Next Article in Journal
Effect of Calcination Temperature in Large-Aperture Medium-Entropy Oxide (FeCoCuZnNa)O on CO2 Hydrogenation for Light Olefins
Previous Article in Journal
Computational Studies of Molybdenum-Containing Metal–Sulfur and Metal–Hydride Clusters
Previous Article in Special Issue
Catalytic Biolubricant Production from Canola Oil Through Double Transesterification with Methanol and Neopentyl Glycol
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Co-Carbonization of Straw and ZIF-67 to the Co/Biomass Carbon for Electrocatalytic Nitrate Reduction

Institute for Energy Research, School of Chemistry and Chemical Engineering, Jiangsu University, Zhenjiang 212013, China
*
Authors to whom correspondence should be addressed.
Catalysts 2024, 14(11), 817; https://doi.org/10.3390/catal14110817
Submission received: 26 October 2024 / Revised: 8 November 2024 / Accepted: 11 November 2024 / Published: 13 November 2024
(This article belongs to the Collection Catalytic Conversion of Biomass to Bioenergy)

Abstract

:
Electrocatalytic nitrate reduction enables the recovery of nitrate from water under mild conditions and generates ammonia for nitrogen fertilizer feedstock in an economical and green means. In this paper, Co/biomass carbon (Co/BC) composite catalysts were prepared by co-carbonization of straw and metal–organic framework material ZIF-67 for electrocatalytic nitrate reduction using hydrothermal and annealing methods. The metal–organic framework structure disperses the catalyst components well and provides a wider specific surface area, which is conducive to the adsorption of nitrate and the provision of more reactive active sites. The introduction of biomass carbon additionally enhances the electrical conductivity of the catalyst and facilitates electron transport. After electrochemical testing, Co/BC-100 exhibited the best performance in electrocatalytic nitrate reduction to ammonia, with an ammonia yield of 3588.92 mmol gcat.−1 h−1 and faradaic efficiency of 97.01% at −0.5 V vs. RHE potential. This study provides a promising approach for the construction of other efficient cobalt-based electrocatalysts.

1. Introduction

The development of urbanization, industrialization, and agricultural modernization has led to severe nitrate (NO3) pollution in water bodies and eutrophication, posing threats to human health, particularly that of infants [1,2,3,4,5,6]. The United States Environmental Protection Agency (EPA) and the World Health Organization (WHO) have established that the maximum concentration of NO3 in drinking water should be below 10.0 mg L−1 [7,8,9]. However, the concentration of NO3 in industrial and agricultural wastewater seriously exceeds this threshold. NO3 is a non-ligand oxygen anion with high solubility and mobility in water [10,11,12,13]. While NO3 naturally exists in geological structures, its pollution is primarily caused by human activities such as agricultural runoff and industrial wastewater discharge [14,15,16,17,18]. With the growing awareness of environmental protection, tackling NO3 pollution has emerged as a major challenge for humanity [19,20,21]. Currently, the main methods used for NO3 removal include biochemical and physical–chemical methods such as reverse osmosis, ion-exchange resins, and hydrogen reduction. These traditional processes not only require sufficient organic carbon sources but also produce carbon dioxide (CO2) [22,23,24]. Due to the recently increasing emphasis on carbon neutrality and peak carbon emissions, there is an urgent need for a strategy that is more compatible with the green concepts of energy saving, emission reduction, and resource recycling simultaneously [25,26].
Electrochemical reduction is a wastewater treatment method that combines electrochemistry and catalytic chemistry [27,28,29]. Electrochemical NO3 reduction reaction (NO3RR) has garnered significant attention due to its controllable reaction rates and selectivity without the need for chemical additives, along with low costs. Self-degradation is achieved by electrochemical reduction of NO3 to ammonia (NH3) in the cathode chamber. The reduction of NO3 to NH3 is considerably easier than the reduction of nitrogen (N2) to NH3 since the bond energy of the N≡N bond (941 KJ mol−1) is much higher than that of the N=O bond (204 KJ mol−1), and nitrogen (0.02 g L−1) is less soluble than nitrate (880 g L−1) in water [30,31,32,33]. Therefore, the activation of N2 molecules during electrocatalytic nitrogen reduction requires a large overpotential and is accompanied by complex gas–liquid–solid interfacial interactions [34,35]. Additionally, NH3 serves as a nitrogen source for the production of ammonium fertilizers at a relatively low cost [36,37]. Current industrial-scale production is still dominated by the Haber-Bosch method, with high energy consumption and harsh process conditions, and the large amount of CO2 emitted during production exacerbates environmental problems [14,38]. Therefore, electrochemical production will become a competitive production method due to its low cost, energy saving, and high efficiency.
However, the electrochemical reduction of NO3 to NH3 involves a complex process of eight electron transfer reactions and exhibits slow reaction kinetics. Therefore, the development of efficient and selective catalysts is essential for enhancing the nitrate reduction reaction’s performance. Noble metal catalysts exhibit high activity in NO3RR, but their prohibitive cost limits their large-scale application [39]. The transition metals have attracted much attention for their low cost and variety among the non-precious metals [40,41]. Recent studies have reported that cobalt (Co) has lower adsorption energy for NO3 than other transition metals, such as manganese (Mn), copper (Cu), and iron (Fe), suggesting that Co is more favorable for NO3 adsorption. Niu et al. developed a Cu-Co3O4 catalyst for NO3RR, but its poor conductivity prevented it from achieving both high selectivity and high yield [42]. Li et al. developed a single-atom catalyst Co-CNP, but its limited active sites resulted in low yield at low concentrations [43]. Kuznetsova et al. designed a bimetallic electrocatalyst based on iron and cobalt nanoparticles, but both yield and selectivity need to be improved [44]. However, the recently reported CoNiO2 and Co3O4-Cu2+1O/CF electrocatalysts both exhibit high catalytic activity and selectivity, which are meaningful for subsequent catalyst designs [45,46].
In this work, we synthesized a Co/biomass carbon (Co/BC) composite catalyst with favorable electrical conductivity and abundant active sites through hydrothermal and annealing using straw and ZIF-67 as precursors. Biomass carbon exhibits promising electrical conductivity, while ZIF features a stable three-dimensional structure that provides abundant metal active sites [47,48,49,50], making it an excellent catalyst for electrochemical NO3RR under mild conditions. The test results demonstrated that Co/BC-100 achieves the highest electrochemical NO3RR performance, with a yield up to 3588.9 mmol gcat.−1 h−1 at −0.5 V vs. RHE and a corresponding faradaic efficiency of 97.01%. The faradaic efficiency could be maintained above 88.02% after five cycles of stability tests, exhibiting good stability.

2. Results and Discussion

Co/BC composites were prepared as shown in Figure 1. Co/BC composites were synthesized by annealing at 800 °C for 1 h in a nitrogen atmosphere, with the addition of hydrothermal carbon at concentrations of 0 mg, 50 mg, and 200 mg. The morphology, composition, and structure of the catalyst can be characterized by transmission electron microscopy (TEM) and high-resolution transmission electron microscopy (HRTEM) images. The TEM image of Co/BC-100 shows Co nanoparticles embedded in the carbon layer, and the lattice stripes are distinctly exhibited in the HRTEM image (Figure 2a–c). specifically, the lattice spacing of 0.46 nm corresponds to the (3 1 1) crystalline surface of Co. Additionally, the elemental mapping analysis also indicates the uniform distribution of Co, C, O, and N elements (Figure 2d–h).
To further investigate the crystal structures of the obtained BC, Co/BC-0, Co/BC-50, Co/BC-100, and Co/BC-200, the materials were characterized using X-ray powder diffraction (XRD). As shown in the XRD patterns from Figure 3a, three distinct diffraction peaks exist for Co/BC-0, Co/BC-50, Co/BC-100, and Co/BC-200. Specifically, the characteristic peaks at 44.2°, 51.5°, and 75.8° can be attributed to the (1 1 1), (2 2 2), and (2 2 0) crystal planes of Co, respectively. In addition, a broad peak at 22° can be observed for BC, Co/BC-50, Co/BC-100, and Co/BC-200, which is the (0 0 2) crystallographic plane of disordered carbon. Subsequently, Raman spectroscopic characterization of Co/BC-100 was carried out to obtain information about the vibrational modes of the material and thus assist in determining the composition and structure of the material (Figure 3b). Two characteristic peaks (D-band and G-band) are visible at 1350 cm−1 and 1600 cm−1 and are attributed to the sp3 and sp2 hybridization of carbon [51]. The broad peaks in this band from 600 to 900 cm−1 may be related to the D and G bands or related low-frequency modes (e.g., the D’ peak). The D’ peak is a low-frequency mode related to the D and G peaks, and it usually appears near the G peak. In the case of highly aggregated defects or at high frequencies, the D’ peak may combine with the intensity of the D and G peaks to affect the Raman spectrum, resulting in broad peaks in the range of 600–900 cm−1. In addition, the contact angle of Co/BC-100 was tested to examine its hydrophilicity. The red line represents the baseline of the liquid or the level of the solid surface, and the blue line shows the outline of the droplet, demonstrating the contact angle between the droplet and the solid surface. The contact angle of Co/BC-100 is 11.4°, as revealed in Figure 3c. The narrow contact angle may be attributed to the capillary phenomenon produced by the presence of numerous micropores on the BC, which results in a better affinity for water and this facilitates the adsorption of NO3 from the electrolyte by the catalyst.
The elemental composition and surface chemical properties of the obtained materials were studied using X-ray photoelectron spectroscopy (XPS). As shown in Figure S1, the XPS spectrum reveals the presence of four elements (Co, C, N, and O) in Co/BC-100. The high-resolution C 1s spectrum suggests that the peak at 284.8 eV is assigned to the C-C bond, and the characteristic peaks at 286 eV and 288.2 eV are attributed to C-O and C=O, respectively (Figure 4a). The elemental N content is relatively low and mainly originates from biomass carbon. In the N 1s spectrum (Figure 4b), the peak located at 399.9 eV belongs to pyridine nitrogen, and the peak at 401.2 eV corresponds to graphitic nitrogen. The high-resolution Co 2p spectrum demonstrates the presence of three species, Co0, Co2+, and Co3+, with Co2+ and Co3+ possibly arising from the oxidation of the catalyst surface (Figure 4c). Therein, Co0 is located at 779.7 eV, and Co2+ and Co3+ are located at 779 eV and 780 eV, respectively. Co 2p3/2 and Co 2p1/2 exhibit a separation of 14.99 eV, which is in accordance with the previous results [52,53]. Two characteristic peaks can be observed in the high-resolution O 1s spectra, with C=O located at 532.2 eV and Co-O at 529 eV (Figure 4d).
The Co/BC and BC samples were loaded on carbon paper as working electrodes separately to test their NO3RR performance using a three-electrode system. The cyclic voltammetry (CV) tests were performed on the working electrodes at a sweep rate of 50 mV s−1 at the potential ranging from −0.73 to 0.27 V vs. RHE before carrying out the electrocatalytic NO3RR aiming at complete activation of the working electrodes (Figure 5a). Subsequently, linear scanning voltammetry (LSV) tests were performed on Co/BC-0, Co/BC-50, Co/BC-100, and Co/BC-200 at the potential ranging from −0.73 to 0.27 V vs. RHE with a sweep rate of 10 mV s−1 to compare their current densities as well as onset potentials. As shown in Figure 5b, the starting potential of Co/BC catalysts is relatively lower and the current density of Co/BC-100 at the same potential is larger than that of Co/BC-0, Co/BC-50, and Co/BC-200, indicating that Co/BC-100 has a better response to electrocatalytic NO3RR.
Additionally, the activity, selectivity, and economics of the prepared catalysts can be further reflected by quantitative analysis of the products. The amperometric tests for nitrate reduction were firstly carried out at the five different potentials of −0.2, −0.3, −0.4, −0.5, and −0.6 V vs. RHE, and the duration of the tests was 1 h for each potential. The i-t curves at different potentials are shown in Figure 5c, where the current densities steadily increase with increasing potential. Subsequently, the electrolyte from the cathode cell was collected, and the concentration of NH3 in the electrolyte was determined using Nessler’s reagent spectrophotometric method. The UV absorption spectra are presented in Figure 5d, and the NH3 yield and faradaic efficiency (FE) can be calculated from the UV absorption value at 420 nm (Figure 5e). The FE reaches the maximum level (97.01%) at the potential of −0.5 V vs. RHE, and the NH3 yield is up to 3588.92 mmol gcat.−1 h−1, exceeding many of the values reported in the literature (Table S1). The NH3 yield reaches 4369.97 mmol gcat.−1 h−1 at the potential of −0.6 V vs. RHE, and the FE decreases to 94.78%, which is attributed to the competition of hydrogen evolution reaction at high potentials.
Five i-t tests were performed on Co/BC-100 (Figure 6a), and the corresponding UV absorption spectra are shown in Figure 6b. After five cyclic stability tests at a potential of −0.4 V vs. RHE, the NH3 yields could still be maintained above 3267.66 mmol gcat.−1 h−1, and the FEs remained above 87.9% (Figure 6c). Moreover, the NO2 yields and FEs arising from the five cyclic stability tests were also examined (Figure 6e), and their UV absorption spectra are shown in Figure 6d. The NO2 yields and FEs increase with the cycling times, but they still remain at a relatively low level, with little influence on the selectivity of NH3. After a long stability test of 12 h (Figure 6f), the current density was virtually unchanged from the beginning of the test, indicating the excellent stability of the Co/BC-100.
To compare the electrocatalytic performance of different catalysts in nitrate reduction reaction, the electrocatalytic NO3RR performance tests were conducted on Co/BC-0, Co/BC-50, and Co/BC-200 within the potential range of −0.6 to −0.2 V vs. RHE (Figures S2–S4). Co/BC-0 achieved the highest FE of 54.76% at −0.4 V vs. RHE, with a corresponding yield of 959.86 mmol gcat.−1 h−1 (Figure S5). Co/BC-50 reached the highest FE of 68.14% at −0.3 V vs. RHE, with a yield of 1104.29 mmol gcat.−1 h−1 (Figure S6). Co/BC-200 also achieved its highest FE of 83.79% at −0.3 V vs. RHE, with a yield of 1077.76 mmol gcat.−1 h−1 (Figure S7). At the same potential of −0.5 V vs. RHE, the yield of Co/BC-100 was 3.1, 3.4, and 2.3 times that of Co/BC-0, Co/BC-50, and Co/BC-200, and the FE was 2.05, 2.02, and 2.26 times that of Co/BC-0, Co/BC-50, and Co/BC-200, respectively. Therefore, excessive biomass carbon can cause the active sites of the catalyst to be covered, thus reducing the effective catalytic reaction area and lowering the reaction rate. In addition, the surface nature of the catalyst in contact with the reactants is an important factor affecting the catalytic efficiency. Excessive addition of biomass carbon may result in the formation of a more compact structure, limiting the diffusion and penetration of reactants, which in turn reduces the catalytic performance.
A series of control experiments was conducted (using empty carbon paper and electrolytes without NO3) to study the role of Co/BC-100 in the electrocatalytic reduction of NO3 (NO3RR). As shown in Figure 7a, the linear sweep voltammetry (LSV) curves were tested under the conditions of empty carbon paper and electrolytes without NO3, revealing that the initial current density was significantly lower than that measured with Co/BC-100. This indicates that the catalytic activity is quite low under these conditions. In Figure 7b, it is shown that the amount of NH3 produced by Co/BC-100 is 816.9 mmol in the electrolyte containing NO3, whereas nearly no NH3 was generated in the electrolyte without NO3. There was also almost no NH3 generated in the electrolyte after testing with empty carbon paper as the working electrode. The results demonstrate that the nitrogen source for the electrocatalytic NO3RR synthesis of NH3 is derived from KNO3 in the electrolyte rather than from the catalyst or atmospheric nitrogen and that Co/BC-100 contributes predominantly to the catalytic reaction. Subsequently, isotope labeling experiments were performed to confirm whether the nitrogen in the generated NH3 came from the electrolyte. K14NO3 was replaced with K15NO3 in the electrolyte, and the resulting electrolyte was collected for 1H NMR analysis. It can be observed that the 1H NMR spectrum of the electrolyte using K15NO3 as the reactant exhibits characteristic double peaks of 15N with a coupling constant J = 72 Hz, whereas the spectrum with K14NO3 shows characteristic triplets of 14N with a coupling constant J = 50 Hz (Figure 7c). The 1H NMR analysis clearly indicates that the nitrogen source in the produced NH3 derives from the applied nitrogen source in the electrolyte rather than from other contaminants [54].
To investigate the reasons behind the excellent performance of Co/BC-100 in electrocatalytic NO3RR reduction, the as-prepared catalysts were subjected to Capacitance of Double Layer (Cdl) testing. The electrochemically active surface area was studied through comparative analysis. As shown in Figure 8a–e, the CV curves of Co/BC-0, Co/BC-50, Co/BC-100, and Co/BC-200 were measured at different scan rates (20, 40, 60, 80, and 100 mV s−1). The calculated Cdl for Co/BC-100 is 18.6 mF cm−2, which is 2.62 times, 1.15 times, and 93 times that of Co/BC-0, Co/BC-50, and Co/BC-200, respectively. This indicates that Co/BC-100 possesses more reactive sites that facilitate the electrocatalytic NO3RR reaction. Furthermore, the electrochemical impedance spectroscopy (EIS) results show that Co/BC-200 exhibits low electrochemical impedance, suggesting good conductivity (Figure 8f). The interface effect between Co and BC enhances electron transfer, resulting in a faster electron transfer rate and better reaction kinetics, which are favorable for the electrocatalytic NO3RR process.
The intermediates were analyzed through in situ infrared technique to gain insight into the reaction process and mechanism of Co/BC-100 electrocatalytic reduction of nitrate to ammonia. The variations of reactants, products, and intermediates were monitored to understand the activity and selectivity of the catalyst. Surface-enhanced in situ FT-IR taken in 0.1 M KOH and 0.1 M KNO3 electrolytes from a potential ranging from 0 to −1 V vs. RHE are shown in Figure 9a. Positive and negative bands in the in situ FT-IR spectra represent the production and depletion of active intermediates, respectively. Specifically, the upward bands at 1140 cm−1 and 1550 cm−1 can be attributed to NH2OH and *NO species, respectively, and the gradual enhancement of the upward band intensities with decreasing potentials attests to the production of NH2OH and *NO intermediates. The broad and strong upward band at 1660 cm−1 corresponds to the dissociation of water, and the reactive hydrogen produced during the dissociation of hydrolysis can be involved in the stepwise hydrogenation of *NOx in solution. The upward band at 1460 cm−1 strengthens as the potential decreases, indicating the gradual production of NH4+, whereas the downward band at 1345 cm−1 proves the depletion of NO3. From the above results, it can be deduced that the nitrate reduction reaction follows the following path: *NO3 → *NO2 → *NO → *NOH → *NH2O → *NH2OH → *NH → *NH2 → NH3 (Figure 9b) [38,55,56,57].
Based on the above analysis, we speculate that the performance of catalysts in enhancing nitrate reduction to ammonia can be described as the coordinated action of each component in the catalyst. ZIF features a stable three-dimensional structure, which allows Co/BC-100 to display large specific surface area, which is conducive to the adsorption of nitrate and the provision of more reactive active sites. Biomass carbon exhibits excellent electrical conductivity, which promotes the rate of charge transfer in the material. The interfacial effect of Co and BC enhances the electron transfer and facilitates the conversion of nitrate to ammonia, ultimately achieving a highly active and selective nitrate reduction electrocatalyst.

3. Materials and Methods

3.1. Materials and Characterizations

All materials, such as 2-methylimidazole (C4H6N2), methanol (CH3OH), cobalt dinitrate hexahydrate (Co(NO3)2·6H2O), potassium hydroxide (KOH), potassium nitrate (KNO3), and ethanol (C2H6O), were analytical-grade reagents and purchased from Sigma Aldrich (St. Louis, MO, USA) and Sinopharm (Beijing, China). All reagents and chemicals were used as received without further purification.
The morphology and microstructure of the samples were observed in a scanning electron microscope (SEM, JEOF-JSM-7800F, Japan Electron Optical Laboratory, Tokyo, Japan) and a transmission electron microscope (TEM, FEI-Talos-F200s, Thermo Fisher Scientific, Waltham, MA, USA) equipped with an energy dispersive spectroscopy (EDS) detector. X-ray diffraction (XRD, D2 Phaser, Bruker, Ettlingen, Germany) data were obtained with a wide-angle diffractometer with Cu Kα radiation (λ = 1.5418 Å) at a generator voltage of 30 kV and current of 10 mA. X-ray photoelectron spectroscopy (XPS) analysis was collected on a Thermo Scientific K-Alpha+ X-ray photoelectron spectrometer (Thermo Fisher Scientific, Waltham, MA, USA) using Al Kα radiation. In situ ATR-FTIR spectra were obtained by Thermo Scientific Nicolet iS50 (Thermo Fisher Scientific, Waltham, MA, USA).

3.2. Synthesis of BC

Corn stover powder was added to 50 mL of deionized water, stirred well, and then transferred to an autoclave reactor to be hydrothermalized at 180 °C for 12 h. After drying, it was added to 1 M KOH to be ultrasonicated for 30 min and impregnated for 24 h. Subsequently, it was transferred to a quartz boat and calcined in a tube furnace under an argon atmosphere at 800 °C for 120 min and then dried to obtain hydrothermal carbon powder. The defect level is an important indicator of the quality of biomass carbon materials. In the lower temperature range, increasing the carbonization temperature can significantly increase the defects in the material. However, this increase does not always take place, and the material becomes saturated with defects at temperatures above 700 °C [58]. Therefore, the carbonization temperature of 800 °C that we chose is an appropriate temperature in combination with experience and practice.

3.3. Synthesis of Co/BC

Next, 209.03 mg of (Co(NO3)2·6H2O) was added to 40 mL of methanol and stirred for 5 min; then, 100 mg of hydrothermal charcoal was added, with continued stirring for 10 min. Subsequently, 0.6568 g of 2-methylimidazole was dissolved in 40 mL of methanol and stirred for 5 min and then slowly added dropwise to cobalt nitrate solution and stirred for 24 h. It was washed with methanol and deionized water, respectively, and then annealed at 800 °C under N2 atmosphere for 1 h to obtain Co/BC-100. Co/BC-0, Co/BC-50, and Co/BC-200 were prepared in the same way as described above, except that the masses of added hydrothermal carbon were 0 mg, 50 mg, and 200 mg, respectively.

3.4. In Situ FT-IR Testing

A layer of 100 nm Au film was deposited onto the reflecting plane of a Si prism by vacuum evaporation using a thermal evaporator (PuDi vacuum PD-400, Wuhan Puti Vacuum Technology Co., Wuhan, China). Before the Au deposition, the Si prism was polished with 0.05 μm Al2O3 suspension and cleaned by sonication sequentially in baths of acetone and deionized water. The working electrode was made by airbrushing the catalyst ink onto the above-prepared Au film. The in situ FT-IR measurements were conducted in a two-compartment spectro–electrochemical cell comprising three electrodes including the working electrode, a platinum-wire as the counter electrode, and a standard Ag/AgCl electrode as the reference. All the in situ FT-IR spectra were acquired using a Fourier Transform Infrared Spectrophotometer (FT-IR, Nicolet iS50, Thermo Fischer scientific, Waltham, MA, USA) equipped with a mercury cadmium telluride (MCT) detector. In a typical test, the working electrode was subjected to an initial activation by running CV cycles between 0 and −0.8 V vs. RHE at a scan rate of 0.05 V S−1 until the system was stabilized. The spectrum under open-circuit voltage was then collected as the background. The cathode potential was then swept from 0 V to −0.8 V vs. RHE, with each potential lasting 2 min for spectra acquisition. All measurements were performed at a spectral resolution of 4 cm−1 and presented in transmission units after subtracting the background.

3.5. Electrochemical Measurement

All the electrochemical measurements were carried out on a CHI 660E electrochemical workstation (CH Instrument, Shanghai, China) at room temperature (20 °C). The H-type electrochemical cell was used as a measuring device and Co/BC, Pt sheet (1 cm × 2 cm, 1 mm in thickness) and HgCl2/Hg (saturated KCl) as cathode, anode, and reference electrode, respectively. The anion exchange membrane (FAB-PK-130, FuMATech, Bietigheimer Schloss, Germany) was used to separate the cathode and anode of the H-cell. The working electrode area was controlled at 1 cm2, and 2.5 mg catalyst was dispersed into 500 μL isopropanol/water mixture (V:V = 3:1) containing 50 μL of 5% Nafion and kept ultrasonic for 1 h. Subsequently, 50 μL of the catalyst ink was loaded on the CP and dried at room temperature (catalyst loading: 0.25 mg·cm−2). The equation of E (vs. RHE) = E (vs. SCE) + 0.244 V + 0.0591 × PH was used to convert electrode potentials to the reversible hydrogen electrode (RHE). The electrolyte in all electrochemical tests was 1 M KOH presence of 0.1 M KNO3. And electrolyte volume was 35.0 mL in each test. The linear sweep voltammetry was scanned at a rate of 20 mV s−1. The nitrate electroreduction performance of all catalysts was evaluated by potentiostatic electrolysis for 1 h in the electrolyte at ambient conditions. The linear sweep voltammetry (LSV) curves were measured at a scan rate 5 mV s−1.

3.6. Isotope Labeling Experiments

1H NMR was used to perform isotope tracing experiments on a H-nuclear magnetic resonance (H-NMR) technic (Bruker AVANCEⅢ400 MHz) as well as for the quantifying the content of NH3 in liquid products.

3.7. Determination of NH3

Nessler’s reagent and potassium sodium tartrate solution (ρ = 500 g/L) were used as color reagents for NH3-N. First, 0.1 mL potassium sodium tartrate solution and 0.1 mL Nessler’s reagent was added to 5 mL diluted electrolyte. Then, standing for 20 min, the absorbance was measured by UV–vis spectrophotometry at a wavelength of 420 nm. The calibration curve was obtained by using a series of standard NH4Cl solutions. The faradaic efficiency (FE) was measured by the following formula:
F E = 8 F × c N H 3 × V Q × 100 %
where c is the concentration of NH3 in mmol L−1; V is the volume of electrolyte solution added in mL; m is the mass of catalyst loaded on the CP in mg; F is Faraday’s constant of 96,485 C mol−1; and Q is the total charge consumed in C.

3.8. Determination of NO2

To prepare the color reagents, 0.4 g p-aminobenzenesulfonamide and 0.02 g N-(1-naphthyl) ethylenediamine dihydrochloride were dissolved in a mixed solution containing 5 mL deionized water and 1 mL phosphoric acid (ρ = 1.70 g/mL). Next, 0.1 mL color reagent was added to the 5 mL diluted electrolyte. The solution was allowed to stand for 20 min before testing the absorbance by UV–vis spectrophotometry at a wavelength of 540 nm. The calibration curve was obtained by using a series of standard NaNO2 solutions. The faradaic efficiency (FE) was measured by the following formula:
F E = 2 F × c N O 2 × V Q × 100 %
where c is the concentration of NH3 in mmol L−1; V is the volume of electrolyte solution added in mL; m is the mass of catalyst loaded on the CP in mg; F is Faraday’s constant of 96,485 C mol−1; and Q is the total charge consumed in C.

3.9. Contact Angle Testing

The Contact Angle Measuring Instrument (SZ-CAMD3, Shanghai Sunzern Instrument Co., Shanghai, China) was used to test the contact angle of materials. Specifically, the angle between the liquid droplet and the solid surface was measured, thus reflecting the hydrophilicity or hydrophobicity of the solid surface to the liquid.

4. Conclusions

In summary, Co/BC-100 composites were successfully synthesized for efficient electrocatalytic nitrate reduction to ammonia by in situ carbonization using ZIF-67/biomass carbon as the precursor. Under mild reaction conditions, the yield of Co/BC-100 reached up to 3588.9 mmol gcat.−1 h−1 at −0.5 V vs. RHE potential, corresponding to a faradaic efficiency of 97.01%. Numerous studies have demonstrated that biomass carbon displays outstanding electrical conductivity, which enhances the charge transfer rate within the material. ZIF features a stable three-dimensional structure, which allows Co/BC-100 to display a large specific surface area and rich exposed metal active sites. The interfacial effect of Co and BC enhances the electron transfer and facilitates the conversion of nitrate to ammonia. Additionally, the intermediate products of the reaction were analyzed through in situ characterization, and the reaction pathway for electrocatalytic nitrate reduction to ammonia was determined. This paper provides a feasible approach for the design of Co-based composites for electrocatalytic nitrate reduction to ammonia.

Supplementary Materials

The following supporting information can be downloaded at https://www.mdpi.com/article/10.3390/catal14110817/s1, Figure S1: XPS spectra of Co/BC-100; Figure S2: UV absorption spectra of ammonia for Co/BC-0 at different potentials; Figure S3: UV absorption spectra of ammonia for Co/BC-50 at different potentials; Figure S4: UV absorption spectra of ammonia for Co/BC-200 at different potentials; Figure S5: Ammonia yields and faradaic efficiencies for Co/BC-0; Figure S6: Ammonia yields and faradaic efficiencies for Co/BC-50; Figure S7: Ammonia yields and faradaic efficiencies for Co/BC-200. Tabel S1: Reported catalysts and their performance for the nitrate reduction to ammonia. References [10,59,60,61,62,63,64,65,66,67,68,69,70,71,72,73] are cited in the Supplementary Materials.

Author Contributions

Conceptualization, J.Y. and H.L.; data curation, N.L. and S.L.; formal analysis, Y.D. and X.Y.; funding acquisition, N.L. and H.L.; investigation, Y.L. (Yaxi Li) and P.C.; methodology, Y.L. (Yunliang Liu); project administration, H.L.; resources, H.L.; software, Y.L. (Yixian Liu); supervision, H.L.; validation, Y.L. (Yaxi Li), Y.C. and X.Z.; visualization, H.L.; writing—original draft, J.Y. and Y.D.; writing—review and editing, N.L. and Y.L. (Yixian Liu). All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the National Natural Science Foundation of China (Grants 52072152, 51802126), the Jiangsu University Jinshan Professor Fund, the Jiangsu Specially-Appointed Professor Fund, the Open Fund from Guangxi Key Laboratory of Electrochemical Energy Materials, Zhenjiang “Jinshan Talents” Project 2021, China PostDoctoral Science Foundation (2022M721372), “Doctor of Entrepreneurship and Innovation” in Jiangsu Province (JSSCBS20221197), the Postgraduate Research & Practice Innovation Program of Jiangsu Province (Nos. KYCX22_3645 and KYCX24_3964), and Student Research Project of Jiangsu University (Nos. 23A586 and 23A806).

Data Availability Statement

Data are contained within the article and Supplementary Materials.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Fu, W.; Hu, Z.; Zheng, Y.; Su, P.; Zhang, Q.; Jiao, Y.; Zhou, M. Tuning mobility of intermediate and electron transfer to enhance electrochemical reduction of nitrate to ammonia on Cu2O/Cu interface. Chem. Eng. J. 2022, 433, 133680. [Google Scholar] [CrossRef]
  2. Fan, K.; Xie, W.; Li, J.; Sun, Y.; Xu, P.; Tang, Y.; Li, Z.; Shao, M. Active hydrogen boosts electrochemical nitrate reduction to ammonia. Nat. Commun. 2022, 13, 7958. [Google Scholar] [CrossRef] [PubMed]
  3. Li, Y.; Liu, Y.; Liu, X.; Liu, Y.; Cheng, Y.; Zhang, P.; Deng, P.; Deng, J.; Kang, Z.; Li, H. Fe-doped SnO2 nanosheet for ambient electrocatalytic nitrogen reduction reaction. Nano Res. 2022, 15, 6026–6035. [Google Scholar] [CrossRef]
  4. Deng, P.; Liu, Y.; Liu, Y.; Li, Y.; Wu, R.; Meng, L.; Liang, K.; Gan, Y.; Qiao, F.; Liu, N.; et al. Microwave Regenerable Nickel, Zinc Co-doped Nitrogen-Coordinated Porous Carbon Catalyst for Nitrogen Fixation. ACS Appl. Mater. Interfaces 2023, 15, 44809–44819. [Google Scholar] [CrossRef]
  5. Chen, D.; Zhang, S.; Yin, D.; Li, W.; Bu, X.; Quan, Q.; Lai, Z.; Wang, W.; Meng, Y.; Liu, C.; et al. Tailored p-Orbital Delocalization by Diatomic Pt-Ce Induced Interlayer Spacing Engineering for Highly-Efficient Ammonia Electrosynthesis. Adv. Energy Mater. 2023, 13, 2203201. [Google Scholar] [CrossRef]
  6. Wen, Y.; Wang, T.; Hao, J.; Zhuang, Z.; Gao, G.; Lai, F.; Lu, S.; Wang, X.; Kang, Q.; Wu, G.; et al. A Coherent Pd–Pd16B3 Core–Shell Electrocatalyst for Controlled Hydrogenation in Nitrogen Reduction Reaction. Adv. Funct. Mater. 2024, 34, 2400849. [Google Scholar] [CrossRef]
  7. Fang, J.; Zheng, Q.; Lou, Y.Y.; Zhao, K.; Hu, S.; Li, G.; Akdim, O.; Huang, X.; Sun, S. Ampere-level current density ammonia electrochemical synthesis using CuCo nanosheets simulating nitrite reductase bifunctional nature. Nat. Commun. 2022, 13, 7899. [Google Scholar] [CrossRef]
  8. Li, Z.; Zhuang, Z.; Lv, F.; Zhu, H.; Zhou, L.; Luo, M.; Zhu, J.; Lang, Z.; Feng, S.; Chen, W.; et al. The Marriage of the FeN4 Moiety and MXene Boosts Oxygen Reduction Catalysis: Fe 3d Electron Delocalization Matters. Adv. Mater. 2018, 30, 1803220. [Google Scholar] [CrossRef]
  9. Liu, Z.; Du, Y.; Yu, R.; Zheng, M.; Hu, R.; Wu, J.; Xia, Y.; Zhuang, Z.; Wang, D. Tuning Mass Transport in Electrocatalysis Down to Sub-5 nm through Nanoscale Grade Separation. Angew. Chem. Int. Ed. 2023, 62, e202212653. [Google Scholar] [CrossRef]
  10. Du, H.; Guo, H.; Wang, K.; Du, X.; Beshiwork, B.A.; Sun, S.; Luo, Y.; Liu, Q.; Li, T.; Sun, X. Durable electrocatalytic reduction of nitrate to ammonia over defective pseudobrookite Fe2TiO5 nanofibers with abundant oxygen vacancies. Angew. Chem. Int. Ed. 2023, 62, e202215782. [Google Scholar] [CrossRef]
  11. Cheng, Y.; Liu, Y.; Chu, C.; Liu, Y.; Li, Y.; Wu, R.; Wu, J.; Zhuang, C.; Kang, Z.; Li, H. Carbon armour with embedded carbon dots for building better supercapacitor electrodes. Nano Res. 2023, 16, 6815–6824. [Google Scholar] [CrossRef]
  12. Cheng, Y.; Jabeen, S.; Lei, S.; Liu, N.; Liu, Y.; Liu, Y.; Li, Y.; Wu, X.; Tong, Z.; Yu, J.; et al. N-doped carbon dots-modulated interfacial charge transfer and surface structure in FeNbO4 photocatalysts for enhanced CO2 conversion selectivity to CH4. Chem. Eng. J. 2024, 498, 155576. [Google Scholar] [CrossRef]
  13. Xia, J.; Xu, J.; Yu, B.; Liang, X.; Qiu, Z.; Li, H.; Feng, H.; Li, Y.; Cai, Y.; Wei, H.; et al. A Metal–Sulfur–Carbon Catalyst Mimicking the Two-Component Architecture of Nitrogenase. Angew. Chem. Int. Ed. 2024, 136, e202412740. [Google Scholar] [CrossRef]
  14. Zhuang, C.; Li, W.; Chang, Y.; Li, S.; Zhang, Y.; Li, Y.; Gao, J.; Chen, G.; Kang, Z. Coordination environment dominated catalytic selectivity of photocatalytic hydrogen and oxygen reduction over switchable gallium and nitrogen active sites. J. Mater. Chem. A 2024, 12, 5711–5718. [Google Scholar] [CrossRef]
  15. Liu, Y.; Xing, Y.; Xu, S.; Lu, Y.; Sun, S.; Jiang, D. Interfacing Co3Mo with CoMoOx for synergistically boosting electrocatalytic hydrogen and oxygen evolution reactions. Chem. Eng. J. 2022, 431, 133240. [Google Scholar] [CrossRef]
  16. Zhao, K.; Yang, W.; Li, L.; Wang, S.; Wang, L.; Qi, Z.; Yang, Y.; Chen, Z.; Zhuang, J.; Hao, J.; et al. Discharge Induced-Activation of Phosphorus-Doped Nickel Oxyhydroxide for Oxygen Evolution Reaction. Chem. Eng. J. 2022, 435, 135049. [Google Scholar]
  17. Wang, S.; Zhao, K.; Chen, Z.; Wang, L.; Qi, Z.; Hao, J.; Shi, W. New insights into cations effect in oxygen evolution reaction. Chem. Eng. J. 2022, 433, 133518. [Google Scholar] [CrossRef]
  18. Cheng, Y.; Liu, Y.; Liu, Y.; Li, Y.; Wu, R.; Du, Y.; Askari, N.; Liu, N.; Qiao, F.; Sun, C.; et al. A core-satellite structured type II heterojunction photocatalyst with enhanced CO2 reduction under visible light. Nano Res. 2022, 15, 8880–8889. [Google Scholar] [CrossRef]
  19. 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. 2022, 34, 2205767. [Google Scholar] [CrossRef]
  20. Zhuang, Z.; Li, Y.; Li, Z.; Lv, F.; Lang, Z.; Zhao, K.; Zhou, L.; Moskaleva, L.; Guo, S.; Mai, L. MoB/g-C3N4 Interface Materials as a Schottky Catalyst to Boost Hydrogen Evolution. Angew. Chem. Int. Ed. 2018, 57, 496–500. [Google Scholar] [CrossRef]
  21. Zhuang, Z.; Xia, L.; Huang, J.; Zhu, P.; Li, Y.; Ye, C.; Xia, M.; Yu, R.; Lang, Z.; Zhu, J.; et al. Continuous Modulation of Electrocatalytic Oxygen Reduction Activities of Single-Atom Catalysts through p-n Junction Rectification. Angew. Chem. Int. Ed. 2023, 62, e202212335. [Google Scholar] [CrossRef] [PubMed]
  22. Wu, X.; Tong, Z.; Liu, Y.; Li, Y.; Cheng, Y.; Yu, J.; Cao, P.; Zhuang, C.; Shi, Q.; Liu, N.; et al. Modification of the CuO electronic structure for enhanced selective electrochemical CO2 reduction to ethylene. Nano Res. 2024, 17, 7194–7202. [Google Scholar] [CrossRef]
  23. Tong, Z.; Liu, Y.; Wu, X.; Cheng, Y.; Yu, J.; Zhang, X.; Liu, N.; Liu, X.; Li, H. Carbon Quantum Dots/Cu2O Photocatalyst for Room Temperature Selective Oxidation of Benzyl Alcohol. Nanomaterials 2024, 14, 212. [Google Scholar] [CrossRef] [PubMed]
  24. Zhu, H.; Sun, S.; Hao, J.; Zhuang, Z.; Zhang, S.; Wang, T.; Kang, Q.; Lu, S.; Wang, X.; Lai, F.; et al. A high-entropy atomic environment converts inactive to active sites for electrocatalysis. Energy Environ. Sci. 2023, 16, 619–628. [Google Scholar] [CrossRef]
  25. Wang, S.; Zhuang, Z.; Xu, J.; Fu, C.; Qiu, Z.; Feng, H.; Xiang, H.; Chen, Z.; Li, H.; Zhang, L.; et al. Non-rigid metal–oxygen bonding empowered nitrate reduction on ruthenium catalysts. Nano Energy 2024, 130, 110088. [Google Scholar] [CrossRef]
  26. Zhuang, Z.; Li, Y.; Yu, R.; Xia, L.; Yang, J.; Lang, Z.; Zhu, J.; Huang, J.; Wang, J.; Wang, Y.; et al. Reversely trapping atoms from a perovskite surface for high-performance and durable fuel cell cathodes. Nat. Catal. 2022, 5, 300–310. [Google Scholar] [CrossRef]
  27. Zhuang, Z.; Li, Y.; Li, Y.; Huang, J.; Wei, B.; Sun, R.; Ren, Y.; Ding, J.; Zhu, J.; Lang, Z.; et al. Atomically dispersed nonmagnetic electron traps improve oxygen reduction activity of perovskite oxides. Energy Environ. Sci. 2021, 14, 1016–1028. [Google Scholar] [CrossRef]
  28. Hao, J.; Wang, T.; Yu, R.; Cai, J.; Gao, G.; Zhuang, Z.; Kang, Q.; Lu, S.; Liu, Z.; Wu, J.; et al. Integrating few-atom layer metal on high-entropy alloys to catalyze nitrate reduction in tandem. Nat. Commun. 2024, 15, 9020. [Google Scholar] [CrossRef]
  29. Hao, J.; Zhuang, Z.; Cao, K.; Gao, G.; Wang, C.; Lai, F.; Lu, S.; Ma, P.; Dong, W.; Liu, T.; et al. Unraveling the electronegativity-dominated intermediate adsorption on high-entropy alloy electrocatalysts. Nat. Commun. 2022, 13, 2662. [Google Scholar] [CrossRef]
  30. 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. 2020, 59, 10479–10483. [Google Scholar] [CrossRef]
  31. Liu, N.; Liu, Y.; Liu, Y.; Li, Y.; Cheng, Y.; Li, H.T. Modulation of photogenerated holes for enhanced photoelectrocatalytic performance. Microstructures 2023, 3, 2023001. [Google Scholar] [CrossRef]
  32. Liu, Y.; Li, Q.; Guo, X.; Kong, X.; Ke, J.; Chi, M.; Li, Q.; Geng, Z.; Zeng, J. A Highly Efficient Metal-Free Electrocatalyst of F-Doped Porous Carbon toward N2 Electroreduction. Adv. Mater. 2020, 32, 1907690. [Google Scholar] [CrossRef] [PubMed]
  33. Zhuang, C.; Chang, Y.; Li, W.; Li, S.; Xu, P.; Zhang, H.; Zhang, Y.; Zhang, C.; Gao, J.; Chen, G.; et al. Light-Induced Variation of Lithium Coordination Environment in g-C3N4 Nanosheet for Highly Efficient Oxygen Reduction Reactions. ACS Nano 2024, 18, 5206–5217. [Google Scholar] [CrossRef] [PubMed]
  34. Liu, Y.; Wu, R.; Liu, Y.; Deng, P.; Li, Y.; Cheng, Y.; Du, Y.; Li, Z.; Yan, X.; Liu, N.; et al. Interface engineering of CoS/MoS2 heterostructure for the electrocatalytic reduction of N2 to NH3. Inorg. Chem. Front. 2023, 10, 5700–5709. [Google Scholar] [CrossRef]
  35. Ling, C.; Zhang, Y.; Li, Q.; Bai, X.; Shi, L.; Wang, J. New Mechanism for N2 Reduction: The Essential Role of Surface Hydrogenation. J. Am. Chem. Soc. 2019, 141, 18264–18270. [Google Scholar] [CrossRef]
  36. Liu, Y.; Zheng, Z.; Jabeen, S.; Liu, N.; Liu, Y.; Cheng, Y.; Li, Y.; Yu, J.; Wu, X.; Yan, N.; et al. Mechanochemical route to fabricate an efficient nitrate reduction electrocatalyst. Nano Res. 2024, 17, 4889–4897. [Google Scholar] [CrossRef]
  37. Liu, N.; Wu, R.; Liu, Y.; Liu, Y.; Deng, P.; Li, Y.; Du, Y.; Cheng, Y.; Zhuang, Z.; Kang, Z.; et al. Oxygen Vacancy Engineering of Fe-Doped NiMoO4 for Electrocatalytic N2 Fixation to NH3. Inorg. Chem. 2023, 62, 11990–12000. [Google Scholar] [CrossRef]
  38. Liu, Y.; Zhuang, Z.; Liu, Y.; Liu, N.; Li, Y.; Cheng, Y.; Yu, J.; Yu, R.; Li, H.; Wang, D. Shear-strained Pd Single-atom Electrocatalysts for Nitrate Reduction to Ammonia. Angew. Chem. Int. Ed. 2024, 63, e202411396. [Google Scholar]
  39. Bhuvanendran, N.; Ravichandran, S.; Jayaseelan, S.S.; Xu, Q.; Khotseng, L.; Su, H. Improved bi-functional oxygen electrocatalytic performance of Pt–Ir alloy nanoparticles embedded on MWCNT with Pt-enriched surfaces. Energy 2020, 211, 118695. [Google Scholar] [CrossRef]
  40. Gao, J.; Jiang, B.; Ni, C.; Qi, Y.; Bi, X. Enhanced reduction of nitrate by noble metal-free electrocatalysis on P doped three-dimensional Co3O4 cathode: Mechanism exploration from both experimental and DFT studies. Chem. Eng. J. 2020, 382, 123034. [Google Scholar] [CrossRef]
  41. Chen, G.-F.; Yuan, Y.; Jiang, H.; Ren, S.-Y.; Ding, L.-X.; Ma, L.; Wu, T.; Lu, J.; Wang, H. Electrochemical reduction of nitrate to ammonia via direct eight-electron transfer using a copper–molecular solid catalyst. Nat. Energy 2020, 5, 605–613. [Google Scholar] [CrossRef]
  42. Niu, Z.; Fan, S.; Li, X.; Liu, Z.; Wang, J.; Duan, J.; Tadé, M.O.; Liu, S. Facile tailoring of the electronic structure and the d-band center of copper-doped cobaltate for efficient nitrate electrochemical hydrogenation. ACS Appl. Mater. Interfaces 2022, 14, 35477–35484. [Google Scholar] [CrossRef] [PubMed]
  43. Li, J.; Li, M.; An, N.; Zhang, S.; Song, Q.; Yang, Y.; Li, J.; Liu, X. Boosted ammonium production by single cobalt atom catalysts with high Faradic efficiencies. Proc. Natl. Acad. Sci. USA 2022, 119, e2123450119. [Google Scholar] [CrossRef] [PubMed]
  44. Kuznetsova, I.; Lebedeva, O.; Kultin, D.; Mashkin, M.; Kalmykov, K.; Kustov, L. Enhancing Efficiency of Nitrate Reduction to Ammonia by Fe and Co Nanoparticle-Based Bimetallic Electrocatalyst. Int. J. Mol. Sci. 2024, 25, 7089. [Google Scholar] [CrossRef] [PubMed]
  45. Zhang, Y.; Xiong, J.; Liu, B.; Yan, S. Electrocatalytic nitrate reduction to ammonia by sea-urchin-like CoNiO2 under mild conditions. CRPS 2024, 5, 101994. [Google Scholar] [CrossRef]
  46. Zhang, J.; Xu, D.; Lu, D.; Wang, H. The synergistic tandem effect of Cu2+1O and Co3O4 enhances the activity and selectivity of nitrate reduction to ammonia in neutral solution. Appl. Catal. A 2024, 677, 119695. [Google Scholar] [CrossRef]
  47. Dong, W.; Zhou, H.; Mao, B.; Zhang, Z.; Liu, Y.; Liu, Y.; Li, F.; Zhang, D.; Zhang, D.; Shi, W. Efficient MOF- derived V–Ni3S2 nanosheet arrays for electrocatalytic overall water splitting in alkali. Int. J. Hydrogen Energy 2021, 46, 10773–10782. [Google Scholar] [CrossRef]
  48. Yu, Z.; Si, C.; Lagrow, A.P.; Tai, Z.; Caliebe, W.A.; Tayal, A.; Sampaio, M.J.; Sousa, J.P.S.; Amorim, I.; Araujo, A.; et al. Iridium–Iron Diatomic Active Sites for Efficient Bifunctional Oxygen Electrocatalysis. ACS Catal. 2022, 12, 9397–9409. [Google Scholar] [CrossRef]
  49. Yu, Z.; Xia, G.; Diaconescu, V.M.; Simonelli, L.; Lagrow, A.P.; Tai, Z.; Xiang, X.; Xiong, D.; Liu, L. Atomically dispersed dinuclear iridium active sites for efficient and stable electrocatalytic chlorine evolution reaction. Chem. Sci. 2024, 15, 9216–9223. [Google Scholar] [CrossRef]
  50. Li, C.; Zhao, D.-H.; Long, H.-L.; Li, M. Recent advances in carbonized non-noble metal–organic frameworks for electrochemical catalyst of oxygen reduction reaction. Rare Met. 2021, 40, 2657–2689. [Google Scholar] [CrossRef]
  51. Liu, D.; Qiao, L.; Chen, Y.; Zhou, P.; Feng, J.; Leong, C.C.; Ng, K.W.; Peng, S.; Wang, S.; Ip, W.F.; et al. Electrocatalytic reduction of nitrate to ammonia on low-cost manganese-incorporated Co3O4 nanotubes. Appl. Catal. B 2023, 324, 122293. [Google Scholar] [CrossRef]
  52. Wang, Z.; Peng, S.; Hu, Y.; Li, L.; Yan, T.; Yang, G.; Ji, D.; Srinivasan, M.; Pan, Z.; Ramakrishna, S. Cobalt nanoparticles encapsulated in carbon nanotube-grafted nitrogen and sulfur co-doped multichannel carbon fibers as efficient bifunctional oxygen electrocatalysts. J. Mater. Chem. A 2017, 5, 4949–4961. [Google Scholar] [CrossRef]
  53. Wu, Y.-j.; Wu, X.-h.; Tu, T.-x.; Zhang, P.-f.; Li, J.-t.; Zhou, Y.; Huang, L.; Sun, S.-g. Controlled synthesis of FeNx-CoNx dual active sites interfaced with metallic Co nanoparticles as bifunctional oxygen electrocatalysts for rechargeable Zn-air batteries. Appl. Catal. B 2020, 278, 119259. [Google Scholar] [CrossRef]
  54. Wang, J.; Wang, Y.; Cai, C.; Liu, Y.; Wu, D.; Wang, M.; Li, M.; Wei, X.; Shao, M.; Gu, M. Cu-doped Iron oxide for the efficient electrocatalytic nitrate reduction reaction. Nano Lett. 2023, 23, 1897–1903. [Google Scholar] [CrossRef]
  55. Zhu, W.J.; Zhang, X.Y.; Yao, F.; Huang, R.P.; Chen, Y.; Chen, C.L.; Fei, J.W.; Chen, Y.Q.; Wang, Z.C.; Liang, H.F. A Hydrazine-Nitrate Flow Battery Catalyzed by a Bimetallic RuCo Precatalyst for Wastewater Purification along with Simultaneous Generation of Ammonia and Electricity. Angew. Chem. Int. Ed. 2023, 62, e202300390. [Google Scholar] [CrossRef]
  56. Zhu, W.J.; Yao, F.; Wu, Q.F.; Jiang, Q.; Wang, J.X.; Wang, Z.C.; Liang, H.F. Weakened d–p orbital hybridization in in situ reconstructed Ru/β-Co(OH)2 heterointerfaces for accelerated ammonia electrosynthesis from nitrates. Energy Environ. Sci. 2023, 16, 2483–2493. [Google Scholar] [CrossRef]
  57. Jang, W.; Oh, D.; Lee, J.; Kim, J.; Matthews, J.E.; Kim, H.; Lee, S.W.; Lee, S.; Xu, Y.; Yu, J.M.; et al. Homogeneously Mixed Cu–Co Bimetallic Catalyst Derived from Hydroxy Double Salt for Industrial-Level High-Rate Nitrate-to-Ammonia Electrosynthesis. J. Am. Chem. Soc. 2024, 146, 27417–27428. [Google Scholar] [CrossRef]
  58. Liu, L.; Xiao, T.; Fu, H.; Chen, Z.; Qu, X.; Zheng, S. Construction and identification of highly active single-atom Fe1-NC catalytic site for electrocatalytic nitrate reduction. Appl. Catal. B 2023, 323, 122181. [Google Scholar] [CrossRef]
  59. Zhang, J.; He, W.; Quast, T.; Junqueira, J.R.C.; Saddeler, S.; Schulz, S.; Schuhmann, W. Single-entity Electrochemistry Unveils Dynamic Transformation during Tandem Catalysis of Cu2O and Co3O4 for Converting NO3 to NH3. Angew. Chem. Int. Ed. 2023, 62, e202214830. [Google Scholar] [CrossRef]
  60. Fu, Y.; Wang, S.; Wang, Y.; Wei, P.; Shao, J.; Liu, T.; Wang, G.; Bao, X. Enhancing Electrochemical Nitrate Reduction to Ammonia over Cu Nanosheets via Facet Tandem Catalysis. Angew. Chem. Int. Ed. 2023, 62, e202303327. [Google Scholar] [CrossRef]
  61. Chen, D.; Zhang, S.; Bu, X.; Zhang, R.; Quan, Q.; Lai, Z.; Wang, W.; Meng, Y.; Yin, D.; Yip, S.; et al. Synergistic modulation of local environment for electrochemical nitrate reduction via asymmetric vacancies and adjacent ion clusters. Nano Energy 2022, 98, 107338. [Google Scholar] [CrossRef]
  62. 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. 2022, 34, 2204306. [Google Scholar] [CrossRef]
  63. Gao, W.; Xie, K.; Xie, J.; Wang, X.; Zhang, H.; Chen, S.; Wang, H.; Li, Z.; Li, C. Alloying of Cu with Ru Enabling the Relay Catalysis for Reduction of Nitrate to Ammonia. Adv. Mater. 2023, 35, 2202952. [Google Scholar] [CrossRef]
  64. Zhang, S.; Wu, J.; Zheng, M.; Jin, X.; Shen, Z.; Li, Z.; Wang, Y.; Wang, Q.; Wang, X.; Wei, H.; et al. Fe/Cu diatomic catalysts for electrochemical nitrate reduction to ammonia. Nat. Commun. 2023, 14, 3634. [Google Scholar] [CrossRef]
  65. Fang, Z.; Jin, Z.; Tang, S.; Li, P.; Wu, P.; Yu, G. Porous Two-dimensional Iron-Cyano Nanosheets for High-rate Electrochemical Nitrate Reduction. ACS Nano 2022, 16, 1072–1081. [Google Scholar] [CrossRef]
  66. Wu, Z.-Y.; Karamad, M.; Yong, X.; Huang, Q.; Cullen, D.A.; Zhu, P.; Xia, C.; Xiao, Q.; Shakouri, M.; Chen, F.-Y.; et al. Electrochemical ammonia synthesis via nitrate reduction on Fe single atom catalyst. Nat. Commun. 2021, 12, 2870. [Google Scholar] [CrossRef]
  67. Xu, Y.; Ren, K.; Ren, T.; Wang, M.; Wang, Z.; Li, X.; Wang, L.; Wang, H. Ultralow-content Pd in-situ incorporation mediated hierarchical defects in corner-etched Cu2O octahedra for enhanced electrocatalytic nitrate reduction to ammonia. Appl. Catal. B 2022, 306, 121094. [Google Scholar] [CrossRef]
  68. Xu, Y.; Ren, K.; Ren, T.; Wang, M.; Liu, M.; Wang, Z.; Li, X.; Wang, L.; Wang, H. Cooperativity of Cu and Pd active sites in CuPd aerogels enhances nitrate electroreduction to ammonia. Chem. Commun. 2021, 57, 7525–7528. [Google Scholar] [CrossRef]
  69. Gao, Q.; Pillai, H.S.; Huang, Y.; Liu, S.; Mu, Q.; Han, X.; Yan, Z.; Zhou, H.; He, Q.; Xin, H.; et al. Breaking adsorption-energy scaling limitations of electrocatalytic nitrate reduction on intermetallic CuPd nanocubes by machine-learned insights. Nat. Commun. 2022, 13, 2338. [Google Scholar] [CrossRef]
  70. Xie, M.; Tang, S.; Li, Z.; Wang, M.; Jin, Z.; Li, P.; Zhan, X.; Zhou, H.; Yu, G. Intermetallic Single-Atom Alloy In–Pd Bimetallene for Neutral Electrosynthesis of Ammonia from Nitrate. J. Am. Chem.Soc. 2023, 145, 13957–13967. [Google Scholar] [CrossRef]
  71. Sun, L.; Liu, B. Mesoporous PdN Alloy Nanocubes for Efficient Electrochemical Nitrate Reduction to Ammonia. Adv. Mater. 2023, 35, 2207305. [Google Scholar] [CrossRef] [PubMed]
  72. Chen, K.; Ma, Z.; Li, X.; Kang, J.; Ma, D.; Chu, K. Single-Atom Bi Alloyed Pd Metallene for Nitrate Electroreduction to Ammonia. Adv. Funct. Mater. 2023, 33, 2209890. [Google Scholar] [CrossRef]
  73. Han, Y.; Zhang, X.; Cai, W.; Zhao, H.; Zhang, Y.; Sun, Y.; Hu, Z.; Li, S.; Lai, J.; Wang, L. Facet-controlled palladium nanocrystalline for enhanced nitrate reduction towards ammonia. J. Colloid Interface Sci. 2021, 600, 620–628. [Google Scholar] [CrossRef]
Figure 1. Schematic illustration for the synthesis of Co/BC-100.
Figure 1. Schematic illustration for the synthesis of Co/BC-100.
Catalysts 14 00817 g001
Figure 2. (a) TEM, (b) HRTEM, and (c) local magnification of the red dashed box for Co/BC-100. (dh) TEM of Co/BC-100 and the corresponding elemental distribution mapping.
Figure 2. (a) TEM, (b) HRTEM, and (c) local magnification of the red dashed box for Co/BC-100. (dh) TEM of Co/BC-100 and the corresponding elemental distribution mapping.
Catalysts 14 00817 g002
Figure 3. (a) XRD patterns of BC, Co/BC-0, Co/BC-50, Co/BC-100, and Co/BC-200; (b) Raman spectra; (c) contact angle test plots of Co/BC-100.
Figure 3. (a) XRD patterns of BC, Co/BC-0, Co/BC-50, Co/BC-100, and Co/BC-200; (b) Raman spectra; (c) contact angle test plots of Co/BC-100.
Catalysts 14 00817 g003
Figure 4. XPS spectra of Co/BC-100: (a) high-resolution C 1s, (b) high-resolution N 1s, (c) high-resolution Co 2p, and (d) high-resolution O 1s.
Figure 4. XPS spectra of Co/BC-100: (a) high-resolution C 1s, (b) high-resolution N 1s, (c) high-resolution Co 2p, and (d) high-resolution O 1s.
Catalysts 14 00817 g004
Figure 5. (a) CV curves of Co/BC-100 in the range of −0.73 to 0.27 V vs. RHE and (b) LSV curves of Co/BC-0, Co/BC-50, Co/BC-100, and Co/BC-200 in the range of −0.73 to 0.27 V vs. RHE. (c) I-t curves of Co/BC-100 measured at different potentials and (d) UV absorption spectra of the corresponding products, (e) ammonia yield and faradaic efficiency, and (f) nitrite yield and faradaic efficiency.
Figure 5. (a) CV curves of Co/BC-100 in the range of −0.73 to 0.27 V vs. RHE and (b) LSV curves of Co/BC-0, Co/BC-50, Co/BC-100, and Co/BC-200 in the range of −0.73 to 0.27 V vs. RHE. (c) I-t curves of Co/BC-100 measured at different potentials and (d) UV absorption spectra of the corresponding products, (e) ammonia yield and faradaic efficiency, and (f) nitrite yield and faradaic efficiency.
Catalysts 14 00817 g005
Figure 6. (a) Five i-t curves of Co/BC-100 measured at −0.4 V vs. RHE potential and corresponding (b) UV absorption spectra of ammonia, (c) ammonia yield and faradaic efficiency, (d) UV absorption spectra of nitrite, (e) nitrite yield and faradaic efficiency, and (f) i-t curves of Co/BC-100 after 12 h long stability test.
Figure 6. (a) Five i-t curves of Co/BC-100 measured at −0.4 V vs. RHE potential and corresponding (b) UV absorption spectra of ammonia, (c) ammonia yield and faradaic efficiency, (d) UV absorption spectra of nitrite, (e) nitrite yield and faradaic efficiency, and (f) i-t curves of Co/BC-100 after 12 h long stability test.
Catalysts 14 00817 g006
Figure 7. (a) LSV curves measured under the conditions of Co/BC-100, empty carbon paper, and electrolyte without NO3; (b) corresponding ammonia yields; (c) 1H NMR spectra of Co/BC-100 using 15NO3 and 14NO3 as electrolyte.
Figure 7. (a) LSV curves measured under the conditions of Co/BC-100, empty carbon paper, and electrolyte without NO3; (b) corresponding ammonia yields; (c) 1H NMR spectra of Co/BC-100 using 15NO3 and 14NO3 as electrolyte.
Catalysts 14 00817 g007
Figure 8. CV curves measured at different sweep speeds (20, 40, 60, 80, and 100 m V s−1) for (a) Co/BC-0, (b) Co/BC-50, (c) Co/BC-100, (d) and Co/BC-200; (e) comparison of Cdl; (f) comparison of EIS.
Figure 8. CV curves measured at different sweep speeds (20, 40, 60, 80, and 100 m V s−1) for (a) Co/BC-0, (b) Co/BC-50, (c) Co/BC-100, (d) and Co/BC-200; (e) comparison of Cdl; (f) comparison of EIS.
Catalysts 14 00817 g008
Figure 9. (a) In situ FT-IR spectra of CV from 0 to −1.0 V vs. RHE sweep at Co/BC-00 electrode in 0.1 M KOH and 0.1 M KNO3 electrolyte. The reference spectra were measured at open circuit voltage. (b) Reaction mechanism diagram.
Figure 9. (a) In situ FT-IR spectra of CV from 0 to −1.0 V vs. RHE sweep at Co/BC-00 electrode in 0.1 M KOH and 0.1 M KNO3 electrolyte. The reference spectra were measured at open circuit voltage. (b) Reaction mechanism diagram.
Catalysts 14 00817 g009
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Yu, J.; Du, Y.; Liu, S.; Liu, Y.; Li, Y.; Cheng, Y.; Cao, P.; Zhang, X.; Yuan, X.; Liu, N.; et al. Co-Carbonization of Straw and ZIF-67 to the Co/Biomass Carbon for Electrocatalytic Nitrate Reduction. Catalysts 2024, 14, 817. https://doi.org/10.3390/catal14110817

AMA Style

Yu J, Du Y, Liu S, Liu Y, Li Y, Cheng Y, Cao P, Zhang X, Yuan X, Liu N, et al. Co-Carbonization of Straw and ZIF-67 to the Co/Biomass Carbon for Electrocatalytic Nitrate Reduction. Catalysts. 2024; 14(11):817. https://doi.org/10.3390/catal14110817

Chicago/Turabian Style

Yu, Jingwen, Yongchao Du, Shuaiqi Liu, Yunliang Liu, Yaxi Li, Yuanyuan Cheng, Peng Cao, Xinyue Zhang, Xinya Yuan, Naiyun Liu, and et al. 2024. "Co-Carbonization of Straw and ZIF-67 to the Co/Biomass Carbon for Electrocatalytic Nitrate Reduction" Catalysts 14, no. 11: 817. https://doi.org/10.3390/catal14110817

APA Style

Yu, J., Du, Y., Liu, S., Liu, Y., Li, Y., Cheng, Y., Cao, P., Zhang, X., Yuan, X., Liu, N., Liu, Y., & Li, H. (2024). Co-Carbonization of Straw and ZIF-67 to the Co/Biomass Carbon for Electrocatalytic Nitrate Reduction. Catalysts, 14(11), 817. https://doi.org/10.3390/catal14110817

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop