Next Article in Journal
Investigating the Effect of Nudges on Consumers’ Willingness to Pay for Genetically Modified Corn Oil
Previous Article in Journal
Would You Accept Virtual Tourism? The Impact of COVID-19 Risk Perception on Technology Acceptance from a Comparative Perspective
Previous Article in Special Issue
Biochar-Added Cementitious Materials—A Review on Mechanical, Thermal, and Environmental Properties
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

The Potential of rGO@TiO2 Photocatalyst for the Degradation of Organic Pollutants in Water

1
Faculty of Mechanical Engineering and Naval Architecture, University of Zagreb, Ivana Lučića 5, 10000 Zagreb, Croatia
2
Centre for Mechanical Technology and Automation, Department of Mechanical Engineering, University of Aveiro, 3810-193 Aveiro, Portugal
3
Intelligent Systems Associate Laboratory (LASI), 3810-193 Aveiro, Portugal
4
Department for Nanostructured Materials, Jožef Stefan Institute, Jamova Cesta 39, SI-1000 Ljubljana, Slovenia
*
Authors to whom correspondence should be addressed.
Sustainability 2022, 14(19), 12703; https://doi.org/10.3390/su141912703
Submission received: 10 August 2022 / Revised: 20 September 2022 / Accepted: 29 September 2022 / Published: 6 October 2022
(This article belongs to the Special Issue Sustainable Functional Materials)

Abstract

:
The availability of clean water is essential for humans wellbeing and the diverse biotic population in the environment. Menkind imposes a significant pressure on food supplies, natural resources, and other commodities. Large-scale anthropogenic activities, such as agriculture and industry, which are practiced to ensure population growth and survival, have caused several harmful environmental effects, including the discharge of pollutants into the aquatic environment. Among them organic micropollutants (OMPs) are considered a significant threat to aquatic ecosystems. The efficient removal of this persistent pollutants from wastewater is nowadays commonly considered in water treatment technologies. Utilizing photocatalysis by harvesting solar energy as an inexhaustible source, represents a facile and easy to upscale solution, for removing persistent pollutants and other emerging contaminants. In the recent decade, graphene-based titanium dioxide catalyst (rGO@TiO2) has received a lot of attention as an effective material for the degradation in the enviroment. This review summarizes the recent progress in preparing rGO@TiO2 nanocomposites and their utilization for purification purposes. Its main advantage over many other photocatalysts is its easy preparation, low toxicity, and reasonable photo-activity in a specific solar light spectrum.

1. Introduction

Today, clean water sources represent the highest priority for the sustainability of the entire ecosystem [1,2,3]. Environmental issues that are related to water pollution raise a serious concern over the last couple of decades due to the expanding industrial and agriculture activities [4,5,6,7]. Toxic organic pollutants amassed in our proximity have a negative effect on the biotic habitat and even human health. It is still extremely difficult to envision the elimination of these persistent pollutants which were found virtually in all water bodies [8,9,10]. Among them, persistent organic micropollutants (OMPs) such as organic dyes (i.e., methylene blue (MB), methylene orange, congo red), pesticides, pharmaceuticals, and their metabolites were proven extremely toxic [11]. The waste consisting of OMPs is also deemed problematic due to its difficulty for removal, low biodegradability, and long degradation lifetime in the ecosystem [12].
Conventional wastewater methods such as adsorption, biological treatment, filtration, chemical treatment, flocculation, etc., are not efficient enough to degrade especialy aromatic and heterocyclic groups in their moleculear structure [13,14]. The remediation should consider degradation and mineralization of this contaminants to its highest stable oxidation state such as: water, carbon dioxide, and the oxidized inorganic anions or into less toxic and also easily degradable species, that can be remediated through naturally pathways [15]. Therefore, extensive efforts have been directed globally toward the development of suitable, efficient, and eco-friendly processes for the removal of persistent pollutants. Nevertheless, strategies are being continuously developed for their remediation [16,17]. The task at hand is to choose an efficient and environmentally benign method, which would clean wastewater of its toxic ingredients. Recent research is focused on the advanced oxidation process (AOP) utilizing semiconductor nanoparticles with appropriate physical and chemical properties which would enable photocatalysis [18,19,20].

1.1. Challenges of TiO2 Semiconductor Photocatalysts

Titanium dioxide (TiO2) represents a well-studied photocatalyst that has been widely considered in research activities and is also commercialized. Because of its appealing physical and chemical properties, it was found useful in a wide spectrum of applications. Its thermal and chemical stability makes it resistant and unproblematic in the environment, it has good mechanical properties, and with respect to photocatalytic processes, it exhibits great activity under ultraviolet (UV) light irradiation (<387 nm) [21,22]. However, utilization of TiO2 in a wider sunlight spectrum, which has its maximum in the visible range, requires a decrease of its bandgap energy, whilst inhibiting the recombination of photogenerated electron (e)—hole (h+) pairs. Among the many approaches composite based on graphene and graphene-like materials (GO, rGO) addresses major limitations of TiO2 [23,24,25,26]. Coupling TiO2 with narrow bandgap materials with the visible light response is important to obtain an effective nanocomposite for photocatalytic applications. The synergistic effect between materials constituting nanocomposite improves visible light response, charge separation, and thus photocatalytic activity. On top of all both species were found way less harmful to the environment.
The mechanism of semiconductor photocatalysis can be divided into three steps (Figure 1). Firstly, using irradiation with the required energy (wavelength), electron (e)—hole (h+) pairs are produced within the TiO2 semiconductor particles. Furthermore, photo-generated charges must diffuse to the surface of the TiO2 hence the low recombination rate. Lastly, generated electrons (e) and holes (h+) must have the right reductive and oxidative potential to initiate chemical reactions driven through the surface of the catalyst. These reactions subsequently transform OMPs into low or non-toxic compounds [27,28].
While papers related to rGO-based TiO2 photocatalyst are continuously being published, a review in front of you is aimed at summarizing recent work on nanocomposites. More precisely, the main objective is to highlight few synthesis methods of graphene-based TiO2 nanocomposites formation and their operation characteristics for efficient photocatalytic water purification treatment. This work also summarizes few research conducted on diverse water bodies, which represents an important step towards practical application.

1.2. Why Coupling TiO2 with Graphene?

Graphene is a single layer of sp2-bonded carbon atoms firmly packed into a two-dimensional honeycomb structure. It is an allotrope of carbon, with extraordinary properties, such as large theoretical specific surface area (2630 m2 g−1), optical transparency, high Young’s modulus (~1 TPa), high carrier mobility at room temperature (~10,000 cm2 V−1 s−1), and excellent thermal conductivity (3000–5000 W m−1 K−1) [26,29,30]. Graphene has been widely used to improve the catalytic efficiency of photocatalysts. The photogenerated charge carriers are separated and the recombination rate of electrons/holes is reduced, thus the lifespan of charge carriers is extended. As a result, more reductive electrons and oxidative holes are accessible for the reaction, and the photocatalytic activity is enhanced [29,31,32,33].
Graphene Oxide (GO) is a valuable derivative of graphene. It consists of also oxygen and hydrogen and their functional groups which are attached to the hexagonal carbon skeleton thus disrupting the perfect sp2 hybridization. The process of partial sp2 bond recovery produces another important hybrid of carbon named reduced graphene oxide (rGO). Ideally, GO consists of a single carbon layer structure with oxygen containing functional groups positioned in or out of the hexagonal skeleton plane. However, most often GO is composed of multilayers. In contrast to perfect graphite consisting of single graphene layers connected through van der walls bonds, carbon layers are intercalated with oxygen functional groups and their bonds. GO is also considered two-dimensional carbon material; however, its properties are far from that of graphene. It does not absorb visible light, has very low electric and thermal conductivity compared to graphene, and demonstrates significantly higher chemical reactivity due to the presence of oxygen species [33,34,35]. Graphene can be oxidized to consist of several different oxygen functional groups such as epoxy, carbonyl, carboxyl, and hydroxyl, which play a relevant role in catalytic applications. The physicochemical properties of GO can be significantly changed according to the oxidation process applied [36]. By this, we can control the concentration of defects in the carbon structure Ref. [37]. By exploring several chemical [38] or physical treatments, ref. [39] GO can be converted in rGO which is in terms of physic-chemical properties an intermediate material of GO and graphene and deserve its own labeling. The molecular structures of graphene and its derivates (GO and rGO) are shown in Figure 2.
GO and rGO have been usually synthesized using natural graphite flakes as a precursor material and in the following section, the most common approaches to synthesizing them will be discussed.

1.2.1. Synthesis of Graphene Oxide (GO)

Carbon has fascinated chemists already a very long time ago and in 1859 Benjamin Brody synthesized GO from graphite flakes using KClO3 and HNO3 [41], even though graphene was not discovered at that time yet. In 1898, Staudenmaier [42] has slightly modified Brody’s method by using a concentrated H2SO4 in the presence of KClO3 and HNO3. In 1958, Hummers [43] has improved the existing process of GO production by using strong oxidizing agents such as KMnO4 and NaNO3 and the obtained product was a higher yield and better purity in comparison to the product from previous methods. The reported methods are very simple for synthesis, however, a major drawback is the formation of several toxic components (NO2, N2O4, and ClO2) during the process [44]. In 2010, Marcano et al. [45], prepared GO material by implementing several changes to Hummers’ method, called modified Hummers’ method and improved Hummers’ method, using natural graphite flakes (~150 µm).
In the last decade, many researchers worldwide have given attention to GO used for many applications obtained via different approaches of the Hummers´ method due to its higher yield in the production of GO [46,47,48,49,50]. All these reports implicate that excellent knowledge of the synthesis of GO material is of paramount importance for obtaining a material with desirable properties for targeted applications.

1.2.2. Reduction of GO to rGO

Through the reduction of GO to rGO, functional groups of oxygen (hydroxyl, carbonyl, carboxyl, and epoxy) can be removed. Since these methods can greatly transform the microstructure and properties of GO, several successful reductions have been performed by chemical [51,52], thermal [51,53], sonochemical [54], electrochemical [55], and photochemical [56] methods.
The most common transformation of GO into rGO is through thermal annealing. The oxygen-containing functional groups decompose during high temperatures into gases. These gases can produce high pressure between the stacked layers. According to the reported research [57], pressures of 40 MPa and 130 MPa are produced at 300 °C and at 1000 °C, respectively.

1.2.3. Preparation of rGO-Based TiO2 Photocatalyst

rGO-based TiO2 material is commonly used in light-driven photocatalysis of dyes in an aqueous medium. Because of exceptional properties, rGO-based oxide semiconductors promote electron separation, which results in boosting photo-driven reactions such as the degradation of carcinogenic dyes (e.g., methylene blue) and solar-fuel (hydrogen) production. Preparation of rGO-based TiO2 photocatalysts increases the specific surface area of the nanocomposite, consequently increasing the photocatalytic activity, which is why rGO-based semiconductor photocatalysts have been found to be promising in several applications. In the last decade, many studies have reported on improved photocatalytic activity on OMPs degradation. In these studies, rGO has been coupled with TiO2, the well-recognized semiconductor material, to enhance their performance in the photodegradation processes of OMPs from water bodies [58,59,60].
Nowadays, the rGO-based TiO2 nanoparticles are synthesized by physical as well as chemical methods such as sol-gel, hydrothermal and solvothermal techniques and by calcination treatment, hydrolysis, etc. The sol-gel method is a simple and cheap synthesis process that provides homogeneity to the prepared nanocomposite and is performed at very low temperatures. Preetha et al. [61] has reported the synthesis of rGO@TiO2 nanocomposite by the sol-gel method using titanium (IV) isopropoxide (TTIP) as a precursor in an ethanol medium followed by calcination treatment. Different weight percentages of GO were dispersed before the sol-gel reaction. The hydrothermal/solvothermal technique is one of the simplest synthesis techniques widely adopted for the preparation of nanocomposite photocatalysts. The morphology and physicochemical properties are significantly affected by the hydrothermal conditions such as duration of synthesis, temperature, cooling conditions, used precursors, etc. [26]. The high pressure used in the hydrothermal/solvothermal technique leads to good interactions between used precursors and can be performed at even lower temperatures. The synthesized nanocomposite exhibits synergy in properties of individual components which are attractive in photocatalytic degradation. During the hydrothermal/solvothermal technique, GO is mixed with a precursor in solution, where the latter transforms into small crystals on the surface of graphene [62]. Several authors reported how enhanced degradation of organic pollutants can be achieved using such composite prepared by this simple and cost-effective hydrothermal technique followed by calcination treatment (Table 1). Another case of enhanced photocatalytic degradation, of methylene blue and rhodamine B in an aqueous medium [63], was observed using self-assembled rGO@TiO2 nanocomposite synthesized by a one-pot hydrothermal technique using TTIP and GO at 180 °C for 6 h followed by calcination treatment at 300 °C for 2 h. Wanag et al. [25] reported a two-step solvothermal technique followed by calcination treatment for the preparation of rGO@TiO2 photocatalyst nanocomposites with improved photocatalytic degradation. Sohail et al. [64] reported a simple hydrolysis method for preparing the same composite using titanium butoxide as a precursor in ethanol which was added to the GO suspension after which the obtained precipitates were calcined at 500 °C for 2 h.

2. Photodegradation of Organic Pollutants

In the recent decade, photocatalytic degradation of persistent OMPs such as dyes, pharmaceuticals, and pesticides into non-toxic and innoxious compounds (water and carbon dioxide) has become a hot research topic. Conventional wastewater treatment techniques present significate limitations in terms of complete degradation of the harmful OMPs as well as its high cost associated with the operational conditions. Indeed, with these techniques, it is very difficult to achieve the quality of treated water to a level required by the WHO organization [85].
Many studies using rGO@TiO2 nanocomposite were made to improve photocatalytic efficiency. Liu [86] treated Methylene Orange (MO) using rGO@TiO2 nanocomposite for 240 min exposing it to visible light (λ > 400 nm) irradiation and reported ~90% photodegradation of overall organic pollutant. Several authors report the photocatalytic degradation of methylene blue (MB) using rGO@TiO2 nanocomposites, among them Deshmukh et al. [65] who got maximum degradation of MB equal to 91.3% within 30 min of sunlight irradiation. In another study by Mohammadi et al. [87], photocatalytic degradation of MB using rGO@TiO2 composite was even better. 95% and 93% of the overall organic pollutant were removed within 30 min using irradiation from a 200 W Mercury short arc and Osram 500 W Xenon lamp with a cut-off UV filter at 400 nm, respectively. Furthermore, Garrafa-Gálvez et al. [66] studied the photocatalytic degradation of MB while using the same rGO@TiO2 nanocomposites under natural sunlight and UV irradiation. They reported high effectiveness, 100% and 85% of MB of overall organic content under 60 min of UV and natural sunlight irradiation, respectively. Kusiak-Nejman et al. [88] investigated the same system but used UV-Vis light with higher UV intensity. The authors reported 91.48% of MB degradation after 60 min of UV-Vis light irradiation. Wu et al. [89] studied photocatalytic degradation of rhodamine B (RhB) using rGO@TiO2 and reported ~95% removal within 90 min of UV irradiation (250 W high-pressure Hg lamp) at room temperature. The authors also studied nanocomposite reusability; the results obtained after five cycles indicate that the prepared nanocomposite is rather stable, with above 90% removal capabilities.
Reported studies also showed that the tailoring of rGO@TiO2 nanocomposite morphology promotes improved photocatalytic efficiency for pharmaceutical and pesticide degradation. Balsamo et al. [84] reported 97% degradation of 2,4-dichlorophenoxyacetic acid (2,4-D) in 180 min using rGO@TiO2 under solar light irradiation (Figure 3). The ion-spray ionization mass spectrometric (ISI-MS) analysis was utilized to show which by-products are formed. Figure 4. displays the ISI-MS spectra of prepared 2.4-D solution before (upper) and after performed photocatalytic test (bottom) using rGO@TiO2 solar catalyst. After 3 h of continuous irradiation, 96% of removal was confirmed with the Total Organic Carbon (TOC) analysis. The ISI-MS spectra are done to investigate the formation of by-products. After 3 h of continuous irradiation, 96% of removal was detected with the Total Organic Carbon (TOC) analysis. As confirmed by the ion-spray ionization mass spectrometry (ISI-MS) 2,4-dichlorophenol (2,4-DCP) and 2,4-dichlororesorcinol (2,4-DCR) formation was observed as the main by-products.
In another study, Luna-Sanguino et al. [90] prepared a rGO@TiO2 nanocomposite that was applied for the removal of methomyl, isoproturon, and alachlor. It was found that rGO@TiO2 exhibits a high degradation rate after 25 min of natural sunlight irradiation. Deepthi et al. [67] investigated the application of synthesized rGO@TiO2 nanocomposite for solar-driven degradation of diclofenac (DCF). Besides high effectiveness (>98% DCF degradation and 65% mineralization of organic content within 60 min irradiation), the synthesized rGO@TiO2 nanocomposite catalyst mineralizes DCF completely in ~100 min. Similar efficiency can be found also for many other degradations [64,68,69,91,92].

Reusability of rGO@TiO2

The investigation of the long-term stability and reusability of prepared rGO@TiO2 photocatalyst is a crucial parameter for its practical application. The stability of prepared rGO@TiO2 photocatalyst is investigated between several consecutive cycles with the same photocatalytic tests. Wanag et al. [25] investigated the stability of prepared rGO@TiO2 photocatalyst under seven cycles. The obtained results show very high activity after five cycles. A substantial decrease in the photoactivity is noted after seventh cycles. Prepared rGO@TiO2 photocatalyst showed high stability during the photodegradation of MB dye. In another study, Balsamo et al. [84] investigated the reusability of prepared rGO@TiO2 solar catalyst under five cycles with very high photodegradation of pesticide. A slight decrease in photoactivity is attributed to the loss of catalyst powder between cycles. Deepthi et al. [67] prepared rGO@TiO2 catalyst with excellent reusability after six cycles of DCF degradation. Similar reusability of rGO@TiO2 photocatalyst has been reported in other photodegradation investigations [88,93].
According to published results on photocatalytic degradation of OMPs (dyes, pharmaceuticals, pesticides) and high reusability after several cycles, rGO-based TiO2 photocatalysts are expected to be commercialized in the near future.

3. Factors Affecting the Photodegradation of Organic Pollutants

Several parameters which affect the photodegradation process are GO to TiO2 weight ratio, catalyst loading, initial pollutant concentration, pH, water matrix composition, the intensity of irradiation, scavengers, and wavelength of light irradiation. The impact of these parameters is shown in Table 2.

3.1. The Effect of GO to TiO2 Weight Ratio

The weight ratio of GO in the rGO@TiO2 nanocomposite plays a crucial role in obtaining a high surface area. Good TiO2-decorated rGO sheets can effectively absorb irradiated light and convert it to electron-hole pairs by decreasing the bandgap. The photocatalyst with the most suitable TiO2 nanoparticles distributed on rGO sheets has the highest charge transfer rate and consequently improves pollutant degradation. Kocijan et al. [63] reported that rGO@TiO2 nanocomposite with 15 wt.% of GO shows the fastest photodegradation rate of MB and RhB dyes under simulated solar light irradiation. According to Kusiak-Nejman et al. [88], the rGO@TiO2 with 8 wt.% rGO shows a high rate of MB removal under UV light irradiation.

3.2. Effect of Catalyst Loading

The amount of prepared rGO@TiO2 photocatalyst can significantly influence the photodegradation rate of the organic pollutant from the aqueous medium. Maruthamani et al. [77] have investigated different concentrations of 20% rGO@TiO2 in the range of 0.5 to 2.0 g/L to observe the effect of photocatalyst concentration on the degradation rate of RhB (1.25 × 10−4 M) under UV irradiation. The obtained experimental results showed that the catalyst amount has both positive and negative impacts on the photodecomposition rate. The catalysts amount up to 1.5 g/L demonstrated increased photodegradation rate of RhB from 70 to 93%, which could be explained by an increased number of active sites from the catalyst [77]. On the other side, increasing the catalyst amount above, from 1.5 to 2.0 g/L, resulted in a significant drop in photodegradation of RhB. According to their agglomeration of catalyst nanoparticles coupled with increase of suspension turbidity caused light scattering that obstructed UV light penetration during the reaction [83]. Similar results can also be found elsewhere [55,63,84].

3.3. Effect of Initial Pollutant Concentration

Li et al. [93] studied the effect of initial concentration on the degradation of acid orange 7 (AO7) at different concentrations such as 10, 20, and 30 mg/L. The degradation rate decreased with increase of the initial AO7 concentration of up to 30 mg/L. Other papers also reported similar results [68,74]. The effect of initial concentration of RhB dyes on the photodegradation was studied by Maruthamani et al. [77]. They modified the initial concentration of RhB between 0.625 × 10−4 and 5.0 × 10−4 M. They found that the decolourisation percentage of RhB initially increased and later decreased with increase of initial dye concentration. Deepthi et al. [67] investigated the removal rate of diclofenac (DCF) in the initial concentration range of 10–30 mg/L. It was found that the degradation rate of DCF increased steadily for up to 25 mg/L of DCF and thereafter was stabilized or slightly decreased. Hence, as the organic pollutant concentration increases, more molecules of pollutant are adsorbed on the photocatalyst surface which reduces the surface of active sites, thus fewer photons can reach the catalyst surface. This phenomenon can significantly decrease photocatalytic degradation efficiency due to the reduced production of oxygen reactive species such as OH and O2•− radicals [94].

3.4. Effect of Initial pH

The pH value can influence the photodegradation efficiency through several possible reaction mechanisms. Parameters responsible for the changes are substrate and surface chemistry, extent of adsorption, catalyst surface charge, types of surface interactions, substrate nature, solvent molecules, the numerous intermediates formed during the progress of the reaction, etc. [68]. For instance, Deepthi et al. [67] investigated DCF degradation under natural sunlight irradiation, using different initial pH of the DCF solution. A maximum of 68.4% degradation was obtained (using the same rGO@TiO2 nanocomposite) at pH 6. The point of zero charge (PZC) of the rGO@TiO2 nanocomposite was found to be at 4.59, meaning that above this pH value the catalyst surface is negatively charged and capable of adsorbing cationic species, whereas below it is positively charged and hence attracts anionic species onto the surface. Thus, the ionic state of the substrate species is vital in determining the adsorption degree and consequently the photodegradation rate. Furthermore, it was observed that pH changes during photocatalytic reactions. According to Balsamo et al. [84], at 3.5 pH (surface with a small positive potential) rGO@TiO2 nanocomposite showed the highest rate of 2,4-D degradation under solar irradiation. Maruthamani et al. [77] investigated RhB degradation between pH 3.5 and 10.0 and noted that the highest RhB decolorization is under acidic conditions. For pH 3.5 after 3 h under UV light irradiation, the RhB decolorization was at 93%. At pH 10 and above there was also an increase in RhB photodegradation, which was attributed to the fact that the catalyst surface is sufficiently negatively charged. At pH 7 RhB decolorization value was the lowest, which was attributed to the decrease of TiOH2+ concentration. Deshmukh et al. [65] reported MB degradation using four pH values: 2.3, 4.0, 12.1, and 13.2, as shown in Figure 5. According to the obtained results, as the pH increased from 2.3 to 13.2, the amount of decolorization/degradation of MB dye increased from 17.4 to 91.3% after 30 min of reaction time. The point of zero charge (PZC) of rGO@TiO2 nanocomposite suspension was measured with the help of zeta potential values at several different pH values and is shown in Figure 5 (inset). The point of zero charge was found to be 2.62 for the ultrasonically prepared rGO@TiO2 catalyst. At lower pH, the ultrasonically prepared nanocomposite surface got positively charged which led to the protonation of active sites. MB dye is a cationic dye; thus, electrostatic repulsion is generated between the positively charged MB dye molecules and the catalyst surface. Contrarily, at higher pH, the surface of the nanocomposite becomes alkaline in nature which leads to deprotonation. In this case, there is a higher number of negative sites on the surface of the photocatalyst, leading to the interaction of the dye with one of these negative sites on the surface of the nanocomposite.

3.5. Effect of Water Matrix on the Photocatalytic Degradation of Pollutant

Usually, the photocatalytic degradation of organic pollutants throughout the literature was carried out with distilled or ultrapure water, which contains a minimum number of ions, hence also the conductivity was very low. The conductivity of ultrapure and distilled water is between 0.15 and 0.055 µS/cm and 0.5 and 3 µS/cm, respectively [95,96]. The presence of these ions may negatively or positively affect the photodegradation efficiency. For industrial scale-up, the photodegradation efficiency of the system should be tested in various natural water matrices, such as tap water, river water, lake water, seawater, industrial wastewater, etc.
Deepthi et al. [67] investigated the photocatalytic degradation of diclofenac in several water matrices (ultrapure water, river water, dug well water, filtered effluent, and unfiltered effluent) under sunlight irradiation using rGO@TiO2 nanocomposite. According to their report, they found that DCF can effectively be degraded in all investigated water systems. The photodegradation efficiency decreases in the following order: ultrapure water  >  river water  >  dug well water  >  filtered effluent  >  unfiltered effluent. Substantial photocatalytic degradation in ultrapure water was finished in 60 min, while it took 90 and 120 min for river and well water, respectively. Four hours were required for the degradation efficiency of only 90% in effluent medium. According to their analysis of the water matrices, the effluent water contained significant amounts of ions (Cl, SO42−, and PO43−). These ions inhibited the photodegradation process. As they argued, ions from effluent water were adsorbed on the catalyst surface, blocking the reactive sites. Furthermore, these ions scavenged holes and hydroxyl radicals producing less powerful oxidants such as NO3, Cl, PO42−, and CO3−.

3.6. Effect of Intensity and Wavelength of Light Irradiation

TiO2 has a wide band gap energy (3.0–3.20 eV) which limits its absorption only in the UV region of the solar spectrum. The wavelengths and intensities of UV light irradiation significantly affect the photodegradation of pollutants in an aqueous medium. UV irradiation is thus more frequently practiced than sunlight as it has higher efficiency in the degradation of pollutants. Expanding the photocatalytic degradation of pollutants to visible irradiation is an important aspect to recon with if we want to commercialize the process. Such a system should be functional under natural sunlight as the irradiation source [97]. The intensity of the light also affects the transition rate of electrons from the valence band (VB) to the conduction band (CB). Higher intensity usually leads to significantly higher degradation rates of the photocatalytic process. After saturation when the amount of photons is equal to TiO2 active sites, the rate of photogeneration becomes less dependent on the increase of the light intensity. Therefore, appropriate photon energy distribution contributes to the photodegradation rate [9]. A surplus of photons of given energy cannot contribute to a higher photocatalytic degradation rate because of the limited amounts of active sites on the surface of the catalyst [98].
Coupling TiO2 with materials such as graphene which can shift the response to visible light is crucial for obtaining an effective nanocomposite for photocatalytic applications using natural sunlight irradiation as a cost-effective irradiation source. The synergistic effect between materials in nanocomposites can improve visible light response, charge separation, and photodegradation rate. The reported investigations based on rGO@TiO2 photocatalyst with different light irradiation are shown in Table 1.

3.7. Effect of Scavengers

The photocatalytic reaction can be enhanced by increasing the number of radicals, which leads to a better separation of charge carriers. These conditions can be created on rGO material which also alters the recombination rate of charge carriers. Furthermore, free oxidative holes can directly react with the organic pollutant compounds or indirectly through the OH radicals, which are strong oxidants [99]. In an aqueous medium, active components such as hydroxyl radical (OH), holes (h+), and electrons (e) can play significant roles in the photodegradation of organic pollutants.
For instance, Mohammadi et al. [87] investigated photodegradation mechanisms of MB by rGO@TiO2 under visible light. Several scavengers were introduced. Among them were EDTA for holes (h+) generation, AgNO3 and K2S2O8 for electrons, p-benzoquinone for O2•−, t-butanol for OH radical scavenger, and NaN3 for singlet oxygen, 1O2 species. According to their results, adding t-butanol, EDTA and NaN3 had no considerable influence on the photodegradation of MB dye. They subsequently concluded that OH radicals, h+, and 1O2 do not play a significant role as active species in MB photodegradation. In contrast, adding e through introduction of K2S2O8 proved to be the main acetifier of photocatalytic degradation of MB dye. Photocatalytic degradation efficiency decreased in the presence of p-benzoquinone, superoxide radical ion, O2•−, which also represents an important finding. Pastrana-Martínez et al. [70] investigated the photodegradation of diphenhydramine under near-UV/Vis and visible light irradiation. EDTA or t-BuOH were used as hole and radical trapping agents. It was found that photogenerated holes and hydroxyl radicals are the main reactive species in the investigated process.
The photocatalytic mechanism of the rGO@TiO2 nanocomposite is shown in Figure 6. rGO in the rGO@TiO2 nanocomposite can improve photocatalytic degradation of pollutants by promoting separation and charge transfer. The electron-hole pairs have been excited within the TiO2 molecule after irradiation. The e quickly shifted from the CB of the TiO2 molecule to the surface of the rGO material. Transferred e react with dissolved oxygen resulting in the high generation of OH radicals. Furthermore, the h+ from the VB of the TiO2 molecule has the potential to oxidize H2O/OH to form OH radicals. From the oxidation and reduction process hydroxyl (OH) and superoxide radical (O2•−), are obtained respectively. They react with pollutants (MB dye) to transform pollutants into non-hazardous compounds such as H2O and CO2 [66,88,100]. Thus, because of multiple responses of rGO, a careful investigation is crucial to determine under what photocatalytic reaction conditions (such as type of irradiation exposure, pH solution, the concentration of rGO in the nanocomposite, and nature of oxidizing agent), the photodegradation mechanism of the rGO@TiO2 nanocomposites can be improved [26,101].

4. Challenges and Future Perspectives

With the growth of population and various industries, many pollutants which can have a negative influence on the environment and human health, enter the ecosystem. There is much ongoing research to partially remove pollutants from water and finding an appropriate and effective method is of vital importance. Nowadays, conventional treatments have many drawbacks such as high cost or the formation of toxic by-products. The use of catalysts represents an appealing method to remove pollutants from the water, however, there are still many costs related issues in terms of energy spending.
Advanced oxidation processes especially heterogeneous photocatalytic degradation can be utilized as an appropriate remediation strategy. Heterogeneous catalysts have many advantages over homogeneous catalysts, such as easy separation from the reaction mixture and their reusability. As mentioned above and shown in Table 1, various organic pollutants can be efficiently degraded in natural waters or wastewaters. Among them, rGO@TiO2 nanocomposite proved to be highly efficient for photodegradation of these contaminants. These novel nanocomposite materials are considered green photocatalysts, and they can be activated by solar light, which avoids pitfalls in consuming energy for the generation of UV light.
In order to realize their full potential, the research gap needs to be closed. First, optimize the integration process of rGO with TiO2 to further increase the photocatalytic activity using solar irradiation. Fine-tuning the synthesis and photocatalytic parameters is crucial for the large-scale synthesis required for technological implementation. In addition, further research is needed to determine the reversibility of the electronic properties of rGO, as its influence on the photochemical process has not yet been fully elucidated. The use of theoretical models to support empirical models could contribute to obtaining new inputs into the interfacial analysis of the rGO@TiO2 nanocomposite. Finally, future research should be devoted to investigate the photodegradation of pollutants under realistic conditions and their toxicity in contrast to many published studies under ideal experimental conditions in a short reaction time.

5. Conclusions

This review summarized the progress performed on rGO@TiO2 catalysts for the photodegradation of organic pollutants from different types of water bodies. Introduced photocatalyst nanocomposite show several advantages. Many studies report different synthesis conditions and there are new strategies for further development. rGO@TiO2 catalyst is very versatile in terms of removing different types of contaminants, but also from the applications point of view since it can be prepared as thin films, coatings, etc. Such immobilization of the rGO@TiO2 catalyst plays an important role for easy recyclability, which is important from the economic aspects. Although many efforts and studies have been made using rGO@TiO2 as a catalyst there are still some challenges that need to be addressed in the future. Among them and probably the most relevant is the conversion of this technology from a lab to a pilot scale project to finally industrialize this important application.

Author Contributions

Conceptualization, M.K., L.Ć., M.P. and G.G.; writing—original draft preparation, M.K., L.Ć., M.P. and G.G.; writing—review and editing L.Ć., G.G. and M.P. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Acknowledgments

Gil Gonçalves thank to the Portuguese Science Foundation for the Program Stimulus of Scientific Employment—Individual Support (CEECIND/01913/2017) and the financial support to TEMA (UIDB/00481/2020 and UIDP/00481/2020) and CENTRO-01-0145-FEDER-022083—Centro Portugal Regional Operational Programme (Centro2020), under the PORTUGAL 2020 Partnership Agreement, through the European Regional Development Fund. The financial support of the Slovenian Research Agency is gratefully acknowledged (Project No. L2-1830 and Program No. P2-0084).

Conflicts of Interest

The authors declare no conflict of interest.

Abbreviations

OMPsorganic micropollutants
MBmethylene blue
AOPsadvanced oxidation processes
TiO2titanium dioxide
UVultraviolet
eelectron
h+hole
GOgraphene oxide
rGOreduced graphene oxide
λlambda
HNO3nitric acid
KClO3potassium chlorate
H2SO4sulfuric acid
KMnO4potassium permanganate
NaNO3sodium nitrate
NO2nitrogen dioxide
N2O4dinitrogen tetroxide
ClO2chlorine dioxide
TTIPtitanium (IV) isopropoxide
WHOWorld Health Organization
MOmethylene orange
Visvisible
RhBrhodamine B
Hgmercury
2,4-D2,4-dichlorophenoxyacetic acid
ISI-MSion-spray ionization mass spectrometry
2,4-DCP2,4-dichlorophenol
2,4-DCRdichlororesorcinol
TOCtotal organic carbon
DCFdiclofenac
AO7acid orange 7
OHhydroxyl radical
O2•−superoxide radical
PZCpoint of zero charge
Clchloride ion
SO42−sulphate ion
PO43−phosphate ion
NO3nitrate ion
VBvalence band
CBconduction band
EDTAethylenediaminetetraacetic acid
AgNO3silver nitrate
K2S2O8potassium persulfate
NaN3sodium azide
t-BuOHtert-butyl alcohol

References

  1. Hojjati-Najafabadi, A.; Mansoorianfar, M.; Liang, T.; Shahin, K.; Karimi-Maleh, H. A review on magnetic sensors for monitoring of hazardous pollutants in water resources. Sci. Total Environ. 2022, 824, 153844. [Google Scholar] [CrossRef]
  2. Karaman, C.; Karaman, O.; Show, P.-L.; Orooji, Y.; Karimi-Maleh, H. Utilization of a double-cross-linked amino-functionalized three-dimensional graphene networks as a monolithic adsorbent for methyl orange removal: Equilibrium, kinetics, thermodynamics and artificial neural network modeling. Environ. Res. 2022, 207, 112156. [Google Scholar] [CrossRef] [PubMed]
  3. Karaman, C.; Karaman, O.; Show, P.-L.; Karimi-Maleh, H.; Zare, N. Congo red dye removal from aqueous environment by cationic surfactant modified-biomass derived carbon: Equilibrium, kinetic, and thermodynamic modeling, and forecasting via artificial neural network approach. Chemosphere 2022, 290, 133346. [Google Scholar] [CrossRef] [PubMed]
  4. Orooji, Y.; Tanhaei, B.; Ayati, A.; Tabrizi, S.H.; Alizadeh, M.; Bamoharram, F.F.; Karimi, F.; Salmanpour, S.; Rouhi, J.; Afshar, S.; et al. Heterogeneous UV-Switchable Au nanoparticles decorated tungstophosphoric acid/TiO2 for efficient photocatalytic degradation process. Chemosphere 2021, 281, 130795. [Google Scholar] [CrossRef] [PubMed]
  5. Karimi-Maleh, H.; Ayati, A.; Ghanbari, S.; Orooji, Y.; Tanhaei, B.; Karimi, F.; Alizadeh, M.; Rouhi, J.; Fu, L.; Sillanpää, M. Recent advances in removal techniques of Cr(VI) toxic ion from aqueous solution: A comprehensive review. J. Mol. Liq. 2021, 329, 115062. [Google Scholar] [CrossRef]
  6. Karimi-Maleh, H.; Ranjbari, S.; Tanhaei, B.; Ayati, A.; Orooji, Y.; Alizadeh, M.; Karimi, F.; Salmanpour, S.; Rouhi, J.; Sillanpää, M.; et al. Novel 1-butyl-3-methylimidazomohammadim bromide impregnated chitosan hydrogel beads nanostructure as an efficient nanobio-adsorbent for cationic dye removal: Kinetic study. Environ. Res. 2021, 195, 110809. [Google Scholar] [CrossRef]
  7. Yaqoob, A.A.; Parveen, T.; Umar, K.; Mohamad Ibrahim, M.N. Role of Nanomaterials in the Treatment of Wastewater: A Review. Water 2020, 12, 495. [Google Scholar] [CrossRef] [Green Version]
  8. Fernández, C.; Larrechi, M.S.; Callao, M.P. An analytical overview of processes for removing organic dyes from wastewater effluents. TrAC-Trends Anal. Chem. 2010, 29, 1202–1211. [Google Scholar] [CrossRef]
  9. Chen, D.; Cheng, Y.; Zhou, N.; Chen, P.; Wang, Y.; Li, K.; Huo, S.; Cheng, P.; Peng, P.; Zhang, R.; et al. Photocatalytic degradation of organic pollutants using TiO2-based photocatalysts: A review. J. Clean. Prod. 2020, 268, 121725. [Google Scholar] [CrossRef]
  10. Yaqoob, A.A.; Ahmad, H.; Parveen, T.; Ahmad, A.; Oves, M.; Ismail, I.M.I.; Qari, H.A.; Umar, K.; Mohamad Ibrahim, M.N. Recent Advances in Metal Decorated Nanomaterials and Their Various Biological Applications: A Review. Front. Chem. 2020, 8, 341. [Google Scholar] [CrossRef]
  11. Feng, Z.; Yuan, R.; Wang, F.; Chen, Z.; Zhou, B.; Chen, H. Preparation of magnetic biochar and its application in catalytic degradation of organic pollutants: A review. Sci. Total Environ. 2021, 765, 142673. [Google Scholar] [CrossRef]
  12. Gusain, R.; Gupta, K.; Joshi, P.; Khatri, O.P. Adsorptive removal and photocatalytic degradation of organic pollutants using metal oxides and their composites: A comprehensive review. Adv. Colloid Interface Sci. 2019, 272, 102009. [Google Scholar] [CrossRef]
  13. Adeola, A.O.; Abiodun, B.A.; Adenuga, D.O.; Nomngongo, P.N. Adsorptive and photocatalytic remediation of hazardous organic chemical pollutants in aqueous medium: A review. J. Contam. Hydrol. 2022, 248, 104019. [Google Scholar] [CrossRef]
  14. Maroudas, A.; Pandis, P.K.; Chatzopoulou, A.; Davellas, L.R.; Sourkouni, G.; Argirusis, C. Synergetic decolorization of azo dyes using ultrasounds, photocatalysis and photo-fenton reaction. Ultrason. Sonochem. 2021, 71, 105367. [Google Scholar] [CrossRef]
  15. Coha, M.; Farinelli, G.; Tiraferri, A.; Minella, M.; Vione, D. Advanced oxidation processes in the removal of organic substances from produced water: Potential, configurations, and research needs. Chem. Eng. J. 2021, 414, 128668. [Google Scholar] [CrossRef]
  16. Guerra, F.D.; Attia, M.F.; Whitehead, D.C.; Alexis, F. Nanotechnology for environmental remediation: Materials and applications. Molecules 2018, 23, 1760. [Google Scholar] [CrossRef] [Green Version]
  17. Yaqoob, A.A.; Noor, N.H.; Binti, M.; Umar, K.; Adnan, R.; Ibrahim, M.N.M.; Rashid, M. Graphene oxide–ZnO nanocomposite: An efficient visible light photocatalyst for degradation of rhodamine B. Appl. Nanosci. 2021, 11, 1291–1302. [Google Scholar] [CrossRef]
  18. Liu, Y.; Su, G.; Zhang, B.; Jiang, G.; Yan, B. Nanoparticle-based strategies for detection and remediation of environmental pollutants. Analyst 2011, 136, 872–877. [Google Scholar] [CrossRef]
  19. Abdelbasir, S.M.; Shalan, A.E. An overview of nanomaterials for industrial wastewater treatment. Korean J. Chem. Eng. 2019, 36, 1209–1225. [Google Scholar] [CrossRef]
  20. Yaqoob, A.A.; Habibah, N.; Serr, A.; Nasir, M.; Ibrahim, M. Advances and Challenges in Developing Efficient Graphene Oxide-Based ZnO Photocatalysts for Dye Photo-Oxidation. Nanomaterials 2020, 10, 932. [Google Scholar] [CrossRef]
  21. Du, X.; Luo, J.; Qin, Q.; Zhang, J.; Fu, D. Modified TiO2-rGO Binary Photo-Degradation Nanomaterials: Modification, Mechanism, and Perspective. Catal. Surv. Asia 2022, 26, 16–34. [Google Scholar] [CrossRef]
  22. Schneider, J.; Matsuoka, M.; Takeuchi, M.; Zhang, J.; Horiuchi, Y.; Anpo, M.; Bahnemann, D.W. Understanding TiO2 Photocatalysis: Mechanisms and Materials. Chem. Rev. 2014, 114, 9919–9986. [Google Scholar] [CrossRef] [PubMed]
  23. Pelaez, M.; Nolan, N.T.; Pillai, S.C.; Seery, M.K.; Falaras, P.; Kontos, A.G.; Dunlop, P.S.M.; Hamilton, J.W.J.; Byrne, J.A.; O’Shea, K.; et al. A review on the visible light active titanium dioxide photocatalysts for environmental applications. Appl. Catal. B Environ. 2012, 125, 331–349. [Google Scholar] [CrossRef] [Green Version]
  24. Ren, G.; Han, H.; Wang, Y.; Liu, S.; Zhao, J.; Meng, X.; Li, Z. Recent Advances of Photocatalytic Application in Water Treatment: A Review. Nanomaterials 2021, 11, 1804. [Google Scholar] [CrossRef] [PubMed]
  25. Wanag, A.; Kusiak-Nejman, E.; Czyżewski, A.; Moszyński, D.; Morawski, A.W. Influence of rGO and preparation method on the physicochemical and photocatalytic properties of TiO2/reduced graphene oxide photocatalysts. Catalysts 2021, 11, 1333. [Google Scholar] [CrossRef]
  26. Padmanabhan, N.T.; Thomas, N.; Louis, J.; Mathew, D.T.; Ganguly, P.; John, H.; Pillai, S.C. Graphene coupled TiO2 photocatalysts for environmental applications: A review. Chemosphere 2021, 271, 129506. [Google Scholar] [CrossRef]
  27. Tong, H.; Ouyang, S.; Bi, Y.; Umezawa, N.; Oshikiri, M.; Ye, J. Nano-photocatalytic Materials: Possibilities and Challenges. Adv. Mater. 2012, 24, 229–251. [Google Scholar] [CrossRef]
  28. Tatarchuk, T.; Peter, A.; Al-Najar, B.; Vijaya, J.; Bououdina, M. Photocatalysis: Activity of Nanomaterials. In Nanotechnology in Environmental Science; Wiley: Hoboken, NJ, USA, 2018; pp. 209–292. [Google Scholar]
  29. Xie, X.; Kretschmer, K.; Wang, G. Advances in graphene-based semiconductor photocatalysts for solar energy conversion: Fundamentals and materials engineering. Nanoscale 2015, 7, 13278–13292. [Google Scholar] [CrossRef] [Green Version]
  30. Xiang, Q.; Yu, J.; Jaroniec, M. Graphene-based semiconductor photocatalysts. Chem. Soc. Rev. 2012, 41, 782–796. [Google Scholar] [CrossRef]
  31. Allen, J.M.; Vincent, T.C.; Richard, K.B. Honeycomb carbon: A Review of Graphene What is graphene? Chem. Rev. 2010, 110, 132–145. [Google Scholar] [CrossRef]
  32. Pastrana-Martínez, L.M.; Morales-Torres, S.; Figueiredo, J.L.; Faria, J.L.; Silva, A.M.T. Graphene Derivatives in Photocatalysis. In Graphene-Based Energy Devices; Wiley: Hoboken, NJ, USA, 2015; pp. 249–276. [Google Scholar] [CrossRef]
  33. Razaq, A.; Bibi, F.; Zheng, X.; Papadakis, R.; Jafri, S.H.M.; Li, H. Review on Graphene-, Graphene Oxide-, Reduced Graphene Oxide-Based Flexible Composites: From Fabrication to Applications. Materials 2022, 15, 1012. [Google Scholar] [CrossRef]
  34. Dideikin, A.T.; Vul, A.Y. Graphene oxide and derivatives: The place in graphene family. Front. Phys. 2019, 6. [Google Scholar] [CrossRef]
  35. Bianco, A.; Cheng, H.M.; Enoki, T.; Gogotsi, Y.; Hurt, R.H.; Koratkar, N.; Kyotani, T.; Monthioux, M.; Park, C.R.; Tascon, J.M.D.; et al. All in the graphene family-A recommended nomenclature for two-dimensional carbon materials. Carbon 2013, 65, 1–6. [Google Scholar] [CrossRef]
  36. Eigler, S. Mechanistic insights into the reduction of graphene oxide addressing its surfaces. Phys. Chem. Chem. Phys. 2014, 16, 19832–19835. [Google Scholar] [CrossRef] [Green Version]
  37. Li, F.; Jiang, X.; Zhao, J.; Zhang, S. Graphene oxide: A promising nanomaterial for energy and environmental applications. Nano Energy 2015, 16, 488–515. [Google Scholar] [CrossRef] [Green Version]
  38. Guex, L.G.; Sacchi, B.; Peuvot, K.F.; Andersson, R.L.; Pourrahimi, A.M.; Ström, V.; Farris, S.; Olsson, R.T. Experimental review: Chemical reduction of graphene oxide (GO) to reduced graphene oxide (rGO) by aqueous chemistry. Nanoscale 2017, 9, 9562–9571. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  39. Agarwal, V.; Zetterlund, P.B. Strategies for reduction of graphene oxide–A comprehensive review. Chem. Eng. J. 2021, 405, 127018. [Google Scholar] [CrossRef]
  40. Oliveira, A.M.L.; Machado, M.; Silva, G.A.; Bitoque, D.B.; Ferreira, J.T.; Ferreira, Q. Graphene Oxide Thin Films with Drug Delivery Function. Nanomaterials 2022, 12, 1149. [Google Scholar] [CrossRef] [PubMed]
  41. Brodie, B.C. On the Atomic Weight of Graphite. Philos. Trans. R. Soc. Lond. 1859, 149, 249–259. [Google Scholar]
  42. Staudenmaier, L. Method for the preparation of the graphite acid. Eur. J. Inorg. Chem. 1898, 31, 1481–1487. [Google Scholar]
  43. Hummers, W.S.; Offeman, R.E. Preparation of Graphitic Oxide. J. Am. Chem. Soc. 1958, 80, 1339. [Google Scholar] [CrossRef]
  44. Al Kausor, M.; Chakrabortty, D. Graphene oxide based semiconductor photocatalysts for degradation of organic dye in waste water: A review on fabrication, performance enhancement and challenges. Inorg. Chem. Commun. 2021, 129, 108630. [Google Scholar] [CrossRef]
  45. Marcano, D.C.; Kosynkin, D.V.; Berlin, J.M.; Sinitskii, A.; Sun, Z.; Slesarev, A.; Alemany, L.B.; Lu, W.; Tour, J.M. Improved synthesis of graphene oxide. ACS Nano 2010, 4, 4806–4814. [Google Scholar] [CrossRef]
  46. Sabzevari, M.; Cree, D.; Wilson, L. Preparation and Characterization of Graphene Oxide Cross-Linked Composites. In Proceedings of the CSME Conference Proceedings, Toronto, ON, Canada, 27–30 May 2018. [Google Scholar] [CrossRef]
  47. Zaaba, N.I.; Foo, K.L.; Hashim, U.; Tan, S.J.; Liu, W.W.; Voon, C.H. Synthesis of Graphene Oxide using Modified Hummers Method: Solvent Influence. Procedia Eng. 2017, 184, 469–477. [Google Scholar] [CrossRef]
  48. Alam, S.N.; Sharma, N.; Kumar, L. Synthesis of Graphene Oxide (GO) by Modified Hummers Method and Its Thermal Reduction to Obtain Reduced Graphene Oxide (rGO). Graphene 2017, 6, 1–18. [Google Scholar] [CrossRef] [Green Version]
  49. Yu, H.; Zhang, B.; Bulin, C.; Li, R.; Xing, R. High-efficient Synthesis of Graphene Oxide Based on Improved Hummers Method. Sci. Rep. 2016, 6, 36143. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  50. Paulchamy, B.; Arthi, G.; Lignesh, B.D. A Simple Approach to Stepwise Synthesis of Graphene Oxide Nanomaterial. J. Nanomed. Nanotechnol. 2015, 6, 1–4. [Google Scholar] [CrossRef]
  51. Gao, X.; Jang, J.; Nagase, S. Hydrazine and thermal reduction of graphene oxide: Reaction mechanisms, product structures, and reaction design. J. Phys. Chem. C 2010, 114, 832–842. [Google Scholar] [CrossRef]
  52. Chua, C.K.; Pumera, M. Chemical reduction of graphene oxide: A synthetic chemistry viewpoint. Chem. Soc. Rev. 2014, 43, 291–312. [Google Scholar] [CrossRef]
  53. Sengupta, I.; Chakraborty, S.; Talukdar, M.; Pal, S.K.; Chakraborty, S. Thermal reduction of graphene oxide: How temperature influences purity. J. Mater. Res. 2018, 33, 4113–4122. [Google Scholar] [CrossRef]
  54. Anandan, S.; Manivel, A.; Ashokkumar, M. One-step sonochemical synthesis of reduced graphene oxide/Pt/Sn hybrid materials and their electrochemical properties. Fuel Cells 2012, 12, 956–962. [Google Scholar] [CrossRef]
  55. Viinikanoja, A.; Wang, Z.; Kauppila, J.; Kvarnström, C. Electrochemical reduction of graphene oxide and its in situ spectroelectrochemical characterization. Phys. Chem. Chem. Phys. 2012, 14, 14003–14009. [Google Scholar] [CrossRef]
  56. Quezada Renteria, J.A.; Ruiz-Garcia, C.; Sauvage, T.; Chazaro-Ruiz, L.F.; Rangel-Mendez, J.R.; Ania, C.O. Photochemical and electrochemical reduction of graphene oxide thin films: Tuning the nature of surface defects. Phys. Chem. Chem. Phys. 2020, 22, 20732–20743. [Google Scholar] [CrossRef] [PubMed]
  57. McAllister, M.J.; Li, J.-L.; Adamson, D.H.; Schniepp, H.C.; Abdala, A.A.; Liu, J.; Herrera-Alonso, M.; Milius, D.L.; Car, R.; Prud’homme, R.K.; et al. Single Sheet Functionalized Graphene by Oxidation and Thermal Expansion of Graphite. Chem. Mater. 2007, 19, 4396–4404. [Google Scholar] [CrossRef]
  58. Stepić, K.; Ljupković, R.; Ickovski, J.; Zarubica, A. A short review of titania-graphene oxide-based composites as a photocatalysts. Adv. Technol. 2021, 10, 51–60. [Google Scholar] [CrossRef]
  59. Das, A.; Adak, M.K.; Mahata, N.; Biswas, B. Wastewater treatment with the advent of TiO2 endowed photocatalysts and their reaction kinetics with scavenger effect. J. Mol. Liq. 2021, 338, 116479. [Google Scholar] [CrossRef]
  60. Hu, Y.; Zhou, C.; Wang, H.; Chen, M.; Zeng, G.; Liu, Z.; Liu, Y.; Wang, W.; Wu, T.; Shao, B.; et al. Recent advance of graphene/semiconductor composite nanocatalysts: Synthesis, mechanism, applications and perspectives. Chem. Eng. J. 2021, 414, 128795. [Google Scholar] [CrossRef]
  61. Preetha, S.; Pillai, R.; Ramamoorthy, S.; Mayeen, A.; Archana, K.M.; Kalarikkal, N.; Narasimhamurthy, B.; Lekshmi, I.C. TiO2–rGO Nanocomposites with high rGO content and Luminescence Quenching through Green Redox Synthesis. Surf. Interfaces 2022, 30, 101812. [Google Scholar] [CrossRef]
  62. Dutta, V.; Singh, P.; Shandilya, P.; Sharma, S.; Raizada, P.; Saini, A.K.; Gupta, V.K.; Hosseini-Bandegharaei, A.; Agarwal, S.; Rahmani-Sani, A. Review on advances in photocatalytic water disinfection utilizing graphene and graphene derivatives-based nanocomposites. J. Environ. Chem. Eng. 2019, 7, 103132. [Google Scholar] [CrossRef]
  63. Kocijan, M.; Ćurković, L.; Ljubas, D.; Mužina, K.; Bačić, I.; Radošević, T.; Podlogar, M.; Bdikin, I.; Otero-Irurueta, G.; Hortigüela, M.J.; et al. Graphene-Based TiO2 Nanocomposite for Photocatalytic Degradation of Dyes in Aqueous Solution under Solar-Like Radiation. Appl. Sci. 2021, 11, 3966. [Google Scholar] [CrossRef]
  64. Sohail, M.; Xue, H.; Jiao, Q.; Li, H.; Khan, K.; Wang, S.; Zhao, Y. Synthesis of well-dispersed TiO2@reduced graphene oxide (rGO) nanocomposites and their photocatalytic properties. Mater. Res. Bull. 2017, 90, 125–130. [Google Scholar] [CrossRef]
  65. Deshmukh, S.P.; Kale, D.P.; Kar, S.; Shirsath, S.R.; Bhanvase, B.A.; Saharan, V.K.; Sonawane, S.H. Ultrasound assisted preparation of rGO/TiO2 nanocomposite for effective photocatalytic degradation of methylene blue under sunlight. Nano-Struct. Nano-Objects 2020, 21, 100407. [Google Scholar] [CrossRef]
  66. Garrafa-Gálvez, H.E.; Alvarado-Beltrán, C.G.; Almaral-Sánchez, J.L.; Hurtado-Macías, A.; Garzon-Fontecha, A.M.; Luque, P.A.; Castro-Beltrán, A. Graphene role in improved solar photocatalytic performance of TiO2-RGO nanocomposite. Chem. Phys. 2019, 521, 35–43. [Google Scholar] [CrossRef]
  67. Deepthi, J.; Rajalakshmi, A.S.; Lopez, R.M.; Achari, V.S. TiO2-reduced graphene oxide nanocomposites for the trace removal of diclofenac. SN Appl. Sci. 2020, 2, 480. [Google Scholar] [CrossRef] [Green Version]
  68. Xu, L.; Yang, L.; Johansson, E.M.J.; Wang, Y.; Jin, P. Photocatalytic activity and mechanism of bisphenol a removal over TiO2−x/rGO nanocomposite driven by visible light. Chem. Eng. J. 2018, 350, 1043–1055. [Google Scholar] [CrossRef]
  69. Zhou, Q.; Zhong, Y.H.; Chen, X.; Huang, X.J.; Wu, Y.C. Mesoporous anatase TiO2/reduced graphene oxide nanocomposites: A simple template-free synthesis and their high photocatalytic performance. Mater. Res. Bull. 2014, 51, 244–250. [Google Scholar] [CrossRef]
  70. Pastrana-Martínez, L.M.; Morales-Torres, S.; Likodimos, V.; Figueiredo, J.L.; Faria, J.L.; Falaras, P.; Silva, A.M.T. Advanced nanostructured photocatalysts based on reduced graphene oxide-TiO2 composites for degradation of diphenhydramine pharmaceutical and methyl orange dye. Appl. Catal. B Environ. 2012, 123–124, 241–256. [Google Scholar] [CrossRef]
  71. Tolosana-Moranchel, Á.; Manassero, A.; Satuf, M.L.; Alfano, O.M.; Casas, J.A.; Bahamonde, A. Influence of TiO2-rGO optical properties on the photocatalytic activity and efficiency to photodegrade an emerging pollutant. Appl. Catal. B Environ. 2019, 246, 1–11. [Google Scholar] [CrossRef]
  72. Leal, J.F.; Cruz, S.M.A.; Almeida, B.T.A.; Esteves, V.I.; Marques, P.A.A.P.; Santos, E.B.H. TiO2-rGO nanocomposite as an efficient catalyst to photodegrade formalin in aquaculture’s waters, under solar light. Environ. Sci. Water Res. Technol. 2020, 6, 1018–1027. [Google Scholar] [CrossRef]
  73. Sun, X.; Ji, S.; Wang, M.; Dou, J.; Yang, Z.; Qiu, H.; Kou, S.; Ji, Y.; Wang, H. Fabrication of porous TiO2-rGO hybrid aerogel for high-efficiency, visible-light photodegradation of dyes. J. Alloys Compd. 2020, 819, 153033. [Google Scholar] [CrossRef]
  74. H.H. Ali, M.; E Goher, M.; D.G. Al-Afify, A. Kinetics and Adsorption Isotherm Studies of Methylene Blue Photodegradation Under UV Irradiation Using Reduced Graphene Oxide-TiO2 Nanocomposite in Different Wastewaters Effluents. Egypt. J. Aquat. Biol. Fish. 2019, 23, 253–263. [Google Scholar] [CrossRef] [Green Version]
  75. Zhang, Y.; Fu, F.; Zhou, F.; Yang, X.; Zhang, D.; Chen, Y. Synergistic effect of rGO/TiO2 nanosheets with exposed (001) facets for boosting visible light photocatalytic activity. Appl. Surf. Sci. 2020, 510, 145451. [Google Scholar] [CrossRef]
  76. Duan, Y.; Gou, M.L.; Guo, Y.; Cai, J.; Song, W.; Liu, Z.; Zhou, E. In situ hydrothermal synthesis of TiO2–rGO nanocomposites for 4-nitrophenol degradation under sunlight irradiation. J. Mater. Res. 2021, 36, 906–915. [Google Scholar] [CrossRef]
  77. Maruthamani, D.; Divakar, D.; Kumaravel, M. Enhanced photocatalytic activity of TiO2 by reduced graphene oxide in mineralization of Rhodamine B dye. J. Ind. Eng. Chem. 2015, 30, 33–43. [Google Scholar] [CrossRef]
  78. Abd Elnaby, M.; Ahmed, M.A.; Abdel-Samad, H.S.; Hesham Medien, M. Exceptional removal of methylene blue dye over novel TiO2/rGO nanocomposites by tandem adsorption-photocatalytic processes. Mater. Sci. Energy Technol. 2022, 5, 217–231. [Google Scholar] [CrossRef]
  79. Fan, H.; Yi, G.; Zhang, Z.; Zhang, X.; Li, P.; Zhang, C.; Chen, L.; Zhang, Y.; Sun, Q. Binary TiO2/rGO photocatalyst for enhanced degradation of phenol and its application in underground coal gasification wastewater treatment. Opt. Mater. 2021, 120, 111482. [Google Scholar] [CrossRef]
  80. Zhou, X.; Zhou, S.; Ma, F.; Xu, Y. Synergistic effects and kinetics of rGO-modified TiO2 nanocomposite on adsorption and photocatalytic degradation of humic acid. J. Environ. Manag. 2019, 235, 293–302. [Google Scholar] [CrossRef]
  81. Sun, N.; Ma, J.; Wang, C.; Xue, J.; Qiang, L.; Tang, J. A facile and efficient method to directly synthesize TiO2/rGO with enhanced photocatalytic performance. Superlattices Microstruct. 2018, 121, 1–8. [Google Scholar] [CrossRef]
  82. Ruidíaz-Martínez, M.; Álvarez, M.A.; López-Ramón, M.V.; Cruz-Quesada, G.; Rivera-Utrilla, J.; Sánchez-Polo, M. Hydrothermal Synthesis of rGO-TiO2 Composites as High-Performance UV Photocatalysts for Ethylparaben Degradation. Catalysts 2020, 10, 520. [Google Scholar] [CrossRef]
  83. Yu, F.; Bai, X.; Yang, C.; Xu, L.; Ma, J. Reduced Graphene Oxide–P25 Nanocomposites as Efficient Photocatalysts for Degradation of Bisphenol A in Water. Catalysts 2019, 9, 607. [Google Scholar] [CrossRef]
  84. Balsamo, S.A.; Fiorenza, R.; Condorelli, M.; Pecoraro, R.; Brundo, M.V.; Presti, F.L.; Scir, S. One-Pot Synthesis of TiO2-rGO Photocatalysts for the Degradation of Groundwater Pollutants. Materials 2021, 14, 5938. [Google Scholar] [CrossRef]
  85. Sohail, M.; Saleem, M.; Ullah, S.; Saeed, N.; Afridi, A.; Khan, M.; Arif, M. Modified and improved Hummer’s synthesis of graphene oxide for capacitors applications. Mod. Electron. Mater. 2017, 3, 110–116. [Google Scholar] [CrossRef]
  86. Liu, Y. Hydrothermal synthesis of TiO2-RGO composites and their improved photocatalytic activity in visible light. RSC Adv. 2014, 4, 36040–36045. [Google Scholar] [CrossRef]
  87. Mohammadi, M.; Rezaee Roknabadi, M.; Behdani, M.; Kompany, A. Enhancement of visible and UV light photocatalytic activity of rGO-TiO2 nanocomposites: The effect of TiO2/Graphene oxide weight ratio. Ceram. Int. 2019, 45, 12625–12634. [Google Scholar] [CrossRef]
  88. Kusiak-Nejman, E.; Wanag, A.; Kapica- Kozar, J.; Kowalczyk, Ł.; Zgrzebnicki, M.; Tryba, B.; Przepiórski, J.; Morawski, A.W. Methylene blue decomposition on TiO2/reduced graphene oxide hybrid photocatalysts obtained by a two-step hydrothermal and calcination synthesis. Catal. Today 2019, 357, 630–637. [Google Scholar] [CrossRef]
  89. Wu, S.; Jia, Q.; Dai, W. Synthesis of rGO/TiO2 hybrid as a high performance photocatalyst. Ceram. Int. 2017, 43, 1530–1535. [Google Scholar] [CrossRef]
  90. Luna-Sanguino, G.; Ruíz-Delgado, A.; Tolosana-Moranchel, A.; Pascual, L.; Malato, S.; Bahamonde, A.; Faraldos, M. Solar photocatalytic degradation of pesticides over TiO2-rGO nanocomposites at pilot plant scale. Sci. Total Environ. 2020, 737, 140286. [Google Scholar] [CrossRef]
  91. Ton, N.N.T.; Dao, A.T.N.; Kato, K.; Ikenaga, T.; Trinh, D.X.; Taniike, T. One-pot synthesis of TiO2/graphene nanocomposites for excellent visible light photocatalysis based on chemical exfoliation method. Carbon 2018, 133, 109–117. [Google Scholar] [CrossRef]
  92. Liang, D.; Cui, C.; Hub, H.; Wang, Y.; Xu, S.; Ying, B.; Li, P.; Lu, B.; Shen, H. One-step hydrothermal synthesis of anatase TiO2/reduced graphene oxide nanocomposites with enhanced photocatalytic activity. J. Alloys Compd. 2014, 582, 236–240. [Google Scholar] [CrossRef]
  93. Li, T.; Wang, T.; Qu, G.; Liang, D.; Hu, S. Synthesis and photocatalytic performance of reduced graphene oxide–TiO2 nanocomposites for orange II degradation under UV light irradiation. Environ. Sci. Pollut. Res. 2017, 24, 12416–12425. [Google Scholar] [CrossRef]
  94. Zhao, D.; Sheng, G.; Chen, C.; Wang, X. Enhanced photocatalytic degradation of methylene blue under visible irradiation on graphene@TiO2 dyade structure. Appl. Catal. B Environ. 2012, 111–112, 303–308. [Google Scholar] [CrossRef]
  95. Zhang, X.; Yang, Y.; Ngo, H.H.; Guo, W.; Wen, H.; Wang, X.; Zhang, J.; Long, T. A critical review on challenges and trend of ultrapure water production process. Sci. Total Environ. 2021, 785, 147254. [Google Scholar] [CrossRef] [PubMed]
  96. Díez, A.M.; Pazos, M.; Sanromán, M.A. Synthesis of magnetic-photo-Fenton catalyst for degradation of emerging pollutant. Catal. Today 2019, 328, 267–273. [Google Scholar] [CrossRef]
  97. Reza, K.M.; Kurny, A.; Gulshan, F. Parameters affecting the photocatalytic degradation of dyes using TiO2: A review. Appl. Water Sci. 2017, 7, 1569–1578. [Google Scholar] [CrossRef] [Green Version]
  98. Cheng, H.Y.; Chang, K.C.; Lin, K.L.; Ma, C.M. Study on isopropanol degradation by UV/TiO2 nanotube. AIP Conf. Proc. 2018, 1946, 020006. [Google Scholar] [CrossRef]
  99. Kocijan, M.; Ćurković, L.; Bdikin, I.; Otero-Irurueta, G.; Hortigüela, M.J.; Gonçalves, G.; Radošević, T.; Vengust, D.; Podlogar, M. Immobilised rGO/TiO2 Nanocomposite for Multi-Cycle Removal of Methylene Blue Dye from an Aqueous Medium. Appl. Sci. 2022, 12, 385. [Google Scholar] [CrossRef]
  100. Muthirulan, P.; Devi, C.N.; Sundaram, M.M. TiO2 wrapped graphene as a high performance photocatalyst for acid orange 7 dye degradation under solar/UV light irradiations. Ceram. Int. 2014, 40, 5945–5957. [Google Scholar] [CrossRef]
  101. Mondal, A.; Prabhakaran, A.; Gupta, S.; Subramanian, V.R. Boosting Photocatalytic Activity Using Reduced Graphene Oxide (rGO)/Semiconductor Nanocomposites: Issues and Future Scope. ACS Omega 2021, 6, 8734–8743. [Google Scholar] [CrossRef]
Figure 1. Photocatalytic reaction mechanism on TiO2 semiconductor material.
Figure 1. Photocatalytic reaction mechanism on TiO2 semiconductor material.
Sustainability 14 12703 g001
Figure 2. Graphical representation of the molecular structures of graphene material and its derivates [40].
Figure 2. Graphical representation of the molecular structures of graphene material and its derivates [40].
Sustainability 14 12703 g002
Figure 3. Photodegradation of 2,4-D [84].
Figure 3. Photodegradation of 2,4-D [84].
Sustainability 14 12703 g003
Figure 4. ISI-MS spectra were acquired before the irradiation (upper spectrum) and after 3 h of irradiation using the rGO@TiO2 solar sample (bottom spectrum) [84].
Figure 4. ISI-MS spectra were acquired before the irradiation (upper spectrum) and after 3 h of irradiation using the rGO@TiO2 solar sample (bottom spectrum) [84].
Sustainability 14 12703 g004
Figure 5. Effect of pH on degradation of MB dye using rGO@TiO2 nanocomposite. Photocatalyst and zeta potential values at different pH of rGO/TiO2 suspension (inset). Reprinted with permission from ref. [65]. 2022, Copyright Elsevier.
Figure 5. Effect of pH on degradation of MB dye using rGO@TiO2 nanocomposite. Photocatalyst and zeta potential values at different pH of rGO/TiO2 suspension (inset). Reprinted with permission from ref. [65]. 2022, Copyright Elsevier.
Sustainability 14 12703 g005
Figure 6. The photocatalytic mechanisms of rGO@TiO2 nanocomposite for MB dye degradation [99].
Figure 6. The photocatalytic mechanisms of rGO@TiO2 nanocomposite for MB dye degradation [99].
Sustainability 14 12703 g006
Table 1. Photocatalytic degradation of organic pollutants by rGO@TiO2 nanocomposite.
Table 1. Photocatalytic degradation of organic pollutants by rGO@TiO2 nanocomposite.
Synthesis
Method
Catalyst
Dosage
Light SourcePollutant and Initial
Concentration
Irradiation TimeKinetic
/10−3
Removal
Efficiency
Reference
Solvothermal0.8 g/LUV lightMethylene blue,
10 mg/L
90 min-100.0%[25]
Sol-gel method0.5 g/LNatural
sunlight
Alizarin Yellow GG,
10 mg/L
150 min-100.0%[61]
Hydrothermal0.5 g/LSimulated solar lightRhodamine B,
10 mg/L
120 min53.1 min−199.8%[63]
Hydrothermal0.5 g/LSimulated solar lightMethylene blue,
10 mg/L
120 min32.84 min−198.1%[63]
Sonochemical2.0 g/LNatural
sunlight
Methylene blue,
20 mg/L
30 min18.8 min−191.3%[65]
Solvothermal0.1 g/LNatural
sunlight
Methylene blue, 15 mg/L60 min21.2 min−185.0%[66]
Solvothermal0.1 g/LUV lampMethylene blue, 15 mg/L60 min89.1 min−1100.0%[66]
Hydrothermal0.075 g/LNatural
sunlight
Diclofenac,
25 mg/L
60 min50.4 min−1~100.0%[67]
Hydrothermal0.1 g/LUV lightMethylene blue,
12 mg/L
180 min13.1 min−1~90.0%[68]
Hydrothermal0.0004 g/LUV lightMethyl orange,
30 mg/L
30 min122.0 min−195.0%[69]
Hydrothermal0.5 g/LVisible
light
Methyl orange,
10 mg/L
240 min126.0 min−187.4%[70]
Hydrothermal0.1 g/LUV-A lampClofibric acid,
20 mg/L
360 min11.2 min−1100.0%[71]
Hydrothermal0.18 g/LSimulated sunlightFormalin,
40 mg/L
90 min17.9 min−193.8%[72]
Freeze-drying0.4 g/LSimulated solar lampRhodamine B,
10 mg/L
240 min-84.6%[73]
Ultrasonication0.75 g/LUV light sourceMethylene blue, 30 mg/L120 min-99.3%[74]
Solvothermal0.1 g/LVisible lightMethyl orange,
10 mg/L
180 min-94.1%[75]
Hydrothermal0.25 g/LSimulated sunlight4-nitrophenol,
45 mg/L
360 min11.8 min−195.0%[76]
Solvothermal1.0 g/LUV lightRhodamine B,
0.125 mM
300 min6.3 min−194.8%[77]
Hydrothermal0.5 g/LUV-A lampMethylene blue,
6.4 mg/L
60 min23.0 min−186.0%[78]
Hydrothermal0.5 g/LNatural
sunlight
Methylene blue,
6.4 mg/L
60 min20.0 min−181.0%[78]
Sol-anaerobic calcination0.4 g/LUV lightPhenol,
10 mg/L
180 min-97.9%[79]
Hydrothermal1.2 g/LVisible lightHumic acid,
30 mg/L
180 min-88.0%[80]
Hydrothermal0.5 g/LVisible lightMethylene blue,
0.01 mM
130 min37.4 min−1~99.0%[81]
Hydrothermal0.7 g/LUV lightEthylparaben,
0.3 mM
40 min96.0 min−198.6%[82]
Hydrothermal0.5 g/LNatural
sunlight
Bisphenol A,
0.01 mol/L
30 min-100.0%[83]
Hydro-solvothermal1.0 g/LSolar light2,4-Dichlorophenoxyacetic acid, 0.5 mM180 min9.6 min−197.0%[84]
Table 2. The impact of parameters on the rGO@TiO2 photocatalytic process.
Table 2. The impact of parameters on the rGO@TiO2 photocatalytic process.
ParameterImpact on Photocatalysis
Catalyst concentrationThe photodegradation rate increases with the increase of the amount of catalyst.
Above a certain amount of catalyst, the photodegradation rate decreases as the catalyst amount increases.
Initial pollutant concentrationThe photodegradation rate increases with the increase of the initial concentration of pollutant.
Above a certain initial concentration of pollutant, the photodegradation rate decreases as the initial concentration of pollutant increases.
pHThe photodegradation rate depends significantly on the pH value.
Light sourceLight source supplies irradiation of different wavelengths (UV-C, UV-B, UV-A, visible light, simulated solar light, natural sunlight).
Intensity of irradiationLight intensity depends on the light source and enhances the photocatalytic reaction.
Water matrixPresence of different pollutants/organic matter that can act as inhibitors or competitors of the photodegradation rate.
Weight ratio of graphene oxide to TiO2 The photodegradation rate increases with the increasing amount of GO to TiO2.
Above a certain threshold the photodegradation rate decreases as the amount of GO increases.
ScavengersImprove the separation rate of photogenerated holes and electrons and enhance the photocatalytic reaction.
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Kocijan, M.; Ćurković, L.; Gonçalves, G.; Podlogar, M. The Potential of rGO@TiO2 Photocatalyst for the Degradation of Organic Pollutants in Water. Sustainability 2022, 14, 12703. https://doi.org/10.3390/su141912703

AMA Style

Kocijan M, Ćurković L, Gonçalves G, Podlogar M. The Potential of rGO@TiO2 Photocatalyst for the Degradation of Organic Pollutants in Water. Sustainability. 2022; 14(19):12703. https://doi.org/10.3390/su141912703

Chicago/Turabian Style

Kocijan, Martina, Lidija Ćurković, Gil Gonçalves, and Matejka Podlogar. 2022. "The Potential of rGO@TiO2 Photocatalyst for the Degradation of Organic Pollutants in Water" Sustainability 14, no. 19: 12703. https://doi.org/10.3390/su141912703

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