Next Article in Journal
Current Progress of Hederagenin and Its Derivatives for Disease Therapy (2017–Present)
Previous Article in Journal
Treatment of Produced Water Using a Pilot-Scale Advanced Electrochemical Oxidation Unit
Previous Article in Special Issue
New Cyclam-Based Fe(III) Complexes Coatings Targeting Cobetia marina Biofilms
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

In Situ Formation of FeNi Nanoparticles on Polypyrrole Hydrogel for Efficient Electrocatalytic Nitrate Reduction to Ammonia

1
School of Environment and Safety Engineering, Jiangsu University, Zhenjiang 212013, China
2
School of Chemistry and Chemical Engineering, Jiangsu University, Zhenjiang 212013, China
3
College of Mechanical Engineering, Yanshan University, Qinhuangdao 066004, China
*
Authors to whom correspondence should be addressed.
These authors contributed equally to this work.
Molecules 2025, 30(6), 1271; https://doi.org/10.3390/molecules30061271
Submission received: 7 January 2025 / Revised: 1 March 2025 / Accepted: 3 March 2025 / Published: 12 March 2025

Abstract

:
The electrocatalytic reduction of nitrate to ammonia (NH3) under mild environmental conditions is attracting increasing attention, in which efficient and inexpensive transition metal catalysts, with the advantages of abundancy and low cost, play a key role. However, synergistic activity and selectivity promotion are still highly challenging. Herein, we developed a hydrogel-assisted strategy to prepare FeNi nanoparticles via the in situ adsorption and reduction of Fe/Ni precursors on a polypyrrole hydrogel. After optimization, the maximum NH3 yield reached 0.166 mmol h−1 cm−2, with a Faradaic efficiency of 88.9% and a selectivity of 86.6%. This excellent electrochemical performance was attributed to the mesoporous hydrophilic structure of the polypyrrole hydrogel, which facilitates the homogeneous loading of FeNi nanoparticles and provides a channel for both charge and mass transfer during nitrate reduction, which is important for the conversion of NO3 to NH3. Electrochemical active surface area determination and impedance spectroscopy showed that the introduction of hydrogel increased the active sites and improved the charge transfer. This study provides an effective strategy for improving the selectivity and yield of NH3 in transition metal electrocatalysts by utilizing the three-dimensional hydrogel network and electrical conductivity.

1. Introduction

Ammonia (NH3), as a one of the most widely used chemicals, not only plays an important role in fertilizer and pharmaceuticals but also as acts as a promising environmentally friendly and carbon-free energy carrier for sustainable applications [1,2,3]. The production of NH3 in an industrial context mainly occurs through the Haber–Bosch process from nitrogen and hydrogen at high temperatures and pressure, which leads to greenhouse gas emission, huge energy consumption, and security problems [4,5]. Therefore, it is crucial to seek new strategies to replace traditional methods. At present, the electrocatalytic reduction of nitrogen species, such as through the nitrate reduction reaction (NO3RR), is as an efficient method for ammonia synthesis under mild environmental conditions [6,7,8]. The fact that NO3 is more readily reduced to ammonia compared to N2 can be attributed to the lower dissociation energy of N=O bonds (204 kJ mol−1) [9] in contrast with the significantly higher dissociation energy of the N≡N triple bonds (941 kJ mol−1) [10,11]. However, the complex NO3RR process involves eight-electron transfer and the production of many N intermediates [12], which seriously hinders the catalytic activity and selectivity of NH3 [13] and poses a great challenge for developing efficient catalysts.
Currently, noble metals such as gold, ruthenium, and palladium are widely utilized for chemical detection and electrochemical NO3RR due to their excellent properties, but their scarcity and high cost prevent their large-scale practical application [14,15,16,17]. Therefore, cheap transition metals such as iron, nickel, and copper have been extensively researched owing to their low cost, abundancy, and unique electronic structure in order to provide active sites for NO3RR [18,19,20,21,22]. However, the low selectivity of NH3 and competitive reactions, especially hydrogen evolution, severely hinder the development of efficient transition metal catalysts [23,24,25]. To enhance NH3‘s selectivity and inhibit competitive reactions, new strategies need to be developed to replace the use of single transition metal catalysts. In recent years, bimetallic electrocatalytic systems have been widely reported due to their synergistic effects, which can effectively improve ammonia yield and selectivity. For example, CuCo hybrid oxides can achieve a high yield and Faraday efficiency (FE) of NH3 at lower potentials by taking advantage of synergistic effects and internal spin states [26]. Bimetallic FeNi electrocatalysts contribute to the transfer of nitrite and allow effective relay catalysis, thus improving selectivity during the reduction of nitrite to NH3 [27]. Despite already possessing the above advantages, the stability of bimetallic catalysts for NO3RR still needs to be further improved.
In recent years, nD (n = 0, 1, 2, 3) nanomaterials formed by composite metals and multidimensional materials have gradually begun to be studied owing to their unique structural properties and enhanced performance in applications such as catalysis, energy storage, and sensing [28]. A hydrogel is a polymer material with a three-dimensional mesoporous structure and great hydrophilicity. Owning to their good electrical conductivity and structural stability, hydrogels have been widely explored in the field of electrocatalysis in recent years, including for efficient water oxidation, the preparation of high-capacitance supercapacitor electrodes, and the reduction of nitrate to ammonia [29,30,31]. For example, phytic acid (PA)-based conductive hydrogels and graphene/polyaniline composite hydrogels for efficient water oxidation and high-capacitance supercapacitor electrodes have high specific capacitance and excellent cycling stability, enabling fast charge and mass transfer [32,33]. In addition, nitrogen-coordinated Fe prepared using the polymer-hydrogel method improved NH3 yield and FE by making the atoms on the carbon uniformly dispersed [34]. Although hydrogels have been used in electrocatalysis in recent years, more research input is still required for NO3RR to achieve high NH3 yields and high selectivity by employing the structural and electronic advantages of hydrogels.
To achieve the effective conversion of nitrate to ammonia, in this work, we prepared an iron–nickel-loaded sodium dodecyl sulfate-pyrrole (PPy) hydrogel from an electrocatalytic perspective. The resulting Fe/Ni-PPy allowed high conversion of NO3 (93.59%) and selectivity of NH3 (86.6%) at −0.9 V vs. a reversible hydrogen electrode (RHE). Importantly, the yield of ammonia reached 0.166 mmol h−1 cm−2, approximately three times higher than that of the Ni-Fe catalyst, with an FE of 88.9%, indicating the high NO3RR activity and selectivity of this catalyst. Electrochemical active surface area (ECSA) and electrochemical impedance spectroscopy (EIS) tests indicated that the introduction of the hydrogel into the bimetallic Fe/Ni system effectively exposed more active sites and enhanced the charge transfer to improve catalytic activity. This study provides a new hydrogel-mediated strategy for the construction of bimetallic electrocatalysts for efficient ammonia production.

2. Results and Discussion

The structure of catalysts was observed using scanning electron microscopy (SEM) and transmission electron microscopy (TEM). As shown in the SEM image in Figure 1a, the hydrogel is successfully attached to the carbon paper and the porous structure provides a larger active area for the catalyst. The TEM image of Fe/Ni-PPy (Figure S1) shows the presence of PPy loaded with FeNi nanoparticles. Meanwhile, the high-resolution TEM (HRTEM) image of Fe/Ni-PPy shows clear lattice stripes, indicating good lattice structures of the FeNi catalyst nanoparticles (Figure 1b) [35,36]. These results show that the Fe/Ni-PPy catalyst can expose more active sites, which is crucial to improve the property of the catalyst. Moreover, the corresponding element mapping images in Figure 1c–g demonstrate the presence of C, N, Ni, and Fe elements in the catalyst with a uniform distribution, further demonstrating that Fe/Ni-PPy has been successfully synthesized and uniformly attached to the carbon paper substrate [37].
The crystal structure of the materials was further investigated by powder X-ray diffraction (XRD). As shown in Figure 2a, the characteristic peaks of the PPy hydrogel material are observed at 22.8° [38], and the diffraction peaks at 44.5, 51.8, and 76.4° are ascribed to the (111), (200), and (220) planes of Ni (PDF#04-0850) of Fe/Ni-PPy, respectively. The Fe/Ni-PPy also exhibits two peaks at 44.7 and 65.0°, which can be ascribed to the (110) and (200) planes of crystalline Fe (PDF#06-0696) [39]. The above results demonstrate the successful preparation of Fe/Ni-PPy [40]. Using the Debye–Scherrer formula (Formula (S1)), the average grain size of the FeNi nanoparticles in the composite was estimated to be 17.43 nm [41]. Figure 2b shows the Fourier Transform Infrared (FTIR) spectra of bare PPy and Fe/Ni-PPy, in which the absorption peaks at 1538, 1720, and 3384 cm−1 correspond to the five-membered C-N heterocyclic vibration, N-H stretching vibration of the pyrrole ring, and the broad band of the hydroxyl group, respectively. Compared to bare PPy, Fe/Ni-PPy exhibits an o-disubstituted C-H vibration at 790 cm−1 and a benzene ring C=C stretching vibration at 1479 cm−1 [42,43]. After the introduction of FeNi nanoparticles, the peak intensity of PPy hydrogel was significantly reduced, which was due to the coordination of Fe/Ni with the functional groups on the surface of the PPy hydrogel [44]. This also indicates the successful introduction of Fe and Ni in the hydrogel.
A comprehensive analysis of the valence state of the Fe/Ni-PPy catalyst was carried out through X-ray photoelectron spectroscopy (XPS). In Figure 2c, the binding energy of C 1s at 284.6, 286.2, and 288.3 eV belong to the characteristic peaks of C-C, C-N, and C=O structures, respectively [45,46]. The characteristic peaks at 398.7, 400.1, 401.2, and 402.7 eV of binding energy are associated with pyridinic N, pyrrolic N, graphitic N, and oxidized N, respectively (Figure 2d) [46,47]. Significantly, C and N from PPy form a variety of conjugation structures to improve the electrical conductivity of Fe/Ni-PPy, which thus enhances the intrinsic activity of the catalyst [48]. In the O 1s spectrum (Figure S2), the peaks at 531.4 and 530.2 eV correspond to the surface hydroxyl group and O22−/O species, respectively. These species facilitate the NO3 adsorption and proton transfer on the Fe/Ni-PPy surface, thereby enhancing the NO3 reduction performance [49,50,51]. Meanwhile, Figure 2e shows two peaks at 710.2 and 722.3 eV, which correspond to the Fe 2p3/2 and 2p1/2 of Fe2+, respectively. The two peaks at 713.9 and 724.3 eV are assigned to Fe 2p3/2 and 2p1/2 of Fe3+, respectively. Most importantly, the peaks at 703.4 and 720.5 eV illustrate the existence of Fe0, which indicates that the Fe0 was produced in the catalyst from the reduction reactions of Fe3+ and Fe2+ during the synthesis of the material [46,52]. The XPS spectrum of Ni 2p (Figure 2f) also clearly shows the characteristic peaks of Ni0 at 852.8 (Ni 2p3/2) and 870.4 eV (Ni 2p1/2), while the peaks at 853.8 and 871.9 eV are assigned to Ni 2p3/2 and 2p1/2 of Ni2+, respectively, with two satellite peaks observed at 860.1 and 877.9 eV [53]. The above results suggest that the Fe and Ni species present in the catalyst exist in a mixed valence state, derived from the reduction reaction during the synthesis of materials. And the mixed valence states of Ni can inhibit the oxidation of Fe0 and ensure the stability of the catalyst [35]. The synergistic effect of Fe and Ni enhances the adsorption of nitrates and disrupts the strong correlation associated with the collaborative hydrogenation and *NH separation processes, thereby enhancing the kinetics of ammonia production [54,55].
The electrochemical experiments were conducted in a Na2SO4 solution (0.5 M) with KNO3 (200 ppm) by using a three-electrode system. As shown in the linear sweep voltammetry (LSV) curves (Figure 3a), the Fe/Ni-PPy electrocatalyst displays an evidently lower onset potential than those of pure FeNi particles and hydrogel, indicating the excellent electrochemical properties of Fe/Ni-PPy. First, the NO3 conversion of different samples is explored in Figure 3b, in which the Fe/Ni-PPy exhibits a higher NO3 conversion (93.59%) at −0.9 V vs. RHE compared to PPy (70.51%) and FeNi (83.76%). Similarly, the NH3 yield of Fe/Ni-PPy was more than doubled compared with bimetallic FeNi particles (Figure 3c), while the FE of NH3 of Fe/Ni-PPy is also much higher than that of FeNi electrocatalysts. Moreover, the NO2 selectivity of Fe/Ni-PPy is only found in a quarter of FeNi electrocatalysts (Figure 3d), indicating that PPy plays a pluripotent role in promoting the adsorption of NO2 and its further conversion towards NH3. The results show that the addition of PPy provides better conditions for NO3RR, and the prepared Fe/Ni-PPy has better electrochemical properties than those of pure FeNi particles and PPy hydrogel electrocatalysts. This phenomenon can be attributed to the introduction of PPy, which promoted the adsorption of NO3 and NO2 and accelerated the conversion of intermediates to NH3 [56].
To investigate the optimal Fe:Ni ratio, PPy hydrogels with Fe:Ni ratios of 2:1, 1:1, 1:2, 1:3, 1:0, and 0:1 were synthesized for electrochemical tests. The current densities of different catalysts were compared by LSV curves, as shown in Figure 4a. At the same current density, the composite of FeNi with PPy can effectively reduce the onset potential of NO3RR. The lowest onset potential was observed at an Fe:Ni ratio of 1:1, indicating that the catalysts exhibited the highest electrochemical reaction rate under this condition. Notably, the introduction of Ni synergistically enhances the electrochemical performance with Fe, whereas an excessive amount of Ni shifts the reaction towards HER (Figure S3), thereby adversely affecting NO3RR [57,58]. The above observations suggest that Fe and Ni are most effective in synergizing NO3RR at a ratio of 1:1. At the same time, the NO3 conversion ratio of Fe/Ni-PPy (Fe:Ni = 1:1) reaches 93.59% at −0.9 V vs. RHE, which shows an increase over 10% compared to Fe-PPy or Ni-PPy (Figure 4b). The optimized Fe/Ni-PPy catalyst also achieves the highest NH3 selectivity of 86.6% and the lowest NO2 selectivity of 8.24% (Figure 4c), superior to those of pure Fe-PPy or Ni-PPy catalysts. The excellent catalyst performance is also reflected in the NH3 yield and FE (Figure 4d), which reaches 0.166 mmol h−1 cm−2 and 88.9%, respectively. Figure S4 shows a comprehensive and systematic comparison of Fe-PPy, Ni-PPy, PPy hydrogels, and Fe/Ni-PPy. Both the NO3 conversion and NH3 selectivity of Fe/Ni-PPy were higher than those of other samples, while the selectivity of NO2 was the lowest in all samples. These performance promotions can be ascribed to the fact that the introduction of PPy enables the uniform loading of the FeNi particles on its 3D scaffold and the fully exposed active sites.
To eliminate the interference of external factors and elements contained in the Fe/Ni-PPy, LSV tests were performed in the presence and absence of the NO3 electrolyte [59]. The experimental result shows that the current density in the presence of KNO3 is significantly higher than the absence of KNO3, indicating that NO3 is involved in the reaction (Figure 5a). To verify the source of N, a series of control experiments concerning NH3 yield were performed at the same potential with the conditions of open circuit, bare carbon cloth without the catalyst, and pure Na2SO4 electrolyte without NO3. The results show that the N source is exclusively from NO3 in the electrolyte, with no contribution from pollutions in the environment or the catalyst (Figure S5). Under different voltages, the NO3 conversion test (Figure S6) indicates a rapid increase in the conversion rate from 74.68% to 93.59% from −0.7 to −0.9 V vs. RHE, followed by a gradual decrease. The varying trend of the selectivity of NO2 and NH3 indicates that NO3 gradually transforms into NH3 at high voltage and the selectivity of NO2 decreases at −0.7 to −1.1 V vs. RHE (Figure S7). It is noteworthy that the selectivity of NH3 undergoes a slight decrease as a result of the hydrogel structure being destructively affected by the high potential [60]. In order to determine the source of nitrogen, K15NO3 and K14NO3 were used as nitrogen sources. The 1H nuclear magnetic resonance (NMR) spectra (Figure 5b) showed characteristic double peaks of 15NH4+ and triple peaks of 14NH4+. It was proved that the NH4+ product was from the electrocatalytic reduction of NO3 by Fe/Ni-PPy, and the measured NH3 concentration is reliable and not contaminated. The volcanic curve in Figure S8 shows that the yield of NH3 first rises and then falls from −0.7 to −1.1 V vs. RHE. However, competitive hydrogen evolution reactions at high potentials result in lower NH3 yield and FE [61]. The time-dependent concentration curves depict the variations in NO3, NO2, and NH3 throughout the NO3RR process (Figure 5c), in which the NH3 concentration continues to increase and the concentration of NO3 steadily decreases and NO2 increases within the first 30 min and then gradually decreases, indicating that the Fe/Ni-PPy possesses the ability to accelerate the conversion of NO2 to NH3. Additionally, the NH3 yield reached 0.166 mmol h−1 cm−2 and the FE was stable above 80% throughout six consecutive tests at −0.9 V vs. RHE (Figure 5d), suggesting the outstanding stability of Fe/Ni-PPy. Furthermore, the structure and morphology of Fe/Ni-PPy can be well reserved after the stability test, as evidenced by TEM (Figure S9). All measurements indicate that Fe/Ni-PPy possesses good structural stability. In conclusion, post-test analyses confirm that Fe/Ni-PPy exhibits adequate chemical stability for NO3RR, thereby establishing a solid basis for the resource utilization of NO3.
To explore the intrinsic reasons for the outstanding catalytic performance of the Fe/Ni-PPy electrocatalysts, the ECSA was calculated using the electrochemical double-layer capacitance (Cdl) method. The cyclic voltammetry (CV) tests (Figure 6a,b) were performed on Fe/Ni-PPy and bimetallic FeNi electrocatalysts at different scan rates within the potential range of 0.2–0.3 V vs. RHE. The results (Figure 6c) show that Fe/Ni-PPy reached 5.2 mF cm−2, while FeNi nanoparticles reached 3.7 mF cm−2, indicating a larger ECSA for Fe/Ni-PPy. To further evaluate the impact of PPy on the kinetics during the reaction process, EIS measurements of FeNi nanoparticles and Fe/Ni-PPy were carried out under the open circuit potential. As shown in Figure 6d, FeNi nanoparticles display a large impedance arc, which reveals the slow electrochemical behaviors of the sample. In contrast, Fe/Ni-PPy exhibits a much smaller resistance arc, indicating that the addition of the conductive PPy hydrogel can efficiently reduce the charge transfer resistance and enhance the charge transfer rate [62]. The above results illustrate that introducing PPy enhances the NO3 reduction ability and accelerates the charge transfer in the NO3RR process.
Therefore, based on the above experiments and characterizations, a plausible mechanism for the NO3RR on Fe/Ni-PPy is proposed in Figure 6e. With Fe/Ni-PPy and the applied bias, NO3 is transformed into *NO3, which undergoes hydrogenation with *H atoms from water to form NO2. Eventually, NO2 undergoes further sequential hydrogenation to produce NH3. The synergy of Fe and Ni enables low energy barriers for NH3 production, while the combination of the conductive hydrogel with the bimetallic FeNi nanoparticles can effectively expose more active sites and enhance the charge collection to improve the catalytic activity. This study provides a useful hydrogel-mediated strategy for the construction of composite electrocatalysts for efficient ammonia production.

3. Experimental Section

3.1. Chemicals and Materials

Sodium dodecyl sulfate (SDS, C12H25SO4Na), pyrrole (C4H5N), potassium nitrate (KNO3), sodium nitrite (NaNO2), ammonium persulfate ((NH4)2S2O8), ferric chloride hexahydrate (FeCl3·6H2O), nickel chloride hexahydrate (NiCl2·6H2O), sodium borohydride (NaBH4), sodium sulfate (Na2SO4), hydrochloric acid (HCl), salicylic acid (C7H6O3), sodium citrate dehydrate (C6H5Na3O7), sodium hydroxide (NaOH), sodium hypochlorite (NaClO), sulfamic acid (H3NO3S), sulfanilamide (C6H8N2O2S), N-(1-naphthyl) ethyldiamine dihydrochloride (C12H14N2·2HCl), and phosphoric acid (H3PO4) were purchased from Sinopharm Group Chemical Reagent Co., Ltd., Shanghai, China. Ethanol (C2H5OH) and sodium nitroprusside (C5FeN6Na2O·2H2O) were purchased from Shanghai Aladdin Bio-Chem Technology Co., Ltd., Shanghai, China. The Nafion film was supplied by the Du Pont China Holding Co., Ltd., Shanghai, China.

3.2. Synthesis of the Catalysts

Pyrrole (416 μL) and SDS (0.58 g) were added to 10 mL of deionized water and stirred for 40 min until a homogeneous solution was obtained (solution A). Meanwhile, 1.5 g of (NH4)2S2O8 was added into 10 mL of deionized water and sonicated until dissolved (solution B). The two solutions were sonicated and then mixed to form a black solution, which was left to stand for 1 h. After the polymerization reaction, the unreacted monomer was washed off with deionized water to obtain the PPy hydrogel.
For Fe/Ni-PPy, FeCl3 (0.05 M) and NiCl2 (0.05 M) were dissolved in deionized water (40 mL), sonicated for 5 min, and poured into the prepared hydrogel to form a mixed solution, which was stirred for 10 h to allow the adsorption of the metal ions. At the end of the process, Fe/Ni-PPy with an Fe:Ni ratio of 1:1 was obtained by adding the NaBH4 solution dropwise under nitrogen atmosphere to reduce the metal salts to FeNi nanoparticles on PPy. It was washed three times with deionized water and ethanol, and then dried in a vacuum oven at 40 °C for 12 h, and then ground into a powder and stored at room temperature. The Fe/Ni-PPy samples with different Fe/Ni ratios (1:3, 1:2, 2:1, 1:0, and 0:1) were also prepared using this method.
Pure FeNi nanoparticles were also prepared by using the same procedure but without PPy.
For the preparation of the working electrode, rectangular sheets of carbon paper (0.5 cm × 2 cm) were pretreated with 0.5 M sulfuric acid for 4 h, and then rinsed with deionized water. The Fe/Ni-PPy powder (10 mg) was mixed ultrasonically with the prepared solution (980 µL of ethanol and 20 µL of 5% Nafion solution). Then, 20 µL of the mixture was dropped on both sides of the carbon paper and dried naturally for further use.

3.3. Material Characterizations

The morphology of the samples was analyzed through a scanning electron microscope (SEM, JSM-7800F, JEOL, Tokyo, Japan) at 20 kV voltage. Transmission electron microscopy (TEM), high-resolution transmission electron microscopy (HRTEM) and energy dispersive X-ray spectroscopy (EDS) were tested using a Tecnai G2 F30 S-Twin microscope (FEI, Hillsboro, OR, USA). With a D/MAX-2500 diffractometer (Rigaku, Tokyo, Japan), X-ray diffraction (XRD) patterns were obtained utilizing Cu Kα radiation (λ = 1.5406 Å) as the source, with a 2θ angle range of 5–80°, operated at 40 kV and 30 mA. Fourier Transform Infrared (FTIR) spectroscopy was tested by Nexus 470 (Nicolet, Waltham, MA, USA). The chemical valence and composition of the catalysts were determined by X-ray photoelectron spectroscopy (XPS, Thermo Scientific ESCALAB 250X, Waltham, MA, USA), and the absorbance of the products was quantified by an ultraviolet visible spectrophotometer (Shimadzu UV2600, Kyoto, Japan). The isotopic labeling experiments were measured by 1H nuclear magnetic resonance (NMR) analysis (AC-P400, BRUKER, Baden-Württemberg, Germany).

3.4. Electrocatalytic Measurements

All electrochemical experiments were conducted utilizing a CHI 760E electrochemical workstation (Chenhua, Shanghai, China). The electrocatalytic reaction system consists of a three-electrode structure separated by a Nafion 117 membrane and an H-type electrolytic cell. The Fe/Ni-PPy loaded on carbon paper, Ag/AgCl, and graphite rod were used as the working, reference, and counter electrodes, respectively. As an electrolyte, 70 mL of 0.5 M Na2SO4 solution was added to the anode chamber, and 70 mL of 0.5 M Na2SO4 containing 200 ppm KNO3 was introduced to the cathode chamber, serving as the nitrogen source. The primary electrochemical evaluations encompass linear sweep voltammetry (LSV), electrochemical impedance spectroscopy (EIS), and cyclic voltammetry (CV). The EIS measurements were conducted in the frequency range of 0.1  to 1 MHz at the open circuit potential [63]. Argon was introduced before the experiment to eliminate other gas interference. The reaction was performed at a stirring rate of 350 rpm for 1 h at different potentials. The performance calculation formulae are shown in Equations (1)–(5).
NO 3   conversion   ( % ) = ( c N O 3 / c 0 ) × 100 %
NO 2   selectivity   ( % ) = ( c N O 2 / 46 ) / ( Δ c N O 3 / 62 ) × 100 %
NH 3   selectivity   ( % ) = ( c N H 3 / 17 ) / ( Δ c N O 3 / 62 ) × 100 %
Yield   of   NH 3   ( mmol   h 1   cm 2 ) = ( c N H 3 × V ) / ( M N H 3 × S × t )
Faradaic   efficiency   ( % )   =   8 × F × c N H 3 × V / ( M N H 3 × Q )
ΔcNO3(mg L−1) is the concentration difference before and after the NO3RR reaction and c0 is the initial concentration of nitrate (mg L−1). The cNO2 and cNH3 are the measured concentrations of NO2 and NH3, respectively. V is the volume of the electrolyte in the electrolytic cell (70 mL). S is the electrode area (1 cm2). t is the electrolysis time (1 h). F is the Faraday constant (96,485 C mol−1). Q is the total charge through the electrode (C). MNH3 is the relative molecular mass of NH3 [64].
NO3, NO2, and NH3 were determined by different chromogenic reagents. The chromogenic reagents of NO3 are HCl and H3NO3S. The chromogenic reagents of NO2 are a mixed solution of C12H14N2·2HCl, C6H8N2O2S, and H3PO4 [65]. The concentration of NH3 was determined by the indophenol blue method [66]. Supporting Information shows the specific details of all the methods. Finally, the standard curves were obtained by drawing the relationship between the absorbance and the measured objects (Figures S10–S12).

4. Conclusions

In this paper, a series of Fe/Ni-PPy electrocatalysts with varying Fe/Ni ratios were constructed by using the PPy hydrogel as a scaffold for the in situ adsorption and formation of FeNi nanoparticles, which shows promoted NH3 yield and FE in NO3RR. The results show that, with optimized Fe:Ni = 1:1, the Fe/Ni-PPy catalyst has an NH3 yield of 0.166 mmol h−1 cm−2 and a FE of 88.9% at −0.9 V vs. RHE. Moreover, the ECSA and EIS indicate that the combination of FeNi bimetallic nanoparticles and PPy can effectively expose more active sites and enhance the charge transfer to improve the catalytic activity and selectivity. The mesoporous hydrophilic structure of the PPy hydrogel enables Fe and Ni to be uniformly dispersed and provides a favorable environment for bimetallic synergism, which enhances the ability of the catalysts to convert NO3 to NH3. By further analyzing the N-species variation during electrolysis, it was found that Fe/Ni-PPy accelerates the rate-limiting step of NO3 to NO2 during NO3RR, owing to the improved catalyst particle loading and substrate adsorption with the introduction of the hydrogel. This study provides an interesting inspiration for the development of hydrogel-based electrocatalysts for NO3RR.

Supplementary Materials

The following supporting information can be downloaded at https://www.mdpi.com/article/10.3390/molecules30061271/s1, Figure S1: TEM image of Fe/Ni-PPy (Fe:Ni = 1:1); Figure S2: XPS spectrum of O 1s; Figure S3: LSV curves of different catalysts with 0.5 M Na2SO4; Figure S4: The performance comparison of different catalysts at −0.9 V vs. RHE for 1 h; Figure S5: NH3 yield in different conditions; Figure S6: The conversion of NO3 by the Fe/Ni-PPy (Fe:Ni=1:1) at different potentials; Figure S7: The selectivity of NH3 and NO2 by the Fe/Ni-PPy (Fe:Ni=1:1) catalyst at different potentials; Figure S8: The NH3 yield and FE by the Fe/Ni-PPy (Fe:Ni=1:1) catalyst at different potentials; Figure S9: TEM image of Fe/Ni-PPy (Fe:Ni=1:1) after 6 continuous cycle tests; Figure S10: Standard curve of NO3; Figure S11: Standard curve of NO2; Figure S12: Standard curve of NH3.

Author Contributions

L.L.: Conceptualization, Methodology, Writing—original draft, Funding acquisition. P.Y.: Writing—original draft, Writing—review and editing. Q.G.: Methodology, Investigation. D.Z.: Methodology, Data curation. C.M.: Supervision, Funding acquisition. Q.Y.: Data curation. H.S.: Data curation. M.L.: Data curation. Y.L.: Supervision, Writing—review and editing, Funding acquisition. B.M.: Writing—review and editing, Supervision, Project administration, Funding acquisition. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by the National Natural Science Foundation of China (22278194 and 21908081), Jiangsu Collaborative Innovation Center of Technology and Material of Water Treatment at Suzhou University of Science and Technology, and the Natural Science Foundation of Hebei Province (E2022203066).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Data will be made available on request.

Conflicts of Interest

The authors declare that there are no commercial or related competitive interests that conflict with the submitted work.

References

  1. Yüzbaşıoğlu, A.E.; Avşar, C.; Gezerman, A.O. The current situation in the use of ammonia as a sustainable energy source and its industrial potential. Curr. Res. Green Sustain. Chem. 2022, 5, 100307. [Google Scholar] [CrossRef]
  2. Xu, Y.; Ma, Y.; Cayuela, M.L.; Sánchez-Monedero, M.A.; Wang, Q. Compost biochemical quality mediates nitrogen leaching loss in a greenhouse soil under vegetable cultivation. Geoderma 2020, 358, 113984. [Google Scholar] [CrossRef]
  3. Xu, Z.; Zhong, S.; Yu, Y.; Wang, Y.; Cheng, H.; Du, D.; Wang, C. Rhus typhina L. triggered greater allelopathic effects than Koelreuteria paniculata Laxm under ammonium fertilization. Sci. Hortic. 2023, 309, 111703. [Google Scholar] [CrossRef]
  4. Ren, Y.; Li, S.; Yu, C.; Zheng, Y.; Wang, C.; Qian, B.; Wang, L.; Fang, W.; Sun, Y.; Qiu, J. NH3 Electrosynthesis from N2 Molecules: Progresses, Challenges, and Future Perspectives. J. Am. Chem. Soc. 2024, 146, 6409–6421. [Google Scholar] [CrossRef]
  5. Wei, W.; Hu, X.; Yang, S.; Wang, K.; Zeng, C.; Hou, Z.; Cui, H.; Liu, S.; Zhu, L. Denitrifying halophilic archaea derived from salt dominate the degradation of nitrite in salted radish during pickling. Food Res. Int. 2022, 152, 110906. [Google Scholar] [CrossRef]
  6. Guo, X.; Du, H.; Qu, F.; Li, J. Recent progress in electrocatalytic nitrogen reduction. J. Mater. Chem. A 2019, 7, 3531–3543. [Google Scholar] [CrossRef]
  7. Dai, H.; Liu, L.; Zhao, H.; Zhou, P.; Ying, Y.; Yin, M.; Wang, X.; Fan, W.; Bai, H. Efficient electrocatalytic reduction of nitrate to ammonia using Cu–CeO2 solid solution. Int. J. Hydrogen Energy 2024, 73, 257–264. [Google Scholar] [CrossRef]
  8. Estudillo-Wong, L.A.; Santillán-Díaz, G.; Arce-Estrada, E.M.; Alonso-Vante, N.; Manzo-Robledo, A. Electroreduction of NOxz species in alkaline medium on Pt nanoparticles. Electrochim. Acta 2013, 88, 358–364. [Google Scholar] [CrossRef]
  9. Yao, J.; Yan, J. Efficient nitrate-to-ammonia transformation through a direct eight-electron reduction. Sci. China Chem. 2020, 63, 1737–1739. [Google Scholar] [CrossRef]
  10. Tang, C.; Qiao, S.-Z. How to explore ambient electrocatalytic nitrogen reduction reliably and insightfully. Chem. Soc. Rev. 2019, 48, 3166–3180. [Google Scholar] [CrossRef]
  11. Cui, X.; Tang, C.; Zhang, Q. A Review of Electrocatalytic Reduction of Dinitrogen to Ammonia under Ambient Conditions. Adv. Energy Mater. 2018, 8, 1800369. [Google Scholar] [CrossRef]
  12. Wang, Y.; Yu, Y.; Jia, R.; Zhang, C.; Zhang, B. Electrochemical synthesis of nitric acid from air and ammonia through waste utilization. Natl. Sci. Rev. 2019, 6, 730–738. [Google Scholar] [CrossRef]
  13. Zeng, Y.; Priest, C.; Wang, G.; Wu, G. Restoring the Nitrogen Cycle by Electrochemical Reduction of Nitrate: Progress and Prospects. Small Methods 2020, 4, 2000672. [Google Scholar] [CrossRef]
  14. Peng, W.; Luo, M.; Xu, X.; Jiang, K.; Peng, M.; Chen, D.; Chan, T.-S.; Tan, Y. Spontaneous Atomic Ruthenium Doping in Mo2CTX MXene Defects Enhances Electrocatalytic Activity for the Nitrogen Reduction Reaction. Adv. Energy Mater. 2020, 10, 2001364. [Google Scholar] [CrossRef]
  15. Sun, L.; Liu, B. Mesoporous PdN Alloy Nanocubes for Efficient Electrochemical Nitrate Reduction to Ammonia. Adv. Mater. 2023, 35, 2207305. [Google Scholar] [CrossRef]
  16. Han, E.; Li, L.; Gao, T.; Pan, Y.; Cai, J. Nitrite determination in food using electrochemical sensor based on self-assembled MWCNTs/AuNPs/poly-melamine nanocomposite. Food Chem. 2024, 437, 137773. [Google Scholar] [CrossRef]
  17. Wu, H.; Xie, R.; Hao, Y.; Pang, J.; Gao, H.; Qu, F.; Tian, M.; Guo, C.; Mao, B.; Chai, F. Portable smartphone-integrated AuAg nanoclusters electrospun membranes for multivariate fluorescent sensing of Hg2+, Cu2+ and l-histidine in water and food samples. Food Chem. 2023, 418, 135961. [Google Scholar] [CrossRef]
  18. Gibert, O.; Abenza, M.; Reig, M.; Vecino, X.; Sánchez, D.; Arnaldos, M.; Cortina, J.L. Removal of nitrate from groundwater by nano-scale zero-valent iron injection pulses in continuous-flow packed soil columns. Sci. Total Environ. 2022, 810, 152300. [Google Scholar] [CrossRef]
  19. Wang, X.; Pan, Y.; Wang, X.; Guo, Y.; Ni, C.; Wu, J.; Hao, C. High performance hybrid supercapacitors assembled with multi-cavity nickel cobalt sulfide hollow microspheres as cathode and porous typha-derived carbon as anode. Ind. Crops Prod. 2022, 189, 115863. [Google Scholar] [CrossRef]
  20. Iarchuk, A.; Dutta, A.; Broekmann, P. Novel Ni foam catalysts for sustainable nitrate to ammonia electroreduction. J. Hazard. Mater. 2022, 439, 129504. [Google Scholar] [CrossRef]
  21. 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] [PubMed]
  22. Wu, Z.; Song, Y.; Guo, H.; Xie, F.; Cong, Y.; Kuang, M.; Yang, J. Tandem catalysis in electrocatalytic nitrate reduction: Unlocking efficiency and mechanism. Interdiscip. Mater. 2024, 3, 245–269. [Google Scholar] [CrossRef]
  23. 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]
  24. 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]
  25. 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]
  26. Sun, S.; Dai, C.; Zhao, P.; Xi, S.; Ren, Y.; Tan, H.R.; Lim, P.C.; Lin, M.; Diao, C.; Zhang, D.; et al. Spin-related Cu-Co pair to increase electrochemical ammonia generation on high-entropy oxides. Nat. Commun. 2024, 15, 260. [Google Scholar] [CrossRef]
  27. Ma, X.; Zhong, J.; Huang, W.; Wang, R.; Li, S.; Zhou, Z.; Li, C. Tuning the d-band centers of bimetallic FeNi catalysts derived from layered double hydroxides for selective electrocatalytic reduction of nitrates. Chem. Eng. J. 2023, 474, 145721. [Google Scholar] [CrossRef]
  28. Estudillo-Wong, L.A.; Guerrero-Barajas, C.; Vázquez-Arenas, J.; Alonso-Vante, N. Revisiting Current Trends in Electrode Assembly and Characterization Methodologies for Biofilm Applications. Surfaces 2023, 6, 2–28. [Google Scholar] [CrossRef]
  29. Gao, D.; Fabiano, S. Conductive hydrogels put electrons in charge. Science 2024, 384, 509–510. [Google Scholar] [CrossRef]
  30. Li, P.; Sun, W.; Li, J.; Chen, J.-P.; Wang, X.; Mei, Z.; Jin, G.; Lei, Y.; Xin, R.; Yang, M.; et al. N-type semiconducting hydrogel. Science 2024, 384, 557–563. [Google Scholar] [CrossRef]
  31. Zhang, H.; Gan, X.; Yan, Y.; Zhou, J. A Sustainable Dual Cross-Linked Cellulose Hydrogel Electrolyte for High-Performance Zinc-Metal Batteries. Nano Micro Lett. 2024, 16, 106. [Google Scholar] [CrossRef] [PubMed]
  32. Hu, Q.; Li, G.; Liu, X.; Zhu, B.; Chai, X.; Zhang, Q.; Liu, J.; He, C. Superhydrophilic Phytic-Acid-Doped Conductive Hydrogels as Metal-Free and Binder-Free Electrocatalysts for Efficient Water Oxidation. Angew. Chem. Int. Ed. 2019, 58, 4318–4322. [Google Scholar] [CrossRef] [PubMed]
  33. Ji, J.; Li, R.; Li, H.; Shu, Y.; Li, Y.; Qiu, S.; He, C.; Yang, Y. Phytic acid assisted fabrication of graphene/polyaniline composite hydrogels for high-capacitance supercapacitors. Compos. Part B 2018, 155, 132–137. [Google Scholar] [CrossRef]
  34. Li, P.; Jin, Z.; Fang, Z.; Yu, G. A single-site iron catalyst with preoccupied active centers that achieves selective ammonia electrosynthesis from nitrate. Energy Environ. Sci. 2021, 14, 3522–3531. [Google Scholar] [CrossRef]
  35. Shan, A.; Idrees, A.; Zaman, W.Q.; Mohsin, A.; Abbas, Z.; Stadler, F.J.; Lyu, S. Synthesis of CaCO3 supported nano zero-valent iron-nickel nanocomposite (nZVI-Ni@CaCO3) and its application for trichloroethylene removal in persulfate activated system. Environ. Res. 2024, 245, 118050. [Google Scholar] [CrossRef]
  36. Muthusamy, T.; Sethuram Markandaraj, S.; Shanmugam, S. Nickel nanoparticles wrapped in N-doped carbon nanostructures for efficient electrochemical reduction of NO to NH3. J. Mater. Chem. A 2022, 10, 6470–6474. [Google Scholar] [CrossRef]
  37. Wang, C.; Yang, H.; Zhang, Y.; Wang, Q. NiFe Alloy Nanoparticles with hcp Crystal Structure Stimulate Superior Oxygen Evolution Reaction Electrocatalytic Activity. Angew. Chem. Int. Ed. 2019, 58, 6099–6103. [Google Scholar] [CrossRef]
  38. Zheng, H.; Chen, M.; Sun, Y.; Zuo, B. Self-Healing, Wet-Adhesion silk fibroin conductive hydrogel as a wearable strain sensor for underwater applications. Chem. Eng. J. 2022, 446, 136931. [Google Scholar] [CrossRef]
  39. Yang, Y.; Xu, Y.; Zhong, D.; Qiao, Q.; Zeng, H. Efficient removal of Cr(VI) by chitosan cross-linked bentonite loaded nano-zero-valent iron composite: Performance and mechanism. J. Hazard. Mater. 2024, 480, 136183. [Google Scholar] [CrossRef]
  40. Vigneshwaran, J.; Jose, J.; Thomas, S.; Gagliardi, A.; Narayan, R.L.; Jose, S.P. PPy-PdO modified MXene for flexible binder-free electrodes for asymmetric supercapacitors: Insights from experimental and DFT investigations. Chem. Eng. J. 2024, 487, 150555. [Google Scholar] [CrossRef]
  41. Khalil, K.D.; Riyadh, S.M.; Alkayal, N.S.; Bashal, A.H.; Alharbi, K.H.; Alharbi, W. Chitosan-Strontium Oxide Nanocomposite: Preparation, Characterization, and Catalytic Potency in Thiadiazoles Synthesis. Polymers 2022, 14, 2827. [Google Scholar] [CrossRef] [PubMed]
  42. Chen, S.; Guo, B.; Yu, J.; Yan, Z.; Liu, R.; Yu, C.; Zhao, Z.; Zhang, H.; Yao, F.; Li, J. A polypyrrole-dopamine/poly(vinyl alcohol) anisotropic hydrogel for strain sensor and bioelectrodes. Chem. Eng. J. 2024, 486, 150182. [Google Scholar] [CrossRef]
  43. Guan, D.; Xu, H.; Huang, Y.-C.; Jing, C.; Tsujimoto, Y.; Xu, X.; Lin, Z.; Tang, J.; Wang, Z.; Sun, X.; et al. Operando Studies Redirect Spatiotemporal Restructuration of Model Coordinated Oxides in Electrochemical Oxidation. Adv. Mater. 2024, 37, 2413073. [Google Scholar] [CrossRef]
  44. Wang, J.; Du, P.; Hsu, Y.-I.; Uyama, H. Smart versatile hydrogels tailored by metal-phenolic coordinating carbon and polypyrrole for soft actuation, strain sensing and writing recognition. Chem. Eng. J. 2024, 493, 152671. [Google Scholar] [CrossRef]
  45. Zhang, Y.; Chen, L.; Yan, B.; Zhang, F.; Shi, Y.; Guo, X. Single Cu atoms confined in N-doped porous carbon networks by flash nanocomplexation as efficient trifunctional electrocatalysts for Zn-air batteries and water splitting. Compos. Part B 2023, 253, 110575. [Google Scholar] [CrossRef]
  46. Wang, Z.; Ang, J.; Liu, J.; Ma, X.Y.D.; Kong, J.; Zhang, Y.; Yan, T.; Lu, X. FeNi alloys encapsulated in N-doped CNTs-tangled porous carbon fibers as highly efficient and durable bifunctional oxygen electrocatalyst for rechargeable zinc-air battery. Appl. Catal. B 2020, 263, 118344. [Google Scholar] [CrossRef]
  47. Guo, Y.; Yao, S.; Gao, L.; Chen, A.; Jiao, M.; Cui, H.; Zhou, Z. Boosting bifunctional electrocatalytic activity in S and N co-doped carbon nanosheets for high-efficiency Zn–air batteries. J. Mater. Chem. A 2020, 8, 4386–4395. [Google Scholar] [CrossRef]
  48. Tang, S.; Yu, C.; Liu, X.; Fu, D.; Liao, W.; Xu, F.; Zhong, W. 2-Methylimidazole assisted synthesis of nanocrystalline shell reinforced PPy hydrogel with high mechanical and electrochemical performance. Chem. Eng. J. 2022, 430, 133033. [Google Scholar] [CrossRef]
  49. Cai, J.; Huang, J.; Cao, A.; Wei, Y.; Wang, H.; Li, X.; Jiang, Z.; Waterhouse, G.I.N.; Lu, S.; Zang, S.-Q. Interfacial hydrogen bonding-involved electrocatalytic ammonia synthesis on OH-terminated MXene. Appl. Catal. B 2023, 328, 122473. [Google Scholar] [CrossRef]
  50. Abdelghafar, F.; Xu, X.; Guan, D.; Lin, Z.; Hu, Z.; Ni, M.; Huang, H.; Bhatelia, T.; Jiang, S.P.; Shao, Z. New Nanocomposites Derived from Cation-Nonstoichiometric Bax(Co, Fe, Zr, Y)O3−δ as Efficient Electrocatalysts for Water Oxidation in Alkaline Solution. ACS Mater. Lett. 2024, 6, 2985–2994. [Google Scholar] [CrossRef]
  51. Gong, Z.; Zhong, W.; He, Z.; Liu, Q.; Chen, H.; Zhou, D.; Zhang, N.; Kang, X.; Chen, Y. Regulating surface oxygen species on copper (I) oxides via plasma treatment for effective reduction of nitrate to ammonia. Appl. Catal. B 2022, 305, 121021. [Google Scholar] [CrossRef]
  52. Huang, C.-C.; Pourzolfaghar, H.; Huang, C.-L.; Liao, C.-P.; Li, Y.-Y. FeNi nanoalloy-carbon nanotubes on defected graphene as an excellent electrocatalyst for lithium-oxygen batteries. Carbon 2024, 222, 118973. [Google Scholar] [CrossRef]
  53. Zhang, C.; Gao, Y.; Zhang, J.; Jiao, Y.; Zhu, Q.; Wang, J.; Chen, Y.; Li, X. Regulating Ni microchemical state for enhanced C-C/H bonds adsorption and activation towards methylcyclohexane steam reforming. Chem. Eng. J. 2025, 506, 160152. [Google Scholar] [CrossRef]
  54. Ma, L.; Xu, F.; Zhang, L.; Nie, Z.; Xia, K.; Guo, M.; Li, M.; Ding, X. Breaking the linear correlations for enhanced electrochemical nitrogen reduction by carbon-encapsulated mixed-valence Fe7(PO4)6. J. Energy Chem. 2022, 71, 182–187. [Google Scholar] [CrossRef]
  55. Qiao, H.; Han, Y.; Yao, L.; Xu, X.; Ma, J.; Wen, B.; Hu, J.; Huang, H. Coating zero valent iron onto hollow carbon spheres as efficient electrocatalyst for N2 fixation and neutral Zn-N2 battery. Chem. Eng. J. 2023, 464, 142628. [Google Scholar] [CrossRef]
  56. Chen, H.; Zhang, Z.; Li, Y.; Yu, L.; Chen, F.; Li, L. Conductive polymer protection strategy to promote electrochemical nitrate reduction to ammonia in highly acidic condition over Cu-based catalyst. Chem. Eng. J. 2024, 481, 148596. [Google Scholar] [CrossRef]
  57. Luo, H.; Li, S.; Wu, Z.; Liu, Y.; Luo, W.; Li, W.; Zhang, D.; Chen, J.; Yang, J. Modulating the Active Hydrogen Adsorption on Fe–N Interface for Boosted Electrocatalytic Nitrate Reduction with Ultra-Long Stability. Adv. Mater. 2023, 35, 2304695. [Google Scholar] [CrossRef]
  58. Xiong, Y.; Sun, M.; Wang, S.; Wang, Y.; Zhou, J.; Hao, F.; Liu, F.; Yan, Y.; Meng, X.; Guo, L.; et al. Atomic Scale Cooperativity of Alloy Nanostructures for Efficient Nitrate Electroreduction to Ammonia in Neutral Media. Adv. Funct. Mater. 2024, 2420153. [Google Scholar] [CrossRef]
  59. Kou, M.; Yuan, Y.; Zhao, R.; Wang, Y.; Zhao, J.; Yuan, Q.; Zhao, J. Insights into the Origin of Activity Enhancement via Tuning Electronic Structure of Cu2O towards Electrocatalytic Ammonia Synthesis. Molecules 2024, 29, 2261. [Google Scholar] [CrossRef]
  60. Deng, K.; Lian, Z.; Wang, W.; Yu, J.; Mao, Q.; Yu, H.; Wang, Z.; Wang, L.; Wang, H. Hydrogen spillover effect tuning the rate-determining step of hydrogen evolution over Pd/Ir hetero-metallene for industry-level current density. Appl. Catal. B Environ. Energy 2024, 352, 124047. [Google Scholar] [CrossRef]
  61. Niu, Z.; Fan, S.; Li, X.; Wang, P.; Liu, Z.; Wang, J.; Bai, C.; Zhang, D. Bifunctional copper-cobalt spinel electrocatalysts for efficient tandem-like nitrate reduction to ammonia. Chem. Eng. J. 2022, 450, 138343. [Google Scholar] [CrossRef]
  62. Mondal, S.; Dilly Rajan, K.; Patra, L.; Rathinam, M.; Ganesh, V. Sulfur Vacancy-Induced Enhancement of Piezocatalytic H2 Production in MoS2. Small 2025, 2411828. [Google Scholar] [CrossRef] [PubMed]
  63. Zhang, X.; Kohler, H.; Schwotzer, M.; Guth, U. Stability improvement of layered Au,Pt-YSZ mixed-potential gas sensing electrodes by cathodic polarization: Studies by steady state and dynamic electrochemical methods. Sens. Actuators B 2021, 342, 130065. [Google Scholar] [CrossRef]
  64. Zhang, D.; Liu, Y.; Li, D.; Jiang, T.; Chen, Q.; Mao, C.; Li, L.; Jiang, D.; Mao, B. Carbon dots-boosted active hydrogen for efficient electrocatalytic reduction of nitrate to ammonia. J. Alloys Compd. 2025, 1014, 178694. [Google Scholar] [CrossRef]
  65. Zhang, K.; Sun, P.; Huang, Y.; Tang, M.; Zou, X.; Pan, Z.; Huo, X.; Wu, J.; Lin, C.; Sun, Z.; et al. Electrochemical Nitrate Reduction to Ammonia on CuCo Nanowires at Practical Level. Adv. Funct. Mater. 2024, 34, 2405179. [Google Scholar] [CrossRef]
  66. Huang, L.; Cheng, L.; Ma, T.; Zhang, J.-J.; Wu, H.; Su, J.; Song, Y.; Zhu, H.; Liu, Q.; Zhu, M.; et al. Direct Synthesis of Ammonia from Nitrate on Amorphous Graphene with Near 100% Efficiency. Adv. Mater. 2023, 35, 2211856. [Google Scholar] [CrossRef]
Figure 1. (a) SEM, (b) HRTEM, (c) HAADF, and (dg) EDS mapping images of Fe/Ni−PPy.
Figure 1. (a) SEM, (b) HRTEM, (c) HAADF, and (dg) EDS mapping images of Fe/Ni−PPy.
Molecules 30 01271 g001
Figure 2. (a) XRD patterns (the shadows are the characteristic peaks of PPy) and (b) FTIR spectra of PPy and Fe/Ni−PPy. The high-resolution XPS curves of (c) C 1s, (d) N 1s, (e) Fe 2p, and (f) Ni 2p of Fe/Ni−PPy.
Figure 2. (a) XRD patterns (the shadows are the characteristic peaks of PPy) and (b) FTIR spectra of PPy and Fe/Ni−PPy. The high-resolution XPS curves of (c) C 1s, (d) N 1s, (e) Fe 2p, and (f) Ni 2p of Fe/Ni−PPy.
Molecules 30 01271 g002
Figure 3. (a) LSV curves of PPy, FeNi, and Fe/Ni−PPy. (b) Conversion of NO3, (c) NH3 yield and FE, and (d) selectivity of NH3 and NO2 with the samples at −0.9 V vs. RHE for 1 h.
Figure 3. (a) LSV curves of PPy, FeNi, and Fe/Ni−PPy. (b) Conversion of NO3, (c) NH3 yield and FE, and (d) selectivity of NH3 and NO2 with the samples at −0.9 V vs. RHE for 1 h.
Molecules 30 01271 g003
Figure 4. (a) LSV curves of different catalysts. (b) Conversion of NO3, (c) selectivity of NH3 and NO2, and (d) NH3 yield and FE with the Fe/Ni−PPy samples with different Fe:Ni ratios at −0.9 V vs. RHE for 1 h.
Figure 4. (a) LSV curves of different catalysts. (b) Conversion of NO3, (c) selectivity of NH3 and NO2, and (d) NH3 yield and FE with the Fe/Ni−PPy samples with different Fe:Ni ratios at −0.9 V vs. RHE for 1 h.
Molecules 30 01271 g004
Figure 5. (a) LSV curves of Fe/Ni−PPy (Fe:Ni = 1:1) in electrolyte with or without NO3. (b) The 1H NMR spectra of 14NH4+ and 15NH4+. (c) The time-dependent concentration curves of NO3, NO2, and NH3, and (d) the 6 consecutive cycle tests of Fe/Ni−PPy (Fe:Ni = 1:1).
Figure 5. (a) LSV curves of Fe/Ni−PPy (Fe:Ni = 1:1) in electrolyte with or without NO3. (b) The 1H NMR spectra of 14NH4+ and 15NH4+. (c) The time-dependent concentration curves of NO3, NO2, and NH3, and (d) the 6 consecutive cycle tests of Fe/Ni−PPy (Fe:Ni = 1:1).
Molecules 30 01271 g005
Figure 6. CV curves with different scan rates of (a) Fe/Ni−PPy and (b) FeNi nanoparticles. (c) Cdl diagram fitted from the CV curves. (d) EIS patterns of FeNi and Fe/Ni−PPy. (e) Schematic diagram of NO3RR on Fe/Ni−PPy (*H and *NO3 are the hydrogen atoms and NO3 adsorbed on the catalyst surface, respectively).
Figure 6. CV curves with different scan rates of (a) Fe/Ni−PPy and (b) FeNi nanoparticles. (c) Cdl diagram fitted from the CV curves. (d) EIS patterns of FeNi and Fe/Ni−PPy. (e) Schematic diagram of NO3RR on Fe/Ni−PPy (*H and *NO3 are the hydrogen atoms and NO3 adsorbed on the catalyst surface, respectively).
Molecules 30 01271 g006
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

Li, L.; Yan, P.; Guo, Q.; Zhang, D.; Mao, C.; Yuan, Q.; Sun, H.; Liu, M.; Liu, Y.; Mao, B. In Situ Formation of FeNi Nanoparticles on Polypyrrole Hydrogel for Efficient Electrocatalytic Nitrate Reduction to Ammonia. Molecules 2025, 30, 1271. https://doi.org/10.3390/molecules30061271

AMA Style

Li L, Yan P, Guo Q, Zhang D, Mao C, Yuan Q, Sun H, Liu M, Liu Y, Mao B. In Situ Formation of FeNi Nanoparticles on Polypyrrole Hydrogel for Efficient Electrocatalytic Nitrate Reduction to Ammonia. Molecules. 2025; 30(6):1271. https://doi.org/10.3390/molecules30061271

Chicago/Turabian Style

Li, Lixia, Paihao Yan, Qinkai Guo, Dongxu Zhang, Chunliang Mao, Quan Yuan, Hongtao Sun, Mingze Liu, Yanhong Liu, and Baodong Mao. 2025. "In Situ Formation of FeNi Nanoparticles on Polypyrrole Hydrogel for Efficient Electrocatalytic Nitrate Reduction to Ammonia" Molecules 30, no. 6: 1271. https://doi.org/10.3390/molecules30061271

APA Style

Li, L., Yan, P., Guo, Q., Zhang, D., Mao, C., Yuan, Q., Sun, H., Liu, M., Liu, Y., & Mao, B. (2025). In Situ Formation of FeNi Nanoparticles on Polypyrrole Hydrogel for Efficient Electrocatalytic Nitrate Reduction to Ammonia. Molecules, 30(6), 1271. https://doi.org/10.3390/molecules30061271

Article Metrics

Back to TopTop