Sandhyawasini Kumariab,
Amrita Tripathyab,
Vishalakshi Gurumurthy DileepKumar*ac,
Afaq Ahmad Khan
ad,
Ashoka Siddaramannae,
John Kiwif,
Mysore Sridhar Santosh*ab,
Sami Rtimi
*f,
Khushwant Singhg and
Sai Smaran S. B.h
aCoal to Hydrogen Energy for Sustainable Solutions (CHESS) Division, CSIR–Central Institute of Mining and Fuel Research (CIMFR), Digwadih Campus, PO: FRI, Dhanbad-828108, Jharkhand, India. E-mail: santoshms@cimfr.res.in; deepuvg29@gmail.com
bAcademy of Scientific and Innovative Research (AcSIR), Ghaziabad-201002, India
cDepartment of Chemistry, HKBK College of Engineering, Bangalore-560045, Karnataka, India
dGreenCat Lab, Department of Chemical Engineering, Indian Institute of Technology (ISM), Dhanbad-826004, Jharkhand, India
eDepartment of Chemistry, School of Applied Sciences, REVA University, Bangalore-500064, Karnataka, India
fGlobal Institute for Water Environment and Health, 1201 Geneva, Switzerland. E-mail: rtimi.sami@gmail.com
gDepartment of Chemistry, Pradhanmantri College of Excellence, SLP Govt PG College, Morar, Gwalior-474006, Madhya Pradesh, India
hSai Smaran S.B., Duke University, Box 90300, Hudson Hall, Durham, NC 27708-0287, USA
First published on 15th July 2025
Granular MgNiO2 has emerged as a promising catalyst owing to its remarkable electrocatalytic activity and photodegradation efficiency under visible light. In this work, granular surface-engineered MgNiO2 nanoparticles were synthesized using the precipitation method. The interaction of Mg and Ni, forming Mg–Ni–O structures during high-temperature MgNiO2 synthesis, was investigated through X-ray photoelectron spectroscopy (XPS) analysis. The presence of Ni3+ species in the ionic form indicated charge transfer reactions in the catalyst. The band gaps of the as-prepared MgNiO2 and NiO were determined to be 2.2 eV and 3.7 eV, respectively. The first-order transverse optical (TO) phonon modes observed at 536 cm−1 indicated the presence of NiO, which was identified as the primary contributor to the Raman peaks. Further, the photocatalytic degradation of caffeine under visible light achieved a removal efficiency of 95.5% within 180 minutes. The intermediate reactive oxidative species (ROS) leading to MgNiO2 degradation were identified, and their lifetime and diffusion length in the solution were reported. Superoxide (O2−˙) and hydroxyl (˙OH) radicals were identified as the main ROS contributing to caffeine degradation. The electrocatalytic oxygen evolution reaction (OER) indicated a high density of oxygen vacancies in MgNiO2 compared to NiO, suggesting the promoter role of Mg species in the photocatalyst. These insights provide a holistic understanding of MgNiO2 as a catalyst and its pivotal role in green and efficient caffeine photodegradation and the electrocatalytic OER.
Caffeine has been reported to be only partially degraded in secondary biological wastewater treatment plants (WWTPs) and passes consequently into lakes, rivers and the sub-surface underground streams. The persistence of caffeine in water bodies has been observed to increase steadily in natural ecosystems over the past few decades. Some articles have recently addressed the methods used to degrade caffeine.1–3 Furthermore, caffeine in relatively high concentrations is difficult to degrade through secondary biological wastewater treatment in municipal treatment plants. Therefore, the need to explore new technical approaches to degrade caffeine is highly imperative. Recently, electro-Fenton, anodic oxidation, and light-activated processes have been reported to lead to caffeine degradation.4–6 Photocatalysis is a highly promising water treatment technology7,8 and appears to suit this objective. In this process, illumination prompts the excitation of photocatalysts, generating electron–hole pairs directly. The efficiency hinges on the recombination rate or lifetime of these pairs; faster recombination translates to a shorter reaction time. Photoluminescence (PL) spectra are vital for probing this recombination rate.
In the past decade, heterogeneous photocatalysis significantly advanced solar energy conversion and improved environmental remediation, garnering widespread attention. Future mass production necessitates the meticulous selection of environmentally friendly, cost-effective, and potent photocatalysts for practical applications, which plays a pivotal role in the development of advanced oxidation processes. Moreover, several countries have reduced the permissible levels of caffeine in drinking water.9 Thus, innovative photocatalytic materials should be explored to increase the degree of water quality.6
In the photodegradation of organic compounds by a photocatalyst, three primary reaction pathways are envisaged: photocatalysis, photolysis, and dye photosensitization. During photolysis, a photoinduced electron directly interacts with O2, generating singlet oxygen (1O2) as an oxidant for the subsequent photolysis of the organic compound, outlining key mechanisms in the process.10,11 Nanomaterials, with their dual functionality as photocatalysts and electrocatalysts, have been recently explored for energy conversion and storage.12,13 One such material, MgNiO2, has found applications in processes related to methane reforming, sensors, supercapacitors, H2 storage, and environmental cleaning. Recent work has also explored MgNiO2 for sensing applications due to its flower-structured flakes with high porosity.14
Another application of MgNiO2 is in gas separation membranes, where its incorporation into mixed matrix membranes (MMMs) has revolutionized gas permeation performance. When embedded in a polysulfone (PSf) matrix,15 MgNiO2 nanoparticles introduced nanoscale voids, facilitating faster gas transport and selective gas permeation while improving the membrane stability and efficiency. Beyond its role in catalysis, MgNiO2 has shown remarkable potential in lithium-metal batteries.16,17 Its incorporation as a ceramic additive improves the protective film quality on lithium anodes, significantly reducing interfacial resistance and enhancing cycling stability. Additionally, MgNiO2-based composites have been explored in zinc-ion hybrid supercapacitor batteries (ZHSBs), where its hollow micro-sphere morphology facilitates fast ion insertion and extraction, contributing to a high energy density of 157–89 Wh kg−1 and a long-term cycling stability of 10000 cycles with 97% retention. Further, a critical application of MgNiO2 is in hydrogen storage. Incorporating MgNiO2 into MgH2 significantly lowers the hydrogen desorption temperature,18 improves absorption kinetics, and enhances storage capacity by facilitating synergetic catalytic effects. Additionally, its high electrochemical stability and superior capacitance retention make it an ideal supercapacitor electrode material.
Hydrothermal preparation to synthesize MgNiO2 in the form of chrysanthemum flowers (CFs) has been reported to show effective sensing characteristics. The Ni-ions in the flower-like MgNiO2 presented multiple Ni3+/Ni2+ oxidation states inducing the electronic behavior.19 Chrysanthemum flower-like nanostructured MgNiO2 electrodes have also been recently reported for the removal of Hg2+ ions in an aquatic medium.14 MgNiO2 has been synthesized using different methods, and different morphologies have been identified. Many factors such as pH, surfactant, and calcination temperature affect the morphology during the precipitation process. A pH of 7 to 9 would be more favourable for obtaining granular morphology.
This work focuses mainly on the preparation of granular MgNiO2 to degrade caffeine under visible light irradiation. This material also shows an electrocatalytic oxygen evolution reaction (OER). Linear sweep voltammetry (LSV) and Tafel slopes have been employed to determine the kinetics, rate of the OER and the reaction mechanism. The MgNiO2 synthesis was designed in such a way that a large surface area is avoided similar to flakes. This prevents significant adsorption of pollutants in the aqueous phase while enabling catalytic and photocatalytic surface processes to proceed. This work presents a detailed study of the MgNiO2 surface and the photoinduced degradation kinetics for caffeine. The mode of intervention of the reactive oxygen species (ROS) in solutions during caffeine degradation is addressed herein. Considering the above-mentioned aspects, this work provides better insights and an understanding of the photocatalytic degradation process and the oxygen evolution reaction mediated by MgNiO2.
To further analyze the morphology of the synthesized nanoparticles, the TEM micrographs of MgNiO2 were taken by dispersing the catalyst in EtOH and drop-cast onto a copper grid. The TEM micrographs shown in Fig. 1 display the well-distributed granular-shaped nanoparticles with irregular edges of the catalyst. These edges of particles indicate a well-ordered crystalline structure obtained in a controlled growth environment during the synthesis. For the sake of comparison, the TEM micrographs of NiO from the literature were used as a reference, which reveal the formation of agglomerates, resembling the MgNiO2 clusters, indicating similar structural characteristics.20 The STEM-EDS (Fig. 1b–e) mapping was performed to investigate the sample's composition, providing evidence for the presence of Mg, Ni and O in the as-prepared material. The micrographs display the uniform distribution of Mg, Ni and O, whereas the SAED pattern shows the polycrystalline nature of the prepared MgNiO2 nanoparticles (Fig. 1f).
The average crystallite size was estimated from the broadening of the diffraction peaks by the Scherrer method.21 The Scherrer equation in eqn (1) relates the size of sub-micrometre crystallites with the observed peak broadening obtained during the XRD diffraction runs.
D = kλ/β![]() | (1) |
The FTIR spectrum of the MgNiO2 sample displayed in Fig. 2d shows a vibrational frequency at 3433 cm−1 corresponding to the vibrational frequency of –OH bonds. The moisture on the sample's surface leads to the absorption of surface water molecules on MgNiO2, giving rise to the observed –OH bond peak. The –OH bond peak is derived from surface water rather than structural chemical bonds within MgNiO2. The peak at 519 cm−1 arises from the vibrational bond frequency of Mg–O,24 while the peak at 443 cm−1 is associated with the vibrational bond frequency of Ni–O.25 The peaks shown in Fig. 2d observed for the MgNiO2 sample agree well with that of the literature ref. 26. The peak at 939 cm−1 likely corresponds to the metal–oxygen (M–O) stretching vibration bonds of Mg and Ni, since both are bonded to oxygen and involve stretching of the bonds between the metal ions and oxygen. The peak at 1563 cm−1 is associated with the bending vibrations of oxygen atoms. This is linked to the oxide-structure deformation or the bending of oxygen atoms in MgNiO2.
The Raman spectroscopy findings of MgNiO2 are given in Fig. 2e. A peak observed at 536 cm−1 can be attributed to first-order transverse optical (TO) phonon modes, while the peak at 1042 cm−1 could have originated from longitudinal optical (LO) phonon modes. The significant presence of NiO in the sample probably plays a crucial role in the emergence of these distinct Raman peaks, highlighting its influence on the vibrational dynamics of MgNiO2. Moreover, the results from the Raman spectra suggest that the catalyst MgNiO2 not only exhibits a substantial degree of phase purity but also reflects well-defined phonon behaviour, which is essential for its potential applications in catalytic processes.27 The 536 cm−1 Raman peaks in MgNiO2 appears higher than the typical TO phonon modes of pure NiO and MgO (446 cm−1). The shift in the Raman peak to 536 cm−1 can be explained by modifications in the lattice dynamics caused by the coexistence of Mg2+ and Ni2+ cations. This mixed-cation environment alters the bond strengths within the crystal, increasing the force constants and consequently increasing the phonon frequencies.28 The 1042 cm−1 peak lies close to the 2LO mode of NiO (1009 cm−1) and the surface mode of MgO (1080 cm−1), suggesting that it probably originates from a second-order phonon process or a surface-related vibrational mode influenced by Mg–Ni interactions in the mixed oxide lattice.14,29
![]() | ||
Fig. 3 Oxidation states of the synthesized catalyst by XPS deconvolution of (a) Mg 1s, (b) Ni 2p, and (c) O 1s in the prepared MgNiO2. |
The Ni 2p peaks in Fig. 3b indicate the presence of Ni2+ in MgNiO2. The binding energy of Ni 2p3/2 in the peak at 861.7 eV (satellite peak) was found to be Ni(III) oxyhydroxide (NiOOH), an oxidation state corresponding to Ni+3. This provides evidence for higher Ni3+ states in the MgNiO2 structure that can only proceed through Mg–O–Ni formation during the catalyst synthesis.
The spin–orbit splitting values (Δ) for Ni2+/Ni3+ are:
Δ = 872.74 − 855.13 = 17.61 eV for Ni2+ and Δ = (874.21 − 856.60) = 17.61 eV for Ni3+ |
Different forms of oxygen-containing species were observed in MgNiO2 like (–OH), surface absorbed oxygen (O2−) and surface lattice oxygen, as shown in Fig. 3c. The 1s deconvolution confirms the presence of oxygen vacancies31 in MgNiO2, modifying the electronic properties and reactivity of the material. According to the reported literature data, the binding energy range for oxygen vacancies in MgNiO2 lies between 530 and 537 eV. Peaks above 530 eV are characteristic of O2− anions in MgNiO2, whereas peaks above this range correspond to oxygen vacancies. The percentage of oxygen vacancies in MgNiO2 particles was estimated to be 22% by integrating the areas of peaks. Correlating the 22% oxygen vacancies identified in MgNiO2 through XPS with the EDS results showing 54.2 atm% of oxygen, the oxygen vacancies represent 11.9% of the total oxygen content of the MgNiO2 material. This is huge and can explain the photocatalytic activity of the material.
Fig. S2† shows the photoluminescence (PL) studies of an aqueous suspension of MgNiO2, performed at different excitation wavelengths (λex) ranging from 280 to 600 nm at room temperature. It is observed from the PL spectra at λex of 330 nm that the maximum emission wavelength (λem) is 389 nm. When the λex value increases, the λem values are red-shifted, and at λex of 400 nm, a λem value of 460 nm is observed. The presence of a single, well-defined peak in the emission spectra indicates a uniform distribution of particles within the aqueous suspension, suggesting homogeneity in their size and optical properties. Further, the pronounced and broad peaks observed in the PL spectra result from the direct recombination of electrons from the conduction band with holes in the valence band. This direct recombination process is crucial for understanding the material's photophysical behaviour and suggests efficient energy transfer mechanisms at play within the MgNiO2 system.
To further comprehend the influence of pollutant concentration on caffeine degradation, concentrations ranging from 10 to 40 ppm were investigated. Notable differences were not observed between the 10 and 20 ppm concentrations of the pollutant. However, a decrease in degradation was noted when the concentration was increased to 30 ppm and 40 ppm (Fig. 5b). This phenomenon may be attributed to the likelihood of increased pollutant concentration obstructing light penetration, thereby affecting the overall degradation process.
Fig. 5c shows the photocatalytic degradation of caffeine under different pH conditions ranging from 5 to 11. The aqueous pH (7.9) was found to be most effective for the degradation of caffeine under visible light using the synthesized catalyst MgNiO2. At this initial pH, the degradation reached 95.47%, showcasing optimal conditions for the process. As the reaction conditions varied from the initial pH, a decrease in the degradation was observed. The efficiency of caffeine degradation is primarily influenced by the degree of ionization of caffeine and the characteristics of the catalyst. Considering the weakly acidic nature of caffeine with a dissociation constant of 8.3, the initial pH proves to be more favourable compared to acidic or alkaline pH levels for effective degradation under visible light conditions.
The total organic carbon (TOC) provides an insight into the overall degradation or mineralization of organic components, including caffeine, and offers a broader perspective on the degradation process by considering changes in the overall organic content. Hence, using a UV-visible spectrophotometer, real-time monitoring of concentration variations throughout the reaction process was performed, and subsequently, the ultimate concentration of the caffeine solution post-mineralization was determined. An analysis of total organic carbon (TOC) was carried out to validate the efficacy of the mineralization process (Fig. 5d). The calculated TOC removal percentage revealed a substantial efficacy of 92.6%. The high TOC removal percentage supports the conclusion that a significant portion of the organic carbon in caffeine molecules has undergone mineralization, resulting in a cleaner and less carbon-intensive solution.
Fig. 5e illustrates the ROS-mediated photodegradation mechanism, as outlined in reactions (1)–(6). This mechanism highlights the reactive oxygen species (ROS) generated in the solution along with their associated potentials:
The ROS-mediated photodegradation mechanism is suggested next by reactions (1)–(6) noting the ROS generated in solution with their respective potentials:
[MgNiO2] + light → [MgNiO2]* → caffeine+ + (MgNiO2cbe−) | (1) |
(MgNiO2cbe−) + O2 + H+ → HO2˙ E0 − 0.05 NHE | (2) |
MgNiO2cbe− + O2ads+ → O2ads E0 0.16 NHE | (3) |
MgNiO2+ + OH−ads+ → ˙OH− E0 1.90 NHE | (4) |
MgNiO2+ + H2Oads+ → ·OHads + H+ | (5) |
O2˙− + H+ → HO2˙ | (6) |
The following section addresses the diffusion length in the solution of the most important radicals since it is a controlling feature of the performance of these radicals during caffeine photodegradation. Taking the reaction rate of the ˙OH-radical with caffeine (∼109 M−1 s−1) and the caffeine concentration of ∼10−4 M (caffeine MW 194), the ˙OH radical diffusion distance (x) away from the MgNiO2 photocatalyst surface can be estimated by taking the reaction rate between ˙OH and caffeine in aqueous solutions following the simplified Smoluchowski equation.35,36 The lifetime of the reaction pair made up of ˙OH radicals and caffeine was estimated using eqn (7):
![]() | (7) |
x2 = Dτ | (8) |
Reactions (2)–(6) allow for two comments pertaining to caffeine photodegradation: (a) that mass transfer does not play a role in the photocatalytic process as known for radical or ion-molecule reactions, and (b) no activation energy is required for ROS radical reactions.
Basically, the reduction in caffeine concentration shown in Fig. 5 consists of the degradation and discolouration of caffeine. The phenolic caffeine molecule nicotine and aliphatic acids during the caffeine photodegradation are colourless. The coffee stain contains colloidal coloured components. They consist of quaternary amines and cationic surfactants. When roasting coffee beans, functionalities are produced from the transformation of saccharides into caramel. These functionalities can also be generated from the transformation of amino acids into melanoidins and chlorogenic acids (CGA) transforming into quinolactones, melanoidins and humic acids. The phenolic acids present in green coffee beans are mostly chlorogenic acid, a pseudo-tannin, made up of low-molecular-weight hydroxycinnamic and quinic acids (implicated in the perceived acidity of coffee).39 The results reported in Fig. 5 for the photodegradation/mineralization of caffeine imply the dissociation of allylic and benzylic bonds. Single CH-bonds dissociate at 85–100 kcal mol−1. Caffeine components containing double C–H bonds need up to 174 kcal mol−1 bond dissociation energy and triple C–H bonds need a bond dissociation energy >200 kcal mol−1. The high-energy requirements needed to dissociate the multicomponent caffeine is the rationale for the slow photodegradation kinetics presented in Fig. 5d.
Scavenger effects were performed using various other scavengers (Fig. S3†), which provides equally substantial evidence to corroborate the emergence of radicals specifically ROS.40 Accordingly, diverse scavengers were employed to capture these entities. Sodium oxalate (4 mM) and sodium bicarbonate (4 mM) were introduced into the solution before the photodegradation of caffeine, serving to quench hole (h+) and hydroxyl radical (˙OH), respectively. Further, the photodegradation of caffeine was conducted under a nitrogen (N2) atmosphere to substantiate the production of oxygen free radicals (O2−˙). The outcomes revealed a significant impact on the photocatalytic efficacy of MgNiO2, which diminished from 95.4% to 86.5% with sodium oxalate, 75.4% with sodium bicarbonate, and 56.4% under N2. This attenuation in activity provides compelling evidence for generating h+, ˙OH and O2−˙ during the photocatalytic process. The higher removal percentage reduction in the caffeine degradation in the presence of N2 indicates that O2−˙ plays a more prominent role in photodegradation.
The adsorption kinetics and isotherm analysis were performed to evaluate the adsorption behavior of the material. The plots in Fig. S4(a) and (c)† show pseudo-first-order kinetic modeling at different pH and adsorbent dosages, respectively, with high correlation coefficients (R2 > 0.97), confirming that the adsorption process follows pseudo-first-order kinetics, indicating chemisorption as the rate-limiting step. The Langmuir isotherm fitting shown in Fig. S4(b) and (d)† exhibits excellent linearity (R2 = 1), suggesting monolayer adsorption on a homogeneous surface. The calculated Langmuir constants (KL) further support the strong affinity of the adsorbate towards the adsorbent surface under varying experimental conditions.
Fig. S5a and b† display the alteration in zeta potential upon incorporating Mg into NiO, indicating enhanced surface basicity. This modification affects the photocatalyst's interaction with polar pollutants and facilitates reactive oxygen species (ROS) generation, ultimately improving its photocatalytic degradation performance. The point of zero charge (PZC) is the pH at which the zeta potential (surface potential) of a material is zero, i.e., the surface carries no net electrical charge, and electrostatic interactions with charged species in solution are minimized. Fig. S5b and c† display the PZC for MgNiO2 and NiO, respectively.
The PZC for a material is defined as the pH at which the net charge of the total particle surface is equal to zero (i.e., concentration of H+ is equal to OH−).41,42 From the obtained results, PZC for MgNiO2 was around 7.9–8.0, and it was 7.4 for NiO. This means that for MgNiO2, its surface shows a net positive charge at pH values less than 7.8 and a negative charge at pH values greater than 7.8. In the case of pure nickel oxide (PZC-7.4), its surface carries a net positive charge at pH values below 7.4 and a net negative charge at pH values above 7.4.
The incorporation of Mg into the NiO lattice increases the PZC from approximately 7.4 to around 7.9–8.0. This shift in the PZC indicates an increase in surface basicity due to the presence of Mg2+ ions (because of the more electropositive nature of Mg2+), which enhance the surface's affinity for hydroxyl (–OH) group adsorption.
At pH < PZC (7.9–8.0), MgNiO2 exhibits a positive net surface charge as the surrounding solution has an excess of H+ ions undergoing surface protonation. OH2+ groups give the surface a positive charge, which is detected in zeta potential measurements, increasing electrostatic attraction for negatively charged or polar contaminants such as caffeine at low pH.43,44
As a result, the modified surface (now functionalized with –OH groups) exhibits enhanced electrostatic interactions with the pollutant, which is a polar molecule, directly influencing adsorption-driven photocatalytic performance under varying pH conditions. Caffeine, a polar molecule with functional groups such as carbonyl CO and amide-like N–H groups, forms hydrogen bond interactions with surface hydroxyl (–OH) groups on MgNiO2. Surface hydroxylation plays a critical role in photocatalytic pollutant degradation by enhancing the adsorption of polar contaminants containing N, O, and F elements through hydrogen bonding.45
At pH > PZC, MgNiO2 shows a negative net surface charge because of an excess of OH− ions in solution. These OH− ions promote the deprotonation of surface –OH groups to give –O−. Therefore, caffeine adsorption reduces due to repulsion, but ROS generation from –OH groups still continues upon light irradiation. The incorporation of Mg2+ creates oxygen vacancies or defect sites due to the size mismatch between Ni and Mg. Mg has a lower electronegativity (1.31) than Ni (1.88). Therefore, MgO bonds are more ionic and weaker than NiO bonds, which triggers the lattice to release oxygen and form vacancies, enhancing surface reactivity and photocatalytic performance.46 These vacancies, along with surface hydroxylation, enhance the MgNiO2 ability in comparison with NiO for ROS generation, like HO2˙ and 1O2, by providing more reactive hydroxyl groups for photocatalytic reactions.47 These defects act as trap sites for photogenerated electrons or holes, reducing their chances of recombination. Mg introduces defects that promote charge separation by trapping photogenerated carriers, extending their lifetime for catalytic reactions.48
The presence of these hydroxyl radicals was further confirmed by the ESR studies presented in Fig. S5c.† The multiple well-resolved peaks observed in the spectrum are due to the interaction of the unpaired electron with nearby nuclei, confirming the presence of hydroxyl radicals in the system.
Next, the main MgNiO2 surface parameters intervening in the caffeine photodegradation under visible light have been summarized. The interaction of caffeine with MgNiO2 occurs through the OH-surface attached groups on MgNiO2. The hydrophilic surface of MgNiO2 is covered by −OH groups present in diverse configurations on Mg and Ni. These −OH groups assume different forms, most of them being covalent terminals, ionic terminals and ionic bridging and found on the MgNiO2 bonding sites. A parameter controlling the density of –OH groups attached to MgNiO2 is the surface potential of MgNiO2. The two last points have not been investigated since they are beyond the scope of the present study. Full surface hydroxylation is noted as MgNiO2 + 2˙OHads → Mg(OH)2 + Ni(OH)2. MgNiO2 surface charges undergo surface migration leading to redox reactions under light irradiation. This reaction can be noted as Mg+2Ni+2O2−2 → Mg+2(O−2H+1)2 + Ni+2(O−2H+1)2.
The stability and reusability of the catalyst were examined by subjecting the used catalyst to multiple degradation cycles of caffeine under identical reaction parameters. The catalyst was initially collected, thoroughly washed with distilled water and then dried in an oven at 60 °C overnight. Subsequently, reusability studies were conducted. Remarkably, after repeated use of the catalyst for four cycles, the degradation decreased by only 4.4% (91.02%). This result highlights the stability of the catalyst and its ability to be reused multiple times without significant loss of efficacy. A comparative study of the catalyst used in this work with other catalysts reported in the literature is presented in Table 1.
Process initial | Catalyst (g L−1) | Catalyst concentration | Pollutant concentration | Removal efficiency | Time (min) | Ref. |
---|---|---|---|---|---|---|
UV | Co-MCM41 | 200 mg L−1 | 0.05 mM | 98.00% | 100 | 49 |
UV mercury lamp (400 W) | ZnO–ZnAl2O4 | 0.3 g L−1 | 20 mg L−1 | 97.32% | 60 | 50 |
UV mercury lamp (400 W) | Mg–ZnO–Al2O3 | 0.6 g L−1 | 20 mg L−1 | 98.9% | 70 | 51 |
Mercury vapor lamp (125 W) | Ag–ZnO | 0.3 g | 20 mg L−1 | 70% | 300 | 52 |
UV-C lamp | BiVO4/rGO | Electrode (20.3 cm2) | 10 μmol L−1 | 94.2% | 480 | 53 |
Visible light | PVA–CuO | 30 mg | 20 mg L−1 | 99.03% | 80 | 6 |
White LED light (50 W) | MgNiO2 | 20 mg | 20 mg L−1 | 95.4% | 180 | This work |
η = b![]() ![]() | (9) |
M + OH− → MOHadsorption + e− | (10) |
MOH + OH− → H2O + MO + e− | (11) |
MO + OH− → MOOH + e− | (12) |
MOOH + OH− → H2O + MOO + e− | (13) |
MOO → O2 + M | (14) |
![]() | ||
Fig. 6 Electrocatalytic OER studies of the as-synthesized MgNiO2 catalyst and bare Ni foam (NF). (a and b) LSV overpotential data vs. RHE and (c) Tafel slopes. |
The Nyquist plot shown in Fig. S6a† compares the charge transfer resistance of NiO and MgNiO2 catalysts. The MgNiO2 sample exhibits a smaller semicircle diameter than that of NiO, indicating a lower charge transfer resistance and improved electrical conductivity. This suggests that magnesium incorporation enhances the electron transport properties of NiO, facilitating faster charge transfer kinetics during the electrochemical reaction, which is beneficial for efficient electrocatalytic performance.
The cycling stability studies were conducted for a period of 22 h, and the same is presented in Fig. S6b.† This chronoamperometric curve demonstrates that the MgNiO2 catalyst exhibits excellent long-term stability for the OER, with only a very slight increase in potential over 22 hours of continuous operation, confirming its durability and suitability for practical electrochemical applications. In addition, post-cycling characterization studies were also carried out to understand the stability or reusability of the material. HRTEM and XPS studies were used to analyse the material, and the same is provided in Fig. S7 and S8,† respectively.
The TEM analysis after electrolysis reveals that the MgNiO2 nanoparticles retain their original granular morphology, indicating excellent structural stability during electrochemical operation. The bright-field TEM image in Fig. S7(a)† and the HAADF-STEM image in Fig. S7(b)† show uniformly dispersed nanoparticles without significant aggregation or structural collapse. The elemental mapping shown in Fig. S7(c–e)† confirms the homogeneous distribution of Mg, Ni and O throughout the sample, demonstrating that the compositional integrity of the material is well preserved even after electrolysis. The SAED pattern (f) displays distinct diffraction rings, suggesting that the crystalline nature of the MgNiO2 nanoparticles remains intact after electrochemical cycling. Overall, these results confirm the robustness of MgNiO2 nanoparticles under electrochemical conditions, maintaining both their granular structure and crystalline phase after electrolysis.
The XPS analysis after electrolysis confirms the chemical stability and retention of key surface states in the MgNiO2 nanoparticles (Fig. S8†). In the Mg 1s spectrum (a), the presence of Mg–O (1301.2 eV), Mg(OH)2 (1302.4 eV) and metallic Mg (1303.4 eV) indicates that magnesium remains largely in its original chemical environment with minor surface hydroxylation after electrolysis. The Ni 2p spectrum (b) shows characteristic peaks corresponding to Ni(OH)2 (853.7 eV), NiO (855.4 eV) and NiOOH (861 eV), suggesting partial surface oxidation during electrolysis, probably due to electrochemical cycling, while the overall nickel oxidation states remain well-defined. The O 1s spectrum (c) shows dominant peaks assigned to lattice oxygen (O2−) and surface hydroxyl species, further confirming the preservation of the oxide framework. These results demonstrate that the MgNiO2 nanoparticles maintain their chemical integrity and surface composition after electrolysis, supporting their stability and suitability for repeated electrochemical operation.
Footnote |
† Electronic supplementary information (ESI) available. See DOI: https://doi.org/10.1039/d5lf00102a |
This journal is © The Royal Society of Chemistry 2025 |