Next Article in Journal
Vinylogous and Arylogous Stereoselective Base-Promoted Phase-Transfer Catalysis
Next Article in Special Issue
Controlling the Structural Properties and Optical Bandgap of PbO–Al2O3 Nanocomposites for Enhanced Photodegradation of Methylene Blue
Previous Article in Journal
Assessment on the Effect of Sulfuric Acid Concentration on Physicochemical Properties of Sulfated-Titania Catalyst and Glycerol Acetylation Performance
Previous Article in Special Issue
Floating Carbon-Doped TiO2 Photocatalyst with Metallic Underlayers Investigation for Polluted Water Treatment under Visible-Light Irradiation
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Removal of Organic Dyes from Water and Wastewater Using Magnetic Ferrite-Based Titanium Oxide and Zinc Oxide Nanocomposites: A Review

by
António B. Mapossa
1,*,
Washington Mhike
2,
José L. Adalima
3 and
Shepherd Tichapondwa
4
1
Department of Chemical Engineering, Institute of Applied Materials University of Pretoria, Pretoria 0002, South Africa
2
Department of Chemical, Metallurgical and Materials Engineering (Polymer Technology Division), Tshwane University of Technology, Pretoria 0183, South Africa
3
Department of Archaeology and Anthropology, Eduardo Mondlane University, Maputo 257, Mozambique
4
Department of Chemical Engineering, Water Utilization and Environmental Engineering Division, University of Pretoria, Pretoria 0002, South Africa
*
Author to whom correspondence should be addressed.
Submission received: 12 November 2021 / Revised: 6 December 2021 / Accepted: 13 December 2021 / Published: 18 December 2021

Abstract

:
Heterogeneous photocatalysis using titanium dioxide (TiO2) and zinc oxide (ZnO) has been widely studied in various applications, including organic pollutant remediation in aqueous systems. The popularity of these materials is based on their high photocatalytic activity, strong photosensitivity, and relatively low cost. However, their commercial application has been limited by their wide bandgaps, inability to absorb visible light, fast electron/hole recombination, and limited recyclability since the nanomaterial is difficult to recover. Researchers have developed several strategies to overcome these limitations. Chief amongst these is the coupling of different semi-conductor materials to produce heterojunction nanocomposite materials, which are both visible-light-active and easily recoverable. This review focuses on the advances made in the development of magnetic ferrite-based titanium oxide and zinc oxide nanocomposites. The physical and magnetic properties of the most widely used ferrite compounds are discussed. The spinel structured material had superior catalytic and magnetic performance when coupled to TiO2 and ZnO. An assessment of the range of synthesis methods is also presented. A comprehensive review of the photocatalytic degradation of various priority organic pollutants using the ferrite-based nanocomposites revealed that degradation efficiency and magnetic recovery potential are dependent on factors such as the chemical composition of the heterojunction material, synthesis method, irradiation source, and structure of pollutant. It should be noted that very few studies have gone beyond the degradation efficiency studies. Very little information is available on the extent of mineralization and the subsequent formation of intermediate compounds when these composite catalysts are used. Additionally, potential degradation mechanisms have not been adequately reported.

1. Introduction

Nowadays, concern around efficient management of water use has become topical, with interest not only limited to agricultural and industrial sectors but also attracting public health and sustainable economic development proponents.
The wide range of anthropogenic activities that use water results in the generation of highly toxic effluents, which are rich in organic pollutants, rendering them unsuitable for reuse in agricultural activities and human consumption. As a result, water decontamination has become the focus of attention for several studies.
Chemical oxidation-based methods have been proposed as possible remediation techniques for the treatment of domestic and industrial effluents since they are rich in toxic organic compounds, resulting in reduced environmental pollution and enabling the recycling of water resources [1,2,3].
Advanced oxidative processes (AOPs), which include photocatalysis, plasma oxidation, Fenton’s reactions, and ozonation, have been used to degrade numerous pollutants in both water and wastewater treatment applications [2]. Among these, photocatalysis is one of the most widely researched and used AOPs.
Photocatalysis typically involves the irradiation of semi-conductor catalysts with UV or visible light, resulting in the transfer of electrons from the valence band to the conduction band. For charge separation to occur, the energy of the photons must be greater than the band gap energy of the catalyst. The holes (h+) are capable of reacting with water molecules, leading to the formation of highly reactive hydroxyl radicals (•OH) [3,4].
Titanium dioxide (TiO2) and zinc oxide (ZnO) are amongst the most widely studied photocatalysts due to their high photocatalytic activity, strong oxidation potential, superhydrophilicity, biological and chemical stability, prolonged durability, non-toxicity, and low cost [4,5,6,7,8,9].
While these catalysts demonstrate excellent photocatalytic activity, their particle sizes, often in the nano range, negatively impact their recovery. As a result, photocatalysis using these catalysts has limited the large-scale application in treatment of polluted water and wastewater due to the costs associated with loss of the material [4,10,11]. This has necessitated the development of new, inexpensive materials with good photocatalytic efficiency, recoverability, and reusability properties.
Nanomaterials that possess good magnetic properties while exhibiting photocatalytic activity comparable to pure TiO2 and ZnO are the most ideal photocatalysts as they can be recovered from the reactor using a magnetic field [4].
In particular, ferrite-based nanoparticles have attracted significant attention from researchers in different applications such as sensors, biomedical, catalysis, and energy storage devices. This is due to their excellent adsorption capacities, high surface area, optical properties [12], and magnetic properties [12,13].
The ferrites also present chemical and thermal stability [14,15]. Therefore, nanocomposites of magnetic ferrites-based metal oxides (TiO2 and ZnO) have been synthesized with the aim of developing photocatalysts with enhanced properties, thereby limiting the deficiencies presented by neat materials.
Several studies show that doping magnetic nanoparticles with ZnO or TiO2 enhanced the photocatalytic performance of the metal oxides through changes in optical properties, increased surface defects, production of surface oxygen vacancies, and impeding recombination of charge carriers [16,17].
A study conducted by Rahmayeni et al. [18] investigated ferromagnetic NiFe2O4 doped with diamagnetic ZnO that resulted in superparamagnetic behavior being imparted to the synthesized photocatalyst. They subsequently demonstrated that these properties made it easier to separate the ZnO/NiFe2O4 nanocomposite from the reaction mixture.
The present work focuses on the development of magnetic spinel ferrite-based zinc oxide and titanium oxide nanocomposites used in the photodegradation of dyes for use in water and wastewater treatment. In particular, principal synthesis and characterization methods are explored.
The recovery and reusability of the magnetic nanocomposite catalysts are reported and discussed. Furthermore, the review also discusses the often overlooked social dimension about the water and wastewater treatment. Finally, conclusions are drawn, and challenges encountered in the use of these catalysts are cited.

2. Magnetic Ferrites Nanoparticles

Ferrites can be divided into four structural groups, which include hexagonal (MFe12O19), orthogonal (MFeO3), and garnet (M3Fe5O12), where M represents metal ions, i.e., Ba2+ and Pb2,+ and spinel (AB2O4) [19]. The cations represented by A and B occupy tetrahedral and octahedral sites. Figure 1a shows both tetrahedral and octahedral sites coordinated to an oxygen atom [12].
In a single unit cell, there are 64 tetrahedral and 32 octahedral sites where only 8 and 24 sites are occupied by cations, respectively [12]. The distribution of cations in the tetrahedral and octahedral sites of spinel structure can be classified into three groups i.e., normal, inverse, and mixed.
Therefore, normal spinel (Figure 1b) is ZnFe2O4 where the tetrahedral sites are filled by Zn2+, and the octahedral sites are occupied by Fe3+ [12]. Furthermore, the common example for inverse spinel ferrite (Figure 1c) is nickel ferrite, wherein the tetrahedral sites are occupied by half of the Fe3+ and octahedral sites are occupied by both Ni2+ cations and the other half of Fe3+ [20].
NiFe2O4 demonstrates ferrimagnetic behavior with a curie temperature (TC) of approximately 858 K, whereas ZnFe2O4 displays antiferromagnetic behavior ordering below the curie temperature of 9 K [21].
An example of a mixed spinel ferrite is MnFe2O4, its structure is represented as follows: ([Mn0.8Fe0.2][Mn0.2Fe1.2]O4), where the +2 and +3 ions are distributed randomly on both sites [22]. CoFe2O4 can be found as an inverse or normal spinel structure; this is dependent on the synthesis method used.
Ferrites have attracted more attention in catalysis for water and wastewater treatment due to their various advantages such as the presence of a band gap capable of absorbing visible light, together with the spinel crystal structure, which increases photocatalytic activity due to the catalytic sites t featured on the crystal lattice [23].
Besides the magnetic behavior, other properties that spinel ferrites possess include spin canting effect, spin-glasslike behavior, and higher heat and corrosion resistance [12,22]. The common band gap energies for some spinel ferrites are as follows: CuFe2O4 (1.32 eV), ZnFe2O4 (1.92 eV), CoFe2O4 (2.27 eV), and NiFe2O4 (2.19 eV) [23,24].
Due to the small band gap energy of ferrites, which makes them effective under absorption of visible light irradiation, they are extremely suitable for the removal of organic pollutants in water and wastewater treatment processes [22,23].

2.1. Methods of Synthesis of Magnetic Spinel Ferrites

Synthesis methods play an important role in the development of magnetic nanoparticles as this controls the electrical, optical, and magnetic properties of the material [15].
Additionally, synthesis methods should pay particular attention to the cost of production. An optimum balance between processing costs and the desired nanomaterial properties is desired [15,26].
Numerous synthesis methods have been used to prepare magnetic nanoparticles (MNPs); these include microwave [27,28,29], sonochemical [30], sol-gel [31,32,33,34], co-precipitation [35,36], combustion [37,38], micro-emulsion [39], and hydrothermal [40,41,42,43,44]. Amongst these, co-precipitation, sol-gel, hydrothermal, and combustion methods are the most widely reported.
Co-precipitation requires careful monitoring of pH in order to obtain pure spinel ferrites [25]. Advantages associated with this method include low cost, short synthesis time, high product yield, and production of uniformly sized particles [45].
For example, El-Okr et al. [46] synthesized magnetic CoFe2O4 using the co-precipitation method and obtained crystallites between 11 and 45 nm in size, with saturation magnetization ranging from 5 to 67 emu/g. The authors reported that the difference in crystallite size and saturation magnetization (Ms) values was associated with the variation of parameters such as pH and calcination temperature.
The hydrothermal synthesis method enables particle size control and flexibility in terms of surface modification. It is based on the wet-chemical synthesis; typically, this occurs in sealed reactors or autoclaves at high vapor pressures (from 0.3 to 4 MPa) and elevated temperatures (130 to 250 °C) [47,48].
Some noteworthy advantages of this method include low temperature for synthesis, high purity, simple reactions, cost-effectiveness, and good dispersibility of the MNPs [49]. For instance, Zhao et al. [41] prepared cobalt ferrite using the hydrothermal method. The resultant material had 70 nm crystallites with a saturation magnetization (Ms) of 86 emu/g.
The sol-gel method is extensively used for the synthesis of spinel ferrites [50,51,52,53,54,55]. The process involves the transition of a system from a liquid phase (sol) to a solid phase (gel), through chemical reactions such as hydrolysis and condensation polymerization of the metallic precursors [25]. Thus, its widespread acceptance is driven by the low cost associated with the method, better homogeneity, composition control, and narrow particle size distribution at relatively low temperatures [25,45].
The sol–gel method also allows for good control of the structural and magnetic properties of MNPs [45]. Sajjia et al. [56] prepared cobalt ferrite nanoparticles by a sol-gel method. Their results demonstrated that the saturation magnetization was 67.3 emu/g and the particle sizes were between 7 and 28 nm, according to the calcination temperature of nanoparticles.
The combustion synthesis method of MNPs is based on the thermodynamics principles and chemistry of propellants and explosives [57,58,59]. The method requires a powdered mixture, typically consisting of an oxidizing agent containing the metal ions of interest such as oxidizing reagents, and a reducing agent such as urea or glycine [57,60,61,62,63].
Upon ignition, the mixture undergoes an exothermic or endothermic redox reaction depending on the material properties. Equation (1) shows the combustion reaction used to produce ZnFe2O4 by Mapossa et al. [64].
Zn(NO3)2·6H2O(s) + 2Fe(NO3)3·9H2O(s) + 6.67(NH2)2CO(s) → ZnFe2O4(s) + 6.67CO2(g) + 37.33H2O(g) + 10.67N2(g)
Using the principles of propellant chemistry, the valences of the elements present in the oxidizing and reducing agents are represented as follows: Zn = +2; Fe = +3; C = +4; H = +1; and O = −2. The total valence of urea is equal to +6, and the total valence of nitrates (Zn(NO3)2/Fe(NO3)3, 1/2) for oxidizing agent zinc ferrite (ZnFe2O4) is equal to −40.
Thus, the combination of total valences of reducing agent (fuel, urea) and oxidizing agent achieved gives the following expression: −40 + 6n = 0, where n is the number of moles of urea (in this case, n = 6.67 mol) for the combustion reaction. Since it is complete combustion, the stoichiometric reaction (Φe = 1) follows the definition described for the oxygen balance and equals zero.
To accomplish this, the content of oxygen from the nitrates is completely oxidized by a reducing agent (fuel) in the mixture [61].
A study conducted by Salunkhe et al. [65] evaluated the magnetic properties of cobalt ferrite nanoparticles prepared using the combustion method using glycine as fuel. The results showed that the crystallite size was 38 nm and saturation magnetization was 67.3 emu/g. Table 1 summarizes the structural and magnetic properties of CoFe2O4 MNPs synthesized using various methods.

2.2. Characterization Methods

Before the synthesized magnetic materials are used as catalysts, an investigation of the various properties, which influence their performance, is essential. Some of the key parameters are size, shape, and surface area.
Therefore, this information can be elucidated using one or a combination of the following techniques: X-ray diffraction (XRD), scanning electron microscopy (SEM), transmission electron microscopy (TEM), BET-N2 analysis, vibrating sample magnetometer (VSM).
X-ray diffraction (XRD) gives information about the structural properties, crystallite size, and crystalline phases of magnetic nanoparticles as catalysts.
The BET-N2 adsorption-desorption isotherm is a technique commonly used to evaluate the porosity and specific surface area of MNPs. For example, smaller particles have a larger surface area, leading to higher photocatalytic activity due to a larger number of active sites [23]. More details are explained in Section 5.
The magnetic behavior of the ferrites (saturation magnetization) is evaluated using a vibrating sample magnetometer (VSM). The information obtained from this technique gives an idea of the recovery potential of the photocatalyst.
Figure 2 shows the magnetic behavior of various ferrites at room temperature. With the exception of zinc ferrite, which has a low magnetization of saturation, the other ferrites have a high magnetization of saturation values, which implies good magnetic and recyclability properties [67].
The morphology, including shape, and particle size of magnetic nanoparticles, can be determined using transmission electron microscopy (TEM) and scanning electron microscopy (SEM). Figure 3a,b demonstrates the difference in morphology and structure of cobalt ferrite affected by different methods of synthesis [68].
Thermal decomposition studies are crucial for the development of nanocatalysts as most of them are synthesized at high temperatures. Thermogravimetric analysis (TGA) or differential scanning calorimetry (DSC) are used to determine the optimum temperature required for the synthesis of magnetic nanoparticles due to the possibility of a loss of activity during their preparation.
Finally, these characterization techniques are also employed to investigate if any changes occur by decomposition or degradation, and whether they retain their magnetic properties after the photocatalytic process.

3. Photocatalytic Application of Magnetic Ferrites and Their Nanocomposites

Photocatalytic degradation is a sequence of chemical reactions promoted by light resulting in the breakdown of the target compound [69].
The photocatalytic activity is effectively dependent on the surface area and electron-holes separation efficiency of the catalyst [69]. Figure 4 shows a schematic of the photocatalytic mechanism of magnetic ZnFe2O4/ZnO nanocomposite during methylene blue and methylene orange dye degradation [69].
It has been reported that nanocatalysts that have high surface areas exhibit higher photocatalytic activity compared to their larger counterparts with lower surface area. The smaller nanoparticles support the easy transition of electrons from the valence to conduction band, thereby generating electron-hole pairs when exposed to UV-visible radiation.
The generated electrons and holes further interact with dissolved oxygen and water to produce highly reactive free radical species capable of degrading the methylene dyes [69,70]. The general photocatalytic degradation mechanism of magnetic nanocomposites towards organic pollutants is demonstrated by Equations (2)–(8).
ZnO + → h+ + e
ZnFe2O4/ZnO → ZnFe2O4/ZnO (e)
e + O2 absorbed → •O2
•O2 + MB dye → degraded products
h+ + H2Oabsorbed → H+ + •OH
h+ + OHabsorbed → •OH
•OH + MB dye → degraded products
Reduction and oxidation take place at the photo-excited surface of the photocatalyst. Recombination between e and h+ can occur for the use of redox reaction. The e and h+ that do not recombine are transferred to the surface of redox reaction and undergo reduction process and oxidation process to form superoxide ion (O2) and ·OH, respectively. OH then leads to the production of strong oxidizing ·OH radicals. Meanwhile, the negative e reacts with the oxygen (O) molecule to form a ·O2. This ·O2 also produces ·OH radicals via the formation of HO2 radicals and H2O2. The radicals formed from the reaction are used to degrade the organic pollutant [71,72].

3.1. Nickel Ferrite and Nanocomposites

NiFe2O4 has generated a lot of interest because of its excellent features. These include being a soft ferrimagnetic or ferrite n-type semiconductor with low coercivity, chemical stability, and electrical resistivity. These make it an excellent material in different applications such as in magnetic resonance imaging enhancement, magnetic recording media, and electronic devices, as well as in catalysis [20].
Following the description in Section 2, NiFe2O4 is completely composed of an inverse spinel structure comprising a face-centered cubic lattice. NiFe2O4 consists of tetrahedral sites occupied by half of the Fe3+ cations, while the rest of the Fe3+ and Ni2+ cations are distributed over the octahedral sites [73,74,75].
Figure 5 shows the XRD spectra and SEM images of neat zinc oxide, nickel ferrite, and their nanocomposites [76]. In the study conducted by Adeleke et al. [76], no secondary peaks or secondary phases of material were observed; this demonstrated the effectiveness of the synthesis method used in this study.
Furthermore, SEM images demonstrated the effect of doping ferrite with zinc oxide. The high degree of agglomeration and different morphologies observed on the ZnO/Fe2O4 catalyst were attributed to the magnetic attraction between nickel ferrite and zinc oxide layers [76].
Several studies have demonstrated that magnetic NiFe2O4 and its nanocomposites are effective photocatalysts for the removal of dye from water and wastewater, due to their high adsorption capacity and strong photocatalytic properties [16,77,78,79,80,81,82].
Khosravi and Eftekhar [83] synthesized magnetic NiFe2O4 using a sol-gel method and evaluated its effectiveness as an adsorbent for the removal of Reactive Blue 5 (RB5) dye. Parameters such as pH, temperatures, and catalyst concentration were evaluated during RB5 degradation [83]. Maximum degradation (90%) was achieved under acidic conditions (pH = 1) at room temperature using an (adsorbent/catalyst loading of 0.03 g/L).
These findings were corroborated by Zhu et al. [16] when they evaluated the photocatalytic degradation of Congo Red dye using NiFe2O4/ZnO as a catalyst. In their study, the NiFe2O4/ZnO nanocomposite resulted in a 94% removal of Congo red solution under simulated solar light irradiation in 10 min. Nickel ferrite was also shown to be effective when coupled with other metal oxides such as TiO2.
In a study done by Hung and Thanh [84], a magnetic nanocomposite of NiFe2O4/TiO2 degraded 98% of methyl orange dye after 14 h of UV or visible light irradiation. Although the reaction time was rather long, the photocatalyst had a high saturation of magnetization (40 emu/g), which makes it easily recyclable for reuse.
The results obtained in these studies demonstrate that nickel ferrite nanocomposites are potential candidates for wastewater treatment in large-scale applications.
Additional studies that illustrate the efficacy of nickel ferrite, nickel ferrite-based titanium oxide, and zinc oxide catalysts in the degradation of variant organic pollutants are summarized in Table 2.

3.2. Zinc Ferrite and Nanocomposites

Zin ferrite has a small bandgap of around 1.9 eV, which gives a good response to the visible light, as well as excellent photochemical stability, considerable magnetism, and cost-effectiveness. As a result, it has also attracted attention by researchers in the photocatalysis process [85].
The compound consists of a fully normal spinel structure, where its tetrahedral sites are occupied only by Zn2+ cations and the Fe3+ ions are distributed in the octahedral sites [86]. Figure 6 shows the XRD patterns and SEM micrographs of the zinc oxide-, zinc ferrite-, and zinc ferrite-based zinc oxide nanocomposites [17].
The photocatalyst ZnFe2O4/ZnO had diffraction peaks similar to those of neat ZnFe2O4 and ZnO. The sharp peaks showed good crystallinity of the nanocomposite [17], which demonstrates the efficacy of the synthesis method used in this study.
The SEM images revealed the effect of incorporating ZnFe2O4 nanoparticles into the pores of the ZnO matrix. The authors observed that a high content of ZnO nanoparticles was formed; these were better defined in the ZnFe2O4/ZnO nanocomposites [17].
Significant efforts have been devoted to investigating ZnFe2O4-based photocatalysts for water and wastewater treatment, with the aim of removing organic pollutants [87,88,89,90].
Yuan et al. [91] investigated the photocatalytic activity of a ZnFe2O4/TiO2 nanocomposite where the pure ZnFe2O4 and TiO2 were obtained via the co-precipitation method. The results showed that the ZnFe2O4/TiO2 and pure TiO2 resulted in 95% and 20% of degradation of phenol respectively during 180 min of irradiation under UV–Visible. This demonstrated that the ZnFe2O4/TiO2 nanocomposite catalyst was more effective than pure TiO2 in the degradation of phenol.
In addition, Shao et al. [85] evaluated the application of ZnFe2O4/ZnO nanoparticles in the photodegradation of methylene blue dye. The findings demonstrated that even after three cycles, the photocatalytic activity of the magnetic nanocomposite ZnFe2O4/ZnO (65%) was better compared to that of pure ZnO (58%), indicating the significance of ZnFe2O4 in the suppression of ZnO photo-corrosion. This was attributed to the photostability of ZnFe2O4 nanoparticles [92].
Further study by Patil et al. [93] investigated the photocatalytic activity of ZnFe2O4 nanoparticles synthesized using a combustion method. Their activity was tested on synthetic wastewater made up of the following dyes: Methylene Blue, Rose Bengal, Evans Blue, and Indigo Carmine.
Degradation efficiencies of 98, 99, 82, and 87% were recorded for each dye, respectively [93]. The antibacterial activity of ZnFe2O4 against diverse gram-negative bacterial strains such as Staphylococcus aureus, Pseudomonas aeruginosa, Escherichia coli, and Bacillus was also investigated. A variation in the antibacterial activity towards the different bacterial strains was observed [93].
Other work conducted by Sripriya et al. [94] reported on the excellent photocatalytic performance of ZnFe2O4 in the degradation of 4-chlorophenol (4-CP). They further reported that factors such as particle size and surface area significantly affected the activity. Further studies on the use of magnetic ZnF2O4 nanocomposites as photocatalysts are listed in Table 2.

3.3. Cobalt Ferrite and Nanocomposites

Cobalt ferrite is a hard ferrimagnetic material that has a face centered cubic structure [95]. It exists as a normal spinel structure or inverse spinel structure, depending on the synthesis method. In the normal spinel structure, the Co2+ ions are distributed in the tetrahedral sites, while the octahedral sites are occupied by Fe3+ ions.
For the inverse spinel structure, the tetrahedral sites are occupied by half of the Fe3+ ions and the rest of the octahedral sites are distributed by Fe3+ and Co2+ ions. It is important to note that the magnetic properties vary depending on the structures [95,96].
Figure 7 shows the XRD patterns and TEM micrographs of the cobalt ferrite and cobalt ferrite-based zinc oxide nanocomposite [97]. Pristine XRD patterns were obtained with no impurities; this demonstrated that the co-precipitation method used for synthesis was highly efficient [97].
Several studies have demonstrated the potential of cobalt spinel ferrite doped with metals oxides (TiO2 or ZnO) as photocatalysts for water and wastewater treatment for organic pollutants removal. This is due to the various attributes that include chemical stability, small band-gap energy that leads to activation by visible light [5,12,98,99,100], magnetic properties, and higher surface area.
A study conducted by de Oliveira et al. [4] demonstrated that the magnetic nanocatalyst of CoFe2O4 coupled to TiO2 resulted in 100% degradation of diuron degradation. This study also observed that CoFe2O4/TiO2 nanoparticles displayed good saturation magnetization, demonstrating that they can be easily separated for reuse.
Furthermore, Li et al. [101] reported good performance of magnetic TiO2/CoFe2O4 nanocomposite for methylene blue (MB) degradation (98%) in 300 min. The good performance of magnetic TiO2/CoFe2O4 nanocomposite was attributed to the presence of CoFe2O4, which not only improved the UV light absorbance but also enhanced the response to the visible light region.
A study done by Chandel et al. [102] reported that the ZnO/CoFe2O4 nanocomposite displayed 94% degradation efficiency towards Methylene Orange (MO) and 92% removal for malachite green (MG) dye. The authors attributed the hydroxyl radicals (OH) and holes (h+VB) as the main reactive species responsible for the degradation of MO and MG dyes.
The photocatalytic activity and stability of ZnO/CoFe2O4 were confirmed by showing 10 cycles for successive reuse. Since ZnO/CoFe2O4 is recyclable and easily recovered magnetically; it is a good candidate for use as a low-cost photocatalyst for water and wastewater treatment.
Additional studies on the photocatalytic performance of cobalt ferrite and their nanocomposites in the degradation of organic pollutants are summarized in Table 2.

3.4. Manganese Ferrite and Nanocomposites

MnFe2O4 is a soft spinel ferrite with high magnetic permeability and moderate saturation magnetization, high chemical stability, high electrical resistance, and special optical properties [103]. A combination of these factors makes it attractive for use in different applications such as biomedical drug delivery and catalysis [103,104].
MnFe2O4 is considered a mixed spinel ferrite in which the tetrahedral and octahedral sites are both are occupied by Mn2+ and Fe3+ ions [104,105]. High reaction temperatures affect the synthesis of the magnetic ferrite, resulting in variation in particle sizes of the material, which in turn affects other parameters such as saturation magnetization.
A study conducted by Chang et al. [106] demonstrated through X-ray diffraction analysis that no other peaks attributed a second phase were observed for manganese ferrite and titanium oxide. Additionally, peaks attributed to pure materials (MnFe2O4 and TiO2) were observed in the XRD pattern of MnFe2O4/TiO2 nanocomposite (Figure 8).
The findings demonstrate that the magnetic nanoparticles were successfully synthesized via a hydrothermal followed by the sol-gel method [106]. The SEM micrograph of MnFe2O4/TiO2 (Figure 9) showed that agglomerated spherical particles were produced.
The MnFe2O4 nanoparticles display a well-defined morphology with particle size around 15–20 nm. Meanwhile, it is clearly visible that the particle size of MnFe2O4/TiO2 is uneven and relatively large [106].
The TEM micrographs of photocatalyst MnFe2O4/TiO2 demonstrated that MnFe2O4 nanoparticles are coated by TiO2 with shape a core-shell structure (see Figure 9) [106].
Numerous studies have demonstrated that manganese ferrite-based metal oxide nanocomposites are effective in the photocatalytic degradation of organic dyes. For example, Zamani et al. [107] evaluated the photocatalytic performance of a magnetic MnFe2O4/ZnO nanocomposite for Congo red dye (CR) removal.
The results showed that 90% degradation of Congo red dye was achieved in 35 min under UV-vis irradiation. Additionally, Arief et al. [103] showed that the same nanocomposite was effective for Rhodamine B dye degradation (95%). This was attributed to the presence of a narrow band gap energy (1.95 eV) of MnFe2O4/ZnO.
Silambarasu et al. [108] tested the performance of MnFe2O4 on the degradation of methylene blue dye. The manganese ferrite achieved 96% dye decolorization and exhibited saturation magnetization (Ms) of 39.7 emu/g. The magnetic properties indicated that the product could be easily recovered for potential reuse.
There are few studies reporting the application of manganese ferrite and manganese ferrite-based zinc oxide and titanium oxide nanocomposites in water and wastewater treatment for organic pollutants removal.

3.5. Copper Ferrite and Nanocomposites

CuFe2O4 is one of the magnetic nanoparticles that has become a promising candidate in the catalysis field due to the presence of surface hydroxyl groups, good chemical, and thermal stabilities, a small band gap, and magnetic properties [109]. These characteristics make it attractive as a photocatalyst for work on water and wastewater treatment for organic pollutants degradation.
Copper ferrite has an inverse spinel structure, where the tetrahedral sites are occupied by half of the Fe3+ ions and the rest of the octahedral sites are occupied by Fe3+ and Cu2+ ions [110]. Therefore, besides the cubic crystal structure, the copper ferrite also presents a tetragonal crystal structure that depends on the synthesis method and annealing temperature [110].
In the XRD patterns of TiO2/CuFe2O4 nanocomposites, the peaks associated with the neat TiO2 and CuFe2O4 were observed without any further secondary phase (Figure 10) [109]. This demonstrated that the nanocomposite photocatalyst was successfully prepared using the Sol-Gel method and the pure titanium oxide and copper ferrite nanoparticles remained with their structure during the synthesis processes.
Figure 11a,b shows the SEM micrographs of CuFe2O4 and TiO2/CuFe2O4, where the presence of agglomerated particles distributed randomly is apparent. The introduction of TiO2 in the CuFe2O4 influenced the morphology of the nanocomposite TiO2/CuFe2O4. The surface of the nanocomposite was much rougher than that of the CuFe2O4.
The TiO2 agglomerated on the surface of CuFe2O4 can provide more active sites for the nanocomposite and improve its photocatalytic activity during organic pollutants degradation [109].
Several studies have reported the photocatalytic activity of copper ferrite and its nanocomposites in water and wastewater treatment [109,111,112].
For instance, a study done by Anandan et al. [110] reported good performance of magnetic CuFe2O4 as photocatalyst for the degradation of methylene blue (MB) dye in the presence of peroxydisulphate under UV–vis light.
The activity of the copper ferrite was attributed to the effect of the peroxydisulphate in the photocatalyst, which improved the photocatalytic degradation of methylene blue (95%) 75 min. Before the addition of the oxidant peroxydisulphate to the cobalt ferrite, the nanoparticles showed 16% MB dye degradation in 75 min.
A recent study conducted by Janani et al. [113] evaluated the magnetic nanocomposite ZnO/CuFe2O4 as a catalyst for methylene blue dye degradation under visible light. In their study, they demonstrated that the ZnO/CuFe2O4 photocatalyst was efficient in the degradation of methylene blue dye (86%) in 77 min.
The authors associated the activity of the nanocomposites with the hydroxyl radicals and holes generated, which play a principal role in the degradation of the dye. Furthermore, the photocatalyst also remained stable after six cycles of reuse. More studies on the activity of copper ferrite nanocomposites for the degradation of organic pollutants are listed in Table 2.

3.6. Mixed-Metal Ferrites and Nanocomposites

The introduction of different cations in the spinel ferrite system is required to improve the physicochemical properties of spinel ferrites. For instance, the substitution of magnetic cations such as Mn2+, Ni2+, Co2+, and Cu2+, and diamagnetic ions such as Zn2+ and Cd2+, in spinel ferrites systems, changes the structural, morphological, opto-magnetic, and catalytic properties [114,115]. In general, this is attributed to the distribution of metallic ions in the tetrahedral and octahedral sites [115,116].
Ciocarlan et al. [117] evaluated the structural, morphological, and photocatalytic properties of magnetic nanoparticles Co0.5Zn0.25M0.25Fe2O4/TiO2, where M represent Ni2+, Cu2+, and Mn2+ ions. XRD patterns for TiO2-based magnetic nanocomposites showed that the introduction of TiO2 into the magnetic nanoparticles affected their structural properties (Figure 12).
TEM micrographs for the two magnetic nanocomposites revealed variations in morphology upon substitution of M2+ cations (Ni2+ and Cu2+) and introduction of TiO2 to the Co0.5Zn0.25M0.25Fe2O4 system (Figure 13).
Finally, in terms of photocatalytic activity, the results demonstrated that approximately 80% and 75% of methylene orange (MO) and methylene blue (MB) were effectively degraded by Co0.5Zn0.25Ni0.25Fe2O4/TiO2. The authors attributed the good photocatalytic activity to the Ni2+ ions and synergistic effect in combination with Co2+ ions [117].
Several other studies have also demonstrated the photocatalytic performance of complex-structured magnetic nanocomposites. For example, a Mn1−xNixFe2O4 catalyst with varying concentrations of nickel (x = 0.1, 0.2, 0.3, 0.4, and 0.5) was evaluated for indigo carmine dye degradation by Jesudoss et al. [118].
Amongst the obtained photocatalysts, the Mn0.5Ni0.5Fe2O4 catalyst exhibited higher photocatalytic performance in the degradation of indigo carmine dye, with 96% degradation within a 180 min period.
The material exhibited excellent saturation magnetization of 35.0 emu/g, demonstrating that this can be recoverable after a catalytic reaction. The authors concluded that the concentration of Ni2+ ions affected the structure of MnFe2O4, and that a high concentration of nickel ions reduced the crystallite size and increased the surface area, thereby affecting photocatalytic activity (Figure 14).
A study by Naik et al. [119] evaluated the performance of nanostructured zinc-doped cobalt ferrites (ZnxCo1−xFe2O4 with (x = 0.0 to 0.6 with the step of 0.2) in the photocatalytic degradation of Congo Red (CR) and Evans Blue (EB) dyes. They established that the photocatalytic performance of cobalt ferrite increased with an increase in Zn-doping up to x = 0.4 then decreased thereafter (Figure 15).
Additional studies revealed that higher Zn2+ doping concentrations increased the bactericidal properties of the CoFe2O4 towards human pathogens. For both Congo red (CR) and Evans Blue (EB) dyes, Zn0.4Co0.6Fe2O4 nanoparticles showed good photocatalytic activity in 150 min of irradiation time.
The study suggests that the synthesized nanoparticles are suitable for photocatalytic applications. More studies of photocatalytic activity of mixed metal ferrites nanocomposites are summarized in Table 2.
It is important to note that, besides the focus of the development of magnetic nanocomposites for water and wastewater treatment, new recent trend of development of new materials is emerging. For example, Mir et al. [138] studied how confining AuNPs in a porous Si template can significantly enhance the photocatalytic activity of MO. The pores prevent agglomeration of nanoparticles and eliminate the need for any functionalization. Confinement of the AuNPs in the Si nanocavities prevents electron–hole recombination and facilitates the transfer of hot carriers from the Si support to accelerate the photocatalytic efficiency.
The results showed that the recyclable, low-band gap photocatalytic system has economic and environmental advantages that promote implementation of catalytic and separation processes in continuous flow mode, with the advantages associated with easier phase separation and product recovery, enhanced safety, and easier operation [138].
Another recent trend of study explored functional elastomeric copolymer membranes designed by nanoarchitectonics approach for Methylene Blue Removal. The results demonstrated specific adsorption abilities up to 18 mg/g of grafted cyclodextrins [139].
The findings obtained in these studies show that more studies can be explored in order to develop new nanomaterials that are sustainable and safer for water and wastewater applications.

4. Factors Affecting the Photocatalytic Activity of Magnetic Nanocomposites

The photocatalytic activity of magnetic nanocomposites is dependent on several factors such as surface area, concentrations of dopant metal ions, pH, and catalyst loading.

4.1. Catalyst Surface Area

Nanomaterials with a high specific surface area exhibit enhanced catalytic activity. A decrease in particle size takes one to the increase of the surface area, which improves the dispersion of the nanomaterials in solution. This results in an enhanced photon absorbance, leading to the retrieval of their photocatalytic performance [140,141,142,143].
A study done by Padmapriya et al. [144] evaluated the effect of surface areas of magnetic nanoparticles system NixZn1−xFe2O4, with different concentrations of Ni2+ ions (0.0 ≤ x ≤ 1.0) in photocatalytic degradation of methylene blue dye (Figure 16a).
The results demonstrated that better photocatalytic activity was found for the nanocatalyst Ni0.6Zn0.4Fe2O4, which had a high surface area (36.6 m2/g) compared to the other samples. It can be understood from this study that surface area is also related to the active sites on the catalytic surface, which enhance the photocatalytic activity.
Manikandan et al. [115] demonstrated that the photocatalytic degradation of 4-Chlorophenol (4-CP) using the photocatalyst ZnFe2O4 was affected by the particle size and morphology of the catalyst.
Additionally, a study conducted by Jia et al. [145] reported that the photocatalytic activity of ZnFe2O4 nanoparticles for methylene blue dye degradation was related to the surface properties and surface defects of the photocatalyst.

4.2. Effect of Catalyst Amount

The efficiency photocatalytic reactions can be influenced by the amount of the catalyst used. Padmapriya et al. [144] reported that the photocatalytic degradation of methylene blue dye was found to increase with an increase in the amount of the magnetic catalyst Ni0.6Zn0.4Fe2O4 until a certain amount of nanocatalyst loading (Figure 16b).
However, further increase in the catalyst loading demonstrated negative influence of the degradation plateaued. The authors also reported that adding an amount of the magnetic nanocatalyst increased the active sites on the catalyst surface, which in turn increased the amount of •OH (hydroxyl) and •O2 (superoxide) radicals and degraded the methylene blue (MB) dye.
Furthermore, the excess of amount of nanocatalyst beyond the optimum may have resulted in the agglomeration of catalyst particles and generated turbidity, which resulted in the decrease of the photocatalytic degradation efficiency [140,142,146].

4.3. Effect of pH

Generally, the solution pH is an important variable in water and wastewater treatment as it has a significant influence on the photocatalytic degradation process of organic compounds [10]. The variation of pH alters the surface charge of heterogeneous catalysts and, consequently, changes the photocatalytic activity of catalyst [147,148].
Figure 16c shows the influence of different pH (3, 4, 5, 6, 7, 8, and 9) on the photodegradation of methylene blue using the nanomagnetic catalyst Ni0.6Zn0.4Fe2O4 in a study conducted by Padmapriya et al. [144]. The results from this study showed that high photocatalytic degradation efficiency was achieved at pH = 3, due to electrostatic attraction between the anionic dye (MB) and the positively charged surface of nanocatalyst.
The authors demonstrated that at pH values above 7, the nanocatalyst surface became negatively charged, leading to electrostatic repulsion between the methylene blue dye and the catalyst, which reduced the photocatalytic degradation efficiency. More studies demonstrating the same behavior were reported by Mirkhani et al. [149] and Suwarnkar et al. [150].

5. Reusability of the Magnetic Nanocatalyst

Heterogeneous photocatalysis technology is always looking for an ideal photocatalyst, one that is reusable and that possesses high photocatalytic efficiency, a large specific surface area, and ability to absorb visible light [138]. Thus, the recyclability of catalysts is one of the key steps towards the sustainable application of photocatalysts and development of heterogeneous photocatalysis technology for water and wastewater treatment. The recyclability of catalysts is also related to their actual operational costs.
Recently, several studies have demonstrated satisfactory recyclability of nanomagnetic nanoparticles via magnetic separation processes using a magnetic field [5,10,150,151,152].
Krishna et al. [100] reported the reusability of the CoFe2O4/TiO2 nanocatalysts for acid blue 113 (AB113) dye degradation through magnetic separation where its photocatalytic activity was found to be retained up to six consecutive cycles and without considerable loss of photocatalytic activity and stability (Figure 17).
Table 3 shows additional results of the activities of reused magnetic nanocomposites for organic photodegradation processes. Most of the magnetic nanocomposites are recyclable up to more than three runs, demonstrating their stability during their application for water and wastewater treatment for organic pollutants removal. Therefore, these studies are indicators for possible industrial or large-scale application.

6. The Overlooked Social Dimension

The focus of most water and wastewater-related research has been on the technical aspects of the problem and improvements in terms of water quality and in minimizing environmental and health impacts, with very limited attention to its basic social and cultural sustainability dimensions [158].
A study done by Wichelns et al. [159] demonstrated that there is a need for a paradigm shift from the ‘treatment for disposal’ to the ‘treatment for reuse’ since wastewater contains pollutants such as organic and inorganic compounds which may pose health risks if not well managed [159].
Additionally, even when wastewater is treated using advanced technologies and health risks are carefully addressed and controlled, irrespective of all scientific evidence, the social perception remains the driver of the success or failure of wastewater reuse schemes [158].
Depending on public perceptions, impressions, and attitudes, the development of a wastewater scheme can be supported or constrained. Negative public perception can prevent well-planned projects from moving forward. On the other hand, positive public perception, which leads to greater acceptance, is the key element for the successful implementation of wastewater treatment [159,160].
Saad et al. [158] reported that various local communities around the world have rejected several water and wastewater treatment projects by their governments due to inadequate community consultation which resulted in negative public perception.
In summary, it can be said that recognizing the role of the social base for wastewater management from risk reduction to reuse can have major implications, for example, on the choice and effectiveness of the technologies employed.
Added to this is the creation of economic incentives for the public and private sector institutions to invest in sanitation and to generate income for private operators as well as secure their sustainability [161]. This section may be divided by subheadings.

7. Conclusions and Recommendation

The development and application of magnetic ferrite-based titanium oxide and zinc oxide nanocomposite as catalysts are extremely promising for the removal of organic pollutants from water and wastewater, as shown by various studies presented in this review. Studies demonstrated that these catalysts can be prepared by different methods such as sol-gel, co-precipitation, hydrothermal, and combustion. However, the methods of synthesis are chosen based on their advantages. The magnetic nanoparticles (MNPs) have several advantages, including that they are easily separated by an external magnetic field without loss of the nanocatalyst, which can be reused up to several runs of experiments. In most of studies, the magnetic based titanium oxide and zinc oxide nanocomposite exhibited an excellent catalytic activity for organic pollutants removal. Additionally, some studies showed that these catalysts were even effective after more three successive cycling runs. The catalytic activity of the MNPs as catalysts is a direct outcome of its intrinsic characteristics as well as of its synthesis method; nevertheless, the catalytic performance can be influenced by conditions that are imposed on these materials to prepare them for a given application. Additionally, the method of synthesis plays a principal role in the physicochemical properties of the catalyst obtained. However, the synthesis of magnetic nanoparticles and their relevance for organic dyes removal from water and wastewater still require more investigation in order to achieve the optimum optimization for large-scale for subsequent practical applications. Finally, studies of the application of MNPs-based oxides nanocomposites in water and wastewater treatment are still few; however, more studies are still required. Additionally, with this technology in progress, scientists have enough supporting theory to upscale and provide a cleaner environment and safe drinking water to human populations.

Author Contributions

Conceptualization, A.B.M.; writing—original draft preparation, A.B.M.; writing—review and editing, A.B.M., W.M., J.L.A., and S.T.; supervision, A.B.M. and S.T.; project administration, S.T. and W.M.; funding acquisition, S.T. and W.M. All authors have read and agreed to the published version of the manuscript.

Funding

No funding grant was used for this work.

Data Availability Statement

Data sharing not applicable—no new data generated.

Acknowledgments

The authors thank the Department of Chemical Engineering, University of Pretoria, and Department of Chemical, Metallurgical and Materials, Tshwane University of Technology South Africa for all the technical and logistical support.

Conflicts of Interest

The authors declare no conflict of interest.

Abbreviations

AB113Acid Blue 113
CRCongo Red
4-CP4-Chlorophenol
DSCDifferential Scanning Calorimetry
EBEvans Blue
MsSaturation Magnetization
MNPsMagnetic Nanoparticles
MBMethylene Blue
MOMethylene Orange
MGMalachite Green
RB5Reactive Blue 5
SEMScanning Electron Microscopy
TGAThermogravimetric Analysis
TEMTransmission Electron Microscopy
TCCurie Temperature
UVUltraviolet
VSMVibrating Sample Magnetometer
XRDX-ray Diffraction

References

  1. Liu, Y.; Sun, L.; Wu, J.; Fang, T.; Cai, R.; Wei, A. Preparation and photocatalytic activity of ZnO/Fe2O3 nanotube composites. Mater. Sci. Eng. B 2015, 194, 9–13. [Google Scholar] [CrossRef]
  2. Cheng, M.; Zeng, G.; Huang, D.; Lai, C.; Xu, P.; Zhang, C.; Liu, Y. Hydroxyl radicals based advanced oxidation processes (AOPs) for remediation of soils contaminated with organic compounds: A review. Chem. Eng. J. 2016, 284, 582–598. [Google Scholar] [CrossRef]
  3. Cahino, A.M.; Loureiro, R.G.; Dantas, J.; Madeira, V.S.; Fernandes, P.C.R. Characterization and evaluation of ZnO/CuO catalyst in the degradation of methylene blue using solar radiation. Ceram. Int. 2019, 45, 13628–13636. [Google Scholar] [CrossRef]
  4. de Oliveira, P.L.; Lima, N.S.; de Melo Costa, A.C.F.; Cavalcanti, E.B.; de Sousa Conrado, L. Obtaining TiO2: CoFe2O4 nanocatalyst by Pechini method for diuron degradation and mineralization. Ceram. Int. 2020, 46, 9421–9435. [Google Scholar] [CrossRef]
  5. Wilson, A.; Mishra, S.R.; Gupta, R.; Ghosh, K. Preparation and photocatalytic properties of hybrid core–shell reusable CoFe2O4–ZnO nanospheres. J. Magn. Magn. Mater. 2012, 324, 2597–2601. [Google Scholar] [CrossRef]
  6. Dette, C.; Pérez-Osorio, M.A.; Kley, C.S.; Punke, P.; Patrick, C.E.; Jacobson, P.; Giustino, F.; Jung, S.J.; Kern, K. TiO2 anatase with a bandgap in the visible region. Nano Lett. 2014, 14, 6533–6538. [Google Scholar] [CrossRef] [PubMed]
  7. Farhadi, A.; Mohammadi, M.R.; Ghorbani, M. On the assessment of photocatalytic activity and charge carrier mechanism of TiO2@SnO2 core-shell nanoparticles for water decontamination. J. Photochem. Photobiol. A Chem. 2017, 338, 171–177. [Google Scholar] [CrossRef]
  8. Kazemi, M.; Mohammadizadeh, M.R. Simultaneous improvement of photocatalytic and superhydrophilicity properties of nano TiO2 thin films. Chem. Eng. Res. Des. 2012, 90, 1473–1479. [Google Scholar] [CrossRef]
  9. Ahmed, M.A.; El-Katori, E.E.; Gharni, Z.H. Photocatalytic degradation of methylene blue dye using Fe2O3/TiO2 nanoparticles prepared by sol–gel method. J. Alloys Compd. 2013, 553, 19–29. [Google Scholar] [CrossRef]
  10. Harraz, F.A.; Mohamed, R.M.; Rashad, M.M.; Wang, Y.C.; Sigmund, W. Magnetic nanocomposite based on titania–silica/cobalt ferrite for photocatalytic degradation of methylene blue dye. Ceram. Int. 2014, 40, 375–384. [Google Scholar] [CrossRef]
  11. Gong, H.; Chu, W. Determination and toxicity evaluation of the generated products in sulfamethoxazole degradation by UV/CoFe2O4/TiO2. J. Hazard. Mater. 2016, 314, 197–203. [Google Scholar] [CrossRef]
  12. Mmelesi, O.K.; Masunga, N.; Kuvarega, A.; Nkambule, T.T.; Mamba, B.B.; Kefeni, K.K. Cobalt ferrite nanoparticles and nanocomposites: Photocatalytic, antimicrobial activity and toxicity in water treatment. Mater. Sci. Semicond. Process. 2021, 123, 105523. [Google Scholar] [CrossRef]
  13. Sundararajan, R.; Srinivasan, V. Catalytic decomposition of nitrous oxide on CuxCo3−xO4 spinels. Appl. Catal. 1991, 73, 165–171. [Google Scholar] [CrossRef]
  14. Mapossa, A.B.; Dantas, J.; Silva, M.R.; Kiminami, R.H.; Costa, A.C.F.; Daramola, M.O. Catalytic performance of NiFe2O4 and Ni0.3Zn0.7Fe2O4 magnetic nanoparticles during biodiesel production. Arab. J. Chem. 2020, 13, 4462–4476. [Google Scholar] [CrossRef]
  15. Hajalilou, A.; Mazlan, S.A. A review on preparation techniques for synthesis of nanocrystalline soft magnetic ferrites and investigation on the effects of microstructure features on magnetic properties. Appl. Phys. A Mater. Sci. Process. 2016, 122, 1–15. [Google Scholar] [CrossRef]
  16. Zhu, H.Y.; Jiang, R.; Fu, Y.Q.; Li, R.R.; Yao, J.; Jiang, S.T. Novel multifunctional NiFe2O4/ZnO hybrids for dye removal by adsorption, photocatalysis and magnetic separation. Appl. Surf. Sci. 2016, 369, 1–10. [Google Scholar] [CrossRef]
  17. Maynez-Navarro, O.D.; Mendez-Rojas, M.A.; Flores-Cervantes, D.X.; Kuri, U.S.; Sanchez-Salas, J.L. Recyclable and Photocatalytic Properties of ZnFe2O4/ZnO for Wastewater Treatment and Disinfection. ChemistrySelect 2020, 5, 15167–15174. [Google Scholar] [CrossRef]
  18. Rahmayeni, R.A.; Stiadi, Y.; Jamarun, N.; Emriadi, A.S. Photocatalytic Performance of ZnO-ZnFe2O4 Magnetic Nanocomposites on Degradation of Congo Red Dye Under Solar Light Irradiation. J. Mater. Environ. Sci. 2017, 8, 1634–1643. [Google Scholar]
  19. Reddy, D.H.K.; Yun, Y.S. Spinel ferrite magnetic adsorbents: Alternative future materials for water purification? Coord. Chem. Rev. 2016, 315, 90–111. [Google Scholar] [CrossRef]
  20. Zhao, H.; Dong, Y.; Wang, G.; Jiang, P.; Zhang, J.; Wu, L.; Li, K. Novel magnetically separable nanomaterials for heterogeneous catalytic ozonation of phenol pollutant: NiFe2O4 and their performances. Chem. Eng. J. 2013, 219, 295–302. [Google Scholar] [CrossRef]
  21. Andjelković, L.; Šuljagić, M.; Lakić, M.; Jeremić, D.; Vulić, P.; Nikolić, A.S. A study of the structural and morphological properties of Ni–ferrite, Zn–ferrite and Ni–Zn–ferrites functionalized with starch. Ceram. Int. 2018, 44, 14163–14168. [Google Scholar] [CrossRef]
  22. Kefeni, K.K.; Mamba, B.B.; Msagati, T.A. Application of spinel ferrite nanoparticles in water and wastewater treatment: A review. Sep. Purif. Technol. 2017, 188, 399–422. [Google Scholar] [CrossRef]
  23. Casbeer, E.; Sharma, V.K.; Li, X.Z. Synthesis and photocatalytic activity of ferrites under visible light: A review. Sep. Purif. Technol. 2012, 87, 1–14. [Google Scholar] [CrossRef]
  24. Yuliantika, D.; Taufiq, A.; Hidayat, A.; Hidayat, N.; Soontaranon, S. Exploring structural properties of cobalt ferrite nanoparticles from natural sand. In IOP Conference Series: Materials Science and Engineering; IOP Publishing: Bristol, UK, 2019; Volume 515, p. 012047. [Google Scholar]
  25. Narang, S.B.; Pubby, K. Nickel spinel ferrites: A review. J. Magn. Magn. Mater. 2021, 519, 167163. [Google Scholar] [CrossRef]
  26. Hazra, S.; Ghosh, N.N. Preparation of nanoferrites and their applications. J. Nanosci. Nanotechnol. 2014, 14, 1983–2000. [Google Scholar] [CrossRef]
  27. Amiri, G.R.; Yousefi, M.H.; Abolhassani, M.R.; Manouchehri, S.; Keshavarz, M.H.; Fatahian, S. Magnetic properties and microwave absorption in Ni–Zn and Mn–Zn ferrite nanoparticles synthesized by low-temperature solid-state reaction. J. Magn. Magn. Mater. 2011, 323, 730–734. [Google Scholar] [CrossRef]
  28. Gupta, N.; Verma, A.; Kashyap, S.C. Micro structural, dielectric and magnetic behavior of spin deposited nanocrystalline nickel-zinc ferrite thin films for microwave applications. J. Magn. Magn. Mater. 2007, 308, 137–142. [Google Scholar] [CrossRef]
  29. Liu, Y.; Li, J.; Min, F.; Zhu, J.; Zhang, M. Microwave-assisted synthesis and magnetic properties of Ni1−xZnxFe2O4 ferrite powder. J. Magn. Magn. Mater. 2014, 354, 295–298. [Google Scholar] [CrossRef]
  30. Shafi, K.; Koltypin, Y.; Gedanken, A.; Prozorov, R.; Balogh, J.; Lendvai, J.; Felner, I. Sonochemical preparation of nano- sized amorphous NiFe2O4 particles. J. Mater. Chem. Phys. 1997, 101, 6409–6414. [Google Scholar]
  31. Chen, D.H.; He, X.R. Synthesis of nickel ferrite nanoparticles by sol–gel method. Mater. Res. Bull. 2001, 36, 1369–1377. [Google Scholar] [CrossRef]
  32. Zahi, S.; Hashim, M.; Daud, A.R. Synthesis, magnetic properties and microstructure of Ni–Zn ferrite by sol–gel technique. J. Magn. Magn. Mater. 2007, 308, 177–182. [Google Scholar] [CrossRef]
  33. Atif, M.; Nadeem, M.; Grössinger, R.; Turtelli, R.S. Studies on the magnetic, magneto strictive and electrical properties of sol–gel synthesized Zn doped nickel ferrite. J. Alloys Compd. 2011, 509, 5720–5724. [Google Scholar] [CrossRef]
  34. Kumar, S.J.; Prameela, P.; Rao, K.S.; Kiran, J.N.; Rao, K.H. Structural and Magnetic Properties of Copper-Substituted Nickel–Zinc Nanoparticles Prepared by Sol-Gel Method. J. Supercond. Nov. Magn. 2020, 33, 693–705. [Google Scholar] [CrossRef]
  35. Yang, J.M.; Tsuo, W.J.; Yen, F.S. Preparation of ultrafine nickel ferrite powders using mixed Ni and Fe tartrates. J. Solid State Chem. 1999, 145, 50–57. [Google Scholar] [CrossRef]
  36. Maaz, K.; Karim, S.; Mumtaz, A.; Hasanain, S.K.; Liu, J.; Duan, J.L. Synthesis and magnetic characterization of nickel ferrite nanoparticles prepared by co-precipitation route. J. Magn. Magn. Mater. 2009, 321, 1838–1842. [Google Scholar] [CrossRef]
  37. Dantas, J.; Leal, E.; Mapossa, A.B.; Cornejo, D.R.; Costa, A.C.F.M. Magnetic nanocatalysts of Ni0.5Zn0.5Fe2O4 doped with Cu and performance evaluation in transesterification reaction for biodiesel production. Fuel 2017, 191, 463–471. [Google Scholar] [CrossRef]
  38. Mapossa, A.B.; Dantas, J.; Diniz, V.C.S.; Silva, M.R.; Kiminami, R.H.G.A.; Costa, A.C.F.M. Síntese e caracterização do ferroespinélio Ni0.7Zn0.3Fe2O4: Avaliação de desempenho na esterificação metílica e etílica. Cerâmica 2017, 63, 223–232. [Google Scholar] [CrossRef] [Green Version]
  39. Abbas, M.K.; Khan, M.A.; Mushtaq, F.; Warsi, M.F.; Sher, M.; Shakir, I.; AlyAboud, M.F. Impact of Dy on structural, dielectric and magnetic properties of Li-Tb-nanoferrites synthesized by micro-emulsion method. Ceram. Int. 2017, 43, 5524–5533. [Google Scholar] [CrossRef]
  40. Rozman, M.; Drofenik, M. Hydrothermal synthesis of manganese zinc ferrites. J. Am. Ceram. Soc. 1995, 78, 2449–2455. [Google Scholar] [CrossRef]
  41. Zhao, L.; Zhang, H.; Xing, Y.; Song, S.; Yu, S.; Shi, W.; Guo, X.; Yang, J.; Lei, Y.; Cao, F. Studies on the magnetism of cobalt ferrite nanocrystals synthesized by hydrothermal method. J. Solid State Chem. 2008, 181, 245–252. [Google Scholar] [CrossRef]
  42. Nejati, K.; Zabihi, R. Preparation and magnetic properties of nano size nickel ferrite particles using hydrothermal method. Chem. Cent. J. 2012, 6, 23. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  43. Feng, J.; Su, L.; Ma, Y.; Ren, C.; Guo, Q.; Chen, X. CuFe2O4 magnetic nanoparticles: A simple and efficient catalyst for the reduction of nitrophenol. Chem. Eng. J. 2013, 221, 16–24. [Google Scholar] [CrossRef]
  44. Chen, C.; Zuo, W.Q.; Yang, J.C.E.; Cui, H.J.; Fu, M.L. Yolk–shell structured CoFe2O4 microspheres as novel catalysts for peroxymonosulfate activation for efficient degradation of butyl paraben. RSC Adv. 2016, 6, 10136. [Google Scholar] [CrossRef]
  45. George, M.; John, A.M.; Nair, S.S.; Joy, P.A.; Anantharaman, M.R. Finite size effects on the structural and magnetic properties of sol–gel synthesized NiFe2O4 powders. J. Magn. Magn. Mater. 2006, 302, 190–195. [Google Scholar] [CrossRef]
  46. El-Okr, M.M.; Salem, M.A.; Salim, M.S.; El-Okr, R.M.; Ashoush, M.; Talaat, H.M. Synthesis of cobalt ferrite nano-particles and their magnetic characterization. J. Magn. Magn. Mater. 2011, 323, 920–926. [Google Scholar] [CrossRef]
  47. Konecny, A.P.; Covarrubias, J.; Wang, H. Magnetic Nanoparticle Design and Application in Magnetic Hyperthermia. In Magnetic Nanomaterials: Applications in Catalysis and Life Sciences; Bossmann, S.H., Wang, H., Eds.; RSC: London, UK, 2017. [Google Scholar]
  48. Ali, A.; Hira Zafar, M.Z.; Ul Haq, I.; Phull, A.R.; Ali, J.S.; Hussain, A. Synthesis, characterization, applications, and challenges of iron oxide nanoparticles. Nanotechnol. Sci. Appl. 2016, 9, 49. [Google Scholar] [CrossRef] [Green Version]
  49. Gopalan Sibi, M.; Verma, D.; Kim, J. Magnetic core–shell nanocatalysts: Promising versatile catalysts for organic and photocatalytic reactions. Catal. Rev. 2020, 62, 163–311. [Google Scholar] [CrossRef]
  50. Bhagwat, V.R.; Humbe, A.V.; More, S.D.; Jadhav, K.M. Sol-gel auto combustion synthesis and characterizations of cobalt ferrite nanoparticles: Different fuels approach. Mater. Sci. Eng. B 2019, 248, 114388. [Google Scholar] [CrossRef]
  51. Venturini, J.; Zampiva, R.Y.S.; Arcaro, S.; Bergmann, C.P. Sol-gel synthesis of substoichiometric cobalt ferrite (CoFe2O4) spinels: Influence of additives on their stoichiometry and magnetic properties. Ceram. Int. 2018, 44, 12381–12388. [Google Scholar] [CrossRef]
  52. Motavallian, P.; Abasht, B.; Abdollah-Pour, H. Zr doping dependence of structural and magnetic properties of cobalt ferrite synthesized by sol–gel based Pechini method. J. Magn. Magn. Mater. 2018, 451, 577–586. [Google Scholar] [CrossRef]
  53. Hossain, M.S.; Alam, M.B.; Shahjahan, M.; Begum, M.H.A.; Hossain, M.M.; Islam, S.; Khatun, N.; Hossain, M.; Alam, M.S.; Al-Mamun, M. Synthesis, structural investigation, dielectric and magnetic properties of Zn2+-doped cobalt ferrite by the sol–gel technique. J. Adv. Dielectr. 2018, 8, 1850030. [Google Scholar] [CrossRef] [Green Version]
  54. Ansari, F.; Sobhani, A.; Salavati-Niasari, M. Simple sol-gel synthesis and characterization of new CoTiO3/CoFe2O4 nanocomposite by using liquid glucose, maltose and starch as fuel, capping and reducing agents. J. Colloid Interface Sci. 2018, 514, 723–732. [Google Scholar] [CrossRef]
  55. Raut, A.V.; Barkule, R.S.; Shengule, D.R.; Jadhav, K.M. Synthesis, structural investigation and magnetic properties of Zn2+ substituted cobalt ferrite nanoparticles prepared by the sol–gel auto-combustion technique. J. Magn. Magn. Mater. 2014, 358, 87–92. [Google Scholar] [CrossRef]
  56. Sajjia, M.; Oubaha, M.; Hasanuzzaman, M.; Olabi, A.G. Developments of cobalt ferrite nanoparticles prepared by the sol–gel process. Ceram. Int. 2014, 40, 1147–1154. [Google Scholar] [CrossRef] [Green Version]
  57. Jain, S.R.; Adiga, K.C.; Verneker, V.P. A new approach to thermochemical calculations of condensed fuel-oxidizer mixtures. Combust. Flame 1981, 40, 71–79. [Google Scholar] [CrossRef]
  58. Mapossa, A.B.; Dantas, J.; Costa, A.C. Transesterification reaction for biodiesel production from soybean oil using Ni0.5Zn0.5Fe2O4 nanomagnetic catalyst: Kinetic study. Int. J. Energy Res. 2020, 44, 6674–6684. [Google Scholar] [CrossRef]
  59. Dantas, J.; Leal, E.; Mapossa, A.B.; Pontes, J.R.; Freitas, N.L.; Fernandes, P.C.; Costa, A.C.F. Biodiesel production on bench scale from different sources of waste oils by using NiZn magnetic heterogeneous nanocatalyst. Int. J. Energy Res. 2021, 45, 10924–10945. [Google Scholar] [CrossRef]
  60. Costa, A.C.F.M.; Morelli, M.R.; Kiminami, R.H.G.A. Combustion synthesis: Effect of urea on the reaction and characteristics of Ni-Zn ferrite powders. J. Mater. Synth. Process. 2002, 9, 347–352. [Google Scholar] [CrossRef]
  61. Hwang, C.C.; Tsai, J.S.; Huang, T.H.; Peng, C.H.; Chen, S.Y. Combustion synthesis of Ni–Zn ferrite powder—influence of oxygen balance value. J. Solid State Chem. 2005, 178, 382–389. [Google Scholar] [CrossRef]
  62. Kombaiah, K.; Vijaya, J.J.; Kennedy, L.J.; Bououdina, M.; Al-Lohedan, H.A.; Ramalingam, R.J. Studies on Opuntia dilenii haw mediated multifunctional ZnFe2O4 nanoparticles: Optical, magnetic and catalytic applications. Mater. Chem. Phys. 2017, 194, 153–164. [Google Scholar] [CrossRef]
  63. Segadães, A.M.; Morelli, M.R.; Kiminami, R.H.G.A. Combustion synthesis of aluminium titanate. J. Eur. Ceram. Soc. 1998, 8, 771–781. [Google Scholar] [CrossRef]
  64. Mapossa, A.B.; Dantas, J.; Kiminami, A.G.H.R.; Silva, M.R.; Costa, A.C.F.M. Síntese do ferroespinélio ZnFe2O4 e avaliação do seu desempenho em reações de esterificação e transesterificação via rota metílica. Rev. Eletrônica Mater. Process. 2015, 10, 137–143. [Google Scholar]
  65. Salunkhe, A.B.; Khot, V.M.; Phadatare, M.R.; Pawar, S.H. Combustion synthesis of cobalt ferrite nanoparticles—Influence of fuel to oxidizer ratio. J. Alloys Compd. 2012, 514, 91–96. [Google Scholar] [CrossRef]
  66. Khorrami, S.A.; Manuchehri, Q.S. Magnetic properties of cobalt ferrite synthesized by hydrothermal and co-precipitation methods: A comparative study. J. Appl. Chem. Res. 2013, 7, 15–23. [Google Scholar]
  67. Ren, Y.; Lin, L.; Ma, J.; Yang, J.; Feng, J.; Fan, Z. Sulfate radicals induced from peroxymonosulfate by magnetic ferrospinel MFe2O4 (M = Co, Cu, Mn, and Zn) as heterogeneous catalysts in the water. Appl. Catal. B Environ. 2015, 165, 572–578. [Google Scholar] [CrossRef]
  68. Soto-Arreola, A.; Huerta-Flores, A.M.; Mora-Hernández, J.M.; Torres-Martínez, L.M. Comparative study of the photocatalytic activity for hydrogen evolution of MFe2O4 (M = Cu, Ni) prepared by three different methods. J. Photochem. Photobiol. A Chem. 2018, 357, 20–29. [Google Scholar] [CrossRef]
  69. Rameshbabu, R.; Kumar, N.; Karthigeyan, A.; Neppolian, B. Visible light photocatalytic activities of ZnFe2O4/ZnO nanoparticles for the degradation of organic pollutants. Mater. Chem. Phys. 2016, 181, 106–115. [Google Scholar] [CrossRef]
  70. Jacinto, M.J.; Ferreira, L.F.; Silva, V.C. Magnetic materials for photocatalytic applications—A review. J. Sol-Gel Sci. Technol. 2020, 96, 1–14. [Google Scholar] [CrossRef]
  71. Hassan, N.S.; Jalil, A.A. A review on self-modification of zirconium dioxide nanocatalysts with enhanced visible-light-driven photodegradation of organic pollutants. J. Hazard. Mater. 2022, 423, 126996. [Google Scholar] [CrossRef]
  72. Suresh, R.; Rajendran, S.; Kumar, P.S.; Vo, D.V.N.; Cornejo-Ponce, L. Recent advancements of spinel ferrite based binary nanocomposite photocatalysts in wastewater treatment. Chemosphere 2021, 274, 129734. [Google Scholar] [CrossRef]
  73. Li, F.; Liu, J.; Evans, D.G.; Duan, X. Stoichiometric synthesis of pure MFe2O4 (M = Mg, Co, and Ni) spinel ferrites from tailored layered double hydroxide (hydrotalcite-like) precursors. Chem. Mater. 2004, 16, 1597–1602. [Google Scholar] [CrossRef]
  74. Tiwari, R.; De, M.; Tewari, H.S.; Ghoshal, S.K. Structural and magnetic properties of tailored NiFe2O4 nanostructures synthesized using auto-combustion method. Results Phys. 2020, 16, 102916. [Google Scholar] [CrossRef]
  75. Vepulanont, K.; Sa-Nguanprang, S.; Buapoon, S.; Bunluesak, T.; Suebsom, P.; Chaisong, K.; Udomsri, N.; Karnchana, N.; Laokae, D.; Chanadee, T. Nickel ferrite ceramics: Combustion synthesis, sintering, characterization, and magnetic and electrical properties. J. Asian Ceram. Soc. 2021, 9, 639–651. [Google Scholar] [CrossRef]
  76. Adeleke, J.T.; Theivasanthi, T.; Thiruppathi, M.; Swaminathan, M.; Akomolafe, T.; Alabi, A.B. Photocatalytic degradation of methylene blue by ZnO/NiFe2O4 nanoparticles. Appl. Surf. Sci. 2018, 455, 195–200. [Google Scholar] [CrossRef] [Green Version]
  77. Hou, X.; Feng, J.; Liu, X.; Ren, Y.; Fan, Z.; Wei, T.; Meng, J.; Zhang, M. Synthesis of 3D porous ferromagnetic NiFe2O4 and using as novel adsorbent to treat wastewater. J. Colloid Interface Sci. 2011, 362, 477–485. [Google Scholar] [CrossRef]
  78. Verma, K.C.; Singh, V.P.; Ram, M.; Shah, J.; Kotnala, R.K. Structural, microstructural and magnetic properties of NiFe2O4, CoFe2O4 and MnFe2O4 nanoferrite thin films. J. Magn. Magn. Mater. 2011, 323, 3271–3275. [Google Scholar] [CrossRef]
  79. Xiong, P.; Fu, Y.; Wang, L.; Wang, X. Multi-walled carbon nanotubes supported nickel ferrite: A magnetically recyclable photocatalyst with high photocatalytic activity on degradation of phenols. Chem. Eng. J. 2012, 195, 149–157. [Google Scholar] [CrossRef]
  80. Peng, T.; Zhang, X.; Lv, H.; Zan, L. Preparation of NiFe2O4 nanoparticles and its visible-light-driven photoactivity for hydrogen production. Catal. Commun. 2012, 28, 116–119. [Google Scholar] [CrossRef]
  81. Ren, A.; Liu, C.; Hong, Y.; Shi, W.; Lin, S.; Li, P. Enhanced visible-light-driven photocatalytic activity for antibiotic degradation using magnetic NiFe2O4/Bi2O3 heterostructures. Chem. Eng. J. 2014, 258, 301–308. [Google Scholar] [CrossRef]
  82. Bhosale, S.V.; Kanhe, N.S.; Bhoraskar, S.V.; Bhat, S.K.; Bulakhe, R.N.; Shim, J.J.; Mathe, V.L. Micro-structural analysis of NiFe2O4 nanoparticles synthesized by thermal plasma route and its suitability for BSA adsorption. J. Mater. Sci. Mater. Med. 2015, 26, 1–15. [Google Scholar] [CrossRef] [PubMed]
  83. Khosravi, I.; Eftekhar, M. Characterization and evaluation catalytic efficiency of NiFe2O4 nano spinel in removal of reactive dye from aqueous solution. Powder Technol. 2013, 250, 147–153. [Google Scholar] [CrossRef]
  84. Hung, D.Q.; Thanh, N.K. Preparation of NiFe2O4-TiO2 nanoparticles and study of their photocatalytic activity. VNU J. Sci. Math.-Phys. 2011, 27, 204–211. [Google Scholar]
  85. Shao, R.; Sun, L.; Tang, L.; Chen, Z. Preparation and characterization of magnetic core–shell ZnFe2O4@ZnO nanoparticles and their application for the photodegradation of methylene blue. Chem. Eng. J. 2013, 217, 185–191. [Google Scholar] [CrossRef]
  86. Fan, G.; Gu, Z.; Yang, L.; Li, F. Nanocrystalline zinc ferrite photocatalysts formed using the colloid mill and hydrothermal technique. Chem. Eng. J. 2009, 155, 534–541. [Google Scholar] [CrossRef]
  87. Cheng, P.; Li, W.; Zhou, T.; Jin, Y.; Gu, M. Physical and photocatalytic properties of zinc ferrite doped titania under visible light irradiationJ. Photochem. Photobiol. A Chem. 2004, 168, 97–101. [Google Scholar] [CrossRef]
  88. Su, M.; He, C.; Sharma, V.K.; Abou Asi, M.; Xia, D.; Li, X.Z.; Deng, H.; Xiong, Y. Mesoporous zinc ferrite: Synthesis, characterization, and photocatalytic activity with H2O2/visible light. J. Hazard. Mater. 2012, 211, 95–103. [Google Scholar] [CrossRef]
  89. Mahmoodi, N.M. Zinc ferrite nanoparticle as a magnetic catalyst: Synthesis and dye degradation. Mater. Res. Bull. 2013, 48, 4255–4260. [Google Scholar] [CrossRef]
  90. Guo, X.; Zhu, H.; Li, Q. Visible-light-driven photocatalytic properties of ZnO/ZnFe2O4 core/shell nanocable arrays. Appl. Catal. B Environ. 2014, 160, 408–414. [Google Scholar] [CrossRef]
  91. Yuan, Z.H. Synthesis, characterization and photocatalyticactivity of ZnFe2O4/TiO2 nanocomposite. J. Mater. Chem. 2001, 11, 1265–1268. [Google Scholar] [CrossRef]
  92. Hou, X.; Feng, J.; Xu, X.; Zhang, M. Synthesis and characterizations of spinel MnFe2O4 nanorod by seed–hydrothermal route. J. Alloys Compd. 2010, 491, 258–263. [Google Scholar] [CrossRef]
  93. Patil, S.B.; Naik, H.B.; Nagaraju, G.; Viswanath, R.; Rashmi, S.K. Sugarcane juice mediated eco-friendly synthesis of visible light active zinc ferrite nanoparticles: Application to degradation of mixed dyes and antibacterial activities. Mater. Chem. Phys. 2018, 212, 351–362. [Google Scholar] [CrossRef]
  94. Sripriya, R.C.; Vigneaswari, B.; Raj, V.A. Comparative Studies of Magneto-Optical and Photocatalytic Properties of Magnetically Recyclable Spinel ZnFe2O4 Nanostructures by Combustion Methods. Int. J. Nanosci. 2019, 18, 1850020. [Google Scholar] [CrossRef]
  95. Dutta, V.; Sharma, S.; Raizada, P.; Hosseini-Bandegharaei, A.; Gupta, V.K.; Singh, P. Review on augmentation in photocatalytic activity of CoFe2O4 via heterojunction formation for photocatalysis of organic pollutants in water. J. Saudi Chem. Soc. 2019, 23, 1119–1136. [Google Scholar]
  96. Sickafus, K.E.; Wills, J.M.; Grimes, N.W. Structure of spinel. J. Am. Ceram. Soc. 1999, 82, 3279–3292. [Google Scholar] [CrossRef]
  97. Zheng, J.; Song, X.; Liu, X.; Chen, W.; Li, Y.; Guo, J. Synthesis of hexagonal CoFe2O4/ZnO nanoparticles and their electromagnetic properties. Mater. Lett. 2012, 73, 143–146. [Google Scholar] [CrossRef]
  98. Sathishkumar, P.; Mangalaraja, R.V.; Anandan, S.; Ashokkumar, M. CoFe2O4/TiO2 nanocatalysts for the photocatalytic degradation of Reactive Red 120 in aqueous solutions in the presence and absence of electron acceptors. Chem. Eng. J. 2013, 220, 302–310. [Google Scholar] [CrossRef]
  99. Moosavi, S.M.; Molla-Abbasi, P.; Haji-Aghajani, Z. Photo-catalyst CoFe2O4–TiO2: Application in photo-degradation of organic dyes and magnetic nanocomposite preparation. J. Mater. Sci. Mater. Electron. 2016, 27, 4879–4886. [Google Scholar] [CrossRef]
  100. Krishna, S.; Sathishkumar, P.; Pugazhenthiran, N.; Guesh, K.; Mangalaraja, R.V.; Kumaran, S.; Gracia-Pinilla, M.A.; Anandan, S. Heterogeneous sonocatalytic activation of peroxomonosulphate in the presence of CoFe2O4/TiO2 nanocatalysts for the degradation of Acid Blue 113 in an aqueous environment. J. Environ. Chem. Eng. 2020, 8, 104024. [Google Scholar]
  101. Li, C.J.; Wang, J.N.; Wang, B.; Gong, J.R.; Lin, Z. A novel magnetically separable TiO2/CoFe2O4 nanofiber with high photocatalytic activity under UV–vis light. Mater. Res. Bull. 2012, 47, 333–337. [Google Scholar] [CrossRef]
  102. Chandel, N.; Sharma, K.; Sudhaik, A.; Raizada, P.; Hosseini-Bandegharaei, A.; Thakur, V.K.; Singh, P. Magnetically separable ZnO/ZnFe2O4 and ZnO/CoFe2O4 photocatalysts supported onto nitrogen doped graphene for photocatalytic degradation of toxic dyes. Arab. J. Chem. 2020, 13, 4324–4340. [Google Scholar] [CrossRef]
  103. Arief, S.; Jamarun, N.; Stiadi, Y. Magnetically separable ZnO-MnFe2O4 nanocomposites synthesized in organic-free media for dye degradation under natural sunlight. Orient. J. Chem. 2017, 33, 2758. [Google Scholar]
  104. Akhlaghi, N.; Najafpour-Darzi, G. Manganese ferrite (MnFe2O4) Nanoparticles: From synthesis to application—A review. J. Ind. Eng. Chem. 2021, 103, 292–304. [Google Scholar] [CrossRef]
  105. Deraz, N.M.; Shaban, S. Optimization of catalytic, surface and magnetic properties of nanocrystalline manganese ferrite. J. Anal. Appl. Pyrol. 2009, 86, 173–179. [Google Scholar] [CrossRef]
  106. Chang, L.; Pu, Y.; Jing, P.; Cui, Y.; Zhang, G.; Xu, S.; Cao, B.; Guo, J.; Chen, F.; Qiao, C. Magnetic core-shell MnFe2O4@TiO2 nanoparticles decorated on reduced graphene oxide as a novel adsorbent for the removal of ciprofloxacin and Cu(II) from water. Appl. Surf. Sci. 2021, 541, 148400. [Google Scholar] [CrossRef]
  107. Zamani, A.; Seyed Sadjadi, M.; Mahjoub, A.R.; Yousefi, M.; Farhadyar, N. Synthesis and characterization ZnFe2O4@MnO and MnFe2O4@ZnO magnetic nanocomposites: Investigation of photocatalytic activity for the degradation of Congo Red under visible light irradiation. Int. J. Nano Dimens. 2020, 11, 58–73. [Google Scholar]
  108. Silambarasu, A.; Manikandan, A.; Balakrishnan, K.; Jaganathan, S.K.; Manikandan, E.; Aanand, J.S. Comparative study of structural, morphological, magneto-optical and photo-catalytic properties of magnetically reusable spinel MnFe2O4 nano-catalysts. J. Nanosci. Nanotechnol. 2018, 18, 3523–3531. [Google Scholar] [CrossRef] [PubMed]
  109. Golshan, M.; Kakavandi, B.; Ahmadi, M.; Azizi, M. Photocatalytic activation of peroxymonosulfate by TiO2 anchored on cupper ferrite (TiO2@CuFe2O4) into 2,4-D degradation: Process feasibility, mechanism and pathway. J. Hazard. Mater. 2018, 359, 325–337. [Google Scholar] [CrossRef] [PubMed]
  110. Anandan, S.; Selvamani, T.; Prasad, G.G.; Asiri, A.M.; Wu, J.J. Magnetic and catalytic properties of inverse spinel CuFe2O4 nanoparticles. J. Magn. Magn. Mater. 2017, 432, 437–443. [Google Scholar] [CrossRef]
  111. You-Yi, Y.; Heng-Qiang, Z. Reduced graphene oxide coupled magnetic CuFe2O4-TiO2 nanoparticles with enhanced photocatalytic activity for methylene blue degradation. Struct Chem. 2016, 35, 472–480. [Google Scholar]
  112. Shevale, V.B.; Dhodamani, A.G.; Delekar, S.D. Catalytic reclamation of silver present in photographic waste using magnetically separable TiO2@CuFe2O4 nanocomposites and thereof its use in antibacterial activity. ACS Omega 2020, 5, 1098–1108. [Google Scholar] [CrossRef] [Green Version]
  113. Janani, B.; Syed, A.; Thomas, A.M.; Al-Rashed, S.; Elgorban, A.M.; Raju, L.L.; Khan, S.S. A simple approach for the synthesis of bi-functional pn type ZnO@CuFe2O4 heterojunction nanocomposite for photocatalytic and antimicrobial application. Phys. E Low-Dimens. Syst. Nanostruct. 2021, 130, 114664. [Google Scholar] [CrossRef]
  114. Zhang, C.F.; Zhong, X.C.; Yu, H.Y.; Liu, Z.W.; Zeng, D.C. Effects of cobalt doping on the microstructure and magnetic properties of Mn–Zn ferrites prepared by the co-precipitation method. Phys. B Condens. Matter. 2009, 404, 2327–2331. [Google Scholar] [CrossRef]
  115. Manikandan, A.; Durka, M.; Antony, S.A. Magnetically recyclable spinel MnxZn1–xFe2O4 (0.0 ≤ x ≤ 0.5) nano-photocatalysts. Adv. Sci. Eng. Med. 2015, 7, 33–46. [Google Scholar] [CrossRef]
  116. Shaikh, P.A.; Kambale, R.C.; Rao, A.V.; Kolekar, Y.D. Structural, magnetic and electrical properties of Co–Ni–Mn ferrites synthesized by co-precipitation method. J. Alloys Compd. 2010, 492, 590–596. [Google Scholar] [CrossRef]
  117. Ciocarlan, R.G.; Seftel, E.M.; Mertens, M.; Pui, A.; Mazaj, M.; Tusar, N.N.; Cool, P. Novel magnetic nanocomposites containing quaternary ferrites systems Co0.5Zn0.25M0.25Fe2O4 (M = Ni, Cu, Mn, Mg) and TiO2-anatase phase as photocatalysts for wastewater remediation under solar light irradiation. Mater. Sci. Eng. B 2018, 230, 1–7. [Google Scholar] [CrossRef]
  118. Jesudoss, S.K.; Vijaya, J.J.; Kennedy, L.J.; Rajan, P.I.; Al-Lohedan, H.A.; Ramalingam, R.J.; Kaviyarasu, K.; Bououdina, M. Studies on the efficient dual performance of Mn1–xNixFe2O4 spinel nanoparticles in photodegradation and antibacterial activity. J. Photochem. Photobiol. B Biol. 2016, 165, 121–132. [Google Scholar] [CrossRef]
  119. Naik, M.M.; Naik, H.B.; Nagaraju, G.; Vinuth, M.; Vinu, K.; Viswanath, R. Green synthesis of zinc doped cobalt ferrite nanoparticles: Structural, optical, photocatalytic and antibacterial studies. Nano-Struct. Nano-Objects. 2019, 19, 100322. [Google Scholar] [CrossRef]
  120. Baig, M.M.; Pervaiz, E.; Afzal, M.J. Catalytic activity and kinetic studies of Core@ Shell nanostructure NiFe2O4@TiO2 for photocatalytic degradation of methyl orange dye. J. Chem. Soc. Pak. 2020, 42, 531. [Google Scholar]
  121. Baldrian, P.; Merhautová, V.; Gabriel, J.; Nerud, F.; Stopka, P.; Hrubý, M.; Beneš, M.J. Decolorization of synthetic dyes by hydrogen peroxide with heterogeneous catalysis by mixed iron oxides. Appl. Catal. B Environ. 2006, 66, 258–264. [Google Scholar] [CrossRef]
  122. Neris, A.M.; Schreiner, W.H.; Salvador, C.; Silva, U.C.; Chesman, C.; Longo, E.; Santos, I.M.G. Photocatalytic evaluation of the magnetic core@ shell system (Co, Mn) Fe2O4@TiO2 obtained by the modified Pechini method. Mater. Sci. Eng. B 2018, 229, 218–226. [Google Scholar] [CrossRef] [Green Version]
  123. Cao, X.; Gu, L.; Lan, X.; Zhao, C.; Yao, D.; Sheng, W. Spinel ZnFe2O4 nanoplates embedded with Ag clusters: Preparation, characterization, and photocatalytic application. Mater. Chem. Phys. 2007, 106, 175–180. [Google Scholar] [CrossRef]
  124. Zhang, G.Y.; Sun, Y.Q.; Gao, D.Z.; Xu, Y.Y. Quasi-cube ZnFe2O4 nanocrystals: Hydrothermal synthesis and photocatalytic activity with TiO2 (Degussa P25) as nanocomposite. Mater. Res. Bull. 2010, 45, 755–760. [Google Scholar] [CrossRef]
  125. Zhang, B.; Zhang, J.; Chen, F. Preparation and characterization of magnetic TiO2/ZnFe2O4 photocatalysts by a sol–gel method. Res. Chem. Intermed. 2008, 34, 375–380. [Google Scholar] [CrossRef]
  126. Li, X.; Hou, Y.; Zhao, Q.; Wang, L. A general, one-step and template-free synthesis of sphere-like zinc ferrite nanostructures with enhanced photocatalytic activity for dye degradation. J. Colloid Interface Sci. 2011, 358, 102–108. [Google Scholar] [CrossRef]
  127. Li, H.; Zhang, Y.; Wang, S.; Wu, Q.; Liu, C. Study on nanomagnets supported TiO2 photocatalysts prepared by a sol–gel process in reverse microemulsion combining with solvent-thermal technique. J. Hazard. Mater. 2009, 169, 1045–1053. [Google Scholar] [CrossRef]
  128. Fu, Y.P.; Chang, W.K.; Wang, H.C.; Liu, C.W.; Lin, C.H. Synthesis and characterization of anatase TiO2 nanolayer coating on Ni–Cu–Zn ferrite powders for magnetic photocatalyst. J. Mater. Res. 2010, 25, 134–140. [Google Scholar] [CrossRef]
  129. Dom, R.; Subasri, R.; Radha, K.; Borse, P.H. Synthesis of solar active nanocrystalline ferrite, MFe2O4 (M: Ca, Zn, Mg) photocatalyst by microwave irradiation. Solid State Commun. 2011, 151, 470–473. [Google Scholar] [CrossRef]
  130. Chen, C.H.; Liang, Y.H.; Zhang, W.D. ZnFe2O4/MWCNTs composite with enhanced photocatalytic activity under visible-light irradiation. J. Alloys Compd. 2010, 501, 168–172. [Google Scholar] [CrossRef]
  131. Rana, S.; Srivastava, R.S.; Sorensson, M.M.; Misra, R.D.K. Synthesis and characterization of nanoparticles with magnetic core and photocatalytic shell: Anatase TiO2–NiFe2O4 system. Mater. Sci. Eng. B 2005, 119, 144–151. [Google Scholar] [CrossRef]
  132. Cheng, P.; Deng, C.; Gu, M.; Shangguan, W. Visible-light responsive zinc ferrite doped titania photocatalyst for methyl orange degradation. J. Mater. Sci. 2007, 42, 9239–9244. [Google Scholar] [CrossRef]
  133. Jadhav, S.D.; Hankare, P.P.; Patil, R.P.; Sasikala, R. Effect of sintering on photocatalytic degradation of methyl orange using zinc ferrite. Mater. Lett. 2011, 65, 371–373. [Google Scholar] [CrossRef]
  134. Xu, S.; Feng, D.; Shangguan, W. Preparations and photocatalytic properties of visible-light-active zinc ferrite-doped TiO2 photocatalyst. J. Phys. Chem. C 2009, 113, 2463–2467. [Google Scholar] [CrossRef]
  135. Arifin, M.N.; Karim, K.M.R.; Abdullah, H.; Khan, M.R. Synthesis of titania doped copper ferrite photocatalyst and its photoactivity towards methylene blue degradation under visible light irradiation. Bull. Chem. React. Eng. Catal. 2019, 14, 219–227. [Google Scholar] [CrossRef] [Green Version]
  136. Zouhier, M.; Tanji, K.; Navio, J.A.; Hidalgo, M.C.; Jaramillo-Páez, C.; Kherbeche, A. Preparation of ZnFe2O4/ZnO composite: Effect of operational parameters for photocatalytic degradation of dyes under UV and visible illumination. J. Photochem. Photobiol. A Chem. 2020, 390, 112305. [Google Scholar] [CrossRef]
  137. To Loan, N.T.; Hien Lan, N.T.; Thuy Hang, N.T.; Quang Hai, N.; Tu Anh, D.T.; Thi Hau, V.; Van Tan, L.; Van Tran, T. CoFe2O4 nanomaterials: Effect of annealing temperature on characterization, magnetic, photocatalytic, and photo-fenton properties. Processes 2019, 7, 885. [Google Scholar] [CrossRef] [Green Version]
  138. Mir, S.H.; Jennings, B.D.; Akinoglu, G.E.; Selkirk, A.; Gatensby, R.; Mokarian-Tabari, P. Enhanced Dye Degradation through Multi-Particle Confinement in a Porous Silicon Substrate: A Highly Efficient, Low Band Gap Photocatalyst. Adv. Opt. Mater. 2021, 9, 2002238. [Google Scholar] [CrossRef]
  139. Sciortino, F.; Sanchez-Ballester, N.M.; Mir, S.H.; Rydzek, G. Functional Elastomeric Copolymer Membranes Designed by Nanoarchitectonics Approach for Methylene Blue Removal. J. Inorg. Organomet. Polym. 2021, 31, 1967–1977. [Google Scholar] [CrossRef]
  140. Velmurugan, R.; Swaminathan, M. An efficient nanostructured ZnO for dye sensitized degradation of Reactive Red 120 dye under solar light. Sol. Energy Mater Sol. Cells. 2011, 95, 942–950. [Google Scholar] [CrossRef]
  141. Selvam, K.; Balachandran, S.; Velmurugan, R.; Swaminathan, M. Mesoporous nitrogen doped nano titania—A green photocatalyst for the effective reductive cleavage of azoxybenzenes to amines or 2-phenyl indazoles in methanol. Appl. Catal. A Gen. 2012, 413, 213–222. [Google Scholar] [CrossRef]
  142. Mathubala, G.; Manikandan, A.; Antony, S.A.; Ramar, P. Photocatalytic degradation of methylene blue dye and magneto-optical studies of magnetically recyclable spinel NixMn1−xFe2O4 (x = 0.0–1.0) nanoparticles. J. Mol. Struct. 2016, 1113, 79–87. [Google Scholar] [CrossRef]
  143. Leonel, A.G.; Mansur, A.A.; Mansur, H.S. Advanced Functional Nanostructures based on Magnetic Iron Oxide Nanomaterials for Water Remediation: A Review. Water Res. 2020, 190, 116693. [Google Scholar] [CrossRef]
  144. Padmapriya, G.; Manikandan, A.; Krishnasamy, V.; Jaganathan, S.K.; Antony, S.A. Spinel NixZn1−xFe2O4 (0.0 ≤ x ≤ 1.0) nano-photocatalysts: Synthesis, characterization and photocatalytic degradation of methylene blue dye. J. Mol. Struct. 2016, 1119, 39–47. [Google Scholar] [CrossRef]
  145. Jia, Z.; Ren, D.; Liang, Y.; Zhu, R. A new strategy for the preparation of porous zinc ferrite nanorods with subsequently light-driven photocatalytic activity. Mater. Lett. 2011, 65, 3116–3119. [Google Scholar] [CrossRef]
  146. Velmurugan, R.; Krishnakumar, B.; Subash, B.; Swaminathan, M. Preparation and characterization of carbon nanoparticles loaded TiO2 and its catalytic activity driven by natural sunlight. Sol. Energy Mater Sol. Cells 2013, 108, 205–212. [Google Scholar] [CrossRef]
  147. Mandal, S.S.; Bhattacharyya, A.J. Electrochemical sensing and photocatalysis using Ag–TiO2 microwires. J. Chem. Sci. 2012, 124, 969–978. [Google Scholar] [CrossRef] [Green Version]
  148. Kumar, A.; Pandey, G. A review on the factors affecting the photocatalytic degradation of hazardous materials. Mater. Sci. Eng. Int. J. 2017, 1, 1–10. [Google Scholar] [CrossRef] [Green Version]
  149. Mirkhani, V.; Tangestaninejad, S.; Moghadam, M.; Habibi, M.H.; Rostami-Vartooni, A. Photocatalytic degradation of azo dyes catalyzed by Ag doped TiO2 photocatalyst. J. Iran. Chem. Soc. 2009, 6, 578–587. [Google Scholar] [CrossRef]
  150. Suwarnkar, M.B.; Dhabbe, R.S.; Kadam, A.N.; Garadkar, K.M. Enhanced photocatalytic activity of Ag doped TiO2 nanoparticles synthesized by a microwave assisted method. Ceram. Int. 2014, 40, 5489–5496. [Google Scholar] [CrossRef]
  151. Zhu, X.; Zhang, F.; Wang, M.; Ding, J.; Sun, S.; Bao, J.; Gao, C. Facile synthesis, structure and visible light photocatalytic activity of recyclable ZnFe2O4/TiO2. Appl. Surf. Sci. 2014, 319, 83–89. [Google Scholar] [CrossRef]
  152. Gupta, V.K.; Eren, T.; Atar, N.; Yola, M.L.; Parlak, C.; Karimi-Maleh, H. CoFe2O4@TiO2 decorated reduced graphene oxide nanocomposite for photocatalytic degradation of chlorpyrifos. J. Mol. Liq. 2015, 208, 122–129. [Google Scholar] [CrossRef]
  153. Haw, C.; Chiu, W.; Rahman, S.A.; Khiew, P.; Radiman, S.; Shukor, R.A.; Hamid, M.A.A.; Ghazali, N. The design of new magnetic-photocatalyst nanocomposites (CoFe2O4–TiO2) as smart nanomaterials for recyclable-photocatalysis applications. New J. Chem. 2016, 40, 1124–1136. [Google Scholar] [CrossRef]
  154. Ahmadpour, N.; Sayadi, M.H.; Sobhani, S.; Hajiani, M. A potential natural solar light active photocatalyst using magnetic ZnFe2O4@TiO2/Cu nanocomposite as a high performance and recyclable platform for degradation of naproxen from aqueous solution. J. Clean. Prod. 2020, 268, 122023. [Google Scholar] [CrossRef]
  155. Nguyen, T.B.; Doong, R.A. Fabrication of highly visible-light-responsive ZnFe2O4/TiO2 heterostructures for the enhanced photocatalytic degradation of organic dyes. RSC Adv. 2016, 6, 103428–103437. [Google Scholar] [CrossRef]
  156. Xu, S.H.; Feng, D.L.; Li, D.X.; Shangguan, W.F. Preparation of magnetic photocatalyst TiO2 supported on NiFe2O4 and effect of magnetic carrier on photocatalytic activity. Chin. J. Chem. 2008, 26, 842–846. [Google Scholar] [CrossRef]
  157. Nguyen, T.B.; Huang, C.P.; Doong, R.A. Photocatalytic degradation of bisphenol A over a ZnFe2O4/TiO2 nanocomposite under visible light. Sci. Total Environ. 2019, 646, 745–756. [Google Scholar] [CrossRef]
  158. Saad, D.; Byrne, D.; Drechsel, P. Social perspectives on the effective management of wastewater. In Physico-Chemical Wastewater Treatment and Resource Recovery; Farooq, R., Ahmad, Z., Eds.; InTech: London, UK, 2017; pp. 253–267. [Google Scholar]
  159. Wichelns, D.; Drechsel, P.; Qadir, M. Wastewater: Economic Asset in an Urbanizing World. In Wastewater; Drechsel, P., Qadir, M., Wichelns, D., Eds.; Springer: Dordrecht, The Netherlands, 2015; pp. 3–14. [Google Scholar]
  160. Friedler, E.; Lahav, O.; Jizhaki, H.; Lahav, T. Study of urban population attitudes towards various wastewater reuse options: Israel as a case study. J. Environ. Manag. 2006, 81, 360–370. [Google Scholar] [CrossRef]
  161. Gebrezgabher, S.; Rao, K.; Hanjra, M.A.; Hernández-Sancho, F. Business models and economic approaches for recovering energy from wastewater and fecal sludge. In Wastewater; Springer: Dordrecht, The Netherlands, 2015; pp. 217–245. [Google Scholar]
Figure 1. Structure of magnetic spinel ferrite showing tetrahedral sites (yellow), octahedral sites (green), and oxygen atoms (red) units (a) [12]. Unit cell structure of (b) normal spinel ferrite, and (c) inverse spinel ferrite [25]. Republished with permission from Elsevier.
Figure 1. Structure of magnetic spinel ferrite showing tetrahedral sites (yellow), octahedral sites (green), and oxygen atoms (red) units (a) [12]. Unit cell structure of (b) normal spinel ferrite, and (c) inverse spinel ferrite [25]. Republished with permission from Elsevier.
Catalysts 11 01543 g001
Figure 2. Magnetic properties behavior of different photocatalysts at room temperature [67]. Republished with permission from Elsevier.
Figure 2. Magnetic properties behavior of different photocatalysts at room temperature [67]. Republished with permission from Elsevier.
Catalysts 11 01543 g002
Figure 3. SEM micrographs for CuFe2O4 obtained by different methods: (a) Sol-Gel method and (b) hydrothermal synthesis. The images show the effect of methods of synthesis for the materials obtained [68]. Republished with permission from Elsevier.
Figure 3. SEM micrographs for CuFe2O4 obtained by different methods: (a) Sol-Gel method and (b) hydrothermal synthesis. The images show the effect of methods of synthesis for the materials obtained [68]. Republished with permission from Elsevier.
Catalysts 11 01543 g003
Figure 4. A general scheme of visible light photodegradation mechanism of magnetic nanocomposite i.e., ZnFe2O4/ZnO for organic pollutants [69]. Republished with permission from Elsevier.
Figure 4. A general scheme of visible light photodegradation mechanism of magnetic nanocomposite i.e., ZnFe2O4/ZnO for organic pollutants [69]. Republished with permission from Elsevier.
Catalysts 11 01543 g004
Figure 5. (a) XRD structural and (b) SEM micrographs of ZnO, NiFe2O4, and magnetic ZnO/NiFe2O4 nanocomposites [76]. Republished with permission from Elsevier.
Figure 5. (a) XRD structural and (b) SEM micrographs of ZnO, NiFe2O4, and magnetic ZnO/NiFe2O4 nanocomposites [76]. Republished with permission from Elsevier.
Catalysts 11 01543 g005
Figure 6. (a) XRD structural and (b) SEM micrographs of ZnO, ZnFe2O4, and magnetic ZnFe2O4/ZnO nanocomposites with different molar ratios (1/1, 1/8) [17]. ◊ corresponds to ZnFe2O4 and * corresponds to ZnO. Republished with permission from Wiley.
Figure 6. (a) XRD structural and (b) SEM micrographs of ZnO, ZnFe2O4, and magnetic ZnFe2O4/ZnO nanocomposites with different molar ratios (1/1, 1/8) [17]. ◊ corresponds to ZnFe2O4 and * corresponds to ZnO. Republished with permission from Wiley.
Catalysts 11 01543 g006
Figure 7. (a) XRD structural and (b) TEM micrographs of CoFe2O4 and magnetic CoFe2O4/ZnO nanocomposites [97]. Republished with permission from Elsevier.
Figure 7. (a) XRD structural and (b) TEM micrographs of CoFe2O4 and magnetic CoFe2O4/ZnO nanocomposites [97]. Republished with permission from Elsevier.
Catalysts 11 01543 g007
Figure 8. XRD patten of MnFe2O4 and MnFe2O4/TiO2 nanocomposites [106]. Republished with permission from Elsevier.
Figure 8. XRD patten of MnFe2O4 and MnFe2O4/TiO2 nanocomposites [106]. Republished with permission from Elsevier.
Catalysts 11 01543 g008
Figure 9. SEM micrographs of: (a) MnFe2O4; (b) MnFe2O4/TiO2 nanocomposite; and TEM images of (c) MnFe2O4 and (d) MnFe2O4/TiO2 nanocomposites [106]. Republished with permission from Elsevier.
Figure 9. SEM micrographs of: (a) MnFe2O4; (b) MnFe2O4/TiO2 nanocomposite; and TEM images of (c) MnFe2O4 and (d) MnFe2O4/TiO2 nanocomposites [106]. Republished with permission from Elsevier.
Catalysts 11 01543 g009
Figure 10. XRD patterns of neat TiO2, neat CuFe2O4, and nanocomposite TiO2/CuFe2O4 [109]. Republished with permission from Elsevier.
Figure 10. XRD patterns of neat TiO2, neat CuFe2O4, and nanocomposite TiO2/CuFe2O4 [109]. Republished with permission from Elsevier.
Catalysts 11 01543 g010
Figure 11. SEM micrographs of: (a) CuFe2O4 and (b) magnetic nanocomposite TiO2/CuFe2O4 [109]. Republished with permission from Elsevier.
Figure 11. SEM micrographs of: (a) CuFe2O4 and (b) magnetic nanocomposite TiO2/CuFe2O4 [109]. Republished with permission from Elsevier.
Catalysts 11 01543 g011
Figure 12. XRD pattern of TiO2 and magnetic ferrite-based TiO2 nanocomposites [117]. Republished with permission from Elsevier.
Figure 12. XRD pattern of TiO2 and magnetic ferrite-based TiO2 nanocomposites [117]. Republished with permission from Elsevier.
Catalysts 11 01543 g012
Figure 13. TEM images of the magnetic nanocomposites: (a) Co0.5Zn0.25Ni0.25Fe2O4/TiO2 and (b) Co0.5Zn0.25Cu0.25Fe2O4/TiO2 [117]. d011 demonstrates the different of diameter of particles. Republished with permission from Elsevier.
Figure 13. TEM images of the magnetic nanocomposites: (a) Co0.5Zn0.25Ni0.25Fe2O4/TiO2 and (b) Co0.5Zn0.25Cu0.25Fe2O4/TiO2 [117]. d011 demonstrates the different of diameter of particles. Republished with permission from Elsevier.
Catalysts 11 01543 g013
Figure 14. Effect of Ni2+ ions doped MnFe2O4 magnetic nanoparticles on the photocatalytic degradation (PCD) efficiency. The conditions used were IC = 150 mg/L, photocatalyst = 50 mg/100 mL, and λ = 365 nm) [118]. Republished with permission from Elsevier.
Figure 14. Effect of Ni2+ ions doped MnFe2O4 magnetic nanoparticles on the photocatalytic degradation (PCD) efficiency. The conditions used were IC = 150 mg/L, photocatalyst = 50 mg/100 mL, and λ = 365 nm) [118]. Republished with permission from Elsevier.
Catalysts 11 01543 g014
Figure 15. Photocatalytic results of: (a) Congo Red (CR) and (b) Evans Blue (EB) dyes using zinc-doped cobalt nanoparticles [119]. Republished with permission from Elsevier.
Figure 15. Photocatalytic results of: (a) Congo Red (CR) and (b) Evans Blue (EB) dyes using zinc-doped cobalt nanoparticles [119]. Republished with permission from Elsevier.
Catalysts 11 01543 g015
Figure 16. (a) Effect of surface area for photocatalytic degradation efficiency of spinel NixZn1−xFe2O4 (x = 0.0, 0.2, 0.4, 0.6, 0.8, and 1.0) nanoparticles. (b) Effect of catalyst amount on the photocatalytic degradation (PCD) efficiency of spinel Ni0.6Zn0.4Fe2O4 nanoparticles. (c) Effect of pH on photodegradation of methylene blue using catalyst Ni0.6Zn0.4Fe2O4 [144]. Republished with permission from Elsevier.
Figure 16. (a) Effect of surface area for photocatalytic degradation efficiency of spinel NixZn1−xFe2O4 (x = 0.0, 0.2, 0.4, 0.6, 0.8, and 1.0) nanoparticles. (b) Effect of catalyst amount on the photocatalytic degradation (PCD) efficiency of spinel Ni0.6Zn0.4Fe2O4 nanoparticles. (c) Effect of pH on photodegradation of methylene blue using catalyst Ni0.6Zn0.4Fe2O4 [144]. Republished with permission from Elsevier.
Catalysts 11 01543 g016
Figure 17. (a) Magnetization hysteresis loops observed for the CoFe2O4 (black dots) and CoFe2O4/TiO2 nanocatalysts (blue line) [100]; (b) photocatalytic degradation of AB113 in the presence of magnetically recyclable CoFe2O4/TiO2 nanocatalysts in cycles [100]; and (c) magnetic responsiveness of CoFe2O4/TiO2 with an external magnetic field [153]. Republished with permission from Elsevier.
Figure 17. (a) Magnetization hysteresis loops observed for the CoFe2O4 (black dots) and CoFe2O4/TiO2 nanocatalysts (blue line) [100]; (b) photocatalytic degradation of AB113 in the presence of magnetically recyclable CoFe2O4/TiO2 nanocatalysts in cycles [100]; and (c) magnetic responsiveness of CoFe2O4/TiO2 with an external magnetic field [153]. Republished with permission from Elsevier.
Catalysts 11 01543 g017
Table 1. Effect of synthesis method in structural and magnetic properties of catalyst cobalt ferrite [65,66,67].
Table 1. Effect of synthesis method in structural and magnetic properties of catalyst cobalt ferrite [65,66,67].
Synthesis MethodSol-GelCo-PrecipitationHydrothermalCombustion
Crystallite size (nm)37.333.015.038.0
Ms (emu·g−1)58.960.956.959.0
Table 2. Previous studies showing the photocatalytic degradation of organic pollutants using magnetic ferrites and ferrites-based zinc oxide and titanium oxides nanocomposites.
Table 2. Previous studies showing the photocatalytic degradation of organic pollutants using magnetic ferrites and ferrites-based zinc oxide and titanium oxides nanocomposites.
Magnetic Nanoparticles (MNPs)Organic PollutantsDye (mg/L)Photocatalyst (mg/L)Irradiation Time (min)Irradiation SourceDegradation (%)References
NiFe2O4/TiO2Methyl Orange UV–Vis90[120]
CoFe2O4Bromophenol Blue502524n/A2[121] a
CuFe2O4Bromophenol blue502524n/A48[121]
FeFe2O4Bromophenol blue502524n/A99[121]
MnFe2O4Bromophenol blue502524n/A0[121]
CoFe2O4Chicago Sky Blue502524n/A93[121]
CuFe2O4Chicago Sky Blue502524n/A95[121]
FeFe2O4Chicago Sky Blue502524n/A98[121]
MnFe2O4Chicago Sky Blue502524n/A91[121]
CoFe2O4Cu Phthalocyanine502524n/A70[121]
CuFe2O4Cu Phthalocyanine502524n/A75[121]
FeFe2O4Cu Phthalocyanine502524n/A92[121]
MnFe2O4Cu Phthalocyanine502524n/A19[121]
CoFe2O4Eosin Yellowish502524n/A25[121]
CuFe2O4Eosin Yellowish502524n/A53[121]
FeFe2O4Eosin Yellowish502524n/A85[121]
MnFe2O4Eosin Yellowish502524n/A12[121]
CoFe2O4Evans Blue502524n/A73[121]
CuFe2O4Evans Blue502524n/A92[121]
FeFe2O4Evans Blue502524n/A99[121]
MnFe2O4Evans Blue502524n/A8[121]
CoFe2O4Naphthol Blue Black502524n/A68[121]
CuFe2O4Naphthol Blue Black502524n/A95[121]
FeFe2O4Naphthol Blue Black502524n/A93[121]
MnFe2O4Naphthol Blue Black502524n/A75[121]
CoFe2O4Phenol Red502524n/A85[121]
CuFe2O4Phenol Red502524n/A86[121]
FeFe2O4Phenol Red502524n/A81[121]
MnFe2O4Phenol Red502524n/A63[121]
CoFe2O4Poly B-411502524n/A0[121]
CuFe2O4Poly B-411502524n/A7[121]
FeFe2O4Poly B-411502524n/A38[121]
MnFe2O4Poly B-411502524n/A0[121]
CoFe2O4Reactive Orange 16502524n/A21[121]
CuFe2O4Reactive Orange 16502524n/A86[121]
FeFe2O4Reactive Orange 16502524n/A77[121]
MnFe2O4Reactive Orange 16502524n/A6[121]
CuFe2O44-chlorophenol2003030UV–Vis81[115]
CuFe2O4-TiO24-chlorophenol2003030UV–Vis84[115]
Cu0.9Mn0.1Fe2O4/TiO24-chlorophenol2003030UV–Vis88[115]
Cu0.8Mn0.2Fe2O4/TiO24-chlorophenol2003030UV–Vis92[115]
Cu0.7Mn0.3Fe2O4/TiO24-chlorophenol2003030UV–Vis94[115]
Cu0.6Mn0.4Fe2O4/TiO24-chlorophenol2003030UV–Vis96[115]
Cu0.5Mn0.5Fe2O4/TiO24-chlorophenol2003030UV–Vis98[115]
(Co,Mn)Fe2O4@TiO2Azo dye1010960UV76[122]
ZnFe2O4Rhodamine B1020150200–700 nm60[123]
TiO2/ZnFe2O4Rhodamine B9.610150λ = 254 nm99.7[124]
TiO2/ZnFe2O4 (1:1)Rhodamine Bn/An/A150UV47[125]
TiO2/ZnFe2O4 (2:1)Rhodamine Bn/An/A150UV58[125]
TiO2/ZnFe2O4 (3:1)Rhodamine Bn/An/A150UV87[125]
TiO2/ZnFe2O4 (4:1)Rhodamine Bn/An/A150UV95[125]
ZnFe2O4Rhodamine B2080300UV–Vis38.4[126]
ZnFe2O4 nanospheresRhodamine B2080300UV–Vis100[126]
ZnFe2O4Rhodamine B20500360UV-light45[86]
ZnFe2O4 bRhodamine B20500360UV-light88[86]
ZnFe2O4 cRhodamine B20500360UV-light75[86]
ZnFe2O4 dRhodamine B20500360UV-light60[86]
ZnFe2O4 bRhodamine B20500360Dark0[86]
TiO2/CoFe2O4 (10%)Methylene Blue50.560UV56[127]
TiO2/CoFe2O4 (20%)Methylene Blue50.560UV60[127]
TiO2/CoFe2O4 (30%)Methylene Blue50.560UV57[127]
TiO2/Ni-Cu-Zn ferriteMethylene Blue2013120UV82[128]
TiO2/Ni-Cu-Zn ferriteMethylene Blue2020120UV98[128]
TiO2/Ni-Cu-Zn ferriteMethylene Blue2026120UV99[128]
TiO2/Ni-Cu-Zn ferriteMethylene Blue2033120UV95[128]
ZnFe2O4Methylene Blue106180UV28[129]
MnFe2O4 eMethylene Blue73001200Visible light15.2[92]
MnFe2O4 fMethylene Blue73001200Visible light67.2[92]
ZnFe2O4Methylene Blue10100360UV–Vis8[130]
ZnFe2O4 + H2O2Methylene Blue10100360UV–Vis52[130]
ZnFe2O4 + H2O2Methylene Blue10100360Dark45[130]
TiO2(57%)/CoFe2O4 (37%)Methyl Orange6n/A250UV0[114]
TiO2(62%)/CoFe2O4 (30%)Methyl Orange6n/A250UV25[114]
CoFe2O4/ZnOMethyl Orange5030300UV93.9[114]
TiO2/ZnFe2O4Methyl Orange880420UV80[131]
TiO2/ZnFe2O4 gMethyl Orange1050180UV–Vis5[87]
TiO2/ZnFe2O4 hMethyl Orange1050180UV–Vis13[87]
TiO2/ZnFe2O4 iMethyl Orange1050180UV–Vis27[87]
TiO2/ZnFe2O4 (0.15%)Methyl Orange255240UV–Vis65[132]
TiO2/ZnFe2O4 (0.30%)Methyl Orange255240UV–Vis75[132]
TiO2/ZnFe2O4 (1.5%)Methyl Orange255240UV–Vis84[132]
TiO2/ZnFe2O4 (3.0%)Methyl Orange255240UV–Vis73[132]
TiO2/ZnFe2O4 (6.05%)Methyl Orange255240UV–Vis55[132]
ZnFe2O4Methyl Orange255240UV–Vis4[132]
ZnFe2O4Methyl Orange10460UV-light75[133]
ZnFe2O4Methyl Orange10100240UV–Vis5[134]
TiO2/ZnFe2O4Methyl Orange10100240UV–Vis40[134]
TiO2/ZnFe2O4 (1.5%)Methyl Orange10100240UV–Vis12[134]
TiO2/ZnFe2O4 (3.0%)Methyl Orange10100240UV–Vis34[134]
TiO2/ZnFe2O4 (4.5%)Methyl Orange10100240UV–Vis24[134]
TiO2/ZnFe2O4 (6.0%)Methyl Orange10100240UV–Vis18[134]
CuFe2O4-TiO2Methylene Blue501000180UV–Vis83.7[135]
ZnFe2O4/ZnOMethylene Blue201000360UV90[136]
ZnFe2O4/ZnORemazol Brilliant Blue201000360UV100[136]
CoFe2O4 + H2O2Rhodamine B10100270UV-Vis90.6[137]
CuFe2O4-TiO2Methylene Blue12100150UV-Vis47[111]
a The reactions of various dyes were evaluated with ferrites and H2O2 with no applied irradiation [121]. b 9 nm average crystal size. c 14 nm average crystal size. d 19 nm average crystal size. e Thermal preparation. f Seed-hydrothermal preparation. g Prepared by citrate–nitrate method. h Prepared by thermal method. i Prepared by thermal method, recovered, and sintered at 800 °C.
Table 3. Reusability of magnetic nanoparticles (MNPs) for organic pollutants degradation.
Table 3. Reusability of magnetic nanoparticles (MNPs) for organic pollutants degradation.
Magnetic NanocompositesOrganic PollutantsDye (mg/L)Photocatalyst (g/L)Irradiation Time (min)pHCycles of Reusability of MNPsDegradation Efficiency (%)Irradiation SourceReferences
CoFe2O4/TiO2-SiO2Methylene blue dye0.30.33306693.2UV[10]
NiFe2O4Reactive blue 50.050.03101–11485.0n/A[83]
NiFe2O4/ZnOCongo red (CR)0.020.05405.5597.0UV–Vis[16]
CoFe2O4/ZnOMethylene blue dye50.02560n/A3-UV–Vis[5]
CoFe2O4/TiO2Methylene blue dyen/An/A360n/A593.8UV[153]
ZnFe2O4/TiO2/CuNaproxen0.030.11204–9572.3Sunlight[154]
ZnFe2O4/TiO2Rhodamine B101303–115>99UV–Vis[155]
CoFe2O4/TiO2/rGOChlorpyrifos50.4605.88-UV[152]
ZnFe2O4/TiO2Methylene blue dye200.053605.6593.2UV–Vis[151]
ZnFe2O4/ZnOMethylene blue dyen/A0.1120n/A3-UV[17]
NiFe2O4/TiO2Methyl Orange1013009.5–103-Visible light[156]
Cu0.5Mn0.5Fe2O4/TiO24-chlorophenol300.23008598.0UV[115]
ZnO/MnFe2O4Methylene blue dyen/An/A3003–9686.0UV–Vis[113]
CuFe2O4/TiO2Methylene blue dye3011202.2381.2UV–Vis[135]
ZnFe2O4/ZnOMethylene blue dye100.12009–10379.0UV[85]
MnFe2O4/TiO2-rGOCiprofloxacin300.063001–6675.0UV–Vis[106]
ZnFe2O4/TiO2Bisphenol A1013009.5–105>90UV–Vis[157]
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Mapossa, A.B.; Mhike, W.; Adalima, J.L.; Tichapondwa, S. Removal of Organic Dyes from Water and Wastewater Using Magnetic Ferrite-Based Titanium Oxide and Zinc Oxide Nanocomposites: A Review. Catalysts 2021, 11, 1543. https://0-doi-org.brum.beds.ac.uk/10.3390/catal11121543

AMA Style

Mapossa AB, Mhike W, Adalima JL, Tichapondwa S. Removal of Organic Dyes from Water and Wastewater Using Magnetic Ferrite-Based Titanium Oxide and Zinc Oxide Nanocomposites: A Review. Catalysts. 2021; 11(12):1543. https://0-doi-org.brum.beds.ac.uk/10.3390/catal11121543

Chicago/Turabian Style

Mapossa, António B., Washington Mhike, José L. Adalima, and Shepherd Tichapondwa. 2021. "Removal of Organic Dyes from Water and Wastewater Using Magnetic Ferrite-Based Titanium Oxide and Zinc Oxide Nanocomposites: A Review" Catalysts 11, no. 12: 1543. https://0-doi-org.brum.beds.ac.uk/10.3390/catal11121543

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