Next Article in Journal
Measurements of Stray Magnetic Fields of Fe-Rich Amorphous Microwires Using a Scanning GMI Magnetometer
Next Article in Special Issue
High-Entropy Materials: Features for Lithium–Sulfur Battery Applications
Previous Article in Journal
Reductions in the Laser Welding Deformation of STS304 Cylindrical Structure Using the Pre-Stress Method
Previous Article in Special Issue
Brief Introduction on Manufacturing and Characterization of Metallic Electrode and Corresponding Modified Materials
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Boosting Electrocatalytic Reduction of Nitrate to Ammonia over Co3O4 Nanosheets with Oxygen Vacancies

1
School of Metallurgy and Environment, Central South University, Changsha 410083, China
2
Chinese National Engineering Research Center for Control & Treatment of Heavy Metal Pollution, Changsha 410083, China
3
State Key Laboratory of Advanced Metallurgy for Non-Ferrous Metals, Changsha 410083, China
*
Authors to whom correspondence should be addressed.
Metals 2023, 13(4), 799; https://doi.org/10.3390/met13040799
Submission received: 16 March 2023 / Revised: 8 April 2023 / Accepted: 17 April 2023 / Published: 18 April 2023
(This article belongs to the Special Issue Manufacturing and Characterization of Metallic Electrode Materials)

Abstract

:
Electrocatalytic nitrate reduction into ammonia is promising for its restricted activity and selectivity in wastewater treatment, however, it remains challenging. In this work, Co3O4 nanosheet electrodes with rich oxygen vacancies (OVs) (Co3O4−x/NF) are prepared and then applied as efficient catalysts for selective electrocatalytic reduction of nitrate to ammonia. The resulting Co3O4−x/NF electrodes exhibit high NO3-N removal efficiency and NH4+-N selectivity, at 93.7% and 85.4%, respectively. X-Ray photoelectron spectroscopy (XPS) and electron paramagnetic resonance spectra (EPR) results clearly reveal the formation of OVs in Co3O4−x/NF. The electrochemical characterization results confirm that OVs can effectively improve electron transfer as well as the electrochemically active area. The Co2+/Co3+ ratio of Co3O4−x/NF increases after the electrocatalytic reduction of nitrate, highlighting the crucial role played by Co2+ in mediating ammonia production via the Co2+/Co3+ cycle. These findings offer valuable guidelines for the development of more efficient and sustainable approaches for nitrate-contaminated wastewater treatment and ammonia synthesis.

1. Introduction

Acid rain deposition of nitrogen oxides from the combustion of nitrogen-containing fuels, bacterial decomposition of nitrogen-containing fertilizers, and discharge of nitrate-containing domestic and industrial wastewater are all highly likely to lead to elevated nitrate concentrations in groundwater [1,2,3]. Nitrate (NO3) levels must be restricted to ensure a balanced ecosystem, otherwise, they can lead to environment pollution (e.g., eutrophication) and threaten the health of human beings (e.g., blue baby syndrome) [4,5,6]. Thus, significant efforts to remove NO3 have led to the development of a number of concepts and techniques, such as ion exchange, reverse osmosis, biological denitrification, electrodialysis, etc. [7,8,9,10,11,12,13,14]. Although these methods have proven to be efficient technologies for treating certain conditions, they may require significant capital investment. For instance, chemical or biological treatment processes may necessitate the use of reducing agents such as H2 or acetic acid that serve as electron donors [15,16]. Recognizing the significance of nitrogen (N) is crucial, as it serves as an essential nutrient for plant growth and plays a key role in the production of industrial products [17,18,19]. As such, recovery or extraction of N from NO3 is greatly beneficial for both sustainable agriculture and industrial production. Unfortunately, current methods for treating waste water result in the conversion of NO3 to nitrogen gas, resulting in the loss of valuable nitrogen resources. Therefore, developing methods for recovery of N resources from NO3-contaminated water is critical for both sustainable development and environmental protection.
Currently, the electrocatalytic nitrate reduction reaction (NO3RR) is generally considered an efficient method for remediating wastewater, with benefits including absence of chemical waste byproduct streams and lower capital costs; thus, it has attracted increasing attention. Importantly, high-value-added products such as ammonia (NH3) can be the harvesting indication for NO3RR process [11,20,21]. It is well known that the Haber–Bosch process is the primary method of NH3 production; this process, however, requires significant energy (~2% of the world’s energy) and emits a substantial amount of greenhouse gases (~1.5% of global carbon dioxide emissions) [22,23]. As a result, NO3RR may be an alternative approach for producing NH3 through a more sustainable and environmentally friendly manner while helping to solve water pollution challenges.
The electrocatalyst, an essential component of NO3RR, directly impacts the rate of NO3 removal and selectivity of the NH3 yield. Therefore, the rational design and development of high-performance electrocatalysts is crucial for achieving efficient and selective NO3RR. Recently, electrocatalysts made of noble metals (e.g., Ru, Pd, and Pt) have been explored and found to exhibit high performance for NH3 production by NO3RR; however, undesirable capital output hinders their application [24,25,26]. Transition metal-based nanocatalysts represent an effective alternative to noble metal catalysts, and have gained significant attention for NH3 production via NO3RR owing to their low cost and abundance [27,28]. Among them, Co3O4 nanocomposite is considered a promising electrocatalyst due to its impressive catalytic properties [29]. Unfortunately, the widespread use of Co3O4 is limited by its poor electrical conductivity [30]. Recently, oxygen vacancies (OVs) have been demonstrated to modulate its electronic states, significantly altering catalytic activity [31]. For instance, Jia and co-workers found that the introduction of O-vacancies into TiO2 nanotubes could significantly enhance the performance of electrodes used for synthesis of NH3 from NO3 through electrochemical reduction [3]. Meanwhile, previous studies have demonstrated that the embedding of OVs affects the proportion of Co2+ and Co3+ in Co3O4 [32]. Notably, it has been reported that Co2+ can serve as an electron donor for NO3 reduction during the electrochemical reaction process. Therefore, it is reasonable to expect introduction of OVs into Co3O4 nanostructures with given facets to be a promising electrocatalyst for NH3 production via NO3RR.
In this work, OV-rich Co3O4 nanosheets grown on the surface of nickel foam (Co3O4−x/NF) were synthesized via NO3 electroreduction for use in NH3 production. X-Ray photoelectron spectroscopy (XPS) and electron paramagnetic resonance spectra (EPR) confirmed that abundant OVs were present in the Co3O4−x/NF. Benefiting from the improved electronic structure and Co2+/Co3+ ratio of Co3O4 after the introduction of OVs, the electrocatalyst exhibited high activity vs. Ag/AgCl at −1.2 V, with 93.7% NO3-N removal efficiency and 85.4% NH4+-N selectivity. When comparing the states before and after the reaction, a significant change in the Co2+/Co3+ ratio was found, indicating that NH3 production from NO3RR was mediated by the Co2+/Co3+ cycle. This study is believed to be a feasible reference that can help to realize utilization of NO3 wastewater resources.

2. Materials and Methods

2.1. Chemicals and Reagents

Analytical-grade NaNO3, NaNO2, (NH4)2SO4, and Na2SO4 were purchased from Kermel Chemical Reagent Co. Ltd. (Tianjin, China) and used to prepare the stock solution. All solutions were prepared using ultrapure water. Nickel foam (NF) purchased from Beijing Pengda Hengtai Technology Co., Ltd. (Beijing, China) was employed as the supporter for growing Co3O4 and Co3O4−x. In order to remove the impurities on the surface of the NF, it was rinsed several times with acetone and 1 M HCl, then again with ultrapure water.

2.2. Preparation of Electrodes

A piece of NF (3 cm × 2 cm × 0.1 cm) was first cleaned several times with 1 M HCl, acetone, ethanol, and H2O to remove impurities. Next, Co3O4/NF was synthesized through a hydrothermal and annealing method. In the typical processes, 0.4202 g of CoSO4·7H2O and 0.36 g of urea were dissolved in 20 mL of H2O and 20 mL of ethanol, forming a wine-red solution. Afterwards, this solution was transferred to a 100 mL Teflon-lined autoclave and the cleaned NF was simultaneously soaked in the mixture. After standing for 12 h without stirring, the Teflon-lined autoclave was heated at 90 °C for 8 h. A pink lump (2CoCO3·3Co(OH)2·H2O/NF) was obtained and then calcined at 250 °C for 2 h under N2 atmosphere to generate Co3O4 nanosheet arrays on the NF surface (Co3O4/NF). Finally, defect-engineered Co3O4−x/NF was prepared via immersing the Co3O4/NF in 1 M NaBH4 solution for 1 h, then rinsed with deionized water and dried under vacuum at 60 °C for 8 h.

2.3. Electrocatalytic NO3 Reduction

Electrocatalytic NO3 reduction tests were performed in a three-electrode electrochemical cell with an effective volume of 100 mL controlled by an electrochemical workstation (CHI760E, Shanghai Chenhua, Shanghai, China). The resulting Co3O4−x/NF material, Ir-Ru/Ti, and saturated Ag/AgCl were used as the working electrode, counter-electrode, and reference electrode, respectively. The gap between each electrode was set to ~1.5 cm. Next, 80 mL of 50 mM Na2SO4 solution containing NO3 was electrolyzed and stirred at 600 rpm. The cathodic potential was controlled in the range of −0.8~−1.4 V (vs. Ag/AgCl). About 2 mL of reaction solution was taken regularly from the reactor (40 min) to detect the variations in NO3, NO2, and NH4+.

2.4. Analytical Methods

An X-Ray photoelectron spectroscopy (XPS) survey of the spectra of Co 2p was carried out using an Escalab 250Xi (Thermol Scientific, Waltham, MA, USA) with an Al/Mg Kα line as the dual X-Ray source. The surface morphology was detected by Scanning Electron Microscopy (SEM, FEI Quanta 200 CZ, FEI, Eindhoven, The Netherlands). X-Ray diffraction (XRD, Ultima IV, Rigaku, Tokyo, Japan) was used to analyze the crystallinity of NF, Co3O4/NF, and Co3O4−x/NF. Materials Data, Inc. (MDI) Jade 5.0 software was used to analyze the diffraction peaks and crystalline phases using the Joint Committee on Powder Diffraction Standards (JCPDS) database as a reference. The electron paramagnetic resonance (EPR) was investigated on a Bruker A300 (Bruker Corporation, Karlsruhe, Germany). Electrochemical testing of the electrochemical impedance spectrum (EIS), linear sweep voltammetry (LSV), and cyclic voltammetry (CV) were carried out using an electrochemical workstation (CHI760E, Shanghai Chenhua, China) in a three-electrode system.
The concentrations of NO3 and nitrite (NO2) were measured by Dionex ion-exchange chromatography (ICS-900) with a 0.5 mL sample loop, conductivity suppressor (ASRS 300 × 4 mm), analytical column (AS23 4 × 250 mm), and precolumn (AG23 4 × 50 mm). The eluent (1.0 mL min−1) was set to 9.4 mM sodium carbonate (Na2CO3) and1.8 mM sodium bicarbonate (NaHCO3). The ammonia was analyzed by ion chromatography (Dionex ICS-900, Massachusetts, USA) with a 20 μL sample loop, cation self-regenerating suppressor (CSRS ULTRA Ⅱ, 4 mm), guard column (CS 12A 4 × 250 mm), and analytical column (CG12A 4 × 50 mm); 20 mM methanesulfonic acid was adopted as the eluent at the rate of 1.0 mL min−1. The production of NO and N2O from the electrocatalytic NO3 reduction for the Co3O4 electrode was negligible [4]. Herein, NO3 reduction ( R N O 3 ( % ) ) and total nitrogen (in this study, the sum of NO3, NO2, and NH4+) were respectively calculated according to the following equations:
R N O 3 ( % ) = ( NO 3 ) 0 ( NO 3 ) t ( N O 3 ) 0
S N H 4 + ( % ) = ( NH 4 + ) t ( NO 3 ) 0 ( NO 3 ) t
where 0 and t stand for the initial and final states, respectively.

2.5. Calculation of Average Current Efficiency

Based on the amount of NO3 removed and the amount of nitrite (NO2), ammonia, and N2 generated, the average current efficiency was calculated using the following equations:
Q ( NO 2 N ) t = 2 F ( C ( NO 2 ) t V M )
Q ( NH 4 + N ) t = 8 F ( C ( NH 4 + ) t V M )
Q ( N 2 N ) t = 5 F ( ( C ( NO 3 ) 0 ( NO 3 ) t C ( NO 2 ) t C ( NH 4 + ) t ) V M )
Q t = 0 t I   d t
η = [ ( Q ( NO 2 ) t +   Q ( N 2 ) t +   Q ( NH 4 + ) t ] / Q t × 100 %  
where η (%) stands for the average current efficiency, Q t is the total number of electrons at reaction time t (min), Q ( NO 2 ) t , Q ( N 2 ) t , and Q ( NH 4 + ) t (C) are the electrons consumed during nitrate reduction to nitrite, N2, and ammonia at time t, respectively, C ( NO 3 ) 0 (mg N L−1) is the initial concentration of nitrate, ( NO 3 ) t ,   C ( NO 2 ) t , and ( NH 4 + ) t (mg N L−1) are the concentrations of nitrate, nitrite, and ammonia at time t, respectively, V is the volume of the solution (0.08 L), M is the molar mass of N (14,000 mg mol−1), and F is the Faraday’s constant (96,500 C mol−1).

2.6. Electrochemical Active Surface Area Measurement

The electrochemical active surface area (ECSA) was implemented in a three-electrode system at an electrochemical station (CHI760E, Shanghai, China). The working electrode was NF, Co3O4/NF, or Co3O4−x/NF, with an effective geometric surface area of 1 cm−2. Platinum and Ag/AgCl served as the counter and reference electrodes, respectively. The ECSA was obtained according to Equations (8) and (9) [33]:
ECSA = R f S
R f = C d l 60   μ F   cm 2
where Rf is the roughness factor, S was generally equal to the geometric area of the working electrode (1 cm−2), Cdl is the double-layer capacitance, and Cdl was estimated by plotting j (mA cm−2) at −0.02 V vs. Ag/AgCl against the scan rate, where j was acquired by Cyclic Voltammetry (CV) measurement under potential windows of −0.08~0 V vs. Ag/AgCl at a given scan rate (5, 10, 15, 20, 25, and 30 mV s−1) (50 mM Na2SO4).

3. Results and Discussion

3.1. Characterization

In this work, pristine Co3O4 nanosheets supported on nickel foam (Co3O4/NF) were synthesized by hydrothermal and annealing methods. To introduce oxygen vacancies into Co3O4 NWs, the obtained Co3O4/NF was further treated by NaBH4 at room temperature [32]. As shown in Figure 1a–c, as compared to pristine NF, the surface of Co3O4/NF was covered with dense nanosheets. It was worth noting that the nanosheet morphology of the as-prepared Co3O4−x/NF electrode exhibited a hexagonal shape. Meanwhile, Figure 1d shows that the thickness of the Co3O4 and Co3O4−x nanosheets was ~110 nm, indicating that the post-treatment process of NaBH4 did not affect the morphology of Co3O4. The crystal structure of the as-prepared samples was analyzed by X-Ray diffraction (XRD). As shown in Figure 2a, the diffraction peaks of Co3O4/NF fully matched the hexagonal Co3O4 (PDF#43-1003). Meanwhile, the Co3O4−x/NF exhibited the same diffraction patterns, showing that our post-treatment process did not damage the phase structure. These results clearly reveal that Co3O4 and Co3O4−x nanosheets with hexagonal morphology were able to be densely grown on the surface of nickel foam. Usually, Co3O4 is partially reduced to CoO in the strongly reducing environment formed by NaBH4 [32], which in turn confirms the advantage of our method of introducing defects. As shown in Figure 2b, the XPS spectra of the wide scan clearly show the constituent elements of Co3O4/NF and Co3O4−x/NF, namely, Ni 2p, Co 2p, and O 1s, which is in accordance with the above XRD results. In the XPS spectra of Co 2p (Figure 2c), the two fitting peaks at 780.5 and 782.1 eV are attributed to Co3+ and Co2+, respectively [4,34]. Noticeably, the Co2+/Co3+ surface ratio of Co3O4−x/NF was determined to be 1.04, which was higher than that of Co3O4/NF (0.78). The electron paramagnetic resonance (EPR) spectra result shows that the value of Co3O4−x/NF (in g) is 2.0038 (Figure 2d), which was due to the OVs in Co3O4 [35]. Therefore, nickel foam electrodes with surfaces covered by Co3O4 nanosheets with OVs were successfully prepared.

3.2. Electrocatalytic NO3 Reduction

The electrocatalytic activities of the electrodes towards NO3 reduction were analyzed by LSV, EIS, and Cdl. As shown in Figure 3a, the current density observed from −1.0 to −1.2 V vs. Ag/AgCl for Co3O4−x/NF was lower than that for Co3O4/NF, indicating that NO3 reduction was more kinetically favorable at the former electrode. EIS can reveal the electron transfer rate in electrochemical reactions based on the arc radius of the Nyquist plot [36]. After reduction by NaBH4, the size of the arc radius decreased (Figure 3b), confirming that the introduction of OVs can reduce the interface impedance and facilitate electron transfer. Accordingly, Co3O4−x/NF showed higher conductivity and electron transfer properties, which are beneficial to electrochemical NO3 reduction through the charge-transfer and mass-transfer processes [37]. It is well known that a larger Cdl reflects a higher electrochemical active surface area (ECSA). Figure 3c and Figure S1 show that Co3O4−x/NF had a Cdl value of 6.20 mF cm−2, which was 1.4 and 2.1 times higher than that of NF (4.37 mF cm−2) and Co3O4/NF (2.99 mF cm−2), respectively. This increased ECSA can result in enhancement of electrocatalytic active sites, thereby improving electrocatalytic activity for NO3 reduction.
The electrocatalytic reduction of NO3 was carried out with NF, Co3O4/NF, and Co3O4−x/NF electrodes. Figure 4a shows that 72.7% of NO3-N was reduced for Co3O4/NF after 280 min of electrolysis. Moreover, NO3-N removal efficiency reached up to 93.7% with the Co3O4−x/NF cathode. On the other hand, only 37.9% of NO3 was removed by bare NF, illustrating that the catalytic active sites for NO3 reduction were Co3O4 and Co3O4−x rather than NF. Noticeably, the electrocatalytic reduction of NO3 with different materials obeyed the pseudo-first-order kinetic model (Figure S2). The corresponding reaction rate constant of NO3 reduction with NF, Co3O4/NF, and Co3O4−x/NF was 0.0017, 0.004, and 0.01 min−1, respectively. These results indicate that the Co3O4−x/NF cathode exhibited the highest electrocatalytic activity for NO3 reduction.
The products transformed via NO3 reduction on different electrodes were determined comparatively (Figure 4b–d). As shown in Figure 4b, generation of NO2 increased first and then decreased, suggesting that NO2 was an intermediate product. However, NH4+ increased steadily with the reaction time, indicating that it was a final product (Figure 4c). Owing to the faster electron transfer rate and higher ECSA induced by OVs, Co3O4−x/NF achieved a remarkable 90% NH4+-N selectivity in less than 40 min, compared to 0% and 50% for NF and Co3O4/NF, respectively. These results fully confirm that Co3O4−x/NF has outstanding electrocatalytic performance for NO3RR into NH3, making it highly suitable for practical applications.

3.3. Effect of Reaction Parameters on Electrocatalytic NO3 Reduction

Electrochemical redox reactions are heavily dependent on the electrode potential-. Several potentials (−1.0, −1.1, −1.2, −1.3, and −1.4 V vs. Ag/AgCl) were applied to disclose the effect mechanism of cathodic potential on NO3RR. As shown in Figure 5a, NO3 removal efficiency was 35.8%, 80.7%, 93.7%, 99.6%, and 98.9% at −1.0, −1.1, −1.2, −1.3, and −1.4 V, respectively, indicating that reducing the potential was beneficial for NO3 removal, though the effect is not significant below −1.2 V. Meanwhile, the process conformed to the pseudo-first-order kinetic model with R2 ≥0.95, and the value of k increased with decreasing voltage. Noticeably, η (Table 1) showed a volcanic pattern with decreasing potential, with a maximum value of 26.9% at −1.2 V. The products were shown to be NH4+-dominated, and the selectivity was 86.8%, 90.5%, 85.4%, 91.2%, and 91.6 at −1.0, −1.1, −1.2, −1.3, and −1.4 V, respectively. As is well known, OVs can weaken N-O bonding, hindering the formation of byproducts [3]. Generally, the optimal applied potential of Co3O4−x/NF for NO3RR was −1.2 V vs. Ag/AgCl.
Initial concentration is an important parameter for evaluating the application prospects of catalysts, and was accordingly investigated. Electrolysis reactions were performed under several initial NO3 concentration (20–200 mg N L−1). Figure 5c shows that the NO3-N removal efficiency was 94.9%, 93.7%, 94.1%, and 94.9% for 20, 50, 100, and 200 mg N L−1, respectively, at different reaction times, indicating that the Co3O4−x/NF electrode can be used for treating different NO3 concentrations in water. As mentioned in Table 1, η values increased with increasing NO3 concentrations; more specifically, the value of η was 12.3, 26.9, 36.5, and 45.9 at 20, 50, 100, and 200 mg N L−1, which is due to the weakened competition in HER at relatively high nitrate concentrations [13]. NH4+ selectivity was consistently maintained ~85% at different NO3 concentrations, indicating that the Co3O4−x/NF electrode has outstanding potential for ammonia production via NO3RR. Notably, as shown in Table 2, the Co3O4−x/NF electrocatalyst showed competitive NO3-N conversion and high NH4+ selectivity. The conversion of nitrate is a key aspect of wastewater treatment, with a high conversion rate indicating more effective purification. In our study, the Co3O4−x/NF electrode consistently achieved an NO3-N conversion rate over 93.7% across a range of nitrate concentrations, demonstrating its great potential for practical applications. The selectivity of NH4+-N is an important factor for the resourceful disposal of NO3 wastewater, as higher values indicate greater potential for useful applications. Our study found that the Co3O4−x/NF electrode consistently maintained an NH4+-N selectivity of approximately 85% across various nitrate concentrations, which exceeds the values reported in many previous studies [38,39,40]. To sum up, our study highlights the substantial practical application value of Co3O4−x/NF electrodes in the management of NO3-containing wastewater. Their consistently high NO3-N conversion rates and superior NH4+-N selectivity make them a promising candidate for resourceful disposal of these types of effluents.

3.4. Proposed Reaction Mechanism

To gain insights into the selective synthesis of NH3 over Co3O4−x/NF, we conducted an XPS analysis to investigate the valence change of Co. In fact, the electrocatalytic activity of cobalt oxide is closely related to its electronic states [30], and oxygen defect engineering as an effective mean for regulating the electronic state of materials has been generally confirmed by previous studies [31]. As shown in Figure 4, the introduction of OVs can significantly enhance both NH3 production performance as well as the kinetics of the electrochemical reduction of NO3. In addition, it is notable that the XPS results clearly show the Co2+/Co3+ ratio increasing from 0.78 to 1.04 after the introduction of OVs. Therefore, it is reasonable to speculate that Co2+ plays a crucial role in NO3 reduction and NH4+ production. As shown in Figure 6a,b, the ratio of Co2+/Co3+ in Co3O4−x/NF decreased to 0.85 after electrocatalytic reduction. Such changes not only indicate that the electrocatalytic reduction of NO3 is mediated by the Co2+/Co3+ cycle, they show that Co2+ acts as the active site for transferring electrons to NO3 toward ammonia production. As a result, the OVs can regulate the conversion of Co3+ to Co2+, which improves the poor conductivity of Co3O4 and provides more active sites, thereby enhancing the catalytic performance of the electrocatalytic reduction of NO3 into NH3.

4. Conclusions

In summary, Co3O4 nanosheets with OVs supported on nickel foam were successfully synthesized and demonstrated to be an efficient catalyst for NH3 synthesis from NO3 electroreduction. At −1.2 V, 93.7% nitrate removal and 85.4% NH3 production selectivity were obtained vs. Ag/AgCl. Compared with Co3O4/NF, the EIS and Cdl results confirmed that the introduction of OVs enhanced the electron transfer capability and ECSA of the electrode. In addition, the presence of OVs influenced the Co2+/Co3+ ratio due to the reduction of Co3+ into Co2+. The Co2+/Co3+ cycle was identified as a key mediator in the electrocatalytic reduction of NO3 for NH3 production. These results can shed light on the importance of the Co2+/Co3+ ratio in the electrocatalytic synthesis of ammonia, and highlight the potential of oxygen-defective catalysts for developing more efficient and sustainable electrocatalytic processes. The findings of our study suggest that electrochemical denitrification may hold promise for ammonia production; however, further research is necessary in order to maximize the utilization of NH4+ and enable its in situ recovery.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/met13040799/s1, Figure S1: CV curves of NF (a) and Co3O4/NF (b) in the range of −0.06~0.06 V vs. Ag/AgCl; Figure S2: The pseudo-first-order kinetic analyses.

Author Contributions

X.W.: Investigation, Formal analysis, Writing—original draft, Writing—review and editing; Z.L. (Zhigong Liu), T.G., J.T. and Z.L. (Zhizhuo Li): Writing—review and editing; Z.S., F.F. and C.Q.: Formal analysis; F.Y.: Formal analysis, Writing—original draft, Funding acquisition; C.T.: Formal analysis, Writing—original draft, Funding acquisition. All authors have read and agreed to the published version of the manuscript.

Funding

This research was financially supported by the Postdoctoral Innovation Talent Support Program of China Project (BX2021378), the China Postdoctoral Science Foundation (2021M703630), the Natural Science Foundation of Hunan Province (Nos. 2022JJ40622), and the Technology Innovation Guidance Project of Jiangxi Province, China (20212BDH81030).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Bhatnagar, A.; Sillanpää, M. A review of emerging adsorbents for nitrate removal from water. Chem. Eng. J. 2011, 168, 493–504. [Google Scholar] [CrossRef]
  2. Duca, M.; Koper, M.T.M. Powering denitrification: The perspectives of electrocatalytic nitrate reduction. Energy Environ. Sci. 2012, 5, 9726–9742. [Google Scholar] [CrossRef]
  3. Jia, R.; Wang, Y.; Wang, C.; Ling, Y.; Yu, Y.; Zhang, B. Boosting Selective Nitrate Electroreduction to Ammonium by Constructing Oxygen Vacancies in TiO2. ACS Catal. 2020, 10, 3533–3540. [Google Scholar] [CrossRef]
  4. Gao, J.; Jiang, B.; Ni, C.; Qi, Y.; Zhang, Y.; Oturan, N.; Oturan, M.A. Non-precious Co3O4-TiO2/Ti cathode based electrocatalytic nitrate reduction: Preparation, performance and mechanism. Appl. Catal. B Environ. 2019, 254, 391–402. [Google Scholar] [CrossRef]
  5. Zhang, S.; Li, M.; Li, J.; Song, Q.; Liu, X. High-ammonia selective metal–organic framework–derived Co-doped Fe/Fe2O3 catalysts for electrochemical nitrate reduction. Proc. Natl. Acad. Sci. USA 2022, 119, e2115504119. [Google Scholar] [CrossRef] [PubMed]
  6. Min, X.; Wu, X.; Shao, P.; Ren, Z.; Ding, L.; Luo, X. Ultra-high capacity of lanthanum-doped UiO-66 for phosphate capture: Unusual doping of lanthanum by the reduction of coordination number. Chem. Eng. J. 2019, 358, 321–330. [Google Scholar] [CrossRef]
  7. Theerthagiri, J.; Park, J.; Das, H.T.; Rahamathulla, N.; Cardoso, E.S.F.; Murthy, A.P.; Maia, G.; Vo, D.N.; Choi, M.Y. Electrocatalytic conversion of nitrate waste into ammonia: A review. Environ. Chem. Lett. 2022, 20, 2929–2949. [Google Scholar] [CrossRef]
  8. 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]
  9. 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]
  10. Wang, Y.; Zhou, W.; Jia, R.; Yu, Y.; Zhang, B. Unveiling the Activity Origin of a Copper-based Electrocatalyst for Selective Nitrate Reduction to Ammonia. Angew. Chem. Int. Ed. 2020, 59, 5350–5354. [Google Scholar] [CrossRef]
  11. Lv, C.; Zhong, L.; Liu, H.; Fang, Z.; Yan, C.; Chen, M.; Kong, Y.; Lee, C.; Liu, D.; Li, S.; et al. Selective electrocatalytic synthesis of urea with nitrate and carbon dioxide. Nat. Sustain. 2021, 4, 868–876. [Google Scholar] [CrossRef]
  12. Daiyan, R.; Tran-Phu, T.; Kumar, P.; Iputera, K.; Tong, Z.; Leverett, J.; Khan, M.H.A.; Esmailpour, A.A.; Jalili, A.; Lim, M.; et al. Nitrate reduction to ammonium: From CuO defect engineering to waste NOx-to-NH3economic feasibility. Energy Environ. Sci. 2021, 14, 3588–3598. [Google Scholar] [CrossRef]
  13. Chen, F.-Y.; Wu, Z.-Y.; Gupta, S.; Rivera, D.J.; Lambeets, S.V.; Pecaut, S.; Kim, J.Y.T.; Zhu, P.; Finfrock, Y.Z.; Meira, D.M.; et al. Efficient conversion of low-concentration nitrate sources into ammonia on a Ru-dispersed Cu nanowire electrocatalyst. Nat. Nanotechnol. 2022, 17, 759–767. [Google Scholar] [CrossRef]
  14. Gao, J.; Shi, N.; Li, Y.; Jiang, B.; Marhaba, T.; Zhang, W. Electrocatalytic Upcycling of Nitrate Wastewater into an Ammonia Fertilizer via an Electrified Membrane. Environ. Sci. Technol. 2022, 56, 11602–11613. [Google Scholar] [CrossRef]
  15. Chaplin, B.P.; Reinhard, M.; Schneider, W.F.; Schüth, C.; Shapley, J.R.; Strathmann, T.J.; Werth, C.J. Critical Review of Pd-Based Catalytic Treatment of Priority Contaminants in Water. Environ. Sci. Technol. 2012, 46, 3655–3670. [Google Scholar] [CrossRef]
  16. Zhao, H.-P.; Ontiveros-Valencia, A.; Tang, Y.; Kim, B.-O.; VanGinkel, S.; Friese, D.; Overstreet, R.; Smith, J.; Evans, P.; Krajmalnik-Brown, R.; et al. Removal of multiple electron acceptors by pilot-scale, two-stage membrane biofilm reactors. Water Res. 2014, 54, 115–122. [Google Scholar] [CrossRef] [PubMed]
  17. Gruber, N.; Galloway, J.N. An Earth-system perspective of the global nitrogen cycle. Nature 2008, 451, 293–296. [Google Scholar] [CrossRef] [PubMed]
  18. Chauhan, R.; Srivastava, V.C. Electrochemical denitrification of highly contaminated actual nitrate wastewater by Ti/RuO2 anode and iron cathode. Chem. Eng. J. 2020, 386, 122065. [Google Scholar] [CrossRef]
  19. Canfield, D.E.; Glazer, A.N.; Falkowski, P.G. The Evolution and Future of Earth’s Nitrogen Cycle. Science 2010, 330, 192–196. [Google Scholar] [CrossRef]
  20. Li, P.; Jin, Z.; Fang, Z.; Yu, G. A single-site iron catalyst with preoccupied active centers that achieves selective ammonia electrosynthesis from nitrate. Energy Environ. Sci. 2021, 14, 3522–3531. [Google Scholar] [CrossRef]
  21. Zeng, Y.; Priest, C.; Wang, G.; Wu, G. Restoring the Nitrogen Cycle by Electrochemical Reduction of Nitrate: Progress and Prospects. Small Methods 2020, 4, 12. [Google Scholar] [CrossRef]
  22. Wang, L.; Xia, M.; Wang, H.; Huang, K.; Qian, C.; Maravelias, C.T.; Ozin, G.A. Greening Ammonia toward the Solar Ammonia Refinery. Joule 2018, 2, 1055–1074. [Google Scholar] [CrossRef]
  23. Soloveichik, G. Electrochemical synthesis of ammonia as a potential alternative to the Haber–Bosch process. Nat. Catal. 2019, 2, 377–380. [Google Scholar] [CrossRef]
  24. Guo, S.; Heck, K.; Kasiraju, S.; Qian, H.; Zhao, Z.; Grabow, L.C.; Miller, J.T.; Wong, M.S. Insights into Nitrate Reduction over Indium-Decorated Palladium Nanoparticle Catalysts. ACS Catal. 2018, 8, 503–515. [Google Scholar] [CrossRef]
  25. Radjenovic, J.; Sedlak, D.L. Challenges and Opportunities for Electrochemical Processes as Next-Generation Technologies for the Treatment of Contaminated Water. Environ. Sci. Technol. 2015, 49, 11292–11302. [Google Scholar] [CrossRef]
  26. Li, T.; Tang, C.; Guo, H.; Wu, H.; Duan, C.; Wang, H.; Zhang, F.; Cao, Y.; Yang, G.; Zhou, Y. In Situ Growth of Fe2O3 Nanorod Arrays on Carbon Cloth with Rapid Charge Transfer for Efficient Nitrate Electroreduction to Ammonia. ACS Appl. Mater. Interfaces 2022, 14, 49765–49773. [Google Scholar] [CrossRef] [PubMed]
  27. Ghalkhani, M.; Irannejad, N.; Sohouli, E.; Keçili, R.; Hussain, C.M. Environmental applications of nanographitic carbon nitride. In Nanoremediation; Elsevier: Amsterdam, The Netherlands, 2023; pp. 187–227. [Google Scholar] [CrossRef]
  28. Farooq, N.; Luque, R.; Len, T.; Osman, S.M.; Qureshi, A.M.; Nazir, M.A.; Rehman, A.U. Design of SrZr0.1Mn0.4Mo0.4Y0.1O3-δ heterostructured with ZnO as electrolyte material: Structural, optical and electrochemical behavior at low temperatures. Ceram. Int. 2023, 49, 2174–2182. [Google Scholar] [CrossRef]
  29. Huang, L.; Zhang, H.; He, Z.; Chen, J.; Song, S. In situ formation of nitrogen-doped carbon-wrapped Co3O4 enabling highly efficient and stable catalytic reduction of p-nitrophenol. Chem. Commun. 2020, 56, 770–773. [Google Scholar] [CrossRef]
  30. Xu, L.; Jiang, Q.; Xiao, Z.; Li, X.; Huo, J.; Wang, S.; Dai, L. Plasma-Engraved Co3O4 Nanosheets with Oxygen Vacancies and High Surface Area for the Oxygen Evolution Reaction. Angew. Chem. Int. Ed. 2016, 55, 5277–5281. [Google Scholar] [CrossRef]
  31. Zhang, Y.; Huang, X.; Li, J.; Bai, J.; Zhou, C.; Li, L.; Wang, J.; Long, M.; Zhu, X.; Zhou, B. Rapid Conversion of Co2+ to Co3+ by Introducing Oxygen Vacancies in Co3O4 Nanowire Anodes for Nitrogen Removal with Highly Efficient H2 Recovery in Urine Treatment. Environ. Sci. Technol. 2022, 56, 9693–9701. [Google Scholar] [CrossRef]
  32. Wang, Y.; Zhou, T.; Jiang, K.; Da, P.; Peng, Z.; Tang, J.; Kong, B.; Cai, W.-B.; Yang, Z.; Zheng, G. Reduced Mesoporous Co3O4 Nanowires as Efficient Water Oxidation Electrocatalysts and Supercapacitor Electrodes. Adv. Energy Mater. 2014, 4, 16. [Google Scholar] [CrossRef]
  33. McCrory, C.C.L.; Jung, S.; Ferrer, I.M.; Chatman, S.M.; Peters, J.C.; Jaramillo, T.F. Benchmarking Hydrogen Evolving Reaction and Oxygen Evolving Reaction Electrocatalysts for Solar Water Splitting Devices. J. Am. Chem. Soc. 2015, 137, 4347–4357. [Google Scholar] [CrossRef]
  34. Wang, X.; Li, X.; Mu, J.; Fan, S.; Chen, X.; Wang, L.; Yin, Z.; Tade, M.; Liu, S. Oxygen Vacancy-rich Porous Co3O4 Nanosheets toward Boosted NO Reduction by CO and CO Oxidation: Insights into the Structure–Activity Relationship and Performance Enhancement Mechanism. ACS Appl. Mater. Interfaces 2019, 11, 41988–41999. [Google Scholar] [CrossRef] [PubMed]
  35. Lin, K.-Y.A.; Chen, B.-C. Efficient elimination of caffeine from water using Oxone activated by a magnetic and recyclable cobalt/carbon nanocomposite derived from ZIF-67. Dalton Trans. 2016, 45, 3541–3551. [Google Scholar] [CrossRef] [PubMed]
  36. Liu, H.; Han, J.-L.; Yuan, J.; Liu, C.; Wang, D.; Liu, T.; Liu, M.; Luo, J.; Wang, A.; Crittenden, J.C. Deep Dehalogenation of Florfenicol Using Crystalline CoP Nanosheet Arrays on a Ti Plate via Direct Cathodic Reduction and Atomic H. Environ. Sci. Technol. 2019, 53, 11932–11940. [Google Scholar] [CrossRef] [PubMed]
  37. Mao, R.; Li, N.; Lan, H.; Zhao, X.; Liu, H.; Qu, J.; Sun, M. Dechlorination of Trichloroacetic Acid Using a Noble Metal-Free Graphene–Cu Foam Electrode via Direct Cathodic Reduction and Atomic H. Environ. Sci. Technol. 2016, 50, 3829–3837. [Google Scholar] [CrossRef]
  38. Xue, Y.; Yu, Q.; Ma, Q.; Chen, Y.; Zhang, C.; Teng, W.; Fan, J.; Zhang, W.-X. Electrocatalytic Hydrogenation Boosts Reduction of Nitrate to Ammonia over Single-Atom Cu with Cu(I)-N3C1 Sites. Environ. Sci. Technol. 2022, 56, 14797–14807. [Google Scholar] [CrossRef]
  39. Wang, Y.; Liu, C.; Zhang, B.; Yu, Y. Self-template synthesis of hierarchically structured Co3O4@NiO bifunctional electrodes for selective nitrate reduction and tetrahydroisoquinolines semi-dehydrogenation. Sci. China Mater. 2020, 63, 2530–2538. [Google Scholar] [CrossRef]
  40. 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]
Figure 1. Characterization of as-prepared material: SEM images of NF (a), Co3O4/NF (b), and Co3O4−x/NF (c,d).
Figure 1. Characterization of as-prepared material: SEM images of NF (a), Co3O4/NF (b), and Co3O4−x/NF (c,d).
Metals 13 00799 g001
Figure 2. XRD patterns of Co3O4/NF and Co3O4−x/NF (a); XPS spectra of wide scan (b) and Co 2p (c); value of Co3O4−x/NF (in g) (d).
Figure 2. XRD patterns of Co3O4/NF and Co3O4−x/NF (a); XPS spectra of wide scan (b) and Co 2p (c); value of Co3O4−x/NF (in g) (d).
Metals 13 00799 g002
Figure 3. LSV curves of Co3O4/NF and Co3O4−x/NF tested with and without NO3-N (a); EIS spectra of different electrodes (b); CV curves of Co3O4−x/NF (c) in the range of −0.06~0.06 V vs. Ag/AgCl; and linear fitting of the capacitive properties of current density vs. scan rate (d).
Figure 3. LSV curves of Co3O4/NF and Co3O4−x/NF tested with and without NO3-N (a); EIS spectra of different electrodes (b); CV curves of Co3O4−x/NF (c) in the range of −0.06~0.06 V vs. Ag/AgCl; and linear fitting of the capacitive properties of current density vs. scan rate (d).
Metals 13 00799 g003
Figure 4. Reduction of NO3 (a), generation of NO2 (b) and NH4+ (c), and TN removal and rate constant (d) for different cathodes. Electrolysis conditions: Cathodic potential = −1.2 V (vs. Ag/AgCl), Initial NO3 concentration = 50 mg N L−1, Solution volume = 80 mL, pH = 7.0, Stirring rate = 600 rpm, Electrolysis time = 280 min.
Figure 4. Reduction of NO3 (a), generation of NO2 (b) and NH4+ (c), and TN removal and rate constant (d) for different cathodes. Electrolysis conditions: Cathodic potential = −1.2 V (vs. Ag/AgCl), Initial NO3 concentration = 50 mg N L−1, Solution volume = 80 mL, pH = 7.0, Stirring rate = 600 rpm, Electrolysis time = 280 min.
Metals 13 00799 g004
Figure 5. Effect of cathodic potential on NO3 reduction (a,b). Electrolysis conditions: Initial NO3 concentration = 50 mg N L−1, Solution volume = 80 mL, pH = 7.0, Stirring rate = 600 rpm, Electrolysis time = 5 h; Effect of initial NO3 concentration on NO3 reduction (c) and removal or generation efficiency (d); Electrolysis conditions: Cathodic potential = −1.2 V (vs. Ag/AgCl), Solution volume = 80 mL, pH = 7.0, Stirring rate = 600 rpm, Electrolysis time = 5 h.
Figure 5. Effect of cathodic potential on NO3 reduction (a,b). Electrolysis conditions: Initial NO3 concentration = 50 mg N L−1, Solution volume = 80 mL, pH = 7.0, Stirring rate = 600 rpm, Electrolysis time = 5 h; Effect of initial NO3 concentration on NO3 reduction (c) and removal or generation efficiency (d); Electrolysis conditions: Cathodic potential = −1.2 V (vs. Ag/AgCl), Solution volume = 80 mL, pH = 7.0, Stirring rate = 600 rpm, Electrolysis time = 5 h.
Metals 13 00799 g005
Figure 6. XPS spectra of Co 2p after the reaction (a), the Co2+/Co3+ ratio before and after the reaction (b), and schematic diagram of the possible reaction mechanism (c).
Figure 6. XPS spectra of Co 2p after the reaction (a), the Co2+/Co3+ ratio before and after the reaction (b), and schematic diagram of the possible reaction mechanism (c).
Metals 13 00799 g006
Table 1. Summary of electrochemical NO3 reduction with Co3O4−x/NF cathode.
Table 1. Summary of electrochemical NO3 reduction with Co3O4−x/NF cathode.
Cathodic Potential (V vs. Ag/AgCl) aNO3 Concentration (mg N L−1) b
−1.0−1.1−1.2−1.3−1.420 c50100 d200 e
R(NO3)35.980.793.799.598.894.993.794.194.9
S(NH4+)86.890.585.491.291.684.685.484.985.7
η f11.524.626.924.522.812.326.936.545.9
k (min−1) g0.00160.00550.01020.01700.02100.01280.01020.00930.0069
R20.980.970.990.980.950.980.990.980.97
a Initial NO3 concentration = 50 mg N L−1, Initial pH = 7.0, Stirring rate = 600 rpm, Solution volume = 80 mL, Reaction time = 280 min; b Cathodic potential = −1.2 V (vs. Ag/AgCl), Initial pH = 7.0, Stirring rate = 600 rpm, Solution volume = 80 mL, Reaction time = 280 min; c Reaction time = 240 min; d Reaction time = 320 min; e Reaction time = 400 min; f η is the average current efficiency; g k is the pseudo-first-order kinetic rate constant.
Table 2. Comparison of NO3RR performance between Co3O4−x/NF and reported catalysts.
Table 2. Comparison of NO3RR performance between Co3O4−x/NF and reported catalysts.
Catalyst NO 3 —N Concentration (ppm) NO 3 —N Conversion (%) NH 4 + —N Selectivity (%) Ref.
Co3O4−x/NF2094.984.6This work
5093.785.4
10094.184.9
20094.985.7
Cu MNC-710094.981.8[38]
TiO2−x/Ti foil5095.287.1[3]
Fe2O3 NRs/CC1400<175.2[26]
Co3O4@NiO2004662.3[39]
Fe SAC7000~769[40]
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

Wu, X.; Liu, Z.; Gao, T.; Li, Z.; Song, Z.; Tang, J.; Feng, F.; Qu, C.; Yao, F.; Tang, C. Boosting Electrocatalytic Reduction of Nitrate to Ammonia over Co3O4 Nanosheets with Oxygen Vacancies. Metals 2023, 13, 799. https://doi.org/10.3390/met13040799

AMA Style

Wu X, Liu Z, Gao T, Li Z, Song Z, Tang J, Feng F, Qu C, Yao F, Tang C. Boosting Electrocatalytic Reduction of Nitrate to Ammonia over Co3O4 Nanosheets with Oxygen Vacancies. Metals. 2023; 13(4):799. https://doi.org/10.3390/met13040799

Chicago/Turabian Style

Wu, Xing, Zhigong Liu, Tianyu Gao, Zhizhuo Li, Zhenhui Song, Jia Tang, Fan Feng, Caiyan Qu, Fubing Yao, and Chongjian Tang. 2023. "Boosting Electrocatalytic Reduction of Nitrate to Ammonia over Co3O4 Nanosheets with Oxygen Vacancies" Metals 13, no. 4: 799. https://doi.org/10.3390/met13040799

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