Introduction

In recent decades, the extensive use of organic dyes, antibiotics, pesticides, and other organic pollutants has caused large-scale environmental pollution [1,2,3,4,5]. To reduce water pollution, these pollutants are removed from wastewater using various technologies including biodegradation, adsorption, ozonation, photocatalytic degradation, physicochemical treatment, catalytic reduction, and coagulation/flocculation [6]. Recently, advanced oxidation processes (AOPs) have attracted much attention for the oxidation treatment of organic pollutants. In different AOPs, multiphase photocatalysis is one of the green chemical methods for the removal of various pollutants with low cost, wide application, and eco-friendly characteristic [7]. Electrons can be excited and transitioned from the valence band (VB) to the conduction band (CB), with irradiation at a certain wavelength on the surface of photocatalytic materials, leaving holes in the VB whose energy is equivalent to the band gap energy of the irradiated light [8]. The electrons and holes oxidize and reduce organic pollutants and mineralize these organic compounds into carbon dioxide, water, and inorganic acids [9]. Photocatalytic degradation of organic compounds using semiconductor materials such as TiO2, ZnO, SnO2, WO3, ZrO2, V2O5, CdS, CuO and MoO3 has been reported [10,11,12,13,14,15,16,17,18]. In addition, semiconductor photocatalysts can also be used in photocatalytic hydrogen production and antimicrobial applications [19,20,21]. Although these traditional semiconductor photocatalysts have the advantages of high efficiency, convenient preparation, high stability, reusability, and environmental friendliness, they have also numerous disadvantages as follows: (i) The surface of metal oxides can only absorb less than 5% of visible and UV radiation for the wide bandgap, resulting in a quick recombination of the photoinduced electron–hole pairs and the limited ability to degrade organic pollutants [22]; (ii) organic pollutants can be completely degraded in the presence of semiconductor photocatalysts, but the formation of a large number of unwanted end products and intermediates is unavoidable in the oxidation of azo dyes [7]; (iii) agglomeration of metal oxide nanoparticles (NPs) decreases the surface area and active sites, reducing the photocatalytic performance and making it difficult to recover the catalyst [23]. To effectively solve the abovementioned problems and improve the photocatalytic activity, conductive materials have been used as electron acceptors to transfer photogenerated electrons, effectively blocking the recombination of photogenerated electron–hole pairs. Among many conductive materials, conductive polymers with excellent processability, including polyaniline (PANI), polypyrrole (PPy), and polythiophene (PT), have been widely studied in basic research and industrial applications [24]. PANI is one of the most attractive members of the intrinsically conductive polymer family because of facile preparation, unique doping mechanism, low toxicity, low cost, large surface area/volume ratio, excellent environmental stability, acid–base properties, and special redox properties [25]. For these reasons, PANI-based materials and composites have been used in numerous fields such as photovoltaics, anticorrosion, adsorption, biomedical equipment, sensors, electrochemistry, and electronics [26]. Particularly, PANI has a high absorption coefficient in the visible-light range and high mobility of charge carriers. In addition, PANI is an excellent electron donor and hole acceptor after irradiation [27]. Owing to these special properties, PANI has attracted much interest in photocatalysis. In this review, recent advanced compositions, preparation methods, testing conditions, possible mechanism, and improvement in photocatalytic efficiency of PANI-based composites are summarized.

Photocatalytic decomposition of dyes

Photocatalytic degradation of organic pollutants mainly involves dyes such as rhodamine B (RhB), methylene blue (MB), methyl orange (MO), malachite green (MG), Congo red (CR), and crystal violet. They have become the primary sources of water pollution, and their toxicity and carcinogenicity have attracted much attention [28,29,30,31,32,33,34]. Degradation of various dyes has been extensively studied in literature via the photocatalysis of pristine PANI or PANI-based composites [35,36,37,38,39,40,41,42,43,44].

Organic dyes can be divided in two basic groups [45]. Cationic dyes include a positive charge, usually localized on the nitrogen atom, which is balanced by a counter ion. MB, RhB, and safranin are important examples of cationic dyes and model pollutants in photodegradation. Anionic dyes usually carry a negative charge (sulfonic or carboxyl groups) and can be delivered as sodium salts. MO and CR are probably the most important members in this family, and they have been used as model substrates to evaluate the degradation reactivity of photocatalysts.

Degradation of MB using PANI composites

Rahman and Kar reported that the incorporation of PANI on TiO2 efficiently degraded organic dye MB under UV-light exposure, and the proposed mechanism is shown in Fig. 1 [46]. Under UV-light irradiation, both PANI and TiO2 generated electron–hole pairs with synergistic effect. The LUMO electrons in PANI were transferred to the CB of TiO2, while the photogenerated holes in the VB of TiO2 migrated to the HOMO of PANI. Moreover, the experimental results show that 0.6 M PANI-TiO2 nanocomposite produced the highest dye degradation efficiency, with an improved efficiency of 2.5 times to pristine PANI and 3.1 times to TiO2.

Fig. 1
figure 1

Mechanism of PANI-TiO2 photocatalytic dye degradation

Novel PANI-SrSnO3 (PANI/SrSnO3) binary nanocomposites with different PANI contents (0 − 10 wt%) were successfully prepared using a simple and direct mechanical grinding process followed by an ultrasonic technique [47]. The XRD and FTIR studies exhibited the formation a standard single-phase of orthorhombic SrSnO3 and the presence of PANI with perovskite SrSnO3. All PANI/SrSnO3 nanocomposites showed better photocatalytic efficiency on the degradation of MB dye than either free PANI or SrSnO3 under UV irradiation. Particularly, 5% PANI/SrSnO3 nanocomposite exhibited excellent catalytic performance with 83% destruction rate of MB within 4 h, which was nearly four folds of activity than free SrSnO3 with sufficient stability and durability.

Besides combining with semiconductor nanomaterials to prepare composite photocatalysts, PANI can also be used to construct metal-free photocatalysts. Graphitic carbon nitride (g-C3N4) nanosheets (CNns) were modified by codoping PANI with an inorganic acid (hydrochloric acid, HCl) and an organic acid (phytic acid, PA) [48]. As a result of the synergistic effect of HCl and PA codoping on PANI through intrachain and interchain connection (Fig. 2), PANI/CNns obtained the characteristics of high electrical conductivity, large specific surface area, inhibition of charge recombination, and rich in free radicals, substantially improving the photocatalytic performance. When 1P1C sample (mPANI:mCNns = 1:1) was used, the degradation efficiency of MB reached 98% within 40 min under simulated sunlight irradiation.

Fig. 2
figure 2

Schematic displaying the preparation of PANI/CNns and the doping behavior of PA and HCl with interchain, intrachain, and interchain/intrachain doping on PANI

Hosseini et al. synthesized a ternary nanocomposite with in-situ oxidative polymerization, camphor sulfonic acid doped PANI-WO3-multiwall carbon nanotube (CSA PANI-WO3-CNT) [49]. The degradation rate of MB dye in 60 min illumination using this nanocomposite reached 91.40%, higher than that of free WO3 (43.45%), free CSA PANI (48.4%), and CSA PANI-WO3 binary nanocomposite (85.15%). A schematic illustration of MB photodegradation by CSA PANI-WO3-CNT is shown in Fig. 3. During light irradiation, the electrons of CSA PANI and WO3 were simultaneously excited. The transfer of electrons from the LUMO of CSA PANI (− 1.63 V vs. NHE) to the CB of WO3 (0.6 V vs. NHE) and the holes from the VB of WO3 (3.58 V vs. NHE) to the HOMO of CSA PANI (0.8 V vs. NHE) occurred for the appropriate band alignment between the inorganic semiconductor WO3 and the polymeric semiconductor CSA PANI. Therefore, the electron–hole recombination was hindered. The transferred holes on the HOMO of CSA PANI reacted with H2O to form hydroxyl radicals (OH), whereas the electrons in the CB of WO3 reacted with oxygen molecules and H+, producing H2O2 to further providing OH radicals. Finally, the OH radicals reacted in turn with MB molecules to produce CO2 and H2O. The CB of WO3 was not conducive to the standard redox potential of O2/O2•− (− 0.046 V vs. NHE); therefore, superoxide anions (O2•−) cannot be obtained. The decomposition percentage of MB further increased in the presence of COOH-MWCNT, which could be the results of more negative energy position of CNT (− 0.1 V vs. NHE) than the standard redox potential of O2/O2•− and more OH radicals were generated.

Fig. 3
figure 3

Schematic presentation of photocatalytic mechanism for MB degradation in the presence of CSA PANI-WO3-CNT

Zhao et al. prepared a novel nontoxic BiVO4-GO-TiO2-PANI (BVGT-PANI) composite with excellent photocatalytic performance in a one-pot hydrothermal reaction [50]. Under visible-light irradiation, PANI-modified BVGTA showed stronger photocatalytic activity for the degradation of MB than BVG (BiVO4-GO) and BVGT (BiVO4-GO-TiO2), indicating a synergistic effect in the hybrid materials of polymer chain, GO flakes, and metal oxides. BVGTA showed the highest kapp rate constant of about 1.06 × 10−2 min−1, which was 1.63 times faster than BVG and 2.94 times faster than BVGT. In addition, in vitro toxicity tests against Bacillus subtilis and Staphylococcus aureus showed that the nanometer photocatalyst was nontoxic. A schematic diagram of dye/phenol degradation using BVGT-PANI is shown in Fig. 4.

Fig. 4
figure 4

Schematic diagram of dye/phenol degradation process

PANI can be combined with semiconductor nanomaterials such as titanium oxide, bismuth-based nanomaterials, ZnO, NiO, WO3, and SnO2 to prepare binary composite photocatalysts. Besides semiconductors, carbon-based materials such as carbon nanotubes (CNTs), graphene oxide (GO), and graphitic carbon nitride (g-C3N4) are suitable options for the hybridization of PANI. To achieve better photocatalysis, metal-oxide NPs, carbon-based materials, and other polymers together with PANI have been used to synthesize ternary or quaternary organo-inorganic photocatalytic nanocomposite materials to degrade MB under visible or UV light. The recently developed PANI-based materials as well as the photocatalytic testing conditions, and photodegradation efficiency for MB are shown in Table 1.

Table 1 Photocatalytic decomposition of MB on polyaniline composites

Degradation of RHB using PANI composites

RhB (C28H31ClN2O3) is widely used as a model molecule in the photodegradation of cationic xanthene class dye.

Steplin Paul Selvin et al. prepared zinc oxide activated charcoal PANI (ZACP) nanocomposite using a simple precipitation method [77]. The as-synthesized photocatalyst exhibited more photocatalytic activity than free ZnO on the degradation of RhB under visible-light irradiation, as the results of photosensitization and electron–hole pair separation by PANI in the composites were obtained. Also, the minimum loss of activity in recycling three times suggested good stability of the composite materials. Moreover, the mineralization of RhB was confirmed by evaluating chemicals and toxicity. A possible reaction mechanism of photocatalytic degradation of RhB in the presence of ZACP under visible-light irradiation is shown in Fig. 5. Compared with the CB and VB positions of ZnO, the LUMO and HOMO levels of PANI are higher. Under visible-light irradiation, PANI transfers excited electrons from the π orbital to the π* orbital. These excited electrons are transferred from the LUMO of PANI to the CB of ZnO and react with water and oxygen to form hydroxyl and superoxide radicals. These hydroxyl and superoxide radicals react with the dye to form less toxic substances. Therefore, the addition of PANI can effectively separate rapid photogenerated electrons, thus improving the photocatalytic activity of ZACP nanocomposites.

Fig. 5
figure 5

Plausible photodegradation mechanism of RhB using ZACP under visible light irradiation

In a pioneering work, Sayed et al. prepared CeO2-PANI and ZrO2-PANI composites in a solvent system of chloroform and 2-butanol [78]. The photocatalytic results showed that the degradation rate of RhB in 60 min of photolysis was 35 and 34% by CeO2-PANI and ZrO2-PANI, respectively. Photosensitization mechanism of PANI–ZrO2/CeO2 composite is shown in Fig. 6. The degradation products of RhB were quantitatively analyzed by LC–MS and GC–MS, and the specific degradation pathways were given. The degradation product at m/z of 415 is the deethyl product of RhB, probably due to the direct photolysis of RhB or the attack of •OH on RhB molecules. The degradation product with an m/z value of 387 can be attributed to the elimination of N substitutions in RhB molecule and the formation of N-deethyl degradation product. Then, it was replaced by •OH digestion to produce a degradation product with an m/z of 122. The degradation product with an m/z of 138 is formed mainly due to the two-step reaction between •OH and RhB molecules. The first step is hydrogen extraction reaction, and the second step is •OH addition reaction. A GC–MS analysis showed that the product with an m/z of 154 was the hydroxylation degradation product. The structure of RhB is further damaged by •OH, and a degradation product with an m/z of 132 is formed, also indicating that the synthetic material makes the molecular degradation of RhB close to mineralization.

Fig. 6
figure 6

Photosensitization mechanism of PANI–ZrO2/CeO2 composite

Feng et al. reported the in-situ polymerization of PANI on the surface of Bi2MoO6 nanosheets to produce PANI/Bi2MoO6 nanocomposites, and the application for the visible-light-driven degradation of RhB [79]. As shown in Fig. 7, the molecular PANI layers covered the surface of flower-like Bi2MoO6 microspheres, which were composed of ultrathin nanosheets (13.8 ± 1.6 nm), and the intrinsic crystallinity of Bi2MoO6 was preserved in the polymerization of PANI. The optimized photodegradation rate for RhB reached up to ~ 100% in 120 min in the presence of PANI0.5/Bi2MoO6. The degradation was consistent with a first-order kinetics with a high apparent rate constant of 0.0335 min−1 and acceptable recycling stability. Mechanism studies showed that both PANI and Bi2MoO6 in the nanocomposites can be excited to produce induced electrons and holes under visible-light irradiation. Bi2MoO6 ultrathin nanosheets covered with PANI molecules provided sufficient active sites for the aggregation of these electrons to capture oxygen molecules and produce superoxide radicals. Both the holes and superoxide radicals can degrade organic pollutants directly and played an important role in improving the photocatalytic performance.

Fig. 7
figure 7

Schematic diagram of the facile synthesis of PANI/Bi2MoO6 nanocomposites for the visible-light-driven photocatalytic degradation of organic pollutants

Recently, Yu et al. studied the removal of organic pollutants such as RhB and phenol in high-salinity wastewater using Ag3PO4/PANI/Cr:SrTiO3 ternary photocatalysts under visible-light irradiation [80]. Under the optimized conditions, the photocatalytic activities of Ag3PO4/PANI/Cr:SrTiO3 composites on RhB and phenol reached 100% within 10 min and 18 min, respectively. The cyclability test showed that the ternary photocatalysts still maintained 92.25% catalytic activity after five cycles. With the increase in SO42− concentration, the activity of Ag3PO4/PANI/Cr:SrTiO3 to RhB remained at a high level, indicating that the catalyst had good tolerance of sulfate. Further analysis indicated the important contribution of photoinduced holes and superoxide radicals to the visible-light photocatalytic activity.

A variety of composite photocatalysts based on PANI doped with semiconductors, carbon-based materials, or other polymers were synthesized for the efficient removal of RhB in wastewater. As shown in Table 2, MoSe2-PANI nanocomposite was synthesized and used as a photocatalyst for the removal of RhB dye [93]. The results indicated that the nanocomposite had the highest degradation rate constant of 1.3205 min−1 under visible-light irradiation when the weight ratio of MoSe2 to PANI is 2:1.

Table 2 Photocatalytic decomposition of RhB on polyaniline composites

Degradation of MO using PANI composites

MO (C14H14N3NaO3S), an anionic azo dye, remains in the environment for a long duration due to its low biodegradability. Thus, MO was always selected as a model dye to assess the photocatalytic ability of composite photocatalysts. Table 3 shows the compositions, photocatalytic testing conditions, and photodegradation efficiency of PANI-based composite photocatalysts on MO degradation in recent years.

Table 3 Photocatalytic decomposition of MO on polyaniline composites

Chen et al. developed a facile two-step route to synthesize one-dimensional (1D) ternary Ag2CO3/Ag/PANI composite nanorods (CNRs) [102]. The structure of the as-prepared composite showed that Ag2CO3 nanorod cores coated with an intermediate layer of AgNPs and a sheath of conducting polymer PANI. By degrading MO under visible-light irradiation, the ternary photocatalyst showed enhanced photochemical current response and photocatalytic activity. The enhanced visible-light-driven photocatalytic activity can be attributed to the intermediate Ag between Ag2CO3 and PANI, which facilitates the separation efficiency of photogenerated carriers, and a Z-scheme charge transfer model was also proposed to understand the charge separation behaviors.

Jung et al. prepared Co0.5Mn0.5Fe2O4-PANI nanofibers through electrostatic spinning, heat treatment, and chemical polymerization, which had a 1D hollow heterostructure with a large surface area [103]. Irradiation of MO under visible-light irradiation showed that the photocatalytic degradation efficiency reached 92% within 120 min, and the kinetic constant was 115 times higher than that of the hollow Co0.5Mn0.5Fe2O4 nanofiber. In addition, the excellent magnetic properties of Co0.5Mn0.5Fe2O4-PANI nanofibers were confirmed by characterizing the spinel structure, which was conducive to the recovery of photocatalyst.

Mousli et al. reported a novel photocatalyst for the mineralization of organic dye pollutants modified with diazonium salts [104], and a TiO2-DPA-PANI nanocomposite was prepared by in-situ oxidation after the diazonium pretreatment of TiO2, which was used for the removal of MO under UV-light irradiation. As shown in Fig. 8, the material was synthesized in two pathways: preparation of PANI in an aqueous solution, and in the presence of pristine and diazonium-modified TiO2. TiO2-DPA-PANI nanocomposite exhibited excellent catalytic performance, and the degradation rate constant was 0.133 min−1, much higher than 0.059 min−1 and 0.085 min−1 of free TiO2 and TiO2-PANI, respectively. Moreover, because the thick coating of PANI protected the following TiO2, TiO2-DPA-PANI could be recycled for five times without losing any photocatalytic activity. However, TiO2-PANI can only be recycled for three times, and bare TiO2 can be reused for one time.

Fig. 8
figure 8

General pathway for TiO2-DPA-PANI and TiO2-PANI nanocomposites

Mousli et al. designed a series of related composites of cotton fabrics (CF) modified with mixed oxides to catalyze the degradation of MO under visible-light irradiation [105]. To be specific, the photocatalyst RuO2-TiO2 was coated on CF using dip-coating method. A layer of PANI was prepared by in-situ polymerization on 4-diphenylamine diazonium salt (DPA) modified RuO2-TiO2 NP coated CF. The CF/RuO2-TiO2/DPA@PANI hybrid photocatalyst exhibited better catalytic performance compared to other catalysts coated on CFs with a photodegradation rate constant of 0.0828 min−1. Owing to the coupling action of the diphenyl amino group from diazonium salts, the PANI film was attached to the surface of RuO2-TiO2 and formed a strong O–N covalent bond with the fabric. The improvement in the catalytic performance was attributed to the strong interfacial interactions between the nanocomposite components and the synergistic effect of charge transfer between different interfaces of CF/RuO2-TiO2/DPA@PANI.

Photocatalytic decomposition of pharmaceuticals

From an environmental viewpoint, even the presence of a low concentration of pharmaceuticals in the wastewater effluent can be hazardous to aquatic organisms and human beings. However, the degradation of drugs is very difficult. The biological methods and physical precipitation methods such as centrifugation and flocculation are common techniques, but they have their own disadvantages [118]. In recent decades, the use of photocatalysts has attracted much attention in the photocatalytic degradation of organic pollutants because of its relatively low cost, environmentally friendly characteristic, sustainable treatment technology, and overcoming the shortcomings of conventional technologies [119]. The photodegradation efficiencies of PANI-based composite photocatalysts on different pharmaceuticals are summarized in Table 4.

Table 4 Photocatalytic decomposition of pharmaceuticals on polyaniline composites

PANI can also form a heterojunction photocatalyst with organic compounds in addition to inorganic semiconductor composite photocatalysts. Wang and Zhu et al. synthesized PANI/perylene diimide with a 3D structure (3D PANI/PDI) using an in-situ growth method, which was applied for the degradation of tetracycline (TC) under visible-light irradiation [120]. 20% PANI/PDI showed excellent catalytic performance and stability, mainly due to the following three aspects: (1) The incorporation of PANI skeleton enhanced the strength of PDI organic hydrogels, thereby increasing the stability of photocatalyst; (2) the 3D structure provided more active sites and electron transport channels; (3) the larger delocalized electron covalent structure and energy matching heterojunction structure formed between PANI and PDI improved the separation efficiency of photogenerated electrons. The proposed degradation mechanism included successive reaction with hydroxylation, dealkylation, aromatization, and ring-opening reactions of TC until complete mineralization under the attack of the main reactive species (H2O2 and h+).

Design and synthesis of green materials capable of green photocatalysis is an important future research direction. Kumar et al. synthesized metal-free carbon-based photocatalysts in a simple way and showed high photocatalytic activity driven by visible light and solar light [121]. The catalysts are based on acidified g-C3N4 (ACN), PANI, reduced GO (RGO), and biochar to form a nanocomposite ACN/PANI/RGO@Biochar (APRB). Among them, biochar acted as an adsorbent, and RGO acted as an electronic medium. Thus, an efficient heterojunction could be produced with the appropriate position of CN, ACN, and PANI, and reduced the recombination of charge carriers to improve the photocatalytic activity. The experimental results show that the degradation rates of ibuprofen (IBN) and 2,4-dichlorophenoxyacetic acid (2,4-D) after 50 min irradiation under xenon lamp were 98.4% and 99.7%, respectively. The degradation pathway was analyzed by LCMS, and 40% and 42% of the total organic carbon was removed in 2 h for IBN and 2,4-D, respectively. Furthermore, the low toxicity of degradation products was determined by the cytotoxicity analysis of human peripheral blood cells.

Photocatalytic decomposition of pesticides

Photocatalytic technology can also be used to deal with environmental pollution caused by the extensive use of pesticides in agriculture, and the recent progress is shown in Table 5.

Table 5 Photocatalytic decomposition of pesticides on polyaniline composites

Photocatalytic degradation of several types of pesticides was achieved using TiO2 NPs modified with PANI (TP nanocomposites) [130]. After simulating solar radiation for 240 min, the degradation rates of thiacloprid, clomazone, quinmerac, and sulcotrione were 13%, 28%, 27%, and 35%, respectively. The cytotoxicity was less than 11% in all the cases, and the photocatalytic degradation efficiency was higher in distilled water than in environmental water. Moreover, pH was the main factor affecting the efficiency of sulcotrione removal. In addition, the addition of H2O2 as an electron acceptor decreased the degradation rate, whereas the addition of KBrO3 increased the degradation rate.

A novel ternary CuO/TiO2/PANI was used as a photocatalyst to degrade 95% of extremely toxic pesticide chlorpyrifos in water within 90 min illumination [131]. The nanocomposite was synthesized using a simple oxidation method and characterized by XRD, EDX, PL, and HR-TEM analyses.

Conclusion

This review shows that PANI is useful to improve the performance of composite photocatalysts for the photocatalytic degradation of hazardous chemicals including dyes, pharmaceuticals, and pesticides, focusing on the roles of PANI. The loading of PANI can substantially improve the photocatalytic activity by enhancing the separation of photogenerated carriers, expanding the light absorption range, increasing the adsorption of reactants, inhibiting photocorrosion, and reducing the formation of large aggregates. This review provides a systematic concept about how PANI can improve the performance of composite photocatalysts [9, 24, 45].