DOI:
10.1039/D4TA04511D
(Paper)
J. Mater. Chem. A, 2025,
13, 4576-4586
Improving nitrate-to-ammonia conversion efficiency on electrodeposited nickel phosphide via surface δ-FeOOH modification†
Received
29th June 2024
, Accepted 2nd January 2025
First published on 2nd January 2025
Abstract
Electrocatalytic nitrate reduction to ammonia is an appealing and promising approach for recovering industry and agricultural wastewater. However, there is a lack of efficient electrocatalysts for NO3− conversion at low overpotentials and a range of nitrate concentrations, which could enable nitrogen recovery from various wastewater streams. Here, we report an affordable and efficient electrodeposited amorphous-Ni3P electrode modified with δ-FeOOH, forming a heterojunction catalyst that efficiently converts nitrate into ammonia. The optimum heterojunction performance achieved a high faradaic efficiency of 98% ± 0.72 at −0.3 V vs. RHE with an NH3 production yield of 8.49 mg h−1 cm−2, ranking Ni3P/δ-FeOOH among the most active electrocatalysts so far reported for this process. Ni3P/δ-FeOOH showed good stability using ethyl cellulose as a green and fluoride-free binder without losing the superficial modification. Mechanistic studies by in situ Raman spectroscopy show that the role of δ-FeOOH is the stabilisation of the NOx intermediates. At the same time, Ni3P provides *H species available for NOx reduction, thus resulting in highly selective NO3 hydrogenation. In contrast, in the absence of δ-FeOOH, Ni3P exhibits significantly lower NH3 faradaic efficiency due to simultaneous H2 evolution. Therefore, the present results show the synergistic effect between δ-FeOOH and Ni3P in the electrocatalytic nitrate reduction, with δ-FeOOH playing a crucial role in stabilising the reaction intermediates and enhancing the ammonia selectivity.
Introduction
Ammonia is one of the essential chemical commodities due to its role as a fertilizer,1,2 and has been responsible for ensuring the world's food supply since the development of the Haber–Bosch process.3,4 Also, NH3 has been considered a potential carbon-free molecule for hydrogen storage3,5 since it is a hydrogen-derived chemical of known storage and transportation infrastructure.6,7 However, current ammonia production methods, such as the traditional Haber–Bosch process,8,9 are energy-intensive and contribute significantly to greenhouse gas emissions. The Haber–Bosch process, for instance, is an N2-converting process that consumes almost 2% of worldwide fossil energy.10,11 It is also responsible for ∼2% of the total greenhouse gas emissions, surpassing 2.16 kg of CO2 per kg of NH3 produced,12 emphasising the urgent need for more sustainable and efficient ammonia production methods, such as the one we propose with our modified electrode.
Considering the current and possible future relevance of NH3, alternative production procedures, particularly related to circular economy and appropriate use of residues, are becoming increasingly important. Wastewater, which is often rich in nitrates (NO3−), serves as a significant source of nitrogeneous oxide species. These include industrial wastewater with high nitrate concentrations,13,14 as well as surface water and underground aquifers with lower nitrate levels from agricultural and biological activity.8,13,15 Currently, state-of-the-art water treatment involves a denitrification process driven by microorganisms, which convert these nitrates to N2.16 Thus, an appealing approach for additional ammonia production by residue valorisation would be the electroreduction of these nitrates.17 Electrocatalytic NO3− reduction (eNitRR) can result in ammonia formation with some N2 and H2 evolution.1,2,9
Despite being promising, eNitRR is a complex multi-electron reduction process that requires nine protons and eight electrons (NO3− + 9H+ + 8e− → NH3 + 3H2O), making the kinetics of the process complex.1,2,18 The parasitic hydrogen evolution reaction (HER) is a competing reaction that can prevail under certain conditions. Therefore, the catalyst must be selective for NO3− reduction, providing the reaction intermediate adequate stabilisation and generating H* on the surface from the water-splitting reaction, avoiding fast H* recombination into H2.2,18 Among the earth-abundant metal catalysts, transition metal phosphides (TMPs) have been reported as promising candidates to replace noble metal-based electrocatalysts in several electrocatalytic reactions19,20 due to the hydrogenase-like catalytic mechanism. TMPs can favour hydrogenation reactions, such as eNitRR, through stabilising H* generated during the Volmer step of the HER.20,21 Moreover, TMPs have been primarily studied for electrocatalytic responses attributed to their intrinsic d-electron configuration favouring the material's high conductivity20,22 and stability.22 However, TMPs frequently exhibit poor faradaic efficiency values at lower overpotential for eNitRR due to the parasitic HER.23
Heterojunction engineering can effectively modulate the local electronic structure of TMPs and regulate the H*-adsorption energy.19,22,24 Iron-modified catalysts have significantly enhanced ammonium conversion rates and selectivity by effectively generating H* (ref. 25) species and suppressing N–N coupling.26 Recently, FeOOH has exhibited distinct catalytic performance and selectivity for eNitRR. Qu et al.27 reported a detailed investigation into the crystal phase engineering of FeOOH and its electrocatalytic performance for NO3− reduction. Based on experimental studies and theoretical calculations, the authors discovered that the δ-FeOOH phase shows a unique behaviour that breaks the scaling relationship between the free energy of *NO3 adsorption and the limiting potential and charge transfer. This specific behaviour derives from the hydrogen bond formation between δ-FeOOH and *NOx intermediates, selectively enhancing the adsorption of these intermediates, thereby favouring higher ammonia selectivity and ammonia yield rate over the HER.
Inspired by the cited previous report, we report an improved eNitRR heterojunction electrode developed by modifying an electrodeposited Ni3P electrode with δ-FeOOH nanoplates using ethyl cellulose as a binder. The surface modification with δ-FeOOH was achieved through a two-step process: first, the nanoplates were deposited via drop-casting, followed by a low-temperature thermal treatment at 220 °C. This modification successfully modulates the activity of the electrodeposited amorphous Ni3P, allowing eNitRR to occur at a lower overpotential. Superior performance is achieved by combining the two materials, reaching a faradaic efficiency of 97% and 8.49 mg h−1 cm−2 NH3 yield, even without using conductive polymeric binders, and the electrodes exhibited outstanding stability. In situ Raman spectroscopy showed that the heterojunction electrode mimics the bimetallic catalytic centre of nitrite reductases, in which the FeOOH is responsible for adsorbing NO3− and stabilising the intermediates by forming H-bonds, and Ni3P provides enough H+ for the nitrate hydrogenation. Furthermore, by simple preparation and modification, the FeOOH efficiently regulates the H* recombination and avoids excessive H2 evolution, enhancing the NO3− electroreduction.
Experimental
Synthesis of the CFP/Ni3P electrodes
Ni3P film electrodeposition was performed in a solution of 0.10 M NiCl2·6H2O, 0.2 M NaCl, and 0.3 M NaH2PO2.20 The pH of the deposition bath was adjusted to 4 by adding a concentrated 2.0 M phosphoric acid solution. Carbon Fiber Paper (CFP) (Thermo Scientific – PTFE treated, TGP-H-60) of 1.0 × 1.5 cm2 was used as the substrate, and a carbon rod electrode was used as the auxiliary electrode. Before deposition, the carbon paper substrate was cleaned with acetone and Milli-Q water in an ultrasonic bath for 10 min and dried at 60 °C for 2 h. The deposition was performed by cyclic voltammetry in a delimited 1.0 cm2 area of the carbon paper substrate, scanning from −0.40 V to −1.60 V vs. Ag/AgCl (KClsat.) at 50 mV s−1 for 80 cycles. After deposition, the nickel phosphide films were rinsed with Milli Q water, dried under argon flow, and stored under vacuum.
Synthesis of the FeOOH lamellar samples
The iron oxide-hydroxide lamellar nanoparticles were prepared using an ultrasound-assisted technique.28 First, 100 mL of a 2 M KOH solution was added drop-wise to 100 mL of a 50 mM Fe(NH4)2(SO4)2·6H2O solution. Then, 2.5 mL of H2O2 was added, and the mixture was treated in an ultrasonic bath for 15 min. The nanoparticles were filtered and washed with Milli Q water until complete removal of the unreacted material, followed by ethanol washing, and then dried at 60 °C overnight.
Synthesis of the CFP/Ni3P/δ-FeOOH electrodes
Following the electrodeposition, the amorphous Ni3P electrode surface was modified by drop-casting a suspension of δ-FeOOH nanoparticles. This suspension was prepared by adding 5 mg of δ-FeOOH in 250 μL of ethylene glycol and 20 μL of a 5% ethyl cellulose solution. The mixture was sonicated for 15 min to ensure uniformity. After drop-casting 10 μL of the iron nanoparticle suspension onto the nickel phosphide films, the electrodes were heated at 220 °C in air for 30 min. Control samples, where the δ-FeOOH sample was directly deposited onto the CFP, were also prepared and labelled as CFP/δ-FeOOH.
Sample characterisation
HRFESEM images, EDS mapping, and spectra of the CFP/Ni3P and CPF/Ni3P/δ-FeOOH electrodes were acquired with a ZEISS GeminiSEM 500 coupled with an X-ray detector EDS apparatus. HRTEM analysis was performed using a JEOL JEM2100F with an accelerated voltage of 200 kV. The HRTEM samples were prepared by dropping 20 μL of the material suspension onto the carbon-coated Cu TEM grid. The δ-FeOOH nanoparticle suspension was prepared using 1 mg of the sample in 1 mL of ethanol. The deposited Ni3P sample was obtained by scratching the CFP/Ni3P film, suspending the material in 500 μL of ethanol, and applying 20 μL onto the TEM grid. XRD analysis of the δ-FeOOH nanoparticles in powder form, CFP/Ni3P, and CFP/Ni3P/δ-FeOOH electrodes was performed using a Cubix-pro PANalytical diffractometer recording from 10 to 80°. XPS surface analysis of the CFP/Ni3P and CFP/Ni3P/δ-FeOOH electrodes was performed on a SPECS spectrometer with a Phoibos 150 MCD-9 detector and an Al X-ray source operating at 200 W. The deconvolution of the elements was performed using Casa XPS software (version 2.3.15). Raman spectra were collected from a freshly prepared electrode surface with a Horiba Jobin Yvon-Labram HR UV-visible-NIR Raman microscope spectrometer using a 514 nm laser.
Electrochemical characterisation
Electrochemical characterisation experiments were performed with a potentiostat/galvanostat Gamry interface 1000 workstation using an H-type cell configuration separated with a 117 Nafion membrane. The electrochemical cell was assembled with Hg/HgO (KOH 1.0 M) and graphite rod electrodes as the reference and counter electrodes. The deposited carbon paper electrodes were directly used as working electrodes with an exposed area of 1 cm2. Before the electrochemical experiments, the electrodes were pre-activated by applying −10 mA cm−2 for 600 s in 1.0 M KOH solution. Linear sweep voltammetry (LSV) curves were recorded at a scan rate of 2 mV s−1 from 0.10 to −0.80 V vs. RHE without iR correction in Ar-purged 1.0 M KOH with the addition of 0.10 M KNO3. The potentials were corrected using the following equation: | E(RHE) = E vs. Hg/HgO(KOH 1.0 M) + 0.0592 × pH + 0.098 V | (1) |
The double-layer capacitance of the CFP/Ni3P and CFP/Ni3P/δ-FeOOH samples was obtained by recording cyclic voltammetry curves with scan rates ranging from 10 to 200 mV s−1 to estimate the electrochemical active surface area (ECSA). Electrochemical impedance spectroscopy (EIS) tests were performed at −0.05 V vs. RHE from 10 kHz to 0.10 Hz with steps of 10 mV in the 1.0 M KOH and 1.0 M KOH + 0.10 M KNO3 electrolytes.
Nitrate electrolysis was also performed in an H-type cell. Before each test, the 117 Nafion membranes were boiled in 5% hydrogen peroxide solution, deionised water, and 0.50 M sulfuric acid solution for 1 h in each solution,9 respectively, and finally rinsed with water. Before the test, the electrochemical cell was purged with Ar (99.99%) for at least 20 min. Constant potential electrolysis experiments were performed for 1 h from −0.10 V to −0.80 V vs. RHE for the un-modified CFP/Ni3P and CFP/Ni3P/δ-FeOOH electrodes to evaluate the performance of the prepared samples. After electrolysis, the samples from the cathodic and anodic compartments were collected for the ammonia and nitrite quantification. Besides investigating the performance at various applied potentials, studies were conducted to evaluate the catalyst's performance at different nitrate concentrations, optimise the applied potential, and determine the electrolyte pH effect. All the electrolysis tests were performed in triplicate to ensure the results were reliable.
Determination of NH3
The ammonia yield in the solution was quantified using the colourimetric Indophenol blue method using a Merck Spectroquant® N–NH4 commercial kit. First, a calibration curve for NH3 was constructed using reference NH3 solutions in 1.0 M KOH with the addition of NH4OH from 0.00 to 3.00 mg L−1. For NH3 quantification with the commercial kit, 0.60 mL of the NH4-1 reagent was added to 5.0 mL of electrolyte aliquots, followed by a micro spoon (supplied in the kit) of the NH4-2 reagent, and the sample was stirred for 5 min. After 10 min, 4 drops of the NH4-3 reagent were added, and the samples were analysed by UV-Vis spectroscopy (Cary 5 G UV-vis-NIR, Varian) in the range of 500 to 800 nm after 20 min. 200 μL of 1.0 M sulfuric acid solution was added as a pH correction step before ammonia quantification to avoid ammonia leakage from the samples. This pH correction step was also performed for the standard curve construction. The standard curve equation (y = 0.2707x + 0.02307, R = 0.9994) showed a good linear relationship of the registered absorbance with the ammonia concentration, as demonstrated in Fig. S1.†
Operando Raman spectroscopy
Operando in situ Raman spectroscopy was conducted using a LabRAM HR Confocal Raman Microscope spectrophotometer. The excitation wavelength was 532 nm, and the electrolysis was carried out in a commercial Raman electrochemical flow cell acquired from Redoxme® using KOH 1 M as electrolyte with 0.1 M of KNO3. The Raman spectra were collected using a working electrode with an active area of 3.5 cm2, a platinum wire as the counter-electrode, and an Ag/AgCl electrode as the reference electrode. Blank samples were collected for each catalyst, and the reaction intermediates were determined from the difference in spectra between the spectra under electrochemical operation conditions and those of the catalyst.
Results and discussion
Synthesis and structural characterisation
The nickel phosphide films were prepared on carbon fibre paper using a co-deposition method employing NiCl2 and HPO2− as precursors. The Ni3P/δ-FeOOH electrodes were subsequently prepared by drop-casting the synthesised δ-FeOOH nanoplates on CFP/Ni3P, as illustrated in Fig. 1a, without any thermal conversion step. The phase structures and crystallinity of the FeOOH sample and the electrodeposited nickel phosphide film were analysed by X-ray diffraction. The FeOOH XRD patterns shown in Fig. 1b reveal the characteristic peaks corresponding to the (100), (002), (102), and (110) planes of the δ-FeOOH phase29 (JCPDS no. 00-013-0518). From the XRD patterns of the Ni3P film Fig. S2,† it was possible to index the recorded diffraction peaks of the carbon paper substrate and a broad, weak band between 40° and 60°, indicating the low crystalline nature of the deposited material.20 After thermal treatment at a low temperature (350 °C under an Ar atmosphere), the XRD pattern of the deposited nickel phosphide (CFP/Ni3P-TT in Fig. S2†) shows peaks corresponding to the planes (031), (231), (141), and (132) of the Ni3P crystalline phase in the region of 35°–55°30 (JCPDS no. 34-0501). Also, diffraction peaks due to metallic Ni were observed due to the low P incorporation capacity of the electrodeposition method.31,32 Raman spectroscopy analysis was performed to identify further structural differences on the surface of the Ni3P electrode (Fig. 1c). In addition to the typical bands of the δ-FeOOH pure phase (297, 400, and 664 cm−1),27,33 a new band at 494 cm−1, assigned to the Ni–O vibration,34,35 was observed in the modified electrode. It is proposed that Ni–O forms by the interaction of the superficial metallic Ni with iron oxyhydroxide and its partial oxidation during the modification step.
 |
| Fig. 1 (a) Schematic diagram of the δ-FeOOH sample synthesis. (b) XRD pattern of the δ-FeOOH sample. (c) Raman spectra of the CFP/Ni3P and CFP/Ni3P/δ-FeOOH electrodes. | |
The morphology of the bare (CFP/Ni3P) and δ-FeOOH-modified nickel phosphide electrodes (CFP/Ni3P/δ-FeOOH) was analysed by HRFESEM. Fig. 2a and b show that the nickel phosphide film is formed by the growth of spherical microparticles, creating a uniform coating on the carbon fibre paper. SEM images of the δ-FeOOH-modified Ni3P electrode (Fig. 2d) revealed homogeneously distributed δ-FeOOH nanoplates over the nickel phosphide coating. From higher magnification images in Fig. 2e–f, it is possible to observe that the δ-FeOOH nanoplates cover the granular structure of the nickel phosphide films through weak van der Waals interaction. Elemental mapping analysis of the δ-FeOOH-decorated Ni3P film (Fig. S3†) demonstrates the homogeneous distribution of Fe, O, Ni, and P in the CFP/Ni3P/δ-FeOOH electrode.
 |
| Fig. 2 High-resolution scanning electron microscopy images of the (a–c) CFP/Ni3P, and (d–f) CFP/Ni3P/δ-FeOOH samples. (g–i) HRTEM images of the δ-FeOOH and Ni3P samples. | |
The structure and morphology of the synthesised δ-FeOOH sample were further analysed by high-resolution transmission electron microscopy (HRTEM). The HRTEM image (Fig. 2g) confirmed the lamellar structure of the δ-FeOOH samples with a nanoplate-like morphology. Lattice fringes observed in Fig. 2h were indexed to δ-FeOOH crystals with d-spacings of 0.23 nm and 0.26 nm for the 002 and 100 facets,33 confirming the successful preparation of the δ-phase. HRTEM analysis confirms the low crystallinity of the nickel phosphide material, a fact that is commonly observed for electrodeposited metal phosphide films, which present few crystalline micro-regions.20,32 In the low-crystalline nickel phosphide sample (Fig. 2i), the lattice stripe spacing of 0.246 nm was indexed to the Ni3P (231) plane, revealing a single-crystalline phase of nickel phosphide. Moreover, HRTEM images of the thermally treated nickel phosphide show the presence of the crystalline micro-regions, which agrees with the XRD pattern, supporting the electrodeposited film's low crystallinity. However, a few crystalline regions can be observed.
X-ray photoelectron spectroscopy (XPS) analysis was performed to study the surface elemental composition of the CFP/Ni3P/δ-FeOOH electrode. Fig. S4† shows the survey spectrum of the δ-FeOOH-modified Ni3P electrode, confirming the presence of Ni, Fe, O, and P in the sample. The Ni 2p, P 2p, Fe 2p, and O 1s high-resolution spectra are shown in Fig. 3a–d. In the Ni 2p spectrum, the two spin–orbit coupling peaks (2p3/2 and 2p1/2) of Ni2+ are located at 855.3 and 873.1 eV,36 respectively, accompanied by the two corresponding satellite peaks at 879 and 860.5 eV.37,38 The doublet peak with a binding energy of 851.2 eV is attributed to the metallic Ni formed by HPO2− reduction of Ni2+ on Ni3P,37,39 as reported by other authors for Ni3P samples.36,39,40
 |
| Fig. 3 XPS analysis of the CFP/Ni3P/δ-FeOOH sample: (a) Ni 2p, (b) P 2p, (c) Fe 2p, and (d) O 1s. | |
XPS analysis of the pure Ni3P with Ar+ sputtering was also performed (Fig. S5†). The etching process significantly reduced the intensity of the P–O peak in the high-resolution P 2p spectra, confirming that the superficial phosphate phase is primarily a result of air exposure after the electrode preparation. Furthermore, the sharpening of the P–Ni bond peaks in the Ni 2p spectra indicates a predominant presence of Ni–P bonds in the bulk structure, further establishing nickel phosphide as the principal component of the catalyst composition.
The deconvolution of the P spectrum (Fig. 3b) revealed the 2p3/2 and 2p1/2 peaks located at 129 and 130.1 eV, which are attributed to Ni–P bonds, thereby supporting the formation of Ni3P. The presence of the peak at 132.8 eV can be ascribed to the P–O bond commonly formed by surface oxidation due to air exposure. Analysis of the Fe 2p spectrum shows peaks at 723.6 and 709.9 eV, corresponding to 2p3/2 and 2p1/2 of Fe3+, with two satellite peaks at 716.5
and 730.6
eV (Fig. 3c).37,41 The high-resolution spectrum of O 1s (Fig. 3d) of the CFP/Ni3P/δ-FeOOH electrode was deconvoluted in four distinctive peaks at 532.1, 530.9, 529.8 and 528.8 eV, which can be assigned, respectively, to O–C
O of the ethyl cellulose, Fe–OH, and Fe–O–Fe species of the δ-FeOOH,42,43 and M–O. The last peak can be attributed either to the Fe–O for the δ-FeOOH surface or the superficial Ni–O formed due to the interaction of Ni with the iron oxyhydroxide44,45 and the partial oxidation of superficial Ni metallic species during the modification with FeOOH.
eNitRR performance
The eNitRR experiments were carried out in an H-type cell under ambient conditions. The electrocatalytic activity of the electrodeposited electrodes was primarily evaluated by recording LSV polarisation curves without iR correction in 1.0 M KOH with and without the addition of 0.1 M KNO3. Fig. 4a presents a pronounced increase in the current density and a significant shift in the onset potential with KNO3 addition for all the evaluated electrodes. In contrast, the carbon paper showed a negligible increase, indicating poor catalytic activity for nitrate reduction and confirming that the activity is related to the electrocatalysts. The individual Ni3P and δ-FeOOH electrodes presented distinct electrochemical behaviour upon incorporating nitrate ions in the solution. The Ni3P electrode displays a much higher current density than the δ-FeOOH sample. However, the positive shift at onset potential is not as significant as in the δ-FeOOH electrode, suggesting different selectivity for the eNitRR. In the absence of nitrate in the electrolyte, dashed lines in Fig. 4a indicate that the individual CFP/Ni3P electrode shows the best performance for the HER with an overpotential for −10 mA cm−2 of −165 mV that is smaller than −240 mV for the CFP/Ni3P/δ-FeOOH electrode. The bare δ-FeOOH electrode shows the worst performance for the HER, usually expected for oxyhydroxides due to the stronger hydrogen bonding with the surface O–H, leading to slow H2 evolution kinetics.46–49 Compared with the individual samples, the CFP/Ni3P/δ-FeOOH electrode showed the most enhanced current density in the KOH + KNO3 electrolyte and a more significant positive shift of the onset potential, suggesting a superior catalytic ability for the nitrate conversion.
 |
| Fig. 4 eNitRR performance: (a) linear sweep voltammetry (at 5 mV s−1) 1.0 M KOH with and without 0.10 M KNO3 of bare CFP, CFP/δ-FeOOH, CFP/Ni3P and CFP/Ni3P/δ-FeOOH. (b) Potential dependent faradaic efficiency and (c) NH3 yield rate for the bare Ni3P and the δ-FeOOH modified Ni3P electrode. (d) Linear regression of the experimental data of charging current vs. scan rate obtained by the capacitance method for ECSA estimation. EIS plots for δ-CFP/FeOOH, CFP/Ni3P, and CFP/Ni3P/δ-FeOOH in 1 M KOH electrolyte (e) without NO3− and (f) with 0.1 M NO3−. | |
Double-layer capacitance (Cdl) values were estimated (Fig. 4b) to understand the synergistic effect of the addition of δ-FeOOH nanoplates on the electrochemically active surface area (ECSA) of the Ni3P electrode. From the estimated Cdl values shown in Fig. 4b, the ECSA of the modified CFP/Ni3P/δ-FeOOH (55 cm2) electrode was more prominent than for the separate CFP/Ni3P (41.7 cm2) and CFP/δ-FeOOH (30 cm2) electrodes, revealing that the integration of both catalysts in the CFP/Ni3P/δ-FeOOH electrode increased the number of active sites for the eNitRR.50 The ECSA-normalized polarisation curves in Fig. S6† demonstrate that the CFP/Ni3P/δ-FeOOH electrode achieves a higher current density in the evaluated potential range and, hence, a superior inherent activity.
Continuous nitrate electrolysis experiments were performed for potentials ranging from −0.10 to −0.80 V vs. RHE to optimise the performance for ammonia production. Fig. 4c and d display the potential-dependent faradaic efficiencies and ammonia yield rates for the un-modified and δ-FeOOH-modified Ni3P electrodes. The corresponding chronoamperometry curves are shown in Fig. S7.† The bare Ni3P sample achieves a faradaic efficiency of 69.0 ± 1.2% (−0.50 V vs. RHE) with a maximum ammonia yield rate of 5.84 mg h−1 cm−2 at −0.70 V vs. RHE. Meanwhile, the CFP/Ni3P/δ-FeOOH electrode exhibited an optimum performance with a higher faradaic efficiency peak of 98.0 ± 0.7% at a low overpotential of −0.3 V vs. RHE and NH3 production yield of 8.49 mg h−1 cm−2, therefore, indicating not only a different catalytic activity tendency for the δ-FeOOH-modified Ni3P electrode but also an enhanced ammonia selectivity for this modified electrode, as can be seen in the faradaic efficiency. The modified Ni3P electrode enhances the efficiency of energy (or charge) utilisation for ammonium production, achieving a higher yield with reduced energy consumption.
Nitrite is undesirably obtained as a byproduct of the direct electrochemical nitrate reduction reaction. After electrolysis, the nitrite (NO2−) content was also quantified using the standard electroanalytical method (Fig. S8†). Fig. S9† shows the variation of the faradaic efficiency for the co-product nitrite within the applied potential. Although the nitrite faradaic efficiency at −0.10 V vs. RHE is higher than 10% for both electrodes, the CFP/Ni3P/δ-FeOOH electrode showed a greater decreasing tendency for more negative applied potentials than the CFP/Ni3P electrode. In both cases, the decline in the nitrite faradaic efficiency was accompanied by a significant increase in the ammonia faradaic efficiency and ammonia yield rate, demonstrating that the nitrate conversion to ammonia is the more selective process for the deposited electrodes. Control experiments of the individual CFP and CFP/δ-FeOOH electrodes were conducted to rule out possible sources of external contamination and verify their contribution to the catalytic performance (Fig. S10†).
In comparison with the Ni3P-containing electrodes, the bare CFP/δ-FeOOH showed the lowest reduction ability at −0.30 V vs. RHE with a low faradaic efficiency for ammonia production of 30.0% ± 0.5 and the lowest ammonia yield of 0.25 mg h−1 cm−2. In addition, the bare carbon fibre paper showed a negligible ammonia yield. The developed electrode outperformed most electrocatalysts reported in the literature for alkaline nitrate electroreduction to ammonia51 regarding faradaic efficiency and NH3 yield rate achieved with low overpotential values, see Fig. S11.†
Furthermore, the influence of the δ-FeOOH modification on the Ni3P electrodes in the kinetics of the samples was studied by electrochemical impedance spectroscopy (EIS) analysis in the presence and absence of nitrate ions. Fig. 4e and f show the Nyquist plots recorded at −0.05 V vs. RHE corresponding to the 1.0 M KOH and 1.0 M KOH + 0.10 M KNO3 electrolytes, respectively. The HER takes place with no nitrate, and the charge transfer resistance mediates the reaction on the electrode surface by reducing the water molecule.52,53 As one can see, the CFP/Ni3P electrode exhibits the smallest semicircle in the 1.0 M KOH electrolyte compared to the CFP/δ-FeOOH and CFP/Ni3P/δ-FeOOH and electrodes, indicating a better charge transport to promote H2 production and hence a better HER performance. Fig. 4f provides the Nyquist plots of the prepared electrodes after introducing the NO3− ions in the electrolyte in which the nitrate reduction reaction and hydrogen evolution reaction co-occur. In this case, there is a catalytic behaviour switch, and a smaller semicircle is observed for the CFP/Ni3P/δ-FeOOH, which indicates faster transfer kinetics for the eNitRR. The effect is also observed for the Ni3P electrode, which, despite lower charge transfer resistance than the CFP/δ-FeOOH sample, is still higher than that of the heterojunction. These results indicate that the incorporation of the δ-FeOOH nanoplates leads to faster kinetics for the nitrate ion conversion over the HER and a slightly higher ECSA, which are responsible for the superior catalytic activity observed for the CFP/Ni3P/δ-FeOOH electrode at low overpotential value.
The developed heterojunction was tested for the eNitRR in different nitrate concentrations to investigate the applicability of the electrocatalyst. The concentrations of 0.005, 0.025, 0.050, 0.075, and 0.10 M of NO3− were chosen because they cover the range of the nitrogen oxide ions (NO3− and NO2−) contained in industrial wastewater and contaminated groundwater.54–56 LSV curves for the CFP/Ni3P/δ-FeOOH under the various nitrate concentrations are presented in Fig. S12a.† As verified, the introduction of higher nitrate concentration in the electrocatalytic system implied a positive shift in the overpotential with a significant increase in current density, indicating that the CFP/Ni3P/δ-FeOOH sample possesses the ability to catalyse the nitrate reduction at different nitrate concentrations. From the ammonia production rates and faradaic efficiencies presented in Fig. S12b,† it is verified that both factors gradually increase with the concentration of nitrate ions in the electrolyte. Although a decrease in the faradaic efficiency is observed in the lowest nitrate concentration, the ammonia production rate, 1.1 mg h−1 cm−2, is comparable to some catalysts54,57–60 reported in an alkaline medium with 20 times higher NO3− concentration and usually more negative applied overpotential. Therefore, these numbers at mM NO3− concentration show the superior performance of CFP/Ni3P/δ-FeOOH and open the possible applicability of the electrode to industrial wastewaters.
Stability performance
Consecutive cycling of eNitRR was performed under the optimised conditions to evaluate the stability of the heterojunction electrocatalyst prepared with natural cellulose as the binder (Fig. 5). Surprisingly, the catalyst resisted more than ten catalytic cycles, retaining the high NH3 yield and faradaic efficiency (∼94% F.E. was retained).
 |
| Fig. 5 (a) Consecutive recycling test of the CFP/Ni3P/δ-FeOOH electrode at −0.3 V vs. RHE. XPS spectra of the CFP/Ni3P/δ-FeOOH sample after 10 consecutive recycling cycles: (b) Fe 2p, (c) Ni 2p, and (d) O 1s. (e) High-resolution scanning electron microscopy images of the CFP/Ni3P/δ-FeOOH sample freshly prepared and after the recycling cycles. | |
Moreover, the simple and fluoride-free preparation of the electrode demonstrated excellent stability during startup and shutdown cycles, suggesting that this modification strategy is promising for producing durable catalysts. However, further long-term performance evaluations are necessary to assess its viability for practical applications. Compared with the literature, the electrode exhibited similar stability behaviour to the self-supported catalyst61–64 and superior electrochemical stability to electrodes prepared using synthetic polymers such as Nafion27,60,65,66 and poly(methyl methacrylate) (PMMA).67
In addition, XPS data were further collected from the CFP/Ni3P/δ-FeOOH electrode after the ten consecutive cycling tests to investigate the catalyst stability and the chemical states. For the Fe 2p region (Fig. 5b), the same peaks corresponding to Fe3+ are identified as in the previous characterisation (Fig. 3c) without indicating the reduction of the superficial iron valence state. Only a slight decrease in the intensity is observed, probably due to the partial FeOOH restructuration over time, and the catalyst activity remains almost the same, demonstrating that only a low amount of iron oxyhydroxide is restructured and the cellulose efficiently retains the FeOOH on the surface of the Ni3P. The spectrum of Ni 2p (Fig. 5c) shows a change in the peak intensity at ∼851 eV, referring to Niδ+ from metal species, due to the surface reconstruction already observed for nickel phosphide at long-term operation in alkaline medium.35,68 No changes in the other peaks were observed, indicating no surface P leaching. However, due to the startup and shutdown cycles, an oxidation process is observed in the P 2p spectra for the superficial phosphide (Fig. S13†). As already described in the literature, the superficial metallic Ni and FeOOH generally suffer from reconstruction during the consecutive cycling process in an alkaline medium and produce an iron-nickel phase.49,69,70 Another peak intensity change is observed in Fig. 5d for the O 1s region. The intensity of the Fe–O–Fe peak is weakened, confirming the restructuration of the δ-FeOOH. However, a considerable amount of δ-FeOOH nanoplates stays on the Ni3P surface, as shown in the HRSEM image (Fig. 5d and e). Furthermore, the intensity of the M–O peak becomes relatively weak due to the reconstruction of δ-FeOOH and Ni0 on the surface since this peak can be associated with the Fe–O and the Ni–O formed during the electrode modification step.
Structural dynamics and active site investigation of CFP/Ni3P/δ-FeOOH
Electrochemical in situ Raman spectroscopy were performed to investigate the selectivity change in the nitrate reduction process with the δ-FeOOH modification on the Ni3P surface. In situ Raman tests were conducted at open circuit potential (OCP) and under optimal eNitRR conditions for CFP/Ni3P/δ-FeOOH (−0.3 V vs. RHE with 0.1 M KNO3) to determine the reaction intermediate formation. The Raman spectra were collected 20 min after the start of the electroreduction reaction at a constant potential to detect the intermediates on the catalyst surface. For both catalysts, with and without the FeOOH modification, a decrease in the intensity of the NO3− stretch peak in solution71–73 located at 1048 cm−1 can be observed in Fig. 6a, showing that even without modification, the Ni3P can reduce nitrate in alkaline solutions. Meanwhile, a broad feature from 800–950 cm−1 appears with the potential applied, corresponding to the H* adsorbed from the water dissociation.74 For the δ-FeOOH modified Ni3P, it is observed that the broad feature exhibits lower Raman intensity, which indicates that the presence of δ-FeOOH on the Ni3P surface limits the H* generation and balances the generation of active hydrogen for efficient nitrate reduction, avoiding H2 evolution by the excess formation of adsorbed H*. In the pure Ni3P Raman spectra, there is evidence for the formation of the HNO intermediate around 1529 cm−1 that can be related to the N
O stretch of HNO,75 implying that the adsorbed NO is converted to HNO even in an alkaline environment due to the local acidic-like environment near the electrode surface and the presence of adsorbed H* from the HER,76 which is usually observed for the Ni-based catalyst.48,77,78
 |
| Fig. 6
Operando Raman spectra without applied potential and at −0.3 V vs. RHE for (a) CFP/Ni3P and (b) CFP/Ni3P/δ-FeOOH. | |
Although the HNO intermediate can be detected in the Ni3P Raman spectra, only low-intensity peaks associated with NH3 can be identified at 714 cm−1 (ref. 71 and 75) and 974 cm−1 (ref. 79), and the peak for the NH3 linked to the catalyst surface at 1574 cm−1 corresponding to the NH3 asymmetric bending73,80 is not seen. These results agree with the lower F.E. and NH3 production yield for pure Ni3P catalysts, indicating low ammonia production selectivity. In contrast, after the modification, more intense peaks for NH3 were observed, with the most prominent peak at 1575 cm−1 that can be assigned to the antisymmetric H–N–H deformation mode of the NH3 linked to the Fe sites with Lewis acid character.75,80 Other intense peaks are seen at 714 cm−1 and 1350 cm−1, related to the rocking vibrations81 and HNH deformation of the NH3,71,80 respectively. Furthermore, two new peaks can be observed at 1235 cm−1 and 950 cm−1, both associated with the adsorbed intermediate NH2OH on the δ-FeOOH surface,73,82 which agrees with the information offered by online DEMS in a previous report for FeOOH phases,27 indicating a change in the mechanism of nitrate reduction with the modification. In addition, when the potential is applied, a new broad and intense peak appears around 664 cm−1, possibly related to the coordination of Fe–NOx,83–87 which shows a Raman shift between 570 and 660 cm−1 in comparison to nitroxyl and heme complexes due the stabilisation of the NOx intermediate (*NO, *NO2, and *NO3) by the hydrogen bond88,89 between the hydroxyl groups exposed on the FeOOH and the oxygen atoms present in the nitrate intermediate. Based on the in situ Raman analysis for the pure Ni3P, the H2 evolution reaction prevails over the NO3− reduction. The HER in an alkaline medium usually takes place through the Volmer and Heyrovsky steps:
H2O + e− → *H + HO− (Volmer step) |
*H + H2O + e− → H2 + HO− (Heyrovsky step) |
It should be noted that after the modification with δ-FeOOH, the reaction selectivity changes from the HER to eNitRR. The primary role of the Ni3P appears to be the generation of adsorbed H* by dissociating H2O (Volmer step) to provide enough available surface hydrogen for the hydrogenation of NO3− rather than the reduction of H2O, thus leading to NH3 instead of H2. The role of the iron oxyhydroxide would be, on the other hand, aside from stabilising the Fe–NOx intermediate, the inhibition of hydrogen evolution (Heyrovsky step) by favouring NH3. The enhanced NO3− selectivity can be attributed to the “relay catalytic effects”90 in the adjacent FeOOH and Ni3P catalytic sites, with FeOOH promoting the adsorption and activation of NO3− and the NOx intermediates and the Ni sites lowering the reaction barrier, promoting the hydrogenation of the adsorbed intermediates. Also, in the modified electrode, Ni3P enhances the charge transfer mechanism between the active sites and adsorbed intermediates, lowering the electron migration to the eNitRR intermediate.91–93
Conclusions
We demonstrated that CFP/Ni3P/δ-FeOOH heterojunctions are efficient electrocatalysts for nitrate conversion to ammonia, with low potential and capable of operating under low NO3− concentration conditions. Modifying the Ni3P electrodeposited electrode surface with the δ-FeOOH nanoplates hinders the HER by adsorbing NOx species that consume adsorbed *H, minimising their excess and thereby increasing the ammonia production rate and the faradaic efficiency. The superficial modification treatment depositing an active δ-FeOOH nanoplate material improved the NO3− active sites, promoting an increase in the ECSA of the electrodeposited Ni3P. Electrochemical in situ Raman spectroscopy analysis indicates that the role of nickel phosphide during the eNitRR changes after the modification. We observed that iron oxyhydroxide facilitates NO3− adsorption, at the same time that its structure stabilises the NOx intermediates, while the principal role of Ni3P is to provide adsorbed *H from the water-splitting reaction available for NOx reduction.
Furthermore, the heterojunction catalyst resisted more than ten consecutive cycles of NitRR, showing good stability over the startup and shutdown cycles while retaining the high NH3 yield and faradaic efficiency (∼94% F.E. was retained) even without synthetic polymer binders. The electrode also delivers an excellent faradaic efficiency yield with substantial NH3 production even at 5 mM NO3− concentration. This work shows the potential application of the CFP/Ni3P/δ-FeOOH catalyst for NH3 electrosynthesis and a feasible alternative for modifying the activity and selectivity of electrodeposited metal phosphide electrodes and modulating their activity.
Data availability
Data are available from the corresponding authors upon reasonable request.
Author contributions
Anelisse B. Silva: investigation, writing – original draft. Eduardo Arizono dos Reis: investigation, writing – original draft. Jiajun Hu: investigation. Josep Albero: investigation, writing – review & editing. Lucia Helena Mascaro: resources, review & editing. Caue Ribeiro: conceptualization, writing – review & editing. Hermenegildo Garcia: supervision, writing – review & editing, resources, conceptualisation.
Conflicts of interest
There are no conflicts to declare.
Acknowledgements
The authors are grateful for the sponsorship and financial support of FAPESP (#2017/11986-5, #2018/01258-5, #2020/11756-2, #2022/15742-1, #2021/12394-0 and # 2013/07296-2), Coordenação de Aperfeiçoamento de Pessoal de Nível Superior – Brazil (CAPES) – Finance Code 001, Conselho Nacional de Desenvolvimento Científico e Tecnológico, CNPq, grant number #152607/2022-6, #311769/2022-5, #405133/2023 and #406156/2022-0, and FINEP-Financiadora de Estudos e Projetos grant number #01.22.0179.00, #01.23.0645.00. Also financial support from the Spanish Ministry of Science and Innovation (CEX-2021-001230 S and PDI2021-0126071-OB-CO21 funded by MCIN/AEI/10.13039/501100011033) and Generalitat Valenciana (Prometeo 2021/038 and Advanced Materials programme Graphica MFA/2022/023 with funding from European Union NextGenerationEU PRTR-C17.I1) is gratefully acknowledged.
References
- L. Sun and B. Liu, Adv. Mater., 2023, 35, 2207305 CrossRef CAS PubMed.
- H. Zhang, H. Wang, X. Cao, M. Chen, Y. Liu, Y. Zhou, M. Huang, L. Xia, Y. Wang, T. Li, D. Zheng, Y. Luo, S. Sun, X. Zhao and X. Sun, Adv. Mater., 2024, 36, 2312746 CrossRef CAS PubMed.
- G. Chehade and I. Dincer, Fuel, 2021, 299, 120845 CrossRef CAS.
- J. W. Erisman, M. A. Sutton, J. Galloway, Z. Klimont and W. Winiwarter, Nat. Geosci., 2008, 1(10), 636–639 CrossRef CAS.
- S. Ghavam, M. Vahdati, I. A. G. Wilson and P. Styring, Front. Energy Res., 2021, 9, 580808 CrossRef.
- K. Adeli, M. Nachtane, A. Faik, D. Saifaoui and A. Boulezhar, Appl. Sci., 2023, 13, 8711 CrossRef CAS.
- N. Salmon, R. Reń, R. Bã Nares-Alcántara and A. Alcántara, Sustain. Energy Fuels, 2021, 5, 2814–2839 RSC.
- L. X. Li, W. J. Sun, H. Y. Zhang, J. L. Wei, S. X. Wang, J. H. He, N. J. Li, Q. F. Xu, D. Y. Chen, H. Li and J. M. Lu, J. Mater. Chem. A, 2021, 9, 21771–21778 RSC.
- G. F. Chen, Y. Yuan, H. Jiang, S. Y. Ren, L. X. Ding, L. Ma, T. Wu, J. Lu and H. Wang, Nat. Energy, 2020, 5(8), 605–613 CrossRef CAS.
- M. Capdevila-Cortada, Nat. Catal., 2019, 2(12), 1055 CrossRef.
- X. Liu, A. Elgowainy and M. Wang, Green Chem., 2020, 22, 5751–5761 RSC.
- S. Ghavam, C. M. Taylor and P. Styring, Front. Energy Res., 2021, 9, 600071 CrossRef.
- J. Theerthagiri, J. Park, H. T. Das, N. Rahamathulla, E. S. F. Cardoso, A. P. Murthy, G. Maia, D.-V. N. Vo and M. Y. Choi, Environ. Chem. Lett., 2022, 20(5), 2929–2949 CrossRef CAS.
- C. Lv, L. Zhong, Y. Yao, D. Liu, Y. Kong, X. Jin, Z. Fang, W. Xu, C. Yan, K. N. Dinh, M. Shao, L. Song, G. Chen, S. Li, Q. Yan and G. Yu, Chem, 2020, 6, 2690–2702 CAS.
- R. Jia, Y. Wang, C. Wang, Y. Ling, Y. Yu and B. Zhang, ACS Catal., 2020, 10, 3533–3540 CrossRef CAS.
- S. Huang, Y. Fu, H. Zhang, C. Wang, C. Zou and X. Lu, Front. Microbiol., 2023, 14, 1284369 CrossRef PubMed.
- N. Zhang, G. Zhang, P. Shen, H. Zhang, D. Ma and K. Chu, Adv. Funct. Mater., 2023, 33, 2211537 CrossRef CAS.
- O. Peng, Q. Hu, X. Zhou, R. Zhang, Y. Du, M. Li, L. Ma, S. Xi, W. Fu, Z. X. Xu, C. Cheng, Z. Chen and K. P. Loh, ACS Catal., 2022, 12, 15045–15055 CrossRef CAS.
- Z. Pu, T. Liu, I. S. Amiinu, R. Cheng, P. Wang, C. Zhang, P. Ji, W. Hu, J. Liu and S. Mu, Adv. Funct. Mater., 2020, 30, 2004009 CrossRef CAS.
- A. B. Silva, M. Medina, L. A. Goulart and L. H. Mascaro, Electrochim. Acta, 2024, 475, 143679 CrossRef CAS.
- T. Chouki, M. Machreki, I. A. Rutkowska, B. Rytelewska, P. J. Kulesza, G. Tyuliev, M. Harb, L. M. Azofra and S. Emin, J. Environ. Chem. Eng., 2023, 11, 109275 CrossRef CAS.
- L. Wang, N. Gong, Z. Zhou, W. Peng, Y. Li, F. Zhang and X. Fan, Int. J. Hydrogen Energy, 2022, 47, 18305–18313 CrossRef CAS.
- J. Hu, H. Huang, M. Yu, S. Wang and J. Li, Nano Res., 2024, 17, 4864–4871 CrossRef CAS.
- Q. Zhou, L. Liao, Q. Bian, F. Yu, D. Li, J. Zeng, L. Zhang, H. Wang, D. Tang, H. Zhou and Z. Ren, Small, 2022, 18, 2105642 CrossRef CAS.
- R. Zhang, Y. Guo, S. Zhang, D. Chen, Y. Zhao, Z. Huang, L. Ma, P. Li, Q. Yang, G. Liang and C. Zhi, Adv. Energy Mater., 2022, 12, 2103872 CrossRef CAS.
- Z. Y. Wu, M. Karamad, X. Yong, Q. Huang, D. A. Cullen, P. Zhu, C. Xia, Q. Xiao, M. Shakouri, F. Y. Chen, J. Y. (Timothy) Kim, Y. Xia, K. Heck, Y. Hu, M. S. Wong, Q. Li, I. Gates, S. Siahrostami and H. Wang, Nat. Commun., 2021, 12(1), 1–10 CrossRef.
- K. Qu, X. Zhu, Y. Zhang, L. Song, J. Wang, Y. Gong, X. Liu and A. L. Wang, Small, 2024, 2401327 CrossRef CAS PubMed.
- C. Kauany, S. Azevedo, E. Arizono Dos Reis, J. Carlos Germino, J. A. Moreto, A. J. Terezo, E. Fernando and J. Quites, Quim. Nova, 2017, 40, 534–540 Search PubMed.
- N. Nishida, S. Amagasa, Y. Kobayashi and Y. Yamada, Appl. Surf. Sci., 2016, 387, 996–1001 CrossRef CAS.
- X. Zhang, H. Xue, J. Sun, N. Guo, T. Song, J. Sun, Y. R. Hao and Q. Wang, Green Chem., 2023, 25, 8606–8614 RSC.
- W. Jo, D. Jeong, J. Jeong, T. Kim, S. Han, M. Son, Y. Kim, Y. H. Park and H. Jung, Front. Chem., 2021, 9, 781838 CrossRef CAS PubMed.
- C. Tian, K. Pratama, A. Sharma, M. Watroba, J. Michler and J. Schwiedrzik, Mater. Sci. Eng., A, 2024, 897, 146342 CrossRef CAS.
- J. Hu, S. Li, J. Chu, S. Niu, J. Wang, Y. Du, Z. Li, X. Han and P. Xu, ACS Catal., 2019, 10705–10711 CrossRef CAS.
- J. Huang, Y. Li, Y. Zhang, G. Rao, C. Wu, Y. Hu, X. Wang, R. Lu, Y. Li and J. Xiong, Angew. Chem., 2019, 131, 17619–17625 CrossRef.
- M. Chen, D. Liu, Y. Chen, D. Liu, X. Du, J. Feng, P. Zhou, B. Zi, Q. Liu, K. H. Lo, S. Chen, S. Wang, W. F. Ip and H. Pan, Appl. Mater. Today, 2022, 26, 101343 CrossRef.
- K. Li, G. Zhou, Y. Tong, Y. Ye and P. Chen, ACS Sustain. Chem. Eng., 2023, 11, 14186–14196 CrossRef CAS.
- Y. Xing, S. Liu, Y. Liu, X. Xiao, Y. Li, Z. Wang, Y. Hu, B. Xin, H. Wang and C. Wang, Nano Energy, 2024, 123, 109402 CrossRef CAS.
- H. Li, J. Xiang and H. Li, J. Mater. Sci.: Mater. Electron., 2020, 31, 11425–11433 CrossRef CAS.
- W. Tang, X. Liu, Y. Li, Y. Pu, Y. Lu, Z. Song, Q. Wang, R. Yu and J. Shui, Nano Res., 2020, 13, 447–454 CrossRef CAS.
- S. Liu, Y. Wang, J. Gao, W. Jin, W. Xiao, L. Xin, Z. Xiao, G. Xu, C. Dai, H. Zhang, Z. Wu and L. Wang, Fuel, 2024, 367, 131445 CrossRef CAS.
- Y. Qiu, Q. Jia, S. Yan, B. Liu, J. Liu and X. Ji, ChemSusChem, 2020, 13, 4911–4915 CrossRef CAS.
- G. Wang, W. Tang, S. Yang, M. Lu, H. Wei, L. Cui and X. Chen, J. Mater. Chem. A, 2023, 11, 26519–26530 RSC.
- C. Wan, Y. Jiao, T. Qiang and J. Li, Carbohydr. Polym., 2017, 156, 427–434 CrossRef CAS PubMed.
- B. P. Payne, M. C. Biesinger and N. S. McIntyre, J. Electron Spectrosc. Relat. Phenom., 2009, 175, 55–65 CrossRef CAS.
- A. R. Blume, W. Calvet, A. Ghafari, T. Mayer, A. Knop-Gericke and R. Schlögl, Phys. Chem. Chem. Phys., 2023, 25, 25552–25565 RSC.
- Z. Wu, L. Huang, H. Liu and H. Wang, ACS Catal., 2019, 9, 2956–2961 CrossRef CAS.
- D. Y. Chung, S. W. Jun, G. Yoon, H. Kim, J. M. Yoo, K. S. Lee, T. Kim, H. Shin, A. K. Sinha, S. G. Kwon, K. Kang, T. Hyeon and Y. E. Sung, J. Am. Chem. Soc., 2017, 139, 6669–6674 CrossRef CAS PubMed.
- N. Guo, H. Xue, J. Sun, T. Song, H. Dong, Z. Zhao, J. Zhang, Q. Wang and L. Wu, Next Mater., 2024, 3, 100109 CrossRef.
- K. Zhu, W. Luo, G. Zhu, J. Wang, Y. Zhu, Z. Zou and W. Huang, Chem.–Asian J., 2017, 12, 2720–2726 CrossRef CAS PubMed.
- X. Y. Ji, K. Sun, Z. K. Liu, X. Liu, W. Dong, X. Zuo, R. Shao and J. Tao, Nano-Micro Lett., 2023, 15, 1–15 CrossRef PubMed.
- T. Zhao, X. Li, J. Hu, J. Zhou, X. Jia and G. Hu, J. Environ. Chem. Eng., 2023, 11, 110122 CrossRef CAS.
- S. Anantharaj, S. Noda, V. R. Jothi, S. C. Yi, M. Driess and P. W. Menezes, Angew. Chem., Int. Ed., 2021, 60, 18981–19006 CrossRef CAS PubMed.
- E. Castañeda-Morales, J. O. Peralta-Cruz, F. Ruiz-Zepeda, A. Susarrey-Arce, M. L. Hernández-Pichardo and A. Manzo-Robledo, Mater. Today Energy, 2024, 41, 101525 CrossRef.
- W. Liao, J. Wang, G. Ni, K. Liu, C. Liu, S. Chen, Q. Wang, Y. Chen, T. Luo, X. Wang, Y. Wang, W. Li, T. S. Chan, C. Ma, H. Li, Y. Liang, W. Liu, J. Fu, B. Xi and M. Liu, Nat. Commun., 2024, 15(1), 1–12 Search PubMed.
- A. Richa, S. Touil and M. Fizir, J. Environ. Manage., 2022, 316, 115265 CrossRef CAS.
- A. Rahman, N. C. Mondal and K. K. Tiwari, Sci. Rep., 2021, 11(1), 1–13 CrossRef PubMed.
- J. Li, Y. Zhang, C. Liu, L. Zheng, E. Petit, K. Qi, Y. Zhang, H. Wu, W. Wang, A. Tiberj, X. Wang, M. Chhowalla, L. Lajaunie, R. Yu and D. Voiry, Adv. Funct. Mater., 2022, 32, 2108316 CrossRef CAS.
- Y. Wang, L. Zhang, Y. Niu, D. Fang, J. Wang, Q. Su and C. Wang, Green Chem., 2021, 23, 7594–7608 RSC.
- Y. Guo, R. Zhang, S. Zhang, Y. Zhao, Q. Yang, Z. Huang, B. Dong and C. Zhi, Energy Environ. Sci., 2021, 14, 3938–3944 RSC.
- Z. Gao, Y. Lai, Y. Tao, L. Xiao, L. Zhang and F. Luo, ACS Cent. Sci., 2021, 7, 1066–1072 CrossRef CAS.
- X. Deng, Y. Yang, L. Wang, X. Z. Fu and J. L. Luo, Adv. Sci., 2021, 8(7), 2004523 CrossRef CAS PubMed.
- Z. Liu, C. Wang, C. Chen, C. Li and C. Guo, Electrochem. Commun., 2021, 131, 107121 CrossRef CAS.
- J. Y. Fang, Q. Z. Zheng, Y. Y. Lou, K. M. Zhao, S. N. Hu, G. Li, O. Akdim, X. Y. Huang and S. G. Sun, Nat. Commun., 2022, 13(1), 7899 CrossRef CAS.
- Y. Wang, A. Xu, Z. Wang, L. Huang, J. Li, F. Li, J. Wicks, M. Luo, D. H. Nam, C. S. Tan, Y. Ding, J. Wu, Y. Lum, C. T. Dinh, D. Sinton, G. Zheng and E. H. Sargent, J. Am. Chem. Soc., 2020, 142, 5702–5708 CrossRef CAS PubMed.
- K. Wu, C. Sun, Z. Wang, Q. Song, X. Bai, X. Yu, Q. Li, Z. Wang, H. Zhang, J. Zhang, X. Tong, Y. Liang, A. Khosla and Z. Zhao, ACS Mater. Lett., 2022, 4, 650–656 CrossRef CAS.
- F. Y. Chen, Z. Y. Wu, S. Gupta, D. J. Rivera, S. V. Lambeets, S. Pecaut, J. Y. T. Kim, P. Zhu, Y. Z. Finfrock, D. M. Meira, G. King, G. Gao, W. Xu, D. A. Cullen, H. Zhou, Y. Han, D. E. Perea, C. L. Muhich and H. Wang, Nat. Nanotechnol., 2022, 17(7), 759–767 CrossRef CAS.
- N. C. Kani, N. H. L. Nguyen, K. Markel, R. R. Bhawnani, B. Shindel, K. Sharma, S. Kim, V. P. Dravid, V. Berry, J. A. Gauthier and M. R. Singh, Adv. Energy Mater., 2023, 13, 2204236 CrossRef CAS.
- C. Sun, H. Wang, J. Ren, X. Wang and R. Wang, Nanoscale, 2021, 13, 13703–13708 RSC.
- M. Gong and H. Dai, Nano Res., 2015, 8, 23–39 CrossRef CAS.
- Y. Feng, S. Wang, H. Wang, Y. Zhong and Y. Hu, J. Mater. Sci., 2020, 55, 13927–13937 CrossRef CAS.
- F. Lei, K. Li, M. Yang, J. Yu, M. Xu, Y. Zhang, J. Xie, P. Hao, G. Cui and B. Tang, Inorg. Chem. Front., 2022, 9, 2734–2740 RSC.
- S. E. Bae, K. L. Stewart and A. A. Gewirth, J. Am. Chem. Soc., 2007, 129, 10171–10180 CrossRef CAS.
- L. R. Sadergaski, T. J. Hager and H. B. Andrews, ACS Omega, 2022, 7, 7287–7296 CrossRef CAS.
- Z. Qiu, C. W. Tai, G. A. Niklasson and T. Edvinsson, Energy Environ. Sci., 2019, 12, 572–581 RSC.
- D. P. Butcher and A. A. Gewirth, Nano Energy, 2016, 29, 457–465 CrossRef CAS.
- J. Theerthagiri, J. Park, H. T. Das, N. Rahamathulla, E. S. F. Cardoso, A. P. Murthy, G. Maia, D.-V. N. Vo and M. Y. Choi, Environ. Chem. Lett., 2022, 20(5), 2929–2949 CrossRef CAS.
- X. Wang, C. Xu, M. Jaroniec, Y. Zheng and S. Z. Qiao, Nat. Commun., 2019, 10(1), 1–8 CrossRef PubMed.
- R. Chen, S. F. Hung, D. Zhou, J. Gao, C. Yang, H. Tao, H. Bin Yang, L. Zhang, L. Zhang, Q. Xiong, H. M. Chen and B. Liu, Adv. Mater., 2019, 31, 1903909 CrossRef CAS PubMed.
- R. L. Aggarwal, L. W. Farrar, S. Di Cecca and T. H. Jeys, AIP Adv., 2016, 6, 25310 CrossRef.
- Y. Peng, K. Li and J. Li, Appl. Catal. B Environ., 2013, 140–141, 483–492 CrossRef CAS.
- F. Lei, K. Li, M. Yang, J. Yu, M. Xu, Y. Zhang, J. Xie, P. Hao, G. Cui and B. Tang, Inorg. Chem. Front., 2022, 9, 2734–2740 RSC.
- H. Liu, X. Lang, C. Zhu, J. Timoshenko, M. Rüscher, L. Bai, N. Guijarro, H. Yin, Y. Peng, J. Li, Z. Liu, W. Wang, B. R. Cuenya and J. Luo, Angew. Chem., Int. Ed., 2022, 61, e202202556 CrossRef CAS.
- B. M. Leu, M. Z. Zgierski, G. R. A. Wyllie, W. R. Scheidt, W. Sturhahn, E. E. Alp, S. M. Durbin and J. T. Sage, J. Am. Chem. Soc., 2004, 126, 4211–4227 CrossRef CAS.
- D. Bykov and F. Neese, J. Biol. Inorg. Chem., 2012, 17, 741–760 CrossRef CAS.
- J. Li, Q. Peng, A. G. Oliver, E. E. Alp, M. Y. Hu, J. Zhao, J. T. Sage and W. R. Scheidt, J. Am. Chem. Soc., 2014, 136, 18100–18110 CrossRef CAS PubMed.
- T. S. Kurtikyan, A. A. Hovhannisyan, A. V. Iretskii and P. C. Ford, Inorg. Chem., 2009, 48, 11236–11241 CrossRef CAS PubMed.
- Z. N. Nilsson, B. L. Mandella, K. Sen, D. Kekilli, M. A. Hough, P. Moënne-Loccoz, R. W. Strange and C. R. Andrew, Inorg. Chem., 2017, 56, 13205–13213 CrossRef CAS.
- F. Li, Z. Men, S. Li, S. Wang, Z. Li and C. Sun, Spectrochim. Acta Mol. Biomol. Spectrosc., 2018, 189, 621–624 CrossRef CAS PubMed.
- A. A. Howard, G. S. Tschumper and N. I. Hammer, J. Phys. Chem. A, 2010, 114, 6803–6810 CrossRef CAS PubMed.
- J. Sun, W. Gao, H. Fei and G. Zhao, Appl. Catal. B Environ., 2022, 301, 120829 CrossRef CAS.
- O. Elmutasim and S. M. Alhassan, Ind. Eng. Chem. Res., 2021, 60, 15525–15539 CrossRef CAS.
- L. Partanen, S. Alberti and K. Laasonen, Phys. Chem. Chem. Phys., 2021, 23, 11538–11547 RSC.
- P. Zhang, H. Qiu, H. Li, J. He, Y. Xu and R. Wang, Nanomaterials, 2022, 12, 1130 CrossRef CAS PubMed.
|
This journal is © The Royal Society of Chemistry 2025 |
Click here to see how this site uses Cookies. View our privacy policy here.