Next Article in Journal
The Influence of Au Loading and TiO2 Support on the Catalytic Wet Air Oxidation of Glyphosate over TiO2+Au Catalysts
Previous Article in Journal
Staphylococcus aureus Alkaline Protease: A Promising Additive for Industrial Detergents
Previous Article in Special Issue
Recent Advances in Advanced Oxidation Processes for Degrading Pharmaceuticals in Wastewater—A Review
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Photocatalytic Application of Polymers in Removing Pharmaceuticals from Water: A Comprehensive Review

by
Sanja J. Armaković
1,2,
Stevan Armaković
2,3,* and
Maria M. Savanović
1,2
1
University of Novi Sad, Faculty of Sciences, Department of Chemistry, Biochemistry and Environmental Protection, 21000 Novi Sad, Serbia
2
Association for the International Development of Academic and Scientific Collaboration (AIDASCO), 21000 Novi Sad, Serbia
3
University of Novi Sad, Faculty of Sciences, Department of Physics, 21000 Novi Sad, Serbia
*
Author to whom correspondence should be addressed.
Catalysts 2024, 14(7), 447; https://doi.org/10.3390/catal14070447
Submission received: 30 May 2024 / Revised: 2 July 2024 / Accepted: 9 July 2024 / Published: 12 July 2024

Abstract

:
This comprehensive review covers recent advancements in utilizing various types of polymers and their modifications as photocatalysts for the removal of pharmaceutical contaminants from water. It also considers polymers that enhance the photocatalytic properties of other materials, highlighting their dual role in improving water purification efficiency. Over the past decades, significant progress has been made in understanding the photocatalytic properties of polymers, including organic, inorganic, and composite materials, and their efficacy in degrading pharmaceuticals. Some of the most commonly used polymers, such as polyaniline, poly(p-phenylene vinylene), polyethylene oxide, and polypyrole, and their properties have been reviewed in detail. Physical modification techniques (mechanical blending and extrusion processing) and chemical modification techniques (nanocomposite formation, plasma modification techniques, surface functionalization, and cross-linking) have been discussed as appropriate for modifying polymers in order to increase their photocatalytic activity. This review examines the latest research findings, including the development of novel polymer-based photocatalysts and their application in the removal of pharmaceutical compounds, as well as optimization strategies for enhancing their performance. Additionally, challenges and future directions in this field are discussed to guide further research efforts.

Graphical Abstract

1. Introduction

The aqueous environment plays a central role in the water cycle and the world’s ecology and is a major component of various forms of human activities [1]. Water pollution with residue amounts of various chemicals, including pharmaceuticals, pesticides [2], dyes, etc., creates a lot of serious environmental problems [3,4,5]. The main reason for the spread of pharmaceuticals is the treatment of various diseases, as well as the so-called pro-pharmaceutical effects [6]. The appearance of pharmaceutical substances in the environment also concerns the emission of veterinary-connected preparations [7,8]. The impact of various factors on the behavior and properties of pharmaceuticals in wastewater, natural bodies (rivers, lakes, seas), and technical water has significantly expanded our understanding [9].
Today, the most common technologies used for the removal of pharmaceutically active compounds include reverse osmosis, ultrafiltration, nanofiltration, activated carbon, and many others [10,11,12]. However, these processes do not entirely remove pharmaceuticals from the water. The inefficiency of conventional methods in wastewater treatment and the removal of pharmaceutical compounds from wastewater have become a global concern [13,14]. Advanced oxidation processes (AOPs) showed great potential and played a prominent role in the complete removal of pharmaceuticals from water sources and the environment. One of the most often used AOPs is photocatalysis [15]. In this process, photocatalytic materials, whose activation is induced by light emission, take a primary role. The application of nanotechnologies for removing pharmaceuticals from water is widely used [14]. Nano-sized photocatalysts, such as TiO2 [16], ZnO [17], etc., are widely used. However, nowadays, new polymer materials [18] and their modifications are used very often, side by side with TiO2 and ZnO [18,19].
Since the number of research studies regarding the efficiency of polymers has expanded in recent years, there is a need for a more comprehensive review of polymers as a promising sustainable material for photocatalytic applications in water treatment [20,21]. Therefore, the current work intends to encompass all significant studies on sustainable polymer materials for the elimination of pharmaceuticals from water and wastewater. This review is divided into several sections. Information on sources, problems, occurrences, and potential risks of pharmaceutical compounds are discussed first. The next section presents a critical examination of various polymer materials and their modifications used in photocatalytic processes from both experimental and computational approaches. Further, the efficiency of different polymer modifications in the removal of pharmaceuticals from the environment is addressed in detail. The most important advantages and challenges of using polymer materials are addressed, as well as their limitations.

1.1. The Problem of Pharmaceutical Pollutants in Water

The production and consumption of pharmaceuticals have increased dramatically in recent decades [22], leading to the emergence of water-soluble and pharmacologically active organic micropollutants [23,24]. These pollutants can be found in water sources and ecosystems as parent compounds or as metabolites [25]. Pharmaceuticals present in the environment can change the composition of the natural flora and fauna [26], as well as the structure of the entire ecosystem in the aquatic environment [27]. They are frequently found in wastewater (Table 1), surface water, and groundwater, where they have raised concerns due to their environmental impact and the ineffectiveness of conventional treatment plants in removing them [28]. The presence of pharmaceutical pollutants can lead to the pollution of ecosystems [29] and the exchange of carbon, nitrogen, and phosphorus within cyclical processes. One of the responses to this environmental challenge is the emergence of new technology and materials, in particular, the use of AOPs [30,31,32,33]. The appearance of pharmaceuticals in the environment can have severe consequences and can generate resistance to certain drugs in aquatic organisms, humans, and animals [34].
From an economic point of view, the cost of removing pharmaceuticals in water treatment plants [35], which requires the development of increasingly efficient treatment technologies, is very high, since pharmaceuticals are present as residues in these waters [36,37]. In order to satisfy current demands without affecting the environment, it is essential to provide a long-term solution for pharmaceutical removal from the environment based on social, technical, economic, and environmental conditions in a sustainable way [38,39,40,41]. Among AOPs, photocatalysis stands out [42]. One of the new materials utilized in photocatalysis is sustainable polymers [43,44]. Photocatalytic polymer-based degradation stands out as a potential approach for the removal of pharmaceuticals from the environment [45] since it typically requires little energy, has low operation costs, and, most importantly, does not produce any toxic by-products that cause secondary contamination [46]. Also, polymer materials can interact with the transformation products made in the photocatalytic degradation process. The advantages of polymer-assisted preparation procedures are mainly the speed of the synthesis and maturation around the matrix surface and improved polymer adhesion/anchorage of the anchored oxides. However, little is known about the effect of the formation of small molecules as products of photocatalytic reactions in situ on the anchored oxides or their interaction with the matrix-to-supported oxide binding. The interactions of photodegradation products with polymers can be of importance for many potential products. The products of photocatalytic degradation of pollutants are often deposited or adhered to matrix materials like polymers. This may affect their properties, determined by chemical as well as physical interaction and binding to the surface. Polymer-modified properties of those materials, as matrices for photocatalytic degradation, have recently been successfully applied for wastewater purification, self-cleaning surfaces, and surfaces with photobiocidal properties. This would allow for an extremely easy one-step preparation of photocatalytic-polymer based reactive materials, used for degradation of both parent compounds and photodegradation products, thus backing up the economy of their implementation [47,48,49].
Table 1. Detection of pharmaceuticals in different wastewater.
Table 1. Detection of pharmaceuticals in different wastewater.
PharmaceuticalLocationAnalysis Methodc (ng/dm3) Detected in Effluents Reference
AmoxicillinKisii and Kabarnet wastewater treatment plants in KenyaHPLC–UV/Vis90[50]
AmpicillinOlomouc region, Czech RepublicUPLC–TQ-S/MS48[51]
AzithromycinSanya, China wastewater treatment plantsHPLC–UV/Vis56,600[52]
CarbamazepineCzech wastewater treatment plantsUHPLC–MS/MS730[53]
CephalexinPharco B
International Pharmaceutical Company in
Borg El Arab, Alexandria, Egypt
HPLC–DAD71,900[54]
CeftazidimePharco B
International Pharmaceutical Company in
Borg El Arab, Alexandria, Egypt
HPLC–DAD94,000[54]
CeftriaxonePharco B
International Pharmaceutical Company in
Borg El Arab, Alexandria, Egypt
HPLC–DAD170,300[54]
ClarithromycinNeringa (Nida) wastewater treatment plantsHPLC–UV/Vis15[55]
ChloramphenicolKanchipuram District, ChennaiGCE39[56]
CiprofloxacinSanya, China wastewater treatment plantsHPLC–UV/Vis24,000[52]
DoxycyclineMachakos, wastewater stabilization pondsLC-ESI–MS/MS1500[57]
ErythromycinSouth Africa, untreated wastewaterLC–MS/MS55[58]
MetoprololCzech wastewater treatment plantsUHPLC–MS/MS960[53]
MetronidazoleCzech wastewater treatment plantsUHPLC–MS/MS7500[53]
NaproxenCzech wastewater treatment plantsUHPLC–MS/MS980[53]
NorfloxacinMachakos, wastewater stabilization ponds LC–ESI–MS/MS2900[57]
OxytetracyclineDhaka wastewater effluents in BangladeshRP–HPLC–UV/Vis670[59]
ParacetamolCzech wastewater treatment plantsUHPLC–MS/MS82,500[53]
PropranololSan Luis Province, ArgentinaHPLC–DAD850[60]
RoxithromycinNicosia wastewater, Cyprus SPE–DEX700[61]
SulfadimidineShanghai wastewater treatment plant UPLC–DAD440[62]
SulfamethoxazoleCzech wastewater treatment plantsUHPLC–MS/MS7500[53]
TetracyclineBouc-bel-air, France wastewaterUPLC–FLD4820[63]
TramadolCzech wastewater treatment plantsUHPLC–MS/MS3500[53]
TrimethoprimCzech wastewater treatment plantsUHPLC–MS/MS3000[53]

1.2. Objectives and Significance of the Study

Photocatalysis by semiconductors has been recognized as one of the most acceptable, sustainable, and eco-friendly techniques to mitigate increasing pollution of the environment [64]. However, commercialization of these advanced technologies is challenging due to the poor activity of pure semiconductor photoactive material, charge recombination, and light absorption only in the ultraviolet region of the solar spectrum [65,66]. Evolution in the materials science area has led to the use of various polymers as adequate support for semi-conductor photocatalysts, providing solutions for the challenges mentioned above [67,68]. This might be attributed to the very low recombination of photogenerated charge carriers (electron–holes, e–h+), which possess a long carrier lifetime due to the heavy adsorption of semiconductor nanocrystals onto polymer support [69]. The introduction of a polymer barrier band can increase photocatalysts’ response to visible light owing to the narrow band gap, and they can also act as effective scavengers of photo-generated holes [70].
Photocatalysis is a green technology [71], and from this perspective, polymers are concurrently applied without employing supporting semiconductors for wastewater purification [72]. Polymers have several distinct advantages as photocatalysts [73] over conventional metal oxides, semiconductors, and inorganic nanocrystals, including turgid backbones, controlled porosity, excellent chemical durability, well-defined micropores, strong chelation sites for favorable interactions with transition metal centers, and flexibility in maintaining conjugation in extended structures [18,74,75]. Organic polymers could provide an alternative to toxic and bioaccumulative heavy-metal-containing semiconductors, especially in biomedical and environmental remediation applications [76]. Still, polymeric materials exhibit low absorption in the visible spectrum of sunlight [77]. This problem can be dealt with by introducing photoactive species, including conducting polymers and TiO2 [78]. Currently, conjugated polymers are recognized for their additional application as visible-light active photocatalysts owing to their long excited-state lifetimes, high quantum yields, and outstanding photostability [79,80]. The major challenge is enhancing the photocatalytic degradation of recalcitrant organic waste in environments where stable, reliable, and benign aromatic (homo)polymers can offer structural and functional sustainability under a wide range of operational conditions [81].
In the history of the development of organic and inorganic chemistry, inorganic nanoparticles, nano capsules, and nano complexes have been used in combination with organic polymers as reaction media for many organic modifications. Their easy methods of preparation have been mediated by chemical and physical procedures [82]. Organic polymers are used as supported/reducing/dispersing/capping agents in most cases of synthesis of such inorganic nano catalysts as reaction media. Polymers are seemingly the best choice as reaction media in the presence of nano catalysts due to their ease of separation, recyclability, low cost, and non-corrosive, reusable, non-toxic nature in mediated synthesis. Polymers have been widely used as reaction media in combination with various nano catalysts in different organic transformations [83]. Organic nano catalysts, polymer nanocomposites, and support polymers have been comprehensively used in combination with molecular catalysts, nano catalysts, and metal-free catalysts. Such polymer-supported nano catalysts have offered a unique benefit in organic synthesis with the development of many magnetic, non-magnetic, and semi-magnetic complex nano hetero molecules, which have been used in many organic transformations and minimize the volume of toxic waste and disposal of products due to their reusability in the organic transformation, and offer economy and safety along with the environmental sustainability of the desired products [84]. The use of TiO2 nanoparticles is attracting increased attention in various technological, environmental, and biomedical applications. In turn, polymers, with their broad spectrum of physical characteristics, processability, tailorable biodegradation, various related substances, responsiveness to external stimuli, and low production cost, are widely used in different areas such as medicine, biotechnologies, environmental uses, drug delivery, dentistry, cosmetics, clothing, packaging, and rarefied gas separation [85]. ZnO is an essential technological material. The popularity of ZnO is due to the fact that zinc oxide nanoparticles possess and manifest various physical properties, as well as the nature and characteristics that are a consequence of the peculiarities of nanoparticle formation and growth. Furthermore, ZnO also has broad applications relative to various sectors and could be used to realize a variety of sensors. ZnO and polymers have a number of properties, the combination of which allows for the realization of a wide range of composite material properties. According to the requirements of protective coatings, the most essential property is protection against ultraviolet and visible light. Some polymers have an enhanced absorption capacity in these particular ranges. In many areas, polymers can effectively compete with more expensive inorganic materials [86,87].

2. Polymers for Photocatalytic Applications

Rapid advancements in interdisciplinary scientific research on hybrid green functional materials, incorporating functional elements for advanced electronic, ionic, and photonic functionalities, continue to facilitate elusive clean energy production, toxic pollutant clean-ups, and safe biologically relevant reductions and hydrogenation chemistries [88,89]. The cost-effective synthesis, facile band manipulation, and straightforward nano-modularity of polymeric photocatalysts are, broadly speaking, the main reasons for increasing research and application attributes [90]. The combination of a synthetic environment and tapping malleable surfaces of photocatalysts often causes an active photocatalytic face to interact charge distribution with electro-photonic and integrated photogenerated product storage and cycles [91].
The versatility of polymers in the industry lies in the ability to combine different compounds to form polymers and mold them into any desired shape, and subsequently apply desired properties, or make chemical modifications to obtain value-added polymers. Polymers can be used in the form of powders, membranes, foams, surface loads, etc. [92,93,94,95]. Polymer powders are defined as finely divided, solid, spherical plastics that are made from combinations of bulk or solution polymerization and processing technology. These spherical particles may have diameters ranging from a few micrometers to over 200 μm. Powdered polymers are essential in terms of form, economics, aesthetics, and chemistry. For practical use, a polymer powder must have a basic elemental composition, suitable physical characteristics, and uniform properties when sampled [96]. Polymer membranes are generally synthesized from organic solvent mixtures involving polymers and additives and need to be prepared in films. According to its intended application, the polymer type/dosage, concentration, crosslinking density, film thickness, and any additives can be chosen to meet different aims, and can include materials like polysulfone, polyetherimide, cellulose acetate, poly(vinylidene fluoride), and others. Utilizing polymer membranes could be preferable owing to their exceptional properties such as structure, surface modifications/enhancements, hydrophobic/hydrophilic balance, size, function, and cost for specific applications. Despite the possibility of using them in different ways, polymer membranes have some disadvantages, such as fouling and low water flux [97]. During the current era of agricultural and environmental concerns, interest in biodegradable and renewable polymer foams is increasing. Importantly, the properties of porous foam are intimately linked with the processing technology used in foam formation. Therefore, manufacturing processes and techniques for novel polymer foams are constantly improving. Due to their unique properties and performance, polymer foams have achieved significant improvements in product performance, cost-effectiveness, weight reduction, etc. [98]. Polymers, having a wide variety of properties and chemical compositions, are currently serviceable in a number of applications, including packaging, electrical insulation, and protective coatings. Many advantageous and valuable properties of polymers are realized in the performance of surface functions. The characteristics of polymers’ surfaces are often complex, qualitatively different, and sometimes unpredictable from data on their bulk state. Polymeric materials are classified into semicrystalline, amorphous, or glassy forms; usually, polymers have complex chemical, physical, and mechanical natures, and exhibit a great range of properties. Polymer surface engineering is both necessary and interesting from a variety of technological perspectives. It can involve structural organization at several length scales, ranging from the small or nanostructured surface topographies (length scale < 100 nm) of coatings and adhesive adhesion devices to macroscopic-scale patterns (length scale > 1000 μm) [99].
Due to the distinctive chemical and physical properties of polymeric materials, they are considered suitable and promising candidates for AOPs [100,101]. The advantage of using polymers as sensitizing agents in photocatalytic composites is that they may contribute to improving photocatalytic efficiency due to π-conjugated systems containing high concentrations of electron-rich species, which are available for transfer in the semiconductor’s conduction band [102]. Also, it has been proved that by using polymer materials, recombination of photo-generated charge carriers is reduced, while conductive polymers tune the semiconductor’s band gap [103]. Based on density functional theory models, conducting polymers such as polyaniline, poly(p-phenylene vinylene), polyethylene oxide, polypyrrole, polythiophene, poly(1-naphthylamine), polycarbazole, polyacetylene, poly(3,4-ethylenedioxythiophene), and poly(o-phenylenediamine) exhibit visible absorption activity due to their low band gap values, which makes them ideal candidates for extending the semiconductor’s range of light absorption [104,105]. Also, conjugated mesoporous polymers and, subsequently, covalent triazine frameworks are widely used for visible light-induced photocatalysis [106]. These bulk systems are large-scale conjugated systems with defined nanostructures, maximizing surface area and, therefore, allowing an efficient photocatalytic process [107].
Over the past few decades, research endeavors have been intensively devoted to finding suitable photocatalysts for practical applications [108]. Among them, polymers have gained attention as promising candidates for many photocatalytic applications (Figure 1) [19]. Some of the most often used polymers in photocatalytic reactions are polyaniline, poly(p-phenylene vinylene), polyethylene oxide, polypyrrole, polythiophene, poly(1-naphthylamine), polycarbazole, polyacetylene, poly (3,4-ethylenedioxythiophene), poly(o-phenylenediamine, etc. [104,105].

2.1. Types of Polymers Used in Photocatalysis

The fundamental classification of polymers:
i.
Inorganic polymers. They represent aluminosilicate-based materials, which are conventionally produced by the activation of solid alkali silicate. Geopolymers are inorganic polymers, considered a subset of inorganic polymers since the aluminosilicate network is coordinated tetrahedrally and primarily has aluminum atoms and silicon [109]. These materials are flammable, durable, resistant to rust, and non-poisonous. Inorganic polymers based on silicon are widely employed in water purification treatment due to their large surface area, electrical conductivity, stability, flexibility, and reusability. The inorganic polymers with silicon and phosphorous in their structures have large bond angles and bond lengths, while their bonds are more ionic [110,111]. Even at low temperatures, they retain elasticity. They can withstand very high temperatures in an oxygenated atmosphere since these polymers have more inorganic elements. Inorganic polymers are highly resistant to homolytic dissociation. As a consequence of their inorganic networks, these polymers are resistant to UV and high-energy irradiation, owing to their optical transparency throughout a narrow wavelength range and stronger inorganic connections [112]. Typically, in an inorganic polymer, the microstructural elements are built from three-dimensional arrays of oxygen, E-O-E, where E is an element such as Si, Al, P, or B, through linkage with either oxygen or other heteroatoms [113]. The mechanism that joins the silicon–oxygen–silicon links, characteristic of siloxanes, has substantially varying hydrophobic or hydrophilic stability according to the size and the polar character of the molecular affine groups, which are silicon polymeric units in a nonpolar environment, differing from the glycol groups. These polymeric membranes polymerize the silicon and oxygen cycles near the parent alkoxide, which is produced with high reactivity, and which can be used in polyester acrylate copolymeric coatings. Inorganic silicon-based polymers must undergo heat treatment for reactivity and their final application (biomaterials, photocataysis, optical devices, energy storage) [114]. Polymers containing a main-chain of boron are an interesting and important class of materials that have applications in several fields such as liquid-crystal materials, photoresist materials, nonlinear optics, boron neutron capture therapy, fire retardants, and inorganic binders. Boron-based polymers and inorganic polyboron clusters with isolable boron–boron, boron–oxygen, and boron–carbon linking atoms have long attracted widespread interest in both pi-conjugated all-boron and pi-conjugated all-boron-containing polymers to expand the palette of functional molecular and polymeric materials, including (opto) electronic, (non) linear optics, magnetic, electrical, and chiral properties. In addition to the inherent upper chalcogenide boron clusters, hexacyclic structures are valuable as a building block for the design of advanced high-level functional building blocks, with the stabilization of the resultant polymers using flexible moieties, porphyrins, and donor-acceptor [115]. While a small number of chemical elements, such as carbon and silicon, have become highly important in the development of new types of polymers, other atoms that are used much less frequently, such as phosphorus, can build polymers with unique and distinct properties. Of particular note for phosphorus-based polymers is the inherent flame-retardant nature of certain phosphinated materials. Such polymers find utility in a number of applications, including the electronics sector and soft materials. Phosphorus-based polymers exhibit a wide range of properties such as flame retardancy, mechanical properties, thermal stability, proton or anion conductivity, and anticorrosive resistance when compared with traditional organic or inorganic polymers [116].
ii.
Organic polymers. Porous organic polymers have a porous framework that is impregnated in various dimensions. These polymers arise from covalent bonds between the organic monomers of different orientations and shapes. They can be categorized as crystalline or amorphous depending on the difference in the arrangement of monomers in their structure. They are mainly used as photocatalysts for modifications because of their extraordinary stability, inherent porosity, economical designs, and flexibility in structure and function. The nature, pore size, functions, and surface area of these polymers can be altered by modifying their side chains. They are used for heterogeneous catalysis, photocatalytic conversion, biochemical sensors, and gas and energy storage. There are different variations in their structure, so they could be divided into polymers that have homogenous catalysts, such as monomers; polymers that are altered post-synthesis; and polymers that have coordinated metal ions [112,117].
iii.
Inorganic–organic hybrids. The coupling of inorganic materials with an organic polymer matrix produces inorganic–organic hybrid materials. Their mechanical, thermal, and physical properties are enhanced to a high extent. The strength of coupling between inorganic and organic depends on the conjugation with the polymer matrix, the number of particles, and the regular dispersal of particles in the 3D network. These polymers’ thermoplastic or thermosetting behaviors and elasticity are improved by inorganic fillers [118]. This new family of polymer composites combining inorganic and organic systems is prepared in situ using sol–gel processes. By optimizing the conditions for their synthesis, their characteristics can be controlled. The inorganic constituents provide stability to heat and friction and enhance chemical and electromagnetic properties and the refractive index, while the organic components are responsible for producing films. Hybrid systems are improved compared to the individual constituents [112].
Conducting polymers belong to the family of conjugated polymers, which include polyaniline, poly(p-phenylene vinylene), polyethylene oxide, polypyrrole, poly(3,4-ethylene dioxythiophene), poly(2,2′ bithiophene), and polythiophene. They have been shown to be valuable materials in high-tech technologies, such as photocatalysis, sensors, biosensors, nanotechnology, and organic electronic devices [104,105].
Polyaniline (PANI) has been one of the most studied conducting polymers due to its numerous possible applications [119]. The terminology of a quasi-metal univocally designates its electronic properties [120]. PANI is an electrically conductive polymer that is easily prepared using chemical or electrochemical oxidation of aniline, usually in aqueous medium [121]. Due to its electrochemical, electrical, and optical properties and excellent thermal and environmental stability, it has emerged as one of the leading conducting polymers for photocatalytic commercial applications [122]. The main disadvantage of PANI is poor processability both in melt and solution processing as a consequence of its backbone stiffness [123]. PANI salt, emeraldine hydrochloride, is not soluble in common organic solvents. It is soluble only in concentrated sulfuric acid. However, recent discoveries have found that functionalized protonic acids like camphor sulfonic and dodecylbenzene sulfonic acid, together with phosphoric acid esters, enable polyaniline salts to dissolve in organic solvents [124]. PANI can have its physical properties precisely tailored to meet the requirements of various applications by applying several synthetic methods. The potential of the PANI thin film on conductive electrode surfaces as an efficient photocatalyst and the possibility of using it as a photocatalyst in different conjunctions have been studied by many authors [125,126,127,128]. In order to achieve suitable environmental and thermal stability, three doping forms of PANI have been designed using post-treatment oxidation methods. Newly synthesized PANI materials (emeraldine base, pernigraniline form, and leucoemeraldine base) have been studied by many characterization techniques such as SEM, TEM, EDS, EPR, XRD, XPS, DRS, and FTIR methods. The results of these studies revealed these materials’ high photocatalytic activity in pharmaceutical degradation. The emeraldine base’s low band gap, policrystallinity, and the fact that it possessed the most active sites made it the best candidate in comparison with the other two materials [129,130,131,132,133].
Poly(p-phenylene vinylene) (PPV) possesses long-term π-conjugation and photophysical properties, which are suitable for photocatalysis [134]. Moreover, the easy chemical modification and synthesis strategy of PPV can obtain various derivatives [135]. Therefore, PPV is a promising candidate for different applications like photocatalytic degradation, CO2 reduction, H2 production, and organic synthesis [136]. PPV is also an essential conjugated polymer used in various optoelectronic devices, such as solar cells, light-emitting diodes, organic transistors, and chemical sensors [134]. PPV-based materials exhibit both p and n doping properties. These materials can be oxidized in the positive potential range [137]. When the polaron is formed, the positive charge is delocalized over the average of the four monomer units. Also, oxidizing PPV units undergo disproportionation processes leading to the formation of localized dications, which are separated by four neutral monomer moieties. The PPV oxidation process is followed by the incorporation of supporting electrolyte anions into the polymeric network. Also, PPV and its derivatives can be reduced at negative potentials to form negative polarons and dope the polymeric phase with the cations of the supporting electrolyte [134]. PPV is utilized in a wide range of applications, including molecular imaging, light-emitting devices, biosensors, and anticancer therapy. Recently, there have been an increasing number of studies on the application of PPV-based materials in the field of photocatalysis [138,139,140,141]. When deposited on an appropriate material, PPV is able to significantly improve its photocatalytic response, allowing the controlled formation (or retention) of exciting matter [142]. PPV and its derivatives have been proven to be important photoactive components for fabricating polymer–inorganic hybrid nanocomposites, which can be easily processed into stable and efficient photoactive coatings that are almost exclusively responsive to visible light within the solar spectrum [138].
Polyethylene oxide (PEO). Hydrophilic polymers have been an essential class of materials for various applications as biocompatible and biodegradable materials [143]. PEO, also known as polyethylene glycol, is a polyether polymer that has recently received a great deal of attention in research and industrial applications in the area of photocatalysis [144]. PEO has been widely used due to its biocompatibility and non-toxicity properties. PEO is a very versatile polymer with a wide range of applications, such as solid polymer electrolytes [145]. As its name indicates, PEO has an oxide group within the main chain; therefore, it is possible to tailor the material according to the requirements of the intended application. The polymer features good dielectric and ion-related properties due to its structure, high solvation power, easy formation of complexes with alkali salts, and the fact that it provides a direct path for cation migration [146]. Considering complex materials design for advanced multifunctional materials for the most strategic sectors, it appears clear that the tailoring of PEO is fundamental for progress in the field of energy storage technologies, like fuel cells and batteries, as well as eco-sustainable solutions for the use of PEOs in water remediation technologies [147].
Polypyrrole (PPy) Electrically conducting polymers have received a considerable amount of attention due to their usability in different applications, such as field-effect transistors, light-emitting diodes, solar cells, and photocatalysis processes [148]. Electrically conducting polymers are amorphous or semi-crystalline organic compounds that exhibit metallic levels of electronic conductivity and high electrochemical and chemical stability, similar to inorganic semiconductors [149]. PPy is a nitrogen-containing five-membered heteroaromatic ring molecule [150]. Five lone-pair electrons are present on the nitrogen atom, which are beneficial for electron transfer, and hence, there is an easy and moderate oxidation of PPy to form PPy+ [151]. This feature is advantageous for the modification and post-synthesis functionalization of the PPy. To improve the potential applications of doped PPy for ions from anionic environmental pollutants, some approaches to reversible tuning of the PPy+ anion exchange membrane to PPy upon removal of the specifically adsorbed ions have been reported in the literature [152,153]. Changes in the electronic structure of PPy can be identified through spectroscopic and electrochemistry-based techniques. Thus, the solid-state near-edge X-ray absorption fine structure spectra revealed an increase in the occupation probability of the electronically conducting π electrons in the partially or fully oxidized PPy [154].

2.2. Modification of Polymers for Enhanced Photocatalytic Activity

The directed modification of synthetic polymers has long been known to significantly expand the range of polymer utility, even when only a few fundamental chemical transformations are available for this purpose [132]. Techniques for polymer modifications to enhance their photocatalytic activity are divided into physical and chemical modification techniques.

2.2.1. Physical Modification Techniques

Mechanical blending. Although a number of methods are available for the preparation of polymer blends, a common one is using a stirrable device in the liquid phase. It is called mechanical blending [155]. In a typical preparation, the polymers to be blended are mixed above the glass transition temperature of the polymers to allow proper mixing and obtain a homogeneous melt. Polymeric blends can also be used to meet requirements for various physical and mechanical characteristics, such as processability, abrasion resistance, impact resistance, and deformability, as well as to meet the needs of biocompatible and environmental formulations with higher biological activity. From these evaluations, toughened plastics, composite materials, films, fibers, elastomers, and gels can be prepared [156]. Therefore, the mechanical blending of polymers is crucial in improving polymers’ properties for the purpose of removing pharmaceutical pollutants by the photocatalytic process. Blending polymers gives stable mixtures that have been used to create a variety of polymers with specific physical and chemical properties for use in some forms of processing. This is particularly attractive because it is generally less expensive to formulate blends than to synthesize entirely new polymers [157].
Extrusion processing. Extrusion and injection molding are widely used methods in polymer processing. In extrusion, polymer melts continuously in a conduit (barrel) and is subsequently forced through a restrictive forming tool. During the extrusion process, polymer melts and is transported through different geometries, takes a complex shape, cools down, and ultimately solidifies into the desired form. For the products of commercial polymers, crystallization usually happens during the extrusion process [158]. In the case of semicrystalline polymers, the most critical factors influencing the properties of the products can be cooling rate, temperature, pressure, and the orientation of the elongated macromolecular chains.
Moreover, extrusion is usually the final process in a complex macromolecular material technology involving operations like the crystallization of newly formed macromolecules [159]. As a result of limiting the operating conditions and the pressure drop in the system (especially at the die hole), the extrudate emerges as a solid, e.g., stock, a profile, a pipe, or a film. Obtained materials are widely used to remove pharmaceutical contaminants from the environment through adsorption or photodegradation. The high value of the die compression ratio means that the effective pressure drop in the die is also enhanced, leading to an increase in viscosity and then an increase in extrusion force [160]. A reduction in the die compression ratio leads to a reduction in the adequate pressure in the die and a correspondingly more significant decrease in viscosity. If speed decreases, a discontinuous flow of material can occur [161].
The ongoing evolution of computer technology has resulted in the ability to perform complex optimization studies on many aspects of single-screw profile extrusion, which would be uneconomical using predictive full 3D numerical simulations. As single-screw extrusion processes of polymer materials are susceptible to numerous influencing parameter combinations, an overall optimization of the distinct technical task is essential [162]. The steady-state computational approach based on a 3D solver has several drawbacks, e.g., computational time and resource use, especially for the simulation of thermal processes with defined temperature boundary conditions. The need for computational optimization to obtain new polymer materials will be illustrated with an example of injection stretch blow molding. The main process steps are illustrated in Figure 2a–f. An injection unit (Figure 2a) fills the cavity of an injection mold (Figure 2b) to produce a pre-form material. This is then transferred to a blow mold (Figure 2c), where it is stretched and blown (Figure 2d) against the contours of the mold cavity. The mold opens (Figure 2e) when the container is sufficiently cold. Modeling of this process to predict the performance of the part for a given set of input conditions (operating conditions, equipment geometry, material properties) typically entails the numerical modeling of each of its steps, followed by expensive process costs. Therefore, before the process starts, modeling the thickness, morphology, and mechanical performance of polymers and their action on pharmaceutical molecules could be predicted. It is suggested that the optimization process be approached backward, as indicated by the curved arrows (Figure 2, from 1 to 5) [159].

2.2.2. Chemical Modification Techniques

Nanocomposite formation. Interaction between a component of the polymer matrix and the nanoparticle surface can be an additional aspect that occurs during the production of classical composites by reinforcing the polymer matrix with filler. Such an interaction, in some aspects, limits the possibility of selecting the filler, significantly changing the internal structure of the polymer nanocomposite and affecting the properties of these materials in the removal of pharmacological compounds from the environment. Nanoparticle dispersion in the polymeric matrix is another essential aspect of research on polymer nanocomposites [163]. Polymer nanocomposites are an important class of materials that are popularly used in various applications due to their characteristic unique behavior in comparison to their pristine polymer forms [164]. The ability to control nanoscale particle dispersion can lead to unique properties (mechanical, electrical, and thermal properties) due either to molecular confinement or control over the structure formation process during nanocomposite production [165]. The most common applications of polymer nanocomposites are in photocatalysis, biomedicine, and other industrial applications [166,167,168].
The dispersion of the nanoparticles, their size, their surface topography, and the chemical properties of the interface with the polymeric matrix all together play an essential role in defining the characteristics of the nanocomposite [169]. These nanocomposites can have rheological, mechanical, and gas barrier properties that surpass the best traditional composites but often also have new functionalities like flame retardancy, transparency, light shielding, photocatalytic activity, or environmental sensors [170]. In the development of complex systems formed by nanoscopic inorganic clusters and organic matrices, it is essential to understand the architecture of the polymer matrix at the level of dynamic and static nanoscopic heterogeneity. In this field, a fundamental issue is to understand how the functional properties of a polymer change due to the complex nanoscale architecture of the matrix formed by the dispersed material [171].
Plasma modification. Process control and plasma generation are the main aspects of polymer modification in many applications, including nanotechnology, material science, and biomedical engineering [172]. The generation of plasma is achieved by the application of an electric field to typically an inert gas, leading to the ionization and excitation of gas molecules. Created plasma is a highly reactive mixture of radicals, electrons, ions, and excited species [173]. To achieve effective polymer modification, control of the plasma generation process and parameters such as exposure time, gas flow rate, power, and pressure are essential. Properties of the plasma, such as its composition, temperature, density, and ultimately the modification of the material surface, can be significantly affected by these parameters [174]. Plasmas offer multiple ways to modify polymer surfaces and thus influence the adhesion environment. Polymer substrates benefit from direct interaction with the plasma medium, and impinging charged and neutral plasma species acquire an organization fitting within two physical processes of interaction, i.e., the formation of permanent polar/aromatic groups or the grafting of other chemical functions via intermediate activation sites [175]. These chemical modifications can affect the polymeric network, generally reducing hydrophobicity and replacing aliphatic molecular moieties with new anchoring points for further chemical surface tailoring [176]. Many techniques are used to generate plasma, such as radiofrequency, alternating current, and direct current. The choice of technique depends on the desired surface properties of the material being modified. For example, radiofrequency plasma is commonly used to modify polymers and remove pharmaceutical compounds through photocatalytic processes [177].
Surface functionalization. Surface functionalization refers to any chemical reaction that alters the surface of a material, or modifies or introduces physicochemical properties without fundamentally changing the bulk characteristics [178]. A properly designed surface can be implemented using various direct surface functionalization processes. The most often used techniques are physisorption and chemisorption of small molecules and manufacturing composite materials [179]. Controlling the surface properties of materials is an essential factor in determining interactions between obtained materials and the environment. The first requirement for surface modification of materials is the introduction of new functionality and physical properties. Surface modification can be performed through various methods like surface-initiated polymerization, high-energy irradiation, and direct immobilization of natural entities [180]. Stereoregular polymers may be produced from optically active monomers. For example, in terms of salient features, chiral polymer production in either a racemic or a scalemic manner may lead to asymmetric surfaces with congruent properties. Another possibility lies in the use of optically active initiators directing the polymerization propagation processes towards the canonical synthesis even of atactic polymers [181]. Fluorescent, electronically active polymers are largely heterogeneous, including the strongly delocalized radical along the polymer chain. These are produced by overcoming the disadvantages of post-deposition functionalization techniques with their confined degree of grafting. Polymer incorporation into a material phase leads to significant changes in the properties of the surface, such as wettability, charge density, or even spatial distribution of functional groups. Various industrial applications are described in the literature: water purification, packaging in pharmaceutical and food industries, wood coatings for protection against sunlight, membranes for gas separation, coatings for textiles and optical sensors, optical information storage, soft contact lenses, etc. [182,183,184].
Cross-linking. Some of the polymers used in composite systems, photocatalytic processes, and other applications are thermoset or cross-linked polymers [185]. Cross-linking polymers containing one or more reactive monomeric subunits can produce materials with three-dimensional net points connected by covalent bonds in a process that is often referred to as curing. Various covalent reactions are used to cross-link polymers, many of which can be considered “click” reactions because they are efficient, regiospecific, and lead to high conversions [186,187]. In particular, they can be used to create environmentally robust, mechanically tough, and heat-resistant materials. By controlling those properties, new materials with unprecedented functionality are expected to emerge, enabling different applications such as all-polymer actuators and novel optical devices. In the last few years, many polymers based on dynamic covalent bonds have been modified and used in many fields because of their superior properties [188,189]. Polymers obtained by cross-linking, with shear self-repairing and thinning properties, are often used in 3D printing, and polymers with antibacterial, conducting, and self-healing properties are often used in photocatalytic reactions [190].

3. Efficiency in Removing Pharmaceutical Pollutants

Photocatalysis is a chemical reaction in which the catalyst is activated by light. Catalysts accelerate a chemical reaction by lowering the activation energy, making the reaction easier. In a conventional chemical reaction, heat is commonly used to activate catalysts. Photocatalysis provides an alternative way to carry out reactions at a reduced temperature, therefore decreasing side products that may be formed. Overall, this yields increased economical and environmentally friendly reactions. The field of photocatalysis has made significant progress, replacing several conventional methods based on its cost-effectiveness and utility [191]. The general mechanism of photocatalysis is based on the use of a semiconductor, which enables the absorption of visible or UV light from the solar spectrum. Absorption of irradiation causes the excitation of electrons from valence band to conduction band and generates e–h+ pairs. In the next step, e and h+ are transferred to the photocatalyst surface for the degradation and oxidization of pollutants to CO2 and H2O [192].
In general, the photocatalysis process involves three significant steps: photocatalyst light absorption, separation of photogenerated charge carriers, and interfacial catalytic redox reactions [193]. When the conducting polymer is irradiated with UV/Visible light, an e from the valence band is excited by a photon of light to the conduction band. This phenomenon is called π-π* electronic transition. The excited e migrates to the surface of the conducting polymer and activates the adsorbed O2 molecules to form superoxide radicals O 2 , which oxidize organic contaminants in the aqueous environment (Figure 3a). Photogenerated h+ can also oxidize organic pollutants directly [194]. In the case of conducting polymer–metal oxide hybrid (Figure 3b), conducting polymer acts as a photosensitizer, and it is able to absorb a wide range of visible irradiation because of the lower band gap compared to metal oxide. The excited e in the lowest unoccupied molecular orbital (LUMO) of conducting polymer chains is injected into a conduction band of transition metal oxide (e.g., TiO2, ZnO), which reacts with an adsorbed water molecule to form O 2 radicals. h+ can react with water to form HO. These polymer modifications, with improved photocatalytic activity, are widely used in the degradation of pharmaceutical compounds by the photodegradation process [81,105].
Researchers are taking a greater interest in the development of new polymeric materials that have a high affinity for harmful organic compounds and also exhibit high fluxes in membrane separation processes for the treatment of natural water, underground, and domestic and industrial sewage. Polymers and polymer-based structures have found a wide range of applications in the removal of pharmaceuticals from the environment [195,196]. Discussions on this topic have been recognized since around 2000. Since the year 2000, 5616 articles related to the use of polymers in photocatalysis have been published, while only 128 articles have referred to the photocatalytic degradation of pharmaceuticals by the use of polymers as catalysts [197]. The main aim of the present work was to comprehensively address studies performed using polymers to achieve effective removal of pharmaceutical contaminants from the environment.
The high synergy between PANI and ZnO was observed by Nosrati et al. [198]. They used PANI/ZnO to remove ampicillin. It was indicated that PANI/ZnO composites could be reused multiple times with a minimal loss of reactivity. ZnO/PANI nanocomposite was also synthesized by Asgari et al. [199] and used in the photocatalytic degradation of metronidazole. The photocatalytic activity of the ZnO/PANI nanocomposite in the degradation of MNZ was 63 times higher than that of ZnO under visible light radiation. Šojić Merkulov et al. [200] investigated the kinetics of photocatalytic degradation of pharmaceuticals in the presence of TiO2/PANI nanocomposite powders. Degradation and mineralization efficiency of propranolol and amitriptyline using TiO2/PANI was high in environmental waters.
The synergistic association of WS2/PANI nanocomposite can degrade pollutants such as antibiotics (nitrofurantoin). This research was conducted by Fatima et al. [201], wherein the WS2/PANI nanocomposite was successfully synthesized via a facile, in situ polymerization technique. By using the in situ oxidative polymerization method, Ujwal et al. [202] synthesized GdFeO3/PANI nanocomposite. The chemical, physical, and structural properties of GdFeO3/PANI enabled electrochemical detection and photocatalytic degradation of an organic pollutant, acetaminophen. Tahir et al. [203] synthesized a novel nanocomposite of Bi2WO6 through PANI intercalation. This nanostructure exhibited effective charge separation and outstanding activity in hydrogen evolution and photocatalytic degradation. Fabricated nanocomposite showed higher stability even after four degradation cycles. PANI/LaFeO3/CoFe2O4 ternary heterojunction was synthesized by in situ polymerization by Kumar et al. [204]. In comparison with the photocatalyst PANI, the synthesized PANI/LaFeO3/CoFe2O4 was shown to be superior. It led to degradation of 92% of the selected pharmaceutical after 120 min of irradiation.
Dai et al. [205] synthesized 3D PANI/Perylene diamide and applied it in pharmaceutical removal using visible irradiation. They provided a promising approach for improving the photocatalytic performance of polymer materials and indicated that 3D PANI/PDI organic heterojunction can be efficiently applied for water treatment remediation. PANi@carbon nanotubes/stainless steel was designed by Peng et al. [206]. Synthesized material showed high efficiency, was environmentally friendly, and stable. It was used and proved to be effective in the degradation of pharmaceutical ibuprofen. Tian et al. [207] synthesized carbonized PANI-activated peracetic acid and used it in the degradation of different pharmaceutical compounds. A high degradation rate of pharmaceuticals was observed, while the primary reactive species for degradation was HO.
Table 2 represents a summary of PANI composites used for photocatalysis of pharmaceutical compounds.
PPy-based composites also exhibit significant photocatalytic properties in pharmaceutical degradation. Silvestri et al. [208] applied PPy-ZnO synthesized via polymerization in the degradation of diclofenac. The PPy-ZnO was demonstrated to be highly active for diclofenac degradation under simulated solar irradiation, even when compared with pure ZnO. Jiadi et al. [209] synthesized TiO2@V2O5-PPy, exhibiting high photocatalytic performance and water stability. This photocatalyst has been used to remove water-soluble organic pollutants with high efficiency under visible light irradiation. The photocatalytic degradation ability of the newly synthesized Fe-TiO2-PPy was investigated under solar light. From the results, it can be concluded that the prepared Fe-TiO2-PPy could be an efficient adsorbent as well as a photocatalyst for the degradation of pharmaceuticals from an aqueous environment [210].
Ultrasound-assisted synthesis of PPy/V2O5 was performed by Zia et al. [211]. The obtained photocatalyst was used in the visible-light-driven photodegradation of erythromycin and ciprofloxacin. The generation of hydroxide and superoxide radicals was enhanced by the use of a newly synthesized composite. Kumar et al. [212] synthesized Fe3O4 nanorods coated in PPy/rGO by a chemical reflux method. Fe3O4@PPy/rGO showed good repeatability, stability, and reproducibility. Photocatalytic studies revealed 84% degradation of acetaminophen by Fe3O4@PPy/rGO in the presence of persulfate. Zhu et al. [213] successfully synthesized high-dispersed PPy@Ag/g-C3N4. The as-prepared PPy@Ag/g-C3N4 was used to remove various organic pollutants from the water. It was established that enhanced photocatalytic activity of PPy@Ag/g-C3N4 came from π-conjugated electronic structures of PPy and Ag species as well as the electron transfer mediator between g-C3N4 and PPy.
The newly synthesized nanocomposite Ag2MoO4/PPy performs as an efficient photocatalyst in the degradation of pharmaceutical ciprofloxacin within 10 min. As prepared, Ag2MoO4/PPy provided a defined morphology, larger surface area, and smaller size, and this improved its electrocatalytic and photocatalytic properties [214]. Das et al. [215] fabricated new PPy-sensitized zinc ferrite/graphitic carbon nitride by an in situ polymerization method. This material showed good photocatalytic performance towards ciprofloxacin degradation and generation of hydrogen energy. It may offer a promising strategy for maximum light absorption that meets environmental claims.
Table 3 represents a summary of PPy composites used for photocatalysis of pharmaceutical compounds.
Besides photocatalytic activity, polymer materials are also used as adsorbents for the efficient removal of pharmaceuticals. The potential and structure of polymeric adsorbents are of much interest in the context of understanding the adsorption phenomenon in solid materials [216,217,218]. The sorption curve or the adsorption isotherm is the most general way to describe the adsorption capacity of the adsorbent or the amount of the adsorbate adsorbed. The specific surface area of polymer materials is the most important parameter characterizing the porous structures of the adsorbent [219]. It is related to the adsorption capacity. In addition, the thermodynamics of the adsorption process are still of great interest, including the enthalpy of adsorption, entropy, and free energy change accompanying the adsorption process. These have provided valuable insight into the adsorption mechanism at the molecular level. Furthermore, the adsorption mechanism of the adsorbent toward the pharmaceuticals in water can be better understood by detecting the surface groups and information about the interactions between the specific groups of the pharmaceutical molecule and the functional ligands on the polymer [220,221].

4. The Role of Atomistic Calculations in Polymer-Assisted Photocatalytic Degradation

The term “atomistic calculations” encompasses a variety of theoretical and computational methods used to perform simulations on molecular and periodic structures. Other names for this class of theoretical tools are molecular modeling, computational chemistry, computational physics, etc. Over the last two decades, scientists have been able to predict a vast number of properties at the atomic level using these calculations and simulations, which are now possible to perform even on simple personal computers [222,223,224]. Atomistic calculations include methods such as molecular dynamics (MD) simulations, ab initio calculations, density functional theory (DFT), semiempirical calculations, and molecular docking, among others [225,226,227]. These techniques allow researchers to study the behavior of materials, predict their properties, and understand complex processes at an unprecedented level of detail. Advancements in computational power and the development of sophisticated algorithms have made it feasible to tackle problems that were once considered highly challenging, leading to significant breakthroughs in materials science, chemistry, and nanotechnology [228].
The power of atomistic calculations is best illustrated by the fact that the first step in the development of novel materials or the improvement of current ones often involves the application of computational methods to predict the properties of structures before they are synthesized. This approach allows researchers to identify the best candidates for physical experimentation, thereby optimizing research and minimizing the use of resources. By using these computational techniques, scientists can efficiently screen a wide range of materials and focus on those with the most promising properties, significantly reducing the time and cost associated with experimental trials. This predictive capability is a fundamental part of modern materials science and engineering, driving innovation and accelerating the discovery of advanced materials.
Over the years, a vast array of methods has been developed for performing calculations on molecules. Nevertheless, three distinct directions of development have emerged, each with its own set of advantages and disadvantages. These methods often intersect and complement each other, forming a comprehensive toolkit for molecular simulations. Their relationships are illustrated in the Venn diagram presented in Figure 4.
In terms of computational modeling, polymers are most frequently treated as molecules rather than periodic structures, so we will refer to molecular calculations hereafter. In general, all basic computational methods for simulations of molecules can be divided into three main groups: classical mechanics, quantum mechanics, and semiempirical methods. Classical mechanics treats molecular systems as classical systems, where a molecule is observed as a collection of balls interconnected with springs. The balls represent atoms, while the springs represent bonds between atoms. This approach utilizes a “force field” to describe the potential energy of the system. Although this approach is relatively simplistic, it is very useful, especially if the equations used to calculate the total energy of a molecule are adequate. However, to derive accurate equations, it is necessary to parameterize the force field using data from a large number of molecules [229].
Quantum mechanical methods offer unprecedented accuracy for theoretical methods, but they come with a significant computational cost. Among these methods, one of the earliest was the Hartree–Fock (HF) method. This method is a typical example of wavefunction methods, meaning that the main task is to find a wavefunction that minimizes the energy of a given system. The main drawback of this method is that it does not account for electron correlations, which is where the density functional theory (DFT) method comes in. DFT is based on electron density, and the main task is to find the electron density that minimizes the energy of a molecule. This method accounts for electron correlation and is by far the most frequently applied computational method for calculations on molecules since it offers the best cost–accuracy ratio. DFT relies on the density functional, a function that determines how the energy of a molecule is calculated. However, based on the fundamental setup of DFT, there is no single, universal density functional suitable for all types of molecules. Literally hundreds of density functionals have been developed so far, and it takes experience to know which functional to apply and when. However, the best-known functional for calculations on molecules is undoubtedly the famous B3LYP [230,231,232,233], while for periodic structure, the best-known functional is certainly the PBE [234]. Post-HF methods offer unprecedented accuracy when it comes to various molecular properties; however, the computational costs associated with these methods prevent them from being applied to a higher extent, compared to DFT methods.
In the last decade, significant effort has been invested in the development of semiempirical methods. These methods are essentially quantum mechanical methods, but with additional approximations in the description of electron density and numerical procedures. For a long time, the parametrized method (PM) family of semiempirical methods [235,236,237,238] was the best choice. However, since 2017, the extended tight binding (xTB) methods developed by the renowned Professor Stefan Grimme and his group have emerged. In many significant aspects, these methods have achieved DFT accuracy at a fraction of the computational time required for calculations. Another important thing to mention is the fact that the GFN family of methods is available in the program called xtb [239,240,241,242,243], completely open source and free to use.
Currently, the latest xTB method is GFN2 [239], and aside from its remarkable accuracy, the superiority of this method over other semiempirical methods lies in its ability to cover the majority of the periodic table of elements (PSE). The parameters for these semiempirical calculations are available for atoms with Z equal up to 86. In terms of xTB methods, the development of GFN methods has also led to the creation of the force field called GFN-FF. This force field stands out not only for its accuracy and speed but also for its automated applicability to the PSE, meaning that the only requirement to run it is the starting molecular structure.
There is a reason why we have devoted significant space to semiempirical methods, particularly to GFN-xTB methods. Polymers are large and highly flexible molecular structures, and their optimization via DFT methods can be quite demanding in terms of computational time. Additionally, highly rotatable functional groups can make vibrational analysis particularly difficult, making it very challenging to identify the ground-state geometry of the polymer chain. In such situations, we warmly recommend a multiscale modeling workflow, i.e., the application of different computational methods (Figure 5).
To achieve faster geometrical optimizations, it is wise to pre-optimize large structures using very fast methods such as GFN2-xTB. This method yields optimized geometries that are very close to those obtained via DFT methods (Figure 6). Ground-state geometries obtained via GFN methods can be further re-optimized with DFT methods. This subsequent optimization is much faster since the pre-optimized structure is already very close to the ground state. This multiscale modeling approach significantly reduces computational time while maintaining high accuracy in the final optimized structures [244].
In Figure 6, we visually compare the optimized geometries of a polymethylmethacrylate (PMMA) polymer chain (consisting of five monomer units) obtained via the GFN2-xTB semiempirical method and the DFT method using the B3LYP-D3 density functional with the 6-31G(d,p) basis set. Both structures appear very similar, with no clearly visible differences in ground-state geometries. However, the GFN2-xTB method took less than five seconds for geometrical optimizations, while the DFT optimization at the specified level took almost two hours.
In the case of very large structures, ground-state geometries obtained via GFN methods can be treated as true ground-state geometries and further subjected to more accurate single-point energy calculations. This approach helps to obtain precise electron distribution, various quantum-molecular descriptors, spectral properties, non-covalent interactions, and more. Such a workflow is illustrated in Figure 5, demonstrating the efficiency and accuracy of combining GFN methods with single-point energy calculations for comprehensive molecular analysis.
One of the most powerful aspects of atomistic calculations in the context of photocatalytic polymers is the ability to estimate light absorption and charge transport properties. These crucial aspects of polymers’ photocatalytic properties can be addressed with molecular orbital theory or time-dependent DFT (TD-DFT) calculations. Accurate DFT or post-HF calculations on geometrically optimized polymer chains yield the energies of the highest occupied and the lowest unoccupied molecular orbitals, HOMO and LUMO, respectively. The energies of these orbitals play a vital role in estimating reactivity and light absorption. The larger the energy gap, the less reactive the molecule, as a larger amount of energy is required to promote electrons from the HOMO to the LUMO. When combining polymers with photocatalytic nanomaterials, it is also important to know the “positions” of the polymers’ HOMO and LUMO orbitals to find compatible combinations of polymers and photocatalytic materials [245].
In terms of improving photocatalytic properties, polymers are frequently used to achieve light sensitization of photocatalytic nanomaterials. In such cases, it is crucial to know the UV/Vis absorption spectrum of a polymer. While UV/Vis spectra can be inexpensively and quickly obtained experimentally, it is necessary to have the polymer physically first. However, synthesizing new structures is often complicated and expensive, making it beneficial to predict the UV/Vis spectrum prior to synthesis to select the best candidates. In this regard, TD-DFT calculations are extremely valuable, as they enable researchers to study excitations and construct UV/Vis spectra. A good illustration of how DFT methods can be applied for comprehensive computational studies of structural, reactive, and optoelectronic properties of polymers can be found in [246,247,248,249].
Regarding computational methods for studying polymers and other molecular structures, it is useful to mention modeling codes and packages currently available. Thanks to the amazing development of both computational codes and computer resources, researchers nowadays have a plethora of computational tools, most of which can be used on simple desktop computers. Computational codes for atomistic calculations can be either open-source, free-for-academics, or commercial [250].
Among open-source solutions, the xtb program of Professor Grimme and coworkers is an amazing tool with extensive possibilities for semiempirical calculations. This and other tools from this group are freely available on the official GitHub page of the Grimme group [251]. In addition to its remarkable functionality, the xtb program is also characterized by very detailed documentation with plenty of tutorials and case studies freely available [252]. Users are also warmly encouraged to visit the official web presentation of the Grimme group to learn more about computational tools and publications arising from them [253].
Online and desktop graphical user interfaces for the xtb program are available within the atomistica.online project. Atomistica.online is a free molecular modeling platform offering online and desktop tools for atomistic calculations, available at atomistica.online. One of the tools is the online GUI for the xtb program, allowing users to run most xtb calculations on a remote server through their web browsers for free. The current limitation is a molecule size of up to 200 atoms, which is sufficient for teaching and most research tasks. Tasks such as molecular dynamics simulations that cannot be run with the online tool can be easily run through the recently developed desktop xtb GUI. The desktop GUI for xtb removes the necessity of running the xtb from the command line and generating running commands or input files [254,255].
Regarding quantum mechanical calculations, the ORCA [256,257,258,259,260,261,262,263,264] code stands as one of the most powerful codes available. It is free for academics and can be obtained through the official forum [265]. This code, developed by Professor Frank Neese and colleagues, implements the latest DFT and wavefunction methods and is interconnected with the xtb code, making it a unique solution on the market. ORCA is well-documented, with plenty of tutorials and case studies, and features user-friendly input file structures. Another tool within the atomistica.online platform is the online input generator, a GUI for easy and user-friendly generation of input files for the ORCA modeling package. Atomistica also offers a desktop GUI for simple running and monitoring of ORCA calculations. All atomistica.online desktop tools are available for both Windows and Linux operating systems [264].
Since surface modifications with polymers play an important role in the functionalization of photocatalytic nanomaterials for water purification, we also mention the possibility of applying a combination of semiempirical and quantum mechanical calculations to study non-covalent interactions between a polymer and photocatalytic material. A detailed analysis of electron density within the molecular system is crucial for this purpose. The famous Multiwfn code [266], developed by Professor Tian Lu and freely available at its official web presentation [267], is extremely useful for this analysis and a vast number of other activities. By examining electron density between atoms, it is possible to identify specific sites where non-covalent interactions are formed and obtain information about their strengths. In particular, the concept of reduced density gradients (RDG) is very useful for this purpose. This approach allows for the identification of areas with specific electron densities indicating attractive or repulsive interactions between atoms [268]. Additionally, with the help of the Multiwfn program, it is possible to obtain RDG scatter plots and RDG surfaces, as shown in Figure 7.
The aforementioned RDG figures help researchers visualize the relationship between electron density and the reduced density gradient to identify non-covalent interactions and regions of weak interactions in a molecular system. The RDG scatter plot, presented in Figure 7a, represents specific points in space within a molecular system, each characterized by its electron density and reduced density gradient values. These dots help identify and visualize regions of non-covalent interactions, such as hydrogen bonds, van der Waals forces, and steric repulsions, by highlighting areas with low density gradients and varying electron densities. For example, a higher number of blue dots indicates a higher number of attractive non-covalent interactions. Similarly, the RDG surface, presented in Figure 7b, represents regions in three-dimensional space where the RDG values indicate the presence of non-covalent interactions. It is color-coded to show the nature and strength of these interactions based on the electron density values.
The atomistica.online project also offers an online tool for easily obtaining RDG scatter plots and files necessary to produce high-quality RDG surfaces through its user-friendly GUI, which runs Multiwfn in the background on a remote server provided by the project developers. Instead of working in the command prompt and combining several programs, users can easily obtain RDG scatter plots and files for RDG surfaces by simply uploading the output file containing the necessary information obtained after quantum mechanical calculations, for example, with ORCA [268].
Commercial solutions come at a price that can be quite high, but the graphical user interfaces, code optimization, and suites of tools for system preparation and data analysis justify the cost and enable users to achieve remarkable productivity. In terms of polymers, the commercial solution offered by Schrödinger, Inc. (New York, NY, USA) [269] within the Materials Modeling Suite [270] provides extremely useful tools for the preparation of polymer systems for quantum mechanical calculations and MD simulations, for running state-of-the-art classical and quantum mechanical calculations, and for the analysis of the obtained data and machine learning. This comprehensive tool enables researchers to study the intrinsic properties of polymers and interactions with photocatalytic materials through both molecular and periodic DFT calculations, allowing them to obtain, for example, band gaps of polymer-functionalized photocatalytic materials, which is of supreme importance for developing novel materials and procedures for water purification. Readers are warmly encouraged to study the case studies and documentation on the corresponding page of the Schrödinger, Inc. website [271].

5. Advantages and Challenges

As already pointed out throughout this manuscript, the use of photocatalytic polymers in the removal of pharmaceuticals from water has drawn significant attention due to their potential for efficient and sustainable water treatment. This chapter explores the advantages and challenges associated with this innovative approach. The benefits of the application of polymers in this area are summarized in the illustration presented in Figure 8.
Due to numerous possibilities during synthesis, polymers can be engineered to enhance photocatalytic efficiency through better light absorption and charge separation. For example, the synthesis of photocatalytically active polymers without any additives or metal co-catalysts is one way to enable the forced degradation of contaminants, as demonstrated in a recent paper by Kumar et al. [272]. In this research, the conductive polymer poly(3,4-ethylenedioxythiophene) (PEDOT) was employed for the photocatalytic removal of metformin, a widely utilized molecule with antidiabetic activity that has been found in wastewater due to overuse.
Another essential advantage of polymer application we would like to emphasize is the potential for developing photocatalytic procedures for the forced degradation of contaminants by combining photocatalytic oxide materials with desired photocatalytic properties with polymers. In this way, photocatalytic nanoparticles are incorporated within a polymer medium, creating hybrid composites. This approach brings numerous improvements, one of them being that photocatalytic nanoparticles are tightly immobilized in the polymer medium [273,274,275], reducing the necessity to filter out nanoparticles later to keep water safe for consumption.
Another advantage essential to showcase is the possibility of functionalizing surfaces with polymers, which allows for targeting specific pharmaceutical pollutants by enhancing the selectivity and effectiveness of the treatment process. For example, in the paper by Park et al. 2013 [276], several polymers were mentioned in the context of modification of TiO2 surface: polyaniline (PANI), polythiophene, poly(3-hexyl-thiophene) (P3HT), and poly(fluorine-co-thiophene) (PFT). For the TiO2 photocatalytic materials, according to a series of studies, these polymers have been demonstrated to be able to increase the absorption of visible light and stability [277], charge transfer [278], charge separation [279], and to sensitize the surface in the visible light area [280,281,282].
Another equally important advantage of polymer applications in the context of photocatalytic removal of pharmaceutical contaminants is related to sustainability and recyclability. Immobilization of photocatalytic nanomaterials within a polymer matrix not only stabilizes the whole system, but also enables multiple uses of the working material. This is crucial because it removes the necessity of filtering out nanoparticles to make the water safe for use. Instead, it is only necessary to remove the entire filter based on the polymer with incorporated photocatalytic nanomaterials [283].
In the context of water purification, it is essential to highlight the non-toxicity of hybrid composites consisting of photocatalytic nanomaterials and polymers. A notable study by Zhao et al. [284] exemplifies this concept by utilizing the well-known PANI to create a novel composite photocatalytic material with BiVO4. This composite not only exhibits superior photocatalytic efficiencies but has also been rigorously tested for toxicity. The material was evaluated against Bacillus subtilis and Staphylococcus aureus, demonstrating non-toxic properties. This finding underscores the potential of such hybrid composites in environmental applications, offering both high efficiency in pollutant degradation and safety for biological systems. The combination of polymers and photocatalytic materials presents a promising approach for developing advanced, eco-friendly photocatalytic materials for effective water purification.
The properties of individual polymers can often be tailored by combination with other materials to achieve a balance of properties that will ensure satisfactory performance under many different conditions. Some applications require a range of mechanical, thermal, electrical, chemical, or optical properties to be achieved in a single product. This diversity of properties is usually achieved by using various additives in the form of discrete particles or fibers. It is generally necessary to modify the surface properties of these particulate or fibrous particles in order to achieve an intimate and enduring association between the additive and the polymer. The large internal surface area available for such associations and the potential increase in macroscopic performance of the filled-polymer systems implies a need for increasing use of heterogeneous oxide fillers. There is also a distinct set of polymers whose degree of crystallinity and mechanical properties are considerably improved with additional phases such as blends or fillers [285,286]. Determining the effect of these different processing techniques, or the properties of the chemically modified phases, on the properties of a heterogeneous polymer often involves the property of interest but also may require extensive measurements such as surface area measurements, surface technique augmentations such as XPS, or the measurement of fundamental packed or porous structures. Characterizing the mechanical properties of these materials is not only of interest from an application perspective, as the most widespread industrial application of polymers is as an engineering material, but it can also tell us more about how the polymers interact at an atomic level [287].
Current challenges include polymer stability, costs, and potential toxicity of degradation products, with one of the greatest being the stability of polymers. Polymers can undergo photo-degradation when exposed to UV or visible light, leading to the breakdown of polymer chains and loss of mechanical and chemical properties. In a recent study by Ran et al. 2018 [288], a well-known conjugated polymer poly(1,3,4-oxadiazole) (POD) was synthesized through readily available synthesis and processing and tested for photocatalytic properties. The authors of the study concluded that although it exhibited excellent photocatalytic properties against selected compounds, its reusability was not ideal because of the relatively poor photostability. However, due to the simplicity of synthesis and processing, further studies and improvements were suggested. In a very interesting study by Chin et al. (2006), different polymeric membranes were tested under conditions typical for photocatalytic degradation, including, among others, the presence of UV light oxidation stress [289]. According to this study, when considering both UV exposure and oxidative test, polytetrafluoroethylene and hydrophobic polyvinylidene fluoride were good choices for photocatalytic applications.
The stability of polymers in water is essential for their use in environmental applications, such as water purification through photocatalysis. Polymers exposed to water, especially under varying pH conditions, can undergo hydrolytic degradation. This involves the breaking of polymer chains due to the reaction with water molecules, which can be accelerated in acidic or alkaline environments. However, in this regard, polymer structures based on polylactic acid (PLA) may offer a solution compared to polymers derived from petroleum [290]. PLA requires much less energy for production compared to petroleum-derived polymers and has the potential to decompose through composting, followed by the release of CO2. On the other hand, PLA has much lower biodegradability, and finding ways to increase the speed of biodegradation was the topic of a recently reported study by Garratt et al. (2023) [291]. In their study, it was demonstrated that the incorporation of α-Fe2O3 into PLA increased the speed of PLA’s degradation, resulting in 32% of the solid transforming to the liquid phase, compared to PLA’s decomposition alone.

6. Conclusions

In conclusion, this comprehensive review highlights the significant advancements and potential of polymer-based photocatalysts in the removal of pharmaceutical contaminants from water. The unique properties of polymers, including their tunable structures, ease of modification, and ability to enhance the photocatalytic efficiency of other materials, make them promising candidates for sustainable water treatment technologies.
The reviewed literature demonstrates that both organic and inorganic polymers, as well as their hybrids, exhibit excellent photocatalytic properties, enabling efficient degradation of pharmaceuticals. Polyaniline, poly(p-phenylene vinylene), polyethylene oxide, and polypyrrole are among the most studied polymers, showing remarkable results in various photocatalytic applications. The modification techniques, including mechanical blending, extrusion processing, nanocomposite formation, plasma modification, surface functionalization, and cross-linking, play a crucial role in enhancing the photocatalytic activity of polymers.
However, despite the promising results, several challenges remain. The stability of polymers under photocatalytic conditions, the cost of synthesis and modification, and the potential toxicity of degradation products are critical issues that need to be addressed. Future research should focus on developing more robust and cost-effective polymer-based photocatalysts with enhanced stability and reduced environmental impact.
Moreover, the integration of computational methods, such as DFT and TD-DFT, can provide valuable insights into the electronic properties and light absorption characteristics of polymers. This approach can guide the design and optimization of new polymeric materials with superior photocatalytic performance.
Polymer-based photocatalysts offer a promising and sustainable solution for the removal of pharmaceuticals from water. Continued interdisciplinary research, combining experimental and computational approaches, will be essential to overcome existing challenges and fully realize the potential of these materials in environmental protection and water purification.

Author Contributions

Conceptualization, S.J.A., S.A. and M.M.S.; Data curation, S.J.A., S.A. and M.M.S.; Funding acquisition, S.J.A. and S.A.; Methodology, S.J.A., S.A. and M.M.S.; Project administration, S.J.A.; Resources, S.J.A., S.A. and M.M.S.; Supervision, S.J.A. and S.A.; Writing—original draft, S.J.A., S.A. and M.M.S. All authors have read and agreed to the published version of the manuscript.

Funding

The authors gratefully acknowledge the financial support of the Ministry of Science, Technological Development and Innovation of the Republic of Serbia (Grants No. 451-03-66/2024-03/200125 & 451-03-65/2024-03/200125).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Gleeson, T.; Wang-Erlandsson, L.; Porkka, M.; Zipper, S.C.; Jaramillo, F.; Gerten, D.; Fetzer, I.; Cornell, S.E.; Piemontese, L.; Gordon, L.J.; et al. Illuminating Water Cycle Modifications and Earth System Resilience in the Anthropocene. Water Resour. Res. 2020, 56, e2019WR024957. [Google Scholar] [CrossRef]
  2. Jazić, J.M.; Đurkić, T.; Bašić, B.; Watson, M.; Apostolović, T.; Tubić, A.; Agbaba, J. Degradation of a Chloroacetanilide Herbicide in Natural Waters Using UV Activated Hydrogen Peroxide, Persulfate and Peroxymonosulfate Processes. Environ. Sci. Water Res. Technol. 2020, 6, 2800–2815. [Google Scholar] [CrossRef]
  3. Sundararaman, S.; Aravind Kumar, J.; Deivasigamani, P.; Devarajan, Y. Emerging Pharma Residue Contaminants: Occurrence, Monitoring, Risk and Fate Assessment—A Challenge to Water Resource Management. Sci. Total Environ. 2022, 825, 153897. [Google Scholar] [CrossRef] [PubMed]
  4. Materon, E.M.; Ibáñez-Redín, G.; Joshi, N.; Gonçalves, D.; Oliveira, O.N.; Faria, R.C. Analytical Detection of Pesticides, Pollutants, and Pharmaceutical Waste in the Environment. In Nanosensors for Environmental Applications; Springer: Cham, Switzerland, 2020; pp. 87–129. [Google Scholar]
  5. Rasheed, T.; Bilal, M.; Nabeel, F.; Adeel, M.; Iqbal, H.M.N. Environmentally-Related Contaminants of High Concern: Potential Sources and Analytical Modalities for Detection, Quantification, and Treatment. Environ. Int. 2019, 122, 52–66. [Google Scholar] [CrossRef]
  6. Dubey, A.; Samra, S.; Sahu, V.K.; Dash, S.L.; Mishra, A. A Screening Models of (In Vivo and In Vitro) Used for the Study of Hepatoprotective Agents. J. Adv. Zool. 2023, 44, 173–187. [Google Scholar] [CrossRef]
  7. Bártíková, H.; Podlipná, R.; Skálová, L. Veterinary Drugs in the Environment and Their Toxicity to Plants. Chemosphere 2016, 144, 2290–2301. [Google Scholar] [CrossRef] [PubMed]
  8. Küster, A.; Adler, N. Pharmaceuticals in the Environment: Scientific Evidence of Risks and Its Regulation. Philos. Trans. R. Soc. B 2014, 369, 20130587. [Google Scholar] [CrossRef] [PubMed]
  9. Litynska, M.; Kyrii, S.; Nosovska, O.; Ryzhenko, N. Problem of Antibiotics in Natural Water: A Review. Water Water Purif. Technol. Sci. Tech. News 2021, 31, 26–34. [Google Scholar] [CrossRef]
  10. Rai, B.; Shrivastav, A. Removal of Emerging Contaminants in Water Treatment by Nanofiltration and Reverse Osmosis. In Development in Wastewater Treatment Research and Processes; Elsevier: Amsterdam, The Netherlands, 2022; pp. 605–628. ISBN 978-0-323-85583-9. [Google Scholar]
  11. Abada, B.; Alivio, T.E.G.; Shao, Y.; O’Loughlin, T.E.; Klemashevich, C.; Banerjee, S.; Jayaraman, A.; Chu, K.-H. Photodegradation of Fluorotelomer Carboxylic 5:3 Acid and Perfluorooctanoic Acid Using Zinc Oxide. Environ. Pollut. 2018, 243, 637–644. [Google Scholar] [CrossRef]
  12. Taheran, M.; Brar, S.K.; Verma, M.; Surampalli, R.Y.; Zhang, T.C.; Valero, J.R. Membrane Processes for Removal of Pharmaceutically Active Compounds (PhACs) from Water and Wastewaters. Sci. Total Environ. 2016, 547, 60–77. [Google Scholar] [CrossRef]
  13. Eniola, J.O.; Kumar, R.; Barakat, M.A.; Rashid, J. A Review on Conventional and Advanced Hybrid Technologies for Pharmaceutical Wastewater Treatment. J. Clean. Prod. 2022, 356, 131826. [Google Scholar] [CrossRef]
  14. Cai, Z.; Dwivedi, A.D.; Lee, W.-N.; Zhao, X.; Liu, W.; Sillanpää, M.; Zhao, D.; Huang, C.-H.; Fu, J. Application of Nanotechnologies for Removing Pharmaceutically Active Compounds from Water: Development and Future Trends. Environ. Sci. Nano 2018, 5, 27–47. [Google Scholar] [CrossRef]
  15. Liu, H.; Wang, C.; Wang, G. Photocatalytic Advanced Oxidation Processes for Water Treatment: Recent Advances and Perspective. Chem. Asian J. 2020, 15, 3239–3253. [Google Scholar] [CrossRef] [PubMed]
  16. Armaković, S.J.; Savanović, M.M.; Armaković, S. Titanium Dioxide as the Most Used Photocatalyst for Water Purification: An Overview. Catalysts 2022, 13, 26. [Google Scholar] [CrossRef]
  17. Borysiewicz, M.A. ZnO as a Functional Material, a Review. Crystals 2019, 9, 505. [Google Scholar] [CrossRef]
  18. Melinte, V.; Stroea, L.; Chibac-Scutaru, A.L. Polymer Nanocomposites for Photocatalytic Applications. Catalysts 2019, 9, 986. [Google Scholar] [CrossRef]
  19. Dai, C.; Liu, B. Conjugated Polymers for Visible-Light-Driven Photocatalysis. Energy Environ. Sci. 2020, 13, 24–52. [Google Scholar] [CrossRef]
  20. Mangaraj, S.; Yadav, A.; Bal, L.M.; Dash, S.K.; Mahanti, N.K. Application of Biodegradable Polymers in Food Packaging Industry: A Comprehensive Review. J. Packag. Technol. Res. 2019, 3, 77–96. [Google Scholar] [CrossRef]
  21. George, A.; Sanjay, M.R.; Srisuk, R.; Parameswaranpillai, J.; Siengchin, S. A Comprehensive Review on Chemical Properties and Applications of Biopolymers and Their Composites. Int. J. Biol. Macromol. 2020, 154, 329–338. [Google Scholar] [CrossRef]
  22. Ojajuni, O.; Saroj, D.; Cavalli, G. Removal of Organic Micropollutants Using Membrane-Assisted Processes: A Review of Recent Progress. Environ. Technol. Rev. 2015, 4, 17–37. [Google Scholar] [CrossRef]
  23. Haider, R. The Future of Pharmaceuticals Industry 2024. J. Pharm. Pharmacol. Res. 2023, 6, 1–11. [Google Scholar] [CrossRef]
  24. Yu, W.; Xiong, L.; Teng, J.; Chen, C.; Li, B.; Zhao, L.; Lin, H.; Shen, L. Advances in Synthesis and Application of Amphoteric Polymer-Based Water Treatment Agents. Desalination 2024, 574, 119456119456. [Google Scholar] [CrossRef]
  25. Jin, J.; Li, C.; He, Y.; Pan, J.; Zhu, J.; Tang, J. Real World Drug Treatment Models for Pregnancy Complicated with Urinary Tract Infection in China from 2018 to 2022: A Cross-Section Analysis. Front. Pharmacol. 2024, 15, 1349121. [Google Scholar] [CrossRef] [PubMed]
  26. Gworek, B.; Kijeńska, M.; Wrzosek, J.; Graniewska, M. Pharmaceuticals in the Soil and Plant Environment: A Review. Water Air Soil Pollut. 2021, 232, 145. [Google Scholar] [CrossRef]
  27. Gorgaslidze, N.; Sulashvili, N.; Gabunia, L.; Pruidze-Liparteliani, N.; Giorgobiani, M. The Impact of Pharmaceuticals on the Ecology and Human Health. World Med. J. Biomed. Sci. 2024, 1, 65–89. [Google Scholar]
  28. Netshithothole, R.; Managa, M.; Botha, T.L.; Madikizel, L.M. Occurrence of Selected Pharmaceuticals in Wastewater and Sludge Samples from Wastewater Treatment Plants in Eastern Cape Province of South Africa. S. Afr. J. Chem. 2024, 78, 7–14. [Google Scholar] [CrossRef]
  29. Kock, A.; Glanville, H.; Law, A.; Stanton, T.; Carter, L.; Taylor, J. Emerging Challenges of the Impacts of Pharmaceuticals on Aquatic Ecosystems: A Diatom Perspective. Sci. Total Environ. 2023, 878, 162939. [Google Scholar] [CrossRef] [PubMed]
  30. Salimi, M.; Esrafili, A.; Gholami, M.; Jonidi Jafari, A.; Rezaei Kalantary, R.; Farzadkia, M.; Kermani, M.; Sobhi, H.R. Contaminants of Emerging Concern: A Review of New Approach in AOP Technologies. Environ. Monit. Assess. 2017, 189, 414. [Google Scholar] [CrossRef] [PubMed]
  31. Garrido-Cardenas, J.A.; Esteban-García, B.; Agüera, A.; Sánchez-Pérez, J.A.; Manzano-Agugliaro, F. Wastewater Treatment by Advanced Oxidation Process and Their Worldwide Research Trends. Int. J. Environ. Res. Public Health 2020, 17, 170. [Google Scholar] [CrossRef]
  32. Kulišťáková, A. Removal of Pharmaceutical Micropollutants from Real Wastewater Matrices by Means of Photochemical Advanced Oxidation Processes–a Review. J. Water Process Eng. 2023, 53, 103727. [Google Scholar] [CrossRef]
  33. Roslan, N.N.; Lau, H.L.H.; Suhaimi, N.A.A.; Shahri, N.N.M.; Verinda, S.B.; Nur, M.; Lim, J.-W.; Usman, A. Recent Advances in Advanced Oxidation Processes for Degrading Pharmaceuticals in Wastewater—A Review. Catalysts 2024, 14, 189. [Google Scholar] [CrossRef]
  34. Bean, T.G.; Chadwick, E.A.; Herrero-Villar, M.; Mateo, R.; Naidoo, V.; Rattner, B.A. Do Pharmaceuticals in the Environment Pose a Risk to Wildlife? Environ. Toxicol. Chem. 2024, 43, 595–610. [Google Scholar] [CrossRef] [PubMed]
  35. Park, J.; Kim, C.; Hong, Y.; Lee, W.; Chung, H.; Jeong, D.-H.; Kim, H. Distribution and Removal of Pharmaceuticals in Liquid and Solid Phases in the Unit Processes of Sewage Treatment Plants. Int. J. Environ. Res. Public Health 2020, 17, 687. [Google Scholar] [CrossRef] [PubMed]
  36. Li, Y.; Li, C.; Wang, Z.; Liu, Y.; Jia, Y.; Li, F.; Ren, R.; Ikhlaq, A.; Kumirska, J.; Siedlecka, E.M. Navigating the Complexity of Pharmaceutical Wastewater Treatment by “Effective Strategy, Emerging Technology, and Sustainable Solution”. J. Water Process Eng. 2024, 63, 105404. [Google Scholar] [CrossRef]
  37. Fdez-Sanromán, A.; Torres-Pinto, A.; Rosales, E.; Silva, C.G.; Faria, J.L.; Pazos, M.; Silva, A.M. Optimisation of a Photoelectrochemical System for the Removal of Pharmaceuticals in Water Using Graphitic Carbon Nitride. Catal. Today 2024, 432, 114578. [Google Scholar] [CrossRef]
  38. Coimbra, R.N.; Escapa, C.; Otero, M. Removal of Pharmaceuticals from Water: Conventional and Alternative Treatments. Water 2021, 13, 487. [Google Scholar] [CrossRef]
  39. Leyva-Díaz, J.C.; Batlles-delaFuente, A.; Molina-Moreno, V.; Sánchez Molina, J.; Belmonte-Ureña, L.J. Removal of Pharmaceuticals from Wastewater: Analysis of the Past and Present Global Research Activities. Water 2021, 13, 2353. [Google Scholar] [CrossRef]
  40. Nguyen, M.-K.; Lin, C.; Bui, X.-T.; Rakib, M.R.J.; Nguyen, H.-L.; Truong, Q.-M.; Hoang, H.-G.; Tran, H.-T.; Malafaia, G.; Idris, A.M. Occurrence and Fate of Pharmaceutical Pollutants in Wastewater: Insights on Ecotoxicity, Health Risk, and State–of–the-Art Removal. Chemosphere 2024, 354, 141678. [Google Scholar] [CrossRef]
  41. Gadore, V.; Mishra, S.R.; Ahmaruzzaman, M. Bandgap Engineering Approach for Synthesising Photoactive Novel Ag/HAp/SnS2 for Removing Toxic Anti-Fungal Pharmaceutical from Aqueous Environment. J. Hazard. Mater. 2024, 461, 132458. [Google Scholar] [CrossRef]
  42. Yu, T.; Chen, H.; Hu, T.; Feng, J.; Xing, W.; Tang, L.; Tang, W. Recent Advances in the Applications of Encapsulated Transition-Metal Nanoparticles in Advanced Oxidation Processes for Degradation of Organic Pollutants: A Critical Review. Appl. Catal. B Environ. 2024, 342, 123401. [Google Scholar] [CrossRef]
  43. Zhang, Y.; Jia, G.; Wang, W.; Jiang, L.; Guo, Z. Photocatalytic Upgrading of 5-Hydroxymethylfurfural–Aerobic or Anaerobic? Green Chem. 2024, 26, 2949–2966. [Google Scholar] [CrossRef]
  44. Bilić, A.; Armaković, S.J.; Savanović, M.M.; Zahović, I.; Dodić, J.; Trivunović, Z.; Savić, I.; Gajo, T.; Armaković, S. Photocatalytic Application of Bacterial-Derived Biopolymer in Removing Pharmaceutical Contaminants from Water. Catal. Commun. 2024, 186, 106821. [Google Scholar] [CrossRef]
  45. Saikia, C.; Gogoi, P. Chitosan: A Promising Biopolymer in Drug Delivery Applications. J. Mol. Genet. Med. S 2015, 4, 899–910. [Google Scholar] [CrossRef]
  46. Muzammal, S.; Ahmad, A.; Sheraz, M.; Kim, J.; Ali, S.; Hanif, M.B.; Hussain, I.; Pandiaraj, S.; Alodhayb, A.; Javed, M.S. Polymer-Supported Nanomaterials for Photodegradation: Unraveling the Methylene Blue Menace. Energy Convers. Manag. X 2024, 22, 100547. [Google Scholar] [CrossRef]
  47. De Oliveira, C.; Renda, C.G.; Moreira, A.J.; Pereira, O.A.P.; Pereira, E.C.; Freschi, G.P.G.; Bertholdo, R. Evaluation of a Graphitic Porous Carbon Modified with Iron Oxides for Atrazine Environmental Remediation in Water by Adsorption. Environ. Res. 2023, 219, 115054. [Google Scholar] [CrossRef] [PubMed]
  48. Renda, C.G.; Contreras Medrano, C.P.; Costa, L.J.D.; Litterst, F.J.; Saitovitch, E.M.B.; Magon, C.J.; Gualdi, A.J.; Venâncio, T.; Bertholdo, R.; Moreira, A.J.; et al. Role of Ferrocene-Derived Iron Species in the Catalytic Graphitization of Novolak Resins. J. Mater. Sci. 2021, 56, 1298–1311. [Google Scholar] [CrossRef]
  49. Moreira, A.J.; Dos Santos, B.R.M.; Dias, J.A.; Rabello, P.T.; Coelho, D.; Mascaro, L.H.; Freschi, G.P.G.; Gobato, Y.G.; Galeti, H.V.A.; Mastelaro, V.R.; et al. Photoactivity of Boron-or Nitrogen-Modified TiO2 for Organic Pollutants Degradation: Unveiling the Photocatalytic Mechanisms and by-Products. J. Environ. Chem. Eng. 2023, 11, 109207. [Google Scholar] [CrossRef]
  50. Wendott, J.K.; Were, M.L.L.; Cherutoi, J.; Odero, M.P. Occurrence and Removal Efficiencies of Four Antibiotics in Kisii and Kabarnet Waste Water Treatment Plants, Kenya. Asian J. Appl. Chem. Res. 2024, 15, 1–6. [Google Scholar] [CrossRef]
  51. Hricová, K.; Röderová, M.; Fryčák, P.; Pauk, V.; Kurka, O.; Mezerová, K.; Štosová, T.; Bardoň, J.; Milde, D.; Kučová, P. Prevalence of Vancomycin-Resistant Enterococci and Antimicrobial Residues in Wastewater and Surface Water. Life 2021, 11, 1403. [Google Scholar] [CrossRef]
  52. Ren, B.; Geng, J.; Wang, Y.; Wang, P. Emission and Ecological Risk of Pharmaceuticals and Personal Care Products Affected by Tourism in Sanya City, China. Environ. Geochem. Health 2021, 43, 3083–3097. [Google Scholar] [CrossRef]
  53. Molnarova, L.; Halesova, T.; Tomesova, D.; Vaclavikova, M.; Bosakova, Z. Monitoring Pharmaceuticals and Personal Care Products in Healthcare Effluent Wastewater Samples and the Effectiveness of Drug Removal in Wastewater Treatment Plants Using the UHPLC-MS/MS Method. Molecules 2024, 29, 1480. [Google Scholar] [CrossRef]
  54. El-Yazbi, A.; Korany, M.A.; Nabil, S. Simple Rapid Eco-Friendly Chromatographic Analytical Method for Ten Residual Cephalosporins in Wastewater after Their Pharmaceutical Production. J. Adv. Pharm. Sci. 2024, 1, 48–57. [Google Scholar] [CrossRef]
  55. Baranauskaite-Fedorova, I.; Dvarioniene, J. Management of Macrolide Antibiotics (Erythromycin, Clarithromycin and Azithromycin) in the Environment: A Case Study of Environmental Pollution in Lithuania. Water 2022, 15, 10. [Google Scholar] [CrossRef]
  56. Rafi, J.; Daniel, M.; Neppolian, B. Ultrasensitive Detection of Chloramphenicol in Water Using Functionalized Polymers with an Aluminium Organic Framework. Chemosphere 2024, 357, 141981. [Google Scholar] [CrossRef]
  57. Kairigo, P.; Ngumba, E.; Sundberg, L.-R.; Gachanja, A.; Tuhkanen, T. Occurrence of Antibiotics and Risk of Antibiotic Resistance Evolution in Selected Kenyan Wastewaters, Surface Waters and Sediments. Sci. Total Environ. 2020, 720, 137580. [Google Scholar] [CrossRef]
  58. Mtetwa, H.N.; Amoah, I.D.; Kumari, S.; Bux, F.; Reddy, P. Optimization of Analytical Methods for Tuberculosis Drug Detection in Wastewater: A Multinational Study. Heliyon 2024, 10, e30720. [Google Scholar] [CrossRef]
  59. Hossain, M.N.; Hossain, Z.Z.; Roy, A.K.; Akhter, H.; Begum, A. Detection of Oxytetracycline and Chloramphenicol in Untreated Wastewater Effluents by Reverse Phase-High Performance Liquid Chromatography (RP-HPLC) Technique. Afr. J. Pharm. Pharmacol. 2022, 16, 11–18. [Google Scholar]
  60. Lanaro, V.M.; Sombra, L.L.; Altamirano, J.C.; Almeida, C.A.; Stege, P.W. Chiral Separation of Propranolol by Electrokinetic Chromatography Using Nanodiamonds and Human Serum Albumin as a Pseudo-stationary Phase in River Water. Chirality 2024, 36, e23640. [Google Scholar] [CrossRef]
  61. Ng, K.; Alygizakis, N.A.; Thomaidis, N.S.; Slobodnik, J. Wide-Scope Target and Suspect Screening of Antibiotics in Effluent Wastewater from Wastewater Treatment Plants in Europe. Antibiotics 2023, 12, 100. [Google Scholar] [CrossRef] [PubMed]
  62. Ji, L.; Wei, X.; Zhu, H.; Wang, X.; Zhou, G.; Huang, M.; He, Y. Feasibility of Ultrafiltration-Based Processes for Simultaneous Removal of Trace-Level Ofloxacin and Sulfamethazine from the Effluent of WWTPs. J. Water Process Eng. 2024, 60, 105118. [Google Scholar] [CrossRef]
  63. Vargas-Muñoz, M.; Boudenne, J.-L.; Coulomb, B.; Robert-Peillard, F.; Palacio, E. Automated Method for the Solid Phase Extraction of Tetracyclines in Wastewater Followed by Fluorimetric Determination. Talanta 2024, 270, 125544. [Google Scholar] [CrossRef]
  64. Habibi-Yangjeh, A.; Pournemati, K. A Review on Emerging Homojunction Photocatalysts with Impressive Performances for Wastewater Detoxification. Crit. Rev. Environ. Sci. Technol. 2024, 54, 290–320. [Google Scholar] [CrossRef]
  65. Zhen, C.; Chen, X.; Chen, R.; Fan, F.; Xu, X.; Kang, Y.; Guo, J.; Wang, L.; Lu, G.Q.; Domen, K.; et al. Liquid Metal-Embraced Photoactive Films for Artificial Photosynthesis. Nat. Commun. 2024, 15, 1672. [Google Scholar] [CrossRef] [PubMed]
  66. Morshedy, A.S.; El-Fawal, E.M.; Zaki, T.; El-Zahhar, A.A.; Alghamdi, M.M.; El Naggar, A.M.A. A Review on Heterogeneous Photocatalytic Materials: Mechanism, Perspectives, and Environmental and Energy Sustainability Applications. Inorg. Chem. Commun. 2024, 163, 112307. [Google Scholar] [CrossRef]
  67. Yang, J.; Zeng, X.; Tebyetekerwa, M.; Wang, Z.; Bie, C.; Sun, X.; Marriam, I.; Zhang, X. Engineering 2D Photocatalysts for Solar Hydrogen Peroxide Production. Adv. Energy Mater. 2024, 14, 2400740. [Google Scholar] [CrossRef]
  68. Khoo, V.; Ng, S.; Haw, C.; Ong, W. Additive Manufacturing: A Paradigm Shift in Revolutionizing Catalysis with 3D Printed Photocatalysts and Electrocatalysts Toward Environmental Sustainability. Small 2024, 2401278. [Google Scholar] [CrossRef] [PubMed]
  69. Elsayed, M.H.; Abdellah, M.; Alhakemy, A.Z.; Mekhemer, I.M.A.; Aboubakr, A.E.A.; Chen, B.-H.; Sabbah, A.; Lin, K.-H.; Chiu, W.-S.; Lin, S.-J.; et al. Overcoming Small-Bandgap Charge Recombination in Visible and NIR-Light-Driven Hydrogen Evolution by Engineering the Polymer Photocatalyst Structure. Nat. Commun. 2024, 15, 707. [Google Scholar] [CrossRef]
  70. Ma, L.; Gao, Y.; Wei, B.; Huang, L.; Zhang, N.; Weng, Q.; Zhang, L.; Liu, S.F.; Jiang, R. Visible-Light Photocatalytic H2O2 Production Boosted by Frustrated Lewis Pairs in Defected Polymeric Carbon Nitride Nanosheets. ACS Catal. 2024, 14, 2775–2786. [Google Scholar] [CrossRef]
  71. Colmenares, J.C.; Xu, Y.-J. Heterogeneous Photocatalysis. In Green Chemistry and Sustainable Technology; Springer: Berlin/Heidelberg, Germany, 2016. [Google Scholar]
  72. Helali, S.; Polo-López, M.I.; Fernández-Ibáñez, P.; Ohtani, B.; Amano, F.; Malato, S.; Guillard, C. Solar Photocatalysis: A Green Technology for E. coli Contaminated Water Disinfection. Effect of Concentration and Different Types of Suspended Catalyst. J. Photochem. Photobiol. A Chem. 2014, 276, 31–40. [Google Scholar] [CrossRef]
  73. Sharma, S.U.; Elsayed, M.H.; Mekhemer, I.M.; Meng, T.S.; Chou, H.-H.; Kuo, S.-W.; Mohamed, M.G. Rational Design of Pyrene and Thienyltriazine-Based Conjugated Microporous Polymers for High-Performance Energy Storage and Visible-Light Photocatalytic Hydrogen Evolution from Water. Giant 2024, 17, 100217. [Google Scholar] [CrossRef]
  74. Parkatzidis, K.; Wang, H.S.; Anastasaki, A. Photocatalytic Upcycling and Depolymerization of Vinyl Polymers. Angew. Chem. 2024, 136, e202402436. [Google Scholar] [CrossRef]
  75. Dey, S.; Manna, K.; Pradhan, P.; Sarkar, A.N.; Roy, A.; Pal, S. Review of Polymeric Nanocomposites for Photocatalytic Wastewater Treatment. ACS Appl. Nano Mater. 2024, 7, 4588–4614. [Google Scholar] [CrossRef]
  76. Xu, R.; Lai, S.; Zhang, Y.; Zhang, X. Research Progress of Heavy-Metal-Free Quantum Dot Light-Emitting Diodes. Nanomaterials 2024, 14, 832. [Google Scholar] [CrossRef]
  77. Han, C.; Kundu, B.K.; Liang, Y.; Sun, Y. Near-Infrared Light-Driven Photocatalysis with an Emphasis on Two-Photon Excitation: Concepts, Materials, and Applications. Adv. Mater. 2024, 36, 2307759. [Google Scholar] [CrossRef] [PubMed]
  78. Tomić, J.; Malinović, N. Titanium Dioxide Photocatalyst: Present Situation and Future Approaches. AIDASCO Rev. 2023, 1, 26–30. [Google Scholar] [CrossRef]
  79. He, J.; Tallman, D.E.; Bierwagen, G.P. Conjugated Polymers for Corrosion Control: Scanning Vibrating Electrode Studies of Polypyrrole-Aluminum Alloy Interactions. J. Electrochem. Soc. 2004, 151, B644. [Google Scholar] [CrossRef]
  80. Yan, M.; Tallman, D.; Rasmussen, S.; Bierwagen, G. Corrosion Control Coatings for Aluminum Alloys Based on Neutral and N-Doped Conjugated Polymers. J. Electrochem. Soc. 2009, 156, C360. [Google Scholar] [CrossRef]
  81. Rohan Khizer, M.; Saddique, Z.; Imran, M.; Javaid, A.; Latif, S.; Mantzavinos, D.; Momotko, M.; Boczkaj, G. Polymer and Graphitic Carbon Nitride Based Nanohybrids for the Photocatalytic Degradation of Pharmaceuticals in Wastewater Treatment—A Review. Sep. Purif. Technol. 2024, 350, 127768. [Google Scholar] [CrossRef]
  82. Pavliuk, M.V.; Wrede, S.; Liu, A.; Brnovic, A.; Wang, S.; Axelsson, M.; Tian, H. Preparation, Characterization, Evaluation and Mechanistic Study of Organic Polymer Nano-Photocatalysts for Solar Fuel Production. Chem. Soc. Rev. 2022, 51, 6909–6935. [Google Scholar] [CrossRef]
  83. Nanda, K.K.; Swain, S.; Satpati, B.; Besra, L.; Mishra, B.; Chaudhary, Y.S. Enhanced Photocatalytic Activity and Charge Carrier Dynamics of Hetero-Structured Organic–Inorganic Nano-Photocatalysts. ACS Appl. Mater. Interfaces 2015, 7, 7970–7978. [Google Scholar] [CrossRef]
  84. Goutham, R.; Badri Narayan, R.; Srikanth, B.; Gopinath, K.P. Supporting Materials for Immobilisation of Nano-Photocatalysts. In Nanophotocatalysis and Environmental Applications; Inamuddin, Sharma, G., Kumar, A., Lichtfouse, E., Asiri, A.M., Eds.; Environmental Chemistry for a Sustainable World; Springer International Publishing: Cham, Switzerland, 2019; Volume 29, pp. 49–82. ISBN 978-3-030-10608-9. [Google Scholar]
  85. Sadowski, R.; Wach, A.; Buchalska, M.; Kuśtrowski, P.; Macyk, W. Photosensitized TiO2 Films on Polymers—Titania-Polymer Interactions and Visible Light Induced Photoactivity. Appl. Surf. Sci. 2019, 475, 710–719. [Google Scholar] [CrossRef]
  86. Wang, Z.; Bockstaller, M.R.; Matyjaszewski, K. Synthesis and Applications of ZnO/Polymer Nanohybrids. ACS Mater. Lett. 2021, 3, 599–621. [Google Scholar] [CrossRef]
  87. Ponnamma, D.; Cabibihan, J.-J.; Rajan, M.; Pethaiah, S.S.; Deshmukh, K.; Gogoi, J.P.; Pasha, S.K.K.; Ahamed, M.B.; Krishnegowda, J.; Chandrashekar, B.N.; et al. Synthesis, Optimization and Applications of ZnO/Polymer Nanocomposites. Mater. Sci. Eng. C 2019, 98, 1210–1240. [Google Scholar] [CrossRef] [PubMed]
  88. Ma, Y.; Yang, Y.; Li, T.; Hussain, S.; Zhu, M. Deep Eutectic Solvents as an Emerging Green Platform for the Synthesis of Functional Materials. Green Chem. 2024, 26, 3627–3669. [Google Scholar] [CrossRef]
  89. Scharf, S.; Notz, S.; Lang, H. Twin Polymerization: A Unique Strategy towards Versatile Functional Materials. Eur. J. Inorg. Chem. 2024, 27, e202300406. [Google Scholar] [CrossRef]
  90. Affrald R, J.; Narayan, S. Intranasal Drug Delivery of Antiviral Agents—A Revisit and WayForward. Curr. Drug Ther. 2024, 19, 130–150. [Google Scholar] [CrossRef]
  91. Karbalaei Akbari, M.; Siraj Lopa, N.; Park, J.; Zhuiykov, S. Plasmonic Nanodomains Decorated on Two-Dimensional Oxide Semiconductors for Photonic-Assisted CO2 Conversion. Materials 2023, 16, 3675. [Google Scholar] [CrossRef]
  92. Sakhadeo, N.N.; Patro, T.U. Exploring the Multifunctional Applications of Surface-Coated Polymeric Foams─A Review. Ind. Eng. Chem. Res. 2022, 61, 5366–5387. [Google Scholar] [CrossRef]
  93. Li, Z.; Zhang, W.; Tao, M.; Shen, L.; Li, R.; Zhang, M.; Jiao, Y.; Hong, H.; Xu, Y.; Lin, H. In-Situ Growth of UiO-66-NH2 in Porous Polymeric Substrates at Room Temperature for Fabrication of Mixed Matrix Membranes with Fast Molecular Separation Performance. Chem. Eng. J. 2022, 435, 134804. [Google Scholar] [CrossRef]
  94. Wu, M.; Zhang, M.; Shen, L.; Wang, X.; Ying, D.; Lin, H.; Li, R.; Xu, Y.; Hong, H. High Propensity of Membrane Fouling and the Underlying Mechanisms in a Membrane Bioreactor during Occurrence of Sludge Bulking. Water Res. 2023, 229, 119456. [Google Scholar] [CrossRef]
  95. Huang, Z.; Liu, J.; Liu, Y.; Xu, Y.; Li, R.; Hong, H.; Shen, L.; Lin, H.; Liao, B.-Q. Enhanced Permeability and Antifouling Performance of Polyether Sulfone (PES) Membrane via Elevating Magnetic Ni@MXene Nanoparticles to Upper Layer in Phase Inversion Process. J. Membr. Sci. 2021, 623, 119080. [Google Scholar] [CrossRef]
  96. Ghali, M.; Brahmi, C.; Benltifa, M.; Vaulot, C.; Airoudj, A.; Fioux, P.; Dumur, F.; Simonnet-Jégat, C.; Morlet-Savary, F.; Jellali, S.; et al. Characterization of Polyoxometalate/Polymer Photo-composites: A Toolbox for the Photodegradation of Organic Pollutants. J. Polym. Sci. 2021, 59, 153–169. [Google Scholar] [CrossRef]
  97. Ulbricht, M. Advanced Functional Polymer Membranes. Polymer 2006, 47, 2217–2262. [Google Scholar] [CrossRef]
  98. Srivastava, V.; Srivastava, R. On the Polymeric Foams: Modeling and Properties. J. Mater. Sci. 2014, 49, 2681–2692. [Google Scholar] [CrossRef]
  99. Shirtcliffe, N.J.; McHale, G.; Newton, M.I. The Superhydrophobicity of Polymer Surfaces: Recent Developments. J. Polym. Sci. Part B Polym. Phys. 2011, 49, 1203–1217. [Google Scholar] [CrossRef]
  100. Liu, H.-Z.; Shu, X.-X.; Huang, M.; Wu, B.-B.; Chen, J.-J.; Wang, X.-S.; Li, H.-L.; Yu, H.-Q. Tailoring D-Band Center of High-Valent Metal-Oxo Species for Pollutant Removal via Complete Polymerization. Nat. Commun. 2024, 15, 2327. [Google Scholar] [CrossRef]
  101. Zhang, P.; Yang, Y.; Duan, X.; Wang, S. Oxidative Polymerization versus Degradation of Organic Pollutants in Heterogeneous Catalytic Persulfate Chemistry. Water Res. 2024, 255, 121485. [Google Scholar] [CrossRef]
  102. Lee, D.-E.; Danish, M.; Alam, U.; Jo, W.-K. Review on Inorganic and Polymeric Materials-Coordinated Metal-Organic-Framework Photocatalysts for Green Hydrogen Evolution. J. Energy Chem. 2024, 92, 322–356. [Google Scholar] [CrossRef]
  103. Guo, X.; Sun, Y.; Sun, X.; Li, J.; Wu, J.; Shi, Y.; Pan, L. Doping Engineering of Conductive Polymers and Their Application in Physical Sensors for Healthcare Monitoring. Macromol. Rapid Commun. 2024, 45, 2300246. [Google Scholar] [CrossRef]
  104. Thien, G.S.H.; Chan, K.-Y.; Marlinda, A.R.; Yap, B.K. Polymer-Enhanced Perovskite Oxide-Based Photocatalysts: A Review. Nanoscale 2023, 15, 19039–19061. [Google Scholar] [CrossRef]
  105. Kumar, R.; Travas-Sejdic, J.; Padhye, L.P. Conducting Polymers-Based Photocatalysis for Treatment of Organic Contaminants in Water. Chem. Eng. J. Adv. 2020, 4, 100047. [Google Scholar] [CrossRef]
  106. Li, H.-Y.; Li, Z.-S.; Qiu, G.-H.; Zhang, R.-R.; Wang, Y.-R.; Wang, F.; Huang, R.-W.; Liu, X.-F.; Zang, S.-Q. Viologen-Based Ionic Conjugated Mesoporous Polymer as the Electron Conveyer for Efficient Polysulfide Trapping and Conversion. Sci. Bull. 2024, 69, 1071–1080. [Google Scholar] [CrossRef] [PubMed]
  107. Mansha, M.; Ahmad, T.; Ullah, N.; Akram Khan, S.; Ashraf, M.; Ali, S.; Tan, B.; Khan, I. Photocatalytic Water-Splitting by Organic Conjugated Polymers: Opportunities and Challenges. Chem. Rec. 2022, 22, e202100336. [Google Scholar] [CrossRef] [PubMed]
  108. Enesca, A.; Cazan, C. Polymer Composite-Based Materials with Photocatalytic Applications in Wastewater Organic Pollutant Removal: A Mini Review. Polymers 2022, 14, 3291. [Google Scholar] [CrossRef] [PubMed]
  109. Ajay, A.; Ramaswamy, K.P.; Thomas, A.V. A Critical Review on the Durability of Geopolymer Composites in Acidic Environment. IOP Conf. Ser. Earth Environ. Sci. 2020, 491, 012044. [Google Scholar] [CrossRef]
  110. Daems, K.; Yadav, P.; Dermenci, K.B.; Van Mierlo, J.; Berecibar, M. Advances in Inorganic, Polymer and Composite Electrolytes: Mechanisms of Lithium-Ion Transport and Pathways to Enhanced Performance. Renew. Sustain. Energy Rev. 2024, 191, 114136. [Google Scholar] [CrossRef]
  111. Xue, M.; Zhang, L.; Wang, X.; Dong, Q.; Zhu, Z.; Wang, X.; Gu, Q.; Kang, F.; Li, X.; Zhang, Q. A Metal-Free Helical Covalent Inorganic Polymer: Preparation, Crystal Structure and Optical Properties. Angew. Chem. Int. Ed. 2024, 63, e202315338. [Google Scholar] [CrossRef] [PubMed]
  112. Intisar, A.; Ramzan, A.; Hafeez, S.; Hussain, N.; Irfan, M.; Shakeel, N.; Gill, K.A.; Iqbal, A.; Janczarek, M.; Jesionowski, T. Adsorptive and Photocatalytic Degradation Potential of Porous Polymeric Materials for Removal of Pesticides, Pharmaceuticals, and Dyes-Based Emerging Contaminants from Water. Chemosphere 2023, 336, 139203. [Google Scholar] [CrossRef]
  113. Caminade, A.-M.; Hey-Hawkins, E.; Manners, I. Smart Inorganic Polymers. Chem. Soc. Rev. 2016, 45, 5144–5146. [Google Scholar] [CrossRef]
  114. Sołoducho, J.; Zając, D.; Spychalska, K.; Baluta, S.; Cabaj, J. Conducting Silicone-Based Polymers and Their Application. Molecules 2021, 26, 2012. [Google Scholar] [CrossRef]
  115. Chauhan, N.P.S.; Hosmane, N.S.; Mozafari, M. Boron-Based Polymers: Opportunities and Challenges. Mater. Today Chem. 2019, 14, 100184. [Google Scholar] [CrossRef]
  116. Ghorai, A.; Banerjee, S. Phosphorus-Containing Aromatic Polymers: Synthesis, Structure, Properties and Membrane-Based Applications. Prog. Polym. Sci. 2023, 138, 101646. [Google Scholar] [CrossRef]
  117. Wang, S.; Li, H.; Huang, H.; Cao, X.; Chen, X.; Cao, D. Porous Organic Polymers as a Platform for Sensing Applications. Chem. Soc. Rev. 2022, 51, 2031–2080. [Google Scholar] [CrossRef] [PubMed]
  118. Kartsonakis, I.A.; Koumoulos, E.P.; Balaskas, A.C.; Pappas, G.S.; Charitidis, C.A.; Kordas, G.C. Hybrid Organic–Inorganic Multilayer Coatings Including Nanocontainers for Corrosion Protection of Metal Alloys. Corros. Sci. 2012, 57, 56–66. [Google Scholar] [CrossRef]
  119. Han, M.G.; Lee, Y.J.; Byun, S.W.; Im, S.S. Physical Properties and Thermal Transition of Polyaniline Film. Synth. Met. 2001, 124, 337–343. [Google Scholar] [CrossRef]
  120. Li, Y.; Zhao, Z.; Zhang, J.; Chen, Z.; Liu, X.; Wang, K.; Sobolev, A.V.; Savilov, S.V.; Chen, M. High Pseudocapacitance Electrode Enabled by Heterovalent Doping and Surface Coating for Rapid Charge Storage. Surf. Coat. Technol. 2024, 476, 130169. [Google Scholar] [CrossRef]
  121. Kazemi, F.; Naghib, S.M.; Zare, Y.; Rhee, K.Y. Biosensing Applications of Polyaniline (PANI)-Based Nanocomposites: A Review. Polym. Rev. 2021, 61, 553–597. [Google Scholar] [CrossRef]
  122. Lee, Y.-J.; Lee, H.S.; Lee, C.-G.; Park, S.-J.; Lee, J.; Jung, S.; Shin, G.-A. Application of PANI/TiO2 Composite for Photocatalytic Degradation of Contaminants from Aqueous Solution. Appl. Sci. 2020, 10, 6710. [Google Scholar] [CrossRef]
  123. Duygun, İ.K.; Altin, Y.; Bedeloglu, A.C. Conducting Polymer Blends, Interpenetrating Polymeric Networks, and Gels Based on Polyvinyl Chloride. In Poly(vinyl chloride)-Based Blends, IPNs, and Gels; Elsevier: Amsterdam, The Netherlands, 2024; pp. 431–455. ISBN 978-0-323-99474-3. [Google Scholar]
  124. Boudjelida, S.; Li, X.; Djellali, S.; Chiappetta, G.; Russo, F.; Figoli, A.; Carraro, M. Synthesis and Characterization of Polyaniline Emeraldine Salt (PANI-ES) Colloids Using Potato Starch as a Stabilizer, to Enhance the Physicochemical Properties and Processibility. Materials 2024, 17, 2941. [Google Scholar] [CrossRef]
  125. Zhu, Y.; Liu, X.; Hu, X.; Wang, T.; Parkin, I.P.; Wang, M.; Boruah, B.D. Polyaniline and Water Pre-Intercalated V2O5 Cathodes for High-Performance Planar Zinc-Ion Micro-Batteries. Chem. Eng. J. 2024, 487, 150384. [Google Scholar] [CrossRef]
  126. Kundu, S.; Majumder, R.; Pal Chowdhury, M. Superior Ammonia Sensing Properties of PET-supported Polyaniline/Reduced Graphene Oxide/Zinc Ferrite Ternary Nanocomposite Thin Film at Room Temperature. J. Appl. Polym. Sci. 2024, 141, e54755. [Google Scholar] [CrossRef]
  127. Bekele Mekonnen, T. Electrochemical Sensor Based on Polyaniline Supported CdS/CeO2/Ag3PO4 Nanocomposite for Malathion Detection. Sens. Int. 2024, 5, 100251. [Google Scholar] [CrossRef]
  128. Ali, M.R.R.; Tigno, S.D.; Honeyman, J.S.; Caldona, E.B. Adhesion-Diffusional-Based Corrosion Protection Mechanisms of Polyaniline-Primed Fluoropolymer Coatings. ACS Appl. Polym. Mater. 2024, 6, 2719–2732. [Google Scholar] [CrossRef]
  129. Atassi, Y.; Tally, M. Electrochemical Polymerization of Anilinium Hydrochloride. arXiv 2013, arXiv:1307.5668. [Google Scholar]
  130. Kalaleh, H.-A.; Tally, M.; Atassi, Y. Preparation of a Clay Based Superabsorbent Polymer Composite of Copolymer Poly(Acrylate-Co-Acrylamide) with Bentonite via Microwave Radiation. arXiv 2013, arXiv:1311.6445. [Google Scholar]
  131. Ullah, R.; Ullah, H.; Shah, A.-H.A.; Bilal, S.; Ali, K. Oligomeric Synthesis and Density Functional Theory of Leucoemeraldine Base Form of Polyaniline. J. Mol. Struct. 2017, 1127, 734–741. [Google Scholar] [CrossRef]
  132. Yoon, S.-B.; Yoon, E.-H.; Kim, K.-B. Electrochemical Properties of Leucoemeraldine, Emeraldine, and Pernigraniline Forms of Polyaniline/Multi-Wall Carbon Nanotube Nanocomposites for Supercapacitor Applications. J. Power Sources 2011, 196, 10791–10797. [Google Scholar] [CrossRef]
  133. Quillard, S.; Louarn, G.; Lefrant, S.; Macdiarmid, A.G. Vibrational Analysis of Polyaniline: A Comparative Study of Leucoemeraldine, Emeraldine, and Pernigraniline Bases. Phys. Rev. B 1994, 50, 12496–12508. [Google Scholar] [CrossRef] [PubMed]
  134. Bobrowska, D.M.; Gdula, K.; Breczko, J.; Basa, A.; Markiewicz, K.H.; Winkler, K. Poly(p-Phenylene Vinylene) Incorporated into Carbon Nanostructures. J. Nanopart. Res. 2022, 24, 222. [Google Scholar] [CrossRef]
  135. Banerjee, J.; Dutta, K. A Short Overview on the Synthesis, Properties and Major Applications of Poly(p-Phenylene Vinylene). Chem. Pap. 2021, 75, 5139–5151. [Google Scholar] [CrossRef]
  136. Blayney, A.J.; Perepichka, I.F.; Wudl, F.; Perepichka, D.F. Advances and Challenges in the Synthesis of Poly(p-phenylene Vinylene)-Based Polymers. Isr. J. Chem. 2014, 54, 674–688. [Google Scholar] [CrossRef]
  137. Basyouni, M.Z.; Abdu, M.E.; Radwan, M.F.; Spring, A.M. From Monomer to Polymer: Controlled Synthesis and Comprehensive Analysis of Poly(p-phenylene vinylene) via ROMP. J. Mol. Struct. 2024, 1310, 138001. [Google Scholar] [CrossRef]
  138. García-Eguizábal, A.; Gomez Mendoza, M.; Barawi, M.; Collado, L.; Sánchez-Ruiz, A.; García Martínez, J.C.; Liras, M.; De La Peña O’Shea, V.A. Nanostructuration Effect of Carbon-Based Phenylene Vinylene Conjugated Porous Polymers on TiO2 Hybrid Materials for Artificial Photosynthesis. Small Struct. 2024, 5, 2300410. [Google Scholar] [CrossRef]
  139. Acelas, M.; Pérez, L.D.; Sierra, C.A. Boosting the Photocatalytic Performance of the Oligo (Phenylene Vinylene) Moiety by Copolymerization for the Heterogeneous Degradation of Indigo Carmine Dye. J. Polym. Eng. 2023, 43, 37–46. [Google Scholar] [CrossRef]
  140. Sun, Z.Y.; Xie, G.; Bai, X.D.; Chen, Y.J. Preparation and Characterisation of Poly(P-phenylene Vinylene)/Semiconductor Nanofibres. Pigment. Resin Technol. 2008, 37, 279–283. [Google Scholar] [CrossRef]
  141. Postacchini, B.B.; Zucolotto, V.; Dias, F.B.; Monkman, A.; Oliveira, O.N. Energy Transfer in Nanostructured Films Containing Poly(p-Phenylene Vinylene) and Acceptor Species. J. Phys. Chem. C 2009, 113, 10303–10306. [Google Scholar] [CrossRef]
  142. Pesco, B.; Rodrigues, R.R.; Siqueira, L.J.A.D.; Péres, L.O. Poly(p-Phenylene Vinylene-b-Ethylene Glycol) Dispersed in Butyltrimethylammonium Bis(Trifluoromethanesulfonyl)Imide as Luminescent Ionic Liquids. Mater. Chem. Phys. 2022, 284, 126021. [Google Scholar] [CrossRef]
  143. Bernhardt, S.; Düring, J.; Haschke, S.; Barr, M.K.S.; Stiegler, L.; Schühle, P.; Bachmann, J.; Hirsch, A.; Gröhn, F. Tunable Photocatalytic Activity of PEO-Stabilized ZnO–Polyoxometalate Nanostructures in Aqueous Solution. Adv. Mater. Interfaces 2021, 8, 2002130. [Google Scholar] [CrossRef]
  144. Liu, Y.; Zong, L.; Zhang, C.; Liu, W.; Fakhri, A.; Gupta, V.K. Design and Structural of Sm-Doped SbFeO3 Nanopowders and Immobilized on Poly(Ethylene Oxide) for Efficient Photocatalysis and Hydrogen Generation under Visible Light Irradiation. Surf. Interfaces 2021, 26, 101292. [Google Scholar] [CrossRef]
  145. Malwal, D.; Gopinath, P. Fabrication and Characterization of Poly(Ethylene Oxide) Templated Nickel Oxide Nanofibers for Dye Degradation. Environ. Sci. Nano 2015, 2, 78–85. [Google Scholar] [CrossRef]
  146. Soitong, T.; Wongsaenmai, S. Photo-Oxidative Degradation Polyethylene Containing Titanium Dioxide and Poly(ethylene oxide). Key Eng. Mater. 2017, 751, 796–800. [Google Scholar] [CrossRef]
  147. Ma, L.; Deng, L.; Chen, J. Applications of Poly(Ethylene Oxide) in Controlled Release Tablet Systems: A Review. Drug Dev. Ind. Pharm. 2014, 40, 845–851. [Google Scholar] [CrossRef] [PubMed]
  148. Duraibabu, D.; Sasikumar, Y. Role of Conducting Polymer Nanostructures in Advanced Photocatalytic Applications. In Green Photocatalysts; Naushad, M., Rajendran, S., Lichtfouse, E., Eds.; Environmental Chemistry for a Sustainable World; Springer International Publishing: Cham, Switzerland, 2020; Volume 34, pp. 189–208. ISBN 978-3-030-15607-7. [Google Scholar]
  149. Uday, S.; Chawla, H.; Chandra, A.; Garg, S. Conducting Polymer Hybrid Nanocomposites-Based Photocatalytic Material for Energy Applications. In Photocatalysis for Environmental Remediation and Energy Production; Garg, S., Chandra, A., Eds.; Green Chemistry and Sustainable Technology; Springer International Publishing: Cham, Switzerland, 2023; pp. 417–438. ISBN 978-3-031-27706-1. [Google Scholar]
  150. Murugan, R. Heterocycles in Polymers. In Progress in Heterocyclic Chemistry; Elsevier: Amsterdam, The Netherlands, 2023; Volume 35, pp. 93–118. ISBN 978-0-443-21936-8. [Google Scholar]
  151. García Rueda, F.C.; González, J.T. Electrochemical Polymerization of Polypyrrole Coatings on Hard-Anodized Coatings of the Aluminum Alloy 2024-T3. Electrochim. Acta 2020, 347, 136272. [Google Scholar] [CrossRef]
  152. Huang, Z.-B.; Yin, G.-F.; Liao, X.-M.; Gu, J.-W. Conducting Polypyrrole in Tissue Engineering Applications. Front. Mater. Sci. 2014, 8, 39–45. [Google Scholar] [CrossRef]
  153. Li, M.; Luo, C.; Zhang, J.; Yang, J.; Xu, J.; Yao, W.; Tan, R.; Duan, X.; Jiang, F. Electrochemical Doping Tuning of Flexible Polypyrrole Film with Enhanced Thermoelectric Performance. Surf. Interfaces 2020, 21, 100759. [Google Scholar] [CrossRef]
  154. Mikalo, R.P.; Hoffmann, P.; Heller, T.; Batchelor, D.R.; Appel, G.; Schmeißer, D. Doping and Defect Inhomogeneities of Polypyrrole Tosylate Films as Revealed by μ-NEXAFS. Mater. Sci. Eng. C 1999, 8–9, 257–265. [Google Scholar] [CrossRef]
  155. Toh, H.W.; Toong, D.W.Y.; Ng, J.C.K.; Ow, V.; Lu, S.; Tan, L.P.; Wong, P.E.H.; Venkatraman, S.; Huang, Y.; Ang, H.Y. Polymer Blends and Polymer Composites for Cardiovascular Implants. Eur. Polym. J. 2021, 146, 110249. [Google Scholar] [CrossRef]
  156. Li, S.; Gao, M.; Zhou, K.; Li, X.; Xian, K.; Zhao, W.; Chen, Y.; He, C.; Ye, L. Achieving Record-High Stretchability and Mechanical Stability in Organic Photovoltaic Blends with a Dilute-absorber Strategy. Adv. Mater. 2023, 36, 2307278. [Google Scholar] [CrossRef] [PubMed]
  157. Wang, J.; Wang, Y.; Xian, K.; Qiao, J.; Chen, Z.; Bi, P.; Zhang, T.; Zheng, Z.; Hao, X.; Ye, L.; et al. Regulating Phase Separation Kinetics for High-Efficiency and Mechanically Robust All-Polymer Solar Cells. Adv. Mater. 2024, 36, 2305424. [Google Scholar] [CrossRef]
  158. Hyvärinen, M.; Jabeen, R.; Kärki, T. The Modelling of Extrusion Processes for Polymers—A Review. Polymers 2020, 12, 1306. [Google Scholar] [CrossRef]
  159. Gaspar-Cunha, A.; Covas, J.A.; Sikora, J. Optimization of Polymer Processing: A Review (Part I—Extrusion). Materials 2022, 15, 384. [Google Scholar] [CrossRef] [PubMed]
  160. Xia, X.; He, C.; Zhang, P. Unified Viscous-to-Inertial Scaling in Liquid Droplet Coalescence. arXiv 2019, arXiv:1906.04970. [Google Scholar]
  161. Wilczyński, K.; Wilczyński, K.J.; Buziak, K. Modeling and Experimental Studies on Polymer Melting and Flow in Injection Molding. Polymers 2022, 14, 2106. [Google Scholar] [CrossRef]
  162. Hilger, D.; Hosters, N. A Data-Driven Reduced Order Modeling Approach Applied in Context of Numerical Analysis and Optimization of Plastic Profile Extrusion. arXiv 2022, arXiv:2209.03121. [Google Scholar]
  163. Carrola, M.; Asadi, A.; Zhang, H.; Papageorgiou, D.G.; Bilotti, E.; Koerner, H. Best of Both Worlds: Synergistically Derived Material Properties via Additive Manufacturing of Nanocomposites. arXiv 2022, arXiv:2202.02378. [Google Scholar] [CrossRef]
  164. Idumah, C.I.; Obele, C.M. Understanding Interfacial Influence on Properties of Polymer Nanocomposites. Surf. Interfaces 2021, 22, 100879. [Google Scholar] [CrossRef]
  165. Ingole, S.B.; Sharma, P.; Verma, R.; Chowdhury, S.; Patil, P.P.; Dwivedi, S.P.; Khan, A.K. Enhancing Mechanical and Thermal Properties of Polymer Matrix Nanocomposites through Tailored Nanomaterial Architectures. E3S Web Conf. 2024, 511, 01016. [Google Scholar] [CrossRef]
  166. Cantarella, M.; Impellizzeri, G.; Di Mauro, A.; Privitera, V.; Carroccio, S.C. Innovative Polymeric Hybrid Nanocomposites for Application in Photocatalysis. Polymers 2021, 13, 1184. [Google Scholar] [CrossRef] [PubMed]
  167. Darwish, M.S.A.; Mostafa, M.H.; Al-Harbi, L.M. Polymeric Nanocomposites for Environmental and Industrial Applications. Int. J. Mol. Sci. 2022, 23, 1023. [Google Scholar] [CrossRef] [PubMed]
  168. Daoudi, W.; Tiwari, A.; Tyagi, M.; Singh, P.; Saxena, A.; Verma, D.K.; Dagdag, O.; Sharma, H.K.; Fuertes, P.O.; El Aatiaoui, A. Carbon Dots Nanoparticles: A Promising Breakthrough in Biosensing, Catalysis, Biomedical and Authers Applications. Nano-Struct. Nano-Objects 2024, 37, 101074. [Google Scholar] [CrossRef]
  169. Genix, A.-C.; Bocharova, V.; Carroll, B.; Dieudonné-George, P.; Chauveau, E.; Sokolov, A.P.; Oberdisse, J. How Tuning Interfaces Impacts the Dynamics and Structure of Polymer Nanocomposites Simultaneously. ACS Appl. Mater. Interfaces 2023, 15, 7496–7510. [Google Scholar] [CrossRef] [PubMed]
  170. Sobiech, M.; Bujak, P.; Luliński, P.; Pron, A. Semiconductor Nanocrystal–Polymer Hybrid Nanomaterials and Their Application in Molecular Imprinting. Nanoscale 2019, 11, 12030–12074. [Google Scholar] [CrossRef] [PubMed]
  171. Diao, Y. Gentler, Nanoscale Ion Implantation. Nat. Nanotechnol. 2024. [Google Scholar] [CrossRef] [PubMed]
  172. Kaushik, N.; Kaushik, N.; Linh, N.; Ghimire, B.; Pengkit, A.; Sornsakdanuphap, J.; Lee, S.-J.; Choi, E. Plasma and Nanomaterials: Fabrication and Biomedical Applications. Nanomaterials 2019, 9, 98. [Google Scholar] [CrossRef] [PubMed]
  173. Bhatt, P.; Kumar, V.; Subramaniyan, V.; Nagarajan, K.; Sekar, M.; Chinni, S.V.; Ramachawolran, G. Plasma Modification Techniques for Natural Polymer-Based Drug Delivery Systems. Pharmaceutics 2023, 15, 2066. [Google Scholar] [CrossRef] [PubMed]
  174. Kaseem, M.; Fatimah, S.; Nashrah, N.; Ko, Y.G. Recent Progress in Surface Modification of Metals Coated by Plasma Electrolytic Oxidation: Principle, Structure, and Performance. Prog. Mater. Sci. 2021, 117, 100735. [Google Scholar] [CrossRef]
  175. Puliyalil, H.; Cvelbar, U. Selective Plasma Etching of Polymeric Substrates for Advanced Applications. Nanomaterials 2016, 6, 108. [Google Scholar] [CrossRef] [PubMed]
  176. Levchenko, I.; Xu, S.; Baranov, O.; Bazaka, O.; Ivanova, E.; Bazaka, K. Plasma and Polymers: Recent Progress and Trends. Molecules 2021, 26, 4091. [Google Scholar] [CrossRef]
  177. Russo, M.; Iervolino, G.; Vaiano, V.; Palma, V. Non-Thermal Plasma Coupled with Catalyst for the Degradation of Water Pollutants: A Review. Catalysts 2020, 10, 1438. [Google Scholar] [CrossRef]
  178. Ramar, P.; Raghavendra, V.; Murugan, P.; Samanta, D. Immobilization of Polymers to Surfaces by Click Reaction for Photocatalysis with Recyclability. Langmuir 2022, 38, 13344–13357. [Google Scholar] [CrossRef]
  179. Borodin, D.; Rahinov, I.; Shirhatti, P.R.; Huang, M.; Kandratsenka, A.; Auerbach, D.J.; Zhong, T.; Guo, H.; Schwarzer, D.; Kitsopoulos, T.N.; et al. Following the Microscopic Pathway to Adsorption through Chemisorption and Physisorption Wells. Science 2020, 369, 1461–1465. [Google Scholar] [CrossRef] [PubMed]
  180. Bakhshandeh, A.; Segala, M.; Escobar Colla, T. Equilibrium Conformations and Surface Charge Regulation of Spherical Polymer Brushes in Stretched Regimes. Macromolecules 2022, 55, 35–48. [Google Scholar] [CrossRef]
  181. Liu, G.; Sun, X.; Li, X.; Wang, Z. The Bioanalytical and Biomedical Applications of Polymer Modified Substrates. Polymers 2022, 14, 826. [Google Scholar] [CrossRef] [PubMed]
  182. Hetemi, D.; Pinson, J. Surface Functionalisation of Polymers. Chem. Soc. Rev. 2017, 46, 5701–5713. [Google Scholar] [CrossRef] [PubMed]
  183. Nie, Z.; Kumacheva, E. Patterning Surfaces with Functional Polymers. Nat. Mater 2008, 7, 277–290. [Google Scholar] [CrossRef] [PubMed]
  184. Booth, J.-P.; Mozetič, M.; Nikiforov, A.; Oehr, C. Foundations of Plasma Surface Functionalization of Polymers for Industrial and Biological Applications. Plasma Sources Sci. Technol. 2022, 31, 103001. [Google Scholar] [CrossRef]
  185. Tillet, G.; Boutevin, B.; Ameduri, B. Chemical Reactions of Polymer Crosslinking and Post-Crosslinking at Room and Medium Temperature. Prog. Polym. Sci. 2011, 36, 191–217. [Google Scholar] [CrossRef]
  186. Lorke, S.; Müller, U.; Meissl, R.; Brüggemann, O. Covalent Cross-Linking of Polymers at Room Temperature. Int. J. Adhes. Adhes. 2019, 91, 150–159. [Google Scholar] [CrossRef]
  187. Huang, S.; Kong, X.; Xiong, Y.; Zhang, X.; Chen, H.; Jiang, W.; Niu, Y.; Xu, W.; Ren, C. An Overview of Dynamic Covalent Bonds in Polymer Material and Their Applications. Eur. Polym. J. 2020, 141, 110094. [Google Scholar] [CrossRef]
  188. Fujiyabu, T.; Sakumichi, N.; Katashima, T.; Liu, C.; Mayumi, K.; Chung, U.; Sakai, T. Tri-Branched Gels: Rubbery Materials with the Lowest Branching Factor Approach the Ideal Elastic Limit. Sci. Adv. 2022, 8, eabk0010. [Google Scholar] [CrossRef]
  189. Yan, C.; Wan, Y.; Long, H.; Luo, H.; Liu, X.; Luo, H.; Chen, S. Improved Capacitive Energy Storage at High Temperature via Constructing Physical Cross-Link and Electron–Hole Pairs Based on P-Type Semiconductive Polymer Filler. Adv. Funct. Mater. 2024, 34, 2312238. [Google Scholar] [CrossRef]
  190. Irzhak, V.I.; Uflyand, I.E.; Dzhardimalieva, G.I. Self-Healing of Polymers and Polymer Composites. Polymers 2022, 14, 5404. [Google Scholar] [CrossRef] [PubMed]
  191. König, B. (Ed.) Chemical Photocatalysis; De Gruyter: Berlin, Germany, 2020; ISBN 978-3-11-057676-4. [Google Scholar]
  192. Tran, V.V.; Nu, T.T.V.; Jung, H.-R.; Chang, M. Advanced Photocatalysts Based on Conducting Polymer/Metal Oxide Composites for Environmental Applications. Polymers 2021, 13, 3031. [Google Scholar] [CrossRef]
  193. Tao, X.; Zhao, Y.; Wang, S.; Li, C.; Li, R. Recent Advances and Perspectives for Solar-Driven Water Splitting Using Particulate Photocatalysts. Chem. Soc. Rev. 2022, 51, 3561–3608. [Google Scholar] [CrossRef] [PubMed]
  194. Bailey, E.J.; Winey, K.I. Dynamics of Polymer Segments, Polymer Chains, and Nanoparticles in Polymer Nanocomposite Melts: A Review. Prog. Polym. Sci. 2020, 105, 101242. [Google Scholar] [CrossRef]
  195. Alipoori, S.; Rouhi, H.; Linn, E.; Stumpfl, H.; Mokarizadeh, H.; Esfahani, M.R.; Koh, A.; Weinman, S.T.; Wujcik, E.K. Polymer-Based Devices and Remediation Strategies for Emerging Contaminants in Water. ACS Appl. Polym. Mater. 2021, 3, 549–577. [Google Scholar] [CrossRef]
  196. Hussain, E.; Ahtesham, A.; Shahadat, M.; Ibrahim, M.N.M.; Ismail, S. Recent Advances of Clay/Polymer-Based Nanomaterials for the Treatment of Environmental Contaminants in Wastewater: A Review. J. Environ. Chem. Eng. 2024, 12, 112401. [Google Scholar] [CrossRef]
  197. Scopus Preview—Scopus—Welcome to Scopus. Available online: https://www.scopus.com/home.uri (accessed on 2 July 2024).
  198. Nosrati, R.; Olad, A.; Maramifar, R. Degradation of Ampicillin Antibiotic in Aqueous Solution by ZnO/Polyaniline Nanocomposite as Photocatalyst under Sunlight Irradiation. Environ. Sci. Pollut. Res. 2012, 19, 2291–2299. [Google Scholar] [CrossRef] [PubMed]
  199. Asgari, E.; Esrafili, A.; Jafari, A.J.; Kalantary, R.R.; Nourmoradi, H.; Farzadkia, M. The Comparison of ZnO/Polyaniline Nanocomposite under UV and Visible Radiations for Decomposition of Metronidazole: Degradation Rate, Mechanism and Mineralization. Process Saf. Environ. Prot. 2019, 128, 65–76. [Google Scholar] [CrossRef]
  200. Šojić Merkulov, D.V.; Despotović, V.N.; Banić, N.D.; Armaković, S.J.; Finčur, N.L.; Lazarević, M.J.; Četojević-Simin, D.D.; Orčić, D.Z.; Radoičić, M.B.; Šaponjić, Z.V.; et al. Photocatalytic Decomposition of Selected Biologically Active Compounds in Environmental Waters Using TiO2/Polyaniline Nanocomposites: Kinetics, Toxicity and Intermediates Assessment. Environ. Pollut. 2018, 239, 457–465. [Google Scholar] [CrossRef]
  201. Fatima, T.; Husain, S.; Khanuja, M. Superior Photocatalytic and Electrochemical Activity of Novel WS2/PANI Nanocomposite for the Degradation and Detection of Pollutants: Antibiotic, Heavy Metal Ions, and Dyes. Chem. Eng. J. Adv. 2022, 12, 100373. [Google Scholar] [CrossRef]
  202. Ujwal, M.P.; Yashas, S.R.; Shivaraju, H.P.; Kumara Swamy, N. Gadolinium Ortho-Ferrite Interfaced Polyaniline: Bi-Functional Catalyst for Electrochemical Detection and Photocatalytic Degradation of Acetaminophen. Surf. Interfaces 2022, 30, 101878. [Google Scholar] [CrossRef]
  203. Tahir, M.B.; Nawaz, T.; Nabi, G.; Sagir, M.; Rafique, M.; Ahmed, A.; Muhammad, S. Photocatalytic Degradation and Hydrogen Evolution Using Bismuth Tungstate Based Nanocomposites under Visible Light Irradiation. Int. J. Hydrogen Energy 2020, 45, 22833–22847. [Google Scholar] [CrossRef]
  204. Kumar, A.; Chandel, M.; Sharma, A.; Thakur, M.; Kumar, A.; Pathania, D.; Singh, L. Robust Visible Light Active PANI/LaFeO3/CoFe2O4 Ternary Heterojunction for the Photo-Degradation and Mineralization of Pharmaceutical Effluent: Clozapine. J. Environ. Chem. Eng. 2021, 9, 106159. [Google Scholar] [CrossRef]
  205. 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]
  206. Xu, P.; Zheng, D.; Xie, Z.; He, Q.; Yu, J. The Degradation of Ibuprofen in a Novel Microbial Fuel Cell with PANi@CNTs/SS Bio-Anode and CuInS2 Photocatalytic Cathode: Property, Efficiency and Mechanism. J. Clean. Prod. 2020, 265, 121872. [Google Scholar] [CrossRef]
  207. Tian, X.; Liu, S.; Zhang, B.; Wang, S.; Dong, S.; Liu, Y.; Feng, L.; Zhang, L. Carbonized Polyaniline-Activated Peracetic Acid Advanced Oxidation Process for Organic Removal: Efficiency and Mechanisms. Environ. Res. 2023, 219, 115035. [Google Scholar] [CrossRef] [PubMed]
  208. Silvestri, S.; Ferreira, C.D.; Oliveira, V.; Varejão, J.M.T.B.; Labrincha, J.A.; Tobaldi, D.M. Synthesis of PPy-ZnO Composite Used as Photocatalyst for the Degradation of Diclofenac under Simulated Solar Irradiation. J. Photochem. Photobiol. A Chem. 2019, 375, 261–269. [Google Scholar] [CrossRef]
  209. Liu, J.; Zhou, S.; Gu, P.; Zhang, T.; Chen, D.; Li, N.; Xu, Q.; Lu, J. Conjugate Polymer-Clothed TiO2@V2O5 Nanobelts and Their Enhanced Visible Light Photocatalytic Performance in Water Remediation. J. Colloid Interface Sci. 2020, 578, 402–411. [Google Scholar] [CrossRef]
  210. Anirudhan, T.S.; Divya, P.L.; Nima, J. Utilization of Polypyrrole Coated Iron-Doped Titania Based Hydrogel for the Removal of Tetracycline Hydrochloride from Aqueous Solutions: Adsorption and Photocatalytic Degradation Studies. Environ. Nanotechnol. Monit. Manag. 2015, 4, 106–117. [Google Scholar] [CrossRef]
  211. Zia, J.; Kashyap, J.; Riaz, U. Facile Synthesis of Polypyrrole Encapsulated V2O5 Nanohybrids for Visible Light Driven Green Sonophotocatalytic Degradation of Antibiotics. J. Mol. Liq. 2018, 272, 834–850. [Google Scholar] [CrossRef]
  212. Kumar, R.S.; Govindan, K.; Ramakrishnan, S.; Kim, A.R.; Kim, J.-S.; Yoo, D.J. Fe3O4 Nanorods Decorated on Polypyrrole/Reduced Graphene Oxide for Electrochemical Detection of Dopamine and Photocatalytic Degradation of Acetaminophen. Appl. Surf. Sci. 2021, 556, 149765. [Google Scholar] [CrossRef]
  213. Zhu, Z.; Tang, X.; Ma, C.; Song, M.; Gao, N.; Wang, Y.; Huo, P.; Lu, Z.; Yan, Y. Fabrication of Conductive and High-Dispersed Ppy@Ag/g-C3N4 Composite Photocatalysts for Removing Various Pollutants in Water. Appl. Surf. Sci. 2016, 387, 366–374. [Google Scholar] [CrossRef]
  214. Abinaya, M.; Rajakumaran, R.; Chen, S.-M.; Karthik, R.; Muthuraj, V. In Situ Synthesis, Characterization, and Catalytic Performance of Polypyrrole Polymer-Incorporated Ag2MoO4 Nanocomposite for Detection and Degradation of Environmental Pollutants and Pharmaceutical Drugs. ACS Appl. Mater. Interfaces 2019, 11, 38321–38335. [Google Scholar] [CrossRef] [PubMed]
  215. Das, K.K.; Patnaik, S.; Mansingh, S.; Behera, A.; Mohanty, A.; Acharya, C.; Parida, K.M. Enhanced Photocatalytic Activities of Polypyrrole Sensitized Zinc Ferrite/Graphitic Carbon Nitride n-n Heterojunction towards Ciprofloxacin Degradation, Hydrogen Evolution and Antibacterial Studies. J. Colloid Interface Sci. 2020, 561, 551–567. [Google Scholar] [CrossRef] [PubMed]
  216. Zhu, S.; Ye, Z.; Liu, Z.; Chen, Z.; Li, J.; Xiang, Z. Adsorption Characteristics of Polymer Solutions on Media Surfaces and Their Main Influencing Factors. Polymers 2021, 13, 1774. [Google Scholar] [CrossRef] [PubMed]
  217. Gupta, A.; Sharma, V.; Mishra, P.K.; Ekielski, A. A Review on Polyacrylonitrile as an Effective and Economic Constituent of Adsorbents for Wastewater Treatment. Molecules 2022, 27, 8689. [Google Scholar] [CrossRef] [PubMed]
  218. Crini, G.; Lichtfouse, E. Advantages and Disadvantages of Techniques Used for Wastewater Treatment. Environ. Chem. Lett. 2019, 17, 145–155. [Google Scholar] [CrossRef]
  219. Akhtar, J.; Amin, N.A.S.; Shahzad, K. A Review on Removal of Pharmaceuticals from Water by Adsorption. Desalination Water Treat. 2016, 57, 12842–12860. [Google Scholar] [CrossRef]
  220. Shaipulizan, N.S.; Md Jamil, S.N.A.; Kamaruzaman, S.; Subri, N.N.S.; Adeyi, A.A.; Abdullah, A.H.; Abdullah, L.C. Preparation of Ethylene Glycol Dimethacrylate (EGDMA)-Based Terpolymer as Potential Sorbents for Pharmaceuticals Adsorption. Polymers 2020, 12, 423. [Google Scholar] [CrossRef]
  221. Kamaraj, M.; Suresh Babu, P.; Shyamalagowri, S.; Pavithra, M.K.S.; Aravind, J.; Kim, W.; Govarthanan, M. β-Cyclodextrin Polymer Composites for the Removal of Pharmaceutical Substances, Endocrine Disruptor Chemicals, and Dyes from Aqueous Solution—A Review of Recent Trends. J. Environ. Manag. 2024, 351, 119830. [Google Scholar] [CrossRef] [PubMed]
  222. Thomas, R.; Mary, Y.S.; Resmi, K.S.; Narayana, B.; Sarojini, S.B.K.; Armaković, S.; Armaković, S.J.; Vijayakumar, G.; Alsenoy, C.V.; Mohan, B.J. Synthesis and Spectroscopic Study of Two New Pyrazole Derivatives with Detailed Computational Evaluation of Their Reactivity and Pharmaceutical Potential. J. Mol. Struct. 2019, 1181, 599–612. [Google Scholar] [CrossRef]
  223. Mary, Y.S.; Mary, Y.S.; Thomas, R.; Narayana, B.; Samshuddin, S.; Sarojini, B.K.; Armaković, S.; Armaković, S.J.; Pillai, G.G. Theoretical Studies on the Structure and Various Physico-Chemical and Biological Properties of a Terphenyl Derivative with Immense Anti-Protozoan Activity. Polycycl. Aromat. Compd. 2021, 41, 825–840. [Google Scholar] [CrossRef]
  224. Beegum, S.; Mary, Y.S.; Mary, Y.S.; Thomas, R.; Armaković, S.; Armaković, S.J.; Zitko, J.; Dolezal, M.; Van Alsenoy, C. Exploring the Detailed Spectroscopic Characteristics, Chemical and Biological Activity of Two Cyanopyrazine-2-Carboxamide Derivatives Using Experimental and Theoretical Tools. Spectrochim. Acta Part A Mol. Biomol. Spectrosc. 2020, 224, 117414. [Google Scholar] [CrossRef] [PubMed]
  225. Haruna, K.; Kumar, V.S.; Armaković, S.J.; Armaković, S.; Mary, Y.S.; Thomas, R.; Popoola, S.A.; Almohammedi, A.R.; Roxy, M.S.; Al-Saadi, A.A. Spectral Characterization, Thermochemical Studies, Periodic SAPT Calculations and Detailed Quantum Mechanical Profiling Various Physico-Chemical Properties of 3,4-Dichlorodiuron. Spectrochim. Acta Part A Mol. Biomol. Spectrosc. 2020, 228, 117580. [Google Scholar] [CrossRef] [PubMed]
  226. Al-Otaibi, J.S.; Mary, Y.S.; Armaković, S.; Thomas, R. Hybrid and Bioactive Cocrystals of Pyrazinamide with Hydroxybenzoic Acids: Detailed Study of Structure, Spectroscopic Characteristics, Other Potential Applications and Noncovalent Interactions Using SAPT. J. Mol. Struct. 2020, 1202, 127316. [Google Scholar] [CrossRef]
  227. Bielenica, A.; Beegum, S.; Mary, Y.S.; Mary, Y.S.; Thomas, R.; Armaković, S.; Armaković, S.J.; Madeddu, S.; Struga, M.; Van Alsenoy, C. Experimental and Computational Analysis of 1-(4-Chloro-3-Nitrophenyl)-3-(3,4-Dichlorophenyl)Thiourea. J. Mol. Struct. 2020, 1205, 127587. [Google Scholar] [CrossRef]
  228. Neto, O.P.V. Intelligent Computational Nanotechnology: The Role of Computational Intelligence in the Development of Nanoscience and Nanotechnology. J. Comput. Theor. Nanosci. 2014, 11, 928–944. [Google Scholar] [CrossRef]
  229. Gooneie, A.; Schuschnigg, S.; Holzer, C. A Review of Multiscale Computational Methods in Polymeric Materials. Polymers 2017, 9, 16. [Google Scholar] [CrossRef]
  230. Stephens, P.J.; Devlin, F.J.; Chabalowski, C.F.; Frisch, M.J. Ab Initio Calculation of Vibrational Absorption and Circular Dichroism Spectra Using Density Functional Force Fields. J. Phys. Chem. 1994, 98, 11623–11627. [Google Scholar] [CrossRef]
  231. Vosko, S.H.; Wilk, L.; Nusair, M. Accurate Spin-Dependent Electron Liquid Correlation Energies for Local Spin Density Calculations: A Critical Analysis. Can. J. Phys. 1980, 58, 1200–1211. [Google Scholar] [CrossRef]
  232. Lee, C.; Yang, W.; Parr, R.G. Development of the Colle-Salvetti Correlation-Energy Formula into a Functional of the Electron Density. Phys. Rev. B 1988, 37, 785. [Google Scholar] [CrossRef] [PubMed]
  233. Becke, A.D. Density-functional Thermochemistry. III. The Role of Exact Exchange. J. Chem. Phys. 1993, 98, 5648–5652. [Google Scholar] [CrossRef]
  234. Perdew, J.P.; Burke, K.; Ernzerhof, M. Generalized Gradient Approximation Made Simple. Phys. Rev. Lett. 1996, 77, 3865–3868, Erratum in Phys. Rev. Lett. 1997, 78, 1396. [Google Scholar] [CrossRef] [PubMed]
  235. Stewart, J.J.P. Optimization of Parameters for Semiempirical Methods II. Applications. J. Comput. Chem. 1989, 10, 221–264. [Google Scholar] [CrossRef]
  236. Stewart, J.J.P. Optimization of Parameters for Semiempirical Methods I. Method. J. Comput. Chem. 1989, 10, 209–220. [Google Scholar] [CrossRef]
  237. Stewart, J.J.P. Optimization of Parameters for Semiempirical Methods V: Modification of NDDO Approximations and Application to 70 Elements. J. Mol. Model. 2007, 13, 1173–1213. [Google Scholar] [CrossRef] [PubMed]
  238. Stewart, J.J.P. Optimization of Parameters for Semiempirical Methods VI: More Modifications to the NDDO Approximations and Re-Optimization of Parameters. J. Mol. Model. 2013, 19, 1–32. [Google Scholar] [CrossRef] [PubMed]
  239. Bannwarth, C.; Caldeweyher, E.; Ehlert, S.; Hansen, A.; Pracht, P.; Seibert, J.; Spicher, S.; Grimme, S. Extended Tight-Binding Quantum Chemistry Methods. WIREs Comput. Mol. Sci. 2021, 11, e1493. [Google Scholar] [CrossRef]
  240. Bannwarth, C.; Ehlert, S.; Grimme, S. GFN2-xTB—An Accurate and Broadly Parametrized Self-Consistent Tight-Binding Quantum Chemical Method with Multipole Electrostatics and Density-Dependent Dispersion Contributions. J. Chem. Theory Comput. 2019, 15, 1652–1671. [Google Scholar] [CrossRef]
  241. Ehlert, S.; Stahn, M.; Spicher, S.; Grimme, S. Robust and Efficient Implicit Solvation Model for Fast Semiempirical Methods. J. Chem. Theory Comput. 2021, 17, 4250–4261. [Google Scholar] [CrossRef] [PubMed]
  242. Pracht, P.; Caldeweyher, E.; Ehlert, S.; Grimme, S. A Robust Non-Self-Consistent Tight-Binding Quantum Chemistry Method for Large Molecules. ChemRxiv 2019. [Google Scholar] [CrossRef]
  243. Grimme, S.; Bannwarth, C.; Shushkov, P. A Robust and Accurate Tight-Binding Quantum Chemical Method for Structures, Vibrational Frequencies, and Noncovalent Interactions of Large Molecular Systems Parametrized for All Spd-Block Elements (Z = 1–86). J. Chem. Theory Comput. 2017, 13, 1989–2009. [Google Scholar] [CrossRef] [PubMed]
  244. Blau, S.; Spotte-Smith, E.W.C.; Wood, B.; Dwaraknath, S.; Persson, K. Accurate, Automated Density Functional Theory for Complex Molecules Using On-the-Fly Error Correction. ChemRxiv 2020. [Google Scholar] [CrossRef]
  245. Yan, Y.; Yu, X.; Shao, C.; Hu, Y.; Huang, W.; Li, Y. Atomistic Structural Engineering of Conjugated Microporous Polymers Promotes Photocatalytic Biomass Valorization. Adv. Funct. Mater. 2023, 33, 2304604. [Google Scholar] [CrossRef]
  246. Mizera, A.; Dubis, A.T.; Łapiński, A. Density Functional Theory Studies of Polypyrrole and Polypyrrole Derivatives; Substituent Effect on the Optical and Electronic Properties. Polymer 2022, 255, 125127. [Google Scholar] [CrossRef]
  247. Adekoya, O.C.; Adekoya, G.J.; Sadiku, E.R.; Hamam, Y.; Ray, S.S. Application of DFT Calculations in Designing Polymer-Based Drug Delivery Systems: An Overview. Pharmaceutics 2022, 14, 1972. [Google Scholar] [CrossRef] [PubMed]
  248. Shah, E.V.; Patel, C.M.; Roy, D.R. Structure, Electronic, Optical and Thermodynamic Behavior on the Polymerization of PMMA: A DFT Investigation. Comput. Biol. Chem. 2018, 72, 192–198. [Google Scholar] [CrossRef]
  249. Franco, F.C.; Padama, A.A.B. DFT and TD-DFT Study on the Structural and Optoelectronic Characteristics of Chemically Modified Donor-Acceptor Conjugated Oligomers for Organic Polymer Solar Cells. Polymer 2016, 97, 55–62. [Google Scholar] [CrossRef]
  250. Nicholls, I.A.; Golker, K.; Olsson, G.D.; Suriyanarayanan, S.; Wiklander, J.G. The Use of Computational Methods for the Development of Molecularly Imprinted Polymers. Polymers 2021, 13, 2841. [Google Scholar] [CrossRef]
  251. Grimme-Lab/Xtb 2024. Available online: https://github.com/grimme-lab/xtb/releases (accessed on 24 June 2024).
  252. User Guide to Semiempirical Tight Binding—Xtb Doc 2023 Documentation. Available online: https://xtb-docs.readthedocs.io/en/latest/ (accessed on 29 May 2024).
  253. Startseite. Available online: https://www.chemie.uni-bonn.de/grimme/de (accessed on 29 May 2024).
  254. Armaković, S.; Armaković, S.J. Online and Desktop Graphical User Interfaces for Xtb Programme from Atomistica.Online Platform. Mol. Simul. 2024, 50, 560–570. [Google Scholar] [CrossRef]
  255. Armaković, S.; Armaković, S.J. Atomistica.Online—Web Application for Generating Input Files for ORCA Molecular Modelling Package Made with the Anvil Platform. Mol. Simul. 2023, 49, 117–123. [Google Scholar] [CrossRef]
  256. Liakos, D.G.; Guo, Y.; Neese, F. Comprehensive Benchmark Results for the Domain Based Local Pair Natural Orbital Coupled Cluster Method (DLPNO-CCSD(T)) for Closed- and Open-Shell Systems. J. Phys. Chem. A 2020, 124, 90–100. [Google Scholar] [CrossRef] [PubMed]
  257. Guo, Y.; Riplinger, C.; Liakos, D.G.; Becker, U.; Saitow, M.; Neese, F. Linear Scaling Perturbative Triples Correction Approximations for Open-Shell Domain-Based Local Pair Natural Orbital Coupled Cluster Singles and Doubles Theory [DLPNO-CCSD(T0/T)]. J. Chem. Phys. 2020, 152, 024116. [Google Scholar] [CrossRef] [PubMed]
  258. Neese, F. The SHARK Integral Generation and Digestion System. J. Comput. Chem. 2022, 44, 381–396. [Google Scholar] [CrossRef] [PubMed]
  259. Teale, A.M.; Helgaker, T.; Savin, A.; Adamo, C.; Aradi, B.; Arbuznikov, A.V.; Ayers, P.W.; Baerends, E.J.; Barone, V.; Calaminici, P.; et al. DFT Exchange: Sharing Perspectives on the Workhorse of Quantum Chemistry and Materials Science. Phys. Chem. Chem. Phys. 2022, 24, 28700–28781. [Google Scholar] [CrossRef]
  260. Neese, F.; Wennmohs, F.; Hansen, A.; Becker, U. Efficient, Approximate and Parallel Hartree–Fock and Hybrid DFT Calculations. A ‘Chain-of-Spheres’ Algorithm for the Hartree–Fock Exchange. Chem. Phys. 2009, 356, 98–109. [Google Scholar] [CrossRef]
  261. Neese, F. Software Update: The ORCA Program System—Version 5.0. WIREs Comput. Mol. Sci. 2022, 12, e1606. [Google Scholar] [CrossRef]
  262. Neese, F. Software Update: The ORCA Program System, Version 4.0. WIREs Comput. Mol. Sci. 2018, 8, e1327. [Google Scholar] [CrossRef]
  263. Neese, F. The ORCA Program System. WIREs Comput. Mol. Sci. 2012, 2, 73–78. [Google Scholar] [CrossRef]
  264. Neese, F.; Wennmohs, F.; Becker, U.; Riplinger, C. The ORCA Quantum Chemistry Program Package. J. Chem. Phys. 2020, 152, 224108. [Google Scholar] [CrossRef] [PubMed]
  265. ORCA Forum—Portal. Available online: https://orcaforum.kofo.mpg.de/app.php/portal (accessed on 29 May 2024).
  266. Lu, T.; Chen, F. Multiwfn: A Multifunctional Wavefunction Analyzer. J. Comput. Chem. 2012, 33, 580–592. [Google Scholar] [CrossRef] [PubMed]
  267. Multiwfn Website. Available online: http://sobereva.com/multiwfn/ (accessed on 30 May 2024).
  268. Saleh, G.; Gatti, C.; Lo Presti, L. Non-Covalent Interaction via the Reduced Density Gradient: Independent Atom Model vs Experimental Multipolar Electron Densities. Comput. Theor. Chem. 2012, 998, 148–163. [Google Scholar] [CrossRef]
  269. Schrödinger—Physics-Based Software Platform for Molecular Discovery & Design. Available online: https://www.schrodinger.com/ (accessed on 30 May 2024).
  270. Materials Science. Available online: https://www.schrodinger.com/materials-science/ (accessed on 30 May 2024).
  271. Schrödinger Platform Application: Polymeric Materials. Available online: https://www.schrodinger.com/materials-science/solutions/polymeric-materials/ (accessed on 30 May 2024).
  272. Kumar, R.; Akbarinejad, A.; Jasemizad, T.; Fucina, R.; Travas-Sejdic, J.; Padhye, L.P. The Removal of Metformin and Other Selected PPCPs from Water by Poly(3,4-Ethylenedioxythiophene) Photocatalyst. Sci. Total Environ. 2021, 751, 142302. [Google Scholar] [CrossRef] [PubMed]
  273. Joseph, A.; Vijayanandan, A. Review on Support Materials Used for Immobilization of Nano-Photocatalysts for Water Treatment Applications. Inorg. Chim. Acta 2023, 545, 121284. [Google Scholar] [CrossRef]
  274. Lei, P.; Wang, F.; Gao, X.; Ding, Y.; Zhang, S.; Zhao, J.; Liu, S.; Yang, M. Immobilization of TiO2 Nanoparticles in Polymeric Substrates by Chemical Bonding for Multi-Cycle Photodegradation of Organic Pollutants. J. Hazard. Mater. 2012, 227–228, 185–194. [Google Scholar] [CrossRef]
  275. Melinte, V.; Culica, M.E.; Chibac-Scutaru, A.L. Cellulose Acetate/Polyurethane Blend as Support Matrix with High Optical Transparency and Improved Mechanical Properties for Photocatalyst CeO2 Nanoparticles Immobilization. Int. J. Biol. Macromol. 2023, 251, 126210. [Google Scholar] [CrossRef]
  276. Park, H.; Park, Y.; Kim, W.; Choi, W. Surface Modification of TiO2 Photocatalyst for Environmental Applications. J. Photochem. Photobiol. C Photochem. Rev. 2013, 15, 1–20. [Google Scholar] [CrossRef]
  277. Li, X.; Wang, D.; Cheng, G.; Luo, Q.; An, J.; Wang, Y. Preparation of Polyaniline-Modified TiO2 Nanoparticles and Their Photocatalytic Activity under Visible Light Illumination. Appl. Catal. B Environ. 2008, 81, 267–273. [Google Scholar] [CrossRef]
  278. Liang, H.; Li, X. Visible-Induced Photocatalytic Reactivity of Polymer–Sensitized Titania Nanotube Films. Appl. Catal. B Environ. 2009, 86, 8–17. [Google Scholar] [CrossRef]
  279. Muktha, B.; Mahanta, D.; Patil, S.; Madras, G. Synthesis and Photocatalytic Activity of Poly(3-Hexylthiophene)/TiO2 Composites. J. Solid State Chem. 2007, 180, 2986–2989. [Google Scholar] [CrossRef]
  280. Wang, D.; Zhang, J.; Luo, Q.; Li, X.; Duan, Y.; An, J. Characterization and Photocatalytic Activity of Poly(3-Hexylthiophene)-Modified TiO2 for Degradation of Methyl Orange under Visible Light. J. Hazard. Mater. 2009, 169, 546–550. [Google Scholar] [CrossRef] [PubMed]
  281. Zhu, Y.; Dan, Y. Photocatalytic Activity of Poly(3-Hexylthiophene)/Titanium Dioxide Composites for Degrading Methyl Orange. Sol. Energy Mater. Sol. Cells 2010, 94, 1658–1664. [Google Scholar] [CrossRef]
  282. Song, L.; Qiu, R.; Mo, Y.; Zhang, D.; Wei, H.; Xiong, Y. Photodegradation of Phenol in a Polymer-Modified TiO2 Semiconductor Particulate System under the Irradiation of Visible Light. Catal. Commun. 2007, 8, 429–433. [Google Scholar] [CrossRef]
  283. Valadez-Renteria, E.; Oliva, J.; Rodriguez-Gonzalez, V. Photocatalytic Materials Immobilized on Recycled Supports and Their Role in the Degradation of Water Contaminants: A Timely Review. Sci. Total Environ. 2022, 807, 150820. [Google Scholar] [CrossRef] [PubMed]
  284. Zhao, J.; Biswas, M.R.U.D.; Oh, W.-C. A Novel BiVO4-GO-TiO2-PANI Composite for Upgraded Photocatalytic Performance under Visible Light and Its Non-Toxicity. Environ. Sci. Pollut. Res. 2019, 26, 11888–11904. [Google Scholar] [CrossRef] [PubMed]
  285. Ok, S.; Vayer, M.; Sinturel, C. A Decade of Innovation and Progress in Understanding the Morphology and Structure of Heterogeneous Polymers in Rigid Confinement. Soft Matter 2021, 17, 7430–7458. [Google Scholar] [CrossRef] [PubMed]
  286. Ren, J.; Liu, Y.; Li, H. Incorporating Polymers within a Single-crystal: From Heterogeneous Structure to Multiple Functions. J. Polym. Sci. 2022, 60, 1151–1173. [Google Scholar] [CrossRef]
  287. Ghobeira, R.; Esbah Tabaei, P.S.; Morent, R.; De Geyter, N. Chemical Characterization of Plasma-Activated Polymeric Surfaces via XPS Analyses: A Review. Surf. Interfaces 2022, 31, 102087. [Google Scholar] [CrossRef]
  288. Ran, X.; Duan, L.; Chen, X.; Yang, X. Photocatalytic Degradation of Organic Dyes by the Conjugated Polymer Poly(1,3,4-Oxadiazole)s and Its Photocatalytic Mechanism. J. Mater. Sci. 2018, 53, 7048–7059. [Google Scholar] [CrossRef]
  289. Chin, S.S.; Chiang, K.; Fane, A.G. The Stability of Polymeric Membranes in a TiO2 Photocatalysis Process. J. Membr. Sci. 2006, 275, 202–211. [Google Scholar] [CrossRef]
  290. Chamas, A.; Moon, H.; Zheng, J.; Qiu, Y.; Tabassum, T.; Jang, J.H.; Abu-Omar, M.; Scott, S.L.; Suh, S. Degradation Rates of Plastics in the Environment. ACS Sustain. Chem. Eng. 2020, 8, 3494–3511. [Google Scholar] [CrossRef]
  291. Garratt, A.; Nguyen, K.; Brooke, A.; Taylor, M.J.; Francesconi, M.G. Photocatalytic Hydrolysis─A Sustainable Option for the Chemical Upcycling of Polylactic Acid. ACS Environ. Au 2023, 3, 342–347. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Overview of various photocatalytic applications of polymers. Reprinted with permission from Ref. [19], copyright Royal Society of Chemistry.
Figure 1. Overview of various photocatalytic applications of polymers. Reprinted with permission from Ref. [19], copyright Royal Society of Chemistry.
Catalysts 14 00447 g001
Figure 2. Optimization of injection stretch blow-molding. (af) illustrate the process steps. Straight arrows follow the process sequence; curved arrows follow the optimization sequence. N—screw speed, Tb—barrel temperature profile, Vinj—injection volume. Reprinted with permission from Ref. [159], copyright MDPI.
Figure 2. Optimization of injection stretch blow-molding. (af) illustrate the process steps. Straight arrows follow the process sequence; curved arrows follow the optimization sequence. N—screw speed, Tb—barrel temperature profile, Vinj—injection volume. Reprinted with permission from Ref. [159], copyright MDPI.
Catalysts 14 00447 g002
Figure 3. The scheme of photocatalysis, based on a (a) conducting polymer (CP) nanocomposite and (b) conducting polymer–metal oxide hybrid modification. Reprinted and adjusted with permission from Ref. [192], copyright MDPI.
Figure 3. The scheme of photocatalysis, based on a (a) conducting polymer (CP) nanocomposite and (b) conducting polymer–metal oxide hybrid modification. Reprinted and adjusted with permission from Ref. [192], copyright MDPI.
Catalysts 14 00447 g003
Figure 4. Venn diagram illustrating the connections between distinct computational methods for molecular simulations.
Figure 4. Venn diagram illustrating the connections between distinct computational methods for molecular simulations.
Catalysts 14 00447 g004
Figure 5. Suggested workflow for geometrical optimization and property calculations of long polymer chains.
Figure 5. Suggested workflow for geometrical optimization and property calculations of long polymer chains.
Catalysts 14 00447 g005
Figure 6. PMMA optimized via (a) GFN2-xTB and (b) B3LYP-D3/6-31G(d,p). Gray spheres represent carbon atoms, white spheres represent hydrogen atoms, red spheres represent oxygen atoms.
Figure 6. PMMA optimized via (a) GFN2-xTB and (b) B3LYP-D3/6-31G(d,p). Gray spheres represent carbon atoms, white spheres represent hydrogen atoms, red spheres represent oxygen atoms.
Catalysts 14 00447 g006
Figure 7. (a) RDG scatter plot and (b) RDG surfaces of PMMA polymer chain.
Figure 7. (a) RDG scatter plot and (b) RDG surfaces of PMMA polymer chain.
Catalysts 14 00447 g007
Figure 8. Advantages of polymers in photocatalytic removal of pharmaceutical contaminants from water.
Figure 8. Advantages of polymers in photocatalytic removal of pharmaceutical contaminants from water.
Catalysts 14 00447 g008
Table 2. Photocatalysis of pharmaceuticals using PANI composite materials.
Table 2. Photocatalysis of pharmaceuticals using PANI composite materials.
Type of Polymer CompositeSynthesis of PhotocatalystContaminantLightRemoval Efficiency (%)Lamp PowerReference
ZnO/PANI nanocompositeChemical adsorptionAmpicillinSunlight41% in 120 min-[198]
ZnO/PANI nanocompositeIn situ chemical polymerizationMetronidazoleVisible97% in 180 min300 W Xenon lamp[199]
TiO2/PANI nanocompositeOxidative polymerizationPropranololUV irradiation34% in 60 min125 W High-pressure mercury lamp[200]
TiO2/PANI nanocompositeOxidative polymerizationAmitriptylineUV irradiation45% in 60 min125 W High-pressure mercury lamp[200]
WS2/PANI nanocompositeIn situ polymerizationNitrofurantoinVisible99% in 120 minXenon arc lamp[201]
GdFeO3/PANIIn situ oxidative polymerizationAcetaminophenVisible88% in 60 min30 W Light emitting diode[202]
Bi2WO6/PANIIntercalationCiprofloxacinVisible98% in 90 min50 W Halogen lamp[203]
PANI/LaFeO3/CoFe2O4 ternary heterojunctionIn situ polymerizationClozapineVisible92% in 120 min50 W Halogen lamp[204]
3D PANI/Perylene diimideIn situ growthTetracyclineVisible70% in 120 min5 W LED lamp (420 nm cut-off filter)[205]
PANi@carbon nanotubes/stainless steelIn situ polymerizationIbuprofenVisible76% in 35 min300 W Xenon lamp[206]
Carbonized PANI-activated peracetic acidInterfacial polymerizationSulfamethoxazoleVisible100% in 60 min30 W Light emitting diode[207]
Carbonized PANI-activated peracetic acidInterfacial polymerizationNaproxenVisible100% in 60 min30 W Light emitting diode[207]
Carbonized PANI-activated peracetic acidInterfacial polymerizationAntipyrineVisible30% in 60 min30 W Light emitting diode[207]
Table 3. Photocatalysis of pharmaceuticals using PPy composite materials.
Table 3. Photocatalysis of pharmaceuticals using PPy composite materials.
Type of Polymer CompositeSynthesis of PhotocatalystContaminantLightRemoval Efficiency (%)Lamp PowerReference
PPy-ZnOPolymerizationDiclofenacUV and visible81% in 60 minXenon lamp (wavelength range 250–800 nm)[208]
TiO2@V2O5-PPyIn situ polymerizationTetracyclineVisible98% in 120 min300 W Xenon lamp[209]
TiO2@V2O5-PPyIn situ polymerizationDoxycyclineVisible96% in 120 min300 W Xenon lamp[209]
TiO2@V2O5-PPyIn situ polymerizationOxytetracyclineVisible85% in 120 min300 W Xenon lamp[209]
Fe-TiO2-PPyPolymerizationTetracycline hydrochlorideSunlight96% in 180 min-[210]
PPy/V2O5Ultrasound-assisted synthesisCiprofloxacinVisible53% in 50 min300 W Xenon arc lamp[211]
PPy/V2O5Ultrasound-assisted synthesisErythromycinVisible76% in 50 min300 W Xenon arc lamp[211]
Fe3O4@PPy/rGOChemical refluxAcetaminophenUV and visible84% in 120 min250 W tungsten-halogen[212]
PPy@Ag/g-C3N4Calcination followed by surface polymerizationDanofloxacinVisible90% in 60 min300 W Xenon lamp[213]
PPy@Ag/g-C3N4Calcination followed by surface polymerizationTetracyclineVisible95% in 60 min300 W Xenon lamp[213]
PPy@Ag/g-C3N4Calcination followed by surface polymerizationCiprofloxacinVisible92% in 60 min300 W Xenon lamp[213]
PPy@Ag/g-C3N4Calcination followed by surface polymerizationGatifloxacinVisible89% in 60 min300 W Xenon lamp[213]
PPy@Ag/g-C3N4Calcination followed by surface polymerizationEnrofloxacin hydrochlorideVisible91% in 60 min300 W Xenon lamp[213]
Ag2MoO4/PPyIn situ synthesisCiprofloxacinSunlight100% in 10 min-[214]
PPy-sensitized zinc ferrite/graphitic carbon nitrideIn situ polymerizationCiprofloxacinVisible92% in 120 min300 W Xenon lamp[215]
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

Armaković, S.J.; Armaković, S.; Savanović, M.M. Photocatalytic Application of Polymers in Removing Pharmaceuticals from Water: A Comprehensive Review. Catalysts 2024, 14, 447. https://doi.org/10.3390/catal14070447

AMA Style

Armaković SJ, Armaković S, Savanović MM. Photocatalytic Application of Polymers in Removing Pharmaceuticals from Water: A Comprehensive Review. Catalysts. 2024; 14(7):447. https://doi.org/10.3390/catal14070447

Chicago/Turabian Style

Armaković, Sanja J., Stevan Armaković, and Maria M. Savanović. 2024. "Photocatalytic Application of Polymers in Removing Pharmaceuticals from Water: A Comprehensive Review" Catalysts 14, no. 7: 447. https://doi.org/10.3390/catal14070447

APA Style

Armaković, S. J., Armaković, S., & Savanović, M. M. (2024). Photocatalytic Application of Polymers in Removing Pharmaceuticals from Water: A Comprehensive Review. Catalysts, 14(7), 447. https://doi.org/10.3390/catal14070447

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