Abstract
Renewable energy-driven electrocatalytic nitrate reduction reaction presents a low-carbon and sustainable route for ammonia synthesis under mild conditions. Yet, the practical application of this process is currently hindered by unsatisfactory electrocatalytic activity and long-term stability. Herein we achieve high-rate ammonia electrosynthesis using a stable amorphous/crystalline dual-phase Cu catalyst. The ammonia partial current density and formation rate reach 3.33 ± 0.005 A cm−2 and 15.5 ± 0.02 mmol h−1 cm−2 at a low cell voltage of 2.6 ± 0.01 V, respectively. Remarkably, the dual-phase Cu catalyst can maintain stable ammonia production with a Faradaic efficiency of around 90% at a high current density of 1.5 A cm−2 for up to 300 h. A scale-up demonstration with an electrode size of 100 cm2 achieves an ammonia formation rate as high as 11.9 ± 0.5 g h−1 at a total current of 160 A. The impressive electrocatalytic performance is ascribed to the presence of stable amorphous Cu domains which promote the adsorption and hydrogenation of nitrogen-containing intermediates, thus improving reaction kinetics for ammonia formation. This work underscores the importance of stabilizing metastable amorphous structures for improving electrocatalytic reactivity and long-term stability.
Subject terms: Electrocatalysis, Materials for energy and catalysis, Energy
The authors develop an efficient and stable amorphous/crystalline dual-phase Cu catalyst towards electrocatalytic nitrate reduction reaction, with an ammonia Faradaic efficiency of around 90% at a high current density of 1.5 A cm−2 for up to 300 h.
Introduction
Ammonia (NH3) is currently produced by the energy-intensive Haber–Bosch process which converts nitrogen (N2) and hydrogen (H2) to NH3 at high temperatures (400−500 °C) and high pressures (10–30 MPa). The Haber–Bosch process accounts for 1–2% of global energy consumption and roughly 1% of global CO2 emission1,2. Alternative routes for green NH3 synthesis under mild conditions are highly desired. The renewable energy-driven electrocatalytic nitrate reduction reaction (NO3−RR), with NO3− (with concentrations from hundreds of ppm to >1 M) from wastewater being as a nitrogen source and water being as a hydrogen source, provides a low-carbon route for green NH3 synthesis under ambient conditions, with promising energy and environmental sustainability3,4.
The NH3 production via NO3−RR involves multiple proton and electron transfer steps (NO3− + 6H2O + 8e− → NH3 + 9OH−), leading to sluggish reaction kinetics5. The NO3−RR rate is governed by the adsorption of nitrogen species present during NO3−RR and the formation of active hydrogen species from water dissociation6,7. An efficient catalyst for NO3−RR should exhibit balanced adsorption of the above intermediates, thus directing the reaction pathway to NH3 at a high reaction rate while minimizing the formation of side products such as nitrite (NO2−) and H28,9. The adsorption of these intermediates is highly dependent on the geometric and electronic structures of metal catalysts, as demonstrated by remarkably distinct NO3−RR performances over noble metal (e.g., Pt, Ru, Pd)6,10,11, non-noble metal (e.g., Fe, Co, Ni, Cu)7,12–18, and bimetallic (e.g., CuPd, CuNi) catalysts19–23. Cu has been recognized as one of the most active non-noble metals, owing to its prominent ability for the conversion of NO3− to NO2−4,24. To date, while Cu-based catalysts have exhibited high NH3 Faradaic efficiency25, their activity and stability are still unsatisfactory for practical application.
Amorphous catalytic materials, with extremely disordered atomic arrangement in long range, possess high density of low-coordinated sites which are essential for the adsorption and activation of reactants and intermediates26,27. However, the amorphous materials are metastable and prone to suffering from crystallization during reaction, especially under highly reducing electrochemical conditions28–30. Here we report a stable amorphous/crystalline dual-phase Cu (a/c-Cu) catalyst for high-rate NH3 electrosynthesis via NO3−RR. The a/c-Cu catalyst exhibits a NH3 partial current density of 3.33 ± 0.005 A cm−2 and a NH3 formation rate of 15.5 ± 0.02 mmol h−1 cm−2 at a low cell voltage of 2.6 ± 0.01 V in an alkaline membrane electrode assembly (MEA) electrolyzer. Remarkably, the NH3 Faradaic efficiency maintains around 90% at an applied current density of 1.5 A cm−2 for 300 h. The impressive NO3−RR performance is rationalized by the presence of stable amorphous Cu domains which promote NH3 generation by optimizing the adsorption of N-containing intermediates and facilitating the formation of active hydrogen species from water dissociation.
Results
NO3−RR performance of a/c-Cu catalyst
The a/c-Cu catalyst was prepared by annealing a commercially available Cu foam in air at 600 °C followed by in situ electrochemical reduction during NO3−RR (Supplementary Fig. 1). The NO3−RR performance of the a/c-Cu catalyst was measured in a home-made alkaline MEA electrolyzer (Supplementary Fig. 2). The anolyte and catholyte were 1 M KOH and 1 M KOH + 0.2 M KNO3 solutions, respectively. The NO3−RR measurements were performed in the galvanostatic mode. The gas product (H2) and liquid products (NH3 and NO2−) were analyzed and quantified using an on-line gas chromatography and an ultraviolet–visible (UV–Vis) spectrophotometer, respectively (Supplementary Figs. 3–5). The quantification of NH3 was further validated by proton nuclear magnetic resonance (1H NMR) analysis (Supplementary Figs. 6–8). Herein, the annealed Cu foam is directly used as a working electrode. Such an integrated electrode is featured with facile preparation and shows significantly improved electrocatalytic performance compared to a conventional electrode with drop-casted catalyst powder (Supplementary Fig. 9). The three-dimensional and highly porous foam structure can remarkably facilitate mass transport especially at high current densities31,32. Figure 1a shows the Faradaic efficiencies and cell voltage as a function of applied current density over the a/c-Cu electrode. The a/c-Cu electrode achieves a peak NH3 Faradaic efficiency of 92 ± 0.87% at an applied current density of 1.5 A cm−2. The NH3 Faradaic efficiency still maintains over 83% at an applied current density up to 4.0 A cm−2, which translates into a NH3 partial current density of 3.33 ± 0.005 A cm−2 and a NH3 formation rate of 15.5 ± 0.02 mmol h−1 cm−2 at a low cell voltage of 2.6 ± 0.01 V (Supplementary Fig. 10). The effect of annealing temperature on the NO3−RR performance was also investigated by further measuring another two Cu foam samples annealed in air at 300 and 900 °C, respectively. The sample annealed at 600 °C (i.e., the a/c-Cu sample) shows the highest NO3−RR performance (Supplementary Fig. 11). The NH3 partial current density over our a/c-Cu electrode measured in 0.2 M KNO3 outperforms a raw Cu foam electrode (Supplementary Fig. 12) as well as previously reported values achieved even with a higher KNO3 concentration up to 1 M (Fig. 1b and Supplementary Table 1)6–8,12,17–20. The peak full-cell energy efficiency for NH3 production reaches 26 ± 1% at 1.0 A cm−2 and outperforms most of previous studies (Fig. 1c, Supplementary Fig. 9 and Table 2)14,23,33–35. The stability test of the a/c-Cu electrode was performed at an applied current density of 1.5 A cm−2. In a course of 300 h, the cell voltage is almost stable and the NH3 Faradaic efficiency maintains around 90% (Fig. 1d, Supplementary Fig. 13). We tried to scale up the NO3−RR process by increasing the geometric electrode area from 1 to 100 cm2 (Supplementary Fig. 14). As shown in Fig. 1e, the NH3 Faradaic efficiency reaches 94 ± 3.9% at an applied current of 160 A (the used direct current power supply has an upper current limit of 170 A) and a cell voltage of 2.23 ± 0.005 V. The highest NH3 formation rate reaches up to 11.9 ± 0.5 g h−1 (Supplementary Fig. 15a) and the corresponding energy consumption for NH3 production is as low as 108.3 ± 0.47 kJ g−1NH3 at 160 A (Supplementary Fig. 15b). The reported NO3−RR performance in terms of NH3 formation rate and energy consumption is superior to state-of-the-art cases in the literature (Fig. 1f)33–36. Overall, by turning a commercially available Cu foam into an active a/c-Cu catalyst through facile annealing treatment, we have achieved very promising selectivity, reaction rate, and stability towards industrial ammonia synthesis via NO3−RR.
Fig. 1. NO3−RR performance of a/c-Cu catalyst measured in MEA electrolyzers.
a Faradaic efficiencies and cell voltage as a function of applied current density. Comparisons of b performance and c full-cell energy efficiency reported in this work measured in a 1-cm2 electrolyzer with literature6–8,12,14,17–20,23,33–35. d Stability test at 1.5 A cm−2. e Scale-up performance with a 100-cm2 electrolyzer and f performance comparison with literature33–36. The error bars represent standard error of the mean and are made based on three fully separate and identical measurements. Source data are provided as a Source Data file.
Structure characterizations
To reveal structure-reactivity correlations of the a/c-Cu catalyst for the impressive NO3−RR performance, a series of structural characterizations and control experiments were conducted. The annealed Cu foam in air at 600 °C was firstly characterized by X-ray diffraction (XRD) and scanning electron microscopy (SEM). The characteristic XRD pattern of a pure CuO phase indicates that the presence of a thick CuO layer on the raw Cu foam after annealing treatment in air (Supplementary Fig. 16). Meanwhile, the original flat surface of the raw Cu foam becomes very rough with irregularly shaped grains (Supplementary Fig. 17). After NO3−RR, these large grains disappear and form a rough and porous layer (Supplementary Fig. 18), along with the reduction of CuO revealed by the XRD pattern of the post-reaction sample (Fig. 2a). The operando Cu K-edge X-ray absorption near edge structure (XANES) spectra of the a/c-Cu catalyst further confirm that CuO is fully reduced to metallic Cu during NO3−RR (Fig. 2b and Supplementary Fig. 19). This is consistent with the results obtained from the Cu LMM Auger spectrum of the quasi in situ X-ray photoelectron spectroscopy (XPS) analysis (Supplementary Fig. 20)37. The Fourier transformation of the extended X-ray absorption fine-structure (EXAFS) spectra show decreased peak intensity of the Cu-Cu coordination at 2.3 Å (Fig. 2c), and the average Cu coordination number estimated by fitting results is around 6.7 ± 0.4 (Supplementary Figs. 21, 22 and Supplementary Table 3). These results indicate the presence of abundant low-coordinated Cu sites that are in situ generated during NO3−RR (Supplementary Fig. 23)38,39. High-resolution transmission electron microscopy (HRTEM) images show that the a/c-Cu catalyst consists of abundant amorphous domains which are separated by crystalline regions (Fig. 2d and Supplementary Fig. 24). The amorphous/crystalline dual-phase structure is further confirmed by the corresponding fast Fourier transform (FFT) patterns acquired from amorphous and crystalline regions (Fig. 2e). High-angle annular dark-field scanning transmission electron microscopy (HAADF-STEM) images and energy-dispersive X-ray spectroscopy (EDS) elemental mapping analysis of the a/c Cu catalyst clearly show the co-existence of disordered Cu domains and crystalline Cu domains (Fig. 2f and Supplementary Fig. 25).
Fig. 2. Structure characterizations.
a XRD patterns of a/c-Cu and a/c-Cu-H2 catalysts after NO3−RR. Operando b normalized XANES and c Fourier-transformed EXAFS spectra of a/c-Cu catalyst at a working potential of –0.8 V (vs. RHE). d HRTEM images of a/c-Cu catalyst and e corresponding FFT patterns acquired from different regions. f HAADF-STEM image and corresponding EDS element map of a/c-Cu catalyst. g Ratios of amorphous Cu domains. The error bars represent standard error of the mean and are made based on ten fully separate and identical measurements. h DSC curves. i CV curves of Pb stripping. j NH3 partial current density at a cell voltage of 2.35 V over a/c-Cu and a/c-Cu-H2 catalysts. The error bars represent standard error of the mean and are made based on three fully separate and identical measurements. Source data are provided as a Source Data file.
To reveal whether the impressive NO3−RR performance correlates with the presence of amorphous domains, we further treated the Cu foam annealed in air at 600 °C by annealing in H2 at 350 °C prior to NO3−RR and the control sample was denoted as a/c-Cu-H2 (Supplementary Figs. 17c and 18c). As indicated by the much sharper XRD characteristic Cu patterns in Fig. 2a, the H2 annealing treatment results in a significant increase in the crystallinity of the a/c-Cu-H2 catalyst, compared with the a/c-Cu catalyst. Accordingly, HRTEM images of the a/c-Cu-H2 catalyst indicate that the amorphous domains reduce drastically (Supplementary Fig. 26). Quantitative and statistical analysis based on ten typical HRTEM images for each sample demonstrates that the ratio of the amorphous domains decreases from 24.1 ± 3.3% over a/c-Cu to 5.5 ± 2.7% over a/c-Cu-H2 (Fig. 2g). The increased crystallinity (thus, decreased amorphization) induced by the H2 annealing treatment is further evidenced by differential scanning calorimetry (DSC) measurements (Fig. 2h). The exothermic peak observed at around 133 °C is assigned to the crystallization of the amorphous Cu domains40. The larger exothermic peak suggests a higher ratio of the amorphous Cu domains in the a/c-Cu catalyst. It is widely accepted that amorphous materials possess abundant low-coordinated sites owing to their long-range atomic disorder26. The Pb underpotential deposition (Pb-UPD) was used to characterize the surfaces of the a/c-Cu and a/c-Cu-H2 catalysts. For the a/c-Cu catalyst, the Pb stripping peak at −0.15 V vs. Ag/AgCl which is assigned to low-coordinated sites is more prominent (Fig. 2i)41, consistent with its higher ratio of the amorphous domains (Fig. 2g). The NO3−RR performance of the a/c-Cu-H2 catalyst was also measured under identical reaction conditions (Supplementary Fig. 27). As shown in Fig. 2j, the NH3 current density over the a/c-Cu-H2 catalyst is much lower than that over the a/c-Cu catalyst at a cell voltage of 2.35 V. These control experiments indicate that the improved ammonia synthesis is positively correlated with the amorphous Cu domains.
It has been reported that amorphous materials are prone to suffering from atomic rearrangement processes such as crystallization under electrochemical reaction conditions29,42. We further conducted Pb-UPD measurements over the a/c-Cu catalyst after different reaction durations (0.5, 1.0, and 2.0 h). As shown in Supplementary Figs. 28 and 29, the ratios of main facets and the amounts of low-coordinated sites (thus, amorphous domains) are almost constant over time, indicative of good structural stability of the dual-phase a/c-Cu catalyst during NO3−RR. In contrast, a pure amorphous Cu catalyst (a-Cu) quickly turns crystalline under NO3−RR conditions (Supplementary Fig. 30). Therefore, it is speculated that the amorphous Cu domains in the a/c Cu catalyst are likely kinetically stabilized by adjacent crystalline Cu domains43,44. The a/c-Cu catalyst after reaction at 1.5 A cm−2 for over 50 h is further characterized by HAADF-STEM (Supplementary Figs. 31 and 32), and the amorphous Cu domains can be still observed, further indicating the good structural stability. Moreover, as evidenced by the electrochemical impedance spectra (EIS) in Supplementary Fig. 33, the poor conductivity of amorphous catalysts45 has also been largely improved owing to the presence of adjacent crystalline Cu domains, resulting in low Ohmic resistance and thus low energy consumption for NH3 production (Fig. 1c, f). Nevertheless, the stable amorphous/crystalline dual-phase structure of the a/c-Cu catalyst, is responsible for the high current density and long-term stability towards NH3 synthesis.
Spectroscopic investigations
The NH3 formation from NO3−RR involves multiple deoxygenation and hydrogenation steps which are closely associated with the adsorption of NO3− and N-containing intermediates as well as the generation of active hydrogen species from water dissociation6,20. To figure out how the dual-phase a/c-Cu catalyst promotes NO3−RR to NH3, thorough spectroscopic characterizations and mechanistic investigation experiments were conducted (Supplementary Figs. 34 and 35). In situ Raman spectroscopy measurements were performed in a home-made flow cell to observe the adsorption and reduction of NO3− on the surface of the a/c-Cu catalyst. As shown in Fig. 3a, upon applying potentials, the NO3− species in the solution (1050 cm−1)46 begin to adsorb onto catalyst surface, as evidenced by the characteristic Raman peaks at 1005 and 1370 cm−1 which correspond to the NO stretching vibration47 and the NO2 antisymmetric stretching vibration48 of the adsorbed NO3− species, respectively. The Raman peak at 1098 cm−1, assigned to copper-nitrate complexes, also suggests an interaction between copper and nitrate47. The Raman peaks at 1120 and 1259 cm−1 are typically assigned to the symmetric and antisymmetric stretching vibration of the important adsorbed NO2− intermediate in a nitro configuration47,48. Furthermore, the presence of the chelating nitrito configuration (*ONO*−, at 1301 cm−1) and the bridging nitro configuration (*N*OO−, at 1435 cm−1) further demonstrating the strong adsorption capacity for NOx intermediates on the a/c-Cu catalyst47,48. The broad Raman peak at 1534 cm−1 is likely be assigned to the ν(N = O) of HNO* (1530 cm−1) and the antisymmetric bending vibration of the HNH of NH3 (1550 cm−1)47,48. However, when the applied potential becomes more negative, the intensities of NOx peaks decrease significantly while that of the Cu-NH2 (387 and 446 cm−1)49 and the N–H bending mode of NH2 (1654 cm−1)50 increase, indicative of the fast conversion of NOx species. On the other hand, these important intermediates are almost invisible over the a/c-Cu-H2 catalyst (Supplementary Fig. 36). This is probably associated with the presence of the amorphous Cu domains with abundant low-coordinated sites that are considered to be active for NO3−RR51,52. The presence of the adsorbed species over the a/c-Cu catalyst was further validated by in situ attenuated total reflectance surface-enhanced infrared absorption spectroscopy (ATR-SEIRAS) measurements (Fig. 3c). The downward and upward IR bands at 1340 and 1256 cm−1 confirm the consumption and generation of NO3− and *NO2, respectively53. At the same time, the N–H bending band at 1442 cm−1 and the NH2 wagging band at 1294 cm−1 are also detected owing to NH3 formation54,55.
Fig. 3. Spectroscopic investigations.
a In situ Raman spectra over a/c-Cu catalyst at different potentials (vs. RHE). b EPR spectra of outlet electrolytes with or without KNO3 using DMPO as radical trapping reagent. c In situ ATR-SEIRAS spectra over a/c-Cu catalyst at different potentials (vs. RHE) and d KIEs over a/c-Cu and a/c-Cu-H2 catalysts. Source data are provided as a Source Data file.
Apart from the adsorbed species, potential-dependent surface structure changes are also observed by in situ Raman measurements. The disappearance of the Raman peak at 298, 343, and 609 cm−1 assigned to CuO at open circuit potential (OCP) indicates the reduction of CuO to metallic Cu during NO3−RR (Fig. 3a), consistent with XRD, quasi in situ XPS, and XANES results. Meanwhile, the two emerging Raman peaks at 539 and 716 cm−1 can be assigned to the symmetric stretching of Cu−OH56,57 and the bending mode of Cu−OH58,59, respectively. The surface-adsorbed –OH is considered to be produced from water dissociation during NO3−RR (Fig. 3a). In contrast, only minor symmetric stretching of Cu−OH Raman peak can be observed over the a/c-Cu-H2 catalyst (Supplementary Fig. 36), indicating the more favorable water dissociation owing to the presence of the amorphous Cu domains. The improved water disassociation is further verified by directly detecting the other dissociation product, adsorbed hydrogen (*H) species, at an applied current density of 2 A cm−2. The *H detection was performed through being trapped with 5,5-dimethyl-1-pyrroline-N-oxide (DMPO) followed by electron paramagnetic resonance (EPR) measurements. Upon electrolysis in the absence of NO3− at 2 A cm−2, the DMPO-H intensity of the a/c-Cu catalyst is much higher than that of the a/c-Cu-H2, indicating the a/c-Cu has stronger ability to drive the H2O dissociation and generate abundant *H species60. After adding NO3−, the DMPO-H intensity of both the a/c-Cu and a/c-Cu-H2 catalyst decrease significantly due to the *H consumption by NO3− reduction. However, the stronger DMPO-H peaks over the a/c-Cu catalyst indicate the more favorable *H generation compared with the a/c-Cu-H2 catalyst61, in agreement with the Raman results. As hydrogenation steps are significantly affected by proton transfer rate, the kinetic isotope effect (KIE) of H/D is determined. When D2O is used instead of H2O in the electrolyte, the current density decreases (Supplementary Fig. 37). The KIE over the a/c-Cu catalyst is ~1.2, much lower than that over the a/c-Cu-H2 catalyst (Fig. 3e), further highlighting the facilitated water disassociation34.
Therefore, we have experimentally shown that the dual-phase a/c-Cu catalyst with rich amorphous Cu domains also plays dual roles in NO3−RR to NH3, namely, improving the adsorption of N-containing intermediates and facilitating the formation of active hydrogen species from water dissociation.
Density functional theory (DFT) calculations
The dual roles of the amorphous Cu domains present in the a/c-Cu catalyst were further validated by DFT calculations. As Cu(100) was the main facet in the a/c-Cu catalyst (Supplementary Fig. 28), the amorphous/crystalline dual-phase Cu structure was modelled by ab initio molecular dynamics (AIMD) simulations using a slab with four Cu(100) layers, of which the top two layers were relaxed and the bottom two layers were fixed62. For the AIMD simulations, a temperature of 800 K was employed to accelerate structural sampling, facilitating the generation of an amorphous Cu surface. The free energy profiles of NO3−RR over Cu(100) and the amorphous phase of the a/c-Cu model were first studied, as shown in Fig. 4a and Supplementary Figs. 38 and 39. Firstly, *NO3 adsorbs on the Cu(100) and a/c-Cu surfaces with energies of −0.08 and −0.01 eV, respectively. On the Cu(100) surface, the subsequent *NO3 hydrogenation to *NO3H undergoes an uphill with an energy of 0.10 eV and the potential-determining step (PDS) is *NO hydrogenation to *NHO with an energy of 0.35 eV (Fig. 4a). In contrast, the a/c-Cu surface significantly improves *NO3 hydrogenation with a distinct downhill trend. In this case, the PDS shifts to the formation of *NO2 from *NO3H with a lower energy of 0.27 eV.
Fig. 4. DFT calculations.
a Reaction pathways of NO3−RR on Cu(100) and a/c-Cu. The inserts are atomic models for Cu(100) and a/c-Cu. b Projected density of states of *NO3 adsorbed on Cu(100) and a/c-Cu. The asterisks represent the sum of other d-orbitals. c Charge density difference between *NO and Cu(100) (c1 and c2), *NO and a/c-Cu (c3 and c4). Isosurface = 2 × 10−3 e/Å3. d Reaction pathway for water dissociation on Cu(100) and a/c-Cu. Source data are provided as a Source Data file.
As the adsorption and activation of NO3− is the initial reaction step63,64, we further estimated the projected density of states (PDOS) of *NO3 (Fig. 4b). For *NO3 adsorbed on Cu(100), the py orbital of O and the dx2-y2 of Cu form a hybrid orbital in the range from −3.89 to −5.21 eV below the Fermi level (The detailed splitting d-orbitals of Cu and p-orbitals of O are displayed in Supplementary Fig. 40). For a/c-Cu, the px+py orbital of O and the dx2-y2 + dxz of Cu form a hybrid orbital in the range from −3.30 to −3.65 eV below the Fermi level (Supplementary Fig. 41). The hybridization region on a/c-Cu is closer to the Fermi level than that on Cu(100), which rationalizes the more favorable *NO3 hydrogenation. The *NO hydrogenation process has been generally considered as the PDS in an alkaline environment63,65,66, but in this work the PDS shifts from *NO hydrogenation on Cu(100) to the formation of *NO2 on a/c-Cu. Thus, we estimated the charge density difference for the *NO intermediate. As shown in Fig. 4c, there is obvious charge transfer between *NO and Cu atoms on both Cu(100) (Fig. 4c1, c2) and a/c-Cu (Fig. 4c3, c4). However, the charge density is substantially discrete for *NO adsorbed on a/c-Cu. This is largely related to the disordered arrangement of the amorphous phase of a/c-Cu, which leads to drastic d-orbital electron splitting and facilitates electron transfer to *NO. Thus, the *NO hydrogenation to *NHO is remarkably improved and becomes an exothermic process on a/c-Cu.
The *H formation from water dissociation was further investigated on Cu(100) and a/c-Cu. As shown in Fig. 4d and Supplementary Fig. 42, the dissociation of H–OH on Cu(100) requires an energy input of 0.34 eV, much more difficult than the downhill process on a/c-Cu, consistent with the experimentally KIE results (Fig. 3e)38. Overall, the above theoretical investigations further demonstrate that the amorphous Cu structure present in the amorphous/crystalline dual-phase Cu catalyst promotes NO3−RR to NH3 by optimizing the adsorption of N-containing intermediates and facilitating water dissociation.
Discussion
In summary, we have developed an a/c-Cu catalyst with stable amorphous/crystalline dual-phase structure, by annealing a commercially available Cu foam. The a/c-Cu catalyst exhibits impressive NO3−RR performance in an alkaline MEA electrolyzer, with a NH3 partial current density of 3.33 ± 0.005 A cm−2 and a NH3 formation rate of 15.5 ± 0.02 mmol h−1 cm−2 at a low cell voltage of 2.6 ± 0.01 V. Moreover, the NH3 Faradaic efficiency maintains around 90% at an applied current density of 1.5 A cm−2 for 300 h. The scale-up demonstration with an electrode size of 100 cm2 achieves a maximum NH3 formation rate up to 11.9 ± 0.5 g h−1 and an energy consumption of 108.3 ± 0.47 kJ g−1NH3 at an applied current of 160 A. The high NH3 formation rate is ascribed to the amorphous Cu domains present in the dual-phase a/c-Cu catalyst, which promote NO3−RR by optimizing the adsorption of N-containing intermediates and facilitating the formation of active hydrogen species from water dissociation. This work highlights the importance of stabilizing metastable amorphous structures for improving electrocatalytic reactivity and long-term stability.
Methods
Chemicals and materials
Cu foams (>99.9%) were purchased from Keshenghe Suzhou. Copper(II) chloride dihydrate (CuCl2·2H2O, >99%) was purchased from Tianjin Kemiou Chemical Reagent Co., Ltd. Sodium hydroxide (NaOH, >96%), potassium nitrate (KNO3, >99%), ammonium chloride (NH4Cl, >99.8%), trisodium citrate dihydrate (>99%), phosphoric acid (H3PO4, >85%), hydrochloric acid (HCl, >36%) and ethylene glycol (>99%) were purchased from Sinopharm Chemical Reagent Co., Ltd. Potassium hydroxide (KOH, >95%), Nafion ionomer (5%), salicylic acid (>99.5%), sulfonamide (>99.8%), lead(II) perchlorate hydrate (Pb(ClO4)2·3H2O, >95%) and tannic acid (>99%) were purchased from Macklin. 5,5-dimethyl-1-pyrroline-N-oxide (DMPO, >97%) and potassium nitrite (KNO2, >97%) were purchased from Aladdin. N-(1-naphthyl) ethylenediamine dihydrochloride (>98%) was purchased from Sigma-Aldrich (Shanghai) Trading Co., Ltd. Sodium nitroferricyanide (C5FeN6Na2O, >99%) was purchased from Shanghai Chemical Reagent Co. Ir black catalyst was purchased from Johnson Matthey Corp. Ultrapure water (18.2 MΩ) was used in all experiments. All the chemicals were used without further purification.
Catalyst preparation
A commercial Cu foam with a geometric area of 1 cm2 and a thickness of 0.2 mm (roughly 30 mg) was firstly annealed in air at 600 °C for 3 h. Subsequently, the annealed Cu foam was in situ electrochemically reduced during NO3−RR to form the a/c-Cu catalyst. The a/c-Cu-H2 catalyst was prepared by treating the above air-annealed Cu foam in 5% H2/Ar (50 mL min−1) at 350 °C for 3 h. Control samples (a/c-Cu-300 and a/c-Cu-900) were also prepared with the same method, but the annealing temperatures were 300 and 900 °C, respectively. An amorphous Cu catalyst was prepared with a method from previous literature67. A homogenous solution was formed by dissolving 0.2 g of CuCl2·2H2O in 40 mL of ethylene glycol and stirring for 30 min. After adding 160 mg of tannic acid, the mixture was stirred for another 30 min. Then, 2 mL of 1.0 M NaOH was gradually added the above solution, followed by stirring for another 10 min. The amorphous Cu catalyst was washed three times with water and acetone, respectively, and then dried in vacuum overnight.
Materials characterizations
XRD patterns were obtained using a PANalytical X’pert PPR diffractometer equipped with a Cu Kα radiation source (λ = 1.5418 Å). SEM images were captured using a JSM-7900F Field-Emission SEM. HRTEM images were obtained using a JEM-2100 with an accelerating voltage of 200 kV. HAADF-STEM images were acquired with a JEM-ARM200F (JEOL, Japan). The ratio of amorphous domains was quantified from HRTEM images using Gatan Microscopy Suite (GMS-3) software and ImageJ software, and was defined as the total amorphous area divided by the overall catalyst area in each image. Error bars were made based on ten independent measurements.
Operando XAS measurements were carried out at the BL11B, BL14W1, and BL20U beamlines of the Shanghai Synchrotron Radiation Facility (SSRF). Energy calibration was performed using the absorption edge of a Cu foil as a reference. A home-made flow cell was used for XAS measurements in fluorescence mode (Supplementary Fig. 19), with a catalyst-coated gas diffusion electrode as working electrode. The anolyte and catholyte were 1 M KOH solution (5 mL min−1) and 1 M KOH + 0.2 M KNO3 solution (5 mL min−1), respectively. The XAS data was analyzed using the ATHENA and ARTEMIS software.
In situ Raman spectroscopy measurements were carried out using a Renishaw inVia Raman microscope with a 785 nm near-infrared laser (Supplementary Fig. 34). The laser power was set to 1% and the exposure time was 50 s. A home-made flow cell with three electrodes was used for the Raman characterization. The cell was equipped with catalyst-coated gas diffusion electrode as the working electrode, a Pt wire as counter electrode, and an Ag/AgCl reference electrode. The cell had a quartz optical window, approximately 6 mm away from the working electrode, and the space in between was filled with catholyte. The anolyte and catholyte were 1 M KOH solution (5 mL min−1) and 1 M KOH + 0.2 M KNO3 solution (5 mL min−1), respectively. The spectra were collected at different potentials after electrolysis for at least 2 min.
In situ ATR-SEIRAS measurement was conducted using an INVENIO S FTIR spectrometer with MCT detector (Supplementary Fig. 35). A home-made H cell was used for ATR-SEIRAS measurement, with Pt wire as a counter electrode and Ag/AgCl electrode as a reference electrode. For working electrode, an Au film was sputtered onto a silicon prism, followed by drop-casting a catalyst ink onto the Au film. All spectra were presented as relative change in absorbance, referenced to background spectra collected at OCP.
Quasi in situ XPS and Auger electron spectroscopy (AES) measurements were conducted using a Thermo Scientific ESCALAB 250Xi spectrometer with an Al Kα X-ray source. The NO3−RR measurements were carried out in a MEA electrolyzer inside a glovebox, maintaining an O2 concentration below 0.01 ppm. After electrolysis for at least 20 min, the electrodes were transferred into the XPS analysis chamber using a mobile transfer chamber, without air exposure during the transfer process.
EPR spectroscopy measurements were conducted using a Bruker A200 spectrometer. A DMPO solution (10 mg mL−1) as a radical trapping reagent was added to the catholyte. After NO3−RR in a MEA electrolyzer, the catholyte was collected, rapidly frozen in liquid nitrogen to prevent degradation, and then thawed for EPR analysis.
DSC measurements was carried out using NETZSCH DSC300 Select with a heat rate of 5 °C min−1 under N2 atmosphere. Both the a/c-Cu and a/c-Cu-H2 powder catalysts were used for measurements.
The hydroxide (OH–) electrosorption experiments were performed through cyclic voltammetry (CV) measurements in an H-cell, using Ar-saturated 1 M KOH aqueous solution. The potential window was set from –0.8 to –0.523 V vs. Ag/AgCl, with a scan rate of 10 mV s–1. Before CV measurements, the electrode was firstly pre-reduced under NO3−RR conditions by applying a potential of –2.0 V vs. Ag/AgCl for 15 min.
The Pb UPD measurements for characterizing low-coordinated Cu sites were referred to in the literature41. Following NO3−RR at an applied potential of –2.0 V vs. Ag/AgCl for 15 min, the catalysts were measured by CV in a solution containing 0.1 M NaClO4, 10 mM HClO4, and 3 mM Pb(ClO4)2. The potential range was set from –0.4 to –0.1 V vs. Ag/AgCl, with a scan rate of 10 mV s–1. The characteristic peak at –0.15 V vs. Ag/AgCl was identified as low-coordinated Cu sites.
The H/D KIE experiments were conducted using the linear sweep voltammetry (LSV) in H-cell at a scan rate of 10 mV s–1, with iR correction. The electrode was firstly pre-reduced under NO3−RR conditions at an applied potential of –2.0 V vs. Ag/AgCl for 15 min. Subsequently, LSV measurements was carried out in the electrolyte, which consisted of 0.2 M KNO3 and 1.0 M KOH, with either H2O or D2O as the solvent. The KIE values were calculated based on the ratio of current density measured in H2O to that measured in D2O.
NO3−RR measurements
NO3−RR performance was assessed using an alkaline MEA electrolyzer at ambient temperature (20–25 °C)68. The electrolyzer was assembled using a graphite flow field plate for catholyte and Pt-coated titanium flow field plates for anolyte (Supplementary Fig. 2). A quaternary ammonia poly(N-methyl-piperidine-co-p-terphenyl) (QAPPT) membrane was used as an anion exchange membrane. The anode was prepared by coating an ink of QAPPT-impregnated Ir black catalyst (1.5 mg cm−2) onto a porous foam. The annealed Cu foam with a geometric area of 1 cm2 was directly used as an integrated porous electrode, if not stated otherwise. For comparison, the annealed foam was also ground into powder and drop-casted onto a carbon paper to form a conventional gas diffusion electrode. The NO3−RR measurements were performed in the galvanostatic mode using an Ivium electrochemical workstation. The anolyte and catholyte were 1 M KOH solution (5 mL min−1) and 1 M KOH + 0.2 M KNO3 solution (3 mL min−1), respectively. In the scale-up experiments, an annealed Cu foam with a geometric area of 100 cm2 was used and the catholyte was 1 M KOH + 0.5 M KNO3 (50 mL min−1). In all the performance measurements including 300-h stability test, all the electrolytes were not recirculated and fresh electrolytes were always used. The NO3−RR measurements were performed in the galvanostatic mode using a Keysight N8940A autoranging system direct current (DC) power supply (0–80 V/0–170 A, 5000 W). The electrolysis durations at each applied current were 15 and 10 min for 1-cm2 and 100-cm2 electrolyzers, respectively. Both catholyte and anolyte were collected at each applied current for subsequent liquid product analysis.
Products analysis
Gas products (i.e., H2 in this work) were quantified using an on-line gas chromatography (Agilent, GC 8860), equipped with a thermal conductivity detector (TCD). Liquid products (NO2− and NH3) were analyzed by UV–vis spectrophotometer (Shimadzu, Uv−2700i).
For NH3 quantification, the indophenol blue method was conducted. Specifically, 2 mL of 1 M NaOH solution containing 5.0 wt% salicylic acid and 5.0 wt% sodium citrate, followed by 1 mL of 0.05 M NaClO, and 0.2 mL of 1 wt% C5FeN6Na2O (sodium nitroferricyanide), were successively introduced into 2 mL of the diluted electrolyte, kept in the dark for 2 h. The NH3 concentration was determined based on the absorbance at 655 nm. Additionally, 1H-NMR spectroscopy was also used to quantify the NH3 (NH4+) production using a JEOL JNM-ECZL400S NMR spectrometer. In a typical procedure, 0.5 mL of standard solution/electrolyte was adjusted to pH 2 by adding 0.6 mL of 1.0 M HCl. Next, 0.5 mL of the above solution was mixed with the 0.1 mL DMSO-d6 (containing 2.5 mg mL–1 C4H4O4 as an internal standard). The quantification of NH4+ was calculated with the peak area ratio of NH4+ versus C4H4O4.
For NO2– quantification, the Griess method was employed. Specifically, after neutralizing 3 mL of the diluted electrolyte with 3 mL of 1.0 M HCl, 0.1 mL of Griess color reagent, containing 4 g of sulfonamide, 10 mL of H3PO4 (85%), and 0.2 g of N-(1-naphthyl) ethylenediamine dihydrochloride in 100 mL of water, was added. The mixture was then kept in the dark for 20 min. The NO2– concentration was determined based on the absorbance at 540 nm.
The Faradaic efficiency of product i is calculated as follows:
1 |
Where, εFaradaic,i: the Faradaic efficiency of product i, Qtotal: the consumed charge, C; Ni: the amount of the product i, mol; ni: the number of transferred charge in the product i; F: Faraday constant, 96485 C mol–1.
The NH3 partial current density is calculated as follows:
2 |
The full-cell energy efficiency (EE) of NH3 is defined as follows:
3 |
Where, EENH3: the full-cell energy efficiency (EE) of NH3, %; 1.23: the equilibrium potential of oxygen evolution reaction, V vs. RHE; E0NH3: the equilibrium potential of NO3– to NH3, which is 0.69 V vs. RHE19; U: cell voltage, V.
The energy consumption for NH3 production is defined as follows33:
4 |
Where, energy consumption: kJ g–1NH3; I: the applied current, A; CNH3: the molar concentration of NH3, mol L–1; v: the flow rate of the catholyte, L s–1; 17: the molecular mass of NH3, g mol–1.
Computational details
All simulations were carried out using the Vienna ab initio simulation package (VASP 6.2.0)69. Electron-ion interactions were described by the projector-augmented wave (PAW) pseudopotentials with a cutoff energy of 450 eV70. For electron exchange and correlation energies, the Perdew–Burke–Ernzerh of generalized gradient approximation (GGA-PBE) was represented71. DFT-D3 with Becke–Jonson (BJ) damping method was used to correct the long-range van der Waals interaction72. The exchange-correlation functional with a Gaussian smearing width term of 0.05 eV was used. The convergence criteria for force and electronic self-consistent iteration were set to 0.01 eV Å–1 and 1 × 10–5 eV, respectively. In addition, all calculations are spin-polarized. For all surface optimization, the sampling of the Brillouin zone was performed using a Monkhorst–Pack scheme of (2 × 2 × 1). For all DOS calculation, the sampling of the Brillouin zone was performed using a Monkhorst–Pack scheme of (5 × 5 × 1)73. The vacuum layer of 15 Å were selected to prevent their periodic images between adjacent layers. The VASPKIT code was used for postprocessing computational data obtained from VASP74.
For the adsorption of *NO3, the energy of HNO3 in the liquid phase is converted to *NO3 proton by the change of entropy as a correction factor, as shown below:
The adsorption energy of NO3− (ΔG*NO3) is approximately expressed as:
5 |
6 |
where G*NO3, G*, GHNO3, and GH2 are the Gibbs free energy of NO3 adsorbed on Cu substrates, Cu substrates, HNO3, and H2 molecules in the gas phase, respectively. According to CRC handbook of chemistry and physics75, ΔGS1 = − 0.075 eV and ΔGS2 = − 0.317 eV. Therefore, ΔGcorrect is set to 0.075 + 0.317 = 0.392 eV.
The Cu(100) model was constructed using a 4 × 4 × 1 supercell comprising four layers and a total of 112 Cu atoms. For the a/c-Cu model, a slab with four Cu(100) layers was generated, with the bottom two layers fixed and the top two layers relaxed during AIMD simulations62. A time step of 2 fs was used, initially under constant-pressure conditions as the temperature was ramped from 300 K to 800 K. Subsequently, a canonical ensemble (NVT) was employed with a Nose–Hoover thermostat set at 800 K to accelerate structural sampling and facilitate the formation of an amorphous Cu surface. The first 15 ps of AIMD served as the equilibration period, followed by a 5 ps production run for data collection and analysis.
Supplementary information
Description of Additional Supplementary Files
Source data
Acknowledgements
This work was supported by the National Key R&D Program of China (2023YFA1508000, G.W.), the National Natural Science Foundation of China (22372171, D.G.; 22125205, G.W.; 22321002, G.W.; 22494711, D.G.), the Fundamental Research Funds for the Central Universities (20720220008, G.W.), the Strategic Priority Research Program of the Chinese Academy of Sciences (XDB0600200, G.W.), the Liaoning Revitalization Talents Program (XLYC2203178, D.G.), the Liaoning Binhai Laboratory (LBLF-2023-02, G.W.; LBLD-2024-02, D.G.), the Dalian Outstanding Young Scientist Foundation (2024RJ003, D.G.), the Dalian Institute of Chemical Physics (DICP I202203, D.G.), the Joint Fund of the Yulin University and the Dalian National Laboratory for Clean Energy (YLU-DNL Fund 2022008, G.W.), the China Postdoctoral Science Foundation (2023M743428, S.W.; GZC20232594, S.W.), the Yanchang Petroleum Group (yc-hw-2023ky-08, D.G.), and the Photon Science Center for Carbon Neutrality (JZHKYPT-2021-07, D.G.). We thank Prof. Jin-Xun Liu at the USTC and Prof. Zhangquan Peng, Dr. Zhiwei Zhao and Dr. Long Pang at the DICP for fruitful discussions. We also thank the staff at the BL11B, BL14W1, and BL20U beamlines of the SSRF for their technical assistance during XAS measurements.
Author contributions
Conceptualization: G.W. and D.G.; methodology: Y.W., S.W., Y.F., J.S., P.W., R.L., and D.G.; investigation: Y.W. and S.W.; visualization: Y.W., S.W., and D.G.; funding acquisition: G.W., D.G., and S.W.; supervision: G.W. and D.G.; writing—original draft: Y.W., S.W., and D.G.; writing—review and editing: D.G., G.W., and X.B.
Peer review
Peer review information
Nature Communications thanks Dong-Hee Lim and the other, anonymous, reviewers for their contribution to the peer review of this work. A peer review file is available.
Data availability
The data that support the findings of this study are available within the paper and the Supplementary Information. Other relevant data are available from the corresponding authors on reasonable request. Source data are provided with this paper.
Competing interests
The authors declare no competing interests.
Footnotes
Publisher’s note Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.
These authors contributed equally: Yi Wang, Shuo Wang.
Contributor Information
Dunfeng Gao, Email: dfgao@dicp.ac.cn.
Guoxiong Wang, Email: wanggx@dicp.ac.cn.
Supplementary information
The online version contains supplementary material available at 10.1038/s41467-025-55889-9.
References
- 1.Chen, J. G. et al. Beyond fossil fuel-driven nitrogen transformations. Science360, eaar6611 (2018). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 2.Suryanto, B. H. et al. Challenges and prospects in the catalysis of electroreduction of nitrogen to ammonia. Nat. Catal.2, 290–296 (2019). [Google Scholar]
- 3.van Langevelde, P. H., Katsounaros, I. & Koper, M. T. Electrocatalytic nitrate reduction for sustainable ammonia production. Joule5, 290–294 (2021). [Google Scholar]
- 4.Xu, H., Ma, Y., Chen, J., Zhang, W. X. & Yang, J. Electrocatalytic reduction of nitrate - a step towards a sustainable nitrogen cycle. Chem. Soc. Rev.51, 2710–2758 (2022). [DOI] [PubMed] [Google Scholar]
- 5.Li, P. et al. Pulsed nitrate-to-ammonia electroreduction facilitated by tandem catalysis of nitrite intermediates. J. Am. Chem. Soc.145, 6471–6479 (2023). [DOI] [PubMed] [Google Scholar]
- 6.Li, J. et al. Efficient ammonia electrosynthesis from nitrate on strained ruthenium nanoclusters. J. Am. Chem. Soc.142, 7036–7046 (2020). [DOI] [PubMed] [Google Scholar]
- 7.Fan, K. et al. Active hydrogen boosts electrochemical nitrate reduction to ammonia. Nat. Commun.13, 7958 (2022). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 8.He, W. et al. Splicing the active phases of copper/cobalt-based catalysts achieves high-rate tandem electroreduction of nitrate to ammonia. Nat. Commun.13, 1129 (2022). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 9.Chen, F. Y. et al. 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]
- 10.Lim, J. et al. Structure sensitivity of Pd facets for enhanced electrochemical nitrate reduction to ammonia. ACS Catal.11, 7568–7577 (2021). [Google Scholar]
- 11.Yang, J., Sebastian, P., Duca, M., Hoogenboom, T. & Koper, M. T. pH dependence of the electroreduction of nitrate on Rh and Pt polycrystalline electrodes. Chem. Commun.50, 2148–2151 (2014). [DOI] [PubMed] [Google Scholar]
- 12.Zhou, J. et al. Regulating active hydrogen adsorbed on grain boundary defects of nano-nickel for boosting ammonia electrosynthesis from nitrate. Energy Environ. Sci.16, 2611–2620 (2023). [Google Scholar]
- 13.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.59, 5350–5354 (2020). [DOI] [PubMed] [Google Scholar]
- 14.Daiyan, R. et al. Nitrate reduction to ammonium: from CuO defect engineering to waste NOx-to-NH3 economic feasibility. Energy Environ. Sci.14, 3588–3598 (2021). [Google Scholar]
- 15.Zhang, G. et al. Tandem electrocatalytic nitrate reduction to ammonia on MBenes. Angew. Chem. Int. Ed.62, e202300054 (2023). [DOI] [PubMed] [Google Scholar]
- 16.Duan, W. et al. In situ reconstruction of metal oxide cathodes for ammonium generation from high-strength nitrate wastewater: Elucidating the role of the substrate in the performance of Co3O4-x. Environ. Sci. Technol.57, 3893–3904 (2023). [DOI] [PubMed] [Google Scholar]
- 17.Wu, Z. Y. et al. Electrochemical ammonia synthesis via nitrate reduction on Fe single atom catalyst. Nat. Commun.12, 2870 (2021). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 18.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]
- 19.Wang, Y. et al. 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]
- 20.Fang, J. Y. et al. Ampere-level current density ammonia electrochemical synthesis using CuCo nanosheets simulating nitrite reductase bifunctional nature. Nat. Commun.13, 7899 (2022). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 21.Liang, S., Teng, X., Xu, H., Chen, L. & Shi, J. H. Species regulation by Mn‐Co(OH)2 for efficient nitrate electro‐reduction in neutral solution. Angew. Chem. Int. Ed.63, e202400206 (2024). [DOI] [PubMed] [Google Scholar]
- 22.Gao, Q. et al. 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]
- 23.Shahid, U. B., Kwon, Y., Yuan, Y., Gu, S. & Shao, M. High‐performance ammonia electrosynthesis from nitrate in a NaOH‐KOH‐H2O ternary electrolyte. Angew. Chem. Int. Ed.63, e202403633 (2024). [DOI] [PubMed] [Google Scholar]
- 24.Reyter, D., Bélanger, D. & Roué, L. Study of the electroreduction of nitrate on copper in alkaline solution. Electrochim. Acta53, 5977–5984 (2008). [Google Scholar]
- 25.Wang, W., Chen, J. & Tse, E. C. Synergy between Cu and Co in a layered double hydroxide enables close to 100% nitrate-to-ammonia selectivity. J. Am. Chem. Soc.145, 26678–26687 (2023). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 26.Pei, Y. et al. Synthesis and catalysis of chemically reduced metal-metalloid amorphous alloys. Chem. Soc. Rev.41, 8140–8162 (2012). [DOI] [PubMed] [Google Scholar]
- 27.Chen, C. et al. Oxidation of metallic Cu by supercritical CO2 and control synthesis of amorphous nano-metal catalysts for CO2 electroreduction. Nat. Commun.14, 1092 (2023). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 28.Bauers, S. R. et al. Structural evolution of iron antimonides from amorphous precursors to crystalline products studied by total scattering techniques. J. Am. Chem. Soc.137, 9652–9658 (2015). [DOI] [PubMed] [Google Scholar]
- 29.Shen, Z. et al. Highly distributed amorphous copper catalyst for efficient ammonia electrosynthesis from nitrate. J. Hazard. Mater.445, 130651 (2023). [DOI] [PubMed] [Google Scholar]
- 30.Zhai, Y. et al. Phase engineering of metal nanocatalysts for electrochemical CO2 reduction. eScience2, 467–485 (2022). [Google Scholar]
- 31.Yang, Y. et al. Hierarchical nanoassembly of MoS2/Co9S8/Ni3S2/Ni as a highly efficient electrocatalyst for overall water splitting in a wide pH range. J. Am. Chem. Soc.141, 10417–10430 (2019). [DOI] [PubMed] [Google Scholar]
- 32.Zheng, W., Liu, M. & Lee, L. Y. S. Best practices in using foam-type electrodes for electrocatalytic performance benchmark. ACS Energy Lett.5, 3260–3264 (2020). [Google Scholar]
- 33.Xu, Z. et al. Continuous ammonia electrosynthesis using physically interlocked bipolar membrane at 1000 mA cm−2. Nat. Commun.14, 1619 (2023). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 34.Hu, Q. et al. Ammonia electrosynthesis from nitrate using a ruthenium-copper cocatalyst system: a full concentration range study. J. Am. Chem. Soc.146, 668–676 (2023). [DOI] [PubMed] [Google Scholar]
- 35.Gan, L. et al. Redirecting surface reconstruction of CoP-Cu heterojunction to promote ammonia synthesis at industrial-level current density. Chem. Eng. J.487, 150429 (2024). [Google Scholar]
- 36.Guo, X. et al. Highly stable perovskite oxides for electrocatalytic acidic NOx− reduction streamlining ammonia synthesis from air. Angew. Chem. Int. Ed.63, e202410517 (2024). [DOI] [PubMed] [Google Scholar]
- 37.Zhang, J. et al. Steering CO2 electroreduction pathway toward ethanol via surface-bounded hydroxyl species-induced noncovalent interaction. Proc. Natl Acad. Sci. USA 120, e2218987120 (2023). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 38.Zhang, B. et al. Defect-induced triple synergistic modulation in copper for superior electrochemical ammonia production across broad nitrate concentrations. Nat. Commun.15, 2816 (2024). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 39.Sang, J. et al. A Reconstructed Cu2P2O7 catalyst for selective CO2 electroreduction to multicarbon products. Angew. Chem. Int. Ed.61, e202114238 (2022). [DOI] [PubMed] [Google Scholar]
- 40.Sheng, H. W., Wilde, G. & Ma, E. The competing crystalline and amorphous solid solutions in the Ag-Cu system. Acta Mater.50, 475–488 (2002). [Google Scholar]
- 41.Yang, Y. et al. Operando studies reveal active Cu nanograins for CO2 electroreduction. Nature614, 262–269 (2023). [DOI] [PubMed] [Google Scholar]
- 42.Zhou, Y. et al. Electronegativity-induced charge balancing to boost stability and activity of amorphous electrocatalysts. Adv. Mater.34, 2100537 (2022). [DOI] [PubMed] [Google Scholar]
- 43.Jeon, S. et al. Reversible disorder-order transitions in atomic crystal nucleation. Science371, 498–503 (2021). [DOI] [PubMed] [Google Scholar]
- 44.Loh, N. D. et al. Multistep nucleation of nanocrystals in aqueous solution. Nat. Chem.9, 77–82 (2016). [DOI] [PubMed] [Google Scholar]
- 45.Jia, B. et al. Construction of amorphous/crystalline heterointerfaces for enhanced electrochemical processes. eScience3, 100112 (2023). [Google Scholar]
- 46.Zhou, Z., Huang, G. G., Kato, T. & Ozaki, Y. Experimental parameters for the SERS of nitrate ion for label‐free semi‐quantitative detection of proteins and mechanism for proteins to form SERS hot sites: a SERS study. J. Raman Spectrosc.42, 1713–1721 (2021). [Google Scholar]
- 47.Bai, L. et al. Electrocatalytic nitrate and nitrite reduction toward ammonia using Cu2O nanocubes: active species and reaction mechanisms. J. Am. Chem. Soc.146, 9665–9678 (2024). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 48.Butcher, D. P. Jr & Gewirth, A. A. Nitrate reduction pathways on Cu single crystal surfaces: Effect of oxide and Cl−. Nano Energy29, 457–465 (2016). [Google Scholar]
- 49.Wang, Y. et al. Wide-pH-range adaptable ammonia electrosynthesis from nitrate on Cu-Pd interfaces. Sci. China Chem.66, 913–922 (2023). [Google Scholar]
- 50.Torreggiani, A., Esposti, A. D., Tamba, M., Marconi, G. & Fini, G. Experimental and theoretical Raman investigation on interactions of Cu(II) with histamine. J. Raman Spectrosc.37, 291–298 (2006). [Google Scholar]
- 51.Hu, Q. et al. 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]
- 52.Huang, K. et al. Boosting nitrate to ammonia via the optimization of key intermediate processes by low-coordinated Cu–Cu Sites. Adv. Funct. Mater. 34, 2315324 (2024).
- 53.Pérez-Gallent, E., Figueiredo, M. C., Katsounaros, I. & Koper, M. T. Electrocatalytic reduction of nitrate on copper single crystals in acidic and alkaline solutions. Electrochim. Acta227, 77–84 (2017). [Google Scholar]
- 54.Yao, Y., Zhu, S., Wang, H., Li, H. & Shao, M. A spectroscopic study of electrochemical nitrogen and nitrate reduction on rhodium surfaces. Angew. Chem. Int. Ed.59, 10479–10483 (2020). [DOI] [PubMed] [Google Scholar]
- 55.Xu, M. et al. Kinetically matched C–N coupling toward efficient urea electrosynthesis enabled on copper single-atom alloy. Nat. Commun.14, 6994 (2023). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 56.Shao, F. et al. Surface water as an initial proton source for the electrochemical CO reduction reaction on copper surfaces. Angew. Chem. Int. Ed.62, e202214210 (2022). [DOI] [PubMed] [Google Scholar]
- 57.Jiao, J. et al. Steering the reaction pathway of CO2 electroreduction by tuning the coordination mumber of copper catalysts. J. Am. Chem. Soc.146, 15917–15925 (2024). [DOI] [PubMed] [Google Scholar]
- 58.Wei, D. et al. Surface adsorbed hydroxyl: a double-edged sword in electrochemical CO2 reduction over oxide-derived copper. Angew. Chem. Int. Ed.62, e202306876 (2023). [DOI] [PubMed] [Google Scholar]
- 59.Zhang, T. et al. Selective increase in CO2 electroreduction to ethanol activity at nanograin-boundary-rich mixed Cu(I)/Cu(0) sites via enriching co-adsorbed CO and hydroxyl species Angew. Chem. Int. Ed.63, e202407748 (2024). [DOI] [PubMed] [Google Scholar]
- 60.Li, X., Shen, P., Li, X., Ma, D. & Chu, K. Sub-nm RuOx clusters on Pd metallene for synergistically enhanced nitrate electroreduction to ammonia. ACS Nano17, 1081–1090 (2023). [DOI] [PubMed] [Google Scholar]
- 61.Wang, Y. et al. Phase-regulated active hydrogen behavior on molybdenum disulfide for electrochemical nitrate-to-ammonia conversion. Angew. Chem. Int. Ed.63, e202315109 (2024). [DOI] [PubMed] [Google Scholar]
- 62.Joutsuka, T. & Tada, S. Adsorption of CO2 on amorphous and crystalline zirconia: a DFT and experimental study. J. Phys. Chem. C.127, 6998–7008 (2023). [Google Scholar]
- 63.Wang, S. et al. Non-noble single-atom alloy for electrocatalytic nitrate reduction using hierarchical high-throughput screening. Nano Energy113, 108543 (2023). [Google Scholar]
- 64.Wang, S., Wang, Y., Fu, Y., Liu, T. & Wang, G. High-throughput mechanistic study of highly selective hydrogen-bonded organic frameworks for electrochemical nitrate reduction to ammonia. J. Energy Chem.87, 408–415 (2023). [Google Scholar]
- 65.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]
- 66.Shu, Z. et al. High-throughput screening of heterogeneous transition metal dual-atom catalysts by synergistic effect for nitrate reduction to ammonia. Adv. Funct. Mater.33, 2301493 (2023). [Google Scholar]
- 67.Yang, P.-P. et al. Highly enhanced chloride adsorption mediates efficient neutral CO2 electroreduction over a dual-phase copper catalyst. J. Am. Chem. Soc.145, 8714–8725 (2024). [DOI] [PubMed] [Google Scholar]
- 68.Wei, P. et al. Coverage-driven selectivity switch from ethylene to acetate in high-rate CO2/CO electrolysis. Nat. Nanotechnol.18, 299–306 (2023). [DOI] [PubMed] [Google Scholar]
- 69.Kresse, G. & Furthmüller, J. Efficient iterative schemes for ab initio total-energy calculations using a plane-wave basis set. Phys. Rev. B54, 11169–11186 (1996). [DOI] [PubMed] [Google Scholar]
- 70.Blöchl, P. E. Projector augmented-wave method. Phys. Rev. B50, 17953–17979 (1994). [DOI] [PubMed] [Google Scholar]
- 71.Perdew, J. P., Burke, K. & Ernzerhof, M. Generalized gradient approximation made simple. Phys. Rev. Lett.78, 1396–1396 (1997). [DOI] [PubMed] [Google Scholar]
- 72.Grimme, S., Ehrlich, S. & Goerigk, L. Effect of the damping function in dispersion corrected density functional theory. J. Comput. Chem.32, 1456–1465 (2011). [DOI] [PubMed] [Google Scholar]
- 73.Pack, J. D. & Monkhorst, H. J. “Special points for Brillouin-zone integrations”—a reply. Phys. Rev. B16, 1748–1749 (1977). [Google Scholar]
- 74.Yang, J. et al. Dynamic behavior of single-atom catalysts in electrocatalysis: Identification of Cu-N3 as an active site for the oxygen reduction reaction. J. Am. Chem. Soc.143, 14530–14539 (2021). [DOI] [PubMed] [Google Scholar]
- 75.Weast, R. C. CRC Handbook of Chemistry and Physics, 71st edn. (CRC Press, 1988).
Associated Data
This section collects any data citations, data availability statements, or supplementary materials included in this article.
Supplementary Materials
Description of Additional Supplementary Files
Data Availability Statement
The data that support the findings of this study are available within the paper and the Supplementary Information. Other relevant data are available from the corresponding authors on reasonable request. Source data are provided with this paper.