Next Article in Journal
Promoting the Electrocatalytic Ethanol Oxidation Activity of Pt by Alloying with Cu
Previous Article in Journal
On Stability of High-Surface-Area Al2O3, TiO2, SiO2-Al2O3, and Activated Carbon Supports during Preparation of NiMo Sulfide Catalysts for Parallel Deoxygenation of Octanoic Acid and Hydrodesulfurization of 1-Benzothiophene
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

One Bicopper Complex with Good Affinity to Nitrate for Highly Selective Electrocatalytic Nitrate Reduction to Ammonia

1
Institute of Materials Science and Devices, School of Material Science and Engineering, Suzhou University of Science and Technology, Suzhou 215009, China
2
School of Chemistry and Chemical Engineering and Anyang Key Laboratory of New Functional Complex Materials, Anyang Normal University, Anyang 455000, China
*
Authors to whom correspondence should be addressed.
Catalysts 2022, 12(12), 1561; https://doi.org/10.3390/catal12121561
Submission received: 15 October 2022 / Revised: 24 November 2022 / Accepted: 25 November 2022 / Published: 2 December 2022
(This article belongs to the Section Electrocatalysis)

Abstract

:
Ammonia (NH3) plays an irreplaceable role in human life as a promising energy carrier and indispensable chemical raw material. Nitrate electroreduction to ammonium (NRA) not only removes nitrate pollutants, but also can be used for efficient NH3 production under ambient conditions. However, achieving high efficiency and selectivity of electrocatalysts is still a great challenge. Herein, a complex Cu2(NO3)4(BMMB)·H2O with a bicopper core is assembled by Cu(NO3)2·3H2O and 1,4-bis{[2-(2’-pyridyl)benzimidazolyl]methyl}benzene (BMMB) for NRA under alkaline conditions. The optimal sample showed excellent nitrate reduction performance with the NO3 conversion rate of 70%, Faradaic efficiency of up to 90%, and NH3 selectivity of more than 95%. The high-catalytic activity is mainly due to the ingeniously designed copper cores with strong affinity for NO3, which accelerates the transferring rate of adsorbed nitrate on the Cu surface and increases the efficiency of rate-determining step (NO3 → NO2) in the whole catalytic process. Therefore, the transformation of surface-exposed nitrate can be rapidly catalyzed by the Cu active sites, facilitating the conversion efficiency of nitrate.

1. Introduction

Currently, water pollution poses a major potential threat to public health and lives. Nitrate (NO3) plays a non-negligible role among the pollutants [1]. In our daily life, the use of fertilizers, chemical production, and fossil fuels causes massive diffusion of nitrate and nitrogen oxides to surface water and groundwater, which resulted in the water quality not meeting the standard [2,3]. In addition, nitrate can be converted into nitrite (NO2) in the human body, which causes severe health problems; for instance, gastric cancer and methemoglobin [4,5,6]. Furthermore, high concentrations of nitrate in the air can lead to acid rain, while in water, high concentrations can contribute to eutrophication [7].
To efficiently denitrify NO3, numerous methods, such as reverse osmosis, adsorption, catalytic reduction, and biological denitrification have been used to treat NO3 [8,9,10,11]. However, these techniques have shown some weaknesses. For example, reverse osmosis and adsorption need a post-treatment of highly concentrated NO3, catalytic reduction is costly, biological denitrification is complicated and requires a long time to treat [12]. Electrocatalytic reduction method is more effective to treat nitrate than the above-mentioned methods. In addition, the highly selective reduction product, ammonia (NH3), can be obtained by an external power supply at room temperature under atmospheric pressure [13,14,15,16]. The most desirable product of electrocatalytic nitrate reduction is NH3, considering that NH3 is an industrially important raw material and a promising liquefied fuel [17]. Industrial NH3 is mainly a synthesis using Haber-Bosch process, which operates at high temperature (400~600 °C) and high pressure (>400 atm) [18]. However, NO3 electrocatalytic reduction to valuable NH3 is an environmentally-friendly, safe, and sustainable process. According to the reported literatures, the rate-determining step of NRA is the conversion of NO3 → NO2; then, the important intermediate NO will be selectively converted into nitrogen or ammonia depending on electrode materials, solution PH, applied voltage, etc. [19,20,21]. The key point affecting the rate-determining step is the rate of nitrate adsorption on the catalyst surface. Therefore, the suitability of an electrocatalytic material for electrocatalytic nitrate reduction is determined by the ability to effectively adsorb NO3 [22]. Among these methods, the NRA technique may be the most suitable for industrial production.
With regard to NRA, electrode materials mainly determine the electrocatalytic performances. At present, noble metal catalysts, such as gold (Au), rubidium (Ru), and platinum (Pt), have been shown to be excellent electrocatalytic materials [23,24,25,26,27,28]. Among these materials, although the performance of precious metals is excellent, the cost is expensive and there is no possibility of large-scale use. Transition metals are more suitable materials due to their low price and high electrocatalytic activity and stability. With regard to transition metal catalysts, copper (Cu) is one of the most desirable metals with high Faradaic efficiency (FE) and selectivity toward NH3 [29,30]. Koper et al. demonstrated that Cu has the highest activity among transition metal catalysts [31]. Wu et al. reported that Cu nanosheets (Cu NSs) have high electric-double-layer-capacitance of 422 mF, indicating a very large electrocatalytic active area. Therefore, Cu NSs exhibit high ammonia selectivity of 99.7% nitrate removal rate [32].
Previous researches declare that some complexes show the ability to produce ammonia by nitrate electrocatalytic reduction. Jakubikowa et al. clarified that [Co(DIM)]3+ (DIM = 2,3-dimethyl-1,4,8,11-tetraazacyclotetradeca-1,3-diene) has an active NO3 electrocatalytic reduction to valuable NH3, and density functional theory (DFT) was used to study the electronic structure and reaction mechanisms [33]. However, improving the NRA efficiency of complexes is still a challenge. Some efforts have been proposed to improve the catalytic activity. Luo et al. reported that Th-BPYDC (BPYDC = 2,2′-bipyridine-5,5′-dicarboxylic acid) has a low catalytic activity of NRA, but once supported by single-site copper catalyst, it presents 225 μmol h−1 cm−2 yield and 92.5% Faradaic efficiency for NRA activated at 0.0 V (vs. RHE) [34]. Zhu et al. used an in-situ synthesis method to populate Cu nanoclusters in CuHHTP (HHTP = 2,3,6,7,10,11-hexahydroxytriphenylene) [35]. In addition, Xu et al. used a similar approach to fill Cu nanoparticles in CeBDC (H2BDC = 1,4-benzenedicarboxylic acid) [36]. All of these methods are based on the loading of metal particles with a complex carrier, and the active center of catalysis is still the metal particle. Could the initial complexes have good catalytic activity?
With the above assumptions, we proposed a strategy for improving catalytic activity through designing a new complex Cu2(NO3)4(BMMB)·H2O (CuBMMB) with a bicopper core, which exhibits high activity for NRA. Specifically, at −0.53 V vs. reversible hydrogen electrode (RHE), CuBMMB achieves NO3 conversion rate and Faradaic efficiency (FE) of up to 70 and 90%, and NH3 selectivity of more than 95% in 1 M KOH solution with 200 ppm NO3-N. The structure and composition are characterized in detail. To study the selectivity, mass activity, and durability, a combination of characterization tools and electrochemical tests was conducted. Finally, the mechanisms for the significant improvement in electrochemical performance are disclosed.

2. Results and Discussion

2.1. Synthesis of Consideration

First, the semirigid organic 1,4-bis{[2-(2′-pyridyl)benzimidazolyl]methyl}benzene (BMMB) chosen as the desirable ligand is attributed to the fixed bidentate bridging mode, and it makes the coordination with metal ions easier to be regulated (Figure S1). Second, the copper ion possesses good NRA performances. Third, it is important to increase the reaction rate in the rate-determining step (NO3 → NO2) of NRA process, and it is a practical strategy to focus on increasing the adsorption efficiency of nitrate on the surface of catalytically active centers. Therefore, we have used copper nitrate as the metal salt, and successfully achieved good affinity of Cu2+ ion with the nitrate ion, which led to the acceleration of the transferring rate of adsorbed nitrate on the Cu surface and improvement of the NRA performances.

2.2. Crystal Structure of CuBMMB

Single-crystal X-ray diffraction analysis reveals that CuBMMB crystallizes in the monoclinic space group C2/c (see Supplementary Material). The asymmetric unit consists of one crystallographically independent Cu2+ ion, two nitrate ions, one BMMB ligand, and one free water molecule. As shown in Figure 1, the Cu2+ ion is coordinated by two BMMB nitrogen atoms and two nitrate oxygen atoms, which yield a square coordination environment. The Cu–N bond lengths are 1.952(3) and 1.978(3) Å, and the Cu–O bond lengths are 1.968(3) and 1.980(3) Å, which are all comparable to those typically observed values in the related copper complexes. Two terminal groups of BMMB connects one Cu2+ ion, and thus each BMMB ligand connects two Cu2+ ions. However, the two other coordination sites of Cu2+ ion are occupied by the monodentate nitrate ion, which hinders the extension of metal node. Finally, the BMMB ligand connects the Cu2+ ion into the complex molecule with a bicopper core.
The X-ray diffraction (XRD) experimental pattern is well-matched with the simulated pattern (Figure 2a), proving the high purity and crystallinity of CuBMMB. With regard to CuBMMB, the distinct peaks at 9.53, 14.57, 18.58, 22.28, and 24.75° can be indexed to the (2 0 2 ¯ ), (4 0 0), (6 0 2 ¯ ), (3 1 2), and (7 1 2 ¯ ) facets. The SEM images (Figure S1, Figure 2b,c) of CuBMMB present a loose structure with a large number of micron pores on the surface. From the magnified Figure 2b, it can be seen that CuBMMB particles are composed of many small rods with flocculent surfaces interlaced and woven. Not surprisingly, this open porous structure could accelerate the mass transfer of nitrate ions and products. SEM elemental mapping (Figure 2d) of CuBMMB shows that C, N, O, and Cu elements are homogeneously dispersed on the surface of CuBMMB. Energy dispersive X-ray spectroscopy (EDS) (Figure S2) calculates the content of each component and the results are shown in Table S1. The proportions of carbon, nitrogen, oxygen, and copper are 73.25, 4.43, 15.05, and 7.27%, respectively.
There are many identical functional groups in the FT-IR spectrum (Figure 3a) of BMMB and CuBMMB. The peak appearing at 1519 cm−1 corresponds to the C=N stretching vibration in the benzimidazole ring, the peaks at 1477 and 3050 cm−1 could be identified as the stretching vibration of the benzene ring skeleton [37]. Some peaks correspond to C=C bending modes (1450–1600 cm−1), C–O stretching modes (1200–1250 cm−1), and C–H stretching modes (680–880 cm−1). In particular, the peak at 1350–1400 cm−1 corresponds to the –N=C in BMMB. The peak at 1250–1300 cm−1 corresponds to the N–O–Cu [38], which matches the results of the O 1s spectrum (Figure S3) of CuBMMB. When the CuBMMB powder was subjected to the TG experiment (Figure 3b), the crystalline water (3.8% observed, 4.0% calculated) was all lost in the region of 81–123 °C. The compounds are stable below 160 °C and decompose rapidly after 160 °C. At 401 °C, a large quantity of carbon begins to be converted into CO2. After almost complete decomposition of carbon at 640 °C, the main component of the final product was Cu−O with mass of 16.3% (17% calculated). XPS characterization of CuBMMB was carried out to investigate the composition and state of the complex. The XPS pattern for the complex displays the presence of C, N, O, and Cu elements (Figure 3c), which corresponds to the above-mentioned element mapping results. Cu 2p XPS spectra were shown in Figure 3d, and could be divided into three pairs of characteristic peaks of Cu2+, Cu+/Cu, and satellites [36,39]. A pair of peaks appearing at 934.39 and 954.25 eV accompanied by a pair of satellite peaks at 943 eV (Cu 2p3/2) and 963 eV (Cu 2p1/2) could be attributed to Cu2+ [36]. Last pair of peaks at 932.42 eV (Cu 2p3/2) and 952.3 eV (Cu 2p1/2) were assigned to Cu+/Cu [40,41]. Interestingly, there is a tendency for Cu2+ to be reduced to Cu+/Cu during the synthesis of the complexes, the ratio of Cu2+ for CuBMMB (50.22%) is almost the same with Cu+/Cu; this phenomenon is consistent with the study by Luo et al. [28]. Furthermore, the high magnification O 1s, C 1s, and N 1s spectra were presented in Figure S3. N 1s spectra are deconvoluted to three peaks of pyridinic-N (398.54 eV), pyrrolic-N (399.86 eV), and Cu−N (405.73 eV) [42], and the contents are 51.9, 30.45, and 17.65%, respectively. The O 1s spectra are fitted to four peaks of H−O (533.3 eV), Cu−O (Cu−O, 532 eV), N−O (532.7 eV), and C−O (531.1 eV) [43,44], and the contents are 11.09, 38.87, 40.98, and 9.06%, respectively; these analyzed bonds correspond to the results of the IR spectra. The relative percentage content regarding each component of multiple nitrogen and oxygen functionalities was listed in Table S1. Complexes containing N- and O- functional groups usually have better wettability and electrical conductivity, which is more conducive to electrocatalytic reactions.
A series of electrochemical tests on the NRA properties of CuBMMB were carried out on 660E electrochemical workstation. Cyclic voltammograms (CV), linear sweep voltammetry (LSV), and chronoamperometry (I-t) curves were evaluated via a three-electrode configuration in an H-type electrolytic cell. The CuBMMB, saturated calomel electrode (SCE), and platinum (Pt) foil acted as working electrode, reference electrode, and counter electrode, respectively. The 1 M potassium hydroxide solution (KOH, 40 mL) was respectively added to the anode and cathode chambers, and KNO3 (200 ppm NO3-N) was added to the cathode chamber. The transformation of NO3 to NH3 was investigated over a reaction time of 5 h via the potentiostatic method with different potentials. The reductive products (NO3-N, NO2-N, NH4+-N) were determined by colorimetric methods, and the calibration curves were depicted in Figure S4. As shown in the Cu coordination environment (Figure 1), the coordinated NO3 and the NO3 on the surface of the electrode material could be converted to NO2 with the electrochemical reaction. The total nitrogen content in the solution could gradually increase with time, and reach 208.7 ppm after 5 h (Table S2). In view of this situation, we use the changed NO3 concentration when calculating the NO3 removal efficiency, Faradaic efficiency, and selectivity toward NH3. As shown in Figure 4a, the LSV curves of CuBMMB showed a rapid increase in current density with the presence of NO3, indicating that CuBMMB has the activity to catalyze the NO3 reduction. Moreover, NRA performance of carbon cloth was investigated (Figures S5 and S6), and showed that carbon cloth exhibited significant lower efficiency in catalyzing the NO3 reduction than CuBMMB in the potential window. Furthermore, the NO3 conversion rate and FE were significantly lower than CuBMMB. Therefore, it can be concluded that the carbon cloth has little effect on the NRA performance of CuBMMB. The NRA performance of CuBMMB at different potentials (−0.63, −0.53, −0.43, and −0.33 V vs. RHE) was investigated. As displayed in Figure 4b, the nitrate removal efficiency gradually increased with the improvement in potential and achieved the maximum value of 89% at −0.63 V. However, Faradaic efficiency toward NH3 production is lowest at 61.4%, which was mainly due to the competing HER, as it played a dominant role at high potential. At −0.53 V, Faradaic efficiency was around 90%, NO3 removal efficiency was 73.4%, NH4+ selectivity and ammonia yield rate were 96.5% and 77 μmol h−1 cm−2, with almost no NO2 in the products (Figure 4c). In addition to the colorimetric methods, the concentration of products was verified using ion chromatography (Figures S7–S9). The calculated concentrations from ion chromatography were almost the same as the colorimetric method (Table S3). As the reaction proceeds, the concentration of NH4+ gradually increased, while the concentration of NO3 decreased, suggesting the reaction of NO3 → NH4+ (Figure S10). After 5 h of reaction, the conversion rate of NO3 became flat, and the growth rate of NH4+ also decreased, implying that the reaction on the surface of CuBMMB reached equilibrium. The results of time-dependent product contents showed that more than 70% of NO3 had converted to NH4+ and NO2, where the NH3 selectivity was 96.5%. The catalytic performance of CuBMMB at different concentrations (50, 100, 200, and 400 ppm) was also investigated, and the LSV curves for each condition were shown in Figure 4d. There is no doubt that the current density increased faster as the nitrate concentration increased. As shown in Figure 4e, the FE of NH4+ production reached a maximum value (97%) at 100 ppm NO3-N, and the 400 ppm NO3-N reached the minimum value (71.7%). Moreover, the impact on selectivity of NH3 production was studied with all NO3-N concentrations, and the values reached the maximum value (96.5%) at 200 ppm NO3-N. Stability is a crucial factor for electrocatalysts in practical applications. As presented in Figure 4f, FE and NO3 conversion rate showed a significant decrease after three consecutive cycles (each cycle lasts 5 h). Due to the coordination of Cu with nitrate, which leads to an unstable material surface during the electrochemical process, the material surface inevitably reconfigures as the reaction time increases, which leads to the exposure of more active sites and acceleration of the hydrogen evolution reaction; therefore, further destroying the crystal structure of the complex. The number of active sites has an important influence on the activity of electrocatalyst. As a result, electrochemical active surface areas (ECSA) were derived from double-layer capacitances (Cdl), which was determined by the functional relationship between capacitive currents and scan rates (Figure S11). The results of linear fitting indicated that the specific capacitance of CuBMMB was 0.28 mF cm−2. Tafel plot obtained from the LSV curves in 1 M KOH electrolyte with 200 ppm NO3-N was shown in Figure S12, with CuBMMB exhibiting a modest Tafel slope of 228.9 mV dec−1.
Characterization of the composition and structure of the CuBMMB after the NRA were conducted by SEM, XRD, and XPS in Figure 5. According to the SEM image (Figure 5a) of the CuBMMB after the reaction, the surface of the material was clearly reconstructed. The small, clear holes on the surface before the reaction have turned into craters, and the spread-out layers have gathered into blocks. Moreover, the XRD image (Figure 5b) proved that the crystal structure of CuBMMB was destroyed. The FE and NO3 conversion rate dropped significantly after three cycles, which was mainly due to the competing HER. By comparing Figure 2c with Figure 5a, we can infer that the disruption of the CuBMMB structure exposes a large number of active sites and breaks the equilibrium between the HER and NRA reactions, which significantly facilitates the HER reaction. Therefore, an increasing number of bubbles exist on the material surface with the reaction. With regard to the high-resolution XPS spectra of Cu 2p (Figure 5c), compared to the initial complex, the Cu 2p1/2 spectra exhibit a slight negative shift, demonstrating that the copper ions in the complex become more stable after NRA, with an increase in the content of Cu2+ (50.22 → 55.48%) [45]. However, two peaks (292.5 and 295.5 eV) at high binding energies in the spectra of C 1s after NRA (Figure S13) indicate a disruption in the material structure, which is consistent with the XRD and SEM characterization findings. The rate-determining step reaction (NO3(ad) → NO2(ad)) first adsorbs NO3(aq) on the surface of the electrocatalytic material to form NO3(ad), and then obtains electrons to generate NO2(ad) [18]. The electron transfer step is very fast, and the key point is the adsorption of NO3. Notably, benefiting from the straightforward and efficient design approach, the adsorption capacity of Cu ions for NO3 is greatly increased, thus accelerating the rate-determining step NO3 + H2O + 2e → NO2 + 2OH and converting almost all of the NO2 adsorbed to NH4+ (NO2+ 8H++ 6e → NH4++ 2H2O) (Figure 5d). The electrocatalytic activity of CuBMMB (initial coordination polymer) is comparable to the MOF-loaded metal particle catalysts (Table S4).

3. Materials and Methods

3.1. Synthesis of CuBMMB Crystals and Powder

All solvents and reagents for the synthesis were commercially available and used as received. The CuBMMB crystals were synthesized by the solvent-diffusing method. First, 6 mL methanol (CH3OH) solution of Cu(NO3)2·xH2O (0.1 mmol) was added dropwise at the bottom of a test tube. Then, the solution of CH3OH and CHCl3 (1:1.8 mL) was carefully added as a buffer layer. Finally, a solution of 1,4-bis{[2-(2’-pyridyl)benzimidazolyl]methyl}benzene (BMMB) in CHCl3 (6 mL) was layered onto the buffer layer, and the test tube was then arranged at an undisturbed place. After ~3 weeks, green single crystals appeared on the wall of the test tube.
The Cu-MOF powder was synthesized as follows. The ligand BMMB and Cu(NO3)2·xH2O were dissolved in CHCl3 and CH3OH to obtain a clear solution, respectively. Subsequently, the two clear liquids were mixed and stirred at room temperature for 12 h. Then, the green powder precipitated and thereafter was centrifuged and rinsed with deionized water and ethanol several times. Finally, the powder was evacuated at 60 °C under vacuum for 6 h.

3.2. Characterization

X-ray diffraction (XRD) was recorded with Bruker D8 Advance powder XRD system (Billerica, MA, USA) using Cu Kα radiation (λ = 1.54 Å). IR spectra were recorded on Thermo-Nicolet 6700 (Thermo Fischer Scientific, Kandel, Germany) with air as a reference. Field emission scanning electron microscopy (FE-SEM, Hitachi SU8010, Tokyo, Japan) has been utilized to obtain morphology information on the samples. X-ray photoelectron spectroscopy (XPS) survey was performed on Thermo Fisher ESCALAB Xi+ instrument and the non-monochromatized Al-Ka X-ray was used. Thermogravimetry analysis was conducted on METTERLER under air atmosphere with the heating rate of 5 °C min−1 in the range of 35–800 °C.

3.3. Electrochemical Measurements

Electrochemical measurements were conducted on a CHI 660E electrochemical workstation (Chinstruments, Shanghai, China) using an H-type electrolytic cell separated by a membrane (Nafion-117). Saturated calomel electrode (SCE), Pt foil electrode, and as-prepared electrode were used as reference electrode, counter electrode, and working electrode, respectively. The 1 M KOH solution (40 mL) with different concentrations of KNO3 was used in the cathode (containing 200 ppm nitrate-N) and anode (containing 0 ppm nitrate-N) compartments. Linear sweep voltammetry (LSV) was performed at a scan rate of 10 mV s−1 to obtain polarization curves from 0.40 to −0.53 V (vs. RHE). The potentiostatic curve tests were performed at different potentials (−0.33, −0.43, −0.53, and −0.63 V vs. RHE) for 5 h with a stirring rate of 300 rpm. The electrochemical double-layer capacitance (Cdl) using the cyclic voltammetry (CV) method was nearly a non-Faradaic potential window (from 0.97 to 1.07 V) at different scan rates of 10, 20, 40, 60, 80, 100 mV s−1. The concentrations of nitrate-N, nitrite-N, and ammonia-N were detected by UV–Vis spectrophotometry (Shimadzu UV-2600i) according to the standard method. Meanwhile, the concentrations of ammonia-N were also confirmed by ion chromatography (Shine CIC-D100). The detailed determination methods are provided in the Supporting Information (SI).

4. Conclusions

In summary, the coordination polymer CuBMMB proved to be a novel NRA electrocatalyst with outstanding activity, good nitrate conversion rate, high NH3 selectivity, and FE of NH3. Under the desirable potential of −0.53 V vs. RHE, CuBMMB in alkaline conditions showed excellent nitrate reduction performance with the NO3 conversion rate of 70%, Faradaic efficiency of up to 90%, and NH3 selectivity of more than 95%. This high-catalytic activity is mainly due to the ingeniously designed copper cores with strong affinity for NO3, which accelerates the transferring rate of adsorbed nitrate on the Cu surface and increases the efficiency of rate-determining step (NO3 → NO2) for the whole catalytic process. Therefore, the transformation of surface-exposed nitrate can be rapidly catalyzed by the Cu active sites, and the high affinity of Cu ion for nitrate greatly promotes the efficiency of nitrate conversion.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/catal12121561/s1, Figure S1: SEM image of the CuBMMB; Figure S2: The EDS images of the CuBMMB; Figure S3: The XPS survey of C 1s, N 1s, and O 1s; Figure S4: Concentration-absorbance calibration curves of (a) ammonia-N, (b) nitrate-N, and (c) nitrite-N; Figure S5: LSV curves in 1 M KOH electrolyte with 200 ppm NO3-N; Figure S6: Faradaic efficiency and nitrate conversion of the CuBMMB and carbon cloth in 1 M KOH electrolyte with 200 ppm NO3-N at −0.53 V (vs. RHE); Figure S7: Ion chromatogram (IC) curves of the time-dependent ammonium production for a series of standard ammonium solutions and CuBMMB at −0.53 V; (b) calibration curve of IC for ammonium concentration; Figure S8: IC curves of the time-dependent nitrate residue for a series of standard nitrate solutions and CuBMMB at −0.53 V; (b) calibration curve of IC for nitrate concentration; Figure S9: IC curves of the time-dependent nitrite production for a series of standard nitrite solutions and CuBMMB at −0.53 V; (b) calibration curve of IC for nitrite concentration; Figure S10: Time-dependent concentration of NO3, NO2, and NH4+ over CuBMMB at −0.53 V vs. RHE; Figure S11: (a) CV curves of the CuBMMB at different scanning rates (10–100 mV s−1); (b) relationship between the scanning rate and the current density of Cu-MOF; Figure S12: Tafel plot of CuBMMB in 1 M KOH electrolyte with 200 ppm NO3-N; Figure S13: The XPS survey of C 1s, N 1s, and O 1s after NRA; Table S1: The element distribution of CuBMMB; Table S2: The value of total nitrogen element at different times; Table S3: The value of ion chromatography and colorimetric methods for the concentration of produced nitrite and ammonium as well as residual nitrate; Table S4: Performance comparison; References [3,5,24,25,29,30,34,35,36] are cited in Supplementary Materials.

Author Contributions

Z.-X.L. conceived the idea, discussed results and revised the manuscript. J.-J.W. analysed crystal data and revised the manuscript. K.-Y.Z. tested electrochemical data, discussed results and prepared the manuscript. X.F. performed syntheses. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by the National Natural Science Foundation of China (21673177, 22171205).

Data Availability Statement

Data openly available in a public repository, which can be free downloaded in the website of mdpi system.

Acknowledgments

K.Y.Z. thanks Feng Du for his assistance in the measurements and discussion of NRA.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Chen, T.; Li, Y.; Li, L.; Zhao, Y.; Shi, S.; Jiang, R.; Ma, H. Cu Modified Pt Nanoflowers with Preferential (100) Surfaces for Selective Electroreduction of Nitrate. Catalysts 2019, 9, 536. [Google Scholar] [CrossRef] [Green Version]
  2. Shi, Z.; Wang, F.; Xiao, Q.; Yu, S.; Ji, X. Selective and Efficient Reduction of Nitrate to Gaseous Nitrogen from Drinking Water Source by UV/Oxalic Acid/Ferric Iron Systems: Effectiveness and Mechanisms. Catalysts 2022, 12, 348. [Google Scholar] [CrossRef]
  3. Eaton, A.D.; Clesceri, L.S.; Greenberg, A.E.; Franson, M.A.H. Standard methods for the examination of water and wastewater. Am. J. Public Health Nations Health 1966, 56, 387–388. [Google Scholar]
  4. Sanchis, I.; Rodriguez, J.J.; Mohedano, A.F.; Diaz, E. Activity and Stability of Pd Bimetallic Catalysts for Catalytic Nitrate Reduction. Catalysts 2022, 12, 729. [Google Scholar] [CrossRef]
  5. Wang, Y.; Yu, Y.; Jia, R.; Zhang, C.; Zhang, B. Electrochemical Synthesis of Nitric Acid from Air and Ammonia through Waste Utilization. Natl. Sci. Rev. 2019, 6, 730. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  6. Zhang, R.; Liu, H.; Jiang, W.; Liu, W. Ordered Mesoporous nZVI/Zr-Ce-SBA-15 Catalysts Used for Nitrate Reduction: Synthesis, Optimization and Mechanism. Catalysts 2022, 12, 797. [Google Scholar] [CrossRef]
  7. Zou, X.; Chen, C.; Wang, C.; Zhang, Q.; Yu, Z.; Wu, H.; Zhuo, C.; Zhang, T.C. Combining electrochemical nitrate reduction and anammox for treatment of nitrate-rich wastewater: A short review. Sci. Total Environ. 2021, 800, 149645. [Google Scholar] [CrossRef]
  8. Kim, D.G.; Hwang, Y.H.; Shin, H.S.; Ko, S.O. Kinetics of nitrate adsorption and reduction by nano-scale zero valent iron (NZVI): Effect of ionic strength and initial pH. KSCE J. Civ. Eng. 2015, 20, 175–187. [Google Scholar] [CrossRef]
  9. Labarca, F.; Bórquez, R. Comparative study of nanofiltration and ion exchange for nitrate reduction in the presence of chloride and iron in groundwater. Sci. Total Environ. 2020, 723, 137809. [Google Scholar] [CrossRef]
  10. Sheydaei, M.; Ayoubi-Feiz, B. Nitrate reduction through the visible-light photoelectrocatalysis and photoelectrocatalysis/reverse osmosis processes: Assessment of graphene/Ag/N-TiO2 nanocomposite. J. Water Process Eng. 2021, 39, 101856. [Google Scholar] [CrossRef]
  11. Zhang, W.; Bai, Y.; Ruan, X.; Yin, L. The biological denitrification coupled with chemical reduction for groundwater nitrate remediation via using SCCMs as carbon source. Chemosphere 2019, 234, 89–97. [Google Scholar] [CrossRef] [PubMed]
  12. Tokazhanov, G.; Ramazanova, E.; Hamid, S.; Bae, S.; Lee, W. Advances in the catalytic reduction of nitrate by metallic catalysts for high efficiency and N2 selectivity: A review. Chem. Eng. J. 2020, 384, 123252. [Google Scholar] [CrossRef]
  13. Shi, M.M.; Bao, D.; Wulan, B.R.; Li, Y.H.; Zhang, Y.F.; Yan, J.M.; Jiang, Q. Au Sub-Nanoclusters on TiO2 toward Highly Efficient and Selective Electrocatalyst for N2 Conversion to NH3 at Ambient Conditions. Adv. Mater. 2017, 29, 1606550. [Google Scholar] [CrossRef] [PubMed]
  14. Luo, Y.; Chen, G.F.; Ding, L.; Chen, X.; Ding, L.X.; Wang, H. Efficient Electrocatalytic N2 Fixation with MXene under Ambient Conditions. Joule 2019, 3, 279–289. [Google Scholar] [CrossRef] [Green Version]
  15. Yu, X.; Han, P.; Wei, Z.; Huang, L.; Gu, Z.; Peng, S.; Ma, J.; Zheng, G. Boron-Doped Graphene for Electrocatalytic N2 Reduction. Joule 2018, 2, 1610–1622. [Google Scholar] [CrossRef] [Green Version]
  16. Lv, C.; Yan, C.; Chen, G.; Ding, Y.; Sun, J.; Zhou, Y.; Yu, G. An Amorphous Noble-Metal-Free Electrocatalyst that Enables Nitrogen Fixation under Ambient Conditions. Angew. Chem. Int. Ed. 2018, 130, 6181–6184. [Google Scholar] [CrossRef]
  17. Jung, W.; Hwang, Y.J. Material strategies in the electrochemical nitrate reduction reaction to ammonia production. Mater. Chem. Front. 2021, 5, 6803–6823. [Google Scholar] [CrossRef]
  18. Wang, J.; Feng, T.; Chen, J.; Ramalingam, V.; Li, Z.; Kabtamu, D.M.; He, J.-H.; Fang, X. Electrocatalytic nitrate/nitrite reduction to ammonia synthesis using metal nanocatalysts and bio-inspired metalloenzymes. Nano Energy 2021, 86, 106088. [Google Scholar] [CrossRef]
  19. Garcia-Segura, S.; Lanzarini-Lopes, M.; Hristovski, K.; Westerhoff, P. Electrocatalytic reduction of nitrate: Fundamentals to full-scale water treatment applications. Appl. Catal. B Environ. 2018, 236, 546–568. [Google Scholar] [CrossRef]
  20. Su, J.F.; Ruzybayev, I.; Shah, I.; Huang, C.P. The electrochemical reduction of nitrate over micro-architectured metal electrodes with stainless steel scaffold. Appl. Catal. B Environ. 2016, 180, 199–209. [Google Scholar] [CrossRef]
  21. Huang, J.; Zhang, Q.; Ding, J.; Zhai, Y. Fe–N–C single atom catalysts for the electrochemical conversion of carbon, nitrogen and oxygen elements. Mater. Rep. Energy 2022, 2, 100141. [Google Scholar] [CrossRef]
  22. Katsounaros, I.; Kyriacou, G. Influence of nitrate concentration on its electrochemical reduction on tin cathode: Identification of reaction intermediates. Electrochim. Acta 2008, 53, 5477–5484. [Google Scholar] [CrossRef]
  23. Polatides, C.; Kyriacou, G. Electrochemical reduction of nitrate ion on various cathodes-reaction kinetics on bronze cathode. J. Appl. Electrochem. 2005, 35, 421–427. [Google Scholar] [CrossRef]
  24. Jiang, M.; Su, J.; Song, X.; Zhang, P.; Zhu, M.; Qin, L.; Jin, Z. Interfacial Reduction Nucleation of Noble Metal Nanodots on Redox-Active Metal–Organic Frameworks for High-Efficiency Electrocatalytic Conversion of Nitrate to Ammonia. Nano Lett. 2022, 22, 2529–2537. [Google Scholar] [CrossRef] [PubMed]
  25. Qin, J.; Wu, K.; Chen, L.; Wang, X.; Zhao, Q.; Liu, B.; Ye, Z. Achieving high selectivity for nitrate electrochemical reduction to ammonia over MOF-supported RuxOy clusters. J. Mater. Chem. A 2022, 10, 3963–3969. [Google Scholar] [CrossRef]
  26. Li, J.; Zhan, G.; Yang, J.; Quan, F.; Mao, C.; Liu, Y.; Wang, B.; Lei, F.; Li, L.; Chan, A.W.M.; et al. Efficient Ammonia Electrosynthesis from Nitrate on Strained Ruthenium Nanoclusters. J. Am. Chem. Soc. 2020, 142, 7036–7046. [Google Scholar] [CrossRef]
  27. Chen, T.; Li, H.; Ma, H.; Koper, M.T. Surface Modification of Pt(100) for Electrocatalytic Nitrate Reduction to Dinitrogen in Alkaline Solution. Langmuir 2015, 31, 3277–3281. [Google Scholar] [CrossRef]
  28. Hasnat, M.A.; Ahamad, N.; Uddin, S.N.; Mohamed, N. Silver modified platinum surface/H+ conducting Nafion membrane for cathodic reduction of nitrate ions. Appl. Surf. Sci. 2012, 258, 3309–3314. [Google Scholar] [CrossRef]
  29. Wang, Y.; Xu, A.; Wang, Z.; Huang, L.; Li, J.; Li, F.; Wicks, J.; Luo, M.; Nam, D.-H.; Tan, C.-S. Enhanced Nitrate-to-Ammonia Activity on Copper–Nickel Alloys via Tuning of Intermediate Adsorption. J. Am. Chem. Soc. 2020, 142, 5702–5708. [Google Scholar] [CrossRef]
  30. Li, J.; Gao, J.; Feng, T.; Zhang, H.; Liu, D.; Zhang, C.; Guo, C. Effect of supporting matrixes on performance of copper catalysts in electrochemical nitrate reduction to ammonia. J. Power Sources 2021, 511, 230463. [Google Scholar] [CrossRef]
  31. Dima, G.E.; De Vooys, A.C.A.; Koper, M.T.M. Electrocatalytic reduction of nitrate at low concentration on coinage and transition-metal electrodes in acid solutions. J. Electroanal. Chem. 2003, 554–555, 15–23. [Google Scholar] [CrossRef]
  32. Wu, T.; Kong, X.; Tong, S.; Chen, Y.; Liu, J.; Tang, Y.; Yang, X.; Chen, Y.; Wan, P. Self-supported Cu nanosheets derived from CuCl-CuO for highly efficient electrochemical degradation of NO3. Appl. Surf. Sci. 2019, 489, 321–329. [Google Scholar] [CrossRef]
  33. Kwon, H.Y.; Braley, S.E.; Madriaga, J.P.; Smith, J.M.; Jakubikova, E. Electrocatalytic nitrate reduction with Co-based catalysts: Comparison of DIM, TIM and cyclam ligands. Dalton Trans. 2021, 50, 12324–12331. [Google Scholar] [CrossRef]
  34. Gao, Z.; Lai, Y.; Tao, Y.; Xiao, L.; Zhang, L.; Luo, F. Constructing Well-Defined and Robust Th-MOF-Supported Single-Site Copper for Production and Storage of Ammonia from Electroreduction of Nitrate. ACS Cent. Sci. 2021, 7, 1066–1072. [Google Scholar] [CrossRef]
  35. Zhu, X.; Huang, H.; Zhang, H.; Zhang, Y.; Shi, P.; Qu, K.; Cheng, S.-B.; Wang, A.-L.; Lu, Q. Filling Mesopores of Conductive Metal–Organic Frameworks with Cu Clusters for Selective Nitrate Reduction to Ammonia. ACS Appl. Mater. Interfaces 2022, 14, 32176–32182. [Google Scholar] [CrossRef] [PubMed]
  36. Xu, Y.T.; Xie, M.Y.; Zhong, H.; Cao, Y. In Situ Clustering of Single-Atom Copper Precatalysts in a Metal-Organic Framework for Efficient Electrocatalytic Nitrate-to-Ammonia Reduction. ACS Catal. 2022, 12, 8698–8706. [Google Scholar] [CrossRef]
  37. Lee, S.Y.; Jung, H.; Kim, N.K.; Oh, H.S.; Min, B.K.; Hwang, Y.J. Mixed Copper States in Anodized Cu Electrocatalyst for Stable and Selective Ethylene Production from CO2 Reduction. J. Am. Chem. Soc. 2018, 140, 8681–8689. [Google Scholar] [CrossRef] [PubMed]
  38. Mao, H.; Xu, J.; Hu, Y.; Huang, Y.; Song, Y. The effect of high external pressure on the structure and stability of MOF α-Mg3(HCOO)6 probed by in situ Raman and FT-IR spectroscopy. J. Mater. Chem. A 2015, 3, 11976–11984. [Google Scholar] [CrossRef]
  39. Zhang, D.F.; Zhang, H.; Guo, L.; Zheng, K.; Han, X.D.; Zhang, Z. Delicate control of crystallographic facet-oriented Cu2O nanocrystals and the correlated adsorption ability. J. Mater. Chem. 2009, 19, 5220–5225. [Google Scholar] [CrossRef]
  40. Yang, L.; Li, J.; Du, F.; Gao, J.; Liu, H.; Huang, S.; Zhang, H.; Li, C.; Guo, C. Interface engineering cerium-doped copper nanocrystal for efficient electrochemical nitrate-to-ammonia production. Electrochim. Acta 2022, 411, 140095. [Google Scholar] [CrossRef]
  41. Xu, H.; Wu, J.; Luo, W.; Li, Q.; Zhang, W.; Yang, J. Dendritic Cell-Inspired Designed Architectures toward Highly Efficient Electrocatalysts for Nitrate Reduction Reaction. Small 2020, 16, 2001775. [Google Scholar] [CrossRef] [PubMed]
  42. Tolman, C.A.; Riggs, W.M.; Linn, W.J.; King, C.M.; Wendt, R.C. Electron spectroscopy for chemical analysis of nickel compounds. Inorg. Chem. 1973, 12, 2770–2778. [Google Scholar] [CrossRef]
  43. Nangan, S.; Ding, Y.; Alhakemy, A.Z.; Liu, Y.; Wen, Z. Hybrid alkali-acid urea-nitrate fuel cell for degrading nitrogen-rich wastewater. Appl. Catal. B: Environ. 2021, 286, 119892. [Google Scholar] [CrossRef]
  44. Nefedov, V.I.; Salyn, Y.V.; Leonhardt, G.; Scheibe, R. A comparison of different spectrometers and charge corrections used in X-ray photoelectron spectroscopy. J. Electron Spectrosc. Relat. Phenom. 1977, 10, 121–124. [Google Scholar] [CrossRef]
  45. Du, F.; Li, J.; Wang, C.; Yao, J.; Tan, Z.; Yao, Z.; Li, C.; Guo, C. Active sites-rich layered double hydroxide for nitrate-to-ammonia production with high selectivity and stability. Chem. Eng. J. 2022, 434, 14641. [Google Scholar] [CrossRef]
Figure 1. The coordination environment of Cu ion and bicopper core in CuBMMB.
Figure 1. The coordination environment of Cu ion and bicopper core in CuBMMB.
Catalysts 12 01561 g001
Figure 2. Structure, morphology, and composition analyses: (a) XRD pattern; (b,c) SEM image of CuBMMB; (d) SEM elemental mapping of C, Cu, N, and O in CuBMMB.
Figure 2. Structure, morphology, and composition analyses: (a) XRD pattern; (b,c) SEM image of CuBMMB; (d) SEM elemental mapping of C, Cu, N, and O in CuBMMB.
Catalysts 12 01561 g002
Figure 3. (a) FT-IR; (b) TG spectra of CuBMMB; and XPS survey of (c) CuBMMB and (d) Cu 2p.
Figure 3. (a) FT-IR; (b) TG spectra of CuBMMB; and XPS survey of (c) CuBMMB and (d) Cu 2p.
Catalysts 12 01561 g003
Figure 4. NRA electrocatalytic performance. (a) LSV curves of CuBMMB in 1 M KOH electrolyte with and without NO3, respectively; (b) potential−dependent FE and selectivity of NH3 over CuBMMB in 1 M KOH electrolyte with 200 ppm NO3; (c) NH3 yield rate and NH3 selectivity of CuBMMB electrodes under different potentials; (d) LSV curves of CuBMMB tested with different NO3-N concentrations; (e) FE and selectivity of NH3 tested at −0.53 V vs. RHE over CuBMMB in 1 M KOH electrolyte with 50, 100, 200, and 400 ppm NO3-N concentrations; (f) the consecutive recycling test of CuBMMB at −0.53 V vs. RHE 1 M KOH electrolyte with 200 ppm NO3.
Figure 4. NRA electrocatalytic performance. (a) LSV curves of CuBMMB in 1 M KOH electrolyte with and without NO3, respectively; (b) potential−dependent FE and selectivity of NH3 over CuBMMB in 1 M KOH electrolyte with 200 ppm NO3; (c) NH3 yield rate and NH3 selectivity of CuBMMB electrodes under different potentials; (d) LSV curves of CuBMMB tested with different NO3-N concentrations; (e) FE and selectivity of NH3 tested at −0.53 V vs. RHE over CuBMMB in 1 M KOH electrolyte with 50, 100, 200, and 400 ppm NO3-N concentrations; (f) the consecutive recycling test of CuBMMB at −0.53 V vs. RHE 1 M KOH electrolyte with 200 ppm NO3.
Catalysts 12 01561 g004
Figure 5. Structure, morphology, and composition analysis after NRA; (a) SEM; (b) XRD; (c) and XPS spectra of CuBMMB; (d) proposed electrochemical process involved in the nitrate reduction in CuBMMB.
Figure 5. Structure, morphology, and composition analysis after NRA; (a) SEM; (b) XRD; (c) and XPS spectra of CuBMMB; (d) proposed electrochemical process involved in the nitrate reduction in CuBMMB.
Catalysts 12 01561 g005aCatalysts 12 01561 g005b
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Zeng, K.-Y.; Wang, J.-J.; Fang, X.; Li, Z.-X. One Bicopper Complex with Good Affinity to Nitrate for Highly Selective Electrocatalytic Nitrate Reduction to Ammonia. Catalysts 2022, 12, 1561. https://doi.org/10.3390/catal12121561

AMA Style

Zeng K-Y, Wang J-J, Fang X, Li Z-X. One Bicopper Complex with Good Affinity to Nitrate for Highly Selective Electrocatalytic Nitrate Reduction to Ammonia. Catalysts. 2022; 12(12):1561. https://doi.org/10.3390/catal12121561

Chicago/Turabian Style

Zeng, Kang-Yu, Jun-Jie Wang, Xiang Fang, and Zuo-Xi Li. 2022. "One Bicopper Complex with Good Affinity to Nitrate for Highly Selective Electrocatalytic Nitrate Reduction to Ammonia" Catalysts 12, no. 12: 1561. https://doi.org/10.3390/catal12121561

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