Open Access Article
This Open Access Article is licensed under a Creative Commons Attribution-Non Commercial 3.0 Unported Licence

Scavenging of reactive oxygen and nitrogen species using nanoparticles and their applications in disease management

Pranita Rananawarea, Sanjay Singh*b and Varsha P. Brahmkhatri*ac
aCentre for Nano and Material Sciences, Jain University, Jain Global Campus, Bengaluru, Karnataka 562112, India. E-mail: brahmkhatri.varsha@gmail.com; varsha.b@cmrit.ac.in
bDBT-National Institute of Animal Biotechnology (NIAB), Opposite Journalist Colony, Near Gowlidoddy, Extended Q-City Road, Gachibowli, Hyderabad, Telangana 500032, India. E-mail: sanjay@niab.org.in
cDepartment of Chemistry, Centre of Excellence in Materials & Sensors, CMR Institute of Technology, Bengaluru 560037, India

Received 21st July 2025 , Accepted 5th November 2025

First published on 4th December 2025


Abstract

Nanomaterials constitute a new trend of disease management that is associated with advanced nanotechnology and bioengineered materials, presenting new solutions for various diseases that were previously problematic to handle with traditional chemical drugs or natural materials. Due to their high surface area, charge, variable size, and other properties, nanomaterials have been broadly used to manage several diseases. Specifically, nanomaterials have appeared with a significant ability to act as RONS scavengers for treatment and disease management. This is a result of their versatility in various applications, controlled release, enhanced reactivity, and unique biochemical properties. Recently, specific nanomaterials for treatment and disease management have been effectively developed into clinical tests. This review article focuses on the different types of nanomaterials that are effective for RONS scavenging and are used for different biomedical applications associated with excessive RONS generation. Nanoparticle-based systems have gained significant attention in recent years for their potential applications in scavenging reactive oxygen and nitrogen species (RONS) as part of disease management strategies. These nanoparticles can be designed to enhance the delivery, stability, and efficacy of antioxidants or other scavenging agents. The current review article provides a complete overview of the anti-inflammatory nature and use of nanoparticle systems by examining the molecular and pathological mechanisms of oxidative stress and the function of this stress in both cell and tissue damage. However, it is important to consider the biocompatibility, stability, and potential toxicity of these nanoparticle systems for therapeutic applications. Additionally, targeted delivery and controlled release mechanisms can enhance their efficacy in scavenging RONS at specific disease sites. RONS play a dual role in biological systems—they are essential for various physiological processes, such as cell signalling and host defence, but their overproduction can lead to oxidative and nitrosative stress, contributing to the development and progression of several diseases. Managing RONS is a key aspect of disease prevention and treatment. This article focuses on the use of nanomaterials for the treatment of various cancers, and in other areas such as tissue engineering, wound healing, osteoclast genesis, inflammation, and neurodegenerative disorders, such as Parkinson's and Alzheimer's disease, through RONS scavenging.


1. Introduction

Oxidative stress disrupts the balanced production of free radicals and antioxidants, leading to neurological and other pathologies. Oxidative-mediated reactions are essential to many basic biological activities, including lipid formation, metal metabolism, phagocytosis of foreign particles, inflammation, immunology, and chemical biosynthesis. Exposure to harmful substances disturbs the subtle equilibrium between reactive oxygen species (ROS) and antioxidants.1–4 ROS is an extremely reactive compound formed from oxygen (O2), the byproduct of the usual oxygen metabolism. ROS comprises oxygen radicals (superoxide and hydroxyl) and a few non-radical derivatives of molecular oxygen with no unpaired electron like H2O2 and lipid peroxide, for example, superoxide (˙O2),5 peroxyl (ROO˙),6 hydroxyl (˙OH),7 alkoxy (RO˙),8 nitric oxide (˙NO),9 and nitrogen dioxide (NO2˙).10

Reactive nitrogen species (RNS) are derived from the interactions of biologically produced free radicals, particularly nitric oxide (˙NO), with other molecules. Enzymes produce ˙NO for assistance in cell signalling. NO plays a significant part in physiology, giving a potential route for several diseases from the perspective of RNS reactions.11 Reactions of ˙NO with various biomolecules lead to oxidation, nitration, nitrosation, and nitrosylation; that is, the addition of O2, NO2, and NO, respectively. As a well-known signalling molecule, NO exhibits its effects through various mechanisms, including reversible covalent binding and nitrosylation. One of the critical targets for nitrosylation is the ferrous ion (Fe2+) in soluble guanylate cyclase (sGC). Nitric oxide (NO) is the prime source of RNS in biological systems.12 At low levels, NO combines with oxygen, other free radicals, and transition-metal centres in proteins, such as iron and copper. NO is known to react with iron centres, but studies on its reactions with proteins that contain copper are still in their beginnings. The reaction of NO with the copper in the cytochrome c oxidase's oxygen-binding site is a crucial and biologically significant process. The quick reactivity of NO with free radicals is one of the main processes in developing RNS. The interaction with superoxide (˙O2) to create peroxynitrite (ONOO) is the most studied and accepted.12 This process was thought to be an extraordinary method of scavenging and neutralising O2 because nitrate is biochemically inactive in mammalian cells. A new viewpoint surfaced as research on this connection progressed: ONOO reacts with most biomolecules and mediates cytotoxicity independent of ˙NO or ˙O2.

Nanomaterials represent a new trend of disease management associated with advanced nanotechnology and bioengineered materials, presenting new solutions for diseases that were previously problematic to handle using traditional chemical drugs or natural materials.13,14 Nanomaterials have found widespread application in treating various diseases because of their large surface area, charge, variable size, and other characteristics.15–18 Specifically, nanomaterials have appeared that have a significant ability to act as RONS scavengers for treatment and disease management. This results from their versatility in various applications, controlled release behaviour, enhanced reactivity, and unique biochemical properties.19–23 Recently, specific nanomaterials for treatment and disease management have been effectively developed into clinical tests.

The current review article focuses on the diverse nanomaterials, including metal nanoparticles and surface-modified and functionalized nanoparticle systems that show effectiveness in RONS scavenging, and various approaches in disease management associated with excessive RONS generation. The design of nanoparticles and their scavenging strategies will be discussed, focusing on managing cancer and other illnesses such as diabetes, osteoclastogenesis, inflammation, and neurological conditions such as Alzheimer's and Parkinson's, and areas such as tissue engineering and wound healing. Fig. 1 describes various RONS entities and their chemical structures.


image file: d5ra05248c-f1.tif
Fig. 1 Chemical structure, composition, and nomenclature of various ROS and RNS. Key ROS include hydrogen peroxide, hydroxyl ion, peroxide, hydroxyl radical, and superoxide anion. Each of these species has distinct chemical structures and characteristics that are critical for their behaviour in biochemical processes. RNS species such as nitrogen oxide, nitryl chloride, nitrous acid, nitric oxide, and dinitrogen oxide emphasise their unique structural compositions and relevance in biological systems. Understanding these compounds is vital for evaluating their roles in oxidative stress, signalling pathways, and various physiological and pathological processes in living organisms.

Since organisms have metabolic defences to counteract their oxidative effects, ROS are considered the main source of tissue damage. The ratio of ROS generation to antioxidant activity frequently tips in favour of ROS in biological systems, resulting in persistently low levels of oxidative damage. Furthermore, ROS are advantageous for several physiological functions, including wound healing, tissue repair, and pathogen defence.24 Mitochondria and cytochrome P450 enzymes (CYP) represent the primary sources of cytosol ROS. ROS comprises oxygen molecules that can either accept or donate a free electron. Therefore, they are unstable and reactive with other molecules. This reaction may produce even more reactive species. The first stage in the complex chain that produces ROS is the reduction of one electron of the oxygen molecule. This may result in the generation of unstable superoxide, which quickly undergoes further reduction to hydrogen peroxide, a reaction that is accelerated and catalysed by superoxide dismutase (SOD).25 Both endogenous and exogenous substances may produce ROS; possible endogenous sources include peroxisomes, mitochondria, inflammatory cell activation, and cytochrome P450 metabolism.26

RNS have been extensively discussed in relation to signal transduction and cellular damage. RNS function as regulatory mediators in signalling processes at low concentrations, but they can be harmful to living organisms by deactivating vital cellular components at moderate or high quantities.

It is evident that reactive species can transition from beneficial to detrimental effects based on their concentration; however, the specific threshold at which this transition happens remains unclear. This complexity is attributed to the fact that cellular oxidants interact with a variety of targets and exhibit different behaviors across various cell types.26 The impact of these species is influenced by the quantity produced, their duration of existence, and their sites of formation within the cell.27 Both reactive oxygen species (ROS) and reactive nitrogen species (RNS) are essential for immune responses and cellular signalling, yet they also pose risks.28

Dendritic cells, neutrophils, and macrophages all react to inflammatory stimuli to produce RONS. Highly reactive RONS are those that have unstable bonds or unpaired valence electrons. Enhancing the immune response while reducing tissue damage requires proper RONS control.

ROS are essential for preserving homeostasis, promoting cell signalling, controlling metabolism, and supporting memory formation through DNA methylation, even though they can contribute to oxidative damage in several diseases.29 Enzymes from a variety of sources, such as the cytoplasmic membrane, mitochondrial respiratory chain enzyme complexes, and different organelles, create ROS in mammalian cells. During ATP biosynthesis in mitochondria, where electron and proton transport eventually lower molecular oxygen levels, ROS are produced. ROS, formed by the ascorbate system, can consecutively produce HO˙, H2O2, and O2˙. In contrast to gamma radiolysis, the reaction can occasionally be interrupted in these systems, but it always proceeds as long as reagents are present.30

The body uses the amino acid L-arginine to make nitric oxide (NO˙), which is a major source of RONS that the body uses to fight bacteria. An enzyme known as nitric oxide synthase (NOS) controls this process by converting L-arginine into L-citrulline and producing NO˙. NOS comes in three varieties: inducible, endothelial, and neuronal. While inducible NOS (iNOS) can function even in low calcium levels and is usually activated by infections, inflammation, or tissue damage, neuronal NOS (nNOS) and endothelial NOS (eNOS) require calcium and calmodulin to function. Depending on the type of cell, different NOS types create different amounts of NO˙. iNOS produces significant amounts of NO˙ during immunological responses, nNOS facilitates nerve cell communication, and eNOS promotes vascular health by assisting blood vessels in relaxing and regulating blood pressure.31

2. Origin of RONS in cells and tissues

RONS can be produced by metabolic activities in the endoplasmic reticulum, mitochondria, and other cell components.32 The electron transport chain (ETC) in mitochondria is the main producer of RONS. In the ETC, most cells and tissues convert a tiny portion of molecular oxygen into RONS. Nitric oxide synthase, xanthine oxidase, and NADPH oxidase are only a few of the cellular enzymes that can be affected by ionising radiation, and RONS are produced by other mechanisms, such as the respiratory burst of activated phagocytes. The generation of RONS, a natural consequence of aerobic metabolism, is vital for sustaining tissue oxygen homeostasis. In the cellular environment, oxidative stress will increase when oxygen homeostasis is not maintained. Hydrogen peroxide, superoxide, and hydroxyl radicals are typical metabolic byproducts continuously produced by the mitochondria in growing cells. RONS play a significant role in conditions such as stroke, hypertension, diabetes, neurodegenerative disease, cancer, and inflammation.

Additionally, RONS are essential for several physiological processes, including controlling oxidative stress and ensuring that vascular cells are working properly. RONS can carry out the process by eliciting a potent immunological response, controlling skeletal muscle's glucose absorption, and potentially serving as signalling molecules. They significantly regulate inflammatory responses and responses to growth factor stimulation. They also control apoptosis, migration, cytoskeletal regulation, differentiation, growth, contraction, and proliferation.33 Apoptosis is the process by which a cell dies by destroying itself. It is a process involving morphological and biochemical characteristics.34 Apoptosis is used in early development to eliminate unwanted cells. RONS production can be induced in mitochondria and endoplasmic reticulum in the form of signalling pathways, as well as NADPH oxidase and cellular-metabolising enzymes, and due to the environmental pollutants, radiation, and chemotherapies. Apoptosis is activated by extrinsic and intrinsic signals.35

2.1 Cellular pathways

2.1.1 Intracellular pathways of RONS generation. The intracellular pathway is the chain of molecules that relay signals inside the cells, known as intracellular signal transduction pathways. For biological activity, RONS are comprised of radicals such as singlet oxygen (1O2), hypochlorous acid (HOCl), O2˙, OH), ONOO, and H2O2.36 H2O2, O2˙, and OH are crucial chemical intermediaries in signal transduction in intracellular pathways during cell proliferation.37 The half-life of O2˙ in the tissue's spontaneous dismutation is catalysed by superoxide dismutase (SOD), making it a signalling molecule. Adding low concentrations of H2O2 or O2˙ stimulates various preparations, including fibroblasts, smooth muscle cells, aortic endothelial cells,38 amnion cells, and prostate cancer cells. It was demonstrated that ionising radiation, hyperoxia, and chemical compounds produced by RONS and H2O2 help cell differentiation.

H2O2 can readily pass through cell membranes because it dissolves in lipids. The extremely reactive OH˙ can be formed when Fe2+ is present.39 When O2˙ combines with NO˙, ONOO is produced. Myeloperoxidase, an enzyme primarily found in neutrophils, produces hypochlorite, a potent antibacterial agent, as shown in eqn (1).40

 
H2O2 + CI → H2O + OCI (1)

2.1.2 Extracellular pathways of RONS generation. A complicated network of multidomain macromolecules arranged according to cell or tissue type makes up the extracellular matrix. The signal transduction pathway, including protein kinase, plays a crucial role in controlling mitosis, meiosis, and postmitotic processes in differentiated cells. The extracellular matrix (ECM) induces ROS, with proteins acting as resistors to maintain ROS homeostasis. The NADPH oxidase family, consisting of seven isoforms, is involved in regulating cell apoptosis. Therapeutic targets for ECM are being explored to control ROS production in tumours and other pathological diseases.41

2.2 Cellular organelles in RONS generation

2.2.1 Mitochondria. Mitochondria are membrane-bound organelles that convert nutrients into chemical energy. They are crucial for apoptosis and are responsible for oxidative stress and respiration, which build up over time and contribute to ageing and degenerative diseases such as cancer. Oxygen is consumed by submitochondrial particles, which convert NADH into H2O2. The oxidation of a protein thiol by H2O2 indicates that mitochondria's function involves redox reactions as dependent signalling.34

When the mitochondrial complex fails, cysteine proteins move to the cytosol and chloroplast, transmitting signals and indicating the state of the mitochondria. New research shows that mitochondria produce H2O2 through enzymatic systems and respiration, highlighting the importance of ROS as signalling molecules. Mitochondrial-mediated oxidative stress may have carcinogenic effects when tumour suppressor mechanisms fail, as ROS and Ca2+ ions work together.35,42,43 However, apoplastic polyamines cause the plasma membrane's Ca2+-ATPases to become active, which removes Ca2+ from the cell. Consequently, polyamines and ROS interact in Ca2+-dependent signalling. Many Ca2+-dependent protein kinases are activated when the cytoplasmic Ca2+ concentration rises. Ca2+ signalling depends on ROS, which attaches to lipid membranes and activates Ca2+. In conjunction with hormones, Ca2+, and electrical impulses, ROS should be viewed as helpful messengers that promote oxidative signalling, non-photochemical quenching of systemic acquired resistance, and acclimatisation.35 Mitochondrial oxidative phosphorylation produces most ATP at the centre of cellular energy metabolism. This process releases electrons from reducing substrates and transfers them to O2, generating an electrochemical gradient that drives the synthesis of ATP. One unpaired electron is produced by oxygen reduction during oxidative phosphorylation, one electron at a time, as represented in eqn (2).

 
O2 → O2˙ → H2O2 → ˙OH → H2O (2)

The most powerful generator of this oxygen radical in mitochondria is superoxide, which is highly potent. Consequently, mitochondria are the primary generator of ROS in mammalian cells.44

2.2.2 Endoplasmic reticulum. The endoplasmic reticulum (ER) in the secretory pathway is an organelle responsible for regulating calcium (Ca2+) flux and lipid production. It plays a key role in protein synthesis, folding, post-translational modifications, and trafficking. Proper regulation of ER processes is essential to ensure optimal protein folding, which is crucial for cellular survival, function, and homeostasis. When ER stress occurs—often due to an overloaded ER—misfolded proteins accumulate within the ER lumen. Conditions such as hypoxia, oxidative stress, viral infections, and calcium depletion can disrupt cellular homeostasis and trigger ER stress. The Unfolded Protein Response (UPR) is a survival mechanism that activates in response to ER stress, helping cells restore balance. The ER-mediated apoptotic pathway also involves the caspase-12/caspase-4 pathway, which can be activated through various mechanisms. In mice, procaspase-12 is anchored to the cytosolic side of the ER membrane, where it can be cleaved by the Ca2+-dependent protease m-calpain.45–48 By moving cytosolic caspase-7 to the ER membrane, caspase-12 can be activated and cleaved in situations of severe ER stress. Under stress, the pro-apoptotic Bcl-2 family protein Bcl may prevent caspase-12 activation by moving to the ER membrane. The ratio of oxidised-to-reduced glutathione (GSSG/GSH) is larger in the ER than in the cytosolic reducing environment. The primary chaperone in the ER that aids in oxidative protein folding is protein disulfide isomerase (PDI). As an electron acceptor during this process, flavin adenine dinucleotide (FAD) transforms into FADH2 and then oxidises with oxygen to produce H2O2. The H2O2 generated may act as a signalling molecule or pose a potential risk. When GSH levels decrease, disulfide bonds in misfolded proteins are reduced, potentially increasing reactive oxygen species (ROS). In response, the UPR promotes cell survival under mild oxidative stress, while severe oxidative stress leads to ER-mediated apoptosis.34,35

3. Classification of ROS

3.1 The superoxide anion (SOD)

An anion is created when oxygen is reduced by one electron; at low pH, this anion is protonated to create the perhydroxyl radical. Single-electron reduction is a more efficient method for reducing oxygen than full reduction, as it can prevent oxidative damage. SOD converts free radical superoxide into peroxide, making it vulnerable to catalase or GPX reactions. During aerobic respiration, mitochondria produce small amounts of superoxide, which releases iron from storage sites and reacts with H2O2 to generate hydroxyl radicals.49,50 SOD converts superoxide into H2O2 and molecular oxygen, as shown in eqn (3).
 
˙O2 + O2 + 2H + SOD H2O2 + O2. (3)

SOD is found in humans in four forms: extracellular SOD (EC-SOD), mitochondrial Mn-SOD, and cytosolic Cu/Zn-SOD. It protects against superoxide inactivation of dehydratases such as fumarases A and B, aconitase, dihydroxy acid dehydratase, and 6-phosphogluconate dehydratase, through dismutation at the active site.50

3.2 Hydrogen peroxide

Molecular oxygen undergoes a two-electron reduction to produce H2O2. The conjugate base HOO is a powerful nucleophile with a high pKa value, meaning it has little bearing on physiological pH. H2O2 is not ideal for thermodynamically preferential biological conditions, but can function as an oxidant if metal ions catalyse the reaction. It is involved in two-electron transfer processes and is more oxidising than hypochlorous acid and peroxynitrite. However, due to its high activation energy barrier, it reacts weakly with biological molecules. Thus, the greatest oxidising power of H2O2 originates indirectly from the Fenton and Haber–Weiss reaction, which is metal-catalysed and converts H2O2 into HO˙ radicals.49 Glutathione peroxidase is an enzyme that uses GSH to reduce hydroperoxides such as ROOH and H2O2, protecting mammalian cells from oxidative damage, as shown in eqn (4).
 
ROOH + 2GSH GPX ROH + GSSG + H2O (4)

The amount of each of the five glutathione peroxidase (GPX) isoenzymes found in mammals varies depending on the kind of tissue. Glutathione is utilised by cytosolic and mitochondrial GPX1, which reduces H2O2 and fatty acid hydroperoxides. GPX1 and PHGPX, present in most tissues, directly reduce hydroperoxides in cholesterol, phospholipids, and fatty acids in oxidised membranes and lipoproteins.51 GPX4 is primarily found in testes and renal epithelial cells, while GPX1 is found in the kidney, liver, and erythrocytes. Most tissues have modest levels of cytosolic GPX2 and extracellular GPX3, except for the kidney and gastrointestinal tract. A newly identified component, GPX5, is expressed exclusively in the mouse epididymis and is independent of selenium.52 GPX, a substrate of catalase, efficiently reacts with lipids and organic hydroperoxides, and its primary defence against mild oxidative damage is the glutathione redox cycle, making catalase crucial for avoiding severe oxidative stress. GPX has long been thought to be the primary antioxidant enzyme in charge of detoxifying H2O2 in both human and animal erythrocytes, since catalase has a lower affinity for H2O2 than GPX.

3.3 Hydroxyl radical (˙OH)

˙OH is a highly reactive radical and readily attacks molecules, particularly on sulfur or rings (aromatic groups). They behave as potent electrophiles as they readily add to double bonds, but also readily abstract electrons or hydrogen atoms from other molecules.49 Oxygen (O2) picks up two electrons to form hydrogen peroxide (H2O2). Its conjugate base (HOO) is a powerful nucleophile but insignificant at a neutral pH. H2O2 can be an oxidant or reductant, itself, but in living cells it generally plays the role of an oxidant – particularly when metal ions (like iron) assist this process. In general, hydrogen peroxide is a powerful oxidising agent, even stronger than hypochlorous acid and peroxynitrite in common biological scenarios.

Because of its high activation energy, H2O2 is not very reactive with biological molecules. However, when metals are present, they acquire indirect oxidative power through Fenton and Haber–Weiss reactions, which result in the production of HO˙ radicals. Protein residues containing sulfur are gradually oxidised to produce sulfenic acid (RSOH), which can then react to produce disulfides or sulfonic acid (RSO2H). With rate constants of about 107 L mol−1 s−1, thiol proteins, such as glutathione peroxidases and peroxiredoxins, react with H2O2 substantially quicker than other proteins. Given its intracellular concentration, the oxidation of pyruvate to acetate and CO2 by H2O2 is also a significant process.53

3.4 Peroxyl radical (HO2)

When carbon-centred radicals are combined with molecular oxygen, peroxyl radicals (HO2) are produced as secondary species.

Peroxisomes, responsible for producing peroxide, play a crucial role in the oxidation of fatty acids and are a significant source of total H2O2 in cells. They are essential for various metabolic processes in mammals, including fatty acid oxidation, ether phospholipid creation, glyoxylate metabolism, amino acid breakdown, polyamine oxidation, and the pentose phosphate pathway oxidative phase.

Peroxisome enzymes, primarily flavoproteins, produce H2O2 as a byproduct of their catalytic cycle, including Acyl-CoA oxidases, D-aspartate oxidases, urate oxidases, L-pipecolic acid oxidases, D-amino acid oxidases, L-α-hydroxy acid oxidases, polyamine oxidases, and xanthine oxidases.54 Reactive oxygen species such as hydroxyl radicals, superoxide, hydrogen peroxide, peroxynitrite, and nitric oxide radicals are produced by peroxisomes, which are involved in many metabolic processes and have a high concentration of ROS-generating enzymes. Catalase is another peroxisomal enzyme that breaks down the H2O2 produced in these organelles. These oxidases create H2O2, which peroxisomal catalase uses to oxidise several other substrates through “peroxidative” processes. Particularly important are these oxidative processes in kidney and liver cells, where peroxisomes detoxify various harmful chemicals (such as ethanol) throughout the bloodstream.55

4. Classification of RNS

4.1 Peroxynitrite

Peroxynitrite (ONOO) is chemically unstable under physiological conditions, which results in the formation of nitrate through isomerisation. ONOO is formed from the addition reaction of ˙NO with superoxide (O2˙−) at a precise rate.56 Nitrosoperoxycarbonate (ONOOCO2) is the product of the reaction between ONOO and CO2. However, this reaction is minimal due to its slower cleavage rate. Because it impacts intermediate retention, the solvent's viscosity affects the rates at which ONOOCO2 and ONOOH decompose. EPR spin trapping confirmed that ONOO, a potent nucleophile, causes β-scission of carbonyls. Peroxynitrite (O2NOO), O2NOOH, and nitrite (NO2) are the products of the reaction between ONOO and ONOOH at neutral pH.57 ONOO can drive a unique electron transfer process to create free radicals, enabling two-electron oxidations in aromatic compounds such as the antioxidant K-tocopherol, without forming radicals from ONOO or the reactant itself. ONOO also reacts with both protein-bound and low-molecular-weight thiols, leading to thiol oxidation via thiol radical formation. The interaction between ONOO and glutathione is a significant defence mechanism against ONOO-induced oxidative damage, given high intracellular thiol concentrations.58

4.2 Nitrogen trioxide radical (˙NO3)

Strong oxidants such as the ˙NO3 radical react with unsaturated gas-phase organic molecules in the troposphere. Its reactivity is achieved either by hydrogen atom abstraction or addition to a double bond.59 Numerous heme proteins and metalloenzymes are bound by the azide anion, which prevents them from functioning. By blocking ATP hydrolysis and cytochrome c oxidase, it also significantly inhibits the mitochondrial electron transport chain.60

4.3 Nitric oxide (NO)

A family of NADPH-dependent enzymes called NO synthase produces NO during the breakdown of arginine to citrulline.61 The uncharged lipophilic molecule is more reactive due to single unpaired electrons reacting further with species such as oxygen, superoxide radicals, and glutathione. The other reactive intermediates of NO, which are not reactive free radicals, affect protein function and the overall functioning of the organism. The transient intermediates can cause nitrosative damage in biomolecules.62

As a result, NO may have antioxidant or oxidative effects. Strong oxidants can be produced by NO, a neurotransmitter and blood pressure regulator, under pathological conditions. Excessive production of NO is linked to ischemia-reperfusion, neurodegeneration, and chronic inflammatory illnesses, for example, inflammatory bowel disease and rheumatoid arthritis. Cytochrome c oxidase is inhibited after generating NO. Oxidative stress, respiration, mitochondrial biogenesis, and other functions can all be affected by the elevated production of ROS and RNS due to the presence of NO in mitochondria.61

4.4 Nitrogen dioxide (NO2˙)

The reaction of peroxyl radicals and NO results in the formation of nitrogen dioxide (NO2˙). The NO2˙ is responsible for lipid peroxidation initiation for the production of free radicals and also for the oxidation of ascorbic acid.63 When tyrosine is oxidised to 3-nitrotyrosine, the NO2˙ radical, a highly reactive oxidant and NO breakdown product, is produced. Additionally, NO2˙ is a substrate for lactoperoxidase and mammalian peroxidase, which undergoes oxidation catalysed by peroxidase to become NO2˙. This demonstrates a different mechanism for the synthesis of 3-nitrotyrosine as well as a pathway linked to enhanced NO generation and cytotoxicity.62

5. RONS-induced diseases

The imbalance in the production of RONS and antioxidant defences is connected to neurological pathologies such as Alzheimer's disease. Mitochondrial dysfunction leads to electron leakage in the respiratory chain, accumulating RONS.

More H2O2 is produced as a result of the uneven additional activity of the enzymes monoamine oxidase B, which breaks down various substances in the brain, and SOD, which catalyses the dismutation or partitioning process alternately. In this process, the combination of superoxide and H2O2 alters iron homeostasis and produces the most harmful hydroxyl radicals.64 The cell membranes may become weaker due to the altered membrane and lipid peroxidation levels, which can disrupt metabolism and cause an ion imbalance. The C-terminal portion of amyloid precursor (A4CT) and beta/A4 protein are both affected by the pro-aggregating impact of RONS. The ageing process is associated with an increase in oxidative stress. An important factor in the aetiology of Alzheimer's disease is oxidative stress, which leads to the death of cells and aggregation. The brain is particularly vulnerable to oxidative stress because it consumes a lot of oxygen for energy production, has an abundance of unsaturated fatty acids, and has limited antioxidant activity.

In Alzheimer's disease (AD), oxidative damage happens before the development of plaques and neurofibrillary tangles. The neurotoxic qualities of ROS aid AD development. The progressive death of neurons is a hallmark of age-related illnesses known as neurodegenerative disorders (NDs), which impact the structure and function of the brain. Multiple sclerosis (MS), Parkinson's disease, Alzheimer's disease, and amyotrophic lateral sclerosis (ALS) are among the NDs that are becoming more prevalent, especially in people over 65 years.65,66

Despite the variations in aetiology and clinical manifestations, ND shares a common feature in cellular events, which can be attributed to pathophysiological mechanisms that induce neurodegeneration at different phases of the neurodegenerative cascade. This comprises mitochondrial failure, proteolytic stress, oxidative stress, neuroinflammation, and excitotoxicity.67–70

The free radicals create superoxide radicals by reducing the amount of molecular oxygen in the water, accumulating electrons to generate H2O2. H2O2 is not a free radical and is less reactive, but it can be considered an oxidant. In the biological system, the reactivity depends on two properties: one is diffusing long distances and crossing membranes, and the other is reacting with transition metals and homolytic cleavage with a highly reactive hydroxyl radical. Hence, ROS impacts the tissues and organs, particularly the brain, which is made of oxidised lipids. In humans, ROS can be generated by both normal and abnormal processes, such as ageing, cancer, joint problems, asthma, and atherosclerosis.

RONS is linked to several illnesses, including cancer, neurological diseases, acute and chronic kidney disease (CKD), biliary diseases, macular degeneration (MD), and cardiovascular diseases (CVDs). Age-related mortality and morbidity are mostly caused by CVDs, with atherosclerosis being a major contributing factor. Studies show that as people age, their heart's capacity to withstand oxidative stress decreases due to decreased antioxidant enzyme concentrations, resulting in cardiovascular changes.1 Chronic kidney disease (CKD) is accelerated by factors such as hypertension, inflammation, endothelial dysfunction, glomerular injury, renal ischemia, and OS. CKD patients produce reactive oxygen species (ROS) due to activated leukocytes and monocytes secreting more MPO and NADPH oxidase. These ROS contribute to carcinogenesis, damaging target cells or attracting more inflammatory cells. High levels of ROS in cancer cells can be attributed to increased metabolic activity, oncogene activity, increased oxidase activity, mitochondrial dysfunction, cyclooxygenases, or immune cell infiltration. ROS activate antioxidant pathways in cancer cells, aiding in carcinogenesis and malignant development. Oxidative stress produces ROS, damaging cellular proteins, lipids, and DNA, causing programmed cell death (PCD), which can be used to help treat cancer cells. Apoptotic cell death is a crucial cancer treatment strategy due to ROS damage. ROS play a crucial role in preventing neuronal degeneration and protecting the brain from aging and degenerative diseases. Diabetes leads to metabolic disorders, causing issues in blood vessels due to an imbalance between free radicals and antioxidants. Type 2 diabetes patients have higher levels of oxidative stress biomarkers but lower levels of antioxidants, indicating older patients experience higher levels of oxidative stress.71–77 Endothelial cells and RPE cells share similar features, but their interaction enhances their proangiogenic potential, including migration and proliferation. TNF-α controls VEGF expression in RPE cells by activating β-catenin in a ROS-dependent manner. ROS influences the autophosphorylation and dimerisation of VEGF-stimulated VEGF receptor 2. Additionally, VEGF stimulates the generation of ROS by activating NOX in endothelial cells.78 The formation of the HIF E3 ubiquitin ligase complex requires suppressing the expression of the scaffolding protein Cullin-2.79 ROS and factor-inhibiting HIF-1α (FIH) inhibit prolyl hydroxylase enzymes by lowering Fe2+ availability.80 High hyperglycemia causes endothelial cells to die, which is caused by overexpressing iNOS in RPE cells, which activates the PKR-like endoplasmic reticulum kinase (PERK) pathway.81

The main mechanisms underlying biological responses mediated by RONS can be summarized as interrelated processes that produce antimicrobial and anticancer effects (Fig. 2). Increased levels of RONS cause oxidative stress, leading to mitochondrial dysfunction, protein oxidation, and lipid peroxidation. DNA damage, such as strand breaks and base modifications, is brought on by this stress and compromises cellular replication and genomic integrity. Apoptotic signalling pathways, which include caspase activation, mitochondrial cytochrome c release, and programmed cell death, are triggered by chronic damage and a loss of repair ability. Together, these pathways show how nitrosative and oxidative stress control cellular destiny, offering a mechanistic foundation for treatment strategies.


image file: d5ra05248c-f2.tif
Fig. 2 Excessive RONS causes stress in cells, which damages DNA, proteins, and fats. It also damages mitochondria, which are the cell's power plants; this activates apoptosis, i.e., cell death. These downstream events contribute to cellular dysfunction, ultimately leading to antimicrobial or anticancer outcomes.

Nanoparticles (NPs) are a promising alternative for in vivo scavenging due to their ability to control and improve antioxidant biodistribution and specificity. NPs may possess inherent antioxidant characteristics and exhibit intrinsic catalytic characteristics due to their large surface areas. They can mimic cellular enzymes' natural scavenging of RONS, allowing direct reactions with RONS.82,83 Biomedical nanotechnology has significantly advanced antioxidant therapy by combining material science with RONS chemistry and biology. Advancements in nanochemistry and nanomanufacturing have enabled the development of nanomaterials such as Se, C, Pt, Ce, Cu, and polymers with exceptional RONS-scavenging properties. These nanomaterials have been used to moderate inflammatory RONS responses, simplifying the treatment of RONS-related illnesses.84–90 This article provides an overview of nano-antioxidants’ application for anti-inflammatory purposes, explaining the molecular and pathophysiological mechanisms of oxidative stress and its role in tissue and cell destruction. It also discusses technological concerns about nanomedicine and RONS-based inflammation resolution, their significance for biomedical applications, and recent advancements in disease treatment. The article suggests that RONS scavenging is based on the efficiency and experiments of nanomedicine, and progress has been made towards different nanomaterials based on RONS scavenging for biomedical applications.

6. Nanomaterials for RONS targeting and scavenging, and biomedical applications

Oxidative stress and excessive production of RONS cause many types of diseases. Conventionally, synthetic, natural, low-molecular-weight compounds are used as RONS scavengers and potential therapeutic agents. However, these scavengers have limitations such as aqueous solubility, ease of metabolism, and selectivity. So, to improve therapeutic effects, NPs have been proposed as RONS scavengers. There are several types of NPs, such as quantum dots and gold, manganese, ceria, and platinum NPs, which can be used for RONS targeting and scavenging for disease management.91

G(TM)PPSP, with a size of 214.0 ± 5.0 nm, or gelatine NPs linked with platinum NPs (PtNPs), were developed by Tianxu et al. for breast cancer therapy and tumour microenvironment (TME) remodelling. Fig. 3A illustrates telmisartan (TM) loaded in gelatin NPs. Due to gelatine degradation in TME, paclitaxel (PTX) is bound to PtNPs via a dual redox-responsive di-selenide bond, and MMP-2 mediates TM release. Because of this, the di-selenide linkage fracture brought on by ROS or glutathione releases intracellular PTX. It has been claimed that PtNPs can stop SOD and H2O2. Using the MTT test on T1 cells, the fluorescence emission was measured at 320 nm to assess the H2O2 cytotoxicity of free drugs and NPs. Following a 24-hours incubation period with various formulations, a negative association was seen between the concentration of PTX and the vitality of the cells. Significant statistical differences exist between drug-loaded NPs and free drugs in terms of their ability to impede 4T1 tumor cells. In tumour tissues, the NPs react to MMP-2 by releasing telmisartan and producing PEG-PtNPs-SeSe-PTX (PPSP). Telmisartan was successful in TME remodelling by suppressing TGF-β. When PPSP is internalised by cancer cells, it scavenges ROS and produces PTX in a redox environment, which kills tumor cells. The ex vivo experimental findings demonstrate that this combination has a substantial anti-tumor impact and TME remodelling capacity.92 In order to investigate its potential as an anti-inflammatory treatment for peritonitis, Jinseong Kim et al. created a polymer/aptamer-integrated gold (Au) nanoconstruct that traps tumour necrosis factor-alpha (TNF-α) and scavenges reactive oxygen species (ROS). They designed a flexible DNA sequence that combined two aptamers (ATP-binding and TNF-α-binding) to build a functioning construct. Fig. 3B illustrates that thiol–gold interactions were used to immobilize these aptamers on the surface of Au nanoparticles (AuNPs). Through particular interactions with the adenosine and cis-diol groups of ATPS, the ATP-binding aptamer and pPBA were connected to produce the pPBA-coated nanoconstruct.93


image file: d5ra05248c-f3.tif
Fig. 3 (A) Schematic of the preparation of G(TM)PPSP.92 Copyright © 2013, American Chemical Society. (B) Au-Apt-ATP-pPBA nanoconstruct for the Fabrication Process. Reprinted with permission from,93 Copyright © 2021, American Chemical Society. (C) Mechanism of the Au-Apt-ATP-pPBA nanoconstruct and the inflammatory effect. Reprinted with permission from,93 Copyright © 2021, American Chemical Society. (D) Synthesis of the Au-PEG-TEMPO NPs for ROS scavenging and control of stem cell differentiation. Reprinted with permission from ref. 94 Copyright © 2017, American Chemical Society. (E) Confocal microscope analysis of intracellular ROS generation in macrophage cells, green fluorescence indicates ROS induced cleavage of fluorescein–HA conjugates from the surface of AuNPs.95 Copyright © 2009 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim.

This nanoconstruct forms a phenylboronic ester that is easily cleaved and known to scavenge ROS. In addition to scavenging excess ROS in the inflammatory region, the phenylboronic ester produced in the Au-Apt-ATP-pPBA also exposes the TNF-α-binding aptamer, allowing TNF-α to be captured. Consequently, the combined effects of TNF-α trapping and ROS scavenging suppress oxidative stress and inflammation. As seen in Fig. 3C, the ATP served as a linking molecule, making it possible for nanoconstruct preparation to be done quickly through particular interactions between ATP and Au-Apt and pPBA.93 Regenerative medicine could greatly benefit from controlling the growth of human mesenchymal stem cells (hMSCs), as shown in Fig. 3D. Combining 40 nm-sized gold nanoparticles (Au NPs) with 2,2,6,6-tetramethylpiperidine-N-oxyl (TEMPO) allows ROS to be scavenged while maintaining the advantageous effects of Au NPs. Researchers are examining how TEMPO-conjugated Au NPs (Au-PEG-TEMPO NPs), which are used to grow hMSCs, affect ROS scavenging osteogenic and adipogenic differentiation as well as proliferation. Human mesenchymal stem cells (hMSCs) are not harmed by Au-PEG-TEMPO nanoparticles (NPs), which also efficiently lower reactive oxygen species (ROS) levels caused by H2O2 exposure. At lower dosages, these NPs, which were produced by a seed-mediated method, exhibit better ROS scavenging capabilities than free TEMPO. They inhibit adipogenic differentiation in hMSCs while promoting osteogenic differentiation, in contrast to free TEMPO. Au-PEG-TEMPO NPs may be useful in cellular therapies since they can improve the regulation of stem cell proliferation and address dysfunctions associated with reactive oxygen species.94

Biomass-derived CDs have shown excellent bioimaging properties such as long-lasting fluorescence, low tendency to photobleach, and good bioconjugation ability without toxicity. Herein, green chili extract is used to synthesise CDs via microwave irradiation. CDs are shown to possess ROS scavenging ability with the potential to modulate gene expression related to ROS scavenging enzymes, thus enhancing wound healing by altering granulation tissue and microvessel formation in animal models. Several assays reveal that CDs reduce oxidative stress and mutagenicity in vivo, which assists in their easy elimination from biological systems. They also help modulate the inflammatory responses of wound healing, making them novel theragnostic probes for cell labelling.96

Innovative gold nanoparticle (AuNP)-based nanoprobes were created by Hyukjin Lee and associates to track intracellular reactive oxygen species (ROS) in living cells. Fluorescein-labelled hyaluronic acid (HA), a naturally occurring polysaccharide that acts as a cleavable substrate sensitive to ROS, was used to inactivate these nanoprobes. HA was immobilised on AuNPs using dopamine to improve the stability within cells and withstand glutathione. This technique was motivated by the sticky qualities of mussel proteins, namely the amino acid L-3,4-dihydroxy-L-phenylalanine (DOPA). As shown in Fig. 3E, the presence of ROS causes HA to be cleaved off the AuNP surface, resulting in a notable fluorescence recovery that permits accurate and quick ROS detection. Dopamine functionalization guarantees better stability of HA on the AuNPs than conventional thiol-based attachments. In addition to offering a useful method for visualising ROS, this novel approach demonstrates the potential of bioinspired materials for creating extremely sensitive and stable intracellular probes for use in bioimaging and diagnostics.95

Traumatic brain injury (TBI) is impacted by ROS and RNS, which are temporary species created by the development of nervous system disorders. So, Mu et al. synthesised ultrasmall fluorescent carbogenic nanozyme (CN), which has higher antioxidant activity and ultrahigh RNS selectivity, showing enzyme mimetic activity. The CN was prepared through microwave heating synthesis from lysine with ascorbic acid (AA), and the synthesised NPs had an average diameter of 2.7 nm. At high dosages, CN effectively eliminates hazardous peroxide and superoxide, and the carbogenic enzymes demonstrate renal clearance without causing adverse effects. It also shows a significant scavenging potential towards RNS such as NO and ONOO in vitro. Scavenging all types of RNS can restore neuron cells damaged by lipopolysaccharide or H2O2, according to in vitro research. Research using 2′,7′-dichlorodihydrofluorescein diacetate (DCFH-DA) labelling on N2a cells with and without enzyme suggests that, following enzyme treatment, the fluorescence from wounded cells is regained and highly effective elimination of ˙NO and ONOO is demonstrated.97

Yim et al. proposed antioxidant nanosheets, precisely 2D transition-metal dichalcogenide (TMD), as a treatment for sepsis instigated by RONS. Liquid-phase exfoliation was used to create these biocompatible TMD nanosheets (WS2, MoSe2, and WSe2), which had significant ROS and RNS scavenging activity, including NO, H2O2, and hydroxyl radicals. They reduce the release of inflammatory cytokines but do not affect anti-inflammatory cytokines. In CLP-induced bacteremia animal models, WS2 nanosheets significantly increased the overall survival rate of septic mice to 90%, demonstrating their effectiveness and biocompatibility.98

Zhong et al. synthesised dopamine melanin (DM) NPs for effective RONS scavenging, including peroxynitrite, superoxides, and hydroxyl radicals. DM NPs show low toxicity and can be effective as an anti-oxidative agent to treat osteoarthritis (OA), overcoming the limits of biocompatibility and cytotoxicity issues. Through the use of the DPPH assay, the authors investigated the efficacy of DM NPs in scavenging RONS radicals. They found that, at 80 µg mL−1, DM NPs suppressed DPPH fluorescence by 64%. Additionally, considerable sequestration of ONOO, superoxide, and hydroxyl radicals was analysed using test-tube methods for NBT reduction. According to this study, melanin nanoparticles effectively scavenge a variety of ROS and RNS. By scavenging ROS and RNS, DM NP prevents cartilage deterioration and the advancement of OA.99

Dowding et al. studied the activity of cerium NPs (CeO2 NPs) towards reduction in ONOO and RNS. CeO2 NPs show better activity due to merits including substantial oxygen storage capacity, cycling between the Ce4+ and Ce3+ redox states. CeO2 NPs scavenge RNS (nitric oxide, ˙NO) in ex vivo conditions. To study the reactivity of CeO2 NPs with ONOO, they followed the oxidation of 3′-(p-aminophenyl)fluorescein (APF) by fluorescence spectrometry in vitro. The scavenging of ONOO studied in the presence of cerium oxide NPs shows compelling, albeit preliminary, evidence that CeO2 NPs readily react with ONOO or one of the reactive oxidants and radicals, eventually resulting in the non-enzymatic breakdown of ONOO.100

Also, solid lipid nanoparticles (SLPs), downconversion (DCNPs), and upconversion nanoparticles have contributed significantly to RONS scavenging. In the work by, mixed PLGA-lipid nanoparticles that exhibit high protection activity from ROS are presented, enabling cell protection.101 Encapsulation of Idebenone (IDE) in PLGA NPs offers potential advantages for drug delivery but also presents technical and practical challenges. A combined in silico and in vitro approach was used to test the effects that the addition of lipid stabilisers DSPE-PEG2000 and Monoolein (MO) had on PLGA-IDE NPs. MO increased the NP stability and guaranteed prominent antioxidant activity. Both DNP-IDE formulations with high and low concentrations of MO were efficient against ROS scavenging activity, showing an increased rate of protection compared to IDE alone. Results have shown that it is possible to develop optimal, stable, nontoxic formulations for the treatment of diseases related to oxidative stress.

Researchers have developed UCNPs-MoS2 nanoassemblies as a sensitive probe for ROS monitoring, demonstrating low cytotoxicity and fast response time. The nanoassemblies were successfully used in a luminescence assay and bioimaging in zebrafish, making them a powerful imaging probe for various physiological and pathological processes.102 Table 1 summarises the clinical translation potential and new nanosystems such as SLPs, DCNPs, and UCNPs.

Table 1 Role of solid lipid nanoparticles (SLPs) and downconversion (DCNPs) and upconversion nanoparticles (UCNPs) in RONS scavenging
Nanoparticle type Composition Functional role Advantages Applications Ref.
Lipid-polymer nanoparticles poly(PMT-co-EGDM) ROS-scavenging in vitro and in vivo Protects neurons against oxidative damage Neuroprotection post spinal cord injury 103
Solid lipid nanoparticles (SLPs) Lipid-based nanocarriers Reactive oxygen/Nitrogen species (RONS) scavenging Significant contribution to oxidative stress reduction and cellular protection Antioxidant therapy, oxidative stress-related diseases 101
PLGA-lipid mixed nanoparticles PLGA + DSPE-PEG2000 + monoolein (MO) + idebenone (IDE) ROS scavenging and drug delivery MO improved nanoparticle stability and antioxidant efficiency; both low and high MO formulations showed higher protection rates than IDE alone Treatment of oxidative stress-related diseases 101
Downconversion nanoparticles (DCNPs) Lanthanide-based NPs (general) RONS scavenging Reported to contribute significantly to ROS control Biomedical imaging and therapy 101
  Ultrasmall Cu5.4O nanoparticles ROS scavenging Exhibited cytoprotective effects against ROS-mediated damage Improved treatment outcomes in acute kidney injury, acute liver injury and wound healing 16
Upconversion nanoparticles (UCNPs)-MoS2 nanoassemblies UCNPs coupled with MoS2 ROS detection and monitoring Exhibited low cytotoxicity, rapid response, and high sensitivity; effective in luminescence assays and zebrafish bioimaging Bioimaging, physiological and pathological process monitoring 102
  CeO2:Yb7Tm0.5 with a butterfly-like structure ROS scavenging Exhibit catalase and glutathione peroxidase-mimicking enzyme properties Treat acute lung injury 104


NPs can limit the disadvantages of conventional RONS scavengers. Scientists have synthesised radical-containing NPs and conjugated antioxidants to give them more functionality. NPs with antioxidant properties, such as ceria and gold NPs, are primarily used as RONS scavengers.

7. Nanomaterial-driven strategies for RONS mitigation in cancer therapy

Nucleic acids are the source of RONS, which include oxygen, superoxide anion, hydroxyl radical, lipid peroxides, hydrogen peroxide, protein peroxides, and peroxides. They are also formed by the iron-catalysed Fenton reaction and enzymatic reactions involving cyclooxygenases, lipoxygenases, xanthine oxidases, and NADPH oxidases (NOXs), which have been specially developed for the creation of RONS. These oxidation-reduction mechanisms control signalling molecules that are crucial for directing cellular pathways and preserving dynamic balance in biological systems.105–109 Nucleic acids and enzymatic processes, such as those involving NOX enzymes and the Fenton reaction, produce RONS, which include hydrogen peroxide, superoxide anion, and hydroxyl radical. NOX enzymes are specifically designed to produce RONS. These substances are essential to signalling pathways that control cellular processes and preserve redox balance.110–112 Aberrant RONS levels at various phases of cancer genesis paradoxically influence cell development and death.113

In cancer cells, NADPH oxidase and mitochondria are important producers of RONS. Fig. 4 highlights the use of nanomaterials in the treatment of cancer by showing the ways in which they scavenge RONS. With an emphasis on molecular mechanisms that raise RONS levels above a crucial threshold to cause cancer cell death, the figure illustrates how nanomaterials can be engineered to either produce or scavenge RONS, supporting the creation of targeted anticancer treatments.


image file: d5ra05248c-f4.tif
Fig. 4 Schematic of RONS generation from mitochondria and NADPH oxidase, and the mechanisms involved in RONS scavenging nanomaterials in cancer cells. NOX: NADPH oxidase; GSSG: glutathione disulfide; SOD: SOD; GR: glutathione reductase; GPX: glutathione peroxidase; CAT: catalase, and e: electron.

7.1 ROS in different types of cancers

Functionalized gold NPs (GNPs-Pep-A) show efficient activity in targeted drug delivery and therapeutic applications. GNPs-Pep-A NPs are specifically targeted toward cancer cells and reduce the side effects of therapeutic applications. It also expresses a novel, promising property. In their report, Sushma Kalmodia and co-workers formulated a gold nano-bio-conjugate with His, Cys, Lys, Arg, Met, and Pep-A, etc. Gold nanoparticles and amino acids were synthesised using thioctic acid linkers. The GNPs-Pep-A formula was tested in the retinoblastoma cancer model, reducing 70% of reactive oxygen species (ROS) in tumorous cells. The combination of organic and inorganic phases showed a 9% to 40% reduction in ROS. The synergistic effect of the novel GNP-Pep-A was confirmed in Vitis vinifera L. The formula also increased radical scavenging and apoptosis of tumour cells by polyphenol-coated GNPs.114

Nanomaterials have shown promise in treating various illnesses, but they struggle with drug delivery in cancer patients due to blood circulation issues and drug leakage. Xiaoding Xu et al. developed a new drug-loaded stable polydrug NP that is responsive to ROS. The NP's formulation improves tissue penetration and tumour targeting through surface-encoded internalising RGD (iRGD). The outer shell of polyethene glycol prolongs blood circulation, and the inner core of the polyprodrug responds to the triggered release of ROS, effectively killing cancer cells and inhibiting cell growth.115

Mena Aioub and colleagues demonstrate the selective ablation of tumour cells using plasmonic photothermal therapy (PPT) based on gold nanorod (AuNR).116 This therapy often causes damage or dysfunction to healthy cells, leading to mutations in the cell membrane, DNA, protein, and lipids due to high heat tolerance and high reactive oxygen species levels. Mena Aioub reported an alternative option for gold nanorods in plasmonic photothermal therapy for the treatment. A platinum-coated gold nanorod with a range of platinum-coating shell thicknesses was effectively developed. The traditional effect of gold nanorods was maintained by the combination of platinum-coated gold nanorods (PtAuNRs). PtAuNRs also protected the surrounding healthy cells from ROS formation. ROS was a byproduct of plasmonic photothermal therapy treatment and caused cell death due to the ROS-scavenging.

The study used fluorescence assays and cell viability tests to examine how heat stress in plasmonic photothermal treatment (PPT) produces ROS. PtAuNRs were used to efficiently suppress tumor cells through hyperthermia and lessen the negative consequences of ROS production. PtAuNRs' formulation exhibits both high photothermal efficiency and ROS-scavenging activity, indicating strong therapeutic efficacy in PPT.116

Recent research has increasingly focused on the role of apoptosis-triggered reactive oxygen species (ROS) in enhancing photodynamic therapy (PDT) for cancer treatment. A study led by Kai Zhang and colleagues demonstrated that ROS can be generated even within low-oxygen tumour environments. This process contributes to elevated oxygen levels inside tumours, representing a novel approach that had not been previously explored. The researchers utilized Fe3O4 nanoparticles exhibiting peroxidase-like activity. These nanoparticles interact efficiently with the tumour's endogenous hydrogen peroxide (H2O2), resulting in the production of both ROS and oxygen. This mechanism mirrors the reliance of traditional PDT on external oxygen sources. In addition, the team developed specialized nanoparticles known as FCCP NPs. These are composed of porphyrin, copper sulfide (CuS), and chitosan-coated Fe3O4. The design of these nanoparticles supports both therapeutic applications and multimodal imaging. FCCP NPs are capable of targeting tumours and facilitating various imaging techniques, including photothermal imaging (PTI), photoacoustic imaging (PAI), magnetic resonance imaging (MRI), and photoluminescence imaging (PLI). In summary, this research underscores the potential of advanced nanotechnology to address oxygen limitations within tumours. By generating both ROS and oxygen directly inside the tumour, these strategies can significantly improve the effectiveness of photodynamic therapy.117

7.2 Breast cancer

RONS play a crucial role in breast cancer therapy, influencing cell death, DNA, immunological responses, and antiangiogenic effects. Understanding and influencing RONS's function can lead to innovative therapeutic approaches. Baojin Ma's team discovered that self-assembled copper-amino acid mercaptide nanoparticles (Cu-Cys NPs) can be strengthened by H2O2 and activated by glutathione, transforming Cu2+ into Cu+ and producing toxic hydroxyl radicals (·OH). This nanoformulation has shown remarkable efficacy in cancer therapy and responds effectively to tumour-induced ROS.118

NPs have high drug loading and targeting ability because they can carry a chemotherapeutic agent. These modified multifunctional NPs enhance anticancer activity as well as therapeutic application. At the same time, the NPs may release the drug at non-specific sites. Most of the drugs are toxic, and they affect healthy cells surrounding the tumours. In their report, Hyeon-Yeol Cho and their team worked on malignant tumours with ROS, which is highly efficient and selective for cancer. ROS generators, such as diethyldithiocarbamate (DDC) and sodium nitroprusside (SNP), were used with the SOD 1 inhibitor. Compared with SNP or DDS alone as a control drug, SNP treatment combined with DDC-loaded THoR-NP showed a promising effect on numerous cancer cell lines by generating ONOO in the cancer cells. Fig. 5 shows the absence of external ROS DDC conjugating with THoR-NP, and the effective removal of cancer cells via magnetic hyperthermia due to the magnetic core (ZnFe2O4) of THoR-NP. A high amount of integrin-expressing tumour cells was selectively and efficiently killed by adding an iRGD peptide as part of THoR-NP. The mouse xenograft model confirmed this experiment. In this approach, NP incorporates a ROS-scavenger-inhibitor with external ROS, which are highly discriminating and effective for tumour therapy.119


image file: d5ra05248c-f5.tif
Fig. 5 Anticancer properties and the effect of magnetic hyperthermia using DDC,119 Copyright © 2019, American Chemical Society.

In this work, Deng et al. developed a ROS scavenging nanoplatform (TECM-NS) functionalized with a peptide that targets the extracellular matrix (ECM). They functionalized the nanoscavenger with dual-benzaldehyde-terminated polyethene glycol (PEG) to generate PEG-TECM-NS via acid-cleavable imine bonds, thereby creating a pH-responsive “stealth” delivery approach. This clever nanoscavenger strengthens T-cell-based customised cancer immunotherapy by eliminating excess ROS in the TME, lowering ROS-mediated immunosuppression, and boosting the ICD triggered by the anticancer medication oleandrin (OLE). By rupturing the chemical bonds at the tumor location, the acidic pH can break the PEG shell and reveal the ECM-targeting peptide, which enables the nanoscavenger to attach to the ECM and constantly remove extracellular ROS. TECM-NS can be oxidized by ROS, which causes it to break down and release OLE under regulated conditions. By encouraging the release of HMGB1, the released OLE penetrates tumor cells and causes ICD. Crucially, HMGB1 maintains its immunostimulatory activity by neutralising extracellular ROS, which leads to dendritic cell activation and effective antigen presentation to T cells, ultimately resulting in a strong antitumor immune response.120

7.3 Cancer cell imaging

Nanoparticles can enhance existing imaging modalities such as MRI, CT scans, or fluorescence imaging by serving as contrast agents that accumulate in areas of high RONS activity, thereby improving the visualisation of tumours and their microenvironment. Monitoring RONS levels using nanoparticle-based imaging can help assess the efficacy of cancer treatments, including chemotherapy and radiation therapy, by evaluating changes in oxidative stress within tumour cells over time. Collecting and leveraging nanoparticles for RONS-based cancer cell imaging offers a multifaceted approach that enhances cancer detection, targeting, therapy, and treatment monitoring.

As functional nanomaterials that resemble natural enzymes, carbon dots (CDs) have become popular. Dehvari et al. synthesised Mn, N, and S integrated CDs (MnNS: CDs) using a one-pot microwave hydrothermal process. These CDs can be used as fluorescence and magnetic resonance imaging probes and can reduce oxidative stress by mimicking catalase. Imaging studies revealed that when the CDs were coupled with hyaluronic acid, they exhibited improved biocompatibility and selective targeting of cancer cells that overexpress CD44. When compared to free DOX, the dual-modal nanoprobe efficiently administered doxorubicin (DOX), increasing cell killing efficiency by 50%. This work highlights quick, scalable synthesis methods with confirmed in vitro and in vivo efficacy, demonstrating the potential of multifunctional CDs as a theranostic platform for dual-modal imaging and targeted treatment.121

Lin et al. discovered that P3+ and Mn2+ can be doped into magnetofluorescent carbon dots (C-dots) (PMn@Cdots) using a pyrolysis technique and microwave heating. The aim was to control the red-emission and free radical scavenging of the PMn@Cdots, using them as antioxidants and dual-modal imaging nanoprobes. The synthetic method is simple, quick, repeatable, and scalable. PMn@Cdots coupled with hyaluronic acid (PMn@Cdots/HA) exhibits good biocompatibility in vitro and in vivo, and shows antioxidant efficacy against superoxide, hydroxyl, and DPPH radicals. MnO2 nanoparticles improve photodynamic treatment by lowering hypoxia in tumour microenvironments. Cdots/HA demonstrated dose-dependent defence against H2O2-induced oxidative stress in B16F1, HeLa, and HEL cells. Cdot-based theranostics may serve as dual-modal probes for accurate diagnosis and possible therapeutic intervention in clinical settings.122

Passi et al. investigated the use of silk fibroin to administer imaging agents and antioxidants. A one-step desolvation procedure was used to create silk fibroin nanoparticles (SFSNPs) loaded with the antioxidant medication sulforaphane. The anionic SFSNPs were then mixed with cationic cerium oxide NPs (CeNPs) and PEI passivated carbon dots (CDs) to form self-assembling CeNP-CD@SFSNPs nanocomposites. PEI was employed as a passivating agent, while mulberry leaves (Morus indica) served as a green source of carbon to produce positively charged CDs. Because CeNPs have redox characteristics, they increased the antioxidant efficiency while the CDs' green fluorescence served as a molecular probe. As a result, the CeNP-CDa@SFSNPs nanocomposite may be a good option for concurrent medication delivery and imaging against oxidative stress.123

Self-regulated hybrid semiconducting polymer nanoparticles (SPNs), a type of optical nano theranostics, have been shown by Zhu et al. to improve photodynamic therapy (PDT) for advanced cancer treatment. The formation of ROS in response to the disease microenvironment is frequently not controlled by traditional nano theranostics, putting healthy tissues at risk of damage. The SPNs solve this by combining nanoceria, which adjust ROS levels according to ambient acidity, with NIR-absorbing polymers, which act as PDT agents. This hybrid structure increases the formation of ROS that target cancer in acidic tumor settings, while minimising harm to normal tissues when exposed to NIR lasers. These SPNs showed improved photodynamic efficacy and imaging capabilities in mouse models, offering a promising method for safe, targeted cancer treatment.124

7.4 Cancer cell sensing and detection

RONS are essential for sensing and identifying cancer cells when nanomaterials are used. Multiple RONS can be detected simultaneously using multiplexing nanomaterial-based sensors, which offer a thorough profile of oxidative stress in cancer cells and their surrounding environment. By utilising the special qualities of nanomaterials, researchers can create extremely sensitive and selective RONS detection platforms. This will enhance our knowledge of cancer biology and increase the capacity for early-stage detection and customized treatment plans.

Reactive oxygen species (ROS) are prevalent in inflammatory tissues and cancer cells. A study by Pandya et al. found that paclitaxel-based nanoparticles (NPs) with pinacol-type boronic ester groups can be formulated and characterised. In an in vivo experiment, ROS-sensitive NPs were found to control the lungs, spleen, and liver, potentially targeting tumorous organs. This bio-distribution demonstrates the potential of ROS-sensitive NPs in cancer treatment. Paclitaxel (PTX) on polymer-based NPs shows anti-tumour efficacy and toxicity and reveals the biodistribution of the NPs. The treatment with PTX containing ROS-sensitive NPs showed effective and improved results in HeLa xenografts as compared to ROS-non-sensitive NPs containing PTX and free PTX in anti-tumour cancer therapy.125

Nadezhda M. Zholobak and associates developed a smart system based on calcein-nanoceria to observe how ROS behave in living cells. This method uses dye-carrying particles (calcein) that, when exposed to ROS, break down and release the dye. When calcein is released, it glows, but if it combines with ROS to create a complex with nanoceria, the glow disappears. Researchers can track ROS activity in real time by using a microscope to observe changes in fluorescence. When high ROS produce fluorescence throughout the cell that eventually concentrates in the nucleus, the approach allows for the tracking of phases of oxidative stress, such as an early viral infection (VSV). This technique offers a sensitive way to track ROS and cerium dioxide behaviour in cells.126

Cancer therapy is receiving more attention because of the production of ROS induced by apoptosis in photodynamic therapy (PDT). Fe3O4 nanoparticles, which resemble peroxidase enzymes, are used by Kai Zhang and colleagues to produce more ROS to improve photodynamic treatment (PDT) in low-oxygen malignancies. When these nanoparticles interact with hydrogen peroxide (H2O2) in tumour cells, additional O2 and ˙OH radicals are produced. This invention creates potent ROS that may damage and kill tumour cells and increase oxygen availability for PDT, representing a previously unexplored approach to cancer treatment.

Synergetic therapy and multimodal diagnostics were achieved with the synthesised multifunctional nano combination modified with porphyrin as well as with CuS, and this modified chitosan-encapsulated Fe3O4 NP (FCCP NPs) improved the efficiency of cancer therapy by using essential peroxidase-related activity to generate O2 from internal H2O2 and produce ROS. The tumour-selecting attributes of FCCP NPs were reported for multimodal diagnostic approaches in vivo, such as photothermal imaging (PTI), MRI, photoacoustic imaging (PAI), and photoluminescence imaging (PLI). This study demonstrated how nanotechnology might eliminate the present inadequacies in tumour therapy by producing ROS and O2 for PDT.127

Cancer cells produce ROS, which is responsible for increased peroxisome activities, elevated metabolic rate, oncogene expression, impaired mitochondria, and elevated cell signalling. Cancer cells need a specific amount of ROS, either higher or lower, to cause cytotoxicity. By taking advantage of this biochemical feature, therapeutic drugs are created to specifically and preferentially target cancer cells.

8. Utilizing nanomaterials for reactive oxygen and nitrogen species scavenging in disease treatment

The human body malfunctions due to a rise in oxidative and nitro oxidative stress, and a reduction in the antioxidant load of the various organs. There are currently few clinical treatments for this condition. Researchers are becoming more interested in using nanomaterials to treat oxidative stress-related diseases. Numerous NP types can scavenge RONS, making them useful for treating various conditions, including diabetes and neurological illness, and for wound healing, tissue engineering, and reducing inflammation.14

8.1 Diabetes

Higher concentrations of RONS—superoxide radicals, hydrogen peroxide, and nitric oxide—all signify oxidative stress and are related to diabetes. This oxidative stress worsens diabetes-related issues, including retinopathy, nephropathy, and neuropathy. In summary, nanomaterials have several applications in diabetes research and treatment, such as RONS tracking and detection, drug administration, and tissue healing assistance. For persons with diabetes, these advancements may result in improved results and an improved standard of living.

Y. Tong and colleagues developed high-dose hollow mesoporous silica nanocomposite particles doped with cerium oxide, which inhibit ROS-associated diabetic nephropathy (DN) pathogenesis through renoprotective activity. The MET-HMSN-CeO2-sized NPs loaded with metformin showed greater kidney accumulation compared to free MET. The system's cyclic transformation of mixed-valence ceria may have antioxidative and ROS-scavenging properties. The nanocarrier HMSN-CeO2 demonstrated renoprotective qualities, making it a promising therapeutic approach for DN prevention.128 Cerium oxide nanoparticles (nanoceria), a water-soluble cerium nanoparticle, benefits diabetes mellitus by scavenging reactive oxygen species (ROS) and promoting wound healing in mouse skin. This process reduces ROS-induced cell death by eliminating intracellular ROS and blocking H2O2-activated apoptosis pathways.129 Chronic wounds, such as diabetic ulcers, face oxidative stress and infection risks. MoS2–CeO2 nanoparticles possess antioxidant and photothermal antibacterial properties that can support wound healing and cell migration.130 Fullerene nanoparticles are utilised to address diabetes complications such as hyperglycemia, which can impair spermatogenic processes and lead to testicular dysfunction. By reducing reactive oxygen species levels, fullerene helps improve male sexual dysfunction and reproductive impairment. Additionally, it alleviates pancreatic dysfunction and hepatic insulin resistance.131

8.2 Tissue engineering

Regulated RONS concentrations could function as signalling molecules that affect cell migration, differentiation, and proliferation. Nanomaterials can be engineered to modulate RONS production in a spatially and temporally controlled manner, guiding cell fate in tissue engineering constructs.

The deterioration of bone and cartilage, evidently intensified by an upsurge in inflammation, was found to promote ROS signalling, as in the case of osteoarthritis (OA). Since cartilaginous tissues do not have blood vessels, chondrocytes are hypoxic and inactive under normal conditions. Chondrocytes become activated when a disruption in the mitochondrial processes regulates the generation of intracellular ROS and increases OA.132

An ROS-scavenging scaffold loaded with rapamycin (Rapa@Gel) may improve intervertebral disk (IVD) tissue regeneration due to the enhanced ROS in the chronic inflammatory environment as reported by Bai et al. By injecting ROS-degradable hydrogel therapeutic scaffold into injured IVD sites, drug can be released in a regulated way, lowering inflammatory reactions. Rats treated with Rapa@Gel had a higher proportion of M2-like macrophages and a lower percentage of M1-like macrophages, which promoted IVD regeneration by lowering inflammation and increasing M2 macrophage activity.133

8.3 Wound healing

The human body's biological process to repair wounds goes through multiple stages, including hemostasis, remodelling, proliferation, and inflammation.134 Compared to normal tissues, injured tissues have higher ROS levels. Overproduction of ROS at wound sites exacerbates oxidative stress, leading to damage to DNA and RNA, apoptosis, and failure of the cellular machinery.135 Several research studies indicate that increased ROS levels impede the body's natural healing process of wounds. Various kinds of ROS-sensitive biomaterials have been developed and used in wound repair to regulate the wound-healing process. Some examples include hydrogels, surgical structures, nanofibrous materials, and drug delivery systems at the nanoscale. Additionally, microneedle patches can directly or indirectly control local ROS production. According to Wang et al., bacterial wounds can be healed by using light-responsive multifunctional nanomaterials. To affect the synergistic chemical and photodynamic therapy for microbially contaminated skin wounds, photosensitizer nano complexes comprising chlorin e6 (Ce6) and magnesium (Mg) were developed.136

8.4 Osteoclastogenesis

The process that produces osteoclasts, which are specialised cells in charge of bone resorption, is known as osteoclastogenesis. RONS can affect osteoclast development and function within the framework of osteoclastogenesis. By influencing the resorptive activity and survival of mature osteoclasts, RONS can potentially affect their function. The accumulation of RONS can cause oxidative stress, which may hamper osteoclast function and exacerbate illnesses related to the bones. Chen et al. discovered that nanomaterials, specifically nano-fluorescent carbon quantum dots (N-CDs), can scavenge reactive oxygen species (ROS). Overactivation of osteoclasts because of excessive ROS can progress to osteolytic diseases. N-CDs counteracted RANKL-induced ROS generation, inhibiting NF-κB and MAPK pathways, thereby suppressing osteoclastogenesis and bone resorption. In vivo studies showed N-CDs protected mice from calvarial bone degradation and tibial bone loss caused by LPS and breast cancer cells, respectively.137 Using glucose-sorbitol-carboxymethyl ether (PSC) as a biological polysaccharide-based antioxidant, Pengjun et al. synthesized Fe2O3@PSC nanoparticles (NPs). These 7.3 nm-sized nanoparticles, which have an elemental composition of O/Fe/Cl/C (190[thin space (1/6-em)]:[thin space (1/6-em)]7[thin space (1/6-em)]:[thin space (1/6-em)]2[thin space (1/6-em)]:[thin space (1/6-em)]88), were created to scavenge excess ROS, suppress osteogenesis, and encourage osteoclast variation to treat osteoporosis caused by iron accumulation (IA). Fe2O3@PSC NPs were found to provide iron levels equivalent to ferric ammonium citrate in ex vivo studies. They were also shown to stimulate ROS scavenging in MC3T3-E1 and Raw 264.7 cells, reduce osteoclast growth by inhibiting the MAPK and NF-κB pathway, and increase osteogenic variation by activating Akt-GSK-3β-β-catenin. By preventing IA-related osteoporosis in a mouse model, the NPs showed potential in treating illnesses associated with iron insufficiency.138

8.5 Inflammation

The link between ROS and inflammation is mutually beneficial. In acute inflammation, ROS produced by neutrophils or macrophages will cause injured cells to undergo apoptosis, which will aid in the healing process.139 Furthermore, ROS can control the inflammatory pathway that links tumor necrosis factor to tumor necrosis factor receptor (TNF-TNFR) in cases of chronic inflammation.140 NOXs are triggered by interactions between RFK and p22phox and the cytoplasmic domain of TNFR. Then ROS are produced by converting extracellular O2 to O2˙−.

Critical, long-standing, or persistent inflammatory diseases are caused by oxidative stress. Enzymes exhibit good reactivity with ROS and biocompatibility when used to treat disorders associated with ROS. In this study, Tengfei Liu et al. described the synthesis of copper oxide NPs using a simple single-step procedure. The fabrication of nano-bio-conjugates with copper oxide NPs was used with multiple enzymes and ROS's wide range of scavenging abilities. Cu5.4O based USNPs depict potent cytoprotective effects by mimicking the activities of ;catalase, SOD, and glutathione peroxidase. The cytoprotective effect, in contrast to ROS, is a very effective treatment for critical liver or kidney injury and wound healing, and it causes damage at extremely low doses. Cu5.4O USNPs' tiny size enables rapid renal clearance of the nanoformulations and demonstrates promising biocompatibility. Cu5.4O USNPs are very biocompatible and have protective properties. The strategy will make it easier to produce next-generation enzymes and streamline the treatment of ROS-related illnesses.141

A synergistic link is also taken into consideration with RNS and inflammation. TNF signaling is initiated by TNF binding to TNFR via a complex known as Receptor interacting protein kinase-1-Tetanus AntiBacilli-Transforming Growth factor beta-activated kinase1 (RIPK1–TAB–TAK1). This complex then stimulates Activator protein-1 (AP-1) transcription and NF-KB (nuclear factor kappa light chain enhancer of activated b-cells). NF-KB, another important inflammatory factor, was released from the NF-KB/IKB complex following IKB phosphorylation.142 Moreover, the mitogen-activated protein kinase (MAPK) pathway activated the inflammatory factor AP-1.143 Upon entering the nucleus, AP-1 and NF-KB activate downstream inflammatory genes. NF-KB drives the production of inducible nitric oxide synthase (iNOS/NOS2), which can produce NO.31

Increased RONS causes mutagenesis, neoplasia, and tumor invasion (e.g., antioxidant gene-GPx4). A persistent inflammatory response stimulates inflammatory cells (neutrophils and macrophages), which results in the continual production of RONS. A self-perpetuating loop is created when oxidative stress causes a prolonged inflammatory response, which, in turn, causes a persistent generation of ROS and RNS.111

8.6 Neurogenerative disorders

The buildup of amyloid-Aβ peptide in the brain could be a factor in the progression of Alzheimer's disease.144 Kim and colleagues presented a novel approach to blood Aβ cleansing, which involves treating blood extracorporeally with multifunctional magnetite/ceria NP assemblies (MCNAs) that are organized like a core and shell. Fig. 6A demonstrates that the MCNA shell is made up of a ceria NP, which is in charge of catalytic activities for scavenging ROS, and the MCNA core is made up of spherical form superparamagnetic NPs for magnetic separation utilizing an external magnetic field. SOD and catalase (CAT) activity tests are used to evaluate the scavenging abilities of CNAs. To validate the blood Aβ cleansing treatment, in vitro investigations were conducted utilizing the 5XFAD transgenic mice model as an age-dependent model of brain amyloidosis development. In this case, the ROS generation in the experiment is reduced with the use of ceria NPs. Of the detected increase in plasma ROS, 42% was attributed to core–shell assemblies of ceria and magnetite NPs. Fig. 6B illustrates how MCNAs, ROS-scavenging NP blood Aβ peptides, and brain Aβ plaques are necessary to treat Alzheimer's illness.145
image file: d5ra05248c-f6.tif
Fig. 6 (A) Schematic of the synthesis of magnetite/ceria NP assemblies (MCNAs) and (B) TEM images of magnetite NPs.145 Copyright 2019, WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim.

Hao et al. proposed that inorganic nanomaterials can scavenge and prevent ROS-generated neurological diseases, such as ameliorating Parkinson's disease. Porous copper oxide nanocrystals (CuxO NCs) of about 65 nm were synthesised, where phenylalanine was used as the structure-directing agent of the nanocrystals. CuxO NCs eliminate ROS because they functionally mimic the activities of peroxidase, SOD, catalase, and glutathione peroxidase and inhibit neurotoxicity in Parkinson's disease. Cell viability testing was used in biocompatibility investigations on the nanocrystals. No cytotoxic effects were observed for the CuxO NCs. The Morris water maze test assessed the PD mice's spatial learning and memory following the experiment. This suggests that the ROS-scavenging properties of CuxO NCs lessened the neuroinflammation in the PD model mice. CuxO exhibits remarkable therapeutic efficacy against neurological problems in PD mice caused by oxidative stress, according to the in vivo investigations.146

Kwon et al. synthesized TPP-conjugated ceria nanoparticles (CeO2) as reusable ROS scavengers that mitigated Aβ-induced mitochondrial ROS and damage in SH-SY5Y neuronal cells and mice. By treating mitochondrial malfunction brought on by aberrant ROS formation, these nanoparticles have the potential to treat Alzheimer's disease by reducing mitochondrial oxidative stress.147 Kim et al. demonstrated that cerium oxide nanoparticles (CONPs) reduce ROS, inflammation, and apoptosis in contusion-injured rat spinal cords, aiding locomotor recovery. CONP treatment suppressed inflammatory and apoptotic molecules, reduced lesion size, and improved BBB and ladder scores. These therapeutic effects, attributed to ROS scavenging, suggest CONPs as a promising cure for acute spinal cord injury.148 In order to mimic SOD, CAT, and GPx, Singh and contemporaries synthesized monodisperse Mn3O4 nanoflowers (Mnf) with multi-enzyme activity. In Parkinson's disease, these nanozymes scavenge hydroxyl radicals, shielding cells from oxidative damage and cytotoxicity. Mnf internalizes into human cells and offers SHSY-5Y cells cytoprotection without harm, showing therapeutic promise against neurological illnesses caused by reactive oxygen species.149 CeNC/IONC/MSN-T807 is a multifunctional nanocomposite that was produced by Qing Chan and colleagues. It contains iron oxide nanocrystals and ceria on MSN, tau protein binding via T807, and tau aggregation inhibition via methylene blue (MB), as shown in Fig. 7A. This nanocomposite is monitored by MR/PET imaging, targets hyperphosphorylated tau, and reduces mitochondrial oxidative stress, as depicted in Fig. 7B. For Alzheimer's disease, CeNC/IONC/MSN-T807-MB is a promising tau-targeted theranostic drug since the combination of MB and CeNC improves therapeutic outcomes.150


image file: d5ra05248c-f7.tif
Fig. 7 (A) Multi-enzyme activity provides efficient cryoprotection in Parkinson's disease model.149 Copyright 2017, Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim. (B) CeNC/IONC/MSN-T807-MB with bimodal imaging capability, and neurons are protected from ROS-mediated apoptosis.150 Copyright 2018, American Chemical Society. (C) Molybdenum disulfide NPs as multifunctional inhibitors151 and (D) schematic of intracellular calcium change induced by Aβ42 with or without MoS2 NPs.151 Copyright 2017, American Chemical Society.

As illustrated in Fig. 7C, Han et al. discovered that molybdenum disulfide (MoS2) nanoparticles inhibit Aβ aggregation, destabilize Aβ fibrils, and reduce oxidative stress and cell toxicity brought on by Aβ. Additionally, they discovered that MoS2 NPs preserve calcium homeostasis in cell membranes by preventing Aβ-induced Ca2+ channel development. These multifunctional effects imply that MoS2 NPs have great promise as therapeutic agents for illnesses linked to amyloids. They have shown that injecting MoS2 NPs might lower the intracellular ROS level caused by Aβ42, as seen in Fig. 7D. These findings, therefore, point to the considerable potential of MoS2 NPs for multifunctional therapy against Alzheimer's disease.151

Here, we have discussed different RONS-scavenging nanomaterials and their different biomedical application, as summarized in Table 2. The antioxidant properties of these nanomaterials are well studied, but their stability and actual physiological behavior need to be investigated further. Some nanomaterials show effective therapeutic effects, while others have some challenges that should be optimized, such as toxicity and bioavailability in diabetes, tissue engineering, wound healing, inflammation, and neurodegenerative diseases. Mortality and impairment among children and young adults are caused by traumatic brain injuries (TBIs) that get worse because no adequate treatments are available; under such conditions, nanomaterials show an ability to improve this biomedical treatment, as shown in Fig. 8.

Table 2 Summary of RONS scavenging nanomaterials and their potential therapeutic applications
N 3anomaterials ROS scavenging mechanism Application Important results References
Melanin nanoparticles (MeNPs) O2˙−, H2O2, ˙OH, ˙NO, and ONOO Brain injury in ischemic stroke MeNPs can also reduce the RONS-triggered inflammatory reactions through suppressing the expression of inflammatory intermediaries and cytokines 152
Cerium oxide nanoparticles (nanoceria) ROS Anticancer treatment FA-nanoceria controlled intracellular ROS to a greater extent than the nanoceria in colon carcinoma cells and reduced their endocytosis and redox activity 153
Cerium oxide nanoparticles (CNPs) ROS Melanoma (malignant skin cancer) Polymer-coated CNPs nontoxic for stromal cells exhibited a cytotoxic, proapoptotic, and anti-invasive capacity on melanoma cells 154
Ceria-zirconia nanoparticles (CZ NPs) O2 and ˙OH Sepsis treatment CZ NPs significantly enhance ROS scavenging activity, hence regulating inflammatory cells at a very low dose 155
Gd-conjugated, oxygen reactive polymer (ORP) H2O2 Traumatic brain injury Oxygen reactive polymer reduces H2O2 levels threefold 156
Manganese dioxide nanoparticles (MnO2 NP) O2 Ovarian cancer PDT with Ce6@MnO2-PEG nanoparticles reduced dose and improved therapeutic efficacy in tumor growth 157
Polysorbate 80 (PS80)NP H2O2 Traumatic brain injury SeNPs decorated with polysaccharide-protein complex (PTW)/PG-6 peptide and loaded with TMP/GM1 158
Poly(propylene sulfide) (PPS) (PPS-NPs) ROS Ischemic stroke Negligible cytotoxicity 159
Ceria NPs ROS Ischemic stroke Uniform 3 nm-sized ceria NPs 160
Selenium nanoparticles (Se NPs) ROS Testicular damage in streptozotocin-induced diabetic SeNPs could extensively reduce the testicular tissue oxidative stress markers, specifically lipid nitric oxide and peroxidation, and improved the glutathione content and antioxidant enzyme activities in testicular tissues 161
Nanoceria (NCe) ROS Ovarian cancer NCe inhibited the production of ROS in A2780 cells, reduced growth factor-mediated cell migration and invasion of SKOV3 cells, without disturbing the cell proliferation 162
Cerium oxide nanoparticles (nanoceria) ROS Ovarian and colon cancer cells The larger nanoceria was found to scavenge intracellular ROS to a better extent than the smaller nanoceria, and ROS scavenging was found to be enhanced with treatment time 163



image file: d5ra05248c-f8.tif
Fig. 8 RONS (reactive oxygen and nitrogen species) generation and neutralization play a critical role in neuronal cells, particularly concerning their implications for neurodegenerative diseases. Within this process, the mitochondrial electron transport chain (mito-ETC) is a significant contributor to RONS production. Antioxidant enzymes such as superoxide dismutase (SOD) and glutathione peroxidase (GPx) work in tandem with reduced glutathione (GSH) to counteract the damaging effects of RONS, while the balance between reduced (GSH) and oxidized glutathione (GSSG) is crucial for maintaining cellular redox homeostasis. Understanding these mechanisms is essential for exploring therapeutic avenues in neurodegenerative disease contexts.

9. Current challenges and aspects

The RONS-savaging reactions based on nanomaterials have value in numerous applications in disease management. However, some challenges, such as chemical stability, biocompatibility, and toxicity of nanomaterials, need to be taken care of to effectively use nanomaterials as RONS scavengers. According to recent research, the low activity and short half-life of traditional antioxidants make it difficult to develop an effective antioxidant therapy or an effective nanomaterial for RONS scavenging.

(a) The design of nanomaterials for RONS scavenging and therapeutic applications is challenging due to the property-specific nanomaterials and their functionalities. The design of nanomaterials should focus on RONS scavenging or elimination.

(b) The design process should consider the structure and corresponding function of nanomaterials. The core of the nanomaterials and functionalities should be given sufficient attention while designing nanomaterials.

(c) Accounting for the clinical and pathological features in the design of a nanomaterial is more challenging because the design of the nanomaterial is fully integrated with the microenvironment characteristics.

(d) Maintaining significant biocompatibility and circulation constancy is moderately challenging for nanomaterials. Systemic and general administration is required for the majority of nanomaterials for disease management. Verification of the developed nanomaterials' stability, circulatory half-life, and toxicity is needed.

(e) Modifications and drug loading on the nanomaterial lead to complex multilayer structures, which may be unstable and become more complicated for functional study.

(f) Numerous nanomaterials encounter issues, including inadequate penetration, a lack of targeted applications, and insufficient delivery of drugs. However, there is an opportunity for us to address such issues. Research on the mechanism of RONS scavenging and disease management is to be strengthened. Such studies need to identify the exact routes of diseases through RONS sequencing and examine the details of disease management.

We summarized and reviewed the significant studies in this area and presume that the concerns outlined above should be addressed in future designs and research into nanomaterials for RONS scavenging and treatment. Nanomaterials for RONS scavenging require careful management of circulatory stability and toxicity in vivo. Future disease management will focus on designing nanomaterials and their functionalities. Strengthening the materials science field is crucial for improving appropriate nanomaterials for disease management. Pre-clinical trials and ethical evaluations will be more challenging for nanomaterials. In the future, nanomaterials will play a significant role in disease management and RONS scavenging.

10. Conclusion and viewpoint

Several methods enhance the therapeutic potential of reactive oxygen and nitrogen species (RONS)-based strategies for antimicrobial and anticancer applications. Surface modifications of nanomaterials, including PEGylation, ligand functionalization, and biomimetic coatings, can improve selective targeting while reducing off-target oxidative damage and enhancing biocompatibility. Hybrid systems, such as nanocomposites combining metallic nanoparticles with metal–organic frameworks (MOFs), photosensitizers, or enzymes, enable controlled ROS production in a synergistic approach. Advanced modulators, such as redox-active drug conjugates, catalytic nanozymes, and stimuli-responsive nanocarriers, facilitate precision in the spatiotemporal regulation of oxidative stress, optimizing therapeutic effectiveness while minimizing associated toxicity. To transition from preclinical to clinical applications, it is essential to establish standardized pathways focusing on toxicity profiling, pharmacokinetics, biodistribution studies, and well-organized clinical trials to ensure therapeutic safety and efficacy.

Reactive oxygen and nitrogen species play a critical role in cell growth and development. However, their levels can increase in neurodegenerative disorders and cancer, leading to increased oxidative stress, which destabilizes redox balance, consequently exacerbating such conditions. Reducing excess RONS is advantageous for recovery in both cancer and neurodegenerative diseases. This review highlights the potential use of RONS scavenging nanomaterials for treating such disorders through different mechanisms of scavenging.

Despite their promise, the long-term stability and physiological behavior of nanomaterials in treating neurological disorders and cancer remain ambiguous. The efficacy of their scavenging action varies depending on multiple factors such as the materials' components, surface coatings, size, shape, morphology, exposed crystal faces, treatment concentration, and environmental conditions. As potent antioxidants, nanomaterials can effectively eliminate RONS in medical applications, including imaging and treatment of cancer and neurologic disorders. However, further research is needed to elucidate their immunological profiles, circulation half-lives, biodistribution, pharmacokinetics, and antioxidant capabilities to fully harness their therapeutic potential.

Author contributions

All of the authors contributed provocatively to the work, whether it was through ideation, study design, data acquisition, analysis, and interpretation, or any combination of these; they also participated in the article's drafting, revision, and critical review; all agreed to take responsibility for the work in its entirety.

Conflicts of interest

The authors declare that there are no conflicts of interest.

Data availability

No primary research results, software or code have been included and no new data were generated or analysed as part of this review.

Acknowledgements

V. B. acknowledges TARE-SERB, TAR/2018/000547 for funding, and CMR Institute of Technology, Bengaluru, India for providing facilities.

References

  1. I. Liguori, G. Russo, F. Curcio, G. Bulli, L. Aran, D. Della-Morte, G. Gargiulo, G. Testa, F. Cacciatore and D. Bonaduce, Oxidative stress, aging, and diseases, Clin. Interv. Aging, 2018, 13, 757 CrossRef CAS PubMed.
  2. S. Zhou, H. Cai, X. He, Z. Tang and S. Lu, Enzyme-mimetic antioxidant nanomaterials for ROS scavenging: Design, classification, and biological applications, Coord. Chem. Rev., 2024, 500, 215536 CrossRef CAS.
  3. D.-B. Cheng, X.-H. Zhang, Y.-J. Gao, L. Ji, D. Hou, Z. Wang, W. Xu, Z.-Y. Qiao and H. Wang, Endogenous Reactive Oxygen Species-Triggered Morphology Transformation for Enhanced Cooperative Interaction with Mitochondria, J. Am. Chem. Soc., 2019, 141, 7235–7239 CrossRef CAS PubMed.
  4. J. Wang, P. Shangguan, X. Chen, Y. Zhong, M. Lin, M. He, Y. Liu, Y. Zhou, X. Pang, L. Han, M. Lu, X. Wang, Y. Liu, H. Yang, J. Chen, C. Song, J. Zhang, X. Wang, B. Shi and B. Z. Tang, A one-two punch targeting reactive oxygen species and fibril for rescuing Alzheimer's disease, Nat. Commun., 2024, 15, 705 CrossRef CAS PubMed.
  5. K. Jakubczyk, K. Dec, J. Kałduńska, D. Kawczuga, J. Kochman and K. Janda, Reactive oxygen species-sources, functions, oxidative damage, Pol Merkur Lek., 2020, 48, 124–127 Search PubMed.
  6. C. M. Bergamini, S. Gambetti, A. Dondi and C. Cervellati, Oxygen, reactive oxygen species and tissue damage, Curr. Pharm. Des., 2004, 10, 1611–1626 CrossRef CAS PubMed.
  7. S. Schlick, M. Danilczuk, A. R. Drews and R. S. Kukreja, Scavenging of Hydroxyl Radicals by Ceria Nanoparticles: Effect of Particle Size and Concentration, J. Phys. Chem. C, 2016, 120, 6885–6890 CrossRef CAS.
  8. S. Kumari, A. K. Badana and R. Malla, Reactive oxygen species: a key constituent in cancer survival, Biomarker Insights, 2018, 13, 1177271918755391 CrossRef PubMed.
  9. J. Tejero, S. Shiva and M. T. Gladwin, Sources of vascular nitric oxide and reactive oxygen species and their regulation, Physiol. Rev., 2019, 99, 311–379 CrossRef CAS PubMed.
  10. X. Li, Y. Wang, X. Liu, J. Li, J. Wu, M. Zeng, J. Yang, N. Hu, H. Zhu, L. Xu and Z. Yang, Trace Detection of Nitrogen Dioxide via Porous Tin Dioxide Nanopods with High Specific Surface Area and Enhanced Charge Transfer, ACS Sens., 2025, 10, 4383–4390 CrossRef CAS PubMed.
  11. M. R. Krishnendu and S. Singh, Reactive oxygen species: Advanced detection methods and coordination with nanozymes, Chem. Eng. J., 2025, 511, 161296 CrossRef CAS.
  12. S. M. Andrabi, N. S. Sharma, A. Karan, S. M. S. Shahriar, B. Cordon, B. Ma and J. Xie, Nitric Oxide: Physiological Functions, Delivery, and Biomedical Applications, Adv. Sci., 2023, 10, e2303259 CrossRef PubMed.
  13. Z. Wang, H. Xiong, Z. Zhai, Y. Yao, T. Zhou, H. Zhang, C. Fan and C. Gao, Reactive oxygen species-scavenging nanoparticles coated with chondroitin sulfate protect cartilage against osteoarthritis in vivo, Nano Res., 2023, 16, 2786–2797 CrossRef CAS.
  14. X. Huang, D. He, Z. Pan, G. Luo and J. Deng, Reactive-oxygen-species-scavenging nanomaterials for resolving inflammation, Mater. Today Bio, 2021, 11, 100124 CrossRef CAS PubMed.
  15. J. Liu, L. Ma, G. Zhang, Y. Chen and Z. Wang, Recent Progress of Surface Modified Nanomaterials for Scavenging Reactive Oxygen Species in Organism, Bioconjugate Chem., 2021, 32, 2269–2289 CrossRef CAS PubMed.
  16. T. Liu, B. Xiao, F. Xiang, J. Tan, Z. Chen, X. Zhang, C. Wu, Z. Mao, G. Luo, X. Chen and J. Deng, Ultrasmall copper-based nanoparticles for reactive oxygen species scavenging and alleviation of inflammation related diseases, Nat. Commun., 2020, 11, 2788 CrossRef CAS PubMed.
  17. P. Pandit, S. Bhagat, P. Rananaware, Z. Mohanta, M. Kumar, V. Tiwari, S. Singh and V. P. Brahmkhatri, Iron oxide nanoparticle encapsulated; folic acid tethered dual metal organic framework-based nanocomposite for MRI and selective targeting of folate receptor expressing breast cancer cells, Microporous Mesoporous Mater., 2022, 340, 112008 CrossRef CAS.
  18. S. Mukherjee, P. Rananaware, V. Brahmkhatri and M. Mishra, Polyvinylpyrrolidone-Curcumin Nanoconjugate as a Biocompatible, Non-toxic Material for Biological Applications, J. Cluster Sci., 2023, 34, 394–414 CrossRef.
  19. S. Zhang, L. Wang, Y. Kang, J. Wu and Z. Zhang, Nanomaterial-based reactive oxygen species scavengers for osteoarthritis therapy, Acta Biomater., 2023, 162, 1–19 CrossRef CAS PubMed.
  20. V. Hegde, P. Pandit, P. Rananaware and V. P. Brahmkhatri, Sulfonic acid-functionalized mesoporous silica catalyst with different morphology for biodiesel production, Front. Chem. Sci. Eng., 2022, 16, 1198–1210 CrossRef CAS.
  21. P. Rananaware, V. P. Brahmkhatri, Fullerene Derivatives for Drug Delivery applications, in: Advanced Porous Biomaterials for Drug Delivery Applications, CRC Press, 2022, pp. 373–393 Search PubMed.
  22. P. Rananaware, V. P. Brahmkhatria, D. Dasgupta and A. Patel, Functionalized mesoporous silica for drug delivery of poorly soluble polyphenols: Synthesis, characterization, and antimicrobial action, J. Solid State Chem., 2023, 326, 124214 CrossRef CAS.
  23. P. Rananaware, P. Pandit, S. Naik, M. Mishra, R. S. Keri and V. P. Brahmkhatri, Anti-amyloidogenic property of gold nanoparticle decorated quercetin polymer nanorods in pH and temperature induced aggregation of lysozyme, RSC Adv., 2022, 12, 23661–23674 RSC.
  24. M. Hunt, M. Torres, E. Bachar-Wikstrom and J. D. Wikstrom, Cellular and molecular roles of reactive oxygen species in wound healing, Commun. Biol., 2024, 7, 1534 CrossRef PubMed.
  25. S. Prasad, P. S. Nayak and P. D’Silva, A histidine-functionalized ROS scavenging hybrid nanozyme for therapeutic application in Parkinson’s disease pathogenesis, Mater. Adv., 2024, 5(6), 2388–2399 RSC.
  26. A. J. P. O. de Almeida, J. C. P. L. de Oliveira, L. V. da Silva Pontes, J. F. de Souza Júnior, T. A. F. Gonçalves, S. H. Dantas, M. S. de Almeida Feitosa, A. O. Silva and I. A. de Medeiros, ROS: Basic Concepts, Sources, Cellular Signaling, and its Implications in Aging Pathways, Oxid. Med. Cell. Longev., 2022, 2022, 1225578 CrossRef PubMed.
  27. U. C. Dash, N. K. Bhol, S. K. Swain, R. R. Samal, P. K. Nayak, V. Raina, S. K. Panda, R. G. Kerry, A. K. Duttaroy and A. B. Jena, Oxidative stress and inflammation in the pathogenesis of neurological disorders: Mechanisms and implications, Acta Pharm. Sin. B, 2025, 15, 15–34 CrossRef CAS PubMed.
  28. N. Üremiş and M. M. Üremiş, Oxidative/Nitrosative Stress, Apoptosis, and Redox Signaling: Key Players, Neurodegener. Dis., 2025, 39, e70133 Search PubMed.
  29. F. R. Palma, B. N. Gantner, M. J. Sakiyama, C. Kayzuka, S. Shukla, R. Lacchini, B. Cunniff and M. G. Bonini, ROS production by mitochondria: function or dysfunction?, Oncogene, 2024, 43, 295–303 CrossRef CAS PubMed.
  30. D. Nolfi-Donegan, A. Braganza and S. Shiva, Mitochondrial electron transport chain: Oxidative phosphorylation, oxidant production, and methods of measurement, Redox Biol., 2020, 37, 101674 CrossRef CAS PubMed.
  31. A. Farahani, A. Farahani, K. Kashfi and A. Ghasemi, Inducible nitric oxide synthase (iNOS): More than an inducible enzyme? Rethinking the classification of NOS isoforms, Pharmacol. Res., 2025, 216, 107781 CrossRef CAS PubMed.
  32. M. Mandal, M. Sarkar, A. Khan, M. Biswas, A. Masi, R. Rakwal, G. K. Agrawal, A. Srivastava and A. Sarkar, Reactive Oxygen Species (ROS) and Reactive Nitrogen Species (RNS) in plants– maintenance of structural individuality and functional blend, Adv. Redox Res., 2022, 5, 100039 CrossRef CAS.
  33. K. Brieger, S. Schiavone, F. J. Miller and K.-H. Krause, Reactive oxygen species: from health to disease, Swiss Med. Wkly., 2012, 142, w13659 CrossRef CAS PubMed.
  34. M. Redza-Dutordoir and D. A. Averill-Bates, Activation of apoptosis signalling pathways by reactive oxygen species, Biochim. Biophys. Acta Mol. Cell Res., 2016, 1863, 2977–2992 CrossRef CAS PubMed.
  35. W. Czarnocka and S. Karpiński, Friend or foe? Reactive oxygen species production, scavenging and signaling in plant response to environmental stresses, Free Radic. Biol. Med., 2018, 122, 4–20 CrossRef CAS PubMed.
  36. A. V. Krylatov, L. N. Maslov, N. S. Voronkov, A. A. Boshchenko, S. V. Popov, L. Gomez, H. Wang, A. S. Jaggi and J. M. Downey, Reactive oxygen species as intracellular signaling molecules in the cardiovascular system, Curr. Cardiol. Rev., 2018, 14, 290–300 CrossRef CAS PubMed.
  37. N. T. Moldogazieva, S. V. Lutsenko and A. A. Terentiev, Reactive Oxygen and Nitrogen Species–Induced Protein Modifications: Implication in Carcinogenesis and Anticancer Therapy, Cancer Res., 2018, 78, 6040–6047 CrossRef CAS PubMed.
  38. J. Ruiz-Gines, S. Lopez-Ongil, M. Gonzalez-Rubio, L. Gonzalez-Santiago, M. Rodriguez-Puyol and D. Rodriguez-Puyol, Reactive oxygen species induce proliferation of bovine aortic endothelial cells, J. Cardiovasc. Pharmacol., 2000, 35, 109–113 CrossRef CAS PubMed.
  39. C. Lennicke and H. M. Cochemé, Redox metabolism: ROS as specific molecular regulators of cell signaling and function, Mol. Cell, 2021, 81, 3691–3707 CrossRef CAS PubMed.
  40. H. Sauer, M. Wartenberg and J. Hescheler, Reactive oxygen species as intracellular messengers during cell growth and differentiation, Cell. Physiol. Biochem., 2001, 11, 173–186 CrossRef CAS PubMed.
  41. H. Huang, W. Du and R. A. Brekken, Extracellular matrix induction of intracellular reactive oxygen species, Antioxidants Redox Signal., 2017, 27, 774–784 CrossRef CAS PubMed.
  42. E. Cadenas and K. J. Davies, Mitochondrial free radical generation, oxidative stress, and aging, Free Radic. Biol. Med., 2000, 29, 222–230 CrossRef CAS PubMed.
  43. J. F. Turrens, Mitochondrial formation of reactive oxygen species, J. Physiol., 2003, 552, 335–344 CrossRef CAS PubMed.
  44. A. Maurel, C. Hernandez, O. Kunduzova, G. Bompart, C. Cambon, A. Parini and B. Francés, Age-dependent increase in hydrogen peroxide production by cardiac monoamine oxidase A in rats, Am. J. Physiol. Heart Circ. Physiol., 2003, 284, H1460–H1467 CrossRef CAS PubMed.
  45. S. Fulda, A. M. Gorman, O. Hori and A. Samali, Cellular stress responses: cell survival and cell death, Int. J. Cell Mol. Biol., 2010, 2010 Search PubMed.
  46. M. Wang and R. J. Kaufman, Protein misfolding in the endoplasmic reticulum as a conduit to human disease, Nature, 2016, 529, 326–335 CrossRef CAS PubMed.
  47. P. Pallepati and D. A. Averill-Bates, Activation of ER stress and apoptosis by hydrogen peroxide in HeLa cells: protective role of mild heat preconditioning at 40 C, Biochim. Biophys. Acta Mol. Cell Res., 2011, 1813, 1987–1999 CrossRef CAS PubMed.
  48. T. Nakagawa and J. Yuan, Cross-talk between two cysteine protease families: activation of caspase-12 by calpain in apoptosis, J. Cell Biol., 2000, 150, 887–894 CrossRef CAS PubMed.
  49. F. Collin, Chemical basis of reactive oxygen species reactivity and involvement in neurodegenerative diseases, Int. J. Mol. Sci., 2019, 20, 2407 CrossRef PubMed.
  50. J. Mates, Effects of antioxidant enzymes in the molecular control of reactive oxygen species toxicology, Toxicology, 2000, 153, 83–104 CrossRef CAS PubMed.
  51. Y. Chu, R. S. Lan, R. Huang, H. Feng, R. Kumar, S. Dayal, K.-S. Chan and D.-F. Dai, Glutathione peroxidase-1 overexpression reduces oxidative stress, and improves pathology and proteome remodeling in the kidneys of old mice, Aging Cell, 2020, 19(6), e13154 CrossRef CAS PubMed.
  52. E. Grignard, J. Morin, P. Vernet and J. R. Drevet, GPX5 orthologs of the mouse epididymis-restricted and sperm-bound selenium-independent glutathione peroxidase are not expressed with the same quantitative and spatial characteristics in large domestic animals, Theriogenology, 2005, 64, 1016–1033 CrossRef CAS PubMed.
  53. C. C. Winterbourn and M. B. Hampton, Thiol chemistry and specificity in redox signaling, Free Radic. Biol. Med., 2008, 45, 549–561 CrossRef CAS PubMed.
  54. M. Fransen, M. Nordgren, B. Wang and O. Apanasets, Role of peroxisomes in ROS/RNS-metabolism: implications for human disease, Biochim. Biophys. Acta Mol. Basis Dis., 2012, 1822, 1363–1373 CrossRef CAS PubMed.
  55. S. Mani, Production of reactive oxygen species and its implication in human diseases, in: Free Radicals in Human Health and Disease, Springer, 2015, pp. 3–15 Search PubMed.
  56. J. S. Beckman, T. W. Beckman, J. Chen, P. A. Marshall and B. A. Freeman, Apparent hydroxyl radical production by peroxynitrite: implications for endothelial injury from nitric oxide and superoxide, Proc. Natl. Acad. Sci. U. S. A., 1990, 87, 1620–1624 CrossRef CAS PubMed.
  57. L. P. Olson, M. D. Bartberger and K. Houk, Peroxynitrate and peroxynitrite: A complete basis set investigation of similarities and differences between these NO x species, J. Am. Chem. Soc., 2003, 125, 3999–4006 CrossRef CAS PubMed.
  58. S. Raj Rai, C. Bhattacharyya, A. Sarkar, S. Chakraborty, E. Sircar, S. Dutta and R. Sengupta, Glutathione: Role in Oxidative/Nitrosative Stress, Antioxidant Defense, and Treatments, ChemistrySelect, 2021, 6, 4566–4590 CrossRef CAS.
  59. I. Batinić-Haberle, I. Spasojević, R. D. Stevens, P. Hambright, P. Neta, A. Okado-Matsumoto and I. Fridovich, New class of potent catalysts of O 2 dismutation. Mn (iii) ortho-methoxyethylpyridyl-and di-ortho-methoxyethylimidazolylporphyrins, Dalton Trans., 2004, 1696–1702 RSC.
  60. M. W. Bowler, M. G. Montgomery, A. G. Leslie and J. E. Walker, How azide inhibits ATP hydrolysis by the F-ATPases, Proc. Natl. Acad. Sci. U. S. A., 2006, 103, 8646–8649 CrossRef CAS PubMed.
  61. S. Bolisetty and E. A. Jaimes, Mitochondria and reactive oxygen species: physiology and pathophysiology, Int. J. Mol. Sci., 2013, 14, 6306–6344 CrossRef CAS PubMed.
  62. B. Drew and C. Leeuwenburgh, Aging and the role of reactive nitrogen species, Ann. N. Y. Acad. Sci., 2002, 959, 66–81 CrossRef CAS PubMed.
  63. J. Lee, N. Koo and D. B. Min, Reactive oxygen species, aging, and antioxidative nutraceuticals, Compr. Rev. Food Sci. Food Saf., 2004, 3, 21–33 CrossRef CAS PubMed.
  64. A. A. Alfadda and R. M. Sallam, Reactive oxygen species in health and disease, J. Biomed. Biotechnol., 2012, 2012 Search PubMed.
  65. T. Chitnis and H. L. Weiner, CNS inflammation and neurodegeneration, J. Clin. Investig., 2017, 127, 3577–3587 CrossRef PubMed.
  66. M. T. Lin and M. F. Beal, Mitochondrial dysfunction and oxidative stress in neurodegenerative diseases, Nature, 2006, 443, 787–795 CrossRef CAS PubMed.
  67. A. W. K. Yeung, N. T. Tzvetkov, M. G. Georgieva, I. V. Ognyanov, K. Kordos, A. Jóźwik, T. Kühl, G. Perry, M. C. Petralia and E. Mazzon, Reactive oxygen species and their impact in neurodegenerative diseases: literature landscape analysis, Antioxidants Redox Signal., 2021, 34, 402–420 CrossRef CAS PubMed.
  68. C. H.-L. Hung, S. S.-Y. Cheng, Y.-T. Cheung, S. Wuwongse, N. Q. Zhang, Y.-S. Ho, S. M.-Y. Lee and R. C.-C. Chang, A reciprocal relationship between reactive oxygen species and mitochondrial dynamics in neurodegeneration, Redox Biol., 2018, 14, 7–19 CrossRef CAS PubMed.
  69. K. A. Jellinger, Basic mechanisms of neurodegeneration: a critical update, J. Cell Mol. Med., 2010, 14, 457–487 CrossRef CAS PubMed.
  70. I. P. Johnson, Age-related neurodegenerative disease research needs aging models, Front. Aging Neurosci., 2015, 7, 168 Search PubMed.
  71. P. Di Rosanna and C. Salvatore, Reactive oxygen species, inflammation, and lung diseases, Curr. Pharm. Des., 2012, 18, 3889–3900 CrossRef PubMed.
  72. B. Uttara, A. V. Singh, P. Zamboni and R. Mahajan, Oxidative stress and neurodegenerative diseases: a review of upstream and downstream antioxidant therapeutic options, Curr. Neuropharmacol., 2009, 7, 65–74 CrossRef CAS PubMed.
  73. A. Popa-Wagner, S. Mitran, S. Sivanesan, E. Chang and A.-M. Buga, ROS and brain diseases: the good, the bad, and the ugly, Oxidative Medicine and Cellular Longevity, 2013 Search PubMed.
  74. W. Ahmad, B. Ijaz, K. Shabbiri, F. Ahmed and S. Rehman, Oxidative toxicity in diabetes and Alzheimer's disease: mechanisms behind ROS/RNS generation, J. Biomed. Sci., 2017, 24, 1–10 CrossRef PubMed.
  75. M. Brownlee, The pathobiology of diabetic complications: a unifying mechanism, Diabetes, 2005, 54, 1615–1625 CrossRef CAS PubMed.
  76. J. Hanus, H. Zhang, Z. Wang, Q. Liu, Q. Zhou and S. Wang, Induction of necrotic cell death by oxidative stress in retinal pigment epithelial cells, Cell Death Dis., 2013, 4, e965 CrossRef CAS PubMed.
  77. D. N. Granger, T. Vowinkel and T. Petnehazy, Modulation of the inflammatory response in cardiovascular disease, Hypertension, 2004, 43, 924–931 CrossRef CAS PubMed.
  78. H. Wang, X. Han, E. S. Wittchen and M. E. Hartnett, TNF-α mediates choroidal neovascularization by upregulating VEGF expression in RPE through ROS-dependent β-catenin activation, Mol. Vis., 2016, 22, 116 CAS.
  79. G. Ghosh, I. V. Subramanian, N. Adhikari, X. Zhang, H. P. Joshi, D. Basi, Y. Chandrashekhar, J. L. Hall, S. Roy and Y. Zeng, Hypoxia-induced microRNA-424 expression in human endothelial cells regulates HIF-α isoforms and promotes angiogenesis, J. Clin. Investig., 2010, 120, 4141–4154 CrossRef CAS PubMed.
  80. T. P. Cash, Y. Pan and M. C. Simon, Reactive oxygen species and cellular oxygen sensing, Free Radic. Biol. Med., 2007, 43, 1219–1225 CrossRef CAS PubMed.
  81. X. Zhang, Y. Fu, X. Xu, M. Li, L. Du, Y. Han and Y. Ge, PERK pathway are involved in NO-induced apoptosis in endothelial cells cocultured with RPE under high glucose conditions, Nitric Oxide, 2014, 40, 10–16 CrossRef CAS PubMed.
  82. K. M. Nash and S. Ahmed, Nanomedicine in the ROS-mediated pathophysiology: Applications and clinical advances, Nanomed. Nanotechnol. Biol. Med., 2015, 11, 2033–2040 CrossRef CAS PubMed.
  83. C. A. Ferreira, D. Ni, Z. T. Rosenkrans and W. Cai, Scavenging of reactive oxygen and nitrogen species with nanomaterials, Nano Res., 2018, 11, 4955–4984 CrossRef CAS PubMed.
  84. Z. T. Rosenkrans, T. Sun, D. Jiang, W. Chen, T. E. Barnhart, Z. Zhang, C. A. Ferreira, X. Wang, J. W. Engle and P. Huang, Selenium-Doped Carbon Quantum Dots Act as Broad-Spectrum Antioxidants for Acute Kidney Injury Management, Adv. Sci., 2020, 7, 2000420 CrossRef CAS PubMed.
  85. K. Fan, J. Xi, L. Fan, P. Wang, C. Zhu, Y. Tang, X. Xu, M. Liang, B. Jiang and X. Yan, In vivo guiding nitrogen-doped carbon nanozyme for tumor catalytic therapy, Nat. Commun., 2018, 9, 1–11 CrossRef CAS PubMed.
  86. A. Watanabe, M. Kajita, J. Kim, A. Kanayama, K. Takahashi, T. Mashino and Y. Miyamoto, In vitro free radical scavenging activity of platinum nanoparticles, Nanotechnology, 2009, 20, 455105 CrossRef PubMed.
  87. M. Soh, D. W. Kang, H. G. Jeong, D. Kim, D. Y. Kim, W. Yang, C. Song, S. Baik, I. Y. Choi and S. K. Ki, Ceria–Zirconia nanoparticles as an enhanced multi-antioxidant for sepsis treatment, Angew. Chem., 2017, 129, 11557–11561 CrossRef.
  88. T. Liu, B. Xiao, F. Xiang, J. Tan, Z. Chen, X. Zhang, C. Wu, Z. Mao, G. Luo and X. Chen, Ultrasmall copper-based nanoparticles for reactive oxygen species scavenging and alleviation of inflammation related diseases, Nat. Commun., 2020, 11, 1–16 CrossRef PubMed.
  89. J. Zhao, W. Gao, X. Cai, J. Xu, D. Zou, Z. Li, B. Hu and Y. Zheng, Nanozyme-mediated catalytic nanotherapy for inflammatory bowel disease, Theranostics, 2019, 9, 2843 CrossRef CAS PubMed.
  90. Y. Dong and Z. Wang, ROS-scavenging materials for skin wound healing: advancements and applications, Front. Bioeng. Biotechnol., 2023, 11, 1304835 CrossRef PubMed.
  91. C. A. Ferreira, D. Ni, Z. T. Rosenkrans and W. J. N. r. Cai, Scavenging of reactive oxygen and nitrogen species with nanomaterials, Nano Res., 2018, 11(10), 4955–4984 CrossRef CAS PubMed.
  92. T. Fang, J. Zhang, T. Zuo, G. Wu, Y. Xu, Y. Yang, J. Yang and Q. Shen, Chemo-Photothermal Combination Cancer Therapy with ROS Scavenging, Extracellular Matrix Depletion, and Tumor Immune Activation by Telmisartan and Diselenide-Paclitaxel Prodrug Loaded Nanoparticles, ACS Appl. Mater. Interfaces, 2020, 12, 31292–31308 CrossRef CAS PubMed.
  93. J. Kim, H. Park, G. Saravanakumar and W. J. Kim, Polymer/Aptamer-Integrated Gold Nanoconstruct Suppresses the Inflammatory Process by Scavenging ROS and Capturing Pro-inflammatory Cytokine TNF-α, ACS Appl. Mater. Interfaces, 2020, 13, 9390–9401 CrossRef PubMed.
  94. J. Li, J. Zhang, Y. Chen, N. Kawazoe and G. Chen, TEMPO-conjugated gold nanoparticles for reactive oxygen species scavenging and regulation of stem cell differentiation, ACS Appl. Mater. Interfaces, 2017, 9, 35683–35692 CrossRef CAS PubMed.
  95. H. Lee, K. Lee, I. K. Kim and T. G. Park, Fluorescent gold nanoprobe sensitive to intracellular reactive oxygen species, Adv. Funct. Mater., 2009, 19, 1884–1890 CrossRef CAS.
  96. B. Das, P. Pal, P. Dadhich, J. Dutta and S. J. A. B. Dhara, In vivo cell tracking, reactive oxygen species scavenging, and antioxidative gene down regulation by long-term exposure of biomass-derived carbon dots, ACS Biomater. Sci. Eng., 2018, 5(1), 346–356 CrossRef PubMed.
  97. X. Mu, H. He, J. Wang, W. Long, Q. Li, H. Liu, Y. Gao, L. Ouyang, Q. Ren, S. Sun, J. Wang, J. Yang, Q. Liu, Y. Sun, C. Liu, X. D. Zhang and W. Hu, Carbogenic Nanozyme with Ultrahigh Reactive Nitrogen Species Selectivity for Traumatic Brain Injury, Nano Lett., 2019, 19, 4527–4534 CrossRef CAS PubMed.
  98. D. Yim, D. E. Lee, Y. So, C. Choi, W. Son, K. Jang, C. S. Yang and J. H. Kim, Sustainable Nanosheet Antioxidants for Sepsis Therapy via Scavenging Intracellular Reactive Oxygen and Nitrogen Species, ACS Nano, 2020, 14, 10324–10336 CrossRef CAS PubMed.
  99. G. Zhong, X. Yang, X. Jiang, A. Kumar, H. Long, J. Xie, L. Zheng and J. Zhao, Dopamine-melanin nanoparticles scavenge reactive oxygen and nitrogen species and activate autophagy for osteoarthritis therapy, Nanoscale, 2019, 11, 11605–11616 RSC.
  100. J. M. Dowding, S. Seal and W. T. Self, Cerium oxide nanoparticles accelerate the decay of peroxynitrite (ONOO), Drug Delivery Transl. Res., 2013, 3, 375–379 CrossRef CAS PubMed.
  101. C. Minnelli, F. Spinozzi, P. Moretti, G. Mobbili, E. Romagnoli, E. Pavoni, E. Mohebbi, B. Romaldi, T. Armeni, P. Stipa and E. Laudadio, Tunable assembly of mixed PLGA-lipid nanoparticles via monoolein with improved reactive oxygen species cell protection, Colloids Surf., A, 2025, 718, 136864 CrossRef CAS.
  102. F. Wang, X. Qu, D. Liu, C. Ding, C. Zhang and Y. Xian, Upconversion nanoparticles-MoS2 nanoassembly as a fluorescent turn-on probe for bioimaging of reactive oxygen species in living cells and zebrafish, Sens. Actuators, B, 2018, 274, 180–187 CrossRef CAS.
  103. T. Zhang, F. Lin, W. Liu, Y. Liu, Z. Guo, C. Xiao, X. Zhuang and X. Chen, Reactive oxide species-scavenging lipid-polymer nanoparticles for neuroprotection after spinal cord injury, Appl. Mater. Today, 2021, 24, 101109 CrossRef.
  104. Y. Li, X. Jin, Y. Sun, S. Guan, B. Chang and Y. Li, Dual-mode nanoprobes for reactive oxygen species scavenging and treatment of acute lung injury, Appl. Surf. Sci., 2023, 610, 155458 CrossRef CAS.
  105. V. I. Lushchak, Free radicals, reactive oxygen species, oxidative stress and its classification, Chem.-Biol. Interact., 2014, 224, 164–175 CrossRef CAS PubMed.
  106. B. Perillo, M. Di Donato, A. Pezone, E. Di Zazzo, P. Giovannelli, G. Galasso, G. Castoria and A. Migliaccio, ROS in cancer therapy: The bright side of the moon, Exp. Mol. Med., 2020, 52, 192–203 CrossRef CAS PubMed.
  107. D. B. Zorov, M. Juhaszova and S. J. Sollott, Mitochondrial reactive oxygen species (ROS) and ROS-induced ROS release, Physiol. Rev., 2014, 94, 909–950 CrossRef CAS PubMed.
  108. J. Zhang, X. Wang, V. Vikash, Q. Ye, D. Wu, Y. Liu and W. Dong, ROS and ROS-mediated cellular signaling, Oxid. Med. Cell. Longevity, 2016, 4350965 CrossRef PubMed.
  109. S. Singh, Antioxidant nanozymes as next-generation therapeutics to free radical-mediated inflammatory diseases: A comprehensive review, Int. J. Biol. Macromol., 2024, 260, 129374 CrossRef CAS PubMed.
  110. F. Scialò, D. J. Fernández-Ayala and A. Sanz, Role of mitochondrial reverse electron transport in ROS signaling: potential roles in health and disease, Front. Physiol., 2017, 8, 428 CrossRef PubMed.
  111. A. El-Kenawi and B. Ruffell, Inflammation, ROS, and mutagenesis, Cancer Cell, 2017, 32, 727–729 CrossRef CAS PubMed.
  112. J. N. Moloney, T. G. Cotter, ROS signalling in the biology of cancer, in: Seminars in Cell & Developmental Biology, Elsevier, 2018, pp. 50–64 Search PubMed.
  113. V. Aggarwal, H. S. Tuli, A. Varol, F. Thakral, M. B. Yerer, K. Sak, M. Varol, A. Jain, M. Khan and G. Sethi, Role of reactive oxygen species in cancer progression: molecular mechanisms and recent advancements, Biomolecules, 2019, 9, 735 CrossRef CAS PubMed.
  114. S. Kalmodia, S. Vandhana, B. T. Rama, B. Jayashree, T. S. Seethalakshmi, V. Umashankar, W. Yang, C. J. Barrow, S. Krishnakumar and S. V. J. C. Elchuri, Bio-conjugation of antioxidant peptide on surface-modified gold nanoparticles: a novel approach to enhance the radical scavenging property in, Cancer Cell, 2016, 7, 1–19 Search PubMed.
  115. X. Xu, P. E. Saw, W. Tao, Y. Li, X. Ji, S. Bhasin, Y. Liu, D. Ayyash, J. Rasmussen and M. J. A. M. Huo, ROS-responsive polyprodrug nanoparticles for triggered drug delivery and effective cancer therapy, Adv. Mater., 2017, 29(33) DOI:10.1002/adma.201700141.
  116. M. Aioub, S. R. Panikkanvalappil and M. A. J. El-Sayed, Platinum-coated gold nanorods: efficient reactive oxygen scavengers that prevent oxidative damage toward healthy, untreated cells during plasmonic photothermal therapy, ACS Nano, 2017, 11(1), 579–586 CrossRef CAS PubMed.
  117. K. Zhang, Z. Yang, X. Meng, Y. Cao, Y. Zhang, W. Dai, H. Lu, Z. Yu, H. Dong and X. J. M. C. Zhang, Peroxidase-like Fe 3 O 4 nanocomposite for activatable reactive oxygen species generation and, Cancer Theranostics, 2018, 2, 1184–1194 CAS.
  118. S. Wu, R. Han, Y. Yan, W. Chen, W. Gao, Y. Hou and K. Tang, Porous CeO2 nanorods loaded with indocyanine green for enhanced tumor-specific therapy, Microporous Mesoporous Mater., 2021, 315, 110905 CrossRef CAS.
  119. H.-Y. Cho, A. Mavi, S.-T. D. Chueng, T. Pongkulapa, N. Pasquale, H. Rabie, J. Han, J. H. Kim, T.-H. Kim, J.-W. Choi and K.-B. Lee, Tumor Homing Reactive Oxygen Species Nanoparticle for Enhanced Cancer Therapy, ACS Appl. Mater. Interfaces, 2019, 11, 23909–23918 CrossRef CAS PubMed.
  120. H. Deng, W. Yang, Z. Zhou, R. Tian, L. Lin, Y. Ma, J. Song and X. Chen, Targeted scavenging of extracellular ROS relieves suppressive immunogenic cell death, Nat. Commun., 2020, 11, 1–12 CrossRef PubMed.
  121. K. Dehvari, S.-H. Chiu, J.-S. Lin, W. M. Girma, Y.-C. Ling and J.-Y. Chang, Heteroatom doped carbon dots with nanoenzyme like properties as theranostic platforms for free radical scavenging, imaging, and chemotherapy, Acta Biomater., 2020, 114, 343–357 CrossRef CAS PubMed.
  122. J.-S. Lin, Y.-W. Tsai, K. Dehvari, C.-C. Huang and J.-Y. Chang, A carbon dot based theranostic platform for dual-modal imaging and free radical scavenging, Nanoscale, 2019, 11, 20917–20931 RSC.
  123. M. Passi, V. Kumar and G. Packirisamy, Theranostic nanozyme: Silk fibroin based multifunctional nanocomposites to combat oxidative stress, Mater. Sci. Eng. C, 2020, 107, 110255 CrossRef CAS PubMed.
  124. H. Zhu, Y. Fang, Q. Miao, X. Qi, D. Ding, P. Chen and K. Pu, Regulating near-infrared photodynamic properties of semiconducting polymer nanotheranostics for optimized cancer therapy, ACS Nano, 2017, 11, 8998–9009 CrossRef CAS PubMed.
  125. A. D. Pandya, E. Jäger, S. B. Fam, A. Höcherl, A. Jäger, V. Sincari, B. Nyström, P. Štěpánek, T. Skotland and K. Sandvig, Paclitaxel-loaded biodegradable ROS-sensitive nanoparticles for cancer therapy, Int. J. Nanomed., 2019, 14, 6269 CrossRef CAS PubMed.
  126. N. M. Zholobak, A. B. Shcherbakov, E. O. Vitukova, A. V. Yegorova, Y. V. Scripinets, I. I. Leonenko, A. Y. Baranchikov, V. P. Antonovich and V. K. J. R. Ivanov, Direct monitoring of the interaction between ROS and cerium dioxide nanoparticles in living cells, RSC Adv., 2014, 4, 51703–51710 RSC.
  127. K. Zhang, Z. Yang, X. Meng, Y. Cao, Y. Zhang, W. Dai, H. Lu, Z. Yu, H. Dong and X. Zhang, Peroxidase-like Fe 3 O 4 nanocomposite for activatable reactive oxygen species generation and cancer theranostics, Mater. Chem. Front., 2018, 2, 1184–1194 RSC.
  128. Y. Tong, L. Zhang, R. Gong, J. Shi, L. Zhong, X. Duan and Y. Zhu, A ROS-scavenging multifunctional nanoparticle for combinational therapy of diabetic nephropathy, Nanoscale, 2020, 12, 23607–23619 RSC.
  129. S. Chigurupati, M. R. Mughal, E. Okun, S. Das, A. Kumar, M. McCaffery, S. Seal and M. P. Mattson, Effects of cerium oxide nanoparticles on the growth of keratinocytes, fibroblasts and vascular endothelial cells in cutaneous wound healing, Biomaterials, 2013, 34, 2194–2201 CrossRef CAS PubMed.
  130. T. Ma, X. Zhai, Y. Huang, M. Zhang, X. Zhao, Y. Du and C. Yan, A Smart Nanoplatform with Photothermal Antibacterial Capability and Antioxidant Activity for Chronic Wound Healing, Adv. Healthcare Mater., 2021, 2100033 CrossRef CAS PubMed.
  131. R. Bal, G. Türk, M. Tuzcu, O. Yilmaz, I. Ozercan, T. Kuloglu, S. Gür, V. S. Nedzvetsky, A. A. Tykhomyrov and G. V. Andrievsky, Protective effects of nanostructures of hydrated C60 fullerene on reproductive function in streptozotocin-diabetic male rats, Toxicology, 2011, 282, 69–81 CrossRef CAS PubMed.
  132. F. J. Blanco, I. Rego and C. Ruiz-Romero, The role of mitochondria in osteoarthritis, Nat. Rev. Rheumatol., 2011, 7, 161–169 CrossRef CAS PubMed.
  133. J. Bai, Y. Zhang, Q. Fan, J. Xu, H. Shan, X. Gao, Q. Ma, L. Sheng, X. Zheng and W. Cheng, Reactive Oxygen Species-Scavenging Scaffold with Rapamycin for Treatment of Intervertebral Disk Degeneration, Adv. Healthcare Mater., 2020, 9, 1901186 CrossRef CAS PubMed.
  134. Q. Chen, J. Wu, Y. Liu, Y. Li, C. Zhang, W. Qi, K. W. Yeung, T. M. Wong, X. Zhao and H. Pan, Electrospun chitosan/PVA/bioglass Nanofibrous membrane with spatially designed structure for accelerating chronic wound healing, Mater. Sci. Eng. C, 2019, 105, 110083 CrossRef CAS PubMed.
  135. V. Chaudhary, E. Mostafavi and A. Kaushik, De-coding Ag as an efficient antimicrobial nano-system for controlling cellular/biological functions, Matter, 2022, 5, 1995–1998 CrossRef CAS.
  136. C. Hu, F. Zhang, Q. Kong, Y. Lu, B. Zhang, C. Wu, R. Luo and Y. Wang, Synergistic chemical and photodynamic antimicrobial therapy for enhanced wound healing mediated by multifunctional light-responsive nanoparticles, Biomacromolecules, 2019, 20, 4581–4592 CrossRef CAS PubMed.
  137. M. Cai, G. Chen, L. Qin, C. Qu, X. Dong, J. Ni and X. Yin, Metal organic frameworks as drug targeting delivery vehicles in the treatment of cancer, Pharmaceutics, 2020, 12, 232 CrossRef CAS PubMed.
  138. P. Yu, L. Zheng, P. Wang, S. Chai, Y. Zhang, T. Shi, L. Zhang, R. Peng, C. Huang and B. Guo, Development of a novel polysaccharide-based iron oxide nanoparticle to prevent iron accumulation-related osteoporosis by scavenging reactive oxygen species, Int. J. Biol. Macromol., 2020, 165, 1634–1645 CrossRef CAS PubMed.
  139. P. Bhushan, V. Kamat, I. Abrol, A. Kaushik and S. Bhansali, Bio-acceptability of wearable sensors: A mechanistic study towards evaluating ionic leaching induced cellular inflammation, Sci. Rep., 2022, 12, 10782 CrossRef CAS PubMed.
  140. H. Blaser, C. Dostert, T. W. Mak and D. Brenner, TNF and ROS crosstalk in inflammation, Trends Cell Biol., 2016, 26, 249–261 CrossRef CAS PubMed.
  141. M. Shen, F. Forghani, X. Kong, D. Liu, X. Ye, S. Chen and T. Ding, Antibacterial applications of metal–organic frameworks and their composites, Compr. Rev. Food Sci. Food Saf., 2020, 19, 1397–1419 CrossRef CAS PubMed.
  142. O. Kutuk and H. Basaga, Inflammation meets oxidation: NF-κB as a mediator of initial lesion development in atherosclerosis, Trends Mol. Med., 2003, 9, 549–557 CrossRef CAS PubMed.
  143. A. Dhar, M. R. Young and N. H. Colburn, The role of AP-1, NF-κB and ROS/NOS in skin carcinogenesis: the JB6 model is predictive, Mol. Cell. Biochem., 2002, 234, 185–193 CrossRef PubMed.
  144. S. D. Hettiarachchi, Y. Zhou, E. Seven, M. K. Lakshmana, A. K. Kaushik, H. S. Chand and R. M. Leblanc, Nanoparticle-mediated approaches for Alzheimer's disease pathogenesis, diagnosis, and therapeutics, J. Contr. Release, 2019, 314, 125–140 CrossRef CAS PubMed.
  145. D. Kim, H. J. Kwon and T. Hyeon, Magnetite/Ceria Nanoparticle Assemblies for Extracorporeal Cleansing of Amyloid-β in Alzheimer's Disease, Adv. Mater., 2019, 31, 1807965 CrossRef PubMed.
  146. C. Hao, A. Qu, L. Xu, M. Sun, H. Zhang, C. Xu and H. Kuang, Chiral Molecule-mediated Porous Cu x O Nanoparticle Clusters with Antioxidation Activity for Ameliorating Parkinson's Disease, J. Am. Chem. Soc., 2018, 141, 1091–1099 CrossRef PubMed.
  147. H. J. Kwon, M.-Y. Cha, D. Kim, D. K. Kim, M. Soh, K. Shin, T. Hyeon and I. Mook-Jung, Mitochondria-targeting ceria nanoparticles as antioxidants for Alzheimer's disease, ACS Nano, 2016, 10, 2860–2870 CrossRef CAS PubMed.
  148. J. W. Kim, C. Mahapatra, J. Y. Hong, M. S. Kim, K. W. Leong, H. W. Kim and J. K. Hyun, Functional Recovery of Contused Spinal Cord in Rat with the Injection of Optimal-Dosed Cerium Oxide Nanoparticles, Adv. Sci., 2017, 4, 1700034 CrossRef PubMed.
  149. N. Singh, M. A. Savanur, S. Srivastava, P. D'Silva and G. Mugesh, A redox modulatory Mn3O4 nanozyme with multi-enzyme activity provides efficient cytoprotection to human cells in a Parkinson's disease model, Angew. Chem., 2017, 129, 14455–14459 CrossRef.
  150. Q. Chen, Y. Du, K. Zhang, Z. Liang, J. Li, H. Yu, R. Ren, J. Feng, Z. Jin and F. Li, Tau-targeted multifunctional nanocomposite for combinational therapy of Alzheimer's disease, ACS Nano, 2018, 12, 1321–1338 CrossRef CAS PubMed.
  151. Q. Han, S. Cai, L. Yang, X. Wang, C. Qi, R. Yang and C. Wang, Molybdenum disulfide nanoparticles as multifunctional inhibitors against Alzheimer's disease, ACS Appl. Mater. Interfaces, 2017, 9, 21116–21123 CrossRef CAS PubMed.
  152. Y. Liu, K. Ai, X. Ji, D. Askhatova, R. Du, L. Lu and J. Shi, 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(2), 856–862 CrossRef CAS PubMed.
  153. J. A. Vassie, J. M. Whitelock and M. S. Lord, Targeted Delivery and Redox Activity of Folic Acid-Functionalized Nanoceria in Tumor Cells, Mol. Pharmaceutics, 2018, 15(3), 994–1004 CrossRef CAS PubMed.
  154. L. Alili, M. Sack, C. von Montfort, S. Giri, S. Das, K. S. Carroll, K. Zanger, S. Seal and P. Brenneisen, Downregulation of tumor growth and invasion by redox-active nanoparticles, Antioxidants Redox Signal., 2013, 19, 765–778 CrossRef CAS PubMed.
  155. M. Soh, D. W. Kang, H. G. Jeong, D. Kim, D. Y. Kim, W. Yang, C. Song, S. Baik, I. Y. Choi, S. K. Ki, H. J. Kwon, T. Kim, C. K. Kim, S. H. Lee and T. Hyeon, Ceria-Zirconia Nanoparticles as an Enhanced Multi-Antioxidant for Sepsis Treatment, Angew. Chem., 2017, 56, 11399–11403 CrossRef CAS PubMed.
  156. J. Xu, M. Ypma, P. A. Chiarelli, J. Park, R. G. Ellenbogen, P. S. Stayton, P. D. Mourad, D. Lee, A. J. Convertine and F. M. Kievit, Theranostic oxygen reactive polymers for treatment of traumatic brain injury, Adv. Funct. Mater., 2016, 26, 4124–4133 CrossRef CAS.
  157. W. Zhu, Z. Dong, T. Fu, J. Liu, Q. Chen, Y. Li, R. Zhu, L. Xu and Z. Liu, Modulation of hypoxia in solid tumor microenvironment with MnO2 nanoparticles to enhance photodynamic therapy, Adv. Funct. Mater., 2016, 26, 5490–5498 CrossRef CAS.
  158. D. Yoo, A. W. Magsam, A. M. Kelly, P. S. Stayton, F. M. Kievit and A. J. Convertine, Core-cross-Linked nanoparticles reduce neuroinflammation and improve outcome in a mouse model of traumatic brain injury, ACS Nano, 2017, 11, 8600–8611 CrossRef CAS PubMed.
  159. O. Rajkovic, C. Gourmel, R. d'Arcy, R. Wong, I. Rajkovic, N. Tirelli and E. Pinteaux, Reactive Oxygen Species-Responsive Nanoparticles for the Treatment of Ischemic Stroke, Adv. Therapeut., 2019, 2, 1900038 CrossRef.
  160. C. K. Kim, T. Kim, I. Y. Choi, M. Soh, D. Kim, Y. J. Kim, H. Jang, H. S. Yang, J. Y. Kim and H. K. Park, Ceria nanoparticles that can protect against ischemic stroke, Angew. Chem., Int. Ed., 2012, 51, 11039–11043 CrossRef CAS PubMed.
  161. M. A. Dkhil, R. Zrieq, S. Al-Quraishy and A. E. Abdel Moneim, Selenium Nanoparticles Attenuate Oxidative Stress and Testicular Damage in Streptozotocin-Induced Diabetic Rats, Molecules, 2016, 21 Search PubMed.
  162. S. Giri, A. Karakoti, R. P. Graham, J. L. Maguire, C. M. Reilly, S. Seal, R. Rattan and V. Shridhar, Nanoceria: a rare-earth nanoparticle as a novel anti-angiogenic therapeutic agent in ovarian cancer, PloS one, 2013, 8, e54578 CrossRef CAS PubMed.
  163. J. A. Vassie, J. M. Whitelock and M. S. Lord, Endocytosis of cerium oxide nanoparticles and modulation of reactive oxygen species in human ovarian and colon cancer cells, Acta Biomater., 2017, 50, 127–141 CrossRef CAS PubMed.

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