Next Article in Journal
3D Embedded Printing of Complex Biological Structures with Supporting Bath of Pluronic F-127
Next Article in Special Issue
Electrospinning of Magnetite–Polyacrylonitrile Composites for the Production of Oxygen Reduction Reaction Catalysts
Previous Article in Journal
Customizing Starch Properties: A Review of Starch Modifications and Their Applications
Previous Article in Special Issue
Influence of Varying Concentrations of Epoxy, Rice Husk, Al2O3, and Fe2O3 on the Properties of Brake Friction Materials Prepared Using Hand Layup Method
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Recent Advances in Applications of Oxidases and Peroxidases Polymer-Based Enzyme Biocatalysts in Sensing and Wastewater Treatment: A Review

by
Hilda Dinah Kyomuhimbo
1,
Usisipho Feleni
2,
Nils H. Haneklaus
3 and
Hendrik Brink
1,*
1
Department of Chemical Engineering, University of Pretoria, Pretoria 0028, South Africa
2
Institute for Nanotechnology and Water Sustainability (iNanoWS), College of Science, Engineering and Technology, University of South Africa, Florida Campus, Roodepoort, Johannesburg 1710, South Africa
3
Transdisciplinarity Laboratory Sustainable Mineral Resources, University for Continuing Education Krems, 3500 Krems, Austria
*
Author to whom correspondence should be addressed.
Polymers 2023, 15(16), 3492; https://doi.org/10.3390/polym15163492
Submission received: 29 June 2023 / Revised: 10 August 2023 / Accepted: 17 August 2023 / Published: 21 August 2023
(This article belongs to the Special Issue Functional Hybrid Polymeric Composites)

Abstract

:
Oxidase and peroxidase enzymes have attracted attention in various biotechnological industries due to their ease of synthesis, wide range of applications, and operation under mild conditions. Their applicability, however, is limited by their poor stability in harsher conditions and their non-reusability. As a result, several approaches such as enzyme engineering, medium engineering, and enzyme immobilization have been used to improve the enzyme properties. Several materials have been used as supports for these enzymes to increase their stability and reusability. This review focusses on the immobilization of oxidase and peroxidase enzymes on metal and metal oxide nanoparticle-polymer composite supports and the different methods used to achieve the immobilization. The application of the enzyme-metal/metal oxide-polymer biocatalysts in biosensing of hydrogen peroxide, glucose, pesticides, and herbicides as well as blood components such as cholesterol, urea, dopamine, and xanthine have been extensively reviewed. The application of the biocatalysts in wastewater treatment through degradation of dyes, pesticides, and other organic compounds has also been discussed.

1. Introduction

Enzymes are globular proteins composed of polypeptide chains with varying sequences of amino acids [1,2]. The structure and property of a particular enzyme depends on the amino acids present and their sequence, presence, or absence of metal ions and conformation of the protein chain [3,4]. Enzymes are naturally occurring catalysts that accelerate life processes including digestion, converting food to appropriate energy, tissue rebuilding and healing, and conversion of toxins and metabolic waste [5,6]. When enzymes are extracted from living organisms, they have proved to retain their catalytic potential and have, therefore, been exploited to catalyze biochemical reactions in various industries such as textile, paper and pulp, pharmaceutical, and food industries [2,7,8].
The catalytic properties of a particular enzyme are determined by the nature of donor groups in the active site and the amino acid spacer length between the coordinating residues [9]. For instance, metal ions, if present in an enzyme, act as templates for binding various domains of the protein, hence bringing reactive groups together in the proper orientation and also activating chemical bonds to make them responsive [10]. Also, the structure and tertiary fold of an enzyme dictates how it binds to its substrate and performs its catalytic reaction [11]. The catalytic efficiency of an enzyme is based on its ability to bind substrates (and cofactors) onto its active site with the scissile bond stereo-specifically oriented in proximity to the catalytic groups that carry out the reactions [12,13].
Oxidases and peroxidases form part of oxidoreductase enzymes with different co-factors such as heme, flavin, and metal ions to catalyze redox reactions [14]. Peroxidases are heme-containing proteins that catalyze a number of oxidative reactions using hydrogen peroxide as the electron acceptor [15,16]. The heme-group is attached to the protein via a histidine residue that acts a proximal ligand [17]. They are mainly obtained from bacteria, fungi, and plants and their physiological functions are associated with defense mechanisms, metabolism, and biosynthesis of cell wall polymers [18,19] They are classified in three classes depending on their original function in their sources. That is, class I comprising intracellular enzymes in plants, bacteria, and yeast, class II comprising secretory fungal enzymes, and class III comprising glycoproteins containing disulphide bridges and calcium ions [15,20,21].
The active site structures of peroxidases are similar and they all follow a similar mechanism that involves the formation of a two-equivalent oxidized intermediate [22]. The catalytic cycle generally involves three reactions. In the first reaction, hydrogen peroxide oxidizes the enzyme by removing an electron each from iron and porphyrin to produce water and a modified or oxidized enzyme (radical). In some peroxidases, an amino acid instead or a porphyrin ring is oxidized. In the second step, the modified enzyme catalyzes the substrate to produce a free radical (modified substrate) and a second modified form of the enzyme. In the final step, the second substrate reduces the second modified form of the enzyme back to its original form [17,23,24,25,26].
Oxidases, on the other hand, catalyze redox reactions using molecular oxygen as the electron acceptor, producing hydrogen peroxide or water as by-products [27,28]. The main substrate classes for oxidases include amino acids, amines, and alcohols [27]. Since amino acids are very poor in mediating the redox reaction, oxidases rely on a tightly bound cofactor for their activity [29]. Oxidases have been classified based on their two main cofactors, that is, metal in mono and trinuclear centers oxidases and Flavin-containing oxidases. Flavin cofactors are present as Flavin adenine dinucleotide or Flavin mononucleotide [30,31]. The metal containing oxidases have a metal or radical complex at the active site that takes part in redox catalysis [14]. The mechanism involves oxidation of the substrate by two-electron transfer to form a fully reduced Flavin (hydroquinone) and an oxidized product intermediate followed by regeneration of the oxidized cofactor by dioxygen [23,29]. Some cofactor independent oxidases have been identified and for the first stage of reaction they use the catalytic base of the enzyme to extract a proton from the organic substrate to form a conjugated anion intermediate [32].
These enzymes, oxidases and peroxidases, have attracted attention in industrial applications due to their specificity, biodegradability, reaction under mild conditions, and minimal byproduct release, hence reduced pollution load [1,33]. They have been explored in kinetic resolution of racemic mixtures to generate enantiomerically enriched stereoisomers in pharmaceutical industries [34,35], in textile industries [36,37,38], wine making [39,40,41], paper and pulp [42,43,44], wastewater treatment [45,46,47], and in sensing [48,49,50].
However, with all the advantages of using these enzymes in industrial settings, challenges such as non-reusability, lack of long term operational stability, insufficient robustness under, particularly harsher, operating conditions, and substrate or product inhibition are encountered [51,52]. Different approaches such as genetic modification of enzymes, medium engineering, and enzyme immobilization have been taken to improve enzyme properties in biochemical reactions [53]. These different approaches used to modify the enzymes for stability and reusability are briefly discussed below.

1.1. Approaches Used to Improve Stability and Reusability of Enzymes

This section gives a brief highlight of the various techniques that have been utilized to improve on the stability and reusability of enzymes for application in industrial settings. The techniques highlighted include medium engineering, protein engineering, and enzyme immobilization.

1.1.1. Medium Engineering

Medium engineering involves changing the media around the enzyme through changes in the hydrophobicity of the media, salt composition of aqueous solution, introduction or removal of non-aqueous organic solvents, and experimental conditions such as pH and pressure [54,55]. The nature of solvent affects the enantio- and regioselectivity of an enzyme. Hydrophobic solvents provide a better microenvironment for enzymes as they have a smaller tendency to strip water from enzymes so that the enzymes tend to have higher activity in them [56]. Several factors such as compatibility with substrates and products, inertness, low density to minimize mass transfer limitations, surface tension, and waste disposal are considered when selecting an appropriate solvent for a given reaction [57]. Medium engineering to improve enzyme activity has, for instance, been reviewed by [58,59].

1.1.2. Protein Engineering

Protein engineering involves changing the amino acid sequence of an enzyme using molecular biology techniques to yield an intrinsically more stable structure with enhanced activity [55]. The major strategies used are directed evolution and rational design to change the molecular composition and active structures of enzymes, thereby changing their functional attributes [60]. Directed evolution can be achieved by either randomly recombining a set of related sequences such as gene shuffling or by introducing random changes in single protein sequences such as error-prone polymerase chain reaction (PCR). This approach takes after the natural evolution of the enzyme and uses mutagenesis to produce mutant libraries that are screened for enzyme variants with desired properties. The structural information of the enzyme is not needed [60,61,62]. Rational design involves use of detailed knowledge of the enzyme structure, function, and mechanism to create changes in the amino acid sequence of the enzyme by site-specific mutagenesis [60,63]. This approach is based on the structural analysis and computational modeling of enzymes to account for the physiochemical properties of their amino acids and simulate their interaction with the environment [61]. Enzyme engineering facilitates development of tailor-made functional biocatalysts with properties far beyond their natural capabilities [64]. However, these techniques are time consuming, expensive, and tedious, so that it is challenging to apply them in an industrial setting [65]. Enzyme engineering techniques, advantages, and shortcomings have, for instance, been extensively reviewed by [66,67,68,69,70].

1.1.3. Immobilization of Enzymes

The difficulty in using soluble enzymes in industrial and environmental applications due to their instability, non-reusability, and challenging recovery, as well as product inhibition has led to enzyme immobilization [71,72]. Immobilization of enzymes involves binding them to a support, entrapment/encapsulation, and crosslinking [73], and is advantageous for commercial applications because of its ease of handling and separation from the reaction, operational stability of the enzyme as the enzyme is dispersed and cannot aggregate, interaction of the enzyme with external interfaces is prevented, and because it ensures rigidification of the enzyme structure [74,75,76]. The benefits associated with immobilization are enzyme thermal and pH stability, specificity, selectivity, operational stability in harsher conditions, ease of separation from the product, and efficient enzyme recovery and reuse [77,78,79]. Since the first step in inactivation of enzymes is the change in their tertiary structure and dissociation of enzyme subunits or loss of their correct assembly, immobilization limits the enzymes conformational variations and leads to protein unfolding resulting in enhanced enzyme stability [54,80].
Different factors such as simplicity of the immobilization procedure, stability and mechanical resistance of the resultant biocatalyst, and possibility of coupling enzyme immobilization with purification are considered when selecting industrial biocatalysts [81]. The biocatalyst should be stable under potentially adverse reaction conditions and at the same time maintain high activity [82,83]. The properties of the biocatalyst are determined by the nature of the enzyme, properties of the supporting material, and the method and conditions of the immobilization procedure [84]. A good immobilization process should be cheap, quick, and enzyme-friendly, resulting in high loading factors, and be applicable to a large scale of biomolecules [85].
Carrier-free immobilization approaches involving crosslinking of enzyme molecules to form crosslinked enzyme crystals (CLECs) and crosslinked enzyme aggregates (CLEAs) have also been explored to improve enzyme properties. In CLECs, the enzyme is first crystallized in an aqueous solution at suitable temperature followed by crosslinking [86]. CLECs often provide higher activity and enantioselectivity in organic solvents than in aqueous solutions. They also retain activity in harsher conditions such as extreme pH and temperature, organic or aqueous-organic mixtures which result from crystallization of the crystal lattice, and its constituent enzyme molecules after chemical crosslinking of enzyme crystals [87,88,89]. However, preparation of CLECs is laborious and requires high enzyme purity [90]. CLEAs are obtained by precipitation of enzymes followed by crosslinking. CLEAs are often too soft and may exhibit poor stability in stirred tanks or in packed bed reactors [91,92]. However, if immobilized in large porous support or rigid polymers, their stability can be improved [86]. Carrier-free crosslinking of enzymes has already been extensively reviewed by [93,94,95,96].
In the case of support required immobilization, the enzyme must be appropriately oriented to prevent inefficient performance, and this is usually achieved by manipulating the structure of the support and immobilization conditions [97,98]. Properties of the support such as pore size, hydrophilic/hydrophobic balance, inertness, biocompatibility, resistance to microbial attack, aquaphilicity, and surface chemistry have an effect on the catalytic properties of the biocatalyst, and have to be considered when selecting an appropriate support [86,99]. The properties of the specific enzyme to be immobilized such as electrostatic, hydrophilic, and hydrogen bonding interactions, and its unfolding mechanisms when in contact with the surface of the target material also need to be considered to retain the activity of the immobilized conjugate [100,101].
Immobilization of enzymes on solid supports can be achieved through physical methods such as adsorption, entrapment, and encapsulation, and chemical methods such as covalent bonding and crosslinking. Physical adsorption is the simplest method and can be performed under mild conditions aided by electrostatic interaction, steric interaction, and changes in the hydration state [102]. Since weak forces are involved in this approach, it is often associated with enzyme leaching, making it inappropriate for applications where stability and long term usage under sub-optimal conditions are necessary [103]. Adsorption of an enzyme can also induce conformational changes which affect the rate and specificity of the catalyst [104].
Covalent coupling methods are adopted because the amino acid residues on an enzyme can be directly bonded to an active group on the support [105]. Multipoint covalent attachment minimizes leakage of enzymes through increased bond strength and protects the enzyme from denaturation by constraining it to the local environment of the support [105,106]. The main challenge associated with covalent immobilization is the need for pure enzymes, and yet purification of enzymes is time consuming and resource intensive. Impurities rather than target enzymes may have a strong tendency to bind to the support which could greatly affect its activity [107]. Moreover, some enzymatic activity may be lost as the active site is hidden or restricted from assuming the conformation needed to initiate catalysis, and in case of irreversible enzyme deactivation, the enzyme and carrier are all rendered useless [73,108].
Different kinds of supports, such as porous and non-porous supports, have been used to immobilize enzymes. Porous materials have high enzyme loading but suffer diffusion limitations [97,109]. Pore sizes equal to or slightly larger than the enzyme dimensions are considered to be more favorable than the larger pore sizes, but they can impose large mass transfer limitations to the substrates and products in and out of the host material [110]. Non-porous materials have minimum diffusion, but with low enzyme loading per unit mass [111,112]. Immobilization of enzymes on non-porous supports may cause enzyme inactivation through interaction with gas bubbles generated by stirring or oxygen [113]. The various enzyme immobilization techniques and supports as well as their advantages and disadvantages have already been reviewed by [114,115,116,117,118].
In this review, our focus is on oxidase and peroxidase enzymes immobilized on metal and metal oxide nanoparticle-polymer composites. Metal and metal oxide-polymer composite materials exhibit metallic and polymer properties with excellent conductivity, high mechanical strength, and ease of synthesis and good chemical and thermal stability [119,120]. Moreover, in the case of enzyme immobilization, these materials can be modified into various shapes, sizes, and compositions to suit the needs of the needed biocatalyst. These characteristics make them great candidates as support materials in biocatalysts for industrial applications. However, the reviews on biocatalysts using metal and metal oxide-polymer composites as enzyme supports are scanty. This review, therefore, focuses on these materials as supports for oxidase and peroxidase enzymes. The review discusses the applications of these biocatalysts in biosensors for hydrogen peroxide, glucose, and other compounds in human blood, pesticides, as well as other organic pollutants. The application of these biocatalysts in waste water treatment, especially degradation of dyes, pharmaceuticals, pesticides, and organic pollutants has also been discussed.

2. Nanoparticle-Polymer Composites

Nanoparticles (NPs) have attracted attention in various applications including enzyme immobilization [121,122], target drug delivery [123,124], bioseparation [125], immunoassays [126,127], and environmental analysis due to their surface properties [128,129]. They have been adopted for enzyme immobilization because they offer relatively large surface areas, high effective enzyme loading, excellent biocompatibility, good conductivity, and high free energy [130,131]. The combination of NPs and enzymes is of considerable importance because NPs offer a friendly platform to assemble enzymes and also enhance the electron transfer process between enzymes and other supports like electrodes [132]. The ability to tailor the properties of NPs offers excellent prospects for enhancing the catalytic performance of the enzyme-based biocatalyst [133]. Enzyme-NPs bioconjugates offer hope for biotechnological applications where high catalytic specificity, prolonged reaction time, and ability to recycle expensive biocatalysts are required [103]. For example, when glucose oxidase was covalently immobilized on amino-modified silica-encapsulated magnetic nanoparticles (MNPs), the biocatalyst demonstrated thermal stability at 80 °C and operational stability of 98% retained activity after 45 days of storage and 90% after 12 reuses [134].
Immobilizing enzymes on planar surfaces can limit their performance due to distortion of native protein configuration, steric hindrance, or slower diffusion rates of an incident substrate towards the bulk surface [65]. Since NPs maintain high radii of curvature due to their small diameters, they allow increased center-to-center distances between adjacent immobilized enzymes while limiting unfavorable protein-to-protein interactions [65,135]. They can also immobilize a considerably higher concentration of enzyme compared to 2D planar surfaces [136]. For example, Husain et al. [137] compared immobilization of galactosidase on native ZnO and ZnO-NPs by physical adsorption. The ZnO-NPs displayed higher adsorption yield (85%) compared to the native ZnO (60%) and retained higher fractions of catalytic activity in acidic and alkaline conditions and a broader optimum temperature. Due to the small sizes of NPs, enzymes attached to them can behave freely as in a soluble state and can easily diffuse through high molecular mass polymeric substrates [138,139]. They, therefore, provide desirable features that balance the contradictory issues of surface area, mass transfer resistance, and effective enzyme loading [140,141]. Enzyme-NPs based biocatalysts are particularly preferred in electrochemical devices that require fast and reversible charge transfer as the NPs help in optimizing electron transfer between the enzyme and the electrode [142,143].
However, due to the large surface area to volume ratio of NPs, they have high reactivity, easily aggregate, and easily undergo degradation upon direct exposure to certain environments, leading to poor stability and dispersity [107,144]. Different organic and inorganic materials including silica [145], alkyl benzenesulfonate [146], tannic acid [147], and polymers such as polyethyleimine (PEI) [148], polyethylene glycol [149], and polyvinyl butyral (PVB) [98] have been used to functionalize the surfaces of NPs to increase enzyme loading amounts and their stability. Coatings can protect NPs from reaction with the external environment and simultaneously serve as a medium for subsequent particle functionalization to render them chemically functional and simultaneously physiologically compatible for biomedical applications [145,150]. They can also enhance mono-dispersity of the NPs by preventing their coagulation [151]. For example, Chen et al. [152] grafted zirconia NPs with carboxylic surfactant modifiers from tween 85 and erucic acid to change its surface from hydrophilic to hydrophobic prior to lipase immobilization. Pan et al. [153] covalently immobilized galactosidase on MNPs-chitosan NPs and the bioconjugate displayed excellent dispersibility compared to when only MNPs were used for immobilization. The biocatalyst retained 92% of its initial activity after 15 cycles and 88% after 60 days of storage.
Amidst the exploration of NP surface modification, hybrids of metal nanoclusters and polymers (metal polymer composites) have been made for enzyme immobilization. The composites comprise either metal nanoparticles dispersed in a polymer matrix or contain a core (metal NP) encapsulated in a shell (polymer), and most of the polymers adhere to the nanoparticle surfaces in a substrate-specific manner [140,151]. Polymer chains offer flexibility and diversity to control the chemical composition and functional groups on the surface of the NPs [154]. Moreover, binding of NPs to polymers occurs through nitrogen in the polymer backbone, hence removing the need for surface modification of the polymer prior to immobilization [155]. On electrode surfaces, polymers also provide an effective immobilization patterning for enzymes and may facilitate electron transfer from enzymes to electrodes, which improves electrode sensitivity [156].

2.1. Polymers Explored in Functionalization of Nanoparticles

One of the polymers most commonly used to functionalize nanoparticles is the natural cationic polysaccharide polymer, chitosan, due to its biocompatibility, non-toxicity, good adhesion, mechanical stability, high permeability towards water, and excellent film-forming ability [85,157,158]. Chitosan molecules are rich with amino groups which provide a hydrophilic environment compatible with biomolecules, nanoparticles, and other polymers, and can, therefore, provide an excellent matrix for the preparation of enzyme electrodes [159,160,161]. In addition, the presence of amino and hydroxyl groups on chitosan enhances its interaction with enzymes and allows simple immobilization techniques such as adsorption and entrapment [162,163]. Chitosan entraps bioactive biomolecules such as enzymes and nucleic acid through inherent chemical crosslinking, ionic complexation mechanisms, and ionic crosslinking [164]. The high positive charge of chitosan solutions enables them to be adsorbed on the surfaces of nanoparticles, thus stabilizing, protecting, and exhibiting special nanometer film effects on the nanoparticles that enhance compatibility of the nanoparticles with enzymes [165,166]. The different functional groups also allow easy chemical modification of chitosan to achieve desired properties of an enzyme carrier and to improve its physiochemical characteristics such as adsorption capacity and mechanical resistance [162,167]. The application of chitosan in the modification of nanoparticles and enzyme immobilization has already been reviewed elsewhere [168,169,170].
Another natural polysaccharide polymer that has been extensively investigated in the functionalization of NPs is alginate, due to its biocompatibility, low toxicity, and mild gelation by addition of divalent cations [171,172]. It is a naturally occurring anionic, hydrophilic, and chain-forming polysaccharide that contains randomly arranged linear unbranched chains of α-l-guluronate (G block) and β-d-mannuronate (M block) residues [173,174]. This polymer contains many free hydroxyl and carboxyl groups, which enables it to form intramolecular hydrogen bonds [175]. Its polymeric chains can easily crosslink in the presence of multivalent cations such as Ca2+, Cu2+, Mn2+, Pb2+, etc., to form insoluble hydrogels ionotropic gelation [176]. The gelation process takes place through ionic cross-linking of negatively charged carboxyl groups of the alginate chain and multivalent metal ions with opposite charges to give a gel network with small pores that can entrap biomolecules such as enzymes [177]. These hydrogels are capable of tolerating high temperature and are biocompatible with biomolecules so that they are used as suitable matrices for the entrapment of enzymes [178]. Most importantly, the hydrogels can be produced at room temperature using simple equipment like beakers and droppers, and the encapsulation can be carried out anywhere, even in the presence of high concentrations of solids. Hence, immobilization of proteins can be carried out under mild and safe conditions [179,180]. The beads/balls formed are of adequate texture, homogeneous, and are porous so that they can allow diffusion of substrates and products to and from the immobilized enzymes [181,182]. The encapsulation of the enzymes and nanomaterials in alginate has been extensively reviewed before [175,183,184,185].
Besides naturally occurring polymers, synthetic polymers such as poly ethylene amine [186,187], polyaniline [188,189], polydopamine [190,191], polypyrrole [192,193,194], polyvinyl butyral [195], nafion [196,197,198], etc., have been utilized in the functionalization of NPs for enzyme immobilization. Conducting polymers, especially polyaniline (PANI) and polypyrrole, have been extensively explored, especially in enzyme biosensor applications, due to their controllable electronic properties, chemical inertness, mechanical stability, limited permeability, and simple preparation procedure [199,200,201]. These polymers contain self π-conjugated systems with alternating single and double bonds along the polymer chain providing a structure with high electronic properties such as high electron affinity, high electrical conductivity, and low ionization potential [202,203]. Additionally, they possess other outstanding properties such as easy preparation and functionalization, biocompatibility, good thermal and electrochemical stability, and thus act as suitable immobilization matrices for biomolecules that also facilitate electron transfer in redox or enzymatic reactions [203,204,205]. Apart from the above mentioned advantages, PANI has demonstrated the ability to couple with oxidoreductase enzymes, cause impressive signal amplification, and eliminate electrode fouling in biosensing applications [206,207].
The different polymers that have been utilized for the immobilization of enzymes and nanoparticles have, for instance, been reviewed by [208,209].

2.2. Metal and Metal Oxide Nanoparticles Explored in Nanocomposites for Enzyme Immobilization

NPs are prepared from a variety of materials including proteins, polysaccharides, polymers, metals, and metal oxides and other inorganic materials [210,211]. The NPs used in enzyme immobilization are commonly classified as carbon nanotubes, dendrimers, quantum dots, liposomes, metallic, and polymeric NPs, and possess different topographies and shapes like nanotubes, nanospheres, nanowires, nanorods, nanorings, and nanofibers, as previously reviewed [212,213]. Their properties are influenced by factors such as size and morphology, surface charge and permeability, degree of biodegradability, and biocompatibility [211]. Enzymes have been immobilized on Silica NPs [214,215,216,217], polymeric nanoparticles [218,219,220], quantum dots [221,222,223,224], carbon nanotubes [225,226,227,228], metal and metal oxide NPs [135,229,230], and bimetallic NPs such as Au-PtNPs [231], TiO-CeONPs [232], and Au-AgNPs [233].
In this review, emphasis is put on enzymes immobilized on metal and metal oxide nanoparticles encapsulated or embedded in polymers. The common metal and metal oxide nanoparticles extensively used in immobilization of oxidase and peroxidase enzymes are AuNPs, AgNPs, MNPs, ZnONPs, and TiO2NPs. These groups are discussed in more detail in the next section of the review. Although not discussed in this review, other metal and metal oxide nanoparticles such as platinum NPs (PtNPs) [234,235,236], copper NPs [237,238], palladium [239,240], nickel [241,242], and nickel oxide NPs [243,244,245,246] have been reported in the literature as supports for enzymes.

2.2.1. Gold Nanoparticles (AuNPs)

AuNPs are good biocompatible materials and provide a mild microenvironment similar to that of redox proteins in native systems, and give the protein molecules more freedom in orientation [247]. They have been used for the immobilization of enzymes for sensor applications because they can act as tiny conduction centers that facilitate electron transfer between enzymes and electrode surfaces [248,249]. AuNPs have high affinity to amine groups and cysteine residues in enzymes, and binding to enzymes occurs through these groups present in the enzyme [136,250]. The AuNPs can also be functionalized with thiolated molecules with carboxylic groups, which in turn, are conjugated with amine groups of the protein [251]. However, immobilizing enzymes to AuNPs is associated with poor reusability, due to the difficulty in separating the bioconjugate from the reaction mixture even under high ultracentrifugation conditions [155,252]. Therefore, the nanoparticles need to be tethered to a more stable structure that can easily be separated from the reaction medium by simple means [144,253].

2.2.2. Silver Nanoparticles (AgNPs)

AgNPs have attracted attention in enzyme immobilization for biosensor activity due to their high electrical conductivity, low cost, biocompatibility, and excellent biocatalytic activity [254]. During adsorption of enzymes onto AgNPs, some hydration water is retained between the adsorbed enzyme layer and the AgNPs surface, which helps form highly hydrated enzyme molecules, thus preserving their activity [98]. For instance, when β-galactose was immobilized on tannic acid-stabilized AgNPs, an immobilization yield of 83.6% was achieved and the biocatalyst demonstrated stability at higher temperatures, acid and alkaline pH, storage at 4 °C (with 77% retained activity after 30 days), and during reusability (with 77% retained activity after 10 cycles). Due to their high conductivity, AgNPs have proved to facilitate more efficient electron transfer in biosensors than other nanoparticles [98,255]. AgNPs do, however, possess a dual effect on the enzymatic activity of certain enzymes. For instance, Ma et al. [256] observed that when glucose oxidase was immobilized on refluxed AgNPs, the refluxing time of AgNPs had an inhibitory effect on the enzyme, which decreased with increase in refluxing time.

2.2.3. Magnetic Nanoparticles (MNPs)

MNPs have found potential applications in biomedical aspects due to their strong magnetic property and low toxicity [138]. Superparamagnetic NPs are preferred to ferromagnetic NPs for practical applications because no residual magnetism is retained after the magnetic field is removed [52,99,257]. Magnetic separation of MNPs offers efficient recovery of the biocatalyst from reaction products, which is especially important in pharmaceutical industries where enzyme contamination of final products can cause detrimental side effects [103,258]. Due to the magnetic property of MNPs, substances attached to them can be separated from the reaction medium or directed by a magnetic field [149,259]. In addition, MNPs present minimal steric hindrance to reactants in solution for accessing the active sites of the biocatalyst, leading to lower mass transfer resistance and less fouling in reactions [34,260]. Immobilization of enzymes on MNPs is associated with less fouling and the bioconjugate can be separated from the mixture by application of a magnetic field [261].
It is noteworthy, though, that MNPs are dispersible in organic solvents in which the enzymes are generally not soluble and tend to agglomerate in liquid media due to strong magnetic dipole-dipole attraction, are susceptible to air oxidation, and do not readily combine with certain enzymes [52,106,153]. For example, no cellulase was adsorbed onto naked MNPs until glutaraldehyde was added [262]. In order to prevent that, their surfaces are often modified with surfactants or polymers with specific functional groups to improve stability and enzyme loading [263,264].

2.2.4. Zinc Oxide Nanoparticles (ZnONPs)

Since Zn compounds have been regarded as generally safe by the U.S. Food and Drug Administration (US FDA), ZnONPs have been extensively applied in several applications including sensors, solar cells, photocatalysis, and biotransformation [265]. ZnONPs have attracted interest as potential materials for biosensing due to their large surface area for strong adsorption, chemical stability, biocompatibility, and high electron communication [266]. They have a high isoelectric point (~9.5) and show greater affinity towards low isoelectric point enzymes with most immobilization procedures achieved through adsorption or crosslinking [267,268]. For example, Antony et al. [269] adsorbed diastase α-amylase on ZnONPs and it was revealed that the enzyme was adsorbed via electrostatic interaction with the functional groups on the surface of the ZnONPs. The resultant biocatalyst demonstrated thermal stability, reusability with 80% retained activity after four cycles, and storage stability of 70% retained activity after 30 days of storage. When tyrosinase was immobilized on ZnONPs for application as a mediator free phenol biosensor, Li et al. [270] observed that the high isoelectric point of ZnONPs did not only provide a conducive microenvironment for negatively charged tyrosinase (pI~4.5) to retain its activity, but also promoted direct electron transfer between the enzyme and electrode. ZnONPs can be surface functionalized with a wide range of metal, semiconductor, and polymer materials, thereby imparting useful properties for a wide range of applications [271]. Moreover, ZnONPs have large excitation binding energy at room temperature, and when exposed to UV radiation, they can release electron-hole pairs which aid catalytic reactions of enzyme electrodes. Hence, the current response can be fine-tuned for the development of photo-controlled enzyme based biosensors [272]. Direct adsorption of enzymes on bare ZnONPs, however, leads to enzyme aggregation due to high enzyme loadings, and the surface of the NPs needs to be functionalized prior to immobilization [273].

2.2.5. Titanium Oxide Nanoparticles (TiO2NPs)

TiO2NPs have gained attention in various applications due to their non-toxicity, photo-corrosion resistance, biocompatibility, photochemical stability, unique electrical and optical properties, and the fact that they can be produced on a large scale under mild conditions [274,275]. These unique properties enable TiO2NPs to create an appropriate microenvironment for immobilizing enzymes without loss of biological activity, and also facilitate electron transfer between enzymes and electrode surfaces in case of biosensing applications [276]. For example, Zhang et al. [277] fabricated a horse radish peroxidase (HRP)-TiO2 film electrode by casting a mixture of HRP solution and aqueous TiO2NPs on pyrolytic graphite (PG) electrodes. The TiO2NPs film greatly enhanced the electron exchange between the enzyme and the PG electrode, and the electrode demonstrated stability and responsiveness in long-time voltammetric experiments. However, the application of bare TiO2NPs is restricted due to their low quantum efficiency resulting from recombination of photo-generated carriers, low stability on electrodes, and a wide band gap. In order to address those issues, the NPs are usually doped with metals and non-metals to improve their properties [278,279]. For instance, Ahmad and Sardar [280] compared physical adsorption of cellulase on TiO2NPs and covalent coupling where the TiO2NPs were modified with aminopropyltriethoxysilane. The covalently immobilized enzymes showed a higher activity (93%) compared to the physically adsorbed enzymes (76%) and demonstrated higher reusability and operational stability.

3. Methods Used to Functionalize Nanoparticles with Polymers on Electrodes for Enzyme Immobilization

This section discusses the specific approaches that have been explored to functionalize nanoparticles with polymers for purposes of enzyme immobilization for biosensor applications and waste water treatment.

3.1. Polymer Grafting

Polymer grafting of NPs via low molecular weight linkers or polymers containing amino or epoxy functional groups is one of the methods used to functionalize NPs for enzyme immobilization [154,281]. NPs have high surface free energy and easily agglomerate when dispersed in the polymer matrix. This thermodynamic instability can be avoided by grafting them with functional polymers prior to their dispersion [282,283,284]. Polymer chains provide flexibility and diversity that control the chemical composition and functional groups on the surface of the NPs [281]. Due to their low molecular weight, monomers can penetrate the aggregated NPs and react with the activated sites on the NPs surface, hence filling the interstitial volume inside the NPs aggregates [285]. This results in steric repulsion between the grafts, thus preventing subsequent aggregation [286]. This also makes the surfaces of the NPs hydrophobic, which is essential for their miscibility in the polymer matrix [285]. The properties of the polymer-grafted NPs can be tailored through a proper selection of the species of the grafting monomers and grafting conditions [282].
The polymer layer can be attached to the NPs in two ways: grafting from and grafting to. The grafting-to approach involves the binding of an active chain end of a polymer with a binding site on the NPs surface (Figure 1) [285,287]. The NPs surfaces are firstly treated with a coupling agent to introduce functional groups that form bonds with both NPs and the polymer, followed by radical grafting polymerization in a suitable medium [282]. The polymer should have functional groups that can react with the surface of the functionalized surface of the NPs, or it can be functionalized by prefabricating its polymer chains via their reactive terminal groups [283,288]. The functionalized polymers are covalently bonded to the existing functional groups on the NPs surfaces through the ligand exchange route (Figure 1) [282,288,289]. For example, Dutta et al. [290] synthesized 3 poly(N-isopropylacrylamide-ran-poly(ethylene glycol) methylether acrylate)-block-poly(acrylic acid) [P(NIPA-r-PEGMEA)-b-PAA] block copolymer for grafting on to amino functionalized MNPs. First, three different di-block copolymers of NIPA, PEGMEA, and tertbutyl alcohol (tBA) were synthesized by a polymerizing mixture of varying molar ratios of NIPA and PEGMEA in the presence of PtBA macro-CTA. The P(NIPA-r-PEGMEA)-b-PtBA copolymers were then hydrolyzed to produce corresponding P(NIPA-r-PEGMEA)-bPAA. Each of the three P(NIPA-r-PEGMEA)-b-PAA copolymers was then covalently linked with NH2-MNPs using 1-Ethyl-3-(3-dimethylaminopropyl)carbodiimide (EDC) and N-Hydroxysuccinimide (NHS) as a coupling agent.
This approach is simple, but its main disadvantage is that the polymer is adsorbed onto the surface of the NPs, producing a monolayer of spherical polymer chains which restricts further adsorption due to diffusion barrier and steric hindrance, leading to low graft density [287,292,293]. The approach is further limited to polymer grafts with defined end groups and the surface of nanoparticles may have unreacted functionality [294]. A more direct method is the grafting-to approach, that can be characterized by self-assembly of the monomer and NPs that are simply mixed with polymerization taking place in a polymerizing agent [295]. Direct incorporation of NPs into block copolymers through direct block copolymer-NPs interaction has also been reported [296,297,298]. The different paths that can be used to achieve self-assembly of polymers on NPs surfaces have already been reviewed by [299,300].
The grafting-from approach involves introducing a monolayer initiator on the NPs surface, followed by growth of polymer chains from the initiator through in situ polymerization via thermal or photochemical means, as illustrated in Figure 1 [283,292,293,301]. This forms a uniform surface coating of the polymer chains on the surface of the NPs [292]. The approach can be used to control the molecular weight, morphology, and composition of the polymer ligands grown from the NPs surface, thus controlling the properties of the nanocomposite [289,302]. For example, Yong et al. [281] modified vinyltriethyoxysilicane (VTES) NPs, followed by the addition of a mixture of glycidyl methacrylate (GMA) and methacryloxethyl trimethyl ammonium chloride (MATAC) monomers (dropwise) in the presence of ethanol and deionized water. The graft polymerization was allowed to stand for 6 h at 70 °C and the products were subsequently collected by magnetic separation, washed with ethanol and distilled water, extracted in ethanol, and dried at room temperature under vacuum.
The binding between the NPs and polymers is strong, and diffusion of smaller monomer is usually easier [289,293]. The thickness of the grafted polymer layer increases with increasing polymerization time at affixed monomer concentration [292]. When the polymer chains are densely grafted to a surface, steric crowding occurs, forcing the chains to stretch away from the surface so that a brush is formed. As a result, this approach provides high grafting density and the NPs can stably disperse in the solvent of interest [292]. The polymer brush length, molecular weight of the polymer brush, molecular weight of the polymer matrix, and grafting density determine the dispersion of the polymer-grafted NPs in a polymer matrix [284].
Synthesis of nanocomposites by polymer grafting has, for instance, been reviewed by [285,289,294,303,304,305].

3.2. Self-Assembled Monolayer Deposition

This technique involves alternate deposition of thin layers (also called self-assembled monolayers) of polymer, nanoparticles, and enzymes on the electrode surface, either by use of voltage power supply, alternate drop-casting of the solutions on the electrode surface [156] and allowing to dry, or alternate dipping of electrodes in respective solutions for a given period of time [306,307]. For example, Luo et al. [250] dipped a gold electrode in a (0.5% w/v, pH 5) chitosan (CS) solution while connected to a 3.0 V DC power supply, allowed it to dry, then immersed it in AuNPs solution for 10 h at 4 °C, and finally incubated the electrode in HRP solution for 12 h at 4 °C. Zhong et al. [308] adsorbed GOx on self-assembled AuNPs and a double-layer 2D network MPS polymer. A gold electrode was immersed in an MPS solution in ethanol for 3 h to produce a self-assembled monolayer, and then dipped in NaOH solution to polymerize the silane networks into a 2D network, followed by immersion in MPS to form a second silane layer. The modified electrode was then dipped in AuNPs solution for 10 h, followed by immersion in GOx overnight. Alternatively, the enzyme solution, polymer, and nanoparticles are mixed together to form a homogenous mixture, which is then dropped onto the electrode and allowed to dry at ambient temperatures [309]. In other instances, the enzyme and nanoparticles are first drop-casted on the electrode surface, allowed to dry at ambient temperatures, and then a polymer solution is also drop-casted on the modified electrode to act as a net that prevents the enzyme and nanoparticles from leaching into the solution [197]. For example, Zou et al. [310] dropped a solution of multiwalled nanotubes (MWNTs) in DMF on the surface of a GCE, followed by electrodeposition of PtNPs on the modified electrode using H2PtCl6 to form Pt/MWNTs/GCE. A solution of GOx was then mixed with chitosan-SiO2 sol-gel by hand, and the mixture was then drop-casted on the Pt/MWNTs/GCE. The electrode was allowed to dry and then nafion solution was drop-casted to form a protective film. Lu et al. [311] drop-casted a mixture of silver nanowires and chitosan solution on a glassy carbon electrode (GCE), allowed to dry, and then immersed the modified electrode in glucose oxidase solution overnight at 4 °C. Biosensors fabricated using the self-assembled monolayers (SAM) technique have proved to possess high sensitivity and short response time [312]. The SAM technique has become a popular, simple, and reliable procedure to immobilize enzymes and molecules on various metal and oxide surfaces due to its simplicity, flexibility, and the formation of a high level of ordered surfaces on a molecular scale [308]. However, the film thickness of the deposited layers is often uncontrollable in this technique [250].

3.3. Electrochemical Deposition

Electrochemical deposition, also known as electrodeposition, involves dipping an electrode in a mixture containing the enzyme, metal salts, and monomer solutions connected to a controlled voltage or current supply [85,235]. Alternatively, the NPs-polymer composite is formed on the electrode prior to immobilization of the enzyme [165]. For example, Perveen et al. [313] drop-casted MnO2-graphene/polythioaniline solution onto a GCE, followed by a ferritin mediator onto the modified electrode. Glucose oxidase in phthalate buffer was then entrapped on the modified electrode by cyclic voltammetry at 100 mV/s between −1 and 1 V. Tan et al. [314] first electrodeposited PtNPs on a gold electrode and then potentiostatically electrodeposited a pre-crosslinked glucose oxidase–glutaraldehyde–chitosan mixture on the modified electrode.
This method is simple, can be performed under mild conditions, and the thickness of the polymer films formed on the electrode can be easily controlled [85,312]. For instance, electrodeposition of chitosan in acidic solution on a gold electrode under a constant voltage led to the formation of hydrogen bubbles in the deposited chitosan hydrogel. Upon drying, the bubbles are turned into nanopores which increase the surface area of the chitosan on the electrode [250].

3.4. Electrospinning

Electrospinning is the production of micro and nanofibers of varying lengths from metal-polymer solutions and melts through an electrically charged jet by use of electrostatic and mechanical force, as illustrated in Figure 2 [315,316]. A strong electric field is applied between the solution droplet and the grounded collector, creating an electrostatic potential that is sufficiently high to overcome the surface tension of the droplet, hence forming a charged liquid jet that is deposited on the collector [317,318]. For example, Golshaei et al. [319] carried out in situ polymerization of anthranilic acid (ANA) monomer, 3-carboxy-N-(2-thenylidene) aniline (CNTA) monomer, and HAuCl4 to form Au/P(ANA-co-CNTA) nanocomposite. The nanocomposite was dispersed in polyvinyl acetate prepared in DMF or acetone solvents and electrospun to produce nanofibers. Glucose oxidase was then immobilized on the activated nanofibers for glucose sensing using an EDC/NHS coupling agent. Sriwichai and Phanichphant [320] dissolved poly (3-aminobenzylamine) (PABA), polyacrylonitrile (PAN), and functionalized carbon nanotubes (f-CNTs) in DMF and electrospun the mixture to obtain a PABA/f-CNTs composite. The fibers were then immersed in GOx solution in the presence of an EDC/NHS coupling agent.
The electrospinning technique can be considered simple, cost effective, flexible, and any soluble polymer can be used to obtain continuous ultra-thin fibers [315]. Electrospun nanofibers have a high surface area to volume ratio, high porosity, and are biocompatible with high numbers of functional groups on the surface, so that they make good matrices for enzyme immobilization [320,322,323]. Polymer-based electrospun nanofibers have been considered the most appropriate form of enzyme support, due to low hindrance of mass transfer, easy recoverability, high enzyme loads, and potential applications for continuous operations [322,324]. Electrospinning of nanofibers for enzyme immobilization has been extensively reviewed by [325].

4. Application of Enzyme-Nanoparticle-Polymer Composites in Biosensors

A combination of enzyme reactions with electrochemical methods allows for the development of different enzyme-based electrochemical biosensors for the detection of environmental pollutants, due to their good selectivity, rapid response, and miniature size [326]. Nanocomposite films have been reported to display three-dimensional superstructures with high electrocatalytic activity, stability, and uniform particle distribution [327]. The polymers act as excellent transducers as the functional groups present in their backbone enable conjugation between enzymes and nanoparticles to form a more electrochemically active structure [312]. For example, Silva and Vieira [328] designed a biosensor for detecting dopamine in pharmaceutical samples using laccase immobilized on AuNPs stabilized in poly(allylamine hydrochloride) (Figure 3). The cyclic voltammetry and electrochemical impedance spectroscopy of the biosensor indicated that the nanocomposite facilitated electron transfer between the enzyme and electrode surface with high selectivity and stability. When German et al. [329] co-immobilized GOx and 1,10-phenanthroline-5,6-dione (PD) (mediator) on AuNPs graphite electrode, a linear range of 0.1–10.0 mM was observed for the detection of glucose. When a polypyrrole layer was added on the electrode through polymerization, the linear range increased to 0.1–25 mM and 0.1–50.0 mM after 22 h and 69 h of polymerization, respectively. Luo et al. [250] adsorbed HRP on a gold electrode modified with AuNPs chemisorbed onto porous chitosan films for the detection of methylene blue. The biosensor showed a wide dynamic range of 8.0 μM–15 mM, LOD of 2.4 μM, storage stability of 85% after 4 weeks storage, and 6% activity loss after 50 reuse cycles.
It is difficult for enzymes to exchange electrons with electrode surfaces directly due to their large and complex structure, since the redox centers are deeply immersed in the bodies and the three-dimensional structures hinder interaction with the electrode. Also, enzymes undergo denaturation upon direct immobilization on bare electrodes, hence, lose their bioactivity. These inhibitions are overcome by modifying electrodes with mediators and promoters, or incorporating enzymes in various films on electrode surfaces [243].

4.1. Biosensing of Hydrogen Peroxide

Hydrogen peroxide (H2O2) is an important analyte due to its significance in various fields such as food processes, textile and paper industries, pharmaceutical research, environmental analysis, disinfecting and cleaning products, mineral processes, clinical laboratory, and medical diagnostics [195,330]. It is involved in several biological processes including cellular signaling, regulation of cell growth, apoptosis, immune activation, stomatal movement and root growth, and is a byproduct or substrate for oxidases [330,331]. H2O2 is a representative of reactive oxygen species in biological systems and its elevated levels have been associated with multiple disease conditions such as cancer, diabetes, asthma, cardiovascular, and oxidative stress-related diseases [332,333]. For instance, monitoring the levels of H2O2 in exhaled breaths provides reliable information about lung injuries, since it is considered a reliable indicator for lung diseases like asthma [334]. Its accumulation in plant cells proved to lead to specific gene expressions which enhance stress and pathogen tolerance [186]. On the other hand, H2O2 has exceptional properties such as oxidizing, gas formation on decomposition, source of energy and free radicals, and effects on biological processes [332]. As a result, it has found many industrial applications such as synthesis of organic compounds, liquid based fuel cells, mediator in pharmaceutical, clinical and environmental research, wastewater treatment, sterilization, and bleaching [330,335,336]. Eventually, H2O2 surfaced as an important contaminant in industrial wastes and products, and at high levels of exposure, it is an irritant to eyes, skin, the brain, and the gastrointestinal tract, causing detrimental effects like cell damage, cancer, and inflammatory diseases [331,332]. Therefore, reliable and economical methods for the determination of H2O2 are of great significance in biological, environmental, and clinical fields [186,337].
Horse radish peroxidase (HRP) is the enzyme that has been used in combination with metal-polymer nanocomposites for the detection of H2O2. HRP enzyme is a peroxidase that contains iron heme prosthetic groups in the polypeptide pockets and can catalyze a variety of substrates by one-electron oxidation when activated by peroxides [336,338]. Four kinds of reactions can be catalyzed by HRP, that is, peroxidation, oxidation, dismutation, and hydroxylation, and as a result, it has often been used in sensors [195]. However, direct electron transfer (DET) between the enzyme and electrode surface is quite slow, probably due to protein denaturation at the electrode surface and limited interaction of the enzymes active site and the electrodes surface due to the enzyme’s three-dimensional structure [339]. It has, therefore, been immobilized on matrices such as polymers, inorganic materials, and sol-gels to achieve DET [336]. In the sensing of H2O2, HRP is converted to its oxidized form, which is reduced at the electrode surface by DET, leading to an increased reduction current [335]. Immobilization of HRP on metal and metal oxide-polymer nanocomposites is a promising venture in the detection of H2O2, with wide detection ranges (20–13,700 μM [340], 10–10,000 μM [195], 8–12,000 μM [250]) and detection limits as low as 0.02 μM [333], as demonstrated in Table 1. These biosensors offer promising applications, especially in the agroindustry where the H2O2 concentrations are usually very low [186,333]. Also, the stability and reusability demonstrated by these biosensors is proof that they can be used as convenient tools for determining H2O2 in various settings; the biosensors retained activities of 100%, 90%, and 90% after 14 days [186], 1 month [336], and 8 weeks [334], respectively. In addition, an activity of 93.9% after 200 cycles [195] and relative standard deviation of 0.45% for 30 cycles [333] was observed.

4.2. Biosensing of Glucose

Glucose is one of the primary energy sources for plants and animals [343]. In humans, it is found in the blood stream and its levels are related to diabetes mellitus, which is a significant threat to human health [344]. Therefore, the measurement of glucose levels in blood has been used as an important clinical test for early diagnosis of diabetes mellitus [345,346]. The detection of glucose concentrations is vital in other areas such as biotechnology and food analysis as well [343,347,348]. Glucose biosensors represent the largest market for biosensors, accounting for roughly 85% of the biosensor market [349]. Glucose-based biosensors have been extensively fabricated and used to measure glucose due to their short response time, low cost, simplicity, and high sensitivity [345]. Glucose oxidase (GOx) is the most commonly used enzyme in the fabrication of glucose sensors due to its excellent stability, high catalytic properties, real time detection, and that it can recognize target molecules quickly and accurately in complicated systems [345,350,351]. GOx is an oxidoreductase that catalyzes oxidation of glucose to gluconolactone following the reduction of the flavine adenine dinucleotide (FAD) prosthetic group. The cofactor is then reoxidized in the second reaction and two protons and two electrons are transferred to molecular oxygen to yield gluconic acid and hydrogen peroxide [192,352,353,354]. The GOx sensor is based on the principle of monitoring the generation of hydronium ions after oxidation of glucose, and the increase in glucose concentrations is observed through potential differences of the electrode [355]. The detection of the signal is obtained from monitoring the increase of anodic current during oxidation of hydrogen peroxide or the decrease of cathodic current during reduction of dissolved oxygen [198,356,357].
However, the FAD redox center (active site) is deeply embedded in the protective protein shell and the structure immobilization matrix is a crucial aspect in terms of maximizing the enzyme activity [358,359]. Metal and metal oxide-polymer nanocomposites are promising immobilization matrices for GOx, especially for glucose sensing, as demonstrated by the wide linear ranges (10–20,000 μM [311], 200–19,900 μM [192], 500–30,000 μM [360], 1.2–40,000 μM [351], and 200–15,000 μM [349]) and low detection limits (0.0001 μM [308], 0.69 μM [356], 0.4 μM [361], 0.33 μM [187], 0.5 μM [159], and 0.9 μM [362]). The stability of GOx is also increased e.g., 100% [312], 90% [351,363], 90% [364] after 3 weeks, 1 month, and 2.5 months, respectively, and 92.6% [308], 94.7% [187], and 99.7% [344] after 160, 300, and 374 assays, respectively (Table 2). It is crucial to note that the concentration of glucose in human blood ranges between 4.1–5.9 mM and 2.0–30 mM for non-diabetic and diabetic patients, respectively, of which these concentrations lie outside the linear ranges of the reported biosensors [192,344]. For application of these biosensors in glucose detection in real human blood samples, sample dilution is required and might lead to dilution errors arising out of sample preparation. Interestingly, when the biosensors were used to measure glucose in actual human blood [189,343,357,360,365,366], urine [192], and beverages [312], good agreement with low relative standard deviations (RSD) was reported between the values obtained and those reported using other techniques. For example, Ren et al. [366] compared the results of glucose detection obtained by GOx/PtNPs/chitosan biosensor and the hospital biochemical analyzer, a relative standard deviation (RSD) of less than 4% was obtained in all three samples. Khumngern et al. [344] compared the GOx/AuNPs/Pty/Prussian blue modified screen-printed carbon electrode with the hexokinase method on 20 human blood samples for detection of glucose. It was observed that there was no significant difference between the two sets of data (p > 0.005). Luo et al. [85] performed a recovery test for glucose in serum samples at different concentrations using GOx/AuNPs/chitosan-modified gold electrode and an RSD of 4.6% was obtained for concentration ranges of 6–16 mM glucose, and recovery of 94–98% was recorded. Similarly, German et al. [192] reported recoveries in the range of 97–99% for glucose in human serum samples using PPy/GOx/AuNPs/graphite electrode with glucose concentrations in the range of 2–8.5 mM with four replicates for each concentration. Hence, these biosensors prove to be reliable for the detection of glucose in real samples.

4.3. Biosensing of Other Compounds in Human Blood

Other than glucose, several compounds can be measured in blood to diagnose and monitor health risks in humans. For example, cholesterol is a parameter used in the diagnosis of clinical lipid disorders, coronary heart disease, hypertension, and arteriosclerosis, and in the assessment of thrombosis and heart attack [374,375,376]. On the other hand, low levels of cholesterol are associated with conditions such as hypothyroidism, anemia, and malabsorption wasting syndrome [377,378]. Urea is another compound whose estimation is important in monitoring kidney functions and disorders associated with kidney failure. High levels of urea in blood serum or urine pose the risk of kidney failure, urinary tract obstruction, and gastrointestinal bleeding. On other hand, low levels are responsible for hepatic failure, nephritic syndrome, and cachexia [379,380,381,382]. It is also used in the food industry to adulterate milk, but beyond a certain limit it causes indigestion, renal failure, and certain cancers [383]. Xanthine is a purine base derived from guanine and adenosine-3-phosphate (ATP) catabolism in the muscle tissues of animals and its accumulation usually results in death [384,385]. The determination of its level in blood and tissue is essential for the diagnosis and management of diseases like gout, renal failure, hyperuricemia, and xanthinuria [386]. It is also an indicator for fish and meat spoilage and freshness determination [387,388]. Creatinine, a metabolic byproduct of amino acids that provide energy to muscles, is a clinical analyte in the diagnosis of kidney disease and muscle dysfunction [389,390]. Triglyceride, a component of very-low-density lipoproteins and chylomicrons is used as a clinical indicator of risk of heart disease and chronic obstructive pulmonary diseases such as bronchitis, bronchopneumia, and Sinusitis larystic [391,392].
Neurotransmitters such as dopamine, acetyl choline, and choline are also monitored for human health purposes. For example, dopamine is a neurotransmitter involved in emotion, reward, endocrine function, and motor control, and its dysregulation is associated with mood and attention deficit hyperactive disorders, schizospermia, and neurodegenerative diseases like Alzheimer’s and Parkinson’s [328]. Acetyl choline is another neurotransmitter found in peripheral and central nerve systems of mammals and its dysregulation in the brain is associated with disorders such as Alzheimer’s, Parkinson’s, and Myasthenia Gravis [393,394].
The detection of the above compounds is fundamental; hence, rapid and accurate measurement systems for their detection are a necessity. Various enzymes have been immobilized on metal and metal-oxide-polymer nanocomposites for the detection of these compounds, as outlined in Table 3. For example, the cholesterol oxidase (Chx) enzyme has been immobilized on various nanoparticles such as Au-NPs [375,395], ZnO-NPs [377], CeO2-NPs [376], SnO2-NPs [374], and NiFe2O4-CuO-FeO-NPs [378] embedded in chitosan polymer for sensing cholesterol. The biosensors showed good reproducibility and reusability. For example, the Chx/CeO2-NPs/chitosan biosensor retained 100% and 90% of its initial activity after 10 assays and 7 weeks, respectively [376], while the Chx/AuNPs/PANI/chitosan biosensor retained 100%, 97%, and 90% of its activity after 20 assays, 2 weeks, and 3 weeks, respectively [395]. However, the linear ranges reported by these biosensors are quite low compared to the concentration range of cholesterol in human blood; hence, sample dilution is required prior to sample analysis. Urease has been immobilized on various nanocomposites such as AuNPs/Boltorn [382], MNPs/chitosan [379,396], ZnO-NPs/polypyrrole/polyamide 6 [383], ZnO-NPs/chitosan [381], Ce3O4-NPs/chitosan [380], and CuO-NPs/PANI/nafion [390] for the detection of urea. The biosensors demonstrate short response times and good stability and reusability (Table 3).

4.4. Biosensing of Pesticides and Other Organic Pollutants

In a bid to increase productivity in agriculture, pest control has been achieved through the use of pesticides [403]. As a result, a large amount of pesticide residues and their metabolites have ended up in the water, soil, and food becoming some of the most important environmental pollutants [404]. These compounds are not only persistent, but highly toxic to humans and aquatic life, and also get more concentrated up the food chain ladder [405]. Their toxicity is due to the irreversible inhibition of the enzyme acetylcholinesterase (AChE), which is responsible for the transmission of nerve impulses to muscles and neuromuscular cells in living organisms. This results in the accumulation of acetylcholine neurotransmitter, leading to respiratory malfunctions, heart attack, and even death [406,407,408].
The main enzyme reported in the fabrication of biosensors for the detection of pesticides is AChE, and the biosensors are based on the inhibition of this enzyme. The enzyme inhibition is determined by amperometric/voltametric detection of thiocoline, an enzymatic oxidation product of acetylthiocholine, at the electrode [405,409]. As observed from Table 4, immobilization of AChE on metal-polymer nanocomposites for the detection of pesticides is a promising venture, as these biosensors can detect amounts as low as 0.3 nM (Malathion) [408], 3 nM (carbofuran) [409], 0.1 nM (chlorpyrifos) [404], 21 nM (oxamyl) [405], and 0.003 nM (paraoxon). The biosensors are also reported to be stable and durable; for example, Du et al. reported 100% and 90% retained activity after 10 and 30 days, respectively, while Kestwal et al. reported 96% and 94% after 20 and 30 days, respectively.
Other enzymes such as sulfite oxidase [410], phenol oxidase [193], HRP [411], laccase [324,412], and tyrosinase [413,414] have also been immobilized on metal-polymer nanocomposites for biosensing of sulfite, phenols, and other organic pollutants such as catechol, bisphenols, p-cresol, and pyrogallol (Table 4). Since most of these pollutants are toxic even at concentrations as low as ng/L, highly sensitive biosensors are required to accurately detect them in the environment [411,414]. Enzyme-based metal and metal oxide-polymer nanocomposite biosensors seem to be up to the task. For example, the Laccase/AuNPs/polyethylene (PEI) biosensor could detect as low as 30 nM catechol, 30 nM guaiacol, 140 nM pyrogallol, and 210 nM hydroquinone [412], while the tyrosinase/AuNPs/dihexadecylphosphate (DHP) biosensor could detect 170 nM catechol [413]. Moreover, these biosensors can be used repetitively over long periods of time. For example, the tyrosinase/AuNP/DHP biosensor only lost 7% of its original activity after 240 measurements of catechol over a period of 1 month [413], and the phenol oxidase/AuNPs/polypyrrole (PPy) biosensor retained 100% activity after 25 measurements of phenol [193].
Table 4. Application of enzyme-nanoparticle-polymer composites in fabrication of biosensors for detection of pesticides and organic pollutants in the environment.
Table 4. Application of enzyme-nanoparticle-polymer composites in fabrication of biosensors for detection of pesticides and organic pollutants in the environment.
Nanocomposite (NC)EnzymeImmobilization MethodElectrode UsedDetected CompoundDetection Range (μM)Limit of Detection (μM)Response Time (s)ReusabilityRef.
AuNPs/PAN membraneAChESequential layer by layer loading of PAN, AuNPs, and AChEPlatinumParaoxon3.6 × 10−7–3.6 × 10−4 *2.69 × 10−7 *590.2% and 75% (9 assays and 20 days, respectively), RSD of 1.68% and 3.5% (6 assays and 6 biosensors)[406]
AuNPs/chitosanAChEElectrochemical depositionPlatinumMalathion and monocrotophos0.003–0.3 and
6.1–60.5 *
0.003 * 100% and 90% (10 and 30 days respectively), RSD of 3.4% and 2.3% (5 biosensors and assays, respectively)[415]
AuNPs/polyethyleneimine (PEI)LaccaseDrop-casting of laccase/AuNPs/PEI solution on GCEGCECatechol0.36–11.000.03 80% (150 assays over 90 days)[412]
Guaiacol0.79–17.420.03
Pyrogallol1.74–19.600.14
Hydroquinone2.90–22.000.21
AuNPs/fenugreek hydrogel-agaroseAChEDrop-casting a homogeneous mixture of agarose, fenugreek hydrogel, AChE, and AuNPs solution-Carbofuran0.002–0.010.002 96% and 94% (20 and 30 days, respectively)[405]
Oxamyl0.01–0.10.021
Methomyl0.1–0.50.113
Carbaryl0.2–10.236
AuNPs/dihexadecylphosphate (DHP)TyrosinaseDrop-casting of a mixture of tyrosinase, AuNPs, and DHPGCECatechol2.5–9500.17 93% (240 assays over 1 month), RSD of 3.8% (3 biosensors)[413]
AgNPs/carboxymethyl cellulose (CMC)/cellulose nanofiberLaccaseElectrospinning of cellulose nanofiber, adsorption of CMC, immersion in AgNO3 solution, incubation in laccase solutionGCECatechol4.98–36501.64 97.6% (3 weeks), RSD of 3.41%, and 1.57% (4 assays and 5 biosensors, respectively)[324]
MNPs/PGMAHRPSelf-assembled deposition of cysteamine-modified electrode with MNPs/PGMA solution followed by HRP solutionGoldp-Cresol500–70026487% and 85% (50 days and 12 assays, respectively)[411]
Aminophenol500–3500133
Catechol500–11,000467
Phenol500–8500283
Pyrogallol500–15,000485
MNPs/Pin5COOHAChEGrafting-from of NC by CV followed by drop-casting of AChE solution on modified electrodeGCEMalathion−0.060.0015 50% (70 days)[404]
Chlorpyrifos0.0015–0.070.0001
MNP/chitosanAChEDrop-casting a mixture of MNPs and chitosan solution followed by AChEGCECarbofuran0.005–0.090.0036900RSD 4.3% and 5.4% (5 assays and biosensors, respectively)[409]
MNPs/chitosanAChEDrop-casting a mixture of MNPs, chitosan, AChE, and glutaraldehydeScreen-printed electrodeMalathion0.0005–0.020.0003 [408]
AuNPs/PPyPhenol oxidaseSequential deposition of HAuCl4, enzyme, and pyrroleGCEPhenol0.05–700.0310100% and 68% (25 and 100 assays, respectively), RSD of 1.36% (6 biosensors)[193]
AuNPs/PA6-poly(allylamine hydrochloride) (PAH)TyrosinaseElectrospun PA6-PAH on FTO, immersion in AuNPs solution, drop-casted tyrosinase solution on modified electrodeFluorine-doped tin oxide (FTO)Bisphenol A0.05–200.011 [414]
PtNPs/PPySulfite oxidaseSequential electropolymerization of K2PtCl6 and pyrrole and immersion in enzyme solutionPlatinumSulfite0.75–650.012596.5%, 92.5%, and 88.2% (10, 11 and 12 weeks, respectively). RSD of 3.2% (9 assays)[410]
* Values convert from mM units.

4.5. Perspectives on Enzyme-Nanoparticle-Polymer Composite Electrodes

The majority of the biocatalysts have demonstrated good response times for the biosensors and good detection ranges for the different analytes. However, for most of these electrodes, storage stability and continuous reusability studies have not been carried out and systematized. Hence, there is need for a detailed storage and reusability study for each of the electrodes to determine their shelf life prior to industrial applications. Although these electrodes display wide linear ranges, their ranges lie outside the occurrences of most of the analytes in real life samples such as blood and wastewater. Therefore, sample dilution would be required for their application, which would lead to dilution errors during sample preparations. Also, the analytes have been detected in distilled water or buffers, which do not depict the actual clinical or environmental conditions; hence, we cannot ascertain the relevance of the electrodes in natural systems.

5. Application of Enzyme-Nanoparticle-Polymer Composites in Wastewater Treatment

In recent decades, the global community has increasingly recognized the formidable challenge posed by water pollution arising from the unregulated release of municipal and industrial waste [416,417]. Many industries including petrochemical, paints and explosives, food, pharmaceutical, leather and textile, pulp and paper, and cosmetics have contributed to this cause [418,419]. These discharges cause serious problems to aquatic life due to their high biochemical oxygen demand (BOD), chemical oxygen demand, and blockage of sunlight [420,421].
One of the industries producing the highest level of toxic chemicals from dyeing, printing, and finishing is the leather and textile industry [416]. The conversion of skin into leather in textile industries generates huge amounts of wastewater containing a variety of organic and inorganic chemicals such as dyes, neutral salts, phenols, and biogenic matter of skins [422,423]. The complex aromatic structures of these chemicals, especially the dyes, make them highly soluble in water and stable against light, aerobic decomposition, and oxidizing reagents [424]. Therefore, their accumulation leads to serious environmental concerns for aquatic life and human beings due to their adverse effects of toxicity, carcinogenicity, and mutagenicity [425]. Another industrial sector that has developed rapidly in the last century is the pesticide industry, as it is an important component of modern global agricultural systems for controlling pests and increasing crop yield [426]. These pesticides are applied in much higher doses than those required to kill the pests, and end up accumulating in water bodies via run off and percolation (Figure 4) [427]. Unfortunately, these agrochemical residues not only pollute the aquatic systems and damage biodiversity, they cause serious health hazards to humans and may even directly or indirectly lead to death [428,429]. Moreover, these compounds have very long half-lives and can remain in the environment for several decades [403,430].
The growth of the pharmaceutical industry (veterinary and human medicines) in the past years has also led to rising amounts of drugs, antibiotics, and hormones. These medicines are not fully metabolized by living organisms and when these end up in wastewater treatment plants, they are difficult to biodegrade, since most of them are fat soluble [431,432,433]. For example a study conducted by Joss et al. [434] indicated that biological degradation of pharmaceuticals using activated sewage sludge from municipal wastewater could only degrade 4 out of 35 compounds by over 90% and 17 compounds by less than 50%. These compounds have increased in the environment due to their increased consumption and direct discharge into the environment, as illustrated in Figure 4. The presence of pharmaceuticals, cosmetics, and their metabolites in municipal waste and industrial effluents presents a significant challenge, as these compounds cannot be effectively eliminated using conventional techniques, and consequently are released to the receiving environment [435,436]. While in the environment, they accumulate or transform into metabolites under certain environmental conditions, and these secondary metabolites may even be more toxic than the parent compounds [427,437]. These make pathogenic organisms develop resistance against them over time, which is a high risk to human health [438].
Figure 4. A schematic presentation of the sources, transport, and exposure routes of pharmaceuticals, personal care products, and pesticides in water systems. Obtained with permission from Okoye et al., 2022 [439].
Figure 4. A schematic presentation of the sources, transport, and exposure routes of pharmaceuticals, personal care products, and pesticides in water systems. Obtained with permission from Okoye et al., 2022 [439].
Polymers 15 03492 g004
The continued release, spread, and accumulation of persistent organic pollutants in the water environment from these industries, including polychlorinated biphenyls and polycyclic aromatic hydrocarbons from the petrochemical industries, have become a major threat to human health due to their toxic, mutagenic, and carcinogenic properties [440,441,442]. The emission of these pollutants occurs at the manufacturing stage, after consumption and disposal of unused products (Figure 4). These products are hard to be tracked or controlled in most situations and are resistant to natural biodegradation [427,443]. Most of these compounds are phenolic and, therefore, bio-recalcitrant, carcinogenic, and easily accumulate in plants and animals. They should, therefore, be removed prior to wastewater discharge [431,444,445].
Different water treatment technologies such as photochemical degradation, biodegradation, electrochemical degradation, reverse osmosis, and membrane separation have been used to get rid of these pollutants. However, these techniques are costly, consist of complicated procedures, do not entirely remove the pollutants, and lead to secondary contaminants that also need to be redisposed of [446,447]. Enzymatic treatments of these pollutants have received great attention due to several advantages compared to physical and chemical treatments, such as mild operating conditions and high catalytic efficiency without harsh side effects [448,449]. Hence, the use of biocatalysts in wastewater treatment has gained momentum due to their ability to target a wide range of pollutants [450]. Enzymes immobilized onto supports are often used in the treatment of wastewaters to ensure improved thermal and pH stability and repeatability, which is rarely achieved with free enzymes [451]. Various pollutants including drugs, dyes, pesticides, polycyclic aromatic hydrocarbons (PAHs), and even heavy metals have been degraded using enzyme/metal-polymer biocatalysts, as demonstrated in Figure 5. Oxidase and peroxidase enzymes from different sources have been immobilized on metal and metal oxide-polymer composites and used in the degradation of pollutants, as observed in Figure 5. This part of the review focuses on the application of oxidases and peroxidases immobilized on metal and metal oxide-polymer composites in wastewater treatment, especially the degradation of dyes, pesticides, pharmaceuticals, and phenolic compounds.

5.1. Laccase-Based Nanocomposite Biocatalysts for Degradation of Pollutants

Laccase is the most explored enzyme in wastewater treatment due to its ability to degrade a wide range of micro pollutants including dyes, pharmaceuticals, and endocrine-disrupting chemicals [452,453,454]. Unlike other oxidoreductases, laccase does not require hydrogen peroxide or other cofactors for substrate cleavage [455,456,457] and its range of compounds for oxidation can be increased with redox mediators [458,459]. Laccase-based composite biocatalysts show great potential in wastewater treatment as they have demonstrated high pollutant degradation rates with high reusability (Table 5). For example, Laccase/Fe2O3/PEI biocatalyst completely degraded sulfa drugs (Sulfadiazine, Sulfamethazine and Sulfamethoxazole) within 30 min and could still degrade 82.8% after 10 cycles in the same time frame [438]. Laccase/Ca-alginate beads degraded 99% bisphenol A [433] and dyes (aniline purple–86%, lanset grey G–85%, and reactive black 5–80%) [460] in 2 h and 24 h, respectively.

5.2. Horse Radish Peroxidase (HRP)-Based Nanocomposite Biocatalysts for Degradation of Pollutants

Another commonly explored peroxidase on nanoparticle-polymer composite materials is horse radish peroxidase (HRP), due to its ability to oxidize a wide range of phenolic compounds in the presence of hydrogen peroxide [473]. It oxidizes phenolic compounds by adding hydrogen peroxide to form corresponding radicals which spontaneously interact to form insoluble polymers that can be easily removed from the wastewater [474]. HRP/nanoparticle-polymer composite biocatalysts have been explored in the degradation of phenols, dyes, and endocrine-disrupting compounds, as illustrated in Table 6. For example, HRP/MNPs/polyvinyl alcohol/poly acrylic acid could completely degrade estrone after 40 min [432], and HRP/TiO2/polydopamine completely removed 2,4-dicholorphenol in Zhaohe wastewater samples in only 30 min [190]. Interestingly, the HRP/TiO2/polydopamine biocatalyst retained 100% and 90% degradation activity after 15 and 25 reuses, respectively.

5.3. Other Oxidase and Peroxidase-Based Nanocomposite Biocatalysts for Degradation of Pollutants

Other enzymes such as chloroperoxidase, manganese peroxidase, and lignin peroxidase immobilized on composite materials, though not very popular, prove that they can offer wonderful materials for pollutant degradation (Table 7). For example, when lignin peroxidase was immobilized on MNPs@SiO2/polydopamine, it was able to degrade tetracycline and other phenolics such as 5-chlorophenol, phenol, and dibutyl phthalate completely within 24 h [447]. Manganese peroxidase immobilized on MNPs/chitosan degraded 96% of methylene blue in synthetic wastewater in just 50 min [417], glucose oxidase immobilized on NiFe2O4/tannin could degrade 98.6% of indigo carmine in presence of UV light within 90 min [446], and chloroperoxidase/TiO2/polydopamine nanocomposites degraded over 95% of aniline blue and crystal violet in 2 min [190].

5.4. Current Limitations

It is worth noting that the majority of the research studies carried out on the degradation of organic pollutants by these biocatalysts have been carried out in buffer solutions or deionized water, which does not depict the actual environmental conditions of the pollutants in wastewater systems and industrial settings. Moreover, most of the studies have been carried out in batch studies, and yet most industries that release these pollutants operate in continuous cycles. The results, therefore, reported in the literature, such as degradation time and operating conditions such as pH and temperature, may be different if these experiments were carried out in real wastewater samples under industrial conditions. For example, when Le et al. [419] (Table 5) used laccase/Fe2O3/Cu-alginate beads to degrade triclosan and RBBR dye in acetate buffer, it required 8 h to remove 89.6% and 75.8%, respectively. However, when the same biocatalyst was used to degrade the same pollutant in cooling system wastewater, the percentage degradation dropped to 53.2% after 8 h and 55% after 25 h for triclosan and RBBR, respectively; and when used to degrade RBBR in sludge pond outlet wastewater, the percentage degradation further dropped to only 35% after 25 h. Similarly, laccase/Ca-alginate beads (Table 5) were tested to remove reactive red 180 and reactive blue 21 dyes from a real textile effluent, and it required up to 11 days to remove 67.2% and 88.05%, respectively [466]. When laccase was entrapped in Cu-alginate beads, it required 24 h to remove only 38% of acid dye in synthetic wastewater [468] (Table 5). Hence, it is necessary to investigate the utilization of these biocatalysts under actual industrial and environmental conditions to determine their suitability in wastewater treatment. This is crucial because industrial effluents involve numerous factors, such as biological and chemical oxygen demand, that must be considered. In an illustrative case, Sondhi et al. [423] employed laccase/Cu-alginate beads in treating textile effluent through a continuous flow packed bead reactor. Their findings indicated a substantial reduction in color (66%), biological oxygen demand (90%), and chemical oxygen demand (98%) at equilibrium, reflecting the effectiveness of this approach in closely resembling real industrial conditions.

6. Conclusions

Numerous strategies have been investigated to enhance the stability and reusability of enzymes, with particular emphasis on enzyme immobilization onto a supportive matrix. Various support materials and immobilization techniques have been examined for different enzymes. Notably, the use of metal-polymer composites has garnered attention in immobilization methods, including physical entrapment and enzyme adsorption. These approaches can be executed under mild conditions with minimal chemical usage, employing straightforward techniques.
Polymers rich in amines, hydroxyls, and carboxylic groups, such as chitosan and alginate, have been identified as suitable platforms for enzyme entrapment through crosslinking, thus enhancing enzyme stability. The polymer’s ability to form hydrogen bonds also facilitates the creation of nanopores, enabling efficient diffusion of substrates and products to and from the enzyme, resulting in minimal loss of enzyme activity. Additionally, in sensing applications, metal and metal oxide nanoparticles facilitate reversible electron transfer between the enzyme and electrode surface, enhancing the selectivity and response of sensing devices.
Consequently, enzyme-metal and metal oxide-polymer composites have been explored in the sensing of various substances, including hydrogen peroxide in surface water, glucose, urea, xanthine, cholesterol, and dopamine in blood, as well as pesticides and herbicides in freshwater systems. The ability of these nanocomposites to form enzyme-entrapped beads using simple techniques is promising, as these beads can be employed in continuous operations resembling industrial settings. It is no surprise that chitosan and alginate polymers have been extensively studied for entrapping enzymes and nanoparticles in wastewater treatment. The inclusion of nanoparticles in the beads provides a larger surface area for enzyme adsorption, enabling high enzyme loads per bead. These beads can be utilized in diverse devices such as packed reactors, filters, and fuel cells. Therefore, enzyme-metal/metal oxide-polymer composites offer promising applications in both wastewater treatment and biosensors.

Author Contributions

Conceptualization, H.D.K., U.F. and H.B.; validation, H.B., U.F. and N.H.H.; investigation, H.D.K.; resources, H.B., U.F. and N.H.H.; writing—original draft preparation, H.D.K.; writing—review and editing, H.D.K., H.B., U.F. and N.H.H.; supervision, H.B. and U.F.; project administration, H.B. and U.F.; funding acquisition, N.H.H., H.B. and H.D.K. All authors have read and agreed to the published version of the manuscript.

Funding

This study was supported by National Research Foundation (NRF) of South Africa [MND210426597525 and CSRP2204204025] and Margaret McNamara Education grants (2023). This work further received support from the Federal Ministry of Education, Science and Research (BMBWF) through Austria’s Agency for Education and Internationalization (OeAD) [Grant Number: Africa UNINET P056].

Institutional Review Board Statement

Not applicable.

Data Availability Statement

Data sharing is not applicable to this article as no new data were created or analyzed in this study.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Roy Choudhury, A.K. Introduction to enzymes. In Sustainable Technologies for Fashion and Textiles; Elsevier: Amsterdam, The Netherlands, 2020; pp. 75–90. [Google Scholar] [CrossRef]
  2. Kornbrust, B.A.; Forman, T.; Matveeva, I. Applications of enzymes in breadmaking. In Breadmaking; Elsevier: Amsterdam, The Netherlands, 2012; pp. 470–498. [Google Scholar] [CrossRef]
  3. Copeland, R.A. Enzymes: A Practical Introduction to Structure, Mechanism, and Data Analysis, 2nd ed.; Wiley: New York, NY, USA, 2000. [Google Scholar]
  4. Auld, D.S. Zinc coordination sphere in biochemical zinc sites. In Zinc Biochemistry, Physiology, and Homeostasis: Recent Insights and Current Trends; Maret, W., Ed.; Springer: Dordrecht, The Netherlands, 2001; pp. 85–127. [Google Scholar] [CrossRef]
  5. Bjerre, J.; Rousseau, C.; Marinescu, L.; Bols, M. Artificial enzymes, “Chemzymes”: Current state and perspectives. Appl. Microbiol. Biotechnol. 2008, 81, 1–11. [Google Scholar] [CrossRef] [PubMed]
  6. Robinson, P.K. Enzymes: Principles and biotechnological applications. Essays Biochem. 2015, 59, 1–41. [Google Scholar] [CrossRef]
  7. Copeland, R.A. Enzymes; Wiley: Hoboken, NJ, USA, 2002. [Google Scholar]
  8. Chapman, J.; Ismail, A.E.; Dinu, C.Z. Industrial Applications of Enzymes: Recent Advances, Techniques, and Outlooks. Catalysts 2018, 8, 238. [Google Scholar] [CrossRef]
  9. Parkin, G. Synthetic Analogues Relevant to the Structure and Function of Zinc Enzymes. Chem. Rev. 2004, 104, 699–768. [Google Scholar] [CrossRef]
  10. Glusker, J.P. Structural Aspects of Metal Liganding to Functional Groups in Proteins. In Advances in Protein Chemistry; Anfinsen, C.B., Edsall, J.T., Richards, F.M., Eisenberg, D.S., Eds.; Academic Press: Cambridge, MA, USA, 1991; Volume 42, pp. 1–76. [Google Scholar] [CrossRef]
  11. Anandan, A.; Vrielink, A. Structure and function of lipid A–modifying enzymes. Ann. N. Y. Acad. Sci. 2020, 1459, 19–37. [Google Scholar] [CrossRef]
  12. Whitaker, J.R.; Voragen, A.G.J.; Wong, D.W.S. (Eds.) Handbook of Food Enzymology; Marcel Dekker: New York, NY, USA, 2003. [Google Scholar]
  13. Stroppolo, M.E.; Falconi, M.; Caccuri, A.M.; Desideri, A. Superefficient enzymes. Cell. Mol. Life Sci. 2001, 58, 1451–1460. [Google Scholar] [CrossRef]
  14. Martínez, A.T.; Ruiz-Dueñas, F.J.; Camarero, S.; Serrano, A.; Linde, D.; Vind, J.; Tovborg, M.; Herold-Majumdar, O.M.; Hofrichter, M. Oxidoreductases on their way to industrial biotransformations. Biotechnol. Adv. 2017, 35, 815–831. [Google Scholar] [CrossRef] [PubMed]
  15. Poulos, T.L. Thirty years of heme peroxidase structural biology. Arch. Biochem. Biophys. 2010, 500, 3–12. [Google Scholar] [CrossRef] [PubMed]
  16. Mäkelä, M.R.; Bredeweg, E.L.; Magnuson, J.K.; Baker, S.E.; de Vries, R.P.; Hildén, K. Fungal Ligninolytic Enzymes and Their Applications. Microbiol. Spectr. 2016, 4, 10–1128. [Google Scholar] [CrossRef]
  17. Poulos, T.L. Heme Enzyme Structure and Function. Chem. Rev. 2014, 114, 3919–3962. [Google Scholar] [CrossRef]
  18. Chen, C.; Li, T. Bacterial dye-decolorizing peroxidases: Biochemical properties and biotechnological opportunities. Phys. Sci. Rev. 2016, 1, 20160051. [Google Scholar] [CrossRef]
  19. Dubrovskaya, E.; Pozdnyakova, N.; Golubev, S.; Muratova, A.; Grinev, V.; Bondarenkova, A.; Turkovskaya, O. Peroxidases from root exudates of Medicago sativa and Sorghum bicolor: Catalytic properties and involvement in PAH degradation. Chemosphere 2017, 169, 224–232. [Google Scholar] [CrossRef] [PubMed]
  20. Lüthje, S.; Meisrimler, C.-N.; Hopff, D.; Möller, B. Phylogeny, topology, structure and functions of membrane-bound class III peroxidases in vascular plants. Phytochemistry 2011, 72, 1124–1135. [Google Scholar] [CrossRef] [PubMed]
  21. Hooda, V.; Gundala, P.B.; Chinthala, P. Sequence analysis and homology modeling of peroxidase from Medicago sativa. Bioinformation 2012, 8, 974. [Google Scholar] [CrossRef]
  22. Gumiero, A.; Murphy, E.J.; Metcalfe, C.L.; Moody, P.C.E.; Raven, E.L. An analysis of substrate binding interactions in the heme peroxidase enzymes: A structural perspective. Arch. Biochem. Biophys. 2010, 500, 13–20. [Google Scholar] [CrossRef] [PubMed]
  23. Taranto, F.; Pasqualone, A.; Mangini, G.; Tripodi, P.; Miazzi, M.M.; Pavan, S.; Montemurro, C. Polyphenol Oxidases in Crops: Biochemical, Physiological and Genetic Aspects. Int. J. Mol. Sci. 2017, 18, 377. [Google Scholar] [CrossRef]
  24. Plácido, J.; Capareda, S. Ligninolytic enzymes: A biotechnological alternative for bioethanol production. Bioresour. Bioprocess. 2015, 2, 23. [Google Scholar] [CrossRef]
  25. Xu, S.; Huang, X.; Chen, Y.; Liu, Y.; Zhao, W.; Sun, Z.; Zhu, Y.; Liu, X.; Wong, C.-P. Silver Nanoparticle-Enzyme Composite Films for Hydrogen Peroxide Detection. ACS Appl. Nano Mater. 2019, 2, 5910–5921. [Google Scholar] [CrossRef]
  26. Meunier, B. 8.11—Heme-peroxidases. In Comprehensive Coordination Chemistry II; McCleverty, J.A., Meyer, T.J., Eds.; Pergamon: Oxford, UK, 2003; pp. 261–280. [Google Scholar] [CrossRef]
  27. Turner, N.J. 7.12 Oxidation: Oxidases. In Comprehensive Chirality; Carreira, E.M., Yamamoto, H., Eds.; Elsevier: Amsterdam, The Netherlands, 2012; pp. 256–274. [Google Scholar] [CrossRef]
  28. Wahart, A.J.C.; Staniland, J.; Miller, G.J.; Cosgrove, S.C. Oxidase enzymes as sustainable oxidation catalysts. R. Soc. Open Sci. 2022, 9, 211572. [Google Scholar] [CrossRef]
  29. Dijkman, W.P.; de Gonzalo, G.; Mattevi, A.; Fraaije, M.W. Flavoprotein oxidases: Classification and applications. Appl. Microbiol. Biotechnol. 2013, 97, 5177–5188. [Google Scholar] [CrossRef]
  30. Phale, P.S.; Sharma, A.; Gautam, K. 11—Microbial degradation of xenobiotics like aromatic pollutants from the terrestrial environments. In Pharmaceuticals and Personal Care Products: Waste Management and Treatment Technology; Prasad, M.N.V., Vithanage, M., Kapley, A., Eds.; Butterworth-Heinemann: Oxford, UK, 2019; pp. 259–278. [Google Scholar] [CrossRef]
  31. Jabłońska, J.; Tawfik, D.S. Innovation and tinkering in the evolution of oxidases. Protein Sci. 2022, 31, e4310. [Google Scholar] [CrossRef] [PubMed]
  32. Fetzner, S.; Steiner, R.A. Cofactor-independent oxidases and oxygenases. Appl. Microbiol. Biotechnol. 2010, 86, 791–804. [Google Scholar] [CrossRef] [PubMed]
  33. Choi, J.-M.; Han, S.-S.; Kim, H.-S. Industrial applications of enzyme biocatalysis: Current status and future aspects. Biotechnol. Adv. 2015, 33, 1443–1454. [Google Scholar] [CrossRef]
  34. Gardimalla, H.M.R.; Mandal, D.; Stevens, P.D.; Yen, M.; Gao, Y. Superparamagnetic nanoparticle -supported enzymatic resolution of racemic carboxylates. Chem. Commun. 2005, 4432–4434. [Google Scholar] [CrossRef]
  35. Kelly, S.A.; Pohle, S.; Wharry, S.; Mix, S.; Allen, C.C.R.; Moody, T.S.; Gilmore, B.F. Application of ω-Transaminases in the Pharmaceutical Industry. Chem. Rev. 2018, 118, 349–367. [Google Scholar] [CrossRef]
  36. Akin, D.E. Linen Most Useful: Perspectives on Structure, Chemistry, and Enzymes for Retting Flax. Int. Sch. Res. Not. 2013, 2013, 186534. [Google Scholar] [CrossRef] [PubMed]
  37. Madhu, A.; Chakraborty, J.N. Developments in application of enzymes for textile processing. J. Clean. Prod. 2017, 145, 114–133. [Google Scholar] [CrossRef]
  38. Chatha, S.A.S.; Asgher, M.; Iqbal, H.M.N. Enzyme-based solutions for textile processing and dye contaminant biodegradation—A review. Environ. Sci. Pollut. Res. 2017, 24, 14005–14018. [Google Scholar] [CrossRef] [PubMed]
  39. Ugliano, M. Enzymes in Winemaking. In Wine Chemistry and Biochemistry; Moreno-Arribas, M.V., Polo, M.C., Eds.; Springer: New York, NY, USA, 2009; pp. 103–126. [Google Scholar] [CrossRef]
  40. Claus, H.; Mojsov, K. Enzymes for Wine Fermentation: Current and Perspective Applications. Fermentation 2018, 4, 52. [Google Scholar] [CrossRef]
  41. Espejo, F. Role of commercial enzymes in wine production: A critical review of recent research. J. Food Sci. Technol. 2021, 58, 9–21. [Google Scholar] [CrossRef]
  42. Yang, S.; Yang, B.; Duan, C.; Fuller, D.A.; Wang, X.; Chowdhury, S.P.; Stavik, J.; Zhang, H.; Ni, Y. Applications of enzymatic technologies to the production of high-quality dissolving pulp: A review. Bioresour. Technol. 2019, 281, 440–448. [Google Scholar] [CrossRef]
  43. Walia, A.; Guleria, S.; Mehta, P.; Chauhan, A.; Parkash, J. Microbial xylanases and their industrial application in pulp and paper biobleaching: A review. 3 Biotech 2017, 7, 11. [Google Scholar] [CrossRef] [PubMed]
  44. Kumar, V.; Marín-Navarro, J.; Shukla, P. Thermostable microbial xylanases for pulp and paper industries: Trends, applications and further perspectives. World J. Microbiol. Biotechnol. 2016, 32, 34. [Google Scholar] [CrossRef] [PubMed]
  45. Kües, U. Fungal enzymes for environmental management. Curr. Opin. Biotechnol. 2015, 33, 268–278. [Google Scholar] [CrossRef] [PubMed]
  46. Lim, J.; Mohamad, Z. Enzymes immobilized polymeric supports for wastewater treatment application: A short review. Mater. Today Proc. 2022, 65, 2946–2952. [Google Scholar] [CrossRef]
  47. Jun, L.Y.; Yon, L.S.; Mubarak, N.M.; Bing, C.H.; Pan, S.; Danquah, M.K.; Abdullah, E.C.; Khalid, M. An overview of immobilized enzyme technologies for dye and phenolic removal from wastewater. J. Environ. Chem. Eng. 2019, 7, 102961. [Google Scholar] [CrossRef]
  48. Kurbanoglu, S.; Ozkan, S.A.; Merkoçi, A. Nanomaterials-based enzyme electrochemical biosensors operating through inhibition for biosensing applications. Biosens. Bioelectron. 2017, 89, 886–898. [Google Scholar] [CrossRef]
  49. Lee, H.; Hong, Y.J.; Baik, S.; Hyeon, T.; Kim, D.-H. Enzyme-Based Glucose Sensor: From Invasive to Wearable Device. Adv. Healthc. Mater. 2018, 7, 1701150. [Google Scholar] [CrossRef]
  50. Nguyen, H.H.; Lee, S.H.; Lee, U.J.; Fermin, C.D.; Kim, M. Immobilized Enzymes in Biosensor Applications. Materials 2019, 12, 121. [Google Scholar] [CrossRef]
  51. Li, G.; Wang, J.; Reetz, M.T. Biocatalysts for the pharmaceutical industry created by structure-guided directed evolution of stereoselective enzymes. Bioorganic Med. Chem. 2018, 26, 1241–1251. [Google Scholar] [CrossRef]
  52. Kim, J.; Lee, J.; Na, H.B.; Kim, B.C.; Youn, J.K.; Kwak, J.H.; Moon, K.; Lee, E.; Kim, J.; Park, J.; et al. A Magnetically Separable, Highly Stable Enzyme System Based on Nanocomposites of Enzymes and Magnetic Nanoparticles Shipped in Hierarchically Ordered, Mesocellular, Mesoporous Silica. Small 2005, 1, 1203–1207. [Google Scholar] [CrossRef] [PubMed]
  53. Kim, J.; Grate, J.W. Single-Enzyme Nanoparticles Armored by a Nanometer-Scale Organic/Inorganic Network. Nano Lett. 2003, 3, 1219–1222. [Google Scholar] [CrossRef]
  54. Fernandez-Lafuente, R. Stabilization of multimeric enzymes: Strategies to prevent subunit dissociation. Enzym. Microb. Technol. 2009, 45, 405–418. [Google Scholar] [CrossRef]
  55. Kim, J.; Grate, J.W.; Wang, P. Nanostructures for enzyme stabilization. Chem. Eng. Sci. 2006, 61, 1017–1026. [Google Scholar] [CrossRef]
  56. Gupta, M.N.; Roy, I. Enzymes in organic media. Eur. J. Biochem. 2004, 271, 2575–2583. [Google Scholar] [CrossRef]
  57. Hari Krishna, S. Developments and trends in enzyme catalysis in nonconventional media. Biotechnol. Adv. 2002, 20, 239–267. [Google Scholar] [CrossRef]
  58. Sheldon, R.A.; Pereira, P.C. Biocatalysis engineering: The big picture. Chem. Soc. Rev. 2017, 46, 2678–2691. [Google Scholar] [CrossRef]
  59. Silva, C.; Martins, M.; Jing, S.; Fu, J.; Cavaco-Paulo, A. Practical insights on enzyme stabilization. Crit. Rev. Biotechnol. 2018, 38, 335–350. [Google Scholar] [CrossRef]
  60. Steiner, K.; Schwab, H. Recent advances in rational approaches for enzyme engineering. Comput. Struct. Biotechnol. J. 2012, 2, e201209010. [Google Scholar] [CrossRef]
  61. Mazurenko, S.; Prokop, Z.; Damborsky, J. Machine Learning in Enzyme Engineering. ACS Catal. 2020, 10, 1210–1223. [Google Scholar] [CrossRef]
  62. Bloom, J.; Meyer, M.; Meinhold, P.; Otey, C.; Macmillan, D.; Arnold, F. Evolving strategies for enzyme engineering. Curr. Opin. Struct. Biol. 2005, 15, 447–452. [Google Scholar] [CrossRef] [PubMed]
  63. Chen, R. Enzyme engineering: Rational redesign versus directed evolution. Trends Biotechnol. 2001, 19, 13–14. [Google Scholar] [CrossRef] [PubMed]
  64. Otte, K.B.; Hauer, B. Enzyme engineering in the context of novel pathways and products. Curr. Opin. Biotechnol. 2015, 35, 16–22. [Google Scholar] [CrossRef]
  65. Ding, S. Increasing the activity of immobilized enzymes with nanoparticle conjugation. Curr. Opin. Biotechnol. 2015, 34, 242–250. [Google Scholar] [CrossRef] [PubMed]
  66. Sharma, A.; Gupta, G.; Ahmad, T.; Mansoor, S.; Kaur, B. Enzyme Engineering: Current Trends and Future Perspectives. Food Rev. Int. 2021, 37, 121–154. [Google Scholar] [CrossRef]
  67. Longwell, C.K.; Labanieh, L.; Cochran, J.R. High-throughput screening technologies for enzyme engineering. Curr. Opin. Biotechnol. 2017, 48, 196–202. [Google Scholar] [CrossRef]
  68. Chowdhury, R.; Maranas, C.D. From directed evolution to computational enzyme engineering—A review. AIChE J. 2020, 66, e16847. [Google Scholar] [CrossRef]
  69. Ali, M.; Ishqi, H.M.; Husain, Q. Enzyme engineering: Reshaping the biocatalytic functions. Biotechnol. Bioeng. 2020, 117, 1877–1894. [Google Scholar] [CrossRef]
  70. Smutok, O.; Kavetskyy, T.; Katz, E. Recent trends in enzyme engineering aiming to improve bioelectrocatalysis proceeding with direct electron transfer. Curr. Opin. Electrochem. 2022, 31, 100856. [Google Scholar] [CrossRef]
  71. Plagemann, R.; Jonas, L.; Kragl, U. Ceramic honeycomb as support for covalent immobilization of laccase from Trametes versicolor and transformation of nuclear fast red. Appl. Microbiol. Biotechnol. 2011, 90, 313–320. [Google Scholar] [CrossRef]
  72. Homaei, A.A.; Sariri, R.; Vianello, F.; Stevanato, R. Enzyme immobilization: An update. J. Chem. Biol. 2013, 6, 185–205. [Google Scholar] [CrossRef] [PubMed]
  73. Sheldon, R.A. Enzyme Immobilization: The Quest for Optimum Performance. Adv. Synth. Catal. 2007, 349, 1289–1307. [Google Scholar] [CrossRef]
  74. Gole, A.; Vyas, S.; Sainkar, S.R.; Lachke, A.; Sastry, M. Enhanced Temperature and pH Stability of Fatty Amine−Endoglucanase Composites:  Fabrication, Substrate Protection, and Biological Activity. Langmuir 2001, 17, 5964–5970. [Google Scholar] [CrossRef]
  75. Vera, M.; Rivas, B.L. Immobilization of Trametes versicolor laccase on different PGMA-based polymeric microspheres using response surface methodology: Optimization of conditions. J. Appl. Polym. Sci. 2017, 134, 45249. [Google Scholar] [CrossRef]
  76. Makas, Y.G.; Kalkan, N.A.; Aksoy, S.; Altinok, H.; Hasirci, N. Immobilization of laccase in κ-carrageenan based semi-interpenetrating polymer networks. J. Biotechnol. 2010, 148, 216–220. [Google Scholar] [CrossRef]
  77. Cao, L. Immobilised enzymes: Science or art? Curr. Opin. Chem. Biol. 2005, 9, 217–226. [Google Scholar] [CrossRef]
  78. Bornscheuer, U.T. Immobilizing Enzymes: How to Create More Suitable Biocatalysts. Angew. Chem. Int. Ed. 2003, 42, 3336–3337. [Google Scholar] [CrossRef]
  79. Mateo, C.; Palomo, J.M.; Fernandez-Lorente, G.; Guisan, J.M.; Fernandez-Lafuente, R. Improvement of enzyme activity, stability and selectivity via immobilization techniques. Enzym. Microb. Technol. 2007, 40, 1451–1463. [Google Scholar] [CrossRef]
  80. Kim, M.I.; Ham, H.O.; Oh, S.-D.; Park, H.G.; Chang, H.N.; Choi, S.-H. Immobilization of Mucor javanicus lipase on effectively functionalized silica nanoparticles. J. Mol. Catal. B Enzym. 2006, 39, 62–68. [Google Scholar] [CrossRef]
  81. Garcia-Galan, C.; Berenguer-Murcia, Á.; Fernandez-Lafuente, R.; Rodrigues, R.C. Potential of Different Enzyme Immobilization Strategies to Improve Enzyme Performance. Adv. Synth. Catal. 2011, 353, 2885–2904. [Google Scholar] [CrossRef]
  82. Bai, Y.; Yang, H.; Yang, W.; Li, Y.; Sun, C. Gold nanoparticles-mesoporous silica composite used as an enzyme immobilization matrix for amperometric glucose biosensor construction. Sens. Actuators B Chem. 2007, 124, 179–186. [Google Scholar] [CrossRef]
  83. Liese, A.; Hilterhaus, L. Evaluation of immobilized enzymes for industrial applications. Chem. Soc. Rev. 2013, 42, 6236–6249. [Google Scholar] [CrossRef] [PubMed]
  84. Jiang, Y.; Guo, C.; Xia, H.; Mahmood, I.; Liu, C.; Liu, H. Magnetic nanoparticles supported ionic liquids for lipase immobilization: Enzyme activity in catalyzing esterification. J. Mol. Catal. B Enzym. 2009, 58, 103–109. [Google Scholar] [CrossRef]
  85. Luo, X.-L.; Xu, J.-J.; Du, Y.; Chen, H.-Y. A glucose biosensor based on chitosan–glucose oxidase–gold nanoparticles biocomposite formed by one-step electrodeposition. Anal. Biochem. 2004, 334, 284–289. [Google Scholar] [CrossRef] [PubMed]
  86. Xie, T.; Wang, A.; Huang, L.; Li, H.; Chen, Z.; Wang, Q.; Yin, X. Recent advance in the support and technology used in enzyme immobilization. Afr. J. Biotechnol. 2009, 8, 4724–4733. [Google Scholar]
  87. Roy, J.J.; Abraham, T.E. Preparation and characterization of cross-linked enzyme crystals of laccase. J. Mol. Catal. B Enzym. 2006, 38, 31–36. [Google Scholar] [CrossRef]
  88. Patel, I.; Ludwig, R.; Mueangtoom, K.; Haltrich, D.; Rosenau, T.; Potthast, A. Comparing soluble Trametes pubescens laccase and cross-linked enzyme crystals (CLECs) for enzymatic modification of cellulose. In Proceedings of the 10th EWLP, Stockholm, Sweden, 25–28 August 2008; De Gruyter: Berlin, Germany, 2009; Volume 63, pp. 715–720. [Google Scholar] [CrossRef]
  89. Sheldon, R.A. Characteristic features and biotechnological applications of cross-linked enzyme aggregates (CLEAs). Appl. Microbiol. Biotechnol. 2011, 92, 467–477. [Google Scholar] [CrossRef]
  90. Hartmann, M.; Jung, D. Biocatalysis with enzymes immobilized on mesoporous hosts: The status quo and future trends. J. Mater. Chem. 2010, 20, 844–857. [Google Scholar] [CrossRef]
  91. Talekar, S.; Joshi, A.; Joshi, G.; Kamat, P.; Haripurkar, R.; Kambale, S. Parameters in preparation and characterization of cross linked enzyme aggregates (CLEAs). RSC Adv. 2013, 3, 12485–12511. [Google Scholar] [CrossRef]
  92. Sheldon, R.A. Cross-Linked Enzyme Aggregates as Industrial Biocatalysts. Org. Process Res. Dev. 2011, 15, 213–223. [Google Scholar] [CrossRef]
  93. Cui, J.D.; Jia, S.R. Optimization protocols and improved strategies of cross-linked enzyme aggregates technology: Current development and future challenges. Crit. Rev. Biotechnol. 2015, 35, 15–28. [Google Scholar] [CrossRef] [PubMed]
  94. Xu, M.-Q.; Wang, S.-S.; Li, L.-N.; Gao, J.; Zhang, Y.-W. Combined Cross-Linked Enzyme Aggregates as Biocatalysts. Catalysts 2018, 8, 460. [Google Scholar] [CrossRef]
  95. Perveen, S.; Noreen, S.; Shahid, S.; Mehboob, H.; Aslam, S.; Iqbal, H.M.N.; Bilal, M. Carrier-Free Cross-linked Laccase Crystals for Biocatalytic Degradation of Textile Industrial Effluents. Appl. Biochem. Biotechnol. 2022, 194, 1775–1789. [Google Scholar] [CrossRef] [PubMed]
  96. Velasco-Lozano, S.; López-Gallego, F.; Mateos-Díaz, J.C.; Favela-Torres, E. Cross-linked enzyme aggregates (CLEA) in enzyme improvement—A review. Biocatalysis 2016, 1, 166–177. [Google Scholar] [CrossRef]
  97. Khoshnevisan, K.; Bordbar, A.-K.; Zare, D.; Davoodi, D.; Noruzi, M.; Barkhi, M.; Tabatabaei, M. Immobilization of cellulase enzyme on superparamagnetic nanoparticles and determination of its activity and stability. Chem. Eng. J. 2011, 171, 669–673. [Google Scholar] [CrossRef]
  98. Ren, X.; Meng, X.; Chen, D.; Tang, F.; Jiao, J. Using silver nanoparticle to enhance current response of biosensor. Biosens. Bioelectron. 2005, 21, 433–437. [Google Scholar] [CrossRef]
  99. Vaghari, H. Application of magnetic nanoparticles in smart enzyme immobilization. Biotechnol. Lett. 2016, 38, 223–233. [Google Scholar] [CrossRef]
  100. You, C.-C.; De, M.; Han, G.; Rotello, V.M. Tunable Inhibition and Denaturation of α-Chymotrypsin with Amino Acid-Functionalized Gold Nanoparticles. J. Am. Chem. Soc. 2005, 127, 12873–12881. [Google Scholar] [CrossRef]
  101. Talbert, J.N.; Goddard, J.M. Enzymes on material surfaces. Colloids Surf. B Biointerfaces 2012, 93, 8–19. [Google Scholar] [CrossRef]
  102. Norde, W.; Zoungrana, T. Surface-induced changes in the structure and activity of enzymes physically immobilized at solid/liquid interfaces. Biotechnol. Appl. Biochem. 1998, 28, 133–143. [Google Scholar]
  103. Johnson, A.K.; Zawadzka, A.M.; Deobald, L.A.; Crawford, R.L.; Paszczynski, A.J. Novel method for immobilization of enzymes to magnetic nanoparticles. J. Nanoparticle Res. 2008, 10, 1009–1025. [Google Scholar] [CrossRef]
  104. Miletić, N.; Abetz, V.; Ebert, K.; Loos, K. Immobilization of Candida antarctica lipase B on Polystyrene Nanoparticles. Macromol. Rapid Commun. 2010, 31, 71–74. [Google Scholar] [CrossRef] [PubMed]
  105. Xie, W.; Ma, N. Immobilized Lipase on Fe3O4 Nanoparticles as Biocatalyst for Biodiesel Production. Energy Fuels 2009, 23, 1347–1353. [Google Scholar] [CrossRef]
  106. Can, K.; Ozmen, M.; Ersoz, M. Immobilization of albumin on aminosilane modified superparamagnetic magnetite nanoparticles and its characterization. Colloids Surf. B Biointerfaces 2009, 71, 154–159. [Google Scholar] [CrossRef]
  107. Xu, J.; Sun, J.; Wang, Y.; Sheng, J.; Wang, F.; Sun, M. Application of Iron Magnetic Nanoparticles in Protein Immobilization. Molecules 2014, 19, 11465–11486. [Google Scholar] [CrossRef]
  108. Dyal, A.; Loos, K.; Noto, M.; Chang, S.W.; Spagnoli, C.; Shafi, K.V.P.M.; Ulman, A.; Cowman, M.; Gross, R.A. Activity of Candida rugosa Lipase Immobilized on γ-Fe2O3 Magnetic Nanoparticles. J. Am. Chem. Soc. 2003, 125, 1684–1685. [Google Scholar] [CrossRef]
  109. Arıca, M.Y.; Altıntas, B.; Bayramoğlu, G. Immobilization of laccase onto spacer-arm attached non-porous poly(GMA/EGDMA) beads: Application for textile dye degradation. Bioresour. Technol. 2009, 100, 665–669. [Google Scholar] [CrossRef]
  110. Kim, J.; Jia, H.; Lee, C.; Chung, S.; Kwak, J.H.; Shin, Y.; Dohnalkova, A.; Kim, B.-G.; Wang, P.; Grate, J.W. Single enzyme nanoparticles in nanoporous silica: A hierarchical approach to enzyme stabilization and immobilization. Enzym. Microb. Technol. 2006, 39, 474–480. [Google Scholar] [CrossRef]
  111. Jia, H.; Zhu, G.; Wang, P. Catalytic behaviors of enzymes attached to nanoparticles: The effect of particle mobility. Biotechnol. Bioeng. 2003, 84, 406–414. [Google Scholar] [CrossRef]
  112. Karagoz, B.; Bayramoglu, G.; Altintas, B.; Bicak, N.; Yakup Arica, M. Amine functional monodisperse microbeads via precipitation polymerization of N-vinyl formamide: Immobilized laccase for benzidine based dyes degradation. Bioresour. Technol. 2011, 102, 6783–6790. [Google Scholar] [CrossRef]
  113. Betancor, L.; Fuentes, M.; Dellamora-Ortiz, G.; López-Gallego, F.; Hidalgo, A.; Alonso-Morales, N.; Mateo, C.; Guisán, J.M.; Fernández-Lafuente, R. Dextran aldehyde coating of glucose oxidase immobilized on magnetic nanoparticles prevents its inactivation by gas bubbles. J. Mol. Catal. B Enzym. 2005, 32, 97–101. [Google Scholar] [CrossRef]
  114. Kyomuhimbo, H.D.; Brink, H.G. Applications and immobilization strategies of the copper-centred laccase enzyme; A review. Heliyon 2023, 9, e13156. [Google Scholar] [CrossRef] [PubMed]
  115. Ahmad, A.; Javed, M.R.; Ibrahim, M.; Sajid, A.; Hussain, K.; Kaleem, M.; Fatima, H.M.; Nadeem, H. Methods of Enzyme Immobilization on Various Supports. In Materials Research Foundations, 1st ed.; Materials Research Forum LLC: Millersville PA, USA, 2019; Volume 44, pp. 1–28. [Google Scholar] [CrossRef]
  116. Bolivar, J.M.; Woodley, J.M.; Fernandez-Lafuente, R. Is enzyme immobilization a mature discipline? Some critical considerations to capitalize on the benefits of immobilization. Chem. Soc. Rev. 2022, 51, 6251–6290. [Google Scholar] [CrossRef] [PubMed]
  117. Guisan, J.M.; Fernandez-Lorente, G.; Rocha-Martin, J.; Moreno-Gamero, D. Enzyme immobilization strategies for the design of robust and efficient biocatalysts. Curr. Opin. Green Sustain. Chem. 2022, 35, 100593. [Google Scholar] [CrossRef]
  118. Alvarado-Ramírez, L.; Rostro-Alanis, M.; Rodríguez-Rodríguez, J.; Castillo-Zacarías, C.; Sosa-Hernández, J.E.; Barceló, D.; Iqbal, H.M.N.; Parra-Saldívar, R. Exploring current tendencies in techniques and materials for immobilization of laccases—A review. Int. J. Biol. Macromol. 2021, 181, 683–696. [Google Scholar] [CrossRef]
  119. Kyomuhimbo, H.D.; Feleni, U. Electroconductive Green Metal-polyaniline Nanocomposites: Synthesis and Application in Sensors. Electroanalysis 2023, 35, e202100636. [Google Scholar] [CrossRef]
  120. Sun, Y.; Zhang, W.; Li, J.; Han, R.; Lu, C. Mechanism and Performance Analysis of Nanoparticle-Polymer Fluid for Enhanced Oil Recovery: A Review. Molecules 2023, 28, 4331. [Google Scholar] [CrossRef]
  121. Ahmad, R.; Sardar, M. Enzyme Immobilization: An Overview on Nanoparticles as Immobilization Matrix. Biochem. Anal. Biochem. 2015, 4, 1000178. [Google Scholar] [CrossRef]
  122. Lee, C.-K.; Au-Duong, A.-N. Enzyme Immobilization on Nanoparticles: Recent Applications. In Emerging Areas in Bioengineering; John Wiley & Sons, Ltd: Hoboken, NJ, USA, 2018; pp. 67–80. [Google Scholar] [CrossRef]
  123. Paul, W.; Sharma, C.P. 13—Inorganic nanoparticles for targeted drug delivery. In Biointegration of Medical Implant Materials, 2nd ed.; Sharma, C.P., Ed.; Woodhead Publishing: Sawston, UK, 2020; pp. 333–373. [Google Scholar] [CrossRef]
  124. Kianfar, E. Magnetic Nanoparticles in Targeted Drug Delivery: A Review. J. Supercond. Nov. Magn. 2021, 34, 1709–1735. [Google Scholar] [CrossRef]
  125. Yang, Q.; Dong, Y.; Qiu, Y.; Yang, X.; Cao, H.; Wu, Y. Design of Functional Magnetic Nanocomposites for Bioseparation. Colloids Surf. B Biointerfaces 2020, 191, 111014. [Google Scholar] [CrossRef]
  126. Tang, D.; Cui, Y.; Chen, G. Nanoparticle -based immunoassays in the biomedical field. Analyst 2013, 138, 981–990. [Google Scholar] [CrossRef]
  127. Urusov, A.E.; Petrakova, A.V.; Zherdev, A.V.; Dzantiev, B.B. Application of Magnetic Nanoparticles in Immunoassay. Nanotechnol. Russ. 2017, 12, 471–479. [Google Scholar] [CrossRef]
  128. Sahoo, B.; Sahu, S.K.; Pramanik, P. A novel method for the immobilization of urease on phosphonate grafted iron oxide nanoparticle. J. Mol. Catal. B Enzym. 2011, 69, 95–102. [Google Scholar] [CrossRef]
  129. Rassaei, L.; Marken, F.; Sillanpää, M.; Amiri, M.; Cirtiu, C.M.; Sillanpää, M. Nanoparticles in electrochemical sensors for environmental monitoring. TrAC Trends Anal. Chem. 2011, 30, 1704–1715. [Google Scholar] [CrossRef]
  130. Zhang, S.; Wang, N.; Niu, Y.; Sun, C. Immobilization of glucose oxidase on gold nanoparticles modified Au electrode for the construction of biosensor. Sens. Actuators B Chem. 2005, 109, 367–374. [Google Scholar] [CrossRef]
  131. Luo, X.-L.; Xu, J.-J.; Zhao, W.; Chen, H.-Y. Glucose biosensor based on ENFET doped with SiO2 nanoparticles. Sens. Actuators B Chem. 2004, 97, 249–255. [Google Scholar] [CrossRef]
  132. Huang, Y.; Zhang, W.; Xiao, H.; Li, G. An electrochemical investigation of glucose oxidase at a CdS nanoparticles modified electrode. Biosens. Bioelectron. 2005, 21, 817–821. [Google Scholar] [CrossRef]
  133. Liu, G.; Lin, Y.; Ostatná, V.; Wang, J. Enzyme nanoparticles -based electronic biosensor. Chem. Commun. 2005, 27, 3481–3483. [Google Scholar] [CrossRef]
  134. Ashtari, K.; Khajeh, K.; Fasihi, J.; Ashtari, P.; Ramazani, A.; Vali, H. Silica-encapsulated magnetic nanoparticles: Enzyme immobilization and cytotoxic study. Int. J. Biol. Macromol. 2012, 50, 1063–1069. [Google Scholar] [CrossRef]
  135. Petkova, G.A.; Záruba, K.; Žvátora, P.; Král, V. Gold and silver nanoparticles for biomolecule immobilization and enzymatic catalysis. Nanoscale Res. Lett. 2012, 7, 287. [Google Scholar] [CrossRef]
  136. Gole, A.; Dash, C.; Ramakrishnan, V.; Sainkar, S.R.; Mandale, A.B.; Rao, M.; Sastry, M. Pepsin−Gold Colloid Conjugates:  Preparation, Characterization, and Enzymatic Activity. Langmuir 2001, 17, 1674–1679. [Google Scholar] [CrossRef]
  137. Husain, Q.; Ansari, S.A.; Alam, F.; Azam, A. Immobilization of Aspergillus oryzae β galactosidase on zinc oxide nanoparticles via simple adsorption mechanism. Int. J. Biol. Macromol. 2011, 49, 37–43. [Google Scholar] [CrossRef] [PubMed]
  138. Wang, T.-H.; Lee, W.-C. Immobilization of proteins on magnetic nanoparticles. Biotechnol. Bioprocess. Eng. 2003, 8, 263–267. [Google Scholar] [CrossRef]
  139. Georgelin, T.; Maurice, V.; Malezieux, B.; Siaugue, J.-M.; Cabuil, V. Design of multifunctionalized γ-Fe2O3@SiO2 core–shell nanoparticles for enzymes immobilization. J. Nanoparticle Res. 2010, 12, 675–680. [Google Scholar] [CrossRef]
  140. Palocci, C.; Chronopoulou, L.; Venditti, I.; Cernia, E.; Diociaiuti, M.; Fratoddi, I.; Russo, M.V. Lipolytic Enzymes with Improved Activity and Selectivity upon Adsorption on Polymeric Nanoparticles. Biomacromolecules 2007, 8, 3047–3053. [Google Scholar] [CrossRef] [PubMed]
  141. Wang, P. Multi-scale Features in Recent Development of Enzymic Biocatalyst Systems. Appl. Biochem. Biotechnol. 2009, 152, 343–352. [Google Scholar] [CrossRef]
  142. Crespilho, F.N.; Iost, R.M.; Travain, S.A.; Oliveira, O.N.; Zucolotto, V. Enzyme immobilization on Ag nanoparticles/polyaniline nanocomposites. Biosens. Bioelectron. 2009, 24, 3073–3077. [Google Scholar] [CrossRef]
  143. Liu, S.; Dai, Z.; Chen, H.; Ju, H. Immobilization of hemoglobin on zirconium dioxide nanoparticles for preparation of a novel hydrogen peroxide biosensor. Biosens. Bioelectron. 2004, 19, 963–969. [Google Scholar] [CrossRef]
  144. Mandal, S.; Phadtare, S.; Sastry, M. Interfacing biology with nanoparticles. Curr. Appl. Phys. 2005, 5, 118–127. [Google Scholar] [CrossRef]
  145. Shukoor, M.I.; Natalio, F.; Therese, H.A.; Tahir, M.N.; Ksenofontov, V.; Panthöfer, M.; Eberhardt, M.; Theato, P.; Schröder, H.C.; Müller, W.E.G.; et al. Fabrication of a Silica Coating on Magnetic γ-Fe2O3 Nanoparticles by an Immobilized Enzyme. Chem. Mater. 2008, 20, 3567–3573. [Google Scholar] [CrossRef]
  146. Ponvel, K.M.; Lee, D.-G.; Woo, E.-J.; Ahn, I.-S.; Lee, C.-H. Immobilization of lipase on surface modified magnetic nanoparticles using alkyl benzenesulfonate. Korean J. Chem. Eng. 2009, 26, 127–130. [Google Scholar] [CrossRef]
  147. Arsalan, A.; Alam, M.F.; Farheen Zofair, S.F.; Ahmad, S.; Younus, H. Immobilization of β-galactosidase on tannic acid stabilized silver nanoparticles: A safer way towards its industrial application. Spectrochim. Acta Part A Mol. Biomol. Spectrosc. 2020, 226, 117637. [Google Scholar] [CrossRef]
  148. Solanki, K.; Gupta, M.N. Simultaneous purification and immobilization of Candida rugosa lipase on superparamagnetic Fe3O4 nanoparticles for catalyzing transesterification reactions. New J. Chem. 2011, 35, 2551–2556. [Google Scholar] [CrossRef]
  149. Yu, C.-C.; Kuo, Y.-Y.; Liang, C.-F.; Chien, W.-T.; Wu, H.-T.; Chang, T.-C.; Jan, F.-D.; Lin, C.-C. Site-Specific Immobilization of Enzymes on Magnetic Nanoparticles and Their Use in Organic Synthesis. Bioconjugate Chem. 2012, 23, 714–724. [Google Scholar] [CrossRef]
  150. Wu, W.; He, Q.; Jiang, C. Magnetic Iron Oxide Nanoparticles: Synthesis and Surface Functionalization Strategies. Nanoscale Res. Lett. 2008, 3, 397. [Google Scholar] [CrossRef]
  151. Gupta, A.K.; Gupta, M. Synthesis and surface engineering of iron oxide nanoparticles for biomedical applications. Biomaterials 2005, 26, 3995–4021. [Google Scholar] [CrossRef]
  152. Chen, Y.Z.; Yang, C.T.; Ching, C.B.; Xu, R. Immobilization of Lipases on Hydrophobilized Zirconia Nanoparticles: Highly Enantioselective and Reusable Biocatalysts. Langmuir 2008, 24, 8877–8884. [Google Scholar] [CrossRef]
  153. Pan, C.; Hu, B.; Li, W.; Sun, Y.; Ye, H.; Zeng, X. Novel and efficient method for immobilization and stabilization of β-d-galactosidase by covalent attachment onto magnetic Fe3O4–chitosan nanoparticles. J. Mol. Catal. B Enzym. 2009, 61, 208–215. [Google Scholar] [CrossRef]
  154. Ren, Y.; Rivera, J.G.; He, L.; Kulkarni, H.; Lee, D.-K.; Messersmith, P.B. Facile, high efficiency immobilization of lipase enzyme on magnetic iron oxide nanoparticles via a biomimetic coating. BMC Biotechnol. 2011, 11, 63. [Google Scholar] [CrossRef]
  155. Phadtare, S.; Kumar, A.; Vinod, V.P.; Dash, C.; Palaskar, D.V.; Rao, M.; Shukla, P.G.; Sivaram, S.; Sastry, M. Direct Assembly of Gold Nanoparticle “Shells” on Polyurethane Microsphere “Cores” and Their Application as Enzyme Immobilization Templates. Chem. Mater. 2003, 15, 1944–1949. [Google Scholar] [CrossRef]
  156. Mazeiko, V.; Kausaite-Minkstimiene, A.; Ramanaviciene, A.; Balevicius, Z.; Ramanavicius, A. Gold nanoparticle and conducting polymer-polyaniline-based nanocomposites for glucose biosensor design. Sens. Actuators B Chem. 2013, 189, 187–193. [Google Scholar] [CrossRef]
  157. Kaushik, A.; Khan, R.; Solanki, P.R.; Pandey, P.; Alam, J.; Ahmad, S.; Malhotra, B.D. Iron oxide nanoparticles–chitosan composite based glucose biosensor. Biosens. Bioelectron. 2008, 24, 676–683. [Google Scholar] [CrossRef] [PubMed]
  158. Nunes, Y.L.; de Menezes, F.L.; de Sousa, I.G.; Cavalcante, A.L.G.; Cavalcante, F.T.T.; da Silva Moreira, K.; de Oliveira, A.L.B.; Mota, G.F.; da Silva Souza, J.E.; de Aguiar Falcão, I.R.; et al. Chemical and physical Chitosan modification for designing enzymatic industrial biocatalysts: How to choose the best strategy? Int. J. Biol. Macromol. 2021, 181, 1124–1170. [Google Scholar] [CrossRef]
  159. Ma, S.; Mu, J.; Jiang, L. Chitosan-Based Glucose Oxidase Electrodes Enhanced by Silver Nanoparticles. J. Dispers. Sci. Technol. 2008, 29, 682–686. [Google Scholar] [CrossRef]
  160. Rafiee, F.; Rezaee, M. Different strategies for the lipase immobilization on the chitosan based supports and their applications. Int. J. Biol. Macromol. 2021, 179, 170–195. [Google Scholar] [CrossRef] [PubMed]
  161. Aranaz, I.; Alcántara, A.R.; Civera, M.C.; Arias, C.; Elorza, B.; Heras Caballero, A.; Acosta, N. Chitosan: An Overview of Its Properties and Applications. Polymers 2021, 13, 3256. [Google Scholar] [CrossRef]
  162. Ribeiro, E.S.; de Farias, B.S.; Sant’Anna Cadaval, T.R., Jr.; de Almeida Pinto, L.A.; Diaz, P.S. Chitosan–based nanofibers for enzyme immobilization. Int. J. Biol. Macromol. 2021, 183, 1959–1970. [Google Scholar] [CrossRef]
  163. Divya, K.; Jisha, M.S. Chitosan nanoparticles preparation and applications. Environ. Chem. Lett. 2018, 16, 101–112. [Google Scholar] [CrossRef]
  164. Verma, M.L.; Kumar, S.; Das, A.; Randhawa, J.S.; Chamundeeswari, M. Chitin and chitosan-based support materials for enzyme immobilization and biotechnological applications. Environ. Chem. Lett. 2020, 18, 315–323. [Google Scholar] [CrossRef]
  165. Du, Y.; Luo, X.-L.; Xu, J.-J.; Chen, H.-Y. A simple method to fabricate a chitosan-gold nanoparticles film and its application in glucose biosensor. Bioelectrochemistry 2007, 70, 342–347. [Google Scholar] [CrossRef]
  166. Zhu, Y.; Zhu, H.; Yang, X.; Xu, L.; Li, C. Sensitive Biosensors Based on (Dendrimer Encapsulated Pt Nanoparticles)/Enzyme Multilayers. Electroanalysis 2007, 19, 698–703. [Google Scholar] [CrossRef]
  167. Krajewska, B. Application of chitin- and chitosan-based materials for enzyme immobilizations: A review. Enzym. Microb. Technol. 2004, 35, 126–139. [Google Scholar] [CrossRef]
  168. Chouhan, D.; Mandal, P. Applications of chitosan and chitosan based metallic nanoparticles in agrosciences—A review. Int. J. Biol. Macromol. 2021, 166, 1554–1569. [Google Scholar] [CrossRef] [PubMed]
  169. Jiang, Y.; Wu, J. Recent development in chitosan nanocomposites for surface-based biosensor applications. Electrophoresis 2019, 40, 2084–2097. [Google Scholar] [CrossRef]
  170. Negm, N.A.; Abubshait, H.A.; Abubshait, S.A.; Abou Kana, M.T.H.; Mohamed, E.A.; Betiha, M.M. Performance of chitosan polymer as platform during sensors fabrication and sensing applications. Int. J. Biol. Macromol. 2020, 165, 402–435. [Google Scholar] [CrossRef]
  171. Lee, K.Y.; Mooney, D.J. Alginate: Properties and biomedical applications. Prog. Polym. Sci. 2012, 37, 106–126. [Google Scholar] [CrossRef]
  172. Maity, C.; Das, N. Alginate-Based Smart Materials and Their Application: Recent Advances and Perspectives. Top. Curr. Chem. (Z) 2021, 380, 3. [Google Scholar] [CrossRef]
  173. Bilal, M.; Iqbal, H.M.N. Lignin peroxidase immobilization on Ca-alginate beads and its dye degradation performance in a packed bed reactor system. Biocatal. Agric. Biotechnol. 2019, 20, 101205. [Google Scholar] [CrossRef]
  174. Gheorghita Puscaselu, R.; Lobiuc, A.; Dimian, M.; Covasa, M. Alginate: From Food Industry to Biomedical Applications and Management of Metabolic Disorders. Polymers 2020, 12, 2417. [Google Scholar] [CrossRef]
  175. Fernando, I.P.S.; Lee, W.; Han, E.J.; Ahn, G. Alginate-based nanomaterials: Fabrication techniques, properties, and applications. Chem. Eng. J. 2020, 391, 123823. [Google Scholar] [CrossRef]
  176. Taemeh, M.A.; Shiravandi, A.; Korayem, M.A.; Daemi, H. Fabrication challenges and trends in biomedical applications of alginate electrospun nanofibers. Carbohydr. Polym. 2020, 228, 115419. [Google Scholar] [CrossRef] [PubMed]
  177. Simó, G.; Fernández-Fernández, E.; Vila-Crespo, J.; Ruipérez, V.; Rodríguez-Nogales, J.M. Research progress in coating techniques of alginate gel polymer for cell encapsulation. Carbohydr. Polym. 2017, 170, 1–14. [Google Scholar] [CrossRef] [PubMed]
  178. Kumar, S.; Haq, I.; Prakash, J.; Raj, A. Improved enzyme properties upon glutaraldehyde cross-linking of alginate entrapped xylanase from Bacillus licheniformis. Int. J. Biol. Macromol. 2017, 98, 24–33. [Google Scholar] [CrossRef] [PubMed]
  179. Leong, J.-Y.; Lam, W.-H.; Ho, K.-W.; Voo, W.-P.; Lee, M.F.-X.; Lim, H.-P.; Lim, S.-L.; Tey, B.-T.; Poncelet, D.; Chan, E.-S. Advances in fabricating spherical alginate hydrogels with controlled particle designs by ionotropic gelation as encapsulation systems. Particuology 2016, 24, 44–60. [Google Scholar] [CrossRef]
  180. Martín, M.C.; López, O.V.; Ciolino, A.E.; Morata, V.I.; Villar, M.A.; Ninago, M.D. Immobilization of enological pectinase in calcium alginate hydrogels: A potential biocatalyst for winemaking. Biocatal. Agric. Biotechnol. 2019, 18, 101091. [Google Scholar] [CrossRef]
  181. Gür, S.D.; İdil, N.; Aksöz, N. Optimization of Enzyme Co-Immobilization with Sodium Alginate and Glutaraldehyde-Activated Chitosan Beads. Appl. Biochem. Biotechnol. 2018, 184, 538–552. [Google Scholar] [CrossRef]
  182. Damayanti, A.; Kumoro, A.C.; Bahlawan, Z.A.S. Review Calcium Alginate Beads as Immobilizing Matrix of Functional Cells: Extrusion Dripping Method, Characteristics, and Application. IOP Conf. Ser. Mater. Sci. Eng. 2021, 1053, 012017. [Google Scholar] [CrossRef]
  183. Chaudhari, S.; Kar, J.; Singhal, R. Immobilization of Proteins in Alginate: Functional Properties and Applications. Curr. Org. Chem. 2015, 19, 1732–1754. [Google Scholar] [CrossRef]
  184. Bibi, A.; Rehman, S.; Yaseen, A. Alginate-nanoparticles composites: Kinds, reactions and applications. Mater. Res. Express 2019, 6, 092001. [Google Scholar] [CrossRef]
  185. Wang, B.; Wan, Y.; Zheng, Y.; Lee, X.; Liu, T.; Yu, Z.; Huang, J.; Ok, Y.S.; Chen, J.; Gao, B. Alginate-based composites for environmental applications: A critical review. Crit. Rev. Environ. Sci. Technol. 2019, 49, 318–356. [Google Scholar] [CrossRef]
  186. Li, J.; Xiao, L.-T.; Liu, X.-M.; Zeng, G.-M.; Huang, G.-H.; Shen, G.-L.; Yu, R.-Q. Amperometric biosensor with HRP immobilized on a sandwiched nano-Au /polymerized m-phenylenediamine film and ferrocene mediator. Anal. Bioanal. Chem. 2003, 376, 902–907. [Google Scholar] [CrossRef] [PubMed]
  187. Tavahodi, M.; Ortiz, R.; Schulz, C.; Ekhtiari, A.; Ludwig, R.; Haghighi, B.; Gorton, L. Direct Electron Transfer of Cellobiose Dehydrogenase on Positively Charged Polyethyleneimine Gold Nanoparticles. ChemPlusChem 2017, 82, 546–552. [Google Scholar] [CrossRef] [PubMed]
  188. Chen, X.; Chen, Z.; Zhu, J.; Xu, C.; Yan, W.; Yao, C. A novel H2O2 amperometric biosensor based on gold nanoparticles/self-doped polyaniline nanofibers. Bioelectrochemistry 2011, 82, 87–94. [Google Scholar] [CrossRef] [PubMed]
  189. Xian, Y.; Hu, Y.; Liu, F.; Xian, Y.; Wang, H.; Jin, L. Glucose biosensor based on Au nanoparticles–conductive polyaniline nanocomposite. Biosens. Bioelectron. 2006, 21, 1996–2000. [Google Scholar] [CrossRef]
  190. Cheng, H.; Hu, M.; Zhai, Q.; Li, S.; Jiang, Y. Polydopamine tethered CPO/HRP-TiO2 nano-composites with high bio-catalytic activity, stability and reusability: Enzyme-photo bifunctional synergistic catalysis in water treatment. Chem. Eng. J. 2018, 347, 703–710. [Google Scholar] [CrossRef]
  191. Chen, C.; Sun, W.; Lv, H.; Li, H.; Wang, Y.; Wang, P. Spacer arm-facilitated tethering of laccase on magnetic polydopamine nanoparticles for efficient biocatalytic water treatment. Chem. Eng. J. 2018, 350, 949–959. [Google Scholar] [CrossRef]
  192. German, N.; Ramanavicius, A.; Ramanaviciene, A. Amperometric Glucose Biosensor Based on Electrochemically Deposited Gold Nanoparticles Covered by Polypyrrole. Electroanalysis 2017, 29, 1267–1277. [Google Scholar] [CrossRef]
  193. Njagi, J.; Andreescu, S. Stable enzyme biosensors based on chemically synthesized Au–polypyrrole nanocomposites. Biosens. Bioelectron. 2007, 23, 168–175. [Google Scholar] [CrossRef]
  194. Xu, G.; Adeloju, S.B.; Wu, Y.; Zhang, X. Modification of polypyrrole nanowires array with platinum nanoparticles and glucose oxidase for fabrication of a novel glucose biosensor. Anal. Chim. Acta 2012, 755, 100–107. [Google Scholar] [CrossRef]
  195. Yuan, R.; Liu, Y.; Li, Q.-F.; Chai, Y.-Q.; Mo, C.-L.; Zhong, X.; Tang, D.-P.; Dai, J.-Y. Electrochemical characteristics of a platinum electrode modified with a matrix of polyvinyl butyral and colloidal Ag containing immobilized horseradish peroxidase. Anal. Bioanal. Chem. 2005, 381, 762–768. [Google Scholar] [CrossRef]
  196. Xiang, C.; Zou, Y.; Sun, L.-X.; Xu, F. Direct electrochemistry and enhanced electrocatalysis of horseradish peroxidase based on flowerlike ZnO–gold nanoparticle–Nafion nanocomposite. Sens. Actuators B Chem. 2009, 136, 158–162. [Google Scholar] [CrossRef]
  197. Zhao, S.; Zhang, K.; Bai, Y.; Yang, W.; Sun, C. Glucose oxidase/colloidal gold nanoparticles immobilized in Nafion film on glassy carbon electrode: Direct electron transfer and electrocatalysis. Bioelectrochemistry 2006, 69, 158–163. [Google Scholar] [CrossRef]
  198. Nor, N.M.; Lockman, Z.; Razak, K.A. Study of ITO Glass Electrode Modified with Iron Oxide Nanoparticles and Nafion for Glucose Biosensor Application. Procedia Chem. 2016, 19, 50–56. [Google Scholar] [CrossRef]
  199. Ul Haque, S.; Inamuddin; Nasar, A.; Asiri, A.M. Fabrication and characterization of electrochemically prepared bioanode (polyaniline/ferritin/glucose oxidase) for biofuel cell application. Chem. Phys. Lett. 2018, 692, 277–284. [Google Scholar] [CrossRef]
  200. Dutta, R.R.; Puzari, P. Amperometric biosensing of organophosphate and organocarbamate pesticides utilizing polypyrrole entrapped acetylcholinesterase electrode. Biosens. Bioelectron. 2014, 52, 166–172. [Google Scholar] [CrossRef]
  201. Sintayehu, Y.D.; Ananda Murthy, H.C. Polypyrrole based biofunctional composite layer for bioelectrocatalytic device system. Adv. Mater. Lett. 2019, 10, 524–532. [Google Scholar] [CrossRef]
  202. Lai, J.; Yi, Y.; Zhu, P.; Shen, J.; Wu, K.; Zhang, L.; Liu, J. Polyaniline-based glucose biosensor: A review. J. Electroanal. Chem. 2016, 782, 138–153. [Google Scholar] [CrossRef]
  203. Osuna, V.; Vega-Rios, A.; Zaragoza-Contreras, E.A.; Estrada-Moreno, I.A.; Dominguez, R.B. Progress of Polyaniline Glucose Sensors for Diabetes Mellitus Management Utilizing Enzymatic and Non-Enzymatic Detection. Biosensors 2022, 12, 137. [Google Scholar] [CrossRef]
  204. Kausaite-Minkstimiene, A.; Mazeiko, V.; Ramanaviciene, A.; Ramanavicius, A. Evaluation of amperometric glucose biosensors based on glucose oxidase encapsulated within enzymatically synthesized polyaniline and polypyrrole. Sens. Actuators B Chem. 2011, 158, 278–285. [Google Scholar] [CrossRef]
  205. Nemzer, L.R.; Epstein, A.J. A polyaniline-based optical biosensing platform using an entrapped oxidoreductase enzyme. Sens. Actuators B Chem. 2010, 150, 376–383. [Google Scholar] [CrossRef]
  206. Nemzer, L.R.; Schwartz, A.; Epstein, A.J. Enzyme Entrapment in Reprecipitated Polyaniline Nano- and Microparticles. Macromolecules 2010, 43, 4324–4330. [Google Scholar] [CrossRef]
  207. Zheng, H.; Liu, M.; Yan, Z.; Chen, J. Highly selective and stable glucose biosensor based on incorporation of platinum nanoparticles into polyaniline-montmorillonite hybrid composites. Microchem. J. 2020, 152, 104266. [Google Scholar] [CrossRef]
  208. Wahab, R.A.; Elias, N.; Abdullah, F.; Ghoshal, S.K. On the taught new tricks of enzymes immobilization: An all-inclusive overview. React. Funct. Polym. 2020, 152, 104613. [Google Scholar] [CrossRef]
  209. John, A.; Benny, L.; Cherian, A.R.; Narahari, S.Y.; Varghese, A.; Hegde, G. Electrochemical sensors using conducting polymer/noble metal nanoparticle nanocomposites for the detection of various analytes: A review. J. Nanostruct. Chem. 2021, 11, 1–31. [Google Scholar] [CrossRef]
  210. Shan, D.; Zhu, M.; Xue, H.; Cosnier, S. Development of amperometric biosensor for glucose based on a novel attractive enzyme immobilization matrix: Calcium carbonate nanoparticles. Biosens. Bioelectron. 2007, 22, 1612–1617. [Google Scholar] [CrossRef]
  211. Sundar, S.; Kundu, J.; Kundu, S.C. Biopolymeric nanoparticles. Sci. Technol. Adv. Mater. 2010, 11, 014104. [Google Scholar] [CrossRef] [PubMed]
  212. Arsalan, A.; Younus, H. Enzymes and nanoparticles: Modulation of enzymatic activity via nanoparticles. Int. J. Biol. Macromol. 2018, 118, 1833–1847. [Google Scholar] [CrossRef]
  213. Chen, M.; Zeng, G.; Xu, P.; Lai, C.; Tang, L. How Do Enzymes ‘Meet’ Nanoparticles and Nanomaterials? Trends Biochem. Sci. 2017, 42, 914–930. [Google Scholar] [CrossRef]
  214. Qhobosheane, M.; Santra, S.; Zhang, P.; Tan, W. Biochemically functionalized silica nanoparticles. Analyst 2001, 126, 1274–1278. [Google Scholar] [CrossRef]
  215. Cruz, J.C.; Würges, K.; Kramer, M.; Pfromm, P.H.; Rezac, M.E.; Czermak, P. Immobilization of Enzymes on Fumed Silica Nanoparticles for Applications in Nonaqueous Media. In Nanoscale Biocatalysis: Methods and Protocols; Wang, P., Ed.; Humana Press: Totowa, NJ, USA, 2011; pp. 147–160. [Google Scholar] [CrossRef]
  216. Xiao, Q.-G.; Tao, X.; Zou, H.-K.; Chen, J.-F. Comparative study of solid silica nanoparticles and hollow silica nanoparticles for the immobilization of lysozyme. Chem. Eng. J. 2008, 137, 38–44. [Google Scholar] [CrossRef]
  217. Wang, F.; Guo, C.; Yang, L.; Liu, C.-Z. Magnetic mesoporous silica nanoparticles: Fabrication and their laccase immobilization performance. Bioresour. Technol. 2010, 101, 8931–8935. [Google Scholar] [CrossRef] [PubMed]
  218. Deepak, V.; Pandian, S.R.b.K.; Kalishwaralal, K.; Gurunathan, S. Purification, immobilization, and characterization of nattokinase on PHB nanoparticles. Bioresour. Technol. 2009, 100, 6644–6646. [Google Scholar] [CrossRef] [PubMed]
  219. Konno, T.; Watanabe, J.; Ishihara, K. Conjugation of Enzymes on Polymer Nanoparticles Covered with Phosphorylcholine Groups. Biomacromolecules 2004, 5, 342–347. [Google Scholar] [CrossRef] [PubMed]
  220. Tang, Z.-X.; Qian, J.-Q.; Shi, L.-E. Preparation of chitosan nanoparticles as carrier for immobilized enzyme. Appl. Biochem. Biotechnol. 2007, 136, 77–96. [Google Scholar] [CrossRef] [PubMed]
  221. Muthurasu, A.; Ganesh, V. Horseradish Peroxidase Enzyme Immobilized Graphene Quantum Dots as Electrochemical Biosensors. Appl. Biochem. Biotechnol. 2014, 174, 945–959. [Google Scholar] [CrossRef] [PubMed]
  222. Yuan, J.; Guo, W.; Wang, E. Quantum dots–bienzyme hybrid system for the sensitive determination of glucose. Biosens. Bioelectron. 2008, 23, 1567–1571. [Google Scholar] [CrossRef]
  223. Agrawal, D.C.; Yadav, A.; Kesarwani, R.; Srivastava, O.N.; Kayastha, A.M. Immobilization of fenugreek β-amylase onto functionalized graphene quantum dots (GQDs) using Box-Behnken design: Its biochemical, thermodynamic and kinetic studies. Int. J. Biol. Macromol. 2020, 144, 170–182. [Google Scholar] [CrossRef]
  224. Wu, P.; He, Y.; Wang, H.-F.; Yan, X.-P. Conjugation of Glucose Oxidase onto Mn-Doped ZnS Quantum Dots for Phosphorescent Sensing of Glucose in Biological Fluids. Anal. Chem. 2010, 82, 1427–1433. [Google Scholar] [CrossRef]
  225. Bilal, M.; Anh Nguyen, T.; Iqbal, H.M.N. Multifunctional carbon nanotubes and their derived nano-constructs for enzyme immobilization—A paradigm shift in biocatalyst design. Coord. Chem. Rev. 2020, 422, 213475. [Google Scholar] [CrossRef]
  226. Ranjan, B.; Pillai, S.; Permaul, K.; Singh, S. Simultaneous removal of heavy metals and cyanate in a wastewater sample using immobilized cyanate hydratase on magnetic-multiwall carbon nanotubes. J. Hazard. Mater. 2019, 363, 73–80. [Google Scholar] [CrossRef]
  227. Ahmad, R.; Khare, S.K. Immobilization of Aspergillus niger cellulase on multiwall carbon nanotubes for cellulose hydrolysis. Bioresour. Technol. 2018, 252, 72–75. [Google Scholar] [CrossRef] [PubMed]
  228. Alagöz, D.; Toprak, A.; Yildirim, D.; Tükel, S.S.; Fernandez-Lafuente, R. Modified silicates and carbon nanotubes for immobilization of lipase from Rhizomucor miehei: Effect of support and immobilization technique on the catalytic performance of the immobilized biocatalysts. Enzym. Microb. Technol. 2021, 144, 109739. [Google Scholar] [CrossRef] [PubMed]
  229. Li, X.; Cao, X.; Xiong, J.; Ge, J. Enzyme–Metal Hybrid Catalysts for Chemoenzymatic Reactions. Small 2020, 16, 1902751. [Google Scholar] [CrossRef] [PubMed]
  230. Solanki, P.R.; Kaushik, A.; Agrawal, V.V.; Malhotra, B.D. Nanostructured metal oxide-based biosensors. NPG Asia Mater. 2011, 3, 17–24. [Google Scholar] [CrossRef]
  231. Upadhyay, S.; Rao, G.R.; Sharma, M.K.; Bhattacharya, B.K.; Rao, V.K.; Vijayaraghavan, R. Immobilization of acetylcholineesterase–choline oxidase on a gold–platinum bimetallic nanoparticles modified glassy carbon electrode for the sensitive detection of organophosphate pesticides, carbamates and nerve agents. Biosens. Bioelectron. 2009, 25, 832–838. [Google Scholar] [CrossRef]
  232. Ansari, A.A.; Sumana, G.; Pandey, M.K.; Malhotra, B.D. Sol-gel-derived titanium oxide–cerium oxide biocompatible nanocomposite film for urea sensor. J. Mater. Res. 2009, 24, 1667–1673. [Google Scholar] [CrossRef]
  233. Sadjadi, M.S.; Farhadyar, N.; Zare, K. Synthesis of bi-metallic Au–Ag nanoparticles loaded on functionalized MCM-41 for immobilization of alkaline protease and study of its biocatalytic activity. Superlattices Microstruct. 2009, 46, 563–571. [Google Scholar] [CrossRef]
  234. Turkmen, E.; Bas, S.Z.; Gulce, H.; Yildiz, S. Glucose biosensor based on immobilization of glucose oxidase in electropolymerized poly(o-phenylenediamine) film on platinum nanoparticles-polyvinylferrocenium modified electrode. Electrochim. Acta 2014, 123, 93–102. [Google Scholar] [CrossRef]
  235. Taurino, I.; Sanzò, G.; Antiochia, R.; Tortolini, C.; Mazzei, F.; Favero, G.; De Micheli, G.; Carrara, S. Recent advances in Third Generation Biosensors based on Au and Pt Nanostructured Electrodes. TrAC Trends Anal. Chem. 2016, 79, 151–159. [Google Scholar] [CrossRef]
  236. Yu, H.; Yu, J.; Li, L.; Zhang, Y.; Xin, S.; Ni, X.; Sun, Y.; Song, K. Recent Progress of the Practical Applications of the Platinum Nanoparticle-Based Electrochemistry Biosensors. Front. Chem. 2021, 9, 677876. [Google Scholar] [CrossRef]
  237. Liu, L.; Chen, Y.; Lv, H.; Wang, G.; Hu, X.; Wang, C. Construction of a non-enzymatic glucose sensor based on copper nanoparticles/poly(o-phenylenediamine) nanocomposites. J. Solid State Electrochem. 2015, 19, 731–738. [Google Scholar] [CrossRef]
  238. Solairaj, D.; Rameshthangam, P.; Muthukumaran, P.; Wilson, J. Studies on electrochemical glucose sensing, antimicrobial activity and cytotoxicity of fabricated copper nanoparticle immobilized chitin nanostructure. Int. J. Biol. Macromol. 2017, 101, 668–679. [Google Scholar] [CrossRef] [PubMed]
  239. Lim, S.H.; Wei, J.; Lin, J.; Li, Q.; KuaYou, J. A glucose biosensor based on electrodeposition of palladium nanoparticles and glucose oxidase onto Nafion-solubilized carbon nanotube electrode. Biosens. Bioelectron. 2005, 20, 2341–2346. [Google Scholar] [CrossRef] [PubMed]
  240. Li, Z.; Wang, X.; Wen, G.; Shuang, S.; Dong, C.; Paau, M.C.; Choi, M.M.F. Application of hydrophobic palladium nanoparticles for the development of electrochemical glucose biosensor. Biosens. Bioelectron. 2011, 26, 4619–4623. [Google Scholar] [CrossRef]
  241. Ganesana, M.; Istarnboulie, G.; Marty, J.-L.; Noguer, T.; Andreescu, S. Site-specific immobilization of a (His)6-tagged acetylcholinesterase on nickel nanoparticles for highly sensitive toxicity biosensors. Biosens. Bioelectron. 2011, 30, 43–48. [Google Scholar] [CrossRef]
  242. Bodelón, G.; Mourdikoudis, S.; Yate, L.; Pastoriza-Santos, I.; Pérez-Juste, J.; Liz-Marzán, L.M. Nickel Nanoparticle-Doped Paper as a Bioactive Scaffold for Targeted and Robust Immobilization of Functional Proteins. ACS Nano 2014, 8, 6221–6231. [Google Scholar] [CrossRef]
  243. Salimi, A.; Sharifi, E.; Noorbakhsh, A.; Soltanian, S. Immobilization of glucose oxidase on electrodeposited nickel oxide nanoparticles: Direct electron transfer and electrocatalytic activity. Biosens. Bioelectron. 2007, 22, 3146–3153. [Google Scholar] [CrossRef]
  244. Bayandori Moghaddam, A.; Ganjali, M.R.; Saboury, A.A.; Moosavi-Movahedi, A.A.; Norouzi, P. Electrodeposition of nickel oxide nanoparticles on glassy carbon surfaces: Application to the direct electron transfer of tyrosinase. J. Appl. Electrochem. 2008, 38, 1233–1239. [Google Scholar] [CrossRef]
  245. Sharifi, E.; Salimi, A.; Shams, E. Electrocatalytic activity of nickel oxide nanoparticles as mediatorless system for NADH and ethanol sensing at physiological pH solution. Biosens. Bioelectron. 2013, 45, 260–266. [Google Scholar] [CrossRef]
  246. Narang, J.; Chauhan, N.; Pundir, C.S. Construction of triglyceride biosensor based on nickel oxide–chitosan/zinc oxide/zinc hexacyanoferrate film. Int. J. Biol. Macromol. 2013, 60, 45–51. [Google Scholar] [CrossRef]
  247. Feng, J.-J.; Zhao, G.; Xu, J.-J.; Chen, H.-Y. Direct electrochemistry and electrocatalysis of heme proteins immobilized on gold nanoparticles stabilized by chitosan. Anal. Biochem. 2005, 342, 280–286. [Google Scholar] [CrossRef]
  248. Wang, H.; Wang, X.; Zhang, X.; Qin, X.; Zhao, Z.; Miao, Z.; Huang, N.; Chen, Q. A novel glucose biosensor based on the immobilization of glucose oxidase onto gold nanoparticles-modified Pb nanowires. Biosens. Bioelectron. 2009, 25, 142–146. [Google Scholar] [CrossRef]
  249. Pingarrón, J.M.; Yáñez-Sedeño, P.; González-Cortés, A. Gold nanoparticle-based electrochemical biosensors. Electrochim. Acta 2008, 53, 5848–5866. [Google Scholar] [CrossRef]
  250. Luo, X.-L.; Xu, J.-J.; Zhang, Q.; Yang, G.-J.; Chen, H.-Y. Electrochemically deposited chitosan hydrogel for horseradish peroxidase immobilization through gold nanoparticles self-assembly. Biosens. Bioelectron. 2005, 21, 190–196. [Google Scholar] [CrossRef]
  251. Mahmoud, K.A.; Male, K.B.; Hrapovic, S.; Luong, J.H.T. Cellulose Nanocrystal/Gold Nanoparticle Composite as a Matrix for Enzyme Immobilization. ACS Appl. Mater. Interfaces 2009, 1, 1383–1386. [Google Scholar] [CrossRef]
  252. Gole, A.; Vyas, S.; Phadtare, S.; Lachke, A.; Sastry, M. Studies on the formation of bioconjugates of Endoglucanase with colloidal gold. Colloids Surf. B Biointerfaces 2002, 25, 129–138. [Google Scholar] [CrossRef]
  253. Phadtare, S.; Vinod, V.p.; Mukhopadhyay, K.; Kumar, A.; Rao, M.; Chaudhari, R.V.; Sastry, M. Immobilization and biocatalytic activity of fungal protease on gold nanoparticle-loaded zeolite microspheres. Biotechnol. Bioeng. 2004, 85, 629–637. [Google Scholar] [CrossRef]
  254. Kumar-Krishnan, S.; Hernandez-Rangel, A.; Pal, U.; Ceballos-Sanchez, O.; Flores-Ruiz, F.J.; Prokhorov, E.; de Fuentes, O.A.; Esparza, R.; Meyyappan, M. Surface functionalized halloysite nanotubes decorated with silver nanoparticles for enzyme immobilization and biosensing. J. Mater. Chem. B 2016, 4, 2553–2560. [Google Scholar] [CrossRef]
  255. Hashemifard, N.; Mohsenifar, A.; Ranjbar, B.; Allameh, A.; Lotfi, A.S.; Etemadikia, B. Fabrication and kinetic studies of a novel silver nanoparticles–glucose oxidase bioconjugate. Anal. Chim. Acta 2010, 675, 181–184. [Google Scholar] [CrossRef]
  256. Ma, S.; Mu, J.; Qu, Y.; Jiang, L. Effect of refluxed silver nanoparticles on inhibition and enhancement of enzymatic activity of glucose oxidase. Colloids Surf. A Physicochem. Eng. Asp. 2009, 345, 101–105. [Google Scholar] [CrossRef]
  257. Netto, C.G.C.M.; Toma, H.E.; Andrade, L.H. Superparamagnetic nanoparticles as versatile carriers and supporting materials for enzymes. J. Mol. Catal. B Enzym. 2013, 85–86, 71–92. [Google Scholar] [CrossRef]
  258. Govan, J.; Gun’ko, Y.K. Recent Advances in the Application of Magnetic Nanoparticles as a Support for Homogeneous Catalysts. Nanomaterials 2014, 4, 222–241. [Google Scholar] [CrossRef] [PubMed]
  259. Kouassi, G.K.; Irudayaraj, J.; McCarty, G. Examination of Cholesterol oxidase attachment to magnetic nanoparticles. J. Nanobiotechnol. 2005, 3, 1. [Google Scholar] [CrossRef]
  260. Johnson, P.A.; Park, H.J.; Driscoll, A.J. Enzyme Nanoparticle Fabrication: Magnetic Nanoparticle Synthesis and Enzyme Immobilization. In Enzyme Stabilization and Immobilization: Methods and Protocols; Minteer, S.D., Ed.; Humana Press: Totowa, NJ, USA, 2011; pp. 183–191. [Google Scholar] [CrossRef]
  261. Liao, M.-H.; Chen, D.-H. Immobilization of yeast alcohol dehydrogenase on magnetic nanoparticles for improving its stability. Biotechnol. Lett. 2001, 23, 1723–1727. [Google Scholar] [CrossRef]
  262. Xu, J.; Huo, S.; Yuan, Z.; Zhang, Y.; Xu, H.; Guo, Y.; Liang, C.; Zhuang, X. Characterization of direct cellulase immobilization with superparamagnetic nanoparticles. Biocatal. Biotransform. 2011, 29, 71–76. [Google Scholar] [CrossRef]
  263. Rossi, L.M.; Quach, A.D.; Rosenzweig, Z. Glucose oxidase–magnetite nanoparticle bioconjugate for glucose sensing. Anal. Bioanal. Chem. 2004, 380, 606–613. [Google Scholar] [CrossRef]
  264. Bilal, M.; Zhao, Y.; Rasheed, T.; Iqbal, H.M.N. Magnetic nanoparticles as versatile carriers for enzymes immobilization: A review. Int. J. Biol. Macromol. 2018, 120, 2530–2544. [Google Scholar] [CrossRef]
  265. Patel, V.; Shah, C.; Deshpande, M.; Madamwar, D. Zinc Oxide Nanoparticles Supported Lipase Immobilization for Biotransformation in Organic Solvents: A Facile Synthesis of Geranyl Acetate, Effect of Operative Variables and Kinetic Study. Appl. Biochem. Biotechnol. 2016, 178, 1630–1651. [Google Scholar] [CrossRef]
  266. Diyanat, S.; Homaei, A.; Mosaddegh, E. Immobilization of Penaeus vannamei protease on ZnO nanoparticles for long-term use. Int. J. Biol. Macromol. 2018, 118, 92–98. [Google Scholar] [CrossRef]
  267. Shi, X.; Gu, W.; Li, B.; Chen, N.; Zhao, K.; Xian, Y. Enzymatic biosensors based on the use of metal oxide nanoparticles. Microchim. Acta 2014, 181, 1–22. [Google Scholar] [CrossRef]
  268. Palanisamy, S.; Karuppiah, C.; Chen, S.-M.; Periakaruppan, P. A Highly Sensitive and Selective Enzymatic Biosensor Based on Direct Electrochemistry of Hemoglobin at Zinc Oxide Nanoparticles Modified Activated Screen Printed Carbon Electrode. Electroanalysis 2014, 26, 1984–1993. [Google Scholar] [CrossRef]
  269. Antony, N.; Balachandran, S.; Mohanan, P.V. Immobilization of diastase α-amylase on nano zinc oxide. Food Chem. 2016, 211, 624–630. [Google Scholar] [CrossRef] [PubMed]
  270. Li, Y.-F.; Liu, Z.-M.; Liu, Y.-L.; Yang, Y.-H.; Shen, G.-L.; Yu, R.-Q. A mediator-free phenol biosensor based on immobilizing tyrosinase to ZnO nanoparticles. Anal. Biochem. 2006, 349, 33–40. [Google Scholar] [CrossRef] [PubMed]
  271. Ansari, S.A.; Husain, Q.; Qayyum, S.; Azam, A. Designing and surface modification of zinc oxide nanoparticles for biomedical applications. Food Chem. Toxicol. 2011, 49, 2107–2115. [Google Scholar] [CrossRef]
  272. Ren, X.; Chen, D.; Meng, X.; Tang, F.; Hou, X.; Han, D.; Zhang, L. Zinc oxide nanoparticles/glucose oxidase photoelectrochemical system for the fabrication of biosensor. J. Colloid Interface Sci. 2009, 334, 183–187. [Google Scholar] [CrossRef] [PubMed]
  273. Verma, N.; Kumar, N.; Upadhyay, L.S.B.; Sahu, R.; Dutt, A. Fabrication and Characterization of Cysteine-Functionalized Zinc Oxide Nanoparticles for Enzyme Immobilization. Anal. Lett. 2017, 50, 1839–1850. [Google Scholar] [CrossRef]
  274. Kaur, G.; Negi, P.; Kaur, M.; Sharma, R.; Konwar, R.J.; Mahajan, A. Morpho-structural and opto-electrical properties of chemically tuned nanostructured TiO2. Ceram. Int. 2018, 44, 18484–18490. [Google Scholar] [CrossRef]
  275. Haghighi, N.; Hallaj, R.; Salimi, A. Immobilization of glucose oxidase onto a novel platform based on modified TiO2 and graphene oxide, direct electrochemistry, catalytic and photocatalytic activity. Mater. Sci. Eng. C 2017, 73, 417–424. [Google Scholar] [CrossRef]
  276. Romero-Arcos, M.; Garnica-Romo, M.G.; Martinez-Flores, H.E.; Vázquez-Marrufo, G.; Ramírez-Bon, R.; González-Hernández, J.; Barbosa-Cánovas, G.V. Enzyme Immobilization by Amperometric Biosensors with TiO2 Nanoparticles Used to Detect Phenol Compounds. Food Eng. Rev. 2016, 8, 235–250. [Google Scholar] [CrossRef]
  277. Zhang, Y.; He, P.; Hu, N. Horseradish peroxidase immobilized in TiO2 nanoparticle films on pyrolytic graphite electrodes: Direct electrochemistry and bioelectrocatalysis. Electrochim. Acta 2004, 49, 1981–1988. [Google Scholar] [CrossRef]
  278. Shetti, N.P.; Bukkitgar, S.D.; Reddy, K.R.; Reddy, C.V.; Aminabhavi, T.M. Nanostructured titanium oxide hybrids-based electrochemical biosensors for healthcare applications. Colloids Surf. B Biointerfaces 2019, 178, 385–394. [Google Scholar] [CrossRef] [PubMed]
  279. Gholivand, M.B.; Shamsipur, M.; Amini, N. Nonenzymatic L-lysine amino acid detection using titanium oxide nanoparticles/multi wall carbon nanotube composite electrodes. Electrochim. Acta 2014, 123, 569–575. [Google Scholar] [CrossRef]
  280. Ahmad, R.; Sardar, M. Immobilization of cellulase on TiO2 nanoparticles by physical and covalent methods: A comparative study. Indian J. Biochem. Biophys. 2014, 51, 7. [Google Scholar]
  281. Yong, Y.; Bai, Y.; Li, Y.; Lin, L.; Cui, Y.; Xia, C. Preparation and application of polymer-grafted magnetic nanoparticles for lipase immobilization. J. Magn. Magn. Mater. 2008, 320, 2350–2355. [Google Scholar] [CrossRef]
  282. Hong, R.Y.; Qian, J.Z.; Cao, J.X. Synthesis and characterization of PMMA grafted ZnO nanoparticles. Powder Technol. 2006, 163, 160–168. [Google Scholar] [CrossRef]
  283. Mahouche-Chergui, S.; Guerrouache, M.; Carbonnier, B.; Chehimi, M.M. Polymer-immobilized nanoparticles. Colloids Surf. A Physicochem. Eng. Asp. 2013, 439, 43–68. [Google Scholar] [CrossRef]
  284. Xu, C.; Ohno, K.; Ladmiral, V.; Composto, R.J. Dispersion of polymer-grafted magnetic nanoparticles in homopolymers and block copolymers. Polymer 2008, 49, 3568–3577. [Google Scholar] [CrossRef]
  285. Kango, S.; Kalia, S.; Celli, A.; Njuguna, J.; Habibi, Y.; Kumar, R. Surface modification of inorganic nanoparticles for development of organic–inorganic nanocomposites—A review. Prog. Polym. Sci. 2013, 38, 1232–1261. [Google Scholar] [CrossRef]
  286. Silmore, K.S.; Gupta, C.; Washburn, N.R. Tunable Pickering emulsions with polymer-grafted lignin nanoparticles (PGLNs). J. Colloid Interface Sci. 2016, 466, 91–100. [Google Scholar] [CrossRef]
  287. Liu, H.; Zhao, H.-Y.; Müller-Plathe, F.; Qian, H.-J.; Sun, Z.-Y.; Lu, Z.-Y. Distribution of the Number of Polymer Chains Grafted on Nanoparticles Fabricated by Grafting-to and Grafting-from Procedures. Macromolecules 2018, 51, 3758–3766. [Google Scholar] [CrossRef]
  288. Gann, J.P.; Yan, M. A Versatile Method for Grafting Polymers on Nanoparticles. Langmuir 2008, 24, 5319–5323. [Google Scholar] [CrossRef] [PubMed]
  289. Hou, Z.; Liu, Y.; Xu, J.; Zhu, J. Surface engineering of magnetic iron oxide nanoparticles by polymer grafting: Synthesis progress and biomedical applications. Nanoscale 2020, 12, 14957–14975. [Google Scholar] [CrossRef] [PubMed]
  290. Dutta, S.; Parida, S.; Maiti, C.; Banerjee, R.; Mandal, M.; Dhara, D. Polymer grafted magnetic nanoparticles for delivery of anticancer drug at lower pH and elevated temperature. J. Colloid Interface Sci. 2016, 467, 70–80. [Google Scholar] [CrossRef] [PubMed]
  291. Wang, S.; Wang, Z.; Li, J.; Li, L.; Hu, W. Surface-grafting polymers: From chemistry to organic electronics. Mater. Chem. Front. 2020, 4, 692–714. [Google Scholar] [CrossRef]
  292. Babu, K.; Dhamodharan, R. Synthesis of Polymer Grafted Magnetite Nanoparticle with the Highest Grafting Density via Controlled Radical Polymerization. Nanoscale Res. Lett. 2009, 4, 1090. [Google Scholar] [CrossRef]
  293. Liu, C.-H.; Pan, C.-Y. Grafting polystyrene onto silica nanoparticles via RAFT polymerization. Polymer 2007, 48, 3679–3685. [Google Scholar] [CrossRef]
  294. Tawade, B.V.; Apata, I.E.; Pradhan, N.; Karim, A.; Raghavan, D. Recent Advances in the Synthesis of Polymer-Grafted Low-K and High-K Nanoparticles for Dielectric and Electronic Applications. Molecules 2021, 26, 2942. [Google Scholar] [CrossRef]
  295. Reddy, K.R.; Lee, K.P.; Gopalan, A.I. Self-assembly approach for the synthesis of electro-magnetic functionalized Fe3O4/polyaniline nanocomposites: Effect of dopant on the properties. Colloids Surf. A Physicochem. Eng. Asp. 2008, 320, 49–56. [Google Scholar] [CrossRef]
  296. Gazit, O.; Cohen, Y.; Tannenbaum, R. Periodic nanocomposites: A simple path for the preferential self-assembly of nanoparticles in block-copolymers. Polymer 2010, 51, 2185–2190. [Google Scholar] [CrossRef]
  297. Harrison, A.; Vuong, T.T.; Zeevi, M.P.; Hittel, B.J.; Wi, S.; Tang, C. Rapid Self-Assembly of Metal/Polymer Nanocomposite Particles as Nanoreactors and Their Kinetic Characterization. Nanomaterials 2019, 9, 318. [Google Scholar] [CrossRef]
  298. Ahmed, S.R.; Kofinas, P. Controlled Room Temperature Synthesis of CoFe2O4 Nanoparticles through a Block Copolymer Nanoreactor Route. Macromolecules 2002, 35, 3338–3341. [Google Scholar] [CrossRef]
  299. Grzelczak, M.; Vermant, J.; Furst, E.M.; Liz-Marzán, L.M. Directed Self-Assembly of Nanoparticles. ACS Nano 2010, 4, 3591–3605. [Google Scholar] [CrossRef] [PubMed]
  300. Cong, H.-P.; Yu, S.-H. Self-assembly of functionalized inorganic–organic hybrids. Curr. Opin. Colloid Interface Sci. 2009, 14, 71–80. [Google Scholar] [CrossRef]
  301. Liu, X.; Lei, L.; Li, Y.; Zhu, H.; Cui, Y.; Hu, H. Preparation of carriers based on magnetic nanoparticles grafted polymer and immobilization for lipase. Biochem. Eng. J. 2011, 56, 142–149. [Google Scholar] [CrossRef]
  302. Kumar, S.K.; Jouault, N.; Benicewicz, B.; Neely, T. Nanocomposites with Polymer Grafted Nanoparticles. Macromolecules 2013, 46, 3199–3214. [Google Scholar] [CrossRef]
  303. Eskandari, P.; Abousalman-Rezvani, Z.; Roghani-Mamaqani, H.; Salami-Kalajahi, M.; Mardani, H. Polymer grafting on graphene layers by controlled radical polymerization. Adv. Colloid Interface Sci. 2019, 273, 102021. [Google Scholar] [CrossRef]
  304. Fan, X.; Yang, J.; Loh, X.J.; Li, Z. Polymeric Janus Nanoparticles: Recent Advances in Synthetic Strategies, Materials Properties, and Applications. Macromol. Rapid Commun. 2019, 40, 1800203. [Google Scholar] [CrossRef]
  305. Messina, M.S.; Messina, K.M.M.; Bhattacharya, A.; Montgomery, H.R.; Maynard, H.D. Preparation of biomolecule-polymer conjugates by grafting-from using ATRP, RAFT, or ROMP. Prog. Polym. Sci. 2020, 100, 101186. [Google Scholar] [CrossRef]
  306. Crespilho, F.N.; Emilia Ghica, M.; Florescu, M.; Nart, F.C.; Oliveira, O.N.; Brett, C.M.A. A strategy for enzyme immobilization on layer-by-layer dendrimer–gold nanoparticle electrocatalytic membrane incorporating redox mediator. Electrochem. Commun. 2006, 8, 1665–1670. [Google Scholar] [CrossRef]
  307. Yang, Y.; Wang, Z.; Yang, M.; Guo, M.; Wu, Z.; Shen, G.; Yu, R. Inhibitive determination of mercury ion using a renewable urea biosensor based on self-assembled gold nanoparticles. Sens. Actuators B Chem. 2006, 114, 1–8. [Google Scholar] [CrossRef]
  308. Zhong, X.; Yuan, R.; Chai, Y.; Liu, Y.; Dai, J.; Tang, D. Glucose biosensor based on self-assembled gold nanoparticles and double-layer 2d-network (3-mercaptopropyl)-trimethoxysilane polymer onto gold substrate. Sens. Actuators B Chem. 2005, 104, 191–198. [Google Scholar] [CrossRef]
  309. Bagal-Kestwal, D.; Kestwal, R.M.; Hsieh, B.-C.; Chen, R.L.C.; Cheng, T.-J.; Chiang, B.-H. Electrochemical β(1→3)-d-glucan biosensors fabricated by immobilization of enzymes with gold nanoparticles on platinum electrode. Biosens. Bioelectron. 2010, 26, 118–125. [Google Scholar] [CrossRef] [PubMed]
  310. Zou, Y.; Xiang, C.; Sun, L.-X.; Xu, F. Glucose biosensor based on electrodeposition of platinum nanoparticles onto carbon nanotubes and immobilizing enzyme with chitosan-SiO2 sol–gel. Biosens. Bioelectron. 2008, 23, 1010–1016. [Google Scholar] [CrossRef]
  311. Lu, Y.; Yang, M.; Qu, F.; Shen, G.; Yu, R. Enzyme-functionalized gold nanowires for the fabrication of biosensors. Bioelectrochemistry 2007, 71, 211–216. [Google Scholar] [CrossRef] [PubMed]
  312. Kesik, M.; Kanik, F.E.; Hızalan, G.; Kozanoglu, D.; Esenturk, E.N.; Timur, S.; Toppare, L. A functional immobilization matrix based on a conducting polymer and functionalized gold nanoparticles: Synthesis and its application as an amperometric glucose biosensor. Polymer 2013, 54, 4463–4471. [Google Scholar] [CrossRef]
  313. Perveen, R.; Nasar, A.; Inamuddin; Asiri, A.M.; Mishra, A.K. Optimization of MnO2-Graphene/polythioaniline (MnO2-G/PTA) hybrid nanocomposite for the application of biofuel cell bioanode. Int. J. Hydrogen Energy 2018, 43, 15144–15154. [Google Scholar] [CrossRef]
  314. Tan, Y.; Deng, W.; Chen, C.; Xie, Q.; Lei, L.; Li, Y.; Fang, Z.; Ma, M.; Chen, J.; Yao, S. Immobilization of enzymes at high load/activity by aqueous electrodeposition of enzyme-tethered chitosan for highly sensitive amperometric biosensing. Biosens. Bioelectron. 2010, 25, 2644–2650. [Google Scholar] [CrossRef]
  315. Inozemtseva, O.A.; Salkovskiy, Y.E.; Severyukhina, A.N.; Vidyasheva, I.V.; Petrova, N.V.; Metwally, H.A.; Stetciura, I.Y.; Gorin, D.A. Electrospinning of functional materials for biomedicine and tissue engineering. Russ. Chem. Rev. 2015, 84, 251–274. [Google Scholar] [CrossRef]
  316. Sharma, D.; Satapathy, B.K. Polymer Substrate-Based Transition Metal Modified Electrospun Nanofibrous Materials: Current Trends in Functional Applications and Challenges. Polym. Rev. 2022, 62, 439–484. [Google Scholar] [CrossRef]
  317. Feng, C.; Khulbe, K.C.; Matsuura, T. Recent progress in the preparation, characterization, and applications of nanofibers and nanofiber membranes via electrospinning/interfacial polymerization. J. Appl. Polym. Sci. 2010, 115, 756–776. [Google Scholar] [CrossRef]
  318. Razmgar, K.; Nasiraee, M. Polyvinyl alcohol-based membranes for filtration of aqueous solutions: A comprehensive review. Polym. Eng. Sci. 2021, 62, 25–43. [Google Scholar] [CrossRef]
  319. Golshaei, R.; Karazehir, T.; Ghoreishi, S.M.; Ates, M.; Sarac, A.S. Glucose oxidase immobilization onto Au/poly[anthranilic acid-co-3-carboxy-N-(2-thenylidene)aniline]/PVAc electrospun nanofibers. Polym. Bull. 2017, 74, 1493–1517. [Google Scholar] [CrossRef]
  320. Sriwichai, S.; Phanichphant, S. Fabrication and characterization of electrospun poly(3-aminobenzylamine)/functionalized multi-walled carbon nanotubes composite film for electrochemical glucose biosensor. Express Polym. Lett. 2022, 16, 439–450. [Google Scholar] [CrossRef]
  321. Long, Y.-Z.; Yan, X.; Wang, X.-X.; Zhang, J.; Yu, M. Chapter 2—Electrospinning: The Setup and Procedure. In Electrospinning: Nanofabrication and Applications; Ding, B., Wang, X., Yu, J., Eds.; William Andrew Publishing: Norwich, NY, USA, 2019; pp. 21–52. [Google Scholar] [CrossRef]
  322. Rodriguez-Abetxuko, A.; Sánchez-deAlcázar, D.; Muñumer, P.; Beloqui, A. Tunable Polymeric Scaffolds for Enzyme Immobilization. Front. Bioeng. Biotechnol. 2020, 8, 830. [Google Scholar] [CrossRef] [PubMed]
  323. Son, H.Y.; Ryu, J.H.; Lee, H.; Nam, Y.S. Bioinspired Templating Synthesis of Metal–Polymer Hybrid Nanostructures within 3D Electrospun Nanofibers. ACS Appl. Mater. Interfaces 2013, 5, 6381–6390. [Google Scholar] [CrossRef]
  324. Fu, J.; Li, D.; Li, G.; Huang, F.; Wei, Q. Carboxymethyl cellulose assisted immobilization of silver nanoparticles onto cellulose nanofibers for the detection of catechol. J. Electroanal. Chem. 2015, 738, 92–99. [Google Scholar] [CrossRef]
  325. Kang, S.; Zhao, K.; Yu, D.-G.; Zheng, X.; Huang, C. Advances in Biosensing and Environmental Monitoring Based on Electrospun Nanofibers. Adv. Fiber Mater. 2022, 4, 404–435. [Google Scholar] [CrossRef]
  326. Du, D.; Chen, S.; Cai, J.; Zhang, A. Immobilization of acetylcholinesterase on gold nanoparticles embedded in sol–gel film for amperometric detection of organophosphorous insecticide. Biosens. Bioelectron. 2007, 23, 130–134. [Google Scholar] [CrossRef]
  327. Wang, B. An amperometric β-glucan biosensor based on the immobilization of bi-enzyme on Prussian blue–chitosan and gold nanoparticles–chitosan nanocomposite films. Biosens. Bioelectron. 2014, 55, 113–119. [Google Scholar] [CrossRef]
  328. Regina Silva, T.; Cruz Vieira, I. A biosensor based on gold nanoparticles stabilized in poly(allylamine hydrochloride) and decorated with laccase for determination of dopamine. Analyst 2016, 141, 216–224. [Google Scholar] [CrossRef]
  329. German, N.; Kausaite-Minkstimiene, A.; Ramanavicius, A.; Semashko, T.; Mikhailova, R.; Ramanaviciene, A. The use of different glucose oxidases for the development of an amperometric reagentless glucose biosensor based on gold nanoparticles covered by polypyrrole. Electrochim. Acta 2015, 169, 326–333. [Google Scholar] [CrossRef]
  330. Chen, S.; Yuan, R.; Chai, Y.; Hu, F. Electrochemical sensing of hydrogen peroxide using metal nanoparticles: A review. Microchim. Acta 2013, 180, 15–32. [Google Scholar] [CrossRef]
  331. Meier, J.; Hofferber, E.M.; Stapleton, J.A.; Iverson, N.M. Hydrogen Peroxide Sensors for Biomedical Applications. Chemosensors 2019, 7, 64. [Google Scholar] [CrossRef]
  332. Pundir, C.S.; Deswal, R.; Narwal, V. Quantitative analysis of hydrogen peroxide with special emphasis on biosensors. Bioprocess Biosyst. Eng. 2018, 41, 313–329. [Google Scholar] [CrossRef] [PubMed]
  333. Yao, Y.; Wen, Y.; Zhang, L.; Xu, J.; Wang, Z.; Duan, X. A Stable Sandwich-Type Hydrogen Peroxide Sensor Based on Immobilizing Horseradish Peroxidase to a Silver Nanoparticle Monolayer Supported by PEDOT:PSS-Nafion Composite Electrode. Int. J. Electrochem. Sci. 2013, 8, 12. [Google Scholar] [CrossRef]
  334. Ansari, A.A.; Sumana, G.; Khan, R.; Malhotra, B.D. Polyaniline-cerium oxide nanocomposite for hydrogen peroxide sensor. J. Nanosci. Nanotechnol. 2009, 9, 4679–4685. [Google Scholar] [CrossRef]
  335. Ahammad, A.J.S. Hydrogen peroxide biosensors based on horseradish peroxidase and hemoglobin. J. Biosens. Bioelectron. 2013, S9, 001. [Google Scholar] [CrossRef]
  336. Liu, X.; Luo, L.; Ding, Y.; Xu, Y.; Li, F. Hydrogen peroxide biosensor based on the immobilization of horseradish peroxidase on γ-Al2O3 nanoparticles/chitosan film-modified electrode. J. Solid State Electrochem. 2011, 15, 447–453. [Google Scholar] [CrossRef]
  337. Rad, A.S.; Mirabi, A.; Binaian, E.; Tayebi, H. A Review on Glucose and Hydrogen Peroxide Biosensor Based on Modified Electrode Included Silver Nanoparticles. Int. J. Electrochem. Sci. 2011, 6, 3671–3683. [Google Scholar] [CrossRef]
  338. Zhang, T.; Wang, W.; Zhang, D.; Zhang, X.; Ma, Y.; Zhou, Y.; Qi, L. Biotemplated Synthesis of Gold Nanoparticle–Bacteria Cellulose Nanofiber Nanocomposites and Their Application in Biosensing. Adv. Funct. Mater. 2010, 20, 1152–1160. [Google Scholar] [CrossRef]
  339. Xiang, C.; Zou, Y.; Sun, L.-X.; Xu, F. Direct Electron Transfer of Horseradish Peroxidase and Its Biosensor Based on Gold Nanoparticles/Chitosan/ITO Modified Electrode. Anal. Lett. 2008, 41, 2224–2236. [Google Scholar] [CrossRef]
  340. Chico, B.; Camacho, C.; Pérez, M.; Longo, M.A.; Sanromán, M.A.; Pingarrón, J.M.; Villalonga, R. Polyelectrostatic immobilization of gold nanoparticles-modified peroxidase on alginate-coated gold electrode for mediatorless biosensor construction. J. Electroanal. Chem. 2009, 629, 126–132. [Google Scholar] [CrossRef]
  341. Xu, Q.; Mao, C.; Liu, N.-N.; Zhu, J.-J.; Sheng, J. Direct electrochemistry of horseradish peroxidase based on biocompatible carboxymethyl chitosan–gold nanoparticle nanocomposite. Biosens. Bioelectron. 2006, 22, 768–773. [Google Scholar] [CrossRef] [PubMed]
  342. Lin, J.; Qu, W.; Zhang, S. Disposable biosensor based on enzyme immobilized on Au–chitosan-modified indium tin oxide electrode with flow injection amperometric analysis. Anal. Biochem. 2007, 360, 288–293. [Google Scholar] [CrossRef] [PubMed]
  343. Lu, W.; Tao, L.; Wang, Y.; Ge, J.; Dong, J.; Qian, W. In-situ synthesis of silver nanoparticles on resin microspheres composed of poly(m-aminophenol), and their application in an enzymatic glucose biosensor. Microchim. Acta 2015, 182, 479–486. [Google Scholar] [CrossRef]
  344. Khumngern, S.; Jirakunakorn, R.; Thavarungkul, P.; Kanatharana, P.; Numnuam, A. A highly sensitive flow injection amperometric glucose biosensor using a gold nanoparticles/polytyramine/Prussian blue modified screen-printed carbon electrode. Bioelectrochemistry 2021, 138, 107718. [Google Scholar] [CrossRef]
  345. Anusha, J.R.; Kim, H.-J.; Fleming, A.T.; Das, S.J.; Yu, K.-H.; Kim, B.C.; Raj, C.J. Simple fabrication of ZnO/Pt/chitosan electrode for enzymatic glucose biosensor. Sens. Actuators B Chem. 2014, 202, 827–833. [Google Scholar] [CrossRef]
  346. Xu, L.; Zhu, Y.; Tang, L.; Yang, X.; Li, C. Dendrimer-encapsulated Pt nanoparticles/polyaniline nanofibers for glucose detection. J. Appl. Polym. Sci. 2008, 109, 1802–1807. [Google Scholar] [CrossRef]
  347. Sanaeifar, N.; Rabiee, M.; Abdolrahim, M.; Tahriri, M.; Vashaee, D.; Tayebi, L. A novel electrochemical biosensor based on Fe3O4 nanoparticles-polyvinyl alcohol composite for sensitive detection of glucose. Anal. Biochem. 2017, 519, 19–26. [Google Scholar] [CrossRef]
  348. Hassan, M.H.; Vyas, C.; Grieve, B.; Bartolo, P. Recent Advances in Enzymatic and Non-Enzymatic Electrochemical Glucose Sensing. Sensors 2021, 21, 4672. [Google Scholar] [CrossRef]
  349. Naderi Asrami, P.; Mozaffari, S.A.; Saber Tehrani, M.; Aberoomand Azar, P. A novel impedimetric glucose biosensor based on immobilized glucose oxidase on a CuO-Chitosan nanobiocomposite modified FTO electrode. Int. J. Biol. Macromol. 2018, 118, 649–660. [Google Scholar] [CrossRef]
  350. Li, J.; Lin, X. Glucose biosensor based on immobilization of glucose oxidase in poly(o-aminophenol) film on polypyrrole-Pt nanocomposite modified glassy carbon electrode. Biosens. Bioelectron. 2007, 22, 2898–2905. [Google Scholar] [CrossRef] [PubMed]
  351. Li, J.; Yuan, R.; Chai, Y.; Che, X.; Li, W. Construction of an amperometric glucose biosensor based on the immobilization of glucose oxidase onto electrodeposited Pt nanoparticles-chitosan composite film. Bioprocess Biosyst. Eng. 2012, 35, 1089–1095. [Google Scholar] [CrossRef] [PubMed]
  352. Maniruzzaman, M.; Jang, S.-D.; Kim, J. Titanium dioxide–cellulose hybrid nanocomposite and its glucose biosensor application. Mater. Sci. Eng. B 2012, 177, 844–848. [Google Scholar] [CrossRef]
  353. Fedorenko, V.; Damberga, D.; Grundsteins, K.; Ramanavicius, A.; Ramanavicius, S.; Coy, E.; Iatsunskyi, I.; Viter, R. Application of Polydopamine Functionalized Zinc Oxide for Glucose Biosensor Design. Polymers 2021, 13, 2918. [Google Scholar] [CrossRef]
  354. Zhai, D.; Liu, B.; Shi, Y.; Pan, L.; Wang, Y.; Li, W.; Zhang, R.; Yu, G. Highly Sensitive Glucose Sensor Based on Pt Nanoparticle/Polyaniline Hydrogel Heterostructures. ACS Nano 2013, 7, 3540–3546. [Google Scholar] [CrossRef]
  355. Shukla, S.K.; Deshpande, S.R.; Shukla, S.K.; Tiwari, A. Fabrication of a tunable glucose biosensor based on zinc oxide/chitosan-graft-poly(vinyl alcohol) core-shell nanocomposite. Talanta 2012, 99, 283–287. [Google Scholar] [CrossRef]
  356. Xue, M.-H.; Xu, Q.; Zhou, M.; Zhu, J.-J. In Situ immobilization of glucose oxidase in chitosan–gold nanoparticle hybrid film on Prussian Blue modified electrode for high-sensitivity glucose detection. Electrochem. Commun. 2006, 8, 1468–1474. [Google Scholar] [CrossRef]
  357. Wang, W.; Li, H.-Y.; Zhang, D.-W.; Jiang, J.; Cui, Y.-R.; Qiu, S.; Zhou, Y.-L.; Zhang, X.-X. Fabrication of Bienzymatic Glucose Biosensor Based on Novel Gold Nanoparticles-Bacteria Cellulose Nanofibers Nanocomposite. Electroanalysis 2010, 22, 2543–2550. [Google Scholar] [CrossRef]
  358. Luo, L.; Li, Q.; Xu, Y.; Ding, Y.; Wang, X.; Deng, D.; Xu, Y. Amperometric glucose biosensor based on NiFe2O4 nanoparticles and chitosan. Sens. Actuators B Chem. 2010, 145, 293–298. [Google Scholar] [CrossRef]
  359. Paik, E.-S.; Kim, Y.-R.; Hong, H.-G. Amperometric Glucose Biosensor Utilizing Zinc Oxide-chitosan- glucose Oxidase Hybrid Composite Films on Electrodeposited Pt-Fe(III). Anal. Sci. 2018, 34, 6. [Google Scholar] [CrossRef] [PubMed]
  360. Santhosh, P.; Manesh, K.M.; Uthayakumar, S.; Komathi, S.; Gopalan, A.I. Fabrication of enzymatic glucose biosensor based on palladium nanoparticles dispersed onto poly(3,4-ethylenedioxythiophene) nanofibers. Bioelectrochemistry 2009, 75, 61–66. [Google Scholar] [CrossRef] [PubMed]
  361. Chaichi, M.J.; Ehsani, M. A novel glucose sensor based on immobilization of glucose oxidase on the chitosan-coated Fe3O4 nanoparticles and the luminol–H2O2–gold nanoparticle chemiluminescence detection system. Sens. Actuators B Chem. 2016, 223, 713–722. [Google Scholar] [CrossRef]
  362. Sapountzi, E.; Braiek, M.; Vocanson, F.; Chateaux, J.-F.; Jaffrezic-Renault, N.; Lagarde, F. Gold nanoparticles assembly on electrospun poly(vinyl alcohol)/poly(ethyleneimine)/glucose oxidase nanofibers for ultrasensitive electrochemical glucose biosensing. Sens. Actuators B Chem. 2017, 238, 392–401. [Google Scholar] [CrossRef]
  363. Wu, B.-Y.; Hou, S.-H.; Yin, F.; Li, J.; Zhao, Z.-X.; Huang, J.-D.; Chen, Q. Amperometric glucose biosensor based on layer-by-layer assembly of multilayer films composed of chitosan, gold nanoparticles and glucose oxidase modified Pt electrode. Biosens. Bioelectron. 2007, 22, 838–844. [Google Scholar] [CrossRef]
  364. Wang, L.; Zeng, Y.; Shen, A.; Fu, Y.; Zeng, L.; Hu, J. Facile and controllable synthesis of triplex Au@Ag–Pt@infinite coordination polymer core–shell nanoparticles for highly efficient immobilization of enzymes and enhanced electrochemical biosensing activity. RSC Adv. 2016, 6, 86025–86033. [Google Scholar] [CrossRef]
  365. Miao, Z.; Wang, P.; Zhong, A.; Yang, M.; Xu, Q.; Hao, S.; Hu, X. Development of a glucose biosensor based on electrodeposited gold nanoparticles–polyvinylpyrrolidone–polyaniline nanocomposites. J. Electroanal. Chem. 2015, 756, 153–160. [Google Scholar] [CrossRef]
  366. Ren, J.; Shi, W.; Li, K.; Ma, Z. Ultrasensitive platinum nanocubes enhanced amperometric glucose biosensor based on chitosan and nafion film. Sens. Actuators B Chem. 2012, 163, 115–120. [Google Scholar] [CrossRef]
  367. Sharma, A.; Kumar, A. Study of structural and electro-catalytic behaviour of amperometric biosensor based on chitosan/polypyrrole nanotubes-gold nanoparticles nanocomposites. Synth. Met. 2016, 220, 551–559. [Google Scholar] [CrossRef]
  368. Bagal-Kestwal, D.R.; Kestwal, R.M.; Hsieh, W.-T.; Chiang, B.-H. Chitosan–guar gum–silver nanoparticles hybrid matrix with immobilized enzymes for fabrication of beta-glucan and glucose sensing photometric flow injection system. J. Pharm. Biomed. Anal. 2014, 88, 571–578. [Google Scholar] [CrossRef]
  369. Yang, Y.; Yang, H.; Yang, M. Amperometric glucose biosensor based on a surface treated nanoporous ZrO2/Chitosan composite film as immobilization matrix. Anal. Chim. Acta 2004, 525, 213–220. [Google Scholar] [CrossRef]
  370. Yang, L.; Ren, X.; Tang, F.; Zhang, L. A practical glucose biosensor based on Fe3O4 nanoparticles and chitosan/nafion composite film. Biosens. Bioelectron. 2009, 25, 889–895. [Google Scholar] [CrossRef] [PubMed]
  371. Buk, V.; Emregul, E.; Emregul, K.C. Alginate copper oxide nano-biocomposite as a novel material for amperometric glucose biosensing. Mater. Sci. Eng. C 2017, 74, 307–314. [Google Scholar] [CrossRef] [PubMed]
  372. Majumdar, S.; Mahanta, D. Deposition of an ultra-thin polyaniline coating on a TiO2 surface by vapor phase polymerization for electrochemical glucose sensing and photocatalytic degradation. RSC Adv. 2020, 10, 17387–17395. [Google Scholar] [CrossRef] [PubMed]
  373. German, N.; Ramanaviciene, A.; Ramanavicius, A. Dispersed Conducting Polymer Nanocomposites with Glucose Oxidase and Gold Nanoparticles for the Design of Enzymatic Glucose Biosensors. Polymers 2021, 13, 2173. [Google Scholar] [CrossRef]
  374. Ansari, A.A.; Kaushik, A.; Solanki, P.R.; Malhotra, B.D. Electrochemical Cholesterol Sensor Based on Tin Oxide-Chitosan Nanobiocomposite Film. Electroanalysis 2009, 21, 965–972. [Google Scholar] [CrossRef]
  375. Gomathi, P.; Ragupathy, D.; Choi, J.H.; Yeum, J.H.; Lee, S.C.; Kim, J.C.; Lee, S.H.; Ghim, H.D. Fabrication of novel chitosan nanofiber/gold nanoparticles composite towards improved performance for a cholesterol sensor. Sens. Actuators B Chem. 2011, 153, 44–49. [Google Scholar] [CrossRef]
  376. Malhotra, B.D.; Kaushik, A. Metal oxide–chitosan based nanocomposite for cholesterol biosensor. Thin Solid Film. 2009, 518, 614–620. [Google Scholar] [CrossRef]
  377. Khan, R.; Kaushik, A.; Solanki, P.R.; Ansari, A.A.; Pandey, M.K.; Malhotra, B.D. Zinc oxide nanoparticles-chitosan composite film for cholesterol biosensor. Anal. Chim. Acta 2008, 616, 207–213. [Google Scholar] [CrossRef]
  378. Singh, J.; Srivastava, M.; Kalita, P.; Malhotra, B.D. A novel ternary NiFe2O4/CuO/FeO-chitosan nanocomposite as a cholesterol biosensor. Process Biochem. 2012, 47, 2189–2198. [Google Scholar] [CrossRef]
  379. Ali, A.; AlSalhi, M.S.; Atif, M.; Ansari, A.A.; Israr, M.Q.; Sadaf, J.R.; Ahmed, E.; Nur, O.; Willander, M. Potentiometric urea biosensor utilizing nanobiocomposite of chitosan-iron oxide magnetic nanoparticles. J. Phys. Conf. Ser. 2013, 414, 012024. [Google Scholar] [CrossRef]
  380. Ali, A.; Israr-Qadir, M.; Wazir, Z.; Tufail, M.; Ibupoto, Z.H.; Jamil-Rana, S.; Atif, M.; Khan, S.A.; Willander, M. Cobalt oxide magnetic nanoparticles–chitosan nanocomposite based electrochemical urea biosensor. Indian J. Phys. 2015, 89, 331–336. [Google Scholar] [CrossRef]
  381. Solanki, P.R.; Kaushik, A.; Ansari, A.A.; Sumana, G.; Malhotra, B.D. Zinc oxide-chitosan nanobiocomposite for urea sensor. Appl. Phys. Lett. 2008, 93, 163903. [Google Scholar] [CrossRef]
  382. Tiwari, A.; Aryal, S.; Pilla, S.; Gong, S. An amperometric urea biosensor based on covalently immobilized urease on an electrode made of hyperbranched polyester functionalized gold nanoparticles. Talanta 2009, 78, 1401–1407. [Google Scholar] [CrossRef] [PubMed]
  383. Migliorini, F.L.; Sanfelice, R.C.; Mercante, L.A.; Andre, R.S.; Mattoso, L.H.C.; Correa Daniel, S. Urea impedimetric biosensing using electrospun nanofibers modified with zinc oxide nanoparticles. Appl. Surf. Sci. 2018, 443, 18–23. [Google Scholar] [CrossRef]
  384. Baş, S.Z.; Gülce, H.; Yıldız, S.; Gülce, A. Amperometric biosensors based on deposition of gold and platinum nanoparticles on polyvinylferrocene modified electrode for xanthine detection. Talanta 2011, 87, 189–196. [Google Scholar] [CrossRef] [PubMed]
  385. Sahyar, B.Y.; Kaplan, M.; Ozsoz, M.; Celik, E.; Otles, S. Electrochemical xanthine detection by enzymatic method based on Ag doped ZnO nanoparticles by using polypyrrole. Bioelectrochemistry 2019, 130, 107327. [Google Scholar] [CrossRef] [PubMed]
  386. Devi, R.; Thakur, M.; Pundir, C.S. Construction and application of an amperometric xanthine biosensor based on zinc oxide nanoparticles–polypyrrole composite film. Biosens. Bioelectron. 2011, 26, 3420–3426. [Google Scholar] [CrossRef]
  387. Dervisevic, M.; Dervisevic, E.; Çevik, E.; Şenel, M. Novel electrochemical xanthine biosensor based on chitosan–polypyrrole–gold nanoparticles hybrid bio-nanocomposite platform. J. Food Drug Anal. 2017, 25, 510–519. [Google Scholar] [CrossRef]
  388. Sadeghi, S.; Fooladi, E.; Malekaneh, M. A nanocomposite/crude extract enzyme-based xanthine biosensor. Anal. Biochem. 2014, 464, 51–59. [Google Scholar] [CrossRef]
  389. Yadav, S.; Devi, R.; Bhar, P.; Singhla, S.; Pundir, C.S. Immobilization of creatininase, creatinase and sarcosine oxidase on iron oxide nanoparticles/chitosan-g-polyaniline modified Pt electrode for detection of creatinine. Enzym. Microb. Technol. 2012, 50, 247–254. [Google Scholar] [CrossRef] [PubMed]
  390. Zhybak, M.; Beni, V.; Vagin, M.Y.; Dempsey, E.; Turner, A.P.F.; Korpan, Y. Creatinine and urea biosensors based on a novel ammonium ion-selective copper-polyaniline nano-composite. Biosens. Bioelectron. 2016, 77, 505–511. [Google Scholar] [CrossRef] [PubMed]
  391. Bodade, A.B.; Taiwade, M.A.; Chaudhari, G.N. Bioelectrode based chitosan-nano copper oxide for application to lipase biosensor. J. Appl. Pharm. Res. 2017, 5, 30–39. [Google Scholar]
  392. Narang, J.; Pundir, C.S. Construction of a triglyceride amperometric biosensor based on chitosan–ZnO nanocomposite film. Int. J. Biol. Macromol. 2011, 49, 707–715. [Google Scholar] [CrossRef]
  393. Chauhan, N.; Pundir, C.S. Amperometric determination of acetylcholine—A neurotransmitter, by chitosan/gold-coated ferric oxide nanoparticles modified gold electrode. Biosens. Bioelectron. 2014, 61, 1–8. [Google Scholar] [CrossRef] [PubMed]
  394. Zhang, Z.; Wang, J.; Wang, X.; Wang, Y.; Yang, X. A sensitive choline biosensor with supramolecular architecture. Talanta 2010, 82, 483–487. [Google Scholar] [CrossRef]
  395. Srivastava, M.; Srivastava, S.K.; Nirala, N.R.; Prakash, R. A chitosan-based polyaniline–Au nanocomposite biosensor for determination of cholesterol. Anal. Methods 2014, 6, 817–824. [Google Scholar] [CrossRef]
  396. Kaushik, A.; Solanki, P.R.; Ansari, A.A.; Sumana, G.; Ahmad, S.; Malhotra, B.D. Iron oxide-chitosan nanobiocomposite for urea sensor. Sens. Actuators B Chem. 2009, 138, 572–580. [Google Scholar] [CrossRef]
  397. Da Silva, W.; Ghica, M.E.; Ajayi, R.F.; Iwuoha, E.I.; Brett, C.M.A. Tyrosinase based amperometric biosensor for determination of tyramine in fermented food and beverages with gold nanoparticle doped poly(8-anilino-1-naphthalene sulphonic acid) modified electrode. Food Chem. 2019, 282, 18–26. [Google Scholar] [CrossRef]
  398. Chawla, S.; Pundir, C.S. An amperometric hemoglobin A1c biosensor based on immobilization of fructosyl amino acid oxidase onto zinc oxide nanoparticles–polypyrrole film. Anal. Biochem. 2012, 430, 156–162. [Google Scholar] [CrossRef]
  399. Singh, K.; Chauhan, R.; Solanki, P.R.; Basu, T. Development of Impedimetric Biosensor for Total Cholesterol Estimation Based on Polypyrrole and Platinum Nanoparticle Multi Layer Nanocomposite. Int. J. Org. Chem. 2013, 3, 262–274. [Google Scholar] [CrossRef]
  400. Sahin, O.G.; Gulce, H.; Gulce, A. Polyvinylferrocenium based platinum electrodeposited amperometric biosensors for lysine detection. J. Electroanal. Chem. 2013, 690, 1–7. [Google Scholar] [CrossRef]
  401. Devi, R.; Pundir, C.S. Construction and application of an amperometric uric acid biosensor based on covalent immobilization of uricase on iron oxide nanoparticles/chitosan-g-polyaniline composite film electrodeposited on Pt electrode. Sens. Actuators B Chem. 2014, 193, 608–615. [Google Scholar] [CrossRef]
  402. Kucherenko, I.S.; Farre, C.; Raimondi, G.; Chaix, C.; Jaffrezic-Renault, N.; Chateaux, J.-F.; Sobolevskyi, M.; Soldatkin, O.O.; Dzyadevych, S.V.; Soldatkin, A.P.; et al. A novel adenosine triphosphate (ATP) biosensor based on electrospun polymer nanofibers with entrapped hexokinase and glucose oxidase. Appl. Nanosci. 2023. [Google Scholar] [CrossRef]
  403. Bucur, B.; Munteanu, F.-D.; Marty, J.-L.; Vasilescu, A. Advances in Enzyme-Based Biosensors for Pesticide Detection. Biosensors 2018, 8, 27. [Google Scholar] [CrossRef]
  404. Chauhan, N.; Narang, J.; Jain, U. Amperometric acetylcholinesterase biosensor for pesticides monitoring utilising iron oxide nanoparticles and poly(indole-5-carboxylic acid). J. Exp. Nanosci. 2016, 11, 111–122. [Google Scholar] [CrossRef]
  405. Kestwal, R.M.; Bagal-Kestwal, D.; Chiang, B.-H. Fenugreek hydrogel–agarose composite entrapped gold nanoparticles for acetylcholinesterase based biosensor for carbamates detection. Anal. Chim. Acta 2015, 886, 143–150. [Google Scholar] [CrossRef]
  406. Marinov, I.; Ivanov, Y.; Gabrovska, K.; Godjevargova, T. Amperometric acetylthiocholine sensor based on acetylcholinesterase immobilized on nanostructured polymer membrane containing gold nanoparticles. J. Mol. Catal. B Enzym. 2010, 62, 66–74. [Google Scholar] [CrossRef]
  407. Pundir, C.S.; Chauhan, N. Acetylcholinesterase inhibition-based biosensors for pesticide determination: A review. Anal. Biochem. 2012, 429, 19–31. [Google Scholar] [CrossRef]
  408. Rodrigues, N.F.M.; Neto, S.Y.; Luz, R.D.C.S.; Damos, F.S.; Yamanaka, H. Ultrasensitive Determination of Malathion Using Acetylcholinesterase Immobilized on Chitosan-Functionalized Magnetic Iron Nanoparticles. Biosensors 2018, 8, 16. [Google Scholar] [CrossRef]
  409. Jeyapragasam, T.; Saraswathi, R. Electrochemical biosensing of carbofuran based on acetylcholinesterase immobilized onto iron oxide–chitosan nanocomposite. Sens. Actuators B Chem. 2014, 191, 681–687. [Google Scholar] [CrossRef]
  410. Adeloju, S.B.; Hussain, S. Potentiometric sulfite biosensor based on entrapment of sulfite oxidase in a polypyrrole film on a platinum electrode modified with platinum nanoparticles. Microchim. Acta 2016, 183, 1341–1350. [Google Scholar] [CrossRef]
  411. Çevik, E.; Şenel, M.; Baykal, A.; Abasıyanık, M.F. A novel amperometric phenol biosensor based on immobilized HRP on poly(glycidylmethacrylate)-grafted iron oxide nanoparticles for the determination of phenol derivatives. Sens. Actuators B Chem. 2012, 173, 396–405. [Google Scholar] [CrossRef]
  412. Brondani, D.; de Souza, B.; Souza, B.S.; Neves, A.; Vieira, I.C. PEI-coated gold nanoparticles decorated with laccase: A new platform for direct electrochemistry of enzymes and biosensingapplications. Biosens. Bioelectron. 2013, 42, 242–247. [Google Scholar] [CrossRef] [PubMed]
  413. Campanhã Vicentini, F.; Garcia, L.L.C.; Figueiredo-Filho, L.C.S.; Janegitz, B.C.; Fatibello-Filho, O. A biosensor based on gold nanoparticles, dihexadecylphosphate, and tyrosinase for the determination of catechol in natural water. Enzym. Microb. Technol. 2016, 84, 17–23. [Google Scholar] [CrossRef] [PubMed]
  414. Mercante, L.A.; Iwaki, L.E.O.; Scagion, V.P.; Oliveira, O.N.; Mattoso, L.H.C.; Correa, D.S. Electrochemical Detection of Bisphenol A by Tyrosinase Immobilized on Electrospun Nanofibers Decorated with Gold Nanoparticles. Electrochem 2021, 2, 41–49. [Google Scholar] [CrossRef]
  415. Du, D.; Ding, J.; Cai, J.; Zhang, A. One-step electrochemically deposited interface of chitosan–gold nanoparticles for acetylcholinesterase biosensor design. J. Electroanal. Chem. 2007, 605, 53–60. [Google Scholar] [CrossRef]
  416. Li, G.; Nandgaonkar, A.G.; Wang, Q.; Zhang, J.; Krause, W.E.; Wei, Q.; Lucia, L.A. Laccase-immobilized bacterial cellulose/TiO2 functionalized composite membranes: Evaluation for photo- and bio-catalytic dye degradation. J. Membr. Sci. 2017, 525, 89–98. [Google Scholar] [CrossRef]
  417. Siddeeg, S.M.; Tahoon, M.A.; Mnif, W.; Ben Rebah, F. Iron Oxide/Chitosan Magnetic Nanocomposite Immobilized Manganese Peroxidase for Decolorization of Textile Wastewater. Processes 2020, 8, 5. [Google Scholar] [CrossRef]
  418. Bayramoğlu, G.; Arıca, M.Y. Enzymatic removal of phenol and p-chlorophenol in enzyme reactor: Horseradish peroxidase immobilized on magnetic beads. J. Hazard. Mater. 2008, 156, 148–155. [Google Scholar] [CrossRef]
  419. Le, T.T.; Murugesan, K.; Lee, C.-S.; Vu, C.H.; Chang, Y.-S.; Jeon, J.-R. Degradation of synthetic pollutants in real wastewater using laccase encapsulated in core–shell magnetic copper alginate beads. Bioresour. Technol. 2016, 216, 203–210. [Google Scholar] [CrossRef] [PubMed]
  420. Lin, C.-W.; Wu, C.-H.; Huang, W.-T.; Tsai, S.-L. Evaluation of different cell-immobilization strategies for simultaneous distillery wastewater treatment and electricity generation in microbial fuel cells. Fuel 2015, 144, 1–8. [Google Scholar] [CrossRef]
  421. Liu, Y.; Yan, M.; Geng, Y.; Huang, J. Laccase Immobilization on Poly(p-Phenylenediamine)/Fe3O4 Nanocomposite for Reactive Blue 19 Dye Removal. Appl. Sci. 2016, 6, 232. [Google Scholar] [CrossRef]
  422. Preethi, S.; Anumary, A.; Ashokkumar, M.; Thanikaivelan, P. Probing horseradish peroxidase catalyzed degradation of azo dye from tannery wastewater. SpringerPlus 2013, 2, 341. [Google Scholar] [CrossRef]
  423. Sondhi, S.; Kaur, R.; Kaur, S.; Kaur, P.S. Immobilization of laccase-ABTS system for the development of a continuous flow packed bed bioreactor for decolorization of textile effluent. Int. J. Biol. Macromol. 2018, 117, 1093–1100. [Google Scholar] [CrossRef]
  424. Shojaat, R.; Saadatjoo, N.; Karimi, A.; Aber, S. Simultaneous adsorption–degradation of organic dyes using MnFe2O4/calcium alginate nano-composites coupled with GOx and laccase. J. Environ. Chem. Eng. 2016, 4, 1722–1730. [Google Scholar] [CrossRef]
  425. Li, Z.; Zhu, Q.; Liu, Z.; Sha, L.; Chen, Z. Improved performance of immobilized laccase for catalytic degradation of synthetic dyes using redox mediators. New J. Chem. 2022, 46, 9792–9798. [Google Scholar] [CrossRef]
  426. Bapat, G.; Labade, C.; Chaudhari, A.; Zinjarde, S. Silica nanoparticle based techniques for extraction, detection, and degradation of pesticides. Adv. Colloid Interface Sci. 2016, 237, 1–14. [Google Scholar] [CrossRef]
  427. Feng, S.; Hao Ngo, H.; Guo, W.; Woong Chang, S.; Duc Nguyen, D.; Cheng, D.; Varjani, S.; Lei, Z.; Liu, Y. Roles and applications of enzymes for resistant pollutants removal in wastewater treatment. Bioresour. Technol. 2021, 335, 125278. [Google Scholar] [CrossRef]
  428. Bilal, M.; Iqbal, H.M.N.; Barceló, D. Persistence of pesticides-based contaminants in the environment and their effective degradation using laccase-assisted biocatalytic systems. Sci. Total Environ. 2019, 695, 133896. [Google Scholar] [CrossRef]
  429. Daumann, L.J.; Larrabee, J.A.; Ollis, D.; Schenk, G.; Gahan, L.R. Immobilization of the enzyme GpdQ on magnetite nanoparticles for organophosphate pesticide bioremediation. J. Inorg. Biochem. 2014, 131, 1–7. [Google Scholar] [CrossRef] [PubMed]
  430. Xiong, S.; Deng, Y.; Zhou, Y.; Gong, D.; Xu, Y.; Yang, L.; Chen, H.; Chen, L.; Song, T.; Luo, A.; et al. Current progress in biosensors for organophosphorus pesticides based on enzyme functionalized nanostructures: A review. Anal. Methods 2018, 10, 5468–5479. [Google Scholar] [CrossRef]
  431. Zdarta, J.; Jankowska, K.; Bachosz, K.; Degórska, O.; Kaźmierczak, K.; Nguyen, L.N.; Nghiem, L.D.; Jesionowski, T. Enhanced Wastewater Treatment by Immobilized Enzymes. Curr. Pollut. Rep. 2021, 7, 167–179. [Google Scholar] [CrossRef]
  432. Xu, R.; Yuan, J.; Si, Y.; Li, F.; Zhang, B. Estrone removal by horseradish peroxidase immobilized on a nanofibrous support with Fe 3 O 4 nanoparticles. RSC Adv. 2016, 6, 3927–3933. [Google Scholar] [CrossRef]
  433. Lassouane, F.; Aït-Amar, H.; Amrani, S.; Rodriguez-Couto, S. A promising laccase immobilization approach for Bisphenol A removal from aqueous solutions. Bioresour. Technol. 2019, 271, 360–367. [Google Scholar] [CrossRef]
  434. Joss, A.; Zabczynski, S.; Göbel, A.; Hoffmann, B.; Löffler, D.; McArdell, C.S.; Ternes, T.A.; Thomsen, A.; Siegrist, H. Biological degradation of pharmaceuticals in municipal wastewater treatment: Proposing a classification scheme. Water Res. 2006, 40, 1686–1696. [Google Scholar] [CrossRef]
  435. Siddeeg, S.M.; Amari, A.; Tahoon, M.A.; Alsaiari, N.S.; Rebah, F.B. Removal of meloxicam, piroxicam and Cd+2 by Fe3O4/SiO2/glycidyl methacrylate-S-SH nanocomposite loaded with laccase. Alex. Eng. J. 2020, 59, 905–914. [Google Scholar] [CrossRef]
  436. Sarkar, S.; Chakraborty, S.; Bhattacharjee, C. Photocatalytic degradation of pharmaceutical wastes by alginate supported TiO2 nanoparticles in packed bed photo reactor (PBPR). Ecotoxicol. Environ. Saf. 2015, 121, 263–270. [Google Scholar] [CrossRef]
  437. Adewuyi, A. Chemically Modified Biosorbents and Their Role in the Removal of Emerging Pharmaceutical Waste in the Water System. Water 2020, 12, 1551. [Google Scholar] [CrossRef]
  438. Shi, L.; Ma, F.; Han, Y.; Zhang, X.; Yu, H. Removal of sulfonamide antibiotics by oriented immobilized laccase on Fe3O4 nanoparticles with natural mediators. J. Hazard. Mater. 2014, 279, 203–211. [Google Scholar] [CrossRef]
  439. Okoye, C.O.; Okeke, E.S.; Okoye, K.C.; Echude, D.; Andong, F.A.; Chukwudozie, K.I.; Okoye, H.U.; Ezeonyejiaku, C.D. Occurrence and fate of pharmaceuticals, personal care products (PPCPs) and pesticides in African water systems: A need for timely intervention. Heliyon 2022, 8, e09143. [Google Scholar] [CrossRef] [PubMed]
  440. Ran, F.; Zou, Y.; Xu, Y.; Liu, X.; Zhang, H. Fe3O4@MoS2@PEI-facilitated enzyme tethering for efficient removal of persistent organic pollutants in water. Chem. Eng. J. 2019, 375, 121947. [Google Scholar] [CrossRef]
  441. Sharma, B.; Dangi, A.K.; Shukla, P. Contemporary enzyme based technologies for bioremediation: A review. J. Environ. Manag. 2018, 210, 10–22. [Google Scholar] [CrossRef]
  442. Xie, H.; Chen, Y.; Wang, C.; Shi, W.; Zuo, L.; Xu, H. The removal of fluoranthene by Agaricus bisporus immobilized in Ca-alginate modified by Lentinus edodes nanoparticles. RSC Adv. 2015, 5, 44812–44823. [Google Scholar] [CrossRef]
  443. Varga, B.; Somogyi, V.; Meiczinger, M.; Kováts, N.; Domokos, E. Enzymatic treatment and subsequent toxicity of organic micropollutants using oxidoreductases—A review. J. Clean. Prod. 2019, 221, 306–322. [Google Scholar] [CrossRef]
  444. Li, J.; Chen, X.; Xu, D.; Pan, K. Immobilization of horseradish peroxidase on electrospun magnetic nanofibers for phenol removal. Ecotoxicol. Environ. Saf. 2019, 170, 716–721. [Google Scholar] [CrossRef]
  445. Alver, E.; Metin, A.Ü. Chitosan based metal-chelated copolymer nanoparticles: Laccase immobilization and phenol degradation studies. Int. Biodeterior. Biodegrad. 2017, 125, 235–242. [Google Scholar] [CrossRef]
  446. Atacan, K.; Güy, N.; Çakar, S.; Özacar, M. Efficiency of glucose oxidase immobilized on tannin modified NiFe2O4 nanoparticles on decolorization of dye in the Fenton and photo-biocatalytic processes. J. Photochem. Photobiol. A Chem. 2019, 382, 111935. [Google Scholar] [CrossRef]
  447. Guo, J.; Liu, X.; Zhang, X.; Wu, J.; Chai, C.; Ma, D.; Chen, Q.; Xiang, D.; Ge, W. Immobilized lignin peroxidase on Fe3O4@SiO2@polydopamine nanoparticles for degradation of organic pollutants. Int. J. Biol. Macromol. 2019, 138, 433–440. [Google Scholar] [CrossRef]
  448. Zhang, K.; Yang, W.; Liu, Y.; Zhang, K.; Chen, Y.; Yin, X. Laccase immobilized on chitosan-coated Fe3O4 nanoparticles as reusable biocatalyst for degradation of chlorophenol. J. Mol. Struct. 2020, 1220, 128769. [Google Scholar] [CrossRef]
  449. Xia, T.-T.; Feng, M.; Liu, C.-L.; Liu, C.-Z.; Guo, C. Efficient phenol degradation by laccase immobilized on functional magnetic nanoparticles in fixed bed reactor under high-gradient magnetic field. Eng. Life Sci. 2021, 21, 374–381. [Google Scholar] [CrossRef] [PubMed]
  450. Morshed, M.N.; Behary, N.; Bouazizi, N.; Guan, J.; Nierstrasz, V.A. An overview on biocatalysts immobilization on textiles: Preparation, progress and application in wastewater treatment. Chemosphere 2021, 279, 130481. [Google Scholar] [CrossRef] [PubMed]
  451. Kołodziejczak-Radzimska, A.; Nghiem, L.D.; Jesionowski, T. Functionalized Materials as a Versatile Platform for Enzyme Immobilization in Wastewater Treatment. Curr. Pollut. Rep. 2021, 7, 263–276. [Google Scholar] [CrossRef]
  452. Jahangiri, E.; Thomas, I.; Schulze, A.; Seiwert, B.; Cabana, H.; Schlosser, D. Characterisation of electron beam irradiation-immobilised laccase for application in wastewater treatment. Sci. Total Environ. 2018, 624, 309–322. [Google Scholar] [CrossRef] [PubMed]
  453. Bilal, M.; Iqbal, H.M.N.; Barceló, D. Mitigation of bisphenol A using an array of laccase-based robust bio-catalytic cues—A review. Sci. Total Environ. 2019, 689, 160–177. [Google Scholar] [CrossRef] [PubMed]
  454. Ren, D.; Wang, Z.; Jiang, S.; Yu, H.; Zhang, S.; Zhang, X. Recent environmental applications of and development prospects for immobilized laccase: A review. Biotechnol. Genet. Eng. Rev. 2020, 36, 81–131. [Google Scholar] [CrossRef]
  455. Unuofin, J.O.; Okoh, A.I.; Nwodo, U.U. Aptitude of Oxidative Enzymes for Treatment of Wastewater Pollutants: A Laccase Perspective. Molecules 2019, 24, 2064. [Google Scholar] [CrossRef]
  456. Singh, J.; Saharan, V.; Kumar, S.; Gulati, P.; Kapoor, R. Laccase grafted membranes for advanced water filtration systems: A green approach to water purification technology. Crit. Rev. Biotechnol. 2017, 38, 883–901. [Google Scholar] [CrossRef]
  457. Daronch, N.A.; Kelbert, M.; Pereira, C.S.; de Araújo, P.H.H.; de Oliveira, D. Elucidating the choice for a precise matrix for laccase immobilization: A review. Chem. Eng. J. 2020, 397, 125506. [Google Scholar] [CrossRef]
  458. Deska, M.; Kończak, B. Immobilized fungal laccase as “green catalyst” for the decolourization process—State of the art. Process Biochem. 2019, 84, 112–123. [Google Scholar] [CrossRef]
  459. Okwara, P.C.; Afolabi, I.S.; Ahuekwe, E.F. Application of laccase in aflatoxin B1 degradation: A review. IOP Conf. Ser. Mater. Sci. Eng. 2021, 1107, 012178. [Google Scholar] [CrossRef]
  460. Daâssi, D.; Rodríguez-Couto, S.; Nasri, M.; Mechichi, T. Biodegradation of textile dyes by immobilized laccase from Coriolopsis gallica into Ca-alginate beads. Int. Biodeterior. Biodegrad. 2014, 90, 71–78. [Google Scholar] [CrossRef]
  461. Hou, J.; Dong, G.; Ye, Y.; Chen, V. Enzymatic degradation of bisphenol-A with immobilized laccase on TiO2 sol–gel coated PVDF membrane. J. Membr. Sci. 2014, 469, 19–30. [Google Scholar] [CrossRef]
  462. Nadaroglu, H.; Mosber, G.; Gungor, A.A.; Adıguzel, G.; Adiguzel, A. Biodegradation of some azo dyes from wastewater with laccase from Weissella viridescens LB37 immobilized on magnetic chitosan nanoparticles. J. Water Process Eng. 2019, 31, 100866. [Google Scholar] [CrossRef]
  463. Qiu, X.; Wang, S.; Miao, S.; Suo, H.; Xu, H.; Hu, Y. Co-immobilization of laccase and ABTS onto amino-functionalized ionic liquid-modified magnetic chitosan nanoparticles for pollutants removal. J. Hazard. Mater. 2021, 401, 123353. [Google Scholar] [CrossRef]
  464. Khakshoor, M.; Makhdoumi, A.; Asoodeh, A.; Hosseindokht, M.R. Co-immobilized spore laccase/TiO2 nanoparticles in the alginate beads enhance dye removal by two-step decolorization. Environ. Sci. Pollut. Res. 2021, 28, 6099–6110. [Google Scholar] [CrossRef]
  465. Tišma, M.; Šalić, A.; Planinić, M.; Zelić, B.; Potočnik, M.; Šelo, G.; Bucić-Kojić, A. Production, characterisation and immobilization of laccase for an efficient aniline-based dye decolourization. J. Water Process Eng. 2020, 36, 101327. [Google Scholar] [CrossRef]
  466. Velu, C.; Veeramani, E.; Suntharam, S.; Kalimuthu, K. Insilico Screening and Comparative Study on the Effectiveness of Textile Dye Decolourization by Crude Laccase Immobilised Alginate Encapsulated Beads from Pleurotus ostreatus. J. Bioprocess. Biotech. 2011, 1, 4. [Google Scholar] [CrossRef]
  467. Noreen, S.; Asgher, M.; Hussain, F.; Iqbal, A. Performance Improvement of Ca-Alginate Bead Cross- Linked Laccase from Trametes versicolor IBL-04. BioResources 2016, 11, 558–572. [Google Scholar] [CrossRef]
  468. Faraco, V.; Pezzella, C.; Miele, A.; Giardina, P.; Sannia, G. Bio-remediation of colored industrial wastewaters by the white-rot fungi Phanerochaete chrysosporium and Pleurotus ostreatus and their enzymes. Biodegradation 2009, 20, 209–220. [Google Scholar] [CrossRef]
  469. Bayramoglu, G.; Yilmaz, M.; Yakup Arica, M. Preparation and characterization of epoxy-functionalized magnetic chitosan beads: Laccase immobilized for degradation of reactive dyes. Bioprocess Biosyst. Eng. 2010, 33, 439–448. [Google Scholar] [CrossRef]
  470. Bayramoğlu, G.; Yilmaz, M.; Yakup Arica, M. Reversible immobilization of laccase to poly(4-vinylpyridine) grafted and Cu(II) chelated magnetic beads: Biodegradation of reactive dyes. Bioresour. Technol. 2010, 101, 6615–6621. [Google Scholar] [CrossRef]
  471. Niladevi, K.N.; Prema, P. Immobilization of laccase from Streptomyces psammoticus and its application in phenol removal using packed bed reactor. World J. Microbiol. Biotechnol. 2008, 24, 1215–1222. [Google Scholar] [CrossRef]
  472. Zhu, Y.; Qiu, F.; Rong, J.; Zhang, T.; Mao, K.; Yang, D. Covalent laccase immobilization on the surface of poly(vinylidene fluoride) polymer membrane for enhanced biocatalytic removal of dyes pollutants from aqueous environment. Colloids Surf. B Biointerfaces 2020, 191, 111025. [Google Scholar] [CrossRef] [PubMed]
  473. Kumar, V.; Misra, N.; Kumar Goel, N.; Thakar, R.; Gupta, J.; Varshney, L. A horseradish peroxidase immobilized radiation grafted polymer matrix: A biocatalytic system for dye waste water treatment. RSC Adv. 2016, 6, 2974–2981. [Google Scholar] [CrossRef]
  474. Wang, S.; Fang, H.; Wen, Y.; Cai, M.; Liu, W.; He, S.; Xu, X. Applications of HRP-immobilized catalytic beads to the removal of 2,4-dichlorophenol from wastewater. RSC Adv. 2015, 5, 57286–57292. [Google Scholar] [CrossRef]
  475. Xu, J.; Tang, T.; Zhang, K.; Ai, S.; Du, H. Electroenzymatic catalyzed oxidation of bisphenol-A using HRP immobilized on magnetic silk fibroin nanoparticles. Process Biochem. 2011, 46, 1160–1165. [Google Scholar] [CrossRef]
  476. Zhai, R.; Zhang, B.; Wan, Y.; Li, C.; Wang, J.; Liu, J. Chitosan–halloysite hybrid-nanotubes: Horseradish peroxidase immobilization and applications in phenol removal. Chem. Eng. J. 2013, 214, 304–309. [Google Scholar] [CrossRef]
  477. Abdulaal, W.H.; Almulaiky, Y.Q.; El-Shishtawy, R.M. Encapsulation of HRP Enzyme onto a Magnetic Fe3O4 Np–PMMA Film via Casting with Sustainable Biocatalytic Activity. Catalysts 2020, 10, 181. [Google Scholar] [CrossRef]
  478. Peng, Q.; Liu, Y.; Zeng, G.; Xu, W.; Yang, C.; Zhang, J. Biosorption of copper(II) by immobilizing Saccharomyces cerevisiae on the surface of chitosan-coated magnetic nanoparticles from aqueous solution. J. Hazard. Mater. 2010, 177, 676–682. [Google Scholar] [CrossRef] [PubMed]
Figure 1. A schematic representation of polymer grafting approaches. Adopted with permission from Wang et al., 2020 [291].
Figure 1. A schematic representation of polymer grafting approaches. Adopted with permission from Wang et al., 2020 [291].
Polymers 15 03492 g001
Figure 2. A schematic representation of the setup and procedure for electrospinning. Reprinted with permission from Long et al., 2019 [321].
Figure 2. A schematic representation of the setup and procedure for electrospinning. Reprinted with permission from Long et al., 2019 [321].
Polymers 15 03492 g002
Figure 3. A schematic representation of a laccase-based enzymatic biosensor for detection of dopamine in pharmaceutical samples. Laccase was immobilized on AuNPs/poly(allylamine hydrochloride) nanocomposite to facilitate electron transfer between electrode and enzyme. Reprinted with permission from Silva and Vieira (2016) [328].
Figure 3. A schematic representation of a laccase-based enzymatic biosensor for detection of dopamine in pharmaceutical samples. Laccase was immobilized on AuNPs/poly(allylamine hydrochloride) nanocomposite to facilitate electron transfer between electrode and enzyme. Reprinted with permission from Silva and Vieira (2016) [328].
Polymers 15 03492 g003
Figure 5. Different pollutants that have been degraded by enzyme-nanoparticle-polymer composites. A—Laccase, B—Horse radish peroxidase, C—Lignin peroxidase, D—Chloroperoxidase, E—Glucose oxidase, F—Glucose oxidase/laccase, G—S. cerevisiae enzyme, H—Glycerophosphodiesterase, I—Manganese peroxidase, * 0–6 h, # 6–24 h, ɸ over 24 h.
Figure 5. Different pollutants that have been degraded by enzyme-nanoparticle-polymer composites. A—Laccase, B—Horse radish peroxidase, C—Lignin peroxidase, D—Chloroperoxidase, E—Glucose oxidase, F—Glucose oxidase/laccase, G—S. cerevisiae enzyme, H—Glycerophosphodiesterase, I—Manganese peroxidase, * 0–6 h, # 6–24 h, ɸ over 24 h.
Polymers 15 03492 g005
Table 1. Application of enzyme-nanoparticle-polymer composites in fabrication of biosensors for detection of hydrogen peroxide.
Table 1. Application of enzyme-nanoparticle-polymer composites in fabrication of biosensors for detection of hydrogen peroxide.
Nano-
Composite (NC)
Immobilization
Method
Electrode UsedDetection Range (μM)Limit of Detection (LOD) (μM)Response Time (s)Reusability and Storage Stability *Ref.
AuNPs/chitosan (CS) hydrogelSelf-assembled monolayer deposition of chitosan, AuNPs and HRPGold8.0–120
500–12,000
2.4 85% (4 weeks), 94% (50 uses)[250]
AuNPs/sodium alginateSelf-assembled monolayer deposition of sodium alginate and HRP-AuNPs solutionGold20–13,70031597% (1 month)[340]
AuNPs/carboxymethyl chitosanDrop-casting a mixture of HRP and AuNPs-CMCS NCGlassy carbon electrode (GCE)5–14000.1045RSD of 1.87% (6 cycles) and 94% (20 days)[341]
AuNPs/bacteria cellulose (BC) nanofibersSelf-assembled monolayer deposition of AuNPs/BC and HRPGCE 1420 [338]
AuNPs/polyaniline (PANI) nanofibersDrop-casting of homogenous mixture of HRP and AuNPs/PANI solutionsGCE10–20001.6595% (15 days) and 85% (1 month)[188]
AgNPs/polyvinyl butyral (PVB)Grafting-to in presence of HRPPlatinum10–10,00021085% (2 weeks), 93.3% (200 cycles) and RSD of 2.1% (9 cycles)[195]
AgNPs/poly(3,4-ethylenedioxythiophene):poly(styrene sulfonate) (PEDOT:PSS)/nafionSelf-assembled monolayer electrodeposition of PEDOT:PSS solution, AgNPs HRP and nafionGCE0.05–200.02 RSD of 0.45% (30 cycles) and 93% (2 weeks)[333]
ZnO-AuNPs/nafionGrafting-to in presence of HRPGCE15–11009 RSD of 2.6% (5 cycles)[196]
AuNPs/chitosanElectrodepositionIndium titanium oxide (ITO)10–5005 [342]
CeO2/PANIElectrodeposition of a mixture of aniline and CeO2 solutionITO50,000–500,00050,000 100%(8 weeks)[334]
AuNPs/chitosanSelf-assembled deposition of chitosan, gold colloid, and HRPITO20–65003.5590% (2 weeks), RSD of 1.7% (5 cycles)[339]
AuNPs/poly (m-phenylenediamine)Sequential grafting-to followed by addition of HRPCarbon paste0.13–1400.1330100% (14 days), 90% (25 days) and 67% (35 days)[186]
γ-Al2O3NPs/chitosanDrop-casting a mixture of HRP, colloid, and chitosan solutionGCE0.5–7000.07590% (1 month)[336]
AgNPs/Poly(γ-glutamic acid) (PGA)Co-assembled HRP with PGA to form colloidal NPs with photo-crosslinking followed by simultaneous electrodeposition with colloidal AgNPsGCE1–500
500–3000
0.351091.43% (30 days) and RSD of 31.4% (3 assays)[25]
* For storage stability, the biocatalysts were stored at 4 °C, RSD—relative standard deviation.
Table 2. Application of enzyme-nanoparticle-polymer composites in fabrication of biosensors for detection of glucose.
Table 2. Application of enzyme-nanoparticle-polymer composites in fabrication of biosensors for detection of glucose.
Nanocomposite (NC)Immobilization MethodElectrode UsedDetection Range (μM)Limit of Detection (μM)Response Time (s)ReusabilityRef.
AuNPs/chitosan hydrogelElectrodepositionGold5.0–24002.77RSD of 3.3% (7 assays), 4.6% (5 sensors), 75% (5 weeks)[85]
AuNPs/MPSSelf-assemblyGold0.00004–0.05280.0001 86.5% (21 days), 92.6% (160 assays)[308]
AuNW/chitosanDrop-castingGCE10–20,0005885% and RSD of 5.1% (1 month)[311]
AuNPs/PAMAM/PVSLbL self-assemblyITO17–15004 [306]
AuNPs/chitosanElectrodepositionGCE50–13001310RSD of 3.3% (10 assays)[165]
AuNPs/NafionDrop-castingGCE34–600034 90% (2 weeks) and RSD of 2.5% (eight assays), 4.5% (5 sensors)[197]
AuNPs/bacteria cellulose nanofibers/PDDASelf-assemblyGCE10–4002.3 90% and RSD of 1.6% (1 week)[357]
AuNPs/chitosanLbL self-assemblyPlatinum500–16,0007890% (1 month), RSD of 3.7% (7 assays) and 5.7% (5 biosensors)[363]
AuNPs/chitosan/prussian blue (PB)ElectrodepositionGCE1–16000.69370% (2 weeks), RSD of 1.1% (5 assays) and 8.3% (10 biosensors)[356]
AuNPs/poly(BEDOA-6)Electrodeposition of poly(BEDOA-6), covalent immobilization of GOx AuNPs, drop-casting of GOx-AuNPs on polymer-modified electrodeGraphite25–125025 100% (3 weeks with daily use)[312]
Nafion/AuNPs/PVP/PANIGrafting-from a mixture of AuNPs, aniline, and PVPGCE50–225010 89.9% (2 weeks), RSD of 3.9% (10 assays) and 5.8% (10 biosensors)[365]
MNP/chitosanGrafting-to of chitosan on MNPsLuminol–H2O2–gold nanoparticle chemiluminescence detection system0.85–1000.4 70% (8 weeks) and 96%, 89%, 81%, and 77% (5, 10, 20, and 25 assays, respectively)[361]
AuNPs/polypyrrole (PPy)Grafting-from of PPy on AuNPsGraphite rod200–19,9002005RSD of 9% (3 assays)[192]
AuNPs/chitosan-PPy nanotubesDrop-casting of PPy-AuNPs composite, and incubation in GOx solutionITO3–2303.1 [367]
AuNPs/Electrospun poly(vinyl alcohol) (PVA)/PEIElectrospinning of GOx, PVA, and PEI, and immersion in AuNPs solutionGold10–2000.9 86.5% (3 weeks) and RSD of <4% (3 assays)[362]
AuNPs/ACGDrop-castingPlatinum0.2–2 *0.060.1 [309]
AuNPs/chitosan/PB-chitosanSelf-assembled electrodeposition of PB-chitosan NC, AuNPs-chitosan, and bi-enzyme mixtureGold6.25–93.751.5610 [327]
AuNPs/PEIDrop-casting of AuNPs/PEI solution on electrode and immersion in enzymeGold1–1000.33594.7% (300 assays), 95% (24 h) and RSD of 4.46% (6 assays)[187]
AuNPs/polytriamineSelf-assembled deposition of tryamine, AuNPs solution, and GOxPB-modified screen-printed carbon1–10001 99.7% and 90% (374 and 411 assays respectively) and 99% and 84% (3 and 4 weeks, respectively)[344]
AgNPs/chitosanImmersion of electrode in a mixture of AgNPs, GOx, and chitosan solutionPlatinum1–80000.55 [159]
AgNPs/guar-gum (GG)/chitosanElectrodeposition of a mixture of silver nitrate, chitosan, GG, and enzyme solutionPhotometric flow injection system analysis1.4–6.9 *0.0003 70% (160 measurements) and 60% (140 days)[368]
0.4–2 *0.0002
AgNPs/poly(m-aminophenol)Drop-casting of a mixture of AgNPs/polymer, GOx, and nafionGCE2000–12,000100397.5% and 87.2% (2 days and 1 week respectively) and RSD of 3.8% (5 assays)[343]
MNPs/chitosanDrop-casting a of mixture of GOx and NC solution on electrodeITO600–22,200 580% (8 weeks)[157]
MNPs/PVADrop-casting a mixture of MNPs, PVA, and GOx on electrodeTin5000–30,00081081% (1 month) and RSD of 4.2% (5 biosensors)[347]
ZnONPs/chitosan-graft-PVASpin-casting a mixture of ZnONPs, chitosan, and PVA, dropped GOx solution on modified electrodeITO2–12002 [355]
ZnO nanorods/polydopamineSelf-assembled deposition of ZnO nanorods, dopamine, and GOxITO15–1206.2 [353]
ZnONPs/chitosanDrop-casting a mixture of ZnO-chitosan on electrode surface ad immersion in enzymePt-Fe(III)/Pt10–11,00011087% (2 weeks) and RSD of 2.8% and 4.1% (10 assays and 7 biosensors, respectively)[359]
ZrO2NPs/chitosanDrop-casting a mixture of GOx and ZrO2NPs/chitosan solutionPlatinum12.5–9500101096.2%, 75.2%, and 60.4% (20, 30, and 40 days, respectively) and RSD of 2.3% and 4.65% (6 assays and 4 biosensors)[369]
PdNPs/PEDOTSequential deposition of PEDOT, PdCl2, and finally GOxITO500–30,00075 75% (12 days) and RSD of 8.5% and 1.85% (6 biosensors and 7 assays, respectively)[360]
AuNPs/PPySequential electrodeposition of HAuCl4, enzyme, and pyrroleGCE2.5–500021060% (2 weeks), 25%, and 68% (25 and 100 assays) and RSD of 1.36% (6 biosensors)[193]
MNPs/nafionSequential drop-casting a mixture of MNPs and GOx, and nafionITO1000–80000.5 [198]
TiO2NPs/celluloseElectrospinning of a mixture of TiO2NPs and cellulose solution, immersion in GOx solutionGlass1000–10,000 [352]
CuONPs/chitosanMagnetic sputtering of CuO on FTO, drop-casting of mixture of GOx and chitosan on modified electrodeFTO200–15,00027487.5% (35 days) and RSD of 1.7% (5 biosensors in real blood serum)[349]
PtNPs/poly(amidoamine)Layer by layer electrodeposition of NC, GOx, and NCPlatinum5–10000.1580% and 86% (30 days and 100 assays, respectively)[166]
ZnO-PtNPs/chitosanSequential drop-casting of ZnONPs, PtNPs, chitosan solution, and enzymeFTO16.6–12216.6 [345]
PtNPs/PPySequential electropolymerization of pyrrole and PtNPs, immersion in GOx solutionAnodized aluminium oxide on a gold disk100–900027.77 [194]
PtNPs/chitosanElectrodeposition of H2PtCl6, CS, and enzymeGCE1.2–40,0000.4593.1% and 89.6% (3 weeks and 1 month, respectively) and RSD of 5.8% (5 biosensors)[351]
PtNPs/PANIDrop-coating of PANI hydrogel, immersion in H2PtCl6 and enzymePlatinum10–80000.73 [354]
PtNPs/PPy/poly(o-aminophenol) (POAP)Sequential electropolymerization of pyrrole, K2PtCl6, and a mixture of OAP and GOxGCE1.5–13,0000.457100%, 89%, and 76% (7, 30, and 60 days, respectively)[350]
MNPs/chitosan/nafionDrop-casting of mixture of GOx and MNPs, immersion in mixture of chitosan and MNPs, drop-casting of nafion solutionPlatinum6–22006 84% and 83% (1 month and 52 assays, respectively)[370]
PtNPs/PDDA/PANI/(PSS)Interfacial polymerization of PANI followed by doping with PSS, absorption of PtNPs/PDDA on PANI/PSS, immersion in GOx solutionGCE10–45000.5585% (20 days) and RSD of 4.4% (5 assays)[346]
AuNPs/PANISequential drop-casting of NC solution and enzymeGCE1–8000.5595% (2 weeks) and RSD of 4.8% (7 biosensors)[189]
NiFe2O4NPs/chitosanDrop-casting a mixture of NPs, chitosan, and GOx solutionGCE100–20,000100490% (30 days)[358]
PtNPs/chitosan/nafionImmersion in a mixture of PtNPs, chitosan, and GOx followed by nafion solutionGCE1–50000.5 90% (20 days) and RSD of 3% (10 assays)[366]
Au@Ag-PtNPs/infinite coordination polymer (ICP)Drop-casting a mixture of GOx and NCPlatinum0.5–33300.06 90% (14 weeks) and RSD of 3.8% and 4.9% (6 assays and 6 biosensors, respectively)[364]
CuONPs/alginateDrop-casting a mixture of CuONPs, alginate, GOx, and EDC/NHSPlatinum40–30004000–35,0001.6 98%, 97%, 93%, 86%, and 78% (2, 4, 6, 10, and 15 days, respectively) and RSD of 0.94% (6 biosensors)[371]
TiO2NPs/PANI/chitosanGrafting-from of aniline vapour on TiO2 and immersion in enzymeGCE20–1405.33 [372]
AuNPs/PANIGrafting-from of AuNPs, aniline, and GOxGraphite70–16,500 *701067.4% (8 days) and RSD of 4.67% (4 assays)[373]
AuNPs/PPyGrafting-from of AuNPs, pyrrole, and GOxGraphite71–16,500 *711071.4% (8 days) and RSD of 5.89% (4 assays)[373]
* The values have been converted from mM, MPS—(3-mercaptopropyl)-trimethoxysilane, ACG—agarose–corn flour–gelatin, PEDOT—poly (3,4-ethylenedioxythiophene) polystyrene, PPy—polypyrrole, PANI—polyaniline.
Table 3. Application of enzyme-nanoparticle-polymer composites in fabrication of biosensors for detection of compounds in human blood.
Table 3. Application of enzyme-nanoparticle-polymer composites in fabrication of biosensors for detection of compounds in human blood.
Nanocomposite (NC)EnzymeImmobilization MethodElectrode UsedDetected CompoundDetection Range (μM)Limit of Detection (μM)Response Time (s)ReusabilityRef.
AuNPs/
chitosan
Cholesterol oxidaseDrop-casting of chitosan nanofibers, electrodeposition of AuNPs, adsorption of enzymeITOCholesterol1–450.5591% (25 days) and RSD of 4.2% (10 assays)[375]
AuNPs/
chitosan/
PPy
Xanthine oxidaseDrop-casting of chitosan-PPy and HAuCl4 mixture, immersion in glutaraldehyde followed by enzyme solutionGCEXanthine1–2000.25892%, 85%, and 78% (13 days, 18 days, and 20 assays, respectively)[387]
AuNPs/poly(allylamine hydrochloride) (PAH)LaccaseMixing laccase solution with AuNPs/PAH composite and graphiteCarbon pasteDopamine0.49 - 23.00.26 [328]
AuNPs/poly(8-anilino-1-naphthalene sulphonic acid) (PANSA)TyrosinaseElectropolymerization of a mixture of AuNPs and ANSA solution, drop-casting of tyrosinase on modified electrodeGCETyramine10–120 0.71 97.5% and 94.3% (20 assays and 20 days, respectively). RSD of 4.3% (3 biosensors)[397]
ZnONPs/
chitosan
Cholesterol oxidaseMixed ZnONPs with chitosan solution and drop-casting mixture on electrode surface, drop-casted enzyme solutionITOCholesterol129.3–7759 *129.3 *1585% and 75% (6 days and 8 weeks, respectively)[377]
CeO2NPs/
chitosan
Cholesterol oxidaseMixed CeO2NPs with chitosan solution and drop-casted mixture on electrode surface followed by enzymeITOCholesterol25–10,345 *129.3 *10100% and 90% (10 assays and 7 weeks, respectively)[376]
SnO2NPs/
chitosan
Cholesterol oxidaseMixed SnO2NPs with chitosan solution and drop-casted mixture on electrode surface followed by enzymeITOCholesterol25–10,345 *129.3 *595% and 90% (6 and 8 weeks, respectively)[374]
MNPs/
chitosan-graft-PANI
Creatininase,
Creatinase
Sarcosine oxidase
Electrodeposition of a mixture of aniline, MNPs, and chitosan solution in HCl, dropping of GA followed by enzyme mixture on electrodePlatinumCreatinine in serum of healthy people1–8001290% (120 assays over 200 days)[389]
Fe@AuNPs/chitosanAChE
Choline oxidase
Electrodeposition of Fe@AuNPs by CV, immersion in chitosan solution, immersion of modified electrode in GA followed by enzyme mixtureGoldAcetyl choline0.005–4000.005350% (100 assays)[393]
ZnONPs/
PPy
Fructosyl amino acid oxidase (FAO)Electropolymerization of PPy followed by ZnONPs by CVGoldHemoglobin A1c100–300050270% (260 assays)[398]
ZnONPs/
PPy
Xanthine oxidaseElectropolymerization of a mixture of PPy and ZnONPs, immersion in enzyme solutionPlatinumXanthine0.8–400.8560% (200 assays in 100 days)[386]
ZnONPs/
chitosan
LipaseGlycerol kinase
Glycerol-3-phosphate oxidase
Immersion in ZnONPs-chitosan mixture, activated with GA, immersion in a mixture of enzyme solutionsPlatinumTriglyceride2839–36,906.7 *1135.6 *675% and 50% (6 and 7 months, respectively)[392]
Ag-ZnONPs/
PPy
Xanthine oxidaseElectropolymerization of PPy followed by Ag-ZnONPs by CV, electrodeposition of enzyme under open circuitPencil graphite electrode (PGE)Xanthine0.06–0.6 0.07 3.277.82% and 77% (20 days and 20 assays, respectively)[385]
AuNPs/
PANI/
chitosan
Cholesterol oxidaseSpin-coated a mixture of chitosan and Au-PANI solution, drop-casted enzyme on modified electrodeITOCholesterol1293–12,931.6 *980 *20100%, 97%, and 90% (20 assays, 2 and 3 weeks, respectively)[395]
NiFe2O4-CuO-FeONPs/chitosanCholesterol oxidaseDrop-casted a mixture of NiFe2O4-CuO-FeONPs and chitosan solutionsITOCholesterol129–12,931.6 *809.5 *1086% and 75% (10 and 90 days, respectively)[378]
PtNPs/PPyCholesterol esterase
Cholesterol oxidase
Electropolymerization of pyrrole, immersion in hexa chloroplatinic acid followed by pyrrole solutionITOCholesterol250–65002502590% (7 weeks)[399]
Au-PTNPs/polyvinylferrocene(PVF)PtNPs/PVFXanthine oxidaseSequential immersion in PVF, KAuCl4 (for Au-PtNPs), PtBr2PlatinumXanthine2–660.6 90% (10 days) and RSD of 3.41% (5 biosensors)[384]
PtNPs/poly (thiolated β–cyclodextrin)HRP
Choline oxidase
Sequential immersion of electrode in mixture of polymer and NPs solution and enzyme mixtureGoldCholine0.001–10,0000.0001 85% (1 month) and RSD of 4.6% (10 assays)[394]
MNPs/PANI/chitosanXanthine oxidaseDispersion of MNPs in aniline, mixed carbon paste, NC, chitosan, and H2PtCl6Carbon pasteXanthine0.2–36.00.1885% (100 uses over 3 months), RSD of 4% (5 assays)[388]
PtNPs/PVFLysine oxidaseElectro-oxidation of PVF, electrodeposition of H2PtCl6, immersion in enzyme solutionPlatinumLysine0.65–30000.653090% and 85% (1 month and 15 assays, respectively)[400]
CuONPs/chitosanLipaseSpin-coating of a mixture of chitosan and CuONPs solution, immersion on lipase enzyme solutionGoldTriglyceride1419.5–17,033.8 *152 [391]
CuONPs/PANI/nafionCreatinine deaminaseCV of copper nitrate, drop-casting of nafion solution, electropolymerization of aniline, drop-casting of enzyme solutionScreen-printed electrodeCreatinine8–900.515 [390]
AuNPs/BoltornUreasePolymer graftingITOUrea10–35,000103100% (10 uses) and RSD of 8% and 6% (5 assays and 10 biosensors, respectively)[382]
MNPs/chitosanUrease
Glutamate dehydrogenase
Dispersion MNPs in chitosan solution, drop mixture of enzymes and NC on electrode surfaceITOUrea833.3–16,666.7 *83.3 *1085% and 75% (8 and 10 weeks, respectively)[396]
MNPs/chitosanureaseDrop-casting of mixture of MNPs and chitosan solutionCopper wireUrea100–80,000 1290% (3 weeks)[379]
ZnONPs/PPy/polyamide 6 (PA6)UreaseElectrospinning of PPy and PA6 on fluorine-modified electrode, immersion in ZnO solution followed by urease solutionTin oxideUrea16.7–41,666.71.83 97% and 80% (2 and 4 weeks, respectively), RSD of 4.4% and 4.5% (8 assays and 3 biosensors, respectively)[383]
ZnONPs/chitosanUrease
Glutamate dehydrogenase
Spin-coating of a mixture of ZnONPs and chitosan, physical adsorption of enzyme on the modified electrodeITOUrea833.3–16,666.7 *500 *10 [381]
MNPs/chitosan-graft-PANIUricaseElectropolymerization of a mixture of aniline, MNPs, and chitosan, immersed modified electrode in GA followed by uricase solutionPlatinumUric acid0.1–8000.1190% (120 assays over 100 days)[401]
Co3O4NPs/chitosanUraseDrop-casted a mixture of NC and chitosan solution, immersion in urase solutionCopper wireUrea100–80,000 1285% (1 month)[380]
CuONPs/PANI/nafionUreaseCV of copper nitrate, drop-casting of nafion solution, electropolymerization of aniline, drop-casting of enzyme solutionScreen-printed electrodeUrea5–500.515 [390]
AuNPs/polyvinyl alcohol (PVA)GOx and hexokinaseElectrospinning of a mixture of enzymes, polymers, and AuNPsPlatinumAdenosine triphosphate (ATP)25–2002515RSD of 3.4% (9 assays) and 86% (10 cycles)[402]
* The values were converted from mM to μM.
Table 5. Application of enzyme-nanoparticle-polymer composites in degradation of organic pollutants for application in wastewater treatment.
Table 5. Application of enzyme-nanoparticle-polymer composites in degradation of organic pollutants for application in wastewater treatment.
Nanocomposite (NC)Immobilization MethodPollutants RemovedDegradation (%)Degradation TimeReusabilityRef.
TiO2/polyvinylidene fluoride (PVDF)Crosslinking of TiO2/PVDF membrane using APTES and glutaraldehyde followed by immersion in laccase solutionBisphenol A955 h91.7% (96 h of continuous use)[461]
TiO2/bacterial cellulose (BC)Physical adsorption of TiO2 on BC followed by crosslinking with glutaraldehyde and immersion in laccase solutionReactive red X-3B in presence of ABTS8060 min70% and 57% (6 and 10 cycles, respectively)[416]
Calcium alginatePhysical entrapment of enzyme in nanocompositeFluoranthene in a fluidized bed reactor81.068 h66.845% (60 days of storage)[442]
Fe2O3/poly(ethylene glycol)/concovalin AChemical co-precipitation followed by crosslinking with glutaraldehyde and immersion in laccase solutionSulfadiazine10030 min82.8% (10 consecutive cycles)[438]
Sulfamethazine
Sulfamethoxazole
(all in presence of syringaldehyde mediator)
MNPs/chitosanPhysical mixing of NPs and chitosan followed by crosslinking with glutaraldehyde and immersion in laccase solutionReactive black 59030 min47% (10 cycles)[462]
Evans blue6030 min
Tryphan blue8040 min
Direct blue 157060 min
MNPs/polydopamineFunctionalized MNP-polydopamine NC with dialdehyde starch followed by immersion in laccase solution2,4-dichlorophenol723 h77% (8 cycles)[191]
9112 h
Fe2O3/Cu-alginatePhysical entrapment of enzyme in nanocompositeTriclosan89.68 h86.9% (3 cycles in acetate buffer)[419]
53.28 h (wastewater)
Remazol Brilliant Blue R (RBBR)75.88 h
5525 h (wastewater)
3525 h (waste water)
Cu (II)-chitosan-graft-poly (glycidyl methacrylate)/poly (ethylene imine)Physical adsorption of laccase on nanocompositesPhenol in presence of ABTS804 h50% (8 cycles)[445]
MNPs/chitosanCrosslinking with glutaraldehyde followed by immersion in laccase solution2,4-Dichlorophenol91.412 h75.8% and 57.4% (2,4-DCP and 4-CP after 10 cycles)[448]
4-Chlorophenol75.5
MNPs/SiO2/poly (glycidyl methacrylate)-S-SHPhysical adsorption of enzyme on the nanocompositeMeloxicam9248 h82.3%, 88.9%, and 87.5% (meloxicam, piroxicam and Cd2+, respectively, after 5 cycles)[435]
Piroxicam95
Cd2+94
MNPs/Poly(p-Phenylenediamine)Covalent immobilization using glutaraldehyde for crosslinkingReactive blue 19801 h43% (8 cycles)[421]
MNPs@MoS2/polyethyleneiminePhysical adsorption of laccase on nanocompositeMalachite green82.7Overnight62% (10 cycles)[440]
Bisphenol A87.6
Bisphenol F
(all in presence of ABTS)
70.6
Cu-alginatePhysical entrapment of enzyme in nanocompositeFuschin blue65 (HOBT)4 h100% and 95% (120 h continuous use and 15 days storage, respectively)[423]
Congo red27 (ABTS)
Tryphan blue51(syringaldehyde)
Malachite green60 (ABTS)
Erichrome black T50 (HOBT)
Crystal violet
(all in different mediators)
32 (HOBT)
Textile effluent in a continuous flow packed bed bioreactor66 (colour)
90 (BOD)
98 (COD)
MNPs/chitosanPhysical entrapment of enzyme in presence of ionic liquid and ABTS2,4-dichlorophenol1004 h93.2% (for 2,4-DCP after 6 cycles)[463]
Bisphenol A10072 h
Indole70.572 h
Anthracene93.372 h
MNPs/polyethylenimineCrosslinking of NPs with PEI using glutaraldehyde followed by chelation of laccase with Cu(II)Phenol in a fixed bed reactor72.93% at a flowrate of 25 μL/min--[449]
MNPs/Cu2+-PEGIn situ oxidation of metal salt using PEG followed by physical adsorption of laccaseMalachite green100 (ABTS)120 min99.9, 90.1, 89.4, 94.6, 76.5, 80.1, 74.6, and 66.1% (respectively, for the dyes after 10 cycles)[425]
Brilliant green96.5 (ABTS)
Crystal violet95.2 (ABTS)
Azophloxine97.7 (TEMPO)
Red MX-5B86.6 (ABTS)
Methyl orange92.7 (VLA)
Reactive blue 1996 (TEMPO)
Alizarin red83.7 (TEMPO)
TiO2/Zn-alginatePhysical entrapment of enzyme in nanocompositeAlizarin red615 h100% (14 cycles)[464]
Tryphan blue96
Malachite green100
Indigo carmine100
Ca-alginatePhysical entrapment with crosslinking of enzyme prior to entrapmentBisphenol A992 h70% (10 successive cycles)[433]
Ca-alginatePhysical entrapment of enzyme in nanocompositeAniline purple86.124 h-[465]
Ca-alginatePhysical entrapment of enzyme in nanocompositeReactive Red 18067.211 days-[466]
Reactive Blue 2188.05
Ca-alginatePhysical entrapment of enzyme in nanocompositeReactive T. Blue9272 h22.3% (6 cycles)[467]
Ca-alginatePhysical entrapment of enzyme in nanocompositeRBBR852 h52.1% and 70% (Bismarck brown and all the others, respectively)[460]
Reactive Black 58024 h
Bismarck Brown R5524 h
Lancet Grey G8524 h
Cu-alginatePhysical entrapment of enzyme in nanocompositeAcid dye38%24 h-[468]
MNPs/chitosanCrosslinking with glutaraldehyde followed by adsorption in laccase solutionReactive yellow 28510 h-[469]
Reactive blue 46012 h
MNPs/poly(GMA-MMA)/Cu-Poly(4-vinyl pyridinePolymer grafting with Cu chelation followed by adsorption of enzymeReactive green 196018 h63%, 76%, and 59% (green, red, and brown dyes, respectively)[470]
Reactive red 288
Reactive brown 1090
Cu-alginatePhysical entrapment of enzyme in nanocompositephenol model solution containing tannic acid, gallic acid, ferulic acid, resorcinol, and pyrogallol756 h35% (8 cycles)[471]
FScubes/PDA@PVDFPrepared the FS/PDA@PVDF membrane using solvothermal process followed by covalent immobilization of laccase using glutaraldehyde as cross linkerCongo red97.13 h85% and 76% (7 days and 5 cycles, respectively)[472]
Table 6. Application of enzyme-nanoparticle-polymer composites in degradation of organic pollutants for application in wastewater treatment.
Table 6. Application of enzyme-nanoparticle-polymer composites in degradation of organic pollutants for application in wastewater treatment.
Nanocomposite (NC)Immobilization MethodPollutants RemovedDegradation (%)Degradation TimeReusabilityRef.
TiO2/polydopamineIn situ polymerization of dopamine on TiO2NPs followed by covalent crosslinking of enzyme with glutaraldehyde2,4-dichlorophenol10030 min100%, 90%, and 63.6% (15, 25, and 40 reuses, respectively)[190]
MNPs/poly(glycidylmethacrylate-co-methylmethacrylate) (poly(GMA-MMA))Crosslinking of enzyme and nanocomposite beads using glutaraldehydephenol862 h84% (8 weeks), 92%, and 79% (phenol and p-chlorophenol, respectively, after 48 h of continuous use)[418]
p-chlorophenol
(in the presence of H2O2)
59
Fe2O3/poly (amido amine) (PAMAM)/silk fibroinCrosslinking of enzyme with nanocomposites using glutaraldehydeBisphenol A in presence of H2O280120 min-[475]
Calcium alginatePhysical entrapment of enzyme in nanocompositeAcid blue 11376240 minCan be recycled up to 3 times[422]
Aluminosilicate halloysite nanotubes/chitosanCrosslinking of enzyme with nanocomposites using glutaraldehydePhenol in presence of hydrogen peroxide98.830 min60% (4 cycles)[476]
MNPs/polyacrylonitrileCrosslinking of enzyme with nanocomposites using glutaraldehydePhenol85.2-52% (5 cycles)[444]
MNPs/poly(vinyl alcohol)/poly(acrylic acid)Physical adsorption of enzyme on nanocompositesEstrone10040 min56.2% (7 cycles)[432]
MNPs/polymethyl methacrylatePhysical entrapment of enzyme in nanocompositePhenol in presence of hydrogen peroxide5550 min-[477]
MNPs/poly(glycidylmethacrylate-co-methylmethacrylate) (poly(GMA-MMA))Crosslinking of enzyme with nanocomposite beads using glutaraldehydePhenol862 h91% and 79% (phenol and chlorophenol, respectively, after 48 h of continuous operation)[418]
p-Chlorophenol
(in presence of hydrogen peroxide in a fluidized bed reactor)
59
Table 7. Application of enzyme-nanoparticle-polymer composites in degradation of organic pollutants for application in wastewater treatment.
Table 7. Application of enzyme-nanoparticle-polymer composites in degradation of organic pollutants for application in wastewater treatment.
Nanocomposite (NC)EnzymeImmobilization MethodPollutants RemovedDegradation (%)Degradation TimeReusabilityRef.
TiO2/polydopamineChloroperoxidase (CPO)Covalent crosslinking of enzyme with nanocomposites using glutaraldehydeAniline blue97.582 min90.3%, 78.2%, and 53.71% (10, 15, and 20 reuses, respectively)[190]
Crystal violet98.982 min
NiFe2O4/tanninGlucose oxidasePhysical adsorption of enzyme on nanocompositeIndigo carmine in presence of UV light98.690 min85.57% (5 cycles)[446]
MnFe2O4/calcium alginateGlucose oxidase and
Laccase
Physical adsorption of enzymes on the nanocompositeMethylene blue82.131 h-[424]
Indigo25.09
Acid red 1420.42
MNPs/PAMAMGlycerophosphodiesterase (GpdQ)Crosslinking of enzyme with nanocomposites using glutaraldehydeOrganophosphate pesticide44.5120 daysUsed as a filter in a Pasteur pipette between two layers of sand[429]
MNPs@SiO2/polydopamineLignin peroxidasePhysical adsorption of enzymes on the nanocompositeTetracycline10024 h80.3% and 67.5% (7 and 14 days of storage), 70% and 30% (4 and 8 cycles, respectively)[447]
Dibutyl phthalate10024 h
5-chlorophenol10024 h
Phenol10024 h
Phenanthrene7924 h
Fluoranthene7324 h
Benzo(a)pyrene6524 h
MNPs/chitosanManganese peroxidaseCrosslinking of enzyme with nanocomposites using glutaraldehydeMethylene blue9650 min91.7% and 86.7% (5 cycles-methylene blue and reactive orange, respectively)[417]
Reactive orange 1698
Fe2O3/chitosanSaccharomyces cerevisiae enzymeAdsorption of chitosan on the NPs surface followed by crosslinking with enzyme using glutaraldehydeCu(II)96.860 min-[478]
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Kyomuhimbo, H.D.; Feleni, U.; Haneklaus, N.H.; Brink, H. Recent Advances in Applications of Oxidases and Peroxidases Polymer-Based Enzyme Biocatalysts in Sensing and Wastewater Treatment: A Review. Polymers 2023, 15, 3492. https://doi.org/10.3390/polym15163492

AMA Style

Kyomuhimbo HD, Feleni U, Haneklaus NH, Brink H. Recent Advances in Applications of Oxidases and Peroxidases Polymer-Based Enzyme Biocatalysts in Sensing and Wastewater Treatment: A Review. Polymers. 2023; 15(16):3492. https://doi.org/10.3390/polym15163492

Chicago/Turabian Style

Kyomuhimbo, Hilda Dinah, Usisipho Feleni, Nils H. Haneklaus, and Hendrik Brink. 2023. "Recent Advances in Applications of Oxidases and Peroxidases Polymer-Based Enzyme Biocatalysts in Sensing and Wastewater Treatment: A Review" Polymers 15, no. 16: 3492. https://doi.org/10.3390/polym15163492

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

Article Metrics

Back to TopTop