Next Article in Journal
A Molecularly Imprinted Polymer-Disposable Pipette Tip Extraction-Capillary Electrophoresis (MISPE-DPX-CE) Method for the Preconcentration and Determination of Scopolamine in Synthetic Urine Samples
Previous Article in Journal
Determination of Pb(II) Ions in Water by Fluorescence Spectroscopy Based on Silver Nanoclusters
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Rational Design of Nanozymes Enables Advanced Biochemical Sensing

1
School of Public Health, Hengyang Medical School, University of South China, Hengyang 421001, China
2
Institute of Green Chemistry and Chemical Technology, School of Chemistry and Chemical Engineering, Jiangsu University, Zhenjiang 212013, China
*
Author to whom correspondence should be addressed.
Chemosensors 2022, 10(10), 386; https://doi.org/10.3390/chemosensors10100386
Submission received: 8 September 2022 / Revised: 20 September 2022 / Accepted: 21 September 2022 / Published: 23 September 2022
(This article belongs to the Section Nanostructures for Chemical Sensing)

Abstract

:
In comparison with bioenzymes, nanozymes exhibit excellent robustness against extreme conditions, a low production cost, and easy-to-adjust properties, as well as potential versatility. These superiorities have attracted abundant interest in the last 15 years, to develop various nanozymes for applications including analytical sensing, environmental engineering, and biomedicine. In particular, for analytical sensing, a lot of nanozyme-involved principles and methods have been explored and applied to clinical diagnosis, environmental monitoring, food safety detection, and forensic analysis. Moreover, rational exploitation and use of nanozyme materials promote the performance of analytical methods. To highlight the latest progress in this attractive field, recent design concepts of nanozymes for advanced biochemical sensing are summarized. The development of single-atom nanozymes, self-cascade nanozymes, structurally biomimetic nanozymes, molecularly imprinted nanozymes, nanozymes breaking the pH limit, and multifunctional nanozymes is discussed in detail, to enhance detection sensitivity and selectivity, as well as expand application scenarios. Finally, some challenges and trends related to nanozyme-based sensors are reported, to satisfy the increasing needs of biochemical analysis with nanozymes.

1. Introduction

Nanozymes are defined as a class of nanoscale materials that are able to exhibit enzyme-like catalytic behaviors, kinetics, and mechanisms [1,2,3]. As an emerging member of artificial enzymes, nanozymes have experienced a period of explosive development in the past 15 years [4]. In contrast to natural bioenzymes, they show several attractive characteristics and benefits, including an excellent stability, even under harsh conditions, easy scale production, low cost, and easy-to-regulate catalytic performance. Compared to the previously developed organic artificial enzymes (cyclodextrins, porphyins, supermolecules, et al.), the unique feature of an inorganic nanostructures can endow nanozymes with potential versatility [5]. These advantages are drawing more and more interest, to explore nanozymes for various applications, such as analytical sensing, environmental engineering, and biomedicine. In the analytical sensing field, the signal amplification offered by catalytic reactions makes nanozymes promising enzyme alternatives for developing a number of analytical principles and methods [6]. Commonly, nanozymes are used as catalysts to trigger a series of chromogenic, fluorescence, or chemiluminescence reactions, thus providing optical signals for the qualitative and quantitative determination of targets. They can replace horseradish peroxidase (HRP) or alkaline phosphatase (ALP) as catalytic labels to fabricate immunoassays [7]. In addition, the conjugation of nanozymes with other sensing elements (DNA chains, aptamers, bioenzymes, et al.) enables the development of multifarious sensors [8]. In addition, interactions between certain species and nanozyme surfaces make it possible to realize the detection of the former [9]. These principles and methods have given nanozymes intensive use in clinical diagnosis, biochemical measurement, environmental monitoring, food safety analysis, and forensic detection [10,11,12].
Undoubtedly, the introduction of nanozymes opens a new path to obtaining high-performance sensing. A variety of chemosensors and biosensors have been fabricated based on nanozyme catalysis in the last decade [13]. Even so, in this field some issues remain that seriously hinder the further development of nanozyme-based biochemical sensing and need to be addressed. For instance, a large number of the nanozymes explored currently lack catalytic specificity, and achieving selective detection with nanozymes becomes a crucial problem [14]. The catalytic activity of nanozymes is still far lower than that of corresponding natural enzymes, and how to utilize nanozymes for a highly sensitive analysis, particularly for ultratrace targets, is challenging. Furthermore, expanding the applicability of nanozyme-based analytical methods in complex scenarios is always desirable.
To solve the above issues, the rational design of nanozymes, based on deeply understanding their structures and mechanisms, to gain the desired catalytic features is necessary [15]. To date, several emerging material strategies have been demonstrated as successful for significantly accelerating the development of chemosensors and biosensors. To highlight the progress in this interesting field, here we aimed at summarizing recent nanozyme concepts for promoting biochemical sensing (Figure 1). The rational design of single-atom nanozymes with extraordinary activity, as well as self-cascade nanozymes with multiple signal amplifications, was reviewed, to enhance detection sensitivity. For selectivity improvement, structurally biomimetic nanozymes and molecularly imprinted nanozymes are introduced in detail. To expand the application scenarios, the development of nanozymes breaking the pH limit, as well as multifunctional nanozymes with luminescent properties, is discussed. At the end of this review, to attract more efforts toward meeting increasing analytical requirements, some challenges and trends in the designing and development of nanozymes for advanced biochemical sensing are presented.

2. Rational Design of Nanozymes for Enhancing Detection Sensitivity

Sensitivity is considered one of the most important analytical parameters for a sensor. In nanozyme-based chemosensors and biosensors, the detection sensitivity closely relies on the catalytic activity of the materials used. Although the signal amplification provided by a nanozyme catalysis is able to offer acceptable sensitivity for conventional sensing, detecting ultratrace analytes requires exploring enzyme mimics with the desired activity. However, the catalytic activity of most nanozymes developed currently is far lower than that of the corresponding bioenzymes. Designing and developing high-activity nanozymes has become a vital topic in the community. To date, several strategies, including structural engineering [16], size regulation [17,18,19], surface modification [9,20], doping effect [21,22,23], and material compositing [24] have been explored to improve the catalytic activity of nanozymes. Given that these traditional strategies have been summarized in some other publications [25,26], we will not repeat these nanozyme activity enhancement methods here. Instead, two emerging material strategies, namely single-atom nanozymes and self-cascade nanozymes, are discussed for improving the detection sensitivity of nanozyme-involved sensors.

2.1. Single-Atom Nanozymes

Since the concept of single atomic catalysts (SACs) was first proposed by Qiao et al. to promote CO oxidation with a high activity [27], it has attracted intensive attention and interest from the academic community [28,29]. SACs consist of atomic active sites highly dispersed on inert carriers and are generally regarded as the limit for the design of nanoscale materials at the atomic level. When SACs are employed as enzyme mimics, namely single-atom nanozymes (SANs), some interesting features and benefits appear [30]. First, the atomic active sites of many SANs are very similar to those of metal-containing bioenzymes [31], and such a structural similarity can inspire the design of nanozymes with high activity and substrate specificity [32]. Second, SANs have well-defined geometric and electronic structures, which help the study of the catalytic processes and mechanisms of nanozymes [33,34]. Third, it is relatively easy to regulate the active site microenvironment of SANs, thus facilitating the study of the structure–activity relationship [35]. Fourth, SANs have a maximum atomic utilization efficiency and can be cycled for reuse, satisfying the green and economical principles of materials. Last but not least, although SANs belong to the heterogeneous catalysts, the high dispersion of active sites enables them to possess some characteristics of homogeneous catalysts [30]. As a result, a number of SACs have been explored as artificial enzymes for various applications in recent years [34,36,37,38], demonstrating that single atomization is an efficient way to gain high-activity nanozymes [39].
With the above features, SANs not only provide ideal models for studying sensing mechanisms, but also endow nanozyme-based sensors with the desired performance [40]. In this regard, several SANs with high activity have been explored and applied to analytical sensing [41,42,43,44,45], and some of them can even provide higher sensitivity and lower limit of detection (LOD) compared to bioenzymes [43]. Cheng et al. developed a CNT/FeNC SAN with high peroxidase-like activity for ultrasensitive biosensing [41]. They first dispersed oxidized carbon nanotubes (CNTs) in a pyrrole solution, where pyrrole could be adsorbed onto CNTs because of the ππ interaction between the carbon planes of CNTs and the pyrrole molecules (Figure 2A). Then, they triggered the polymerization of pyrrole adsorbed on CNTs, to form polypyrrole (PPy), using (NH4)2S2O8 as an oxidant. After that, the obtained CNT/PPy was impregnated in a mixture of Fe(NO3)3 and NaCl to adsorb metal cations, followed by pyrolyzing the collected solid hybrid in N2 and NH3 atmospheres successively, thus obtaining CNT/FeNC SAN. The gained SAN theoretically had 100% atomic Fe–Nx–C moieties, exhibiting excellent peroxidase-mimicking catalytic activity to trigger the oxidation of several chromogenic substrates in the presence of H2O2. With the superior SAN, they developed paper-based bioassays for H2O2 and ascorbic acid (AA). Detection of glucose was also achieved when combining the SAN with glucose oxidase (GOx).
Apart from the above synthetic strategy, calcination of metal–organic frameworks (MOFs) is another widely used method to gain SANs with an outstanding catalytic activity [44,45,47]. Our group synthesized a Fe-N-C SAN with single-atom Fe–Nx entities stabilized by MOFs-derived carbon, which could present peroxidase-like activity almost at the same level as natural HRP [44]. The Fe-N-C SAN was prepared by successively pyrolyzing a FeZn MOF in different atmospheres, followed by washing off large metal and metal oxide particles with acid. It was found that the obtained Fe-N-C SAN exhibited both peroxidase- and oxidase-like catalytic activities in acidic media. By employing 3,3′,5,5′-tetramethylbenzidine (TMB) as a substrate, the catalytic activity of the Fe-N-C SAN was carefully evaluated according to the standard protocol [48]. Consequently, the SAN provided a specific peroxidase-like activity up to 57.76 U/mg, which was comparable to natural HRP (Figure 2B). With this exceptional activity, we further fabricated a smartphone-assisted paper-based bioassay for the monitoring of butyrylcholinesterase (BChE), a typical biomarker of organophosphorus and carbamate pesticide residues. As a result, the bioassay provided linear responses for BChE activity in the range of 2~40 U/L, with a detection limit lower than most previous methods.
In many SACs and SANs, the content of single metal atoms acting as catalytic active sites is very low (less than 1.5 wt%), seriously hindering their mass activity and advanced applications [49]. To further improve the activity of SANs and their sensing sensitivity, exploring effective methods to fabricate SANs with high metal loading is required. In this regard, Zhu’s group designed a Cu-based SAN with abundant Cu sites (up to 5.1 wt%) loaded on carbon nanosheets using a salt-template strategy (Figure 2C), which exhibited excellent enzyme-like activity and enhanced biosensing performance [46]. It should be noted that the obtained Cu-N-C SAN only exhibited peroxidase-like catalytic activity, but no oxidase-like activity. This feature could effectively avoid the background brought by the oxidase-mimetic activity during biosensing. By integrating the Cu-N-C SAN with natural acetylcholinesterase (AChE) and choline oxidase (ChO), they fabricated a triple-enzyme cascade reaction system for the colorimetric determination of acetylcholine (ACh) and organophosphorus (OP) with enhanced sensitivity.

2.2. Self-Cascade Nanozymes

Cascade biocatalysis commonly occurs in living organisms, to control the metabolism and signal transduction. In a cascade biocatalytic system, multiple bioenzymes are often involved in triggering two or more reactions, where each subsequent reaction begins after the previous one is completed. Given multi-enzyme-driven cascade reactions can provide superior activity and selectivity, a lot of systems have been developed as biomimetic reactors for biosensing and biomedical applications [50,51]. With increasing materials being found to exhibit multiple enzyme-like activities [52,53,54,55], instead of using several enzyme mimics to fabricate a cascade catalytic system, it is preferable to design nanozymes with self-cascade catalytic properties, where cascade reactions occur on the surface of a single nanozyme. As a result, the proximity effect of different active sites, as well as the shortened intermediate diffusion, can provide a higher catalytic efficiency [56,57]. With this attractive feature, self-cascade nanozymes have found promising use in biomedicine [58,59,60,61,62].
For analytical sensing, self-cascade catalytic reactions are able to offer multiple amplifications of signals (Figure 3A). Correspondingly, a higher detection sensitivity is supposed to be obtained in comparison with sensors based on single catalysis. As a common example, several Cu-based materials have been explored as self-cascade nanozymes for biosensing [63,64,65,66,67], where their oxidase- and peroxidase-like activities are mainly employed. We designed a Cu-based MOF material (CuBDC) that acted as a self-cascade nanozyme with cysteine oxidase- and peroxidase-mimetic catalytic activities [63]. In this material, the Cu–O clusters not only showed the capacity for catalyzing the oxidation of cysteine with the help of dissolved O2 to produce cystine and H2O2, but also provided the peroxidase-mimetic ability for catalyzing the generated H2O2 to produce radicals under the same conditions. Given that the terephthalic acid (TA) ligand of the MOF could be oxidized to fluorescent 2-hydroxyterephthalic acid (TAOH) by the produced radicals, the CuBDC presented a remarkable fluorescence when cysteine was fed (Figure 3B). Such a system catalyzed by the self-cascade nanozyme offered an efficient method to sensitively detect the cysteine substrate.
Similar enzyme-like activities were also found in CuO NPs by Liu’s group [64]. The single-component CuO NPs not only acted as a glutathione (GSH) oxidase mimic, to oxidize GSH in the solution containing dissolved O2, where H2O2 was produced, but also played a peroxidase-like role in inducing the oxidation of the TA substrate by the produced H2O2, with a high local concentration near CuO NPs. The dual enzyme-like activities of the Cu NPs produced a self-organized cascade reaction (Figure 3C), enabling the determination of GSH with good performance. By using the strong coordination interaction of GSH and metal ions, highly sensitive detection of Ag+ was also achieved with the self-cascade nanozyme, providing a LOD down to 0.037 nM.
According to the above examples, the construction of self-cascade catalytic systems depends on the multi-activity features of a single nanozyme. Although, in recent years, increasing materials with more than one enzyme-mimetic activity have been explored to promote analytical applications, their use might also lead to some negative impacts [52]. For instance, both peroxidase- and oxidase-like activities are able to trigger the oxidation of substrates, even in the same environment. As a result, the oxidase-like catalytic reaction would affect the measurement of a target that originally relied only on the peroxidase-mimicking activity [46]. Therefore, more attention is required to rationally design nanozymes with the desired self-cascade catalytic activities, without introducing other irrelevant activities and interference.

3. Rational Design of Nanozymes to Improve Sensing Selectivity

Selectivity is another important parameter for a sensor. It is widely recognized that a lack of substrate specificity becomes the biggest challenge for nanozymes [26,68]. For this reason, foreign recognition elements are often utilized in combination with nanozymes to gain high-selectivity detection [14]. At present, a few means have been investigated to realize the above goal, such as constructing natural enzyme-nanozyme cascade sensors [51,69], combining nanozymes with DNA chains or aptamers [70,71,72], fabricating nanozyme-linked immunosensors [7], and utilizing the specific interactions between certain nanozymes and analytes [73,74]. To fundamentally solve the selectivity problem, rationally designing nanozymes that have intrinsic catalytic specificity is desirable. Here, structurally biomimetic nanozymes and molecularly imprinted nanozymes are discussed in detail, and their applications to achieve selective sensing are revealed.

3.1. Structurally Biomimetic Nanozymes

It is known that the majority of natural enzymes are able to offer excellent substrate specificity. Such specificity mainly relies on their fine structures. In protein-based enzymes, complex multi-level structures provide clearly defined channels and cavities, with shape, size, and interactions well matching their corresponding substrates, thus endowing them with accurate substrate recognition and catalysis. Differently from bioenzymes, inorganic nanozyme materials inherently lack well-defined structures near active sites, leading to the very poor catalytic specificity of nanozymes. Naturally, the structural bionic concept inspires the design of nanozymes with surface microstructures similar to the corresponding bioenzymes [68], to obtain inorganic enzyme mimics with the desired catalytic selectivity.
In this regard, several studies on designing and developing nanozymes that mimic the fine structures of natural enzymes have been reported [75,76]. For instance, inspired by the active site structure of natural catechol oxidase, Li et al. fabricated a MOF material called MOF-818, which consisted of tri-nuclear copper centers (Figure 4A) [76]. In natural catechol oxidase (PDB code 1BT1), a six-histidine-coordinated binuclear Cu metal center plays the catalytic role in triggering the oxidation of o-diphenol. Correspondingly, the O2 molecule was reduced to H2O2 and further protonated via cleaving the HO–OH bond to generate H2O. Similarly, with tri-nuclear copper centers, MOF-818 could induce the catalytic oxidation of the substrate in the participation of dissolved O2, to finally generate H2O2 instead of H2O. This was because the designed MOF-818 did not have a peroxidase-like activity. Unlike common CeO2 and Pt NPs as nanozymes to induce the catalytic oxidation of TMB or 2,2′-azino-bis(3-ethylbenzothiazoline-6-sulfonic acid) (ABTS) with the presence or absence of H2O2, the proposed MOF-818 only exhibited a catechol oxidase-mimetic activity to trigger the oxidation of 3,5-di-tert-butylcatechol (3,5-DTBC). As a consequence, the MOF-818 acting as a specific artificial catechol oxidase exhibited excellent catalytic selectivity.
As mentioned above, the fine surface structures near active centers play an important role in determining the selectivity of nanozymes toward substrates. This inspires the adjustment of the specificity of nanozymes by tailoring their surface. Typically, chiral catalysis can be obtained when some chiral ligands are anchored on an inorganic nanozyme surface [77,79,80,81]. In general, amino acid ligands and small chiral molecules, as well as DNA chains, are employed to modify the surface of inorganic materials to obtain chiral nanozymes. For example, Qu’s group modified a series of D- or L-amino acids onto a ceria surface, to realize the enantioselective catalysis of 3,4-dihydroxyphenylalanine (DOPA) enantiomers [77], a commonly used drug for treating the Parkinson’s disease. By measuring the reaction kinetics of ceria nanoparticles (CeNPs) decorated by eight different amino acids (phenylalanine, alanine, tryptophan, histidine, glutamic acid, arginine, lysine, and tyrosine), it was concluded that the phenylalanine (Phe)-coated CeNPs were optimal for DOPA oxidation and provided excellent stereoselectivity toward its enantiomers. In detail, L-Phe-decorated CeNPs offered a higher catalytic ability to oxidize D-DOPA, while D-Phe-decorated CeNPs were more effective toward L-DOPA (Figure 4B). Theoretical calculations indicated that both DOPA enantiomers could interact with the chiral Phe modifiers, mainly by forming different hydrogen bonds. Three hydrogen bonds could be formed between D-Phe and L-DOPA, compared to only one hydrogen bond formed between D-Phe and D-DOPA. Moreover, the ππ aromatic packing interaction was able to further intensify the interaction of D-Phe and L-DOPA. On the contrary, in comparison with the interplay of L-Phe and L-DOPA, a stronger interaction could be obtained between L-Phe and D-DOPA by forming two hydrogen bonds. This indicated that the modification of chiral amino acids on nanozyme surface could have a great impact on the stereoselectivity of the latter. These fascinating results are inspiring increasing efforts in designing chiral nanozymes [80].
Given SANs have clearly defined geometric and electronic structures and the microenvironment around active centers plays a significant role in determining their catalytic performance, one can tailor the microenvironment of SANs to obtain enhanced substrate selectivity [32]. As such, Wang et al. reported the development of several Mo-based SANs (MoSA–Nx–C, x = 2, 3 or 4) [78], and it was observed that their peroxidase-mimetic specificity could be adjusted by changing the coordination number of single-atom Mo sites. Compared to MoSA–N2–C and MoSA–N4–C, the optimized MoSA–N3–C exhibited a significantly oppressed oxidase-like activity, but a higher peroxidase-mimicking activity was observed (Figure 4C). This study demonstrated the feasibility of rationally tailoring the microenvironment of SACs to obtain high-specificity nanozymes. Currently, although there have been very few cases of designing structurally biomimetic nanozymes for selective analysis, it is believed that in the near future the structural bionic concept will draw more and more attention to this topic.

3.2. Molecularly Imprinted Nanozymes

Molecularly imprinted polymers (MIPs) have artificial binding sites toward target molecules. They are generally synthesized via the co-polymerization of template molecules and functional monomers, with or without the addition of some cross-linking agents. After the template molecules are removed, some recognition cavities remain in the formed polymer matrix. The remaining cavities are complementary to the template molecules in shape, size, and chemical functional groups, and can selectively bind the template molecules again [82]. Combining MIPs with nanozymes provides a potential path to solve the poor substrate selectivity problem of the latter. For example, Zhang fabricated molecularly imprinted sites on peroxidase-like Fe3O4 surface by utilizing the commonly used enzymatic substrates TMB and ABTS as template molecules [83]. In comparison with bare Fe3O4 nanoparticles, the explored molecularly imprinted nanozymes presented improved catalytic kinetics toward the corresponding substrates. Importantly, the introduction of molecularly imprinted sites endowed the nanozyme with significantly enhanced specific affinities toward the substrates (Figure 5A). Similar phenomena have also been observed in other molecularly imprinted nanozymes [84,85,86,87]. These studies provide the basis of introducing artificial imprinting sites onto a nanozyme surface to solve the poor selectivity issue.
Instead of imprinting enzymatic substrates onto the nanozyme surface, one can use analytes as the template molecules and fabricate molecularly imprinted nanozymes for analytical applications [88,89,90,91,92]. As such, our group prepared molecularly imprinted sites on a peroxidase-mimetic Fe3O4 nanoparticle surface through self-polymerizing dopamine (DA) in a weak alkaline environment [88], and further fabricated a colorimetric sensor for the high-selectivity determination of tetracycline (TC) (Figure 5B). The imprinted nanozyme (Fe3O4@MIP) had abundant cavities and channels for substrates access to the peroxidase-like Fe3O4 core. When TC was fed, it was selectively recognized and captured by the MIP shell and masked the cavities and channels. Consequently, the catalyzed TMB color reaction was suppressed. On the basis of such a principle, colorimetric sensing of TC with good selectivity against some structural analogues (oxytetracycline, chlorotetracycline, erythromycin, penicillin, et al.) was achieved. It should be stated that, although introducing molecular imprinting sites onto a nanozyme surface can improve the substrate specificity for selective detection, the imprinted layer introduced would significantly cover up the active surface and sites of nanozymes, thus leading to a decrease in catalytic activity. Thus, balancing the catalytic specificity and activity turns to be a vital issue. How to rationally design molecularly imprinted nanozymes with good specificity and enough activity for high-performance sensing is the current challenge.

4. Rational Design of Nanozymes to Expand Application Scenarios

In addition to sensitivity and selectivity, expanding the application scenarios of nanozyme-based sensors is another research direction. This mainly includes how to design and develop nanozymes applicable to different environments, as well as how to realize high-performance detection using different sensing modes. Currently, although great progress has been made to serve the analytical sensing field, some intrinsic shortcomings still hinder the wide use of nanozymes. Here, we discuss the rational design of nanozymes that can break the pH limit and multifunctional nanozymes, both of which can greatly expand the applicability of enzyme mimics in analytical sensing.

4.1. Nanozymes Breaking the pH Limit

To date, many materials have been explored as peroxidase-mimetic nanozymes, while the majority of them present their activity only in acidic conditions (pH 3.0~5.0). This is because in acidic media, the H2O2 substrate prefers to undergo a homogeneous dissociation process to form hydroxyl radicals and trigger further reactions, thus exhibiting a peroxidase-like activity [93]. If the solution pH is expanded to the neutral medium, a very weak, or even no, activity is gained. This characteristic greatly limits the wider use of peroxidase mimics in biochemical analysis. Taking glucose detection as an example, two operation steps are often employed, where a peroxidase-like nanozyme is combined with natural GOx. In detail, GOx is first used to trigger the oxidation of glucose in a neutral medium to produce H2O2, and the produced H2O2 is then catalyzed by the peroxidase-like nanozyme in a moderately acidic medium, to trigger some chromogenic reactions. Obviously, such a two-step protocol results in some drawbacks: on the one hand, the H2O2 species produced in the first step will spontaneously decompose, and its loss limits the detection sensitivity; on the other hand, different pH conditions are required for the bioenzyme–nanozyme cascade catalytic system, complicating the detection operation.
To solve the above dilemma, one can try to develop peroxidase-mimicking nanozymes that are able to exhibit the desired activity in physiological environments. To sum up, there mainly are two strategies used to expand the working pH of nanozymes (Figure 6A): one is to explore nanozymes with intrinsically high activity in acidic environments, which can also exhibit a certain activity in neutral media [94,95,96,97,98,99]; the other is to design the surface of nanozymes, where special surface engineering (introducing surface charge, fabricating acidic microenvironment, et al.) enables the extension of the working pH to neutral conditions [100,101,102,103,104].
Thus, we designed a Co-based polyoxometalate nanozyme (CoPW11O39) that could show good peroxidase-mimicking activity at neutral pH [96]. Zeta potential measurements indicated that the surface of the obtained CoPW11O39 was negatively charged. As a result, it exhibited a strong ability to catalyze the oxidation of positively charged TMB, even under neutral conditions (Figure 6B). In contrast, no catalytic activity under the physiological conditions was found when negatively charged ABTS was used as the substrate. Given that the nanozyme and natural GOx had a similar working pH environment, they were used to develop a one-pot colorimetric method for glucose detection.
Another efficient strategy to modulate the working pH of peroxidase mimics is to modify charged ligands onto their surface. For example, Han et al. developed Au@Ag nanorods (NRs) stabilized by positively charged poly(diallyldimethylammonium) (PDDA), which exhibited a high activity level over a wide pH range (pH 4.0~6.5) when using ABTS as the chromogenic substrate (Figure 6C) [101]. At pH 6.5, the peroxidase-like activity of the Au@Ag NRs was stable, showing a strong affinity toward negatively charged ABTS. Given this character, the detection of H2O2 at pH 6.5 was realized based on the PDDA-modified Au@Ag NRs catalyzing the colorimetric reaction of H2O2 and ABTS. Similarly, we engineered the charge of peroxidase-mimetic CuS via amino acid surface modification, to achieve good catalytic activity in a neutral medium [100]. In detail, aspartic acid (Asp), an acidic amino acid with an isoelectric point of 2.97, was used to decorate the surface of CuS during preparation. As the Asp modifier could be hydrolyzed to negatively charged species in the neutral condition, rich negative charges were observed on the nanozyme surface, which could prominently promote the diffusion and adsorption of the positively charged substrate TMB onto the active surface for reaction. As a consequence, differently from bare CuS that had little activity at neutral pH, the Asp modified CuS exhibited excellent peroxidase-mimetic activity in the same environment (Figure 6D). As an application of this, one-pot colorimetric determination of glucose was obtained by integrating GOx with the engineered peroxidase-mimetic nanozyme.

4.2. Multifunctional Nanozymes

Intrinsically, nanozymes are a class of nanoscale materials. When the size of a material is conditioned to the nanometer level, a series of interesting properties emerge. In fact, apart from enzyme-like catalytic features, many materials also possess some attractive optical [105,106,107], magnetic [108,109,110], and electrical [111] properties or act as promising carriers [112,113,114] for biochemical analysis. These multiple properties of nanozymes offer biochemical sensing unparalleled benefits in practical applications. For instance, nanozymes with responsive fluorescence can be employed to self-output signals without introducing external substrates or labels. By combining the catalytic feature of nanozymes with their additional properties, multifunctional nanozymes have huge potential in the biochemical sensing area [115]. In this work, materials with both enzyme-like catalytic activity and luminescent properties are discussed.
Currently, several classes of materials with luminescent properties have been explored, including MOFs [116,117,118,119,120,121,122,123,124,125,126,127], coordination compounds [128,129], metal clusters [130], carbon dots [131,132], and C3N4-based materials [133]. These multifunctional nanozymes enable “turn-on” fluorescence, “turn-off” fluorescence, bimodal fluorescence/colorimetric, and ratiometric fluorescence measurements of analytes (Figure 7A). As a typical method, multifunctional nanozymes with fluorescent properties can be obtained by integrating TA with active metal nodes, to fabricate MOFs-based nanozymes, where TA acts as an organic ligand to bind Fe3+ or Cu2+ and exhibits a remarkable fluorescence when it is catalytically oxidized to TAOH. Thus, Ye’s group prepared a dual-functional MIL-53(Fe) material with peroxidase-like activity via the self-assembly of Fe3+ and TA, to enable the label-free “turn-on” fluorescence detection of H2O2 and glucose [117]. Once H2O2 was fed, the MIL-53(Fe) could provide a notable peroxidase-mimetic catalytic ability for decomposing H2O2 to produce hydroxyl radicals, and the latter oxidized the ligand TA in the MIL-53(Fe) to TAOH, which presented a remarkable fluorescence at approximately 440 nm (Figure 7B). As a result, a self-reporting fluorescence sensor based on the MIL-53(Fe) was fabricated for H2O2 sensing, free of the addition of any other substrate or label. By integrating the bi-functional nanozyme with GOx, convenient detection of glucose was also gained.
Recently, we designed a FeZr bimetal–organic framework material (UiO-66(Fe/Zr)-NH2) with multiple roles (luminescent property, peroxidase-like activity, and target recognition) to construct a ratiometric fluorescence sensor for the high-performance sensing of phosphate ions (Pi) [119]. The employment of a fluorescent ligand (2-aminoterephthalic acid) made the MOF show a strong fluorescence at 435 nm (Figure 7C). The Fe3+/Fe2+ nodes offered a good enzyme-mimicking ability for catalyzing the o-phenylenediamine (OPD) substrate to a fluorescent product (OPDox) at 555 nm, which could quench the fluorescence (435 nm) of UiO-66(Fe/Zr)-NH2 because of the inner filter effect. The Zr4+ nodes in the material acted as sites for Pi recognition and capture. When Pi existed, it was specifically adsorbed onto the UiO-66(Fe/Zr)-NH2 and decreased the latter’s peroxidase-mimicking activity. Consequently, the fluorescence of UiO-66(Fe/Zr)-NH2 at 435 nm was restored, but the fluorescence at 555 nm was reduced. Based on the ratiometric sensing principle, efficient detection of Pi with favorable selectivity and sensitivity was gained. In another work, we developed a bimodal fluorescence/colorimetric approach for sensing pesticides, by coupling stimulus-triggered luminescence with oxidase-mimetic activity in Ce-based coordination polymer nanoparticles (CPNs(Ⅳ)) [128]. The CPNs(Ⅳ) offered an oxidase-mimicking ability to catalyze the oxidation of colorless TMB to blue TMBox, offering a colorimetric signal at around 652 nm (Figure 7D); when ascorbic acid 2-phosphate (AAP) was utilized as an enzymatic substrate of acid phosphatase (ACP) to be hydrolyzed, the AA produced with a certain reducibility could induce the chemical reduction of Ce4+ in the CPNs(Ⅳ) to Ce3+, leading to the production of CPNs(Ⅲ). In comparison with CPNs(Ⅳ), the produced CPNs(Ⅲ) exhibited a greatly reduced oxidase-mimicking activity, leading to a remarkable suppression of the TMB color reaction. In addition, the produced CPNs(Ⅲ) offered a fluorescence peak at 356 nm; with the addition of OP or carbamate pesticides, the activity of ACP was inhibited. As a consequence, the hydrolytic procedure of AAP to AA was suppressed, and the conversion of CPNs(Ⅳ) to CPNs(Ⅲ) was also blocked, thus resulting in the recovery of the TMB color reaction; however, the fluorescence signal of CPNs(Ⅲ) was inhibited. On the basis of this sensing principle, bimodal fluorescence/colorimetric sensing of malathion, a typical OP, was demonstrated. In addition, the satisfactory interference-tolerance reliability and practicability of the method were confirmed.

5. Conclusions and Perspectives

Compared to natural bioenzymes, the attractive benefits of nanozymes have motivated the exploration of the abundant inorganic nanomaterials with enzyme-like properties, especially for sensing applications. On the other side of the coin, several obvious shortcomings have also been exposed in artificial enzyme mimics, including a low catalytic activity and poor substrate selectivity. To overcome these shortcomings and to better serve analytical sensing, the rational design of nanozymes is required. Great progress has been made in this field in the past few years. SANs with a comparable activity have been developed to provide higher sensitivity for nanozyme-based sensors. In addition, self-cascade nanozymes have been designed to fabricate tandem catalytic systems with multistage amplifications for ultrasensitive detection. For selective detection, some efficient strategies, including molecular imprinting and structurally bionic design, have been validated, to make up for the deficiency of catalytic specificity in nanozymes. Moreover, multifunctional nanozymes with luminescent properties, as well as nanozymes that can break the pH limit, have also been applied in advanced analytical sensing. The design of these emerging nanozymes greatly promotes the creation of analytical principles, methods, technologies, and devices. Furthermore, an effectively enhanced analytical performance has been gained by using these nanozymes.
The trend of rationally designing nanozymes for biochemical detection is evident [134]. It is believed that in the near future more interesting nanozymes will be designed and developed, thus bringing extended methods and improved performance for biochemical sensing: (1) more material design strategies are expected to be explored to enhance the analytical performance of nanozyme-based sensors. The design concepts discussed above indeed solve some problems of the nanozymes used in analytical chemistry, while these are not sufficient to meet the increasing demands of modern measurement. As mentioned above, the nanozymes developed currently still have various challenges, including a low catalytic activity, poor substrate specificity, limited catalytic environment, and complex bioconjugation [135]. Thus, exploring new designs of nanozymes is necessary, to compensate for these shortcomings for advanced biochemical sensing. For instance, combining the bioenzyme-inspired design of SANs with the precise regulation of their active sites, as well as increasing the site density may be a promising way to improve the catalytic activity and specificity of nanozymes [46,78]. In order to further expand the applications of nanozymes, designing more nanozymes with multiple functions (e.g., nanozymes with fluorescent, chemiluminescent, Raman, magnetic, or/and electrical properties) is promising; (2) in addition to sensitivity, selectivity, and application scenarios, other analytical parameters should not be overlooked. Advanced sensing requires not only high selectivity, excellent selectivity, and good applicability, but also a rapid response, smart sensing, and simple operation. Although several nanozyme strategies have been demonstrated as efficient for enhancing the first three parameters, the improvement of response speed, detection operation, and smart sensing is still urgently needed, which requires future efforts from the community. For example, fabricating portable detection devices by integrating nanozyme sensing with 3D-printed accessories and smartphone-based sensing technology would enable the on-site rapid analysis of targets [136]. Ingenious integration of nanozymes and their reaction substrates, as well as other sensing elements (such as bioenzymes) in one carrier, makes it possible to simplify the detection operation [56]; (3) effective combination of nanozyme materials and analytical methods is expected. For a sensor, its analytical performance relies on both the sensing material used and the sensing principle and method. Only the rational combination of sensing materials and analytical methods will result in sensors with the desired detection performance. Therefore, more attention should be focused on integrating sensing methods with efficient nanozyme materials for better biochemical analysis, including exploring new nanozyme categories, developing new sensing modes, and expanding the detectable targets.

Author Contributions

Conceptualization, X.N.; methodology, J.L.; software, J.L.; resources, J.L., X.N.; writing—original draft preparation, J.L.; writing—review and editing, X.N. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Acknowledgments

The authors would like to acknowledge the support from the Start-up Research Fund of University of South China. The work was also supported by the Faculty of Agricultural Equipment of Jiangsu University.

Conflicts of Interest

The authors declare no conflict of interest.

Abbreviations

AAAscorbic acid
AAPAscorbic acid 2-phosphate
ABTS2,2′-Azino-bis(3-ethylbenzothiazoline-6-sulfonic acid)
AChAcetylcholine
AChEAcetylcholinesterase
ACPAcid phosphatase
ALPAlkaline phosphatase
AspAspartic acid
BChEButyrylcholinesterase
CeNPsCeria nanoparticles
ChOCholine oxidase
CNTsCarbon nanotubes
CPNs(IV)Ce-based coordination polymer nanoparticles
DADopamine
DOPA3,4-Dihydroxyphenylalanine
GOxGlucose oxidase
GSHGlutathione
HRPHorseradish peroxidase
LODLimit of detection
MIPsMolecularly imprinted polymers
MOFsMetal-organic frameworks
OPOrganophosphorus
OPDo-Phenylenediamine
PDDAPoly(diallyldimethylammonium)
PhePhenylalanine
PiPhosphate ion
PPyPolypyrrole
SACsSingle atomic catalysts
SANsSingle-atom nanozymes
TATerephthalic acid
TAOH2-Hydroxyterephthalic acid
TCTetracycline
TMB3,3′,5,5′-Tetramethylbenzidine

References

  1. Wei, H.; Wang, E. Nanomaterials with Enzyme-like Characteristics (Nanozymes): Next-Generation Artificial Enzymes. Chem. Soc. Rev. 2013, 42, 6060–6093. [Google Scholar] [CrossRef] [PubMed]
  2. Wei, H.; Gao, L.; Fan, K.; Liu, J.; He, J.; Qu, X.; Dong, S.; Wang, E.; Yan, X. Nanozymes: A clear definition with fuzzy edges. Nano Today 2021, 40, 101269. [Google Scholar] [CrossRef]
  3. Liang, M.; Yan, X. Nanozymes: From new concepts, mechanisms, and standards to applications. Acc. Chem. Res. 2019, 52, 2190–2200. [Google Scholar] [CrossRef] [PubMed]
  4. Jiao, J.; Fan, K.; Hu, Z.; Yan, X.; Du, P. Development trend and priority areas of nanozyme. Sci. Sin. Chim. 2019, 49, 1442–1453. [Google Scholar]
  5. Ragg, R.; Tahir, M.N.; Tremel, W. Solids Go Bio: Inorganic Nanoparticles as Enzyme Mimics. Eur. J. Inorg. Chem. 2015, 2016, 1906–1915. [Google Scholar] [CrossRef]
  6. Wang, Q.; Liu, S.; Tang, Z. Recent progress in the design of analytical methods based on nanozymes. J. Mater. Chem. B 2021, 9, 8174–8184. [Google Scholar] [CrossRef]
  7. Niu, X.; Cheng, N.; Ruan, X.; Du, D.; Lin, Y. Review—Nanozyme-based immunosensors and immunoassays: Recent developments and future trends. J. Electrochem. Soc. 2020, 167, 037508. [Google Scholar] [CrossRef]
  8. Wang, Q.; Wei, H.; Zhang, Z.; Wang, E.; Dong, S. Nanozyme: An emerging alternative to natural enzyme for biosensing and immunoassay. TrAC Trends Anal. Chem. 2018, 105, 218–224. [Google Scholar] [CrossRef]
  9. Liu, B.; Liu, J. Surface modification of nanozymes. Nano Res. 2017, 10, 1125–1148. [Google Scholar] [CrossRef]
  10. Li, X.; Wang, L.; Du, D.; Ni, L.; Pan, J.; Niu, X. Emerging applications of nanozymes in environmental analysis: Opportunities and trends. TrAC Trends Anal. Chem. 2019, 120, 115653. [Google Scholar] [CrossRef]
  11. 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]
  12. Huang, L.; Sun, D.; Pu, H.; Wei, Q. Development of Nanozymes for Food Quality and Safety Detection: Principles and Recent Applications. Compr. Rev. Food Sci. Food Saf. 2019, 18, 1496–1513. [Google Scholar] [CrossRef]
  13. Li, S.; Zhang, Y.; Wang, Q.; Lin, A.; Wei, H. Nanozyme-Enabled Analytical Chemistry. Anal. Chem. 2021, 94, 312–323. [Google Scholar] [CrossRef]
  14. Li, X.; Zhu, H.; Liu, P.; Wang, M.; Pan, J.; Qiu, F.; Ni, L.; Niu, X. Realizing selective detection with nanozymes: Strategies and trends. TrAC Trends Anal. Chem. 2021, 143, 116379. [Google Scholar] [CrossRef]
  15. Wang, H.; Wan, K.; Shi, X. Recent Advances in Nanozyme Research. Adv. Mater. 2019, 31, 1805368. [Google Scholar] [CrossRef]
  16. Tian, Z.; Li, J.; Zhang, Z.; Gao, W.; Zhou, X.; Qu, Y. Highly sensitive and robust peroxidase-like activity of porous nanorods of ceria and their application for breast cancer detection. Biomaterials 2015, 59, 116–124. [Google Scholar] [CrossRef]
  17. Shu, Q.W.; Li, C.M.; Gao, P.F.; Gao, M.X.; Huang, C.Z. Porous hollow CuS nanospheres with prominent peroxidase-like activity prepared in large scale by a one-pot controllable hydrothermal step. RSC Adv. 2015, 5, 17458–17465. [Google Scholar] [CrossRef]
  18. Wu, H.; Liu, Y.; Li, M.; Chong, Y.; Zeng, M.; Lo, Y.M.; Yin, J.-J. Size-dependent tuning of horseradish peroxidase bioreactivity by gold nanoparticles. Nanoscale 2015, 7, 4505–4513. [Google Scholar] [CrossRef]
  19. Xi, Z.; Gao, W.; Xia, X. Size effect in Pd−Ir core-shell nanoparticles as nanozymes. ChemBioChem 2020, 21, 2440–2444. [Google Scholar] [CrossRef]
  20. Tang, G.; He, J.; Liu, J.; Yan, X.; Fan, K. Nanozyme for tumor therapy: Surface modification matters. Exploration 2021, 1, 75–89. [Google Scholar] [CrossRef]
  21. Jampaiah, D.; Reddy, T.S.; Kandjani, A.E.; Selvakannan, P.R.; Sabri, Y.M.; Coyle, V.E.; Shukla, R.; Bhargava, S.K. Fe-doped CeO2 nanorods for enhanced peroxidase-like activity and their application towards glucose detection. J. Mater. Chem. B 2016, 4, 3874–3885. [Google Scholar] [CrossRef]
  22. Gao, M.; Lu, X.; Chen, S.; Tian, D.; Zhu, Y.; Wang, C. Enhanced peroxidase-like activity of Mo6+-doped Co3O4 nanotubes for ultrasensitive and colorimetric L-cysteine detection. ACS Appl. Nano Mater. 2018, 1, 4703–4715. [Google Scholar] [CrossRef]
  23. Nagvenkar, A.P.; Gedanken, A. Cu0.89Zn0.11O, A New Peroxidase-Mimicking Nanozyme with High Sensitivity for Glucose and Antioxidant Detection. ACS Appl. Mater. Interfaces 2016, 8, 22301–22308. [Google Scholar] [CrossRef] [PubMed]
  24. Liu, M.; Zhao, H.; Chen, S.; Yu, H.; Quan, X. Stimuli-responsive peroxidase mimicking at a smart graphene interface. Chem. Commun. 2012, 48, 7055–7057. [Google Scholar] [CrossRef] [PubMed]
  25. Huang, Y.; Ren, J.; Qu, X. Nanozymes: Classification, Catalytic Mechanisms, Activity Regulation, and Applications. Chem. Rev. 2019, 119, 4357–4412. [Google Scholar] [CrossRef] [PubMed]
  26. 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. 2018, 48, 1004–1076. [Google Scholar] [CrossRef] [PubMed]
  27. Qiao, B.; Wang, A.; Yang, X.; Allard, L.F.; Jiang, Z.; Cui, Y.; Liu, J.; Li, J.; Zhang, T. Single-atom catalysis of CO oxidation using Pt1/FeOx. Nat. Chem. 2011, 3, 634–641. [Google Scholar] [CrossRef]
  28. Wang, A.; Li, J.; Zhang, T. Heterogeneous single-atom catalysis. Nat. Rev. Chem. 2018, 2, 65–81. [Google Scholar] [CrossRef]
  29. Kaiser, S.K.; Chen, Z.; Akl, D.F.; Mitchell, S.; Pérez-Ramírez, J. Single-Atom Catalysts across the Periodic Table. Chem. Rev. 2020, 120, 11703–11809. [Google Scholar] [CrossRef]
  30. Jiao, L.; Yan, H.; Wu, Y.; Gu, W.; Zhu, C.; Du, D.; Lin, Y. When Nanozymes Meet Single-Atom Catalysis. Angew. Chem. Int. Ed. 2019, 59, 2565–2576. [Google Scholar] [CrossRef]
  31. Ding, S.; Lyu, Z.; Fang, L.; Li, T.; Zhu, W.; Li, S.; Li, X.; Li, J.C.; Du, D.; Lin, Y. Single-atomic site catalyst with heme enzymes-like active sites for electrochemical sensing of hydrogen peroxide. Small 2021, 17, 2100664. [Google Scholar] [CrossRef]
  32. Wu, W.; Huang, L.; Wang, E.; Dong, S. Atomic engineering of single-atom nanozymes for enzyme-like catalysis. Chem. Sci. 2020, 11, 9741–9756. [Google Scholar] [CrossRef]
  33. Zhao, C.; Xiong, C.; Liu, X.; Qiao, M.; Li, Z.; Yuan, T.; Wang, J.; Qu, Y.; Wang, X.; Zhou, F.; et al. Unraveling the enzyme-like activity of heterogeneous single atom catalyst. Chem. Commun. 2019, 55, 2285–2288. [Google Scholar] [CrossRef]
  34. Ma, W.; Mao, J.; Yang, X.; Pan, C.; Chen, W.; Wang, M.; Yu, P.; Mao, L.; Li, Y. A single-atom Fe–N4 catalytic site mimicking bi-functional antioxidative enzymes for oxidative stress cytoprotection. Chem. Commun. 2019, 55, 159–162. [Google Scholar] [CrossRef]
  35. Li, X.; Liu, L.; Ren, X.; Gao, J.; Huang, Y.; Liu, B. Microenvironment modulation of single-atom catalysts and their roles in electrochemical energy conversion. Sci. Adv. 2020, 6, eabb6833. [Google Scholar] [CrossRef]
  36. Xu, B.; Wang, H.; Wang, W.; Gao, L.; Li, S.; Pan, X.; Wang, H.; Yang, H.; Meng, X.; Wu, Q.; et al. A Single-Atom Nanozyme for Wound Disinfection Applications. Angew. Chem. Int. Ed. 2019, 58, 4911–4916. [Google Scholar] [CrossRef]
  37. Huang, L.; Chen, J.; Gan, L.; Wang, J.; Dong, S. Single-atom nanozymes. Sci. Adv. 2019, 5, eaav5490. [Google Scholar] [CrossRef]
  38. Ji, S.; Jiang, B.; Hao, H.; Chen, Y.; Dong, J.; Mao, Y.; Zhang, Z.; Gao, R.; Chen, W.; Zhang, R.; et al. Matching the kinetics of natural enzymes with a single-atom iron nanozyme. Nat. Catal. 2021, 4, 407–417. [Google Scholar] [CrossRef]
  39. Lin, S.; Wei, H. Design of high performance nanozymes: A single-atom strategy. Sci. China Life Sci. 2019, 62, 710–712. [Google Scholar] [CrossRef]
  40. Jin, H.; Ye, D.; Shen, L.; Fu, R.; Tang, Y.; Jung, J.C.-Y.; Zhao, H.; Zhang, J. Perspective for Single Atom Nanozymes Based Sensors: Advanced Materials, Sensing Mechanism, Selectivity Regulation, and Applications. Anal. Chem. 2022, 94, 1499–1509. [Google Scholar] [CrossRef]
  41. 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, 1901485. [Google Scholar] [CrossRef]
  42. Wu, Y.; Jiao, L.; Luo, X.; Xu, W.; Wei, X.; Wang, H.; Yan, H.; Gu, W.; Xu, B.Z.; Du, D.; et al. Oxidase-like Fe-N-C single-atom nanozymes for the detection of acetylcholinesterase activity. Small 2019, 15, 1903108. [Google Scholar] [CrossRef]
  43. Lyu, Z.; Ding, S.; Zhang, N.; Zhou, Y.; Cheng, N.; Wang, M.; Xu, M.; Feng, Z.; Niu, X.; Cheng, Y.; et al. Single-atom nanozymes linked immunosorbent assay for sensitive detection of Aβ 1-40: A biomarker of Alzheimer’s disease. Research 2020, 2020, 4724505. [Google Scholar] [CrossRef]
  44. Niu, X.; Shi, Q.; Zhu, W.; Liu, D.; Tian, H.; Fu, S.; Cheng, N.; Li, S.; Smith, J.N.; Du, D.; et al. Unprecedented peroxidase-mimicking activity of single-atom nanozyme with atomically dispersed Fe–Nx moieties hosted by MOF derived porous carbon. Biosens. Bioelectron. 2019, 142, 111495. [Google Scholar] [CrossRef]
  45. Chen, Q.; Li, S.; Liu, Y.; Zhang, X.; Tang, Y.; Chai, H.; Huang, Y. Size-controllable Fe-N/C single-atom nanozyme with exceptional oxidase-like activity for sensitive detection of alkaline phosphatase. Sensors Actuators B Chem. 2019, 305, 127511. [Google Scholar] [CrossRef]
  46. Wu, Y.; Wu, J.; Jiao, L.; Xu, W.; Wang, H.; Wei, X.; Gu, W.; Ren, G.; Zhang, N.; Zhang, Q.; et al. Cascade Reaction System Integrating Single-Atom Nanozymes with Abundant Cu Sites for Enhanced Biosensing. Anal. Chem. 2020, 92, 3373–3379. [Google Scholar] [CrossRef]
  47. Wang, D.; Zhao, Y. Single-atom engineering of metal-organic frameworks toward healthcare. Chem 2021, 7, 2635–2671. [Google Scholar] [CrossRef]
  48. Jiang, B.; Duan, D.; Gao, L.; Zhou, M.; Fan, K.; Tang, Y.; Xi, J.; Bi, Y.; Tong, Z.; Gao, G.F.; et al. Standardized assays for determining the catalytic activity and kinetics of peroxidase-like nanozymes. Nat. Protoc. 2018, 13, 1506–1520. [Google Scholar] [CrossRef]
  49. Wu, J.; Xiong, L.; Zhao, B.; Liu, M.; Huang, L. Densely Populated Single Atom Catalysts. Small Methods 2019, 4, 2106139. [Google Scholar] [CrossRef]
  50. Zhang, X.; Li, G.; Chen, G.; Wu, D.; Wu, Y.; James, T.D. Enzyme Mimics for Engineered Biomimetic Cascade Nanoreactors: Mechanism, Applications, and Prospects. Adv. Funct. Mater. 2021, 31, 2106139. [Google Scholar] [CrossRef]
  51. Cai, X.; Jiao, L.; Yan, H.; Wu, Y.; Gu, W.; Du, D.; Lin, Y.; Zhu, C. Nanozyme-involved biomimetic cascade catalysis for biomedical applications. Mater. Today 2021, 44, 211–228. [Google Scholar] [CrossRef]
  52. Niu, X.; Liu, B.; Hu, P.; Zhu, H.; Wang, M. Nanozymes with Multiple Activities: Prospects in Analytical Sensing. Biosensors 2022, 12, 251. [Google Scholar] [CrossRef] [PubMed]
  53. Ma, Y.; Tian, Z.; Zhai, W.; Qu, Y. Insights on catalytic mechanism of CeO2 as multiple nanozymes. Nano Res. 2022, 1–15. [Google Scholar] [CrossRef] [PubMed]
  54. Guo, S.; Guo, L. Unraveling the Multi-Enzyme-Like Activities of Iron Oxide Nanozyme via a First-Principles Microkinetic Study. J. Phys. Chem. C 2019, 123, 30318–30334. [Google Scholar] [CrossRef]
  55. Liu, J.; Zhang, W.; Peng, M.; Ren, G.; Guan, L.; Li, K.; Lin, Y. ZIF-67 as a template generating and tuning “raisin pudding”-type nanozymes with multiple enzyme-like activities: Toward online electrochemical detection of 3,4-dihydroxyphenylacetic acid in living brains. ACS Appl. Mater. Interfaces 2020, 12, 29631–29640. [Google Scholar] [CrossRef] [PubMed]
  56. Zhu, H.; Xu, L.; Hu, P.; Liu, B.; Wang, M.; Yin, X.; Pan, J.; Niu, X. Smartphone-assisted bioenzyme-nanozyme-chromogen all-in-one test strip with enhanced cascade signal amplification for convenient paraoxon sensing. Biosens. Bioelectron. 2022, 215, 114583. [Google Scholar] [CrossRef]
  57. Sengupta, P.; Pramanik, K.; Datta, P.; Sarkar, P. Chemically modified carbon nitride-chitin-acetic acid hybrid as a metal-free bifunctional nanozyme cascade of glucose oxidase-peroxidase for “click off” colorimetric detection of peroxide and glucose. Biosens. Bioelectron. 2020, 154, 112072. [Google Scholar] [CrossRef]
  58. Lin, A.; Sun, Z.; Xu, X.; Zhao, S.; Li, J.; Sun, H.; Wang, Q.; Jiang, Q.; Wei, H.; Shi, D. Self-Cascade Uricase/Catalase Mimics Alleviate Acute Gout. Nano Lett. 2021, 22, 508–516. [Google Scholar] [CrossRef]
  59. Meng, X.; Li, D.; Chen, L.; He, H.; Wang, Q.; Hong, C.; He, J.; Gao, X.; Yang, Y.; Jiang, B.; et al. High-Performance Self-Cascade Pyrite Nanozymes for Apoptosis–Ferroptosis Synergistic Tumor Therapy. ACS Nano 2021, 15, 5735–5751. [Google Scholar] [CrossRef]
  60. Tao, N.; Li, H.; Deng, L.; Zhao, S.; Ouyang, J.; Wen, M.; Chen, W.; Zeng, K.; Wei, C.; Liu, Y.-N. A Cascade Nanozyme with Amplified Sonodynamic Therapeutic Effects through Comodulation of Hypoxia and Immunosuppression against Cancer. ACS Nano 2021, 16, 485–501. [Google Scholar] [CrossRef]
  61. Wang, Q.; Cheng, C.; Zhao, S.; Liu, Q.; Zhang, Y.; Liu, W.; Zhao, X.; Zhang, H.; Pu, J.; Zhang, S.; et al. A valence-engineered self-cascading antioxidant nanozyme for the therapy of inflammatory bowel disease. Angew. Chem. Int. Ed. 2022, 61, e202201101. [Google Scholar]
  62. Zhang, X.; Zhang, S.; Yang, Z.; Wang, Z.; Tian, X.; Zhou, R. Self-cascade MoS2 nanozymes for efficient intracellular antioxidation and hepatic fibrosis therapy. Nanoscale 2021, 13, 12613–12622. [Google Scholar] [CrossRef]
  63. Li, X.; Zhou, H.; Qi, F.; Niu, X.; Xu, X.; Qiu, F.; He, Y.; Pan, J.; Ni, L. Three hidden talents in one framework: A terephthalic acid-coordinated cupric metal–organic framework with cascade cysteine oxidase- and peroxidase-mimicking activities and stimulus-responsive fluorescence for cysteine sensing. J. Mater. Chem. B 2018, 6, 6207–6211. [Google Scholar] [CrossRef]
  64. He, L.; Lu, Y.; Gao, X.; Song, P.; Huang, Z.; Liu, S.; Liu, Y. Self-cascade system based on cupric oxide nanoparticles as dual-functional enzyme mimics for ultrasensitive detection of silver ions. ACS Sustain. Chem. Eng. 2018, 6, 12132–12139. [Google Scholar] [CrossRef]
  65. Chen, L.F.; Lin, M.T.; Noreldeen, H.A.A.; Peng, H.P.; Deng, H.H.; He, S.B.; Chen, W. Fructose oxidase-like activity of CuO nanoparticles supported by phosphate for a tandem catalysis-based fructose sensor. Anal. Chim. Acta 2022, 1220, 340064. [Google Scholar] [CrossRef]
  66. He, S.-B.; Balasubramanian, P.; Hu, A.-L.; Zheng, X.-Q.; Lin, M.-T.; Xiao, M.-X.; Peng, H.-P.; Deng, H.-H.; Chen, W. One-pot cascade catalysis at neutral pH driven by CuO tandem nanozyme for ascorbic acid and alkaline phosphatase detection. Sens. Actuators B Chem. 2020, 321, 128511. [Google Scholar] [CrossRef]
  67. Liu, Y.; Jin, H.; Zou, W.; Guo, R. Protein-mediated wool-ball-like copper sulfide as a multifunctional nanozyme for dual fluorescence “turn-on” sensors of cysteine and silver ions. J. Mater. Chem. B 2020, 8, 9075–9083. [Google Scholar] [CrossRef]
  68. Wang, Z.; Zhang, R.; Yan, X.; Fan, K. Structure and activity of nanozymes: Inspirations for de novo design of nanozymes. Mater. Today 2020, 41, 81–119. [Google Scholar] [CrossRef]
  69. Wei, H.; Wang, E. Fe3O4 magnetic nanoparticles as peroxidase mimetics and their applications in H2O2 and glucose detection. Anal. Chem. 2008, 80, 2250–2254. [Google Scholar] [CrossRef]
  70. Wang, C.; Tang, G.; Tan, H. Colorimetric determination of mercury(II) via the inhibition by ssDNA of the oxidase-like activity of a mixed valence state cerium-based metal-organic framework. Mikrochim. Acta 2018, 185, 475. [Google Scholar] [CrossRef]
  71. Weerathunge, P.; Ramanathan, R.; Shukla, R.; Sharma, T.K.; Bansal, V. Aptamer-Controlled Reversible Inhibition of Gold Nanozyme Activity for Pesticide Sensing. Anal. Chem. 2014, 86, 11937–11941. [Google Scholar] [CrossRef]
  72. Li, S.; Zhao, X.; Yu, X.; Wan, Y.; Yin, M.; Zhang, W.; Cao, B.; Wang, H. Fe3O4 nanozymes with aptamer-tuned catalysis for selective colorimetric analysis of ATP in blood. Anal. Chem. 2019, 91, 14737–14742. [Google Scholar] [CrossRef]
  73. Liu, B.; Huang, Z.; Liu, J. Boosting the oxidase mimicking activity of nanoceria by fluoride capping: Rivaling protein enzymes and ultrasensitive F detection. Nanoscale 2016, 8, 13562–13567. [Google Scholar] [CrossRef] [PubMed]
  74. Long, Y.J.; Li, Y.F.; Liu, Y.; Zheng, J.J.; Tang, J.; Huang, C.Z. Visual observation of the mercury-stimulated peroxidase mimetic activity of gold nanoparticles. Chem. Commun. 2011, 47, 11939–11941. [Google Scholar] [CrossRef]
  75. Chen, J.; Huang, L.; Wang, Q.; Wu, W.; Zhang, H.; Fang, Y.; Dong, S. Bio-inspired nanozyme: A hydratase mimic in a zeolitic imidazolate framework. Nanoscale 2019, 11, 5960–5966. [Google Scholar] [CrossRef]
  76. Li, M.; Chen, J.; Wu, W.; Fang, Y.; Dong, S. Oxidase-like MOF-818 nanozyme with high specificity for catalysis of catechol oxidation. J. Am. Chem. Soc. 2020, 142, 15569–15574. [Google Scholar] [CrossRef]
  77. Sun, Y.; Zhao, C.; Gao, N.; Ren, J.; Qu, X. Stereoselective Nanozyme Based on Ceria Nanoparticles Engineered with Amino Acids. Chem.–A Eur. J. 2017, 23, 18146–18150. [Google Scholar] [CrossRef]
  78. Wang, Y.; Jia, G.; Cui, X.; Zhao, X.; Zhang, Q.; Gu, L.; Zheng, L.; Li, L.H.; Wu, Q.; Singh, D.J.; et al. Coordination Number Regulation of Molybdenum Single-Atom Nanozyme Peroxidase-like Specificity. Chem 2020, 7, 436–449. [Google Scholar] [CrossRef]
  79. Zhou, Y.; Wei, Y.; Ren, J.; Qu, X. A chiral covalent organic framework (COF) nanozyme with ultrahigh enzymatic activity. Mater. Horizons 2020, 7, 3291–3297. [Google Scholar] [CrossRef]
  80. Zhang, R.; Zhou, Y.; Yan, X.; Fan, K. Advances in chiral nanozymes: A review. Microchim. Acta 2019, 186, 782. [Google Scholar] [CrossRef] [PubMed]
  81. Dong, K.; Xu, C.; Ren, J.; Qu, X. Chiral Nanozymes for Enantioselective Biological Catalysis. Angew. Chem. Int. Ed. 2022. [Google Scholar] [CrossRef]
  82. BelBruno, J.J. Molecularly imprinted polymers. Chem. Rev. 2019, 119, 94–119. [Google Scholar] [CrossRef] [PubMed]
  83. Zhang, Z.; Zhang, X.; Liu, B.; Liu, J. Molecular Imprinting on Inorganic Nanozymes for Hundred-fold Enzyme Specificity. J. Am. Chem. Soc. 2017, 139, 5412–5419. [Google Scholar] [CrossRef] [PubMed]
  84. Zhang, Z.; Liu, J. Intracellular delivery of a molecularly imprinted peroxidase mimicking DNAzyme for selective oxidation. Mater. Horizons 2018, 5, 738–744. [Google Scholar] [CrossRef]
  85. Zhang, Z.; Liu, B.; Liu, J. Molecular Imprinting for Substrate Selectivity and Enhanced Activity of Enzyme Mimics. Small 2016, 13, 1602730. [Google Scholar] [CrossRef]
  86. Zhang, Z.; Li, Y.; Zhang, X.; Liu, J. Molecularly imprinted nanozymes with faster catalytic activity and better specificity. Nanoscale 2019, 11, 4854–4863. [Google Scholar] [CrossRef]
  87. Zhang, Z.; Liu, Y.; Zhang, X.; Liu, J. A Cell-Mimicking Structure Converting Analog Volume Changes to Digital Colorimetric Output with Molecular Selectivity. Nano Lett. 2017, 17, 7926–7931. [Google Scholar] [CrossRef]
  88. Liu, B.; Zhu, H.; Feng, R.; Wang, M.; Hu, P.; Pan, J.; Niu, X. Facile Molecular Imprinting on Magnetic Nanozyme Surface for Highly Selective Colorimetric Detection of Tetracycline. Sens. Actuators B Chem. 2022, 370, 132451. [Google Scholar] [CrossRef]
  89. Fan, L.; Lou, D.; Wu, H.; Zhang, X.; Zhu, Y.; Gu, N.; Zhang, Y. A novel AuNP-based glucose oxidase mimic with enhanced activity and selectivity constructed by molecular imprinting and O2-containing nanoemulsion embedding. Adv. Mater. Interfaces 2018, 5, 1801070. [Google Scholar] [CrossRef]
  90. Wu, Y.; Chen, Q.; Liu, S.; Xiao, H.; Zhang, M.; Zhang, X. Surface molecular imprinting on g-C3N4 photooxidative nanozyme for improved colorimetric biosensing. Chin. Chem. Lett. 2019, 30, 2186–2190. [Google Scholar] [CrossRef]
  91. Zhang, Y.; Feng, Y.S.; Ren, X.H.; He, X.W.; Li, W.Y.; Zhang, Y.K. Bimetallic molecularly imprinted nanozyme: Dual-mode detection platform. Biosens. Bioelectron. 2022, 196, 113718. [Google Scholar] [CrossRef]
  92. Cardoso, A.R.; Frasco, M.F.; Serrano, V.; Fortunato, E.; Sales, M.G.F. Molecular imprinting on nanozymes for sensing applications. Biosensors 2021, 11, 152. [Google Scholar] [CrossRef]
  93. Li, J.; Liu, W.; Wu, X.; Gao, X. Mechanism of pH-switchable peroxidase and catalase-like activities of gold, silver, platinum and palladium. Biomaterials 2015, 48, 37–44. [Google Scholar] [CrossRef]
  94. Zhang, J.; Wang, J.; Liao, J.; Lin, Y.; Zheng, C.; Liu, J. In Situ Fabrication of Nanoceria with Oxidase-like Activity at Neutral pH: Mechanism and Boosted Bio-Nanozyme Cascades. ACS Appl. Mater. Interfaces 2021, 13, 50236–50245. [Google Scholar] [CrossRef]
  95. Zhang, J.; Wu, S.; Lu, X.; Wu, P.; Liu, J. Manganese as a Catalytic Mediator for Photo-oxidation and Breaking the pH Limitation of Nanozymes. Nano Lett. 2019, 19, 3214–3220. [Google Scholar] [CrossRef]
  96. He, Y.; Li, X.; Xu, X.; Pan, J.; Niu, X. A cobalt-based polyoxometalate nanozyme with high peroxidase-mimicking activity at neutral pH for one-pot colorimetric analysis of glucose. J. Mater. Chem. B 2018, 6, 5750–5755. [Google Scholar] [CrossRef]
  97. Li, D.; Guo, Q.; Ding, L.; Zhang, W.; Cheng, L.; Wang, Y.; Xu, Z.; Wang, H.; Gao, L. Bimetallic CuCo2S4 nanozymes with enhanced peroxidase activity at neutral pH for combating burn infections. ChemBioChem 2020, 21, 2620–2627. [Google Scholar] [CrossRef]
  98. Nguyen, P.T.; Lee, J.; Cho, A.; Kim, M.S.; Choi, D.; Han, J.W.; Kim, M.I.; Lee, J. Rational development of Co-doped mesoporous ceria with high peroxidase-mimicking activity at neutral pH for paper-based colorimetric detection of multiple biomarkers. Adv. Funct. Mater. 2022, 32, 2112428. [Google Scholar] [CrossRef]
  99. Zhu, Q.; Yang, J.; Peng, Z.; He, Z.; Chen, W.; Tang, H.; Li, Y. Selective detection of glutathione by flower-like NiV2O6 with only peroxidase-like activity at neutral pH. Talanta 2021, 234, 122645. [Google Scholar] [CrossRef]
  100. Niu, X.; Xu, X.; Li, X.; Pan, J.; Qiu, F.; Zhao, H.; Lan, M. Surface charge engineering of nanosized CuS via acidic amino acid modification enables high peroxidase-mimicking activity at neutral pH for one-pot detection of glucose. Chem. Commun. 2018, 54, 13443–13446. [Google Scholar] [CrossRef]
  101. Han, L.; Li, C.; Zhang, T.; Lang, Q.; Liu, A. Au@Ag Heterogeneous Nanorods as Nanozyme Interfaces with Peroxidase-Like Activity and Their Application for One-Pot Analysis of Glucose at Nearly Neutral pH. ACS Appl. Mater. Interfaces 2015, 7, 14463–14470. [Google Scholar] [CrossRef]
  102. Chen, Q.; Liu, Y.; Liu, J.; Liu, J. Liposome-boosted peroxidase-mimicking nanozymes breaking the pH limit. Chem. Eur. J. 2020, 26, 16659–16665. [Google Scholar] [CrossRef]
  103. Gu, H.; Huang, Q.; Zhang, J.; Li, W.; Fu, Y. Heparin as a bifunctional biotemplate for Pt nanocluster with exclusively peroxidase mimicking activity at near-neutral pH. Colloids Surf. A Physicochem. Eng. Asp. 2020, 606. [Google Scholar] [CrossRef]
  104. Yin, X.; Liu, P.; Xu, X.; Pan, J.; Li, X.; Niu, X. Breaking the pH limitation of peroxidase-like CoFe2O4 nanozyme via vitriolization for one-step glucose detection at physiological pH. Sens. Actuators B Chem. 2020, 328, 129033. [Google Scholar] [CrossRef]
  105. Wang, L.; Chen, Y. Luminescence-Sensing Tb-MOF Nanozyme for the Detection and Degradation of Estrogen Endocrine Disruptors. ACS Appl. Mater. Interfaces 2020, 12, 8351–8358. [Google Scholar] [CrossRef]
  106. Wu, J.; Qin, K.; Yuan, D.; Tan, J.; Qin, L.; Zhang, X.; Wei, H. Rational design of Au@Pt multibranched nanostructures as bi-functional nanozymes. ACS Appl. Mater. Interfaces 2018, 10, 12954–12959. [Google Scholar] [CrossRef]
  107. Zhu, N.; Gu, L.; Wang, J.; Li, X.; Liang, G.-X.; Zhou, J.; Zhang, Z. Novel and Sensitive Chemiluminescence Sensors Based on 2D-MOF Nanosheets for One-Step Detection of Glucose in Human Urine. J. Phys. Chem. C 2019, 123, 9388–9393. [Google Scholar] [CrossRef]
  108. Lai, X.; Zhang, G.; Zeng, L.; Xiao, X.; Peng, J.; Guo, P.; Zhang, W.; Lai, W. Synthesis of PDA-Mediated Magnetic Bimetallic Nanozyme and Its Application in Immunochromatographic Assay. ACS Appl. Mater. Interfaces 2020, 13, 1413–1423. [Google Scholar] [CrossRef]
  109. Gökçal, B.; Kip, Ç.; Tuncel, A. One-pot, direct glucose detection in human whole blood without using a dilution factor by a magnetic nanozyme with dual enzymatic activity. J. Alloys Compd. 2020, 843, 156012. [Google Scholar] [CrossRef]
  110. Xu, X.; Wu, S.; Guo, D.; Niu, X. Construction of a recyclable oxidase-mimicking Fe3O4@MnOx-based colorimetric sensor array for quantifying and identifying chlorophenols. Anal. Chim. Acta 2020, 1107, 203–212. [Google Scholar] [CrossRef]
  111. Wen, S.H.; Zhong, X.L.; Wu, Y.D.; Liang, R.P.; Zhang, L.; Qiu, J.D. Colorimetric assay conversion to highly sensitive electro-chemical assay for bimodal detection of arsenate based on cobalt oxyhydroxide nanozyme via arsenate absorption. Anal. Chem. 2019, 91, 6487–6497. [Google Scholar] [CrossRef] [PubMed]
  112. Cheng, H.; Zhang, L.; He, J.; Guo, W.; Zhou, Z.; Zhang, X.; Nie, S.; Wei, H. Integrated Nanozymes with Nanoscale Proximity for in Vivo Neurochemical Monitoring in Living Brains. Anal. Chem. 2016, 88, 5489–5497. [Google Scholar] [CrossRef] [PubMed]
  113. Zhao, Z.; Huang, Y.; Liu, W.; Ye, F.; Zhao, S. Immobilized Glucose Oxidase on Boronic Acid-Functionalized Hierarchically Porous MOF as an Integrated Nanozyme for One-Step Glucose Detection. ACS Sustain. Chem. Eng. 2020, 8, 4481–4488. [Google Scholar] [CrossRef]
  114. Shen, H.; Zhou, Z.; He, W.; Chao, H.; Su, P.; Song, J.; Yang, Y. Oligonucleotide-Functionalized Enzymes Chemisorbing on Magnetic Layered Double Hydroxides: A Multimodal Catalytic Platform with Boosted Activity for Ultrasensitive Glucose Detection. ACS Appl. Mater. Interfaces 2021, 13, 14995–15007. [Google Scholar] [CrossRef]
  115. Wu, J.; Li, S.; Wei, H. Multifunctional nanozymes: Enzyme-like catalytic activity combined with magnetism and surface plasmon resonance. Nanoscale Horiz. 2018, 3, 367–382. [Google Scholar] [CrossRef]
  116. Wang, N.; Shi, J.; Liu, Y.; Sun, W.; Su, X. Constructing bifunctional metaleorganic framework based nanozymes with fluorescence and oxidase activity for the dualchannel detection of butyrylcholinesterase. Anal. Chim. Acta 2022, 1205, 339717. [Google Scholar] [CrossRef]
  117. Lin, T.; Qin, Y.; Huang, Y.; Yang, R.; Hou, L.; Ye, F.; Zhao, S. A label-free fluorescence assay for hydrogen peroxide and glucose based on the bifunctional MIL-53(Fe) nanozyme. Chem. Commun. 2018, 54, 1762–1765. [Google Scholar] [CrossRef]
  118. Ye, K.; Wang, L.; Song, H.; Li, X.; Niu, X. Bifunctional MIL-53(Fe) with pyrophosphate-mediated peroxidase-like activity and oxidation-stimulated fluorescence switching for alkaline phosphatase detection. J. Mater. Chem. B 2019, 7, 4794–4800. [Google Scholar] [CrossRef]
  119. Li, X.; Liu, P.; Niu, X.; Ye, K.; Ni, L.; Du, D.; Pan, J.; Lin, Y. Tri-functional Fe–Zr bi-metal–organic frameworks enable high-performance phosphate ion ratiometric fluorescent detection. Nanoscale 2020, 12, 19383–19389. [Google Scholar] [CrossRef]
  120. Guo, J.; Liu, Y.; Mu, Z.; Wu, S.; Wang, J.; Yang, Y.; Zhao, M.; Wang, Y. Label-free fluorescence detection of hydrogen peroxide and glucose based on the Ni-MOF nanozyme–induced self-ligand emission. Mikrochim. Acta 2022, 189, 219. [Google Scholar] [CrossRef]
  121. Sun, Y.; Lin, T.; Zeng, C.; Jiang, G.; Zhang, X.; Ye, F.; Zhao, S. A self-correcting fluorescent assay of tyrosinase based on Fe-MIL-88B-NH2 nanozyme. Mikrochim. Acta 2021, 188, 158. [Google Scholar] [CrossRef]
  122. Liu, P.; Li, X.; Xu, X.; Ye, K.; Wang, L.; Zhu, H.; Wang, M.; Niu, X. Integrating peroxidase-mimicking activity with photoluminescence into one framework structure for high-performance ratiometric fluorescent pesticide sensing. Sens. Actuators B Chem. 2021, 328, 129024. [Google Scholar] [CrossRef]
  123. Xu, X.; Luo, Z.; Ye, K.; Zou, X.; Niu, X.; Pan, J. One-pot construction of acid phosphatase and hemin loaded multifunctional metal–organic framework nanosheets for ratiometric fluorescent arsenate sensing. J. Hazard. Mater. 2020, 412, 124407. [Google Scholar] [CrossRef]
  124. Guo, J.; Wu, S.; Wang, Y.; Zhao, M. A label-free fluorescence biosensor based on a bifunctional MIL-101(Fe) nanozyme for sensitive detection of choline and acetylcholine at nanomolar level. Sens. Actuators B Chem. 2020, 312, 128021. [Google Scholar] [CrossRef]
  125. Hou, L.; Qin, Y.; Lin, T.; Sun, Y.; Ye, F.; Zhao, S. Michael reaction-assisted fluorescent sensor for selective and one step de-termination of catechol via bifunctional Fe-MIL-88NH2 nanozyme. Sens. Actuators B Chem. 2020, 321, 128547. [Google Scholar] [CrossRef]
  126. Li, S.; Hu, X.; Chen, Q.; Zhang, X.; Chai, H.; Huang, Y. Introducing bifunctional metal-organic frameworks to the construction of a novel ratiometric fluorescence sensor for screening acid phosphatase activity. Biosens. Bioelectron. 2019, 137, 133–139. [Google Scholar] [CrossRef]
  127. Xia, Y.; Sun, K.; Zuo, Y.-N.; Zhu, S.; Zhao, X.-E. Fluorescent MOF-based nanozymes for discrimination of phenylenediamine isomers and ratiometric sensing of o-phenylenediamine. Chin. Chem. Lett. 2021, 33, 2081–2085. [Google Scholar] [CrossRef]
  128. Liu, P.; Zhao, M.; Zhu, H.; Zhang, M.; Li, X.; Wang, M.; Liu, B.; Pan, J.; Niu, X. Dual-mode fluorescence and colorimetric detection of pesticides realized by integrating stimulus-responsive luminescence with oxidase-mimetic activity into cerium-based coordination polymer nanoparticles. J. Hazard. Mater. 2022, 423, 127077. [Google Scholar] [CrossRef]
  129. Liu, P.; Zhu, H.; Wang, M.; Wei, M.; Liu, B.; Hu, P.; Lin, J.; Niu, X. Bimodal ratiometric fluorescence and colorimetric sensing of paraoxon based on trifunctional Ce,Tb co-coordinated polymers. Sens. Actuators B Chem. 2022, 360, 131616. [Google Scholar] [CrossRef]
  130. Wu, Y.; Gao, Y.; Du, J. Bifunctional gold nanoclusters enable ratiometric fluorescence nanosensing of hydrogen peroxide and glucose. Talanta 2019, 197, 599–604. [Google Scholar] [CrossRef]
  131. Wan, Y.; Zhao, J.; Deng, X.; Chen, J.; Xi, F.; Wang, X. Colorimetric and Fluorescent Dual-Modality Sensing Platform Based on Fluorescent Nanozyme. Front. Chem. 2021, 9, 774486. [Google Scholar] [CrossRef]
  132. Li, X.; Ding, S.; Lyu, Z.; Tieu, P.; Wang, M.; Feng, Z.; Pan, X.; Zhou, Y.; Niu, X.; Du, D.; et al. Single-Atomic Iron Doped Carbon Dots with Both Photoluminescence and Oxidase-Like Activity. Small 2022, 18, 2203001. [Google Scholar] [CrossRef]
  133. Zhang, W.; Li, X.; Xu, X.; He, Y.; Qiu, F.; Pan, J.; Niu, X. Pd nanoparticle-decorated graphitic C3N4 nanosheets with bifunctional peroxidase mimicking and ON–OFF fluorescence enable naked-eye and fluorescent dual-readout sensing of glucose. J. Mater. Chem. B 2018, 7, 233–239. [Google Scholar] [CrossRef]
  134. Gooding, J.J. Can Nanozymes Have an Impact on Sensing? ACS Sens. 2019, 4, 2213–2214. [Google Scholar] [CrossRef]
  135. Zhou, Y.; Liu, B.; Yang, R.; Liu, J. Filling in the Gaps between Nanozymes and Enzymes: Challenges and Opportunities. Bioconjugate Chem. 2017, 28, 2903–2909. [Google Scholar] [CrossRef]
  136. Ruan, X.; Hulubei, V.; Wang, Y.; Shi, Q.; Cheng, N.; Wang, L.; Lyu, Z.; Davis, W.C.; Smith, J.N.; Lin, Y.; et al. Au@PtPd enhanced immunoassay with 3D printed smartphone device for quantification of diaminochlorotriazine (DACT), the major atrazine biomarker. Biosens. Bioelectron. 2022, 208, 114190. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Diagram for the rational design of nanozymes, enabling advanced biochemical sensing.
Figure 1. Diagram for the rational design of nanozymes, enabling advanced biochemical sensing.
Chemosensors 10 00386 g001
Figure 2. (A) Preparation of CNT/FeNC SAN with a high peroxidase-mimetic activity (reprinted with permission from Ref. [41]). (B) Fe-N-C SAN with prominent peroxidase-mimicking activity for paper-based bioassay of BChE activity (reprinted with permission from Ref. [44]). (C) Design of Cu-N-C SAN with rich Cu sites for organophosphorus sensing (reprinted with permission from Ref. [46]).
Figure 2. (A) Preparation of CNT/FeNC SAN with a high peroxidase-mimetic activity (reprinted with permission from Ref. [41]). (B) Fe-N-C SAN with prominent peroxidase-mimicking activity for paper-based bioassay of BChE activity (reprinted with permission from Ref. [44]). (C) Design of Cu-N-C SAN with rich Cu sites for organophosphorus sensing (reprinted with permission from Ref. [46]).
Chemosensors 10 00386 g002
Figure 3. (A) Comparison of nanozyme-involved cascade catalytic reactions with a single reaction for highly-sensitive sensing. (B) Design of CuBDC with cascade cysteine oxidase- and peroxidase-mimicking activities for cysteine detection (reprinted with permission from Ref. [63]). (C) Self-cascade system based on CuO NPs with GSH oxidase- and peroxidase-like activities (reprinted with permission from Ref. [64]).
Figure 3. (A) Comparison of nanozyme-involved cascade catalytic reactions with a single reaction for highly-sensitive sensing. (B) Design of CuBDC with cascade cysteine oxidase- and peroxidase-mimicking activities for cysteine detection (reprinted with permission from Ref. [63]). (C) Self-cascade system based on CuO NPs with GSH oxidase- and peroxidase-like activities (reprinted with permission from Ref. [64]).
Chemosensors 10 00386 g003
Figure 4. (A) Catechol oxidase-inspired design of MOF-818 showing a high catalytic specificity (reprinted with permission from Ref. [76]). (B) Chiral Phe-modified ceria nanoparticles as nanozymes for stereoselective catalysis (reprinted with permission from Ref. [77]). (C) Coordination number regulation of Mo-based SAN with peroxidase-like specificity (reprinted with permission from Ref. [78]).
Figure 4. (A) Catechol oxidase-inspired design of MOF-818 showing a high catalytic specificity (reprinted with permission from Ref. [76]). (B) Chiral Phe-modified ceria nanoparticles as nanozymes for stereoselective catalysis (reprinted with permission from Ref. [77]). (C) Coordination number regulation of Mo-based SAN with peroxidase-like specificity (reprinted with permission from Ref. [78]).
Chemosensors 10 00386 g004
Figure 5. (A) Molecularly imprinted layers grown on Fe3O4 NPs with substrate binding pockets for enhanced nanozyme specificity (reprinted with permission from Ref. [83]). (B) Molecular imprinting on Fe3O4 particle surface to gain TC colorimetric detection with excellent selectivity (reprinted with permission from Ref. [88]).
Figure 5. (A) Molecularly imprinted layers grown on Fe3O4 NPs with substrate binding pockets for enhanced nanozyme specificity (reprinted with permission from Ref. [83]). (B) Molecular imprinting on Fe3O4 particle surface to gain TC colorimetric detection with excellent selectivity (reprinted with permission from Ref. [88]).
Chemosensors 10 00386 g005
Figure 6. (A) Strategies to obtain nanozymes that can break the pH limit. (B) Co-based nanozyme showing good peroxidase-like activity in a neutral solution (reprinted with permission from Ref. [96]). (C) Au@Ag NRs modified with positively charged PDDA, showing a pH-dependent peroxidase-like activity (reprinted with permission from Ref. [101]). (D) Surface modification of nanoscale CuS with aspartic acid to gain favorable peroxidase-mimetic activity at neutral pH (reprinted with permission from Ref. [100]).
Figure 6. (A) Strategies to obtain nanozymes that can break the pH limit. (B) Co-based nanozyme showing good peroxidase-like activity in a neutral solution (reprinted with permission from Ref. [96]). (C) Au@Ag NRs modified with positively charged PDDA, showing a pH-dependent peroxidase-like activity (reprinted with permission from Ref. [101]). (D) Surface modification of nanoscale CuS with aspartic acid to gain favorable peroxidase-mimetic activity at neutral pH (reprinted with permission from Ref. [100]).
Chemosensors 10 00386 g006
Figure 7. (A) Design of luminescent nanozymes for advanced fluorescence sensing. (B) Bifunctional MIL-53(Fe) nanozyme for label-free fluorescence detection of H2O2 and glucose (reprinted with permission from Ref. [117]). (C) Design of UiO-66(Fe/Zr)-NH2 enabling ratiometric fluorescence detection of Pi (reprinted with permission from Ref. [119]). (D) Bimodal colorimetric and fluorescence sensing of pesticides based on Ce-based coordination polymer nanoparticles with oxidase-like catalytic activity and stimulus-responsive luminescence (reprinted with permission from Ref. [128]).
Figure 7. (A) Design of luminescent nanozymes for advanced fluorescence sensing. (B) Bifunctional MIL-53(Fe) nanozyme for label-free fluorescence detection of H2O2 and glucose (reprinted with permission from Ref. [117]). (C) Design of UiO-66(Fe/Zr)-NH2 enabling ratiometric fluorescence detection of Pi (reprinted with permission from Ref. [119]). (D) Bimodal colorimetric and fluorescence sensing of pesticides based on Ce-based coordination polymer nanoparticles with oxidase-like catalytic activity and stimulus-responsive luminescence (reprinted with permission from Ref. [128]).
Chemosensors 10 00386 g007
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Liu, J.; Niu, X. Rational Design of Nanozymes Enables Advanced Biochemical Sensing. Chemosensors 2022, 10, 386. https://doi.org/10.3390/chemosensors10100386

AMA Style

Liu J, Niu X. Rational Design of Nanozymes Enables Advanced Biochemical Sensing. Chemosensors. 2022; 10(10):386. https://doi.org/10.3390/chemosensors10100386

Chicago/Turabian Style

Liu, Jinjin, and Xiangheng Niu. 2022. "Rational Design of Nanozymes Enables Advanced Biochemical Sensing" Chemosensors 10, no. 10: 386. https://doi.org/10.3390/chemosensors10100386

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