Skip to main content

Reactive oxygen species-scavenging nanomaterials for the prevention and treatment of age-related diseases

Abstract

With increasing proportion of the elderly in the population, age-related diseases (ARD) lead to a considerable healthcare burden to society. Prevention and treatment of ARD can decrease the negative impact of aging and the burden of disease. The aging rate is closely associated with the production of high levels of reactive oxygen species (ROS). ROS-mediated oxidative stress in aging triggers aging-related changes through lipid peroxidation, protein oxidation, and DNA oxidation. Antioxidants can control autoxidation by scavenging free radicals or inhibiting their formation, thereby reducing oxidative stress. Benefiting from significant advances in nanotechnology, a large number of nanomaterials with ROS-scavenging capabilities have been developed. ROS-scavenging nanomaterials can be divided into two categories: nanomaterials as carriers for delivering ROS-scavenging drugs, and nanomaterials themselves with ROS-scavenging activity. This study summarizes the current advances in ROS-scavenging nanomaterials for prevention and treatment of ARD, highlights the potential mechanisms of the nanomaterials used and discusses the challenges and prospects for their applications.

Graphical Abstract

Introduction

With the ameliorating sanitary conditions and the continuous development of the economic status, population aging becomes a major global phenomenon revealing a steady increase in life expectancy among geriatrics. According to the World Population Prospects 2019 issued by United Nations, the global population of people over 65 years is expected to increase from 703 million to 1.5 billion by 2050 (from approximately 9% in 2019 to nearly 16% in 2050) [1]. Aging has become a major health concern in the world as it is the primary driver for most chronic diseases, such as cardiovascular diseases, type 2 diabetes, glaucoma, obesity, Alzheimer’s disease, Parkinson’s disease, macular degeneration, and osteoarthritis. Given the increasing aging population and age-related adverse side effects, it is crucial to understand the molecular mechanisms underlying aging and explore more efficient therapeutic strategies [2, 3].

Scientists have been exploring mechanisms of aging and developing methods to postpone senility, including cross-linkage theory of aging [4], the free radical theory [5], telomere shortening theory [6], and immune senescence theory of aging [7]. The free radical theory of aging posits that excessive reactive oxygen species (ROS) and oxidative stress (OS) cause oxidative damage and abnormal functioning of biomolecules (e.g. DNA, proteins and lipids), leading to damage of cells and tissues. In this process, organisms gradually lose their functional and adaptive capacity to the point of aging [5, 8, 9]. ROS are a class of free radicals, including oxidizing substances such as superoxide anion (·O2·), hydrogen peroxide (H2O2), hydroxyl radical (OH-), and singlet oxygen (1O2). OH- are one of the most reactive ROS. OH- and ·O2· can react with other molecules, leading to lipid peroxidation, protein oxidation, and DNA damage, thus triggering cell damage and death. In addition, ·O2· is a precursor to other ROS and can generate more toxic ROS species. Excessive amounts of H2O2 can lead to cytochrome oxidation, protein oxidation and DNA damage, ultimately leading to cell death. 1O2 can react with double-bond-containing biomolecules in organisms to cause oxidative damage. For example, it can trigger lipid peroxidation, which damages the integrity of cell membranes and affects cell function. Consequently, scavenging ROS by antioxidants may be an effective strategy for slowing the progression of aging-related disease. Antioxidants could approximately be classified as hydrophilic antioxidants like vitamin C, glutathione (GSH) and lipophilic antioxidants like vitamin E, carotene, coenzyme Q10 (CoQ10). Recent studies have revealed the pharmacological properties of antioxidants both in vivo and in vitro, while clinical trials involving antioxidants have produced largely disappointing results [10]. Antioxidants’ medicinal potential is constrained by their poor stability and low utilization [10]. Therefore, developing new methods for the application of antioxidants is essential.

ROS-Scavenging nanotechnology has emerged as an exciting and promising new means of treating age-related disease. Nanomaterials (NMs) are particles between 1 and 100 nm in size. Due to their nanoscale size, these particles have greater surface area and higher surface-to-volume ratios, have higher mechanical strength, and are quite stable [11]. NMs can be utilized as medication carriers. In order to preserve tiny molecules from degradation or to facilitate the absorption and distribution of natural antioxidants, polymeric nanoparticles are used to encapsulate or integrate the molecules. Additionally, NMs offer weak water-soluble antioxidants greater solubility and improved surface functionalization to produce target-specificity. Some nanoparticles (NPs) that have a quenching impact on ROS can be used directly as antioxidants. Fullerene (C60) and its derivatives and other inorganic NMs with inherent catalytic characteristics (such as platinum (Pt) and gold (Au)) are examples of common ROS-detoxifying nanoplatforms [12,13,14]. ROS Scavenging nanotechnology show great potential in the prevention and treatment of ARD.

In this perspective, we investigate the use of ROS-scavenging nanotechnology in ARD, discussing its safety, prospective uses, potential applications, and translational challenges in order to promote progress in the development of new treatments.

Reactive oxygen species and the oxidative stress theory of aging

Source of ROS

The free radical theory of aging is predicated on the premise that age-related functional declines are the result of ROS-induced damage accumulation. ROS are a group of oxygen-containing chemical substances that are highly reactive, mainly generated by redox reactions in the organism. ROS are classified as either free radicals or non-free radicals [15]. Free-radical ROS includes superoxide anion radical (·O2·), hydroxyl radical (OH), peroxyl radical (ROO), and sulfhydryl peroxyl radical (RSOO). Non-free radical ROS includes hydrogen peroxide (H2O2), organic hydroperoxides (ROOH), ozone (O3) and singlet oxygen (1O2). Excessive ROS can cause the disruption of the balance between the pro-oxidant and anti-oxidants, leading to OS [16].

ROS are generated in multiple compartments and by a variety of enzymes within the cell, and there are endogenous and exogenous ROS in body [17] (Fig. 1). Endogenous ROS are mainly produced directly by various organelles such as mitochondria, cytoplasmic membrane, endoplasmic reticulum (ER), peroxisomes, and lysosomes. The most significant source of ROS production occurs mainly in the mitochondrial electron transport chain (ETC) complexes I, II and III, due to electron leakage [18, 19]. ETC transfers electrons from NADH to O2 and generate ·O2·, which can then be rapidly broken down to H2O2 by superoxide dismutase (SOD). When Fe2+ and Cu2+ are present, H2O2 can also be converted to ·OH through the Fenton reaction. The protein misfolding process that occurs in ER is also accompanied by the production of ROS [19]. ROS production on ER is generated by delivering electrons to O2 by NADH-cytochrome P450 reductase to form ·O2·, with electrons delivered to O2 by the electron transport chain on the nuclear membrane, assisted by NADH [20, 21]. In addition, various types of oxidase such as NADPH oxidase (NOX), cytochrome P450 (CYP) enzymes, xanthine oxidase (XO), nitric oxide synthase (NOS), which promote the production of endogenous ROS [22, 23]. Hypoxanthine can be converted to xanthine catalyzed by XO in a process accompanied by the reduction of O2 to ·O2·. Endothelial nitric oxide synthase (eNOS) produce ·O2·. Monoamine oxidase, lipoxygenase and cyclooxygenase enzymes, can also promote the production of ROS in normal biological reactions. Additionally, genetic factors can potentially contribute to oxidative stress. The copper (zinc) superoxide dismutase 1 (SOD-1) gene is the most prevalent genetic contributor to amyotropic lateral sclerosis (ALS), which accounts for 5–10% of cases. The mutations in this gene, which enhance oxidative stress in the cells, promote protein deposition, disrupt intracellular calcium ions, and cause the diffusion of toxicity, are responsible for around 20% of familial ALS and 2% of sporadic ALS [24]. Exogenous ROS are induced by external factors including alcohol, cigarette smoke, heavy metals (lead, chromium), industrial solvents, pesticides, medications like halothane and nonsteroidal anti-inflammatory medicines, radiation, and other pollutants such as air and water pollutants [25]. In addition, ischemia–reperfusion (I/R) damage, infections, and inflammation all lead to increased levels of ROS [26].

Fig. 1
figure 1

Categories and source of ROS. Created with BioRender.com

Oxidative damage of ROS in aging

The aging process is a loss of internal homeostasis due to the accumulation of molecular damage to macromolecules such as DNA, lipids and proteins. Under physiological circumstances, the intracellular generation and scavenging of ROS is usually in homeostasis [27]. At low concentrations, ROS participates in cell growth and survival, immune response, metabolic regulation, and cell signaling process [28, 29]. OS is determined by an imbalance between ROS generation and antioxidant defenses, which gradually damages biomolecules including DNA, lipids, and proteins by oxidation [30] (Fig. 2). Harman formulated the free radical theory of aging, indicating that free-radical associated macromolecular damage may promote senescence [5].

Fig. 2
figure 2

Copyright 2020, Elsevier. Copyright 2002, Elsevier

Oxidative damage of protein, lipid and DNA.

Lipid peroxidation

Polyunsaturated fatty acid (PUFA), especially PUFA with more double bonds, such as arachidonic acid and linoleic acid, are highly susceptible to ROS and free radicals [31]. Since PUFA is the main component of cell membranes, cell membranes are vulnerable to free radical damage; when membrane phospholipids come into contact with an inordinate amount of ROS, lipid peroxidation occurs. This extensive lipid peroxidation alters the membrane's structure, reducing its fluidity and compromising its integrity [32]. Moreover, Lipid peroxides are also extremely reactive substances that have the ability to generate more ROS or breakdown into reactive substances that can crosslink proteins and DNA. They interact with free amino groups in proteins, causing them to covalently modify, cross-link, oligomerize, and aggregate. These mechanisms, which produce intracellular damage, decrease cell activities and induce cell death, have been linked to aging and a variety of ARD.

Protein oxidation

Exposure of proteins to ROS results in multiple changes, including amino acid residues oxidation, protein fragmentation due to oxidative cleavage of the peptide backbone, irreversible production of protein carbonyl compounds and generation of protein–protein cross-linkages [33,34,35,36]. With the accumulation of oxidative damage, proteins are more likely to misfold. Moderately oxidized proteins are degraded by the proteasome [37], while severely oxidized proteins can cross-link with other proteins, thus preventing their degradation [38]. As a result, severely damaged proteins accumulate within the cell, altering physiological properties such as loss of catalytic activity and paralysis of regulation of metabolic pathways. It is known that dysfunctions in the cellular apparatus of protein quality control contribute to aging and ARD, such as neurodegenerative and cardiovascular diseases [39].

DNA oxidation

ROS generate major OS when they react with nitrogenous bases and deoxyribose. DNA oxidation damage mainly include base mutation, strand breaking, DNA–protein cross-links, and formation of DNA-adducts [33]. Direct strand excision and oxidative damage to pyrimidine and purine bases are both effects of hydroxyl radical stress on DNA. In addition to oxidizing DNA bases, ROS may potentially disrupt DNA strands by attacking the DNA backbone with free radicals [40, 41]. Furthermore, adducts to DNA can be formed through the reaction of deoxyguanosine and other macromolecular modifications triggered by ROS [42]. In addition, mitochondrial DNA (mtDNA) is highly susceptible to ROS, and has a significantly higher mutation rate than nuclear DNA. Histones and other chromatin-associated proteins present in the nuclear genome, which function as free radical scavengers, but not in the mitochondrial genome [43]. The persistence and accumulation of damaged mtDNA in the mitochondria inevitably lead to more ROS production, which in turn cause further damage. DNA damage can cause aging by affecting transcription, apoptosis signaling or cellular senescence or through somatic mutations and telomere shortening [44,45,46]. Continuous oxidative damage to mtDNA has been linked to aging, inflammation, carcinogenesis, and the development of malignancy [47]. The DNA damage response which consists of the activation of checkpoint pathways, cell cycle arrest and DNA repair, removes most of ROS-induced DNA damage [48].

8-hydroxyguanosine (8-OHG) is the oxidized base that occurs most frequently in RNA. Guanine is initially reacted with by the extremely reactive hydroxyl radical, which subsequently produces 8-OHG after losing an electron (e −) and proton (H +) [49]. The oxidized RNA is substantially intact, while its translation fidelity has been severely diminished. The oxidative alteration of RNA disrupts the translational process and impairs protein synthesis, causing cell degeneration or even cell death [50].

ROS-scavenging nanotechnology and scavenging mechanisms

A new window of opportunity has opened up for the advancement of conventional antioxidant therapy thanks to the recent proliferation of nanotechnology and nanoscience in the construction of ROS scavengers. It is possible to classify ROS-scavenging NMs as carriers for delivering natural antioxidants or nanomaterials with inherent ROS-scavenging activity (Fig. 3).

Fig. 3
figure 3

ROS-scavenging nanomaterials in treatment of ARD. Created with BioRender.com

Nanomaterials for catalytic generation of ROS-scavenging agents and its mechanisms

Nanomaterials with ROS scavenging activity include metals and metal oxides, carbon-based nanomaterials, enzyme-like nanoparticles, selenium and (2,2,6,6-tetramethylpiperidin-1-yl)oxyl (TEMPO). In the following section, we provide an overview of ROS-scavenging nanomaterials, focusing on their distinctive redox properties and mechanisms. Table 1 outlines the key aspects of select ROS scavengers.

Table 1 Nanomaterials with ROS scavenging activity and its mechanisms

Strong catalytic activity is shown by nanoscale noble metal NMs like palladium (Pd), Au, and Pt, which is primarily ascribed to their huge specific surface area and larger fraction of metal atoms on their surfaces [51]. These nanocatalysts of noble metals have been proposed as possible antioxidants. Although AuNPs are not typically considered to possess redox activity, they serve as an ideal platform for electrochemical biosensors. This is because they can function as redox catalysts, thereby enhancing the electron transfer of various electroactive biological species (primarily redox proteins) without necessitating the use of electron transfer mediators [52]. Pt NMs are a viable choice to treat the oxidative damage due to their potent peroxidase POD-, CAT-, and SOD-like nanozyme activities that catalytically convert O2 to H2O2, H2O2 to H2O and O2 [53,54,55,56].

The high redox potential of Prussian blue (PB) NMs is due in large part to their high electron transfer capacity. Using an inflammatory model, Zhang et al. showed that PB NMs had the capacity to prevent or alleviate ROS-induced damage [57]. The antioxidant enzymes POD, CAT, and SOD are responsible for their catalytic activity and, by extension, their capacity to scavenge reactive oxygen species.

Copper (Cu) NMs possess excellent catalytic activity like POD-, CAT-, SOD-, and GSH-like enzyme activities [58]. It improves the body's capacity to rid itself of free radicals by increasing the efficiency with which SOD and other enzymes function [58,59,60].

Manganese (Mn) is an important element that plays a role in several cellular processes and metabolic reactions in the human body. The strong POD-, SOD-, and CAT-like activities of Mn4+ NMs have been shown in a number of different investigations [61,62,63]. Mn NMs (Mn4+) directly catalyze H2O2 to produce O2 and Mn2+. Then, Mn NMs (Mn2+) may imitate SOD function by reacting with ·O2· to produce H2O2. Mn3O4 NMs mimic the function of glutathione peroxidase (GPx), CAT, and SOD [64].

Due to the existence of Ce3+/Ce4+ (oxidized/reduced) and compensatory oxygen vacancies, cerium-based NMs have emerged as one of the most common ROS scavengers, enabling them to release or abstract an electron to neutralize different types of ROS [65,66,67]. In general, CeO2 NMs possess efficient redox activity to convert ·O2· to O2, react with HO·, catalyze the degradation of H2O2, scavenge ONOO − , exhibiting SOD (Ce3+) and CAT (Ce4+) mimetic activity to prevent oxidative injury to cells [68,69,70,71].

Nanomaterials having carbon frameworks, such as graphene, graphdiyne, and C60 and its derivatives, may be among the most prevalent ROS quenchers [72,73,74]. In the previous publication, the antioxidant capabilities of C60 and their derivatives were ascribed to the effectiveness of the C60 molecule, which can eliminate ROS through the C60's delocalized double bond system [75, 76]. C60 extinguishes ROS by accepting unpaired electrons, capable of receiving up to six electrons and accommodating as many as 34 methyl free radicals on the C60 sphere [75]. C60 has SOD-like activity [76].

Selenium (Se) functions as a redox center for GPx. Supplying with Se may raise GPx levels, increase H2O2 decomposition and decrease cell damage [77]. Selenoprotein P (SELENOP) and GPx are two of the antioxidant enzymes that assimilate Se NMs as selenocysteine (SEC). The redox center of these enzymes is the element Se [78, 79].

(2,2,6,6-tetramethylpiperidin-1-yl)oxyl (TEMPO) is a well-known ROS scavenger because it can capture unpaired electrons from other radicals by a single electron on nitroxide, and the redox reaction switch between oxidation states of nitroxide, oxoammonium cation, and hydroxylamine [80]. TEMPO is a membrane-permeable stable nitroxide radical that can scavenge superoxide and performs Fenton reactions and radical–radical recombination [81].

Applications of nanocomposites in ROS-scavenging nanotechnology

A typical tactic for preserving redox equilibrium and minimizing OS damage is the introduction of extracellular ROS scavengers. Vitamin C, Vitamin E, CoQ10, resveratrol, MLT, quercetin, curcumin, H2 and other natural antioxidants make up the majority of the chemicals employed in the creation of antioxidant nanoparticles (Table 1). Nanomaterials can be used to composite not only natural antioxidants but also nano-enzymes to improve antioxidant properties and functionality. In addition to enhancing the stability and bioavailability of ROS scavenging drugs, NMs as delivery vehicles can also achieve targeted and controlled drug delivery. In the meantime, NMs as carriers may reduce the administered dose of medications, thereby minimizing adverse effects. By using a range of delivery vehicles, including liposomes, nanospheres, nanoemulsions and nanocrystals, the delivery techniques of the aforementioned non-enzymatic antioxidants have up till now been extensively explored (Table 2).

Table 2 Nanomaterials as carriers for delivering ROS scavenging drugs

ROS-Scavenging nanotechnology in prevention and treatment of ARD

Therapeutic interventions towards oxidative stress might allow restoring the health and curing the aged-related diseases that share basal processes. Overproduction of ROS leads to oxidative stress, which has been observed in diabetes, cardiovascular disease, idiopathic pulmonary fibrosis, neurodegenerative diseases, skeletal degenerative diseases, skin aging, reproductive system aging, and ocular aging. We focused on the implications of NPs-mediated ROS scavenger systems in aging and age-related diseases to provide insights into a potential intervention that may affect the aging process, and subsequently promote healthy longevity (Fig. 4).

Fig. 4
figure 4

ROS-Scavenging nanotechnology in prevention and treatment of age-related diseases. Created with BioRender.com

Type 2 diabetes mellitus

Type 2 diabetes mellitus (T2DM) is a burden on the geriatric population's health, afflicting almost 25% of those over 65 years old [115]. There are several recognized contributors to the pathophysiology of T2DM. ROS and OS play roles in all of them, including hyperglycemia, hyperlipidemia, inflammation, insulin resistance, and endothelial dysfunction. In hyperglycemic conditions, the polyol pathway attempts to reduce excess glucose to sorbitol by using NADPH. Thus NADPH is unable to produce the reduced GSH responsible for the inhibition of OS [116]. Under diabetic conditions, glucose is easily oxidized, causing the formation of H2O2 and other reactive byproducts [117]. There is evidence from clinical studies that strong correlations exist between the levels of pro-oxidants and OS-induced tissue damage indicators such as oxidation of DNA bases, 4-hydroxy-2-nonenal (HNE) proteins, hydroperoxides, 8-hydroxy-deoxyguanine, and 8-epi-prostaglandin [118,119,120]. Therefore, one of the greatest options to lessen the negative consequences of T2DM is antioxidant medication.

Se NMs have been utilized in conjunction with other nanomaterials to boost their antioxidant properties. Hanaa et al. treated diabetic mice with liposomes-Se (L-Se) nanoparticles. L-Se reduced serum glucose, pancreatic malondialdehyde (MDA), nitric oxide (NO), tumor necrosis factor-α (TNF-α), and prostaglandin F2α (PGF2α) levels. The treated diabetic mice also had higher serum insulin, pancreatic GSH, SOD, CAT, GPx, and GSH reductase (GR) levels [121]. Polysaccharide (RTFP-3)-functionalized Se nanoparticles (RP3-SeNPs) protected pancreatic islet cells in INS-1 cells from oxidative damage in another study. RTFP-3 owed high biocompatibility and biodegradability, while it exhibited antioxidant and α-glucosidase-inhibiting activities. RTFP-3 could generate synergistic effect with SeNPs [122]. The combination of Nanocerium and sodium selenite was verified that improved antioxidant enzymes and decreased oxidative stress more effectively than either alone [123].

Applications of AuNPs and ZnO NPs are being researched feverishly. AuNPs were discovered that they could inhibit lipid peroxidation and regulate antioxidant enzymes such as SOD, CAT, and GPx in diabetic mice. The AuNPs regulate hyperglycemia by scavenging free radicals, inhibiting the formation of ROS, and boosting antioxidant defense enzymes [83, 124]. Additionally, silver-gold nanoparticles (Ag@AuNP) with a core–shell structure were tested on diabetic rats. The Ag@AuNPs had better effects and lower expenses than AuNPs in reduction of blood glucose level and insulin resistance, as well as increasing insulin level [125].

ZnO NPs exhibit high antioxidant capabilities through the scavenging of ROS and the up-regulation of antioxidant enzyme activities. Furthermore, it had a hypoglycemic impact in diabetic mice via enhanced insulin production and glucose absorption by the liver, skeletal muscles, and adipose regions [126, 127]. Prissana et al. reported the treatment effects of doping silver (Ag) into the ZnO nanorods (ZnO:Ag NR's) in a diabetic murine model. The silver-doping strategy appears to effectively enhance the antioxidant potential of ZnO, as evidenced by their activities in scavenging NO, DPPH, and ·O2· [128].

Nanoparticles have limited use in diabetic treatment. Functionalized gadofullerene was later demonstrated to improve defective glycolipid metabolism in type 2 diabetic mice. However, gadofullerene's effect on clearance of ROS is negligible [129]. To ensure their success, it must be followed by carefully executed parallel bio-distribution and toxicity investigations.

Atherosclerosis

The main pathological manifestation of Atherosclerosis (AS) is lipid deposition in some arterials with smooth muscle cells and fibrous matrix proliferation, which progressively develop into atherosclerotic plaques. There is a correlation between the degree of oxidation and the severity of AS. And it has been shown that oxidative changes of lipids and proteins have been found in vascular lesions [130]. Several processes involved in atherogenesis have been linked to ROS, including adhesion molecule expression, increased proliferation and migration of vascular smooth muscle, endothelial apoptosis, lipid peroxidation, matrix metalloproteinase activation and alterations in vasomotor activity [131]. Vascular endothelial cells experience chronic OS due to a decrease in the production of antioxidant enzymes such as SOD and CAT, leading to an increase in free radicals and ROS [132]. Hence, prevention of vascular OS represents crucial therapeutic strategy of AS.

Research on Nano-modification of traditional Chinese medicine is booming in AS, especially on the intelligent and biomimetic modification of their carriers.

Ginsenoside (Re) is a powerful component with anti-inflammatory and antioxidant characteristics [133, 134], as well as the ability to improve AS [135]. CAT and Re were co-loaded onto the surface of porous poly (lactic-coglycolic acid) (PLGA) NPs to develop a dual targeted model and multi-mechanism therapeutic biomimetic nanosystem (Cat/Re@PLGA@UCM) [108, 136]. The biomimetic nanosystem not only exhibit the ability to scavenge ROS, but also enable escaping macrophage phagocytosis and targeting atherosclerotic plaques, and H2O2-responsive drug release ability. The nanodrugs reduced atherosclerotic area 2.7-fold better than free Re.

Teng Wu et al. established a smart medication delivery device that adapted to the oxidative microenvironment of atherosclerotic plaques [137]. Poly (ethylene glycol) and poly (propylene sulphide) (PEG-PPS) was used to make andrographolide-loaded micelles. Andrographolide-loaded PEG-PPS micelles reduce inflammation and OS simultaneously. After oxidation, PPS becomes hydrophilic, improving medication distribution and effectiveness.

Meili et al. developed a smart system for reacting to the microenvironment of atherosclerotic plaques, which included ROS and shear stress. Red blood cells (RBCs) and simvastatin-loaded micelles (SV MC) comprised the system. RBCs were utilized to extend the circulation and improve the therapeutic effect. SV MC@RBCs micelles were ethylenediamine-functionalized ring-opened poly (glycidyl methacrylate)-poly (propylene sulfide) (PGED-PPS). The micelle ruptured when high ROS made hydrophobic PPS hydrophilic, releasing medication. PPS also reduces ROS, enabling synergistic AS therapy with medicines and materials [138].

Ferulic acid nanoparticles primarily inhibit the production of ROS by suppressing the expression of oxLDL receptors. Rebecca A. Chmielowski et al. developed ferulic acid-based poly (anhydride-ester) nanoparticles to reduce oxLDL absorption and ROS in human monocyte-derived macrophages (HMDMs) [139]. Ferulic acid-based polymer nanoparticles, which were biodegradable, may release ferulic acid sustainably and tunablely to inhibit macrophage foam cell production.

CeO2 nanoparticles could protect endothelial cells (ECs) from oxidative damage by counteracting H2O2-induced ROS [140]. Gao et al. found that the gadolinium doping of CeO2 (Gd/CeO2) nanozymes promoted the surface proportion of Ce3+ and ROS catalytic activity [141]. The optimized Gd/CeO2 nanozyme, which displayed optimal CAT and SOD mimic activities, revealed enhanced efficacy and anti-inflammatory benefits against AS via ROS salvage. Using probucol-loaded mesoporous polydopamine (MPDA) carriers and platelet membranes, Lu Chen et al. created a bionic multifunctional nanoplatform (BM-NP) [142]. BM-NPs selectively aggregated in plaque lesions of the ligated right carotid artery (RCA) animal model due to platelet membrane adherence to damaged blood arteries. BM-NPs' antioxidant properties may synergistically reduce plaque ROS and foamy macrophages, avoiding AS.

Metal NMs like MnO2, Au, and Pt have also been utilized in a wide range of researches. Mesoporous MnO2 nanoparticles with the modification of hyaluronic acid (HA) [143] reached high drug loading capacity of curcumin, which combined the catalytic activity of the nanocarrier and the antioxidant functions of curcumin. MnO2/HA showed intrinsic catalase mimic activity, which catalyzed the endogenous abundant H2O2 into O2 as self-oxygenation agent to relieve hypoxia in AS site. The resulting nanomedicine could also achieve targeting drug delivery by HA modification to bind CD44 receptor overexpressed on diseased macrophages surface.

Wang et al. produced raspberry-like Pt and cerium bimetallic nanostructures with ticagrelor loading and PEGylation (DPTP NRs) for synergistic AS treatment. Pt-cerium bimetallic nano-raspberry prevented foam cell formation by scavenging ROS and lowering plaque oxidized LDLs more effectively. Ticagrelor reduced plaque and platelet aggregation [144].

Another study used a SOD-mimetic agent (Tempol) and a H2O2-eliminating substance of phenylboronic acid pinacol ester covalently conjugated on β-cyclodextrin (β-CD) (TPCD NPs) to treat AS (Fig. 5). TPCD NPs accumulated in atherosclerotic lesions by passive targeting through the dysfunctional endothelium and translocation via inflammatory cells. TPCD NPs reduced systemic and local oxidative stress and inflammation, and eliminated oxidized LDL internalization [145].

Fig. 5
figure 5

Copyright 2018, American Chemical Society

An overview of the design, distribution, and targeting capabilities of a nanoparticle with a wide spectrum ROS scavenging capacity. A The creation of a TPCD NP and its chemical structure as a ROS-scavenging substance. B TPCD is able to remove H2O2, DHHP, ·O2·, and hypochlorite, with the effectiveness depending on the dosage. C Representative transmission electron microscopy (TEM) image, scanning electron microscopy image (SEM), size distribution profile and TEM image after phosphotungstic acid staining of TPCD NPs.

Wu et al. covalently bonded Au nanoparticles (Au NPs) to L-Arginine (LA) and β-cyclodextrin (β-CD) to make a NO-driven nanomotor (CD-LA-Au-aV). Modified anti-VascularCellAdhesionMolecule-1 antibody targets and anchored nanomotors to blood vessel walls. LA reduced ROS, β-CD cleared cholesterol in foam cells, and Au NPs killed inflammatory macrophages. Dual-mode nanomotors improved anti-AS efficiency [146].

In order to treat AS, a unique tetrapod needle-like PdH (TN-PdHs) nanozyme [147] that reacted ROS scavenging, anti-inflammation, and autophagy activation was developed. The oxidative alteration of the confined LDL was prevented by the designed TN-PdHs, which also decreased OS in the vessels. They were quite effective in reducing inflammation, as they reduced levels of pro-inflammatory cytokines such as TNF-α, IL-1, and IL-6. Another study prepared a new type of PdH-Tellurium (PdH-Te) nanozyme. This PdH-Te nanozyme not only exhibited intrinsic CAT and SOD-like activities, but also as worked as an excellent H2 storage material, both of which can reach effective treatment through a combination of scavenging ROS and anti-inflammation [148].

Using porous manganese-substituted prussian blue (PMPB) nanocubes (NC), Zhang et al. [149] developed a theranostic agent loaded with simvastatin (Sim). The two active components PMPB NC and Sim helped reduce atherosclerotic plaques and inflammation by decreasing ROS levels (free radicals and H2O2), pro-inflammatory cytokine secretion, collagen accumulation, fibrous cap thickness, macrophage infiltration, foam cell generation, and LDL internalization. Sim as a model drug, Epigallocatechin gallate (EGCG) as an antioxidant agent, and distearyl phosphatidylcholine (DSPC) as major carriers were used to make liposome nanoparticles (SE-LNPs) in the study of Jun Wan et al. [95]. SE-LNPs had a prolonged release profile, allowing the bulk of medication to accumulate at the targeted atherosclerotic plaque, which might resist oxidation, apoptosis, enhance M2 polarization, and decrease blood lipids and lesions. Yue Dai et al. created GPRD NPs by electrostatically adsorbing Gd-doped Prussian blue (GPB), polymer polyethyleneimine (PEI), fluorescent molecule rhodamine (Rd), and targeted molecule dextran sulfate (DS) [150]. GPRD NPs effectively imaged and inhibited AS susceptible plaque in vivo using GPB’s MR and fluorescence imaging, Rd's nano-enzyme, and DS's targeting abilities. GPB NPs had the action without drug loading, simplifying nanocomplex production. Yan Zhu et al. constructed a Prussian blue-based nanomedical loading system with hyaluronic acid (HA) coating, in which colchicine was encapsulated to create col@PBNP@HA [151]. col@PBNP@HA successfully reduced MDA and MPO levels and increased GSH levels, HA on the drug surface specifically bound to CD44 expressed on inflammatory macrophages, which allowed the drug to target plaques to eliminate inflammation.

Jessica Chavez et al. used carbon nanodots (CNDs) in EA.hy926 Endothelial Cells [152]. CNDs effectively scavenged H2O2 and increased the activity of the antioxidant enzyme NQO1.

Suman Basak et al. drafted novel nitroxide-based nanogels (NGs) crafted through controlled RAFT (Reversible Addition Fragmentation chain Transfer) polymerization to introduce atherosclerosis. Nitroxyl radical-based antioxidants mimic SOD activity, effectively scavenging ROS and reducing LDL oxidation. NGs provided increased surface area, enhanced accessibility of nitroxide groups, higher stability cross-linking, and longer shelf life. NGs effectively reduced foam cell formation and prevents oxidative damage [153].

Many appealing properties of nanoparticles include their tiny size (and consequently huge surface area per volume), relative simplicity of manipulation, and surface components. The survival of nanoparticles in plasma and their permeability in non-targeted organs and tissues must also be explored.

Age-related pulmonary disease

Idiopathic pulmonary fibrosis

Interstitial remodeling is a hallmark of the degenerative lung condition known as idiopathic pulmonary fibrosis (IPF). Telomere shortening, DNA damage response (DDR), and cellular senescence are all linked to pulmonary fibrosis [154, 155]. ROS causes single-stranded DNA damage and breakage, resulting in alveolar epithelial cells (AEC) injury and necrosis via the death receptor route [156], mitochondrial death pathway [157], and endoplasmic reticulum-associated death pathway [158]. Given the compelling evidence connecting OS to the pathophysiology of IPF, targeting ROS may be a successful therapeutic approach.

C60 fullerene has been demonstrated to be capable of scavenging multiple types of free radicals, including ·O2·,1O2, and·OH [159]. At low physiological concentrations, water-soluble C60 is innocuous and possesses significant antioxidant properties. Dong et al. found that water-soluble C60 reduced the severity of bleomycin-induced pulmonary fibrosis in mice [160]. In AEC, water-soluble C60 reduced the concentration of ROS, the expression of TGF-1 and TNF, apoptosis, and/or necrosis. Gadofullerenol (GF-OH m) and fullerenol (C70-OH) NPs were designed as ROS scavengers to inhibit BLM-induced pulmonary fibrosis in a separate study [161]. GF-OH/C70-OH NPs were superior to GF-OH NPs at neutralizing OS and scavenging free radicals.

Yinjuan Lv et al. encapsulated copper-based nanozyme (CuxO NPs) and gold nanoparticles (Au NPs) in oxidation-sensitive dextran (Oxi-Dex) to synthesize ROS-responsive nanocomposites (named as RSNPs) [162]. CuxO NPs showed superior SOD-like and CAT-like activities. RSNPs specifically recognized excess ROS and damaged mesenchymal stem cells (MSCs), released therapeutic nanoenzymes, thereby enhancing the anti-oxidative stress capacity of MSCs and prolonging their survival time in vivo.

Vanadium carbide nanosheets (V4C3 NSs) were reported to serve as a potential antioxidant for treatment of IPF, which triggers multiple antioxidant mechanisms including electron transfer, H atom transfer, and enzyme-like catalysis [163]. V4C3 NSs demonstrated significant therapeutic efficacy by scavenging ROS and RNS (ABTS + •, DPPH•, PTIO•,·OH, ·O2·, H2O2), anti-inflammatory activity, and reestablishment of lung antioxidant defenses.

Chronic obstructive pulmonary disease

External variables, such as cigarette smoking, air pollution exposure, and occupational exposures, are major contributors to the development of chronic obstructive pulmonary disease (COPD). Increases in oxidative load, ROS and reactive nitrogen intermediates (RNI) [164], which are linked to COPD. COPD patients' neutrophils and airway smooth muscle cells have higher amounts of ROS than those of healthy people [165]. Similarly, neutrophils isolated from COPD patients' peripheral blood have been found to produce higher levels of ROS compared to healthy controls [166]. The degradation of elastin in the lung parenchyma might be hastened by OS, which can disrupt the activity of antiproteases such alpha-1 antitrypsin and secretory leukoprotease inhibitor. OS reduces histone deacetylase activity [167, 168] and boosts histone acetyltransferase activity [169], resulting in increased expression of proinflammatory marks. Both chronic bronchitis and small-airway fibrosis have been linked to OS [170, 171].

Multiple materials have been shown to be effective in treating COPD, with NMs as vectors for enhancing functions. Chitosan (CS) and SLNs were used to encapsulate berberine (Ber) [99]. The effects of Ber pretreatment on MPO and SOD activity in cigarette smoke-induced COPD mice were amplified by Ber encapsulated in SLN-chitosan nanoparticles. The aqueous solubility and oral bioavailability of SLN nanoparticles coated with CS improve the pharmacological effects of Ber. Paudel et al. found that treating human broncho-epithelial cells and macrophages with Ber-loaded liquid crystalline nanoparticles (LCNs) improved its physiochemical properties such as high entrapment efficiency and sustained in vitro release. Ber-LCNs inhibited total cellular ROS, modulated genes associated in inflammation and OS [172].

Likewise, lipopolysaccharide (LPS)-induced oxidative damage in human bronchial epithelial cell line (BEAS-2-B) cells was researched using rutin-loaded liquid crystalline nanoparticles (LCNs). LCNs increased transport, biological activity, treatment regime, and patient compliance. Rutin-loaded LCNs dramatically lowered NO and ROS levels in BEAS-2B cells while also preventing apoptosis [173]. Keshav Raj Paudel et al. evaluated the effect of zerumbone-loaded LCNs (ZER-LCNs) in cigarette smoke extract (CSE)-induced models [174]. The antioxidant activity of ZER is exerted by increasing GSH levels to reduce ROS. ZER-LCN showed greater pharmacological and biological benefits in reducing smoking-induced inflammation, oxidative stress, and aging than free ZER alone.

Dimethyl fumarate (DMF) has antioxidant and anti-inflammatory properties in COPD patients [175]. It reduces OS by activating the nuclear factor (erythroid-derived 2) -like 2 (Nrf2) genetic pathway [176]. Priya Muralidharan et al. [177] created respiratory tract-targeted inhalable DMF dry powders. Solid-state respirable microparticles/nanoparticles dispersed aerosols well, which show the potential to reach lower airways.

Kosuke Chikuma et al. developed a co-delivery approach using core–shell type lipid-polymer nanoparticles (LPNs) with a poly lactic acid (PLA) core carrying a potent antioxidant Mn-porphyrin dimer (MnPD) and a cationic lipid (DOTAP) shell that binds HDAC2-encoding plasmid DNA (pHDAC2). The co-delivery system had low toxicity, high serum stabilities, delayed and tuneable drug release, and excellent drug encapsulation efficiency. PLA-MnPD/DOTAP/pHDAC2 decreased ROS and glucocorticoid resistance in COPD patients [178]. S Castellani et al. used SLNs to encapsulate grape seed extract (GSE) with proanthocyanidins. GSE-loaded SLNs had a longer anti-oxidant impact than free GSE in H441 airway epithelial cells. This formulation may reduce ROS-induced inflammation during chronic lung illnesses [179].

Incorporated polyoxalate (HPOX) may reduce respiratory tract inflammation [180]. HPOX NMs scavenged H2O2, reduced intracellular OS, and inhibited the expression of pro-inflammatory mediators like iNOS, cyclooxygenase-2 (COX-2), and IL-1β in stimulated macrophages. HPOX NMs were biocompatible and strong antioxidants and anti-inflammatories for airway inflammatory diseases (Fig. 6).

Fig. 6
figure 6

Copyright 2013, Elsevier

Incorporated polyoxalate (HPOX) nanoparticle structure and antioxidant capacity. A HPOX is an innovative prodrug polymer that uses HBA as its backbone antioxidant and anti-inflammatory properties. HPOX’s medicinal actions come from the release of HBA during the breakdown process. B The ability of HPOX nanoparticles to scavenge H2O2 and to suppress the production of ROS in PMA-stimulated macrophages.

The lungs are unique compared to other systems in that NMs can be administered directly in the lungs to avoid first-pass metabolism, thereby increasing local concentrations in lungs. However, there are still problems such as airway mucus layer barriers, clearance by mucosal ciliary clearance systems, and the need to cross the epithelial barrier for the drug to reach the endothelial cell layer. All these issues need to be considered together in drug design with respect to the chemical-physical properties of the NMs [181]. Currently, research is focused on maximizing delivery efficiency and minimizing toxicity. This includes the PEG-modification on surface and optimization of osmotic pressure gradient for mucus penetration, as well as the optimization of formulation to improve stability, deep lung deposition, and distribution. To successfully transport antioxidants to the lungs, further study is required. The potential for immunogenicity and toxicity to the lungs is an important factor to consider.

Skeletal and muscle degenerative diseases

A crucial regulator of osteoclast development, both bone production and bone resorption is receptor activator of nuclear factor Kappa-B ligand (RANKL). Studies have indicated that the osteoprotegerin (OPG), receptor activator of nuclear factor Kappa-B (RANK), and RANKL system may play a crucial role in the process tying osteoporosis and osteoarthritis together (Fig. 7). Interleukin (IL-6, IL-13), TNF, and other inflammatory substances that are released have high osteoclastogenic activity and can either directly activate osteoclast precursors or stimulate RANKL to promote osteoclast formation. Along with the rise in RANKL, a significant amount of RANKL binds to the usual level of OPG, causing a compensatory drop in OPG and an increase in bone resorption [182].

Fig. 7
figure 7

Pathology of skeletal degenerative diseases induced by ROS. Created with BioRender.com

Due to the denser nature of pathological skeletal tissues, high concentrations of drugs are required to achieve efficacy, which can also have toxic effects on other organs. Therefore, the development of well-targeted, highly permeable, slow-release, low-toxicity, and bone-targeted NMs is in the spotlight [183].

Osteoporosis

Decreased bone density and degradation of bone tissue microstructure characterize osteoporosis (OP), a systemic and metabolic disease of aging. OP is characterized by increased OC activity relative to OB activity [184]. Patients with OP have a bone microenvironment characterized as immune imbalance and increased OS. Excessive ROS under stressful conditions trigger apoptosis in OBs and osteocytes while encouraging the development and function of OCs [185]. Therefore, enhancing the oxidative state is crucial for osteoporosis therapy and prevention.

Yanhai Xi et al. designed PEGylated hollow gold nanoparticles (HGNPs) loaded with α-Lipoic acid (ALA) (mPEG@HGNPs-ALA) [186]. ALA can suppress intracellular oxidative stress levels and promote the proliferation and differentiation of osteoblasts. In addition to a larger drug loading capacity and enhanced photothermal conversion ability, HGNPs are also tiny (only 30–60 nm in diameter), non-toxic, and spherical in form. The antioxidant capacity and biocompatibility of mPEG@HGNPs-ALA indicated its potential for use in osteoporosis therapy.

The nitrogen-doped carbon dots (N-CDs) have therapeutic promise for the treatment of osteoclast-related osteolytic disorders [187]. The N-CDs decreased Nox1 and upregulated Nrf2 to inhibit RANKL-induced ROS production. Inhibiting osteoclastogenesis and bone resorption with N-CDs in vivo partially protected mice against lipopolysaccharide (LPS)-induced calvarial bone degradation and breast cancer-induced tibial bone destruction. Photoluminescent carbon dots (PCDs) from sour apples cured a mouse calvarial osteolysis model induced by ultra-high molecular weight polyethylene (UHMWPE) wear particles. PCDs reduced UHMWPE-induced ROS stress and pro-inflammatory cytokine production to inhibit osteoclastogenesis and bone resorption in vitro [188].

The osteoporosis cell model examined the ROS-scavenger nanoceria encapsulated in mesoporous silica nanoparticles (Ce@MSNs). Self-regenerating nanoceria mimics SOD and CAT activities. The bioactive MSNs and nanoceria in Ce@MSNs NPs stimulate bone repair and reduce osteoclast activity by releasing osteogenic silica and scavenging ROS. The Ce@MSNs showed promise as a therapy for osteoporosis, based on their potential therapeutic efficacy [189].

The polyglucose-sorbitol-carboxymethyl ether (PSC) was employed as the precursor to synthesize Fe2O3@PSC NPs in a mouse model of iron accumulation (IA)-related osteoporosis [190]. Nanoscaled Fe2O3 minimized the generation of free iron ions. PSC protected bone tissues from the damaging effects caused by ROS generation induced by free iron ions. Fe2O3@PSC sustainably released iron ions instead of releasing a great quantity in a short time, which showed promise as a new IA-related osteoporosis treatment. Iron oxide nanoparticles (IONPs) scavenge ROS through the Nrf2-keap1 pathway to ameliorate postmenopausal bone loss. Zheng et al. created bone targeting IONPs (BTNPs) using alendronate. BTNPs targeted bone surfaces and scavenged ROS to treat mice with ovariectomy-induced osteoporosis. BTNPs outperformed IONPs and bisphosphonates, suggesting a viable clinical use [191].

Polyhydroxyalkanoate-encapsulated CaSi2 nanoparticles (CSN)-loading mesoporous bioactive glass (MBG) scaffolds (CSN@PHA-MBG) were designed for releasing H2 in the repair of bone defect of elders [192]. CSN greatly improved H2 release capacity for approximately one week. Sustained treatment of H2 generally attenuated oxidative stress and effectively remodelled the senescence-associated secretory phenotype via anti-inflammatory pathways, supporting damaged aged bone repair.

Nahida Rasool et al. [193] developed thiolated, bioactive mesoporous silica nanoparticles (MSN-SH) for bone tissue engineering/osteoporosis. Functional modification of the surface thiol groups enhanced the osteogenic properties of MSN and confers antioxidant and cell adhesion properties. MSN-SH neutralized ROS and provide protection against ROS-induced cellular damage.

Conventional therapies have the limitation of side effects and poor penetration into skeletal lesions, while NMs could improve drug solubility and stability [194]. NMs in circulation may still be non-specifically phagocytosed by the liver and spleen, limiting the targeted impact, and this is one of the main reasons why biological NMs are not widely used in the treatment of OP.

Osteoarthritis

Osteoarthritis (OA) is a progressive joint disease that is characterized by the deterioration of articular cartilage and oseophyte. OA can affect any joints in the body. Numerous studies point to the role of ROS as primary contributors to the development of OA. The OS caused by ROS is capable of oxidizing cartilage, which will then disturb its homeostasis, encouraging catabolism through the induction of cell death, and harming a variety of components of the joint [195]. ROS operate as inflammatory mediators by activating proteoglycans, collagen molecules, matrix proteins, and membrane proteins directly [196, 197]. These proteins, including IL-1β and TNF-α, are directly responsible for the significant damage that is caused to the joint tissues of OA sufferers. As a result, ROS scavengers have a significant amount of untapped potential for the treatment and remission of OA.

Surface quinone residues in natural melanin efficiently scavenge radicals. Zhong et al. found that dopamine melanin (DM) NPs may scavenge ROS (including ·O2·, ·OH) and reactive nitrogen species (RNS), protecting chondrocytes from OS, inflammation, and cartilage degeneration. DM NPs, which were almost 110 nm, may stay in the joint longer than small molecule scavengers, suppressing ROS/RNS and managing OA [198]. MOF-decorated mesoporous polydopamine was utilized by Song et al. to develop a dual-drug delivery system, with rapamycin (Rap) injected into the mesopores and Bi deposited onto the MOF shell. By coupling the collagen II-targeting peptide (WYRGRL) to the nanocarrier, a cartilage-targeting dual-drug delivery nanoplatform (RB@MPMW) was developed. RB@MPMW effectively eliminated cellular ROS through Br and enhances autophagic activity via Rap [199].

The capacity of chitosan nanoparticles with glutathione (Np-GSH) were evaluated in Rats with OA [200]. GSH can directly interact with ROS or act as a cofactor in enzymatic processes. Chitosan-based grafts were ideal substrates for the proliferation of chondrocytes. The GSH contained within nanoparticles (NPs) can be delivered to chondrocytes, reducing ROS, increasing GSH levels and the activity of GPx, and reducing lipid peroxidation.

Haifeng Liang et al. encapsulated melatonin in poly(lactic-co-glycolic acid) (PLGA), with the type II collagen targeting peptide attached to the surface to prepare a nano-delivery system loaded with melatonin(MT@PLGA-COLBP) [201]. Melatonin enhanced the activity of antioxidant enzymes such as GPx and SOD. It repaired the damaged mitochondrial function in chondrocytes and reduces hydroxyl radicals through its metal chelating activity. The MT@PLGA-COLBP formulation achieved targeted functional release and sustained release of melatonin within the joint space, improving cartilage matrix metabolism and delaying the progression of OA in the body.

Exogenous SOD's poor pharmacokinetics and poor cell permeability may explain why native SOD showed no therapeutic benefits. O-HTCC-SOD is a nanoparticle-like compound of cationic functionalized CS and SOD [202]. Due to its highly cationic nanoparticle-like feature, O-HTCC-SOD may penetrate cells and effectively scavenge intracellular ROS. O-HTCC-SOD protected chondrocytes longer than native SOD from monoiodoacetate (MIA)-induced oxidative damage, which included reducing mechanical allodynia, inhibiting gross morphological and histological cartilage lesions, and increasing antioxidant capacity and anti-inflammatory action. Tao et al. used SOD-loaded porous polymersome nanoparticles (SOD-NPs) to target mouse synovium [102]. SOD-NPs had prolonged mouse joint retention time and minimized oxidative damages.

Zhang et al. loaded calcium boride nanosheets (CBN) as H2 precursors onto dopamine-modified hydrogel platform (CBN@GelDA hydrogel) for OA treatment. CBN@GelDA hydrogel released H2 stably and continuously under physiological conditions, the release process does not affect pH of the microenvironment. CBN@GelDA hydrogel scavenged excessive ROS, alleviated oxidative stress, reduced inflammation and joint destruction, and provided long-lasting relief of OA [203].

Zhao et al. created novel drug-free nanospheres which were self-assembled into spherical aggregates from the block copolymer of P(DMA-b-SBMA) in aqueous solution. The nanospheres' clever construction gave them the capacity to withstand physiological stresses, improve lubrication, and neutralize harmful ROS. In a rat model of temporomandibular joint (TMJ) osteoarthritis, the nanospheres prevented structural damage to the condylar cartilage and subchondral bone, slowed the deterioration and ageing of the cartilage matrix [204].

MnO2 NPs can function as an artificial nanoenzyme to scavenge ROS. The PEG-MnO2 NPs improved chondrocyte viability and extracellular matrix preservation by reducing inflammation-induced OS in cartilage [63]. Chen et al. synthesized an intelligent hollow MnO2 (H-MnO2) modified with NH2-PEG-NH2 to target OA treatment [205]. H-MnO2 NPs had the ability to efficiently eliminate ROS and greatly alleviate the inflammatory response of OA without evident side effects, opening up new treatment avenues for those living with the condition.

Pei et al. treated OA in rats using water-soluble polyhydroxylated C60 (fullerol) NPs [206]. Fullerol reduced OA by preventing synovial membrane inflammation and chondrocyte destruction in OA joints.

Ruiming Liang et al. suggested using nanofibers constructed of poly (ε-caprolactone) (PCL) and PCL-grafted lignin (PCL-g-lignin) copolymer [207]. PCL tailored mechanical properties whereas ligin had inherent and persistent antioxidant action. Biocompatible, biodegradable, and antioxidant-rich PCL-lignin nanofibrous membranes treated OA.

Compared to traditional spherical cerium dioxide nanoparticles, Urchin-like ceria nanoparticles loaded miR-224-5p more effectively delivered miRNA into cells and exhibit superior ROS scavenging capabilities. This enhanced their ability to suppress inflammatory responses and modulate the microenvironment of OA, thereby improving the gene therapy approaches for OA [208].

Degeneration of the whole joint characterizes OA, making intra-articular injection of ROS-responsive nanomedicine an ideal treatment option, since it allows for regulated release and focused therapy without systemic side effects. Furthermore, NMs should be developed to maximize the retention period in the joint cavity because of the quick clearance of the joint cavity.

Sarcopenia

Consistent muscular weakening and atrophy with advancing age was termed sarcopenia [209]. An imbalance between ROS/RNS and the enzymatic antioxidant defence system is a crucial player in the pathophysiological pathways that lead to sarcopenia. Recent studies have shown that compared to young/adult rats, myofibers from elderly rodents contain higher amounts of RONS intracellularly [210]. Muscle mass was negatively impacted by elevated ROS because it facilitated ER stress, which caused cell death in muscle cells. Increased oxidative damage and mitochondrial malfunction, decreased ATP generation, increased protein breakdown, and decreased protein synthesis are all potential outcomes of an overactive redox signaling system in muscle fibers [211, 212].

As a nanocarrier for antioxidants, hydroxyapatite is a material that is often used in sarcopenia. Biocompatibility and biodegradability make hydroxyapatite (HAP) a popular drug delivery system material. The following materials increased curcumin loading surface area. Curcumin-loaded HAP modified with stearic acid (Cur-SHAP) released continuously for over 2 weeks, reducing sarcopenia development or even reversing it [213]. Bletilla striata polysaccharide (BSP) coupled with HAP was employed by Ya-Jyun Liang et al. [213]. BSP is an efficient ROS scavenger. In the current investigation, BSP-HAP administered by intramuscular injection would remain in the muscle tissue, followed by a slow absorption via endocytosis. In the recovery of LPS-induced muscle damage, the created BSP-HAP could decrease LPS-induced ROS formation and improve tissue healing.

Natural antioxidants such as curcumin rather than nanomaterials with ROS scavenging activity are mostly used in the antioxidant treatment of sarcopenia. Nevertheless, the effect of antioxidant supplementation on muscle performance is still highly debatable.

Skin aging

Skin aging is characterized by fine lines and wrinkles, loss of elasticity and volume, sagging, roughness and pallor in appearance. The generation of ROS, which causes DNA, protein, and lipid damage as well as extracellular matrix dis-organization, is a typical hallmark of both intrinsic and extrinsic skin aging [214]. Skin has a greater ROS burden when compared to other organs, which impacts both intrinsic and extrinsic aging [215]. Excessive ROS can boost the expression of pro-inflammatory cytokines including IL-1, TNF-α, IL-6, and COX-2 to regulate the inflammatory response [216, 217], as well as make the MMPs/TIMPs ratio imbalanced by activating MMPs and decreasing TIMP production, decompressing ECM [218]. Meanwhile, ROS can suppress collagen formation and accelerate skin aging via regulating the TGF-β/Smad signaling pathway [219] (Fig. 8). Antioxidants have been demonstrated to dramatically reduce or prevent free radical damage to the skin.

Fig. 8
figure 8

Pathology of skin aging induced by ROS. Created with BioRender.com

Antioxidant nanoparticles have attracted a lot of interest in the cosmetics industry as a possible solution to the effects of skin aging. Investigations on nanoization of conventional medications begun at an early stage, such as EGCG, RSV, CoQ10, quercetin.

Nano-transfersomes loaded with EGCG and hyaluronic acid (HA) were employed by Avadhani et al. [220]. HA's anti-aging qualities, which include biocompatibility, particular viscoelasticity, hydration, and lubrication, make it a promising anti-aging agent [221]. Optimized transfersomes had far greater skin penetration and EGCG deposition than pure EGCG, which improved cell survival, lipid peroxidation, intracellular ROS, and MMP expression in human keratinocyte cell lines (HaCaT).

SLNs and NLCs can provide intimate contact and promote medication absorption via the skin. Incorporating RSV into SLNs and NLCs [96], encapsulating CoQ10 into NLC [100], liposomes (LIPO-Q10) and SLNs (SLN-Q10) [94], ultra-small lipid nanoparticles (usNLC-CoQ10) [222] has been investigated for topical use. All of the above exhibited excellent antioxidant capacity in cells following UVA and UVB irradiation.

A typical dietary flavonoid, quercetin has several physiological benefits including being a powerful antioxidant, scavenger of free radicals, and anti-inflammatory [223]. Tyrosol-incorporated copolyoxalate (TPOX) NPs were synthesized by Kim et al., and they were made up of an H2O2-sensitive peroxalate ester incorporating tyrosol. Then, quercetin (QTPOX) was included into the TPOX NPs. H2O2 may delicately break down TPOX into CO2 and H2O. This sensitive quality helps to target and release in oxidatively damaged cells. In HaCaT cells, the QTPOX NPs demonstrate cytoprotective properties via antioxidative and anti-inflammatory actions [224]. Nisar et al. created Quercetin-loaded zinc oxide nanoparticles (Quercetin@ZnO NPs) in vitro. ZnO NPs release the drug for sunblocking and protecting such as antioxidant, anti-inflammatory, and iron-sequestering properties by delivering maximum quercetin molecules to the targeted site after UVA exposure [225].

The antioxidant capabilities of stable SeNPs stabilized with chitosan of varying molecular weights (Mws) (CS-SeNPs) were investigated. Because of its low toxicity and bioavailability, CS could survive pepsin and pancreatin, and stabilize the Se system in the digestive enzyme environment. In skin-aging mice, all CS-SeNPs penetrated tissues and had antioxidant effects [77].

A fullerene-loaded nanoemulsion was employed to preserve collagen and prevent skin aging [226]. In the HaCaT cell line, Xiao et al. investigated the antioxidant properties of several water-soluble fullerene derivatives. A ROS-scavenging effect against UVB-injuries was demonstrated for PVP/fullerene, CD/fullerene, and hydroxyl group-containing fullerene, indicating the likelihood of skin aging [227].

After UVA radiation, CeO2 NPs decreased pro-inflammatory cytokines, intracellular ROS, senescence-associated β-galactosidase activity, and JNK activation [228]. CeO2 NPs were used to scavenge ROS, protected skin against radiation and inflammation, and helped wounded healing [229, 230].

Enzyme-mimicking Au-Pt nanocomposites (NCs) were produced by Xiong et al. in HaCaT cell lines to scavenge cellular ROS caused by UV irradiation [231].

Chiral manganese dioxide nanoparticles with high sensitivity and selectivity for ROS were engineered. MnO2 NPs eliminated ROS in skin tissues, increased collagen, and showed exciting roles in inhibiting oxidative damage in skin and preventing skin aging [232].

Redox nanoparticles (RNPN) are nitroxide radical-containing polymers that may efficiently remove ROS. Oral RNPN supplementation increased the therapeutic benefits of the core nitroxide radical and decreased UVB-induced skin aging in an inflammatory skin model. The RNPN may protect skin against ROS damage and slow aging [233].

Nanotechnology can improve the performance of cosmetics in a variety of ways, such as by enhancing entrapment efficacy, physical stability and dermal penetration of the active ingredient, regulating the release of the active ingredient. The majority of these bioactive compounds, however, are poorly absorbed by the skin. On the one hand, the skin permeability of nanomedicines needs to be enhanced, and on the other hand, nanoparticles may cross the skin and enter the body circulation, causing unintended toxicity and side effects. Nanoparticles may cause skin irritation or allergic reactions. It is necessary to adjust the size, shape, charge, degradability, and dose of the drugs to make them more absorbable, less toxic, and less allergenic [234].

Neurodegenerative diseases

Alzheimer's disease and Parkinson’s disease are the two most prevalent neurodegenerative diseases, respectively. In terms of mechanisms of OS, there are many commonalities between AD and PD. ROS production played an important role in Amyloid-beta (Aβ) oxidation, mitochondrial dysfunction, upregulation of inflammatory factor expression and selective neuronal degeneration (Fig. 9).

Fig. 9
figure 9

Pathology of neurodegenerative diseases induced by ROS. Created with BioRender.com

The existence of the blood–brain barrier (BBB) hinders the identification and treatment of brain illnesses by limiting the transit of biologically active chemicals and drugs [235]. Drugs were unable to sustain a high enough bioavailability to have an impact on the brain parenchyma pharmacologically. The unique qualities of NMs, including as their tiny size, drug-loading capacity, high blood stability, low immunogenicity, high biodegradability, and the ability to change surface properties, have been employed to treat neurodegenerative illnesses [236].

Although NMs can effectively penetrate the BBB to reach the brain, they may accumulate in the brain and thus cause adverse reactions or toxicity. In addition, the long-term stability and metabolic pathways of nanoparticles in the body are not yet fully understood, and there may be a risk of long-term accumulation and chronic toxicity. NMs may activate the immune system of the human body, triggering an inflammatory response and affecting the health of the nervous system [237, 238].

Translated with DeepL.com (free version).

Alzheimer's disease

AD is a neurodegenerative disease characterized by gradual cognitive decline and behavioral abnormalities, with common clinical symptoms including progressive forgetfulness, loss of recognition, loss of reading, and loss of speaking. Neurodegenerative illnesses like AD are characterized by OS and neuronal death [239]. Accumulation of free radical caused damage and altered expression of antioxidant enzymes are hallmarks of AD [240]. Due to reduced production of major antioxidant enzymes as CAT, SOD, GPx, and GSH reductase, the body is unable to utilize its detoxification mechanisms effectively [241]. Together, the oxidative imbalance, the overexpression of nuclear factor kappa-light-chain-enhancer (NF-kB), and the release of inflammatory mediators (such as IL-1β, IL-6, TNF-α, and TGF-β) create an environment conducive to the development of AD [242, 243]. The activation of N-methyl-D-aspartate receptors (NMDARs) results in ROS production when Aβ accumulates abnormally. This in turn led to OS. ROS triggers a vicious loop that causes the onset and development of AD by increasing the synthesis and aggregation of Aβ and hyperphosphorylated Tau. Antioxidant treatments have emerged as promising possibilities for treating AD, according to preclinical research [244]. Additionally, several types of nanoparticles have been utilized in AD investigations (Table 3).

Table 3 Application of ROS-scavenging nanomaterials in treatment of Alzheimer's disease

Due to site-specific delivery, the ability to cross the BBB, increased drug solubility, and greater therapeutic efficacy, nano-delivery is a preferable alternative. BBB penetration favors particles with a lower dimension. It is essential to use nanoparticles of a reduced size to improve BBB penetration, reduce acute toxicity and adverse effects, and increase drug loading capacity.

Parkinson’s disease

Parkinson's disease (PD) is a degenerative disorder of the central nervous system that slows the mobility of the patient. The early manifestations of the disease include resting tremor, myotonia, slow movement, difficulty in starting movements and abnormal posture. In PD patients, Farias et al. [273]discovered elevated lipid hydroperoxides (LOOH), MDA levels, and SOD activity, as well as reduced CAT activity. ROS-mediated OS is closely related to PD, mainly because the production of large amounts of ROS by activated microglia is accompanied by increased sensitivity to ROS and reduced scavenging capacity of brain tissue in PD patients [274]. Additional pathways involved in PD are neurodegeneration caused by the action of androgen receptors [275], enhanced α-synuclein aggregation and formation of oxidatively modified forms of α-synuclein [276], degradation of quinone oxidoreductase 1 [277], attenuation of protein DJ-1’s deglycase activity [278], activation of gene LRRK2 [279], decreased tetrahydrobiopterin and tyrosine hydroxylase (TH) metabolism [280]. Numerous new pharmaceutical therapeutics targeting the OS pathway have been developed, and they are proved useful in the treatment of PD. Here, we discuss the use of ROS-scavenging nanotechnology for PD therapy (Table 4).

Table 4 Application of nanomaterials ROS-scavenging nanomaterials in treatment of Parkinson’s disease

Nanomaterials have great antioxidant qualities, however their applications raise certain safety concerns. In the neurological system, NMs may cause apoptosis, release ROS, modify the production of pro-inflammatory cytokines, and affect neurotransmitter expression [281]. Protecting the brain's homeostasis against the effects of nanoparticles and their breakdown products is an urgent need.

Reproductive aging

Female reproductive aging

The adult hypothalamic-pituitary system, also known as the hypothalamic-pituitary-ovarian (HPO) axis, coordinates with the follicles in the ovaries to control menstrual cycles and the reproductive lifetime and healthspan. With increasing age, follicles are gradually depleted and their quality declines, leading to reproductive aging and menopause. This process is reflected in a significant age-related increase in the probability of infertility, miscarriage and birth defects in the offspring [304]. ROS is considered to be responsible for the initiation or development of pathological processes affecting ovarian function [305]. Follicle atresia and decreased oocyte quality and quantity may result from excessive ROS, which damage DNA, disturb protein function and homeostasis, promote ER stress, autophagy, and proteasome dysfunction among other detrimental effects [306]. Pathological ROS drive ovarian aging by apoptosis, mitochondrial dysfunction, inflammation, telomere shortening and other aspects [307,308,309]. Related antioxidants, such as MLT, vitamin E, and resveratrol, could improve ovarian function and therefore have potential clinical applications [310, 311]. Unfortunately, there is a dearth of studies on the impact of nano-antioxidants on ovarian aging.

Male reproductive aging

Decrease in sperm quality and a higher chance of birth abnormalities and disorders in progeny are signs of reproductive aging in males [304]. 15% of couples worldwide struggle with male infertility, making it an important health issue that has to be addressed head-on [312]. According to recent research, 25–40% of infertile men have high ROS levels in their semen [313, 314]. The integrity of sperm DNA is similarly compromised by OS, which may have an impact on the ability of embryos to grow and the health of their progeny. Male reproductive potential may be decreased by age-related OS because of deteriorating semen quality, altered endocrine, and sexual dysfunction [315]. Patients with elevated levels of ROS may benefit from antioxidant treatment [316], and it is important to design the reasonable antioxidants for male reproductive aging.

The great majority of antioxidants in male reproductive aging are nanoparticles with their own ROS scavenging activity, and some of these are also being utilized in conjunction with traditional medicines like MLT to maximize their capabilities.

MLT is a powerful antioxidant that is capable of capturing free radicals. Synthesized gold (III) MLT (Au3+/MLT) complexes showed anti-inflammatory and antioxidant properties to protect against testicular injury [317]. MLT is an effective formulation for scavenging ROS, triggering the production of molecules that protect sperm from oxidative stress. The combination of Au + 3/MLT significantly enhances total antioxidant capacity compared to using MLT alone.

By reducing OS, Nanoform Se (NSe) reduced testicular toxicity and apoptosis cause by BPA or NiSO4 [318, 319]. NSe was more protective than Se [319].

(FSH)-conjugated SOD-loaded PLGA NPs designed by Snow-Lisy et al. targeted testis Sertoli cells to combat male infertility caused by high levels of ROS [320].

Ionizing radiation produces ROS through the radiolysis of water in irradiated testicular tissue, which causes spermatogenic cell mutation or death, reduced sperm quantity and motility, and increased sperm deformity rate. Molecular hydrogen (H2) has the potential to be a radioprotective agent due to its ability to scavenge ·OH selectively. The use of MgH2 nanoparticles for hydrogen storage and release have several benefits, including high storage capacity, a smooth release rate, and great stability. Ma et al. [321] observed that MgH2 reduced MDA levels in testis, inhibited ROS formation after irradiation, and removed ·OH. Furthermore, by neutralizing hydroxyl free radicals, MgH2 therapy enhanced male fertility impairment due to irradiation.

Ce NPs' potential protective impact against fipronil-induced testis damage was investigated in a rat model [322]. The Ce NPs mitigated the deleterious effects of fipronil on testicular tissue by reducing lipid peroxidation, apoptosis, inflammation, and boosting antioxidant activity.

Fullerenol C60(OH)24, a hydroxylated derivative of fullerene, is investigated for its NO-scavenging action in mesenchymal cells from rat testicles in a separate research by Mirkov et al. [323]. C60(OH)24 could scavenge ·O2· in xanthine/xanthine oxidase system.

There was promising evidence that antioxidants might slow the aging of the male reproductive system. However, an imbalance between oxidants and antioxidants, known as reductive stress (RS), can be caused by an overabundance of antioxidants. The fertility rate and three fundamental seminal indicators (motility, concentration, and morphology) have all been linked to RS's negative consequences [324]. The fertilization process was decreased owing to the inhibition of significant functional activities of the spermatozoa [325]. Therefore, precision antioxidant may be the way forward for study into the effects of aging on reproduction.

Ocular neurodegenerative disease

Age-related macular degeneration

Age-related macular degeneration (AMD) is a chronic neurodegenerative and progressive disease with a multifactorial aetiology that leads to alterations in the macula region's structure [326]. Non-neovascular (‘‘dry’’) AMD effects approximately 85–90% of patients, whereas neovascular (‘‘wet’’) AMD affects the residual 10–15% of patients. Due to its high oxygen metabolism requirements, high unsaturated fatty acid content, the presence of photosensitive molecules (retinoids and lipofuscin), and protracted exposure to light, the retina is more susceptible to injury induced by ROS and OS [327]. Oxidative damage is a precursor to the development of AMD and is implicated in AMD-related inflammation and neovascularization. Key to secondary oxidative injury in the retina [328] are disturbances in the regulation of OS-related molecular pathways such as autophagy and Nrf2 signaling pathways. Given the importance of OS in the pathogenesis of AMD, excessive ROS-targeting antioxidant therapies have been proposed as the first-line treatment.

In order to better administer medications like polydopamine and lutein, nanomaterials are modified to have an enhanced dosage form and permeability. Jiang et al. produced anti-angiogenetic protein-loaded polydopamine NPs for wet AMD [329]. Polydopamine NPs reduced angiogenic agent expression by scavenging ROS stimulated by external OS. In reaction to OS, the particles controllably released loaded anti-angiogenic medicines to cure wet AMD. Lutein is commonly used as an antioxidant due to its ability to quench singlet oxygen and eliminate ROS [330]. However, lutein's inadequate water solubility limits its absorption and effectiveness. Ying Ge et al. created a penetratin-modified lutein nanoemulsion in-situ gel (P-NE GEL) to cure rat dry AMD produced by NaIO3. GEL solution significantly extends the corneal retention time of drugs. With the aid of penetratin, P-NE is rapidly transported to the posterior segment of the eye and distributed in the retinal area. P-NE GEL strongly inhibited cell apoptosis and ROS in human retinal epithelial cells (ARPE-19), indicating its potential use in AMD therapy [104]. By modulating Nrf2 via the PI3K/AKT/mTOR signaling pathway, astragaloside-IV (ASIV) may reduce OS. Three different sized ASIV lipid nanocapsules (ASIV-LNCs), sized at 20, 50, and 90 nm, were loaded with a phospholipid complex produced from ASIV [331]. LNCs offer reduced toxicity, increased drug loading capacity, and enhanced permeability. In a mouse model of dry AMD caused by NaIO3, the ultra-small-size LNCs (ASIV-LNCs-20) exhibited superior penetration effects, which were able to lower ROS generation and the rate of cell death.

Due to its tiny particle size, NMs with free radical scavenging action offers a distinct advantage in ocular illnesses. Mitra et al. developed water-soluble, biocompatible, trackable nanoceria formulation glycolchitosan-coated ceria nanoparticles (GCCNPs) with enhanced antioxidant ability to scavenge intracellular ROS. In laser-induced AMD, GCCNPs decreased ROS-induced pro-angiogenic vascular endothelial growth factor (VEGF) expression, cumulative oxidative damage, and endothelial precursor cell recruitment without toxicity [332]. Fullerenol (Fol) decreased ROS, normalized GPx activity, and promoted CAT in H2O2-induced RPE senescence [333]. Its nanosize permitted intravitreal injection into the retina and RPE cells. Yong-Su Kwon et al. [334] used PEGylated synthetic melanin-like nanoparticles (MNPs) in the RPE to restore melanin for AMD therapy. Biocompatible MNPs preferentially targeted ROS with significant antioxidant effects. MNPs could also treat AMD pathology with a single treatment (Fig. 10).

Fig. 10
figure 10

Copyright 2022, American Chemical Society

Illustrations and characterizations of MNP schematics. A Schematic of the MNP synthesis and characterization. B TEM, hydrodynamic dimensions, and Zeta-potential of Bare-MNPs and MNPs. C MNPs’ ROS-scavenging activity in ARPE19 cells.

Absorption rates, medication penetration, active solubility, and bioavailability have all been proven to enhance with the use of nanomedicine. The absence of blood flow at the location of sickness is common in the eye since it is a relatively closed organ. The creation of a nanocarrier for topical use in the eye is urgently required. When designing a dosage, it's important to keep nanomaterial complexity and dose to a minimal. The clinical translation of methods for sustained and targeted administration of nanoscale medicines to the posterior portion of the eye to treat AMD is still an area of active research.

Cataract

Cataract is primarily an ARD, with a loss of transparency in the lens of the eye, manifesting as blurred vision or glare. The buildup of primary lipid peroxidation (LPO) products (diene conjugates, cetodienes) was characteristic of the early stages of cataract. The preponderance of end LPO luminous products was characteristic of the later stages [335]. Cataractogenesis has been linked to ROS that cause damage in the lens cell, which may take the form of protein oxidation, DNA damage, and/or lipid peroxidation [336]. Antioxidants are a potentially effective strategy for managing cataracts as well as a variety of ocular disorders of the aging eye caused by ROS.

Ethylene glycol, ethylene glycol monoacetate, and ethylene glycol diacetate (EGCNPs) coated cerium oxide nanoparticles were produced in a work by Hanafy et al. Elevated level of reduced GSH to oxidized GSH (GSH/GSSG) in human lens epithelial cells (HLECs) was a result of EGCNPs displacing POD activity [337].

Renal aging

The aging process is associated with a variety of structural and functional changes in the kidneys and a decreased ability to recover from a kidney injury, both of which contribute to long-term renal outcomes: over 60 percent of people aged 80 and older are diagnosed with chronic kidney disease (CKD) [338, 339]. The loss of renal mass, glomerulosclerosis, glomerular basement membrane hypertrophy, tubular atrophy, interstitial fibrosis, and arteriosclerosis are associated with aging kidneys [340, 341]. Renal aging and CKD are linked to elevated OS levels [342]. Multiple studies have linked an increase in ROS markers to a decline in kidney function beginning in the early phases of CKD in adults and children [343, 344]. As the disease progresses, OS indicators such as mitochondrial superoxide, oxidized LDL, homocysteine, SOD, and GSH deficiencies increase in concentration [345, 346]. This overall increased oxidative burden may contribute to chronic cellular stress, mitochondrial injury, apoptosis, and may induce tubular cell injury [347, 348]. Antioxidants are potential anti-aging strategies for the kidneys.

Yuh-Feng Lin et al. attached anti-kidney injury molecule-1 antibodies to resveratrol-loaded PLGA nanoparticles (KIM-1-Res NPs). The unique KIM-1-Res NPs may accurately control medication release, directly target damaged kidney cells, limit side effects, and improve therapeutic results. Molecular-1-Res NPs decreased creatinine and prevented tubulointerstitial damage in CKD mice [349].

In order to protect renal cells from oxidative damage, Fong-Yu Cheng et al. [350] investigated whether thapsigargin (TG)-encapsulated PLGA nanoparticles (TG-PLGA NPs) might promote autophagy. Nrf2 and forkhead box, class O (FoxO1) were activated by the TG NPs to rescue HK-2 cells from OS-induced cell death. Through the production of ER stress and its downstream pathways, the antibody-conjugated TG NPs reduced kidney dysfunction and damage.

As a result of its ability to shield thiol-containing proteins (antioxidant enzymes), zinc has been touted as a pro-antioxidant agent [351]. In order to tackle CKD [352], researchers employed a combination of spironolactone (SPL) and zinc oxide nanoparticles (ZnO-NPs). The antioxidant and anti-inflammatory effects of ZnO-NPs significantly improved the therapeutic efficacy of SPL in the treatment of CKD.

Although ROS-scavenging nanoparticles have promising anti-aging effects, they may potentially trigger OS and mitochondrial dysfunction in the kidneys if used in excess. Several fundings have shown that multi-walled carbon nanotubes (MWCNTs) [353], AuNPs [354], Silver nanoparticles (AgNPs) [355], copper nanoparticles (CuNPs) [356], Pt NPs [357] could induce renal injury. The trade-off between biological toxicity and therapeutic efficacy of nanoparticles remains to be explored in more depth in future studies. NMs may accumulate in the kidneys due to their small size and unique surface properties. The extent of bioaccumulation due to repeated dosage over long periods of time is still unknown [358] [359].

Clinical trials of ROS-scavenging nanotechnology in treatment of ARD

Even though ROS-scavenging nanotechnology has been the subject for the treatment of ARD, only a handful of these treatments have advanced to the stage of clinical trials (Table 5). The current clinical trial studies about ROS-scavenging nanotechnology suffer from scarcity of trial conduct, small sample size, heterogeneity of study population, diversity of antioxidants, and absence of uniform clinical endpoint indicators. Further studies comparing ROS-scavenging nanotechnology with traditional antioxidants or combinations of them are even more scarce. The efficacy and safety of many antioxidants are currently unknown. and more research, especially clinical trials, are needed to further validate them. The creation, translation, clinical studies, and even the drive toward actual patient usage of nanotechnology still have a great deal of unfinished business.

Table 5 Clinical Trials of ROS-Scavenging Nanotechnology in treatment of ARD

Conclusions and future perspective

In this review, we provided an introduction to ROS-scavenging nanomaterials, discussed their use in the study of aging, and outlined directions for future research. There are significant obstacles to the clinical translation of ROS-scavenging nanotechnology in aging and ARD, despite the encouraging results from preclinical investigations and clinical trials. Nanomaterials that can scavenge ROS have the potential to outperform current antioxidant treatments, increasing human longevity and enhancing quality of life. However, there are still issues to be resolved, such as the effectiveness of nanoparticles for targeted delivery, the safety of nanomaterials, and a dynamic monitoring system for antioxidant nanomedicine. The root cause, location, lesion micro-environment, and gene expression/signaling pathway modifications of each illness are unique. Nanomaterials should be developed with these features in mind.

Overall, current enhancements in nanomedicine primarily focus on:

Precision targeting The development of nanomedicines with targeted capabilities ensures the concentration of drugs at lesion sites, thereby reducing the impact on healthy tissue. This strategy significantly enhances treatment precision. Existing antioxidant-based treatments lack specificity for dysfunctional cells, tissues, and organelles. Antioxidants are frequently not designed to act selectively on senescent cells, which creates uncertainty regarding their actual efficacy and biosafety. In addition, certain biological barriers can impede the accumulation of nanomaterials at disease sites and reduce the efficacy of therapies. Nanodrug delivery may be severely hampered by the non-specific absorption of nanodrugs by healthy organs, one of the common biological barriers. Several strategies have been proposed to combat non-specific absorption by extending the half-life of nanodrugs in circulation. Clinical contexts have utilized PEGylated NP strategies that inhibit clearance by the reticuloendothelial system (RES) or mononuclear phagocyte system (MPS) [102, 111]. Advancements in nanoparticle surface functionalization, such as pH, redox, and light responsiveness. Targeting ligands is also emerging as a promising avenue of research. Nanodrug surface modification of targeting ligands can identify overexpressed receptors in pathological tissues and facilitate site-specific nanodrug delivery [366]. Many targeting ligands such as aptamers, nanoantibodies, small molecules and peptides [367], have been widely used for tumor-targeting nanodrugs. While non-tumor disorders are the principal indication for ROS-eliminating nanodrugs, identifying appropriate ligands will be a fruitful field of study. Drug distribution is greatly hampered by the BBB, a critical barrier in a number of neurological illnesses. The BBB has been approached using a variety of approaches, including chemical alteration of medicines and prodrugs, local distribution mediated by NPs, disruption of the BBB, and different nanocarriers that can cross the BBB [212, 368].

Biodegradable materials There’s a concentrated effort to use natural, biocompatible, and easily biodegradable materials for creating nanoparticles. This approach aims to minimize toxicity and the risk of bioaccumulation, ensuring that these nanoparticles can be safely decomposed and cleared from the body post drug release.

Optimized nanoparticle design Fine-tuning the size, charge, shape, solubility, and surface properties of nanoparticles can improve their distribution and excretion in targeted tissues, reducing systemic toxicity. For example, smaller nanoparticles in the reproductive system have been linked to reduced sperm count and vitality, potentially leading to damage in cumulus cells and hindering egg maturation [369]. It's observed that cationic nanoparticles typically exhibit greater toxicity than their neutral or anionic counterparts. Surface modifications of nanoparticles, such as glycosylation, acetylation, PEGylation, or peptide modification, can enhance biocompatibility, decreasing immune responses and toxicity. Furthermore, the synergistic use of adjuvants, like permeation enhancers in the skin, can temporarily alter the skin's barrier function to boost the transdermal absorption of nanomedicine. For instance, in cardiovascular applications, stimulus-responsive nanoparticles that react to changes within the blood vessels (such as shear stress) or to external stimuli, like magnetic and temperature-sensitive nanoparticles, present innovative therapeutic possibilities [370].

Controlled release systems High reactivity, poor storage ability, and limited bioavailability during in vivo distribution characterize antioxidants. Because encapsulation techniques rely mostly on the passive release or diffusion of antioxidant chemicals, they can't be used for sustained and regulated treatment. Antioxidants in nanomaterials that escape before they reach the site of action may have diminished therapeutic efficacy or even harmful side effects. Enhanced hydrophobic contacts, electrostatic interactions, van der Waals forces, π—π stacking, hydrogen bonding, and covalent bonding are only some of the common interactions used to stabilize nanomaterials for drug delivery platforms [371, 372]. Due to its removal in an acidic intercellular environment, MnO2 and ZnO could be utilized as gatekeepers to efficiently restrict medication leakage. The release of antioxidants may need to be balanced with the biodegradability of the biomaterial. Many loaded antioxidant components are released too quickly, in an incomplete form, or are unstable after release [373]. Nanotechnology-based controlled release systems represent a future-worthy area of development, which enable precise drug delivery, improved bioavailability, targeted therapy with minimal side effects, and the capability for simultaneous multi-drug delivery.

Safety assessment There is also worry over the toxicity caused by ROS that are created by nanomaterials [374]. In particular, metal nanoparticles can affect the expression of neurotransmitters [281], trigger inflammatory responses, and cause OS. Small nanoparticles generate more ROS [375], because they have a larger specific surface area and greater surface reactivity than larger nanomaterials. Nanomaterials' ability to generate ROS is influenced by a number of physical and chemical characteristics. Too much ROS can be produced if the Fenton reaction speeds up (as it could if the concentration of Cu2+ and Cu+ were both raised). By generating oxygen radicals and causing the oxidation and cross-linking of protein thiol groups necessary for cell viability, Se in excess leads to apoptosis [376]. Intensify research into the safety of nanomedicines, encompassing systematic evaluations of their biodistribution, metabolic pathways, long-term stability, and potential toxicity within the body.

The majority of antioxidant clinical trials have been conducted on patients with established pathology. However, once senescent cells manifest, antioxidants are unable to reverse their condition. In our summary of above studies, nowadays research concentrate more on synchronous intervention or post-modelling treatment. Prevention of disease may be more realistic than cure. On the individual oxidative status, the interaction of multiple compounds from diet or supplements, the optimal type of antioxidant, exact dosage, treatment intervals, and total duration of therapy, there are numerous unanswered questions. ROS-based nanomaterials should be combined with other therapeutic methods for improved outcomes. The production or elimination of intracellular ROS is dynamic in space and time when nanomaterials are introduced [377]. The development of techniques to monitor and identify the capabilities of particular ROS in real time is also crucial and essential. It’s best to keep things as straightforward as possible when designing nanomaterials, as more intricate structural and functional designs make manufacturing in bulk more challenging and less reliable.

Availability of data and materials

Data will be made available on request.

References

  1. Zhang Y-S, Jin Y, Rao W-W, Cui L-J, Li J-F, Li L, et al. Prevalence and socio-demographic correlates of major depressive disorder in older adults in Hebei province China. J Affect Disorders. 2020;265:590–4.

    Article  CAS  PubMed  Google Scholar 

  2. López-Otín C, Blasco MA, Partridge L, Serrano M, Kroemer G. The hallmarks of aging. Cell. 2013;153:1194–217.

    Article  PubMed  PubMed Central  Google Scholar 

  3. Davalli P, Mitic T, Caporali A, Lauriola A, D’Arca D. ROS, cell senescence, and novel molecular mechanisms in aging and age-related diseases. Oxid Med Cell Longev. 2016;2016:3565127.

    Article  PubMed  PubMed Central  Google Scholar 

  4. Bjorksten J. The crosslinkage theory of aging. J Am Geriatr Soc. 1968;16:408–27.

    Article  CAS  PubMed  Google Scholar 

  5. Harman D. Aging: a theory based on free radical and radiation chemistry. J Gerontol. 1956;11:298–300.

    Article  CAS  PubMed  Google Scholar 

  6. Kruk PA, Rampino NJ, Bohr VA. DNA damage and repair in telomeres: relation to aging. Proc Natl Acad Sci USA. 1995;92:258–62.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  7. Effros RB. Roy Walford and the immunologic theory of aging. Immun Ageing. 2005;2:7.

    Article  PubMed  PubMed Central  Google Scholar 

  8. Tan BL, Norhaizan ME, Liew W-P-P, Sulaiman RH. Antioxidant and oxidative stress a mutual interplay in age-related diseases. Front Pharmacol. 2018. https://doi.org/10.3389/fphar.2018.01162.

    Article  PubMed  PubMed Central  Google Scholar 

  9. Buzea C, Pacheco II, Robbie K. Nanomaterials and nanoparticles: sources and toxicity. Biointerphases. 2007. https://doi.org/10.1116/1.2815690.

    Article  PubMed  Google Scholar 

  10. Tan BL, Norhaizan ME, Liew WP, Sulaiman RH. Antioxidant and oxidative stress: a mutual interplay in age-related diseases. Front Pharmacol. 2018;9:1162.

    Article  PubMed  PubMed Central  Google Scholar 

  11. Li K, Liang N, Yang H, Liu H, Li S. Temozolomide encapsulated and folic acid decorated chitosan nanoparticles for lung tumor targeting: improving therapeutic efficacy both and. Oncotarget. 2017;8:111318–32.

    Article  PubMed  PubMed Central  Google Scholar 

  12. Yin J-J, Lao F, Fu PP, Wamer WG, Zhao Y, Wang PC, et al. The scavenging of reactive oxygen species and the potential for cell protection by functionalized fullerene materials. Biomaterials. 2009;30:611–21.

    Article  CAS  PubMed  Google Scholar 

  13. Moglianetti M, De Luca E, Pedone D, Marotta R, Catelani T, Sartori B, et al. Platinum nanozymes recover cellular ROS homeostasis in an oxidative stress-mediated disease model. Nanoscale. 2016;8:3739–52.

    Article  CAS  PubMed  Google Scholar 

  14. Nelson BC, Johnson ME, Walker ML, Riley KR, Sims CMJA. Antioxidant cerium oxide nanoparticles in biology and medicine. Antioxidants. 2016;5:15.

    Article  PubMed  PubMed Central  Google Scholar 

  15. Trachootham D, Alexandre J, Huang P. Targeting cancer cells by ROS-mediated mechanisms: a radical therapeutic approach? Nat Rev Drug Discov. 2009;8:579–91.

    Article  CAS  PubMed  Google Scholar 

  16. Zuo L, Zhou T, Pannell BK, Ziegler AC, Best TM. Biological and physiological role of reactive oxygen species–the good, the bad and the ugly. Acta Physiol. 2015;214:329–48.

    Article  CAS  Google Scholar 

  17. Massy ZA, Nguyen-Khoa T. Oxidative stress and chronic renal failure: markers and management. J Nephrol. 2002;15:336–41.

    CAS  PubMed  Google Scholar 

  18. Berry BJ, Trewin AJ, Amitrano AM, Kim M, Wojtovich AP. Use the protonmotive force: mitochondrial uncoupling and reactive oxygen species. J Mol Biol. 2018;430:3873–91.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  19. Handy DE, Loscalzo J. Redox regulation of mitochondrial function. Antioxid Redox Signal. 2012;16:1323–67.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  20. Delaunay-Moisan A, Appenzeller-Herzog C. The antioxidant machinery of the endoplasmic reticulum: protection and signaling. Free Radical Biol Med. 2015;83:341–51.

    Article  CAS  Google Scholar 

  21. Zeeshan HMA, Lee GH, Kim H-R, Chae H-J. Endoplasmic reticulum stress and associated ROS. Int J Mol Sci. 2016;17:327.

    Article  PubMed  PubMed Central  Google Scholar 

  22. Lefranc C, Friederich-Persson M, Palacios-Ramirez R, Cat ND. Mitochondrial oxidative stress in obesity: role of the mineralocorticoid receptor. J Endocrinol. 2018. https://doi.org/10.1530/JOE-18-0163.

    Article  PubMed  Google Scholar 

  23. Bedard K, Krause K-H. The NOX family of ROS-generating NADPH oxidases: physiology and pathophysiology. Physiol Rev. 2007;87:245–313.

    Article  CAS  PubMed  Google Scholar 

  24. Bozzo F, Mirra A, Carrì MT. Oxidative stress and mitochondrial damage in the pathogenesis of ALS: new perspectives. Neurosci Lett. 2017;636:3–8.

    Article  CAS  PubMed  Google Scholar 

  25. Phaniendra A, Jestadi DB, Periyasamy L. Free radicals: properties, sources, targets, and their implication in various diseases. Indian J Clin Biochem IJCB. 2015;30:11–26.

    Article  CAS  PubMed  Google Scholar 

  26. Ullah R, Khan M, Shah SA, Saeed K, Kim MO. Natural antioxidant anthocyanins-a hidden therapeutic candidate in metabolic disorders with major focus in neurodegeneration. Nutrients. 2019. https://doi.org/10.3390/nu11061195.

    Article  PubMed  PubMed Central  Google Scholar 

  27. Espinosa-Diez C, Miguel V, Mennerich D, Kietzmann T, Sánchez-Pérez P, Cadenas S, et al. Antioxidant responses and cellular adjustments to oxidative stress. Redox Biol. 2015;6:183–97.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  28. Sies H, Jones DP. Reactive oxygen species (ROS) as pleiotropic physiological signalling agents. Nat Rev Mol Cell Biol. 2020;21:363–83.

    Article  CAS  PubMed  Google Scholar 

  29. Sies H, Belousov VV, Chandel NS, Davies MJ, Jones DP, Mann GE, et al. Defining roles of specific reactive oxygen species (ROS) in cell biology and physiology. Nat Rev Mol Cell Biol. 2022;23:499–515.

    Article  CAS  PubMed  Google Scholar 

  30. Dai D-F, Chiao YA, Marcinek DJ, Szeto HH, Rabinovitch PS. Mitochondrial oxidative stress in aging and healthspan. Longev Healthspan. 2014;3:6.

    Article  PubMed  PubMed Central  Google Scholar 

  31. Yin H, Xu L, Porter NA. Free radical lipid peroxidation: mechanisms and analysis. Chem Rev. 2011;111:5944–72.

    Article  CAS  PubMed  Google Scholar 

  32. Yadav DK, Kumar S, Choi E-H, Chaudhary S, Kim M-H. Molecular dynamic simulations of oxidized skin lipid bilayer and permeability of reactive oxygen species. Sci Rep. 2019;9:4496.

    Article  PubMed  PubMed Central  Google Scholar 

  33. Luo J, Mills K, le Cessie S, Noordam R, van Heemst D. Ageing, age-related diseases and oxidative stress: What to do next? Ageing Res Rev. 2020;57: 100982.

    Article  CAS  PubMed  Google Scholar 

  34. Davies MJ. Protein oxidation and peroxidation. Biochem J. 2016;473:805–25.

    Article  CAS  PubMed  Google Scholar 

  35. Stadtman ER. Protein oxidation and aging. Free Radical Res. 2006;40:1250–8.

    Article  CAS  Google Scholar 

  36. Berlett BS, Stadtman ER. Protein oxidation in aging, disease, and oxidative stress*. J Biol Chem. 1997;272:20313–6.

    Article  CAS  PubMed  Google Scholar 

  37. Tanaka K. The proteasome: overview of structure and functions. Proc Jpn Acad Ser B Phys Biol Sci. 2009;85:12–36.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  38. Jung T, Bader N, Grune T. Oxidized proteins: intracellular distribution and recognition by the proteasome. Arch Biochem Biophys. 2007;462:231–7.

    Article  CAS  PubMed  Google Scholar 

  39. Zavadskiy S, Sologova S, Moldogazieva N. Oxidative distress in aging and age-related diseases: Spatiotemporal dysregulation of protein oxidation and degradation. Biochimie. 2022;195:114–34.

    Article  CAS  PubMed  Google Scholar 

  40. Salehi F, Behboudi H, Kavoosi G, Ardestani SK. Oxidative DNA damage induced by ROS-modulating agents with the ability to target DNA: a comparison of the biological characteristics of citrus pectin and apple pectin. Sci Rep. 2018;8:13902.

    Article  PubMed  PubMed Central  Google Scholar 

  41. Dizdaroglu M, Gajewski E, Reddy P, Margolis SA. Structure of a hydroxyl radical-induced DNA-protein crosslink involving thymine and tyrosine in nucleohistone. Biochemistry. 1989;28:3625–8.

    Article  CAS  PubMed  Google Scholar 

  42. Marnett LJ. Oxy radicals, lipid peroxidation and DNA damage. Toxicology. 2002;181–182:219–22.

    Article  PubMed  Google Scholar 

  43. Kaufman BA, Durisic N, Mativetsky JM, Costantino S, Hancock MA, Grutter P, et al. The mitochondrial transcription factor TFAM coordinates the assembly of multiple DNA molecules into nucleoid-like structures. Mol Biol Cell. 2007;18:3225–36.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  44. Vijg J. From DNA damage to mutations: all roads lead to aging. Ageing Res Rev. 2021;68: 101316.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  45. Nelson G, Wordsworth J, Wang C, Jurk D, Lawless C, Martin-Ruiz C, et al. A senescent cell bystander effect: senescence-induced senescence. Aging Cell. 2012;11:345–9.

    Article  CAS  PubMed  Google Scholar 

  46. Bohr VA, Anson RM. Mitochondrial DNA repair pathways. J Bioenerg Biomembr. 1999;31:391–8.

    Article  CAS  PubMed  Google Scholar 

  47. Trifunovic A, Hansson A, Wredenberg A, Rovio AT, Dufour E, Khvorostov I, et al. Somatic mtDNA mutations cause aging phenotypes without affecting reactive oxygen species production. Proc Natl Acad Sci USA. 2005;102:17993–8.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  48. Das A. The emerging role of microplastics in systemic toxicity: involvement of reactive oxygen species (ROS). Sci Total Environ. 2023;895: 165076.

    Article  CAS  PubMed  Google Scholar 

  49. Kong Q, Lin C-LG. Oxidative damage to RNA: mechanisms, consequences, and diseases. Cell Mol Life Sci CMLS. 2010;67:1817–29.

    Article  CAS  PubMed  Google Scholar 

  50. Tanaka M, Chock PB, Stadtman ER. Oxidized messenger RNA induces translation errors. Proc Natl Acad Sci USA. 2007;104:66–71.

    Article  CAS  PubMed  Google Scholar 

  51. Shibuya S, Ozawa Y, Watanabe K, Izuo N, Toda T, Yokote K, et al. Palladium and platinum nanoparticles attenuate aging-like skin atrophy via antioxidant activity in mice. PLoS ONE. 2014;9: e109288.

    Article  PubMed  PubMed Central  Google Scholar 

  52. Sims CM, Hanna SK, Heller DA, Horoszko CP, Johnson ME, Montoro Bustos AR, et al. Redox-active nanomaterials for nanomedicine applications. Nanoscale. 2017;9(15226):15251.

    Google Scholar 

  53. Jawaid P, Rehman Mu, Yoshihisa Y, Li P, Ql Z, Hassan MA, et al. Effects of SOD/catalase mimetic platinum nanoparticles on radiation-induced apoptosis in human lymphoma U937 cells. Apoptosis Int J Program Cell Death. 2014;19:1006–16.

    Article  CAS  Google Scholar 

  54. Zhang L, Laug L, Münchgesang W, Pippel E, Gösele U, Brandsch M, et al. Reducing stress on cells with apoferritin-encapsulated platinum nanoparticles. Nano Lett. 2010;10:219–23.

    Article  PubMed  Google Scholar 

  55. Yoshihisa Y, Zhao Q-L, Hassan MA, Wei Z-L, Furuichi M, Miyamoto Y, et al. SOD/catalase mimetic platinum nanoparticles inhibit heat-induced apoptosis in human lymphoma U937 and HH cells. Free Radical Res. 2011;45:326–35.

    Article  CAS  Google Scholar 

  56. Kajita M, Hikosaka K, Iitsuka M, Kanayama A, Toshima N, Miyamoto Y. Platinum nanoparticle is a useful scavenger of superoxide anion and hydrogen peroxide. Free Radical Res. 2007;41:615–26.

    Article  CAS  Google Scholar 

  57. Zhang W, Hu S, Yin J-J, He W, Lu W, Ma M, et al. Prussian blue nanoparticles as multienzyme mimetics and reactive oxygen species scavengers. J Am Chem Soc. 2016;138:5860–5.

    Article  CAS  PubMed  Google Scholar 

  58. Liu T, Xiao B, Xiang F, Tan J, Chen Z, Zhang X, et al. Ultrasmall copper-based nanoparticles for reactive oxygen species scavenging and alleviation of inflammation related diseases. Nat Commun. 2020;11:2788.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  59. Tang Y, Han Y, Zhao J, Lv Y, Fan C, Zheng L, et al. A rational design of metal-organic framework nanozyme with high-performance copper active centers for alleviating chemical corneal burns. Nano-micro Lett. 2023;15:112.

    Article  CAS  Google Scholar 

  60. Yang J, Zhang R, Zhao H, Qi H, Li J, Li J-F, et al. Bioinspired copper single-atom nanozyme as a superoxide dismutase-like antioxidant for sepsis treatment. Exploration. 2022;2:20210267.

    Article  PubMed  PubMed Central  Google Scholar 

  61. Huang Y, Liu Z, Liu C, Ju E, Zhang Y, Ren J, et al. Self-assembly of multi-nanozymes to mimic an intracellular antioxidant defense system. Angew Chem Int Ed Engl. 2016;55:6646–50.

    Article  CAS  PubMed  Google Scholar 

  62. Liu X, Wang Q, Zhao H, Zhang L, Su Y, Lv Y. BSA-templated MnO2 nanoparticles as both peroxidase and oxidase mimics. Analyst. 2012;137:4552–8.

    Article  CAS  PubMed  Google Scholar 

  63. Kumar S, Adjei IM, Brown SB, Liseth O, Sharma B. Manganese dioxide nanoparticles protect cartilage from inflammation-induced oxidative stress. Biomaterials. 2019;224: 119467.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  64. Singh N, Savanur MA, Srivastava S, D’Silva P, Mugesh G. A manganese oxide nanozyme prevents the oxidative damage of biomolecules without affecting the endogenous antioxidant system. Nanoscale. 2019;11:3855–63.

    Article  CAS  PubMed  Google Scholar 

  65. Adebayo OA, Akinloye O, Adaramoye OA. Cerium oxide nanoparticles attenuate oxidative stress and inflammation in the liver of diethylnitrosamine-treated mice. Biol Trace Elem Res. 2020;193:214–25.

    Article  CAS  PubMed  Google Scholar 

  66. Kwon HJ, Cha M-Y, Kim D, Kim DK, Soh M, Shin K, et al. Mitochondria-targeting ceria nanoparticles as antioxidants for Alzheimer’s disease. ACS Nano. 2016;10:2860–70.

    Article  CAS  PubMed  Google Scholar 

  67. Li Y, He X, Yin JJ, Ma Y, Zhang P, Li J, et al. Acquired superoxide-scavenging ability ceria nanoparticles. Angewandte Chemie. 2015;127(1852):1855.

    Google Scholar 

  68. Korsvik C, Patil S, Seal S, Self WT. Superoxide dismutase mimetic properties exhibited by vacancy engineered ceria nanoparticles. Chem Commun. 2007;10:1056–8.

    Article  Google Scholar 

  69. Heckert EG, Karakoti AS, Seal S, Self WT. The role of cerium redox state in the SOD mimetic activity of nanoceria. Biomaterials. 2008;29:2705–9.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  70. Pirmohamed T, Dowding JM, Singh S, Wasserman B, Heckert E, Karakoti AS, et al. Nanoceria exhibit redox state-dependent catalase mimetic activity. Chem Commun. 2010;46:2736–8.

    Article  CAS  Google Scholar 

  71. Dowding JM, Seal S, Self WT. Cerium oxide nanoparticles accelerate the decay of peroxynitrite (ONOO(-)). Drug Deliv Transl Res. 2013;3:375–9.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  72. Wang IC, Tai LA, Lee DD, Kanakamma PP, Shen CK, Luh TY, et al. C(60) and water-soluble fullerene derivatives as antioxidants against radical-initiated lipid peroxidation. J Med Chem. 1999;42:4614–20.

    Article  CAS  PubMed  Google Scholar 

  73. Qiu Y, Wang Z, Owens ACE, Kulaots I, Chen Y, Kane AB, et al. Antioxidant chemistry of graphene-based materials and its role in oxidation protection technology. Nanoscale. 2014;6:11744–55.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  74. Lucente-Schultz RM, Moore VC, Leonard AD, Price BK, Kosynkin DV, Lu M, et al. Antioxidant single-walled carbon nanotubes. J Am Chem Soc. 2009;131:3934–41.

    Article  CAS  PubMed  Google Scholar 

  75. Andrievsky GV, Bruskov VI, Tykhomyrov AA, Gudkov SV. Peculiarities of the antioxidant and radioprotective effects of hydrated C60 fullerene nanostuctures in vitro and in vivo. Free Radical Biol Med. 2009;47:786–93.

    Article  CAS  Google Scholar 

  76. Ali SS, Hardt JI, Quick KL, Kim-Han JS, Erlanger BF, Huang T-T, et al. A biologically effective fullerene (C60) derivative with superoxide dismutase mimetic properties. Free Radical Biol Med. 2004;37:1191–202.

    Article  CAS  Google Scholar 

  77. Zhai X, Zhang C, Zhao G, Stoll S, Ren F, Leng X. Antioxidant capacities of the selenium nanoparticles stabilized by chitosan. J Nanobiotechnol. 2017;15:4.

    Article  Google Scholar 

  78. Bjørklund G, Shanaida M, Lysiuk R, Antonyak H, Klishch I, Shanaida V, et al. Selenium: an antioxidant with a critical role in anti-aging. Molecules. 2022. https://doi.org/10.3390/molecules27196613.

    Article  PubMed  PubMed Central  Google Scholar 

  79. Khurana A, Tekula S, Saifi MA, Venkatesh P, Godugu C. Therapeutic applications of selenium nanoparticles. Biomed Pharmacother Biomed Pharm. 2019;111:802–12.

    Article  CAS  Google Scholar 

  80. Soule BP, Hyodo F, Matsumoto K-I, Simone NL, Cook JA, Krishna MC, et al. The chemistry and biology of nitroxide compounds. Free Radical Biol Med. 2007;42:1632–50.

    Article  CAS  Google Scholar 

  81. Calabrese G, Ardizzone A, Campolo M, Conoci S, Esposito E, Paterniti I. Beneficial effect of tempol, a membrane-permeable radical scavenger, on inflammation and osteoarthritis in in vitro models. Biomolecules. 2011. https://doi.org/10.3390/biom11030352.

    Article  Google Scholar 

  82. Zeng Z, He X, Li C, Lin S, Chen H, Liu L, et al. Oral delivery of antioxidant enzymes for effective treatment of inflammatory disease. Biomaterials. 2021;271: 120753.

    Article  CAS  PubMed  Google Scholar 

  83. Barathmanikanth S, Kalishwaralal K, Sriram M, Pandian SRK, Youn H-S, Eom S, et al. Anti-oxidant effect of gold nanoparticles restrains hyperglycemic conditions in diabetic mice. J Nanobiotechnol. 2010;8:16.

    Article  Google Scholar 

  84. Mahmoudi A, Kesharwani P, Majeed M, Teng Y, Sahebkar A. Recent advances in nanogold as a promising nanocarrier for curcumin delivery. Colloids Surf B. 2022;215: 112481.

    Article  CAS  Google Scholar 

  85. Sathyabhama M, Priya Dharshini LC, Karthikeyan A, Kalaiselvi S, Min T. The credible role of curcumin in oxidative stress-mediated mitochondrial dysfunction in mammals. Biomolecules. 2022. https://doi.org/10.3390/biom12101405.

    Article  PubMed  PubMed Central  Google Scholar 

  86. Lee Y, Kim H, Kang S, Lee J, Park J, Jon S. Bilirubin nanoparticles as a nanomedicine for anti-inflammation therapy. Angew Chem Int Ed Engl. 2016;55:7460–3.

    Article  CAS  PubMed  Google Scholar 

  87. Bao X, Zhao J, Sun J, Hu M, Yang X. Polydopamine nanoparticles as efficient scavengers for reactive oxygen species in periodontal disease. ACS Nano. 2018;12:8882–92.

    Article  CAS  PubMed  Google Scholar 

  88. Motohashi N, Gallagher R, Anuradha V, Gollapudi RJM. Reviews C Co-enzyme Q10 (Ubiquinone): It’s implication in improving the life style of the elderly. Med Clin Rev. 2018. https://doi.org/10.21767/2471-299X.1000052.

    Article  Google Scholar 

  89. Gherardi G, Corbioli G, Ruzza F, Rizzuto R. CoQ and resveratrol effects to ameliorate aged-related mitochondrial dysfunctions. Nutrients. 2022. https://doi.org/10.3390/nu14204326.

    Article  PubMed  PubMed Central  Google Scholar 

  90. Wang Q, Yu Q, Wu M. Antioxidant and neuroprotective actions of resveratrol in cerebrovascular diseases. Front Pharmacol. 2022;13: 948889.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  91. Brewer MS. Natural antioxidants: sources, compounds, mechanisms of action, and potential applications. Comprehensive Rev Food Sci Food Safety. 2011;10:221–47.

    Article  CAS  Google Scholar 

  92. Joshi YB, Praticò D. Vitamin E in aging, dementia, and Alzheimer’s disease. BioFactors. 2012;38:90–7.

    Article  CAS  PubMed  Google Scholar 

  93. Ohsawa I, Ishikawa M, Takahashi K, Watanabe M, Nishimaki K, Yamagata K, et al. Hydrogen acts as a therapeutic antioxidant by selectively reducing cytotoxic oxygen radicals. Nat Med. 2007;13:688–94.

    Article  CAS  PubMed  Google Scholar 

  94. Gokce EH, Korkmaz E, Tuncay-Tanrıverdi S, Dellera E, Sandri G, Bonferoni MC, et al. A comparative evaluation of coenzyme Q10-loaded liposomes and solid lipid nanoparticles as dermal antioxidant carriers. Int J Nanomed. 2012;7:5109–17.

    CAS  Google Scholar 

  95. Wan J, Yang J, Lei W, Xiao Z, Zhou P, Zheng S, et al. Anti-oxidative, anti-apoptotic, and M2 polarized DSPC liposome nanoparticles for selective treatment of atherosclerosis. Int J Nanomed. 2023;18:579–94.

    Article  CAS  Google Scholar 

  96. Gokce EH, Korkmaz E, Dellera E, Sandri G, Bonferoni MC, Ozer O. Resveratrol-loaded solid lipid nanoparticles versus nanostructured lipid carriers: evaluation of antioxidant potential for dermal applications. Int J Nanomed. 2012;7:1841–50.

    Article  CAS  Google Scholar 

  97. Picone P, Bondi ML, Montana G, Bruno A, Pitarresi G, Giammona G, et al. Ferulic acid inhibits oxidative stress and cell death induced by Ab oligomers: improved delivery by solid lipid nanoparticles. Free Radical Res. 2009;43:1133–45.

    Article  CAS  Google Scholar 

  98. Sathya S, Shanmuganathan B, Devi KP. Deciphering the anti-apoptotic potential of alpha-bisabolol loaded solid lipid nanoparticles against a beta induced neurotoxicity in Neuro-2a cells. Coll Surf B-Biointerf. 2020. https://doi.org/10.1016/j.colsurfb.2020.110948.

    Article  Google Scholar 

  99. Liu H, Li Y, Zhang X, Shi M, Li D, Wang Y. Chitosan-coated solid lipid nano-encapsulation improves the therapeutic antiairway inflammation effect of berberine against COPD in cigarette smoke-exposed rats. Can Respir J. 2022;2022:8509396.

    Article  PubMed  PubMed Central  Google Scholar 

  100. Wang J, Wang H, Xia Q. Ubidecarenone-loaded nanostructured lipid carrier (UB-NLC): percutaneous penetration and protective effects against hydrogen peroxide-induced oxidative stress on HaCaT cells. Int J Mol Sci. 2018. https://doi.org/10.3390/ijms19071865.

    Article  PubMed  PubMed Central  Google Scholar 

  101. Kuang Y, Zhang J, Xiong M, Zeng W, Lin X, Yi X, et al. A novel nanosystem realizing curcumin delivery based on feO@Carbon dots nanocomposite for Alzheimer’s disease therapy. Front Bioeng Biotechnol. 2020;8: 614906.

    Article  PubMed  PubMed Central  Google Scholar 

  102. Gui T, Luo L, Chhay B, Zhong L, Wei Y, Yao L, et al. Superoxide dismutase-loaded porous polymersomes as highly efficient antioxidant nanoparticles targeting synovium for osteoarthritis therapy. Biomaterials. 2022;283: 121437.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  103. Dhas N, Mehta T. Cationic biopolymer functionalized nanoparticles encapsulating lutein to attenuate oxidative stress in effective treatment of Alzheimer’s disease: a non-invasive approach. Int J Pharm. 2020;586: 119553.

    Article  CAS  PubMed  Google Scholar 

  104. Ge Y, Zhang A, Sun R, Xu J, Yin T, He H, et al. Penetratin-modified lutein nanoemulsion gel for the treatment of age-related macular degeneration. Expert Opin Drug Deliv. 2020;17:603–19.

    Article  CAS  PubMed  Google Scholar 

  105. Pangeni R, Sharma S, Mustafa G, Ali J, Baboota S. Vitamin E loaded resveratrol nanoemulsion for brain targeting for the treatment of Parkinson’s disease by reducing oxidative stress. Nanotechnology. 2014;25: 485102.

    Article  PubMed  Google Scholar 

  106. Shea TB, Ortiz D, Nicolosi RJ, Kumar R, Watterson AC. Nanosphere-mediated delivery of vitamin E increases its efficacy against oxidative stress resulting from exposure to amyloid beta. J Alzheimer’s Dis JAD. 2005;7:297–301.

    Article  CAS  PubMed  Google Scholar 

  107. Onoue S, Terasawa N, Nakamura T, Yuminoki K, Hashimoto N, Yamada S. Biopharmaceutical characterization of nanocrystalline solid dispersion of coenzyme Q10 prepared with cold wet-milling system. Eur J Pharm Sci. 2014;53:118–25.

    Article  CAS  PubMed  Google Scholar 

  108. Shen J-W, Li C, Yang M-Y, Lin J-F, Yin M-D, Zou J-J, et al. Biomimetic nanoparticles: U937 cell membranes based core-shell nanosystems for targeted atherosclerosis therapy. Int J Pharm. 2022;611: 121297.

    Article  CAS  PubMed  Google Scholar 

  109. Tiwari MN, Agarwal S, Bhatnagar P, Singhal NK, Tiwari SK, Kumar P, et al. Nicotine-encapsulated poly(lactic-co-glycolic) acid nanoparticles improve neuroprotective efficacy against MPTP-induced parkinsonism. Free Radical Biol Med. 2013;65:704–18.

    Article  CAS  Google Scholar 

  110. Singhal A, Morris VB, Labhasetwar V, Ghorpade A. Nanoparticle-mediated catalase delivery protects human neurons from oxidative stress. Cell Death Dis. 2013;4: e903.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  111. Amin FU, Shah SA, Badshah H, Khan M, Kim MO. Anthocyanins encapsulated by PLGA@PEG nanoparticles potentially improved its free radical scavenging capabilities via p38/JNK pathway against Aβ-induced oxidative stress. J Nanobiotechnol. 2017;15:12.

    Article  Google Scholar 

  112. Sathya S, Manogari BG, Thamaraiselvi K, Vaidevi S, Ruckmani K, Devi KP. Phytol loaded PLGA nanoparticles ameliorate scopolamine-induced cognitive dysfunction by attenuating cholinesterase activity, oxidative stress and apoptosis in Wistar rat. Nutr Neurosci. 2022;25:485–501.

    Article  CAS  PubMed  Google Scholar 

  113. Mulik RS, Mönkkönen J, Juvonen RO, Mahadik KR, Paradkar AR. ApoE3 mediated poly(butyl) cyanoacrylate nanoparticles containing curcumin: study of enhanced activity of curcumin against beta amyloid induced cytotoxicity using in vitro cell culture model. Mol Pharm. 2010;7:815–25.

    Article  CAS  PubMed  Google Scholar 

  114. Mirzaie Z, Ansari M, Kordestani SS, Rezaei MH, Mozafari M. Preparation and characterization of curcumin-loaded polymeric nanomicelles to interference with amyloidogenesis through glycation method. Biotechnol Appl Biochem. 2019;66:537–44.

    Article  CAS  PubMed  Google Scholar 

  115. Ogurtsova K, da Rocha Fernandes JD, Huang Y, Linnenkamp U, Guariguata L, Cho NH, et al. IDF Diabetes Atlas: global estimates for the prevalence of diabetes for 2015 and 2040. Diabetes Res Clin Pract. 2017;128:40–50.

    Article  CAS  PubMed  Google Scholar 

  116. Brownlee M. The pathobiology of diabetic complications: a unifying mechanism. Diabetes. 2005;54:1615–25.

    Article  CAS  PubMed  Google Scholar 

  117. Aronson D. Hyperglycemia and the pathobiology of diabetic complications. Adv Cardiol. 2008. https://doi.org/10.1159/000115118.

    Article  PubMed  Google Scholar 

  118. Shin CS, Moon BS, Park KS, Kim SY, Park SJ, Chung MH, et al. Serum 8-hydroxy-guanine levels are increased in diabetic patients. Diabetes Care. 2001;24:733–7.

    Article  CAS  PubMed  Google Scholar 

  119. Sakuraba H, Mizukami H, Yagihashi N, Wada R, Hanyu C, Yagihashi S. Reduced beta-cell mass and expression of oxidative stress-related DNA damage in the islet of Japanese type II diabetic patients. Diabetologia. 2002;45:85–96.

    Article  CAS  PubMed  Google Scholar 

  120. Rehman K, Akash MSH. Mechanism of generation of oxidative stress and pathophysiology of type 2 diabetes mellitus: how are they interlinked? J Cell Biochem. 2017;118:3577–85.

    Article  CAS  PubMed  Google Scholar 

  121. Ahmed HH, Abd El-Maksoud MD, Abdel Moneim AE, Aglan HA. Pre-clinical study for the antidiabetic potential of selenium nanoparticles. Biol Trace Elem Res. 2017;177:267–80.

    Article  CAS  PubMed  Google Scholar 

  122. Wang L, Li C, Huang Q, Fu X. Biofunctionalization of selenium nanoparticles with a polysaccharide from Rosa roxburghii fruit and their protective effect against HO-induced apoptosis in INS-1 cells. Food Funct. 2019;10:539–53.

    Article  CAS  PubMed  Google Scholar 

  123. Pourkhalili N, Hosseini A, Nili-Ahmadabadi A, Hassani S, Pakzad M, Baeeri M, et al. Biochemical and cellular evidence of the benefit of a combination of cerium oxide nanoparticles and selenium to diabetic rats. World J Diabetes. 2011;2:204–10.

    Article  PubMed  PubMed Central  Google Scholar 

  124. Sengani M, Rajeswari D. Gold nanosupplement in selective inhibition of methylglyoxal and key enzymes linked to diabetes. IET Nanobiotechnol. 2017;11:861–5.

    Article  PubMed Central  Google Scholar 

  125. Shaheen TI, El-Naggar ME, Hussein JS, El-Bana M, Emara E, El-Khayat Z, et al. Antidiabetic assessment; in vivo study of gold and core-shell silver-gold nanoparticles on streptozotocin-induced diabetic rats. Biomed Pharmacother Biomed Pharm. 2016;83:865–75.

    Article  CAS  Google Scholar 

  126. Singh TA, Sharma A, Tejwan N, Ghosh N, Das J, Sil PC. A state of the art review on the synthesis, antibacterial, antioxidant, antidiabetic and tissue regeneration activities of zinc oxide nanoparticles. Adv Coll Interf Sci. 2021;295: 102495.

    Article  CAS  Google Scholar 

  127. Umrani RD, Paknikar KM. Zinc oxide nanoparticles show antidiabetic activity in streptozotocin-induced type 1 and 2 diabetic rats. Nanomedicine. 2014;9(1):89.

    Article  CAS  PubMed  Google Scholar 

  128. Robkhob P, Ghosh S, Bellare J, Jamdade D, Tang IM, Thongmee S. Effect of silver doping on antidiabetic and antioxidant potential of ZnO nanorods. J Trace Elements Med Biol Organ Soc MinTrace Elements. 2020;58: 126448.

    Article  CAS  Google Scholar 

  129. Wu J, Chen Y, Li X, Ran L, Liu X, Wang X, et al. Functionalized gadofullerene ameliorates impaired glycolipid metabolism in type 2 diabetic mice. J Genet Genomics. 2022;49:364–76.

    Article  CAS  PubMed  Google Scholar 

  130. Förstermann U, Xia N, Li H. Roles of vascular oxidative stress and nitric oxide in the pathogenesis of atherosclerosis. Circ Res. 2017;120:713–35.

    Article  PubMed  Google Scholar 

  131. Harrison D, Griendling KK, Landmesser U, Hornig B, Drexler H. Role of oxidative stress in atherosclerosis. 2003;91:7–11.

    Google Scholar 

  132. Wang JC, Bennett M. Aging and atherosclerosis: mechanisms, functional consequences, and potential therapeutics for cellular senescence. Circ Res. 2012;111:245–59.

    Article  CAS  PubMed  Google Scholar 

  133. Cai M. Yang EJJTAJoCM Ginsenoside Re attenuates neuroinflammation in a symptomatic ALS animal model. Am J Chinese Med. 2016;44(401):413.

    Google Scholar 

  134. Lim KH, Lim D-J. Kim J-HJJogr ginsenoside-re ameliorates ischemia and reperfusion injury in the heart: a hemodynamics approach. J Ginseng Res. 2013;37:283.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  135. Sun Y, Liu Y, Chen K. Roles and mechanisms of ginsenoside in cardiovascular diseases: progress and perspectives. Sci China Life Sci. 2016;59:292–8.

    Article  CAS  PubMed  Google Scholar 

  136. Liu Z-Q, Luo X-Y, Liu G-Z, Chen Y-P, Wang Z-C, et al. In vitro study of the relationship between the structure of ginsenoside and its antioxidative or prooxidative activity in free radical induced hemolysis of human erythrocytes. J Agric Food Chem. 2003;51(2555):2558.

    Google Scholar 

  137. Wu T, Chen X, Wang Y, Xiao H, Peng Y, Lin L, et al. Aortic plaque-targeted andrographolide delivery with oxidation-sensitive micelle effectively treats atherosclerosis via simultaneous ROS capture and anti-inflammation. Nanomed Nanotechnol Biol Med. 2018;14:2215–26.

    Article  CAS  Google Scholar 

  138. Shen M, Li H, Yao S, Wu X, Liu S, Yang Q, et al. Shear stress and ROS-responsive biomimetic micelles for atherosclerosis via ROS consumption. Mater Sci Eng, C Mater Biol Appl. 2021;126: 112164.

    Article  CAS  PubMed  Google Scholar 

  139. Chmielowski RA, Abdelhamid DS, Faig JJ, Petersen LK, Gardner CR, Uhrich KE, et al. Athero-inflammatory nanotherapeutics: Ferulic acid-based poly(anhydride-ester) nanoparticles attenuate foam cell formation by regulating macrophage lipogenesis and reactive oxygen species generation. Acta Biomater. 2017;57:85–94.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  140. Chen S, Hou Y, Cheng G, Zhang C, Wang S, Zhang J. Cerium oxide nanoparticles protect endothelial cells from apoptosis induced by oxidative stress. Biol Trace Elem Res. 2013;154:156–66.

    Article  CAS  PubMed  Google Scholar 

  141. Gao Y, Liu S, Zeng X, Guo Z, Chen D, Li S, et al. Reduction of reactive oxygen species accumulation using gadolinium-doped ceria for the alleviation of atherosclerosis. ACS Appl Mater Interf. 2023;15:10414–25.

    Article  CAS  Google Scholar 

  142. Chen L, Yang J, Fu X, Huang W, Yu X, Leng F, et al. A targeting mesoporous dopamine nanodrug platform with NIR responsiveness for atherosclerosis improvement. Biomater Adv. 2022;136: 212775.

    Article  CAS  PubMed  Google Scholar 

  143. Sun W, Xu Y, Yao Y, Yue J, Wu Z, Li H, et al. Self-oxygenation mesoporous MnO nanoparticles with ultra-high drug loading capacity for targeted arteriosclerosis therapy. J Nanobiotechnol. 2022;20:88.

    Article  CAS  Google Scholar 

  144. Sw A, Yue ZB, Xl B, Min XB, Nan LB, Kang ZA. Platinum-cerium bimetallic nano-raspberry for atherosclerosis treatment via synergistic foam cell inhibition and P2Y12 targeted antiplatelet aggregation. Chem Eng J. 2021;430:132859.

    Google Scholar 

  145. Wang Y, Li L, Zhao W, Dou Y, An H, Tao H, et al. Targeted therapy of atherosclerosis by a broad-spectrum reactive oxygen species scavenging nanoparticle with intrinsic anti-inflammatory activity. ACS Nano. 2018;12:8943–60.

    Article  CAS  PubMed  Google Scholar 

  146. Wu Z, Wu R, Li X, Wang X, Tang X, Tan K, et al. Multi-pathway microenvironment regulation for atherosclerosis therapy based on beta-cyclodextrin/L-Arginine/Au nanomotors with dual-mode propulsion. Small. 2022;18: e2104120.

    Article  PubMed  Google Scholar 

  147. Hu R, Dai C, Dong C, Ding L, Huang H, Chen Y, et al. Living Macrophage-delivered tetrapod PdH nanoenzyme for targeted atherosclerosis management by ROS scavenging, hydrogen anti-inflammation, and autophagy activation. ACS Nano. 2022;16:15959–76.

    Article  CAS  PubMed  Google Scholar 

  148. Xu M, Zhang X, Dong B, Wang W, Zhao Z. Sustained Release of Hydrogen by PdH-Te nanozyme for anti-inflammatory therapy against atherosclerosis. Particle & Particle Systems Characterization

  149. Zhang Y, Yin Y, Zhang W, Li H, Wang T, Yin H, et al. Reactive oxygen species scavenging and inflammation mitigation enabled by biomimetic prussian blue analogues boycott atherosclerosis. J Nanobiotechnol. 2021;19:161.

    Article  CAS  Google Scholar 

  150. Dai Y, Sha X, Song X, Zhang X, Xing M, Liu S, et al. Targeted therapy of atherosclerosis vulnerable plaque by ROS-scavenging nanoparticles and mr/fluorescence dual-modality imaging tracing. Int J Nanomed. 2022;17:5413–29.

    Article  CAS  Google Scholar 

  151. Zhu Y, Fang Y, Wang Y, Han D, Liu J, Tian L, et al. Cluster of differentiation-44-targeting prussian blue nanoparticles onloaded with colchicine for atherosclerotic plaque regression in a mice model. ACS Biomater Sci Eng. 2024;10:1530–43.

    Article  CAS  PubMed  Google Scholar 

  152. Chavez J, Khan A, Watson KR, Khan S, Si Y, Deng AY, et al. Carbon nanodots inhibit tumor necrosis factor-α-induced endothelial inflammation through scavenging hydrogen peroxide and upregulating antioxidant gene expression in EAhy926 endothelial cells. Antioxidants. 2024. https://doi.org/10.3390/antiox13020224.

    Article  PubMed  PubMed Central  Google Scholar 

  153. Basak S, Mukherjee I, Das TK. Injectable biocompatible RAFT mediated nitroxide nanogels: a robust ROS-reduction antioxidant approach. Colloids Surf, B. 2024;236: 113790.

    Article  CAS  Google Scholar 

  154. Lee S, Islam MN, Boostanpour K, Aran D, Jin G, Christenson S, et al. Molecular programs of fibrotic change in aging human lung. Nat Commun. 2021;12:6309.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  155. Schafer MJ, White TA, Iijima K, Haak AJ, Ligresti G, Atkinson EJ, et al. Cellular senescence mediates fibrotic pulmonary disease. Nat Commun. 2017;8:14532.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  156. Liu G, Beri R, Mueller A, Kamp DW. Molecular mechanisms of asbestos-induced lung epithelial cell apoptosis. Chem Biol Interact. 2010;188:309–18.

    Article  CAS  PubMed  Google Scholar 

  157. Fu Y-q, Fang F, Lu Z-y, Kuang F-w, Xu F. N-acetylcysteine protects alveolar epithelial cells from hydrogen peroxide-induced apoptosis through scavenging reactive oxygen species and suppressing c-Jun N-terminal kinase. Experim Lung Res. 2010;36:352–61.

    Article  CAS  Google Scholar 

  158. Jorgensen E, Stinson A, Shan L, Yang J, Gietl D, Albino AP. Cigarette smoke induces endoplasmic reticulum stress and the unfolded protein response in normal and malignant human lung cells. BMC Cancer. 2008;8:229.

    Article  PubMed  PubMed Central  Google Scholar 

  159. Markovic Z, Trajkovic V. Biomedical potential of the reactive oxygen species generation and quenching by fullerenes (C60). Biomaterials. 2008;29:3561–73.

    Article  CAS  PubMed  Google Scholar 

  160. Dong R, Liu M, Huang X-X, Liu Z, Jiang D-Y, Xiao H-J, et al. Water-soluble C protects against bleomycin-induced pulmonary fibrosis in mice. Int J Nanomed. 2020;15:2269–76.

    Article  CAS  Google Scholar 

  161. Zhou Y, Zhen M, Ma H, Li J, Shu C, Wang C. Inhalable gadofullerenol/[70] fullerenol as high-efficiency ROS scavengers for pulmonary fibrosis therapy. Nanomed Nanotechnol Biol Med. 2018;14:1361–9.

    Article  CAS  Google Scholar 

  162. Lv Y, Yu C, Li X, Bao H, Song S, Cao X, et al. ROS-activatable nanocomposites for CT imaging tracking and antioxidative protection of mesenchymal stem cells in idiopathic pulmonary fibrosis therapy. J Controll Release Off J Controll Release Soc. 2023;357:249–63.

    Article  CAS  Google Scholar 

  163. Liu Q, Ren Y, Jia H, Yuan H, Tong Y, Kotha S, et al. Vanadium carbide nanosheets with broad-spectrum antioxidant activity for pulmonary fibrosis therapy. ACS Nano. 2023;17:22527–38.

    Article  CAS  PubMed  Google Scholar 

  164. Tkacova R, Kluchova Z, Joppa P, Petrasova D, Molcanyiova A. Systemic inflammation and systemic oxidative stress in patients with acute exacerbations of COPD. Respir Med. 2007;101:1670–6.

    Article  PubMed  Google Scholar 

  165. MacNee WJC. Oxidants/antioxidants and COPD. Eur RespirJ. 2000;117(303S):317S.

    Google Scholar 

  166. Wiegman CH, Michaeloudes C, Haji G, Narang P, Clarke CJ, Russell KE, et al. Oxidative stress–induced mitochondrial dysfunction drives inflammation and airway smooth muscle remodeling in patients with chronic obstructive pulmonary disease. J Allergy Clin Immunol. 2015;136:769–80.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  167. Marwick JA, Kirkham PA, Stevenson CS, Danahay H, Giddings J, Butler K, et al. Cigarette smoke alters chromatin remodeling and induces proinflammatory genes in rat lungs. Am J Respir Cell Mol Biol. 2004;31:633–42.

    Article  CAS  PubMed  Google Scholar 

  168. Yang S-R, Wright J, Bauter M, Seweryniak K, Kode A, Rahman I. Sirtuin regulates cigarette smoke-induced proinflammatory mediator release via RelA/p65 NF-kappaB in macrophages in vitro and in rat lungs in vivo: implications for chronic inflammation and aging. Am J Physiol Lung Cell Mol Physiol. 2007;292:L567–76.

    Article  CAS  PubMed  Google Scholar 

  169. Rahman I, Gilmour PS, Jimenez LA, MacNee W. Oxidative stress and TNF-alpha induce histone acetylation and NF-kappaB/AP-1 activation in alveolar epithelial cells: potential mechanism in gene transcription in lung inflammation. Mol Cell Biochem. 2002;234–235:239–48.

    Article  PubMed  Google Scholar 

  170. Gao W, Li L, Wang Y, Zhang S, Adcock IM, Barnes PJ, et al. Bronchial epithelial cells: the key effector cells in the pathogenesis of chronic obstructive pulmonary disease? Respirology. 2015;20:722–9.

    Article  PubMed  Google Scholar 

  171. Takeyama K, Agustí C, Ueki I, Lausier J, Cardell LO, Nadel JA. Neutrophil-dependent goblet cell degranulation: role of membrane-bound elastase and adhesion molecules. Am J Physiol. 1998;275:L294–302.

    CAS  PubMed  Google Scholar 

  172. Paudel KR, Panth N, Manandhar B, Singh SK, Gupta G, Wich PR, et al. Attenuation of cigarette-smoke-induced oxidative stress, senescence, and inflammation by berberine-loaded liquid crystalline nanoparticles in vitro study in 16HBE and RAW264 7 cells. Antioxidants. 2022. https://doi.org/10.3390/antiox11050873.

    Article  PubMed  PubMed Central  Google Scholar 

  173. Paudel KR, Wadhwa R, Mehta M, Chellappan DK, Hansbro PM, Dua K. Rutin loaded liquid crystalline nanoparticles inhibit lipopolysaccharide induced oxidative stress and apoptosis in bronchial epithelial cells in vitro. Toxicol Vitro Int J Publ Assoc With BIBRA. 2020;68: 104961.

    Article  CAS  Google Scholar 

  174. Paudel KR, Clarence DD, Panth N, Manandhar B, De Rubis G, Devkota HP, et al. Zerumbone liquid crystalline nanoparticles protect against oxidative stress, inflammation and senescence induced by cigarette smoke extract in vitro. Naunyn Schmiedebergs Arch Pharmacol. 2024;397:2465–83.

    Article  CAS  PubMed  Google Scholar 

  175. Seidel P, Roth M. Anti-inflammatory dimethylfumarate: a potential new therapy for asthma? Med Inflamm. 2013;2013: 875403.

    Article  Google Scholar 

  176. Lounsbury N, Mateo G, Jones B, Papaiahgari S, Thimmulappa RK, Teijaro C, et al. Heterocyclic chalcone activators of nuclear factor (erythroid-derived 2)-like 2 (Nrf2) with improved in vivo efficacy. Bioorg Med Chem. 2015;23:5352–9.

    Article  CAS  PubMed  Google Scholar 

  177. Muralidharan P, Hayes D, Black SM, Mansour HM. Microparticulate/nanoparticulate powders of a novel Nrf2 activator and an aerosol performance enhancer for pulmonary delivery targeting the lung Nrf2/Keap-1 pathway. Mol Syst Design Eng. 2016;1:48–65.

    Article  CAS  Google Scholar 

  178. Chikuma K, Arima K, Asaba Y, Kubota R, Asayama S, Sato K, et al. The potential of lipid-polymer nanoparticles as epigenetic and ROS control approaches for COPD. Free Radical Res. 2020;54:829–40.

    Article  CAS  Google Scholar 

  179. Castellani S, Trapani A, Spagnoletta A, di Toma L, Magrone T, Di Gioia S, et al. Nanoparticle delivery of grape seed-derived proanthocyanidins to airway epithelial cells dampens oxidative stress and inflammation. J Transl Med. 2018;16:140.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  180. Yoo D, Guk K, Kim H, Khang G, Wu D, Lee D. Antioxidant polymeric nanoparticles as novel therapeutics for airway inflammatory diseases. Int J Pharm. 2013;450:87–94.

    Article  CAS  PubMed  Google Scholar 

  181. Deng Z, Kalin GT, Shi D, Kalinichenko VV. Nanoparticle delivery systems with cell-specific targeting for pulmonary diseases. Am J Respir Cell Mol Biol. 2021;64:292–307.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  182. Yuan XL, Meng HY, Wang YC, Peng J, Guo QY, Wang AY, et al. Bone-cartilage interface crosstalk in osteoarthritis: potential pathways and future therapeutic strategies. Osteoarthr Cartilage. 2014;22:1077–89.

    Article  CAS  Google Scholar 

  183. Chen Y, Wu X, Li J, Jiang Y, Xu K, Su J. Bone-targeted nanoparticle drug delivery system: an emerging strategy for bone-related disease. Front Pharmacol. 2022;13: 909408.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  184. Zhang L, Zheng Y-L, Wang R, Wang X-Q, Zhang H. Exercise for osteoporosis: a literature review of pathology and mechanism. Front Immunol. 2022;13:1005665.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  185. Domazetovic V, Marcucci G, Iantomasi T, Brandi ML, Vincenzini MT. Oxidative stress in bone remodeling: role of antioxidants. Clin Cases Mineral Bone Metab Off J Italian Soc Osteoporosis Mineral Metab Skeletal Dis. 2017;14:209–16.

    Article  Google Scholar 

  186. Xi Y, Pan W, Liu Y, Liu J, Xu G, Su Y, et al. α-Lipoic acid loaded hollow gold nanoparticles designed for osteoporosis treatment: preparation, characterization and in vitro evaluation. Artif Cells Nanomed Biotechnol. 2023;51:131–8.

    Article  CAS  PubMed  Google Scholar 

  187. Chen R, Liu G, Sun X, Cao X, He W, Lin X, et al. Chitosan derived nitrogen-doped carbon dots suppress osteoclastic osteolysis via downregulating ROS. Nanoscale. 2020;12:16229–44.

    Article  CAS  PubMed  Google Scholar 

  188. Li X, Lu Y, Li J, Zhou S, Wang Y, Li L, et al. Photoluminescent carbon dots (PCDs) from sour apple: a biocompatible nanomaterial for preventing UHMWPE wear-particle induced osteolysis via modulating Chemerin/ChemR23 and SIRT1 signaling pathway and its bioimaging application. J Nanobiotechnol. 2022;20:301.

    Article  Google Scholar 

  189. Pinna A, Torki Baghbaderani M, Vigil Hernández V, Naruphontjirakul P, Li S, McFarlane T, et al. Nanoceria provides antioxidant and osteogenic properties to mesoporous silica nanoparticles for osteoporosis treatment. Acta Biomater. 2021;122:365–76.

    Article  CAS  PubMed  Google Scholar 

  190. Yu P, Zheng L, Wang P, Chai S, Zhang Y, Shi T, et al. 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;165:1634–45.

    Article  CAS  PubMed  Google Scholar 

  191. Zheng L, Zhuang Z, Li Y, Shi T, Fu K, Yan W, et al. Bone targeting antioxidative nano-iron oxide for treating postmenopausal osteoporosis. Bioactive Mater. 2022;14:250–61.

    Article  CAS  Google Scholar 

  192. Chen S, Yu Y, Xie S, Liang D, Shi W, Chen S, et al. Local H2 release remodels senescence microenvironment for improved repair of injured bone. Nat Commun. 2023;14:7783.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  193. Rasool N, Negi D, Singh Y. Thiol-functionalized, antioxidant, and osteogenic mesoporous silica nanoparticles for osteoporosis. ACS Biomater Sci Eng. 2023;9:3535–45.

    Article  CAS  PubMed  Google Scholar 

  194. Dayanandan AP, Cho WJ, Kang H, Bello AB, Kim BJ, Arai Y, et al. Emerging nano-scale delivery systems for the treatment of osteoporosis. Biomater Res. 2023;27:68.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  195. Hosseinzadeh A, Kamrava SK, Joghataei MT, Darabi R, Shakeri-Zadeh A, Shahriari M, et al. Apoptosis signaling pathways in osteoarthritis and possible protective role of melatonin. J Pineal Res. 2016;61:411–25.

    Article  CAS  PubMed  Google Scholar 

  196. Wojdasiewicz P, Poniatowski ŁA, Szukiewicz D. The role of inflammatory and anti-inflammatory cytokines in the pathogenesis of osteoarthritis. Med Inflamm. 2014;2014: 561459.

    Article  Google Scholar 

  197. Ansari MY, Ahmad N, Haqqi TM. Oxidative stress and inflammation in osteoarthritis pathogenesis: role of polyphenols. Biomed Pharmacother. 2020;129: 110452.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  198. Zhong G, Yang X, Jiang X, Kumar A, Long H, Xie J, et al. Dopamine-melanin nanoparticles scavenge reactive oxygen and nitrogen species and activate autophagy for osteoarthritis therapy. Nanoscale. 2019;11:11605–16.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  199. Xue S, Zhou X, Sang W, Wang C, Lu H, Xu Y, et al. Cartilage-targeting peptide-modified dual-drug delivery nanoplatform with NIR laser response for osteoarthritis therapy. Bioactive Mater. 2021;6:2372–89.

    Article  CAS  Google Scholar 

  200. Ramírez-Noguera P, Zetina Marín I, Gómez Chavarin BM, Valderrama ME, López-Barrera LD, Díaz-Torres R. Study of the early effects of chitosan nanoparticles with glutathione in rats with osteoarthrosis. Pharmaceutics. 2023;15:2172.

    Article  PubMed  PubMed Central  Google Scholar 

  201. Liang H, Yan Y, Sun W, Ma X, Su Z, Liu Z, et al. Preparation of melatonin-loaded nanoparticles with targeting and sustained release function and their application in osteoarthritis. Int J Mol Sci. 2023;24:8740.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  202. Gao X, Ma Y, Zhang G, Tang F, Zhang J, Cao J, et al. Targeted elimination of intracellular reactive oxygen species using nanoparticle-like chitosan—superoxide dismutase conjugate for treatment of monoiodoacetate-induced osteoarthritis. Int J Pharm. 2020;590: 119947.

    Article  CAS  PubMed  Google Scholar 

  203. Zhang W, Zeng L, Yu H, He Z, Huang C, Li C, et al. Injectable spontaneous hydrogen-releasing hydrogel for long-lasting alleviation of osteoarthritis. Acta Biomater. 2023;158:163–77.

    Article  CAS  PubMed  Google Scholar 

  204. Zhao W, Yu Y, Zhang Z, He D, Zhang H. Bioinspired nanospheres as anti-inflammation and antisenescence interfacial biolubricant for treating temporomandibular joint osteoarthritis. ACS Appl Mater Interf. 2022;14:35409–22.

    Article  CAS  Google Scholar 

  205. Chen L, Tiwari SR, Zhang Y, Zhang J, Sun Y. Facile synthesis of hollow MnO nanoparticles for reactive oxygen species scavenging in osteoarthritis. ACS Biomater Sci Eng. 2021;7:1686–92.

    Article  CAS  PubMed  Google Scholar 

  206. Pei Y, Cui F, Du X, Shang G, Xiao W, Yang X, et al. Antioxidative nanofullerol inhibits macrophage activation and development of osteoarthritis in rats. Int J Nanomed. 2019;14:4145–55.

    Article  CAS  Google Scholar 

  207. Liang R, Zhao J, Li B, Cai P, Loh XJ, Xu C, et al. Implantable and degradable antioxidant poly(ε-caprolactone)-lignin nanofiber membrane for effective osteoarthritis treatment. Biomaterials. 2020;230: 119601.

    Article  CAS  PubMed  Google Scholar 

  208. Chen H, Ye T, Hu F, Chen K, Li B, Qiu M, et al. Urchin-like ceria nanoparticles for enhanced gene therapy of osteoarthritis. Sci Adv. 2023;9:0988.

    Google Scholar 

  209. Jackson MJ. Reactive oxygen species in sarcopenia: Should we focus on excess oxidative damage or defective redox signalling? Mol Aspects Med. 2016;50:33–40.

    Article  CAS  PubMed  Google Scholar 

  210. Sakellariou GK, Pearson T, Lightfoot AP, Nye GA, Wells N, Giakoumaki II, et al. Mitochondrial ROS regulate oxidative damage and mitophagy but not age-related muscle fiber atrophy. Sci Rep. 2016;6:33944.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  211. Deldicque L. Endoplasmic reticulum stress in human skeletal muscle: any contribution to sarcopenia? Front Physiol. 2013;4:236.

    Article  PubMed  PubMed Central  Google Scholar 

  212. Powers SK. Can antioxidants protect against disuse muscle atrophy? Sports Med. 2014;44(Suppl 2):S155–65.

    Article  PubMed  Google Scholar 

  213. Liang Y-J, Yang IH, Lin Y-W, Lin J-N, Wu C-C, Chiang C-Y, et al. Curcumin-loaded hydrophobic surface-modified hydroxyapatite as an antioxidant for sarcopenia prevention. Antioxidants. 2016;10:616.

    Article  Google Scholar 

  214. Papaccio F, Arino D. Focus on the contribution of oxidative stress in skin aging. Antioxidants. 2022;11:1121.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  215. Rinnerthaler M, Bischof J, Streubel MK, Trost A, Richter K. Oxidative stress in aging human skin. Biomolecules. 2015;5:545–89.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  216. Bell S, Degitz K, Quirling M, Jilg N, Page S, Brand K. Involvement of NF-κB signalling in skin physiology and disease. Cell Signal. 2003;15:1–7.

    Article  CAS  PubMed  Google Scholar 

  217. Wang Y, Wang L, Wen X, Hao D, Zhang N, He G, et al. NF-κB signaling in skin aging. Mech Ageing Dev. 2019;184: 111160.

    Article  CAS  PubMed  Google Scholar 

  218. Hall M-C, Young DA, Waters JG, Rowan AD, Chantry A, Edwards DR, et al. The comparative role of activator protein 1 and smad factors in the regulation of <em>Timp-1</em> and <em>MMP-1</em> gene expression by transforming growth factor-β1 *. J Biol Chem. 2003;278:10304–13.

    Article  CAS  PubMed  Google Scholar 

  219. Kim J-A, Ahn B-N, Kong C-S, Kim S-K. Chitooligomers inhibit UV-A-induced photoaging of skin by regulating TGF-β/Smad signaling cascade. Carbohyd Polym. 2012;88:490–5.

    Article  CAS  Google Scholar 

  220. Avadhani KS, Manikkath J, Tiwari M, Chandrasekhar M, Godavarthi A, Vidya SM, et al. Skin delivery of epigallocatechin-3-gallate (EGCG) and hyaluronic acid loaded nano-transfersomes for antioxidant and anti-aging effects in UV radiation induced skin damage. Drug Delivery. 2017;24:61–74.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  221. Atrux-Tallau N, Lasselin J, Han S-H, Delmas T, Bibette J. Quantitative analysis of ligand effects on bioefficacy of nanoemulsion encapsulating depigmenting active. Colloids Surf, B. 2014;122:390–5.

    Article  CAS  Google Scholar 

  222. Lohan SB, Bauersachs S, Ahlberg S, Baisaeng N, Keck CM, Müller RH, et al. Ultra-small lipid nanoparticles promote the penetration of coenzyme Q10 in skin cells and counteract oxidative stress. Eur J Pharmaceutics Biopharmaceutics. 2015;89:201–7.

    Article  CAS  Google Scholar 

  223. Rathod S, Arya S, Kanike S, Shah SA, Bahadur P, Tiwari S. Advances on nanoformulation approaches for delivering plant-derived antioxidants: a case of quercetin. Int J Pharm. 2022;625: 122093.

    Article  CAS  PubMed  Google Scholar 

  224. Kim AY, Ha JH, Park SN. Selective release system for antioxidative and anti-inflammatory activities using HO-responsive therapeutic nanoparticles. Biomacromol. 2017;18:3197–206.

    Article  CAS  Google Scholar 

  225. Nisar MF, Yousaf M, Saleem M, Khalid H, Niaz K, Yaqub M, et al. Development of iron sequester antioxidant quercetin@ZnO nanoparticles with photoprotective effects on UVA-Irradiated HaCaT cells. Oxid Med Cell Longev. 2021;2021:6072631.

    Article  PubMed  PubMed Central  Google Scholar 

  226. Ngan CL, Basri M, Tripathy M, Abedi Karjiban R, Abdul-Malek E. Skin intervention of fullerene-integrated nanoemulsion in structural and collagen regeneration against skin aging. Eur J Pharmaceutical Sci Off J Eur Federation Pharmaceutical Sci. 2015;70:22–8.

    CAS  Google Scholar 

  227. Xiao L, Takada H, Maeda K, Haramoto M, Miwa N. Antioxidant effects of water-soluble fullerene derivatives against ultraviolet ray or peroxylipid through their action of scavenging the reactive oxygen species in human skin keratinocytes. Biomed Pharmacother Biomed Pharm. 2005;59:351–8.

    Article  CAS  Google Scholar 

  228. Li Y, Hou X, Yang C, Pang Y, Li X, Jiang G, et al. Photoprotection of cerium oxide nanoparticles against UVA radiation-induced senescence of human skin fibroblasts due to their antioxidant properties. Sci Rep. 2019;9:2595.

    Article  PubMed  PubMed Central  Google Scholar 

  229. Peloi KE, Contreras Lancheros CA, Nakamura CV, Singh S, Neal C, Sakthivel TS, et al. Antioxidative photochemoprotector effects of cerium oxide nanoparticles on UVB irradiated fibroblast cells. Colloids Surf B. 2020;191: 111013.

    Article  CAS  Google Scholar 

  230. Ribeiro FM, de Oliveira MM, Singh S, Sakthivel TS, Neal CJ, Seal S, et al. Ceria nanoparticles decrease UVA-induced fibroblast death through cell redox regulation leading to cell survival, migration and proliferation. Front Bioeng Biotechnol. 2020. https://doi.org/10.3389/fbioe.2020.577557.

    Article  PubMed  PubMed Central  Google Scholar 

  231. Xiong B, Xu R, Zhou R, He Y, Yeung ES. Preventing UV induced cell damage by scavenging reactive oxygen species with enzyme-mimic Au-Pt nanocomposites. Talanta. 2014;120:262–7.

    Article  CAS  PubMed  Google Scholar 

  232. Qu A, Chen Q, Sun M, Xu L, Hao C, Xu C, et al. Sensitive and selective dual-mode responses to reactive oxygen species by chiral manganese dioxide nanoparticles for antiaging skin. Adv Mater. 2024;36: e2308469.

    Article  PubMed  Google Scholar 

  233. Feliciano CP, Nagasaki Y. Oral nanotherapeutics: redox nanoparticles attenuate ultraviolet B radiation-induced skin inflammatory disorders in Kud:Hr- hairless mice. Biomaterials. 2017;142:162–70.

    Article  CAS  PubMed  Google Scholar 

  234. Bhatia E, Kumari D, Sharma S, Ahamad N, Banerjee R. Nanoparticle platforms for dermal antiaging technologies: insights in cellular and molecular mechanisms. Wiley Interdiscipl Rev Nanomed Nanobiotechnol. 2022;14: e1746.

    Article  Google Scholar 

  235. Wohlfart S, Gelperina S, Kreuter J. Transport of drugs across the blood-brain barrier by nanoparticles. J Controlled Release Off J Controll Release Soc. 2012;161:264–73.

    Article  CAS  Google Scholar 

  236. Shakeri S, Ashrafizadeh M, Zarrabi A, Roghanian R, Afshar EG, Pardakhty A, et al. Multifunctional polymeric nanoplatforms for brain diseases diagnosis. Ther Ther Biomed. 2020. https://doi.org/10.3390/biomedicines8010013.

    Article  Google Scholar 

  237. Sim TM, Tarini D, Dheen ST, Bay BH, Srinivasan DK. Nanoparticle-based technology approaches to the management of neurological disorders. Int J Mol Sci. 2020. https://doi.org/10.3390/ijms21176070.

    Article  PubMed  PubMed Central  Google Scholar 

  238. Martinelli C, Pucci C, Battaglini M, Marino A, Ciofani G. Antioxidants and nanotechnology: promises and limits of potentially disruptive approaches in the treatment of central nervous system diseases. Adv Healthcare Mater. 2020;9: e1901589.

    Article  Google Scholar 

  239. Lin MT, Beal MF. Mitochondrial dysfunction and oxidative stress in neurodegenerative diseases. Nature. 2006;443:787–95.

    Article  CAS  PubMed  Google Scholar 

  240. Praticò D. Oxidative stress hypothesis in Alzheimer’s disease: a reappraisal. Trends Pharmacol Sci. 2008;29:609–15.

    Article  PubMed  Google Scholar 

  241. Elfawy HA, Das B. Crosstalk between mitochondrial dysfunction, oxidative stress, and age related neurodegenerative disease: etiologies and therapeutic strategies. Life Sci. 2019;218:165–84.

    Article  CAS  PubMed  Google Scholar 

  242. Zheng C, Zhou X-W, Wang J-Z. The dual roles of cytokines in Alzheimer’s disease: update on interleukins, TNF-α TGF-β and IFN-γ. Transl Neurodegenerat. 2016;5:7.

    Article  Google Scholar 

  243. Winiarska-Mieczan A, Baranowska-Wójcik E, Kwiecień M, Grela ER, Szwajgier D, Kwiatkowska K, et al. The role of dietary antioxidants in the pathogenesis of neurodegenerative diseases and their impact on cerebral oxidoreductive balance. Nutrients. 2020;12:435.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  244. Halevas E, Mavroidi B, Nday CM, Tang J, Smith GC, Boukos N, et al. Modified magnetic core-shell mesoporous silica nano-formulations with encapsulated quercetin exhibit anti-amyloid and antioxidant activity. J Inorg Biochem. 2020;213: 111271.

    Article  CAS  PubMed  Google Scholar 

  245. Chonpathompikunlert P, Yoshitomi T, Han J, Toh K, Isoda H, Nagasaki Y. Chemical nanotherapy: nitroxyl radical-containing nanoparticle protects neuroblastoma SH-SY5Y cells from Abeta-induced oxidative stress. Ther Deliv. 2011;2:585–97.

    Article  CAS  PubMed  Google Scholar 

  246. Chonpathompikunlert P, Yoshitomi T, Vong LB, Imaizumi N, Ozaki Y, Nagasaki Y. Recovery of cognitive dysfunction via orally administered redox-polymer nanotherapeutics in SAMP8 mice. PLoS ONE. 2015;10: e0126013.

    Article  PubMed  PubMed Central  Google Scholar 

  247. Gao F, Zhao J, Liu P, Ji D, Zhang L, Zhang M, et al. Preparation and in vitro evaluation of multi-target-directed selenium-chondroitin sulfate nanoparticles in protecting against the Alzheimer’s disease. Int J Biol Macromol. 2020;142:265–76.

    Article  CAS  PubMed  Google Scholar 

  248. Jia Z, Yuan X, Wei J-A, Guo X, Gong Y, Li J, et al. A functionalized octahedral palladium nanozyme as a radical scavenger for ameliorating Alzheimer’s disease. ACS Appl Mater Interf. 2021;13:49602–13.

    Article  CAS  Google Scholar 

  249. Li C, Wang N, Zheng G, Yang L. Oral administration of resveratrol-selenium-peptide nanocomposites alleviates alzheimer’s disease-like pathogenesis by inhibiting aβ aggregation and regulating gut microbiota. ACS Appl Mater Interfaces. 2021;13:46406–20.

    Article  CAS  PubMed  Google Scholar 

  250. Liu P, Zhang T, Chen Q, Li C, Chu Y, Guo Q, et al. Biomimetic dendrimer-peptide conjugates for early multi-target therapy of alzheimer’s disease by inflammatory microenvironment modulation. Adv Mater (Deerfield Beach Fla). 2021;33: e2100746.

    Article  Google Scholar 

  251. Ma M, Gao N, Sun Y, Du X, Ren J, Qu X. Redox-activated near-infrared-responsive polyoxometalates used for photothermal treatment of Alzheimer’s disease. Adv Healthcare Mater. 2018;7: e1800320.

    Article  Google Scholar 

  252. Yang L, Yin T, Liu Y, Sun J, Zhou Y, Liu J. Gold nanoparticle-capped mesoporous silica-based HO-responsive controlled release system for Alzheimer’s disease treatment. Acta Biomater. 2016;46:177–90.

    Article  CAS  PubMed  Google Scholar 

  253. Muller AP, Ferreira GK, Pires AJ, de Bem Silveira G, de Souza DL, Brandolfi JdA, et al. Gold nanoparticles prevent cognitive deficits, oxidative stress and inflammation in a rat model of sporadic dementia of Alzheimer’s type. Mater Sci Eng Mater Biol Appl. 2017;77(4):76–483.

    Google Scholar 

  254. Zhang J, Liu R, Zhang D, Zhang Z, Zhu J, Xu L, et al. Neuroprotective effects of maize tetrapeptide-anchored gold nanoparticles in Alzheimer’s disease. Colloids Surf, B. 2021;200: 111584.

    Article  CAS  Google Scholar 

  255. Zhang Y, Wang ZY, Li XJ, Wang L, Yin M, Wang LH, et al. Dietary iron oxide nanoparticles delay aging and ameliorate neurodegeneration in Drosophila. Adv Mater. 2016;28:1387–93.

    Article  CAS  PubMed  Google Scholar 

  256. Zhao Y, Xu Q, Xu W, Wang D, Tan J, Zhu C, et al. Probing the molecular mechanism of cerium oxide nanoparticles in protecting against the neuronal cytotoxicity of Aβ1-42 with copper ions. Metall Integr Biometal Sci. 2016;8:644–7.

    Article  CAS  Google Scholar 

  257. Yu DQ, Ma MM, Liu ZW, Pi ZF, Du XB, Ren JS, et al. MOF-encapsulated nanozyme enhanced siRNA combo: Control neural stem cell differentiation and ameliorate cognitive impairments in Alzheimer’s disease model. Biomaterials. 2020;255:120160.

    Article  CAS  PubMed  Google Scholar 

  258. Ge K, Mu Y, Liu M, Bai Z, Liu Z, Geng D, et al. Gold Nanorods with spatial separation of CeO deposition for plasmonic-enhanced antioxidant stress and photothermal therapy of Alzheimer’s disease. ACS Appl Mater Interf. 2022;14:3662–74.

    Article  CAS  Google Scholar 

  259. Zhong G, Long H, Zhou T, Liu Y, Zhao J, Han J, et al. Blood-brain barrier Permeable nanoparticles for Alzheimer’s disease treatment by selective mitophagy of microglia. Biomaterials. 2022;288: 121690.

    Article  CAS  PubMed  Google Scholar 

  260. Zhao C, Chen J, Ye J, Li Z, Su L, Wang J, et al. Structural transformative antioxidants for dual-responsive anti-inflammatory delivery and photoacoustic inflammation imaging. Angew Chem Int Ed Engl. 2021;60:14458–66.

    Article  CAS  PubMed  Google Scholar 

  261. Frozza RL, Bernardi A, Hoppe JB, Meneghetti AB, Matté A, Battastini AMO, et al. Neuroprotective effects of resveratrol against Aβ administration in rats are improved by lipid-core nanocapsules. Mol Neurobiol. 2013;47:1066–80.

    Article  CAS  PubMed  Google Scholar 

  262. Lu X, Ji C, Xu H, Li X, Ding H, Ye M, et al. Resveratrol-loaded polymeric micelles protect cells from Aβ-induced oxidative stress. Int J Pharm. 2009;375:89–96.

    Article  CAS  PubMed  Google Scholar 

  263. Han Y, Chu X, Cui L, Fu S, Gao C, Li Y, et al. Neuronal mitochondria-targeted therapy for Alzheimer’s disease by systemic delivery of resveratrol using dual-modified novel biomimetic nanosystems. Drug Delivery. 2020;27:502–18.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  264. Fan S, Zheng Y, Liu X, Fang W, Chen X, Liao W, et al. Curcumin-loaded PLGA-PEG nanoparticles conjugated with B6 peptide for potential use in Alzheimer’s disease. Drug Delivery. 2018;25:1091–102.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  265. Tiwari SK, Agarwal S, Seth B, Yadav A, Nair S, Bhatnagar P, et al. Curcumin-loaded nanoparticles potently induce adult neurogenesis and reverse cognitive deficits in Alzheimer’s disease model via canonical Wnt/β-catenin pathway. ACS Nano. 2014;8:76–103.

    Article  CAS  PubMed  Google Scholar 

  266. Doggui S, Sahni JK, Arseneault M, Dao L, Ramassamy C. Neuronal uptake and neuroprotective effect of curcumin-loaded PLGA nanoparticles on the human SK-N-SH cell line. J Alzheimers Dis. 2012;30:377–92.

    Article  CAS  PubMed  Google Scholar 

  267. Sadegh Malvajerd S, Izadi Z, Azadi A, Kurd M, Derakhshankhah H, Sharifzadeh M, et al. Neuroprotective potential of curcumin-loaded nanostructured lipid carrier in an animal model of Alzheimer’s disease: behavioral and biochemical evidence. J Alzheimer’s Dis JAD. 2019;69:671–86.

    Article  CAS  PubMed  Google Scholar 

  268. Ray B, Bisht S, Maitra A, Maitra A, Lahiri DK. Neuroprotective and neurorescue effects of a novel polymeric nanoparticle formulation of curcumin (NanoCurc™) in the neuronal cell culture and animal model: implications for Alzheimer’s disease. J Alzheimer’s Disease JAD. 2011;23:61–77.

    Article  CAS  PubMed  Google Scholar 

  269. Pan Q, Ban Y, Xu L. Silibinin-albumin nanoparticles: characterization and biological evaluation against oxidative stress-stimulated neurotoxicity associated with Alzheimer’s disease. J Biomed Nanotechnol. 2021;17:1123–30.

    Article  CAS  PubMed  Google Scholar 

  270. Sathya S, Shanmuganathan B, Devi KP. Deciphering the anti-apoptotic potential of α-bisabolol loaded solid lipid nanoparticles against Aβ induced neurotoxicity in Neuro-2a cells. Colloids Surf B. 2020;190: 110948.

    Article  CAS  Google Scholar 

  271. Singh A, Ujjwal RR, Naqvi S, Verma RK, Tiwari S, Kesharwani P, et al. Formulation development of tocopherol polyethylene glycol nanoengineered polyamidoamine dendrimer for neuroprotection and treatment of Alzheimer disease. J Drug Target. 2022;30:777–91.

    Article  CAS  PubMed  Google Scholar 

  272. Prakashkumar N, Sivamaruthi BS, Chaiyasut C, Suganthy N. Decoding the neuroprotective potential of methyl gallate-loaded starch nanoparticles against beta amyloid-induced oxidative stress-mediated apoptosis an in vitro study. Pharmaceutics. 2021;13:299.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  273. de Farias CC, Maes M, Bonifácio KL, Bortolasci CC, de Souza NA, Brinholi FF, et al. Highly specific changes in antioxidant levels and lipid peroxidation in Parkinson’s disease and its progression: disease and staging biomarkers and new drug targets. Neurosci Lett. 2016;617:66–71.

    Article  PubMed  Google Scholar 

  274. Dionísio PA, Amaral JD, Rodrigues CMP. Oxidative stress and regulated cell death in Parkinson’s disease. Ageing Res Rev. 2021;67: 101263.

    Article  PubMed  Google Scholar 

  275. Tenkorang MAA, Duong P, Cunningham RL. NADPH oxidase mediates membrane androgen receptor-induced neurodegeneration. Endocrinology. 2019;160:947–63.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  276. Musgrove RE, Helwig M, Bae E-J, Aboutalebi H, Lee S-J, Ulusoy A, et al. Oxidative stress in vagal neurons promotes parkinsonian pathology and intercellular α-synuclein transfer. J Clin Investig. 2019;129:3738–53.

    Article  PubMed  PubMed Central  Google Scholar 

  277. Luo S, Kang SS, Wang Z-H, Liu X, Day JX, Wu Z, et al. Akt phosphorylates NQO1 and triggers its degradation, abolishing its antioxidative activities in Parkinson’s disease. J Neurosci Off J Soc Neurosci. 2019;39:7291–305.

    Article  CAS  Google Scholar 

  278. Sharma N, Rao SP, Kalivendi SV. The deglycase activity of DJ-1 mitigates α-synuclein glycation and aggregation in dopaminergic cells: role of oxidative stress mediated downregulation of DJ-1 in Parkinson’s disease. Free Radical Biol Med. 2019;135:28–37.

    Article  CAS  Google Scholar 

  279. Russo I, Kaganovich A, Ding J, Landeck N, Mamais A, Varanita T, et al. Transcriptome analysis of LRRK2 knock-out microglia cells reveals alterations of inflammatory- and oxidative stress-related pathways upon treatment with α-synuclein fibrils. Neurobiol Dis. 2019;129:67–78.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  280. Kurosaki H, Yamaguchi K, Man-Yoshi K, Muramatsu S-I, Hara S, Ichinose H. Administration of tetrahydrobiopterin restored the decline of dopamine in the striatum induced by an acute action of MPTP. Neurochem Int. 2019;125:16–24.

    Article  CAS  PubMed  Google Scholar 

  281. Bencsik A, Lestaevel P, Guseva CI. Nano- and neurotoxicology: an emerging discipline. Prog Neurobiol. 2018;160:45–63.

    Article  CAS  PubMed  Google Scholar 

  282. Hao C, Qu A, Xu L, Sun M, Zhang H, Xu C, et al. Chiral molecule-mediated porous Cu O nanoparticle clusters with antioxidation activity for ameliorating Parkinson’s disease. J Am Chem Soc. 2019;141:1091–9.

    Article  CAS  PubMed  Google Scholar 

  283. Maghsoudi A, Fakharzadeh S, Hafizi M, Abbasi M, Kohram F, Sardab S, et al. Neuroprotective effects of three different sizes nanochelating based nano complexes in MPP(+) induced neurotoxicity. Apoptosis Int J Programmed Cell Death. 2015;20:298–309.

    Article  CAS  Google Scholar 

  284. Pichla M, Pulaski Ł, Kania KD, Stefaniuk I, Cieniek B, Pieńkowska N, et al. Nitroxide radical-containing redox nanoparticles protect neuroblastoma SH-SY5Y cells against 6-hydroxydopamine toxicity. Oxid Med Cell Longev. 2020;2020:9260748.

    Article  PubMed  PubMed Central  Google Scholar 

  285. Vernekar AA, Sinha D, Srivastava S, Paramasivam PU, D’Silva P, Mugesh G. An antioxidant nanozyme that uncovers the cytoprotective potential of vanadia nanowires. Nature Commun. 2014;5(1):5301.

    Article  CAS  Google Scholar 

  286. Liu Y-Q, Mao Y, Xu E, Jia H, Zhang S, Dawson VL, et al. Nanozyme scavenging ROS for prevention of pathologic α-synuclein transmission in Parkinson’s disease. Nano Today. 2021;36: 101027.

    Article  CAS  Google Scholar 

  287. Dugan LL, Lovett EG, Quick KL, Lotharius J, Lin TT, O’Malley KL. Fullerene-based antioxidants and neurodegenerative disorders. Parkinsonism Relat Disord. 2001;7:243–6.

    Article  PubMed  Google Scholar 

  288. Dugan LL, Tian L, Quick KL, Hardt JI, Karimi M, Brown C, et al. Carboxyfullerene neuroprotection postinjury in Parkinsonian nonhuman primates. Ann Neurol. 2014;76:393–402.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  289. Umarao P, Bose S, Bhattacharyya S, Kumar A, Jain S. Neuroprotective potential of superparamagnetic iron oxide nanoparticles along with exposure to electromagnetic field in 6-OHDA rat model of Parkinson’s Disease. J Nanosci Nanotechnol. 2016;16:261–9.

    Article  CAS  PubMed  Google Scholar 

  290. Ruotolo R, De Giorgio G, Minato I, Bianchi MG, Bussolati O, Marmiroli N. Cerium oxide nanoparticles rescue α-synuclein-induced toxicity in a yeast model of Parkinson’s Disease. Nanomaterials. 2020. https://doi.org/10.3390/nano10020235.

    Article  PubMed  PubMed Central  Google Scholar 

  291. Kwon HJ, Kim D, Seo K, Kim YG, Han SI, Kang T, et al. Ceria nanoparticle systems for selective scavenging of mitochondrial, intracellular, and extracellular reactive oxygen species in Parkinson’s Disease. Angew Chem Int Ed Engl. 2018;57:9408–12.

    Article  CAS  PubMed  Google Scholar 

  292. Dillon CE, Billings M, Hockey KS, DeLaGarza L, Rzigalinski BA: Cerium Oxide Nanoparticles Protect Against MPTP-Induced Dopaminergic Neurodegeneration In A Mouse Model For Parkinson's Disease. In NSTI Nanotechnology Conference and Expo; Jun 13–16; Boston, MA. 2011: 451–454.

  293. Hegazy MA, Maklad HM, Samy DM, Abdelmonsif DA, El Sabaa BM, Elnozahy FY. Cerium oxide nanoparticles could ameliorate behavioral and neurochemical impairments in 6-hydroxydopamine induced Parkinson’s disease in rats. Neurochem Int. 2017;108:361–71.

    Article  CAS  PubMed  Google Scholar 

  294. Li Y, Li Y, Wang H, Liu R. Yb, Er codoped cerium oxide upconversion nanoparticles enhanced the enzymelike catalytic activity and antioxidative activity for parkinson’s disease treatment. ACS Appl Mater Interf. 2021;13:13968–77.

    Article  CAS  Google Scholar 

  295. Singh N, Savanur MA, Srivastava S, D’Silva P, Mugesh G. A redox modulatory Mn O nanozyme with multi-enzyme activity provides efficient cytoprotection to human cells in a Parkinson’s disease model. Angew Chem Int Ed Engl. 2017;56:14267–71.

    Article  CAS  PubMed  Google Scholar 

  296. Xu Z, Qu A, Wang W, Lu M, Shi B, Chen C, et al. Facet-dependent biodegradable Mn O nanoparticles for ameliorating parkinson’s disease. Adv Healthcare Mater. 2021;10: e2101316.

    Article  Google Scholar 

  297. Nanjwade BK, Kadam VT, Manvi FV. Formulation and characterization of nanostructured lipid carrier of ubiquinone (Coenzyme Q10). J Biomed Nanotechnol. 2013;9:450–60.

    Article  CAS  PubMed  Google Scholar 

  298. Haney MJ, Klyachko NL, Zhao Y, Gupta R, Plotnikova EG, He Z, et al. Exosomes as drug delivery vehicles for Parkinson’s disease therapy. J Control Release. 2015;207:18–30.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  299. da Rocha LG, Bonfanti Santos D, Colle D, Gasnhar Moreira EL, Daniel Prediger R, Farina M, et al. Improved neuroprotective effects of resveratrol-loaded polysorbate 80-coated poly(lactide) nanoparticles in MPTP-induced Parkinsonism. Nanomedicine. 2015;10:1127–38.

    Article  Google Scholar 

  300. Chiu S, Terpstra KJ, Bureau Y, Hou J, Raheb H, Cernvosky Z, et al. Liposomal-formulated curcumin [Lipocurc™] targeting HDAC (Histone Deacetylase) prevents apoptosis and improves motor deficits in Park 7 (DJ-1)-knockout rat model of Parkinson’s disease: implications for epigenetics-based nanotechnology-driven drug platform. J Complement Integr Med. 2013;10:75–88.

    Article  CAS  Google Scholar 

  301. Bollimpelli VS, Kumar P, Kumari S, Kondapi AK. Neuroprotective effect of curcumin-loaded lactoferrin nano particles against rotenone induced neurotoxicity. Neurochem Int. 2016;95:37–45.

    Article  CAS  PubMed  Google Scholar 

  302. Singhal NK, Agarwal S, Bhatnagar P, Tiwari MN, Tiwari SK, Srivastava G, et al. Mechanism of nanotization-mediated improvement in the efficacy of caffeine against 1-Methyl-4-Phenyl-1,2,3,6-tetrahydropyridine-induced Parkinsonism. J Biomed Nanotechnol. 2015;11:2211–22.

    Article  CAS  PubMed  Google Scholar 

  303. Liu H, Han Y, Wang T, Zhang H, Xu Q, Yuan J, et al. Targeting microglia for therapy of parkinson’s disease by using biomimetic ultrasmall nanoparticles. J Am Chem Soc. 2020;142:21730–42.

    Article  CAS  PubMed  Google Scholar 

  304. Liu Y, Gao J. Reproductive aging: biological pathways and potential interventive strategies. J Genet Genom Yi Chuan Xue Bao. 2022. https://doi.org/10.1016/j.jgg.2022.07.002.

    Article  Google Scholar 

  305. Lu J, Wang Z, Cao J, Chen Y, Dong Y. A novel and compact review on the role of oxidative stress in female reproduction. Reprod Biol Endocrinol RB&E. 2018;16:80.

    Article  Google Scholar 

  306. Sasaki H, Hamatani T, Kamijo S, Iwai M, Kobanawa M, Ogawa S, et al. Impact of oxidative stress on age-associated decline in oocyte developmental competence. Front Endocrinol. 2019;10:811.

    Article  Google Scholar 

  307. Agarwal A, Aponte-Mellado A, Premkumar BJ, Shaman A, Gupta S. The effects of oxidative stress on female reproduction: a review. Reprod Biol Endocrinol RB&E. 2012;10:49.

    Article  Google Scholar 

  308. Dong L, Teh DBL, Kennedy BK, Huang Z. Unraveling female reproductive senescence to enhance healthy longevity. Cell Res. 2023;33:11–29.

    Article  PubMed  PubMed Central  Google Scholar 

  309. Showell MG, Mackenzie-Proctor R, Jordan V, Hart RJ. Antioxidants for female subfertility. Cochrane Database Syst Rev. 2020;8:CD007807.

    PubMed  Google Scholar 

  310. Bjelakovic G, Nikolova D, Gluud LL, Simonetti RG. Gluud CJCdosr Antioxidant supplements for prevention of mortality in healthy participants and patients with various diseases. Cochrane Database Syst Rev. 2012. https://doi.org/10.1002/14651858.CD007176.pub2.

    Article  PubMed  PubMed Central  Google Scholar 

  311. Ochiai A, Kuroda K, Ikemoto Y, Ozaki R, Nakagawa K, Nojiri S, et al. Influence of resveratrol supplementation on IVF–embryo transfer cycle outcomes. Reprod BioMed Online. 2019;39:205–10.

    Article  CAS  PubMed  Google Scholar 

  312. Agarwal A, Mulgund A, Hamada A, Chyatte MR. A unique view on male infertility around the globe. Reprod Biol Endocrinol RB&E. 2015;13:37.

    Article  Google Scholar 

  313. Deepinder F, Chowdary HT, Agarwal A. Role of metabolomic analysis of biomarkers in the management of male infertility. Expert Rev Mol Diagn. 2007;7:351–8.

    Article  CAS  PubMed  Google Scholar 

  314. Zini A, Libman J. Sperm DNA damage: importance in the era of assisted reproduction. Current Opin Urol. 2006. https://doi.org/10.1097/01.mou.0000250283.75484.dd.

    Article  Google Scholar 

  315. Leisegang K, Henkel R, Agarwal AJCPD. Redox regulation of fertility in aging male and the role of antioxidants: a savior or stressor. Current Pharm Design. 2017;23(4438):4450.

    Google Scholar 

  316. Agarwal A, Parekh N, Panner Selvam MK, Henkel R, Shah R, Homa ST, et al. Male oxidative stress infertility (MOSI): proposed terminology and clinical practice guidelines for management of idiopathic male infertility. World J Men’s Health. 2019;37:296–312.

    Article  Google Scholar 

  317. El-Megharbel SM, Almalki AS, Hamza RZ, Gobouri AA, Alhadhrami AA, Al-Humaidi JY, et al. Synthesis and suggestion of a new nanometric gold(III) melatonin drug complex: an interesting model for testicular protection. Future Med Chem. 2018;10:1693–704.

    Article  CAS  PubMed  Google Scholar 

  318. Zhang X, Gan X, et al. Ameliorative effects of nano-selenium against NiSO-induced apoptosis in rat testes. Toxicol Mechan Methods. 2019;29:467–77.

    Article  CAS  Google Scholar 

  319. Khalaf AA, Ahmed W, Moselhy WA, Abdel-Halim BR, Ibrahim MA. Protective effects of selenium and nano-selenium on bisphenol-induced reproductive toxicity in male rats. Hum Exp Toxicol. 2019;38:398–408.

    Article  CAS  PubMed  Google Scholar 

  320. Snow-Lisy DC, Sabanegh ES, Samplaski MK, Morris VB, Labhasetwar V. Superoxide dismutase-loaded biodegradable nanoparticles targeted with a follicle-stimulating hormone peptide protect Sertoli cells from oxidative stress. Fertil Steril. 2014;101:560–7.

    Article  CAS  PubMed  Google Scholar 

  321. Ma J, Dong SH, Lu HT, Chen ZM, Yu HJ, Sun XJ, et al. The hydrogen storage nanomaterial MgH2 improves irradiation-induced male fertility impairment by suppressing oxidative stress. Biomater Res. 2022. https://doi.org/10.1186/s40824-022-00266-6.

    Article  PubMed  PubMed Central  Google Scholar 

  322. Saleh H, Nassar AMK, Noreldin AE, Samak D, Elshony N, Wasef L, et al. Chemo-protective potential of cerium oxide nanoparticles against fipronil-induced oxidative stress apoptosis inflammation and reproductive dysfunction in male white albino rats. Molecules. 2020. https://doi.org/10.3390/molecules25153479.

    Article  PubMed  PubMed Central  Google Scholar 

  323. Mirkov SM, Djordjevic AN, Andric NL, Andric SA, Kostic TS, Bogdanovic GM, et al. Nitric oxide-scavenging activity of polyhydroxylated fullerenol, C60(OH)24. Nitric Oxide Biol Chem. 2004;11:201–7.

    Article  CAS  Google Scholar 

  324. Henkel R, Sandhu IS, Agarwal A. The excessive use of antioxidant therapy: a possible cause of male infertility? Andrologia. 2019;51: e13162.

    Article  PubMed  Google Scholar 

  325. Panner Selvam MK, Henkel R, Sharma R, Agarwal A. Calibration of redox potential in sperm wash media and evaluation of oxidation-reduction potential values in various assisted reproductive technology culture media using MiOXSYS system. Andrology. 2018;6:293–300.

    Article  CAS  PubMed  Google Scholar 

  326. Kaarniranta K, Salminen A, Haapasalo A, Soininen H, Hiltunen M. Age-related macular degeneration (AMD): Alzheimer’s disease in the eye? J Alzheimer’s Dis JAD. 2011;24:615–31.

    Article  CAS  PubMed  Google Scholar 

  327. Fleckenstein M, Keenan TDL, Guymer RH, Chakravarthy U, Schmitz-Valckenberg S, Klaver CC, et al. Age-related macular degeneration. Nat Rev Dis Primers. 2021;7:31.

    Article  PubMed  Google Scholar 

  328. Ozawa Y. Oxidative stress in the light-exposed retina and its implication in age-related macular degeneration. Redox Biol. 2020;37: 101779.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  329. Jiang PF, Choi A, Swindle-Reilly KE. Controlled release of anti-VEGF by redox-responsive polydopamine nanoparticles. Nanoscale. 2020;12:17298–311.

    Article  CAS  PubMed  Google Scholar 

  330. Yang C, Fischer M, Kirby C, Liu R, Zhu H, Zhang H, et al. Bioaccessibility, cellular uptake and transport of luteins and assessment of their antioxidant activities. Food Chem. 2018;249:66–76.

    Article  CAS  PubMed  Google Scholar 

  331. Sun R, Zhang A, Ge Y, Gou J, Yin T, He H, et al. Ultra-small-size astragaloside-IV loaded lipid nanocapsules eye drops for the effective management of dry age-related macular degeneration. Expert Opin Drug Deliv. 2020;17:1305–20.

    Article  CAS  PubMed  Google Scholar 

  332. Mitra RN, Gao R, Zheng M, Wu M-J, Voinov MA, Smirnov AI, et al. Glycol chitosan engineered autoregenerative antioxidant significantly attenuates pathological damages in models of age-related macular degeneration. ACS Nano. 2017;11:4669–85.

    Article  CAS  PubMed  Google Scholar 

  333. Zhuge C-C, Xu J-Y, Zhang J, Li W, Li P, Li Z, et al. Fullerenol protects retinal pigment epithelial cells from oxidative stress-induced premature senescence via activating SIRT1. Invest Ophthalmol Vis Sci. 2014;55:4628–38.

    Article  PubMed  Google Scholar 

  334. Kwon Y-S, Zheng M, Zhang AY, Han Z. Melanin-like nanoparticles as an alternative to natural melanin in retinal pigment epithelium cells and their therapeutic effects against age-related macular degeneration. ACS Nano. 2022;16:19412–22.

    Article  CAS  PubMed  Google Scholar 

  335. Babizhayev MA, Deyev AI, Linberg LF. Lipid peroxidation as a possible cause of cataract. Mech Ageing Dev. 1988;44:69–89.

    Article  CAS  PubMed  Google Scholar 

  336. Babizhayev MA, Yegorov YE. Reactive oxygen species and the aging eye: specific role of metabolically active mitochondria in maintaining lens function and in the initiation of the oxidation-induced maturity onset cataract–a novel platform of mitochondria-targeted antioxidants with broad therapeutic potential for redox regulation and detoxification of oxidants in eye diseases. Am J Ther. 2016;23:e98-117.

    Article  PubMed  Google Scholar 

  337. Hanafy BI, Cave GWV, Barnett Y, Pierscionek BK. Nanoceria prevents glucose-induced protein glycation in eye lens cells. Nanomaterials. 2021;11:1473.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  338. O’Sullivan ED, Hughes J, Ferenbach DA. Renal aging: causes and consequences. J Am Soc Nephrol. 2017;28:407–20.

    Article  CAS  PubMed  Google Scholar 

  339. Stevens LA, Li S, Wang C, Huang C, Becker BN, Bomback AS, et al. Prevalence of CKD and comorbid illness in elderly patients in the United States: results from the kidney early evaluation program (KEEP). Am J Kidney Dis. 2010;55:S23–33.

    Article  PubMed  PubMed Central  Google Scholar 

  340. Sobamowo H, Prabhakar SS. The kidney in aging: physiological changes and pathological implications. Prog Mol Biol Transl Sci. 2017;146:303–40.

    Article  CAS  PubMed  Google Scholar 

  341. Sturmlechner I, Durik M, Sieben CJ, Baker DJ, van Deursen JM. Cellular senescence in renal ageing and disease. Nat Rev Nephrol. 2017;13:77–89.

    Article  CAS  PubMed  Google Scholar 

  342. Vlassara H, Torreggiani M, Post JB, Zheng F, Uribarri J, Striker GE. Role of oxidants/inflammation in declining renal function in chronic kidney disease and normal aging. Kidney International Supplement. 2009: S3–11.

  343. Tbahriti HF, Kaddous A, Bouchenak M, Mekki KJBRI. Effect of different stages of chronic kidney disease and renal replacement therapies on oxidant-antioxidant balance in uremic patients. 2013, 2013.

  344. J Rysz B Franczyk J Ławiński AJA Gluba-Brzózka 2020 Oxidative stress in ESRD patients on dialysis and the risk of cardiovascular diseases 9 1079

  345. Drożdż D, Kwinta P, Sztefko K, Kordon Z, Drożdż T, Łątka M, et al. Oxidative stress biomarkers and left ventricular hypertrophy in children with chronic kidney disease. 2016, 2016.

  346. Chien S-J, Lin I-C, Hsu C-N, Lo M-H. Tain Y-LJCJ Homocysteine and arginine-to-asymmetric dimethylarginine ratio associated with blood pressure abnormalities in children with early chronic kidney disease. Circ J. 2015;79:2031–7.

    Article  CAS  PubMed  Google Scholar 

  347. Qiao X, Chen X, Wu D, Ding R, Wang J, Hong Q, et al. Mitochondrial pathway is responsible for aging-related increase of tubular cell apoptosis in renal ischemia/reperfusion injury. J Gerontol A Biol Sci Med Sci. 2005;60:830–9.

    Article  PubMed  Google Scholar 

  348. Small DM, Bennett NC, Roy S, Gabrielli BG, Johnson DW, Gobe GC. Oxidative stress and cell senescence combine to cause maximal renal tubular epithelial cell dysfunction and loss in an in vitro model of kidney disease. Nephron Exp Nephrol. 2012;122:123–30.

    Article  CAS  PubMed  Google Scholar 

  349. Lin Y-F, Lee Y-H, Hsu Y-H, Chen Y-J, Lin Y-F, Cheng F-Y, et al. Resveratrol-loaded nanoparticles conjugated with kidney injury molecule-1 as a drug delivery system for potential use in chronic kidney disease. Nanomedicine. 2017;12:2741–56.

    Article  CAS  PubMed  Google Scholar 

  350. Cheng F-Y, Lee Y-H, Hsu Y-H, Chiu IJ, Chiu Y-J, Lin Y-F, et al. Promising therapeutic effect of thapsigargin nanoparticles on chronic kidney disease through the activation of Nrf2 and FoxO1. Aging. 2019;11:9875–92.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  351. Prasad AS, Bao B. Molecular mechanisms of Zinc as a pro-antioxidant mediator: clinical therapeutic implications. Antioxidants. 2019;8:164.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  352. Awadalla A, Hamam ET, El-Senduny FF, Omar NM, Mahdi MR, Barakat N, et al. Zinc oxide nanoparticles and spironolactone-enhanced Nrf2/HO-1 pathway and inhibited Wnt/β-catenin pathway in adenine-induced nephrotoxicity in rats. Redox Rep Commun Free Radical Res. 2022;27:249–58.

    CAS  Google Scholar 

  353. Reddy ARN, Reddy YN, Krishna DR, Himabindu V. Multi wall carbon nanotubes induce oxidative stress and cytotoxicity in human embryonic kidney (HEK293) cells. Toxicology. 2010;272:11–6.

    Article  CAS  PubMed  Google Scholar 

  354. Ferreira GK, Cardoso E, Vuolo FS, Michels M, Zanoni ET, Carvalho-Silva M, et al. Gold nanoparticles alter parameters of oxidative stress and energy metabolism in organs of adult rats. Biochem Cell Biol Biochimie Et Biol Cell. 2015;93:548–57.

    Article  CAS  Google Scholar 

  355. Vasanth SB, Kurian GA. Toxicity evaluation of silver nanoparticles synthesized by chemical and green route in different experimental models. Artif Cells Nanomed Biotechnol. 2017;45:1721–7.

    Article  CAS  PubMed  Google Scholar 

  356. Sarkar A, Das J, Manna P, Sil PC. Nano-copper induces oxidative stress and apoptosis in kidney via both extrinsic and intrinsic pathways. Toxicology. 2011;290:208–17.

    Article  PubMed  Google Scholar 

  357. Almeer RS, Ali D, Alarifi S, Alkahtani S, Almansour M. Green platinum nanoparticles interaction with HEK293 cells: cellular toxicity, apoptosis, and genetic damage. Dose-Response Publ Int Hormesis Soc. 2018;16:1559325818807382.

    CAS  Google Scholar 

  358. Cheng H-T, Ngoc Ta Y-N, Hsia T, Chen Y. A quantitative review of nanotechnology-based therapeutics for kidney diseases. Wiley Interdiscipl Rev Nanomed Nanobiotechnol. 2024;16: e1953.

    Article  CAS  Google Scholar 

  359. Shang S, Li X, Wang H, Zhou Y, Pang K, Li P, et al. Targeted therapy of kidney disease with nanoparticle drug delivery materials. Bioactive Mater. 2024;37:206–21.

    Article  CAS  Google Scholar 

  360. Varma K, Amalraj A, Divya C, Gopi S. The efficacy of the novel bioavailable curcumin (Cureit) in the management of sarcopenia in healthy elderly subjects: a randomized, placebo-controlled, double-blind clinical study. J Med Food. 2021;24:40–9.

    Article  CAS  PubMed  Google Scholar 

  361. Decha P, Kanokwan K, Jiraporn T, Pichaya J, Pisittawoot A. Phonopheresis associated with nanoparticle gel from phyllanthus amarus relieves pain by reducing oxidative stress and proinflammatory markers in adults with knee osteoarthritis. Chin J Integr Med. 2019;25:691–5.

    Article  CAS  PubMed  Google Scholar 

  362. Alizadeh F, Javadi M, Karami AA, Gholaminejad F, Kavianpour M, Haghighian HK. Curcumin nanomicelle improves semen parameters, oxidative stress, inflammatory biomarkers, and reproductive hormones in infertile men: a randomized clinical trial. Phytotherapy Research : PTR. 2018;32:514–21.

    Article  CAS  PubMed  Google Scholar 

  363. Qu Y, Wang Z, Zhou H, Kang M, Dong R, Zhao J. Oligosaccharide nanomedicine of alginate sodium improves therapeutic results of posterior lumbar interbody fusion with cages for degenerative lumbar disease in osteoporosis patients by downregulating serum miR-155. Int J Nanomed. 2017;12:8459–69.

    Article  CAS  Google Scholar 

  364. Miljkovic S, Jeftic B, Sarac D, Matovic V, Slavkovic M, Koruga D. Influence of hyper-harmonized fullerene water complex on collagen quality and skin function. J Cosmet Dermatol. 2020;19:494–501.

    Article  PubMed  Google Scholar 

  365. Dastani M, Rahimi HR, Askari VR, Jaafari MR, Jarahi L, Yadollahi A, et al. Three months of combination therapy with nano-curcumin reduces the inflammation and lipoprotein (a) in type 2 diabetic patients with mild to moderate coronary artery disease: evidence of a randomized, double-blinded, placebo-controlled clinical trial. BioFactors. 2023;49:108–18.

    Article  CAS  PubMed  Google Scholar 

  366. Xu S, Olenyuk BZ, Okamoto CT, Hamm-Alvarez SF. Targeting receptor-mediated endocytotic pathways with nanoparticles: rationale and advances. Adv Drug Deliv Rev. 2013;65:121–38.

    Article  CAS  PubMed  Google Scholar 

  367. Talelli M, Oliveira S, Rijcken CJF, Pieters EHE, Etrych T, Ulbrich K, et al. Intrinsically active nanobody-modified polymeric micelles for tumor-targeted combination therapy. Biomaterials. 2013;34:1255–60.

    Article  CAS  PubMed  Google Scholar 

  368. Koo Y-EL, Reddy GR, Bhojani M, Schneider R, Philbert MA, Rehemtulla A, et al. Brain cancer diagnosis and therapy with nanoplatforms. Adv Drug Delivery Rev. 2006;58:1556–77.

    Article  CAS  Google Scholar 

  369. Ajdary M, Keyhanfar F, Moosavi MA, Shabani R, Mehdizadeh M, Varma RS. Potential toxicity of nanoparticles on the reproductive system animal models: a review. J Reprod Immunol. 2021;148: 103384.

    Article  CAS  PubMed  Google Scholar 

  370. Mohamed NA, Marei I, Crovella S, Abou-Saleh H. Recent developments in nanomaterials-based drug delivery and upgrading treatment of cardiovascular diseases. Int J Mol Sci. 2022. https://doi.org/10.3390/ijms23031404.

    Article  PubMed  PubMed Central  Google Scholar 

  371. Doane T, Burda C. Nanoparticle mediated non-covalent drug delivery. Adv Drug Deliv Rev. 2013;65:607–21.

    Article  CAS  PubMed  Google Scholar 

  372. Wang N, Cheng X, Li N, Wang H, Chen H. Nanocarriers and their loading strategies. Adv Healthcare Mater. 2019;8: e1801002.

    Article  Google Scholar 

  373. Marrazzo P, O'Leary C. Repositioning Natural Antioxidants for Therapeutic Applications in Tissue Engineering. Bioengineering (Basel). 2020, 7.

  374. Tee JK, Ong CN, Bay BH, Ho HK, Leong DT. Oxidative stress by inorganic nanoparticles. Wiley Interdiscipl Rev Nanomed Nanobiotechnol. 2016;8:414–38.

    Article  CAS  Google Scholar 

  375. Zhang C, Wang X, Du J, Gu Z, Zhao Y. Reactive oxygen species-regulating strategies based on nanomaterials for disease treatment. Adv Sci. 2021;8:2002797.

    Article  CAS  Google Scholar 

  376. Coskun M, Kayis T, Gulsu E, Alp E. Effects of selenium and vitamin E on enzymatic, biochemical, and immunological biomarkers in Galleria mellonella L. Sci Rep. 2020;10:9953.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  377. Woolley JF, Stanicka J, Cotter TG. Recent advances in reactive oxygen species measurement in biological systems. Trends Biochem Sci. 2013;38:556–65.

    Article  CAS  PubMed  Google Scholar 

Download references

Funding

This work was financially supported by the National Key Research and Development Program of China (grant number 2022YFC2704100), National Natural Science Foundation of China (no. 22104040, 82001514).

Author information

Authors and Affiliations

Authors

Contributions

Jun Dai, Shixuan Wang: Conceptualization, Revision, Funding acquisition. Yun Dai, Meng Wu: Literature search, Visualization, Original article. Weicheng Tang, Dan Chen, Liru Xue, Yican Guo: Editing, Accessing quality. Yifan Guo, Simin Wei: Visualization, Revision. Final approval was passed through all authors.

Corresponding authors

Correspondence to Meng Wu, Jun Dai or Shixuan Wang.

Ethics declarations

Ethics approval and consent to participate

Not applicable.

Consent for publication

Not applicable.

Competing interests

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

Additional information

Publisher's Note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Rights and permissions

Open Access This article is licensed under a Creative Commons Attribution 4.0 International License, which permits use, sharing, adaptation, distribution and reproduction in any medium or format, as long as you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons licence, and indicate if changes were made. The images or other third party material in this article are included in the article's Creative Commons licence, unless indicated otherwise in a credit line to the material. If material is not included in the article's Creative Commons licence and your intended use is not permitted by statutory regulation or exceeds the permitted use, you will need to obtain permission directly from the copyright holder. To view a copy of this licence, visit http://creativecommons.org/licenses/by/4.0/. The Creative Commons Public Domain Dedication waiver (http://creativecommons.org/publicdomain/zero/1.0/) applies to the data made available in this article, unless otherwise stated in a credit line to the data.

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Dai, Y., Guo, Y., Tang, W. et al. Reactive oxygen species-scavenging nanomaterials for the prevention and treatment of age-related diseases. J Nanobiotechnol 22, 252 (2024). https://doi.org/10.1186/s12951-024-02501-9

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1186/s12951-024-02501-9

Keywords