Advances in nanozymes with peroxidase-like activity for biosensing and disease therapy applications

Xiaohua Yuan ab, Xun He c, Jiwen Fan ab, Yunze Tai ab, Yongchao Yao ab, Yao Luo ab, Jie Chen ab, Han Luo ab, Xingli Zhou c, Fengming Luo cd, Qian Niu *ab, Wenchuang (Walter) Hu *ab, Xuping Sun *ce and Binwu Ying *ab
aDepartment of Laboratory Medicine/Clinical Laboratory Medicine Research Center, West China Hospital, Sichuan University, Chengdu 610041, Sichuan, China. E-mail: yingbinwu@scu.edu.cn; huwenchuang@wchscu.cn; niuqian991@scu.edu.cn
bFrontiers Science Center for Disease-related Molecular Network, West China Hospital, Sichuan University, Chengdu 610041, Sichuan, China
cCenter for High Altitude Medicine, West China Hospital, Sichuan University, Chengdu 610041, Sichuan, China. E-mail: sunxp@wchscu.edu.cn
dDepartment of Pulmonary and Critical Care Medicine, West China Hospital, Sichuan University, Chengdu 610041, Sichuan, China
eCollege of Chemistry, Chemical Engineering and Materials Science, Shandong Normal University, Jinan 250014, Shandong, China

Received 14th October 2024 , Accepted 9th December 2024

First published on 20th December 2024


Abstract

Natural enzymes are crucial in biological systems and widely used in biomedicine, but their disadvantages, such as insufficient stability and high cost, have limited their widespread application. Since discovering the enzyme-like activity of Fe3O4 nanoparticles, extensive research progress in diverse nanozymes has been made with their in-depth investigation, resulting in rapid development of related nanotechnologies. Nanozymes can compensate for the defects of natural enzymes and show higher stability with lower costs. Among them, peroxidase (POD)-like nanozymes have attracted extensive attention in biomedical applications owing to their efficient catalytic performance and diverse structures. This review explores different types of nanozymes with POD-like activity and discusses their activity regulation, particularly emphasizing their latest development trends and advances in biosensing and disease treatment. Finally, the challenges and prospects for the development of POD-like nanozymes and their potential future applications in the biomedical field are also provided.


1. Introduction

Natural enzymes are powerful biocatalysts that are used to selectively and efficiently catalyze various biochemical reactions. Natural enzymes have high selectivity and efficient catalytic activity and play an important role in many fields, such as food industry, disease diagnosis and treatment, and environmental monitoring.1 POD is a natural enzyme that enables the conversion of H2O2 into H2O, during which the substrate is oxidized to its corresponding oxide. Horseradish peroxidase (HRP), as a typical peroxidase, has been studied over the years as a useful tool to catalyze H2O2. However, disadvantages such as low operational stability, high cost, difficult storage, easy inactivation, cumbersome and time-consuming preparation have hindered its widespread application.2 Therefore, the development of new artificial enzymes to replace POD is of great significance to broaden their applications in biology.

In recent years, various nanomaterials have exhibited excellent catalytic activity by mimicking the function or structure of peroxidase, such as metal/gold composite nanoparticles, carbon-based nanomaterials, metal–organic frameworks (MOFs), and single-atom nanozymes (SAzymes). Because at least one of the three dimensions of nanomaterials is in the nanoscale range, their small size endows them with many specific effects such as small size effect, surface effect and quantum size effect, compared with other materials. In addition, nanomaterials have the advantages of flexible structural design and composition, as well as high active sites, which can enhance their catalytic activity. Since Yan et al. discovered that Fe3O4 magnetic nanoparticles have POD-like properties, they have been widely used in many fields, such as the separation and capture of analytes, sensing, and treatment.3

POD-like nanozymes are dual-substrate nanozymes, which can catalyze the hydrogen acceptor to produce a large number of oxidative free radicals in the presence of the first substrate, which is a hydrogen acceptor (such as H2O2), and then rapidly oxidize the hydrogen donor, which is the second substrate. Depending on the hydrogen donor, POD-like nanozymes can mimic the activities of various enzymes, such as peroxidase, glutathione (GSH) peroxidase, and lipid peroxidase. There are various hydrogen receptors for peroxidases. For example, the hydrogen acceptor for HRP is H2O2, and the hydrogen acceptor for lipid peroxidase is lipid peroxide. The free radicals catalyzed by hydrogen acceptors can oxidize various hydrogen donor substrates, including small-molecule metabolites (formic acid, methanol, ethanol, etc.) and biological macromolecules (nucleic acids, proteins, polysaccharides, lipids, etc.). Taking the earliest horseradish POD-like nanozyme as an example, its activity detection method typically employs the same approach as that used for natural HRP.4 When the reaction is catalyzed by the POD-like nanozyme using H2O2 as the hydrogen acceptor, numerous intermediate ˙OH radicals are generated, which then oxidize hydrogen donors. In particular, the catalytic pathway of POD-like nanozyme mainly involves reactive oxygen species (ROS) formation and electron transfer. For ROS production, the nanozyme reacts with H2O2 to produce ROS (˙OH, O2−˙, HO2˙, etc.), which is responsible for further oxidation. Fe3O4 NPs are the first proposed nanozymes with intrinsic POD-like activity, in which the surface Fe2+ plays a leading role in activating H2O2 and producing ROS via the Fenton or Haber–Weiss reaction.5 For the electron transfer process, the substrate is first adsorbed onto the surface of the nanozyme, providing electrons. Large specific surface areas with abundant active sites or variable oxidation states confer the nanozyme with increased electron density and excellent mobility, resulting in redox reactions between the substrate and H2O2. Rapid electron transfer from the substrate to H2O2 further accelerates substrate oxidation and H2O2 reduction.6

While some reviews have been published on POD-like nanomaterials, it is worth noting that some of them have a narrow focus, discussing only a specific type of POD-like nanozyme or highlighting a limited range of biomedical applications. Based on this consideration, this article will provide a comprehensive review of POD-like nanomaterials and their biomedical applications (Fig. 1). First, we briefly introduce different types of nanozymes with POD-like activity. Then, strategies for endowing POD-like activity to nanozymes are summarized. Furthermore, the recent research progress of POD-like nanozymes in the field of biosensing and disease treatment is comprehensively reviewed. Finally, our insights into the prospects and challenges of the development of POD-like nanozymes are presented. The aim is to provide valuable and comprehensive information to scientists in the fields of materials science and biomedical research, and promote the development of interdisciplinary fields.


image file: d4tb02315c-f1.tif
Fig. 1 Overview of POD-like nanozymes for biomedical applications.

2. Classification of POD-like nanozymes

The simulated enzyme activity of nanomaterials is generally believed to be produced by the atoms on the surface and in the core of the material. Although it has been widely reported that size, morphology, surface modification, pH, and temperature affect the nanozyme activity, the elemental composition is still the kernel factor that determines the catalytic activity of the nanozyme. It has been reported that more than 40 elements and about 100 different types of nanozymes have been discovered. Its classification can be made from different angles. Based on material types, they are mainly classified into the following categories. Table 1 summarizes the synthesis, structure, and applications of various POD-like nanozymes.
Table 1 Different types of POD-like nanozymes
Nanozyme Preparation method Structure Application Ref.
Metal nanozymes GNPs General reduction method Nanoparticles The structure–activity relationship of catalytic properties of nanozyme was studied 7
Au/Pt NPs Coreduction method Nanoparticles Au/Pt alloy nanoshells with different catalytic properties were prepared 8
Au/Pt NPs Electrodeposition method Nanoparticles The size and morphology of nanozymes were controlled by controlling electrodeposition parameters 9
Ag/Au NPs Replacement reaction method Core–shell The solvent and temperature determine the structure and composition of the nanoparticles formed by the displacement reaction 10
Pd Nanocrystals Solvothermal method Either cubic or octahedral shapes To investigate the relationship between the catalytic activity of Pd nanozyme and the exposed crystal surface 11
Metal compound nanozymes Fe3O4 NPs Solvothermal ion-exchange methods Nanosphere To prepare different types of iron oxide nanozymes 12
Fe3O4@TiO2/TP Solvothermal and ion-exchange methods Nanoparticle preparation of efficient catalysts 13
GO NPs Hummers’ method and ultrasonic stripping method Clustered layer To prepare nanomaterials with POD-like activity for the determination of heavy metal ions 14
Fe3C@FeN-CM Hydrothermal method and high temperature calcination method Polyhedral structure Multi-enzyme activity and rapid detection of food quality and safety 15
IO-MC Impregnation method and high-temperature calcination method Rod-like morphology With POD-like activity, they realize highly sensitive detection of glucose 16
Carbon-based nanozymes N-Doped Carbon Sheets High temperature pyrolytic cracking Sheet structure For the determination of hydrogen peroxide, glucose and ascorbic acid 17
Metal–organic framework nanozymes MOF-818 Direct synthesis technique Octahedron in shape Specific catechol oxidase activity 18
AuPt/MOF Post-synthetic modification sheets structure A new metal ionizer or ligand is introduced into MOFs to give them enzyme-like catalytic activity 19
ZIF8 Composite preparation method Rhombic dodecahedral MOFs act as carriers to support other nanozymes 20
Single-atom Nanozymes FeN5 SA/CNF Spatial constraint synthesis strategy Mesoporous The active site of Fe–N5 is similar to the active site of cytochrome P450 and has good oxidase-like activity 21


2.1. Metal nanozymes

Metal nanomaterials such as gold (Au), silver (Ag), iridium (Ir), palladium (Pd), platinum (Pt), ruthenium (Ru), copper (Cu) and lanthanides have been reported to have enzyme-like catalytic activity. The main synthesis methods include the general reduction method,22 the co-reduction method,23 and the displacement reaction method.24 Metal nanozymes exhibit POD-like nanozyme activities, facilitating their incorporation into diverse research domains encompassing chemical analysis, biological sensing, antimicrobial strategies, and disease treatment methodologies.25 Furthermore, these metal-based nanomaterials boost remarkable versatility, allowing for facile modification with chemical functionalities such as –NH2, –COOH, and –SH. Their ease of conjugation to biomolecules such as antibodies, enzymes, and oligonucleotides underscores their immense potential in the development of advanced biosensing platforms.26 In addition, advances in the field have seen the emergence of bimetallic and even trimetallic nanozymes, including Pt–Ir nanocubes and Pt–Pd hybrid nanozymes, which have attracted considerable attention and have been extensively reported for their unique properties and potential applications. Compared with single-metal nanozymes, bimetal or trimetal nanozymes tend to exhibit higher stability and catalytic activity due to the synergistic catalytic effect, so they have potential applications in the field of biosensing and therapy.27,28 For example, Fu et al. prepared a porous Au@Pt nanoparticle with ultra-high POD-like activity using a seed-mediated growth method that enables colorimetric detection of the SARS-CoV2 spike protein.29 Li et al. ingeniously developed a Ag/Pd bimetallic nanozyme that exhibits intrinsic POD-like activity,30 efficiently decomposing H2O2 into hydroxyl radicals (˙OH). By harnessing the synergistic interplay of this enzymatic property with its intrinsic photothermal effects, the nanozyme achieves remarkable enhancement in its antitumor capabilities (Fig. 2a–c). Furthermore, based on ultra-small Ir/Ru alloy nanoparticles, Wei et al. designed a multi-enzyme active therapeutic platform with tumor microenvironment-responsive properties and achieved synergies between tumor starvation therapy and catalytic therapy by additionally loading glucose oxidase.31 Besides the synergistic effect of Au NPs and other materials, the fabricated nanozyme exhibits enhanced POD-like activity, making it a potential alternative to peroxidase. Chen et al. designed a hybrid nanozyme, gold nanoparticles/N-doped porous carbon (Au NPs/NPC), which is fabricated via a supramolecular assembly-assisted pyrolysis strategy and engineered as a peroxidase mimic.32 It exhibits broad prospects in biosensing by mimicking enzymes in catalytic fields and thus helping in clinical diagnosis.
image file: d4tb02315c-f2.tif
Fig. 2 (a) Schematic of the synthesis of AgPd@BSA/DOX. (b) TEM images during the synthesis process: (i) Ag NPs, (ii) AgPd NPs, (iii) AgPd@BSA, and (iv) AgPd@BSA/DOX. (c) SEM image and elemental mapping images of AgPd NPs. Adapted with permission.30 Copyright 2020, Elsevier. (d) TEM images of V2O5 nanowires. Adapted with permission.33 Copyright 2010, Wiley. (e) Schematic of the preparation of COF-CNT. Adapted with permission.34 Copyright 2021, Wiley. (f) PCN-222(Fe) structure with POD-like activity. Adapted with permission.35 Copyright 2012, Wiley. (g) Schematic of the synthesis process of AuPt/ZIF-8-rGO. Adapted with permission.36 Copyright 2019, the American Chemical Society. (h) Illustration of the preparation process of FeN3P-SAzyme. (i) TEM image of HAADF-STEM. (j) Corresponding EDS mapping images of FeN3P-SAzyme. Adapted with permission.37 Copyright 2021, Nature Portfolio.

2.2. Metal compound nanozymes

Metal compound nanozymes mainly include metal oxide nanozymes, metal sulfide nanozymes, metal selenide nanozymes, and metal phosphide nanozymes, and the main synthesis methods include solvothermal method, liquid-phase precipitation method, and thermal decomposition method. Metal oxides are the most studied and applied nanozymes. Among them, Fe3O4 is the most desired by researchers. Its activity was first reported in 2007.38,39 Fe3O4 exhibits similar catalytic kinetics to natural HRP in the enzyme-like catalytic process, and kinetic studies show that the reaction process of Fe3O4 follows the ping-pong catalytic mechanism. That is, Fe3O4 first reacts with H2O2 to form ˙OH, and ˙OH combines with Fe3O4 to form a complex, which further interacts with the reducing substrate to catalyze the oxidation of the substrate, and finally, Fe3O4 returns to its initial state. In 2011, the Tremel group reported that V2O5 has POD-like activity, just like Fe3O4. They found that V2O5 works best at a pH of 4.0. When tested in kinetic experiments using Michaelis–Menten principles and ABTS (2,2′-azino-bis(3-ethylbenzothiazoline-6-sulfonic acid)) as the substrate, V2O5 exhibited a similar catalytic behavior to natural vanadium-dependent haloperoxidase. Moreover, the catalytic mechanism of V2O5 follows the ping-pong mechanism (Fig. 2d).

Metal sulfides, including MoS, CuS, and FeS, can also be used as peroxidase mimics. Xu et al. used a solvothermal method to convert garlic-derived organosulfur compounds. Compared with organic sulfides, the antimicrobial efficiency of nano-iron sulfides is more than 500-fold higher when acting on pathogenic and drug-resistant bacteria, and the mechanism suggests that the POD-like activity of nano-iron sulfides can accelerate the release of polysulfides, thereby effectively enhancing antimicrobial activity.40 In addition, bimetallic sulfides exhibit good POD-like activity, Li et al. reported a bimetallic CuCo2S4 nanozyme that show significant POD-like activity under neutral conditions, and its catalytic activity is much higher than that of single-metal CuS and CoS nanoparticles.41

2.3. Carbon-based nanozymes

Carbon-based nanomaterials including carbon nanodots, graphenes, diamonds, carbon nanotubes, fullerenes, and graphene quantum dots have also been found to have enzyme-like activities. The different hybridization states of carbon atoms (sp, sp2 and sp3) lead to great differences in the structure and physicochemical properties of carbon materials. As a non-metallic nanozyme, carbon-based nanozymes have a wide range of applications in the field of enzyme-like catalysis due to their high stability and low cost.42,43

The POD-like activity of graphenes was first reported by Qu et al. in 2010 and used for the detection of glucose.44 In the enzyme-like catalytic process of graphenes and graphene derivatives, oxygen-containing functional groups such as –OH, –COOH, and epoxides, rich in surface or edge sites, play a very important role in catalytic activity. In 2010, Song et al. reported that single-walled carbon nanotubes have potential peroxidase activity to catalyze the oxidation of 3,3′,5,5′-tetramethylbenzidine (TMB) in the presence of H2O2, enabling detection of DNA up to 1 nM.45 Since carbon materials such as carbon quantum dots and carbon nanotubes also have similar oxygen-containing functional groups, this conclusion also applies to the catalytic process of these materials. Similar to graphenes, carbon nanotubes are mainly used to mimic peroxidase during catalysis (Fig. 2e). In the study of amorphous carbon mimetic enzymes, the activity of amorphous carbon nanozymes can be regulated by adjusting the amount and type of heteroatoms incorporated.46,47 One of the most common methods is N-doping. For example, Fan and his team created a nanozyme with multiple enzymatic activities using N-doped porous carbon spheres. The introduction of N renders this nanozyme POD-like activity, which allows it to regulate the intracellular levels of reactive oxygen species (ROS).48

2.4. Metal–organic framework nanozymes

MOFs are a class of porous coordination polymers formed by the coordination of inorganic metal centers and organic ligands. MOFs have the characteristics of unique chemical composition, orderly porous structure, large specific surface area and adjustable structure. MOFs are highly diverse due to the variety of metal nodes and organic ligands they can contain. As a result, they have been applied in various fields including chemical catalysis, gas separation, and drug delivery. In the area of enzyme-like catalysis, MOFs can exhibit different catalytic activities and types of catalysis by carefully selecting the metal centers and organic ligands used in their construction.49,50 In 2012, Feng et al. first reported the POD-like activity of the zirconium-based porphyrin MOFs (PCN-222), the MOF with different open metal sites and a super-large one-dimensional pore structure formed by the coordination of Zr6 clusters with porphyrins, and they introduced different metal sites (Fe, Co, Mn, Ni, Cu, and Zn) into the metal porphyrin ligands and explored their catalytic activity. PCN-222(Fe) was found to exhibit the highest POD-like activity in the solution system, enabling biomimetic oxidation of a range of substrates (Fig. 2f). Zhang et al. prepared ZIF-8-graphene oxide (ZIF-8-GO) by a wet chemical method, further introduced AuCl4 and PtCl42−, and obtained ZIF-8-rGO with AuPt nanoparticles by a reduction method, which showed enhanced peroxidase activity for the detection of H2O2 in human serum samples with ideal performance (Fig. 2g). Compared to metal-based, metal oxide/sulfide-based, and carbon-based nanozymes, MOF-based nanozymes offer unique advantages in mimicking the types of natural enzyme activities. This is due to their adjustable metal nodes and ligands, which allow them to exhibit a wider range of catalytic activities beyond just oxidoreductase activities. This characteristic is of great significance for expanding the potential applications of nanozymes.51 The materials based on MOFs are becoming a powerful tool in the detection of common biomolecules. For example, Lu et al. developed a simple yet effective biosensor based on ZIF-67 loaded with bi-enzymes of glucose and hexokinase (HEX) for effective detection of ATP.52 It paved the path for real-time monitoring of ATP released by cells, which will aid in understanding tumor cell glycolysis and immune responses. Zhao et al. synthesized CuO/C at different temperatures using metal–organic frameworks as sacrificial templates to achieve the detection of glucose in serum/saliva samples.53 These make MOF materials attract widespread attention in the biomedical field.

2.5. Single-atom nanozymes

Single-atomic nanozymes (SAzymes) refer to a new type of catalyst that anchors isolated metal atoms with catalytic activity to solid supports. SAzymes have uniformly dispersed catalytically active sites, which provide a good opportunity to reveal the structure–activity relationship of catalytically active sites and precisely control the geometric and electronic properties of catalytically active sites. It has the advantages of maximizing the utilization efficiency of metal atoms, high catalytic activity and high stability. Most of the SAzymes developed at this stage use carbon-based materials as carriers to anchor metal atoms through nitrogen to form M–N–C SAzymes.54 M–N–C SAzymes have an active site of M–Nx, which is very similar to that of native metalloproteinases, and many researchers have focused on the development of various types of M–N–C-type SAzymes for biomimetic enzyme catalysis.55,56 This class of SAzymes with a definite active structure of M–Nx has inherent advantages in elucidating the catalytic mechanism of nanozymes, overcoming the shortcomings of traditional nanozymes prepared based on trial-and-error strategies.57

By adjusting the types of metal atoms and coordination atoms, the geometric structure and electronic structure of the active sites of SAzymes can also be optimized, so as to achieve controllable regulation of the catalytic activity of enzymes. It has attracted increasingly more attention from researchers. Wang et al. used ZIF-8 as a precursor and introduced a molybdenum metal (Mo) source component into ZIF-8 through an in situ assembly strategy, and Mo-doped ZIF-8 calcined at different temperatures to obtain single-atom Mo nanozymes with different coordination numbers (MoSA–Nx–C, x = 2, 3, or 4). The comparison of enzyme kinetics showed that MoSA–N3–C exhibited the highest POD-like activity.58 To further optimize the coordination environment of metal sites, Ji et al. also used ZIF-8-derived carbon carriers to design Fe SAzymes with a Fe–N3P1 site, which exhibited catalytic activity and enzyme kinetics comparable to that of natural peroxidase (Fig. 2h–j).

3. Strategies for regulating the activity of POD-like nanozymes

3.1. Composition

The composition control of nanozymes opens up the possibility of tuning their catalytic activity. The methods to modulate the activity of POD-like nanozymes by changing the composition can be broadly classified into two categories: (1) coating or growing materials with high POD-like activity on less active surfaces, and (2) doping another element into the lattice of the nanomaterial to form a solid solution. Some noble metal nanoparticles (e.g., iridium, ruthenium, and platinum) have been shown to have excellent POD-like activity.59 By coating or growing highly active noble metal nanoparticles on less active materials, not only their POD-like activity can be enhanced, but also the atomic utilization efficiency of noble metals can be improved. For example, Xia et al. obtained Pa–Ir core–shell cubes by depositing a layer of Ir atoms on the outer layer of Pa nanocubes. The POD-like activity of Pa–Ir nanocubes was enhanced by 20-fold and 400-fold compared to that of Pa cubes and HRP, respectively (Fig. 3a and b). In addition, noble metal nanoparticles have been used to modulate the POD-like activity of carbon-based or metal oxide-based nanomaterials, which not only enhances the POD-like activity of less active carbon or metal oxide-based nanomaterials, but also improves the catalytic cycling stability of metal nanoparticles.60 In addition, a large number of studies have demonstrated synergistic effects between different components in composite nanomaterials. For example, Pt/cubic cerium oxide was prepared by in situ deposition of Pt nanoparticles on the surface of cubic cerium oxide, and the composite Pt/cubic cerium oxide exhibited a higher catalytic activity than a mixture of Pt and CeO2 of the same mass, thanks to the metal-carrier interactions between the Pt particles and cerium oxide.61
image file: d4tb02315c-f3.tif
Fig. 3 (a) Structural and compositional analyses of the Pd–Ir core–shell cubes. (i) TEM image of the sample. (ii) and (iii) HAADF-STEM image of an individual particle. (iv) Line-scan EDX spectra of elemental Pd and Ir. (b) Detection of PSA with Pd–Ir cube-based ELISA and conventional HRP-based ELISA. Adapted with permission.62 Copyright 2015, the American Chemical Society. (c) Schematic of the POD-like activity of N-rGO. (d) Normalized peroxidase-mimicking activities of rGO and N-rGO-5. Adapted with permission.46 Copyright 2018, the American Chemical Society. (e) TEM images of Fe3O4 NPs with different average diameters. Adapted with permission.63 Copyright 2018, Elsevier. (f) Schematic of the one-step preparation of Grp-Au NP hybrids using Graphene flakes, HCOONa and HAuCl4. Adapted with permission.64 Copyright 2017, Elsevier.

Doping of other elements into the lattice of nanomaterials is another effective method to enhance the POD-like activity65,66 and the enhanced POD-like activity is attributed to the alteration of the electronic structure of the original material by the doped elements. Hu et al. doped heteroatomic nitrogen into reduced graphene oxide (rGO) and mesoporous carbon, and the POD-like activity was increased by more than 100 and 60 times, respectively (Fig. 3c and d). In order to further improve the POD-like activity of carbon nanomaterials, researchers have successfully synthesized sulfur/nitrogen co-doped or nitrogen/boron co-doped carbon nanomaterials, which have a stronger POD-like activity than that of nitrogen doping alone.67 In addition to carbon materials, the POD-like activity of metal oxides can be enhanced by elemental doping. For example, transition metals were doped into the lattice of cerium oxide (CeO2) nanoparticles to form MxCe1−xO2−δ solid solutions (M stands for transition metal elements). In these CeO2-based doped materials, the Mn-doped cerium solid solution exhibits significantly enhanced POD-like activity.68

3.2. Size

Generally speaking, size has an important effect on the properties of different nanomaterials. In most cases, nanozymes with smaller particle sizes show more activity in catalytic reactions due to their larger specific surface area. Therefore, the POD-like activity of nanomaterials is closely related to their size.69 Many studies have shown that the smaller the size, the higher the POD-like activity of the material. Gao et al. demonstrated that the catalytic activity of Fe3O4 nanoparticles is size-dependent, and the corresponding POD activity of Fe3O4 nanoparticles decreases as the size of Fe3O4 nanoparticles (30, 150, and 300 nm) increases. Smaller Fe3O4 nanoparticles have higher catalytic activity due to their larger specific surface area, which, in turn, allows them to expose more catalytically active sites. Peng et al. prepared Fe3O4 nanoparticles of three sizes (11, 20 and 150 nm) and similarly found that the POD-like activity of Fe3O4 decreases with the increase in its size (Fig. 3e).63 Ahmed et al. obtained Au nanoparticles with particle sizes of 10, 80, 200 and 500 nm by in situ reduction on the surface of graphenes as a substrate, and it was also found that the smaller the size of Au nanoparticles, the higher the POD-like activity (Fig. 3f).64

Therefore, the size can change the specific surface area as well as other surface properties of nanomaterials, which, in turn, affects the POD-like activity.70 The phenomenon has been similarly found in some noble metal nanozymes. For example, compared to Pt nanozymes synthesized from guanine-rich oligonucleotides (1.8 nm in size), Pt nanozymes synthesized from cytosine-rich oligonucleotides (2.9 nm in size) showed better POD-like activity. This study showed that more metal Pt0 was present on the surface of 2.9 nm Pt nanozymes, resulting in high POD-like activity, whereas more Pt2+ was present on the surface of the 1.8 nm Pt nanozymes.

3.3. Crystal surfaces

In terms of regulating the catalytic activity of enzymes, controlling the exposure of the crystal plane of materials is also a very effective means. Generally speaking, adjusting the morphology of materials is a common strategy in the process of crystal plane manipulation, and different morphologies often expose different crystal planes. These crystals with different atomic arrangements and unsaturated coordination sites exhibit different adsorption energy, reaction energy and activation energy barriers to the same substrate, and have different catalytic activities. Since nanomaterials with different morphologies have different crystal planes, the POD-like activity of nanozymes can be adjusted by regulating the morphology. Liu et al.71 obtained three types of Fe3O4 nanocrystals with different morphologies (clustered spheres, octahedra and triangles) by a hydrothermal method, respectively. This study showed that the clustered spheres with the exposed {311} facets exhibited the highest POD-like activity, the triangular nanocrystals with the {200} facets exhibited relatively weak POD-like activity, and the octahedral Fe3O4 with the {111} facets had the lowest catalytic activity. Many studies have shown that metal nanoparticles with these facets are usually more reactive than those with low-index facets due to the abundance of ligand-unsaturated atomic steps and frameworks, etc., on the high-index facets72 Wu et al. prepared Pd–Pt core–frame nanodendrites by growing dense arrays of Pt branches on a Pd core by wet etching (Fig. 4a). In order to further improve the catalytic activity, the less active Pd nanocore structures therein were selectively etched away to form Pt frame structures capable of exposing more active sites and high index facets (i.e., {110} and {311} facets). In addition, density functional theory calculations confirmed the different POD-like activities of the materials with different crystallographic facets. Li et al. investigated the effect of different crystal facets on the POD-like activity of Au nanocrystals by density-functional theory calculations and found that Au nanozymes with the {111}, {110} and {211} facets followed the same catalytic mechanism and had the same rate-determining step. Based on the activation energy barrier of the rate-determining step, the {111} facet with the highest energy barrier had the lowest catalytic activity, while the Au nanozymes with the {211} facet had the highest catalytic activity (Fig. 4b).
image file: d4tb02315c-f4.tif
Fig. 4 (a) Construction of Pt hollow nanodendrites from Pd nano-cubes. Adapted with permission.73 Copyright 2018, Wiley. (b) Structures of Au(111), Au(110), and Au(211) surfaces. Adapted with permission.74 Copyright 2015, Elsevier. (c) Co3O4 nanozymes modified with different functional groups. Adapted with permission.75 Copyright 2021, the American Chemical Society. (d) Schematic of engineering nanoceria by transition metal ion chelation for enhanced peroxidase mimics. Adapted with permission.28 Copyright 2021, Elsevier. (e) Schematic of the proposed mechanism for Au gel preparation. (f) TEM images of as-prepared OAm-Au. TEM images of aerogels produced in (g) 15 mM and (h) 150 mM (NH4)2S. Adapted with permission.76 Copyright 2019, the American Chemical Society.

3.4. Surface modification

Since the reaction between nanozymes and substrates generally occurs on the surface of the material, properties such as surface charge, exposure of active sites, and substrate binding ability can be modulated by modifying the surface of the material.75 The diverse modifications encompass a broad spectrum, including group modification, ionic modification, and polymer modification, among others. Each of these strategies offers unique advantages in tailoring the properties and functionalities of the material. Through surface modification, nanozymes can be endowed with new types of catalytic activities, improved enzyme-like catalytic specificity, or inhibition of a particular catalytic reaction activity. Therefore, surface modification is a highly flexible and controllable strategy to tune the activity of nanozymes. Liu et al. reported that the catalytic efficiency of DNA-encapsulated iron oxide nanoparticles as peroxidase mimics was approximately 10-fold higher than that of bare NPs.77 The DNA membrane layer not only enhances the binding ability to the TMB amino group via hydrogen bonding but also provides a π–π stacking interaction between nucleobases and the TMB benzene ring, which effectively strengthens the affinity of the Fe3O4 NPs toward TMB. Huo et al. modified Co3O4 nanoplates with amino (NH2–Co3O4), carboxyl (COOH–Co3O4), hydroxyl (OH–Co3O4), and sulfhydryl (SH–Co3O4) groups, respectively, and systematically investigated their catalytic activities (Fig. 4c). Except for the hydroxyl group, all functional groups showed positive enhancement of POD-like activity, with NH2–Co3O4 nanoplates showing the greatest enhancement of POD-like activity. According to Huo et al., the effect of functional groups on the electron transfer capacity of nanozymes is crucial for modulating their catalytic performance. Yue et al. prepared functionalized ceria nanorod catalysts M/CeO2 (M = Fe3+, Co2+, Mn2+, Ni2+, Cu2+, and Zn2+) by chelating ceria nanorods on CeO2 surfaces with metal ions.28 All of these metal-chelated nanoceramics exhibited enhanced POD-like properties, with Mn(II)/CeO2 displaying the optimal catalytic performance (Fig. 4d).

3.5. Surface valence state

The valence of elements on the surface of nanozymes is closely related to the catalytic activity of the enzyme, and the valence of different elements leads to changes in catalytic activity and catalytic type. For instance, Fan et al. achieved the first surface valence control of Au-based nanozymes (Fig. 4e–h). In their system, the catalytic efficiency of substrate oxidation (TMB and H2O2) decreased as the proportion of Au(I) complexes in the Au aerogel decreased.76 By tuning the copper state from Cu0 to Cu2+, Xi et al. found that the POD-like activity of copper/carbon nanozymes was closely related to the Cu state.69 Wang et al. reported that the POD mimetic activity of Ni-based nanozymes was related to the oxidation state of Ni.78 In their study, the catalytic performance of porous LaNiO3 chalcogenides was 58-fold higher than that of NiO and 22-fold higher than that of Ni NPs, suggesting that Ni-based nanomaterials exhibit POD-like properties that depend on the Ni oxidation state. Furthermore, they demonstrated the modulation of catalytic activity by varying the proportion of Ni3+ through the comparison of LaNiO3–H2 and LaNiO3 nanocubes.

3.6. External factors

In addition to the aforementioned regulatory strategies, several external factors, including pH, temperature, biologically relevant molecules, salt ion concentration, and light/acoustic stimulation, can significantly influence the POD-like catalytic activity. For instance, the incorporation of ATP resulted in significant enhancement in the POD-like activity of Au NPs.79,80 In addition, the POD-like activity exhibited pH dependence. Based on this, Sanjay Singh group reported that nanoparticles showed POD-like activity at acidic pH, and the POD-like activity disappeared under neutral conditions. Loading them in pH-responsive microcapsules becomes a vital strategy to avoid exposure to the acidic environment of stomach.81 Under acidic conditions, certain oxides such as Fe3O4 exhibit properties reminiscent of POD-like activity. Conversely, under neutral and alkaline conditions, the activities of catalase and superoxide dismutase predominate. The introduction of external light can manipulate the nano-enzymatic reaction temperature towards the optimal range via the photothermal effect of the material, thereby accelerating the catalytic reaction rate. Alternatively, it can directly modulate the electron population and transfer rate on the surface of materials, such as through localized plasmonic resonance to stimulate hot electron generation or light-induced electron–hole separation to enhance carrier concentration. These strategies, which leverage external factors to optimize the catalytic performance of nanozymes, represent highly efficacious regulatory approaches.

4. Biomedical applications

4.1. Biosensing

POD-like nanozymes play an important role in signal generation and amplification of biosensors. Thus far, various types of biosensing technologies including colorimetry, fluorescence, chemiluminescence, electrochemistry, and surface-enhanced Raman technology have been successfully constructed based on different signal modes catalyzed by nanozymes. The sensing strategies of these sensors can be broadly classified into two categories: the first utilizes nanozymes as signaling units to catalyze the reaction between various chromogenic substrates and targets or their oxidation products, with the color intensity of the chromogenic substrates being proportional to the concentration of the target. The second category involves inhibiting or enhancing the catalytic efficiency of nanozymes through the presence of the target, allowing the analyte content to be indirectly reflected by either the initial color change of the chromogenic substrate or the reaction residue. We summarize the applications of POD-like nanozymes in biosensing in Table 2.
Table 2 POD-like nanozymes for biosensing applications
Materials Sensing method Sensing object Linear range Limit of detection Ref.
MOFs-MIPs Surface-enhanced Raman scattering Trace diazinon 10 nmol L−1–1.0 mmol L−1 3.6 nmol L−1 34
Ag NPs@CS-Hemin/rGO Electrochemistry CEA 20 fg mL−1–200 ng mL−1 6.7 fg mL−1 36
Co-MOFs Electrochemistry miRNA 1.0 × 10−12–1.0 × 10−6 mol L−1 1.2 × 10−13 mol L−1 49
Ir(III)/GO Colorimetry PIB 1.0 × 10−8–3.0 × 10−7 mol L−1 2.81 × 10−9 mol L−1 55
MIL-3DG-75 Fluorescence Xanthine 0–2.0 × 10−4 mol L−1 1.4 × 10−9 mol L−1 56
PdNPs@Fe-MOFs Electrochemistry miR-122 1.0 × 10−17–1.0 × 10−14 mol L−1 3.0 × 10−18 mol L−1 61
Au NCs Colorimetry GSH 2.0 × 10−6–2.5 × 10−5 mol L−1 4.2 × 10−7 mol L−1 81
GK-Pd NPs Colorimetry Glucose 1.0 × 10−5–1.0 × 10−3 mol L−1 6.0 × 10−6 mol L−1 82
MoS2NRs-Au NPs Colorimetry Cholesterol 4.0 × 10−4–1.0 × 10−3 mol L−1 1.5 × 10−5 mol L−1 83
Fe-g-C3N4 Colorimetry Glucose 5.0 × 10−7–1.0 × 10−5 mol L−1 5.0 × 10−7 mol L−1 84
CS-MoSe2 NS Colorimetry Hg2+ 0.025–2.5 μmol L−1 0.0035 μmol L−1 85
MOF-808 Colorimetry AA 5.0 × 10−5–1.03 × 10−3 mol L−1 1.5 × 10−5 mol L−1 86
MIL-53 (Fe) MOF Colorimetry Glucose 2.5 × 10−7–2.0 × 10−5 mol L−1 2.5 × 10−7 mol L−1 87
GO@Au@MnO2 Surface-enhanced Raman scattering Glucose 1–10 μmol L−1 0.114 μmol L−1 88
Au@AuAg NPs Fluorescence H2O2 1.0 × 10−6–4.0 × 10−4 mol L−1 8.65 × 10−7 mol L−1 89
Cu2O@Fe(OH)3 Fluorescence Ochratoxin A 1 ng L−1–10 μg L−1 0.56 ng L−1 90
V2O5 Fluorescence AA 0.01–2.5 μmol L−1 3 nmol L−1 91
N-CDs-MnO2 Fluorescence Acid phosphatase 5–40 U L−1 0.1 U L−1 92
Borophene QDs Fluorescence Oxytetracycline 0–1.0 × 10−4 mol L−1 1.14 × 10−6 mol L−1 93
hemin@CD Fluorescence H2O2 1.7 × 10−7–3.33 × 10−4 mol L−1 1.5 × 10−7 mol L−1 94
N, Fe-CDs Fluorescence UA 8.0 × 10−7–1.33 × 10−4 mol L−1 4.9 × 10−7 mol L−1 95
Cu-MOF Fluorescence Glucose 1.0 × 10−5 – 1.0 × 10−4 mol L−1 4.1 × 10−7 mol L−1 96
Cu–N/C Fluorescence AA 1.0 × 10−5–1.3 × 10−4 mol L−1 7.0 × 10−7 mol L−1 97
BSA@AuNCs Chemiluminescence GSH 50.0–5000.0 nmol L−1 5.2 nmol L−1 98
Pt/Ag Chemiluminescence Glucose 1–500 μmol L−1 0.35[thin space (1/6-em)] μmol L−1 99
Au NPs Chemiluminescence AFP 5–70 ng mL−1 2.5 ng mL−1 100
Co(OH)2 Chemiluminescence Enrofloxacin 0.0001–1000 ng mL−1 0.041 pg mL−1 101
CoMoO4 Chemiluminescence Dopamine 10–500 nmol L−1 1.98 nmol L−1 102
Fe3O4M NPs Chemiluminescence CA125 0–10 μmU mL−1 8.0 μU mL−1 103
HRP-CdTe Chemiluminescence Triphosadenine 50–231 μmol L−1 185 nmol L−1 104
Ni/Co-MOF Chemiluminescence Florfenicol 0.0001–1000 ng mL−1 0.033 pg mL−1 105
Cu-TCPP(Co) MOF Chemiluminescence Thrombin 8.93 × 10−13–5.9 × 10−10 mol L−1 2.18 × 10−13 mol L−1 106
Co SAzyme Chemiluminescence 5-Fluorouracil 0.001–1000 ng mL−1 0.29 pg mL−1 107
Co SADCs Chemiluminescence Carbendazim 1.0 pg mL−1–25 ng mL−1 0.33 pg mL−1 108
Pt/PdCu Electrochemistry SCCA 0.0001–1 ng mL−1, 1–30 ng mL−1 25 fg mL−1 109
Ni–Co–P NSs Electrochemistry DA 0.3–50 μmol L−1 0.016 μmol L−1 110
NiCo-P@NiCo-LDH Electrochemistry Isoprenaline 0.5–2110 μmol L−1 0.17 μmol L−1 111
MoS2/Au Electrochemistry Tau protein 0.1 pg mL−1–100 ng mL−1 33.4 fg mL−1 112
Fe3O4@SiO2 Electrochemistry CEA 0.001–80 ng mL−1 0.0002 ng mL−1 113
Au–Fe3O4 Electrochemistry α-Fetoprotein 0.01–40 ng mL−1 2.3 pg mL−1 114
MWCNT-PDDA-AuPd Electrochemistry Bisphenol A 0.18–18 μmol L−1 60 nmol L−1 115
AuNFs/Fe3O4@ZIF-8-MoS2 Electrochemistry H2O2 5.0 × 10−6–1.2 × 10−1 mol L−1 0.9 × 10−6 mol L−1 116
Fe3C@C/Fe–N–C Electrochemistry H2O2 1.0 × 10−6–6.0 × 10−3 mol L−1 2.6 × 10−7 mol L−1 117
Co-AcNC-3 Electrochemistry Catechol 4.0–300.0 μmol L−1 0.072 μmol L−1 118
Au@Pt NPs Surface-enhanced Raman scattering Dichlorvos 20–2000 μg L−1 20[thin space (1/6-em)] μg L−1 119
Cu2O@Au Surface-enhanced Raman scattering Aflatoxin B1 0.001–100 ng mL−1 0.7 pg mL−1 120
SiO2@Au@Ag Surface-enhanced Raman scattering α-Fetoprotein 0.2–22 ng mL−1 0.1 ng mL−1 121
GO Surface-enhanced Raman scattering PSA 0.5–500 pg mL−1 0.23 pg mL−1 122


4.1.1. Colorimetric biosensing. Colorimetric biosensing is a method of detecting biomolecules using the color changes produced by chemical reactions. Its principle is based on the color change produced after biological molecules react with specific reagents. Among the various sensing methods for nanozymes, colorimetric assays are the most prevalent, and these assays leverage the ability of nanozymes to catalyze the chemical reactions involving chromogenic substrates such as TMB and ABTS, leading to a visible color change that serves as an output signal. These signals can be qualitatively assessed through visual inspection or quantitatively measured using equipment like UV spectrophotometers or microplate readers. Recently, with the ascendancy of point-of-care detection technologies, the outputs of such sensors have been integrated with more affordable portable devices, such as smartphones.

Colorimetric sensors based on POD-like activity represent the most widely reported category. For instance, Zheng et al. successfully synthesized a novel and stable MOF (MOF-808) that exhibits remarkable POD-mimicking catalytic activity under alkaline, neutral, and acidic conditions.86 This material effectively catalyzes the oxidation of TMB in the presence of H2O2, generating distinct color changes. Building upon the excellent catalytic performance of the MOF-808 nanozyme, a novel, simple, and sensitive colorimetric biosensor was successfully developed. Moreover, Huang et al. prepared chitosan-functionalized molybdenum selenide nanosheets and employed them as recognition elements. Intriguingly, Hg2+ ions can activate the nanozyme activity of these nanosheets, thereby facilitating the oxidation of TMB and yielding a colorimetric response. This breakthrough enabled the establishment of a colorimetric biosensor for the detection of Hg2+ (Fig. 5a). Additionally, Zhang et al. utilized synthetic Fe3O4 nanoparticle clusters (Fe3O4 NPC) as nanozymes to amplify colorimetric signals.123 By incorporating aptamers and the glycopeptide antibiotic vancomycin (Van), they developed a colorimetric biosensor capable of recognizing Listeria monocytogenes (L. monocytogenes), a Gram-positive bacterium (Fig. 5b). Furthermore, Xia et al. devised a Ni–Pt nanozyme that boosts the catalytic activity 10[thin space (1/6-em)]000 times higher than that of natural peroxidase.124 When applied to the enzyme-linked immunosorbent assay (ELISA) for the carcinoembryonic antigen, this nanozyme achieved an ultrasensitive detection limit of 1.1 pg mL−1, which is hundreds of times lower than the limit achieved by traditional ELISA immunoassays relying on natural enzymes (376 pg mL−1). This demonstrates that enhancing the catalytic activity of nanozymes is pivotal in improving the detection sensitivity (Fig. 5c).


image file: d4tb02315c-f5.tif
Fig. 5 (a) Illustration of the CS-MoSe2 NS-based versatile colorimetric detection of mercury ions. Adapted with permission.85 Copyright 2019, the American Chemical Society. (b) (i) Schematic of the preparation of Fe3O4 NPC. (ii) Principle of the Fe3O4 NP-based biosensor and (iii) Fe3O4 NPC-catalyzed signal amplification biosensor. Adapted with permission.123 Copyright 2016, Elsevier. (c) High-performance Ni–Pt nanozyme applied in the detection of carcinoembryonic antigen by ELISA. Adapted with permission.124 Copyright 2021, the American Chemical Society.
4.1.2. Fluorescence biosensing. Fluorescence biosensing is a sensing technology that uses the properties of fluorescent substances to detect specific biomolecules. When a fluorescent substance binds to a target biomolecule, its fluorescence properties (such as fluorescence intensity, emission wavelength, or fluorescence lifetime) change. By detecting these changes, the high-sensitivity detection of target biomolecules can be achieved. It is primarily based on target analyte-mediated fluorescence enhancement (“on”) or fluorescence quenching (“off”). In recent years, nanozyme-based fluorescent biosensors have also gained extensive research interest due to their excellent performance in fluorescence signal generation and amplification. They have been widely used in many fields.125 For example, Lin et al. synthesized a MIL-53(Fe) nanozyme with POD-like catalytic activity (Fig. 6a). Under the catalysis of H2O2 by the MIL-53(Fe) nanozyme, a fluorescent product was generated and the fluorescence intensity of the sensing system is related to the concentration of H2O2. Furthermore, a simple, stable, cost-effective and sensitive method for the detection of glucose in normal and diabetic human serum samples is constructed by using the label-free MIL-53(Fe) nanozyme and glucose oxidase for the construction of an enzyme cascade reaction. Fu et al. constructed a ratiometric fluorescence sensing platform based on POD-like nanozymes for the detection of glucose (Fig. 6b).106 First, glucose oxidase catalyzes glucose and provides H2O2 for subsequent reactions, followed by the conversion of the fluorescent substrate o-phenylenediamine (OPD) to the fluorescent species 2,3-diamino phenylhydrazine (DAP) catalyzed by peroxidase activity of copper-doped carbon-based nanozymes with maximum emission at 558 nm. Due to the presence of fluorescence internal filtration, DAP quenched the fluorescence of Mg/N-doped carbon quantum dots (Mg-N-CQDs) at 444 nm, and within a certain range, the proportional fluorescence signal I558/I444 increased linearly with the glucose concentration, from which the concentration of glucose could be quantified. The method was validated in human serum samples with a detection limit of 1.56 μmol L−1.
image file: d4tb02315c-f6.tif
Fig. 6 (a) Principle of fluorescence detection of glucose based on the bifunctional MIL-53(Fe) nanozyme. Adapted with permission.126 Copyright 2018, the Royal Society of Chemistry. (b) Illustration of the ratiometric fluorescence sensing platform for glucose. Adapted with permission.127 Copyright 2022, the American Chemical Society. (c) Schematic exhibition of fabrication process of the immunosensor and label-free CL immunoassay protocol. Adapted with permission.26 Copyright 2016, the American Chemical Society. (d) Schematic of the nanozyme chemiluminescence paper test for SARS-CoV2 S-RBD antigen. Adapted with permission.14 Copyright 2021, Elsevier.
4.1.3. Chemiluminescence biosensing. Chemiluminescence sensing is a kind of sensing technology based on the light signal generated via chemical reactions. This luminescence phenomenon is usually caused by the energy released in a chemical reaction to stimulate electrons to jump to a higher energy level, and when these electrons fall back to a lower energy level, they release photons, resulting in a light signal. These light signals can be captured by sensors and converted into electrical signals for quantitative or qualitative analysis. Chemiluminescence biosensors have emerged as promising analytical tools due to their independence from excitation light sources and complex spectroscopic systems, making detection equipment relatively simple and cost-effective. The advent of nanozymes has sparked interest in advancing chemiluminescence sensors. Luminol, the prevalent chemiluminescent substrate, emits intense blue light upon oxidation, lasting for up to several minutes. Its oxidation necessitates the presence of H2O2, making it an ideal substrate for nanozymes with POD-like activity. For example, Au NPs, Ag NPs, and MIL-type MOFs, endowed with POD-like catalytic activity, have been harnessed as biomimetic components to catalyze the generation of fluorescent signals in the H2O2-luminol system, fostering the development of chemiluminescent biosensors.128 Luo et al. developed a novel MOF-based solid catalyst, Hemin@HKUST-l, by encapsulating heme chloride (Hemin).129

This catalyst exhibits robust POD-like activity, efficiently catalyzing the production of fluorescence signals in the H2O2-luminol system, thereby enabling the creation of a chemiluminescent biosensor with high selectivity and sensitivity.129 As shown in Fig. 6c, Yang et al. constructed an off-chemiluminescence immunosensor for the detection of alpha-fetoprotein antigen. First, CuS nanoparticles were used as peroxidase mimics and carrier materials to be immobilized on glass slides by chitosan crosslinking. Subsequently, streptavidin was modified on the slides to capture biotinylated antibodies, resulting in an interface that exhibited good catalytic activity and triggered intense chemiluminescence in the H2O2-luminol system. Upon capturing the alpha-fetoprotein antigen as it flowed through the sensing interface, the formed antibody–antigen complex hindered the free diffusion of the chemiluminescent substrate to the signaling interface, leading to a decrease in chemiluminescence intensity with the increase in antigen concentration. Furthermore, in view of the rapid spread of epidemic virus globally, Yan et al. established a highly sensitive chemiluminescence rapid test strip platform using a Co–Fe@hemin peroxide nanozyme. This platform enables rapid and sensitive detection of SARS-CoV2 virus within 16 minutes, with a detection limit of 360 TCID50 per mL, comparable to that of the ELISA method (Fig. 6d). This advancement is instrumental in early virus screening and epidemic prevention and control efforts.

4.1.4. Electrochemical biosensing. Electrochemical sensing is a detection technology based on electrochemical principles, which uses the electrical signals generated by substances during chemical reactions to detect and analyze target substances. As a subclass of chemical sensors, electrochemical biosensors include two components: target recognition unit and electrochemical signal transduction unit. Typically, in the presence of the analyte to be measured, the target recognition unit (e.g., enzyme, antigen, etc.) first reacts with the analyte and subsequently transduces that recognition event into a detectable electrical signal. This electrical signal is typically positively or negatively correlated with the analyte concentration, allowing for qualitative or quantitative detection. Electrochemical biosensors have been widely used in pharmaceutical analysis, environmental monitoring, chemical analysis, bioanalysis, and clinical diagnosis due to their simple operation, low cost, good stability, and sensitive response. In order to further improve the analytical sensitivity of electrochemical biosensors, the modification of electrodes with more uniformly dispersed catalysts and electrocatalytic sites is a very effective strategy.130 In recent years, various nanozymes (including Au NPs, Ag NPs, Pt NPs, Au@Pt NPs, AuPd@Au NPs, and Cu-MOF) have good enzyme-like catalytic activity, and have been widely used to catalyze electrochemical signal amplification and construct novel electrochemical biosensors. Based on nickel oxide nanoplate-modified screen-printed electrodes, Khairy et al. developed a new, simple and highly sensitive nanozyme electrochemical biosensor. This electrochemical biosensor has demonstrated good analytical performance (including good stability, high selectivity, and sensitivity) in the determination of OPs (p-thio) in vegetables, water, and human urine samples.131 As shown in Fig. 7a–c, Feng et al. developed an ultra-sensitive sandwich electrochemical immunosensor for the quantitative detection of prostate-specific antigen (PSA) using PtCu@rGO/g-C3N4 nanozyme as a catalytic tag, with a detection limit of 16.6 fg mL−1, demonstrating good stability and selectivity.132
image file: d4tb02315c-f7.tif
Fig. 7 (a) Schematic of the fabrication procedure of Au@Th/GO. (b) and (c) PtCu@rGO/g-C3N4/Ab2 and the sandwich-type electrochemical immunosensor for PSA detection. Adapted with permission.132 Copyright 2017, Elsevier. (d) Schematic of Au NPs@MIL-101@oxidases for efficient enzymatic cascade reactions. Adapted with permission.133 Copyright 2017, the American Chemical Society. (e) Schematic of the triggered self-assembly of plasmonic nanostructures for miR-107 detection. Adapted with permission.134 Copyright 2019, Wiley. (f) Copper oxide nanozymes as efficient multifunctional signal markers of photocurrent quenching of ZnO/Au/AgSbS2 hybrid material. Adapted with permission.135 Copyright 2022, the American Chemical Society. (g) Schematic of the colorimetric sensing strategy based Ag3PO4 nanozymes. Adapted with permission.136 Copyright 2018, Elsevier.
4.1.5. Surface-enhanced Raman scattering biosensing. Surface-enhanced Raman scattering (SERS) biosensing is an advanced analytical method that combines surface-enhanced Raman scattering and biosensing techniques. When the molecules adsorb onto the surface of nanostructures of some specially prepared metals (such as Au, Ag, and Cu), the Raman scattering signal will be greatly enhanced, and this phenomenon is called SERS. Biosensing technology uses biomolecules (such as enzymes, antibodies, and nucleic acids) as recognition elements to detect and analyze target substances through their specific interactions with target molecules. Since Fleischmann et al. first found the abnormally enhanced Raman scattering of pyridine on a rough silver electrode, SERS technology has attracted a great deal of attention in biomedicine. SERS biosensors typically quantify targets by measuring changes in the Raman-labeled molecular signals. The analysis process involves target identification and capture, Raman nanoprobe labeling, and signal reading. In recent years, various SERS-based analytical platforms have been successfully established for the sensitive detection of various target molecules, using nanozymes to enhance SERS performance. For example, Jiang et al. developed a novel and efficient SERS substrate by modifying Ag NPs onto the surface of MIL-101(Fe) with POD-like activity using an in situ manufacturing strategy. The interaction between the MIL-101(Fe) nanozyme and high-density Ag NPs creates numerous Raman hotspots, resulting in high-performance SERS biosensors.137 Hu et al. prepared Au NPs@MIL-101 nanozymes by growing Au NPs in situ within the thermostable and porous MIL-101 framework.133 This nanozyme exhibits POD-like catalytic activity, enabling it to not only catalyze the oxidation of an inactive Raman reporter (recessive malachite green) into a SERS-active form (malachite green) in the presence of H2O2 but also serve as a robust SERS substrate for generating strong Raman signals. Leveraging its exceptional properties, the Au NPs@MIL-101 nanozyme was employed to construct SERS biosensors for monitoring glucose and lactate levels in living tissues (Fig. 7d). Furthermore, Trau et al. designed a nanozyme-based SERS sensing platform for detecting biomarker miRNAs. In the presence of the target miRNA, nucleic acid probe-labeled Au POD-like nanozymes are assembled onto a hollow nano-framework through base complementarity, forming plasmonic nanostructures.134 The subsequent addition of TMB is catalyzed by the POD-like nanozymes, producing oxTMB as a Raman reporter. This strategy was validated in clinical urine samples, achieving a minimum detection limit of 0.7 fmol L−1 (Fig. 7e).
4.1.6. Other biosensing. In addition to the above-mentioned biosensors, other nanozyme-based biosensors such as photoelectrochemistry (PEC) and surface plasmon resonance (SPR) biosensors have been successfully developed. For example, Wu et al. developed a novel dual-mode microfluidic biosensing platform for the detection of ultrasensitive neuron-specific enolases using CuO nanozymes as PEC and fluorescence signal labels.135 The PEC biosensors have a linear range of 0.0001–150 ng mL−1 and a limit of detection of 0.028 pg mL−1, whereas the FL biosensors have a linear range of 0.001–150 ng mL−1 and a limit of detection of 0.25 pg mL−1 (Fig. 7f). In another study, Wu et al. reported an SPR biosensing strategy based on the precisely regulated Ag3PO4 nanozyme signal amplification of ascorbic acid (AA) for the screening of α-glucosidase inhibitors. This SPR biosensor provides ultra-sensitive colorimetric detection of α-glucosidase inhibitors with a detection limit of 10 nM and allows for visual detection of concentrations up to 1 μM. This study provides a new perspective for the construction of biosensors using nanozyme-mediated, highly stable, and efficient cascade signal amplification (Fig. 7g).

4.2. Disease therapy

In recent years, the application research of nanozymes in disease therapy has become increasingly in-depth, and their unique properties and broad application prospects have shown great potential in multiple disease fields. At present, their high POD-like activity has been explored in clinical anti-tumor, antibacterial, antiviral therapy and other fields.
4.2.1. Cancer therapy. Cancer has emerged as a major threat to human life and health. Traditional cancer treatments including surgery, radiotherapy, and chemotherapy often struggle to eradicate cancer, are associated with high side effects, and suffer from poor patient compliance. Consequently, it is important to identify novel and effective therapies that can supersede these traditional approaches. In recent years, the emergence of the field of nanozymes has ushered in new avenues for the development of therapeutic strategies that are responsive to the tumor microenvironment. Table 3 summarises the applications of nanozymes with POD-like activity in cancer therapy.
Table 3 Nanozymes with POD-like activity used for cancer therapy
Materials Application Target Methods Ref.
Ag@GQD Cancer therapy 549, MCF-7 Synergistic therapy of chemotherapy drugs and nanozymes 138
IrOx Cancer therapy A549 human lung adenocarcinoma Mild photothermal-enhanced nanocatalysis therapy 139
SA-Rh Cancer therapy 4T1 tumor-bearing BALB/c mice By synergistic CDT and PTT modalities to Induce apoptosis in cancer cells 140
Ir–N5 SA Cancer therapy Mouse breast cancer cells Generate ROS 141
GQDs Cancer therapy A549 tumor Deep tumor penetration of nanozymes in biomedical application 142
Cu–Ag Cancer therapy Mouse breast tumor cells The production of ˙OH induced by the Cu–Ag nanozyme triggered ferroptosis in tumor cells 143
Bi–NC Cancer therapy 4T1 tumors Catalyze the production of ˙OH and depletes antioxidant GSH 144
FeOx Cancer therapy 4T1 cells Catalyze the decomposition of H2O2 to O2, allowing to alleviate the hypoxic environment of the tumors 145
Enzyme-MOF Cancer therapy Melanoma cells Generate ROS and deplete GSH 146
Ir@MnFe2O4 NP Cancer therapy HeLa cells Catalyze the conversion of H2O2 to cytotoxic ˙OH 147
FeSi Cancer therapy LM3 cells Catalyze H2O2 into ˙OH 148


When nanozymes are applied to antitumor therapy, the overarching core mechanism involves the generation of ROS to eliminate tumor cells.72 Among these, POD-like activity is the most frequently used for tumor therapy. For instance, Zeng et al. developed a brand-new peroxidase-mimetic nanozyme based on biodegradable boron nitride (BON) structure, which can in vitro and in vivo catalyze hydrogen peroxide to efficiently generate hydroxyl radicals.149 Upon intratumoral injection, this material induced significant tumor cell apoptosis, demonstrating therapeutic efficacy comparable to that of anticancer drugs (Fig. 8a–f). Furthermore, Fan et al. developed nitrogen-doped porous carbon spheres for tumour catalytic therapy (Fig. 8g–m). In catalytic property assessment experiments, it was observed that these nitrogen-doped porous carbon spheres exhibited pH-dependent multi-enzyme activity. In acidic environments, they displayed POD-like and oxidase activities, tending to decompose H2O2 to generate toxic ˙OH or convert O2 to ˙O2−. Conversely, under neutral conditions, they exhibited superoxide dismutase and catalase activities, decomposing and converting H2O2 and ˙O2−, among others, into non-toxic H2O and O2. This unique property renders nitrogen-doped porous carbon spheres a highly safe and efficient nano-enzymatic therapeutic agent. Upon accumulation at the tumor site, they generate toxic ROS using locally available O2 and overexpressed H2O2, leading to significant tumor suppression and effectively prolonging survival in mice post-treatment.48 In addition, the combination of nanomaterials and drugs is also promising in cancer treatment. Zhou group designed and constructed MOF-coated MnO2 nanosheets to realize the co-delivery of a survivin inhibiting DNAzyme and doxorubicin for chemo-gene combinatorial treatment of cancer.149 This strategy also paves the way for the future use of nanomaterials in disease treatment.


image file: d4tb02315c-f8.tif
Fig. 8 In vivo evaluations of the BON nanozyme for breast cancer therapy. (a) Schematic of the therapeutic mechanism of BON. (b) Tumor growth curves of 4T1 tumors in BALB/c mice treated under different conditions. (c) Body weights of different groups of mice, (d) Photographs of exfoliated tumors of different groups. (e) Optical photographs of tumor sections stained with H&E. (f) Corresponding fluorescent images of tumor sections subjected to TUNEL staining. Adapted with permission.149 Copyright 2021, Wiley. Tumor therapy based on N-PCNSs using HFn coordination. (g) Schematic of N-PCNS-induced tumor cell destruction via ferritin-mediated specific delivery. (h) TEM image of HFn assembled onto N-PCNSs and TEM image of cancer cells treated with HFn-PCNSs. (i) Quantification of HFn-enhanced internalization of N-PCNSs by flow cytometry analysis. (j) Photographs of tumors after treatment in different groups. (k) Enrichment of the material at the tumor site. (l) Changes in the tumor volume during treatment in different groups. (m) Evaluation of the survival rate in mice. Adapted with permission.48 Copyright 2021, Nature portfolio.

Besides, as treatment techniques continue to be optimized, the development of nanomaterials with synergistic tumor ablation has also gained considerable interest from researchers. For instance, Zhang et al. have developed pH-responsive Ce6-conjugated gold nanorods (Ce6-PEG-AuNRs), which can achieve fluorescence-guided PTT/PDT ablation of tumors.150 Song et al. synthesized a Bi2Se3@HA-doped PPy/ZnPc nanodisc complex using bismuth as a raw material, which could simultaneously exert photothermal and photodynamic effects and had an excellent tumor-growth inhibition ratio (96.4%).151 Liu et al. reported a novel Bi-based nanocomposite (UCNP@NBOF-FePc-PFA) that could not only enable the upconversion luminescence/CT bioimaging of living beings but also be applied for photothermal/photodynamic/radio synergistic tumor ablation.152 These materials have great antitumor effects and potential applications in biomedicine. Luo group prepared silver-modified polyvinyl alcohol pyrrolidone nanoparticles with nano-heterostructures by a two-step method (Bi–Ag@PVP NPs).153 After modification of PVP, the Bi–Ag@PVP NPs possessed low cytotoxicity and caused little damage to normal organs. Notably, under 808 nm laser irradiation, the Bi–Ag@PVP NPs exhibited a better photothermal effect than single bismuth NPs (Bi@PVP NPs) in vivo and in vitro. Importantly, the Bi–Ag@PVP NPs could employ NIR to generate ROS, thereby synergizing PDT and PTT to ablate tumor cells, achieving an enhanced therapeutic effect. Meanwhile, the BiAg@PVP NPs also possessed computed tomography (CT) and photoacoustic (PA) dual-modal imaging properties, enhancing the CT and PA contrast of tumor sites. Interestingly, the Bi–Ag@PVP NPs could be effectively used to kill Escherichia coli, indicating that the Bi–Ag@PVP NPs had an excellent antibacterial function and could be used to avoid bacterial infections during cancer therapy.

4.2.2. Antibacterial therapy. In recent years, excessive use of antibiotics has resulted in a large number of multidrug-resistant bacteria, which has had a serious impact on public health. The existence of multidrug-resistant bacteria significantly limits the drug options available for patients with bacterial infections, and in some cases, leaves them without effective treatment. Nanozyme antibacterial therapy is a new antibacterial therapy after photodynamic antibacterial therapy. The emergence of nanozyme antibacterial therapy provides a new means to resist multidrug-resistant bacteria. Table 4 summarises the applications of nanozymes with POD-like activity in antibacterial therapy. Nanozyme exhibiting POD-like activity can catalyze the production or consumption of ROS. Excessive reactive oxygen radicals can damage cell membranes, proteins and nucleic acids, and kill pathogenic microorganisms, including bacteria and fungi. In recent years, nanozymes have shown promising applications in antimicrobial therapy. Pan et al. designed Au3+-functionalized UIO-67 MOF nanoparticles as mimics of POD and oxidase double enzymes, which catalyze H2O2 and oxygen to generate ˙OH and ˙O2− to kill Escherichia coli and Staphylococcus aureus and promote wound healing (Fig. 9a–e).
Table 4 Nanozymes with POD-like activity used for antibacterial therapy
Materials Application Target Methods Ref.
Fe@HCMS Antibacterial therapy Gram-positive S. aureus and Gram-negative E. coli ROS mediated biocatalytic therapy 154
Au@Cu2−xS Antibacterial therapy Gram-positive Enterococcus faecalis and Gram-negative Fusobacterium Synergistic antimicrobial system 155
CuTAQ Antibacterial therapy S. aureus, E. coli Metal-polyphenol nanomaterials in efficient antimicrobial 156
CuCo2O4 Antibacterial therapy E. coli Generate ˙OH and ˙O2− 157
Fe3O4 Antibacterial therapy S. aureus To catalyze the production of more ˙OH for killing bacteria 158
Cu2O/Pt Antibacterial therapy Gram-positive S. aureus and Gram-negative E. coli Photothermal properties of the Cu2O/Pt nanozyme for enhanced CDT 159
Fe–N–C SAzyme Antibacterial therapy E. coli and S. aureus Spherical mesoporous SAzymes as an antibiotic 160
Fe3O4@MOF@Au NPs Antibacterial therapy Gram-positive S. aureus and Gram-negative E. coli Generate ˙OH 161
Fe-CDs Antibacterial therapy Gram-positive and Gram-negative bacteria Carbon dots as the antibiotics-free nano platforms 162
VOxC NSs Antibacterial therapy E. coli and S. aureus Generate hydroxyl radicals for bacterial killing 163
Cu SASs/NPC Antibacterial therapy E. coli, S. aureus Antibacterial and anti-infective bio-applications of Cu single-atom-containing catalysts 164



image file: d4tb02315c-f9.tif
Fig. 9 (a) Use of Au3+-NMOFs as nanozymes for antibacterial therapy. (b)–(e) Probing the wound-healing efficiencies and cytotoxicity of the Au3+-NMOFs and control systems. Adapted with permission.165 Copyright 2022, Wiley. (f) Comparison of the catalytic activity of different materials. (g) ESR demonstrating the generation of ˙OH. (h) Wound photographs of different treatment groups at different days and (i) corresponding wound diameter size. (j) H&E staining of wounds in different groups at the end of treatment. Adapted with permission.166 Copyright 2022, Elsevier.

At the same time, because metal ions from metals and oxides can be toxic, how to cut down the amount of metals is another important issue. Therefore, SAzymes have gradually been developed for antibacterial therapy in recent years.

Compared with traditional nanozymes, SAzymes have the advantages of high atomic utilization efficiency and high catalytic activity. Fan et al. loaded single-atom Pt on graphitized carbon nitride (SA-Pt/g–C3N4–K) and constructed an efficient single-atom peroxidase mimic.166 Theoretical calculations showed that SA-Pt/g–C3N4–K with a Pt–N–C structure significantly increased the production of ˙OH by reducing the desorption energy of OH* active intermediate at the active site during H2O2 activation. In the mouse wound model infected by Escherichia coli, SA-Pt/g–C3N4–K + H2O2 group effectively promoted wound healing in the past 6 days, and compared with the other groups, SA-Pt/g–C3N4–K + H2O2 group had optimal healing efficiency (Fig. 9f–j).

In addition, nanozyme-catalyzed therapy combined with PTT is a promising approach. Cheng group prepared a pyrolytic-etch-adsorption-pyrolysis (PEAP) process that can be used as a photothermal catalytic antibacterial nanozyme (Cu SASs/NPC). Compared with undoped Cu NPCS, Cu SASs/NPCS have stronger catalytic activity, GSH consumption and photothermal effect.167 The prepared Cu SASs/NPC can be used as peroxidase nanase to catalyze H2O2 to generate ˙OH efficiently, thus achieving obvious bactericidal effects. The photothermal effect of Cu SASs/NPC under laser irradiation can further improve the catalytic action of peroxide-like enzymes, thus producing more ROS and achieving better antibacterial effects in vitro. More importantly, Cu SASs/NPC effectively eradicates internal bacterial infections transmitted by MRSA pathogens at the wound site, resulting in better internal wound healing. The application of SAzymes in the biomedical field is further expanded.

Nanozymes with photothermal and photodynamic properties have also been reported. Yang group designed and synthesized a nanozyme CeO2/FePPOP composite of a porphyrin-based porous organic polymer, FePPOP, and an inorganic material, CeO2, for multimodal antimicrobial activity.168 By subjecting CeO2/FePPOP to a single near-infrared (NIR) light, it is possible to catalyze the production of a significant amount of ROS. This material also has excellent light-to-heat conversion efficiency, thereby inflicting dual damage on bacteria. It provides innovative insights for the creation of nanomaterial antimicrobials suitable for phototherapy interventions.

4.2.3. Antiviral therapy. At present, the strategies to combat viral diseases are prevention (vaccine) and treatment (antiviral drugs and antibodies). However, drug resistance caused by high mutations in the viral genome is a major challenge in the development of antiviral drugs. For this reason, nanozymes use their catalytic and physicochemical properties to directly or indirectly inactivate viruses, providing a new direction for the fight against viral infection. The role of nanozymes in virus therapy primarily involves increasing ROS levels, regulating the immune environment, inhibiting virus replication, and oxidizing viral lipid envelopes and proteins. For the new coronavirus (SARS-CoV2) which is still prevalent at present, Wang et al. Successfully designed a single-atom silver nanozyme (Ag–TiO2) with titanium dioxide as the carrier, and proved that the SAzyme has good potential anti-SARS-CoV2 activity.169

The new SAzyme showed strong adsorption ability to SARS-CoV2 pseudovirus in vitro, and the adsorption rate was as high as 99.65%. The pseudovirus highly expressed SARS-CoV2 spike protein on the surface of HIV-1. Theoretical calculations and experimental evidence indicate that the Ag atom in Ag–TiO2 is strongly bound to cysteine and asparagine, which are the most abundant amino acids on the surface of the spike protein. In vivo, the nanozymes can effectively facilitate the phagocytosis of virus by macrophages. The intrinsic POD-like activity of Ag–TiO2 exhibited a potent antiviral effect within lysosomes. Through the virus elimination model, it was found that Ag–TiO2 could significantly eliminate SARS-CoV2 pseudoviruses and protect host cells from infection (Fig. 10a). In addition, the influenza virus is a very widespread virus. Based on this, Qin et al. reported a strategy using nanozymes against influenza virus.170 Influenza A viruses (IAVs) are characterized by a granular shape with a diameter of 80–120 nm, which is composed of segmented RNA and glycoprotein. In the process of action, the Fe3O4 nanozyme directly contacts IAV particles and collapses their lipid envelopes by elevating lipid peroxidation levels. This, in turn, generates free radicals that damage adjacent proteins such as hemagglutinin, neuraminidase, and matrix protein 1, ultimately impairing the structure and function of the viruses and preventing their infection of host cells (Fig. 10b).


image file: d4tb02315c-f10.tif
Fig. 10 (a) Schematic of Ag–TiO2 against SARS-CoV2 virus. Adapted with permission.169 Copyright 2021, Elsevier. (b) Schematic of the Fe3O4 nanozyme against influenza virus. Adapted with permission.170 Copyright 2019, Spring International Publisher.

5. Conclusion and prospects

The advances in nanotechnology offer huge opportunities for the development of POD-like nanomaterials. Some specific breakthroughs have allowed various strategies to give nanomaterial POD-like activity. With the unique advantages of low cost, high stability and easy preparation, POD-like nanozymes have already become effective substitutes for natural enzymes. This review provided an overview of the classification and activity regulation strategy of POD-like nanozymes. Furthermore, we comprehensively reviewed the latest advances in nanomaterials with POD-like activity in biosensing and disease treatment.

Although breakthroughs have been accomplished in the research of POD-like nanozymes, their development still faces challenges: (1) the catalytic mechanism and key catalytic sites of nanozymes need to be analyzed in depth. At present, most of the research on nanozymes focuses on the optimization of catalytic activity, but less on the analysis of catalytic mechanisms. Although some studies have used advanced mathematical methods such as quantum mechanical computational simulation, molecular dynamics simulation, and machine learning to predict the catalytic pathways of nanozymes, few studies have been able to clarify the key sites of nanozymes that are playing the main catalytic role. Furthermore, little attention has been paid to understanding the dynamic changes in the nanozyme structure during catalysis through experimental verification. Therefore, it is of great importance to employ advanced characterization methods, which can not only help us deeply analyze the catalytic mechanism of nanozymes but also understand the entire process of catalytic reactions from the atomic/molecular level. (2) Catalytic activity and specificity of nanozymes. Generally speaking, the catalytic activity of nanozymes is derived from the active site located on the surface, which means that the more exposed the surface, the higher the catalytic activity. However, excessive exposure will lead to the lack of surface modification of nanozymes and reduce their stability. Additionally, nanozymes are inherently nonspecific. Nonetheless, research has demonstrated that nanozymes can be specific by surface modification of antibodies, nucleic acids, small molecules, and some structural polymers. Nevertheless, most of the surface modifications also weaken the catalytic activity of nanozymes. Thus, the balance between catalytic activity and specificity, as well as the choice of surface modification, remains yet to be studied. (3) The biosafety of nanozymes needs to be further explored. Furthermore, similar to other nanomaterials, toxicological studies of nanozymes including long-term cytotoxicity, administration routes, metabolism processes, excretion mechanisms, biodistribution patterns, pharmacokinetics, pharmacodynamics in large animals, and in vivo experiments, remain largely unproven. Furthermore, the targeting, specificity, in vivo activity and localization of nanozymes are still unclear for catalytic therapeutic effects. Given the complexity of the in vivo microenvironment, the catalytic pathways of nanozymes often become uncontrollable. At the same time, the mechanisms of potential in in vivo biocatalytic reactions and biological interfaces are also difficult to predict, all of which will be factors affecting biosecurity. Therefore, systematic studies and analyses are needed on the in vivo stability, pharmacokinetics (such as uptake, distribution, and metabolism), duration of therapeutic effects, and toxicity (including hemolysis and coagulation analysis) of nanozymes. (4) Strengthen interdisciplinary integration. We believe that the future of POD-like nanomaterials requires interdisciplinary collaboration, including chemistry, physics, biology, and materials science. Furthermore, devices derived from 3D-printed POD-like nanomaterials, especially biosensors, hold the promise of meeting the portability and functional needs of biomedical applications. However, it is acknowledged that numerous existing applications of high-performance POD nanomaterials are still at the proof-of-concept stage, having only been demonstrated in laboratory environments. We believe that there is still a long way to go before commercial applications. The long-term stability and durability of POD-like nanomaterial systems need to be verified prior to practical application. Through the efforts of these different research areas, POD-like nanomaterials can provide more opportunities for personalized disease diagnosis, high-precision bioimaging, biosensing, and drug delivery. In summary, we anticipate that addressing these unresolved issues will significantly contribute to expanding the applications of POD-like nanozymes.

Data availability

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

Conflicts of interest

The authors declare that they have no known competing financial interests or personal relationships that could have appeared to influence the work reported in this paper.

Acknowledgements

This research was funded by the National Natural Science Foundation of China (No. 82102466, 82102466, 81871713 and 82003262), the National Organ Transplantation Special Support Project (No. 2019JYJH08), the Key Research and Development Project of Technology Department of Sichuan Province (No. 2023YFS0108, 2021YFS0160 and 2020YFS0228), and the 1·3·5 Incubation Project of West China Hospital of Sichuan University (No. 19HXFH023).

References

  1. X. Lian, Y. Fang, E. Joseph, Q. Wang, J. Li, S. Banerjee, C. Lollar, X. Wang and H.-C. Zhou, Chem. Soc. Rev., 2017, 46, 3386–3401 RSC.
  2. J. Wu, X. Wang, Q. Wang, Z. Lou, S. Li, Y. Zhu, L. Qin and H. Wei, Chem. Soc. Rev., 2019, 48, 1004–1076 RSC.
  3. L. Gao, J. Zhuang, L. Nie, J. Zhang, Y. Zhang, N. Gu, T. Wang, J. Feng, D. Yang, S. Perrett and X. Yan, Nat. Nanotechnol., 2007, 2, 577–583 CrossRef CAS PubMed.
  4. J. Shah, A. Pandya, P. Goyal, S. K. Misra and S. Singh, ACS Appl. Energy Mater., 2020, 3, 3355–3370 CAS.
  5. D. Yang, Q. Li, S. K. Tammina, Z. Gao and Y. Yang, Sens. Actuators, B, 2020, 319, 128273 CrossRef CAS.
  6. X. Wu, T. Chen, Y. Chen and G. Yang, J. Mater. Chem. B, 2020, 8, 2650–2659 RSC.
  7. R. Das, A. Dhiman, A. Kapil, V. Bansal and T. K. Sharma, Anal. Bioanal. Chem., 2019, 411, 1229–1238 CrossRef CAS PubMed.
  8. W. He, X. Han, H. Jia, J. Cai, Y. Zhou and Z. Zheng, Sci. Rep., 2017, 7, 7 CrossRef PubMed.
  9. U. Jain, S. Gupta and N. Chauhan, Int. J. Biol. Macromol., 2017, 105, 549–555 CrossRef CAS PubMed.
  10. J. Yang, J. Y. Lee and H. P. Too, J. Phys. Chem. B, 2005, 109, 19208 CrossRef CAS PubMed.
  11. C. Ge, G. Fang, X. Shen, Y. Chong, W. G. Wamer, X. Gao, Z. Chai, C. Chen and J.-J. Yin, ACS Nano, 2016, 10, 10436–10445 CrossRef CAS PubMed.
  12. X. Yuan, S. Cheng, L. Chen, Z. Cheng, J. Liu, H. Zhang, J. Yang and Y. Li, Talanta, 2023, 258, 124407 CrossRef CAS PubMed.
  13. X. He, J. Li, R. Li, D. Zhao, L. Zhang, X. Ji, X. Fan, J. Chen, Y. Wang, Y. Luo, D. Zheng, L. Xie, S. Sun, Z. Cai, Q. Liu, K. Ma and X. Sun, Inorg. Chem., 2022, 62, 25–29 CrossRef PubMed.
  14. N. N. Nghia, B. T. Huy and Y.-I. Lee, Microchim. Acta, 2018, 186, 1 Search PubMed.
  15. R. Li, X. Qiao, H. Ma, H. Li, C. Li and L. Jin, J. Mater. Chem. B, 2022, 10, 3311–3319 RSC.
  16. M. A. Wahab, S. M. A. Hossain, M. K. Masud, H. Park, A. Ashok, M. Mustapić, M. Kim, D. Patel, M. Shahbazi, M. S. A. Hossain, Y. Yamauchi and Y. V. Kaneti, Sens. Actuators, B, 2022, 366, 131980 CrossRef CAS.
  17. Z. Lou, S. Zhao, Q. Wang and H. Wei, Anal. Chem., 2019, 91, 15267–15274 CrossRef CAS PubMed.
  18. M. Li, J. Chen, W. Wu, Y. Fang and S. Dong, J. Am. Chem. Soc., 2020, 142, 15569–15574 CrossRef CAS PubMed.
  19. T. Zhang, Y. Xing, Y. Song, Y. Gu, X. Yan, N. Lu, H. Liu, Z. Xu, H. Xu, Z. Zhang and M. Yang, Anal. Chem., 2019, 91, 10589–10595 CrossRef CAS PubMed.
  20. M. Wang, X. Zhou, Y. Li, Y. Dong, J. Meng, S. Zhang, L. Xia, Z. He, L. Ren, Z. Chen and X. Zhang, Bioact. Mater., 2022, 17, 289–299 CAS.
  21. L. Huang, J. Chen, L. Gan, J. Wang and S. Dong, Sci. Adv., 2019, 5, eaav5490 CrossRef CAS PubMed.
  22. R. Das, A. Dhiman, A. Kapil, V. Bansal and T. K. Sharma, Anal. Bioanal. Chem., 2019, 411, 1229–1238 CrossRef CAS PubMed.
  23. S. Cai, C. Qi, Y. Li, Q. Han, R. Yang and C. Wang, J. Mater. Chem. B, 2016, 4, 1869–1877 RSC.
  24. Q. Wang, L. Zhang, C. Shang, Z. Zhang and S. Dong, Chem. Commun., 2016, 52, 5410–5413 RSC.
  25. R. Pan, G. Li, S. Liu, X. Zhang, J. Liu, Z. Su and Y. Wu, Trends Anal. Chem., 2021, 145, 1164642 CrossRef.
  26. Z. Yang, Y. Cao, J. Li, M. Lu, Z. Jiang and X. Hu, ACS Appl. Mater. Interfaces, 2016, 8, 12031–12038 CrossRef CAS PubMed.
  27. C. Huang, F. Zhang, Q. Wang, Y. Lin and J. Huang, Anal. Methods, 2020, 12, 1988–1994 RSC.
  28. Y. Yue, H. Wei, J. Guo and Y. Yang, Colloids Surf., A, 2021, 610, 127515 CrossRef.
  29. Z. Fu, W. Zeng, S. Cai, H. Li, J. Ding, C. Wang, Y. Chen, N. Han and R. Yang, J. Colloid Interface Sci., 2021, 604, 113–121 CrossRef CAS PubMed.
  30. L. Li, H. Liu, J. Bian, X. Zhang, Y. Fu, Z. Li, S. Wei, Z. Xu, X. Liu, Z. Liu, D. Wang and D. Gao, Chem. Eng. J., 2020, 397, 125438 CrossRef CAS.
  31. C. Wei, Y. Liu, X. Zhu, X. Chen, Y. Zhou, G. Yuan, Y. Gong and J. Liu, Biomaterials, 2020, 238, 119848 CrossRef CAS PubMed.
  32. Z. Chen, T. Zhang, Y. Liu, X. Zhang, L. Chen, Z. Zhang and N. Lu, ACS Appl. Nano Mater., 2024, 7, 3645–3655 CrossRef CAS.
  33. R. André, F. Natálio, M. Humanes, J. Leppin, K. Heinze, R. Wever, H. C. Schröder, W. E. G. Müller and W. Tremel, Adv. Funct. Mater., 2010, 21, 501–509 CrossRef.
  34. S. Yao, X. Zhao, X. Wang, T. Huang, Y. Ding, J. Zhang, Z. Zhang, Z. L. Wang and L. Li, Adv. Mater., 2022, 34, 2109568 CrossRef CAS PubMed.
  35. D. Feng, Z. Y. Gu, J. R. Li, H. L. Jiang, Z. Wei and H. C. Zhou, Angew. Chem., Int. Ed., 2012, 51, 10307–10310 CrossRef CAS PubMed.
  36. T. Zhang, Y. Xing, Y. Song, Y. Gu, X. Yan, N. Lu, H. Liu, Z. Xu, H. Xu, Z. Zhang and M. Yang, Anal. Chem., 2019, 91, 10589–10595 CrossRef CAS PubMed.
  37. S. Ji, B. Jiang, H. Hao, Y. Chen, J. Dong, Y. Mao, Z. Zhang, R. Gao, W. Chen, R. Zhang, Q. Liang, H. Li, S. Liu, Y. Wang, Q. Zhang, L. Gu, D. Duan, M. Liang, D. Wang, X. Yan and Y. Li, Nat. Catal., 2021, 4, 407–417 CrossRef CAS.
  38. N. Alizadeh and A. Salimi, J. Nanobiotechnol., 2021, 19, 26 CrossRef CAS PubMed.
  39. D. Liu, C. Ju, C. Han, R. Shi, X. Chen, D. Duan, J. Yan and X. Yan, Biosens. Bioelectron., 2021, 173, 112817 CrossRef CAS PubMed.
  40. Z. Xu, Z. Qiu, Q. Liu, Y. Huang, D. Li, X. Shen, K. Fan, J. Xi, Y. Gu, Y. Tang, J. Jiang, J. Xu, J. He, X. Gao, Y. Liu, H. Koo, X. Yan and L. Gao, Nat. Commun., 2018, 9, 3713 CrossRef CAS PubMed.
  41. D. Li, Q. Guo, L. Ding, W. Zhang, L. Cheng, Y. Wang, Z. Xu, H. Wang and L. Gao, ChemBioChem, 2020, 21, 2620–2627 CrossRef CAS PubMed.
  42. H. Ding, B. Hu, B. Zhang, H. Zhang, X. Yan, G. Nie and M. Liang, Nano Res., 2020, 14, 570–583 CrossRef.
  43. W. Ma, Y. Xue, S. Guo, Y. Jiang, F. Wu, P. Yu and L. Mao, Chem. Commun., 2020, 56, 5115–5118 RSC.
  44. Y. Song, K. Qu, C. Zhao, J. Ren and X. Qu, Adv. Mater., 2010, 22, 2206–2210 CrossRef CAS PubMed.
  45. Y. Song, X. Wang, C. Zhao, K. Qu, J. Ren and X. Qu, Chem. – Eur. J., 2010, 16, 3617–3621 CrossRef CAS PubMed.
  46. Y. Hu, X. J. Gao, Y. Zhu, F. Muhammad, S. Tan, W. Cao, S. Lin, Z. Jin, X. Gao and H. Wei, Chem. Mater., 2018, 30, 6431–6439 CrossRef CAS.
  47. Z. Lou, S. Zhao, Q. Wang and H. Wei, Anal. Chem., 2019, 91, 15267–15274 CrossRef CAS PubMed.
  48. K. Fan, J. Xi, L. Fan, P. Wang, C. Zhu, Y. Tang, X. Xu, M. Liang, B. Jiang, X. Yan and L. Gao, Nat. Commun., 2018, 9, 1440 CrossRef PubMed.
  49. Q. Wang and D. Astruc, Chem. Rev., 2019, 120, 1438–1511 CrossRef PubMed.
  50. D. Ni, J. Lin, N. Zhang, S. Li, Y. Xue, Z. Wang, Q. Liu, K. Liu, H. Zhang, Y. Zhao, C. Chen and Y. Liu, Wiley Nanomed. Nanobiotechnol., 2022, 14, 1773 CrossRef PubMed.
  51. M. Wang, X. Zhou, Y. Li, Y. Dong, J. Meng, S. Zhang, L. Xia, Z. He, L. Ren, Z. Chen and X. Zhang, Bioact. Mater., 2022, 17, 289–299 CAS.
  52. Y. Lu, J. Li, Y. Liu, L. Zhu, S. Xiao, M. Bai, D. Chen and T. Xie, Microchim. Acta, 2023, 190, 71 CrossRef CAS PubMed.
  53. C. Zhao, X. Tang, J. Zhao, J. Cao, Z. Jiang and J. Qin, J. Nanobiotechnol., 2022, 20, 507 CrossRef PubMed.
  54. B. Jiang and M. Liang, Chin. J. Chem., 2020, 39, 174–180 CrossRef.
  55. Q. Chen, S. Li, Y. Liu, X. Zhang, Y. Tang, H. Chai and Y. Huang, Sens. Actuators, B, 2020, 305, 127511 CrossRef CAS.
  56. B. Xu, H. Wang, W. Wang, L. Gao, S. Li, X. Pan, H. Wang, H. Yang, X. Meng, Q. Wu, L. Zheng, S. Chen, X. Shi, K. Fan, X. Yan and H. Liu, Angew. Chem., Int. Ed., 2019, 58, 4911–4916 CrossRef CAS PubMed.
  57. S. Lin and H. Wei, Sci. China: Life Sci., 2019, 62, 710–712 CrossRef PubMed.
  58. Y. Wang, G. Jia, X. Cui, X. Zhao, Q. Zhang, L. Gu, L. Zheng, L. H. Li, Q. Wu, D. J. Singh, D. Matsumura, T. Tsuji, Y.-T. Cui, J. Zhao and W. Zheng, Chem, 2021, 7, 436–449 CAS.
  59. X. Wang, L. Qin, M. Zhou, Z. Lou and H. Wei, Anal. Chem., 2018, 90, 11696–11702 CrossRef CAS PubMed.
  60. M. N. Ivanova, E. D. Grayfer, E. E. Plotnikova, L. S. Kibis, G. Darabdhara, P. K. Boruah, M. R. Das and V. E. Fedorov, ACS Appl. Mater. Interfaces, 2019, 11, 22102–22112 CrossRef CAS PubMed.
  61. Z. Li, X. Yang, Y. Yang, Y. Tan, Y. He, M. Liu, X. Liu and Q. Yuan, Chem. Eur. J., 2017, 24, 409–415 CrossRef PubMed.
  62. X. Xia, J. Zhang, N. Lu, M. J. Kim, K. Ghale, Y. Xu, E. McKenzie, J. Liu and H. Ye, ACS Nano, 2015, 9, 9994–10004 CrossRef CAS PubMed.
  63. F. F. Peng, Y. Zhang and N. Gu, Chin. Chem. Lett., 2008, 19, 730–733 CrossRef CAS.
  64. S. R. Ahmed, K. Takemeura, T.-C. Li, N. Kitamoto, T. Tanaka, T. Suzuki and E. Y. Park, Biosens. Bioelectron., 2017, 87, 558–565 CrossRef CAS PubMed.
  65. M. S. Kim, S. Cho, S. H. Joo, J. Lee, S. K. Kwak, M. I. Kim and J. Lee, ACS Nano, 2019, 13, 4312–4321 CrossRef CAS PubMed.
  66. Q. Liang, J. Xi, X. J. Gao, R. Zhang, Y. Yang, X. Gao, X. Yan, L. Gao and K. Fan, Nano Today, 2020, 35, 100935 CrossRef CAS.
  67. Y. Zhu, J. Wu, L. Han, X. Wang, W. Li, H. Guo and H. Wei, Anal. Chem., 2020, 92, 7444–7452 CrossRef CAS PubMed.
  68. W. Guo, M. Zhang, Z. Lou, M. Zhou, P. Wang and H. Wei, ChemCatChem, 2018, 11, 737–743 CrossRef.
  69. Z. Xi, W. Gao and X. Xia, ChemBioChem, 2020, 21, 2440–2444 CrossRef CAS PubMed.
  70. Y. Fu, X. Zhao, J. Zhang and W. Li, J. Phys. Chem. C, 2014, 118, 18116–18125 CrossRef CAS.
  71. S. Liu, F. Lu, R. Xing and J. J. Zhu, Chem. – Eur. J., 2010, 17, 620–625 CrossRef PubMed.
  72. J. Yang, H. Yao, Y. Guo, B. Yang and J. Shi, Angew. Chem., Int. Ed., 2022, 61, 202200480 CrossRef PubMed.
  73. C. Ge, R. Wu, Y. Chong, G. Fang, X. Jiang, Y. Pan, C. Chen and J. J. Yin, Adv. Funct. Mater., 2018, 28, 1801484 CrossRef.
  74. J. Li, W. Liu, X. Wu and X. Gao, Biomaterials, 2015, 48, 37–44 CrossRef CAS PubMed.
  75. J. Huo, J. Hao, J. Mu and Y. Wang, ACS Appl. Bio Mater., 2021, 4, 3443–3452 CrossRef CAS PubMed.
  76. X. Fan, B. Cai, R. Du, R. Hübner, M. Georgi, G. Jiang, L. Li, M. Samadi Khoshkhoo, H. Sun and A. Eychmüller, Chem. Mater., 2019, 31, 10094–10099 CrossRef CAS.
  77. B. Liu and J. Liu, Nanoscale, 2015, 7, 13831–13835 RSC.
  78. X. Wang, W. Cao, L. Qin, T. Lin, W. Chen, S. Lin, J. Yao, X. Zhao, M. Zhou, C. Hang and H. Wei, Theranostics, 2017, 7, 2277–2286 CrossRef CAS PubMed.
  79. J. Shah, R. Purohit, R. Singh, A. S. Karakoti and S. Singh, J. Colloid Interface Sci., 2015, 456, 100–107 CrossRef CAS PubMed.
  80. N. V. S. Vallabani, A. S. Karakoti and S. Singh, Colloids Surf., B, 2017, 153, 52–60 CrossRef CAS PubMed.
  81. S. Bhagat and S. Singh, Adv. Funct. Mater., 2024, 34, 2408043 CrossRef CAS.
  82. L. Rastogi, D. Karunasagar, R. B. Sashidhar and A. Giri, Sens. Actuators, B, 2017, 240, 1182–1188 CrossRef CAS.
  83. H. Sun, A. Zhao, N. Gao, K. Li, J. Ren and X. Qu, Angew. Chem., Int. Ed., 2015, 54, 7176–7180 CrossRef CAS PubMed.
  84. J. Tian, Q. Liu, A. M. Asiri, A. H. Qusti, A. O. Al-Youbi and X. Sun, Nanoscale, 2013, 5, 11604–11609 RSC.
  85. L. Huang, Q. Zhu, J. Zhu, L. Luo, S. Pu, W. Zhang, W. Zhu, J. Sun and J. Wang, Inorg. Chem., 2019, 58, 1638–1646 CrossRef CAS PubMed.
  86. H.-Q. Zheng, C.-Y. Liu, X.-Y. Zeng, J. Chen, J. Lü, R.-G. Lin, R. Cao, Z.-J. Lin and J.-W. Su, Inorg. Chem., 2018, 57, 9096–9104 CrossRef CAS PubMed.
  87. M. Zhao, Y. Li, X. Ma, M. Xia and Y. Zhang, Talanta, 2019, 200, 293–299 CrossRef CAS PubMed.
  88. T. Zhang, S. Zhu, J. Wang, Z. Liu, M. Wang, S. Li and Q. Huang, Spectrochim. Acta, Part A, 2023, 291, 122307 CrossRef CAS PubMed.
  89. X. Qu, J. Zou, Y. Shen, B. Zhao, J. Liang, Z. Wang, Y. Zhang and L. Niu, Analyst, 2022, 147, 1808–1814 RSC.
  90. H. Zhu, Y. Cai, A. Qileng, Z. Quan, W. Zeng, K. He and Y. Liu, J. Hazard. Mater., 2021, 411, 125090 CrossRef CAS PubMed.
  91. R. Zhu, W. Huang, X. Ma, Y. Zhang, C. Yue, W. Fang, Y. Hu, J. Wang, J. Dang, H. Zhao and Z. Li, Anal. Chim. Acta, 2019, 1089, 131–143 CrossRef CAS PubMed.
  92. Z. Zhu, X. Lin, L. Wu, C. Zhao, Y. Zheng, A. Liu, L. Lin and X. Lin, Sens. Actuators, B, 2018, 274, 609–615 CrossRef CAS.
  93. D. Gogoi, C. Hazarika, G. Neog, P. Mridha, H. K. Bora, M. R. Das, S. Szunerits and R. Boukherroub, ACS Appl. Mater. Interfaces, 2024, 16, 14645–14660 CrossRef CAS PubMed.
  94. L. Su, Y. Cai, L. Wang, W. Dong, G. Mao, Y. Li, M. Zhao, Y. Ma and H. Zhang, Microchim. Acta, 2020, 187, 132 CrossRef CAS PubMed.
  95. G. Yue, S. Li, W. Liu, F. Ding, P. Zou, X. Wang, Q. Zhao and H. Rao, Sens. Actuators, B, 2019, 287, 408–415 CrossRef CAS.
  96. M.-Y. Shi, M. Xu and Z.-Y. Gu, Anal. Chim. Acta, 2019, 1079, 164–170 CrossRef CAS PubMed.
  97. S. Dong, Y. Dong, T. Jia, S. Liu, J. Liu, D. Yang, F. He, S. Gai, P. Yang and J. Lin, Adv. Mater., 2020, 32, 2002439 CrossRef CAS PubMed.
  98. M. I. Halawa, Q. Xia and B. S. Li, J. Mater. Chem. B, 2021, 9, 8038–8047 RSC.
  99. M. Tian, L. Zhao, Y. Wang, G. Liu and P. Zhang, Anal. Lett., 2022, 56, 643–655 CrossRef.
  100. X. Huang and J. Ren, Anal. Chim. Acta, 2011, 686, 115–120 CrossRef CAS PubMed.
  101. Y. Pei, L. Zeng, C. Wen, K. Wu, A. Deng and J. Li, Microchim. Acta, 2021, 188, 194 CrossRef CAS PubMed.
  102. Z. Dong, S. Xia, A. A. M. A. Alboull, I. M. Mostafa, A. Abdussalam, W. Zhang, S. Han and G. Xu, ACS Appl. Nano Mater., 2024, 7, 2983–2991 CrossRef CAS.
  103. Q. Xu, J. Li, S. Li and H. Pan, J. Solid State Electrochem., 2012, 16, 2891–2898 CrossRef CAS.
  104. Z.-M. Zhou, Y. Yu and Y.-D. Zhao, Analyst, 2012, 137, 4262–4266 RSC.
  105. X. Zeng, H. Liu, K. Wu, A. Deng and J. Li, Analyst, 2022, 147, 1321–1328 RSC.
  106. Y. Lin, Y. Sun, Y. Dai, W. Sun, X. Zhu, H. Liu, R. Han, D. Gao, C. Luo and X. Wang, Talanta, 2020, 207, 120300 CrossRef CAS PubMed.
  107. J. Li, Y. Li, K. Wu, A. Deng and J. Li, Talanta, 2023, 265, 124870 CrossRef CAS PubMed.
  108. S. Luo, J. Gao, J. Xian, H. Ouyang, L. Wang and Z. Fu, Anal. Chem., 2022, 94, 13533–13539 CrossRef CAS PubMed.
  109. Y. Liu, H. Ma, J. Gao, D. Wu, X. Ren, T. Yan, X. Pang and Q. Wei, Biosens. Bioelectron., 2016, 79, 71–78 CrossRef CAS PubMed.
  110. Z.-Y. Wang, Z.-Y. Tsai, H.-W. Chang and Y.-C. Tsai, Micromachines, 2024, 15, 105 CrossRef PubMed.
  111. S. Farokhi, M. Roushani and H. Hosseini, Electrochim. Acta, 2020, 362, 137218 CrossRef CAS.
  112. R. Tang, S. Wang, H. Shao, S. Yang, Q. Liu, X. Chen, Y. Huang, N. Gan and S. Huang, Sens. Actuators, B, 2024, 410, 137514 CrossRef.
  113. T. Feng, X. Qiao, H. Wang, Z. Sun and C. Hong, Biosens. Bioelectron., 2016, 79, 48–54 CrossRef CAS PubMed.
  114. G. K. Parshetti, F.-H. Lin and R.-A. Doong, Sens. Actuators, B, 2013, 186, 34–43 CrossRef CAS.
  115. F. Mo, J. Xie, T. Wu, M. Liu, Y. Zhang and S. Yao, Food Chem., 2019, 292, 253–259 CrossRef CAS PubMed.
  116. J. Lu, Y. Hu, P. Wang, P. Liu, Z. Chen and D. Sun, Sens. Actuators, B, 2020, 311, 127909 CrossRef CAS.
  117. X. Wei, S. Song, W. Song, W. Xu, L. Jiao, X. Luo, N. Wu, H. Yan, X. Wang, W. Gu, L. Zheng and C. Zhu, Anal. Chem., 2021, 93, 5334–5342 CrossRef CAS PubMed.
  118. Y. Niu, K. Kang, B. Wang, L. Wang, C. Li, X. Gao, Z. Zhao and X. Ji, Talanta, 2024, 268, 125349 CrossRef CAS PubMed.
  119. H. Yu, M. Wang, J. Cao, Y. She, Y. Zhu, J. Ye, J. Wang, S. Lao and A. M. Abd El-Aty, Anal. Lett., 2021, 55, 427–437 CrossRef.
  120. P. Chen, S. Li, C. Jiang, Z. Wang and X. Ma, Food Biosci., 2023, 54, 102885 CrossRef CAS.
  121. S. Wang, Y. Qin and Z. Zou, Anal. Lett., 2015, 49, 1209–1220 CrossRef.
  122. L. Yang, S. J. Zhen, Y. F. Li and C. Z. Huang, Nanoscale, 2018, 10, 11942–11947 RSC.
  123. L. Zhang, R. Huang, W. Liu, H. Liu, X. Zhou and D. Xing, Biosens. Bioelectron., 2016, 86, 1–7 CrossRef CAS PubMed.
  124. Z. Xi, K. Wei, Q. Wang, M. J. Kim, S. Sun, V. Fung and X. Xia, J. Am. Chem. Soc., 2021, 143, 2660–2664 CrossRef CAS PubMed.
  125. D. P. Cormode, L. Gao and H. Koo, Trends Biotechnol., 2018, 36, 15–29 CrossRef CAS PubMed.
  126. T. Lin, Y. Qin, Y. Huang, R. Yang, L. Hou, F. Ye and S. Zhao, Chem. Commun., 2018, 54, 1762–1765 RSC.
  127. Q. Fu, X. Zhou, M. Wang and X. Su, Anal. Chim. Acta, 2022, 1216, 339993 CrossRef CAS PubMed.
  128. Y. He, B. Xu, W. Li and H. Yu, J. Agric. Food Chem., 2015, 63, 2930–2934 CrossRef CAS PubMed.
  129. F. Luo, Y. Lin, L. Zheng, X. Lin and Y. Chi, ACS Appl. Mater. Interfaces, 2015, 7, 11322–11329 CrossRef CAS PubMed.
  130. H. Chen, T. Yang, F. Liu and W. Li, Sens. Actuators, B, 2019, 286, 401–407 CrossRef CAS.
  131. M. Khairy, H. A. Ayoub and C. E. Banks, Food Chem., 2018, 255, 104–111 CrossRef CAS PubMed.
  132. J. Feng, Y. Li, M. Li, F. Li, J. Han, Y. Dong, Z. Chen, P. Wang, H. Liu and Q. Wei, Biosens. Bioelectron., 2017, 91, 441–448 CrossRef CAS PubMed.
  133. Y. Hu, H. Cheng, X. Zhao, J. Wu, F. Muhammad, S. Lin, J. He, L. Zhou, C. Zhang, Y. Deng, P. Wang, Z. Zhou, S. Nie and H. Wei, ACS Nano, 2017, 11, 5558–5566 CrossRef CAS PubMed.
  134. J. Li, K. M. Koo, Y. Wang and M. Trau, Small, 2019, 13, 1904689 CrossRef PubMed.
  135. T. Wu, S. Yu, L. Dai, J. Feng, X. Ren, H. Ma, X. Wang, Q. Wei and H. Ju, ACS Sens., 2022, 7, 1732–1739 CrossRef CAS PubMed.
  136. D. Wu, N. Hu, J. Liu, G. Fan, X. Li, J. Sun, C. Dai, Y. Suo, G. Li and Y. Wu, Talanta, 2018, 190, 103–109 CrossRef CAS PubMed.
  137. Z. Jiang, P. Gao, L. Yang, C. Huang and Y. Li, Anal. Chem., 2015, 87, 12177–12182 CrossRef CAS PubMed.
  138. Y. Xuan, Y. Gao, Y. Zhao, W. Zhang, X. Bian, M. Zhang, R. Zhang and S. Zhang, Sens. Actuators, B, 2024, 411, 135692 CrossRef CAS.
  139. W. Zou, L. Wang, J. Hao, L. Jiang, W. Du, T. Ying, X. Cai, H. Ran, J. Wu and Y. Zheng, Composites, Part B, 2022, 234, 109707 CrossRef CAS.
  140. N. Tao, S. Chen, S. Mahdinloo, Q. Zhang, T. Lan, Q. Saiding, S. Chen, Y. Xiong, W. Tao and J. Ouyang, Nano Today, 2024, 57, 102371 CrossRef CAS.
  141. Y. Liu, B. Wang, J. Zhu, X. Xu, B. Zhou and Y. Yang, Adv. Mater., 2023, 35, 2208512 CrossRef CAS PubMed.
  142. X. Liu, Z. Liu, K. Dong, S. Wu, Y. Sang, T. Cui, Y. Zhou, J. Ren and X. Qu, Biomaterials, 2020, 258, 120263 CrossRef CAS PubMed.
  143. C. Cao, N. Yang, Y. Su, Z. Zhang, C. Wang, X. Song, P. Chen, W. Wang and X. Dong, Adv. Mater., 2022, 34, 2203236 CrossRef CAS PubMed.
  144. S. Zhong, Z. Zhang, Y. Zhao, S. Wang, Q. Hu and L. Li, Nanoscale, 2023, 15, 16619–16625 RSC.
  145. C. Qi, W.-H. Wang, J.-F. Zheng, L.-W. Jiang, C. Meng, H. Liu and J.-J. Wang, J. Mater. Sci., 2023, 58, 5773–5787 CrossRef CAS.
  146. W. He, X. Han, H. Jia, J. Cai, Y. Zhou and Z. Zheng, Sci. Rep., 2017, 7, 7 CrossRef PubMed.
  147. J. Shen, T. W. Rees, Z. Zhou, S. Yang, L. Ji and H. Chao, Biomaterials, 2020, 251, 120079 CrossRef CAS PubMed.
  148. S. Zhu, H. Xu, W. Guo, M. Yang, H. Tan, S. Hou, J. Yao, H. Luo, Y. Yao, J. Zhao, Y. Wei, X. Sun and B. Ying, Adv. Healthcare Mater., 2024, 202403002 Search PubMed.
  149. L. Zeng, Y. Han, Z. Chen, K. Jiang, D. Golberg and Q. Weng, Adv. Sci., 2021, 8, 2101184 CrossRef CAS PubMed.
  150. C. Zhang, X. Cheng, M. Chen, J. Sheng, J. Ren, Z. Jiang, J. Cai and Y. Hu, Colloids Surf., B, 2017, 160, 345–354 CrossRef CAS PubMed.
  151. Y. Song, J. Wang, L. Liu, Q. Sun, Q. You, Y. Cheng, Y. Wang, S. Wang, F. Tan and N. Li, Mol. Pharm., 2018, 15, 1941–1953 CrossRef CAS PubMed.
  152. J. Liu, J. Zhang, F. Huang, Y. Deng, B. Li, R. Ouyang, Y. Miao, Y. Sun and Y. Li, Acta Biomater., 2020, 113, 570–583 CrossRef CAS PubMed.
  153. Z. Zhou, J. Xie, S. Ma, X. Luo, J. Liu, S. Wang, Y. Chen, J. Yan and F. Luo, ACS Omega, 2021, 6, 10723–10734 CrossRef CAS PubMed.
  154. S. Zhang, Z. Yang, J. Hao, F. Ding, Z. Li and X. Ren, Chem. Eng. J., 2022, 432, 134309 CrossRef CAS.
  155. J. Cao, Q. Sun, A.-G. Shen, B. Fan and J.-M. Hu, Chem. Eng. J., 2021, 422, 130090 CrossRef CAS.
  156. T. Yao, X. Zeng, H. Li, T. Luo, X. Tao and H. Xu, Int. J. Biol. Macromol., 2024, 269, 132115 CrossRef CAS PubMed.
  157. W. Wang, Y. Cui, X. Wei, Y. Zang, X. Chen, L. Cheng and X. Wang, ACS Nano, 2024, 18, 15845–15863 CrossRef CAS PubMed.
  158. J. Guo, W. Wei, Y. Zhao and H. Dai, Regen. Biomater., 2022, 9, rbac041 CrossRef CAS PubMed.
  159. T. Wang, D. Dong, T. Chen, J. Zhu, S. Wang, W. Wen, X. Zhang, H. Tang, J. Liang, S. Wang and H. Xiong, Chem. Eng. J., 2022, 446, 137142 CrossRef.
  160. Y. Feng, J. Qin, Y. Zhou, Q. Yue and J. Wei, J. Colloid Interface Sci., 2022, 606, 826–836 CrossRef PubMed.
  161. Y. Liu, B. Wang, J. Zhu, X. Xu, B. Zhou and Y. Yang, Adv. Mater., 2023, 35, 2208512 CrossRef CAS PubMed.
  162. Y. Liu, B. Xu, M. Lu, S. Li, J. Guo, F. Chen, X. Xiong, Z. Yin, H. Liu and D. Zhou, Bioact. Mater., 2022, 12, 246–256 CAS.
  163. H. Geng, X. Li, X. J. Gao, Y. Cong, Q. Liu, J. Li, Y. Guan, L. Wang and W. He, Nano Today, 2023, 52, 101989 CrossRef CAS.
  164. X. Wang, Q. Shi, Z. Zha, D. Zhu, L. Zheng, L. Shi, X. Wei, L. Lian, K. Wu and L. Cheng, Bioact. Mater., 2021, 6, 4389–4401 CAS.
  165. M. M. Pan, Y. Ouyang, Y. L. Song, L. Q. Si, M. Jiang, X. Yu, L. Xu and I. Willner, Small, 2022, 18, e2200548 CrossRef PubMed.
  166. Y. Fan, X. Gan, H. Zhao, Z. Zeng, W. You and X. Quan, Chem. Eng. J., 2022, 427, 131572 CrossRef CAS.
  167. X. Wang, Q. Shi, Z. Zha, D. Zhu, L. Zheng, L. Shi, X. Wei, L. Lian, K. Wu and L. Cheng, Bioact. Mater., 2021, 6, 4389–4401 CAS.
  168. Y. Liu, F. Yang, P. Liu, C. Lou, L. Zhu and Y. Yang, ACS Appl. Nano Mater., 2024, 7, 23076–23086 CrossRef CAS.
  169. D. Wang, B. Zhang, H. Ding, D. Liu, J. Xiang, X. J. Gao, X. Chen, Z. Li, L. Yang, H. Duan, J. Zheng, Z. Liu, B. Jiang, Y. Liu, N. Xie, H. Zhang, X. Yan, K. Fan and G. Nie, Nano Today, 2021, 40, 101243 CrossRef CAS PubMed.
  170. T. Qin, R. Ma, Y. Yin, X. Miao, S. Chen, K. Fan, J. Xi, Q. Liu, Y. Gu, Y. Yin, J. Hu, X. Liu, D. Peng and L. Gao, Theranostics, 2019, 9, 6920–6935 CrossRef CAS PubMed.

This journal is © The Royal Society of Chemistry 2025
Click here to see how this site uses Cookies. View our privacy policy here.