Next Article in Journal
Synthesis of SAPO-34 Nanoplates with High Si/Al Ratio and Improved Acid Site Density
Next Article in Special Issue
Manifestations of Laser-Induced Graphene under Ultraviolet Irradiation of Polyimide with Varied Optical Fluence
Previous Article in Journal
Comparison of the Morphological and Structural Characteristic of Bioresorbable and Biocompatible Hydroxyapatite-Loaded Biopolymer Composites
Previous Article in Special Issue
Photoinitiated Polymerization of Hydrogels by Graphene Quantum Dots
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Graphene Family Nanomaterials (GFN)-TiO2 for the Photocatalytic Removal of Water and Air Pollutants: Synthesis, Characterization, and Applications

by
Chih-Hsien Lin
1 and
Wei-Hsiang Chen
1,2,3,*
1
Institute of Environmental Engineering, National Sun Yat-sen University, Kaohsiung 804, Taiwan
2
Aerosol Science and Research Center, National Sun Yat-sen University, Kaohsiung 804, Taiwan
3
Department of Public Health, Kaohsiung Medical University, Kaohsiung 807, Taiwan
*
Author to whom correspondence should be addressed.
Nanomaterials 2021, 11(12), 3195; https://doi.org/10.3390/nano11123195
Submission received: 23 October 2021 / Revised: 19 November 2021 / Accepted: 21 November 2021 / Published: 25 November 2021
(This article belongs to the Special Issue Synthesis, Modification and Application of Graphene)

Abstract

:
Given the industrial revolutions and resource scarcity, the development of green technologies which aims to conserve resources and reduce the negative impacts of technology on the environment has become a critical issue of concern. One example is heterogeneous photocatalytic degradation. Titanium dioxide (TiO2) has been intensively researched given its low toxicity and photocatalytic effects under ultraviolet (UV) light irradiation. The advantages conferred by the physical and electrochemical properties of graphene family nanomaterials (GFN) have contributed to the combination of GFN and TiO2 as well as the current variety of GFN-TiO2 catalysts that have exhibited improved characteristics such as greater electron transfer and narrower bandgaps for more potential applications, including those under visible light irradiation. In this review, points of view on the intrinsic properties of TiO2, GFNs (pristine graphene, graphene oxide (GO), reduced GO, and graphene quantum dots (GQDs)), and GFN-TiO2 are presented. This review also explains practical synthesis techniques along with perspective characteristics of these TiO2- and/or graphene-based materials. The enhancement of the photocatalytic activity by using GFN-TiO2 and its improved photocatalytic reactions for the treatment of organic, inorganic, and biological pollutants in water and air phases are reported. It is expected that this review can provide insights into the key to optimizing the photocatalytic activity of GFN-TiO2 and possible directions for future development in these fields.

1. Introduction

A circular economy promises a comprehensive solution to resource efficiency given the concern of non-renewable energy scarcity. Besides raw materials, energy sources such as renewables are becoming increasingly viable alternatives to fossil fuels. These factors have combined to increase the research activity into the circular economy and renewable resources, as shown in Figure 1. To date, it is estimated that more than 10,000 studies associated with renewables have been reported, as part of these have focused on their applications in the fields of pollution prevention and control. For instance, ethanol is currently known as an alternative fuel from agricultural, industrial and urban residues [1,2]. Electrochemical technologies replace or reduce hazardous materials used in conventional chemical treatment processes [3,4]. Among these discussions, photocatalysis demands an approach associated with the intermittent nature of sunlight, which is considered a renewable energy source, or the assistance of ultraviolet (UV) light. The recognition of the interesting fact, including chemical reactions enabled and/or powered by the energy delivered by photons, relatively higher reaction rates with lower energy requirement, and repeatability of catalysts, has inspired generations of scientists to develop technologies more efficient and less costly to meet the needs for environmental treatment and remediation.
Titanium dioxide (TiO2) is one of the typical catalysts that has been used for pollution control [5]. Graphene is a two-dimensional carbon allotrope with premium thermal and electrical properties [6,7]. As combining TiO2 with graphene has shown promising results in the research of photocatalysis, our goal in writing this review is to provide a broad overview of this field. We aim to offer a historical account of the development and uses of TiO2 and graphene family nanomaterials (GFN) for photocatalysis. Most importantly of all, we hope to unify the discussion of these two materials and provide readers an opportunity to focus on the critical scientific merits of combining TiO2 and GFN for enhanced photocatalysis. The article is organized in the following ways. We start with the basic principles that govern photocatalysis and then move on to introduce the historical views of TiO2 and GFNs. Their emergences and preparation methods are summarized, followed by a discussion on their physical and electrochemical properties. Afterward, the review examines the integration of TiO2 and GFNs as an emerging photocatalyst for the treatment of different water and air pollutants that have been reported in the studies. At the end of this review, we present our perspectives on where the research field of this integrated photocatalysis could be headed.

2. TiO2

2.1. Background

TiO2 is a naturally occurring oxide of titanium with structural stability and corrosion resistance [8,9,10]. Although TiO2 is typically considered to be of low toxicity, the development of TiO2 nanotechnologies has resulted in increased human and environmental exposure, putting TiO2 nanoparticles under toxicological scrutiny. Evidence in experimental animals for the carcinogenicity of TiO2 has been reported [11]. The International Agency for Research on Cancer (IARC) has indicated that TiO2 is possibly carcinogenic to humans [12]. Fujishima and Honda (1972) first discovered UV-light induced electrocatalysis for the splitting of water by using TiO2 as a photoanode in an electrochemical cell [13]. Frank and Bard (1977) reported the heterogeneous photocatalytic oxidation of cyanide in water using TiO2 powder [14,15]. Since then, photocatalysis using TiO2 has achieved a burst of interest to researchers due to the potential implications in the fields of environmental treatment and pollution control [16,17]. TiO2 is commonly present in the structures of anatase, brookite, and rutile [18]. Although rutile is the most abundant form of TiO2 with thermal stability [19], anatase TiO2 has improved photosensitive properties due to its excellent charge-carrier mobility and a greater number of surface hydroxyl groups [20]. To date, TiO2-based photocatalysis has become a viable technology for various purposes, including treatment of a wide range of environmental pollutants and eco-friendly green processes of organic synthesis.

2.2. Photocatalysis

Photocatalysis occurs by utilizing light and semiconductors as the substrate [21], as illustrated in Figure 2. The substrate absorbs light and alters the rate of a chemical reaction. In this phenomenon, when a substrate adsorbs photons with the energy exceeding the bandgap energy, an electron-hole (e−h+) pair is formed by exciting electrons from the valence band to the conduction band. The existence of the valence band holes (hVB+) and conduction band electrons (eCB) is typically transient and rapidly removed by recombination with heat or light emission. For certain materials, namely photocatalysts, the lifetime of the e−h+ pair is extended, allowing a fraction of the e−h+ pairs to migrate through the substrate to the surface, performing redox reactions in the surrounding medium [22,23]. The hVB+ can oxidize water and hydroxyl anions to generate hydroxyl radicals (OH), while dissolved oxygen can be reduced by the eCB, leading to the formation of superoxide radical anions (O2) or hydroperoxyl radicals (OOH) with further protonation [24]. These strong oxidizing radical species allow the degradation or mineralization of pollutants in the environment upon the exposure of a photocatalyst to light.

2.3. Synthesis

The structural, electronic, and optical properties of TiO2 are affected by using materials with different sizes, shapes, or phases for synthesis. However, the method used for synthesis is the key to determining the TiO2 product characteristics. To date, TiO2 is synthesized by using methods including the sol-gel method, micelle and inverse micelle method, sol method, hydrothermal method, solvothermal method, direct oxidation method, chemical vapor deposition, physical vapor deposition, electrodeposition, sonochemical method, flame pyrolysis, and microwave method [25]. Table 1 lists the methods that are widely used and provides more detailed descriptions. The sol-gel method is one of the most commonly used approaches. This approach produces TiO2 particles with high crystallinity, limited agglomeration, as well as good size distribution and dispersity. Additionally, the formation of rutile can be controlled by temperature adjustment in this procedure, as anatase materials are effectively obtained at low temperatures.

2.4. Properties between Different Polymorphs

TiO2 is typically recognized to occur in three different polymorphs, including rutile, anatase, and brookite. The latter is rarely used as a catalyst because it is difficult to synthesize. The photocatalytic activities of rutile and anatase TiO2 are dependent upon the crystal structure, size distribution, surface area, pore structure, etc. Despite its low bandgap (Table 2), the lower photocatalytic activity of rutile TiO2 is correlated to the intrinsic recombination of photogenerated e−h+ pairs [35]. It has been reported that the bulk transport of excitons to the surface contributed to the different photocatalytic activities between the rutile and anatase TiO2, as charger carriers excited deeper in the bulk contribute to more efficient photocatalysis in anatase than in rutile [36]. Furthermore, compared to the rutile structure, the photocatalytic activity of anatase TiO2 is improved by its smaller particle size [37], higher surface area [38], and more importantly, higher surface-adsorbed hydroxyl radicals and the slower so-called photoinduced charge-carrier recombination in anatase relative to rutile [39,40]. The increased lifetime of the e--h+ pair can predominate over the charge-hole recombination process. The lower effective mass of the photogenerated charge carrier can increase the mobility of electron transfer. These characteristics enhance the photocatalytic activity of the crystalize anatase, thereby making it the most active catalyst compared to rutile and brookite. Table 2 compares the properties imperative to TiO2 in its anatase, rutile, and brookite crystalline phases.
Although the discussion above promotes the use of the anatase as the catalyst of preference compared to rutile, a larger intrinsic bandgap of anatase TiO2 (3.2 and 3.0 eV for anatase and rutile structures, respectively) only allows a smaller portion of the solar spectrum in the UV light region to be adsorbed, thereby negatively affecting the applicability of this technology. One solution is the doping of different ions that contributes to the improved activities of TiO2 in different ways. For example, doping with Fe or Zn improves the conductivity of TiO2 and the mobility of charge carriers, slowing recombination and more efficiently separating photogenerated electrons and holes [49]. Recently, the non-metal doping of C has been extensively investigated due to its improved response to visible light and high photostability. The replacement of O in the TiO2 lattice with C narrows the bandgap and promotes the adsorption of the main region of the solar spectrum. Furthermore, impurity states formed near the valence band edge along with C-doping can act as shallow traps and extend the occurrence of photogenerated electron-hole pairs [49,50], as graphene represents one emerging material increasingly used for this purpose.

3. Graphene Family Nanomaterials (GFN)

3.1. Graphene and Its Derivatives

Since its successful extraction from graphite in 2004 [51], research with this material stems from its exceptional electrical, mechanical, and optical properties and the potential applications employing these properties. GFN includes graphene oxide (GO), reduced GO (rGO), and graphene quantum dots (GQDs) [52], as illustrated in Figure 3. Graphene is a two-dimensional carbon allotrope, as the sp2 hybridization results in the extreme of such properties including high conductivity, remarkable optical features, and mechanical strength along two dimensions [51,53]. GO is the sheet of a defective sp2 carbon network that incorporates oxygenated groups, including hydroxyl, epoxy, carbonyl, and carboxyl groups, at the interior and on the edge [7,54]. These groups tend to change the surface properties of GO from being hydrophobic to hydrophilic. By reducing the oxygen content and generating different defects in GO, a material with intermediate features between pristine graphene and GO, namely reduced or partially rGO, is produced [55,56]. Different methodologies applied for GO reduction affect the types and numbers of defects and thereby the chemical properties of the final product. GQDs that contain both sp2 and sp3 hybridizations are separated from the single to several layers of graphene sheets to several nm in lateral size [57]. The features of GQDs include the size-dependent optical band gap, high electron mobility, excellent solubility, and easy functionalization.

3.2. Synthesis

Particular emphasis is directed toward the effects of different synthesis methods on the properties of GFN products and the characterization imperative to determine the quality of the synthesis [53,58]. Different synthesis methods and operational factors are known to change the distances between the layers (d-spacing), layer number, stacking order, and structure completeness, which further influences the quality of GFNs. For example, Wu et al. have revealed that the number of graphene layers was effectively tuned by selecting suitable starting materials in the chemical exfoliation method [59]. Artificial graphite, flake graphite powder, Kish graphite, and natural flake graphite were used as starting materials to produce single-layer, single- and double-layer, double- and triple-layer, and few-layer (4–10 layers) graphene final products, respectively. Table 3 discusses the GFNs, including graphene, GO, rGO, and GQD, prepared by different synthesis methods and the associated pros and cons.
Graphene is typically prepared by using mechanical exfoliation, oxidative exfoliation-reduction (OER), liquid-phase exfoliation (LPE), and chemical vapor deposition, as listed in Table 3. Other emerging methods include arc plasmas [60], unzipping of carbon nanotubes [61], epitaxial graphene growth [62], substrate-free gas-phase synthesis (SFGP) [63], the soft-hard template approach [64], and total organic synthesis [65]. Lee et al. [58] evaluated aspects of product quality, process safety and complexity, yield efficiency, environmental impacts, cost-effectiveness, and scalability among different approaches for graphene synthesis (Figure 4). The popularity of the OER and LPE are explained by their relatively higher scores in each category of comparison.
The synthetic methods of GO were continuously modified in recent decades. The methods that are widely used and frequently discussed include the Brodie method [66], Staudenmaier method [67], Hofmann method [68], and Hummers method [69], as listed in Table 3. The Hummers is typically recognized as one popular method for its efficiency, safety, effective oxidation and crystallinity, and scalable production of large-area and high-quality products. Recently modified Hummers approaches that are more environmentally friendly have emerged as one of the most popular methods for GO production for different purposes. For example, hazardous chemicals such as NaNO3 used in conventional Hummers methods that form toxic NO2/N2O4 gases were replaced without a yield decrease in an improved Hummers method [70,71]. Zaaba et al. improved the method by carrying it out at room temperature and without NaNO3 [72]. The chemical recipe of the Hummers method was adjusted (e.g., the increase of KMnO4 used and change of the H2SO4/H3PO4 mixing ratio) to enhance the efficiency of the oxidation process [73]. A different oxidant, K2FeO4, was studied for its potential to reduce the formation of toxic gases, to enable the recycling of sulfuric acid, and to increase the reaction efficiency [74].
rGO is typically processed by chemical, thermal, and other methods [75,76]. Chemical reduction is commonly used given its merits of fine product quality and scalable production [77,78,79,80]. Thermal reduction is another method for rGO production. These processes are straightforward and cost-effective. However, the needs of certain hazardous reductants or capital costs and energy in chemical and thermal reduction, respectively, resulted in the rise of other emerging methods such as electrochemical reduction [81], microwave and thermal reduction [82]. These technologies provide alternatives with high yield efficiencies, fine product qualities, and the potential for green chemistry.
GQDs are one of a few layers of graphene with a size smaller than 100 nm [83]. GQDs exploit the intrinsic characteristics of graphene nanomaterials and increase their applications with their enhanced and tunable photoluminescence, unique photo-induced redox properties, and high biocompatibility [84,85]. The preparation methods of GQDs include the “top-down” and “bottom-up” methods. The “top-down” methods, which mainly prepare GQDs by chemically, electrochemically, or physically cutting the crystallites of graphene, include hydrothermal synthesis [86], the solvent thermal method [87], chemical oxidation [88], electrochemical exfoliation [89,90,91,92], electron beam lithography [93], the microwave-assisted method [94], and ultra-sonication exfoliation [95,96]. The “bottom-up” methods offer new strategies to fabricate GQDs by pyrolysis of small organic compounds or by chemical fusion of small aromatic compounds. These methods include the soft template method [97], acid and solvent-free synthesis [98], and metal catalysis [99]. Considering the need for hours to develop low-dimensional GQD, the manufacturing of GQDs for industrial-scale applications is still being investigated and increasingly discussed.

3.3. Properties

Table 4 lists some properties of GFNs that have been reported in the studies. Graphene has aroused wide attention because of its unique electronic, optical, thermal, and mechanical properties, as the properties of derivatives were known to be changed by different functional groups, structural defects, and stacking layers and sizes. Graphene displays ultrahigh mobility of electrons (e.g., 15,000 cm2 v−2 s−1) which depends weakly on temperature, remarkable mechanical strength (Young’s modulus of 1.0 TPa and fracture strength of 130 GPa), high-frequency optical conductivity from the infrared through the visible range of the spectrum, high thermal conductivity (~4000 Wm−1 K−1), and the capability of easily converting electrical currents to heat [53].
Although the conjugated regions of GO that are partially destructed by the oxygen-containing functional groups negatively affected its electrical mobility and mechanical strength (the average elastic modulus was 32 GPa, while the highest fracture strength was 120 MPa), GO is stable in water, and this property has provided opportunities for possible applications in solutions [100,101]. The introduction of chemicals such as divalent polyallylamine or metal ions that cross-link between GO layers has improved the mechanical properties of GO [53,102].
Table 4. Properties of GFNs that have been reported in studies.
Table 4. Properties of GFNs that have been reported in studies.
PropertiesGrapheneGOrGOGQD
Functional groupNo functional groupEpoxy, carboxyl, hydroxyl, and carboxylEpoxy, carboxyl, and hydroxyl Epoxy, carbonyl, hydroxyl, and carboxyl
Nature Hydrophobic Hydrophilic Hydrophilic -
C:O ratio No oxygen 2-4 8-246 3
d-spacing (nm) 0.335 0.737 0.368 0.381
Surface area (m2/g) 2600 487 466 -
Electron mobility (cm2V/s) 10,000–50,000 Insulator 0.05–200 -
Resistance (Ω) 7200 0.514±0.236 2.01 ± 1.6 -
Optics 2.3% absorption(visible light) - ~20% adsorption (400–1800 nm) -
Thermal conductivity (W/m·K) ~5000 2.94 61.8 -
Zeta potential (mV) - −33~−21.46 −23.5~−26.5 8
Young’s modulus 1 0.2 0.25 -
Reference[79,103,104,105,106,107,108][103,104,109,110,111,112][77,103,104,110,113,114,115][103,104,113,116,117]
As a form of GO that is reduced to destruct the conjugates and to form defects, the structural flexibility (e.g., higher stiffness and tensile strength) and excellent conductivity of rGO have been examined. Sun et al. have reported that adding 0.30 wt% rGO increased the yield strength and ultimate tensile strength of an rGO/Al composite by 15.6% and 11.7% compared with the bare Al material, respectively [118]. The excellent electrochemical properties of rGO-containing metal oxides has allowed them to be suitable candidates for anode materials in battery applications [119].
Because of the quantum confinement and edge effects, GQDs revealed superior luminescence properties, chemical stability, and biocompatibility [120]. There has been much interest in the use of GQDs for applications in microelectronic, sensing, and biomedical technologies [113]. Overall, composites based on graphene or rGO are of increasing interest as the materials for the synthesis of photocatalysts since they have suitable physicochemical and optical properties, such as high specific surface areas, superior electron mobility, and excellent light transmissivity.

4. GFN-TiO2

4.1. Synthesis

Many materials and methods can be used to synthesize TiO2-containing composites. It has been reported that these composites can be produced in many different forms, such as nanoparticles [121,122,123,124], nanofibers [125,126,127], and nanosheets [128,129]. The forms affect the physicochemical properties of these composites, such as the specific surface areas, influencing their photocatalytic activities. For example, the synthesis of TiO2-containing nanowires [130,131,132,133,134], nanorods [135,136,137,138], and nanotubes [139,140,141,142] with high specific areas that are associated with their improved efficiencies have been revealed. Table 5 lists the selected physicochemical properties of TiO2-containing composites prepared in different dimensions. Besides the forms of the catalysts, the materials added in the synthesis of composites are another key. Among various materials, including carbonaceous materials and metal oxides, that are commonly used to enhance their photocatalytic performance [143,144], GFNs have aroused substantial attention recently due to their unique characteristics described above. Table 6 lists the methods that have been reported for the synthesis of GFN-TiO2. These methods include ion implantation, sintering at high temperatures, plasma processes, the hydrothermal method, the sol-gel method, hydrolysis, chemical modification, and low-temperature carbonization [32]. The hydrothermal method is the most frequently used method, given the advantages comprising the adjustable crystal form, GFN content, and variable reduction level of an rGO-TiO2 [145]. This method is known to avoid the high-temperature destruction of carbonaceous structures and successfully preserve stable and complete crystal forms.

4.2. Characterization

Different approaches have been used to study the different surface characteristics and chemical structures of GFN-TiO2 (Table 7). Scanning electron microscopy (SEM) [151,163,164,165,166], transmission electron microscopy (TEM) [163,165,166], and atomic force microscopy (AFM) [164] are typically used for morphological observation. The results indicated that GFN was well embedded or covered by TiO2. The composites with lower GFN ratios tended to aggregate, forming large spherical-shaped particles [151,163,164,165,166]. Adding graphene increased and then decreased the crystallite size of composites. The initial augmentation was caused by accelerating the crystallization of TiO2. Excess H2O by the dispersion of graphene promoted the hydrolysis of titanium isopropoxide. Continuously increasing the graphene content enhanced incorporation between the nucleation centers, delaying crystallization and decreasing the crystallite size [151,163,164,165,166]. Composites could exhibit non-spherical structures, such as platelet- or flower-like morphology with elevated GFN ratios [151,163,164,165,166,167]. The TEM studies indicated that GFN-TiO2 exhibited two-dimensional structures [163,165,166,167]. An AFM study showed a significant increase in the thickness when excess graphene was added during composite preparation [164].
The chemical constitutions of GFN-TiO2 were investigated by using Fourier transform infrared spectrometry (FTIR) [151,165,168], X-ray photoelectron spectroscopy (XPS) [163,164], X-ray diffraction (XRD) [151,163,164,165,166,168], Raman spectrometry [163,164,165], and electron paramagnetic resonance (EPR) [166]. The FTIR results showed that the peak at 400–900 cm−1 was broadened or shifted due to the presence of Ti-O-C in the Ti-O-Ti adsorption peak. The original peaks of carbonyl (C=O, 1700 cm−1) and epoxy (C-O, 1230 cm−1) groups of GO became negligible in the results of GFN-TiO2 [151,165,168]. The XPS studies observed the bands of 463.2 and 458.5 eV in GFN-TiO2, indicating a chemical state of Ti4+ (TiO2) in GFN-TiO2 [163,164]. The identification of the peaks associated with Ti and GFN indicated the presence of Ti-C, O=C-O-Ti, and C-O-Ti in TiO2-GFNs, as the C1s spectrum showed peaks attributed to C=C/C-C, epoxy (C-O)/hydroxyl (C-OH), and carboxyl groups (C(=O)OH). [163,164]. The XRD studies have revealed the peak areas of anatase (25.3°) and a few rutile phases (27.4°), indicating TiO2 was well mixed with GFN with limited phase changes [151,163,164,165,166,168]. The Raman spectra of GFN-TiO2 exhibited bands of Eg(1) (149 cm−1), B1g(1) (395 cm−1), A1g+B1g(2) (517 cm−1), and Eg(2) (640 cm−1), attributable to the symmetric stretching and symmetric/asymmetric bending vibrations of the O-Ti-O group. The spectra also exhibited D (1384 cm−1) and G bands (1596 cm−1) of GFN, as the D/G intensity ratio was higher than that of GFN [163,164,165]. The EPR study showed increasing intensities of the hydroxyl and superoxide radicals by increasing the ratio of GFN to TiO2 [166].
Physicochemical properties including the surface charge, thermal stability, surface area, pore size, and pore volume of TiO2-GFN have been investigated by Zeta potential analysis [164,167], thermal gravity analysis (TGA) [164], and Brunauer–Emmett–Teller (BET) analysis [151,163,164,165,168], respectively. The nucleation of TiO2 on GFN masked the functional groups on the surface and lowered the zeta potential of GFN-TiO2 [164,167]. The TGA study showed a better heat resistance of GFN-TiO2, as TiO2 stabilized GO by the interaction between oxygen-containing groups of GFN and TiO2 [164]. Most studies have indicated a higher surface area of GFN-TiO2 compared to that of TiO2 [151,163,164,165,168], whereas an opposite trend has also been reported in a few studies [151,163,164,165,168]. GFN-TiO2 typically exhibited mesopore size distribution with averages near 10 nm [151,163,164,165,168].
Potentiostat, photoluminescence (PL), and ultraviolet-visible spectroscopy (UV-Vis) are useful tools to investigate the optical characteristics of GFN-TiO2 [151,164,165,166,168]. A study has reported that an optimal ratio of GFN to TiO2 increased the current density of GFN-TiO2, because the two-dimensional conjugation structure of GFN accepted and transported the excited electron from TiO2 [168]. Pallotti et al. used photoluminescence (PL) spectroscopy for real-time analysis to trace the time dynamics of the photoreduction of GO [169]. It was found in real-time that the photocatalysis induced by the presence of TiO2 contributed to GO photoreduction. By adding GFN into TiO2, the absorption edge of GFN-TiO2 displayed an increase in wavelength (known as redshift) that indicated a bandgap narrowing. Its light absorption intensity in the UV region was also increased [151,164,165,166,168]. Table 8 lists some examples of TiO2-GFN prepared for photocatalysis and battery storage.

4.3. Photocatalysis Enhancement

Studies have demonstrated the enhanced photocatalysis activity of GFN-TiO2, as illustrated in Figure 5. An optimal graphene addition content (e.g., 0.05 wt%) showed photocatalytic activity higher than that of pure TiO2 by a factor of 1.7 [163]. The excellent acceptance and transport of electrons by graphene reduced the recombination of charge carriers during photocatalysis. It has been indicated that the excellent conductivity of GFN suppressed the recombination of e−h+ pairs, enhancing radical formation and pollutant degradation [151,163,164,165,166,168]. The formation of the Ti-O-C bond of GFN-TiO2 effectively reduced the bandgap energy (e.g., 2.66–3.18 eV) [151,166,168]. Compared to pure TiO2, GFN-TiO2 was more efficient to absorb photons for the generation of e−h+ pair due to the shift of the absorption edge toward the visible region [173].

5. Photocatalytic Removal of Pollutants

5.1. Water-Phase Pollutants

GFN-TiO2 has been used for the photocatalytic removal of inorganic, organic, and biological pollutions in the water phase (Table 9). The photocatalytic reduction of inorganic pollutants such as metal ions was one example. Jiang et al. investigated the reduction of Cr(VI) to Cr(III) in water by using GFN-TiO2 [164]. The reduction rate constant was 0.0691 min−1, exceeding that of using pure TiO2 (0.0174 min−1) by a factor of 3.9. In another Cr(VI) removal study, the Cr(VI) concentration was adsorbed (~55%) by using TiO2-GO for 1 h, and with UV irradiation, nearly all Cr(VI) concentration was reduced in 7 h [174]. In the same system using TiO2 with UV irradiation, the Cr(VI) concentrations were limitedly adsorbed (23%) and reduced (30%).
Graphene-TiO2 has been frequently investigated for its potentials for photocatalytic degradation of organic pollutants. Homolytic cleavage is typically the first chemical step in photodegradation. Free radicals are formed in this step and rapidly react with any oxygen present in the system. Li et al. investigated the photocatalytic activity of graphene-TiO2 towards representative aqueous persistent organic pollutants (POPs) [170]. The POPs included rhodamine B, norfloxacin, and aldicarb. The presence of graphene-TiO2 significantly enhanced the removal of these POPs. While the compound concentrations were negligibly changed during pure photolysis, the presence of GFN-TiO2 (0.86% w/w of graphene) resulted in 79.7% and 86.2% of total organic carbon (TOC) removal in the experiments of rhodamine B and norfloxacin, respectively, after 10 h of simulated sunlight irradiation (λ > 320 nm). Only 36.8% of TOC removal was observed in the aldicarb experiment after 25 h of visible light irradiation (λ > 400 nm). Zhang et al. investigated photodegradation of methylene blue by using TiO2, carbon nanotube (CNT)-TiO2, and graphene-TiO2 as photocatalysts [175]. In 1 h of UV irradiation, the removal efficiency of graphene-TiO2 (85%) was significantly higher than TiO2 (25%) and CNT-TiO2 (71%). Using visible light reduced the performance of TiO2 by a factor of 2, whereas the removal efficiency of graphene-TiO2 (65%) was less affected.
GO represents another material that can work well with TiO2, forming an efficient photocatalyst. Perera et al. compared the photodegradation of malachite green by using TiO2, GO, and GO-TiO2 [145]. Pseudo-first-order reactions were found when TiO2 and GO-TiO2 were used as catalysts. The rate constant of GO-TiO2 (0.0674min−1) exceeded that of TiO2 (0.0281 min−1) by a factor of 3. No photodegradation of malachite green occurred in the presence of GO. Another study investigated the photodegradation of rhodamine B by using three different nanosphere catalysts (amine-modified TiO2–SiO2, graphene-TiO2, and GO-TiO2–SiO2) [176]. In 1.5 h of irradiation, the removal efficiencies of graphene-TiO2 (91%) and GO-TiO2–SiO2 (71%) were significantly higher than that of amine-modified TiO2–SiO2 (65%), indicating the synergistic effect between graphene or GO and TiO2 for the enhanced catalysis activity.
The use of rGO-TiO2 for the enhanced photocatalytic degradation of organic pollutants has also been demonstrated. Increasing the rGO content (from 0 to 1% w/w) in rGO-TiO2 enhanced the photocatalytic decomposition of phenol (the 1st-order rate constant was increased from 0.0039 to 0.0048 min−1) [177]. rGO-TiO2 exhibited fine photocatalytic performance after 5 cycles; however, a high rGO content (e.g., 5% w/w) potentially shielded the catalyst surface from light absorption, reducing the photocatalytic activity. Ng et al. investigated the removal of 2,4-dichlorophenolyxacetic acid (2,4-D), a commonly used herbicide, by photocatalytic reduction using TiO2 and rGO-TiO2 [178]. The pseudo-first-order rate constants of using TiO2 and rGO-TiO2 were 0.002 and 0.008 min−1, respectively. Adding rGO increased the response of the photocurrent by a factor of 2 and the availability of 2,4-D on the surface of rGO-TiO2, improving the whole photocatalytic reaction by a factor of 4.
Photocatalysis is capable of being adopted for use in many applications for disinfection in water matrices. Adding graphene in Ag3PO4-TiO2 effectively improved the synergistic photocatalytic disinfection of E. coli, S.aureus, S.typhi, P. aeruginosa, B. subtilis, and B. pumilus [180]. Fernández-Ibáñez et al. have reported effective solar photocatalytic disinfection of E. coli and F. solani spores by using rGO–TiO2. The presence of rGO significantly enhanced the performance of photocatalytic disinfection of E. coli. Increasing rGO–TiO2 from 0 to 500 mg/L accelerated the inactivation of E. coli (106 colony-forming units (CFU)/mL) from more than 100 to 10 min and reduced the solar UV dosage needed from 123 to 11 kJ/m2. Although both rGO-TiO2 and pure TiO2 exhibited excellent disinfection of F. solani spores, rGO significantly reduced the solar energy required from 336.2 to 42.1 kJ/m2 [179]. A certain ratio between rGO and TiO2 significantly enhanced the photocatalytic disinfection under UV and solar irradiation [182]. Another study has also demonstrated that GO, which effectively separated photo-generated e−h+ pairs for more ▪OH production, improved the photocatalytic disinfection of E. coli. In 30 min, the disinfection efficiencies of using pure TiO2, GO, GO-TiO2 were 39.27%, 73.82%, 99.60%, respectively [181]. More detailed information concerning the removal of different inorganic, organic, and biological pollutants by using GFN-TiO2 is available in Table 8.

5.2. Air-Phase Pollutants

Similar to the removal of pollutants in the water phase, GFN-TiO2 has been adopted for use in removing a wide range of air pollutants. Shorter contact times and the complexity of the heterogeneous photocatalytic reactions (e.g., photon absorbance and radical reactions) between pollutants and catalyst surfaces represent two typical challenges in this field [151].
In the aspect of inorganic removal, the treatment efficiencies of gaseous NOx (from NO(g) to NO2(g) to NO3(s)) by using pure TiO2, graphene-TiO2, and rGO-TiO2 were compared [165]. An appreciable level of GFN (e.g., 0.01–0.1% graphene or rGO) in TiO2 improved the removal of NOx under UV and visible light. The NOx removal efficiencies were 25.45%, 26.26–35.40%, and 39.38–42.86% by using TiO2, graphene-TiO2, and rGO-TiO2 under UV light, respectively, while under visible light the removal efficiencies using TiO2, graphene-TiO2, and rGO-TiO2 were 9.35%, 15.20–22.75%, and 19.88–22.34%, respectively. Giampiccolo et al. prepared graphene-TiO2 by using the sol-gel method for electrochemical sensing and photocatalytic degradation of NOx in the air [183]. Interestingly, the performances of graphene-TiO2 prepared by using the same method but with different step orders were compared (adding graphene to the reaction before initiating the sol-gel reaction followed by annealing (GTiO2S) and adding graphene to TiO2 which had already been annealed (GTiO2M)). The addition of graphene significantly improved the performance of the catalysts under solar irradiation (280–780 nm) (e.g., the pseudo-first-order rate constants of NOx removal by GTiO2S, GTiO2M, and TiO2 were 6.7, 5.6, and 4.3/min, respectively.). The thermal treatment helped synthesize graphene and TiO2 in more intimate contact and improved the exhibition. Besides NOx, photodegradation of CO by using GO-TiO2, which was functionalized by attaching a cobalt (Co)-imidazole (Im) complex on GO, was investigated [109]. The results revealed that the bandgaps of this functionalized GO-TiO2 (with Co and Im), GO-TiO2, and pure TiO2 were 2.78, 2.96, and 3.10 eV, respectively. The removal efficiencies of CO and NOx were improved from 10% to 46% and from 16% to 51% when the catalyst was changed from TiO2 to the functionalized GO-TiO2, respectively. Xu et al. added graphene into TiO2 to enhance the photocatalytic CO2 conversion to chemical fuels [184]. The addition of graphene inhibited the recombination of e−h+ pairs and raised the surface temperature, improving the CO2 conversion efficiency. The conversion rates of CO2 to CH4 and CO by using graphene-TiO2 were higher than those using TiO2 by factors of 5.1 and 2.8, respectively.
Studies have demonstrated the photocatalytic degradation of organic pollutants in the air phase by using GFN-TiO2. Zang et al. have reported that adding graphene into TiO2 with a specific ratio (e.g., 0.5% w/w) exhibited a synergetic effect on the UV light photodegradation of benzene (the mineralization rates of GFN-TiO2 and TiO2 were 76.2% in 10 h and 1.2% in 28 h, respectively). The adsorption of benzene and intermediates during benzene degradation negatively affecting TiO2 adsorbing UV light was decreased by the presence of graphene. However, excess graphene could adsorb extra compounds and impact light absorption. Benzene removal was limitedly found when visible light was used [166]. Similarly, in a study that focused on the photocatalytic degradation of acetone in the air, graphene-TiO2 exhibited a better activity (the pseudo-1st-order rate constant was 10.2 × 10−3/min) exceeding that of pure TiO2 (5.99 × 10−3/min) by a factor of 1.7 and a good reproducibility after three cycles of illumination [163].
Adding other materials to graphene-TiO2 has been investigated to further enhance its photocatalytic activity. Photocatalytic degradation of formaldehyde by using graphene-TiO2-Fe3+ has been reported [168]. Under UV light, both graphene-TiO2-Fe3+ and graphene-TiO2 revealed better performances than pure TiO2, as only the photolytic activity of graphene-Fe3+-TiO2 was better under visible light irradiation. The photocatalyst with a TiO2/graphene ratio of 50 and a ratio of Fe3+/graphene-TiO2 of 0.12% revealed the optimal performance. Nitrogen has been doped into reduced graphene-TiO2 to change the polarity of the catalyst and to influence the adsorption and photodegradation of polar acetaldehyde and nonpolar ethylene [185]. Both reduced graphene-TiO2 and N-doped reduced graphene-TiO2 exhibited higher treatment efficiencies than pure TiO2. One explanation was that nitrogen doping improved the polarity of the catalyst, further enhancing the removal efficiency of polar acetaldehyde.
The feasibility of adding GO into TiO2 for the photocatalytic degradation of organic pollutants has been reported. A study used GO-TiO2 as a photocatalyst to accelerate the degradation of benzene, toluene, ethylbenzene, and xylene (BTEX) in the air [151]. Under UV irradiation, the removal of these compounds by using GO-TiO2 was higher than that of using pure TiO2 by a factor of 1.2, while GO-TiO2 exhibited an excellent treatment efficiency exceeding that of pure TiO2 by a factor of 12 under visible light irradiation. GO-TiO2 has also been used for the photocatalytic degradation of methyl ethyl ketone in indoor air [171]. The addition of GO in TiO2 has improved the removal efficiency from 32.7% to 96.8% under visible light irradiation. Proper humidity (e.g., 40%), flow rate (e.g., 50 mL/min), and pollutant concentration (e.g., 30 ppmv) were the key to optimal performance. Note that the use of nanostructured membranes based on polymeric nanofibers using TiO2 and GFNs, including GO, rGO, and few-layer graphene, for the photocatalytic oxidation of gas-phase methanol has been reported. As the photocatalytic activity was greatly changed by the membrane structure and affected by the affinity of GFN to the polymer matrix, rGO exhibited better performance due to its more enhanced electron mobility [186]. Table 10 summarizes the applications of GFN-TiO2 for the photodegradation of organic pollutants in the air in these studies.

6. Conclusions and Future Work

TiO2 has been intensively investigated in early studies given its photocatalytic effects for radical production degrading a wide range of pollutants in the environment. TiO2 with an anatase-crystal structure generally exhibited higher photocatalytic activity than rutile TiO2. Its intrinsic properties, including the surface area, adsorption capacity, bandgap, and lifetime of the e−h+ pair, have provided opportunities for applications under UV light irradiation. However, these properties could be improved to guarantee a wider range of applications, such as those for visible-light or solar irradiation. The advantages conferred by the physical, optical, and electrochemical properties of GFN have contributed to the current variety of GFN-TiO2 catalysts that exhibit improved characteristics, such as higher surface areas, more rapid electron transfer, and narrower bandgap. Although the physicochemical properties and photocatalytic activity could be different between GFN-TiO2 prepared by different methods, many studies presented in this review have demonstrated that the applications of using GFN-TiO2 have greatly improved photocatalytic reactions for the treatment of organic, inorganic, and biological pollutants in water and air phases. GFN-TiO2 exhibited better photocatalytic activity than pure TiO2 under UV light irradiation, as the improvement is more significant under visible-light irradiation.
Note that the ratio of GFN and TiO2 in GFN-TiO2 is typically the key to optimizing the photocatalytic reactions in many studies. Excess GFN could increase the opacity of GFN-TiO2, limiting the light absorption of TiO2 and negatively affecting the formation of e−h+ pairs. Besides the type of GFN (e.g., graphene, GO, and rGO), different preparation methods affected the properties of GFN-TiO2 products. Recently, studies have turned their attention to green chemistry that could use fewer chemicals or energy for the preparation of GFN and GFN-TiO2. Examples include the electrochemical exfoliation of graphene and the UV-assisted photoreduction of GO. The applications of GFN-TiO2 for the removal of inorganic pollutants in the water, such as photocatalytic reactions of ammonium and nitrite and inactivation of biological pollutants in the air, were relatively limitedly examined and represent other directions of technological innovation and possible future development in these fields.

Author Contributions

Conceptualization, C.-H.L. and W.-H.C.; methodology, C.-H.L.; formal analysis, C.-H.L.; investigation, C.-H.L. and W.-H.C.; resources, W.-H.C.; data curation, C.-H.L. and W.-H.C.; writing—original draft preparation, C.-H.L.; writing—review and editing, W.-H.C.; visualization, C.-H.L. and W.-H.C.; supervision, W.-H.C.; project administration, C.-H.L. and W.-H.C.; funding acquisition, W.-H.C. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the Ministry of Science and Technology (MOST) in Taiwan grant number MOST 110-2628-E-110-001-, MOST108-2622-E-110-013-CC3, MOST 107-2221-E-110-003-MY3, and MOST 106-2621-M-110-003.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Prasad, S.; Singh, A.; Joshi, H. Ethanol as an alternative fuel from agricultural, industrial and urban residues. Resour. Conserv. Recycl. 2007, 50, 1–39. [Google Scholar] [CrossRef]
  2. Taghizadeh-Alisaraei, A.; Motevali, A.; Ghobadian, B. Ethanol production form date wastes: Adapted technologies, challenges, and global potential. Renew. Energy 2019, 143, 1094–1110. [Google Scholar] [CrossRef]
  3. Bai, Y.; Zhou, L.; Ma, K.; Zhou, H.; Xin, Y. Study on catalytic performances and reaction mechanisms of graphene electroactive membrane in wastewater treatment. Sep. Purif. Technol. 2019, 226, 278–285. [Google Scholar] [CrossRef]
  4. Shih, Y.-J.; Huang, C.-P.; Chan, Y.-H.; Huang, Y.-H. Electrochemical degradation of oxalic acid over highly reactive nano-textured γ-and α-MnO2/carbon electrode fabricated by KMnO4 reduction on loofah sponge-derived active carbon. J. Hazard. Mater. 2019, 379, 120759. [Google Scholar] [CrossRef] [PubMed]
  5. Wellia, D.V.; Xu, Q.C.; Sk, M.A.; Lim, K.H.; Lim, T.M.; Tan, T.T.Y. Experimental and theoretical studies of Fe-doped TiO2 films prepared by peroxo sol-gel method. Appl. Catal. A Gen. 2011, 401, 98–105. [Google Scholar] [CrossRef]
  6. Chen, W.-H.; Huang, J.-R.; Lin, C.-H.; Huang, C.-P. Catalytic degradation of chlorpheniramine over GO-Fe3O4 in the presence of H2O2 in water: The synergistic effect of adsorption. Sci. Total Environ. 2020, 736, 139468. [Google Scholar] [CrossRef]
  7. Lin, C.H.; Li, C.M.; Chen, C.H.; Chen, W.H. Removal of chlorpheniramine and variations of nitrosamine formation potentials in municipal wastewaters by adsorption onto the GO-Fe3O4. Environ. Sci. Pollut. Res. 2019, 26, 20701–20711. [Google Scholar] [CrossRef]
  8. Kisch, H. Semiconductor photocatalysis—Mechanistic and synthetic aspects. Angew. Chem. Int. Ed. 2013, 52, 812–847. [Google Scholar] [CrossRef] [PubMed]
  9. Ibhadon, A.; Fitzpatrick, P. Heterogeneous photocatalysis: Recent advances and applications. Catalysts 2013, 3, 189–218. [Google Scholar] [CrossRef] [Green Version]
  10. Gupta, S.M.; Tripathi, M. A review of TiO2 nanoparticles. Chin. Sci. Bull. 2011, 56, 1639. [Google Scholar] [CrossRef] [Green Version]
  11. Trouiller, B.; Reliene, R.; Westbrook, A.; Solaimani, P.; Schiestl, R.H. Titanium Dioxide Nanoparticles Induce DNA Damage and Genetic Instability In vivo in Mice. Cancer Res. 2009, 69, 8784–8789. [Google Scholar] [CrossRef] [Green Version]
  12. IARC. IARC Monographs on the Evaluation of Carcinogenic Risks to Humans: Carbon Black, Titanium Dioxide, and Talc; IARC: Lyon, France, 2010. [Google Scholar]
  13. Fujishima, A.; Honda, K. Electrochemical photolysis of water at a semiconductor electrode. Nature 1972, 238, 37–38. [Google Scholar] [CrossRef] [PubMed]
  14. Frank, S.N.; Bard, A.J. Heterogeneous photocatalytic oxidation of cyanide ion in aqueous solutions at titanium dioxide powder. J. Am. Chem. Soc. 1977, 99, 303–304. [Google Scholar] [CrossRef]
  15. Frank, S.N.; Bard, A.J. Heterogeneous photocatalytic oxidation of cyanide and sulfite in aqueous solutions at semiconductor powders. J. Phys. Chem. 1977, 81, 1484–1488. [Google Scholar] [CrossRef]
  16. Hoffmann, M.R.; Martin, S.T.; Choi, W.; Bahnemann, D.W. Environmental applications of semiconductor photocatalysis. Chem. Rev. 1995, 95, 69–96. [Google Scholar] [CrossRef]
  17. Fox, M.A.; Dulay, M.T. Heterogeneous photocatalysis. Chem. Rev. 1993, 93, 341–357. [Google Scholar] [CrossRef]
  18. Carp, O.; Huisman, C.L.; Reller, A. Photoinduced reactivity of titanium dioxide. Prog. Solid State Chem. 2004, 32, 33–177. [Google Scholar] [CrossRef]
  19. Banerjee, S.; Pillai, S.C.; Falaras, P.; O’shea, K.E.; Byrne, J.A.; Dionysiou, D.D. New insights into the mechanism of visible light photocatalysis. J. Phys. Chem. Lett. 2014, 5, 2543–2554. [Google Scholar] [CrossRef] [Green Version]
  20. Pelaez, M.; Nolan, N.T.; Pillai, S.C.; Seery, M.K.; Falaras, P.; Kontos, A.G.; Dunlop, P.S.; Hamilton, J.W.; Byrne, J.A.; O’shea, K. 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]
  21. Gaya, U.I.; Abdullah, A.H. Heterogeneous photocatalytic degradation of organic contaminants over titanium dioxide: A review of fundamentals, progress and problems. J. Photochem. Photobiol. C Photochem. Rev. 2008, 9, 1–12. [Google Scholar] [CrossRef]
  22. Banerjee, S.; Dionysiou, D.D.; Pillai, S.C. Self-cleaning applications of TiO2 by photo-induced hydrophilicity and photocatalysis. Appl. Catal. B Environ. 2015, 176, 396–428. [Google Scholar] [CrossRef] [Green Version]
  23. Keane, D.A.; McGuigan, K.G.; Ibáñez, P.F.; Polo-López, M.I.; Byrne, J.A.; Dunlop, P.S.; O’Shea, K.; Dionysiou, D.D.; Pillai, S.C. Solar photocatalysis for water disinfection: Materials and reactor design. Catal. Sci. Technol. 2014, 4, 1211–1226. [Google Scholar] [CrossRef] [Green Version]
  24. Fujishima, A. Photocatalytic and self-cleaning functions of TiO2 coatings. In Sustainable Energy and Environmental Technologies; World Scientific: Singapore, 2001; pp. 1–5. [Google Scholar]
  25. Chen, X.; Mao, S.S. Titanium dioxide nanomaterials: Synthesis, properties, modifications, and applications. Chem. Rev. 2007, 107, 2891–2959. [Google Scholar] [CrossRef] [PubMed]
  26. Sharma, S.D.; Singh, D.; Saini, K.; Kant, C.; Sharma, V.; Jain, S.; Sharma, C. Sol–gel-derived super-hydrophilic nickel doped TiO2 film as active photo-catalyst. Appl. Catal. A Gen. 2006, 314, 40–46. [Google Scholar] [CrossRef]
  27. Catauro, M.; Tranquillo, E.; Dal Poggetto, G.; Pasquali, M.; Dell’Era, A.; Vecchio Ciprioti, S. Influence of the heat treatment on the particles size and on the crystalline phase of TiO2 synthesized by the sol-gel method. Materials 2018, 11, 2364. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  28. Dubey, R.; Krishnamurthy, K.V.; Singh, S. Experimental studies of TiO2 nanoparticles synthesized by sol-gel and solvothermal routes for DSSCs application. Results Phys. 2019, 14, 102390. [Google Scholar] [CrossRef]
  29. Li, G.; Li, L.; Boerio-Goates, J.; Woodfield, B.F. High purity anatase TiO2 nanocrystals: Near room-temperature synthesis, grain growth kinetics, and surface hydration chemistry. J. Am. Chem. Soc. 2005, 127, 8659–8666. [Google Scholar] [CrossRef]
  30. Hirano, M.; Nakahara, C.; Ota, K.; Tanaike, O.; Inagaki, M. Photoactivity and phase stability of ZrO2-doped anatase-type TiO2 directly formed as nanometer-sized particles by hydrolysis under hydrothermal conditions. J. Solid State Chem. 2003, 170, 39–47. [Google Scholar] [CrossRef]
  31. Wang, L.Q.; Yang, X.N.; Zhao, X.L.; Zhang, R.J.; Yang, Y.L. Preparation of TiO2 nanoparticles in the solvothermal method. In Proceedings of the Key Engineering Materials; Trans Tech Publications Ltd.: Freienbach, Switzerland, 2011; pp. 1672–1677. [Google Scholar]
  32. Hong, S.-S.; Lee, M.S.; Park, S.S.; Lee, G.-D. Synthesis of nanosized TiO2/SiO2 particles in the microemulsion and their photocatalytic activity on the decomposition of p-nitrophenol. Catal. Today 2003, 87, 99–105. [Google Scholar] [CrossRef]
  33. Volz, H.G. Pigments, Inorganic Ullman’s Encyclopedia of Industrial Chemistry; Wiley-VCH: Weinheim, Germany, 2006. [Google Scholar]
  34. Teleki, A.; Pratsinis, S.E.; Kalyanasundaram, K.; Gouma, P. Sensing of organic vapors by flame-made TiO2 nanoparticles. Sens. Actuators B Chem. 2006, 119, 683–690. [Google Scholar] [CrossRef] [Green Version]
  35. Jung, H.S.; Kim, H. Origin of Low Photocatalytic Activity of Rutile TiO2. Electron. Mater. Lett. 2009, 5, 73–76. [Google Scholar] [CrossRef]
  36. Luttrell, T.; Halpegamage, S.; Tao, J.G.; Kramer, A.; Sutter, E.; Batzill, M. Why is anatase a better photocatalyst than rutile?—Model studies on epitaxial TiO2 films. Sci. Rep. 2014, 4, 4043. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  37. Zhang, Q.; Gao, L.; Guo, J. Effects of calcination on the photocatalytic properties of nanosized TiO2 powders prepared by TiCl4 hydrolysis. Appl. Catal. B Environ. 2000, 26, 207–215. [Google Scholar] [CrossRef]
  38. Kesselman, J.M.; Shreve, G.A.; Hoffmann, M.R.; Lewis, N.S. Flux-matching conditions at TiO2 photoelectrodes: Is interfacial electron transfer to O2 rate-limiting in the TiO2-catalyzed photochemical degradation of organics? J. Phys. Chem. 1994, 98, 13385–13395. [Google Scholar] [CrossRef]
  39. Hanaor, D.A.; Sorrell, C.C. Review of the anatase to rutile phase transformation. J. Mater. Sci. 2011, 46, 855–874. [Google Scholar] [CrossRef] [Green Version]
  40. Sclafani, A.; Herrmann, J. Comparison of the photoelectronic and photocatalytic activities of various anatase and rutile forms of titania in pure liquid organic phases and in aqueous solutions. J. Phys. Chem. 1996, 100, 13655–13661. [Google Scholar] [CrossRef]
  41. Asahi, R.; Taga, Y.; Mannstadt, W.; Freeman, A.J. Electronic and optical properties of anatase TiO2. Phys. Rev. B 2000, 61, 7459. [Google Scholar] [CrossRef]
  42. Koelsch, M.; Cassaignon, S.; Minh, C.T.T.; Guillemoles, J.-F.; Jolivet, J.-P. Electrochemical comparative study of titania (anatase, brookite and rutile) nanoparticles synthesized in aqueous medium. Thin Solid Film. 2004, 451, 86–92. [Google Scholar] [CrossRef]
  43. Amtout, A.; Leonelli, R. Optical properties of rutile near its fundamental band gap. Phys. Rev. B 1995, 51, 6842. [Google Scholar] [CrossRef]
  44. Zhu, S.-C.; Xie, S.-H.; Liu, Z.-P. Nature of rutile nuclei in anatase-to-rutile phase transition. J. Am. Chem. Soc. 2015, 137, 11532–11539. [Google Scholar] [CrossRef]
  45. Zhang, H.; Banfield, J.F. Structural characteristics and mechanical and thermodynamic properties of nanocrystalline TiO2. Chem. Rev. 2014, 114, 9613–9644. [Google Scholar] [CrossRef] [PubMed]
  46. Zhang, J.; Zhou, P.; Liu, J.; Yu, J. New understanding of the difference of photocatalytic activity among anatase, rutile and brookite TiO2. Phys. Chem. Chem. Phys. 2014, 16, 20382–20386. [Google Scholar] [CrossRef] [PubMed]
  47. Cromer, D.T.; Herrington, K. The structures of anatase and rutile. J. Am. Chem. Soc. 1955, 77, 4708–4709. [Google Scholar] [CrossRef]
  48. Mo, S.-D.; Ching, W. Electronic and optical properties of three phases of titanium dioxide: Rutile, anatase, and brookite. Phys. Rev. B 1995, 51, 13023. [Google Scholar] [CrossRef] [PubMed]
  49. Kang, X.L.; Liu, S.H.; Dai, Z.D.; He, Y.P.; Song, X.Z.; Tan, Z.Q. Titanium Dioxide: From Engineering to Applications. Catalysts 2019, 9, 191. [Google Scholar] [CrossRef] [Green Version]
  50. Hamal, D.B.; Klabunde, K.J. Synthesis, characterization, and visible light activity of new nanoparticle photocatalysts based on silver, carbon, and sulfur-doped TiO2. J. Colloid Interface Sci. 2007, 311, 514–522. [Google Scholar] [CrossRef]
  51. Geim, A.K.; Novoselov, K.S. The rise of graphene. Nat. Mater. 2007, 6, 183–191. [Google Scholar] [CrossRef]
  52. Suvarnaphaet, P.; Pechprasarn, S. Graphene-based materials for biosensors: A review. Sensors 2017, 17, 2161. [Google Scholar] [CrossRef] [Green Version]
  53. Zhu, Y.; Murali, S.; Cai, W.; Li, X.; Suk, J.W.; Potts, J.R.; Ruoff, R.S. Graphene and graphene oxide: Synthesis, properties, and applications. Adv. Mater. 2010, 22, 3906–3924. [Google Scholar] [CrossRef]
  54. Li, C.M.; Chen, C.H.; Chen, W.H. Different influences of nanopore dimension and pH between chlorpheniramine adsorptions on graphene oxide-iron oxide suspension and particle. Chem. Eng. J. 2017, 307, 447–455. [Google Scholar] [CrossRef]
  55. Bo, Z.; Shuai, X.R.; Mao, S.; Yang, H.C.; Qian, J.J.; Chen, J.H.; Yan, J.H.; Cen, K. Green preparation of reduced graphene oxide for sensing and energy storage applications. Sci. Rep. 2014, 4, 4684. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  56. Liu, S.B.; Zeng, T.H.; Hofmann, M.; Burcombe, E.; Wei, J.; Jiang, R.R.; Kong, J.; Chen, Y. Antibacterial Activity of Graphite, Graphite Oxide, Graphene Oxide, and Reduced Graphene Oxide: Membrane and Oxidative Stress. ACS Nano 2011, 5, 6971–6980. [Google Scholar] [CrossRef] [PubMed]
  57. Zheng, P.; Wu, N.Q. Fluorescence and Sensing Applications of Graphene Oxide and Graphene Quantum Dots: A Review. Chem. Asian J. 2017, 12, 2343–2353. [Google Scholar] [CrossRef] [PubMed]
  58. Lee, X.J.; Hiew, B.Y.Z.; Lai, K.C.; Lee, L.Y.; Gan, S.; Thangalazhy-Gopakumar, S.; Rigby, S. Review on graphene and its derivatives: Synthesis methods and potential industrial implementation. J. Taiwan Inst. Chem. Eng. 2019, 98, 163–180. [Google Scholar] [CrossRef]
  59. Wu, Z.S.; Ren, W.C.; Gao, L.B.; Liu, B.L.; Jiang, C.B.; Cheng, H.M. Synthesis of high-quality graphene with a pre-determined number of layers. Carbon 2009, 47, 493–499. [Google Scholar] [CrossRef]
  60. Shashurin, A.; Keidar, M. Synthesis of 2D materials in arc plasmas. J. Phys. D Appl. Phys. 2015, 48, 314007. [Google Scholar] [CrossRef] [Green Version]
  61. Kosynkin, D.V.; Higginbotham, A.L.; Sinitskii, A.; Lomeda, J.R.; Dimiev, A.; Price, B.K.; Tour, J.M. Longitudinal unzipping of carbon nanotubes to form graphene nanoribbons. Nature 2009, 458, 872. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  62. Mishra, N.; Boeckl, J.; Motta, N.; Iacopi, F. Graphene growth on silicon carbide: A review. Phys. Status Solidi 2016, 213, 2277–2289. [Google Scholar] [CrossRef]
  63. Dato, A.; Radmilovic, V.; Lee, Z.; Phillips, J.; Frenklach, M. Substrate-free gas-phase synthesis of graphene sheets. Nano Lett. 2008, 8, 2012–2016. [Google Scholar] [CrossRef]
  64. Yang, Y.; Liu, R.; Wu, J.; Jiang, X.; Cao, P.; Hu, X.; Pan, T.; Qiu, C.; Yang, J.; Song, Y. Bottom-up fabrication of graphene on silicon/silica substrate via a facile soft-hard template approach. Sci. Rep. 2015, 5, 13480. [Google Scholar] [CrossRef] [Green Version]
  65. Yang, X.; Dou, X.; Rouhanipour, A.; Zhi, L.; Räder, H.J.; Müllen, K. Two-dimensional graphene nanoribbons. J. Am. Chem. Soc. 2008, 130, 4216–4217. [Google Scholar] [CrossRef] [PubMed]
  66. Brodie, B.C. On the Atomic Weight of Graphite. Philos. Trans. R. Soc. Lond. 1859, 149, 249–259. [Google Scholar]
  67. Staudenmaier, L. Verfahren zur Darstellung der Graphitsäure. Ber. Dtsch. Chem. Ges. 1898, 31, 1481–1499. [Google Scholar] [CrossRef] [Green Version]
  68. Hofmann, U.; König, E. Untersuchungen über Graphitoxyd. Z. Anorg. Allg. Chem. 1937, 234, 311–336. [Google Scholar] [CrossRef]
  69. Hummers, W.S.; Offeman, R.E. Preparation of graphitic oxide. J. Am. Chem. Soc. 1958, 80, 1339. [Google Scholar] [CrossRef]
  70. Chen, J.; Yao, B.; Li, C.; Shi, G. An improved Hummers method for eco-friendly synthesis of graphene oxide. Carbon 2013, 64, 225–229. [Google Scholar] [CrossRef]
  71. 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] [Green Version]
  72. Zaaba, N.; Foo, K.; Hashim, U.; Tan, S.; Liu, W.-W.; Voon, C. Synthesis of graphene oxide using modified hummers method: Solvent influence. Procedia Eng. 2017, 184, 469–477. [Google Scholar] [CrossRef]
  73. 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]
  74. Peng, L.; Xu, Z.; Liu, Z.; Wei, Y.; Sun, H.; Li, Z.; Zhao, X.; Gao, C. An iron-based green approach to 1-h production of single-layer graphene oxide. Nat. Commun. 2015, 6, 5716. [Google Scholar] [CrossRef] [Green Version]
  75. Gómez-Navarro, C.; Weitz, R.T.; Bittner, A.M.; Scolari, M.; Mews, A.; Burghard, M.; Kern, K. Electronic transport properties of individual chemically reduced graphene oxide sheets. Nano Lett. 2007, 7, 3499–3503. [Google Scholar] [CrossRef]
  76. Zhu, Y.; Stoller, M.D.; Cai, W.; Velamakanni, A.; Piner, R.D.; Chen, D.; Ruoff, R.S. Exfoliation of graphite oxide in propylene carbonate and thermal reduction of the resulting graphene oxide platelets. ACS Nano 2010, 4, 1227–1233. [Google Scholar] [CrossRef]
  77. Stankovich, S.; Dikin, D.A.; Piner, R.D.; Kohlhaas, K.A.; Kleinhammes, A.; Jia, Y.; Wu, Y.; Nguyen, S.T.; Ruoff, R.S. Synthesis of graphene-based nanosheets via chemical reduction of exfoliated graphite oxide. Carbon 2007, 45, 1558–1565. [Google Scholar] [CrossRef]
  78. Zhang, J.; Yang, H.; Shen, G.; Cheng, P.; Zhang, J.; Guo, S. Reduction of graphene oxide via L-ascorbic acid. Chem. Commun. 2010, 46, 1112–1114. [Google Scholar] [CrossRef] [PubMed]
  79. Stankovich, S.; Dikin, D.A.; Dommett, G.H.; Kohlhaas, K.M.; Zimney, E.J.; Stach, E.A.; Piner, R.D.; Nguyen, S.T.; Ruoff, R.S. Graphene-based composite materials. Nature 2006, 442, 282–286. [Google Scholar] [CrossRef]
  80. Fernandez-Merino, M.J.; Guardia, L.; Paredes, J.I.; Villar-Rodil, S.; Solis-Fernandez, P.; Martinez-Alonso, A.; Tascon, J.M.D. Vitamin C Is an Ideal Substitute for Hydrazine in the Reduction of Graphene Oxide Suspensions. J. Phys. Chem. C 2010, 114, 6426–6432. [Google Scholar] [CrossRef]
  81. Guo, H.-L.; Wang, X.-F.; Qian, Q.-Y.; Wang, F.-B.; Xia, X.-H. A green approach to the synthesis of graphene nanosheets. ACS Nano 2009, 3, 2653–2659. [Google Scholar] [CrossRef] [PubMed]
  82. Chen, W.; Yan, L.; Bangal, P.R. Preparation of graphene by the rapid and mild thermal reduction of graphene oxide induced by microwaves. Carbon 2010, 48, 1146–1152. [Google Scholar] [CrossRef]
  83. Pan, D.Y.; Zhang, J.C.; Li, Z.; Wu, M.H. Hydrothermal Route for Cutting Graphene Sheets into Blue-Luminescent Graphene Quantum Dots. Adv. Mater. 2010, 22, 734. [Google Scholar] [CrossRef]
  84. Liang, W.X.; Bunker, C.E.; Sun, Y.P. Carbon Dots: Zero-Dimensional Carbon Allotrope with Unique Photoinduced Redox Characteristics. Acs Omega 2020, 5, 965–971. [Google Scholar] [CrossRef]
  85. Wang, Z.F.; Yu, J.F.; Zhang, X.; Li, N.; Liu, B.; Li, Y.Y.; Wang, Y.H.; Wang, W.X.; Li, Y.Z.; Zhang, L.C.; et al. Large-Scale and Controllable Synthesis of Graphene Quantum Dots from Rice Husk Biomass: A Comprehensive Utilization Strategy. ACS Appl. Mater. Interfaces 2016, 8, 1434–1439. [Google Scholar] [CrossRef] [PubMed]
  86. Van Tam, T.; Kang, S.G.; Babu, K.F.; Oh, E.-S.; Lee, S.G.; Choi, W.M. Synthesis of B-doped graphene quantum dots as a metal-free electrocatalyst for the oxygen reduction reaction. J. Mater. Chem. A 2017, 5, 10537–10543. [Google Scholar] [CrossRef]
  87. Zhang, J.; Ma, Y.-Q.; Li, N.; Zhu, J.-L.; Zhang, T.; Zhang, W.; Liu, B. Preparation of graphene quantum dots and their application in cell imaging. J. Nanomater. 2016, 2016, 9245865. [Google Scholar] [CrossRef]
  88. Li, Y.; Shu, H.; Niu, X.; Wang, J. Electronic and optical properties of edge-functionalized graphene quantum dots and the underlying mechanism. J. Phys. Chem. C 2015, 119, 24950–24957. [Google Scholar] [CrossRef]
  89. Lu, J.; Yang, J.-X.; Wang, J.; Lim, A.; Wang, S.; Loh, K.P. One-pot synthesis of fluorescent carbon nanoribbons, nanoparticles, and graphene by the exfoliation of graphite in ionic liquids. ACS Nano 2009, 3, 2367–2375. [Google Scholar] [CrossRef] [PubMed]
  90. Wang, J.; Xin, X.; Lin, Z. Cu2ZnSnS4 nanocrystals and graphene quantum dots for photovoltaics. Nanoscale 2011, 3, 3040–3048. [Google Scholar] [CrossRef]
  91. Zheng, X.T.; Ananthanarayanan, A.; Luo, K.Q.; Chen, P. Glowing graphene quantum dots and carbon dots: Properties, syntheses, and biological applications. Small 2015, 11, 1620–1636. [Google Scholar] [CrossRef]
  92. Peng, J.; Zhao, Z.; Zheng, M.; Su, B.; Chen, X.; Chen, X. Electrochemical synthesis of phosphorus and sulfur co-doped graphene quantum dots as efficient electrochemiluminescent immunomarkers for monitoring okadaic acid. Sens. Actuators B Chem. 2020, 304, 127383. [Google Scholar] [CrossRef]
  93. Wang, L.; Li, W.; Wu, B.; Li, Z.; Pan, D.; Wu, M. Room-temperature synthesis of graphene quantum dots via electron-beam irradiation and their application in cell imaging. Chem. Eng. J. 2017, 309, 374–380. [Google Scholar] [CrossRef]
  94. Singh, R.; Kumar, R.; Singh, D.; Savu, R.; Moshkalev, S. Progress in microwave-assisted synthesis of quantum dots (graphene/carbon/semiconducting) for bioapplications: A review. Mater. Today Chem. 2019, 12, 282–314. [Google Scholar] [CrossRef]
  95. Hassan, M.; Haque, E.; Reddy, K.R.; Minett, A.I.; Chen, J.; Gomes, V.G. Edge-enriched graphene quantum dots for enhanced photo-luminescence and supercapacitance. Nanoscale 2014, 6, 11988–11994. [Google Scholar] [CrossRef] [PubMed]
  96. Choi, J.U.; Jo, W.-K. FeWO4/g-C3N4 heterostructures decorated with N-doped graphene quantum dots prepared under various sonication conditions for efficient removal of noxious vapors. Ceram. Int. 2020, 46, 11346–11356. [Google Scholar] [CrossRef]
  97. Li, R.; Liu, Y.; Li, Z.; Shen, J.; Yang, Y.; Cui, X.; Yang, G. Bottom-up fabrication of single-layered nitrogen-doped graphene quantum dots through intermolecular carbonization arrayed in a 2D plane. Chem. Eur. J. 2016, 22, 272–278. [Google Scholar] [CrossRef] [PubMed]
  98. Xia, C.; Hai, X.; Chen, X.-W.; Wang, J.-H. Simultaneously fabrication of free and solidified N, S-doped graphene quantum dots via a facile solvent-free synthesis route for fluorescent detection. Talanta 2017, 168, 269–278. [Google Scholar] [CrossRef]
  99. Lu, J.; Yeo, P.S.E.; Gan, C.K.; Wu, P.; Loh, K.P. Transforming C 60 molecules into graphene quantum dots. Nat. Nanotechnol. 2011, 6, 247. [Google Scholar] [CrossRef]
  100. Dreyer, D.R.; Park, S.; Bielawski, C.W.; Ruoff, R.S. The chemistry of graphene oxide. Chem. Soc. Rev. 2010, 39, 228–240. [Google Scholar] [CrossRef]
  101. Nethravathi, C.; Rajamathi, J.T.; Ravishankar, N.; Shivakumara, C.; Rajamathi, M. Graphite oxide-intercalated anionic clay and its decomposition to graphene− inorganic material nanocomposites. Langmuir 2008, 24, 8240–8244. [Google Scholar] [CrossRef]
  102. Lin, C.H.; Chen, W.H. Influence of water, H2O2, H2SO4, and NaOH filtration on the surface characteristics of a graphene oxide-iron (GO-Fe) membrane. Sep. Purif. Technol. 2021, 262, 118317. [Google Scholar] [CrossRef]
  103. Carmalin, A.S.; Lima, E.C.; Allaudeen, N.; Rajan, S. Application of graphene based materials for adsorption of pharmaceutical traces from water and wastewater—A review. Desalination Water Treat. 2016, 57, 27573–27586. [Google Scholar] [CrossRef]
  104. Singh, V.; Joung, D.; Zhai, L.; Das, S.; Khondaker, S.I.; Seal, S. Graphene based materials: Past, present and future. Prog. Mater. Sci. 2011, 56, 1178–1271. [Google Scholar] [CrossRef]
  105. Bolotin, K.I.; Sikes, K.J.; Jiang, Z.; Klima, M.; Fudenberg, G.; Hone, J.; Kim, P.; Stormer, H. Ultrahigh electron mobility in suspended graphene. Solid State Commun. 2008, 146, 351–355. [Google Scholar] [CrossRef] [Green Version]
  106. Nair, R.R.; Blake, P.; Grigorenko, A.N.; Novoselov, K.S.; Booth, T.J.; Stauber, T.; Peres, N.M.; Geim, A.K. Fine structure constant defines visual transparency of graphene. Science 2008, 320, 1308. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  107. Balandin, A.A.; Ghosh, S.; Bao, W.; Calizo, I.; Teweldebrhan, D.; Miao, F.; Lau, C.N. Superior thermal conductivity of single-layer graphene. Nano Lett. 2008, 8, 902–907. [Google Scholar] [CrossRef] [PubMed]
  108. Li, J.-L.; Kudin, K.N.; McAllister, M.J.; Prud’homme, R.K.; Aksay, I.A.; Car, R. Oxygen-driven unzipping of graphitic materials. Phys. Rev. Lett. 2006, 96, 176101. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  109. Seifvand, N.; Kowsari, E. Novel TiO2/graphene oxide functionalized with a cobalt complex for significant degradation of NOx and CO. RSC Adv. 2015, 5, 93706–93716. [Google Scholar] [CrossRef]
  110. Renteria, J.D.; Ramirez, S.; Malekpour, H.; Alonso, B.; Centeno, A.; Zurutuza, A.; Cocemasov, A.I.; Nika, D.L.; Balandin, A.A. Strongly anisotropic thermal conductivity of free-standing reduced graphene oxide films annealed at high temperature. Adv. Funct. Mater. 2015, 25, 4664–4672. [Google Scholar] [CrossRef]
  111. Baskoro, F.; Wong, C.-B.; Kumar, S.R.; Chang, C.-W.; Chen, C.-H.; Chen, D.W.; Lue, S.J. Graphene oxide-cation interaction: Inter-layer spacing and zeta potential changes in response to various salt solutions. J. Membr. Sci. 2018, 554, 253–263. [Google Scholar] [CrossRef]
  112. Suk, J.W.; Piner, R.D.; An, J.; Ruoff, R.S. Mechanical properties of monolayer graphene oxide. ACS Nano 2010, 4, 6557–6564. [Google Scholar] [CrossRef]
  113. Shen, J.; Zhu, Y.; Yang, X.; Zong, J.; Zhang, J.; Li, C. One-pot hydrothermal synthesis of graphene quantum dots surface-passivated by polyethylene glycol and their photoelectric conversion under near-infrared light. New J. Chem. 2012, 36, 97–101. [Google Scholar] [CrossRef]
  114. Becerril, H.A.; Mao, J.; Liu, Z.; Stoltenberg, R.M.; Bao, Z.; Chen, Y. Evaluation of solution-processed reduced graphene oxide films as transparent conductors. ACS Nano 2008, 2, 463–470. [Google Scholar] [CrossRef]
  115. Sadhukhan, S.; Ghosh, T.K.; Rana, D.; Roy, I.; Bhattacharyya, A.; Sarkar, G.; Chakraborty, M.; Chattopadhyay, D. Studies on synthesis of reduced graphene oxide (RGO) via green route and its electrical property. Mater. Res. Bull. 2016, 79, 41–51. [Google Scholar] [CrossRef]
  116. Wang, S.; Cole, I.S.; Zhao, D.; Li, Q. The dual roles of functional groups in the photoluminescence of graphene quantum dots. Nanoscale 2016, 8, 7449–7458. [Google Scholar] [CrossRef] [Green Version]
  117. Zhu, S.; Zhang, J.; Qiao, C.; Tang, S.; Li, Y.; Yuan, W.; Li, B.; Tian, L.; Liu, F.; Hu, R. Strongly green-photoluminescent graphene quantum dots for bioimaging applications. Chem. Commun. 2011, 47, 6858–6860. [Google Scholar] [CrossRef]
  118. Sun, Y.; Qi, Y.; Peng, B.; Li, W. NTCP-Reconstituted In Vitro HBV Infection System. In Hepatitis B Virus; Guo, H., Cuconati, A., Eds.; Humana Press: New York, NY, USA, 2016; Volume 1540. [Google Scholar]
  119. Neikov, O.D.; Naboychenko, S.S.; Yefimov, N.A. Handbook of Non-Ferrous Metal Powders: Technologies and Applications; Elsevier: Amsterdam, The Netherlands, 2019. [Google Scholar]
  120. Tajik, S.; Dourandish, Z.; Zhang, K.; Beitollahi, H.; Le, Q.V.; Jang, H.W.; Shokouhimehr, M. Carbon and graphene quantum dots: A review on syntheses, characterization, biological and sensing applications for neurotransmitter determination. RSC Adv. 2020, 10, 15406–15429. [Google Scholar] [CrossRef] [Green Version]
  121. Zhang, Y.; Jiang, Z.; Huang, J.; Lim, L.Y.; Li, W.; Deng, J.; Gong, D.; Tang, Y.; Lai, Y.; Chen, Z. Titanate and titania nanostructured materials for environmental and energy applications: A review. RSC Adv. 2015, 5, 79479–79510. [Google Scholar] [CrossRef]
  122. Jiang, L.C.; Zhang, W.D. Electrodeposition of TiO2 nanoparticles on multiwalled carbon nanotube arrays for hydrogen peroxide sensing. Electroanal. Int. J. Devoted Fundam. Pract. Asp. Electroanal. 2009, 21, 988–993. [Google Scholar]
  123. Cheng, Y.H.; Huang, Y.; Kanhere, P.D.; Subramaniam, V.P.; Gong, D.; Zhang, S.; Highfield, J.; Schreyer, M.K.; Chen, Z. Dual-Phase Titanate/Anatase with Nitrogen Doping for Enhanced Degradation of Organic Dye under Visible Light. Chem. Eur. J. 2011, 17, 2575–2578. [Google Scholar] [CrossRef] [PubMed]
  124. Ren, Y.; Liu, Z.; Pourpoint, F.; Armstrong, A.R.; Grey, C.P.; Bruce, P.G. Nanoparticulate TiO2 (B): An anode for lithium-ion batteries. Angew. Chem. 2012, 124, 2206–2209. [Google Scholar] [CrossRef]
  125. Rongan, H.; Haijuan, L.; Huimin, L.; Difa, X.; Liuyang, Z. S-scheme photocatalyst Bi2O3/TiO2 nanofiber with improved photocatalytic performance. J. Mater. Sci. Technol. 2020, 52, 145–151. [Google Scholar] [CrossRef]
  126. Xu, D.; Li, L.; He, R.; Qi, L.; Zhang, L.; Cheng, B. Noble metal-free RGO/TiO2 composite nanofiber with enhanced photocatalytic H2-production performance. Appl. Surf. Sci. 2018, 434, 620–625. [Google Scholar] [CrossRef]
  127. Wu, M.-C.; Wu, P.-Y.; Lin, T.-H.; Lin, T.-F. Photocatalytic performance of Cu-doped TiO2 nanofibers treated by the hydrothermal synthesis and air-thermal treatment. Appl. Surf. Sci. 2018, 430, 390–398. [Google Scholar] [CrossRef]
  128. Meng, A.; Zhang, J.; Xu, D.; Cheng, B.; Yu, J. Enhanced photocatalytic H2-production activity of anatase TiO2 nanosheet by selectively depositing dual-cocatalysts on {101} and {001} facets. Appl. Catal. B Environ. 2016, 198, 286–294. [Google Scholar] [CrossRef]
  129. Chen, J.; Wang, M.; Han, J.; Guo, R. TiO2 nanosheet/NiO nanorod hierarchical nanostructures: P–n heterojunctions towards efficient photocatalysis. J. Colloid Interface Sci. 2020, 562, 313–321. [Google Scholar] [CrossRef] [PubMed]
  130. Safajou, H.; Khojasteh, H.; Salavati-Niasari, M.; Mortazavi-Derazkola, S. Enhanced photocatalytic degradation of dyes over graphene/Pd/TiO2 nanocomposites: TiO2 nanowires versus TiO2 nanoparticles. J. Colloid Interface Sci. 2017, 498, 423–432. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  131. Lin, Y.; Wu, G.; Yuan, X.; Xie, T.; Zhang, L. Fabrication and optical properties of TiO2 nanowire arrays made by sol–gel electrophoresis deposition into anodic alumina membranes. J. Phys. Condens. Matter 2003, 15, 2917. [Google Scholar] [CrossRef]
  132. Makal, P.; Das, D. Self-doped TiO2 nanowires in TiO2-B single phase, TiO2-B/anatase and TiO2-anatase/rutile heterojunctions demonstrating individual superiority in photocatalytic activity under visible and UV light. Appl. Surf. Sci. 2018, 455, 1106–1115. [Google Scholar] [CrossRef]
  133. Al-Hajji, L.A.; Ismail, A.A.; Al-Hazza, A.; Ahmed, S.A.; Alsaidi, M.; Almutawa, F.; Bumajdad, A. Impact of calcination of hydrothermally synthesized TiO2 nanowires on their photocatalytic efficiency. J. Mol. Struct. 2020, 1200, 127153. [Google Scholar] [CrossRef]
  134. Dou, H.; Long, D.; Rao, X.; Zhang, Y.; Qin, Y.; Pan, F.; Wu, K. Photocatalytic Degradation Kinetics of Gaseous Formaldehyde Flow Using TiO2 Nanowires. ACS Sustain. Chem. Eng. 2019, 7, 4456–4465. [Google Scholar] [CrossRef]
  135. Lim, Y.W.L.; Tang, Y.; Cheng, Y.H.; Chen, Z. Morphology, crystal structure and adsorption performance of hydrothermally synthesized titania and titanate nanostructures. Nanoscale 2010, 2, 2751–2757. [Google Scholar] [CrossRef]
  136. Miao, L.; Tanemura, S.; Toh, S.; Kaneko, K.; Tanemura, M. Fabrication, characterization and Raman study of anatase-TiO2 nanorods by a heating-sol–gel template process. J. Cryst. Growth 2004, 264, 246–252. [Google Scholar] [CrossRef]
  137. Liu, B.; Wang, J.; Yang, J.; Zhao, X. Charge carrier interfacial transfer pathways from TiO2 and Au/TiO2 nanorod arrays to electrolyte and the association with photocatalysis. Appl. Surf. Sci. 2019, 464, 367–375. [Google Scholar] [CrossRef]
  138. Chen, W.; Wang, Y.; Liu, S.; Gao, L.; Mao, L.; Fan, Z.; Shangguan, W.; Jiang, Z. Non-noble metal Cu as a cocatalyst on TiO2 nanorod for highly efficient photocatalytic hydrogen production. Appl. Surf. Sci. 2018, 445, 527–534. [Google Scholar] [CrossRef] [Green Version]
  139. Ge, M.-Z.; Li, S.-H.; Huang, J.-Y.; Zhang, K.-Q.; Al-Deyab, S.S.; Lai, Y.-K. TiO2 nanotube arrays loaded with reduced graphene oxide films: Facile hybridization and promising photocatalytic application. J. Mater. Chem. A 2015, 3, 3491–3499. [Google Scholar] [CrossRef]
  140. Kasuga, T.; Hiramatsu, M.; Hoson, A.; Sekino, T.; Niihara, K. Formation of titanium oxide nanotube. Langmuir 1998, 14, 3160–3163. [Google Scholar] [CrossRef]
  141. Huy, T.H.; Bui, D.P.; Kang, F.; Wang, Y.-F.; Liu, S.-H.; Thi, C.M.; You, S.-J.; Chang, G.-M.; Pham, V.V. SnO2/TiO2 nanotube heterojunction: The first investigation of NO degradation by visible light-driven photocatalysis. Chemosphere 2019, 215, 323–332. [Google Scholar] [CrossRef] [PubMed]
  142. Hou, J.; Zhou, J.; Liu, Y.; Yang, Y.; Zheng, S.; Wang, Q. Constructing Ag2O nanoparticle modified TiO2 nanotube arrays for enhanced photocatalytic performances. J. Alloys Compd. 2020, 849, 156493. [Google Scholar] [CrossRef]
  143. Parida, K.M.; Mohapatra, L. Carbonate intercalated Zn/Fe layered double hydroxide: A novel photocatalyst for the enhanced photo degradation of azo dyes. Chem. Eng. J. 2012, 179, 131–139. [Google Scholar] [CrossRef]
  144. Tobajas, M.; Belver, C.; Rodriguez, J.J. Degradation of emerging pollutants in water under solar irradiation using novel TiO2-ZnO/clay nanoarchitectures. Chem. Eng. J. 2017, 309, 596–606. [Google Scholar] [CrossRef]
  145. Perera, S.D.; Mariano, R.G.; Vu, K.; Nour, N.; Seitz, O.; Chabal, Y.; Balkus, K.J., Jr. Hydrothermal synthesis of graphene-TiO2 nanotube composites with enhanced photocatalytic activity. ACS Catal. 2012, 2, 949–956. [Google Scholar] [CrossRef]
  146. Cho, T.-Y.; Han, C.-W.; Jun, Y.; Yoon, S.-G. Formation of artificial pores in nano-TiO 2 photo-electrode films using acetylene-black for high-efficiency, dye-sensitized solar cells. Sci. Rep. 2013, 3, 1–7. [Google Scholar] [CrossRef] [Green Version]
  147. Wei, N.; Cui, H.; Wang, C.; Zhang, G.; Song, Q.; Sun, W.; Song, X.; Sun, M.; Tian, J. Bi2O3 nanoparticles incorporated porous TiO2 films as an effective p-n junction with enhanced photocatalytic activity. J. Am. Ceram. Soc. 2017, 100, 1339–1349. [Google Scholar] [CrossRef]
  148. Lv, X.; Wang, T.; Jiang, W. Preparation of Ag@AgCl/g-C3N4/TiO2 porous ceramic films with enhanced photocatalysis performance and self-cleaning effect. Ceram. Int. 2018, 44, 9326–9337. [Google Scholar] [CrossRef]
  149. Kim, S.; Chang, H.-K.; Kim, K.B.; Kim, H.-J.; Lee, H.-N.; Park, T.J.; Park, Y.M. Highly Porous SnO2/TiO2 Heterojunction Thin-Film Photocatalyst Using Gas-Flow Thermal Evaporation and Atomic Layer Deposition. Catalysts 2021, 11, 1144. [Google Scholar] [CrossRef]
  150. Low, F.W.; Lai, C.W.; Hamid, S.B.A. Surface modification of reduced graphene oxide film by Ti ion implantation technique for high dye-sensitized solar cells performance. Ceram. Int. 2017, 43, 625–633. [Google Scholar] [CrossRef]
  151. Jo, W.-K.; Kang, H.-J. Titanium dioxide–graphene oxide composites with different ratios supported by Pyrex tube for photocatalysis of toxic aromatic vapors. Powder Technol. 2013, 250, 115–121. [Google Scholar] [CrossRef]
  152. Chowdhury, S.; Parshetti, G.K.; Balasubramanian, R. Post-combustion CO2 capture using mesoporous TiO2/graphene oxide nanocomposites. Chem. Eng. J. 2015, 263, 374–384. [Google Scholar] [CrossRef]
  153. Zhang, C.; Chaudhary, U.; Lahiri, D.; Godavarty, A.; Agarwal, A. Photocatalytic activity of spark plasma sintered TiO2–graphene nanoplatelet composite. Scr. Mater. 2013, 68, 719–722. [Google Scholar] [CrossRef]
  154. Liang, D.; Cui, C.; Hu, 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]
  155. Shen, J.; Yan, B.; Shi, M.; Ma, H.; Li, N.; Ye, M. One step hydrothermal synthesis of TiO2-reduced graphene oxide sheets. J. Mater. Chem. 2011, 21, 3415–3421. [Google Scholar] [CrossRef]
  156. Chang, B.Y.S.; Huang, N.M.; An’amt, M.N.; Marlinda, A.R.; Norazriena, Y.; Muhamad, M.R.; Harrison, I.; Lim, H.N.; Chia, C.H. Facile hydrothermal preparation of titanium dioxide decorated reduced graphene oxide nanocomposite. Int. J. Nanomed. 2012, 7, 3379. [Google Scholar]
  157. Li, W.; Wang, F.; Feng, S.; Wang, J.; Sun, Z.; Li, B.; Li, Y.; Yang, J.; Elzatahry, A.A.; Xia, Y. Sol–gel design strategy for ultradispersed TiO2 nanoparticles on graphene for high-performance lithium ion batteries. J. Am. Chem. Soc. 2013, 135, 18300–18303. [Google Scholar] [CrossRef]
  158. Li, W.; Wang, F.; Liu, Y.; Wang, J.; Yang, J.; Zhang, L.; Elzatahry, A.A.; Al-Dahyan, D.; Xia, Y.; Zhao, D. General strategy to synthesize uniform mesoporous TiO2/graphene/mesoporous TiO2 sandwich-like nanosheets for highly reversible lithium storage. Nano Lett. 2015, 15, 2186–2193. [Google Scholar] [CrossRef]
  159. Williams, G.; Seger, B.; Kamat, P.V. TiO2-graphene nanocomposites. UV-assisted photocatalytic reduction of graphene oxide. Acs Nano 2008, 2, 1487–1491. [Google Scholar] [CrossRef]
  160. Li, B.; Zhang, X.; Li, X.; Wang, L.; Han, R.; Liu, B.; Zheng, W.; Li, X.; Liu, Y. Photo-assisted preparation and patterning of large-area reduced graphene oxide–TiO2 conductive thin film. Chem. Commun. 2010, 46, 3499–3501. [Google Scholar] [CrossRef]
  161. Cai, C.-J.; Xu, M.-W.; Bao, S.-J.; Ji, C.-C.; Lu, Z.-J.; Jia, D.-Z. A green and facile route for constructing flower-shaped TiO2 nanocrystals assembled on graphene oxide sheets for enhanced photocatalytic activity. Nanotechnology 2013, 24, 275602. [Google Scholar] [CrossRef]
  162. Štengl, V.; Bakardjieva, S.; Grygar, T.M.; Bludská, J.; Kormunda, M. TiO2-graphene oxide nanocomposite as advanced photocatalytic materials. Chem. Cent. J. 2013, 7, 41. [Google Scholar] [CrossRef] [Green Version]
  163. Wang, W.; Yu, J.; Xiang, Q.; Cheng, B. Enhanced photocatalytic activity of hierarchical macro/mesoporous TiO2–graphene composites for photodegradation of acetone in air. Appl. Catal. B Environ. 2012, 119, 109–116. [Google Scholar] [CrossRef]
  164. Jiang, G.; Lin, Z.; Chen, C.; Zhu, L.; Chang, Q.; Wang, N.; Wei, W.; Tang, H. TiO2 nanoparticles assembled on graphene oxide nanosheets with high photocatalytic activity for removal of pollutants. Carbon 2011, 49, 2693–2701. [Google Scholar] [CrossRef]
  165. Trapalis, A.; Todorova, N.; Giannakopoulou, T.; Boukos, N.; Speliotis, T.; Dimotikali, D.; Yu, J. TiO2/graphene composite photocatalysts for NOx removal: A comparison of surfactant-stabilized graphene and reduced graphene oxide. Appl. Catal. B Environ. 2016, 180, 637–647. [Google Scholar] [CrossRef]
  166. Zhang, Y.; Tang, Z.-R.; Fu, X.; Xu, Y.-J. TiO2—Graphene nanocomposites for gas-phase photocatalytic degradation of volatile aromatic pollutant: Is TiO2− graphene truly different from other TiO2—Carbon composite materials? ACS Nano 2010, 4, 7303–7314. [Google Scholar] [CrossRef] [PubMed]
  167. Lambert, T.N.; Chavez, C.A.; Hernandez-Sanchez, B.; Lu, P.; Bell, N.S.; Ambrosini, A.; Friedman, T.; Boyle, T.J.; Wheeler, D.R.; Huber, D.L. Synthesis and characterization of titania− graphene nanocomposites. J. Phys. Chem. C 2009, 113, 19812–19823. [Google Scholar] [CrossRef]
  168. Low, W.; Boonamnuayvitaya, V. Enhancing the photocatalytic activity of TiO2 co-doping of graphene–Fe3+ ions for formaldehyde removal. J. Environ. Manag. 2013, 127, 142–149. [Google Scholar] [CrossRef]
  169. Pallotti, D.; Gesuele, F.; Palomba, M.; Longo, A.; Carotenuto, G.; Maddalena, P.; Lettieri, S. Photoluminescence-based real-time monitoring of graphene oxide photoreduction: Demonstrations and application to graphene oxide/titanium dioxide composites. J. Lumin. 2017, 188, 129–134. [Google Scholar] [CrossRef]
  170. Li, K.; Xiong, J.; Chen, T.; Yan, L.; Dai, Y.; Song, D.; Lv, Y.; Zeng, Z. Preparation of graphene/TiO2 composites by nonionic surfactant strategy and their simulated sunlight and visible light photocatalytic activity towards representative aqueous POPs degradation. J. Hazard. Mater. 2013, 250, 19–28. [Google Scholar] [CrossRef] [PubMed]
  171. Tri, N.L.M.; Jitae, K.; Van Thuan, D.; Huong, P.T.; Al Tahtamouni, T. Improved photocatalytic decomposition of methyl ethyl ketone gas from indoor air environment by using TiO2/graphene oxide. Mater. Res. Express 2019, 6, 105509. [Google Scholar]
  172. Najafi, M.; Kermanpur, A.; Rahimipour, M.R.; Najafizadeh, A. Effect of TiO2 morphology on structure of TiO2-graphene oxide nanocomposite synthesized via a one-step hydrothermal method. J. Alloys Compd. 2017, 722, 272–277. [Google Scholar] [CrossRef]
  173. Fotiou, T.; Triantis, T.M.; Kaloudis, T.; Pastrana-Martínez, L.M.; Likodimos, V.; Falaras, P.; Silva, A.N.M.; Hiskia, A. Photocatalytic Degradation of Microcystin-LR and Off-Odor Compounds in Water under UV-A and Solar Light with a Nanostructured Photocatalyst Based on Reduced Graphene Oxide–TiO2 Composite. Identification of Intermediate Products. Ind. Eng. Chem. Res. 2013, 52, 13991–14000. [Google Scholar] [CrossRef]
  174. Hu, X.; Zhao, Y.; Wang, H.; Cai, X.; Hu, X.; Tang, C.; Liu, Y.; Yang, Y. Decontamination of Cr (VI) by graphene oxide@ TiO2 in an aerobic atmosphere: Effects of pH, ferric ions, inorganic anions, and formate. J. Chem. Technol. Biotechnol. 2018, 93, 2226–2233. [Google Scholar] [CrossRef]
  175. Zhang, H.; Lv, X.; Li, Y.; Wang, Y.; Li, J. P25-graphene composite as a high performance photocatalyst. ACS Nano 2009, 4, 380–386. [Google Scholar] [CrossRef]
  176. Zhang, J.; Zhu, Z.; Tang, Y.; Feng, X. Graphene encapsulated hollow TiO2 nanospheres: Efficient synthesis and enhanced photocatalytic activity. J. Mater. Chem. A 2013, 1, 3752–3756. [Google Scholar] [CrossRef]
  177. Wang, P.; Wang, J.; Wang, X.; Yu, H.; Yu, J.; Lei, M.; Wang, Y. One-step synthesis of easy-recycling TiO2-rGO nanocomposite photocatalysts with enhanced photocatalytic activity. Appl. Catal. B Environ. 2013, 132, 452–459. [Google Scholar] [CrossRef]
  178. Ng, Y.H.; Lightcap, I.V.; Goodwin, K.; Matsumura, M.; Kamat, P.V. To what extent do graphene scaffolds improve the photovoltaic and photocatalytic response of TiO2 nanostructured films? J. Phys. Chem. Lett. 2010, 1, 2222–2227. [Google Scholar] [CrossRef]
  179. Fernandez-Ibanez, P.; Polo-López, M.; Malato, S.; Wadhwa, S.; Hamilton, J.; Dunlop, P.; D’sa, R.; Magee, E.; O’shea, K.; Dionysiou, D. Solar photocatalytic disinfection of water using titanium dioxide graphene composites. Chem. Eng. J. 2015, 261, 36–44. [Google Scholar] [CrossRef]
  180. Yang, X.; Qin, J.; Jiang, Y.; Li, R.; Li, Y.; Tang, H. Bifunctional TiO2/Ag3 PO4/graphene composites with superior visible light photocatalytic performance and synergistic inactivation of bacteria. Rsc Adv. 2014, 4, 18627–18636. [Google Scholar] [CrossRef]
  181. Chang, Y.-N.; Ou, X.-M.; Zeng, G.-M.; Gong, J.-L.; Deng, C.-H.; Jiang, Y.; Liang, J.; Yuan, G.-Q.; Liu, H.-Y.; He, X. Synthesis of magnetic graphene oxide—TiO2 and their antibacterial properties under solar irradiation. Appl. Surf. Sci. 2015, 343, 1–10. [Google Scholar] [CrossRef]
  182. Cruz-Ortiz, B.R.; Hamilton, J.W.; Pablos, C.; Díaz-Jiménez, L.; Cortés-Hernández, D.A.; Sharma, P.K.; Castro-Alférez, M.; Fernández-Ibañez, P.; Dunlop, P.S.; Byrne, J.A. Mechanism of photocatalytic disinfection using titania-graphene composites under UV and visible irradiation. Chem. Eng. J. 2017, 316, 179–186. [Google Scholar] [CrossRef]
  183. Giampiccolo, A.; Tobaldi, D.M.; Leonardi, S.G.; Murdoch, B.J.; Seabra, M.P.; Ansell, M.P.; Neri, G.; Ball, R.J. Sol gel graphene/TiO2 nanoparticles for the photocatalytic-assisted sensing and abatement of NO2. Appl. Catal. B Environ. 2019, 243, 183–194. [Google Scholar] [CrossRef]
  184. Xu, M.; Hu, X.; Wang, S.; Yu, J.; Zhu, D.; Wang, J. Photothermal effect promoting CO2 conversion over composite photocatalyst with high graphene content. J. Catal. 2019, 377, 652–661. [Google Scholar] [CrossRef]
  185. Dai, X.; Wang, Y.; Wang, X.; Tong, S.; Xie, X. Polarity on adsorption and photocatalytic performances of N-GR/TiO2 towards gaseous acetaldehyde and ethylene. Appl. Surf. Sci. 2019, 485, 255–265. [Google Scholar] [CrossRef]
  186. Roso, M.; Boaretti, C.; Pelizzo, M.G.; Lauria, A.; Modesti, M.; Lorenzetti, A. Nanostructured photocatalysts based on different oxidized graphenes for VOCs removal. Ind. Eng. Chem. Res. 2017, 56, 9980–9992. [Google Scholar] [CrossRef]
Figure 1. SCI journals focusing on the circular economy, renewable energy, and photocatalysis. Environment-related represents the articles associated with the categories of environmental engineering, environmental science, or environmental studies.
Figure 1. SCI journals focusing on the circular economy, renewable energy, and photocatalysis. Environment-related represents the articles associated with the categories of environmental engineering, environmental science, or environmental studies.
Nanomaterials 11 03195 g001
Figure 2. Scheme illustration of a particulate photocatalyst for the mineralization of pollutants in the environment.
Figure 2. Scheme illustration of a particulate photocatalyst for the mineralization of pollutants in the environment.
Nanomaterials 11 03195 g002
Figure 3. Chemical structures of (a) graphene; (b) graphene oxide (GO); (c) reduced GO (rGO); and (d) graphene quantum dots (GQDs).
Figure 3. Chemical structures of (a) graphene; (b) graphene oxide (GO); (c) reduced GO (rGO); and (d) graphene quantum dots (GQDs).
Nanomaterials 11 03195 g003
Figure 4. Score evaluation of different methods for graphene synthesis (ME, OER, LPE, CVD, AP, UZ, EG, SFGP, SFT, and TOS represent mechanical exfoliation, oxidative exfoliation-reduction, liquid-phase exfoliation, chemical vapor deposition, arc plasmas, unzipping of carbon nanotubes, epitaxial graphene growth, substrate-free gas-phase synthesis, soft-hard template approach, and total organic synthesis, respectively) (Scoring system: 1-low, 2-medium, and 3-high) [58].
Figure 4. Score evaluation of different methods for graphene synthesis (ME, OER, LPE, CVD, AP, UZ, EG, SFGP, SFT, and TOS represent mechanical exfoliation, oxidative exfoliation-reduction, liquid-phase exfoliation, chemical vapor deposition, arc plasmas, unzipping of carbon nanotubes, epitaxial graphene growth, substrate-free gas-phase synthesis, soft-hard template approach, and total organic synthesis, respectively) (Scoring system: 1-low, 2-medium, and 3-high) [58].
Nanomaterials 11 03195 g004
Figure 5. Scheme illustration of enhanced photocatalysis activity of GFN-TiO2.
Figure 5. Scheme illustration of enhanced photocatalysis activity of GFN-TiO2.
Nanomaterials 11 03195 g005
Table 1. Summary of the methods widely used for the synthesis of TiO2.
Table 1. Summary of the methods widely used for the synthesis of TiO2.
MethodMechanismPhase of FormationPros and ConsReference
Sol-gelHydrolysis and condensation of TiCl4 or an organometallic compoundAmorphous and rutileHigh purity, fine particle sizes, good size distribution, high surface areas, but the ease of agglomeration and long reaction time[25,26,27,28]
HydrothermalPrecipitation of TiO2 from aqueous solution at elevated temperature and pressureAnatase and rutileHigh crystallinity, low defects, fine particle size, good size distribution, limited agglomeration, control of crystal shape by temperature adjustment, but relatively higher costs[25,29,30]
SolvothermalPrecipitation of TiO2 from organic solution at elevated temperature and pressureAnatase and rutileHigh crystallinity, low defects, suitability for materials unstable at high temperature, but organic solvents needed[25,31]
Micelle and inverse micelle Aggregation of TiO2 in a liquid colloidAmorphousHigh crystallinity, low defects, fine particle sizes, but relatively high costs and high crystallization temperatures[25,32]
Flame pyrolysisCombustion of TiCl4 with oxygen; used in industrial processesAnatase and rutileRapid and mass production, but high energy needed and ease of rutile formation[25,33,34]
Table 2. Comparison of different polymorphic forms of TiO2.
Table 2. Comparison of different polymorphic forms of TiO2.
PropertiesAnataseBrookiteRutile
Crystal structureTetragonalOrthorhombicTetragonal
Density (g/cm3)3.793.994.13
Band gap (eV)3.2 a~3.2 b3.0 c
Light absorption (nm)<390-<415
Dielectric constant6.047.896.62
Lattice energy (kJ/mol) d24.7518.530
Surface enthalpy (J/m2) e1.341.661.93
Photocatal. activity (mol/h) f3.5 × 10−5-1.1 × 10−5
Effective electron mass (me*/m0) g0.09480.09491.4640
Effective hole mass (mh*/m0) g0.19950.56200.4345
Ti-O bond length (Å) h1.94 (shorter); 1.97 (longer)1.87–2.041.95 (shorter); 1.98 (longer)
O-Ti-O bond angle (degree)77.7; 92.677.0–10581.2; 90.0
Reference sources: a [41]; b [42]; c [43]; d [44]; e [45]; and f [40]; and g [46]. The other numbers are sourced from [47,48]. h Anatase and rutile TiO2 have two different interatomic distances, while brookite TiO2 has six different Ti-O bonds with a distance ranging from 1.87 to 2.04 A.
Table 3. Comparison of different synthesis methods of GFNs.
Table 3. Comparison of different synthesis methods of GFNs.
MethodMajor ApproachPros and ConsCost
GrapheneMechanical exfoliationMicro-mechanical cleavage, sonication, ball milling, and fluid dynamicsStraightforward and eco-friendly processes, fine product qualities, but relatively higher costs and limits of scalable productionHigh
Oxidative exfoliation-reductionChemical reduction, thermal reduction, and electrochemical reductionStraightforward processes, cost-effectiveness, scalable production, but possible structural damage due to mal exfoliation, and potential use of hazardous chemicalsLow
Liquid phase exfoliationSonication with proper solventsStraightforward and eco-friendly processes (solvents recyclable), fine product qualities, scalable production, but parameters (e.g., solvent and ultra-sonication) critical to avoid physical deformation and defectsModerate
Chemical vapor deposition (CVD)Thermal CVD, plasma-enhanced CVD, and thermal decompositionHighly connected products with low defects and high surface areas, but relatively higher costs, limited yields, and high technical thresholdsModerate
Graphene oxideBrodieGraphite + H2CO3 (C/O ratio = 2.23)Adjustable oxidation states, but potentials of long reaction time and production of explosive ClO2 and acid fog Low
StaudenmaierGraphite + HNO3 (fuming) + H2SO4 + KClO3 (C/O ratio = 2.52)Adjustable oxidation state, but long reaction time and low temperatures to avoid exothermic reactionsLow
HofmannGraphite + HNO3 + H2SO4 + KClO3 (C/O ratio = 2.52)Low
HummersGraphite+NaNO3 +H2SO4+ KMnO4 (C/O ratio = 2.1-2.9)Safe and fast reactions, but more parameters to controlLow
Reduced graphene oxideChemical reductionVarious reductantsFine product qualities, scalable production, but the potential of using hazardous reductants. Lower product qualities and removal of excess chemicals with the use of green reductantsLow
Thermal reduction1000–1100 °C for 30–45 s in
the absence of air
Straightforward and eco-friendly processes, cost-effectiveness, but high capital costs and energy neededModerate
Electrochemical reductionThe cathodic potential of 1–1.5 VLow-defect products, rapid and eco-friendly processes, cost-effectiveness, but lower reduction levels and limited scalable productionLow
Microwave and photo-reductionMicrowave reaction with visible or UV lightFast reactions, no chemicals needed, and high yield efficienciesLow
Graphene quantum dotTop-downHydrothermal synthesis, solvent thermal method, chemical oxidation, electrochemical exfoliation, electron beam lithography, microwave-assisted method, and ultra-sonication exfoliationScalable production, but difficulty of effective size controlHigh
Bottom-upSoft template method, acid- and solvent-free synthesis, and metal catalysisEffective size control, but long reaction time and limited scalable productionHigh
Table 5. Selected physicochemical properties of TiO2-containing composites prepared in different dimensions.
Table 5. Selected physicochemical properties of TiO2-containing composites prepared in different dimensions.
DimensionStructureSurface AreaLight Absorption WavelengthCurrent DensityReference
0Nanoparticle (less than 100 nm)180–250 m2/gUltraviolet to infrared radiationNot available[121,122,123,124]
1Nanofiber52–55 m2/g<510 nm0.06 mA/cm2[125,126,127]
Nanowire61.5–92.6 m2/g250–540 nm1.6 mA/cm2[130,131,132,133,134]
Nanorod104.6 m2/g~380 nm0.8 mA/cm2[135,136,137,138]
Nanotube400 m2/g<500 nm0.02 mA/cm2[139,140,141,142]
2Nanosheet31–146 m2/g200–900 nm0.03 mA/cm2[128,129]
3Porous film36.4–70.8 m2/g200–700 nm18.54 mA/cm2[146,147,148,149]
Table 6. The synthesis methods of TiO2-GFN composites.
Table 6. The synthesis methods of TiO2-GFN composites.
MethodsCrystal FormGFN RatioPros and ConsReference
Ion implantationAnataseNot availableFast production, few interfacial defects, great optical character, but high energy costs[150]
Colloidal blending processAnatase or rutileadjustableAging at room temperature and vacuum drying needed[151,152]
Spark plasma sinteringRutile1% v/vFast production, but high energy costs and increased rutile form[153]
Hydrothermal methodAnataseadjustableAdjustable doping ratio, but high pressure needed[154,155,156]
Sol-gel methodAnatase48% w/wAging at room temperature, long reaction time, and calcination needed[157]
HydrolysisAnatase16% w/wGreat heterogeneous nucleation, but longer reaction time and calcination needed[158]
UV-assisted photo-reductionNot availableNot availableFast production and few collapses during reduction, but extra light source needed[159,160]
In-situ assemblyAnataseNot availableNo calcination and full anatase formation, but long synthesis time[161,162]
Table 7. Methods and outcomes of characterization of TiO2-graphene composites.
Table 7. Methods and outcomes of characterization of TiO2-graphene composites.
CategoryTechnologyDescriptionRef.
MorphologySEMSpherical and non-spherical (platelet- or flower-like) shapes were observed with low and high GFN contents, respectively. [151,163,164,165,166,167]
TEMA fine dispersion of TiO2 in GFN with low- and nano-dimensions was reported. [163,165,166,167]
AFMThe thickness of GFN-TiO2 was increased to a scale of μm after preparation. [164]
Chemical constitutionFTIRThe peak of Ti-O-Ti at 400–900 cm−1 was broadened or shifted by the influence of Ti-O-C. The signals of carbonyl and epoxy groups were reduced. [151,165,168]
XPSThe formation of C-Ti, O=C-O-Ti, and C-O-Ti bonds in GFN-TiO2 was observed. [163] [164]
XRDThe signals due to the presence of anatase and rutile were reported. [151,163,164,165,166,168]
RamanThe signals of both TiO2 and GFN were reported. The D/G intensity ratio of GFN-TiO2 was higher than that of GFN.[163,164,165]
EPRThe formation of hydroxyl and superoxide radical species was observed in GFN-TiO2.[166]
Physicochemical propertiesZeta potentialThe zeta potential of GFN-TiO2 ranged between those of GFN and TiO2.[164]
TGAThe irregular mass loss occurred at high temperatures. [164]
BETThe surface area of GFN-TiO2 was significantly increased at a certain ratio of GFN to TiO2.[151,163,164,165,168]
ACMThe current density of GFN-TiO2 was significantly increased at a certain ratio of GFN to TiO.[168]
PLThe time dynamics of the TiO2-induced photoreduction of GO were observed. [169]
UV-VisA shift to larger wavelengths in the absorption edge was observed, indicating bandgap narrowing.[151,164,165,166,168]
Table 8. Properties of TiO2-GFN prepared for photocatalysis and battery storage in various studies.
Table 8. Properties of TiO2-GFN prepared for photocatalysis and battery storage in various studies.
MaterialsAverage Size (nm)Functional GroupBandgap (eV)Wavelength (nm)Surface Area (m2/g)Reference
Graphene-TiO23.8C-O, C=O, O=C-O, and O-TiNA 1600176[170]
Graphene-TiO2~6C-O and O-C=ONANA252[158] 2
GO-TiO2NAC-O, Ti-O-Ti, Ti-O-C, and OHNA~80069.2[151]
GO-Co-TiO2NAC-O, C-N, O-C=O2.77421206[109]
GO-TiNANA2.9~55068.9[171]
rGO-TiO235NANA~360212.75[172]
rGO-TiO2~8NANANA229[157] 2
1 NA denotes not available. 2 The materials were prepared for battery storage.
Table 9. Removal of water-phase pollutants by GFN-TiO2 in selected studies.
Table 9. Removal of water-phase pollutants by GFN-TiO2 in selected studies.
PollutantCatalystLight SourceRemovalRef.
InorganicCr(VI) (0.2 mM)GO-TiO2 (0.5 g/L)254 nm, 20 W, UV lamp90%[164]
Cr(VI)(10 mg/L)GO-TiO2 (0.5 g/L)365 nm, 8 W, UV lamp99%[174]
OrganicMethylene blue (0.01 g/L)Graphene-TiO2 (0.75 g/L)365 nm, 100 W, high-pressure Hg lamp
>400 nm, 500W, Xe lamp
85%
65%
[175]
Rhodamine B (20 mg/L)Graphene-TiO2 (0.1 g/L)11 W, low-pressure Hg lamp91%[176]
Rhodamine B (20 mg/L)
Norfloxacin (20 mg/L)
Aldicarb (10.5 mg/L)
Graphene-TiO2 (1 g/L)>400 nm, Xe lamp79.7%
86.2%
36.8%
[170]
Malachite green oxalate (13.1 mg/L)GO-TiO2 (0.2 g/L)450 W, water-cooled Hg lamp80%[145]
Phenol (10 mg/L)rGO-TiO2 (5 g/L)310-400 nm, UV lampNot given[177]
2,4-D (15 mM)rGO-TiO2 (film)<320 nm, 450 W, Xe lamp~87%[178]
BiologicalE. coli (106 CFU/mL), F. solani spores (103 CFU/mL)rGO-TiO2 (0.5 g/L)Sunlight~100%[179]
E. coli, S.aureus, S.typhi, P. aeruginosa, B. subtilis, B. pumilus (106 CFU/mL)Graphene-Ag3PO4-TiO2>420 nm, 350 W, Xe lamp~100%[180]
E. coli (105–106 CFU/mL)GO-TiO2 (0.2 g/L)Xe lamp~100%[181]
E. coli (106 CFU/mL)rGO-TiO2 (18 mg/L)>285 nm, UV-visible light; >420 nm, visible light~100%[182]
Table 10. Removal of air-phase pollutants by GFN-TiO2 in selected studies.
Table 10. Removal of air-phase pollutants by GFN-TiO2 in selected studies.
PollutantCatalystLight SourceHumidity or Flow RateRemovalRef.
InorganicNOx (1 ppm)Graphene-TiO2
rGO-TiO2
15 W, UVA
8 W, visible light
50% humidity, 3 L/min42%
49%
[165]
NOx (200 ppb)Graphene-TiO2280–780 nm, 300 W, solar lamp1 L/min77%[183]
CO (50 ppm)
NOx (1 ppm)
Graphene-TiO28 W, UV lamp0.2 L/min46%
51%
[109]
OrganicAcetone (300 ± 20 ppm)Graphene-TiO2365 nm, 15 W, UV lamp1 L/min~60%[163]
Acetaldehyde (500 ppm)
Ethylene (50 ppm)
Graphene-TiO2260 W, fluorescent lamp
500 W, Xenon lamp
20 cm3/min~82%
~90%
[185]
Benzene (250 ppm)Graphene-TiO2254 nm, 4 W, UV lamp20 mL/min6.4%[166]
Formaldehyde (3000 ppm)Graphene-TiO2365 nm, 8 W, black light blue lamp
>420 nm, 8 W, fluorescent lamp
Not specified50.3%
25.5%
[168]
Methanol (4,000 ppm)Graphene-TiO2
GO-TiO2
rGO-TiO2
254 nm, 16 W, UV lamp155 cm3/min80%
99%
99%
[186]
BTEX (1 ppm)GO-TiO2400–720 nm, 8 W, daylight lamp55% humidity, 1 L/min96%[151]
MEKT (30 ppm)GO-TiO280 W, Xe lamp40% humidity, 50 mL/min96.8%[171]
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Lin, C.-H.; Chen, W.-H. Graphene Family Nanomaterials (GFN)-TiO2 for the Photocatalytic Removal of Water and Air Pollutants: Synthesis, Characterization, and Applications. Nanomaterials 2021, 11, 3195. https://doi.org/10.3390/nano11123195

AMA Style

Lin C-H, Chen W-H. Graphene Family Nanomaterials (GFN)-TiO2 for the Photocatalytic Removal of Water and Air Pollutants: Synthesis, Characterization, and Applications. Nanomaterials. 2021; 11(12):3195. https://doi.org/10.3390/nano11123195

Chicago/Turabian Style

Lin, Chih-Hsien, and Wei-Hsiang Chen. 2021. "Graphene Family Nanomaterials (GFN)-TiO2 for the Photocatalytic Removal of Water and Air Pollutants: Synthesis, Characterization, and Applications" Nanomaterials 11, no. 12: 3195. https://doi.org/10.3390/nano11123195

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