. G. Waugh and J. Lawrence, in Polymer Surface Modification to Enhance Adhesion, 2024, pp. 365–388, DOI:10.1002/9781394231034.ch9.
- L. Hao and J. Lawrence, Mater. Sci. Eng., A, 2004, 364, 171–181 CrossRef.
- O. C. Compton, B. Jain, D. A. Dikin, A. Abouimrane, K. Amine and S. T. Nguyen, ACS Nano, 2011, 5, 4380–4391 CrossRef CAS PubMed.
- A. Mukhopadhyay, Y. Yang, Y. Li, Y. Chen, H. Li, A. Natan, Y. Liu, D. Cao and H. Zhu, Adv. Funct. Mater., 2019, 29, 1903192 CrossRef CAS.
- J. C. Silva Filho, E. C. Venancio, S. C. Silva, H. Takiishi, L. G. Martinez and R. A. Antunes, SN Appl. Sci., 2020, 2, 1450 CrossRef CAS.
- G. S. dos Reis, S. H. Larsson, M. Mathieu, M. Thyrel and T. N. Pham, Biomass Convers. Biorefin., 2023, 13, 10113–10131 CrossRef CAS.
- Y. Wang, Y. Jiao, H. Yan, G. Yang, C. Tian, A. Wu, Y. Liu and H. Fu, Angew. Chem., Int. Ed., 2022, 61, e202116233 CrossRef CAS.
- Y. Sun, T. Liu, Z. Li, A. Meng, G. Li, L. Wang and S. Li, Chem. Eng. J., 2022, 433, 133684 CrossRef CAS.
- C. Wu, Y. Yang, D. Dong, Y. Zhang and J. Li, Small, 2017, 13, 1602873 CrossRef.
- H. Huang, C. Li, F. Yan, F. Yuan, X. Liang, W. Zhou and J. Guo, Appl. Surf. Sci., 2023, 623, 157079 CrossRef CAS.
- D. C. Phillips, S. J. Sawhill, R. Self and M. E. Bussell, J. Catal., 2002, 207, 266–273 CrossRef CAS.
- J. Wang, W. Yang and J. Liu, J. Mater. Chem. A, 2016, 4, 4686–4690 RSC.
- T. Binninger, T. J. Schmidt and D. Kramer, Phys. Rev. B, 2017, 96, 165405 CrossRef.
- X. Long, T. Zhong, F. Huang, P. Li, H. Zhao, J. Fang, D. Shu and C. He, Appl. Catal., B, 2025, 365, 124944 CrossRef CAS.
- Y. Pan, Y. Lin, Y. Chen, Y. Liu and C. Liu, J. Mater. Chem. A, 2016, 4, 4745–4754 RSC.
- S. Han, H. Li, T. Li, F. Chen, R. Yang, Y. Yu and B. Zhang, Nat. Catal., 2023, 6, 402–414 CrossRef CAS.
- J. Xiang, H. Zhao, K. Chen, X. Yang and K. Chu, J. Colloid Interface Sci., 2024, 659, 432–438 CrossRef CAS PubMed.
- Y. Zhu, H. Duan, C. G. Gruber, W. Qu, H. Zhang, Z. Wang, J. Zhong, X. Zhang, L. Han, D. Cheng, D. D. Medina, E. Cortés and D. Zhang, Angew. Chem., Int. Ed., 2025, 64(11), e202421821 CrossRef CAS PubMed.
- L. Fan, J. Xu and Y. Hong, RSC Adv., 2022, 12, 6772–6782 RSC.
- B. Ouyang, S. Xiong, Z. Yang, Y. Jing and Y. Wang, Nanoscale, 2017, 9, 8126–8132 RSC.
- D. Arumugam, M. Subramani, D. Subramani and S. Ramasamy, Int. J. Hydrogen Energy, 2024, 68, 545–558 CrossRef CAS.
- Y. Han, J. Zhao, X. Guo and T. Jiao, Langmuir, 2023, 39, 7648–7659 CrossRef CAS PubMed.
- D. Liu, Y. Shi, L. Tao, D. Yan, R. Chen and S. Wang, Chin. Chem. Lett., 2019, 30, 207–210 CrossRef CAS.
- Z. Liang, X. Zhong, T. Li, M. Chen and G. Feng, ChemElectroChem, 2019, 6, 260–267 CrossRef CAS.
- M. Subramani, D. Arumugam and S. Ramasamy, Int. J. Hydrogen Energy, 2023, 48, 4016–4034 CrossRef CAS.
- P. Lv, D. Wu, B. He, X. Li, R. Zhu, G. Tang, Z. Lu, D. Ma and Y. Jia, J. Mater. Chem. A, 2022, 10, 9707–9716 RSC.
- D. Arumugam, D. Subramani, A. Muralidharan and S. Ramasamy, ACS Appl. Mater. Interfaces, 2024, 16, 64916–64928 CrossRef CAS.
- K. Yang, S.-H. Han, C. Cheng, C. Guo, T. Li and Y. Yu, J. Am. Chem. Soc., 2024, 146, 12976–12983 CrossRef CAS PubMed.
- N. Sathishkumar, S.-Y. Wu and H.-T. Chen, Appl. Surf. Sci., 2022, 598, 153829 CrossRef CAS.
|
This journal is © The Royal Society of Chemistry 2025 |
Click here to see how this site uses Cookies. View our privacy policy here.