Application of Polydopamine-Based Photocatalysts in Energy and Environmental Systems

Review Open Access

Application of Polydopamine-Based Photocatalysts in Energy and Environmental Systems

Author Information
1
Department of Chemical and Petroleum Engineering, University of Calgary, 2500 University Drive, NW, Calgary, AB T2N1N4, Canada
2
School of Chemistry and Chemical Engineering, Huaiyin Normal University, Huai’an 223300, China
*
Authors to whom correspondence should be addressed.
Views:532
Downloads:75
Photocatalysis: Research and Potential 2024, 1 (3), 10007;  https://doi.org/10.70322/prp.2024.10007

Received: 29 July 2024 Accepted: 04 September 2024 Published: 09 September 2024

Creative Commons

© 2024 The authors. This is an open access article under the Creative Commons Attribution 4.0 International License (https://creativecommons.org/licenses/by/4.0/).

ABSTRACT: Polydopamine (PDA) is also widely sought after in photocatalytic applications due to its fascinating properties such as simple preparation, templating agent, near-infrared absorption, high photothermal conversion efficiency, abundant functional groups, and strong chelating effect of metal ions. This review will present the structural features and synthetic methods of PDA, the advantages of PDA for photocatalytic applications (templating agent effect, light absorption properties, film-forming properties, hydrophilicity, conductivity, etc.), the modulation strategies of PDA for photocatalytic applications, and the use of PDA-based photocatalytic materials for solar-powered water purification (heavy metal adsorption and reduction, catalytic degradation of organic pollutants, and antimicrobial properties), hydrogen production, hydrogen peroxide production, CO2 reduction, and organic conversion. Finally, this review will provide valuable information for the design and development of PDA-based photocatalytic materials.
Keywords: Polydopamine; Photocatalyst; Energy; Environment

Graphical Abstract

1. Introduction

Photocatalytic technology has important applications, especially in the fields of environmental remediation, energy conversion, and synthetic chemistry, by using light energy to excite semiconductor materials to generate electron and hole pairs and initiate chemical reactions [1,2,3,4]. The photocatalytic process involves light absorption, separation and migration of electron-hole pairs, and reaction with molecules at the material surface [5]. Despite the challenges of limited light absorption range, and material stability, high photogenerated carrier complexity, the development of new materials will promote its application in the environmental and energy fields [6,7]. Polydopamine (PDA), as an emerging material, has gained much attention in photocatalysis in recent years. Further, PDA exhibited great potential in the field of photocatalysis as a melanoidin-like material with structural diversity (rich functional groups such as phenolic hydroxyl, amino, and amide groups), biocompatibility, good adhesion (stable coating and film-forming properties), and broad-spectrum absorption (capable of absorbing ultraviolet, visible, and near-infrared light) [8,9]. Currently, PDAs have the potential to be used in the effective degradation of organic pollutants, decomposition of water to produce hydrogen, and reduction of carbon dioxide, etc. To improve its photocatalytic performance, researchers have mainly used the strategy of constructing heterojunctions to modulate the electronic structure of PDA and enhance the separation efficiency of photogenerated carriers [10,11,12,13]. However, PDA also has drawbacks such as photocorrosion and structural instability, especially in strong acid or alkali environments. In addition, the poor dispersion of PDA nanoparticles will affect the modification of the photocatalyst and reduce the catalytic efficiency [14,15]. Despite some challenges, the continuous optimization and design of PDA makes it a promising prospect for applications in the fields of environmental pollution control, clean energy conversion, and green chemical synthesis. Future research will focus on enhancing their efficiency and stability for a wider range of applications. In recent years, the preparation and extensive applications of PDA-based materials on photocatalytic have rapidly developed, as shown by the growing publications over past 10 years (Figure 1a). As impressive as the volume of publications on photocatalysis is its broad interdisciplinary contribution. An analysis of the same dataset of photocatalysis publications by discipline revealed that nearly 75% of the research focused in the fields of water purification and water splitting. In addition to that, significant contributions came from CO2 reduction and H2O2 production (Figure 1b).
Figure 1. (<b>a</b>) Number of publications on PDA-based photocatalysts during the past 10 years (source: Web of Science; date: 30 July. 2024; keywords: polydopamine photocatalytic). (<b>b</b>) Pie chart of “PDA-based photocatalysts” by different research areas using the same data as in application.

2. Molecular Structure and Polymerization Mechanisms of PDA

PDA (Polydopamine) is a synthetic polymer inspired by the adhesive proteins found in mussels. Its structure and polymerization process are of significant interest due to its adhesive properties and biocompatibility. The specific process includes dissolving dopamine hydrochloride in a Tris-HCl buffer solution, usually at a pH of 8.5. The solution is stirred at room temperature, allowing the dopamine to undergo self-oxidation in the presence of oxygen in the air, gradually forming polydopamine [16,17]. In PDA, dopamine functions as a building block, consisting of an amine group (a primary amine attached to an ethyl chain) and a catechol group (an ethyl ring). The multiple chemical reaction sites in PDA, complex redox processes, and generates numerous intermediates, resulting in a diverse synthesis mechanism with no definitive conclusion yet. Lee et al. proposed two polymerization mechanisms for PDA: non-covalent self-assembly and covalent polymerization, which have been widely accepted [8,18]. In the covalent reaction pathway, dopamine is first oxidized to a quinone structure. The quinone structure may then be further oxidized to form 5,6-dihydroxyindole (DHI) (Figure 2a), or undergo intermolecular reactions through catechol-catechol coupling. Finally, cross-linking reactions between these intermediates occur, forming the final PDA structure. In addition to covalent polymerization, the formation of PDA also involves a non-covalent self-assembly pathway [19]. As illustrated in Figure 2b, PDA is considered to be an aggregate of monomers that are primarily cross-linked through strong intrachain and interchain non-covalent forces, encompassing hydrogen bonding, charge transfer and π-π stacking, which is similar to other synthetic or biological supramolecular polymers [8,17].
Figure 2. (<b>a</b>). Dopamine oxidation and subsequent cyclization and rearrangement; (<b>b</b>) The different pathways of PDA formation.

3. Advantages of PDA in the Field of Photocatalysis

3.1. Light Absorption and Excellent Photothermal Conversion Properties PDA, a black polymer generated through the self-assembly of dopamine, exhibits an expansive light absorption spectrum that spans from the ultraviolet to the near-infrared spectrum. This broad absorption capability stems from the intricate conjugated system within its molecular structure, which consists of aromatic rings intertwined with nitrogen-containing moieties. Additionally, the synergistic conjugation of phenolic hydroxyl and amino groups further facilitates electron transition at diminished energy levels [17]. The optical absorption ability could be boosted by changing the bandgap of PDA, for widely using in photothermal conversion and energy harvesting applications. As reported by Zou et al., doping TEMPO moiety into the PDA microstructure by covalently connecting with DHI/IQ oligomers could enhance the light absorption across a broad spectrum by decreasing the energy bandgap and increasing electron delocalization (Figure 3a) [20]. Later, Bai and co-workers proposed that introducing nitrogen-containing heterocycles on PDA significantly lowers the LUMO, promoting electron delocalization in the local regions, which also resulted in the visible light absorption [21]. Moreover, the broad-spectrum light absorption capability of PDA is highly beneficial in photocatalysis. Utilizing PDA in semiconductor modification or constructing heterojunctions can expand the absorption range of semiconductors and promote charge separation. Some researches demonstrated that PDA could be used as photonic sensitizers for enhanced photocatalysis of ZnO in the range of visible light. For instance, Dong et al. fabricated a water purification membrane (PDA/ZnO-NWs/PVDF) by DA polymerization on the ZnO-NWs/PVDF surface [22]. The UV-vis-DRS was shown in Figure 3b, the introduction of PDA has significantly improved the material’s absorbance to light in the range of 396−1750 nm. Similarly, by modifying PDA, the light utilization range can be extended from UV to visible light, which can also be applied to other semiconductors, such as TiO2 and graphitized carbon nitride (g-C3N4) [23,24,25].
Figure 3. (<b>a</b>) The schematic illustration of the bandgap and light absorption ability of PDA and TEMPO-doped PDA [20]. (<b>b</b>) Light absorption ability of PDA/ZnO-NWs/PVDF and ZnO-NWs/PVDF [22]. (<b>c</b>) TEM image of CuSnO-DopaCube and distribution of corresponding elements. (<b>d</b>,<b>e</b>) UV-Vis absorbance spectrum and the evaluation of photothermal properties of CuSnO-DopaCube [26]. (<b>f</b>) Schematic illustration of ZIF-8@PDA/C<sub>3</sub>N<sub>4</sub> coating on the wood hydrogel. (<b>g</b>) IR images of the ZP-C-W in different solar illuminations. (<b>h</b>) The corresponding evaporation rates and solar evaporation efficiencies of different photocatalysts [27].
PDA excels in photothermal conversion due to its remarkable light absorption capacity, effectively absorbing light from the visible to the near-infrared spectrum. It converts absorbed light into heat energy through non-radiative relaxation processes, where electron excitation energy is transformed into molecular vibrational energy, subsequently released as heat. The molecular structure of PDA not only facilitates efficient photothermal conversion but also enables rapid heat transfer via intermolecular interactions. Additionally, its robust thermal stability allows it to retain structural integrity at elevated temperatures, establishing PDA as a potent material for photothermal applications [28]. Gao and co-workers developed a Cu-Sn-PDA photocatalytic Fenton system [26]. As the TEM images shown in Figure 3c, the heterojunction structure fabricated by Cu2O and SnO2 significantly as a core, which was closely wrapped by PDA. The mapping images also confirmed the successful synthesis of the three-layered heterostructure, the element mapping showed a uniform mixture of Cu, Sn, O, and C elements, indicating that the Cu/Sn heterostructure was tightly wrapped by PDA. Firstly, the CuSnO-DopaCube’s optical density (OD) values were detected through UV full spectrum (Figure 3d). The CuSnO-DopaCube (500 μg·mL−1) had an OD value of 1.04 at 808 nm and 0.82 at 980 nm, indicating better photo-thermal conversion efficiency in the NIR region. In addition, the PDA coating in CuSnO-DopaCube significantly improved the photothermal performance (Figure 3e). Based on the excellent light absorption properties across a wide spectrum of light and high photothermal conversion efficiency, PDA as a light absorption layer was well developed in solar evaporator. The solar thermal evaporation process offers a promising technology for producing thermal energy for promoting reaction dynamics and thermodynamics [29]. Lu et al. designed and fabricated a wood/ZIF-8/PDA composite and applied to the solar evaporation and water purification with the photothermal conversion efficiency of 92%. Among them, the black PDA coating on the wood surface used as light absorption layer [30]. Based on the wood/ZIF-8/PDA composite, the group members designed a wood-based hydrogel evaporator modified by C3N4 and PVA/chitosan hydrogel, which accelerate solar desalination and remove pollutants from wastewater (Figure 3f) [27]. As displayed in Figure 3g,h, the evaporator surface temperature increased with the increasing solar intensity. The highest surface temperature and energy conversion efficiency under 6 kW/cm−2 after 60 min could reach up to 82.7 °C and 94.42%, respectively. The excellent solar evaporation performance was contributed to the solar spectra utilization of PDA. In brief, the great light absorption and photothermal conversion properties allow PDA to improve solar energy utilization. Apart from PDA, the light absorption capacities of other common photocatalytic polymers include polyaniline (PANI) [19,31], polypyrrole (PPy) [32] and poly(3,4-ethylenedioxythiophene):poly(styrene sulfonate) (PEDOT:PSS) [33,34] are more limited than that of PDA, and it has fewer active sites [35]. For example, PANI and PPy have good light absorption ability in the UV region, but their performance in the visible region is relatively weak. Some polymers can enhance their photocatalytic properties in the form of doping or composites, but in general additional modifications are required to achieve the effect of PDA. 3.2. Surface Adhesion The catechol and amine groups on the PDA surface make it a good substrate or template for the chelating metal ions [36]. Wang et al. synthesized PDA/CdS nanocomposite for Rhodamine B (RhB) removal from aqueous solutions [37]. First, the PDA nanoparticles were synthesized uniformly with a spherical structure (Figure 4a). Then followed by depositing CdS nanoparticles on the PDA substrate. A possible explanation may be that Cd2+ ions chelated with the PDA substrate, which promoted a heterogeneous nucleation of CdS nanoparticles (Figure 4b). In photocatalytic in removing RhB process, the PDA NPs in nanocomposite with the functional groups, such as −OH and −NH2, exhibited the excellent absorption effect for RhB. In conclusion, it was inferred that there was a significant synergy between adsorption and degradation to efficiently remove RhB from the reactions. Additionally, PDA served as an effective shell material for constructing core-shell structures, encapsulating nanoparticles to enhance their stability. Throughout its formation, dopamine monomers polymerize at various sites to create a cross-linked network, boosting the mechanical stability and adhesive properties of the polymer film. PDA demonstrated superior interfacial activity, capable of self-assembling onto diverse material surfaces to form a uniform coating. Its molecules naturally orient at the liquid-solid interface, creating a dense film that underpins its exceptional adhesive qualities [38]. As shown in Figure 4c, the PDA coating of the core-shell structure was spontaneously formed by immersing the core structure material in an alkaline aqueous solution of dopamine [39]. The sandwich structure was a three-layer concentric heterostructure composed of two or more materials, which exhibited stronger synergistic effects comparing with traditional two-component core-shell structures [10,40,41,42]. Shang et al. synthesized the novel recyclable Fe3O4@PDA/CuS core-shell structure photocatalyst through by in situ reduction method [43]. Among them, the Fe3O4 was used as a magnetic core to provide recovery ability, then the PDA coated on Fe3O4 by self-polymerization. The PDA coating can prevent Fe3O4 from being oxidized, and the abundant -OH and -NH2 can chelate copper ions. The Fe3O4@PDA/CuS exhibited excellent photocatalytic performance and cycle stability for MB degradation. Therefore, the rich active groups on the surface of PDA enable it to form strong adhesion with a variety of material surfaces, which can be used as a bonding layer to immobilize the photocatalyst on various substrates and increase the stability and service life of the photocatalyst. Compared with PDA, PANI, PPy and PEDOT: PSS exhibited weak adhesion, especially in wet environment, but it can still be enhanced through chemical modification or by introducing a hydrophilic polymer adhesive layer to improve its surface adhesion properties [44,45].
Figure 4. (<b>a</b>) The preparation process of PDA NPs/CdS nanocomposite and (<b>b</b>) the RhB degradation process [37]. (<b>c</b>) Preparation process of the PDA@WO<sub>3</sub> nanocomposite [39].
3.3. Film Forming Properties As inspired by mussels’ foot proteins mfp, a strong attachment to almost all substrate types have made PDA one of the most popular coating methods. So far, PDA coatings have been used as mechanical strength coatings combined with porous materials, such as hydrogel [46], foam [47], sponges [48,49] and cotton fabrics [50,51]. On the one hand, this ability to uniformly coat different materials ensures that the light absorption layer is stable and durable, maintaining its efficiency over time. On the other hand, the modification of PDA on the substrates will further enable strong binding of photocatalysts on the substrate, benefiting the stability and recyclability of the photocatalyst. Li et al. reported a novel recyclable cotton fabric/PDA/BiOBr (CFPB) photocatalyst prepared by immobilizing the BiOBr component on PDA modified cotton fabric, which was used for RhB removal and Cr(VI) reduction in visible light irradiation (Figure 5a) [52]. Gong et al. synthesized a novel coated photocatalyst based on the polydopamine modified melamine sponge and then adhering Ag-AgCl/WO3/g-C3N4 (AWC) nanoparticles (Figure 5b) [53]. The AWC/PDA/MS removed 99.9% of trimethoprim (TMP) under visible light in 90 min. The recyclable photocatalyst based on a PDA modified flexible substrate provides a green alternative to powdered photocatalysts with low recyclability. PDA have been used as a component in a number of different filtration membranes, such as PAN and PVDF [54,55]. For instance, Wang et al. designed a 3D layered photocatalytic fiber membrane PAN@PDA/Tb-g-C3N4/ZnIn2S4 (PPTZ), by electrospinning and coating process (Figure 5c). First, the PAN@TCN fiber films (layer 1) were obtained via electrostatic spinning. Then PDA was deposited on the PT membrane via in situ self-polymerization, which was acted as an intermediate linker. Lastly, the ZnIn2S4 was uniformly deposited on the membrane surface, then the three-dimensional layered membrane was labelled as PPTZ. Such a fibrous film had excellent photocatalytic degradation efficiency of antibiotics. In conclusion, PDA forms stable coatings on various substrates with its simple film formation process, excellent adhesion, and surface uniformity, while the film formation of other polymers is more complex, often requiring specific control of conditions, and has advantages in electrical conductivity and mechanical properties [45]. Therefore, PDA is relatively more advantageous in the field of photocatalysis.
Figure 5. (<b>a</b>) The preparation process of photocatalyst fabrics CFPB [52]. (<b>b</b>) The schematic of photocatalytic degradation of TMP with AWC/PDA/MS photocatalyst coatings [53]. (<b>c</b>) The synthetic pathway of the PPTZ film [56].
3.4. Conductivity Polydopamine (PDA), as an organic polymer, is not a conductive material in the traditional sense, but it has certain conductivity. This is mainly due to the aromatic ring structure contained in the PDA molecule forming a certain conjugated system. These conjugated systems allow π electrons to move easily between molecules, thereby giving PDA a certain conductivity [57]. Second, PDA exhibits significant π−π interactions due to its unique molecular structure, which contains aromatic rings. These π−π interactions contribute to several important properties and applications of PDA [38]. For example, the π−π interactions enable PDA to strongly adsorb onto surfaces containing aromatic rings or π-conjugated systems [58]. Additionally, the strong interactions facilitate electron transfer processes, contributing to the overall catalytic performance. Graphitized carbon nitride (g-C3N4) is a 2D metal-free π-conjugated polymer semiconductor photocatalyst with a wide range of applications in photocatalysis [59]. The conjugated π structure shared between PDA and g-C3N4 allows π-π electrons delocalization, which synergistically enhanced carrier transport. Therefore, modifying g-C3N4 with PDA can enhance the activity [15,60]. Furthermore, the bridge-like channel structure formed by the PDA could induce charge transfer through the additional electron transfer channel (PDA), forming the direct Z-scheme charge transfer mechanism [61,62,63]. Wu and colleagues constructed a Z-Scheme Bi@Bi2O2CO3/PDA/g-C3N4 (BPCN) heterojunction [64]. As illustrated in Figure 6a, during the polymerization process, PDA adhered to g-C3N4 through strong covalent and non-covalent interactions, as well as π-π stacking, and provided additional pathways for electron transfer. Thereby enhancing photocatalytic performance through synergistic effects on light absorption and charge carrier separation. The possible photocatalytic mechanism of BPCN heterojunction was proposed in Figure 6b. In the internal electric field formed between Bi2O2CO3 and g-C3N4, electrons rapidly flow from the conduction band of Bi2O2CO3 to the valence band of g-C3N4. They then recombine with the holes on g-C3N4 via PDA acting as an electronic bridge. The conjugated π-π structure between g-C3N4 and PDA facilitated electron delocalization and carrier transfer, balancing the Fermi levels between the semiconductors. Similarly, PDA bridge acted as cross-link of the active phase on carbon nitrates (g-C3N4) to ZnAl layered double hydroxide nanosheets to construct VL-driven Z-type heterojunction catalysts [61].
Figure 6. (<b>a</b>) The preparation process of PDA/g-C<sub>3</sub>N<sub>4</sub> and (<b>b</b>) the possible mechanism of charge transfer of BPCN under visible light irradiation [64]. (<b>c</b>) The sandwich structure of CdS@PDA@SnO<sub>2−x</sub> and (<b>d</b>) the photoelectron transfer of CdS@PDA@SnO<sub>2−x</sub> [65].
In sandwich structure composites, PDA can not only prevent photocorrosion of the nuclear structure, but also served as photosensitizer and electron transfer mediator in the heterojunction [51,66]. Liu and co-workers constructed a CdS@PDA@SnO2−x heterojunction with sandwich structure (Figure 6c) [65]. In this structure, PDA molecules wraped around the surface of CdS, protecting the core from photocorrosion. Simultaneously, the catechol structure of PDA can anchor Cd2+ and Sn4+ ions at specific positions, leading to quick electron transfer, as illustrated in Figure 6d. The as- prepared photocatalyst exhibited an excellent photocatalytic hydrogen (H2) production activity. Qin et al. synthesized a CdS@PDA/NiS ternary composite for H2 production [40]. The strong Cd–O chemical bond between CdS and PDA served as high-speed electron transmission channels, while the Schottky junction accelerated the spatial separation of carriers and improved the photocatalytic hydrogen production performance. After four experiment cycles, the obtained CdS@PDA/NiS maintained good photocatalytic activity and structural integrity. PDA with conjugated systems has excellent electron transport capability, which can reduce the recombination of electron-hole pairs during photocatalysis. Apart from PDA, other common photocatalytic polymers (such as PANI and PPy) also exhibited excellent electrical conductivity and stability [67,68]. This property allows them to conduct electrons efficiently and promote the separation of electrons from holes in the photocatalytic reaction, thereby improving the photocatalytic efficiency.

4. Regulation Strategies of PDA in the Field of Photocatalysis (Mainly Composite Materials with Various Semiconductors)

4.1. Directly Construct Composite Materials with PDA and Other Semiconductors PDA is a kind of polymer material with unique chemical structure and functional groups, which has received wide attention for its potential application in photocatalysis. At present, the photocatalytic performance of PDA can be improved by metal ion doping and constructing composites with other semiconductors (or plasma metal monomers) or surface modification [69,70,71,72,73,74,75]. In the construction of photocatalytic composites, PDAs are often mainly used to form heterojunctions with other semiconductor materials (the main types include: direct to electron conduction type, type II, Z-scheme and S-scheme) to improve the photocatalytic performance. Commonly, g-C3N4 [58,76,77,78,79], metal oxides (TiO2, ZnO, Cu2O, CeO2, etc.) [80,81,82,83,84,85,86], Bi-based materials (BiOCl, Bi2.15WO6 etc.) [87,88,89,90], metal sulfides (CdS, ZnS, ZnCdS, ZnIn2S4, etc.) [91,92,93,94], MOFs [95,96], metal alloys [97,98,99] and carbon dots [100] are used to construct composites directly with PDA. By compositing PDA with these semiconductor materials, it is possible to enhance the photocatalytic efficiency and at the same time extend its application in the visible light range, thus realizing more efficient photocatalytic reactions. Jing and co-workers designed a 1D CdS/PDA heteronanotubes (HNTs) photocatalyst for antibiotic degradation (Figure 7a) [93]. Under the light excitation, the oxygen was reduced into •O2 by photoexcited electrons or continuously converted into •OH to attack the antibiotics. On the other hand, h+ in the valence band can also attack the antibiotics directly and trigger continuous degradation. A Cu2O@PDA nanospheres composite was designed by Zhou and co-workers for methylene orange (MO) dye degradation (Figure 7b) [84]. Cu₂O and PDA exhibit energy level alignment suitable for the construction of a type II heterojunction. Based on this mechanism, the electrons accumulated on the PDA surface and reacted with water and oxygen to produce ROS (including •OH and •O2), while the h+ focused on the valence band (VB) of Cu2O. Under the oxidation of ROS and h+, MO was degraded into CO2, H2O, etc. Shi et al. constructed a polydopamine-coated tungsten oxide (PDA@WO3) nanomaterial with special yolk-shell structures (Figure 7c) [39]. The WO3 and PDA can form a Z-Scheme transport mechanism in photocatalytic reactions and lead to a lifetime extension of the carriers. Specifically, under light illumination, the PDA in PDA@WO3 induces a π–π transition, where excited-state electrons are transferred from the highest occupied molecular orbital (HOMO) to the lowest unoccupied molecular orbital (LUMO) of PDA, leaving behind a large number of holes (h+). The electrons generated from WO3 are transferred to the HOMO orbital of PDA through interfacial transport under the effect of built-in electric field, and these electrons are combined with holes in HOMO. The electrons in the LUMO orbital of PDA could reduce O2 to •O2, which react directly with the MB. Zhang et al. synthesized a novel S-scheme heterojunction photocatalyst comprised of ultrathin g-C3N4 (U-CN) and PDA [12] (Figure 7d). Under light irradiation, electrons in U-CN were excited from their VB to conduction band (CB) and eventually recombined with the holes in the HOMO of PDA driven by the built-in electric field, manifesting the S-scheme mechanism.
Figure 7. (<b>a</b>) Schematic diagram of the reaction mechanism during the catalytic decomposition of antibiotics by CdS/PDA [93]. (<b>b</b>) Schematic diagram of the possible photodegradation mechanism of Cu<sub>2</sub>O@PDA via visible light irradiation [84]. (<b>c</b>) Schematic diagram of the photodegradation mechanism of MB on PDA@WO<sub>3</sub> sample [39]. (<b>d</b>) Schematic diagram of the S-type photocatalytic mechanism of C<sub>3</sub>N<sub>4</sub>/PDA [12].
4.2. Composite Materials Constructed with PDA as an Electronic Bridge Structure PDA as an electronic bridge plays multiple advantages in photocatalytic composites, including improving the separation efficiency of photogenerated electrons and holes and forming multifunctional composite structures with a variety of semiconductors to realize diversified photocatalytic reactions, which shows its potential for application in building highly efficient and stable photocatalytic materials [101,102,103,104]. Zhang et al. developed a ZnO@PDA-Au photocatalytic system for organic pollutant removal (Figure 8a) [105]. When Au contacts ZnO, a Schottky barrier forms due to work function differences, and ZnO under UV-visible light excitation generates electron-hole pairs. Electrons transfer from ZnO to Au and PDA, reducing dissolved oxygen to create •O2, while holes oxidize water to form •OH. These radicals contribute to organic molecule degradation. The PDA and local surface plasmon resonance (LSPR) effect of Au enhanced visible light absorption. The Schottky barrier served as an electron trap utilizing charge separation efficiency and significantly and boosting the catalytic performance of ZnO@PDA-Au. Figure 8b illustrates that AgNPs/PDA/g-C3N4 enhances the charge step transfer route via the built-in electric field, with PDA capturing molecular oxygen to generate free radicals and improve pollutant adsorption. The LSPR effect of AgNPs was triggered at a wavelength of 390–420 nm, producing hot electrons and reducing the electron-hole recombination rate, which increased the generation rate of charge carriers. This synergy between PDA and AgNPs significantly boosted the photocatalytic MB degradation activity of g-C3N4 [106]. Figure 8c illustrated the photocatalytic hydrogen production mechanism developed by Meng et al. using a g-C3N4 nanotube@PDA/NiCo-LDH composite. Under visible light, g-C3N4 and NiCo-LDH generated electron-hole pairs, with PDA acting as a photosensitizer and interface medium to boost light absorption and facilitate charge transport. The electrons from g-C3N4 were drawn to the higher energy levels of NiCo-LDH, while the holes migrated to g-C3N4. Pt nanoparticles catalyzed the reaction of electrons on NiCo-LDH with H+ ions in water to produce hydrogen, whereas holes on g-C3N4 were quenched by TEOA. The combined action of the type II heterojunction and PDA significantly enhances carrier separation and transport, greatly improving the efficiency of hydrogen production [107]. Yang et al. developed a type II heterostructure by combining carbon nitride nanosheets (CNNS) with CdS, which under light exposure, facilitated the production of photogenerated electron-hole pairs (Figure 8d) [15]. Electrons were transferred from the CB of CNNS to the CB of CdS, while holes migrated from the VB of CdS to the VB of CNNS. The electrons in CdS reacted with dissolved oxygen to produce superoxide radicals (•O2), and the holes in CNNS directly participated in the degradation of RhB. The inclusion of polydopamine (PDA) in this setup enhanced charge migration, minimized electron-hole recombination, and thus significantly bolstered the photocatalytic efficiency. A novel g-C3N4@PDA/BiOBr Z-scheme photocatalytic system was constructed by Guo et al. (Figure 8e) [62]. Here, electrons move from the CB of BiOBr to the VB of g-C3N4 through PDA, leading to effective charge separation. This movement resulted in a higher concentration of electrons in the CB of g-C3N4 and holes in the VB of BiOBr. The holes in BiOBr’s VB are potent enough to directly oxidize sulfamethoxazole (SMX), while the accumulated electrons in g-C3N4’s CB reduce O2 to form superoxide radicals (•O2). Concurrently, the VB holes in BiOBr can also oxidize H2O/OH− to generate hydroxyl radicals (•OH). These radicals collaborate to decompose SMX into CO2 and H2O. PDA serves a crucial role as an electron mediator, enhancing carrier separation and thereby boosting the photocatalytic redox capabilities of the composite material.
Figure 8. (<b>a</b>) Mechanism diagram of ZnO@PDA-Au under UV-visible light irradiation [105]. (<b>b</b>) Schematic diagram of charge migration of AgNPs/PDA/g-C<sub>3</sub>N<sub>4</sub> [106]. (<b>c</b>) Possible mechanism of photocatalytic hydrogen production by g-C<sub>3</sub>N<sub>4</sub>@pDA/NiCo-LDH [107]. (<b>d</b>) Possible mechanism of photocatalytic removal of RhB by CNNS/PDA/CdS [15]. (<b>e</b>) Possible mechanism of photodegradation of SMX by Z-scheme heterojunction g-C<sub>3</sub>N<sub>4</sub>@PDA/BiOBr under visible light irradiation [62].
4.3. Comprehensive Photocatalytic System Constructed by Combining PDA, Semiconductor and Membrane Materials (Other Carrier Materials) The photocatalytic system fabricated by combining PDA, semiconductor and membrane materials utilizes the excellent film-forming and surface modification capabilities of PDA to form a uniform protective layer on the surface of the semiconductor material, which enhances stability and durability [108,109,110,111,112,113,114]. The hydrophilicity and anti-pollution properties of the PDA-modified membranes enhance the efficiency of wastewater treatment and simplify the cleaning and regeneration process. The biocompatibility and environmentally friendly synthesis process of PDA support the construction of sustainable photosynthetic systems. Membrane materials provide physical protection for the photocatalysts and prolong the service life, while PDA enhances the surface activity and reaction sites to accelerate the photocatalytic reaction rate. The composite system of PDA, semiconductor, and membrane materials (PES, PVDF, cellulose membrane, PVA, cellulose acetate membrane, PET, poly (arylene ether nitrile) (PEN), polyacrylonitrile (PAN), polytetrafluoroethylene (PTFE), etc.) is simple to prepare, easy to produce on a large scale, and multifunctional. Overall, the photocatalytic system constructed by PDA, semiconductor and membrane materials significantly improves the photocatalytic efficiency by improving the photocatalytic activity, stability, enhanced the surface properties and multifunctionality, and shown a wide range of application potential [115,116,117,118,119,120,121]. Shi et al. developed a poly (vinylidene fluoride) (PVDF) composite membrane enhanced with polydopamine and titanium dioxide nanoparticles, designed for efficient and eco-friendly treatment of organic wastewater (Figure 9a) [55]. After UV irradiation for 60 min, this membrane exhibited a water flux three times higher than that of the standard PVDF membrane. It showed remarkable pollutant retention, removing approximately 99.0% of Congo red and Coomassie Brilliant Blue G250. Post-treatment, the water flux of the membrane recovered over 87.2%. Additionally, the composite membrane demonstrated excellent retention and antifouling capabilities in oil-water emulsion filtration, underscoring its versatility in various wastewater treatment and separation applications. Zheng et al. developed a novel photocatalytic nanofiltration (NF) membrane by applying interfacial polymerization on PVDF membranes to introduce a polyamide layer and incorporating polydopamine with BaTiO3/Ti3C2Tx nanoparticles (Figure 9b) [110]. This combination achieved both physical separation and photocatalytic degradation functionalities. The membrane demonstrated high water permeability (34.0 L·m−2·h−1·bar−1) and excellent Na2SO4 retention (94.7%). Under irradiation from a 300 W Xe lamp for 180 min, it showed substantial photocatalytic degradation efficiencies for various azo dyes: 95.3% for methyl orange (MO), 92.7% for alizarin yellow GG, and 83.5% for methyl red. This research introduces an innovative approach for creating bifunctional photocatalytic nanofiltration membranes with effective degradation capabilities. In addition to the membrane materials mentioned above, more types of carrier materials (polyurethane sponge, cellulose fibers, nickel foam, hydrogel, woven cotton fabric, wood, melamine foam, filter paper, etc.) can replace membrane materials [122,123,124,125,126,127,128], further promoting the practical application value of PDA-based photocatalytic materials.
Figure 9. (<b>a</b>) PVDF/PDA-TiO<sub>2</sub> composite membrane for filtration of pollutants photodegradation [55]. (<b>b</b>) Schematic diagram of the BaTiO<sub>3</sub>/Ti<sub>3</sub>C<sub>2</sub>Tx@PDA membrane manufacturing process and the photodegradation mechanism of the BaTiO<sub>3</sub>/Ti<sub>3</sub>C<sub>2</sub>Tx@PDA membrane [110].

5. Application of PDA in the Field of Photocatalysis

As a typical organic polymer material, PDA itself has very poor photocatalytic activity. Nevertheless, PDA-based photocatalysts have attracted much attentions owing to their excellent properties, such as light harvesting abilities, interfacial properties and electron transfer capabilities. Hence, PDA can synergize with semiconductor photocatalysts to enhance photocatalytic performance in various application fields. In this part, we give a brief overview of their application areas in water purification, water splitting, carbon dioxide reduction and other photocatalytic reactions. 5.1. Water Purification In recent years, the shortage of water resources has become increasingly serious, and water pollution has exacerbated the shortage problem. The main pollutants include organic pollutants, heavy metal ions, microbes and bacteria. Photocatalytic technology has become an important, economical and effective method to solve environmental pollution problems. 5.1.1. Organic Pollutants Degradation PDA, with rich functional groups and unique structure, exhibits significant absorptivity for organic pollutants, making it a valuable material in environmental remediation. For example, the aromatic rings of PDA molecular structure can engage in π-π interactions with the aromatic rings of organic pollutants molecules, improving adsorption capacity. This interaction is particularly effective for adsorbing organic pollutants with aromatic ring structures, such as azo dyes and aromatic compounds [129,130,131]. Besides, electrostatic attraction or repulsion between the surface charge of the catalyst and the pollutant molecules affects adsorption performance. For example, a positively charged catalyst surface is more likely to adsorb negatively charged pollutant molecules. The surface charges of PDA are controllable by adjusting the pH medium [63]. Recently, an Ag/PDA@Ag2S smart pH-heterojunction photocatalyst with multi interface band structure modulation was designed and proposed through a simple and rapid in situ polymerization method (Figure 10a) [132]. The photocatalytic process was carried out with Ag/PDA@Ag2S for the four cationic and anionic dyes and, at the same time, the PDA-free photocatalyst was used as a reference. As shown in Figure 10b,c, it was observed that the Ag/PDA@Ag2S s could successfully degrade cationic dyes with remarkably higher efficiency than anionic dyes. Specifically, methylene blue (MB) was degraded by 80% after 10 min, while MO was degraded by about 4.0%, which was proven visually by the color change as well. In addition to improving charge transport and separation, the ultra-thin layer shell introduced by PDA increases the photostability of Ag2S/Ag photocatalysts. The PDA has the unique capability to switch on and off Ag/Ag2S hybrid photocatalysts depending on the pH or ionic strength of the reaction medium, resulting in selective photodegradation of organic dyes. Recently, Daniel et al. explored the influence of PDA coating on the electronic effects of Au particles and tested the photocatalytic performance with Rhodamine 6G (Rh6G) [133]. As shown in Figure 10d, the Au/PDA hybrid nanoplatform was synthesized by an AuNR core followed by the further self-polymerization of PDA. The thickness of the PDA was influenced by the concentration of dopamine hydrochloride. Figure 10e illustrated the RhG6 degradation efficiency under different photocatalysts, indicating the higher photocatalytic efficiency of Au/PDA composites than that of Au and PDA particles alone and increased with the with the PDA thickness, where the degradation of AuNRs/PDA3 could reach up to 50% in 60 min. It was analyzed that the Au/PDA nanoparticles could generate reactive oxygen species (ROS) by the absorbed heat both in the visible and NIR regime (Figure 10f). When combined with the Au, the heat pump into the PDA, which could strengthen and drive the production of ROS of •OH and •O2. In addition, due to the photothermal effect and the release of ROS, the PDA modified plasmonic metal photocatalyst has excellent antibacterial properties in waste water treatment.
Figure 10. (<b>a</b>) The Ag/PDA@Ag<sub>2</sub>S smart heterojunction photocatalyst for dyes photocatalytic degradation. (<b>b</b>) Photocatalytic degradation efficiency of MB and MO at pH = 10. (<b>c</b>) Plots of photodegradation of MB and MO mixture against Ag/PDA@Ag<sub>2</sub>S smart photocatalyst at pH = 2 [132]. (<b>d</b>) PDA shell thickness (with SD) versus log<sup>2</sup>[DA]. (<b>e</b>) Percentage of Rh6G concentration depleted versus time of different photocatalysis. (<b>f</b>) Band model of the general production of ROS during VIS irradiation [133].
5.1.2. Heavy Metal Ion Reduction (Cr(VI)) PDA has a large number of hydroxyl and amino functional groups, which can chelate with metal ions, effectively adsorbing and removing heavy metal ions from water, thereby improving water treatment efficiency. The reduction of Cr(VI) ions typically involves converting to Cr(III), which is less toxic and more stable in the environment [134]. For example, Lu et al. synthesized PDA/AgNPs through the self-polymerization of PDA and the in-situ reduction of Ag+ to AgNPs via PDA [135]. The reaction process of Cr(VI) removal was shown in Figure 11a, the large amino groups of PDA adsorbed Cr(VI) into the surface of photocatalyst, then improving the degradation efficiency under illumination. As demonstrated in Figure 11b, the Cr(VI) removal under dark conditions and Cr(VI) concentration after illumination indicated that PDA absorption was also beneficial to Cr(VI) removal. For the existence of plentiful amino and catechol groups, the homogeneous composites combined the PDA and semiconductors through coordination and DA self-polymerization. Additionally, PDA can absorb a broad range of wavelengths, increasing light capture capabilities, thereby enhancing the activity of photocatalysts under visible light. Li and colleagues reported a part of the TiO2-like structure Ti-PDA NPs composite for Cr(VI) reduction (Figure 11c) [69]. As shown in Figure 11d, the photoreduction rate of Cr(VI) by Ti-PDA 3 was near 99% in 30 min, which was contributed to the sensitization effect of PDA, which kept at 95.02% after three cycles. Li and co-workers reported a 0D/1D heterostructure of PDA modified hexagonal carbon nitride tubes (HCN) for the photocatalytic Cr(VI) reduction [77]. 0D PDA nanoparticles were widely distributed inside and outside the open channels of 1D HCN to promote the sufficiently photoreduction of Cr(VI) on dual surfaces, and can achieve nearly 100% reduction of Cr(VI) after 75 min. This highly efficient 0D/1D heterostructure may serve as a guide for photocatalysts of other conjugated polymers, thereby enhancing their potential for environmental remediation and energy applications.
Figure 11. (<b>a</b>) The photocatalytic process of Cr(VI) through PDA/AgNPs composites. (<b>b</b>) Removal effiency of Cr(VI) under illumination or dark [135]. (<b>c</b>) Synthesis of Ti-PDA NPs and the process of Cr(VI) absorption and photoreduction. (<b>d</b>) The photoreduction curves of Cr(VI) by different photocatalyst [69]. (<b>e</b>) The possible mechanistic for tetracycline (TC) degradation and antibacterial by ZnO@PDA/Cu-POMs [136].
5.1.3. Sterilization As be mentioned before, PDA exhibited high antibacterial efficiency under near infrared irradiation owing to the photothermal effect and photodynamic-induced ROS [42,137]. Recently, the excellent physiochemical properties and effectiveness against pathogenic microorganisms of metallic oxides have enabled them to be used for preparing photocatalyst composites for environmental remediation [138,139,140]. Cui and co-workers synthesized a ZnO@PDA/Cu-POMs, with high-efficiency photocatalytic degradation of TC and extraordinary antibacterial activities [136]. As shown in Figure 11e, the ZnO@PDA/Cu-POMs heterostructure conformed to the Z-type charge transfer mechanism. Then under the attack of the reactive species, 90.75% of TC was effectively degraded. Additionally, the ZnO@PDA/Cu-POMs nanocomposite exhibited excellent antibacterial activity against 4 kinds of Gram-positive/negative bacteria, and the antibacterial rates were higher than 99% for all bacterial strains. The possible mechanistic of antibacterial action was that the ROS of ZnO@PDA/Cu-POMs could cause damage to the membrane and DNA of bacteria, then collectively induced bacterial cell death. PDA is rich in functional groups such as hydroxyl, amine, and catechol groups, which can interact with various photocatalysts like TiO2, ZnO, and g-C3N4 to form composite materials. In water purification, PDA synergistically works with photocatalysts to significantly enhance the degradation efficiency of organic pollutants such as MO and MB. The presence of PDA facilitates the generation of ROS (e.g., •OH, •O2), which rapidly oxidize and decompose contaminants in water. PDA not only degrades organic pollutants but also chelate with metal ions, effectively adsorbing and removing heavy metal ions from water, thereby im-proving water treatment efficiency. Furthermore, PDA exhibits outstanding antibacterial properties, effectively inhibiting the growth of pathogenic microorganisms in water, thus achieving multifunctional water purification. In short, under synergy with the semiconductor photocatalyst (such as metal sulfide or metal oxide), PDA shows excellent photocatalytic effects in the fields of water purification, including organic pollutants degradation, Cr(VI) reduction and sterilization. 5.2. Water Splitting Hydrogen, as an efficient and pollution-free energy source, can be widely used in industry, transportation, and energy storage. By producing hydrogen through photocatalytic water splitting, we can harness the inexhaustible resource of solar energy, reduce dependence on fossil fuels, decrease greenhouse gas emissions, and help address global climate change and energy crises [141,142,143]. As one of the most widely used water splitting materials, g-C3N4 has attracted considerable attention for its adjustable bandgaps and excellent physiochemical properties. In addition, the hollow g-C3N4 nanospheres (HCNS) stand out in the research on various novel structures of g-C3N4, owing to their unique hollow structure and abundant active sites. Wang et al. reported a strategy comprising the g-C3N4 and PDA to a hollow flower-like nanosphere (g-C3N4@PDA), where monodisperse polystyrene microsphere using as template [144]. The photocatalyst construction process was shown in Figure 12a. the PDA@PS nanospheres was formed via DA self-polymerizes. Specifically, the g-C3N4 nanosheets were then introduced onto the outer surface of the PDA@PS nanospheres via interfacial self-assembly, then removing the PS template by in CHCl3. A series of photocatalysts were used to evaluate the H2 production rates (HER) as shown in Figure 12b. A significant improvement in the HER was observed after exposure to visible light irradiated with g-C3N4@PDA nanospheres, which was contributed by the conjugated system and the hollow nanostructure. In addition, the HER matched well with its light absorption curve (Figure 12c). When the incident light wavelength was extended to 700 nm, the g-C3N4@PDA still showed good photocatalytic activity (HER = 5 µmol h−1). The mechanism of optical activity was illustrated in the Figure 12d,e, based on the synergistic effect, PDA continuously supplied electrons to the CB of g-C3N4 while accepting holes from g-C3N4, thus reducing electron-hole recombination. Additionally, the hollow nanospheres with multifaceted open surface areas can facilitate surface-dependent reactions by effectively shuttling electrons and chemicals. These factors collaboratively enhance the photocatalytic performance of the nanomaterials. Similar, Yang et al. fabricated a PDA/defective ultrathin mesoporous g-C3N4 (PDA/DCN) Z-scheme organic for the H2 evolution and organic degradation (Figure 12f) [145]. It was demonstrated that N-vacancy and π-π interactions provide strong connection and Z-scheme fast charge transfer.
Figure 12. (<b>a</b>) The preparation process of g-C<sub>3</sub>N<sub>4</sub>@PDA nanosphere. (<b>b</b>) The H<sub>2</sub> evolution efficiency of different samples in light irradiation. (<b>c</b>) Wavelength dependence of H<sub>2</sub> evolution rate on g-C<sub>3</sub>N<sub>4</sub>@PDA photocatalyst. (<b>d</b>) Band positions of PDA and g-C<sub>3</sub>N<sub>4</sub> and (<b>e</b>) electron transfer in the g-C<sub>3</sub>N<sub>4</sub>@PDA system (<sup>&#x2217;</sup> means that the electron transuition to excited state) [144]. (<b>f</b>) Scheme diagram of PDA/DCN photocatalytic H<sub>2</sub> production and organic contaminant degradation [145]. (<b>g</b>) The photocatalytic H<sub>2</sub> production effiency under different PDA thickness. (<b>h</b>) Schematic representation of the ZnS/PDA heterojunction [94].
In addition to g-C3N4, transition metal sulfides, such as such as ZnS, are widely applied as an efficient photocatalyst in photocatalytic H2 evolution, due to their high reduction potential, as well as the rapid generation of photocarriers abilities. However, the application of ZnS is hindered by photocorrosion and instability due to oxidation of sulfide ions by photogenerated holes. PDA has good adhesion and interface modification capabilities, allowing it to form uniform coatings on various material surfaces, thereby improving the surface activity and stability of the photocatalyst. Recently, nanometer-controlled PDA coatings have been shown to enhance photocatalytic activation wavelength, rather than hinder them. Kim and co-workers reported on a strongly coupled heterojunction photocatalyst comprising by a ZnS nanorod and ultrathin PDA (ZnS/PDA). Figure 12g,h shown that the highest H2 production rate was 2162.5 μmol h−1 g−1 with PDA layer of 1.2 nm, and remained 78.7% of initial photoactivity after three cycling. In addition, since the effective carrier transmission length was limited to less than 2 nm, the excited carriers cannot be transmitted to the surface if the PDA layer was thicker than the transmission length. This research results may be applicable to other sulfonium catalyst and sensitive nanomaterials and light catalyst, such as selenide, boride, telluride, etc. PDA has shown significant potential in the application of photocatalytic water splitting, primarily due to its unique physicochemical properties and its synergistic effects with photocatalysts. The reductive properties and excellent conductivity of PDA facilitate the transfer of photogenerated electrons and reduce the recombination of electron-hole pairs, thereby increasing the rate of hydrogen production during water splitting. Moreover, the PDA coating can be used as a protective layer to prevent photocorrosion or chemical corrosion of the catalyst, thereby improving the photocatalytic stability. 5.3. CO2 Reduction The burning of fossil fuels and industrial activities have increased carbon dioxide (CO2) concentration in the atmosphere, which in turn has caused climate change and a series of environmental problems. Photocatalytic CO2 reduction is considered a promising method for reducing carbon dioxide emissions and producing renewable fuels. It utilizes light energy to convert CO2 into valuable chemical products, such as methanol, methane or carbon monoxide [2,146]. This process typically involves the use of a semiconductor photocatalyst, which absorbs light and generates electron-hole pairs. These carriers then participate in the reduction of CO2 and the oxidation of a sacrificial agent or water. It is essential to use photocatalysts that can separate charge efficiently and adsorb CO2 [147,148]. Research has shown that the amine-functionalized metal oxides can effectively capture CO2 by forming strong chemical adsorptive species, which not only promoted the activation of CO2 but also enhanced the CO2 reduction in photocatalytic. A great deal of amino groups in PDA, can easily be anchored to most surfaces, and enhance the CO2 adsorption and activation capacity. For example, it was demonstrated that the C3N4 nanospheres modified with PDA had significantly enhanced CO2 adsorption capacity and high photocatalytic efficiency [149]. It has been reported that dopamine modification on TiO2 nanoparticles can not only improve the charge separation efficiency, but also increase the reduction power of electrons by 100 mV [150]. In addition, compared with TiO2 bulks, hierarchical structure of TiO2 has a shorter distance charge transfer, thereby increasing the charge transfer, the duty carrier recombination. Meng et al. reported a series of PDA-modified TiO2 hollow spheres (TiO2@PDA) for photocatalytic CO2 reduction (Figure 13a) [151]. As shown in Figure 13b, when photocatalytic CO2 reduction reactions were performed using photocatalysts, CH4 and CH3OH were detected as products. For pristine TiO2, the generating rates of CH4 and CH3OH were 0.30 and 0.63 μmol h−1 g−1. With the deposition of 0.05% PDA, the CH4 yield increased to 1.14 μmol h−1 g−1, which was nearly four times as many as the generation rate of sample T. The higher yields and outstanding methane selectivity were contributed to the enhanced CO2 adsorption capacity as well as efficient separation and transfer of charge carriers induced between TiO2 and PDA (Figure 13c).
Figure 13. (<b>a</b>) Schematic illustration of TiO<sub>2</sub>@PDA composites. (<b>b</b>) Photocatalytic yields of pure TiO<sub>2</sub> and TiO<sub>2</sub>@PDA photocatalysts. (<b>c</b>) The possible mechanism of S-scheme charge transfer of TiO<sub>2</sub>@PDA [151]. (<b>d</b>) The synthesis of PDA/ZnO/Co<sub>3</sub>O<sub>4</sub> photocatalysts. (<b>e</b>,<b>f</b>) The HRTEM images of ZnO/Co<sub>3</sub>O<sub>4</sub>and PDA<sub>15</sub>/ZnO/Co<sub>3</sub>O<sub>4</sub>. (<b>g</b>) The CO and H<sub>2</sub> evolution rate of different photocatalysts under UV-vis illumination (the arrow points to the CO selectivity). (<b>h</b>) The recycling performance of on PDA<sub>15</sub>/ZnO/Co<sub>3</sub>O<sub>4</sub>. (<b>i</b>) The mechanism diagram for photocatalytic CO<sub>2</sub> reduction mechanism diagram via PDA<sub>15</sub>/ZnO/Co<sub>3</sub>O<sub>4</sub> [152].
Band position and intrinsic properties make the metal oxide can combine well matching the heterojunction formation. Additionally, the MOF-derived metal oxide also inherits the porous structure and large specific surface area, making it a promising candidate material for efficient photocatalytic CO2 reduction [66]. For instance, a hollow ZnO/Co3O4 photocatalyst was constructed through pyrolyzing bimetallic ZnCo-ZIFs and modifying them with PDA (Figure 13d–f) [152]. Figure 13g shown shows the production of CO on PDAm/ZnO/Co3O4, with the amount of PDA increasing from 3 mg to 120 mg. Among them, the PDA15/ZnO/Co3O4 (the contentPDA = 15 mg) showed the highest CO production in the photocatalytic CO2 reduction, with a rate of 537.5 μmol/g/h and a selectivity of 97.7%. The enhanced CO2 adsorption/activation capability was attributed to the PDA modification. Further increasing PDA to 30 mg and 120 mg resulted in a decrease in CO production, which might be attributed to the reduced light absorption by excess covering of semiconductor ZnO/Co3O4 surface. As a result of recycling experiments, the high photocatalytic activity and CO selectivity were maintained after five cycles (Figure 13h). PDA shows great potential in photocatalytic CO2 reduction through its synergistic effects with photocatalysts, particularly in enhancing photocatalytic efficiency, improving stability, and controlling product selectivity. For example, PDA combined with the amine-functionalized metal oxides can effectively capture CO2. By optimizing the thickness and surface chemistry of PDA, it is possible to improve the selectivity of specific products. 5.4. Photocatalytic H2O2 Production As an environmentally friendly oxidant and ideal carrier of energy, hydrogen peroxide (H2O2) has been widely used chemistries, such as chemical production, healthcare, and environmental treatment for many years [153]. As the industrial production of one of the primary methods of hydrogen peroxide, anthraquinone method is through the cycle of anthraquinone compounds REDOX reaction, high concentrations of hydrogen peroxide production. Nevertheless, the process of anthraquinone depends on precious metals and organic solvents, and consumes a lot of energy [154]. Photocatalytic H2O2 strategy is based on the most abundant water (H2O) and oxygen (O2) in nature, which uses a photocatalyst to break down water or other compounds to form hydrogen peroxide under light, is considered an emerging research direction [155,156]. Photocatalytic H2O2 production in the main pathways including direct photocatalytic decomposition of water oxidation reaction (WOR) and oxygen reduction reaction (ORR). Among them, ORR is considered light catalytic production of H2O2 main ways, where oxygen is acting as an electron acceptor. Therefore, improving the photocatalyst’s ability to effectively adsorb and desorb oxygen can prevent electron-hole recombination and enhance the efficiency of photocatalytic H2O2 generation. As reported that the reversible conversion between catechol and o-ben zoquinone in PDA during the photocatalytic H2O2 production, which was similarly to the transformation of anthraquinone. Wei et al. designed the CdS-PDA hybrid catalyst via coating PDA on CdS for H2O2 production (Figure 14a) [157]. The adsorption energy of O2 on CdS (002) and CdS (002)-PDA models was estimated by performing density functional theory (DFT) calculations (Figure 14b), which was consistent with the O2TPD result in that O2 was more strongly adsorbed on CdS (002)-PDA (E(O2)ads = −2.41 eV) than that on CdS (002) (E(O2)ads = −1.25 eV). The H2O2 selectivity of CdS -PDA was significantly enhanced from 30% to 82%, achieving a 13.7-fold catalytic yield over the CdS (Figure 14c). In addition, the Kf and Kd values (corresponding to the formation and decomposition rate constant of H2O2) in Figure 14d indicated that the PDA coating could not only improve the ability of O2 adsorption, but also inhibit the decomposition of H2O2, and enhance the stability of CdS light corrosion. Similarly, Huang and co-workers constructed a PDA-loaded ZnIn2S4 (ZIS/PDA) for H2O2 production [91]. The S-scheme photocatalyst ZIS/PDA exhibited outstanding H2O2 production performance under visible light in pure water, where the highest H2O2 yield (1747 μmol·g–1·h–1), which was 39 times the yield of pure ZnIn2S4 (Figure 14e). The transition intermediates and the corresponding free energies were shown in Figure 14f, the ΔG of OOH generated by ZIS/PDA0.1 was much lower than that of ZIS, suggesting that the ZIS/PDA0.1 could reduce the energy barrier and improve the overall reaction rate. The possible mechanism of photocatalytic H2O2 by the ZIS/PDA0.1 was illustrated in Figure 14g. Under the influence of the internal electric field (IEF), the constructed ZIS/PDA0.1 S scheme heterojunction rapidly separated and transfered photogenerated electrons and holes. The photogenerated electrons, which accumulated on the CB of PDA, are then consumed by oxygen and protons. The PDA coating enhanced the selectivity of H2O2, increased the overall reaction rate, and significantly inhibited the decomposition of H2O2, thereby promoting the accumulation of H2O2.
Figure 14. (<b>a</b>) The synthesis process of CdS–PDA. (<b>b</b>) Computational models for O<sub>2</sub> adsorption on photocatalysts. (<b>c</b>) Long-term photocatalytic H<sub>2</sub>O<sub>2</sub> production and (<b>d</b>) K<sub>f</sub> and K<sub>d</sub> values of photocatalysts [157]. (<b>e</b>) H<sub>2</sub>O<sub>2</sub> production rate of ZnIn<sub>2</sub>S<sub>4</sub> and ZIS/PDAx. (<b>f</b>) Free energy of the intermediate state in the H<sub>2</sub>O<sub>2</sub> production of ZIS and ZIS/PDA<sub>0.1</sub>. (<b>g</b>) Mechanism of photocatalytic H<sub>2</sub>O<sub>2</sub> by the sustainable transformation of ZIS/PDA<sub>0.1</sub> [91]. (<b>h</b>) The possible reaction mechanism of H<sub>2</sub>O<sub>2</sub> production under CN-PDA [158].
In photocatalytic H2O2 production process, Superoxide free radical (•O2) is generated photocatalytic oxygen reduction of important intermediates, has the strong oxidizing properties, could undermine the stability of the catalyst. Therefore, avoiding the accumulation of •O2 resulting in its rapid conversion to H2O2 is crucial to improve catalyst stability and H2O2 yield. Deng et al. developed a heterojunction of PDA and g-C3N4 (CN-PDA) for H2O2 synthesis [158]. The possible mechanism of CN-PDA photocatalytic synthesis of H2O2 was shown in Figure 14h, which included two synthesis pathways. In WOR part, two adjacent C-NH groups on DHI absorb two water molecules. Then, 2e WOR generated H2O2, leaving behind two protonated amine groups (NH2+). On the other hand, NH2+ on PDA acted as a key position for the adsorption and reduction of •O2, which can rapidly convert •O2 into H2O2. This prevented the accumulation of •O2 that could lead to the destruction of the catalyst structure and promotes the regeneration of the C-NH group sites, thereby achieving efficient continuous production of H2O2. During the photocatalytic production of hydrogen peroxide, the electrons in the conduction band react with oxygen in the solution, reducing it to O2. The superoxide anion further reacts with H+ to eventually form H2O2. PDA has strong adsorption capabilities, effectively capturing dissolved oxygen and inhibiting H2O2 decomposition. In addition, the catechol in PDA reacted with O2 to produce H2O2 and obenzoquinone. Studies on the photocatalytic synthesis of hydrogen peroxide by PDA have the potential to develop green and sustainable hydrogen peroxide synthesis methods. This method can be used to produce hydrogen peroxide under environmentally friendly conditions using renewable energy sources such as sunlight and has broad application prospects. 5.5. Applications in Other Photocatalytic Reactions 5.5.1. N2 Fixation As an essential energy source, ammonia (NH3) has a significant impact in industrial and agricultural production. In recent years, the photochemical synthesis of NH3 from atmospheric nitrogen (N2) has become a hot research frontier [159]. However, photocatalytic nitrogen fixation materials face limitations such as the difficulty of N2 adsorption and activation, and the challenge of preventing the recombination of photogenerated carriers [160]. PDA has shown promising application in photocatalytic nitrogen fixation due to its unique properties. Wang et al. constructed a ternary SiW9Co3/PDA/BWO photocatalyst, where PDA acting as an adhesive bridge between SiW9Co3 with BWO [161]. In this work, PDA had a broad light absorption range and acted as an electron transport chain to facilitate electron–hole migration through π–π electronic delocalization (Figure 15a). In addition, the SiW9Co3/PDA/BWO provided suitable binding sites for molecular nitrogen activation due to the synergy between three photocatalysts. The catalytic efficiency of Z-scheme ternary SiW9Co3/PDA/BWO was 6.0 times higher than that of BWO.
Figure 15. (<b>a</b>) Schematic illustration of separation and migration of photo-generated electron-hole pairs over the Z-scheme SiW<sub>9</sub>Co<sub>3</sub>/PDA/BWO heterojunction [161]. (<b>b</b>) The plausible mechanism for selective aerobic oxidation of thioanisole under visible light by combining TiO<sub>2</sub>@PDA photocatalysis and TEMPO catalysis [162].
5.5.2. Organic Synthesis PDA is extensively utilized in organic synthesis owing to its chemical properties and functionality. It exhibits catalytic activity and can generate active free radicals to facilitate oxidation, reduction, and coupling reactions. PDA possesses abundant surface functional groups, which allow for easy functionalization, making it an excellent substrate for developing catalysts and adsorbents. Additionally, PDA can serve as a support to immobilize metal or organic catalysts. Its functional groups can engage in reactions or alter substrate adsorption, thereby influencing reactivity. The biocompatibility of PDA also renders it particularly valuable in bioorganic synthesis, making it suitable for the synthesis of bioactive molecules or drug precursors [163,164,165]. Many studies have shown that PDA, as a multifunctional coating on the surface of metal oxide semiconductors, can significantly improve the photoelectric performance of these materials due to the presence of π-π electron transitions. Shi et al. constructed the TiO2@PDA with a large π-conjugated PDA system and used for aerobic oxidation reaction of thioanisole [162]. The possible mechanism was shown in Figure 15b, it was worth noting that PDA was activated to an unstable intermediate PDA radical•+, which can regenerate into PDA by transforming TEMPO into TEMPO+. Thioanisole, under the action of TEMPO+, was converted into S-centered radical cation, which prefer to react with O2•− to form persulfoxide. Then the persulfoxide combined with CH3OH to yield the sulfoxide product. Hence, the strategy of surface modification of semiconductor by PDA had a great potential for photocatalytic organic transformations selectively. In addition, the synthesis of amide compounds holds great significant importance in organic synthesis. Xu and co-workers reported a magnetic Fe3O4/PDA/CdS active photocatalyst for the amidation of aromatic aldehydes under air, where PDA improved the photocatalytic activity by its light energy conversion [166]. In a word, PDA photocatalytic nitrogen fixation is an advanced research topic focusing on using polydopamine as a photocatalyst to convert N2 into NH3 under light irradiation. PDA can be used as a conductive adhesive to bridge to synthesize ternary heterojunctions with other catalysts. The PDA facilitates the separation of photogenerated electron-hole pairs from the other two semiconductors, accelerates electron transfer, and reduces recombination, which is essential for improving photocatalytic reaction efficiency. Alternatively, PDA photocatalysis system can be used to synthesize various organic compounds, possibly providing new routes to synthesize complex molecules.

6. Conclusions and Prospects

This review provides an in-depth discussion on the structural characteristics and synthesis methods of polydopamine (PDA). It highlights PDA’s advantages in photocatalytic applications, such as its templating agent effect, light absorption performance, film-forming ability, hydrophilicity, and conductivity. The article also addresses strategies for enhancing PDA’s photocatalytic performance and explores the application of PDA-based materials in various fields, including solar water purification (targeting heavy metal reduction, organic pollutant degradation, and antibacterial activity), H2 production, H2O2 production, CO2 reduction, and organic transformation. The ultimate goal of the review is to offer valuable insights for the design and development of PDA-based photocatalytic materials. While PDA holds significant promise in the field of photocatalysis, it also faces certain challenges that need to be addressed. (1) To address the complexity of the PDA synthesis process and the difficulty of functionalization, the following strategies can be used to improve the process: simplified synthesis methods, such as one-step synthesis and direct oxidation, as well as the use of microwave-assisted synthesis to accelerate the reaction; simplified methods of functionalization, including the use of pre-functionalized dopamine precursors and in-situ functionalization techniques; highly efficient methods of functionalization, such as surface self-assembly and cross-linking functionalization; green synthesis methods, such as bio templating and enzymatic polymerization; and template-assisted synthesis, including hard and soft template methods, to control the morphology and structure of PDA. These strategies help to improve the preparation efficiency and performance of PDA and enhance their maneuverability and economy in practical applications. (2) To improve the limited light absorption range and weak absorption capacity of PDA, it can be improved by various strategies to enhance its potential for photocatalytic applications. These strategies include: doping of metallic or nonmetallic elements to modulate the energy band structure and electronic structure; dye sensitization (e.g., phthalocyanine and porphyrin dyes) to broaden the light absorption range; construction of composites of PDA with other inorganic semiconductors to fabricate heterojunctions to improve light absorption capacity; surface modification and functionalization to introduce conjugated groups (e.g., benzene rings, alkynyl groups, etc.) or grafting of photosensitizing compounds to change the light absorption properties; design of different nanostructures such as nanoparticles and porous structures to improve light absorption efficiency; and other approaches such as composite with surface plasmon resonance effect materials and copolymer design. Together, these approaches can help to enhance the effectiveness of PDA in applications, such as environmental remediation and energy conversion. (3) To solve the problem of high PDA photogenerated carrier complexation rate, various strategies can be adopted, including metallic or nonmetallic doping to modulate the energy band structure and promote carrier separation, dye sensitization to enhance light absorption and assist in carrier separation, construction of heterojunctions to promote the transfer of electrons and holes, surface modification and functionalization to change the electronic structure, design of novel nanostructures to increase the carrier transport path, composite with surface plasmon resonance effect materials to enhance the electromagnetic field and promote carrier separation, optimization of synthesis methods to improve the electronic structure of the materials, application of external electric or magnetic fields to guide the directional movement of carriers, and optimization of photocatalytic reaction conditions to promote carrier separation. The combined application of these methods can significantly reduce the photogenerated carrier complexation rate in PDAs and enhance the photocatalytic efficiency. (4) To enhance the stability of PDA, the reaction conditions can be adjusted to optimize the pH to a weakly basic range and synthesize at low temperature to reduce the occurrence of side reactions; doping with metallic or non-metallic elements to enhance the stability; surface modification to reduce hydrolysis and degradation through the introduction of hydrophobic groups; the use of chemical cross-linking agents to form a stable three-dimensional network structure; designing the nanostructures of PDA, such as nanoparticles and porous structures, to enhance stability; optimization of the synthesis process such as shortening the reaction time and adjusting the stirring speed to reduce side reactions and improve the homogeneity of the reaction. The implementation of these strategies helps to improve the stability of PDA, which in turn provides better performance and wider application prospects in areas such as photocatalysis and environmental remediation. (5) PDA holds significant potential in the field of environmental remediation and energy conversion due to its unique chemical properties and biocompatibility. For instance, (a): PDA combined with sporous materials can serve as support materials for photocatalysts, enhancing their stability and recyclability. Its excellent adsorption properties allow it to effectively capture pollutants in water, such as organic dyes and heavy metal ions, facilitating their degradation under light irradiation; (b): In the process of water splitting for hydrogen production, PDA can act as a co-catalyst, promoting the separation of photogenerated electrons and holes, thereby enhancing hydrogen production efficiency. (c): PDA also has potential applications in the photocatalytic reduction of CO2. It can interact with CO2 molecules, increasing their adsorption on the photocatalyst surface, which facilitates the reduction reaction and improves the selectivity and efficiency of the process; (d): Leveraging PDA’s photocatalytic properties, self-cleaning and antimicrobial coatings can be developed. When exposed to light, PDA coatings can catalyze the degradation of organic pollutants and bacteria, maintaining surface cleanliness. This application is particularly promising in building materials, medical devices, and textiles; (e) The excellent biocompatibility, tunable photothermal conversion performance, and extensive surface modification ability make PDA become a potential material in the fields of biomedicine (photothermal therapy, antibacterial applications), energy (energy storage), and environmental science (solar-driven evaporation and desalination). In summary, improving the comprehensive performance of PDA includes the use of simplified synthesis, functionalization strategies, doping, sensitization, surface and nanostructure design to enhance their preparation efficiency, light absorption ability and photocatalytic potential. Meanwhile, the stability of PDA can be effectively enhanced by adjusting the reaction conditions, doping, surface modification, cross-linking, nano structuring, and synthesis process, thus expanding its application in photocatalysis.

Acknowledgments

The authors would like to thank Canada First Research Excellence Fund (CFREF) and Natural Sciences and Engineering Research Council of Canada-Discovery Grant for their funding. H.W. acknowledges the financial support from the China Scholarship Council.

Author Contributions

H.W.: Writing–original draft, Supervision, Formal analysis. L.J.: Writing–review & editing, Supervision. A.V.: Supervision. Y.Y.: Supervision. H.Z.: Supervision. Q.G.: Supervision. J.H.: Writing–review& editing, Supervision, Funding acquisition.

Ethics Statement

Not applicable.

Informed Consent Statement

Not applicable.

Funding

This research was funded by Natural Sciences and Engineering Research Council of Canada, Discovery Grant [grant number: 10040079].

Declaration of Competing Interest

The authors declare that they have no known competing financial interests or personal relationships that could have appeared to influence the work reported in this paper.

References

1.
Kumari H, Sonia, Suman, Ranga R, Chahal S, Devi S, et al. A review on photocatalysis used for wastewater treatment: Dye degradation.  Water Air Soil Pollut. 2023, 234, 349. [Google Scholar]
2.
Fang S, Rahaman M, Bharti J, Reisner E, Robert M, Ozin GA, et al. Photocatalytic CO2 reduction.  Nat. Rev. Method Prime 2023, 3, 61. [Google Scholar]
3.
Cao W, Zhang W, Dong L, Ma Z, Xu J, Gu X, et al. Progress on quantum dot photocatalysts for biomass valorization. Exploration 2023, 3, 20220169. [Google Scholar]
4.
Kong Y, Yang C, Cai Y, Mu X, Li L. Photocatalytic aerobic conversion of methane. Photocatal. Res. Potential 2023, 1, 10005. [Google Scholar]
5.
Huang Z, Luo N, Zhang C, Wang F. Radical generation and fate control for photocatalytic biomass conversion.  Nat. Rev. Chem. 2022, 6, 197–214. [Google Scholar]
6.
Wang J, Wang X, Zhao H, Van Humbeck JF, Richtik BN, Dolgos MR, et al. Selective C3–C4 cleavage via glucose photoreforming under the effect of nucleophilic dimethyl sulfoxide.  ACS Catal. 2022, 12, 14418–14428. [Google Scholar]
7.
Wang J, Zhao H, Chen L, Björk J, Rosen J, Kumar P, et al. Selective cellobiose photoreforming for simultaneous gluconic acid and syngas production in acidic conditions. Appl. Catal. B Environ. 2024, 344, 123665. [Google Scholar]
8.
Yang P, Zhu F, Zhang Z, Cheng Y, Wang Z, Li Y. Stimuli-responsive polydopamine-based smart materials. Chem. Soc. Rev. 2021, 50, 8319–8343. [Google Scholar]
9.
Aguilar-Ferrer D, Szewczyk J, Coy E. Recent developments in polydopamine-based photocatalytic nanocomposites for energy production: Physico-chemical properties and perspectives.  Catal. Today 2022, 397–399, 316–349. [Google Scholar]
10.
Zhu X, Chen C, Shi Y, Wang J, Wang Y, Pan L, et al. Dual-functional, highly efficient CaIn2S4/PDA@SnO2 photocatalyst with Z-Scheme for photocatalytic hydrogen production from water splitting and organic pollutant degradation. J. Alloys Compd. 2024, 976, 173151. [Google Scholar]
11.
Han G, Xu F, Cheng B, Li Y, Yu J, Zhang L. Enhanced photocatalytic H2O2 production over inverse opal ZnO@Polydopamine S-scheme heterojunctions. Acta Phys. Chim. Sin. 2022, 38, 2112037. [Google Scholar]
12.
Zhang X, Yu J, Macyk W, Wageh S, Al-Ghamdi AA, Wang L. C3N4/PDA S-Scheme heterojunction with enhanced photocatalytic H2O2 production performance and its mechanism. Adv. Sustain. Syst. 2022, 7, 2200113. [Google Scholar]
13.
Feng S, Wang S, Lei C, Liu S, Jin Q. Photocatalytic degradation of antibiotics by S-Scheme heterojunctions constructed by thermally sheared flower-like TiO2-loaded PDA. Catal. Lett. 2023, 153, 3783–3794. [Google Scholar]
14.
da Silva AFV, da Silva J, Vicente R, Ambrosi A, Zin G, Di Luccio M, et al. Recent advances in surface modification using polydopamine for the development of photocatalytic membranes for oily wastewater treatment. J. Water Process Eng. 2023, 53, 103743. [Google Scholar]
15.
Yang D, Wang W, Zhao X, Zhou Z, Ren H, Chen Y, et al. Synthesis of high-efficient g-C3N4/polydopamine/CdS nanophotocatalyst based on bioinspired adhesion and chelation.  Mater. Res. Bull. 2020, 131, 110970. [Google Scholar]
16.
Bernsmann F, Ball V, Addiego F, Ponche A, Michel M, Gracio JJ, et al. Dopamine-melanin film deposition depends on the used oxidant and buffer solution. Langmuir 2011, 27, 2819–2825. [Google Scholar]
17.
Liu Y, Ai K, Lu L. Polydopamine and its derivative materials: Synthesis and promising applications in energy, environmental, and biomedical fields.  Chem. Rev. 2014, 114, 5057–5115. [Google Scholar]
18.
Hong S, Na YS, Choi S, Song IT, Kim WY, Lee H. Non-covalent self-assembly and covalent polymerization Co-contribute to polydopamine formation. Adv. Funct. Mater. 2012, 22, 4711–4717. [Google Scholar]
19.
Li M, Xuan Y, Zhang W, Zhang S, An J. Polydopamine-containing nano-systems for cancer multi-mode diagnoses and therapies: A review. Int. J. Biol. Macromol. 2023, 247, 125826. [Google Scholar]
20.
Zou Y, Chen X, Yang P, Liang G, Yang Y, Gu Z, et al. Regulating the absorption spectrum of polydopamine. Sci. Adv. 2020, 6, eabb4696. [Google Scholar]
21.
Bai W, Yang P, Liu H, Zou Y, Wang X, Yang Y, et al. Boosting the optical absorption of melanin-like polymers. Macromolecules 2022, 55, 3493–3501. [Google Scholar]
22.
Dong W, Zhao Z, Liu F, Li P, Wang L, Zhou Y, et al. PVDF nanofiber modified with ZnO nanowires/polydopamine for the treatment of sewage containing heavy metals, organic dyes, and bacteria.  ACS Appl. Mater. Interfaces 2023, 15, 58994–59004. [Google Scholar]
23.
Sun X, Yan L, Xu R, Xu M, Zhu Y. Surface modification of TiO2 with polydopamine and its effect on photocatalytic degradation mechanism.  Colloids Surf. A 2019, 570, 199–209. [Google Scholar]
24.
Nie N, He F, Zhang L, Cheng B. Direct Z-scheme PDA-modified ZnO hierarchical microspheres with enhanced photocatalytic CO2 reduction performance. Appl. Surf. Sci. 2018, 457, 1096–1102. [Google Scholar]
25.
Yu Z, Li F, Yang Q, Shi H, Chen Q, Xu M. Nature-mimic method to fabricate polydopamine/graphitic carbon nitride for enhancing photocatalytic degradation performance. ACS Sustain. Chem. Eng. 2017, 5, 7840–7850. [Google Scholar]
26.
Gao J, He Q, Liang G, Luo L, Liu X, Cai K, et al. Heterogeneous alloyed CuSnO-DopaCube mediated photo-Fenton and photothermal synergistic catalysis for dye elimination. J. Alloys Compd. 2023, 960, 170737. [Google Scholar]
27.
Shen Z, Wang X, Fan D, Xu X, Lu Y. Wood–hydrogel composites coated with C3N4 photocatalyst for synchronous solar steam generation and photocatalytic degradation.  J. Mater. Sci. 2023, 58, 13154–13164. [Google Scholar]
28.
Usta DD, Celebi N, Soysal F, Saglam ASY, Yildiz N, Salimi K. Bio-inspired NIR responsive Au-pDA@pDA sandwich nanostructures with excellent photo-thermal performance and stability.  Colloids Surf. A 2021, 611, 125758. [Google Scholar]
29.
Zhao X, Zha X, Tang L, Pu J, Ke K, Bao R.-Y, et al. Self-assembled core-shell polydopamine@MXene with synergistic solar absorption capability for highly efficient solar-to-vapor generation. Nano Res. 2020, 13, 255–264. [Google Scholar]
30.
Lu Y, Fan D, Shen Z, Zhang H, Xu H, Yang X. Design and performance boost of a MOF-functionalized-wood solar evaporator through tuning the hydrogen-bonding interactions.  Nano Energy 2022, 95, 107016. [Google Scholar]
31.
Pang AL, Arsad A, Ahmad Zaini MA, Garg R, Saqlain Iqbal M, Pal U, et al. A comprehensive review on photocatalytic removal of heavy metal ions by polyaniline-based nanocomposites.  Chem. Eng. Commun. 2023, 211, 275–299. [Google Scholar]
32.
Jamali Alyani S, Dadvand Koohi A, Ashraf Talesh SS, Ebrahimian Pirbazari A. Investigation of TiO2/PPy nanocomposite for photocatalytic applications; synthesis, characterization, and combination with various substrates: A review.  Environ. Sci. Pollut. Res. Int. 2024, 31, 42521–42546. [Google Scholar]
33.
Yang Y, Cao Z, Shen T, Zhang J, Gu Z, Zhao M, et al. Synergistic photocatalytic-photothermal contribution induced by Pd-TiO2/PEDOT immobilized in PVDF membrane for enhanced water purification. Chem. Eng. J. 2024, 494, 153168. [Google Scholar]
34.
Jasemizad T, Malmström J, Padhye LP. Mechanistic investigation of the photocatalytic activity of PEDOT for aqueous contaminant removal: The role of iron and hydroxyl radicals.  Environ. Sci. Wat. Res. 2024, 10, 1256–1270. [Google Scholar]
35.
Dey S, Manna K, Pradhan P, Sarkar AN, Roy A, Pal S. Review of polymeric nanocomposites for photocatalytic wastewater treatment.  ACS Appl. Nano Mater. 2024, 7, 4588–4614. [Google Scholar]
36.
Barclay TG, Hegab HM, Clarke SR, Ginic-Markovic M. Versatile surface modification using polydopamine and related polycatecholamines: Chemistry, structure, and applications. Adv. Mater. Interfaces 2017, 4, 1601192. [Google Scholar]
37.
Wang W, Dong X, Huang X, Wu H, Zhu D. Polydopamine-mediated CdS crystal growth behavior and the synergistic photocatalytic–adsorption mechanisms in removing Rhodamine B. Energ. Fuels 2023, 37, 18184–18193. [Google Scholar]
38.
Ryu JH, Messersmith PB, Lee H. Polydopamine surface chemistry: A decade of discovery. ACS Appl. Mater. Interfaces 2018, 10, 7523–7540. [Google Scholar]
39.
Shi J, Zhang J, Jia Z, Zhao R, Abubakr Yasin UA, Du J. Construction of yolk-shell structural polydopamine@WO3 micro/nanospheres for enhanced photocatalytic degradation.  ChemistrySelect 2023, 8, e202301084. [Google Scholar]
40.
Qin Y, Xu L, Zhu Z, Wong W-Y. Strongly coupled interface facilitating charge separation to the improved visible light-driven hydrogen production on CdS@polydopamine/NiS photocatalyst.  J. Mater. Chem. A 2023, 11, 11840–11848. [Google Scholar]
41.
Qu R, Zhang W, Liu N, Zhang Q, Liu Y, Li X, et al. Antioil Ag3PO4 nanoparticle/polydopamine/Al2O3 sandwich structure for complex wastewater treatment: Dynamic catalysis under natural light.  ACS Sustain. Chem. Eng. 2018, 6, 8019–8028. [Google Scholar]
42.
Zhong J, Xiao F, Jin B, Zhang G, Hou H, Bi J, et al. Construction of sandwich-structured V2CTx@PDA@Ag nanosheet composite for enhanced near-infrared photothermocatalytic sterilization performance.  Mater. Chem. Phys. 2024, 314, 128931. [Google Scholar]
43.
Shang L, Li W, Wang X, Ma L, Li L, Duan Q, et al. Preparation of magnetic Fe3O4@PDA/CuS core-shell nanocomposite as a green photocatalyst.  Synth. Met. 2023, 292, 117230. [Google Scholar]
44.
Inoue A, Yuk H, Lu B, Zhao X. Strong adhesion of wet conducting polymers on diverse substrates. Sci. Adv. 2020, 6, eaay5394. [Google Scholar]
45.
Beygisangchin M, Abdul Rashid S, Shafie S, Sadrolhosseini AR, Lim HN. Preparations, properties, and applications of polyaniline and polyaniline thin films-a review. Polymers 2021, 13, 2003. [Google Scholar]
46.
Wen J, Li X, Zhang H, Zheng S, Yi C, Yang L, et al. Architecting Janus hydrogel evaporator with polydopamine-TiO2 photocatalyst for high-efficient solar desalination and purification.  Sep. Purif. Technol. 2023, 304, 122403. [Google Scholar]
47.
Feng X, Liu C, Long R, Liu S, Liu X. Polydopamine-modified magnetic foam composites synergistically activate peroxymonosulfate with visible light for sulfadiazine degradation and oil–water separation. Chem. Eng. J. 2023, 478, 147252. [Google Scholar]
48.
Chen X, Ma H, Ji X, Han R, Pang K, Yang Z, et al. Engineering green MOF-based superhydrophobic sponge for efficiently synchronous removal of microplastics and pesticides from high-salinity water.  Water Res. 2023, 243, 120314. [Google Scholar]
49.
Duan J, Fang X, Li C, Qu J, Guo L, Zou Y, et al. Efficient and stable monolithic microreactor with Ag/AgCl photocatalysts coated on polydopamine modified melamine sponge for photocatalytic water purification. Colloids Surf. A 2023, 659, 130759. [Google Scholar]
50.
Ran J, Bi S, Jiang H, Telegin F, Bai X, Yang H, et al. Core–shell BiVO4@PDA composite photocatalysts on cotton fabrics for highly efficient photodegradation under visible light. Cellulose 2019, 26, 6259–6273. [Google Scholar]
51.
Ran J, Chen H, Bai X, Bi S, Jiang H, Cai G, et al. Immobilizing CuO/BiVO4 nanocomposite on PDA-templated cotton fabric for visible light photocatalysis, antimicrobial activity and UV protection. Appl. Surf. Sci. 2019, 493, 1167–1176. [Google Scholar]
52.
Zhao J, Chen J, Chen Z, Zhang Y, Xia D, Wang Q. Flexible cotton fabrics/PDA/BiOBr composite photocatalyst using bioinspired polydopamine as electron transfer mediators for dye degradation and Cr(VI) reduction under visible light.  Colloids Surf. A 2020, 593, 124623. [Google Scholar]
53.
Fan G, Ning R, Li X, Lin X, Du B, Luo J, et al. Mussel-inspired immobilization of photocatalysts with synergistic photocatalytic-photothermal performance for water remediation. ACS Appl. Mater. Interfaces 2021, 13, 31066–31076. [Google Scholar]
54.
Zhou J, Zhao Z, Wang Y, Ding Z, Xu X, Peng W, et al. BiOCl0.875Br0.125/polydopamine functionalized PVDF membrane for highly efficient visible-light-driven photocatalytic degradation of roxarsone and simultaneous arsenic immobilization. Chem. Eng. J. 2020, 402, 126048. [Google Scholar]
55.
Shi Y, Chen X, Wu Q, Zhen H, Wang S, Dong H, et al. Enhance organic pollutants removal of wastewater by a PVDF/PDA-TiO2 composite membrane with photocatalytic property. J. Environ. Chem. Eng. 2023, 11, 110389. [Google Scholar]
56.
Wang B, Cao Q, Cheng M, Li G, Zhang J, Jiang H. Photocatalytic degradation of antibiotics in water by pollution-free photocatalytic films with a three-dimensional layered structure and the reaction mechanism study. J. Water Process Eng. 2023, 52, 103550. [Google Scholar]
57.
Huang Z, Zeng Q, Hui Y, Alahi MEE, Qin S, Wu T. Fast polymerization of polydopamine based on titanium dioxide for high-performance flexible electrodes.  ACS Appl. Mater. Interfaces 2020, 12, 14495–14506. [Google Scholar]
58.
Zhou Z, Yang D, Wang W, Chen Y, Zhao Z, An K, et al. Bioinspired construction of g-C3N4 nanolayers on a carbonized polydopamine nanosphere surface with excellent photocatalytic performance. Ind. Eng. Chem. Res. 2020, 59, 12389–12398. [Google Scholar]
59.
Li Y, He Z, Liu L, Jiang Y, Ong W-J, Duan Y, et al. Inside-and-out modification of graphitic carbon nitride (g-C3N4) photocatalysts via defect engineering for energy and environmental science.  Nano Energy 2023, 105, 108032. [Google Scholar]
60.
Huo R, Yang X-L, Yang J-Y, Yang S-Y, Xu Y-H. Self-assembly synthesis of BiVO4/Polydopamine/g-C3N4 with enhanced visible light photocatalytic performance. Mater. Res. Bull. 2018, 98, 225–230. [Google Scholar]
61.
Li X, Yu Z, Shao L, Zeng H, Liu Y, Feng X. A novel strategy to construct a visible-light-driven Z-scheme (ZnAl-LDH with active phase/g-C3N4) heterojunction catalyst via polydopamine bridge (a similar “bridge” structure).  J. Hazard. Mater. 2020, 386, 121650. [Google Scholar]
62.
Guo F, Chen J, Zhao J, Chen Z, Xia D, Zhan Z, et al. Z-scheme heterojunction g-C3N4@PDA/BiOBr with biomimetic polydopamine as electron transfer mediators for enhanced visible-light driven degradation of sulfamethoxazole. Chem. Eng. J. 2020, 386, 124014. [Google Scholar]
63.
Lam Y, Fan S, He L, Ho Y, Fei B, Xin JH. Charge-controllable mussel-inspired magnetic nanocomposites for selective dye adsorption and separation.  Chemosphere 2022, 300, 134404. [Google Scholar]
64.
Wu C, Zuo H, Zhang S, Zhao S, Du H, Yan Q. A novel strategy to construct a direct Z-Scheme Bi@Bi2O2CO3/g-C3N4 heterojunction catalyst via PDA electronic bridge. Sep. Purif. Technol. 2022, 294, 121242. [Google Scholar]
65.
Liu X, Zhang T, Li Y, Zhang J, Du Y, Yang Y, et al. CdS@Polydopamine@SnO2−x sandwich structure with electrostatic repulsion effect and oxygen deficiency: Enhanced photocatalytic hydrogen evolution activity and inhibited photo-corrosion. Chem. Eng. J. 2022, 434, 134602. [Google Scholar]
66.
Gautam S, Agrawal H, Thakur M, Akbari A, Sharda H, Kaur R, et al. Metal oxides and metal organic frameworks for the photocatalytic degradation: A review.  J. Environ. Chem. Eng. 2020, 8, 103726. [Google Scholar]
67.
Majeed AH, Mohammed LA, Hammoodi OG, Sehgal S, Alheety MA, Saxena KK, et al. A review on polyaniline: Synthesis, properties, nanocomposites, and electrochemical applications.  Int. J. Polym. Sci. 2022, 2022, 9047554. [Google Scholar]
68.
Pang AL, Arsad A, Ahmadipour M. Synthesis and factor affecting on the conductivity of polypyrrole: A short review. Polym. Advan. Technol. 2020, 32, 1428–1454. [Google Scholar]
69.
Li L, Chi L, Zhang H, Wu S, Wang H, Luo Z, et al. Fabrication of Ti-PDA nanoparticles with enhanced absorption and photocatalytic activities for hexavalent chromium Cr(VI) removal. Appl. Surf. Sci. 2022, 580, 152168. [Google Scholar]
70.
Li W, Shang L, You D, Li H, Wang H, Ma L, et al. Fabrication of Fe3O4@Ti-PDA nanoparticles with enhanced photocatalytic activities for degradation of organic dye. J. Phys. Chem. Solids 2023, 172, 111047. [Google Scholar]
71.
Wang H, Yan R, Zou Y, Xing D, Zhong K. Light-driven self-healing polyurethane based on PDA@ Ag nanoparticles with improved mechanical and antibacterial properties.  J. Mater. Chem. B 2022, 10, 1085–1093. [Google Scholar]
72.
Hou X, Cai Y, Song X, Wu Y, Zhang J, Wei Q. Electrospun TiO2 nanofibers coated with polydopamine for enhanced sunlight-driven photocatalytic degradation of cationic dyes.  Surf. Interface Anal. 2019, 51, 169–176. [Google Scholar]
73.
Han D, Li Y, Liu X, Yeung KWK, Zheng Y, Cui Z, et al. Photothermy-strengthened photocatalytic activity of polydopamine-modified metal-organic frameworks for rapid therapy of bacteria-infected wounds. J. Mater. Sci. Technol. 2021, 62, 83–95. [Google Scholar]
74.
Wang W, Li M, Huang X, Fang J, Peng F, Huang H. Structural evolution mechanisms of Polydopamine/CdS and photothermal effect boosted photocatalytic H2 production activity.  Appl. Surf. Sci. 2022, 601, 154114. [Google Scholar]
75.
Mei H, Gao Z, Wang Q, Sun H, Zhao K, Zhang P, et al. Ultrasound expands the versatility of polydopamine coatings. Ultrason.  Sonochem. 2021, 74, 105571. [Google Scholar]
76.
Gong Z, Chen L, Chen K, Gou S, Zhao X, Song L, et al. Significantly improved photocatalytic H2O2 generation over PDA-modified g-C3N4 via promoting charge-carriers separation and oxygen adsorption. J. Environ. Chem. Eng. 2023, 11, 109405. [Google Scholar]
77.
Li Y, Wei W, Guo Z, Zou L, Li M, Ai L, et al. Double-sided assembly of 0D polydopamine on 1D hexagonal tubular carbon nitride for boosting photocatalytic Cr(VI) reduction.  Sep. Purif. Technol. 2022, 299, 121730. [Google Scholar]
78.
Chu J, Li W, Cao Z, Bai X, Rao X, Zheng S, et al. Benzene bridged carbon nitride for efficient photocatalytic hydrogen evolution.  Photocatal. Res. Potential 2024, 1, 10001. [Google Scholar]
79.
Zhang H, Zhao H, Zhai S, Zhao R, Wang J, Cheng X, et al. Electron-enriched Lewis acid-base sites on red carbon nitride for simultaneous hydrogen production and glucose isomerization. Appl. Catal. B Environ. 2022, 316, 121647. [Google Scholar]
80.
Chen Y, Dong S, Zhang H, Wei Y, Wang B, Zhang Y. Boosting electron transfer in Cu2O@PDA hollow fiber membranes to decompose oxytetracycline under visible light with persulfate.  Sep. Purif. Technol. 2023, 318, 124024. [Google Scholar]
81.
Chen F, Zhao L, Yu W, Wang Y, Zhang H, Guo L-H. Dynamic monitoring and regulation of pentachlorophenol photodegradation process by chemiluminescence and TiO2/PDA. J. Hazard. Mater. 2020, 399, 123073. [Google Scholar]
82.
Sridharan M, Kamaraj P, Vennila R, Huh YS, Arthanareeswari M. Bio-inspired construction of melanin-like polydopamine-coated CeO2 as a high-performance visible-light-driven photocatalyst for hydrogen production. New J. Chem. 2020, 44, 15223–15234. [Google Scholar]
83.
Chen F, Yu W, Qie Y, Zhao L, Zhang H, Guo L-H. Enhanced photocatalytic removal of hexavalent chromium through localized electrons in polydopamine-modified TiO2 under visible irradiation.  Chem. Eng. J. 2019, 373, 58–67. [Google Scholar]
84.
Zhou X, Jin B, Luo J, Gu X, Zhang S. Photoreduction preparation of Cu2O@polydopamine nanospheres with enhanced photocatalytic activity under visible light irradiation.  J. Solid State Chem. 2017, 254, 55–61. [Google Scholar]
85.
Yang F, Wang F, Guo, Z. Characteristics of binary WO3@ CuO and ternary WO3@ PDA@ CuO based on impressive sensing acetone odor.  J. Colloid Interface Sci. 2018, 524, 32–41. [Google Scholar]
86.
Zhao R, Zhang H, Zhao R, Liu W, Liu Q, Hu H. Dynamic Cu(I)/Cu(II) redox shuttle for maltose and lactose isomerization under pulsed electric field. ACS Sustain. Chem. Eng. 2024, 12, 6232–6241. [Google Scholar]
87.
Chilivery R, Zhang R, Chen G, Yao D, Fan D, Lu F, et al. Facile in situ construction of novel hybrid 3D-BiOCl@PDA heterostructures with vacancy induced charge transfer for efficient visible light driven photocatalysis and antibacterial activity. Colloids Surf. A 2023, 656, 130415. [Google Scholar]
88.
Huang X, Niu Y, Peng Z, Hu W. Core–shell structured BiOCl@polydopamine hierarchical hollow microsphere for highly efficient photocatalysis. Colloids Surf. A 2019, 580, 123747. [Google Scholar]
89.
Ren H-T, Jing M-Z, Li D-S, Li T-T, Lou C-W, Lin J-H, et al. Photocatalytic reduction of Cr(VI) by Bi2.15WO6 complexed with polydopamine: Contribution of the ligand-to-metal charge transfer path.  J. Colloid Interface Sci. 2022, 622, 50–61. [Google Scholar]
90.
Zhao X, Zhang N, Yu Y, Fang T, Hu J, Feng J, et al. Exploring Bi4V2O11 as photoanode for water splitting with a wide range of solar light capture and suitable band potential. Photocatal.  Res. Potential 2023, 1, 10002. [Google Scholar]
91.
Huang S, Gao J, Zhou L, Lei J, Wang L, Liu Y, et al. Polydopamine-loaded ZnIn2S4 photocatalyst for H2O2 production in water.  ACS Appl. Nano Mater. 2024, 7, 4481–4490. [Google Scholar]
92.
Feng C, Zhang L. Microdroplet assisted hollow ZnCdS@ PDA nanocages’ synergistic confinement effect for promoting photocatalytic H2O2 production. Mater. Horiz. 2024, 11, 1515–1527. [Google Scholar]
93.
Lu J, Fang C, Wang G, Zhu L. Design of one-Dimensional cadmium sulfide/polydopamine heteronanotube photocatalysts for ultrafast degradation of antibiotics. Ind. Eng. Chem. Res. 2022, 61, 1100–1110. [Google Scholar]
94.
Kim Y, Coy E, Kim H, Mrówczyński R, Torruella P, Jeong D-W, et al. Efficient photocatalytic production of hydrogen by exploiting the polydopamine-semiconductor interface. Appl. Catal. B Environ. 2021, 280, 119423. [Google Scholar]
95.
Meng S, Zeng W, Wang M, Niu L, Hu S, Su B, et al. Nature-mimic fabricated polydopamine/MIL-53 (Fe): Efficient visible-light responsive photocatalysts for the selective oxidation of alcohols. New J. Chem. 2020, 44, 2102–2110. [Google Scholar]
96.
Celebi N, Aydin MY, Soysal F, Yıldız N, Salimi K. Core/shell PDA@UiO-66 metal–organic framework nanoparticles for efficient visible-light photodegradation of organic dyes.  ACS Appl. Nano Mater. 2020, 3, 11543–11554. [Google Scholar]
97.
Huang H, Hou L, Du H, Wei H, Liu X, Wang Q, et al. Efficient dual defense: PDA-Cu coating for simultaneous corrosion resistance and antibacterial protection of Mg alloys. Corros. Sci. 2024, 233, 112103. [Google Scholar]
98.
Tan L, Zhu T, Huang Y, Yuan H, Shi L, Zhu Z, et al. Ozone-induced rapid and green synthesis of polydopamine coatings with high uniformity and enhanced stability. Adv. Sci. 2024, 11, e2308153. [Google Scholar]
99.
Hu S, Du H, Huang H, Wei Y, Hou L, Wang Q, et al. Deposition of modifiable MAO-PDA coatings on magnesium alloy based on photocatalytic effect. Appl. Surf. Sci. 2024, 669, 160522. [Google Scholar]
100.
Zhang HD, Chen AY, Gan B, Jiang H, Gu LJ. Corrosion protection investigations of carbon dots and polydopamine composite coating on magnesium alloy.  J. Magnes. Alloy 2022, 10, 1358–1367. [Google Scholar]
101.
Gao J, Cheng S, Zhou P, Zhang J, Li H, Zhang Y, et al. Fenton-driven photothermal enhanced degradation of methylene blue by using Cu2O@SnO2@PDA cubic heterostructured catalyst.  Mater. Lett. 2023, 334, 133724. [Google Scholar]
102.
Wen L, Wanpei H, Qian L, Xu L, Rongsheng C, Hongwei N, et al. Antibacterial properties of Ag/TiO2/PDA nanofilm on anodized 316L stainless steel substrate under illumination by a normal flashlight.  J. Mater. Sci. 2020, 55, 9538–9550. [Google Scholar]
103.
Jiang J, Chen Z, Wang P, Gu P-Y, Liu J, Zhang Z, et al. Preparation of black hollow TiO2 nanotube-coated PDA@Ag2S heterostructures for efficient photocatalytic reduction of Cr(VI).  J. Solid State Chem. 2022, 307, 122865. [Google Scholar]
104.
Chen F, Yu W, Wang Y, Wang S, Liang Y, Wang L, et al. Dynamic control of pentachlorophenol photodegradation process using P25/PDA/BiOBr through regulation of photo-induced active substances and chemiluminescence.  Chemosphere 2022, 307, 135914. [Google Scholar]
105.
Zhang Z, Li M, Wang H. ZIF-8 derived ZnO decorated with polydopamine and Au nanoparticles for efficient photocatalytic degradation of Rhodamine B.  ChemistrySelect 2021, 6, 5356–5365. [Google Scholar]
106.
Shahsavandi F, Amirjani A, Reza Madaah Hosseini H. Plasmon-enhanced photocatalytic activity in the visible range using AgNPs/polydopamine/graphitic carbon nitride nanocomposite.  Appl. Surf. Sci. 2022, 585, 152728. [Google Scholar]
107.
Meng F, Qin Y, Lu J, Lin X, Meng M, Sun G, et al. Biomimetic design and synthesis of visible-light-driven g-C3N4 nanotube @polydopamine/NiCo-layered double hydroxides composite photocatalysts for improved photocatalytic hydrogen evolution activity.  J. Colloid Interface Sci. 2021, 584, 464–473. [Google Scholar]
108.
Yang Y, Lai Q, Mahmud S, Lu J, Zhang G, Huang Z, et al. Potocatalytic antifouling membrane with dense nano-TiO2 coating for efficient oil-in-water emulsion separation and self-cleaning. J. Membr. Sci. 2022, 645, 120204. [Google Scholar]
109.
Sun X-H, Chen S-H, Guo Q-Z, Shen Z-C, Wu J, Du Z-X. Convenient recycling and efficient photocatalytic degradation of disinfection by-products by in-situ growth MOF on PET.  Sep. Purif. Technol. 2023, 324, 124506. [Google Scholar]
110.
Zheng H, Meng X, Yang Y, Chen J, Huo S. Bifunctional photocatalytic nanofiltration membranes with immobilized BaTiO3/Ti3C2Tx catalysts for the simultaneous separation and degradation of azo compounds. J. Environ. Chem. Eng. 2023, 11, 110064. [Google Scholar]
111.
Sun A, Zhan Y, Chen X, Jia H, Zhu F, Zhao C, et al. Design of photocatalytic self-cleaning poly (arylene ether nitrile)/nitrogen-doped Bi2O2CO3 composite membrane for emulsified oily wastewater purification. J. Environ. Chem. Eng. 2023, 11, 110810. [Google Scholar]
112.
Hu Y, Zhao P, Liu H, Yi X, Song W, Wang X. Photocatalytic thin film composite forward osmosis membrane for mitigating organic fouling in active layer facing draw solution mode. Chin. Chem. Lett. 2023, 34, 107931. [Google Scholar]
113.
Liu Y, Liao M, Gan D, Chen M, Ma L, Yang B, et al. Photocatalytic RGO membrane with carbon nitride nanotube intercalation for enhanced wastewater purification. Colloids Surf. A 2023, 663, 131080. [Google Scholar]
114.
Ren S, Guo N, Li J, Wang Y. Integration of antibacterial and photocatalysis onto polyethersulfone membrane for fouling mitigation and contaminant degradation.  J. Environ. Chem. Eng. 2023, 11, 110401. [Google Scholar]
115.
Govardhanan B, GokulaKrishnan SA, Arthanareeswaran G, Ashok M, Chen Y-S, Taweepreda W. Photocatalytic bio-inspired PDA-RGO/g-CN nanocomposite incorporated PES membrane for removal of environmental antibiotic pollutants. J. Chem. Technol. Biotechnol. 2024, 99, 1401–1413. [Google Scholar]
116.
da Silva AFV, Della Rocca DG, Moreira RFPM, Oliveira JV, Ambrosi A, Di Luccio M. Synergistic effects of PVDF membrane surface functionalization with polydopamine and titanium dioxide for enhanced oil/water separation and photocatalytic behavior. J. Environ. Chem. Eng. 2024, 12, 111854. [Google Scholar]
117.
Ren H-T, Li D-S, Jing M-Z, Li T-T, Lou C-W, Lin J-H. Fabrication of sandwich structure membrane and its performance for photocatalytic reduction of Cr(VI). Cellulose 2023, 30, 3165–3183. [Google Scholar]
118.
Li R, Xiao G, Chen C, Chen C, Hu X, Li Y, et al. Self-assembly graphene oxide membrane embedded with CPU@PAA@TiO2 for enhancing the dyes separation and photocatalytic self-cleaning performance.  J. Membr. Sci. 2023, 683, 121816. [Google Scholar]
119.
Zhang H, Zhang J, Luo J, Wan Y. A novel paradigm of photocatalytic cleaning for membrane fouling removal. J. Membr. Sci. 2022, 641, 119859. [Google Scholar]
120.
Le TMH, Wang R, Sairiam S. Self-protecting PVDF-PDA-TiO2 membranes towards highly efficient and prolonged dye wastewater treatment by photocatalytic membranes. J. Membr. Sci. 2023, 683, 121789. [Google Scholar]
121.
Liu L, Wang D, Huang J, Huang Z, Zhang Y, Li L. Multicomponent composite membrane with three-phase interface heterostructure as photocatalyst for organic dye removal. ACS Omega 2022, 7, 17128–17143. [Google Scholar]
122.
Sui S, Quan H, Yang X, Dong X, Ji Y, Liu C, et al. Multifunctional modified polyurethane sponge for recovery of oil spills and photocatalytic degradation. Compos. Part B Eng. 2024, 271, 111176. [Google Scholar]
123.
Zheng L, Zhong K, Huang X, Qian X. Cellulose fibers/polydopamine/zinc sulfide composite paper: Predoping in situ fabrication for biphase interface photocatalytic hydrogen production from water. Cellulose 2024, 31, 2523–2540. [Google Scholar]
124.
Peng L, Wang H, Li G, Zhang W, Liang Z, An T. Photocatalytic inactivation of airborne bacteria onto g-C3N4/TiO2/Ni-polydopamine/Ni bifunctional coupling filter with non-size dependent capture effect.  Appl. Catal. B 2023, 329, 122580. [Google Scholar]
125.
Yuan M, Xue J, Li J, Ma S, Wang M. PCN-222/Ag2O-Ag p-n heterojunction modified fabric as recyclable photocatalytic platform for boosting bacteria inactivation and organic pollutant degradation. Colloids Surf. A 2023, 656, 130474. [Google Scholar]
126.
Wan J, Li H, Xu L, Yan J, Liao Y, Wang X. PDA/PEI-induced in-situ growth of a lotus leaf-like TiO2 nanoparticle film on N-halamine cotton fabric for photocatalytic, self-cleaning and efficient antibacterial performance. Cellulose 2023, 30, 3953–3972. [Google Scholar]
127.
Zhao CH, Zhang YP, Wan L, Chen XX, Yuan P, Qu LB. Versatile application of TiO2@PDA modified filter paper for oily wastewater treatment.  Molecules 2023, 28, 7903. [Google Scholar]
128.
Cui Z, Marcelle SSA, Zhao M, Wu J, Liu X, Si J, et al. Thermoplastic polyurethane/titania/polydopamine (TPU/TiO2/PDA) 3-D porous composite foam with outstanding oil/water separation performance and photocatalytic dye degradation. Adv. Compos. Hybrid Mater. 2022, 5, 2801–2816. [Google Scholar]
129.
Liu R, Dai L, Si C-L. Mussel-inspired cellulose-based nanocomposite fibers for adsorption and photocatalytic degradation. ACS Sustain. Chem. Eng. 2018, 6, 15756–15763. [Google Scholar]
130.
Zou Y, Wang Z, Chen Z, Zhang Q-P, Zhang Q, Tian Y, et al. Synthetic melanin hybrid patchy nanoparticle photocatalysts.  J. Phys. Chem. C 2019, 123, 5345–5352. [Google Scholar]
131.
Zango ZU, Jumbri K, Sambudi N, Ramli A, Abu Bakar N, Saad, B, et al. A critical review on metal-organic frameworks and their composites as advanced materials for adsorption and photocatalytic degradation of emerging organic pollutants from wastewater. Polymers 2020, 12, 2648. [Google Scholar]
132.
Nekouei F, Wang T, Keshtpour F, Liu Y, Li H, Nekouei S. Effective redox reaction in a three-body smart photocatalyst through multi-interface modulation of organic semiconductor junctioned with metal and inorganic semiconductor.  Appl. Catal. B Environ. Energ. 2024, 351, 123974. [Google Scholar]
133.
Aguilar-Ferrer D, Vasileiadis T, Iatsunskyi I, Ziółek M, Żebrowska K, Ivashchenko O, et al. Understanding the photothermal and photocatalytic mechanism of polydopamine coated gold nanorods. Adv. Funct. Mater. 2023, 33, 2304208. [Google Scholar]
134.
Zhang Z, Zhang N, Liu Y, Fang Q, Xi J, Xiao Y, et al. Efficient degradation of organic dyes and reduced Cr(VI) in environmental water purification by in-situ deposition of silver nanoparticles on polydopamine-modified M-ATP/PCN.  Catal. Commun. 2022, 172, 106528. [Google Scholar]
135.
Lu X, Gao Z, Wang G, Chen S, Gu Y, Yan B, et al. Mussel-inspired PDA/Ag nanocomposite catalyst for highly-efficient Cr(vi) removal via visible light-induced reduction and absorption. New J. Chem. 2023, 47, 9066–9076. [Google Scholar]
136.
Xing C, Ma M, Chang J, Ji Z, Wang P, Sun L, et al. Polyoxometalate anchored zinc oxide nanocomposite as a highly effective photocatalyst and bactericide for wastewater decontamination. Chem. Eng. J. 2023, 464, 142632. [Google Scholar]
137.
Ding T, Zhang L, Han J, Zhou J, Han Y. Photo-responded antibacterial therapy of reinfection in percutaneous implants by nanostructured bio-heterojunction.  Small 2023, 19, e2206265. [Google Scholar]
138.
Zhang F, Wu C, Zhou Z, Wang J, Bao W, Dong L, et al. Blue-light -activated nano-TiO2@PDA for highly effective and nondestructive tooth whitening.  ACS Biomater. Sci. Eng. 2018, 4, 3072–3077. [Google Scholar]
139.
Huang Q, Chen J, Liu M, Huang H, Zhang X, Wei Y. Polydopamine-based functional materials and their applications in energy, environmental, and catalytic fields: State-of-the-art review. Chem. Eng. J. 2020, 387, 124019. [Google Scholar]
140.
Teng X, Liu X, Cui Z, Zheng Y, Chen D-f, Li Z, et al. Rapid and highly effective bacteria-killing by polydopamine/IR780@MnO2–Ti using near-infrared light.  Prog. Nat. Sci- Mater. 2020, 30, 677–685. [Google Scholar]
141.
Dong W, Zhou S-A, Ma Y, Chi D-J, Chen R, Long H-M, et al. N-doped C-coated MoO2/ZnIn2S4 heterojunction for efficient photocatalytic hydrogen production.  Rare Met. 2023, 42, 1195–1204. [Google Scholar]
142.
Ma Y, Fang H-X, Chen R, Chen Q, Liu S-J, Zhang K, et al. 2D-MOF/2D-MOF heterojunctions with strong hetero-interface interaction for enhanced photocatalytic hydrogen evolution.  Rare Met. 2023, 42, 3993–4004. [Google Scholar]
143.
Yan Y, Zhou L, Wang H, Song Z, Jing L, Hu J. Photothermal ReO2/ReS2/Zn2In2S5 heterojunction for enhanced photocatalytic hydrogen evolution.  Sep. Purif. Technol. 2024, 346, 127473. [Google Scholar]
144.
Wang H, Lin Q, Yin L, Yang Y, Qiu Y, Lu C, et al. Biomimetic design of hollow Flower-like g-C3N4@PDA organic framework nanospheres for realizing an efficient photoreactivity.  Small 2019, 15, e1900011. [Google Scholar]
145.
Yang F, Wang S, Li Z, Xu Y, Yang W, Yv C, et al. Polydopamine/defective ultrathin mesoporous graphitic carbon nitride nanosheets as Z-scheme organic assembly for robust photothermal-photocatalytic performance. J. Colloid Interface Sci. 2022, 613, 775–785. [Google Scholar]
146.
Ran J, Jaroniec M, Qiao SZ. Cocatalysts in semiconductor-based photocatalytic CO2 reduction: Achievements, challenges, and opportunities. Adv. Mater. 2018, 30, 1704649. [Google Scholar]
147.
Xu S-R, Xiao F-X. Metal halide perovskites quantum dots: Synthesis and modification strategies for solar CO2 conversion. Chin. J. Struct. Chem. 2023, 42, 100173. [Google Scholar]
148.
Xu Q-J, Jiang J-W, Wang X-F, Duan L-Y, Guo H. Understanding oxygen vacant hollow structure CeO2@In2O3 heterojunction to promote CO2 reduction. Rare Met. 2023, 42, 1888–1898. [Google Scholar]
149.
Hayat A, Sohail M, Iqbal W, Taha TA, Alenad AM, Al-Sehemi AG, et al. Molecular engineering optimized carbon nitride photocatalyst for CO2 reduction to solar fuels. J. Sci-Adv. Mater. Dev. 2022, 7, 100483. [Google Scholar]
150.
Dimitrijevic NM, Saponjic ZV, Bartels DM, Thurnauer MC, Tiede DM, Rajh T. Revealing the nature of trapping sites in nanocrystalline titanium dioxide by selective surface.  J. Phys. Chem. B 2003, 107, 7368–7375. [Google Scholar]
151.
Meng A, Cheng B, Tan H, Fan J, Su C, Yu J. TiO2/polydopamine S-scheme heterojunction photocatalyst with enhanced CO2-reduction selectivity. Appl. Catal. B Environ. 2021, 289, 120039. [Google Scholar]
152.
Li M, Zhang S, Li L, Han J, Zhu X, Ge Q, et al. Construction of highly active and selective polydopamine modified hollow ZnO/Co3O4 p-n heterojunction catalyst for photocatalytic CO2 reduction. ACS Sustain. Chem. Eng. 2020, 8, 11465–11476. [Google Scholar]
153.
Hu J, Chen C, Yang H, Yang F, Qu J, Yang X, et al. Tailoring well-ordered, highly crystalline carbon nitride nanoarrays via molecular engineering for efficient photosynthesis of H2O2 Appl. Catal. B Environ. 2022, 317, 121723. [Google Scholar]
154.
Ma J, Peng X, Zhou Z, Yang H, Wu K, Fang Z, et al. Extended conjugation tuning carbon nitride for non-sacrificial H2O2 photosynthesis and hypoxic tumor therapy.  Angew. Chem. Int. Ed Engl. 2022, 61, e202210856. [Google Scholar]
155.
Zhao Y, Gao J, Yang Z, Li L, Cui J, Zhang P, et al. Efficient exciton dissociation in ionically interacted methyl viologen and polymeric carbon nitride for superior H2O2 photoproduction.  ACS Catal. 2023, 13, 2790–2801. [Google Scholar]
156.
Wang J, Zhao H, Larter SR, Kibria MG, Hu J. One-pot sequential cascade reaction for selective gluconic acid production from cellulose photobiorefining. Chem. Commun. 2023, 59, 3451–3454. [Google Scholar]
157.
Wei Z, Zhao S, Li W, Zhao X, Chen C, Phillips DL, et al. Artificial photosynthesis of H2O2 through reversible photoredox transformation between catechol and o-benzoquinone on polydopamine-coated CdS. ACS Catal. 2022, 12, 11436–11443. [Google Scholar]
158.
Deng Y, Liu W, Xu R, Gao R, Huang N, Zheng Y, et al. Reduction of superoxide radical intermediate by polydopamine for efficient hydrogen peroxide photosynthesis.  Angew. Chem. Int. Ed Engl. 2024, 63, e202319216. [Google Scholar]
159.
Wei Y, Jiang W, Liu Y, Bai X, Hao D, Ni BJ. Recent advances in photocatalytic nitrogen fixation and beyond. Nanoscale 2022, 14, 2990–2997. [Google Scholar]
160.
Ziegenbalg D, Zander J, Marschall R. Photocatalytic nitrogen reduction: Challenging materials with reaction engineering.  ChemPhotoChem 2021, 5, 792–807. [Google Scholar]
161.
Wang T, Liu J, Wu P, Feng C, Wang D, Hu H, et al. Direct utilization of air and water as feedstocks in the photo-driven nitrogen reduction reaction over a ternary Z-scheme SiW9Co3/PDA/BWO heterojunction. J. Mater. Chem. A 2020, 8, 16590–16598. [Google Scholar]
162.
Shi J-L, Lang X. Assembling polydopamine on TiO2 for visible light photocatalytic selective oxidation of sulfides with aerial O2 Chem. Eng. J. 2020, 392, 123632. [Google Scholar]
163.
Ghorbani F, Zamanian A, Behnamghader A, Joupari MD. A facile method to synthesize mussel-inspired polydopamine nanospheres as an active template for in situ formation of biomimetic hydroxyapatite. Mater. Sci. Eng. C 2019, 94, 729–739. [Google Scholar]
164.
Boecker M, Micheel M, Mengele AK, Neumann C, Herberger T, Marchesi D’Alvise T, et al. Rhodium-complex-functionalized and polydopamine-coated CdSe@CdS nanorods for photocatalytic NAD+ reduction.  ACS Appl. Nano Mater. 2021, 4, 12913–12919. [Google Scholar]
165.
Wang Y-M, Li W-J, Wang M-M, Zhang M, Zhang Z-H. Magnetic MoS2 efficient heterogeneous photocatalyst for the α-methoxymethylation and aminomethylation of aromatic ketones.  Catal. Sci. Technol. 2023, 13, 665–674. [Google Scholar]
166.
Xu L, Zhang SZ, Li W, Zhang ZH. Visible-light-mediated oxidative amidation of aldehydes by using magnetic CdS quantum dots as a photocatalyst.  Chem. Eur. J. 2021, 27, 5483–5491. [Google Scholar]
TOP