Next Article in Journal
The Course and the Effects of Agricultural Biomass Pyrolysis in the Production of High-Calorific Biochar
Next Article in Special Issue
The Advances and Challenges of Liposome-Assisted Drug Release in the Presence of Serum Albumin Molecules: The Influence of Surrounding pH
Previous Article in Journal
The Effect of Active Additives and Coarse Aggregate Granulometric Composition on the Properties and Durability of Pervious Concrete
Previous Article in Special Issue
Acrylate and Methacrylate Polymers’ Applications: Second Life with Inexpensive and Sustainable Recycling Approaches
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Carbon Nanomaterials (CNMs) and Enzymes: From Nanozymes to CNM-Enzyme Conjugates and Biodegradation

by
Petr Rozhin
1,
Jada Abdel Monem Gamal
2,3,
Silvia Giordani
2,* and
Silvia Marchesan
1,*
1
Department of Chemical and Pharmaceutical Sciences, University of Trieste, 34127 Trieste, Italy
2
School of Chemical Sciences, Faculty of Science & Health, Dublin City University, D09 E432 Dublin, Ireland
3
Department of Chemistry, Faculty of Mathematical, Physical and Natural Sciences, University Sapienza of Rome, 00185 Rome, Italy
*
Authors to whom correspondence should be addressed.
Materials 2022, 15(3), 1037; https://doi.org/10.3390/ma15031037
Submission received: 18 December 2021 / Revised: 20 January 2022 / Accepted: 26 January 2022 / Published: 28 January 2022
(This article belongs to the Special Issue Functional Nanomaterials for a Better Life)

Abstract

:
Carbon nanomaterials (CNMs) and enzymes differ significantly in terms of their physico-chemical properties—their handling and characterization require very different specialized skills. Therefore, their combination is not trivial. Numerous studies exist at the interface between these two components—especially in the area of sensing—but also involving biofuel cells, biocatalysis, and even biomedical applications including innovative therapeutic approaches and theranostics. Finally, enzymes that are capable of biodegrading CNMs have been identified, and they may play an important role in controlling the environmental fate of these structures after their use. CNMs’ widespread use has created more and more opportunities for their entry into the environment, and thus it becomes increasingly important to understand how to biodegrade them. In this concise review, we will cover the progress made in the last five years on this exciting topic, focusing on the applications, and concluding with future perspectives on research combining carbon nanomaterials and enzymes.

Graphical Abstract

1. Introduction

Carbon nanomaterials (CNMs) and enzymes belong to two scientific fields that have been traditionally widely separated, and research at their interface bears a number of challenges to overcome, both experimentally and culturally [1]. Nevertheless, in recent years a number of studies have allowed great progress in this research area, pushing the limits of what was considered possible, and demonstrating great innovation potential for a wide number of applications. Enzymes can be covalently conjugated onto nanocarbons, adsorbed on their surface, or even encapsulated in those with a hollow structure, as reviewed elsewhere [2,3,4,5]. The interactions between proteins and nanocarbons can play an important role in determining their fate in vivo, including their biodegradation [6,7,8], formation of a biocorona [9,10,11], and consequent modulation of the immune response [12].
Besides the exciting area of biointegration of nanocarbons, for instance for the regeneration of conductive tissues [13], and (flexible) bioelectronics’ development [14], the fields of sensing and diagnostics [15] certainly benefit from the inclusion of both nanocarbons and enzymes. Finally, considering the current emergency we are facing in terms of environmental impact of human activities and climate change that requires urgent action, the use of enzymes is very attractive for the production, modification, and degradation of nanocarbons to render the relevant processes more sustainable.
To provide a general overview of the diverse members of each family, this minireview will briefly introduce the classification of enzymes and nanocarbons, followed by an overview of the literature describing their combination. We will then discuss the recent progress in a variety of applications, focusing on the last five years, and directing the reader to reviews of each topic where older research can be found. Finally, we will conclude with a perspective towards the future of this exciting research area.

1.1. Enzyme Classification

Enzymes have attracted scientists’ attention for a long time. It is imperative to adopt an unambiguous nomenclature and classification for this enormous and diverse group of biological catalysts to permit their rigorous study and accelerate scientific progress in this area. The International Union of Biochemistry and Molecular Biology (IUBMB) has introduced the IUBMB ExplorEnz website [16], where it is possible to navigate the Enzyme Commission (EC) system, which uses four-component numbers to identify each enzyme (i.e., EC X.X.X.X):
  • The first component refers to the general type of reaction being catalyzed. For instance, EC 1 indicates oxidoreductases that catalyze redox reactions, and EC 3 identifies hydrolases that catalyze hydrolytic reactions (Figure 1).
  • The second number indicates the subclass based on the type of compound or functional group involved in the reaction. For example, EC 1.13 refers to oxygenases that insert oxygen on the substrate, and EC 2.3 indicates acyl-transferases that transfer acyl groups, etc.).
  • The third component denotes the sub-subclass, by further specifying the reaction being catalyzed, for instance in terms of acceptors, or specific groups being transferred. As an example, EC 2.1.1 indicates methyl transferases.
  • The fourth component is simply a serial number that refers to the specific enzyme.
However, the system presents room for improvement, because it is not always unambiguous. In fact, new classes are constantly being added to the EC system, and it should also be noted that non-physiologically occurring enzymes are not included [17]. For these reasons, many other databases can be useful to complement the ExplorEnz information. For instance, helpful information can be found in the relevant metabolic pathways described on KEGG [18], kinetic data on BRENDA [19], thermodynamic data on NIST [20], and human gene names curated by HUGO [21] and NCBI [22]. New databases are continuously being created, including some that use different classification criteria, such as MEROPS for peptidases [23], and others that more generally collect various types of data pertaining to proteins, such as Expasy [24].

1.2. Carbon Nanomaterials (CNMs)

The family of CNMs is very diverse, with new members continuously being discovered. In this minireview, we will examine the most popular types of CNMs (Figure 2), as well as their general properties. Those features are briefly discussed in the following sub-sections, where readers are referred to reviews that describe in more depth the various CNM types that are mainly composed of carbon atoms (Figure 2), and their general properties. We will briefly mention their main features in the following sub-sections, where readers are referred to recent reviews that describe more in depth the various CNM types [25].
Generally, sp2-hybridized carbon allotropes share common features, such as high mechanical strength and electronic conductivity, together with low density, and a high surface area; these can be functionalized to tailor their properties ad hoc for the intended application, as reviewed elsewhere [15,28]. In particular, the various modes of interaction and conjugation with proteins is a widely covered topic in the literature, and thus it will not be discussed here [1,2,29]. CNMs composed mainly of sp3-hybridized carbon atoms exist also, such as nanodiamonds (NDs) [30]; they find scope especially in sensing and catalysis [31], although they have not attracted scientists’ attention to a great extent, likely due to their more limited accessibility relative to other CNMs.
Each carbon nanostructure has a specific size, morphology, curvature, and reactivity. However, despite the many research efforts on their modification for tailored applications, it is not always straightforward to predict a priori which one will give the best performance, with great variation across allotropes [32,33,34,35,36]. Nevertheless, their conductivity is particularly promising to enhance (bio)materials’ properties [37,38,39,40], for the regeneration of conductive tissues such as the challenging nerve [41,42,43,44,45] and cardiac tissues [46,47,48,49,50,51], in sensing [52] and various types of catalyses [53,54,55], and in the field of energy sustainable production [56], conversion [57], and storage [58]. Furthermore, the performance in the latter fields can be enhanced by the generation of highly porous materials. Indeed, carbon foams [59] and porous carbon [60] are highly advantageous since they can be produced at low cost and from a variety of precursors [61], such as polymers [62] and biomass [63], bearing high potential especially in gas [64,65,66] and energy storage [67,68,69,70,71], and in catalysis [72,73,74]. However, in this review we will focus on the traditional CNM types, as described in the following sub-sections.

1.2.1. Fullerenes

Buckminsterfullerene, a discrete molecule composed of 60 carbon atoms (C60), was the first nano-allotrope to be discovered, in 1985 [75]. Fullerenes can also be composed of more carbon atoms, such as C76, C84, and C90 [76]. After C60, the rugby-ball shaped C70 is the second most known, and it is easier to isolate in sufficient amounts for research than higher fullerenes [77]. The advent of 100-milligram scale preparation of C60 [78] opened the way to decades of research that have provided several routes for its chemical functionalization, as recently reviewed [79]. Further tailoring of C60 properties can be attained through doping with heteroatoms [80]. Molecular confinement also offers the possibility to encase other elements in their interior to provide endohedral fullerenes [81] with interesting properties [82].
Fullerenes are fascinating molecules that have attracted researchers’ interest as they can be formed in space [83]. On Earth, they have been proposed for various applications, including the targeted delivery of therapeutics [84] and innovative antiviral therapies [85]. Despite the many potential uses in medicine [86], especially in photodynamic cancer therapy thanks to their ability to generate and modulate reactive oxygen species (ROS) levels [87], their electronic properties thus far found translation mainly in the field of photovoltaics, thanks to their electron-acceptor nature [88] and electron-transport ability [89].

1.2.2. Carbon Nano-Onions (CNOs)

Multiple fullerenes can be organized one inside another in carbon nano-onions (CNOs) [27]. These concentric, multi-layered fullerenes can have a size ranging from 2 to 100 nm, depending on the method of synthesis [90]. Generally, multi-fullerenes display a decreasing reactivity with increasing size, corresponding to decreasing curvature, and therefore, the associated strain on the CNO surface. Consequently, small CNOs present good reactivity, although it should be noted that also other factors, such as the production method and the consequent amount of defect sites, affect their reactivity [91].
In the last decade, their biological [92,93] and electrochemical [94,95] applications have significantly expanded due to the favorable properties of the nanomaterial, including their small size, large accessible surface area, and high biocompatibility [96,97,98], especially once they are rendered soluble through covalent and non-covalent functionalization [99,100,101,102,103]. It is worth noting that recent findings described white-light luminescence arising when CNOs were produced through pyrolysis and underwent oxidative treatments [104].

1.2.3. Carbon Nanohorns (CNHs)

Carbon nanocones [105] can form clusters termed carbon nanohorns (CNHs) [106]. Like many other CNMs, pristine nanocones unfortunately tend to aggregate in many solvents, and they do so in various morphologies termed dahlia-, bud-, or seed-like CNHs [107]. Despite this limitation, they can find promising applications in biosensing [108], medicine [109], and electrocatalysis [110], sometimes even outperforming other CNMs [111], and with relevance to clean energy [112] and carbon dioxide fixation [113]. There are not many studies on CNHs relative to the other CNMs, possibly also due to a more limited number of commercial sources. As a result, this type of nanomaterial remains an underexplored opportunity for innovation in numerous fields.

1.2.4. Carbon Nanodots (CNDs)

In the last few years, luminescent carbon dots (CNDs) have emerged for their innovation potential, especially in sensing and biomedicine [114,115], thanks to their low-toxicity, chemical inertness, ease of preparation, environmental friendliness, and interesting physico-chemical properties [116,117]. Another key advantage relative to many other CNMs, is the excellent solubility of CNDs in a large variety of solvents, both organic and aqueous, depending on their chemical nature [118].
These nanomaterials have attracted great attention for their cost-efficient and sustainable production [119], using, for instance, natural products [120,121] or biomass waste as carbon source [122]. However, the fine tailoring of their desired optical properties has been challenging [123]. Indeed, the exact chemical structure of CNDs still poses many unanswered questions [124,125], and the thorough use of spectroscopic methods is key to providing an accurate characterization of the emittive species [126]. As we advance our understanding of CNDs’ nature, further prospects open up [127], expanding their applications to organocatalysis [128], sensitizers for photocatalysis [129] and pollutant degradation [130], and to energy conversion and storage [131], as capacitor electrodes, for instance [132].

1.2.5. Nanodiamonds (NDs)

NDs are characterized by sp3-hybridization of their carbon core. NDs come in different sizes, morphologies, and surface types, depending on the method used for their preparation and functionalization [133]. They display attractive physico-chemical properties such as hardness, biocompatibility, and chemical inertness, leading to research on a variety of potential biological uses, especially delivery of therapeutics and imaging [134,135], but also sensing [136], tissue regeneration [137], skin products’ formulations [138], and in polished or active coatings with antimicrobial, antifriction, and mechanical reinforcing properties [139].
Recently, NDs have been considered also for their use in theranostics as applied to neurodegenerative diseases, thanks to their additional benefit of crossing the blood-brain-barrier [140]. Other emerging areas of application include primarily biological use in cells [141,142] and in vivo [143], but also catalysis [144].

1.2.6. Carbon Nanotubes (CNTs)

Carbon nanotubes can be composed of one graphitic layer rolled up in single-walled CNTs (SWCNTs) [145], or multiple coaxial graphitic nanotubes called multi-walled CNTs (MWCNTs) [146], which can be grown with branches also [147]. CNTs can have very different properties; for instance, they can be semiconducting or metallic, depending on their type and chirality [148]. Over the years, several functionalization strategies have been developed [149]; oxidation is by far the most popular way to increase their polarity and dispersibility in various solvents, including water [150]. Therefore, fine control over their synthesis, purity and, thus, homogeneity is critical to enable their translation into large-scale use [151] and to fill the gap between their properties as individual CNTs and those of their bundled aggregates—often a practical limitation for industrial use [152].
CNTs have attracted great interest although there are still challenges to overcome to enable a wider commercial use of their unique properties [153]. Key areas of application include various types of high-performing composite materials, where demands of conductivity, robustness, flexibility, and mechanical resistance are high [154]. These include artificial neuromuscular prostheses [155,156], and more generally nano-bioelectronics [157] and wearable electronics [158], but also sensing [159] and imaging [160], orthopedic devices [161], tissue regeneration and biomedical use [162,163,164], electroactive materials for environmental and energy technology [165,166,167,168,169,170,171], electronics and computing [172,173], and various forms of catalysis [174,175,176,177,178,179].

1.2.7. Graphene (G) and Graphene-Based Materials

In the last decade, the most popular carbon allotrope has been graphene (G), which can be considered as a 2D layer of sp2-hybridized carbon atoms arranged in a honeycomb lattice. It is worth noting that G can come in many forms, in terms of size, layers, level of oxidation, etc. which will all affect its physico-chemical properties [180]. In particular, graphene oxide (GO) [181,182] or its reduced form (rGO) [183,184] are often used for their improved dispersibility, relative to pristine G. G properties have also been tailored through topology [185], such as twists and nanoribbons [186]. Given the large heterogeneity in size, number of layers and of defects of G flakes, the general term of graphene-based materials is preferred over just G, to refer to this sub-class of CNMs [187].
Applications are similar to CNTs, and they include composite reinforcement [188], wearable [189] and flexible electronics [190,191], including memory devices [192] and even stretchable batteries [193], energy storage [194] and conversion [195,196,197,198], environmental remediation [199,200], varying types of catalysis [201,202,203,204], and innovative uses in the healthcare sector [205], such as regenerative medicine [206] and sensing [207,208]. In this case, large-scale, cost-effective production of high-quality G [209,210] and standardization are key for the translation of G properties into commodity products at a global level [211].

1.3. Bibliometric Analysis of CNMs and Enzymes

A literature search for the term “enzyme” in conjunction with each one of the most popular CNMs shown in Figure 2, in the title, abstract, or keywords, has revealed that the vast majority of scientific articles pertain to CNTs (4.3 × 103 documents), followed by G (3.4 × 103 documents). However, in the last decade (Figure 3), scientific works on either one averaged about 300 per year, with a slight decrease for CNTs after 2017, opposed to a continuous increase for G up to 2019, surpassing CNTs in 2015. This trend could be related to concerns over CNTs’ toxicity, as suggested by the fact that scientific documents on this topic peaked at nearly 300 in 2015, and since then held steady. However, given their high innovation potential in medicine [212,213], and the high number of variables that affect their biocompatibility [214,215], alarming generalizations to ban their use are best avoided [216].
In comparison to CNTs and G, all the other CNMs lagged behind, each one reaching far less than 100 in total—with two exceptions—and representing today a missed opportunity for research. The first exception was fullerenes, possibly since they were the first to be discovered and thus have had more years of related research, reaching just over 600 records. The second exception was CNDs, which are among the most recent CNMs to be discovered. Scientific papers on CNDs have been increasing steadily year after year, since the first ones appeared in 2004. Therefore, it is foreseeable that CNDs will keep rising in popularity in the immediate future, although there is still a long way ahead to approach the numbers seen for CNTs and G.
In terms of applications, most studies pertain biosensing and biofuel cells (Table 1), although biocatalysis and biomedical applications other than sensing have been pursued with these systems. Interestingly, CNMs have also been envisaged as enzyme mimics, inhibitors, or detectors. We will concisely cover the progress made over the last five years on all these topics in the following sections.

2. CNMs for Enzyme Mimicry, Inhibition, or Monitoring

2.1. CNMs for Enzyme Mimicry

A plethora of works describe the use of CNMs as nanozymes, meaning nanostructures that mimic enzymes as they display catalytic activity [289]. Research in this area is intended to overcome some of the common limitations of enzymes, particularly, the limited physico-chemical resistance against solvents and changes in temperature, pH or other experimental conditions [290]. Potential applications range from various biomedical applications [291], including innovative therapy [292] biosensing [293,294], and disinfection [295], to environmental monitoring and remediation [296]. In particular, peroxidase mimicry by CNMs (Figure 4) has been widely studied [297], especially for the development of glucose biosensors [298], although hydrolase mimicry is also attracting increasing interest [299].
In particular, CNDs have been functionalized with Fe (III) to mimic peroxidases and exert antimicrobial activity through generation of hydroxyl radicals [301]. In contrast, no hydroxyl radicals were generated when they were derivatized with glucose or cyclodextrin to mimic peroxidases, indicating a different mechanism [302]. The peroxidase mimicry activity can be correlated to the phosphorescence quantum yield and can inhibit bacterial growth under light irradiation, an activity that was envisaged for photodynamic antimicrobial chemotherapy applications [303]. Alternatively, the use of light could trigger radical oxygen species generation by the CND nanozymes, in an effort to mimic nuclease activity and cleave DNA [304].
Chemical functionalization has been successfully employed to attain CNDs with switchable fluorescence too. In this case, the fluorescence of amino-derivatized CNDs can be quenched by chelation with Fe(II) ions as nanozymes, and restored upon treatment with hydrogen peroxidase with a concomitant shift from yellow to green [305]. Furthermore, addition of other divalent metal cations can lead to additional advantages. For instance, Mn(II) extends the peroxidase mimicry by CNDs to neutral pH values, which is otherwise rather uncommon [306]. Despite the fact that mimicking enzymes’ enantioselectivity is a grand challenge, recent reports are demonstrating it is possible, in topoisomerase mimicry, for example [307]. Heteroatom-doped CNDs have been developed for theranostics as well, thanks to nanozyme activity [308]. Many other examples of nanozymes based on CNDs have been reported, through addition of other components, such as hemin [309], metal nanoparticles (NPs) [310,311], co-doping with various elements [312,313,314], MOFs [315], carbon nitride [316], and metal oxides [317].
Both graphene oxide (GO) and its reduced derivative (rGO) have also demonstrated peroxidase mimicking ability, which has been ascribed to the presence of carbonyl groups on the surface of the nanomaterial that get activated by hydrogen peroxide as a key step in the catalytic cycle [318]. Interestingly, rGO co-doping with N and B allowed development of nanozymes for the selective mimicry of peroxidases (but not oxidases) with enhanced catalytic performance for the development of biosensors [319].
Analogously, oxidized CNTs demonstrated peroxidase-like activity, which was envisaged for the treatment of bacterial infections [320]. Different oxygen-bearing functional groups exert competing interactions with hydrogen peroxide (Figure 5), and thus control over oxidation is important [320]. Combination of CNTs with other chemical components is a popular strategy to tailor nanozyme activity to the intended application. In a recent example, SWCNTs have been functionalized with a nickel complex for the biomimicry of oxidase for H2 oxidation, and subsequent integration in fuel cells [267]. Alternatively, MWCNTs were coated with polypyrrole to introduce N-based ligands for Fe to be used in single-atom catalysis as peroxidase mimics [321]. CNTs were combined with hemin for peroxidase mimicry also [322,323]. They have been derivatized with polyoxometallate-based metal-organic frameworks (MOFs) for the selective sensing of cysteine [324], or with copper complexes [325], MOFs [326], and NPs [327] to develop nanozymes. For this type of application, many types of metal NPs have been used [328,329,330], as well as polymers to mimic phosphodiesterases [331].
CNTs can be further assembled into macroscopic materials, such as carbon nanofibers (CNFs), which find many uses, especially in high-performance composites and energy devices [332,333,334,335,336]. In this case, they have been decorated with Fe(III) complexes to mimic oxidases [337], peroxidases, and catalases for sensing and environmental technology [338]. Analogously to CNTs, CNFs can be oxidized, although gas-phase methods are preferable to the liquid-phase methods typically used for CNTs, to preserve the CNF macroscopic morphology, with the additional advantage of being virtually waste-free [339].
Peroxidase mimicry can be exerted by other CNMs as well, and for various uses. CNHs have been used as peroxidase mimics for the detection of drug traces as environmental pollutants [340]. They have been combined with nanosized ceria to detect hydrogen peroxide in commodity products, such as washing liquids and milk [341]. In the case of CNOs, nitrogen doping has been successfully applied to improve their catalytic performance in the electrochemical generation of molecular oxygen from hydrogen peroxide [342]. Boron and nitrogen co-doped CNOs showed great performance as electrocatalysts for the oxygen reduction reaction [95]. Interestingly, in the case of NDs, they were envisaged for redox-enzyme mimicry, with an activity that could be selectively tailored depending on the pH. At acidic pH, NDs catalyzed the reduction of molecular oxygen and hydrogen peroxide. At alkaline pH, they catalyzed the dismutation decomposition of hydrogen peroxide to produce molecular oxygen. It was proposed that the molecular mechanism of their peroxidase-like activity is electron-transfer acceleration, the source of which is likely derived from oxygen-containing functional groups on their surface [343].
Finally, besides peroxidases, and, generally, redox-active enzymes, which represent the vast majority of nanozyme mimicry studies on CNMs, hydrolases have started to attract scientists’ attention. In a recent report, fullerene derivatives were applied to this end through the presentation of multiple functional groups inspired from the natural enzymes’ catalytic sites [344]. Analogously to the other CNMs, fullerenes could also act as peroxidase mimics at acidic pH, and were thus envisaged for the eradication of Helycobacter pylori in vivo [345].

2.2. CNMs as Enzyme Inhibitors

Fullerene derivatives have been envisaged for the inhibition of a variety of enzymes, including recent examples of HIV-1 protease [346], ribonuclease A [347], glycosidases [347], ubiquitin-activating enzyme 1 [348], and acetylcholinesterase [349]. Their size, hydrophobic nature, and spherical morphology appear very suitable for hydrophobic interactions with lipophilic sites on the target enzymes (Figure 6), whilst C60 functionalization can add hydrophilic appendages for more specific interactions.
CNDs have been used to inhibit tyrosinase for cosmetic and food applications, thanks to hydrophobic interactions between the CNDs and the enzymatic surface, as well as chelation by the CND COOH groups of the enzyme copper ions. Tyrosinase is involved in the browning process of fruits and vegetables, and its overexpression has been linked to skin pigmentation disorders and tumorigenesis. Therefore, its inhibition could find several useful applications [350]. Furthermore, CNDs were found to tune glucose oxidase activity, depending on their functionalization type [224], and inhibit maltase, an effect that was envisaged as an innovative means to control physiological glucose levels [230].
Enzyme inhibition has been studied for CNTs also. SWCNTs demonstrated the ability to act as competitive inhibitors for proteases, such as chymotrypsin, thanks to hydrophobic interactions between the curved CNT surface and a morphologically complementary crevice on the enzyme surface, without alteration of the enzyme secondary structure or active site [351].

2.3. CNMs for Enzyme Monitoring

CNMs can be engineered to monitor enzymatic activity. Several examples have been reported especially using CNDs, whose fluorescence is initially quenched through interaction with a second component, and then restored upon a chemical transformation triggered by the activity of the target enzyme [352,353,354]. To this end, graphene CNDs have been functionalized with a cobalt derivative to allow for redox-dependent fluorescence that can be used to detect alkaline phosphatase activity, in serum, through the dephosphorylation of a substrate on the CNDs that releases ascorbic acid, which restores fluorescence [352]. Through a similar principle, silver NPs have been applied to quench the fluorescence of CNDs, so that, in the presence of enzymatic activity that generates hydrogen peroxide as a byproduct (e.g., through an oxidase), the silver NP structure decomposes, and fluorescence is restored. Applications in the health sector were envisaged, in particular for the monitoring of relevant biomolecules, such as glucose or cholesterol that could act as substrates for the corresponding oxidase [353]. Alternatively, glycosidases could be monitored through a similar principle, by functionalizing the CNDs to favor interaction with p-nitrophenol, which is generated through enzymatic activity on a glycosylated derivative [354,355]. Another target enzyme was thioredoxin reductase, overexpressed in many cancer cells [356].
SWCNTs have been applied to enzyme biosensing. Recently, SWCNTs were coated with a peptide to develop a biosensor for trypsin detection in urine samples, exploiting variations in CNT near-infrared photoluminescence upon enzymatic degradation of the peptide coating [357]. In another example, CNTs were envisaged for applications in cancer diagnostics, through the detection of matrix metalloproteinase-7, which is overexpressed in cancer cells [358]. Finally, CNT-fibers have been used to develop highly sensitive (54 µA·cm−2·mM−1) photoelectrodes for the detection of NADH, which is a key cofactor in many biocatalytic processes; its quantification correlates to specific enzyme activity [359].

3. Applications of CNM-Enzyme Conjugates

3.1. Biosensing

Biosensors typically comprise three elements, which are: (1) an element for biological recognition, such as an enzyme; (2) a transducer, to convert energy from the biorecognition event into another form (electrical, thermal, optical, etc.); (3) a signal processing system for the response readout and/or recording [360]. Biosensors often rely on enzyme inhibition, thus being ideal to monitor inhibitors that are relevant to human health, such as drugs or pollutants [361]. Enzymes are ideal components for biosensing, thanks to high sensitivity, specificity, low cost, and accessibility [360]. Coupling a semiconductor to enzymes can be exploited in photobiocatalysis, which is inspired by natural photosynthesis, but does not necessarily involve light for activation [362]. In general, inclusion of nanomaterials allows for better performance in a variety of analytical parameters, such as sensitivity, detection limit, stability, and response rate [363].
In particular, CNMs are ideal supports especially for biosensors that require multiple layers of enzymes, but also for providing a good electronic contact through the layers and with the electrodes [364]. CNMs can be good active supports for oxidoreductases as they may facilitate electron transfer to enhance catalysis, whilst offering a high surface area for high-level loading of enzymes [365]. However, the occurrence of direct electron transfer (DET) is a matter of ongoing debate, depending on the type of enzymes under consideration, the accessibility of their redox-active site to the CNMs, and the type of direct or indirect contact between CNMs and enzymes [265]. The electronic properties of CNMs render them attractive building blocks for electrochemical biosensors, besides the more traditional optical alternatives [366,367].
Graphene is one of the most studied CNMs for a variety of biosensing devices (Figure 7) thanks to its exceptional electronic and mechanical properties, as recently reviewed in detail elsewhere [368]. Other less studied CNMs, such as CNOs, can also make attractive electrode components for the development of low-cost, simple to use, and highly sensitive sensors [237,369]. CNOs (mean size of 30 nm) were employed as electrochemical sensors by covalently immobilizing the glucose oxidase enzyme (GOx) on their surface via carbodiimide chemistry. GOx selectively catalyzed the oxidation of glucose, giving a sensor with high sensitivity and selectivity. However, the catalytic activity of GOx on the sensor electrode was highly sensitive to environmental conditions such as temperature, pH and humidity. Furthermore, the performance of the sensor was limited by enzymatic stability. Thus, an enzyme-free glucose sensor was designed, using Pt-decorated CNOs (Pt@CNOs) that outperformed many other CNMs previously studied for the same application [237].
In 2020, Cumba et al. [369] described the preparation of the first ink that was based on CNOs to produce cheap and disposable electrodes, yielding sensors with elevated performance (Figure 8). Careful selection and optimization of all the components was a key step to attain a suitable formulation for the ink to be screen-printed. They included the conducting nanocarbons (i.e., graphite (GRT) and CNOs), the polymer binder, the plasticizer, and the organic solvent. The electrodes were screen-printed and they consisted of a conducting network of interconnected CNMs with a uniform distribution. The system displayed a heterogeneous electron transfer rate constant corresponding to 1.3 ± 0.7 × 10−3 cm·s−1 and also a current density that was higher than the ferrocene/ferrocenium coupled to a GRT screen-printed electrode that was commercially available. Furthermore, the CNO/GRT electrode allowed for the detection of dopamine in micromolar concentration (i.e., 10.0–99.9 µM), and with a 0.92 µM detection limit. The analytical sensitivity thus revealed a notable 4-fold increase relative to the commercial reference electrode based on GRT. Overall, this study opened the way to the use of CNO-based electrodes for high-performance sensing, electrocatalysis and battery research [369].
As can be seen from Table 1, the vast majority of CNM-enzyme conjugates have been studied for biosensing applications. The most popular target molecule is glucose for biometric health monitoring (Table 2) [224,227,228,237,249,269,271,273]. However, biosensors have been developed to detect many other bioactive compounds too, such as cholesterol [312,370,371] and triglycerides [282], lactose [220] and lactate [225,280], neurotransmitters [234,240,246,278,279,372] and hormones [239], various disease biomarkers [244,257,261], microRNAs [223], drugs [218,287], pathogens [219] and toxins [221], xanthine [262] and caffeic acid [373], p-coumaric acid [232], ferulic acid [233], trace metals [274], and oxygen [268].

3.2. Biofuel Cells

Biofuel cells are electrochemical devices that typically use redox enzymes as sustainable catalysts for the conversion of chemical energy into electrical energy (Figure 9); they consist of two-electrode cells that are separated by a proton-conducting medium. At the bioanode, fuels are oxidized, freeing electrons that flow to the biocathode through the external electrical circuit. At the biocathode, oxidants such as oxygen or peroxide are reduced to water [374]. Redox-active enzymes have attracted great interest for their use in the electrochemical production of fuels as sustainable alternatives in the field of clean energy, such as water splitting reactions [375].
Conjugation with CNMs allows for high-performance devices. They have been coupled to enzymes to serve as anodes [248,251,258,259,269,285], cathodes [253,267,276,277,286], or both [247,284].
An electrochemical reaction of particular interest is the molecular oxygen reduction (ORR) at the cathode. In this case, use of CNT-laccase as biocathode allowed reaching current densities >1.8 mA·cm−2, a direct electron transfer efficiency as high as 70–100%, and a turnover frequency of 5.0·103 s−1 [253]. When bilirubin oxidase was used coupled to CNTs at the cathode, a maximum current density of 5.5 mA·cm−2 was found, and a power density of 1.85 mW·cm−2 at 0.6 V was attained, relative to 2.46 mW·cm−2 at 0.32 V with Pt/C as counter electrode [267]. Addition of catalase to a glucose oxidase (GOx)-CNT conjugate was thought to be another convenient strategy for ORR. In this case, GOx catalyzes the oxidation of glucose to gluconolactone with the concomitant consumption of molecular oxygen to produce hydrogen peroxide, which is then converted by the catalase into water and molecular oxygen that feeds back into the GOx reaction. As a result, this type of catalyst reached a maximum power density of 0.18 mW·cm−2 and a current density of 59 μA·cm−2 [276]. Another additive that can assist with catalytic performance in ORR is 2,2′-azino-di-(3-ethylbenzthiazoline sulfonic acid) or ABTS, which is a common substrate for hydrogen peroxidase and acts as an efficient electron transfer mediator between the enzyme and the electrode surface. With ABTS, a maximum power density of 1.12 mW·cm−2 at 0.45 V was obtained, which after two weeks had decreased just to 0.928 mW·cm−2, indicating good stability over time [277].
Wearable CNT-based biofuel cells were developed on a cotton textile that allowed illumination of an LED on the cloth [376]. Amongst the CNMs that have been used with enzymes in biofuel cells as summarized in Table 1, CNTs are certainly the mostly studied [247,248,251,253,258,259,267,269,276,277,284,285,286]. Recently, scientists are recognizing innovation opportunities also in other types of CNMs, such as CNDs [229], GO [242] or rGO [275], although reports in this direction are still very limited.

3.3. Biocatalysis

Thanks to great progress on biotechnology and protein engineering, biocatalysis has emerged as a green solution to increase the efficiency of industrial processes in a sustainable way [377]. Its importance and societal impact has been recognized through the Nobel Prize in chemistry in 2018 to Arnold, who pioneered the directed evolution that enabled development of resistant enzymes of industrial interest [378].
CNMs can be envisaged as active supports to immobilize enzymes and facilitate their recycling [379]. Besides, their electronic properties may favor the catalytic performance of redox-active enzymes. To this end, enzymes have been coupled to CNTs to enable asymmetric hydrogenation in flow [380]. Furthermore, bioelectrocatalysis involving direct electron transfer (DET) can benefit from the use of CNMs as active supports for redox enzymes, and the role played by their surface functionalization in the process has recently been reviewed [381].

3.4. Water Remediation and Environmental Monitoring

Enzymes supported on nanomaterials can be very convenient to detect pollutants for environmental monitoring through the development of sensitive sensors, but also for their removal from polluted waters [382]. For example, CNMs coupled to enzymes can be applied for the electrochemical monitoring of chromium [383]. CNDs’ fluorescence has also been envisaged for the optical detection of organic pesticides through coupling with an enzymatic reaction [384]. In addition, rGO has been envisaged for the detection of pesticides through the immobilization of an esterase on a biocomposite containing fibrin and thrombin, which was assembled taking inspiration form the blood coagulation process [245]. Finally, CNOs were coupled to a peroxidase in a cyclodextrin polymer matrix for the detection of herbicides, as tested in soil and river water samples [238].

3.5. Innovative Therapy and Theranostics

The rise of smart materials that can respond and adapt to stimuli and changes in the local microenvironment has opened new avenues that are enabling great progress especially in the biomedical field [385]. Enzymes can be used as convenient stimuli for the design of responsive materials [386], with great potential in the development of combined therapy and diagnosis, for instance through activation on a target pathological site characterized by the selective overexpression of certain enzymes [387]. The coupling of enzyme-responsive materials with nanostructures can be convenient to develop photodynamic therapies for cancer treatment [388].
Alternatively, enzymes can be supported onto CNMs for combined chemodynamic therapy (CDT). For example, MWCNTs were functionalized with Fe3O4 and glucose oxidase, so that the enzyme could convert glucose into gluconate and hydrogen peroxide. Conversion of the latter through the iron oxide-mediated Fenton reaction into hydroxyl radicals induces tumor cell death, and the reaction is favored by the lowered pH of the local microenvironment due to gluconate production. Finally, near-infrared (NIR) light irradiation can further boost the overall process at the target pathological site through generation of hyperthermia [252].
With the rise of biologics, enzymes have found applications also as therapeutic agents. As an example, laronidase can be used as replacement therapy for a type of mucopolysaccharidosis that is associated with deficiency of the natural enzyme, which hydrolyses glucosaminoglycans, causing their pathological accumulation in lysosomes. MWCNTs were thus envisaged as vectors for laronidase, which was covalently conjugated onto the CNMs [254]. There are clearly many unexplored opportunities in this research area that are worth future investigation.

4. Enzymatic Biodegradation of CNMs

The possibility of biodegrading CNMs through enzymatic activity is very appealing for various reasons, including lowering their persistence in the environment after use, but also avoiding or reducing their bioaccumulation in living organisms. Furthermore, the breaking down of larger CNMs, such as GO, into smaller components, can be envisaged as a green production method of graphene quantum dots (Figure 10) [389].
In general, the aromatic nature of CNTs renders them persistent in the environment, with little or no degradation by microorganisms [390], yet their oxidized forms appear to be biodegradable by microorganisms, whose enzymes are likely to use hydroxyl groups on the CNT surface as attackable sites that can be processed through enzymatic activity [391]. Various peroxidases have been found to be able to biodegrade CNTs and G derivatives, as recently reviewed [392]. They are mainly horseradish peroxidase (HRP), myeloperoxidase (MPO), manganese peroxidase (MnP) and lignine peroxidase (LiP). These four enzymes require hydrogen peroxide to participate in the degradation of CNMs. In the enzymatic degradation process of CNMs, molecular docking technology is used to predict possible binding sites, which helps to understand the degradation mechanism [393]. Recently, oxidases were reported to biodegrade MWCNTs [394], CNDs [8,395], and fullerenes [396]. It is not surprising to see that nanozymes are being developed for the same purpose, for instance as applied to the degradation of GO [397].
It is worth noting that besides the type of CNM, the level and type of functionalization is one of the factors playing a key role in determining the CNM biodegradation. Whilst it is accepted that oxidation generally favors biodegradation [150], other types of functionalization can have the opposite effect. In particular, chemical reduction of GO [398] and/or coating with bovine serum albumin or polyethylene glycol [399] rendered the CNM resistant to peroxidase-mediated biodegradation.
CNM biodegradation mediated by bacteria typically involves electron-transfer processes, which lead to the breaking of C–C covalent bonds. As a result, numerous pores arise on the surface of CNMs which lose structural integrity. Electrons can flow in either direction at the CNM-bacteria interface. In particular, cationic and anionic CNMs act as electron acceptors and donors, respectively [400]. Furthermore, oxygen interference can occur at the point of electron transfer between bacteria and CNMs [401]. In general, the functionalization of CNMs with anionic species on the surface of CNMs favors the electrostatic interaction with enzymes, which often display cationic amino acids on their surface, but also the catalytically-active heme group in redox-active enzymes plays a role in the interaction with CNMs. Clearly, pristine CNMs may be more challenging to degrade, and defect sites offer typical locations for the beginning of their structural deterioration [402].
Currently, fullerene biodegradation is still a largely unexplored research topic. It is known that this nanocarbon is challenging to degrade when exposed to soil bacteria [403]. However, the situation is notably improved in the case of organics-rich clay, such that more than half of the fullerene present can be mineralized just over two months, and even more so in the case of functionalized fullerol. Its structural deterioration can be notably accelerated through the combination of biodegradation with photochemistry, which likely mediates the destruction of the stable aromatic core [404]. Likewise, C60 photodegradation using UV light was facilitated by hydroxylation [405]. In another study, fullerene aggregates decreased in volume upon exposure to bacteria, with occurrence of hydroxylation, although the structural deterioration of the nanocarbon was slow and no significant production of carbon dioxide from C60 was noted, using isotope labelling [406]. In general, the efficiency of photodegradation can be relatively high, but it should be noted that only UV light can degrade CNMs. In natural environments, CNMs will react with other substances too, and their degradation by UV light will be affected by all these factors. There is still a knowledge gap in the detailed understanding of biodegradation of several CNMs, especially in realistic experimental conditions pertaining to those found in the environment, including soil and water.

5. Conclusions and Future Perspectives

Combining CNMs and enzymes requires a diverse skill set that is rare to find and represents a multidisciplinary research area that bears many technical and scientific challenges. However, a growing number of scientists are trying to innovate in this exciting field. The focus of our review has been to provide a concise overview from which it is evident how most studies have been focused on CNTs and, more recently, on graphene-based materials and CNDs, for applications in biosensing and biofuel cells. Nonetheless, CNMs offer far more benefits, and the multivarious members of the nanocarbon family still present today a valuable innovation opportunity that is worth exploring. Among other aspects that deserve further examination is their environmental fate, especially how biodegradation and photodegradation processes can improve the efficiency of CNM degradation.
Further research potential can be found in the development of computational methods to enhance enzymatic performance and robustness [407], including machine learning for enzyme engineering [408], potentially coupled to directed evolution approaches [409]. The range of enzymatic activity can be further expanded through the incorporation of unnatural amino acids [410], thanks to the emergence of robust methods for their genetic encoding [411]. Higher levels of complexity for the development of the next-generation devices can be attained with the incorporation of multienzymatic cascade reactions [412], also in confined environments [413], in an attempt to mimic, or go even beyond, the mesmerizing performance of biochemical cascades in living organisms. To this end, advancing electrochemical techniques for the characterization of enzymes at the electrode interface will be key [414], especially to leverage the unique electronic properties of CNMs and their application to further enhancing enzymatic activity. In particular, an attractive area is the development of wearable, flexible bioelectronics for the harvesting of bioenergy and its use in self-powered biosensing for health monitoring [415].

Author Contributions

Writing—original draft preparation, P.R. and J.A.M.G.; writing—review and editing, S.G. and S.M; visualization, P.R. and J.A.M.G.; supervision, S.G. and S.M. All authors have read and agreed to the published version of the manuscript.

Funding

Part of the research described in this work received funding from the Italian Ministry of University and Research (MIUR) through the PRIN grant n. 2015TWP83Z to S.M., J.A.M.G. acknowledges funding through the Erasmus+ program.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Acknowledgments

This article is based upon work from COST Action EsSENce CA19118, supported by COST (European Cooperation in Science and Technology). Available online: www.cost.eu (accessed on 15 December 2021). The authors wish to thank Michał Bartkowski for proofreading the manuscript.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Calvaresi, M.; Zerbetto, F. The Devil and Holy Water: Protein and Carbon Nanotube Hybrids. Acc. Chem. Res. 2013, 46, 2454–2463. [Google Scholar] [CrossRef] [PubMed]
  2. Marchesan, S.; Prato, M. Under the lens: Carbon nanotube and protein interaction at the nanoscale. Chem. Commun. 2015, 51, 4347–4359. [Google Scholar] [CrossRef] [PubMed]
  3. Wang, X.; Zhu, Y.; Chen, M.; Yan, M.; Zeng, G.; Huang, D. How do proteins ‘response’ to common carbon nanomaterials? Adv. Colloid Interface Sci. 2019, 270, 101–107. [Google Scholar] [CrossRef]
  4. Chaudhary, K.; Kumar, K.; Venkatesu, P.; Masram, D.T. Protein immobilization on graphene oxide or reduced graphene oxide surface and their applications: Influence over activity, structural and thermal stability of protein. Adv. Colloid Interface Sci. 2021, 289, 102367. [Google Scholar] [CrossRef]
  5. Botta, L.; Bizzarri, B.M.; Crucianelli, M.; Saladino, R. Advances in biotechnological synthetic applications of carbon nanostructured systems. J. Mater. Chem. B 2017, 5, 6490–6510. [Google Scholar] [CrossRef] [PubMed]
  6. Hou, J.; Wan, B.; Yang, Y.; Ren, X.-M.; Guo, L.-H.; Liu, J.-F. Biodegradation of single-walled carbon nanotubes in macrophages through respiratory burst modulation. Int. J. Mol. Sci. 2016, 17, 409. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  7. Zhang, M.; Yang, M.; Bussy, C.; Iijima, S.; Kostarelos, K.; Yudasaka, M. Biodegradation of carbon nanohorns in macrophage cells. Nanoscale 2015, 7, 2834–2840. [Google Scholar] [CrossRef]
  8. Srivastava, I.; Sar, D.; Mukherjee, P.; Schwartz-Duval, A.S.; Huang, Z.; Jaramillo, C.; Civantos, A.; Tripathi, I.; Allain, J.P.; Bhargava, R.; et al. Enzyme-catalyzed biodegradation of carbon dots follows sequential oxidation in a time dependent manner. Nanoscale 2019, 11, 8226–8236. [Google Scholar] [CrossRef]
  9. Czarnecka, J.; Kwiatkowski, M.; Wiśniewski, M.; Roszek, K. Protein corona hinders N-CQDs oxidative potential and favors their application as nanobiocatalytic system. Int. J. Mol. Sci. 2021, 22, 8136. [Google Scholar] [CrossRef]
  10. Bhattacharya, K.; Mukherjee, S.P.; Gallud, A.; Burkert, S.C.; Bistarelli, S.; Bellucci, S.; Bottini, M.; Star, A.; Fadeel, B. Biological interactions of carbon-based nanomaterials: From coronation to degradation. Nanomedicine 2016, 12, 333–351. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  11. Pinals, R.L.; Yang, D.; Lui, A.; Cao, W.; Landry, M.P. Corona exchange dynamics on carbon nanotubes by multiplexed fluorescence monitoring. J. Am. Chem. Soc. 2020, 142, 1254–1264. [Google Scholar] [CrossRef] [PubMed]
  12. Marchesan, S.; Kostarelos, K.; Bianco, A.; Prato, M. The winding road for carbon nanotubes in nanomedicine. Mater. Today 2015, 18, 12–19. [Google Scholar] [CrossRef]
  13. Wang, W.; Hou, Y.; Martinez, D.; Kurniawan, D.; Chiang, W.-H.; Bartolo, P. Carbon nanomaterials for electro-active structures: A review. Polymers 2020, 12, 2946. [Google Scholar] [CrossRef] [PubMed]
  14. Yu, Y.; Nyein, H.Y.Y.; Gao, W.; Javey, A. Flexible electrochemical bioelectronics: The rise of in situ bioanalysis. Adv. Mater. 2020, 32, e1902083. [Google Scholar] [CrossRef] [PubMed]
  15. Speranza, G. Carbon nanomaterials: Synthesis, functionalization and sensing applications. Nanomaterials 2021, 11, 967. [Google Scholar] [CrossRef] [PubMed]
  16. ExplorEnz—The Enzyme Database. Available online: https://www.enzyme-database.org/ (accessed on 19 January 2022).
  17. McDonald, A.G.; Tipton, K.F. Enzyme nomenclature and classification: The state of the art. FEBS J. 2021; in press. [Google Scholar] [CrossRef] [PubMed]
  18. KEGG PATHWAY Database. Available online: https://www.genome.jp/kegg/pathway.html (accessed on 19 January 2022).
  19. Institute of Biochemistry and Bioinformatics at the Technical University of Braunschweig. BRENDA The Comprehensive Enzyme Information System. Available online: https://www.brenda-enzymes.org/ (accessed on 19 January 2022).
  20. NIST National Institute of Science and Technology. Thermodynamics Research Center. Available online: https://trc.nist.gov/ (accessed on 19 January 2022).
  21. The European Society of Human Genetics. ESHG Home. Available online: https://www.eshg.org/index.php?id=home (accessed on 19 January 2022).
  22. NCBI. Gene Search Database. Available online: https://www.ncbi.nlm.nih.gov/gene (accessed on 19 January 2022).
  23. Rawlings, N.D.; Waller, M.; Barrett, A.J.; Bateman, A. MEROPS: The database of proteolytic enzymes, their substrates and inhibitors. Nucleic Acids Res. 2014, 42, D503–D509. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  24. Swiss Insitute of Bioinformatics. Expasy Swiss Bioinformatics Resource Portal. Available online: https://www.expasy.org/ (accessed on 19 January 2022).
  25. Georgakilas, V.; Perman, J.A.; Tucek, J.; Zboril, R. Broad family of carbon nanoallotropes: Classification, chemistry, and applications of fullerenes, carbon dots, nanotubes, graphene, nanodiamonds, and combined superstructures. Chem. Rev. 2015, 115, 4744–4822. [Google Scholar] [CrossRef]
  26. Adorinni, S.; Cringoli, M.C.; Perathoner, S.; Fornasiero, P.; Marchesan, S. Green approaches to carbon nanostructure-based biomaterials. Appl. Sci. 2021, 11, 2490. [Google Scholar] [CrossRef]
  27. Ugarte, D. Onion-like graphitic particles. Carbon 1995, 33, 989–993. [Google Scholar] [CrossRef]
  28. Marchesan, S.; Melchionna, M.; Prato, M. Carbon nanostructures for nanomedicine: Opportunities and challenges. Fuller. Nanotub. Carbon Nanostructures 2014, 22, 190–195. [Google Scholar] [CrossRef]
  29. Oliveira, S.F.; Bisker, G.; Bakh, N.A.; Gibbs, S.L.; Landry, M.P.; Strano, M.S. Protein functionalized carbon nanomaterials for biomedical applications. Carbon 2015, 95, 767–779. [Google Scholar] [CrossRef] [Green Version]
  30. Mochalin, V.N.; Shenderova, O.; Ho, D.; Gogotsi, Y. The properties and applications of nanodiamonds. Nat. Nanotechnol. 2012, 7, 11–23. [Google Scholar] [CrossRef] [PubMed]
  31. Basso, L.; Cazzanelli, M.; Orlandi, M.; Miotello, A. Nanodiamonds: Synthesis and application in sensing, catalysis, and the possible connection with some processes occurring in space. Appl. Sci. 2020, 10, 4094. [Google Scholar] [CrossRef]
  32. Tonellato, M.; Piccione, M.; Gasparotto, M.; Bellet, P.; Tibaudo, L.; Vicentini, N.; Bergantino, E.; Menna, E.; Vitiello, L.; Di Liddo, R.; et al. Commitment of autologous human multipotent stem cells on biomimetic poly-L-lactic acid-based scaffolds is strongly influenced by structure and concentration of carbon nanomaterial. Nanomaterials 2020, 10, 415. [Google Scholar] [CrossRef] [Green Version]
  33. Piovesana, S.; Iglesias, D.; Melle-Franco, M.; Kralj, S.; Cavaliere, C.; Melchionna, M.; Laganà, A.; Capriotti, A.L.; Marchesan, S. Carbon nanostructure morphology templates nanocomposites for phosphoproteomics. Nano Res. 2020, 13, 380–388. [Google Scholar] [CrossRef]
  34. Vicentini, N.; Gatti, T.; Salerno, M.; Gomez, Y.S.H.; Bellon, M.; Gallio, S.; Marega, C.; Filippini, F.; Menna, E. Effect of different functionalized carbon nanostructures as fillers on the physical properties of biocompatible poly (l-lactic acid) composites. Mater. Chem. Phys. 2018, 214, 265–276. [Google Scholar] [CrossRef]
  35. Iglesias, D.; Melle-Franco, M.; Kurbasic, M.; Melchionna, M.; Abrami, M.; Grassi, M.; Prato, M.; Marchesan, S. Oxidized nanocarbons-tripeptide supramolecular hydrogels: Shape matters! ACS Nano 2018, 12, 5530–5538. [Google Scholar] [CrossRef]
  36. Eissa, S.; Alshehri, N.; Rahman, A.M.A.; Dasouki, M.; Abu-Salah, K.M.; Zourob, M. Electrochemical immunosensors for the detection of survival motor neuron (SMN) protein using different carbon nanomaterials-modified electrodes. Biosens. Bioelectron. 2018, 101, 282–289. [Google Scholar] [CrossRef]
  37. Iglesias, D.; Bosi, S.; Melchionna, M.; Da Ros, T.; Marchesan, S. The glitter of carbon nanostructures in hybrid/composite hydrogels for medicinal use. Curr. Top. Med. Chem. 2016, 16, 1976–1989. [Google Scholar] [CrossRef] [Green Version]
  38. Maziukiewicz, D.; Maciejewska, B.M.; Litowczenko, J.; Kościński, M.; Warowicka, A.; Wychowaniec, J.K.; Jurga, S. Designing biocompatible spin-coated multiwall carbon nanotubes-polymer composite coatings. Surf. Coat. Technol. 2020, 385, 125199. [Google Scholar] [CrossRef]
  39. Tadyszak, K.; Wychowaniec, J.K.; Litowczenko, J. Biomedical applications of graphene-based structures. Nanomaterials 2018, 8, 944. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  40. Tofail, S.A.M.; Koumoulos, E.P.; Bandyopadhyay, A.; Bose, S.; O’Donoghue, L.; Charitidis, C. Additive manufacturing: Scientific and technological challenges, market uptake and opportunities. Mater. Today 2018, 21, 22–37. [Google Scholar] [CrossRef]
  41. Marchesan, S.; Ballerini, L.; Prato, M. Nanomaterials for stimulating nerve growth. Science 2017, 356, 1010–1011. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  42. Aydin, T.; Gurcan, C.; Taheri, H.; Yilmazer, A. Graphene based materials in neural tissue regeneration. Adv. Exp. Med. Biol. 2018, 1107, 129–142. [Google Scholar] [CrossRef] [PubMed]
  43. Bei, H.P.; Yang, Y.; Zhang, Q.; Tian, Y.; Luo, X.; Yang, M.; Zhao, X. Graphene-based nanocomposites for neural tissue engineering. Molecules 2019, 24, 658. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  44. Farokhi, M.; Mottaghitalab, F.; Saeb, M.R.; Shojaei, S.; Zarrin, N.K.; Thomas, S.; Ramakrishna, S. Conductive biomaterials as substrates for neural stem cells differentiation towards neuronal lineage cells. Macromol. Biosci. 2021, 21, e2000123. [Google Scholar] [CrossRef] [PubMed]
  45. Monaco, A.M.; Giugliano, M. Carbon-based smart nanomaterials in biomedicine and neuroengineering. Beilstein J. Nanotechnol. 2014, 5, 1849–1863. [Google Scholar] [CrossRef]
  46. Amin, D.R.; Sink, E.; Narayan, S.P.; Abdel-Hafiz, M.; Mestroni, L.; Peña, B. Nanomaterials for cardiac tissue engineering. Molecules 2020, 25, 5189. [Google Scholar] [CrossRef]
  47. Marchesan, S.; Bosi, S.; Alshatwi, A.; Prato, M. Carbon nanotubes for organ regeneration: An electrifying performance. Nano Today 2016, 11, 398–401. [Google Scholar] [CrossRef]
  48. Ashtari, K.; Nazari, H.; Ko, H.; Tebon, P.; Akhshik, M.; Akbari, M.; Alhosseini, S.N.; Mozafari, M.; Mehravi, B.; Soleimani, M.; et al. Electrically conductive nanomaterials for cardiac tissue engineering. Adv. Drug Deliv. Rev. 2019, 144, 162–179. [Google Scholar] [CrossRef]
  49. Ghai, P.; Mayerhofer, T.; Jha, R.K. Exploring the effectiveness of incorporating carbon nanotubes into bioengineered scaffolds to improve cardiomyocyte function. Expert Rev. Clin. Pharmacol. 2020, 13, 1347–1366. [Google Scholar] [CrossRef] [PubMed]
  50. Mousavi, A.; Vahdat, S.; Baheiraei, N.; Razavi, M.; Norahan, M.H.; Baharvand, H. Multifunctional conductive biomaterials as promising platforms for cardiac tissue engineering. ACS Biomater. Sci. Eng. 2021, 7, 55–82. [Google Scholar] [CrossRef] [PubMed]
  51. Esmaeili, H.; Patino-Guerrero, A.; Hasany, M.; Ansari, M.O.; Memic, A.; Dolatshahi-Pirouz, A.; Nikkhah, M. Electroconductive biomaterials for cardiac tissue engineering. Acta Biomater. 2021; in press. [Google Scholar] [CrossRef] [PubMed]
  52. Sainio, S.; Leppänen, E.; Mynttinen, E.; Palomäki, T.; Wester, N.; Etula, J.; Isoaho, N.; Peltola, E.; Koehne, J.; Meyyappan, M.; et al. Integrating carbon nanomaterials with metals for bio-sensing applications. Mol. Neurobiol. 2020, 57, 179–190. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  53. Melchionna, M.; Prato, M.; Fornasiero, P. Mix and match metal oxides and nanocarbons for new photocatalytic frontiers. Catal. Today 2016, 277, 202–213. [Google Scholar] [CrossRef]
  54. Rosso, C.; Filippini, G.; Criado, A.; Melchionna, M.; Fornasiero, P.; Prato, M. Metal-free photocatalysis: Two-dimensional nanomaterial connection toward advanced organic synthesis. ACS Nano 2021, 15, 3621–3630. [Google Scholar] [CrossRef]
  55. Melchionna, M.; Fornasiero, P.; Prato, M. Into the carbon: A matter of core and shell in advanced electrocatalysis. APL Mater. 2020, 8, 020905. [Google Scholar] [CrossRef] [Green Version]
  56. Antonietti, M.; Bandosz, T.; Centi, G.; Costa, R.; Cruz-Silva, R.; Di, J.; Feng, X.; Frank, B.; Gebhardt, P.; Guldi, D.M.; et al. Nanocarbon-Inorganic Hybrids: Next Generation Composites for Sustainable Energy Applications; Eder, D., Schlogl, R., Eds.; Walter de Gruyter GmbH & Co KG: Berlin, Germany, 2014. [Google Scholar]
  57. Lu, H.; Tournet, J.; Dastafkan, K.; Liu, Y.; Ng, Y.H.; Karuturi, S.K.; Zhao, C.; Yin, Z. Noble-metal-free multicomponent nanointegration for sustainable energy conversion. Chem. Rev. 2021, 121, 10271–10366. [Google Scholar] [CrossRef]
  58. Chen, S.; Qiu, L.; Cheng, H.-M. Carbon-based fibers for advanced electrochemical energy storage devices. Chem. Rev. 2020, 120, 2811–2878. [Google Scholar] [CrossRef]
  59. Liu, H.; Wu, S.; Tian, N.; Yan, F.; You, C.; Yang, Y. Carbon foams: 3D porous carbon materials holding immense potential. J. Mater. Chem. A 2020, 8, 23699–23723. [Google Scholar] [CrossRef]
  60. Tian, W.; Zhang, H.; Duan, X.; Sun, H.; Shao, G.; Wang, S. Porous carbons: Structure-oriented design and versatile applications. Adv. Funct. Mater. 2020, 30, 1909265. [Google Scholar] [CrossRef]
  61. Perovic, M.; Qin, Q.; Oschatz, M. From molecular precursors to nanoparticles—Tailoring the adsorption properties of porous carbon materials by controlled chemical functionalization. Adv. Funct. Mater. 2020, 30, 1908371. [Google Scholar] [CrossRef]
  62. Wang, T.; Okejiri, F.; Qiao, Z.-A.; Dai, S. Tailoring polymer colloids derived porous carbon spheres based on specific chemical reactions. Adv. Mater. 2020, 32, e2002475. [Google Scholar] [CrossRef]
  63. Niu, J.; Shao, R.; Liu, M.; Zan, Y.; Dou, M.; Liu, J.; Zhang, Z.; Huang, Y.; Wang, F. Porous carbons derived from collagen-enriched biomass: Tailored design, synthesis, and application in electrochemical energy storage and conversion. Adv. Funct. Mater. 2019, 29, 1905095. [Google Scholar] [CrossRef]
  64. Zhang, Z.; Cano, Z.P.; Luo, D.; Dou, H.; Yu, A.; Chen, Z. Rational design of tailored porous carbon-based materials for CO2 capture. J. Mater. Chem. A 2019, 7, 20985–21003. [Google Scholar] [CrossRef]
  65. Singh, G.; Lakhi, K.S.; Sil, S.; Bhosale, S.V.; Kim, I.; Albahily, K.; Vinu, A. Biomass derived porous carbon for CO2 capture. Carbon 2019, 148, 164–186. [Google Scholar] [CrossRef]
  66. Kumar, K.V.; Preuss, K.; Titirici, M.-M.; Rodriguez-Reinoso, F. Nanoporous materials for the onboard storage of natural gas. Chem. Rev. 2017, 117, 1796–1825. [Google Scholar] [CrossRef]
  67. Bi, Z.; Kong, Q.; Cao, Y.; Sun, G.; Su, F.; Wei, X.; Li, X.; Ahmad, A.; Xie, L.; Chen, C.-M. Biomass-derived porous carbon materials with different dimensions for supercapacitor electrodes: A review. J. Mater. Chem. A 2019, 7, 16028–16045. [Google Scholar] [CrossRef]
  68. Liu, T.; Liu, G. Block copolymer-based porous carbons for supercapacitors. J. Mater. Chem. A 2019, 7, 23476–23488. [Google Scholar] [CrossRef]
  69. Tian, H.; Wang, T.; Zhang, F.; Zhao, S.; Wan, S.; He, F.; Wang, G. Tunable porous carbon spheres for high-performance rechargeable batteries. J. Mater. Chem. A 2018, 6, 12816–12841. [Google Scholar] [CrossRef]
  70. Xu, M.; Yu, Q.; Liu, Z.; Lv, J.; Lian, S.; Hu, B.; Mai, L.; Zhou, L. Tailoring porous carbon spheres for supercapacitors. Nanoscale 2018, 10, 21604–21616. [Google Scholar] [CrossRef] [PubMed]
  71. He, Y.; Zhuang, X.; Lei, C.; Lei, L.; Hou, Y.; Mai, Y.; Feng, X. Porous carbon nanosheets: Synthetic strategies and electrochemical energy related applications. Nano Today 2019, 24, 103–119. [Google Scholar] [CrossRef]
  72. Matos, I.; Bernardo, M.; Fonseca, I. Porous carbon: A versatile material for catalysis. Catal. Today 2017, 285, 194–203. [Google Scholar] [CrossRef]
  73. Cao, Y.; Mao, S.; Li, M.; Chen, Y.; Wang, Y. Metal/porous carbon composites for heterogeneous catalysis: Old catalysts with improved performance promoted by N-doping. ACS Catal. 2017, 7, 8090–8112. [Google Scholar] [CrossRef]
  74. Benzigar, M.R.; Talapaneni, S.N.; Joseph, S.; Ramadass, K.; Singh, G.; Scaranto, J.; Ravon, U.; Al-Bahily, K.; Vinu, A. Recent advances in functionalized micro and mesoporous carbon materials: Synthesis and applications. Chem. Soc. Rev. 2018, 47, 2680–2721. [Google Scholar] [CrossRef]
  75. Kroto, H.W.; Heath, J.R.; O’Brien, S.C.; Curl, R.F.; Smalley, R.E. C60: Buckminsterfullerene. Nature 1985, 318, 162–163. [Google Scholar] [CrossRef]
  76. Diederich, F.; Ettl, R.; Rubin, Y.; Whetten, R.L.; Beck, R.; Alvarez, M.; Anz, S.; Sensharma, D.; Wudl, F.; Khemani, K.C.; et al. The higher fullerenes: Isolation and characterization of C76, C84, C90, C94, and C70O, an oxide of D5h-C70. Science 1991, 252, 548–551. [Google Scholar] [CrossRef]
  77. Howard, J.B.; McKinnon, J.T.; Makarovsky, Y.; Lafleur, A.L.; Johnson, M.E. Fullerenes C60 and C70 in flames. Nature 1991, 352, 139–141. [Google Scholar] [CrossRef]
  78. Krätschmer, W.; Lamb, L.D.; Fostiropoulos, K.; Huffman, D.R. Solid C60: A new form of carbon. Nature 1990, 347, 354–358. [Google Scholar] [CrossRef]
  79. Guldi, D.M.; Martin, N. 1.07—Functionalized fullerenes: Synthesis and functions. In Comprehensive Nanoscience and Nanotechnology, 2nd ed.; Andrews, D.L., Lipson, R.H., Nann, T., Eds.; Academic Press: Oxford, UK, 2011; pp. 187–204. [Google Scholar]
  80. Von Delius, M.; Hirsch, A. Heterofullerenes: Doped buckyballs. In Chemical Synthesis and Applications of Graphene and Carbon Materials; Antonietti, M., Müllen, K., Eds.; Wiley: Hoboken, NJ, USA, 2017; pp. 191–216. [Google Scholar]
  81. Ceron, M.R.; Maffeis, V.; Stevenson, S.; Echegoyen, L. Endohedral fullerenes: Synthesis, isolation, mono- and bis-functionalization. Inorg. Chim. Acta 2017, 468, 16–27. [Google Scholar] [CrossRef]
  82. Jalife, S.; Arcudia, J.; Pan, S.; Merino, G. Noble gas endohedral fullerenes. Chem. Sci. 2020, 11, 6642–6652. [Google Scholar] [CrossRef] [PubMed]
  83. Candian, A. A fresh mechanism for how buckyballs form in space. Nature 2019, 574, 490–491. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  84. Kazemzadeh, H.; Mozafari, M. Fullerene-based delivery systems. Drug Discov. Today 2019, 24, 898–905. [Google Scholar] [CrossRef]
  85. Illescas, B.M.; Rojo, J.; Delgado, R.; Martin, N. Multivalent glycosylated nanostructures to inhibit ebola virus infection. J. Am. Chem. Soc. 2017, 139, 6018–6025. [Google Scholar] [CrossRef] [Green Version]
  86. Goodarzi, S.; Da Ros, T.; Conde, J.; Sefat, F.; Mozafari, M. Fullerene: Biomedical engineers get to revisit an old friend. Mater. Today 2017, 20, 460–480. [Google Scholar] [CrossRef] [Green Version]
  87. Li, J.; Chen, L.; Su, H.; Yan, L.; Gu, Z.; Chen, Z.; Zhang, A.; Zhao, F.; Zhao, Y. The pharmaceutical multi-activity of metallofullerenol invigorates cancer therapy. Nanoscale 2019, 11, 14528–14539. [Google Scholar] [CrossRef]
  88. Collavini, S.; Delgado, J.L. Fullerenes: The stars of photovoltaics. Sustain. Energy Fuels 2018, 2, 2480–2493. [Google Scholar] [CrossRef]
  89. Deng, L.-L.; Xie, S.-Y.; Gao, F. Fullerene-based materials for photovoltaic applications: Toward efficient, hysteresis-free, and stable perovskite solar cells. Adv. Electron. Mater. 2018, 4, 1700435. [Google Scholar] [CrossRef] [Green Version]
  90. Camisasca, A.; Giordani, S. Carbon nano-onions for bioimaging and cancer therapy applications. In Nanooncology; Springer: Berlin/Heidelberg, Germany, 2018; pp. 417–455. [Google Scholar] [CrossRef]
  91. Palkar, A.; Melin, F.; Cardona, C.M.; Elliott, B.; Naskar, A.K.; Edie, D.D.; Kumbhar, A.; Echegoyen, L. Reactivity differences between carbon nano onions (CNOs) prepared by different methods. Chem.-Asian J. 2007, 2, 625–633. [Google Scholar] [CrossRef]
  92. Camisasca, A.; Giordani, S. Carbon nano-onions in biomedical applications: Promising theranostic agents. Inorg. Chim. Acta 2017, 468, 67–76. [Google Scholar] [CrossRef]
  93. Lettieri, S.; Camisasca, A.; d’Amora, M.; Diaspro, A.; Uchida, T.; Nakajima, Y.; Yanagisawa, K.; Maekawa, T.; Giordani, S. Far-red fluorescent carbon nano-onions as a biocompatible platform for cellular imaging. RSC Adv. 2017, 7, 45676–45681. [Google Scholar] [CrossRef] [Green Version]
  94. Zeiger, M.; Jäckel, N.; Mochalin, V.N.; Presser, V. Review: Carbon onions for electrochemical energy storage. J. Mater. Chem. A 2016, 4, 3172–3196. [Google Scholar] [CrossRef] [Green Version]
  95. Camisasca, A.; Sacco, A.; Brescia, R.; Giordani, S. Boron/nitrogen-codoped carbon nano-onion electrocatalysts for the oxygen reduction reaction. ACS Appl. Nano Mater. 2018, 1, 5763–5773. [Google Scholar] [CrossRef]
  96. Marchesano, V.; Ambrosone, A.; Bartelmess, J.; Strisciante, F.; Tino, A.; Echegoyen, L.; Tortiglione, C.; Giordani, S. Impact of carbon nano-onions on Hydra vulgaris as a model organism for nanoecotoxicology. Nanomaterials 2015, 5, 1331–1350. [Google Scholar] [CrossRef] [Green Version]
  97. D’Amora, M.; Maffeis, V.; Brescia, R.; Barnes, D.; Scanlan, E.; Giordani, S. Carbon nano-onions as non-cytotoxic carriers for cellular uptake of glycopeptides and proteins. Nanomaterials 2019, 9, 1069. [Google Scholar] [CrossRef] [Green Version]
  98. Lettieri, S.; d’Amora, M.; Camisasca, A.; Diaspro, A.; Giordani, S. Carbon nano-onions as fluorescent on/off modulated nanoprobes for diagnostics. Beilstein J. Nanotechnol. 2017, 8, 1878–1888. [Google Scholar] [CrossRef]
  99. D’Amora, M.; Camisasca, A.; Boarino, A.; Arpicco, S.; Giordani, S. Supramolecular functionalization of carbon nano-onions with hyaluronic acid-phospholipid conjugates for selective targeting of cancer cells. Colloids Surf. B 2020, 188, 110779. [Google Scholar] [CrossRef]
  100. Bartelmess, J.; Giordani, S. Carbon nano-onions (multi-layer fullerenes): Chemistry and applications. Beilstein J. Nanotechnol. 2014, 5, 1980–1998. [Google Scholar] [CrossRef]
  101. Flavin, K.; Chaur, M.N.; Echegoyen, L.; Giordani, S. Functionalization of multilayer fullerenes (carbon nano-onions) using diazonium compounds and “click” chemistry. Org. Lett. 2010, 12, 840–843. [Google Scholar] [CrossRef]
  102. Bartelmess, J.; Frasconi, M.; Balakrishnan, P.B.; Signorelli, A.; Echegoyen, L.; Pellegrino, T.; Giordani, S. Non-covalent functionalization of carbon nano-onions with pyrene–BODIPY dyads for biological imaging. RSC Adv. 2015, 5, 50253–50258. [Google Scholar] [CrossRef]
  103. Frasconi, M.; Maffeis, V.; Bartelmess, J.; Echegoyen, L.; Giordani, S. Highly surface functionalized carbon nano-onions for bright light bioimaging. Methods Appl. Fluoresc. 2015, 3, 044005. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  104. Revuri, V.; Cherukula, K.; Nafiujjaman, M.; Cho, K.J.; Park, I.-K.; Lee, Y.-K. White-light-emitting carbon nano-onions: A tunable multichannel fluorescent nanoprobe for glutathione-responsive bioimaging. ACS Appl. Nano Mater. 2018, 1, 662–674. [Google Scholar] [CrossRef]
  105. Lin, C.-T.; Lee, C.-Y.; Chiu, H.-T.; Chin, T.-S. Graphene structure in carbon nanocones and nanodiscs. Langmuir 2007, 23, 12806–12810. [Google Scholar] [CrossRef] [PubMed]
  106. Iijima, S.; Yudasaka, M.; Yamada, R.; Bandow, S.; Suenaga, K.; Kokai, F.; Takahashi, K. Nano-aggregates of single-walled graphitic carbon nano-horns. Chem. Phys. Lett. 1999, 309, 165–170. [Google Scholar] [CrossRef]
  107. Yuge, R.; Yudasaka, M.; Toyama, K.; Yamaguchi, T.; Iijima, S.; Manako, T. Buffer gas optimization in CO2 laser ablation for structure control of single-wall carbon nanohorn aggregates. Carbon 2012, 50, 1925–1933. [Google Scholar] [CrossRef]
  108. Liu, X.; Ying, Y.; Ping, J. Structure, synthesis, and sensing applications of single-walled carbon nanohorns. Biosens. Bioelectron. 2020, 167, 112495. [Google Scholar] [CrossRef] [PubMed]
  109. Moreno-Lanceta, A.; Medrano-Bosch, M.; Melgar-Lesmes, P. Single-walled carbon nanohorns as promising nanotube-derived delivery systems to treat cancer. Pharmaceutics 2020, 12, 850. [Google Scholar] [CrossRef] [PubMed]
  110. Kagkoura, A.; Tagmatarchis, N. Carbon nanohorn-based electrocatalysts for energy conversion. Nanomaterials 2020, 10, 1407. [Google Scholar] [CrossRef]
  111. Iglesias, D.; Giuliani, A.; Melchionna, M.; Marchesan, S.; Criado, A.; Nasi, L.; Bevilacqua, M.; Tavagnacco, C.; Vizza, F.; Prato, M.; et al. N-doped graphitized carbon nanohorns as a forefront electrocatalyst in highly selective O2 reduction to H2O2. Chem 2018, 4, 106–123. [Google Scholar] [CrossRef]
  112. Melchionna, M.; Beltram, A.; Montini, T.; Monai, M.; Nasi, L.; Fornasiero, P.; Prato, M. Highly efficient hydrogen production through ethanol photoreforming by a carbon nanocone/Pd@TiO2 hybrid catalyst. Chem. Commun. 2016, 52, 764–767. [Google Scholar] [CrossRef] [PubMed]
  113. Melchionna, M.; Bracamonte, M.V.; Giuliani, A.; Nasi, L.; Montini, T.; Tavagnacco, C.; Bonchio, M.; Fornasiero, P.; Prato, M. Pd@TiO2/carbon nanohorn electrocatalysts: Reversible CO2 hydrogenation to formic acid. Energy Environ. Sci. 2018, 11, 1571–1580. [Google Scholar] [CrossRef] [Green Version]
  114. Liu, J.; Li, R.; Yang, B. Carbon dots: A new type of carbon-based nanomaterial with wide applications. ACS Cent. Sci. 2020, 6, 2179–2195. [Google Scholar] [CrossRef] [PubMed]
  115. Du, J.; Xu, N.; Fan, J.; Sun, W.; Peng, X. Carbon dots for in vivo bioimaging and theranostics. Small 2019, 15, e1805087. [Google Scholar] [CrossRef]
  116. Liu, M.L.; Chen, B.B.; Li, C.M.; Huang, C.Z. Carbon dots: Synthesis, formation mechanism, fluorescence origin and sensing applications. Green Chem. 2019, 21, 449–471. [Google Scholar] [CrossRef]
  117. Anwar, S.; Ding, H.; Xu, M.; Hu, X.; Li, Z.; Wang, J.; Liu, L.; Jiang, L.; Wang, D.; Dong, C.; et al. Recent advances in synthesis, optical properties, and biomedical applications of carbon dots. ACS Appl. Bio Mater. 2019, 2, 2317–2338. [Google Scholar] [CrossRef]
  118. Zhao, P.; Zhu, L. Dispersibility of carbon dots in aqueous and/or organic solvents. Chem. Commun. 2018, 54, 5401–5406. [Google Scholar] [CrossRef]
  119. Sharma, V.; Tiwari, P.; Mobin, S.M. Sustainable carbon-dots: Recent advances in green carbon dots for sensing and bioimaging. J. Mater. Chem. B 2017, 5, 8904–8924. [Google Scholar] [CrossRef]
  120. Zhang, X.; Jiang, M.; Niu, N.; Chen, Z.; Li, S.; Liu, S.; Li, J. Natural-product-derived carbon dots: From natural products to functional materials. ChemSusChem 2018, 11, 11–24. [Google Scholar] [CrossRef]
  121. Bag, P.; Maurya, R.K.; Dadwal, A.; Sarkar, M.; Chawla, P.A.; Narang, R.K.; Kumar, B. Recent development in synthesis of carbon dots from natural resources and their applications in biomedicine and multi-sensing platform. ChemistrySelect 2021, 6, 2774–2789. [Google Scholar] [CrossRef]
  122. Kang, C.; Huang, Y.; Yan, X.F.; Yang, H.; Chen, Z.P. A review of carbon dots produced from biomass wastes. Nanomaterials 2020, 10, 2316. [Google Scholar] [CrossRef] [PubMed]
  123. Shamsipur, M.; Barati, A.; Karami, S. Long-wavelength, multicolor, and white-light emitting carbon-based dots: Achievements made, challenges remaining, and applications. Carbon 2017, 124, 429–472. [Google Scholar] [CrossRef]
  124. Xiong, Y.; Schneider, J.; Ushakova, E.V.; Rogach, A.L. Influence of molecular fluorophores on the research field of chemically synthesized carbon dots. Nano Today 2018, 23, 124–139. [Google Scholar] [CrossRef]
  125. Yao, B.; Huang, H.; Liu, Y.; Kang, Z. Carbon Dots: A small conundrum. Trends Chem. 2019, 1, 235–246. [Google Scholar] [CrossRef]
  126. Righetto, M.; Carraro, F.; Privitera, A.; Marafon, G.; Moretto, A.; Ferrante, C. The elusive nature of carbon nanodot fluorescence: An unconventional perspective. J. Phys. Chem. C 2020, 124, 22314–22320. [Google Scholar] [CrossRef]
  127. He, C.; Xu, P.; Zhang, X.; Long, W. The synthetic strategies, photoluminescence mechanisms and promising applications of carbon dots: Current state and future perspective. Carbon 2022, 186, 91–127. [Google Scholar] [CrossRef]
  128. Rosso, C.; Filippini, G.; Prato, M. Carbon dots as nano-organocatalysts for synthetic applications. ACS Catal. 2020, 10, 8090–8105. [Google Scholar] [CrossRef]
  129. Hutton, G.A.M.; Martindale, B.C.M.; Reisner, E. Carbon dots as photosensitisers for solar-driven catalysis. Chem. Soc. Rev. 2017, 46, 6111–6123. [Google Scholar] [CrossRef] [Green Version]
  130. Akbar, K.; Moretti, E.; Vomiero, A. Carbon dots for photocatalytic degradation of aqueous pollutants: Recent advancements. Adv. Opt. Mater. 2021, 9, 2100532. [Google Scholar] [CrossRef]
  131. Hu, C.; Li, M.; Qiu, J.; Sun, Y.-P. Design and fabrication of carbon dots for energy conversion and storage. Chem. Soc. Rev. 2019, 48, 2315–2337. [Google Scholar] [CrossRef]
  132. Wang, C.; Strauss, V.; Kaner, R.B. Carbon nanodots for capacitor electrodes. Trends Chem. 2019, 1, 858–868. [Google Scholar] [CrossRef]
  133. Reina, G.; Zhao, L.; Bianco, A.; Komatsu, N. Chemical functionalization of nanodiamonds: Opportunities and challenges ahead. Angew. Chem. Int. Ed. 2019, 58, 17918–17929. [Google Scholar] [CrossRef] [PubMed]
  134. Gao, G.; Guo, Q.; Zhi, J. Nanodiamond-based theranostic platform for drug delivery and bioimaging. Small 2019, 15, e1902238. [Google Scholar] [CrossRef] [PubMed]
  135. Prabhakar, N.; Rosenholm, J.M. Nanodiamonds for advanced optical bioimaging and beyond. Curr. Opin. Colloid Interface Sci. 2019, 39, 220–231. [Google Scholar] [CrossRef]
  136. Torelli, M.D.; Nunn, N.A.; Shenderova, O.A. A perspective on fluorescent nanodiamond bioimaging. Small 2019, 15, e1902151. [Google Scholar] [CrossRef] [PubMed]
  137. Whitlow, J.; Pacelli, S.; Paul, A. Multifunctional nanodiamonds in regenerative medicine: Recent advances and future directions. J. Control. Release 2017, 261, 62–86. [Google Scholar] [CrossRef] [PubMed]
  138. Namdar, R.; Nafisi, S. Nanodiamond applications in skin preparations. Drug Discov. Today 2018, 23, 1152–1158. [Google Scholar] [CrossRef]
  139. Kumar, S.; Nehra, M.; Kedia, D.; Dilbaghi, N.; Tankeshwar, K.; Kim, K.-H. Nanodiamonds: Emerging face of future nanotechnology. Carbon 2019, 143, 678–699. [Google Scholar] [CrossRef]
  140. Saraf, J.; Kalia, K.; Bhattacharya, P.; Tekade, R.K. Growing synergy of nanodiamonds in neurodegenerative interventions. Drug Discov. Today 2019, 24, 584–594. [Google Scholar] [CrossRef]
  141. Chipaux, M.; van der Laan, K.J.; Hemelaar, S.R.; Hasani, M.; Zheng, T.; Schirhagl, R. Nanodiamonds and their applications in cells. Small 2018, 14, e1704263. [Google Scholar] [CrossRef]
  142. Hui, Y.Y.; Hsiao, W.W.-W.; Haziza, S.; Simonneau, M.; Treussart, F.; Chang, H.-C. Single particle tracking of fluorescent nanodiamonds in cells and organisms. Curr. Opin. Solid State Mater. Sci. 2017, 21, 35–42. [Google Scholar] [CrossRef]
  143. Van der Laan, K.J.; Hasani, M.; Zheng, T.; Schirhagl, R. Nanodiamonds for in vivo applications. Small 2018, 14, e1703838. [Google Scholar] [CrossRef] [PubMed]
  144. Lin, Y.; Sun, X.; Su, D.S.; Centi, G.; Perathoner, S. Catalysis by hybrid sp2/sp3 nanodiamonds and their role in the design of advanced nanocarbon materials. Chem. Soc. Rev. 2018, 47, 8438–8473. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  145. Iijima, S.; Ichihashi, T. Single-shell carbon nanotubes of 1-nm diameter. Nature 1993, 363, 603–605. [Google Scholar] [CrossRef]
  146. Iijima, S. Helical microtubules of graphitic carbon. Nature 1991, 354, 56–58. [Google Scholar] [CrossRef]
  147. Malik, S.; Marchesan, S. Growth, properties, and applications of branched carbon nanostructures. Nanomaterials 2021, 11, 2728. [Google Scholar] [CrossRef]
  148. Qiu, L.; Ding, F. Understanding single-walled carbon nanotube growth for chirality controllable synthesis. Acc. Mater. Res. 2021, 2, 828–841. [Google Scholar] [CrossRef]
  149. Melchionna, M.; Prato, M. 3—Functionalization of carbon nanotubes. In Nanocarbon-Inorganic Hybrids: Next Generation Composites for Sustainable Energy Applications; Eder, D., Schlögl, R., Eds.; De Gruyter: Berlin, Germany, 2014; pp. 43–70. [Google Scholar] [CrossRef]
  150. Deline, A.R.; Frank, B.P.; Smith, C.L.; Sigmon, L.R.; Wallace, A.N.; Gallagher, M.J.; Goodwin, D.G., Jr.; Durkin, D.P.; Fairbrother, D.H. Influence of oxygen-containing functional groups on the environmental properties, transformations, and toxicity of carbon nanotubes. Chem. Rev. 2020, 120, 11651–11697. [Google Scholar] [CrossRef]
  151. Yang, F.; Wang, M.; Zhang, D.; Yang, J.; Zheng, M.; Li, Y. Chirality pure carbon nanotubes: Growth, sorting, and characterization. Chem. Rev. 2020, 120, 2693–2758. [Google Scholar] [CrossRef]
  152. Qian, L.; Xie, Y.; Zou, M.; Zhang, J. Building a bridge for carbon nanotubes from nanoscale structure to macroscopic application. J. Am. Chem. Soc. 2021, 143, 18805–18819. [Google Scholar] [CrossRef]
  153. Rao, R.; Pint, C.L.; Islam, A.E.; Weatherup, R.S.; Hofmann, S.; Meshot, E.R.; Wu, F.; Zhou, C.; Dee, N.; Amama, P.B.; et al. Carbon nanotubes and related nanomaterials: Critical advances and challenges for synthesis toward mainstream commercial applications. ACS Nano 2018, 12, 11756–11784. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  154. Wu, Y.; Zhao, X.; Shang, Y.; Chang, S.; Dai, L.; Cao, A. Application-driven carbon nanotube functional materials. ACS Nano 2021, 15, 7946–7974. [Google Scholar] [CrossRef] [PubMed]
  155. Bruschi, A.; Donati, D.M.; Choong, P.; Lucarelli, E.; Wallace, G. Dielectric elastomer actuators, neuromuscular interfaces, and foreign body response in artificial neuromuscular prostheses: A review of the literature for an in vivo application. Adv. Healthc. Mater. 2021, 10, 2100041. [Google Scholar] [CrossRef] [PubMed]
  156. Mirvakili, S.M.; Hunter, I.W. Artificial muscles: Mechanisms, applications, and challenges. Adv. Mater. 2018, 30, 1704407. [Google Scholar] [CrossRef] [PubMed]
  157. Gibney, S.; Hicks, J.M.; Robinson, A.; Jain, A.; Sanjuan-Alberte, P.; Rawson, F.J. Toward nanobioelectronic medicine: Unlocking new applications using nanotechnology. WIREs Nanomed. Nanobiotechnol. 2021, 13, e1693. [Google Scholar] [CrossRef]
  158. Wang, C.; Xia, K.; Wang, H.; Liang, X.; Yin, Z.; Zhang, Y. Advanced carbon for flexible and wearable electronics. Adv. Mater. 2019, 31, e1801072. [Google Scholar] [CrossRef]
  159. Farrera, C.; Torres, F.A.; Feliu, N. Carbon nanotubes as optical sensors in biomedicine. ACS Nano 2017, 11, 10637–10643. [Google Scholar] [CrossRef]
  160. Pan, J.; Li, F.; Choi, J.H. Single-walled carbon nanotubes as optical probes for bio-sensing and imaging. J. Mater. Chem. B 2017, 5, 6511–6522. [Google Scholar] [CrossRef]
  161. Aoki, K.; Ogihara, N.; Tanaka, M.; Haniu, H.; Saito, N. Carbon nanotube-based biomaterials for orthopaedic applications. J. Mater. Chem. B 2020, 8, 9227–9238. [Google Scholar] [CrossRef]
  162. Vashist, A.; Kaushik, A.; Sagar, V.; Nair, M.; Vashist, A.; Ghosal, A.; Gupta, Y.K.; Ahmad, S. Advances in carbon nanotubes-hydrogel hybrids in nanomedicine for therapeutics. Adv. Healthc. Mater. 2018, 7, e1701213. [Google Scholar] [CrossRef]
  163. Simon, J.; Flahaut, E.; Simon, J.; Golzio, M. Overview of carbon nanotubes for biomedical applications. Materials 2019, 12, 624. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  164. Negri, V.; Pacheco-Torres, J.; Calle, D.; Lopez-Larrubia, P. Carbon nanotubes in biomedicine. Top. Curr. Chem. 2020, 378, 15. [Google Scholar] [CrossRef] [PubMed]
  165. Liu, Y.; Gao, G.; Vecitis, C.D. Prospects of an electroactive carbon nanotube membrane toward environmental applications. Acc. Chem. Res. 2020, 53, 2892–2902. [Google Scholar] [CrossRef]
  166. Thamaraiselvan, C.; Wang, J.; James, D.K.; Narkhede, P.; Singh, S.P.; Jassby, D.; Tour, J.M.; Arnusch, C.J. Laser-induced graphene and carbon nanotubes as conductive carbon-based materials in environmental technology. Mater. Today 2020, 34, 115–131. [Google Scholar] [CrossRef]
  167. Bati, A.S.R.; Yu, L.; Batmunkh, M.; Shapter, J.G. Recent Advances in Applications of Sorted Single-Walled Carbon nanotubes. Adv. Funct. Mater. 2019, 29, 1902273. [Google Scholar] [CrossRef]
  168. Yang, Z.; Tian, J.; Yin, Z.; Cui, C.; Qian, W.; Wei, F. Carbon nanotube- and graphene-based nanomaterials and applications in high-voltage supercapacitor: A review. Carbon 2019, 141, 467–480. [Google Scholar] [CrossRef]
  169. Boncel, S.; Kolanowska, A.; Kuziel, A.W.; Krzyzewska, I. Carbon nanotube wind turbine blades: How Far are we today from laboratory tests to industrial implementation? ACS Appl. Nano Mater. 2018, 1, 6542–6555. [Google Scholar] [CrossRef]
  170. Jeon, I.; Xiang, R.; Shawky, A.; Matsuo, Y.; Maruyama, S. Single-walled carbon nanotubes in emerging solar cells: Synthesis and electrode applications. Adv. Energy Mater. 2019, 9, 1801312. [Google Scholar] [CrossRef]
  171. Sun, L.; Wang, X.; Wang, Y.; Zhang, Q. Roles of carbon nanotubes in novel energy storage devices. Carbon 2017, 122, 462–474. [Google Scholar] [CrossRef]
  172. He, M.; Zhang, S.; Zhang, J. Horizontal single-walled carbon nanotube arrays: Controlled synthesis, characterizations, and applications. Chem. Rev. 2020, 120, 12592–12684. [Google Scholar] [CrossRef]
  173. Gaviria Rojas, W.A.; Hersam, M.C. Chirality-enriched carbon nanotubes for next-generation computing. Adv. Mater. 2020, 32, 1905654. [Google Scholar] [CrossRef] [PubMed]
  174. Zhao, K.; Quan, X. Carbon-based materials for electrochemical reduction of CO2 to C2+ oxygenates: Recent progress and remaining challenges. ACS Catal. 2021, 11, 2076–2097. [Google Scholar] [CrossRef]
  175. Sheng, J.; Li, Y. Applications of carbon nanotubes in oxygen electrocatalytic reactions. ACS Appl. Mater. Interfaces, 2021; in press. [Google Scholar] [CrossRef] [PubMed]
  176. Paul, R.; Zhu, L.; Chen, H.; Qu, J.; Dai, L. Recent advances in carbon-based metal-free electrocatalysts. Adv. Mater. 2019, 31, e1806403. [Google Scholar] [CrossRef]
  177. Iglesias, D.; Melchionna, M. Enter the tubes: Carbon nanotube endohedral catalysis. Catalysts 2019, 9, 128. [Google Scholar] [CrossRef] [Green Version]
  178. Yang, J.; Ganesan, P.; Ishihara, A.; Nakashima, N. Carbon Nanotube-based non-precious metal electrode catalysts for fuel cells, water splitting and zinc-air batteries. ChemCatChem 2019, 11, 5929–5944. [Google Scholar] [CrossRef]
  179. Chen, Y.; Wei, J.; Duyar, M.S.; Ordomsky, V.V.; Khodakov, A.Y.; Liu, J. Carbon-based catalysts for Fischer-Tropsch synthesis. Chem. Soc. Rev. 2021, 50, 2337–2366. [Google Scholar] [CrossRef]
  180. Wick, P.; Louw-Gaume, A.E.; Kucki, M.; Krug, H.F.; Kostarelos, K.; Fadeel, B.; Dawson, K.A.; Salvati, A.; Vázquez, E.; Ballerini, L.; et al. Classification framework for graphene-based materials. Angew. Chem. Int. Ed. 2014, 53, 7714–7718. [Google Scholar] [CrossRef] [Green Version]
  181. Du, W.; Wu, H.; Chen, H.; Xu, G.; Li, C. Graphene oxide in aqueous and nonaqueous media: Dispersion behaviour and solution chemistry. Carbon 2020, 158, 568–579. [Google Scholar] [CrossRef]
  182. Nishina, Y.; Eigler, S. Chemical and electrochemical synthesis of graphene oxide—A generalized view. Nanoscale 2020, 12, 12731–12740. [Google Scholar] [CrossRef]
  183. Jakhar, R.; Yap, J.E.; Joshi, R. Microwave reduction of graphene oxide. Carbon 2020, 170, 277–293. [Google Scholar] [CrossRef]
  184. De Silva, K.K.H.; Huang, H.-H.; Joshi, R.; Yoshimura, M. Restoration of the graphitic structure by defect repair during the thermal reduction of graphene oxide. Carbon 2020, 166, 74–90. [Google Scholar] [CrossRef]
  185. Liu, J.; Feng, X. Synthetic tailoring of graphene nanostructures with zigzag-edged topologies: Progress and perspectives. Angew. Chem. Int. Ed. 2020, 59, 23386–23401. [Google Scholar] [CrossRef] [PubMed]
  186. Jolly, A.; Miao, D.; Daigle, M.; Morin, J.-F. Emerging bottom-up strategies for the synthesis of graphene nanoribbons and related structures. Angew. Chem. Int. Ed. 2020, 59, 4624–4633. [Google Scholar] [CrossRef] [PubMed]
  187. Huang, X.; Yin, Z.; Wu, S.; Qi, X.; He, Q.; Zhang, Q.; Yan, Q.; Boey, F.; Zhang, H. Graphene-based materials: Synthesis, characterization, properties, and applications. Small 2011, 7, 1876–1902. [Google Scholar] [CrossRef] [PubMed]
  188. Papageorgiou, D.G.; Li, Z.; Liu, M.; Kinloch, I.A.; Young, R.J. Mechanisms of mechanical reinforcement by graphene and carbon nanotubes in polymer nanocomposites. Nanoscale 2020, 12, 2228–2267. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  189. Kim, J.; Lee, Y.; Kang, M.; Hu, L.; Zhao, S.; Ahn, J.-H. 2D Materials for skin-mountable electronic devices. Adv. Mater. 2021, 33, 2005858. [Google Scholar] [CrossRef] [PubMed]
  190. Chen, X.; Shehzad, K.; Gao, L.; Long, M.; Guo, H.; Qin, S.; Wang, X.; Wang, F.; Shi, Y.; Hu, W.; et al. Graphene hybrid structures for integrated and flexible optoelectronics. Adv. Mater. 2020, 32, e1902039. [Google Scholar] [CrossRef] [PubMed]
  191. You, R.; Liu, Y.-Q.; Hao, Y.-L.; Han, D.-D.; Zhang, Y.-L.; You, Z. Laser fabrication of graphene-based flexible electronics. Adv. Mater. 2020, 32, e1901981. [Google Scholar] [CrossRef] [PubMed]
  192. Sattari-Esfahlan, S.M.; Kim, C.-H. Flexible graphene-channel memory devices: A review. ACS Appl. Nano Mater. 2021, 4, 6542–6556. [Google Scholar] [CrossRef]
  193. Kim, S.D.; Sarkar, A.; Ahn, J.-H. Graphene-based nanomaterials for flexible and stretchable batteries. Small 2021, 17, 2006262. [Google Scholar] [CrossRef] [PubMed]
  194. Wang, C.; Muni, M.; Strauss, V.; Borenstein, A.; Chang, X.; Huang, A.; Qu, S.; Sung, K.; Gilham, T.; Kaner, R.B. Graphene’s role in emerging trends of capacitive energy storage. Small 2021, 17, 2006875. [Google Scholar] [CrossRef] [PubMed]
  195. Zhang, H.; Yang, D.; Lau, A.; Ma, T.; Lin, H.; Jia, B. Hybridized graphene for supercapacitors: Beyond the limitation of pure graphene. Small 2021, 17, 2007311. [Google Scholar] [CrossRef] [PubMed]
  196. Han, J.; Li, H.; Yang, Q.-H. Compact energy storage enabled by graphenes: Challenges, strategies and progress. Mater. Today 2021, 51, 552–565. [Google Scholar] [CrossRef]
  197. Lu, B.; Jin, X.; Han, Q.; Qu, L. Planar graphene-based microsupercapacitors. Small 2021, 17, 2006827. [Google Scholar] [CrossRef]
  198. Kumar, R.; Sahoo, S.; Joanni, E.; Singh, R.K.; Maegawa, K.; Tan, W.K.; Kawamura, G.; Kar, K.K.; Matsuda, A. Heteroatom doped graphene engineering for energy storage and conversion. Mater. Today 2020, 39, 47–65. [Google Scholar] [CrossRef]
  199. Bhol, P.; Yadav, S.; Altaee, A.; Saxena, M.; Misra, P.K.; Samal, A.K. Graphene-based membranes for water and wastewater treatment: A review. ACS Appl. Nano Mater. 2021, 4, 3274–3293. [Google Scholar] [CrossRef]
  200. Zhao, G.; Zhu, H. Cation-π interactions in graphene-containing systems for water treatment and beyond. Adv. Mater. 2020, 32, 1905756. [Google Scholar] [CrossRef]
  201. Ahmad, M.S.; Nishina, Y. Graphene-based carbocatalysts for carbon-carbon bond formation. Nanoscale 2020, 12, 12210–12227. [Google Scholar] [CrossRef]
  202. Cui, H.; Guo, Y.; Zhou, Z. Three-dimensional graphene-based macrostructures for electrocatalysis. Small 2021, 17, 2005255. [Google Scholar] [CrossRef]
  203. Bie, C.; Yu, H.; Cheng, B.; Ho, W.; Fan, J.; Yu, J. Design, fabrication, and mechanism of nitrogen-doped graphene-based photocatalyst. Adv. Mater. 2021, 33, e2003521. [Google Scholar] [CrossRef] [PubMed]
  204. Zhuo, H.-Y.; Zhang, X.; Liang, J.-X.; Yu, Q.; Xiao, H.; Li, J. Theoretical understandings of graphene-based metal single-atom catalysts: Stability and catalytic performance. Chem. Rev. 2020, 120, 12315–12341. [Google Scholar] [CrossRef] [PubMed]
  205. Reina, G.; Iglesias, D.; Samori, P.; Bianco, A. Graphene: A disruptive opportunity for COVID-19 and future pandemics? Adv. Mater. 2021, 33, 2007847. [Google Scholar] [CrossRef] [PubMed]
  206. Bellet, P.; Gasparotto, M.; Pressi, S.; Fortunato, A.; Scapin, G.; Mba, M.; Menna, E.; Filippini, F. Graphene-based scaffolds for regenerative medicine. Nanomaterials 2021, 11, 404. [Google Scholar] [CrossRef] [PubMed]
  207. Zhang, X.; Jing, Q.; Ao, S.; Schneider, G.F.; Kireev, D.; Zhang, Z.; Fu, W. Ultrasensitive Field-Effect Biosensors Enabled by the Unique Electronic Properties of Graphene. Small 2020, 16, e1902820. [Google Scholar] [CrossRef] [PubMed]
  208. Zheng, Q.; Lee, J.-H.; Shen, X.; Chen, X.; Kim, J.-K. Graphene-based wearable piezoresistive physical sensors. Mater. Today 2020, 36, 158–179. [Google Scholar] [CrossRef]
  209. Wyss, K.M.; Luong, D.X.; Tour, J.M. Large-scale syntheses of 2-D Materials: Flash joule heating and other methods. Adv. Mater. 2021, 2106970. [Google Scholar] [CrossRef] [PubMed]
  210. Sun, Z.; Hu, Y.H. Ultrafast, low-cost, and mass production of high-quality graphene. Angew. Chem. Int. Ed. 2020, 59, 9232–9234. [Google Scholar] [CrossRef]
  211. Zhu, Y.; Qu, B.; Andreeva, D.V.; Ye, C.; Novoselov, K.S. Graphene standardization: The lesson from the East. Mater. Today 2021, 47, 9–15. [Google Scholar] [CrossRef]
  212. Adorinni, S.; Rozhin, P.; Marchesan, S. Smart hydrogels meet carbon nanomaterials for new frontiers in medicine. Biomedicines 2021, 9, 570. [Google Scholar] [CrossRef] [PubMed]
  213. Rahamathulla, M.; Bhosale, R.R.; Osmani, R.A.M.; Mahima, K.C.; Johnson, A.P.; Hani, U.; Ghazwani, M.; Begum, M.Y.; Alshehri, S.; Ghoneim, M.M.; et al. Carbon nanotubes: Current perspectives on diverse applications in targeted drug delivery and therapies. Materials 2021, 14, 6707. [Google Scholar] [CrossRef] [PubMed]
  214. Aoki, K.; Saito, N. Biocompatibility and carcinogenicity of carbon nanotubes as biomaterials. Nanomaterials 2020, 10, 264. [Google Scholar] [CrossRef] [PubMed]
  215. Saleemi, M.A.; Fouladi, M.H.; Yong, P.V.C.; Chinna, K.; Palanisamy, N.K.; Wong, E.H. Toxicity of carbon nanotubes: Molecular mechanisms, signaling cascades, and remedies in biomedical applications. Chem. Res. Toxicol. 2021, 34, 24–46. [Google Scholar] [CrossRef] [PubMed]
  216. Heller, D.A.; Jena, P.V.; Pasquali, M.; Kostarelos, K.; Delogu, L.G.; Meidl, R.E.; Rotkin, S.V.; Scheinberg, D.A.; Schwartz, R.E.; Terrones, M.; et al. Banning carbon nanotubes would be scientifically unjustified and damaging to innovation. Nat. Nanotechnol. 2020, 15, 164–166. [Google Scholar] [CrossRef] [PubMed]
  217. Zygouri, P.; Spyrou, K.; Mitsari, E.; Barrio, M.; Macovez, R.; Patila, M.; Stamatis, H.; Verginadis, I.I.; Velalopoulou, A.P.; Evangelou, A.M.; et al. A facile approach to hydrophilic oxidized fullerenes and their derivatives as cytotoxic agents and supports for nanobiocatalytic systems. Sci. Rep. 2020, 10, 8244. [Google Scholar] [CrossRef] [PubMed]
  218. Beilinson, R.M.; Yavisheva, A.A.; Medyantseva, E.P.; Budnikov, H.C. Amperometric tyrosinase biosensors modified by nanomaterials of different nature for determining diclofenac. J. Anal. Chem. 2021, 76, 653–659. [Google Scholar] [CrossRef]
  219. Chen, Y.; Guo, S.; Zhao, M.; Zhang, P.; Xin, Z.; Tao, J.; Bai, L. Amperometric DNA biosensor for Mycobacterium tuberculosis detection using flower-like carbon nanotubes-polyaniline nanohybrid and enzyme-assisted signal amplification strategy. Biosens. Bioelectron. 2018, 119, 215–220. [Google Scholar] [CrossRef]
  220. Sharma, S.K.; Micic, M.; Li, S.; Hoar, B.; Paudyal, S.; Zahran, E.M.; Leblanc, R.M. Conjugation of carbon dots with β-galactosidase enzyme: Surface chemistry and use in biosensing. Molecules 2019, 24, 3275. [Google Scholar] [CrossRef] [Green Version]
  221. Singh, R.; Kashayap, S.; Singh, V.; Kayastha, A.M.; Mishra, H.; Saxena, P.S.; Srivastava, A.; Singh, R.K. QPRTase modified N-doped carbon quantum dots: A fluorescent bioprobe for selective detection of neurotoxin quinolinic acid in human serum. Biosens. Bioelectron. 2018, 101, 103–109. [Google Scholar] [CrossRef]
  222. Hetmann, A.; Wujak, M.; Bolibok, P.; Zięba, W.; Wiśniewski, M.; Roszek, K. Novel biocatalytic systems for maintaining the nucleotide balance based on adenylate kinase immobilized on carbon nanostructures. Mater. Sci. Eng. C 2018, 88, 130–139. [Google Scholar] [CrossRef]
  223. Liu, Q.; Ma, C.; Liu, X.-P.; Wei, Y.-P.; Mao, C.-J.; Zhu, J.-J. A novel electrochemiluminescence biosensor for the detection of microRNAs based on a DNA functionalized nitrogen doped carbon quantum dots as signal enhancers. Biosens. Bioelectron. 2017, 92, 273–279. [Google Scholar] [CrossRef] [PubMed]
  224. Zhang, M.; Ma, Y.; Wang, H.; Wang, B.; Zhou, Y.; Liu, Y.; Shao, M.; Huang, H.; Lu, F.; Kang, Z. Chiral control of carbon dots via surface modification for tuning the enzymatic activity of glucose oxidase. ACS Appl. Mater. Interfaces 2021, 13, 5877–5886. [Google Scholar] [CrossRef] [PubMed]
  225. Bravo, I.; Gutierrez-Sanchez, C.; Garcia-Mendiola, T.; Revenga-Parra, M.; Pariente, F.; Lorenzo, E. Enhanced performance of reagent-less carbon nanodots based enzyme electrochemical biosensors. Sensors 2019, 19, 5576. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  226. Zhang, M.; Wang, W.; Wu, F.; Zheng, T.; Ashley, J.; Mohammadniaei, M.; Zhang, Q.; Wang, M.; Li, L.; Shen, J.; et al. Biodegradable poly(γ-glutamic acid)@glucose oxidase@carbon dot nanoparticles for simultaneous multimodal imaging and synergetic cancer therapy. Biomaterials 2020, 252, 120106. [Google Scholar] [CrossRef]
  227. Nashruddin, S.N.A.; Abdullah, J.; Haniff, M.A.S.M.; Zaid, M.H.M.; Choon, O.P.; Wee, M.F.M.R. Label free glucose electrochemical biosensor based on poly(3,4-ethylenedioxy thiophene): Polystyrene sulfonate/titanium carbide/graphene quantum dots. Biosensors 2021, 11, 267. [Google Scholar] [CrossRef]
  228. Cho, M.-J.; Park, S.-Y. Carbon-dot-based ratiometric fluorescence glucose biosensor. Sens. Actuators B 2019, 282, 719–729. [Google Scholar] [CrossRef]
  229. Wu, G.; Gao, Y.; Zhao, D.; Ling, P.; Gao, F. Methanol/oxygen enzymatic biofuel cell using laccase and NAD+-dependent dehydrogenase cascades as biocatalysts on carbon nanodots electrodes. ACS Appl. Mater. Interfaces 2017, 9, 40978–40986. [Google Scholar] [CrossRef]
  230. Zhang, M.; Wang, H.; Wang, B.; Ma, Y.; Huang, H.; Liu, Y.; Shao, M.; Yao, B.; Kang, Z. Maltase decorated by chiral carbon dots with inhibited enzyme activity for glucose level control. Small 2019, 15, e1901512. [Google Scholar] [CrossRef]
  231. Kim, J.; Lee, S.H.; Tieves, F.; Choi, D.S.; Hollmann, F.; Paul, C.E.; Park, C.B. Biocatalytic C=C bond reduction through carbon nanodot-sensitized regeneration of NADH analogues. Angew. Chem. Int. Ed. 2018, 57, 13825–13828. [Google Scholar] [CrossRef]
  232. Bounegru, A.V.; Apetrei, C. Development of a novel electrochemical biosensor based on carbon nanofibers-cobalt phthalocyanine-laccase for the detection of p-coumaric acid in phytoproducts. Int. J. Mol. Sci. 2021, 22, 9302. [Google Scholar] [CrossRef]
  233. Bounegru, A.V.; Apetrei, C. Development of a novel electrochemical biosensor based on carbon nanofibers-gold nanoparticles-tyrosinase for the detection of ferulic acid in cosmetics. Sensors 2020, 20, 6724. [Google Scholar] [CrossRef] [PubMed]
  234. Ford, R.; Devereux, S.J.; Quinn, S.J.; O’Neill, R.D. Carbon nanohorn modified platinum electrodes for improved immobilisation of enzyme in the design of glutamate biosensors. Analyst 2019, 144, 5299–5307. [Google Scholar] [CrossRef] [PubMed]
  235. Ai, Y.; Li, X.; Zhang, L.; Zhong, W.; Wang, J. Highly sensitive electrochemiluminescent immunoassay for neuron-specific enolase amplified by single-walled carbon nanohorns and enzymatic biocatalytic precipitation. J. Electroanal. Chem. 2018, 818, 257–264. [Google Scholar] [CrossRef]
  236. Sok, V.; Fragoso, A. Preparation and characterization of alkaline phosphatase, horseradish peroxidase, and glucose oxidase conjugates with carboxylated carbon nano-onions. Prep. Biochem. Biotechnol. 2018, 48, 136–143. [Google Scholar] [CrossRef] [PubMed]
  237. Mohapatra, J.; Ananthoju, B.; Nair, V.; Mitra, A.; Bahadur, D.; Medhekar, N.V.; Aslam, M. Enzymatic and non-enzymatic electrochemical glucose sensor based on carbon nano-onions. Appl. Surf. Sci. 2018, 442, 332–341. [Google Scholar] [CrossRef]
  238. Sok, V.; Fragoso, A. Carbon nano-onion peroxidase composite biosensor for electrochemical detection of 2,4-D and 2,4,5-T. Appl. Sci. 2021, 11, 6889. [Google Scholar] [CrossRef]
  239. Povedano, E.; Cincotto, F.H.; Parrado, C.; Diez, P.; Sanchez, A.; Canevari, T.C.; Machado, S.A.S.; Pingarron, J.M.; Villalonga, R. Decoration of reduced graphene oxide with rhodium nanoparticles for the design of a sensitive electrochemical enzyme biosensor for 17β-estradiol. Biosens. Bioelectron. 2017, 89, 343–351. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  240. Albishri, H.M.; El-Hady, D.A. Hyphenation of enzyme/graphene oxide-ionic liquid/glassy carbon biosensors with anodic differential pulse stripping voltammetry for reliable determination of choline and acetylcholine in human serum. Talanta 2019, 200, 107–114. [Google Scholar] [CrossRef] [PubMed]
  241. Anton-Millan, N.; Garcia-Tojal, J.; Marty-Roda, M.; Garroni, S.; Cuesta-Lopez, S.; Tamayo-Ramos, J.A. Influence of three commercial graphene derivatives on the catalytic properties of a lactobacillus plantarum α-L-rhamnosidase when used as immobilization matrices. ACS Appl. Mater. Interfaces 2018, 10, 18170–18182. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  242. Kang, Z.; Jiao, K.; Xu, X.; Peng, R.; Jiao, S.; Hu, Z. Graphene oxide-supported carbon nanofiber-like network derived from polyaniline: A novel composite for enhanced glucose oxidase bioelectrode performance. Biosens. Bioelectron. 2017, 96, 367–372. [Google Scholar] [CrossRef]
  243. Fotiadou, R.; Patila, M.; Hammami, M.A.; Enotiadis, A.; Moschovas, D.; Tsirka, K.; Spyrou, K.; Giannelis, E.P.; Avgeropoulos, A.; Paipetis, A.; et al. Development of effective lipase-hybrid nanoflowers enriched with carbon and magnetic nanomaterials for biocatalytic transformations. Nanomaterials 2019, 9, 808. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  244. Pakchin, P.S.; Ghanbari, H.; Saber, R.; Omidi, Y. Electrochemical immunosensor based on chitosan-gold nanoparticle/carbon nanotube as a platform and lactate oxidase as a label for detection of CA125 oncomarker. Biosens. Bioelectron. 2018, 122, 68–74. [Google Scholar] [CrossRef] [PubMed]
  245. Zhang, L.; Liu, Z.; Xie, Q.; Li, Y.; Ying, Y.; Fu, Y. Bioinspired assembly of reduced graphene oxide by fibrin fiber to prepare multi-functional conductive bionanocomposites as versatile electrochemical platforms. Carbon 2019, 153, 504–512. [Google Scholar] [CrossRef]
  246. Magar, H.S.; Ghica, M.E.; Abbas, M.N.; Brett, C.M.A. A novel sensitive amperometric choline biosensor based on multiwalled carbon nanotubes and gold nanoparticles. Talanta 2017, 167, 462–469. [Google Scholar] [CrossRef]
  247. Hyun, K.; Kang, S.; Kim, J.; Kwon, Y. New biocatalyst including a 4-nitrobenzoic acid mediator embedded by the cross-linking of chitosan and genipin and its use in an energy device. ACS Appl. Mater. Interfaces 2020, 12, 23635–23643. [Google Scholar] [CrossRef]
  248. Kim, B.C.; Lee, I.; Kwon, S.-J.; Wee, Y.; Kwon, K.Y.; Jeon, C.; An, H.J.; Jung, H.-T.; Ha, S.; Dordick, J.S.; et al. Fabrication of enzyme-based coatings on intact multi-walled carbon nanotubes as highly effective electrodes in biofuel cells. Sci. Rep. 2017, 7, 40202. [Google Scholar] [CrossRef]
  249. Luo, X.; Shi, W.; Liu, Y.; Sha, P.; Chu, Y.; Cui, Y. A smart tongue depressor-based biosensor for glucose. Sensors 2019, 19, 3864. [Google Scholar] [CrossRef] [Green Version]
  250. Huang, H.; Li, T.; Jiang, M.; Wei, C.; Ma, S.; Chen, D.; Tong, W.; Huang, X. Construction of flexible enzymatic electrode based on gradient hollow fiber membrane and multi-wall carbon tubes meshes. Biosens. Bioelectron. 2020, 152, 112001. [Google Scholar] [CrossRef]
  251. Inamuddin; Shakeel, N.; Ahamed, M.I.; Kanchi, S.; Kashmery, H.A. Green synthesis of ZnO nanoparticles decorated on polyindole functionalized-MCNTs and used as anode material for enzymatic biofuel cell applications. Sci. Rep. 2020, 10, 5052. [Google Scholar] [CrossRef] [Green Version]
  252. Liu, X.; Liu, Y.; Wang, J.; Wei, T.; Dai, Z. Mild hyperthermia-enhanced enzyme-mediated tumor cell chemodynamic therapy. ACS Appl. Mater. Interfaces 2019, 11, 23065–23071. [Google Scholar] [CrossRef]
  253. Gentil, S.; Rousselot-Pailley, P.; Sancho, F.; Robert, V.; Mekmouche, Y.; Guallar, V.; Tron, T.; Le Goff, A. Efficiency of Site-specific clicked laccase-carbon nanotubes biocathodes towards O2 reduction. Chem. Eur. J. 2020, 26, 4798–4804. [Google Scholar] [CrossRef] [PubMed]
  254. Da Ros, T.; Ostric, A.; Andreola, F.; Filocamo, M.; Pietrogrande, M.; Corsolini, F.; Stroppiano, M.; Bruni, S.; Serafino, A.; Fiorito, S. Carbon nanotubes as nanovectors for intracellular delivery of laronidase in Mucopolysaccharidosis type I. Nanoscale 2018, 10, 657–665. [Google Scholar] [CrossRef] [PubMed]
  255. Fan, Y.; Su, F.; Li, K.; Ke, C.; Yan, Y. Carbon nanotube filled with magnetic iron oxide and modified with polyamidoamine dendrimers for immobilizing lipase toward application in biodiesel production. Sci. Rep. 2017, 7, 45643. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  256. Jun, L.Y.; Mubarak, N.M.; Yon, L.S.; Bing, C.H.; Khalid, M.; Jagadish, P.; Abdullah, E.C. Immobilization of peroxidase on functionalized MWCNTs-buckypaper/polyvinyl alcohol nanocomposite membrane. Sci. Rep. 2019, 9, 1–15. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  257. Rezaei, B.; Shoushtari, A.M.; Rabiee, M.; Uzun, L.; Mak, W.C.; Turner, A.P.F. An electrochemical immunosensor for cardiac Troponin I using electrospun carboxylated multi-walled carbon nanotube-whiskered nanofibres. Talanta 2018, 182, 178–186. [Google Scholar] [CrossRef] [PubMed]
  258. Kim, J.H.; Hong, S.-G.; Wee, Y.; Hu, S.; Kwon, Y.; Ha, S.; Kim, J. Enzyme precipitate coating of pyranose oxidase on carbon nanotubes and their electrochemical applications. Biosens. Bioelectron. 2017, 87, 365–372. [Google Scholar] [CrossRef] [PubMed]
  259. Sahin, S.; Wongnate, T.; Chuaboon, L.; Chaiyen, P.; Yu, E.H. Enzymatic fuel cells with an oxygen resistant variant of pyranose-2-oxidase as anode biocatalyst. Biosens. Bioelectron. 2018, 107, 17–25. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  260. Wee, Y.; Park, S.; Kwon, Y.H.; Ju, Y.; Yeon, K.-M.; Kim, J. Tyrosinase-immobilized CNT based biosensor for highly-sensitive detection of phenolic compounds. Biosens. Bioelectron. 2019, 132, 279–285. [Google Scholar] [CrossRef] [PubMed]
  261. Yang, M.; Wang, H.; Liu, P.; Cheng, J. A 3D electrochemical biosensor based on Super-Aligned Carbon NanoTube array for point-of-care uric acid monitoring. Biosens. Bioelectron. 2021, 179, 113082. [Google Scholar] [CrossRef]
  262. Dalkiran, B.; Erden, P.E.; Kilic, E. Amperometric biosensors based on carboxylated multiwalled carbon nanotubes-metal oxide nanoparticles-7,7,8,8-tetracyanoquinodimethane composite for the determination of xanthine. Talanta 2017, 167, 286–295. [Google Scholar] [CrossRef]
  263. Zappi, D.; Caminiti, R.; Ingo, G.M.; Sadun, C.; Tortolini, C.; Antonelli, M.L. Biologically friendly room temperature ionic liquids and nanomaterials for the development of innovative enzymatic biosensors. Talanta 2017, 175, 566–572. [Google Scholar] [CrossRef] [PubMed]
  264. Zappi, D.; Gabriele, S.; Gontrani, L.; Dini, D.; Sadun, C.; Marini, F.; Antonelli, M.L. Biologically friendly room temperature ionic liquids and nanomaterials for the development of innovative enzymatic biosensors: Part II. Talanta 2019, 194, 26–31. [Google Scholar] [CrossRef] [PubMed]
  265. Blazek, T.; Gorski, W. Oxidases, carbon nanotubes, and direct electron transfer: A cautionary tale. Biosens. Bioelectron. 2020, 163, 112260. [Google Scholar] [CrossRef] [PubMed]
  266. Neupane, S.; Patnode, K.; Li, H.; Baryeh, K.; Liu, G.; Hu, J.; Chen, B.; Pan, Y.; Yang, Z. Enhancing enzyme immobilization on carbon nanotubes via metal-organic frameworks for large-substrate biocatalysis. ACS Appl. Mater. Interfaces 2019, 11, 12133–12141. [Google Scholar] [CrossRef] [PubMed]
  267. Gentil, S.; Lalaoui, N.; Dutta, A.; Nedellec, Y.; Cosnier, S.; Shaw, W.J.; Artero, V.; Le Goff, A. Carbon nanotube-supported bio-inspired nickel catalyst and its integration in hybrid hydrogen/air fuel cells. Angew. Chem. Int. Ed. 2017, 56, 1845–1849. [Google Scholar] [CrossRef] [PubMed]
  268. Zelechowska, K.; Trawinski, B.; Draminska, S.; Majdecka, D.; Bilewicz, R.; Kusz, B. Oxygen biosensor based on carbon nanotubes directly grown on graphitic substrate. Sens. Actuators B Chem. 2017, 240, 1308–1313. [Google Scholar] [CrossRef]
  269. Boussema, F.; Gross, A.J.; Hmida, F.; Ayed, B.; Majdoub, H.; Cosnier, S.; Maaref, A.; Holzinger, M. Dawson-type polyoxometalate nanoclusters confined in a carbon nanotube matrix as efficient redox mediators for enzymatic glucose biofuel cell anodes and glucose biosensors. Biosens. Bioelectron. 2018, 109, 20–26. [Google Scholar] [CrossRef] [PubMed]
  270. Dang, X.; Zhao, H. Bimetallic Fe/Mn metal-organic-frameworks and Au nanoparticles anchored carbon nanotubes as a peroxidase-like detection platform with increased active sites and enhanced electron transfer. Talanta 2020, 210, 120678. [Google Scholar] [CrossRef]
  271. Xu, S.; Zhang, Y.; Zhu, Y.; Wu, J.; Li, K.; Lin, G.; Li, X.; Liu, R.; Liu, X.; Wong, C.-P. Facile one-step fabrication of glucose oxidase loaded polymeric nanoparticles decorating MWCNTs for constructing glucose biosensing platform: Structure matters. Biosens. Bioelectron. 2019, 135, 153–159. [Google Scholar] [CrossRef]
  272. Amatatongchai, M.; Sroysee, W.; Chairam, S.; Nacapricha, D. Amperometric flow injection analysis of glucose using immobilized glucose oxidase on nano-composite carbon nanotubes-platinum nanoparticles carbon paste electrode. Talanta 2017, 166, 420–427. [Google Scholar] [CrossRef]
  273. Chen, C.; Ran, R.; Yang, Z.; Lv, R.; Shen, W.; Kang, F.; Huang, Z.-H. An efficient flexible electrochemical glucose sensor based on carbon nanotubes/carbonized silk fabrics decorated with Pt microspheres. Sens. Actuators B Chem. 2018, 256, 63–70. [Google Scholar] [CrossRef]
  274. Da Silva, W.; Ghica, M.E.; Brett, C.M.A. Biotoxic trace metal ion detection by enzymatic inhibition of a glucose biosensor based on a poly(brilliant green)-deep eutectic solvent/carbon nanotube modified electrode. Talanta 2020, 208, 120427. [Google Scholar] [CrossRef] [PubMed]
  275. Zhang, H.; Zhang, L.; Han, Y.; Yu, Y.; Xu, M.; Zhang, X.; Huang, L.; Dong, S. RGO/Au NPs/N-doped CNTs supported on nickel foam as an anode for enzymatic biofuel cells. Biosens. Bioelectron. 2017, 97, 34–40. [Google Scholar] [CrossRef] [PubMed]
  276. Christwardana, M.; Chung, Y.; Kwon, Y. Co-immobilization of glucose oxidase and catalase for enhancing the performance of a membraneless glucose biofuel cell operated under physiological conditions. Nanoscale 2017, 9, 1993–2002. [Google Scholar] [CrossRef]
  277. Kang, Z.; Jiao, K.; Cheng, J.; Peng, R.; Jiao, S.; Hu, Z. A novel three-dimensional carbonized PANI1600@CNTs network for enhanced enzymatic biofuel cell. Biosens. Bioelectron. 2018, 101, 60–65. [Google Scholar] [CrossRef]
  278. Maity, D.; Kumar, R.T.R. Highly sensitive amperometric detection of glutamate by glutamic oxidase immobilized Pt nanoparticle decorated multiwalled carbon nanotubes(MWCNTs)/polypyrrole composite. Biosens. Bioelectron. 2019, 130, 307–314. [Google Scholar] [CrossRef]
  279. Coelho, J.H.; Eisele, A.P.P.; Valezi, C.F.; Mattos, G.J.; Schirmann, J.G.; Dekker, R.F.H.; Barbosa-Dekker, A.M.; Sartori, E.R. Exploring the exocellular fungal biopolymer botryosphaeran for laccase-biosensor architecture and application to determine dopamine and spironolactone. Talanta 2019, 204, 475–483. [Google Scholar] [CrossRef]
  280. Bollella, P.; Sharma, S.; Cass, A.E.G.; Antiochia, R. Microneedle-based biosensor for minimally-invasive lactate detection. Biosens. Bioelectron. 2019, 123, 152–159. [Google Scholar] [CrossRef] [Green Version]
  281. Wan, X.; Xiang, X.; Tang, S.; Yu, D.; Huang, H.; Hu, Y. Immobilization of Candida antarctic lipase B on MWNTs modified by ionic liquids with different functional groups. Colloids Surf. B 2017, 160, 416–422. [Google Scholar] [CrossRef]
  282. Di Tocco, A.; Robledo, S.N.; Osuna, Y.; Sandoval-Cortez, J.; Granero, A.M.; Vettorazzi, N.R.; Martinez, J.L.; Segura, E.P.; Ilina, A.; Zon, M.A.; et al. Development of an electrochemical biosensor for the determination of triglycerides in serum samples based on a lipase/magnetite-chitosan/copper oxide nanoparticles/multiwalled carbon nanotubes/pectin composite. Talanta 2018, 190, 30–37. [Google Scholar] [CrossRef]
  283. Markiton, M.; Boncel, S.; Janas, D.; Chrobok, A. Highly active nanobiocatalyst from lipase noncovalently immobilized on multiwalled carbon nanotubes for baeyer-villiger synthesis of lactones. ACS Sustain. Chem. Eng. 2017, 5, 1685–1691. [Google Scholar] [CrossRef]
  284. Gentil, S.; Mansor, S.M.C.; Jamet, H.; Cosnier, S.; Cavazza, C.; Le Goff, A. Oriented immobilization of [NiFeSe] hydrogenases on covalently and noncovalently functionalized carbon nanotubes for H2/air enzymatic fuel cells. ACS Catal. 2018, 8, 3957–3964. [Google Scholar] [CrossRef]
  285. Franco, J.H.; Klunder, K.J.; Lee, J.; Russell, V.; de Andrade, A.R.; Minteer, S.D. Enhanced electrochemical oxidation of ethanol using a hybrid catalyst cascade architecture containing pyrene-TEMPO, oxalate decarboxylase and carboxylated multi-walled carbon nanotube. Biosens. Bioelectron. 2020, 154, 112077. [Google Scholar] [CrossRef]
  286. Bocanegra-Rodriguez, S.; Molins-Legua, C.; Campins-Falco, P.; Giroud, F.; Gross, A.J.; Cosnier, S. Monofunctional pyrenes at carbon nanotube electrodes for direct electron transfer H2O2 reduction with HRP and HRP-bacterial nanocellulose. Biosens. Bioelectron. 2021, 187, 113304. [Google Scholar] [CrossRef] [PubMed]
  287. Chokkareddy, R.; Bhajanthri, N.K.; Redhi, G.G. An enzyme-induced novel biosensor for the sensitive electrochemical determination of isoniazid. Biosensors 2017, 7, 21. [Google Scholar] [CrossRef] [Green Version]
  288. Yaari, Z.; Cheung, J.M.; Baker, H.A.; Frederiksen, R.S.; Jena, P.V.; Horoszko, C.P.; Jiao, F.; Scheuring, S.; Luo, M.; Heller, D.A. Nanoreporter of an enzymatic suicide inactivation pathway. Nano Lett. 2020, 20, 7819–7827. [Google Scholar] [CrossRef]
  289. Ding, H.; Hu, B.; Zhang, B.; Zhang, H.; Yan, X.; Nie, G.; Liang, M. Carbon-based nanozymes for biomedical applications. Nano Res. 2021, 14, 570–583. [Google Scholar] [CrossRef]
  290. Wu, J.; Wang, X.; Wang, Q.; Lou, Z.; Li, S.; Zhu, Y.; Qin, L.; Wei, H. Nanomaterials with enzyme-like characteristics (nanozymes): Next-generation artificial enzymes (II). Chem. Soc. Rev. 2019, 48, 1004–1076. [Google Scholar] [CrossRef]
  291. Shang, Y.; Liu, F.; Wang, Y.; Li, N.; Ding, B. Enzyme mimic nanomaterials and their biomedical applications. ChemBioChem 2020, 21, 2408–2418. [Google Scholar] [CrossRef]
  292. Liu, X.; Gao, Y.; Chandrawati, R.; Hosta-Rigau, L. Therapeutic applications of multifunctional nanozymes. Nanoscale 2019, 11, 21046–21060. [Google Scholar] [CrossRef]
  293. Wu, Y.; Darland, D.C.; Zhao, J.X. Nanozymes—Hitting the biosensing “target”. Sensors 2021, 21, 5201. [Google Scholar] [CrossRef] [PubMed]
  294. Song, W.; Zhao, B.; Wang, C.; Ozaki, Y.; Lu, X. Functional nanomaterials with unique enzyme-like characteristics for sensing applications. J. Mater. Chem. B 2019, 7, 850–875. [Google Scholar] [CrossRef] [PubMed]
  295. Fan, X.; Yang, F.; Nie, C.; Ma, L.; Cheng, C.; Haag, R. Biocatalytic nanomaterials: A new pathway for bacterial disinfection. Adv. Mater. 2021, 33, 2100637. [Google Scholar] [CrossRef] [PubMed]
  296. Sun, H.; Zhou, Y.; Ren, J.; Qu, X. Carbon nanozymes: Enzymatic Properties, catalytic mechanism, and applications. Angew. Chem. Int. Ed. 2018, 57, 9224–9237. [Google Scholar] [CrossRef] [PubMed]
  297. Hu, Y.; Hojamberdiev, M.; Geng, D. Recent advances in enzyme-free electrochemical hydrogen peroxide sensors based on carbon hybrid nanocomposites. J. Mater. Chem. C 2021, 9, 6970–6990. [Google Scholar] [CrossRef]
  298. Liu, X.; Huang, D.; Lai, C.; Qin, L.; Zeng, G.; Xu, P.; Li, B.; Yi, H.; Zhang, M. Peroxidase-Like activity of smart nanomaterials and their advanced application in colorimetric glucose biosensors. Small 2019, 15, e1900133. [Google Scholar] [CrossRef] [PubMed]
  299. Nothling, M.D.; Xiao, Z.; Bhaskaran, A.; Blyth, M.T.; Bennett, C.W.; Coote, M.L.; Connal, L.A. Synthetic catalysts inspired by hydrolytic enzymes. ACS Catal. 2019, 9, 168–187. [Google Scholar] [CrossRef]
  300. Wang, X.; Wang, H.; Zhou, S. Progress and perspective on carbon-based nanozymes for peroxidase-like applications. J. Phys. Chem. Lett. 2021, 12, 11751–11760. [Google Scholar] [CrossRef]
  301. Li, Y.; Ma, W.; Sun, J.; Lin, M.; Niu, Y.; Yang, X.; Xu, Y. Electrochemical generation of Fe3C/N-doped graphitic carbon nanozyme for efficient wound healing in vivo. Carbon 2020, 159, 149–160. [Google Scholar] [CrossRef]
  302. Lv, Y.; Ma, M.; Huang, Y.; Xia, Y. Carbon dot nanozymes: How to be close to natural enzymes. Chem. Eur. J. 2019, 25, 954–960. [Google Scholar] [CrossRef]
  303. Zhang, J.; Lu, X.; Tang, D.; Wu, S.; Hou, X.; Liu, J.; Wu, P. Phosphorescent carbon dots for highly efficient oxygen photosensitization and as photo-oxidative nanozymes. ACS Appl. Mater. Interfaces 2018, 10, 40808–40814. [Google Scholar] [CrossRef] [PubMed]
  304. Zhang, J.; Wu, S.; Ma, L.; Wu, P.; Liu, J. Graphene oxide as a photocatalytic nuclease mimicking nanozyme for DNA cleavage. Nano Res. 2020, 13, 455–460. [Google Scholar] [CrossRef]
  305. Zhan, Y.; Yang, S.; Luo, F.; Guo, L.; Zeng, Y.; Qiu, B.; Lin, Z. Emission wavelength switchable carbon dots combined with biomimetic inorganic nanozymes for a two-photon fluorescence immunoassay. ACS Appl. Mater. Interfaces 2020, 12, 30085–30094. [Google Scholar] [CrossRef] [PubMed]
  306. Zhang, J.; Wu, S.; Lu, X.; Wu, P.; Liu, J. Manganese as a catalytic mediator for photooxidation and breaking the pH limitation of nanozymes. Nano Lett. 2019, 19, 3214–3220. [Google Scholar] [CrossRef] [PubMed]
  307. Li, F.; Li, S.; Guo, X.; Dong, Y.; Yao, C.; Liu, Y.; Song, Y.; Tan, X.; Gao, L.; Yang, D. Chiral carbon dots mimicking topoisomerase I to mediate the topological rearrangement of supercoiled DNA enantioselectively. Angew. Chem. Int. Ed. 2020, 59, 11087–11092. [Google Scholar] [CrossRef]
  308. Dehvari, K.; Chiu, S.-H.; Lin, J.-S.; Girma, W.M.; Ling, Y.-C.; Chang, J.-Y. Heteroatom doped carbon dots with nanoenzyme like properties as theranostic platforms for free radical scavenging, imaging, and chemotherapy. Acta Biomater. 2020, 114, 343–357. [Google Scholar] [CrossRef]
  309. Chen, J.; Yan, J.; Feng, Q.; Miao, X.; Dou, B.; Wang, P. Label-free and enzyme-free fluorescence detection of microRNA based on sulfydryl-functionalized carbon dots via target-initiated hemin/G-quadruplex-catalyzed oxidation. Biosens. Bioelectron. 2021, 176, 112955. [Google Scholar] [CrossRef]
  310. Gan, H.; Han, W.; Fu, Z.; Wang, L. The chain-like Au/carbon dots nanocomposites with peroxidase-like activity and their application for glucose detection. Colloids Surf. B 2021, 199, 111553. [Google Scholar] [CrossRef]
  311. Xue, T.; Sheng, Y.; Xu, J.; Li, Y.; Lu, X.; Zhu, Y.; Duan, X.; Wen, Y. In-situ reduction of Ag+ on black phosphorene and its NH2-MWCNT nanohybrid with high stability and dispersibility as nanozyme sensor for three ATP metabolites. Biosens. Bioelectron. 2019, 145, 111716. [Google Scholar] [CrossRef]
  312. Zhao, L.; Wu, Z.; Liu, G.; Lu, H.; Gao, Y.; Liu, F.; Wang, C.; Cui, J.; Lu, G. High-activity Mo, S co-doped carbon quantum dot nanozyme-based cascade colorimetric biosensor for sensitive detection of cholesterol. J. Mater. Chem. B 2019, 7, 7042–7051. [Google Scholar] [CrossRef]
  313. Tripathi, K.M.; Ahn, H.T.; Chung, M.; Le, X.A.; Saini, D.; Bhati, A.; Sonkar, S.K.; Kim, M.I.; Kim, T. N, S, and P-co-doped carbon quantum dots: Intrinsic peroxidase activity in a wide pH range and its antibacterial applications. ACS Biomater. Sci. Eng. 2020, 6, 5527–5537. [Google Scholar] [CrossRef] [PubMed]
  314. Xue, S.; Zhang, T.; Wang, X.; Zhang, Q.; Huang, S.; Zhang, L.; Zhang, L.; Zhu, W.; Wang, Y.; Wu, M.; et al. Cu, Zn dopants boost electron transfer of carbon dots for antioxidation. Small 2021, 17, 2102178. [Google Scholar] [CrossRef] [PubMed]
  315. Upadhyay, S.S.; Gadhari, N.S.; Srivastava, A.K. Biomimetic sensor for ethambutol employing β-cyclodextrin mediated chiral copper metal organic framework and carbon nanofibers modified glassy carbon electrode. Biosens. Bioelectron. 2020, 165, 112397. [Google Scholar] [CrossRef] [PubMed]
  316. Lv, S.; Li, Y.; Zhang, K.; Lin, Z.; Tang, D. Carbon dots/g-C3N4 nanoheterostructures-based signal-generation tags for photoelectrochemical immunoassay of cancer biomarkers coupling with copper nanoclusters. ACS Appl. Mater. Interfaces 2017, 9, 38336–38343. [Google Scholar] [CrossRef]
  317. Luo, N.; Yang, Z.; Tang, F.; Wang, D.; Feng, M.; Liao, X.; Yang, X. Fe3O4/carbon nanodot hybrid nanoparticles for the indirect colorimetric detection of glutathione. ACS Appl. Nano Mater. 2019, 2, 3951–3959. [Google Scholar] [CrossRef]
  318. Wang, D.; Song, X.; Li, P.; Gao, X.J.; Gao, X. Origins of the peroxidase mimicking activities of graphene oxide from first principles. J. Mater. Chem. B 2020, 8, 9028–9034. [Google Scholar] [CrossRef]
  319. Kim, M.S.; Cho, S.; Joo, S.H.; Lee, J.; Kwak, S.K.; Kim, M.I.; Lee, J. N- and B-codoped graphene: A strong candidate to replace natural peroxidase in sensitive and selective bioassays. ACS Nano 2019, 13, 4312–4321. [Google Scholar] [CrossRef]
  320. Wang, H.; Li, P.; Yu, D.; Zhang, Y.; Wang, Z.; Liu, C.; Qiu, H.; Liu, Z.; Ren, J.; Qu, X. Unraveling the enzymatic activity of oxygenated carbon nanotubes and their application in the treatment of bacterial infections. Nano Lett. 2018, 18, 3344–3351. [Google Scholar] [CrossRef]
  321. Cheng, N.; Li, J.-C.; Liu, D.; Lin, Y.; Du, D. Single-atom nanozyme based on nanoengineered Fe-N-C catalyst with superior peroxidase-like activity for ultrasensitive bioassays. Small 2019, 15, e1901485. [Google Scholar] [CrossRef]
  322. Wu, W.; Wang, Q.; Chen, J.; Huang, L.; Zhang, H.; Rong, K.; Dong, S. Biomimetic design for enhancing the peroxidase mimicking activity of hemin. Nanoscale 2019, 11, 12603–12609. [Google Scholar] [CrossRef]
  323. Xu, M.; Xing, S.; Zhao, Y.; Zhao, C. Peptide nucleic acid-assisted colorimetric detection of single-nucleotide polymorphisms based on the intrinsic peroxidase-like activity of hemin-carbon nanotube nanocomposites. Talanta 2021, 232, 122420. [Google Scholar] [CrossRef] [PubMed]
  324. Li, X.; Yang, X.-Y.; Sha, J.-Q.; Han, T.; Du, C.-J.; Sun, Y.-J.; Lan, Y.-Q. POMOF/SWNT nanocomposites with prominent peroxidase-mimicking activity for L-cysteine “on-off switch” colorimetric biosensing. ACS Appl. Mater. Interfaces 2019, 11, 16896–16904. [Google Scholar] [CrossRef] [PubMed]
  325. Wang, Z.; Chen, Y.; Dong, W.; Zhou, J.; Han, B.; Jiao, J.; Lan, L.; Miao, P.; Chen, Q. Copper (II)-ploy-L-histidine functionalized multi walled carbon nanotubes as efficient mimetic enzyme for sensitive electrochemical detection of salvianic acid A. Biosens. Bioelectron. 2018, 121, 257–264. [Google Scholar] [CrossRef]
  326. Wu, L.; Lu, Z.; Ye, J. Enzyme-free glucose sensor based on layer-by-layer electrodeposition of multilayer films of multi-walled carbon nanotubes and Cu-based metal framework modified glassy carbon electrode. Biosens. Bioelectron. 2019, 135, 45–49. [Google Scholar] [CrossRef] [PubMed]
  327. He, Z.; Cai, Y.; Yang, Z.; Li, P.; Lei, H.; Liu, W.; Liu, Y. A dual-signal readout enzyme-free immunosensor based on hybridization chain reaction-assisted formation of copper nanoparticles for the detection of microcystin-LR. Biosens. Bioelectron. 2019, 126, 151–159. [Google Scholar] [CrossRef] [PubMed]
  328. Gallay, P.; Eguilaz, M.; Rivas, G. Designing electrochemical interfaces based on nanohybrids of avidin functionalized-carbon nanotubes and ruthenium nanoparticles as peroxidase-like nanozyme with supramolecular recognition properties for site-specific anchoring of biotinylated residues. Biosens. Bioelectron. 2020, 148, 111764. [Google Scholar] [CrossRef] [PubMed]
  329. Xu, D.; Hou, B.; Qian, L.; Zhang, X.; Liu, G. Non-enzymatic electrochemical sensor based on sliver nanoparticle-decorated carbon nanotubes. Molecules 2019, 24, 3411. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  330. Song, H.; Ma, C.; Wang, L.; Zhu, Z. Platinum nanoparticle-deposited multi-walled carbon nanotubes as a NADH oxidase mimic: Characterization and applications. Nanoscale 2020, 12, 19284–19292. [Google Scholar] [CrossRef] [PubMed]
  331. Dong, J.; Lee, M.A.; Rajan, A.G.; Rahaman, I.; Sun, J.H.; Park, M.; Salem, D.P.; Strano, M.S. A synthetic mimic of phosphodiesterase type 5 based on corona phase molecular recognition of single-walled carbon nanotubes. Proc. Natl. Acad. Sci. USA 2020, 117, 26616–26625. [Google Scholar] [CrossRef] [PubMed]
  332. Mikhalchan, A.; Vilatela, J.J. A perspective on high-performance CNT fibres for structural composites. Carbon 2019, 150, 191–215. [Google Scholar] [CrossRef]
  333. Konstantopoulos, G.; Semitekolos, D.; Koumoulos, E.P.; Charitidis, C. Carbon fiber reinforced composites: Study of modification effect on weathering-induced ageing via nanoindentation and deep learning. Nanomaterials 2021, 11, 2631. [Google Scholar] [CrossRef] [PubMed]
  334. Semitekolos, D.; Trompeta, A.-F.; Husarova, I.; Man’ko, T.; Potapov, A.; Romenskaya, O.; Liang, Y.; Li, X.; Giorcelli, M.; Dong, H.; et al. Comparative physical-mechanical properties assessment of tailored surface-treated carbon fibres. Materials 2020, 13, 3136. [Google Scholar] [CrossRef]
  335. Mikhalchan, A.; Vila, M.; Arévalo, L.; Vilatela, J.J. Simultaneous improvements in conversion and properties of molecularly controlled CNT fibres. Carbon 2021, 179, 417–424. [Google Scholar] [CrossRef]
  336. Senokos, E.; Rana, M.; Vila, M.; Fernandez-Cestau, J.; Costa, R.D.; Marcilla, R.; Vilatela, J.J. Transparent and flexible high-power supercapacitors based on carbon nanotube fibre aerogels. Nanoscale 2020, 12, 16980–16986. [Google Scholar] [CrossRef] [PubMed]
  337. Song, N.; Ma, F.; Zhu, Y.; Chen, S.; Wang, C.; Lu, X. Fe3C/nitrogen-doped carbon nanofibers as highly efficient biocatalyst with oxidase-mimicking activity for colorimetric sensing. ACS Sustain. Chem. Eng. 2018, 6, 16766–16776. [Google Scholar] [CrossRef]
  338. Chen, S.; Wang, Y.; Zhong, M.; Yu, D.; Wang, C.; Lu, X. Fe(III)-tannic acid complex derived Fe3C decorated carbon nanofibers for triple-enzyme mimetic activity and their biosensing application. ACS Biomater. Sci. Eng. 2019, 5, 1238–1246. [Google Scholar] [CrossRef] [PubMed]
  339. Iglesias, D.; Senokos, E.; Alemán, B.; Cabana, L.; Navío, C.; Marcilla, R.; Prato, M.; Vilatela, J.J.; Marchesan, S. Gas-Phase functionalization of macroscopic carbon nanotube fiber assemblies: Reaction control, electrochemical properties, and use for flexible supercapacitors. ACS Appl. Mater. Interfaces 2018, 10, 5760–5770. [Google Scholar] [CrossRef] [PubMed]
  340. Yang, M.; Wu, X.; Hu, X.; Wang, K.; Zhang, C.; Gyimah, E.; Yakubu, S.; Zhang, Z. Electrochemical immunosensor based on Ag+-dependent CTAB-AuNPs for ultrasensitive detection of sulfamethazine. Biosens. Bioelectron. 2019, 144, 111643. [Google Scholar] [CrossRef]
  341. Bracamonte, M.V.; Melchionna, M.; Giuliani, A.; Nasi, L.; Tavagnacco, C.; Prato, M.; Fornasiero, P. H2O2 sensing enhancement by mutual integration of single walled carbon nanohorns with metal oxide catalysts: The CeO2 case. Sens. Actuators B 2017, 239, 923–932. [Google Scholar] [CrossRef]
  342. Mykhailiv, O.; Zubyk, H.; Brzezinski, K.; Gras, M.; Lota, G.; Gniadek, M.; Romero, E.; Echegoyen, L.; Plonska-Brzezinska, M.E. Improvement of the structural and chemical properties of carbon nano-onions for electrocatalysis. ChemNanoMat 2017, 3, 583–590. [Google Scholar] [CrossRef]
  343. Chen, T.M.; Tian, X.M.; Huang, L.; Xiao, J.; Yang, G.W. Nanodiamonds as pH-switchable oxidation and reduction catalysts with enzyme-like activities for immunoassay and antioxidant applications. Nanoscale 2017, 9, 15673–15684. [Google Scholar] [CrossRef] [PubMed]
  344. Gülseren, G.; Saylam, A.; Marion, A.; Özçubukçu, S. Fullerene-based mimics of biocatalysts show remarkable activity and modularity. ACS Appl. Mater. Interfaces 2021, 13, 45854–45863. [Google Scholar] [CrossRef] [PubMed]
  345. Zhang, J.; Chen, Z.; Kong, J.; Liang, Y.; Chen, K.; Chang, Y.; Yuan, H.; Wang, Y.; Liang, H.; Li, J.; et al. Fullerenol nanoparticles eradicate Helicobacter pylori via pH-responsive peroxidase activity. ACS Appl. Mater. Interfaces, 2020; in press. [Google Scholar] [CrossRef] [PubMed]
  346. Barzegar, A.; Naghizadeh, E.; Zakariazadeh, M.; Azamat, J. Molecular dynamics simulation study of the HIV-1 protease inhibit ion using fullerene and new fullerene derivatives of carbon nanostructures. Mini-Rev. Med. Chem. 2017, 17, 633–647. [Google Scholar] [CrossRef] [PubMed]
  347. Roy, P.; Bag, S.; Chakraborty, D.; Dasgupta, S. Exploring the inhibitory and antioxidant effects of Fullerene and Fullerenol on Ribonuclease A. ACS Omega 2018, 3, 12270–12283. [Google Scholar] [CrossRef] [Green Version]
  348. Singh, R.K.; Kazansky, Y.; Wathieu, D.; Fushman, D. Hydrophobic patch of Ubiquitin is important for its optimal activation by Ubiquitin activating enzyme E1. Anal. Chem. 2017, 89, 7852–7860. [Google Scholar] [CrossRef] [Green Version]
  349. Liu, Y.; Yan, B.; Winkler, D.A.; Fu, J.; Zhang, A. Competitive inhibition mechanism of acetylcholinesterase without catalytic active site interaction: Study on functionalized C60 nanoparticles via in vitro and in silico assays. ACS Appl. Mater. Interfaces 2017, 9, 18626–18638. [Google Scholar] [CrossRef]
  350. Barros, M.R.; da Silva, L.P.; Menezes, T.M.; Garcia, Y.S.; Neves, J.L. Efficient tyrosinase nano-inhibitor based on carbon dots behaving as a gathering of hydrophobic cores and key chemical group. Colloids Surf. B 2021, 207, 112006. [Google Scholar] [CrossRef]
  351. Di Giosia, M.; Marforio, T.D.; Cantelli, A.; Valle, F.; Zerbetto, F.; Su, Q.; Wang, H.; Calvaresi, M. Inhibition of α-chymotrypsin by pristine single-wall carbon nanotubes: Clogging up the active site. J. Colloid Interface Sci. 2020, 571, 174–184. [Google Scholar] [CrossRef]
  352. Liu, J.J.; Tang, D.; Chen, Z.; Yan, X.; Zhong, Z.; Kang, L.; Yao, J. Chemical redox modulated fluorescence of nitrogen-doped graphene quantum dots for probing the activity of alkaline phosphatase. Biosens. Bioelectron. 2017, 94, 271–277. [Google Scholar] [CrossRef]
  353. Li, H.; Yan, X.; Qiao, S.; Lu, G.; Su, X. Yellow-emissive carbon dot-based optical sensing platforms: Cell imaging and analytical applications for biocatalytic reactions. ACS Appl. Mater. Interfaces 2018, 10, 7737–7744. [Google Scholar] [CrossRef] [PubMed]
  354. Tang, C.; Zhou, J.; Qian, Z.; Ma, Y.; Huang, Y.; Feng, H. A universal fluorometric assay strategy for glycosidases based on functional carbon quantum dots: β-galactosidase activity detection in vitro and in living cells. J. Mater. Chem. B 2017, 5, 1971–1979. [Google Scholar] [CrossRef] [PubMed]
  355. Ao, H.; Feng, H.; Huang, X.; Zhao, M.; Qian, Z. A reversible fluorescence nanoswitch based on dynamic covalent B-O bonds using functional carbon quantum dots and its application for α-glucosidase activity monitoring. J. Mater. Chem. C 2017, 5, 2826–2832. [Google Scholar] [CrossRef]
  356. Sidhu, J.S.; Singh, A.; Garg, N.; Singh, N. Carbon dot based, Naphthalimide coupled FRET pair for highly selective ratiometric detection of thioredoxin reductase and cancer screening. ACS Appl. Mater. Interfaces 2017, 9, 25847–25856. [Google Scholar] [CrossRef] [PubMed]
  357. Shumeiko, V.; Paltiel, Y.; Bisker, G.; Hayouka, Z.; Shoseyov, O. A paper-based near-infrared optical biosensor for quantitative detection of protease activity using peptide-encapsulated SWCNTs. Sensors 2020, 20, 5247. [Google Scholar] [CrossRef] [PubMed]
  358. Palomar, Q.; Xu, X.; Selegaard, R.; Aili, D.; Zhang, Z. Peptide decorated gold nanoparticle/carbon nanotube electrochemical sensor for ultrasensitive detection of matrix metalloproteinase-7. Sens. Actuators B 2020, 325, 128789. [Google Scholar] [CrossRef]
  359. Del Barrio, M.; Rana, M.; Vilatela, J.J.; Lorenzo, E.; De Lacey, A.L.; Pita, M. Photoelectrocatalytic detection of NADH on n-type silicon semiconductors facilitated by carbon nanotube fibers. Electrochim. Acta 2021, 377, 138071. [Google Scholar] [CrossRef]
  360. Nguyen, H.H.; Lee, S.H.; Lee, U.J.; Fermin, C.D.; Kim, M. Immobilized enzymes in biosensor applications. Materials 2019, 12, 121. [Google Scholar] [CrossRef] [Green Version]
  361. Amine, A.; Arduini, F.; Moscone, D.; Palleschi, G. Recent advances in biosensors based on enzyme inhibition. Biosens. Bioelectron. 2016, 76, 180–194. [Google Scholar] [CrossRef]
  362. Maciá-Agulló, J.A.; Corma, A.; Garcia, H. Photobiocatalysis: The power of combining photocatalysis and enzymes. Chem. Eur. J. 2015, 21, 10940–10959. [Google Scholar] [CrossRef] [Green Version]
  363. Kucherenko, I.S.; Soldatkin, O.O.; Kucherenko, D.Y.; Soldatkina, O.V.; Dzyadevych, S.V. Advances in nanomaterial application in enzyme-based electrochemical biosensors: A review. Nanoscale Adv. 2019, 1, 4560–4577. [Google Scholar] [CrossRef] [Green Version]
  364. Lisdat, F. Trends in the layer-by-layer assembly of redox proteins and enzymes in bioelectrochemistry. Curr. Opin. Electrochem. 2017, 5, 165–172. [Google Scholar] [CrossRef]
  365. Zdarta, J.; Meyer, A.S.; Jesionowski, T.; Pinelo, M. Developments in support materials for immobilization of oxidoreductases: A comprehensive review. Adv. Colloid Interface Sci. 2018, 258, 1–20. [Google Scholar] [CrossRef] [PubMed]
  366. Campuzano, S.; Yanez-Sedeno, P.; Pingarro, J.M. Carbon dots and graphene quantum dots in electrochemical biosensing. Nanomaterials 2019, 9, 634. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  367. Cajigas, S.; Orozco, J. Nanobioconjugates for signal amplification in electrochemical biosensing. Molecules 2020, 25, 3542. [Google Scholar] [CrossRef] [PubMed]
  368. Han, Q.; Pang, J.; Li, Y.; Sun, B.; Ibarlucea, B.; Liu, X.; Gemming, T.; Cheng, Q.; Zhang, S.; Liu, H.; et al. Graphene Biodevices for Early Disease Diagnosis Based on Biomarker Detection. ACS Sens. 2021, 6, 3841–3881. [Google Scholar] [CrossRef]
  369. Cumba, L.R.; Camisasca, A.; Giordani, S.; Forster, R.J. Electrochemical properties of screen-printed carbon nano-onion electrodes. Molecules 2020, 25, 3884. [Google Scholar] [CrossRef]
  370. Priyadarshini, E.; Rawat, K. Au@carbon dot nanoconjugates as a dual mode enzyme-free sensing platform for cholesterol. J. Mater. Chem. B 2017, 5, 5425–5432. [Google Scholar] [CrossRef]
  371. Hayat, A.; Haider, W.; Raza, Y.; Marty, J.L. Colorimetric cholesterol sensor based on peroxidase like activity of zinc oxide nanoparticles incorporated carbon nanotubes. Talanta 2015, 143, 157–161. [Google Scholar] [CrossRef]
  372. Isoaho, N.; Peltola, E.; Sainio, S.; Wester, N.; Protopopova, V.; Wilson, B.P.; Koskinen, J.; Laurila, T. Carbon Nanostructure Based Platform for Enzymatic Glutamate Biosensors. J. Phys. Chem. C 2017, 121, 4618–4626. [Google Scholar] [CrossRef]
  373. Bounegru, A.V.; Apetrei, C. Voltamperometric sensors and biosensors based on carbon nanomaterials used for detecting caffeic acid—A review. Int. J. Mol. Sci. 2020, 21, 9275. [Google Scholar] [CrossRef] [PubMed]
  374. Xiao, X.; Xia, H.-Q.; Wu, R.; Bai, L.; Yan, L.; Magner, E.; Cosnier, S.; Lojou, E.; Zhu, Z.; Liu, A. Tackling the challenges of enzymatic (bio)fuel cells. Chem. Rev. 2019, 119, 9509–9558. [Google Scholar] [CrossRef] [PubMed]
  375. Ruth, J.C.; Spormann, A.M. Enzyme electrochemistry for industrial energy applications—A perspective on future areas of focus. ACS Catal. 2021, 11, 5951–5967. [Google Scholar] [CrossRef]
  376. Yin, S.; Jin, Z.; Miyake, T. Wearable high-powered biofuel cells using enzyme/carbon nanotube composite fibers on textile cloth. Biosens. Bioelectron. 2019, 141, 111471. [Google Scholar] [CrossRef] [PubMed]
  377. Pyser, J.B.; Chakrabarty, S.; Romero, E.O.; Narayan, A.R.H. State-of-the-art biocatalysis. ACS Cent. Sci. 2021, 7, 1105–1116. [Google Scholar] [CrossRef] [PubMed]
  378. Yang, Y.; Arnold, F.H. Navigating the unnatural reaction space: Directed evolution of heme proteins for selective Carbene and Nitrene transfer. Acc. Chem. Res. 2021, 54, 1209–1225. [Google Scholar] [CrossRef]
  379. Ramakrishna, T.R.B.; Nalder, T.D.; Yang, W.; Marshall, S.N.; Barrow, C.J. Controlling enzyme function through immobilization on graphene, graphene derivatives and other two dimensional nanomaterials. J. Mater. Chem. B 2018, 6, 3200–3218. [Google Scholar] [CrossRef]
  380. Zor, C.; Reeve, H.A.; Quinson, J.; Thompson, L.A.; Lonsdale, T.H.; Dillon, F.; Grobert, N.; Vincent, K.A. H2-Driven biocatalytic hydrogenation in continuous flow using enzyme-modified carbon nanotube columns. Chem. Commun. 2017, 53, 9839–9841. [Google Scholar] [CrossRef] [Green Version]
  381. Xia, H.; Zeng, J. Rational surface modification of carbon nanomaterials for improved direct electron transfer-type bioelectrocatalysis of redox enzymes. Catalysts 2020, 10, 1447. [Google Scholar] [CrossRef]
  382. Wang, M.; Mohanty, S.K.; Mahendra, S. Nanomaterial-supported enzymes for water purification and monitoring in point-of-use water supply systems. Acc. Chem. Res. 2019, 52, 876–885. [Google Scholar] [CrossRef]
  383. Hilali, N.; Mohammadi, H.; Amine, A.; Zine, N.; Errachid, A. Recent advances in electrochemical monitoring of chromium. Sensors 2020, 20, 5153. [Google Scholar] [CrossRef] [PubMed]
  384. Li, H.; Yan, X.; Lu, G.; Su, X. Carbon dot-based bioplatform for dual colorimetric and fluorometric sensing of organophosphate pesticides. Sens. Actuators B Chem. 2018, 260, 563–570. [Google Scholar] [CrossRef]
  385. Kaspar, C.; Ravoo, B.J.; van der Wiel, W.G.; Wegner, S.V.; Pernice, W.H.P. The rise of intelligent matter. Nature 2021, 594, 345–355. [Google Scholar] [CrossRef] [PubMed]
  386. Cook, A.B.; Decuzzi, P. Harnessing endogenous stimuli for responsive materials in theranostics. ACS Nano 2021, 15, 2068–2098. [Google Scholar] [CrossRef] [PubMed]
  387. Mu, J.; Lin, J.; Huang, P.; Chen, X. Development of endogenous enzyme-responsive nanomaterials for theranostics. Chem. Soc. Rev. 2018, 47, 5554–5573. [Google Scholar] [CrossRef]
  388. Du, B.; Tung, C.-H. Enzyme-assisted photodynamic therapy based on nanomaterials. ACS Biomater. Sci. Eng. 2020, 6, 2506–2517. [Google Scholar] [CrossRef] [PubMed]
  389. He, X.; Sorescu, D.C.; Star, A. Composition and structure of fluorescent graphene quantum dots generated by enzymatic degradation of graphene oxide. J. Phys. Chem. C 2021, 125, 13361–13369. [Google Scholar] [CrossRef]
  390. Parks, A.N.; Chandler, G.T.; Ho, K.T.; Burgess, R.M.; Ferguson, P.L. Environmental biodegradability of [14C] single-walled carbon nanotubes by Trametes versicolor and natural microbial cultures found in New Bedford Harbor sediment and aerated wastewater treatment plant sludge. Environ. Toxicol. Chem. 2015, 34, 247–251. [Google Scholar] [CrossRef]
  391. Zhang, L.; Petersen, E.J.; Habteselassie, M.Y.; Mao, L.; Huang, Q. Degradation of multiwall carbon nanotubes by bacteria. Environ. Pollut. 2013, 181, 335–339. [Google Scholar] [CrossRef] [PubMed]
  392. Chen, M.; Qin, X.; Zeng, G. Biodegradation of carbon nanotubes, graphene, and their derivatives. Trends Biotechnol. 2017, 35, 836–846. [Google Scholar] [CrossRef]
  393. Liu, Z.; Liu, Y.; Zeng, G.; Shao, B.; Chen, M.; Li, Z.; Jiang, Y.; Liu, Y.; Zhang, Y.; Zhong, H. Application of molecular docking for the degradation of organic pollutants in the environmental remediation: A review. Chemosphere 2018, 203, 139–150. [Google Scholar] [CrossRef] [PubMed]
  394. Gonzalez-Lavado, E.; Iturrioz-Rodriguez, N.; Padin-Gonzalez, E.; Gonzalez, J.; Garcia-Hevia, L.; Heuts, J.; Pesquera, C.; Gonzalez, F.; Villegas, J.C.; Valiente, R.; et al. Biodegradable multi-walled carbon nanotubes trigger anti-tumoral effects. Nanoscale 2018, 10, 11013–11020. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  395. Martin, C.; Jun, G.; Schurhammer, R.; Reina, G.; Chen, P.; Bianco, A.; Menard-Moyon, C. Enzymatic degradation of graphene quantum dots by human peroxidases. Small 2019, 15, 1905405. [Google Scholar] [CrossRef]
  396. Piotrovskiy, L.B.; Litasova, E.V.; Sokolov, A.V.; Iljin, V.V.; Utsal, V.A.; Zhurkovich, I.K. Degradation of fullerene C60 by human myeloperoxidase and some reaction products. Fuller. Nanotub. Carbon Nanostructures 2020, 28, 196–201. [Google Scholar] [CrossRef]
  397. Kurapati, R.; Bianco, A. Peroxidase mimicking DNAzymes degrade graphene oxide. Nanoscale 2018, 10, 19316–19321. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  398. Kotchey, G.P.; Allen, B.L.; Vedala, H.; Yanamala, N.; Kapralov, A.A.; Tyurina, Y.Y.; Klein-Seetharaman, J.; Kagan, V.E.; Star, A. The enzymatic oxidation of graphene oxide. ACS Nano 2011, 5, 2098–2108. [Google Scholar] [CrossRef] [PubMed]
  399. Li, Y.; Feng, L.; Shi, X.; Wang, X.; Yang, Y.; Yang, K.; Liu, T.; Yang, G.; Liu, Z. Surface coating-dependent cytotoxicity and degradation of graphene derivatives: Towards the design of non-toxic, degradable nano-graphene. Small 2014, 10, 1544–1554. [Google Scholar] [CrossRef]
  400. Liu, L.; Zhu, C.; Fan, M.; Chen, C.; Huang, Y.; Hao, Q.; Yang, J.; Wang, H.; Sun, D. Oxidation and degradation of graphitic materials by naphthalene-degrading bacteria. Nanoscale 2015, 7, 13619–13628. [Google Scholar] [CrossRef]
  401. Salas, E.C.; Sun, Z.; Lüttge, A.; Tour, J.M. Reduction of graphene oxide via bacterial respiration. ACS Nano 2010, 4, 4852–4856. [Google Scholar] [CrossRef]
  402. Allen, B.L.; Kotchey, G.P.; Chen, Y.; Yanamala, N.V.K.; Klein-Seetharaman, J.; Kagan, V.E.; Star, A. Mechanistic investigations of horseradish peroxidase-catalyzed degradation of single-walled carbon nanotubes. J. Am. Chem. Soc. 2009, 131, 17194–17205. [Google Scholar] [CrossRef]
  403. Avanasi, R.; Jackson, W.A.; Sherwin, B.; Mudge, J.F.; Anderson, T.A. C60 fullerene soil sorption, biodegradation, and plant uptake. Environ. Sci. Technol. 2014, 48, 2792–2797. [Google Scholar] [CrossRef] [PubMed]
  404. Berry, T.D.; Filley, T.R.; Clavijo, A.P.; Gray, M.B.; Turco, R. Degradation and microbial uptake of C60 fullerols in contrasting agricultural soils. Environ. Sci. Technol. 2017, 51, 1387–1394. [Google Scholar] [CrossRef] [PubMed]
  405. Hwang, Y.S.; Li, Q. Characterizing photochemical transformation of aqueous nC60 under environmentally relevant Conditions. Environ. Sci. Technol. 2010, 44, 3008–3013. [Google Scholar] [CrossRef] [PubMed]
  406. Chae, S.R.; Hunt, D.E.; Ikuma, K.; Yang, S.; Cho, J.; Gunsch, C.K.; Liu, J.; Wiesner, M.R. Aging of fullerene C60 nanoparticle suspensions in the presence of microbes. Water Res. 2014, 65, 282–289. [Google Scholar] [CrossRef] [PubMed]
  407. Musil, M.; Konegger, H.; Hon, J.; Bednar, D.; Damborsky, J. Computational design of stable and soluble biocatalysts. ACS Catal. 2019, 9, 1033–1054. [Google Scholar] [CrossRef] [Green Version]
  408. Mazurenko, S.; Prokop, Z.; Damborsky, J. Machine learning in enzyme engineering. ACS Catal. 2020, 10, 1210–1223. [Google Scholar] [CrossRef]
  409. Chen, K.; Arnold, F.H. Engineering new catalytic activities in enzymes. Nat. Catal. 2020, 3, 203–213. [Google Scholar] [CrossRef]
  410. Nelson, A. Catalytic machinery of enzymes expanded. Nature 2019, 570, 172–173. [Google Scholar] [CrossRef] [Green Version]
  411. Drienovska, I.; Roelfes, G. Expanding the enzyme universe with genetically encoded unnatural amino acids. Nat. Catal. 2020, 3, 193–202. [Google Scholar] [CrossRef]
  412. Hwang, E.T.; Lee, S. Multienzymatic cascade reactions via enzyme complex by immobilization. ACS Catal. 2019, 9, 4402–4425. [Google Scholar] [CrossRef]
  413. Vazquez-Gonzalez, M.; Wang, C.; Willner, I. Biocatalytic cascades operating on macromolecular scaffolds and in confined environments. Nat. Catal. 2020, 3, 256–273. [Google Scholar] [CrossRef]
  414. Kornienko, N.; Ly, K.H.; Robinson, W.E.; Heidary, N.; Zhang, J.Z.; Reisner, E. Advancing techniques for investigating the enzyme-electrode interface. Acc. Chem. Res. 2019, 52, 1439–1448. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  415. Jeerapan, I.; Sempionatto, J.R.; Wang, J. On-body bioelectronics: Wearable biofuel cells for bioenergy harvesting and self-powered biosensing. Adv. Funct. Mater. 2020, 30, 1906243. [Google Scholar] [CrossRef]
Figure 1. Seven enzyme classes constituting the first component of the Enzyme Commission (EC) number.
Figure 1. Seven enzyme classes constituting the first component of the Enzyme Commission (EC) number.
Materials 15 01037 g001
Figure 2. Main types of carbon nanostructures (not to scale). Reproduced from [26] under a Creative Commons license. The CNO schematic structure is adapted with permission from [27], copyright ©1996, Elsevier.
Figure 2. Main types of carbon nanostructures (not to scale). Reproduced from [26] under a Creative Commons license. The CNO schematic structure is adapted with permission from [27], copyright ©1996, Elsevier.
Materials 15 01037 g002
Figure 3. A literature search on carbon nanomaterials and enzymes focused on the last decade (Source: Scopus 14 November 2021).
Figure 3. A literature search on carbon nanomaterials and enzymes focused on the last decade (Source: Scopus 14 November 2021).
Materials 15 01037 g003
Figure 4. CNMs typically used for peroxidase mimicry (left) and a possible reaction mechanism that ultimately generates hydroxyl radicals for the oxidation of colorless 3,3′,5,5′-tetramethylbenzidine (TMB) to a colored product (oxTMB). Reprinted with permission from [300], Copyright © 2022, American Chemical Society.
Figure 4. CNMs typically used for peroxidase mimicry (left) and a possible reaction mechanism that ultimately generates hydroxyl radicals for the oxidation of colorless 3,3′,5,5′-tetramethylbenzidine (TMB) to a colored product (oxTMB). Reprinted with permission from [300], Copyright © 2022, American Chemical Society.
Materials 15 01037 g004
Figure 5. (a) Optimized binding modes between hydrogen peroxide and CNTs either in their pristine form (p-CNTs) or with different oxygen-bearing functional groups. (b) Binding energies between hydrogen peroxide and the various CNT types shown in (a). (c) Schematic illustration of peroxidase mimicry by oxidized CNTs. Reprinted with permission from [320], Copyright © 2022 American Chemical Society.
Figure 5. (a) Optimized binding modes between hydrogen peroxide and CNTs either in their pristine form (p-CNTs) or with different oxygen-bearing functional groups. (b) Binding energies between hydrogen peroxide and the various CNT types shown in (a). (c) Schematic illustration of peroxidase mimicry by oxidized CNTs. Reprinted with permission from [320], Copyright © 2022 American Chemical Society.
Materials 15 01037 g005
Figure 6. Front-view (A) and vertical section (B) of the surface of the active pocket of acetylcholinesterase, with the peripheral anionic site (PAS) giving access, through a narrow gorge, to the catalytic active site (CAS). (C) Fullerene (brown sphere) can interact with the PAS through hydrophobic interactions with the enzyme surface, whose lipophilic potential (LP) is color-coded from brown (highest hydrophobicity) to blue (highest hydrophilicity). Reproduced with permission from [349], Copyright © 2022, American Chemical Society.
Figure 6. Front-view (A) and vertical section (B) of the surface of the active pocket of acetylcholinesterase, with the peripheral anionic site (PAS) giving access, through a narrow gorge, to the catalytic active site (CAS). (C) Fullerene (brown sphere) can interact with the PAS through hydrophobic interactions with the enzyme surface, whose lipophilic potential (LP) is color-coded from brown (highest hydrophobicity) to blue (highest hydrophilicity). Reproduced with permission from [349], Copyright © 2022, American Chemical Society.
Materials 15 01037 g006
Figure 7. Graphene is one of the most popular CNMs, employed in a variety of biosensing devices thanks to its exceptional electronic and mechanical properties. Reproduced with permission from [368], Copyright © 2022 American Chemical Society.
Figure 7. Graphene is one of the most popular CNMs, employed in a variety of biosensing devices thanks to its exceptional electronic and mechanical properties. Reproduced with permission from [368], Copyright © 2022 American Chemical Society.
Materials 15 01037 g007
Figure 8. Screen-printed electrode preparation using a conductive ink based on graphite and CNOs. Reproduced from [369], under a Creative Commons license.
Figure 8. Screen-printed electrode preparation using a conductive ink based on graphite and CNOs. Reproduced from [369], under a Creative Commons license.
Materials 15 01037 g008
Figure 9. Schematic representation of a biofuel cells with enzymes at the bioanode, where the fuel is oxidized, and the biocathode, where oxidants are reduced. Reproduced with permission from [374], Copyright © 2022, American Chemical Society.
Figure 9. Schematic representation of a biofuel cells with enzymes at the bioanode, where the fuel is oxidized, and the biocathode, where oxidants are reduced. Reproduced with permission from [374], Copyright © 2022, American Chemical Society.
Materials 15 01037 g009
Figure 10. Enzymatic biodegradation of GO as a green production method of graphene quantum dots. Reproduced with permission from [389], Copyright © 2022, American Chemical Society.
Figure 10. Enzymatic biodegradation of GO as a green production method of graphene quantum dots. Reproduced with permission from [389], Copyright © 2022, American Chemical Society.
Materials 15 01037 g010
Table 1. Scientific literature since 2017 on carbon nanomaterials and enzymes for various applications. C60 = buckminsterfullerene. CNDs = carbon nanodots. CNFs = carbon nanofibers. CNHs = carbon nanohorns. CNOs = carbon nano-onions. DHase = dehydrogenase. GO = graphene oxide. rGO = reduced GO. MWCNTs = multi-walled CNTs. SWCNTs = single-walled CNTs.
Table 1. Scientific literature since 2017 on carbon nanomaterials and enzymes for various applications. C60 = buckminsterfullerene. CNDs = carbon nanodots. CNFs = carbon nanofibers. CNHs = carbon nanohorns. CNOs = carbon nano-onions. DHase = dehydrogenase. GO = graphene oxide. rGO = reduced GO. MWCNTs = multi-walled CNTs. SWCNTs = single-walled CNTs.
CNM TypeConjugationEnzymeEC ClassApplicationRef.
C60Non-covalentLaccase1Water purification[217]
CovalentLaccase1Water purification[217]
Tyrosinase1Biosensing[218]
C60, MWCNTsNon-covalentEndonuclease3Biosensing[219]
CNDsCovalentGalactosidase3Biosensing[220]
Quinolinate phosphoribosyl transferase2Biosensing[221]
Non-covalentAdenylate kinase2Biocatalysis[222]
Endonucleases3Biosensing[223]
Glucose oxidase1Biosensing/therapy[224,225,226,227]
Glucose oxidase, peroxidase1Biosensing[228]
Laccase,
NAD+-dependent DHase,
alcohol DHase, aldehyde DHase,
formate DHase
1Biofuel cell[229]
Lactate oxidase1Biosensing[225]
Maltase3Therapy[230]
Old yellow enzyme1Biocatalysis[231]
Uricase1Biosensing[225]
CNFsCovalentLaccase1Biosensing[232]
Non-covalentTyrosinase3Biosensing[233]
CNHsCovalentGlutamate oxidase1Biosensing[234]
Non-covalentPeroxidase1Biosensing[235]
CNOsCovalentAlkaline phosphatase3Biosensing[236]
Glucose oxidase1Biosensing[236,237]
Peroxidase1Biosensing[236,238]
Non-covalentAdenylate kinase2Biocatalysis[222]
GOCovalentLaccase1Biosensing[239]
Non-covalentCholine oxidase/acetylcholine esterase1/3Biosensing[240]
Rhamnosidase3Biocatalysis[241]
Adenylate kinase2Biocatalysis[222]
Glucose Oxidase1 Biosensing/fuel cells[242]
Lipase3Biocatalysis[243]
GO, MWCNTsNon-covalentLactate oxidase1Biosensing[244]
rGONon-covalentAcetylcholine esterase3Biosensing[245]
Thrombin1Biosensing[245]
MWCNTsCovalentCholine Oxidase1Biosensing[246]
Glucose oxidase1Biosensing/fuel cells[247,248,249,250,251]
Glucose oxidase1Cancer therapy[252]
Laccase1Biofuel cells[253]
Laronidase3Therapy[254]
Lipase3Biofuel cells[255]
Peroxidase1Sensing/membranes[256,257]
Pyranose oxidase1Biosensing/fuel cells[258,259]
Tyrosinase3Biosensing[218,260]
Uricase1Biosensing[261]
Xanthine oxidase1Biosensing[262]
Non-covalentAlcohol dehydrogenase1Biosensing[263,264]
Alcohol oxidase1Biosensing[265]
Amino acid oxidase1Biosensing[265]
Amylase, lysozyme3Biocatalysis[266]
Bilirubin oxidase1Biosensing/fuel cells[267,268]
Choline oxidase1Biosensing[265]
Glucose dehydrogenase1Biosensing/fuel cells[269]
Glucose oxidase1Biosensing/fuel cells[247,270,271,272,273,274,275]
Glucose oxidase/catalase1Biofuel cells[276]
Glucose oxidase, laccase1Biosensing/fuel cells[277]
Glutamate oxidase1Biosensing[278]
Laccase1Biosensing[279]
Lactate oxidase1Biosensing[280]
Lipase3Biosensing/biocatalysis[263,264,281,282,283]
[NiFeSe]-hydrogenase1Biofuel cells[284]
Oxalate decarboxylase4Biofuel cells[285]
Peroxidase1Biosensing/fuel cells[286,287]
Pyranose oxidase1Biosensing/fuel cells[258,265]
SWCNTsCovalentTyrosinase1Biosensing[288]
Non-covalentCholine oxidase1Biosensing[265]
Table 2. Comparison of recent reports (since 2017) that described CNMs-based sensors for glucose detection.
Table 2. Comparison of recent reports (since 2017) that described CNMs-based sensors for glucose detection.
CNM TypeConjugationLinear RangeDetection LimitSensitivityRef.
CNDsNon-covalent250–3000 μMn.a.n.a.[224]
0.1–500 μM65 µM21.6 µA·mM−1·cm−2[227]
Covalent0.1–500 μM0.04 μMn.a.[228]
CNOsCovalent1000–10,000 μM210 μM26.5 µA·mM1·cm−2[237]
CNTsCovalent100–6000 μM9.01 μMn.a.[249]
n.a.n.a.0.27–3.7 µA·mM1·cm2 [258]
Non-covalent1000–20,000 μMn.a.0.198 µA·mM1·cm2[269]
1–5000 μM0.36 μMn.a.[271]
0–5000 μM50 μM289 μA·mM−1·cm−2[273]
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Rozhin, P.; Abdel Monem Gamal, J.; Giordani, S.; Marchesan, S. Carbon Nanomaterials (CNMs) and Enzymes: From Nanozymes to CNM-Enzyme Conjugates and Biodegradation. Materials 2022, 15, 1037. https://doi.org/10.3390/ma15031037

AMA Style

Rozhin P, Abdel Monem Gamal J, Giordani S, Marchesan S. Carbon Nanomaterials (CNMs) and Enzymes: From Nanozymes to CNM-Enzyme Conjugates and Biodegradation. Materials. 2022; 15(3):1037. https://doi.org/10.3390/ma15031037

Chicago/Turabian Style

Rozhin, Petr, Jada Abdel Monem Gamal, Silvia Giordani, and Silvia Marchesan. 2022. "Carbon Nanomaterials (CNMs) and Enzymes: From Nanozymes to CNM-Enzyme Conjugates and Biodegradation" Materials 15, no. 3: 1037. https://doi.org/10.3390/ma15031037

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

Article Metrics

Back to TopTop