Open Access Article
This Open Access Article is licensed under a Creative Commons Attribution-Non Commercial 3.0 Unported Licence

Recent developments in hydrogels containing copper and palladium for the catalytic reduction/degradation of organic pollutants

Jaber Dadashia, Mohammad Ali Ghasemzadeh*b and Masoud Salavati-Niasaric
aCatalysts and Organic Synthesis Research Laboratory, Department of Chemistry, Iran University of Science and Technology, Tehran, Iran
bDepartment of Chemistry, Qom Branch, Islamic Azad University, Qom, Iran. E-mail: ma.ghasemzadeh@iau.ac.ir; qasemzade.a@gmail.com
cInstitute of Nano Science and Nano Technology, University of Kashan, Kashan, Iran

Received 1st June 2022 , Accepted 27th July 2022

First published on 18th August 2022


Abstract

The elimination of toxic and hazardous contaminants from different environmental media has become a global challenge, causing researchers to focus on the treatment of pollutants. Accordingly, the elimination of inorganic and organic pollutants using sustainable, effective, and low-cost heterogeneous catalysts is considered as one of the most essential routes for this aim. Thus, many efforts have been devoted to the synthesis of novel compounds and improving their catalytic performance. Recently, palladium- and copper-based hydrogels have been used as catalysts for reduction, degradation, and decomposition reactions because they have significant features such as high mechanical strength, thermal stability, and high surface area. Herein, we summarize the progress achieved in this field, including the various methods for the synthesis of copper- and palladium-based hydrogel catalysts and their applications for environmental remediation. Moreover, palladium- and copper-based hydrogel catalysts, which have certain advantages, including high catalytic ability, reusability, easy work-up, and simple synthesis, are proposed as a new group of effective catalysts.


1. Introduction

Environmental pollution induced by industrial development and increase in population affects the life of living organisms by increasing the production of dangerous and toxic compounds such as organic dyes and food waste. Thus, this global problem has received much attention from researchers.1

Among the various pollutants, oil spills, heavy metals, herbicides, organic compounds, fertilizers, and sewage are considered the most important.2 Nitro compounds as one of the main pollutant groups are usually found in water. For instance, 4-nitrophenol, a key industrial pollutant, poses a threat to the ecosystem and human health when released into the environment owing to its high chemical stability and toxicity.3–5 Organic dyes are another group of compounds that pollute the environment. Synthetic organic dyes, which are poisonous, bio refractory, and carcinogenic pollutants, have become a significant source of water contamination. Because these dyes are chemically and physiologically persistent in the environment, the typical degradation processes are unsuitable for their separation.6–8 Furthermore, although there are various methods to remove pollutants from water and the environment, some of them are specific to certain uses. For example, coagulation, chemical reduction, reverse osmosis, and photodegradation procedures are used to remove nitro compounds and organic dyes.9–11 Scheme 1 provides the different approaches for the elimination of pollutants including absorption, chemical reactions, adsorption, photocatalysis, and filtration.12–23 Some of these methods are simple, inexpensive, with low energy consumption and high yield, and do not produce toxic products or require complex instruments; however, eco-friendly approaches are preferable.24 Chemical degradation is one of the most extensively utilized strategies for removing pollution from the environment, including ozone/UV, irradiation/H2O2 oxidation, photocatalytic degradation, supercritical water oxidation, and electrochemical methods.25


image file: d2ra03418b-s1.tif
Scheme 1 Different methods employed for environmental remediation.

Recently, there have been many attempts to build efficient and novel catalytic systems for environmental applications.26 Also, the use of homogeneous catalysts in industry is a challenge, and thus heterogeneous catalysts are more suitable because of their high stability, recyclability, and simple separation from the reaction media.27 Metal complexes are widely employed as catalysts for eco-friendly applications.28–30 Furthermore, the catalytic yield of metal complexes can be improved by increasing their active surface area by decreasing their size to the nanoscale (1–100 nm), causing fascinating properties and differences from bulk materials to emerge.31,32 Metal nanoparticles (MNPs) such as gold, palladium, silver, platinum, and copper have recently been explored as catalysts for removing organic dyes and nitro chemicals.33–37

2. Hydrogel

Hydrogels with three-dimensional networks are insoluble polymeric materials comprised of numerous hydrophilic groups, which can load a vast amount of water. It has been reported that hydrogels are versatile carriers of active compounds including enzymes, drugs, and nanoparticles. Recently, smart hydrogels with the ability to respond to stimulation from the external environment, such as additional chemicals, pH, temperature, and light, have received increasing attention.38–43 Thus, due to these attractive properties, these smart hydrogels can be potential materials with promising applications in drug delivery, chemical sensors, catalysts, and optics.44–50

2.1. Classification of hydrogels

Hydrogels can be divided into numerous types based on their preparation procedure, ionic charge, and physical structures, including homopolymer hydrogels, co-polymer hydrogels, multi-polymer hydrogels, and interpenetrating network hydrogels. Homo-polymer hydrogels are cross-linked networks of hydrophilic monomer units of one type, whereas co-polymer hydrogels, are made by cross linking chains made up of two co-monomer units, one of which must be hydrophilic for them to swell in water. Three or more comonomers are reacted together to form multi-polymer hydrogels. Finally, interpenetrating network hydrogels can be made in two ways, i.e., in a premade network and solution. The typical strategy is the polymerization of one monomer within various cross-linked hydrogel networks. The monomer polymerizes to form a polymer or a second cross-linked network that is intermeshed with the first network.

Hydrogels can also be categorized in a variety of ways. Ionic hydrogels with ionic charges on the backbone polymers are classed as follows: (A) uncharged, (B) anionic hydrogels (containing only negative charges), (C) cationic hydrogels (having only positive charges) and (D) ampholytic hydrogels (having both positive and negative charges). The net charge on these last gels can be harmful, positive, or neutral.

Hydrogels can be characterized as either (A) amorphous hydrogels (with covalent cross-links) or (B) semicrystalline hydrogels, depending on the physicochemical structural properties of their network (may or may not have covalent cross-links). The macromolecular chains in amorphous hydrogels are randomly organized. Alternatively, semicrystalline hydrogels (crystallites) consist of self-assembled regions of organized macromolecular chains.38–40

2.2. Preparation of hydrogels

Small multifunctional molecules such as monomers and oligomers are frequently combined, reacting to form a network structure, for the fabrication of covalently cross-linked hydrogels. Large polymer molecules can sometimes be cross-linked with small multifunctional molecules. The reaction of two chemical groups on two distinct molecules, which can be initiated by catalysts, photopolymerization or radiation cross-linking, can produce this type of cross-linking.

Free radical reactions are used various ways to make cross-linked hydrogels. One approach is a copolymerization-cross-linking reaction involving one or more monomers and one multifunctional monomer in small amounts. Two water-soluble polymers can be cross-linked using a similar mechanism involving the generation of free radicals on both polymer molecules, which combine to create a cross-link. These reactions are known as free radical polymerization or cross-linking reactions. Moreover, they can be triggered by the decomposition of peroxides or azo compounds, or by ionizing radiation or UV light. Ionizing radiation involves the use of gamma rays, X-rays, or electron beams to excite a polymer, causing free radical reactions to generate a cross-linked structure. These free radical reactions, which normally occur in the absence of oxygen or air (note: some polymers are damaged by radiation, especially in air), result in the rapid development of a three-dimensional network.38

Chemical cross-linking is another method that involves the reaction of a linear or branched polymer directly with a difunctional or multifunctional low-molecular-weight cross-linking agent. This agent frequently connects two higher-molecular-weight chains through its di- or multifunctional groups. There are several well-known processes for joining hydrophilic polymers together or with cross-linkers to generate hydrogels, including some newer techniques that are gaining traction, such as the Michael addition of dithiol compounds with divinyl compounds. A related method involves the cross-linking of a tiny bifunctional molecule with linear polymeric chains that have pendant or terminal reactive groups such as –OH, –NH2, NCO, and –COOH. Enzymes can also be used to cross-link natural polymers such as proteins in a similar way.

2.3. Formation mechanism of hydrogels

The different types of gelation mechanisms are summarized in Table 1. Gelation can take place either by physical cross-linking (physical gelation) or by chemical cross-linking (chemical gelation) and radiation cross-linking. Physical gels can be sub-categorized as strong physical gels and weak gels. Strong physical gels have strong physical bonds between their polymer chains and effectively permanent under a given set of experimental conditions. Weak physical gels have reversible links formed from temporary associations between their chains. Alternatively, chemical gelation involves the formation of covalent bonds and always results in a strong gel.51
Table 1 The general methods to produce physical and chemical gels
Physical cross-linking Chemical cross-linking Radiation cross-linking
Heating/cooling a polymer Chemical cross-linkers Aqueous-state radiation
Ionic interaction Grafting (chemical grafting and radiation grafting) Radiation in paste
Complex coacervation   Solid-state radiation (cross-linking in the solid state, natural polymers and synthetic/natural polymer blends)
H-bonding  
Maturation (heat-induced aggregation)  
Freeze-thawing    


2.4. The techniques for the characterization of hydrogels

To design hydrogels with the desired performance and structure, the determination and characterization of their network parameters are significant. Table 2 presents the techniques for the characterization of hydrogels including physico-chemical, structural and thermal characterization, and the tools for their physicochemical characterization such as infrared spectroscopy, atomic force microscopy, X-ray diffraction analysis, electron microscopy, and many other techniques are described to characterize their structure.52
Table 2 Techniques for the characterization of hydrogels
Physico-chemical characterization Structural characterization Thermal characterization
Solubility Transmission electron microscopy Differential scanning calorimetry
Swelling measurement Scanning electron microscopy Thermogravimetric analysis
Sol gel analysis Atomic force microscopy Dynamic mechanical thermal analysis
Rheology Dynamic force microscopy  
UV-Vis absorption spectroscopy Confocal microscopy  
Infrared spectroscopy    
Mass spectroscopy    
Nuclear magnetic resonance    
Light scattering    
X-ray diffraction analysis    


2.5. The reaction mechanism for the catalytic reduction/degradation of organic pollutants by hydrogels

Contaminants may be adsorbed on hydrogels through various methods including complexation, coordination, chelation, ion exchange, adsorption by physical forces, and ion trapping in their inter- and intra-fibrillar capillaries.53 These methods are always influenced by the properties of the hydrogels, the physical and chemical features of the specific type of pollutant being targeted, and the micro-environmental conditions (such as the contact solution pH and ionic strength).54 Hydrogels have the ability to remove pollutants through a variety of mechanisms or by combining the impacts of several different mechanisms, for example, the reduction of nitrophenols to aminophenols.

The analysis of the reaction mechanism is beneficial for the development of more effective catalysts. An example of a heterogeneous catalytic reaction is the catalytic reduction of nitrophenols using hydrogel-supported metal nanoparticles. The Eley–Rideal and Langmuir–Hinshelwood models can both be used to predict the course of a heterogeneous catalytic reaction. In the case of the Eley–Rideal model, only one of the educts will adhere to the surface of the catalyst and interact with the other educt in the solution. According to the Langmuir–Hinshelwood model, both educts adsorb on the surface of the catalyst and undergo reactions there. Accordingly, by conducting two sets of tests, namely, examining the dependence of the rate constant on the concentration of PNP and sodium borohydride, the appropriate reaction mechanism for the reduction of nitrophenols can be proposed.

In the Eley–Rideal model, the rate increases with an increase in the concentration of PNP, which is not adsorbed on the surface. Alternatively, the rate constant decreases with an increase in the concentration of PNP according to the Langmuir–Hinshelwood model.55,56

In the present review, we describe the advantages and properties of copper and palladium-based hydrogels. In addition, the role of copper- and palladium-based hydrogels in the development of novel, safe, and efficient catalysts is explained.

3. Catalytic applications of synthesized palladium- and copper-based hydrogels

Some hydrogels have a large number of polar functional groups such as hydroxyl and carboxyl groups, which can strongly chelate with transition metals (Au,57 Ag,58 Cu,59 Pd,60 Ni,61 Co,62 etc.), leading to the effective loading of metal ions/nanoparticles and preventing their aggregation. Furthermore, they can increase the ion/nanoparticle loading and number of catalytic sites for catalytic reactions. Also, some hydrogels are biocompatible and biodegradable, and thus can be applied as green, low-cost, and suitable candidates for the removal of pollutants from the environment.63,64 Generally, nanocomposite hydrogels can be prepared by immersing a hydrogel in aqueous solutions of the respective metal salts. Subsequently, the metal ions are converted to metal nanoparticles via reduction by a typical reducing agent, such as NaBH4. Nitro compounds, phenols, and heavy metals are the most common pollutants, and thus an overwhelming number of works has been focused on their destruction.65–67 Au-,64 Pt-,68 and Ag-based69 hydrogels are commonly used for the catalytic reduction of 4-nitrophenol (4-NP) despite their high cost. Alternatively, copper (Cu) is a good candidate to replace these expensive metals because it is an abundant element in the Earth's crust.70–75 Also, palladium nanoparticles (Pd NPs) are one of the most promising candidates for pollutant removal due to their outstanding catalytic activity. Thus, among the different transition metals, Pd and Cu nanoparticles or Cu and Pd-based complexes have been significantly used in catalytic applications, especially considering their exclusive features and low cost. Besides, Cu NPs and Pd NPs play critical roles in degrading organic pollutants.76–79 Recently, Pt- and Cu-based hydrogels have been regarded as active metals with excellent performances for different chemical transformations. Different synthetic (such as polyacrylamide and polyacrylic acid) and natural-based polymers (such as chitosan and cellulose) are used as supports. The purpose of the extensive research in this area is to develop low-cost approaches for the scalable production of biocompatible catalysts. In this section, we present the catalytic uses of these compounds in the degradation of environmental pollutants.

3.1. Reduction and degradation of environmental pollutants using Cu-based hydrogels

3.1.1. Synthetic polymer-based catalysts. In 2011, Sahiner et al. reported a facile technique for the synthesis of some polymers with various behaviors, which have a core/shell structure (acrylonitrile (AN) as the core and acrylic acid (AAc) and methacrylic acid (MA) as the shell monomers) through microemulsion polymerization using hexadecyltrimethylammonium bromide (CTAB) as the cationic surfactant (Scheme 2).80 The prepared nanoparticles exhibited significant potential for use in various fields; firstly, as drug targeting vehicles in the biological field. Secondly, for the removal of organic pollutants such as the toxic organic species 4-nitrophenol (4-NP) and herbicide (Paraquat) in eco-friendly uses, and thirdly as a model for preparing copper catalysts to reduce 4-NP to 4-aminophenol (4-AP). The created p(AN-co-AAc) and p(AN-co-MA) particles were further chemically modified to make them more hydrophilic by converting their nitrile groups to amidoxime groups. The amidoximated particles presented improved absorption capacity for metal and organic species and enhanced catalytic behaviors.
image file: d2ra03418b-s2.tif
Scheme 2 Preparation of Cu NPs inside p(AN-co-MA) and p(AN-co-AAc) particles.

In 2012, Ozay et al. reported a simple route for the preparation of Cu(0) nanoparticles inside p(2-acrylamido-2-methyl-1-propansulfonic acid) (p(AMPS)) hydrogel networks.81 The p(2-acrylamido-2-methyl-1-propansulfonic acid)–Cu composites were utilized for the reduction of 4-NP as a model reaction in the presence of excess NaBH4, exhibiting superior catalytic activity (Scheme 3). Even in the presence of a significantly lower quantity of catalyst (1.1 mg Cu(0)), the degradation rate of 4-NP by the p(2-acrylamido-2-methyl-1-propansulfonic acid)–Cu composites at a temperature close to room temperature (30 °C) led to a much higher value (0.1032 min−1) compared to Ni(0) and Co(0) catalysts synthesized under very similar conditions. At various activation energies and reaction temperatures, the kinetic parameters defined for 4-NP and 4-AP were 39.2 and 28.2 kJmol−1, respectively.


image file: d2ra03418b-s3.tif
Scheme 3 Schematic of the creation of Cu NPs inside p(AMPS) hydrogel network and their use for the reduction of 4-NP to 4-AP.

In 2013, Sahiner and Sagbas synthesized bulk poly(vinyl phosphonic acid) (p(VPA)) hydrogels through a photo-polymerization method utilizing polyethylene glycol diacrylate with various molecular weights as the cross-linker (Scheme 4).82 The prepared hydrogel was used as a template for the in situ synthesis of Ni, Co, and Cu nanoparticles. Initially, the equivalent ions from aqueous media were loaded in the poly(vinyl phosphonic acid) hydrogel complexes. Subsequently, the metal ions were reduced by sodium borohydride to the (p(VPA)) hydrogel matrices. They investigated the catalytic performance of p(VPA)–M (M: Co, Ni, Cu, etc.) for the reduction of 4-NP to 4-AP.


image file: d2ra03418b-s4.tif
Scheme 4 Bulk p(VPA) hydrogel crosslinking by p(EGDA).

In 2017, Guin and co-workers synthesized a thermally stable mixed elastic poly(acrylic acid–sodium-sulfonate-styrene graphene oxide) (poly(AAc–SSS–GO)) superabsorbent hydrogel via addition polymerization induced by free gamma radiation after crosslinking (Fig. 1).83 The composite material adsorbed 98 mg g−1 copper(II) from an aqueous solution of copper(II) at room temperature at pH 5 with the chemi-adsorption of copper(II) ions by its various functional groups with high energy. Through the chemical reduction of the pre-adsorbed copper(II) ions, Cu NPs with a size of 12 ± 8 nm were prepared in situ. They confirmed the excellent catalytic activity of Cu NP–poly(AAc–SSS–GO) for the reduction of 4-NP to 4-AP.


image file: d2ra03418b-f1.tif
Fig. 1 Preparation of the blended poly(AAc10–SSS15–GO0.01) composite hydrogel via a crosslinking process based on gamma radiation. This figure has been reproduced from ref. 83 with permission from John Wiley & Sons Inc. Copyright 5316550528241.

In 2017, Naseer and co-workers synthesized poly(acrylic acid-co-acrylamide) (p(AAc-co-AAm)) hydrogels, which were subsequently used as templates for the synthesis of Cu nanoparticles (Scheme 5).84 The Cu NPs were prepared via the in situ reduction of Cu(II) ions. The synthesized p(AAc-co-AAm)–Cu composites were utilized as catalysts for the reduction of 4-NP and as an adsorbent in aqueous media to remove malachite. Moreover, the p(AAc-co-AAm)–Cu catalyst could be reutilized 4 times with no decline in its catalytic activity.


image file: d2ra03418b-s5.tif
Scheme 5 Synthesis of p(AAc-co-AAm)–Cu.

In 2019, Li and colleagues synthesized some Cu2O/Cu/reduced graphene oxide @carbon nanomaterial (Cu2O/Cu/rGO@CN) photocatalysts via a new sodium alginate hydrogel technique (Fig. 2).85 Cu2O NPs (∼50 nm) were synthesized via a calcination method under the protection of a nitrogen atmosphere. Inevitably, Cu nanoparticles (∼6 nm) were found over the surface of Cu2O. Thus, a Cu2O/Cu heterostructure was formed, which was identified as a Schottky junction. Synchronously, graphene oxide was reduced in situ by sodium alginate during the synthetic procedure, and eventually its 3-D structure contributed to the formation of the hydrogel skeleton. The photocatalytic activity and adsorption capacity of Cu2O/Cu/rGO@CN improved considerably when it was modified by 3-D rGO. Subsequently, the synthesized Cu2O/Cu/rGO@CN was employed as an effective catalyst for the degradation of p-nitrochlorobenzene. The advantages of this method included facile synthesis, no need for complex instruments, and use of nontoxic raw materials.


image file: d2ra03418b-f2.tif
Fig. 2 (a) Formation of Cu2O/Cu/rGO@CN photocatalysts using sodium alginate hydrogel as a template. SEM images of (b) Cu2O/Cu@CN, (c) Cu2O/Cu/rGO@CN-4 and (d) rGO. This figure has been reproduced from ref. 85 with permission from Elsevier. Copyright 5316550031989.

Li et al. synthesized nanostructures of steady flexible electrospun hydrogel nanofibers comprised of poly(allylamine hydrochloride) and poly(acrylic acid).86 The metal NPs were formed by loading the hydrogel fibers with metal ions such as Ag and Cu and immersing them in solutions with metal ions after chemical reduction. Finally, the metal ions were absorbed from the medium by the hydrogel fibers. They presented a viscous environment for promoting the creation of alloy particles with a size of less than 25 nm. The reduction of methylene blue (MB) and 4-NP was carried out to examine and compare the catalytic activity of Ag–Cu bimetallic NPs and monometallic NPs (Scheme 6). It was found that the Ag–Cu bimetallic NPs presented desirable selectivity to reduce 4-NP and better catalytic activity for the degradation of MB.


image file: d2ra03418b-s6.tif
Scheme 6 Chemical reduction of 4-NP and MB.
3.1.2. Natural polymer-based catalysts. In 2016, Kamal and colleagues reported a simple method for the preparation of CS–CMM by adhering chitosan (CS) on a cellulose microfiber mat (CMM).87 The synthesized CS–CMM was utilized as a host for the preparation of Cu NPs. The metal ions encapsulated in the chitosan layer were reacted with NaBH4 for the synthesis of copper NPs loaded on CS–CMM (Cu/CS–CMM), followed by adsorbing copper(II) ions from an aqueous solution of CuSO4. The created Cu/CS–CMM was used as a catalyst for the reduction of the aromatic rings in 2-NP, 4-NP, and the organic cresyl blue dye (CB) (Scheme 7). The rate constants for the reduction of cresyl blue, 2-NP and 4-NP were 1.3 × 10−3 s−1, 1.2 × 10−3 s−1, and 2.1 × 10−3 s−1, respectively.
image file: d2ra03418b-s7.tif
Scheme 7 Degradation of 2-NP, 4-NP and CB by Cu/CS–CMM catalyst.

In the same year, a simple procedure was reported for the synthesis of Cu NPs in a chitosan/fluorescein (CS/F) hydrogel and film matrix by immersing the particular matrix in an aqueous medium comprised of a metallic salt precursor and reducing agent under an N2 atmosphere at 75 °C for 3 h (Scheme 8).88 The high catalytic yield of the prepared nanocomposites was demonstrated by reducing methylene blue using N2H4 as a reducing agent. According to the experimental results, the order of the degradation reaction of the MB–N2H4 system and catalytic activity were strongly dependent on the type of nanocomposite used.


image file: d2ra03418b-s8.tif
Scheme 8 (A) Preparation of CS/F–Cu film and hydrogel nanocomposites and (B) proposed mechanism for the reduction of Cu2+ to Cu0.

Sharma and colleagues synthesized a non-toxic biocompatible chitosan-based hydrogel nanocomposite with Cu NPs through in situ free radical polymerization utilizing acrylamide as the monomer. Chitosan was employed as the substrate, glutaraldehyde as the cross-linking agent, and Cu NPs and ammonium persulphate as the initiator (CH–PAmCu–NP catalyst).89 They employed the prepared hydrogel nanocomposite as a recyclable catalyst with high efficiency for the reduction of 4-NP to 4-AP (Scheme 9).


image file: d2ra03418b-s9.tif
Scheme 9 Reduction of 4-nitrophenol by CS-PAmCu-NP catalyst.

In 2018, Xiao and co-workers established an internal solidification procedure induced by a rigid template for the preparation of ultrafine MNPs immobilized on calcium alginate, polyethylene imine, and glutaraldehyde beads (Cu@GLA–PEI–CA). During the synthetic procedure, CaCO3 nanoparticles with bi-functional properties were applied as a hard skeleton for adjusting the structure of the catalyst and also as a subsidiary cross-linking agent to obtain hydrogel beads.90 Copper nanoparticles with a narrow size distribution of 2–3 nm were homogeneously supported on the surface of the GLA–PEI–CA beads with a well-developed hierarchical pore structure and high specific surface area. They observed that the Cu@GLA–PEI–CA beads demonstrated good catalytic activity for the reduction of 4-NP to 4-AP with sodium borohydride in aqueous solution (Scheme 10). Furthermore, this catalytic system could be simply separated and reused.


image file: d2ra03418b-s10.tif
Scheme 10 Reduction of 4-nitrophenol by Cu@GLA–PEI–CA.

In 2020, a new bionanocomposite of gelatin as a polypeptide hydrogel comprised of CuO nanomaterial (GL–CuO) was synthesized by Ahmad and co-workers.91 After their dispersal, the synthesized CuO NPs were cross-linked with the addition of a CH2O solution in the GL aqueous solution. The synthesized GL–CuO with an average size of about 60–80 nm showed durable and extraordinary activity for the reduction of Congo red (CR) and methyl orange (MO) in an eco-friendly medium (Scheme 11). This method had the advantages of simple catalyst synthesis, high efficiency, and easy work up.


image file: d2ra03418b-s11.tif
Scheme 11 Reducing methyl orange and Congo red by GL–CuO hydrogel nanocomposite catalyst.

In 2020, a simple route was reported by Kamal et al. for the preparation of carbon nitride (g-C3N4)-incorporated sodium alginate (Alg) beads. The Alg–CNBs (alginate carbon nitride beads) were synthesized by adding an Alg aqueous solution containing dispersed g-C3N4 to an aqueous solution containing calcium(II) ions (Fig. 3).92 Then, Alg–CNBs were further treated with various transition metal salt solutions. Subsequently, employing sodium borohydride as a potent reducing agent, the transition metal ions covering the surface of Alg–CNBs were reduced to the zero-valent state (M0) and utilized for the catalytic degradation of various organic dyes. Finally, the synthesized M0/Alg–CNBs (where M0 = Ag0, Fe0, Cu0, and Ni0) catalysts were examined for the reduction of 4-NP to determine the best one. Among them, it was found that Cu0/Alg–CNBs reduced 4-nitrophenol to 4-aminophenol efficiently. Also, Cu0/Alg–CNBs was further employed for the reduction of CR, MO, 2-NP and 2,6-DNP.


image file: d2ra03418b-f3.tif
Fig. 3 Preparation of Cu0/Alg–CNB catalytic beads and their catalytic utilization. This figure has been reproduced from ref. 92 with permission from Elsevier. Copyright 5316550823396.

In 2021, Xiao et al. successfully prepared an N-doped sandwich-like carbon layer coating on copper NPs supported on N-doped porous carbon (N–C/Cu/N–C) by the deposition of a polydopamine (PDA) layer over alginate-based hydrogel beads after high-temperature pyrolysis in N2.93 The results demonstrated that copper nanoparticles with a narrow size distribution in the range of 10–30 nm were immobilized uniformly over the porous carbon bead framework by protecting the outer N–C layer. Moreover, the mechanical strength was enhanced effectively by the layer extracted from PDA on the porous beads, therefore preventing the loss of copper NPs during the catalytic reaction. Subsequently, the performance of the prepared catalyst (N–C/Cu/N–C) was assessed for the reduction of 4-NP to 4-AP (Scheme 12).


image file: d2ra03418b-s12.tif
Scheme 12 Reduction of 4-NP in the presence of N–C/Cu/N–C.

In 2022, Andou and co-workers designed a facile method for the in situ preparation of a new three-dimensional photocatalyst comprised of Cu2O/TiO2 supported on a cellulose nanofiber (CNF)/reduced graphene hydrogel (rGH) through a simple freeze-drying and hydrothermal treatment.94 The three-dimensional macrostructure acted as an appropriate template for anchoring TiO2 and Cu2O, providing an effective electron transport path to improve the photocatalytic activity. According to their findings, TiO2 and Cu2O were homogeneously incorporated in the aerogel template, resulting in structures with a larger surface area in the uncovered active sites. In comparison to the bare rGH, Cu2O/CNF/rGH, CNF/rGH, and TiO2/CNF/rGH, Cu2O/TiO2/CNF/rGH presented enhanced activity for the removal of methyl orange (Scheme 13).


image file: d2ra03418b-s13.tif
Scheme 13 Proposed degradation path for MO utilizing Cu2O/TiO2/CNF/rGH.

The reports on the use of Cu-based hydrogels for environmental remediation are presented in Table 3.

Table 3 The catalytic elimination of inorganic/organic contaminants utilizing Cu-based hydrogels
Cu-based hydrogels Size of Cu nanoparticles Application Catalytic activity/benefits Year Ref.
a Discussed in the review.
Cu NPs inside p(AN-co-MA) and p(AN-co-AAc) particles Reduction of 4-NP to 4-AP Multipurpose 2011 80a
p(AMPS)–Cu 30–35   High conductivity 2012 81a
p (VPA)–M   Potential applications in bone mineralization 2013 82a
Cu NW–Ag (CuAg-3) 100   Superior catalytic activity 2014 96
DNA–Cu-hydrogel 2–3   Potential for green organic synthesis in aqueous media and for biomedical in vivo applications 2016 100
      Clean synthesis of the catalyst    
Cu NP–poly(AAc–SSS–GO) 12   Economically viable 2017 83a
CH/f-Cu hydrogel   Utilized 4 times with no loss in catalytic activity 2017 88a
p(AAc-co-AAm)–Cu   Separable with a magnetic field 2017 84a
CH–PAmCu–NPs   Good chelation with Cu NPs    
Multi-functional 2017 89a
Selective adsorption and recycling copper element for waste disposal industry    
Cu NPs-loaded SPI/PEI 10–80   Retained 70% activity after being stored for 30 days 2017 101
WSC-g-PAA/PVA–Cu 20–30   Beyond 90% conversion in 10 cycles 2017 102
      Tunable structure and properties    
Cu@GLA–PEI–CA beads 2–3   Cost-effective and biocompatible 2018 90a
Cu NP-impregnated cellulose 4–16   Scalable reduction and reusable for 6 times 2019 112
Chitosan–Cu   Enhanced mechanical strength and stability 2020 110
Alg–CNB beads     2020 92a
N–C/Cu/N–C 10–30     2021 93a
Cu/CHCMM 48 2-NP, 4-NP and cresyl blue Economical and environmentally friendly 2016 87a
Cu2O/Cu/rGO@CN photocatalysts ∼6 Degradation of p-nitrochlorobenzene Suitable chemical stability during recycling 2019 86a
GL–CuO 83 Reduction of Congo red and methyl orange Good catalytic reduction properties toward different environmental pollutants and easy separation through filtration 2020 91a
p(EP-g-AA)–Cu 100 Reduction of MB and 4-NP Maintained 80–90% catalytic activity after 5 times use and 30 days storage 2018 103
poly(allylamine hydrochloride) and poly(acrylic acid) nanofibers <20   Simplicity, controllability, and versatility 2021 86a
Cu(II)–poly(EGDE–DA) Reduction of MO Oxidation of pharmaceutical product (epinephrine) 2014 97
Cu2O/TiO2/CNF/rGH 4.3–6.3   Enhanced photoactivity 2021 94a
PEI–Cu Reduction of 4-NP and 2-NP Very fast reduction times of 1.2 and 0.67 min to 4-AP and 2-AP, respectively 2015 99
Amid–p(Mac–Cu–AN)–M 10–50 Reduction of 4-NP, 2-NP, MO, MB, eosin Y Individual and simultaneous degradation of cationic/anionic dyes double triple compounds 2016 95
Amid–p(AAm)–Cu Reduction of 4-NP, 2-NP, MB, eosin Y Fast response to external stimuli, good mechanical strength, and durability 2015 98
WSC-g-PAA/PVA–Cu 20–50 Degradation of chloramphenicol Great potential for future applications for heavy metal recovery and reuse 2018 104
m-WSC/FP–Cu 2–5 Reduction of 2-aminobenzoic acid Maintained 99.02% over 5 cycles utilization and 90.60% after 30 days storage 2019 105
AG–CuO 15–92 Reduction of 4-NP, 2-NP, 2,6-DNP Easy recyclability and high catalytic activity 2019 106
BSA–Cu Reduction of 4-NP, CR, MB Robust hydrogel incorporating metal ions 2020 107
Fe3O4–CuO@C Reduction of MB Enhanced photocatalytic activity 2020 108
Cu/MC 25 Reduction of 4-NP and rhodamine B Magnetic responsiveness to an external magnetic field 2020 109
CuS@CNs 5 2,4-Dichlorphenol Improved the separation of photogenerated charge carriers 2021 111


3.2. Reduction and degradation of environmental pollutants using Pd-based hydrogels

3.2.1. Synthetic polymer-based catalysts. In 2014, a facile technique was reported by Wang et al. for the synthesis of a novel bifunctional nanocatalyst composed of 3D macroscopic carbon nanotube (CNT)–graphene hydrogel (GH) supported palladium NPs (Pd–CNT–GH) and its application for the catalytic reduction of 4-NP to 4-AP (Scheme 14).113 A facile and one-pot self-assembly method was used to synthesize three-dimensional Pd–CNT–GH via the hydrothermal treatment of a precursor mixed aqueous solution containing graphene oxide (GO), PdCl42−, and CNT. Under suitable conditions, the spontaneous redox reaction between the precursor CNT–GO and PdCl42− occurs simultaneously with the self-assembly of macroscopic CNT–GH, resulting in the formation of three-dimensional Pd–CNT–GH. The catalytic process was accomplished within about 30 s utilizing this catalyst at room temperature. Furthermore, the catalyst could be reused 20 times without considerable loss in its catalytic activity.
image file: d2ra03418b-s14.tif
Scheme 14 Reduction of 4-NP in the presence of 3D Pd–CNT–GH.113

In 2017, Shen and colleagues designed and synthesized new three-dimensional graphene oxide nanosheets reduced by molybdenum disulfide and supported by Pd (Pd/MoS2–rGO) through a convenient one pot self-assembly process (Fig. 4).114 The restacking of rGO nanosheets was prevented by the presence of MoS2 and their specific surface area increased. Moreover, a further transport platform was also afforded for palladium NPs to enhance their catalytic features. Excellent catalytic activity for the reduction of 4-NP to 4-AP was demonstrated by the synthesized Pd/MoS2–rGO NPs in water in the presence of borohydride.


image file: d2ra03418b-f4.tif
Fig. 4 Procedure for the preparation of the 3D Pd/MoS2–rGO composite hydrogel. This figure has been reproduced from ref. 114 with permission from Elsevier. Copyright 5316560270510.

In the same year, Wu and co-workers utilized a two-step technique to synthesize a reduced graphene oxide nanosheet/magnetite–palladium (rGSs/Fe3O4–Pd) aerogel with superior recyclability and catalytic performance.115 In the first step, graphene oxide nanosheet hydrogels were prepared via the self-assembly of GSs during the hydrothermal procedure. Meanwhile, the palladium nanoparticles (NPs) and hematite (α-Fe2O3) were synthesized and anchored on the hydrogel surface. Subsequently, upon heat-treatment, the graphene oxide nanosheets were reduced to rGSs, whereas the nonmagnetic α-Fe2O3 NPs changed to magnetite Fe3O4 nanoparticles. The as-achieved rGSs/Fe3O4–Pd aerogel possessed a 3D interconnected porous hierarchical architecture rich in macro-pores and mesopores. This structure was appropriate for catalysis because it not only enhanced the transport and mass diffusion but also simply exposed the catalytic palladium nanoparticles to the substrates. The potential catalytic performance of the synthesized Pd NPs was demonstrated for the reduction of 4-nitrophenol (Scheme 15).


image file: d2ra03418b-s15.tif
Scheme 15 Reduction of 4-NP in the presence of rGSs/Fe3O4–Pd.

In 2018, Yao and co-workers synthesized a nanobead-based polypyrrole (PPy) hydrogel via a reactive-template technique in one step (Scheme 16).116 Fe3O4 NPs were chosen as the reactive template and used as the oxidant for initiating the polymerization of the pyrrole monomer and also guiding the growth of the polymer chains. Given that the PPy hydrogels were formed by the cross-linking of the individual nanobeads, they bridged the macro-dimension and nano-dimension; thus, a three-dimensional hierarchical porous structure was formed. The PPy hydrogels possessed a large surface area and several interconnected channels, and by taking advantage of their structural merits, they acted as superior candidates for adsorbing rhodamine B. The PPy hydrogels for loading Pd nanoclusters could serve as a support and used as adsorbents.


image file: d2ra03418b-s16.tif
Scheme 16 Possible mechanism for the formation of PPy hydrogel.

In the same year, Haraguchi and Varade described a novel approach to prepare discrete monometallic (silver, gold, and palladium) and bimetallic (Au–Pd, Pt–Pd) nanoparticles using a nanocomposite hydrogel comprised of a polymer–clay network (Fig. 5).117 Thermo-responsive nanocomposite hydrogels (NC gels) were prepared via the in situ free radical polymerization of N-isopropylacrylamide in the presence of clay (synthetic hectorite) nanosheets (CNSs). Given that the CNSs have robust affinities for metal ions, the concentration of the surrounding metal precursors was balanced and non-aggregated and well dispersed. Spherical bimetallic and monometallic NPs were obtained by reducing metal ions using ascorbic acid in NC gels, which were strongly immobilized in the polymer–clay network. The resultant hybrid NP–NC gels, which contained monometallic or bimetallic nanoparticles, exhibited high catalytic activities for the reduction of NP to AP.


image file: d2ra03418b-f5.tif
Fig. 5 Schematic showing the formation of the hybrid NP–NC gel: (a) metal ions penetrate the NC gel, (b) metal ions interact with the silanol groups on the clay surface, and (c) metal NPs are formed by the ascorbic-acid reduction of ions, which are subsequently trapped near the clay surface.

In 2020, PEI–Pd and PEI–Ag composite hydrogels were successfully synthesized by Jiao et al.118 Furthermore, the prepared PEI–Pd and PEI–Ag showed extraordinary and durable activity for the reduction of 4-nitrophenol to 4-aminophenol in an environmentally friendly medium (Fig. 6). This process has many advantages such as the high effectiveness of the composite hydrogels, simple process, and easy workup.


image file: d2ra03418b-f6.tif
Fig. 6 Preparation and catalytic application of PEI–Ag/PEI–Pd composites and PEI hydrogel.

Lee and co-workers described a simple technique for immobilizing palladium NPs with high physicochemical stability in uniform hydrogel microparticles. Simultaneously, it resulted in the high physicochemical stability and dispersion of the immobilized palladium nanoparticles inside the hydrogel microparticles. The PVP (polyvinylpyrrolidone)-stabilized palladium nanoparticles were synthesized via a microwave-assisted method. The palladium nanoparticles were immobilized utilizing a micro-molding approach in the UV enhanced hydrogel precursor.119 The immobilized palladium nanoparticles were physically stabilized because they were trapped in the porous hydrogel framework. Meanwhile, they were chemically stabilized by the H-bonding induced among the PEGDA (polyethylene glycol diacrylate) hydrogel matrix, PVP, and PEG (Fig. 7). Also, the stability of the immobilized palladium nanoparticles against oxidation was confirmed by the formed hydrogen bonding. They investigated the catalytic performance of the synthesized Pd NPs for the reduction of 4-NP to 4-AP.


image file: d2ra03418b-f7.tif
Fig. 7 Postulated stabilization mechanism for Pd NPs immobilized on hydrogels. This figure has been reproduced from ref. 119 with permission from Elsevier. Copyright 5318331364158.

In the same year, Wang and colleagues synthesized Pd-immobilized rGO and black phosphorus nanosheet-based (rGO–BP–Pd) hydrogels via a facile self-assembly method.120 Precisely, GO was reduced to the rGO hydrogel by addition of L-ascorbic acid. Furthermore, the rGO hydrogel was utilized as a skeleton for loading palladium nanoparticles and the catalytic features of the obtained template hydrogels were investigated (Fig. 8). They also investigated the effects of the hydrogels prepared by adding various quantities of black phosphorus nanosheets (BPNS) on the catalytic reduction of 4-nitrophenol. The outcomes demonstrated that increasing the quantity of BPNS could improve the catalytic activity owing to the interaction between Pd NPs and BPNS.


image file: d2ra03418b-f8.tif
Fig. 8 Catalytic performances and process for the synthesis of rGO-based composite hydrogels. This figure has been reproduced from ref. 120 with permission from Elsevier. Copyright 5318330689362.
3.2.2. Natural polymer-based catalysts. In 2019, Xie and colleagues prepared a novel cellulosic hydrogel anchoring a 1,1,3,3-tetramethyl guanidinium–cellulose-based ionic liquid moiety (GC) by simply adding bicyclic anhydrides to the recently established 1,1,3,3-tetramethyl guanidine/DMSO/CO2 solvent system for cellulose (Scheme 17).63 The amount and presence of bicyclic anhydrides as a cross-linker were demonstrated to be a key factor for the formation of the hydrogel. Then, the palladium NPs@GC structure was formed by reducing the absorbed palladium(II) ions on GC using sodium borohydride as a reducing agent. The CS provided decent immobilization capacity for palladium nanoparticles, and the obtained Pd NPs@GC presented good reusability and high catalytic performance for the degradation of 4-NP in H2O.
image file: d2ra03418b-s17.tif
Scheme 17 Synthesis of cellulosic hydrogel anchoring 1,1,3,3-tetramethylguanidine poly-ionic liquid moiety.

In the same year, Park and co-workers assessed the catalytic efficiency of alginate hydrogels as a support for catalysts.121 An alginate solution comprised of magnetite NPs and Pd was continuously dropped via a tapered glass capillary utilizing a syringe pump, and the created alginate droplets were gathered in an aqueous solution containing divalent cations. The created Pd NP/MNP-embedded alginate particles were considered as the catalyst support, catalyst, and magnetic stirrer for the reduction of 4-NP (Scheme 18). It was found that the catalytic reactions depended on the magnetic stirring of the alginate hydrogels.


image file: d2ra03418b-s18.tif
Scheme 18 Reduction of 4-NP to 4-AP applying Pd NP-embedded alginate hydrogel.

In 2020, Jiao and co-workers synthesized a series of palladium NP-incorporated chitosan (CS)-based composite hydrogels by adding carbon nanotubes, graphene oxide, and layered double hydrotalcites (LDHs) to a CS hydrogel with glutaraldehyde as the cross-linking agent and in situ growth of palladium nanoparticles.122 The catalytic reduction of 4-nitrophenol was carried out to evaluate the performance of the chitosan-based composite hydrogel (Scheme 19). The prepared chitosan-based composite hydrogel catalysts presented a higher catalytic performance in comparison with the chitosan hydrogel catalyst-doped palladium nanoparticles (CS-PdNPs) and recoverability for reducing 4-NP in aqueous media. Furthermore, the chitosan-based composite hydrogel catalysts comprised of CNTs possessed the highest catalytic activity among the three types of chitosan-based composite hydrogel catalysts because the particle size of the palladium nanoparticles was the smallest.


image file: d2ra03418b-s19.tif
Scheme 19 Reduction of 4-NP to 4-AP applying Pd NPs. This figure has been reproduced from ref. 122 with permission from Elsevier. Copyright 5316560975161.

In 2021, An and co-workers fabricated N-doped mesoporous carbon modified by highly dispersion nickel NPs (Ni–N@C) through the pyrolysis of Ni2+/histidine cross-linked alginate hydrogels.123 Next, the obtained Ni–N@C nano-catalyst was treated using a solution of palladium(II). Small amounts of the deposited palladium nanoparticles were found on the surface of nickel by the reduction of nickel to obtain a higher distribution of the valuable metal material (Fig. 9). The catalyst presented a good performance, where the pollutants were completely degraded in the presence of highly concentration 4-NP within a very short time. It was shown that the pre-embedded nickel nanoparticles could not only enhance the activity of palladium nanoparticles but also endow simple separation features to the catalyst. Also, this catalyst maintained favorable catalytic activity even after 5 reaction cycles.


image file: d2ra03418b-f9.tif
Fig. 9 Synthesis of Pd Ni–N@C. This figure has been reproduced from ref. 123 with permission from Elsevier. Copyright 5357550496279.

The reports on the use of Pd-based hydrogels for environmental remediation are presented in Table 4.

Table 4 The catalytic elimination of inorganic/organic contaminants utilizing Pd-based hydrogels
Pd-based hydrogels Size of Cu nanoparticles Application Catalytic activity/benefits Year Ref.
a Discussed in the review.
Pd–CNT–GH 2–3 Reduction of 4-NP to 4-AP Catalyzation over about 30 s at room temperature 2014 113a
Pd/MoS2–rGO 5   Catalyzation within 30 s and 20 successive cycles without obvious deactivation 2017 114a
rGSs/Fe3O4–Pd   Magnetically separable 2017 115a
NP–NC gels   Environmentally friendly catalyst 2018 117a
Pd NPs@GC 2–4   Good immobilization capacity of Pd NPs 2019 63a
Pd NPs/MNP-embedded alginate particles   At least 10 consecutive catalytic reactions 2019 121a
PEI–Pd   Easily separated 2020 118a
Pd-immobilized on PEGDA-based hydrogel 2–4   Enhanced stability 2020 119a
CH–PdNPs   High stability and easy recyclability 2020 122a
rGO–BP–Pd 5–10   Enhanced catalytic performance 2020 120a
Ni–N@C   Favorable catalytic activity for 5 reaction cycles 2021 123
PPy/Pd hydrogel Adsorption of rhodamine B Good TOF rate constant 2017 116a
Pd_TMV_PEG 2–3 Reduction of Cr(VI) Robust, and scalable synthesis 2013 124
Pd–TMV nano-biocomplexes 1–2   Retaining catalytic activity for at least 3 days 2019 126
Pd–GHJ 10 Reduction of nitroarenes Use of ammonia borane as a hydrogen storage material for the first time 2017 125
MarGO–Pd 30 Reduction of 4-NP, CR, MB Suitable for several biochemical applications 2020 127
Pd NPs@CNCC–AHB 9 Reduction of MB Suitable catalyst for C–C coupling along with MB reduction 2020 128


4. Conclusion and future prospects

The industrial development and increasing population have resulted in an increase in the release of pollutants in the environment, threatening the health of all living organisms. These pollutants can be removed from the environment in different ways; however, it is vital to find an effective way to remove these pollutants. Catalysts are compounds that can degrade pollutants by different mechanisms, which are safe and effective for this aim. The high diversity of compounds possessing catalytic activities and possibilities of improving their catalytic yield by decreasing their size or increasing their surface area have resulted in significant progress in this field. However, although homogeneous catalysts have various advantages, including high selectivity and reactivity, they are not generally commercialized owing their difficult separation, leading to fouling and corrosion. To create heterogeneous catalysts, homogeneous catalysts can be loaded on different supports, among which hydrogels are suitable owing to their numerous advantages including high thermal stability, high mechanical strength, easy modification, and high surface area.

By applying catalytic reduction or degradation reactions, different organic pollutants such as including nitro compounds, organic dyes, hexavalent chromium, and medicinal compounds can be removed. As stated above, copper and palladium-based hydrogels can be used as an effective catalyst to remove these contaminants. Copper and palladium with fascinating features such as antibacterial, antifungal, and antimicrobial activities can be used for environmental remediation. Herein, we summarized the recent progress on the use of copper- and palladium-based hydrogels for environmental remediation. Despite the many studies on copper- and palladium-based hydrogels, there are still some challenges for their future application. In the future, hydrogel-based carriers can be an excellent candidate for the successful administration of drugs at the desired rate and site in the body. With the creation of new hydrogels with various structural properties and hydrophobicity/hydrophilicity ratios, specific release rates and dissolution profiles may be achieved. Accordingly, these methods can be used to treat pathological illnesses such as diabetes and even cancer and enhance the delivery of more sensitive molecules. More advancements are anticipated, specifically, in the delivery of therapeutic proteins and peptides via hydrogels.

• Despite the progress in environmental remediation, some compounds have to be still eliminated from the environment.

• Use of different natural-based polymers such as lignin, chitin, chitosan, gum, cellulose, starch, and gelatin as supports.

• Use of low-cost approaches for the scalable production of catalysts.

• Manufacture of industrial-scale copper- and palladium-based hydrogels can be useful to eliminate environmental pollutants.

• Magnetic copper- and palladium-based hydrogels can be upgraded to exhibit high recyclability and catalytic activity to remove environmental pollutants.

Abbreviations

MNPsMetal nanoparticles
4-NP4-Nitrophenol
PNPpara-Nitrophenol
CuCopper
Pd NPsPalladium nanoparticles
ANAcrylonitrile
AAcAcrylic acid
MAMethacrylic acid
CTABHexadecyl trimethyl ammonium bromide
4-AP4-Aminophenol
(P(AMPS))P(2-acrylamido-2-methyl-1-propansulfonic acid)
P(VPA)P(vinyl phosphonic acid)
P(EGDA)Polyethylene glycol diacrylate
(Poly(AAc–SSS–GO))Poly(acrylic acid–sodium-sulfonate-styrene graphene oxide)
(P(AAc-co-AAm))Poly(acrylic acid-co-acrylamide)
Cu2O/Cu/rGO@CNCu2O/Cu/reduced graphene oxide@carbon nanomaterial
MBMethylene blue
CSChitosan
CMMCellulose microfiber mat
CBCresyl blue dye
CS/Fchitosan/fluorescein
CH–PAmCu–NPsCu NPs and ammonium persulphate
Cu@GLA–PEI–CACalcium alginate, polyethylene imine, and glutaraldehyde
GLGelatin
CRCongo red
MOMethyl orange
g-C3N4Carbon nitride
AlgSodium alginate
Alg–CNBsAlginate carbon nitride beads
N–C/Cu/N–CN-doped porous carbon
PDAPolydopamine
CNFCellulose nanofiber
rGHReduced graphene hydrogel
CNTCarbon nanotube
GHGraphene hydrogel
GOGraphene oxide
Pd/MoS2-rGOMolybdenum disulfide and supported by Pd
rGSs/Fe3O4–PdReduced graphene oxide nanosheets/magnetite-palladium
NPsNanoparticles
α-Fe2O3Hematite
PPyPolypyrrole
NC gelsNanocomposite hydrogels
CNSsClay (synthetic hectorite) nanosheets
PVPPolyvinylpyrrolidone
PEGDAPolyethylene glycol diacrylate
rGO–BP–PdPd-immobilized rGO and black phosphorus nanosheets
BPNSBlack phosphorus nanosheets
GCGuanidinium-cellulose
LDHsLayered double hydrotalcites
CS–Pd NPsChitosan hydrogel catalyst doped palladium nanoparticles
Ni–N@C NPsCarbon modified by highly dispersion nickel nanoparticles

Conflicts of interest

There are no conflicts to declare.

Acknowledgements

Here, the partial support of the Research Council of Islamic Azad University, Qom Branch are acknowledged.

References

  1. F. Han, V. S. R. Kambala, M. Srinivasan, D. Rajarathnam and R. Naidu, Appl. Catal., A, 2009, 359, 25–40 CrossRef CAS.
  2. J. A. Tkaczyk, K. Mitrowska and A. Posyniak, Sci. Total Environ., 2020, 717, 137222 CrossRef PubMed.
  3. O. P. Zhao, X. Feng, D. Huang, G. Yang and D. Astruc, Coord. Chem. Rev., 2015, 287, 114–136 CrossRef.
  4. M. Zhang, X. Su, L. Ma, A. Khan, L. Wang, J. Wang, A. Maloletnev and C. Yang, J. Hazard. Mater., 2021, 403, 123870 CrossRef CAS PubMed.
  5. M. A. Zakaria, A. Menazea, A. M. Mostafa and E. A. Al-Ashkar, Surf. Interfaces, 2020, 19, 100438 CrossRef CAS.
  6. S. Xia, L. Zhang, G. Pan, P. Qian and Z. Ni, Phys. Chem. Chem. Phys., 2015, 17, 5345–5351 RSC.
  7. V. C. Lops, A. Ancona, K. Di Cesare, B. Dumontel, N. Garino, G. Canavese, S. Hérnandez and V. Cauda, Appl. Catal., B, 2019, 243, 629–640 CrossRef PubMed.
  8. V. A. Sakkas, M. A. Islam, C. Stalikas and T. A. Albanis, J. Hazard. Mater., 2010, 175, 33–44 CrossRef CAS PubMed.
  9. D. Ayodhya and G. Veerabhadram, Mater. Today Energy, 2018, 9, 83–113 CrossRef.
  10. S. Nataraj, K. Hosamani and T. Aminabhavi, Desalination, 2009, 249, 12–17 CrossRef CAS.
  11. B. Shi, G. Li, D. Wang, C. Feng and H. Tang, J. Hazard. Mater., 2007, 143, 567–574 CrossRef CAS PubMed.
  12. J. Ma, X. Tang, Y. He, Y. Fan and J. Chen, Desalination, 2020, 480, 114328 CrossRef CAS.
  13. I. C. da Costa Soares, Á. R. L. da Silva, E. C. M. de Moura Santos, E. V. dos Santos, D. R. da Silva and C. A. Martínez-Huitle, J. Solid State Electrochem., 2020, 24, 3245–3256 CrossRef CAS.
  14. K. B. Tan, M. Vakili, B. A. Horri, P. E. Poh, A. Z. Abdullah and B. Salamatinia, Sep. Purif. Technol., 2015, 150, 229–242 CrossRef CAS.
  15. G. Liu, M. R. Abukhadra, A. M. El-Sherbeeny, A. M. Mostafa and M. A. Elmeligy, J. Environ. Manage., 2020, 254, 109799 CrossRef CAS PubMed.
  16. M. Wawrzkiewicz, M. Wiśniewska, A. Wołowicz, V. M. Gun'ko and V. I. Zarko, Microporous Mesoporous Mater., 2017, 250, 128–147 CrossRef CAS.
  17. G. Liao, Y. Gong, L. Zhong, J. Fang, L. Zhang, Z. Xu, H. Gao and B. Fan, Nano Res., 2019, 12, 2407–2436 CrossRef CAS.
  18. G. Liao, J. Fang, Q. Li, S. Li, Z. Xu and B. Fang, Nanoscale, 2019, 11, 7062–7096 RSC.
  19. G. Liao, J. Chen, W. Zeng, C. Yu, C. Yi and Z. Xu, J. Phys. Chem. C, 2016, 45, 25935–25944 CrossRef.
  20. G. Liao, Q. Li, W. Zhao, Q. Pang, H. Gao and Z. Xu, Appl. Catal., A, 2018, 549, 102–111 CrossRef CAS.
  21. G. Liao, W. Zhao, Q. Li, Q. Pang and Z. Xu, Chem. Lett., 2017, 46, 1631–1634 CrossRef CAS.
  22. Q. Wu, J. Wang, Z. Wang, Y. Xu, Z. Xing, X. Zhang, Y. Guan, G. Liao and X. Li, J. Mater. Chem. A, 2020, 8, 13685–13693 RSC.
  23. H. Zhu, S. Cai, G. Liao, Z. F. Gao, X. Min, Y. Huang, S. Jin and X. Fan, ACS Catal., 2021, 11(24), 14751–14771 CrossRef CAS.
  24. R. Sankar, P. Manikandan, V. Malarvizhi, T. Fathima, K. S. Shivashangari and V. Ravikumar, Spectrochim. Acta, Part A, 2014, 121, 746–750 CrossRef CAS PubMed.
  25. M. M. Khin, A. S. Nair, V. J. Babu, R. Murugan and S. Ramakrishna, Energy Environ. Sci., 2012, 5, 8075–8109 RSC.
  26. W.-J. Liu, T.-T. Qian and H. Jiang, Chem. Eng. J., 2014, 236, 448–463 CrossRef CAS.
  27. M. D. Argyle and C. H. Bartholomew, Catalysts, 2015, 5, 145–269 CrossRef CAS.
  28. M. Nasrollahzadeh, M. Sajjadi, J. Dadashi and H. Ghafuri, Adv. Colloid Interface Sci., 2020, 276, 102103 CrossRef CAS PubMed.
  29. M. Chen, P. Liu, C. Wang, W. Ren and G. Diao, New J. Chem., 2014, 38, 4566–4573 RSC.
  30. K. M. Lee, C. W. Lai, K. S. Ngai and J. C. Juan, Water Res., 2016, 88, 428–448 CrossRef CAS PubMed.
  31. J. K. H. Wong, H. K. Tan, S. Y. Lau, P.-S. Yap and M. K. Danquah, J. Environ. Chem. Eng., 2019, 7, 103261 CrossRef CAS.
  32. R. K. Ibrahim, M. Hayyan, M. A. AlSaadi, A. Hayyan and S. Ibrahim, Environ. Sci. Pollut. Res., 2016, 23, 13754–13788 CrossRef CAS PubMed.
  33. N. Y. Nadaf and S. S. Kanase, Arabian J. Chem., 2019, 12, 4806–4814 CrossRef CAS.
  34. H. Dabhane, S. Chatur, G. Jadhav, P. Tambade and V. Medhane, Environ. Chem. Ecotoxicol., 2021, 3, 160–171 CrossRef CAS.
  35. A. J. Kora and L. Rastogi, Arabian J. Chem., 2018, 11, 1097–1106 CrossRef CAS.
  36. Z. Gong, T. Ma and F. Liang, J. Alloys Compd., 2021, 873, 159802 CrossRef CAS.
  37. S. Joseph and B. Mathew, J. Mol. Liq., 2015, 204, 184–191 CrossRef CAS.
  38. N. A. Peppas, J. Z. Hilt, A. Khademhosseini and R. Langer, Adv. Mater., 2006, 18, 1345–1360 CrossRef CAS.
  39. S. Wang, T. Vincent, C. Faur and E. Guibal, Appl. Sci., 2018, 8, 264–281 CrossRef.
  40. B. V. Slaughter, S. S. Khurshid, O. Z. Fisher, A. Khademhosseini and N. A. Peppas, Adv. Mater., 2009, 21, 3307–3329 CrossRef CAS PubMed.
  41. G.-J. Jiao, J. Ma, Y. Li, D. Jin, J. Zhou and R. Sun, J. Hazard. Mater., 2022, 421, 126722 CrossRef CAS PubMed.
  42. Q. Luo, H. Yuan, M. Zhang, P. Jiang, M. Liu, D. Xu, X. Guo and Y. Wu, J. Hazard. Mater., 2021, 401, 123432 CrossRef CAS PubMed.
  43. D. Dou, D. Wei, X. Guan, Z. Liang, L. Lan, X. Lan, P. Liu, H. Mo and P. Lan, J. Hazard. Mater., 2022, 423, 127137 CrossRef CAS PubMed.
  44. Y. Meng, J. Yin, T. Jiao, J. Bai, L. Zhang, J. Su, S. Liu, Z. Bai, M. Cao and Q. Peng, J. Mol. Liq., 2020, 298, 112010 CrossRef CAS.
  45. M. Liao, H. Qin, W. Guo, P. Gao and H. Xiao, Ceram. Int., 2021, 47, 33667–33677 CrossRef CAS.
  46. Y. Zhang, S. Liu, T. Li, L. Zhang, U. Azhar, J. Ma, C. Zhai, C. Zong and S. Zhang, Carbohydr. Polym., 2020, 236, 116021 CrossRef CAS PubMed.
  47. A. L. Mohamed, A. A. Soliman, E. A. Ali, N. Y. Abou-Zeid and A. A. Nada, Int. J. Biol. Macromol., 2020, 163, 888–897 CrossRef CAS PubMed.
  48. J. F. Souza, G. P. Costa, R. Luque, D. Alves and A. R. Fajardo, Catal. Sci. Technol., 2019, 9, 136–145 RSC.
  49. H. Ozay, P. Ilgin and O. Ozay, Int. J. Hydrogen Energy, 2020, 45, 17613–17624 CrossRef CAS.
  50. D. George, P. U. Maheswari and K. M. S. Begum, Int. J. Biol. Macromol., 2019, 132, 784–794 CrossRef CAS PubMed.
  51. S. K. H Gulrez, S. Al-Assaf and G. O Phillips, Prog. Mol. Environ. Bioeng.: Anal. Model. Technol. Appl., 2011, 117150, 117–150 Search PubMed.
  52. E. M. Ahmed, J. Adv. Res., 2015, 6, 105–121 CrossRef CAS PubMed.
  53. D. Zhou, L. Zhang and S. Guo, Water Res., 2005, 39, 3755–3762 CrossRef CAS PubMed.
  54. B. Benguella and H. Benaissa, Water Res., 2002, 36, 2463–2474 CrossRef CAS PubMed.
  55. S. Wunder, F. Polzer, Y. Lu, Y. Mei and M. Ballauff, J. Phys. Chem. C, 2010, 114, 8814–8820 CrossRef CAS.
  56. P. Herves, M. Perez-Lorenzo, L. M. Liz-Marzan, J. Dzubiella, Y. Lu and M. Ballauff, Chem. Soc. Rev., 2012, 41, 5577–5587 RSC.
  57. A. Zinchenko, Y. Miwa, L. I. Lopatina, V. G. Sergeyev and S. Murata, ACS Appl. Mater. Interfaces, 2014, 6, 3226–3232 CrossRef CAS PubMed.
  58. A. Lim Teik Zheng, T. Phromsatit, S. Boonyuen and Y. Andou, FlatChem, 2020, 23, 100174 CrossRef CAS.
  59. K. Akhtar, S. B. Khan, E. M. Bakhsh, T. Kamal, S. Ahmad, A. M. Asiri and Y. Anwar, Int. J. Biol. Macromol., 2022, 208, 56–69 CrossRef CAS PubMed.
  60. L. Ge, M. Zhang, R. Wang, N. Li, L. Zhang, S. Liu and T. Jiao, RSC Adv., 2020, 10, 15091–15097 RSC.
  61. T. Kamal, S. B. Khan and A. M. Asiri, Environ. Pollut., 2016, 218, 625–633 CrossRef CAS PubMed.
  62. A. Khalil, N. Ali, A. Khan, A. M. Asiri and T. Kamal, Int. J. Biol. Macromol., 2020, 164, 2922–2930 CrossRef CAS PubMed.
  63. X. Li, F. Dong, L. Zhang, Q. Xu, X. Zhu, S. Liang, L. Hu and H. Xie, Chem. Eng. J., 2019, 372, 516–525 CrossRef CAS.
  64. J. Li, C.-y. Liu and Y. Liu, J. Mater. Chem., 2012, 22, 8426–8430 RSC.
  65. N. Sahiner, H. Ozay, O. Ozay and N. Aktas, Appl. Catal., A, 2010, 385, 201–207 CrossRef CAS.
  66. N. Jabeen, Z. H. Farooqi, A. Shah, A. Ali, M. Khurram, K. Mahmood, N. Sahiner and M. Ajmal, J. Porous Mater., 2021, 28, 1563–1576 CrossRef CAS.
  67. J. Dhiman, S. O. Prasher, E. ElSayed, R. M. Patel, C. Nzediegwu and A. Mawof, J. Cleaner Prod., 2021, 311, 127644 CrossRef CAS.
  68. D. Berillo and A. Cundy, Carbohydr. Polym., 2018, 192, 166–175 CrossRef CAS PubMed.
  69. N. Alhokbany, T. Ahama, Ruksana, M. Naushad and S. M. Alshehri, Composites, Part B, 2019, 173, 106950 CrossRef CAS.
  70. Z. Khorsandi, A. R. Hajipour, M. R. Sarfjoo and R. S. Varma, Green Chem., 2021, 23, 5222–5229 RSC.
  71. H.-C. Li, W.-J. Liu, H.-X. Han and H.-Q. Yu, J. Mater. Chem. A, 2016, 4, 11680–11687 RSC.
  72. N. Ali, T. Kamal, M. Ul-Islam, A. Khan, S. J. Shah and A. Zada, Int. J. Biol. Macromol., 2018, 111, 832–838 CrossRef CAS PubMed.
  73. M. Mifsud, K. V. Parkhomenko, I. W. Arends and R. A. Sheldon, Tetrahedron, 2010, 66, 1040–1044 CrossRef CAS.
  74. J. Zhong, Q. Wang, J. Zhou, D. Chen and Z. Ji, Appl. Surf. Sci., 2016, 367, 342–346 CrossRef CAS.
  75. C. Zhou, J. Bai, Y. Zhang, J. Li, Z. Li, P. Jiang, F. Fang, M. Zhou, X. Mei and B. Zhou, J. Hazard. Mater., 2021, 401, 123232 CrossRef CAS PubMed.
  76. M. Noman, M. Shahid, T. Ahmed, M. B. K. Niazi, S. Hussain, F. Song and I. Manzoor, Environ. Pollut., 2020, 257, 113514 CrossRef CAS PubMed.
  77. M. I. Din, R. Khalid, Z. Hussain, T. Hussain, A. Mujahid, J. Najeeb and F. Izhar, Crit. Rev. Anal. Chem., 2020, 50, 322–338 CrossRef CAS PubMed.
  78. R. Kottappara, S. C. Pillai and B. K. Vijayan, Inorg. Chem. Commun., 2020, 121, 108181 CrossRef CAS.
  79. A. J. Kora and L. Rastogi, Ind. Crops Prod., 2016, 81, 1–10 CrossRef CAS.
  80. N. Sahiner, S. Butun and P. Ilgin, Colloids Surf., A, 2011, 386, 16–24 CrossRef CAS.
  81. N. Sahiner and O. Ozay, Curr. Nanosci., 2012, 8, 367–374 CrossRef CAS.
  82. N. Sahiner and S. Sagbas, Colloids Surf., A, 2013, 418, 76–83 CrossRef CAS.
  83. J. Paul Guin, Y. K. Bhardwaj and L. Varshney, J. Appl. Polym. Sci., 2018, 135, 46200–46212 CrossRef.
  84. F. Naseer, M. Ajmal, F. Bibi, Z. H. Farooqi and M. Siddiq, Polym. Compos., 2018, 39, 3187–3198 CrossRef CAS.
  85. R. Su, S. Ge, H. Li, Y. Su, Q. Li, W. Zhou, B. Gao and Q. Yue, Sci. Total Environ., 2019, 693, 133657 CrossRef CAS PubMed.
  86. Y. Y. Li Sip, D. W. Fox, L. R. Shultz, M. Davy, H.-S. Chung, D.-X. Antony, Y. Jung, T. Jurca and L. Zhai, ACS Appl. Nano Mater., 2021, 4, 6045–6056 CrossRef CAS.
  87. S. Haider, T. Kamal, S. B. Khan, M. Omer, A. Haider, F. U. Khan and A. M. Asiri, Appl. Surf. Sci., 2016, 387, 1154–1161 CrossRef CAS.
  88. C. Saldías, D. D. Díaz, S. Bonardd, C. Soto-Marfull, A. Cordoba, S. Saldías, C. Quezada, D. Radic and Á. Leiva, Carbohydr. Polym., 2018, 180, 200–208 CrossRef PubMed.
  89. S. Sharma, Deepak, A. Kumar, S. Afgan and R. Kumar, ChemistrySelect, 2017, 2, 11281–11287 CrossRef CAS.
  90. S. Wang, Z. Xiao, X. Ma, Z. Zhao, D. Guo, Y. Chen, S. Zhai, Q. An and D. Yang, Appl. Catal., A, 2018, 568, 105–113 CrossRef CAS.
  91. S. Ahmad, S. B. Khan, A. M. Asiri, H. M. Marwani and T. Kamal, J. Sol-Gel Sci. Technol., 2020, 96, 382–394 CrossRef CAS.
  92. S. B. Khan, S. Ahmad, T. Kamal, A. M. Asiri and E. M. Bakhsh, Int. J. Biol. Macromol., 2020, 164, 1087–1098 CrossRef CAS PubMed.
  93. Z. Zhao, Z. Xiao, C. Qin, H. Lv, L. Qin, W. Niu, S. Zhai and Q. An, Ind. Crops Prod., 2021, 164, 113413 CrossRef CAS.
  94. A. L. T. Zheng, S. Sabidi, T. Ohno, T. Maeda and Y. Andou, Chemosphere, 2022, 286, 131731 CrossRef CAS PubMed.
  95. M. Ajmal, S. Demirci, M. Siddiq, N. Aktas and N. Sahiner, New J. Chem., 2016, 40, 1485–1496 RSC.
  96. Y. Sun, F. Zhang, L. Xu, Z. Yin and X. Song, J. Mater. Chem. A, 2014, 2, 18583–18592 RSC.
  97. L. V. Lombardo Lupano, J. M. Lázaro Martínez, L. L. Piehl, E. Rubín de Celis, R. M. Torres Sánchez and V. Campo Dall’ Orto, Langmuir, 2014, 30, 2903–2913 CrossRef CAS PubMed.
  98. N. Sahiner, F. Seven and H. Al-lohedan, Water, Air, Soil Pollut., 2015, 226, 122–133 CrossRef.
  99. S. Demirci and N. Sahiner, Water, Air, Soil Pollut., 2015, 226, 64–76 CrossRef.
  100. A. Zinchenko, Y. Che, S. Taniguchi, L. I. Lopatina, V. G. Sergeyev and S. Murata, J. Nanopart. Res., 2016, 18, 179–187 CrossRef.
  101. J. Liu, D. Su, J. Yao, Y. Huang, Z. Shao and X. Chen, J. Mater. Chem. A, 2017, 5, 4163–4171 RSC.
  102. J. Ding, Q. Li, L. Zhao, X. Li, Q. Yue and B. Gao, RSC Adv., 2017, 7, 17599–17611 RSC.
  103. R. Su, Q. Li, R. Huang, L. Zhao, Q. Yue, B. Gao and Y. Chen, J. Taiwan Inst. Chem. Eng., 2018, 91, 235–242 CrossRef CAS.
  104. J. Ding, Q. Li, X. Xu, X. Zhang, Y. Su, Q. Yue and B. Gao, Carbohydr. Polym., 2018, 190, 12–22 CrossRef CAS PubMed.
  105. R. Su, F. Wang, J. Ding, Q. Li, W. Zhou, Y. Liu, B. Gao and Q. Yue, Carbohydr. Polym., 2019, 220, 202–210 CrossRef CAS PubMed.
  106. T. Kamal, Polym. Test., 2019, 77, 105896–105903 CrossRef CAS.
  107. A. Upadhyay, A. Narula and C. P. Rao, ACS Appl. Bio Mater., 2020, 3, 8619–8626 CrossRef CAS PubMed.
  108. L. Qin, R. Ru, J. Mao, Q. Meng, Z. Fan, X. Li and G. Zhang, Appl. Catal., B, 2020, 269, 118754 CrossRef CAS.
  109. P. Xu, C. Cen, M. Zheng, Y. Wang, Z. Wu and Z. Teng, Mater. Chem. Phys., 2020, 253, 123444 CrossRef CAS.
  110. A. Intanin, P. Inpota, T. Chutimasakul, J. Tantirungrotechai, P. Wilairat and R. Chantiwas, Molecules, 2020, 25, 1798 CrossRef CAS PubMed.
  111. Y. Chen, R. Su, F. Wang, W. Zhou, B. Gao, Q. Yue and Q. Li, Chemosphere, 2021, 270, 129295 CrossRef CAS PubMed.
  112. W. Peng, Y. Yan, D. Zhang, Y. Zhou, D. Na, C. Xiao, C. Yang, G. Wen and J. Zhang, Colloids Surf., A, 2021, 624, 126809 CrossRef CAS.
  113. Z. Zhang, T. Sun, C. Chen, F. Xiao, Z. Gong and S. Wang, ACS Appl. Mater. Interfaces, 2014, 6, 21035–21040 CrossRef CAS PubMed.
  114. J. Ji, Y. Li, W. Fu, Z. Cui, J. Shen and M. Ye, J. Colloid Interface Sci., 2017, 505, 983–994 CrossRef CAS PubMed.
  115. Y. Feng, H. Zhang, B. Xin and J. Wu, J. Colloid Interface Sci., 2017, 506, 154–161 CrossRef CAS PubMed.
  116. T. Yao, W. Jia, X. Tong, Y. Feng, Y. Qi, X. Zhang and J. Wu, J. Colloid Interface Sci., 2018, 527, 214–221 CrossRef CAS PubMed.
  117. D. Varade and K. Haraguchi, eXPRESS Polym. Lett., 2018, 12, 996–1004 CrossRef CAS.
  118. Y. Feng, J. Yin, S. Liu, Y. Wang, B. Li and T. Jiao, ACS Omega, 2020, 5, 3725–3733 CrossRef CAS PubMed.
  119. K.-K. Kang, K. Shim and C.-S. Lee, Colloids Surf., A, 2020, 593, 124607 CrossRef CAS.
  120. R. Wang, M. Zhang, B. Ge, L. Zhang, J. Zhou, S. Liu and T. Jiao, J. Mol. Liq., 2020, 310, 113083 CrossRef CAS.
  121. M. Xia, S.-M. Kang, G.-W. Lee, Y. S. Huh and B. J. Park, J. Ind. Eng. Chem., 2019, 73, 306–315 CrossRef CAS.
  122. J. Zhu, X. Zhang, Z. Qin, L. Zhang, Y. Ye, M. Cao, L. Gao and T. Jiao, Colloids Surf., A, 2021, 611, 125889 CrossRef CAS.
  123. H. Liu, Y.-n. Qin, H.-y. Li, L.-x. Gai, Q.-d. An, S.-r. Zhai, Z.-y. Xiao and L. Cui, Int. J. Biol. Macromol., 2021, 173, 160–167 CrossRef CAS PubMed.
  124. C. Yang, C.-H. Choi, C.-S. Lee and H. Yi, ACS Nano, 2013, 7, 5032–5044 CrossRef CAS PubMed.
  125. P. Eghbali, B. Nişancı and Ö. Metin, Pure Appl. Chem., 2018, 90, 327–335 CrossRef CAS.
  126. I. Sadeghi, E. Y. Liu, H. Yi and A. Asatekin, ACS Appl. Nano Mater., 2019, 2, 5233–5244 CrossRef CAS.
  127. L. Minati, G. Speranza, V. Micheli, M. Dalla Serra and M. Clamer, Dalton Trans., 2020, 49, 3333–3340 RSC.
  128. B. Wang, M. Ran, G. Fang, T. Wu, Q. Tian, L. Zheng, L. Romero-Zerón and Y. Ni, Cellulose, 2020, 27, 6995–7008 CrossRef CAS.

This journal is © The Royal Society of Chemistry 2022