Next Article in Journal
Facile and Low-Cost Fabrication of ZnO/Kaolinite Composites by Modifying the Kaolinite Composition for Efficient Degradation of Methylene Blue Under Sunlight Illumination
Previous Article in Journal
Co-Catalyst-Free Al6Si2O13/Cd8.05Zn1.95S10 Nanocomposites for Visible-Light-Driven Stable H2 Evolution and DDVP Degradation
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Recent Advances in PDI-Based Heterojunction Photocatalysts for the Degradation of Organic Pollutants and Environmental Remediation

1
Department of Biological Engineering, Xianning Vocational Technical College, Xianning 437100, China
2
Hubei Province Etnan Special Agricultural Industry Technology Research Institute, Xianning 437100, China
3
School of Nuclear Technology and Chemistry & Biology, Hubei Key Laboratory of Radiation Chemistry and Functional Materials, Hubei University of Science and Technology, Xianning 437100, China
4
Hubei Provincial Key Laboratory of Green Materials for Light Industry, Hubei University of Technology, Wuhan 437100, China
*
Authors to whom correspondence should be addressed.
Catalysts 2025, 15(6), 565; https://doi.org/10.3390/catal15060565
Submission received: 6 May 2025 / Revised: 25 May 2025 / Accepted: 28 May 2025 / Published: 6 June 2025

Abstract

:
With the rapid advancement of industrialization, the adverse impacts of organic pollutants on the water environment of aquatic ecosystems have become increasingly concerning. Consequently, the development of efficient and environmentally friendly photocatalytic degradation technologies has attracted considerable research attention. Perylene diimide (PDI)-based heterojunction photocatalysts have demonstrated remarkable potential in degrading organic pollutants, attributed to their broad spectral response, high charge separation efficiency, and exceptional stability. In recent years, substantial progress has been achieved in the field of PDI-based heterojunction photocatalysts. This paper provides an in-depth review of the existing research on PDI-based heterojunction photocatalysts. Specifically, it elucidates the principles and types of heterojunction construction, as well as the design and synthesis strategies for PDI-based heterojunction photocatalysts. Furthermore, this paper provides a comprehensive summary of the latest advancements in performance optimization and catalytic mechanisms. Finally, the existing challenges and future prospects of PDI-based heterojunction photocatalytic materials are discussed, with the aim of offering innovative solutions for the purification of resource-oriented wastewater.

Graphical Abstract

1. Introduction

The rapid advancement of industrialization has resulted in the discharge of wastewater into natural waterways, thereby causing significant ecological and environmental damage [1,2,3,4,5]. These water bodies are often contaminated with a broad spectrum of organic pollutants, including dyes, pesticides, and pharmaceutical residues [6,7,8]. To mitigate the adverse effects of wastewater, numerous methods for removing organic pollutants from water have been developed. These methodologies can be broadly classified into physical, chemical, and biological approaches. Traditional dye removal techniques encompass physical methods (e.g., filtration and sedimentation), chemical methods (e.g., redox reactions and adsorption), and biological methods (e.g., microbial degradation). From the standpoint of wastewater treatment (such as adsorption or chemical precipitation), these conventional methods exhibit limitations in terms of their efficacy, cost-effectiveness, and environmental impact. Consequently, there is an increasing demand for environmentally friendly alternative technologies capable of effectively degrading organic pollutants [9,10]. Photocatalysis, as a highly promising “green” technology, demonstrates considerable potential for environmental remediation [11,12]. Semiconductor-based photocatalysis offers advantages such as high photosensitivity, environmental compatibility, and relatively low cost [13,14,15,16]. Photocatalysts can directly degrade organic pollutants into secondary products or mineralize them into carbon dioxide and water via the generation of reactive oxygen species (ROS), thereby enabling the treatment of refractory organic compounds [17,18]. However, most semiconductor photocatalysts encounter bottleneck issues, such as low separation efficiency of photogenerated carriers, limited surface area, narrow absorption spectrum, and inadequate cycling stability, which significantly impede their photocatalytic activity and practical application [19,20,21,22,23]. Therefore, designing and developing novel broad-spectrum photocatalysts plays a crucial role in promoting sustainable societal development.

2. Principles and Types of Heterojunctions

Photocatalytic heterojunctions are considered a highly effective and promising strategy for enhancing the separation efficiency of photogenerated charges and improving the oxidation capacity of holes [24,25,26]. In recent years, the development of low-cost heterojunctions with high photogenerated charge separation efficiency has become a significant trend in the field. Efficient charge separation is critically promoted by photocatalysts featuring heterojunction structures, thereby enabling their extensive application in wastewater treatment and related areas [27,28,29,30]. Semiconductor heterojunctions are constructed by combining two semiconductors with appropriately aligned conduction band (CB) and valence band (VB) potentials. This approach effectively accelerates the transfer and separation of photogenerated carriers, while simultaneously prolonging their lifetimes [31,32,33,34].
Researchers categorize various types of heterojunctions, including type-II, Z-scheme, and S-scheme heterojunctions, based on the relative positions of their band gaps [35,36,37]. The enhanced photocatalytic performance, characterized by a narrowed band gap, improved separation of photoinduced charge carriers, and reduced recombination rate, is attributed to the synergistic interaction between semiconductors [38,39]. However, in type-II heterojunctions, interfacial charge transfer is hindered by the inherent Coulomb attraction and electrostatic repulsion at the interface, leading to limitations in subsequent redox reactions and a negative impact on photocatalytic performance [40,41]. Compared to traditional heterojunctions, the separation of electron (e)–hole (h+) pairs within different semiconductors can be effectively promoted by Z-scheme heterojunctions [42,43]. Nevertheless, issues such as the inefficient photogenerated charge transfer, weak redox capabilities of carriers, and inadequate light absorption characteristics have been identified as factors that impede the improvement of photocatalytic performance in this system [44]. Additionally, due to thermodynamic and kinetic limitations, the efficiency of traditional Z-scheme heterojunction photocatalysts is significantly compromised [45,46,47,48].
To address these challenges, the S-scheme (step-type) heterojunction theory was first proposed by Yu et al. based on direct Z-scheme heterojunctions, aiming to resolve the ambiguities between traditional type-II and Z-scheme heterojunctions and overcome their inherent limitations [49]. Typically, an S-scheme heterojunction is formed by coupling two semiconductors—one with a lower Fermi level functioning as an oxidative photocatalyst and the other with a higher Fermi level acting as a reductive photocatalyst [50,51,52,53]. In S-scheme heterojunctions, interfacial charge transfer is facilitated by the built-in electric field between semiconductors and the band bending at the interface, leading to a significant enhancement in photocatalytic activity [54,55].

3. Design and Construction of PDI-Based Heterojunction Photocatalysts

3.1. Properties of PDI

PDI is a derivative of polycyclic aromatic hydrocarbons [56,57], characterized by a molecular structure with a central perylene core connected to two imide groups (-CONHCO-) on either side [58]. PDI is known as a deep red organic dye [59] that is inherently non-toxic [60,61]. In its natural form, PDI exhibits a relatively low specific surface area; however, this limitation can be overcome through self-assembly processes, forming structures such as supramolecular nanofibers, layered assemblies, and three-dimensional frameworks [62,63,64]. Despite its poor water solubility caused by π–π stacking interactions between perylene units [65,66], researchers have successfully enhanced its solubility by introducing hydrophilic groups at specific positions, including the bay, imide, or ortho regions. For instance, ionic groups such as cationic ammonium salts [67,68,69], anionic carboxylic acids [70], sulfonic acids [71], and phosphonic acids [72] have been directly incorporated into the perylene core to produce water-soluble PDI derivatives. Additionally, nonionic substituents with multiple polar groups, such as polyethylene glycol (PEG), polyglycerol (PG) dendrimers, and dendritic carbohydrate derivatives, have been attached to the bay or imide sites of PDI [73]. The bandgap energy (Eg) of PDI typically ranges from approximately 1.5 to 2.5 eV, enabling it to absorb visible light with wavelengths up to around 750 nm [74]. For instance, Li et al. [75] reported that the valence band potential of PDI is +1.81 eV (vs. NHE), while the conduction band potential is –0.36 eV (vs. NHE), which meets the requirements for simultaneously photocatalytically decomposing water into hydrogen and oxygen, reducing carbon dioxide, and removing pollutants. PDI and its derivatives, as representative n-type semiconductors, exhibit superior electronic and optical properties and have been extensively investigated in the fields of solar cells and sensors [76,77]. Self-assembled supramolecular materials based on PDI demonstrate significant potential in photocatalytic applications owing to their exceptional charge transport characteristics, remarkable photothermal stability, and strong electron affinity, and have been widely applied in the fields of pollutant degradation, hydrogen evolution, oxygen evolution, and carbon dioxide reduction, among others [78,79,80]. Nevertheless, the practical application of PDI is hindered by issues such as the rapid recombination of photogenerated carriers, poor solubility, and insufficient adsorption capacity [81].

3.2. The Role of PDI in Heterojunctions and Construction Strategies

3.2.1. The Role of PDI in Heterojunctions

At present, the photocatalytic modification of perylene diimide primarily focuses on structural regulation and the development of composite materials [82,83,84]. Among these approaches, the construction of composite materials leverages the inherent advantages of PDI while mitigating its limitations, such as self-aggregation, thereby significantly enhancing the photocatalytic performance of the catalyst. The formation of heterojunctions is widely recognized as an effective strategy to improve the separation efficiency of photogenerated carriers [85,86]. In recent years, researchers have successfully developed heterojunction-based photocatalysts by combining PDI with other semiconductors, including rGO/PDI [87], Bi2WO6/PDI [88], PDI/g-C3N4 [89], and PANI/PDI [90]. These composites have demonstrated superior photocatalytic activity. The integration of PDI into heterojunction structures with other catalysts represents a highly promising approach to enhance photocatalytic performance, improve light responsiveness, and accelerate the transfer and separation of charge carriers.

3.2.2. The Interface Coupling of PDI Functional Groups with Heterogeneous Materials

PDI molecules can form coordination bonds with metal active sites on the surface of metal oxides. For example, Wang et al. [91] successfully connected PDI to MIL-125(Ti)-NH2 heterojunctions using an acid-catalyzed method based on the blocked Lewis pair strategy. In this process, the amino groups in MIL-125(Ti)-NH2 acted as nano-scaffolds via amidation reactions, enabling their interaction with PDI. Specifically, the unsaturated Ti-O clusters functioned as Lewis acid sites to accept photogenerated electrons, while PDI served as Lewis base sites for capturing these electrons. Furthermore, the edge substituents of PDI can establish stable coupling structures with other materials through covalent or hydrogen bonding. As an illustration, Chen et al. [92] introduced carboxyl-terminated organic supramolecular PDI during the synthesis of MIL-53(Fe) crystals, thereby constructing a PDI/MIL-53(Fe) (PM) Z-scheme heterojunction. In this structure, PDI nanofibers were uniformly dispersed and firmly anchored within MIL-53(Fe) via covalent bonds. In addition, composite materials can be prepared by imidization reactions involving the PDI precursor, perylene tetracarboxylic dianhydride (PTCDA). For instance, Wang et al. [93] fabricated a PDI/g-C3N4 heterojunction through a one-step imidization reaction between PTCDA and g-C3N4. During this process, PTCDA and g-C3N4 formed a hybrid structure via O=C-N-C=O covalent bonds at the heterojunction interface, significantly enhancing the material’s interfacial interactions and charge transport properties. In addition, PDI can form heterojunctions with other materials through π–π stacking interactions, thereby enhancing interfacial charge transfer and material performance. PDI molecules can self-assemble into supramolecular materials through π–π stacking interactions. Larger π–π stacking enhances π–electron delocalization and increases the overlap of π–electron clouds, which is beneficial for carrier migration and separation [94]. Based on the π–π stacking interactions in PDI supramolecular materials, researchers have attempted to combine PDI with other π–conjugated organic materials to construct advanced π–π composite systems. Leveraging the π–π stacking interactions between PDI molecules and organic materials, highly stable self-assembled composite materials can be obtained [87]. For instance, Dai et al. [90] constructed a heterojunction with well-matched energy levels through π–π interactions between PANI and PDI. This heterojunction exhibits enhanced π conjugation and stronger electron delocalization, which have been proven to facilitate carrier migration and separation. As a result, the photocatalytic performance and stability of the material were significantly improved.

3.2.3. Preparation Methods for PDI-Based Composite Photocatalysts

Various methods are currently employed for the preparation of PDI-based photocatalysts, including electrostatic adsorption, π–π stacking self-assembly, water bath heating, hydrothermal synthesis, and electrochemical approaches. The primary objective of optimizing these preparation methods is to fine-tune the band structure, particle size, specific surface area, morphology, and crystal planes of the photocatalyst, thereby improving its overall photocatalytic performance.
Electrostatic Adsorption Method: Miao et al. [95] successfully fabricated PDI@AuNPs (nanogold rod) composite materials via the electrostatic adsorption method. The visible light-induced degradation rate of phenol by PDI@AuNPs is 1.7 times higher than that of PDI nanowires. This enhanced photocatalytic activity of PDI@AuNPs can be attributed to the surface plasmon resonance (SPR) effect of AuNPs. Specifically, the SPR effect of AuNPs and the resonance energy transfer between AuNPs and PDI facilitate the utilization of visible light. Moreover, the lower Fermi level of Au quantum dots promotes the efficient transfer of photogenerated carriers from PDI to AuNPs. Yang et al. [96] synthesized GQD (carbon quantum dot)/PDI photocatalysts through electrostatic interactions. Under visible light irradiation, the rate constant for phenol photocatalytic degradation by GQD/PDI-14% (0.018 min−1) was 4.73 times higher than that of nano-PDI. The superior photocatalytic activity of GQD/PDI is primarily due to the π–π interactions between GQDs and PDI, which enhance electron delocalization. Additionally, the quantum confinement effect of GQDs facilitates electron transfer from GQDs to PDI, shifting the conduction band of PDI to a more negative position and thereby enhancing its reduction capability. Wang et al. [97] developed an efficient Co-NG/PDI photocatalyst using an in situ electrostatic adsorption method. The Co-NG/PDI composite exhibited superior photocatalytic degradation performance for bisphenol A compared to pure Co-NG and self-assembled PDI. Notably, the 7% Co-NG/PDI composite demonstrated the best photocatalytic performance, with an apparent rate constant 6.64 times higher than self-assembled PDI and 7.6 times higher than Co-NG.
π–π Stacking Self-Assembly Method: Wei et al. [98] prepared PTCDI–C60 composite materials by self-assembling PTCDI (perylene diimide) and C60 (carbon 60) through π–π stacking interactions, enabling rapid transfer of photogenerated carriers. The π–π interactions effectively lowered the valence band position and narrowed the band gap, thereby enhancing redox capabilities and broad-spectrum responsiveness. The PTCDI–C60 organic photocatalyst achieved a phenol degradation rate constant of 0.216 h−1, which is 8.24 times higher than that of PTCDI alone. Furthermore, the presence of C60 improved the stability of the composite material and reduced the accumulation of negative charges. Yang et al. [87] constructed 3D(three-dimensional) rGO/PDI composite materials via a simple water bath heating method. Self-assembled PDI nanofibers and rGO were effectively combined through non-covalent π–π stacking interactions. These π–π interactions enhanced long-range π–electron delocalization and electronic coupling effects, significantly improving carrier mobility and the separation efficiency of photogenerated electron–hole pairs. As a result, the 3D rGO/PDI composite material exhibited superior photocatalytic activity compared to pure PDI.
Water Bath Heating Method: Zhang et al. [88] successfully fabricated a Bi2WO6/PDI composite material by combining a Bi2WO6 photocatalyst with PDI through the water bath heating method. This composite material was utilized in the photocatalytic oxidation of phenol under simulated visible light irradiation. The results demonstrated that, under visible light irradiation, the degradation efficiency of phenol by the Bi2WO6/PDI composite material was significantly higher than that of the PDI self-assembled body and Bi2WO6 alone. In the photocatalytic process of the Bi2WO6/PDI composite material under visible light irradiation, the energy level alignment between the conduction and valence bands of Bi2WO6 and PDI facilitated the effective separation of photogenerated carriers.
Hydrothermal Method: Zhang et al. [99] synthesized PDI/TiO2 photocatalysts via the hydrothermal method. The effects of introducing different structural forms of PDI into TiO2 were evaluated by measuring the photocatalytic degradation rate of MB (methylene blue). The photocatalytic activities of the PDI-1/TiO2 and PDI-2/TiO2 catalysts were superior to that of PDI-3/TiO2. This was attributed to the large surface areas of PDI-1 nanorods and PDI-2 nanobelts, which expanded one-dimensional photogenerated carrier channels, thereby promoting electron transfer to the TiO2 surface and enhancing the photocatalytic activity of the composite material. Notably, the PDI-1/TiO2 composite material exhibited the highest photocatalytic activity, maintaining 86.4% of its initial activity after four repeated uses. The extended π–π stacking of self-assembled PDI-1 and the strong interaction between self-assembled PDI-1 and TiO2 played a critical role in accelerating charge transfer and reducing the recombination of photogenerated electron–hole pairs.
Electrochemical Co-deposition Method: Sheng et al. [100] prepared PDI/rGO composite materials using a simple electrodeposition–impregnation method. The removal rate of MB by the PDI/rGO catalyst in photoelectrocatalysis was 8.2 times higher than that of single photocatalysis and 4.5 times higher than that of single electrocatalysis. In dynamic photocatalytic degradation experiments, the degradation rate of MB by the PDI/rGO catalyst reached 88.5%, significantly surpassing that of PDI alone (56.3%). The PDI/rGO composite material exhibited exceptional stability, with no noticeable decrease in activity over 55 h of continuous photocatalytic degradation.

4. Application of PDI-Based Heterojunctions in Pollutant Degradation

4.1. Antibiotics and Drug Residues

The long-term environmental impacts of antibiotics and drug residues on water systems have emerged as a global environmental challenge. Their complex chemical structures and inherent resistance to degradation render traditional treatment technologies insufficient for achieving efficient removal. In recent years, PDI-based heterojunction photocatalytic technology has demonstrated significant advantages in antibiotic degradation due to its unique broad-spectrum light response capability and superior charge separation efficiency. Researchers have innovatively designed type-II, Z-scheme, and S-scheme heterojunction architectures by efficiently coupling PDI with semiconductors such as g-C3N4, metal–organic frameworks, and metal oxides. These designs not only broaden the light absorption range, but also markedly enhance photocatalytic activity through interfacial charge transfer mechanisms. This section systematically reviews the latest advancements in PDI-based heterojunctions for the degradation of antibiotics and drug residues, elucidates the underlying mechanisms driving performance enhancement, and explores potential directions for future technological optimization.
Wu et al. [101] successfully synthesized CNPDI (g-C3N4/PDI) type-II heterojunction catalysts via in situ self-assembly of PDI and g-C3N4 at controlled mass ratios. Under visible light irradiation for 60 min with potassium persulfate (PMS) activation, the CNPDI-7 photocatalyst achieved a degradation efficiency of 83.6% for tetracycline (TC), as shown in Table 1. This enhanced photocatalytic performance was primarily attributed to the formation of an effective type-II heterojunction between self-assembled PDI and g-C3N4, which significantly promoted the separation of electron–hole pairs and optimized charge density redistribution at the interface. In another study, Li et al. [102] developed a novel g-C3N4/PDI@NH2-MIL-53(Fe) (CPM) type-II heterojunction through a combination of thermal polymerization, surface growth technology, and solvothermal synthesis. Structural characterization using scanning electron microscopy (SEM) and transmission electron microscopy (TEM) revealed that spindle-shaped NH2-MIL-53(Fe) particles were in intimate contact with the g-C3N4/PDI surface (Figure 1a). High-resolution TEM analysis of the CPM-2 interface (highlighted in Figure 1a) confirmed the establishment of efficient interfacial electron transfer pathways. The CPM heterojunction exhibited remarkable photocatalytic activity for the degradation of various aqueous organic pollutants under visible light irradiation in the presence of H2O2, achieving removal efficiencies of 90.0% for TC within 1 h, 78.0% for carbamazepine (CBZ) within 2.5 h, and complete degradation (100.0%) of both bisphenol A (BPA) and p-nitrophenol (PNP) within 10 and 30 min, respectively. This superior photocatalytic performance was ascribed to the formation of an optimized type-II heterojunction, facilitated by the close interfacial contact and well-aligned band structures between g-C3N4/PDI and NH2-MIL-53(Fe), which enhanced charge separation efficiency and accelerated photodegradation kinetics. The photocatalytic degradation of organic pollutants in the CPM-2/visible light/H2O2 system is governed by a type-II heterojunction-mediated electron transfer mechanism. As illustrated in Figure 1b, this process involves the generation and subsequent migration of electron–hole pairs within the CPM-2 composite. Upon exposure to visible light irradiation, electron–hole pairs are initially formed on the CPM material. Subsequently, electrons from the conduction band of g-C3N4/PDI migrate to the conduction band of NH2-MIL-53(Fe), while holes from the valence band of NH2-MIL-53(Fe) transfer to the valence band of g-C3N4PDI. This charge transfer mechanism effectively suppresses electron–hole recombination and enhances the free electron density in the conduction band of NH2-MIL-53(Fe). The Fe2+ ions generated via the reduction of Fe3+ by these electrons subsequently catalyze the decomposition of H2O2 to produce hydroxyl radicals (•OH), which play a critical role in the oxidative degradation of organic pollutants. The photocatalytic efficiency of this system is further enhanced by the high electron affinity of PDI, which positively modulates the redox potential of g-C3N4, thereby optimizing the band structure alignment between g-C3N4/PDI and NH2-MIL-53(Fe). This optimized alignment facilitates the formation of an efficient type-II heterojunction, significantly improving the separation and transfer of photogenerated charge carriers. Notably, the activation of H2O2 by Fe2+ in NH2-MIL-53(Fe) to generate •OH demonstrates substantially higher efficiency compared to the direct production of •OH through electron transfer from g-C3N4/PDI under visible light irradiation. Therefore, the establishment of an effective heterojunction between g-C3N4/PDI and NH2-MIL-53(Fe) markedly enhances the Fenton-like photocatalytic degradation of organic pollutants under visible light conditions. Wu et al. [103] successfully achieved the modification of PDI on the surface of NH2-MIL-88B(Fe, Mn) (FM88B) via a water bath heating process, leading to the formation of a type-II heterojunction between PDI and FM88B (PDI/FM88B). This modification was enabled by the formation of an amide bond between PDI and FM88B, as shown in Figure 1c. The amide bond not only enhanced the structural stability of the heterogeneous system but also provided an efficient pathway for interfacial electron transfer. In the Fenton system, under visible light irradiation for 30 min, the degradation efficiency of TC by 6.0% PDI/FM88B reached 89.0%. The improved photocatalytic performance was mainly ascribed to the combined effects of photocatalysis and the Fenton reaction. Moreover, the type-II heterojunction enabled the directional movement of photoinduced electrons from PDI (electron donor) to FM88B (electron acceptor). The built-in electric field (BIEF) existing at the PDI/FM88B interface significantly contributed to the separation of photoinduced e–h+ pairs, which played a vital role in the TC degradation process within this photo–Fenton system, as depicted in Figure 1d. When exposed to visible light, the PDI/FM88B heterostructure, with its remarkable light absorption capacity, guided the transfer of photoinduced electrons from PDI to FM88B. The BIEF related to the type-II heterojunction further improved the separation and directional movement of photoinduced charge carriers. Furthermore, the amide bond served as a nanoscale rapid pathway, decreasing the transfer distance and energy barrier for photoinduced electrons moving from PDI to FM88B, thus substantially enhancing the efficiency.
Table 1. The photocatalytic degradation performance of antibiotics by a PDI-based heterojunction.
Table 1. The photocatalytic degradation performance of antibiotics by a PDI-based heterojunction.
PhotocatalystSynthesis MethodAntibioticsLight SourceTime
(min)
Efficiency
(%)
TypeReference
CNPDIIn situ self-assemblyTCVisible6083.6IIWu et al. [101]
CPMSurface growth, solvothermalTCVisible6090.0IILi et al. [102]
PDI/FM88BWater bath heatingTCVisible3089.0IIWu et al. [103]
I-PDI/PEDOT-MSurface self-corrosion- assisted rapid spin- coatingTCVisible6073.7IILu et al. [104]
PDI/BiOCl-BiPO4Ultrasonic, hydrothermalTCSimulated sunlight15081.0ZZhuang
et al. [105]
PDI/FePcIn situ self-assemblyTCVisible6078.6ZShi et al. [106]
WO3@Cu@PDIPhoto-deposition,
water bath heating
TCVisible1575.0ZZeng et al. [107]
PDI/WO3/α-Fe2O3PLALTC254 nm15094.2ZMao et al. [108]
PDI/MIL-53(Fe)Solvent thermalTCVisible3094.1ZChen et al. [92]
TMOPElectrospinningTCSimulated sunlight8091.2ZSun et al. [109]
Ag3PO4/PDIsmSelf-assemblyTCVisible882.8ZCai et al. [110]
PANI/PDIIn situ growthTCVisible12070.0ZDai et al. [90]
PDI-Ala/S-C3N4In situ self-assemblyTCVisible9090.0SLi et al. [111]
IM-NSH-PMImprintTCSimulated sunlight6071.9SLu et al. [112]
PDIs/C, N, S-CeO2Oil bath heatingTCVisible3080.1SJing et al. [113]
PDIs/Fe2O3@CHydrothermal calcination,
oil bath heating
TCVisible878.9SJing et al. [114]
PDI/ZnFe2O4UltrasonicTCVisible6066.7SXu et al. [115]
Ag/PCN/UPDISelf-assembly, photoreductionOTCVisible15097.5SXiao et al. [116]
g-C3N4/PDI/Co-FeCo-precipitationDOXVisible6096.1ZLi et al. [117]
PDI-Urea/BiOBrSolvent thermal, in situ growthOFLOVisible15093.0SWang et al. [118]
Bis-PDI-T@TiO2Double-solvent phase transferCBZVisible30100.0ZYang et al. [119]
β-PDI/MIL-101(Fe)GrindingSMXVisible699.7ZJia et al. [120]
MIL-101(Fe)-NH2/PDIGrindingSMXVisible599.2ZJia et al. [121]
Figure 1. (a) SEM images of CPM-2 (inset: enlarged image of the yellow frame). (b) Schematic illustration for the possible photocatalytic mechanism of the CPM-2 composite under visible light irradiation [102]. (c) Ammoniation reaction between PDI and FM88B. (d) Degradation mechanism of TC in the system of visible light/6% PDI/FM88B/H2O2 [103]. (e) TEM images of WO3@Cu@PDI composite; (f) Z-scheme electron transfer mechanism [107].
Figure 1. (a) SEM images of CPM-2 (inset: enlarged image of the yellow frame). (b) Schematic illustration for the possible photocatalytic mechanism of the CPM-2 composite under visible light irradiation [102]. (c) Ammoniation reaction between PDI and FM88B. (d) Degradation mechanism of TC in the system of visible light/6% PDI/FM88B/H2O2 [103]. (e) TEM images of WO3@Cu@PDI composite; (f) Z-scheme electron transfer mechanism [107].
Catalysts 15 00565 g001
In a related study, Lu et al. [104] developed an imprinted PDI/PEDOT type-II heterojunction photocatalyst anchoring film (I-PDI/PEDOT-M) using an N-methylpy- rrolidone (NMP)-induced surface self-corrosion-assisted rapid spin-coating method. Under visible light irradiation for 1 h, the degradation efficiencies of I-PDI/PEDOT for TC and ciprofloxacin (CIP) were 73.7% and 5.0%, respectively. Notably, the degradation efficiency of TC by I-PDI/PEDOT was 14.65 times higher than that of CIP. The construction of the type-II heterojunction between PDI and PEDOT effectively promoted the rapid separation of photogenerated electrons and holes, thereby maintaining the superior photocatalytic activity of I-PDI/PEDOT-M. Recent advancements in Z-scheme heterojunction photocatalysts have demonstrated significant improvements in photocatalytic performance through innovative material design and synthesis strategies. Zhuang et al. [105] developed a ternary Z-scheme PDI/BiOCl-BiPO4 composite via ultrasonic dispersion and hydrothermal methods. The experimental results revealed that under 150 min of simulated solar irradiation, the PDI (5.0%)/BiOCl-BiPO4 composite achieved remarkable degradation efficiencies of 98.0% for rhodamine B (RhB) and 81.0% for TC. The enhanced photocatalytic activity was primarily attributed to the Z-scheme heterojunction structure, which facilitated efficient separation and migration of photogenerated electron–hole pairs. In a parallel development, Shi et al. [106] engineered a novel PDI/iron phthalocyanine (FePc) Z-scheme heterojunction through a self-assembly approach, leveraging strong π–π interactions. This configuration demonstrated a tetracycline hydrochloride removal efficiency of 78.6% after 60 min of visible light exposure. The superior photocatalytic performance was ascribed to the robust π–π interactions between PDI and FePc, which effectively minimized the interlayer spacing of the supramolecular structure and enhanced charge carrier separation and transfer within the built-in electric field. A further innovation was demonstrated by Zeng et al. [107], who constructed a Z-scheme heterojunction by photo-depositing Cu onto WO3, followed by coupling with PDI through a water bath method. The resulting 30% WO3@Cu@PDI photocatalyst exhibited a 75.0% tetracycline removal efficiency within 15 min of visible light irradiation. Structural characterization (Figure 1e) revealed a uniform distribution of WO3 nanoparticles on the PDI supramolecular surface, with dispersed Cu nanoparticles. The Z-scheme architecture enabled efficient electron transfer from WO3 to PDI’s donor level, facilitating rapid recombination with h+. This process resulted in electron accumulation on PDI’s conduction band and hole accumulation on WO3’s valence band, thereby enhancing redox activity. The electron transfer mechanism (Figure 1f) involved Cu-mediated electron transfer from WO3’s conduction band to PDI’s valence band, followed by the generation of •O2 (superoxide radical) from PDI’s conduction band, confirming •O2 as the primary reactive species in the photocatalytic process. Mao et al. [108] successfully synthesized a perylene diimide/tungsten trioxide/hematite (PDI/WO3/α-Fe2O3, PWF) composite photocatalyst with a dual Z-scheme heterojunction via the pulsed laser ablation in liquid (PLAL) technique. Photocatalytic performance evaluation revealed that the PWF composite achieved a 94.2% TC removal efficiency under 180 min of irradiation using a 15 W low-pressure mercury lamp (λ = 254 nm). The enhanced photocatalytic activity can be primarily attributed to two key factors: (1) the increased number of adsorption sites and strengthened surface charge interactions, which facilitated rapid TC adsorption; and (2) the engineered dual Z-scheme heterojunction, which optimized the band structure, thereby improving photovoltaic conversion efficiency, promoting efficient charge carrier separation, and ultimately enhancing photocatalytic performance. In a separate study, Chen et al. [92] developed a perylene diimide/metal–organic framework (PDI/MIL-53(Fe), PM) Z-scheme heterojunction composite photocatalyst through solvothermal synthesis. The 5PM composite demonstrated superior photocatalytic activity, achieving a 94.1% tetracycline degradation efficiency within 30 min of visible light irradiation. This performance represents a 4-fold enhancement compared to pristine PDI and a 33-fold improvement over MIL-53(Fe) alone. The exceptional photocatalytic efficiency can be ascribed to the synergistic effects of the Z-scheme heterojunction, which facilitates effective charge separation, and the intrinsic semiconductor properties that provide strong redox potential.
Sun et al. [109] successfully developed a novel self-supported tricolor microfiber directional heterojunction photocatalyst, designated as TMOP, via a three-axis parallel electrospinning technique. The TMOP structure consists of three distinct layers: [g-C3N4/poly(methyl methacrylate) (PMMA)]//[TiO2/polyaniline (PANI)/PMMA]//[self- assembled 3,4,9,10-perylene tetracarboxylic diimide (PDI)/PMMA]. Under simulated solar light irradiation, the optimized TMOP exhibited exceptional photocatalytic degradation efficiencies: 91.2% for tetracycline hydrochloride (within 80 min), 89.0% for ciprofloxacin (CIP) (within 90 min), 77.6% for chlortetracycline (CTC) (within 150 min), 69.5% for levofloxacin (within 150 min), and 92.5% for methylene blue (within 50 min). The enhanced photocatalytic performance can be primarily attributed to three critical factors: (1) the unique morphology and heterojunction structure of the tricolor microfibers, which significantly increased active sites; (2) the broadband spectral characteristics of PDI, which expanded the solar spectrum absorption range; and (3) the synergistic effect between conductive polyaniline and the one-dimensional double Z-scheme directional heterojunction structure, which facilitated directional and rapid carrier transport through multiple pathways. This synergistic effect substantially reduced carrier recombination probability while maintaining strong redox capability. Furthermore, the self-supporting nature of TMOP ensured excellent recoverability and practical applicability. In a related study, Cai et al. [110] fabricated an Ag3PO4/PDI supramolecular Z-scheme photocatalyst using a two-step self-assembly strategy. As illustrated in Figure 2a, the synthesis process involved the initial formation of PDIsm (PDI supramolecular) rods through self-assembly, stabilized by π–π stacking and hydrogen bonding interactions. The polar nature of PDI molecules established an intramolecular electric field from the perylene core to the carboxyl groups, enabling the adsorption of silver ions (Ag+) onto the carboxyl groups of PDIsm rods. The subsequent reaction with NaH2PO4 resulted in the in situ loading of Ag3PO4 nanoparticles on the PDIsm surface. This spatial configuration facilitated the formation of an efficient Z-scheme system, achieving a degradation efficiency of 82.8% for tetracycline hydrochloride under visible light irradiation within eight minutes. The detailed photocatalytic mechanism and stability enhancement are further elaborated in the subsequent section. Comparative analysis of the photoluminescence (PL) spectra between the Ag3PO4 and Ag3PO4/PDIsm composites, as shown in Figure 2b, revealed a significant reduction in PL intensity for the Ag3PO4/PDIsm system. This observation indicates a marked improvement in charge carrier separation efficiency. Photocurrent response measurements (Figure 2c) further corroborated this finding, demonstrating that the Ag3PO4/PDIsm composite exhibited superior photocurrent density under visible light irradiation compared to pristine Ag3PO4 and PDIsm alone, thereby confirming more efficient migration of photogenerated charge carriers to the electrode interface. Electrochemical impedance spectroscopy (EIS) analysis, as depicted in the Nyquist plot (Figure 2d), revealed that the Ag3PO4/PDIsm composite possessed the lowest charge transfer resistance among the three samples, suggesting enhanced electrical conductivity and accelerated electron transfer kinetics following the incorporation of PDIsm, which collectively contribute to improved charge carrier separation efficiency. These experimental findings collectively demonstrate that the integration of PDIsm with Ag3PO4 significantly enhances charge carrier separation efficiency, thereby facilitating photocatalytic processes. The superior photocatalytic performance of the Ag3PO4/PDI composite can be mechanistically attributed to three primary factors: (1) The implementation of a Z-scheme charge transfer mechanism facilitates efficient spatial separation of charge carriers, wherein the negative conduction band potential of PDI promotes the reduction of molecular oxygen to generate •O2. (2) The inherent strong oxidative capacity of Ag3PO4 is preserved, enabling direct oxidation of tetracycline hydrochloride. (3) The establishment of an internal electric field effectively inhibits the reduction of Ag+, mitigates photochemical corrosion, and enhances the overall photostability of the composite system. Through in situ growth technology, Dai et al. [90] successfully fabricated a 3D polyaniline/perylene diimide (PANI/PDI) Z-scheme heterojunction photocatalyst. The experimental results demonstrated that after two hours of visible light irradiation, the 20% PANI/PDI composite material achieved a tetracycline degradation efficiency of 70.0%. The exceptional photocatalytic performance and structural stability of this system can be primarily attributed to three critical factors: first, the incorporation of the polyaniline polymer framework reinforced the PDI hydrogel structure, thereby enhancing the durability of the catalytic system; second, the three-dimensional network structure not only provided abundant active sites but also optimized the medium transport channels; and finally, the π–π interaction between polyaniline and perylene diimide formed an extensive delocalized π–electron conjugated system and constructed precisely aligned heterojunction structures, significantly improving the migration and separation efficiency of the photogenerated carriers. In another study, Li et al. [111] successfully developed an S-scheme heterojunction photocatalyst composed of the organic semiconductor PDI-Ala (N,N’-dipropionyl-perylene-3,4,9,10-tetracarboxylic diimide) and sulfur-doped g-C3N4 (S-C3N4) via an in situ self-assembly approach. Under visible light irradiation, the 30% PDI-Ala/S-C3N4 composite material achieved a TC removal efficiency of up to 90.0% within 90 min. This S-scheme heterojunction features ideal band alignment and robust interfacial bonding, which not only accelerates intermolecular electron transfer but also broadens the visible light absorption range of the heterojunction, thereby significantly enhancing its photocatalytic activity. Lu et al. [112] developed a novel photocatalytic membrane (IM-NSH-PM) through the combination of molecular imprinting technology and dopamine-assisted surface modification. Upon exposure to simulated sunlight for 60 min, the non-metallic S-scheme heterojunction photocatalytic membrane (IM-NSH-PM) demonstrated photocatalytic degradation efficiencies of 71.9% for TC and 25.5% for CIP. The membrane’s stability was ensured by the strong interfacial bonding between the polyvinylidene fluoride (PVDF) substrate and the imprinted non-metal S-scheme heterojunction (IM-NSH), achieved through dopamine cross-linking and polymerization. The construction of an S-scheme heterojunction between PDI and g-C3N4 significantly reduced electron–hole pair recombination, thereby boosting the photocatalytic activity. The proposed reaction mechanism of IM-NSH-PM is depicted in Figure 2e. The membrane’s selective TC adsorption capability stems from the synergistic combination of imprinted cavities and antibiotic molecular diversity. When TC molecules are specifically captured by the imprinted cavities, a series of redox reactions are triggered under simulated sunlight. The S-scheme heterojunction mechanism prevents e recombination on PDI and h+ recombination on g-C3N4. Due to the lower LUMO (lowest unoccupied molecular orbital) level of g-C3N4 compared to the standard reduction potential (E0), electrons react with O2 to produce •O2, which then convert to •OH. These reactive species, along with the holes in PDI’s HOMO (highest occupied molecular orbital), actively participate in pollutant degradation. The three-dimensional structural matching between TC and imprinted cavities promotes selective interaction, leading to the complete mineralization of TC into CO2, H2O, and other small molecules through radical-mediated reactions. In a separate study, Jing et al. [113] fabricated a 10% PDI/C, N, and S-CeO2 S-scheme heterojunction by incorporating organic supramolecular self-assembled PDI (perylene diimide) semiconductors onto a three-dimensional metal oxide framework (C, N, and S-CeO2) using oil bath heating. The material’s photocatalytic performance was assessed through organic pollutant degradation under visible light. Remarkably, the heterojunction achieved 80.1% TC degradation within 30 min of visible light exposure, with a reaction rate constant 15.83 times greater than pure C, N, and S-CeO2. Upon 60 min of visible light irradiation, the 10% PDI/C, N, and S-CeO2 composite material achieved degradation efficiencies of 72.9%, 83.5%, 44.9%, and 56.5% for CTC, CIP, RhB, and methylene blue, respectively. This enhanced performance resulted from optimized S-scheme band alignment, multidimensional charge transfer pathways, improved visible light absorption, and superior photothermal properties. In a subsequent study, Jing et al. [114] developed a novel composite photocatalytic material, a perylene diimide/iron oxide@carbon (PDI/Fe2O3@C) S-scheme heterojunction, through a facile hydrothermal calcination approach combined with an oil bath process. Within the PMS activation system, the 20% PDI/Fe2O3@C composite exhibited remarkable tetracycline degradation efficiency, reaching 78.9% within eight minutes of visible light irradiation. Upon 12 min of visible light irradiation, the 20% PDI/Fe2O3@C composite material achieved degradation efficiencies of 71.59%, 60.93%, and 45.61% for CTC, CIP, and sulfadiazine (SDZ), respectively. The enhanced photocatalytic performance was primarily ascribed to the synergistic integration of PDIs and Fe2O3@C, which effectively mitigated PDIs’ aggregation during self-assembly, increased the specific surface area of Fe2O3@C, exposed additional active sites, and facilitated the activation of the Fe2+/Fe³+ redox cycle in PMS. Furthermore, the hybrid organic–inorganic interface promoted efficient photogenerated electron transfer and electron–hole pair separation, establishing a novel S-scheme electron transport pathway that significantly augmented PMS activation efficiency. Xu et al. [115] employed an ultrasonic-assisted approach to fabricate a one-dimensional (1D) S-scheme heterojunction composed of PDI and ZnFe2O4. The composite’s photocatalytic efficiency was thoroughly assessed under visible light exposure for one hour, demonstrating a tetracycline degradation rate of 66.7%. This outcome indicates a remarkable improvement, surpassing the performance of pure PDI by 9.18-fold (7.3%) and ZnFe2O4 by 9.73-fold (6.9%). The enhanced photocatalytic activity is mainly due to the establishment of an S-scheme heterojunction between PDI and ZnFe2O4, which efficiently separates photogenerated electron–hole pairs and enhances redox potential. Furthermore, the rod-shaped structure of PDI and the nanoscale dimensions of ZnFe2O4 work in concert to promote charge carrier movement to the surface, thereby improving the overall photocatalytic efficacy.
In a related study, Xiao et al. [116] developed a ternary S-scheme heterojunction photocatalyst (Ag/PCN/UPDI) through a multi-step synthesis process. The fabrication involved electrostatic self-assembly of urea-modified perylene diimide (UPDI) with graphitic carbon nitride (PCN), followed by surface modification with silver nanoparticles (Ag) via photoreduction. The preparation methodology is schematically illustrated in Figure 3a. The photocatalytic efficiency of Ag/PCN/UPDI was comprehensively assessed through the degradation of oxytetracycline (OTC) under visible light irradiation for 150 min, achieving a remarkable degradation efficiency of 97.5%. The underlying photocatalytic mechanism for organic chloride degradation is comprehensively depicted in Figure 3b,c. The superior photocatalytic performance of Ag/PCN/UPDI is primarily ascribed to the integration of multiple electric fields (MEF, IEF, and PEF), which establish a continuous and efficient charge transfer pathway within the composite catalyst. Specifically, MEF enhances carrier separation in UPDI, while the generated free electrons migrate to PCN under the influence of IEF. Furthermore, PEF promotes charge separation on the PCN surface and increases the concentration of high-energy electrons. The enhanced reaction kinetics can be ascribed to two principal factors: (1) the spatial charge separation facilitated by the series electric field, which significantly prolongs carrier lifetime; and (2) the superior photothermal effect that accelerates near-field chemical reactions. Li et al. [117] successfully synthesized a ternary composite material, g-C3N4/PDI/Co-Fe, via a co-precipitation method, leading to the formation of a double Z-scheme heterojunction. Under visible light irradiation for 60 min in the presence of PMS, the degradation efficiency of doxycycline hydrochloride (DOX) by the g-C3N4/PDI/Co-Fe composite reached 96.1%. The enhanced degradation efficiency was primarily attributed to the double Z-scheme heterojunction, which significantly promoted the separation and migration of photogenerated charge carriers. The mechanism by which g-C3N4/PDI/Co-Fe PBA activates PMS for DOX degradation under visible light is illustrated in Figure 3d. Upon visible light irradiation, the composite catalyst generates photogenerated charge carriers and establishes a double Z-scheme heterojunction among g-C3N4, PDI, and Co-Fe PBA. This heterojunction facilitates the transfer of photogenerated electrons from the conduction band of PDI to the valence bands of g-C3N4 and Co-Fe PBA, thereby inhibiting the recombination of photogenerated electron–hole pairs and enhancing the photocatalytic activity. Furthermore, photogenerated h+ directly oxidizes DOX. Photogenerated electrons activate PMS to produce sulfate radicals (SO4), some of which react with hydroxide ions (OH) to generate •OH and bisulfate ions (HSO4). Bisulfate ions subsequently react with photogenerated holes to regenerate SO4. Transition metals in Co-Fe PBA also contribute to the activation of PMS, generating additional SO4 that can react with photogenerated electrons to form •O2 and O2. •O2 then combines with photogenerated holes to form singlet oxygen (1O2). The synergistic action of 1O2, SO4, •O2, and photogenerated h+ collectively drives the degradation process. Additionally, •OH plays a pivotal role in the degradation and mineralization of DOX. A novel organic–inorganic stepped S-scheme heterojunction PDI-Urea/BiOBr composite photocatalyst was successfully fabricated by Wang et al. [118] through a simple solvothermal method coupled with an in situ growth technique. Under visible light irradiation, the 15% PDI-Urea/BiOBr composite demonstrated a degradation efficiency of 93.0% for ofloxacin (OFLO) within 150 min and achieved a removal rate of 58% for TC within 90 min. This significant improvement in photocatalytic efficiency mainly resulted from the construction of an S-scheme heterojunction, which effectively modified the band structure, broadened the spectral response range, and promoted the separation of photogenerated charge carriers. In a related study, Yang et al. [119] developed a 2,5-bis(tri-n-butylstannyl)thiophene-perylene diimide–TiO2 (Bis-PDI-T @TiO2) composite photocatalyst through a dual-solvent phase transfer method. This composite demonstrated exceptional peroxymonosulfate (PS) activation capability under visible light irradiation. Specifically, the 0.1% Bis-PDI-T@TiO2 system, when combined with 1.0 mM PS at a catalyst concentration of 1.0 g/L, achieved complete degradation (100.0%) of carbamazepine (CBZ) within 30 min. The superior photocatalytic performance was ascribed to the formation of a Z-scheme heterojunction between Bis-PDI-T and TiO2, which significantly enhanced the separation efficiency of photogenerated electron–hole pairs. Jia et al. [120] successfully developed a novel Z-scheme photocatalyst (MPx) through a mechanochemical synthesis approach, wherein β-alanine-modified perylene diimide derivatives (β-PDI) were integrated with MIL-101(Fe) via electrostatic interactions. The MP50/PS/visible light system demonstrated exceptional photocatalytic performance, achieving a 99.7% removal efficiency for sulfamethoxazole (SMX) within a remarkably short duration of six minutes. The reaction rate constant was determined to be 0.7532 min¹, representing a 69.1-fold and 175.2-fold enhancement compared to the individual components, β-PDI and MIL-101(Fe), respectively. Mechanistic investigations revealed that PS played dual roles in the system: (1) as an activator generating SO4, and (2) as an electron scavenger, which effectively promoted the separation of photogenerated electron–hole pairs within the photocatalyst, thereby significantly enhancing the overall photocatalytic efficiency. In a subsequent study, Jia et al. [121] engineered MNPx(MIL-101(Fe)-NH2/PDI) composites featuring Z-scheme heterojunctions through the mechanical grinding of MIL-101(Fe)-NH2 and PDI. These composites were employed for the activation of peroxydisulfate (PDS) under visible light irradiation, demonstrating remarkable efficacy in the degradation of SMX. The MNP30/PDS/Vis system achieved a 99.2% degradation efficiency of SMX within five minutes. The Z-scheme heterojunction architecture in the MNPx photocatalysts facilitated efficient electron transfer, effectively suppressed the recombination of electron–hole pairs, and significantly enhanced the activation efficiency of PDS. Furthermore, the large specific surface area of MIL-101(Fe)-NH2 provided abundant active sites for catalytic reactions. These synergistic effects collectively contributed to the rapid and efficient removal of SMX from the system.
In summary, PDI-based heterojunction photocatalysts have achieved remarkable technological advancements and demonstrated significant application potential in the degradation of antibiotics and pharmaceutical residues. Numerous researchers have successfully developed various types of PDI-based heterojunction photocatalysts using diverse preparation methods, including type-II, Z-scheme, S-scheme, and double Z-scheme heterojunctions. These photocatalysts exhibit superior degradation capabilities for a wide range of antibiotics and organic pollutants under visible light irradiation, achieving removal efficiencies of TC ranging from 70.0% to 99.7% within relatively short periods. The enhancement in photocatalytic performance is primarily attributed to the effective promotion of photogenerated carrier separation and migration by the heterojunction structure, which strengthens the oxidation–reduction ability of the photocatalyst, broadens the light absorption spectrum, and increases the number of active sites. These research findings offer critical insights and valuable references for the practical application of photocatalytic technology in addressing pollution from antibiotics and pharmaceutical residues.

4.2. Phenolic Compounds

In recent years, phenolic compounds—as representative refractory organic pollutants—have garnered significant attention in the development of efficient removal technologies. Among these materials, PDI-based heterojunctions have demonstrated remarkable advantages due to their distinctive photoelectric properties. Scholars have achieved a series of breakthroughs in enhancing photocatalytic performance through innovative heterojunction designs and advanced interface regulation strategies. Notably, researchers have realized directional carrier migration by modulating the heterojunction type (e.g., type II, Z-scheme, and S-scheme), combined with strategies such as plasma enhancement, multi-interface synergy, and three-dimensional structural design. These approaches not only significantly broaden the light response range, but also enhance the yield of reactive oxygen species by three to seven times. Such innovative systems exhibit universal applicability in the deep mineralization of bisphenol A and estrogenic pollutants, offering a novel solution for the efficient treatment of complex phenolic contaminants.
Chen et al. [122] fabricated a novel plasmonic Z-scheme heterojunction photocatalyst, Ag@AgCl/PDI, utilizing a stepwise approach that combines in situ deposition with photoreduction techniques. The fabrication process, as shown in Figure 4a, commenced with dissolving PDI in an alkaline triethanolamine (TEA) solution, subsequently acidified with HCl. Through π–π interactions and hydrogen bonding of carboxyl groups, self-assembled supramolecular PDI (SA-PDI) was formed. Silver chloride nanoparticles were then electrostatically assembled onto the SA-PDI surface through a precipitation reaction between Ag+ from silver nitrate and chloride ions (Cl) from sodium chloride. The synthesis was completed by partially reducing Ag+ to metallic silver (Ag0) through in situ photoreduction, yielding the final Ag@AgCl/PDI composite. Figure 4b illustrates the electron transfer mechanism in the composite system. Electrons naturally flow from SA-PDI to AgCl through the silver component until Fermi level equilibrium is achieved among the three components. This electron migration decreases electron density in SA-PDI while increasing it in AgCl, creating positive charges on SA-PDI and negative charges on AgCl. Consequently, an internal electric field and band bending are established at the heterojunction interface. When exposed to light, this arrangement enables the movement of free electrons from the downward-bent conduction band of AgCl to the upward-bent conduction band of SA-PDI via the silver bridge. Figure 4c schematically presents this Ag-bridged Z-scheme heterojunction mechanism. The Ag@AgCl/PDI-3% composite exhibited remarkable photocatalytic performance, achieving 92.6% phenol removal efficiency after three hours of visible light exposure, as shown in Table 2. This enhanced photocatalytic oxidation capability stems from three key aspects: (1) The integration of Ag@AgCl nanoparticles substantially improves the light absorption of SA-PDI through surface plasmon resonance (SPR). (2) Silver nanoparticles act as efficient electron traps, capturing photogenerated electrons from SA-PDI’s conduction band. Additionally, the Schottky barrier formed by silver promotes the transfer of SPR-excited electrons from Ag to AgCl, boosting charge separation efficiency. (3) The composite generates increased concentrations of reactive oxygen species (•O2 and •OH) and h+, significantly contributing to its photocatalytic activity.
Table 2. The photocatalytic degradation performance of a phenolic by a PDI-based heterojunction.
Table 2. The photocatalytic degradation performance of a phenolic by a PDI-based heterojunction.
PhotocatalystSynthesis MethodPhenolicLight SourceTime
(min)
Efficiency
(%)
TypeReference
Ag@AgCl/PDIIn situ deposition, photoreductionPhOHVisible18092.6ZChen et al. [122]
BiOCl/PDIWater bath heatingPhOHSimulated sunlight18087.0ZGao et al. [123]
TOC-PDI-POSS/g-C3N4Solvent exchangePhOHSimulated sunlight6097.0SDai et al. [124]
BiOBr/Bi4O5Br2/PDISelf-assemblyBPAVisible7590.0IIWang et al. [125]
PDIBr/A10Solvent exchangeBPAVisible6071.0ZZha et al. [126]
Bi12O15Cll6@W18O49
@g-C3N4/PDI
Solvent thermal, calcinationBPASimulated sunlight30100.0ZZhang
et al. [127]
Figure 4. (a) Schematic illustration of synthesizing Ag@AgCl/PDI composite photocatalysts; (b,c) schematic diagram of the charge transfer process in Ag@AgCl/PDI [122]. (d,e) TEM images of BiOBr/Bi4O5Br2/PDI; (f) UV–Vis DRS spectra of as-prepared samples; (g) photocatalytic degradation curves of BPA; (h) pseudo-first-order kinetic model fitting the kinetic degradation curves of BPA [125].
Figure 4. (a) Schematic illustration of synthesizing Ag@AgCl/PDI composite photocatalysts; (b,c) schematic diagram of the charge transfer process in Ag@AgCl/PDI [122]. (d,e) TEM images of BiOBr/Bi4O5Br2/PDI; (f) UV–Vis DRS spectra of as-prepared samples; (g) photocatalytic degradation curves of BPA; (h) pseudo-first-order kinetic model fitting the kinetic degradation curves of BPA [125].
Catalysts 15 00565 g004
In a parallel investigation, Gao et al. [123] prepared PDI/BiOCl Z-scheme heterojunctions using a water bath heating technique. The 30% BiOCl/PDI composite demonstrated superior photocatalytic performance, achieving 87.0% phenol degradation efficiency after 180 min of simulated sunlight exposure. This represents a 2.2-fold enhancement compared to pure BiOCl and a 1.6-fold improvement over PDISA. The composite’s outstanding photocatalytic properties are ascribed to its enhanced generation of reactive species, comprehensive light absorption across the full spectrum, and efficient interfacial charge separation. Dai et al. [124] successfully synthesized a soluble hybrid material, titanium oxocluster–pyrene diimide–polyhedral oligomeric silsesquioxane (TOC-PDI-POSS), via the solvent exchange method. Exploiting π–π stacking interactions with g-C3N4, they successfully fabricated a novel TOC-PDI-POSS/g-C3N4 S-scheme heterojunction. Under simulated sunlight irradiation for 60 min, the TOC-PDI-POSS/g-C3N4 photocatalyst achieved a phenol removal efficiency of 97.0%. This exceptional photocatalytic performance is primarily attributed to the efficient separation of photogenerated carriers enabled by the S-scheme heterojunction in TOC-PDI-POSS/g-C3N4. Wang et al. [125] constructed a BiOBr/Bi4O5Br2/PDI photocatalyst featuring a type-II double heterojunction through alkaline dehalogenation and PDI self-assembly techniques. As illustrated in Figure 4d, the thin belt-like structure of PDI is uniformly coated on the surface of BiOBr/Bi4O5Br2, leveraging the superior photosensitivity of PDI to extend the visible light response range. High-resolution transmission electron microscopy (HRTEM) analysis (Figure 4e) revealed distinct lattice fringes of BiOBr and Bi4O5Br2, with interplanar spacings of 0.276 nm for the (110) plane of BiOBr and 0.365 nm for the (310) plane of Bi4O5Br2, while PDI exhibited an amorphous state. The crystalline structures of BiOBr and Bi4O5Br2 in the BiOBr/Bi4O5Br2/PDI composite remained intact post-solvothermal reaction and PDI self-assembly. Ultraviolet–visible diffuse reflectance spectroscopy (UV–Vis DRS) characterization (Figure 4f) demonstrated that BiOBr/Bi4O5Br2/PDI exhibited a stronger visible light response compared to BiOBr/Bi4O5Br2, with the light absorption wavelength extending from 479 to 721 nm upon PDI incorporation. The construction of the Bi4O5Br2 heterojunction and the integration of PDI synergistically enhanced photocatalytic activity. Under visible light irradiation, BiOBr/Bi4O5Br2/PDI degraded 90.0% of BPA within 75 min, completely degraded 17α-ethynylestradiol (EE2) within 15 min, and fully removed 17β-estradiol (E2) within the same duration. The BPA removal efficiency followed the order BiOBr/Bi4O5Br2/PDI (20.0%) > Bi4O5Br2/PDI > BiOBr/Bi4O5Br2 > PDI > BiOBr/PDI > Bi4O5Br2 > BiOBr (Figure 4g). Pseudo-first-order kinetic fitting of the degradation curves (Figure 4h) revealed that BiOBr/Bi4O5Br2/PDI exhibited the highest reaction rate constant (K = 0.0376), significantly surpassing that of pure BiOBr/Bi4O5Br2 (K = 0.0157) and pure BiOBr (K = 0.0056). The superior photocatalytic activity of the BiOBr/Bi4O5Br2/PDI composite is ascribed to the matched band structure between Bi4O5Br2 and BiOBr, which facilitates heterojunction formation and enhances spatial charge separation. Furthermore, the strong photosensitivity of PDI combined with BiOBr/Bi4O5Br2 not only augments visible light photocatalytic activity but also broadens the light-harvesting range. Zha et al. [126] successfully demonstrated the synthesis of a novel Z-scheme heterojunction PDIBr/A10 composite material via a solvent exchange method, wherein perylene diimide derivatives (PDIs) were self-assembled on the surface of anatase TiO2 (A10). The photocatalytic performance of the composite was systematically evaluated under visible light irradiation for one hour, achieving a degradation efficiency of 71.1% for BPA without a sacrificial agent and 71.7% with persulfate as a sacrificial agent. The enhanced photocatalytic activity was primarily attributed to the π–π orbital overlap in H-aggregates facilitated by PDIBr, which formed an interface layer on the TiO2 surface. This interface layer promoted efficient charge transfer and suppressed electron–hole recombination. Additionally, the Z-scheme heterojunction mechanism was confirmed as the primary pathway for the photocatalytic degradation of BPA, using the PDIBr/A10 composite as a model system. In a related study, Zhang et al. [127] developed a plasmonic Bi12O15Cl6@ W18O49@graphitic carbon nitride/poly(1,4-phenylene terephthalamide) (g-C3N4/PDI) double Z-scheme heterojunction photocatalyst featuring dual charge transfer pathways. The synthesis involved a solvothermal method, where W18O49 nanowires were grown on Bi12O15Cl6 nanosheets, followed by a calcination process to load g-C3N4/PDI onto the Bi12O15Cl6@ W18O49 structure. The resulting heterojunction exhibited exceptional photocatalytic performance for BPA degradation under simulated solar light irradiation, achieving complete removal efficiency within 30 min. The superior photocatalytic activity was ascribed to the synergistic effects of the double Z-scheme heterojunction and plasmonic resonance, which significantly enhanced the separation efficiency of photogenerated carriers and the overall photocatalytic activity of the composite material.
In a series of studies on the degradation of phenolic compounds, PDI-based heterojunction photocatalysts have exhibited outstanding performance. Through various preparation techniques, such as in situ precipitation, in situ deposition photoreduction, water bath heating, and solvent exchange, researchers have successfully fabricated diverse heterojunction structures, including type II, Z-scheme, and S-scheme. These heterojunctions not only enhance the light absorption capability of the photocatalysts but also significantly improve the separation and transfer efficiency of photogenerated carriers. For example, photocatalysts such as Ag@AgCl/PDI, PDI/BiOCl, and TOC-PDI-POSS/g-C3N4 demonstrate remarkable photocatalytic activity under visible light irradiation. These research achievements provide versatile strategies for photocatalyst development and offer robust technical support for the efficient removal of phenolic pollutants from water.

4.3. Industrial Dyes and Other Pollutants

With the rapid advancement of the printing and dyeing industry and modern agriculture, the treatment of complex pollutants, such as rhodamine B, atrazine (ATZ), malathion (MA), sodium lignosulfonate (SL), and iodinated organic halides (IOH), has emerged as a significant challenge in the environmental domain. To address the limitations of traditional photocatalytic materials, such as poor universality and low response efficiency for multiple pollutant types, researchers have achieved breakthroughs in co-degradation by precisely designing multi-component PDI-based heterojunction systems. These advancements indicate that through heterojunction interface engineering (e.g., π–π conjugation enhancement and stepwise bandgap modulation) and multi-mechanism coupling (e.g., plasmonic resonance and the piezoelectric effect), PDI-based composite materials are evolving into a new generation of environmental remediation materials with broad-spectrum degradation capabilities and substantial engineering application potential.
Through ultrasonic-assisted self-assembly technology, Zhang et al. [128] successfully fabricated PDI/BiO2−x type-II heterojunctions. The experimental results demonstrated that the 20% self-assembled PDI/BiO2−x composite materials exhibited remarkable photocatalytic performance under visible light irradiation, achieving a degradation efficiency of RhB up to 98.7% within 20 min, as shown in Table 3. This enhanced performance is primarily attributed to the formation of type-II heterojunctions, wherein the built-in electric field effectively facilitates the spatial separation of photogenerated carriers. Xu et al. [129] developed PDISA/AgBr type-II heterojunction organic–inorganic hybrid photocatalysts via a chemical co-precipitation method. The experimental findings revealed that the PDISA/AgBr-40 composite materials achieved a degradation efficiency of RhB of 97.8% within 20 min under visible light irradiation. This superior photocatalytic performance can be ascribed to the synergistic effects of type-II heterojunctions, which not only enhance the separation of photogenerated electron–hole pairs but also broaden the light absorption spectrum. Furthermore, Zhang et al. [130] successfully synthesized PDI/Bi2O4 type-II heterojunction photocatalysts by integrating water bath heating and ultrasonic dispersion techniques. The 5% PDISA/Bi2O4 composites achieved degradation efficiencies of 98.6% for RhB and 97.0% for methylene blue within 25 min of visible light irradiation. The underlying mechanism involves the formation of an internal electric field at the PDI/Bi2O4 interface, which significantly promotes the separation and migration of photogenerated carriers, thereby enhancing the overall photocatalytic activity.
Mardiroosi et al. [131] developed an innovative type-II heterojunction material, PDI@g-C3N4@UiO-66 (PCN@UiO-66), through solvothermal synthesis. This composite exhibited exceptional photocatalytic properties when degrading rhodamine B under visible light exposure for 140 min, with an impressive removal rate of 99.0%. The outstanding performance of 30PCN@UiO-66 is mainly due to its improved visible light harvesting capability and the effective segregation and movement of photoinduced charge carriers, enabled by its type-II heterojunction configuration. Zhang et al. [132] constructed a novel SAPDI/Bi4O7 (PB) type-II heterojunction by integrating self-assembled perylene diimide (SAPDI) with Bi4O7. This catalytic material showed an 87.6% decomposition rate for rhodamine B when exposed to visible light for 20 min. The improved photocatalytic efficiency stems from the type-II photoinduced charge transfer occurring between Bi4O7 and SAPDI, which promotes electron transfer and increases the separation of photoinduced charge carriers. Through a combination of two-step heating and oil bath techniques, Zhu et al. [133] fabricated g-C3N4/PDI@ZnIn2S4 Z-scheme heterojunctions. The g-C3N4/PDI@ZnIn2S4-0.7 sample achieved an 83.9% removal rate for RhB after 120 min of visible light exposure, outperforming pure g-C3N4, g-C3N4/PDI, and ZnIn2S4 by factors of 7.0, 2.2, and 1.4, respectively. The superior photocatalytic activity results from the close interfacial connection and optimized band alignment between g-C3N4/PDI and ZnIn2S4, which enables the creation of an effective Z-scheme heterojunction, improving charge carrier separation and visible light utilization. Tang et al. [134] developed a novel π–π stacked PDI/g-C3N4/TiO2@Ti3C2(PCT) photocatalyst featuring S-scheme heterojunction through a calcination process. As illustrated in Figure 5a, the photocatalytic degradation efficiency of ATZ was significantly enhanced through the activation of PMS. Under visible light irradiation for one hour, the PDI/g-C3N4/TiO2@Ti3C2 photocatalyst achieved a 75.0% removal rate of ATZ. The pseudo-first-order kinetic curve in Figure 5b and the calculated reaction rate constants revealed that the initial reaction rate constant of ATZ was negligible (k1 = 0.0002). The synergistic effect of g-C3N4 and PMS improved the degradation of ATZ, achieving a removal rate of 31.0% (k2 = 0.0056). The incorporation of TiO2@Ti3C2 or PDI into g-C3N4 further enhanced ATZ degradation, with no significant difference observed in the reaction constants between the PC(PDI/g-C3N4) and CT(g-C3N4/TiO2@Ti3C2) samples (k3 = 0.0089 and k4 = 0.0100, respectively). This enhancement is ascribed to the intermolecular π–π interactions characteristic of the PC composites and the heterojunctions formed between TiO2@Ti3C2 and g-C3N4. The combination of the intermolecular π–π interactions of PC with TiO2@Ti3C2 to form an S-scheme heterojunction significantly increased the photocatalytic degradation rate of ATZ to 75.0% in the PCT/PMS/visible light system. Electron paramagnetic resonance (EPR) spectroscopy confirmed the presence of •OH and SO4 radicals through the detection of the DMPO spectrum (Figure 5c). The results indicated that both •OH and SO4 were suppressed in the dark, but strong •OH peaks (intensity ratio of 1:2:2:1) and weak SO4 peaks were observed after visible light irradiation. This suggests that •OH and SO4 may be activated during the photocatalytic process, but do not play a dominant role in ATZ degradation. The presence of 1O2 was also investigated using EPR spectroscopy (Figure 5d), with the DMPO-1O2 signal observed only after visible light irradiation. It is hypothesized that either a type-II or S-scheme heterojunction may have formed between PC and TiO2@Ti3C2 (Figure 5e). In the case of a type-II heterojunction formation (illustrated on the left side of Figure 5e), visible light excitation would cause photogenerated electrons to transfer to the valence band of PC, while holes would move to the conduction band of TiO2@Ti3C2. Nevertheless, the valence band energy level of PC (2.05 eV) is inadequate for •OH production, given the high oxidation–reduction potential of H2O/•OH (2.37 eV). Consequently, the electron transfer process cannot be accounted for by a conventional type-II heterojunction. Considering the formation of •OH radicals, an S-scheme heterojunction offers a more convincing explanation for the electron transfer dynamics in the PCT photocatalyst (depicted on the right side of Figure 5e). The intimate interface between PC and TiO2@Ti3C2 promotes the establishment of an S-scheme heterojunction. When exposed to visible light, photogenerated electrons emerge in the valence bands of both PC and TiO2@Ti3C2, subsequently being excited to their respective conduction bands. Electrons from PC are then transferred to TiO2@Ti3C2 until the Fermi levels of both materials achieve equilibrium. This charge redistribution induces bending of the valence band in PC and the conduction band edge in TiO2@Ti3C2, thereby creating an internal electric field within the PCT composite. The S-scheme heterojunction mechanism elucidates the production of •OH and, in conjunction with the conduction band of PC, the formation of •O2 (with an oxidation–reduction potential of –0.33 eV). The π–π interactions within PDI/g-C3N4 facilitate the delocalization of photogenerated electrons, enhancing their mobility. Additionally, the close interfacial contact and staggered band alignment between PDI/g-C3N4 and TiO2@Ti3C2 establish an S-scheme heterojunction, thereby minimizing the recombination of photogenerated charge carriers.
Through the hydrothermal synthesis method, Ren et al. [135] successfully fabricated an H-PDI supramolecular/NH2-MIL-101(Fe) Z-scheme heterojunction. The experimental results demonstrated that the 40% H-PDI/NH2-MIL-101(Fe) composite material achieved a degradation efficiency of 91.1% for malathion after 180 min of simulated sunlight exposure. This remarkable photocatalytic performance is primarily attributed to the formation of the Z-scheme heterojunction structure, which significantly enhances the spatial separation efficiency of photogenerated electron–hole pairs. Chen et al. [136] developed a novel organic–inorganic hybrid photocatalyst, In2O3/PDI/In2S3 (IO/PDI/IS), by integrating solvent-induced self-assembly and electrostatic interaction mechanisms. This catalyst features a unique double S-scheme structure. Under 420 nm LED light irradiation (5 W), the mineralization efficiency of sodium lignosulfonate reached 80.9% within 80 min. This superior photocatalytic performance can be ascribed to two critical factors: first, the incorporation of layered PDI optimizes the charge transfer pathway in the IO/IS type-II heterojunction, establishing a double S-scheme charge transfer mechanism; second, the extended π–π conjugation system in layered PDI significantly enhances the internal electric field strength, thereby improving the migration and separation efficiency of photogenerated carriers. Ji et al. [137] successfully synthesized a new Z-scheme perylene diimide/MIL-101(Cr) (PM) heterojunction via the water bath heating method. Figure 5f provides a detailed illustration of the preparation process of this hybrid material. The PM heterojunction was employed for visible light-assisted persulfate activation, constructing the PM/PS/Vis photocatalytic system for the degradation of IOH in aqueous environments. In the PM-7/PS/Vis system, IOH was nearly completely degraded within 35 min, achieving a removal efficiency close to 100.0%. Additionally, under conditions simulating actual wastewater treatment plant samples, the degradation rate of IOH reached 77.5% within the same time frame. Figure 5g illustrates the reaction mechanism of the PM-7/PS/Vis system. Unlike the traditional type-II heterojunction model, where photogenerated holes migrate from MIL-101(Cr)’s valence band to SA-PDI’s valence band and electrons transfer from SA-PDI’s conduction band to MIL-101(Cr)’s conduction band, this system exhibits unique characteristics. The conduction band potential of MIL-101(Cr) (–0.046 eV) is insufficiently negative to reduce O2 to •O2, while SA-PDI’s valence band potential (1.54 eV) lacks the necessary positive potential to oxidize H2O into •OH. These limitations led to the development of a Z-scheme heterojunction mechanism for the PM photocatalyst. In this configuration, excited electrons from MIL-101(Cr)’s conduction band interface with photogenerated holes in SA-PDI’s valence band, maintaining the robust reducing capacity of SA-PDI’s conduction band electrons and the potent oxidizing capability of MIL-101(Cr)’s valence band holes, ultimately boosting photocatalytic performance. PS integration serves as an efficient approach to minimize electron–hole recombination in SA-PDI. This compound demonstrates dual capabilities: directly accepting electrons from SA-PDI’s conduction band to form SO4 and being activated by MIL-101(Cr) to produce reactive species. Within the PDI/PS/Vis framework, radical chain reactions occur, with SA-PDI’s conduction band potential (–0.046 eV vs. O2/•O2) enabling •O2 formation. Conduction band electrons engage in various reduction processes, transforming persulfate (S2O82) into SO4 and H2O into •OH. Moreover, photogenerated electrons can directly break PS’s O-O bonds, generating additional SO4. These reactive species (SO4, •OH, •O2, and 1O2) work in concert with h+ to decompose IOH. The PM system’s Z-scheme heterojunction structure exhibits remarkable charge transfer efficiency, enabling electron movement from MIL-101(Cr)’s conduction band to SA-PDI’s valence band. This mechanism, combined with PS’s function as an electron acceptor, improves charge carrier mobility and stimulates reactive oxygen species production, thereby enhancing IOH oxidation. In a related development, Ren et al. [138] successfully synthesized NH2-UiO-66(Zr)/PDI (NUPDI) Z-scheme heterojunction photocatalysts via a facile self-assembly approach. Under visible light irradiation for 120 min, the NUPDI photocatalyst demonstrated a remarkable hexavalent chromium (Cr(VI)) removal efficiency of 99.5%, representing 24.9-fold and 1.3-fold enhancements compared to pristine PDI and NH2-UiO-66(Zr), respectively. The superior photocatalytic performance of the NUPDI heterojunction in Cr(VI) reduction can be primarily attributed to two key factors: (1) the π–π interactions between NH2-UiO-66(Zr) and PDI, and (2) the energy-matched Z-scheme heterojunction structure. These both collectively contribute to enhanced charge separation and transfer efficiency.
Herein, we thoroughly investigated the application of PDI-based heterojunction photocatalysts in the degradation of industrial dyes and the treatment of various pollutants. Through diverse preparation methods, such as ultrasonic-assisted self-assembly, chemical co-precipitation, and water bath heating, researchers have successfully constructed type-II, S-scheme, and Z-scheme heterojunction structures. These heterojunction photocatalysts have markedly improved the degradation efficiency of pollutants, including rhodamine B, methylene blue, and hexavalent chromium, by promoting the separation of photogenerated carriers, extending the light absorption range, and enhancing the generation of reactive species. These studies not only highlight the superior performance of PDI-based heterojunction photocatalysts in degrading industrial dyes but also broaden their potential applications in heavy metal ion treatment and other fields, offering novel strategies and approaches for environmental pollution control.

5. Conclusions

This article provided a systematic review of the existing research on PDI-based heterojunction photocatalysts in the field of organic pollutant degradation. It highlighted that PDI-based heterojunction photocatalysts, characterized by their wide spectral response, high charge separation efficiency, and excellent stability, have been demonstrated to possess significant advantages for the efficient degradation of various pollutants, including antibiotics, phenolic compounds, and industrial dyes. By constructing heterojunction structures such as type II, Z-scheme, and S-scheme (e.g., PDI/g-C3N4, PDI/MOFs, and PDI/BiOCl), researchers have effectively mitigated the issues of high carrier recombination rates and narrow light absorption ranges commonly found in traditional photocatalysts. These heterojunctions significantly enhance photocatalytic activity through interfacial charge transfer mechanisms. For example, the degradation rate of tetracycline using PDI-based heterojunctions can reach 70.0% to 99.7%, while the degradation efficiency for pollutants such as bisphenol A and rhodamine B exceeds 90.0%. This article further summarized the core mechanisms underlying heterojunction design, including band alignment, optimization of built-in electric fields, and enhancement of π–π conjugation effects. Additionally, it emphasized the promoting roles of synergistic effects, such as plasmonic resonance and persulfate activation, in enhancing the generation of reactive oxygen species. Future research should focus on material functionalization strategies (e.g., group modification and supramolecular assembly), expanding the light absorption range (via narrow bandgap semiconductor composites), improving interface stability (through carbon coating or MOF encapsulation), and developing multi-pollutant co-degradation systems. These efforts will facilitate the transition of PDI-based heterojunction photocatalysts from laboratory-scale studies to large-scale environmental remediation applications.

6. Prospects

6.1. Functionalization and Modification of Photocatalytic Materials

In the context of molecular engineering regulation, the water dispersibility and pollutant adsorption capacity of PDI derivatives can be enhanced through functional group modification (e.g., sulfonic acid groups and amino groups). In terms of supramolecular self-assembly technology, ordered nanostructures such as nanofibers and micelles can be constructed via π–π stacking or hydrogen bonding interactions, thereby increasing the exposure of active sites. Integration of stimuli-responsive moieties (e.g., pH-sensitive or redox-active groups) could enable smart photocatalytic systems for adaptive environmental remediation under varying wastewater conditions.

6.2. Optimization of Light Absorption and Quantum Efficiency

Despite the wide spectral response capabilities of existing PDI-based materials, there remains potential for improvement in their utilization of the full solar spectrum, particularly in the visible and near-infrared regions. By incorporating narrow bandgap semiconductors (e.g., Ag2CO3 and CdS) to construct heterojunctions, the light absorption range can be expanded and the separation efficiency of photogenerated carriers can be significantly enhanced. Incorporating upconversion nanoparticles (e.g., NaYF4:Yb3+/Tm3+) could extend light harvesting to the near-infrared region, enhancing energy utilization in turbid or shaded water bodies. Furthermore, element doping (e.g., non-metals or transition metals) and surface plasmon resonance effects (e.g., noble metal nanoparticles) can be employed to fine-tune the band structure and improve the quantum efficiency of light, thereby reducing the energy consumption of photocatalytic reactions.

6.3. Enhancement of Heterojunction Interfaces and Stability

The heterojunction interface plays a critical role in determining photocatalytic performance. Future research should focus on in-depth investigations of the interfacial electron transfer mechanisms between PDI and other semiconductors (e.g., TiO2 and g-C3N4) to optimize interface contact and minimize carrier recombination. Additionally, addressing the stability issues of PDI-based materials, such as photo-corrosion and chemical degradation, is essential. This can be achieved through introducing protective layers (e.g., carbon coatings or MOF encapsulation) or developing self-healing material systems to extend the cyclic service life of the catalysts. Self-healing materials (e.g., dynamic covalent networks) could autonomously repair defects during operation, extending catalyst lifespan. For industrial viability, modular reactor designs incorporating flow-through catalytic membranes or immobilized PDI composites should be developed to ensure stability under continuous operation.

6.4. Development and Expansion of Synergistic Catalytic Technologies

Hybrid systems integrating PDI photocatalysis with complementary technologies (e.g., electrocatalysis, piezocatalysis, or bioaugmentation) could unlock multifunctional remediation platforms. In photoelectrocatalytic coupling, the synergistic interaction between an external electric field and light excitation enhances the generation efficiency of reactive species such as •OH and SO4, thereby improving the degradation efficiency of organic pollutants. Photoelectrocatalytic reactors combining PDI heterojunctions with conductive substrates (e.g., carbon cloth or Ti mesh) can enhance •OH yield via bias-assisted charge separation. Microbial fuel cell–PDI hybrids enable simultaneous organic pollutant degradation and bioelectricity generation. Plasmonic photothermal–PDI systems utilize localized heat to accelerate reaction kinetics in viscous or high-salinity wastewater.

6.5. Commercialization-Oriented Innovation

To accelerate the transition of photocatalytic technology from the laboratory to industrialization, future research should prioritize the following areas: In terms of low-cost raw materials, precious metals should be replaced with abundant alternatives (such as copper or iron-based cocatalysts) to reduce material costs and enhance sustainability. In the aspect of circular economy strategies, recyclable PDI composites (such as magnetic Fe3O4@PDI) should be developed or closed-loop recycling schemes should be established to improve resource utilization and environmental compatibility. In the field of artificial intelligence-driven process optimization, digital twin technology should be utilized to simulate large-scale photocatalytic reactors and predict their performance under dynamic operating conditions, thereby enhancing process efficiency and scalability.
The future advancement of PDI-based photocatalytic technology is expected to be centered on the improvement of material performance via molecular engineering approaches (e.g., functional group modification and supramolecular assembly), achieving full-spectrum light utilization through the design of narrow-bandgap heterostructures and upconversion materials, and overcoming stability limitations by incorporating interface engineering and self-healing mechanisms. Concurrently, efforts will be directed toward developing light–electromagnetic synergy catalytic systems to enable the simultaneous degradation of multiple pollutants. Furthermore, by integrating circular economy strategies (such as magnetic recovery and closed-loop regeneration) with artificial intelligence-driven digital twin technology, a cost-effective, high-efficiency industrial-scale photocatalytic solution may be established, thereby promoting the intelligent and sustainable development of environmental remediation technologies.

Author Contributions

X.S.: investigation, writing—original draft preparation, review and editing. J.L.: visualization. Y.H.: methodology. Y.C.: investigation, editing. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the Doctoral Startup Project of Hubei University of Science and Technology, funded by the Hubei University of Science and Technology (Grant No. BK202501); the Horizontal Project of Hubei University of Science and Technology, funded by the Hubei University of Science and Technology (Grant No. 2025HX041); and the 2024 Open Fund Project of Hubei Provincial Key Laboratory of Green Materials for Light Industry, Hubei University of Technology, funded by the Hubei University of Technology (Grant No. 202409A02).

Data Availability Statement

Raw data are available upon request.

Acknowledgments

The authors have reviewed and edited the output and take full responsibility for the content of this publication.

Conflicts of Interest

The authors declare no conflicts of interest in publishing these results.

References

  1. Alsulmi, A.; Mohammed, N.N.; Soltan, A.; Messih, M.A.; Ahmed, M. Engineering S-scheme CuO/ZnO heterojunctions sonochemically for eradicating RhB dye from wastewater under solar radiation. RSC Adv. 2023, 13, 13269–13281. [Google Scholar] [CrossRef] [PubMed]
  2. Khudhair, E.M.; Ammar, S.H.; Al-Najjar, S.Z.; Al-Jubouri, S.M.; Mahdi, A.S.; Jabbar, Z.H. Facile construction of g-C3N4/MnWO4/NiS heterostructures for photocatalytic degradation of organic contaminates under visible light irradiation. Mater. Lett. 2023, 347, 134599. [Google Scholar] [CrossRef]
  3. Ponomarev, A.A.; Nurullina, T.S.; Zavatsky, M.D. Remediation of Cr (vi) in water using biosynthesized palladium nano-materials loaded (Shewanella oneidensis) MR-1. Water Conserv. Manag. 2022, 6, 146–153. [Google Scholar] [CrossRef]
  4. Al-Jubouri, S.M.; Sabbar, H.A.; Khudhair, E.M.; Ammar, S.H.; Batty, S.A.; Khudhair, S.Y.; Mahdi, A.S. Silver oxide-zeolite for removal of an emerging contaminant by simultaneous adsorption-photocatalytic degradation under simulated sunlight irradiation. J. Photochem. Photobiol. A Chem. 2023, 442, 114763. [Google Scholar] [CrossRef]
  5. Yaseen, D.; Scholz, M. Textile dye wastewater characteristics and constituents of synthetic effluents: A critical review. Int. J. Environ. Sci. Technol. 2019, 16, 1193–1226. [Google Scholar] [CrossRef]
  6. Zhao, Y.; Guo, H.; Liu, J.; Xia, Q.; Liu, J.; Liang, X.; Liu, E.; Fan, J. Effective photodegradation of rhodamine B and levofloxacin over CQDs modified BiOCl and BiOBr composite: Mechanism and toxicity assessment. J. Colloid Interface Sci. 2022, 627, 180–193. [Google Scholar] [CrossRef]
  7. Wang, J.; Wang, S. Toxicity changes of wastewater during various advanced oxidation processes treatment: An overview. J. Clean. Prod. 2021, 315, 128202. [Google Scholar] [CrossRef]
  8. Al-Jubouri, S.M.; Al-Jendeel, H.A.; Rashid, S.A.; Al-Batty, S. Green synthesis of porous carbon cross-linked Y zeolite nanocrystals material and its performance for adsorptive removal of a methyl violet dye from water. Micropor. Mesopor. Mater. 2023, 356, 112587. [Google Scholar] [CrossRef]
  9. Lan, D.; Zhu, H.; Zhang, J.; Li, S.; Chen, Q.; Wang, C.; Wu, T.; Xu, M. Adsorptive removal of organic dyes via porous materials for wastewater treatment in recent decades: A review on species, mechanisms and perspectives. Chemosphere 2022, 293, 133464. [Google Scholar] [CrossRef]
  10. Zhou, T.; Hou, J.; Tai, M.; Shi, J.; Mi, X.; Hu, B.; Liu, C.; Yan, L.; Liu, L. Polyethyleneimine-induced in-situ chemical epitaxial growth ultrathin 2D/2D graphene carbon nitride intralayer heterojunction with elevating photocatalytic activity: Performances and mechanism insight. Int. J. Hydrogen Energy 2024, 51, 884–896. [Google Scholar] [CrossRef]
  11. Guo, M.; Ma, Y.; Liu, Z.; Wang, D.; Yang, Y.; Li, X.; Liu, E. Electron, hole and radical competition mechanism of layered porous g-C3N4 for hydrogen generation and organic pollutant degradation. J. Catal. 2024, 430, 115332. [Google Scholar] [CrossRef]
  12. Hu, X.; Zhang, Z.; Lu, P.; Zhou, Y.; Zhou, Y.; Bai, Y.; Yao, J. Cyano-deficient g-C3N4 for round-the-clock photocatalytic degradation of tetracycline: Mechanism and application prospect evaluation. Water Res. 2024, 260, 121936. [Google Scholar] [CrossRef]
  13. Hayat, A.; Ajmal, Z.; Alzahrani, A.Y.A.; Moussa, S.B.; Khered, M.; Almuqati, N.; Alshammari, A.; Al-Hadeethi, Y.; Ali, H.; Orooji, Y. The photocatalytic H2O2 production: Design strategies, Photocatalyst advancements, environmental applications and future prospects. Coord. Chem. Rev. 2025, 522, 216218. [Google Scholar] [CrossRef]
  14. Khoo, V.; Ng, S.F.; Haw, C.Y.; Ong, W.J. Additive manufacturing: A paradigm shift in revolutionizing catalysis with 3D printed photocatalysts and electrocatalysts toward environmental sustainability. Small 2024, 20, 2401278. [Google Scholar] [CrossRef]
  15. Zhang, Y.; Ma, D.; Li, J.; Zhi, C.; Zhang, Y.; Liang, L.; Mao, S.; Shi, J.W. Recent research advances of metal organic frameworks (MOFs) based composites for photocatalytic H2 evolution. Coord. Chem. Rev. 2024, 517, 215995. [Google Scholar] [CrossRef]
  16. Li, H.; Li, L.; Tang, Y.; Zhang, X.; Ji, S.; Luo, L.; Jiang, F. Photoinduced RhB-sensitized effect on a novel AgI/BiOCl/biochar photocatalyst to boost its photocatalytic performance for 17α-ethinyl estradiol degradation. Sep. Purif. Technol. 2024, 332, 125774. [Google Scholar] [CrossRef]
  17. Chen, Y.; Liu, Y.; Gong, X.; Wang, J. Photocatalytic degradation of chlorinated organic pollutants by ZnS@ ZIF-8 composite through hydrogen peroxide generation by activating dioxygen under simulated sunlight irradiation. J. Colloid Interface Sci. 2024, 654, 1417–1430. [Google Scholar] [CrossRef]
  18. Wang, Z.; Hu, N.; Wang, L.; Zhao, H.; Zhao, G. In Situ Production of Hydroxyl Radicals via Three-Electron Oxygen Reduction: Opportunities for Water Treatment. Angew. Chem. Int. Ed. 2024, 136, e202407628. [Google Scholar] [CrossRef]
  19. Bao, Z.; Jiang, Y.; Zhang, Z.; Lv, J.; Shen, W.; Dai, J.; Wang, J.; Cai, J.; Wu, Y. Visible-light-responsive S-vacancy ZnIn2S4/N- doped TiO2 nanoarray heterojunctions for high-performance photoelectrochemical water splitting. J. Mater. Chem. A 2024, 12, 15902–15913. [Google Scholar] [CrossRef]
  20. Xie, Z.; Saad, A.; Shang, Y.; Wang, Y.; Luo, S.; Wei, Z. Enhanced degradation of micropollutants by visible light photocatalysts with strong oxygen activation ability. Water Res. 2023, 247, 120785. [Google Scholar] [CrossRef]
  21. Deng, Y.; Shi, Y.; Li, L.; Tang, R.; Zhou, Z.; Xiong, S.; Li, W.; Liu, J.; Huang, Y. Molecular modification: A promising strategy for the design of donor-acceptor-type organic polymers photocatalyst. Appl. Catal. B Environ. 2024, 352, 124043. [Google Scholar] [CrossRef]
  22. Wu, W.; Zhang, N.; Wang, Y. Construction of Au/ZnWO4/CdS ternary photocatalysts with oxygen vacancy modification for efficient photocatalytic hydrogen production. Adv. Funct. Mater. 2024, 34, 2316604. [Google Scholar] [CrossRef]
  23. Li, R.; Li, Y.; Jia, X.; Yang, J.; Miao, X.; Shao, D.; Wu, J.; Song, H. 2D/2D ultrathin polypyrrole heterojunct aerogel with synergistic photocatalytic-photothermal evaporation performance for efficient water purification. Desalination 2024, 574, 117295. [Google Scholar] [CrossRef]
  24. Zhang, J.; Wang, X.; Wang, X.; Li, C. Heterophase Junction Effect on Photogenerated Charge Separation in Photocatalysis and Photoelectrocatalysis. Acc. Chem. Res. 2025, 58, 787–798. [Google Scholar] [CrossRef]
  25. Dhakshinamoorthy, A.; Li, Z.; Yang, S.; Garcia, H. Metal–organic framework heterojunctions for photocatalysis. Chem. Soc. Rev. 2024, 53, 3002–3035. [Google Scholar] [CrossRef]
  26. Li, F.; Zhu, G.; Jiang, J.; Yang, L.; Deng, F.; Li, X. A review of updated S-scheme heterojunction photocatalysts. J. Mater. Sci. Technol. 2024, 177, 142–180. [Google Scholar] [CrossRef]
  27. Ma, C.; Qin, L.; Zhou, T.; Zhang, J. Customized structures of hydrogen-bonded organic frameworks towards photocatalysis. Energy Environ. Sci. 2024, 17, 8992–9026. [Google Scholar] [CrossRef]
  28. Nie, C.; Wang, X.; Lu, P.; Zhu, Y.; Li, X.; Tang, H. Advancements in S-scheme heterojunction materials for photocatalytic environmental remediation. J. Mater. Sci. Technol. 2024, 169, 182–198. [Google Scholar] [CrossRef]
  29. Lu, Y.; Dong, Y.; Liu, W.; Jin, Q.; Lin, H. Piezo-photocatalytic enhanced microplastic degradation on hetero-interpenetrated Fe1−xS/FeMoO4/MoS2 by producing H2O2 and self-Fenton action. Chem. Eng. J. 2025, 508, 160935. [Google Scholar] [CrossRef]
  30. Li, J.; Xu, N.; Zhang, Y.; Dong, H.; Li, C. Research progress of heterogeneous photocatalyst for H2O2 production: A mini review. Chin. Chem. Lett. 2024, 00, 110470. [Google Scholar] [CrossRef]
  31. Wang, G.; Lv, S.; Shen, Y.; Li, W.; Lin, L.; Li, Z. Advancements in heterojunction, cocatalyst, defect and morphology engineering of semiconductor oxide photocatalysts. J. Mater. 2024, 10, 315–338. [Google Scholar] [CrossRef]
  32. Miao, J.; Yang, Y.; Cui, P.; Ru, C.; Zhang, K. Improving charge transfer beyond conventional heterojunction photoelectrodes: Fundamentals, strategies and applications. Adv. Funct. Mater. 2024, 34, 2406443. [Google Scholar] [CrossRef]
  33. Wang, D.; Xu, Y.; Yu, W.; Yin, L.; Liu, X.; Xia, J.; Zhang, N.; Fu, Y.; Yang, G.; Ni, B. Modulating charge carrier transfer channel by 2D/2D Schottky heterojunction of Ti3C2/BiOIO3 for effective photocatalytic degradation of typical antibiotics. Sep. Purif. Technol. 2024, 337, 126393. [Google Scholar] [CrossRef]
  34. Zhang, B.; Gao, H.; Kang, Y.; Li, X.; Li, Q.; Zhai, P.; Hildebrandt, D.; Liu, X.; Wang, Y.; Qiao, S. Molecular and Heterojunction Device Engineering of Solution-Processed Conjugated Reticular Oligomers: Enhanced Photoelectrochemical Hydrogen Evolution through High-Effective Exciton Separation. Adv. Sci. 2024, 11, 2308535. [Google Scholar] [CrossRef] [PubMed]
  35. Jiang, G.; Zheng, C.; Yan, T.; Jin, Z. Cd0.8Mn0.2S/MoO3 composites with an S-scheme heterojunction for efficient photocatalytic hydrogen evolution. Dalton Trans. 2021, 50, 5360–5369. [Google Scholar] [CrossRef]
  36. Li, Y.; Zhou, M.; Cheng, B.; Shao, Y. Recent advances in g-C3N4-based heterojunction photocatalysts. J. Mater. Sci. Technol. 2020, 56, 1–17. [Google Scholar] [CrossRef]
  37. Du, H.; Liu, Y.; Shen, C.; Xu, A. Nanoheterostructured photocatalysts for improving photocatalytic hydrogen production. Chin. J. Catal. 2017, 38, 1295–1306. [Google Scholar] [CrossRef]
  38. Shu, A.; Qin, C.; Li, M.; Zhao, L.; Shangguan, Z.; Shu, Z.; Yuan, X.; Zhu, M.; Wu, Y.; Wang, H. Electric effects reinforce charge carrier behaviour for photocatalysis. Energy Environ. Sci. 2024, 17, 4907–4928. [Google Scholar] [CrossRef]
  39. Lu, N.; Yan, X.; Wu, B.; Kobayashi, H.; Li, R. A universal molecular oxygen-mediated photocatalysis strategy to boost visible-light induced hydrogen evolution through partial water splitting. Appl. Catal. B Environ. 2025, 360, 124536. [Google Scholar] [CrossRef]
  40. Ahmad, I.; Shukrullah, S.; Naz, M.; Ahmad, M.; Ahmed, E.; Liu, Y.; Hussain, A.; Iqbal, S.; Ullah, S. Recent advances and challenges in 2D/2D heterojunction photocatalysts for solar fuels applications. Adv. Colloid Interface Sci. 2022, 304, 102661. [Google Scholar] [CrossRef]
  41. Zhu, B.; Sun, J.; Zhao, Y.; Zhang, L.; Yu, J. Construction of 2D S-scheme heterojunction photocatalyst. Adv. Mater. 2024, 36, 2310600. [Google Scholar] [CrossRef] [PubMed]
  42. Akinoglu, E.M.; Hoogeveen, D.A.; Cao, C.; Simonov, A.N.; Jasieniak, J.J. Prospects of Z-scheme photocatalytic systems based on metal halide perovskites. ACS Nano 2021, 15, 7860–7878. [Google Scholar] [CrossRef] [PubMed]
  43. Pan, S.; Li, J.; Wen, Z.; Lu, R.; Zhang, Q.; Jin, H.; Zhang, L.; Chen, Y.; Wang, S. Halide perovskite materials for photo (electro) chemical applications: Dimensionality, heterojunction, and performance. Adv. Energy Mater. 2022, 12, 2004002. [Google Scholar] [CrossRef]
  44. Wang, L.; Bie, C.; Yu, J. Challenges of Z-scheme photocatalytic mechanisms. Trends Chem. 2022, 4, 973–983. [Google Scholar] [CrossRef]
  45. Huang, X.; Du, R.; Zhang, Y.; Ren, J.; Yang, Q.; Wang, K.; Ni, Y.; Yao, Y.; Soomro, R.; Guo, L. Modulating charge oriented accumulation via interfacial chemical-bond on In2O3/Bi2MoO6 heterostructures for photocatalytic nitrogen fixation. J. Colloid Interface Sci. 2024, 664, 33–44. [Google Scholar] [CrossRef]
  46. Chang, P.; Wang, Y.; Wang, Y.; Zhu, Y. Current trends on In2O3 based heterojunction photocatalytic systems in photocatalytic application. Chem. Eng. J. 2022, 450, 137804. [Google Scholar] [CrossRef]
  47. Leelavathi, H.; Muralidharan, R.; Abirami, N.; Tamizharasan, S.; Sankeetha, S.; Kumarasamy, A.; Arulmozhi, R. Construction of step-scheme g-C3N4/Co/ZnO heterojunction photocatalyst for aerobic photocatalytic degradation of synthetic wastewater. Colloids Surf. A Physicochem. Eng. Asp. 2023, 656, 130449. [Google Scholar] [CrossRef]
  48. Zhang, B.; Hu, X.; Liu, E.; Fan, J. Novel S-scheme 2D/2D BiOBr/g-C3N4 heterojunctions with enhanced photocatalytic activity. Chin. J. Catal. 2021, 42, 1519–1529. [Google Scholar] [CrossRef]
  49. Zhang, L.; Zhang, J.; Yu, H.; Yu, J. Emerging S-scheme photocatalyst. Adv. Mater. 2022, 34, 2107668. [Google Scholar] [CrossRef]
  50. Qi, K.; Imparato, C.; Almjasheva, O.; Khataee, A.; Zheng, W. TiO2-based photocatalysts from type-II to S-scheme heterojunction and their applications. J. Colloid Interface Sci. 2024, 675, 150–191. [Google Scholar] [CrossRef]
  51. Zhou, T.; Ma, Y.; Feng, H.; Lu, Y.; Che, G.; Liu, C.; Lan, Y. COFs-Based Metal-Free Heterojunctions for Solar-to-Chemical Energy Conversion. Adv. Funct. Mater. 2024, 34, 2409396. [Google Scholar] [CrossRef]
  52. Chen, R.; Gan, W.; Guo, J.; Lu, Y.; Ding, S.; Liu, R.; Zhang, M.; Sun, Z. Internal electric field and oxygen vacancies synergistically boost S-scheme VO/BiOCl-TiO2 heterojunction film for photocatalytic degradation of norfloxacin. Chem. Eng. J. 2024, 489, 151260. [Google Scholar] [CrossRef]
  53. Rao, V.N.; Ahn, C.W.; Lee, Y.; Shankar, M.V.; Kwon, H.; Kim, K.; Rezakazemi, M.; Kim, S.j.; Yang, J.M. Insights into excitons manipulation in metal chalcogenides based Nano-heterojunction Photocatalysts: A breakthrough in green hydrogen production. Coord. Chem. Rev. 2025, 522, 216176. [Google Scholar] [CrossRef]
  54. Li, Y.; Xia, Z.; Yang, Q.; Wang, L.; Xing, Y. Review on g-C3N4-based S-scheme heterojunction photocatalysts. J. Mater. Sci. Technol. 2022, 125, 128–144. [Google Scholar] [CrossRef]
  55. Xu, Q.; Zhang, L.; Cheng, B.; Fan, J.; Yu, J. S-scheme heterojunction photocatalyst. Chem 2020, 6, 1543–1559. [Google Scholar] [CrossRef]
  56. Wang, Z.; Peng, Q.; Huang, X.; Ma, Q.; Shao, J.; Shen, Q. Recent progress of acenaphthylene-imide-fused polycyclic aromatic hydrocarbons: Synthesis and application. Dye. Pigment. 2021, 185, 108877. [Google Scholar] [CrossRef]
  57. Tan, J.; Zhang, G.; Ge, C.; Liu, J.; Zhou, L.; Liu, C.; Gao, X.; Narita, A.; Zou, Y.; Hu, Y. Electron-deficient contorted polycyclic aromatic hydrocarbon via one-pot annulative π-extension of perylene diimide. Org. Lett. 2022, 24, 2414–2419. [Google Scholar] [CrossRef]
  58. Chen, S.; Slattum, P.; Wang, C.; Zang, L. Self-assembly of perylene imide molecules into 1D nanostructures: Methods, morphologies, and applications. Chem. Rev. 2015, 115, 11967–11998. [Google Scholar] [CrossRef]
  59. Zhou, C.; Xia, W.; Huang, D.; Cheng, M.; Zhang, H.; Cai, T.; Xiong, W.; Yang, Y.; Song, B.; Wang, W.; et al. Strategies for enhancing the perylene diimide photocatalytic degradation activity: Method, effect factor, and mechanism. Environ. Sci. Nano 2021, 8, 602–618. [Google Scholar] [CrossRef]
  60. Fan, Q.; Cheng, K.; Yang, Z.; Zhang, R.; Yang, M.; Hu, X.; Ma, X.; Bu, L.; Lu, X.; Xiong, X. Perylene-diimide-based nanoparticles as highly efficient photoacoustic agents for deep brain tumor imaging in living mice. Adv. Mater. 2014, 27, 843. [Google Scholar] [CrossRef]
  61. Chang, Z.Y.; Wang, Z.Y.; Zhang, R.; Yu, L. Acceleration of biotic decolorization and partial mineralization of methyl orange by a photo-assisted n-type semiconductor. Chemosphere 2022, 291, 132846. [Google Scholar] [CrossRef] [PubMed]
  62. Chen, P.; Blaney, L.; Cagnetta, G.; Huang, J.; Wang, B.; Wang, Y.; Deng, S.; Yu, G. Degradation of ofloxacin by perylene diimide supramolecular nanofiber sunlight-driven photocatalysis. Environ. Sci. Technol. 2019, 53, 1564–1575. [Google Scholar] [CrossRef] [PubMed]
  63. Li, M.; Zajaczkowski, W.; Velpula, G.; Jänsch, D.; Graf, R.; Marszalek, T.; Parekh, S.H.; Zagranyarski, Y.; Mali, K.; Wagner, M. Transformation from helical to layered supramolecular organization of asymmetric perylene diimides via multiple intermolecular hydrogen bonding. Chem. Sci. 2020, 11, 4960–4968. [Google Scholar] [CrossRef]
  64. Zhang, Q.; Jiang, L.; Wang, J.; Zhu, Y.; Pu, Y.; Dai, W. Photocatalytic degradation of tetracycline antibiotics using three-dimensional network structure perylene diimide supramolecular organic photocatalyst under visible-light irradiation. Appl. Catal. B Environ. 2020, 277, 119122. [Google Scholar] [CrossRef]
  65. Langhals, H.; Jona, W.; Einsiedl, F.; Wohnlich, S. Self-dispersion: Spontaneous formation of colloidal dyes in water. Adv. Mater. 1998, 10, 1022–1024. [Google Scholar] [CrossRef]
  66. Goerl, D.; Zhang, X.; Würthner, F. Molecular assemblies of perylene bisimide dyes in water. Angew. Chem. Int. Ed. 2012, 51, 6328–6348. [Google Scholar] [CrossRef]
  67. Liu, K.; Xu, Z.; Yin, M.; Yang, W.; He, B.; Wei, W.; Shen, J. A multifunctional perylenediimide derivative (DTPDI) can be used as a recyclable specific Hg2+ ion sensor and an efficient DNA delivery carrier. J. Mater. Chem. B 2014, 2, 2093–2096. [Google Scholar] [CrossRef]
  68. Xu, Z.; Guo, K.; Yu, J.; Sun, H.; Tang, J.; Shen, J.; Müllen, K.; Yang, W.; Yin, M. A unique perylene-based DNA intercalator: Localization in cell nuclei and inhibition of cancer cells and tumors. Small 2014, 10, 4087–4092. [Google Scholar] [CrossRef]
  69. Kohl, C.; Weil, T.; Qu, J.; Müllen, K. Towards highly fluorescent and water-soluble perylene dyes. Chem. Eur. J. 2004, 10, 5297–5310. [Google Scholar] [CrossRef]
  70. Zhong, L.; Xing, F.; Shi, W.; Yan, L.; Xie, L.; Zhu, S. Synthesis, spectra, and electron-transfer reaction of aspartic acid-functionalized water-soluble perylene bisimide in aqueous solution. ACS Appl. Mater. Interfaces 2013, 5, 3401–3407. [Google Scholar] [CrossRef]
  71. Peneva, K.; Mihov, G.; Nolde, F.; Rocha, S.; Hotta, J.-i.; Braeckmans, K.; Hofkens, J.; Uji-i, H.; Herrmann, A.; Müllen, K. Water-soluble monofunctional perylene and terrylene dyes: Powerful labels for single-enzyme tracking. Angew. Chem. Int. Ed. 2008, 47, 3372–3375. [Google Scholar] [CrossRef] [PubMed]
  72. Battagliarin, G.; Davies, M.; Mackowiak, S.; Li, C.; Müllen, K. Ortho-functionalized perylenediimides for highly fluorescent water-soluble dyes. Chem Phys Chem 2012, 13, 923–926. [Google Scholar] [CrossRef] [PubMed]
  73. Sun, M.; Müllen, K.; Yin, M. Water-soluble perylenediimides: Design concepts and biological applications. Chem. Soc. Rev. 2016, 45, 1513–1528. [Google Scholar] [CrossRef] [PubMed]
  74. Zhou, W.; Liu, G.; Yang, B.; Ji, Q.; Xiang, W.; He, H.; Xu, Z.; Qi, C.; Li, S.; Yang, S. Review on application of perylene diimide (PDI)-based materials in environment: Pollutant detection and degradation. Sci. Total Environ. 2021, 780, 146483. [Google Scholar] [CrossRef]
  75. Sun, T.; Song, J.; Jia, J.; Li, X.; Sun, X. Real roles of perylenetetracarboxylic diimide for enhancing photocatalytic H2-production. Nano Energy 2016, 26, 83–89. [Google Scholar] [CrossRef]
  76. Yu, Y.; Zhu, G.; Lan, L.; Chen, J.; Zhu, X.; Duan, J.; Cong, S.; Li, Z.; Wang, Y.; Wang, Z.J. n-Type glycolated imide-fused polycyclic aromatic hydrocarbons with high capacity for liquid/solid-electrolyte-based electrochemical devices. Adv. Funct. Mater. 2023, 33, 2300012. [Google Scholar] [CrossRef]
  77. Hao, Y.; Zhu, X.; Dong, Y.; Zhang, N.; Wang, H.; Li, X.; Ren, X.; Ma, H.; Wei, Q. Self-assembled perylene diimide (PDI) nanowire sensitized In2O3@MgIn2S4 S-scheme heterojunction as photoelectrochemical biosensing platform for the detection of CA15–3. Anal. Chem. 2024, 96, 13197–13206. [Google Scholar] [CrossRef]
  78. Liu, L.; Wu, Y.; Song, R.; Zhang, Y.; Ma, Y.; Wan, J.; Zhang, M.; Cui, H.; Yang, H.; Chen, X. Morphology engineering and photothermal effect derived from perylene diimide based derivative for boosting photocatalytic hydrogen evolution of ZnIn2S4. J. Colloid Interface Sci. 2022, 628, 701–711. [Google Scholar] [CrossRef] [PubMed]
  79. Gao, W.; Bai, Y.; Wang, X.; Fu, H.; Zhao, P.; Zhu, P.; Yu, J. Self-standing perylene diimide covalent organic framework membranes for trace TMA sensing at room temperature. J. Colloid Interface Sci. 2024, 663, 262–269. [Google Scholar] [CrossRef]
  80. Sun, Y.; Wang, D.; Zhu, Y. Deep degradation of pollutants by perylene diimide supramolecular photocatalyst with unique bi-planar π-π conjugation. Chem. Eng. J. 2022, 438, 135667. [Google Scholar] [CrossRef]
  81. Che, W.; Sun, C.; Wu, Z.; Sun, Y.; Shang, Q. Efficient separation of photo-generated carriers for in-situ induction of PDI cation radicals to enhance the photocatalytic performance of PDI supramolecules. J. Clean. Prod. 2024, 453, 142235. [Google Scholar] [CrossRef]
  82. Li, Z.; Liu, F.; Lu, Y.; Hu, J.; Feng, J.; Shang, H.; Sun, B.; Jiang, W. Molecular design of perylene diimide derivatives for photocatalysis. ACS Catal. 2025, 15, 1829–1840. [Google Scholar] [CrossRef]
  83. Fan, Y.; Kong, C.; Zhang, L.; Wu, H.; Li, J.; Guo, J.; Yi, Q. Enhancing photocatalytic hydrogen evolution performance for D-π-A conjugated polymers based on the perylene diimide. Sep. Purif. Technol. 2025, 355, 129721. [Google Scholar] [CrossRef]
  84. Zhao, P.; Hu, Y.; An, X.; Ji, R.; Liu, H.; Zhao, H.; Song, W.; Dong, Y.; Wang, X. Polymeric PDI-based photocatalytic nanoarchitectures promoting the performance of thin film composite membrane for forward osmosis water purification. Chem. Eng. J. 2023, 476, 146747. [Google Scholar] [CrossRef]
  85. Dong, T.; Dong, G.; Han, K.; Chen, C.; Hu, J.; Uvdal, K. All-organic heterojunctions used for the excellent photocatalytic H2O2 synthesis: The key role of bay-position Cl in PDI. Appl. Catal. B Environ. 2024, 354, 124144. [Google Scholar] [CrossRef]
  86. Yang, S.; Deng, X.; Chen, P.; Li, G.; Wang, Q.; Wang, Q.; Yin, S.-F. Bridges engineering manipulated exciton dissociation and charge separation in small acceptors of PDI supramolecular for boosting photocatalytic nitrogen fixation. Chem. Eng. J. 2022, 441, 136084. [Google Scholar] [CrossRef]
  87. Yang, J.; Miao, H.; Wei, Y.; Li, W.; Zhu, Y. π-π Interaction between self-assembled perylene diimide and 3D graphene for excellent visible-light photocatalytic activity. Appl. Catal. B Environ. 2019, 240, 225–233. [Google Scholar] [CrossRef]
  88. Zhang, K.; Wang, J.; Jiang, W.; Yao, W.; Yang, H.; Zhu, Y. Self-assembled perylene diimide based supramolecular heterojunction with Bi2WO6 for efficient visible-light-driven photocatalysis. Appl. Catal. B Environ. 2018, 232, 175–181. [Google Scholar] [CrossRef]
  89. Gao, Q.; Xu, J.; Wang, Z.; Zhu, Y. Enhanced visible photocatalytic oxidation activity of perylene diimide/g-C3N4 n-n heterojunction via π-π interaction and interfacial charge separation. Appl. Catal. B Environ. 2020, 271, 118933. [Google Scholar] [CrossRef]
  90. Dai, W.; Jiang, L.; Wang, J.; Pu, Y.; Zhu, Y.; Wang, Y.; Xiao, B. Efficient and stable photocatalytic degradation of tetracycline wastewater by 3D polyaniline/perylene diimide organic heterojunction under visible light irradiation. Chem. Eng. J. 2020, 397, 125476. [Google Scholar] [CrossRef]
  91. Wang, S.; Xia, Y.; Yan, G.; Chen, M.; Wang, X.; Wu, L.; Liang, R. PDI bridged MIL-125(Ti)-NH2 heterojunction with frustrated Lewis pairs: A promising photocatalyst for Cr(VI) reduction and antibacterial application. Appl. Catal. B Environ. 2022, 317, 121798. [Google Scholar] [CrossRef]
  92. Chen, H.; Zeng, W.; Liu, Y.; Dong, W.; Cai, T.; Tang, L.; Li, J.; Li, W. Unique MIL-53(Fe)/PDI Supermolecule Composites: Z-Scheme Heterojunction and Covalent Bonds for Uprating Photocatalytic Performance. ACS Appl. Mater. Interfaces 2021, 13, 16364–16373. [Google Scholar] [CrossRef] [PubMed]
  93. Wang, X.; Meng, J.; Yang, X.; Hu, A.; Yang, Y.; Guo, Y. Fabrication of a perylene tetracarboxylic diimide-graphitic carbon nitride heterojunction photocatalyst for efficient degradation of aqueous organic pollutants. ACS Appl. Mater. Interfaces 2018, 11, 588–602. [Google Scholar] [CrossRef]
  94. Wang, L.; Liu, X.; Ji, L.; Luo, Q.; Duan, Y.; An, J.; Chen, X.; Zhang, Y.; Ren, J.; Wang, D. Resin with short-range π-π stacking aggregates for an efficient photocatalyst. Chem. Eng. J. 2022, 433, 134502. [Google Scholar] [CrossRef]
  95. Miao, H.; Yang, J.; Wei, Y.; Li, W.; Zhu, Y. Visible-light photocatalysis of PDI nanowires enhanced by plasmonic effect of the gold nanoparticles. Appl. Catal. B Environ. 2018, 239, 61–67. [Google Scholar] [CrossRef]
  96. Yang, J.; Miao, H.; Jing, J.; Zhu, Y.; Choi, W. Photocatalytic activity enhancement of PDI supermolecular via π-π interaction and energy level adjusting with graphene quantum dots. Appl. Catal. B Environ. 2021, 281, 119547. [Google Scholar] [CrossRef]
  97. Wang, R.; Liu, J.; Wang, B.; Yang, R.; Zhu, S.; Song, Y.; Hua, Y.; Yan, J.; Cheng, M.; Xu, H.; et al. Noble-metal-free Co-N-graphene/PDI for significant enhancement of photocatalytic performance. J. Alloys Compd. 2022, 925, 166370. [Google Scholar] [CrossRef]
  98. Wei, Y.; Ma, M.; Li, W.; Yang, J.; Miao, H.; Zhang, Z.; Zhu, Y. Enhanced photocatalytic activity of PTCDI-C60 via π-π interaction. Appl. Catal. B Environ. 2018, 238, 302–308. [Google Scholar] [CrossRef]
  99. Zhang, F.; Li, W.; Jiang, T.; Li, X.; Shao, Y.; Ma, Y.; Wu, J. Real roles of perylene diimides for improving photocatalytic activity. RSC Adv. 2020, 10, 23024–23037. [Google Scholar] [CrossRef]
  100. Sheng, Y.; Miao, H.; Jing, J.; Yao, W.; Zhu, Y. Perylene diimide anchored graphene 3D structure via π-π interaction for enhanced photoelectrochemical degradation performances. Appl. Catal. B Environ. 2020, 272, 118897. [Google Scholar] [CrossRef]
  101. Wu, G.; Tai, G.; Li, G.; Lu, J.; Pan, Y.; Han, J.; Xing, W. Self-assembled perylene diimide decorated g-C3N4 heterojunction catalyst with strong interfacial charge transfer through π-π interaction for efficient boosted photocatalytic degradation of tetracycline. Surf. Interfaces 2024, 53, 105008. [Google Scholar] [CrossRef]
  102. Li, Y.; Fang, Y.; Cao, Z.; Li, N.; Chen, D.; Xu, Q.; Lu, J. Construction of g-C3N4/PDI@MOF heterojunctions for the highly efficient visible light-driven degradation of pharmaceutical and phenolic micropollutants. Appl. Catal. B Environ. 2019, 250, 150–162. [Google Scholar] [CrossRef]
  103. Wu, M.; Yang, H.; Wu, Q.; He, Z.; Wang, S. Directional and rapid electron transfer in perylene diimide modified iron- manganese bimetallic metal-organic frameworks for enhanced photo-Fenton process. J. Environ. Chem. Eng. 2024, 12, 112246. [Google Scholar] [CrossRef]
  104. Lu, Z.; Li, B.; Wei, B.; Zhou, G.; Xu, Y.; Zhang, J.; Chen, H.; Hua, S.; Wu, C.; Liu, X. NMP-induced surface self-corrosion- assisted rapid spin-coating method for synthesizing imprinted heterojunction photocatalyst anchored membrane towards high-efficiency selective degradation tetracycline. Sep. Purif. Technol. 2023, 314, 123609. [Google Scholar] [CrossRef]
  105. Zhuang, H.; Wang, F.; Shi, K.; Yang, K. Designed Synthesis of PDI/BiOCl-BiPO4 Composited Material for Boosted Photocatalytic Contaminant Degradation. Catalysts 2023, 13, 688. [Google Scholar] [CrossRef]
  106. Shi, K.; Zhou, M.; Wang, F.; Li, X.; Huang, W.; Lu, K.; Yang, K.; Yu, C. Perylene diimide/iron phthalocyanine Z-scheme heterojunction with strong interfacial charge transfer through π-π interaction: Efficient photocatalytic degradation of tetracycline hydrochloride. Chemosphere 2023, 329, 138617. [Google Scholar] [CrossRef]
  107. Zeng, W.; Cai, T.; Liu, Y.; Wang, L.; Dong, W.; Chen, H.; Xia, X. An artificial organic-inorganic Z-scheme photocatalyst WO3@Cu@PDI supramolecular with excellent visible light absorption and photocatalytic activity. Chem. Eng. J. 2020, 381, 122691. [Google Scholar] [CrossRef]
  108. Mao, Z.; Luo, P.; Ling, J.; Zhu, X.; Sun, K.; Cao, Y.; Zhu, D.; Liu, W. Laser preparation of dual Z-scheme heterojunctions PDI/WO3/α-Fe2O3 to enhance adsorption-photocatalytic synergistic degradation. J. Alloys Compd. 2025, 1011, 178195. [Google Scholar] [CrossRef]
  109. Sun, F.; Xie, Y.; Xu, D.; Liu, F.; Qi, H.; Ma, Q.; Yang, Y.; Yu, H.; Yu, W.; Dong, X. Electrospun self-supporting double Z-scheme tricolor-typed microfiber oriented-heterostructure photocatalyst with highly effective hydrogen evolution and organic pollutants degradation. J. Environ. Chem. Eng. 2023, 11, 109169. [Google Scholar] [CrossRef]
  110. Cai, T.; Zeng, W.; Liu, Y.; Wang, L.; Dong, W.; Chen, H.; Xia, X. A promising inorganic-organic Z-scheme photocatalyst Ag3PO4/PDI supermolecule with enhanced photoactivity and photostability for environmental remediation. Appl. Catal. B Environ. 2020, 263, 118327. [Google Scholar] [CrossRef]
  111. Li, X.; Liu, J.; Huang, J.; He, C.; Feng, Z.; Chen, Z.; Wan, L.; Deng, F. All Organic S-Scheme Heterojunction PDI-Ala/S-C3N4 Photocatalyst with Enhanced Photocatalytic Performance. Acta Phys. Chim. Sin. 2020, 37, 2010030. [Google Scholar] [CrossRef]
  112. Lu, Z.; Ren, Y.; Wang, P.; Xu, Y.; Zhang, J.; Wei, B.; Zhou, G.; Liu, X.; Huang, Y.; Wu, C. High-throughput imprinted non-metal S-scheme heterojunction self-cleaning membrane with tight adhesion via dopamine for selective photodegradation of TC. J. Environ. Chem. Eng. 2023, 11, 109745. [Google Scholar] [CrossRef]
  113. Jing, L.; Xu, Y.; Xie, M.; Liu, Y.; Du, X.; Hu, J. Photothermal-assisted S-scheme PDIs/C, N, S-CeO2 derived from MOF-808 (Ce) heterojunction for photocatalytic removal of antibiotics. J. Alloys Compd. 2024, 979, 173568. [Google Scholar] [CrossRef]
  114. Jing, L.; Xu, Y.; Xie, M.; Liu, Y.; Du, X.; Hu, J. Rational construction of visible-light-driven perylene diimides/Fe2O3@C derived from MIL-88A (Fe) heterojunction with S-scheme electron transfer pathway to activate peroxymonosulfate for degradation of antibiotics. J. Colloid Interface Sci. 2024, 659, 520–532. [Google Scholar] [CrossRef] [PubMed]
  115. Xu, Y.; Zhu, X.; Yan, H.; Wang, P.; Song, M.; Ma, C.; Chen, Z.; Chu, J.; Liu, X.; Lu, Z. Hydrochloric acid-mediated synthesis of ZnFe2O4 small particle decorated one-dimensional Perylene Diimide S-scheme heterojunction with excellent photocatalytic ability. Chin. J. Catal. 2022, 43, 1111–1122. [Google Scholar] [CrossRef]
  116. Xiao, Y.; Wang, Z.; Li, M.; Liu, Q.; Liu, X.; Wang, Y. Efficient Charge Separation in Ag/PCN/UPDI Ternary Heterojunction for Optimized Photothermal-Photocatalytic Performance via Tandem Electric Fields. Small 2024, 20, 2306692. [Google Scholar] [CrossRef]
  117. Li, D.; Zhang, Y.; Gao, C.; Wen, Q.; Ma, X.; Song, F.; Zhou, J. Photocatalysis and peroxymonosulfate activation by dual Z-scheme g-C3N4/PDI/Co-Fe Prussian blue analogue for doxycycline hydrochloride removal. J. Environ. Chem. Eng. 2025, 13, 115422. [Google Scholar] [CrossRef]
  118. Wang, W.; Li, X.; Deng, F.; Liu, J.; Gao, X.; Huang, J.; Xu, J.; Feng, Z.; Chen, Z.; Han, L. Novel organic/inorganic PDI-Urea/BiOBr S-scheme heterojunction for improved photocatalytic antibiotic degradation and H2O2 production. Chin. Chem. Lett. 2022, 33, 5200–5207. [Google Scholar] [CrossRef]
  119. Yang, L.; Hao, X.; Yu, D.; Zhou, P.; Peng, Y.; Jia, Y.; Zhao, C.; He, J.; Zhan, C.; Lai, B. High visible-light catalytic activity of Bis-PDI-T@TiO2 for activating persulfate toward efficient degradation of carbamazepine. Sep. Purif. Technol. 2021, 263, 118384. [Google Scholar] [CrossRef]
  120. Jia, Y.; Li, H.; Duan, L.; Gao, Q.; Zhang, H.; Li, S.; Li, M. Activation of persulfate by β-PDI/MIL-101(Fe) photocatalyst under visible light toward efficient degradation of sulfamethoxazole. Chem. Eng. J. 2024, 481, 148588. [Google Scholar] [CrossRef]
  121. Jia, Y.; Duan, L.; Li, H.; Zhang, C.; Gao, Q.; Zhang, H.; Li, S.; Li, M. Fast removal of sulfamethoxazole by MIL-101(Fe)–NH2/perylene diimide activated persulfate under visible light. Sep. Purif. Technol. 2025, 358, 130292. [Google Scholar] [CrossRef]
  122. Chen, X.; Wang, Z.; Shen, X.; Zhang, Y.; Lou, Y.; Pan, C.; Zhu, Y.; Xu, J. A plasmonic Z-scheme Ag@AgCl/PDI photocatalyst for the efficient elimination of organic pollutants, antibiotic resistant bacteria and antibiotic resistance genes. Appl. Catal. B Environ. 2023, 324, 122220. [Google Scholar] [CrossRef]
  123. Gao, X.; Gao, K.; Li, X.; Shang, Y.; Fu, F. Hybrid PDI/BiOCl heterojunction with enhanced interfacial charge transfer for a full-spectrum photocatalytic degradation of pollutants. Catal. Sci. Technol. 2020, 10, 372–381. [Google Scholar] [CrossRef]
  124. Dai, S.; Xu, Y.; Zhang, W.; Li, S.; Guo, Q.Y.; Cui, J.; Song, Y.; Yuan, J.; Peng, W.; Huang, M. S-scheme enhanced photocatalysis on titanium oxide clusters functionalized with soluble perylene diimides. J. Mater. Chem. A 2022, 10, 20248–20253. [Google Scholar] [CrossRef]
  125. Wang, H.; Zhou, Y.; Wang, J.; Li, A.; Corvini, P. BiOBr/Bi4O5Br2/PDI constructed for visible-light degradation of endocrine disrupting chemicals: Synergistic effects of bi-heterojunction and oxygen evolution. Chem. Eng. J. 2022, 433, 133622. [Google Scholar] [CrossRef]
  126. Zha, K.; Li, L.; Zhang, J.; Tang, S.; Li, X.; Hai, J.; Fan, D.; Li, M.; Liu, Y.; Lu, Z. Investigation the influence of bay substitution with perylene diimide on the photocatalytic performance of perylene–diimide/TiO2 composites. J. Photochem. Photobiol. A Chem. 2024, 451, 115517. [Google Scholar] [CrossRef]
  127. Zhang, Z.; Liu, J.; Gu, P.Y.; Ji, R.; Jin, L.; Zhou, S.; He, J.; Chen, D.; Xu, Q.; Lu, J. Preparation of a Bi12O15Cl6@W18O49@g-C3N4/PDI heterojunction with dual charge transfer paths and its photocatalytic performance for phenolic pollutants. Sep. Purif. Technol. 2022, 287, 120539. [Google Scholar] [CrossRef]
  128. Zhang, X.; Shi, L.; Zhang, Y. Preparation of organic-inorganic PDI/BiO2-x photocatalyst with boosted photocatalytic performance. J. Taiwan Inst. Chem. Eng. 2022, 132, 104111. [Google Scholar] [CrossRef]
  129. Xu, T.; Zhang, S.; Zhang, W.; Shi, L. Facile preparation of PDI nano-rods coupled with AgBr for enhanced photocatalytic performance. Opt. Mater. 2024, 147, 114656. [Google Scholar] [CrossRef]
  130. Zhang, X.; Shi, L.; Yao, L.; Cui, L. The boosted photocatalytic activity over perylene diimide modified Bi2O4 hybrid photocatalyst with internal electric field. Mater. Res. Bull. 2022, 146, 111589. [Google Scholar] [CrossRef]
  131. Mardiroosi, A.; Mahjoub, A.R.; Fakhri, H.; Boukherroub, R. Design and fabrication of a perylene dimiide functionalized g-C3N4@UiO-66 supramolecular photocatalyst: Insight into enhancing the photocatalytic performance. J. Mol. Struct. 2021, 1246, 131244. [Google Scholar] [CrossRef]
  132. Zhang, S.; Zhang, X.; Wang, Z.; Yao, L.; Shi, L. Preparation of 1D/1D perylene diimide nano-rod modified Bi4O7 rod with highly active photocatalytic performance. Opt. Mater. 2023, 138, 113734. [Google Scholar] [CrossRef]
  133. Zhu, L.; Chen, Y.; Shen, Y.; Zhang, Y.; Men, D.; Qiu, L.; Xu, X.; Xi, J.; Li, P.; Duo, S. g-C3N4/PDI@ZnIn2S4 2D/2D organic– inorganic hybrid heterojunction with enhanced visible light photocatalytic property. Chem. Phys. Lett. 2023, 833, 140936. [Google Scholar] [CrossRef]
  134. Tang, R.; Gong, D.; Deng, Y.; Xiong, S.; Deng, J.; Li, L.; Zhou, Z.; Zheng, J.; Su, L.; Yang, L. π-π Stacked step-scheme PDI/g-C3N4/TiO2@Ti3C2 photocatalyst with enhanced visible photocatalytic degradation towards atrazine via peroxymono- sulfate activation. Chem. Eng. J. 2022, 427, 131809. [Google Scholar] [CrossRef]
  135. Ren, Y.; Tian, Y.; Lu, Y.; Nie, D.; Zhu, H.; Yang, X. Z-scheme H-PDI supermolecule/NH2-MIL-101(Fe) for enhanced malathion degradation: Mechanism, pathway, and toxicity assessment. J. Environ. Chem. Eng. 2024, 12, 114358. [Google Scholar] [CrossRef]
  136. Chen, R.; Lou, H.; Pang, Y.; Yang, D.; Qiu, X. Enhancing Pollutant Mineralization through Organic–Inorganic Defect-Transit Dual S-scheme with a Robust Internal Electric Field. Small 2024, 20, 15. [Google Scholar] [CrossRef]
  137. Ji, Q.; Cheng, X.; Kong, X.; Sun, D.; Wu, Y.; Xu, Z.; Liu, Y.; Duan, X.; He, H.; Li, S.; et al. Visible-light activation of persulfate ions by Z-scheme perylene diimide/MIL-101(Cr) heterojunction photocatalyst towards efficient degradation of iohexol. Chem. Eng. J. 2022, 435, 134947. [Google Scholar] [CrossRef]
  138. Ren, J.; Meng, Y.; Zhang, X.; Gao, Y.; Liu, L.; Zhou, X.; Zhang, Z.; Zeng, L.; Ke, J. Self-assembled perylene diimide modified NH2-UiO-66 (Zr) construct n-n heterojunction catalysts for enhanced Cr (VI) photocatalytic reduction. Sep. Purif. Technol. 2022, 296, 121423. [Google Scholar] [CrossRef]
Figure 2. (a) Schematic illustration of Ag3PO4/PDIsm synthesis; (b) PL spectra of samples; (c) transient photocurrent responses of samples; (d) EIS Nyquist plots of samples [110]. (e) Possible selective photodegradation mechanism of TC by IM-NSH-PM [112].
Figure 2. (a) Schematic illustration of Ag3PO4/PDIsm synthesis; (b) PL spectra of samples; (c) transient photocurrent responses of samples; (d) EIS Nyquist plots of samples [110]. (e) Possible selective photodegradation mechanism of TC by IM-NSH-PM [112].
Catalysts 15 00565 g002
Figure 3. (a) Schematic illustrating the synthesis procedures of Ag/PCN/UPDI; (b) the schematic of photocatalytic OTC degradation; (c) schematic of charge transfer in Ag/PCN/UPDI upon light irradiation [116]. (d) Schematic diagram of g-C3N4/PDI/Co-Fe PBA activating PMS to degrade DOX under visible light irradiation [117].
Figure 3. (a) Schematic illustrating the synthesis procedures of Ag/PCN/UPDI; (b) the schematic of photocatalytic OTC degradation; (c) schematic of charge transfer in Ag/PCN/UPDI upon light irradiation [116]. (d) Schematic diagram of g-C3N4/PDI/Co-Fe PBA activating PMS to degrade DOX under visible light irradiation [117].
Catalysts 15 00565 g003
Figure 5. (a) The photocatalytic removal of ATZ in the PCT/PMS/Vis system and (b) the corresponding reaction rate constant. (c) ESR spectra of the •OH and SO4 radical; (d) ESR spectra of the 1O2; (e) the schematics illustrate the type-II heterojunction and S-scheme heterojunction mechanism [134]. (f) The synthesis process of PM, the cage structure of MIL-101(Cr), and the colors of MIL-101(Cr), SA-PDI, and PM-7; (g) the possible mechanism of the PM/PS/Vis system [137].
Figure 5. (a) The photocatalytic removal of ATZ in the PCT/PMS/Vis system and (b) the corresponding reaction rate constant. (c) ESR spectra of the •OH and SO4 radical; (d) ESR spectra of the 1O2; (e) the schematics illustrate the type-II heterojunction and S-scheme heterojunction mechanism [134]. (f) The synthesis process of PM, the cage structure of MIL-101(Cr), and the colors of MIL-101(Cr), SA-PDI, and PM-7; (g) the possible mechanism of the PM/PS/Vis system [137].
Catalysts 15 00565 g005
Table 3. The photocatalytic degradation performance of other pollutants by a PDI-based heterojunction.
Table 3. The photocatalytic degradation performance of other pollutants by a PDI-based heterojunction.
PhotocatalystSynthesis MethodPollutantsLight SourceTime
(min)
Efficiency
(%)
TypeReference
PDI/BiO2-xUltrasonic,
self-assembly
RhBVisible2098.7IIZhang et al. [128]
PDISA/AgBrCo-precipitationRhBVisible2097.8IIXu et al. [129]
PDI/Bi2O4Water bath heating, ultrasonicRhBVisible2598.6IIZhang et al. [130]
PCN@UiO-66Solvent thermalRhBVisible14099.0IIMardiroosi
et al. [131]
SAPDI/Bi4O7Self-assemblyRhBVisible2087.6IIZhang et al. [132]
g-C3N4/PDI@ZnIn2S4Oil bath heatingRhBVisible12083.9ZZhu et al. [133]
PDI/g-C3N4/TiO2@Ti3C2CalcinationATZVisible6075.0STang et al. [134]
H-PDI/NH2-MIL-101(Fe)HydrothermalMASimulated sunlight18091.1ZRen et al. [135]
In2O3/PDI/In2S3Self-assemblySL420 nm8080.9SChen et al. [136]
PDI/MIL-101(Cr)Water bath heatingIOHVisible35100.0ZJi et al. [137]
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Song, X.; Lou, J.; Huang, Y.; Chen, Y. Recent Advances in PDI-Based Heterojunction Photocatalysts for the Degradation of Organic Pollutants and Environmental Remediation. Catalysts 2025, 15, 565. https://doi.org/10.3390/catal15060565

AMA Style

Song X, Lou J, Huang Y, Chen Y. Recent Advances in PDI-Based Heterojunction Photocatalysts for the Degradation of Organic Pollutants and Environmental Remediation. Catalysts. 2025; 15(6):565. https://doi.org/10.3390/catal15060565

Chicago/Turabian Style

Song, Xiaofang, Jiahui Lou, Yaqiong Huang, and Yijiang Chen. 2025. "Recent Advances in PDI-Based Heterojunction Photocatalysts for the Degradation of Organic Pollutants and Environmental Remediation" Catalysts 15, no. 6: 565. https://doi.org/10.3390/catal15060565

APA Style

Song, X., Lou, J., Huang, Y., & Chen, Y. (2025). Recent Advances in PDI-Based Heterojunction Photocatalysts for the Degradation of Organic Pollutants and Environmental Remediation. Catalysts, 15(6), 565. https://doi.org/10.3390/catal15060565

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop