Skip to main content
Advanced Science logoLink to Advanced Science
. 2024 Sep 3;11(40):2407301. doi: 10.1002/advs.202407301

Spin Manipulation of Co sites in Co9S8/Nb2CTx Mott–Schottky Heterojunction for Boosting the Electrocatalytic Nitrogen Reduction Reaction

Shuai Zhang 1, Weihua Zhao 1, Jiameng Liu 2, Zheng Tao 1, Yinpeng Zhang 1, Shuangrun Zhao 1, Zhihong Zhang 1,, Miao Du 1,
PMCID: PMC11516103  PMID: 39225309

Abstract

Regulating the adsorption of an intermediate on an electrocatalyst by manipulating the electron spin state of the transition metal is of great significance for promoting the activation of inert nitrogen molecules (N2) during the electrocatalytic nitrogen reduction reaction (eNRR). However, achieving this remains challenging. Herein, a novel 2D/2D Mott–Schottky heterojunction, Co9S8/Nb2CTx‐P, is developed as an eNRR catalyst. This is achieved through the in situ growth of cobalt sulfide (Co9S8) nanosheets over a Nb2CTx MXene using a solution plasma modification method. Transformation of the Co spin state from low (t2g 6eg 1) to high (t2g 5eg 2) is achieved by adjusting the interface electronic structure and sulfur vacancy of Co9S8/Nb2CTx‐P. The adsorption ability of N2 is optimized through high spin Co(II) with more unpaired electrons, significantly accelerating the *N2→*NNH kinetic process. The Co9S8/Nb2CTx‐P exhibits a high NH3 yield of 62.62 µg h−1 mgcat. −1 and a Faradaic efficiency (FE) of 30.33% at −0.40 V versus the reversible hydrogen electrode (RHE) in 0.1 m HCl. Additionally, it achieves an NH3 yield of 41.47 µg h−1 mgcat. −1 and FE of 23.19% at −0.60 V versus RHE in 0.1 m Na2SO4. This work demonstrates a promising strategy for constructing heterojunction electrocatalysts for efficient eNRR.

Keywords: ammonia synthesis, cobalt sulfide, Nb2CTx MXene, nitrogen reduction reaction, spin manipulation


A novel 2D/2D Mott–Schottky heterojunction, Co9S8/Nb2CTx‐P is preparaed by a solution plasma modification method. The transformation of the Co 3d orbital electron configuration frow low (t2g 6eg 1) to high (t2g 5eg 2) spin is induced by Sv and interface effects. The highly spin‐polarized state of the Co9S8/Nb2CTx‐P demonstrated superior eNRR performance.

graphic file with name ADVS-11-2407301-g006.jpg

1. Introduction

As an excellent alternative to fossil fuels, ammonia (NH3) is regarded as a clean energy, offering advantages including low cost, easy availability, volatility, convenient storage, low pollution, high burning value, a high‐octane number, relatively safe operation, and compatibility with common materials.[ 1 ] Despite the high N2 content in air (78%), converting N2 from air to NH3 and related compounds is challenging, owing to the inertness and high dissociation energy of molecular N2.[ 2 ] Commercial NH3 is synthesized using the Haber–Bosch process, which requires harsh conditions, thereby leading to substantial energy consumption and significant CO2 emissions.[ 3 ] In contrast to the traditional high‐energy Haber‐Bosch process, the electrocatalytic nitrogen reduction reaction (eNRR) offers an emerging alternative strategy for NH3 generation under environmental conditions, using renewable energy sources.[ 4 ] However, the eNRR is typically conducted in acidic media. This is because the presence of abundant hydrogen protons (H+) promotes the formation of the *NNH intermediate, which is the rate‐determining step (RDS) in NH3 synthesis.[ 5 ] Given the growing environmental concerns, there is a high demand for the conversion of N2 into NH3 in neutral media. Therefore, the development of advanced NRR electrocatalysts for the production of NH3 in pH‐universal media is highly desirable.

Transition metal atoms comprise incompletely filled d orbitals that can effectively activate N2. Consequently, various transition metal materials have been developed as effective catalysts for the eNRR.[ 6 ] Nevertheless, the high electronegativity between the transition metal sites and neighboring N atoms renders the free energy of the intermediate (e.g., *NNH) unsuitable for adsorption.[ 7 ] Electrocatalysts in a highly spin‐polarized state can promote N2 adsorption in the eNRR and regulate the bonding strength of the adsorption/desorption intermediates on the electrocatalyst surface, affording tunable surface electrochemical reaction kinetics.[ 8 ] Transition metal spin states are closely related to the occupancy of the eg and t2g orbitals. Therefore, adjusting the electron configuration of these two orbitals can accelerate the eNRR and improve NH3 production efficiency. Various strategies for heteroatom doping,[ 7 , 9 ] element (e.g., S, O, N, or metal) vacancy engineering,[ 10 ] and interface engineering[ 11 ] have been exploited to regulate the electron configuration of the eg and t2g orbitals of active metal sites. For example, oxygen vacancies have been introduced into metal oxides (e.g., oxygen‐rich MoO2 [ 12 ] and TiO2 [ 13 ]) to regulate the electronic state and adsorption properties of the active sites. This strategy weakens the N─N bond, thereby facilitating NH3 production.[ 14 ] In addition, NH3 production using heterojunction electrocatalysts can be enhanced through the accumulation of reactants, formation of specific intermediates,[ 15 ] and inhibition of byproducts.[ 16 ] Current investigations on the spin states of active metal sites in heterojunctions mainly focus on revealing the origin of intrinsic activities.[ 16 , 17 ] However, research exploring the working mechanism of the electron spin states in such electrocatalysts and their impact on effective spin‐related electron transfer during the eNRR remains scarce. In particular, regulating the bonding strength of intermediates on electrocatalysts by manipulating the electron spin states of metal atomic active sites remains a significant challenge for promoting the activation of inert N2 during the eNRR.

In this study, a novel S‐vacancy (Sv)‐enriched Co9S8 and Nb2CTx MXene (Tx = surface terminal group) Mott–Schottky (M–S) heterojunction, Co9S8/Nb2CTx‐P, was prepared in situ using a solution plasma (SP) modification method for the efficient NH3 synthesis via the eNRR in acid and neutral electrolyte. During the SP irradiation procedure, a large number of free radicals and high‐energy particles can be generated, which further attack the electrocatalyst surface.[ 18 ] Rich lattice defects and element vacancies thus formed on the electrocatalyst surface thanks to the dissociation of chemical bonds of active component, which enhanced the electron delocalization and spin state transition of the Co active sites. The highly spin‐polarized state stimulated the production of more unpaired electrons in the d orbitals of the Co sites, which effectively promoted the adsorption and activation of N2. Moreover, owing to the M–S effect, the strong electron coupling between the resulting heterojunctions further reduced the energy barrier for N2 protonation. Density functional theory (DFT) calculations revealed that the upward shift of the dx2dy2 band center of the Co sites in Co9S8/Nb2CTx‐P energetically favored N2 adsorption and N─H bond formation in *NNH. The constructed Sv‐enriched Co9S8/Nb2CTx‐P M–S heterojunction exhibited superior eNRR performances, yielding a high NH3 yield of 62.62 µg h−1 mgcat. −1 at −0.40 V versus the reversible hydrogen electrode (RHE) in 0.1 m HCl, with a Faradaic efficiency (FE) of 30.33%. Additionally, it achieved an NH3 yield of 41.47 µg h−1 mgcat. −1 at −0.60 V versus RHE in 0.1 m Na2SO4, with an FE of 23.19%. This work systematically confirms the application of spin engineering to eNRR activity, paving the way for the design of a new type of heterojunction electrocatalyst for the eNRR.

2. Results and Discussion

Figure  1a illustrates the formation process of the Co9S8/Nb2CTx‐P M–S heterojunction. First the Co9S8/Nb2CTx M–S heterojunction was constructed through the in situ growth of Co9S8 nanosheets (NSs) around Nb2CTx MXene nanoflakes using a hydrothermal method. Subsequently, the obtained heterojunction was modified by SP treatment to form the Co9S8/Nb2CTx‐P M–S heterojunction comprising abundant Svs. Figure S1a,b (Supporting Information) clearly depicts that in the Co9S8/Nb2CTx, small‐sized NSs were vertically grown on the multilayered nanoflakes. Further, the edge of Co9S8/Nb2CTx was identified as Co9S8NSs, which is marked with red dashed lines in the TEM image (Figure S1b, Supporting Information). The high‐resolution transmission electron microscopy (HR‐TEM) image of the Co9S8/Nb2CTx (Figure S1c, Supporting Information) shows the connected structure of Co9S8 and Nb2CTx, along with a clear lattice spacing of 0.248 nm attributed to the (400) plane of Co9S8.

Figure 1.

Figure 1

a) Schematic illustration of the preparation procedure of the Co9S8/Nb2CTx‐P M–S heterojunction. b) SEM, c) TEM, d) HR‐TEM images of the Co9S8/Nb2CTx‐P. e) HR‐TEM images and corresponding lattice line scanning of Co9S8/Nb2CTx, f) HR‐TEM images and corresponding lattice line scanning of Co9S8/Nb2CTx‐P. g) The EDS mapping of Co9S8/Nb2CTx‐P (blue: Co, orange: Nb, and red: S).

Similar results were observed for the heterojunction following SP treatment (Figure 1b–d). However, a few Svs were observed in the HR‐TEM image of the Co9S8/Nb2CTx‐P (Figure 1f, encircled in red), which were absent in that of the pristine heterojunction (Figure 1e). Additionally, the energy‐dispersive X‐ray spectroscopy mapping images of the Co9S8/Nb2CTx and Co9S8/Nb2CTx‐P (Figure S1d, Supporting Information; Figure 1g, respectively) revealed the co‐existence and homogeneous distribution of Nb, Co, S, C, and O throughout the selected region, indicating the full coverage of Nb2CTx with Co9S8 NSs. For comparison, the individual components, Co9S8 and Nb2CTx, before and after SP modification were also prepared, and their surface morphologies and nanostructures were analyzed (Figures S2 and S3, respectively). The Co9S8/Nb2CTx demonstrated a larger Brunauer–Emmett–Teller surface area of 80.1 m2 g−1 compared to those of Co9S8 (51.9 m2 g−1) and Nb2CTx (24.7 m2 g−1) (Figure S4a and Table S1, Supporting Information). This superior result is conducive to the penetration of the electrolyte and sufficient contact between the reactants and catalyst.

Figure S5 (Supporting Information) depicts the X‐ray diffraction (XRD) patterns of the Co9S8/Nb2CTx M–S heterojunction before and after SP modification, revealing low crystallinity indicated by weak signals. Specifically, three weak diffraction peaks at 2θ = 29.7°, 47.8°, and 52.0° were observed in the XRD pattern of the heterojunction, corresponding to the (311), (511), and (440) facets of Co9S8, respectively. Furthermore, the diffraction peak at 2θ = 7.7°, which was attributed to the (002) plane of Nb2CTx, was relatively weak due to coverage by the Co9S8 NSs.

The Fourier‐transform infrared (FT‐IR) spectra of the heterojunction before and after the SP treatment were similar (Figure S6, Supporting Information). The characteristic band at 613 cm−1 was assigned to the stretching vibration of Co‐S bond in Co9S8. The peaks located at 483, 1107, 1363, and 1734 cm−1 were assigned to the Nb─C stretching, C─F, C─H, and C═O, respectively.

The X‐ray photoelectron spectroscopy (XPS) survey scan spectra of the Co9S8/Nb2CTx and Co9S8/Nb2CTx‐P (Figure S7, Supporting Information) displayed clear Nb 3d (207.3 eV), Co 2p (780.7 eV), S 2p (162.7 eV), C 1s (284.1 eV), and O 1s (531.6 eV) signals, further verifying the combination of the elemental signals of Co9S8 and Nb2CTx. The Co 2p XPS spectrum of Co9S8/Nb2CTx‐P (Figure  2a) was deconvoluted into two pairs of peaks corresponding to Co 2p 3/2 (781.1 eV) and Co 2p 1/2 (797.1 eV). The pairs comprised peaks attributed to Co0 (779.2 and 796.2 eV), Co3+ (781.1 and 797.3 eV), and Co2+ (782.9 and 798.7 eV), indicating Co0/Co2+/Co3+mixed valence states. Notably, the binding energy (BE) positions of the Co 2p 3/2 (781.3 eV) and Co 2p 1/2 (797.1 eV) peaks of Co9S8/Nb2CTx‐P shifted negatively, by 0.2 and 0.4 eV, relative to those of the Co9S8/Nb2CTx (781.1 and 797.5 eV), respectively. The peak positions of Co 2p 3/2 (781.3 eV) and Co 2p 1/2 (797.1 eV) in the Co9S8/Nb2CTx spectrum (curve ii, Figure 2a) also exhibited a slight negative shift of 0.22 eV, compared to those of pristine Co9S8 (curve i, Figure 2a). Specifically, the electrons of the Co atom in Co9S8/Nb2CTx can be transferred to the surrounding coordination environment.[ 19 ] Moreover, the Co2+/Co3+ intensity ratio of the Co9S8/Nb2CTx‐P M–S heterojunction (0.69) was higher than that of pristine Co9S8/Nb2CTx (0.54), and both were lower than that of Co9S8 (0.81) (Figure 2b).

Figure 2.

Figure 2

a) The Co 2p XPS spectra, and b) corresponding the different Co valence intensity ratio, c) S 2p XPS spectra of i) Co9S8, ii) Co9S8/Nb2CTx and iii) Co9S8/Nb2CTx‐P. d) Co K‐edge XANES and e) FT‐EXAFS spectra of Co foil, CoO, Co9S8, Co9S8/Nb2CTx, and Co9S8/Nb2CTx‐P. f) FT‐EXAFS fitting results of Co9S8, Co9S8/Nb2CTx, and Co9S8/Nb2CTx‐P. WT‐EXAFS contour map of g) Co9S8, h) Co9S8/Nb2CTx, and I) Co9S8/Nb2CTx‐P.

During plasma irradiation in water, a high flux of hydrogen radicals, reactive oxygen species, electrons, and ions is produced via water splitting at the open end of the discharge zone.[ 20 ] These active sites can then attack the Co─S bond in Co9S8, leading to its dissociation. Consequently, Co9S8/Nb2CTx‐P remains enriched with Svs with very low energy at the electrocatalyst surface, without altering the bulk properties. The pre‐occupied S 2p orbital electrons become delocalized around the Co3+ ions neighboring the S vacancies. Subsequently, the unpaired electrons resulting from the removal of S atoms can be accommodated in the orbitals of Co3+ ions, leading to the partial reduction of Co3+ to Co2+.[ 21 ] Additionally, the S 2p XPS spectrum of Co9S8/Nb2CTx‐P (curve iii, Figure 2c) included peaks corresponding to Co─S (163.9 eV), C═S (162.7 eV), and S─O (168.9 eV). The peak positions of S 2p 3/2 (160.7 eV) and S 2p 1/2 (161.9 eV) in the Co9S8/Nb2CTx spectrum (curve ii, Figure 2c) were negatively shifted by 0.7 and 0.8 eV compared to those of Co9S8 (161.7 and 162.8 eV) (curve i, Figure 2c), respectively. The M–S junction interface generated between Co9S8 and Nb2CTx can modulate the density of Co active sites, resulting in electron transfer from Nb2CTx to Co9S8.[ 22 ] Moreover, the BE positions of S 2p 3/2 and S 2p 1/2 in the Co9S8/Nb2CTx‐P spectrum exhibited negative shifts of 0.2 and 0.12 eV, respectively, relative to those of the Co9S8/Nb2CTx spectrum. This was attributed to the generation of abundant Svs in Co9S8/Nb2CTx‐P, which reduces the electron density on the S atom. In addition, the atomic composition, determined through XPS analysis, revealed a smaller S/Co atomic ratio of Co9S8/Nb2CTx‐P (0.53) compared to those of Co9S8 (0.69) and Co9S8/Nb2CTx (0.62; Figures S8–S10, Supporting Information), indicating the formation of more Svs.

To further probe the local coordination geometry at the atomic level of Co9S8/Nb2CTx before and after SP treatment, X‐ray absorption near‐edge structure (XANES) and extended X‐ray absorption fine structure (EXAFS) analyses were conducted on Co9S8, Co foil, and CoO for comparison. The Co9S8, Co9S8/Nb2CTx, and Co9S8/Nb2CTx‐P XANES spectra (Figure 2d) displayed similar peak shapes and adsorption edges, indicating their close valence states. Moreover, the energy absorption threshold of Co9S8/Nb2CTx‐P fell between those of the standard Co foil and CoO. These results indicated that the valence state of the Co atom in Co9S8/Nb2CTx‐P is higher than that of Co0 but lower than that of Co2+. The FT k2‐weighted phase‐uncorrected Co K‐edge EXAFS spectrum of Co9S8/Nb2CTx‐P (Figure 2e) displayed its main peak at 1.75 Å, attributed to the single scattering path of Co─S. Furthermore, the pristine heterojunction demonstrated a similar single scattering path of Co─S, which, however, was slightly shorter than that of Co9S8 (1.79 Å). Notably, the peak intensity of the Co─S bond in Co9S8/Nb2CTx‐P was lower than those in Co9S8/Nb2CTx and Co9S8. This weaker Co─S peak intensity of Co9S8/Nb2CTx‐P was primarily attributed to the presence of Svs.

The quantitative least‐squares EXAFS curves were fitted to analyze the local coordination of Co in the three catalysts (Figures 2f and S11 and Table S2, Supporting Information). Co9S8 presented two different sets, namely Co─S1 with bond length of 2.08 Å and Co─S2 with bond length of 2.21 Å, suggesting the coexistence of two different unit cells. The coordination numbers for Co─S1 and Co─S2 were fitted to be 1.53 and 3.19, respectively. For the Co9S8/Nb2CTx‐P heterojunction, the coordination number for Co─S1 was determined to be 1.26, which is lower than those of Co9S8 (1.53) and Co9S8/Nb2CTx (1.47) owing to the abundant formation of Svs. The decreased length of the Co─S bond and the low coordination number of Co9S8/Nb2CTx‐P demonstrate negligible structural deformation attributed to the high population of Svs. Moreover, the wavelet transform analysis of the three samples (Figure 2g–I) revealed that the Co─S path exhibited maximum intensity at R ≈ 2.15 Å and k ≈ 6.7 Å−1.

The normalized secondary electron cutoff energies of the samples were measured using ultraviolet photoelectron spectroscopy (UPS) (Figure  3a), enabling the deduction of the work function (Φ) of the electrocatalyst. As illustrated in Figure 3b, the Φ of Co9S8/Nb2CTx is ≈4.26 eV, which is marginally smaller than those of Co9S8 (4.58 eV) and Nb2CTx (4.49 eV). The semiconductor type of Co9S8 and Co9S8/Nb2CTx was probed using M–S curves (Figure S12, Supporting Information), revealing characteristics consistent with n‐type semiconductors. Ultraviolet‐visible (UV–vis) absorption spectra were recorded for all the prepared electrocatalysts (Figure S12c, Supporting Information). Additionally, the forbidden bands of Nb2CTx, Co9S8, and Co9S8/Nb2CTx were calculated as 0.84, 0.87, and 0.78 eV, respectively, from Tauc plots (Figure S12d–f, Supporting Information).

Figure 3.

Figure 3

a) UPS spectra in the normalized secondary electron cutoff energy (E cutoff) regions of i) Co9S8, ii) Nb2CTx, and iii) Co9S8/Nb2CTx. b) Calculated Φ stemming from UPS spectra. c) Energy band structures of Co9S8, Nb2CTx, and Co9S8/Nb2CTx (EF: Fermi level, EVB: valence band level, and Evac: vacuum level). d) Energy band diagrams of Nb2CTx and Co9S8 before and after the Mott‐Schottky junction formation. e) Schematic illustration of the interfacial electronic couple effect between Nb2CTx and Co9S8. f) EPR spectra i) Co9S8, ii) Co9S8/Nb2CTx, iii) Co9S8/Nb2CTx‐P. The electron localization function of g) Co9S8/Nb2CTx and h) Co9S8/Nb2CTx‐P. I) Magnetic susceptibilities and j) The number of unpaired electrons of i) Co9S8, ii) Co9S8/Nb2CTx, iii) Co9S8/Nb2CTx‐P. k) Illustration of Co2+ spin states in Co9S8/Nb2CTx‐P.

The combination of M–S plots, UPS, and UV–vis results confirmed the formation of an M–S heterojunction between Nb2CTx and Co9S8. Furthermore, the energy band plots and band structures of Nb2CTx and Co9S8 (Figure 3c,d) revealed the redistribution of electrons at the interface between Nb2CTx and Co9S8 owing to the Schottky barrier. When Nb2CTx and Co9S8 come into contact and form a heterojunction interface, electrons from Nb2CTx spontaneously flow to Co9S8 until dynamic equilibrium is reached in terms of the Fermi levels. At this moment, the holes in the valence band of Co9S8 flow to Nb2CTx, facilitating charge transfer at the interface. This induces band bending and generates an internal electric field at the M–S interfaces, thereby promoting the flow of electrons from Nb2CTx into Co9S8 (Figure 3e). In the presence of the Fermi level and interfacial built‐in electric field, interfacial electron transfer accelerates the catalytic process of the eNRR. This acceleration is attributed to the tendency of the Sv‐enriched Co9S8/Nb2CTx‐P to donate more electrons to the intermediates in the eNRR process.

Theoretically, the eNRR performance of transition metal‐based electrocatalysts is related to the occupancy rates of their d orbitals.[ 23 ] Therefore, to investigate the electronic configurations of the Co center, we conducted electron paramagnetic resonance (EPR), zero‐field cooling (ZFC) temperature‐dependent magnetization (χm). Figure 3f illustrates the signals of the Co─S dangling bonds at g = 2.003 in the EPR spectra of Co9S8/Nb2CTx before and after SP modification, which are proportional to the concentrations of dangling bonds from the Svs in Co9S8/Nb2CTx and Co9S8/Nb2CTx‐P. The intensity of the g signal of Co9S8/Nb2CTx‐P was higher than that of the pristine heterojunction, indicating an increase in the number of defects. Specifically, the Svs in Co9S8/Nb2CTx‐P result from a region of electron delocalization at the S position, further causing a slight localization of electrons in the surrounding S atom (Figure 3g,h). Consequently, the electronic structure of the surrounding Co active site can be modulated, facilitating electron transfer from the catalyst to N2 and the intermediates generated in the eNRR process. This enhances the NRR performance of the constructed electrocatalyst.

Given that charge redistribution is accompanied by the transition of the 3d electron spin configuration, temperature‐dependent magnetization in the ZFC mode was conducted at 1000 Oe in the temperature range of 5–300 K to gain deeper insight into the effect of the magnetic moments of Co9S8, Co9S8/Nb2CTx, and Co9S8/Nb2CTx‐P. From the χm versus T plots (Figure 3I), the effective magnetic moment (µeff) of Co9S8/Nb2CTx‐P was estimated to be 4.34 µB, which is markedly higher than those of Co9S8 (2.03 µB) and Co9S8/Nb2CTx (2.01 µB). Therefore, with an increase in the number of Svs, electrons can fill the high‐energy orbital (e g), thereby improving the spin state of Co. According to the equation 2.828χmT = µeff = n(n+2), [ 24 ] the number of unpaired d electrons (n) in Co9S8/Nb2CTx‐P (Figure 3j) was calculated to be ≈3.45, which is significantly larger than those in Co9S8/Nb2CTx (≈1.24) and Co9S8 (≈1.26). The result indicated that Co9S8/Nb2CTx‐P had three unpaired electrons, leading to a high spin state (t2g 5eg 2). However, Co9S8/Nb2CTx and Co9S8 only had one unpaired electron, forming a low spin state (t2g 6eg 1). As a result, it determines the electronic structure of Co 3d in Co9S8/Nb2CTx‐P (Figure 3k).

Owing to quantum spin exchange interactions, a greater number of unpaired electrons occupying the active orbital can enhance the catalytic activity of the electrocatalyst by weakening the binding of the reaction intermediates.[ 25 ] Accordingly, the strong interaction between the Svs and the formation of the heterojunction effectively reshaped the electronic structure of Co, resulting in a Co 3d electron spin configuration transition from low to high. Thus, Co9S8/Nb2CTx‐P exhibited more unpaired electrons in the 3d orbitals, which enhanced its catalytic activity compared to those of Co9S8/Nb2CTx and Co9S8 (Figure  4a).

Figure 4.

Figure 4

a) The schematic illustration of the spin regulation of Co 3d for the acceleration of the protonation process. b) LSV polarization curves of Co9S8/Nb2CTx‐P in an N2‐ and Ar‐saturated 0.1 m HCl electrolyte. Dependence of the NH3 yield and FE of Co9S8/Nb2CTx‐P at each applied potential in N2‐saturated c) 0.1 m HCl and d) 0.1 m Na2SO4 with the NRR measurement time of 6000 s. e) The eNRR performances of different electrocatalysts of Co9S8, Nb2CTx, and Co9S8/Nb2CTx before and after the SP treatment at −0.40 V versus RHE in 0.1 m HCl. f) Comparison of the NH3 yield rate of Co9S8/Nb2CTx‐P with some reported electrocatalysts.

The primary product of eNRR is NH3, which can be captured by electrolyte solution. Additionally, N2H2 as an unexpected by‐product also affects the performance of eNRR. Therefore, the production of NH3 and N2H4 in the cathode compartment in each of the two solutions was measured after each electrolysis procedure, according to the calibration curves standardized by the indophenol blue and Watt–Chrisp method methods (Figures S13–S16, Supporting Information). The eNRR performance of the Co9S8/Nb2CTx‐P M–S heterojunction was investigated in both 0.1 m HCl and 0.1 m Na2SO4, using Co9S8 and pristine Co9S8/Nb2CTx for comparison. Amongst, the influence of the Co9S8 content in Co9S8/Nb2CTx‐P on the NRR performances was systematically probed (Figure S17, Supporting Information). Linear sweep voltammetry polarization curves of Co9S8/Nb2CTx‐P were recorded separately in Ar‐ and N2‐saturated 0.1 m HCl (Figure 4b) and 0.1 m Na2SO4 (Figure S18, Supporting Information) at a scan rate of 5 mV s−1. A significant current density gap of Co9S8/Nb2CTx‐P was observed within the potential window of −0.23 to −0.70 V and −0.68 to −1.10 V versus RHE in 0.1 m HCl and 0.1 m Na2SO4, respectively. The onset potential of Co9S8/Nb2CTx‐P in 0.1 m HCl electrolyte is −0.40 V, while it in 0.1 m Na2SO4 is −0.68 V, in which the onset potential is determined at the current density of −0.5 mA cm−2. As results, the current gap of Co9S8/Nb2CTx‐P between N2 and Ar saturated electrolyte indicates that enhanced eNRR may occur at the nitrogen/electrocatalyst/electrolyte three‐phase interface. Additionally, cyclic voltammetry curves and Tafel plots of Co9S8/Nb2CTx‐P were investigated in Ar‐ and N2‐saturated 0.1 m HCl or Na2SO4 electrolyte (Figure S19, Supporting Information). Subsequently, chronoamperometry tests were conducted for 6000 s in 0.1 m HCl and 0.1 m Na2SO4 to evaluate the NH3 yield and FE of Co9S8/Nb2CTx‐P (Figure S20, Supporting Information). The NH3 yield was calculated from the UV–vis absorption spectra of the electrolyte (Figure S21, Supporting Information). Notably, no significant N2H4 byproduct was detected during the eNRR (Figure S22, Supporting Information). These results indicate that the Co9S8/Nb2CTx‐P M–S heterojunction exhibits good selectivity for NH3 production in both acidic and neutral media. As depicted in Figure 4c, Co9S8/Nb2CTx‐P exhibited a maximum NH3 yield rate of 62.62 µg h−1 mgcat. −1 and a high FE of 30.33% at −0.40 V versus RHE in 0.1 m HCl. However, when a more negative potential was applied, both the NH3 yield and FE of Co9S8/Nb2CTx‐P decreased markedly, owing to the competing hydrogen evolution reaction (HER). In addition, the electrocatalyst displayed extraordinary eNRR activity in 0.1 m Na2SO4 (Figure 4d), achieving a maximum NH3 yield of 41.47 µg h−1 mgcat. −1 at −0.60 V versus RHE and an FE of 23.19%.

As illustrated in Figure 4e, the NH3 yield and FE of the Co9S8/Nb2CTx‐P heterojunction are also higher than those of Co9S8 (NH3 yield, 7.85 µg h−1 mgcat. −1; FE, 16.62%) and Co9S8/Nb2CTx (NH3 yield, 16.89 µg h−1 mgcat. −1; FE, 13.83%) at −0.40 V versus RHE in 0.1 HCl. This performance also surpasses those of other Co‐ or MXene‐based electrocatalysts (Table S3, Supporting Information; Figure 4f), such as Co‐SAs/NC,[ 26 ] CoP3/CC,[ 27 ] 1T‐MoS2/g‐C3N4,[ 28 ] MnO2‐Ti3C2Tx,[ 29 ] and C@CoS@TiO2.[ 30 ] Similarly, the NH3 production rate and FE of Co9S8/Nb2CTx‐P in a neutral medium exceed those of most state‐of‐art electrocatalysts under the same conditions (Figure 4f; Table S4, Supporting Information), such as ZnO‐CoS QD,[ 31 ] FeCoOOH HNCs,[ 32 ] O‐CoP/CNT@G,[ 33 ] and CoS@S‐Mas.[ 34 ] This highlights the potential application of the constructed Co9S8/Nb2CTx‐P electrocatalyst for NH3 production under ambient conditions. The superior eNRR ability of Co9S8/Nb2CTx‐P is ascribed to the strong electronic coupling of Co9S8 with Nb2CTx and the abundant Svs, which significantly adjust the electronic structure of the Co active sites. The presence of more unpaired electrons in the 3d orbital of the Co active sites enhances the activation ability toward N2 molecules, thereby boosting the eNRR performance.

A control experiment was conducted under the same conditions using bare carbon paper (CP) at an open‐circuit voltage in N2‐saturated 0.1 m HCl (Figure S23a, Supporting Information) and 0.1 m Na2SO4 (Figure S23b Supporting Information). However, no significant NH3 production was observed with bare CP, indicating that NH3 is generated entirely from the eNRR catalyzed by Co9S8/Nb2CTx‐P. The origin of the N source was further determined using isotopic labeling measurements. As illustrated in the NMR spectra in Figures 5a and S24–S27 (Supporting Information), a distinct triplet (or doublet) coupling of 14NH4 + (or 15NH4 +) was observed when 14N2 (or 15N2) was used as the feeding gas. In contrast, no significant 14NH4 + (or 15NH4 +) signals were detected when the feeding gas was Ar (Figure S28, Supporting Information). The corresponding yields of 14N‐ and 15N‐labeled NH3 obtained using 1H NMR show that the amounts of NH3 produced by Co9S8/Nb2CTx‐P in 0.1 m HCl were 65.75 and 59.39 µg h−1 mgcat. −1, respectively. In 0.1 m Na2SO4, these values decreased to 44.74 and 38.29 µg h−1 mgcat. −1, respectively. These results are similar to those obtained using the indophenol blue method (Figure 5b). The introduced gas and experimental environment were carefully controlled throughout the experiments. Therefore, we supposed that the difference between the indophenol blue and isotopic labeling methods originated from experimental errors.

Figure 5.

Figure 5

15N2, 14N2 isotope labeling and indophenol blue experiments in 0.1 m HCl and 0.1 m Na2SO4. a) Baseline‐subtracted 1H NMR spectra of the post‐electrolyte with 15N2 and 14N2 compared with the reference 14NH4Cl and 15NH4Cl. b) Comparison of the NH3 yields calculated by NMR and indophenol blue tests. c) Cycling stability of Co9S8/Nb2CTx‐P at −0.40 V versus RHE in 0.1 m HCl. d) The NH3 yield and FE of original and post‐eNRR electrolysis in 0.1 M HCl and 0.1 M Na2SO4. e) The corresponding electrochemical double layer capacitances and f) the EIS Nyquist plots of Co9S8, Co9S8/Nb2CTx, and Co9S8/Nb2CTx ‐P at −0.4 V versus RHE in 0.1 m HCl. Inset: equivalent circuit model, where R s is the electrolytic resistance, R ct is charge‐transfer resistance, and CPE (constant phase angle element) is the double‐layer capacitance.

The stability of Co9S8/Nb2CTx‐P was assessed in 0.1 m HCl (Figure 5c) and 0.1 m Na2SO4 (Figure S29, Supporting Information). No substantial variations were observed in either the NH3 yields or FEs during five cycles of chronoamperometric runs, indicating the excellent stability of Co9S8/Nb2CTx‐P. Moreover, the chronoamperometry curves indicated that the current density of Co9S8/Nb2CTx‐P toward the eNRR remained stable without significant fluctuation, for 24 h, in both 0.1 m HCl (Figure S30a, Supporting Information) and 0.1 m Na2SO4 (Figure S30b, Supporting Information). Consequently, no significant changes were observed in the NH3 production rate or FE before and after 24 h (Figure 5d), verifying the high stability of Co9S8/Nb2CTx‐P toward the eNRR. In addition, the pH value of the post‐electrolyte after eNRR was studied, revealing a negligible difference in pH between the 0.1 m HCl and 0.1 m Na2SO4 electrolytes (Figure S31, Supporting Information). The SEM and TEM images and XRD patterns of Co9S8/Nb2CTx‐P after the long‐term tests in 0.1 m HCl (Figure S32, Supporting Information) and 0.1 m Na2SO4 (Figure S33, Supporting Information) also displayed no substantial changes, further confirming its good structural stability. The XPS characterizations of the spent electrocatalyst showed that the BE positions of the Co 2p (Figure S34a, Supporting Information) and S 2p XPS (Figure S34b, Supporting Information) spectra displayed positive shifts in both 0.1 m HCl and 0.1 m Na2SO4. This indicated that the electron density of the Co active sites in Co9S8 increased after use in the eNRR. In contrast, the Nb 3d XPS spectrum (Figure S34c, Supporting Information) displayed a negative shift, revealing a decrease in the electron density of the Nb sites. Collectively, these observations suggest that electron transfer occurred between Co9S8 and Nb2CTx, indicating charge reconstruction during the eNRR process.

To elucidate the catalytic mechanism of the eNRR, we conducted electrochemical active surface area and electrochemical impedance spectroscopy (EIS) measurements for all the catalysts in 0.1 m HCl and 0.1 m Na2SO4. Cyclic voltammetry curves of all samples (Figure S35, Supporting Information), obtained within the narrow potential window of 0.10–0.20 V versus Ag/AgCl (saturated KCl solution), were constructed to evaluate the double‐layer capacitance (C dl), which represents the electrochemical surface area. Co9S8/Nb2CTx‐P displayed a C dl value of 2.14 mF cm−2, which was higher than that of Co9S8 (1.34 mF cm−2) and lower than that of Co9S8/Nb2CTx (3.52 mF cm−2) (Figure 5e). Moreover, the EIS Nyquist plots (Figure 5f) revealed that Co9S8/Nb2CTx‐P exhibited a smaller R ct (21.3 Ω) compared to those of Co9S8 (41.2 Ω) and Co9S8/Nb2CTx (60.0 Ω). In 0.1 m Na2SO4, the C dl of Co9S8/Nb2CTx‐P (4.20 mF cm−2) was slightly higher than that of Co9S8/Nb2CTx (3.60 mF cm−2) (Figure S36, Supporting Information). Similarly, the EIS Nyquist plot of Co9S8/Nb2CTx‐P in 0.1 m Na2SO4 (Figure S37, Supporting Information) indicated a smaller R ct (54.6 Ω) compared to those of Co9S8 (110.0 Ω) and Co9S8/Nb2CTx (71.2 Ω). These results suggest that Co9S8/Nb2CTx‐P facilitates faster electron transfer and possesses more catalytically active sites than the other catalysts, thereby enhancing the eNRR performance at a relatively low overpotential.

Given the superior eNRR ability demonstrated by the constructed Co9S8/Nb2CTx‐P M–S junction, assembly of a Zn–N2 aqueous battery for NH3 synthesis and simultaneous electricity generation is envisioned.[ 35 ] To explore this possibility, we conceptually manufactured a home‐made Zn–N2 battery composed of a cathode of Co9S8/Nb2CTx‐P coated on the CP and an anode of Zn foil. A photograph of the reaction cell is shown in Figure  6a. N2 was continuously bubbled into the cathode within an asymmetric electrolyte system comprising an acid catholyte and alkaline anolyte separated by a bipolar membrane.

Figure 6.

Figure 6

a) Schematic diagram of a rechargeable Zn–N2 battery. b) Discharging polarization curves and the power density plot of the Co9S8/Nb2CTx‐P‐assembled Zn–N2 battery. c) Comparison of the power densities of the Zn–N2 battery assembled with different catalysts. d) The discharge curve of the constructed Zn–N2 battery with Co9S8/Nb2CTx‐P at 0.01 mA cm−2 for 7200 s. e) The NH3 yields produced by the Zn–N2 battery with Co9S8/Nb2CTx‐P for five repeated tests. f) The discharge curve of the manufactured Zn–N2 battery at the current density of 0.01 mA cm−2.

The rate–discharge curve of Co9S8/Nb2CTx‐P in Figure 6b displays the voltage changes at different current densities. As intended, the Zn–N2 battery assembled with Co9S8/Nb2CTx‐P delivered a high discharge current density, peaking at 48.8 mA cm−2 at 0.10 V versus Zn2+/Zn. Furthermore, Figure 6b shows that the Co9S8/Nb2CTx‐P‐based Zn–N2 cell can deliver a power density of 11.17 mW cm−2, significantly outperforming some reported Zn–N2 batteries (Figure 6c), including CoxNi3‐x(HITP)2/BNSs‐P (2.5 mW cm−2),[ 36 ] CoPi/NPCS (0.49 mW cm−2),[ 37 ] and CoPi/HSNPC (0.31 mW cm−2).[ 38 ] After the constructed cell was discharged at 0.01 mA cm−2 for 7200 s (Figure 6d), an average NH3 yield of 11.8 µg h−1 cm−2 (Figure 6e) was determined based on five repeated tests. Additionally, a high energy density of 124 mA h g−1 was achieved at 0.01 mA cm−2 (Figure 6f).

To thoroughly elucidate the electrocatalytic pathway of the eNRR on Co9S8/Nb2CTx‐P, time‐dependent in situ ATR‐FTIR spectra were recorded in 0.1 m N2‐saturated HCl at −0.40 V versus RHE. As depicted in Figure  7a, the appearance of the absorption peak at 1104 cm−1, corresponding to the N─N tensile bond, suggests the dissociation of the N≡N bond of the adsorbed N2 molecules on the working electrode surface.[ 39 ] Additionally, the three adsorption peaks at 1272, 1442, and 3278 cm−1 were assigned to ─NH2 vibration, H─N─H bending, and N─H stretching vibrations, respectively.[ 40 ] Notably, the intensities of these peaks increased with increasing reaction time. Consequently, we supposed that the reaction intermediates accumulate during the eNRR process. Most importantly, the absorption peak at 1637 cm−1, attributed to N2 molecules adsorbed on the catalyst surface, gradually decreases in intensity, indicating the consumption of N2 molecules.[ 36 ] These results suggest that the eNRR over the Co9S8/Nb2CTx‐P surface may follow an alternative pathway. Similar results were observed in the in situ ATR‐FTIR spectra of the electrocatalyst recorded in 0.1 m Na2SO4 (Figure S38, Supporting Information), indicating the same catalytic pathway.

Figure 7.

Figure 7

a) In situ electrochemical FT‐IR spectra and b) in situ Raman spectra on Co9S8/Nb2CTx‐P during the eNRR in 0.1 m HCl. c) The N2 adsorption energy of Nb2CTx, Co9S8, Co9S8‐P, Co9S8/Nb2CTx, and Co9S8/Nb2CTx‐P. d) The projected density of states of Co 3dx2dy2 orbitals for Co9S8, Co9S8/Nb2CTx, and Co9S8/Nb2CTx‐P. e) The schematic illustration of the shift of dx2dy2 orbital toward E F on the facilitation of the protonation of Co9S8/Nb2CTx‐P, E H, and E BC represent the dx2dy2 highest occupied crystal orbital and band center, respectively. f) Partial density of states of N2 adsorption on Co9S8, Co9S8‐P, Co9S8/Nb2CTx, and Co9S8/Nb2CTx‐P. g) Charge density difference of N2 adsorption on Co9S8/Nb2CTx‐P (red sphere: S, blue sphere: Co, brown sphere: Nb, and gray sphere: C. Yellow and cyan‐colored iso‐surfaces show electron gain and loss, respectively). h) Calculated Gibbs free energies of the eNRR on Co9S8‐P, Co9S8, Co9S8/Nb2CTx, and Co9S8/Nb2CTx‐P along alternating pathways.

To identify the operative active sites of Co9S8/Nb2CTx‐P, the eNRR was further monitored via electrochemical in situ Raman spectroscopy at various potentials (Figure 7b). Notable Co─N bond formation was observed at 263 cm−1 with increasing applied potential,[ 41 ] indicating that N2 molecules were predominately captured by the Co site (Figure 7b). Furthermore, the significant peak at ≈1578.5 cm−1 was assigned to H─N─H tensile vibration (Figure 7b).[ 42 ] Notably, the intensity of this peak decreased with increasing potential, possibly due to NH3 accumulation. The strongest Co─N bond was observed at −0.40 V versus RHE in the potential‐dependent in situ Raman spectra, consistent with experimental findings.

To further elucidate the activation process of N2 at the Co site and the evolution of intermediates during the eNRR process, DFT calculations were performed to provide deeper insight into the electrocatalytic mechanism of the prepared catalyst toward the NRR. As depicted in Figure 7c, the adsorption energy of N2 molecules on the Co9S8/Nb2CTx‐P surface was higher compared to that of the plasma‐modified Co9S8 (Co9S8‐P) and markedly higher than those of pristine Co9S8, Nb2CTx, and Co9S8/Nb2CTx. This finding reveals that the formation of a microinterface between the diverse components and the generation of rich S vacancies through plasma modification can enhance N2 adsorption, thereby improving the eNRR performance. In terms of the density of states (DOS) (Figure S39, Supporting Information), Co9S8/Nb2CTx‐P displayed an increased integral area of the net spin‐up compared to both Co9S8 and Co9S8/Nb2CTx. This observation confirms that the Svs and interface effects lead to an increase in the number of spin‐up electrons, which is consistent with the ZFC temperature‐dependent magnetic susceptibility experimental results. In addition, the broader electronic states of the e g orbitals close to the Fermi level can improve the electron transfer mobility and provide a lower adsorption energy between the active site and reaction intermediate, boosting the eNRR performance.[ 7 , 43 ] As illustrated in Figure 7d,e, the broader dx2dy2 electron state enabled more efficient electron transfer between N2 and the electrocatalyst, thereby facilitating the N2 protonation step. This was verified by examining the projected DOS before and after N2 adsorption (Figure 7f). The projected DOS (Figures 7f and S39, Supporting Information) revealed that the Co‐3d orbital of Co9S8/Nb2CTx‐P and the *N2‐2p orbital can overlap more effectively compared to those of Co9S8 (Figure 7f), Co9S8‐P (Figure 7f), and Co9S8/Nb2CTx (Figure 7f). Therefore, the Co9S8/Nb2CTx‐P M–S heterojunction is more conducive to the adsorption and activation of N2. Charge difference analysis (Figure S40, Supporting Information) demonstrated that electrons can accumulate at the edge of the interface between Co9S8 and Nb2CTx. Both negative (charge accumulation) and positive (charge depletion) charges are simultaneously present around *N2 after N2 adsorption (Figure 7g).[ 44 ] Thus, the Svs and their neighboring Co return their 3d electrons to the π* orbital of *N2, further activating and polarizing N2 molecules through an “acceptance–donation” mechanism.

We further calculated the Gibbs free energies for the eNRR catalyzed by Co9S8‐P, Co9S8/Nb2CTx‐P, Co9S8, and Co9S8/Nb2CTx (Figure  7h ). Remarkably, the high adsorption ability of N2 molecules onto Co9S8/Nb2CTx‐P led to a low energy barrier (0.27 eV) for the first protonation RDS of the *N2→*NNH formation process. In addition, Co9S8/Nb2CTx and Co9S8 exhibited the energy barriers of 2.19 and 2.27 eV in the first protonation step, respectively, which were markedly higher than those of Co9S8/Nb2CTx‐P (0.27 eV) and Co9S8‐P (0.34 eV) (Figure 7h). Moreover, the RDS barrier of Co9S8/Nb2CTx‐P is lower than those of most previously reported catalysts.[ 45 ] The competing HER kinetics in the eNRR procedure driven by Co9S8/Nb2CTx and Co9S8/Nb2CTx‐P was calculated using DFT, as depicted in Figure S41 (Supporting Information). The result manifested that the Gibbs free energy of hydrogen (*H) adsorption on the Co9S8/Nb2CTx‐P (−0.26 eV) was remarkably lower than that of Co9S8/Nb2CTx (0.80 eV). Thereby, the *H species can be spontaneously adsorbed on the Co9S8/Nb2CTx‐P surface, finally forming the 1/2H product. These results hinted that the generated rich sulfur vacancies remarkably accelerated the formation of the transition intermediate of active *H species, thus boosting the rate‐determining step of the eNRR, i.e., the production of *NNH. In addition, Co9S8/Nb2CTx‐P showed lower energy of N2 adsorption (Figure 7c, −0.87 eV) than that of hydrogen adsorption (−0.26 eV). It suggested the efficient HER suppression, thereby improving Faradaic efficiency and NH3 yield of Co9S8/Nb2CTx‐P. These results indicate that the Sv abundance together with the synergistic effect of Co9S8 with Nb2CTx can promote the highly spin‐polarized state of the Co 3d orbital electrons. This activates N2 and promotes the formation of *NNH, greatly enhancing the eNRR activity.

3. Conclusion

In summary, a new Co9S8/Nb2CTx‐P hybrid was synthesized by coupling Co9S8 with Nb2CTx, in which the spin state of the Co2+ catalytic center was successfully regulated to improve the efficiency and selectivity of the eNRR. The transformation of the Co 3d orbital electron configuration from low (t2g 6eg 1) to high (t2g 5eg 2) spin was induced by Sv and interface effects. The resulting highly spin‐polarized state of the Co9S8/Nb2CTx‐P hybrid demonstrated superior eNRR performance in acid and neutral electrolyte. Compared to the pristine Co9S8 and Co9S8/Nb2CTx hybrids, Co9S8/Nb2CTx‐P exhibited high NH3 yields of 62.62, and 41.47 µg h−1 mgcat. −1 coupled with FEs of 30.33% and 23.19% in 0.1 m HCl and 0.1 m Na2SO4, respectively, surpassing the performances of most reported NRR electrocatalysts. DFT calculations indicated that this improvement in catalytic performance is related to the increase and upward shift of the Co dx2dy2 orbitals, which significantly promotes the adsorption and activation of N2 and reduces its protonation energy barrier. This study demonstrates the potential of heterostructured materials in designing high‐performance spin‐based catalysts, highlighting the importance of optimizing the spin states for the eNRR.

Conflict of Interest

The authors declare no conflict of interest.

Supporting information

Supporting Information

Acknowledgements

The authors are grateful for the financial support from the National Natural Science Foundation of China (No. U21A20286) and the Key Research Project of University of Henan Province (No. 23ZX001).

Zhang S., Zhao W., Liu J., Tao Z., Zhang Y., Zhao S., Zhang Z., Du M., Spin Manipulation of Co sites in Co9S8/Nb2CTx Mott–Schottky Heterojunction for Boosting the Electrocatalytic Nitrogen Reduction Reaction. Adv. Sci. 2024, 11, 2407301. 10.1002/advs.202407301

Contributor Information

Zhihong Zhang, Email: 2006025@zzuli.edu.cn.

Miao Du, Email: dumiao@zzuli.edu.cn.

Data Availability Statement

The data that support the findings of this study are available from the corresponding author upon reasonable request.

References

  • 1. Tian F., Zhou N., Chen W., Zhan J., Tang L., Wu M., Adv. Sustainable Syst. 2024, 8, 2300618. [Google Scholar]
  • 2.a) Ma J. L., Bao D., Shi M. M., Yan J. M., Zhang X. B., Chem 2017, 2, 525; [Google Scholar]; b) Tanabe Y., Nishibayashi Y., Chem. Soc. Rev. 2021, 50, 5201. [DOI] [PubMed] [Google Scholar]
  • 3.a) Yang B., Ding W., Zhang H., Zhang S., Energy Environ. Sci. 2021, 14, 672; [Google Scholar]; b) Zhao S., Lu X., Wang L., Gale J., Amal R., Adv. Mater. 2019, 31, 1805367. [DOI] [PubMed] [Google Scholar]
  • 4.a) Wu H., Singh‐Morgan A., Qi K., Zeng Z., Mougel V., Voiry D., ACS Catal. 2023, 13, 5375; [DOI] [PMC free article] [PubMed] [Google Scholar]; b) Wen Y., Wang T., Hao J., Zhuang Z., Gao G., Lai F., Lu S., Wang X., Kang Q., Wu G., Du M., Zhu H., Adv. Funct. Mater. 2024, 34, 2400849; [Google Scholar]; c) Choe S., Kim N., Jang Y. J., EcoEnergy 2023, 1, 3. [Google Scholar]
  • 5. Mhadeshwar A. B., Kitchin J. R., Barteau M. A., Vlachos D. G., Catal. Lett. 2004, 96, 13. [Google Scholar]
  • 6. Shi L., Yin Y., Wang S., Sun H., ACS Catal. 2020, 10, 6870. [Google Scholar]
  • 7. Li Y., Ji Y., Zhao Y., Chen J., Zheng S., Sang X., Yang B., Li Z., Lei L., Wen Z., Feng X., Hou Y., Adv. Mater. 2022, 34, 2202240. [DOI] [PubMed] [Google Scholar]
  • 8. Bu F., Chen C., Yu Y., Hao W., Zhao S., Hu Y., Qin Y., Angew. Chem., Int. Ed. 2023, 62, e202216062. [DOI] [PubMed] [Google Scholar]
  • 9.a) Yang Y., Zhang L., Hu Z., Zheng Y., Tang C., Chen P., Wang R., Qiu K., Mao J., Ling T., Qiao S. Z., Angew. Chem., Int. Ed. 2020, 59, 4525; [DOI] [PubMed] [Google Scholar]; b) Chen X., Guo Y., Du X., Zeng Y., Chu J., Gong C., Huang J., Fan C., Wang X., Xiong J., Adv. Energy Mater. 2020, 10, 1903172; [Google Scholar]; c) Sun Y., Sun Z., Zhang W., Li W., Liu C., Zhao Q., Huang Z., Li H., Wang J., Ma T., EcoEnergy 2023, 1, 186. [Google Scholar]
  • 10.a) Guo H., Yang P., Yang Y., Wu H., Zhang F., Huang Z. F., Yang G., Zhou Y., Small 2024, 20, 2309007; [DOI] [PubMed] [Google Scholar]; b) Yao J. X., Bao D., Zhang Q., Shi M. M., Wang Y., Gao R., Yan J. M., Jiang Q., Small Methods 2019, 3, 1800333; [Google Scholar]; c) Mu J., Gao X. W., Yu T., Zhao L. K., Luo W. B., Yang H., Liu Z. M., Sun Z., Gu Q.‐F., Li F., Adv. Sci. 2024, 11, 2308979. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 11. Ye T., Ba K., Yang X., Xiao T., Sun Y., Liu H., Tang C., Ge B., Zhang P., Duan T., Sun Z., Chem. Eng. J. 2023, 452, 139515. [Google Scholar]
  • 12. Wen Y., Liu J., Zhang F., Li Z., Wang P., Fang Z., He M., Chen J., Song W., Si R., Wang L., Nano Res. 2023, 16, 4664. [Google Scholar]
  • 13. Song G., Gao R., Zhao Z., Zhang Y., Tan H., Li H., Wang D., Sun Z., Feng M., Appl. Catal. B: Environ. 2022, 301, 120809. [Google Scholar]
  • 14. Yang X., Ling F., Su J., Zi X., Zhang H., Zhang H., Li J., Zhou M., Wang Y., Appl. Catal. B: Environ. 2020, 264, 118477. [Google Scholar]
  • 15. Chu K., Luo Y., Shen P., Li X., Li Q., Guo Y., Adv. Energy Mater. 2022, 12, 2103022. [Google Scholar]
  • 16. Li X., Hai G., Liu J., Zhao F., Peng Z., Liu H., Leung M. K. H., Wang H., Appl. Catal. B: Environ. 2022, 314, 121531. [Google Scholar]
  • 17. Yang X., Tian Y., Mukherjee S., Li K., Chen X., Lv J., Liang S., Yan L.‐K., Wu G., Zang H.‐Y., Angew. Chem., Int. Ed. 2023, 62, e202304797. [DOI] [PubMed] [Google Scholar]
  • 18. Luo Y., Wu Y., Huang C., Menon C., Feng S.‐P., Chu P. K., EcoMat 2022, 4, e12197. [Google Scholar]
  • 19. Song K., Zhang H., Lin Z., Wang Z., Zhang L., Shi X., Shen S., Chen S., Zhong W., Adv. Funct. Mater. 2024, 34, 2312672. [Google Scholar]
  • 20.a) Lee I., Kang M., Park S., Park C., Lee H., Bae S., Lim H., Kim S., Hong W., Choi S. Y., Small 2024, 20, 2305143; [DOI] [PubMed] [Google Scholar]; b) Yu F., Wang C., Li Y., Ma H., Wang R., Liu Y., Suzuki N., Terashima C., Ohtani B., Ochiai T., Fujishima A., Zhang X., Adv. Sci. 2020, 7, 2000204. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 21. Tang Y., Shen K., Zheng J., He B., Chen J., Lu J., Ge W., Shen L., Yang P., Deng S., Chem. Eng. J. 2022, 427, 130915. [Google Scholar]
  • 22. He K., Li W., Tang L., Chen L., Wang G., Liu Q., Xin X., Yang C., Wang Z., Lv S., Xing D., Appl. Catal. B: Environ. 2023, 338, 123077. [Google Scholar]
  • 23. Liu T., Qu X., Zhang Y., Wang X., Dang Q., Li X., Wang B., Tang S., Luo Y., Jiang J., Chem. Eng. J. 2023, 457, 141187. [Google Scholar]
  • 24. Wei X., Song S., Cai W., Luo X., Jiao L., Fang Q., Wang X., Wu N., Luo Z., Wang H., Zhu Z., Li J., Zheng L., Gu W., Song W., Guo S., Zhu C., Chem 2023, 9, 181. [Google Scholar]
  • 25. Huang C. Y., Lin H. M., Chiang C. H., Chen H. A., Liu T. R., Vishnu D. S. K., Chiou J. W., Sankar R., Tsai H. M., Pong W. F., Chen C. W., Adv. Funct. Mater. 2023, 33, 2305792. [Google Scholar]
  • 26. Zhang S., Jiang Q., Shi T., Sun Q., Ye Y., Lin Y., Zheng L. R., Wang G., Liang C., Zhang H., Zhao H., ACS Appl. Energy Mater. 2020, 3, 6079. [Google Scholar]
  • 27. Gao J., Lv X., Wang F., Luo Y., Lu S., Chen G., Gao S., Zhong B., Guo X., Sun X., J. Mater. Chem. A 2020, 8, 17956. [Google Scholar]
  • 28. Xu X., Tian X., Sun B., Liang Z., Cui H., Tian J., Shao M., Appl. Catal. B: Environ. 2020, 272, 118984. [Google Scholar]
  • 29. Kong W., Gong F., Zhou Q., Yu G., Ji L., Sun X., Asiri A. M., Wang T., Luo Y., Xu Y., J. Mater. Chem. A 2019, 7, 18823. [Google Scholar]
  • 30. Liu Y. T., Chen X., Yu J., Ding B., Angew. Chem., Int. Ed. 2019, 58, 18903. [DOI] [PubMed] [Google Scholar]
  • 31. Ji Y., Hu Q., Yang M., Liu X., J. Mater. Chem. A 2023, 11, 25247. [Google Scholar]
  • 32. Han X., Liu C., Tang Y., Meng Q., Zhou W., Chen S., Deng S., Wang J., J. Mater. Chem. A 2023, 11, 14424. [Google Scholar]
  • 33. Meng Q., Hou Y., Yang F., Cao C., Zou Z., Luo J., Zhou W., Tong Z., Chen S., Zhou S., Wang J., Deng S., Appl. Catal. B: Environ. 2022, 303, 120874. [Google Scholar]
  • 34. Li Q., Song T., Wang Z., Wang X., Zhou X., Wang Q., Yang Y., Small 2021, 17, 2103305. [DOI] [PubMed] [Google Scholar]
  • 35. Lv X. W., Liu Y., Wang Y. S., Liu X. L., Yuan Z. Y., Appl. Catal. B: Environ. 2021, 280, 119434. [Google Scholar]
  • 36. Liu J., He L., Zhao S., Li S., Hu L., Tian J. Y., Ding J., Zhang Z., Du M., Adv. Sci. 2023, 10, 2205786. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 37. Ren J. T., Chen L., Wang H. Y., Yuan Z. Y., ACS Appl. Mater. Interfaces 2021, 13, 12106. [DOI] [PubMed] [Google Scholar]
  • 38. Ren J. T., Chen L., Liu Y., Yuan Z. Y., J. Mater. Chem. A 2021, 9, 11370. [Google Scholar]
  • 39. Sun Y., Han Y., Zhang X., Cai W., Zhang Y., Zhang Y., Li Z., Li B., Lai J., Wang L., Appl. Catal. B: Environ. 2022, 319, 121933. [Google Scholar]
  • 40.a) Yu Y., Lv Z., Liu Z., Sun Y., Wei Y., Ji X., Li Y., Li H., Wang L., Lai J., Angew. Chem., Int. Ed. 2024, 63, e202402236; [DOI] [PubMed] [Google Scholar]; b) Yu Y., Sun Y., Han J., Guan Y., Li H., Wang L., Lai J., Energy Environ. Sci. 2024, 17, 5183. [Google Scholar]
  • 41. Fang J. Y., Zheng Q. Z., Lou Y. Y., Zhao K. M., Hu S. N., Li G., Akdim O., Huang X. Y., Sun S. G., Nat. Commun. 2022, 13, 7899. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 42. Ren Y., Yu C., Wang L., Tan X., Wang Z., Wei Q., Zhang Y., Qiu J., J. Am. Chem. Soc. 2022, 144, 10193. [DOI] [PubMed] [Google Scholar]
  • 43.a) Xu G., Xue R., Stuard S. J., Ade H., Zhang C., Yao J., Li Y., Li Y., Adv. Mater. 2021, 33, 2006753; [DOI] [PubMed] [Google Scholar]; b) Li J., Wu C., Wang Z., Meng H., Zhang Q., Tang Y., Zou A., Zhang Y., Zhong H., Xi S., Xue J., Wang X., Wu J., Angew. Chem., Int. Ed. 2024, 63, e202404730. [DOI] [PubMed] [Google Scholar]
  • 44. Tan Y., Zhao Y., Chen X., Zhai S., Wang X., Su L., Yang H., Deng W.‐Q., Ho G. W., Wu H., EcoEnergy 2024, 2, 258. [Google Scholar]
  • 45.a) Cheng Y., Li X., Shen P., Guo Y., Chu K., Energy Environ. Mater. 2023, 6, e12268; [Google Scholar]; b) Chu K., Li X., Li Q., Guo Y., Zhang H., Small 2021, 17, 2102363. [DOI] [PubMed] [Google Scholar]

Associated Data

This section collects any data citations, data availability statements, or supplementary materials included in this article.

Supplementary Materials

Supporting Information

Data Availability Statement

The data that support the findings of this study are available from the corresponding author upon reasonable request.


Articles from Advanced Science are provided here courtesy of Wiley

RESOURCES