Next Article in Journal
Effects of Support and Synthetic Procedure for Sol-Immobilized Au Nanoparticles
Next Article in Special Issue
Photocatalytic Removal of Microbiological Consortium and Organic Matter in Greywater
Previous Article in Journal
Rh(III)-Catalyzed, Highly Selectively Direct C–H Alkylation of Indoles with Diazo Compounds
Previous Article in Special Issue
Kinetic Model for Simultaneous Adsorption/Photodegradation Process of Alizarin Red S in Water Solution by Nano-TiO2 under Visible Light
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Advances in Magnetically Separable Photocatalysts: Smart, Recyclable Materials for Water Pollution Mitigation

Nanotechnology and Water Sustainability Research Unit, College of Science, Engineering and Technology, University of South Africa, Florida 1709, Johannesburg, South Africa
*
Author to whom correspondence should be addressed.
Catalysts 2016, 6(6), 79; https://doi.org/10.3390/catal6060079
Submission received: 16 April 2016 / Revised: 12 May 2016 / Accepted: 18 May 2016 / Published: 20 June 2016
(This article belongs to the Special Issue Photocatalytic Wastewater Treatment)

Abstract

:
Organic and inorganic compounds utilised at different stages of various industrial processes are lost into effluent water and eventually find their way into fresh water sources where they cause devastating effects on the ecosystem due to their stability, toxicity, and non-biodegradable nature. Semiconductor photocatalysis has been highlighted as a promising technology for the treatment of water laden with organic, inorganic, and microbial pollutants. However, these semiconductor photocatalysts are applied in powdered form, which makes separation and recycling after treatment extremely difficult. This not only leads to loss of the photocatalyst but also to secondary pollution by the photocatalyst particles. The introduction of various magnetic nanoparticles such as magnetite, maghemite, ferrites, etc. into the photocatalyst matrix has recently become an area of intense research because it allows for the easy separation of the photocatalyst from the treated water using an external magnetic field. Herein, we discuss the recent developments in terms of synthesis and photocatalytic properties of magnetically separable nanocomposites towards water treatment. The influence of the magnetic nanoparticles in the optical properties, charge transfer mechanism, and overall photocatalytic activity is deliberated based on selected results. We conclude the review by providing summary remarks on the successes of magnetic photocatalysts and present some of the future challenges regarding the exploitation of these materials in water treatment.

Graphical Abstract

1. Introduction

Rapid development in terms of industrialisation and population growth over the years has put pressure on sustainable clean water supply. This is partly due to the ever increasing demand and the significant contribution towards compromised fresh water quality that comes with population growth and industrialisation. Similarly, climate change (drought conditions and elevated temperatures) has also impacted negatively on the amount of fresh water available. Consequently, the development of effective water treatment technologies remains an important area of research. This would ensure adequate utilisation of the available fresh water sources, which are constantly polluted through various domestic and industrial activities. Moreover, it would allow efficient treatment of industrial effluent water and enable it to be recycled which would lessen the burden on constant fresh water supply. This is particularly key to the smooth operations of the industries in water scarce localities where the amount of fresh water allocated per industry per day is highly restricted [1].
Over the years, a number of water treatment technologies such as membrane processes [1,2], biological methods [3,4], advanced oxidation processes (AOPs) [5,6], adsorption [7,8], electrochemical methods [9,10], etc. have been developed and explored for the destruction/removal of organic, inorganic, and microbial pollutants. These methods have also been explored in various combinations in order to improve their efficiency [11,12,13,14]. Despite the impressive level of success observed with these technologies in water treatment, they often suffer a number of setbacks, which relate to inefficiency, high operational costs, or both. For example, in membrane processes, fouling (organic, inorganic and biofouling) is a major problem that results in the alteration of the membrane flux and rejection properties [15,16,17]. Toxicity of the effluent matrix and the non-biodegradable nature of some pollutants often render biodegradation ineffective. Moreover, some of the pollutants may be transformed into more toxic compounds during biological treatment, resulting in secondary pollution [18,19,20]. The generation of highly concentrated sludge which poses further treatment and disposal challenges is a major setback associated with adsorption and coagulation–flocculation [21].
Advanced oxidation processes such as ozonation, hydrogen peroxide oxidation, and their combination with UV light have been extensively exploited in water treatment; however, problems such as the short half-life of ozone, activation by UV light, incomplete mineralisation of organic compounds, and the costs associated with these tools are usually cause for concern [12,22]. Semiconductor photocatalysis has rapidly emerged as the most attractive AOP in recent years for the destruction of organic, inorganic, and microbial pollutants due to a number of advantages it bears over other water treatment protocols. Such merits include its versatile nature in terms of the type of pollutants that can be removed (organic, inorganic, and microbial) and the medium in which the process is applicable (liquid and aqueous medium) [23,24,25,26,27]. Moreover, the semiconductor photocatalysts can also be exploited in other applications such as sensors [28], solar cells [29], pollutant adsorption [30], catalysts for organic synthesis [31], water splitting [32], etc. Apart from its versatile nature, semiconductor photocatalysis has the potential to completely mineralise organic pollutants into carbon dioxide, water, and inorganic ions, thereby eliminating the problem of sludge formation which otherwise cause secondary pollution. Semiconductor photocatalysis provides a stable reaction site/medium since the photocatalyst is not used up during the process and can be re-used several times, and the toxicity of some of the pollutants or their matrices does not affect their destruction by photocatalysis as would be the case with biodegradation [33,34]. In addition, photocatalysis has the advantage of being applicable at ambient conditions (room temperature and standard pressure).
Despite all the highlighted merits associated with semiconductor photocatalysis for water treatment, practical exploitation of this technology has remained a daunting challenge. The first challenge relates to the size of the band gap of the semiconductor since this determines the photoresponse of the photocatalyst. Some of the widely exploited semiconductors such as TiO2, ZnO, SnO2, etc. have wide band gaps of 3.0–3.2, 3.2, and 3.5 eV, respectively, which means these semiconductors can only be activated by UV light irradiation. However, artificial UV sources are costly, and UV light itself is toxic and requires protective gear when working under it. The sun does supply UV light but it only makes up about 4% of the solar spectrum, making practical exploitation unsustainable [35,36]. The second challenge in semiconductor photocatalysis is the fast recombination rate of the photogenerated charge carriers, which ultimately lowers the photocatalytic activity of the semiconductor [37,38]. Ideally, subsequent to light absorption and electron excitation in the semiconductor, the electrons, which are promoted to the conduction band, and the positive holes, which remain in the valence band of the semiconductor, must remain far apart so that they can react with oxygen and water or hydroxide ions to form the superoxide and hydroxyl radicals, respectively. These are the active species that together with the holes may be responsible for the degradation of the pollutants [39,40]. Thirdly, the slow degradation kinetics also hamper the full exploitation of photocatalysis in water treatment. The residence times are usually long and sometimes accompanied by low degradation efficiencies [41]. Incomplete mineralisation of organic pollutants is another problem usually observed in photocatalytic experiments and may lead to the formation of toxic by-products which are more toxic than the parent compound [42]. Lastly, the photocatalysts are usually applied in powdered form, which may undergo aggregation in the aqueous medium thereby lowering its activity. Moreover, after the treatment has been completed, the photocatalyst needs to be separated and recovered from the water, and this presents an expensive and time-consuming exercise, especially when it comes to nanometre-sized particles. This often leads to inefficient recovery of the photocatalyst, which not only affects its recyclability but also causes secondary pollution and escalates the treatment costs.
Over the past years, tremendous work has been done focusing on improving the visible light response of various photocatalysts, enhancing charge carrier separation and the overall photocatalytic activity. Efficient utilisation of visible light would enable the exploitation of sunlight as a source of energy, which consists of about 40% visible light. Generally, strategies such as doping the semiconductor with metal ions (alkali, alkaline earth, noble, rare earth, and transition metals) [43,44,45,46,47], non-metal ions (oxygen, sulphur, nitrogen, phosphorus, halogens, carbon, and boron) [48,49,50,51], codoping/multidoping (metal + metal, non-metal + metal, non-metal + non-metal) [52,53,54,55], coupling semiconductors with carbon nanomaterials (reduce graphene oxide (RGO), graphene oxide (GO), multiwalled carbon nanotubes (MWCNTs), single-walled carbon nanotubes (SWCNTs), carbon nanospheres (CNS), fullerenes, etc.) [56,57,58,59], sensitisation (dye, polymer and surface complex sensitisation) [60,61,62], and coupling two or more semiconductors [63,64,65] have been extensively investigated for improving the photocatalytic activity of various semiconductors. Encouraging results have been reported, especially for the nanocomposite photocatalysts, which showed improved visible light utilisation and efficient charge separation, which could be ascribed to the combined contribution of the individual components of the nanocomposite.
Engineering of materials with high visible light photocatalytic activity that are easily separable and recoverable from the treated water remains the ultimate goal of ongoing research. With the increased usage of nanophotocatalysts, efficient separation and recovery of these materials from water is becoming increasingly difficult. Immobilising the photocatalyst particles on solid supports such as glass, ceramics, polymers, zeolites, sand, etc. has been widely investigated and has successfully enabled easy separation of the photocatalyst [66,67,68,69,70,71]. However, in the process of attaining easy separation, the photocatalytic activity of the immobilised semiconductors decreased remarkably. This was due to a decrease in the surface area, poor interaction with the reaction medium, and limited exposure of the photocatalyst to light [72,73]. Since neither the photocatalytic activity nor the easy separation can be sacrificed, researchers turned their interests to magnetic nanoparticles as integral ingredients for developing photocatalysts, which were easily separated from water with an external magnetic field [74,75,76]. Magnetic nanoparticles such as haematite (α-Fe2O3), maghemite (γ-Fe2O3), magnetite (Fe3O4), ferrites (MFe2O4, M = Mg, Ni, Zn, Cu, Co, Cd, etc.), etc. have been successfully coupled with various photocatalysts to induce magnetic behaviour in the composite material and allow easy magnetic separation [39,77,78,79]. Since the photocatalyst remains powdery upon incorporation of the magnetic nanoparticles, this does not compromise its surface area and the distribution of the photocatalyst in the aquatic medium. In addition, some of these magnetic nanoparticles have good visible light absorption and photocatalytic properties, which further enhance the overall pollutant degradation kinetics over the magnetic nanocomposite [39,40,64,80,81,82].
Considering the future prospects of magnetic photocatalysts, the design and preparation of magnetic visible-light-active nanocomposites has emerged as a hot topic in photocatalytic environmental pollution control. Therefore, it is necessary to bring into perspective the current developments in the tailoring and application of such materials in water treatment. A few reviews are available mainly on the synthesis, structure, and various applications of magnetic materials [83,84,85,86,87,88]. This review exclusively discusses the most recent developments in terms of the synthesis and exploitation of various magnetic nanocomposite photocatalysts in water pollution mitigation. Details on the synthesis routes and structure of the various magnetic nanoparticles have been well-reviewed and have been excluded in this review. Furthermore, this review only focusses on iron oxide magnetic nanoparticles resulting in four categories of magnetic nanocomposite photocatalysts: haematite, maghemite, magnetite, and ferrite-based magnetic photocatalysts. Under each category, we discuss the influence of the magnetic nanoparticle on optical properties, charge transfer mechanism, and the overall photocatalytic properties of the nanocomposites. We conclude the review by providing some summary remarks and highlight the future challenges towards tailoring and utilising magnetic photocatalysts in water treatment.

2. Recent Developments in Magnetic Photocatalysts

2.1. Haematite (α-Fe2O3)-Based Magnetic Photocatalysts

As a pristine material, haematite is a narrow band gap semiconductor (2–2.2 eV) with good visible light response, good chemical stability, low-cost, ferromagnetic behaviour, abundantly available, high resistance to corrosion, and environmentally friendly [89,90,91]. On its own, haematite shows poor photocatalytic activity due to the high recombination rate of the charge carriers and poor conductivity [90,91]. Despite its poor photocatalytic activity, the narrow band gap of haematite makes it an ideal sensitiser for improving the visible light response of various semiconductors. Zhao et al. developed a nanostructured α-Fe2O3-AgBr nonwoven cloth (Figure 1a,b) via a simple electrospinning-calcination method and studied its visible light photocatalytic behaviour towards rhodamine B (RhB) and para-chlorophenol (4-CP) decomposition. For both pollutants, the α-Fe2O3-AgBr nonwoven cloth showed superior activity over α-Fe2O3 and AgBr clothes individually, attaining 91.8% 60 min and 74.2% in 120 min for RhB and 4-CP decomposition, respectively [81]. Coupling AgBr with α-Fe2O3 resulted in a significant red-shift and resulted in the formation of heterojunctions, which facilitated charge separation, thereby improving the overall photocatalytic activity of the α-Fe2O3-AgBr nonwoven cloth. In addition, the cloth was easy to handle and recycle during photocatalysis by simple dipping and removal or by using an external magnet. It was found that the composite cloth was stable over four cycles with only a slight decrease in photocatalytic performance [81].
Xue and coworkers fabricated magnetically separable α-Fe2O3/graphitic carbon nitride (g-C3N4) nanocomposite photocatalysts with varying α-Fe2O3 loadings (5%, 10%, 30%, 50%) and investigated their performance for degradation of RhB. The highest degradation efficiency (96.7% in 4 h) was recorded using the sample containing 10% α-Fe2O3, which suggests optimum composition [92]. The introduction of α-Fe2O3 nanoparticles into g-C3N4 matrix enhanced its visible light utilisation and charge separation efficiency. Moreover, the composite photocatalyst displayed good stability and could be easily separable and recycled using an external magnet [92]. Similarly, core-shell magnetically separable Ag/AgCl/Fe2O3 showed remarkable photocatalytic activity towards bisphenol A (BPA) and E. coli under visible light exposure. At a 5% α-Fe2O3 loading, the composite photocatalyst reached the highest activity, which was 13 times that of Ag/AgCl for BPA degradation, and complete destruction of E. coli was observed after just 30 min of irradiation [39]. The nanocomposite displayed good stability and sufficient magnetic response to an external magnet.
In terms of charge transfer mechanism, the path followed by the electrons and holes depends on the semiconductor coupled with haematite and how their band structures match. In any case, under visible light illumination, haematite is excited and electrons accumulate in the conduction band while leaving positive holes in the valence band. In the case of α-Fe2O3-AgBr, both semiconductors are excited by visible light, and the electrons from AgBr transfer to the conduction band of α-Fe2O3, while the holes transfer in the opposite direction (Figure 1e). The holes will directly attack the organic pollutants to form the degradation products, while the electrons may react with adsorbed oxygen leading to the eventual formation of hydroxyl radicals, which are strong oxidising agents [81]. In the Ag/AgCl/Fe2O3 ternary nanostructure, the charge transfer mechanism (Figure 1f) is a bit complex. Only Ag and α-Fe2O3 are excited upon visible light absorption, and the plasmon-generated electrons in Ag will be injected into the conduction band of AgCl and further transfer to the conduction band of α-Fe2O3. Meanwhile, the holes in α-Fe2O3 can directly decompose the pollutants, while the holes left in Ag react with Cl ions to form Cl0, which then attacks the pollutants [39].
Other haematite-based magnetically separable nanostructured photocatalysts such as graphene (GR)-α-Fe2O3-ZnO [93], α-Fe2O3/TiO2 nanofibres [94], α-Fe2O3/RGO [95], and α-Fe2O3/ZnO [96] have been fabricated, have shown remarkable visible light utilisation, and have been easily separable using an external magnetic field. Coupling the various semiconductors with α-Fe2O3 resulted in synergy between the materials, leading to enhanced visible light response, improved charge separation, and an overall increase in photocatalytic activity. Moreover, the incorporation of haematite allowed for the easy separation of the photocatalysts after the water treatment. However, α-Fe2O3 generally shows weak ferromagnetic behaviour, and, most importantly, its magnetic properties largely depend on the synthesis method, which influences its particle size and shape [92]. Several α-Fe2O3-incorporating heterostructures such as α-Fe2O3/Ag3VO4 [91], g-C3N4/Ag/α-Fe2O3 [97], g-C3N4/α-Fe2O3 [98,99,100], α-Fe2O3/TiO2 [101] Au/g-C3N4/α-Fe2O3 [102], and α-Fe2O3/CdS [90] have been fabricated, showing enhanced visible light photocatalytic properties, and no magnetic behaviour was reported. Therefore, when the main aim of incorporating the iron oxide nanoparticles is to induce magnetic response, haematite may not be the ideal choice.

2.2. Maghemite (γ-Fe2O3)-based Magnetic Photocatalysts

Maghemite is another important form of iron oxide but, unlike haematite, it shows good magnetic properties with saturation magnetisation values of up to 76 emu/g [103]. Consequently, the incorporation of maghemite into various photocatalyst matrices has been explored to fabricate magnetically separable composite photocatalysts for easy recovery and recycling [104,105]. Interestingly, despite the sensitivity to an external magnetic field, when the magnetic field is removed, the γ-Fe2O3-incorporating nanocomposites do not retain any significant residual magnetism that would otherwise cause problems by creating large clusters of magnetised particles during water treatment [106]. Maghemite shows good visible light response owing to its narrow band gap (2.2 eV) and possesses good adsorption properties. However, its photocatalytic activity is poor as a result of the high recombination rate of the photogenerated charge carriers and photodissolution [74,107]. Two approaches are usually adopted when coupling maghemite with other semiconductors. In the first approach, maghemite is directly coupled with other semiconductors to form heterojunctions, and it contributes towards visible light response of the nanocomposite photocatalyst as well as the overall charge transfer process [74,108,109]. According to the second approach, maghemite and the other semiconductor do not form heterojunctions (not in direct contact), but they are separated by an inert layer or shell usually silica (SiO2). This is aimed at preventing photodissolution of the iron oxide and charge transfer between the two materials, which could otherwise result in the recombination on the maghemite surface. In this case, maghemite nanoparticles do not play any part in the photocatalysis process but function solely in the magnetic separation of the photocatalyst from the reaction medium [104,106,110].
A magnetically separable quaternary γ-Fe2O3/N and Fe-codoped TiO2 nanocomposite was prepared using a sol-gel method (as illustrated in Figure 2a) and employed for the decomposition of reactive blue 4 (RB 4) under visible light illumination. Significantly higher photocatalytic activity was observed over the quaternary nanocomposite photocatalyst compared to N,Fe-TiO2, N-TiO2, and TiO2, reaching 100% degradation in 180 min (Figure 2b) [111]. The remarkable improvement in activity of the nanocomposite could be explained in terms of the combined contribution of codoping TiO2 and coupling with γ-Fe2O3, which resulted in significant band gap narrowing, efficient visible light response, and efficient charge separation and transfer. Furthermore, the multicomponent photocatalyst demonstrated good stability and recyclability over four cycles without any significant loss in activity or magnetic properties, which allowed for easy magnetic separation [111]. Yu and coworkers probed the adsorption and UV photocatalytic oxidation of As(III) over TiO2 decorated on magnetic (γ-Fe2O3) mesoporous SBA-15 nanocomposite (γ-Fe2O3/SBA-15/TiO2) derived from simple inner-pore hydrolysis combined with solvent evaporation route [112]. The nanocomposite photocatalyst displayed a dual function of adsorption and photocatalytic degradation. Firstly, As(III) ions were adsorbed onto the photocatalyst and then upon UV irradiation, undergo photocatalytic oxidation to As(V), which is less toxic and also removable by the photocatalyst. In addition, the photocatalyst showed good stability and could be regenerated for five cycles using NaOH (0.01 M) and still maintained over 90% As(III) removal and good magnetic properties for easy separation [112]. The incorporation of the SBA-15 ensured good dispersibility of the nanoparticles and improved the surface area of the nanocomposite. All three components of the nanocomposite were equally important in the overall removal of As(III).
Improved optical and photocatalytic properties have also been reported for other magnetic photocatalysts such as ZnO/γ-Fe2O3 [113], g-C3N4/γ-Fe2O3 [109], and γ-Fe2O3/TiO2 [74] towards the degradation of various pollutants including RhB, methylene blue (MB), and methyl orange (MO). Moreover, these nanostructures showed good stability and could be magnetically separated and recycled without a significant loss in activity. However, the maghemite nanoparticle loading needs to be carefully controlled in order to strike a good balance between sufficient magnetism to facilitate separation and the detrimental effect of recombination of the charge carriers on the maghemite nanoparticles. This means that, while enjoying the luxury of easy separation, the detrimental effect of recombination must not be ignored in order to positively benefit from the incorporation of the maghemite nanoparticles.
According to the proposed charge transfer route in γ-Fe2O3/N,Fe-TiO2 (Figure 3a), both maghemite and codoped titania absorb visible light, and electrons are excited from the valence band to the conduction band in γ-Fe2O3, leaving holes in the valence band. Similarly, in the codoped titania, electrons are excited from the valence band of TiO2 to the impurity states formed by Fe doping and also from N sub-band gap states to the conduction band of TiO2. Subsequently, electrons transfer from the conduction band of maghemite to the Fe impurity states in TiO2, while the holes migrate to the N states in TiO2. The electrons and holes will be captured by molecular oxygen and water to form the superoxide and hydroxyl radicals, respectively [111]. These radical species are important oxidising agents for pollutant decomposition.
The strategy of preventing the formation of heterojunctions between maghemite and the semiconductor coupled with it has been widely explored using silica as the barrier/shell surrounding the magnetic core. For example, Szeto et al. used a layer-by-layer assembly route to fabricate urchin-like γ-Fe2O3@SiO2@TiO2 composite microparticles (Figure 3b–e) and observed enhanced phenol decomposition under UV irradiation [114]. The tailored composite photocatalyst was shown to possess a large surface area, a prolonged life of the charge carriers, remarkable permeability, and significantly higher photocatalytic activity compared to commercial P25. In addition, incorporation of the maghemite nanoparticles ensured easy and efficient separation of the photocatalyst using an external magnetic field [114]. Similar results have been reported for other magnetic heterostructures such as γ-Fe2O3@SiO2@Ce-TiO2 [110], γ-Fe2O3@SiO2@AgBr:Ag [104], and γ-Fe2O3@SiO2@TiO2 [106], whereby a silica coating separates the maghemite nanoparticles from direct contact with the other components of the heterostructure. In these nanocomposites, the maghemite nanoparticles were only incorporated to facilitate the magnetic separation and play no part in the charge transfer process and overall photocatalytic decomposition of the pollutants. Although this approach is logical and has yielded some positive results, a critical comparison between the heterojunction type of photocatalyst and the core-shell structure of the same nanocomposite is rarely documented. This could map a way forward in terms of the most efficient approach towards engineering the maghemite-based magnetic photocatalyst and benefitting from the properties of this material.

2.3. Magnetite (Fe3O4)-based Magnetic Photocatalysts

Magnetite is by far one of the most widely exploited iron oxide magnetic nanoparticle in the fabrication of magnetic photocatalysts due to its low toxicity, good magnetic properties, biocompatibility, and remarkable adsorption properties [115,116]. The magnetisation saturation for magnetite can reach values of up to 92 emu/g, making it the ideal candidate for synthesis of magnetically separable photocatalysts [103]. However, unlike maghemite, magnetite shows no photocatalytic properties, and, like maghemite, the formation of heterojunctions with other semiconductors in nanocomposite photocatalysts is undesirable because magnetite may suppress the overall photocatalytic efficiency by acting as a recombination centre for the charge carriers [117,118]. Silica is commonly utilised to enclose the magnetite core, which is wrapped with the other semiconductor nanoparticles to form the magnetic composite photocatalysts. The silica barrier prevents electron transfer between magnetite and the other semiconductor, additionally preventing photodissolution of the iron oxide [119,120].
Despite the possible negative effect of magnetite in heterojunction nanocomposites, several reports on successful tailoring and improvement in photocatalytic activity over heterojunction nanocomposites have emerged, and careful control of the Fe3O4 loading in the nanocomposite is an important point of consideration. For example, Zhao et al. observed enhanced photocatalytic activity towards ampicillin (AMP) over Fe3O4/TiO2/Ag with a sea urchin-like morphology, both under visible and UV light irradiation. The composite photocatalyst reached 98.7% and 91.5% AMP degradation after 360 min of UV and visible light illumination, respectively [116]. Furthermore, the nanocomposite possessed good magnetic behaviour with a magnetisation saturation of 26.5 emu·g−1 (Figure 4), which was high enough to facilitate magnetic separation (Figure 4 insert). In addition to the degradation of AMP, the nanocomposite photocatalyst showed remarkable photocatalytic properties towards S. aureus (G+), E. coli (G), and A. niger [116]. In another example, improved MO decomposition was realised over Fe3O4/N-TiO2/Ag hollow nanospheres, reaching 99.5% in 80 min of visible light exposure when the Ag loading was 1.0% [121]. Ma and coworkers employed a low-temperature crystallisation method to wrap Fe3O4 spheres with TiO2 nanoparticles and subsequently immobilised Au nanoparticles on the core-shell Fe3O4/TiO2 to yield a Fe3O4/TiO2/Au plasmonic magnetic photocatalyst. RhB was used as a model pollutant to evaluate the photocatalytic properties of the nanocomposite, and it showed superior performance over Fe3O4 and Fe3O4/TiO2, with good magnetic properties (magnetisation saturation (Ms), 44.6 emu·g−1), and recyclability [122]. The heterojunctions formed between Fe3O4 and titania allow electrons to transfer between the two materials, which improves charge separation. However, the Fe3O4 loading needs to be kept in check as excess Fe3O4 will cause detrimental effects. Therefore, Fe3O4 nanoparticles have a dual function in the heterojunction nanostructures—electron trapping and enabling magnetic separation.
Li et al. fabricated Fe3O4 magnetic core coated with SiO2 prior to the deposition of TiO2 and surface modification with lysine, and the quaternary nanocomposite photocatalyst (Fe3O4@SiO2@TiO2-lysine) showed multifunctional properties as it was employed as an adsorbent, photocatalyst, and sensor for dissolved organic and inorganic phosphorus in seawater [123]. The photocatalyst displayed remarkable stability and could be magnetically separated and recycled for 10 cycles without any significant loss in its adsorption and photocatalytic properties [123]. Other core-shell Fe3O4@SiO2 nanostructures decorated with TiO2 have been fabricated and evaluated for the decomposition of various pollutants such as RhB, 2-chlorophenol (2-CP), phenol, MB, and MO. These materials showed improved photocatalytic properties, good stability, and sufficiently high magnetic response that enabled separation using an external magnetic field [124,125,126,127]. Table 1 provides a summary of the synthesis and photocatalytic properties of various other magnetite-incorporating titania-based photocatalysts towards water treatment applications. In the magnetic core-shell nanostructures, the Fe3O4 nanoparticles encapsulated with the silica shell do not partake in the charge transfer process but only in the magnetic separation of the photocatalyst.
Chidambaram and coworkers prepared a novel plasmonic, magnetic Fe3O4/ZnO/Ag nanocomposite and employed it for the photocatalytic degradation of MB under visible light illumination. Complete MB removal was observed after 6 h of visible light exposure on Fe3O4/ZnO/Ag, while only 20% MB removal was realised over Fe3O4/ZnO during the same irradiation time. The improved activity could be explained in terms of improved charge separation and the surface plasmon resonance (SPR) effect of Ag nanoparticles, which is responsible for the visible light activity of the nanocomposite photocatalyst [139]. Fe3O4 nanoparticles played a dual role of facilitating charge separation and enabling magnetic separation. In addition, the photocatalyst showed great stability over 10 cycles without much loss in activity or magnetic properties. Upon visible light irradiation, the SPR-generated electrons are injected into the conduction band of ZnO and further transfer to the Fe3O4 nanoparticles. Subsequently, the electrons reduce Fe3+ to Fe2+, which in turn is oxidised back to Fe3+ upon reaction with adsorbed oxygen forming the superoxide radicals. Meanwhile, the SPR-generated holes in Ag could react with water to form the hydroxyl radicals and together with the superoxide radicals are responsible for pollutant degradation [139]. Yan et al. observed higher visible light photocatalytic performance over urchin-like Fe3O4@SiO2@ZnO/CdS core-shell microspheres (Figure 5a,b) towards RhB compared to Fe3O4@SiO2@ZnO microspheres. Coating Fe3O4 with SiO2 prevented the formation of heterojunctions with ZnO and CdS, and the nanocomposite showed good magnetic properties with magnetisation saturation value of 18.97 emu/g [140]. On visible light exposure, only CdS nanoparticles were excited and played the role of a sensitiser for ZnO, with electrons transferred to the conduction band of ZnO while the holes remained in the valence band of CdS (Figure 5c), resulting in efficient charge separation and the formation of the oxidising species [140].
Habibi-Yangjeh’s group has recently prepared and investigated the visible light photocatalytic properties of various magnetic nanocomposites-based on Fe3O4 and ZnO such as ZnO/AgI/Fe3O4 [141], ZnO/AgBr/Fe3O4/Ag3VO4 [142], ZnO/Ag3VO4/Fe3O4 [143], Fe3O4/ZnO/AgCl [144], and Fe3O4/AgBr-ZnO [145] towards model pollutants such as RhB, phenol, MB, and MO. All of the nanocomposite photocatalysts showed good photocatalytic properties, which can be credited to efficient charge separation and improved visible light absorption as a result of the heterojunctions formed between ZnO and the co-catalysts. Moreover, it was shown that despite the formation of the composite photocatalysts, the presence of Fe3O4 nanoparticles ensured that the photocatalysts still retained sufficient magnetic response to allow for efficient separation using an external magnetic field. Obviously, the magnetisation saturation of Fe3O4 decreased sharply upon the formation of the nanocomposites due to their interaction and coverage by the non-magnetic components of the photocatalysts [141,142,143,144,145]. Under visible light irradiation, ZnO remains inactive and only the co-catalyst with a narrow band gap can be excited. For example, in ZnO/AgI/Fe3O4 and ZnO/AgBr/Fe3O4/Ag3VO4, only AgI, AgBr, and Ag3VO4, respectively, are excited, and electrons are promoted to the conduction band, while the holes remain in the valence band. Subsequently, the electrons transfer to the conduction band of ZnO where they react with adsorbed oxygen to form the radical species (superoxide and hydroxyl radicals). Accordingly, the holes are directly involved in the degradation of the pollutants since they are not positive enough to oxidise water for the hydroxyl radicals [141,142]. The electrons and holes occupied different sites in the nanocomposite photocatalysts, which ensured efficient charge separation and the formation of the reactive species for pollutant degradation.
Owing to its meta-free nature, polymeric structure, non-toxicity, good visible light response, good photocatalytic properties, and high thermal and chemical stability, tailoring graphitic carbon nitride (g-C3N4)-based magnetic photocatalysts is an important exploit towards potential practical application of these materials in water treatment. Using a combination of calcination and co-precipitation, Yang and coworkers prepared Fe3O4/g-C3N4 nanocomposite photocatalyst and probed its photocatalytic behaviour towards 2,4,6-trichlorophenol (2,4,6-TCP) under visible light illumination [77]. In a similar work, Jia et al. fabricated Fe3O4/g-C3N4 via a hydrothermal route and studied its visible light photocatalytic performance towards RhB decomposition [146]. In both studies, higher photocatalytic performance was observed over the nanocomposites compared to Fe3O4 and g-C3N4 individually, reaching 95.5% in 60 min [146] and 96.5% in 100 min [77] for RhB and 2,4,6-TCP, respectively. Coupling g-C3N4 with Fe3O4 improved charge separation since only g-C3N4 is excited by visible light, and the electrons in the conduction band quickly transfer to Fe3O4, where they are captured by oxygen to form the radical species. Meanwhile, the holes remain in the valence band of g-C3N4 and directly oxidise the pollutants to form the degradation products. Therefore, the holes and electrons occupy different locations in the nanocomposite. Moreover, magnetic separation was made possible by the presence of the Fe3O4 nanoparticles [77,146].
Habibi-Yangjeh’s group investigated the visible light photocatalytic performance of novel magnetic nanocomposites: g-C3N4/Fe3O4/BiOI [147], g-C3N4/Fe3O4/AgCl [148], and g-C3N4/AgBr/Fe3O4 [149] towards the decomposition of RhB, MO, and MB. Significantly higher photocatalytic activity was observed over the ternary heterostructures compared to single and binary nanocomposites, and this could be ascribed to the synergistic effect of the various components of the ternary nanocomposites, resulting in improved visible light response and charge separation efficiency. Moreover, the presence of Fe3O4 in the heterostructures endowed them with magnetic response, which allowed separation using an external magnetic field [147,148,149]. In terms of charge transfer, only g-C3N4 is excited by visible light in the case of g-C3N4/Fe3O4/AgCl, while both g-C3N4 and BiOI are excited in g-C3N4/Fe3O4/BiOI, and the electrons are promoted to their conduction bands, leaving positive holes in the valence bands. This is followed by the transfer of electrons from the conduction band of g-C3N4, which is more negative, to the conduction bands of AgCl and BiOI, which are less negative, whereby photoreduction reactions involving adsorbed oxygen take place and result in the formation of the radical species [147,148]. Accordingly, in both materials, the holes accumulate in the valence band of g-C3N4 where they directly oxidise the pollutants to form the degradation products.
Generally, the incorporation of Fe3O4 nanoparticles in the photocatalysts can serve as a means to prolong the life of the photogenerated charge carriers by acting as electron traps and induce magnetic response in the nanocomposites, which enables easy separation using an external magnetic field. In both magnetic core-shell and heterojunction nanostructures, the magnetic influence of the Fe3O4 nanoparticles was significant despite showing lower magnetisation saturation compared to neat Fe3O4 due to the presence of the non-magnetic components of the photocatalysts. In the heterojunction nanostructures, Fe3O4 nanoparticles played a dual role of charge separation and magnetic separation, but the detrimental effect of recombination becomes a problem if the optimum Fe3O4 loading is not carefully controlled. Meanwhile, the silica shell in the core-shell nanostructures serve as a barrier to prevent the formation of heterojunctions and charge transfer, limiting the role of the Fe3O4 nanoparticles only to magnetic separation of the photocatalysts. A host of other Fe3O4-based magnetic photocatalyst have been designed, prepared, and exploited in water treatment applications and are summarised in Table 2.

2.4. Ferrites (MFe2O4)-based Magnetic Photocatalysts

The spinel ferrites with the chemical formula MFe2O4, where M is a cation such as Mg2+, Cu2+, Zn2+, Ni2+, Co2+, etc., are popular materials due to their narrow band gaps, which endow them with remarkable visible light response. Moreover, ferrites show good chemical, thermal, and ferromagnetic properties, numerous oxygen vacancies, and a high density of surface hydroxyl groups [166,167]. In terms of the crystal structure of ferrites, the oxygen atoms form a face-centred cubic (fcc) close packing, while the M2+ and Fe2+ cations occupy either octahedral or tetrahedral interstitial sites [168]. Generally, ferrites have attracted tremendous interests in various applications such as fabrication of memory devices, biotechnology, catalysts for organic reactions, water splitting, pollution remediation, lithium ion batteries, etc. [169,170,171,172]. Among the ferrites, zinc ferrite (ZnFe2O4) is one of the most attractive ferrite due to its photochemical and thermal stability, low toxicity, narrow band gap (1.9 eV), natural abundance, good magnetic properties, and easy preparation [64,173,174]. However, as a pristine photocatalyst, ZnFe2O4 show low photocatalytic activity (low quantum efficiency) due to the high recombination rate of the photogenerated charge carriers, unconvincing photoelectric conversion, and low valence band potential [64].
In an attempt to improve the photocatalytic performance of ZnFe2O4, laser ablation obtained ZnOx(OH)x and FeOx colloids were used as precursors to uniformly deposit ZnFe2O4 nanoparticles on RGO sheets to form a magnetic RGO/ZnFe2O4 nanocomposite. The RGO/ZnFe2O4 nanocomposite exhibited improved charge separation and photocatalytic activity towards MB degradation under visible light irradiation [175]. In a similar work by Yang et al., ZnFe2O4 nanocrystals were confined within an interconnected graphene network, which served as dispersing agents, as transport channels for electrons, and as electron scavengers and reservoirs to minimise recombination. Enhanced visible light photocatalytic degradation of MB was observed over the nanocomposite photocatalyst [176]. In the RGO/ZnFe2O4 nanocomposite, electrons are excited by visible light from the valence band to the conduction band of ZnFe2O4, and these electrons quickly transfer to the RGO skeleton where they are trapped by adsorbed oxygen to form superoxide radicals (Figure 6a). The holes are confined to the valence band of ZnFe2O4 where they directly oxidise the pollutants since they are not positive enough to react with water to form the hydroxyl radicals [175,176]. More interestingly, the RGO/ZnFe2O4 nanocomposites showed good stability and recyclability and were easily separable using an external magnetic field owing to the good magnetic response (Figure 6b). The presence of RGO ensured that the electrons and holes are located at different parts of the nanocomposite, thereby ensuring efficient charge separation.
Enhanced visible light photocatalytic degradation of 17α-ethinylestradiol (EE 2) was observed over magnetic ZnFe2O4–Ag/RGO nanocomposite prepared via a hydrothermal method. In terms of degradation kinetics, the nanocomposite photocatalyst showed the highest activity, which was 14.6 and 5.6 times that of ZnFe2O4 and ZnFe2O4/RGO, respectively. This could be credited to the combined contribution of RGO, Ag, and ZnFe2O4 in the nanocomposite, resulting in an improved surface area, efficient charge separation and transportation, minimal aggregation, and improved visible light utilization [177]. Chen and coworkers used a combination of solvothermal and in situ precipitation routes to tailor a magnetic RGO/ZnFe2O4/Ag3PO4 nanocomposite and examined its photocatalytic behaviour towards 2,4-DCP under visible light exposure. The ternary nanocomposite exhibited significantly superior activity over pure Ag3PO4 [178]. In the ZnFe2O4–Ag/RGO heterostructure, electrons excited by visible light from the valence band of ZnFe2O4 to its conduction band transfer to the Ag nanoparticles and, further, to the RGO skeleton where they are captured by adsorbed oxygen to form the radical species (Figure 6c). The holes directly attack the pollutant to form the degradation products [177]. Similarly, in the RGO/ZnFe2O4/Ag3PO4 ternary heterostructure, a heterojunction type of charge transfer mechanism was the most plausible route (Figure 6d). Both ZnFe2O4 (band gap, 1.92 eV) and Ag3PO4 (band gap, 2.42 eV) are excited by visible light, and electrons are promoted to the conduction bands leaving positive holes in the valence bands. The conduction band potential (−0.39 eV) and valence band potential (+1.49 eV) of ZnFe2O4 are more negative than those of Ag3PO4; therefore, electrons transfer to the conduction band of Ag3PO4 and RGO, while the holes transfer in the opposite direction to the valence band of ZnFe2O4 [178]. This results in efficient charge separation and the formation of the reactive species.
Several other interesting ZnFe2O4-based magnetic nanocomposite photocatalysts such as ZnFe2O4/ZnO/Ag3PO4 [64], ZnFe2O4/ZnO [179], N-TiO2/ZnFe2O4 [180], ZnFe2O4/Ag3PO4 [38], and Ag/ZnO/ZnFe2O4 [181] have recently been fabricated, and their photocatalytic properties were examined for the degradation of various pollutants in water. Notably, coupling the various semiconductors with ZnFe2O4 not only improved the overall visible response and photocatalytic properties of the photocatalysts but also introduced magnetic behaviour, which then allowed for easy separation using an external magnetic field.
Nickel ferrite (NiFe2O4) is another commonly exploited ferrite with a narrow band gap (1.7 eV), good magnetic properties, and high chemical and thermal stability. Structurally, NiFe2O4 has a typical inverse spinel structure and is ferromagnetic in nature. Its magnetism stems from the magnetic moment of antiparallel spins between Fe3+ cations occupying tetrahedral sites and Ni2+ cations sitting at octahedral sites [172,182,183]. Pure NiFe2O4 exhibits very low photocatalytic activity due to a high recombination rate of the charge carriers and the aggregation of the nanoparticles. However, NiFe2O4 is an attractive semiconductor to couple with other semiconductors or carbon nanomaterials and form nanocomposite photocatalysts with improved activity and, most importantly, could easily be separated using a magnet [167,184,185]. For example, magnetically separable AgBr/NiFe2O4 nanocomposite photocatalyst prepared via a combination of hydrothermal and an ultrasound-assisted precipitation method showed improved photocatalytic performance towards RhB under visible light illumination [186]. Ji and coworkers observed a significant improvement in the photocatalytic activity of magnetic g-C3N4/NiFe2O4 towards MB degradation, reaching 87% in 4 h of visible light irradiation [172]. In another example, magnetic Ag3PO4/NiFe2O4 nanocomposite tailored using an in situ precipitation method exhibited better visible light-driven photocatalytic activity towards MB compared to Ag3PO4 and NiFe3O4, individually [184]. NiFe2O4-based magnetic nanocomposites incorporating carbon nanomaterials such as Pd-NiFe2O4/RGO [27] and NiFe2O4/MWCNTs [187] have also shown remarkable enhancement in photocatalytic activity owing to the synergy between the carbon nanomaterials and NiFe2O4, which resulted in improved optical properties, charge separation, and transportation and specific surface area [27,187].
In all the NiFe2O4-incorporating photocatalysts, NiFe2O4 is excited upon the absorption of visible light. The charge transfer process depends on the other semiconductors involved in the nanocomposites. For example in AgBr/NiFe2O4 and Ag3PO4/NiFe2O4, in addition to NiFe2O4, AgBr and Ag3PO4 are also excited by visible light, and electrons are promoted to their conduction bands, leaving holes in the valence bands. The electrons transfer from the conduction band of NiFe2O4, since it is more negative, to the conduction bands of AgBr and Ag3PO4. Accordingly, the holes transfer from the valence bands of AgBr and Ag3PO4 to the valence band of NiFe2O4 where they are involved in the degradation of the pollutants [184,186]. Meanwhile, in g-C3N4/NiFe2O4 (both semiconductors are visible active), the electrons transfer from the conduction band of g-C3N4 to the conduction band of NiFe2O4, and the holes transfer in the opposite direction to the valence band of g-C3N4 (Figure 7a) [172]. Charge separation efficiency is greatly improved upon coupling NiFe2O4 with the other semiconductors since the holes and electrons are located at different sites in the nanocomposites. Moreover, the presence of NiFe2O4 nanoparticles in the nanocomposites endowed them with good magnetic response, which allowed for easy magnetic separation.
Cobalt ferrite (CoFe2O4) is another important magnetic material with several attractive properties such as a crystallisation temperature similar to that of TiO2, remarkable electromagnetic behaviour, high cubic magnetocrystalline anisotropy, good chemical stability, and high mechanical hardness [188,189,190,191]. Coupling CoFe2O4 with other semiconductors, carbon nanomaterials and conducting polymers are some of the strategies explored in order to benefit from its optical and magnetic properties. Pristine CoFe2O4 shows poor photocatalytic efficiency due to the high recombination rate of the charge carriers [61,192,193]. Kim et al. prepared hollow cobalt ferrite-polyaniline (CoFe2O4-PANI) nanofibres using the electrospinning method, calcination, and in situ chemical oxidative polymerisation, in succession. It was observed that the nanocomposite was magnetically separable and also showed significantly higher visible light photocatalytic degradation of MO, which was 80 times the activity of neat CoFe2O4 [61]. Hydrothermally prepared magnetic CoFe2O4/graphene (GR) nanocomposites with varying GR loadings exhibited improved visible light photocatalytic properties towards MB degradation compared to pristine CoFe2O4. At an optimum GR loading of 10%, the nanocomposite reached 100% MB degradation in 60 min, while the neat CoFe2O4 could only manage a low 34% degradation in twice the irradiation time [192]. The enhanced activity could be ascribed to the formation of heterojunctions between CoFe2O4 and GR, which allowed for the efficient separation of the charge carriers. Upon excitation by visible light, the electrons in the conduction band of CoFe2O4 transfer to GR, where they are trapped by molecular oxygen adsorbed on the surface [192]. A similar scenario was highlighted as responsible for the improved activity of CoFe2O4-PANI, whereby the electrons are transferred to the conduction band of CoFe2O4 from PANI while the holes transfer in the opposite direction to PANI [61]. The electrons and holes are kept apart, which ensures their availability to form the active species responsible for pollutant degradation.
Enhanced visible light photocatalytic decomposition of MB was reported over CoFe2O4 decorated CdS nanorods (CoFe2O4/CdS, prepared via a soft chemical method. The composite nanorods exhibited better activity than CdS and CoFe2O4, individually, and were magnetically separable [193]. Three-dimensional (3D) urchin-like TiO2, prepared hydrothermally, was decorated with CoFe2O4 using a co-precipitation strategy to yield the nanocomposite photocatalyst CoFe2O4/TiO2 with improved photocatalytic properties compared to the individual materials. Upon UV light illumination, the nanocomposite reached 98.9% MB degradation, while 79.9% MB decomposition was realised when the urchin-like TiO2 was used as the photocatalyst [194]. Moreover, the nanocomposite photocatalyst showed good stability and recyclability, reaching 93.8% MB degradation in the fifth cycles while still maintaining good magnetic properties [194]. Impressive photocatalytic activity and versatility was observed for CoFe2O4/BiOX (X = Cl, Br and I) nanocomposites towards the UV and visible light photocatalytic decomposition of MO, RhB, MB, and their mixture (MO + MB + RhB) [195]. Similarly, Xu and coworkers examined the photocatalytic properties and versatile nature of the Ag/AgCl/CoFe2O4 nanocomposite by employing the photocatalyst for the decomposition of MO, bisphenol A (BPA), ciprofloxacin (CIP), and the deactivation of E. coli. The nanocomposite photocatalyst displayed outstanding photocatalytic performance towards both coloured (MO) and colourless (BPA and CIP) pollutants as well as microbial pollutants, reaching 93.4% in 150 min for MO and 100% in 90 min for CIP, and almost all of the E. coli were deactivated in just 30 min of visible light illumination [196]. In all the CoFe2O4-incorporating nanocomposites, the enhancement in activity could be ascribed to the formation of a heterojunction between CoFe2O4 and the other semiconductor, which allowed efficient charge separation and formation of the reactive species responsible for pollutant degradation. In addition, the presence of CoFe2O4 endowed the nanocomposite photocatalysts with sufficient magnetic sensitivity for easy separation using an external magnetic field [193,194,195,196]. Similar results have been reported for other CoFe2O4-based magnetic photocatalysts such as CoFe2O4/MCM-41/TiO2 [197], Fe,N-TiO2/CoFe2O4 [198], and BiOBr/CoFe2O4 [65].
A host of other ferrite-based magnetic nanocomposite photocatalysts (Table 3) have been designed and tailored for the removal of various pollutants in water, for which their photocatalytic properties were also examined. Enhanced photocatalytic performance has been observed upon the incorporation of the ferrites. Moreover, the presence of these ferrites induced magnetic response in the photocatalyst, which aids the separation process for easy recovery and recycling of the photocatalysts. Despite the decrease in magnetisation saturation of the nanocomposites compared to the neat ferrites, the photocatalyst retained enough magnetic response to enable separation using an external magnet.

3. Summary Remarks and Future Outlook

The development of efficient visible light photocatalytic degradation of pollutants and easy recovery and recyclability of the photocatalyst is an exciting prospect in water treatment. The incorporation of magnetic nanoparticles such as haematite, maghemite, magnetite, and the ferrites into various photocatalyst matrices provides an attractive strategy to enhance the optical properties, charge separation efficiency, and the overall photocatalytic activity. Most interestingly, the presence of the magnetic nanoparticles endow the nanocomposite photocatalyst with sufficient magnetic response to enable separation of the photocatalyst using an external magnetic field. However, careful control of the magnetic nanoparticle, especially loading maghemite and magnetite, is important in ensuring a balance between good magnetic response and the detrimental effect of recombination that occurs on the surface of the nanoparticles. Alternatively, silica is used as a barrier between the magnetic core and the other semiconductor coupled with maghemite or magnetite. This prevents charge transfer between the coupled materials that could otherwise result in photodissolution and recombination. The incorporation of the ferrites in various photocatalysts provides more beneficial properties than just good magnetic response; they contribute towards visible light absorption, charge separation, and the photocatalytic performance of the nanocomposite photocatalyst. Therefore, the ferrites are the most attractive nanoparticles compared to haematite, maghemite, and magnetite, in terms of their photocatalytic performance and multifaceted contribution in the magnetic nanocomposites. However, a systematic, comparative study of the photocatalytic performance of the ferrites, haematite, magnetite, and maghemite is seldom reported, and this makes it difficult to map a way forward in terms of exploiting the best materials for photocatalytic applications.
Despite the promising results regarding visible-light-active magnetic photocatalysts, there are still some challenges that possibly hinder the practical exploitation of these materials. The most obvious problem relates to the reaction kinetics; most of the degradation experiments are slow (may take several hours) despite the improvements in visible light absorption of the photocatalyst, and this presents a massive challenge for practical applications. Secondly, work still needs to be done to develop synthesis routes that will ensure uniform shape and size of the magnetic particles as well as their uniform distribution within the nanocomposite matrix in order to ensure good magnetic response and efficient recovery. Strong contact between the magnetic nanoparticles and the other semiconductor particles is crucial in ensuring minimal leaching out and loss of the photocatalytic activity and magnetic response. In the magnetic core-shell structures, careful control of the thickness of the silica shell is important, as this will affect the magnetic response of the composite material. Extensive theoretical studies are needed in order to understand the interactions between the magnetic nanoparticles and the other semiconductor nanoparticles in the nanocomposites; this can help design and tailor materials with good heterojunctions that promote efficient charge separation. Moreover, theoretical studies can provide a good idea of the optimum loadings of the magnetic nanoparticles in the photocatalyst to ensure good magnetic response and minimal recombination. In most cases, the magnetisation saturation of the magnetic nanoparticles decrease significantly upon coupling with the non-magnetic components of the nanocomposite photocatalyst, which has a negative bearing on the separation efficiency and recovery of the photocatalyst. Work still needs to be done to ensure that nanocomposites retain sufficiently high magnetic response in order to enable efficient magnetic separation.

Acknowledgments

Financial support from the College of Science, Engineering and Technology (CSET) of the University of South Africa is highly appreciated.

Author Contributions

Gcina Mamba conceived the idea of writing a review on magnetically separable photocatalysts and wrote the introduction and Section 2.1, Section 2.2, and Section 2.3. Ajay Mishra is the project leader and was responsible for structuring the review, writing Section 2.4, and editing the entire review article.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Fersi, C.; Gzara, L.; Dhahbi, M. Treatment of textile effluents by membrane technologies. Desalination 2005, 185, 399–409. [Google Scholar] [CrossRef]
  2. Pendergast, M.M.; Hoek, E.M.V. A review of water treatment membrane nanotechnologies. Energy Environ. Sci. 2011, 4, 1946–1971. [Google Scholar] [CrossRef]
  3. Wang, H.; Zheng, X.W.; Su, J.Q.; Tian, Y.; Xiong, X.J.; Zheng, T.L. Biological decolorization of the reactive dyes Reactive Black 5 by a novel isolated bacterial strain Enterobacter sp. EC3. J. Hazard. Mater. 2009, 171, 654–659. [Google Scholar] [CrossRef] [PubMed]
  4. Kumari, K.; Abraham, T.E. Biosorption of anionic textile dyes by nonviable biomass of fungi and yeast. Bioresour. Technol. 2007, 98, 1704–1710. [Google Scholar] [CrossRef] [PubMed]
  5. Soutsas, K.; Karayannis, V.; Poulios, I.; Riga, A.; Ntampegliotis, K.; Spiliotis, X.; Papapolymerou, G. Decolorization and degradation of reactive azo dyes via heterogeneous photocatalytic processes. Desalination 2010, 250, 345–350. [Google Scholar] [CrossRef]
  6. Wu, J.; Doan, H.; Upreti, S. Decolorization of aqueous textile reactive dye by ozone. Chem. Eng. J. 2008, 142, 156–160. [Google Scholar] [CrossRef]
  7. Harsha, N.; Krishna, K.V.S.; Renuka, N.K.; Shukla, S. Facile synthesis of g–Fe2O3 nanoparticles integrated H2Ti3O7 nanotubes structure as a magnetically recyclable dye-removal catalyst. RSC Adv. 2015, 5, 30354–30362. [Google Scholar] [CrossRef]
  8. Li, K.; Chai, F.; Zhao, Y.; Guo, X. Facile synthesis of magnetic Fe3O4/CeCO3OH composites with excellent adsorption capability for small cationic dyes. RSC Adv. 2015, 5, 94397–94404. [Google Scholar] [CrossRef]
  9. Song, S.; Fan, J.; He, Z.; Zhan, L.; Liu, Z.; Chen, J.; Xu, X. Electrochemical degradation of azo dye C.I. Reactive Red 195 by anodic oxidation on Ti/SnO2-Sb/PbO2 electrodes. Electrochim. Acta 2010, 55, 3606–3613. [Google Scholar] [CrossRef]
  10. Faouzi, A.M.; Nasr, B.; Abdellatif, G. Electrochemical degradation of anthraquinone dye Alizarin Red S by anodic oxidation on boron-doped diamond. Dyes Pigments 2007, 73, 86–89. [Google Scholar] [CrossRef]
  11. Rodrigues, C.S.D.; Madeira, L.M.; Boaventura, R.A.R. Synthetic textile dyeing wastewater treatment by integration of advanced oxidation and biological processes—Performance analysis with costs reduction. J. Environ. Chem. Eng. 2014, 2, 1027–1039. [Google Scholar] [CrossRef]
  12. Basha, C.A.; Selvakumar, K.V.; Prabhu, H.J.; Sivashanmugam, P.; Lee, C.W. Degradation studies for textile reactive dye by combined electrochemical, microbial and photocatalytic methods. Sep. Purif. Technol. 2011, 79, 303–309. [Google Scholar] [CrossRef]
  13. Berberidou, C.; Avlonitis, S.; Poulios, I. Dyestuff effluent treatment by integrated sequential photocatalytic oxidation and membrane filtration. Desalination 2009, 249, 1099–1106. [Google Scholar] [CrossRef]
  14. Khouni, I.; Marrot, B.; Moulin, P.; Ben, R. Decolourization of the reconstituted textile effluent by different process treatments: Enzymatic catalysis, coagulation/flocculation and nanofiltration processes. Desalination 2011, 268, 27–37. [Google Scholar] [CrossRef]
  15. Trzcinski, A.P.; Stuckey, D.C. Inorganic fouling of an anaerobic membrane bioreactor treating leachate from the organic fraction of municipal solid waste (OFMSW) and a polishing aerobic membrane bioreactor. Bioresour. Technol. 2016, 204, 17–25. [Google Scholar] [CrossRef] [PubMed]
  16. Schulz, M.; Soltani, A.; Zheng, X.; Ernst, M. Effect of inorganic colloidal water constituents on combined low-pressure membrane fouling with natural organic matter (NOM). J. Membr. Sci. 2016, 507, 154–164. [Google Scholar] [CrossRef]
  17. Hegab, H.M.; Elmekawy, A.; Barclay, T.G.; Michelmore, A.; Zou, L.; Saint, C.P.; Ginic-Markovic, M. Effective in-situ chemical surface modification of forward osmosis membranes with polydopamine-induced graphene oxide for biofouling mitigation. Desalination 2016, 385, 126–137. [Google Scholar] [CrossRef]
  18. Gottlieb, A.; Shaw, C.; Smith, A.; Wheatley, A.; Forsythe, S. The toxicity of textile reactive azo dyes after hydrolysis and decolourisation. J. Biotechnol. 2003, 101, 49–56. [Google Scholar] [CrossRef]
  19. Kariyajjanavar, P.; Jogttappa, N.; Nayaka, Y.A. Studies on degradation of reactive textile dyes solution by electrochemical method. J. Hazard. Mater. 2011, 190, 952–961. [Google Scholar] [CrossRef] [PubMed]
  20. Faouzi Elahmadi, M.; Bensalah, N.; Gadri, A. Treatment of aqueous wastes contaminated with Congo Red dye by electrochemical oxidation and ozonation processes. J. Hazard. Mater. 2009, 168, 1163–1169. [Google Scholar] [CrossRef] [PubMed]
  21. Mohammad, N.; Arami, M. Degradation and toxicity reduction of textile wastewater using immobilized titania nanophotocatalysis. J. Photochem. Photobiol. B 2009, 94, 20–24. [Google Scholar]
  22. Aksu, Z.; Dönmez, G. Combined effects of molasses sucrose and reactive dye on the growth and dye bioaccumulation properties of Candida tropicalis. Process Biochem. 2005, 40, 2443–2454. [Google Scholar] [CrossRef]
  23. Yu, W.; Xu, D.; Peng, T. Enhanced photocatalytic activity of g-C3N4 for selective CO2 reduction to CH3OH via facile coupling of ZnO: A direct Z-scheme mechanism. J. Mater. Chem. A 2015, 3, 19936–19947. [Google Scholar] [CrossRef]
  24. Lu, D.; Zhang, G.; Wan, Z. Visible-light-driven g-C3N4/Ti3+-TiO2 photocatalyst co-exposed {001} and {101} facets and its enhanced photocatalytic activities for organic pollutant degradation and Cr(VI) reduction. Appl. Surf. Sci. 2015, 358, 223–230. [Google Scholar] [CrossRef]
  25. Li, G.; Nie, X.; Chen, J.; Jiang, Q.; An, T.; Wong, P.K.; Zhang, H.; Zhao, H.; Yamashita, H. Enhanced visible-light-driven photocatalytic inactivation of Escherichia coli using g-C3N4/TiO2 hybrid photocatalyst synthesized using a hydrothermal-calcination approach. Water Res. 2015, 86, 17–24. [Google Scholar] [CrossRef] [PubMed]
  26. Zhong, Y.; Ni, Y.; Li, S.; Wang, M. Chain-like Fe3O4@resorcinol-formaldehyde resins-Ag composite microstructures: Facile construction and applications in antibacterial and catalytic fields. RSC Adv. 2016, 6, 15831–15837. [Google Scholar] [CrossRef]
  27. Li, Y.; Zhang, Z.; Pei, L.; Li, X.; Fan, T.; Ji, J.; Shen, J.; Ye, M. Multifunctional photocatalytic performances of recyclable Pd-NiFe2O4/reduced graphene oxide nanocomposites via different co-catalyst strategy. Appl. Catal. B Environ. 2016, 190, 1–11. [Google Scholar] [CrossRef]
  28. Zeng, B.; Zhang, L.; Wan, X.; Song, H.; Lv, Y. Fabrication of α-Fe2O3/g-C3N4 composites for cataluminescence sensing of H2S. Sens. Actuators Chem. 2015, 211, 370–376. [Google Scholar] [CrossRef]
  29. Wang, G.; Kuang, S.; Zhang, J.; Hou, S.; Nian, S. Graphitic carbon nitride/multiwalled carbon nanotubes composite as Pt-free counter electrode for high-efficiency dye-sensitized solar cells. Electrochim. Acta 2016, 187, 243–248. [Google Scholar] [CrossRef]
  30. Wang, M.; Yang, X.; Bi, W. Application of magnetic graphitic carbon nitride nanocomposites for the solid-phase extraction of phthalate esters in water samples. J. Sep. Sci. 2015, 38, 445–452. [Google Scholar] [CrossRef] [PubMed]
  31. Baig, R.B.N.; Verma, S.; Varma, R.S.; Nadagouda, M.N. Magnetic Fe@g-C3N4: A photoactive catalyst for the hydrogenation of alkenes and alkynes. ACS Sustain. Chem. Eng. 2016, 4, 1661–1664. [Google Scholar] [CrossRef]
  32. Preethi, V.; Kanmani, S. Photocatalytic hydrogen production using Fe2O3-based core shell nano particles with ZnS and CdS. Int. J. Hydrogen Energy 2014, 39, 1613–1622. [Google Scholar] [CrossRef]
  33. Cao, M.; Wang, P.; Ao, Y.; Wang, C.; Hou, J.; Qian, J. Visible light activated photocatalytic degradation of tetracycline by a magnetically separable composite photocatalyst: Graphene oxide/magnetite/cerium-doped titania. J. Colloid Interface Sci. 2016, 467, 129–139. [Google Scholar] [CrossRef] [PubMed]
  34. Bao, X.; Qiang, Z.; Chang, J.; Ben, W.; Qu, J. Synthesis of carbon-coated magnetic nanocomposite (Fe3O4@C) and its application for sulfonamide antibiotics removal from water. J. Environ. Sci. 2014, 26, 962–969. [Google Scholar] [CrossRef]
  35. Zhou, L.; Deng, J.; Zhao, Y.; Liu, W.; An, L.; Chen, F. Preparation and characterization of N–I co-doped nanocrystal anatase TiO2 with enhanced photocatalytic activity under visible-light irradiation. Mater. Chem. Phys. 2009, 117, 522–527. [Google Scholar] [CrossRef]
  36. Nolan, N.T.; Synnott, D.W.; Seery, M.K.; Hinder, S.J.; van Wassenhoven, A.; Pillai, S.C. Effect of N-doping on the photocatalytic activity of sol-gel TiO2. J. Hazard. Mater. 2012, 211–212, 88–94. [Google Scholar] [CrossRef] [PubMed]
  37. Pany, S.; Parida, K.M. A facile in situ approach to fabricate N,S-TiO2/g-C3N4 nanocomposite with excellent activity for visible light induced water splitting for hydrogen evolution. Phys. Chem. Chem. Phys. 2015, 17, 8070–8077. [Google Scholar] [CrossRef] [PubMed]
  38. Ge, M.; Chen, Y.; Liu, M.; Li, M. Synthesis of magnetically separable Ag3PO4/ZnFe2O4 composite photocatalysts for dye degradation under visible LED light irradiation. J. Environ. Chem. Eng. 2015, 3, 2809–2815. [Google Scholar] [CrossRef]
  39. Xu, Y.; Huang, S.; Xie, M.; Li, Y.; Jing, L.; Xu, H. Core-shell magnetic Ag/AgCl@Fe2O3 photocatalysts with enhanced photoactivity for eliminating bisphenol A and microbial contamination. New J. Chem. 2016, 40, 3413–3422. [Google Scholar] [CrossRef]
  40. Yao, Y.; Lu, F.; Zhu, Y.; Wei, F.; Liu, X.; Lian, C.; Wang, S. Magnetic core-shell CuFe2O4@C3N4 hybrids for visible light photocatalysis of Orange II. J. Hazard. Mater. 2015, 297, 224–233. [Google Scholar] [CrossRef] [PubMed]
  41. Lin, H.-P.; Chen, C.-C.; Lee, W.W.; Lai, Y.-Y.; Chen, J.; Chen, Y.; Fu, J.-Y. Synthesis of a SrFeO3−x/g-C3N4 heterojunction with improved visible-light photocatalytic activities in chloramphenicol and crystal violet degradation. RSC Adv. 2015, 6, 2323–2336. [Google Scholar] [CrossRef]
  42. Li, G.; Nie, X.; Gao, Y.; An, T. Can environmental pharmaceuticals be photocatalytically degraded and completely mineralized in water using g-C3N4/TiO2 under visible light irradiation?—Implications of persistent toxic intermediates. Appl. Catal. B Environ. 2016, 180, 726–732. [Google Scholar] [CrossRef]
  43. Zhang, J.; Hu, S.; Wang, Y. A convenient method to prepare a novel alkali metal sodium doped carbon nitride photocatalyst with a tunable band structure. RSC Adv. 2014, 4, 62912–62919. [Google Scholar] [CrossRef]
  44. Jin, R.; Hu, S.; Gui, J.; Liu, D. A convenient method to prepare novel rare earth metal Ce-doped carbon nitride with enhanced photocatalytic activity under visible light. Bull. Korean Chem. Soc. 2015, 36, 17–23. [Google Scholar] [CrossRef]
  45. Akpan, U.G.; Hameed, B.H. Enhancement of the photocatalytic activity of TiO2 by doping it with calcium ions. J. Colloid Interface Sci. 2011, 357, 168–178. [Google Scholar] [CrossRef] [PubMed]
  46. Inturi, S.N.R.; Boningari, T.; Suidan, M.; Smirniotis, P.G. Visible-light-induced photodegradation of gas phase acetonitrile using aerosol-made transition metal (V, Cr, Fe, Co, Mn, Mo, Ni, Cu, Y, Ce, and Zr) doped TiO2. Appl. Catal. B Environ. 2014, 144, 333–342. [Google Scholar] [CrossRef]
  47. Li, Z.; Wang, J.; Zhu, K.; Ma, F.; Meng, A. Ag/g-C3N4 composite nanosheets: Synthesis and enhanced visible photocatalytic activities. Mater. Lett. 2015, 145, 167–170. [Google Scholar] [CrossRef]
  48. Hu, S.; Ma, L.; You, J.; Li, F.; Fan, Z.; Wang, F.; Liu, D.; Gui, J. A simple and efficient method to prepare a phosphorus modified g-C3N4 visible light photocatalyst. RSC Adv. 2014, 4, 21657–21663. [Google Scholar] [CrossRef]
  49. Sun, H.; Wang, S.; Ang, H.M.; Tadé, M.O.; Li, Q. Halogen element modified titanium dioxide for visible light photocatalysis. Chem. Eng. J. 2010, 162, 437–447. [Google Scholar] [CrossRef]
  50. Dong, G.; Ai, Z.; Zhang, L. Efficient anoxic pollutant removal with oxygen functionalized graphitic carbon nitride under visible light. RSC Adv. 2014, 4, 5553–5560. [Google Scholar] [CrossRef]
  51. Sharotri, N.; Sud, D. Ultrasound-assisted synthesis and characterization of visible light responsive nitrogen-doped TiO2 nanomaterials for removal of 2-Chlorophenol. Desalin. Water Treat. 2016, 57, 8776–8788. [Google Scholar] [CrossRef]
  52. Reszczyńska, J.; Grzyb, T.; Sobczak, J.W.; Lisowski, W.; Gazda, M.; Ohtani, B.; Zaleska, A. Visible light activity of rare earth metal doped (Er3+, Yb3+ or Er3+/Yb3+) titania photocatalysts. Appl. Catal. B: Environ. 2015, 163, 40–49. [Google Scholar] [CrossRef]
  53. Kaur, N.; Kaur, S.; Singh, V. Preparation, characterization and photocatalytic degradation kinetics of Reactive Red dye 198 using N, Fe codoped TiO2 nanoparticles under visible light. Desalination Water Treat. 2016, 57, 9237–9246. [Google Scholar] [CrossRef]
  54. Xu, X.; Zhou, X.; Zhang, L.; Xu, L.; Ma, L.; Luo, J.; Li, M.; Zeng, L. Photoredox degradation of different water pollutants (MO, RhB, MB, and Cr(VI)) using Fe-N-S-tri-doped TiO2 nanophotocatalyst prepared by novel chemical method. Mater. Res. Bull. 2015, 70, 106–113. [Google Scholar] [CrossRef]
  55. Antonopoulou, M.; Vlastos, D.; Konstantinou, I. Photocatalytic degradation of pentachlorophenol by N-F-TiO2: Identification of intermediates, mechanism involved, genotoxicity and ecotoxicity evaluation. Photochem. Photobiol. Sci. 2015, 14, 520–527. [Google Scholar] [CrossRef] [PubMed]
  56. Mamba, G.; Mbianda, X.Y.; Mishra, A.K. Photocatalytic degradation of the diazo dye naphthol blue black in water using MWCNT/Gd,N,S-TiO2 nanocomposites under simulated solar light. J. Environ. Sci. 2015, 33, 219–228. [Google Scholar] [CrossRef] [PubMed]
  57. Mamba, G.; Mamo, M.A.; Mbianda, X.Y.; Mishra, A.K. Nd,N,S-TiO2 decorated on reduced graphene oxide for a visible light active photocatalyst for dye degradation: Comparison to its MWCNT/Nd,N,S-TiO2 analogue. Ind. Eng. Chem. Res. 2014, 53, 14329–14338. [Google Scholar] [CrossRef]
  58. Mamba, G.; Mbianda, X.Y.; Mishra, A.K. Enhanced visible light photocatalytic degradation of eriochrome black T and eosin blue shade in water using tridoped titania decorated on SWCNTs and MWCNTs: Effect of the type of carbon nanotube incorporated. Mater. Chem. Phys. 2015, 149–150, 734–742. [Google Scholar] [CrossRef]
  59. Apostolopoulou, V.; Vakros, J.; Kordulis, C.; Lycourghiotis, A. Preparation and characterization of [60] fullerene nanoparticles supported on titania used as a photocatalyst. Colloids Surf. A Physicochem. Eng. Asp. 2009, 349, 189–194. [Google Scholar] [CrossRef]
  60. Shang, J.; Zhao, F.; Zhu, T.; Li, J. Photocatalytic degradation of rhodamine B by dye-sensitized TiO2 under visible-light irradiation. Sci. China Chem. 2011, 54, 167–172. [Google Scholar] [CrossRef]
  61. Nam, K.; Jung, H.; Lee, W. Hollow cobalt ferrite-polyaniline nanofibers as magnetically separable visible-light photocatalyst for photodegradation of methyl orange. J. Photochem. Photobiol. A Chem. 2016, 321, 257–265. [Google Scholar]
  62. Kim, G.; Choi, W. Charge-transfer surface complex of EDTA-TiO2 and its effect on photocatalysis under visible light. Appl. Catal. B Environ. 2010, 100, 77–83. [Google Scholar] [CrossRef]
  63. Shi, W.; Lu, D.; Wang, L.; Teng, F.; Zhang, J. Core-shell structured Fe3O4@SiO2@CdS nanoparticles with enhanced visible-light photocatalytic photocatalytic activities. RSC Adv. 2015, 5, 106038–106043. [Google Scholar] [CrossRef]
  64. Li, J.; Liu, Z.; Zhu, Z. Enhanced photocatalytic activity in ZnFe2O4-ZnO-Ag3PO4 hollow nanospheres through the cascadal electron transfer with magnetical separation. J. Alloys Compd. 2015, 636, 229–233. [Google Scholar] [CrossRef]
  65. Jiang, R.; Zhu, H.; Li, J.; Fu, F.; Yao, J.; Jiang, S.; Zeng, G.-M. Fabrication of novel magnetically separable BiOBr/CoFe2O4 microspheres and its application in the efficient removal of dye from aqueous phase by an environment-friendly and economical approach. Appl. Surf. Sci. 2016, 364, 604–612. [Google Scholar] [CrossRef]
  66. Mi, J.; Hwan, Y.; Keun, B.; Ho, S.; Soo, B.; Bin, J. Novel immobilization of titanium dioxide (TiO2) on the fluidizing carrier and its application to the degradation of azo-dye. J. Hazard. Mater. 2006, 134, 230–236. [Google Scholar]
  67. Fernhndez, A.; Lassaletta, G.; Jimknez, V.M.; Justo, A. Preparation and characterization of TiO2 photocatalysts supported on various rigid supports (glass, quartz and stainless steel). Comparative studies of photocatalytic activity in water purification. Appl. Catal. B Environ. 1995, 7, 49–63. [Google Scholar] [CrossRef]
  68. Carlson, P.J.; Pretzer, L.A.; Boyd, J.E. Solvent deposition of titanium dioxide on acrylic for photocatalytic application. Ind. Eng. Chem. Res. 2007, 46, 7970–7976. [Google Scholar] [CrossRef]
  69. Arabatzis, I.M.; Antonaraki, S.; Stergiopoulos, T.; Hiskia, A. Preparation, characterization and photocatalytic activity of nanocrystalline thin film TiO2 catalysts towards 3,5-dichlorophenol degradation. J. Photochem. Photobiol. A Chem. 2002, 149, 237–245. [Google Scholar] [CrossRef]
  70. Anpo, M.; Zhang, S.G.; Mishima, H.; Matsuoka, M.; Yamashita, H. Design of photocatalysts encapsulated within the zeolite framework and cavities for the decomposition of NO into N2 and O2 at normal temperature. Catal. Today 1997, 39, 159–168. [Google Scholar] [CrossRef]
  71. Rao, N.N.; Chaturvedi, V. Photoactivity of TiO2-coated pebbles. Ind. Eng. Chem. Res. 2007, 46, 4406–4414. [Google Scholar] [CrossRef]
  72. Gad-allah, T.A.; Kato, S.; Satokawa, S.; Kojima, T. Treatment of synthetic dyes wastewater utilizing a magnetically separable photocatalyst (TiO2/SiO2/Fe3O4): Parametric and kinetic studies. Desalination 2009, 244, 1–11. [Google Scholar] [CrossRef]
  73. Wang, W.; Yao, J. Synthesis of magnetically separable Sn doped magnetite/silica core-shell structure and photocatalytic property. Mater. Res. Bull. 2010, 45, 710–716. [Google Scholar] [CrossRef]
  74. Belessi, V.; Lambropoulou, D.; Konstantinou, I.; Zboril, R.; Tucek, J.; Jancik, D. Structure and photocatalytic performance of magnetically separable titania photocatalysts for the degradation of propachlor. Appl. Catal. B Environ. 2009, 87, 181–189. [Google Scholar] [CrossRef]
  75. Linley, S.; Leshuk, T.; Gu, F.X. Synthesis of magnetic rattle-type nanostructures for use in water treatment. ACS Appl. Mater. Interfaces 2013, 5, 2540–2548. [Google Scholar] [CrossRef] [PubMed]
  76. Kumar, S.; Kumar, B.; Baruah, A.; Shanker, V. Synthesis of magnetically separable and recyclable g-C3N4–Fe3O4 hybrid nanocomposites with enhanced photocatalytic performance under visible-light irradiation. J. Phys. Chem. C 2013, 117, 26135–26143. [Google Scholar] [CrossRef]
  77. Yang, J.; Chen, H.; Gao, J.; Yan, T.; Zhou, F.; Cui, S.; Bi, W. Synthesis of Fe3O4/g-C3N4 nanocomposites and their application in the photodegradation of 2,4,6-trichlorophenol under visible light. Mater. Lett. 2016, 164, 183–189. [Google Scholar] [CrossRef]
  78. Shih, Y.; Su, C.; Chen, C.; Dong, C. Synthesis of magnetically recoverable ferrite (MFe2O4, M = Co, Ni and Fe)-supported TiO2 photocatalysts for decolorization of methylene blue. Catal. Commun. 2015, 72, 127–132. [Google Scholar] [CrossRef]
  79. Pang, Y.L.; Lim, S.; Ong, H.C.; Chong, W.T. Synthesis, characteristics and sonocatalytic activities of calcined α-Fe2O3 and TiO2 nanotubes/α-Fe2O3 magnetic catalysts in the degradation of Orange G. Ultrason. Sonochem. 2016, 29, 317–327. [Google Scholar]
  80. Zhang, D.; Pu, X.; Du, K.; Moon, Y.; Jeong, J.; Cai, P.; Kim, S., II; Seo, H.J. Combustion synthesis of magnetic Ag/NiFe2O4 composites with enhanced visible-light photocatalytic properties. Sep. Purif. Technol. 2014, 137, 82–85. [Google Scholar] [CrossRef]
  81. Zhao, H.; Zhang, L.; Gu, X.; Li, S.; Li, B.; Wang, H. Fe2O3-AgBr nonwoven cloth with hierarchical nanostructures as efficient and easily recyclable macroscale photocatalysts. RSC Adv. 2015, 5, 10951–10959. [Google Scholar] [CrossRef]
  82. Zhang, D.; Wang, Q.; Wang, L.; Zhang, L. Magnetically separable CdFe2O4/graphene catalyst and its enhanced photocatalytic properties. J. Mater. Chem. A 2015, 3, 3576–3585. [Google Scholar] [CrossRef]
  83. Liu, J.; Qiao, S.Z.; Hu, Q.H.; Qing, G.; Lu, M. Magnetic nanocomposites with mesoporous structures: Synthesis and applications. Small 2011, 7, 425–443. [Google Scholar] [CrossRef] [PubMed]
  84. Polshettiwar, V.; Luque, R.; Fihri, A.; Zhu, H.; Bouhrara, M.; Basset, J. Magnetically recoverable nanocatalysts. Chem. Rev. 2011, 111, 3036–3075. [Google Scholar] [CrossRef] [PubMed]
  85. Zhang, D.; Zhou, C.; Sun, Z.; Wu, L.-Z.; Tung, C.-H.; Zhang, T. Magnetically recyclable nanocatalysts (MRNCs): A versatile integration of high catalytic activity and facile recovery. Nanoscale 2012, 4, 6244–6255. [Google Scholar] [CrossRef] [PubMed]
  86. Wang, D.; Astruc, D. Fast-growing field of magnetically recyclable nanocatalysts. Chem. Rev. 2014, 114, 6949–6985. [Google Scholar] [CrossRef] [PubMed]
  87. Linley, S.; Leshuk, T.; Gu, F.X. Magnetically separable water treatment technologies and their role in future advanced water treatment: A patent review. CLEAN Soil Air Water 2013, 41, 1152–1156. [Google Scholar] [CrossRef]
  88. Yao, H.; Fan, M.; Wang, Y. Magnetic titanium dioxide based nanomaterials: Synthesis, characteristics, and photocatalytic application in pollutant degradation. J. Mater. Chem. A 2015, 3, 17511–17524. [Google Scholar] [CrossRef]
  89. Liu, D.; Li, Z.; Wang, W.; Wang, G.; Liu, D. Hematite doped magnetic TiO2 nanocomposites with improved photocatalytic activity. J. Alloy Compd. 2016, 654, 491–497. [Google Scholar] [CrossRef]
  90. Shi, Y.; Li, H.; Wang, L.; Shen, W.; Chen, H. Novel α-Fe2O3/CdS cornlike nanorods with enhanced photocatalytic performance. ACS Appl. Mater. Interfaces 2012, 4, 4800–4806. [Google Scholar] [CrossRef] [PubMed]
  91. Xu, Y.; Jing, L.; Chen, X.; Ji, H.; Xu, H.; Li, H. Novel visible-light-driven Fe2O3/Ag3VO4 composite with enhanced photocatalytic activity toward organic pollutants degradation. RSC Adv. 2016, 6, 3600–3607. [Google Scholar] [CrossRef]
  92. Xu, Y.; Huang, S.; Xie, M.; Li, Y.; Xu, H.; Huang, L.; Zhang, Q.; Li, H. Magnetically separable Fe2O3/g-C3N4 catalyst with enhanced photocatalytic activity. RSC Adv. 2015, 5, 95727–95735. [Google Scholar] [CrossRef]
  93. Kumar, S.V.; Huang, N.M.; Yusoff, N.; Lim, H.N. High performance magnetically separable graphene/zinc oxide nanocomposite. Mater. Lett. 2013, 93, 411–414. [Google Scholar] [CrossRef]
  94. Wang, H.; Fei, X.; Wang, L.; Li, Y.; Xu, S. Magnetically separable iron oxide nanostructures-TiO2 nanofibers hierarchical heterostructures: Controlled fabrication and photocatalytic activity. New J. Chem. 2011, 35, 1795–1802. [Google Scholar] [CrossRef]
  95. Muthukrishnaraj, A.; Vadivel, S.; Kamalakannan, V.P.; Balasubramanian, N. α-Fe2O3/reduced graphene oxide nanorod as efficient photocatalyst for methylene blue degradation. Mater. Res. Innov. 2016, 19, 258–264. [Google Scholar] [CrossRef]
  96. Wu, W.; Zhang, S.; Xiao, X.; Zhou, J.; Ren, F.; Sun, L.; Jiang, C. Controllable synthesis, magnetic properties, and enhanced photocatalytic activity of spindle-like mesoporous α-Fe2O3/ZnO core-shell heterostructures. ACS Appl. Mater. Interfaces 2012, 4, 3602–3609. [Google Scholar] [CrossRef] [PubMed]
  97. Xue, J.; Ma, S.; Zhou, Y.; Zhang, Z.; Liu, X. Fabrication of porous g-C3N4/Ag/Fe2O3 composites with enhanced visible light photocatalysis performance. RSC Adv. 2015, 5, 58738–58745. [Google Scholar] [CrossRef]
  98. Liu, X.; Jin, A.; Jia, Y.; Jiang, J.; Hu, N.; Chen, X. Facile synthesis and enhanced visible-light photocatalytic activity of graphitic carbon nitride decorated with ultrafine Fe2O3 nanoparticles. RSC Adv. 2015, 5, 92033–92041. [Google Scholar] [CrossRef]
  99. Theerthagiri, J.; Senthil, R.A.; Priya, A.; Madhavan, J.; Michael, R.J.V.; Ashokkumar, M. Photocatalytic and photoelectrochemical studies of visible-light-active α-Fe2O3-g-C3N4 nanocomposites. RSC Adv. 2014, 4, 38222–38229. [Google Scholar] [CrossRef]
  100. Liu, Y.; Yu, Y.X.; Zhang, W. Photoelectrochemical study on charge transfer properties of nanostructured Fe2O3 modified by g-C3N4. Int. J. Hydrogen Energy 2014, 39, 9105–9113. [Google Scholar] [CrossRef]
  101. Peng, L.; Xie, T.; Lu, Y.; Fan, H.; Wang, D. Synthesis, photoelectric properties and photocatalytic activity of the Fe2O3/TiO2 heterogeneous photocatalysts. Phys. Chem. Chem. Phys. 2010, 12, 8033–8041. [Google Scholar] [CrossRef] [PubMed]
  102. Pawar, R.C.; Pyo, Y.; Ahn, S.H.; Lee, C.S. Photoelectrochemical properties and photodegradation of organic pollutants using hematite hybrids modified by gold nanoparticles and graphitic carbon nitride. Appl. Catal. B Environ. 2015, 176–177, 654–666. [Google Scholar] [CrossRef]
  103. Yamauraa, M.; Camilo, R.L.; Sampaio, L.C.; Macedo, M.A.; Nakamurad, M.; Toma, H.E. Preparation and characterization of (3-aminopropyl) triethoxysilane-coated magnetite nanoparticles. J. Magn. Magn. Mater. 2004, 279, 210–217. [Google Scholar] [CrossRef]
  104. Tian, B.; Wang, T.; Dong, R.; Bao, S.; Yang, F.; Zhang, J. Core-shell structured–Fe2O3@SiO2@AgBr: Ag composite with high magnetic separation efficiency and excellent visible light activity for acid orange 7 degradation. Appl. Catal. B Environ. 2014, 147, 22–28. [Google Scholar] [CrossRef]
  105. Idris, A.; Hassan, N.; Rashid, R.; Ngomsik, A. Kinetic and regeneration studies of photocatalytic magnetic separable beads for chromium (VI) reduction under sunlight. J. Hazard. Mater. 2011, 186, 629–635. [Google Scholar] [CrossRef] [PubMed]
  106. Yu, X.; Liu, S.; Yu, J. Superparamagnetic γ-Fe2O3@SiO2@TiO2 composite microspheres with superior photocatalytic properties. Appl. Catal. B Environ. 2011, 104, 12–20. [Google Scholar] [CrossRef]
  107. Lin, Y.; Chang, C. Magnetic mesoporous iron oxide/carbon aerogel photocatalysts with adsorption ability for organic dye removal. RSC Adv. 2014, 4, 28628–28631. [Google Scholar] [CrossRef]
  108. Zhang, Y.; Zhang, Y.; Tan, J. Novel magnetically separable AgCl/iron oxide composites with enhanced photocatalytic activity driven by visible light. J. Alloy. Compd. 2013, 574, 383–390. [Google Scholar] [CrossRef]
  109. Ye, S.; Qiu, L.-G.; Yuan, Y.-P.; Zhu, Y.-J.; Xia, J.; Zhu, J.-F. Facile fabrication of magnetically separable graphitic carbon nitride photocatalysts with enhanced photocatalytic activity under visible light. J. Mater. Chem. A 2013, 1, 2929–3188. [Google Scholar] [CrossRef]
  110. He, M.; Li, D.; Jiang, D.; Chen, M. Magnetically separable γ-Fe2O3@SiO2@Ce-doped TiO2 core-shell nanocomposites: Fabrication and visible-light-driven photocatalytic activity. J. Solid State Chem. 2012, 192, 139–143. [Google Scholar] [CrossRef]
  111. Kaur, N.; Shahi, S.K.; Singh, V. Synthesis, characterization and photocatalytic activity of magnetically separable γ-Fe2O3/N,Fe codoped TiO2 heterojunction for degradation of Reactive Blue 4 dye. RSC Adv. 2015, 5, 61623–61630. [Google Scholar] [CrossRef]
  112. Yu, L.; Yang, X.; Wang, D. TiO2 incorporated in magnetic mesoporous SBA-15 by a facile inner-pore hydrolysis process toward enhanced adsorption-photocatalysis performances for As(III). J. Colloid Interface Sci. 2015, 448, 525–532. [Google Scholar] [CrossRef] [PubMed]
  113. Liu, Y.; Yu, L.; Hu, Y.; Guo, C.; Wen, X.; Lou, D. A magnetically separable photocatalyst based on nest-like γ-Fe2O3/ZnO double-shelled hollow structures with enhanced photocatalytic activity. Nanoscale 2012, 4, 183–187. [Google Scholar] [CrossRef] [PubMed]
  114. Szeto, W.; Li, J.; Huang, H.; Xuan, J.; Leung, D.Y.C. Novel urchin-like Fe2O3@SiO2@TiO2 microparticles with magnetically separable and photocatalytic properties. RSC Adv. 2015, 5, 55363–55371. [Google Scholar] [CrossRef]
  115. Kumar, A.; Guo, C.; Sharma, G.; Pathania, D.; Naushad, M.; Kalia, S.; Dhiman, P. Magnetically recoverable ZrO2/Fe3O4/chitosan nanomaterials for enhanced sunlight driven photoreduction of carcinogenic Cr(VI) and dechlorination and mineralization of 4-chlorophenol from simulated waste water. RSC Adv. 2016, 6, 13251–13263. [Google Scholar] [CrossRef]
  116. Zhao, Y.; Tao, C.; Xiao, G.; Wei, G.; Li, L.; Liu, C.; Su, H. Controlled synthesis and photocatalysis of sea urchin-like Fe3O4@TiO2@Ag nanocomposites. Nanoscale 2016, 8, 5313–5326. [Google Scholar] [CrossRef]
  117. Li, G.; Wong, K.H.; Zhang, X.; Hu, C.; Yu, J.C.; Chan, R.C.Y.; Wong, P.K. Degradation of Acid Orange 7 using magnetic AgBr under visible light: The roles of oxidizing species. Chemosphere 2009, 76, 1185–1191. [Google Scholar] [CrossRef] [PubMed]
  118. Álvarez, P.M.; Jaramillo, J.; López-pi, F.; Plucinski, P.K. Preparation and characterization of magnetic TiO2 nanoparticles and their utilization for the degradation of emerging pollutants in water. Appl. Catal. B Environ. 2010, 100, 338–345. [Google Scholar] [CrossRef]
  119. Zhang, L.; Wang, W.; Sun, S.; Sun, Y.; Gao, E.; Zhang, Z. Elimination of BPA endocrine disruptor by magnetic BiOBr@SiO2@Fe3O4 photocatalyst. Appl. Catal. B Environ. 2014, 148–149, 164–169. [Google Scholar] [CrossRef]
  120. Chen, F.; Yan, F.; Chen, Q.; Wang, Y.; Han, L.; Chen, Z.; Fang, S. Fabrication of Fe3O4@SiO2@TiO2 nanoparticles supported by graphene oxide sheets for the repeated adsorption and photocatalytic degradation of rhodamine B under UV irradiation. Dalton Trans. 2014, 43, 13537–13544. [Google Scholar] [CrossRef] [PubMed]
  121. Wang, H.; Wang, W. Synthesis of Fe3O4/N-TiO2/Ag hollow nanospheres and their application as recyclable photocatalysts. J. Cluster Sci. 2016, 27, 403–415. [Google Scholar] [CrossRef]
  122. Ma, J.; Guo, S.; Guo, X.; Ge, H. A mild synthetic route to Fe3O4@TiO2-Au composites: Preparation, characterization and photocatalytic activity. Appl. Surf. Sci. 2015, 353, 1117–1125. [Google Scholar] [CrossRef]
  123. Li, S.; Liang, W.; Zheng, F.; Zhou, H.; Lin, X. Lysine surface modified Fe3O4@SiO2@TiO2 microspheres-based preconcentration and photocatalysis for in situ selective determination of nanomolar dissolved organic and inorganic phosphorus in seawater. Sens Actuators B Chem. 2016, 224, 48–54. [Google Scholar] [CrossRef]
  124. Hongfei, L.I.U.; Shengfu, J.I.; Yuanyuan, Z.; Ming, L.I. Porous TiO2-coated magnetic core-shell nanocomposites: Preparation and enhanced photocatalytic activity. Chin. J. Chem. Eng. 2013, 21, 569–576. [Google Scholar]
  125. Ma, J.; Guo, S.; Guo, X.; Ge, H.-G. Liquid-phase deposition of TiO2 nanoparticles on core-shell Fe3O4@SiO2 spheres: Preparation, characterization, and photocatalytic activity. J. Nanopart. Res. 2015, 17, 1–11. [Google Scholar] [CrossRef]
  126. Rashid, J.; Barakat, M.A.; Ruzmanova, Y.; Chianese, A. Fe3O4/SiO2/TiO2 nanoparticles for photocatalytic degradation of 2-chlorophenol in simulated wastewater. Environ. Sci. Pollut. Res. 2015, 22, 3149–3157. [Google Scholar] [CrossRef] [PubMed]
  127. Hamzezadeh-Nakhjavani, S.; Tavakoli, O.; Akhlaghi, S.P. Efficient photocatalytic degradation of organic pollutants by magnetically recoverable nitrogen-doped TiO2 nanocomposite photocatalysts under visible light irradiation. Environ. Sci. Pollut. Res. 2015, 22, 18859–18873. [Google Scholar] [CrossRef] [PubMed]
  128. Cao, M.; Wang, P.; Ao, Y.; Wang, C.; Hou, J.; Qian, J. Photocatalytic degradation of tetrabromobisphenol A by a magnetically separable graphene-TiO2 composite photocatalyst: Mechanism and intermediates analysis. Chem. Eng. J. 2015, 264, 113–124. [Google Scholar] [CrossRef]
  129. Li, C.; Tan, J.; Fan, X.; Zhang, B.; Zhang, H.; Zhang, Q. Magnetically separable one dimensional Fe3O4/P(MAA-DVB)/TiO2 nanochains: Preparation, characterization and photocatalytic activity. Ceram. Int. 2015, 41, 3860–3868. [Google Scholar] [CrossRef]
  130. Xiang, Y.; Wang, H.; He, Y.; Song, G. Efficient degradation of methylene blue by magnetically separable Fe3O4/chitosan/TiO2 nanocomposites. Desalin. Water Treat. 2015, 55, 1018–1025. [Google Scholar] [CrossRef]
  131. Su, W.; Zhang, T.; Li, L.; Xing, J.; He, M.; Zhong, Y.; Li, Z. Synthesis of small yolk-shell Fe3O4@TiO2 nanoparticles with controllable thickness as recyclable photocatalysts. RSC Adv. 2014, 4, 8901–8906. [Google Scholar] [CrossRef]
  132. Dong, X.; Shao, Y.; Zhang, X.; Ma, H.; Zhang, X.; Shi, F.; Ma, C.; Xue, M. Synthesis and properties of magnetically separable Fe3O4/TiO2/Bi2O3 photocatalysts. Res. Chem. Intermed. 2014, 40, 2953–2961. [Google Scholar] [CrossRef]
  133. Li, C.; Younesi, R.; Cai, Y.; Zhu, Y.; Ma, M.; Zhu, J. Photocatalytic and antibacterial properties of Au-decorated Fe3O4@mTiO2 core-shell microspheres. Appl. Catal. B Environ. 2014, 156–157, 314–322. [Google Scholar] [CrossRef]
  134. Su, J.; Zhang, Y.; Xu, S.; Wang, S.; Ding, H.; Pan, S. Highly efficient and recyclable triple-shelled Ag@Fe3O4@SiO2@TiO2 photocatalysts for degradation of organic pollutants and reduction of hexavalent chromium ions. Nanoscale 2014, 6, 5181–5192. [Google Scholar] [CrossRef] [PubMed]
  135. Chen, L.; Li, L.; Wang, T.; Zhang, L.; Xing, S. A novel strategy to fabricate multifunctional Fe3O4@C@TiO2 yolk-shell structures as magnetically recyclable photocatalysts. Nanoscale 2014, 6, 6603–6608. [Google Scholar] [CrossRef] [PubMed]
  136. Chi, Y.; Yuan, Q.; Li, Y.; Zhao, L.; Li, N.; Li, X.; Yan, W. Magnetically separable Fe3O4@SiO2@TiO2-Ag microspheres with well-designed nanostructure and enhanced photocatalytic activity. J. Hazard. Mater. 2013, 262, 404–411. [Google Scholar] [CrossRef] [PubMed]
  137. Tang, Y.; Zhang, G.; Liu, C.; Luo, S.; Xu, X.; Chen, L.; Wang, B. Magnetic TiO2-graphene composite as a high-performance and recyclable platform for efficient photocatalytic removal of herbicides from water. J. Hazard. Mater. 2013, 252–253, 115–122. [Google Scholar] [CrossRef] [PubMed]
  138. Shojaei, F.A.; Shams-Nateri, A.; Ghomashpasand, M. Comparative study of photocatalytic activities of magnetically separable WO3/TiO2/Fe3O4 nanocomposites and TiO2, WO3/TiO2 and TiO2/Fe3O4 under visible light irradiation. Superlattice Microstruct. 2015, 88, 211–224. [Google Scholar] [CrossRef]
  139. Chidambaram, S.; Pari, B.; Kasi, N.; Muthusamy, S. ZnO/Ag heterostructures embedded in Fe3O4 nanoparticles for magnetically recoverable photocatalysis. J. Alloy Compd. 2016, 665, 404–410. [Google Scholar] [CrossRef]
  140. Yang, J.; Wang, J.; Li, X.; Wang, D.; Song, H. Synthesis of urchin-like Fe3O4@SiO2@ZnO/CdS core-shell microspheres for the repeated photocatalytic degradation of rhodamine B under visible light. Catal. Sci. Technol. 2016. [Google Scholar] [CrossRef]
  141. Shekofteh-Gohari, M.; Habibi-Yangjeh, A. Ultrasonic-assisted preparation of novel ternary ZnO/AgI/Fe3O4 nanocomposites as magnetically separable visible-light-driven photocatalysts with excellent activity. J. Colloid Interface Sci. 2016, 461, 144–153. [Google Scholar] [CrossRef] [PubMed]
  142. Shekofteh-Gohari, M.; Habibi-Yangjeh, A. Novel magnetically separable ZnO/AgBr/Fe3O4/Ag3VO4 nanocomposites with tandem n–n heterojunctions as highly efficient visible-light-driven photocatalysts. RSC Adv. 2015, 6, 2402–2413. [Google Scholar]
  143. Shekofteh-Gohari, M.; Habibi-Yangjeh, A. Ternary ZnO/Ag3VO4/Fe3O4 nanocomposites: Novel magnetically separable photocatalyst for efficiently degradation of dye pollutants under visible-light irradiation. Solid State Sci. 2015, 48, 177–185. [Google Scholar] [CrossRef]
  144. Shekofteh-Gohari, M.; Habibi-Yangjeh, A. Novel magnetically separable Fe3O4@ZnO/AgCl nanocomposites with highly enhanced photocatalytic activities under visible-light irradiation. Sep. Purif. Technol. 2015, 147, 194–202. [Google Scholar] [CrossRef]
  145. Shekofteh-Gohari, M.; Habibi-Yangjeh, A. Facile preparation of Fe3O4@AgBr-ZnO nanocomposites as novel magnetically separable visible-light-driven photocatalysts. Ceram. Int. 2015, 41, 1467–1476. [Google Scholar] [CrossRef]
  146. Jia, X.; Dai, R.; Sun, Y.; Song, H.; Wu, X. One-step hydrothermal synthesis of Fe3O4/g-C3N4 nanocomposites with improved photocatalytic activities. J. Mater. Sci. 2016, 27, 3791–3798. [Google Scholar] [CrossRef]
  147. Mousavi, M.; Habibi-Yangjeh, A. Magnetically separable ternary g-C3N4/Fe3O4/BiOI nanocomposites: Novel visible-light-driven photocatalysts based on graphitic carbon nitride. J. Colloid Interface Sci. 2016, 465, 83–92. [Google Scholar] [CrossRef] [PubMed]
  148. Akhundi, A.; Habibi-Yangjeh, A. Novel magnetic g-C3N4/Fe3O4/AgCl nanocomposites: Facile and large-scale preparation and highly efficient photocatalytic activities under visible-light irradiation. Mater. Sci. Semicond. Process. 2015, 39, 162–171. [Google Scholar] [CrossRef]
  149. Akhundi, A.; Habibi-Yangjeh, A. Novel magnetically separable g-C3N4/AgBr/Fe3O4 nanocomposites as visible-light-driven photocatalysts with highly enhanced activities. Ceram. Int. 2015, 41, 5634–5643. [Google Scholar] [CrossRef]
  150. Feng, X.; Lou, X. The effect of surfactants-bound magnetite (Fe3O4) on the photocatalytic properties of the heterogeneous magnetic zinc oxides nanoparticles. Sep. Purif. Technol. 2015, 147, 266–275. [Google Scholar] [CrossRef]
  151. Li, W.; Tian, Y.; Li, P.; Zhang, B.; Zhang, H.; Geng, W.; Zhang, Q. Synthesis of rattle-type magnetic mesoporous and investigation of its photoactivity in the degradation of methylene blue. RSC Adv. 2015, 5, 48050–48059. [Google Scholar] [CrossRef]
  152. Li, X.; Wang, G.; Cheng, Y. Preparation and characteristics of Fe3O4-BaTiO3 heterostructural nanocomposite as photocatalyst. Res. Chem. Intermed. 2015, 41, 3031–3039. [Google Scholar] [CrossRef]
  153. Singh, K.K.; Senapati, K.K.; Borgohain, C.; Sarma, K.C. Newly developed Fe3O4-Cr2O3 magnetic nanocomposite for photocatalytic decomposition of 4-chlorophenol in water. J. Environ. Sci. 2015. [Google Scholar] [CrossRef]
  154. Cao, C.; Xiao, L.; Chen, C.; Cao, Q. Magnetically separable Cu2O/chitosan–Fe3O4 nanocomposites: Preparation, characterization and visible-light photocatalytic performance. Appl. Surf. Sci. 2015, 333, 110–118. [Google Scholar] [CrossRef]
  155. Cao, Y.; Li, C.; Li, J.; Li, Q.; Yang, J. Magnetically separable Fe3O4/AgBr hybrid materials: Highly efficient photocatalytic activity and good stability. Nanoscale Res. Lett. 2015, 10, 952. [Google Scholar] [CrossRef] [PubMed]
  156. Zhong, S.; Jiang, W.; Han, M.; Liu, G.; Zhang, N.; Lu, Y. Graphene supported silver@silver chloride and ferroferric oxide hybrid, a magnetically separable photocatalyst with high performance under visible light irradiation. Appl. Surf. Sci. 2015, 347, 242–249. [Google Scholar] [CrossRef]
  157. Jing, J.; Zhang, Y.; Feng, J.; Li, W.; Yu, W.W. Facile preparation and high performance of magnetically separable metalloporphyrin. Chem. Eng. J. 2015, 263, 385–391. [Google Scholar] [CrossRef]
  158. Wang, B.; Zhang, M.; Li, W.; Wang, L.; Zheng, J.; Gan, W. Fabrication of Au(Ag)/AgCl/Fe3O4@PDA@Au nanocomposites with enhanced visible-light-driven photocatalytic activity. Dalton Trans. 2015, 44, 17020–17025. [Google Scholar] [CrossRef] [PubMed]
  159. Yang, S.; Ye, C.; Song, X.; He, L.; Liao, F. Theoretical calculation based synthesis of a poly(p-phenylenediamine)–Fe3O4 composite: A magnetically recyclable photocatalyst with high selectivity for acid dyes. RSC Adv. 2014, 4, 54810–54818. [Google Scholar] [CrossRef]
  160. An, L.; Wang, G.; Shi, X.; Su, M.; Gao, F.; Cheng, Y. Recyclable Fe3O4/ZnO/PPy composite photocatalyst: Fabrication and photocatalytic activity. Russ. J. Phys. Chem. A 2014, 88, 2419–2423. [Google Scholar] [CrossRef]
  161. Wang, C.; Cao, M.; Wang, P.; Ao, Y.; Hou, J.; Qian, J. Preparation of a magnetic graphene oxide-Ag3PO4 composite photocatalyst with enhanced photocatalytic activity under visible light irradiation. J. Taiwan Inst. Chem. Eng. 2014, 45, 1080–1086. [Google Scholar] [CrossRef]
  162. Peik-See, T.; Pandikumar, A.; Ngee, H.; Ming, N. Magnetically separable reduced graphene oxide/iron oxide nanocomposite materials for environmental remediation. Catal. Sci. Technol. 2014, 4, 4396–4405. [Google Scholar] [CrossRef]
  163. Li, Z.; Wen, Y.; Shang, J.; Wu, M.; Wang, L.; Guo, Y. Magnetically recoverable Cu2O/Fe3O4 composite photocatalysts: Fabrication and photocatalytic activity. Chin. Chem. Lett. 2014, 25, 287–291. [Google Scholar] [CrossRef]
  164. Cao, X.; Chen, Y.; Jiao, S.; Fang, Z.; Xu, M.; Liu, X.; Li, L.; Pang, G.; Feng, S. Magnetic photocatalysts with a p–n junction: Fe3O4 nanoparticle and FeWO4 nanowire heterostructures. Nanoscale 2014, 6, 12366–12370. [Google Scholar] [CrossRef] [PubMed]
  165. Wang, H.; Wei, Z.; Matsui, H.; Zhou, S. Fe3O4/carbon quantum dots hybrid nanoflower for highly active and recyclable visible-light driven. J. Mater. Chem. A 2014, 2, 15740–15745. [Google Scholar] [CrossRef]
  166. Zhang, D.; Pu, X.; Gao, Y.; Su, C.; Li, H.; Li, H. One-step combustion synthesis of CoFe2O4-graphene hybrid materials for photodegradation of methylene blue. Mater. Lett. 2013, 113, 179–181. [Google Scholar] [CrossRef]
  167. Zhu, H.; Jiang, R.; Huang, S.; Yao, J.; Fu, F.; Li, J. Novel magnetic NiFe2O4/multi-walled carbon nanotubes hybrids: Facile synthesis, characterization, and application to the treatment of dyeing wastewater. Ceram. Int. 2015, 41, 11625–11631. [Google Scholar] [CrossRef]
  168. Bai, S.; Shen, X. One-pot solvothermal preparation of magnetic reduced graphene oxide-ferrite hybrids for organic dye removal. Carbon 2012, 50, 2337–2346. [Google Scholar] [CrossRef]
  169. Sun, J.; Fu, Y.; Xiong, P.; Sun, X.; Xu, B.; Wang, X. A magnetically separable P25/CoFe2O4/graphene catalyst with enhanced adsorption capacity and visible-light-driven photocatalytic activity. RSC Adv. 2013, 3, 22490–22497. [Google Scholar] [CrossRef]
  170. Xiong, P.; Zhu, J.; Wang, X. Cadmium sulfide-ferrite nanocomposite as a magnetically recyclable photocatalyst with enhanced visible-light-driven photocatalytic activity and photostability. Ind. Eng. Chem. Res. 2013, 52, 17126–17133. [Google Scholar] [CrossRef]
  171. Fu, Y.; Chen, Q.; He, M.; Wan, Y.; Sun, X.; Xia, H.; Wang, X. Copper ferrite-graphene hybrid: A multifunctional heteroarchitecture for photocatalysis and energy storage. Ind. Eng. Chem. Res. 2012, 51, 11700–11709. [Google Scholar] [CrossRef]
  172. Ji, H.; Jing, X.; Xu, Y.; Yan, J.; Li, H.; Li, Y.; Huang, L.; Zhang, Q.; Xu, H.; Li, H. Magnetic g-C3N4/NiFe2O4 hybrids with enhanced photocatalytic activity. RSC Adv. 2015, 5, 57960–57967. [Google Scholar] [CrossRef]
  173. Zhang, W.; Wang, M.; Zhao, W.; Wang, B. Magnetic composite photocatalyst ZnFe2O4/BiVO4: Synthesis, characterization, and visible-light photocatalytic activity. Dalton Trans. 2013, 42, 15464–15474. [Google Scholar] [CrossRef] [PubMed]
  174. Yao, Y.; Cai, Y.; Lu, F.; Qin, J.; Wei, F.; Xu, C.; Wang, S. Magnetic ZnFe2O4-C3N4 hybrid for photocatalytic degradation of aqueous organic pollutants by visible light. Ind. Eng. Chem. Res. 2014, 53, 17294–17302. [Google Scholar] [CrossRef]
  175. Wu, S.; Wang, P.; Cai, Y.; Liang, D.; Ye, Y.; Tian, Z.; Liu, J.; Linag, C. Reduced graphene oxide anchored magnetic ZnFe2O4 nanoparticles with enhanced visible-light photocatalytic activity. RSC Adv. 2015, 5, 9069–9074. [Google Scholar] [CrossRef]
  176. Yang, D.; Feng, J.; Jiang, L.; Wu, X.; Sheng, L.; Jiang, Y.; Wei, T.; Fan, Z. Photocatalyst interface engineering: Spatially confined growth of ZnFe2O4 within graphene networks as excellent visible-light-driven photocatalysts. Adv. Funct. Mater. 2015, 25, 7080–7087. [Google Scholar] [CrossRef]
  177. Khadgi, N.; Li, Y.; Upreti, A.R.; Zhang, C.; Zhang, W.; Wang, Y.; Wang, D. Enhanced photocatalytic degradation of 17 α-ethinylestradiol exhibited by multifunctional ZnFe2O4-Ag/rGO nanocomposite under visible light. Photochem. Photobiol. 2016, 92, 238–246. [Google Scholar] [CrossRef] [PubMed]
  178. Chen, X.; Dai, Y.; Guo, J.; Liu, T.; Wang, X. Novel magnetically separable reduced graphene oxide (RGO)/ZnFe2O4/Ag3PO4 nanocomposites for enhanced photocatalytic performance toward 2,4-dichlorophenol under visible light. Ind. Eng. Chem. Res. 2016, 55, 568–578. [Google Scholar] [CrossRef]
  179. Kulkarni, S.D.; Kumbar, S.; Menon, S.G.; Choudhari, K.S.; Santhosh, C. Magnetically separable core-shell ZnFe2O4@ZnO nanoparticles for visible light photodegradation of methyl orange. Mater. Res. Bull. 2016, 77, 70–77. [Google Scholar] [CrossRef]
  180. Yao, Y.; Qin, J.; Chen, H.; Wei, F.; Liu, X.; Wang, J.; Wang, S. One-pot approach for synthesis of N-doped TiO2/ZnFe2O4 hybrid as an efficient photocatalyst for degradation of aqueous organic pollutants. J. Hazard. Mater. 2015, 291, 28–37. [Google Scholar] [CrossRef] [PubMed]
  181. Wu, S.; Shen, X.; Zhu, G.; Zhou, H.; Ji, Z.; Chen, K. Synthesis of ternary Ag/ZnO/ZnFe2O4 porous and hollow nanostructures with enhanced photocatalytic activity. Appl. Catal. B Environ. 2016, 184, 328–336. [Google Scholar] [CrossRef]
  182. 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]
  183. Fu, Y.; Chen, H.; Sun, X.; Wang, X. Graphene-supported nickel ferrite: A magnetically separable photocatalyst with high activity under visible light. Am. Inst. Chem. Eng. J. 2012, 58, 3298–3305. [Google Scholar] [CrossRef]
  184. Patil, S.S.; Tamboli, M.S.; Deonikar, V.G.; Umarji, G.G.; Ambekar, J.D.; Kulkarni, M.V.; Kolekar, S.S.; Kale, B.B.; Patil, D.R. Magnetically separable Ag3PO4/NiFe2O4 composites with enhanced photocatalytic activity. Dalton Trans. 2015, 44, 20426–20434. [Google Scholar] [CrossRef] [PubMed]
  185. Kim, S.H.; Kim, D.; Kwak, S.B.; Han, B.G.; Um, M.; Kang, M. Synthesis of magnetically separable core@shell structured NiFe2O4@TiO2 nanomaterial and its use for photocatalytic hydrogen production by methanol/water splitting. Chem. Eng. J. 2014, 243, 272–279. [Google Scholar] [CrossRef]
  186. Ge, M.; Hu, Z. Novel magnetic AgBr/NiFe2O4 composite with enhanced visible light photocatalytic performance. Ceram. Int. 2016, 42, 6510–6514. [Google Scholar] [CrossRef]
  187. 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, 195–196, 149–157. [Google Scholar] [CrossRef]
  188. Li, C.; Wang, J.; Wang, B.; Ru, J.; 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]
  189. 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]
  190. Pongwan, P.; Inceesungvorn, B.; Phanichphant, S.; Wetchakun, N. Synthesis and characterization of a magnetically separable CoFe2O4/TiO2 nanocomposite for the photomineralization of formic acid. Ferroelectrics 2013, 453, 133–140. [Google Scholar] [CrossRef]
  191. Wetchakun, N.; Chaiwichain, S.; Wetchakun, K. Synthesis and characterization of novel magnetically separable CoFe2O4/CeO2 nanocomposite photocatalysts. Mater. Lett. 2013, 113, 76–79. [Google Scholar] [CrossRef]
  192. Gan, L.; Shang, S.; Wah, C.; Yuen, M.; Jiang, S.; Hu, E. Hydrothermal synthesis of magnetic CoFe2O4/graphene nanocomposites with improved photocatalytic activity. Appl. Surf. Sci. 2015, 351, 140–147. [Google Scholar] [CrossRef]
  193. Singh, S.; Khare, N. Magnetically separable, CoFe2O4 decorated CdS nanorods for enhanced visible light driven photocatalytic activity. Mater. Lett. 2015, 161, 64–67. [Google Scholar] [CrossRef]
  194. 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]
  195. Choi, Y.I.; Kim, Y.; Cho, D.W.; Kang, J.; Leung, K.T.; Sohn, Y. Recyclable magnetic CoFe2O4/BiOX (X = Cl, Br and I) microflowers for photocatalytic treatment of water contaminated with methyl orange, rhodamine B, methylene blue, and a mixed dye. RSC Adv. 2015, 5, 79624–79634. [Google Scholar] [CrossRef]
  196. Xu, Y.; Zhou, T.; Huang, S.; Xie, M.; Li, H.; Xu, H. Preparation of magnetic Ag/AgCl/CoFe2O4 composites with high photocatalytic and antibacterial ability. RSC Adv. 2015, 5, 41475–41483. [Google Scholar] [CrossRef]
  197. Shi, B.; Wan, J.; Liu, C.; Yu, X. Synthesis of CoFe2O4/MCM-41/TiO2 composite microspheres and its performance in degradation of phenol. Mater. Sci. Semicond. Process. 2015, 37, 241–249. [Google Scholar] [CrossRef]
  198. Gaikwad, P.N.; Hankare, P.P.; Wandre, T.M.; Garadkar, K.M.; Sasikala, R. Photocatalytic performance of magnetically separable Fe,N co-doped TiO2-cobalt ferrite nanocomposite. Mater. Sci. Eng. B 2016, 205, 40–45. [Google Scholar] [CrossRef]
  199. Manikandan, A.; Durka, M.; Seevakan, K.; Antony, S.A. A novel one-pot combustion synthesis and opto-magnetic properties of magnetically separable spinel MnxMg1−xFe2O4 (0.0 ≤ x ≤ 0.5) nanophotocatalysts. J. Supercond. Novel Magn. 2015, 28, 1405–1416. [Google Scholar] [CrossRef]
  200. Tio, A.; Nife, O.; Nicola, D.; Vanags, M. Ag sensitized TiO2 and NiFe2O4 three-component nanoheterostructures: Synthesis, electronic structure and strongly enhanced visible light. RSC Adv. 2016, 6, 18834–18842. [Google Scholar]
  201. Chen, C.-C.; Fu, Y.-P.; Hu, S.-H. Characterizations of TiO2/SiO2/Ni-Cu-Zn ferrite composite for magnetic photocatalysts. J. Am. Chem. Soc. 2015, 98, 2803–2811. [Google Scholar] [CrossRef]
  202. Sathishkumar, P.; Viswanathan, R. 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]
  203. Kadi, M.W.; Mohamed, R.M. Environmental remediation of aqueous cyanide by photocatalytic oxidation using a NiFe2O4/TiO2-SiO2 core-shell nanocomposite. Desalin. Water Treat. 2016, 56, 1940–1948. [Google Scholar] [CrossRef]
  204. Su, R.N.; Lv, P.; Li, M.; Zhang, X.; Li, M.; Niu, J. Fabrication of MgFe2O4-ZnO heterojunction photocatalysts for application of organic pollutants. Mater. Lett. 2014, 122, 201–204. [Google Scholar] [CrossRef]
  205. Leng, C.; Wei, J.; Liu, Z. Facile synthesis of PANI-modified CoFe2O4-TiO2 hierarchical flower-like nanoarchitectures with high photocatalytic activity. J. Nanopart. Res. 2013, 15, 1643. [Google Scholar] [CrossRef]
  206. Sun, A.; Chen, H.; Song, C.; Jiang, F.; Wang, X.; Fu, Y. Magnetic Bi25FeO40-graphene catalyst and its high visible-light photocatalytic performance. RSC Adv. 2013, 3, 4332–4340. [Google Scholar] [CrossRef]
  207. Xie, T.; Xu, L.; Liu, C.; Yang, J.; Wang, M. Magnetic composite BiOCl-SrFe12O19:A novel p–n type heterojunction with enhanced photocatalytic activity. Dalton Trans. 2014, 43, 2211–2220. [Google Scholar] [CrossRef] [PubMed]
  208. Bhukal, S.; Bansal, S.; Singhal, S. Magnetic Mn substituted cobalt zinc ferrite systems: Structural, electrical and magnetic properties and their role in photo-catalytic degradation of methyl orange azo dye. Physica B 2014, 445, 48–55. [Google Scholar] [CrossRef]
  209. Qasim, M.; Asghar, K.; Raj, B.; Prathapani, S.; Khan, W.; Naqvi, A.H.; Das, D. Magnetically recyclable Ni0.5Zn0.5Fe2O4/Zn0.95Ni0.05O nano-photocatalyst: Structural, optical, magnetic and photocatalytic properties. Spectrochim. Acta A 2015, 137, 1348–1356. [Google Scholar] [CrossRef] [PubMed]
  210. Manikandan, A.; Antony, S.A. Magnetically separable MnxZn1−xFe2O4; (0.0 ≤ x ≤ 0.5) nanostructures: Structural, morphological, opto-magnetic and photocatalytic properties. Synth. React. Inorg. Met-Org. Nano-Met. Chem. 2016. [Google Scholar] [CrossRef]
  211. Shakir, I.; Sarfraz, M.; Ali, Z.; Aboud, M.F.A. Magnetically separable and recyclable graphene-MgFe2O4 nanocomposites for enhanced photocatalytic applications. J. Alloy. Compd. 2016, 660, 450–455. [Google Scholar] [CrossRef]
  212. Bhukal, S.; Singhal, S. Magnetically separable copper substituted cobalt-zinc nano-ferrite photocatalyst with enhanced photocatalytic activity. Mater. Sci. Semicond. Process. 2014, 26, 467–476. [Google Scholar] [CrossRef]
  213. Liu, F.; Xie, Y.; Yu, C.; Liu, X.; Dai, Y. Novel hybrid Sr-doped TiO2/magnetic Ni0.6Zn0.4Fe2O4 for enhanced separation and photodegradation of organics under visible light. RSC Adv. 2015, 5, 24056–24063. [Google Scholar] [CrossRef]
  214. Vignesh, K.; Suganthi, A.; Min, B.-K.; Kang, M. Photocatalytic activity of magnetically recoverable MnFe2O4/g-C3N4/TiO2 nanocomposite under simulated solar light irradiation. J. Mol. Catal. A Chem. 2014, 395, 373–383. [Google Scholar] [CrossRef]
  215. He, H.-Y. Photocatalytic degradations of dyes on magnetically separable Ni1−xCoxFe2O4 nanoparticles synthesized by a hydrothermal process. Part. Sci. Technol. 2016, 34, 143–151. [Google Scholar] [CrossRef]
  216. Zhang, Z.; Xu, L.; Liu, C. Preparation and characterization of composite magnetic photocatalyst MnxZn1−xFe2O4/β-Bi2O3. RSC Adv. 2015, 5, 79997–80004. [Google Scholar] [CrossRef]
  217. He, H.; Yan, Y.; Huang, J.; Lu, J. Rapid photodegradation of methyl blue on magnetic Zn1−xCoxFe2O4 nanoparticles synthesized by hydrothermal process. Sep. Purif. Technol. 2014, 136, 36–41. [Google Scholar] [CrossRef]
  218. Shi, Y.; Zhou, K.; Wang, B.; Jiang, S.; Qian, X.; Gui, Z.; Yuen, R.K.K.; Hu, Y. Ternary graphene-CoFe2O4/CdS nanohybrids: Preparation and application as recyclable photocatalysts. J. Mater. Chem. A 2014, 2, 535–544. [Google Scholar] [CrossRef]
  219. Sathishkumar, P.; Pugazhenthiran, N. ZnO supported CoFe2O4 nanophotocatalysts for the mineralization of Direct Blue 71 in aqueous environments. J. Hazard. Mater. 2013, 252–253, 171–179. [Google Scholar] [CrossRef] [PubMed]
Figure 1. (a,b) SEM images of α-Fe2O3-AgBr nonwoven cloth; (c,d) illustration of magnetic separation of α-Fe2O3-AgBr cloth and charge transfer mechanism in (e) α-Fe2O3-AgBr cloth (Reproduced with permission from [81], Copyright 2015, RSC); and (f) Ag/AgCl/Fe2O3. Reproduced with permission from [39]. Copyright 2016, RSC.
Figure 1. (a,b) SEM images of α-Fe2O3-AgBr nonwoven cloth; (c,d) illustration of magnetic separation of α-Fe2O3-AgBr cloth and charge transfer mechanism in (e) α-Fe2O3-AgBr cloth (Reproduced with permission from [81], Copyright 2015, RSC); and (f) Ag/AgCl/Fe2O3. Reproduced with permission from [39]. Copyright 2016, RSC.
Catalysts 06 00079 g001
Figure 2. (a) Schematic illustration of the sol-gel preparation of γ-Fe2O3/N.Fe-TiO2; and (b) photocatalytic degradation of RB 4 over γ-Fe2O3/N.Fe-TiO2 and the controls. Reproduced with permission from [111]. Copyright 2015, RSC.
Figure 2. (a) Schematic illustration of the sol-gel preparation of γ-Fe2O3/N.Fe-TiO2; and (b) photocatalytic degradation of RB 4 over γ-Fe2O3/N.Fe-TiO2 and the controls. Reproduced with permission from [111]. Copyright 2015, RSC.
Catalysts 06 00079 g002
Figure 3. (a) Charge transfer mechanism in γ-Fe2O3/Fe,N-TiO2 (Reproduced with permission from [111], Copyright 2015, RSC) and (be) SEM images of urchin-like γ-Fe2O3@SiO2@TiO2 composite microparticles. Reproduced with permission from [114]. Copyright 2015, RSC.
Figure 3. (a) Charge transfer mechanism in γ-Fe2O3/Fe,N-TiO2 (Reproduced with permission from [111], Copyright 2015, RSC) and (be) SEM images of urchin-like γ-Fe2O3@SiO2@TiO2 composite microparticles. Reproduced with permission from [114]. Copyright 2015, RSC.
Catalysts 06 00079 g003
Figure 4. Magnetization curves and insert showing magnetic separation of Fe3O4/TiO2/Ag. Reproduced with permission from [116]. Copyright 2016, RSC.
Figure 4. Magnetization curves and insert showing magnetic separation of Fe3O4/TiO2/Ag. Reproduced with permission from [116]. Copyright 2016, RSC.
Catalysts 06 00079 g004
Figure 5. (a,b) SEM images of urchin-like Fe3O4@SiO2@ZnO/CdS microspheres and (c) charge transfer route in Fe3O4@SiO2@ZnO/CdS. Reproduced with permission from [140]. Copyright 2016, RSC.
Figure 5. (a,b) SEM images of urchin-like Fe3O4@SiO2@ZnO/CdS microspheres and (c) charge transfer route in Fe3O4@SiO2@ZnO/CdS. Reproduced with permission from [140]. Copyright 2016, RSC.
Catalysts 06 00079 g005
Figure 6. (a) Charge transfer route in RGO/ZnFe2O4; (b) magnetisation curves for pure ZnFe2O4 and RGO/ZnFe2O4, insert showing magnetic separation of RGO/ZnFe2O4 (Reproduced with permission from [175], Copyright 2014, RSC); charge transfer mechanism in (c) ZnFe2O4–Ag/RGO (Reproduced with permission from [177], Copyright 2016, Wiley) and (d) RGO/ZnFe2O4/Ag3PO4. (Reproduced with permission from [178], Copyright 2016, ACS.
Figure 6. (a) Charge transfer route in RGO/ZnFe2O4; (b) magnetisation curves for pure ZnFe2O4 and RGO/ZnFe2O4, insert showing magnetic separation of RGO/ZnFe2O4 (Reproduced with permission from [175], Copyright 2014, RSC); charge transfer mechanism in (c) ZnFe2O4–Ag/RGO (Reproduced with permission from [177], Copyright 2016, Wiley) and (d) RGO/ZnFe2O4/Ag3PO4. (Reproduced with permission from [178], Copyright 2016, ACS.
Catalysts 06 00079 g006
Figure 7. Charge transfer route in g-C3N4/NiFe2O4. Reproduced with permission from [172]. Copyright 2015, RSC.
Figure 7. Charge transfer route in g-C3N4/NiFe2O4. Reproduced with permission from [172]. Copyright 2015, RSC.
Catalysts 06 00079 g007
Table 1. Summary of the photocatalytic properties of Fe3O4-incorporating TiO2-based photocatalysts.
Table 1. Summary of the photocatalytic properties of Fe3O4-incorporating TiO2-based photocatalysts.
SamplePreparation MethodApplicationCatalyst DosePollutant ConcentrationDegradation EfficiencySaturation MagnetisationReference
Fe3O4/GO/Ce-TiO2Low-temperature sol-gel/ultrasonicationDegradation of tetracycline (TC)/Vis/300 W Xe lamp/400 nm UV filter0.050 g25 ppm82.9% in 60 minNo data[33]
Fe3O4/RGO/TiO2Reduction-depositionDegradation of MB and tetrabromobisphenol A (TBBPA)/UV/230 W Hg lamp0.025 g10 ppm99.5% in 60 min/TBBPA, 95.9% in 60 min/MB1.14 emu/g[128]
Fe3O4/P(MAA-DVB)/TiO2Magnetic field induced assembly/precipitation-polymerisationDegradation of RhB/UV/250 W Hg lamp0.040 g4.43 ppm71.5% in 120 min35.2 emu/g[129]
Fe3O4@SiO2@N-TiO2Sol-gelDegradation of phenol/Vis/15 W florescent lamp0.800 g100 ppm46% in 480 min~2 emu/g[127]
Fe3O4/chitosan/TiO2Hydrothermal/crosslinkingDegradation of MB/UV/8 W UV lamp1.00 g1.28 ppm93% in 40 min4.2 emu/g[130]
Fe3O4/TiO2Reverse microemulsion/sol-gelDegradation of RhB/UV/50 W Xe lamp0.004 g22.12 ppm100% in 100 min21.68 emu/g[131]
Ag3PO4/TiO2/Fe3O4In situ hydrolysis/depositionDegradation of acid orange 7 (AO 7)/50 mW diode blue laser and E. coli/Vis/300 W Xe lamp/420 nm UV filter0.100 g5.25 ppm/AO7 and 107 CFU/mL/E. coli~100% in 2.5 min/AO 7, 99.8% in 5 min/E. coliNo data[38]
Fe3O4/TiO2/Bi2O3Sol-gelDegradation of MO/Simulated solar light/350 W Xe lamp0.200 gNo data69% in 150 minNo data[132]
Fe3O4/TiO2/AuSol-gel/hydrothermalDegradation of MB and E. coli/UV/4 × 9 W black lights0.015 g/MB and 0.010 g/E. coli8.00 ppm/MB and 108 CFU/mL/E. coli78% in 4 h/MB, 89.3% in 60 min/E. coliNo data[133]
Ag@Fe3O4@SiO2@TiO2Solvothermal Degradation of MB and Cr(VI) reduction/Vis/500 W Xe lamp/425 UV filter0.020 g50 ppm/MB and 22.24 ppm/K2Cr2O799.9% in 4 h/Cr(VI), ~90 mg/g/MB13.92 emu/g[134]
Fe3O4@SiO2@TiO2@GOReverse microemulsion/sol-gel/amide conjugationDegradation of RhB/UV/400 W column high pressure Hg lamp0.050 g8.85 ppm92% in 120 min16.90 emu/g[120]
Fe3O4@C@TiO2Solvothermal/calcinationDegradation of RhB/UV/125 W high pressure Hg lamp0.020 g10 ppm~100% in 80 min6.04 emu/g[135]
Fe3O4@SiO2@TiO2Stobber process/sol-gel/chemical precipitationDegradation of RhB/UV/50 W high pressure Hg lamp0.010 g8.85 ppm100% in 10 min30.60 emu/g[136]
Fe3O4/TiO2Hydrothermal Degradation of MB/UV/9 W UV lamp0.002 g1.00 ppm100% in 10 minNo data[75]
Fe3O4@TiO2@GRSol-gel/assembly routeDegradation of 2,4-dichlorophenoxy-acetic acid (2,4-D)/simulated solar light/500 W Xe arc lamp0.020 g20 ppm100% in 40 min13.00 emu/g[137]
WO3/TiO2/Fe3O4Sol-gelDegradation of direct blue 71 (DB 71)/Vis/200 W Xe lamp0.030 g50 ppm98% in 35 min18.20 emu/g[138]
Table 2. Summary of the photocatalytic properties of various Fe3O4-based photocatalyst.
Table 2. Summary of the photocatalytic properties of various Fe3O4-based photocatalyst.
SamplePreparation MethodApplicationCatalyst DosePollutant ConcentrationDegradation EfficiencyMagnetisation SaturationReference
Fe3O4@SiO2@CdSStobber method/chelating assisted growthDegradation of MB and TC/Vis0.010 g10 ppm/MB and 100 ppm/TC80% in 21 min/TC, 94% in 140 min/MB22.00 emu/g[63]
Fe3O4(TAMH)/ZnOSolvothermal Degradation of phenol/Vis/575 W MSR metal halide lamp0.325 g20 ppm71.3% in 150 minNo data[150]
Fe3O4@mSiO2@BiOBrSolvothermal Degradation of MB/Vis/500 W Xe lamp/420 nm UV filter0.100 g20 ppm96% in 120 min40.00 emu/g[151]
Fe3O4/BaTiO3Solvothermal/sol-gelDegradation of MO and orange II/UV/150 W UV lamp0.002 g10 ppm71.2% in 20 h/MO, 43.7% in 20 h/orange II60.50 emu/g[152]
Fe3O4/Cr2O3Wet chemical/ultrasonicationDegradation of 4-CP/UV/12 W low-pressure Hg lamp0.100 g1.29 ppm100% in 150 min~20 emu/g[153]
Cu2O/chitosan/Fe3O4Precipitation-reductionDegradation of Reactive Brilliant red X-3B (X-3B)/Vis/500 W W-halogen lamp0.100 g50 ppm99.7% in 50 min15.1 emu/g[154]
Fe3O4/AgBrPrecipitation routeDegradation of MO/Vis/300 W Xe arc lamp/420 nm UV filter0.100 g20 ppm85% in 12 minNo data[155]
RGO/Ag/AgCl/Fe3O4Solvothermal/deposition-precipitationDegradation of MB and RhB/Vis/500 W Xe arc/10 cm water filter0.100 g10 ppm97.4% in 100 min/MB, 97.9% in 120 min/RhB18.8 emu/g[156]
Metalloporphyrin/Fe3O4Covalent conjugationDegradation of AO 7/Vis/125 W W-halogen lamp0.444 g17.5 ppm69.0% in 5 h61.45 emu/g[157]
Au(Ag)/AgCl/Fe3O4@PDA@AuSolvothermal/galvanic replacementDegradation of MB/Vis0.004 g5 ppm100% in 20 min5.40 emu/g[158]
Fe3O4@resorcinol-formaldehyde–AgIn situ polymerisation/reductionInactivation of S. aureus, E. coli and degradation of 4-nitrophenol (4-NP), MB and RhB/Vis/No data on light source0.010 g16 ppm/RhB/MB/4-NP and 107 CFU/mL/ bacteria1.5 < OD in 6 h/E. coli, ~1.0 OD in 6 h, S. aureus, ~100% in 3 min/4-NP, ~100% in 4 min/MB and ~100% in 3 min/RhB~30 emu/g[26]
Poly(p-phenylenediamine)–Fe3O4Chemical oxidation polymerisationDegradation of bromocresol green (BG), blue (BB), purple (BP), RhB, neutral red (NR), MB, Sudan III (SIII), MO and Congo red (CR)/UV (500 W Hg)/Vis (500 W Xe lamp/420 nm UV filter)0.025 g50 ppm~95% in 1100 min/BB/UV, ~80% in 1100 min/BB/Vis, ~100% in 1100 min/BG/UV, ~90% in 1100 min/BG/VisNo data[159]
Polypyrrole (PPY)/Fe3O4/ZnOIn situ polymerisationDegradation of MB/UV/8 W germicidal lamp0.200 g10 ppm85.2% in 4 hNo data[160]
Fe3O4/GO/Ag3PO4Co-precipitation/ultrasonicationDegradation of MB/Vis/250 W halogen lamp/400 nm UV filter0.025 g20 ppm100% in 10 min12.56 emu/g[161]
RGO/Fe3O4In situ chemical synthesisDegradation of MB/natural sunlight/bright sunny days/9.00 am to 2.00 pm0.002 g10 ppm100% in 60 min30.30 emu/g[162]
ZrO2/Fe3O4/chitosanCo-precipitation/refluxingReduction of Cr(VI) and degradation of 4-CP/natural sunlight0.050 g/Cr(III), 0.010 g/4-CP70 ppm/K2Cr2O7, 20 ppm/4-CP 88.6% in 180 min/4-CP, 90.2% in 180 min/Cr(VI)42.00 emu/g[115]
Cu2O/Fe3O4Solvothermal/precipitationDegradation of MO/Vis/500 W Xe lamp/420 nm UV filter0.100 g30 ppm90% in 90 min41.70 emu/g[163]
FeWO4/Fe3O4Hydrothermal Degradation of MB/UV-Vis/500 W Xe lamp0.020 g20 ppm97.1% in 60 min9.00 emu/g[164]
Fe3O4@carbon quantum dots (CQDs)Hydrothermal Degradation of MB/Vis75 W Xe lamp/420 UV filter0.001 g1.00 ppm94.4% in 30 min33.80 emu/g[165]
BiOBr@SiO2@Fe3O4Hydrothermal/Stobber method2,2-bis(4-hydroxyphenyl)propane (BPA)/500 W Xe lamp/420 UV filter0.100 g20 ppm87.0% in 50 minNo data[119]
Table 3. Summary of the photocatalytic properties of ferrite-based photocatalysts.
Table 3. Summary of the photocatalytic properties of ferrite-based photocatalysts.
SamplePreparation MethodApplicationCatalyst DosePollutant ConcentrationDegradation EfficiencyMagnetisation SaturationReference
P25/CoFe2O4/RGOHydrothermal Degradation of MB, MO, neutral dark yellow (NDY)/Vis/500 W Xe lamp/420 nmUV filter0.010 g40 ppm~90 mg/g in 60 min/MB, ~20 mg/g in 60 min and 10 mg/g in 60 min~35 emu/g[169]
MnxMg1−xFe2O4 (0.0 ≤ x ≤ 0.5)Microwave-assisted combustionDegradation of 4-CP/Vis/8 × 8 W medium pressure Hg lamps0.030 g200 ppm~98% in 300 min64.67 emu/g[199]
Ag/TiO2/NiFe2O4Photodeposition/solvothermalDegradation of MO/Vis/100 W LED0.100 g10 ppm100% in 120 minNo data[200]
TiO2@SiO2@Ni-Cu-ZnFe2O4Chemical co-precipitationDegradation of MB/simulated solar light/35 W Xe arc lamp0.400 g10 ppm83.9% in 6 h37.45 emu/g[201]
CoFe2O4/TiO2Co-precipitationDegradation of Reactive Red 120 (RR 120)/Vis/150 W W-halogen lamp0.400 g14.70 ppm4.98 × 10−9 S−1~0.2 emu/g[202]
Ag/NiFe2O4Combustion Degradation of MB/Vis/300 W Xe lamp/450 nm UV filter0.025 g20 ppm~80% in 120 min~20 emu/g[80]
NiFe2O4/TiO2-SiO2Modified sol-gel/solvothermalDegradation of cyanide/Vis/150 W blue fluorescent lamp/420 nm UV filterNo data100 ppm100% in 60 min42.7 emu/g[203]
MgFe2O4-ZnOSolution method/chemical co-precipitationDegradation of RhB/Vis/500 W Xe lamp/420 nm UV filter0.050 g4.40 ppm100% in 120 min21.37 emu/g[204]
PANI-CoFe2O4-TiO2Hydrothermal/in situ chemical depositionDegradation of MB/UV (500 W Xe lamp)/Vis/500 W Xe lamp/400 nm UV filter0.100 g50 ppm0.0962/min/UV and 0.0110/min/Vis11.4 emu/g[205]
Bi25FeO40-RGOHydrothermal Degradation of MB/Vis/500 W Xe lamp/400 nm UV filter0.080 g28.80 ppm92.8% in 180 min10.50 emu/g[206]
BiOCl-SrFe12O19Hydrothermal Degradation of MB/UV/Vis/No data on light source0.400 g10 ppm99% in 50 min/UV, 67.8% in 8 h/Vis15.13 emu/g[207]
CuFe2O4@C3N4Self-assembly routeDegradation of orange II/Vis/500 W Xe lamp/420 nm UV filter0.010 g9.81 ppm100% in 90 minNo data[40]
Co0.6Zn0.4MnxFe2-xO4 (x = 0.2–1.0)Sol-gel auto-combustionDegradation of MO/Vis/400 W Hg lamp0.100 g30 ppm94% in 60 min39.51 emu/g[208]
Ni0.5Zn0.5Fe2O4/Zn0.95Ni0.05Egg albumen assisted sol-gelDegradation of RhB/natural sunlight/sunny days/11 am to 3 pm.0.010 g20 ppm~90% in 4 h8.00 emu/g[209]
CdFe2O4/GRHydrothermal Degradation of MB/Vis/500 W Xe lamp/420 nm UV filter0.100 g10 ppm89.2% in 4 h14.26 emu/g[82]
MnxZn1−xFe2O4-TiO2 (0.0 ≤ x ≤ 0.5)Auto-combustionDegradation of 4-CP/UV/8 × 8 W low pressure Hg lamps0.030 g200 ppm~98% in 270 min45.17 emu/g[210]
Co0.6Zn0.4CuxFe2-xO4Sol-gel auto-combustionDegradation of MO/Vis/400 W Hg lamp0.100 g30 ppm86% in 60 min38.02 emu/g[211]
CoxZn1−xFe2O4-GRChemical co-precipitation/ultrasonicationDegradation of MB/Vis/No data on light source0.100 g5 ppm>95% in 60 minNo data[212]
Sr-TiO2/Ni0.6Zn0.4Fe2O4Combustion/sol-gelDegradation of BPA/UV (10 W low pressure Hg lamp)/Vis/500 W Xe arc lamp0.150 g10 ppm100% in 4 h/UV, 90% in 4 h/Vis19.04 emu/g[213]
CoFe2O4-GRCombustion Degradation of MB/Vis/300 W Xe lamp/450 nm UV filter0.025 g20 ppm100% in 120 min5.3 emu/g[166]
MnFe2O4/g-C3N4/TiO2Chemical impregnationDegradation of MO/Simulated solar light/150 W Xe arc lamp0.050g10 ppm99.3% in 180 min0.065 emu/g[214]
Ni1−xCoxFe2O4Hydrothermal Degradation of malachite green (MG)/natural sunlight (200–250 Wm−2)0.025 g0.365 ppm~100% in 60 min~50.7–64.2 emu/g.[215]
MnxZn1−xFe2O4/β-Bi2O3Dip-calcinationDegradation of RhB/Simulated solar light/300 W Xe lamp0.200g10 ppm99.1% in 150 min7.01 emu/g[216]
Zn1−xCoxFe2O4Hydrothermal Degradation of MB/Natural sunlight0.050 g3.20 ppm~88% in 270 min~50 emu/g[217]
SrFeO3−x/g-C3N4Sintering methodDegradation of chloramphenicol (CAP) and crystal violet (CV)/Vis/150 W Xe arc lamp0.010 g10 ppm91.3% in 96 h/CAP, 99.9% in 12 h/CV~0.17 emu/g[41]
CoFe2O4/GR/CdSSolvothermal Degradation of MB/daylight/40 W daylight lamp0.025 g20 ppm80% in 180 min18.00 emu/g[218]
ZnO/CoFe2O4Co-precipitationDegradation of direct blue 71 (BD 71)/150 W W-halogen lamp0.160 g38.64 ppm~100% in 30 min~0.040 emu/g[219]

Share and Cite

MDPI and ACS Style

Mamba, G.; Mishra, A. Advances in Magnetically Separable Photocatalysts: Smart, Recyclable Materials for Water Pollution Mitigation. Catalysts 2016, 6, 79. https://doi.org/10.3390/catal6060079

AMA Style

Mamba G, Mishra A. Advances in Magnetically Separable Photocatalysts: Smart, Recyclable Materials for Water Pollution Mitigation. Catalysts. 2016; 6(6):79. https://doi.org/10.3390/catal6060079

Chicago/Turabian Style

Mamba, Gcina, and Ajay Mishra. 2016. "Advances in Magnetically Separable Photocatalysts: Smart, Recyclable Materials for Water Pollution Mitigation" Catalysts 6, no. 6: 79. https://doi.org/10.3390/catal6060079

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