0% found this document useful (0 votes)
24 views19 pages

1 s2.0 S235294072300207X Main

This article reviews the use of polymeric materials as catalysts in medical, environmental, and energy applications, highlighting their advantages such as high surface area and chemical inertness. It discusses specific types of catalytic polymers, including graphitic carbon nitride and covalent organic frameworks, and their roles in photocatalysis for processes like pollutant degradation and energy generation. The review also identifies challenges and future research directions in the field of polymeric catalysts.

Uploaded by

Madhan
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
0% found this document useful (0 votes)
24 views19 pages

1 s2.0 S235294072300207X Main

This article reviews the use of polymeric materials as catalysts in medical, environmental, and energy applications, highlighting their advantages such as high surface area and chemical inertness. It discusses specific types of catalytic polymers, including graphitic carbon nitride and covalent organic frameworks, and their roles in photocatalysis for processes like pollutant degradation and energy generation. The review also identifies challenges and future research directions in the field of polymeric catalysts.

Uploaded by

Madhan
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd

Applied Materials Today 35 (2023) 101937

Contents lists available at ScienceDirect

Applied Materials Today


journal homepage: [Link]/locate/apmt

Polymer materials as catalysts for medical, environmental, and


energy applications
Federico Mazur, Andy-Hoai Pham, Rona Chandrawati *
School of Chemical Engineering and Australian Centre for Nanomedicine (ACN), The University of New South Wales, Sydney, NSW 2052, Australia

A R T I C L E I N F O A B S T R A C T

Keywords: Catalysts are the backbone of almost all aspects in life, increasing the rate of chemical and biological reactions for
Catalytic polymeric materials greater efficiency. They are ubiquitous in industrial processes, turning raw materials into useful products like
Photocatalysis milk to yogurt or petroleum into plastic jugs. In the human body, they are called enzymes, and help in creating
Graphitic carbon nitride
signals for bodily functions and digesting food, among other tasks. Polymers have commonly been used in
Covalent organic frameworks
catalytic applications as supports for catalytic compounds because of their high surface area, porosity, surface
functionality, and chemical inertness towards various harsh conditions. Recently a greater emphasis has been
placed on using polymer materials themselves as catalysts, or in conjunction with other compounds to syner­
gically enhance catalytic activity. These types of polymeric catalysts have been used in a range of fields including
medical-based applications like antibacterial and wound healing, environmental remediation applications like
pollutant degradation and sustainable synthesis, as well as energy applications like sustainable energy generation
and storage. In this review, we discuss the various platforms developed for these three main fields and outline
challenges and emerging research areas going forward.

1. Introduction expensive [4]. Heterogenous catalysts are in a different phase than the
reactants, making them readily separated and recycled, however
Catalysis is a key aspect in industrial and biological processes that generally at the cost of lower rates and selectivity [5]. Finally, bio­
allows for reactions to occur more quickly and efficiently [1]. This is catalysts involve natural proteins such as enzymes, which catalyze
achieved via the use of catalysts, which are substances that accelerate a chemical reactions with high selectivity and efficiency, but tend to
chemical reaction without themselves being consumed in the process. denature at harsh reaction conditions [6]. Although catalysts can be
They accomplish this by altering the reaction pathway, reducing the reused, they typically have reduced effectiveness due to fouling,
activation energy of the reaction, and selectively promoting specific poisoning, thermal degradation, leaching of active metal species, and/or
reactions over others. The global catalyst market was valued at $USD other wear and tear effects [7]. Catalysts provide a transient weak bond
35.5 billion in 2020, and is expected to reach $USD 57.5 billion by 2030, with the reactant which compensates for the energy required to break
with most of the market value share applied in petroleum refining, the old bond until the new bond is formed to stabilize the product. The
chemical synthesis, polymers and petrochemicals, and environmental specific mechanism by which catalysts achieve this depends on the
uses [2]. The increase in demand for catalysts in these applications, catalyst and reaction. As such, catalysts can also be categorized by
amongst others, is driving the need for catalysts which allow for mechanisms of action, such as acid-base, oxidation-reduction (redox),
increased process optimization, yield, as well as reduced costs and en­ photo-, and coordination catalysis. Acid-base catalysis relies on mole­
ergy consumption. cules which can donate (acid) or accept (base) a proton (H+) to increase
Catalysts can be broadly categorized into three main groups: ho­ the rate of the chemical reaction [8]. On the other hand, redox catalysis
mogenous, heterogenous, and biocatalysts (Fig. 1) [3]. For the former, involves the transfer of electrons (e− ) between reactants in a chemical
the catalyst compound is in the same phase as the reactants, most reaction [9]. In other words, the catalyst donates or accepts e− to/from
commonly liquid phase, and can achieve high rates, with good selec­ the reactants to facilitate the reaction. Photocatalysis also involves the
tivity. However, recovery of the catalytic element is complex and transfer of electrons but instead relies on the absorption of light to

* Corresponding author.
E-mail address: [Link]@[Link] (R. Chandrawati).

[Link]
Received 25 July 2023; Received in revised form 29 August 2023; Accepted 13 September 2023
Available online 18 September 2023
2352-9407/© 2023 The Author(s). Published by Elsevier Ltd. This is an open access article under the CC BY license ([Link]
F. Mazur et al. Applied Materials Today 35 (2023) 101937

generate electron-hole pairs, usually via a semiconductor material [10]. research avenues will be given.
Finally, in coordination catalysis, binding between the reactant and the
catalyst takes place to improve the reaction rate [11]. 2. Photocatalytic polymeric materials
Of the various types of catalysts available, polymeric materials have
emerged as a type of catalyst, and are termed catalytic polymers. They Upon reviewing the literature, the aforementioned criterion nar­
can withstand harsh reaction conditions, are highly tuneable, can be rowed down polymeric catalysts to those which primarily rely on the
heterogenous or homogenous, and can be less toxic due to the lack of mechanism of action of photocatalysis. This means that some catalytic
metals. Polymers can act as catalysts by having certain functional groups polymers like graphitic carbon nitride, covalent organic frameworks,
incorporated within their chain to promote specific reactions. Addi­ and amine-rich polymers are consistently referenced within this review
tionally, they can provide a large surface area for reactant molecules to (Fig. 2). Consequently, a brief introduction of these most common
adsorb onto, and act as support materials for other catalytic substances polymeric materials will be included within the following section to
[12], however these latter applications are outside the scope of this familiarize the reader with each.
review.
Herein, the use of polymeric materials as catalysts in the last 4 years
2.1. Graphitic carbon nitride (g-C3N4)
in the field of medicine, environmental remediation, and energy gen­
eration will be evaluated. First, a brief overview of the most commonly
Graphitic carbon nitride (g-C3N4) is a synthetic polymer produced by
used polymeric materials discussed within this review will be outlined.
the thermal polymerization of precursors such as cyanamide, dicyan­
Then, the application of these catalytic polymeric materials in different
diamide, and melamine. In contrast to the well-known metal-based
fields will be evaluated. Specifically, cases where the polymers them­
photocatalyst, titanium dioxide (TiO2) [13], g-C3N4 emerged as a
selves act as a catalyst or synergistically perform catalysis in conjunction
promising metal-free, stable, and low-toxic alternative for photo­
with other compounds will be discussed. Finally, an outlook on potential
catalysis. g-C3N4 has been widely investigated for the generation of

Fig. 1. Catalyst classification. Catalysts can be divided into three main categories: (i) homogenous catalysts where the reaction mixture and catalyst are in the same
phase, (ii) heterogenous catalysts where the reaction mixture and catalysts are in different phases, and (iii) biocatalysts where natural proteins or nucleic acids
catalyze the reaction. The benefits and limitations of each are also outlined.

2
F. Mazur et al. Applied Materials Today 35 (2023) 101937

Fig. 2. Commonly researched photocatalytic polymers. Illustration of: (i) graphitic carbon nitride formation using different nitrogen containing precursor
subunits, (ii) covalent organic framework formation using knots (organic molecules with various reactive points to enable branching of polymeric backbone) and
linkers (organic molecules that connect knots), (iii) a triazine-based conjugated microporous polymer with a 3D network structure, and (iv) various natural (poly
(histidine), poly(arginine), chitosan) and synthetic (polyaniline, polypyrrole) polymeric amines.

clean fuel (H2) [14], carbon dioxide (CO2) reduction [15], and the including sensing [23], catalysis [24], energy storage [25],
degradation of organic pollutants [16]. It is a type of semiconductor photo-energy conversion [26], and others [27], mainly due to their or­
photocatalyst whose conduction (− 1.4 eV) and valence (1.3 eV) band dered structure, accessible surface area, porosity, thermal stability, and
positions promote the harvesting of visible light towards oxidation and tunable functionalities [28]. As with g-C3N4, they employ similar
reduction processes, which is ideal for photocatalytic applications. For mechanisms of action towards energy generation and pollutant degra­
energy applications, the catalytic process generally consists of light dation, i.e. light-triggered generation, migration, and separation of
harvesting, charge separation, and surface redox reactions on the pho­ charge carriers for surface redox reactions and ROS generation,
tocatalyst. On the other hand, their main mechanism of action for respectively [29]. They have effective light-harvesting capabilities due
pollutant degradation is via the light-triggered generation of reactive to π-conjugation and coherent band structures. Their light-harvesting
oxygen species (ROS) including hydroxyl radicals (⋅OH), superoxide capacity can be improved by modifying building blocks, the structure
anions (O−2 ), hydrogen peroxide (H2O2), and singlet oxygen (O2), using and interlayer interactions, as well as transmission distance, bestowing a
water as an endogenous source. Bulk g-C3N4 is limited by a low specific highly controllable design for photocatalytic applications [30]. COFs are
surface area, high photoexcited electron and hole recombination rate, as generally limited by their poor electrical conductivity and low stability
well as restricted light absorption in the visible range [17]. As such, in aqueous/organic conditions. As such, research has been conducted to
various strategies have been employed over the last decade to increase improve their performance via the addition of functional groups [31], by
photocatalytic efficiency. These include: (i) morphological control increasing their surface area and pore size [32], adjusting molecular
which aims to provide more active sites, expand the visible-light design [33], as well as the addition of active sites [34].
response range, and shorten the carrier diffusion path [18], (ii)
elemental doping with non-metal or metal atoms to generate interme­
diate band gaps and regulate the band structure of g-C3N4 [19], (iii) 2.3. Conjugated microporous polymers (CMPs)
heterojunction construction to improve the separation and transport
efficiency of electron-hole pairs [20], and (iv) nanomaterial composition Conjugated microporous polymers (CMPs) are the amorphous ana­
to enhance charge transfer kinetics and electrical conductivity [21]. logues of COFs [35]. CMPs are a 3D semiconducting polymer with rigid
aromatic groups linked together, forming π-conjugated microporous
networks [36]. The increase in research surrounding CMPs arises from
2.2. Covalent organic frameworks (COFs) the variety of building blocks available, which allows for specific func­
tionalities and structures [37]. The alternation between single and
Covalent organic frameworks (COFs) are a class of covalently linked double-/triple-bonds within the network gives CMPs unique electronic
porous polymer networks with high crystallinity synthesized via properties, and have therefore been used in a range of areas including
condensation reactions [22]. They are widely employed in various fields gas separation [38], chemical encapsulation [39], catalysis [40], sensing

3
F. Mazur et al. Applied Materials Today 35 (2023) 101937

[41], as well as energy storage [42]. 3.1. Medical-based applications

The use of metal catalysts for physiological applications is chal­


2.4. Polymeric amine
lenging as they can be a source of toxicity due to leaching, even when
immobilized [53]. Zeolite catalysts, though generally considered
Polymeric amine refers to a class of polymers which contain a high
biocompatible, are not naturally found in the body, and may stimulate
concentration of primary (R-NH2), secondary (RR’-NH), and/or tertiary
the immune system by promoting an inflammatory response [54]. The
(RR’R’’-N) amine groups in their backbone and/or side branches. These
use of polymeric materials as catalysts circumvents this as they are
include natural polymers like chitosan, polydopamine (PDA), and poly
organic molecules that are biocompatible and biodegradable. The
(amino acids), as well as synthetic polymers including polyaniline
following section will discuss various medical-based applications where
(PANI), polypyrrole (PPy), and polyethyleneimine (PEI) [43]. In most
polymers have been used for a catalytic process.
cases, the amine groups act as an absorption site to increase the
pollutant concentration near a photocatalyst’s microenvironment. This
3.1.1. Antibacterial
is achieved via their cationic [44] and anion-exchange [45] behavior,
With the increasing reliance on antibiotics in recent years, the
coordination/complexation sites [46], and hydrogen bonding [47],
resistance of bacteria towards them has almost tripled in the last two
among others [48,49]. In terms of inherent catalytic activity, not many
decades [55]. As such, alternative strategies for effective sterilization
polymeric amine possess it, however, they are usually used in
have been considered. Traditional methods include chlorination and UV
conjunction with other materials in photocatalytic applications. The
irradiation [56,57], however photocatalysis has emerged as a preferred
nitrogen of the amino groups have been shown to contribute to heter­
choice whereby the bactericidal effect arises from the catalytic genera­
ojunctions which promote charge separation and shift the
tion of ROS or heat [58]. A highly investigated polymeric material in this
photo-response towards the visible range [50]. In other words, the
regard involves g-C3N4. Despite their antibacterial properties without
introduction of these polymeric amine results in significantly enhanced
the addition of additives [59], several strategies have been employed to
visible-light driven photocatalytic activity due to improved separation
further increase their performance. Most commonly, these include the
of photoexcited charge carriers, wider light absorption range, and
addition of nanoparticles [60], silver-based substances [61], doping
increased reactant adsorption [51]. The mechanism of action involves
[62], different morphologies [63] and heterostructures [64], as well as
the generation of radicals which target organic pollutants via oxidative
the addition of amine-rich polymers like PEI [65] and PANI [66]. These
degradation reactions [52].
approaches have proven effective against Pseudomonas putida, Staphy­
lococcus aureus, and/or Escherichia coli, achieving high sterilization rates
3. Polymeric materials as catalysts in various fields of >80% within 0.3–2 h. In fact, some studies have reported bactericidal
properties towards chloramphenicol-resistant E. coli and
Regardless of the field, i.e. medical, environmental, or energy, the methicillin-resistant S. aureus (MRSA) of 100% within 30 min
mechanism of action for polymeric material to act as catalysts are (Fig. 4ai-ii) [67]. This was achieved by the synergy between photo­
similar. It often involves light harvesting to photocatalytically generate catalysis and nanozyme catalytic activity. For the former, doping g-C3N4
heat, reactive oxygen species (ROS), and/or surface redox reactions with sulfur (g-SCN) leads to a broadening of the absorption spectrum
(Fig. 3). The following section will outline the use of these catalysts in which promotes photocarrier separation, thus increased photocatalytic
more detail.

Fig. 3. Mechanism of actions of polymer catalysts for various applications. The use of polymer catalysts in medical-, environmental-, or energy-based appli­
cations involves the use of light to trigger the photocatalytic generation of heat, ROS, and/or redox surface reactions. In most cases, medical-based applications such
as antimicrobial applications rely on heat and/or ROS generation, environmental-based applications such as pollutant degradation rely on ROS generation, while
energy-based applications such as energy generation rely on surface redox reactions.

4
F. Mazur et al. Applied Materials Today 35 (2023) 101937

Fig. 4. Catalytic polymeric materials (g-C3N4) for antibacterial applications. a) Relative survival rate of (i) chloramphenicol-resistant Escherichia coli and (ii)
methicillin-resistant Staphylococcus aureus under different conditions with and without light. Photographs of bacterial colonies for (iii) chloramphenicol-resistant
Escherichia coli and (iv) methicillin-resistant Staphylococcus aureus. (I) Control, (II) H2O2, (III) g-SCN, (IV) g-SCN + H2O2, (V) control + light, (VI) H2O2 + light,
(VII) g-SCN + light, and (VIII) g-SCN + H2O2 + light. b) Schematic illustration of generation of ⋅OH during H2O2 decomposition using g-SCN. Reprinted with
permission from [67]. Copyright 2023, Elsevier.

performance. For the latter, the sulfur element coordinates with nitro­ most popular approaches, where the generation of ROS for bacteria
gen in g-C3N4, endowing them with peroxidase-like catalytic activity. inactivation is improved by the inclusion of porphyrins [72], triazine
Collectively, this led to a greater generation of ⋅OH radicals which have [73], 2D layered transition metal carbides or carbonitrides [74], as well
high bactericidal properties (Fig. 4b). This was also visually observed as non-metals [75]. The antibacterial effect for all these platforms to­
from the bacterial culture, where no growth of bacterial colonies were wards Pseudomonas aeruginosa, Enterococcus faecalis, S. aureus, and/or
observed when using g-SCN + H2O2 + light (Fig. 4aiii-iv). Furthermore, E. coli was observed upon irradiation with white light, achieving ster­
the increased specific surface area of g-SCN also contributed to more ilization efficiencies of >90% within 1–4 h. Recent studies have even
effective trapping of bacteria. Similar doped formats, like phosphorus involved synergistic bacterial inhibition by integrating photodynamic,
doping [68], have also achieved high rates of sterilization. photothermal, and peroxidase-like enzymatic activities [76]. To achieve
Antibacterial attempts employing COFs have also been developed this, a covalent organic framework (TAPP-BDP) with a unique
relying on various bactericidal mechanisms including photocatalytic donor-acceptor structure was synthesized consisting of a narrow energy
sterilization [69], cationic sterilization [70], and biomolecular com­ gap and extended π-conjugated skeleton for improved ROS generation
posites [71]. Photocatalytic antibacterial platforms are also one of the and enhanced absorption in the near infrared (NIR), respectively

5
F. Mazur et al. Applied Materials Today 35 (2023) 101937

Fig. 5. Catalytic polymeric materials for antibacterial applications. a) (i) Illustration of triple synergistic antibacterial effect of TAPP-BDP under irradiation with
a single NIR light. (ii) Wound healing images of mice under different conditions. (iii) Photographs of bacterial colonies from the homogenized tissue dispersions of the
infection sites of mice for each condition. (I) PBS, (II) PBS + NIR, (III) TAPP, (IV) TAPP + NIR, (V) BDP, (VI) BDP + NIR, (VII) TAPP-BDP, and (VIII) TAPP-BDP +
NIR. Reprinted with permission from [76]. Copyright 2022, American Chemical Society. b) (i) Antibacterial activity of red fluorescence polymers toward
ampicillin-resistant Escherichia coli TOP10 with and without irradiation of white light. (ii) Photographs of bacterial colonies of ampicillin-resistant Escherichia coli
TOP10 treated with red fluorescence polymers with and without irradiation of white light. (iii) Antibacterial activity of red fluorescence polymers toward
ampicillin-resistant Escherichia coli TOP10 under different intensities of light irradiation. (iv) Confocal laser scanning microscopy images of ampicillin-resistant
Escherichia coli TOP10 stained by PI (propidium iodide) and SYTO9 after treating with red fluorescence polymers with and without irradiation of white light. v)
Photographs of mice infected with Staphylococcus aureus within 12 days postinjection under treatment of PBS or the red fluorescent polymer. Reprinted with
permission from [77]. Copyright 2018, American Chemical Society.

6
F. Mazur et al. Applied Materials Today 35 (2023) 101937

(Fig. 5ai). The antibacterial properties of TAPP-BDP was demonstrated bacterial colonies were observed for the TAPP-BDP/NIR group
in vivo, by evaluating wound healing conditions at different time in­ (Fig. 5aiii). Alternatively, red fluorescence polymers bearing quaternary
tervals in mice (Fig. 5aii). After a 5-day treatment, redness and swelling ammonium groups can also be developed for bactericidal applications
was reduced for the NIR exposed TAPP-BDP treatment, with both the [77]. These polymers have shown to be effective towards Gram positive
wound area and infection degree being lower. On the 7th day, infection and negative bacteria as well as fungi, with 100% bactericidal effect
recovery and wound healing was superior compared to other treatments. towards ampicillin-resistant E. coli TOP10 (Fig. 5bi-ii). The antimicrobe
Furthermore, wound skin tissues were harvested and minimal number of activity arises from “light” and “dark” toxicity. For the former, the

Fig. 6. Catalytic polymeric materials for wound healing applications. a) (i) Illustration of synthesis of chitosan-based C3N4-PDA-Ag composite dressing (C3N4-
PDA-Ag@CS). Antimicrobial rate towards (ii) Staphylococcus aureus and (iii) Pseudomonas aeruginosa after 10 min of irradiation. (iv) Wound healing photographs of
the control, C0, C3, and C3 + hv groups. Reprinted with permission from [86]. Copyright 2023, Elsevier. b) (i) Schematic diagram of CTCS hydrogel synthesis. (ii)
Schematic diagram of the photocatalysis mechanism of CuTCPP-Cur nanoparticle. (iii) Wound healing photographs for the control, 3M wound dressing, and CTCS
hydrogel groups after 0, 2, 5, and 10 days. Reprinted with permission from [88]. Copyright 2023, Elsevier.

7
F. Mazur et al. Applied Materials Today 35 (2023) 101937

backbone structure of the polymer consists of fluorene, benzene, thio­ E. coli and has shown in vivo wound healing and tissue regeneration via
phene, and benzothiadiazole units, which can produce ROS upon light the change in expression of inflammatory factors (Fig. 6biii). It is also
irradiation (Fig. 5biii-iv). The “dark” toxicity arises from the quaternary possible to load COFs with drugs which compound the antibacterial
ammonium groups in the side chains, which sensitize oxygen molecules benefits [89]. Recently, a biodegradable porphyrin-based COF nano­
to generate ROS. Moreover, their applicability for wound healing and sheet was synthesized which synergized photodynamic and photo­
disinfection in vivo without causing damage to tissue was demonstrated, thermal therapy under red light irradiation [90]. The red light also
providing efficient antimicrobial and clinical pathogenic treatment triggered the release of nitric oxide (NO; a known antibacterial agent)
(Fig. 5bv). for gaseous therapy. The developed platform could efficiently generate
ROS and high temperatures, release NO under laser irradiation, exhibit
3.1.2. Wound healing antibacterial properties against E. coli and S. aureus, inhibit inflamma­
There are two main reasons for delayed wound healing: (i) excessive tory cytokines, and promote wound healing in mice, all while being
inflammatory response and/or (ii) bacterial presence which make the biodegradable and biocompatible.
environment unsuitable for cell proliferation [78]. As bacteria inhibition
is key, the previously described antibacterial properties of catalytic 3.1.3. Antiviral
polymeric materials are closely linked with enhanced wound healing. A The tuneable chemical structure and composition of polymers com­
recently investigated approach involves light-to-heat conversion from pounded with their high molecular weight and buffering effect, endow
polydopamine (PDA). PDA has NIR response photothermal catalytic them with unique antiviral capabilities. Most commonly, they have been
characteristics which provide antibacterial properties, i.e. the high used for this application as drug carriers [91], lipid-membrane dis­
temperatures generated by NIR accelerate cell death. This has been rupting agents [92], coatings [93], and ion-releasing platforms [94],
demonstrated with high sterilization efficiencies (>95%) for E. coli [79], among others [95–97]. The antiviral effect for these systems is carried
S. aureus [80], and methicillin-resistant S. aureus [81]. It should be noted out via two different approaches. The polymer is used as a stabilizing
that for these cases, PDA is usually complexed with bactericidal sub­ matrix which protects and delivers an antiviral agent, or the polymer
stances, whereby the synergy between both leads to an enhanced anti­ itself is used as the antiviral agent which binds to the virus surface and
bacterial effect. Nevertheless, studies have also been shown to achieve inhibits infection. Polymers can also be used for antiviral catalytic ap­
high sterilization rates (>96% within 10 min) towards E. coli and plications; however, these are not as common. This is likely due to the
S. aureus when employing only PDA nanoparticles within food gum inactivation behavior of ROS which have been shown to be
matrices [82]. In this approach the role of PDA is twofold. Firstly, it microorganism-dependent [98]. As with bacteria, the inactivation
interacts with the food gum matrix to form a hydrogel with excellent mechanism can proceed via three main routes: (i) physical damage to
mechanical strength and high elasticity, enabling it to adapt to the the virus, (ii) photocatalytic release of metal ions inducing metal ion
wound shape. Secondly, it acts as a photothermal transduction agent to toxicity, and (iii) chemical oxidation via ROS to degrade the cell wall
convert NIR laser radiation to heat, triggering bacterial death. This [99]. Chemical oxidation is the most widely used approach due to its
hydrogel demonstrated accelerated wound healing in vivo by reducing high efficiency. With the advent of COVID-19, face masks have become
the inflammatory response and promoting vascular reconstruction. As the recommended prophylactic measure to contain the virus trans­
expected, g-C3N4 has also been extensively investigated for wound mission, with several reports attempting to improve upon their intended
healing applications [83], with various examples described within the use by increasing their re-usability [100], providing therapeutic benefit
previous section, and have shown highly efficient (>99% within 10–15 [101], and of relevance to this review, adding polymeric-based antiviral
min) sterilization of bacteria with successful in vivo wound healing [84, coatings [102]. For instance, COFs loaded with silver nanoparticles
85]. In a recent approach, a carbon nitride-PDA-silver (C3N4-PDA-Ag) (AgNPs) can be deposited onto the outer layer of non-woven PET fabric
complex compounded with chitosan (CS) was developed to achieve a face masks endowing them with solar-powered antibacterial and anti­
highly effective antimicrobial effect (C3N4-PDA-Ag@CS; Fig. 6ai) [86]. viral properties [103]. This is achieved via photothermal (PTT) and
In this design, PDA has two functions: (i) it enhances the photocatalytic photodynamic therapy (PDT), which can induce hyperthermia and
activity of C3N4 through π-π* interactions and (ii) it reduces Ag+ to generate ROS under light irradiation, respectively, leading to microbial
AgNPs. Thus, a synergistic antibacterial effect was achieved with the damage. Moreover, AgNPs are known to act against bacteria, viruses,
enhanced photocatalyst and the metal nanoparticles. To assess the and fungi by interacting with their cell membranes, whilst enhancing
antibacterial activity towards S. aureus and Pseudomonas aeruginosa, the PTT and PDT due to their inherent plasmonic nature. Virucidal perfor­
OD measurement of the bacterial solution was used. It was observed that mance was also observed with various metal-free photocatalytic-based
the OD did not differ significantly between the control and CS only (C0). graphitic carbon nitride platforms [104]. By oxygen doping g-C3N4 and
However, as the concentration of the C3N4-PDA-Ag complex increased integrating it with hydrothermal carbonation carbon, human adenovirus
(C1, C2, C3), the absorbance value decreased, suggesting high antibac­ type 2 can be inactivated [105]. This is due to the strong visible light
terial activity. The optimised complex (C3N4-PDA-Ag@CS) was assessed absorption and narrow band gap of the complex, which enhances charge
against S. aureus and Pseudomonas aeruginosa and showed high inhibi­ separation and enables the production of ROS as strong antiviral agents.
tion (>80%) within 10 min (Fig. 6aii-iii). More importantly, it also As such, complete antiviral effects within 2 h in drinking water can be
promoted wound healing in infected mice by facilitating collagen achieved. Other similar efficient strategies relying on ROS generation
deposition and accelerating epidermal regeneration (Fig. 6aiv). As with have also been reported with rapid (<2 h) and effective virucidal ca­
g-C3N4, COFs have also been extensively applied for wound healing pabilities [106].
applications [87]. For instance, curcumin (Cur) can be linked to tet­
ra-(4-carboxyphenyl) porphyrins (TCPP) and then combined with so­ 3.1.4. Cancer therapy
dium alginate (SA) and Cu2+ to form a hydrogel with photocatalytic and Cancer is one of the leading causes of death worldwide and various
anti-inflammatory properties (CTCS; Fig. 6bi) [88]. The mechanism strategies have been devised to tackle this issue, the most common of
behind it involves the electron-rich Cur and the photosensitizer TCPP which include chemotherapy, surgery, and radiotherapy [107]. Ad­
forming a π-conjugated donor-acceptor COF-like structure. Under irra­ vances in the STEM fields have led to the development of alternative
diation, electrons easily transfer from the Cur to the TCPP, inhibiting the approaches, including PDT and PTT. The former relies on the generation
complexation of the photogenerated electron-hole pair. Introducing of ROS to cause oxidative damage to cellular components while the
Cu2+ reduces the band gap and promotes charge separation, thus latter involves heat generation to ablate the tumor.
improving the photocatalytic effect (Fig. 6bii). The hydrogel can achieve Semiconductor-based polymers such as COFs are ideal candidates for
high sterilization rates (>98% within 20 min) against S. aureus and these emerging strategies. Although commonly used for chemotherapy

8
F. Mazur et al. Applied Materials Today 35 (2023) 101937

drug delivery applications via pH- [108], redox- [109], and cations [128] have also been investigated to enhance their therapeutic
GSH-responsive platforms [110], reports have also shown the polymers benefit. As with COFs, the development of multifunctional nanoplat­
themselves exhibiting tumor inhibiting capabilities via PDT [111,112] forms is becoming more common [129]. For instance, ultrathin graphitic
or PTT [113,114] strategies. Moreover, some of these photocatalytic carbon nitride nanosheets loaded with a hypoxia-targeting platinum(IV)
approaches have shown efficient anticancer properties whilst allowing prodrug can be developed which generate ROS in addition to DNA
for imaging applications [115,116]. For instance, a dye-labelled oligo­ binding platinum species to enhance antiproliferation of regulatory
nucleotide can be integrated onto porphyrin-based COF nanoparticles transcription factors [130]. Even further, g-C3N4 QDs can be used for
(COF-survivin) for effective cancer diagnosis and therapy [117]. In this imaging and chemo-photodynamic combination therapy [131]. This is
case, the dye is quenched by the COF due to FRET (fluorescence reso­ achieved by first introducing carboxyamino-terminated oligomeric
nance energy transfer), but in the presence of the cancer biomarker polyethylene glycol and folic acid onto the surface of the g-C3N4 QDs,
survivin mRNA, the fluorescence signal is recovered (Fig. 7ai). Upon endowing them with physiological stability, blue fluorescence,
irradiation with NIR light, enhanced ROS generation is observed due to biocompatibility, and ROS generating capabilities. Non-covalently
the crystalline reticular structure of the COF. The performance of this loaded doxorubicin enables its chemotherapy function. Compared
platform was demonstrated in vivo, allowing for selective cancer imaging with free doxorubicin, g-C3N4 QDs demonstrate more efficient chemo­
and elimination of the malignant tumor (Fig. 7aii-iii). In essence, these therapy to HeLa cells. This is due to improved folate receptor-mediated
polymeric catalysts allow for the integration of diagnosis and treatment. cellular uptake and intracellular pH-triggered drug release. In addition,
In fact, this “dual-purpose” approach also inspired other synergistic the complex shows enhanced inhibition of cancer cell growth compared
antitumour strategies. By combining different therapies, such as to sole chemotherapy or PDT.
chemotherapy with PDT/PTT [118,119] or PDT with PTT [120,121], an
enhanced therapeutic effect can be achieved. For instance, a non­
porphyrin containing COF nanoparticle can be developed with inherent 3.2. Environmental-based applications
PDT capabilities and be further conjugated with CuSe, a photothermal
agent, to form a dual functional COF with synergistic phototherapeutic The increase in standard of living, population, and industrialization,
effects in vivo (Fig. 7bi) [122]. This platform shows enhanced thera­ has led to significant environmental pollution, with severe negative
peutic effect in killing cancer cells and inhibiting tumor growth impacts on human health, wildlife, and natural resources. These effects
compared to its pristine counterpart (Fig. 7bii). Unexpectedly, g-C3N4 are compounded if the pollutants accumulate in the environment,
has not been as thoroughly investigated compared to COFs for anti­ potentially leading to permanent damages. Most of these pollutants are
cancer applications [123], with most studies focusing on g-C3N4 in the anthropogenic, including toxic organic dyes, agricultural and pharma­
form of NPs [124], QDs [125], or nanosheets [126] for PDT and PTT ceutical wastage, heavy metals, as well as toxic gasses (CO, SOx, NOx)
applications. The addition of elements such as AuNPs [127] or metal [132–137]. With the current demand for energy efficient systems, a shift
has occurred towards the development of techniques that can degrade

Fig. 7. Catalytic polymeric materials for cancer therapy applications. a) (i) In vivo fluorescence imaging of tumor-bearing nude mice after injected with COF-
survivin. (ii) Relative tumor volume after different treatments (PBS, PBS+laser, COF-survivin, COF-survivin+laser). (iii) Survival curves of tumor-bearing mice after
different treatments (PBS, PBS+laser, COF-survivin, COF-survivin+laser). Reprinted with permission from [117]. Copyright 2020, Royal Society of Chemistry. b) (i)
Schematic illustration of the synthesis of COF–CuSe nanoparticles for combined PTT and PDT. (ii) Photographs of tumor-bearing mice and (ii) tumors dissected from
mice. I: control; ii: PDT; iii: PTT; iv: COF; v: COF–CuSe; vi: COF + PDT; vii: COF + PTT; viii: COF–CuSe + PDT; ix: COF–CuSe + PTT; x: COF–CuSe + PDT + PTT.
Reprinted with permission from [122]. Copyright 2019, American Chemical Society.

9
F. Mazur et al. Applied Materials Today 35 (2023) 101937

these compounds while consuming little energy. Polymeric catalysts various organic dye pollutants, most commonly methylene blue [149],
have been used in this regard for a range of different pollutants including methylene orange [150], and rhodamine B [151]. In all cases, the
flame retardants [138], plasticizers (BPA) [139], sulfur-rich compounds degradation mechanism is the same, whereby ROS are
[140], polymers found in oilfield sewage [141], hydroxyl‑containing photo-catalytically generated to degrade the pollutant. Instead, the
pollutants [142], and nitrobenzene [143]. However, the most variable that researchers investigate is the ability to enhance the
researched pollutants involve organic dyes, pharmaceutical drugs, and degradation efficiency by modifying g-C3N4 via the addition of nano­
toxic gasses. As such, the following section will focus on strategies particles [152], polymers like polydopamine (PDA) [153], doping
employed to degrade these latter contaminants. [154], as well as morphological changes [155]. For instance, highly
condensed few-layered nanosheets of g-C3N4 (CM-CN) can be developed
3.2.1. Organic dye degradation by mixing cyanuric acid and melamine in water and calcinating the
The annual worldwide production of synthetic dyes is estimated at mixture. The nanosheet structure leads to a higher charge mobility due
70 million tonnes and is a major contributing factor towards water to the reduced charge diffusion length within the smaller and thinner
contamination [144]. These dyes are categorized based on their struc­ nanosheets compared to pristine g-C3N4 (M-CN; Fig. 9ai). Furthermore,
ture and application, ranging from cosmetics, leather, wool, and silk, to the reduced arc diameter of the Nyquist plot for the nanosheet structure
medicine, ink, and paper. Regardless of their category, they can act as depicts a decrease in charge transfer at the catalyst interface when the
ecotoxicological threats with negative effects on living organisms [145]. charge carriers transfer from the surface to the electrolyte (Fig. 9aii).
This is because these toxic chemicals that are discharged as effluent into These photoelectrochemical characteristics compounded with the high
wastewater remain in water and soil over extended periods of time, surface area and pore volume, lead to complete degradation of rhoda­
resulting in reduced soil fertility and photosynthetic activity of plants as mine B within 15 min, even after five successive runs, as well as tetra­
well as increased biochemical and chemical oxygen demand. Moreover, cycline within 90 min (Fig. 9aiii-iv) [156]. Alternatively, a g-C3N4/TiO2
they can enter the food chain, leading to bioaccumulation and thus nanocomposite can be synthesized which show enhanced catalytic
promoting toxicity, mutagenicity, and carcinogenicity [146]. Tradi­ degradation towards methylene blue compared to its pure counterparts,
tional purification systems like flocculation or electrochemical treat­ i.e. TiO2 or g-C3N4. However, the addition of boron nanoparticles
ment produce harmful by-products which contribute to secondary further improves this degradation. This is attributed to the electron
pollution [147] or high energy consumption [148], respectively. As transfer from the composite to the elemental boron due to the hetero­
such, catalytic polymeric materials have been considered as attractive junction between them, achieving efficiencies of >80% within 3 h
alternatives for this application, specifically the advanced oxidation [157].
technology: photocatalysis. Three techniques are commonly used, Similarly, COF-based platforms also rely on ROS generation for the
among others, to evaluate the performance of these systems including: degradation of dye pollutants including methylene blue [158], methy­
(a) the photocurrent density, (b) the Nyquist plot, and (c) the catalytic lene orange [159], rhodamine B [160], orange II [161], and various
activity (Fig. 8). The former refers to the rate of flow of electrical current nitroaromatic compounds [162], among others [163,164]. Recently,
produced by the photoexcited charge carriers under visible light illu­ imine-linked COFs have been shown to exhibit efficient photo­
mination. A higher photocurrent density indicates better light absorp­ degradation of azo dyes, whereby the N = N bonds are broken by su­
tion, efficient charge separation, and enhanced photoelectrochemical peroxide radicals within 1.5 h [165]. Moreover, the catalyst can be
activity. In other words, the catalyst’s ability to convert light energy into recovered from the dye solution and used at least three times with
useful chemical reactions, such as water splitting or pollutant degrada­ minimal loss in activity. This performance is attributed to their densely
tion (Fig. 8a). The Nyquist plot, also known as electrochemical imped­ conjugated structure, large absorbance range, excellent dispersibility in
ance spectroscopy (EIS), provides information about charge transfer water, and high surface area, which allows for ROS generation towards
resistance, surface recombination rates, and other kinetic processes dye degradation. The catalytic degradation was demonstrated for
occurring at the electrode-electrolyte interface. A well-defined Nyquist mordant black 3, mordant black 17, mordant red 7, eriochrome black A,
plot with a semicircular arc indicates efficient charge transfer, while and eriochrome black T, with efficiencies ranging from 40 to 90%
deviations from this behavior might suggest limitations in charge sep­ depending on the dye. Of greater interest is the ability of these polymeric
aration and transfer efficiency. In addition, a smaller arc radii implies a catalysts to perform multiple functions, i.e. degrade different types of
higher electron transfer efficiency and a smaller interfacial resistance pollutants at the same time [166,167]. An example involves a AgI
(Fig. 8b). Finally, the catalytic activity of the polymer is evaluated by modified COF (COF-PD/AgI) which was fabricated to photocatalytically
measuring the decomposition rate by monitoring the change in the degrade bacteria, organic pollutants, as well as antipyretic drugs, spe­
contaminant concentration over time under visible light irradiation cifically E. coli, rhodamine B, and acetaminophen, respectively [168].
(Fig. 8c). For the former, a negligible decrease in viable E. coli was observed when
Graphitic carbon nitride has been applied for the degradation of exposed to either COF-PD or a physical mixture of COF-PD and AgI with

Fig. 8. Performance indicators for polymeric catalysts in pollutant degradation. Representative (a) transient photocurrent response, (b) electrochemical
impedance spectroscopy plot (Nyquist plot), and (c) photodegradation kinetics of an unmodified polymeric catalyst (e.g. pristine g-C3N4) and a modified one (e.g. g-
C3N4 nanosheets).

10
F. Mazur et al. Applied Materials Today 35 (2023) 101937

Fig. 9. Catalytic polymeric materials for organic dye degradation. a) (i) Transient photocurrent response and (ii) electrochemical impedance spectroscopy plot
of pristine g-C3N4 (M-CN) and graphitic carbon nitride nanosheets (CM-CN). Photolysis and photodegradation kinetics for (iii) rhodamine B and (iv) tetracycline.
Reprinted with permission from [156]. Copyright 2020, American Chemical Society. b) Photocatalytic performance of AgI modified COF towards (i) Escherichia coli
disinfection, (ii) rhodamine B degradation, and (iii) acetaminophen degradation. Reprinted with permission from [168]. Copyright 2020, Elsevier. c) (i) Absorption
spectra, (ii) change in concentration vs. irradiation time, and (iii) comparison of degradation rate of different nanocomposite photocatalysts (ZnTiO3, 10%PAN­
I/ZnTiO3, and 1%Ag/10%PANI/ZnTiO3) for the degradation of methylene blue. Reprinted with permission from [185]. Copyright 2021, Elsevier.

visible light irradiation (Fig. 9bi). On the other hand, complete disin­ (PPS) [173] and polythiophene (PTh) [174] have also been evaluated.
fection was achieved by COF-PD/AgI within 40 min with visible light They have been used for the degradation of methylene orange [175],
irradiation, indicating a greatly enhanced bactericidal effect. It should methylene blue [176], rhodamine B, other dyes [177,178], as well as
be noted that due to the small amount of silver ions released from other pollutants like PVC [179] and phenol [180]. Although amine-rich
COF-PD/AgI, they cannot be considered the main contributor to the polymers show effective catalytic activity by themselves [181], many
photocatalytic disinfection [169]. Instead, the reactive species (⋅O−2 , h+, reports combine them with other semiconductor materials to enhance
and ⋅OH) generated during the photocatalytic process are responsible for their performance. The most popular involves TiO2 [182], however
the photocatalytic disinfection. For rhodamine B, a catalytical degra­ other materials like SrSnO3 [183] and MoSe2 [184] have also been re­
dation of 84.2% within 80 min was observed when using a mixture of ported. As with g-C3N4 and COFs, the degradation mechanism relies on
the unmodified COF and AgI, with most of the catalytic degradation photogenerated ROS to break down C-C and C-H bonds of organic pol­
attributed to AgI. However, the performance was increased by the het­ lutants. For instance, ZnTiO3 nanostructures which already possess
erojunction formed between the COF and AgI, which enhanced the catalytic properties can be doped with PANI and Ag nanoparticles via a
separation of photogenerated h+ and e− , with complete removal within modified sol-gel method to enhance their activity towards environ­
80 min (Fig. 9bii). On the other hand, 96.3% degradation of acetamin­ mental remediation applications, specifically methylene blue degrada­
ophen was observed when using the unmodified COF, which increased tion [185]. When observing the change in absorption spectra for the MB
to 97.8% when using a mixture of the unmodified COF and AgI, and was dye molecule in the presence of Ag/PANI/ZnTiO3 under visible light
completely removed when employing COF/AgI after 160 min illumination, a decrease in the peaks can be seen, with a flattening of the
(Fig. 9biii). Although less obvious, the data indicates that the modifi­ curve within 25 min (Fig. 9ci). This highlights the effective and
cation of COF-PD with AgI can also enhance the photocatalytic degra­ target-oriented performance of the photocatalyst. It is important to note
dation of ACTP. that pristine ZnTiO3 showed reduced catalytic performance compared to
Amine-rich polymers have also been extensively used in organic dye the hybrid frameworks, whereby 75.8% of the dye was degraded under
degradation applications. These commonly include PANI [170], PDA bare ZnTiO3, 81.7% in the presence of 10%PANI/ZnTiO3, and 95.6%
[171], and PPy [172], however others like poly(phenylene sulfide) when employing the 1%Ag/10%PANI/ZnTiO3 hybrid nanostructure

11
F. Mazur et al. Applied Materials Today 35 (2023) 101937

(Fig. 9cii). The increased photodegradation performance was attributed respectively [186]. The incorporation of L-tyrosine in both morphologies
to the heterojunction formed between Ag, PANI, and ZnTiO3, which results in a decrease in band gap as well as additional absorption edges
suppressed the high energy requirements for electron-hole pair recom­ at wavelengths higher than 400 nm, allowing for more light absorption
bination. This leads to the formation of radicals (⋅O−2 and ⋅OH) which are in the visible range. As a result, a 2.2x increase in photocatalytic activity
highly capable of pollutant degradation. Moreover, the surface plasmon is observed compared to the L-tyrosine free titania.
resonance phenomenon of AgNPs in the nanostructure led to the
lowering of the band gap energy for the hybrid structures. Ag-based 3.2.2. Pharmaceutical drug degradation
plasmon resonance could capture many visible light photons to Pharmaceutical drugs have recently been categorized as an emerging
redshift the light absorbance, thus enhancing degradation. In combi­ pollutant due to their high consumption, biorefractory nature, and
nation, these characteristics lead to greatly enhanced photodegradation widespread distribution in the environment [187]. Their unrestricted
rates (Fig. 9ciii). Alternatively, amino acid-based hybrid organotitanias use causes them to be a key contributor towards pollution of surface
can be formed by in situ incorporation of L-tyrosine during synthesis. water, wastewater, and soil, as well as promoting the emergence of
The crystalline structure, shape, and size can be tuned by varying the antibiotic-resistant bacterial strains [188]. Their stable chemical struc­
pH, resulting in nanoparticles or nanorods at a pH of 2.2 or 0, ture makes them difficult to remove via conventional wastewater

Fig. 10. Catalytic polymeric materials for pharmaceutical drug degradation. a) (i) Schematic illustration of synthesis of g-C3N4@ZIF-8. (ii) Photocatalytic
degradation of tetracycline, rhodamine B, and Cr(VI) over g-C3N4@ZIF-8. Reprinted with permission from [215]. Copyright 2021, Elsevier. b) Photocatalytic
degradation of (i) ciprofloxacin and (ii) rhodamine B over a POM/g-C3N4 complex. Reprinted with permission from [217]. Copyright 2022, Elsevier. c) (i) Schematic
illustration of the synthesis of the PoPD/AgCl/CN photocatalyst. (ii) Transient photocurrent response of CN, AgCl-35/CN, and PoPD/AgCl-35/CN. (iii) Photocatalytic
degradation of tetracycline over AgCl, CN, AgCl-35/CN, PoPD/AgCl-35/CN. Reprinted with permission from [221]. Copyright 2019, Elsevier.

12
F. Mazur et al. Applied Materials Today 35 (2023) 101937

treatment options and biodegradation [189], therefore, the residues of Several studies have also integrated conductive polymers with
the incomplete absorption and metabolism of pharmaceutical drugs can semiconductors like g-C3N4 or TiO2, enhancing the electrical conduc­
be released into the environment and cause secondary pollution. It is tivity, cycle utilization, and electronic transmission rate [220]. In other
then imperative to develop strategies which can address the impact they words, coupling conductive polymers with semiconductors is an effec­
have on human health and the environment. Although some tive way to increase photocatalytic activity. For instance, graphitic
polymeric-based platforms have been developed which can degrade nanosheets (CN) inherently show improved performance due to their
drugs including anti-inflammatory [190], anti-convulsant [191], increased specific surface area and reaction sites. However, their cata­
anti-depressant drugs [192], as well as pharmaceutical preservatives lytic activity can be increased by the incorporation of
[193], the most researched pharmaceutical drugs involve antibiotics poly-o-phenylenediamine (PoPD) and AgCl, allowing for 3x the photo­
[194]. As such, this section will focus on their degradation. current density compared to the pure nanosheet (Fig. 10c) [221]. The
The polymeric platform that has been leading the field in this regard PoPD/AgCl/CN hybrid exhibits the highest photocurrent density, indi­
is graphitic carbon nitride, primarily relying on the photocatalytic cating effective transfer of photogenerated electron-hole pairs
generation of ROS to degrade them [195]. Tetracycline is the most (Fig. 10cii), and a smaller radius arc in the impedance curve, indicating
investigated antibiotic since it is a typical broad-spectrum antibiotic lower resistance to charge transfer during the reaction. In combination,
used to prevent human and animal infections and is commonly found in these properties can explain the improved photocatalytic activity of the
various water bodies [196]. However, the degradation of other antibi­ complex, achieving a degradation rate of 74.7% towards tetracycline
otics like ciprofloxacin [197], berberine hydrochloride [198], and within 30 min for AgCl-CN and 83.1% for PoPD/AgCl/CN (Fig. 10ciii).
ofloxacin [199], among others [200], have also been investigated which Similar catalytic performances were shown for other platforms
rely on similar ROS-generating mechanisms. As with many studies employing other semiconductors and amine-rich polymers [222,223]. In
involving this semiconductor-based polymeric catalyst, strategies have all cases, the degradation mechanism is attributed to the generation of
been employed which aim to increase their light-absorption capacity ROS which degrade the pollutant. Alternatively, nanofiber membranes
and thus increase performance. These include morphology changes can be loaded with tungsten oxide (WO3) nanoparticles surface coated
[201], heterojunction choice [202], the addition of quantum dots [202] with PDA, and tested for oil/water separation, photothermal water
or nanoparticles [203], carbon or nitrogen deficient complexes evaporation, as well as ampicillin degradation [224]. For the latter case,
[204–206], MOF-based composites [207], and various others the inherent photocatalytic activity of WO3 was enhanced by the addi­
[208–212]. It is promising to see however, the rise in reports which aim tion of a thin layer of PDA, generating radicals more efficiently and
to develop catalysts which can degrade various types of pollutants allowing for the degradation of the pollutant. Although this catalyst
simultaneously, rather than just one [213]. This is because in many showed high performance towards oil/water separation and water
cases, pollutants are synchronous and it is therefore beneficial if they evaporation, only 40% ampicillin was reduced after 3 h. However, the
can be removed simultaneously [214]. For instance, MOF-based nano­ platform could be improved by fine-tuning the absorption wavelength of
reactors with photocatalyst inclusion (e.g., g-C3N4) can lead to improved the nanocomposite, increasing the loading and exposure of WO3 on the
performance by enhancing areas of heterojunctions, increasing the surface, optimizing the initial concentration of the ampicillin, adjusting
amount of active sites, and confining active species within a limited the pH, and removing the interfering radical scavengers.
porous space [214]. A core-shelled g-C3N4@ZIF-8 photocatalysts with
Zn-exposure can be designed with not only enhanced tetracycline 3.2.3. Toxic gas removal
degradation compared to pure g-C3N4 (4.8 times), but also superior The relationship between increased standard of living and pollution
degradation rate of rhodamine B (99.3%) and Cr(VI) (96.6%) is complex and multivariable. On the one hand, it leads to an increased
(Fig. 10ai-ii) [215]. The photocatalytic activity is due to the synergistic awareness of environmental issues and willingness to reduce pollution
effect that is produced when combining g-C3N4 and ZIF-8, resulting in an and invest in “cleaner” alternatives [225]. On the other hand, it results
enhanced transient photocurrent response and a decreased electro­ in increased industrialization, urbanization, and consumption of goods
chemical impedance spectra than either single material. In other words, and services, in other words, more energy consumed, waste generated,
the hybrid material exhibits improved carriers’ separation and trans­ and pollutants emitted into the environment. An example of this is the
portation. Similarly, a synergy can be achieved when decorating g-C3N4 increase in motor vehicles which release harmful gasses to the envi­
with polyoxometalates (POM), the latter of which have demonstrated ronment and cause severe air pollution [226]. One of these gasses
excellent photocatalytic properties [216]. By doing so, a POM/g-C3N4 include nitrogen oxides (NOx), with nitrogen oxide (NO) accounting for
(CPM-CN) complex can exhibit enhanced photocatalytic activity 90% of NOx in ambient air [227]. Several efficient de-NOx methods
compared to its pure counterparts. This was demonstrated for the including biological [228], chemical [229], photochemical [230],
degradation of antibiotics (tetracycline and ciprofloxacin) as well as electrochemical [231], and photocatalytic have been developed [232].
organic dyes (rhodamine B) (Fig. 10b) [217]. As with other platforms, The latter approach is currently considered the most attractive due to its
the increased photocatalytic performance was attributed to: (i) the cost-effective, simple, and environmentally friendly nature [233]. As
suppressed recombination of photoinduced charge carriers indicating an expected, polymer-based photocatalytic platforms have also been eval­
increased separation ability of electron-hole pairs, thus more effective uated to remove NO from air, with the most common examples utilizing
photocatalytic performance, (ii) a higher transient photocurrent g-C3N4 and COFs. The mechanism for photocatalytic degradation of NO
response compared to either single material, indicating a greater pho­ relies on the generation of ROS [234] and is commonly improved via
tocatalytic behavior as the separation rate of photo-generated elec­ various strategies such as using different stoichiometric ratios of C:N
tron-hole pairs was enhanced, and (iii) a smaller radius of the arc on the [235], the addition of natural minerals [236] or their oxides [237],
electrochemical impedance spectra suggesting an effective interfacial developing models with N-defects [238], as well as different types of
charge transfer and elevated separation ability of photogenerated heterojunctions [239,240]. For instance, a N-deficient mesoporous
electron-hole pairs. A modified carbon nitride framework with a 3:5 C:N carbon nitride can be synthesized with a greater number of exposed
stoichiometry can also be developed with a reported narrower bandgap active sites, resulting in a wider band gap, thus enhanced redox ability
and higher potential of conductive band than that of g-C3N4, where the [241]. The N defects create catalytic active sites and induce effective
higher N ratio reduces adsorption energy and facilitates electron transfer separation of photogenerated carriers. This synergistic effect results in
[218]. This modified carbon nitride can be used to activate perox­ the generation of more ROS, achieving 60.6% higher activity compared
ymonosulfate (oxidizing agent), which generates ROS and degrades to its pure counterpart (g-C3N4). Alternatively, K- and B-doped platforms
various micropollutants including ciprofloxacin, tetracycline, sulfa­ have also been developed with the aim of improving the recombination
methoxazole, and diclofenac sodium [219]. of photogenerated carriers and low concentration of oxidizing species

13
F. Mazur et al. Applied Materials Today 35 (2023) 101937

[242]. B is introduced by replacing C atoms, while K is inserted into the increased optical responsivity. These paints showed a 50% decrement in
interlayer by binding to the N and C atoms of adjacent layers. This re­ benzene removal under UV light, which increased by 23% under VIS
sults in increased electron delocalization and expands the π-conjugate light. It is worth noting that although PANI itself was able to remove
system. As such, expanded light absorption and a reduced band gap can benzene under both UV and VIS light, the performance was decreased
be achieved. This leads to a system 2x more efficient compared to an comparatively.
un-doped variant which can be re-used even after 5 cycles and be used
for hydrogen evolution. Interestingly photocatalytic paints containing a 3.3. Energy-based applications
PANI/TiO2 composite have also been developed [243]. The band gap of
this composite is decreased compared to TiO2, whilst showing an More than two-thirds of worldwide energy demands are currently

Fig. 11. Catalytic polymeric materials for energy generation. a) (i) H2 generation rate of g-C3N4 (CN) and Ni3S2-NiS2/CN with different loading amount of Ni3S2-
NiS2. (ii) Photocurrent density of CN and Ni3S2-NiS2/CN-3, with the higher density implying a superior charge transfer efficiency. (iii) Electrochemical impedance
spectroscopy of CN and Ni3S2-NiS2/CN-3, with the smaller radius indicating faster electron transfer. Reprinted with permission from [266]. Copyright 2022, Elsevier.
b) (i) CO production rates of PCN, O-PCN, A-PCN, and OA-PCN depicting a ~17x increase for OA-PCN. (ii) Contact-angle measurements and (iii) electrochemical
impedance spectroscopy reinforcing the enhanced electron and mass transfer of the polymeric carbon nitride modified with hydroxyl and amino groups on its surface.
Reprinted with permission from [300]. Copyright 2019, American Chemical Society. c) (i) Illustration of the functionalization of C3N4 nanosheets via an electrostatic
assembly process. (ii) H2O2 production by PEI/C3N4 and C3N4. (iii) Photocurrent density of PEI/C3N4 and C3N4, with the higher density implying a superior charge
transfer efficiency. (iv) Electrochemical impedance spectroscopy of PEI/C3N4 and C3N4, with the smaller radius indicating faster electron transfer for the modified
polymer material. Reprinted with permission from [314]. Copyright 2020, American Chemical Society.

14
F. Mazur et al. Applied Materials Today 35 (2023) 101937

met by non-renewable energy sources [244], therefore, great efforts [Link]. O2 evolution. It is possible to alter the band structure of g-C3N4
have been placed to transition towards more sustainable energy alter­ to make it more effective towards O2 evolution reactions [277]. For
natives. Polymeric catalysts can play a significant role in addressing this instance, calcining it with sodium borohydride results in a boron-doped
issue, mainly via energy generation [245], conversion [246], and stor­ and nitrogen-deficient g-C3N4 [278]. This leads to the conduction and
age [247]. The following section will briefly discuss the role of poly­ valence band being tuned to enable an increase in optical absorption in
meric catalysts in these areas. visible light and driving force for water oxidation, resulting in a 6x in­
crease in O2 evolution compared to pristine carbon nitride [279].
3.3.1. Energy generation Similarly, a g-C3N4/bismuth vanadate composite can be doped with
sulfur to facilitate charge separation and enhance water oxidation [280].
[Link]. Hydrogen evolution. Hydrogen (H2) has emerged as an alter­ This is because sulfur doping narrows the band gap of g-C3N4 by
native fuel option to transition the energy industry away from fossil fuels changing the band structure via staking of 2p orbitals on the valence
and promote decarbonization. Currently, 85% of H2 is produced from band. The complex exhibits an oxygen evolution rate under visible light
fossil fuels, however solar-driven polymeric catalysts can be used to irradiation which is 129% greater compared to that of pristine bismuth.
generate H2 in a sustainable manner [248]. g-C3N4 is a prime polymeric Microporous conjugated polymers (CMPs) can also photoelectrochemi­
example that has been extensively evaluated for this application. cally generate O2 from water via doping. By using metal-free pyr­
Various strategies have been employed to increase the photocatalytic H2 ene-based nitrogen and sulfur containing CMPs, a low bandgap is
generation from g-C3N4. These include different types of heterojunctions achieved, which allows for highly efficient catalytic activity [281]. This
such as type-II [249], Z-scheme [250], S-scheme [251], and Schottky increased activity is attributed to the pyrene units in the π-conjugated
heterojunctions [252], metal [253] and non-metal [254] doping, the use polymer skeleton, the heteroatoms N and S, as well as the high specific
of cocatalysts [255], as well as morphology (0D, 1D, 2D, 3D) [256–260] surface area with inherent microporosity. Water splitting to produce
and composition changes [261,262]. In all cases, the primary goal is to hydrogen and oxygen simultaneously is difficult to achieve given that
overcome some of the limitations of pristine g-C3N4 and generate more the band-edges of photocatalysts must allow for both water reduction
H2 comparatively. These improvements can range from a 1–2x increase and oxidation for H2 and O2, respectively [282]. In other words, a
in H2 generation [263,264], to up to triple digit increases [265]. For catalyst with suitable band alignments is required which has a con­
instance, binary Ni3S2-NiS2 can be in-situ grown on g-C3N4 (CN) to duction band (CB) minimum more negative than the H+/H2 energy
achieve hydrogen evolution efficiencies 300x that of pure g-C3N4 level, and a valance band (VB) maximum more positive than the O2/H2O
(Fig. 11ai) [266]. In this case, the increase is attributed to the faster energy level [283]. A way of achieving this could be by introducing
photogenerated charge separation, as characterized by the photocurrent cobalt phosphate (Co3(PO4)2) within g-C3N4 nanosheets, which could
response curve (Fig. 11aii). The photocurrent density is higher for the allow for H2 and O2 evolution without the need for a cocatalyst. A
complex compared to pristine CN, implying superior charge transfer similar platform was recently demonstrated for a type-II heterojunction
efficiency. Furthermore, the electrochemical impedance spectra showed photocatalyst based on Co3(PO4)2 and reduced graphene oxide (rGO)
a smaller radius for Ni3S2-NiS2 indicating faster electron transfer [284]. In this case, CB or VB edges of rGO and Co3(PO4)2 had sufficient
(Fig. 11aiii). These properties, in conjunction with the increased light overpotential for water reduction or oxidation. Moreover, the formation
absorption capacity of the complex, explain the enhanced hydrogen of a type-II staggered band heterojunction with RGO promoted charge
evolution performance of the catalyst. COFs are also recognized as a collection and transport, resulting in increased catalytic activity.
promising crystalline porous organic material for photocatalytic H2
generation. As with g-C3N4, several strategies have been employed to [Link]. CO2 reduction. With the increasing concerns regarding green­
improve H2 evolution, mainly involving functional building blocks and house gas emissions, specifically CO2, many strategies have been
linkages [267,268], heterojunction choice [269], metal-doping [270], employed to address this issue, one of which involves carbon capture
and even by developing graphene-like COFs [271]. These improve their and utilization (CCU). CCU is the process by which CO2 emissions are
electron-deficiency, π-conjugation, or nitrogen/metal-containing cata­ used as a resource rather than released into the environment, thereby
lytic active sites. Furthermore, the type and dispersion of cocatalysts mitigating the greenhouse gas effect [285]. As with H2 and O2 evolution,
within COFs has also been evaluated to enhance H2 generation. For a highly researched polymeric catalyst involves g-C3N4, which can carry
instance, adsorption sites for platinum species can be constructed by out the catalytic reduction of CO2 to various hydrocarbons [286], most
introducing adjacent hydroxyl group and imine-N within a COF struc­ commonly CH3OH [287], CH4 [288], HCOOH [289], and CO [290].
tural unit, leading to the in situ reduction of adsorbed platinum species However, due to its poor ability to activate the C-O bond, strategies
into metal clusters [272]. The well-dispersed deposition allows for including doping [291], cocatalyst loading [292], different hetero­
optimal utilization of the Pt co-catalyst, leading to high hydrogen evo­ junctions [293,294], morphologies [295,296], and others [297,298],
lution rates. Other polymeric catalysts have also been investigated have been used to improve the catalytic performance [299]. Some
which aim to improve effective separation, transport, and tunneling of strategies have even relied on activating polymeric carbon nitride (PCN)
charge carriers. For example, the H2 generation of pure zinc sulfide by modifying its surface with hydroxyl (O) and amino (A) groups [300].
(ZnS) can be improved by coating with ultrathin polydopamine, without These superficial functional groups synergistically activate g-C3N4 for
compromising stability and activity [273]. The improved performance CO2 electroreduction, achieving CO production rates 17x greater
arises from the effective separation, transport, and tunneling of charge compared to the unmodified polymer (Fig. 11bi). In this platform, the
carriers of the complex. Moreover, the polydopamine layer improves the surface contact angle of O-PCN, A-PCN, and OA-PCN decreased
wettability of the catalyst while also providing a passivating layer. compared to that of PCN, suggesting the -OH and -NH2 functional groups
Additionally, other amine-rich polymers can be incorporated onto could reinforce water adsorption (Fig. 11bii). Moreover, the electro­
semiconductors to increase their performance like PANI, PPy, or PEDOT chemical impedance spectroscopy displayed a smaller radius for
[273–275]. For instance, the surface of rGO can be functionalized with OA-PCN compared to PCN, implying decreased charge transfer resis­
TiO2 and attached on polymerized aniline to obtain a catalyst with tance when functionalised (Fig. 11biii). Simultaneously, these proper­
efficient charge separation [276]. This is because it shows a narrower ties suggest hydroxyl and amino groups can promote electron and mass
bandgap energy, increased light absorption in the visible range, rapid transfer, thus enhancing catalytic performance. Similarly, COFs can also
photoinduced charge separation, and suppressed charge recombination. be used for CCU, specifically via photocatalytic [301], electrocatalytic
Consequently, a H2 evolution rate of 1.8x greater than TiO2/rGO and [302], or photo-electrocatalytic [303] CO2 reduction, to generate CO
TiO2 can be obtained. [304], HCHO [305], HCOOH [306], CH4 [307], as well as C1-C3

15
F. Mazur et al. Applied Materials Today 35 (2023) 101937

chemicals [303]. Most reports rely on hybrid materials, e.g. combining oxyethyl-N-methylamino-pyridine) as the shell, and be used in a one-pot
with metals [308] to generate higher yields. However, it is also possible acid/base-catalysed deacetalisation–Knoevenagel cascade reaction
to enhance their catalytic activity by increasing their available active [330]. The CSMs were produced via a two-step surfactant-free emulsion
sites via porous three-dimensional (3D) designs which have spatially polymerization, without resulting in mutual quenching. These CSMs
separated building blocks [309]. showed excellent performance in various solvents and recyclability (up
to 6 times), whilst also having high potential for scalability. The
[Link]. Other fuels. Alternative strategies for fuel generation have also developed colloidal cascade catalyst could be synthesised on a 25 g scale
been evaluated to transition away from fossil-fuels. Another promising and produce 10 g of the final product. In fact, some catalytic polymeric
liquid fuel involves hydrogen peroxide (H2O2), which has also been materials have shown 100% substrate conversion and yield of the final
extensively evaluated [310,311]. Due to the typical low yields of H2O2 product such as in the one-pot production of β-nitrostyrene from benz­
from water, proton donors can be added to increase catalytic perfor­ aldehyde dimethyl acetal under cascade reactions (acid-catalyzed
mance, and thus yield [312]. It has also been shown that boron-doped deacetalization and base-catalyzed Henry reactions) [331]. To achieve
g-C3N4 containing nitrogen vacancies can also efficiently catalyze this, microporous polymer supports are modified with ethylenediamine
H2O2 evolution from water [313]. This is attributed to the effect these followed by 1,3-propane sultone. This results in the simultaneous
variables have on the electronic structure of g-C3N4, which leads to immobilization of an acid and base catalyst on a single polymeric sup­
different dissociation barriers for the proton donor and promotes the port, providing a catalytic synergism. Moreover, the catalyst could easily
generation of different reactive oxygen species. Similarly, photocatalytic be recovered via filtration and be reused up to 4 times with approxi­
H2O2 production can be boosted by the incorporation of PEI (Fig. 11ci-ii) mately 10% loss in performance. Recently, it has also been demon­
[314]. Briefly, PEI can be grafted onto g-C3N4, thereby tuning the local strated that photocatalytic chromophores can be incorporated into
electronic environment. The interaction between PEI and g-C3N4 leads polymeric structural frameworks to allow for covalent heterogenization
to an increase in photocurrent density (Fig. 11ciii) as well as a smaller of a catalyst, specifically towards economical organic synthesis [332].
radius on the Nyquist plot (Fig. 11civ), achieving generation rates 25x This class of photoactive organic polymers combines the flexibility of
higher than pristine g-C3N4. small-molecule dyes with the recyclability of solid-phase catalysts. The
active sites within this porous microstructure display an increased
3.3.2. Energy storage reactivity, which is further supported by the mobility of excited states
Catalytic polymeric materials have mainly been applied in the field and charged species. This is important as it results in increased stability
of energy storage to improve power density and stability [315]. Early and catalytic efficiencies. In theory, this approach could allow for any
reports have focussed on lithium-ion batteries, by synergistically pro­ organic or organometallic photocatalyst to be heterogenized through
moting electron and ion transmission [316], preventing the formation of copolymerization with porosity-promoting bulky organic monomers. On
an unstable solid electrolyte interphase [317], improving structural the other hand, great efforts have also focussed on the development of
stability [318] and storage capacity [319], as well as reducing the catalysts with high potential in sustainable catalysis. CMPs are widely
activation energy of deposition [320]. More recently, efforts have been used for heterogeneous Knoevenagel condensation, a strategy for C-C
placed towards sodium-ion batteries, specifically using catalytic poly­ bond formation in organic synthesis [333]. These reactions are usually
mers to improve electrochemical reaction processes [321,322]. For carried out in harsh conditions like pyridine or sodium ethoxide. How­
instance, annealing g-C3N4 at different temperatures results in varying ever, a Salen-based CMP can be synthesised which can catalyze the
layer spacings, allowing for increased Na+ storage capacity [323]. This Knoevenagel condensation reaction between aromatic aldehyde and
is because annealing affects the content of pyrrole-N, which has a very malononitrile in a H2O-ethanol mixture. This solvent is considered more
active p-orbital lone pair. Metal-air batteries have also been receiving sustainable from an environmental point of view. Moreover, the CMP
significant attention since they use oxygen in the air as a reaction ma­ exhibited excellent thermal stability as well as high porosity, CO2 up­
terial and are thus an attractive alternative for energy saving and low take, reusability (up to 8 times) and catalytic performance for the
carbon emissions [324]. By introducing a catalytic polymeric material to cycloaddition of CO2 with epoxides [334].
the electrode, the charge–discharge capacity/performance, energy effi­
ciency, and cycle stability of the batteries can be improved [325]. For 4. Concluding remarks
instance, by using single-atom catalysts, such as via the introduction of
Pt onto ultrathin g-C3N4, the stability and access to multiple active sites It is not often that materials transcend a single field or function and
can be increased [326]. The single-atom Pt can achieve high dis­ demonstrate high applicability in a wide range of fields. Within this
persibility which increases the utilization efficiency and enhances review, polymer-based catalysts have been shown to be such a material.
electrochemical activity. When employed as Li-O2 batteries’ cathode A focus was placed on studies where polymers themselves performed a
catalyst, Pt/g-C3N4 exhibits higher discharge specific capacities catalytic process, or synergistically, in conjunction with other materials,
compared to pure g-C3N4. Alternatively, organo-catalysts like led to enhanced activity. As a result, polymers which met this criterion
redox-active organic polymers can be developed that enhance diffusion were predominantly photocatalytic-based, relying on light absorption to
and cycling performance in lithium-ion batteries [327]. To achieve this, create electron-hole pairs that react with absorbed molecules or species
triptycene benzoquinone structural units are used to achieve a high in the environment. The performance of photocatalysts are significantly
density of redox-active sites, while tetramethyl groups provide sufficient affected by three main areas: (i) light absorption, (ii) separation of
steric hindrance to restrict rotation about the C–N imide single bond. photogenerated carriers, and (iii) activation and reaction of reactant
molecules. Described and referenced within this review are a plethora of
3.3.3. Energy efficiency studies which aim to enhance one or more of these steps, with consid­
In the field of organic chemistry, a great emphasis has been placed on erable improvements over the last 1–2 decades. However, there are
“one-pot” cascade reactions to improve reaction efficiency and minimize several key areas which could drive the field towards the development of
material loss from multiple iterations of reaction and purification steps a new generation of catalytic polymers, which will be discussed in this
[328]. Catalytic polymeric materials have also been used to tackle this section.
issue, such as in deacetalisation–Henry [329] and deacetalisa­ Fundamental understanding: Despite research into these photo­
tion–Knoevenagel reactions. For instance, a core-shell microparticle catalytic polymers started 1–2 decades ago, their reaction mechanism is
(CSM) can be developed with an acid catalyst (PSSA; polystyrene sul­ still not fully understood. In many cases, researchers outline a range of
fonic acid) as the core and a base catalyst (VEMAP; 4-N-(4-vinylbenzyl) possible pathways that could be causing the catalytic activity being
observed. The lack of fundamental understanding of the reaction

16
F. Mazur et al. Applied Materials Today 35 (2023) 101937

mechanism is severely constraining new design concepts. This is because catalysts have the potential to play a crucial role in tackling the issues
it becomes difficult to identify the exact area which is limiting the currently faced in medicine, environmental remediation, and energy.
platform. As such, it is important to understand fundamental concepts Considering the rapid development of this field in the last decade, the
before continuing forward. This idea was put forward sarcastically in a next generation of catalyst that will tackle many of these mentioned
recent study where bird feces were introduced into graphene to improve issues are likely to appear soon.
electrocatalytic performance [335].
Performance in real samples: Many studies evaluate the activity of Declaration of Competing Interest
their catalyst in pristine conditions rather than environments more
closely resembling their intended final use. This could lead to misleading The authors declare that they have no known competing financial
performance indicators, as various contaminants present in “real” interests or personal relationships that could have appeared to influence
samples undergoing the catalytic process could affect activity and the work reported in this paper.
recyclability of the platform. This is a key factor which can significantly
increase future studies’ overall impact and promote industry collabo­ Data availability
ration opportunities.
Multifunctional catalysts: In line with “real-sample” performance, No data was used for the research described in the article.
recently, studies have shown catalysts are able to perform efficient
photocatalytic degradation towards various pollutant types. This is an
attractive quality as pollutants are commonly synchronous, therefore Acknowledgments
developing a catalyst which can efficiently degrade various types of
contaminants without loss in performance or recyclability is ideal. R.C. acknowledges support from the National Health and Medical
Catalyst recyclability/stability: A major focus with all studies reviewed Research Council Emerging Leadership Investigator Grant (NHMRC
herein is on increasing catalytic activity, either for pollutant degrada­ APP1173428) and the UNSW Scientia Fellowship. F.M. acknowledges
tion, energy generation, or other applications. Although a very impor­ Angie Davina Tjandra for designing and creating Figs. 1–3.
tant factor, the recyclability of the catalyst, in other words, the ability of
a catalyst to be recovered, regenerated, and reused multiple times
References
without significant loss of activity or selectivity, is often overlooked,
with many studies evaluating <5 cyclic uses. The catalyst recyclability is [1] Y.H. Lin, et al., J. Am. Soc. Mass Spectrom. 34 (2023) 109–118.
highly dependent on the industry and sample, but in all cases has a high [2] Allied Market Research, Catalyst Market Statistics, Trends: Industry Analysis
impact on costs. As such, demonstrating catalysts can be re-used mul­ 2030. [Link] 2023 (accessed
22 March 2023).
tiple times with little loss in performance whilst in “real” sample envi­ [3] E. Roduner, Chem. Soc. Rev. 43 (2014) 8226–8239.
ronments would be highly attractive. Furthermore, considering some [4] A.Z. Fadhel, et al., Molecules 15 (2010) 8400–8424.
microbes remain in the environment for extended periods, catalytic [5] C.M. Friend, et al., Acc. Chem. Res. 50 (2017) 517–521.
[6] Y.P. Xue, et al., Chem. Soc. Rev. 47 (2018) 1516–1561.
stability is another important factor to evaluate. [7] M. Miceli, et al., Catalysts 11 (2021) 591.
Large-scale production: With the growing concern regarding sustain­ [8] P. Peng, et al., Acc. Chem. Res. 50 (2017) 1171–1183.
able practices, catalyst synthesis should shift towards using more cost- [9] Y.M. Lee, et al., Chem. Sci. (2023) 4205–4218.
[10] L. Candish, et al., Chem. Rev. 122 (2022) 2907–2980.
effective strategies and environmentally friendly solvents. This will [11] A. Ubaldo-Alarcón, et al., Polymers 14 (2022) 2352.
lead to compounded benefits as a holistic mindset is used which con­ [12] T.L. Nghiem, et al., Polymers 12 (2020) 2190.
siders all aspects of the catalyst life cycle, especially when considering [13] Q. Guo, et al., Adv. Mater. 31 (2019), 1901997.
[14] H. Dong, et al., Int. J. Hydrogen Energy 46 (2021) 32044–32054.
the previous suggestions.
[15] Y. Liu, et al., Mater. Today Energy 31 (2023), 101211.
Other research avenues: Currently, photocatalysis is the main mech­ [16] A. Alaghmandfard, et al., Nanomaterials 12 (2022) 294.
anism employed for most polymer-based catalysts. In these cases, light is [17] Y. Xing, et al., Chin. Chem. Lett. 32 (2021) 13–20.
[18] S. Gao, et al., Appl. Catal. B: Environ. 295 (2021), 120272.
required to trigger the activity. Other mechanisms which do not require
[19] H. Zhou, et al., Chem 8 (2022) 465–479.
an external trigger could also be investigated. For instance, we have [20] X. Ruan, et al., Chem. Eng. J. 428 (2022), 132579.
recently demonstrated that amine-rich polymers can efficiently cata­ [21] L. Ai, et al., Small 17 (2021), 2007523.
lytically generate nitric oxide (NO; therapeutic agent) from an endoge­ [22] P. Yadav, et al., Mater. Adv. 3 (2022) 1432–1458.
[23] Z. Chen, et al., ACS Appl. Mater. Interfaces 13 (2021) 1145–1151.
nous prodrug, GSNO (S-nitrosoglutathione) [336]. This means that [24] N. Haque, et al., ACS Appl. Nano Mater. 4 (2021) 7663–7674.
unmodified polymers can open new avenues for NO-generating mate­ [25] S. Jin et al., InfoMat 4 (2022) e12277.
rials for various therapeutic applications (antibacterial, antiviral, anti­ [26] T. Sick, et al., J. Am. Chem. Soc. 140 (2018) 2085–2092.
[27] L. Ma, et al., Chin. Chem. Lett. 27 (2016) 1383–1394.
cancer, antithrombotic, wound healing, etc.). This largely unexplored [28] R.K. Sharma, et al., Mater. Horiz. 7 (2020) 411–454.
diversity of polymer catalysts can be traced back almost 50 years ago, [29] V. Hasija, et al., Coord. Chem. Rev. 452 (2022), 214298.
where modified PEI was used to catalyze the decarboxylation of oxal­ [30] H. Wang, et al., Chem. Soc. Rev. 49 (2020) 4135–4165.
[31] H.L. Qian, et al., Angew. Chem. Int. Ed. 59 (2020) 17607–17613.
acetate [337]. In fact, some researchers have begun to notice this largely [32] S. Ma, et al., Angew. Chem. Int. Ed. 61 (2022), e202208919.
unexplored field and developed polymers with enzyme mimicking [33] Y. Xing, et al., ACS Appl. Mater. Interfaces 12 (2020) 51555–51562.
qualities by incorporating the hydroxyl group of a serine residue, an [34] H.S. Xu, et al., Nat. Commun. 11 (2020) 1434.
[35] X. Wang, et al., Nat. Chem. 10 (2018) 1180–1189.
imidazole group of a histidine residue, and a carboxyl group of an
[36] A.I. Cooper, Adv. Mater. 21 (2009) 1291–1295.
aspartic acid, onto a polymer scaffold for esterolytic applications [338]. [37] S. Kim, et al., ACS Nano 16 (2022) 17041–17048.
On the other hand, antimicrobial applications for polymer-based cata­ [38] S.N. Talapaneni, et al., Chem. Mater. 28 (2016) 4460–4466.
[39] [Link] Rao, et al., Phys. Chem. Chem. Phys. 18 (2016) 156–163.
lysts have mainly focussed on bacteria, most likely due to their relatively
[40] C.A. Wang, et al., RSC Adv. 7 (2017) 408–414.
simple structure and well-developed detection methods. It would then [41] T.M. Geng, et al., J. Mater. Sci. 51 (2016) 4104–4114.
follow that a greater focus be placed on other microbes like fungi and [42] C. Zhang, et al., Adv. Funct. Mater. 28 (2018), 1705432.
protozoa, as well as other pollutants. This could also provide funda­ [43] M.E. Elhalwagy, et al., Mater. Today Chem. 27 (2023), 101344.
[44] G.G. Celestino, et al., Environ. Sci. Pollut. Res. 26 (2019) 28444–28454.
mental understanding of degradation mechanisms. Finally, recent [45] L. Zhang, et al., Chemosphere 248 (2020), 126102.
studies have also allowed for the automated synthesis of large libraries [46] M.R. Abukhadra, et al., Int. J. Biol. Macromol. 173 (2021) 435–444.
of peptides with potential catalytic properties utilizing the unique [47] S. Senguttuvan, et al., Chemosphere 287 (2022), 132164.
[48] S. Bao, et al., J. Hazard. Mater. 409 (2021), 124470.
chemistry and properties of selenium [339]. [49] Y.R. Lee, et al., Chem. Eng. J. 378 (2019), 122133.
When taking these improvements into account, polymer-based [50] Y. Ma, et al., Sci. Total Environ. 821 (2022), 153453.

17
F. Mazur et al. Applied Materials Today 35 (2023) 101937

[51] S.L. Lee, et al., Polymers 11 (2019) 206. [137] V. Kumar, et al., Chemosphere 216 (2019) 449–462.
[52] A. Yadav, et al., Colloid Interface Sci. Commun. 40 (2021), 100339. [138] M. Yu, et al., ACS Appl. Mater. Interfaces 14 (2022) 5376–5383.
[53] K.S. Egorova, et al., Angew. Chem. Int. Ed. 55 (2016) 12150–12162. [139] C. Sun, et al., Sep. Purif. Technol. 271 (2021), 118873.
[54] Z. Grądzki, et al., Poult. Sci. 99 (2020) 2424–2437. [140] Y. Zeng, et al., Chem. Eng. J. 435 (2022), 134918.
[55] T.P. Van Boeckel, et al., Science 365 (2019) eaaw1944. [141] X. Zhang, et al., Environ. Sci. Water Res. Technol. 8 (2022) 1965–1975.
[56] V. Mahmoodi, et al., Environ. Sci. Pollut. Res. 25 (2018) 8268–8285. [142] Y. Wang, et al., J. Hazard. Mater. 408 (2021), 124818.
[57] A. Kumar, et al., Chem. Eng. J. 382 (2020), 122937. [143] C.W. Zheng, et al., Chem. Eng. J. 403 (2021), 126400.
[58] G.N. Bosio, et al., Photodiagnosis Photodyn. Ther. 37 (2022), 102683. [144] R. Al-Tohamy, et al., Ecotoxicol. Environ. Saf. 231 (2022), 113160.
[59] J.H. Thurston, et al., J. Colloid Interface Sci. 505 (2017) 910–918. [145] S. Parmar, et al., Chapter 8 - Microorganism: an ecofriendly tool for waste
[60] H. Zhang, et al., Rare Metals 41 (2022) 1570–1582. management and environmental safety, in: M.P. Shah, et al. (Eds.), Development
[61] A. Balapure, et al., ACS Appl. Mater. Interfaces 12 (2020) 21481–21493. in Wastewater Treatment Research and Processes, Elsevier, 2022, pp. 175–193.
[62] W. Wang, et al., Chem. Eng. J. 378 (2019), 122132. [146] R. Patil, et al., Chapter 8 - Constructed wetland: a promising technology for the
[63] Y. Yin, et al., Nanoscale 14 (2022) 2686–2695. treatment of hazardous textile dyes and effluent, in: M. Shah, et al. (Eds.),
[64] H. Liu, et al., Appl. Catal. B: Environ. 261 (2020), 118201. Development in Wastewater Treatment Research and Processes, Elsevier, 2022,
[65] X. Zeng, et al., Appl. Catal. B: Environ. 274 (2020), 119095. pp. 173–198.
[66] Y. Cheng, et al., J. Hazard. Mater. 383 (2020), 121166. [147] Y. Zhou, et al., Environ. Pollut. 252 (2019) 352–365.
[67] L. Wang, et al., Appl. Catal. B: Environ. 325 (2023), 122345. [148] T.S. Anantha Singh, et al., Environ. Eng. Sci. 30 (2013) 333–349.
[68] X. Zhang, et al., Chem. Eng. J. 429 (2022), 132588. [149] T.P. Vijayakumar, et al., Diamond Relat. Mater. 121 (2022), 108735.
[69] F. Liu, et al., Environ. Sci. Technol. 55 (2021) 5371–5381. [150] M. Ding, et al., Chin. Chem. Lett. 31 (2020) 71–76.
[70] A. Mal, et al., Angew. Chem. Int. Ed. 59 (2020) 8713–8719. [151] I. Rabani, et al., J. Hazard. Mater. 407 (2021), 124360.
[71] S. Qiao, et al., ACS Appl. Mater. Interfaces 13 (2021) 32058–32066. [152] J. Wang, et al., J. Phys. Chem. Solids 136 (2020), 109164.
[72] F.L. Meng, et al., Talanta 233 (2021), 122536. [153] F. Shahsavandi, et al., Appl. Surf. Sci. 585 (2022), 152728.
[73] T. Liu, et al., J. Photochem. Photobiol. B: Biol. 175 (2017) 156–162. [154] M. Jourshabani, et al., Chem. Eng. J. 427 (2022), 131710.
[74] C. Liu, et al., Chem. Eng. J. 430 (2022), 132663. [155] H. Cong, et al., ChemistrySelect 6 (2021) 9458–9466.
[75] Y. Zhang, et al., J. Mater. Chem. B 10 (2022) 3285–3292. [156] H. Fattahimoghaddam, et al., J. Hazard. Mater. 403 (2021), 123703.
[76] G.P. Yang, et al., ACS Appl. Mater. Interfaces 14 (2022) 28289–28300. [157] S. Faryad, et al., Chemosphere 320 (2023), 138002.
[77] Y. Wang, et al., ACS Appl. Mater. Interfaces 10 (2018) 34878–34885. [158] Y. Zhang, et al., J. Mater. Chem. A 7 (2019) 16364–16371.
[78] Z. Yuan, et al., Small 17 (2021), 2007522. [159] S. He, et al., Appl. Catal. B: Environ. 247 (2019) 49–56.
[79] B. Tao, et al., Chem. Eng. J. 403 (2021), 126182. [160] Y. Hou, et al., Dalton Trans. 48 (2019) 14989–14995.
[80] D. Han, et al., J. Mater. Sci. Technol. 62 (2021) 83–95. [161] Y. Yao, et al., J. Colloid Interface Sci. 554 (2019) 376–387.
[81] J. Song, et al., ACS Appl. Mater. Interfaces 12 (2020) 8915–8928. [162] R.L. Wang, et al., Dalton Trans. 48 (2019) 1051–1059.
[82] Q. Zeng, et al., Bioact. Mater. 6 (2021) 2647–2657. [163] H. Xue, et al., Ind. Eng. Chem. Res. 60 (2021) 8687–8695.
[83] D. Wang, et al., Commun. Mater. 3 (2022) 98. [164] Y. Li, et al., ACS Appl. Nano Mater. 2 (2019) 7290–7298.
[84] Y. Xiang, et al., J. Mater. Sci. Technol. 57 (2020) 1–11. [165] H. Xue, et al., Green Energy Environ. 8 (2023) 194–199.
[85] A. Ahmed, et al., Eur. Polym. J. 130 (2020), 109650. [166] K.K. Khaing, et al., Inorg. Chem. 59 (2020) 6942–6952.
[86] C. Liu, et al., Carbohydr. Polym. 303 (2023), 120436. [167] T. Bhoyar, et al., ACS Appl. Nano Mater. 6 (2023) 3484–3496.
[87] B. Sun, et al., Small 18 (2022), 2200895. [168] F. Liu, et al., J. Hazard. Mater. 398 (2020), 122865.
[88] P. Li, et al., Compos. B: Eng. 252 (2023), 110506. [169] J. Liang, et al., Water Res. 123 (2017) 632–641.
[89] Y. Zou, et al., ACS Appl. Mater. Interfaces 14 (2022) 8680–8692. [170] J. Yu, et al., Appl. Surf. Sci. 470 (2019) 84–90.
[90] B. Sun, et al., ACS Appl. Mater. Interfaces 13 (2021) 42396–42410. [171] X. Sun, et al., Colloids Surf. Physicochem. Eng. Aspects 570 (2019) 199–209.
[91] H. Elkateb, et al., Int. J. Pharm. 588 (2020), 119794. [172] V.E. Podasca, et al., J. Photochem. Photobiol. A: Chem. 371 (2019) 188–195.
[92] Y. Sun, et al., Adv. Mater. 34 (2022), 2109580. [173] C. Yang, et al., Chem. Eng. J. 374 (2019) 1382–1393.
[93] P. Chen, et al., Sci. Adv. 8 (2022) eabl8812. [174] M. Faisal, et al., Mater. Today Commun. 24 (2020), 101048.
[94] N.Y. Steinman, et al., Biomacromolecules 22 (2021) 4357–4364. [175] R. Tanwar, et al., RSC Adv. 9 (2019) 8977–8993.
[95] T. Yadavalli, et al., ACS Infect. Dis. 6 (2020) 2926–2937. [176] S. Saha, et al., Int. Nano Lett. 9 (2019) 127–139.
[96] L. Soria-Martinez, et al., J. Am. Chem. Soc. 142 (2020) 5252–5265. [177] J. Bhadra, et al., J. Environ. Chem. Eng. 8 (2020), 103746.
[97] S.C. Günther, et al., Sci. Rep. 10 (2020) 768. [178] M. Mitra, et al., ACS Omega 4 (2019) 1623–1635.
[98] C. Zhang, et al., Chemosphere 208 (2018) 84–92. [179] V. Najafi, et al., J. Polym. Environ. 27 (2019) 784–793.
[99] V. Hasija, et al., Catalysts 11 (2021) 1448. [180] H. Wang, et al., Mater. Lett. 260 (2020), 126906.
[100] N. El-Atab, et al., ACS Nano 14 (2020) 7659–7665. [181] [Link] Taymaz, et al., Polym. Bull. 78 (2021) 2849–2865.
[101] F. Mazur, et al., Adv. Mater. Technol. 8 (2023), 2201916. [182] Y.J. Lee, et al., Appl. Sci. 10 (2020) 6710.
[102] I. Foffa, et al., J. Appl. Biomater. Funct. Mater. 20 (2022), 22808000221076326. [183] M. Faisal, et al., Ceram. Int. 45 (2019) 20484–20492.
[103] L.G. Ding, et al., J. Mater. Chem. A 10 (2022) 3346–3358. [184] H. Mittal, et al., Sep. Purif. Technol. 254 (2021), 117508.
[104] Y. Li, et al., Water Res. 106 (2016) 249–258. [185] M. Faisal, et al., Sep. Purif. Technol. 256 (2021), 117847.
[105] C. Zhang, et al., Appl. Catal. B: Environ. 248 (2019) 11–21. [186] G. Sarigul, et al., Adv. Sustain. Syst. 5 (2021), 2100076.
[106] R. Cheng, et al., Catalysts 8 (2018) 406. [187] A. Gogoi, et al., Groundw. Sustain. Dev. 6 (2018) 169–180.
[107] A.R. Bagheri, et al., Biorg. Med. Chem. 51 (2021), 116493. [188] F. Mazur, et al., Nat. Rev. Bioeng. 1 (2023) 180–192.
[108] Y. Jia, et al., Mater. Sci. Eng. C 117 (2020), 111243. [189] I.B. Seiple, et al., Nature 533 (2016) 338–345.
[109] S. Liu, et al., Macromol. Rapid Commun. 41 (2020), 1900570. [190] Z. Zhao, et al., Chem. Eng. J. 407 (2021), 127167.
[110] R. Anbazhagan, et al., Mater. Sci. Eng. C 120 (2021), 111704. [191] Y. Zhou, et al., J. Alloys Compd. 890 (2022), 161786.
[111] S.B. Wang, et al., Biomaterials 234 (2020), 119772. [192] J. Luo, et al., J. Colloid Interface Sci. 610 (2022) 280–294.
[112] X. Wan, et al., Chem. Commun. 57 (2021) 6145–6148. [193] C. Sandoval, et al., Mater. Res. Bull. 116 (2019) 8–15.
[113] P. Wang, et al., Chem. Sci. 11 (2020) 1299–1306. [194] Z. Dong, et al., Chem. Eng. J. 403 (2021), 126383.
[114] F. Shu, et al., J. Nanobiotechnol. 19 (2021) 4. [195] X. Li, et al., Chem. Eng. J. 429 (2022), 132130.
[115] Z. Mi, et al., J. Am. Chem. Soc. 141 (2019) 14433–14442. [196] J. Guo, et al., Catal. Commun. 122 (2019) 63–67.
[116] D. Wang, et al., Biomaterials 223 (2019), 119459. [197] M. Yu, et al., Chin. Chem. Lett. 32 (2021) 2155–2158.
[117] P. Gao, et al., Chem. Sci. 11 (2020) 6882–6888. [198] Y. Yu, et al., J. Hazard. Mater. 404 (2021), 124171.
[118] K. Wang, et al., ACS Appl. Mater. Interfaces 11 (2019) 39503–39512. [199] Z. Huang, et al., New J. Chem. 46 (2022) 3727–3737.
[119] S. Liu, et al., ACS Appl. Mater. Interfaces 12 (2020) 43456–43465. [200] S. Li, et al., Chem. Eng. J. 382 (2020), 122394.
[120] Q. Guan, et al., ACS Nano 13 (2019) 13304–13316. [201] Y. Wu, et al., J. Hazard. Mater. 414 (2021), 125547.
[121] C. Hu, et al., Chem. Commun. 55 (2019) 9164–9167. [202] F. Guo, et al., Sep. Purif. Technol. 210 (2019) 608–615.
[122] C. Hu, et al., ACS Appl. Mater. Interfaces 11 (2019) 23072–23082. [203] D. Jia, et al., Env. Sci. Nano 8 (2021) 415–431.
[123] H. Taheri, et al., Small 16 (2020), 1904619. [204] Q. Shen, et al., J. Hazard. Mater. 413 (2021), 125376.
[124] N.S. Heo, et al., Mater. Sci. Eng. C 90 (2018) 531–538. [205] H. Sun, et al., Chem. Eng. J. 406 (2021), 126844.
[125] X. Wu, et al., ACS Appl. Bio Mater. 2 (2019) 1998–2005. [206] J. Jiang, et al., J. Hazard. Mater. 414 (2021), 125528.
[126] X. Liu, et al., Spectrochim. Acta A Mol. Biomol. Spectrosc. 250 (2021), 119363. [207] Q. Su, et al., Chem. Eng. J. 427 (2022), 131594.
[127] J. Dai, et al., ACS Appl. Mater. Interfaces 11 (2019) 10589–10596. [208] X. Zhang, et al., J. Hazard. Mater. 418 (2021), 126333.
[128] Y. Wu, et al., Appl. Catal. B: Environ. 269 (2020) 118848. [209] Y. Wang, et al., Ceram. Int. 48 (2022) 2459–2469.
[129] M. Xu, et al., Chem. Eng. J. 360 (2019) 866–878. [210] C. Jin, et al., Ind. Eng. Chem. Res. 59 (2020) 2860–2873.
[130] Z. Xu, et al., Inorg. Chem. 61 (2022) 20346–20357. [211] G.Q. Zhao, et al., Colloids Surf. Physicochem. Eng. Aspects 652 (2022), 129870.
[131] W. Zhang, et al., Colloids Surf. B. Biointerfaces 199 (2021), 111549. [212] K. Ren, et al., Carbon 186 (2022) 355–366.
[132] Y. Yuan, et al., Adv. Mater. 30 (2018), 1800069. [213] X. Li, et al., Nano Energy 81 (2021), 105671.
[133] R. Wang, et al., Eur. Polym. J. 110 (2019) 1–8. [214] M. Jourshabani, et al., Appl. Catal. B: Environ. 302 (2022), 120839.
[134] L. Wang, et al., J. Hazard. Mater. 444 (2023), 130332. [215] X. Yuan, et al., Chem. Eng. J. 416 (2021), 129148.
[135] Y. Chae, et al., Environ. Pollut. 240 (2018) 387–395. [216] J. Lan, et al., Nanoscale Adv. 3 (2021) 4646–4658.
[136] M. Patel, et al., Chem. Rev. 119 (2019) 3510–3673. [217] H. Shi, et al., J. Solid State Chem. 311 (2022), 123069.

18
F. Mazur et al. Applied Materials Today 35 (2023) 101937

[218] P. Kumar, et al., J. Am. Chem. Soc. 141 (2019) 5415–5436. [278] P. Yang, et al., Appl. Catal. B: Environ. 256 (2019), 117794.
[219] J. Zhang, et al., Appl. Catal. B: Environ. 289 (2021), 120023. [279] D. Zhao, et al., Adv. Mater. 31 (2019), 1903545.
[220] X. Yuan, et al., J. Environ. Chem. Eng. 8 (2020), 104178. [280] H.J. Kong, et al., Chem. Mater. 28 (2016) 1318–1324.
[221] L. Sun, et al., Chin. J. Catal. 40 (2019) 80–94. [281] S.K. Das, et al., Front. Chem. 9 (2021) 803860.
[222] J. Wang, et al., Mater. Sci. Semicond. Process. 121 (2021), 105329. [282] X. Chen, et al., Chem. Rev. 110 (2010) 6503–6570.
[223] L. Ye, et al., Mater. Lett. 236 (2019) 370–373. [283] S. Cao, et al., Trends Chem. 2 (2020) 57–70.
[224] M.O. Mavukkandy, et al., Chem. Eng. J. 427 (2022), 131021. [284] A. Samal, et al., ChemSusChem 9 (2016) 3150–3160.
[225] IEA, Global EV Outlook 2022 - Data product. [Link] [285] L. Ran, et al., J. Am. Chem. Soc. 144 (2022) 17097–17109.
istics/data-product/global-ev-outlook-2022, 2023 (accessed 22 March 2023). [286] Z. Zhu, et al., J. Chin. Chem. Soc. 67 (2020) 1654–1660.
[226] H. Wang, et al., ACS Appl. Mater. Interfaces 12 (2020) 40176–40185. [287] P. Madhusudan, et al., Appl. Catal. B: Environ. 282 (2021), 119600.
[227] X. Hu, et al., Appl. Surf. Sci. 556 (2021), 149817. [288] R.R. Ikreedeegh, et al., J. Environ. Chem. Eng. 9 (2021), 105600.
[228] K.A. Brown, et al., Science 352 (2016) 448–450. [289] S. Kumar, et al., J. Photochem. Photobiol. A: Chem. 439 (2023), 114591.
[229] A. Eizawa, et al., Nat. Commun. 8 (2017) 14874. [290] X. Jiang, et al., Appl. Catal. B: Environ. 281 (2021), 119473.
[230] T. Oshikiri, et al., Angew. Chem. Int. Ed. 55 (2016) 3942–3946. [291] Y.N. Gong, et al., Nano Res. 15 (2022) 551–556.
[231] V. Kyriakou, et al., Catal. Today 286 (2017) 2–13. [292] X. Zhang, et al., Appl. Catal. B: Environ. 286 (2021), 119879.
[232] D. Liu, et al., Small 15 (2019), 1902291. [293] Y. Huo, et al., ACS Appl. Energy Mater. 4 (2021) 956–968.
[233] J. Lasek, et al., J. Photochem. Photobiol. C: Photochem. Rev. 14 (2013) 29–52. [294] N. Su, et al., Inorg. Chem. 61 (2022) 15600–15606.
[234] L. Hu, et al., Carbon 200 (2022) 187–198. [295] Q. Liu, et al., Nano Energy 77 (2020), 105104.
[235] J. Zhang, et al., Appl. Catal. B: Environ. 296 (2021), 120372. [296] K. Chen, et al., Chem. Eng. J. 418 (2021), 129476.
[236] X. Li, et al., Environ. Sci. Nano 7 (2020) 1990–1998. [297] H. Li, et al., Chem. Commun. 56 (2020) 5641–5644.
[237] P. Van Viet, et al., J. Materiomics 8 (2022) 1–8. [298] W. Yang, et al., J. Mater. Chem. A 11 (2023) 2225–2232.
[238] J. Liao, et al., Chem. Eng. J. 379 (2020), 122282. [299] Q. Wang, et al., Adv. Fiber Mater. 4 (2022) 342–360.
[239] Y. Ren, et al., Chin. J. Catal. 42 (2021) 69–77. [300] N. Meng, et al., ACS Catal. 9 (2019) 10983–10989.
[240] M. Fu, et al., J. Alloys Compd. 906 (2022), 164371. [301] Y. Huang, et al., Appl. Catal. B: Environ. 288 (2021), 120001.
[241] P. Lu, et al., J. Alloys Compd. 934 (2023), 167825. [302] Y. Zhao, et al., Appl. Catal. B: Environ. 291 (2021), 119915.
[242] X. Xia, et al., Langmuir 39 (2023) 1250–1261. [303] B. Wang, et al., Chem. Eng. J. 396 (2020), 125255.
[243] A. Hashemi Monfared, et al., Prog. Org. Coat. 136 (2019), 105257. [304] Z. Fu, et al., Chem. Sci. 11 (2020) 543–550.
[244] IEA, World – World Energy Balances: Overview – Analysis. [Link] [305] [Link] Chowdhury, et al., New J. Chem. 44 (2020) 11720–11726.
g/reports/world-energy-balances-overview/world, 2023 (accessed 22 March [306] K. Guo, et al., Chem. Eng. J. 405 (2021), 127011.
2023). [307] M. Liu, et al., Sci. Bull. 66 (2021) 1659–1668.
[245] J. Chen, et al., Appl. Catal. B: Environ. 262 (2020), 118271. [308] H. Dong, et al., Appl. Catal. B: Environ. 303 (2022), 120897.
[246] K. Wang, et al., J. Catal. 348 (2017) 168–176. [309] S.Y. Chi, et al., J. Mater. Chem. A 10 (2022) 4653–4659.
[247] J.H. Park, et al., Adv. Funct. Mater. 31 (2021), 2101727. [310] Y. Xie, et al., Appl. Catal. B: Environ. 265 (2020), 118581.
[248] Q. Wang, et al., Molecules 28 (2023) 432. [311] H. Zhao, et al., Chem Catal. 2 (2022) 1720–1733.
[249] M. Kombo, et al., Catal. Sci. Technol. 9 (2019) 2196–2202. [312] Y. Kofuji, et al., ChemCatChem 10 (2018) 2070–2077.
[250] W. Lei, et al., Int. J. Hydrogen Energy 47 (2022) 10877–10890. [313] J. Luo, et al., Appl. Catal. B: Environ. 301 (2022), 120757.
[251] D.E. Lee, et al., J. Mater. Sci. Technol. 148 (2023) 19–30. [314] X. Zeng, et al., ACS Catal. 10 (2020) 3697–3706.
[252] R. Shen, et al., Appl. Catal. B: Environ. 291 (2021), 120104. [315] M.C. Kim, et al., J. Power Sources 453 (2020), 227850.
[253] H. Zhou, et al., J. Colloid Interface Sci. 610 (2022) 126–135. [316] Y. Hou, et al., Nano Energy 8 (2014) 157–164.
[254] L. Acharya, et al., New J. Chem. 46 (2022) 3493–3503. [317] X. Li, et al., Adv. Funct. Mater. 25 (2015) 6858–6866.
[255] S. Meng, et al., J. Colloid Interface Sci. 585 (2021) 108–117. [318] M. Hankel, et al., J. Phys. Chem. C 119 (2015) 21921–21927.
[256] C. Li, et al., Mater. Lett. 308 (2022), 131232. [319] J. Chen, et al., ACS Nano 11 (2017) 12650–12657.
[257] Z. Mo, et al., Appl. Catal. B: Environ. 225 (2018) 154–161. [320] Z. Zhuang, et al., Sustain. Energy Fuels 5 (2021) 2433–2440.
[258] Z. Zhao, et al., Langmuir 38 (2022) 11054–11067. [321] A. Roy, et al., Batter. Supercaps 4 (2021) 978–988.
[259] A. Cao, et al., ACS Sustain. Chem. Eng. 7 (2019) 2492–2499. [322] M. Huang, et al., Small Structures 2 (2021), 2000085.
[260] Q. Li, et al., Chem. Eng. J. 450 (2022), 138010. [323] J. Liu, et al., Adv. Mater. 31 (2019), 1901261.
[261] J. Li, et al., Carbon 172 (2021) 602–612. [324] C. Liu, et al., ACS Nano 15 (2021) 3309–3319.
[262] W. Shi, et al., Chem. Eng. J. 382 (2020), 122960. [325] C. Wang, et al., J. Mater. Chem. A 7 (2019) 1451–1458.
[263] K. Qi, et al., Chin. J. Catal. 41 (2020) 114–121. [326] W. Zhao, et al., J. Colloid Interface Sci. 564 (2020) 28–36.
[264] H. Yu, et al., Solar RRL 5 (2021), 2000372. [327] A. Wang, et al., J. Am. Chem. Soc. 144 (2022) 17198–17208.
[265] G. Liang, et al., Appl. Surf. Sci. 504 (2020), 144448. [328] T.L. Lohr, et al., Nat. Chem. 7 (2015) 477–482.
[266] C. Xi, et al., Mater. Lett. 307 (2022), 131012. [329] X. Wang, et al., Appl. Surf. Sci. 478 (2019) 221–229.
[267] Y. Li, et al., ChemSusChem 15 (2022), e202200901. [330] C. Chen, et al., Angew. Chem. Int. Ed. 60 (2021) 237–241.
[268] Y. Wang, et al., Nat. Commun. 13 (2022) 100. [331] S. Ghanooni, et al., ACS Omega 7 (2022) 30989–31002.
[269] D. Shang, et al., ACS Sustain. Chem. Eng. 9 (2021) 14238–14248. [332] R.Y. Liu, et al., Nat. Commun. 13 (2022) 2775.
[270] J. Chen, et al., Appl. Catal. B: Environ. 241 (2019) 461–470. [333] R. Gao, et al., Front. Chem. 9 (2021) 687183.
[271] J. Xu, et al., Angew. Chem. Int. Ed. 59 (2020) 23845–23853. [334] P. Ju, et al., Catal. Lett. (2022) 2125–2136.
[272] Y. Li, et al., Nat. Commun. 13 (2022) 1355. [335] L. Wang, et al., ACS Nano 14 (2020) 21–25.
[273] Y. Kim, et al., Appl. Catal. B: Environ. 280 (2021), 119423. [336] Z. Luo, et al., Small 19 (2023), 2200502.
[274] V.D. Dao, et al., J. Alloys Compd. 775 (2019) 942–949. [337] W.J. Spetnagel, et al., J. Am. Chem. Soc. 98 (1976) 8199–8204.
[275] H.M. El-Bery, et al., RSC Adv. 11 (2021) 13229–13244. [338] A. Bhaskaran, et al., Polymer 225 (2021), 123735.
[276] J. Ma, et al., Renew. Energy 156 (2020) 1008–1018. [339] G. Wu, et al., Nat. Synth. 2 (2023) 515–526.
[277] S. Chu, et al., ACS Catal. 3 (2013) 912–919.

19

You might also like