Next Article in Journal
Multiple-Beam Steering Using Graphene-Based Coding Metasurfaces
Next Article in Special Issue
Biosensors for the Isolation and Detection of Circulating Tumor Cells (CTCs) in Point-of-Care Settings
Previous Article in Journal
Fabrication Methods and Chronic In Vivo Validation of Mechanically Adaptive Microfluidic Intracortical Devices
Previous Article in Special Issue
A Noninvasive Sweat Glucose Biosensor Based on Glucose Oxidase/Multiwalled Carbon Nanotubes/Ferrocene-Polyaniline Film/Cu Electrodes
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Antioxidant Nanozymes: Mechanisms, Activity Manipulation, and Applications

by
Nguyen Thi My Thao
1,†,
Hoang Dang Khoa Do
2,†,
Nguyen Nhat Nam
3,
Nguyen Khoi Song Tran
4,
Thach Thi Dan
5,* and
Kieu The Loan Trinh
6,*
1
School of Medicine and Pharmacy, Tra Vinh University, Tra Vinh City 87000, Vietnam
2
NTT Hi-Tech Institute, Nguyen Tat Thanh University, Ward 13, District 04, Ho Chi Minh City 70000, Vietnam
3
Biotechnology Center, School of Agriculture and Aquaculture, Tra Vinh University, Tra Vinh City 87000, Vietnam
4
College of Korean Medicine, Gachon University, 1342 Seongnam-daero, Sujeong-gu, Seongnam-si 13120, Republic of Korea
5
Tra Vinh University, Tra Vinh City 87000, Vietnam
6
Department of BioNano Technology, Gachon University, 1342 Seongnam-daero, Sujeong-gu, Seongnam-si 13120, Republic of Korea
*
Authors to whom correspondence should be addressed.
These authors contributed equally to this work.
Micromachines 2023, 14(5), 1017; https://doi.org/10.3390/mi14051017
Submission received: 12 April 2023 / Revised: 6 May 2023 / Accepted: 8 May 2023 / Published: 9 May 2023
(This article belongs to the Special Issue Biosensors for Biomedical and Environmental Applications)

Abstract

:
Antioxidant enzymes such as catalase, superoxide dismutase, and glutathione peroxidase play important roles in the inhibition of oxidative-damage-related pathological diseases. However, natural antioxidant enzymes face some limitations, including low stability, high cost, and less flexibility. Recently, antioxidant nanozymes have emerged as promising materials to replace natural antioxidant enzymes for their stability, cost savings, and flexible design. The present review firstly discusses the mechanisms of antioxidant nanozymes, focusing on catalase-, superoxide dismutase-, and glutathione peroxidase-like activities. Then, we summarize the main strategies for the manipulation of antioxidant nanozymes based on their size, morphology, composition, surface modification, and modification with a metal-organic framework. Furthermore, the applications of antioxidant nanozymes in medicine and healthcare are also discussed as potential biological applications. In brief, this review provides useful information for the further development of antioxidant nanozymes, offering opportunities to improve current limitations and expand the application of antioxidant nanozymes.

1. Introduction

During the metabolism processes and normal physiological activities of most aerobic organisms, oxygen undergoes a series of chemical reactions resulting in the formation of toxic by-products, such as superoxide anion radicals (O2•−), hydroxyl radicals, and hydrogen peroxide (H2O2). These by-products are called reactive oxygen species (ROS) and are known as one of the main causes of oxidative damage to some important biological molecules (e.g., DNAs, RNAs, lipids, and proteins) [1,2]. Although ROS are required for numerous essential functions, such as signaling cascade and redox-governing activities, an excess of ROS can cause severe diseases, such as neurodegenerative diseases, digestive diseases, respiratory diseases, and cancer [3,4]. To balance the level of ROS, organisms build an antioxidant system comprised of antioxidant enzymes that can neutralize or degrade the excessive ROS. The natural antioxidant enzymes mainly include catalase (CAT), superoxide dismutase (SOD), and glutathione peroxidase (GPx). While antioxidant enzymes have been widely used to inhibit oxidative damage-related diseases, they have certain limitations, such as high cost, low stability, difficult storage, and poor reusability [5,6,7].
Taking advantage of nanotechnology, nanozymes have emerged as a neoteric approach to improve the limitations of natural antioxidant enzymes. Nanozymes are known as materials with sizes of 100 nm or smaller that display enzyme-like activities. Similar to natural antioxidant enzymes, antioxidant nanozymes catalyze the degradation or decomposition of oxidative agents. In addition, antioxidant nanozymes tend to become popular and promising artificial enzymes owing to their advantages. First, nanozymes can be easily manufactured and do not require expensive storage compared to natural enzymes [8]. Second, the activities of nanozymes can be rationally designed [9]. Third, nanozymes are supposed to be more stable and durable [10]. The mechanisms of antioxidant nanozymes mainly rely on intrinsic CAT, SOD, and GPx mimics to decrease the level of ROS [11,12,13]. The antioxidant activities of nanozymes are highly affected by various factors, including the size and morphology of materials, surface modification, composition, etc. Although dozens of excellent reviewers have discussed the mechanisms and regulation of nanozymes, an overview of antioxidant nanozyme profiles is still lacking. Therefore, in this review, we first discuss the three key mechanisms of antioxidant nanozymes, which are CAT, SOD, and GPx activities (Figure 1). Afterward, we summarize the factors for the manipulation of antioxidant nanozymes. Finally, we review the application of antioxidant nanozymes in several fields, such as medicine, healthcare, diagnostics, and analytics.

2. Mechanisms of Nanozymes for Antioxidant Effects

2.1. Catalase-Like Activity

Catalases (CAT) are natural enzymes containing iron porphyrin at their active sites and are known as antioxidant enzymes because they can catalyze the degradation of hydrogen peroxide to form molecular oxygen and water [14,15,16]. Although hydrogen peroxides play an important role in biological systems, an excess of hydrogen peroxides in the cytoplasm can undergo the Fenton reaction in the presence of transition metal ions and generate hydroxyl radicals, which are strong oxidants [17,18]. Thanks to the feature of CATs that guards living systems against potential oxidative damage, two hydrogen peroxide molecules are degraded to generate two water molecules and one oxygen molecule in a two-step reaction. First, the iron porphyrin group (Fe3+) of CAT reacts with one hydrogen peroxide to form an oxoferryl porphyrin cation radical (Fe4+). The radical then degrades a second hydrogen peroxide to generate water and oxygen molecules, while Fe4+ is reduced back to the Fe3+ state [19,20].
2 H 2 O 2 C a t a l a s e   O 2 + 2 H 2 O
Inspired by the nature of CATs, metal/metal-oxide-based nanomaterials, such as cerium oxide, cobalt oxide, iron oxide, and gold nanoparticles, have been investigated to mimic CAT [21,22,23,24]. Among the several types of reported metal and metal oxide-based nanozymes, cerium-based nanomaterials have been researched in detail regarding the mechanisms of their antioxidant effects. Generally, the antioxidant effects of nanoceria involve a reaction between hydrogen peroxide and Ce4+, in which hydrogen peroxide is degraded to molecular oxygen and Ce4+ is reduced to Ce3+. Interestingly, Ce3+ can be oxidized by another hydrogen peroxide and returned to the Ce4+ state. In this way, the cycle of Ce4+/Ce3+ is created, which is similar to the catalytic reaction of natural CATs (Figure 2) [25,26,27,28]. The cycling between Ce4+ and Ce3+ oxidation states in the presence of hydrogen peroxides can be summarized as the following reaction:
2 Ce 4 + +   H 2 O 2   2 Ce 3 + +   O 2 + 2 H +
2 Ce 3 + +   H 2 O 2 + 2 H +   2 H 2 O   + 2 Ce 4 +
The discovery of ferromagnetic nanoparticles possessing intrinsic peroxidase-like activity was reported in 2007, and these nanoparticles are considered the first inorganic nanoparticle used as an enzyme mimetic for biomedical applications [29]. Later, ferritin–platinum nanoparticles were observed to have CAT-like activity in basic and neutral pH conditions by Nie’s group in 2011 [30]. They found that ferritin–platinum nanoparticles could facilitate the decomposition of hydrogen peroxide to generate oxygen and water, which is similar to natural CAT activity. The decomposition of hydrogen peroxide by iron-based nanozymes at basic and neutral pH levels can be summarized as follows [31]:
Fe 3 + +   H 2 O 2   FeOOH 2 + +   H 2 O
FeOOH 2 +   Fe 2 + +   HO 2
Fe 2 + +   H 2 O 2   Fe 3 + + HO   OH
HO 2   H + +   O 2
OH + HO 2 / O 2   H 2 O   +   O 2
Although the CAT-like mechanisms focusing on the substrates and products in the catalytic process have been described, our knowledge of the dominating surface structures of nanozymes is still limited. Recently, Fan and co-workers reported the CAT-like activity of ferrihydrite (Fe5HO8) and demonstrated that the abundant iron-associated hydroxyl groups on the surface of Fe5HO8 have a critical influence on hydrogen peroxide decomposition [32]. The catalytic process of Fe5HO8 can be described with three main steps. First, hydrogen peroxide is adsorbed on the surface of Fe5HO8. Second, the base-like decomposition of hydrogen peroxide is then catalyzed by Fe5HO8, releasing Fe5HO8 with an H-vacancy. Third, the Fe5HO8 with an H-vacancy degrades another hydrogen peroxide through acid-like decomposition. The three steps of hydrogen peroxide decomposition by Fe5HO8 can be summarized as follows:
Fe 5 HO 8 +   H 2 O 2   H 2 O 2 *
H 2 O 2 * +   H - OFe     HO * +   H 2 O * + H - v a c a n c y
H 2 O 2 +   HO * + H - v a c a n c y     O 2 * +   H 2 O * + H - O F e
* : s p e c i e s   a d s o r b e d   o n   t h e   s u r f a c e   o f   F e 5 H O 8
As a whole, the mechanism of CAT-like nanozymes can be summarized as below:
2 H 2 O 2 C A T l i k e   n a n o z y m e   O 2 + 2 H 2 O

2.2. Superoxide-Dismutase-Like Activity

Superoxide radicals are key elements of ROS generated during the metabolisms of living systems as by-products. Superoxide radicals are intimately involved in oxidative stress and easily converted to other ROS, making it complicated to evaluate their pathogenic pathways [33]. High levels of ROS, especially superoxide radicals, have been implicated in cardiovascular diseases (heart failure, hypertension, diabetes, hypercholesterolemia, and atherosclerosis), Parkinson’s disease, and cancers [34,35]. As a natural antioxidant enzyme, superoxide dismutase (SOD) uses metal as a cofactor and catalyzes the dismutation of superoxide radicals to hydrogen peroxide and oxygen.
2 O 2 + 2 H +   S u p e r o x i d e   d i s m u t a s e   H 2 O 2 +   O 2
SOD is known as a major cellular defense against superoxide radicals, protecting the biological body from oxidative stress. There are three isoforms of SOD that have been found in mammals (CuZnSOD, MnSOD, and ecSOD). Each is produced by distinct genes but catalyzes the same reaction. The mechanism of action of SOD relies on the cycle of the reduction and oxidation states of redox-active transition metals, such as copper and manganese, at the active site of SOD [36,37].
O x i d i z e d   m e t a l - S O D + O 2     R e d u c e d   m e t a l - S O D + O 2
R e d u c e d   m e t a l - S O D + O 2 + 2 H +   O x i d i z e d - S O D + H 2 O 2
Nanozymes possessing SOD-like activity play an important role in the protection against oxidative stress caused by superoxide radicals [38,39]. In 1991, a cluster consisting of 60 carbon atoms was reported as a scavenger of free radicals by Krusic et al. [40]. This research provided early proof that synthesized materials can exhibit antioxidant activity. Since then, various nanozymes have been reported to have SOD-like activity. Most superoxide dismutase nanozymes consist of transition metals (copper, iron, cerium, etc.) and elements such as nitrogen, oxygen, carbon, and sulphur. Among the various nanozymes, SOD-like cerium nanoparticles have been widely developed due to their high biocompatibility and deeply researched mechanism of action [41,42,43]. In 2007, the first cerium oxide nanoparticle exhibiting SOD-like activity was introduced by Korsvil’s group [44]. Similar to nanocerium exhibiting CAT activity, the SOD-like activity of cerium oxide is also involved in the conversion between Ce3+ and Ce4+. The change in the oxidation states of cerium oxide generates oxygen vacancies in the crystal lattice structure by giving out oxygen and electrons. The oxygen vacancies play a key role in the exhibition of SOD-like activity, allowing cerium nanoparticles to uptake or release oxygen. The presence of Ce3+ is a result of oxygen vacancy, and thus the high Ce3+/Ce4+ ratio provides more oxygen vacancies, which increases the SOD-like activity (Figure 3) [45,46]. The dismutation of superoxide radicals by cerium nanoparticles is displayed below:
O 2 +   Ce 4 +     Ce 3 + +   O 2
O 2 +   Ce 3 + + 2 H +   Ce 4 + +   H 2 O 2
Similar to cerium, other transition metals, such as copper (Cu) [47], gold (Au) [48], iron (Fe) [49], manganese (Mn) [50], platinum (Pt) [51], cobalt (Co) [52], and silver (Ag) [53], have been used as the main elements to fabricate nanozymes that exhibit SOD-like activity. Generally, a superoxide radical consists of a Brønsted base with pKb = 9.12 [54], and thus, superoxide radicals can capture protons from H2O to form HO2 and HO. The adsorption of HO2 on the facets of Au, Ag, Pd, and Pt can trigger the conversion of HO2 to O2 and H2O2 [55].
Generally, the mechanism of SOD-like nanozymes can be summarized as below:
2 O 2 + 2 H +   S u p e r o x i d e   d i s m u t a s e   H 2 O 2 +   O 2
Figure 3. The illustration of CAT-like activity for (a) H2O2 decomposition, (b) SOD-like activity for superoxide radical scavenging, and (c) GPx-like activity for H2O2 decomposition. Reprinted with permission from [11]. Copyright (2023) ACS publications. (df) The fabrication and mechanism of selenium-containing pentapeptide (Sec-Arg-Gly-Asp-Cys)-modified gold nanozyme exhibiting GPx activity in the presence of GSH. Reprinted with permission from [56]. Copyright (2020) Royal Society of Chemistry.
Figure 3. The illustration of CAT-like activity for (a) H2O2 decomposition, (b) SOD-like activity for superoxide radical scavenging, and (c) GPx-like activity for H2O2 decomposition. Reprinted with permission from [11]. Copyright (2023) ACS publications. (df) The fabrication and mechanism of selenium-containing pentapeptide (Sec-Arg-Gly-Asp-Cys)-modified gold nanozyme exhibiting GPx activity in the presence of GSH. Reprinted with permission from [56]. Copyright (2020) Royal Society of Chemistry.
Micromachines 14 01017 g003

2.3. Glutathione-Peroxidase-Like Activity

Glutathione peroxidase (GPx) is involved in the termination of the ROS pathway, resulting in a decrease in oxidative stress [57,58,59]. Similar to natural CAT, GPx catalyzes the degradation of H2O2 to H2O and contributes to the cellular antioxidant activity of many organisms. GPx contains seleno-cysteine in its active site, in which selenol (ESeH) degrades H2O2 to H2O through a redox reaction and is oxidized to form selenenic acid (ESeOH). ESeOH then undergoes redox reactions with two reduced glutathiones (GSHs) to recover to ESeH. In this redox cycle, two GSHs are oxidized to form two glutathione disulfides (GSSG) [60,61,62].
Similar to natural GPx, nanozymes exhibit glutathione-peroxidase-like activity using two GSHs to generate redox cycles for H2O2 decomposition [63,64,65,66]. As a typical example, orthorhombic V2O5 nanocrystals catalyze the decomposition of H2O2 in the presence of GSH [64,67]. The glutathione-peroxidase-like activity of V2O5 involves the interaction of hydrogen atoms in H2O2 with the oxygen atoms of the V=O and V−O−V groups, whereas one of the oxygen atoms in H2O2 interacts with vanadium atoms. After reacting with the first H2O2, V=O is converted into a V-peroxide intermediate, which is then converted into V−OH by reacting with GSH. The V−OH reacts with the second H2O2 and GSH to return to the V=O form.
Other nanomaterials exhibit glutathione peroxidase-like activity, such as manganese (II, III) oxide [68], CuxO nanoparticles [69], ultrasmall Cu5.4O nanoparticles [70], citrate functionalized Mn3O4 nanoparticles [71], and Pt/CeO2 nanozymes [72]. All nanozymes mimicking glutathione peroxidase use GSH as one of the substrates to perform their catalytic activity. GSH acts as a reductant and can nucleophilically attack the peroxide bond of peroxide species to form GSSG [67]. Zhang et al. synthesized a selenium-containing pentapeptide (Sec-Arg-Gly-Asp-Cys)-modified gold nanozyme exhibiting GPx activity in the presence of GSH [56]. In this system, gold nanoparticles served as a scaffold that decreased the mobility and constrained the conformation of the peptides. The synthesized nanozyme exhibited GPx activity following an ordered mechanism in which the reaction requires multiple substrates to react sequentially with the enzyme before releasing the first products. Meanwhile, the free peptides follow a ping-pong mechanism in which the first substrate binds to the enzyme and releases the first product before binding the second substrate.

3. Manipulation of Nanozyme Activities for Antioxidant Effects

3.1. Size

It is well-known that the catalytic activity of nanozymes is affected by the size of the materials. For example, cerium oxide nanoparticles with a size of 4.2 nm have been reported to have higher antioxidant activity than the same nanoparticles with a size of 14 nm [73]. Baldim et al. reported that the size of cerium oxide nanoparticles is almost inversely proportional to the SOD-like activity [13]. To explain, a decrease in the particle size results in an increase in the Ce3+/Ce4+ ratio on the surface of particles, which increases the number of defects caused by oxygen vacancy. Oxygen vacancy plays a key role in SOD-like activity, in which a higher number of oxygen vacancies provide more active sites for reactions with substrates [42,46,74,75]. As a result, cerium oxide nanoparticles with a size of ≈5 nm exhibited the highest SOD-like activity, followed by ≈8 nm, ≈23 nm, and ≈28 nm, respectively. Similar to cerium nanoparticles, carbon dot nanozymes with a size of 2 nm and a large specific surface area also provided abundant binding and active sites for catalytic reactions resulting in an increase in SOD-like activity [76]. Similarly, CeVO4 nanorods with sizes of ≈50 nm, 100 nm and 150 nm exhibited SOD-like activity of 4.12 ± 0.19, 3.93 ± 0.20, and 2.57 ± 0.07 ng/µL, respectively [77].
The CAT-like and GPx-like activity of nanozymes is also affected by the size of the materials. At a cerium concentration of 200 µM, the CAT-like activities of cerium nanoparticles with particle sizes of 23 and 28 nm were found to be lower than those with particle sizes of 4.5 and 7.8 nm [13]. Interestingly, Zhang et al. synthesized a single-atom nanozyme of RhN4 with a 20-fold higher affinity for CAT-like activity compared to natural CAT [65]. CoO nanoparticles with sizes between 25 and 30 nm exhibited CAT-like activity of 19.01 U/mL [78]. A single-atom nanozyme, VN4, showed a 7-fold higher GPx-like activity than natural GPx [65]. Gold nanoparticles with an average size of about 20 nm coupled with selenium-containing pentapeptide (Sec-Arg-Gly-Asp-Cys) can enhance GPx activity about 14 times compared to free selenopeptide [56]. Ultrasmall Ru nanoparticles with a size of ≈2 nm increased surface-oxidized Ru atoms, hence exhibiting a higher antioxidant activity than medium-sized (≈3.9 nm) and large-size (≈5.9 nm) Ru nanoparticles [79]. In general, the smaller the nanozyme is, the higher the level of enzyme activity it has.

3.2. Morphology

Previous studies have demonstrated that the morphology of nanomaterials highly affects the catalytic performance. The morphology of nanomaterials is closely related to their specific surface area, pore size, and volume [42,80,81,82]. Most nanozymes with large specific surface areas have more exposed active sites on their surface, resulting in an increase in catalytic activity, including antioxidant activity. For example, Yang et al. synthesized cerium-based nanomaterials with different morphologies, such as nanoclusters, nanoparticles, and nanochains [83]. Among them, cerium nanoclusters showed the highest level of SOD-like activity, and cerium nanochains exhibited higher SOD-like activity than cerium nanoparticles at cerium concentrations ranging from 5 to 100 mg/L. Singh et al. compared the SOD-, CAT-, and GPx-like activity of Mn3O4 nanozymes with different shapes, such as a flower-like morphology, flake-like morphology, hexagonal plates, polyhedrons, and cubes [84]. The flower-like Mn3O4 nanozymes showed the highest SOD-, CAT-, and GPx-like activity, followed by the flake-like Mn3O4 nanozymes. The specific surface area of the flower-like Mn3O4 nanozyme was 97.7 m2/g, higher than that of other morphologies. Ge et al. reported that (111)-faceted Pd octahedrons with low surface energy showed higher antioxidant activity than (100)-faceted nanocubes with high surface energy (Figure 4) [85]. Mn/Zr-co-doped CeO2 tandem nanozymes with hollow mesopores possessed SOD-like and peroxidase-like activity, while their CAT-like activity was inhibited [86]. VO2 nanofibers have been reported that possess higher peroxidase-like activity than VO2 nanosheets and VO2 nanorods [87]. Co3O4 nanozymes with different morphologies exhibited different levels of CAT-like activity following the order Co3O4 nanoplates > Co3O4 nanorods > Co3O4 nanocubes [88]. In general, nanozymes with large specific surface areas, including flower-like shapes, nanoclusters, nanofibers, and mesopores, tend to exhibit higher enzyme activities.

3.3. Composition

The enzyme-like activity of a nanozyme can be economically and efficiently regulated by adjusting the proportion and design of its scomponents [21,48,89]. One of the most effective strategies to enhance the enzymatic activity of a nanozyme is a combination between metals and organic or inorganic materials [90,91,92,93,94]. For example, polyvinylpyrrolidone–platinum–copper nanoparticle clusters (PVP-PtCuNCs) were synthesized by Liu et al. and exhibited 10-fold higher SOD-like activity and 4-fold higher CAT-like activity than PVP-PtNCs [51]. As another example, apoferritin was used as a cage to limit the growth field of and confine Au–Ag nanoparticles in a homogeneous distribution [53]. As a result, the wrapped Au–Ag nanoparticles inside apoferritins possessed 6 × 104 times higher SOD-like activity than that of unwrapped Au–Ag nanoparticles. Generally, Au nanoparticles alone did not exhibit SOD-like activity, even though the concentration of Au nanoparticles was very high (300 µg/mL) [21]. However, Au core/Ce shell-based nanozymes exhibited SOD-like activity at a wide range of pH values (2−11) and temperatures up to 90 °C. Another nanocomposite consisting of Au, Cu, and cysteine (Au-Cu-Cys) showed three times higher IC50 values for SOD-like activity than did the natural SOD enzyme [95]. Matysik et al. conjugated natural SOD on multimetallic nanocomposites, such as ZnO–MnO, ZnO–CuO, and ZnO–MnO–CuO [96]. The SOD-like activities of such nanocomposites were ranked as ZnO–MnO+SOD > ZnO–CuO+SOD > ZnO–MnO–CuO+SOD > SOD enzyme. Zhou et al. fabricated a nanocomposite containing polyvinylpyrrolidone, a Cu single-atom nanozyme, L-cysteine, and MoOx nanoparticles (denoted as MCCP) (Figure 5a) [97]. MCCP achieved impressive CAT-like activity, which was 138-fold stronger than that of typical MnO2.
Curcumin and graphene oxide have become the focal point of cancer studies for their strong antioxidant activity, which correlates with their anticancer property. One of the limitations of curcumin and graphene oxide is they may cause severe damage to not only cancer cells but also normal cells because of their low selectivity. In response, Al-Ani et al. used curcumin and graphene oxide along with biocompatible gold nanoparticle functionalization (CAG) to enhance their selectivity [98]. As a result, the CAG nanocomposite enhanced the selectivity index by ≥30% for SW-948 and ≥90% for HT-29 cancer cells as compared to traditional reduced graphene oxide–gold nanoparticles. The high selectivity of antioxidant nanozymes is critical for clinical applications because it can minimize toxicity.
In terms of CAT-like activities, Zhang et al., compared the enzymatic activities of ten different iron oxide nanomaterials, including 2-line ferrihydrite, 6-line ferrihydrite, akageneite, feroxyhyte, goethite, lepidocrocite, schwertmannite, maghemite, magnetite, and hematite [32]. Among the ten mentioned iron-based nanomaterials, 2-line ferrihydrite possessed the highest CAT-like activity. The CAT-like activities of other iron-based nanomaterials were ranked as 6-line ferrihydrite > feroxyhyte > other iron oxide nanomaterials. The iron-based nanomaterials possessing hydroxyl groups in their stoichiometric composition (2-line ferrihydrite, 6-line ferrihydrite, and feroxyhyte) tended to exhibit higher CAT-like activity than iron-based nanomaterials possessing no hydroxyl groups (hematite, maghemite, and magnetite). It was reported that hydroxyl groups provided the sites for ion exchange to occur on the surface of metal oxides, and thus, the presence of hydroxyl groups could greatly enhance CAT-like activity [99,100]. However, goethite with a low specific surface area still exhibited low CAT-like activity despite its high surface hydroxyl site density. Akageneite possessing a low surface hydroxyl site density and a high specific surface area also exhibited low CAT-like activity.
Most nanoparticles are vulnerable to aggregation, leading to a decrease in their antioxidant activity. Zhou et al., addressed the problem of aggregation using 2D graphdiynes (GDY) to immobilize ultrasmall CeO2 nanoparticles as well as enhance CAT-like activity [101]. As a result, GDY-CeO2 nanocomposites showed a 4.2-fold greater rate constant in H2O2 decomposition than CeO2 nanoparticles. Remarkably, a nanozyme consisting of a V2O5 nanowire, MnO2 nanoparticles, and dopamine was designed and synthesized to mimic multiple oxidant enzymes, such as SOD, CAT, and GPx [102]. In this nanocomposite model, V2O5 nanowires possessed GPx-like activity, while MnO2 nanoparticles possessed SOD- and CAT-like activities. Dopamine was used to assemble the V2O5 and MnO2 nanomaterials. The combination between V2O5, MnO2, and dopamine exhibited significantly higher ROS scavenging capacity than V2O4 and MnO2 nanomaterials. As another example of nanocomposite exhibiting multiple antioxidant enzymes, copper–tannic acid nanozymes were reported to be a thermostable and highly active SOD mimic and CAT mimic [103]. The copper–tannic acid nanozymes with a concentration of 1 µg/mL exhibited almost the same SOD-like activity compared to the natural SOD enzyme with a concentration of 10 U/mL. However, the copper–tannic acid nanozyme could degrade superoxide radicals at a wide range of temperatures (25–80 °C), while natural SOD could not exhibit enzymatic activity at temperatures higher than 40 °C. The CAT-like activity of copper–tannic acid was strongly stable even at 80 °C.
As a whole, the composition of antioxidant nanozymes can highly influence antioxidant performance. Adjusting their proportions and regulating their chemical components, such as breaking the working pH and temperature limits, enabling multi-target function, avoiding self-aggregation, etc., improves the current limitations of antioxidant nanozymes and controls their enzyme-like properties.

3.4. Surface Modification

Surface modification is another factor that highly affects antioxidant activity because most antioxidant reactions occur on the surface of nanomaterials. Numerous studies have reported that surface properties such as thickness, surface charges, and functional groups are important factors in manipulating the activity of nanozymes [104,105,106]. For example, gold nanoparticles capped with N-acetylcysteine (Au-NAC) achieved an ultrasmall size with an average diameter of 2 nm and possessed much higher SOD-like activity than free N-acetylcysteine and gold nanoparticles [107]. A bacteria-like nanozyme was fabricated by decorating ultrasmall CeO2 nanoparticles into a dendritic mesoporous silica-coated Bi2S3 (Ce-Bi@DMSN). Then, the Ce-Bi@DMSN was coated with polyethylene glycol (PEG) [108]. CeO2 endowed the nanozyme with high CAT-like activity. Bi@DMSN acted as a stabilizer to prevent CeO2 from precipitation. The PEG-coated surface endowed the nanocomposite with high hydrophilicity. In addition, it was reported that glycinated and PEGylated ceria nanoparticles exhibited 1.75–1.9 times higher SOD-like activity than uncoated ceria nanoparticles (Figure 5b) [43]. The high physiological stability, biocompatibility, and monodisperisty of nanozymes can be achieved through coating with polyvinylpyrrolidone (PVP) [109], amine-terminated PAMAM dendrimer [110], mercaptoundecanoic acid (11-MUA) [111], PEG [112,113], etc.
Strain effects are a neoteric approach for enhancing the antioxidant activity of nanozymes. Strain effects arise from the torsional angles, distortion of chemical bonds, or mismatched lattices [114,115,116]. Han et al. generated a strained Mn3O4 layer on CeO2 nanoparticles [117]. O 1s XPS spectra results indicated that the ratio of oxygen defects to lattice oxygen in the strained Mn3O4 on CeO2 nanozymes (0.85) was higher than that of CeO2 nanozymes (0.60), leading to the enhancement of antioxidant activities involved in SOD- and CAT-like activity.

3.5. Modification with Metal–Organic Framework

Recently, metal–organic framework (MOF) nanozymes emerged as a huge, remarkable class of functional materials, especially for mimicking antioxidant activity [118,119,120,121]. MOFs are porous materials that are obtained through the self-assembly of organic ligands and metal nodes. One of the most attractive features of MOF nanozymes is their great synthetic tunability, which allows fine chemical and structural control. Depending on their specific purposes, different MOF nanozymes can be rationally designed with properties such as stability, porosity, and particle morphology [122]. It has been reported that a high Ce3+/Ce4+ ratio increases SOD-like activity [45,46]. However, two monovalent Ce-based MOF constructed by 1,3,5-benzenetricarboxylic acid (CeIIIBTC and CeIVBTC) showed unexpected SOD-like activity in which CeIVBTC exhibited higher SOD-like activity than CeIIIBTC. Furthermore, both CeIIIBTC and CeIVBTC showed high selectivity toward superoxide radical elimination [121]. Magnesium gallate MOF with micropores exhibited remarkable antioxidant activity as well as biocompatibility [123]. Copper-based MOF was synthesized to decompose H2O2 and considered a CAT mimic with high catalytic activity [124]. For mimicking glutathione peroxidase, selenium-containing molecules (PhSeBr) were grafted to a Zr(IV)-based UiO-66-NH2 framework. In this mimic of the glutathione peroxidase system, PhSeBr acted as a donator, while a Zr(IV)-based UiO-66-NH2 framework with a high surface area and uniform porosity provided more catalytic active centers, resulting in a high enzyme-like activity [125]. Sang et al. synthesized PZIF67-AT nanoparticles through coordination between a 2-methylimidazole organic linker and cobalt ions [126]. The SOD-mimicking activity of the synthesized PZIF67-AT nanoparticles endowed them with the ability to efficiently produce H2O2 from superoxide radicals. Simultaneously, PZIF67-AT suppresses CAT and GPx activity, preventing the transformation of H2O2 to water. In this way, PZIF67-AT increases H2O2 accumulation, which facilitates Fenton reaction-based chemodynamic therapy. Wang et al. prepared a self-assembled photodynamic therapy nanoagent (OxgeMCC-r single-atom enzyme) consisting of single-atom ruthenium doped into Mn3[Co(CN)6]2 MOF and encapsulated with chlorin e6 [127]. The ruthenium serves as an active center for the conversion of H2O2 into O2, endowing oxgeMCC-r nanozyme with a high level of CAT-like activity. The oxgeMCC-r nanozyme could selectively accumulate in the tumor sites to increase the performance of photodynamic therapy. A cerium-based metal–organic framework (Ce-MOF) was synthesized employing cerium as the active center and 4,4′,4′′-nitrilotribenzoic acid as the linker [128]. With CAT-, SOD-, and peroxidase-like enzymatic activities, Ce-MOF could effectively eliminate fungi such as Rhodotorula glutinis, Aspergillus terreus, Aspergillus flavus, Aspergillus niger, and Candida albicans. In brief, the porous structure and multiple channels of MOFs assist the contact of ROS with the active catalytic sites of MOF nanozymes. The antioxidant activity of MOF nanozymes can be controlled by changing the size of pores, and different enzyme-like activities can be rationally designed by changing the combination of organic ligands and metal nodes. MOF-based antioxidants and their derivatives hold great potential for replacing natural enzymes thanks to their notable features, including their highly specific surface area, the ability to adjust their pore size, and their tunable porosity, as shown in Table 1. Moreover, MOF-based antioxidants demonstrate high thermal stability, flexibility, biocompatibility, and versatile functionality, making them become a feasible strategy for scavenging ROS.

4. Applications

4.1. Applications in Medicine and Healthcare

Antioxidant nanozymes provide a wide variety of benefits to the medical field, and they have been proven for their therapeutic potential, especially for oxidative damage-related pathological diseases. The excess ROS level in our body can be a deleterious effect that leads to various types of cancer by inducing genetic mutations, raising abnormal protein functions, and promoting tumor growth [129,130]. In response, various antioxidant nanozymes have been investigated to balance ROS levels for cancer prevention. Ismail et al. analyzed the potential of platinum nanoparticles to decrease oxidative stress conditions in lung cancer. The data supported that epithelial lung cancer cells did not survive when they were exposed to platinum nanoparticles for 3 h [131]. Cerium oxide nanoparticles also possess anticancer properties by inducing antioxidant activity. Cerium oxide nanozymes with mimetic antioxidant activities, such as CAT-like and SOD-like activities, demonstrate ROS regulation abilities, which can be used to enhance antitumor therapies, including photodynamic and photothermal therapy [28]. It is worth noting that cerium oxide nanoparticles protect cells against oxidative stress at the neutral pH of healthy cells, though they can also induce oxidative stress in the acidic conditions of cancer cells [132,133]. Due to their high flexibility, cerium oxide nanozymes can be used as cytotoxic drugs to kill cancer cells and as protective agents for normal cells.
Age-related diseases, including Alzheimer’s and Parkinson’s disease, are closely associated with ROS accumulation [134,135,136]. According to a recent study, MoS2 nanosheets have been investigated regarding their SOD-like and CAT-like activity to quench NO, O2•– and DPPH free radials [137]. Ceria (CeO2) additionally play an SOD-like role by shuttling between the Ce3+ and Ce4+ states of the mitochondria and certain pathways in Alzheimer’s disease mouse models [34,138,139]. With their ROS-scavenging activity, ceria also take part in therapeutic applications for Parkinson’s disease and depression [140,141].
Moreover, the supporting role of antioxidant-like nanozymes has been demonstrated in research on the circulatory system, inflammation, and metabolic diseases. A study on hydrogel/nanoparticles-based stem cells proved the ROS-suppressing ability of fullerenol by activating the ERK and p38 pathways [142]. Another report suggested the anti-inflammation action of graphene oxide via controlling ROS generation during the M1 macrophage polarization process in the cardiac infraction region [143]. In addition, certain nanozymes, such as copper-based nanoparticles, ceria, and Mn3O4 [70,144,145], have been applied for the therapeutic examination of injuries. Since oxidative stress leads to many metabolic disorders, the ROS-scavenging effect of nanoparticles has been applied to the treatment of various diseases as well. To be specific, polydopamine nanoparticles, fenozyme, and self-cascade MoS2 nanozymes were investigated to improve periodontal disease, cerebral malaria, and hepatic fibrosis, respectively [146,147,148].
Hepatic ischemia-reperfusion injury (IRI) is a pathophysiological process that disturbs liver metabolism and damages other tissues and organs. IRI mainly results from an excess of ROS promoting multiple deleterious processes in DNA, proteins, and lipids, which cause cell death, inflammation, and liver dysfunction and failure. Due to their ROS scavenging ability, hydrophilic carbohydrate-derived nanoparticles (C-NPs) were synthesized and used as nanoantioxidants to prevent hepatic IRI [149]. C-NPs possessed excellent properties, such as nontoxicity, good colloidal stability, ROS scavenging ability, selective delivery to the liver, good circulation lifetime, and slow degradability, all of which are required to maintain a healthy liver status. Liu et al. investigated melanin nanoparticles for antioxidative therapy to protect the brain from ischemic stroke [150]. Melanin nanoparticles exhibiting broad antioxidant activities against O2•−, OH, H2O2, NO, and ONOO can significantly decrease the severity of the brain injury caused by an ischemic stroke. Other studies used antioxidant nanozymes to alleviate acute kidney injury (AKI), another oxidative stress-related pathological disease [107,151,152]. The adopted adjuvant treatment of AKI uses small antioxidants such as amifostine, ι-carnitine, and N-acetylcysteine [153,154,155]. However, these antioxidants have a low ability to target the kidney and large sizes, which decreases their effectiveness for AKI treatment and makes them a challenge to accumulate in renal tissue. Antioxidant nanozymes with diverse surface modifications, sizes, and shapes allow them to specifically target the kidney and easily accumulate in renal tissue. Cerium oxide nanozymes exhibiting multiple antioxidant activities, including SOD- and CAT-like activities, provided efficient protection of renal tissue against oxidative stress caused by an excess of H2O2 in vitro and in vivo (Figure 6) [151,156]. Other metal and metal oxide nanozymes, such as platinum [157], gold [107], iridium [158], copper [70], and manganese [159] nanozymes, showed significant improvements in AKI treatment.

4.2. Applications in Diagnostics and Analytics

In addition to the application of antioxidant nanozymes in medicine, antioxidant nanozymes also provide broad benefits for diagnostic and analytic purposes. With the rapid development of biosensors, antioxidant nanozyzmes have been found to be promising tools in targeted substance detection. Combinations between nanozymes and conventional detection techniques, such as electrochemical-, colorimetric-, and fluorescence-based analysis, offer opportunities to develop new strategies for diagnostics and analytics [160,161,162]. The past decade has witnessed the broad application of nanozyme-based biosensors to detect specific targets. For example, Liang et al. introduced polyethyleneimine-coated nanocubes (PEI NCs) to detect illegal additives [163]. This system provided a pH-switchable dual mode for detecting rosiglitazone, which can exhibit catalase-like and peroxidase-like activity in alkaline and acidic condition, respectively. In alkaline conditions, the catalase-like activity of PEI NCs reduced the fluorescence intensity of 4-chloro-1-naphthol in the presence of rosiglitazone. In acidic conditions, the peroxidase-like activity of PEI NCs triggered a colorimetric reaction that changed the color of 3,3′,5,5′-tetramethylbenzidine (TMB) in the presence of rosiglitazone. The CAT-like and SOD-like activity of antioxidant nanozymes, such as Co3O4 [164], vanadium-based nanozymes [165], and copper nanozymes [166], endows them the ability to detect biomarkers, microbiological pathogens, and toxic compounds in the presence of H2O2 by changing the color of TMB. In general, antioxidant nanozymes catalyze the oxidation of TMB, which changes TMB’s color from colorless to blue in the presence of H2O2. However, the presence of targets interrupts the oxidation of TMB, preventing this color change. This strategy was used to detect several targets, such as glutathione [164], superoxide anion [165], ascorbic acid [166], and amyloid-β peptide [167].
Antioxidant nanozymes mainly contribute to electrochemical analysis by exhibiting SOD-like activity to analyze superoxide radicals, which are highly related to the multi-stage cancerization of healthy tissue, diabetes, Parkinson’s disease, etc. [168,169]. In this strategy, SOD-mimicking nanozymes, such as CeO2 [170], Mnx(PO4)y [171], CeO2–TiO2 [172], and Co3(PO4)2 [173], scavenge superoxide radicals to produce O2 on the electrode of an electrochemical sensor. The O2 accumulation on the electrode increases the current density to a higher level. By changing the electric current obtained at the electrode in the presence of the target, the electrochemical sensor not only detects superoxide radicals but also quantifies them. The integration of SOD-like nanozymes and electrochemical sensors provides fast, precise, sensitive, and selective analytical tools for superoxide radical analysis.

5. Conclusions and Future Perspectives

In this review, we summarized recent nanozymes with antioxidant effects, focusing on the mechanisms and neoteric approaches for activity regulation. Generally, nanozymes exhibit antioxidant activity by mimicking three main natural enzymes: catalase, superoxide dismutase, and glutathione peroxidase. Catalase- and glutathione peroxidase-mimicking nanozymes decompose H2O2, while superoxide dismutase-mimicking nanozymes catalyze thesuperoxide radical decomposition. In addition, the enzymatic activities of nanozymes can be regulated through the modification of their size, morphology, surface, and composition, as well as modification with MOF. The modification of size and morphology mainly manipulates antioxidant activity because size and morphology strongly affect the specific surface area. Similarly, the stability, biocompatibility, selectivity, and multi-target ability of antioxidant nanozymes can be manipulated through the modification of their surface and composition, as well as through MOF. Although antioxidant nanozymes offer many advantages in the improvement of natural antioxidant enzymes, there are still limitations that need to be further addressed. First, numerous nanozymes are poorly biocompatible as compared to natural enzymes because of the use of toxic agents in nanozyme fabrications. Therefore, the green synthesis of antioxidant nanozymes should be investigated to extend the application of antioxidant nanozymes, especially in clinical fields. Second, antioxidant nanozymes can become pro-oxidants, which have the opposite action of antioxidants, causing oxidative stress depending on the internal environment. For example, a nanozyme can act as an antioxidant at a high pH level but can act as a pro-oxidant at a low pH level. Fortunately, this property can be used to specifically target oxidative-stressed sites, which are mainly acidic. On the other hand, antioxidant nanozymes can produce ROS when they meet acidic but healthy body parts, such as the stomach, skin, and large intestine. Therefore, studies on the response of antioxidant nanozymes to different areas of the human body should be thoroughly investigated, thus extending the applications of antioxidant nanozymes.

Author Contributions

Conceptualization, N.T.M.T., H.D.K.D., N.N.N., N.K.S.T., T.T.D. and K.T.L.T.; writing—original draft preparation, N.T.M.T., H.D.K.D., N.N.N. and N.K.S.T.; writing—review and editing, K.T.L.T. and T.T.D.; supervision, K.T.L.T. and T.T.D. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Data Availability Statement

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

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Auten, R.L.; Davis, J.M. Oxygen Toxicity and Reactive Oxygen Species: The Devil Is in the Details. Pediatr. Res. 2009, 66, 121–127. [Google Scholar] [CrossRef] [PubMed]
  2. Covarrubias, L.; Hernández-García, D.; Schnabel, D.; Salas-Vidal, E.; Castro-Obregón, S. Function of Reactive Oxygen Species during Animal Development: Passive or Active? Dev. Biol. 2008, 320, 1–11. [Google Scholar] [CrossRef]
  3. Liu, Z.; Ren, Z.; Zhang, J.; Chuang, C.-C.; Kandaswamy, E.; Zhou, T.; Zuo, L. Role of ROS and Nutritional Antioxidants in Human Diseases. Front. Physiol. 2018, 9, 477. [Google Scholar] [CrossRef] [PubMed]
  4. Yang, S.; Lian, G. ROS and Diseases: Role in Metabolism and Energy Supply. Mol. Cell. Biochem. 2020, 467, 1–12. [Google Scholar] [CrossRef]
  5. Liang, M.; Yan, X. Nanozymes: From New Concepts, Mechanisms, and Standards to Applications. Acc. Chem. Res. 2019, 52, 2190–2200. [Google Scholar] [CrossRef]
  6. Ashrafi, A.M.; Bytesnikova, Z.; Barek, J.; Richtera, L.; Adam, V. A Critical Comparison of Natural Enzymes and Nanozymes in Biosensing and Bioassays. Biosens. Bioelectron. 2021, 192, 113494. [Google Scholar] [CrossRef]
  7. Zhang, R.; Yan, X.; Fan, K. Nanozymes Inspired by Natural Enzymes. Acc. Mater. Res. 2021, 2, 534–547. [Google Scholar] [CrossRef]
  8. Li, A.; Long, L.; Liu, F.; Liu, J.; Wu, X.; Ji, Y. Antigen-Labeled Mesoporous Silica-Coated Au-Core Pt-Shell Nanostructure: A Novel Nanoprobe for Highly Efficient Virus Diagnosis. J. Biol. Eng. 2019, 13, 87. [Google Scholar] [CrossRef]
  9. Liu, J.; Niu, X. Rational Design of Nanozymes Enables Advanced Biochemical Sensing. Chemosensors 2022, 10, 386. [Google Scholar] [CrossRef]
  10. Karpova, E.V.; Shcherbacheva, E.V.; Komkova, M.A.; Eliseev, A.A.; Karyakin, A.A. Core–Shell Nanozymes “Artificial Peroxidase”: Stability with Superior Catalytic Properties. J. Phys. Chem. Lett. 2021, 12, 5547–5551. [Google Scholar] [CrossRef] [PubMed]
  11. Zeng, J.; Ding, C.; Chen, L.; Yang, B.; Li, M.; Wang, X.; Su, F.; Liu, C.; Huang, Y. Multienzyme-Mimicking Au@Cu2O with Complete Antioxidant Capacity for Reactive Oxygen Species Scavenging. ACS Appl. Mater. Interfaces 2023, 15, 378–390. [Google Scholar] [CrossRef]
  12. Feng, N.; Liu, Y.; Dai, X.; Wang, Y.; Guo, Q.; Li, Q. Advanced Applications of Cerium Oxide Based Nanozymes in Cancer. RSC Adv. 2022, 12, 1486–1493. [Google Scholar] [CrossRef] [PubMed]
  13. Baldim, V.; Bedioui, F.; Mignet, N.; Margaill, I.; Berret, J.-F. The Enzyme-like Catalytic Activity of Cerium Oxide Nanoparticles and Its Dependency on Ce3+ Surface Area Concentration. Nanoscale 2018, 10, 6971–6980. [Google Scholar] [CrossRef]
  14. Liu, X.; Kokare, C. Microbial Enzymes of Use in Industry. In Biotechnology of Microbial Enzymes; Elsevier: Amsterdam, The Netherlands, 2017; pp. 267–298. ISBN 978-0-12-803725-6. [Google Scholar]
  15. Haider, M.S.; Jaskani, M.J.; Fang, J. Overproduction of ROS: Underlying Molecular Mechanism of Scavenging and Redox Signaling. In Biocontrol Agents and Secondary Metabolites; Elsevier: Amsterdam, The Netherlands, 2021; pp. 347–382. ISBN 978-0-12-822919-4. [Google Scholar]
  16. Sharma, I.; Ahmad, P. Catalase. In Oxidative Damage to Plants; Elsevier: Amsterdam, The Netherlands, 2014; pp. 131–148. ISBN 978-0-12-799963-0. [Google Scholar]
  17. Zenin, V.; Ivanova, J.; Pugovkina, N.; Shatrova, A.; Aksenov, N.; Tyuryaeva, I.; Kirpichnikova, K.; Kuneev, I.; Zhuravlev, A.; Osyaeva, E.; et al. Resistance to H2O2-Induced Oxidative Stress in Human Cells of Different Phenotypes. Redox Biol. 2022, 50, 102245. [Google Scholar] [CrossRef] [PubMed]
  18. Wijeratne, S.S.K.; Cuppett, S.L.; Schlegel, V. Hydrogen Peroxide Induced Oxidative Stress Damage and Antioxidant Enzyme Response in Caco-2 Human Colon Cells. J. Agric. Food Chem. 2005, 53, 8768–8774. [Google Scholar] [CrossRef] [PubMed]
  19. Alfonso-Prieto, M.; Biarnés, X.; Vidossich, P.; Rovira, C. The Molecular Mechanism of the Catalase Reaction. J. Am. Chem. Soc. 2009, 131, 11751–11761. [Google Scholar] [CrossRef]
  20. Galasso, M.; Gambino, S.; Romanelli, M.G.; Donadelli, M.; Scupoli, M.T. Browsing the Oldest Antioxidant Enzyme: Catalase and Its Multiple Regulation in Cancer. Free Radic. Biol. Med. 2021, 172, 264–272. [Google Scholar] [CrossRef]
  21. Bhagat, S.; Srikanth Vallabani, N.V.; Shutthanandan, V.; Bowden, M.; Karakoti, A.S.; Singh, S. Gold Core/Ceria Shell-Based Redox Active Nanozyme Mimicking the Biological Multienzyme Complex Phenomenon. J. Colloid Interface Sci. 2018, 513, 831–842. [Google Scholar] [CrossRef]
  22. Zhang, W.; Dong, J.; Wu, Y.; Cao, P.; Song, L.; Ma, M.; Gu, N.; Zhang, Y. Shape-Dependent Enzyme-like Activity of Co3O4 Nanoparticles and Their Conjugation with His-Tagged EGFR Single-Domain Antibody. Colloids Surf. B Biointerfaces 2017, 154, 55–62. [Google Scholar] [CrossRef]
  23. Wang, X.; Yang, Q.; Cao, Y.; Hao, H.; Zhou, J.; Hao, J. Metallosurfactant Ionogels in Imidazolium and Protic Ionic Liquids as Precursors to Synthesize Nanoceria as Catalase Mimetics for the Catalytic Decomposition of H2O2. Chem. Eur. J. 2016, 22, 17857–17865. [Google Scholar] [CrossRef] [PubMed]
  24. Gao, L.; Fan, K.; Yan, X. Iron Oxide Nanozyme: A Multifunctional Enzyme Mimetic for Biomedical Applications. Theranostics 2017, 7, 3207–3227. [Google Scholar] [CrossRef]
  25. Nicolini, V.; Gambuzzi, E.; Malavasi, G.; Menabue, L.; Menziani, M.C.; Lusvardi, G.; Pedone, A.; Benedetti, F.; Luches, P.; D’Addato, S.; et al. Evidence of Catalase Mimetic Activity in Ce3+/Ce4+ Doped Bioactive Glasses. J. Phys. Chem. B 2015, 119, 4009–4019. [Google Scholar] [CrossRef]
  26. Celardo, I.; Pedersen, J.Z.; Traversa, E.; Ghibelli, L. Pharmacological Potential of Cerium Oxide Nanoparticles. Nanoscale 2011, 3, 1411. [Google Scholar] [CrossRef]
  27. Huang, Y.; Ren, J.; Qu, X. Nanozymes: Classification, Catalytic Mechanisms, Activity Regulation, and Applications. Chem. Rev. 2019, 119, 4357–4412. [Google Scholar] [CrossRef]
  28. Zeng, L.; Cheng, H.; Dai, Y.; Su, Z.; Wang, C.; Lei, L.; Lin, D.; Li, X.; Chen, H.; Fan, K.; et al. In Vivo Regenerable Cerium Oxide Nanozyme-Loaded PH/H2O2-Responsive Nanovesicle for Tumor-Targeted Photothermal and Photodynamic Therapies. ACS Appl. Mater. Interfaces 2021, 13, 233–244. [Google Scholar] [CrossRef]
  29. Gao, L.; Zhuang, J.; Nie, L.; Zhang, J.; Zhang, Y.; Gu, N.; Wang, T.; Feng, J.; Yang, D.; Perrett, S.; et al. Intrinsic Peroxidase-like Activity of Ferromagnetic Nanoparticles. Nat. Nanotechnol. 2007, 2, 577–583. [Google Scholar] [CrossRef]
  30. Fan, J.; Yin, J.-J.; Ning, B.; Wu, X.; Hu, Y.; Ferrari, M.; Anderson, G.J.; Wei, J.; Zhao, Y.; Nie, G. Direct Evidence for Catalase and Peroxidase Activities of Ferritin–Platinum Nanoparticles. Biomaterials 2011, 32, 1611–1618. [Google Scholar] [CrossRef] [PubMed]
  31. Li, M.; Zhang, H.; Hou, Y.; Wang, X.; Xue, C.; Li, W.; Cai, K.; Zhao, Y.; Luo, Z. State-of-the-Art Iron-Based Nanozymes for Biocatalytic Tumor Therapy. Nanoscale Horiz. 2020, 5, 202–217. [Google Scholar] [CrossRef]
  32. Zhang, R.; Chen, L.; Liang, Q.; Xi, J.; Zhao, H.; Jin, Y.; Gao, X.; Yan, X.; Gao, L.; Fan, K. Unveiling the Active Sites on Ferrihydrite with Apparent Catalase-like Activity for Potentiating Radiotherapy. Nano Today 2021, 41, 101317. [Google Scholar] [CrossRef]
  33. Fujii, J.; Homma, T.; Osaki, T. Superoxide Radicals in the Execution of Cell Death. Antioxidants 2022, 11, 501. [Google Scholar] [CrossRef] [PubMed]
  34. Hayyan, M.; Hashim, M.A.; AlNashef, I.M. Superoxide Ion: Generation and Chemical Implications. Chem. Rev. 2016, 116, 3029–3085. [Google Scholar] [CrossRef] [PubMed]
  35. Fukai, T. Extracellular Superoxide Dismutase and Cardiovascular Disease. Cardiovasc. Res. 2002, 55, 239–249. [Google Scholar] [CrossRef]
  36. Liu, M.; Sun, X.; Chen, B.; Dai, R.; Xi, Z.; Xu, H. Insights into Manganese Superoxide Dismutase and Human Diseases. Int. J. Mol. Sci. 2022, 23, 15893. [Google Scholar] [CrossRef] [PubMed]
  37. Fukai, T.; Ushio-Fukai, M. Superoxide Dismutases: Role in Redox Signaling, Vascular Function, and Diseases. Antioxid. Redox Signal. 2011, 15, 1583–1606. [Google Scholar] [CrossRef] [PubMed]
  38. Sheng, Y.; Abreu, I.A.; Cabelli, D.E.; Maroney, M.J.; Miller, A.-F.; Teixeira, M.; Valentine, J.S. Superoxide Dismutases and Superoxide Reductases. Chem. Rev. 2014, 114, 3854–3918. [Google Scholar] [CrossRef] [PubMed]
  39. Weydert, C.J.; Cullen, J.J. Measurement of Superoxide Dismutase, Catalase and Glutathione Peroxidase in Cultured Cells and Tissue. Nat. Protoc. 2010, 5, 51–66. [Google Scholar] [CrossRef]
  40. Krusic, P.J.; Wasserman, E.; Keizer, P.N.; Morton, J.R.; Preston, K.F. Radical Reactions of C60. Science 1991, 254, 1183–1185. [Google Scholar] [CrossRef]
  41. Heckert, E.G.; Karakoti, A.S.; Seal, S.; Self, W.T. The Role of Cerium Redox State in the SOD Mimetic Activity of Nanoceria. Biomaterials 2008, 29, 2705–2709. [Google Scholar] [CrossRef] [PubMed]
  42. Zhao, H.; Zhang, R.; Yan, X.; Fan, K. Superoxide Dismutase Nanozymes: An Emerging Star for Anti-Oxidation. J. Mater. Chem. B 2021, 9, 6939–6957. [Google Scholar] [CrossRef]
  43. Damle, M.A.; Jakhade, A.P.; Chikate, R.C. Modulating Pro- and Antioxidant Activities of Nanoengineered Cerium Dioxide Nanoparticles against Escherichia Coli. ACS Omega 2019, 4, 3761–3771. [Google Scholar] [CrossRef]
  44. Korsvik, C.; Patil, S.; Seal, S.; Self, W.T. Superoxide Dismutase Mimetic Properties Exhibited by Vacancy Engineered Ceria Nanoparticles. Chem. Commun. 2007, 1056–1058. [Google Scholar] [CrossRef] [PubMed]
  45. D’Angelo, A.M.; Liu, A.C.Y.; Chaffee, A.L. Oxygen Uptake of Tb–CeO2: Analysis of Ce3+ and Oxygen Vacancies. J. Phys. Chem. C 2016, 120, 14382–14389. [Google Scholar] [CrossRef]
  46. Dutta, P.; Pal, S.; Seehra, M.S.; Shi, Y.; Eyring, E.M.; Ernst, R.D. Concentration of Ce 3+ and Oxygen Vacancies in Cerium Oxide Nanoparticles. Chem. Mater. 2006, 18, 5144–5146. [Google Scholar] [CrossRef]
  47. Yang, J.; Zhang, R.; Zhao, H.; Qi, H.; Li, J.; Li, J.; Zhou, X.; Wang, A.; Fan, K.; Yan, X.; et al. Bioinspired Copper Single-atom Nanozyme as a Superoxide Dismutase-like Antioxidant for Sepsis Treatment. Exploration 2022, 2, 20210267. [Google Scholar] [CrossRef]
  48. Sharifi, M.; Faryabi, K.; Talaei, A.J.; Shekha, M.S.; Ale-Ebrahim, M.; Salihi, A.; Nanakali, N.M.Q.; Aziz, F.M.; Rasti, B.; Hasan, A.; et al. Antioxidant Properties of Gold Nanozyme: A Review. J. Mol. Liq. 2020, 297, 112004. [Google Scholar] [CrossRef]
  49. Guo, S.; Guo, L. Unraveling the Multi-Enzyme-Like Activities of Iron Oxide Nanozyme via a First-Principles Microkinetic Study. J. Phys. Chem. C 2019, 123, 30318–30334. [Google Scholar] [CrossRef]
  50. Singh, N.; Geethika, M.; Eswarappa, S.M.; Mugesh, G. Manganese-Based Nanozymes: Multienzyme Redox Activity and Effect on the Nitric Oxide Produced by Endothelial Nitric Oxide Synthase. Chem. Eur. J. 2018, 24, 8393–8403. [Google Scholar] [CrossRef]
  51. Liu, Y.; Qing, Y.; Jing, L.; Zou, W.; Guo, R. Platinum–Copper Bimetallic Colloid Nanoparticle Cluster Nanozymes with Multiple Enzyme-like Activities for Scavenging Reactive Oxygen Species. Langmuir 2021, 37, 7364–7372. [Google Scholar] [CrossRef]
  52. Dong, J.; Song, L.; Yin, J.-J.; He, W.; Wu, Y.; Gu, N.; Zhang, Y. Co3O4 Nanoparticles with Multi-Enzyme Activities and Their Application in Immunohistochemical Assay. ACS Appl. Mater. Interfaces 2014, 6, 1959–1970. [Google Scholar] [CrossRef]
  53. Dashtestani, F.; Ghourchian, H.; Najafi, A. Silver-Gold-Apoferritin Nanozyme for Suppressing Oxidative Stress during Cryopreservation. Mater. Sci. Eng. C 2019, 94, 831–840. [Google Scholar] [CrossRef]
  54. Bielski, B.H.J.; Cabelli, D.E.; Arudi, R.L.; Ross, A.B. Reactivity of HO2/O2 Radicals in Aqueous Solution. J. Phys. Chem. Ref. Data 1985, 14, 1041–1100. [Google Scholar] [CrossRef]
  55. Shen, X.; Liu, W.; Gao, X.; Lu, Z.; Wu, X.; Gao, X. Mechanisms of Oxidase and Superoxide Dismutation-like Activities of Gold, Silver, Platinum, and Palladium, and Their Alloys: A General Way to the Activation of Molecular Oxygen. J. Am. Chem. Soc. 2015, 137, 15882–15891. [Google Scholar] [CrossRef]
  56. Zhang, D.; Shen, N.; Zhang, J.; Zhu, J.; Guo, Y.; Xu, L. A Novel Nanozyme Based on Selenopeptide-Modified Gold Nanoparticles with a Tunable Glutathione Peroxidase Activity. RSC Adv. 2020, 10, 8685–8691. [Google Scholar] [CrossRef] [PubMed]
  57. Kendall, A.; Woolcock, A.; Brooks, A.; Moore, G.E. Glutathione Peroxidase Activity, Plasma Total Antioxidant Capacity, and Urinary F2- Isoprostanes as Markers of Oxidative Stress in Anemic Dogs. J. Vet. Intern. Med. 2017, 31, 1700–1707. [Google Scholar] [CrossRef] [PubMed]
  58. Leonel, C.; Gelaleti, G.B.; Jardim, B.V.; Moschetta, M.G.; Regiani, V.R.; Oliveira, J.G.; Zuccari, D.A. Expression of Glutathione, Glutathione Peroxidase and Glutathione S-Transferase Pi in Canine Mammary Tumors. BMC Vet. Res. 2014, 10, 49. [Google Scholar] [CrossRef]
  59. Huchzermeyer, B.; Menghani, E.; Khardia, P.; Shilu, A. Metabolic Pathway of Natural Antioxidants, Antioxidant Enzymes and ROS Providence. Antioxidants 2022, 11, 761. [Google Scholar] [CrossRef] [PubMed]
  60. Lubos, E.; Loscalzo, J.; Handy, D.E. Glutathione Peroxidase-1 in Health and Disease: From Molecular Mechanisms to Therapeutic Opportunities. Antioxid. Redox Signal. 2011, 15, 1957–1997. [Google Scholar] [CrossRef]
  61. Flohe, L.; Günzler, W.A.; Schock, H.H. Glutathione Peroxidase: A Selenoenzyme. FEBS Lett. 1973, 32, 132–134. [Google Scholar] [CrossRef]
  62. Kraus, R.J.; Foster, S.J.; Ganther, H.E. Identification of Selenocysteine in Glutathione Peroxidase by Mass Spectroscopy. Biochemistry 1983, 22, 5853–5858. [Google Scholar] [CrossRef] [PubMed]
  63. Wu, J.; Yu, Y.; Cheng, Y.; Cheng, C.; Zhang, Y.; Jiang, B.; Zhao, X.; Miao, L.; Wei, H. Ligand-Dependent Activity Engineering of Glutathione Peroxidase-Mimicking MIL-47(V) Metal–Organic Framework Nanozyme for Therapy. Angew. Chem. Int. Ed. 2021, 60, 1227–1234. [Google Scholar] [CrossRef]
  64. Ghosh, S.; Roy, P.; Karmodak, N.; Jemmis, E.D.; Mugesh, G. Nanoisozymes: Crystal-Facet-Dependent Enzyme-Mimetic Activity of V 2 O 5 Nanomaterials. Angew. Chem. Int. Ed. 2018, 57, 4510–4515. [Google Scholar] [CrossRef]
  65. Zhang, S.; Li, Y.; Sun, S.; Liu, L.; Mu, X.; Liu, S.; Jiao, M.; Chen, X.; Chen, K.; Ma, H.; et al. Single-Atom Nanozymes Catalytically Surpassing Naturally Occurring Enzymes as Sustained Stitching for Brain Trauma. Nat. Commun. 2022, 13, 4744. [Google Scholar] [CrossRef] [PubMed]
  66. Hou, C.; Luo, Q.; Liu, J.; Miao, L.; Zhang, C.; Gao, Y.; Zhang, X.; Xu, J.; Dong, Z.; Liu, J. Construction of GPx Active Centers on Natural Protein Nanodisk/Nanotube: A New Way to Develop Artificial Nanoenzyme. ACS Nano 2012, 6, 8692–8701. [Google Scholar] [CrossRef]
  67. Vernekar, A.A.; Sinha, D.; Srivastava, S.; Paramasivam, P.U.; D’Silva, P.; Mugesh, G. An Antioxidant Nanozyme That Uncovers the Cytoprotective Potential of Vanadia Nanowires. Nat. Commun. 2014, 5, 5301. [Google Scholar] [CrossRef]
  68. Singh, N.; Savanur, M.A.; Srivastava, S.; D’Silva, P.; Mugesh, G. A Manganese Oxide Nanozyme Prevents the Oxidative Damage of Biomolecules without Affecting the Endogenous Antioxidant System. Nanoscale 2019, 11, 3855–3863. [Google Scholar] [CrossRef]
  69. Hao, C.; Qu, A.; Xu, L.; Sun, M.; Zhang, H.; Xu, C.; Kuang, H. Chiral Molecule-Mediated Porous CuxO Nanoparticle Clusters with Antioxidation Activity for Ameliorating Parkinson’s Disease. J. Am. Chem. Soc. 2019, 141, 1091–1099. [Google Scholar] [CrossRef]
  70. Liu, T.; Xiao, B.; Xiang, F.; Tan, J.; Chen, Z.; Zhang, X.; Wu, C.; Mao, Z.; Luo, G.; Chen, X.; et al. Ultrasmall Copper-Based Nanoparticles for Reactive Oxygen Species Scavenging and Alleviation of Inflammation Related Diseases. Nat. Commun. 2020, 11, 2788. [Google Scholar] [CrossRef]
  71. Adhikari, A.; Mondal, S.; Das, M.; Biswas, P.; Pal, U.; Darbar, S.; Bhattacharya, S.S.; Pal, D.; Saha-Dasgupta, T.; Das, A.K.; et al. Incorporation of a Biocompatible Nanozyme in Cellular Antioxidant Enzyme Cascade Reverses Huntington’s Like Disorder in Preclinical Model. Adv. Healthc. Mater. 2021, 10, 2001736. [Google Scholar] [CrossRef] [PubMed]
  72. Yan, R.; Sun, S.; Yang, J.; Long, W.; Wang, J.; Mu, X.; Li, Q.; Hao, W.; Zhang, S.; Liu, H.; et al. Nanozyme-Based Bandage with Single-Atom Catalysis for Brain Trauma. ACS Nano 2019, 13, 11552–11560. [Google Scholar] [CrossRef] [PubMed]
  73. Shlapa, Y.; Solopan, S.; Sarnatskaya, V.; Siposova, K.; Garcarova, I.; Veltruska, K.; Timashkov, I.; Lykhova, O.; Kolesnik, D.; Musatov, A.; et al. Cerium Dioxide Nanoparticles Synthesized via Precipitation at Constant PH: Synthesis, Physical-Chemical and Antioxidant Properties. Colloids Surf. B Biointerfaces 2022, 220, 112960. [Google Scholar] [CrossRef] [PubMed]
  74. Charbgoo, F.; Ahmad, M.; Darroudi, M. Cerium Oxide Nanoparticles: Green Synthesis and Biological Applications. Int. J. Nanomed. 2017, 12, 1401–1413. [Google Scholar] [CrossRef] [PubMed]
  75. Turin-Moleavin, I.-A.; Fifere, A.; Lungoci, A.-L.; Rosca, I.; Coroaba, A.; Peptanariu, D.; Nastasa, V.; Pasca, S.-A.; Bostanaru, A.-C.; Mares, M.; et al. In Vitro and In Vivo Antioxidant Activity of the New Magnetic-Cerium Oxide Nanoconjugates. Nanomaterials 2019, 9, 1565. [Google Scholar] [CrossRef] [PubMed]
  76. Gao, W.; He, J.; Chen, L.; Meng, X.; Ma, Y.; Cheng, L.; Tu, K.; Gao, X.; Liu, C.; Zhang, M.; et al. Deciphering the Catalytic Mechanism of Superoxide Dismutase Activity of Carbon Dot Nanozyme. Nat. Commun. 2023, 14, 160. [Google Scholar] [CrossRef]
  77. Singh, N.; NaveenKumar, S.K.; Geethika, M.; Mugesh, G. A Cerium Vanadate Nanozyme with Specific Superoxide Dismutase Activity Regulates Mitochondrial Function and ATP Synthesis in Neuronal Cells. Angew. Chem. Int. Ed. 2021, 60, 3121–3130. [Google Scholar] [CrossRef] [PubMed]
  78. Mao, M.; Guan, X.; Wu, F.; Ma, L. CoO Nanozymes with Multiple Catalytic Activities Regulate Atopic Dermatitis. Nanomaterials 2022, 12, 638. [Google Scholar] [CrossRef]
  79. Xia, F.; Hu, X.; Zhang, B.; Wang, X.; Guan, Y.; Lin, P.; Ma, Z.; Sheng, J.; Ling, D.; Li, F. Ultrasmall Ruthenium Nanoparticles with Boosted Antioxidant Activity Upregulate Regulatory T Cells for Highly Efficient Liver Injury Therapy. Small 2022, 18, 2201558. [Google Scholar] [CrossRef] [PubMed]
  80. Wang, J.; Tao, H.; Lu, T.; Wu, Y. Adsorption Enhanced the Oxidase-Mimicking Catalytic Activity of Octahedral-Shape Mn3O4 Nanoparticles as a Novel Colorimetric Chemosensor for Ultrasensitive and Selective Detection of Arsenic. J. Colloid Interface Sci. 2021, 584, 114–124. [Google Scholar] [CrossRef]
  81. Liu, Q.; Zhang, A.; Wang, R.; Zhang, Q.; Cui, D. A Review on Metal- and Metal Oxide-Based Nanozymes: Properties, Mechanisms, and Applications. Nano-Micro Lett. 2021, 13, 154. [Google Scholar] [CrossRef]
  82. Wu, H.; Liao, H.; Li, F.; Lee, J.; Hu, P.; Shao, W.; Li, X.; Ling, D. Bioactive ROS-scavenging Nanozymes for Regenerative Medicine: Reestablishing the Antioxidant Firewall. Nano Sel. 2020, 1, 285–297. [Google Scholar] [CrossRef]
  83. Yang, Z.; Luo, S.; Zeng, Y.; Shi, C.; Li, R. Albumin-Mediated Biomineralization of Shape-Controllable and Biocompatible Ceria Nanomaterials. ACS Appl. Mater. Interfaces 2017, 9, 6839–6848. [Google Scholar] [CrossRef]
  84. Singh, N.; Savanur, M.A.; Srivastava, S.; D’Silva, P.; Mugesh, G. A Redox Modulatory Mn3O4 Nanozyme with Multi-Enzyme Activity Provides Efficient Cytoprotection to Human Cells in a Parkinson’s Disease Model. Angew. Chem. Int. Ed. 2017, 56, 14267–14271. [Google Scholar] [CrossRef] [PubMed]
  85. Ge, C.; Fang, G.; Shen, X.; Chong, Y.; Wamer, W.G.; Gao, X.; Chai, Z.; Chen, C.; Yin, J.-J. Facet Energy versus Enzyme-like Activities: The Unexpected Protection of Palladium Nanocrystals against Oxidative Damage. ACS Nano 2016, 10, 10436–10445. [Google Scholar] [CrossRef]
  86. Dong, S.; Dong, Y.; Liu, B.; Liu, J.; Liu, S.; Zhao, Z.; Li, W.; Tian, B.; Zhao, R.; He, F.; et al. Guiding Transition Metal-Doped Hollow Cerium Tandem Nanozymes with Elaborately Regulated Multi-Enzymatic Activities for Intensive Chemodynamic Therapy. Adv. Mater. 2022, 34, 2107054. [Google Scholar] [CrossRef]
  87. Tian, R.; Sun, J.; Qi, Y.; Zhang, B.; Guo, S.; Zhao, M. Influence of VO2 Nanoparticle Morphology on the Colorimetric Assay of H2O2 and Glucose. Nanomaterials 2017, 7, 347. [Google Scholar] [CrossRef] [PubMed]
  88. Mu, J.; Zhang, L.; Zhao, M.; Wang, Y. Catalase Mimic Property of Co3O4 Nanomaterials with Different Morphology and Its Application as a Calcium Sensor. ACS Appl. Mater. Interfaces 2014, 6, 7090–7098. [Google Scholar] [CrossRef] [PubMed]
  89. Wu, J.; Wang, X.; Wang, Q.; Lou, Z.; Li, S.; Zhu, Y.; Qin, L.; Wei, H. Nanomaterials with Enzyme-like Characteristics (Nanozymes): Next-Generation Artificial Enzymes (II). Chem. Soc. Rev. 2019, 48, 1004–1076. [Google Scholar] [CrossRef] [PubMed]
  90. Zhu, Z.; Guan, Z.; Jia, S.; Lei, Z.; Lin, S.; Zhang, H.; Ma, Y.; Tian, Z.-Q.; Yang, C.J. Au@Pt Nanoparticle Encapsulated Target-Responsive Hydrogel with Volumetric Bar-Chart Chip Readout for Quantitative Point-of-Care Testing. Angew. Chem. Int. Ed. 2014, 53, 12503–12507. [Google Scholar] [CrossRef]
  91. Hu, X.; Saran, A.; Hou, S.; Wen, T.; Ji, Y.; Liu, W.; Zhang, H.; He, W.; Yin, J.-J.; Wu, X. Au@PtAg Core/Shell Nanorods: Tailoring Enzyme-like Activities via Alloying. RSC Adv. 2013, 3, 6095. [Google Scholar] [CrossRef]
  92. Boujakhrout, A.; Díez, P.; Martínez-Ruíz, P.; Sánchez, A.; Parrado, C.; Povedano, E.; Soto, P.; Pingarrón, J.M.; Villalonga, R. Gold Nanoparticles/Silver-Bipyridine Hybrid Nanobelts with Tuned Peroxidase-like Activity. RSC Adv. 2016, 6, 74957–74960. [Google Scholar] [CrossRef]
  93. Li, J.; Lv, L.; Zhang, G.; Zhou, X.; Shen, A.; Hu, J. Core–Shell Fructus Broussonetia-like Au@Ag@Pt Nanoparticles as Highly Efficient Peroxidase Mimetics for Supersensitive Resonance-Enhanced Raman Sensing. Anal. Methods 2016, 8, 2097–2105. [Google Scholar] [CrossRef]
  94. Long, L.; Liu, J.; Lu, K.; Zhang, T.; Xie, Y.; Ji, Y.; Wu, X. Highly Sensitive and Robust Peroxidase-like Activity of Au–Pt Core/Shell Nanorod-Antigen Conjugates for Measles Virus Diagnosis. J. Nanobiotechnol. 2018, 16, 46. [Google Scholar] [CrossRef] [PubMed]
  95. Dashtestani, F.; Ghourchian, H.; Eskandari, K.; Rafiee-Pour, H.-A. A Superoxide Dismutase Mimic Nanocomposite for Amperometric Sensing of Superoxide Anions. Microchim. Acta 2015, 182, 1045–1053. [Google Scholar] [CrossRef]
  96. Matysik, J.; Długosz, O.; Loureiro, J.; da Silva Pereira, M. do C.; Banach, M. Multioxide-Superoxide Dismutase Enzyme-Nanocomplexes and Their Antioxidant Activity. J. Mater. Sci. 2022, 57, 15954–15966. [Google Scholar] [CrossRef]
  97. Zhou, J.; Xu, D.; Tian, G.; He, Q.; Zhang, X.; Liao, J.; Mei, L.; Chen, L.; Gao, L.; Zhao, L.; et al. Coordination-Driven Self-Assembly Strategy-Activated Cu Single-Atom Nanozymes for Catalytic Tumor-Specific Therapy. J. Am. Chem. Soc. 2023, 145, 4279–4293. [Google Scholar] [CrossRef]
  98. Al-Ani, L.A.; Yehye, W.A.; Kadir, F.A.; Hashim, N.M.; AlSaadi, M.A.; Julkapli, N.M.; Hsiao, V.K.S. Hybrid Nanocomposite Curcumin-Capped Gold Nanoparticle-Reduced Graphene Oxide: Anti-Oxidant Potency and Selective Cancer Cytotoxicity. PLoS ONE 2019, 14, e0216725. [Google Scholar] [CrossRef] [PubMed]
  99. Tamura, H.; Mita, K.; Tanaka, A.; Ito, M. Mechanism of Hydroxylation of Metal Oxide Surfaces. J. Colloid Interface Sci. 2001, 243, 202–207. [Google Scholar] [CrossRef]
  100. Tamura, H.; Tanaka, A.; Mita, K.; Furuichi, R. Surface Hydroxyl Site Densities on Metal Oxides as a Measure for the Ion-Exchange Capacity. J. Colloid Interface Sci. 1999, 209, 225–231. [Google Scholar] [CrossRef]
  101. Zhou, X.; You, M.; Wang, F.; Wang, Z.; Gao, X.; Jing, C.; Liu, J.; Guo, M.; Li, J.; Luo, A.; et al. Multifunctional Graphdiyne–Cerium Oxide Nanozymes Facilitate MicroRNA Delivery and Attenuate Tumor Hypoxia for Highly Efficient Radiotherapy of Esophageal Cancer. Adv. Mater. 2021, 33, 2100556. [Google Scholar] [CrossRef] [PubMed]
  102. Huang, Y.; Liu, Z.; Liu, C.; Ju, E.; Zhang, Y.; Ren, J.; Qu, X. Self-Assembly of Multi-Nanozymes to Mimic an Intracellular Antioxidant Defense System. Angew. Chem. Int. Ed. 2016, 55, 6646–6650. [Google Scholar] [CrossRef]
  103. Lin, S.; Cheng, Y.; Zhang, H.; Wang, X.; Zhang, Y.; Zhang, Y.; Miao, L.; Zhao, X.; Wei, H. Copper Tannic Acid Coordination Nanosheet: A Potent Nanozyme for Scavenging ROS from Cigarette Smoke. Small 2020, 16, 1902123. [Google Scholar] [CrossRef]
  104. Wang, S.; Chen, W.; Liu, A.-L.; Hong, L.; Deng, H.-H.; Lin, X.-H. Comparison of the Peroxidase-Like Activity of Unmodified, Amino-Modified, and Citrate-Capped Gold Nanoparticles. ChemPhysChem 2012, 13, 1199–1204. [Google Scholar] [CrossRef]
  105. Abu Tarboush, N.; Jensen, L.M.R.; Feng, M.; Tachikawa, H.; Wilmot, C.M.; Davidson, V.L. Functional Importance of Tyrosine 294 and the Catalytic Selectivity for the Bis-Fe(IV) State of MauG Revealed by Replacement of This Axial Heme Ligand with Histidine. Biochemistry 2010, 49, 9783–9791. [Google Scholar] [CrossRef]
  106. Mohammad, M.; Ahmadpoor, F.; Shojaosadati, S.A. Mussel-Inspired Magnetic Nanoflowers as an Effective Nanozyme and Antimicrobial Agent for Biosensing and Catalytic Reduction of Organic Dyes. ACS Omega 2020, 5, 18766–18777. [Google Scholar] [CrossRef]
  107. Zhang, D.-Y.; Tu, T.; Younis, M.R.; Zhu, K.S.; Liu, H.; Lei, S.; Qu, J.; Lin, J.; Huang, P. Clinically Translatable Gold Nanozymes with Broad Spectrum Antioxidant and Anti-Inflammatory Activity for Alleviating Acute Kidney Injury. Theranostics 2021, 11, 9904–9917. [Google Scholar] [CrossRef] [PubMed]
  108. Dong, S.; Dong, Y.; Jia, T.; Liu, S.; Liu, J.; Yang, D.; He, F.; Gai, S.; Yang, P.; Lin, J. GSH-Depleted Nanozymes with Hyperthermia-Enhanced Dual Enzyme-Mimic Activities for Tumor Nanocatalytic Therapy. Adv. Mater. 2020, 32, 2002439. [Google Scholar] [CrossRef] [PubMed]
  109. Mu, X.; Wang, J.; Li, Y.; Xu, F.; Long, W.; Ouyang, L.; Liu, H.; Jing, Y.; Wang, J.; Dai, H.; et al. Redox Trimetallic Nanozyme with Neutral Environment Preference for Brain Injury. ACS Nano 2019, 13, 1870–1884. [Google Scholar] [CrossRef] [PubMed]
  110. Liu, C.-P.; Wu, T.-H.; Lin, Y.-L.; Liu, C.-Y.; Wang, S.; Lin, S.-Y. Tailoring Enzyme-Like Activities of Gold Nanoclusters by Polymeric Tertiary Amines for Protecting Neurons Against Oxidative Stress. Small 2016, 12, 4127–4135. [Google Scholar] [CrossRef]
  111. Jain, V.; Bhagat, S.; Singh, M.; Bansal, V.; Singh, S. Unveiling the Effect of 11-MUA Coating on Biocompatibility and Catalytic Activity of a Gold-Core Cerium Oxide-Shell-Based Nanozyme. RSC Adv. 2019, 9, 33195–33206. [Google Scholar] [CrossRef]
  112. Wang, M.; Liang, Y.; Liao, F.; Younis, M.R.; Zheng, Y.; Zhao, X.; Yu, X.; Guo, W.; Zhang, D.-Y. Iridium Tungstate Nanozyme-Mediated Hypoxic Regulation and Anti-Inflammation for Duplex Imaging Guided Photothermal Therapy of Metastatic Breast Tumors. ACS Appl. Mater. Interfaces 2022, 14, 56471–56482. [Google Scholar] [CrossRef]
  113. Yan, B.C.; Cao, J.; Liu, J.; Gu, Y.; Xu, Z.; Li, D.; Gao, L. Dietary Fe3O4 Nanozymes Prevent the Injury of Neurons and Blood–Brain Barrier Integrity from Cerebral Ischemic Stroke. ACS Biomater. Sci. Eng. 2021, 7, 299–310. [Google Scholar] [CrossRef]
  114. Jørgensen, M.; Grönbeck, H. Strain Affects CO Oxidation on Metallic Nanoparticles Non-Linearly. Top. Catal. 2019, 62, 660–668. [Google Scholar] [CrossRef]
  115. Ellaby, T.; Varambhia, A.; Luo, X.; Briquet, L.; Sarwar, M.; Ozkaya, D.; Thompsett, D.; Nellist, P.D.; Skylaris, C.-K. Strain Effects in Core–Shell PtCo Nanoparticles: A Comparison of Experimental Observations and Computational Modelling. Phys. Chem. Chem. Phys. 2020, 22, 24784–24795. [Google Scholar] [CrossRef]
  116. Zhang, S.; Zhang, X.; Jiang, G.; Zhu, H.; Guo, S.; Su, D.; Lu, G.; Sun, S. Tuning Nanoparticle Structure and Surface Strain for Catalysis Optimization. J. Am. Chem. Soc. 2014, 136, 7734–7739. [Google Scholar] [CrossRef] [PubMed]
  117. Han, S.I.; Lee, S.; Cho, M.G.; Yoo, J.M.; Oh, M.H.; Jeong, B.; Kim, D.; Park, O.K.; Kim, J.; Namkoong, E.; et al. Epitaxially Strained CeO2 /Mn3O4 Nanocrystals as an Enhanced Antioxidant for Radioprotection. Adv. Mater. 2020, 32, 2001566. [Google Scholar] [CrossRef]
  118. Li, L.; Li, H.; Shi, L.; Shi, L.; Li, T. Tin Porphyrin-Based Nanozymes with Unprecedented Superoxide Dismutase-Mimicking Activities. Langmuir 2022, 38, 7272–7279. [Google Scholar] [CrossRef]
  119. Niu, X.; Li, X.; Lyu, Z.; Pan, J.; Ding, S.; Ruan, X.; Zhu, W.; Du, D.; Lin, Y. Metal–Organic Framework Based Nanozymes: Promising Materials for Biochemical Analysis. Chem. Commun. 2020, 56, 11338–11353. [Google Scholar] [CrossRef]
  120. Zhang, L.; Zhang, Y.; Wang, Z.; Cao, F.; Sang, Y.; Dong, K.; Pu, F.; Ren, J.; Qu, X. Constructing Metal–Organic Framework Nanodots as Bio-Inspired Artificial Superoxide Dismutase for Alleviating Endotoxemia. Mater. Horiz. 2019, 6, 1682–1687. [Google Scholar] [CrossRef]
  121. Liu, Y.; Li, H.; Liu, W.; Guo, J.; Yang, H.; Tang, H.; Tian, M.; Nie, H.; Zhang, X.; Long, W. Design of Monovalent Cerium-Based Metal Organic Frameworks as Bioinspired Superoxide Dismutase Mimics for Ionizing Radiation Protection. ACS Appl. Mater. Interfaces 2022, 14, 54587–54597. [Google Scholar] [CrossRef] [PubMed]
  122. Baumann, A.E.; Burns, D.A.; Liu, B.; Thoi, V.S. Metal-Organic Framework Functionalization and Design Strategies for Advanced Electrochemical Energy Storage Devices. Commun. Chem. 2019, 2, 86. [Google Scholar] [CrossRef]
  123. Cooper, L.; Hidalgo, T.; Gorman, M.; Lozano-Fernández, T.; Simón-Vázquez, R.; Olivier, C.; Guillou, N.; Serre, C.; Martineau, C.; Taulelle, F.; et al. A Biocompatible Porous Mg-Gallate Metal–Organic Framework as an Antioxidant Carrier. Chem. Commun. 2015, 51, 5848–5851. [Google Scholar] [CrossRef]
  124. Wang, J.; Li, W.; Zheng, Y.-Q. Nitro-Functionalized Metal–Organic Frameworks with Catalase Mimic Properties for Glutathione Detection. Analyst 2019, 144, 6041–6047. [Google Scholar] [CrossRef]
  125. Zhou, W.; Li, H.; Xia, B.; Ji, W.; Ji, S.; Zhang, W.; Huang, W.; Huo, F.; Xu, H. Selenium-Functionalized Metal-Organic Frameworks as Enzyme Mimics. Nano Res. 2018, 11, 5761–5768. [Google Scholar] [CrossRef]
  126. Sang, Y.; Cao, F.; Li, W.; Zhang, L.; You, Y.; Deng, Q.; Dong, K.; Ren, J.; Qu, X. Bioinspired Construction of a Nanozyme-Based H2O2 Homeostasis Disruptor for Intensive Chemodynamic Therapy. J. Am. Chem. Soc. 2020, 142, 5177–5183. [Google Scholar] [CrossRef]
  127. Wang, D.; Wu, H.; Phua, S.Z.F.; Yang, G.; Qi Lim, W.; Gu, L.; Qian, C.; Wang, H.; Guo, Z.; Chen, H.; et al. Self-Assembled Single-Atom Nanozyme for Enhanced Photodynamic Therapy Treatment of Tumor. Nat. Commun. 2020, 11, 357. [Google Scholar] [CrossRef]
  128. Abdelhamid, H.N.; Mahmoud, G.A.-E.; Sharmouk, W. A Cerium-Based MOFzyme with Multi-Enzyme-like Activity for the Disruption and Inhibition of Fungal Recolonization. J. Mater. Chem. B 2020, 8, 7548–7556. [Google Scholar] [CrossRef]
  129. Reuter, S.; Gupta, S.C.; Chaturvedi, M.M.; Aggarwal, B.B. Oxidative Stress, Inflammation, and Cancer: How Are They Linked? Free. Radic. Biol. Med. 2010, 49, 1603–1616. [Google Scholar] [CrossRef]
  130. Zahra, K.F.; Lefter, R.; Ali, A.; Abdellah, E.-C.; Trus, C.; Ciobica, A.; Timofte, D. The Involvement of the Oxidative Stress Status in Cancer Pathology: A Double View on the Role of the Antioxidants. Oxidative Med. Cell. Longev. 2021, 2021, 9965916. [Google Scholar] [CrossRef] [PubMed]
  131. Ismail, N.A.S.; Lee, J.X.; Yusof, F. Platinum Nanoparticles: The Potential Antioxidant in the Human Lung Cancer Cells. Antioxidants 2022, 11, 986. [Google Scholar] [CrossRef] [PubMed]
  132. Alpaslan, E.; Yazici, H.; Golshan, N.H.; Ziemer, K.S.; Webster, T.J. PH-Dependent Activity of Dextran-Coated Cerium Oxide Nanoparticles on Prohibiting Osteosarcoma Cell Proliferation. ACS Biomater. Sci. Eng. 2015, 1, 1096–1103. [Google Scholar] [CrossRef]
  133. Asati, A.; Santra, S.; Kaittanis, C.; Perez, J.M. Surface-Charge-Dependent Cell Localization and Cytotoxicity of Cerium Oxide Nanoparticles. ACS Nano 2010, 4, 5321–5331. [Google Scholar] [CrossRef]
  134. Giorgio, M.; Trinei, M.; Migliaccio, E.; Pelicci, P.G. Hydrogen Peroxide: A Metabolic by-Product or a Common Mediator of Ageing Signals? Nat. Rev. Mol. Cell. Biol. 2007, 8, 722–728. [Google Scholar] [CrossRef] [PubMed]
  135. Bartley, M.G.; Marquardt, K.; Kirchhof, D.; Wilkins, H.M.; Patterson, D.; Linseman, D.A. Overexpression of Amyloid-β Protein Precursor Induces Mitochondrial Oxidative Stress and Activates the Intrinsic Apoptotic Cascade. J. Alzheimer’s Dis. 2012, 28, 855–868. [Google Scholar] [CrossRef] [PubMed]
  136. Westermann, B. Nitric Oxide Links Mitochondrial Fission to Alzheimer’s Disease. Sci. Signal. 2009, 2, pe29. [Google Scholar] [CrossRef] [PubMed]
  137. Chen, T.; Zou, H.; Wu, X.; Liu, C.; Situ, B.; Zheng, L.; Yang, G. Nanozymatic Antioxidant System Based on MoS2 Nanosheets. ACS Appl. Mater. Interfaces 2018, 10, 12453–12462. [Google Scholar] [CrossRef]
  138. Kwon, H.J.; Cha, M.-Y.; Kim, D.; Kim, D.K.; Soh, M.; Shin, K.; Hyeon, T.; Mook-Jung, I. Mitochondria-Targeting Ceria Nanoparticles as Antioxidants for Alzheimer’s Disease. ACS Nano 2016, 10, 2860–2870. [Google Scholar] [CrossRef]
  139. Chen, Q.; Du, Y.; Zhang, K.; Liang, Z.; Li, J.; Yu, H.; Ren, R.; Feng, J.; Jin, Z.; Li, F.; et al. Tau-Targeted Multifunctional Nanocomposite for Combinational Therapy of Alzheimer’s Disease. ACS Nano 2018, 12, 1321–1338. [Google Scholar] [CrossRef]
  140. Kwon, H.J.; Kim, D.; Seo, K.; Kim, Y.G.; Han, S.I.; Kang, T.; Soh, M.; Hyeon, T. Ceria Nanoparticle Systems for Selective Scavenging of Mitochondrial, Intracellular, and Extracellular Reactive Oxygen Species in Parkinson’s Disease. Angew. Chem. Int. Ed. 2018, 57, 9408–9412. [Google Scholar] [CrossRef]
  141. Fu, S.; Chen, H.; Yang, W.; Xia, X.; Zhao, S.; Xu, X.; Ai, P.; Cai, Q.; Li, X.; Wang, Y.; et al. ROS-Targeted Depression Therapy via BSA-Incubated Ceria Nanoclusters. Nano Lett. 2022, 22, 4519–4527. [Google Scholar] [CrossRef]
  142. Hao, T.; Li, J.; Yao, F.; Dong, D.; Wang, Y.; Yang, B.; Wang, C. Injectable Fullerenol/Alginate Hydrogel for Suppression of Oxidative Stress Damage in Brown Adipose-Derived Stem Cells and Cardiac Repair. ACS Nano 2017, 11, 5474–5488. [Google Scholar] [CrossRef]
  143. Han, J.; Kim, Y.S.; Lim, M.-Y.; Kim, H.Y.; Kong, S.; Kang, M.; Choo, Y.W.; Jun, J.H.; Ryu, S.; Jeong, H.; et al. Dual Roles of Graphene Oxide to Attenuate Inflammation and Elicit Timely Polarization of Macrophage Phenotypes for Cardiac Repair. ACS Nano 2018, 12, 1959–1977. [Google Scholar] [CrossRef]
  144. Li, F.; Qiu, Y.; Xia, F.; Sun, H.; Liao, H.; Xie, A.; Lee, J.; Lin, P.; Wei, M.; Shao, Y.; et al. Dual Detoxification and Inflammatory Regulation by Ceria Nanozymes for Drug-Induced Liver Injury Therapy. Nano Today 2020, 35, 100925. [Google Scholar] [CrossRef]
  145. Yao, J.; Cheng, Y.; Zhou, M.; Zhao, S.; Lin, S.; Wang, X.; Wu, J.; Li, S.; Wei, H. ROS Scavenging Mn3O4 Nanozymes for in Vivo Anti-Inflammation. Chem. Sci. 2018, 9, 2927–2933. [Google Scholar] [CrossRef] [PubMed]
  146. Bao, X.; Zhao, J.; Sun, J.; Hu, M.; Yang, X. Polydopamine Nanoparticles as Efficient Scavengers for Reactive Oxygen Species in Periodontal Disease. ACS Nano 2018, 12, 8882–8892. [Google Scholar] [CrossRef]
  147. Zhao, S.; Duan, H.; Yang, Y.; Yan, X.; Fan, K. Fenozyme Protects the Integrity of the Blood–Brain Barrier against Experimental Cerebral Malaria. Nano Lett. 2019, 19, 8887–8895. [Google Scholar] [CrossRef] [PubMed]
  148. Zhang, X.; Zhang, S.; Yang, Z.; Wang, Z.; Tian, X.; Zhou, R. Self-Cascade MoS2 Nanozymes for Efficient Intracellular Antioxidation and Hepatic Fibrosis Therapy. Nanoscale 2021, 13, 12613–12622. [Google Scholar] [CrossRef] [PubMed]
  149. Long, Y.; Wei, H.; Li, J.; Li, M.; Wang, Y.; Zhang, Z.; Cao, T.; Carlos, C.; German, L.G.; Jiang, D.; et al. Prevention of Hepatic Ischemia-Reperfusion Injury by Carbohydrate-Derived Nanoantioxidants. Nano Lett. 2020, 20, 6510–6519. [Google Scholar] [CrossRef]
  150. Liu, Y.; Ai, K.; Ji, X.; Askhatova, D.; Du, R.; Lu, L.; Shi, J. Comprehensive Insights into the Multi-Antioxidative Mechanisms of Melanin Nanoparticles and Their Application to Protect Brain from Injury in Ischemic Stroke. J. Am. Chem. Soc. 2017, 139, 856–862. [Google Scholar] [CrossRef]
  151. Zhang, D.-Y.; Liu, H.; Li, C.; Younis, M.R.; Lei, S.; Yang, C.; Lin, J.; Li, Z.; Huang, P. Ceria Nanozymes with Preferential Renal Uptake for Acute Kidney Injury Alleviation. ACS Appl. Mater. Interfaces 2020, 12, 56830–56838. [Google Scholar] [CrossRef]
  152. Wang, K.; Zhang, Y.; Mao, W.; Feng, W.; Lu, S.; Wan, J.; Song, X.; Chen, Y.; Peng, B. Engineering Ultrasmall Ferroptosis-Targeting and Reactive Oxygen/Nitrogen Species-Scavenging Nanozyme for Alleviating Acute Kidney Injury. Adv Funct Mater. 2022, 32, 2109221. [Google Scholar] [CrossRef]
  153. Su, X.; Xie, X.; Liu, L.; Lv, J.; Song, F.; Perkovic, V.; Zhang, H. Comparative Effectiveness of 12 Treatment Strategies for Preventing Contrast-Induced Acute Kidney Injury: A Systematic Review and Bayesian Network Meta-Analysis. Am. J. Kidney Dis. 2017, 69, 69–77. [Google Scholar] [CrossRef]
  154. Fraga, C.M.; Tomasi, C.D.; de Castro Damasio, D.; Vuolo, F.; Ritter, C.; Dal-Pizzol, F. N-Acetylcysteine plus Deferoxamine for Patients with Prolonged Hypotension Does Not Decrease Acute Kidney Injury Incidence: A Double Blind, Randomized, Placebo-Controlled Trial. Crit. Care 2016, 20, 331. [Google Scholar] [CrossRef]
  155. Liesenfeld, L.F.; Wagner, B.; Hillebrecht, H.C.; Brune, M.; Eckert, C.; Klose, J.; Schmidt, T.; Büchler, M.W.; Schneider, M. HIPEC-Induced Acute Kidney Injury: A Retrospective Clinical Study and Preclinical Model. Ann. Surg. Oncol. 2022, 29, 139–151. [Google Scholar] [CrossRef] [PubMed]
  156. Weng, Q.; Sun, H.; Fang, C.; Xia, F.; Liao, H.; Lee, J.; Wang, J.; Xie, A.; Ren, J.; Guo, X.; et al. Catalytic Activity Tunable Ceria Nanoparticles Prevent Chemotherapy-Induced Acute Kidney Injury without Interference with Chemotherapeutics. Nat. Commun. 2021, 12, 1436. [Google Scholar] [CrossRef]
  157. Zhang, D.-Y.; Liu, H.; Younis, M.R.; Lei, S.; Yang, C.; Lin, J.; Qu, J.; Huang, P. Ultrasmall Platinum Nanozymes as Broad-Spectrum Antioxidants for Theranostic Application in Acute Kidney Injury. Chem. Eng. J. 2021, 409, 127371. [Google Scholar] [CrossRef]
  158. Zhang, D.-Y.; Younis, M.R.; Liu, H.; Lei, S.; Wan, Y.; Qu, J.; Lin, J.; Huang, P. Multi-Enzyme Mimetic Ultrasmall Iridium Nanozymes as Reactive Oxygen/Nitrogen Species Scavengers for Acute Kidney Injury Management. Biomaterials 2021, 271, 120706. [Google Scholar] [CrossRef]
  159. Meng, L.; Feng, J.; Gao, J.; Zhang, Y.; Mo, W.; Zhao, X.; Wei, H.; Guo, H. Reactive Oxygen Species- and Cell-Free DNA-Scavenging Mn3O4 Nanozymes for Acute Kidney Injury Therapy. ACS Appl. Mater. Interfaces 2022, 14, 50649–50663. [Google Scholar] [CrossRef]
  160. Dinu Gugoasa, L.A.; Pogacean, F.; Kurbanoglu, S.; Tudoran, L.-B.; Serban, A.B.; Kacso, I.; Pruneanu, S. Graphene-Gold Nanoparticles Nanozyme-Based Electrochemical Sensor with Enhanced Laccase-Like Activity for Determination of Phenolic Substrates. J. Electrochem. Soc. 2021, 168, 067523. [Google Scholar] [CrossRef]
  161. Dinu, L.A.; Kurbanoglu, S.; Romanitan, C.; Pruneanu, S.; Serban, A.B.; Stoian, M.C.; Pachiu, C.; Craciun, G. Electrodeposited Copper Nanocubes on Multi-Layer Graphene: A Novel Nanozyme for Ultrasensitive Dopamine Detection from Biological Samples. Appl. Surf. Sci. 2022, 604, 154392. [Google Scholar] [CrossRef]
  162. Dinu, L.A.; Baracu, A.M.; Brincoveanu, O. The Non-Enzymatic Detection of the Pollutant Bisphenol A Using S-Graphene as Nanozyme Material. In Proceedings of the 2022 International Semiconductor Conference (CAS), Poiana Brasov, Romania, 12–14 October 2022; IEEE: New York, NY, USA; pp. 95–98. [Google Scholar]
  163. Liang, H.; Liu, Y.; Qileng, A.; Shen, H.; Liu, W.; Xu, Z.; Liu, Y. PEI-Coated Prussian Blue Nanocubes as PH-Switchable Nanozyme: Broad-PH-Responsive Immunoassay for Illegal Additive. Biosens. Bioelectron. 2023, 219, 114797. [Google Scholar] [CrossRef] [PubMed]
  164. Li, W.; Wang, J.; Zhu, J.; Zheng, Y.-Q. Co 3 O 4 Nanocrystals as an Efficient Catalase Mimic for the Colorimetric Detection of Glutathione. J. Mater. Chem. B 2018, 6, 6858–6864. [Google Scholar] [CrossRef] [PubMed]
  165. Niu, L.; Cai, Y.; Dong, T.; Zhang, Y.; Liu, X.; Zhang, X.; Zeng, L.; Liu, A. Vanadium Nitride@carbon Nanofiber Composite: Synthesis, Cascade Enzyme Mimics and Its Sensitive and Selective Colorimetric Sensing of Superoxide Anion. Biosens. Bioelectron. 2022, 210, 114285. [Google Scholar] [CrossRef]
  166. Liu, C.; Cai, Y.; Wang, J.; Liu, X.; Ren, H.; Yan, L.; Zhang, Y.; Yang, S.; Guo, J.; Liu, A. Facile Preparation of Homogeneous Copper Nanoclusters Exhibiting Excellent Tetraenzyme Mimetic Activities for Colorimetric Glutathione Sensing and Fluorimetric Ascorbic Acid Sensing. ACS Appl. Mater. Interfaces 2020, 12, 42521–42530. [Google Scholar] [CrossRef]
  167. Liu, C.; You, X.; Lu, D.; Shi, G.; Deng, J.; Zhou, T. Gelsolin Encountering Ag Nanorods/Triangles: An Aggregation-Based Colorimetric Sensor Array for in Vivo Monitoring the Cerebrospinal Aβ 42% as an Indicator of Cd2+ Exposure-Related Alzheimer’s Disease Pathogenesis. ACS Appl. Bio Mater. 2020, 3, 7965–7973. [Google Scholar] [CrossRef]
  168. Nakamura, H.; Takada, K. Reactive Oxygen Species in Cancer: Current Findings and Future Directions. Cancer Sci. 2021, 112, 3945–3952. [Google Scholar] [CrossRef]
  169. Dias, V.; Junn, E.; Mouradian, M.M. The Role of Oxidative Stress in Parkinson’s Disease. J. Park. Dis. 2013, 3, 461–491. [Google Scholar] [CrossRef]
  170. Wang, Z.; Zhao, H.; Chen, K.; Zhou, F.; Magdassi, S.; Lan, M. Two-Dimensional Mesoporous Nitrogen-Rich Carbon Nanosheets Loaded with CeO2 Nanoclusters as Nanozymes for the Electrochemical Detection of Superoxide Anions in HepG2 Cells. Biosens. Bioelectron. 2022, 209, 114229. [Google Scholar] [CrossRef] [PubMed]
  171. Cai, X.; Shi, L.; Sun, W.; Zhao, H.; Li, H.; He, H.; Lan, M. A Facile Way to Fabricate Manganese Phosphate Self-Assembled Carbon Networks as Efficient Electrochemical Catalysts for Real-Time Monitoring of Superoxide Anions Released from HepG2 Cells. Biosens. Bioelectron. 2018, 102, 171–178. [Google Scholar] [CrossRef] [PubMed]
  172. Wang, Z.; Zhao, H.; Gao, Q.; Chen, K.; Lan, M. Facile Synthesis of Ultrathin Two-Dimensional Graphene-like CeO2–TiO2 Mesoporous Nanosheet Loaded with Ag Nanoparticles for Non-Enzymatic Electrochemical Detection of Superoxide Anions in HepG2 Cells. Biosens. Bioelectron. 2021, 184, 113236. [Google Scholar] [CrossRef] [PubMed]
  173. Zou, Z.; Chen, J.; Shi, Z.; Yuan, C.; Zhou, G.; Liu, Q.; Chen, H.; Zeng, Q.; Liang, T.; Tang, K.L.; et al. Cobalt Phosphates Loaded into Iodine-Spaced Reduced Graphene Oxide Nanolayers for Electrochemical Measurement of Superoxide Generated by Cells. ACS Appl. Nano Mater. 2021, 4, 3631–3638. [Google Scholar] [CrossRef]
Figure 1. Manipulation of antioxidant nanozymes exhibiting CAT-, SOD-, and GPx-like activity.
Figure 1. Manipulation of antioxidant nanozymes exhibiting CAT-, SOD-, and GPx-like activity.
Micromachines 14 01017 g001
Figure 2. Cerium-based nanozyme exhibited CAT-like activity through cycling between Ce4+ and Ce3+ oxidation states. Reprinted with permission from [28]. Copyright (2020) ACS publications.
Figure 2. Cerium-based nanozyme exhibited CAT-like activity through cycling between Ce4+ and Ce3+ oxidation states. Reprinted with permission from [28]. Copyright (2020) ACS publications.
Micromachines 14 01017 g002
Figure 4. Effect of morphology on the SOD-like and CAT-like activities of Pd nanocrystals. Quantitative analysis of the ROS levels by flow cytometry. Data are represented as the mean fluorescence intensity. p-values compared to H2O2-treated cells were calculated by Student’s t test: ** p < 0.01, *** p < 0.001. Reprinted with permission from [85]. Copyright (2016) ACS publications.
Figure 4. Effect of morphology on the SOD-like and CAT-like activities of Pd nanocrystals. Quantitative analysis of the ROS levels by flow cytometry. Data are represented as the mean fluorescence intensity. p-values compared to H2O2-treated cells were calculated by Student’s t test: ** p < 0.01, *** p < 0.001. Reprinted with permission from [85]. Copyright (2016) ACS publications.
Micromachines 14 01017 g004
Figure 5. (a) Illustration of nanocomposite consisting of multiple elements increasing CAT-like activity. Reprinted with permission from [97]. Copyright (2023) ACS publications. (b) SOD-like activity of PEGylated, glycinated, and uncoated cerium nanoparticles. Reprinted with permission from [43]. Copyright (2019) ACS publications.
Figure 5. (a) Illustration of nanocomposite consisting of multiple elements increasing CAT-like activity. Reprinted with permission from [97]. Copyright (2023) ACS publications. (b) SOD-like activity of PEGylated, glycinated, and uncoated cerium nanoparticles. Reprinted with permission from [43]. Copyright (2019) ACS publications.
Micromachines 14 01017 g005
Figure 6. Schematic illustration of the design and characterization of catalytic-activity-tunable ceria nanoparticles (CNPs) that protect against chemotherapy-induced acute kidney injury (AKI). (a) CNPs switch their antioxidant activity in a pH-dependent manner. (b) CNPs scavenge the excessive chemotherapeutics-induced ROS. (c) The acidic tumor microenvironment suppresses their ROS scavenging capability without causing interference with the chemotherapeutic efficiency. Reprinted with permission from [156]. Copyright (2021) Springer Nature.
Figure 6. Schematic illustration of the design and characterization of catalytic-activity-tunable ceria nanoparticles (CNPs) that protect against chemotherapy-induced acute kidney injury (AKI). (a) CNPs switch their antioxidant activity in a pH-dependent manner. (b) CNPs scavenge the excessive chemotherapeutics-induced ROS. (c) The acidic tumor microenvironment suppresses their ROS scavenging capability without causing interference with the chemotherapeutic efficiency. Reprinted with permission from [156]. Copyright (2021) Springer Nature.
Micromachines 14 01017 g006
Table 1. Manipulation of antioxidant nanozymes.
Table 1. Manipulation of antioxidant nanozymes.
FactorNanozymeCatalytic
Activity
Outcome
SizeCeO2 nanoparticle
(~5–28 nm) [13]
CAT, SODCatalytic activities are enhanced for smaller particles and for the particles with larger Ce3+ fractions
Carbon dot
(2 nm) [76]
SODSOD activity of over 10,000 U/mg
CeVO4 nanorod
(50–150 nm) [77]
SODThe inhibition of formazan production by rods with sizes of 50, 100, and 150 were 4.12 ± 0.19, 3.93 ± 0.20, and 2.57 ± 0.07 ng/µL, respectively
MorphologyCeria nanomaterial
(nanocluster, nanoparticle, nanochain) [83]
SODSOD activities: nanoclusters > nanochains > nanoparticles
Mn3O4
(flower-like morphology, flake-like morphology, hexagonal plates, polyhedrons, cubes) [84]
CAT, SOD, GPxFlower-like Mn3O4 with a specific surface area of 97.7 m2/g exhibited the highest CAT, SOD, and GPx activities
Co3O4
(nanoplates, nanorods, nanocubes) [88]
CATCAT activity: Co3O4 nanoplates > Co3O4 nanorods > Co3O4 nanocubes
CompositionPVP–PtCuNCs [51]CAT, SODPVP–PtCuNCs exhibited 10-fold higher SOD-like activity and 4-fold higher CAT-like activity than PVP–PtNCs
Au core/Ce shell-based nanozyme [21]SODCan work under extreme conditions: pH (2–11) and temperatures up to 90 °C
CeO2 nanoparticles immobilized on 2D graphdiynes [101]CAT4.2-fold greater rate constant in H2O2 decomposition than CeO2 nanoparticles
Prevents CeO2 aggregation
Surface modificationAuNP capped with N-acetylcysteine [107]SODHigher SOD-like activity than free N-acetylcysteine and gold nanoparticles
PEG-coated Ce-Bi@DMSN [108]CATPEG-coated surface endows the nanocomposite with high hydrophilicity
MOFGrafting PhSeBr to a Zr(IV)-based UiO-66-NH2 framework [125]GPxPhSeBr acted as a donator, while the Zr(IV)-based UiO-66-NH2 framework with high surface area and uniform porosity provided more catalytic active centers, resulting in a high enzyme-like activity
PZIF67-AT [126]SODSimultaneously increased SOD activity and suppressed CAT and GPx activity. Effectively accumulated H2O2 for Fenton reaction-based chemodynamic therapy
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Thao, N.T.M.; Do, H.D.K.; Nam, N.N.; Tran, N.K.S.; Dan, T.T.; Trinh, K.T.L. Antioxidant Nanozymes: Mechanisms, Activity Manipulation, and Applications. Micromachines 2023, 14, 1017. https://doi.org/10.3390/mi14051017

AMA Style

Thao NTM, Do HDK, Nam NN, Tran NKS, Dan TT, Trinh KTL. Antioxidant Nanozymes: Mechanisms, Activity Manipulation, and Applications. Micromachines. 2023; 14(5):1017. https://doi.org/10.3390/mi14051017

Chicago/Turabian Style

Thao, Nguyen Thi My, Hoang Dang Khoa Do, Nguyen Nhat Nam, Nguyen Khoi Song Tran, Thach Thi Dan, and Kieu The Loan Trinh. 2023. "Antioxidant Nanozymes: Mechanisms, Activity Manipulation, and Applications" Micromachines 14, no. 5: 1017. https://doi.org/10.3390/mi14051017

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

Article Metrics

Back to TopTop