Next Article in Journal
Carbon-Coated Ni-Fe Nanocatalysts: Bridging the Gap in Cinnamaldehyde Hydrogenation Performance and Durability
Next Article in Special Issue
Catalytic Reductive Degradation of 4-Nitrophenol and Methyl orange by Novel Cobalt Oxide Nanocomposites
Previous Article in Journal
A CoFe Bimetallic Catalyst for the Direct Conversion of Syngas to Olefins
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Tuning the Magnetic and Catalytic Properties of Manganese Ferrite through Zn2+ Doping: Gas Phase Oxidation of Octanol

1
Department of Chemistry, University of Malakand, Chakdara 18800, Pakistan
2
Department of Chemistry, Faculty of Science and Arts at Sharurah, Najran University, Sharurah 68342, Saudi Arabia
3
Department of Chemistry, Bacha Khan University, Charsadda 24420, Pakistan
4
Department of Physics, University of Malakand, Chakdara 18800, Pakistan
*
Authors to whom correspondence should be addressed.
Catalysts 2023, 13(12), 1473; https://doi.org/10.3390/catal13121473
Submission received: 13 September 2023 / Revised: 6 November 2023 / Accepted: 7 November 2023 / Published: 27 November 2023
(This article belongs to the Special Issue Design and Synthesis of Nanostructured Catalysts, 2nd Edition)

Abstract

:
Spinel ferrites, ZnFe2O4, MnFe2O4, and ZnMnFe2O4, were synthesized using the sol–gel method and thoroughly investigated for their potential as catalytic and magnetic materials. Experiments unveiled that ZnMnFe2O4 exhibited excellent catalytic and magnetic properties, whereas the Density Functional Theory (DFT) calculations provided insight into the excellent performance of ZnMnFe2O4 compared with ZnFe2O4 and MnFe2O4. The catalytic efficiencies of the synthesized spinel ferrites were evaluated against a model reaction, i.e., the gas-phase oxidation of octanol to a corresponding aldehyde, utilizing molecular oxygen as an oxidant. The results indicated that the order of catalytic activity was ZnMnFe2O4 > MnFe2O4 > ZnFe2O4. The reaction was found to follow Langmuir Hinshelwood’s mechanism for dissociative adsorption of molecular oxygen. Owing to their superb catalytic and magnetic properties, mixed ferrites can be extended to a variety of organic transformation reactions.

Graphical Abstract

1. Introduction

Spinel ferrite nanoparticles (SFNPs) have magnificent dielectric, magnetic, catalytic, and optical properties [1,2,3]. Owing to these important properties, ferrites are applicable in various fields like catalysis, organic synthesis, drug delivery, magnetic fluids, information and electronic materials, and wastewater treatment [4,5,6,7]. These SFNPs possess the general formula “MFe2O4” and contain M as a divalent metal cation, i.e., Mn(II), Ni(II), Cu(II), Co(II), or Zn(II). SFNPs can generally provide more fascinating catalytic activities compared with single-component metal oxides. Metal cations coupled to oxygen atoms at tetrahedral and octahedral sites, respectively, are considered most appropriate for oxidation reactions due to their structural arrangement [8]. The distribution of cations on octahedral and tetrahedral positions is dependent on ligand attachment and is explained with the help of ligand field theory (LFT) and ligand field stabilization energy (LFSE). SFNPs have incredible potential to be used as a heterogeneous catalyst for different chemical transformation reactions. The advantages of an extended life span (recycling and recovery) make the heterogeneous catalysis method superior to the homogeneous one in industrial applications [9]. SFNPs can be used for novel synthetic processes due to their unique magnetic properties under the protocol of green chemistry. The exhibition of dielectric and magnetic properties makes the application of SFNPs more diversified, including chemical and electronic applications and communication devices, such as computer memory chips, antennas, converters, inductors, and frequency filter coils in mobile boasters and radio centers [10,11,12]. SFNPs are frequently used in the biomedical field, magnetic resonance imaging, supercapacitors, sensing of DNA and RNA, and hyperthermia [13,14,15]. Apart from these applications, inverse spinel Ni ferrites are used as insulators, circulators, and phase shifters [16,17].
The use of mixed ferrites (containing two metal ions other than iron, (MnZnFe2O4)) is also notorious among nanocatalysis. Their size varies in a broad range and can be synthesized using various methods like sol–gel, microwave heating, and co-precipitation [18]. Both SFNPs and mixed ferrites have been frequently used as catalysts. Cobalt ferrite nanoparticles and their mixed ferrites, like Cu1−xCoxFe2O4, were applied as catalysts for the production of hydrogen from methanol [19]. Martins et al. [20] reported solvent-free oxidation of alcohol using different ferrites under microwave irradiation. Similarly, Ni-ZnFe2O4 and Nd-doped Ni-ZnFe2O4 have been successfully utilized for photodegradation of methylene blue while possessing high-magnetic properties [21].
The catalytic oxidation of alcohol to the corresponding carbonyl compound is one of the important transformation reactions in organic synthesis. SFNPs have been used for the oxidation of alcohol to carbonyl compounds with H2O2 as an oxidant while using CuFe2O4/CoFe2O4 as a catalyst, respectively [20,21]. Similarly, the literature reports the enhancing effect of nickel hydroxide (Bronsted base) for the catalytic properties of CoFe2O4 up to 87% in an alcohol oxidation reaction [22], while elsewhere the effect of solvent was reported in the oxidation of benzyl alcohol to benzaldehyde with NiFe2O4 in the presence of acetonitrile at 80 °C with molecular oxygen resulting in a 100% selectivity and a 77% overall activity [23]. These results compelled us to investigate the potential applications of mixed ferrites for gas-phase oxidation of alcohol under mild reaction conditions in a self-designed reactor, Finger Projected Fixed-Bed Reactor (FPFBR).
Here, we investigated the catalytic and magnetic properties of MnFe2O4, ZnFe2O4, and Zn mixed with MnFe2O4 through experiments and theoretical calculation (DFT). To the best of our knowledge, these ferrites have not been investigated for gas-phase oxidation of octanol to the corresponding aldehyde.

2. Results and Discussion

2.1. Characterization

SEM micrograms (Figure 1a,b) revealed the smooth morphology of the nanoparticles. The magnified image shows that the particles are bunched together. The EDX spectrum (Figure 1c) obtained for Mn-ferrite demonstrates that Mn (24 wt%), Fe (46 wt%), and O (29 wt%) are the major elements and reflect the absence of contamination. The diffraction patterns of the Mn-ferrite as shown in Figure 1d are well indexed and have a pure cubic phase of spinel manganese ferrite (JCPDS card no. 38-0430) [24]. Furthermore, the particle size was calculated (68 nm) via XRD, which is in good agreement with the SEM results (Figure 1b). Similarly, 44.3 m2/g of surface area was obtained via Brunauer–Emmett–Teller (BET) adsorption/desorption using the N2 method.
Figure 2a,b show the SEM images of the ZnFe2O4 nanoparticles in which the agglomeration of the nanoparticles was revealed, while EDX analysis (Figure 2c) confirmed the presence of Zn, Fe, and O. Figure 2d represents the XRD patterns of ZnFe2O4 with characteristic planes (220), (311), (222), (400), (422), (511), and (440) for the cubic phase spinel structure (JCPDS No. 77-0011) [25]. The particle size (43 nm) from the most intense peak of XRD (Scherrer equation) is in good agreement with the SEM results (Figure 2b). Furthermore, the BET surface area of the sample was found to be 56 m2/g [26], which is perhaps due to the small size of the particles.
Similarly, SEM images presented in Figure 3a,b show that ZnMnFe2O4 nanoparticles are chimed together; therefore, the particle size cannot be measured from the SEM microgram using the average grain intercept (AGI) method. The EDX spectra (Figure 3c) confirmed the presence of a stoichiometric amount of Zn, Mn, Fe and O, while it confirmed the absence of impurities. The XRD pattern (Figure 3d) shows that the ferrites retained the fundamental cubic phase spinel structure with characteristic planes (111), (220), (311), (400), (422), (511) and (440) [27]. The average particle size (29 nm) was calculated using Equation (6) utilizing the most intense peak, and it was found to be in agreement with the reported literature of the JCPDS card no 10-0319, showing a face-centered cubic crystal structure [28]. Similarly, the BET surface area (31.5 m2/g) was found to decrease with the incorporation of Zn into the matrix of MnFe2O4.

2.2. Magnetic Behavior

Zero-field cooled (ZFC) measurements were performed from 5 to 300 K in a 0.1 T magnetic field to collect data of the magnetic characteristics as shown in Figure 4a. The magnetization continuously displays a net irreversibility, which is compatible with the superparamagnetic condition for MnFe2O4, ZnFe2O4 and ZnMnFe2O4. When the temperature was gradually increased, the magnetization showed the maxima known as the blocking temperature (TB) of the sample. Temperature TB is defined as the point at which thermal activation breaks through the magnetic anisotropy energy barrier of a magnetic single domain, causing the magnetization direction to fluctuate. TB for all the samples is about 35 K.
The magnetization increases with the doping of a Zn ion into the matrix of MnFe2O4 as reflected by Figure 4a–e, which is, of course, one of the prime objectives of the study. Magnetization decreases as temperature exceeds the blocking temperature for all the samples, which is in agreement with the super paramagnetic behavior. The Curie–Weiss law states that the magnetization drop observed above TB is caused by the super paramagnetic behavior of nanoparticles. Initially, the increase in magnetization at a temperature below TB can be explained by frozen magnetic moments along the easy axis of nanoparticles because of the considerable anisotropy energy available at low temperatures. The nanoparticles accumulate thermal energy as the temperature rises, resisting the anisotropic energy barrier, making it much simpler to align with the field. The relationship between the temperature (TB) at which magnetization is best and the volume of the crystalline particles (V) is KV/25 kB, where K is the magneto crystalline anisotropy constant.

2.3. DFT Calculation

The structural properties of the ZnFe2O4, ZnMnFe2O4 and MnFe2O4 spinals are calculated through GGA+U, and their crystal structure is shown in Figure 5. The unit cell of each compound is optimized using the fitted Birch–Murnaghan equation of state [29] as shown in Figure 6 in order to find out the ground state structure parameters of these compounds. From Figure 7, the calculated ground state energies (E0), lattice constants (a0), volume (V) and bulk modulus (B0) compound along with experimental work are listed in Table 1. The calculated lattice constants of these compounds are 8.438, 8.441 and 8.447 Å, respectively.
It can be seen from Table 1 that calculated lattice constants are consistent with the experiments and are reliable. The calculated E0 for ZnFe2O4, ZnMnFe2O4 and MnFe2O4 is −2.17 × 105, −2.35 × 105 and −2.52 × 105 eV, respectively. The volume (V) of these spinels is 599.07, 601.425 and 602.751 Å3, respectively, and bulk modulus B0 (GPa) is 148.167, 174.039 and 216.738, which show that MnFe2O4 is harder than the other compounds.
To find the stable magnetic structure of these spinal ferrites, the (1 × 1 × 2) super cell of each compound is optimized using the fitted Birch–Murnaghan equation of state [33] in paramagnetic (PM)/non-magnetic (NM), ferromagnetic (FM) and in Anti-ferromagnetic (AFM) phases as shown in Figure 7. From Figure 7, the calculated ground state energies for the NM phase of ZnFe2O4, ZnMnFe2O4 and MnFe2O4 are −4.35 × 105, −4.7 × 105 and −5 × 105 eV; for FM phases, they are −4.35 × 105, −4.7 × 105 and −5 × 105 eV; and for the AFM phase, they are −4.35 × 105, −4.7 × 105 and −5 × 105 eV, respectively. The comparison of the calculated energies in different phases reveals the AFM nature of all these compounds in which the energy is lower compared to that of other phases. The calculated AFM phase of these compounds is consistent with experimental results. This shows that present results are logical and rational.
As the understudy compounds are anti-ferromagnetic, the electronic properties of these compounds were performed in the anti-ferromagnetic phase using the GGA+U potential. The total density of states (TDOS) and the band structure for the understudy compounds are computed in Figure 8. The TDOS for these compounds reveal the presence of the energy gap flanked by the valance band (VB), and conduction bands (CBs) are non-overlapping with densities around the Fermi Level, which reveal the semiconductor nature of MnFe2O4, ZnFe2O4 and ZnMnFe2O4. From TDOS, the estimated band gap values are 1.6, 1.9 and 1.1 eV, respectively, consistent with the reported results [29,35]. Additionally, the band structures for MnFe2O4, ZnFe2O4 and ZnMnFe2O4 are also calculated and displayed in Figure 9 to further validate their semiconducting nature. Figure 9 shows that there is non-overlapping at the Fermi level, and the maxima of VB align with the minima of CB at the central symmetry for compounds, which indicates their direct band gap character. The enhanced catalytic activity of ZnMnFe2O4 can be explained on the basis of reduction in band gap energy, which further endorses the experimental observation (order of catalytic activity: ZnMnFe2O4 > MnFe2O4 > ZnFe2O4).

2.4. Catalytic Properties of MnFe2O4, ZnFe2O4 and ZnMnFe2O4

All the ferrite NPs were screened for catalytic activity in a model reaction (gas-phase oxidation of octanol to octanal) at a pre-optimized set of reaction conditions for example catalysts: 2 g/chamber, vapor pressure of reactant, 80 Torr; temperature, 140 °C; and flow, 40 mL/min. The data were collected after 30 min, when the steady state established. Figure 10 shows that the rate of reaction increases with increase in reaction duration, following the same trend for MnFe2O4, ZnFe2O4 and ZnMnFe2O4. However, the order of activity was found, ZnMnFe2O4 > MnFe2O4 > ZnFe2O4, which points to the synergistic effect of Zn doping in the matrix of MnFe2O4. Similar results for Zn2+ doping over MnFe2O4 were reported by Manohar et al. [36] for magnetic hypothermia treatment. The increase in Zn2+ content was found to increase the magnetic properties linearly. The doping also has an effect on the size of the crystallite, as increase in the content of dopant causes the crystallite size to decrease, and agglomeration of NPs occurs. Perhaps this synergism arises from the rearrangement of Fe3+ ions on the octahedral sites due to the occupancy of tetrahedral sites by Zn2+ ions, which generates electron hopping between Fe2+ and Fe3+ ions on octahedral sites. Similarly, Figure 11 elucidates the demission of diffusion and kinetic control regime. The variation in the rate of reaction with vapor pressure of reactant revealed that at a lower vapor pressure (20–40 Torr), the reaction is diffusion control, while at a high vapor pressure (80 Torr), the rate shows a linear trend with an increase in flow feed gases pointing to the reaction being kinetic control. Similarly, the flow effect and activation energy (56 kJ/mole) calculated from the Arrhenius equation, also confirming the kinetic control regime (inset of Figure 11).

2.4.1. Kinetics Analysis

The experimental data obtained from the gas phase oxidation of octanol over spinel ferrites were exposed to the Langmuir–Hinshelwood (L-H) model for kinetic analysis [37].
The L-H model considered for the adsorbed reactants (C8H18O: A and O2: B) over active sites of the catalyst can be formulated as
r = k K 1 K 2 C 8 H 18 O [ O 2 ] g n [ 1 + K 1 C 8 H 18 O + K 2 [ O 2 ] g n ] 2 , L-H   model ,
where k = rate coefficient, K1 = equilibrium adsorption constant of Reactant A, and K2 = equilibrium adsorption constants of Reactant B. Figure 14a shows that the reaction rate for octanol dehydrogenation in the presence of oxygen is higher than that of direct dehydrogenation, which confirms the involvement of oxygen. However, to understand the nature of oxygen adsorption on the surface of the catalyst, Equation (1) was modified to the following equation by considering low partial pressure of O2 where k2[O2] in the denominator that can be neglected:
r = k K 1 K 2 C 8 H 18 O [ O 2 ] g [ 1 + K 1 C 8 H 18 O ] 2 , L-H   model ,
where at constant vapor pressure of Reactant A, Equation (2) was modified to Equation (3),
r = a O 2 , Non-dissociative ,
while a is a constant used instead of all constant values.
Equation (1) can also be modified to Equation (4) for a dissociative adsorption of O2 on the surface of the catalyst.
r = a [ O 2 ] [ b + c [ O 2 ] 2 ] , Dissociative ,
where the linear form of Equation (4) can be written as
O 2 4 r a t e = b a + c a [ O 2 ] ,   Linear   form   of   dissociative .
The experimental data were fitted into Equations (3) and (5), and a better regression value was observed for Equation (5) (0.9906) than for the Equation (3) (0.9490), which revealed that the reaction follows the L-H form with a dissociative adsorption of oxygen rather than adsorption of molecular oxygen as shown in Figure 12, but direct activation of the gas phase O2 on the surface of the catalyst has a much high barrier (cal: 2.64 eV), which restricted the direct cleavage of the O2 bond under ambient conditions.
Keeping in view the previous reports regarding the activation of O2 [38,39,40], we assumed that the co-adsorbed alcohol [C8H17OH---O2] on the surface of catalyst acts as an activator for the dissociation of O2, where oxygen abstracts hydrogen from alcohol and produces an important intermediate hydroperoxyl species; {(C8H17OH---O2)* → C8H17O* + OOH*} has a barrier of 0.88 eV (TS1, Figure 13), which is lower than the direct dissociation of O—O bond, although including the possibility of H2O formation as a by-product during the progress of reaction further decrease the barrier up to 0.56 eV (TS2, Figure 13), which shows that H2O is acting as promoter for O2 dissociation. The subsequent oxidation of C8H17OH can be achieved by OOH*, O* and OH* as shown in Scheme 1.

2.4.2. Comparative Study

The catalytic efficiency of MnFe2O4, ZnFe2O4 and ZnMnFe2O4 in terms of oxidation for different substrates were tabulated and compared with recent results as shown in Table 2. The MnFe2O4 for the oxidation of benzyl alcohol (Table 2, entry 1) in acetonitrile showed a 100% selectivity with good activity in the presence of an oxidant (H2O2) at a mild temperature and a 10 h reaction time. Similarly, the oxidation of 5-hydroxy methyl furfural (HMF) (Table 2, entry 2) has been reported with an 85% selectivity of the reaction, which shows that MnFe2O4 is a potent catalyst for oxidation reactions. The benzyl alcohol oxidation to benzophenone and benzaldehyde are reported (Table 2, Entries 3 and 4) with a mild reaction condition and remarkable selectivity and activity. Similarly, phenol has been successfully converted to organic carbon and water in the presence of potassium peroxy disulfate at room temperature for a period of 6 h (Table 2, Entry 5). The results can be used for the potential use of MnFe2O4 for oxidation reactions with remarkable efficiency and selectivity in the liquid phase which is limited by an extended reaction duration.
However, ZnFe2O4 was effectively utilized for the gas phase oxidation of 1-butene to 1,3-butadiene at very high temperature >350 °C (Table 2, Entries 6 and 7). However, the prolonged reaction time and high temperature challenge the applicability of the catalyst in these particular reactions. Similarly, MnZnFe2O4 was also used for the formation of Acetophenone, utilizing t-BuOOH as an oxidant with a 99% selectivity and a mild reaction condition (Table 2, Entry 8) with a 96.7% selectivity of the reaction; but still, the use organic oxidants possess problems. The tabulated data (Table 2, Entry 1–9) show importance of the present work.

3. Experimental

3.1. Materials and Methods

All the chemicals were purchased from Sigma Aldrich (Louis, MO, USA) and used without any further treatment or purification, while gases (N2, O2, CO2 and H2) were delivered by Pakistan Oxygen Company (BOC) Ltd. (Taxila, Pakistan) and passed through proper filters. Flow rate and pressure of gases and reactants were controlled through analogue gauges, T-valves and needle valves.

3.1.1. Preparation of Metal Doped Ferrite

The sol–gel method was employed to synthesize spinel ferrites and mixed-ferrite NPs. A stoichiometric amount of precursor salts was dissolved in deionized water/ethanol and sonicated for 30 min. To the preheated solution (50 °C), citric acid (5 mL) was added dropwise under vigorous stirring (1200 rpm). The pH of the mixture was maintained at 8–9 by adding aqueous ammonia. The mixture was aged for 6 h at 80 °C. The temperature of the mixture was increased to 220 °C with 1 °C/min to obtain the gel and further dried until flake formation. Finally, the particles were calcined for 6 h at 800 °C (0.5 °C/min).

3.1.2. Characterization

The synthesized samples were morphologically studied by Scanning Electron Microscopy (JSM 5910, JEOL, Tokyo, Japan), and Energy Dispersive X-rays (EDX, JSM 5910, JEOL, Tokyo, Japan) were used for the determination of elemental composition. XRD spectra were recorded by an X-ray Diffractometer (Model: JDX; (3532) JEOL, Tokyo, Japan) with a radiation source (CuKα) and 2θ = 0–80°. The most intense peak was used for calculating particle size (Scherrer equation (Equation (6)),
D = k λ β C o s θ
where λ = 1.5406 Å, β is the full width at half maximum (FWHM) in the radian. The surface area and pore size were determined by BET surface area analyzer (Quantachrome Nova 2200e, Boynton Beach, FL, USA). Magnetic properties like remanence magnetization (Mr), saturation magnetization (Ms) and coercivity (Hc) were determined by a Vibrating Sample Magnetometer (VSM; Cryogenic Ltd., London, UK).

3.2. DFT Study

In this work, DFT calculations were executed with the method of full-potential Linearized Augmented plane wave (FP-LAPW) employed in the WIEN2k code [48]. The generalized gradient approximation along with Hubbard potential (GGA+U) was used to calculate the exchange correlation energy for determination of the electronic properties [49]. In the FP-LAPW method, the charge density expansion and wave function for potential is composed of the combination of radial function multiplied by spherical harmonics inside atomic radii and plan waves in the interstitial region. The cut-off value for planned wave RMT × Kmax = 8 Ry is selected, while the charge density is Fourier expanded up to Gmax = 14 (Ryd). For the integration of the Brillouin Zone, we used a 7 × 7 × 7 k-point. The self-consistent convergence was achieved up to 0.5 mRy/Bohr. For GGA+U calculations, the values for divalent metals were taken from the literature [50].

3.3. Catalytic Test

The synthesized catalysts were loaded to a self-designed reactor (FPFBR; Glass Blowing Department, University of Engineering and Technology, Peshawar, Pakistan) as shown in Figure 14b. Ferrites were held inside the reactor by externally fitted magnetic beads. Continuous flows of reactants with desired vapor pressure (20–80 Torr) were simultaneously allowed into the reactor (FPFBR) through a saturator at a preheated temperature of 413 ± 5 K. A general formula (n = PVf/RT) was used to calculate the number of moles of reactants, where Vf is the flow rate (mL) of volume. The mixture of gases was allowed passing through reaction cores (beds of catalyst in the reaction chamber) at the desired temperature, maintained through heating taps, of one to three chambered reactors, each chamber consisting of 18 catalyst beds as reported in previous study [51]. The six-port valve was used for the direct analysis of outflow through GC (Perkin Elmer, Clarus; 580, Shelton, CT, USA). The effect of oxygen flow on/off was demonstrated for the oxidation of octanol as shown in Figure 14a.

4. Conclusions

Here, we tuned the catalytic and magnetic properties of spinal ferrite through incorporation of other divalent metal atoms. A combinatorial approach of experiments and theory was adapted to understand the phenomena behind the superb properties of mixed ferrites. Experiments showed that ZnMnFe2O4 has better catalytic and magnetic properties than MnFe2O4 and ZnFe2O4, which was endorsed by DFT calculation. DFT calculation showed that the Zn metal alters the electronic structure and gap state of MnFe2O4, decreasing the Fermi energy level and band gap up to 1.1 eV. On the basis of catalysis, a ranking was established for the model reaction (oxidation of octanol), ZnMnFe2O4 > MnFe2O4 > ZnFe2O4, with a dissociative adsorption L-H mechanism of O2 at the surface of the catalyst. High magnetic properties of ZnMnFe2O4 showed its extended use for a variety of gas phase catalysis in a finger projected reactor (FPFBR). Furthermore, stable nature and magnetic and excellent catalytic properties revealed that ZnMnFe2O4 is a potent catalyst for industrial important organic transformation.

Author Contributions

Investigation, methodology and writing—original draft preparation, M.B.; conceptualization, supervision and writing—review and editing, M.S.; software and theoretical calculation, Z.A.; writing—review and editing, Z.I., M.A.R. and R.A.A.; funding acquisition, M.A.R. and R.A.A. All authors have read and agreed to the published version of the manuscript.

Funding

This research is supported by the financial support of the Deputy for Research and Innovation Ministry of Education, Kingdom of Saudi Arabia through a grant (NU/IFC/2/SERC/-/7) under the Institutional Funding Committee at Najran University, Kingdom of Saudi Arabia.

Data Availability Statement

Raw data will be made available on request.

Acknowledgments

Authors appreciate and acknowledge the Deputy for Research and Innovation Ministry of Education and Institutional Funding Committee at Najran University, Kingdom of Saudi Arabia.

Conflicts of Interest

Authors declare no conflict of interest.

References

  1. Almessiere, M.A.; Güner, S.; Slimani, Y.; Baykal, A.; Shirsath, S.E.; Korkmaz, A.D.; Badar, R.; Manikandan, A. Investigation on the structural, optical, and magnetic features of Dy3+ and Y3+ co-doped Mn0.5Zn0.5Fe2O4 spinel ferrite nanoparticles. J. Mol. Struct. 2022, 1248, 131412. [Google Scholar] [CrossRef]
  2. Kanna, R.R.; Lenin, N.; Sakthipandi, K.; Kumar, A.S. Structural, optical, dielectric and magnetic studies of gadolinium-added Mn-Cu nanoferrites. J. Magn. Magn. Mater. 2018, 453, 78–90. [Google Scholar] [CrossRef]
  3. Amiri, M.; Eskandari, K.; Salavati-Niasari, M. Magnetically retrievable ferrite nanoparticles in the catalysis application. Adv. Colloid Interface Sci. 2019, 271, 101982. [Google Scholar] [CrossRef]
  4. Kefeni, K.K.; Msagati, T.A.M.; Nkambule, T.T.; Mamba, B.B. Spinel ferrite nanoparticles and nanocomposites for biomedical applications and their toxicity. Mater. Sci. Eng. C 2020, 107, 110314. [Google Scholar] [CrossRef] [PubMed]
  5. Ma, H.; Liu, C. A mini-review of ferrites-based photocatalyst on application of hydrogen production. Front. Energy 2021, 15, 621–630. [Google Scholar] [CrossRef]
  6. Gupta, N.K.; Ghaffari, Y.; Kim, S.; Bae, J.; Kim, K.S.; Saifuddin, M. Photocatalytic Degradation of Organic Pollutants over MFe2O4 (M  =  Co, Ni, Cu, Zn) Nanoparticles at Neutral pH. Sci. Rep. 2020, 10, 4942. [Google Scholar] [CrossRef] [PubMed]
  7. Dippong, T.; Levei, E.A.; Cadar, O. Recent Advances in Synthesis and Applications of MFe2O4 (M = Co, Cu, Mn, Ni, Zn) Nanoparticles. Nanomaterials 2021, 11, 1560. [Google Scholar] [CrossRef] [PubMed]
  8. Prasad, S.A.V.; Deepty, M.; Ramesh, P.N.; Prasad, G.; Srinivasarao, K.; Srinivas, C.; Babu, K.V.; Kumar, E.R.; Mohan, N.K.; Sastry, D.L. Synthesis of MFe2O4 (M = Mg2+, Zn2+, Mn2+) spinel ferrites and their structural, elastic and electron magnetic resonance properties. Ceram. Int. 2018, 44, 10517–10524. [Google Scholar] [CrossRef]
  9. Pham, T.N.; Huy, T.Q.; Le, A.T. Spinel ferrite (AFe2O4)-based heterostructured designs for lithium-ion battery, environmental monitoring, and biomedical applications. RSC Adv. 2020, 10, 31622–31661. [Google Scholar] [CrossRef]
  10. Jung, S.; Kim, S.; Cho, W.; Lee, K. Development of highly efficient energy harvester based on magnetic field emanating from a household power line and its autonomous interface electronics. IEEE Sens. J. 2023, 23, 6607–6615. [Google Scholar] [CrossRef]
  11. Hu, W.; Qin, N.; Wu, G.; Lin, Y.; Li, S.; Bao, D. Opportunity of spinel ferrite materials in nonvolatile memory device applications based on their resistive switching performances. J. Am. Chem. Soc. 2012, 134, 14658–14661. [Google Scholar] [CrossRef] [PubMed]
  12. Cai, H.L.; Zhan, J.; Yang, C.; Chen, X.; Yang, Y.; Chi, B.Y.; Wang, A.; Ren, T.L. Application of ferrite nanomaterial in RF on-chip inductors. J. Nanomater. 2013, 2013, 832401. [Google Scholar] [CrossRef]
  13. Silvestri, N.; Gavilán, H.; Guardia, P.; Brescia, R.; Fernandes, S.; Samia, A.C.S.; Teran, F.J.; Pellegrino, T. Di- and tri-component spinel ferrite nanocubes: Synthesis and their comparative characterization for theranostic applications. Nanoscale 2021, 13, 13665–13680. [Google Scholar] [CrossRef] [PubMed]
  14. Abdel Maksoud, M.I.A.; Ghobashy, M.M.; Kodous, A.S.; Fahim, R.A.; Osman, A.I.; Al-Muhtaseb, A.H.; Rooney, D.W.; Mamdouh, M.A.; Nady, N.; Ashour, A.H. Insights on magnetic spinel ferrites for targeted drug delivery and hyperthermia applications. Nanotechnol. Rev. 2022, 11, 372–413. [Google Scholar] [CrossRef]
  15. Alghamdi, N.A.; Hankiewicz, J.H.; Anderson, N.R.; Stupic, K.F.; Camley, R.E.; Przybylski, M.; Zukrowski, J.; Celinski, Z. Development of Ferrite-Based Temperature Sensors for Magnetic Resonance Imaging: A Study of Cu1–xZnxFe2O4. Phys. Rev. Appl. 2018, 9, 054030. [Google Scholar] [CrossRef] [PubMed]
  16. Valenzuela, R. Novel applications of ferrites. Phys. Res. Int. 2012, 2012, 591839. [Google Scholar] [CrossRef]
  17. Khan, M.Z.; Gul, I.H.; Baig, M.M.; Khan, A.N. Comprehensive study on structural, electrical, magnetic and photocatalytic degradation properties of Al3+ ions substituted nickel ferrites nanoparticles. J. Alloys Compd. 2020, 848, 155795. [Google Scholar] [CrossRef]
  18. Lai, Z.; Xu, G.; Zheng, Y. Microwave assisted low temperature synthesis of MnZn ferrite nanoparticles. Nanoscale Res. Lett. 2007, 2, 40–43. [Google Scholar] [CrossRef]
  19. Mathew, T.; Vijayaraj, M.; Pai, S.; Tope, B.B.; Hegde, S.G.; Rao, B.S.; Gopinath, C.S. A mechanistic approach to phenol methylation on Cu1−xCoxFe2O4: FTIR study. J. Catal. 2004, 227, 175–185. [Google Scholar] [CrossRef]
  20. Martins, N.M.R.; Martins, L.M.D.R.S.; Amorim, C.O.; Amaral, V.S.; Pombeiro, A.J.L. Solvent-Free Microwave-Induced Oxidation of Alcohols Catalyzed by Ferrite Magnetic Nanoparticles. Catalysts 2017, 7, 222. [Google Scholar] [CrossRef]
  21. Dhiman, P.; Rana, G.; Dawi, E.A.; Kumar, A.; Sharma, G.; Kumar, A.; Sharma, J. Tuning the Photocatalytic Performance of Ni-Zn Ferrite Catalyst Using Nd Doping for Solar Light-Driven Catalytic Degradation of Methylene Blue. Water 2023, 15, 187. [Google Scholar] [CrossRef]
  22. Bhat, P.B.; Inam, F.; Bhat, B.R. Nickel hydroxide/cobalt-ferrite magnetic nanocatalyst for alcohol oxidation. ACS Comb. Sci. 2014, 16, 397–402. [Google Scholar] [CrossRef]
  23. Iraqui, S.; Kashyap, S.S.; Rashid, M.H. NiFe2O4 nanoparticles: An efficient and reusable catalyst for the selective oxidation of benzyl alcohol to benzaldehyde under mild conditions. Nanoscale Adv. 2020, 2, 5790–5802. [Google Scholar] [CrossRef]
  24. Ranjith Kumar, E.; Jayaprakash, R.; Arunkumar, T.; Kumar, S. Effect of reaction time on particle size and dielectric properties of manganese substituted CoFe2O4 nanoparticles. J. Phys. Chem. Solids 2013, 74, 110–114. [Google Scholar] [CrossRef]
  25. Vergis, B.R.; Kottam, N.; Krishna, R.H.; Nagabhushana, B.M. Removal of Evans Blue dye from aqueous solution using magnetic spinel ZnFe2O4 nanomaterial: Adsorption isotherms and kinetics. Nano-Struct. Nano-Objects 2019, 18, 100290. [Google Scholar] [CrossRef]
  26. Zhang, X.; Chen, Z.; Liu, J.; Cui, S. Synthesis and characterization of ZnFe2O4 nanoparticles on infrared radiation by xerogel with sol-gel method. Chem. Phys. Lett. 2021, 764, 138265. [Google Scholar] [CrossRef]
  27. Godara, S.K.; Kaur, V.; Narang, S.B.; Singh, G.; Singh, M.; Bhadu, G.R.; Chaudhari, J.C.; Babu, P.D.; Sood, A.K. Tailoring the magnetic properties of M-type strontium ferrite with synergistic effect of co-substitution and calcinations temperature. J. Asian Ceram. Soc. 2021, 9, 686–698. [Google Scholar] [CrossRef]
  28. Karaagac, O.; Köçkar, H. The effects of temperature and reaction time on the formation of manganese ferrite nanoparticles synthesized by hydrothermal method. J. Mater. Sci. Mater. Electron. 2020, 31, 2567–2574. [Google Scholar] [CrossRef]
  29. Yao, J.; Li, X.; Li, Y.; Le, S. Density Functional Theory Investigations on the Structure and Electronic Properties of Normal Spinel ZnFe2O4. Integr. Ferroelectr. 2013, 145, 17–23. [Google Scholar] [CrossRef]
  30. Ali, S.Y.; Eid, O.I.; Siddig, M.A. The Influence of Cu on the Dielectric Properties of NiZnFe2O4 Synthesized by Solid State Reaction Method. J. Mater. Sci. Chem. Eng. 2020, 8, 14–23. [Google Scholar]
  31. Mozaffari, M.; Arani, M.; Amighian, J. The effect of cation distribution on magnetization of ZnFe2O4 nanoparticles. J. Magn. Magn. Mater. 2010, 322, 3240–3244. [Google Scholar] [CrossRef]
  32. Yadav, A.; Varshney, D. Structural and temperature dependent dielectric behavior of Cr and Zn doped MnFe2O4 nano ferrites. Superlattices Microstruct. 2018, 113, 153–159. [Google Scholar] [CrossRef]
  33. Ansari, M.M.N.; Khan, S. Structural, electrical and optical properties of sol-gel synthesized cobalt substituted MnFe2O4 nanoparticles. Phys. B Condens. Matter 2017, 520, 21–27. [Google Scholar] [CrossRef]
  34. Pışırır, O.M.; Bingöl, O. Industrial PC based heliostat control for solar power towers. Acta Phys. Pol. Ser. A 2016, 130, 36–40. [Google Scholar] [CrossRef]
  35. Simon, C.; Blösser, A.; Eckardt, M.; Kurz, H.; Weber, B.; Zobel, M.; Marschall, R. Magnetic properties and structural analysis on spinel MnFe2O4 nanoparticles prepared via non-aqueous microwave synthesis. Z. Anorg. Allg. Chem. 2021, 647, 2061–2072. [Google Scholar] [CrossRef]
  36. Manohar, A.; Vijayakanth, V.; Chintagumpala, K.; Manivasagan, P.; Jang, E.S.; Kim, K.H. Zn-doped MnFe2O4 nanoparticles for magnetic hyperthermia and their cytotoxicity study in normal and cancer cell lines. Colloids Surf. A Physicochem. Eng. Asp. 2023, 675, 132037. [Google Scholar] [CrossRef]
  37. Ilyas, M.; Sadiq, M. Liquid-Phase Aerobic Oxidation of Benzyl Alcohol Catalyzed by Pt/ZrO2. Chem. Eng. Technol. 2007, 30, 1391–1397. [Google Scholar] [CrossRef]
  38. Chang, C.R.; Wang, Y.G.; Li, J. Theoretical investigations of the catalytic role of water in propene epoxidation on gold nanoclusters: A hydroperoxyl-mediated pathway. Nano Res. 2011, 4, 131–142. [Google Scholar] [CrossRef]
  39. Chang, C.R.; Yang, X.F.; Long, B.; Li, J. A water-promoted mechanism of alcohol oxidation on a Au(111) surface: Understanding the catalytic behavior of bulk gold. ACS Catal. 2013, 3, 1693–1699. [Google Scholar] [CrossRef]
  40. Zhou, M.; Zhao, Y.; Gong, Y.; Li, J. Formation and characterization of the XeOO+ cation in solid argon. J. Am. Chem. Soc. 2006, 128, 2504–2505. [Google Scholar] [CrossRef]
  41. Jacintha, A.M.; Manikandan, A.; Chinnaraj, K.; Antony, S.A.; Neeraja, P. Comparative studies of Spinel MnFe2O4 nanostructures: Structural, morphological, optical, magnetic and catalytic properties. J. Nanosci. Nanotechnol. 2015, 15, 9732–9740. [Google Scholar] [CrossRef] [PubMed]
  42. Gawade, A.B.; Nakhate, A.V.; Yadav, G.D. Selective synthesis of 2,5-furandicarboxylic acid by oxidation of 5-hydroxymethylfurfural over MnFe2O4 catalyst. Catal. Today 2018, 309, 119–125. [Google Scholar] [CrossRef]
  43. Burange, A.S.; Kale, S.R.; Zboril, R.; Gawande, M.B.; Jayaram, R.V. Magnetically retrievable MFe2O4 spinel (M = Mn, Co, Cu, Ni, Zn) catalysts for oxidation of benzylic alcohols to carbonyls. RSC Adv. 2014, 4, 6597–6601. [Google Scholar] [CrossRef]
  44. Mary Jacintha, A.; Umapathy, V.; Neeraja, P.; Rex Jeya Rajkumar, S. Synthesis and comparative studies of MnFe2O4 nanoparticles with different natural polymers by sol-gel method: Structural, morphological, optical, magnetic, catalytic and biological activities. J. Nanostructure Chem. 2017, 7, 375–387. [Google Scholar] [CrossRef]
  45. Stoia, M.; Muntean, C.; Militaru, B. MnFe2O4 nanoparticles as new catalyst for oxidative degradation of phenol by peroxydisulfate. J. Environ. Sci. 2017, 53, 269–277. [Google Scholar] [CrossRef] [PubMed]
  46. Zeng, T.; Sun, G.; Miao, C.; Yan, G.; Ye, Y.; Yang, W.; Sautet, P. Stabilizing Oxidative Dehydrogenation Active Sites at High Temperature with Steam: ZnFe2O4-Catalyzed Oxidative Dehydrogenation of 1-Butene to 1,3-Butadiene. ACS Catal. 2020, 10, 12888–12897. [Google Scholar] [CrossRef]
  47. Lee, H.; Jung, J.C.; Kim, H.; Chung, Y.M.; Kim, T.J.; Lee, S.J.; Oh, S.H.; Kim, Y.S.; Song, I.K. Effect of pH in the preparation of ZnFe2O4 for oxidative dehydrogenation of n-butene to 1,3-butadiene: Correlation between catalytic performance and surface acidity of ZnFe2O4. Catal. Commun. 2008, 9, 1137–1142. [Google Scholar] [CrossRef]
  48. Ambrosch-Draxl, C.; Sofo, J.O. Linear optical properties of solids within the full-potential linearized augmented planewave method. Comput. Phys. Commun. 2006, 175, 1–14. [Google Scholar] [CrossRef]
  49. Perdew, J.P.; Burke, K.; Ernzerhof, M. Generalized Gradient Approximation Made Simple. Phys. Rev. Lett. 1996, 77, 3865. [Google Scholar] [CrossRef]
  50. Zhang, Y.J.; Tang, H.M.; Gao, S.P. Density Functional Theory Study of ZnIn2S4 and CdIn2S4 Polymorphs Using Full-Potential Linearized Augmented Plane Wave Method and Modified Becke–Johnson Potential. Phys. Status Solidi 2020, 257, 1900485. [Google Scholar] [CrossRef]
  51. Bibi, M.; Ullah, R.; Sadiq, M.; Sadiq, S.; Khan, I.; Saeed, K.; Zia, M.A.; Iqbal, Z.; Ullah, I.; Iqbal, Z.; et al. Catalytic hydrogenation of carbon dioxide over magnetic nanoparticles: Modification in fixed-bed reactor. Catalysts 2021, 11, 592. [Google Scholar] [CrossRef]
Figure 1. (a,b) SEM images = at different magnifications (1700 and 5500); (c) EDX spectra and elemental composition; (d) XRD pattern of MnFe2O4.
Figure 1. (a,b) SEM images = at different magnifications (1700 and 5500); (c) EDX spectra and elemental composition; (d) XRD pattern of MnFe2O4.
Catalysts 13 01473 g001
Figure 2. (a,b) SEM images at different magnifications (2700 and 5500); (c) EDX spectra and elemental composition; (d) XRD pattern of ZnFe2O4.
Figure 2. (a,b) SEM images at different magnifications (2700 and 5500); (c) EDX spectra and elemental composition; (d) XRD pattern of ZnFe2O4.
Catalysts 13 01473 g002
Figure 3. (a,b) SEM images at different magnifications (2700 and 5500); (c) EDX spectra and elemental composition; (d) XRD pattern of ZnMnFe2O4.
Figure 3. (a,b) SEM images at different magnifications (2700 and 5500); (c) EDX spectra and elemental composition; (d) XRD pattern of ZnMnFe2O4.
Catalysts 13 01473 g003
Figure 4. (a) Magnetic properties and (be) demonstration of magnetic properties in the presence of external magnet.
Figure 4. (a) Magnetic properties and (be) demonstration of magnetic properties in the presence of external magnet.
Catalysts 13 01473 g004
Figure 5. (Left) side figure is the crystal structure of ZnFe2O4 and MnFe2O4; (right) side is ZnMnFe2O4.
Figure 5. (Left) side figure is the crystal structure of ZnFe2O4 and MnFe2O4; (right) side is ZnMnFe2O4.
Catalysts 13 01473 g005
Figure 6. Energy versus volume optimization curves of MnFe2O4, ZnFe2O4 and ZnMnFe2O4.
Figure 6. Energy versus volume optimization curves of MnFe2O4, ZnFe2O4 and ZnMnFe2O4.
Catalysts 13 01473 g006
Figure 7. Optimization curves of the stable magnetic phases of MnFe2O4, ZnFe2O4 and ZnMnFe2O4.
Figure 7. Optimization curves of the stable magnetic phases of MnFe2O4, ZnFe2O4 and ZnMnFe2O4.
Catalysts 13 01473 g007
Figure 8. Corresponding total DOS for MnFe2O4, ZnFe2O4 and ZnMnFe2O4.
Figure 8. Corresponding total DOS for MnFe2O4, ZnFe2O4 and ZnMnFe2O4.
Catalysts 13 01473 g008
Figure 9. Plot for band gap structure of MnFe2O4, ZnFe2O4 and ZnMnFe2O4.
Figure 9. Plot for band gap structure of MnFe2O4, ZnFe2O4 and ZnMnFe2O4.
Catalysts 13 01473 g009
Figure 10. Effect of time on the rate of reaction and number of moles trapped. Conditions: Catalysts, 2 g/chamber; Vapor pressure of reactant, 80 Torr; Temperature, 140 °C; Flow, 40 mL/min; Data collection, after 30 min. r = rate of reaction (molg−1 min−1), n = no. of moles trapped (product: mol).
Figure 10. Effect of time on the rate of reaction and number of moles trapped. Conditions: Catalysts, 2 g/chamber; Vapor pressure of reactant, 80 Torr; Temperature, 140 °C; Flow, 40 mL/min; Data collection, after 30 min. r = rate of reaction (molg−1 min−1), n = no. of moles trapped (product: mol).
Catalysts 13 01473 g010
Figure 11. Effect of flow on the rate of reaction, while the inset represents the Arrhenius plot for activation energy. Conditions: Catalysts, 2 g/chamber; Vapor pressure of reactant, 80 Torr; Temperature, 140 °C; Flow, 40 mL/min; Data collection, after 30 min. Ea; activation energy, K; kelvin, r = rate of reaction (molg−1 min−1).
Figure 11. Effect of flow on the rate of reaction, while the inset represents the Arrhenius plot for activation energy. Conditions: Catalysts, 2 g/chamber; Vapor pressure of reactant, 80 Torr; Temperature, 140 °C; Flow, 40 mL/min; Data collection, after 30 min. Ea; activation energy, K; kelvin, r = rate of reaction (molg−1 min−1).
Catalysts 13 01473 g011
Figure 12. Langmuir–Hinshelwood mechanism for dissociative and non-dissociative adsorption of O2 on the surface of catalysts. Conditions: Catalysts, 2 g/chamber; Vapor pressure of reactant, 80 Torr; Temperature, 140 °C; Flow, 40 mL/min; Data collection, after 30 min.
Figure 12. Langmuir–Hinshelwood mechanism for dissociative and non-dissociative adsorption of O2 on the surface of catalysts. Conditions: Catalysts, 2 g/chamber; Vapor pressure of reactant, 80 Torr; Temperature, 140 °C; Flow, 40 mL/min; Data collection, after 30 min.
Catalysts 13 01473 g012
Scheme 1. Different pathways for the oxidation of octanol. * represent the activated species.
Scheme 1. Different pathways for the oxidation of octanol. * represent the activated species.
Catalysts 13 01473 sch001
Figure 13. Energy profile for the activation of oxygen on the surface of the catalyst by co-adsorbed alcohol with (blue) or without (red) the promoting effect of by-product H2O during the progress of reaction. * represent the activated species.
Figure 13. Energy profile for the activation of oxygen on the surface of the catalyst by co-adsorbed alcohol with (blue) or without (red) the promoting effect of by-product H2O during the progress of reaction. * represent the activated species.
Catalysts 13 01473 g013
Figure 14. (a) Effect of oxygen on/off on the progress of reaction in the flow of O2/N2 (Orange: Oxidative Dehydrogenation, Blue: Direct Dehydrogenation); (b) experimental setup conditions: Catalysts, 2 g/chamber; Vapor pressure of reactant, 80 Tor; Temperature, 140 °C; Flow, 40 mL/min; Data collection, after 30 min.
Figure 14. (a) Effect of oxygen on/off on the progress of reaction in the flow of O2/N2 (Orange: Oxidative Dehydrogenation, Blue: Direct Dehydrogenation); (b) experimental setup conditions: Catalysts, 2 g/chamber; Vapor pressure of reactant, 80 Tor; Temperature, 140 °C; Flow, 40 mL/min; Data collection, after 30 min.
Catalysts 13 01473 g014
Table 1. Calculated lattice constants (ao), ground state energy (E0), ground state volume (V0), and bulk moduli (B) of the ZnFe2O4, ZnMnFe2O4 and MnFe2O4 ferrites.
Table 1. Calculated lattice constants (ao), ground state energy (E0), ground state volume (V0), and bulk moduli (B) of the ZnFe2O4, ZnMnFe2O4 and MnFe2O4 ferrites.
CompoundPresentLiterature
ZnFe2O4
a (Å)8.438* 8.441 [30,31]
E0 (eV)−2.17 × 105
V (Å)3599.07
B0 (Gpa)148.167
ZnMnFe2O4
a (Å)8.4410* 8.368 [32]
E0 (eV)−2.35 × 105
V (Å)3601.425
B0 (Gpa)174.039
MnFe2O4
a (Å)8.4472* 8.4512 [33]
E0 (eV)−2.52 × 105
V (Å)3602.751* 603.608 [33]
B0 (Gpa)216.738** 146 [34]
* experimental, ** theoretical.
Table 2. Efficiency of MnFe2O4, ZnFe2O4, ZnMnFe2O4 as catalysts for a variety of organic substrates.
Table 2. Efficiency of MnFe2O4, ZnFe2O4, ZnMnFe2O4 as catalysts for a variety of organic substrates.
EntryReaction ConditionsSubstrate/
Reaction
Product* C** SRef.
1MnFe2O4 (1 g), acetonitrile, H2O2 (10 mmole), temp (80 °C), time (10 h)BzOH/
Oxidation
BzH70100[41]
2MnFe2O4 (0.015 g), acetonitrile (20 mL) HMF (1 mmole), TBHB (9 mmole), stirring (1000 rpm), temp (100 °C), time (5 h)5-hydroxy methylfurfural/oxidation2,5-furandicarboxylic acid(FDCA)6885[42]
3MnFe2O4 (10 mmole), TBHP 70%, solvent, DMSO, temp (90 °C),
time (5 h)
Benzylic alcohol/oxidation benzophenone90100[43]
4MnFe2O4 (1 g), H2O2 (10 mole), acetonitrile (10 mL), temp (80 °C),
time (10 h)
benzyl alcohol/oxidationbenzaldehyde8089.47[44]
5MnFe2O4 (3 g/L), potassium peroxydisulfate (5 mg/L),
time (6 h), room temperature pH (3–3.5)
Phenol/
oxidation
Organic carbon and H2O 10095 [45]
6ZnFe2O4 (1 g), absolute ethanol (200 mL), temp (355–380 °C), time (4 h) 1-butene/oxidation 1,3-butadiene 7594[46]
7ZnFe2O4, temp (450 °C),
Time (6 h), pH (6–10), air (oxygen source)
n-butane/oxidation1,3-butadiene 4580.2[47]
8MnZnFe2O4 (30 mg, 0.13 mmole), temp (80 °C),
t-BuOOH (5.0 mmol)
1-phenylethanol/
Oxidation
Acetophenone8099[20]
9i. MnFe2O4
ii. ZnFe2O4
iii. ZnMnFe2O4
Reaction conditions: catalysts, 2 g/chamber; Vapor pressure of reactant, 80 Torr; Temperature, 140 °C; Flow, 40 mL/min; Data collection; after 30 min.
Octanol/
Oxidation
Octanali. 74
ii. 59
iii. 91
i. 100
ii. 100
iii. 100
Recent work
* C; conversion and ** S; selectivity.
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Bibi, M.; Sadiq, M.; Rizk, M.A.; Alsaiari, R.A.; Iqbal, Z.; Ali, Z. Tuning the Magnetic and Catalytic Properties of Manganese Ferrite through Zn2+ Doping: Gas Phase Oxidation of Octanol. Catalysts 2023, 13, 1473. https://doi.org/10.3390/catal13121473

AMA Style

Bibi M, Sadiq M, Rizk MA, Alsaiari RA, Iqbal Z, Ali Z. Tuning the Magnetic and Catalytic Properties of Manganese Ferrite through Zn2+ Doping: Gas Phase Oxidation of Octanol. Catalysts. 2023; 13(12):1473. https://doi.org/10.3390/catal13121473

Chicago/Turabian Style

Bibi, Mehnaz, Muhammad Sadiq, Moustafa A. Rizk, Raiedhah A. Alsaiari, Zaffar Iqbal, and Zahid Ali. 2023. "Tuning the Magnetic and Catalytic Properties of Manganese Ferrite through Zn2+ Doping: Gas Phase Oxidation of Octanol" Catalysts 13, no. 12: 1473. https://doi.org/10.3390/catal13121473

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop