Abstract
The electrochemical nitric oxide reduction reaction (NORR) holds a great potential for removing environmental pollutant NO and meanwhile generating high value-added ammonia (NH3). Herein, we tactfully design and synthesize a ternary Co/Co3O4/CoB heterostructure that displays a high NH3 Faradaic efficiency of 98.8% in NORR with an NH3 yield rate of 462.18 µmol cm−2 h−1 (2.31 mol h−1 gcat−1) at −0.5 V versus reversible hydrogen electrode, outperforming most of the reported NORR electrocatalysts to date. The superior NORR performance is attributed to the enhanced charge and proton transfer over the ternary Co/Co3O4/CoB heterostructure. The charge transfer between CoB and Co/Co3O4 yields electron-deficient Co and electron-rich Co3O4. The electron-deficient Co sites boost H2O dissociation to generate *H while the electron-rich low-coordination Co3O4 sites promote NO adsorption. The *H formed on electron-deficient Co sites is more favorable to transfer to electron-rich Co3O4 sites adsorbed with NO, facilitating the selective hydrogenation of NO. This study paves the way for designing and developing highly efficient electrocatalysts for electrochemical reduction of NO to NH3.
Similar content being viewed by others
Introduction
Ammonia (NH3) is an indispensable chemical for agriculture, chemical industry, refrigerant, hydrogen carrier, and healthcare1,2,3,4,5. Unfortunately, NH3 production by traditional Haber-Bosch method requires harsh conditions of high temperature and high pressure, emitting a large quantity of greenhouse gases6,7. One of the alternatives to the Haber-Bosch method for NH3 synthesis is the electrocatalytic reduction of nitrogen-containing species, including nitrogen gas (N2), nitric oxide (NO), nitrite (NO2−), and nitrate (NO3−)8,9,10,11. The nitrogen reduction reaction (NRR) suffers from low NH3 Faradaic efficiency (FENH3) and NH3 yield rate because of the high dissociation energy (941 kJ mol−1) of N ≡ N and the competitive hydrogen evolution reaction (HER) (the reduction potential of N2 (0.093 V vs. RHE) and H2O (0 V vs. RHE) are close to each other)12,13,14. The electroreduction of NO2− and NO3− to NH3 (NO2−RR and NO3−RR) have more complicated reaction pathways, and generate more types of N-containing by-products15,16. Compared with NRR, the electrochemical NO reduction reaction (NORR) is kinetically and thermodynamically more favorable because NO possesses a lower dissociation energy (204 kJ mol−1) and a more positive reduction potential (0.71 V vs. RHE)17,18,19. The approach to couple NO oxidation with NO3− reduction is also considered as an alternative to NORR. But NORR to produce NH3 generally requires lower energy consumption20,21. In recent years, NORR is becoming increasingly attractive because NORR not only can remove environmental pollutant NO from industrial exhaust gas, but also produces value-added chemicals.
The electrochemical reduction of NO involves chemical activation of NO followed by NO hydrogenation22,23,24. Currently, the FENH3 and NH3 yield rate in electrochemical NORR are still impeded by the weak adsorption of NO and the sluggish proton supply. To improve the NORR performance, researchers have made endeavors to enhance the adsorption of NO by transition metal oxide modification or vacancy engineering25,26. The proton supply could be enhanced by using metal-based electrocatalysts that would facilitate H2O dissociation24,27. Constructing single atoms in amorphous metal oxides with oxygen vacancies to form metal-O moieties could accelerate hydrogenation of NO to produce NH₃28,29,30. However, till now, it is still intractable to acquire both high FENH3 and NH3 yield rate simultaneously because of the complex balance of the NO adsorption and H2O dissociation. Early studies have shown fairly high NO adsorption capacity over Co3O4 surface31,32,33,34. However, the poor activation and dissociation of H2O on Co3O4 greatly restricts the hydrogenation reaction in NORR35. Metallic Co on Co3O4 heterostructure is designed to promote H2O dissociation to increase available *H36. But this concurrently accelerates HER, leading to a low NORR FENH3 and NH3 yield rate27,30,37,38,39. Tuning charge distribution on catalyst’s surface is able to directionally drive *H transfer to enhance NO hydrogenation during NORR40,41. Boron (B) possesses a 2s2p1 electronic structure, which exhibits flexible valence states during catalytic reactions. Introducing B into the Co/Co3O4 heterostructure is anticipated to optimize NO adsorption and H2O dissociation, and in the meantime induce *H transfer to promote NO hydrogenation and suppress HER during NORR, thus greatly enhancing NORR activity and selectivity.
In this study, we tactfully designed and synthesized a ternary Co/Co3O4/CoB heterostructure by reducing Co3O4 with NaBH4. The ternary Co/Co3O4/CoB catalyst exhibited an excellent electrochemical NORR performance, achieving an NH3 yield rate of 462.18 µmol cm−2 h−1 (2.31 mol h−1 gcat−1) and an FENH3 of 98.80% at −0.5 V vs. RHE, outperforming most of the reported NORR electrocatalysts to date. To demonstrate the application potential, a Zn-NO battery was assembled using the Co/Co3O4/CoB as the cathode, which delivered a high power density of 10.06 mW cm−2. A series of experiments and density functional theory (DFT) calculations revealed that the enhanced charge and proton transfer over the ternary Co/Co3O4/CoB effectively boosted the electrochemical NORR. In detail, the charge transfer between CoB and Co/Co3O4 yielded electron-deficient Co and electron-rich Co3O4. The electron-deficient Co sites boosted H2O dissociation to generate *H, while the electron-rich low-coordination Co3O4 sites promoted NO adsorption and *H transfer. Thanks to the enhanced charge and proton transfer, the energy barrier of the NORR potential-determining step from *NO to *HNO over the ternary Co/Co3O4/CoB was greatly reduced. Furthermore, the Co/Co3O4/CoB could also facilitate adsorption of *NH and *NH2, beneficial for NO reduction to produce NH3.
Results
Catalyst synthesis and structure characterization
The ternary Co/Co3O4/CoB heterostructured electrocatalyst was synthesized by reducing Co3O4 using NaBH4 in an Ar atmosphere at 500 °C for 4 h (Fig. 1a). By applying the same method, Co/Co3O4 could also be obtained by decreasing the quality ratio of NaBH4/Co3O4. The as-prepared Co/Co3O4/CoB shows the morphology of nanoparticles supported on nanosheets based on the transmission electron microscope (TEM, Supplementary Fig. 1) measurement. The average thickness of the nanosheets is ∼3.3 nm determined by atomic force microscopy (AFM, Fig. 1b). The nanoparticles on the nanosheets are composed of large nanoparticles surrounded by many small nanoparticles (HRTEM, Supplementary Fig. 2). Co/Co3O4/CoB shows Co3O4 (311) and CoB (021) crystalline phases from the HRTEM image (Fig. 1c). To precisely determine the structure of Co/Co3O4/CoB, aberration-corrected HAADF-STEM measurement was further performed. As shown in Fig. 1d, abundant interfaces among Co, Co3O4, and CoB phases can be clearly observed, suggesting the formation of a ternary Co/Co3O4/CoB heterostructure. The electron energy loss spectroscopy (EELS) mapping (Fig. 1e) combined with the HRTEM image (Supplementary Fig. 2) clarifies that the Co/Co3O4/CoB heterostructure is composed of a CoB nanosheet supported with large Co and small Co3O4 nanoparticles. For comparison, Co/Co3O4 shows a heterostructure of Co nanoparticles supported on Co3O4 nanosheets (Supplementary Figs. 3 and 4).
a Schematic diagram showing the synthesis process of Co/Co3O4/CoB. b AFM image of Co/Co3O4/CoB. The inset shows the height profile along the red line in (b). c HRTEM image of Co/Co3O4/CoB. d Aberration-corrected HAADF-STEM image of Co/Co3O4/CoB. e HAADF-STEM image and the corresponding EELS mappings of Co, O, and B elements in Co/Co3O4/CoB. Source data for Fig. 1 are provided as a Source Data file.
The structure and valence state of Co/Co3O4/CoB were analyzed by X-ray powder diffraction (XRD) and X-ray photoelectron spectroscopy (XPS). As shown in Fig. 2a, the XRD patterns show clear diffraction peaks of Co3O4 (PDF #42-1467), Co (PDF #15-0806) and CoB (PDF #03-0959) for Co/Co3O4/CoB, and Co3O4 (PDF #42-1467) and Co (PDF #15-0806) for Co/Co3O4. The B 1s XPS spectrum of Co/Co3O4/CoB also evidences the formation of CoB (Supplementary Fig. 5). In Fig. 2b, the Co 2p3/2 XPS spectrum of Co/Co3O4 exhibits three peaks at binding energies of 779.7, 781.0 and 783.1 eV, which can be attributed to Co0, Co3+ and Co2+, respectively42,43,44. For Co/Co3O4/CoB, the Co0 peak slightly shifts to a higher binding energy (780.1 eV), while the Co3+ (780.7 eV) and Co2+ (782.2 eV) peaks shift toward lower binding energies as compared to those of Co/Co3O4. Because Co0 and Co3+/Co2+ mainly exist in Co and Co3O4, respectively, it can be inferred that the introduction of CoB leads to more electron loss in Co and more electron gain in Co3O4. To clarify the electron transfer among different interfaces in Co/Co3O4/CoB and Co/Co3O4, DFT calculations were conducted. Herein, by balancing the computation power and calculation accuracy, the Co/Co3O4/CoB was modeled by Co (111) and Co3O4 (311) clusters supported on CoB (021), while Co/Co3O4 was modeled by Co (111) clusters supported on Co3O4 (311) (for details see DFT calculations in Method section). Firstly, the Bader charge was analyzed over Co/Co3O4 and Co/Co3O4/CoB models. For the Co/Co3O4 model, 3.64 |e| is transferred from Co to Co3O4 (Supplementary Fig. 6, Supplementary Table 1, and Supplementary Data 1). In Co/Co3O4/CoB, Co3O4 gets 2.23 |e| and 2.36 |e| from Co and CoB, respectively (Supplementary Fig. 7, Supplementary Table 1, and Supplementary Data 1). The charge density difference analysis of Co/Co3O4/CoB indicates a significant charge redistribution at the interface between CoB and Co/Co3O4 (Fig. 2c), with electron transfer from Co to CoB and from CoB to Co3O4. Furthermore, the average numbers of electron transfer per Co atom in Co/Co3O4/CoB and Co/Co3O4 were analyzed. As compared in Supplementary Table 1, Co loses more electrons (−0.22 e vs. −0.13 e) and Co3O4 gains more electrons (+0.51 e vs. +0.08 e) in Co/Co3O4/CoB as compared to Co/Co3O4, evidencing that CoB introduction can modulate the electronic structure of Co and Co3O4 in Co/Co3O4/CoB. X-ray absorption spectroscopy (XAS), including X-ray absorption near-edge structure (XANES) and extended X-ray absorption fine structure (EXAFS), was performed to examine the electronic structure and coordination environment of Co in Co/Co3O4/CoB and Co/Co3O4. In the Co K-edge XANES spectra (Fig. 2d), the position of the absorption edge for Co/Co3O4/CoB and Co/Co3O4 are located between that for Co foil and Co3O4 references, indicating that the average valence state of Co atom in Co/Co3O4/CoB and Co/Co3O4 are within 0 and +2/+3, owing to co-existence of both metallic Co and Co3O4. To understand the local structure around Co atom in these samples, the EXAFS spectra are transformed into R-space. The R-space EXAFS spectra (Fig. 2e) display dominant peaks of Co-B (1.60 Å)45, Co-Co (2.17 Å)46, and Co-O (1.44 Å)47 coordination bonds in CoB, Co foil and Co3O4, respectively. The FT-EXAFS spectra of Co/Co3O4/CoB and Co/Co3O4 can be well-fitted using backscattering paths of Co-Co, Co-B and Co-O (Fig. 2f, Supplementary Tables 2 and 3). Notably, the Co-O coordination number (2.2) in Co/Co3O4/CoB is significantly lower than that (5.7) in Co/Co3O4. The low-coordinated Co-O structure in Co/Co3O4/CoB is beneficial for NO adsorption48. The wavelet transform (WT) contour plot of Co/Co3O4/CoB (Fig. 2g) displays three intensity maximums at around 7.2, 5.9, and 5.4 Å−1, corresponding to the Co-Co, Co-B, and Co-O coordination, respectively47,49. For comparison, Co/Co3O4 only shows Co-Co and Co-O coordinations (Supplementary Fig. 8).
a XRD patterns of Co/Co3O4 and Co/Co3O4/CoB. b Co 2p3/2 XPS spectra of Co/Co3O4 and Co/Co3O4/CoB. c The charge density difference analysis for Co/Co3O4/CoB. Yellow and blue colors represent electron accumulation and depletion regions, respectively. The isosurface value is 0.004 e/Å3. d Co K-edge XANES spectra of Co/Co3O4, Co/Co3O4/CoB, Co foil, CoB, and Co3O4. e The FT-EXAFS spectra of Co/Co3O4, Co/Co3O4/CoB, Co foil, CoB, and Co3O4. f FT-EXAFS fitting curves for Co/Co3O4 and Co/Co3O4/CoB in R space. g WT contour plots of Co/Co3O4/CoB, Co foil, CoB, and Co3O4. Source data for Fig. 2 are provided as a Source Data file.
Electrochemical NORR performance
The electrochemical NO reduction performance over Co/Co3O4/CoB was assessed in an H-type electrolytic cell filled with 0.1 M phosphate buffer saline (PBS) aqueous solution. As shown in Fig. 3a and Supplementary Fig. 9, the linear sweep voltammetry (LSV) curves (iR-corrected or not) reveal that both Co/Co3O4/CoB and Co/Co3O4 exhibit a significantly increased current density in 10 vol.% NO/Ar atmosphere as compared to pure Ar atmosphere, suggesting preferred NORR on Co/Co3O4/CoB and Co/Co3O4 surface. Chronoamperometric (1 h electrolysis) and ultraviolet-visible (UV-vis) colorimetric measurements were used to quantitatively determine the NH3 Faradaic efficiency (FE) and yield rate (Supplementary Figs. 10 and 11). At −0.5 V vs. RHE, the Co/Co3O4/CoB displays the optimum FENH3 of 98.80% and NH3 yield rate of 462.18 µmol h−1 cm−2 (2.31 mol h−1 gcat−1), much higher than Co/Co3O4 (FENH3 of 91.28% and NH3 yield rate of 252.34 µmol h−1 cm−2, Fig. 3b, c). 1H nuclear magnetic resonance (NMR) measurements give comparable results (Supplementary Figs. 12–14). The possible by-products produced over Co/Co3O4/CoB during NORR were checked by differential electrochemical mass spectroscopy (DEMS), UV-vis absorption spectroscopy, and gas chromatography (Supplementary Figs. 15–18). The hydroxylamine (NH2OH), N2 and N2O by-products were not detected. Moreover, we conducted the NORR experiment over Co/Co3O4/CoB in 0.1 M KOH aqueous electrolyte. The NORR performance was poorer as compared with the experiment conducted in 0.1 M PBS (Supplementary Fig. 19). The better NORR performance of Co/Co3O4/CoB as compared to that of Co/Co3O4 can be further reflected by the smaller semicircle radius in the electrochemical impedance spectroscopy (EIS) spectrum (Supplementary Fig. 20), indicating a smaller charge transfer resistance (Rct) over Co/Co3O4/CoB surface. A series of control experiments indicate that the nitrogen in the NH3 product originates solely from NO (Fig. 3d). Carbon paper itself shows negligible NORR activity, disclosing that NH3 is mainly produced on Co/Co3O4/CoB (Fig. 3d). Using isotopic 15NO as the reactant, the characteristic double peaks (inset in Fig. 3d) appear with a separation of 72 Hz, corresponding to 15NH4+, indicating that 15NO is the sole nitrogen source of NH3 production. Besides excellent catalytic activity, Co/Co3O4/CoB also displays good catalytic stability. During five consecutive electrolysis cycles at −0.5 V vs. RHE, the FENH3 and NH3 yield rate remain nearly unchanged (Fig. 3e). Moreover, this catalyst exhibits good current stability for ~100 h at −0.5 V vs. RHE (Supplementary Fig. 21). Co/Co3O4/CoB shows a good structural stability during the NORR based on the XRD, HRTEM and XPS characterizations (Supplementary Figs. 22–24). The ternary Co/Co3O4/CoB achieves a high FENH3 of 98.80% and NH3 yield rate of 462.18 µmol h−1 cm−2 as well as long-time stability at a low cathodic potential of −0.5 V vs. RHE, superior to most of the reported NORR electrocatalysts (Fig. 3f and Supplementary Tables 4 and 5). To demonstrate the practical application potential of the Co/Co3O4/CoB catalyst, a Zn-NO battery with Co/Co3O4/CoB cathode and Zn plate anode was assembled (Supplementary Fig. 25), which could deliver an open circuit voltage (OCV) of 2.04 V (Supplementary Fig. 26) and a maximum power density of 10.06 mW cm−2, outperforming all reported results in the literature (Supplementary Figs. 27 and 28, Supplementary Table 6). The output discharge current density increases continuously from 0.5 to 8 mA cm−2 and each step exhibits a stable discharging plateau, indicating that our cell has an excellent discharge capability (Supplementary Fig. 29). Moreover, the NH3 yield rate exhibits a maximum of 1627.67 µg h−1 mgcat−1 at 8 mA cm−2 (Supplementary Fig. 30). Significantly, the Zn-NO battery can remove NO pollutant, produce NH3 and generate electricity at the same time.
a LSV curves of Co/Co3O4 and Co/Co3O4/CoB with a scanning of 5 mV/s recorded in 0.1 M Ar-saturated and NO-saturated PBS (pH = 7.3 ± 0.1, with 90% iR-correction). b FENH3 and NH3 yield rate over Co/Co3O4/CoB at various cathodic potentials with a mass loading of 0.2 mg cm−2 at 25 °C. c Comparison of NORR performance between Co/Co3O4 and Co/Co3O4/CoB at -0.5 V vs. RHE. d NORR performance over Co/Co3O4/CoB and carbon paper (CP) under different conditions (O/C represents open circuit condition). The inset shows the 1H nuclear magnetic resonance spectra to quantify NH4+. e Stability test of Co/Co3O4/CoB at −0.5 V vs. RHE. f Comparison of electrocatalytic NORR performance of Co/Co3O4/CoB with other reported electrocatalysts in literature. The error bars denote the standard deviation derived from three independent measurements, with the central marker indicating the mean value. Source data for Fig. 3 are provided as a Source Data file.
Reaction mechanism
To understand the excellent NORR performance of Co/Co3O4/CoB, Brunauer-Emmett-Teller (BET) surface area and electrochemically active surface area (ECSA) of the as-prepared catalysts were first measured. Compared to Co/Co3O4, Co/Co3O4/CoB displays a higher BET surface area (Supplementary Fig. 31) and ECSA (Supplementary Fig. 32), which shall offer more active sites to catalyze electrochemical NORR. To clarify the reactive sites for NO adsorption/activation and H2O dissociation as well as the protonation process of adsorbed intermediates on Co/Co3O4/CoB, a series of DFT calculations, NO breakthrough and electron paramagnetic resonance (EPR) measurements were conducted. The NO adsorption energies with N-end/O-end on the Co sites of Co (111), Co3O4 (311), and CoB (021) and the B sites of CoB (021) were computed by DFT (Fig. 4a and Supplementary Figs. 33–35). The results indicate that NO prefers to adsorb on Co sites of Co3O4 (311) in the form of N-end adsorption, having the most negative adsorption energy of −3.64 eV. Compared to Co/Co3O4, Co/Co3O4/CoB binds to NO more strongly (Fig. 4b and Supplementary Data 1). The low-coordinated Co-O structure in Co/Co3O4/CoB (Fig. 2f and Supplementary Table 2) contributes to NO adsorption as demonstrated by NO breakthrough analysis (Fig. 4c)47. A crystal orbital Hamilton population (COHP) was further employed to investigate the NO adsorption strength on Co/Co3O4/CoB and Co/Co3O4 (Fig. 4d)50,51,52. Compared to Co/Co3O4, Co/Co3O4/CoB displays a more negative integrated COHP (ICOHP) value of the Co-N bond, indicating a stronger NO adsorption on the Co/Co3O4/CoB surface53. DFT calculation was further performed to explore the water dissociation process. The energy barrier for water transformation from the adsorbed state (H2O*) to the transition state (TS) on Co/Co3O4/CoB and Co/Co3O4 are 0.50 eV and 0.66 eV, respectively (Fig. 4e, Supplementary Figs. 36 and 37, Supplementary Table 6). This result indicates that the dissociation of H2O is more kinetically favorable on Co/Co3O4/CoB. In order to further investigate the H2O dissociation ability of Co/Co3O4/CoB and Co/Co3O4, in-situ Raman spectroscopy measurements were conducted at different applied cathodic potentials. Three types of O-H stretching modes belonging to 4-HB·H2O, 2-HB·H2O, and M·H2O (M = Na+/K+), respectively54, are observed. The relative proportion of M·H2O displays a little increasing trend along with increasing the applied cathodic potential, suggesting that M·H2O is closer to the catalyst’s surface and could be more affected by the electric field. The M·H2O proportion over Co/Co3O4/CoB is always higher than that over Co/Co3O4 at all studied cathodic potentials, suggesting better water dissociation ability over Co/Co3O4/CoB (Supplementary Figs. 38 and 39)55,56. Consequently, Co/Co3O4/CoB could generate more *H as reflected in the liquid EPR measurement using 5,5-dimethyl-1-pyrroline-N-oxude (DMPO) as the *H trapping regent (Fig. 4f)57. The DMPO-H signal completely disappeared when NO was introduced into the reaction system (Supplementary Fig. 40), indicating that the generated *H could be rapidly consumed by NO on the Co/Co3O4/CoB surface. This further verifies that *H is more favorable for bonding with NO to form NHx intermediates rather than generating H258. The *H generated on the metallic Co sites in Co/Co3O4/CoB can be quickly transferred to the neighboring Co3O4 surface to trigger the *NO hydrogenation reaction (Fig. 4g), thanks to the relatively electron-richer Co3O4 in Co/Co3O4/CoB than that in Co/Co3O4. As a result, the enhanced charge and *H transfer over Co/Co3O4/CoB promote electrochemical NO reduction to produce NH3. The COHP analysis was performed to investigate the adsorption capacity of nitrogen-containing intermediates on Co/Co3O4/CoB and Co/Co3O4. The ICOHP value of the Co-NH/Co-NH2 bond is −3.83/3.20 and −2.65/1.70 on Co/Co3O4/CoB and Co/Co3O4, respectively (Fig. 4h, i, Supplementary Figs. 41 and 42), indicating a higher bonding state energy of *NH/*NH2 on Co/Co3O4/CoB than Co/Co3O4. The projected density of states (PDOS) of Co/Co3O4/CoB and Co/Co3O4 upon adsorption of *NH and *NH2 were analyzed (Supplementary Figs. 43 and 44). The Ep (the highest peak below the Fermi level) position of Co/Co3O4/CoB is higher than that of Co/Co3O4 upon adsorption of *NH and *NH2, indicating that the antibonding states of NORR intermediates are located at higher energies with lower occupancies on Co/Co3O4/CoB surface, which results in stronger bindings of intermediates on Co/Co3O4/CoB during NORR59. Both the COHP and PDOS results indicate enhanced adsorption of *NH and *NH2 intermediates over Co/Co3O4/CoB, which is conducive to NORR for producing NH3.
a The calculated NO adsorption energy of nitrogen end on Co (111), CoB (021) and Co3O4 (311). b The calculated NO adsorption energy on Co/Co3O4/CoB and Co/Co3O4. The blue circle marks the NO molecule. Atom color-coding: Purple, cobalt; red, oxygen; pink, boron; nitrogen, orange. c NO adsorption breakthrough curves of Co/Co3O4/CoB, Co/Co3O4 and Co3O4 catalysts at 25 °C. d The COHP analysis of Co-N bond during NO adsorption on Co/Co3O4/CoB and Co/Co3O4. e Gibbs free energy for H2O dissociation on Co/Co3O4/CoB and Co/Co3O4. f EPR spectra recorded over Co/Co3O4/CoB, Co/Co3O4 and Co3O4 upon electrolysis in the absence of NO. g Comparison of Gibbs free energy for *H adsorption on Co and Co3O4 sites of Co/Co3O4/CoB and Co/Co3O4. The COHP analysis of h *NH and i *NH2 on Co/Co3O4/CoB and Co/Co3O4. Source data for Fig. 4 are provided as a Source Data file.
To reveal the reaction pathway of electrochemical NORR, in-situ attenuated total reflection surface enhanced infrared absorption spectroscopy (ATR-SEIRAS) measurements were conducted to probe the reaction intermediates in the potential range from OCP to −0.7 V (vs. RHE). Over the Co/Co3O4/CoB surface (Fig. 5a), the bands at 1103 and 3224 cm−1 belong to *HNO species while the bands at 1167, 1245, and 1504 cm−1 are related to N-H stretching, -NH2 wagging, and -H-N-H bending, respectively60,61. The gradually increased bands at 1456 and 1635 cm−1 can be attributed to NH4+ and -OH species accumulation on the Co/Co3O4/CoB surface62. Additionally, a wide overlapped absorption at around 3000 ~ 3800 cm−1 is observed, in which the bands at 3380 and 3731 cm−1 are assigned to the stretching of -OH and -NH, respectively63. The in-situ ATR-SEIRAS spectra as a function of time also display the same reaction intermediates over Co/Co3O4/CoB surface (Supplementary Fig. 45). Compared to Co/Co3O4 (Fig. 5b), a large number of important reactive intermediates including *NH and *NH2 can be observed over Co/Co3O4/CoB surface at less negative cathodic potentials, suggesting much enhanced NORR reaction kinetics on Co/Co3O4/CoB. On the other hand, a larger amount of *NH and *NH2 intermediates were observed accumulating on Co3O4 surface compared with Co/Co3O4 and Co/Co3O4/CoB, implying the weaker ability of supplying *H on Co3O4 (Supplementary Fig. 46). To further confirm the key reaction intermediates, the DEMS measurement was performed over Co/Co3O4/CoB at −0.5 V vs. RHE. As displayed in Fig. 5c and Supplementary Fig. 47, it can be observed that NO decreases with a decrease of NOH and HNOH intermediates, and meanwhile NH, NH2, and NH3 increase. These intermediates observed in DEMS align well with the results obtained from in-situ ATR-SEIRAS measurements. Therefore, the NORR pathway taking place on the Co/Co3O4/CoB surface is speculated to follow *NO → *HNO→ *NHOH → *NH → *NH2 → *NH3 → NH3. Additionally, the signal of NH3 in the DEMS measurement is much stronger than that of H2, N2, NH2OH, and N2O, explaining the high FENH3 in NORR over Co/Co3O4/CoB. Figure 5d compares the reaction energy diagram of NORR over Co/Co3O4/CoB and Co/Co3O4 surface (the structural models of intermediates are shown in Supplementary Figs. 48 and 49). Compared with *NOH, the first protonation of NO is more energetically favorable to form *HNO (Supplementary Figs. 50 and 51), which is the potential-determining step (PDS) of NORR on Co/Co3O4/CoB and Co/Co3O4. Clearly, Co/Co3O4/CoB shows a lower reaction energy barrier (0.81 eV) in PDS than that (1.11 eV) of Co/Co3O4, owing to the accelerated protonation process of NO over Co/Co3O4/CoB. The second proton preferentially reacts with *HNO to generate *HNOH with a ΔG of −1.88 eV, rather than *N with a ΔG of −0.82 eV. Finally, three successive protonation processes occur until NH3 is formed. Besides, the reaction energy of NORR was also calculated over Co3O4, Co, and CoB (Supplementary Figs. 52–55 and Table 7). The reaction energy of the potential-determining step (*NO to *HNO) is 1.25, 2.81 and 1.6 eV over Co3O4, Co, and CoB, respectively, much higher than that over Co/Co3O4/CoB. This result evidences that the first step of NO hydrogenation can be notably boosted over the ternary Co/Co3O4/CoB heterostructure.
In-situ ATR-SEIRAS spectra recorded over Co/Co3O4/CoB a and Co/Co3O4 b in 0.1 M PBS at the applied cathodic potential from OCP to −0.7 V vs. RHE. c Online DEMS measurement results over Co/Co3O4/CoB at −0.5 V vs. RHE. d The calculated NORR Gibbs free energy diagrams over Co/Co3O4 and Co/Co3O4/CoB at 0 V vs. RHE. Atom color-coding: Purple, cobalt; red, oxygen; orange, nitrogen; gray, hydrogen. e Schematic illustration showing the NORR mechanism over Co/Co3O4/CoB electrocatalyst. Atom color-coding: Light blue, cobalt; pink, boron; red, oxygen; deep blue, nitrogen; gray, hydrogen. Source data for Fig. 5 are provided as a Source Data file.
Discussion
In summary, we have designed and prepared a ternary Co/Co3O4/CoB heterostructure through a thermal solid-state reduction reaction of Co3O4 by NaBH4. The prepared Co/Co3O4/CoB displayed a high FENH3 of 98.80% with NH3 yield rate of 462.18 µmol cm−2 h−1 (2.31 mol h−1 gcat−1) and long-time stability at a low cathodic potential of −0.5 V vs. RHE, outperforming most of the reported NORR electrocatalysts in the literature. The extraordinary NORR performance of Co/Co3O4/CoB was resulted from the unique charge and mass transfer within the Co/Co3O4/CoB heterostructure (Fig. 5e). The enhanced electron transfer in Co/Co3O4/CoB yielded electron-deficient Co and electron-rich Co3O4. The electron-deficient Co sites could effectively boost H2O dissociation to generate *H while the electron-rich low coordination Co3O4 sites promoted NO adsorption. The *H formed on electron-deficient Co site was more favorable to transfer to electron-rich Co3O4 site adsorbed with NO, facilitating the selective hydrogenation of NO. Thanks to the enhanced charge and proton transfer, the energy barrier of the potential-determining step from *NO to *HNO in NORR over Co/Co3O4/CoB was significantly reduced. Moreover, the introduction of CoB in Co/Co3O4/CoB could also facilitate the adsorption of *NH and *NH2 intermediates. All of which greatly promoted electrochemical NORR. This study paves the way for designing and developing highly efficient electrocatalysts for reducing NO to NH3.
Methods
Reagents and materials
Carbon paper (99.99%) and Ketjen carbon were purchased from Suzhou Sinero Technology Co., Ltd. Cobalt(II) acetate tetrahydrate (Co(CH3COO)2.4H2O, ≥99%), polyvinyl pyrrolidone (PVP, ≥99%), sodium borohydride (NaBH4, ≥98.0%), chloride (NH4Cl, 99.99%), ethanol (C2H5OH, ≥99.5%), were purchased from Shanghai Titan Scientific Co., Ltd. Hydroxylammonium chlorideammonium (NH2OH·HCl, ≥99%), dimethyl Sulfoxide-D6 (C2D6OS, 99.8%), sodium hydroxide (NaOH, ≥98%), trisodium citrate (C6H5Na3O7, 98%), sodium nitroprusside dihydrate (C5H4FeN6Na2O3·H2O, ≥99.98%), sulfuric acid (H2SO4, >95%), hydrogen peroxide (H2O2, 30%), and phosphate buffer solution (PBS, 10X, pH = 7.3 ± 0.1) were purchased from Shanghai Adamas Reagents Co., Ltd. Sodium hypochlorite solution (NaClO, available chlorine 4.0%), salicylic acid (C7H6O3, 99%) were purchased from Macklin Biochemical Technology Co., Ltd. Deionized water (18.25 MΩ cm resistivity) was obtained via an ultrapure water equipment in laboratory. All the reagents are analytical-grade and directly used without further purification.
Synthesis of Co3O4 nanosheets
The Co3O4 nanosheets were prepared by a hydrothermal method as follows: Co(CH3COO)2.4H2O (0.75 g) and PVP (2.4 g) were dissolved in 70 mL methanol under stirring for 30 min. The mixture was transferred into a 100 mL Teflon-lined stainless-steel autoclave that was sealed and heated to 190 °C for 12 h. After cooling to 25 °C, the products were harvested by centrifugation, washed three times with ethanol, dried at 60 °C, and calcined at 350 °C for 5 h in air64.
Synthesis of Co/Co3O4/CoB and Co/Co3O4
In a typical synthesis of Co/Co3O4/CoB, Co3O4 and NaBH4 were uniformly mixed in a mortar at a mass ratio of 1:1.5. The mixture was then placed in a horizontal quartz tube and heated to 500 °C at a ramp rate of 2 °C min−1 under an Ar atmosphere, which was maintained at 500 °C for 4 h. The resulting products were washed extensively by an ethanol/deionized water mixture and collected by filtration. Finally, the washed Co/Co3O4/CoB was dried at 80 °C for 12 h in a vacuum oven. For preparing Co/Co3O4, the mass ratio of Co3O4 and NaBH4 was adjusted to 1.5:1, while other preparation parameters were kept unchanged.
Electrochemical measurements
Electrochemical measurements were performed on a CHI660E electrochemical workstation in an H-type electrochemical cell (Supplementary Fig. 56) separated by a Nafion 117 membrane (10*10 cm, thickness was 183 µm, Dupont) in a three-electrode configuration at 25 °C. We chose 0.1 M PBS buffer solution (pH = 7.3 ± 0.1) to maintain the pH of the electrolyte during NORR. The membrane was sequentially treated in a H2O2 (5 wt.%) aqueous solution at 80 °C for 1 h, then in a 0.5 M H2SO4 solution at 80 °C for 2 h and finally in deionized water at 25 °C for 6 h. The catalyst coated on carbon paper (CP,1 cm2) was used as the working electrode, a platinum plate (1 cm2) was used as the counter electrode, and an Ag/AgCl (saturated 3.5 M KCl aqueous solution) electrode was used as the reference electrode. The Ag/AgCl reference electrode was calibrated in H2-saturated 0.1 M PBS using a symmetric Pt electrode system.
To prepare the working electrode, 2.5 mg of electrocatalyst and 2.5 mg of Ketjen carbon were ultrasonically dispersed in 300 μL ethanol, 150 μL ultrapure water, and 25 μL Nafion solution (5 wt.%, Du Pont) to form a homogeneous catalyst ink followed by dropping 38 µL of the catalyst ink onto a piece of CP (1 cm2) that was dried at 25 °C, the catalyst mass loading was 0.2 mg cm−2. Before all electrochemical tests, 30 min of high-purity Ar gas (99.999%) and 30 min of NO/Ar gas (10 vol.%) saturated the electrolyte to exclude air in the reaction system.
Linear sweep voltammetry (LSV) was performed at a scanning rate of 5 mV/s prior to 50 cycles of cyclic voltammetry at a scan rate of 50 mV/s to obtain a stable curve with 90% iR-correction. The non-iR corrected data for all catalysts are provided in the Supplementary Information. The electrochemical impedance spectroscopy (EIS) was obtained without iR-correction in the frequency range from 0.01 Hz to 100 kHz upon an AC voltage amplitude of 5 mV at an open-circuit potential under 25 °C. The chronoamperometry was operated to evaluate the stability under continuous stirring (1600 rpm) at different current densities. All potentials in this study were converted to the reversible hydrogen electrode (RHE) scale according to the following equation:
Calculation of Faradaic efficiency (FE) and yield rate (Y)
Where Qi is the charge of product i, Qtotal is the total charge, ci is the concentration of products (µg mL−1), V is the volume of electrolyte in the cathode compartment (25 mL), n: the number of electron transfer in products; F: Faraday constant (96485 C mol−1); Mi: the molar mass of products (g mol−1); s: the area of the electrode (1 cm2); t: the reaction time (1 h).
ECSA analysis
Electrochemical active surface area (ECSA) was evaluated through cyclic voltammetry (CV) measurements conducted within the non-Faradaic potential region in 0.1 M PBS electrolyte, where iR drop is negligible. The electrochemical double-layer capacitance (Cdl) was determined from the linear slope of capacitive current (Δj = 0.5 × |jcharge − jdischarge|) plotted against scan rate (20–120 mV s−1). The ECSA of the catalyst was determined by normalizing the double-layer capacitance (Cdl) against the standard specific capacitance (Cs) for smooth metal surfaces in 1.0 M netural electrolyte, following the relation: ECSA = Cdl/Cs.
Quantification of NH3
The yield rate of NH3 in NORR was quantitatively determined by the indophenol blue method and NMR. For the indophenol blue method, the concentration-absorbance curves were calibrated using a standard NH3 solution with a series of concentrations. The fitting curve (y = 0.4149x – 0.0264, R2 = 0.9999) shows a good linear relationship between absorbance and NH3 concentration. For NMR measurement, after NORR, the electrolyte was taken out, whose pH was adjusted to 3 using a 0.5 M HCl aqueous solution, and then sent for quantification by 1H NMR (600 MHz) with internal standard of maleic acid (3.5 M). The number of scans was 128 for all NMR measurements. The fitting curve (y = 2.7375x – 1.9166, R2 = 0.9993) shows a good linear relationship.
15N isotope labeling experiment
The 15N isotope labeling experiment proceeded on the 1H NMR spectroscopy (600 MHz) to identify the nitrogen source for NORR. Replace 14NO with 15NO to conduct the NORR experiment at −0.5 V vs. RHE. After the experiment, take 2 mL of the catholyte were acidized by H2SO4 (2 mL, 1 M). 60 μL of the above solutions were respectively added to a NMR tube and mixed with maleic acid aqueous solution (20 μL, 3.6 mM), H2SO4 aqueous solution (20 μL, 4 M) and DMSO-d6 (500 μL), and then analyzed by 1H NMR measurements.
Quantification of hydroxylamine (NH2OH)
To quantify NH2OH, the following literature procedures were executed: 1.0 mL of the catholyte (or 0.1 mM~0.5 mM standard NH2OH solution) was added 1.0 mL PBS buffer (pH = 7.3 ± 0.1) and 1.0 mL 1% 8-hydroxylquinoline. Under vigorous shaking, 1.0 mL 0.1 M Na2CO3 was added and the mixture was heated at 100 °C for 1 min. In the presence of NH2OH, the solution would turn from light yellow to blue-green, showing an absorption peak at ~705 nm. Plotting the peak absorbance with NH2OH concentration yielded a calibration curve.
Zn-NO battery
A Co/Co3O4/CoB coated on carbon paper was employed as the cathode to perform the NORR in a cathodic electrolyte (0.1 M PBS). A polished Zn plate was applied as the anode in an anodic electrolyte (1 M KOH), and a Nafion 117 membrane was used to separate the cathode from the anode. The Zn-NO battery was assessed on a CHI660E electrochemical workstation under an ambient atmosphere at 25 °C.
Characterizations
The morphological information was characterized by atomic force microscopy (AFM, Park NX10), field emission scanning electron microscopy (SEM, Zeiss-sigma-300), and transmission electron microscopy (TEM, JEOL JEM-1400). High-resolution transmission electron microscopy (HRTEM) and aberration-corrected high-angle annular dark-field scanning transmission electron microscopy (HAADF-STEM) were conducted on a JEOL JEM 2100 F and JEM-ARM 300 F Grand ARM. X-ray diffraction (XRD) patterns were recorded on a Bruker D8 Advance Diffractometer (Cu-Kα radiation: λ = 0.15406 nm). The surface valence states were studied by X-ray photoelectron spectroscopy (XPS, PHI-5300) with Mg Kα radiation (hν = 1486.6 eV). The X-ray absorption near edge structure (XANES) and extended X-ray absorption fine structure (EXAFS) spectra at the Co K-edge were collected at the HXMA beamline of the Canadian Light Source (CLS) using Fluorescence mode and Si (111) monochromator. The samples were pressed into wafers for collecting the data at 25 °C. The EXAFS raw data were background subtracted, normalized and Fourier transformed by standard procedures using the ATHENA program. The acquired EXAFS data were processed according to the standard procedures using the ATHENA module implemented in the IFEFFIT software packages. The k3-weighted EXAFS spectra were obtained by subtracting the post-edge background from the overall absorption and then normalizing for the edge-jump step. The ultraviolet-visible (UV‒vis) absorbance spectra were measured on an Agilent S3 Cary 5000. A Bruker 600 M NMR instrument with water suppression was used to record the 1H NMR spectra. The Brunauer-Emmett-Teller (BET) surface area was measured on a U.S. Quantachrome ASAP 2020 M at 77 K.
NO adsorption breakthrough measurements
The NO adsorption breakthrough measurements were carried out on a dynamic sorption analyzer (mixSorb S) equipped with a thermal conductivity detector (TCD) combined with a mass spectrometer (OMNISTARTM). Firstly, 40 mg of catalyst was pre-treated in a high purity He atmosphere at a flow rate of 30 mL/min for 1 h. Afterwards, the sample was exposed to NO gas (0.02 vol.% NO balanced by He) at a total gas flow rate of 20 mL/min at 25 °C accompanied by recording the TCD signal and the MS signal of NO at m/z = 30 until the outlet mass signal of NO achieved saturation.
In-situ attenuated total reflection-surface enhanced infrared absorption spectroscopy (ATR-SEIRAS) measurements
In-situ ATR-SEIRAS spectra were recorded on an INVENIO-R FTIR spectrometer (Bruker) equipped with a liquid nitrogen-cooled mercury cadmium telluride (MCT) detector (Supplementary Fig. 57). The catalyst ink was prepared by mixing 5 mg of catalyst, 300 μL ethanol, 150 μL ultrapure water, and 25 μL Nafion solution under sonication for 30 min. Next, 40 μL catalyst ink was slowly dropped onto a face-angled Si crystal to prepare the working electrode of the custom-made spectroelectrochemical cell fixed on the ATR accessory. Ag/AgCl electrode was employed as the reference electrode and a Pt wire was used as the counter electrode. Electrolyte was added in advance with continuous Ar flow at a flow rate of 50 mL/min for 30 minutes to remove interference of H2O and O2 prior to ATR-SEIRAS measurement. The background was measured at open circuit potential. Subsequently, 10% NO/Ar gas was injected and the absorption spectra were recorded at different applied potentials from OCP to −0.7 V vs. RHE. Afterwards, spectrum was collected at −0.5 V vs. RHE every 5 min for 60 min.
Online differential electrochemical mass spectroscopy (DEMS) measurements
In-situ DEMS measurement was conducted on Linglu (Supplementary Fig. 58). A Teflon layer was applied to separate the electrolyte from the vacuum system. The vacuum system, comprising of two dry pumps and one turbo pump, kept the vacuum degree below 10−7 Pa. The catalyst ink was prepared by mixing 3 mg of catalyst, 460 μL ethanol, 500 μL ultrapure water, and 40 μL Nafion solution under sonication for 30 min. Next, 40 μL catalyst ink was slowly dropped onto gold film to prepare the working electrode. Ag/AgCl electrode was employed as the reference electrode and a Pt wire was used as the counter electrode. Before the DEMS measurement, NO/Ar was bubbled into the electrolyte until saturation. I-t test for 100 s was used to conduct DEMS investigations.
In-situ Raman spectroscopy
The in-situ Raman measurements were carried out jointly by an RXN1 Raman instrument (KAISER OPTICAL SYSTEM) and a CHI660E electrochemical workstation. A custom-made spectroelectrochemical cell was used as a reactor to enable the in-situ measurements. The obtained Co/Co₃O₄/CoB catalyst, Ag/AgCl, and platinum wire served as the working electrode, reference electrode, and counter electrode, respectively. The working electrode was immersed in the electrolyte and positioned such that the electrode plane was perpendicular to the laser. In-situ Raman spectra were obtained while the electrodes were under potentiostatic control. The experiment was conducted for 200 s under each fixed potential.
EPR measurements
5,5-dimethyl-1-pyrroline N-oxide (DMPO) was used to capture the unstable hydrogen radical to form the DMPO-H adduct to generate EPR spectrum. Briefly, 5 mL of electrolyte was mixed with 100 μL of DMPO and the mixture was deoxygenated by bubbling Ar. The constant current electrolysis was carried out for 10 min in the H-type cell under the protection of Ar. EPR measurement was performed on a JES X320, JEOL Co. spectrometer operated at a frequency near 9.5 GHz, sweep width of 200 G, and power of 20 mW. Co/Co3O4/CoB fitting parameters: g = 2.0045, AN = 15.5 G, AH = 21.0 G, lwpp = 0.3. Co/Co3O4 fitting parameters: g = 2.0043, AN = 15.6 G, AH = 20.8 G, lwpp = 0.3.
Density functional theory (DFT) calculations
Considering the limitations of computational power, we opted for a simplified mode, namely CoB supported neighboring Co and Co3O4 clusters, combininng with the morphology and three interfaces of Co/Co3O4/CoB heterostructure. The choice of using crystal facets of Co (111), Co3O4 (311), and CoB (021) to model Co/Co3O4/CoB heterostructure is primarily based on the XRD characterizations. These crystal facets exhibit the strongest diffraction in the XRD patterns and are also the thermodynamically stable facets. It is noted that the Co (111) and Co3O4 (311) clusters are taken from (111) facet of Co single crystal and (311) facet of Co3O4 single crystal, respectively.
All calculations were performed in the framework of the density functional theory with the projector augmented plane-wave method, as implemented in the Vienna ab initio simulation package (VASP)65. The generalized gradient approximation proposed by Perdew, Burke, and Ernzerhof was selected for the exchange-correlation potential66. The Grimme D3 correction used a coordination number dependent dispersion correction67. The cut-off energy for the plane wave was set to 450 eV. The energy criterion was set to 10−5 eV in the iterative solution of the Kohn-Sham equation. A vacuum layer of 15 Å was added perpendicular to the sheet to avoid artificial interaction between periodic images. The Brillouin zone integration was performed using a 2 × 2 × 1 k-mesh68. All the structures were relaxed until the residual forces on the atoms had declined to less than 0.05 eV/Å. The maximum atomic force in Co/Co3O4/CoB is 0.047, which meets the convergence requirement. In addition, we also evaluated the structural stability by using ab initio molecular dynamics (AIMD). The Brillouin zone was sampled using 1 × 1 × 1 k-point grid. Self-consistent calculations were conducted with an energy convergence threshold of 10−5 eV. The AIMD was performed within the canonical (NVT) ensemble by Nosé-Hoover thermostats with a time step of 1.0 fs at a finite temperature of 300 K69. In the AIMD simulation up to 10 ps, the heterostructure of Co/Co3O4/CoB was not destroyed, and the basic crystal structure was stable, proving the rationality of the structure (Supplementary Figs. 59–61 and Supplementary Data 2). Spin polarization was included in the calculations and the default setting of magnetic moment (MAGMOM= number of atoms of Co*1.0) was chosen for all calculations. The Gibbs free energy (ΔG) of reaction intermediates was calculated by the following:
where ΔE is the adsorption energy. ΔEZPE and ΔS are the difference for the zero-point energy and entropy, respectively. The zero-point energy and entropy were calculated at the standard conditions corresponding to the pressure of 101325 Pa ( ~ 1 bar) of H2 at the temperature of 298.15 K. The climbing image nudged elastic band (cNEB) method was used to search the reaction path and transition state, and the vibration frequency calculation was used to confirm it further70.
Data availability
Source data for all the figures and tables generated in this study are provided as a Source Data file. Source data are provided with this paper.
References
Li, K. et al. Enhancement of lithium-mediated ammonia synthesis by addition of oxygen. Science 374, 1593–1597 (2021).
Tayyebi, A. et al. Bias-free solar NH3 production by perovskite-based photocathode coupled to valorization of glycerol. Nat. Catal. 7, 510–521 (2024).
Mehta, P. et al. Overcoming ammonia synthesis scaling relations with plasma-enabled catalysis. Nat. Catal. 1, 269–275 (2018).
Fu, X. et al. Continuous-flow electrosynthesis of ammonia by nitrogen reduction and hydrogen oxidation. Science 379, 707–712 (2023).
Suryanto, B. H. R. et al. Challenges and prospects in the catalysis of electroreduction of nitrogen to ammonia. Nat. Catal. 2, 290–296 (2019).
Weng, G. et al. A high-efficiency electrochemical proton-conducting membrane reactor for ammonia production at intermediate temperatures. Joule 7, 1333–1346 (2023).
Hawtof, R. et al. Catalyst-free, highly selective synthesis of ammonia from nitrogen and water by a plasma electrolytic system. Sci. Adv. 5, eaat5778 (2019).
Zhao, X. et al. Comprehensive understanding of the thriving ambient electrochemical nitrogen reduction reaction. Adv. Mater. 33, 2007650 (2021).
Han, S. et al. Ultralow overpotential nitrate reduction to ammonia via a three-step relay mechanism. Nat. Catal. 6, 402–414 (2023).
Zhang, R. et al. A Zn–nitrite battery as an energy-output electrocatalytic system for high-efficiency ammonia synthesis using carbon-doped cobalt oxide nanotubes. Energy Environ. Sci. 15, 3024–3032 (2022).
Pan, D. et al. Integrated electrocatalytic synthesis of ammonium nitrate from dilute NO gas on metal organic frameworks-modified gas diffusion electrodes. Nat. Commun. 15, 7243 (2024).
Shen, H. et al. Electrochemical ammonia synthesis: mechanistic understanding and catalyst design. Chem 7, 1708–1754 (2021).
Ren, Y. et al. Strategies to suppress hydrogen evolution for highly selective electrocatalytic nitrogen reduction: challenges and perspectives. Energy Environ. Sci. 14, 1176–1193 (2021).
Li, Y. et al. Advanced electrocatalysis for energy and environmental sustainability via water and nitrogen reactions. Adv. Mater. 33, 2000381 (2021).
Chen, G.-F. et al. Electrochemical reduction of nitrate to ammonia via direct eight-electron transfer using a copper–molecular solid catalyst. Nat. Energy 5, 605–613 (2020).
Li, S. et al. ITO@TiO2 nanoarray: an efficient and robust nitrite reduction reaction electrocatalyst toward NH3 production under ambient conditions. eScience 2, 382–388 (2022).
Long, J. et al. Direct electrochemical ammonia synthesis from nitric oxide. Angew. Chem. Int. Ed. 59, 9711–9718 (2020).
Li, Y. et al. Electrocatalytic reduction of low-concentration nitric oxide into ammonia over Ru nanosheets. ACS Energy Lett. 7, 1187–1194 (2022).
Liu, Y., Wang, Y., Liu, Z. & Wang, Q. Oxidation removal of nitric oxide from flue gas using UV photolysis of aqueous hypochlorite. Environ. Sci. Technol. 51, 11950–11959 (2017).
Wang, D. et al. Coupling electrocatalytic nitric oxide oxidation over carbon cloth with hydrogen evolution reaction for nitrate synthesis. Angew. Chem. Int. Ed. 60, 24605–24611 (2021).
Zhang, R. et al. Electrochemical nitrate reduction in acid enables high-efficiency ammonia synthesis and high-voltage pollutes-based fuel cells. Nat. Commun. 14, 8036 (2023).
Liang, J. et al. Amorphous boron carbide on titanium dioxide nanobelt arrays for high-efficiency electrocatalytic NO reduction to NH3. Angew. Chem. Int. Ed. 61, e202202087 (2022).
Meng, J., Cheng, C., Wang, Y., Yu, Y. & Zhang, B. Carbon support enhanced mass transfer and metal–support interaction promoted activation for low-concentrated nitric oxide electroreduction to ammonia. J. Am. Chem. Soc. 146, 10044–10051 (2024).
Zhang, H. et al. Isolated electron-rich ruthenium atoms in intermetallic compounds for boosting electrochemical nitric oxide reduction to ammonia. Angew. Chem. Int. Ed. 62, e202213351 (2023).
Guo, X. et al. Aqueous electroreduction of nitric oxide to ammonia at low concentration via vacancy engineered FeOCl. Angew. Chem. Int. Ed. 63, e202318792 (2024).
Liu, P. et al. High-performance NH3 production via NO electroreduction over a NiO nanosheet array. Chem. Commun. 57, 13562–13565 (2021).
Wang, D. et al. Hexagonal cobalt nanosheets for high-performance electrocatalytic NO reduction to NH3. J. Am. Chem. Soc. 145, 6899–6904 (2023).
Chen, K., Wang, G., Guo, Y., Ma, D. & Chu, K. Iridium single-atom catalyst for highly efficient NO electroreduction to NH3. Nano Res. 16, 8737–8742 (2023).
Chen, K., Wang, J., Zhang, H., Ma, D. & Chu, K. Self-tandem electrocatalytic NO reduction to NH3 on a W single-atom catalyst. Nano Lett. 23, 1735–1742 (2023).
Chen, K. et al. p-Block antimony single-atom catalysts for nitric oxide electroreduction to ammonia. ACS Energy Lett. 8, 1281–1288 (2023).
Ma, L. et al. Catalytic performance and reaction mechanism of NO oxidation over Co3O4 catalysts. Appl. Catal., B 267, 118371 (2020).
Yung, M. M., Holmgreen, E. M. & Ozkan, U. S. Cobalt-based catalysts supported on titania and zirconia for the oxidation of nitric oxide to nitrogen dioxide. J. Catal. 247, 356–367 (2007).
Liu, D. et al. ZIF-67-derived 3D hollow mesoporous crystalline Co3O4 wrapped by 2D g-C3N4 nanosheets for photocatalytic removal of nitric oxide. Small 15, 1902291 (2019).
Cao, H. et al. Rational design of Cu-doped tetrahedron of spinel oxide for high-performance nitric oxide electrochemical sensor. ACS Appl. Mater. Interfaces 15, 23489–23500 (2023).
Yang, B. et al. Electron-deficient cobalt nanocrystals for promoted nitrate electrocatalytic reduction to synthesize ammonia. Nano Energy 117, 108901 (2023).
Liu, Y. et al. Cobalt-based Co3Mo3N/Co4N/Co metallic heterostructure as a highly active electrocatalyst for alkaline overall water splitting. Angew. Chem. Int. Ed. 63, e202319239 (2024).
Zhang, L. et al. High-performance electrochemical NO reduction into NH3 by MoS2 nanosheet. Angew. Chem. Int. Ed. 60, 25263–25268 (2021).
Wu, B. et al. Boron-modulated electronic-configuration tuning of cobalt for enhanced nitric oxide fixation to ammonia. Nano Lett. 23, 7120–7128 (2023).
Xiao, L. et al. Identification of Cu(111) as superior active sites for electrocatalytic NO reduction to NH3 with high single-pass conversion efficiency. Angew. Chem. Int. Ed. 63, e202319135 (2024).
Garrido-Barros, P., Derosa, J., Chalkley, M. J. & Peters, J. C. Tandem electrocatalytic N2 fixation via proton-coupled electron transfer. Nature 609, 71–76 (2022).
Chalkley, M. J., Garrido-Barros, P. & Peters, J. C. A molecular mediator for reductive concerted proton-electron transfers via electrocatalysis. Science 369, 850–854 (2020).
Dai, J. et al. Cu(II) ions induced structural transformation of cobalt selenides for remarkable enhancement in oxygen/hydrogen electrocatalysis. ACS Catal. 9, 10761–10772 (2019).
Song, D. et al. Acidic media regulated hierarchical cobalt compounds with phosphorous doping as water splitting electrocatalysts. Adv. Energy Mater. 11, 2100358 (2021).
Li, H. et al. Earthworm-inspired Co/Co3O4/CoF2@NSC nanofibrous electrocatalyst with confined channels for enhanced ORR/OER performance. Adv. Mater. 36, 2311272 (2024).
Jiang, B. et al. A mesoporous non-precious metal boride system: synthesis of mesoporous cobalt boride by strictly controlled chemical reduction. Chem. Sci. 11, 791–796 (2020).
Guo, S. et al. Electrocatalytic hydrogenation of quinolines with water over a fluorine-modified cobalt catalyst. Nat. Catal. 13, 5297 (2022).
Lin, Q. et al. Tuning the interface of Co1–xS/Co(OH)F by atomic replacement strategy toward high-performance electrocatalytic oxygen evolution. ACS Nano 16, 15460–15470 (2022).
Jakub, Z. et al. Local structure and coordination define adsorption in a model Ir1/Fe3O4 single-atom catalyst. Angew. Chem. Int. Ed. 58, 13961–13968 (2019).
Yu, P. et al. Co Nanoislands rooted on Co–N–C nanosheets as efficient oxygen electrocatalyst for Zn–air batteries. Adv. Mater. 31, 1901666 (2019).
Li, M. et al. Proximity electronic effect of Ni/Co diatomic sites for synergistic promotion of electrocatalytic oxygen reduction and hydrogen evolution. Adv. Funct. Mater. 33, 2210867 (2023).
Gu, Y. et al. Boosting selective nitrogen reduction via geometric coordination engineering on single-tungsten-atom catalysts. Adv. Mater. 33, 2100429 (2021).
Yue, X., Cheng, L., Li, F., Fan, J. & Xiang, Q. Highly strained Bi-MOF on bismuth oxyhalide support with tailored intermediate adsorption/desorption capability for robust CO2 photoreduction. Angew. Chem. Int. Ed. 61, e202208414 (2022).
Dong, G., Huang, X. & Bi, Y. Anchoring black phosphorus quantum dots on Fe-doped W18O49 nanowires for efficient photocatalytic nitrogen fixation. Angew. Chem. Int. Ed. 61, e202204271 (2022).
Wang, Y.-H. et al. In situ Raman spectroscopy reveals the structure and dissociation of interfacial water. Nature 600, 81–85 (2021).
Chen, X. et al. Revealing the role of interfacial water and key intermediates at ruthenium surfaces in the alkaline hydrogen evolution reaction. Nat. Commun. 14, 5289 (2023).
Ni, J. et al. Atomic Co─P catalytic pair drives efficient electrochemical nitrate reduction to ammonia. Adv. Energy Mater. 14, 2400065 (2024).
Li, J. et al. Efficient ammonia electrosynthesis from nitrate on strained ruthenium nanoclusters. J. Am. Chem. Soc. 142, 7036–7046 (2020).
Liu, H. et al. Efficient electrochemical nitrate reduction to ammonia with copper-supported rhodium cluster and single-atom catalysts. Angew. Chem. Int. Ed. 61, e202202556 (2022).
Chu, K., Li, Q. -q, Liu, Y. -p, Wang, J. & Cheng, Y. -h Filling the nitrogen vacancies with sulphur dopants in graphitic C3N4 for efficient and robust electrocatalytic nitrogen reduction. Appl. Catal. B. 267, 118693 (2020).
Li, Y. et al. Local spin-state tuning of iron single-atom electrocatalyst by S-coordinated doping for kinetics-boosted ammonia synthesis. Adv. Mater. 34, 2202240 (2022).
Liu, S. et al. Proton-filtering covalent organic frameworks with superior nitrogen penetration flux promote ambient ammonia synthesis. Nat. Catal. 4, 322–331 (2021).
Li, P. et al. Visible-light-driven nitrogen fixation catalyzed by Bi5O7Br nanostructures: enhanced performance by oxygen vacancies. J. Am. Chem. Soc. 142, 12430–12439 (2020).
Li, H., Shang, J., Ai, Z. & Zhang, L. Efficient visible light nitrogen fixation with BiOBr nanosheets of oxygen vacancies on the exposed {001} facets. J. Am. Chem. Soc. 137, 6393–6399 (2015).
Qin, C. et al. Crystal-plane-dependent Fischer–Tropsch performance of cobalt catalysts. ACS Catal. 8, 9447–9455 (2018).
Kresse, G. & Joubert, D. From ultrasoft pseudopotentials to the projector augmented-wave method. Phys. Rev. B Condens. Matter 59, 1758–1775 (1999).
Perdew, J. P., Burke, K. & Ernzerhof, M. Generalized gradient approximation made simple. Phys. Rev. Lett. 77, 3865–3868 (1996).
Grimme, S., Antony, J., Ehrlich, S. & Krieg, H. A consistent and accurate ab initio parametrization of density functional dispersion correction (DFT-D) for the 94 elements H-Pu. J. Chem. Phys. 132, 154104 (2010).
Monkhorst, H. J. & Pack, J. D. Special points for Brillouin-zone integrations. Phys. Rev. B 13, 5188–5192 (1976).
Gerry, C. C. Coherent states and the Kepler-Coulomb problem. Phys. Rev. A 33, 6–11 (1986).
Henkelman, G., Uberuaga, B. P. & Jónsson, H. A climbing image nudged elastic band method for finding saddle points and minimum energy paths. J. Chem. Phys. 113, 9901–9904 (2000).
Acknowledgements
The authors gratefully acknowledge financial support from the National Natural Science Foundation of China (22125604 and 22436003 to D.S.Z.), Shanghai Rising-Star Program (22QA1403700 to L.P.H.) and Science and Technology Commission of Shanghai Municipality (23230713700 to D.S.Z.), the City University of Hong Kong startup fund (9020003 to B.L.), ITF–RTH - Global STEM Professorship (9446006 to B.L.), and JC STEM lab of Advanced CO2 Upcycling (9228005 to B.L.). X.Y.W. acknowledges the receipt of support from the CLSI Student Travel Support Program. XAFS spectra were tested at the HXMA beamline of the Canadian Light Source (CLS), a national research facility of the University of Saskatchewan, which is supported by the Canada Foundation for Innovation (CFI), the Natural Sciences and Engineering Research Council (NSERC), the National Research Council (NRC), and the Government of Saskatchewan. The authors thank Shenzhen HUASUAN Technology Co., Ltd for assistance on DFT calculations (https://huasuankeji.com).
Author information
Authors and Affiliations
Contributions
X.X.F. performed the experiments, analyzed the data, and wrote the first draft of the manuscript. Z.Y.T., Y.J.S., W.Q.Q., J.L.S., Z.L.W. and H.Y.D. helped with the measurements and analysis, discussed and revised the manuscript. X.Y.W. and Y.A.W. performed the XAFS measurements and analysis. L.P.H., B.L. and D.S.Z. supervised the project, manuscript preparation and discussion.
Corresponding authors
Ethics declarations
Competing interests
The authors declare no competing interests.
Peer review
Peer review information
Nature Communications thanks Liang Chu, and the other, anonymous, reviewer(s) for their contribution to the peer review of this work. A peer review file is available.
Additional information
Publisher’s note Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.
Source data
Rights and permissions
Open Access This article is licensed under a Creative Commons Attribution 4.0 International License, which permits use, sharing, adaptation, distribution and reproduction in any medium or format, as long as you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons licence, and indicate if changes were made. The images or other third party material in this article are included in the article's Creative Commons licence, unless indicated otherwise in a credit line to the material. If material is not included in the article's Creative Commons licence and your intended use is not permitted by statutory regulation or exceeds the permitted use, you will need to obtain permission directly from the copyright holder. To view a copy of this licence, visit http://creativecommons.org/licenses/by/4.0/.
About this article
Cite this article
Fan, X., Teng, Z., Han, L. et al. Boosted charge and proton transfer over ternary Co/Co3O4/CoB for electrochemical nitric oxide reduction to ammonia. Nat Commun 16, 4874 (2025). https://doi.org/10.1038/s41467-025-60043-6
Received:
Accepted:
Published:
DOI: https://doi.org/10.1038/s41467-025-60043-6