Open Access Article
Keyvan
Mirehbar
ab,
Santosh
Kumar
c,
Ignasi
Sirés
d,
Jaime S.
Sánchez
ae,
Georg
Held
c,
Jesús
Palma
a and
Julio J.
Lado
*a
aElectrochemical Processes Unit, IMDEA Energy Institute, Avda. Ramón de La Sagra 3, 28935, Móstoles, Madrid, Spain. E-mail: julio.lado@imdea.org
bEscuela de Doctorado UAM, Centro de Estudios de Posgrado, Universidad Autónoma de Madrid, C/Francisco Tomás y Valiente, n° 2. Ciudad Universitaria de Cantoblanco, Spain
cDiamond Light Source, Harwell Science and Innovation Campus, Didcot, Oxfordshire OX11 0DE, UK
dLaboratori d’Electroquímica dels Materials i del Medi Ambient, Departament de Ciència de Materials i Química Física, Secció de Química Física, Facultat de Química, Universitat de Barcelona, Martí i Franquès 1-11, 08028 Barcelona, Spain
eSungrow Power Supply Co., Ltd, No. 1699 Xiyou Rd, High-tech Industry Development Zone, Hefei, Anhui Province, China
First published on 30th October 2025
Electro-oxidation is one of the most promising and eco-friendly technologies for water decontamination. However, its industrial application is still limited by the high cost, poor faradaic efficiency, low durability, and potential toxicity of common high-power oxidation anodes. These challenges have been addressed by developing a novel composite comprising a mixed metal oxide (NiMnO3) and reduced graphene oxide (rGO). The NiMnO3–rGO anode allowed the fast and complete removal of phenol. Among different highly porous substrates, graphite felt (GF) led to the highest energy efficiency, since the GF/NiMnO3–rGO anode yielded 100% phenol removal within only 30 min at a current density as low as 10 mA cm−2, which was accompanied by 85% COD removal at 120 min. This anode demonstrated excellent stability, maintaining 100% phenol removal efficiency across five consecutive cycles while also showing low energy consumption (60–65 Wh (kg COD)−1). Operando X-ray photoelectron spectroscopy (XPS) and near-edge X-ray absorption fine structure (NEXAFS) analysis provided mechanistic insights. It is demonstrated that rGO shifts the ˙OH production pathway towards the lattice oxygen mechanism (LOM), in contrast to the adsorbate evolution mechanism (AEM) observed for NiMnO3 alone. This mechanistic shift supports the enhanced stability and sustained electrocatalytic activity, contributing to the high performance of the GF/NiMnO3–rGO composite anode in the context of a more sustainable technology for treating organic contaminants.
Electro-oxidation (EO) process is one of the most prominent EAOPs due to its simplicity and rapid pollutant degradation, in which the oxygen evolution reaction (OER) plays a key. This is an electrochemical process that produces ˙OH as an intermediate, proceeding through diverse mechanisms that depend on the nature of the catalyst surface.11,12 In the adsorbate evolution mechanism (AEM), ˙OH is formed as intermediate during the stepwise water oxidation reaction to yield O2. This mechanism involves the adsorption of water molecules onto the catalyst surface, where hydroxyl ion (OH−) is electrochemically oxidized to form adsorbed hydroxyl radical (*OH). This radical is further oxidized to yield highly reactive oxygen species (*O˙), which can then interact with water molecules to form *OOH as an intermediate and, finally, oxygen (O2).13 The AEM is advantageous for pollutant degradation, as it continuously produces reactive species that directly contribute to organic pollutant cleavage. A different pathway is based on the so-called lattice oxygen mechanism (LOM), which involves the participation of lattice oxygen atoms within the catalyst material. Here, lattice oxygen (Olatt) atoms couple with adsorbed *OH or *O˙ species, forming O–O bonds critical to OER. The LOM differs from AEM in the fact that it consumes lattice oxygen, which is subsequently replenished through electrochemical oxidation and deprotonation of OH− ions.13 This mechanism is highly dependent on the structural integrity and oxidation states of the catalyst, as lattice oxygen regeneration is essential for sustaining the OER activity. The electrocatalytic efficiency in both AEM and LOM mechanisms relies significantly on metal centers within the catalyst material. Transition metals such as Fe, Ni, and Mn are known for their ability to facilitate electron transfer and stabilize intermediate radical species.14,15 Such oxidation flexibility of metals supports the stepwise electron transfer needed for OER, as exemplified in biological systems like those where a Mn4CaO5 cluster plays a crucial role in the stepwise oxidation of water, resulting in O–O bond formation.16 In synthetic catalysts, Ni and Mn provide a similar redox versatility, cycling through oxidation states that stabilize intermediates like *O˙ and *OOH, thereby enhancing the radical formation and eventually boosting the pollutant degradation efficiency.
In the EO process, the anode material is crucial, as it influences the efficiency of OER and the availability of reactive species. Conventional anode materials, such as PbO2, boron-doped diamond (BDD), SnO2, and other metal oxides are known as high-oxidation power anodes with great effectiveness,17–22 but they face limitations due to their high cost, scarcity, and environmental concerns, especially toxicity.23,24 Considering these constraints, there is a demand for new anode materials that not only maintain high pollutant degradation efficiency over prolonged use but also address economic and environmental challenges.
This study aims to meet these demands by proposing a novel anode design, consisting of a composite prepared with NiMnO3, acting as an electrocatalytic mixed metal oxide (MMO), and reduced graphene oxide (rGO) to improve the electrical conductivity and accommodate the possible volume change.25 Furthermore, the enhanced surface area provided by rGO is expected to increase the redox activity of NiMnO3, thereby stabilizing the active intermediates over extended periods. The ultimate goal is to improve pollutant degradation efficiency and electrode durability. In previous research, we demonstrated that a nickel foam (NF) substrate coated with NiMnO3 could achieve a high degree and rate of organic pollutant degradation; however, it exhibited limitations in long-term stability.26 Additionally, in the present work, we have explored various substrates, including graphite felt (GF), aluminum foam (AF), and reticulated vitreous carbon (RVC), to optimize the structural composition further. Notably, few investigations have employed in situ or operando spectroscopic techniques to elucidate the mechanisms governing the EO process, particularly for the degradation of organic pollutants. Here, direct analysis using operando X-ray photoelectron spectroscopy (XPS) and near-edge X-ray absorption fine structure (NEXAFS) has been performed on NiMnO3 and NiMnO3–rGO coatings under reaction conditions, thereby revealing the dominant route of ˙OH production. This systematic strategy not only underscores the novelty of such real-time mechanistic insights but also demonstrates the strong performance and durability of the coating, key features for effective and long-lasting water treatment applications.
Finally, the surface and near-surface composition of the NiMnO3–rGO samples was investigated through ex situ and in situ operando XPS and NEXAFS spectroscopy. These experiments were performed at the Versatile Soft X-ray (VerSoX) beamline B07 at the Diamond Light Source, the UK's synchrotron radiation facility in Harwell Campus in Oxfordshire, UK. Near ambient pressure (NAP)-XPS and NEXAFS spectroscopy measurements were carried out in the Teacup Endstation of Branch C at the B07 beamline. A Nafion/C/Catalyst as a working electrode was mounted on the electrochemical flow cell for operando measurements, and NEXAFS spectra were recorded in total electron yield (TEY) mode at room temperature under high vacuum or 2 mbar water vapor using the 600 L mm−1 monochromator grating and an exit slit y-gap of 0.05. The photon energies were calibrated and normalized using I0 characteristics (mesh current at Branch C). Furthermore, background correction was performed at lower photon energies and normalized to 1 at higher photon energies. On the other hand, for operando NAP-XPS measurements, a fixed photon energy of 1200 eV was used, using the 600 L mm−1 monochromator grating and an exit slit y-gap of 0.05 mm. This resulted in a beamline resolution (the broadening introduced by the monochromator) below 0.5 eV. A pass energy of 20 eV was used for high-resolution scans. The Fermi edge was used to calibrate the binding energies. The background was normalized to 1 for all spectra.
The cell comprised an expanded graphite plate (Sigracell TF6) as the current collector for the anode, positioned in the anolyte channel, and a thin stainless-steel plate (0.2 mm thickness, Goodfellow) in the cathodic compartment. A Masterflex® L/S peristaltic pump maintained a constant liquid flow rate, with a 4 mm interelectrode gap and a Nafion 117 membrane (Ion Power) separating the anolyte and catholyte. This design, as reported in our earlier study, significantly advances electrochemical cell technology for water treatment applications.26
To study the electrochemical oxidation of phenol, experiments were conducted using a 0.1 M Na2SO4 (Sigma-Aldrich) solution as the supporting electrolyte. A variety of electrodes, namely NF/NiMnO3, NF/NiMnO3–rGO, GF/NiMnO3, GF/NiMnO3–rGO, AF/NiMnO3, and RVC/NiMnO3 were employed as anodes. Each of them was consistently paired with a stainless-steel sheet serving as the cathode. The other experimental parameters included an initial phenol (Sigma-Aldrich) concentration of 100 mg L−1, equal anolyte and catholyte volumes of 100 mL, a flow rate maintained at 30 mL min−1, and a solution pH of approximately 6.0 (i.e., natural pH of the treated solution).
The chemical oxygen demand (COD) of the solutions under study was determined by employing commercial kits from Merck (ISO 15705) on a Spectroquant® Prove 100 VIS spectrophotometer. In turn, the phenol concentration was measured by UV/Vis spectroscopy on a PerkinElmer (model LAMBDA 1050 WB InGaAs) spectrophotometer that was set at λ between 190 to 350 nm.
.27 Notably, similar peaks were also clearly visible in the NiMnO3–rGO nanocomposite sample. The prominent first peak in NiMnO3–rGO results from overlap with rGO characteristic diffraction at a similar 2θ position, making the (110) reflection appear more intense than in pure NiMnO3. rGO typically exhibits a broad reflection around 24–26°, depending on reduction conditions and stacking order. These findings confirm that the crystalline structure remained essentially unchanged during the hybridization process. Further insights into the structural nuances were gained through Raman spectroscopy, as illustrated in Fig. 2b. The spectra for both NiMnO3 and the NiMnO3–rGO nanocomposite exhibited characteristic peaks within the 200–700 cm−1 range, attributable to metal–oxygen bonding vibrations. Notably, the NiMnO3–rGO nanocomposite spectrum exhibited additional bands at 1329 cm−1 and 1583 cm−1, corresponding to the D and G bands of carbonaceous structures, respectively. The D band is associated with structural defects in sp2-hybridized carbon domains, whereas the G band is attributed to the first-order scattering of the E2g vibrational mode in graphitic carbon.28 A noteworthy observation is the lower ID/IG ratio in the nanocomposite (0.90) as compared to that of pure rGO (0.98), suggesting a successful reduction of GO to rGO. The distinct peaks at 542 and 611 cm−1 are attributed to the MnO vibrational modes, and those at approximately 370 and 1050 cm−1 likely correspond to Ni–O bonding.29 TGA curves provided quantitative insights into the composition of the nanocomposites (Fig. 2c). The curve for pure NiMnO3 exhibited a marginal mass loss of approximately 1%, likely due to the release of lattice-incorporated water or surface-adsorbed hydroxyl groups. Conversely, the pure rGO sample demonstrated a near-complete mass loss upon heating to 500 °C, confirming its decomposition. For the NiMnO3–rGO nanocomposite, an abrupt mass loss was observed between 500 °C and 600 °C, attributable to the combustion and decomposition of the rGO component, allowing for an estimation of the NiMnO3 mass content at around 80% in the nanocomposite.
The high-resolution XPS spectra of Ni 2p, Mn 2p, C 1s, and O 1s of NiMnO3–rGO sample, obtained from the survey scan (Fig. S3), are shown in Fig. 2d–f. Ni 2p scan (Fig. 2d) indicates that nickel exists in its mixed state Ni2+ (854.7 eV and 857.9 eV)/Ni3+ (855.9 eV) at the surface of NiMnO3–rGO sample. In addition, the shakeup satellite peak was also observed at 7 eV above the main emission line for Ni 2p3/2 due to the presence of Ni2+. The Mn 2p scan (Fig. 2e) confirms the presence of Mn3+ (639.5 eV) and Mn4+ (642.0 eV). Similarly, the binding energy peaks in the O 1s scan were positioned at 529.9, 531.3, and 533.2 eV corresponding to metal–O–metal (M–O–M), hydroxides (OH−), and adsorbed water molecules (H2O), respectively (Fig. 2f). Thus, the resultant phases are the oxyhydroxides of Ni and Mn. The apparent higher quality of the O 1s spectra arises primarily from the stronger signal of surface-adsorbed oxygen species, compared to the relatively weaker lattice oxygen contribution. In contrast, the Ni 2p and Mn 2p spectra inherently display lower signal-to-noise ratios, which can be attributed to the lower photoemission cross-sections of Ni and Mn and the presence of complex multiplet splitting. These factors make the transition-metal spectra appear noisier, even though they were acquired under identical measurement conditions. The C 1s spectra of rGO (NiMnO3–rGO) exhibits three major components at 288.8, 286.1, 284.6 eV, which can be indexed to the O
C–O, C–O, and C–C, respectively (Fig. S4). The XPS results provide evidence of the surface oxidation states of the metallic centers in the synthesized NiMnO3–rGO and the existing bonds on the surface of this material. Note that Fig. 2f shows the clear overlapping contributions such as O
C/M–O and C–O/M–H2O.
SEM imaging of the NiMnO3–rGO nanocomposite is presented in Fig. 3a, offering a detailed morphological perspective. The micrographs reveal that the composite predominantly comprises particles of a few micrometers, exhibiting irregular shapes. This morphological characteristic is likely a result of the attachment of small metal oxide nanoparticles onto larger rGO flakes, approximately 20 nm in size. Notably, a certain degree of particle agglomeration was observed, which is hypothesized to have occurred during the drying phase of the sample preparation. EDS analysis further complemented the SEM observations, confirming an equitable distribution of nickel (Ni), manganese (Mn), and oxygen (O) elements across the surface of the nanocomposite. Additionally, a carbon (C) content of approximately 15% was identified through EDS, corroborating the findings obtained from TGA curves. This good agreement between EDS and TGA results underscores the reliability of our compositional analysis and provides a comprehensive understanding of the elemental distribution in the nanocomposite.
TEM analyses were conducted to elucidate the distribution of oxide nanoparticles on the rGO nanosheets within the NiMnO3–rGO nanocomposite. Fig. 3b reveals the rGO nanosheets, characterized by their thin-layered structure with noticeable surface wrinkling. However, identifying distinct crystalline domains on these nanosheets proved challenging. Enhanced magnification, as displayed in Fig. 3c, provides a clearer view of the carbon network surface. It is densely and uniformly covered with NiMnO3 nanoparticles, each approximately 2.5 nm in diameter (Fig. 3d). This uniformity suggests a robust anchoring of nanoparticles across the surface of the conductive rGO layers. The nanoparticles appear to lack defined boundaries, and their agglomeration contributes to the nanoporous architecture of the nanocomposite. An intriguing observation from Fig. 3e is the uneven decoration of the rGO edges with NiMnO3 nanoparticles, potentially due to a preferential attachment of metal oxides to the basal plane of rGO, which is rich in epoxide and hydroxyl groups, as opposed to the edges featuring carboxyl and carbonyl groups. This distribution and the potential interaction of the nanoparticles with the rGO matrix is expected to become crucial to boost the electrochemical performance of the material.30 The intimate association between the nanoparticles and rGO enhances conductivity and shortens ion transport pathways.31 Such a configuration facilitates redox reactions, significantly impacting the overall electrochemical response.32 This microscopic insight into the nanocomposite structure provides a valuable understanding of its functional properties and potential applications in electrochemical systems.
The comparative analysis of substrates highlights the superior performance of the GF/NiMnO3–rGO electrodes, achieving nearly complete phenol removal (∼100%) within 30 min, whereas NF/NiMnO3–rGO electrodes required up to 60 min to reach similar efficiency (Fig. 5a and c). This result suggests a synergistic effect when using GF as a substrate in combination with rGO. Additionally, COD removal trends mirror these findings (Fig. 5b and d), indicating that GF/NiMnO3–rGO is the most efficient in pollutant degradation.
In terms of energy consumption, all electrodes demonstrated low energy requirements, falling within the range of 60–65 Wh per kg of COD removed (Fig. 5e). In particular, the rGO-containing electrodes exhibited the lowest energy consumption values, underscoring the positive impact of rGO in reducing resistance and enhancing electron mobility, which translates into lower operation costs. One of the most striking outcomes of this study was the exceptional stability demonstrated by the GF/NiMnO3–rGO electrode. Stability assessments revealed an astonishing level of robustness, with the electrode maintaining its phenol removal efficiency virtually unchanged over five consecutive cycles of 2 h (Fig. 5f). This outstanding performance contrasts with the NF/NiMnO3 electrodes, which suffered a significant efficiency decline of over 25% under identical conditions. These results highlight the superior stability advantages imparted by the GF substrate and rGO modification, establishing the GF/NiMnO3–rGO electrode as a highly reliable and durable solution for phenol removal applications.
Based on these results, the GF/NiMnO3–rGO electrode stands out as the most promising candidate, offering high removal efficiency, low energy consumption, and superior durability, making it an ideal choice for advanced EOPs targeting organic pollutants.
All experiments were conducted in 0.1 M Na2SO4 electrolyte (neutral pH) with an initial phenol concentration of 100 mg L−1, operated at a current density of 10 mA cm−2.
![]() | ||
| Fig. 6 Operando XPS spectra of NiMnO3 (a) Ni 2p; (b) Mn 2p; (c) O 1s, and NiMnO3–rGO (d) Ni 2p; (e) Mn 2p; (f) O 1s. | ||
Based on our operando XPS and NEXAFS analysis, the prevailing OER pathway on the pure NiMnO3 electrocatalyst is consistent with the adsorbate evolution mechanism (AEM, Fig. 7a). This interpretation is supported by the observed redox transitions of Ni2+/Ni3+ and Mn3+/Mn4+ in Fig. 6, which indicate that oxygen evolution proceeds via surface-adsorbed intermediates. The AEM involves four concerted proton–electron transfer steps centered on the metal ions: initial OH− adsorption at a surface vacancy, deprotonation to form O*, O–O bond formation via reaction with a second OH−, and eventual release of O2 while regenerating the active site.36,37 In contrast, for GF/NiMnO3–rGO, the operando spectra reveal a higher degree of lattice oxygen participation, as evidenced by the reversible changes in the O 1s and Mn L-edge features (Fig. 6d–f). These observations are consistent with the lattice oxygen mechanism (LOM, Fig. 7b), where O–O bond formation proceeds via coupling between lattice oxygen atoms. The free energy diagram (Fig. 7c) illustrates the difference: while the AEM pathway relies on metal-centered adsorbates, the LOM pathway involves lattice oxygen redox, leading to a distinct energetic profile that rationalizes the improved kinetics and long-term stability observed for the rGO-containing electrode.
![]() | ||
| Fig. 7 Proposed OER mechanism over (a) NiMnO3 and (b) NiMnO3–rGO electrocatalysts. (c) Comparison of the OER mechanisms via the AEM and LOM. | ||
X-ray absorption spectroscopy (XAS), particularly L-edge XAS (or L XAS), is one of the most effective methods for investigating the oxidation states of 3d transition metals using absorption edge positions and spectral characteristics. The NEXAFS spectra are interpreted qualitatively on the basis of line-shape evolution, edge shifts, and relative intensity changes, as full multiplet/crystal-field fitting of transition-metal L-edges is non-trivial and often non-unique; this qualitative analysis, supported by XPS results, is sufficient to track oxidation state and local coordination changes under operando conditions. Fig. 8 shows the L-edge XAS spectra of the observed, Ni, Mn L-edge, and O K-edge spectra, during operando measurements under real-time OER conditions as a function of applied potentials for pure NiMnO3 (Fig. 8a–c) and NiMnO3–rGO (Fig. 8e and f). The O K-edge spectra in Fig. 8c and f provide complementary information on the involvement of lattice oxygen and the formation of surface hydroxyl terminations (M–OH) and peroxo-like (μ-O) species as the potential increases. The appearance and growth of these species at specific photon energies confirm the progressive oxidation of lattice oxygen and the transformation of adsorbed intermediates, supporting a mechanism where lattice and adsorbed oxygen species interchangeably contribute to O–O bond formation. Very interestingly, as shown in Fig. 8f, at OCP, the intensity of t2g and eg changes, which is related to the removal of the electrons of oxygen and the increase of holes in the t2g and eg orbitals, indicating the formation of M–O–O–M bond. Further increase in the applied potential to slightly above OER potential, we observe the oxidation of O moiety as we can see the lowering intensity of t2g orbitals.36,37 The transition between oxidation states of Ni (Fig. 8a and d) and Mn (Fig. 8b and e), facilitated by their redox flexibility,38,39 is critical for stabilizing the high-energy intermediates involved in the OER, thereby enhancing the overall catalytic efficiency of the NiMnO3–rGO composite for phenol oxidation.40
Overall, these operando XAS observations provide deep insight into the dynamic electronic restructuring of the NiMnO3–rGO catalyst under OER conditions, underscoring the role of LOM. The spectral changes confirm the simultaneous involvement of Ni and Mn centers in stabilizing reactive oxygen species, which promotes sustained ˙OH generation and is expected to improve the electrocatalyst efficiency and durability in pollutant degradation.
In contrast, the NF/NiMnO3 electrode shows a higher initial voltage, followed by fluctuations before reaching a relatively stable value near 3.5 V. The initial instability and the higher starting potential observed with NF/NiMnO3 may be attributed to the less conductive and less stable nature of the NF substrate, which could lead to higher internal resistance and gradual degradation of the electrochemical performance. These fluctuations could also indicate a slower stabilization of the catalyst surface or possible degradation of active sites over time, affecting the electrode efficiency in maintaining a consistent OER rate. The results clearly demonstrate the superior stability of the GF/NiMnO3–rGO material. This finding aligns with the aforementioned enhanced catalytic activity and efficiency of the GF/NiMnO3–rGO electrode in pollutant degradation, further reinforcing its suitability for long-term applications in electrochemical oxidation processes.
To evaluate the structural stability of the GF/NiMnO3–rGO electrode after extended electrochemical testing, SEM images were captured before and after 400 h of continuous operation (the previous test of 250 h was extended for an additional 150 h). The first set of SEM images (Fig. 9b and c) provides higher-magnification views of the particle morphology before and after the stability test. Notably, the individual NiMnO3–rGO particles maintain their original morphology, showing no signs of structural collapse or significant agglomeration after 400 h of operation. The unchanged particle morphology suggests that the NiMnO3–rGO composite itself possesses high structural stability, with the rGO matrix potentially playing a role in preserving particle integrity by mitigating particle agglomeration or collapse under prolonged reaction conditions. In the second set of SEM images (Fig. 9d), taken before testing, the electrode surface exhibits a well-distributed layer of NiMnO3–rGO particles on the GF substrate. These particles are uniformly coated and exhibit a cohesive structure, with no apparent signs of detachment or degradation. This initial morphology indicates a robust integration of the NiMnO3–rGO active material on the substrate. After 400 h of operation, however, the SEM images in Fig. 9e reveal evidence of material loss from the electrode surface. The originally continuous layer of NiMnO3–rGO appears thinner, and certain regions show partial exposure of the underlying GF fibers. This suggests that prolonged electrochemical cycling led to some detachment of the active material, possibly due to gradual wear or mechanical stress under reaction conditions. Despite this material loss, the remaining active layer retains its integrity, covering a substantial portion of the surface, which likely contributed to the electrode sustained performance observed over the test period. These SEM analyses indicate that while there is some material loss on the macro-scale, the microstructure and morphology of the NiMnO3–rGO particles remain stable, which could be attributed to the stabilizing influence of rGO. This partial retention of active material, combined with preserved particle morphology, likely underpins the electrode's overall stability and efficiency, making GF/NiMnO3–rGO a promising candidate for long-term applications despite minor material detachment.
Finally, the EDS analysis performed on the GF/NiMnO3–rGO electrode after 400 h of continuous operation shows that Ni and Mn are uniformly distributed, which suggests that the NiMnO3 active layer remains well-integrated with the rGO and GF substrate, even after prolonged electrochemical cycling (Fig. 9f). This distribution indicates that the NiMnO3–rGO composite structure is stable under extended reaction conditions, as no significant agglomeration or segregation of metal species is observed. The presence of oxygen is consistent with the anticipated formation of oxide layers, which contribute to the OER mechanism through the production of hydroxyl radicals. The uniform oxygen distribution supports the prevalence of the LOM mechanism in OE, as discussed earlier, where lattice oxygen and adsorbed intermediates interact to stabilize O–O bonds and facilitate radical generation.13,36,41
Note that, additionally, trace amounts of sodium (Na) and sulfur (S) were detected on the electrode surface, originating from the Na2SO4 supporting electrolyte used during the reaction.
Supplementary information: procedures used to calculate phenol and COD removal efficiencies, a detailed description of the anode preparation process, and the textural and structural characterization of the active materials. Furthermore, it outlines the rationale for selecting the electrode substrates, as well as a comprehensive description of the electrochemical cell design and the equipment employed during the synchrotron-based XPS and NEXAFS experiments. See DOI: https://doi.org/10.1039/d5ta05337d.
| This journal is © The Royal Society of Chemistry 2025 |