Skip to main content
3 Biotech logoLink to 3 Biotech
. 2021 Jul 5;11(8):365. doi: 10.1007/s13205-021-02913-6

Engineered tyrosinases with broadened bio-catalysis scope: immobilization using nanocarriers and applications

Asim Hussain 1, Hamza Rafeeq 1, Muhammad Qasim 2, Zara Jabeen 1, Muhammad Bilal 3,, Marcelo Franco 4, Hafiz M N Iqbal 5,
PMCID: PMC8257883  PMID: 34290948

Abstract

Enzyme immobilization is a widely used technology for creating more stable, active, and reusable biocatalysts. The immobilization process also improves the enzyme's operating efficiency in industrial applications. Various support matrices have been designed and developed to enhance the biocatalytic efficiency of immobilized enzymes. Given their unique physicochemical attributes, including substantial surface area, rigidity, semi-conductivity, high enzyme loading, hyper catalytic activity, and size-assisted optical properties, nanomaterials have emerged as fascinating matrices for enzyme immobilization. Tyrosinase is a copper-containing monooxygenase that catalyzes the o-hydroxylation of monophenols to catechols and o-quinones. This enzyme possesses a wide range of uses in the medical, biotechnological, and food sectors. This article summarizes an array of nanostructured materials as carrier matrices for tyrosinase immobilization. Following a detailed background overview, various nanomaterials, as immobilization support matrices, including carbon nanotubes (CNTs), carbon dots (CDs), carbon black (CB), nanofibers, Graphene nanocomposite, platinum nanoparticles, nano-sized magnetic particles, lignin nanoparticles, layered double hydroxide (LDH) nanomaterials, gold nanoparticles (AuNPs), and zinc oxide nanoparticles have been discussed. Next, applied perspectives have been spotlights with particular reference to environmental pollutant sensing, phenolic compounds detection, pharmaceutical, and food industry (e.g., cereal processing, dairy processing, and meat processing), along with other miscellaneous applications.

Keywords: Tyrosinase, Bio-catalysis, Immobilization, Enzyme engineering, Support matrices, Nanomaterials, Applications

Introduction

Enzymes are effective biocatalysts that provide competitive processes than traditional (chemical) catalysts and significantly increase reaction rates in various chemical and biological processes (Choi et al. 2015). There is a need for reliable biocatalysts as enzymatic bioremediation is recently gaining high research interest as a "green" solution (Das et al. 2018). Owing to the unique characteristics, including stability, refinement, and specificity, several forms of enzymes and related materials with high catalytic potentials are commonly used in several industries, including dairy, fabric, medicines, water management, etc. (Du et al. 2017; Al-Rawi et al. 2020; Cunha et al. 2020). Biocatalyst stability is an essential constraint for their manufacturing applications since they must withstand intense reaction circumstances, such as high temperature and excessive pH (Madhavan et al. 2017). However, due to certain limitations of free enzymes, such as susceptibility to harsh environment, poor activity and stability, no recoverability, limited reusability and so on, the use of solvable enzymes in the chemical and biotechnology industries is restricted (Liu et al. 2017).

Tyrosinase (EC 1.14.18.1) is a monooxygenase that contains a Cu molecule in its catalytic site (Fig. 1) (Zaidi et al. 2014). This enzyme is involved in the o-hydroxylation and oxidation of monophenols resulting in catechols and o-quinones, respectively (Fujieda et al. 2020). It catalyzes the ortho-hydroxylation of tyrosine (monophenol) to 3,4-dihydroxyphenylalanine or DOPA (o-diphenol) and the oxidation of DOPA to dopaquinone. It is also active in the biosynthetic pathways of melanin (o-quinone) that can be converted into melanin pigments (Zaidi et al. 2014). In addition to the biochemical substrates tyrosine and L-DOPA, tyrosinase converts a variety of phenols and diphenols to the related diphenols and quinones (Matoba et al. 2018). CuO2 (II) or a bis (µ-oxo) dicopper (III) is thought to be involved in the transition of an oxygen atom to the phenol substrate (Fig. 2) (Pillaiyar et al. 2018). The resonance Raman spectra of these intermediates are distinctive. Versions of individual intermediate complexes and confirmation for prompt stability among the two modes are provided by synthetic studies (Rolff et al. 2011).

Fig. 1.

Fig. 1

Structure of tyrosinase. Reprinted from Zaidi et al. (2014) with permission under the terms of the creative commons attribution license

Fig. 2.

Fig. 2

An overview of catalytic cycles of tyrosinase. Reprinted from Pillaiyar et al. (2018) with permission from American chemical society

Tyrosinase activities (monophenol hydroxylase and diphenoloxidase) are the foundation of many industrial applications involving the detoxification of phenol containing wastewater and soil (dos Santosa et al. 2013). In therapeutic sectors, it is also used to manufacture o-diphenols as an indicator for melanoma patients (Gradilone et al. 2010), prodrugs activator (Jawaid et al. 2009), and food manufacturing for amendment of diet proteins (Monogioudi et al. 2011). There is a lot of evidence that this enzyme has a lot of promise in the foodstuff, medication, agronomic sectors, diagnostic and ecological activities (Wang et al. 2005). However, these widespread industrial applications are often hindered by a lack of operational stability, long-term shelf life, and cumbersome recovery and re-utilization. Immobilization of enzymes might solve these drawbacks by developing robust, stable, and insoluble biocatalysts (Das et al. 2017).

Enzymes can be quickly isolated and extracted from the reaction substrate, reducing the chance of free enzyme contamination of the finished product. Furthermore, using immobilized enzymes, it is possible to reduce the expense of the related processes. An effective immobilization can significantly improve enzyme's characteristics, such as stability at higher temperature and pH and increased catalytic efficiency (Tacias-Pascacio et al. 2017). It appears that new functionalized supports with low biodegradability, nontoxicity, and robust mechanical properties are needed for convenient and effective immobilization (Wahba 2017). As a result, great emphasis has been given to the immobilization of enzymes, including tyrosinase on various carriers. The synthesis of an appropriate and robust support matrices is crucial for effective immobilization. Traditional immobilization techniques, which are divided into chemical and physical processes, may be used to accomplish the said purpose (Khan et al. 2017). The weak attachment force between the enzyme and the support substance is one of the most detrimental physical approaches. This is why chemical immobilization and covalent bonding, which improve the stability of modified surfaces, are often suggested (Chung et al. 2017).

Strong support structures usually stabilize the enzymes' configuration, and thus, their functions are maintained (Husain 2017). Immobilized enzymes are thus more stable and resistant to environmental modifications than free enzymes (Wang et al. 2020). Furthermore, heterogeneous immobilized enzyme systems allow fast recovery of enzymes, multiple enzyme reuse, hyper-activity and stability, rapid reaction termination, and a wider range of bioreactor designs (Bernal et al. 2018). In recent years, there has been a lot of curiosity and focus on the ability of immobilized enzymes (Bommarius and Riebel-Bommarius 2004). Immobilized enzymes usually are more stable and simpler to manage than their free versions. Furthermore, the reaction products are not polluted with the enzyme (which is particularly beneficial in the food and pharmaceutical industries). In proteases, the rate of autolysis can be significantly slowed while they are immobilized (Massolini and Calleri 2005). When it comes to the efficiency of the immobilized enzyme mechanism, the matrix properties are crucial. Physical tolerance to compression, hydrophilicity, inertness to enzymes, ease of derivatization, biocompatibility, resistance to microbial attack, and low-cost are all desirable properties (Barry and O'Riordan 2016).

Various methodologies have been established for the immobilization of tyrosinase. Some of which include silica sol–gel composite films (Wang et al. 2000), tyrosinase–modified boron-doped diamond electrodes (Notsu et al. 2002; Zehani et al. 2015), tyrosinase entrapped in an agarose-guar gum (Tembe et al. 2007), glassy carbon electrode modified by gold nanoparticles (Sanz et al. 2005), tyrosinase-FeO3 nanoparticles-chitosan (Wang et al. 2008a, b), use of polyaniline-carbon nanofibers or nanotubes (Zhang et al. 2009; Branzoi et al. 2010; Yin et al. 2010; Rodríguez Méndez et al. 2013; Apetrei and Apetrei 2015), diazonium salt (Cortina-Puig et al. 2010) or titania sol–gel matrix (Kochana et al. 2015).

This work summarizes recent advancements in the use of a variety of nanostructured materials as carrier matrices for tyrosinase immobilization. The review also spotlights the biocatalytic capabilities and applications of the immobilized tyrosinase-based engineered nano-biocatalytic systems in the medical, nutritional, and biotechnological sectors.

Nanomaterials for enzyme immobilization

Because of the exponential advancement in nanotechnology, nanomaterials and enzymes are increasingly being combined for use in biomedicine, biochemical engineering, ecological control, and biosensors (Chen et al. 2017a, b). Since 2014, there has been significant development in the field of enzyme nanomaterial interactions, including enzymatic alteration or deprivation of nanomaterials, and enzyme immobilization (Chen et al. 2017a, b).

Nanoparticles typically have a diameter of 1–100 nm and are made up of hundreds of atoms. Clusters of particles are described as particles with a diameter of less than 1 nm. Nanoparticles with less than 10 nm diameter are especially intriguing because they can be called almost entirely surface. After all, all atoms are on or near the surface (Shi et al. 2015). Nanostructures have been used to promote enzyme immobilization via various mechanisms, including adsorption, covalent binding, encapsulation, cross-linking and others (Yang et al. 2016). It should be noted that controlling enzyme–support interactions is critical for understanding the various implementations and improvements that can be made, and maintaining the orientation of the enzyme can be particularly important (large substrates, immobilization area) (Barbosa et al. 2013).

Nanomaterials have proven to be one of the most effective immobilization supports owing to their distinctive features, like their large surface area (Sharifi et al. 2020; Liang et al. 2020). Although the high surface-to-volume ratio is advantageous, Brownian motion will affect the action of enzymes immobilized on these matrices (Gupta et al. 2011). It can overcome the diffusional limitations of traditional reaction systems and enhance enzyme stacking per unit mass of support (Zhang et al. 2016). Micron-sized particles have been used as carriers or matrices for enzyme immobilization. The size employed by the carriers/matrices is much greater than that required for the enzyme molecules, which has been a source of concern. One of the main motivations for prompt effort on enzyme aggregates (without carriers) was the need to reduce bioreactor sizes (Bilal et al. 2019). Crosslinked enzyme clusters are a more modern variant of such enzyme aggregates (CLEA). The growing nanotechnology-based research interests and the focus on miniaturization have further strengthened the enzyme immobilization using different nanomaterials in various shapes and sizes (Orfanakis et al. 2018). The implication is that nanomaterials can aid in the reduction of bioreactor size. This is because a greater amount of biocatalyst should be loaded per unit weight of the carrier (Cho et al. 2020). This article summarizes recent progress on using an array of nanostructured materials as carrier matrices for tyrosinase immobilization, biocatalytic properties, and applications of the resulting immobilized tyrosinase-based engineered nano-biocatalytic systems in the medical, nutritional, and biotechnological sectors. The nanomaterials used for tyrosine immobilization are listed in Table 1.

Table 1.

Nanomaterials for immobilization of tyrosinase

Nanomaterial Application References
Carbon dots Biosensor

Samdani et al. (2016)

Mohammadi et al. (2011)

Baluta et al. (2020)

Platinum nanoparticles along with graphene Biosensor Liu et al. (2011)
Carbon black Biosensor

Nadifiyine et al. (2013)

Ibáñez-Redín et al. (2018)

Carbon nanotubes Biosensor Kim et al. (2010)
Lignin nanoparticles/nanocapsules Capecchi et al. (2018, 2019)
Magnetic nanoparticles Degradation of phenolic compounds

Liu et al. (2018)

Abdollahi et al. (2019)

Layered Double Hydroxide (LDH) nanomaterials Biosensor Soussou et al. (2017)
Nanofibers Biosensor Mercante et al. (2021)
Graphene nanocomposite Biosensor

Yang et al. (2015)

Tang et al. (2013)

Reza et al. (2015)

Poly(diallyldimethylammonium Chloride) Capped Gold Nanoparticles Biosensor Chaimuangyong et al. (2019)
Zinc oxide Biosensor

Liu et al. (2006)

Gu et al. (2009)

Carbon nanotubes

Because of their unique electronic and mechanical qualities, carbon nanotubes (CNTs) have been intensively researched as electrode materials in recent years (Basiuk et al. 2019). CNTs are supposed to have intrinsic electrochemical qualities comparable to other carbon electrodes commonly applied in numerous electrochemical uses from a chemical standpoint. Compared to other carbon-based nanomaterials like C60 and C70, CNTs have somewhat distinct electrochemical characteristics (Schroeder et al. 2018). When used as an electrode supporter, CNTs' sensitive electrical properties enable them to arbitrate electron transfer reactions with electroactive species in solution. As a result, CNT-based electrodes have become commonplace in electrochemical sensing (Gupta et al. 2018). Electrodes modified with carbon nanotubes have been designed for biosensors (Tang et al. 2017). The CNT-coated electrodes are the most straightforward to create, as long as stable CNT suspensions are preserved. For the fabrication of the CNT/surfactant-modified electrode, suitable surfactants were selected to achieve high catalytic efficiency and strong selectivity against the analytes. However, there is little information about making polymer-modified CNT electrodes extremely catalytic and selective for analytes (Yin et al. 2017). Kim et al. (2010) used PAAc-g-MWNT and PMAn-g-MWNT to make tyrosinase-immobilized biosensors. The detection limits of the tyrosinase-immobilized biosensor dependent on PAAc-g-MWNT and PMAn in the phosphate-buffered solution are 0.2–0.9 mM concentration and 0.1–0.5 mM for phenol.

Carbon dots

Despite their unique qualities, such as hydrophilic nature, less toxic, comparatively high strength, and biochemically stability over time, carbon dots (CDs) are a promising sensor medium (Bonet-San-Emeterio et al. 2020; Ding et al. 2020). The use of the CDs medium to evaluate different compounds, emphasizing their fluorescent qualities, is recorded in the literature, for example, hydrazine in water (Hiremath et al. 2020). CDs are being used in electrochemical sensors for a diversity of uses, including the assessment of ascorbic acid (Zhou et al. 2020), riboflavin (Priyadarshini et al. 2020) and neurotransmitter (Mphuthi et al. 2016). Using a glassy carbon electrode (GCE) updated with FeMoO4 nanorods, Samdani et al. (2016) anticipated an electrochemical approach for norepinephrine (NE) determination. Adopting cyclic voltammetry (CV) and differential pulse voltammetry (DPV) techniques, synthesized nanorods were used to oxidase NE. The spectrophotometric response of NE revealed an immediate increase in current. Mohammadi et al. (2011) tested a different approach for detecting NE. SWCNTs and restrained tyrosinase were added to a glassy carbon electrode. For NE assessment, the DPV technique was used. The changed electrode's detection limit for NE was determined to be 0.1 M.

Baluta et al. (2020) used nontoxic reagents to develop a modern biosensor-based method to measure neurotransmitters as easily as possible. The Au-E biosensor was developed and built using tyrosinase immobilized on an electroactive coating of cysteamine and carbon nanoparticles covering the gold electrode. This sensing system used voltammetric techniques, such as CV and DPV, to calculate the catalytic oxidation of norepinephrine (NE) to NE quinone. The organized bio-system had excellent constraints, including a comprehensive direct range (1–200 M), a 196 nM limit of detection, a 312 nM limit of quantification, and high discrimination and sensitivity abilities. It's worth noting that the mentioned approach was effective in determining NE in real samples.

Carbon black

Owing to its lower price and exceptional physicochemical qualities, Carbon Black (CB) has risen to prominence as a carbon nanomaterial to apply electrochemical biosensors (Talarico et al. 2015). Since these attributes are heavily reliant on external interaction, the impact of functionalization has been extensively researched for a variety of uses. In the interim, the impact of CB fabrication on its characteristics in electroanalytical uses is still understudied (Silva et al. 2017). Carbon paste electrodes are common in electroanalysis because of their low surface current, impending high range, biochemical resistance, and ease of preparation from low-cost materials. Carbon has been used in various ways, from graphite, glassy carbon, carbon nanotubes, and more recently, carbon black (Arduini et al. 2010a, b; Deng et al. 2011). Carbon black was commonly used as a carbon substantial to manufacture fuel cells, lithium batteries, sodium batteries, and sensors for identifying gaseous phases (Martinez-Alvarez and Miranda-Hernandez 2008). Conversely, only a few carbon black-based biosensors have been identified (Razumiene et al. 2005). Arduini et al. (2011, 2012) conducted research that showed the appropriateness of carbon black as an electrochemical transducer.

A carbon black paste electrode was developed. This device showed significantly reduced noise compared to enzymatic sensors based on traditional graphite paste electrodes. Using catechol as a substrate, the response was linear in the range 0.013 to 150 µmol L−1 with a detection limit of 6 nmol L−1 (based on three times the S/N ratio). Peroxidase and laccase biosensors were also prepared. The responses of these three enzymatic biosensors to twenty different phenolic compounds were investigated, taking into account their molecular structure and their specific enzyme activity relationship. In addition, another sensor was reported based on semipurified tyrosinase. There was a good statistical correlation between the results obtained with tyrosinase biosensor and Folin–Ciocalteu spectrophotometric methods for phenol determination in olive oils.

Nadifiyine et al. designed a prototype biosensor established by the halt of commercially available tyrosinase onto a carbon black paste electrode (2013). As compared to enzymatic sensors using conventional graphite paste electrodes, this instrument produced significantly less noise. The output was linear in the range of 0.013 to 150 mol/1 using catechol as the substrate, with a diagnosis limit of 6 nmol/L. Biosensors for peroxidase and laccase were also created. The contributions of biosensors to 20 distinct phenolic substances were studied, with their structure and activity association taken into account. Ibáez-Redn et al. (2018) identified the production of responsive electrochemical biosensors. The chemical pre-treatment improved the hydrophilicity of the substance by increasing the amount of exterior oxygenated functional groups, as measured by elemental analysis and FTIR. Compared to unfunctionalized CB, the electron transfer rate constant (k0) was observed to be nearly 100-fold higher, indicating a significant increase in electrocatalytic properties. They developed a tyrosinase biosensor for catechol determination in water samples using functionalized CB and di-hexadecyl phosphate (DHP). Their findings show that functionalized CB is an ideal medium to produce biosensors.

Nanofibers

Nanofibers are effective platforms for immobilizing biomolecules for further use in biosensing (Andre et al. 2021). Electrospinning is a novel tool that blends low price, easiness of processing, and implementation for a wide range of materials, among other techniques for manufacturing such nanofibers (Mercante et al. 2017). Electrospinning is an electrohydrodynamic method that produces fibers with diameters ranging from nm to m by uniaxially stretching and elongating a viscoelastic solution. The nanofibrous membranes have a high surface-to-volume ratio, a porous integrated configuration, a low diffusion barrier, and amendable flexibility (Li et al. 2019). These characteristics allow for the development of hybrid and composite nanofibrous films with high sensitivity and selectivity for various analytes (Yang et al. 2020). Furthermore, the nanofibrous structure allows for high enzyme loadings, which is advantageous for biosensing efficiency. Regardless of the appropriateness of electrospun nanofibers, adjusting the configuration of the nanofibrous film to attain high performance for an assumed analyte requires significant effort (Jankowska et al. 2020).

Mercante et al. (2021) identified a polymeric electrospun nanofiber of polyamide 6 (PA6) and poly(allylamine hydrochloride) (PAH) ornamented with gold nanoparticles (AuNPs), referred to as PA6/PAH@AuNPs, that was dumped onto a fluorine-doped tin oxide (FTO) substrate for perceiving low levels of Bisphenol A (BPA). The immobilization of tyrosinase by the hybrid layer was fantastic, allowing for amperometric identification of BPA with a maximum recognition of 0.011 M in the concentration of 0.05 to 20 M. The collective outcome of the bulky surface area and permeability of PA6/PAH nanofibers, the catalytic action of AuNPs, and Tyr's oxidoreductase capability is credited with the improved sensing efficiency. Such discoveries exposed the way to new biosensing architectures for monitoring BPA and other endocrine-disrupting chemicals (EDCs) in water supplies.

Graphene nanocomposite

Graphene has perfect and unusual properties, such as high charge transport agility, wide surface area, strong chemical stability, and mechanical strength (Du et al. 2011). As a result, it has sparked considerable interest in various technical areas, including nanoelectronics, nanocomposite, biosensing, and bioelectronics (Sun et al. 2011). To satisfy the desired condition in each application, graphene must be chemically functionalized, which can affect considerable reductions in the electrical conductivity of the functionalized graphene (Kuilla et al. 2010). This could be attributed to an increase in blemish quantity and drugging of shielding resources on the graphene surface, all of which would impair electron transport and affect electrical conductivity. As a result, research into combining graphene with other nanomaterials to attain graphene exterior amendment to improve conductivity is also required.

Several environmentally safe processes, such as electrodeposition, an association of metal nanoparticles on functionalized graphene top, and microwave therapy, have been employed to anchor noble metal nanoparticles on graphene nanosheets (Yin et al. 2015). For instance, Yang et al. (2015) used a green approach to anchor AuNPs onto carboxylic graphene nanosheets without adding any reductant or shielding particle. Tang et al. (2013) demonstrated the electrochemical deposition of reduced graphene oxide and gold nanoparticles onto a glassy carbon electrode (GCE) surface without using a harmful reductant. AuNPs have a suitable atmosphere for enzyme immobilization and promote electron transmission between the immobilized enzyme and the electrode, which is why they were chosen as the nanomaterial (Makhotkina and Kilmartin, 2010). Furthermore, graphene is an auspicious support matrix for metal nanoparticles owing to its wide surface area, electronic properties, and mechanical strength. Since these special requirements reduce AuNP agglomeration and improve their stability in the graphene nanosheet matrix, future use in biosensing has been investigated (Li et al. 2010).

Applications of graphene nanocomposites have recently piqued researchers' curiosity in various fields, especially in the construction of biosensors. Reza et al. (2015) developed a tyrosinase biosensor based on rGO/Chit to detect bisphenol A electrochemically. The reduced GO-AuNPs film for tyrosinase halt for phenol discovery has been investigated. Graphene–polyaniline and tyrosinase-modified GCE were also used to study electrochemical biosensors for phenolic derivatives. Additionally, graphene/Co3O4 was successfully employed as a medium for tyrosinase immobilization and catechol detection in fruit samples (Liang et al. 2016).

Platinum nanoparticles

Despite its peculiar qualities (Li et al. 2008; Wang et al. 2008a, b), like strong density, versatile electrical conductivity, and extraordinary heat resistance and physical attributes, graphene has piqued interest as a good resource (Shan et al. 2010). Furthermore, graphene's plane structure provides a large area for charging nanoparticles ideal for electrode amendment (Wang and Hu 2009). Because the incorporation of carbon-based substances and metal nanoparticles, such as CNT–Pt composites, has been shown to have a synergic effect in electrocatalytic applications, (Yao et al. 2006) reason to believe that graphene and platinum nanoparticles would have the same impact. Besides, when the pH is greater than 5, tyrosinase is negatively charged. As a result, electrostatic interference among a positively charged self-assembled mono-layer (SAM) and the negatively charged enzyme particles under an appropriate pH range is a viable way to immobilize tyrosinase (Love et al. 2005). For the identification of organophosphorus pesticides, Liu et al. (2011) created an amperometric biosensor. A glassy carbon electrode adapted with graphene and the enzyme tyrosinase restrained on platinum nanoparticles were used. Different pesticides can be identified in catechol as a substrate since they suppress the enzyme that mediates the oxidation of catechol to o-quinone. Platinum nanoparticles and graphene improve the performance of the electrochemical removal of o-quinone, resulting in increased sensitivity.

Magnetic nanoparticles

The development of an effective method for recovering biocatalyst from reaction mixtures has received a lot of attention. Nano-sized magnetic particles have recently become popular as an excellent help for immobilizing biomolecules since they can be quickly retrieved from the reaction medium by inserting a powerful magnet close to the mixture (Xie and Zang 2017). Magnetic nanoparticles (MNPs) are a viable carrier for biomolecule immobilization artifacts due to their ease of separation with the aid of a magnetic field and their low toxicity (Khoshnevisan et al. 2017; Wang et al. 2016). MNPs may also be isolated from the media with an additional magnetic field (Bezerra et al. 2017).

Because of their excellent advantages of fast synthesis and isolation, magnetic nanoparticles are commonly employed as carriers for enzyme arrest (Feng et al. 2016). For more modifications, they are normally amino-functionalized. Amino-functionalization agents include 3-aminopropyl) triethoxysilane (APTES), polyethyleneimine (PEI), and dopamine (Li et al. 2015). However, these amino-functionalization processes are time-consuming and necessitate the use of assignable agents. The one-pot solvothermal synthesis of aminated magnetic nanoparticles (Fe3O4-NH2) appears to be a promising approach for avoiding the time-consuming amino-functionalization process (Martn et al. 2014).

To overcome the limitations of both the noncovalent and covalent processes, Liu et al. (2018) developed a procedure that fused the immobilization technique of physical adsorption with covalent crosslinking. Abdollahi et al. (2019) immobilized tyrosinase onto functionalized nanoparticles. Superparamagnetic nanoparticles were synthesized, and silica-coated using a simple process, and their ability for tyrosinase immobilization was demonstrated. A combination of tyrosinase and cyanuric chloride functionalized magnetic nanoparticles (Cyc/EMNPs) was shaken at room temperature to immobilize the enzyme.

Lignin nanoparticles

Because of their unique environmentally friendly qualities, green resources are thought to play a positive role in the Circular Economy (Velte and Steinhilper 2016). Paper and biorefinery industries produce more than 50 million tons of lignin per year, making it the most abundant organic polyphenol in nature (Ragauskas et al. 2014). Lignin derivatives are low-price ingredients that work like consequent petroleum equivalents, exhibiting antioxidant activity (Kaur and Uppal 2015), UV-shielding safety, anticorrosion possessions, and electrochemical sensitivity in all charge storing compounds and electrochemical sensors due to their aromatic character (Ding et al. 2016). As a result of the direct consequence of the "size-effect" phenomenon, lignin is incipient as an innovative and affordable base substance for synthesizing nanocapsules with improved physicochemical characteristics (Xiong et al. 2017). In theory, these nanoparticles may be effective polyphenol-based platforms for enzyme immobilization. The supramolecular connection between lignin and proteins has been investigated in-depth to recognize affinity lignin-binding peptide structures and determine the part of polymer structural variations in the molecular classification stage (Leskinen et al. 2017). Protein folding composition and Coulombic forces in the non-ionic medium have occurred as critical constraints in protein self-assembly on the lignin surface (Strobel et al. 2015; Westereng et al. 2015). Lignin and tyrosinase work together to achieve effective oxidative processes. Tyrosinase has been extensively researched due to its applications in skin health, cosmetics, and agriculture industries. When used in the immobilized state, it performs well in industrial applications due to its reusability, increased flexibility, and simplified purification procedures (Dinçer et al. 2012).

Guazzaroni et al. (2012) identified the processing of heterogeneous catalysts for the production of bioactive catechol derivatives by immobilizing tyrosinase on microcapsules of the epoxy resin Eupergit, or MWCNTs. The production of neuroactive 3,4-dihydroxyphenylalanine (DOPA)-containing peptides and lipophilic catechols with extensive spectrum antiviral activity against DNA and RNA viruses was aided by these catalysts, which had a high catalytic ability (Bizzarri et al. 2017; Botta et al. 2015). In general, nanostructured catalysts outperformed microcapsule counterparts in terms of activity, storage life, and reusability (Botta et al. 2017).

Capecchi et al. (2018) identified different methods for immobilizing tyrosinase in organosolv lignin (OL) nanoparticles. OL has a low molecular weight and a discrete solubility in organic solvents compared to native lignin and other scientific lignins. The novel catalysts have been characterized in terms of structural and kinetic properties, and they have been used to convert selected phenols to bioactive catechols in a series of experiments. Notably, behavior and kinetic parameters for lignin-dependent tyrosinase catalysts prepared by layer-by-layer assembly were on par with or better than those previously recorded for catalysts based on conventional inorganic supports.

Layered double hydroxide nanomaterials

Layered Double Hydroxide (LDH) nanomaterials are desirable candidates for electrochemical investigation in broad-spectrum and for enzyme arrest (Tonelli et al. 2013) between host materials previously used for the immobility of enzymes distinguishing polyphenols (Zehani et al. 2015). Subsequently, they display several strategic qualities in the production of hybrid biomaterials. LDHs are a unique type of substance made up of positively charged layers with charge-balancing anions sandwiched. Like their strong ion exchange potential or encapsulation efficiency, special characteristics are due to their unique nature. Finally, these substances have excellent biocompatibility, which is significant in the sense of enzyme immobilization (Shan et al. 2003). In addition, the immobilization of various types of biomolecules, such as amino acids or DNA, onto LDH films has been successfully tested. LDH's properties cause an enzyme to be immobilized while maintaining its automatic configuration without disrupting catalytic sites, minimizing any inhibition consequence (Baccar et al. 2012). Thus, they are also well suited to the production of electrochemical biosensors. Antibodies and enzymes have also been suggested as applications for this kind of substance (Temani et al. 2014). Hybrid matrices composed of chitosan and LDHs were identified in the case of tyrosinase (Han et al. 2007). The use of tyrosinase-MgAlCO3 LDH to functionalize glassy carbon electrodes has also been documented (Han et al. 2012). Furthermore, the chemical composition of LDH material can affect enzymatic behavior. Previous experiments found that tyrosinase is triggered in the existence of Co or Zn and inhibited in the presence of Ni, so they intended to immobilize it using an LDH content containing cobalt (Co1.57Al (OH) xSO4, shortened hereafter as CoAl) (Hidouri et al. 2011). Finally, the electrochemical activity of cobalt-based LDHs has been well studied, as cobalt has excellent electrochemical properties (Han et al. 2011).

For the construction of an electrochemical tyrosinase-based biosensor used for the identification of a hybrid combination of polyphenols isolated from green tea, Soussou et al. (2017) investigated gold screen-printed electrodes glazed with thin films of Layered Double Hydroxides (LDHs) containing Co and Al (Co1.57Al (OH) xSO4, shortened as CoAl). According to physicochemical studies, the resulting biosensor has excellent accuracy, a wide vibrant range, and very small detection confines, 0.33 pg/mL and 0.03 pg/mL for oxidation and reduction.

Gold nanoparticles

Gold nanoparticles (AuNPs) are an excellent option for biosensing electrodes because of their conductivity, durability, biocompatibility, and broad surface region. Gold nanoparticles (AuNPs) and carbon nanotubes (CNTs) were used for their outstanding physio-chemical qualities. AuNPs were discovered to have excellent conductivity, surface area, and catalytic properties (Chen et al. 2018). Rolling graphite sheets create carbon nanotubes (CNTs), which have quasi-one-dimensional (1D) shapes. It has good chemical stability, high surface area, excellent adsorption potential, excellent biocompatibility and can be used to promote electron transfer between electroactive species and electrodes (Amatatongchai et al. 2015). Furthermore, some polymer materials are used in conjunction with other materials. The polymer poly-(diallyl dimethylammonium chloride) (PDDA) was electrochemically dumped on the electrode exterior to create PDDA-capped AuNPs (PDDA@AuNPs), a polymer with high electrical conductivity and ionic power. These materials are being used to produce dopamine biosensors that are accurate, clear, inexpensive, and quick to analyze (Liu et al. 2006).

Chaimuangyong et al. (2019) developed dopamine biosensors with tyrosinase (Tyr)-deposited poly-(diallyldimethylammonium chloride)-capped gold nanoparticles (PDDA@AuNPs) composite and carbon nanotubes (CNTs)-adapted glassy carbon electrode (GCE) (Tyr/PDDA@AuNPs-CNTs/GCE). Dopamine was detected using cyclic voltammetric and amperometric techniques. The parameters of the applied potential, Tyr loading, PDDA@AuNPs number, and carbon nanotubes amount were all optimized. Under mild conditions, the modified electrode had high immunity with no intervention from uric acid, ascorbic acid, cholesterol, ethanol, or aspirin. It demonstrated high reliability for more than 30 years of use. With a maximum sense of 8.50 nM (S/N = 3), the dopamine biosensor showed two linear ranges of 10 nM to 100 nM (R2 = 0.9997) and 100 nM to 1 mM (R2 = 0.9979).

Zinc oxide nanoparticles

Apart from optoelectronic properties, nanostructural ZnO has several benefits for biosensors, including a high aspect ratio, a polar surface along the c-axis, strong electron contact, and nontoxicity (Wei et al. 2006). Notably, ZnO has a high isoelectric point (IEP) of around 9.5, making it ideal for immobilizing biomolecules with low IEP, such as enzymes and proteins, using electrostatic attraction at the suitable pH (Ahmad et al. 2010). Biosensors made of ZnO nanoparticles, porous film, nanocombs, and nanorods have been used to sense cytochrome c, protein, uric acid, glucose, and phenolic acid correspondingly (Wang et al. 2006; Taratula et al. 2006). The semiconductor ZnO is a very appealing material. Rubber, ceramics, paints, pharmaceuticals, and sensors all use bulk zinc oxide (Liu and Zeng, 2003). Topoglidis and colleagues described the immobilization and bioelectrochemistry of proteins on nanoporous TiO2 and ZnO films, claiming ZnO with a high isoelectric point (9.5) was ideal for the adsorption of proteins with low isoelectric points (Topoglidis et al. 2001). As a result, ZnO nanoparticles (nano-ZnO) merit auxiliary study as a potential candidate for support material in biosensor fabrication.

Liu et al. (2006) established a phenol biosensor that is free of mediators. Tyrosinase with a low isoelectric point was adsorbed on the surface of ZnO nanoparticles (nano-ZnO) with a high isoelectric point, aided by electrostatic connections, and then incapacitated on the glassy carbon electrode by chitosan film formation. The nano-ZnO matrix was found to have a favorable microenvironment for tyrosinase loading due to its desirable isoelectric point and the immobilized tyrosinase maintaining its operation to a large degree. Furthermore, no other electron mediators are needed. The undeviating reduction of biocatalytically produced quinone species at 200 mV was used to evaluate phenolic compounds (vs. saturated calomel electrode). The enzyme electrode's fabrication parameters, as well as the different experimental variables, were augmented. The subsequent biosensor has a sensitivity of 182 lA mmol−1 L and can attain 95% of steady-state current in 10 s. The phenol sensing range was 1.5 * 10–7 to 6.5 * 10–5 mol L−1, with a sensing limit of 5.0 * 10–8 mol L−1 at a signal/noise ratio of 3. The obvious Michaelis–Menten constant (KMapp) and the enzyme electrode's stability were also calculated. The established biosensor's efficiency was likened to that of biosensors based on different immobilization matrices.

Gu et al. (2009) used zinc powders as raw material and hydrothermally fabricated ZnO nanorods on a gold wire without using some additional surfactant or alleviating agent. The gold wire was expertly handled to increase nucleation for ZnO nanostructure progress and boost the biosensor's efficiency, which was developed by immobilizing tyrosinase (Tyr) on ZnO nanorods for phenol discovery. According to electrochemical measurements, FTIR and SEM studies, the Tyr was stably adsorbed on the ZnO nanorod surface with bioactivity for phenol oxidation. The biosensor achieved a 95% steady-state current of less than 5 s, with a sensitivity of 103.08 A/mM at Cphenol > 20 M and 40.76 A/mM Cphenol 20 M. At a signal/noise ratio of 3, the sensing limit of 0.623 M was obtained.

Applications of tyrosinases

Environmental pollutant sensing

Because of their high sensitivity, quick reaction, and ease of use, enzyme-based biosensors are a promising option for detecting pesticide residues (Amine et al. 2006). Biosensors that use potentiometric or amperometric transducers to inhibit the cholinesterase (ChE) enzyme are the most often used for recognizing and quantifying organophosphates (Arduini et al. 2010a, b). In recent times, biosensors focused on the restriction of tyrosinase action by a compound (monophenol or o-diphenol) have been established to detect contaminants (Anh et al. 2004). Tyrosinase is a catechol oxidase with a dinuclear copper core that is found in a wide range of microorganisms, plants, and animals (Chen and Jin 2010). In the presence of oxygen as an electron acceptor, tyrosinase can catalyze the conversion of monophenols to o-diphenols, which can then be oxidized to o-quinones (Cortina-Puig et al. 2010). On the electrode surface, the produced o-quinones can be electrochemically reduced to o-diphenols again. The analysis of improvements in electrochemical reduction signals of o-quinones resulting from the inhibition of tyrosinase production can be used to detect contaminants. As a result, one of the research's main goals is to create a biosensor with high sensitivity (Liu et al. 2011).

Numerous amperometric biosensors dependent on the inhibition of tyrosinase enzyme action have been employed to determine the presence of triazine and phenylurea herbicides in the atmosphere in recent years (Martinazzo et al. 2018). Herbicides are commonly used despite the risk they pose to the ecosystem since they have less ecological perseverance, leading to severe toxicity. Since their existence might be found in external and earth water, they pose a severe hazard to human health. Vaz et al. (2021) developed a carbon nanosphere-based sensor for herbicide detection (Fig. 3). For the identification and quantifiable assessment of lethal intensities, a variety of techniques are used. Usually, LC or GC are employed, but owing to their simpler sample preparation, electrochemical enzyme sensors are also considered a substitute to traditional spectrometric approaches for pollutant detection (Bucur et al. 2018). Fe3O4 nanoparticles have been inserted into a sol–gel/chitosan biosensor membrane to improve biosensor sensitivity and reaction, along with immobilized alkaline phosphates. These nanoparticles, which are used in electrochemical biosensors, may be responsible for a desirable microenvironment for biomolecules like proteins to directly interchange electrons with an electrode (Abdel-Aziz and Heikal 2021). Electrochemical biosensors have also been constructed using enzyme-based biosensors, such as laccase, tyrosinase, glucose oxidase, horseradish peroxidase, and Fe3O4 nanoparticles. The tyrosinase electrode is often used to track phenolic and catecholic substances, and it is one of the most powerful transducers for phenol or catechol discovery (Tucci et al. 2019).

Fig. 3.

Fig. 3

Carbon nanosphere-based sensor used for herbicide detection. Reprinted from Vaz et al. (2021) with permission from Elsevier

Pharmaceutical application

L-Dihydroxy phenylalanine is a naturally befalling nutritional complement and psychoactive agent present in many foods and herbs and is synthesized in the mammalian body and brain from the amino acid L-tyrosine. L-DOPA is a predecessor of the central nervous system's release of dopamine. As a result, L-DOPA is used as a potent medication to cure Parkinson's disease and regulate neurogenic myocardium damage (Min et al. 2019). The enzyme tyrosinase was used as a biocatalyst to manufacture L-DOPA using L-tyrosine as a substrate and L-ascorbate as a reducing agent (Fig. 4) (Zaidi et al. 2014). The amount of L-DOPA generated in batch reactors using microbial tyrosinases varies from 1.44 to 54 mg (Tan et al. 2019). However, efficiency appears to be significantly lower due to two factors. First, the addition of L-tyrosine seems to be insufficient, with less than 30% of it absorbed throughout the progression. Second, dopaquinone, leukodopachrome, and dopachrome are formed with melanin due to lateral reactions and reversible progression intermediates, which are eliminated by adding L-ascorbate in a concentration equivalent tyrosine (Ilesanmi and Adewale 2020). In continuous and batch processes, Ates et al. (2007) used microbial tyrosinase in Cu-alginate gels. L-DOPA has a global sales share of $101 billion per year, and therefore, new manufacturing processes are also being investigated. The global demand for L-DOPA is shown by an increase of 250 tons per year (Etemadi et al. 2018).

Fig. 4.

Fig. 4

L-DOPA production by tyrosinase using L-tyrosine as a substrate and L-ascorbate as a reducing agent. Reprinted from Zaidi et al. (2014) with permission under the terms of the Creative Commons Attribution License

Phenolic compounds detection

Because of the significant concerns about toxicity, the accurate quantification of phenols in composite environmental matrices has received a lot of interest. Wastewater from various industries, including coal conversion, resin and plastics, petroleum refineries, textiles, dyes, iron and steel, and pulp and paper, contains phenolic compounds (Mishra and Chiang 2020). Phenols are harmful pollutants found in industrial waste that pose several health risks, with some of the potential carcinogens. Furthermore, phenol induces coloration of the receiving streams, making it necessary to decontaminate the compound (Fig. 5) (Wee et al. 2019). A number of researchers have investigated the use of enzymes in wastewater management. Elimination of phenol and its byproducts using a polyphenol oxidase enzyme such as tyrosinase has become a very important and efficient process. S. antibioticus tyrosinase, for example, was active against industrial toxins, including 3-and 4-chlorophenols and 3-and 4-fluorophenols (Camargo et al. 2018). The use of bacterial tyrosinase to treat polluted waste has recently been appraised, and it could be achieved using tyrosinase generating stains or the enzyme in an immobilized state as the protagonist. Tyrosinase benefits over other enzyme mechanisms employed for phenol elimination in that it uses molecular oxygen as the oxidant rather than H2O2, thereby lowering the cost of implementation (Laranjo et al. 2019). Electrochemical biosensors focused on immobilized tyrosinase have gained the most interest among the various diagnostic approaches employed to quickly check these phenolic compounds. For the identification of phenolic compounds, spectrophotometric or chromatographic methods have traditionally been used. However, new techniques have been introduced that may provide higher precision, lower costs, and quicker and easier sample handling (Mishra and Chiang 2020).

Fig. 5.

Fig. 5

Schematic illustrations of (A) EAPC protocol, consisting of enzyme adsorption, precipitation, and crosslinking, and (B) electrochemical reactions occurring in the EAPC electrode on glassy carbon electrode. Reprinted from Wee et al. (2019) with permission from Elsevier

Tyrosinases catalyze the transfer of phenolic substrate to quinine species, electrochemically reduced for phenolic analyte detection at low potentials. Adsorption, crosslinking, on the surface of electrodes, entrapment in silicone films and hydrogels, carbon paste matrix, and graphite-epoxy composite electrodes are all used to immobilize tyrosinase an electrochemical transducer (Yin et al. 2019). Immobilization of tyrosinase to identify phenolic compounds has also been accomplished using electropolymerization, self-assembled monolayers, silica sol–gel, alumina sol–gel, and nanoparticles (Cerrato-Alvarez et al. 2019).

Applications in the food industry

Montereali et al. (2010) used a biosensor based on tyrosinase and laccase to identify polyphenols found in musts and wines. Both enzymes were halted on ferrocene-modified graphite screen-printed electrodes. These biosensors had strong specimen performance associated with spectrophotometric analysis, but due to SO2, enzymatic activity was inhibited. Tyrosinases are also extracted as a toxin byproduct in industrial fermentation processes (Osma et al. 2010).

Cereal processing

Tyrosinases have been extensively explored for their capability to catalyze the oxidation of phenolic compounds found in breakfast cereal proteins and polysaccharides by forming connections in or from polysaccharides, proteins and polysaccharides, or proteins themselves (Glusac et al. 2018). Tyrosinase (mushroom extract rich in polyphenol oxidase) has been used to study the role of tyrosinase in the wheat dough and the development of 2-S-cysteinyl-DOPA 2,5-di-S-cysteinyl-DOPA, 6-S-cysteinyl-3, 4-DOPA, and di-DOPA crosslinks in gluten proteins (Nawaz et al. 2017).

Dairy processing

To create a smooth consistency securer, crosslinking may be used in dairy products to avoid syneresis. Hetero-cross-linking cereal, milk, and meat biopolymers could create innovative diet yields with specific qualities and properties. Tyrosinase enhances the functionalization of milk-based products by modifying the antioxidation attributes. Crosslinking of casein proteins and oxidation of tyrosyl residues in dairy proteins have previously been identified. Tyrosinases have also been shown to cause partial crosslinking of whey proteins; for example, A. bisporus tyrosinase has been shown to cross-link-lactalbumin (Nawaz et al. 2017).

Meat processing

Crosslinking enzymes are crucial in customizing the gelation qualities of meat, as the ability to shape gels and and the textural and binding properties of meat are all crucial in the production of meat products. Pork and chicken proteins have recently been evaluated for tyrosinase processing. Tyrosinase increases the gel shape quality of a 4 percent chicken breast myofibrillar protein suspension in the presence of 0.35 M NaCl and the insistence of homogenate gels comprising a lower level of phosphate-free beef (Tedeschi et al. 2021; Nawaz et al. 2017).

Other applications

The widespread acceptance of tyrosinase's catalytic property has been attributed to its use in cell culture, which aids nerve cell development. The tyrosinase enzyme is stamped onto plastic sheets, resulting in thin films of melanin in situ. Melanin has a bacteriostatic benefit, so it may help prevent bacterial infection. Entrapment of tyrosinase in polymer, a natural polymer, or transformed polystyrene, or adsorption on nylon zeolite, glass beads, fuller soil, and chitin stimulated with hexamethylenediamine are all used to change L-tyrosine to L-DOPA (Carvalho et al. 2000). Several studies have described the embedding of silk proteins onto chitosan via tyrosinase reactions, indicating that polymers can be tailored. Similarly, L-DOPA has been successfully grafted onto the proteins of wool fibers. It is also used to make hydrogels for skin replacements, medicine transport matrices, and tissue engineering (Mattinen et al. 2008). Tyrosinase was used by Aberg et al. (2004) for biosynthetic splicing of phenolic moieties or protein onto chitosan. The morphology and reusability of the two unique heterogeneous biocatalysts have been studied. Under moderate and environmentally safe laboratory conditions, these biocatalysts have been used to efficiently and selectively synthesize bioactive catechols. Lipophilic catechols are also made with tyrosinase and layer-by-layer assisted tyrosinases. These derivatives with antioxidant activity and long carbon alkyl side chains have shown substantial antiviral activity, indicating the probability of an innovative inhibition contrivance dependent on both redox and lipophilic qualities (Bozzini et al. 2013).

Chen et al. (2002) defined a new method for conjugating gelatin to the polysaccharide chitosan in vitro using tyrosinase. Hydroxytyrosol, a powerful antioxidant found in olives, was also made from tyrosol using tyrosinase. Tyrosinase has been employed as a possible prodrug in treating melanoma patients who have responded well to tyrosinase action. Anghileri et al. (2007) use microbial tyrosinase to make clusters from sericin, a peptide presented in silk textile industry wastewater. EMPA, multidisciplinary research and support organization, recently developed approaches for the recombinant synthesis of bacterial tyrosinase, making biomaterials including crosslinked proteins and melanin, such as recombinant Verrucomicrobium. The protein Spinosum tyrosinase is used to make custom-made melanin and other polyphenolic materials from various phenols and catechols as primary resources. These materials have an extensive range of uses, comprising the manufacture of organic semiconductors and photovoltaics. The tyrosinase enzyme can also be used to make crosslinked proteins, making it possible to recycle enzyme biocatalysts, including lipase (Wang et al. 2021).

Reactive lignin nanocapsules accelerate a pigmenting response to provide a durable, versatile bioink of a new kind. The creation of several pigments/dyes by natural or synthetic method is catalyzed by tyrosinase and laccase. Lignin nanocapsules that work with oxidation enzymes may be a biological catalyst for pigment formation and a newly manufactured pigment as carrier, adhesive, and additive (Polak and Jarosz‐Wilkolazka 2012). Capecchi et al. (2019) identified numerous roles of lignin nanocapsules in the support and enzyme activation needed for pigment production. In addition, lignin nanocapsules were discovered to preserve the melanin pigment from alkaline and UV degradation.

Conclusion and perspectives

Since the first industrial use of immobilized enzymes in the 1960s, enzyme immobilization harmony has advanced rapidly. Currently, the strategy of immobilized enzymes that fit various precise implementations has abandoned the traditional trial-and-error methodology and gradually migrated to a coherent scheme. This has also been distinguished by the fact that enzyme immobilization is now used to anticipate the reuse of expensive enzymes. The presence of a stable immobilized enzyme is now widely recognized as providing prompt perception into method progress and saving costs in process development and manufacturing. One of the most important classes of industrial enzymes is tyrosinase enzymes. These enzymes are widely used in manufacturing processes, including pharmaceuticals, cosmetics, and food. Several studies suggest that this enzyme has a lot of promise in medicine, agriculture, and analytical and environmental applications. The immobilized tyrosinase had a wider pH and temperature operating range and improved reusability and storage flexibility. Furthermore, since the immobilized tyrosinase was stabilized in its active conformation by electrostatic interaction before covalent crosslinking, it had a higher affinity for the substrate. Therefore, sufficient immobilization matrices for tyrosinase immobilization are needed to improve its usability.

Acknowledgements

Consejo Nacional de Ciencia y Tecnología (MX) is thankfully acknowledged for partially supporting this work under Sistema Nacional de Investigadores (SNI) program awarded to Hafiz M. N. Iqbal (CVU: 735340).

Declarations

Conflict of interests

The authors declare no conflicting interests.

Contributor Information

Muhammad Bilal, Email: bilaluaf@hyit.edu.cn.

Hafiz M. N. Iqbal, Email: hafiz.iqbal@tec.mx

References

  1. Abdel-Aziz HM, Heikal YM. Nanosensors for the detection of fertilizers and other agricultural applications. Nanosens Environ Food Agric. 2021;1:157–168. doi: 10.1007/978-3-030-63245-8_7. [DOI] [Google Scholar]
  2. Abdollahi K, Yazdani F, Panahi R. Fabrication of the robust and recyclable tyrosinase-harboring biocatalyst using ethylenediamine functionalized superparamagnetic nanoparticles: nanocarrier characterization and immobilized enzyme properties. J Biol Inorg Chem. 2019;24(7):943–959. doi: 10.1007/s00775-019-01690-1. [DOI] [PubMed] [Google Scholar]
  3. Aberg CM, Chen T, Olumide A, Raghavan SR, Payne GF. Enzymatic grafting of peptides from casein hydrolysate to chitosan. Potential for value-added byproducts from food-processing wastes. J Agric Food Chem. 2004;52(4):788–793. doi: 10.1021/jf034626v. [DOI] [PubMed] [Google Scholar]
  4. Ahmad M, Pan C, Luo Z, Zhu J. A single ZnO nanofiber-based highly sensitive amperometric glucose biosensor. J Phys Chem C. 2010;114(20):9308–9313. doi: 10.1021/jp102505g. [DOI] [Google Scholar]
  5. Al-Rawi UA, Sher F, Hazafa A, Rasheed T, Al-Shara NK, Lima EC, Shanshool J. Catalytic activity of Pt loaded zeolites for hydroisomerization of n-hexane using supercritical CO2. Ind Eng Chem Res. 2020;59(51):22092–22106. doi: 10.1021/acs.iecr.0c05184. [DOI] [Google Scholar]
  6. Amatatongchai M, Sroysee W, Chairam S, Nacapricha D. Simple flow injection for determination of sulfite by amperometric detection using glassy carbon electrode modified with carbon nanotubes–PDDA–gold nanoparticles. Talanta. 2015;133:134–141. doi: 10.1016/j.talanta.2014.07.055. [DOI] [PubMed] [Google Scholar]
  7. Amine A, Mohammadi H, Bourais I, Palleschi G. Enzyme inhibition-based biosensors for food safety and environmental monitoring. Biosens Bioelectron. 2006;21(8):1405–1423. doi: 10.1016/j.bios.2005.07.012. [DOI] [PubMed] [Google Scholar]
  8. Andre RS, Mercante LA, Facure MH, Pavinatto A, Correa DS. Electrospun polymers and composites. Woodhead Publishing; 2021. Electrospun composite nanofibers as sensors for food analysis; pp. 261–286. [Google Scholar]
  9. Anghileri A, Lantto R, Kruus K, Arosio C, Freddi G. Tyrosinase-catalyzed grafting of sericin peptides onto chitosan and production of protein–polysaccharide bioconjugates. J Biotechnol. 2007;127(3):508–519. doi: 10.1016/j.jbiotec.2006.07.021. [DOI] [PubMed] [Google Scholar]
  10. Anh TM, Dzyadevych SV, Van MC, Renault NJ, Duc CN, Chovelon JM. Conductometric tyrosinase biosensor for the detection of diuron, atrazine and its main metabolites. Talanta. 2004;63(2):365–370. doi: 10.1016/j.talanta.2003.11.008. [DOI] [PubMed] [Google Scholar]
  11. Apetrei IM, Apetrei C. The biocomposite screen-printed biosensor based on immobilization of tyrosinase onto the carboxyl functionalised carbon nanotube for assaying tyramine in fish products. J Food Eng. 2015;149:1–8. doi: 10.1016/j.jfoodeng.2014.09.036. [DOI] [Google Scholar]
  12. Arduini F, Amine A, Majorani C, Di Giorgio F, De Felicis D, Cataldo F, Palleschi G. High performance electrochemical sensor based on modified screen-printed electrodes with cost-effective dispersion of nanostructured carbon black. Electrochem Commun. 2010;12(3):346–350. doi: 10.1016/j.elecom.2009.12.028. [DOI] [Google Scholar]
  13. Arduini F, Amine A, Moscone D, Palleschi G. Biosensors based on cholinesterase inhibition for insecticides, nerve agents and aflatoxin B 1 detection. Microchim Acta. 2010;170(3–4):193–214. doi: 10.1007/s00604-010-0317-1. [DOI] [Google Scholar]
  14. Arduini F, Majorani C, Amine A, Moscone D, Palleschi G. Hg2+ detection by measuring thiol groups with a highly sensitive screen-printed electrode modified with a nanostructured carbon black film. Electrochim Acta. 2011;56(11):4209–4215. doi: 10.1016/j.electacta.2011.01.094. [DOI] [Google Scholar]
  15. Arduini F, Di Nardo F, Amine A, Micheli L, Palleschi G, Moscone D. Carbon black-modified screen-printed electrodes as electroanalytical tools. Electroanalysis. 2012;24(4):743–751. doi: 10.1002/elan.201100561. [DOI] [Google Scholar]
  16. Ates S, Cortenlioglu E, Bayraktar E, Mehmetoglu U. Production of L-DOPA using Cu-alginate gel immobilized tyrosinase in a batch and packed bed reactor. Enzyme Microb Technol. 2007;40(4):683–687. doi: 10.1016/j.enzmictec.2006.05.031. [DOI] [Google Scholar]
  17. Baccar ZM, Caballero D, Eritja R, Errachid A. Development of an impedimetric DNA-biosensor based on layered double hydroxide for the detection of long ssDNA sequences. Electrochim Acta. 2012;74:123–129. doi: 10.1016/j.electacta.2012.04.031. [DOI] [Google Scholar]
  18. Baluta S, Lesiak A, Cabaj J. Simple and cost-effective electrochemical method for norepinephrine determination based on carbon dots and tyrosinase. Sensors. 2020;20(16):4567. doi: 10.3390/s20164567. [DOI] [PMC free article] [PubMed] [Google Scholar]
  19. Barbosa O, Torres R, Ortiz C, Berenguer-Murcia A, Rodrigues RC, Fernandez-Lafuente R. Heterofunctional supports in enzyme immobilization: from traditional immobilization protocols to opportunities in tuning enzyme properties. Biomacromol. 2013;14(8):2433–2462. doi: 10.1021/bm400762h. [DOI] [PubMed] [Google Scholar]
  20. Barry S, O'Riordan A. Electrochemical nanosensors: advances and applications. Rep Electrochem. 2016;6:1–14. [Google Scholar]
  21. Basiuk EV, Huerta L, Basiuk VA. Noncovalent bonding of 3d metal (II) phthalocyanines with single-walled carbon nanotubes: a combined DFT and XPS study. Appl Surf Sci. 2019;470:622–630. doi: 10.1016/j.apsusc.2018.11.159. [DOI] [Google Scholar]
  22. Bernal C, Rodriguez K, Martinez R. Integrating enzyme immobilization and protein engineering: an alternative path for the development of novel and improved industrial biocatalysts. Biotechnol Adv. 2018;36(5):1470–1480. doi: 10.1016/j.biotechadv.2018.06.002. [DOI] [PubMed] [Google Scholar]
  23. Bezerra RM, Neto DMA, Galvão WS, Rios NS, Carvalho ACLDM, Correa MA, Gonçalves LR. Design of a lipase-nano particle biocatalysts and its use in the kinetic resolution of medicament precursors. Biochem Eng J. 2017;125:104–115. doi: 10.1016/j.bej.2017.05.024. [DOI] [Google Scholar]
  24. Bilal M, Asgher M, Cheng H, Yan Y, Iqbal HM. Multi-point enzyme immobilization, surface chemistry, and novel platforms: a paradigm shift in biocatalyst design. Crit Rev Biotechnol. 2019;39(2):202–219. doi: 10.1080/07388551.2018.1531822. [DOI] [PubMed] [Google Scholar]
  25. Bizzarri BM, Martini A, Serafini F, Aversa D, Piccinino D, Botta L, Saladino R. Tyrosinase mediated oxidative functionalization in the synthesis of DOPA-derived peptidomimetics with anti-Parkinson activity. RSC Adv. 2017;7(33):20502–20509. doi: 10.1039/C7RA03326E. [DOI] [Google Scholar]
  26. Bommarius AS, Riebel-Bommarius BR. Biocatalysis: fundamentals and applications. John Wiley & Sons; 2004. [Google Scholar]
  27. Bonet-San-Emeterio M, Algarra M, Petković M, Del Valle M. Modification of electrodes with N-and S-doped carbon dots Evaluation of the electrochemical response. Talanta. 2020;212:120806. doi: 10.1016/j.talanta.2020.120806. [DOI] [PubMed] [Google Scholar]
  28. Botta G, Bizzarri BM, Garozzo A, Timpanaro R, Bisignano B, Amatore D, Saladino R. Carbon nanotubes supported tyrosinase in the synthesis of lipophilic hydroxytyrosol and dihydrocaffeoyl catechols with antiviral activity against DNA and RNA viruses. Bioorg Med Chem. 2015;23(17):5345–5351. doi: 10.1016/j.bmc.2015.07.061. [DOI] [PMC free article] [PubMed] [Google Scholar]
  29. Botta L, Bizzarri BM, Crucianelli M, Saladino R. Advances in biotechnological synthetic applications of carbon nanostructured systems. J Mater Chem B. 2017;5(32):6490–6510. doi: 10.1039/C7TB00764G. [DOI] [PubMed] [Google Scholar]
  30. Bozzini T, Botta G, Delfino M, Onofri S, Saladino R, Amatore D, Palamara AT. Tyrosinase and layer-by-layer supported tyrosinases in the synthesis of lipophilic catechols with antiinfluenza activity. Bioorg Med Chem. 2013;21(24):7699–7708. doi: 10.1016/j.bmc.2013.10.026. [DOI] [PubMed] [Google Scholar]
  31. Branzoi V, Branzoi F, Cioaca B, Pilan LN (2010) Highly sensitive amperometric biosensors for phenols based on polyaniline/carbon nanotube composite modified electrodes. In: ECS Meeting Abstracts (No. 42, p. 1827). IOP Publishing.
  32. Bucur B, Munteanu FD, Marty JL, Vasilescu A. Advances in enzyme-based biosensors for pesticide detection. Biosensors. 2018;8(2):27. doi: 10.3390/bios8020027. [DOI] [PMC free article] [PubMed] [Google Scholar]
  33. Camargo JR, Baccarin M, Raymundo-Pereira PA, Campos AM, Oliveira GG, Fatibello-Filho O, Janegitz BC. Electrochemical biosensor made with tyrosinase immobilized in a matrix of nanodiamonds and potato starch for detecting phenolic compounds. Anal Chim Acta. 2018;1034:137–143. doi: 10.1016/j.aca.2018.06.001. [DOI] [PubMed] [Google Scholar]
  34. Capecchi E, Piccinino D, Delfino I, Bollella P, Antiochia R, Saladino R. Functionalized tyrosinase-lignin nanoparticles as sustainable catalysts for the oxidation of phenols. Nanomaterials. 2018;8(6):438. doi: 10.3390/nano8060438. [DOI] [PMC free article] [PubMed] [Google Scholar]
  35. Capecchi E, Piccinino D, Bizzarri BM, Avitabile D, Pelosi C, Colantonio C, Saladino R. Enzyme-Lignin nanocapsules are sustainable catalysts and vehicles for the preparation of unique polyvalent bioinks. Biomacromol. 2019;20(5):1975–1988. doi: 10.1021/acs.biomac.9b00198. [DOI] [PubMed] [Google Scholar]
  36. Carvalho GM, Alves TLM, Freire DM. L-DOPA production by immobilized tyrosinase. Appl Biochem Biotechnol. 2000;84(1):791–800. doi: 10.1385/ABAB:84-86:1-9:791. [DOI] [PubMed] [Google Scholar]
  37. Cerrato-Alvarez M, Bernalte E, Bernalte-García MJ, Pinilla-Gil E. Fast and direct amperometric analysis of polyphenols in beers using tyrosinase-modified screen-printed gold nanoparticles biosensors. Talanta. 2019;193:93–99. doi: 10.1016/j.talanta.2018.09.093. [DOI] [PubMed] [Google Scholar]
  38. Chaimuangyong S, Changkawprom M, Preechaworapun A, Tangkuaram T (2019) Tyrosinase immobilized on poly (diallyldimethylammonium chloride) capped gold nanoparticles composite with carbon nanotubes as dopamine biosensor.
  39. Chen J, Jin Y. Sensitive phenol determination based on co-modifying tyrosinase and palygorskite on glassy carbon electrode. Microchim Acta. 2010;169(3–4):249–254. doi: 10.1007/s00604-010-0320-6. [DOI] [Google Scholar]
  40. Chen T, Embree HD, Wu LQ, Payne GF. In vitro protein–polysaccharide conjugation: Tyrosinase-catalyzed conjugation of gelatin and chitosan. Biopolymers. 2002;64(6):292–302. doi: 10.1002/bip.10196. [DOI] [PubMed] [Google Scholar]
  41. Chen M, Qin X, Zeng G. Biodegradation of carbon nanotubes, graphene, and their derivatives. Trends Biotechnol. 2017;35(9):836–846. doi: 10.1016/j.tibtech.2016.12.001. [DOI] [PubMed] [Google Scholar]
  42. Chen M, Zeng G, Xu P, Lai C, Tang L. How do enzymes ‘meet’nanoparticles and nanomaterials? Trends Biochem Sci. 2017;42(11):914–930. doi: 10.1016/j.tibs.2017.08.008. [DOI] [PubMed] [Google Scholar]
  43. Chen Y, Zheng G, Shi Q, Zhao R, Chen M. Preparation of thiolated calix [8] arene/AuNPs/MWCNTs modified glassy carbon electrode and its electrocatalytic oxidation toward paracetamol. Sens Actuators, B Chem. 2018;277:289–296. doi: 10.1016/j.snb.2018.09.012. [DOI] [Google Scholar]
  44. Cho IH, Kim DH, Park S. Electrochemical biosensors: perspective on functional nanomaterials for on-site analysis. Biomaterials Research. 2020;24(1):1–12. doi: 10.1186/s40824-019-0181-y. [DOI] [PMC free article] [PubMed] [Google Scholar]
  45. Choi JM, Han SS, Kim HS. Industrial applications of enzyme biocatalysis: Current status and future aspects. Biotechnol Adv. 2015;33(7):1443–1454. doi: 10.1016/j.biotechadv.2015.02.014. [DOI] [PubMed] [Google Scholar]
  46. Chung Y, Christwardana M, Tannia DC, Kim KJ, Kwon Y. Biocatalyst including porous enzyme cluster composite immobilized by two-step crosslinking and its utilization as enzymatic biofuel cell. J Power Sources. 2017;360:172–179. doi: 10.1016/j.jpowsour.2017.06.012. [DOI] [Google Scholar]
  47. Cortina-Puig M, Muñoz-Berbel X, Calas-Blanchard C, Marty JL. Diazonium-functionalized tyrosinase-based biosensor for the detection of tea polyphenols. Microchim Acta. 2010;171(1–2):187–193. doi: 10.1007/s00604-010-0425-y. [DOI] [Google Scholar]
  48. Cunha MR, Lima EC, Lima DR, da Silva RS, Thue PS, Seliem MK, Larsson SH. Removal of captopril pharmaceutical from synthetic pharmaceutical-industry wastewaters: use of activated carbon derived from Butia catarinensis. J Environ Chem Eng. 2020;8(6):104506. doi: 10.1016/j.jece.2020.104506. [DOI] [Google Scholar]
  49. Das R, Mishra H, Srivastava A, Kayastha AM. Covalent immobilization of β-amylase onto functionalized molybdenum sulfide nanosheets, its kinetics and stability studies: a gateway to boost enzyme application. Chem Eng J. 2017;328:215–227. doi: 10.1016/j.cej.2017.07.019. [DOI] [Google Scholar]
  50. Das R, Talat M, Srivastava ON, Kayastha AM. Covalent immobilization of peanut β-amylase for producing industrial nano-biocatalysts: a comparative study of kinetics, stability and reusability of the immobilized enzyme. Food Chem. 2018;245:488–499. doi: 10.1016/j.foodchem.2017.10.092. [DOI] [PubMed] [Google Scholar]
  51. Deng P, Fei J, Zhang J, Feng Y. Determination of trace aluminum by anodic adsorptive stripping voltammetry using a multi-walled carbon nanotube modified carbon paste electrode. Anal Lett. 2011;44(8):1521–1535. doi: 10.1080/00032719.2010.520382. [DOI] [Google Scholar]
  52. Dinçer A, Becerik S, Aydemir T. Immobilization of tyrosinase on chitosan–clay composite beads. Int J Biol Macromol. 2012;50(3):815–820. doi: 10.1016/j.ijbiomac.2011.11.020. [DOI] [PubMed] [Google Scholar]
  53. Ding J, Gu L, Dong W, Yu H. Epoxidation modification of renewable lignin to improve the corrosion performance of epoxy coating. Int J Electrochem Sci. 2016;11(7):6256–6265. doi: 10.20964/2016.07.56. [DOI] [Google Scholar]
  54. Ding X, Niu Y, Zhang G, Xu Y, Li J. Electrochemistry in carbon-based quantum dots. Chemistry. 2020;15(8):1214–1224. doi: 10.1002/asia.202000097. [DOI] [PubMed] [Google Scholar]
  55. dos Santosa VP, Silvaa LM, Salgadoa AM, Pereirab KS. Application of Agaricus bisporus extract for benzoate sodium detection based on tyrosinase inhibition for biosensor development. Chem Eng. 2013;32:2. [Google Scholar]
  56. Du H, Ye J, Zhang J, Huang X, Yu C. A voltammetric sensor based on graphene-modified electrode for simultaneous determination of catechol and hydroquinone. J Electroanal Chem. 2011;650(2):209–213. doi: 10.1016/j.jelechem.2010.10.002. [DOI] [Google Scholar]
  57. Du Y, Gao J, Zhou L, Ma L, He Y, Huang Z, Jiang Y. Enzyme nanocapsules armored by metal-organic frameworks: A novel approach for preparing nanobiocatalyst. Chem Eng J. 2017;327:1192–1197. doi: 10.1016/j.cej.2017.07.021. [DOI] [Google Scholar]
  58. Etemadi F, Hashemi M, Randhir R, ZandVakili O, Ebadi A. Accumulation of L-DOPA in various organs of faba bean and influence of drought, nitrogen stress, and processing methods on L-DOPA yield. Crop J. 2018;6(4):426–434. doi: 10.1016/j.cj.2017.12.001. [DOI] [Google Scholar]
  59. Feng J, Yu S, Li J, Mo T, Li P. Enhancement of the catalytic activity and stability of immobilized aminoacylase using modified magnetic Fe3O4 nanoparticles. Chem Eng J. 2016;286:216–222. doi: 10.1016/j.cej.2015.10.083. [DOI] [Google Scholar]
  60. Fujieda N, Umakoshi K, Ochi Y, Nishikawa Y, Yanagisawa S, Kubo M, Itoh S. Copper-oxygen dynamics in the tyrosinase mechanism. Angew Chem. 2020;132(32):13487–13492. doi: 10.1002/ange.202004733. [DOI] [PubMed] [Google Scholar]
  61. Glusac J, Davidesko-Vardi I, Isaschar-Ovdat S, Kukavica B, Fishman A. Gel-like emulsions stabilized by tyrosinase-crosslinked potato and zein proteins. Food Hydrocolloids. 2018;82:53–63. doi: 10.1016/j.foodhyd.2018.03.046. [DOI] [Google Scholar]
  62. Gradilone A, Cigna E, Agliano AM, Frati L. Tyrosinase expression as a molecular marker for investigating the presence of circulating tumor cells in melanoma patients. Curr Cancer Drug Targets. 2010;10(5):529–538. doi: 10.2174/156800910791517136. [DOI] [PubMed] [Google Scholar]
  63. Gu BX, Xu CX, Zhu GP, Liu SQ, Chen LY, Li XS. Tyrosinase immobilization on ZnO nanorods for phenol detection. J Phys Chem B. 2009;113(1):377–381. doi: 10.1021/jp808001c. [DOI] [PubMed] [Google Scholar]
  64. Guazzaroni M, Pasqualini M, Botta G, Saladino R. A novel synthesis of bioactive catechols by layer-by-layer immobilized tyrosinase in an organic solvent medium. ChemCatChem. 2012;4(1):89–99. doi: 10.1002/cctc.201100229. [DOI] [Google Scholar]
  65. Gupta MN, Kaloti M, Kapoor M, Solanki K. Nanomaterials as matrices for enzyme immobilization. Artif Cells Blood Subst Biotechnol. 2011;39(2):98–109. doi: 10.3109/10731199.2010.516259. [DOI] [PubMed] [Google Scholar]
  66. Gupta S, Murthy CN, Prabha CR. Recent advances in carbon nanotube based electrochemical biosensors. Int J Biol Macromol. 2018;108:687–703. doi: 10.1016/j.ijbiomac.2017.12.038. [DOI] [PubMed] [Google Scholar]
  67. Han E, Shan D, Xue H, Cosnier S. Hybrid material based on chitosan and layered double hydroxides: characterization and application to the design of amperometric phenol biosensor. Biomacromol. 2007;8(3):971–975. doi: 10.1021/bm060897d. [DOI] [PubMed] [Google Scholar]
  68. Han J, Xu X, Rao X, Wei M, Evans DG, Duan X. Layer-by-layer assembly of layered double hydroxide/cobalt phthalocyanine ultrathin film and its application for sensors. J Mater Chem. 2011;21(7):2126–2130. doi: 10.1039/C0JM02430A. [DOI] [Google Scholar]
  69. Han R, Cui L, Ai S, Yin H, Liu X, Qiu Y. Amperometric biosensor based on tyrosinase immobilized in hydrotalcite-like compounds film for the determination of polyphenols. J Solid State Electrochem. 2012;16(2):449–456. doi: 10.1007/s10008-011-1352-5. [DOI] [Google Scholar]
  70. Hidouri S, Baccar ZM, Abdelmelek H, Noguer T, Marty JL, Campàs M. Structural and functional characterisation of a biohybrid material based on acetylcholinesterase and layered double hydroxides. Talanta. 2011;85(4):1882–1887. doi: 10.1016/j.talanta.2011.07.026. [DOI] [PubMed] [Google Scholar]
  71. Hiremath SD, Priyadarshi B, Banerjee M, Chatterjee A. Carbon dots-MnO2 based turn-on fluorescent probe for rapid and sensitive detection of hydrazine in water. J Photochem Photobiol A Chem. 2020;389:112258. doi: 10.1016/j.jphotochem.2019.112258. [DOI] [Google Scholar]
  72. Husain Q. High yield immobilization and stabilization of oxidoreductases using magnetic nanosupports and their potential applications: an update. Curr Catal. 2017;6(3):168–187. doi: 10.2174/2211544706666170704141828. [DOI] [Google Scholar]
  73. Ibáñez-Redín G, Silva TA, Vicentini FC, Fatibello-Filho O. Effect of carbon black functionalization on the analytical performance of a tyrosinase biosensor based on glassy carbon electrode modified with dihexadecylphosphate film. Enzyme Microb Technol. 2018;116:41–47. doi: 10.1016/j.enzmictec.2018.05.007. [DOI] [PubMed] [Google Scholar]
  74. Ilesanmi OS, Adewale IO. Physicochemical properties of free and immobilized tyrosinase from different species of yam (Dioscorea spp) Biotechnol Rep. 2020;27:e00499. doi: 10.1016/j.btre.2020.e00499. [DOI] [PMC free article] [PubMed] [Google Scholar]
  75. Jankowska K, Zdarta J, Grzywaczyk A, Kijeńska-Gawrońska E, Biadasz A, Jesionowski T. Electrospun poly (methyl methacrylate)/polyaniline fibres as a support for laccase immobilisation and use in dye decolourisation. Environ Res. 2020;184:109332. doi: 10.1016/j.envres.2020.109332. [DOI] [PubMed] [Google Scholar]
  76. Jawaid S, Khan TH, Osborn HM, Williams NAO. Tyrosinase activated melanoma prodrugs. Anti-Cancer Agents Med Chem. 2009;9(7):717–727. doi: 10.2174/187152009789056886. [DOI] [PubMed] [Google Scholar]
  77. Kaur R, Uppal SK. Structural characterization and antioxidant activity of lignin from sugarcane bagasse. Colloid Polym Sci. 2015;293(9):2585–2592. doi: 10.1007/s00396-015-3653-1. [DOI] [Google Scholar]
  78. Khan M, Husain Q, Bushra R. Immobilization of β-galactosidase on surface modified cobalt/multiwalled carbon nanotube nanocomposite improves enzyme stability and resistance to inhibitor. Int J Biol Macromol. 2017;105:693–701. doi: 10.1016/j.ijbiomac.2017.07.088. [DOI] [PubMed] [Google Scholar]
  79. Khoshnevisan K, Vakhshiteh F, Barkhi M, Baharifar H, Poor-Akbar E, Zari N, Bordbar AK. Immobilization of cellulase enzyme onto magnetic nanoparticles: applications and recent advances. Mol Catal. 2017;442:66–73. doi: 10.1016/j.mcat.2017.09.006. [DOI] [Google Scholar]
  80. Kim KI, Lee JC, Robards K, Choi SH. Immobilization of tyrosinase in carboxylic and carbonyl group-modified MWNT electrode and its application for sensing phenolics in red wines. J Nanosci Nanotechnol. 2010;10(6):3790–3798. doi: 10.1166/jnn.2010.1982. [DOI] [PubMed] [Google Scholar]
  81. Kochana J, Wapiennik K, Kozak J, Knihnicki P, Pollap A, Woźniakiewicz M, Kościelniak P. Tyrosinase-based biosensor for determination of bisphenol A in a flow-batch system. Talanta. 2015;144:163–170. doi: 10.1016/j.talanta.2015.05.078. [DOI] [PubMed] [Google Scholar]
  82. Kuilla T, Bhadra S, Yao D, Kim NH. Saswata Bose, Joong Hee Lee. Prog Polym Sci. 2010;35:1350. doi: 10.1016/j.progpolymsci.2010.07.005. [DOI] [Google Scholar]
  83. Laranjo MT, Morawski FM, Dias SL, Benvenutti EV, Arenas LT, Costa TM. Silica/titania graphite composite modified with chitosan and tyrosinase employed as a sensitive biosensor for phenolic compounds. J Braz Chem Soc. 2019;30(12):2660–2672. [Google Scholar]
  84. Leskinen T, Witos J, Valle-Delgado JJ, Lintinen K, Kostiainen M, Wiedmer SK, Mattinen ML. Adsorption of proteins on colloidal lignin particles for advanced biomaterials. Biomacromol. 2017;18(9):2767–2776. doi: 10.1021/acs.biomac.7b00676. [DOI] [PubMed] [Google Scholar]
  85. Li X, Zhang G, Bai X, Sun X, Wang X, Wang E, Dai H. Highly conducting graphene sheets and Langmuir-Blodgett films. Nat Nanotechnol. 2008;3(9):538–542. doi: 10.1038/nnano.2008.210. [DOI] [PubMed] [Google Scholar]
  86. Li Y, Schluesener HJ, Xu S. Gold nanoparticle-based biosensors. Gold Bull. 2010;43(1):29–41. doi: 10.1007/BF03214964. [DOI] [Google Scholar]
  87. Li T, Shen X, Chen Y, Zhang C, Yan J, Yang H, Liu Y. Polyetherimide-grafted Fe3O4@ SiO2 nanoparticles as theranostic agents for simultaneous VEGF siRNA delivery and magnetic resonance cell imaging. Int J Nanomed. 2015;10:4279. doi: 10.2147/IJN.S85095. [DOI] [PMC free article] [PubMed] [Google Scholar]
  88. Li Y, Abedalwafa MA, Tang L, Wang L. Electrospinning: Nanofabrication and Applications. William Andrew Publishing; 2019. Electrospun nanofibers for sensors; pp. 571–601. [Google Scholar]
  89. Liang K, Fu X, Wu L, Qin Y, Song Y. A novel tyrosinase biosensor based on graphene and Co3O4 nanocomposite materials for rapid determining catechol. Int J Electrochem Sci. 2016;11:250–258. [Google Scholar]
  90. Liang S, Wu XL, Xiong J, Zong MH, Lou WY. Metal-organic frameworks as novel matrices for efficient enzyme immobilization: an update review. Coord Chem Rev. 2020;406:213149. doi: 10.1016/j.ccr.2019.213149. [DOI] [Google Scholar]
  91. Liu B, Zeng HC. Hydrothermal synthesis of ZnO nanorods in the diameter regime of 50 nm. J Am Chem Soc. 2003;125(15):4430–4431. doi: 10.1021/ja0299452. [DOI] [PubMed] [Google Scholar]
  92. Liu Z, Tang HL, Pan M. Synthesis and characterization of PDDA-stabilized Pt nanoparticles for direct methanol fuel cells. Electrochim Acta. 2006;51(26):5721–5730. doi: 10.1016/j.electacta.2006.03.004. [DOI] [Google Scholar]
  93. Liu T, Xu M, Yin H, Ai S, Qu X, Zong S. A glassy carbon electrode modified with graphene and tyrosinase immobilized on platinum nanoparticles for sensing organophosphorus pesticides. Microchim Acta. 2011;175(1):129–135. doi: 10.1007/s00604-011-0665-5. [DOI] [Google Scholar]
  94. Liu S, Höldrich M, Sievers-Engler A, Horak J, Lämmerhofer M. Papain-functionalized gold nanoparticles as heterogeneous biocatalyst for bioanalysis and biopharmaceuticals analysis. Anal Chim Acta. 2017;963:33–43. doi: 10.1016/j.aca.2017.02.009. [DOI] [PubMed] [Google Scholar]
  95. Liu DM, Chen J, Shi YP. Tyrosinase immobilization on aminated magnetic nanoparticles by physical adsorption combined with covalent crosslinking with improved catalytic activity, reusability and storage stability. Anal Chim Acta. 2018;1006:90–98. doi: 10.1016/j.aca.2017.12.022. [DOI] [PubMed] [Google Scholar]
  96. Love JC, Estroff LA, Kriebel JK, Nuzzo RG, Whitesides GM. Self-assembled monolayers of thiolates on metals as a form of nanotechnology. Chem Rev. 2005;105(4):1103–1170. doi: 10.1021/cr0300789. [DOI] [PubMed] [Google Scholar]
  97. Madhavan A, Sindhu R, Binod P, Sukumaran RK, Pandey A. Strategies for design of improved biocatalysts for industrial applications. Biores Technol. 2017;245:1304–1313. doi: 10.1016/j.biortech.2017.05.031. [DOI] [PubMed] [Google Scholar]
  98. Makhotkina O, Kilmartin PA. The use of cyclic voltammetry for wine analysis: determination of polyphenols and free sulfur dioxide. Anal Chim Acta. 2010;668(2):155–165. doi: 10.1016/j.aca.2010.03.064. [DOI] [PubMed] [Google Scholar]
  99. Martín M, Salazar P, Villalonga R, Campuzano S, Pingarrón JM, González-Mora JL. Preparation of core–shell Fe 3 O 4@ poly (dopamine) magnetic nanoparticles for biosensor construction. J Mater Chem B. 2014;2(6):739–746. doi: 10.1039/C3TB21171A. [DOI] [PubMed] [Google Scholar]
  100. Martinazzo J, Muenchen DK, Brezolin AN, Cezaro AM, Rigo AA, Manzoli A, Steffens C. Cantilever nanobiosensor using tyrosinase to detect atrazine in liquid medium. J Environ Sci Health B. 2018;53(4):229–236. doi: 10.1080/03601234.2017.1421833. [DOI] [PubMed] [Google Scholar]
  101. Martinez-Alvarez O, Miranda-Hernandez M. Characterisation of carbon pastes as matrices in composite electrodes for use in electrochemical capacitors. Carbon-Sci Tech. 2008;1:30–38. [Google Scholar]
  102. Massolini G, Calleri E. Immobilized trypsin systems coupled on-line to separation methods: Recent developments and analytical applications. J Sep Sci. 2005;28(1):7–21. doi: 10.1002/jssc.200401941. [DOI] [PubMed] [Google Scholar]
  103. Matoba Y, Kihara S, Bando N, Yoshitsu H, Sakaguchi M, Kayama KE, Sugiyama M. Catalytic mechanism of the tyrosinase reaction toward the Tyr98 residue in the caddie protein. PLoS Biol. 2018;16(12):e3000077. doi: 10.1371/journal.pbio.3000077. [DOI] [PMC free article] [PubMed] [Google Scholar]
  104. Mattinen ML, Lantto R, Selinheimo E, Kruus K, Buchert J. Oxidation of peptides and proteins by Trichoderma reesei and Agaricus bisporus tyrosinases. J Biotechnol. 2008;133(3):395–402. doi: 10.1016/j.jbiotec.2007.10.009. [DOI] [PubMed] [Google Scholar]
  105. Mercante LA, Scagion VP, Migliorini FL, Mattoso LH, Correa DS. Electrospinning-based (bio) sensors for food and agricultural applications: A review. TrAC, Trends Anal Chem. 2017;91:91–103. doi: 10.1016/j.trac.2017.04.004. [DOI] [Google Scholar]
  106. Mercante LA, Iwaki LE, Scagion VP, Oliveira ON, Mattoso LH, Correa DS. Electrochemical detection of bisphenol a by tyrosinase immobilized on electrospun nanofibers decorated with gold nanoparticles. Electrochem. 2021;2(1):41–49. doi: 10.3390/electrochem2010004. [DOI] [Google Scholar]
  107. Min K, Park GW, Yoo YJ, Lee JS. A perspective on the biotechnological applications of the versatile tyrosinase. Bioresour Technol. 2019;289:121730. doi: 10.1016/j.biortech.2019.121730. [DOI] [PubMed] [Google Scholar]
  108. Mishra SK, Chiang KS. Phenolic-compounds sensor based on immobilization of tyrosinase in polyacrylamide gel on long-period fiber grating. Opt Laser Technol. 2020;131:106464. doi: 10.1016/j.optlastec.2020.106464. [DOI] [Google Scholar]
  109. Mohammadi A, Moghaddam AB, Hosseini S, Kazemzad M, Dinarvand R. A norepinephrine biosensor based on a glassy carbon electrode modified with carbon nanotubes. Anal Methods. 2011;3(10):2406–2411. doi: 10.1039/c1ay05291h. [DOI] [Google Scholar]
  110. Monogioudi E, Faccio G, Lille M, Poutanen K, Buchert J, Mattinen ML. Effect of enzymatic crosslinking of β-casein on proteolysis by pepsin. Food Hydrocolloids. 2011;25(1):71–81. doi: 10.1016/j.foodhyd.2010.05.007. [DOI] [Google Scholar]
  111. Montereali MR, Della Seta L, Vastarella W, Pilloton R. A disposable laccase-tyrosinase based biosensor for amperometric detection of phenolic compounds in must and wine. J Mol Catal B Enzym. 2010;64(3–4):189–194. doi: 10.1016/j.molcatb.2009.07.014. [DOI] [Google Scholar]
  112. Mphuthi NG, Adekunle AS, Ebenso EE. Electrocatalytic oxidation of epinephrine and norepinephrine at metal oxide doped phthalocyanine/MWCNT composite sensor. Sci Rep. 2016;6(1):1–20. doi: 10.1038/srep26938. [DOI] [PMC free article] [PubMed] [Google Scholar]
  113. Nadifiyine S, Haddam M, Mandli J, Chadel S, Blanchard CC, Marty JL, Amine A. Amperometric biosensor based on tyrosinase immobilized on to a carbon black paste electrode for phenol determination in olive oil. Anal Lett. 2013;46(17):2705–2726. doi: 10.1080/00032719.2013.811679. [DOI] [Google Scholar]
  114. Nawaz A, Shafi T, Khaliq A, Mukht H. Tyrosinase: sources, structure and applications. Int J Biotech Bioeng. 2017;3(5):142–148. [Google Scholar]
  115. Notsu H, Tatsuma T, Fujishima A. Tyrosinase-modified boron-doped diamond electrodes for the determination of phenol derivatives. J Electroanal Chem. 2002;523(1–2):86–92. doi: 10.1016/S0022-0728(02)00733-7. [DOI] [Google Scholar]
  116. Orfanakis G, Patila M, Catzikonstantinou AV, Lyra KM, Kouloumpis A, Spyrou K, Stamatis H. Hybrid nanomaterials of magnetic iron nanoparticles and graphene oxide as matrices for the immobilization of β-glucosidase: synthesis, characterization, and biocatalytic properties. Front Mater. 2018;5:25. doi: 10.3389/fmats.2018.00025. [DOI] [Google Scholar]
  117. Osma JF, Toca-Herrera JL, Rodríguez-Couto S. Uses of laccases in the food industry. Enzyme Res. 2010;2010(2):2. doi: 10.4061/2010/918761. [DOI] [PMC free article] [PubMed] [Google Scholar]
  118. Pillaiyar T, Namasivayam V, Manickam M, Jung SH. Inhibitors of melanogenesis: an updated review. J Med Chem. 2018;61(17):7395–7418. doi: 10.1021/acs.jmedchem.7b00967. [DOI] [PubMed] [Google Scholar]
  119. Polak J, Jarosz-Wilkolazka A. Structure/Redox potential relationship of simple organic compounds as potential precursors of dyes for laccase-mediated transformation. Biotechnol Prog. 2012;28(1):93–102. doi: 10.1002/btpr.713. [DOI] [PubMed] [Google Scholar]
  120. Priyadarshini E, Rawat K, Bohidar HB. Multimode sensing of riboflavin via Ag@ carbon dot conjugates. Appl Nanosci. 2020;10(1):281–291. doi: 10.1007/s13204-019-01090-6. [DOI] [Google Scholar]
  121. Ragauskas AJ, Beckham GT, Biddy MJ, Chandra R, Chen F, Davis MF, Dixon R. Valorization: improving lignin processing in the biorefinery. Science. 2014;344(6185):1246843. doi: 10.1126/science.1246843. [DOI] [PubMed] [Google Scholar]
  122. Razumiene J, Barkauskas J, Kubilius V, Meškys R, Laurinavičius V. Modified graphitized carbon black as transducing material for reagentless H2O2 and enzyme sensors. Talanta. 2005;67(4):783–790. doi: 10.1016/j.talanta.2005.04.004. [DOI] [PubMed] [Google Scholar]
  123. Reza KK, Ali MA, Srivastava S, Agrawal VV, Biradar AM. Tyrosinase conjugated reduced graphene oxide based biointerface for bisphenol A sensor. Biosens Bioelectron. 2015;74:644–651. doi: 10.1016/j.bios.2015.07.020. [DOI] [PubMed] [Google Scholar]
  124. Rodríguez Méndez ML, Apetrei IM, Apretei C, Saja Sáez JAD (2013) Enzyme sensor based on carbon nanotubes/cobalt (II) phthalocyanine and tyrosinase used in pharmaceutical analysis
  125. Rolff M, Schottenheim J, Decker H, Tuczek F. Copper–O 2 reactivity of tyrosinase models towards external monophenolic substrates: molecular mechanism and comparison with the enzyme. Chem Soc Rev. 2011;40(7):4077–4098. doi: 10.1039/c0cs00202j. [DOI] [PubMed] [Google Scholar]
  126. Samdani KJ, Samdani JS, Kim NH, Lee JH. FeMoO4 based, enzyme-free electrochemical biosensor for ultrasensitive detection of norepinephrine. Biosens Bioelectron. 2016;81:445–453. doi: 10.1016/j.bios.2016.03.029. [DOI] [PubMed] [Google Scholar]
  127. Sanz VC, Mena ML, González-Cortés A, Yanez-Sedeno P, Pingarrón JM. Development of a tyrosinase biosensor based on gold nanoparticles-modified glassy carbon electrodes: Application to the measurement of a bioelectrochemical polyphenols index in wines. Anal Chim Acta. 2005;528(1):1–8. doi: 10.1016/j.aca.2004.10.007. [DOI] [Google Scholar]
  128. Schroeder V, Savagatrup S, He M, Lin S, Swager TM. Carbon nanotube chemical sensors. Chem Rev. 2018;119(1):599–663. doi: 10.1021/acs.chemrev.8b00340. [DOI] [PMC free article] [PubMed] [Google Scholar]
  129. Shan D, Cosnier S, Mousty C. Layered double hydroxides: an attractive material for electrochemical biosensor design. Anal Chem. 2003;75(15):3872–3879. doi: 10.1021/ac030030v. [DOI] [PubMed] [Google Scholar]
  130. Shan C, Yang H, Han D, Zhang Q, Ivaska A, Niu L. Electrochemical determination of NADH and ethanol based on ionic liquid-functionalized graphene. Biosens Bioelectron. 2010;25(6):1504–1508. doi: 10.1016/j.bios.2009.11.009. [DOI] [PubMed] [Google Scholar]
  131. Sharifi M, Sohrabi MJ, Hosseinali SH, Hasan A, Kani PH, Talaei AJ, Falahati M. Enzyme immobilization onto the nanomaterials: application in enzyme stability and prodrug-activated cancer therapy. Int J Biol Macromol. 2020;143:665–676. doi: 10.1016/j.ijbiomac.2019.12.064. [DOI] [PubMed] [Google Scholar]
  132. Shi D, Yang H, Ji S, Jiang S, Liu X, Zhang D. Preparation and characterization of core-shell structure Fe3O4@ C magnetic nanoparticles. Proc Eng. 2015;102:1555–1562. doi: 10.1016/j.proeng.2015.01.291. [DOI] [Google Scholar]
  133. Silva TA, Moraes FC, Janegitz BC, Fatibello-Filho O (2017) Electrochemical biosensors based on nanostructured carbon black: a review. J Nanomater
  134. Soussou A, Gammoudi I, Moroté F, Kalboussi A, Cohen-Bouhacina T, Grauby-Heywang C, Baccar ZM. Efficient immobilization of tyrosinase enzyme on layered double hydroxide hybrid nanomaterials for electrochemical detection of polyphenols. IEEE Sens J. 2017;17(14):4340–4348. doi: 10.1109/JSEN.2017.2709342. [DOI] [Google Scholar]
  135. Strobel KL, Pfeiffer KA, Blanch HW, Clark DS. Structural insights into the affinity of Cel7A carbohydrate-binding module for lignin. J Biol Chem. 2015;290(37):22818–22826. doi: 10.1074/jbc.M115.673467. [DOI] [PMC free article] [PubMed] [Google Scholar]
  136. Sun X, Hu S, Li L, Xiang J, Sun W. Sensitive electrochemical detection of hydroquinone with carbon ionogel electrode based on BMIMPF6. J Electroanal Chem. 2011;651(1):94–99. doi: 10.1016/j.jelechem.2010.10.019. [DOI] [Google Scholar]
  137. Tacias-Pascacio VG, Virgen-Ortíz JJ, Jiménez-Pérez M, Yates M, Torrestiana-Sanchez B, Rosales-Quintero A, Fernandez-Lafuente R. Evaluation of different lipase biocatalysts in the production of biodiesel from used cooking oil: critical role of the immobilization support. Fuel. 2017;200:1–10. doi: 10.1016/j.fuel.2017.03.054. [DOI] [Google Scholar]
  138. Talarico D, Arduini F, Constantino A, Del Carlo M, Compagnone D, Moscone D, Palleschi G. Carbon black as successful screen-printed electrode modifier for phenolic compound detection. Electrochem Commun. 2015;60:78–82. doi: 10.1016/j.elecom.2015.08.010. [DOI] [Google Scholar]
  139. Tan D, Zhao JP, Ran GQ, Zhu XL, Ding Y, Lu XY. Highly efficient biocatalytic synthesis of l-DOPA using in situ immobilized V errucomicrobium spinosum tyrosinase on polyhydroxyalkanoate nano-granules. Appl Microbiol Biotechnol. 2019;103(14):5663–5678. doi: 10.1007/s00253-019-09851-7. [DOI] [PubMed] [Google Scholar]
  140. Tang Y, Huang R, Liu C, Yang S, Lu Z, Luo S. Electrochemical detection of 4-nitrophenol based on a glassy carbon electrode modified with a reduced graphene oxide/Au nanoparticle composite. Anal Methods. 2013;5(20):5508–5514. doi: 10.1039/c3ay40742j. [DOI] [Google Scholar]
  141. Tang R, Shi Y, Hou Z, Wei L. Carbon nanotube-based chemiresistive sensors. Sensors. 2017;17(4):882. doi: 10.3390/s17040882. [DOI] [PMC free article] [PubMed] [Google Scholar]
  142. Taratula O, Galoppini E, Wang D, Chu D, Zhang Z, Chen H, Lu Y. Binding studies of molecular linkers to ZnO and MgZnO nanotip films. J Phys Chem B. 2006;110(13):6506–6515. doi: 10.1021/jp0570317. [DOI] [PubMed] [Google Scholar]
  143. Tedeschi T, Anzani C, Ferri M, Marzocchi S, Caboni MF, Monari S, Tassoni A. Enzymatic digestion of calf fleshing meat by-products: antioxidant and anti-tyrosinase activity of protein hydrolysates, and identification of fatty acids. Foods. 2021;10(4):755. doi: 10.3390/foods10040755. [DOI] [PMC free article] [PubMed] [Google Scholar]
  144. Temani M, Baccar ZM, Mansour HB. Activity of cholesterol oxidase immobilized on layered double hydroxide nanomaterials for biosensor application: Acacia salicina scavenging power of hypercholesterolemia therapy. Microelectron Eng. 2014;126:165–168. doi: 10.1016/j.mee.2014.07.009. [DOI] [Google Scholar]
  145. Tembe S, Inamdar S, Haram S, Karve M, D'Souza SF. Electrochemical biosensor for catechol using agarose–guar gum entrapped tyrosinase. J Biotechnol. 2007;128(1):80–85. doi: 10.1016/j.jbiotec.2006.09.020. [DOI] [PubMed] [Google Scholar]
  146. Tonelli D, Scavetta E, Giorgetti M. Layered-double-hydroxide-modified electrodes: electroanalytical applications. Anal Bioanal Chem. 2013;405(2):603–614. doi: 10.1007/s00216-012-6586-2. [DOI] [PubMed] [Google Scholar]
  147. Topoglidis E, Cass AE, O'Regan B, Durrant JR. Immobilisation and bioelectrochemistry of proteins on nanoporous TiO2 and ZnO films. J Electroanal Chem. 2001;517(1–2):20–27. doi: 10.1016/S0022-0728(01)00673-8. [DOI] [Google Scholar]
  148. Tucci M, Bombelli P, Howe CJ, Vignolini S, Bocchi S, Schievano A. A storable mediatorless electrochemical biosensor for herbicide detection. Microorganisms. 2019;7(12):630. doi: 10.3390/microorganisms7120630. [DOI] [PMC free article] [PubMed] [Google Scholar]
  149. Vaz FC, Silva TA, Fatibello-Filho O, Assumpção MH, Vicentini FC. A novel carbon nanosphere-based sensor used for herbicide detection. Environ Technol Innov. 2021;2:101529. doi: 10.1016/j.eti.2021.101529. [DOI] [Google Scholar]
  150. Velte CJ, Steinhilper R (2016) Complexity in a circular economy: a need for rethinking complexity management strategies. In Proceedings of the World Congress on Engineering, London, UK (Vol. 29).
  151. Wahba MI. Porous chitosan beads of superior mechanical properties for the covalent immobilization of enzymes. Int J Biol Macromol. 2017;105:894–904. doi: 10.1016/j.ijbiomac.2017.07.102. [DOI] [PubMed] [Google Scholar]
  152. Wang F, Hu S. Electrochemical sensors based on metal and semiconductor nanoparticles. Microchim Acta. 2009;165(1–2):1–22. doi: 10.1007/s00604-009-0136-4. [DOI] [Google Scholar]
  153. Wang B, Zhang J, Dong S. Silica sol–gel composite film as an encapsulation matrix for the construction of an amperometric tyrosinase-based biosensor. Biosens Bioelectron. 2000;15(7–8):397–402. doi: 10.1016/S0956-5663(00)00096-8. [DOI] [PubMed] [Google Scholar]
  154. Wang YS, Tian SP, Xu Y. Effects of high oxygen concentration on pro-and anti-oxidant enzymes in peach fruits during postharvest periods. Food Chem. 2005;91(1):99–104. doi: 10.1016/j.foodchem.2004.05.053. [DOI] [Google Scholar]
  155. Wang JX, Sun XW, Wei A, Lei Y, Cai XP, Li CM, Dong ZL. Zinc oxide nanocomb biosensor for glucose detection. Appl Phys Lett. 2006;88(23):233106. doi: 10.1063/1.2210078. [DOI] [Google Scholar]
  156. Wang S, Tan Y, Zhao D, Liu G. Amperometric tyrosinase biosensor based on Fe3O4 nanoparticles–chitosan nanocomposite. Biosens Bioelectron. 2008;23(12):1781–1787. doi: 10.1016/j.bios.2008.02.014. [DOI] [PubMed] [Google Scholar]
  157. Wang X, Zhi L, Müllen K. Transparent, conductive graphene electrodes for dye-sensitized solar cells. Nano Lett. 2008;8(1):323–327. doi: 10.1021/nl072838r. [DOI] [PubMed] [Google Scholar]
  158. Wang L, Chen G, Zhao J, Cai N. Catalase immobilization on amino-activated Fe3O4@ SiO2 nanoparticles: loading density affected activity recovery of catalase. J Mol Catal B Enzym. 2016;133:S468–S474. doi: 10.1016/j.molcatb.2017.03.011. [DOI] [Google Scholar]
  159. Wang Y, Wang H, Wang X, Xiao Y, Zhou Y, Su X, Sun F. Resuscitation, isolation and immobilization of bacterial species for efficient textile wastewater treatment: a critical review and update. Sci Total Environ. 2020;2:139034. doi: 10.1016/j.scitotenv.2020.139034. [DOI] [PubMed] [Google Scholar]
  160. Wang F, Xu Z, Wang C, Guo Z, Yuan Z, Kang H, Liu Y. Biochemical characterization of a tyrosinase from Bacillus aryabhattai and its application. Int J Biol Macromol. 2021;176:37–46. doi: 10.1016/j.ijbiomac.2021.02.042. [DOI] [PubMed] [Google Scholar]
  161. Wee Y, Park S, Kwon YH, Ju Y, Yeon KM, Kim J. Tyrosinase-immobilized CNT based biosensor for highly-sensitive detection of phenolic compounds. Biosens Bioelectron. 2019;132:279–285. doi: 10.1016/j.bios.2019.03.008. [DOI] [PubMed] [Google Scholar]
  162. Wei A, Sun XW, Wang JX, Lei Y, Cai XP, Li CM, Huang W. Enzymatic glucose biosensor based on ZnO nanorod array grown by hydrothermal decomposition. Appl Phys Lett. 2006;89(12):123902. doi: 10.1063/1.2356307. [DOI] [Google Scholar]
  163. Westereng B, Cannella D, Agger JW, Jørgensen H, Andersen ML, Eijsink VG, Felby C. Enzymatic cellulose oxidation is linked to lignin by long-range electron transfer. Sci Rep. 2015;5(1):1–9. doi: 10.1038/srep18561. [DOI] [PMC free article] [PubMed] [Google Scholar]
  164. Xie W, Zang X. Covalent immobilization of lipase onto aminopropyl-functionalized hydroxyapatite-encapsulated-γ-Fe2O3 nanoparticles: a magnetic biocatalyst for interesterification of soybean oil. Food Chem. 2017;227:397–403. doi: 10.1016/j.foodchem.2017.01.082. [DOI] [PubMed] [Google Scholar]
  165. Xiong F, Han Y, Wang S, Li G, Qin T, Chen Y, Chu F. Preparation and formation mechanism of size-controlled lignin nanospheres by self-assembly. Ind Crops Prod. 2017;100:146–152. doi: 10.1016/j.indcrop.2017.02.025. [DOI] [Google Scholar]
  166. Yang L, Zhao H, Li Y, Li CP. Electrochemical simultaneous determination of hydroquinone and p-nitrophenol based on host–guest molecular recognition capability of dual β-cyclodextrin functionalized Au@ graphene nanohybrids. Sens Actuators B Chem. 2015;207:1–8. doi: 10.1016/j.snb.2014.10.083. [DOI] [Google Scholar]
  167. Yang D, Wang X, Shi J, Wang X, Zhang S, Han P, Jiang Z. In situ synthesized rGO–Fe3O4 nanocomposites as enzyme immobilization support for achieving high activity recovery and easy recycling. Biochem Eng J. 2016;105:273–280. doi: 10.1016/j.bej.2015.10.003. [DOI] [Google Scholar]
  168. Yang T, Zhan L, Huang CZ. Recent insights into functionalized electrospun nanofibrous films for chemo-/bio-sensors. TrAC Trends Anal Chem. 2020;124:115813. doi: 10.1016/j.trac.2020.115813. [DOI] [Google Scholar]
  169. Yao YL, Ding Y, Ye LS, Xia XH. Two-step pyrolysis process to synthesize highly dispersed Pt–Ru/carbon nanotube catalysts for methanol electrooxidation. Carbon. 2006;44(1):61–66. doi: 10.1016/j.carbon.2005.07.010. [DOI] [Google Scholar]
  170. Yin H, Zhou Y, Xu J, Ai S, Cui L, Zhu L. Amperometric biosensor based on tyrosinase immobilized onto multiwalled carbon nanotubes-cobalt phthalocyanine-silk fibroin film and its application to determine bisphenol A. Anal Chim Acta. 2010;659(1–2):144–150. doi: 10.1016/j.aca.2009.11.051. [DOI] [PubMed] [Google Scholar]
  171. Yin PT, Shah S, Chhowalla M, Lee KB. Design, synthesis, and characterization of graphene–nanoparticle hybrid materials for bioapplications. Chem Rev. 2015;115(7):2483–2531. doi: 10.1021/cr500537t. [DOI] [PMC free article] [PubMed] [Google Scholar]
  172. Yin Z, Wang H, Jian M, Li Y, Xia K, Zhang M, Zhang Y. Extremely black vertically aligned carbon nanotube arrays for solar steam generation. ACS Appl Mater Interfaces. 2017;9(34):28596–28603. doi: 10.1021/acsami.7b08619. [DOI] [PubMed] [Google Scholar]
  173. Yin CM, Fan X, Liu C, Fan Z, Shi DF, Yao F, Gao H. The antioxidant properties, tyrosinase and α-glucosidase inhibitory activities of phenolic compounds in different extracts from the golden oyster mushroom, Pleurotus citrinopileatus (Agaricomycetes) Int J Med Mushrooms. 2019;21:9. doi: 10.1615/IntJMedMushrooms.2019031857. [DOI] [PubMed] [Google Scholar]
  174. Zaidi KU, Ali AS, Ali SA, Naaz I. Microbial tyrosinases: promising enzymes for pharmaceutical, food bioprocessing, and environmental industry. Biochem Res Int. 2014;2014(2):2. doi: 10.1155/2014/854687. [DOI] [PMC free article] [PubMed] [Google Scholar]
  175. Zehani N, Fortgang P, Lachgar MS, Baraket A, Arab M, Dzyadevych SV, Jaffrezic-Renault N. Highly sensitive electrochemical biosensor for bisphenol A detection based on a diazonium-functionalized boron-doped diamond electrode modified with a multi-walled carbon nanotube-tyrosinase hybrid film. Biosens Bioelectron. 2015;74:830–835. doi: 10.1016/j.bios.2015.07.051. [DOI] [PubMed] [Google Scholar]
  176. Zhang J, Lei J, Liu Y, Zhao J, Ju H. Highly sensitive amperometric biosensors for phenols based on polyaniline–ionic liquid–carbon nanofiber composite. Biosens Bioelectron. 2009;24(7):1858–1863. doi: 10.1016/j.bios.2008.09.012. [DOI] [PubMed] [Google Scholar]
  177. Zhang Q, Kang J, Yang B, Zhao L, Hou Z, Tang B. Immobilized cellulase on Fe3O4 nanoparticles as a magnetically recoverable biocatalyst for the decomposition of corncob. Chin J Catal. 2016;37(3):389–397. doi: 10.1016/S1872-2067(15)61028-2. [DOI] [Google Scholar]
  178. Zhou X, Qu Q, Wang L, Li L, Li S, Xia K. Nitrogen dozen carbon quantum dots as one dual function sensing platform for electrochemical and fluorescent detecting ascorbic acid. J Nanopart Res. 2020;22(1):1–13. doi: 10.1007/s11051-019-4718-8. [DOI] [Google Scholar]

Articles from 3 Biotech are provided here courtesy of Springer

RESOURCES