Next Article in Journal
Mechanical Performance of Granite Fine Fly Dust-Filled Basalt/Glass Polyurethane Polymer Hybrid Composites
Next Article in Special Issue
Preparation of Large Conjugated Polybenzimidazole Fluorescent Materials and Their Application in Metal Ion Detection
Previous Article in Journal
Extraction of Microcrystalline Cellulose from Washingtonia Fibre and Its Characterization
Previous Article in Special Issue
Enhanced Chemical and Electrochemical Stability of Polyaniline-Based Layer-by-Layer Films
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Advanced Photocatalysts Based on Conducting Polymer/Metal Oxide Composites for Environmental Applications

1
Alan G. MacDiarmid Energy Research Institute, Chonnam National University, Gwangju 61186, Korea
2
Advanced Institute of Science and Technology, University of Danang, Danang 50000, Vietnam
3
Industry-University Cooperation Foundation, Chonnam National University, Gwangju 61186, Korea
4
Department of Polymer Engineering, Graduate School, Chonnam National University, Gwangju 61186, Korea
5
School of Polymer Science and Engineering, Chonnam National University, Gwangju 61186, Korea
*
Authors to whom correspondence should be addressed.
These authors contributed equally to this work.
Polymers 2021, 13(18), 3031; https://doi.org/10.3390/polym13183031
Submission received: 13 August 2021 / Revised: 5 September 2021 / Accepted: 5 September 2021 / Published: 8 September 2021

Abstract

:
Photocatalysts provide a sustainable method of treating organic pollutants in wastewater and converting greenhouse gases. Many studies have been published on this topic in recent years, which signifies the great interest and attention that this topic inspires in the community, as well as in scientists. Composite photocatalysts based on conducting polymers and metal oxides have emerged as novel and promising photoactive materials. It has been demonstrated that conducting polymers can substantially improve the photocatalytic efficiency of metal oxides owing to their superior photocatalytic activities, high conductivities, and unique electrochemical and optical properties. Consequently, conducting polymer/metal oxide composites exhibit a high photoresponse and possess a higher surface area allowing for visible light absorption, low recombination of charge carriers, and high photocatalytic performance. Herein, we provide an overview of recent advances in the development of conducting polymer/metal oxide composite photocatalysts for organic pollutant degradation and CO2 conversion through photocatalytic processes.

Graphical Abstract

1. Introduction

Photocatalysis plays a critical role in the development of emerging technologies for environmental applications, such as wastewater treatment and CO2 reduction [1,2,3]. Currently, advanced photocatalytic materials consist mainly of metal oxides, such as TiO2, SnO2, ZnO, and spinel ferrites [2,4,5]. Among them, TiO2 is still the most used semiconductor, comprising about 25% of semiconductors used in the photocatalyst field [6]. Nonetheless, TiO2 possesses a wide band-gap energy (~3.2 eV), and, thus, it only absorbs ultraviolet (UV) light, which accounts for only 4% of solar energy [7]. Moreover, TiO2 also shows a fast electron–hole recombination rate [4]. These two drawbacks are common challenges in the development of other metal oxides for photocatalytic applications. Therefore, many strategies have been investigated for tailoring and modulating the light adsorption ability, as well as enhancing the charge separation, in metal oxide semiconductors, including (i) band-gap engineering approaches [8]; (ii) developing co-photocatalysts [9]; (iii) doping with metal nanoparticles (i.e., Au, Pt, Cu) [10]; and (iv) design of different composite photocatalysts [11]. Among them, the development of composite photocatalysts has been regarded as one of the most promising approaches due to several advantages. First, the surface properties of metal oxide-based composite structures can be tuned to achieve mid-band-gap electronic states and produce a high absorption in the visible light spectrum [12]. Metal oxide composites can also enhance photocatalyst stability and enable the fabrication of advanced photocatalysts with novel structures, such as microporous, hollow shells, or hierarchical structures.
Conducting polymers have been commonly used as electrocatalysts and photocatalysts as promising alternatives to traditional inorganic semiconductors in various applications (i.e., energy storage, sensors, and environmental protection) because of their superior photocatalytic activities, good conductivities, and unique electrochemical and optical properties [13,14,15]. Techniques for the preparation of conducting polymers are also simple and easy to scale up for large-scale production by using chemical or electrochemical approaches. In addition, conducting polymers with narrow bandgaps enable the absorption of visible light from the sun [14]. Polyaniline (PANI), poly (3,4-ethylenedioxythiophene) (PEDOT), polypyrrole (PPy), and their derivatives are common conducting polymers used in photocatalytic applications for wastewater treatment and CO2 reduction. However, there are some problems that severely limit the practical applications of conducting polymers. Most conducting polymers exhibit a low mechanical strength, high brittleness and poor processability [16]. In order to overcome these limitations, the fabrication and development of organic–inorganic composite photocatalysts based on organic conducting polymers and metal oxides has been considered as a promising approach for environmental applications [17]. In these composites, conducting polymers play a role as the supporting matrix for the intercalation of catalytically metal oxide nanoparticles and photosensitizers for the enhancement of light adsorption in the visible spectrum, which improves the photocatalytic performance and stability of the composite photocatalysts.
To date, many studies on the synthesis and development of novel conducting polymer/metal oxide composites have been reported. However, there has not yet been a review article that provides a systematic overview of the reported studies on these composites. Therefore, this review aims to summarize recent advancements in photocatalytic conducting polymer/metal oxide composites for environmental applications. We also explain and discuss the photocatalytic mechanisms and outline some problems related to the use of these composites in practical applications.

2. Synthesis and Properties of Conjugated Polymer/Metal Oxide Composites

Composites of conjugated polymers (CPs) and transitional metal oxides exhibit a significant improvement in photocatalytic performance [14,18,19]. The hybrid material possesses a high photoresponse and delayed recombination of charge carriers due to distinct exciton-plasmon interactions and a higher surface area for light absorption. Consequently, composites of CPs and wide band-gap inorganic semiconductors, especially metal oxides, have been attracting attention for their potential use in photocatalytic and photoelectric conversion applications. It has been reported that CP/metal oxide composites often act as p-n junction semiconductors that can be tailored by combining a p-type CP with an n-type metal oxide semiconductor in order to overcome the problems related to a high rate of electron–hole recombinations, poor activation with visible light, leaching and thermal decomposition. In combination with metal oxides, PANI, PEDOT, and PPy are widely used to prepare composite photocatalysts for the photocatalysis processes needed in environmental applications and CO2 reduction, as summarized in Table 1 and discussed in following sections.
In general, there are two common types of the CP/metal oxide composite-based photocatalysts: (i) binary composite and (ii) ternary composite. Binary composites contain one transition metal oxide and one CP [21]. Ternary composites are made of one binary metal oxide composite (e.g., TiO2/Fe3O4 and Cu2O/ZnO) and one CP or one metal oxide and one binary composite of CP with another material, such as carbon nanotube and graphene [39].

2.1. Binary Composites of CPs and Metal Oxides

2.1.1. PANI/Metal Oxide Composites

Polyaniline (PANI) is one of the most common CPs in photocatalytic applications due to its high stability, high processability, and tunability of the conducting and optical properties [42]. Generally, PANI has a prolonged alternate σ and π bond electronic cloud system, resulting in a significant energy band gap of 2.8 eV [43]. When irradiated by UV–Vis and ultraviolet light, PANI can work as an extraordinary e and h (donor-acceptor) photosensitizer [44]. PANI can be blended with a wide range of metal oxides, such as TiO2, ZnO, Fe3O4, SnO2, Sn3O4, and spinel ferrites to prepare composite photocatalysts.
PANI/TiO2 composites: Due to a wideband gap, TiO2 primarily absorbs UV-light, which accounts for only about 3–5% of sunlight, and limits its photocatalytic ability in the visible region [45,46]. However, PANI exhibits high absorption coefficients in the visible light range, charge carriers with a high mobility, non-toxicity, and low-cost synthesis, and in addition is an excellent electron donor [47]. As a result, PANI is often used to delay the recombination of electron–hole pairs, enhance the charge separation efficiency, and improve the photocatalysis efficiency of TiO2 [21,22]. Generally, PANI/TiO2 composites are often fabricated through an in-situ chemical oxidation process [48]. The photocatalytic activity of PANI/TiO2 can be significantly reduced due to aggregations caused by TiO2 particle collisions. However, the aniline molecules tend to create a barrier to the aggregation processes of TiO2 nanoparticles and PANI protects the TiO2 surface from blockage by intermediates. It has been supposed that the synergetic effect between PANI and TiO2 increases the photocatalytic activity of the obtained composites. It has been reported that there are different TiO2 nanostructures that can be incorporated with PANI, such as nanorods, mesoporous, nanobelts, and nanotubes [20,21,22,23]. The band-gap energy of PANI/TiO2 composites can be reduced to 2.77–3.1 eV.
PANI/ZnO composite: Like TiO2, ZnO is also a widely investigated semiconductor due to its abundance, low cost, and low toxicity [49]. Nonetheless, the recombination of photogenerated charge carriers remains a dominant barrier which significantly limits the practical applications of ZnO, especially on a large scale. Defect-related mediation has been considered as a potential approach in order to modulate the activity of ZnO, and defects on the outer surface of ZnO which can boost its photocatalytic activity remarkably in degradation and photoreduction reactions [50,51]. Pei et al. successfully developed a composite of PANI and defect-rich ZnO using the chemisorption method [52]. The monomolecular-layered PANI can stabilize the surface of ZnO. Additionally, it is expected that a surface of ZnO coated by PANI molecules results in a PANI-based ZnO catalyst with higher photocatalytic activity, while preventing the photo-corrosion of ZnO, even for the monomolecular form of PANI [24]. Under visible-light irradiation, PANI molecules generate a π-π* transition and deliver the excited electrons to the conduction band (CB) of ZnO, which reduces the band-gap energy of ZnO [25]. The PANI/ZnO composite photocatalysts are expected to be promising candidates for the design of high-activity, high-stability, and visible-light-driven photocatalysts in the future.
PANI/magnetic iron oxide composites: PANI and magnetic iron oxide hybrid materials have been demonstrated to be good candidates for the photodegradation of organic dyes in wastewater, due to their electrical conductivity properties associated with superparamagnetism [26]. For example, Alves et al. prepared a composite of PANI and magnetic iron oxide by in-situ chemical polymerization for the adsorption/photodegradation of blue methylene dye. In this photocatalyst, iron oxide nanoparticles (10–15 nm) were embedded in/on the polymer matrix and the synergistic effects of the iron oxide particles and polymer phases were considered to be responsible for the photocatalytic action and the high absorption behavior. In another study, Yang et al. also developed a composite of PANI and Fe3O4 nanoparticles, in which the polymerization of aniline was catalyzed by Fe3O4 nanoparticles [53]. In general, PANI/Fe3O4 composites often have a core/shell structure (Figure 1a) [54].
PANI/Tin oxide composites: SnO2 is considered to be a potential alternative to TiO2 owing to being a better electron acceptor and having a more positive CB potential [55]. However, SnO2 has showed low efficiency in the utilization of solar energy and low photocatalytic efficiency because of the fast recombination of its photogenerated electrons and holes [56,57]. It has been indicated that PANI has a matching electronic band structure with SnO2 and that they can be combined to obtain a type-II SnO2/PANI heterojunction. In this composite, PANI transfers its photogenerated electrons to the CB of SnO2; thus, it plays an important role as a photosensitizer for SnO2 under visible light [58]. Moreover, the metal oxide and the hydroxyl group of SnO2 can be altered slightly with the substitution of PANI. Compared with bare SnO2 nanoparticles, the crystallite size of SnO2 in the composite can be significantly decreased, while its surface area was increased due to the inclusion of PANI [27] Consequently, the substitution of PANI reduced the reflectance and band-gap energy (3.1–2.7 eV) of SnO2, resulting in the SnO2/PANI composite working effectively in the visible light range [59]. In addition to common PANI/SnO2 composites, Manfei et al. recently reported the coupling of a p-type PANI with an n-type Sn3O4 for photocatalytic applications [28]. In that study, a Sn3O4 nanosheet composite was modified by PANI nanofibers and a p–n PANI/Sn3O4 heterojunction was successfully prepared via a mechanical milling method. The photocatalytic activity and stability of the PANI/Sn3O4 composite were highly improved compared with single Sn3O4.
PANI/spinel ferrite composites: Spinel ferrite nanoparticles (SFNPs) are defined as metal oxides with the spinel structure, with the general formula of MFe2O4, where M = divalent cations [60]. The physicochemical properties of SFNPs depend mainly on the types, amounts, and positions of the M cations in the crystallographic structure. MFe2O4 nanoparticles exhibit some advantages, such as stability, biocompatibility, low-cost, excellent magnetic properties, and easy separation [61]. Therefore, they have been widely used in the development of binary nanocomposites as photocatalysts for the photodegradation of pollutants. Several SFNPs, including CoFe2O4, CuFe2O4, MnFe2O4, NiFe2O4 and ZnFe2O4 have been reported as potential candidates for the development of composite photocatalysts with CPs [62,63,64]. Combinations of MFe2O4 with different concentrations of CPs have been demonstrated to be an effective approach for improving the photocatalytic performance, with PANI being especially effective. For example, Kim et al. fabricated CoFe2O4/PANI hollow core-double shell nanostructures as a composite photocatalyst using the electrospinning technique and in-situ chemical oxidative polymerization (Figure 1b) [29]. The results showed that owing to the heterojunction built between CoFe2O4 and PANI, the hollow CoFe2O4/PANI composite easily captured visible light and exhibited effective charge separation, which resulted in a significant improvement in visible light photocatalysis.

2.1.2. Composites of PEDOT and Metal Oxides

PEDOT is known as a conducting polymer that exhibits a narrow band gap (E = 1.69 eV) and an excellent ability to absorb light in the visible and near infrared regions [65]. PEDOT based photocatalysts exhibit good stability, good recyclability, and reusability [30]. Therefore, PEDOT has been commonly coupled with a wide range of metal oxides as composite photocatalysts. Among them, TiO2 and ZnO are often combined with PEDOT in the preparation of composites for photocatalytic applications.
PEDOT/TiO2 composite: PEDOT is regarded as an attractive CP for coupling with TiO2 in composite materials for visible-light-driven photocatalytic applications. Similar to PANI, it can be photoexcited under visible-light irradiation to transfer electrons into the CB of TiO2, which leads to effectively separate holes (h+) and electrons (e) and increases the number of photoexcited charges available to drive photoreactions substantially [30]. The PEDOT/TiO2 composite was developed to overcome disadvantages related to the lower photon transport of the TiO2 surface by incorporation into the polymer layer. Generally, the PEDOT coating thickness is considered to be an important parameter during the fabrication process and must be controlled, so that photogenerated charge carriers can be easily transported from the external polymer interface to the inner TiO2 layer [66]. Recently, Liu et al. demonstrated the enhanced photocatalytic performance of PEDOT to TiO2 nanofibers by improving the rate of transformation of photogenerated holes (Figure 2) [67]. The PEDOT/TiO2 nanofiber composite was fabricated via electrospinning and calcination to form TiO2, followed by the introduction of PEDOT using vapor phase polymerization.
PEDOT/ZnO composite: PEDOT was used to increase the photocatalytic activity of ZnO due to its efficient electron donor and good electron transporters upon visible-light irradiation. For instance, Abdiryim et al. introduced a simple solid-state heating method to prepare PEDOT/ZnO nanocomposites in powder form with the content of ZnO varying between 10 and 20 wt% [68]. The photocatalytic activity of these nanocomposites can be enhanced by the incorporation of ZnO nanoparticles under both UV and visible-light irradiation, which can be ascribed to the high charge separation of electron and hole pairs in the obtained composite.

2.1.3. Composites of PPy and Metal Oxides

Owing to superior conductivity, high charge carrier mobility, high absorption coefficient in the visible light, and good environmental stability, polypyrrole (PPy) is one of the most promising candidates for the development of stable photosensitizers to improve the photocatalytic activity and solar light conversion efficiency of metal oxides [69].
PPy/TiO2 composite: PPy/TiO2 nanocomposites have been mainly applied in the photocatalytic degradation of organic species. Currently, there are numerous methods that can be used to synthesize PPy/TiO2 nanocomposites, such as anodic co-deposition [70], self-assembly techniques [71], photo-electrochemical polymerization [72], and hydrothermal methods [31]. For practical environmental applications, however, the in-situ chemical oxidation method has been considered as the most promising technique due to its simplicity, good reproducibility, and possibility for large-scale production [32]. For instance, Gao et al. successfully prepared PPy/TiO2 nanocomposites using a facile chemical oxidation of pyrrole in a prepared TiO2 sol solution [32]. For this composite, a PPy film of about 2–3 nm was coated onto the TiO2 surface, which is supposed to increase the photocatalytic activity for the degradation of rhodamine B and the reduction of CO2. Moreover, PPy nanostructures (i.e., nanofibers and nanospheres), can be synthesized by directly oxidizing pyrrole (Py) monomers in a solution under mild oxidation conditions and a low temperature [73]. Based on this approach, Dimitrijevic et al. developed a simple one-step hydrothermal method for fabricating PPy/TiO2 nanocomposites [31]. According to this method, 4.5 nm TiO2 nanoparticles were electronically coupled to 200−300 nm PPy granules to form a stable composite, which is capable of efficient visible-light photocatalysis. In this composite, PPy molecules act as visible-light photosensitizers, and the photocatalytic activity of the composite increases through the enhanced electron transfer from excited PPy to TiO2 nanoparticles. Importantly, it has been shown that a high concentration of TiO2 nanoparticles used in the composite can significantly increase the photocatalytic efficiency of the PPy/TiO2 composite.
PPy/ZnO composite: With regards to photocatalytic activity, PPy donates photon-induced electrons to ZnO under visible-light irradiation, which results in an improvement in the photocatalytic activity and a reduction in the recombination of charge carriers [74,75]. To create a flexible photocatalytic film, a novel composite of ZnO-microrod arrays and electrodeposited PPy was recently developed (Figure 3) [33]. To prepare the composite film, the upper section of a ZnO microrod is first covered with a very thin PPy shell, while the lower section of the ZnO microrod is coated with a thick PPy base layer. In this composite, the upper PPy shell works as a photosensitizer through the absorption of visible light and then, the conversion of photons into free carriers (i.e., electrons and holes) [76,77,78], whereas the lower PPy base layer will stabilize the ZnO-microrods on the flexible substrate and facilitate the electron transport to the substrate [79]. It has been suggested that the accelerated carrier separation at the ZnO/PPy interface leads to a considerable enhancement in the photocatalytic activity of ZnO/PPy composite films [80]. Moreover, due to the unique structure integrating flexibility, sunlight-driven photocatalytic properties, and high mechanical strength, ZnO/PPy composite films show a high potential for use in flexible electronics and other applications in the environmental field.

2.2. Ternary Composites of CP/Metal Oxides

Based on the above section, it can be concluded that binary composites of CPs and metal oxides can significantly enhance the photocatalytic activity of individual semiconductors in the visible light region. However, recovery and reusability of photocatalysts are also considered as important factors for practical applications. Recently, ternary composites based on CPs and metal oxides have led to new insights into the design and development of novel multicomponent photocatalysts with versatile and extraordinary properties [37,39,41,81,82]. Therefore, the preparation and design of multicomponent nanocomposites for further improving catalytic performances is of great interest. It has been demonstrated that the formation of a Z-scheme heterojunction can effectively improve carrier mobility, while the synergetic interactions of the components could also maintain the redox ability of the generated electrons and holes for a very long time [83]. Currently, the preparation of ternary nanocomposites based on CPs and metal oxides for creating Z-scheme heterojunctions has attracted further attention for improvements in the properties of metal oxides and binary composite photocatalysts due to two superior advantages: (i) inhibition of corrosion and stabilization of CPs; (ii) synergistic enhancement of the three components [37]. It is strongly believed that a ternary nanocomposite of CP, metal oxide, and another compound will show an enhanced photocatalytic activity in terms of a low band-gap energy, minimized recombination rate and strong absorption of visible light due to the synergism effect between the constituents [39].
As mentioned earlier, the combination of PANI and TiO2 can enhance the photo activity of TiO2 into the visible light region. Moreover, incorporating magnetic nanoparticles into the binary composites of PANI/TiO2 ensures that the photocatalyst can easily handle magnetic separation [84] and provides an effective approach for achieving a more efficient charge separation, resulting in increased photocatalytic activity [85,86]. Taking all these points into consideration, Xiong et al. prepared a magnetically recyclable ternary TiO2-CoFe2O4-PANI composite photocatalyst using in-situ oxidative polymerization [34]. The results indicate that ternary TiO2-CoFe2O4-PANI composites show highly enhanced photocatalytic activity for the visible light region compared with binary TiO2-CoFe2O4, CoFe2O4-PANI, or TiO2-PANI composites. Additionally, this ternary composite photocatalyst can be easily separated out and reused by simply using an external magnetic field after the reaction, owing to the good magnetic properties of CoFe2O4 nanoparticles. Using a similar approach, Li et al. also confirmed the enhanced electrical conductivities and photocatalytic activity of ternary ZnFe2O4-TiO2-PANI composites with different amounts of PANI [35]. Moreover, it has been indicated that interactions between individual components in these ternary composites result in an enhancement of their electrical conductivities and photocatalytic activities, which is an especially important purpose of PANI coating. As the mass fraction of aniline was up to 50%, the ternary composite exhibited considerable photocatalytic activity and displayed excellent reusability.
Graphene, which is well-known as a 2D layered hexagonal lattice of carbon nanomaterials [87], is a potential material for increasing the photocatalytic efficiency and stability of composite photocatalysts, due to its superior electronic and transport properties, high surface area, and zero band-gap energy [88]. In addition, it has been demonstrated that graphene can accept electrons and therefore inhibit recombination and increase the absorption properties and stability of composite catalysts [89,90,91]. Therefore, nanocomposites of various semiconductors with graphene (GN) have been investigated as advanced photocatalysts [92,93,94]. It has been reported that the fabrication of ternary composites based on graphene and reduced graphene oxide (rGO) with transition metal oxides and conducting polymers can be considered to be a promising approach for overcoming the problems related to recombination losses and developing novel photocatalysts with perfect photocatalytic activity [36,95,96,97]. According to this methodology, there is a diversity in ternary composites in combination with graphene or graphene oxide, PANI, and metal oxides that have been recently introduced in the degradation and absorption of organic pollutants. For example, Kumar et al. successfully developed a conducting ternary PANI/TiO2/graphene nanocomposite through an in-situ oxidative polymerization method, in which aniline molecules were polymerized in the presence of TiO2 and graphene nanoparticles (Figure 4a) [39]. UV–Vis absorption and PL spectra showed that PANI/TiO2/graphene exhibited a higher visible light absorption and a lower recombination rate than PANI/TiO2 (Figure 4b,c). Using a combination of rGO, PANI, and spinel ferrite, Feng et al. synthesized a rGO/ZnFe2O4/PANI ternary photocatalyst via a simple and low-cost method (Figure 4d) [36]. In the first step, spinel ZnFe2O4 nanoparticles were deposited onto the surface of rGO to form the binary rGO/ZnFe2O4 composite. This binary composite was then coated with PANI to obtain the ternary composite in the next step. It was shown that this ternary composite structure exhibited three main advantages: (i) all spinel ZnFe2O4 nanoparticles in the binary rGO/ZnFe2O4 composite were completely coated with PANI, and thus the photoinduced e-h+ pairs produced by ZnFe2O4 could be stabilized by PANI; (ii) the photoinduced e-h+ pairs produced by PANI were also stabilized by rGO; (iii) the synergistic effect between the three components significantly enhanced the photocatalytic activity of the ternary composite photocatalyst. In another study, Miao et al. reported a ternary hybrid through a combination of rGO and a binary PANI/Cu2O composite via a one-pot method [97]. In this composite, PANI and Cu2O nanoparticles were embedded in the rGO nanosheets, which highly increased the photoactivities of ternary nanocomposites compared with the binary ones of PANI/Cu2O, rGO/Cu2O, or PANI/rGO. In summary, it is expected that the preparation of ternary photocatalysts based on graphene or rGO, and PANI with metal oxides will be a new and promising pathway for the development of advanced photocatalysts.
Recently, the combination of TiO2, graphitic carbon nitride (g-C3N4), and PANI has also been investigated in an attempt to discover novel ternary composite photocatalysts by taking advantage of each component [98]. These composites are expected to have a higher interfacial charge transfer, which can greatly enhance the photodecomposition of organic pollutants. For example, Alenizi et al. reported a ternary g-C3N4/TiO2/PANI nanocomposite photocatalyst for the degradation of organic dyes in wastewater [41]. Regarding the synthesis process, a defect-rich TiO2 lattice and lamellar structures were first generated from TiO2 powders mixed with 10 M NaOH. It was indicated that mixed phase titania and sodium titanate lamellar structures resulted in a better surface area for binding with g-C3N4 and PANI, which promoted the higher interfacial charge separation and improved the adsorption-photocatalytic properties. Moreover, the g-C3N4/TiO2/PANI nanocomposite showed a higher photocatalytic activity under direct sunlight irradiation. In another study, such a novel ternary heterostructure also showed a considerable increase in charge separation efficiency, specific surface area and visible light harvesting, which can be attributed to the synergetic effects of PANI and ZnO and the exfoliated two dimensional CN nanosheets in the roles of catalysts and supporting materials, respectively [98].

3. Visible-Light-Responsive Photocatalysis Mechanisms of Conducting Polymer/Metal Oxide Composites

In terms of the general mechanism, a semiconductor photocatalyst enables the absorption of visible light from the solar spectrum, which causes the excitation of electrons from the valence band (VB) to the CB and generates electron–hole pairs. These electrons and holes are then transferred to the composite photocatalyst surface for the degradation and oxidization of CO2 or pollutants.
In the binary composites of conducting polymers and metal oxides, conducting polymers work as visible photosensitizers to generate photoelectrons from the VB to the CB, which can be transferred to the CB of the metal oxides [99]. This is especially the case in relation to the PANI/TiO2 composite, and the processes of photoexcitation, charge separation and reaction in the composite under visible-light irradiation is presented in Figure 5a [100,101]. It has been demonstrated that the energy levels of TiO2 and PANI can be matched to each other [23]. The CB of TiO2 is slightly lower than the LUMO of PANI, and thus TiO2 can work as a sink for the photogenerated electrons in the composite photocatalyst. Furthermore, the HOMO of PANI is higher than the VB of TiO2, and thus PANI can act as an acceptor for the photogenerated holes in the composite photocatalyst. Consequently, the adsorption and electrical conductivity of the binary PANI–TiO2 composite under visible-light irradiation is significantly boosted, and larger numbers of electron–hole pairs are generated. Specially, under visible-light irradiation, PANI will absorb photons to induce electrons into the LUMO, while TiO2 will absorb the UV–Vis light to excite the electrons into the CB. Due to the different potentials of PANI and TiO2, as mentioned previously, the excited electrons in the LUMO of PANI can be transferred to the CB of TiO2 and the generated h+ can move from the VB of TiO2 to the HOMO of PANI. The photoelectrons on the surface of the composite will reduce H+ to form H2 or react with the surface-adsorbed O2 to generate ●OH radicals, which play a main role in the degradation of pollutants. Meanwhile, photo-holes enable pollutants to oxidize mineralized products.
For ternary composites of conducting polymers and metal oxides, the mechanism can be explained based on two different types of composites: (i) conducting polymer/metal oxide/metal oxide; (ii) conducting polymer/metal oxide/another compound. Regarding the conducting polymer/metal oxide/metal oxide composites, it has been indicated that the formation of Z-scheme composites is likely to effectively enhance the carrier mobility and generation of electrons and holes for a very long time [83]. In these Z-scheme composite photocatalysts, photogenerated electrons from a metal oxide recombine with the photogenerated holes in another coupled metal oxide [99,103]. For example, the ternary composite Cu2O/ZnO- PANI showed Z-scheme heterojunction properties, which resulted in super-fast photocatalytic activities and high stability [38]. The photocatalytic mechanism of the Z-scheme Cu2O/ZnO- PANI composite is presented in Figure 5b. It is obvious that the photogenerated electrons in the CB of ZnO are rapidly transferred to the VB of Cu2O and the HOMO of PANI, where they recombine with the photogenerated holes. At the same time, the photogenerated electrons in the LUMO of PANI and the CB of Cu2O are separated and migrate to the surface to react with surface-adsorbed O2 in order to generate ●O2 radicals, while the photogenerated holes in the VB of ZnO are also transferred to the surface for the photocatalytic evolution of ●OH radicals. These formed radials are mainly responsible for the photodegradation and photoreduction of pollutants.
Regarding ternary composites of conducting polymer/metal oxide/another compound, the cooperation between the three components is expected to improve the photoinduced charge separation and suppress charge recombination, resulting in an enhanced photocatalytic performance. For instance, the highly enhanced photocatalytic activity of a ternary PANI/TiO2/rGO composite is significantly contributed to by the interfacial charge transfer in PANI and rGO due to p-conjugated groups [102]. This detailed photocatalytic mechanism is shown in Figure 5c. In this composite, the ●O2 radicals for the oxidization and degradation of the pollutants are generated by similar pathways as in a binary PANI/TiO2 composite. Moreover, the hydrogen bonds between TiO2 and rGO also promote the migration of photogenerated electrons and holes to the CB of TiO2. Moreover, photoinduced electrons can easily be transferred in the composite due to the π-π stacking of rGO, which leads to a long-term maintenance of electron and hole separation, thereby resulting in the significantly enhanced photocatalytic activity of the composite.

4. Photocatalytic Applications of CP/Metal Oxide Composites in the Environment Field

Generally, composite photocatalysts are typical semiconductors with an adequate band-gap energy that can strongly absorb light from the solar spectrum, leading to the excitation of electrons from the VB to the CB and, then, the formation of electron–hole pairs. Finally, the electrons and holes move to the photocatalyst surface and react with the absorbed pollutants. As advanced photocatalysts, CP/metal oxide composites can be used in various applications in the environmental field. Four main applications will be discussed in the following section, including the decomposition of organic pollutants, solar water splitting, CO2 reduction, and N2 reduction.

4.1. Decomposition of Organic Pollutants

As mentioned previously, the photocatalytic performances of single components of CPs (PPy, PANI, and PEDOT) and metal oxides (e. g., TiO2, ZnO, Fe3O4, and ZnFe2O4) are poor. However, their combination can greatly enhance performance due to inhibiting the recombination of charge carriers. Actually, in case of the binary composites of CP and metal oxides, CPs play an important role as photosensitizers in the absorption of visible light, while transition metal ions (Co2+, Fe2+, Mn2+, Ni+, etc.) can activate peroxymonosulfate to generate sulfate radicals (SO4) and hydroxyl radicals (OH) for the degradation of organic pollutants [104,105]. Most binary composites of CPs and metal oxides have a high photocatalytic activity in the degradation of organic dyes such as bisphenol A, direct blue 15, RhB, methyl orange (MO), methylene blue (MB), and congo red, as presented in Table 1. Taking PANI/SnO2 as an example, these binary composites have been commonly used to enhance the photocatalytic degradation of various organic pollutants in wastewater [106,107]. It has been shown that the inclusion of PANI in SnO2 results in a decreased crystallite size and an increased surface area. Recently, the photocatalytic efficiencies of SnO2/PANI and PANI/Sn3O4 nanocomposites were evaluated with direct blue 15 [27] and rhodamine B [28]. The results indicate that the PANI/Sn3O4 composite can reach a maximum degradation efficiency of around 97% for rhodamine B under visible-light irradiation, which is 2.27 times higher than that of Sn3O4 alone. In addition, the photocatalytic performance of PANI/Sn3O4 exhibited a relative stability during RhB photodegradation after three runs and the photodegradation efficiency could still be maintained at >90%.
Recently, PANI and magnetic iron oxide composite photocatalysts have been considered as good candidates for the adsorption and photodegradation of organic dyes in the treatment of wastewater due to their high charge transport properties and superparamagnetism at room temperature [108,109]. For instance, Alves et al. reported a PANI/magnetic iron oxide composite for MB removal [26]. Owing to the dispersion of magnetic iron oxide particles in the polymer phase, the composite showed a high electrical conductivity (10−2 S·cm−1) and good adsorption properties. Interestingly, the PANI/magnetic iron oxide composite showed both adsorption and photodegradation properties in the pollutant removal process. The composite exhibited a reduction efficiency of 99% of the initial dye concentration because of the synergism effect between the iron oxide-polymer phases in the photocatalytic action. Importantly, due to the super-paramagnetic behavior, the PANI/magnetic iron oxide composite can be easily collected by applying a magnetic field and, therefore, can be reused in numerous cycles. Using a similar approach, Kharazi et al. introduced the novel binary composite of copper spinel ferrite (CuFe2O4) and PANI as an adsorbent and photocatalyst for the treatment of MO in wastewater [108]. With the presence of pre-synthesized CuFe2O4 nanoparticles, the nanocomposite has a mesoporous structure with a BET surface area of 20.3668 m2/g, which showed excellent adsorption of the dye (capacity of 345.9 mg/g) and excellent magnetic properties.
Nowadays, repeated recovery and reusability of photocatalysts are of great significance for practical applications in the environmental field. Therefore, it is important to synthesize and develop multicomponent nanocomposites based on metal oxides and conducting polymers for better catalytic performances in adsorption and photodegradation of organic dyes [34,37,110,111]. Moreover, the incorporation of magnetic nanoparticles in the composite photocatalysts, especially spinel ferrites and iron oxides, enables collection and reuse via magnetic separation [84]. Taking this into consideration, Xiong et al. prepared a magnetically recyclable photocatalyst based on a ternary composite of TiO2-CoFe2O4-PANI for the degradation of various dyes [34]. The results show that the ternary composite photocatalyst exhibits a high photocatalytic activity in the degradation of anionic dyes, such as MO, trypan blue (TB), and Brilliant Blue R (BBR), while only poor activity for cationic dyes, such as MB, Malachite Green (MG), and Neutral Red (NR) (Figure 6a). The higher efficiency in the degradation of anionic dyes can be attributed to an electrostatic attraction between the negatively charged groups of anionic dyes and the positively charged backbone of PANI, which greatly promotes the degradation. Moreover, the ternary composite showed a high photodegradation property rather than adsorption, where the photobleaching of MO mainly come from the photodegradation process (Figure 6b). Due to the good magnetic properties of CoFe2O4, the ternary TiO2-CoFe2O4-PANI photocatalyst also enables collection by a simple magnet or an applied magnetic field (see inset of Figure 6c). The stability of photocatalysts is also one of the most important factors for their use in practical applications [24,112,113]. Ternary composites of metal oxides and CPs demonstrated a high stability. For example, TiO2-CoFe2O4-PANI photocatalysts still retained a high rate of photodegradation after three cycles (Figure 6c), as well as structural stability (Figure 6d).
In summary, the decomposition of organic pollutants, especially dyes, is an important consideration in the application of both binary and ternary composites of metal oxides and CPs. Current strategies are focused on the development of multicomponent composites possessing a high photocatalytic activity along with a perfect recyclability and reusability.

4.2. CO2 Reduction

Over the past century, CO2 (recognized as a major greenhouse gas) levels have been rapidly increasing because of human activities, which has caused a global warming problem [114]. Among the various strategies developed for reducing CO2 emissions, photocatalytic CO2 reduction using semiconductors is one of the most viable approaches [115,116,117]. CO2 photoreduction is defined as the process of using light irradiation-induced energy to convert CO2 to reduced C1 and C2 hydrocarbon compounds [118]. However, CO2 is a very stable molecule and cannot absorb in the sunlight spectrum, and thus the photoreduction process needs support from suitable photosensitizers [99]. These photosensitizers are semiconductors that can generate electron–hole pairs and their subsequent transfer to CO2 and a reductant, respectively. A wide range of metal oxides, such as TiO2, Ga2O3, W18O49, SrTiO3, ZnGa2O4, Zn2GeO4, and Bi2WO6, have been used as semiconductors for the photocatalytic reduction of CO2 with H2O [119,120,121,122]. Nonetheless, it has been indicated that their poor photocatalytic performance is caused by the rapid recombination of photogenerated electrons and holes, low CO2 adsorption ability, and low CO2 reactivity.
It has to be noted that the application of using conducting polymers as a single photocatalyst for CO2 photoreduction is relatively limited mainly due to their low photostability. Therefore, conducting polymers, such as PANI, PPy, and PEDOT, have recently been combined with metal oxides in binary or ternary composites for CO2 photo-reducing applications. Liu et al. reported a PANI/TiO2 composite photocatalyst that demonstrated a considerable enhancement in the photoreduction of CO2 with H2O [118]. This enhancement was ascribed to an increase in CO2 chemisorption and the facilitated separation of photogenerated electron–hole pairs. The specific mechanism for CO2 photoreduction is presented in Figure 7. It has been indicated that the LUMO level or CB edge of TiO2 has the same level of 0.18 V under N2 and CO2 conditions. However, the LUMO level of PANI is different under N2 and CO2 conditions, at −0.42 V and −0.15 V, respectively. Consequently, electrons were transferred from the CB of TiO2 to PANI under a CO2 atmosphere due to the change in the LUMO level of PANI. It was proposed that this phenomenon makes a large contribution to the improvement in the separation of photogenerated electron–hole pairs in the binary PANI/TiO2 composite, which leads to a higher performance in the photoreduction of CO2 and H2O to CH4 and H2. In particular, the binary composite showed higher rates of CO, CH4, and H2 formation (2.8, 3.8, and 2.7 times, respectively) from CO2 photoreduction, as compared with TiO2. It was also suggested that the synergistic effect between TiO2 and PANI largely reduced CO2 in the presence of H2O. This work may be regarded as the first related to the use of binary conducting polymer/metal oxide composite for CO2 photoreduction. In another study, Gao et al. used another conducting polymer (PPy) in combination with TiO2 in a binary PPy/TiO2 nanocomposite for the reduction of CO2 [32]. The composite was synthesized using simple oxidative polymerization of pyrrole using ferric chloride (FeCl3) as the oxidant in the presence of TiO2 nanoparticles. The nanocomposite showed better photoreduction efficiency for CO2 than pure TiO2 under simulated solar light irradiation. In addition, the PPy/TiO2 photocatalyst has high potential for practical applications because of its high stability. In summary, these studies indicate that the composites of conducting polymers and metal oxides are potential candidates for the development of advanced materials for the reduction of CO2.

4.3. Photocatalytic Oxidation of Heavy Metals

Arsenic (As), is one of the heavy metals present in water environments, and it can pose a serious threat to humans and other species because of its high toxicity [123,124]. Generally, As(III) (arsenite) and As(V) (arsenate) are the two most common types of As present in groundwater, of which As(III) shows a higher toxicity. Current removal technologies exhibit poor performance with regard to As(III), due to its very high mobility [125]. It has been proposed that one of the most efficient approaches for removal is the conversion of As(III) to As(V), which then enables the removal of As(V) by an adsorption process [126]. Recently, photocatalyst-based oxidation methods have attracted great attention for the removal of heavy metals [127,128]. Therefore, there has been increasing attention devoted to the development of novel materials for an efficient oxidation of As(III) to As(V) and simultaneous removal of As(V). Due to an effective adsorption of As(V) onto the surface, good magnetic properties, and easy separation, γ-Fe2O3 has recently been used in fabricating photocatalysts for As(III) removal, in combination with TiO2 [129,130]. Taking advantage of the synergistic effects of TiO2 and γ-Fe2O3 along with the support of PANI, Wang et al. successfully developed a novel ternary composite (γ-Fe2O3/PANI/TiO2) and significantly enhanced the photocatalytic adsorption of As(III) [131]. Regarding the preparation process, γ-Fe2O3 was first synthesized by annealing Fe3O4 at 300 °C for 2 h under air atmosphere (Figure 8a). A magnetic γ-Fe2O3 core–shell ternary nanocomposite was then obtained by hydrothermal crystallization of TiO2 on the surface of a magnetic core–shell loaded with PANI. The results indicate that the γ-Fe2O3/PANI/TiO2 composite effectively removed aqueous As(III) via a coupled photocatalytic oxidation/adsorption process. Specifically, the As(V) concentration in the solution increased and accounted for 54% of the total arsenic after 300 min of treatment with the γ-Fe2O3/PANI/TiO2 composite under visible-light irradiation. The adsorption process of As(III) was almost balanced, and its photocatalytic oxidation efficiency reached 75%. Notably, the photocatalytic oxidation of As (III) is affected by the synergic effects of some active substances, especially superoxide free radicals and photogenerated holes. The nanocomposite also showed excellent stability and good reusability.
Chromium (Cr) is another heavy metal that has a high toxicity to human beings. Cr can exist in various valance states but occurs mainly as Cr(VI) and Cr(III) [133]. Cr(VI) has a much higher toxicity for humans than Cr(III) [134]. Accordingly, it is necessary to design and develop advanced materials which can convert Cr(VI) into Cr(III). Recently, the photocatalytic reduction method has been considered as a promising approach for reducing toxic Cr(VI) to nontoxic Cr(III), due to its efficiency, simplicity, low cost, and convenience [45,135,136]. It has been demonstrated that photoreduction efficiency largely depends on the adsorption and diffusion of Cr(VI) ions on the surface of the photocatalyst. Due to its high adsorption ability, it has been suggested that PANI is a material that shows good potential for facilitating the photocatalytic reduction process of Cr(VI) through incorporation with other photocatalysts [137,138]. Taking this into consideration, Deng et al. developed a binary PANI/TiO2 composite for improving the photocatalytic reduction performance and stability of Cr(VI) ions [100]. This enhancement was due to two reasons: (i) PANI possesses many positively charged amino groups that enable the effective adsorption of Cr(VI) but make Cr(III) leave the reaction interface quickly, resulting in a high promotion of the photocatalytic reduction process and improvement in the photocatalyst stability; and (ii) modification of the TiO2 surface by PANI promotes the separation of photogenic charges on the TiO2 surface, leading to a great increase in photocatalytic activity. Specifically, the photocatalytic reduction performance of a binary PANI/TiO2 composite for different concentrations of Cr(VI) is presented in Figure 8b. Photocatalytic reduction is inversely proportional to the increase in concentration of Cr(VI), and Cr(VI) was completely removed to concentrations below 20 ppm after 30 min of light irradiation. It is difficult to separate the reactant product Cr(III) from the catalyst surface for a high concentration of Cr(VI), resulting in a covered photoreduction active site and reduced photocatalytic activity.
Moreover, the binary PANI/TiO2 composite showed a high stability while retaining 100% of the reduction performance after ten cycles. In order to further enhance the photocatalytic reduction and adsorption of binary PANI/TiO2 composites for high concentrations of Cr(VI), Vellaichamy et al. synthesized a ternary PANI/MnO2/TiO2 nanocomposite via a one-pot oxidative polymerization method at room temperature [132]. In this composite, TiO2 plays an important role as efficient linker between PANI and MnO2 [139,140]. Due to the synergistic effects between the three components, the ternary PANI/MnO2/TiO2 nanocomposite showed a superior photocatalytic activity in the reduction of toxic Cr(VI) to benign Cr(III) in comparison with single photocatalysts (PANI, MnO2, and TiO2) and binary composites (PANI/MnO2, PANI/TiO2, and MnO2/TiO2) (Figure 8c). Especially, PANI/MnO2/TiO2 revealed an excellent photocatalytic performance in the reduction of Cr(VI), with a transformation efficiency of 99.9% within 5 min and a rate constant of 15.97 × 10−2 min−1. The reduction rate was found to depend on the initial Cr(VI) concentration, oxidant (HCOOH), pH, and temperature. In addition, the PANI/MnO2/TiO2 nanocomposite also displayed good stability and retained a high photocatalytic efficiency, even after use in five cycles.
In summary, the results of these studies demonstrate that both binary and ternary composites of conducting polymers and metal oxides can be used as effective and economically viable photocatalysts for the reduction of toxic heavy metals in water.

5. Conclusions

The development of highly active composite photocatalysts for use in environmental applications is considered to be a sustainable way of eliminating organic pollutants and heavy metals. Among them, composite photocatalysts based on inorganic semiconductors, especially metal oxides, and CPs emerge as novel promising photoactive materials. It was demonstrated that these composite photocatalysts have several outstanding characteristics, such as light absorption in the visible range of the spectrum, high photocatalytic activity and stability, good reusability, low cost, convenience, and scalability of production. Over the past few decades, numerous composite photocatalysts based on conducting polymers and metal oxides were prepared and developed. These can be classified into two types based on the number of components in the composites, including (i) binary composite photocatalysts involving the use of one conducting polymer (i.e., PANI, PEDOT, and PPy) and one metal oxide (i.e., TiO2, Fe3O4, SnO2, and ZnO); and (ii) ternary composite photocatalysts, which often include one conducting polymer and two different metal oxides or one conducting polymer, one metal oxide, and another semiconductor. Both binary and ternary conducting polymer/metal oxide composites show promising photocatalytic activity for the degradation, reduction, and adsorption of organic pollutants, CO2 gas, and heavy metals. However, ternary composites were confirmed as superior to binary ones due to the synergistic effects of three components in promoting photocatalytic activity and improving photocatalyst stability. Therefore, conducting polymer/metal oxide composites are currently one of the most promising candidates for photocatalysts in environmental applications.

Author Contributions

Conceptualization, V.V.T. and M.C.; writing—original draft preparation, V.V.T., T.T.V.N. and H.-R.J.; writing—review and editing, V.V.T., T.T.V.N., H.-R.J. and M.C.; visualization, H.-R.J. and M.C.; supervision, H.-R.J. and M.C.; project administration H.-R.J. and M.C.; funding acquisition, H.-R.J. and M.C. All authors have read and agreed to the published version of the manuscript.

Funding

This work was financially supported by the Basic Science Research Program through the National Research Foundation of Korea (NRF) funded by the Ministry of Education (NRF-2020R1I1A306849711) and the Technology development program (S2829951) funded by the Ministry of SMEs and Startups (MSS, Korea).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Fresno, F.; Portela, R.; Suárez, S.; Coronado, J.M. Photocatalytic materials: Recent achievements and near future trends. J. Mater. Chem. A 2014, 2, 2863–2884. [Google Scholar] [CrossRef]
  2. Park, H.; Park, Y.; Kim, W.; Choi, W. Surface modification of TiO2 photocatalyst for environmental applications. J. Photochem. Photobiol. 2013, 15, 1–20. [Google Scholar] [CrossRef]
  3. Xie, S.; Zhang, Q.; Liu, G.; Wang, Y. Photocatalytic and photoelectrocatalytic reduction of CO2 using heterogeneous catalysts with controlled nanostructures. Chem. Commun. 2016, 52, 35–59. [Google Scholar] [CrossRef]
  4. Schneider, J.; Matsuoka, M.; Takeuchi, M.; Zhang, J.; Horiuchi, Y.; Anpo, M.; Bahnemann, D.W. Understanding TiO2 Photocatalysis: Mechanisms and Materials. Chem. Rev. 2014, 114, 9919–9986. [Google Scholar] [CrossRef] [PubMed]
  5. Zimbone, M.; Cacciato, G.; Spitaleri, L.; Egdell, R.G.; Grimaldi, M.G.; Gulino, A. Sb-Doped Titanium Oxide: A Rationale for Its Photocatalytic Activity for Environmental Remediation. ACS Omega 2018, 3, 11270–11277. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  6. Fiorenza, R.; Di Mauro, A.; Cantarella, M.; Gulino, A.; Spitaleri, L.; Privitera, V.; Impellizzeri, G. Molecularly imprinted N-doped TiO2 photocatalysts for the selective degradation of o-phenylphenol fungicide from water. Mater. Sci. Semicond. Process. 2020, 112, 105019. [Google Scholar] [CrossRef]
  7. Ma, Y.; Wang, X.; Jia, Y.; Chen, X.; Han, H.; Li, C. Titanium Dioxide-Based Nanomaterials for Photocatalytic Fuel Generations. Chem. Rev. 2014, 114, 9987–10043. [Google Scholar] [CrossRef] [PubMed]
  8. Chen, S.; Takata, T.; Domen, K. Particulate photocatalysts for overall water splitting. Nat. Rev. Mater. 2017, 2, 17050. [Google Scholar] [CrossRef]
  9. Ran, J.; Jaroniec, M.; Qiao, S.-Z. Cocatalysts in Semiconductor-based Photocatalytic CO2 Reduction: Achievements, Challenges, and Opportunities. Adv. Mater. 2018, 30, 1704649. [Google Scholar] [CrossRef]
  10. Zhang, Y.; He, S.; Guo, W.; Hu, Y.; Huang, J.; Mulcahy, J.R.; Wei, W.D. Surface-Plasmon-Driven Hot Electron Photochemistry. Chem. Rev. 2018, 118, 2927–2954. [Google Scholar] [CrossRef] [PubMed]
  11. Maeda, K. Z-Scheme Water Splitting Using Two Different Semiconductor Photocatalysts. ACS Catal. 2013, 3, 1486–1503. [Google Scholar] [CrossRef]
  12. Dahl, M.; Liu, Y.; Yin, Y. Composite Titanium Dioxide Nanomaterials. Chem. Rev. 2014, 114, 9853–9889. [Google Scholar] [CrossRef] [PubMed]
  13. Tran, V.V.; Tran, N.H.T.; Hwang, H.S.; Chang, M. Development strategies of conducting polymer-based electrochemical biosensors for virus biomarkers: Potential for rapid COVID-19 detection. Biosens. Bioelectron. 2021, 182, 113192. [Google Scholar] [CrossRef] [PubMed]
  14. Zhou, Q.; Shi, G. Conducting Polymer-Based Catalysts. J. Am. Chem. Soc. 2016, 138, 2868–2876. [Google Scholar] [CrossRef]
  15. Ognibene, G.; Gangemi, C.M.A.; Spitaleri, L.; Gulino, A.; Purrello, R.; Cicala, G.; Fragalà, M.E. Role of the surface composition of the polyethersulfone–TiiP–H2T4 fibers on lead removal: From electrostatic to coordinative binding. J. Mater. Sci. 2019, 54, 8023–8033. [Google Scholar] [CrossRef]
  16. Zang, L.; Qiu, J.; Yang, C.; Sakai, E. Preparation and application of conducting polymer/Ag/clay composite nanoparticles formed by in situ UV-induced dispersion polymerization. Sci. Rep. 2016, 6, 20470. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  17. Gangopadhyay, R.; De, A. Conducting Polymer Nanocomposites: A Brief Overview. Chem. Mater. 2000, 12, 608–622. [Google Scholar] [CrossRef]
  18. Jana, B.; Bhattacharyya, S.; Patra, A. Conjugated polymer P3HT–Au hybrid nanostructures for enhancing photocatalytic activity. Phys. Chem. Chem. Phys. 2015, 17, 15392–15399. [Google Scholar] [CrossRef]
  19. Xu, S.; Gu, L.; Wu, K.; Yang, H.; Song, Y.; Jiang, L.; Dan, Y. The influence of the oxidation degree of poly(3-hexylthiophene) on the photocatalytic activity of poly(3-hexylthiophene)/TiO2 composites. Sol. Energy Mater. Sol. Cells 2012, 96, 286–291. [Google Scholar] [CrossRef]
  20. Hidalgo, D.; Bocchini, S.; Fontana, M.; Saracco, G.; Hernández, S. Green and low-cost synthesis of PANI–TiO2 nanocomposite mesoporous films for photoelectrochemical water splitting. RSC Adv. 2015, 5, 49429–49438. [Google Scholar] [CrossRef] [Green Version]
  21. Sambaza, S.S.; Maity, A.; Pillay, K. Polyaniline-Coated TiO2 Nanorods for Photocatalytic Degradation of Bisphenol A in Water. ACS Omega 2020, 5, 29642–29656. [Google Scholar] [CrossRef]
  22. Chen, X.; Li, H.; Wu, H.; Wu, Y.; Shang, Y.; Pan, J.; Xiong, X. Fabrication of TiO2@PANI nanobelts with the enhanced absorption and photocatalytic performance under visible light. Mater. Lett. 2016, 172, 52–55. [Google Scholar] [CrossRef]
  23. Li, C.; Zhou, T.; Zhu, T.; Li, X. Enhanced visible light photocatalytic activity of polyaniline–crystalline TiO2–halloysite composite nanotubes by tuning the acid dopant in the preparation. RSC Adv. 2015, 5, 98482–98491. [Google Scholar] [CrossRef]
  24. Zhang, H.; Zong, R.; Zhu, Y. Photocorrosion Inhibition and Photoactivity Enhancement for Zinc Oxide via Hybridization with Monolayer Polyaniline. J. Phys. Chem. C 2009, 113, 4605–4611. [Google Scholar] [CrossRef]
  25. Gilja, V.; Vrban, I.; Mandić, V.; Žic, M.; Hrnjak-Murgić, Z. Preparation of a PANI/ZnO Composite for Efficient Photocatalytic Degradation of Acid Blue. Polymers 2018, 10, 940. [Google Scholar] [CrossRef] [Green Version]
  26. Alves, F.H.d.O.; Araújo, O.A.; de Oliveira, A.C.; Garg, V.K. Preparation and characterization of PAni(CA)/Magnetic iron oxide hybrids and evaluation in adsorption/photodegradation of blue methylene dye. Surf. Interfaces 2021, 23, 100954. [Google Scholar] [CrossRef]
  27. Rajaji, U.; Eva Gnana Dhana Rani, S.; Chen, S.-M.; Rajakumar, K.; Govindasamy, M.; Alzahrani, F.M.; Alsaiari, N.S.; Ouladsmane, M.; Sharmila Lydia, I. Synergistic photocatalytic activity of SnO2/PANI nanocomposite for the removal of direct blue 15 under UV light irradiation. Ceram. Int. 2021. [Google Scholar] [CrossRef]
  28. Lv, M.; Yang, L.; Wang, X.; Cheng, X.; Song, Y.; Yin, Y.; Liu, H.; Han, Y.; Cao, K.; Ma, W.; et al. Visible-light photocatalytic capability and the mechanism investigation of a novel PANI/Sn3O4 p–n heterostructure. RSC Adv. 2019, 9, 40694–40707. [Google Scholar] [CrossRef] [Green Version]
  29. Kim, K.N.; Jung, H.-R.; Lee, W.-J. Hollow cobalt ferrite–polyaniline nanofibers as magnetically separable visible-light photocatalyst for photodegradation of methyl orange. J. Photochem. Photobiol. A 2016, 321, 257–265. [Google Scholar] [CrossRef]
  30. Katančić, Z.; Chen, W.-T.; Waterhouse, G.I.N.; Kušić, H.; Lončarić Božić, A.; Hrnjak-Murgić, Z.; Travas-Sejdic, J. Solar-active photocatalysts based on TiO2 and conductive polymer PEDOT for the removal of bisphenol A. J. Photochem. Photobiol. A 2020, 396, 112546. [Google Scholar] [CrossRef]
  31. Dimitrijevic, N.M.; Tepavcevic, S.; Liu, Y.; Rajh, T.; Silver, S.C.; Tiede, D.M. Nanostructured TiO2/Polypyrrole for Visible Light Photocatalysis. J. Phys. Chem. C 2013, 117, 15540–15544. [Google Scholar] [CrossRef]
  32. Gao, F.; Hou, X.; Wang, A.; Chu, G.; Wu, W.; Chen, J.; Zou, H. Preparation of polypyrrole/TiO2 nanocomposites with enhanced photocatalytic performance. Particuology 2016, 26, 73–78. [Google Scholar] [CrossRef]
  33. Yan, B.; Wang, Y.; Jiang, X.; Liu, K.; Guo, L. Flexible Photocatalytic Composite Film of ZnO-Microrods/Polypyrrole. ACS Appl. Mater. Interfaces 2017, 9, 29113–29119. [Google Scholar] [CrossRef]
  34. Xiong, P.; Wang, L.; Sun, X.; Xu, B.; Wang, X. Ternary Titania–Cobalt Ferrite–Polyaniline Nanocomposite: A Magnetically Recyclable Hybrid for Adsorption and Photodegradation of Dyes under Visible Light. Ind. Eng. Chem. Res. 2013, 52, 10105–10113. [Google Scholar] [CrossRef]
  35. Li, J.; Xiao, Q.; Li, L.; Shen, J.; Hu, D. Novel ternary composites: Preparation, performance and application of ZnFe2O4/TiO2/polyaniline. Appl. Surf. Sci. 2015, 331, 108–114. [Google Scholar] [CrossRef]
  36. Feng, J.; Hou, Y.; Wang, X.; Quan, W.; Zhang, J.; Wang, Y.; Li, L. In-depth study on adsorption and photocatalytic performance of novel reduced graphene oxide-ZnFe2O4-polyaniline composites. J. Alloys Compd. 2016, 681, 157–166. [Google Scholar] [CrossRef]
  37. Wu, H.; Lin, S.; Chen, C.; Liang, W.; Liu, X.; Yang, H. A new ZnO/rGO/polyaniline ternary nanocomposite as photocatalyst with improved photocatalytic activity. Mater. Res. Bull. 2016, 83, 434–441. [Google Scholar] [CrossRef]
  38. Mohammed, A.M.; Mohtar, S.S.; Aziz, F.; Aziz, M.; Ul-Hamid, A. Cu2O/ZnO-PANI ternary nanocomposite as an efficient photocatalyst for the photodegradation of Congo Red dye. J. Environ. Chem. Eng. 2021, 9, 105065. [Google Scholar] [CrossRef]
  39. Kumar, R.; Ansari, M.O.; Parveen, N.; Oves, M.; Barakat, M.A.; Alshahri, A.; Khan, M.Y.; Cho, M.H. Facile route to a conducting ternary polyaniline@TiO2/GN nanocomposite for environmentally benign applications: Photocatalytic degradation of pollutants and biological activity. RSC Adv. 2016, 6, 111308–111317. [Google Scholar] [CrossRef]
  40. Park, Y.; Numan, A.; Ponomarev, N.; Iqbal, J.; Khalid, M. Enhanced photocatalytic performance of PANI-rGO-MnO2 ternary composite for degradation of organic contaminants under visible light. J. Environ. Chem. Eng. 2021, 9, 106006. [Google Scholar] [CrossRef]
  41. Alenizi, M.A.; Kumar, R.; Aslam, M.; Alseroury, F.A.; Barakat, M.A. Construction of a ternary g-C3N4/TiO2@polyaniline nanocomposite for the enhanced photocatalytic activity under solar light. Sci. Rep. 2019, 9, 12091. [Google Scholar] [CrossRef]
  42. Riaz, U.; Ashraf, S.M.; Kashyap, J. Enhancement of photocatalytic properties of transitional metal oxides using conducting polymers: A mini review. Mater. Res. Bull. 2015, 71, 75–90. [Google Scholar] [CrossRef]
  43. Wang, Q.; Hui, J.; Li, J.; Cai, Y.; Yin, S.; Wang, F.; Su, B. Photodegradation of methyl orange with PANI-modified BiOCl photocatalyst under visible light irradiation. Appl. Surf. Sci. 2013, 283, 577–583. [Google Scholar] [CrossRef]
  44. Lee, K.; Cho, S.; Heum Park, S.; Heeger, A.J.; Lee, C.-W.; Lee, S.-H. Metallic transport in polyaniline. Nature 2006, 441, 65–68. [Google Scholar] [CrossRef]
  45. Nguyen, T.N.; Tran, V.V.; Bui, V.K.H.; Kim, M.; Park, D.; Hur, J.; Kim, I.T.; Lee, H.U.; Ko, S.; Lee, Y.-C. A Novel Photocatalyst Composite of Magnesium Aminoclay and TiO2 Immobilized into Activated Carbon Fiber (ACF) Matrix for Pollutant Removal. J. Nanosci. Nanotech. 2020, 20, 6844–6849. [Google Scholar] [CrossRef]
  46. Bui, V.K.H.; Tran, V.V.; Moon, J.-Y.; Park, D.; Lee, Y.-C. Titanium Dioxide Microscale and Macroscale Structures: A Mini-Review. Nanomaterials 2020, 10, 1190. [Google Scholar] [CrossRef]
  47. Yadav, A.; Kumar, H.; Sharma, R.; Kumari, R. Influence of polyaniline on the photocatalytic properties of metal nanocomposites: A review. Colloids Interface Sci. Commun. 2021, 40, 100339. [Google Scholar] [CrossRef]
  48. Gilja, V.; Novaković, K.; Travas-Sejdic, J.; Hrnjak-Murgić, Z.; Kraljić Roković, M.; Žic, M. Stability and Synergistic Effect of Polyaniline/TiO2 Photocatalysts in Degradation of Azo Dye in Wastewater. Nanomaterials 2017, 7, 412. [Google Scholar] [CrossRef] [Green Version]
  49. Mirzaeifard, Z.; Shariatinia, Z.; Jourshabani, M.; Rezaei Darvishi, S.M. ZnO Photocatalyst Revisited: Effective Photocatalytic Degradation of Emerging Contaminants Using S-Doped ZnO Nanoparticles under Visible Light Radiation. Ind. Eng. Chem. Res. 2020, 59, 15894–15911. [Google Scholar] [CrossRef]
  50. Guo, M.Y.; Ng, A.M.C.; Liu, F.; Djurišić, A.B.; Chan, W.K.; Su, H.; Wong, K.S. Effect of Native Defects on Photocatalytic Properties of ZnO. J. Phys. Chem. C 2011, 115, 11095–11101. [Google Scholar] [CrossRef]
  51. Pei, Z.; Ding, L.; Hu, J.; Weng, S.; Zheng, Z.; Huang, M.; Liu, P. Defect and its dominance in ZnO films: A new insight into the role of defect over photocatalytic activity. Appl. Catal. B 2013, 142–143, 736–743. [Google Scholar] [CrossRef]
  52. Pei, Z.; Ding, L.; Lu, M.; Fan, Z.; Weng, S.; Hu, J.; Liu, P. Synergistic Effect in Polyaniline-Hybrid Defective ZnO with Enhanced Photocatalytic Activity and Stability. J. Phys. Chem. C 2014, 118, 9570–9577. [Google Scholar] [CrossRef]
  53. Yang, C.; Du, J.; Peng, Q.; Qiao, R.; Chen, W.; Xu, C.; Shuai, Z.; Gao, M. Polyaniline/Fe3O4 Nanoparticle Composite: Synthesis and Reaction Mechanism. J. Phys. Chem. B 2009, 113, 5052–5058. [Google Scholar] [CrossRef] [PubMed]
  54. Xuan, S.; Wang, Y.-X.J.; Leung, K.C.-F.; Shu, K. Synthesis of Fe3O4@Polyaniline Core/Shell Microspheres with Well-Defined Blackberry-Like Morphology. J. Phys. Chem. C 2008, 112, 18804–18809. [Google Scholar] [CrossRef]
  55. Luo, Q.; Wang, L.; Wang, D.; Yin, R.; Li, X.; An, J.; Yang, X. Preparation, characterization and visible-light photocatalytic performances of composite films prepared from polyvinyl chloride and SnO2 nanoparticles. J. Environ. Chem. Eng. 2015, 3, 622–629. [Google Scholar] [CrossRef]
  56. Babu, B.; Cho, M.; Byon, C.; Shim, J. Sunlight-driven photocatalytic activity of SnO2 QDs-g-C3N4 nanolayers. Mater. Lett. 2018, 212, 327–331. [Google Scholar] [CrossRef]
  57. Chen, Y.; Sun, F.; Huang, Z.; Chen, H.; Zhuang, Z.; Pan, Z.; Long, J.; Gu, F. Photochemical fabrication of SnO2 dense layers on reduced graphene oxide sheets for application in photocatalytic degradation of p-Nitrophenol. Appl. Catal. B 2017, 215, 8–17. [Google Scholar] [CrossRef]
  58. Ma, H.; Li, C.; Yin, J.; Pu, X.; Zhang, D.; Su, C.; Wang, X.; Shao, X. Polyoxometalate enhances the photocatalytic performance of polyaniline/SnO2 composites. Mater. Lett. 2016, 168, 103–106. [Google Scholar] [CrossRef]
  59. Li, J.; Peng, T.; Zhang, Y.; Zhou, C.; Zhu, A. Polyaniline modified SnO2 nanoparticles for efficient photocatalytic reduction of aqueous Cr(VI) under visible light. Sep. Purif. Technol. 2018, 201, 120–129. [Google Scholar] [CrossRef]
  60. Kefeni, K.K.; Mamba, B.B. Photocatalytic application of spinel ferrite nanoparticles and nanocomposites in wastewater treatment: Review. Sustain. Mater. Technol. 2020, 23, e00140. [Google Scholar] [CrossRef]
  61. Suresh, R.; Rajendran, S.; Kumar, P.S.; Vo, D.-V.N.; Cornejo-Ponce, L. Recent advancements of spinel ferrite based binary nanocomposite photocatalysts in wastewater treatment. Chemosphere 2021, 274, 129734. [Google Scholar] [CrossRef]
  62. Rashad, M.M.; Mohamed, R.M.; Ibrahim, M.A.; Ismail, L.F.M.; Abdel-Aal, E.A. Magnetic and catalytic properties of cubic copper ferrite nanopowders synthesized from secondary resources. Adv. Powder Technol. 2012, 23, 315–323. [Google Scholar] [CrossRef]
  63. Li, X.; Lu, H.; Zhang, Y.; He, F. Efficient removal of organic pollutants from aqueous media using newly synthesized polypyrrole/CNTs-CoFe2O4 magnetic nanocomposites. Chem. Eng. J. 2017, 316, 893–902. [Google Scholar] [CrossRef]
  64. Laokul, P.; Arthan, S.; Maensiri, S.; Swatsitang, E. Magnetic and Optical Properties of CoFe2O4 Nanoparticles Synthesized by Reverse Micelle Microemulsion Method. J. Supercond. Nov. Magn. 2015, 28, 2483–2489. [Google Scholar] [CrossRef]
  65. Ghosh, S.; Kouame, N.A.; Remita, S.; Ramos, L.; Goubard, F.; Aubert, P.-H.; Dazzi, A.; Deniset-Besseau, A.; Remita, H. Visible-light active conducting polymer nanostructures with superior photocatalytic activity. Sci. Rep. 2015, 5, 18002. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  66. O’Neal Tugaoen, H.; Garcia-Segura, S.; Hristovski, K.; Westerhoff, P. Compact light-emitting diode optical fiber immobilized TiO2 reactor for photocatalytic water treatment. Sci. Total Environ. 2018, 613–614, 1331–1338. [Google Scholar] [CrossRef]
  67. Liu, J.; McCarthy, D.L.; Tong, L.; Cowan, M.J.; Kinsley, J.M.; Sonnenberg, L.; Skorenko, K.H.; Boyer, S.M.; DeCoste, J.B.; Bernier, W.E.; et al. Poly(3,4-ethylenedioxythiophene) (PEDOT) infused TiO2 nanofibers: The role of hole transport layer in photocatalytic degradation of phenazopyridine as a pharmaceutical contaminant. RSC Adv. 2016, 6, 113884–113892. [Google Scholar] [CrossRef]
  68. Abdiryim, T.; Ali, A.; Jamal, R.; Osman, Y.; Zhang, Y. A facile solid-state heating method for preparation of poly(3,4-ethelenedioxythiophene)/ZnO nanocomposite and photocatalytic activity. Nanoscale Res. Lett. 2014, 9, 89. [Google Scholar] [CrossRef] [Green Version]
  69. Yan, H.; Zhang, L.; Shen, J.; Chen, Z.; Shi, G.; Zhang, B. Synthesis, property and field-emission behaviour of amorphous polypyrrole nanowires. Nanotechnology 2006, 17, 3446–3450. [Google Scholar] [CrossRef]
  70. Ferreira, C.A.; Domenech, S.C.; Lacaze, P.C. Synthesis and characterization of polypyrrole/TiO2 composites on mild steel. J. Appl. Electrochem. 2001, 31, 49–56. [Google Scholar] [CrossRef]
  71. Tai, H.; Jiang, Y.; Xie, G.; Yu, J.; Zhao, M. Self-assembly of TiO2/polypyrrole nanocomposite ultrathin films and application for an NH3 gas sensor. Int. J. Environ. Anal. Chem. 2007, 87, 539–551. [Google Scholar] [CrossRef]
  72. Liu, Y.; Zhao, C.; Wang, X.; Xu, H.; Wang, H.; Zhao, X.; Feng, J.; Yan, W.; Ren, Z. Preparation of PPy/TiO2 core-shell nanorods film and its photocathodic protection for 304 stainless steel under visible light. Mater. Res. Bull. 2020, 124, 110751. [Google Scholar] [CrossRef]
  73. Liu, Z.; Liu, Y.; Poyraz, S.; Zhang, X. Green-nano approach to nanostructured polypyrrole. Chem. Commun. 2011, 47, 4421–4423. [Google Scholar] [CrossRef]
  74. Ahmad, N.; Sultana, S.; Faisal, S.M.; Ahmed, A.; Sabir, S.; Khan, M.Z. Zinc oxide-decorated polypyrrole/chitosan bionanocomposites with enhanced photocatalytic, antibacterial and anticancer performance. RSC Adv. 2019, 9, 41135–41150. [Google Scholar] [CrossRef] [Green Version]
  75. Chougule, M.A.; Sen, S.; Patil, V.B. Facile and efficient route for preparation of polypyrrole-ZnO nanocomposites: Microstructural, optical, and charge transport properties. J. Appl. Polym. Sci. 2012, 125, E541–E547. [Google Scholar] [CrossRef]
  76. Yang, Y.; Wen, J.; Wei, J.; Xiong, R.; Shi, J.; Pan, C. Polypyrrole-Decorated Ag-TiO2 Nanofibers Exhibiting Enhanced Photocatalytic Activity under Visible-Light Illumination. ACS Appl. Mater. Interfaces 2013, 5, 6201–6207. [Google Scholar] [CrossRef]
  77. Wang, B.; Li, C.; Pang, J.; Qing, X.; Zhai, J.; Li, Q. Novel polypyrrole-sensitized hollow TiO2/fly ash cenospheres: Synthesis, characterization, and photocatalytic ability under visible light. Appl. Surf. Sci. 2012, 258, 9989–9996. [Google Scholar] [CrossRef]
  78. Pruna, A.; Shao, Q.; Kamruzzaman, M.; Li, Y.Y.; Zapien, J.A.; Pullini, D.; Busquets Mataix, D.; Ruotolo, A. Effect of ZnO core electrodeposition conditions on electrochemical and photocatalytic properties of polypyrrole-graphene oxide shelled nanoarrays. Appl. Surf. Sci. 2017, 392, 801–809. [Google Scholar] [CrossRef]
  79. Ong, W.L.; Low, Q.X.; Huang, W.; van Kan, J.A.; Ho, G.W. Patterned growth of vertically-aligned ZnO nanorods on a flexible platform for feasible transparent and conformable electronics applications. J. Mater. Chem. 2012, 22, 8518–8524. [Google Scholar] [CrossRef]
  80. Silvestri, S.; Ferreira, C.D.; Oliveira, V.; Varejão, J.M.T.B.; Labrincha, J.A.; Tobaldi, D.M. Synthesis of PPy-ZnO composite used as photocatalyst for the degradation of diclofenac under simulated solar irradiation. J. Photochem. Photobiol. A 2019, 375, 261–269. [Google Scholar] [CrossRef]
  81. Balakumar, V.; Sekar, K.; Chuaicham, C.; Manivannan, R.; Sasaki, K. Synergistic ternary porous CN–PPy–MMt nanocomposite for efficient photocatalytic metronidazole mineralization: Performance, mechanism, and pathways. Environ. Sci. Nano 2021, 8, 2261–2276. [Google Scholar] [CrossRef]
  82. Ali, H. Facile synthesis of mesoporous TiO2-CdS-polyaniline ternary system with improved optical properties. Mater. Res. Express 2019, 6, 115529. [Google Scholar] [CrossRef]
  83. Meng, S.; Zhang, J.; Chen, S.; Zhang, S.; Huang, W. Perspective on construction of heterojunction photocatalysts and the complete utilization of photogenerated charge carriers. Appl. Surf. Sci. 2019, 476, 982–992. [Google Scholar] [CrossRef]
  84. Shylesh, S.; Schünemann, V.; Thiel, W.R. Magnetically Separable Nanocatalysts: Bridges between Homogeneous and Heterogeneous Catalysis. Angew. Chem. Int. Ed. 2010, 49, 3428–3459. [Google Scholar] [CrossRef]
  85. Zhang, Q.; Fan, W.; Gao, L. Anatase TiO2 nanoparticles immobilized on ZnO tetrapods as a highly efficient and easily recyclable photocatalyst. Appl. Catal. B 2007, 76, 168–173. [Google Scholar] [CrossRef]
  86. Zhou, M.; Yu, J.; Liu, S.; Zhai, P.; Jiang, L. Effects of calcination temperatures on photocatalytic activity of SnO2/TiO2 composite films prepared by an EPD method. J. Hazard. Mater. 2008, 154, 1141–1148. [Google Scholar] [CrossRef]
  87. Berger, C.; Song, Z.; Li, X.; Wu, X.; Brown, N.; Naud, C.; Mayou, D.; Li, T.; Hass, J.; Marchenkov, A.N.; et al. Electronic Confinement and Coherence in Patterned Epitaxial Graphene. Science 2006, 312, 1191–1196. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  88. Dong, S.; Ding, X.; Guo, T.; Yue, X.; Han, X.; Sun, J. Self-assembled hollow sphere shaped Bi2WO6/RGO composites for efficient sunlight-driven photocatalytic degradation of organic pollutants. Chem. Eng. J. 2017, 316, 778–789. [Google Scholar] [CrossRef]
  89. Akhavan, O. Graphene Nanomesh by ZnO Nanorod Photocatalysts. ACS Nano 2010, 4, 4174–4180. [Google Scholar] [CrossRef]
  90. 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]
  91. Wang, J.; Tsuzuki, T.; Tang, B.; Hou, X.; Sun, L.; Wang, X. Reduced Graphene Oxide/ZnO Composite: Reusable Adsorbent for Pollutant Management. ACS Appl. Mater. Interfaces 2012, 4, 3084–3090. [Google Scholar] [CrossRef]
  92. 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]
  93. Xie, X.; Kretschmer, K.; Wang, G. Advances in graphene-based semiconductor photocatalysts for solar energy conversion: Fundamentals and materials engineering. Nanoscale 2015, 7, 13278–13292. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  94. Han, C.; Zhang, N.; Xu, Y.-J. Structural diversity of graphene materials and their multifarious roles in heterogeneous photocatalysis. Nano Today 2016, 11, 351–372. [Google Scholar] [CrossRef]
  95. Ameen, S.; Seo, H.-K.; Shaheer Akhtar, M.; Shin, H.S. Novel graphene/polyaniline nanocomposites and its photocatalytic activity toward the degradation of rose Bengal dye. Chem. Eng. J. 2012, 210, 220–228. [Google Scholar] [CrossRef]
  96. Yang, Y.; Ma, Z.; Xu, L.; Wang, H.; Fu, N. Preparation of reduced graphene oxide/meso-TiO2/AuNPs ternary composites and their visible-light-induced photocatalytic degradation n of methylene blue. Appl. Surf. Sci. 2016, 369, 576–583. [Google Scholar] [CrossRef]
  97. Miao, J.; Xie, A.; Li, S.; Huang, F.; Cao, J.; Shen, Y. A novel reducing graphene/polyaniline/cuprous oxide composite hydrogel with unexpected photocatalytic activity for the degradation of Congo red. Appl. Surf. Sci. 2016, 360, 594–600. [Google Scholar] [CrossRef]
  98. Pandiselvi, K.; Fang, H.; Huang, X.; Wang, J.; Xu, X.; Li, T. Constructing a novel carbon nitride/polyaniline/ZnO ternary heterostructure with enhanced photocatalytic performance using exfoliated carbon nitride nanosheets as supports. J. Hazard. Mater. 2016, 314, 67–77. [Google Scholar] [CrossRef]
  99. Liras, M.; Barawi, M.; de la Peña O’Shea, V.A. Hybrid materials based on conjugated polymers and inorganic semiconductors as photocatalysts: From environmental to energy applications. Chem. Soc. Rev. 2019, 48, 5454–5487. [Google Scholar] [CrossRef]
  100. Deng, X.; Chen, Y.; Wen, J.; Xu, Y.; Zhu, J.; Bian, Z. Polyaniline-TiO2 composite photocatalysts for light-driven hexavalent chromium ions reduction. Sci. Bull. 2020, 65, 105–112. [Google Scholar] [CrossRef] [Green Version]
  101. Xu, S.; Han, Y.; Xu, Y.; Meng, H.; Xu, J.; Wu, J.; Xu, Y.; Zhang, X. Fabrication of polyaniline sensitized grey-TiO2 nanocomposites and enhanced photocatalytic activity. Sep. Purif. Technol. 2017, 184, 248–256. [Google Scholar] [CrossRef]
  102. Ma, J.; Dai, J.; Duan, Y.; Zhang, J.; Qiang, L.; Xue, J. Fabrication of PANI-TiO2/rGO hybrid composites for enhanced photocatalysis of pollutant removal and hydrogen production. Renew. Energy 2020, 156, 1008–1018. [Google Scholar] [CrossRef]
  103. Wu, J.; Feng, Y.; Logan, B.E.; Dai, C.; Han, X.; Li, D.; Liu, J. Preparation of Al–O-Linked Porous-g-C3N4/TiO2-Nanotube Z-Scheme Composites for Efficient Photocatalytic CO2 Conversion and 2,4-Dichlorophenol Decomposition and Mechanism. ACS Sustain. Chem. Eng. 2019, 7, 15289–15296. [Google Scholar] [CrossRef]
  104. Ding, Y.; Zhu, L.; Wang, N.; Tang, H. Sulfate radicals induced degradation of tetrabromobisphenol A with nanoscaled magnetic CuFe2O4 as a heterogeneous catalyst of peroxymonosulfate. Appl. Catal. B 2013, 129, 153–162. [Google Scholar] [CrossRef]
  105. Guan, Y.-H.; Ma, J.; Ren, Y.-M.; Liu, Y.-L.; Xiao, J.-Y.; Lin, L.-q.; Zhang, C. Efficient degradation of atrazine by magnetic porous copper ferrite catalyzed peroxymonosulfate oxidation via the formation of hydroxyl and sulfate radicals. Water Res. 2013, 47, 5431–5438. [Google Scholar] [CrossRef] [PubMed]
  106. Nath, B.K.; Chaliha, C.; Kalita, E.; Kalita, M.C. Synthesis and characterization of ZnO:CeO2:nanocellulose:PANI bionanocomposite. A bimodal agent for arsenic adsorption and antibacterial action. Carbohydr. Polym. 2016, 148, 397–405. [Google Scholar] [CrossRef]
  107. Faisal, M.; Harraz, F.A.; Ismail, A.A.; Alsaiari, M.A.; Al-Sayari, S.A.; Al-Assiri, M.S. Novel synthesis of Polyaniline/SrSnO3 nanocomposites with enhanced photocatalytic activity. Ceram. Int. 2019, 45, 20484–20492. [Google Scholar] [CrossRef]
  108. Kharazi, P.; Rahimi, R.; Rabbani, M. Copper ferrite-polyaniline nanocomposite: Structural, thermal, magnetic and dye adsorption properties. Solid State Sci. 2019, 93, 95–100. [Google Scholar] [CrossRef]
  109. Zhou, Q.; Wang, Y.; Xiao, J.; Zhan, Y. Preparation of magnetic core-shell Fe3O4@polyaniline composite material and its application in adsorption and removal of tetrabromobisphenol A and decabromodiphenyl ether. Ecotoxicol. Environ. Saf. 2019, 183, 109471. [Google Scholar] [CrossRef]
  110. Hayashi, H.; Lightcap, I.V.; Tsujimoto, M.; Takano, M.; Umeyama, T.; Kamat, P.V.; Imahori, H. Electron Transfer Cascade by Organic/Inorganic Ternary Composites of Porphyrin, Zinc Oxide Nanoparticles, and Reduced Graphene Oxide on a Tin Oxide Electrode that Exhibits Efficient Photocurrent Generation. J. Am. Chem. Soc. 2011, 133, 7684–7687. [Google Scholar] [CrossRef]
  111. Iwase, A.; Ng, Y.H.; Ishiguro, Y.; Kudo, A.; Amal, R. Reduced Graphene Oxide as a Solid-State Electron Mediator in Z-Scheme Photocatalytic Water Splitting under Visible Light. J. Am. Chem. Soc. 2011, 133, 11054–11057. [Google Scholar] [CrossRef] [PubMed]
  112. Zhang, H.; Zhu, Y. Significant Visible Photoactivity and Antiphotocorrosion Performance of CdS Photocatalysts after Monolayer Polyaniline Hybridization. J. Phys. Chem. C 2010, 114, 5822–5826. [Google Scholar] [CrossRef]
  113. Lin, Y.; Li, D.; Hu, J.; Xiao, G.; Wang, J.; Li, W.; Fu, X. Highly Efficient Photocatalytic Degradation of Organic Pollutants by PANI-Modified TiO2 Composite. J. Phys. Chem. C 2012, 116, 5764–5772. [Google Scholar] [CrossRef]
  114. Mikkelsen, M.; Jørgensen, M.; Krebs, F.C. The teraton challenge. A review of fixation and transformation of carbon dioxide. Energy Environ. Sci. 2010, 3, 43–81. [Google Scholar] [CrossRef]
  115. Mao, J.; Li, K.; Peng, T. Recent advances in the photocatalytic CO2 reduction over semiconductors. Catal. Sci. Technol. 2013, 3, 2481–2498. [Google Scholar] [CrossRef]
  116. Dhakshinamoorthy, A.; Navalon, S.; Corma, A.; Garcia, H. Photocatalytic CO2 reduction by TiO2 and related titanium containing solids. Energy Environ. Sci. 2012, 5, 9217–9233. [Google Scholar] [CrossRef]
  117. Wang, S.; Han, X.; Zhang, Y.; Tian, N.; Ma, T.; Huang, H. Inside-and-Out Semiconductor Engineering for CO2 Photoreduction: From Recent Advances to New Trends. Small Struct. 2021, 2, 2000061. [Google Scholar] [CrossRef]
  118. Liu, G.; Xie, S.; Zhang, Q.; Tian, Z.; Wang, Y. Carbon dioxide-enhanced photosynthesis of methane and hydrogen from carbon dioxide and water over Pt-promoted polyaniline–TiO2 nanocomposites. Chem. Commun. 2015, 51, 13654–13657. [Google Scholar] [CrossRef]
  119. Indrakanti, V.P.; Kubicki, J.D.; Schobert, H.H. Photoinduced activation of CO2 on Ti-based heterogeneous catalysts: Current state, chemical physics-based insights and outlook. Energy Environ. Sci. 2009, 2, 745–758. [Google Scholar] [CrossRef]
  120. Roy, S.C.; Varghese, O.K.; Paulose, M.; Grimes, C.A. Toward Solar Fuels: Photocatalytic Conversion of Carbon Dioxide to Hydrocarbons. ACS Nano 2010, 4, 1259–1278. [Google Scholar] [CrossRef]
  121. Califano, M.; Zhou, Y. Inverse-designed semiconductor nanocatalysts for targeted CO2 reduction in water. Nanoscale 2021, 13, 10024–10034. [Google Scholar] [CrossRef]
  122. Fiorenza, R.; Bellardita, M.; Balsamo, S.A.; Spitaleri, L.; Gulino, A.; Condorelli, M.; D’Urso, L.; Scirè, S.; Palmisano, L. A solar photothermocatalytic approach for the CO2 conversion: Investigation of different synergisms on CoO-CuO/brookite TiO2-CeO2 catalysts. Chem. Eng. J. 2021, 428, 131249. [Google Scholar] [CrossRef]
  123. Kim, Y.; Choi, K.; Jung, J.; Park, S.; Kim, P.-G.; Park, J. Aquatic toxicity of acetaminophen, carbamazepine, cimetidine, diltiazem and six major sulfonamides, and their potential ecological risks in Korea. Environ. Int. 2007, 33, 370–375. [Google Scholar] [CrossRef]
  124. Van Tran, V.; Park, D.; Lee, Y.-C. Hydrogel applications for adsorption of contaminants in water and wastewater treatment. Environ. Sci. Pollut. Res. Int. 2018, 25, 24569–24599. [Google Scholar] [CrossRef]
  125. Zhu, J.; Zhu, Z.; Zhang, H.; Lu, H.; Zhang, W.; Qiu, Y.; Zhu, L.; Küppers, S. Calcined layered double hydroxides/reduced graphene oxide composites with improved photocatalytic degradation of paracetamol and efficient oxidation-adsorption of As(III). Appl. Catal. B 2018, 225, 550–562. [Google Scholar] [CrossRef]
  126. Nicomel, N.R.; Leus, K.; Folens, K.; Van Der Voort, P.; Du Laing, G. Technologies for Arsenic Removal from Water: Current Status and Future Perspectives. Int. J. Environ. Res. Public Health 2016, 13, 62. [Google Scholar] [CrossRef] [PubMed]
  127. Liu, J.; Fang, W.; Wang, Y.; Xing, M.; Zhang, J. Gold-loaded graphene oxide/PDPB composites for the synchronous removal of Cr(VI) and phenol. Chinese J. Catal. 2018, 39, 8–15. [Google Scholar] [CrossRef]
  128. Liu, F.; Yu, J.; Tu, G.; Qu, L.; Xiao, J.; Liu, Y.; Wang, L.; Lei, J.; Zhang, J. Carbon nitride coupled Ti-SBA15 catalyst for visible-light-driven photocatalytic reduction of Cr (VI) and the synergistic oxidation of phenol. Appl. Catal. B 2017, 201, 1–11. [Google Scholar] [CrossRef]
  129. Yang, X.; Wang, X.; Liu, X.; Zhang, Y.; Song, W.; Shu, C.; Jiang, L.; Wang, C. Preparation of graphene-like iron oxide nanofilm/silica composite with enhanced adsorption and efficient photocatalytic properties. J. Mater. Chem. A 2013, 1, 8332–8337. [Google Scholar] [CrossRef]
  130. Mou, F.; Guan, J.; Xiao, Z.; Sun, Z.; Shi, W.; Fan, X.-a. Solvent-mediated synthesis of magnetic Fe2O3 chestnut-like amorphous-core/γ-phase-shell hierarchical nanostructures with strong As(v) removal capability. J. Mater. Chem. 2011, 21, 5414–5421. [Google Scholar] [CrossRef]
  131. Wang, Y.; Zhang, P.; Zhang, T.C.; Xiang, G.; Wang, X.; Pehkonen, S.; Yuan, S. A magnetic γ-Fe2O3@PANI@TiO2 core–shell nanocomposite for arsenic removal via a coupled visible-light-induced photocatalytic oxidation–adsorption process. Nanoscale Adv. 2020, 2, 2018–2024. [Google Scholar] [CrossRef] [Green Version]
  132. Vellaichamy, B.; Periakaruppan, P.; Nagulan, B. Reduction of Cr6+ from Wastewater Using a Novel in Situ-Synthesized PANI/MnO2/TiO2 Nanocomposite: Renewable, Selective, Stable, and Synergistic Catalysis. ACS Sustain. Chem. Eng. 2017, 5, 9313–9324. [Google Scholar] [CrossRef]
  133. Lin, C.J.; Wang, S.L.; Huang, P.M.; Tzou, Y.M.; Liu, J.C.; Chen, C.C.; Chen, J.H.; Lin, C. Chromate reduction by zero-valent Al metal as catalyzed by polyoxometalate. Water Res. 2009, 43, 5015–5022. [Google Scholar] [CrossRef]
  134. Guo, L.; Xiao, Y.; Wang, Y. Hexavalent Chromium-induced Alteration of Proteomic Landscape in Human Skin Fibroblast Cells. J. Proteome Res. 2013, 12, 3511–3518. [Google Scholar] [CrossRef] [Green Version]
  135. Wang, L.; Jin, P.; Duan, S.; She, H.; Huang, J.; Wang, Q. In-situ incorporation of Copper(II) porphyrin functionalized zirconium MOF and TiO2 for efficient photocatalytic CO2 reduction. Sci. Bull. 2019, 64, 926–933. [Google Scholar] [CrossRef] [Green Version]
  136. Bui, V.K.H.; Park, D.; Tran, V.V.; Lee, G.-W.; Oh, S.Y.; Huh, Y.S.; Lee, Y.-C. One-Pot Synthesis of Magnesium Aminoclay-Titanium Dioxide Nanocomposites for Improved Photocatalytic Performance. J. Nanosci. Nanotech. 2018, 18, 6070–6074. [Google Scholar] [CrossRef]
  137. Qiu, B.; Xu, C.; Sun, D.; Yi, H.; Guo, J.; Zhang, X.; Qu, H.; Guerrero, M.; Wang, X.; Noel, N.; et al. Polyaniline Coated Ethyl Cellulose with Improved Hexavalent Chromium Removal. ACS Sustain. Chem. Eng. 2014, 2, 2070–2080. [Google Scholar] [CrossRef]
  138. Xiong, S.; Phua, S.L.; Dunn, B.S.; Ma, J.; Lu, X. Covalently Bonded Polyaniline−TiO2 Hybrids: A Facile Approach to Highly Stable Anodic Electrochromic Materials with Low Oxidation Potentials. Chem. Mater. 2010, 22, 255–260. [Google Scholar] [CrossRef]
  139. Ansari, M.O.; Khan, M.M.; Ansari, S.A.; Raju, K.; Lee, J.; Cho, M.H. Enhanced Thermal Stability under DC Electrical Conductivity Retention and Visible Light Activity of Ag/TiO2@Polyaniline Nanocomposite Film. ACS Appl. Mater. Interfaces 2014, 6, 8124–8133. [Google Scholar] [CrossRef] [PubMed]
  140. Qiao, Y.; Bao, S.-J.; Li, C.M.; Cui, X.-Q.; Lu, Z.-S.; Guo, J. Nanostructured Polyaniline/Titanium Dioxide Composite Anode for Microbial Fuel Cells. ACS Nano 2008, 2, 113–119. [Google Scholar] [CrossRef] [PubMed]
Figure 1. (a) SEM image of the as-prepared Fe3O4/PANI core/shell composite [54]. (b) CoFe2O4/PANI hollow nanofibers [29].
Figure 1. (a) SEM image of the as-prepared Fe3O4/PANI core/shell composite [54]. (b) CoFe2O4/PANI hollow nanofibers [29].
Polymers 13 03031 g001
Figure 2. The preparation process of PEDOT infused TiO2 nanofibers via electrospinning and calcination [67]: (A) Sol–gel solution containing PMMA, precursor TTIP and the solvents; (B) post-electrospinned polymer fibers containing PMMA, amorphous TiO2 and solvent without evaporating; (C) post-calcined TiO2 nanofibers under different calcination temperatures with different phase compositions; (D) stirring of post-calcined TiO2 nanofibers with yellow oxidant Fe(PTS)3 solution; (E) dried TiO2 nanofibers with adsorbed Fe(PTS)3 on surface were put in a heated chamber for VPP reaction; (F) bluish dried PEDOT infused TiO2 nanofibers on filter paper.
Figure 2. The preparation process of PEDOT infused TiO2 nanofibers via electrospinning and calcination [67]: (A) Sol–gel solution containing PMMA, precursor TTIP and the solvents; (B) post-electrospinned polymer fibers containing PMMA, amorphous TiO2 and solvent without evaporating; (C) post-calcined TiO2 nanofibers under different calcination temperatures with different phase compositions; (D) stirring of post-calcined TiO2 nanofibers with yellow oxidant Fe(PTS)3 solution; (E) dried TiO2 nanofibers with adsorbed Fe(PTS)3 on surface were put in a heated chamber for VPP reaction; (F) bluish dried PEDOT infused TiO2 nanofibers on filter paper.
Polymers 13 03031 g002
Figure 3. Schematic illustration of the preparation process of a flexible ZnO-microrod/PPy composite film. (a) Flexible base polyester film. (b) ZnO-microrod arrays were grown on the base polyester film. (c) PPy was electrodeposited onto the ZnO-microrod arrays. (d) Optical image of highly flexible ZnO-microrod/PPy composite film (20 mm × 20 mm) [33].
Figure 3. Schematic illustration of the preparation process of a flexible ZnO-microrod/PPy composite film. (a) Flexible base polyester film. (b) ZnO-microrod arrays were grown on the base polyester film. (c) PPy was electrodeposited onto the ZnO-microrod arrays. (d) Optical image of highly flexible ZnO-microrod/PPy composite film (20 mm × 20 mm) [33].
Polymers 13 03031 g003
Figure 4. (a) Schematic representation of the synthesis of PANI/TiO2/graphene nanocomposite (b) PL and (c) UV–Vis diffuse absorbance spectra of TiO2, TiO2/GN and PANI/TiO2/graphene nanocomposite [39]. (d) Microstructure and photocatalytic mechanism diagram of ternary GO-ZnFe2O4-PANI composite [36].
Figure 4. (a) Schematic representation of the synthesis of PANI/TiO2/graphene nanocomposite (b) PL and (c) UV–Vis diffuse absorbance spectra of TiO2, TiO2/GN and PANI/TiO2/graphene nanocomposite [39]. (d) Microstructure and photocatalytic mechanism diagram of ternary GO-ZnFe2O4-PANI composite [36].
Polymers 13 03031 g004
Figure 5. (a) Proposed mechanism for the photocatalytic degradation using a binary PANI/TiO2 composite under visible-light irradiation [101]. (b) Schematic diagram of the charge transfer pathway in a ternary Z-scheme Cu2O/ZnO-PANI composite under visible-light irradiation [38]. (c) Mechanism of photocatalytic activity of ternary PANI/TiO2/rGO composite [102].
Figure 5. (a) Proposed mechanism for the photocatalytic degradation using a binary PANI/TiO2 composite under visible-light irradiation [101]. (b) Schematic diagram of the charge transfer pathway in a ternary Z-scheme Cu2O/ZnO-PANI composite under visible-light irradiation [38]. (c) Mechanism of photocatalytic activity of ternary PANI/TiO2/rGO composite [102].
Polymers 13 03031 g005
Figure 6. (a) Photodegradation of various dyes on TiO2-CoFe2O4-PANI. (b) Photodegradation of MO on TiO2-CoFe2O4-PANI without irradiation and under visible-light irradiation. (c) Photodegradation of MO on TiO2-CoFe2O4-PANI over several cycles. Inset shows the magnetic separation after photodegradation. (d) XRD patterns of TiO2-CoFe2O4-PANI before and after photocatalysis [34].
Figure 6. (a) Photodegradation of various dyes on TiO2-CoFe2O4-PANI. (b) Photodegradation of MO on TiO2-CoFe2O4-PANI without irradiation and under visible-light irradiation. (c) Photodegradation of MO on TiO2-CoFe2O4-PANI over several cycles. Inset shows the magnetic separation after photodegradation. (d) XRD patterns of TiO2-CoFe2O4-PANI before and after photocatalysis [34].
Polymers 13 03031 g006
Figure 7. Schematic illustration of the mechanism for CO2 photoreduction using a binary PANI/TiO2 composite [99].
Figure 7. Schematic illustration of the mechanism for CO2 photoreduction using a binary PANI/TiO2 composite [99].
Polymers 13 03031 g007
Figure 8. (a) Schematic illustration of the preparation procedure for ternary γ-Fe2O3/PANI/TiO2 composite for As(III) photoreduction [131]. (b) Liquid-phase photocatalytic Cr(VI) anion reduction by the binary PANI/TiO2 composite [100]. (c) Photographic images of Cr(VI) reduction by ternary PANI/MnO2/TiO2 nanocomposite [132].
Figure 8. (a) Schematic illustration of the preparation procedure for ternary γ-Fe2O3/PANI/TiO2 composite for As(III) photoreduction [131]. (b) Liquid-phase photocatalytic Cr(VI) anion reduction by the binary PANI/TiO2 composite [100]. (c) Photographic images of Cr(VI) reduction by ternary PANI/MnO2/TiO2 nanocomposite [132].
Polymers 13 03031 g008
Table 1. Summarized properties and applications of conducting polymer/metal oxide composites as photocatalysts.
Table 1. Summarized properties and applications of conducting polymer/metal oxide composites as photocatalysts.
CompositeBand-Gap Energy (eV)Photocatalytic PropertiesApplicationsReference
Binary Composite of CPs and Metal Oxides
Mesoporous PANI/TiO2-Enhanced water oxidation efficiency under sunlight irradiation, reaching about two-fold higher photocurrent densities than pure TiO2 nanoparticles Water splitting[20]
PANI/TiO2 nanorods3.1
-
Quantum yield: 9.86 × 10−5 molecules/photon and 2.82 × 10−5 molecules/photon for PANI/TiO2 and TiO2, respectively.
-
PANI/TiO2 showed better performance than TiO2 with a rate constant of 4.46 × 10−2 min−1 compared with 2.18 × 10−2 min−1, respectively
Degradation of organic pollutants (Bisphenol A)[21]
PANI nanobelt/TiO22.77The photocatalytic degradation rate of rhodamine B was 99%Degradation of rhodamine B[22]
PANI/TiO2 nanotubes-The photocatalytic activity can easily be tuned using a particular type and concentration of the acid dopant in the redoping processDegradation of rhodamine B[23]
PANI/ZnO-
-
The composite exhibits a dramatic photocatalytic activity both under ultraviolet and visible-light irradiation
-
The photo-corrosion of ZnO was successfully inhibited
Degradation of methylene blue (MB)[24]
PANI/ZnO2.13–2.22The composite photocatalysts’ activity was broadened into the Vis regionDegradation of acid blue[25]
PANI/Fe3O4-The adsorption process prevails in relation to photodegradationDegradation of MB[26]
PANI/SnO22.7Increased photocatalytic activity for visible light is due to its electrical conductivity and efficient charge separationDegradation of direct blue 15[27]
PANI/Sn3O42.06Photocatalytic activity for visible light is 2.27 times higher than that of Sn3O4 aloneDegradation of rhodamine B[28]
PANI/CoFe2O4-Photocatalytic activity under visible-light irradiation with CoFe2O4/PANI was 80 times greater than for CoFe2O4Degradation of methyl orange (MO)[29]
PEDOT/TiO23.01–3.05PEDOT infused TiO2 nanofiber, exhibits the highest degradation enhancement (125%)Degradation of phenazopyridine[30]
PPy/TiO2-The photoactivity of the nanocomposite arises from the electron transfer from excited PPy to TiO2 nanoparticles and further across the nanocomposite interfaceDegradation of MB[31]
PPy/TiO23.08–3.11The photoactivity of nanocomposites increased by 41% compared with pure TiO2Degradation of RhB and CO2[32]
ZnO-microrods/PPy1.7Composite films achieve a much higher photocatalytic efficiency in comparison with pure ZnO-microrod arrays (a rate of 22%/min MB degradation)Degradation of MB[33]
Ternary Composites
TiO2-CoFe2O4-PANI-The ternary TiO2-CoFe2O4-PANI composite shows a highly enhanced photocatalytic activity in the range of visible light, compared with the binary TiO2-CoFe2O4, CoFe2O4-PANI, or TiO2-PANI compositesDegradation of methyl orange[34]
ZnFe2O4-TiO2-PANI-The decontaminating efficiency of composites on MO and RhB reached up to 98%Degradation and adsorption of MO and RhB[35]
rGO-ZnFe2O4-PANI-The photocatalytic activity still stays above 90% after five recyclesDegradation of RhB[36]
ZnO/rGO/PANI-The photocatalyst shows an enhanced photocatalytic performance in the photodegradation of MO (almost 100%)Degradation of methyl orange[37]
Cu2O/ZnO-PANI2.68The ternary composite with Z-scheme heterojunction properties displayed outstanding adsorption properties, super-fast photocatalytic activities as well as enhanced stabilityDegradation of congo red[38]
PANI/TiO2/graphene2.1High photocatalytic activity is partly due to the sensitizing effect of PANI and the low recombination rate due to the graphene electron scavenging propertyDegradation of MB[39]
PANI-rGO-MnO21.92The ternary composite exhibited significantly enhanced catalytic and photocatalytic activity under visible-light irradiation within 2 hDegradation of MB[40]
g-C3N4/TiO2/PANI2.58Greatly enhanced photocatalytic degradation and high reusabilityDegradation of congo red[41]
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Tran, V.V.; Nu, T.T.V.; Jung, H.-R.; Chang, M. Advanced Photocatalysts Based on Conducting Polymer/Metal Oxide Composites for Environmental Applications. Polymers 2021, 13, 3031. https://doi.org/10.3390/polym13183031

AMA Style

Tran VV, Nu TTV, Jung H-R, Chang M. Advanced Photocatalysts Based on Conducting Polymer/Metal Oxide Composites for Environmental Applications. Polymers. 2021; 13(18):3031. https://doi.org/10.3390/polym13183031

Chicago/Turabian Style

Tran, Vinh Van, Truong Thi Vu Nu, Hong-Ryun Jung, and Mincheol Chang. 2021. "Advanced Photocatalysts Based on Conducting Polymer/Metal Oxide Composites for Environmental Applications" Polymers 13, no. 18: 3031. https://doi.org/10.3390/polym13183031

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