Electrocatalytic nitrate reduction on Fe, Fe3O4, and Fe@Fe3O4 cathodes: Elucidating structure-sensitive mechanisms of direct electron versus hydrogen atom transfer

Yuwei Liu Yihui Zhu Weijian Duan Yizhuo Yang Haorui Tuo Chunhua Feng

Citation:  Yuwei Liu, Yihui Zhu, Weijian Duan, Yizhuo Yang, Haorui Tuo, Chunhua Feng. Electrocatalytic nitrate reduction on Fe, Fe3O4, and Fe@Fe3O4 cathodes: Elucidating structure-sensitive mechanisms of direct electron versus hydrogen atom transfer[J]. Chinese Chemical Letters, 2025, 36(6): 110347. doi: 10.1016/j.cclet.2024.110347 shu

Electrocatalytic nitrate reduction on Fe, Fe3O4, and Fe@Fe3O4 cathodes: Elucidating structure-sensitive mechanisms of direct electron versus hydrogen atom transfer

English

  • Nitrate, a highly soluble and toxic pollutant, has been increasingly detected in various water sources, including agricultural runoff, industrial wastewater, and drinking water. This pollutant is directly linked to the adverse eutrophication phenomenon and other serious environmental issues that threaten both ecosystem balance and human health [1-3]. The development of highly effective denitrification strategies is desirable and receives considerable attention. Among these, the electrochemical nitrate reduction reaction (NO3RR) emerges as a promising method to achieve this goal in a greener and more sustainable manner, due to the remarkable advantages of its renewable electricity-driven nature, application feasibility, and high efficiency [4-7]. Additionally, ammonia (NH3), as the principal product during NO3RR, has been widely proposed as the most potential candidate for a sustainable hydrogen energy carrier because of its high energy density (4.3 kW/kg) and economic affordability for storage and transport [8]. Therefore, the advancement of NO3RR technology has gain increasing attention from both the environmental and energy communities.

    Electrocatalysts play a crucial role in the emerging technology of nitrate reduction, as they significantly influence the activity and product distribution of NO3RR. A wide range of materials with unique structures and compositions has been explored to enhance the NH3 yield rate and Faradaic efficiency (FE) of NO3RR, including pure metals (e.g., Ru and Co) [9,10], metal alloys (e.g., CuPd) [11-13], metal oxides (e.g., Co3O4 and Fe3O4) [14,15], and carbon-based materials (e.g., single atomic catalysts) [16-18]. Despite considerable progress in developing materials for NH3 production via NO3RR, the critical relationship between the structure of catalysts and their performance in NO3RR remains unclear. This lack of clarity is due to insufficient understanding of the reaction process, which significantly impedes the rational design and advancement of catalysts for efficient NH3 production. It is generally recognized that NO3RR can proceed via either a direct electron transfer mechanism (DET) [19,20] or an indirect hydrogen atom (H) transfer (HAT) mechanism [21,22], with each expected to lead to different NO3RR performance. Previous studies have shown that a balance between the consumption and supply of H generated from the Volmer step during the hydrogen evolution reaction (HER) within the HAT mechanism is critical for superior NH3 yield via NO3RR [23]. An excessive supply of H may lead to a preferred occurrence of competitive HER, thereby reducing the FE of NH3. Theoretically, the DET mechanism does not involve hydrogen atoms as key intermediates for the parasitic HER, potentially leading to higher NO3RR efficiency due to reduced H generation. The dominance of a particular mechanism is highly dependent on the catalyst property. Therefore, a comprehensive understanding of the intricate relationship between catalyst structure, reaction mechanism, and NO3RR performance is of great importance.

    Herein, we designed and fabricated three types of free-standing electrocatalysts, building upon our previous work with minor modifications [24]. These catalysts included pure metallic Fe nanoparticles in-situ grown on iron foil (denoted as Fe/FF), pure Fe3O4 nanosheets on carbon cloth (Fe3O4/CC), and a mixture of metallic Fe nanoparticles and Fe3O4 nanosheets in-situ constructed on iron foil (Fe@Fe3O4/FF). By carefully examining and comparing their NO3RR performance in terms of NH3 yield, NH3 selectivity, and FE, we employed a combination of electrochemical characterizations, quenching experiments, electron spin resonance (ESR) measurements, and theoretical calculations to discern the differences in catalyst structure and reaction mechanisms. Through this comprehensive analysis, we were able to elucidate the structure-sensitive mechanisms of DET versus HAT and their impact on NO3RR performance. This investigation provides valuable insights into catalyst design for the electrochemical treatment of nitrate-laden wastewater, offering an important perspective on optimizing the efficiency and selectivity of the NO3RR process.

    The experimental section is elaborated in Supporting information, which includes chemicals and reagents, preparation of catalysts, experimental setup, physicochemical and electrochemical characterizations, analysis methods, and density functional theory (DFT) calculations. The electrodes used in this study were fabricated using a similar method as described previously [24], which involved initial annealing in a static air atmosphere followed by electrochemical activation. Briefly, Fe@Fe3O4/FF was prepared by annealing Fe foil at 500 ℃ in a static air atmosphere, followed by electrochemical cathode activation at 10 mA/cm2. Meanwhile, Fe/FF was produced using the same procedure but with the annealing temperature adjusted to 350 ℃. A two-step protocol was employed for synthesizing Fe3O4/CC, which consisted of electrodeposition of Fe(OH)3 on carbon cloth followed by air annealing at 350 ℃. Physicochemical characterization techniques including scanning electron microscopy (SEM), transmission electron microscopy (TEM), and diffractometry (XRD) were utilized to identify differences in composition and structure among the Fe/FF, Fe3O4/CC, and Fe@Fe3O4/FF cathodes. Fig. 1a and Fig. S1 (Supporting information) illustrate that the surface of Fe@Fe3O4/FF appeared rougher in comparison to the smoother surface of Fe3O4/FF prior to electrochemical treatment, with a high density of nanoparticles dispersed on the original Fe3O4 nanosheets. This observation aligns with our previous findings that electrochemical reduction treatment facilitates the formation of Co nanoparticles from Co3O4 nanosheets on cobalt foil (CF) under the influence of an applied cathode potential [24]. Fig. 1b shows that Fe3O4/CC exhibited flat nanosheet morphology, similar to Fe3O4/FF before electrochemical treatment. A lower calcination temperature of 350 ℃ during Fe/FF synthesis resulted in fewer Fe3O4 nanosheets forming on FF (Fig. S2 in Supporting information), in contrast to the abundant and dense formation of Fe3O4 nanosheets on Fe3O4/FF (calcination at 500 ℃), which were then entirely reduced to metallic Fe without any remaining Fe3O4 through the same electrochemical treatment, as depicted in Fig. 1c. Moreover, TEM analyses (Figs. 1d and e) further supported these findings, revealing lattice fringes with interplanar spacings of 0.25 nm and 0.215 nm, corresponding to the (311) plane of Fe3O4 and the (110) plane of Fe, respectively, indicating the coexistence of metallic Fe and Fe3O4 in Fe@Fe3O4/FF. However, for Fe3O4/CC, only Fe3O4 lattice fringes were observed (Fig. S3 in Supporting information). The composition of the three electrocatalysts was also confirmed through XRD analysis (Fig. 1f), where crystalline phases of Fe and Fe3O4 were identified for Fe/FF and Fe3O4/CC, respectively. In contrast, Fe@Fe3O4/FF exhibited crystalline phases of both Fe and Fe3O4, suggesting that Fe@Fe3O4/FF comprises both metallic Fe and Fe3O4 nanosheets.

    Figure 1

    Figure 1.  SEM images of (a) Fe@Fe3O4/FF, (b) Fe3O4/CC, and (c) Fe/FF. (d, e) HRTEM images of Fe@Fe3O4/FF. (f) XRD patterns of three electrocatalysts.

    To compare the effectiveness in nitrate removal and NH3 production among the three specially prepared cathodes and a bare FF cathode, the chronoamperometry method was utilized for evaluation. Fig. S4 (Supporting information) reveals that after 6 h of operation at −1.3 V (vs. SCE) with 100 mg/L NO3-N, the nitrate removal efficiencies were 100%, 56.7%, and 33.6% for Fe@Fe3O4/FF, Fe/FF, and Fe3O4/CC, respectively. This data indicates the superior efficacy of Fe@Fe3O4/FF in NO3RR, outperforming the other tested cathodes. Of particular note, all NH3 yield rates were normalized against the electrochemically active surface area (ECSA), which represents the maximum number of available active sites during electrolysis. The ECSA was measured using the well-established double-layer capacitance method, as detailed in Figs. S5 and S6 (Supporting information). As illustrated in Figs. 2a and b, the Fe@Fe3O4/FF cathode exhibited significantly higher NH3 yield performance across all tested potentials compared to its counterparts. It is noticeable that the NH3 yield rate for Fe@Fe3O4/FF increased progressively from 0.056 mmol h−1 cm−2 to 0.28 mmol h−1 cm−2 as the potential shifted more negatively from −1.1 V to −1.6 V (vs. SCE), achieving a high FE of 95.7% even at the high overpotential of −1.6 V (vs. SCE). This highlights the exceptional capability of Fe@Fe3O4/FF in facilitating NH3 production via NO3RR. In contrast, the peak NH3 yield rates and FE at −1.6 V (vs. SCE) for the other cathodes were significantly lower, recording 0.10 mmol h−1 cm−2/53.0%, 0.14 mmol h−1 cm−2/82.7%, and 0.076 mmol h−1 cm−2/44.1% for the FF, Fe/FF, and Fe3O4/CC, respectively, which were all markedly inferior to the performance of Fe@Fe3O4/FF. In addition, the NH3 yield response to varying nitrate concentrations exhibited distinct patterns for Fe/FF, Fe3O4/CC, and Fe@Fe3O4/FF, as depicted in Fig. 2c, where each catalyst showed a volcano-shaped yield curve, albeit with different nitrate concentration turning points. The NH3 yield rates for Fe3O4/CC and Fe@Fe3O4/FF followed a similar trend, increasing with nitrate concentration from 25 mmol/L to 300 mmol/L before showing a decline, with peak NH3 yields of 0.11 mmol h−1 cm−2 and 0.23 mmol h−1 cm−2, respectively. Fe/FF exhibited a decrease in the NH3 yield rate at a lower nitrate concentration of 100 mmol/L, indicating its higher sensitivity to nitrate concentration. The suppression of NH3 yield at high nitrate concentrations occurred likely because the nitrite produced could not be continuously reduced to ammonia. This was due to the high concentration of nitrate in the bulk solution, which competed with the intermediate nitrite for electrons on the cathode.

    Figure 2

    Figure 2.  Effects of applied potential on (a) NH3 yield rate and (b) FE of NH3, and (c) the impact of nitrate concentration on NH3 yield rate across all the investigated cathodes. Nitrate concentration-dependent FE of various products during NO3RR for (d) Fe@Fe3O4/FF, (e) Fe3O4/CC, and (f) Fe/FF. Unless otherwise specified, reaction conditions are as follows: applied potential of −1.3 V (vs. SCE), 100 mmol/L Na2SO4, 100 mmol/L NaNO3, and 1 h of electrolysis.

    As illustrated in Figs. 2d-f, a remarkable increase in the accumulation of the undesired nitrite byproduct was observed for Fe/FF, with the FE(NO2) values increasing across nitrate concentrations, reaching up to 69.3% at 500 mmol/L nitrate and −1.3 V (vs. SCE). Meanwhile, the maximum FE(NO2) for Fe@Fe3O4/FF and Fe3O4/CC was significantly lower, indicating a more controlled nitrite formation during NO3RR, especially at elevated nitrate concentrations. The values of FE(H2) resulting from the competitive HER were also examined. It is revealed that this competitive reaction predominantly occurred at lower nitrate concentrations, with the Fe3O4/CC cathode more inclined towards favoring HER. Specifically, at 25 mmol/L nitrate, the FE(H2) values were ranked as follows: Fe3O4/CC (44.42%) > Fe@Fe3O4/FF (13%) > Fe/FF (2%). Overall, the FE(NH3) of Fe@Fe3O4/FF outperformed Fe3O4/CC and Fe/FF under all tested conditions, highlighting its superior ability for the selective production of NH3.

    Comprehensive evidence was presented to elucidate the mechanism underlying the exceptional performance of Fe@Fe3O4/FF in NO3RR, particularly its remarkable NH3 yield and substantially decreased nitrite production. The analysis involved three cathodes, Fe@Fe3O4/FF, Fe/FF, and Fe3O4/CC, and a control electrode, Fe3O4/nickel foam (NF), which was prepared in a manner analogous to Fe3O4/CC but utilized NF as the substrate. Figs. 3a-c and Fig. S7 (Supporting information) show the effects of tertiary butanol (TBA), a recognized scavenger of H, on the linear sweep voltammetry (LSV) responses of various electrodes. The differences in current density reduction observed in the LSV curves upon the addition of TBA implied the varied influence of H in NO3RR among various electrodes. Specifically, Fe@Fe3O4/FF exhibited a reduction in current density from 12.32 mA/cm2 to 10.93 mA/cm2 at −1.3 V (vs. SCE) in the presence of TBA, indicating a current density inhibition of 10.1%. Meanwhile, the suppression ratios of current density after TBA addition for the other electrodes were significantly higher, at 45.2% for Fe/FF, 77.4% for Fe3O4/CC, and 84.7% for Fe3O4/NF, respectively. This differential response suggests a predominant role of H in the NO3RR for these electrodes, while DET may primarily facilitate NO3RR on Fe@Fe3O4/FF.

    Figure 3

    Figure 3.  Current density variations in LSV for (a) Fe@Fe3O4/FF, (b) Fe/FF, and (c) Fe3O4/CC electrodes following the addition of 300 mmol/L TBA and 100 mg/L NaNO3. (d) Impact of TBA concentration on the nitrate removal efficiency across four electrodes, and (e) corresponding calculations of the contributions from different mechanisms involved in NO3RR. (f) Correlation between the NH3 yield rate and the proportion of DET in NO3RR, as determined from (e). (g) Measurement of the OH adsorptive peak in 1 mol/L NaOH solution. (h) EIS spectra of the three electrodes at −1.1 V (vs. SCE) in the presence of 100 mg/L NaNO3. (i) Influence of different alkali metal cations in the electrolyte on the NH3 yield rate for the three electrodes, evaluated in a solution of 100 mmol/L M2SO4 and 100 mmol/L MNO3 at −1.3 V (vs. SCE).

    Further investigation into the reaction mechanisms focused on evaluating nitrate removal efficiency in the presence of varying concentrations of TBA. Fig. 3d shows that Fe@Fe3O4/FF had minimal sensitivity to changes in TBA concentration, with removal efficiency slightly diminishing from 99.5% to 80.2% as the molar ratio of TBA to nitrate (C(TBA)/C(N)) was elevated from 0 to 50. This modest decline suggests a relatively low H contribution of 19.4% to the NO3RR process for Fe@Fe3O4/FF, as illustrated in Fig. 3e. In contrast, the other electrodes experienced a more marked decrease in nitrate removal efficiency with the same increase in TBA concentration, with efficiencies dropping to 59.9%, 38.0%, and 19.5% for Fe/FF, Fe3O4/CC, and Fe3O4/NF, respectively. This pattern, coupled with the distinct reduction in current density observed in LSV curves after TBA was added, infers H contributions of 38.5%, 60.5%, and 88.6% to NO3RR for Fe/FF, Fe3O4/CC, and Fe3O4/NF, respectively. In addition, after performing reactions with a C(TBA)/C(N) ratio of 50 for all electrodes, the nitrate removal efficiency was reevaluated without the addition of TBA. Interestingly, the reassessment resulted in efficiency values that were almost identical to those recorded in the initial tests under the same conditions for all electrodes. This consistency indicates that the inhibitory effect of TBA stems from its scavenging activity rather than a poisoning effect on the electrodes [25,26]. This conclusion highlights the distinctive pathways through which each electrode facilitates NO3RR, with Fe@Fe3O4/FF exhibiting a particularly efficient and less H-dependent reaction mechanism.

    The ESR technology was utilized to semi-quantitatively measure the amount of electrochemically generated H across the four electrodes under identical conditions. As illustrated in Fig. S8 (Supporting information), the characteristic peaks for 5,5-dimethyl-1-pyrroline N-oxide (DMPO)-H, a marker for H, were detected for all four electrodes, albeit with varying signal intensities. Fe3O4/NF exhibited the highest peak intensity of DMPO-H, followed by Fe3O4/CC, Fe/FF, and Fe@Fe3O4/FF, respectively. This order aligns with the previously determined H contributions to the NO3RR, again delineating the different predominant reaction mechanisms for each electrode. Combining these findings with earlier experimental results, it is conclusively determined that Fe@Fe3O4/FF primarily facilitates NO3RR via the DET mechanism. In contrast, HAT predominantly drives the reaction for Fe3O4/CC and Fe3O4/NF, while both DET and HAT mechanisms are active in Fe/FF. Intriguingly, when correlating the NH3 yield rate via NO3RR at −1.3 V (vs. SCE) with the contribution of the DET mechanism, a nearly linear relationship appeared (Fig. 3f). This suggests that the superior NH3 yield rate and substantially reduced nitrite production observed for Fe@Fe3O4/FF may be attributed to the dominance of the DET mechanism and minimal involvement of HAT. Therefore, a significant contribution from the DET mechanism is fundamentally crucial for the exceptional performance of Fe@Fe3O4/FF in achieving high NH3 generation.

    Based on the established relationship between reaction mechanisms and NO3RR performance, the next step was to elucidate how the inherent properties of electrocatalysts influence these mechanisms. It is hypothesized that the strong adsorption of NO3 and other nitrogen-containing intermediates onto the surface of electrocatalysts is crucial for the NO3RR to proceed via the DET mechanism [20,27,28], thereby leading to enhanced NO3RR performance. To test this hypothesis, the adsorption of OH serving as a proxy for NO3 and other nitrogen-containing intermediates generated during NO3RR was measured using the cyclic voltammetry (CV) technology. The results in Fig. 3g show that Fe@Fe3O4/FF exhibited the strongest OH adsorption, as indicated by the most negative potential of the OH adsorptive peak at −0.813 V (vs. SCE). This suggests a potentially strong bond between nitrogen-containing species and Fe@Fe3O4/FF, facilitating direct electron transfer. In comparison, the OH adsorptive peaks for Fe/FF and Fe3O4/CC were observed at −0.774 and −0.755 V (vs. SCE), respectively, indicating a better adsorption affinity for nitrogen-containing species on Fe/FF than on Fe3O4/CC, but still inferior to Fe@Fe3O4/FF. No peak corresponding to OH adsorption was detected on Fe3O4/NF, indicating inferior nitrate adsorption for Fe3O4/NF. The measured order of OH adsorptive peak potentials correlated well with the contribution of the DET mechanism in NO3RR for the electrodes, suggesting that enhanced adsorption of nitrogen-containing species is favorable for NO3RR via the DET mechanism. Moreover, the interfacial charge transfer resistance for Fe@Fe3O4/FF, Fe/FF, Fe3O4/CC and Fe3O4/NF was determined using electrochemical impedance spectroscopy (EIS) (Fig. 3h). As anticipated, Fe@Fe3O4/FF displayed the lowest resistance to charge transfer, as evidenced by the smallest diameter of the corresponding semi-circle in the EIS spectra. This result supports the most favorable interfacial electron transfer for Fe@Fe3O4/FF, followed by Fe/FF, Fe3O4/CC and Fe3O4/NF, and aligns with the previously mentioned contribution of the DET mechanism in NO3RR.

    The role of protonation kinetics, alongside enhanced adsorption of nitrogen-containing species, is also crucial for achieving high NH3 yield, especially given the nature of electrochemical hydrogenation in aqueous media. This study examined the protonation kinetics of three electrocatalysts by altering the alkali metal cation of the electrolyte from Li+ to Na+, K+, and Cs+, a method thoroughly documented in existing literature [29-31]. Fig. 3i shows that Fe@Fe3O4/FF, Fe3O4/CC, and Fe3O4/NF all exhibited progressively improved NH3 yield rates with the transition from Li+ to Cs+, increasing from 0.074, 0.03, and 0.0058 mmol h−1 cm−2 to 0.15, 0.08, and 0.02 mmol h−1 cm−2, respectively. This improvement indicates the favorable protonation kinetics of Fe@Fe3O4/FF, Fe3O4/CC, and Fe3O4/NF, consistent with trends reported previously [30,32]. In contrast, Fe/FF demonstrated a minimal improvement in NH3 yield rate, from mmol h−1 cm−2 0.03 to 0.04 mmol h−1 cm−2 under the same conditions, indicating weaker protonation kinetics and manifesting the critical role of Fe3O4 in enhancing NO3RR protonation. Interestingly, the superior protonation kinetics of Fe@Fe3O4/FF and Fe3O4/CC were closely associated with a significantly reduced accumulation of nitrite, as observed in Figs. 2d-f. This suggests that the presence of Fe3O4 in these electrocatalysts facilitates the continuous protonation of intermediate nitrite into NH3. This effect is likely due to the inherent ability of common metal oxides, such as Fe3O4, to adsorb and dissociate water effectively [33,34].

    DFT calculations were employed to deepen the understanding of the relationship between catalyst structure and NO3RR performance, particularly focusing on the advantages of the DET mechanism over the HAT mechanism. Three representative models (Fig. 4a), Fe(211) for Fe/FF, Fe3O4(220) for Fe3O4/CC, and Fe@Fe3O4(220) for Fe@Fe3O4/FF, were developed based on the experimental characterizations presented in Fig. 1, to closely reflect the structures of the actual catalysts. The adsorption free energy of NO3 on these models was calculated to identify which structure favored NO3 binding, a significant aspect in determining the reaction mechanism [35]. The results, shown in Fig. 4b, indicate that Fe@Fe3O4(220) exhibited the highest ΔG for NO3 adsorption at −0.772 eV, followed by −0.526 eV for Fe(221) and −0.415 eV for Fe3O4(220). This suggests that the Fe component in Fe@Fe3O4/FF plays a pivotal role in enhancing NO3 adsorption, thereby contributing to its superior NO3RR performance. The slightly stronger NO3 adsorption on Fe@Fe3O4(220) compared to Fe(221) also highlighted the beneficial effect of the coexisting Fe3O4 in Fe@Fe3O4/FF, aligning with the experimentally observed trends in NO3 adsorption. In addition, the kinetic energy barrier for the initial step of NO3RR (NO3 → NO2), often considered the rate-limiting step [36], was evaluated for both HAT and DET mechanisms. The HAT mechanism involves the direct hydrogenation of adsorbed NO3 (*NO3) by adjacent H, while DET facilitates this transformation through sequential proton and electron transfer to adsorbed NO3 without the involvement of neutral H. The energy profiles for NO3 transformation into NO2 + OH via both mechanisms were analyzed for all three models (Figs. 4c and d). Notably, Fe3O4(220) exhibited the lowest energy barrier of 1.01 eV for the HAT mechanism, followed by 1.28 eV for Fe(221) and 1.70 eV for Fe@Fe3O4(220), consistent with the experimentally measured contributions of the HAT mechanism. Conversely, Fe@Fe3O4(220) presented the lowest energy barrier of 0.42 eV for the DET mechanism, followed by 0.63 eV for Fe(221) and 0.99 eV for Fe3O4(220), theoretically supporting the preference for NO3RR via DET on Fe@Fe3O4/FF and suggesting a significant contribution of the HAT mechanism for NO3RR over Fe3O4/CC. These theoretical outcomes confirm the DET mechanism's favorable position relative to the HAT mechanism, especially since Fe@Fe3O4/FF, dominated by DET, exhibited the best nitrate conversion capability among the three investigated electrodes. This comprehensive analysis not only corroborates the experimental observations but also provides a mechanistic understanding of why Fe@Fe3O4/FF outperforms its counterparts in NO3RR, highlighting the critical role of catalyst structure in dictating the reaction pathway and efficiency.

    Figure 4

    Figure 4.  (a) Representations of Fe/FF, Fe3O4/CC, and Fe@Fe3O4/FF models used for theoretical calculations. (b) Comparative analysis of NO3 adsorption free energy across the three models. Evaluation of the kinetic energy barrier for (c) HAT and (d) DET mechanisms on each of the models.

    The collaborative experimental and theoretical analyses provided profound insights into the mechanics and effects of different components within electrocatalysts on NO3RR. Fig. 5 encapsulates these insights, highlighting how the interplay between metallic Fe and Fe3O4 within Fe@Fe3O4/FF enhances the NO3RR process. The role of metallic Fe is pivotal due to its strong adsorption capabilities for nitrogen-containing species, facilitated by intensive D-π interactions between the electron-rich metallic Fe surface and N atoms [37]. This interaction fosters a preference for the DET mechanism in NO3RR, attributed to reduced interfacial resistance for charge transfer. Concurrently, the inclusion of Fe3O4 in Fe@Fe3O4/FF is advantageous for promoting favorable protonation kinetics, thanks to its high affinity for water adsorption and dissociation. This is attributed to the strong interaction between the empty d orbitals of the metal oxide and the lone-pair electrons of the oxygen atom in water molecules [33,38]. This characteristic significantly mitigates nitrite accumulation and boosts NH3 production. In contrast, while Fe/FF also demonstrates robust nitrogen-containing species binding and low interfacial charge transfer resistance, its sole reliance on metallic Fe without the presence of Fe3O4 limits its protonation kinetics. This limitation could lead to considerable nitrite accumulation, especially at higher nitrate concentrations. Fe3O4/CC, characterized by effective protonation kinetics, however, exhibits poor NH3 yield and slightly reduced nitrite accumulation, primarily influenced by the HAT mechanism. Therefore, the synergy between Fe and Fe3O4 in Fe@Fe3O4/FF is crucial for achieving superior NH3 yield through the DET mechanism. This insight is useful for the design and development of highly effective electrocatalysts for NO3RR, emphasizing the need to balance components that enhance nitrogen-containing species adsorption and promote efficient protonation kinetics to minimize undesirable byproducts and maximize desired outcomes.

    Figure 5

    Figure 5.  Schematic representation of reaction mechanisms on (a) Fe@Fe3O4/FF, (b) Fe/FF, and (c) Fe3O4/CC.

    The performance of Fe@Fe3O4/FF for complete denitrification under various conditions, including the presence of NaCl, was thoroughly assessed. The addition of NaCl aimed to initiate electrochemically mediated breakpoint chlorination, oxidizing the NH4+ generated from NO3RR into harmless N2 [39-42]. Fig. S9a (Supporting information) shows that 500 mg/L NO3-N can be fully removed at 30 mA/cm2 after 3 h of electrolysis, achieving a pseudo-first-order rate constant of 1.39 h−1. This rate constant was observed to increase incrementally to 1.43, 1.49, and 1.52 h−1 as the current density was raised to 35, 40, and 45 mA/cm2, respectively. Negligible amounts of NH4+-N and NO2-N were detected after the electrolysis process, indicating the capability of Fe@Fe3O4/FF for complete denitrification with the assistance of NaCl.

    The impacts of initial NO3-N concentration and NaCl concentration on the denitrification efficiency of Fe@Fe3O4/FF were explored, as presented in Figs. S9b and c (Supporting information). An observed trend showed a gradual decline in the kinetics as the initial NO3-N concentration increased from 500 mg/L to 800 mg/L, attributed to the limitation in active sites [43]. Nevertheless, nearly 100% denitrification efficiency was still achieved after 3 h of electrolysis, even at the elevated NO3-N concentration of 800 mg/L. Complete removal of 500 mg/L NO3-N was observed, irrespective of the NaCl amount. The denitrification kinetics exhibited a slight improvement with increasing concentrations of NaCl. This enhancement may be attributed to the refreshment of active sites occupied by the NH3 product and the alleviation of pH increase in the solution through electrochemically mediated breakpoint chlorination. This suggests a strong resistance to oxidative corrosion by active chlorine species produced from Cl oxidation at the anode [41,42]. The Fe@Fe3O4/FF cathode maintained exceptional denitrification efficiency across a pH range from 7 to 12 (Fig. S9d in Supporting information), suggesting the promising potential for practical applications. The acidic conditions were not examined in the present study due to inevitable dissolution of Fe3O4 under these conditions. Material stability, a critical performance metric, was evaluated over 20 cycles. As shown in Fig. S10 (Supporting information), the Fe@Fe3O4/FF cathode showed consistent denitrification efficiency of approximately 100% removal across all cycles, indicating its exceptional robustness. The post-XRD pattern of Fe@Fe3O4/FF in Fig. S11 (Supporting information) displayed a profile similar to that of the pristine Fe@Fe3O4/FF. These findings highlight the effectiveness and durability of Fe@Fe3O4/FF in complex water treatment scenarios, offering a promising approach to achieving complete denitrification in the presence of challenging constituents like NaCl.

    Given the advantages of Fe@Fe3O4/FF, such as cost-effectiveness and ease of scale-up, the practical feasibility of using Fe@Fe3O4/FF as a cathode material for treating real wastewater with high nitrate concentrations was investigated. To facilitate complete denitrification through electrochemically mediated breakpoint chlorination, an additional 100 mmol/L NaCl was supplemented into the wastewater. An experimental setup featuring an internal circulation mode was constructed, as depicted in Fig. S12a (Supporting information). This setup was operated in an internal-circulation mode, including a peristaltic pump for flow rate control, an electrolyzer equipped with a larger-sized Fe@Fe3O4/FF cathode (4 cm × 3 cm) for denitrification, and a container for storing and mixing 500 mL of the original wastewater with the effluent from the electrolyzer. Operating at a current density of 30 mA/cm2 and a flow rate of 1 mL/min, the NO3-N concentration in the wastewater was steadily reduced from 3815 mg/L to zero over 120 h of electrolysis, as shown in Fig. S12b (Supporting information). The byproducts of NH4+-N and NO2-N were scarcely accumulated by the end of the process, demonstrating the effectiveness of active chlorine species generated from Cl oxidation at the MMO anode in ensuring efficient denitrification. In stark contrast, when using Fe/FF as the cathode under identical operating conditions, the denitrification performance was significantly inferior, achieving only a 29.3% removal efficiency.

    In conclusion, through detailed investigations of NO3RR on Fe/FF, Fe3O4/CC, and Fe@Fe3O4/FF cathodes, we have clearly established that the catalyst structure directly influences the reaction mechanism and, consequently, the NO3RR performance. The Fe@Fe3O4/FF cathode demonstrated a strong affinity for NO3 and a moderate capability for water dissociation, being attributed to its metallic Fe and Fe3O4 components, respectively. These characteristics are key to the preferred occurrence of the DET mechanism during NO3RR, leading to the outstanding denitrification performance of Fe@Fe3O4/FF. Conversely, the absence of metallic oxide Fe3O4 in Fe/FF, which is crucial for the protonation process, resulted in poor NO3RR performance and significant accumulation of the intermediate product nitrite. Fe3O4/CC, containing only the metallic oxide Fe3O4 component, exhibited inferior denitrification performance. The lack of metallic Fe, which is essential for strong bonding with NO3, leads to a preference for the HAT mechanism during NO3RR. This preference accelerates the competitive HER and limits denitrification performance, despite a noticeable reduction in the accumulation of NO2 due to the enhanced protonation capability of Fe3O4. These experimental insights are crucial for a fundamental understanding of the DET and HAT mechanisms involved in NO3RR. They also provide a solid foundation for the rational design and development of advanced catalysts for effective nitrate reduction, highlighting the importance of carefully considering catalyst composition and structure to optimize performance.

    The authors declare that they have no known competing financial interests or personal relationships that could have appeared to influence the work reported in this paper.

    Yuwei Liu: Writing – original draft, Investigation, Formal analysis, Data curation. Yihui Zhu: Investigation, Formal analysis, Data curation. Weijian Duan: Writing – original draft, Methodology, Data curation, Conceptualization. Yizhuo Yang: Resources, Data curation. Haorui Tuo: Investigation, Data curation. Chunhua Feng: Writing – review & editing, Supervision, Methodology, Formal analysis, Conceptualization.

    We gratefully acknowledge financial support from the National Natural Science Foundation of China (Nos. U21A2034 and 21876052), the Guangdong Special Support Plan for Innovation Teams (No. 2019BT02L218), the Guangdong Special Support Plan for Young Top-notch Talents (No. 2019TQ05L179), and the Natural Science Foundation of Guangdong Province, China (No. 2021B1515120077).

    Supplementary material associated with this article can be found, in the online version, at doi:10.1016/j.cclet.2024.110347.


    1. [1]

      M. Duca, M.T.M. Koper, Energy Environ. Sci. 5 (2012) 9726–9742. doi: 10.1039/c2ee23062c

    2. [2]

      H. Xu, Y. Ma, J. Chen, W.X. Zhang, J. Yang, Chem. Soc. Rev. 51 (2022) 2710–2758. doi: 10.1039/d1cs00857a

    3. [3]

      Y. Xiong, Y. Wang, J. Zhou, et al., Adv. Mater. 36 (2024) 2304021. doi: 10.1002/adma.202304021

    4. [4]

      Y. Wang, C. Wang, M. Li, Y. Yu, B. Zhang, Chem. Soc. Rev. 50 (2021) 6720–6733. doi: 10.1039/d1cs00116g

    5. [5]

      W. Zheng, L. Zhu, Z. Yan, et al., Environ. Sci. Technol. 55 (2021) 13231–13243.

    6. [6]

      T. Zhu, Q. Chen, P. Liao, et al., Small 16 (2020) 2004526. doi: 10.1002/smll.202004526

    7. [7]

      Y. Ren, S. You, Y. Wang, J. Yang, Y. Liu, Environ. Sci. Technol. 58 (2024) 2144–2152. doi: 10.1021/acs.est.3c09759

    8. [8]

      J. Guo, P. Chen, Chem 3 (2017) 709–712. doi: 10.1016/j.chempr.2017.10.004

    9. [9]

      J. Li, G. Zhan, J. Yang, et al., J. Am. Chem. Soc. 142 (2020) 7036–7046. doi: 10.1021/jacs.0c00418

    10. [10]

      M. Fu, Y. Mao, H. Wang, et al., Chin. Chem. Lett. 35 (2024) 108341. doi: 10.1016/j.cclet.2023.108341

    11. [11]

      Y. Xu, K. Shi, T. Ren, et al., Small 18 (2022) 2203335. doi: 10.1002/smll.202203335

    12. [12]

      Y. Wang, A. Xu, Z. Wang, et al., J. Am. Chem. Soc. 142 (2020) 5702–5708. doi: 10.1021/jacs.9b13347

    13. [13]

      H. Yin, X. Mao, S. Bell, D. Golberg, A. Du, Chem. Mater. 35 (2023) 2884–2891. doi: 10.1021/acs.chemmater.2c03788

    14. [14]

      X. Fan, L. Xie, J. Liang, et al., Nano Res. 15 (2022) 3050–3055. doi: 10.1007/s12274-021-3951-5

    15. [15]

      L. Fang, S. Wang, S. Lu, et al., Chin. Chem. Lett. 35 (2024) 108864. doi: 10.1016/j.cclet.2023.108864

    16. [16]

      Z.Y. Wu, M. Karamad, X. Yong, et al., Nat. Commun. 12 (2021) 2870. doi: 10.1038/s41467-021-23115-x

    17. [17]

      X.F. Cheng, J.H. He, H.Q. Ji, et al., Adv. Mater. 34 (2022) 2205767. doi: 10.1002/adma.202205767

    18. [18]

      Y. Xue, Q. Yu, Q. Ma, et al., Environ. Sci. Technol. 56 (2022) 14797–14807. doi: 10.1021/acs.est.2c04456

    19. [19]

      J. Gao, B. Jiang, C. Ni, et al., Appl. Catal. B 254 (2019) 391–402. doi: 10.1016/j.apcatb.2019.05.016

    20. [20]

      Y. Li, J. Ma, Z. Wu, Z. Wang, Environ. Sci. Technol. 56 (2022) 8673–8681. doi: 10.1021/acs.est.1c05841

    21. [21]

      X. Li, P. Shen, X. Li, D. Ma, K. Chu, ACS Nano 17 (2023) 1081–1090. doi: 10.1021/acsnano.2c07911

    22. [22]

      N. Zhang, G. Zhang, P. Shen, et al., Adv. Funct. Mater. 33 (2023) 2211537. doi: 10.1002/adfm.202211537

    23. [23]

      K. Fan, W. Xie, J. Li, et al., Nat. Commun. 13 (2022) 7958. doi: 10.1038/s41467-022-35664-w

    24. [24]

      W. Duan, Y. Chen, H. Ma, et al., Environ. Sci. Technol. 57 (2023) 3893–3904. doi: 10.1021/acs.est.2c09147

    25. [25]

      R. Mao, N. Li, H. Lan, et al., Environ. Sci. Technol. 50 (2016) 3829–3837. doi: 10.1021/acs.est.5b05006

    26. [26]

      H. Lan, R. Mao, Y. Tong, et al., Environ. Sci. Technol. 50 (2016) 11872–11878. doi: 10.1021/acs.est.6b02822

    27. [27]

      W. Fu, Y. Du, J. Jing, C. Fu, M. Zhou, Appl. Catal. B 324 (2023) 122201. doi: 10.1016/j.apcatb.2022.122201

    28. [28]

      X. Zhao, Y. Jiang, M. Wang, et al., Adv. Energy Mater. 13 (2023) 2301409. doi: 10.1002/aenm.202301409

    29. [29]

      S. Xue, B. Garlyyev, S. Watzele, et al., ChemElectroChem 5 (2018) 2326–2329. doi: 10.1002/celc.201800690

    30. [30]

      W. Ma, S. Xie, T. Liu, et al., Nat. Catal. 3 (2020) 478–487. doi: 10.1038/s41929-020-0450-0

    31. [31]

      X. Lu, W. Tu, Y. Zhou, Z. Zou, Adv. Energy Mater. 13 (2023) 2300628. doi: 10.1002/aenm.202300628

    32. [32]

      W. Ma, S. Xie, X.G. Zhang, et al., Nat. Commun. 10 (2019) 892. doi: 10.1038/s41467-019-08805-x

    33. [33]

      H. Sun, Z. Yan, C. Tian, et al., Nat. Commun. 13 (2022) 3857. doi: 10.1038/s41467-022-31561-4

    34. [34]

      K.L. Zhou, Z. Wang, C.B. Han, et al., Nat. Commun. 12 (2021) 3783. doi: 10.1038/s41467-021-24079-8

    35. [35]

      Y. Hua, N. Song, Z. Wu, et al., Adv. Funct. Mater. 34 (2024) 2314461. doi: 10.1002/adfm.202314461

    36. [36]

      H. Jiang, G.F. Chen, O. Savateev, et al., Wang, Angew. Chem. Int. Ed. 62 (2023) 18717.

    37. [37]

      Y. Dong, J.B. Chen, J. Ying, et al., Chem. Mater. 34 (2022) 8271–8279. doi: 10.1021/acs.chemmater.2c01738

    38. [38]

      W. Shen, Y. Zheng, Y. Hu, et al., J. Am. Chem. Soc. 146 (2024) 5324–5332. doi: 10.1021/jacs.3c11861

    39. [39]

      Z. Yan, W.J. Kuang, Y. Lei, et al., Environ. Sci. Technol. 57 (2023) 20915–20928. doi: 10.1021/acs.est.3c06326

    40. [40]

      C. Zhang, D. He, J. Ma, T.D. Waite, Water Res. 145 (2018) 220–230. doi: 10.1016/j.watres.2018.08.025

    41. [41]

      F. Li, L. Sun, Y. Liu, et al., J. Hazard. Mater. 400 (2020) 123246. doi: 10.1016/j.jhazmat.2020.123246

    42. [42]

      F. Li, X. Peng, Y. Liu, et al., Chemosphere 229 (2019) 383–391. doi: 10.1016/j.chemosphere.2019.04.180

    43. [43]

      Z.Y. Wu, P. Zhu, D.A. Cullen, et al., Nat. Synth. 1 (2022) 658–667. doi: 10.1038/s44160-022-00129-x

  • Figure 1  SEM images of (a) Fe@Fe3O4/FF, (b) Fe3O4/CC, and (c) Fe/FF. (d, e) HRTEM images of Fe@Fe3O4/FF. (f) XRD patterns of three electrocatalysts.

    Figure 2  Effects of applied potential on (a) NH3 yield rate and (b) FE of NH3, and (c) the impact of nitrate concentration on NH3 yield rate across all the investigated cathodes. Nitrate concentration-dependent FE of various products during NO3RR for (d) Fe@Fe3O4/FF, (e) Fe3O4/CC, and (f) Fe/FF. Unless otherwise specified, reaction conditions are as follows: applied potential of −1.3 V (vs. SCE), 100 mmol/L Na2SO4, 100 mmol/L NaNO3, and 1 h of electrolysis.

    Figure 3  Current density variations in LSV for (a) Fe@Fe3O4/FF, (b) Fe/FF, and (c) Fe3O4/CC electrodes following the addition of 300 mmol/L TBA and 100 mg/L NaNO3. (d) Impact of TBA concentration on the nitrate removal efficiency across four electrodes, and (e) corresponding calculations of the contributions from different mechanisms involved in NO3RR. (f) Correlation between the NH3 yield rate and the proportion of DET in NO3RR, as determined from (e). (g) Measurement of the OH adsorptive peak in 1 mol/L NaOH solution. (h) EIS spectra of the three electrodes at −1.1 V (vs. SCE) in the presence of 100 mg/L NaNO3. (i) Influence of different alkali metal cations in the electrolyte on the NH3 yield rate for the three electrodes, evaluated in a solution of 100 mmol/L M2SO4 and 100 mmol/L MNO3 at −1.3 V (vs. SCE).

    Figure 4  (a) Representations of Fe/FF, Fe3O4/CC, and Fe@Fe3O4/FF models used for theoretical calculations. (b) Comparative analysis of NO3 adsorption free energy across the three models. Evaluation of the kinetic energy barrier for (c) HAT and (d) DET mechanisms on each of the models.

    Figure 5  Schematic representation of reaction mechanisms on (a) Fe@Fe3O4/FF, (b) Fe/FF, and (c) Fe3O4/CC.

  • 加载中
计量
  • PDF下载量:  0
  • 文章访问数:  157
  • HTML全文浏览量:  4
文章相关
  • 发布日期:  2025-06-15
  • 收稿日期:  2024-03-19
  • 接受日期:  2024-08-16
  • 修回日期:  2024-07-18
  • 网络出版日期:  2024-08-17
通讯作者: 陈斌, bchen63@163.com
  • 1. 

    沈阳化工大学材料科学与工程学院 沈阳 110142

  1. 本站搜索
  2. 百度学术搜索
  3. 万方数据库搜索
  4. CNKI搜索

/

返回文章