Mingrui
Wu
a,
Quan
Li
a,
Zhengting
Xiao
ab,
Dongcai
Shen
b,
Minghui
Hao
a and
Wentai
Wang
*a
aKey Laboratory of Marine Chemistry Theory and Technology, Ministry of Education; College of Chemistry and Chemical Engineering, Ocean University of China, Qingdao, 266100, China. E-mail: wentaiwang@ouc.edu.cn
bKey Laboratory of Biofuels, Qingdao Institute of Bioenergy and Bioprocess Technology, Chinese Academy of Sciences, Qingdao 266101, PR China
First published on 9th October 2025
The development of efficient electrocatalysts for the nitrogen oxidation reaction (NOR) under mild conditions is crucial for sustainable nitrate synthesis. Mo-doped δ-MnO2 electrocatalysts with varying Mo concentrations were successfully prepared for the NOR. Structural and electrochemical analyses revealed that Mo doping simultaneously enhanced the conductivity and electrochemically active surface area (ECSA) while promoting N2 adsorption and activation through electronic structure modulation. The optimized 2.5% Mo-doped δ-MnO2 (denoted as MM2.5) exhibited superior NOR performance in 0.1 M KOH, delivering a NO3− production rate of 116.75 μg h−1 mgcat−1 with a faradaic efficiency (FE) of 7.04% and excellent long-term stability. In addition, a Zn–N2 device was formed with MM2.5 as the anode and a Zn plate as the cathode, and the NO3− yield obtained in this device was even higher than 144.5 μg h−1 mgcat−1. However, structural characterization revealed that excessive Mo doping disrupted the δ-MnO2 crystal structure, reducing specific surface area and active site density. Density functional theory (DFT) calculations demonstrated that Mo doping lowered the Gibbs free energy of the rate-determining step (*N2 → *NNOH) from 2.41 eV to 1.94 eV by facilitating electron transfer, thereby optimizing the reaction pathway. This study provides a new strategy for the design of transition metal oxide-based electrocatalysts, as well as the application in artificial nitrogen fixation.
N bond (941 kJ mol−1) that leads to sluggish activation kinetics. This fundamental limitation has positioned the rational design of high-performance electrocatalysts as the pivotal research frontier in advancing nitrogen electrooxidation technologies.6–8
Transition metals facilitate electron transfer and modulate the electronic structure, thereby enhancing catalytic performance and are extensively utilized in electrocatalytic synthesis.9–11 In addition, transition metal oxides have garnered significant attention in nitrogen fixation catalysis owing to their versatile electronic configurations and abundant exposed active sites.12–15 Particularly, manganese dioxide (MnO2) demonstrates exceptional potential through its distinctive δ-phase layered architecture and dynamic Mn3+/Mn4+ redox couples, which create metastable surface states conducive to nitrogen activation.16,17 The engineered oxygen vacancies in MnO2 nanosheets can facilitate the formation of high-spin Mn3+–Mn3+ dimers.15 These coordinatively unsaturated pairs exhibit enhanced d-orbital overlap with N2 antibonding orbitals, effectively lowering the activation barrier for N
N bond cleavage through optimized electron transfer pathways. A manganese phthalocyanine catalyst with hollow nanotube structures for the electrocatalytic NOR was successfully prepared by Ghorai et al., who found that Mn atoms not only act as the main active sites of the reaction, but also significantly promote the generation of reactive species, which effectively enhances the NOR performance.18 Hou et al. prepared a Ru/Mn1.04–RuO2 catalyst for the electrocatalytic NOR, and enhanced the catalytic selectivity and activity for the NOR by constructing a Mn → Ru electron transfer pathway to modulate the electronic structure of the catalyst surface and increase the electron density of the Ru center.19
However, unmodified MnO2 still suffers from insufficient electrical conductivity issues, including weak adsorption of intermediates and structural instability. These limitations result in a significant gap between its catalytic performance and practical application requirements.20,21 Doping of molybdenum (Mo) has been demonstrated as an effective strategy to modulate the electronic structure of catalysts. For instance, the doping of high electronegativity Mo in TiO2 induces electron transfer from adjacent Ti atoms upon substituting Ti sites, significantly enhancing the adsorption and activation of N2.22 The Mo atoms in FeP-doped systems exhibit electron-deficient characteristics, enabling them to accept electrons from adjacent N atoms. This interaction amplifies the charge disparity between the two N atoms, thereby polarizing and activating the N
N bond.23 The doping of Mo into MnO2 can enhance structural stability and induce local electronic reconstruction, thereby optimizing the reaction pathway.17,24,25 Among several MnO2 crystal types, δ-MnO2 has attracted much attention due to its unique layered structure, and this open crystalline framework not only facilitates the homogeneous distribution of dopant elements, but also provides an ideal channel for reactant diffusion and electron transport.26–28 The doping of Mo in δ-MnO2 facilitates electron injection from electron-deficient Mo sites into the antibonding orbitals of N2, and therefore, directly promotes the cleavage of the strong N
N triple bond.13
To investigate the constitutive relationship and application potential of highly efficient electrocatalytic nitrate reduction (NOR) materials, five types of δ-MnO2 nanoflowers with different Mo doping concentrations were synthesized by the hydrothermal method in this study. Through combined XPS and DFT analyses, it was demonstrated that Mo doping could modulate the electronic structure and active sites of δ-MnO2, enhance the electrical conductivity, promote an increase in the proportion of low-valent Mn species, and enhance the adsorption capacity and reduce the thermodynamic barrier of the NOR through electron cloud density reconfiguration. The electron cloud density reconstruction enhances the N2 adsorption capacity and lowers the thermodynamic barrier of the NOR, showing a NOR-specific gradient mechanism (Mo doping → increase in low-valent Mn → electron cloud reconstruction → enhanced N2 adsorption). The Zn–N2 battery system was further developed to achieve the synergistic coupling of the NOR and power output, and its NO3− yield is higher than that of the traditional H-type electrolytic cell, and it enables the dual functionality of “nitrate synthesis-energy storage” through the cell discharge. This study offers novel insights and technical approaches for efficient NOR material design and distributed green nitrate synthesis.
![]() | ||
| Fig. 2 (a) SEM images of MM2.5. (b and c) TEM images of MM2.5. (d) HRTEM images of MM2.5. (e) Lattice stripe diagram of MM2.5. (f) EDX elemental mapping images of Mn, Mo and O for MM2.5. | ||
The X-ray diffraction (XRD) patterns, as shown in Fig. 3a, reveal that all samples display typical characteristic peaks of δ-MnO2 at 12.55° (001), 25.25° (002), 37.31° (−111), and 65.62° (020) (JCPDS #80-1098). Notably, when the Mo doping amount was ≤5%, there were no significant shifts in the positions and intensities of the diffraction peaks, indicating that low-concentration doping did not disrupt the main crystal structure. As shown in Fig. S1, when the doping amount was raised to 10% (MM10), new diffraction peaks appeared at 26.3°, 37.2°, and 54.8°, which matched with the monoclinic phase MoO2 (JCPDS #32-0671), confirming that excessive doping led to the formation of a molybdenum oxide heterogeneous phase.
![]() | ||
| Fig. 3 (a) XRD patterns of MM0 and MM2.5. (b and c) Nitrogen adsorption–desorption isotherms and pore size distribution curves for MM0, MM2.5 and MM10. (d–f) XPS spectra of Mn 2p, Mo 3d and O 1s. | ||
The specific surface areas of MM0, MM2.5 and MM10 were determined by the Brunauer–Emmett–Teller (BET) method (Fig. 3b). The results show that the three samples can be categorized as those with typical II isotherms with hysteresis loops, indicating the presence of mesoporous structures.29 Among them, the moderately optimized sample MM2.5 possessed a specific surface area of 76.43 m2 g−1, which is significantly higher than those of the undoped sample MM0 (57.46 m2 g−1) and the highly doped sample MM10 (21.22 m2 g−1). This indicates that moderate doping of Mo (MM2.5) increases the specific surface area of δ-MnO2, while the decrease in the specific surface area of the sample with excessive addition of Mo (MM10) may be due to the generation of new phases of MoO2, which is consistent with the XRD analysis results.30,31 More detailed information about the pore structure of Mo-doped δ-MnO2 was obtained using the pore size distribution curve determined by the DFT method (Fig. 3c). The pore size distribution curves show that all three samples have a mesoporous structure of 3–8 nm, with a greater proportion of MM2.5 pores located in this range. This hierarchical porous characteristic is favorable for increasing the exposure density of active sites and facilitating the electrolyte permeation and mass transfer process.32
The high-resolution XPS characterization system revealed the modulation mechanism of doped Mo on the electronic structure of the material surface. The full-spectrum XPS scanning results (Fig. S2) clearly indicate the presence of characteristic peaks corresponding to Mn and O in all synthesized samples, suggesting the consistent retention of the base material composition. Notably, in the series of Mo-doped samples (MM1, MM2.5, MM5, and MM10), well-defined Mo 3d photoelectron peaks emerged within the binding energy range of 230–238 eV. This spectral feature aligns precisely with the anticipated values for the Mo6+ oxidation state, confirming the successful incorporation of molybdenum into the lattice structure. Conversely, the undoped Mo sample (MM0) shows no evidence of Mo 3d spectral features.
The high-resolution spectral analysis of the Mn 2p orbitals (Fig. 3d) shows that all samples exhibit the typical structure of two deconvolution peaks, Mn3+ (641.8 eV, 653.3 eV) and Mn4+ (643.4 eV, 654.7 eV), which correspond to the 2p3/2 and 2p1/2 spin orbitals, respectively.33 Mo doping changed the valence distribution of Mn, and the atomic ratios of Mn3+/Mn4+ were 1.002
:
1, 1.145
:
1, 1.227
:
1, 1.425
:
1, and 1.472
:
1 for samples MM0, MM1, MM2.5, MM5, and MM10, respectively, with a gradual increase in the percentage of Mn3+. This suggests that the addition of Mo promotes the shift of elemental manganese to the lower valence state; this change occurs moderately and is likely to greatly improve the N2 activation capacity of δ-MnO2.34 The Mn 2p peak positions of the MM2.5 to MM10 samples were negatively shifted by about 0.1, 0.3, 0.5, and 0.8 eV, respectively, compared with that of MM0.
As shown in Fig. 3e, the O 1s spectrum of MM0 can be decomposed into three characteristic peaks at 529.5, 531.8, and 533.2 eV, corresponding to Mn–O–Mn, Mn–O–H and H–O–H, respectively.35 The peak positions of Mo-doped samples show negative shifts similar to those of the Mn 2p peaks as compared with that of MM0, which may be due to electron transfer effects induced by the high electronegativity of Mo6+, resulting in the negative shift of the Mn 2p peak position of MM10 samples. These changes may be due to the electron transfer effect caused by the high electronegativity of Mo6+, resulting in an increase in the electron cloud density at the Mn and O sites.36,37 It is noteworthy that the peak of the Mo–O bond (530.1 eV) can also be obtained in the O 1s spectrum of MM10, which further suggests that the excess addition of Mo elements leads to the generation of MoO2, which is consistent with the XRD results.38
Fine spectral analysis of the Mo 3d orbitals (Fig. 3f) confirms that the characteristic double peaks at 231.6 eV (3d5/2) and 234.8 eV (3d3/2) are exclusively observed in the Mo-doped sample (MM2.5).39 This finding is consistent with the typical characteristics of Mo6+ and suggests that Mo is stably present in the δ-MnO2 lattice with substitutional doping. In contrast, these characteristic peaks were not detected in the undoped pristine MM0 sample. Notably, unlike the pronounced negative shift in the Mn 2p peaks, all Mo 3d doublet peaks show a consistent but obvious positive binding energy shift. This shift likely originates from a decrease in localized electron density around Mo, possibly due to charge transfer or changes in the coordination environment.40,41
The NOR activity of MM2.5 was evaluated by linear scanning voltammetry (LSV) in N2-saturated and Ar-saturated 0.1 M KOH electrolytes, respectively. As shown in Fig. 4c, the current density of MM2.5 increased significantly in the N2-saturated electrolyte, confirming its significant NOR activity, and the potential interval of 1.8–2.2 V (vs. RHE) was chosen as the range for further NOR tests. The chronoamperometry test (I–T) curves depicted in Fig. 4d revealed that within the potential range of 1.8–2.2 V, the current densities of MM2.5 increased substantially as the voltage rose (LSV and I–T curves for the remaining four samples are shown in Fig. S4). The quantitative analysis in Fig. 4e shows that the NO3− yield of MM2.5 is 116.75 μg h−1 mgcat−1 at 2.0 V, which is about 2.4 times that of the unmodified sample (MM0), while the FE of MM2.5 is 7.04%, which is up to 2.8 times that of MM0. Upon analyzing the by-products, as shown in Fig. 4f, it was revealed that there were no detectable NOx species within the gas phase products. The FE values of MM0 and MM2.5 for O2 were 78.86% and 79.37%, respectively. The liquid products H2O2 and NO2− were determined by the standard curves as shown in Fig. S5. MM2.5 showed a decreased H2O2 yield of 13.47%, compared to the 17.06% for MM0, while the yield of NO2− only showed a slight decrease (0.71% → 0.63%). Therefore, it is reasonable to speculate that doped Mo can improve the selectivity of the NOR by inhibiting the pathway of H2O2 generation. Furthermore, the inherent characteristics of the NOR were validated through systematically conducted controlled experiments, as illustrated in Fig. 4g. The results indicated that when a voltage was applied and N2 was passed through, the blank CP produced virtually no NO3−. Conversely, the CP loaded with MM2.5 (M) generated a substantial amount of NO3− under the same conditions. At an open circuit potential (OCP), there was almost no NO3− at the MM2.5 electrode. Upon replacing N2 with Ar, almost no NO3− was detected at a voltage of 2.0 V. These results verified the origin of N in NO3− from the feedstock N2. When compared with the electrocatalytic NOR catalysts reported previously (Fig. 4h), MM2.5 demonstrated a higher NO3− yield and FE (detailed data comparisons along with relevant references are presented in Table S1).
The practical utility potential of MM2.5 was assessed via multidimensional stability tests. As illustrated in Fig. 5a–c, MM2.5 exhibited remarkable cycling stability, reproducibility, and long-term stability. Specifically, it endured five consecutive NOR test cycles, with each cycle lasting for 1 h, and the fluctuations in the NO3− yield and FE were less than 5%. For five batches of electrodes prepared independently, there was no noticeable decline in the NO3− yield, which verified the outstanding reproducibility of the synthesis method. When the electrodes were operated at a constant potential of 2.0 V versus RHE for 24 h, only a minimal decay in the current density was observed. This observation indicated the exceptional long-term durability of the MM2.5 electrode. Fig. 5d–f present the XPS spectra of Mn 2p, O 1s, and Mo 3d for MM2.5 before and after the NOR. It is evident that no significant shifts in peak positions were observed for any of the elements. This observation further supports the remarkable structural stability of MM2.5.
We developed a Zn–N2 battery system based on a Mo-doped δ-MnO2 (MM2.5) anode, which successfully realized the synergistic coupling of the NOR and electrical energy output (Fig. 6a and b). The device used MM2.5-modified composite carbon paper (CP) as the anode (the catalyst came in contact the with N2-saturated 0.1 M KOH electrolyte) and a zinc plate as the cathode (0.5 M Na2SO4 electrolyte), and separation of the two chambers was achieved using a bipolar membrane. MM2.5 was coated on the conductive side of the CP, and N2 was permeated from the water-resistant side. A peristaltic pump was connected to both sides of the anode chamber for better circulation of the electrolyte saturated with nitrogen to facilitate the NOR. The formation of a gas–liquid–solid three phase interface significantly improves the mass transfer efficiency of gaseous N2. During the charging process, the NOR occurs at the anode (N2 + 12OH− → 2NO3− + 10e− + 6H2O),39 accompanied by hydrogen precipitation at the cathode (2H2O + 2e− → H2↑ + 2OH−). The open circuit potential (OCP) of the system was found to be stable at 1.420 V using a digital multimeter (Fig. 6c), which is in great agreement with the results of the electrochemical workstation test.44 Electrochemical tests showed that the device voltage was quite stable over a wide current range of 0.25 to 1.25 mA cm−2 (Fig. 6d), and the NO3− yield steadily increased with increasing current density, reaching 144.5 μg h−1 mgcat−1 at 0.7 mA cm−2, which is an enhancement of about 23.7% over the yield of the conventional H-type cell (Fig. 7d). Discharge mode was achieved by NO3− cathodic reduction (NO3− + 7H2O + 8e− → NH4OH + 9OH−) with Zn anodic oxidation (Zn → Zn2+ + 2e−) to realize chemical–electrical energy conversion,40 which can continuously light up 22 LED lamps (Fig. 6e), verifying the utility of its dual function (nitrate synthesis and energy storage). This study provides an innovative technological path for distributed nitrate green synthesis and residual power recovery.
Based on the differential charge density information and the data in the literature,13 as shown in Fig. 7c, three characteristic sites, the Mn atom active site (Mn site) on the pristine δ-MnO2 (001) surface, the Mo atom active site (Mo site) on the Mo-doped surface, and the Mn atom active site (MnMo site) coordinated with Mo on the Mo-doped surface, were selected for Crystal Orbital Hamilton Populations (COHP) analysis to analyze the strength of the interaction between the active sites and the N atoms. The anti-bonding (grey area) and bonding states (white area) correspond to the negative and positive values of –ICOHP, respectively. The metal–N bond strength was assessed from the magnitude of the integral of the –ICOHP value. Fig. 7d reveals that the chemical bonding between the Mo site and the adsorbed N2 molecules exhibited a remarkably negative COHP integral value of −2.645 eV, which was about two times higher than that of the original Mn site (−1.189 eV). This change implies that Mo doping significantly strengthened the antibonding orbitals between the catalyst surface and the terminal nitrogen atoms of N2.45 Such a strong interaction is capable of effectively reducing the bonding energy of the N
N triple bond, thereby decreasing its dissociation energy barrier. Notably, the COHP integral value of the MnMo site, which was −1.410 eV, lay between the values of the Mo site and the original Mn site. This indicates that the doped Mo atoms can enhance the activation ability of the adjacent Mn atoms toward N2via the electronic modulation effect.
In addition, we explored the crucial modulation of the adsorption energies of N2 molecules, the lengths of the N
N bonds and the potential determining step (PDS) intermediates, specifically the intermediate NNOH, by different adsorption sites on the catalyst surface, namely the Mn site, the MnMo site, and the Mo site (Fig. 7e). The computational results demonstrate that the Mo site exhibits superior synergistic catalytic effects. Notably, the Mo site induces the most pronounced bond elongation in both N2 (1.153 Å) and NNOH (1.154 Å) species, compared to the reference bond length of free N2 molecules (1.102 Å). This substantial lengthening of the N
N triple bond indicates a significant weakening of bond strength, which effectively reduces the activation energy barrier required for nitrogen bond cleavage. Simultaneously, the Mo active center demonstrates the most pronounced adsorption energies for both N2 (−0.517 eV) and NNOH (−0.628 eV), reflecting its superior thermodynamic stabilization of these intermediates. This strong adsorption capability directly facilitates the reduction of energy barriers in the rate determining step of the reaction pathway.46 In contrast, Mn sites exhibit weaker interactions, with adsorption energies of −0.349 eV (*N2) and −0.485 eV (*NNOH), accompanied by less effective N
N bond elongation (1.114 Å and 1.124 Å, respectively). Interestingly, under Mo doping modulation, the Mn sites show marginally improved performance in both bond stretching (1.122 Å and 1.126 Å for N2 and NNOH) and adsorption energies (−0.352 eV and −0.507 eV) compared to undoped Mn. This enhancement correlates with the increased electron-deficient character of Mn atoms when coordinated to Mo, which likely promotes the critical nitrogen–oxygen bond activation step.
The density of states (DOS) analysis shows (Fig. 7f) that Mo doping introduces new electronic states near the Fermi level, effectively bridges a part of the band gap of δ-MnO2, and enhances the electrical conductivity of the catalysts, which is in agreement with the EIS results.25,47 Crucially, Mo doping induces a 0.097 eV upward shift in the material's d-band center, from −4.836 eV in pristine δ-MnO2 to −4.739 eV, positioning it closer to the Fermi level. This positive d-band shift aligns with d-band center theory, which posits that such electronic modulation strengthens hybridization between catalyst surface d-orbitals and N2 π*-antibonding orbitals, thereby enhancing nitrogen adsorption capacity.48 The Gibbs free energy diagram of the NOR, constructed based on DFT calculations, as shown in Fig. 7g, reveals that both the δ-MnO2 (001) and Mo-doped δ-MnO2 (001) surfaces adhere to the sequential “+OH, −H” process:
| This journal is © the Partner Organisations 2026 |