Abstract
The electrocatalytic nitrate/nitrite reduction reaction (eNOx−RR) to ammonia (NH3) is thermodynamically more favorable than the eye-catching nitrogen (N2) electroreduction. To date, the high eNOx−RR-to-NH3 activity is limited to strong alkaline electrolytes but cannot be achieved in economic and sustainable neutral/near-neutral electrolytes. Here, we construct a copper (Cu) catalyst encapsulated inside the hydrophilic hierarchical nitrogen-doped carbon nanocages (Cu@hNCNC). During eNOx−RR, the hNCNC shell hinders the diffusion of generated OH− ions outward, thus creating a self-enhanced local high pH environment around the inside Cu nanoparticles. Consequently, the Cu@hNCNC catalyst exhibits an excellent eNOx−RR-to-NH3 activity in the neutral electrolyte, equivalent to the Cu catalyst immobilized on the outer surface of hNCNC (Cu/hNCNC) in strong alkaline electrolyte, with much better stability for the former. The optimal NH3 yield rate reaches 4.0 moles per hour per gram with a high Faradaic efficiency of 99.7%. The strong-alkalinity-free advantage facilitates the practicability of Cu@hNCNC catalyst as demonstrated in a coupled plasma-driven N2 oxidization with eNOx−RR-to-NH3.
A confined catalyst gets rid of the dependence on strong alkaline electrolyte in electroreduction of nitrate/nitrite to ammonia.
INTRODUCTION
Ammonia (NH3) is one of the most important fundamental feedstocks in the world, which has wide applications in chemical and fertilizer industries and is also a promising energy carrier of the hydrogen economy (1, 2). The Haber-Bosch synthesis of ammonia is an energy-, carbon-, and capital-intensive process, which accounts for 1 to 2% of global energy consumption and is also poorly compatible with distributed agriculture (3, 4). Electrocatalytic nitrogen reduction reaction (eNRR) to NH3 powered by renewable electricity is a potential alternative to the Haber-Bosch process (5–7). However, the strong N ≡ N bonding and the poor solubility of N2 in electrolytes are unfavorable in thermodynamics and kinetics for NH3 production (8–10). In contrast, the NH3 production via electrocatalytic nitrate/nitrite reduction reaction (eNOx−RR) is more favorable in thermodynamics and kinetics, demonstrating the superior NH3 yield rate and Faradaic efficiency (FE) to eNRR in aqueous electrolytes (11–13). In addition, the eNOx−RR-to-NH3 provides a win-win opportunity for distributed fertilizer synthesis and nitrogen pollution control in water with renewable electricity (fig. S1) (14–17).
The eNOx−RR is accompanied by the competitive hydrogen evolution reaction (HER) due to their overlapping reaction potentials, and they both are pH-sensitive (18, 19). Usually, the strong alkaline condition could accelerate the reaction kinetics of the eNOx−RR and suppress the HER (20–23). Hence, most of the excellent eNOx−RR-to-NH3 performances to date have been achieved in strong alkaline electrolytes (24–26). For example, for copper catalysts, theoretical and experimental studies verified the suitable adsorption of intermediates (e.g., *NOH), the favorable kinetics, and the suppressed HER in alkaline conditions, in comparison with the cases in acidic or neutral conditions, leading to better eNOx−RR-to-NH3 selectivity and activity for the former (22, 23). However, the use of strong alkaline electrolytes is accompanied by many drawbacks unfavorable for practical applications such as the side reaction with ubiquitous CO2, high corrosion to equipment, and high cost. Therefore, it is desirable to develop highly efficient eNOx−RR-to-NH3 catalysts suitable for neutral/near-neutral electrolytes.
As is known, the catalytic performance highly depends on the local chemical environment around the catalyst (27–32). eNOx−RR-to-NH3 is accompanied by the OH− formation via the reaction of NOx− + (x + 3)H2O + (2x + 2)e− → NH3 + (2x + 3)OH− (12). In principle, this process could increase the local pH around the catalytic sites and inhibit the competing HER, thus enhancing the eNOx−RR-to-NH3 selectivity and activity. However, for conventional catalysts, usually, the in situ formed OH− will rapidly diffuse into the bulk electrolyte driven by the concentration gradient, leading to the slow enhancement of local pH and catalytic performance thereof. Increasing the surface roughness of catalysts, e.g., by nanosizing the particles or constructing surface microstructures, could decelerate the OH− diffusion and enhance the local pH to some extent (19, 33), which is still not enough due to the open structures, especially under stirring or flowing conditions. Hence, how to effectively confine the OH− around the catalytic sites to increase the local pH is a key yet challenging topic.
In recent years, we have developed hierarchical carbon nanocages (hCNC) featuring three-dimensional (3D) porous structure, high conductivity, large specific surface area, and convenient doping, which can facilitate the mass/charge synergetic transfer, thus becoming an advanced platform for energy conversion and storage (34–36). This unique platform, especially the large cavities of carbon nanocages and abundant microchannels across the shells, provides an ideal “semi-closed” environment to encapsulate the eNOx−RR-to-NH3 active species (e.g., Cu, the most promising catalyst with low-cost and inferior HER competition) inside and prevents the in situ generated OH− from escaping out of the nanocages, as illustrated in Fig. 1. With this strategy, in this study, we encapsulated Cu nanoparticles inside the hydrophilic hierarchical nitrogen-doped carbon nanocages (hNCNC), and the so-constructed confined catalyst of Cu@hNCNC demonstrates a self-enhanced localized high pH environment around the inside Cu nanoparticles. Consequently, the Cu@hNCNC catalyst exhibits an excellent eNOx−RR-to-NH3 activity even in the neutral electrolyte, comparable to the Cu catalyst immobilized on the outer surface of hNCNC (Cu/hNCNC) with an “open” environment in the strong alkaline electrolyte, with much better stability for the former. In the neutral electrolyte with 1 M NOx−, Cu@hNCNC catalyst exhibits an excellent NH3 yield of up to 4.0 mol hour−1 g−1 with a high FE of 99.7%. The strong-alkalinity-free advantage suggests the potential application, and the practicability of the Cu@hNCNC catalyst is demonstrated by the coupled plasma-driven N2 oxidization with eNOx−RR-to-NH3 for NH3 synthesis.
Fig. 1. Schematic illustration of the self-enhanced localized alkalinity within the cavities of hNCNC for Cu@hNCNC catalyst during eNOx−RR.
(A) Cu nanoparticles are encapsulated inside the hierarchical nitrogen-doped carbon nanocages (hNCNC). (B) The in situ generated OH− ions accumulate within the cavities of hNCNC due to the hindered escaping, leading to the self-enhanced local high OH− concentration (COH−) inside the nanocages. The COH− outside the hNCNC increases slowly due to the hindered OH− escaping from the inside and the bulk of neutral electrolyte outside.
RESULTS
Structural characterization and pH-dependent eNO3−RR-to-NH3 performances
The hydrophilic hNCNC with an N content of 8.6 atomic % is used to ensure the convenient entrance of the aqueous electrolyte into the cavities of nanocages (fig. S2). Cu@hNCNC and Cu/hNCNC catalysts, with Cu nanoparticles inside and outside the hNCNC respectively, were constructed to evaluate the effect of the catalytic microenvironment on the eNO3−RR-to-NH3 performance, as shown in Fig. 2. The Cu nanoparticles were encapsulated inside the hNCNC via vacuum filling or immobilized on the outer surface of hNCNC by microwave-assisted ethylene glycol (EG) reduction (see the “Materials synthesis” section in Materials and Methods). Scanning electron microscopy (SEM) and transmission electron microscopy (TEM) characterizations show that, for Cu@hNCNC, most cavities of hNCNC are filled with Cu nanoparticles of ~10.7 nm in average size without observable Cu nanoparticles on the outer surface of hNCNC (Fig. 2, A and B, and fig. S3). The lattice fringes with the interplanar distance of 2.1 Å can be assigned to the (111) planes of metallic Cu (JCPDS no. 04-0836) (Fig. 2B). In addition to the Cu nanoparticles, there is still enough space inside the nanocages for catalytic reactions. For Cu/hNCNC, numerous high-dispersed Cu nanoparticles of ~9.8 nm are immobilized on the outer surface of hNCNC (Fig. 2, C and D, and fig. S3). The mass loadings of Cu nanoparticles in Cu@hNCNC and Cu/hNCNC are nearly identical, about 20 wt %, as revealed by the thermogravimetric analysis (fig. S4). The two catalysts show similar resistances (fig. S5).
Fig. 2. Structural characterizations of Cu@hNCNC and Cu/hNCNC.
(A and B) Scanning electron microscopy (SEM) and transmission electron microscopy (TEM) images of Cu@hNCNC. Inset in (A) is the low-magnification view of Cu@hNCNC, and inset in (B) is the typical high-resolution TEM image of Cu nanoparticles. (C and D) SEM and TEM images of Cu/hNCNC. (E) Operando Raman spectra. (F and G) Operando x-ray absorption spectroscopy analysis with the first derivative of x-ray absorption near edge structure (F) and k3-weighted Fourier transformed (FT) extended x-ray absorption fine structure spectra of Cu K-edge (G). Note: The data of Cu, Cu2O, and CuO in (E) and (F) are presented for comparison. a.u., arbitrary units.
The Cu@hNCNC and Cu/hNCNC maintain the hierarchical 3D porous structure of the hNCNC support (Fig. 2, A to D, and fig. S3, A to C). All the hNCNC, Cu@hNCNC, and Cu/hNCNC display the IV-type N2 adsorption-desorption isotherms with obvious hysteresis, showing the coexisting micro-meso-macropores (fig. S6). For Cu@hNCNC, the mesoporous distribution shifts to the small-sized side in comparison with the case for Cu/hNCNC and hNCNC as expected since partial inside cavities of hNCNC are occupied by the Cu nanoparticles. Such an integrated architecture of hierarchical carbon scaffold and interconnected multiscale pores is beneficial to the synergic charge/mass transfer during the electrocatalysis, favorable for the reaction kinetics (34–36). The micropores across the nanocage wall provide the diffusion channels of reactants into the nanocages for electrocatalytic reaction, meanwhile hinders the diffusion of products from the inside to outside of the nanocages, i.e., the “blocking effect” (figs. S6 and S7 and table S1) (37). As a result, the OH− concentrations (COH−) inside the nanocages are much higher than those outside the nanocages, thus significantly affecting the catalytic performance (Fig. 1).
Operando Raman and operando x-ray absorption spectroscopy (XAS) were conducted to characterize the electronic structure and chemical state of Cu species in Cu@hNCNC and Cu/hNCNC in analogy to the pretreatment condition before electrochemical tests of eNOx−RR (figs. S8 and S9). In the Raman spectra of the two catalysts, the initially sharp Raman bands (ascribed to Cu2O and/or CuO) disappear after the pretreatment at an applied potential of −1.5 V (Fig. 2E). For XAS analysis, the energy of the maximum peak in the first derivative of x-ray absorption near edge structure spectrum is proportional to the valence state of the element. At open circuit potential (OCP), the valence states of Cu in the two catalysts are close to Cu2+, which are reduced to Cu0 at −1.5 V (Fig. 2F). The extended x-ray absorption fine structure (EXAFS) spectra show that both catalysts feature the Cu-O bonds at OCP, while only the Cu-Cu bond at −1.5 V (Fig. 2G). The fitting results of the EXAFS spectra exhibit similar coordination numbers and bond lengths of Cu-Cu for Cu@hNCNC and Cu/hNCNC after pretreatment, indicating the highly similar structures (table S2). The x-ray photoelectron spectroscopy (XPS) analysis is in high agreement with these operando results (fig. S10). The preceding characterization results indicate that the Cu species is reduced to metallic Cu after pretreatment and the two catalysts have identical electronic structures and chemical states.
pH-dependent eNO3−RR-to-NH3 performances
The Cu@hNCNC and Cu/hNCNC catalysts have similar active surface areas (fig. S11). The electrochemical performances of the two catalysts were evaluated by the glassy carbon electrode (GCE) with a diameter of 3 mm (working area of 0.07 cm2) (fig. S12). All potentials were versus the Ag/AgCl electrode directly, unless explicitly stated. The linear sweep voltammetry (LSV) curves obtained in the electrolytes with 50 mM NO3− exhibit much larger current densities for Cu@hNCNC and Cu/hNCNC than those for hNCNC in all pH, with ~0.5-V positive shift of the onset potential, indicating the catalytic function of Cu species to eNO3−RR (fig. S13). The eNO3−RR performance was first evaluated in potentiostatic mode. The dissolved products including NH3 and NO2− were quantitatively measured using the colorimetric method (figs. S14 and S15). The eNO3−RR-to-NH3 performance of Cu@hNCNC and Cu/hNCNC are directly compared taking into account almost the same content and small contribution of the hNCNC support (figs. S4 and S16). In all the electrolytes with pH 7 to 13, the FEs of NH3 for the two catalysts display “volcano-like” evolutions with increasing the applied potential, reaching the maximum at −1.5 V (fig. S17). A similar performance is obtained on the larger GCE with a working area of 1 cm2 (fig. S18). The decays of FEs at −1.6 V are attributed to the increased HER and the decreased catalyst-electrolyte interface thereof due to the coverage of H2 bubbles (fig. S19). Hence, the optimal potential of −1.5 V was taken for evaluating the eNO3−RR-to-NH3 performance.
For the Cu/hNCNC, the FE and yield rate of NH3 increase with increasing pH from 7 to 12, and then markedly decrease at pH 13, with the maximum FE (92%) and NH3 yield rate (3.4 mol hour−1 g−1) at pH 12 (Fig. 3, A and B). In contrast, the Cu@hNCNC reaches the FE of 90% and the NH3 yield rate of ~3 mol hour−1 g−1 even at pH 7, close to the optimal performance of the Cu/hNCNC at pH 12 (Fig. 3, A and B). Moreover, the Cu@hNCNC presents a high NH3 yield rate of 3.5 mol hour−1 g−1 in the pH range of 11 to 12, with the top-level NH3 partial current densities (jNH3) (Fig. 3B and table S3). For the cases of pH ≤ 11, Cu@hNCNC also shows better suppression of HER than Cu/hNCNC, as learned by the lower H2 FE (FEH2) and H2 partial current density (jH2) for the former (Fig. 3C and fig. S20). The FEH2 of the two catalysts almost drop to 1% at pH 12 and 13, indicating the suppressed HER at high pH (20–23, 38). The decrease of FENH3 at pH 13 could be ascribed to the competitive adsorption of OH− on the Cu surface, which results in the decrease of adsorption capacities for NO3− and thereby the sluggish eNO3−RR kinetics (Fig. 3 and figs. S21 and S22). The kinetics of eNO3−RR is deduced on the basis of the time-dependent concentration change of NO3− in the electrolyte, which obeys pseudo–first-order kinetics (Fig. 3D and fig. S23). The apparent rate constant (k) for Cu@hNCNC at pH 7 is close to that for Cu/hNCNC at pH 12, indicating a comparable reaction rate. Hence, the Cu@hNCNC catalyst can efficiently express its excellent activity toward eNO3−RR-to-NH3 in neutral/near-neutral electrolytes, which avoids the unfavorable strong alkaline electrolytes. It is noteworthy that the Cu@hCNC catalyst by encapsulating Cu inside hydrophobic hCNC shows a much lower FE and NH3 yield rate than Cu@hNCNC catalysts, indicating the crucial role of hydrophilicity of hNCNC from N-doping to ensure the full access of the electrolyte to the encapsulated catalyst (figs. S2 and S24).
Fig. 3. eNO3−RR-to-NH3 performances of Cu@hNCNC and Cu/hNCNC.
(A to C) Evolutions of NH3 FE (A), NH3 yield rate and NH3 partial current density (B), and H2 FE (C) with pH of 7 to 13 at −1.5 V (versus Ag/AgCl). (D) Linearized pseudo–first-order kinetic profiles. Note: The error bars in (A) to (C) represent the data range obtained in the repeated tests. The mixture of 0.5 M Na2SO4 and 50 mM NaNO3 was used as the electrolyte.
The catalytic stability at pH 7 and 12 was evaluated at −1.5 V for eight consecutive cycles. The FE and yield rate of NH3 show negligible changes for Cu@hNCNC while presenting obvious decay after three cycles for Cu/hNCNC. The TEM characterizations show that, after cycling eight times, the Cu nanoparticles are still encapsulated inside hNCNC for Cu@hNCNC while agglomerated severely for Cu/hNCNC (fig. S25). As expected, after testing in the electrolyte for 1 hour, the Cu concentration in the electrolyte for the case of Cu@hNCNC is obviously lower than that for Cu/hNCNC (fig. S26). The superb stability of Cu@hNCNC is attributed to the confinement of hNCNC, which protects the active-Cu species from dissolution and agglomeration (fig. S25).
Evaluation and theoretical simulation of local pH
To get insight into the excellent eNO3−RR-to-NH3 performance of Cu@hNCNC in neutral/near-neutral electrolytes, we monitored the pH evolution at the electrode surface during the eNO3−RR to deduce the local pH situation inside hNCNC, in combination with finite element simulations, as shown in Fig. 4. A pH meter was placed against the working electrode surface to monitor the pH change, with or without stirring (Fig. 4, A and B). In the potentiostatic test, without stirring, the COH− at the surface of the Cu@hNCNC electrode is lower than that of the Cu/hNCNC electrode in the initial stage (~15 min), and reversely later (Fig. 4C). The lower COH− for the former in the initial stage results from the retarded OH− ions outward diffusion due to the confinement of hNCNC (the red arrows for Cu@hNCNC in Fig. 4B), and the later reversal results from the more generated OH− ions for the former due to the higher activity of Cu resulting from the local high pH within the cavities of hNCNC. With stirring, the OH− diffusion is accelerated (the red and green arrows in Fig. 4B), leading to the unobvious difference of the COH− for Cu@hNCNC and Cu/hNCNC electrodes in the initial stage (~8 min). The subsequent increasing difference of the COH− results from the more generated OH− ions for the former due to the higher local pH within the cavities of hNCNC (Fig. 4C and fig. S27). The obvious lower COH− with stirring than without stirring is attributed to the more rapid diffusion and dilution of OH− ions into the bulk electrolyte. The pH evolution in a periodic intermittent eNO3−RR with six cycles further confirms these analyses and results (fig. S28).
Fig. 4. Evaluation and theoretical simulation of the local pH for Cu@hNCNC and Cu/hNCNC catalysts.
(A) Configuratioin neutral/near-neutraln for measuring the pH value at the electrode surface (the contact area in the highlighted orange box). (B) Schematic illustration for OH− generation, diffusion, and pH detection. The red, green, and black arrows represent the OH− diffusion out of the cavities of hNCNC, in-between the hNCNC sheets toward the electrode surface, and to the bulk electrolyte, respectively. The black dashed line and the blue arc represent the electrode surface and the pH meter, respectively. (C) The detection of COH− at electrode surface during eNO3−RR at −1.5 V (potentiostatic mode) without and with stirring. (D) eNO3−RR-to-NH3 performances in different intervals without stirring at 80 mA cm−2 (galvanostatic mode). (E) COH− maps after eNO3−RR for 30 min simulated with finite element method (taken from fig. S35 for the electrolyte with an initial pH of 7). Insets at the bottom are the enlarged rectangular areas demonstrating the spatial distribution of COH−. The black dots denote different distances from the Cu surface (1 to 9 nm). (F) The simulated spatial distributions of COH−. Note: The eNO3−RR was performed in a 30-ml H-type electrolytic cell. The electrolyte is composed of 0.5 M Na2SO4 and 50 mM NaNO3 with an initial pH of 7. The catalyst layer has an area of 0.5 cm × 0.4 cm, with a mass loading of 0.12 mg cm−2. The stirring rate is 300 rpm.
The preceding experimental results clearly indicate the OH− accumulation within the cavities of hNCNC for Cu@hNCNC due to the hindered outward diffusion of OH− by the nanocage shell with abundant micropores. Taking the galvanostatic test to control the current density in eNO3−RR (fig. S29), using the rotating ring-disk electrode (RRDE) to monitor the pH evolution on the electrode surface (fig. S30), or using phosphate-buffered (PB) solution to exclude the interference of pH fluctuation (figs. S31 to S32), the experimental results also indicate the delayed pH increase on the electrode surface of Cu@hNCNC, which further confirm the “blocking effect” of hNCNC and the local high pH environment around the encapsulated Cu during eNO3−RR thereof. Such a nanocage geometry with self-enhanced local alkalinity makes the Cu@hNCNC an ideal catalyst for eNO3−RR in neutral/near-neutral electrolytes, which is difficult to realize for the supported catalysts, e.g., the Cu/hNCNC, with an open environment (figs. S27 to S30 and S33). Actually, for eNO3−RR-to-NH3 with an initial pH of 7, the Cu@hNCNC demonstrates superior performances to the Cu/hNCNC in different reaction intervals (Fig. 4D and fig. S34). For Cu@hNCNC, the local high pH leads to high performance in the whole reaction period with a slightly increasing FE and yield rate of NH3 in the first two intervals. For Cu/hNCNC, the initial low pH leads to low performance in the early period. The FE and yield rate of NH3 gradually increase with increasing pH during eNO3−RR-to-NH3 and are close to those of Cu@hNCNC in the later period (approximately the interval of 20 to 30 min) due to the similar pH level. These results demonstrate the strong connection of the catalytic activity with the local pH environment.
The blocking effect of hNCNC is further elucidated by finite element simulations with two 2D models corresponding to Cu@hNCNC and Cu/hNCNC (fig. S35). After eNO3−RR of 30 min, the COH− inside hNCNC for Cu@hNCNC is much higher than that outside with a sharp gradient of COH− across the shell, while the COH− for Cu/hNCNC shows a gradual decrease with increasing the distance from Cu surface (Fig. 4E). Specifically, the COH− inside hNCNC is approximately five times higher than that outside for the former (Fig. 4F). This is the effect of a single nanocage in bulk electrolyte, and the COH− enhancement should be much strengthened in practical situation with hierarchical porous structure.
On the basis of the above results, the picture for Cu@hNCNC in eNO3−RR could be well figured out. With the initial pH <12, when eNO3−RR is started, OH− is generated. The self-enhanced localized pH inside the hNCNC in turn increases the eNO3−RR activity. Such “positive feedback” leads to the quick increase of the local pH to ~12. The further increase of the local pH (over 12) leads to the drop of the eNO3−RR activity (Fig. 3), i.e., the OH− generation slows down quickly and cannot offset the amount of outward diffusion. Such “negative feedback” results in a stable local high pH of around 12 eventually. This picture is highly consistent with the results of pH measurements and theoretical simulations. For eNO2−RR-to-NH3, the Cu@hNCNC also exhibited a high FE (94.0%) and yield rate of NH3 (2.9 mol hour−1 g−1) at −1.5 V in neutral electrolyte, higher than the corresponding data for Cu/hNCNC (fig. S36).
eNOx−RR-to-NH3 performance of Cu@hNCNC catalyst in neutral electrolyte
The preceding experimental and theoretical results demonstrate the excellent performance of Cu@hNCNC catalyst for both eNO3−RR- and eNO2−RR-to-NH3 in neutral electrolytes due to the self-enhanced localized high pH around Cu nanoparticles inside hNCNC. Hence, we have further evaluated the eNOx−RR-to-NH3 performance of Cu@hNCNC in neutral electrolyte with the fixed NO3−/NO2− concentration ratio of 1:3. With increasing the concentration of NOx−, the current density increases accordingly in the range of 0.1 to 1000 mM, followed by a slight fluctuation in 1000 to 2000 mM (Fig. 5A and fig. S37A). Similarly, both the FE and yield rate of NH3 monotonously increase in the range of 0.1 to 1000 mM, reaching the maximum of 99.7% and 4.0 mol hour−1 g−1, respectively, and then remain at the high level of >99% and >3.8 mol hour−1 g−1 in 1000 to 2000 mM. The increase of the NH3 yield rate follows the positive- and zero-order reactions with the low (≤1000 mM) and high (≥1000 mM) concentrations of NOx−, respectively (Fig. 5B and fig. S37B) (39). Importantly, in neutral/near-neutral electrolytes, Cu@hNCNC exhibits high catalytic activity in a wide range of NOx− concentration, which is suitable for various application scenarios, e.g., the eNO3−RR-to-NH3 from the wastewater with low NOx− concentrations, and the industrial-scale NH3 synthesis by the coupled plasma-driven N2 oxidization with eNOx−RR-to-NH3 with tunable NOx− concentrations. The durability of Cu@hNCNC catalyst in eNOx−RR was evaluated by 50 electrocatalysis cycles in 1000 mM NOx−, which remains the high FE and yield rate of NH3 with a small fluctuation, indicating the excellent electrocatalytic stability of Cu@hNCNC. In contrast, Cu/hNCNC in strong alkaline electrolytes only presents a comparable yield rate of NH3 in the initial few cycles, with gradual decay in the following cycles (Fig. 5C and fig. S38). The excellent performance of Cu@hNCNC reaches the top level in neutral electrolytes, which also outperforms those of most non-noble metal electrocatalysts in strong alkaline electrolytes (Fig. 5D and table S3) (15, 20, 21, 24–26, 39–58). Hence, the Cu@hNCNC catalyst can efficiently express its high activity toward eNOx−RR-to-NH3 in neutral/near-neutral electrolytes with high stability, which avoids the use of unfavorable strong alkaline electrolytes.
Fig. 5. The eNOx−RR-to-NH3 performance of Cu@hNCNC with different concentrations of NOx− at −1.5 V in the neutral electrolyte (pH 7).
(A) i-t curves. (B) FE and yield rate of NH3. (C) Electrocatalytic stability of Cu@hNCNC (red) in 1000 mM NOx−. The corresponding data of Cu/hNCNC (blue) in strong alkaline electrolytes (pH 12) are also presented for comparison. (D) Comparison of the eNOx−RR-to-NH3 performance of Cu@hNCNC with those of typical electrocatalysts in literature. Note: The NO3−/NO2− ratio in NOx− is fixed at 1:3.
Practicability of Cu@hNCNC catalyst in eNOx−RR-to-NH3
The preceding results exhibit the excellent eNOx−RR-to-NH3 performance of Cu@hNCNC in neutral electrolytes due to the self-enhanced high local pH around Cu nanoparticles within the cavities of hNCNC. This result indicates the great potential for sustainable NH3 synthesis as demonstrated by coupling plasma-driven N2 oxidization with eNOx−RR-to-NH3 in Fig. 6 (fig. S39). The NOx gas was generated by the dielectric barrier discharge (DBD) in the air and converted into NO3−/NO2− in the PB-containing absorber (pH 7), which was then exchanged into the electrolyzer for electroreduction to NH3 (Fig. 6A). The concentration of NOx− increases linearly with the plasma operation time and reaches 42.2 mM after 30 min of plasma treatment (Fig. 6B). The subsequent eNOx−RR-to-NH3 exhibits the high FE (90.5%) and yield rate of NH3 (2.44 mol hour−1 g−1) for Cu@hNCNC, much better than the corresponding 74.3% and 1.80 mol hour−1 g−1 for Cu/hNCNC, with larger current density for the former (Fig. 6, C and D). This performance of Cu@hNCNC in the coupled system in the neutral condition is comparable to the corresponding ones in literature in the strong alkaline electrolytes, suggesting the potential applications (table S4).
Fig. 6. Demonstration of sustainable NH3 synthesis by coupling plasma-driven N2 oxidization with eNOx−RR-to-NH3.
(A) Schematic illustration of the coupled reaction system. (B) Plots of NOx− concentration versus plasma operation time. (C) i-t curves. (D) FE and yield rate of NH3 after 30-min reaction. Note: The air flow rate was 2000 sccm. PB (0.2 M; pH 7) with 0.3 M Na2SO4 was used as the absorption solution. The applied potential is −1.6 V (fig. S31).
The self-enhanced local high pH inside hNCNC was also observed in the CO2 electroreduction reaction for Cu@hNCNC relative to Cu/hNCNC and in the oxygen reduction reaction for Pt@hNCNC relative to Pt/hNCNC (fig. S40). These results indicate the general validity of regulating the local pH environment via carbon nanocage confinement, which provides a new approach to enhance catalytic performances via microenvironment engineering.
DISCUSSION
In summary, we have designed and constructed the confined Cu@hNCNC catalyst by encapsulating Cu nanoparticles inside the hydrophilic hNCNC. Such Cu@hNCNC catalyst can prevent the in situ generated OH− ions during eNOx−RR-to-NH3 from escaping out of nanocages and thus lead to a self-enhanced localized high pH environment around the Cu nanoparticles, which is confirmed by the experimental results and finite element simulations. Consequently, the Cu@hNCNC catalyst with a semi-closed environment exhibits an outstanding eNOx−RR-to-NH3 activity even in the neutral electrolyte, equivalent to the Cu/hNCNC catalyst with an open environment in the strong alkaline electrolyte. An excellent NH3 yield rate of 4.0 mol hour−1 g−1 with FE of 99.7% is achieved in the neutral electrolytes, which outperforms those of most non-noble metal electrocatalysts in strong alkaline electrolytes. The confinement also effectively inhibits the loss and agglomeration of Cu species during eNOx−RR, leading to the excellent stability of Cu@hNCNC catalyst much superior to Cu/hNCNC. The practicability of Cu@hNCNC catalyst is demonstrated by a coupled plasma-driven N2 oxidization with the eNOx−RR-to-NH3 process. The integrated advantages of high activity, high FE, and high stability with alkali-free electrolytes suggest an economic and sustainable alternative for ammonia synthesis. Taking into account the unique architecture and properties of the hierarchical carbon nanocages, the convenient strategy in this study, i.e., encapsulating catalytic active species inside the nanocages for achieving a local high pH environment, could be a general way for microenvironment engineering to explore advanced catalysts for some key reactions.
MATERIALS AND METHODS
Materials synthesis
The hCNC and hNCNC were prepared by the in situ MgO template method at 800°C with benzene and pyridine precursor, respectively (59, 60).
The Cu@hNCNC was prepared by a vacuum-filling method (61). In a typical procedure, hNCNC (50 mg) was evacuated to ~10 Pa in a sealed flask, and then, 30 ml of 0.4 M Cu(NO3)2 solution was quickly injected into the flask, followed by magnetic stirring for 2 hours. After that, the sample was filtrated, freeze-dried, and then washed three times with deionized water to remove the Cu(NO3)2 outside the hNCNC. After the second vacuum-drying process, the sample was reduced to 50 sccm of H2/Ar (5%/95%) at 220°C for 1 hour, leading to the formation of Cu@hNCNC.
The Cu/hNCNC was synthesized by a microwave-assisted EG reduction method (62). Typically, 20 mg of hNCNC was dispersed ultrasonically into 50 ml of EG, and then 1 ml of 0.4 M Cu(NO3)2/EG and 4 ml of 0.2 M NaOH/EG solution was successively added to the suspension and this mixture was stirred for 4 hours. The suspension was irradiated in a microwave oven at 700 W for 40 s followed by aging for 5 min. After filtering, washing with ethanol and water, and drying at 70°C, the Cu/hNCNC was obtained by the reduction in 50 sccm of H2/Ar (5%/95%) at 220°C for 1 hour. Similarly, the Cu@hCNC and Cu/hCNC were prepared by replacing hNCNC with hCNC for comparison.
Characterization
The morphology and microstructure of the samples were examined by SEM (Hitachi, S-4800, operating at 10 kV) and TEM (JEM-2100, operating at 200 kV). The crystalline structure was determined by an x-ray diffractometer (Bruker, D8 Advance A25, Cu Ka1, λ = 0.154056 nm). The surface composition and element valence were analyzed by XPS (ULVAC-PHI INC, PHI 5000 VersaProbe, Al Kα). N2 adsorption-desorption isotherms were obtained on Thermo Fisher Scientific Surfer Gas Adsorption Porosimeter at 77 K. hNCNC was degassed at 300°C, while Cu@hNCNC and Cu/hNCNC were degassed at 180°C. The pore size distribution was calculated from the corresponding adsorption branch of N2 isotherm by the Barrett-Joyner-Halenda method for mesopore and the Horvath-Kawazoe method for micropore. The thermogravimetry (Netzsch, STA449F3) analysis was carried out in 20 vol % O2-containing Ar of 20 sccm at a rate of 10.0°C min−1. The concentration of Cu dissolved in the electrolyte was determined by inductively coupled plasma optical emission spectrometry (SHIMADZU, ICPE-9810) and the linear correlation of concentration intensity was calibrated using a series of standard copper sulfate solutions with ultrapure water as the solvent. pH was measured by a microelectrode (STMICRO5) on a pH meter (OHAUS, STARTER 3100) and an RRDE (Pine Instrument Co.).
Electrochemical measurements
Five milligrams of catalysts (Cu@hNCNC, Cu/hNCNC, or hNCNC) was dispersed ultrasonically into a mixed solution including 0.8 ml of ultrapure water (18.2 megohms·cm), 0.2 ml of ethanol, and 150 μl of 5% Nafion solution, followed by stirring overnight. Two microliters of the catalyst ink was dropped onto the GCE (3-mm diameter) for natural drying at room temperature. The catalyst mass loading on the working electrode was 0.12 mg cm−2. Ag/AgCl electrode (3 M KCl) and platinum sheet (1 cm × 1 cm) were used as the reference and counter electrode, respectively. The 0.5 M Na2SO4 solution was used as the electrolyte which contained 50 mM NaNO3 to provide NO3−, with different pH values (7 to 13) adjusted by NaOH or H2SO4 solution.
The electrochemical measurements were conducted on a VMP3 electrochemical workstation (BioLogic). The eNO3−RR experiments were performed in the potentiostatic mode using a three-electrode system in a 10-ml H-type electrolytic cell separated by a pretreated proton exchange membrane (Nafion 117). All potentials were versus the Ag/AgCl electrode directly, unless explicitly stated. Under stable pH conditions, the Ag/AgCl potential can be converted to a reversible hydrogen electrode (RHE) scale as follows: ERHE = EAg/AgCl + 0.0592 × pH + 0.209. Ar gas (20 sccm) flowed through the cathodic compartment for 15 min before each electrochemical test and continued in the whole eNO3−RR experiment. Before the first electrochemical measurement, the working electrodes with Cu-based catalysts were pretreated with two potentiostatic cycles of eNO3−RR at −1.5 V for 0.5 hours per cycle. The LSV tests were performed at a scan rate of 20 mV s−1. The chronoamperometric tests were conducted at different potentials for 0.5 hours with a rotation rate of 300 rpm, and the cyclic stability tests were conducted at −1.5 V for 0.5 hours per cycle. Electrochemical impedance spectroscopy measurements were conducted at OCP, and the frequency scan range was from 0.1 Hz to 100 kHz. iR compensation was not conducted in all electrochemical results. The absorption solution of 0.05 M H2SO4 was used to collect the evaporated NH3 in outflowing gas from the cathodic compartment. The NH3 amounts in the cathodic and anodic compartment electrolytes and in the H2SO4 solution were detected and quantified, respectively. The sum was used to calculate the NH3 yield rate and FE.
Practically, NO3− and NO2− coexist in most cases, e.g., in the wastewater, or the NOx− produced by plasma-driven N2 oxidization. Hence, the NO3−/NO2− mixture was used as the raw material to evaluate the potential applications of the catalysts. The NO3−/NO2− ratio of 1:3 was selected by referring to that in the NOx− produced by plasma-driven N2 oxidization in our experiment (Fig. 6B).
Operando Raman test
A single-cell system was used in the operando Raman test. Catalyst-coated GCE was used as the working electrode. Ag/AgCl electrode (3 M KCl) and platinum wire were used as the reference and counter electrode, respectively. The solution of 0.5 M Na2SO4 and 50 mM NaNO3 (pH 7) was used as the electrolyte. The electrolyte was first bubbled by argon for 30 min. Then, it is injected into the electrolytic cell with a circulating flow rate of 10 sccm. eNO3−RR was performed at −1.5 V, and the corresponding Raman spectra were measured using a Horiba JY Evolution spectrometer with a 633-nm laser source.
Operando XAS test
A flow cell system was used in the operando XAS test. The catalyst was coated on the A side of a carbon paper, with the A side facing the chamber and the B side facing outside. The carbon paper was sealed on the outer wall of the flow cell, allowing the x-ray to irradiate directly onto the working electrode. Ag/AgCl electrode (3 M KCl) and nickel foam were used as the reference and counter electrode, respectively. Nafion 117 proton exchange membrane was used as the separator. The solution with 0.5 M Na2SO4 and 50 mM NaNO3 (pH 7) was used as the electrolyte. The flow rate of electrolytes in both the cathode and anode chambers was 10 sccm. eNO3−RR was performed at an OCP of −1.5 V, and the corresponding XAS spectrum at the Cu K-edge was obtained on the BL11B beamline at the Shanghai Synchrotron Radiation Facility. The XAS data were analyzed by using ATHENA.
Determination of products
The possible gas products for eNO3−RR, including H2, N2, and O2, were determined qualitatively by online mass spectrometry (AMETEK DYCOR) and H2 was quantified by online gas chromatography with a thermal conductivity detector (LU GEN ANALYSIS GC-9860). The liquid ones were quantified by the colorimetry method. The ultraviolet-visible (UV-vis) absorbance spectra were measured on a high-speed spectrometer (PG2000, ideaoptics).
Quantification of NH3
The NH3 produced by eNO3−RR was quantified by the indophenol blue method. First, a specific amount of electrolyte in the two compartments (cathodic and anodic) and absorption solution was respectively diluted to 2 ml, to ensure the NH3 concertation in the detection range. Next, 2 ml of 1 M NaOH solution containing salicylic acid (5 wt %) and sodium citrate (5 wt %) was added and mixed thoroughly, and then 1 ml of 0.05 M sodium hypochlorite and 0.2 ml of sodium nitroprusside (1 wt %) were added. After sitting for 1 hour, the absorption intensity at a wavelength of 653 nm was recorded. The concentration-absorbance curve was calibrated using a series of standard ammonium chloride solutions with the electrolyte (0.5 M Na2SO4 and 50 mM NaNO3) as the solvent.
Quantification of NO2−
A mixture of p-aminobenzenesulfonamide (4 g), N-(1-naphthyl)ethylenediamine dihydrochloride (0.2 g), ultrapure water (50 ml), and phosphoric acid (10 ml, ρ = 1.70 g ml−1) was used as a color reagent. A specific amount of electrolyte was taken out from the cathodic compartment and added into 5 ml of ultrapure water to dilute the detection range. Next, 0.1 ml of color reagent was added to the aforementioned solution and mixed uniformly. After sitting for 20 min, the absorption intensity at a wavelength of 540 nm was recorded. The concentration-absorbance curve was calibrated using a series of standard sodium nitrite solutions with ultrapure water as the solvent.
Quantification of NO3−
A certain amount of electrolyte in the cathodic compartment was diluted to 5 ml to match the range of calibration curves for NO3−. Then, 0.1 ml of HCl (1 M) and 0.01 ml of sulfamic acid solution (0.8 wt %) were added successively and let sit for 30 min. The absorption intensities at 220 and 275 nm were recorded by UV-vis absorption spectroscopy. The final absorbance value was calculated by the following equation: A = A220nm − 2A275nm. The concentration of NO3− was calibrated using a series of standard NaNO3 solutions with different concentrations.
Calculation of yield rate and FE of NH3 and NO2−
The yield rate of NH3 and NO2− was calculated as
The FE of NH3 and NO2− was calculated as
where CNH3 and CNO2− are the concentrations of NH3 (aq) and NO2− (aq), respectively; V is the volume of electrolyte in the cathodic or anodic compartment (10 ml); MNH3 and MNO2− are the molar masses of NH3 (17 g mol−1) and NO2− (46 g mol−1), respectively; t is the electrolysis time (0.5 hours); mcat. is the catalyst mass on GCE; F is the Faradaic constant (96,485 C mol−1); and Q is the total charge passing the electrode calculated by evaluating the integral over the i-t curves.
For the calculation of FENH3 in eNOx−RR, both NO3− and NO2− concentrations were quantified before and after the electrolysis. The amount of NO2− converted from NO3− was negligible because it was far less than that in the electrolyte. The CNH3(from NO3−) and CNH3(from NO2−) were calculated as
and the FENH3 here was calculated as
Calculation of the partial current density for NH3
The partial current density for NH3 (jNH3) was calculated as
where FENH3 is the FE of NH3, i is the current during the chronoamperometric tests at given potentials, and t is the electrolysis time.
Calculation of turnover frequency
The turnover frequency (TOF) was calculated as
where Jk represents the kinetic current (in amperes) calculated by evaluating the integral over the i-t curves in Fig. 6C, α is the number of transfer electrons (8), nCu is the number of active Cu, and e is an elementary charge (1.6 × 10−19 C). nCu is estimated from the radius of Cu nanoparticle (~5 nm from fig. S3), density (8.96 g cm−3) (63), and surface atomic concentration (1.4 × 1019 m−2) (64).
NMR determination of NH3
The produced NH4+ was quantitatively determined by 1H nuclear magnetic resonance (1H-NMR). Dimethyl sulfoxide (DMSO)–d6 was used as the solvent and C4H4O4 as the internal standard. The calibration curve was made by a series of standard NH4Cl solutions with known concentrations. The standard NH4Cl solutions were prepared in 0.01 M HCl containing 50 mM NaNO3. The solution (0.5 ml) was mixed with 0.1 ml of DMSO-d6 containing 0.04 wt % C4H4O4 (20 mg C4H4O4 in 50 g of DMSO-d6). The mixture was tested by a Bruker Avance III 400 MHz spectrometer at room temperature. The calibration curve was achieved using the peak area ratio between NH4+ and C4H4O4. For testing the produced NH4+ from eNO3−RR, the pH of the obtained electrolyte was adjusted to weak acid by 0.01 M HCl before the test. Then, the testing processes are the same as that for making the calibration curve. The amount of produced NH4+ can be calculated from the peak area using the calibration curve.
Isotope labeling experiments
To clarify the N source of NH3, Na15NO3 was used as the feeding 15N source. Then, the produced 15NH4+ is determined by the same procedure as that in the NMR determination of the NH3 section.
pH detection
pH detection with a pH meter
The catalyst ink (5.6 μl) was dropped onto a double side of a carbon paper (0.5 cm × 0.4 cm), followed by natural drying at room temperature. The catalyst mass loading on the carbon paper (working electrode) was 0.12 mg cm−2. The eNO3−RR experiments were performed using a three-electrode system with the same conditions as the electrochemical measurements above, except for the use of a 30-ml H-type electrolytic cell with controlled stirring. The eNO3−RR was carried out in the potentiostatic mode at −1.5 V or in the galvanostatic mode at 80 mA cm−2 for 30 min. The pH on the working electrode surface during eNO3−RR was directly detected with a pH meter. For the pH detection in the flow cell system, the electrolyte of 500 ml and the carbon paper of 0.5 cm × 1 cm with the same catalyst loading of 0.12 mg cm−2 were used, and the pH meter was immersed in the bulk electrolyte.
pH detection with an RRDE
The pH on the electrode surface was measured using the RRDE technique, using a disk electrode (0.2475 cm2) with a Pt ring electrode (0.1866 cm2) (65). The potential of the Pt ring is sensitive to pH change and serves as an indicator for monitoring pH evolution on the surface of the disk electrode. A three-electrode system was used, including the disk electrode in RRDE, Ag/AgCl electrode (3 M KCl), and graphite rod, serving as the working, reference, and counter electrodes, respectively. The pH dependence of the OCP (EOCP) was then evaluated in H2-saturated electrolytes including 0.5 M Na2SO4 and 50 mM NaNO3 using the Pt ring electrode. The EOCP of the Pt ring electrode indicates the equilibrium potential of H+/H2, varying with electrolyte pH
where R, T, and F are the gas constant, absolute temperature, and Faraday constant, respectively.
The pH measurements were conducted in both single and H-type cells with an H2-saturated electrolytes including 0.5 M Na2SO4 and 50 mM NaNO3. The catalysts were coated on the disk electrode (working electrode) with a mass loading of 0.12 mg cm−2. The working electrode rotated at a speed of 0 or 1600 rpm. A galvanostatic test with a current density of 80 mA cm−2 for 30 min was performed on the disk electrode to obtain a constant number of electrons. EOCP was simultaneously recorded on the Pt ring electrode. The pH on the disk electrode coated by the catalyst can be deduced from the pH on the Pt ring electrode by the following equation (65)
where crt,H+ and crt,OH− are the concentrations of H+ and OH− on the Pt ring electrode, respectively; cd,H+ and cd,OH− are the concentrations of H+ and OH− on the disk electrode, respectively; c∞,H+ and c∞,OH− are the concentrations of H+ and OH− in the bulk electrolyte recorded by a pH meter, respectively; and ND = 0.37 is the collection efficiency of the Pt ring electrode (66).
Plasma electrocatalysis experiments
The plasma electrocatalysis reaction system consists of three parts: a plasma system, an absorber, and an H-type electrolyzer. A DBD system was used to activate and dissociate N2 and O2 molecules to produce NOx. The experimental setup consists of a plasma reactor, a power supply, and an electrical measurement system. The plasma reactor is composed of a quartz chamber and two metal plates that serve as the high-voltage and ground electrodes. The discharge was driven by a plasma generator with a power supply (CTP-2000 K, Nanjing Suman Plasma Technology Co., Ltd). The applied voltage is 80 V. Standard air was fed into the plasma reactor at a constant gas flow of 2000 sccm. The outlet gas was fed into a cylindrical absorber containing electrolytes with 0.2 M PB and 0.3 M Na2SO4 (pH 7). This electrolyte was exchanged for the 10-ml H-type electrolyzer with the same electrolyte through a reverse double pump system. The electrochemical measurements are similar to those used in the eNO3−RR, with an optimum potential of −1.6 V (fig. S31). The yield of NOx− and NH3 was quantized by UV-vis spectrophotometer measurements.
Theoretical simulation
Theoretical simulations were performed with the COMSOL Multiphysics 5.6 model. The Transport of Diluted Species module was used to estimate the COH− near the electrode surface. Two 2D models of the Cu nanoparticles in confined and free states were designed, corresponding to the Cu@hNCNC and Cu/hNCNC (fig. S35). The OH− is constantly generated on the Cu nanoparticle with a surface reaction rate of 3 mol m−2 s−1, which is estimated by the measured pH on the surface of electrodes in fig. S27. The generated OH− then diffuses to the boundary of the simulation box due to the fixed COH− of 10−7 M (pH 7). The OH− diffusion coefficient is 5.293 × 10−9 m2 s−1, and the simulation box is 40 nm × 40 nm with an initial COH− of 10−7 M (pH 7) inside (67). All the simulations were equilibrated with a time range of 30 min, which is long enough to get the stationary states.
Acknowledgments
We thank the Shanghai Synchrotron Radiation Facility (BL11B, SSRF) and J. Li for help in characterizations.
Funding: This work was jointly supported by the National Key Research and Development Program of China (no. 2021YFA1500900), the National Natural Science Foundation of China (nos. 21832003, 21972061, and 52071174), and the Natural Science Foundation of Jiangsu Province (Major Project: BK20212005). The numerical calculation was done on the computing facilities in the High Performance Computing Center (HPCC) of Nanjing University.
Author contributions: Z.H., Q.W., and Z.S. conceived and designed the experiments. Z.S. performed the experiments. G.C. synthesized the catalyst samples. X.C. and F.X. helped with the electrochemical measurements. L.Y. conducted theoretical simulations. Z.S., Z.H., Q.W., X.W., and H.H. analyzed the experimental data and cowrote the paper. All the authors discussed the results and commented on the manuscript.
Competing interests: The authors declare that they have no competing interests.
Data and materials availability: All data needed to evaluate the conclusions in the paper are present in the paper and/or the Supplementary Materials.
Supplementary Materials
This PDF file includes:
Figs. S1 to S40
Tables S1 to S4
References
REFERENCES AND NOTES
- 1.Erisman J. W., Sutton M. A., Galloway J., Klimont Z., Winiwarter W., How a century of ammonia synthesis changed the world. Nat. Geosci. 1, 636–639 (2008). [Google Scholar]
- 2.Guo J. P., Chen P., Catalyst: NH3 as an energy carrier. Chem 3, 709–712 (2017). [Google Scholar]
- 3.Chen J. G., Crooks R. M., Seefeldt L. C., Bren K. L., Bullock R. M., Darensbourg M. Y., Holland P. L., Hoffman B., Janik M. J., Jones A. K., Kanatzidis M. G., King P., Lancaster K. M., Lymar S. V., Pfromm P., Schneider W. F., Schrock R. R., Beyond fossil fuel-driven nitrogen transformations. Science 360, eaar6611 (2018). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 4.Wang M., Khan M. A., Mohsin I., Wicks J., Ip A. H., Sumon K. Z., Dinh C.-T., Sargent E. H., Gates I. D., Kibria M. G., Can sustainable ammonia synthesis pathways compete with fossil-fuel based Haber–Bosch processes? Energy Environ. Sci. 14, 2535–2548 (2021). [Google Scholar]
- 5.Soloveichik G., Electrochemical synthesis of ammonia as a potential alternative to the Haber-Bosch process. Nat. Catal. 2, 377–380 (2019). [Google Scholar]
- 6.Zhao X., Hu G., Chen G. F., Zhang H., Zhang S., Wang H., Comprehensive understanding of the thriving ambient electrochemical nitrogen reduction reaction. Adv. Mater. 33, e2007650 (2021). [DOI] [PubMed] [Google Scholar]
- 7.Qing G., Ghazfar R., Jackowski S. T., Habibzadeh F., Ashtiani M. M., Chen C. P., Smith M. R. III, Hamann T. W., Recent advances and challenges of electrocatalytic N2 reduction to ammonia. Chem. Rev. 120, 5437–5516 (2020). [DOI] [PubMed] [Google Scholar]
- 8.Andersen S. Z., Čolić V., Yang S., Schwalbe J. A., Nielander A. C., McEnaney J. M., Enemark-Rasmussen K., Baker J. G., Singh A. R., Rohr B. A., Statt M. J., Blair S. J., Mezzavilla S., Kibsgaard J., Vesborg P. C. K., Cargnello M., Bent S. F., Jaramillo T. F., Stephens I. E. L., Nørskov J. K., Chorkendorff I., A rigorous electrochemical ammonia synthesis protocol with quantitative isotope measurements. Nature 570, 504–508 (2019). [DOI] [PubMed] [Google Scholar]
- 9.Suryanto B. H. R., Du H. L., Wang D. B., Chen J., Simonov A. N., MacFarlane D. R., Challenges and prospects in the catalysis of electroreduction of nitrogen to ammonia. Nat. Catal. 2, 290–296 (2019). [Google Scholar]
- 10.Tang C., Qiao S. Z., How to explore ambient electrocatalytic nitrogen reduction reliably and insightfully. Chem. Soc. Rev. 48, 3166–3180 (2019). [DOI] [PubMed] [Google Scholar]
- 11.Choi J., Suryanto B. H. R., Wang D., Du H. L., Hodgetts R. Y., Ferrero Vallana F. M., MacFarlane D. R., Simonov A. N., Identification and elimination of false positives in electrochemical nitrogen reduction studies. Nat. Commun. 11, 5546 (2020). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 12.Wang Y., Wang C., Li M., Yu Y., Zhang B., Nitrate electroreduction: Mechanism insight, in situ characterization, performance evaluation, and challenges. Chem. Soc. Rev. 50, 6720–6733 (2021). [DOI] [PubMed] [Google Scholar]
- 13.Van Langevelde P. H., Katsounaros I., Koper M. T. M., Electrocatalytic nitrate reduction for sustainable ammonia production. Joule 5, 290–294 (2021). [Google Scholar]
- 14.Wang Y., Zhou W., Jia R., Yu Y., Zhang B., Unveiling the activity origin of a copper-based electrocatalyst for selective nitrate reduction to ammonia. Angew. Chem. Int. Ed. Engl. 59, 5350–5354 (2020). [DOI] [PubMed] [Google Scholar]
- 15.Li P. P., Jin Z. Y., Fang Z. W., Yu G. H., A single-site iron catalyst with preoccupied active centers that achieves selective ammonia electrosynthesis from nitrate. Energy Environ. Sci. 14, 3522–3531 (2021). [Google Scholar]
- 16.Lehnert N., Dong H. T., Harland J. B., Hunt A. P., White C. J., Reversing nitrogen fixation. Nat. Rev. Chem. 2, 278–289 (2018). [Google Scholar]
- 17.Galloway J. N., Townsend A. R., Erisman J. W., Bekunda M., Cai Z., Freney J. R., Martinelli L. A., Seitzinger S. P., Sutton M. A., Transformation of the nitrogen cycle: Recent trends, questions, and potential solutions. Science 320, 889–892 (2008). [DOI] [PubMed] [Google Scholar]
- 18.Katsounaros I., On the assessment of electrocatalysts for nitrate reduction. Curr. Opini. Electrochem. 28, 100721 (2021). [Google Scholar]
- 19.Chen C., Li Y., Yang P., Address the “alkalinity problem” in CO2 electrolysis with catalyst design and translation. Joule 5, 737–742 (2021). [Google Scholar]
- 20.Hu Q., Qin Y. J., Wang X. D., Wang Z. Y., Huang X. W., Zheng H. J., Gao K. R., Yang H. P., Zhang P. X., Shao M. H., He C. X., Reaction intermediate-mediated electrocatalyst synthesis favors specified facet and defect exposure for efficient nitrate-ammonia conversion. Energy Environ. Sci. 14, 4989–4997 (2021). [Google Scholar]
- 21.Deng X., Yang Y., Wang L., Fu X. Z., Luo J. L., Metallic Co nanoarray catalyzes selective NH3 production from electrochemical nitrate reduction at current densities exceeding 2 A cm−2. Adv. Sci. 8, 2004523 (2021). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 22.Wang Y., Qin X. P., Shao M. H., First-principles mechanistic study on nitrate reduction reactions on copper surfaces: Effects of crystal facets and pH. J. Catal. 400, 62–70 (2021). [Google Scholar]
- 23.Hu T., Wang C., Wang M., Li C. M., Guo C., Theoretical insights into superior nitrate reduction to ammonia performance of copper catalysts. ACS Catal. 11, 14417–14427 (2021). [Google Scholar]
- 24.Chen F. Y., Wu Z. Y., Gupta S., Rivera D. J., Lambeets S. V., Pecaut S., Kim J. Y. T., Zhu P., Finfrock Y. Z., Meira D. M., King G., Gao G. H., Xu W. Q., Cullen D. A., Zhou H., Han Y., Perea D. E., Muhich C. L., Wang H. T., Efficient conversion of low-concentration nitrate sources into ammonia on a Ru-dispersed Cu nanowire electrocatalyst. Nat. Nanotechnol. 17, 759–767 (2022). [DOI] [PubMed] [Google Scholar]
- 25.Yang J., Qi H., Li A., Liu X., Yang X., Zhang S., Zhao Q., Jiang Q., Su Y., Zhang L., Li J. F., Tian Z. Q., Liu W., Wang A., Zhang T., Potential-driven restructuring of Cu single atoms to nanoparticles for boosting the electrochemical reduction of nitrate to ammonia. J. Am. Chem. Soc. 144, 12062–12071 (2022). [DOI] [PubMed] [Google Scholar]
- 26.Gao Q., Pillai H. S., Huang Y., Liu S. K., Mu Q. M., Han X., Yan Z. H., Zhou H., He Q., Xin H. L., Zhu H. Y., Breaking adsorption-energy scaling limitations of electrocatalytic nitrate reduction on intermetallic CuPd nanocubes by machine-learned insights. Nat. Commun. 13, 2338 (2022). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 27.Wagner A., Sahm C. D., Reisner E., Towards molecular understanding of local chemical environment effects in electro- and photocatalytic CO2 reduction. Nat. Catal. 3, 775–786 (2020). [Google Scholar]
- 28.Huang J. E., Li F., Ozden A., Sedighian Rasouli A., García de Arquer F. P., Liu S., Zhang S., Luo M., Wang X., Lum Y., Xu Y., Bertens K., Miao R. K., Dinh C.-T., Sinton D., Sargent E. H., CO2electrolysis to multicarbon products in strong acid. Science 372, 1074–1078 (2021). [DOI] [PubMed] [Google Scholar]
- 29.Wakerley D., Lamaison S., Ozanam F., Menguy N., Mercier D., Marcus P., Fontecave M., Mougel V., Bio-inspired hydrophobicity promotes CO2 reduction on a Cu surface. Nat. Mater. 18, 1222–1227 (2019). [DOI] [PubMed] [Google Scholar]
- 30.Liu M., Pang Y., Zhang B., De Luna P., Voznyy O., Xu J., Zheng X., Dinh C. T., Fan F., Cao C., de Arquer F. P., Safaei T. S., Mepham A., Klinkova A., Kumacheva E., Filleter T., Sinton D., Kelley S. O., Sargent E. H., Enhanced electrocatalytic CO2 reduction via field-induced reagent concentration. Nature 537, 382–386 (2016). [DOI] [PubMed] [Google Scholar]
- 31.Yu S., Kim D., Qi Z., Louisia S., Li Y., Somorjai G. A., Yang P., Nanoparticle assembly induced ligand interactions for enhanced electrocatalytic CO2Conversion. J. Am. Chem. Soc. 143, 19919–19927 (2021). [DOI] [PubMed] [Google Scholar]
- 32.Wang X., Jiao Y., Li L., Zheng Y., Qiao S. Z., Local environment determined reactant adsorption configuration for enhanced electrocatalytic acetone hydrogenation to propane. Angew. Chem. Int. Ed. 61, e202114253 (2022). [DOI] [PubMed] [Google Scholar]
- 33.Veenstra F. L. P., Ackerl N., Martín A. J., Pérez-Ramírez J., Laser-microstructured copper reveals selectivity patterns in the electrocatalytic reduction of CO2. Chem 6, 1707–1722 (2020). [Google Scholar]
- 34.Wu Q., Yang L., Wang X., Hu Z., From carbon-based nanotubes to nanocages for advanced energy conversion and storage. Acc. Chem. Res. 50, 435–444 (2017). [DOI] [PubMed] [Google Scholar]
- 35.Wu Q., Yang L., Wang X., Hu Z., Carbon-based nanocages: A new platform for advanced energy storage and conversion. Adv. Mater. 32, e1904177 (2020). [DOI] [PubMed] [Google Scholar]
- 36.Wu Q., Yang L. J., Wang X. Z., Hu Z., Mesostructured carbon-based nanocages: An advanced platform for energy chemistry. Sci. China Chem. 63, 665–681 (2020). [Google Scholar]
- 37.Yu Z. H., Ji N., Li X. Y., Zhang R., Qiao Y. N., Xiong J., Liu J., Lu X. B., Kinetics driven by hollow nanoreactors: An opportunity for controllable catalysis. Angew. Chem. Int. Ed. 62, e2022136 (2022). [DOI] [PubMed] [Google Scholar]
- 38.Anantharaj S., Noda S., Jothi V. R., Yi S., Driess M., Menezes P. W., Strategies and perspectives to catch the missing pieces in energy-efficient hydrogen evolution reaction in alkaline media. Angew. Chem. Int. Ed. Engl. 60, 18981–19006 (2021). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 39.Wu Z. Y., Karamad M., Yong X., Huang Q., Cullen D. A., Zhu P., Xia C., Xiao Q., Shakouri M., Chen F. Y., Kim J. Y. T., Xia Y., Heck K., Hu Y., Wong M. S., Li Q., Gates I., Siahrostami S., Wang H., Electrochemical ammonia synthesis via nitrate reduction on Fe single atom catalyst. Nat. Commun. 12, 2870 (2021). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 40.Chen G. F., Yuan Y. F., Jiang H. F., Ren S. Y., Ding L. X., Ma L., Wu T. P., Lu J., Wang H. H., Electrochemical reduction of nitrate to ammonia via direct eight-electron transfer using a copper-molecular solid catalyst. Nat. Energy 5, 605–613 (2020). [Google Scholar]
- 41.Guo Y., Zhang R., Zhang S. C., Zhao Y. W., Yang Q., Huang Z. D., Dong B. B., Zhi C. Y., Pd doping-weakened intermediate adsorption to promote electrocatalytic nitrate reduction on TiO2 nanoarrays for ammonia production and energy supply with zinc-nitrate batteries. Energy Environ. Sci. 14, 3938–3944 (2021). [Google Scholar]
- 42.Jia R. R., Wang Y. T., Wang C. H., Ling Y. F., Yu Y. F., Zhang B., Boosting selective nitrate electroreduction to ammonium by constructing oxygen vacancies in TiO2. ACS Catal. 10, 3533–3540 (2020). [Google Scholar]
- 43.Wang C., Liu Z., Hu T., Li J., Dong L., Du F., Li C., Guo C., Metasequoia-like nanocrystal of iron-doped copper for efficient electrocatalytic nitrate reduction into ammonia in neutral media. ChemSusChem 14, 1825–1829 (2021). [DOI] [PubMed] [Google Scholar]
- 44.Ren T., Yu Z., Yu H., Deng K., Wang Z., Li X., Wang H., Wang L., Xu Y., Interfacial polarization in metal-organic framework reconstructed Cu/Pd/CuOx multi-phase heterostructures for electrocatalytic nitrate reduction to ammonia. Appl. Catal. B: Environ. 318, 121805 (2022). [Google Scholar]
- 45.Sun W. J., Ji H. Q., Li L. X., Zhang H. Y., Wang Z. K., He J. H., Lu J. M., Built-in Electric field triggered interfacial accumulation effect for efficient nitrate removal at ultra-low concentration and electroreduction to ammonia. Angew. Chem. Int. Ed. 60, 22933–22939 (2021). [DOI] [PubMed] [Google Scholar]
- 46.Qiu W. X., Chen X. J., Liu Y. T., Xiao D., Wang P. F., Li R., Liu K., Jin Z. Y., Li P. P., Confining intermediates within a catalytic nanoreactor facilitates nitrate-to-ammonia electrosynthesis. Appl. Catal. B: Environ. 315, 121548 (2022). [Google Scholar]
- 47.Jiang M., Su J., Song X., Zhang P., Zhu M., Qin L., Tie Z., Zuo J. L., Jin Z., Interfacial reduction nucleation of noble metal nanodots on redox-active metal-organic frameworks for high-efficiency electrocatalytic conversion of nitrate to ammonia. Nano Lett. 22, 2529–2537 (2022). [DOI] [PubMed] [Google Scholar]
- 48.Cheng X. F., He J. H., Ji H. Q., Zhang H. Y., Cao Q., Sun W. J., Yan C. L., Lu J. M., Coordination symmetry breaking of single-atom catalysts for robust and Efficient nitrate electroreduction to ammonia. Adv. Mater. 34, e2205767 (2022). [DOI] [PubMed] [Google Scholar]
- 49.Yu Y., Wang C. H., Yu Y. F., Wang Y. T., Zhang B., Promoting selective electroreduction of nitrates to ammonia over electron-deficient Co modulated by rectifying Schottky contacts. Sci. China Chem. 63, 1469–1476 (2020). [Google Scholar]
- 50.Gao P., Xue Z. H., Zhang S. N., Xu D., Zhai G. Y., Li Q. Y., Chen J. S., Li X. H., Schottky barrier-induced surface electric field boosts universal reduction of NOx– in water to ammonia. Angew. Chem. Int. Ed. Engl. 60, 20711–20716 (2021). [DOI] [PubMed] [Google Scholar]
- 51.Zhang W.-D., Dong H., Zhou L., Xu H., Wang H.-R., Yan X., Jiang Y., Zhang J., Gu Z.-G., Fe single-atom catalysts with pre-organized coordination structure for efficient electrochemical nitrate reduction to ammonia. Appl. Catal. B: Environ. 317, 121750 (2022). [Google Scholar]
- 52.Li L., Tang C., Cui X., Zheng Y., Wang X., Xu H., Zhang S., Shao T., Davey K., Qiao S. Z., Efficient nitrogen fixation to ammonia through integration of plasma oxidation with electrocatalytic reduction. Angew. Chem. Int. Ed. Engl. 60, 14131–14137 (2021). [DOI] [PubMed] [Google Scholar]
- 53.Gao Z., Lai Y., Tao Y., Xiao L., Zhang L., Luo F., Constructing well-defined and robust Th-MOF-supported single-site copper for production and storage of ammonia from electroreduction of nitrate. ACS Cent. Sci. 7, 1066–1072 (2021). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 54.Ren Y., Yu C., Wang L., Tan X., Wang Z., Wei Q., Zhang Y., Qiu J., Microscopic-level insights into the mechanism of enhanced NH3 synthesis in plasma-enabled cascade N2 oxidation–electroreduction system. J. Am. Chem. Soc. 144, 10193–10200 (2022). [DOI] [PubMed] [Google Scholar]
- 55.Ye S., Chen Z., Zhang G., Chen W., Peng C., Yang X., Zheng L., Li Y., Ren X., Cao H., Xue D., Qiu J., Zhang Q., Liu J., Elucidating the activity, mechanism and application of selective electrosynthesis of ammonia from nitrate on cobalt phosphide. Energy Environ. Sci. 15, 760–770 (2022). [Google Scholar]
- 56.Liu H. M., Lang X. Y., Zhu C., Timoshenko J., Rüscher M., Bai L. C., Guijarro N., Yin H. B., Peng Y., Li J. H., Liu Z., Wang W. C., Cuenya B. R., Luo J. S., Efficient electrochemical nitrate reduction to ammonia with copper-supported rhodium cluster and single-atom catalysts. Angew. Chem. Int. Ed. Engl. 61, e202202556 (2022). [DOI] [PubMed] [Google Scholar]
- 57.Song Z., Liu Y., Zhong Y., Guo Q., Zeng J., Geng Z., Efficient electroreduction of nitrate into ammonia at ultralow concentrations via an enrichment effect. Adv. Mater. 34, e2204306 (2022). [DOI] [PubMed] [Google Scholar]
- 58.Wang J., Wu D., Li M., Wei X., Yang X., Shao M., Gu M., Bismuth ferrite as an electrocatalyst for the electrochemical nitrate reduction. Nano Lett. 22, 5600–5606 (2022). [DOI] [PubMed] [Google Scholar]
- 59.Xie K., Qin X., Wang X., Wang Y., Tao H., Wu Q., Yang L., Hu Z., Carbon nanocages as supercapacitor electrode materials. Adv. Mater. 24, 347–352 (2012). [DOI] [PubMed] [Google Scholar]
- 60.Zhao J., Lai H., Lyu Z., Jiang Y., Xie K., Wang X., Wu Q., Yang L., Jin Z., Ma Y., Liu J., Hu Z., Hydrophilic hierarchical nitrogen-doped carbon nanocages for ultrahigh supercapacitive performance. Adv. Mater. 27, 3541–3545 (2015). [DOI] [PubMed] [Google Scholar]
- 61.Shen L., Sun T., Zhuo O., Che R., Li D., Ji Y., Bu Y., Wu Q., Yang L., Chen Q., Wang X., Hu Z., Alcohol-tolerant platinum electrocatalyst for oxygen reduction by encapsulating platinum nanoparticles inside nitrogen-doped Carbon nanocages. ACS Appl. Mater. Interfaces 8, 16664–16669 (2016). [DOI] [PubMed] [Google Scholar]
- 62.Zhang Z., Wu Q., Mao K., Chen Y., Du L., Bu Y., Zhuo O., Yang L., Wang X., Hu Z., Efficient ternary synergism of platinum/tin oxide/nitrogen-doped carbon leading to high-performance ethanol oxidation. ACS Catal. 8, 8477–8483 (2018). [Google Scholar]
- 63.J. G. Speight, Lange's Handbook of Chemistry (McGraw-Hill Education, 2017). [Google Scholar]
- 64.Parris G. E., Klier K., The specific copper surface areas in Cu/ZnO methanol synthesis catalysts by oxygen and carbon monoxide chemisorption: Evidence for irreversible CO chemisorption induced by the interaction of the catalyst components. J. Catal. 97, 374–384 (1986). [Google Scholar]
- 65.Yokoyama Y., Miyazaki K., Miyahara Y., Fukutsuka T., Abe T., In situ measurement of local pH at working electrodes in neutral pH solutions by the rotating ring-disk electrode technique. ChemElectroChem 6, 4750–4756 (2019). [Google Scholar]
- 66.Zheng X. Z., Shi X. Y., Ning H. H., Yang R., Lu B., Luo Q., Mao S. J., Xi L. L., Wang Y., Tailoring a local acid-like microenvironment for efficient neutral hydrogen evolution. Nat. Commun. 14, 4209 (2023). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 67.Welch A. J., Fenwick A. Q., Bohme A., Chen H. Y., Sullivan I., Li X. Q., DuChene J. S., Xiang C. X., Atwater H. A., OperandoLocal pH measurement within gas diffusion electrodes performing electrochemical carbon dioxide reduction. J. Phys. Chem. C 125, 20896–20904 (2021). [Google Scholar]
- 68.Yang Y., Louisia S., Yu S., Jin J. B., Roh I., Chen C. B., Guzman M. V. F., Feijóo J. L., Chen P. C., Wang H. S., Pollock C. J., Huang X., Shao Y. T., Wang C., Muller D. A., Abruña H. D., Yang P. D., Operando studies reveal active Cu nanograins for CO2 electroreduction. Nature 614, 262–269 (2023). [DOI] [PubMed] [Google Scholar]
- 69.Wang X., Klingan K., Klingenhof M., Möller T., Ferreira de Araújo J., Martens I., Bagger A., Jiang S., Rossmeisl J., Dau H., Strasser P., Morphology and mechanism of highly selective Cu(II) oxide nanosheet catalysts for carbon dioxide electroreduction. Nat. Commun. 12, 794 (2021). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 70.Giri S. D., Sarkar A., Electrochemical study of bulk and monolayer copper in alkaline solution. J. Electrochem. Soc. 163, H252–H259 (2016). [Google Scholar]
- 71.Feng T., Wang J., Wang Y., Yu C. F., Zhou X., Xu B. C., Laszlo K., Li F. T., Zhang W. X., Selective electrocatalytic reduction of nitrate to dinitrogen by Cu2O nanowires with mixed oxidation-state. Chem. Eng. J. 433, 133495 (2022). [Google Scholar]
- 72.De Luna P., Quintero-Bermudez R., Dinh C.-T., Ross M. B., Bushuyev O. S., Todorović P., Regier T., Kelley S. O., Yang P. D., Sargent E. H., Catalyst electro-redeposition controls morphology and oxidation state for selective carbon dioxide reduction. Nat. Catal. 1, 103–110 (2018). [Google Scholar]
- 73.Winter L. R., Chen J. G. G., N2 fixation by plasma-activated processes. Joule 5, 300–315 (2021). [Google Scholar]
- 74.Yang K. W., Han S. H., Wang Y. T., Zhang B., Yu Y. F., Sustainable production and in-place utilization of a liquid nitrogenous fertilizer. Joule 7, 1948–1955 (2023). [Google Scholar]
- 75.Rouwenhorst K. H. R., Jardali F., Bogaerts A., Lefferts L., From the Birkeland–Eyde process towards energy-efficient plasma-based NOXsynthesis: A techno-economic analysis. Energy Environ. Sci. 14, 2520–2534 (2021). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 76.Zhai G.-Y., Li Q.-Y., Zhang S.-N., Xu D., Xia S.-Y., Gao P., Lin X., Lin Y.-X., Cheng J.-H., Hu W.-Y., Sun L.-H., Li X.-H., Chen J.-S., Accelerating the activation of NOx- on Ru nanoparticles for ammonia production by tuning their electron deficiency. CCS Chem. 4, 3455–3462 (2022). [Google Scholar]
- 77.Gong Z., Zhong W., He Z., Liu Q., Chen H., Zhou D., Zhang N., Kang X., Chen Y., Regulating surface oxygen species on copper(I) oxides via plasma treatment for effective reduction of nitrate to ammonia. Appl. Catal. B: Environ. 305, 121021 (2022). [Google Scholar]
- 78.Jiang M., Zhu Q., Song X., Gu Y., Zhang P., Li C., Cui J., Ma J., Tie Z., Jin Z., Batch-scale synthesis of nanoparticle-agminated three-dimensional porous Cu@Cu2O microspheres for highly selective electrocatalysis of nitrate to ammonia. Environ. Sci. Technol. 56, 10299–10307 (2022). [DOI] [PubMed] [Google Scholar]
- 79.Li J., Zhan G., Yang J., Quan F., Mao C., Liu Y., Wang B., Lei F., Li L., Chan A. W. M., Xu L., Shi Y., Du Y., Hao W., Wong P. K., Wang J., Dou S. X., Zhang L., Yu J. C., Efficient ammonia electrosynthesis from nitrate on strained ruthenium nanoclusters. J. Am. Chem. Soc. 142, 7036–7046 (2020). [DOI] [PubMed] [Google Scholar]
- 80.Fu X. B., Zhao X. G., Hu X. B., He K., Yu Y. N., Li T., Tu Q., Qian X., Yue Q., Wasielewski M. R., Kang Y. J., Alternative route for electrochemical ammonia synthesis by reduction of nitrate on copper nanosheets. Appl. Mater. Today 19, 100620 (2020). [Google Scholar]
- 81.Wang Y., Xu A., Wang Z., Huang L., Li J., Li F., Wicks J., Luo M., Nam D. H., Tan C. S., Ding Y., Wu J., Lum Y., Dinh C. T., Sinton D., Zheng G., Sargent E. H., Enhanced nitrate-to-ammonia activity on copper-nickel alloys via tuning of intermediate adsorption. J. Am. Chem. Soc. 142, 5702–5708 (2020). [DOI] [PubMed] [Google Scholar]
- 82.Daiyan R., Tran-Phu T., Kumar P., Iputera K., Tong Z. Z., Leverett J., Khan M. H. A., Esmailpour A. A., Jalili A., Lim M., Tricoli A., Liu R. S., Lu X. Y., Lovell E., Amal R., Nitrate reduction to ammonium: From CuO defect engineering to waste NOx-to-NH3 economic feasibility. Energy Environ. Sci. 14, 3588–3598 (2021). [Google Scholar]
- 83.McEnaney J. M., Blair S. J., Nielander A. C., Schwalbe J. A., Koshy D. M., Cargnello M., Jaramillo T. F., Electrolyte engineering for efficient electrochemical nitrate reduction to ammonia on a titanium electrode. ACS Sustain. Chem. Eng. 8, 2672–2681 (2020). [Google Scholar]
- 84.Ge Z. X., Wang T. J., Ding Y., Yin S. B., Li F. M., Chen P., Chen Y., Interfacial engineering enhances the electroactivity of frame-like concave RhCu bimetallic nanocubes for nitrate reduction. Adv. Energy Mater. 12, 2103916 (2022). [Google Scholar]
- 85.Lv Y., Ke S. W., Gu Y. M., Tian B. L., Tang L. Y., Ran P., Zhao Y., Ma J., Zuo J.-L., Ding M. N., Highly efficient electrochemical nitrate reduction to ammonia in strong acid conditions with Fe2M-trinuclear-cluster metal–organic frameworks. Angew. Chem. Int. Ed. Engl. 62, e202305246 (2023). [DOI] [PubMed] [Google Scholar]
- 86.Zhang R., Li C., Cui H. L., Wang Y. B., Zhang S. C., Li P., Hou Y., Guo Y., Liang G. J., Huang Z. D., Peng C., Zhi C. Y., Electrochemical nitrate reduction in acid enables high-efficiency ammonia synthesis and high-voltage pollutes-based fuel cells. Nat. Commun. 14, 8036 (2023). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 87.Ren S. J., Gao R. T., Nguyen N. T., Wang L., Enhanced charge carrier dynamics on Sb2Se3 photocathodes for efficient photolectrochemical nitrate reduction to ammonia. Angew. Chem. Int. Ed. Engl. 63, e202317414 (2024). [DOI] [PubMed] [Google Scholar]
- 88.Meng S.-L., Zhang C., Ye C., Li J.-H., Zhou S., Zhu L., Li X.-B., Tung C.-H., Wu L.-Z., Cobaloximes: Selective nitrite reduction catalysts for tandem ammonia synthesis. Energy Environ. Sci. 16, 1590–1596 (2023). [Google Scholar]
- 89.Meng Z., Yao J. X., Sun C. N., Kang X., Gao R., Li H. R., Bi B., Zhu Y. F., Yan J. M., Jiang Q., Efficient ammonia production beginning from enhanced air activation. Adv. Energy Mater. 12, 2202105 (2022). [Google Scholar]
- 90.Gao R., Dai T. Y., Meng Z., Sun X. F., Liu D. X., Shi M. M., Li H. R., Kang X., Bi B., Zhang Y. T., Xu T. W., Yan J. M., Jiang Q., A bifunctional catalyst for green ammonia synthesis from ubiquitous air and water. Adv. Mater. 35, e2303455 (2023). [DOI] [PubMed] [Google Scholar]
- 91.Wu A. J., Yang J., Xu B., Wu X. Y., Wang Y. H., Lv X. J., Ma Y. C., Xu A. N., Zheng J. G., Tan Q. H., Peng Y. Q., Qi Z. F., Qi H. F., Li J. F., Wang Y. L., Harding J., Tu X., Wang A. Q., Yan J. H., Li X. D., Direct ammonia synthesis from the air via gliding arc plasma integrated with single atom electrocatalysis. Appl. Catal. B: Environ. 299, 120667 (2021). [Google Scholar]
Associated Data
This section collects any data citations, data availability statements, or supplementary materials included in this article.
Supplementary Materials
Figs. S1 to S40
Tables S1 to S4
References






