Anshu
Shrivastava
,
Uttam
Kumar
and
Indrajit
Sinha
*
Department of Chemistry, Indian Institute of Technology (Banaras Hindu University), Varanasi 221005, India. E-mail: isinha.apc@iitbhu.ac.in
First published on 28th April 2025
The poor photo-Fenton properties of NiFe2O4, a magnetic material, can be altered by appropriate metal ion doping. We carried out preliminary DFT and TD-DFT calculations and found that the photoexcited molybdenum doped NiFe2O4 interacts effectively with H2O2 for hydroxyl radical generation. Based on this prediction, we prepared Mo-doped NiFe2O4 and tested its photo-Fenton activity for tetracycline (TC) degradation. The Mo-doped NiFe2O4 nanoparticles were significantly finer in size and superparamagnetic, enabling easy after-use separation. Experimental and DFT calculation results showed that the Mo-dopant substitutes the Fe3+ occupying the octahedral site in NiFe2O4. Mo-doping extended the absorption edge and decreased the interfacial charge transfer resistance of NiFe2O4. The Mo-doped NiFe2O4 nanoparticles demonstrated excellent TC photo-Fenton degradation activity, along with appropriate recyclability. A combination of experimental and DFT calculation results indicated the possible photo-Fenton mechanism.
In contrast, the photo-driven heterogeneous Fenton reaction, relying on efficient hydroxyl radicals (˙OH) production from H2O2, is a green method favored for chemically degrading a wide range of non-biodegradable pollutants to their non-toxic fragments in the presence of iron-based photocatalysts.7 Light of suitable wavelength photo-excites the semiconductor photocatalyst. The photo-excited electrons in the photocatalyst's conduction band (CB) reduce H2O2 to produce ˙OH radicals, while oxidation occurs at the holes in the valence band (VB).8 The separate oxidation and reduction sites in photo-Fenton photocatalysts enhance efficiency by overcoming the rate-limiting issues of conventional Fenton processes. Moreover, the process offers high catalytic efficiency, facilitated by the robust regeneration of Fe2+ by photoexcited electrons (Fe2+ ⇌ Fe3+), ensuring continuous redox cycling. Additionally, it demonstrates broader pH tolerance, with simple operational requirements, mild reaction conditions, and overall higher efficiency.
The spinel NiFe2O4 semiconductor is a prospective candidate for the photo-Fenton degradation of organic pollutants due to its moderate visible light band gap (1.9–2.2 eV) and good photochemical stability.9 Crucially, the superparamagnetic behavior of NiFe2O4 nanoparticles allows their easy separation from aqueous solutions and re-dispersal by simple application and removal of the applied magnetic field. Nonetheless, pristine NiFe2O4 characteristically exhibits low surface reaction rates, rapid recombination of photogenerated electrons and holes, and poor visible light absorption.10 Various strategies have been employed to address these shortcomings and enhance its catalytic efficiency. A promising approach is metal ion doping to alter the physical and chemical properties of spinel NiFe2O4. Typically, the dopant level introduced in the band gap enhances charge separation and slows the recombination process. There are a few instances of the metal-doped NiFe2O4 photocatalysis. Thus, Alzahrani et al. reported the synthesis of Nd-doped NiFe2O4 and Gd-doped NiFe2O4 photocatalysts for photocatalytic degradation of organic effluents. They observed the superior photocatalytic activity of doped NiFe2O4 samples than bare NiFe2O4 due to reduction in the recombination of electrons and holes.11,12 Similarly, enhancement in the photocatalytic activity of Co-doped NiFe2O4 and Mg-doped NiFe2O4 were ascribed to dopant energy states slowing the recombination of electrons and holes.13,14
In this regard, molybdenum (Mo) is a particularly suitable dopant, as it can modify the electronic properties (carrier concentration or CB position) and reduce the recombination of photogenerated charge carriers. Mo exists in multiple oxidation states (+4, +5, and +6), and the dopant energy level in the NiFe2O4 band gap can improve the iron cycle (Fe2+ ⇌ Fe3+) in Fenton catalysis, enhancing catalytic efficiency. For instance, Ying et al. found that the TiO2 doped by Mo6+ extends the absorption edge and decreases the interfacial charge transfer resistance. The photocatalytic activity of Mo-doped TiO2 was much better than pure TiO2.15 Likewise, in a recent report, authors have shown that Mo-doped FeS2 demonstrates improved stability and photo-Fenton activity.16 Considering the advantages mentioned and the variable oxidation states of Mo, which are crucial for the Fenton reaction, this study explores the photo-Fenton activity of Mo-doped NiFe2O4. Notably, there has been no investigation into the design and effects of Mo-doping in spinel NiFe2O4 on its photo-Fenton TC degradation activity, making this investigation a novel contribution to the field.
Previous studies,17–19 combining experimental and computational approaches, indicate that for a photo-Fenton catalyst to be effective, its photo-excited CB must interact favourably with H2O2, to facilitate ˙OH radical production. Guided by this principle, in the present work, we employed density functional theory (DFT) and time-dependent DFT (TD-DFT) calculations to qualitatively evaluate whether the Mo doped NiFe2O4 satisfies this criterion, prior to experimental investigations. The following description highlights the novelty of this research work. There is no previous investigation on photocatalytic properties of Mo-doped NiFe2O4. The Mo-dopant can act as a donor dopant level and a shallow trap in the NiFe2O4 band gap preventing photoexcited electron–hole recombination, improving charge separation and extended carrier lifetimes. Moreover, favourable Mo6+/Mo5+ redox potential also implies the possibility of increased Fe3+ to Fe2+ back reaction rate. Critically, we use DFT and TD-DFT calculations to predict that Mo doping results in charge redistribution, altered electronic transitions and facilitates efficient ˙OH generation through enhanced H2O2 activation. Hence, such doping can substantially enhance the photocatalytic activity of Mo-doped NiFe2O4. The improved activity, along with the superparamagnetic properties of the doped NiFe2O4 nanoparticles (enabling easy recovery and redispersal of the nanoparticles for reuse), could result in an industrially attractive photocatalyst.
Satisfied by the DFT calculation results and the possibility of enhanced photocatalytic activity of this novel doped nanomaterial, we synthesized Mo-doped NiFe2O4 nanoparticles by a one-step hydrothermal protocol. Comprehensive materials characterizations were conducted to examine the physicochemical properties, microstructure, chemical composition, optical magnetic properties of the prepared Mo-doped NiFe2O4 catalysts. Mott–Schottky and electrochemical impedance spectroscopy (EIS) were used to calculate the flat band potential, semiconductor type, and charge transfer resistance. The effects of Mo-dopant amount, H2O2 concentration, and pH on TC degradation were systematically investigated. Under visible light irradiation, the synthesized Mo-doped NiFe2O4 nanoparticles demonstrated enhanced photo-Fenton degradation of tetracycline. Scavenger experiments identified the key reactive species responsible for the photocatalytic activity. Experimental and DFT calculation results were combined to understand the photocatalytic mechanism in this photo-Fenton process.
The above-mentioned synthesis procedure outlines the synthesis of 1% Mo doped NiFe2O4 nanoparticles. The 2% and 4% Mo-doped NiFe2O4 nanoparticles were synthesised in a similar manner, with the only difference being the amount of (NH4)6Mo7O24·4H2O utilised in the synthesis. The synthesized Mo-doped NiFe2O4 samples were abbreviated as 1MNFO, 2MNFO, and 4MNFO in accordance with the increasing at% corresponding to 1 at%, 2 at%, and 4 at% Mo-dopant in NiFe2O4, respectively. Pure NiFe2O4 was prepared by an analogues protocol but without the addition of (NH4)6Mo7O24·4H2O salt.
Several control experiments were carried out under various conditions. For instance, experiments were conducted at different pH values to determine the optimal reaction pH. The effect of H2O2 dosage on the photo-Fenton activity was also investigated by evaluating the photo-Fenton activity of Mo-doped NiFe2O4 sample at various H2O2 concentrations in the reaction mixture. Finally, the best photocatalyst with the highest photo-Fenton activity among the Mo-doped NiFe2O4 samples was assessed under optimal pH and H2O2 conditions.
Separate scavenger experiments were carried out to investigate OH˙ radical, photogenerated electrons/holes, and superoxide anion radical generation in the photo-Fenton degradation process. Isopropyl alcohol (IPA), AgNO3, and triethanolamine (TEOA) were used to capture OH˙, electrons, and holes, respectively. For the active species determination experiment, 100 μL of 0.185 mM IPA or 20 μL of 0.5 M TEOA, or 10 μL of 0.82 mM AgNO3 were added to the reaction mixture. The rest of the experiment remained the same. On the other hand, the nitro blue tetrazolium (NBT) test detected the superoxide radical anion formation (O2˙−). All the photo Fenton degradation experiments were conducted at ∼27 °C temperature. The degradation efficiency (%) of each photocatalyst was calculated by eqn (1),
![]() | (1) |
In eqn (1), C0 and Ce represent the initial and equilibrium concentrations of TC in the reaction mixture (mg L−1).
Ef = Edefect(N1) − (Eperfect(NiFe2O4) − μFe + μMo) | (2) |
Ef = Edefect(N2) − (Eperfect(NiFe2O4) − μFe + μMo) | (3) |
Ef = Edefect(N3) − (Eperfect(NiFe2O4) − μNi + μMo) | (4) |
Ef = Edefect(N4) − (Eperfect(NiFe2O4) + μMo) | (5) |
In eqn (2)–(5)Ef is the defect formation energy, Eperfect represents the energy of N0 model, Edefect depicts the energy of the Mo doped optimized model; μNi, μFe, and μMo are the chemical potentials of Ni, Fe, and Mo atom i.e. the energy per atom. The unit cell of Ni (COD #9012985), Fe (COD #9015972), Mo (COD #9008474) was taken from Crystallography Open Database (COD). The geometry of the optimized model having most negative defect formation energy gives the location of Mo dopant in NiFe2O4 lattice.
The optimized Mo-doped NiFe2O4 model with the most negative formation energy was subjected to further processing. A cluster of Mo-doped NiFe2O4 was cleaved from a suitable supercell of this model using the MAPS 4.4.1 software (released by Scienomics).22 The unsaturated terminal oxygens of this cluster were passivated by hydrogens in the GaussView 6.0 software. In this cluster (having total 33 atoms), Ni occupies an octahedral site (abbreviated as Ni(Oct)), Mo occupies another octahedral site (denoted as Mo(Oct)), one Fe occupies an octahedral site (abbreviated as Fe(Oct)) and another Fe is in a tetrahedral site (denoted by Fe(Td)). Since, the cluster has been carved out of an optimized supercell, therefore, a single point DFT calculation was performed on this cluster using the DGDZVP basis set with the B3LYP (Becke-3-parameter-Lee–Yang–Parr) functional in the Gaussian 16 software.23 This Mo-doped NiFe2O4 cluster model is denoted by the abbreviation ‘MoNIF’ cluster throughout this research work.
In the next step, the H2O2 molecule was built by GaussView 6.0 software. This molecule was optimized by Gaussian 16 software at the same level of theory as earlier. Fig. S4† shows the optimized structure of H2O2 molecule. DFT calculations were conducted to investigate the interaction of H2O2 with the MoNIF cluster in the ground state. All DFT calculations were performed at the same level of theory as discussed earlier. Natural bond orbital (NBO) analysis gave the charge transfer/delocalization due to the H2O2-MoNIF interaction. All calculations were performed in the gas phase.
Fig. 1a and b present the powder XRD patterns of pure NiFe2O4 alongside the Mo-doped samples, i.e., 1MNFO, 2MNFO, and 4MNFO nanoparticles. The XRD pattern for pure NiFe2O4 exhibits distinct diffraction peaks at 18.4°, 30.2°, 35.7°, 37.3°, 43.3°, 53.8°, 57.2°, 62.9°, and 74.4° (2θ) values. These correspond to the (111), (220), (311), (222), (400), (422), (511), (440), and (533) crystal planes of the inverse spinel NiFe2O4 structure (JCPDS file 74-2081).24 Similarly, the XRD patterns of 1MNFO, 2MNFO, and 4MNFO match those of NiFe2O4 (JCPDS file 74-2081). No additional peaks were detected, confirming the incorporation of Mo into the NiFe2O4 lattice.
Notably, the XRD patterns of Mo-doped samples exhibit a slight shift to lower 2θ values, indicating an increase in d-spacing (interplanar distance) as the Mo content rises (Fig. 1b). Table 1 gives the magnitude of lattice expansion (details of lattice parameter calculation given in ESI†). The expansion is due to the dopant's lattice position in the NiFe2O4 inverse spinel structure. To understand this, we compare the ionic radii of the metal ions in NiFe2O4, along with Mo in its +5 and +6 oxidation states in Table S1.†25 The XPS results later show that the dopant Mo in the doped NiFe2O4 samples exist in +5 and +6 oxidation states. For this reason, Table S1† also displays the ionic radii of Mo in its +5 and +6 oxidation states. Mo doping into the NiFe2O4 structure could occur in two ways: if Mo substitutes a larger cation, such as Ni2+, in the octahedral position, lattice contraction would be expected. Conversely, lattice expansion should occur if Mo replaces a smaller cation like Fe3+ in an octahedral position. Thus, lattice expansion implies Mo could be substituting Fe3+ in the octahedral sites of the NiFe2O4 lattice.
Nanoparticle | Crystallite size (nm) | Lattice parameter (Å) |
---|---|---|
Section S3† provides the equation's details used for calculation of structural parameters. | ||
Pure NiFe2O4 | 42.87 | 8.343 |
1MNFO | 27.02 | 8.344 |
2MNFO | 17.34 | 8.347 |
4MNFO | 28.05 | 8.358 |
Column 2 of Table 1 shows the crystallite sizes of both doped and un-doped NiFe2O4 samples. The crystallite size decreases from 42.87 nm to 17.34 nm as the Mo-dopant increases to 2% (2MNFO). However, Mo doping beyond 2% increases the crystallite size. This is possibly due to the dopant limiting crystallite growth by slowing the nucleation rate during crystallization (Table 1).26
The morphology and microstructure of pristine NiFe2O4 and the 1MNFO samples were examined using TEM and SEM techniques. TEM images (Fig. 2a) reveal that pure NiFe2O4 particles exhibit a combination of octahedral (Oct) and nanosheet (NS) like morphologies. The high-resolution TEM (HR-TEM) analysis (Fig. 2b) shows a lattice d-spacing of 0.294 nm, corresponding to the (220) crystal plane of FCC NiFe2O4. This observation conforms to the XRD results, further confirming the crystalline structure of NiFe2O4 (JCPDS no.-74-2081). The d-spacing between adjacent lattice planes was measured using Digital Micrograph (Gatan) software. Similarly, Fig. 2c illustrates the morphology of 1MNFO particles, which retain an approximately octahedral shape even after Mo-doping. The HR-TEM image of 1MNFO (Fig. 2d) shows lattice fringes with a d-spacing of 0.309 nm, consistent with the (220) crystal plane. This slight increase in d-spacing from 0.294 nm to 0.309 nm aligns with the XRD results, confirming the successful incorporation of molybdenum into the NiFe2O4 nanoparticles. The particle size distribution of NiFe2O4 and 1MNFO samples have been calculated and given in the Fig. S5 of ESI.† Mo doping of NiFe2O4 decreases the average particle size from 75–95 nm (NiFe2O4) to 25–45 nm (1MNFO). The polydispersity index (PDI) was determined using eqn (S3) and (S4), provided in the section S4 of ESI.† Before Mo doping, pristine NiFe2O4 exhibited a PDI of 0.099, indicating a monodisperse nanoparticle distribution. In contrast, the 1MNFO sample displayed a PDI of 0.126, suggesting a transition to a moderately polydisperse nanoparticle system.
Fig. 3a and b show a SEM micrograph of the as-synthesized 1MNFO nanoparticles and its elemental mapping. The latter shows that the Mo-dopant is homogeneously distributed in the 1MNFO sample (Fig. 3b).
The VSM analysis gives the magnetic characteristics of the photocatalyst. Fig. 4 shows the VSM analysis for pure NiFe2O4 and 1MNFO by hysteresis loop plotted between magnetization (emu g−1) versus applied magnetic field (Oe). The saturation magnetization for pure NiFe2O4 is 39.70 emu g−1. After Mo doping, the saturation magnetization of 1MNFO was reduced to 20.30 emu g−1. The absence of a hysteresis loop indicates that the pure and doped sample is superparamagnetic. The magnetization of both samples is sufficient for their separation from water by a magnet.
The XPS of pure NiFe2O4 and 1MNFO samples were carried out to determine the chemical composition of the photocatalyst surface and the oxidation states of the elements present. Fig. 5a displays the Ni 2p spectrum of NiFe2O4 and 1MNFO samples. The NiFe2O4 spectrum has Ni 2p3/2, and Ni 2p1/2 spin orbit doublets at 854.97 eV and 872.58 eV binding energies (B.E.) with two shake-up satellite peaks (denoted as ‘sat.’ in Fig. 5a), indicating Ni2+ occupies octahedral sites.19 The fitted Ni 2p core-level spectrum of 1MNFO (Fig. 5a) also exhibits two shake-up satellites and spin–orbit doublets (Ni 2p3/2, Ni 2p1/2). The Ni 2p3/2 and Ni 2p1/2 spin–orbit doublets at 855.39 and 873.02 eV show that Ni2+ occupies octahedral sites in the 1MNFO crystal structure.19
![]() | ||
Fig. 5 High resolution comparative XPS spectra of, (a) Ni 2p, (b) Fe 2p, (c) O 1s, and (d) Mo 3d in pure NiFe2O4 and 1MNFO. |
Fig. 5b compares the Fe 2p regions of bare NiFe2O4 and 1MNFO samples. The Fe 2p spectrum has two spin–orbit doublets (Fe 2p3/2, Fe 2p1/2) and two satellite peaks (denoted as ‘Sat.’ in Fig. 5b). Further deconvolution gave peaks at 709.98, 712.14, and 723.65, 726.28 eV, with two satellite peaks at 718.46 and 732.66 eV. These show that Fe3+ occupies octahedral and tetrahedral sites of the NiFe2O4 lattice.27–30 The Fe 2p spectrum of 1MNFO also displays the Fe 2p3/2 and Fe 2p1/2 spin–orbit doublets. The Fe 2p3/2 part is deconvoluted into 710.10 and 712.30 eV, while the Fe 2p1/2 part fits peaks at 723.69 and 726.59 eV. These peaks, in combination with two satellite peaks at 718.59 and 732.51 eV binding energies, show that Fe3+ occupies octahedral and tetrahedral sites of the 1MNFO sample.27,30 A comparison of the spectrum of NiFe2O4 and 1MNFO reveals that Mo-doping shifts the deconvoluted Fe2p peaks to slightly higher binding energies.
Fig. 5c presents the O 1s spectra of pure NiFe2O4 and 1MNFO samples. For pure NiFe2O4, three distinct deconvoluted peaks are observed at binding energies of 529.73, 531.0, and 532.18 eV, corresponding to metal–oxygen bonds (Ni–O or Fe–O),31,32 oxygen vacancies,33,34 and chemisorbed oxygen in –OH groups,34,35 respectively (Fig. 5c). Upon Mo doping, the lattice oxygen deconvoluted O 1s peak shifts to a slightly lower 529.71 eV binding energy.31,32 The other two O 1s peaks shift to slightly higher 531.27 and 532.34 eV binding energies. These are also attributed to oxygen vacancies33,34 and chemisorbed oxygen34,35 species, respectively (Fig. 5c). However, the area of oxygen vacancy peak and chemisorbed oxygen peak of pure NiFe2O4 decreased after Mo doping. This indicates that Mo dopants utilized chemisorbed oxygen to compensate for the oxygen vacancy and maintain charge balance in the lattice.
Fig. 5d displays the Mo 3d fitted spectrum region. Two major peaks are observed at 232.25 (Mo 3d5/2) and 235.18 eV (Mo 3d3/2) binding energies. Each of these peaks could be deconvoluted into two more peaks. The peaks at 231.78 and 235.06 eV indicate the presence of Mo in the +5 oxidation state, while the peaks at 232.35 and 235.76 eV correspond to Mo in the +6 state.36–39 The higher valent Mo (in +6 oxidation state) dopant substitutes a Fe site in the NiFe2O4 lattice, causing oxygen deficiency. Oxygen vacancies are electron-rich, enabling the reduction of the higher-valent Mo6+ ions to Mo5+ ions.38–40 Since higher-valent dopants require additional oxygen to maintain charge balance within the lattice, they tend to utilize chemisorbed oxygen to compensate for the oxygen deficiency. Consequently, this process leads to a reduction in the overall percentage of oxygen vacancies in the lattice.
Overall, Mo doping causes a slight shift in the Ni 2p and Fe 2p peaks towards higher binding energies. Concurrently, the lattice oxygen peak shifts to lower binding energy. These observations suggest that Mo doping facilitates charge transfer from the Ni and Fe sites to the lattice oxygen.
Fig. S6(a–d)† shows the band gap determination of pure NiFe2O4 and different Mo-doped NiFe2O4 samples, calculated by the Tauc equation. The graph was plotted between (αhν)1/nversus photon energy (hν) to determine the band gap (where n = 1/2 for direct band). The symbols ‘α’, ‘h’, and ‘ν’ are referred to as absorption coefficient, Planck's constant, and photon's frequency in the Tauc relation. For direct transition, n equals 1/2, and 2 for indirect transitions.41 The band gap (Eg) decreases from 2.014 (NiFe2O4) to 1.939 eV (2MNFO) after Mo-doping. The band gap of Mo-doped NiFe2O4 initially decreases and then increases with increasing Mo doping percentage after 2MNFO.
Electrochemical impedance spectroscopy (EIS) observations are related to the kinetics of charge carriers’ recombination. Fig. 6a and b show the Nyquist plot of pure NiFe2O4, 1MNFO, 2MNFO, and 4MNFO samples. The diameter of semicircle of the Nyquist plot is directly proportional to the charge transfer resistance and inversely proportional to charge separation.42 Among all samples, the Nyquist plot of 1MNFO had least semicircle diameter, indicating best charge separation during reaction.
Mott Schottky (MS) data were acquired at 500 Hz to ascertain the semiconductor type and calculate the band edge positions of NiFe2O4 and 1MNFO semiconductors. The equation V(NHE) = V(Ag/AgCl) + 0.059 pH + 0.197 was used to convert the potential (V) to the normal hydrogen electrode (NHE) scale for the MS plots.43 The pH of the electrolyte was approximately 6.8. Fig. 7a and b illustrate the MS plot of the pure NiFe2O4 and 1MNFO particles. Their MS plots show a positive slope, so pure NiFe2O4 and 1MNFO are n-type semiconductors. The pure NiFe2O4 and 1MNFO have CB edge potentials of −0.71 and −0.43 V, respectively. Equation Eg = EVB − ECB was utilized to determine the VB edges of pure NiFe2O4 and 1MNFO using these CB edge values. The calculated VB edge is 1.31 V for pure NiFe2O4 and 1.535 V for the 1MNFO sample.
The photo-Fenton activities of Mo-doped and undoped NiFe2O4 for TC degradation have been investigated. Photo Fenton activities of the 1MNFO doped sample were studied at pH ∼1, 3 and 6 (near neutral). Fig. S7c† shows that the best photo Fenton activity of 1MNFO was observed at pH 3. Another set of experiments were conducted to optimize H2O2 amount in the photocatalysis reaction mixture at pH 3 (Fig. S7d†). The volume of H2O2 used in the experiments were 50 μl, 100 μl, and 150 μl of 0.5M H2O2. The best photo-Fenton activity on the 1MNFO sample was with 100 μl of H2O2 amount. Nearly, ∼99% of photo-Fenton degradation of TC by 1MNFO was completed in 130 min (Fig. S7d†). Only 60% of TC degradation was observed when 150 μl of 0.5 M H2O2 was used. This reduction occurs because excess (unreacted) H2O2 acts as a hydroxyl radical (˙OH) scavenger and generate hydroperoxyl radical (˙HO2) (shown in eqn (6) and (7)).44 This ˙HO2 radical is a considerably weaker oxidising agent than ˙OH radical.44
H2O2 + ˙OH → ˙HO2 + H2O | (6) |
˙OH + ˙HO2 → H2O + O2 | (7) |
Fig. 8a displays the concentration versus time plots of TC degradation under optimum photo Fenton conditions on different Mo-doped and undoped NiFe2O4 samples. Among all photocatalysts, 1MNFO shows the best TC degradation photo-Fenton activity and undoped NiFe2O4 sample shows poor photo-Fenton activity. Fig. S7e† shows the UV-visible absorbance spectra of photo-Fenton degradation of TC by 1MNFO photocatalyst at pH 3. Within 130 min, ∼99% of TC was degraded under visible light irradiation. Fig. 8b shows that the TC degradation concentration versus time data fits best pseudo-first-order kinetics (0.988 < R2 < 0.999). Table S2† compares the photo-Fenton activity of our photocatalyst for tetracycline degradation with previous research work. The activity of the Mo-doped NiFe2O4 is among best reported in literature for doped semiconductors.
Separate active species trapping tests were conducted with IPA, AgNO3, and TEOA. IPA is a hydroxyl radical scavenger, AgNO3 is an electron scavenger, and TEOA serves as a hole scavanger.45 The photo-Fenton reaction on the 1MNFO catalyst almost completely degraded tetracycline in 130 min without any scavenging agent. In comparison, the photo-Fenton TC degradation efficiencies on the 1MNFO sample with either IPA, or AgNO3, or TEOA were ∼15.9%, ∼14.6%, and 95.5% (Fig. S7f†) respectively. This demonstrates that hydroxyl radicals and photogenerated electrons are the predominant active species for photo Fenton TC degradation on the 1MNFO sample.
In addition to the above, the Nitro blue tetrazolium (NBT) test was used to detect the existence of superoxide radicals in the aqueous solution of 1MNFO photocatalyst.46 Following the standard procedure, 3 ml of a 2.5 × 10−2 mM NBT solution was placed in a quartz cuvette, and 100 microliters of the dispersed photocatalyst nanoparticles (1 mg ml−1) were added under (cool white LED) visible light irradiation. The supernatant was examined by UV-visible absorbance spectroscopy after the photocatalyst was separated using a magnet after fixed time intervals (Fig. S7g†). No change in the NBT molecule's absorption maximum at ∼265 nm with time confirms that the superoxide radical (˙O2−) generation is negligible and is not an active species in TC degradation.
The recyclability test of 1MNFO was conducted for five consecutive cycles of photo-Fenton degradation of TC. The detailed experimental procedure is provided in section S5 of the ESI.†Fig. 8c illustrates the reusability of 1MNFO via consecutive experiments (5 cycles). The photo-Fenton activity of 1MNFO was ∼91.5% after 5 cycles. Fig. 8d compares the XRD patterns of fresh and recycled 1MNFO samples. All XRD peaks of recycled 1MNFO match those of the unused fresh 1MNFO sample. The 1MNFO photocatalyst shows good photostability after five photocatalytic cycles.
The leaching concentration of Ni, Fe, and Mo metals of the recycled 1MNFO sample in the solution after the 5th photo Fenton degradation cycle was investigated using inductively coupled plasma-mass spectrometry (ICP-MS). The percentage concentrations of leached metal ions in the solution were Ni (0.98%), Fe (0.45%), and Mo (0.008%), respectively. The minor leaching of metal ions during the reuse of the 1MNFO sample results in a decrease in photo-Fenton activity of 1MNFO after five cycles.
XPS analysis was also done on the recycled 1MNFO sample (Fig. S8†). Fig. S8a† displays two spin–orbit doublet peaks at 854.26 (Ni 2p3/2) and 872.0 eV (Ni 2p1/2) binding energies, indicating Ni2+ in octahedral sites. Similarly, Fig. S8b† shows two spin–orbit doublet peaks (Fe 2p3/2 and Fe 2p1/2) and two satellite peaks in the fitted Fe 2p XPS spectrum. Further deconvolution confirmed the presence of Fe3+ in both octahedral and tetrahedral sites. Fig. S8c† represents the O 1s spectrum region. The three deconvoluted peaks obtained were similar to the fresh 1MNFO sample. These peaks indicate the presence of lattice oxygen (or metal–oxygen), oxygen vacancy, and chemisorbed oxygen species. A comparison of the XPS spectra of fresh and recycled 1MNFO samples (Fig. 5 and Fig. S8†) shows no significant change in Ni 2p, Fe 2p, and O 1s spectrum regions. Fig. S8d† shows deconvoluted Mo 3d5/2 peaks at 230.15 eV, 231.35 eV, and 232.50 eV due to Mo in the +4, +5, and +6 oxidation states. The Mo 3d3/2 peaks at 232.99 eV, 233.50 eV, and 234.70 eV are also due to Mo in the +4, +5, and +6 states in the recycled 1MNFO sample. Notably, the generation of Mo4+ after the fifth consecutive cycle of photo-Fenton TC degradation is likely due to the reduction of Mo6+/Mo5+ by photogenerated electrons and oxygen vacancies.
HOMO and LUMO locations give us the nucleophilic and electrophilic sites on the photocatalyst. In this context, ground state HOMO and LUMO positions indicate the nucleophilic and electrophilic sites before photoexcitation, while the excited state HOMO and LUMO positions on the photocatalyst give the nucleophilic and electrophilic positions after its photoexcitation. Fig. 9a displays the ground state HOMO LUMO locations on MoNIF cluster. The ground state HOMO is mainly on Ni(Oct) and Fe(Oct), while the LUMO is located partially on Fe(Oct), Fe(Td) and Mo(Oct) in the MoNIF cluster.
![]() | ||
Fig. 9 The calculated HOMO/LUMO (a) ground state, and (b) excited state (552 nm) using DFT and TDDFT methods. |
Next, single point TD-DFT calculation was carried out on the optimized MoNIF cluster. Fig. S9† displays the MoNIF cluster's simulated absorption spectra obtained from TD-DFT calculation. The vertical lines depict the excitations associated with electronic transitions at different wavelengths. The solid black curve in the figure displays the experimental solid-state UV-visible spectrum of the MoNIF cluster. The excited state HOMO–LUMO analysis was implemented on the oscillator frequency corresponding to the 552 nm (Fig. S9†) excited state. In the photo excited state, HOMO shifted from Ni(Oct), Fe(Oct) to Fe(Td) and lattice oxygens adjacent to the Mo metal centre (Fig. 9b).
After this DFT calculations were conducted to investigate the interaction of H2O2 with the Mo doped NiFe2O4 (MoNIF) cluster in the ground state and excited state. It was found that H2O2 interacts best with the octahedral Fe position when MoNIF is in its ground state. In contrast, H2O2 gives the lowest interaction energy with the tetrahedral Fe position (excited state HOMO location) when MoNIF is in the photoexcited state. The optimized H2O2 interaction structures are abbreviated as H2O2_Fe(Oct)_MoNIF_GS (H2O2 interacts with Fe(Oct) of MoNIF cluster in the ground state) and H2O2_Fe(Td)_MoNIF_ES (H2O2 interacts with Fe(Td) of MoNIF cluster in excited state) in the rest of the publication (Fig. 10a and b). Section S6 of ESI† displays the details of the H2O2 interaction potential energy calculations with MoNIF cluster in ground and excited states. Table S4† shows that the interaction energy for the H2O2_Fe(Td)_MoNIF_ES model has a negative value (−0.0053 Hartree) in contrast to positive value (0.0881 Hartree) found for the H2O2_Fe(Oct)_MoNIF_GS model.
Fig. 11(a–d) show the change in different bond lengths in H2O2_Fe(Oct)_MoNIF_GS, and H2O2_Fe(Td)_MoNIF_ES interaction systems. Fig. 11c shows that oxygen of H2O2 interacts with Fe(Oct) of MoNIF system in H2O2_Fe(Oct)_MoNIF_GS model. Fig. 11d shows that oxygen of H2O2 interacts with Fe(Td) and the terminal hydrogen of H2O2 interacts with oxygen coordinated to Fe(Td) in H2O2_Fe(Td)_MoNIF_ES system. The Oα–Oβ bond length of H2O2 part in H2O2_Fe(Oct)_MoNIF_GS, and H2O2_Fe(Td)_MoNIF_ES systems is 1.686 Å, and 1.763 Å. The longer or more weakened Oα–Oβ bond of H2O2, signifies greater bond activation. Thus, Oα–Oβ bond activation is substantially more in the H2O2_Fe(Td)_MoNIF_ES interaction system, in line with the negative (or lower) interaction energy for this system.
Natural bond orbital (NBO) calculations at the same level of theory (Table 2) as earlier were performed on optimized H2O2_Fe(Oct)_MoNIF_GS, and H2O2_Fe(Td)_MoNIF_ES conformations to calculate the second-order stabilization energy E(2) due to electron delocalization. The abbreviations for atom types in Fig. 11c and d systems have also been used to describe the NBO analysis results. Eqn (8) employs second-order perturbation theory to calculate this energy E(2).47
![]() | (8) |
System | Donor NBO | Acceptor NBO | E (2) (kcal mol−1) |
---|---|---|---|
H2O2_Fe(Oct)_MoNIF_GS | LP(FeOct) | BD*(Oα–Oβ) | 0.52 |
BD(Oα–Oβ) | LP*(FeOct) | 4.51 | |
BD(FeOct–Oμ) | LP*(Mo) | 24.67 | |
LP(Mo) | BD*(FeOct–Oμ) | 3.08 | |
H2O2_Fe(Td)_MoNIF_ES | LP(FeTd) | BD*(Oα–Oβ) | 17.38 |
BD(Oα–Oβ) | LP*(FeTd) | 16.31 | |
BD(Mo–Oη) | LP*(FeTd) | 20.79 | |
LP(Oγ) | BD*(Hβ–Oβ) | 7.57 |
In eqn (8), qi = donor orbital occupancy, εi = energy of accepter NBO, εj = energy of donar NBO, Fi,j = off diagonal Fock matrix element. E(2) values quantify the delocalization adjustments to the corresponding zeroth-order natural Lewis structure.
Table 2 shows the significantly larger charge transfers (or delocalization) in the H2O2_Fe(Td)_MoNIF_ES system compared to the H2O2_Fe(Oct)_MoNIF_GS interaction system. The H2O2_Fe(Oct)_MoNIF_GS system, displays a weak charge transfer from LP(FeOct) to BD*(Oα–Oβ) with 0.52 kcal mol−1 interaction energy (E(2) value). Simultaneously, there is a reverse charge transfer from BD(Oα–Oβ) to LP*(FeOct) with 4.51 kcal mol−1E(2) value. These weak charge transfers cause only a slight activation of Oα–Oβ bond of H2O2 molecule. In contrast, the BD (FeOct–Oμ) charge transfer to the LP*(Mo) occurs with a considerably higher E(2) value i.e. 24.67 kcal mol−1. On the contrary, the charge transfer from LP(Mo) to BD*(FeOct–Oμ) only has 3.08 kcal mol−1E(2) value.
In H2O2_Fe(Td)_MoNIF_ES system, charge transfer takes place from LP (FeTd) to BD*(Oα–Oβ) with 17.38 kcal mol−1 delocalization energy. Concurrently, a back donation occurs from BD(Oα–Oβ) to LP*(FeTd) with 16.31 kcal mol−1E(2) value. Both charge transfers are the reasons for the significantly increased weakening and activation of the Oα–Oβ bond of H2O2 molecule in the photoexcited state. On the other hand, charge transfer occurs from BD(Mo–Oη) to LP*(FeTd) with 20.79 kcal mol−1E(2) value. This charge transfer promotes the earlier two charge transfers between Fe(2) and Oα–Oβ related NBOs. Complementary to this, another charge transfer takes place from LP (Oγ) to BD*(Hβ–Oβ) with 7.57 kcal mol−1 interaction energy. Hence, the delocalization in the H2O2_Fe(Td)_MoNIF_ES system is considerably larger. Consequently, the bond activation of Oα–Oβ bond of H2O2 molecule is more in photoexcited state than in ground state, which leads to the efficient formation of OH˙ radicals during the photocatalytic reaction.
DFT studies reveal that HOMO in the MoNIF (a cluster model of 1MNFO) system shifts from Fe(Oct) in the ground state to Fe(Td) in the photoexcited state. TDDFT calculations indicate that the excited state HOMO is localized on the Fe(Td) center. Effectively, the Fe3+ occupying the Td site in 1MNFO is reduced to Fe2+ by the photoexcited electrons. H2O2 interacts more effectively with the photoexcited model through the Fe(Td) site causing greater activation of its O–O bond and OH˙ radical generation. Scavenger experiments confirm that OH˙ radicals are the active species responsible for the photo-Fenton activity. These OH˙ radicals and photogenerated holes subsequently react with tetracycline (TC), oxidizing it into smaller aliphatic fragments (eqn (14)–(16)).
Overall, suitable visible light irradiation photogenerates electron/hole pairs in the photocatalyst (eqn (9)). The photoexcited electrons can reduce Fe3+ to Fe2+ on the surface of the photocatalyst (eqn (10)). Because the redox potential of Mo6+/Mo5+ is approximately 0.4 eV,15,38 the Mo dopant energy level just below the CB of 1MNFO can trap photogenerated carriers, enhancing charge separation and prolonging the charge carrier lifetimes (eqn (11)). Once Mo6+ captures electrons, it is reduced to Mo5+, forming a redox-active species capable of further reactions. Mo5+ then interacts with H2O2 and facilitates the decomposition of H2O2 into ˙OH. Through this reaction, Mo5+ is oxidized back to Mo6+, ensuring the continuous regeneration of active sites and maintaining the catalytic cycle. The detection of Mo4+ in the Mo 3d XPS of recycled 1MNFO sample further implies a redox cycle involving Mo in +6, +5, and +4 oxidation states, potentially contributing to enhanced catalytic performance. The dopant level electrons can also get photoexcited to convert Fe3+ to Fe2+ in the 1MNFO photocatalyst (eqn (12)). The latter interacts and reduces H2O2 to generate ˙OH radicals (eqn (13) and (14)).
1MNFO + hν → 1MNFO(ecb− + hvb+) | (9) |
Fe3+ + ecb− → Fe2+ | (10) |
Mo6+/Mo5+ + ecb− → Mo5+/Mo4+ | (11) |
Mo5+/Mo4+ + Fe3+ → Fe2+ + Mo6+/Mo5+ | (12) |
e− + H2O2 → ˙OH + OH− | (13) |
Fe2+ + H2O2 → Fe3+ + ˙OH + OH− | (14) |
˙OH + tetracycline → TC fragments/CO2 + H2O | (15) |
h+ + TC fragments → CO2 + H2O | (16) |
Mo-doped NiFe2O4 has a reduced band gap (and the shifting of its band edges) show that the dopant Mo acts both as an electron trapper and supplier, improving charge separation and carrier lifetimes during the photo-Fenton process. DFT and TDDFT studies reveal that in the Mo-doped NiFe2O4 (MoNIF) system, the HOMO shifts from Fe(Oct) in the ground state to Fe(Td) in the excited state. Additionally, Fe(Td) of Mo doped NiFe2O4 activates the O–O bond of H2O2 more effectively in the excited state, producing hydroxyl radicals. DFT and experimental results enable us to propose an effective mechanism for photo-Fenton TC degradation.
Footnote |
† Electronic supplementary information (ESI) available: Instruments used for material characterizations, equations for structural parameters calculations, ionic radius table, particle size distribution graph, Tauc plots, photo-Fenton experiment results, TOF comparison table, potential energy DFT calculations, DFT calculated UV visible spectrum, interaction energy DFT calculations. See DOI: https://doi.org/10.1039/d4dt03359k |
This journal is © The Royal Society of Chemistry 2025 |