Next Article in Journal
A Non-Invasive Fetal QRS Complex Detection Method Based on a Multi-Feature Fusion Neural Network
Previous Article in Journal
Exploring the Effectiveness of Advanced Chatbots in Educational Settings: A Mixed-Methods Study in Statistics
Previous Article in Special Issue
Contamination with Pharmaceuticals in Aquatic Environment: Focus on Analytical Methodologies
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Electrocatalytic Nitrate Reduction for Brackish Groundwater Treatment: From Engineering Aspects to Implementation

1
Laboratory of Process and Environmental Engineering, Higher School of Technology, Hassan II University of Casablanca, Casablanca 20000, Morocco
2
Clermont Auvergne INP, Institut Pascal, Clermont Auvergne University, CNRS, 63000 Clermont–Ferrand, France
3
International Water Research Institute, Mohammed VI Polytechnic University, Benguerir 43150, Morocco
*
Authors to whom correspondence should be addressed.
Appl. Sci. 2024, 14(19), 8986; https://doi.org/10.3390/app14198986 (registering DOI)
Submission received: 21 August 2024 / Revised: 24 September 2024 / Accepted: 1 October 2024 / Published: 5 October 2024

Abstract

:
In recent years, nitrate has emerged as a significant groundwater pollutant due to its potential ecotoxicity. In particular, nitrate contamination of brackish groundwater poses a serious threat to both ecosystems and human health and remains difficult to treat. A promising, sustainable, and environmentally friendly solution when biological treatments are not applicable is the conversion of nitrate to harmless nitrogen (N2) or ammonia (NH3) as a nutrient by electrocatalytic nitrate reduction (eNO3R) using solar photovoltaic energy. This review provides a comprehensive overview of the current advances in eNO3R for the production of nitrogen and ammonia. The discussion begins with fundamental concepts, including a detailed examination of the mechanisms and pathways involved, supported by Density Functional Theory (DFT) to elucidate specific aspects of ammonium and nitrogen formation during the process. Furthermore, the integration of artificial intelligence (AI) and machine learning (ML) offers promising advancements in enhancing the predictive power of DFT, accelerating the discovery and optimization of novel catalysts. In this review, we also explore various electrode preparation methods and emphasize the importance of in situ characterization techniques to investigate surface phenomena during the reaction process. The review highlights numerous examples of copper-based catalysts and analyses their feasibility and effectiveness in ammonia production. It also explores strategies for the conversion of nitrate to N2, focusing on nanoscale zerovalent iron as a selective material and the subsequent oxidation of the produced ammonia. Finally, this review addresses the implementation of the eNO3R process for the treatment of brackish groundwater, discussing various challenges and providing reasonable opinions on how to overcome these obstacles. By synthesizing current research and practical examples, this review highlights the potential of eNO3R as a viable solution to mitigate nitrate pollution and improve water quality.

1. Introduction

The water crisis is considered to be the most critical problem that requires attention and practical solutions, as more than one billion people do not have access to drinking water [1,2]. The reasons for the limited access to drinking water are usually the unequal distribution of water on Earth, with only 2.5% freshwater in the total water, and water pollution by dissolved organic and mineral substances [3]. Groundwater is typically regarded as one of the purest sources of drinking water in many countries, such as India [4], and the Anti-Atlas in Morocco [5]. However, these water resources are facing serious problems due to rapid urbanization and industrialization. One of these issues is nitrate contamination, which presents a serious threat to ecosystems and human health. Broadly speaking, nitrate accumulation in groundwater is caused by various anthropogenic and geogenic activities [6]. In fact, it is commonly agreed that agricultural activities are the most important source of nitrate pollution in groundwater [7,8,9,10]. Many regions of the world report a significant percentage of nitrates in groundwater that exceeds World Health Organization (WHO) standards. Hence, anthropogenic nitrate is an environmental handicap with serious consequences when it reaches the human body through water ingestion [11], especially a disease that occurs in the bloodstream of babies under 6 months of age, called infant cyanosis, methemoglobinemia, or blue babies [12]. The conversion process of nitrates to nitrites and then to nitrosamines in the human body leads to harmful effects, such as gastric cancer [13]. That is why the WHO considers drinking water safe in terms of nitrates when it has a quality ratio less than or equal to one ( QR = C nitrate 50 1 ) [14].
To meet the permissible standard of nitrate concentration in drinking water, many physicochemical techniques have been adopted to remove nitrates from the water, such as adsorption [15,16], ion exchange [17,18], biological denitrification [19,20], anaerobic ammonium oxidation (Anammox) [21,22], chemical denitrification [23,24], catalytic denitrification [25,26,27], nanofiltration, and reverse osmosis [28,29]. Nevertheless, these processes suffer from serious limitations, including high installation and maintenance costs, biofouling, brine generation, membrane sensitivity, and the need for post-treatment [30]. Currently, eNO3R can be perceived as the best way to treat contaminated water containing nitrogen compounds owing to their advantages, such as easy operation, remote control, minimal generation of secondary waste generation, and no potential for failure associated with high salinity, and these treatments can also be driven by renewable energy such as solar energy, hydropower, and wind power [31,32]. On the other hand, some disadvantages can minimize their effectiveness, such as low efficiency for higher nitrate contents and very long reaction time [33,34]. Numerous research investigations have centered their attention on the electrochemical synthesis of ammonia ( NH 3 ) through the reduction in nitrate ( NO 3 ) to ammonium ( NH 4 + ) in nitrate-contaminated water. This focus emerges as a promising avenue to alleviate the substantial energy demands entailed by the conventional Haber–Bosch industrial process, which produces NH 3 from nitrogen ( N 2 ) and hydrogen ( H 2 ) [35]. Simultaneously, it offers a sustainable means of nitrogen recovery. The second strategy is eliminating NO 3 from groundwater while concurrently converting it into nitrogen gas. This pursuit assumes significance due to its pivotal role in maintaining equilibrium within the nitrogen cycle (N-cycle). From a thermodynamic standpoint, N 2 boasts the lowest solubility in water, simplifying its separation from aqueous mediums. Consequently, the electroreduction in NO 3 into N 2 emerges as a highly desirable approach, representing an efficient pathway for the generation of potable water and the restoration of equilibrium within the nitrogen cycle [36]. Thus, the present review focuses on introducing the mechanism of eNO3R and providing a basic understanding of the complex multi-electron transfer required to successfully carry out the process, by discussing the trends in the catalytic activity of the eNO3R to ammonia and nitrogen using Density Functional Theory (DFT). In addition, we explore the emerging role of machine learning (ML) and artificial intelligence (AI) in enhancing DFT methodologies, highlighting their potential to accelerate the discovery and optimization of efficient catalysts for nitrate reduction. As well, we review the various preparation methods for electrocatalysts and discuss advanced characterization techniques, with particular emphasis on in situ methods, to better understand catalyst behavior under reaction conditions. Furthermore, the most recent advancements of electrocatalysts for NH 3 production by copper-based catalysts as a case study have been summarized, and the strategies to convert nitrate to nitrogen by selective materials as the case of nanoscale zero valent iron and the oxidation of ammonium generated during electrolysis are highlighted. Additionally, parameters that are influencing the reactivity and selectivity such as pH, coexistence ions, current density, and applied voltage are presented. Finally, the application of the process with the perspective, challenges, and technoeconomic analysis for brackish groundwater is analyzed.
As a result, the objective of this review is to propose a comprehensive overview of the current advances in eNO3R for the production of nitrogen and ammonia. This includes a discussion of fundamental concepts, a detailed examination of the mechanisms and the potential catalysts, and an analysis of the pathways of ammonium and nitrogen formation with a focus on strategies for the conversion of nitrate to N2. Finally, this review addresses the implementation of the eNO3R process for the treatment of brackish groundwater as a typical example. By synthesizing current research and practical examples, this review highlights the potential of eNO3R as a sustainable solution to mitigate nitrate pollution and improve water quality.

2. Reaction Mechanisms and Kinetics of Electrocatalytic Nitrate Reduction

The electrochemical nitrate reduction mechanism is a complicated multi-electron-transfer process, with 5-electron to form N 2 and 8-electron to generate NH 4 + which involves a variety of reaction intermediates and products. According to the relevant standard redox potential of nitrate reduction and the respective possible products (Table 1), the most stable compounds are N 2 and NH 4 + which shows the highest equilibrium potential among all the possible products of NO 3 reduction [37].

2.1. Mechanistic Insight into the Electrochemical Nitrate Reduction

From an electrochemical standpoint, several investigations have confirmed that eNO3R can involve different pathways, which produce a variety of products that enclose nitrogen gas and ammonia as well as undesired species such as nitrite, hydrazine, hydroxylamine, nitric oxide, nitrous oxide, and others as depicted in Figure 1.
The electroreduction process of nitrates, as shown in various studies, initiates with the conversion of NO 3 to NO 2 , which is commonly referred to as the rate-determining step (RDS). This step starts with the adsorption of nitrate as the electron is completely donated to the electrocatalyst surface. The adsorbed nitrate ( NO 3 ads) could be reduced to the unstable nitrate dianion radical ( NO 3 2 ), which is hydrolyzed to form a nitrogen dioxide radical ( NO 2 ) that is finally reduced to adsorbed nitrite ( NO 2   ads ) playing a critical role in the overall reaction pathway as given in Equations (1)–(4) [38].
NO 3   aq +   *   NO 3   ads  
NO 3   ads   + e NO 3   ads   2
NO 3   ads   2 + H 2 O NO 2   ads   + 2   OH
NO 2   ads   + e NO 2   ads   + H 2 O  
The next step is the reduction in intermediate NO 2   ads into adsorbed nitric oxide ( NO ads ) that is considered an important intermediate that opens an alternative reaction into three pathways (Figure 2) [39,40].
Duca et al. [41] indicated that the hydrogenation of NO ads to NH 2   ads , then NH 2   ads reacts with NO ads and desorbs to form N 2 (Equations (5) and (6)).
NO   ads + 4 e + 3   H 2 O   NH 2 ,   ads + 4   OH
NO   ads + NH 2 ,   ads N 2 +   H 2 O
Furthermore, De Vooys et al. [40] showed that the mechanism proceeds by dimerization of NO ads to form N 2 O 2 , followed by protonation to HN 2 O 2 which is then reduced electrochemically to N 2 O   ads as the following equations (Equations (7) and (8)).
NO   ads + NO   ( aq ) +   H + +   e   HN 2 O 2 ,   ads
HN 2 O 2 ,   ads +   H + +   e   N 2 O   aq +   H 2 O
Then, Chumanov et al. [42], confirmed that N 2 O   aq can then react with the hydrated electron and undergo the following reactions to form N 2 (Equations (9)–(11)).
N 2 O   aq + e N 2 O
N 2 O + H 2 O N 2 + OH + OH
OH + e OH  
The third proposed pathway is the formation of hydroxylamine ( NH 2 OH ) and NH 4 + due to the prevention of desorption of NO ads which allows its hydrogenation to HNO , H 2 NO , NH 2 OH , and finally NH 4 + (Equations (12)–(16)) [40,43,44,45,46,47].
NO   ads +   H + + e HNO   ads
HNO   ads +   H + + e NH 2 OH   ads
NH 2 OH + H + H 3 NOH +
NH 2 OH + 2   H + + e NH 3 +   H 2 O
NH 3 + H + NH 4 +
On the other hand, Vetter and Schmid proposed two autocatalytic mechanisms (Equations (17)–(24)) due to the protonation of the unstable NO 2   ads   to nitrous acid in highly acidic media, which creates a sufficient amount of HNO 2 that dominates over the NO 2   , leading to the spontaneous formation of several products, including nitric oxide ( NO ), nitrogen dioxide ( NO 2 ), and HNO 2 as a result of disproportionation reactions [48].
Vetter mechanism
NO 2 + e NO 2
NO 2 + H + HNO 2
HNO 2 + HNO 3 N 2 O 4 +   H 2 O
N 2 O 4 2   NO 2
Schmid mechanism
NO + + e NO
HNO 2 + H + NO + +   H 2 O
HNO 2 + HNO 3 N 2 O 4 +   H 2 O
N 2 O 4 + 2   NO   + 2   H 2 O   4   HNO 2
Another important pathway was proposed by Katsounaros et al. [44], who confirmed the formation of N 2 and N 2 O from NH 2 OH in the presence of HNO 2 and HNO (Equations (25) and (26)).
NH 2 OH   +   HNO N 2 + 2   H 2 O
NH 2 OH   +   HNO 2 N 2 O + 2   H 2 O
In addition to the kind of nitrate reduction by the electron generated from the cathode as described above, nitrate can also be mediated by the atomic hydrogen ( H ads ) (Figure 3).
The process starts with the reduction in water H 2 O via the Volmer process, to form H ads . Then, H ads which is considered a strong reducing agent ( E H + / H 0 = −2.31 V vs. SHE) directly reduces nitrate to form the dominant end-product NH 4 + . Notably, two N ads can combine to produce N 2 but the formation of the N–N bond is kinetically less favorable than the N–H bond because the calculated migration barrier (Δ E a ) of H ads is 0.10 eV, which is much lower than that of N ads (0.75 eV) [49]. The following electrochemical reactions described the mechanistic formation of N 2 and NH 3 (Equations (27)–(33)).
H 2 O   +   e H ads +   OH
NO 3 ,   ads   + 2   H ads NO 2 ,   ads   +   H 2 O
NO 2 ,   ads   +   H ads NO ads +   OH
NO ads + 2   H ads N ads +   H 2 O
N ads +   H ads NH ads
NH ads +   H ads NH 2 ,   ads
NH 2 ,   ads +   H ads NH 3 ,   ads
Based on the previously discussed reactions, it is evident that eNO3R to produce N 2 and NH 3 involves intricate and multifaceted pathways. The main hurdle in this process lies in enhancing the kinetics of nitrite formation, known as RDS (rate-determining step). To overcome this challenge and optimize the overall performance of the process, it is imperative to gain a deeper understanding of the intricate mechanisms governing electrochemical nitrate reduction. This includes a comprehensive exploration of intermediates and reaction kinetics. Consequently, it becomes apparent that further investigation using advanced methodologies and theoretical studies is indispensable to validate the proposed reaction mechanism.

2.2. DFT Calculation as an Advanced Tool for eNO3R

In recent years, significant advancements have been made in using Density Functional Theory (DFT) to simulate the mechanisms of nitrate reduction, providing new insights into this process. DFT calculations have revealed critical details about the adsorption energies of various nitrogen species (e.g., NO 3 , NO 2 , NO , N 2 , etc.), which are necessary for assessing the catalytic activity of different materials. The adsorption of nitrates plays a crucial role in electrocatalytic processes as it represents the first step. Employing DFT calculation, it has been demonstrated that NO 3 can be adsorbed on the active site via chemisorption (2 O and 1 O patterns) or physisorption (Figure 4) [50].
Furthermore, the presence of oxygen species at the catalytic sites is significant because it can facilitate this bond-breaking process, making it easier for the subsequent steps of nitrate reduction to occur. The subsequent steps of the mechanism following nitrate adsorption are generally exothermic, indicating the presence of multiple rate-determining steps within the process [51]. DFT, in conjunction with the use of descriptors, has proven instrumental in predicting novel materials by identifying specific RDS, as illustrated in Table 2.
This approach offers deeper insights into the selective catalytic reduction mechanisms of nitrates into gaseous nitrogen or ammonium. Similar advancements have been observed in hydrogen evolution reactions (HER), highlighting the competitive nature of these processes with eNO3R. Furthermore, the integration of artificial intelligence (AI) and machine learning (ML) provides a promising avenue for predicting catalytic performance and significantly reducing the computational demands of DFT, thereby accelerating material discovery and optimization.

2.2.1. Insights into eNO3R: Descriptors and Rate Determining Steps (RDS)

Descriptors are simplified parameters connected to fundamental surface properties that control both adsorption energies and activation energies for reactions. Moreover, a more suitable descriptor is the strength of surface-adsorbate bonds, which can be estimated from experimental data. This estimation facilitates the construction of a Sabatier-type ‘volcano’ plot, thus helping to identify optimal catalysts by correlating bond strength with catalytic activity [64]. For example, the binding energies of nitrate adsorbed (ΔG*NO3) serve as the first descriptor in determining catalytic efficiency, as nitrate adsorption is recognized as the initial step involving a solution-mediated proton transfer without electron transfer, as previously described (This step has been previously identified as RDS) [65]. In addition to * NO 3 , the formation of adsorbed * NO 2 and * NO represent additional key descriptors, with * NO 2 identified as the RDS and * NO as the potential-limiting step (PLS) in the process. Weak adsorption of NO 2 and NO can result in their desorption into the solution, thereby negatively impacting the overall efficiency of the reaction. Figure 5A presents the binding energies of * NO 3 , * NO 2 , and * NO on different transition metals. The weak binding of * NO 3 and * NO 2 on Au results in poor catalytic performance for nitrate reduction. Moreover, most transition metals show a high desorption barrier for * NO except for Ag and Au [66]. To confirm these results, Karamad et al. [66] used the volcano framework as illustrated in Figure 5B, based on the Sabatier principle to analyze the catalytic activity of transition metals in the eNO3R. They focus on the binding energies of nitrogen which correlate with the binding energies of * NO adsorbates (ΔG*NO) and the binding energy of oxygen which correlates with *OH adsorbates (ΔG*OH). They also identify the PLS as the reaction with the most negative limiting potential, revealing that for metals like Cu, Ni, Rh, Pt, and Pd, the key steps are * NO → * NOH or * NO H * NO which means that the protonation of * NO or * NO 3 is the slowest and most difficult step in the overall reaction pathway. Furthermore, Cu demonstrates superior catalytic activity despite a significant overpotential and is the most active and selective catalyst for eNO3R to ammonia among the other transition metals (It will be useful in the discussion in Section 4 for copper-based electrocatalyst). They also proposed strategies for enhancing nitrate reduction catalysis including (a) alloying with oxygen-affinitive elements to stabilize intermediates like * NOH and H * NO , though care must be taken to avoid * OH poisoning; (b) tethering ligands that geometrically interact with * NO / H * NO to stabilize the latter without significantly affecting * NO binding; (c) employing promoters to modify adsorption sites and binding energies; and (d) leveraging hydrogen bonding with intermediates such as * NHO /* NOH to achieve stronger stabilization and improved catalytic performance [66]. Another study by Liu et al. [67] predicted the activity of nitrate reduction electrocatalysts under different applied potentials (−0.2, 0, 0.2, and 0.4 V). They also constructed theoretical volcano plots using nitrogen binding energy (ΔEN) and oxygen binding energy (ΔEO) which serve as key descriptors for evaluating eNO3R performance on metal surfaces, obtained via DFT-based mean-field microkinetic modeling. The volcano plots reveal how the nitrate reduction rate depends on the adsorption strengths of N and oxygen O species at four different applied potentials. RDS in nitrate reduction varies with applied potential and metal type. At low potentials (−0.2 to 0 V), the RDS for Pt and Pd is the dissociation of * NO 3 . As the potential increases to 0.4 V, * NO desorption becomes rate-controlling for these metals. For Rh, * N 2 formation is rate-controlling above 0.2 V. In contrast, for Fe and Co, the RDS shifts from * NO 3 dissociation to N2 formation as the potential surpasses −0.2 V. On Cu, the RDS transitions from * NO 3 dissociation to * NO 2 dissociation at 0.2 V due to increased * NO 2 coverage. For Ag and Au, * NO 3 dissociation remains the RDS across all potentials, attributable to their weak adsorption and low nitrate coverage [67]. As previously demonstrated, descriptors and volcano plots play a critical role in identifying RDS and determining optimal catalysts. These approaches have been extensively utilized to predict new catalysts and effectively distinguish between different materials based on their catalytic activity and selectivity, as exemplified by the study conducted by Shathishkumar et al. [68] who employed first-principles calculations to assess the electrocatalytic performance of dual-atom catalysts (DACs) based on TM 2 / N 6 G (DACs supported by an N-doped graphene nanosheet where TM 2 ranges from Sc 2 to Zn 2 ). Through high-throughput calculations, they identified Cr 2 / N 6 G , Mn 2 / N 6 G , and Cu 2 / N 6 G as prominent candidates near the volcano plot as illustrated in Figure 5D, exhibiting impressive limiting potentials of −0.46, −0.45, and −0.36 V, respectively. Furthermore, significant energy barriers for byproduct formation ( NO 2 , NO , and N 2 O ) which confirm their high selectivity [68].

2.2.2. Nitrate to Ammonium

For ammonia generation, research endeavors have shifted their focus towards investigating the subsequent step through DFT studies, specifically examining the hydrogenation of NO ads . This pathway is widely acknowledged as the primary route for the synthesis of ammonium; however, it is complex to study due to the addition of hydrogen atoms, which can occur at both the nitrogen (N) and oxygen (O) atoms of NO , two possible products can be formed: H NO (hydrogenated nitrogen) or NO H (hydrogenated oxygen) [69]. Furthermore, recent studies have shown that the hydrogenation process occurs in four distinct steps. These steps correspond to different adsorption configurations of NO , including O-end, O-side, N-end, and N-side pathways (Figure 6) [53]. These hydrogenation steps are also known for their significant energy barriers, which are influenced by several factors, with the electrocatalyst being the key determinant. Lim et al. [70] showed that the reduction in nitrate by different facets of Pd nanoparticles changes the selectivity and reduction rate. Energetics DFT calculations demonstrated that the reaction energy associated with the dissociation of * NO 3 to * NO 2 + * O was favorable on the Pd(111) facet than on the Pd(100) facet. In contrast, * NO 2 adsorbs more strongly on Pd(100) surfaces which favors strong ammonium formation compared to Pd(111) surfaces. Another study by Liu et al. [55] selected the Pd(111) plane as the Pd facet to build Pd ( WO 3 ) 3 catalyst and they have carried out DFT calculations to investigate the enhanced eNO3R activity. The comparison between the calculated adsorption energy of NO 3 on pure Pd and Pd ( WO 3 ) 3 showed that the energy barrier of the hydrogenation *N to *NH on pure Pd is the largest by 3.38 eV. However, the doping of the surface with ( WO 3 ) 3 reduced this energy barrier. This study also reported that the reaction undergoes a * NO 3 → * NO 2 → * NO → * N → * NH → * NH 2 → * NH 3 pathway remaining the same but the RDS changed from the hydrogenation of * N to * NH on pure Pd to the hydrogenation of * NH 2 to * NH 3 on Pd ( WO 3 ) 3 . Furthermore, it is important to note that RDS changes as a function of facet and material composition, as reported by Liu et al. [55].

2.2.3. Nitrate to Nitrogen

In the context of nitrogen formation, DFT computes the adsorption energy of nitric oxide on catalytic surfaces. Equally essential is the understanding of the intricate process of NO transfer from surface sites to the bulk solution, involving the assessment of energy profiles associated with desorption. Furthermore, the energetics of nitrogen–nitrogen (N–N) bond formation are of paramount significance [71]. The rational modification of catalyst structures can regulate the adsorption strengths between the active intermediates and the catalyst sites, suppressing competitive reactions and promoting the selectivity of the target product. Fan et al. [63] developed an electrocatalytic material called Meso-Fe-N-C, consisting of atomically dispersed iron coordinated with nitrogen on a mesoporous carbon framework, which greatly improves the selectivity of eNO3R. They found that the generated * NO , which is a key intermediate for the formation of N 2 or NH 3 / NH 4 + , reacts with dissolved NO in solution to produce * N 2 O . Then, the latter can be easily reduced to * N 2 and desorbed from the active sites. They confirmed that the successive hydrogenation of * NO forms * NO H, * NH 2 OH , and finally * NH 3 . Moreover, the study concluded that the Fe–N active site is more thermodynamically favorable than Fe–Fe for N 2 emergence. Furthermore, the confinement effect of mesoporous carbon is also a significant factor contributing to selective N 2 generation, mainly through two aspects: geometry distortion of surface Fe–N–C active sites and modified the distribution of NO concentration in mesopores [63].

2.2.4. eNO3R vs. HER

Generally, an ideal catalyst for eNO3R should maintain high stability and catalytic activity while should effectively suppress the Hydrogen Evolution Reaction (HER). For this reason, Strong competition from HER presents a significant challenge in aqueous systems [72]. Notably, Hu et al. [51] found that HER competes strongly with nitrate reduction; at pH = 0, the formation of H2 requires only 0.25 eV, lower than the 0.37 eV needed for nitrate reduction on Cu(111), making eNO3R less favorable in acidic conditions. In neutral environments (pH = 7), eNO3R outcompetes HER, while in alkaline conditions (pH = 14), HER predominates [51]. This indicates that pH significantly influences the competitive dynamics between HER and eNO3R. However, it is not desirable to fully suppress HER, as the Adsorption strengths of the H atom or active hydrogen species are essential for the next step—nitrogen hydrogenation to produce ammonia, as discussed in the previous section of nitrate to ammonium. Typically, the binding energy of *H serves as a useful descriptor of hydrogen evolution activity across a variety of catalyst materials. To analyze selectivity trends for the NO3RR among transition metals, Karamad et al. [66] used an approach similar to that used for CO 2 reduction. Accordingly, they plotted the difference UL(eNO3R) − UL( H 2 ) against the limiting potential UL(eNO3R) for various transition metals as illustrated in Figure 7a. A more positive value of UL(eNO3R) − UL( H 2 ) correlates with increased selectivity for eNO3R over HER, suggesting that catalysts in the upper right region exhibit both high activity and selectivity for eNO3R. Importantly, this analysis focuses on selectivity for the eNO3R relative to HER rather than specifically towards NH 3 or N 2 . The results suggest that Cu emerges as the most active and selective catalyst for the eNO3R, aligning with previous studies identifying Cu as a leading candidate. Additionally, while Ir shows higher catalytic activity than Cu, its lower selectivity towards eNO3R products leads to H 2 being the dominant product under eNO3R conditions. Furthermore, it is critical to develop strategies to effectively manage this competition. Therefore, active sites should preferentially bind NO3 over H atoms, with ΔGNO3 > ΔGH, and the formation of HER byproducts can be minimized with UL( H 2 ) > UL( NH 3 ) [66]. In this context, Sathishkumar et al. [68] considered TM 2 / N 6 G ( TM 2 = Sc 2 , Ti 2 , V 2 , Cr 2 , Mn 2 , Fe 2 , Co 2 , Cu 2 , and Zn 2 ) as potential catalysts. The adsorption capabilities of NO 3 and the suppression of HER (illustrated in Figure 7b) was assessed. Among the nine candidates, Cr 2 / N 6 G , Mn 2 / N 6 G , and Cu 2 / N 6 G met the selectivity criteria: ΔGNO3 > ΔG*H and UL( H 2 ) > UL( NH 3 ). The free energy diagrams for HER indicated uphill ΔGPDS values of 0.59, 0.24, and 0.37 eV for Cr 2 / N 6 G , Mn 2 / N 6 G , and Cu 2 / N 6 G , respectively, at their corresponding limiting potentials of −0.46 V, −0.45 V, and −0.36 V. Thus, these three systems were identified as promising candidates for nitrate reduction to ammonia [68].

2.2.5. Analyzing Experimental Data vs. DFT for Nitrate Reduction

The comparison between experimental data and DFT-based simulations reveals notable discrepancies in the detailed energetics of intermediate steps within eNO3R. Regarding that, Long et al. [73] reported that product selectivity and Faradaic efficiencies (FE) for N 2 O , NH 3 , and H 2 obtained from a nanostructured Ag electrode in phosphate-buffered saline (PBS) show distinct potential windows, where NH 3 production becomes dominant at lower potentials, and N2O prevails at higher potentials. This experimental observation aligns with predictions from DFT-based microkinetic simulations, as shown in Figure 8a,b, which highlight potential-dependent shifts in selectivity. However, discrepancies in the energetics of intermediate steps arise; while DFT predicts the RDS for NH 3 production as the electrochemical protonation of H*NOH, experimental results indicate deviations in the exact potential for product transitions. Similarly, the competition between NH 2 OH and NH 3 production is captured by DFT, though slight variations in NH 2 OH suppression onset are noted experimentally [73]. Despite these differences, microkinetic modeling plays a crucial role in bridging the gap between theory and experiment. As shown by Carvalho et al. [74] who compared experimental observations with model predictions to elucidate nitrate reduction reaction (NRA) dynamics on transition metal surfaces (Figure 8c,d). The experimental rate-order profiles exhibited a potential-dependent peak, characteristic of a competitive Langmuir–Hinshelwood mechanism, indicating interdependence between H* and nitrate adsorption. By calibrating a microkinetic model to incorporate the rate-limiting steps of nitrate reduction and hydrogen evolution kinetics, they achieved agreement between experimental and modeled nitrate rate orders by adjusting thermodynamic parameters (ΔGH*–ΔGNO3). The model effectively captured observed trends, including shifts in peak rate order and mass transfer limitations, reinforcing its validity and providing key insights into the competitive adsorption and reaction kinetics across various transition metal catalysts [74]. Furthermore, the integration of machine learning and Artificial Intelligence techniques holds significant potential to further facilitate these modeling efforts by optimizing parameter selection and improving predictive accuracy. This will be explored in the next section.

2.3. Machine Learning and Artificial Intelligence for Efficient eNO3R

For eNO3R, computational methods like DFT, molecular dynamics, and microkinetics have been invaluable in predicting catalyst behaviors and screening materials as discussed above. However, traditional atomistic simulations, while accurate, are computationally intensive, limiting their ability to explore large catalyst design spaces. Machine learning (ML) offers a solution by rapidly predicting catalyst performance using scaling relationships like Brønsted–Evans–Polanyi (BEP) and adsorption-energy trends. By leveraging data-driven approaches, ML can identify optimal catalytic properties for eNO3R [75]. As demonstrated by Chen et al. [76] ML models like Gradient Boosting Regression (GBR) were trained using key intrinsic properties of transition metals—such as d-electron number, electronegativity, and atomic radius—to predict adsorption energies of nitrogen intermediates. This approach significantly reduced computational costs, offering reliable predictions for both side-on and end-on adsorption modes, with the potential to accelerate the discovery of highly efficient eNO3R catalysts. The model’s interpretability further enhances the understanding of catalyst behavior, highlighting the role of ML in advancing catalytic design for nitrate reduction [76]. In addition, a study by Liu et al. [77] discovered that there was no linear correlation between the adsorption-free energies of reaction intermediates, making traditional free energy descriptors ineffective. For this reason, they employed Random Forest (RF) algorithm to predict the limiting potentials (UL) for electrocatalytic NO reduction (NORR) to NH 3 , utilizing 22 atomic, electronic, and custom features. Through Pearson’s correlation, 11 key features were selected for further analysis, including atomic structure and bond lengths. The RF model demonstrated excellent performance with an R2 value of 0.96 and low RMSE, highlighting important features such as bond length, atomic number, and d-band center in influencing UL. Furthermore, they applied the SISSO approach (Sure Independence Screening and Sparsifying Operator) to derive a simplified mathematical expression that links UL to the key features, which aligned well with DFT predictions. As a result, Cu and Pt/Cu single-atom catalysts (SAACs) were identified as promising candidates for NORR to NH 3 . [77]. Another study by Yang et al. [78] applied multiple ML algorithms—SVR, GPR, LASSO, RFR, and XGBR—to further investigate the structure-activity relationship in NORR. A key component of this analysis involved the use of a normalized dataset, with a 10-fold cross-validation approach used to identify optimal hyperparameters for the ML models. The Random Forest Regression (RFR) and XGBR models outperformed others, demonstrating excellent generalization ability with R2 values of 0.95 and 0.98, and RMSE values of 0.30 eV and 0.19 eV, respectively. These models provided reliable predictions of Gibbs free energy changes for the reaction intermediates [78].
Other models have been developed to improve the prediction of adsorption energies and active sites, such as the TinNet (Theory-Infused Neural Network) model and the Bayesian Statistical Learning model (Bayeschem). TinNet integrates traditional theoretical principles with neural networks (AI-based machine learning) which integrates domain-specific physical theories, including the d-band theory of chemisorption, into machine learning frameworks, particularly graph neural networks (GNNs). By embedding these physical principles, TinNet effectively captures complex interactions between atoms and surfaces, resulting in more accurate predictions of adsorption energies and chemical reactivity. On the other hand, the Bayeschem model used a statistical approach to optimize relationships between variables, like material composition and electronic properties, through global optimization techniques such as Markov Chain Monte Carlo (MCMC) [75]. For instance, Gao et al. [79] employed the Bayeschem model to analyze the adsorption properties of * N and * NO 3 on selected pristine metal surfaces. Their study revealed a mechanism to bypass linear adsorption-energy scaling limitations by exploiting site-specific Pauli repulsion interactions. This work highlighted how an enhanced understanding of the interactions between adsorbates and catalyst surfaces can lead to improved electrocatalytic performance. Ultimately, the Bayeschem model plays a crucial role in bridging theoretical insights with practical applications, facilitating the design of more efficient electrocatalysts for nitrate reduction [79].
By employing ML, researchers can efficiently evaluate and optimize new catalysts for electrochemical nitrate reduction, thereby minimizing the reliance on costly simulations and facilitating a more effective discovery of high-performing catalytic solutions [80].

3. Recent Developments in Electrocatalyst Fabrication and Characterization Techniques for Effective eNO3R Process

Currently, numerous techniques including electrodeposition, plasma treatment, thermal treatment, and so on, have been developed for the synthesis of diverse electrocatalysts for nitrate reduction, which plays a critical role in determining the purity and homogeneity of the electrode surface, which are critical parameters influencing the eNO3R efficiency and the catalyst’s selectivity. Furthermore, there are several well-known methods for material characterization, which can be used to characterize electrocatalysts both before and after reactions, including SEM, TEM, XRD, and XPS. Notably, in situ characterization techniques are particularly important for observing changes during electrolysis. Therefore, in this section, we will discuss the different types of electrode preparation methods, followed by an overview of advanced in situ characterization techniques, highlighting their advantages and limitations.

3.1. Advances in Electrocatalyst Preparation Techniques for Efficient Nitrate Reduction

3.1.1. Electrodeposition

Electrodeposition involves depositing material onto an electrode by applying an electrical current or potential in an electrochemical cell. The performance of electrodeposited electrocatalysts is highly dependent on the specific electrodeposition parameters used. To optimize these electrocatalysts, a detailed investigation is needed to understand how different parameters affect the deposition process [81]. Key factors, such as the choice of electrolyte, electrodeposition time, mode of electrodeposition, and the presence of additives, play a crucial role in shaping the performance of the deposited electrocatalyst. Regarding that, Zurita et al. [82] investigated how the electrolyte in the electrodeposition process affects the morphology of copper nanoparticles, specifically on Vitreous Carbon (VC) substrate, observing that Na 2 SO 4 solution led to flower-shaped dendritic structures; however, H 2 SO 4 produced randomly distributed hemispherical particles. The differences in morphology on VC were linked to substrate defects interacting with the electrolytic media, affecting the nucleation and growth of the crystals [82]. Another factor which is the electrodeposition time, plays a crucial role in the performance and electrocatalysts stability for nitrate reduction. Specifically, longer deposition times can lead to issues with electrode stability, as seen in the case of Cu-BDD electrodes. For instance, a copper oxide layer formed after 600 s of electrodeposition was prone to detachment during electrochemical nitrate reduction. This instability can hinder the electrocatalyst’s performance and longevity. In contrast, a deposition time of 300 s was found to be optimal. At this duration, copper particles formed on the Cu-BDD surface were unevenly distributed with diameters ranging from 72.7 to 139 nm, and an average size of 106 nm. The shorter deposition time resulted in better particle adhesion and electrode stability, making it more suitable for eNO3R [83]. Another study by Zhang et al. [84] demonstrated that longer electrodeposition times result in higher Cu loading on the electrode, enhancing NO 3 removal due to increased catalytic activity. However, after 15 min, Cu is evenly distributed on the Ni foam surface, and additional electrodeposition time does not significantly increase active sites. Additionally, extended electrodeposition times (25 and 35 min) can lead to Cu agglomeration, which reduces the degradation rate of NO 3 [84]. The advantage of the electrochemical deposition method is its ability to synthesize electrocatalysts with specific crystallographic facets tailored to needs. For example, Chen et al. [85] synthesized Tetrahexahedral Cu nanocrystals (THH Cu NCs) with {5 2 0} facets using a programmed square-wave potential (SWP) method, which allows precise control over the exposed crystallographic facets by adjusting potential limits. The synthesized THH Cu{5 2 0} nanocrystals demonstrated exceptional catalytic activity and selectivity in the electrochemical reduction in nitrate, achieving a Faradaic efficiency of 98.3% at −0.90 V (vs. RHE) in neutral solution [85]. However, the drawback of this electrocatalyst preparation method is the instability of the electrodes. According to Chen et al. [86] who developed Pd/Cu electrocatalysts with electrodeposition durations between 60 and 360 s, Pd/Cu 180 s demonstrated superior nitrate removal efficiency and nitrogen selectivity. However, ICP analysis identified a Pd concentration of 0.063 ppm after prolonged reaction times, highlighting the necessity for future improvements in electrode stability, possibly by adding other treatments [86].

3.1.2. Plasma Treatment

Plasma treatment uses an ionized gas to modify the surface properties of the materials and specifically to control oxygen vacancies in electrocatalyst materials [87]. different plasma treatments with N 2 , Ar, and O 2 are employed to achieve high and low concentrations of oxygen vacancies in the samples. Jang et al. [88] used a GVTech GVPT–100 plasma system, with a 100 W power supply and at a pressure of 180 Pa. MnCuO x samples were exposed to Ar and O 2 plasma for 240 s, generating different concentrations of oxygen vacancies, with Ar creating a higher concentration and O 2 a lower concentration. Their study confirms that plasma treatment modifies the oxygen vacancy concentration of electrocatalysts without altering morphology or bulk structure. Furthermore, This concentration of oxygen vacancies has a significant effect in improving the adsorption energy of nitrate and intermediates which significantly improves the NH 3 yield rate and FE as illustrated in Figure 9 [88].

3.1.3. Thermal Treatment (Calcination)

The calcination treatment requires thermal treatment at controlled temperatures for a defined time [89]. Zhang et al. [90] prepared a Copper-modified titanate cathode (Cal-Ti/Cu) at 350 °C for 30 min for nitrate reduction in boiling water. They demonstrated that Nitrate reduction efficiency with Cal-Ti/Cu (65.37%) is much higher than Ti (5.3%) and Ti/Cu (5.21%) at 80 °C. Furthermore, the calcination of electrocatalysts increases surface area, active sites, and corrosion resistance, preventing surface deactivation [90]. Another study by Liu et al. [91] prepared Ni foam/Carbon nanotubes (CNTs)/Cu electrodes at different temperatures (350 °C, 400 °C, and 450 °C). The maximum nitrate removal was achieved at 450 °C due to better binding of CNTs to Ni foam and improved CNT film properties with this specific temperature of calcination [91].

3.1.4. Thermal Treatment (Annealing)

Thermal treatment (annealing) involves heating a material to a precise temperature for a defined period and then cooling it under controlled conditions. Annealing is generally used in the middle of catalyst preparation and specifically for catalyst decomposition such as the conversion of Cu(OH)2 to CuO [92]. Electrocatalyst materials subjected to annealing showed significant improvements in nitrate reduction performance compared to those without annealing. Kuang et al. [83] showed that annealing increased BDD’s surface energy, allowing better copper attachment due to numerous dangling bonds, additionally the annealing process consolidated copper oxides on the BDD surface, which significantly enhanced electrode stability. The surface morphology of the BDD electrode remained unchanged before and after copper oxide removal, indicating its suitability for long-term use. Furthermore, the BDD electrode could be reused multiple times without significant changes in copper oxide morphology [83].

3.1.5. Hydrothermal Treatment

The hydrothermal treatment method is generally used in a high-temperature and high-pressure aqueous environment with an autoclave to grow or synthesize the specific size and shape of a monodispersed catalyst [93]. This treatment is generally followed by other specific treatments, as illustrated in the work of Chen et al. [94] who used immersed carbon cloth in a solution containing Bi ( NO 3 ) 3 · 5 H 2 O and then subjected to hydrothermal treatment in an autoclave with an initial impregnation step in a solution containing metal precursors ( Cu ( NO 3 ) 3 · 3 H 2 O ) to incorporate the metal into the material followed by annealing in the 350 °C muffle furnace for 10 min, to synthesize Cu 2 O   NPs / BiO 2 x   nanosheets. Hydrothermal treatment has a great influence on improving the oxygen vacancy-rich BiO2-x which enhances eNO3R efficiency [94].

3.1.6. Impregnation Method

The preparation of the electrocatalyst materials using the impregnation method requires immersing the noble metal in a solution, followed by drying and calcination to complete the process [95]. The impregnation method is commonly used to optimize the amount of noble metal to be added in order to minimize the loss of noble metals, as demonstrated in the study by Zhang et al. [96] who synthesized Pd Cu / γ Al 2 O 3 catalysts with different metallic palladium loading rates. As reported in their study, a palladium loading rate of 5% demonstrated the best performance among the various tested rates and was, therefore, selected for subsequent experiments [96].

3.2. Advanced in Situ Characterization Techniques

Electrode materials gained particular attention for their ability to ensure the eNO3R mechanism. Hence, to establish a link between the electrode surface structure, (including its chemical composition, crystal structure, coordination number, and interatomic distance) and the eNO3R performance, in situ characterization techniques have been employed during the electrochemical reaction. These techniques mainly include X-ray absorption spectroscopy (XAS), X-ray photoelectron spectroscopy (XPS), in situ electrochemical scanning tunneling microscopy (EC-STM), grazing-incidence X-ray diffraction (GIXRD), and shell isolated nanoparticle enhanced Raman spectroscopy (SHINERS) [57,97,98,99,100,101]. These techniques allow the real-time control of the reaction and the characterization of the catalysts under reaction conditions. Thanks to the use of these techniques, a qualitative understanding of the eNO3R reaction mechanism is possible. In addition, they can provide information on dynamic changes in the catalyst’s stereo geometry and electronic structure. Thus, the development of in situ techniques is of utmost importance to identify the active species involved in the eNO3R mechanism and provide a qualitative understanding of the reaction mechanism. The various in situ techniques are summarized in the following Table 3, which outlines the detected products and intermediates, principles, advantages, and disadvantages of each method.

4. Advanced Strategies for eNO3R Toward Ammonia and Nitrogen

eNO3R process has been extensively studied using a wide range of materials, including platinoids which have been used several times for nitrate reduction such as Pt [112,113,114,115], Ru [116], Pb [117], Rh [118,119], Ir [120]. Thus there are other materials that have been studied such as Sn [121,122], Zn [117], Ni [115,117]. Additionally, many researchers have studied alloys containing at least two components and having to contain at least one metal, such as platinoid alloys including Pt-Ir [120], Pt-Sn [123,124,125], Pd-Sn [125,126,127], or Pt-Rh [128]. Furthermore, non-metallic materials, including graphene, graphite, [129,130], carbon fiber [131], Boron doped diamond (BDD) [132,133], and so on, have emerged as promising catalysts for eNO3R, offering advantages such as low-cost, durability, and tunable electronic structures. In the search for efficient eNO3R catalysts, researchers’ efforts diverge to synthesize materials effective in reducing nitrates to nitrogen and ammonium. Figure 10 and Table 4 showed that copper-based catalysts have demonstrated interesting results in yield rates and faradaic efficiency (FE) compared to the value required for practical application (∼57 mg cm−2 h−1, faradaic efficiency FE > 90%) [134]. Various strategies have been explored to enhance Cu-based catalysts including facet manipulation, defect engineering, geometric regulation, fabricating copper composites with other metals or non-metals, bimetallic catalysts, encapsulation of active species, nanostructuring, and so on. Therefore, in the following section, the strategies used to convert nitrate to non-toxic gaseous nitrogen, in particular with copper and iron nanomaterials, will be outlined. As a matter of fact, these latter have yielded interesting results, as shown in Table 4, with the emphasis on their advantages. The engineering of zero-valent iron to advance the process is particularly addressed. Additionally, the strategy to convert ammonia generated through indirect oxidation by chloride-electrochemical advanced oxidation is discussed.

4.1. Copper-Based Electrocatalyst for the eNO3R: A Case Study for Ammonia Production

4.1.1. Copper-Based Single Atom Catalyst (Cu-Based SACs)

Cu-based SACs have demonstrated exceptional performance in the eNO3R owing to the intrinsic electronic configuration of Cu, which facilitates strong interaction with nitrogen oxyanions such as NO 3 . For instance, Cu incorporated into 3,4,9,10-perylenetetracarboxylic dianhydride (Cu-PTCDA) exhibits high selectivity (FE = 85.9%) and a substantial production rate (436 μ g . h 1 . cm 2 ) for NH 3 , surpassing other transition metals (Ag, Bi, Ir, Pt, Co, Fe, Ni). The Cu atom in Cu-PTCDA forms strong bonds with NO 3 , promoted by the efficient orbital overlap between NO 3 ’s 2p orbitals and Cu’s d orbitals [159]. Cu incorporated into nitrogen-doped carbon (Cu-N-C) further enhances eNO3R activity, particularly Cu-N-C-800, which shows superior NH4⁺–N selectivity (80%) compared to bulk Cu. The Cu-N-C-800 catalyst features Cu–N2 and Cu–N4 moieties that exhibit high adsorption energies for NO 3 and NO 2 , crucial for efficient eight-electron transfer in the eNO3R. This high adsorption capability enables Cu-N-C-800 to reduce NO 3 to NH 2 without releasing NO 2 into the electrolyte, highlighting its role in accelerating NH 3 production [160]. Contrasting Cu SACs with other transition metals, SACs reveal Cu’s unique advantages in the eNO3R. While Fe SACs also show promising performance, particularly in terms of NH 3 FE and yield rate, Cu SACs exhibit superior thermodynamic properties and kinetics due to stronger NO 3 coordination with Cu–N4 moieties [161]. Cu-based SACs represent a promising avenue for advancing electrocatalysts in the eNO3R, leveraging their unique electronic properties and strong interactions with nitrogen species to achieve high selectivity and efficiency in NH 3 production.

4.1.2. Facet Engineering

Copper-based catalysts exhibit notable catalytic activity, especially when possessing a special electronic structure. Among these, Cu has demonstrated the highest activity, primarily yielding ammonia as the main product. The performance of Cu electrodes in nitrate electroreduction is closely linked to their structural characteristics. Additionally, the specific atomic structure of Cu catalysts, such as Cu(100) and Cu(111) surfaces, significantly influences their electrochemical behavior. In alkaline media, Cu(100) is more active in forming hydroxylamine compared to Cu(111), and the selectivity for nitrate reduction products can be controlled by adjusting the ratio of Cu(200) to Cu(111), which can be further tuned using surfactants to optimize nitrogen selectivity [162,163,164]. Studies on Cu2O electrocatalysts have shown that manipulating exposed facets, particularly the (100) facet, enhances ammonia yield compared to other facets like (111), demonstrating the importance of surface structure in catalytic performance [165]. Bae and Gewirth [166] compared eNO3R activity on Cu(111) and Cu(100) surfaces in acidic electrolytes. As a result, Cu(111) demonstrated superior activity thanks to the easier formation of surface oxides, as evidenced by cyclic voltammetry (CV) showing significant reduction currents at less negative potentials compared to Cu(100) (Figure 11a). In addition, in situ electrochemical scanning tunneling microscopy (EC-STM) revealed that surface oxidation on Cu(111) enhanced its deoxygenation activity towards NO 3 . The Cu2O formationon Cu surfaces influences their NO 3 -to- NO 2 conversion efficiency which exhibits higher susceptibility to oxidation and facilitates oxygen transfer from NO 3 and subsequent reduction to NO 2 at less negative potentials. Thus, highlighting Cu2O role in catalyzing the eNO3R process [166].

4.1.3. Copper-Based Bimetallic Catalysts

Copper bimetallic catalysts, particularly CuNi alloys and Cu/Ni heterostructures, have shown remarkable potential in electrochemical nitrate reduction due to their synergistic effects and enhanced performance [167]. CuNi alloy cathodes, such as Cu80Ni20 and Cu50Ni50, have demonstrated superior activity in nitrate reduction compared to mono-component metals. The synergistic mechanism between Cu and Ni, where Cu facilitates nitrate reduction while Ni preferentially adsorbs hydrogen atoms, enhances the reduction process by hindering competitive hydrogen adsorption [65,168]. Operando spectroscopy studies have elucidated the structural-performance relationship, revealing the upshifting of Cu d-band center towards the Fermi level upon alloying with Ni. This enhances the adsorption energies of intermediates, reducing overpotential and promoting nitrate reduction [65]. Porous CuNi alloy electrodes, prepared via hydrogen evolution-assisted electrodeposition, exhibit high stability and effective nitrate reduction, attributed to their large surface area and enhanced mass transport [169]. The mechanism proposed by Zhang et al. [84] involves the formation of infinite galvanic cells, facilitating electron transfer and adsorption of nitrate on Cu sites, leading to enhanced reaction kinetics and selectivity (Figure 12). While copper bimetallic catalysts offer promising advantages in nitrate reduction, challenges such as leaching, and particle aggregation need to be addressed for prolonged operation.

4.1.4. Copper Oxide

Copper oxides exhibit significant electrocatalytic activity in the eNO3R, particularly when engineered with oxygen vacancies (OVs). These vacancies serve as chemisorption sites and facilitate charge transfer, enhancing the catalytic performance of the materials [170]. Daiyan et al. [171] prepared CuO nanoparticles via scalable flame spray pyrolysis (FSP) followed by mild plasma treatment to controllably induce defects. The optimized plasma-treated CuO demonstrated enhanced NH 4 + yield (∼292 μ mol . h 1 . cm 2 at −0.6 V vs. RHE) compared to non-defective CuO. The presence of OVs in CuOx not only strengthens NO3⁻ adsorption energy but also inhibits the hydrogen evolution reaction (HER), crucial for improving selectivity in the eNO3R. DFT studies further indicate that most NO3R intermediates are more stabilized on CuOx with two OVs compared to Cu2O, highlighting the effectiveness of defect engineering in optimizing catalytic performance [171].

4.1.5. Cu-Based Heterostructure Engineering

Cu-based heterostructures have been extensively studied for their electrocatalytic performance in the conversion of NO 3 to NH 3 . For instance, Ren et al. [172] designed Cu/Pd/CuOx heterostructures with a porous morphology, which facilitated electron transfer between Cu, Pd, and CuOx components. This heterostructure enhanced NH 3 production significantly, achieving a NH 3 production rate of 1510.3 μg h−1 mg−1, FE of 86.1%, and NH 3 selectivity of 90.06% (Figure 13a). The electronic interactions within the Cu/Pd/CuOx interface promoted the adsorption of intermediates, thus suppressing competitive hydrogen evolution reactions and improving overall catalytic efficiency [172]. Another study by Xu et al. [173] explored Cu2O corner-etched octahedra supporting Pd nanoparticles (Pd-Cu2O), where Pd acted as the active center for *H adsorption and catalyzing NH 3 production, while Cu2O provided additional active sites through electrochemical reduction to Cu0/Cu+. This system exhibited outstanding NH 3 formation characteristics, including a NH 3 formation rate of 925.11 μg h−1 mg−1, selectivity of 95.31%, and FE of 96.56% (Figure 13b). The oxygen defects in Cu2O were crucial for enhancing NO 3 adsorption and weakening of the N–O bond, thereby promoting efficient NH 3 synthesis [173].

4.1.6. Copper-Based Transition Metal Oxide Composite

Transition metal oxide composite electrodes, particularly those incorporating Cu, have shown promising electrocatalytic activity in nitrate reduction. For example, combining copper oxide (CuO) with cobalt oxide (Co3O4) on titanium (Ti) electrodes results in a dramatic increase in nitrate reduction efficiency, reaching nearly 100%. The unoccupied π * orbital and the unclosed d-orbital shells of CuO facilitate nitrate adsorption and its subsequent reduction to nitrites, accelerating the overall nitrate reduction process [174]. Yang et al. [175] proposed a Ti/Cu5ZnOₓ composite electrode that offers more active sites for the degradation of intermediate nitrogen compounds, especially in alkaline solutions compared to Ti/CuOₓ. The addition of zinc enhances the reduction in nitrite to nitrogen gas (N2), achieving reduction efficiencies comparable to those of precious metal catalysts like palladium [175]. The process of creating composite electrodes, including those involving copper, is complex and has a low success rate. This complexity limits their large-scale preparation and practical application.

4.1.7. Copper-Based Inorganic/Organic Material

Loading inorganic materials such as graphene and phosphorus onto copper increases the electrode’s stability and specific surface area, which in turn improves the electrocatalytic reduction efficiency of nitrate. Graphene-modified Cu electrodes exhibited a 3.5-fold increase in reduction peak densities compared to pure Cu electrodes, showcasing improved electrocatalytic performance [176]. Polypyrrole (Ppy) coating on copper electrodes enhances the selectivity for ammonia production during nitrate reduction. The mechanism demonstrated that in the low electrode potential range (0.4–0.8 V), Ppy facilitates proton adsorption and electron transfer, which generates atomic hydrogen to convert nitrate to ammonia, improving selectivity and reducing unwanted by-products like nitrite [177]. Additionally, flower-like polycrystalline Cu grown on carbon paper via electrodeposition exhibited remarkable electrochemical activity with an NH 3 formation rate of 101.4 μmol h−1 cm−2 and an FE of 93.91% in neutral solution. The structure facilitated efficient mass transfer of reactants, ensuring sustained performance over multiple cycles [178].

4.1.8. Copper-Based Nanomaterials

Advances in the rational design of micro-/nanostructures have enhanced their efficacy. Notably, copper nanomaterials have gained attraction toward nitrate reduction to ammonia. Cu-nanomaterials used to date are nanosheets, nanowire arrays, nanocubes, nanoparticles, 3D nanobelts, and nanoclusters. Cu-nanosheets (CuNSs) have demonstrated significant efficiency in several studies such as Fu et al. [137] who investigated CuNSs with {111} basal planes, Cu nanocubes ({100}), and irregular Cu nanoparticles. The results indicate that CuNSs displayed exceptional NH 3 production selectivity (99.7% FE) at a low overpotential of −0.15 V (vs. RHE). Furthermore, they exhibited distinct reduction peaks in linear sweep voltammetry (LSV), attributed to their specific facets and efficient NO 2 generation, crucial for high eNO3R performance [137]. Another nanomaterial that has exhibited notable efficiency is nanowire arrays (NWAs), NWAs have been effective due to their reconstructed surface during catalysis, which suppresses HER and achieves high faradaic efficiencies towards ammonia production (up to 95.8%) [58]. Incorporating metals like ruthenium (Ru) on Cu nanowires (Ru-CuNW) has led to exceptionally active catalysts, achieving reduction currents of 1 A cm−2 at low overpotentials (−0.13 V vs. RHE) and converting over 99% of nitrate into ammonia [179]. Cu-nanobelts have been found to substantially increase the specific surface area, thus boosting the catalytic activity by facilitating better nitrate transfer and accelerating electrocatalytic kinetics. These nanostructured copper electrodes exhibit a high FE of 95.35% and a maximum NH 3 production rate of 650 mmol h−1 g−1cat with a potential 0.15 V vs. RHE, outperforming traditional foam copper electrodes. The enhanced efficiency is attributed to their large surface area and high charge transfer capability, which promotes effective nitrate adsorption and reduction [163]. The advantage of using nanomaterials is their operation at low potentials, which minimizes energy consumption. However, scalability remains a significant barrier to the use of Cu-nanomaterials for nitrate reduction to ammonia.

4.2. Advanced Strategies Toward eNO3R to Nitrogen

The challenge facing electrochemical nitrate reduction to nitrogen is the low selectivity compared to HER, which means the use of high energy and hinders industrial implementation. In this section, we will discuss the two strategies to advance the process starting with the use of specific materials as in the case of iron-based materials and specifically nanoscale zero-valent iron. Secondly, we address the removal of ammonium by-product generated by electrochemical nitrate reduction with the indirect advanced oxidation with chloride compound.

4.2.1. Nanoscale Zero-Valent Iron (nZVI)

nZVI for nitrate reduction to nitrogen, leveraging various strategies to enhance their catalytic performance. Nanoscale zero-valent iron supported on ordered mesoporous carbon (nZVI/OMC) has demonstrated excellent performance, achieving a maximum nitrate removal capacity of 315 mg N g−1 Fe with a high selectivity towards N 2 of up to 74%. The uniform dispersion of iron nanoparticles within the carbon matrix and the large surface area of the electrode contributed to its superior electrocatalytic activity. Additionally, the presence of hydrogen generated during the reduction process not only acted as a reductant but also protected the nZVI from oxidation, thereby enhancing its stability and performance over extended operation periods [143]. Moreover, efforts have been made to explore the mechanism of nitrate reduction mediated by iron-based electrodes, such as polymeric beads supported nZVI (nZVI/D201), which further elucidated the role of in situ formed Fe0@FexOy-Fe(II) structures in facilitating the electrochemical reduction process. These advancements underscore the potential of iron-based materials in developing efficient and stable electrodes for nitrate reduction, although challenges related to complex preparation processes remain a consideration for practical applications [144]. Bimetallic nanoparticles, such as iron–nickel (Fe/Ni) and iron-palladium (Fe/Pd), as well as trimetallic nanoparticles like palladium–copper–iron (Pd–Cu–nZVI), represent advanced strategies to boost nitrate reduction efficacy and facilitate the production of non-toxic nitrogen gas. These nanoparticles function synergistically, where the iron component of nZVI acts as an electron donor, initiating nitrate reduction. In bimetallic systems, metals like palladium enhance the conversion of nitrite to nitrogen gas, thereby promoting a more complete detoxification of the water [180]. Furthermore, adjusting the composition ratios, such as maintaining a 2:1 ratio of Pd to Cu in Pd-Cu-NZVI, optimizes the selectivity toward nitrogen gas formation [180]. Durability remains a challenge for nZVI catalysts due to leaching and oxidation. Zhang et al. [181] addressed this by protecting Fe nanoparticles with ultrathin graphene nanosheet layers (Fe@Gnc), maintaining over 96% NO 3 conversion rate and N 2 selectivity after 960 h of cycling. This graphene nano chainmail prevents active site agglomeration and accelerates electron/ion transport, suggesting a potential strategy for enhancing the durability and efficiency of Fe-based catalysts for nitrogen conversion processes [181].

4.2.2. Indirect eNO3R to Nitrogen via Cl-Electrochemical Advanced Oxidation

The ammonia generated during eNO3R can be oxidized to nitrogen gas by free radicals ( HO , Cl , ClO ) or active chlorine, generated during electrolysis [182]. Studies confirm that HO alone does not affect ammonia specifically in the absence of chloride [183,184]. In contrast, Cl , ClO , and active chlorine exhibits high ammonia oxidation capacity, achieving nitrogen selectivity exceeding 90% [183,185,186]. Ammonia oxidation in the Cl-EAO process primarily occurs via active chlorine and chlorine radicals. Initially, Cl ions adsorb onto the anode surface and lose electrons, forming Cl . These radicals may combine to form Cl2 or react with Cl to produce Cl2, which then diffuses into the solution and reacts further to form HClO or ClO . Additionally, HO is generated on the anode surface, some of which reacts with Cl in the solution to form Cl . Both HO and Cl can further react with HClO or ClO to form ClO . These species react with ammonia through breakpoint chlorination and chlorine radical oxidation mechanisms, as illustrated in Figure 14 [187]. In the Cl-EAO process, ammonia oxidation occurs at the electrode-solution interface, where the production rate of active chlorine and chlorine radicals is crucial and often limits the overall reaction rate [188]. The rate of chlorine radical production, particularly the production of Cl , is identified as a slower step compared to other reaction rates [189]. During breakpoint chlorination, the mass transfer of active chlorine to water becomes another rate-limiting factor, influencing the rate of chlorine radical oxidation more than breakpoint chlorination itself. At the cathode, active chlorine in the solution is reduced back to Cl . Over time, the concentration of active chlorine in the solution stabilizes, maintaining a stable Cl concentration due to this reduction process. Despite some Cl2 escaping from the solution, the impact on active chlorine production remains minimal [187]. Furthermore, the ammonium generated during the process can be influenced by different factors, as demonstrated in Table 5.
The debatable aspect of this process is the formation of chloramine intermediates, which include NH2Cl, NHCl2, and NCl3, as well as chlorine-containing byproducts (CBPs) such as ClO 2 ,   ClO 3 , and ClO 4 [192,196]. Researchers have found that chloramine concentrations are always below certain specific levels, such as 0.44 mg/L [197]. A study using a graphite/PbO2 anode showed that the total chloramine concentration was always less than 2.2% of the initial ammonia concentration [192]. Therefore, it is important to focus on the concentration and behavior of chloramines and CBPs in the Cl-EAO process to enhance the understanding and application for the removal of ammonium generated during the electrochemical reduction in nitrates.

5. Application of eNO3R to Brackish Groundwater: A Case Study

The groundwater is typified by elevated levels of chloride, potassium, sodium, magnesium, calcium, sulfate, bicarbonate, and carbonate, culminating in the formation of a brackish environment. Furthermore, it is important to acknowledge that the quality of groundwater is subject to various influences, including environmental and geological factors such as residence periods, climate conditions, soil properties, flow patterns, terrains, the chemistry of the replenishment zone, local geology, and water intermixing [198,199]. Additionally, nitrate derived from fertilizers easily leaches through soils and enters groundwater, especially following periods of heavy rainfall. The application of electrochemical processes remains the most rational and applicable compared to other technologies. However, before applying this method to large-scale water treatment, it is essential to discuss the parameters that significantly affect the nitrate reduction process, including pH, nitrate concentration, coexisting ions, current density, and applied voltage analyze the water matrix, and identify potential issues that may arise.

5.1. pH

During the reduction process, electrocatalytic nitrate reduction consumes protons in both the rate-limiting step and the subsequent reduction steps, which means that the pH increases with the electrolysis process. A thorough investigation was carried out to identify the optimal circumstances in which the pH could exert an influence. eNO3R process indicated that the best removal rates were achieved under acidic conditions, in contrast to neutral conditions that led to poor results [105]. Lowering the pH of the electrolyte solution, used in electrochemical nitrate reduction processes, can also provide additional benefits as reported by Yao et al. [200] where an electrolyte with a lower pH facilitates the formation of ions (e.g., Ti 2 + or / and   Ti 3 + ), which can reduce nitrates and nitrites. On the other hand, the Pourbaix diagram of the N 2 H 2 O system (Figure 15) indicated that the hydrogen evolution reaction (HER) cannot be ignored as it competes with the eNO3R due to electron consumption. The FE of HER is higher in acidic media than in alkaline media, which results in a significantly lower FE of nitrate reduction in acidic media compared to alkaline media. Due to competition with H 2 , the direct charge transfer process has low efficiency. However, the indirect electrochemical autocatalytic process (Vetter and/or Schmid) can be improved under acidic conditions due to the compensation of this effect. Zhang et al. [181] found that a neutral pH of 7, was the most favorable for electrocatalytic nitrate reduction, as acidic conditions caused hydrogen bubbles to form and prevent contact between the active site and the electrolyte, while higher pH led to the active surface being occupied by adsorbed hydroxide. On the other hand, Beltrame et al. [201] indicated that solutions with pH values between 6.0 and 6.5 induced a lower conversion of nitrate than in alkaline media. Notably, nitrite formation was significantly minimized, and gaseous compound generation was higher in solutions at given pH values due to the availability of H ( aq ) + , which in turn promotes gaseous compound generation.

5.2. Effect of Chloride Anions

The primary ions present are chlorides, which are found in substantial concentrations in this type of water. Generally, the pathway for ammonium production by eNO3R is limited in this case, as it necessitates a cathode with high faradaic efficiency (FE) and an anode with the highest oxygen evolution reaction (OER) compared to chlorine ion oxidation reactions (ClOR). Additionally, research on this topic is currently limited. Thermodynamically, ClOR has higher potentials compared to OER, making it prone to activation under high cell voltages required for industrial current densities. In acidic conditions, ClOR manifests as a chlorine evolution reaction (ClER), while in alkaline conditions, it forms hypochlorite, both of which are corrosive and detrimental to catalyst longevity [203]. Materials such as Mn-based oxides and non-noble metal catalysts like transition metal hexacyanometallates [204], cobalt phosphates [205], selenides [206], and borates [207] have shown promise in stabilizing OER in NaCl-containing electrolytes. However, they struggle to deliver industrially relevant current densities at overpotentials below 490 mV. Recently, electrocatalysts resistant to chloride ion oxidation for industrially relevant current densities and low overpotentials have been extensively studied, even though their primary application is aimed at hydrogen production from seawater. Researchers such as Dong et al. [208] have developed a robust electrocatalyst based on nickel–iron hydroxide with low nucleophilicity. The EA-Ni0.82Fe0.18-OH electrocatalyst provides industrially relevant current densities of 400 and 1200 mA cm−2 in simulated alkaline seawater, with respective overpotentials of 312 mV and 411 mV. Moreover, it exhibits impressive durability at current densities of 500 mA cm−2 over 50 h of simulated alkaline seawater electrolysis [208]. We need to advance research in developing anodes that will play a crucial role in the large-scale application of ammonium production from brackish waters. On the other hand, the pathway for nitrogen selectivity remains reasonable and motivates its large-scale application, especially with the presence of a high chloride concentration in this type of water, which plays a significant role in converting generated ammonium to gaseous nitrogen, as discussed in the previous section (Cl-electrochemical advanced oxidation).

5.3. Coexistence Ions

The influence of the nature of the coexisting ions has been largely reported. For instance, Katsuunaros et al. [122] noted that the reduction in nitrate increases slightly with an increase in NaCl concentration. Additionally, according to the Li + < Na + < K + < Cs +   series of the supporting electrolyte, the reduction rate also increases significantly. Likewise, they demonstrated that the rate of reduction is notably higher in the presence of multivalent cations, namely Ca 2 + and La 3 + as well as NH 4 + [122]. Moreover, Shen et al. [209] showed that when the concentration of Na 2 SO 4 was high, the rate of removal of NO 3 was rapid and this can be explained by the fact that a higher concentration of electrolytes induced a higher rate of electron transfer and the effect of electrolytes on the selectivity to N 2 was remarkable. The effect of the supporting electrolyte cation on nitrate reduction has been explained by the theory of “cationic catalysis”, which suggests that the cation of the supporting electrolyte acts as an attracting center for the reduced anion by forming an instantly neutral ion pair. The latter is not repelled by the negatively charged electrode, increasing the rate of nitrate reduction [122]. Manzo-Robledo et al. [210] evaluated the effect of K + and Na + cations on nitrate reduction in an electrolyte solution using real-time on-line differential electrochemical mass spectrometry. The results indicated that the presence of K + led to the production of N 2 as the main product, while Na + favored the production of N 2 O . The study also reported a moderate HER in K + -containing electrolyte solutions, promoting the reaction between NO 3 and H 2 species, while the HER kinetics was more dominant in Na + -containing electrolyte solutions, leading to a suppression of molecular nitrogen generation. These results highlight the significant influence of different cations in the supporting electrolyte on the products and kinetics of nitrate reduction [210]. Fedurco et al. [211] conducted also a study to investigate the influence of La 3 + and Al 3 + cations on nitrate reduction at a polycrystalline silver electrode. The outcomes showed that the presence of La 3 + caused a positive shift in the onset potential, indicating an enhancement of nitrate reduction. This may be attributed to acid-base catalysis, with La(III) species acting as proton donors and/or OH ion acceptors at the electrode surface. The research studies also exhibited that films deposited onto the electrode surface consisted of La ( OH ) 3 and/or Al ( OH ) 3 in the solutions containing La 3 + and Al 3 + ions, respectively. The La(III) species in the solution then diffused to the electrode surface as nitrate complexes or ion pairs, thereby promoting the reduction reaction. In contrast, Al 3 + cations, which do not form ion pairs with NO 3 anions were found to be less effective in promoting the nitrate reduction reaction [211]. In contrast, the rate of reduction was reduced by the anion of the supporting electrolyte. Specifically, I exhibited the greatest inhibitory effect, followed by Br , Cl , and F . Beltrame et al. [201] noticed that pH control by sulfuric acid affected nitrate reduction, as the SO 4 2 anion present in the solution can interfere with nitrate adsorption in the active sites of the catalyst, leading to a diminution of nitrate reduction.

5.4. Cathode/Anode Passivation

The presence of calcium, magnesium, and carbonates within brackish groundwater constitutes a significant impediment to the development of a large-scale process. In high pH environments, cations like Ca2⁺ and Mg2⁺ can contribute to the formation of a passivating surface layer on the cathode duo to electrophoretic migration phenomena. Furthermore, calcium and magnesium readily react with hydroxides formed during electrolysis and carbonates to produce insoluble compounds that form a layer that deposits on the cathode surface, thereby reducing the efficiency of nitrate reduction to nitrogen and ammonium [212]. Indeed, among the strategies proposed to reduce cathode passivation is using pulsed current instead of direct current [213]. So far, there have been few studies aimed at investigating the effect of pulsed current such as Li et al. [101] who demonstrated that the use of pulsed electrolysis (PE) effectively inhibited the HER across all PE experiments compared to constant potential electrolysis (CE), accompanied by lower energy consumption. Additionally, Cu single-atom-modified gels (Cu SAGs) used in the study also exhibited good stability after ten successive catalytic cycles, with a slight performance decrease noted after the seventh cycle. Therefore, it is crucial to focus on this topic in future research concerning this issue, while also emphasizing the stability of the cathode [101]. Now the question arises: Is this strategy advantageous for the anode? The clearest aspect is that pulsed currents can momentarily disrupt the accumulation of ions at the electrode surface, which contributes to delaying the decay of the anode potential. However, the question of whether the application of dual-pulsed or multi-pulsed electro-oxidation is more beneficial has already been addressed in the research by Jiang et al. [214] in which they showed that Multi-Pulse Electro-Oxidation has the highest total nitrogen removal compared to Direct Current Electro-Oxidation and Dual-Pulse Electro-Oxidation. This mode of MP-EO perfectly improves the mass transfer and specifically during the Toff period between pulses, the oxidants and intermediates produced can efficiently transfer into other areas, thereby improving the overall efficiency of the process. However, the dual-pulsed approach, which was expected to yield favorable results, showed the opposite in this study. This outcome is attributed to the type of anode used, specifically stainless steel. It is evident that this material is not designed for the specific chlorine evolution reactions required [214]. Other specific types of materials, such as Ti/Ru-Ir, have not yet been tested in this context and for the same specific conditions to determine whether the dual-pulsed effect on ammonium removal efficiency is truly superior to that of direct current.

5.5. Nitrate Concentration

The variation in nitrate concentration also caused changes in the kinetic order of the electrode during electrochemical reduction. In acidic solutions, Dima et al. [39] studied the electrocatalytic reduction in nitrate at low concentrations using noble and transition-metal electrodes. They found that cathodic currents increased as the concentration of nitrate augmented. The Tafel slopes indicated that the rate-determining step involved the first electron transfer and was close to or greater than 120 mV. dec−1. The nitrate reaction order was found to be approximately 0.51 for Pt and 0.34 for Rh from the logarithm linear fitting of the current densities versus the logarithm of nitrate concentration [39]. Nitrate adsorbates occupied the active surface area, causing the butterfly-like hydrogen adsorption and desorption pattern to disappear. Also, Taguchi et al. [215] assessed the impact of nitrate concentration on the cyclic voltammogram of Pt(110) in 0.1 M HClO 4 + 10 mM KNO 3 , and concluded that the reaction follows a zero-order dependence on nitrate concentration. The adsorbed NO 3 is believed to be in a state of quasi-equilibrium with the NO 3 in the bulk solution before the rate-determining step of the reduction reaction. The average Tafel slope value of −66 ± 2 mV. dec−1 indicating that the RDS does not involve any electron transfer. Instead, a pure chemical reaction occurs on Pt(100), wherein NO 3 reacts with H ads [215]. In another study, Katsounaros et al. [122] studied the kinetics of nitrate reduction at high concentrations and determined that the reduction follows Langmuir–Hinshelwood kinetics. Whereas the reduction is characterized by first-order kinetics when the concentration is below 0.3 M. However, at higher concentrations, zero-order kinetics dominate due to the blockage of surface sites, indicating that the regeneration of active sites controls the reaction. Also, the selectivity towards N 2 increased from 70% to 83% with an increase in nitrate concentration from 100 to 1500 mg L−1 and remained relatively stable at higher nitrate concentrations. Conversely, the selectivity of ammonia showed a declining trend from 25% to 11%. Furthermore, the faradic efficiency increased from 25% to 78% with an increase in nitrate concentration, and this trend was observed when 95% of nitrate was reduced [122].

5.6. Influence of Applied Voltage and Current Density

The applied voltage or potential had a relatively strong effect on nitrate reduction. Ding et al. [216] demonstrated that the selectivity of nitrate reduction in products is strongly dependent on the potential applied. In addition, Reyter et al. [217] showed that at a potential of −0.9 V, nitrite is the only substance generated at the cathode, which is then oxidized at the anode, thus decreasing the process efficiency [217]. Also, Shen et al. [209] found that the application of increasingly negative cathodic potentials resulted in an improvement in nitrate removal efficiency. Specifically, a more negative potential facilitated the rapid removal of nitrate due to the thermodynamically favorable reduction in nitrite to N 2 occurring at more negative potentials. On the other hand, Zhao et al. [218] showed a decrease in selectivity for N 2 due to the reduction in Cu 2 O to metallic Cu caused by the application of a high cathodic potential. It is worth noting that the application of very negative potentials may lead to the formation of undesirable by-products such as ammonia or hydrogen gas. Hence, the cathodic potential should be optimized to achieve both high nitrate removal efficiency and high selectivity for nitrogen gas.
In addition to the cathodic potential, the current density also plays a significant role in the performance of nitrate removal. In this regard, Yao et al. [200] demonstrated that nitrate removal efficiency increased progressively with increasing current density. It also caused an acceleration of electrode corrosion and produced reducing agents that enhanced nitrate reduction. Gao et al. [219] found that as the current density increased, more energy was utilized to produce hydrogen bubbles. This, in turn, resulted in a reduced selectivity of gaseous nitrogen and an increased selectivity of ammonia, ultimately leading to a lower overall rate of total nitrogen removal.

5.7. Challenge to Move from Lab Scale to Pilot eNO3R

The eNO3R reactors in most of all previous research do not exceed 500 mL. However, large-scale application requires several factors to ensure successful applicability, specifically for brackish groundwater. One of the initial considerations in designing a reactor is selecting between divided and undivided reactors. For nitrate removal from brackish groundwater, the choice between a single chamber cell (SCC) and a double chamber cell (DCC) is directly linked to the objective. If the goal is to convert nitrates into gaseous nitrogen, the choice should be an Undivided Flow electrochemical reactor (UFECR). However, if the objective is ammonium production and an inert anode for chlorides is used, as discussed in the previous section, the preference remains for UFECRs. On the other hand, if an anode with higher ClOR is used, then the choice is a Divided Flow Electrochemical reactor (DFECR), as explained in Figure 16.
Studies employing continuous mode which allows for easier scale-up and industrial application are few. Typical continuous reactors are illustrated in Figure 17. For example, Abdulah et al. [220] used a homemade flow cell with a porous modified copper cathode and platinum anode to convert nitrate to ammonium; the results showed a maximum NH 4 + selectivity of 96% and a high faradaic efficiency for NH 4 + formation (FE 72%) was recorded with a solution of pH 7.2, and a flow rate of 2 mL min−1 [220]. Even though the water matrix used differs from that of brackish groundwater, the results remain significant and encourage the development of DFECRs for ammonia production. Another study by Makover et al. [221] demonstrated significant results regarding the use of continuous mode; however, the application of periodic current reversal complicates the use of these processes for large-scale applications, particularly for natural waters, due to cathode dissolution [221]. Another study utilized a 500-L pilot continuous-flow reactor for nitrate reduction to ammonia [222]. The stable operation achieved a NH 4 + –N selectivity of 92.9% at a hydraulic retention time (HRT) of 12 h and a current density of 20 mA cm−2. However, the long hydraulic retention time required for a significant decrease in the initial NO 3 –N concentration is one of the weaknesses of this operation. This is due to non-uniform flow and the presence of numerous dead zones in the reactors, demonstrating that using a continuous-flow reactor for nitrates is not always straightforward [222]. However, the major challenge is transitioning to continuous mode with good hydraulic flow and minimal dead zones to ensure uniform nitrate reduction. Therefore, selecting the appropriate geometry and operating conditions for electrochemical reactors (ECRs) to achieve high current efficiencies (close to 100%) toward nitrate reduction to nitrogen or ammonium in brackish groundwater, requires intensive research into transport phenomena such as hydrodynamics, potential-current distributions, heat, and mass transfer. Mathematical modeling and Computational Fluid Dynamics (CFD) simulations have facilitated the design, characterization, and scaling up of ECRs. Fluids dynamics phenomena, fluid pressure drop, and current-potential distribution for ECRs have been widely studied using CFD. However, mass transfer phenomenon is an important aspect to be analyzed with CFD for eNO3R [223]. Many mass transfer issues in electrochemical reactors have been solved by neglecting the migration term, which simplifies practical assumptions, as in a recent study by Oriol et al. [224] who utilized a rotating cylinder electrode reactor (RCE) for nitrate reduction. The results indicate that 90% nitrate was removed in 10 min for a solution comprising 10 mM NO3 and 500 mM K2SO4, at 1000 rpm and 25 °C confirming the RCE reactor’s effective performance for nitrate reduction. Furthermore, CFD showed that mass transport in RCE reactor is linked to the hydrodynamic pattern, highlighting the critical role of fluid dynamics in reactors, particularly in nitrate reduction, to ensure effective mass transfer. Therefore, thorough studies using CFD are necessary to design future types of electrochemical reactors [224]. Another strategy was recently innovated by Zhou et al. [225] to address the mass transfer issue for nitrate reduction, utilizing a flow-through zero-gap electrochemical (ZGEC) reactor. The system demonstrated significant success, achieving 100% nitrate conversion and a rapid reduction kinetics of 0.07676 min−1 [225]. These two studies confirm that focusing on mass transfer is crucial to ensure the success of eNO3R for large-scale applications.

5.8. Economic Analysis

Most publications in nitrate removal by electrocatalysis focus on the performance of electrode material but rarely discuss capital expenditures (CAPEX) and operational expenditures (OPEX). However, it is obvious that since the process is still in the R&D phase, significant steps remain before industrial application, such as the construction of efficient reactors [226]. Nonetheless, this does not dismiss the importance of making estimates to determine whether the process is economically viable or not. Generally, 90% of synthetic ammonia is produced using the Haber–Bosch process, but there are other methods such as electrocatalytic and photocatalytic N 2 reduction. Therefore, eNO3R remains innovative compared to existing methods [227,228]. The comparison between energy consumption and energy source for the different processes shows that eNO3R is the most promising due to the total energy demand during the process. Furthermore, the possibility of using solar energy remains a strong point of the process. The production cost of ammonia by the Haber–Bosch process is $560 per metric ton, which is close to the estimated value for production by eNO3R in the case of using the strained ruthenium nanocluster ($776 per metric ton) [229]. Additionally, Joshua M. McEnaney et al. [230] conducted an economic study based on NH4NO3 as the final product instead of NH 3 . They validated the production of this type of fertilizer at a cost that matches or is lower than the typical United States Department of Agriculture (USDA) cost range for NH4NO3. Nonetheless, strictly with low electricity costs and reasonable Faradaic efficiencies (Figure 18) [230]. On the other hand, the techno-economic analysis by Jianan et al. [231] confirmed the relationship between the energy cost per kilogram of (NH4)2SO4 and the Faradaic efficiency/energy consumption. They also emphasized that the concentration of nitrates in effluents plays a crucial role in minimizing costs. So, a high concentration of 1.0 M nitrate found in wastewater is preferable to 0.01 M found in some brackish groundwater to minimize the cost of ammonia production [231]. Overall, eNO3R for ammonia or even fertilizers production has potential economic benefits and is a promising candidate to replace the Haber–Bosch process specifically for high concentrated nitrate contaminated water. Regarding the concentration of nitrate in brackish groundwater, it is preferable to convert it to gaseous nitrogen, which is economically viable. However, the cost of electrode materials and energy consumption are critical to the application of this process. So far, there have been no economic analyses of this process because it is still in the stage of researching selective materials. However, it is essential to consider the cost of the materials used in the process, focusing on non-noble or non-metallic materials in future research to compete with existing processes.

5.9. Sustainability Analysis

As previously discussed, electrocatalytic nitrate reduction (eNO3R) processes are gaining attraction as alternative processes for the removal of nitrate from various water effluents, including wastewater and groundwater [232]. Indeed, the main reason for this is that these techniques are regarded as environmentally friendly and efficient approaches. Nonetheless, the main stumbling blocks to implementing these technologies on an industrial scale are energy consumption and cost [232,233,234].
Thus, renewable energies (RE), particularly solar energy, are seen as an alternative to conventional electricity, as they could convert these processes into truly sustainable, eco-friendly, and autonomous technologies, as well as lessen the environmental impacts related to fossil energy [232,235]. In fact, solar energy can be mobilized to power the electrochemical cells employed in nitrate reduction, which on the one hand will lead to an increase in the NH3 faradaic efficiency, and on the other hand, reduce reliance on non-renewable energy sources and subsequently decrease the overall carbon footprint. In this context, Kani et al. [236] harnessed solar energy for the NO 3 electroreduction to NH3 using oxide-derived Co as the catalyst. As a result, they attained a high NH3 faradaic efficiency of 92.37 ± 6.7% current density of 565.26 mA.cm−2 at −0.8 V vs. RHE, and a high solar-to-fuel efficiency of 11%. Furthermore, hybrid energy systems that couple solar power with other renewable sources like wind or hydroelectric power could guarantee a more consistent and reliable energy source, thus overcoming the intermittency problems inherent in solar power alone. Additionally, energy storage solutions including batteries or supercapacitors can also boost system stability and effectiveness, maintaining continuous operation during times when solar power is unavailable [232].
The integration of solar and wind RE enhances the environmental and sustainability aspects of electrochemical processes but can also have adverse environmental impacts. For instance, water and land use, habitat loss, and the use of harmful materials in manufacturing can be viewed as potential negative effects in the case of photovoltaic (PV) solar panels [237]. Similarly, the deployment of wind turbines can lead to the depletion or degradation of habitat for wildlife, fish, and plants. Rotating turbine blades can also pose a threat to flying wildlife, and cause “noise pollution” [238]. Whilst many studies have documented these environmental impacts, it remains the case that renewable energies contribute to reducing conventional energy consumption and carbon emissions, as well as supporting local development [239,240]. It is, therefore, vital to carry out life cycle assessment (LCA) studies to list and evaluate all potential environmental impacts, so as to be able to draw up an in-depth action plan to reduce the harmful impacts listed above and take advantage of the positive points of implementing renewable energies.
In parallel, a Life Cycle Assessment of electrocatalytic nitrate reduction is essential to fully assess its environmental sustainability. LCA evaluates the environmental emissions and impacts on human health associated with the entire life cycle of the process, from raw material extraction to final disposal [241,242]. This analysis must consider factors such as energy consumption, emissions, and waste production. When comparing these impacts with those of conventional nitrate removal processes, i.e., ion exchange, reverse osmosis, or electrodialysis processes, we can better identify the relative advantages and limitations of electrocatalytic reduction. Furthermore, key sustainability metrics, such as carbon footprint and water use, can provide quantitative information on the environmental performance of the process and emphasize areas for enhancement and optimization. To the best of the authors’ knowledge, there are no previous literature studies on the LCA of electrocatalytic nitrate reduction for brackish groundwater treatment. However, Mosalpuri et al. [243] carried out a techno-economic analysis and an LCA of the electrochemical conversion of nitrate ions present in wastewater to hydroxylamine (NH2OH), which is a valuable chemical intermediate. They calculated the NH2OH production costs and determined the life cycle emissions for a small-scale facility (producing 1500 kgNH2OH/day) and a large-scale facility (producing 50,000 kgNH2OH/day) integrated into an Iowa wastewater treatment plant. Thus, they found that the NH2OH production costs for the small- and large-scale facilities are estimated at $6.14/kg-NH2OH and $5.37/kg-NH2OH, respectively. The parameters dominating the electrochemical reactor cost are electrolyte, separations, and fixed cost. Meanwhile, LCA results revealed that the proposed electrochemical pathway to produce NH2OH has lower life cycle impacts than the conventional pathway, especially when using solar energy. Bearing in mind that the Haber–Bosch process entails significant environmental impacts, such as high energy consumption and high CO2 emissions (400 t of CO2 per year, or 1.6% of global CO2 emissions), mostly due to the production of H2 by steam methane reforming (SMR), which releases CO2 stoichiometrically [244,245].
In light of what has been described above, it can be deduced that a great deal of research is still required, both at the laboratory and pilot scale, to bring electrocatalytic processes to a high level of technological maturity, but they are also attractive in terms of prospects and applications. Accordingly, there are both challenges and opportunities in implementing renewable and sustainable energy solutions in electrocatalytic nitrate reduction processes. The major challenge is to scale up these systems to the industrial scale, whilst keeping economic feasibility by considering the initial costs of renewable energy infrastructure and storage technologies, which can be significant. Technological innovations, including advances in electrode materials, catalysts, and energy conversion systems, are key to enhancing efficiency, reducing costs, and making sustainable solutions economically viable. In the future, the integration of emerging trends in renewable energy and environmental regulations is likely to shape the future of sustainable nitrate reduction technologies, offering both opportunities for progress and challenges.

6. Summary and Outlook

This review provides an overview of the electrocatalytic nitrate reduction (eNO3R) to nitrogen and ammonia for application to brackish groundwater. A comprehensive analysis of the eNO3R mechanism and the theoretical simulations by DFT were presented to predict the reaction pathway of electrocatalysts for NH 3 production and N 2 selectivity. An overview of copper-based materials for eNO3R to ammonia as efficient transition metal-based catalysts is given in terms of single-atom copper catalysts, facet engineering, bimetallic, Cu oxide, Cu transition metal oxide composite, Cu-based inorganic/organic material, and Cu-based nanomaterials. Strategies to convert nitrate to nitrogen with iron-based catalysts specifically nanoscale zero-valent iron (nZVI) and strategies to convert ammonia generated with chlorine radical and active chlorine are discussed. A comprehensive analysis of eNO3R for brackish groundwater as a case study with the various operational parameters that influence the process with the challenge of moving towards large-scale application alongside an economic analysis was reported. This comprehensive analysis is essential for researchers and practitioners, as it provides critical information for the design and optimization of nitrate removal systems for this kind of natural water. eNO3R holds great promise for removing nitrate from contaminated water sources, including brackish groundwater, but it should be noted that this technology is still in its infancy and several challenges must be addressed. The following recommendations and information could be considered for the design and optimization of an efficient electrochemical cell unit and the selection of the preferred pathway for this type of water:
  • The mechanism for the eNO3R process is not well understood, which could be a reason to continue using DFT calculations. Farther, the complexity of the nitrate reduction process may make it difficult to accurately model using DFT. On the other hand, the DFT calculations require significant computational resources, which can be time-consuming and expensive. Additionally, DFT calculations rely on several approximations and assumptions, which may not always accurately reflect the real-world system. Despite these challenges, the use of DFT and microkinetic modeling can still provide valuable insights into the mechanism of electrocatalytic nitrate reduction and guide the tailored design of more efficient and selective electrode materials.
  • Machine learning (ML) and artificial intelligence (AI) play an increasingly important role in optimizing catalyst design, by enhancing the efficiency of DFT. These tools reduce the complexity of high-dimensional catalytic activity maps by identifying key descriptors, such as adsorption energies of critical species. This approach allows for more efficient screening of potential catalysts across vast chemical spaces, accelerating material discovery while significantly lowering the computational cost of DFT calculations.
  • Developing more efficient electrocatalysts for eNO3R is needed to find non-precious metal electrode materials that are highly efficient and durable, as the scarcity and cost of noble metals like Pd make them unsuitable for practical use. This requires extensive research and testing of various materials and modifications to improve their stability and durability, as well as their selectivity and activity for electrochemical nitrate reduction. Furthermore, advanced research in nanomaterials for Cu-nanomaterials or Fe-nanomaterials is required due to their operational low potential which reduces energy consumption for the process.
  • Brackish groundwater with high chloride concentration can be advantageous for nitrate conversion to nitrogen; an efficient anode of Ti/Ru-Ir is widely used for this purpose. However, by-product formation such as chlorine-containing byproducts (CBPs) can be the primary drawback, especially with water-containing organic compounds. Focusing on a future hybrid process that includes photo-electrocatalysis or electrocoagulation may be an effective strategy for addressing this special issue.
  • Optimizing operational parameters such as pH, current density, and applied voltage is crucial to improve eNO3R and their effects should be examined in real environments as in the case of brackish groundwater. Additionally, further studies are needed, especially on the challenge of anode and cathode passivation which results in fouling of the electrodes and seriously decrease the efficiency of nitrate reduction, notably on the application of electric pulses and the specific type to ensure the performance of the process and prevent metal leaching. This can enhance nitrate removal efficiency, minimize by-product formation, and promote a sustainable process.
  • Selecting the configuration type for the cell, divided or undivided, can greatly impact the efficiency and selectivity of the reduction process. Divided cells are more effective at nitrate reduction, and they could be the suitable choice for ammonia generation as they prevent re-oxidation of the reduced nitrate in the anode and concurrently, we recover the ammonium that migrates through the cation exchange membrane (CEM). However, the drawbacks associated with divided cells are fouling and scaling of the membrane in complex water matrices as the case in brackish groundwater, additionally, the limited availability of ammonium-selective membranes needs to be addressed in future research. The alternative pathway for eNO3R to ammonium in brackish groundwater can be employed in undivided cells, providing a specific anode that prevents the production of chlorine radicals and active chlorine. These anodes are currently being developed, especially for large-scale applications. On the other hand, undivided cells remain the optimal choice for eNO3R to N 2 due to their effectiveness in minimizing energy consumption and operational costs, while also simplifying system design and maintenance.
  • Reactor design and the components of the cells, such as shapes, dimensions of electrodes, and their placement in the flow channel are essential for the success of the process due to their role in ensuring fluid dispersion, current distribution, and mass transport. Mathematical approaches to designing electrochemical reactors have been adopted for different technology using numerical tools such as computational fluids dynamics (CFD) since we can test the various reactors with different components before the experimental characterization in laboratory-scale cells and then, during the scale-up of pilot plants. Therefore, the use of CFD simulation to visualize surface-phase changes during electrode reactions, including the mass transport and current distribution during eNO3R is crucial to increase the knowledge and performance of the process. Furthermore, designing a new reactor to lower the ohmic resistance and thus reduce unnecessary energy consumption is also a desirable approach.
  • Researchers who focus on the conversion of eNO3R for ammonia production to pursue subsequent utilization should consider the concentration of nitrate in water. Using renewable energies for cost minimization remains advantageous. However, in waters with low nitrate concentrations, it is preferable to prioritize the production of drinking water, specifically through the pathway of converting nitrates into nitrogen. Nevertheless, the generation of ammonium from chemical industry waste, sewage, and the site of caliche ore with high nitrate concentration is the most effective approach to address “two birds in one stone”, removing nitrate and producing ammonia.
  • The operating cost requires special consideration for large-scale applications, especially raw materials, transportation, and post-treatment costs. Also, the electrical energy consumed during nitrate electroreduction represents a large proportion of the total cost. In fact, coupling thermodynamically favorable reactions such as the oxidation of organics is promising, which can not only decrease the electrical energy needed but can also raise the additional value of electrochemical water treatment.
  • The inclusion of renewable energies, notably solar power, in electrocatalytic nitrate reduction (eNO3R) promotes sustainable development by minimizing dependence on non-renewable sources and minimizing the carbon footprint. Moreover, the use of Life Cycle Assessment (LCA) is essential to assess the environmental impact of eNO3R compared to conventional methods, concentrating on energy consumption, emissions, and waste. Yet in-depth research studies are called for to address challenges such as scale-up, high initial costs, and the requirement for technological innovation for sustainable implementation, with the potential support of economic incentives and emerging renewable energy trends.

Author Contributions

Conceptualization, H.O., C.V. and B.G.; writing—original draft preparation, H.O. and S.K.; writing—review and editing, F.A., C.V. and B.G.; supervision, F.A., C.V. and B.G.; project administration, F.A. and B.G.; funding acquisition, F.A. and B.G. All authors have read and agreed to the published version of the manuscript.

Funding

The research was financially supported by the French ministry of Europe and foreign affairs (MEAE), of higher education, research and innovation (MESRI), and the Moroccan ministry of higher education, scientific research and professional training via the Hubert Curien partnership (PHC) Toubkal (grant N° TBK/21/120).

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Ighalo, J.O.; Adeniyi, A.G. A Comprehensive Review of Water Quality Monitoring and Assessment in Nigeria. Chemosphere 2020, 260, 127569. [Google Scholar] [CrossRef] [PubMed]
  2. Ekere, N.R.; Agbazue, V.E.; Ngang, B.U.; Ihedioha, J.N. Hydrochemistry and Water Quality Index of Groundwater Resources in Enugu North District, Enugu, Nigeria. Environ. Monit. Assess. 2019, 191, 150. [Google Scholar] [CrossRef] [PubMed]
  3. Howe, C.W.; Hecht, A. Book Reviews. Clim. Chang. 1995, 31, 119–126. [Google Scholar] [CrossRef]
  4. Malik, M.S.; Shukla, J.P. GIS Modeling Approach for Assessment of Groundwater Vulnerability in Parts of Tawa River Catchment Area, Hoshangabad, Madhya Pradesh, India. Groundw. Sustain. Dev. 2019, 9, 100249. [Google Scholar] [CrossRef]
  5. Heiß, L.; Bouchaou, L.; Tadoumant, S.; Reichert, B. Index-Based Groundwater Vulnerability and Water Quality Assessment in the Arid Region of Tata City (Morocco). Groundw. Sustain. Dev. 2020, 10, 100344. [Google Scholar] [CrossRef]
  6. Singh, S.; Anil, A.G.; Kumar, V.; Kapoor, D.; Subramanian, S.; Singh, J.; Ramamurthy, P.C. Nitrates in the Environment: A Critical Review of Their Distribution, Sensing Techniques, Ecological Effects and Remediation. Chemosphere 2022, 287, 131996. [Google Scholar] [CrossRef] [PubMed]
  7. Zahra HAFIANE, F.; Tahri, L.; Nouayti, N.; El Jarmouni, M.; Arifi, K.; Elamrani Idrissi, A.; Fekhaoui, M. Assessment of spatial and seasonal nitrate variation of groundwater in the irrigated perimeter (Tadla Plain-Morocco). Agric. For. 2020, 66, 203–214. [Google Scholar] [CrossRef]
  8. Lahjouj, A.; El Hmaidi, A.; Bouhafa, K.; Boufala, M. Mapping Specific Groundwater Vulnerability to Nitrate Using Random Forest: Case of Sais Basin, Morocco. Model. Earth Syst. Environ. 2020, 6, 1451–1466. [Google Scholar] [CrossRef]
  9. Wagh, V.; Panaskar, D.; Muley, A.; Mukate, S.; Gaikwad, S. Neural Network Modelling for Nitrate Concentration in Groundwater of Kadava River Basin, Nashik, Maharashtra, India. Groundw. Sustain. Dev. 2018, 7, 436–445. [Google Scholar] [CrossRef]
  10. Adimalla, N.; Taloor, A.K. Hydrogeochemical Investigation of Groundwater Quality in the Hard Rock Terrain of South India Using Geographic Information System (GIS) and Groundwater Quality Index (GWQI) Techniques. Groundw. Sustain. Dev. 2020, 10, 100288. [Google Scholar] [CrossRef]
  11. IARC Working Group on the Evaluation of Carcinogenic Risks to Humans; International Agency for Research on Cancer; World Health Organization Ingested Nitrate and Nitrite, and Cyanobacterial Peptide Toxins. IARC Monographs on the Evaluation of Carcinogenic Risks to Humans; International Agency for Research on Cancer; Distributed by WHO Press: Lyon, France; Geneva, The Netherland, 2010; ISBN 978-92-832-1294-2. [Google Scholar]
  12. Greer, F.R. Infant Methemoglobinemia: The Role of Dietary Nitrate in Food and Water. Pediatrics 2005, 116, 784–786. [Google Scholar] [CrossRef] [PubMed]
  13. Nitrate Contamination: Exposure, Consequence, and Control; Bogárdi, I.; Kuzelka, R.D.; Ennenga, W.G. (Eds.) NATO ASI Series; Series G, Ecological Sciences; Springer-Verlag: New York, NY, USA, 1991; ISBN 978-0-387-53088-8. [Google Scholar]
  14. World Health Organization (WHO). Nitrate and Nitrite in Drinking-Water: Background Document for Development of WHO Guidelines for Drinking-Water Quality; World Health Organization: Geneva, Switzerland, 2011. [Google Scholar]
  15. Zhang, Z.; Huang, G.; Zhang, P.; Shen, J.; Wang, S.; Li, Y. Development of Iron-Based Biochar for Enhancing Nitrate Adsorption: Effects of Specific Surface Area, Electrostatic Force, and Functional Groups. Sci. Total Environ. 2023, 856, 159037. [Google Scholar] [CrossRef] [PubMed]
  16. Loganathan, P.; Vigneswaran, S.; Kandasamy, J. Enhanced Removal of Nitrate from Water Using Surface Modification of Adsorbents—A Review. J. Environ. Manag. 2013, 131, 363–374. [Google Scholar] [CrossRef] [PubMed]
  17. Samatya, S.; Kabay, N.; Yüksel, Ü.; Arda, M.; Yüksel, M. Removal of Nitrate from Aqueous Solution by Nitrate Selective Ion Exchange Resins. React. Funct. Polym. 2006, 66, 1206–1214. [Google Scholar] [CrossRef]
  18. Korak, J.A.; Mungan, A.L.; Watts, L.T. Critical Review of Waste Brine Management Strategies for Drinking Water Treatment Using Strong Base Ion Exchange. J. Hazard. Mater. 2023, 441, 129473. [Google Scholar] [CrossRef]
  19. Ding, X.; Wei, D.; Guo, W.; Wang, B.; Meng, Z.; Feng, R.; Du, B.; Wei, Q. Biological Denitrification in an Anoxic Sequencing Batch Biofilm Reactor: Performance Evaluation, Nitrous Oxide Emission and Microbial Community. Bioresour. Technol. 2019, 285, 121359. [Google Scholar] [CrossRef]
  20. Yu, L.; Zhang, E.; Yang, L.; Liu, S.; Rensing, C.; Zhou, S. Combining Biological Denitrification and Electricity Generation in Methane-Powered Microbial Fuel Cells. J. Environ. Sci. 2023, 130, 212–222. [Google Scholar] [CrossRef]
  21. Cao, S.; Zhou, Y. New Direction in Biological Nitrogen Removal from Industrial Nitrate Wastewater via Anammox. Appl. Microbiol. Biotechnol. 2019, 103, 7459–7466. [Google Scholar] [CrossRef]
  22. Pang, S.; Zhang, G.; Hu, Y.; Luo, X.; Li, N.; Zhu, J.; Jiang, J.; Han, B.; Shen, T. Enhancing the Total Nitrogen Removal Efficiency of Anaerobic Ammonium Oxidation (Anammox) with Two Types of Biochars. J. Environ. Chem. Eng. 2023, 11, 109108. [Google Scholar] [CrossRef]
  23. Vilardi Giorgio Bimetallic Nzvi-Induced Chemical Denitrification Modelling Using the Shrinking Core Model. Chem. Eng. Trans. 2018, 70, 235–240. [CrossRef]
  24. Jiang, M.; Wu, Y.; He, P.; Hu, S.; Li, Q.; Chen, S. Assembled Denitrifying Consortia for Efficient Nitrate Removal under Low-COD/N Conditions. Chem. Eng. J. 2023, 460, 141655. [Google Scholar] [CrossRef]
  25. Yun, Y.; Wen, X. High Catalytic Activity and N2 Selectivity on Catalytic Denitrification by Fe0 and Bimetallic Catalyst. J. Environ. Eng. 2020, 146, 04020104. [Google Scholar] [CrossRef]
  26. Xue, S.; Wen, X.; Yun, Y. Pathway and Mechanism Study on Improvement of N2 Selectivity of Catalytic Denitrification. J. Saudi Chem. Soc. 2023, 27, 101577. [Google Scholar] [CrossRef]
  27. Patra, S.; Pranudta, A.; Chanlek, N.; Nguyen, T.T.; Nhat, N.H.; El-Moselhy, M.M.; Padungthon, S. Denitrification of Nitrate in Regeneration Waste Brine Using Hybrid Cation Exchanger Supported Nanoscale Zero-Valent Iron with/without Palladium Nanoparticles. Chemosphere 2023, 310, 136851. [Google Scholar] [CrossRef] [PubMed]
  28. Epsztein, R.; Nir, O.; Lahav, O.; Green, M. Selective Nitrate Removal from Groundwater Using a Hybrid Nanofiltration–Reverse Osmosis Filtration Scheme. Chem. Eng. J. 2015, 279, 372–378. [Google Scholar] [CrossRef]
  29. Mehenktaş, C.; Arar, Ö. Application of Membrane Processes for Nitrate (NO3-) Removal. Curr. Chin. Sci. 2023, 3, 42–56. [Google Scholar] [CrossRef]
  30. Tyagi, S.; Rawtani, D.; Khatri, N.; Tharmavaram, M. Strategies for Nitrate Removal from Aqueous Environment Using Nanotechnology: A Review. J. Water Process Eng. 2018, 21, 84–95. [Google Scholar] [CrossRef]
  31. Wang, X.; Zhu, M.; Zeng, G.; Liu, X.; Fang, C.; Li, C. A Three-Dimensional Cu Nanobelt Cathode for Highly Efficient Electrocatalytic Nitrate Reduction. Nanoscale 2020, 12, 9385–9391. [Google Scholar] [CrossRef]
  32. Shahedi, A.; Darban, A.K.; Taghipour, F.; Jamshidi-Zanjani, A. A Review on Industrial Wastewater Treatment via Electrocoagulation Processes. Curr. Opin. Electrochem. 2020, 22, 154–169. [Google Scholar] [CrossRef]
  33. Moussa, D.T.; El-Naas, M.H.; Nasser, M.; Al-Marri, M.J. A Comprehensive Review of Electrocoagulation for Water Treatment: Potentials and Challenges. J. Environ. Manag. 2017, 186, 24–41. [Google Scholar] [CrossRef]
  34. Zou, S.; Guan, L.; Taylor, D.P.; Kuhn, D.; He, Z. Nitrogen Removal from Water of Recirculating Aquaculture System by a Microbial Fuel Cell. Aquaculture 2018, 497, 74–81. [Google Scholar] [CrossRef]
  35. Spatolisano, E.; Pellegrini, L.A. Haber-Bosch Process Intensification: A First Step towards Small-Scale Distributed Ammonia Production. Chem. Eng. Res. Des. 2023, 195, 651–661. [Google Scholar] [CrossRef]
  36. Nie, S.; Mo, S.; Gao, T.; Yan, B.; Shen, P.; Kashif, M.; Zhang, Z.; Li, J.; Jiang, C. Coupling Effects of Nitrate Reduction and Sulfur Oxidation in a Subtropical Marine Mangrove Ecosystem with Spartina Alterniflora Invasion. Sci. Total Environ. 2023, 862, 160930. [Google Scholar] [CrossRef] [PubMed]
  37. Katsounaros, I. On the Assessment of Electrocatalysts for Nitrate Reduction. Curr. Opin. Electrochem. 2021, 28, 100721. [Google Scholar] [CrossRef]
  38. Anastasiadou, D.; van Beek, Y.; Hensen, E.J.M.; Costa Figueiredo, M. Ammonia Electrocatalytic Synthesis from Nitrate. Electrochem. Sci. Adv. 2022, 3, e2100220. [Google Scholar] [CrossRef]
  39. 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]
  40. de Vooys, A.C.A.; Beltramo, G.L.; van Riet, B.; van Veen, J.A.R.; Koper, M.T.M. Mechanisms of Electrochemical Reduction and Oxidation of Nitric Oxide. Electrochim. Acta 2004, 49, 1307–1314. [Google Scholar] [CrossRef]
  41. Duca, M.; Figueiredo, M.C.; Climent, V.; Rodriguez, P.; Feliu, J.M.; Koper, M.T.M. Selective Catalytic Reduction at Quasi-Perfect Pt(100) Domains: A Universal Low-Temperature Pathway from Nitrite to N2. J. Am. Chem. Soc. 2011, 133, 10928–10939. [Google Scholar] [CrossRef]
  42. Zheng, J.; Lu, T.; Cotton, T.M.; Chumanov, G. Photoinduced Electrochemical Reduction of Nitrite at an Electrochemically Roughened Silver Surface. J. Phys. Chem. B 1999, 103, 6567–6572. [Google Scholar] [CrossRef]
  43. Bartberger, M.D.; Liu, W.; Ford, E.; Miranda, K.M.; Switzer, C.; Fukuto, J.M.; Farmer, P.J.; Wink, D.A.; Houk, K.N. The Reduction Potential of Nitric Oxide (NO) and Its Importance to NO Biochemistry. Proc. Natl. Acad. Sci. USA 2002, 99, 10958–10963. [Google Scholar] [CrossRef]
  44. 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]
  45. Kuwabata, S.; Uezumi, S.; Tanaka, K.; Tanaka, T. Assimilatory and Dissimilatory Reduction of Nitrate and Nitrite with a Tris(Tetrabutylammonium) Nonakis(Benzenethiolato)Octasulfidohexaferratedimolybdate(3-) Modified Glassy-Carbon Electrode in Water. Inorg. Chem. 1986, 25, 3018–3022. [Google Scholar] [CrossRef]
  46. Reyter, D.; Bélanger, D.; Roué, L. Elaboration by High-Energy Ball Milling of Copper/Palladium Composite Materials—Characterization and Electrocatalytic Activity for the Reduction of Nitrate in Alkaline Medium. J. Electroanal. Chem. 2008, 622, 64–72. [Google Scholar] [CrossRef]
  47. Reyter, D.; Bélanger, D.; Roué, L. Elaboration of Cu−Pd Films by Coelectrodeposition: Application to Nitrate Electroreduction. J. Phys. Chem. C 2009, 113, 290–297. [Google Scholar] [CrossRef]
  48. Lange, R.; Maisonhaute, E.; Robin, R.; Vivier, V. On the Kinetics of the Nitrate Reduction in Concentrated Nitric Acid. Electrochem. Commun. 2013, 29, 25–28. [Google Scholar] [CrossRef]
  49. 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]
  50. Wang, S.; Li, L.; Hui, K.S.; Dinh, D.A.; Lu, Z.; Zhang, Q.; Hui, K.N. Non−noble Single−atom Alloy for Electrocatalytic Nitrate Reduction Using Hierarchical High−throughput Screening. Nano Energy 2023, 113, 108543. [Google Scholar] [CrossRef]
  51. Hu, T.; Wang, C.; Wang, M.; Li, C.M.; Guo, C. Theoretical Insights into Superior Nitrate Reduction to Ammonia Performance of Copper Catalysts. ACS Catal. 2021, 11, 14417–14427. [Google Scholar] [CrossRef]
  52. Cheng, J.; Sun, W.; Dai, G.; Yang, X.; Xia, R.; Xu, Y.; Yang, X.; Tu, W. Electroreduction of Nitrate to Ammonia on Atomically-Dispersed Cu-N4 Active Sites with High Efficiency and Stability. Fuel 2023, 332, 126106. [Google Scholar] [CrossRef]
  53. Zheng, X.; Yan, Y.; Li, X.; Liu, Y.; Yao, Y. Theoretical Insights into Dissociative-Associative Mechanism for Enhanced Electrochemical Nitrate Reduction to Ammonia. J. Hazard. Mater. 2023, 446, 130679. [Google Scholar] [CrossRef]
  54. Li, J.; Zhao, D.; Zhang, L.; Ren, Y.; Yue, L.; Li, Z.; Sun, S.; Luo, Y.; Chen, Q.; Li, T.; et al. Boosting Electrochemical Nitrate-to-Ammonia Conversion by Self-Supported MnCo2O4 Nanowire Array. J. Colloid Interface Sci. 2023, 629, 805–812. [Google Scholar] [CrossRef] [PubMed]
  55. Liu, F.; Chen, C.; Jiang, X.; Guo, J.-Y.; Wei, Y.; Li, J.-W.; Sheng, T.; Zhao, X.; Wei, L. High-Index Surface Structure Engineering of Au–Pd Concave Triple-Octahedrons for Boosting Electrocatalytic Nitrate Reduction to Ammonia. ACS Sustain. Chem. Eng. 2023, 11, 1631–1637. [Google Scholar] [CrossRef]
  56. Chen, Y.; He, J.; Jiang, P.; Qu, F.; Yu, D.; Hao, X.; Zhang, J. Electrochemical Denitrification by a Recyclable Nanoporous Co3o4/Co Cathode: Rapid Recovery and Selective Catalysis. Available online: https://ssrn.com/abstract=4327262 (accessed on 20 August 2024).
  57. Li, X.; Zhang, G.; Zhang, N.; Luo, Y.; Shen, P.; Li, X.; Chu, K. Regulating Pd Nanosheets by W-Doping for Electrochemical Nitrate Reduction to Ammonia. New J. Chem. 2022, 46, 14724–14730. [Google Scholar] [CrossRef]
  58. 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]
  59. Tao, W.; Wang, P.; Li, H.; Huang, R.; Zhou, G. Engineering Sulfur Vacancies Optimization in Ni3Co6S8 Nanospheres toward Extraordinarily Efficient Nitrate Electroreduction to Ammonia. Appl. Catal. B Environ. 2023, 324, 122193. [Google Scholar] [CrossRef]
  60. Lei, F.; Xu, W.; Yu, J.; Li, K.; Xie, J.; Hao, P.; Cui, G.; Tang, B. Electrochemical Synthesis of Ammonia by Nitrate Reduction on Indium Incorporated in Sulfur Doped Graphene. Chem. Eng. J. 2021, 426, 131317. [Google Scholar] [CrossRef]
  61. Gou, F.; Wang, H.; Fu, M.; Jiang, Y.; Shen, W.; He, R.; Li, M. Boron-Induced Electron Localization in Cu Nanowires Promotes Efficient Nitrate Reduction to Ammonia in Neutral Media. Appl. Surf. Sci. 2023, 612, 155872. [Google Scholar] [CrossRef]
  62. Niu, Z.; Fan, S.; Li, X.; Duan, J.; Chen, A. Interfacial Engineering of CoMn2O4/NC Induced Electronic Delocalization Boosts Electrocatalytic Nitrogen Oxyanions Reduction to Ammonia. Appl. Catal. B Environ. 2023, 322, 122090. [Google Scholar] [CrossRef]
  63. Fan, J.; Chen, Y.; Chen, X.; Wu, Z.; Teng, W.; Zhang, W. Atomically Dispersed Iron Enables High-Efficiency Electrocatalytic Conversion of Nitrate to Dinitrogen on a N-Coordinated Mesoporous Carbon Architecture. Appl. Catal. B Environ. 2023, 320, 121983. [Google Scholar] [CrossRef]
  64. Calle-Vallejo, F.; Koper, M.T.M.; Bandarenka, A.S. Tailoring the Catalytic Activity of Electrodes with Monolayer Amounts of Foreign Metals. Chem. Soc. Rev. 2013, 42, 5210. [Google Scholar] [CrossRef]
  65. Wang, Y.; Xu, A.; Wang, Z.; Huang, L.; Li, J.; Li, F.; Wicks, J.; Luo, M.; Nam, D.-H.; Tan, C.-S.; et al. 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] [PubMed]
  66. Karamad, M.; Goncalves, T.J.; Jimenez-Villegas, S.; Gates, I.D.; Siahrostami, S. Why Copper Catalyzes Electrochemical Reduction of Nitrate to Ammonia. Faraday Discuss. 2023, 243, 502–519. [Google Scholar] [CrossRef] [PubMed]
  67. Liu, J.-X.; Richards, D.; Singh, N.; Goldsmith, B.R. Activity and Selectivity Trends in Electrocatalytic Nitrate Reduction on Transition Metals. ACS Catal. 2019, 9, 7052–7064. [Google Scholar] [CrossRef]
  68. Sathishkumar, N.; Chen, H.-T. Mechanistic Understanding of the Electrocatalytic Nitrate Reduction Activity of Double-Atom Catalysts. J. Phys. Chem. C 2023, 127, 994–1005. [Google Scholar] [CrossRef]
  69. McIntyre, S.M.; Ennis, C.; Garden, A.L. A Computational Investigation into the Hydrogenation of NO on Water Ice Surfaces to Rationalize the NO:HNO:NOH Disparity in Space. ACS Earth Space Chem. 2023, 7, 1195–1206. [Google Scholar] [CrossRef]
  70. Lim, J.; Liu, C.-Y.; Park, J.; Liu, Y.-H.; Senftle, T.P.; Lee, S.W.; Hatzell, M.C. Structure Sensitivity of Pd Facets for Enhanced Electrochemical Nitrate Reduction to Ammonia. ACS Catal. 2021, 11, 7568–7577. [Google Scholar] [CrossRef]
  71. Chun, H.-J.; Zeng, Z.; Greeley, J. DFT Insights into NO Electrochemical Reduction: A Case Study of Pt(211) and Cu(211) Surfaces. ACS Catal. 2022, 12, 1394–1402. [Google Scholar] [CrossRef]
  72. Yu, P.; Wang, F.; Shifa, T.A.; Zhan, X.; Lou, X.; Xia, F.; He, J. Earth Abundant Materials beyond Transition Metal Dichalcogenides: A Focus on Electrocatalyzing Hydrogen Evolution Reaction. Nano Energy 2019, 58, 244–276. [Google Scholar] [CrossRef]
  73. Long, J.; Guo, C.; Fu, X.; Jing, H.; Qin, G.; Li, H.; Xiao, J. Unveiling Potential Dependence in NO Electroreduction to Ammonia. J. Phys. Chem. Lett. 2021, 12, 6988–6995. [Google Scholar] [CrossRef]
  74. Carvalho, O.Q.; Marks, R.; Nguyen, H.K.K.; Vitale-Sullivan, M.E.; Martinez, S.C.; Árnadóttir, L.; Stoerzinger, K.A. Role of Electronic Structure on Nitrate Reduction to Ammonium: A Periodic Journey. J. Am. Chem. Soc. 2022, 144, 14809–14818. [Google Scholar] [CrossRef]
  75. Xin, H.; Mou, T.; Pillai, H.S.; Wang, S.-H.; Huang, Y. Interpretable Machine Learning for Catalytic Materials Design toward Sustainability. Acc. Mater. Res. 2024, 5, 22–34. [Google Scholar] [CrossRef]
  76. Chen, C.; Xiao, B.; Li, Z.; Li, W.; Li, Q.; Yu, X. High Throughput Screening for Electrocatalysts for Nitrogen Reduction Reaction Using Metal-Doped Bilayer Borophene: A Combined Approach of DFT and Machine Learning. Mol. Catal. 2024, 557, 113972. [Google Scholar] [CrossRef]
  77. Liu, J.; Guo, H.; Xiong, Y.; Chen, X.; Yu, Y.; Wang, C. Rational Design of Pt-Anchored Single-Atom Alloy Electrocatalysts for NO-to-NH3 Conversion by Density Functional Theory and Machine Learning. Chin. J. Catal. 2024, 62, 243–253. [Google Scholar] [CrossRef]
  78. Yang, L.; Fan, J.; Xiao, B.; Zhu, W. Unveiling “Sabatier Principle” for Electrocatalytic Nitric Oxide Reduction on Single Cluster Catalysts: A DFT and Machine Learning Guideline. Chem. Eng. J. 2023, 468, 143823. [Google Scholar] [CrossRef]
  79. 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]
  80. Wan, X.; Zhang, Z.; Yu, W.; Guo, Y. A Density-Functional-Theory-Based and Machine-Learning-Accelerated Hybrid Method for Intricate System Catalysis. Mater. Rep. Energy 2021, 1, 100046. [Google Scholar] [CrossRef]
  81. Walsh, F.C.; Wang, S.; Zhou, N. The Electrodeposition of Composite Coatings: Diversity, Applications and Challenges. Curr. Opin. Electrochem. 2020, 20, 8–19. [Google Scholar] [CrossRef]
  82. Zurita, N.; García, S.G. Comparative Study of Electrodeposited Copper Nanoparticles on Different Substrates for Their Use in the Reduction of Nitrate Ions. Results Eng. 2023, 17, 100800. [Google Scholar] [CrossRef]
  83. Kuang, P.; Natsui, K.; Einaga, Y.; Feng, C.; Cui, Y.; Zhang, W.; Deng, Y. Annealing Enhancement in Stability and Performance of Copper Modified Boron-Doped Diamond (Cu-BDD) Electrode for Electrochemical Nitrate Reduction. Diam. Relat. Mater. 2021, 114, 108310. [Google Scholar] [CrossRef]
  84. Zhang, Y.; Zhao, Y.; Chen, Z.; Wang, L.; Wu, P.; Wang, F. Electrochemical Reduction of Nitrate via Cu/Ni Composite Cathode Paired with Ir-Ru/Ti Anode: High Efficiency and N2 Selectivity. Electrochim. Acta 2018, 291, 151–160. [Google Scholar] [CrossRef]
  85. Chen, L.-F.; Xie, A.-Y.; Lou, Y.-Y.; Tian, N.; Zhou, Z.-Y.; Sun, S.-G. Electrochemical Synthesis of Tetrahexahedral Cu Nanocrystals with High-Index Facets for Efficient Nitrate Electroreduction. J. Electroanal. Chem. 2022, 907, 116022. [Google Scholar] [CrossRef]
  86. Chen, K.-L.; Ahmad, M.S.; Chen, C.-L. Enhanced Nitrate Reduction over Functionalized Pd/Cu Electrode with Tunable Conversion to Nitrogen and Sodium Hydroxide Recovery. Sci. Total Environ. 2023, 869, 161849. [Google Scholar] [CrossRef] [PubMed]
  87. Zhang, C.; Liu, G.; Geng, X.; Wu, K.; Debliquy, M. Metal Oxide Semiconductors with Highly Concentrated Oxygen Vacancies for Gas Sensing Materials: A Review. Sens. Actuators A Phys. 2020, 309, 112026. [Google Scholar] [CrossRef]
  88. Jang, D.; Maeng, J.; Kim, J.; Han, H.; Park, G.H.; Ha, J.; Shin, D.; Hwang, Y.J.; Kim, W.B. Boosting Electrocatalytic Nitrate Reduction Reaction for Ammonia Synthesis by Plasma-Induced Oxygen Vacancies over MnCuOx. Appl. Surf. Sci. 2023, 610, 155521. [Google Scholar] [CrossRef]
  89. Kawashima, K.; Márquez, R.A.; Smith, L.A.; Vaidyula, R.R.; Carrasco-Jaim, O.A.; Wang, Z.; Son, Y.J.; Cao, C.L.; Mullins, C.B. A Review of Transition Metal Boride, Carbide, Pnictide, and Chalcogenide Water Oxidation Electrocatalysts. Chem. Rev. 2023, 123, 12795–13208. [Google Scholar] [CrossRef]
  90. Zang, B.; Zhang, Z.; Yuan, H.; Li, R.; Li, S.; Lu, J. Electrochemical Reduction of Nitrate in an Electric Kettle Equipped with Copper-Modified Titanate Cathode under High-Temperature. J. Water Process Eng. 2023, 56, 104415. [Google Scholar] [CrossRef]
  91. Liu, X.; Guo, R.; Qin, X.; Xu, C. Enhanced Selective Nitrate-to-Nitrogen Electrocatalytic Reduction by CNTs Doped Ni Foam/Cu Electrode Coupled with Cl−. J. Water Process Eng. 2023, 54, 104067. [Google Scholar] [CrossRef]
  92. Liu, H.; Li, J.; Du, F.; Yang, L.; Huang, S.; Gao, J.; Li, C.; Guo, C. A Core–Shell Copper Oxides-Cobalt Oxides Heterostructure Nanowire Arrays for Nitrate Reduction to Ammonia with High Yield Rate. Green Energy Environ. 2023, 8, 1619–1629. [Google Scholar] [CrossRef]
  93. Wang, J.; Tang, J. Fe-Based Fenton-like Catalysts for Water Treatment: Preparation, Characterization and Modification. Chemosphere 2021, 276, 130177. [Google Scholar] [CrossRef]
  94. Chen, M.; Huang, X.; Wang, N.; Wang, T.; Yang, J.; Wei, Y.; Dao, X.; Zhou, L.; Hao, H. Cu2O Nanoparticles Modified BiO2-x Nanosheets for Efficient Electrochemical Reduction of Nitrate-N and Nitrobenzene from Wastewater. Sep. Purif. Technol. 2022, 289, 120728. [Google Scholar] [CrossRef]
  95. Pham, X.-H.; Ashik, U.P.M.; Hayashi, J.-I.; Pérez Alonso, A.; Pla, D.; Gómez, M.; Pham Minh, D. Review on the Catalytic Tri-Reforming of Methane—Part II: Catalyst Development. Appl. Catal. A Gen. 2021, 623, 118286. [Google Scholar] [CrossRef]
  96. Zhang, Z.; Xu, Y.; Shi, W.; Wang, W.; Zhang, R.; Bao, X.; Zhang, B.; Li, L.; Cui, F. Electrochemical-Catalytic Reduction of Nitrate over Pd–Cu/γAl2O3 Catalyst in Cathode Chamber: Enhanced Removal Efficiency and N2 Selectivity. Chem. Eng. J. 2016, 290, 201–208. [Google Scholar] [CrossRef]
  97. Liu, M.J.; Guo, J.; Hoffman, A.S.; Stenlid, J.H.; Tang, M.T.; Corson, E.R.; Stone, K.H.; Abild-Pedersen, F.; Bare, S.R.; Tarpeh, W.A. Catalytic Performance and Near-Surface X-Ray Characterization of Titanium Hydride Electrodes for the Electrochemical Nitrate Reduction Reaction. J. Am. Chem. Soc. 2022, 144, 5739–5744. [Google Scholar] [CrossRef] [PubMed]
  98. Yang, X.; Wang, R.; Wang, S.; Song, C.; Lu, S.; Fang, L.; Yin, F.; Liu, H. Sequential Active-Site Switches in Integrated Cu/Fe-TiO2 for Efficient Electroreduction from Nitrate into Ammonia. Appl. Catal. B Environ. 2023, 325, 122360. [Google Scholar] [CrossRef]
  99. Braley, S.E.; Xie, J.; Losovyj, Y.; Smith, J.M. Graphite Conjugation of a Macrocyclic Cobalt Complex Enhances Nitrite Electroreduction to Ammonia. J. Am. Chem. Soc. 2021, 143, 7203–7208. [Google Scholar] [CrossRef]
  100. Li, J.; Yue, M.-F.; Wei, Y.-M.; Li, J.-F. Synthetic Strategies of Single-Atoms Catalysts and Applications in Electrocatalysis. Electrochim. Acta 2022, 409, 139835. [Google Scholar] [CrossRef]
  101. Li, P.; Li, R.; Liu, Y.; Xie, M.; Jin, Z.; Yu, G. Pulsed Nitrate-to-Ammonia Electroreduction Facilitated by Tandem Catalysis of Nitrite Intermediates. J. Am. Chem. Soc. 2023, 145, 6471–6479. [Google Scholar] [CrossRef]
  102. Wang, C.; Ye, F.; Shen, J.; Xue, K.-H.; Zhu, Y.; Li, C. In Situ Loading of Cu 2 O Active Sites on Island-like Copper for Efficient Electrochemical Reduction of Nitrate to Ammonia. ACS Appl. Mater. Interfaces 2022, 14, 6680–6688. [Google Scholar] [CrossRef]
  103. Butcher, D.P.; Gewirth, A.A. Nitrate Reduction Pathways on Cu Single Crystal Surfaces: Effect of Oxide and Cl−. Nano Energy 2016, 29, 457–465. [Google Scholar] [CrossRef]
  104. Gong, Z.; Zhong, W.; He, Z.; Jia, C.; Zhou, D.; Zhang, N.; Kang, X.; Chen, Y. Improving Electrochemical Nitrate Reduction Activity of Layered Perovskite Oxide La2CuO4 via B-Site Doping. Catal. Today 2022, 402, 259–265. [Google Scholar] [CrossRef]
  105. Yao, J.; Mei, Y.; Yuan, T.; Chen, J.; Pan, H.; Wang, J. Electrochemical Removal of Nitrate from Wastewater with a Ti Cathode and Pt Anode for High Efficiency and N2 Selectivity. J. Electroanal. Chem. 2021, 882, 115019. [Google Scholar] [CrossRef]
  106. Dutta, A. Fourier Transform Infrared Spectroscopy. In Spectroscopic Methods for Nanomaterials Characterization; Elsevier: Amsterdam, The Netherlands, 2017; pp. 73–93. ISBN 978-0-323-46140-5. [Google Scholar]
  107. Yang, Q.; Yang, C.-C.; Lin, C.-H.; Jiang, H.-L. Metal-Organic-Framework-Derived Hollow N-Doped Porous Carbon with Ultrahigh Concentrations of Single Zn Atoms for Efficient Carbon Dioxide Conversion. Angew. Chem. Int. Ed. 2019, 58, 3511–3515. [Google Scholar] [CrossRef] [PubMed]
  108. Timoshenko, J.; Roldan Cuenya, B. In Situ/Operando Electrocatalyst Characterization by X-Ray Absorption Spectroscopy. Chem. Rev. 2021, 121, 882–961. [Google Scholar] [CrossRef] [PubMed]
  109. Fang, L.; Seifert, S.; Winans, R.E.; Li, T. Understanding Synthesis and Structural Variation of Nanomaterials through In Situ/Operando XAS and SAXS. Small 2022, 18, 2106017. [Google Scholar] [CrossRef] [PubMed]
  110. Zheng, W.; Lee, L.Y.S. Observing Electrocatalytic Processes via In Situ Electrochemical Scanning Tunneling Microscopy: Latest Advances. Chem. Asian J. 2022, 17, e202200384. [Google Scholar] [CrossRef]
  111. Ma, X.; Yu, Z.; Han, Y.; Song, X.; Sun, Q. In Situ Scanning Electron Microscopy Observation of Deformation and Fracture Behavior of Ti-3Zr-2Sn-3Mo-25Nb Alloy. Rare Met. 2012, 31, 318–322. [Google Scholar] [CrossRef]
  112. Petrii, O.A.; Safonova, T.Y. Electroreduction of Nitrate and Nitrite Anions on Platinum Metals: A Model Process for Elucidating the Nature of the Passivation by Hydrogen Adsorption. J. Electroanal. Chem. 1992, 331, 897–912. [Google Scholar] [CrossRef]
  113. de Groot, M.T.; Koper, M.T.M. The Influence of Nitrate Concentration and Acidity on the Electrocatalytic Reduction of Nitrate on Platinum. J. Electroanal. Chem. 2004, 562, 81–94. [Google Scholar] [CrossRef]
  114. Safonova, T.Y.; Petrii, O.A. Effect of Inorganic Cations on the Electroreduction of Nitrate Anions on Pt|Pt Electrodes in Sulfuric Acid Solutions. J. Electroanal. Chem. 1998, 448, 211–216. [Google Scholar] [CrossRef]
  115. Li, H.; Robertson, D.H.; Chambers, J.Q.; Hobbs, D.T. Electrochemical Reduction of Nitrate and Nitrite in Concentrated Sodium Hydroxide at Platinum and Nickel Electrodes. J. Electrochem. Soc. 1988, 135, 1154–1158. [Google Scholar] [CrossRef]
  116. 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] [PubMed]
  117. Li, H.L.; Chambers, J.Q.; Hobbs, D.T. Electroreduction of Nitrate Ions in Concentrated Sodium Hydroxide Solutions at Lead, Zinc, Nickel and Phthalocyanine-Modified Electrodes. J. Appl. Electrochem. 1988, 18, 454–458. [Google Scholar] [CrossRef]
  118. Yang, J.; Sebastian, P.; Duca, M.; Hoogenboom, T.; Koper, M.T.M. pH Dependence of the Electroreduction of Nitrate on Rh and Pt Polycrystalline Electrodes. Chem. Commun. 2014, 50, 2148–2151. [Google Scholar] [CrossRef] [PubMed]
  119. Wasberg, M.; Horányi, G. Electrocatalytic Reduction of Nitric Acid at Rhodized Electrodes and Its Inhibition by Chloride Ions. Electrochim. Acta 1995, 40, 615–623. [Google Scholar] [CrossRef]
  120. Ureta-Zañartu, S.; Yáñez, C. Electroreduction of Nitrate Ion on Pt, Ir and on 70:30 Pt:Ir Alloy. Electrochim. Acta 1997, 42, 1725–1731. [Google Scholar] [CrossRef]
  121. Katsounaros, I.; Ipsakis, D.; Polatides, C.; Kyriacou, G. Efficient Electrochemical Reduction of Nitrate to Nitrogen on Tin Cathode at Very High Cathodic Potentials. Electrochim. Acta 2006, 52, 1329–1338. [Google Scholar] [CrossRef]
  122. Katsounaros, I.; Kyriacou, G. Influence of the Concentration and the Nature of the Supporting Electrolyte on the Electrochemical Reduction of Nitrate on Tin Cathode. Electrochim. Acta 2007, 52, 6412–6420. [Google Scholar] [CrossRef]
  123. Tada, K.; Shimazu, K. Kinetic Studies of Reduction of Nitrate Ions at Sn-Modified Pt Electrodes Using a Quartz Crystal Microbalance. J. Electroanal. Chem. 2005, 577, 303–309. [Google Scholar] [CrossRef]
  124. Shimazu, K.; Kawaguchi, T.; Tada, K. Preparation of Binary Metal Electrocatalysts by Self-Assembly of Precursor Ionic Species on Gold and Reduction of Nitrate Ions. J. Electroanal. Chem. 2002, 529, 20–27. [Google Scholar] [CrossRef]
  125. Shimazu, K.; Goto, R.; Tada, K. Electrochemical Reduction of Nitrate Ions on Tin-Modified Platinum and Palladium Electrodes. Chem. Lett. 2002, 31, 204–205. [Google Scholar] [CrossRef]
  126. Shimazu, K.; Goto, R.; Piao, S.; Kayama, R.; Nakata, K.; Yoshinaga, Y. Reduction of Nitrate Ions on Tin-Modified Palladium Thin Film Electrodes. J. Electroanal. Chem. 2007, 601, 161–168. [Google Scholar] [CrossRef]
  127. Casella, I.G.; Contursi, M. An Electrochemical and XPS Study of the Electrodeposited Binary Pd–Sn Catalyst: The Electroreduction of Nitrate Ions in Acid Medium. J. Electroanal. Chem. 2006, 588, 147–154. [Google Scholar] [CrossRef]
  128. da Cunha, M.C.P.M.; De Souza, J.P.I.; Nart, F.C. Reaction Pathways for Reduction of Nitrate Ions on Platinum, Rhodium, and Platinum−Rhodium Alloy Electrodes. Langmuir 2000, 16, 771–777. [Google Scholar] [CrossRef]
  129. Brylev, O.; Sarrazin, M.; Belanger, D.; Roue, L. Rhodium Deposits on Pyrolytic Graphite Substrate: Physico-Chemical Properties and Electrocatalytic Activity towards Nitrate Reduction in Neutral Medium. Appl. Catal. B Environ. 2006, 11, 243–253. [Google Scholar] [CrossRef]
  130. Brylev, O.; Sarrazin, M.; Roue, L.; Belanger, D. Nitrate and Nitrite Electrocatalytic Reduction on Rh-Modified Pyrolytic Graphite Electrodes. Electrochim. Acta 2007, 52, 6237–6247. [Google Scholar] [CrossRef]
  131. De, D.; Kalu, E.E.; Tarjan, P.P.; Englehardt, J.D. Kinetic Studies of the Electrochemical Treatment of Nitrate and Nitrite Ions on Iridium-Modified Carbon Fiber Electrodes. Chem. Eng. Technol. 2004, 27, 56–64. [Google Scholar] [CrossRef]
  132. Georgeaud, V.; Diamand, A.; Borrut, D.; Grange, D.; Coste, M. Electrochemical Treatment of Wastewater Polluted by Nitrate: Selective Reduction to N2 on Boron-Doped Diamond Cathode. Water Sci. Technol. 2011, 63, 206–212. [Google Scholar] [CrossRef]
  133. Ghazouani, M.; Akrout, H.; Bousselmi, L. Nitrate and Carbon Matter Removals from Real Effluents Using Si/BDD Electrode. Environ. Sci. Pollut. Res. 2017, 24, 9895–9906. [Google Scholar] [CrossRef]
  134. Ye, L.; Nayak-Luke, R.; Bañares-Alcántara, R.; Tsang, E. Reaction: “Green” Ammonia Production. Chem 2017, 3, 712–714. [Google Scholar] [CrossRef]
  135. Yang, J.; Qi, H.; Li, A.; Liu, X.; Yang, X.; Zhang, S.; Zhao, Q.; Jiang, Q.; Su, Y.; Zhang, L.; et al. Potential-Driven Restructuring of Cu Single Atoms to Nanoparticles for Boosting the Electrochemical Reduction of Nitrate to Ammonia. J. Am. Chem. Soc. 2022, 144, 12062–12071. [Google Scholar] [CrossRef]
  136. Wu, K.; Sun, C.; Wang, Z.; Song, Q.; Bai, X.; Yu, X.; Li, Q.; Wang, Z.; Zhang, H.; Zhang, J.; et al. Surface Reconstruction on Uniform Cu Nanodisks Boosted Electrochemical Nitrate Reduction to Ammonia. ACS Mater. Lett. 2022, 4, 650–656. [Google Scholar] [CrossRef]
  137. Fu, X.; Zhao, X.; Hu, X.; He, K.; Yu, Y.; Li, T.; Tu, Q.; Qian, X.; Yue, Q.; Wasielewski, M.R.; et al. Alternative Route for Electrochemical Ammonia Synthesis by Reduction of Nitrate on Copper Nanosheets. Appl. Mater. Today 2020, 19, 100620. [Google Scholar] [CrossRef]
  138. Bi, B.; Dong, A.-Q.; Shi, M.-M.; Sun, X.-F.; Li, H.-R.; Kang, X.; Gao, R.; Meng, Z.; Chen, Z.-Y.; Xu, T.-W.; et al. Efficient Ammonia Synthesis from Nitrate Catalyzed by Au/Cu with Enhanced Adsorption Ability. Small Struct. 2023, 4, 2200308. [Google Scholar] [CrossRef]
  139. Fang, J.-Y.; Zheng, Q.-Z.; Lou, Y.-Y.; Zhao, K.-M.; Hu, S.-N.; Li, G.; Akdim, O.; Huang, X.-Y.; Sun, S.-G. Ampere-Level Current Density Ammonia Electrochemical Synthesis Using CuCo Nanosheets Simulating Nitrite Reductase Bifunctional Nature. Nat. Commun. 2022, 13, 7899. [Google Scholar] [CrossRef] [PubMed]
  140. Chavez, M.E.; Biset-Peiró, M.; Murcia-López, S.; Morante, J.R. Cu2O–Cu@Titanium Surface with Synergistic Performance for Nitrate-to-Ammonia Electrochemical Reduction. ACS Sustain. Chem. Eng. 2023, 11, 3633–3643. [Google Scholar] [CrossRef] [PubMed]
  141. He, W.; Zhang, J.; Dieckhöfer, S.; Varhade, S.; Brix, A.C.; Lielpetere, A.; Seisel, S.; Junqueira, J.R.C.; Schuhmann, W. Splicing the Active Phases of Copper/Cobalt-Based Catalysts Achieves High-Rate Tandem Electroreduction of Nitrate to Ammonia. Nat. Commun. 2022, 13, 1129. [Google Scholar] [CrossRef]
  142. Li, M.; Feng, C.; Zhang, Z.; Sugiura, N. Efficient Electrochemical Reduction of Nitrate to Nitrogen Using Ti/IrO2–Pt Anode and Different Cathodes. Electrochim. Acta 2009, 54, 4600–4606. [Google Scholar] [CrossRef]
  143. Teng, W.; Bai, N.; Liu, Y.; Liu, Y.; Fan, J.; Zhang, W. Selective Nitrate Reduction to Dinitrogen by Electrocatalysis on Nanoscale Iron Encapsulated in Mesoporous Carbon. Environ. Sci. Technol. 2018, 52, 230–236. [Google Scholar] [CrossRef]
  144. Liu, Z.; Dong, S.; Zou, D.; Ding, J.; Yu, A.; Zhang, J.; Shan, C.; Gao, G.; Pan, B. Electrochemically Mediated Nitrate Reduction on Nanoconfined Zerovalent Iron: Properties and Mechanism. Water Res. 2020, 173, 115596. [Google Scholar] [CrossRef]
  145. Teng, W.; Fan, J.; Zhang, W. Iron-Catalyzed Selective Denitrification over N-Doped Mesoporous Carbon. ACS Appl. Mater. Interfaces 2020, 12, 28091–28099. [Google Scholar] [CrossRef]
  146. Liu, X.; Wang, Y.; Smith, R.L.; Liu, L.; Qi, X. Synthesis of Self-Renewing Fe(0)-Dispersed Ordered Mesoporous Carbon for Electrocatalytic Reduction of Nitrates to Nitrogen. Sci. Total Environ. 2022, 836, 155640. [Google Scholar] [CrossRef] [PubMed]
  147. Ni, F.; Chen, J.; Tai, Z.; Wang, L.; Liu, L.; Yang, J. Iron Nanoparticles Confined in Periodic Mesoporous Organosilicon as Nanoreactors for Efficient Nitrate Reduction. ACS Appl. Nano Mater. 2022, 5, 5149–5157. [Google Scholar] [CrossRef]
  148. Wang, J.; Ling, L.; Deng, Z.; Zhang, W. Nitrogen-Doped Iron for Selective Catalytic Reduction of Nitrate to Dinitrogen. Sci. Bull. 2020, 65, 926–933. [Google Scholar] [CrossRef] [PubMed]
  149. Duan, W.; Li, G.; Lei, Z.; Zhu, T.; Xue, Y.; Wei, C.; Feng, C. Highly Active and Durable Carbon Electrocatalyst for Nitrate Reduction Reaction. Water Res. 2019, 161, 126–135. [Google Scholar] [CrossRef] [PubMed]
  150. Chen, X.; Zhang, T.; Kan, M.; Song, D.; Jia, J.; Zhao, Y.; Qian, X. Binderless and Oxygen Vacancies Rich FeNi/Graphitized Mesoporous Carbon/Ni Foam for Electrocatalytic Reduction of Nitrate. Environ. Sci. Technol. 2020, 54, 13344–13353. [Google Scholar] [CrossRef]
  151. Sun, J.; Gao, W.; Fei, H.; Zhao, G. Efficient and Selective Electrochemical Reduction of Nitrate to N2 by Relay Catalytic Effects of Fe-Ni Bimetallic Sites on MOF-Derived Structure. Appl. Catal. B Environ. 2022, 301, 120829. [Google Scholar] [CrossRef]
  152. Niu, Q.; Yang, S.; Song, Z.; Liu, C.; Li, Z.; Wang, X.; Song, L. Fabrication of Efficient and Robust Fe0/Ni2P/CC Composite and the Employment for Electrochemical Reduction of Nitrate. J. Environ. Chem. Eng. 2021, 9, 106412. [Google Scholar] [CrossRef]
  153. Jonoush, Z.A.; Rezaee, A.; Ghaffarinejad, A. Electrocatalytic Nitrate Reduction Using Fe0/Fe3O4 Nanoparticles Immobilized on Nickel Foam: Selectivity and Energy Consumption Studies. J. Clean. Prod. 2020, 242, 118569. [Google Scholar] [CrossRef]
  154. Su, L.; Han, D.; Zhu, G.; Xu, H.; Luo, W.; Wang, L.; Jiang, W.; Dong, A.; Yang, J. Tailoring the Assembly of Iron Nanoparticles in Carbon Microspheres toward High-Performance Electrocatalytic Denitrification. Nano Lett. 2019, 19, 5423–5430. [Google Scholar] [CrossRef]
  155. Su, L.; Zhang, F.; Wang, L.; Fang, X.; Jiang, W.; Yang, J. Flexible Electrocatalysts: Interfacial-Assembly of Iron Nanoparticles for Nitrate Reduction. Chem. Commun. 2021, 57, 6740–6743. [Google Scholar] [CrossRef]
  156. Ni, F.; Ma, Y.; Chen, J.; Luo, W.; Yang, J. Boron-Iron Nanochains for Selective Electrocatalytic Reduction of Nitrate. Chin. Chem. Lett. 2021, 32, 2073–2078. [Google Scholar] [CrossRef]
  157. Hong, W.; Su, L.; Wang, J.; Jiang, M.; Ma, Y.; Yang, J. Boosting the Electrocatalysis of Nitrate to Nitrogen with Iron Nanoparticles Embedded in Carbon Microspheres. Chem. Commun. 2020, 56, 14685–14688. [Google Scholar] [CrossRef] [PubMed]
  158. Lan, Y.; Chen, J.; Zhang, H.; Zhang, W.; Yang, J. Fe/Fe3C Nanoparticle-Decorated N-Doped Carbon Nanofibers for Improving the Nitrogen Selectivity of Electrocatalytic Nitrate Reduction. J. Mater. Chem. A 2020, 8, 15853–15863. [Google Scholar] [CrossRef]
  159. 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]
  160. Zhu, T.; Chen, Q.; Liao, P.; Duan, W.; Liang, S.; Yan, Z.; Feng, C. Single-Atom Cu Catalysts for Enhanced Electrocatalytic Nitrate Reduction with Significant Alleviation of Nitrite Production. Small 2020, 16, 2004526. [Google Scholar] [CrossRef]
  161. 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]
  162. Pérez-Gallent, E.; Figueiredo, M.C.; Katsounaros, I.; Koper, M.T.M. Electrocatalytic Reduction of Nitrate on Copper Single Crystals in Acidic and Alkaline Solutions. Electrochim. Acta 2017, 227, 77–84. [Google Scholar] [CrossRef]
  163. Hu, Q.; Qin, Y.; Wang, X.; Wang, Z.; Huang, X.; Zheng, H.; Gao, K.; Yang, H.; Zhang, P.; Shao, M.; et al. Reaction Intermediate-Mediated Electrocatalyst Synthesis Favors Specified Facet and Defect Exposure for Efficient Nitrate–Ammonia Conversion. Energy Environ. Sci. 2021, 14, 4989–4997. [Google Scholar] [CrossRef]
  164. Shih, Y.-J.; Wu, Z.-L.; Lin, C.-Y.; Huang, Y.-H.; Huang, C.-P. Manipulating the Crystalline Morphology and Facet Orientation of Copper and Copper-Palladium Nanocatalysts Supported on Stainless Steel Mesh with the Aid of Cationic Surfactant to Improve the Electrochemical Reduction of Nitrate and N2 Selectivity. Appl. Catal. B Environ. 2020, 273, 119053. [Google Scholar] [CrossRef]
  165. Zhao, J.; Shen, Z.; Yu, J.; Guo, Y.; Mushtaq, M.A.; Ding, Y.; Song, Z.; Zhang, W.; Huang, X.; Li, Y.; et al. Constructing Cu-CuO Heterostructured Skin on Cu Cubes to Promote Electrocatalytic Ammonia Production from Nitrate Wastewater. J. Hazard. Mater. 2022, 439, 129653. [Google Scholar] [CrossRef]
  166. Bae, S.-E.; Gewirth, A.A. Differential Reactivity of Cu(111) and Cu(100) during Nitratereduction in Acid Electrolyte. Faraday Discuss. 2009, 140, 113–123. [Google Scholar] [CrossRef] [PubMed]
  167. Shih, Y.-J.; Wu, Z.-L.; Huang, Y.-H.; Huang, C.-P. Electrochemical Nitrate Reduction as Affected by the Crystal Morphology and Facet of Copper Nanoparticles Supported on Nickel Foam Electrodes (Cu/Ni). Chem. Eng. J. 2020, 383, 123157. [Google Scholar] [CrossRef]
  168. Mattarozzi, L.; Cattarin, S.; Comisso, N.; Guerriero, P.; Musiani, M.; Vázquez-Gómez, L.; Verlato, E. Electrochemical Reduction of Nitrate and Nitrite in Alkaline Media at CuNi Alloy Electrodes. Electrochim. Acta 2013, 89, 488–496. [Google Scholar] [CrossRef]
  169. Mattarozzi, L.; Cattarin, S.; Comisso, N.; Gambirasi, A.; Guerriero, P.; Musiani, M.; Vázquez-Gómez, L.; Verlato, E. Hydrogen Evolution Assisted Electrodeposition of Porous Cu-Ni Alloy Electrodes and Their Use for Nitrate Reduction in Alkali. Electrochim. Acta 2014, 140, 337–344. [Google Scholar] [CrossRef]
  170. Zhang, J.; Yin, R.; Shao, Q.; Zhu, T.; Huang, X. Oxygen Vacancies in Amorphous InOx Nanoribbons Enhance CO2 Adsorption and Activation for CO2 Electroreduction. Angew. Chem. Int. Ed. 2019, 58, 5609–5613. [Google Scholar] [CrossRef]
  171. Daiyan, R.; Tran-Phu, T.; Kumar, P.; Iputera, K.; Tong, Z.; Leverett, J.; Khan, M.H.A.; Asghar Esmailpour, A.; Jalili, A.; Lim, M.; et al. Nitrate Reduction to Ammonium: From CuO Defect Engineering to Waste NOx-to-NH3 Economic Feasibility. Energy Environ. Sci. 2021, 14, 3588–3598. [Google Scholar] [CrossRef]
  172. Ren, T.; Yu, Z.; Yu, H.; Deng, K.; Wang, Z.; Li, X.; Wang, H.; Wang, L.; Xu, Y. Interfacial Polarization in Metal-Organic Framework Reconstructed Cu/Pd/CuOx Multi-Phase Heterostructures for Electrocatalytic Nitrate Reduction to Ammonia. Appl. Catal. B Environ. 2022, 318, 121805. [Google Scholar] [CrossRef]
  173. 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 Environ. 2022, 306, 121094. [Google Scholar] [CrossRef]
  174. Yang, M.; Wang, J.; Shuang, C.; Li, A. The Improvement on Total Nitrogen Removal in Nitrate Reduction by Using a Prepared CuO–Co3O4/Ti Cathode. Chemosphere 2020, 255, 126970. [Google Scholar] [CrossRef]
  175. Yang, S.; Wang, L.; Jiao, X.; Li, P. Electrochemical Reduction of Nitrate on Different Cu-Zn Oxide Composite Cathodes. Int. J. Electrochem. Sci. 2017, 12, 4370–4383. [Google Scholar] [CrossRef]
  176. Öznülüer, T.; Özdurak, B.; Öztürk Doğan, H. Electrochemical Reduction of Nitrate on Graphene Modified Copper Electrodes in Alkaline Media. J. Electroanal. Chem. 2013, 699, 1–5. [Google Scholar] [CrossRef]
  177. Çirmi, D.; Aydın, R.; Köleli, F. The Electrochemical Reduction of Nitrate Ion on Polypyrrole Coated Copper Electrode. J. Electroanal. Chem. 2015, 736, 101–106. [Google Scholar] [CrossRef]
  178. Zhao, Y.; Liu, Y.; Zhang, Z.; Mo, Z.; Wang, C.; Gao, S. Flower-like Open-Structured Polycrystalline Copper with Synergistic Multi-Crystal Plane for Efficient Electrocatalytic Reduction of Nitrate to Ammonia. Nano Energy 2022, 97, 107124. [Google Scholar] [CrossRef]
  179. 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]
  180. Lubphoo, Y.; Chyan, J.; Grisdanurak, N.; Liao, C. Influence of Pd–Cu on Nanoscale Zero–Valent Iron Supported for Selective Reduction of Nitrate. J. Taiwan Inst. Chem. Eng. 2016, 59, 285–294. [Google Scholar] [CrossRef]
  181. Zhang, H.; Wang, C.; Luo, H.; Chen, J.; Kuang, M.; Yang, J. Iron Nanoparticles Protected by Chainmail-structured Graphene for Durable Electrocatalytic Nitrate Reduction to Nitrogen. Angew. Chem. 2023, 135, e202217071. [Google Scholar] [CrossRef]
  182. Yan, C.; Liu, L. Oxidation of Gas Phase Ammonia via Accelerated Generation of Radical Species and Synergy of Photo Electrochemical Catalysis with Persulfate Activation by CuO-Co3O4 on Cathode Electrode. J. Hazard. Mater. 2020, 388, 121793. [Google Scholar] [CrossRef]
  183. Li, F.; Peng, X.; Liu, Y.; Mei, J.; Sun, L.; Shen, C.; Ma, C.; Huang, M.; Wang, Z.; Sand, W. A Chloride-Radical-Mediated Electrochemical Filtration System for Rapid and Effective Transformation of Ammonia to Nitrogen. Chemosphere 2019, 229, 383–391. [Google Scholar] [CrossRef]
  184. Zheng, W.; Zhu, L.; Liang, S.; Ye, J.; Yang, X.; Lei, Z.; Yan, Z.; Li, Y.; Wei, C.; Feng, C. Discovering the Importance of ClO in a Coupled Electrochemical System for the Simultaneous Removal of Carbon and Nitrogen from Secondary Coking Wastewater Effluent. Environ. Sci. Technol. 2020, 54, 9015–9024. [Google Scholar] [CrossRef]
  185. Huang, X.; Zhang, Y.; Bai, J.; Li, J.; Li, L.; Zhou, T.; Chen, S.; Wang, J.; Rahim, M.; Guan, X.; et al. Efficient Degradation of N-Containing Organic Wastewater via Chlorine Oxide Radical Generated by a Photoelectrochemical System. Chem. Eng. J. 2020, 392, 123695. [Google Scholar] [CrossRef]
  186. Li, F.; Sun, L.; Liu, Y.; Fang, X.; Shen, C.; Huang, M.; Wang, Z.; Dionysiou, D.D. A ClO -Mediated Photoelectrochemical Filtration System for Highly-Efficient and Complete Ammonia Conversion. J. Hazard. Mater. 2020, 400, 123246. [Google Scholar] [CrossRef] [PubMed]
  187. Zhang, C.; He, D.; Ma, J.; Waite, T.D. Active Chlorine Mediated Ammonia Oxidation Revisited: Reaction Mechanism, Kinetic Modelling and Implications. Water Res. 2018, 145, 220–230. [Google Scholar] [CrossRef] [PubMed]
  188. Sun, S.; Chen, R.; Wen, Y.; Li, Y.; Zhan, W.; Ma, F.; Jiang, X.; He, W.; Ni, H. Trace Amount of RuO2 Loaded on TiO2 Nanowires for Efficient Electrocatalytic Degradation of Ammonia Nitrogen in Wastewater. J. Alloys Compd. 2022, 928, 167058. [Google Scholar] [CrossRef]
  189. Hansen, H.A.; Man, I.C.; Studt, F.; Abild-Pedersen, F.; Bligaard, T.; Rossmeisl, J. Electrochemical Chlorine Evolution at Rutile Oxide (110) Surfaces. Phys. Chem. Chem. Phys. 2010, 12, 283–290. [Google Scholar] [CrossRef]
  190. Song, X.; Chen, X.; Zhang, S.; Wu, D. Treatment of High Chlorine-Containing Composting Leachate Biochemical Effluent by Ti/RuO2-IrO2 Anodic Electrochemical Oxidation: Optimization and Evolution of Pollutants. J. Environ. Chem. Eng. 2023, 11, 109674. [Google Scholar] [CrossRef]
  191. 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]
  192. Mandal, P.; Yadav, M.K.; Gupta, A.K.; Dubey, B.K. Chlorine Mediated Indirect Electro-Oxidation of Ammonia Using Non-Active PbO2 Anode: Influencing Parameters and Mechanism Identification. Sep. Purif. Technol. 2020, 247, 116910. [Google Scholar] [CrossRef]
  193. Zhang, Z.; Liu, J.; Ai, H.; Chen, A.; Xu, L.; Labiadh, L.; Fu, M.-L.; Yuan, B. Construction of the Multi-Layer TiO2-NTs/Sb-SnO2/PbO2 Electrode for the Highly Efficient and Selective Oxidation of Ammonia in Aqueous Solution: Characterization, Performance and Mechanism. J. Environ. Chem. Eng. 2023, 11, 109834. [Google Scholar] [CrossRef]
  194. Romano, A.; Ortiz, I.; Urtiaga, A.M. Comprehensive Kinetics of Electrochemically Assisted Ammonia Removal in Marine Aquaculture Recirculating Systems. J. Electroanal. Chem. 2021, 897, 115619. [Google Scholar] [CrossRef]
  195. Zhang, Y.; Li, J.; Bai, J.; Li, L.; Chen, S.; Zhou, T.; Wang, J.; Xia, L.; Xu, Q.; Zhou, B. Extremely Efficient Decomposition of Ammonia N to N2 Using ClO from Reactions of HO and HOCl Generated in Situ on a Novel Bifacial Photoelectroanode. Environ. Sci. Technol. 2019, 53, 6945–6953. [Google Scholar] [CrossRef]
  196. Garcia-Segura, S.; Mostafa, E.; Baltruschat, H. Electrogeneration of Inorganic Chloramines on Boron-Doped Diamond Anodes during Electrochemical Oxidation of Ammonium Chloride, Urea and Synthetic Urine Matrix. Water Res. 2019, 160, 107–117. [Google Scholar] [CrossRef] [PubMed]
  197. Sun, J.; Liu, L.; Yang, F. Electro-Enhanced Chlorine-Mediated Ammonium Nitrogen Removal Triggered by an Optimized Catalytic Anode for Sustainable Saline Wastewater Treatment. Sci. Total Environ. 2021, 776, 146035. [Google Scholar] [CrossRef] [PubMed]
  198. Wang, Y.; Zhang, Y.; Sun, W.; Zhu, L. The Impact of New Urbanization and Industrial Structural Changes on Regional Water Stress Based on Water Footprints. Sustain. Cities Soc. 2022, 79, 103686. [Google Scholar] [CrossRef]
  199. Gao, Y.; Chen, J.; Qian, H.; Wang, H.; Ren, W.; Qu, W. Hydrogeochemical Characteristics and Processes of Groundwater in an over 2260 Year Irrigation District: A Comparison between Irrigated and Nonirrigated Areas. J. Hydrol. 2022, 606, 127437. [Google Scholar] [CrossRef]
  200. Yao, F.; Yang, Q.; Zhong, Y.; Shu, X.; Chen, F.; Sun, J.; Ma, Y.; Fu, Z.; Wang, D.; Li, X. Indirect Electrochemical Reduction of Nitrate in Water Using Zero-Valent Titanium Anode: Factors, Kinetics, and Mechanism. Water Res. 2019, 157, 191–200. [Google Scholar] [CrossRef]
  201. Beltrame, T.F.; Carvalho, D.; Marder, L.; Ulla, M.A.; Marchesini, F.A.; Bernardes, A.M. Comparison of Different Electrode Materials for the Nitrate Electrocatalytic Reduction in a Dual-Chamber Cell. J. Environ. Chem. Eng. 2020, 8, 104120. [Google Scholar] [CrossRef]
  202. Paul, S.; Adalder, A.; Ghorai, U.K. Progress of Electrocatalytic Urea Synthesis: Strategic Design, Reactor Engineering, Mechanistic Details and Techno-Commercial Study. Mater. Chem. Front. 2023, 7, 3820–3854. [Google Scholar] [CrossRef]
  203. Mao, Q.; Deng, K.; Yu, H.; Xu, Y.; Wang, Z.; Li, X.; Wang, L.; Wang, H. In Situ Reconstruction of Partially Hydroxylated Porous Rh Metallene for Ethylene Glycol-Assisted Seawater Splitting. Adv. Funct. Mater. 2022, 32, 2201081. [Google Scholar] [CrossRef]
  204. Avila, Y.; Acevedo-Peña, P.; Reguera, L.; Reguera, E. Recent Progress in Transition Metal Hexacyanometallates: From Structure to Properties and Functionality. Coord. Chem. Rev. 2022, 453, 214274. [Google Scholar] [CrossRef]
  205. Hoa, V.H.; Austeria, M.; Thi Dao, H.; Mai, M.; Kim, D.H. Dual-Phase Cobalt Phosphide/Phosphate Hybrid Interactions via Iridium Nanocluster Interfacial Engineering toward Efficient Overall Seawater Splitting. Appl. Catal. B Environ. 2023, 327, 122467. [Google Scholar] [CrossRef]
  206. Sun, J.; Li, J.; Li, Z.; Li, C.; Ren, G.; Zhang, Z.; Meng, X. Modulating the Electronic Structure on Cobalt Sites by Compatible Heterojunction Fabrication for Greatly Improved Overall Water/Seawater Electrolysis. ACS Sustain. Chem. Eng. 2022, 10, 9980–9990. [Google Scholar] [CrossRef]
  207. Fan, J.; Fu, C.; Liang, R.; Lv, H.; Fang, C.; Guo, Y.; Hao, W. Mild Construction of “Midas Touch” Metal–Organic Framework-Based Catalytic Electrodes for Highly Efficient Overall Seawater Splitting. Small 2022, 18, 2203588. [Google Scholar] [CrossRef] [PubMed]
  208. Dong, J.; Yu, C.; Wang, H.; Chen, L.; Huang, H.; Han, Y.; Wei, Q.; Qiu, J. A Robust & Weak-Nucleophilicity Electrocatalyst with an Inert Response for Chlorine Ion Oxidation in Large-Current Seawater Electrolysis. J. Energy Chem. 2024, 90, 486–495. [Google Scholar] [CrossRef]
  209. Shen, Z.; Liu, D.; Peng, G.; Ma, Y.; Li, J.; Shi, J.; Peng, J.; Ding, L. Electrocatalytic Reduction of Nitrate in Water Using Cu/Pd Modified Ni Foam Cathode: High Nitrate Removal Efficiency and N2-Selectivity. Sep. Purif. Technol. 2020, 241, 116743. [Google Scholar] [CrossRef]
  210. Manzo-Robledo, A.; Lévy-Clément, C.; Alonso-Vante, N. The Interplay between Hydrogen Evolution Reaction and Nitrate Reduction on Boron-Doped Diamond in Aqueous Solution: The Effect of Alkali Cations. Electrochim. Acta 2014, 117, 420–425. [Google Scholar] [CrossRef]
  211. Fedurco, M.; Kedzierzawski, P.; Augustynski, J. Effect of Multivalent Cations upon Reduction of Nitrate Ions at the Ag Electrode. J. Electrochem. Soc. 1999, 146, 2569–2572. [Google Scholar] [CrossRef]
  212. Forero-Saboya, J.; Davoisne, C.; Dedryvère, R.; Yousef, I.; Canepa, P.; Ponrouch, A. Understanding the Nature of the Passivation Layer Enabling Reversible Calcium Plating. Energy Environ. Sci. 2020, 13, 3423–3431. [Google Scholar] [CrossRef]
  213. Liu, F.; Zhou, R.; Zhang, C.; Wu, Z.; Ren, H.; Ng, H.Y. Critical Review on the Pulsed Electrochemical Technologies for Wastewater Treatment: Fundamentals, Current Trends, and Future Studies. Chem. Eng. J. 2024, 479, 147588. [Google Scholar] [CrossRef]
  214. Jiang, Z.; Cheng, Z.; Yan, C.; Zhang, X.; Tian, Y.; Zhang, X.; Quan, X. Simultaneous Removal of Nitrogen and Refractory Organics from a Biologically Treated Leachate by Pulse Electrochemical Oxidation in a Multi-Channel Flow Reactor. ACS Omega 2021, 6, 25539–25550. [Google Scholar] [CrossRef]
  215. Taguchi, S.; Feliu, J.M. Electrochemical Reduction of Nitrate on Pt(S)[n(111)×(111)] Electrodes in Perchloric Acid Solution. Electrochim. Acta 2007, 52, 6023–6033. [Google Scholar] [CrossRef]
  216. Ding, J.; Li, W.; Zhao, Q.-L.; Wang, K.; Zheng, Z.; Gao, Y.-Z. Electroreduction of Nitrate in Water: Role of Cathode and Cell Configuration. Chem. Eng. J. 2015, 271, 252–259. [Google Scholar] [CrossRef]
  217. Reyter, D.; Bélanger, D.; Roué, L. Nitrate Removal by a Paired Electrolysis on Copper and Ti/IrO2 Coupled Electrodes—Influence of the Anode/Cathode Surface Area Ratio. Water Res. 2010, 44, 1918–1926. [Google Scholar] [CrossRef] [PubMed]
  218. Zhao, Q.; Tang, Z.; Chen, B.; Zhu, C.; Tang, H.; Meng, G. Efficient Electrocatalytic Reduction of Nitrate to Nitrogen Gas by a Cubic Cu2O Film with Predominant (111) Orientation. Chem. Commun. 2022, 58, 3613–3616. [Google Scholar] [CrossRef] [PubMed]
  219. 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]
  220. Abdallah, R.; Geneste, F.; Labasque, T.; Djelal, H.; Fourcade, F.; Amrane, A.; Taha, S.; Floner, D. Selective and Quantitative Nitrate Electroreduction to Ammonium Using a Porous Copper Electrode in an Electrochemical Flow Cell. J. Electroanal. Chem. 2014, 727, 148–153. [Google Scholar] [CrossRef]
  221. Makover, J.; Hasson, D.; Semiat, R.; Shemer, H. Electrochemical Removal of Nitrate from High Salinity Waste Stream in a Continuous Flow Reactor. J. Environ. Chem. Eng. 2020, 8, 103727. [Google Scholar] [CrossRef]
  222. Zheng, W.; Zhu, L.; Yan, Z.; Lin, Z.; Lei, Z.; Zhang, Y.; Xu, H.; Dang, Z.; Wei, C.; Feng, C. Self-Activated Ni Cathode for Electrocatalytic Nitrate Reduction to Ammonia: From Fundamentals to Scale-Up for Treatment of Industrial Wastewater. Environ. Sci. Technol. 2021, 55, 13231–13243. [Google Scholar] [CrossRef]
  223. Rivera, F.F.; Pérez, T.; Castañeda, L.F.; Nava, J.L. Mathematical Modeling and Simulation of Electrochemical Reactors: A Critical Review. Chem. Eng. Sci. 2021, 239, 116622. [Google Scholar] [CrossRef]
  224. Oriol, R.; Nava, J.L.; Brillas, E.; Cornejo, O.M.; Sirés, I. Modeling the Electrocatalytic Nitrate Removal in a Rotating Cylinder Electrode Reactor. Sep. Purif. Technol. 2024, 340, 126714. [Google Scholar] [CrossRef]
  225. Zhou, J.; Zhu, Y.; Wen, K.; Pan, F.; Ma, H.; Niu, J.; Wang, C.; Zhao, J. Efficient and Selective Electrochemical Nitrate Reduction to N2 Using a Flow-Through Zero-Gap Electrochemical Reactor with a Reconstructed Cu(OH)2 Cathode: Insights into the Importance of Inter-Electrode Distance. Environ. Sci. Technol. 2024, 58, 4824–4836. [Google Scholar] [CrossRef]
  226. Garcia-Segura, S.; Qu, X.; Alvarez, P.J.J.; Chaplin, B.P.; Chen, W.; Crittenden, J.C.; Feng, Y.; Gao, G.; He, Z.; Hou, C.-H.; et al. Opportunities for Nanotechnology to Enhance Electrochemical Treatment of Pollutants in Potable Water and Industrial Wastewater—A Perspective. Environ. Sci. Nano 2020, 7, 2178–2194. [Google Scholar] [CrossRef]
  227. Qiu, W.; Liu, Y.; Xie, M.; Jin, Z.; Li, P.; Yu, G. Structural Engineering of Catalysts for Ammonia Electrosynthesis from Nitrate: Recent Advances and Challenges. EES. Catal. 2024, 2, 202–219. [Google Scholar] [CrossRef]
  228. Wang, Y.; Wang, C.; Li, M.; Yu, Y.; Zhang, B. Nitrate Electroreduction: Mechanism Insight, in Situ Characterization, Performance Evaluation, and Challenges. Chem. Soc. Rev. 2021, 50, 6720–6733. [Google Scholar] [CrossRef] [PubMed]
  229. Shahid, U.B.; Siddharth, K.; Shao, M. Electrifying the Nitrogen Cycle: An Electrochemical Endeavor. Curr. Opin. Electrochem. 2021, 30, 100790. [Google Scholar] [CrossRef]
  230. McEnaney, J.M.; Blair, S.J.; Nielander, A.C.; Schwalbe, J.A.; Koshy, D.M.; Cargnello, M.; Jaramillo, T.F. Electrolyte Engineering for Efficient Electrochemical Nitrate Reduction to Ammonia on a Titanium Electrode. ACS Sustain. Chem. Eng. 2020, 8, 2672–2681. [Google Scholar] [CrossRef]
  231. Gao, J.; Shi, N.; Guo, X.; Li, Y.; Bi, X.; Qi, Y.; Guan, J.; Jiang, B. Electrochemically Selective Ammonia Extraction from Nitrate by Coupling Electron- and Phase-Transfer Reactions at a Three-Phase Interface. Environ. Sci. Technol. 2021, 55, 10684–10694. [Google Scholar] [CrossRef]
  232. Ganiyu, S.O.; Martínez-Huitle, C.A. The Use of Renewable Energies Driving Electrochemical Technologies for Environmental Applications. Curr. Opin. Electrochem. 2020, 22, 211–220. [Google Scholar] [CrossRef]
  233. Martínez-Huitle, C.A.; Rodrigo, M.A.; Scialdone, O. (Eds.) Electrochemical Water and Wastewater Treatment; Butterworth-Heinemann: Oxford, UK; Cambridge, MA, USA, 2018; ISBN 978-0-12-813160-2. [Google Scholar]
  234. Lacasa, E.; Cotillas, S.; Saez, C.; Lobato, J.; Cañizares, P.; Rodrigo, M.A. Environmental Applications of Electrochemical Technology. What Is Needed to Enable Full-Scale Applications? Curr. Opin. Electrochem. 2019, 16, 149–156. [Google Scholar] [CrossRef]
  235. Millán, M.; Fernández-Marchante, C.M.; Lobato, J.; Cañizares, P.; Rodrigo, M.A. Management of Solar Energy to Power Electrochemical Wastewater Treatments. J. Water Process Eng. 2021, 41, 102056. [Google Scholar] [CrossRef]
  236. Kani, N.C.; Gauthier, J.A.; Prajapati, A.; Edgington, J.; Bordawekar, I.; Shields, W.; Shields, M.; Seitz, L.C.; Singh, A.R.; Singh, M.R. Solar-Driven Electrochemical Synthesis of Ammonia Using Nitrate with 11% Solar-to-Fuel Efficiency at Ambient Conditions. Energy Environ. Sci. 2021, 14, 6349–6359. [Google Scholar] [CrossRef]
  237. Tawalbeh, M.; Al-Othman, A.; Kafiah, F.; Abdelsalam, E.; Almomani, F.; Alkasrawi, M. Environmental Impacts of Solar Photovoltaic Systems: A Critical Review of Recent Progress and Future Outlook. Sci. Total Environ. 2021, 759, 143528. [Google Scholar] [CrossRef] [PubMed]
  238. Gonçalves, S.; Rodrigues, T.P.; Chagas, A.L.S. The Impact of Wind Power on the Brazilian Labor Market. Renew. Sustain. Energy Rev. 2020, 128, 109887. [Google Scholar] [CrossRef]
  239. Strielkowski, W.; Civín, L.; Tarkhanova, E.; Tvaronavičienė, M.; Petrenko, Y. Renewable Energy in the Sustainable Development of Electrical Power Sector: A Review. Energies 2021, 14, 8240. [Google Scholar] [CrossRef]
  240. Shahsavari, A.; Akbari, M. Potential of Solar Energy in Developing Countries for Reducing Energy-Related Emissions. Renew. Sustain. Energy Rev. 2018, 90, 275–291. [Google Scholar] [CrossRef]
  241. Sun, Y.; Bai, S.; Wang, X.; Ren, N.; You, S. Prospective Life Cycle Assessment for the Electrochemical Oxidation Wastewater Treatment Process: From Laboratory to Industrial Scale. Environ. Sci. Technol. 2023, 57, 1456–1466. [Google Scholar] [CrossRef]
  242. Nath, S. Electrochemical Wastewater Treatment Technologies Through Life Cycle Assessment: A Review. ChemBioEng Rev. 2024, 11, e202400016. [Google Scholar] [CrossRef]
  243. Mosalpuri, M.; Li, W.; Mba Wright, M. Techno-Economic Analysis and Life Cycle Assessment of Hydroxylamine Eco-Manufacturing via Wastewater Electrochemical Reduction. ACS Sustain. Chem. Eng. 2023, 11, 13636–13645. [Google Scholar] [CrossRef]
  244. Erisman, J.W.; Sutton, M.A.; Galloway, J.; Klimont, Z.; Winiwarter, W. How a Century of Ammonia Synthesis Changed the World. Nat. Geosci. 2008, 1, 636–639. [Google Scholar] [CrossRef]
  245. Liu, H. Ammonia Synthesis Catalyst 100 Years: Practice, Enlightenment and Challenge. Chin. J. Catal. 2014, 35, 1619–1640. [Google Scholar] [CrossRef]
Figure 1. Electrochemical nitrate reduction mechanisms and pathways in water.
Figure 1. Electrochemical nitrate reduction mechanisms and pathways in water.
Applsci 14 08986 g001
Figure 2. Proposed Pathways for Nitrate Reduction by Duca−Felio−Koper, Voys−Kopper−Chumanov, and ammonia−hydroxylamine formation.
Figure 2. Proposed Pathways for Nitrate Reduction by Duca−Felio−Koper, Voys−Kopper−Chumanov, and ammonia−hydroxylamine formation.
Applsci 14 08986 g002
Figure 3. The proposed direct electron reduction and indirect adsorbed hydrogen reduction in nitrate electroreduction.
Figure 3. The proposed direct electron reduction and indirect adsorbed hydrogen reduction in nitrate electroreduction.
Applsci 14 08986 g003
Figure 4. Three possible configurations for the adsorption of NO 3 .
Figure 4. Three possible configurations for the adsorption of NO 3 .
Applsci 14 08986 g004
Figure 5. (A) The binding energies of (a) * NO 3 , (b) * NO 2 , and (c) * NO adsorbates on different transition metals and the horizontal dashed line indicates the chemical potential NO 3 [66].(B) Volcano plot showing limiting potentials for (a) Ni, Rh, Pt, Pd, and Cu based on binding energies of NO and (b) Co, Ru, and Ir based on binding energies of OH. The horizontal dashed lines in (a) and (b) depict the equilibrium potential for the eNO3R to ammonia. The vertical dashed line in (a) depicts the Gibbs free energy of NO(g) [66]. (C) Theoretical volcano plots of the TOF as a function of atomic oxygen (ΔEO) and nitrogen (ΔEN) adsorption energies for electrocatalytic nitrate reduction on transition metal surfaces based on DFT-based microkinetic simulations at (a) −0.2 V, (b) 0 V, (c) 0.2 V, and (d) 0.4 V vs. RHE. Reaction conditions are T = 300 K with a H + / NO 3 Molar ratio of 1:1. White indicates unphysical regions where the activation energies of some elementary steps are negative. Ag falls within the white region due to errors associated with using linear scaling relationships [67]. (D) (a) Theoretical limiting potential UL for nine double-atom catalysts supported on N-doped graphene ( TM 2 / N 6 G ) screened out by stability analysis. (b) Volcano plot for the NO3R limiting potential on TM 2 / N 6 G as a function of the descriptor φ. (c) Volcano plot for the NO3R limiting potential on TM 2 / N 6 G as a function of the d-orbital (number electrons in d-orbital). Cr 2 , Mn 2 , Fe 2 , and Cu 2 / N 6 G stands near the top of the volcano plot [68].
Figure 5. (A) The binding energies of (a) * NO 3 , (b) * NO 2 , and (c) * NO adsorbates on different transition metals and the horizontal dashed line indicates the chemical potential NO 3 [66].(B) Volcano plot showing limiting potentials for (a) Ni, Rh, Pt, Pd, and Cu based on binding energies of NO and (b) Co, Ru, and Ir based on binding energies of OH. The horizontal dashed lines in (a) and (b) depict the equilibrium potential for the eNO3R to ammonia. The vertical dashed line in (a) depicts the Gibbs free energy of NO(g) [66]. (C) Theoretical volcano plots of the TOF as a function of atomic oxygen (ΔEO) and nitrogen (ΔEN) adsorption energies for electrocatalytic nitrate reduction on transition metal surfaces based on DFT-based microkinetic simulations at (a) −0.2 V, (b) 0 V, (c) 0.2 V, and (d) 0.4 V vs. RHE. Reaction conditions are T = 300 K with a H + / NO 3 Molar ratio of 1:1. White indicates unphysical regions where the activation energies of some elementary steps are negative. Ag falls within the white region due to errors associated with using linear scaling relationships [67]. (D) (a) Theoretical limiting potential UL for nine double-atom catalysts supported on N-doped graphene ( TM 2 / N 6 G ) screened out by stability analysis. (b) Volcano plot for the NO3R limiting potential on TM 2 / N 6 G as a function of the descriptor φ. (c) Volcano plot for the NO3R limiting potential on TM 2 / N 6 G as a function of the d-orbital (number electrons in d-orbital). Cr 2 , Mn 2 , Fe 2 , and Cu 2 / N 6 G stands near the top of the volcano plot [68].
Applsci 14 08986 g005
Figure 6. Schematic illustration of the possible hydrogenation of nitrous oxide including O-end, O-side, N-end, and N-side pathway to NH 3 .
Figure 6. Schematic illustration of the possible hydrogenation of nitrous oxide including O-end, O-side, N-end, and N-side pathway to NH 3 .
Applsci 14 08986 g006
Figure 7. (a) The differences between the limiting potentials for the eNO3R and HER, i.e., UL(eNO3R) − UL( H 2 ), are plotted against the limiting potentials for the eNO3R, UL(eNO3R), for different transition metals. UL(eNO3R) − UL( H 2 ) shows the trend in selectivity for the eNO3R over the HER, and UL(eNO3R) reflects the trend in eNO3R activity. The most promising catalysts lie in the upper-right corner of the plot [66]. (b) Gibbs free energy diagram of the hydrogen evolution reaction (HER) on the TM 2 / N 6 G (TM = Sc to Zn) surface [68].
Figure 7. (a) The differences between the limiting potentials for the eNO3R and HER, i.e., UL(eNO3R) − UL( H 2 ), are plotted against the limiting potentials for the eNO3R, UL(eNO3R), for different transition metals. UL(eNO3R) − UL( H 2 ) shows the trend in selectivity for the eNO3R over the HER, and UL(eNO3R) reflects the trend in eNO3R activity. The most promising catalysts lie in the upper-right corner of the plot [66]. (b) Gibbs free energy diagram of the hydrogen evolution reaction (HER) on the TM 2 / N 6 G (TM = Sc to Zn) surface [68].
Applsci 14 08986 g007
Figure 8. Theoretical (a) and experimental (b) Faradaic efficiencies of different products for eNORR on Ag [73]. (c) Microkinetically modeled and (d) Experimentally measured potential-dependent nitrate rate order, measured by steady-state chronoamperometry in 0.1 M Na x H 3 xPO 4 with a series of sodium nitrate concentrations for Cu (golden), Ni 0.68 Cu 0.32 (light blue), and Ni (orange) [74].
Figure 8. Theoretical (a) and experimental (b) Faradaic efficiencies of different products for eNORR on Ag [73]. (c) Microkinetically modeled and (d) Experimentally measured potential-dependent nitrate rate order, measured by steady-state chronoamperometry in 0.1 M Na x H 3 xPO 4 with a series of sodium nitrate concentrations for Cu (golden), Ni 0.68 Cu 0.32 (light blue), and Ni (orange) [74].
Applsci 14 08986 g008
Figure 9. Schematic illustration of NH3 yield rate and FE for MnCuO x catalyst without plasma treatment, with Ar plasma treatment, and with O 2 plasma treatment [88].
Figure 9. Schematic illustration of NH3 yield rate and FE for MnCuO x catalyst without plasma treatment, with Ar plasma treatment, and with O 2 plasma treatment [88].
Applsci 14 08986 g009
Figure 10. Comparison between the performance of different transition metal-based electrocatalysts for electrochemical nitrate reduction to ammonia (yellow zone is required for application).
Figure 10. Comparison between the performance of different transition metal-based electrocatalysts for electrochemical nitrate reduction to ammonia (yellow zone is required for application).
Applsci 14 08986 g010
Figure 11. (A) Cyclic voltammograms obtained from Cu(111) (a) and Cu(100) (b) surfaces in 0.1 M HClO4 + 1 mM HNO3. (B) In situ EC-STM images of the Cu(111) surface were obtained in 1 mM HNO3 + 0.1 M HF during cathodic potential sweeping between −0.35 V–−0.47 V (a), −0.57 V–−0.69 V (b), and −0.79 V–−0.80 V. (c) The image size is 150 nm × 150 nm. The tip bias and tunneling current are 10 mV and 10 nA, respectively [166].
Figure 11. (A) Cyclic voltammograms obtained from Cu(111) (a) and Cu(100) (b) surfaces in 0.1 M HClO4 + 1 mM HNO3. (B) In situ EC-STM images of the Cu(111) surface were obtained in 1 mM HNO3 + 0.1 M HF during cathodic potential sweeping between −0.35 V–−0.47 V (a), −0.57 V–−0.69 V (b), and −0.79 V–−0.80 V. (c) The image size is 150 nm × 150 nm. The tip bias and tunneling current are 10 mV and 10 nA, respectively [166].
Applsci 14 08986 g011
Figure 12. The possible reaction mechanism of enhanced nitrate reduction on the Cu/Ni foam electrode [84].
Figure 12. The possible reaction mechanism of enhanced nitrate reduction on the Cu/Ni foam electrode [84].
Applsci 14 08986 g012
Figure 13. Schematic illustration showing the electrocatalytic nitrate-to-ammonia process over the (a) Cu/Pd/CuOx [172] and (b) Pd-Cu2O [173].
Figure 13. Schematic illustration showing the electrocatalytic nitrate-to-ammonia process over the (a) Cu/Pd/CuOx [172] and (b) Pd-Cu2O [173].
Applsci 14 08986 g013
Figure 14. Mechanisms for ammonia oxidation by breakpoint chlorination and chlorine radical.
Figure 14. Mechanisms for ammonia oxidation by breakpoint chlorination and chlorine radical.
Applsci 14 08986 g014
Figure 15. Pourbaix diagram of nitrogenous forms [202].
Figure 15. Pourbaix diagram of nitrogenous forms [202].
Applsci 14 08986 g015
Figure 16. Schematic diagrams for (a) undivided flow electrochemical reactor for nitrate reduction with selective cathode for nitrogen selectivity; (b) undivided flow electrochemical reactor for nitrate reduction with selective cathode for ammonia selectivity and anode with highest OER; and (c) undivided flow electrochemical reactor for nitrate reduction with selective cathode for ammonia selectivity and cationic exchange membrane (CEM).
Figure 16. Schematic diagrams for (a) undivided flow electrochemical reactor for nitrate reduction with selective cathode for nitrogen selectivity; (b) undivided flow electrochemical reactor for nitrate reduction with selective cathode for ammonia selectivity and anode with highest OER; and (c) undivided flow electrochemical reactor for nitrate reduction with selective cathode for ammonia selectivity and cationic exchange membrane (CEM).
Applsci 14 08986 g016
Figure 17. Schematic representations of (a) the flow cell with porous electrode for nitrate reduction to ammonia [220]. (b) Electrochemical continuous flow system with copper as cathode and DSA as anode [221]. (c) The flow-through electrocatalytic system for nitrate reduction with the zero-gap electrochemical reactor [225]. (d) Schematic diagram of the pilot-scale electrochemical reactor [222].
Figure 17. Schematic representations of (a) the flow cell with porous electrode for nitrate reduction to ammonia [220]. (b) Electrochemical continuous flow system with copper as cathode and DSA as anode [221]. (c) The flow-through electrocatalytic system for nitrate reduction with the zero-gap electrochemical reactor [225]. (d) Schematic diagram of the pilot-scale electrochemical reactor [222].
Applsci 14 08986 g017
Figure 18. (a) Preliminary techno-economic analysis considering only the electricity cost required to produce ammonium nitrate from NO 3 based on a given price of electricity, cell efficiency, and total cell potential applied. The brown region shows USDA data for the cost of ammoniumnitrate from 2004 to 2014, with the average cost per metric ton in this period as an inset brown dashed line [230]. (b) Electric energy cost as a function of energy-related parameters and unit renewable energy cost [231]. (c) Energy-related parameters and the corresponding electric energy cost as a function of NO 3 concentration [231].
Figure 18. (a) Preliminary techno-economic analysis considering only the electricity cost required to produce ammonium nitrate from NO 3 based on a given price of electricity, cell efficiency, and total cell potential applied. The brown region shows USDA data for the cost of ammoniumnitrate from 2004 to 2014, with the average cost per metric ton in this period as an inset brown dashed line [230]. (b) Electric energy cost as a function of energy-related parameters and unit renewable energy cost [231]. (c) Energy-related parameters and the corresponding electric energy cost as a function of NO 3 concentration [231].
Applsci 14 08986 g018
Table 1. Standard redox potentials for different NO 3 reduction reactions.
Table 1. Standard redox potentials for different NO 3 reduction reactions.
Reduction ReactionStandard Redox Potential
2 NO 3 (aq) + 12 H+ + 10 e N 2 (g) + 6 H2O(l) E NO 3 / N 2 0 = + 1.25 V
NO 3 (aq) + 10 H+ + 8 e NH 4 + (aq) + 3 H2O(l) E NO 3 / NH 4 + 0 = + 1.20 V
2 NO 3 (aq) + 10 H+ + 8 e N 2 O(g) + 5 H2O(l) E NO 3 / N 2 O 0 = + 1.12 V
NO 3 (aq) + 4 H+ + 3 e ⇌ NO(g) + 2 H2O(l) E NO 3 / NO 0 = + 0.96 V
2   NO 3 (aq) + 10 H+ + 8 e H 2 N 2 O 2 (g) + 4 H2O(l) E NO 3 / H 2 N 2 O 2 0 = + 0.90 V
NO 3 (aq) + 2 H+ + 2 e NO 2 (aq) + H2O(l) E NO 3 / NO 2 0 = + 0.89 V
NO 3 (aq) + 9 H+ + 8e NH 3 (aq) + 3H2O(l) E NO 3 / NH 3 0 = + 0.82 V
NO 3 (aq) + 2 H+ + e NO 2 (g) + H2O(l) E NO 3 / NO 2 0 = + 0.80 V
NO 3 (aq) + 7 H+ + 6 e NH 2 OH (aq) + 2 H2O(l) E NO 3 / NH 2 OH 0 = + 0.67 V
NO 3 (aq) + 7 H2O + 8 e NH 4 OH (aq) + 9 OH (l) E NO 3 / NH 4 OH   0 = −0.12 V
2   NO 3 (aq) + 10 H2O + 14 e N 2 H 4 (aq) + 16 OH (l) E NO 3 / NH 4 OH   0 = −0.15 V
Table 2. Pathways and limiting steps for novel categories synthesizing electrode materials for eNO3R using DFT calculations.
Table 2. Pathways and limiting steps for novel categories synthesizing electrode materials for eNO3R using DFT calculations.
Electrode MaterialsFree Energy DiagramRate Determining StepsRefs.
Cu catalyst anchored on porous N-doped carbon.
(Cu−NC)
Applsci 14 08986 i001*NO to *N[52]
CuO(111) surface with two oxygen vacancy (OVs)
(CuO−2OVs)
Applsci 14 08986 i002The protonation of *O-*N to*N-*OH[53]
MnCo 2 O 4 (3 1 1) facetApplsci 14 08986 i003*NO to *NOH[54]
MnCo 2 O 4 (2 2 0) facetApplsci 14 08986 i004* N 2 O 2 H to * N 2 O [54]
Au (111) (in yellow),
Au (331) (in blue),
Pd modified Au(331) (in red).
Applsci 14 08986 i005 NO 3 (ads) NO 2 (ads) + O(ads)
Au (111) = 0.72 eV
Au (331) = 0. 39 eV
Pd-Au (331) = 0.37 eV
NO 2 (ads) NO (ads) + O(ads)
Au (111) = 1.74 eV
Au (331) = 0.83 eV
Pd-Au (331) = 0.23 eV
[55]
nanoporous Co 3 O 4 / Co Applsci 14 08986 i006* NO 2 to * NO [56]
Pd (in blue),
Pd ( WO 3 ) 3 (in red).
Applsci 14 08986 i007*N to *NH
Pd = 3.38 eV
* NH 2 to * NH 3
Pd ( WO 3 ) 3 = 2.99 eV
[57]
Cu NWAs (in red), Cu / Cu 2 O NWAs (in yellow)Applsci 14 08986 i008The protonation of *O-*N to*N-*OH [58]
Ni 3 Co 6 S 8 (NCS) (in blue)
Ni 3 Co 6 S 8 nanospheres with sulfur-rich vacancies (NCS-2) (in red)
Applsci 14 08986 i009* HNO 3 to * NO 2
NCS-2 = 0.24 eV
NCS-2 = 0.19 eV
[59]
The incorporation of indium in sulfur-doped graphene
(In-S-G)
Applsci 14 08986 i010 N 2 formation pathway
*NO to * N 2 O 3 (0.90 eV)
NH 3 production
* NO 3 to * NO 2 OH (0.63 eV)
[60]
Boron-doped copper nanowire electrocatalyst (B-Cu)Applsci 14 08986 i011The hydrogenation of *NO to form H*NO[61]
Cobalt manganese spinel nanoparticles ( CoMn 2 O 4 ) Applsci 14 08986 i012The desorption of * NH 3 [62]
Cobalt manganese spinel nanoparticles embedded in multichannel carbon fibers ( CoMn 2 O 4 / NC )Applsci 14 08986 i013*NO-to-*NHO[62]
Iron coordinated with nitrogen on an ordered mesoporous carbon framework (Meso-Fe-N-C)Applsci 14 08986 i014 N 2 formation pathway
N 2 desorption
NH 3 production
*NO-to-*NHO
[63]
Table 3. In situ Characterization Techniques for eNO3R.
Table 3. In situ Characterization Techniques for eNO3R.
Characterization
Techniques
Detected Products and IntermediatesPrinciplesAdvantagesDisadvantagesRefs.
Shell isolated nanoparticle enhanced Raman spectroscopy (SHINERS)All the intermediates adsorbed on different faces of catalyst material.The principle is based on measuring the inelastic scattering of light caused by molecular vibrations involving changes in the polarization of the molecule. It consisted of a gold nanoparticle with an insulating thin SiO2 shell (Au/SiO2) to eliminate the interference of deposited Au or other substrates (Ag, Au), which have a surface plasmon resonance effect, and to inhibit the signal from the analytes adsorbed on the Au core.High sensitivity, rapid, less destructive, and noninvasive technique.It cannot be used for the detection of atomic cations and anions directly, complicated process, and has a weak pH tolerance.[102,103]
Fourier-transformed infrared spectroscopy (FT-IR)All intermediates formed at the surface of the electrode.The incident infrared ray is guided by the prism and optical stage toward the chemical bonds of intermediates formed at the surface of the electrode, which was then diffused and reflected in the detector. The FT-IR measurements were performed on a spectrometer equipped with a liquid-nitrogen-cooled detector.Rapid determination of adsorbed intermediates, applied in wide concentration ranges, nondestructive, and repeatable.FTIR instruments have only a single beam.[104,105,106]
X-Ray Absorption SpectroscopyActive sites at the surface of the electrode.The method involves selecting the desired energy of X-ray radiation using a monochromator and measuring the resulting XAS spectrum, which is composed of two parts: the X-ray absorption near-edge structure (XANES) and the extended X-ray absorption fine structure (EXAFS). The XANES spectrum provides information on the electronic structure of the element being studied, while the EXAFS spectra give information on the local geometry of the atom. This technique is well suited for identifying active sites in multi-element systems and can measure elements at low concentrations due to their high flux properties.No crystal and ordered structure is needed for the sample additionally the method works for single-atom materials too.XAS data represents an average of the entire sample and Soft XAS requires an ultra-high or high vacuum environment. XAS may not provide detailed information about specific regions of the sample.[107,108,109]
Electrochemical Scanning Tunneling Microscopysurface morphologyIn situ, electrochemical scanning tunneling microscopy (EC-STM) is a form of scanning probe microscopy that uses a metal needle tip and an electrochemical control circuit to acquire high-resolution images of a sample surface. It operates on the quantum tunneling effect and can work in various environments, including ultra-high vacuum and liquid environments.Applicable to a wide range of environments, Wide working temperature range, and high resolution.Damage risk in constant height mode on undulating surfaces, cannot detect bulk structure information of the sample, cannot directly observe insulator materials, and the resolution of the image to the real surface is limited if the sample surface is covered with a conductive layer.[110,111]
Table 4. Ammonia production with different Cu-based materials and nitrogen selectivity with various Fe-based catalysts.
Table 4. Ammonia production with different Cu-based materials and nitrogen selectivity with various Fe-based catalysts.
CathodeAmmonia ProductionNitrogen SelectivityRefs.
ConditionsPerformanceConditionsPerformance
Copper-Based Electrocatalysts
Cu Single Atom Gel (SAG)0.1 M phosphate buffer solution (PBS), pH 7, 20 mM KNO3, and Overpotential −0.9 V vs. Reference Hydrogen Electrode (RHE) NH 3 yield rate: 0.53 mg cm−2 h−1; FE: 95.3%--[101]
Cu-N-C SAC
Copper into porous nitrogen-doped carbon
0.1 M KOH, 0.1 M KNO3, and Overpotential − 1.00 V vs. RHE NH 3 yield rate: 4.5 mg cm−2 h¹; FE: 84.7%--[135]
Cu nanodisks (111)0.1 M KOH, 10 mM KNO3, 10 mM KNO2 and Overpotential −0.5 V vs. RHE NH 3 yield rate: 2.18 mg mgcat−1 h¹; FE: 81.1% [136]
Cu nanosheets0.1 M KOH
170 ppm NO3N and Overpotential −0.15 V vs. RHE
NH 3 production rate: 0.39 mg mgcat−1 h¹; FE: 99.7%--[137]
Au/Cu0.5 M NaNO3, 1 M NaOH, and Overpotential −0.7 V vs. RHE NH 3 yield rate: 73.4 mg cm−2 h−1; FE: 98.02%--[138]
Cu50Co501 M KOH, 100 mM KNO3, and Overpotential −0.2 V vs. RHE NH 3 yield rate: 81.7 mg cm−2 h¹; FE: 100 ± 1%--[139]
Cu/Cu2O0.5 M Na2SO4
14.3 mM NO 3 and Overpotential −0.85 V vs. RHE
NH 3 production rate: 4.088 mg cm−2 h−1; FE: 95.8%--[58]
Cu2O-Cu/Ti0.1 M KNO3, 1 M KOH, and Overpotential 0.5 V vs. RHE NH 3 yield rate: 4.76 mg cm−2 h−1; FE: 92%--[140]
Cu/CuOx/Co/CoO0.1 M NO 3 , pH 13, 0.1 M KNO3, and Overpotential −0.175 V RHE NH 3 yield rate: 19.94 mg cm−2 h−1; FE: 93.3%--[141]
Iron-based electrocatalysts
Fe--3.5 mM Na2SO4 + 7 mM NaNO3 electrolysis duration of 3 h.87% nitrate reduction efficiency with 100% nitrogen selectivity.[142]
nZVI@OMC
mesoporous carbon supported nanoscale zerovalent iron
--50 mg/L NO 3
N;0.02 M NaCl and 24 h of electrolysis.
65% nitrate conversion with 74% nitrogen selectivity.[143]
nZVI@D201
polymeric beads supported by nZVI
--50 mg/L NO 3 –N;
1.0 M NaCl and 60 h of electrolysis.
80% nitrate conversion with 95% nitrogen selectivity.[144]
FeNC/MC
Iron over N-doped mesoporous Carbon
--100 mg/L NO 3 –N;
0.1 M Na2SO4 and 24 h of electrolysis.
87% nitrate conversion with 81% nitrogen selectivity.[145]
Fe#OMC
Fe(0)-dispersed ordered mesoporous carbon
--50 mg/L NO 3 –N;
0.1 M Na2SO4;
0.02 M NaCl and 24 h of electrolysis.
86.9% nitrate conversion with 100% nitrogen selectivity.[146]
Fe@PMO
Iron Nanoparticles Confined in Periodic Mesoporous Organosilicon
--100 mg/L NO 3 –N;
0.02 M Na2SO4;
0.02 M NaCl and 24 h of electrolysis.
90% nitrate conversion with 99% nitrogen selectivity.[147]
FeN-NC
N-doped Fe nanoparticles
--100 mg/L NO 3 –N;
0.1 M Na2SO4;
0.02 M NaCl and 24 h of electrolysis.
91% nitrogen selectivity[148]
Fe@N-C
N-doped graphitic carbon-encapsulated iron nanoparticles
--50 mg/L NO 3 –N;
1.0 g/L NaCl and 24 h of electrolysis.
83% nitrate conversion with 100% nitrogen selectivity[149]
FeNi/g-mesoC/NF
binderless FeNi/graphitized mesoporous carbon on Ni Foam
--50 mg/L NO 3 –N;
0.05 M Na2SO4;
0.02 M NaCl and 24 h of electrolysis.
75% nitrate conversion with 100% nitrogen selectivity[150]
Fe3Ni-N-C
Fe/Ni bimetallic nitrogen-doped porous carbon
--100 ppm NO 3 -N;
0.1 M Na2SO4 and 0.5 h of electrolysis.
97.9% nitrate conversion with 99.3% nitrogen selectivity.[151]
Fe0/Ni2P/CC--15 mg/L NO 3 –N;
0.3 M Na2SO4 and 4 h of electrolysis.
89.81% nitrate conversion with 95.55% nitrogen selectivity.[152]
Ni-Fe0@Fe3O4
Fe0@Fe3O4 nanoparticles immobilized on nickel foam
--0.5 g/L Cl-, 50 mg/L NO 3 , Current density of 5 mA/cm2, electrolysis duration of 4 h, and pH of 6.2.90.19% nitrate conversion with 88.09% nitrogen selectivity.[153]
CL-Fe@C
Iron Nanoparticles in Carbon Microspheres
--100 mg/L NO 3 –N;
0.02 M NaCl and 48 h of electrolysis.
54% nitrate conversion with 98% nitrogen selectivity.[154]
CC/Fe@C
carbon-coated iron nanoparticles on a functionalized carbon cloth
--50 mg/L NO 3 –N;
0.01 M Na2SO4;
0.01 M NaCl and 24 h of electrolysis.
92% nitrate conversion with 82% nitrogen selectivity.[155]
B-Fe NCs
boron-iron nanochains
--100 mg/L NO 3 –N;
0.02 M Na2SO4;
0.02 M NaCl and 24 h of electrolysis.
80% nitrate conversion with 99% nitrogen selectivity.[156]
Fe@C-1
iron nanoparticles uniformly embedded in the carbon microsphere
--100 mg/L NO 3 –N;
0.02 M NaCl and 48 h of electrolysis.
75.9% nitrate conversion with 98% nitrogen selectivity[157]
Fe/Fe3C-NCNF-2
Fe/Fe3C nanoparticle-decorated N-doped carbon nanofibers
100 mg/L NO 3 –N;
0.02 M NaCl and 24 h of electrolysis.
95% nitrate conversion with 100% nitrogen selectivity.[158]
Table 5. Parameters Affecting the Oxidation Process of Ammonium generated.
Table 5. Parameters Affecting the Oxidation Process of Ammonium generated.
Influence FactorsDescriptionsRefs.
Initial chloride concentrationhigh initial concentration of chlorides increases the production of active chlorine and Cl , thereby enhancing Ammonia oxidation[185,190,191]
pHpH decreases during the process.
Acidic and neutral media are preferable for ammonia oxidation.
Alkaline media decrease ammonia oxidation efficiency
[185,192,193]
Current densityIncreasing current density improves ammonia oxidation efficiency (rule applied for lower current density than 37.5 mA·cm−2)[186,192]
Ammonia concentrationHigh ammonia in solution inhibits the formation of active chlorine and Cl [194,195]
TemperatureIncreasing temperature promotes the oxidation of ammonia. However, at temperatures higher than 55 °C the efficiency decreases due to the lose of active chlorine.[193]
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

Outaleb, H.; Kouzbour, S.; Audonnet, F.; Vial, C.; Gourich, B. Electrocatalytic Nitrate Reduction for Brackish Groundwater Treatment: From Engineering Aspects to Implementation. Appl. Sci. 2024, 14, 8986. https://doi.org/10.3390/app14198986

AMA Style

Outaleb H, Kouzbour S, Audonnet F, Vial C, Gourich B. Electrocatalytic Nitrate Reduction for Brackish Groundwater Treatment: From Engineering Aspects to Implementation. Applied Sciences. 2024; 14(19):8986. https://doi.org/10.3390/app14198986

Chicago/Turabian Style

Outaleb, Hamza, Sanaa Kouzbour, Fabrice Audonnet, Christophe Vial, and Bouchaib Gourich. 2024. "Electrocatalytic Nitrate Reduction for Brackish Groundwater Treatment: From Engineering Aspects to Implementation" Applied Sciences 14, no. 19: 8986. https://doi.org/10.3390/app14198986

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