Next Article in Journal
The Effects of Photosensitizing Dyes Fagopyrin and Hypericin on Planktonic Growth and Multicellular Life in Budding Yeast
Next Article in Special Issue
Profiling the Concentration of Reduced and Oxidized Glutathione in Rat Brain Using HPLC/DAD Chromatographic System
Previous Article in Journal
Design of Liposomes Carrying HelixComplex Snail Mucus: Preliminary Studies
Previous Article in Special Issue
Reaction of Chalcones with Cellular Thiols. The Effect of the 4-Substitution of Chalcones and Protonation State of the Thiols on the Addition Process. Diastereoselective Thiol Addition
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Detection of Oxidative Stress Induced by Nanomaterials in Cells—The Roles of Reactive Oxygen Species and Glutathione

Department of Biological and Biochemical Sciences, Faculty of Chemical Technology, University of Pardubice, Studentska 573, 532 10 Pardubice, Czech Republic
*
Author to whom correspondence should be addressed.
Molecules 2021, 26(16), 4710; https://doi.org/10.3390/molecules26164710
Submission received: 1 July 2021 / Revised: 22 July 2021 / Accepted: 2 August 2021 / Published: 4 August 2021
(This article belongs to the Special Issue Glutathione: Chemistry and Biochemistry)

Abstract

:
The potential of nanomaterials use is huge, especially in fields such as medicine or industry. Due to widespread use of nanomaterials, their cytotoxicity and involvement in cellular pathways ought to be evaluated in detail. Nanomaterials can induce the production of a number of substances in cells, including reactive oxygen species (ROS), participating in physiological and pathological cellular processes. These highly reactive substances include: superoxide, singlet oxygen, hydroxyl radical, and hydrogen peroxide. For overall assessment, there are a number of fluorescent probes in particular that are very specific and selective for given ROS. In addition, due to the involvement of ROS in a number of cellular signaling pathways, understanding the principle of ROS production induced by nanomaterials is very important. For defense, the cells have a number of reparative and especially antioxidant mechanisms. One of the most potent antioxidants is a tripeptide glutathione. Thus, the glutathione depletion can be a characteristic manifestation of harmful effects caused by the prooxidative-acting of nanomaterials in cells. For these reasons, here we would like to provide a review on the current knowledge of ROS-mediated cellular nanotoxicity manifesting as glutathione depletion, including an overview of approaches for the detection of ROS levels in cells.

1. Introduction

Molecular oxygen (O2) has a significant effect on numerous chemical reactions and biological processes. O2 reductions are one of the most critical electrocatalytic reactions that function in electrochemical energy conversion [1]. Free radicals contain an unpaired electron mostly bound to oxygen atoms. Conversely, the group of compounds named reactive oxygen species (ROS) also contains molecules without an unpaired electron, e.g., hydrogen peroxide [2,3]. Thus, the group of ROS also contains oxygen free radicals such as superoxide or hydroxyl, alkoxyl, peroxyl, and nitroxyl radicals [4,5]. The production of ROS is commonly linked with mitochondria, where the electrons are transferred through the respiratory chain to O2 forming water [6,7]. Mitochondrial ROS production depends on many factors such as the membrane potential of mitochondria [8], concentration of mitochondrial respiratory substrates, or a type of cells [9]. Mitochondria are the most important sources of superoxide and hydrogen peroxide in mammalian cells. The production of these ROS occurs mainly on the mitochondrial respiratory complex I and III [7,10]. In addition to mitochondrial complexes, ROS is also produced in mammalian cells by the participation of other enzymes such as flavoproteins [11] and other enzymes involved in nutrient metabolism [12]. As ROS plays important roles in the regulation of cell death processes, i.e., apoptosis [13] or necrosis [14,15,16], their pathological roles have been identified in a number of diseases including cancer and other age-related degenerative processes [17,18]. Given their deleterious effects, ROS production is usually finely tuned by ROS-scavenging systems [9].
Nanomaterials (NMs) exhibit great potential for use in the biomedical, optical, and electronic fields [19,20,21,22,23]. However, nanomaterials have been considered as potentially toxic due to their unique properties. They have extremely high surface-to-volume ratios, making them very reactive and catalytically active [24]. Their toxic potential in cells is also supported by their small size, enabling them to easily penetrate cell membranes [25]. TiO2 is one of the most commonly used nanomaterials in the chemical industry (e.g., cosmetics and pigments) [26]. In addition to white lead properties, TiO2 can be very active in photocatalytic reactions with organic compounds, providing the formation of ROS including OH, O2, H2O2 [27]. In addition to TiO2, other nanomaterials of different chemical compositions can produce ROS. The overview of NMs capable of ROS production is summarized in Table 1 including the lifetime.
Nanomaterials or nanoparticles (NPs) can expose transition metals on their surface, which can generate ROS through Fenton or Haber-Weiss reactions [44]. During these reactions, hydrogen peroxide is reduced in the presence of transition metals (Fe2+, Cu+) to form a highly active and toxic hydroxyl radical. Thus, the role of nanomaterials in ROS-mediated cell damage is significant and ROS production induced by NMs can lead to the modulation of various intracellular pathways, e.g., NF-κB, caspases, MAPK, etc., involving the activation of cell death processes [45,46].
In this study, we aimed to provide a recent and detailed view on ROS production induced by nanomaterials. The importance of our review can be also supported by the role of increased ROS levels that can lead to glutathione depletion and to the activation of cellular signaling pathways, resulting in changes in cellular metabolism, cell damage, or even in cell death.

2. Reactive Oxygen Species

2.1. Superoxide

Superoxide radical is formed during enzymatic and non-enzymatic reactions in biological systems [1,47]. In atoms and molecules, paired electrons occur usually as antiparallel, which strongly limits the oxidation properties of O2. After one-electron reduction of molecular oxygen, the superoxide radical (O2) forms. This reaction is thermodynamically very unfavorable and the interaction of O2 with another paramagnetic center is important for overcoming spin restriction [48]. Although the reactivity of O2 is mild, the crucial role of superoxide is that it enables the formation of other ROS (Figure 1), playing important roles in the pathology of various diseases.
Superoxide radical (O2) is formed mainly in mitochondria and its reactivity with biomolecules is relatively low. Superoxide can be produced after the reaction of molecular oxygen with divalent metals catalyzing a single-electron reduction under their simultaneous oxidation (equation 1).
O 2 +   Fe 2 +     O 2 . +   Fe 3 +
Another formation can be catalyzed by enzymes including xanthine oxidase, lipoxygenase, or cyclooxygenase [49]. The superoxide radical may exist in two possible forms: either in the form of O2 at physiological pH or as a hydroperoxyl radical (HO2) at low pH levels [50]. Hydroperoxyl radical penetrates better through phospholipid bilayers compared to the charged form O2 [28,51]. The superoxide radical may react with another superoxide radical to form hydrogen peroxide and O2 (equation 2). The reaction is catalyzed by the enzyme superoxide dismutase (SOD) [52,53]. A product of the dismutation reaction is H2O2 which becomes an important factor in the formation of the most reactive ROS, i.e., hydroxyl radical (OH) [54].
O 2 +   O 2 . + 2 H 2 O   Cu ,   Zn ,   Mn SOD   H 2 O 2 +   O 2
The mitochondrial electron transport chain (ETC) has been attributed to the role as the main ROS generator in cells. When transporting electrons, some of the electrons from the ETC can reduce molecular oxygen to O2 [55]. The resulting O2 is rapidly dismissed by mitochondrial superoxide dismutase (Mn-SOD) forming H2O2 [56]. Mitochondrial ETC consists of several electron transporters (flavoproteins, proteins containing iron and sulfur, ubiquinone, and cytochromes) with redox potentials ranging from −0.200 to +0.600 V [57,58]. According to the respective redox potentials, the individual electron carriers are arranged in individual complexes of the respiratory chain I–IV. Electrons that are transported into the respiratory chain as reducing equivalents of NADH or FADH2 enter the ETC through mitochondrial Complexes I and II. Then, the electrons are transferred through ETC to Complex IV which reduces O2 to H2O. From the thermodynamical perspective, all these electron transport systems could transfer the electrons directly to O2 to form O2. However, there are only two major sites of the respiratory chain where ROS can be generated, i.e., at Complexes I and III [59,60].
In Complex I, a reaction occurs between O2 and the reduced form of the flavinmononucleotide (FMN), leading to production of O2. The amount of reduced FMN depends on the NADH/NAD+ ratio [61]. In Complex III, two specific binding sites for coenzyme Q10 are known, i.e., Qi and Qo. Superoxide production is located in Qo. When antimycin A is added as an inhibitor of the Qi site, O2 production increases [62], while the addition of a myxothiazole inhibitor for the Qo site decreases ROS production [63]. Under physiological conditions, the production of ROS in Complex III depends on the ∆Ψ. The rate of O2 formation may increase exponentially with increasing ∆Ψ. This directly correlates with the fact that due to ∆Ψ fluctuations, the transport of electrons from heme bL to heme bH slows down, which then increases superoxide generation [64].

2.1.1. Role of Superoxide in Nanomaterial Toxicity

Damage to mitochondria and subsequent ROS leakage is a commonly accepted mechanism of nanoparticles toxicity. Damaged mitochondria release O2 into the inter-membrane space which can ultimately damage the cell [65]. Across different types of nanomaterials, their involvement in the ROS generation can be found. Far more often than in size, their possible cytotoxic effects are chemically dependent. Despite the similar size and crystal shape of ZnO NPs and SiO2 NPs, higher toxicity of ZnO NPs is observed, where cell viability is reduced and O2 generation is reduced, due to which glutathione (GSH) depletion occurs [29]. TiO2 nanoparticles generate O2 [30] both in solution and in cells, and intracellular O2 reduces the expression of histone deacetylase 9 (HDAC9), an epigenetic modifier [66]. Cellular internalization of TiO2 NPs has been shown to activate macrophages and neutrophils contributing to the production of O2 by the NADPH oxidase [67]. Oxidative stress induced by excessive O2 production is an important mechanism of the CuO NPs toxicity [31]. CuO NPs can enter HepG2 cells, where they are capable of inducing cellular toxicity by generating O2 leading to GSH depletion [68]. Activation of mitogen-activated protein kinases (MAPKs) and redox-sensitive transcription factors was demonstrated, suggesting that MAPK pathways and redox-sensitive transcription factors could be major factors of CuO NPs toxicity [69].
Analysis of mouse fibroblasts and human hepatocytes revealed that an increase in ROS levels induced by Ag NPs is accompanied by a reduction of mitochondrial membrane potential, release of cytochrome c into the cytosol, JNK activation, and translocation of Bax to mitochondria [32]. After exposure to Ag nanoparticles, GSH depletion occurs in liver cells, which is directly related to ROS production [70]. Ag NPs appear to induce DNA damage through a mechanism involving ROS production.

2.1.2. Methods for the Detection of Superoxide

MitoSox

Hydroethidium (HE) is a selective O2 detection probe (Figure 2) that reacts very rapidly to changes in O2 concentration, forming a red fluorescent product with 2-hydroxyethidium cation (2-OH-E+). Hydroethidine is a reduced form of ethidium that can be oxidized to ethidium in cells. The resulting ethidium intercalates nucleic acids and significantly increases its fluorescence, emitted at 610 nm (excitation = 535 nm) [23,71].
A new hydroethidine analog was synthesized for the purposes of O2 detection, which is produced in mitochondria. This analog carries a charged triphenylphosphonium residue (Mito-HE; Mito-Sox Red). As the phosphonium residue is positively charged and surrounded by three lipophilic phenyl groups, it penetrates very easily through cell membranes, mainly through the inner mitochondrial membrane [72]. After they cross the cell membranes, they accumulate in mitochondria depending on the negative ∆Ψ [73]. Importantly, redistribution of MitoSox from mitochondria is dependent on decreasing ∆Ψ based on various stimuli, which may not be ROS. For this reason, the use of MitoSox is a semi-quantitative test. Very important is the fact that MitoSox is transferred from mitochondria to the cytoplasm. Here, the supply of nucleic acids is higher and the increasing fluorescence is independent to mitochondrial ROS production, which may distort the results of individual measurements. The formation of MitoSox oxidation products in mitochondria may result in changes of values, which may reduce the passage of other MitoSox molecules into the mitochondria and generally affect measurements due to decreased MitoSox and ROS concentrations that are not produced by breathing chain breakage. The fluorescent product emits radiation at 580 nm with excitation at 540 nm [74,75,76].

1,3–Diphenylisobenzofuran

The 1,3-diphenylisobenzofuran (DPBF) probe is a molecule that, when incorporated into liposome phospholipids, acquires fluorescent properties. It is used for the detection of O2 and 1O2. After reaction with oxygen radicals, it produces a decrease of fluorescence, thus the fluorescence rates correlate inversely with increasing concentrations of O2 and 1O2 [77,78]. The reaction of DPBF with ROS such as singlet oxygen, hydroxyl, alkoxy and alkyl peroxy radicals gives 1,2-dibenzoylbenzene. In contrast, only reaction with H2O2 produces 9-hydroxyanthracen-10-(9H)-one. This product can be detected using fluorescence spectroscopy, NMR spectroscopy, or HPLC [79].

2.2. Hydroxyl Radical

The hydroxyl radical is a neutral form of the hydroxide ion. It belongs among the most reactive ROS because it can react with a variety of organic and inorganic compounds including DNA, proteins, and lipids, resulting in serious cell damage. The hydroxyl radical may be formed as a product of the Fenton or Haber–Weiss reaction [80,81,82,83].
The Fenton reaction is based on the reaction between H2O2 and Fe2+. Iron is an essential component of many proteins involved in the transport or metabolism of oxygen due to its ability to undergo cyclic oxidation and reduction. Iron has to be present for the ongoing synthesis of iron-containing proteins. As such, it can directly lead to the formation of free radicals, which can cause cellular damage of large extent. The reaction of Fe2+ with H2O2 produces an oxidized form of iron (Fe3+), as well as OH and OH (Equation (3)).
Fe 2 + +   H 2 O 2     Fe 3 +   [ H 2 O 2 ]     OH +   OH
O 2 . +   H 2 O 2 O 2 +   OH +   OH
Another possible reaction to form OH is the Haber–Weiss reaction. In this reaction, less reactive O2 and H2O2 react with each other (Equation (4)). As in the case of the Fenton reaction, very toxic OH is formed. Very unfavorable thermodynamic conditions are applied to this reaction, in which the rate constant in the aqueous solution is close to zero. The presence of a transition metal catalyst is required to ensure the reaction. The iron atom serves as the catalyst. Both reactions produce highly reactive OH, which ultimately severely damages cells [84,85,86,87]. The Fenton reaction can be used to induce apoptosis in cancer cells, where OH is formed on a copper ion [88,89].

2.2.1. Role of Hydroxyl Radical in Nanomaterial Toxicity

TiO2 and ZnO NPs are widely used in cosmetics and industry [22]. Under the influence of UV radiation, ZnO NPs generate reactive oxygen species such as OH or H2O2, causing GSH depletion [33,90]. The rate of OH generation and the total photocatalytic activity depends on the physical properties of the nanomaterial used, e.g., TiO2 NPs [34]. Cu NPs play an important role as a cofactor in a number of enzymes such as cytochrome c oxidase [91]. However, they exhibit significant toxicity and can induce ROS production, including largely reactive OH. Copper can catalyze electron transfer (Cu2+ and Cu+). This can give rise to O2 reduction to H2O2 in cells, leading to GSH depletion [35]. Other particles that induce OH production include Fe3O4 [92], silica nanoparticle [93], and silver nanoparticles [94].

2.2.2. Methods for the Detection of Hydroxyl Radical

Terephthalic acid (TA) can be hydroxylated in presence of OH to give the highly fluorescent product 2-hydroxy-TA [95]. TA has a configuration of two carboxylate anion (COO) side groups attached to a six-carbon ring at positions 1 and 4 to form a structurally symmetrical compound. Reaction of OH with any of the four unsubstituted carbons will form only one hydroxylated product, 2-hydroxy-TA (2-OH-TA). TA is non-fluorescent, whereas 2-OH-TA is highly fluorescent. Neither TA nor 2-OH-TA is present in tissues physiologically. In addition, none of them is known to be involved in cellular functions, thus they exhibit no cellular toxicity [96].
Fluorogenic spin probes can be used to detect OH. Their signal can be detected both fluorometrically and using EPR spectroscopy. The rhodamine nitroxide probe is a non-fluorescent substance reacting quantitatively with OH (Ex/Em = 560/588 nm) [97].
The HKOH-1 probe was designed for better uptake and longer retention in cells. The HKOH-1 probe has excellent sensitivity, selectivity, and extremely rapid turn-on response toward OH in live cells in both confocal imaging and flow cytometry experiments [98].

2.3. Singlet Oxygen

Singlet oxygen (1O2), the highest energy state of molecular oxygen, has been extensively studied to oxidize toxic persistent organic contaminants [99]. Singlet oxygen is a highly reactive form of oxygen. It is produced during photochemical reactions or even physiologically in the respiratory chain of mitochondria. In excitation, molecular oxygen is excited to the first state (1∆g) and then to the higher excited state (1∑g). In the first excited state, O2 has two counter-spin electrons in a π orbital, while in the second excited state, O2 has one counter-spin electron in two π orbitals [100,101]. The first excited state is highly reactive. 1∆g 1O2 is also produced physiologically, e.g., in the activation of neutrophils and macrophages [102,103]. It is a highly potent oxidizing agent that can cause fatal damage of DNA [104] or cell death [105,106].
Singlet oxygen reacts with several biological molecules including DNA, RNA, lipids, sterols, and especially proteins [107]. Amino acid residues of proteins can react with 1O2 by direct chemical reaction or physical quenching. Physical quenching causes de-excitation of the singlet state of oxygen proved in proteins through the interaction with tryptophan residues [108].

2.3.1. Role of Singlet Oxygen in Nanomaterial Toxicity

Reactive oxygen species are formed by the reaction of photoinduced binding electrons with oxygen molecules. After the release of photoinduced electrons, valence band holes are formed on the surface of TiO2 NPs that cannot oxidize water [109]. Another type of ROS that occurs during photocatalytic reactions on the surface of TiO2 NPs is 1O2 (Figure 3) [41]. Nanomaterials that can induce singlet oxygen production also include Ag NPs [42]. Nanomaterials-bound generation of 1O2 can be also used in the treatment of tumors [43]. An activatable system has been developed that enables tumor-specific 1O2 generation, based on a Fenton-like reaction between linoleic acid hydroperoxide (LAHP), tethered on FeO NPs and Fe2+ ions released from FeO NPs under acidic pH conditions [43]. After increased production of 1O2 in cells, the intracellular concentration of GSH decreases [110,111,112].

2.3.2. Methods for the Detection of Singlet Oxygen

The DPAX-1 fluorescent probe (9-[2-(3-carboxy-9,10-diphenyl)-anthryl]-6-hydroxy-3H-xanthen-3-one) has been used to detect 1O2 forming endoperoxide as a reaction product. The probe is based on 9,10-diphenylanthracene (DPA), conjugated to fluorescein. The high quantum yield and wavelength of the excitation radiation are suitable for biological applications [113]. The DMAX 9-[2-(3-carboxy-9,10-dimethyl)anthryl]-6-hydroxy-3H-xanthen-3-one has been also used to detect 1O2. The DMAX probe reacts much more specifically and faster with 1O2 compared to the DPAX-1 probe [114].
Other approach for singlet oxygen detection are amino-functionalized nanoparticles covalently linked to Singlet Oxygen Sensor Green® (SOSG) which is an anthracene-fluorescein dye. The fluorescence of the SOSG molecule is inhibited by photoinduced intramolecular electron transfer. When anthracene is endoperoxidized in the presence of 1O2, the electron transfer is blocked and fluorescein self-fluorescence is restored [115].

2.4. Hydrogen Peroxide

Hydrogen peroxide is formed directly through SOD-catalyzed dismutation from superoxide [116]. It belongs among ROS but it is not a free radical. The relatively long lifespan and size of H2O2 allows it to pass through cell membranes to different parts of the cell, which facilitates signaling reactions [117]. It causes cell damage at concentrations higher than 100 nM. Concentration of H2O2 in the range of 1–10 nM acts physiologically in the process of redox signaling [116]. It does not cause direct DNA damage but DNA damage is ensured due to OH presence, which arises from H2O2 in the presence of transition metal ions [118]. Enzymes eliminating H2O2 include catalase, glutathione peroxidase, and peroxiredoxins [119].
In peroxisomes, the main metabolic process producing H2O2 is the β-oxidation of fatty acids through acyl-CoA-oxidase. Other enzymes involved in the formation of ROS include urate oxidase [120], D-aspartate oxidase [121], or xanthine oxidase [28].

2.4.1. Role of Hydrogen Peroxide in Nanomaterial Toxicity

Most nanomaterials that induce the production of O2 also induce the production of H2O2. In a study [36], colorectal cancer cells were exposed to polystyrene NPs (20 and 40 nm) with two surfactants (amino and carboxylic acid). After the exposure of cells to polystyrene NPs, a decrease in cell viability was observed and the induction of the apoptosis process was reduced by decreased H2O2 production by catalase. In another study [37], the authors observed a decrease in intracellular GSH concentration after the exposure of cells to 8 nm Au NPs. Subsequently, it was found that there was a decrease in mitochondrial membrane potential (∆Ψ) and cell apoptosis deepened after 48 h of incubation of cells with Au NPs. Then, a decreased mitochondrial GSH concentration and increased H2O2 production were observed. Other nanomaterials capable of induction of H2O2 formation are e.g., TiO2 NPs [38], ZnO NPs [39], and Ag NPs [40].

2.4.2. Methods for the Detection of Hydrogen Peroxide

2′,7′-Dichlorodihydrofluorescein

The 2′,7′-dichlorodihydrofluorescein (DCFH) probe is a specific indicator of the presence of H2O2. The diacetate form of DCFH (DCFH-DA) has been used to detect ROS in cells due to its ability to penetrate cell membranes. Two acetate groups are hydrolyzed by intracellular esterases after DCFH-DA transfer into cells. Then, the presence of peroxidases is important for the oxidation of DCFH by H2O2. Other agents capable of oxidizing DCFH include hematin or cytochrome c [122,123] which may increase the fluorescence of the probe without any H2O2 production [124]. DCFH can be also oxidized with H2O2 in the presence of Fe2+ but this is most likely due to the formation of OH. In contrast, O2 is unable to oxidize the DCFH probe [125]. In the presence of visible light or ultraviolet radiation, a DCF photoreduction can occur (Figure 4). The fluorescent product exhibits fluorescence at 522 nm (excitation at 498 nm).
The oxidation of the probe produces a semichinone radical (DCF●−) that, when reacted with O2, gives rise to O2. Dismutation of O2 produces H2O2 that then artificially increases the oxidation of DCFH. The oxidation of DCFH results in the formation of a fluorescent product DCF exhibiting strong fluorescence. However, this reaction can increase the fluorescence intensity of the DCF product and give false-positive results [126,127,128]. In the case of the measurement of ROS production in tested nanomaterials, the form of DCFH-DA has been mostly used in ZnO2 NMs [33,129,130,131,132] and TiO2 NMs [133,134,135,136].

Amplex Red

Amplex Red (N-acetyl-3,7-dihydroxyphenoxazine) is a non-fluorescent molecule that can be specifically oxidized by H2O2 in the presence of horseradish peroxidase (HRP) to the highly fluorescent resorufin product (Figure 5), EX/EM 563/587 nm [137]. At excessive H2O2 concentrations, the fluorescent product resorufin can be further oxidized to non-fluorescent resazurin [138]. Amplex Red reacts with H2O2 stoichiometrically. It can also be used for the detection of O2 in a mixture with SOD converting O2 to H2O2. The background fluorescence during the measurement is very low and the fluorescent product is very stable. These features increase the sensitivity of the measurement. Significant loss of fluorescence may be due to the oxidation of resorufin to the non-fluorescent resazurin product that can be catalyzed by HRP [139,140].

HyPer Ratiometric Sensor

The H2O2 concentration can be measured using the expression of a HyPer genetically encoded ratio sensor. HyPer consists of the bacterial H2O2-sensitive transcription factor OxyR, fused to the circular fluorescent protein YFP. Cysteine oxidation of the OxyR moiety induces a conformational change that results in an increase in YFP fluorescence intensity excited at 500 nm and a decrease in YFP emission excited at 420 nm. This reversible change can monitor the intracellular concentration of H2O2 [141].

Pentafluorobenzenesulfonyl Fluoresceins

Perhydrolysis of acyl resorufins is a reaction that acts as a fluorescent indicator for the determination of H2O2. This method is based on deprotection rather than oxidation, which enables the fluorescence of resorufin and fluorescein. The selectivity of this method for H2O2 detection is higher compared to DCFH. For the above reasons, pentafluorobenzenesulfonyl fluoresceins have been proposed as selective fluorescent probes for H2O2 detection. Importantly, sulfonates are more stable to hydrolysis than esters. Fluoresceins have high fluorescence yields and the pentafluorobenzene ring increases the reactivity of sulfonates with H2O2 [142].

Europium Ion

The method is based on the binding of Eu3+-tetracycline [Eu (tc)] linked to propanesulfonic acid (MOPS) in an aqueous solution to H2O2. After binding, a strongly fluorescent complex ([Eu (hp) (tc)]) is formed (λEX/EM = 390-405 /616 nm). The increase in fluorescence is up to 15x after H2O2 binding and it is strongly dependent on the pH value. The increase in fluorescence is most pronounced at the physiological pH environment. The fluorescence of the probe [Eu (tc)] is not affected by ammonium, chloride, sulphate, or nitrate ions. However, citrate and phosphate can interfere with the assay [143].

Homovanilic Acid

Recently, homovanillic acid (3-methoxy-4-hydroxyphenylacetic acid) has been increasingly used instead of scopoletin for H2O2 detection in mitochondria. In contrast to the fluorescent scopoletin indicating the presence of H2O2 by a fluorescence decrease, homovanillic acid becomes a fluorescent through H2O2-induced oxidation in the presence of HRP [144]. The product of this reaction is a highly fluorescent dimer 2,2′-dihydroxy-3,3′-dimethoxydiphenyl-5,5′-diacetic acid [145]. In the following Table 2, an overview of all described fluorescent probes for ROS detection are summarized.

3. Role of Reactive Oxygen Species Induced by Nanoparticles in Cell Signaling

Nanomaterials are capable of interfering with cell signaling pathways. Recently, three main pathways participating in the apoptosis process have been identified (Figure 6). The first pathway is the direct NMs occupation of the FADD receptor. The second pathway is the modulation of the function of mitochondria in the presence of NMs and the third is the localization of NMs pacting in the endoplasmic reticulum. All of these pathways converge upon caspase activation, thereby the mitochondria produce higher levels of ROS, increase production of Bid protein, and activate Bax or Bak1 proteins, which can ultimately lead to organelle damage, DNA cleavage, and cell death [146].
The dynamic and rapid nature of ROS signaling is the result of ROS production and removal. The balance between the production and removal of ROS is balanced due to their interaction. This causes rapid changes in ROS levels [147]. ROS play an important role in activating many cellular proteins and factors, e.g., NF-κB, MAPK, Keap1-Nrf2-ARE, or PI3K-Akt [148,149].
The NF-κB family is a family of transcriptional proteins consisting of five members, i.e., NF-κB1, NF-κB2, RelA, RelB, and c-Rel [150]. The activation of the transcription factor NF-κB involves signal-dependent degradation of phosphorylated inhibitors such as IκBα. The mechanism of NF-κB activation by H2O2 [151] or O2 [152] is different from the activation in the presence of cytokines or mitogens. Serines 32 and 36 play a key role in the activation of NF-κB by cytokines, while tyrosine residues 42 and serine/threonine in the PEST domain of the IκBα protein play a key role in the activation by H2O2 [153]. H2O2 activates IκBα kinase without subsequent serine phosphorylation of IκBα. In contrast, H2O2, similar to TNF, induces serine phosphorylation of the p65 subunit of NF-kB, leading to its nuclear translocation [154]. Nanoparticles participate directly in the activation of the factor NF-κB through increased ROS production which was confirmed by the translocation of the high-mobility group box 1 (HMGB1) protein from the nucleus to the cytoplasm observed in cells after exposure to silica nanoparticles [155]. Subsequently, HMGB1 binds to the TLR4 receptor; this complex regulates the expression of the myeloid differentiation factor and activates the NF-κB-signaling pathway.
In eukaryotic cells, signaling by MAPK kinases is very important. Various MAPK pathways can be activated by different stimuli. Ultimately, activated MAPK pathways coordinate gene transcription activation, acting in the regulation of protein synthesis, cell cycle, cell death, and cell differentiation [156]. The MAPK cascade is composed of three distinct signaling modules, i.e., the c-Jun N-terminal kinase cascade, the p38 MAPK cascade, and the extracellular signal-regulated kinase ERK [157]. Several cellular stimuli activating ROS production can also activate MAPK activation itself [158]. For instance, MAPK kinases can be activated by H2O2 [159]. MAPK activation occurs by activating growth factor receptors in several cell types [160]. Another mechanism of MAPK activation by ROS is the inactivation of the MKP protein by its oxidation [161]. The physiological FEM protein keeps the MAPK signaling pathway inactive. In addition to the activation of MAPK, the JNK pathway is also activated during the oxidation of the FEM protein [162]. A number of studies have demonstrated the activation of a variety of kinases by ROS, including ASK1 [163], MEKK1 [164], c-Src [165], and EGFR [166]. These activated kinases ultimately can activate the MAPK cascade [167]. Cerium oxide particles have been shown to activate ROS production and to reduce SOD and glutathione peroxidase activities. This results in increased phosphorylation levels of p38 MAPK as well as ERK1/2 and JNK [168]. The nanoparticles that can damage cells through p38 MAPK activation are silica NPs [169,170], polystyrene NPs [171], and TiO2 NPs [172]. Conversely, the exposure to Au [173] and iron oxide [174] NPs causes the osteogenetic differentiation through the activation of relevant genes by p38 MAPK.
The tumor suppressor protein p53 induces apoptotic cell death in response to oncogenic stress. Malignant progression is dependent on the loss of p53 function by mutations in the TP53 gene itself or defects in signaling pathways. Phosphorylation of p53 regulates the ability to activate the expression of apoptotic target genes [175]. Overexpression of p53 transactivates a number of p53 genes. Many of these genes encode redox active proteins including enzymes (quinone oxidoreductase and proline oxidase) generating ROS. Ultimately, this regulation of ROS production leads to oxidative stress that can induce apoptosis [176]. Increasing the intracellular concentration of ROS leads to the activation of the p38 protein, which increases the expression and transcriptional activity of p53 [177]. The p53 protein transcriptionally activates the PUMA gene encoding two proteins, PUMA-α and PUMA-β, of similar activity. These proteins bind to Bcl-2 and integrate into the mitochondria, where they induce the release of cytochrome c [178,179,180].
Last but not least, ROS activate the JNK kinase pathway, which plays an important role in the apoptosis process [4,181]. During intracellular ROS production, there is a permanent activation of JNK [182]. This is due to the inactivation of MAPK phosphatases (FEM) by oxidation of their catalytic cysteine in the presence of intracellularly accumulated H2O2. Expression of catalytically inactive FEMs prolongs JNK activation [183].

4. Current Trends in the Evaluation of Nanotoxicity In Vitro

The number of studies focusing on nanotoxicity testing has been growing very rapidly in the last two decades. The cause of that can be also found in the perpetual production of new nanomaterials for its following use in industry or medicine. Conversely, especially in medicine, nanomaterials raise some concerns regarding their cytotoxicity or biocompatibility. Thus, a number of scientific projects have been assessing the toxicity of the selected nanomaterials and creating the risk management framework for the use of nanomaterials in medical applications [184].
Recent studies on nanotoxicity have been using basic assays for the evaluation of cell function changes, e.g., cell viability, membrane integrity, and enzyme activities measurements. To estimate the oxidative status in cells, the levels of antioxidants can be measured using a number of methods. In addition to the most frequently used methods, other approaches have been used to characterize the cellular nanotoxicity recently. These methods include scanning electron microscopy [185], liquid cell transmission electron microscopy [186], atomic force microscopy [187], and hyperspectral and laser confocal microscopy applied to cell-nanoparticles interactions [185]. All these microscopic methods are very sensitive and specific, which allows for a very detailed description of the function state of the cells after nanomaterials treatment. To understand the toxicity of nanomaterials, we need to develop new and innovative methods that will provide us with information about the changes in the intracellular environment after exposure to nanomaterials. In addition, there is a need to develop methods that are fast, robust, and combine several biological tests. In contrast to conventional assays using lipophilic fluorescent probes detecting ROS levels, a nanoelectrode has been developed to study the toxicity of magnetic nanoparticles. The nanoelectrode is composed of individual platinum nanoelectrodes with a cavity at the tip. It is part of an upright microscope and is used to measure intracellular ROS [188].
A further topic of interest in nanotoxicity testing is the use of newly developed relevant biological models. In comparison to two-dimensional (2D) cultured cell lines, those new biogical models ought to provide accurate predictions of nanomaterials effects in vivo. Thus, some new scientific studies described the use of pulmonary fibrosis models [189], organ on-chip technology bridging the differences between 2D in vitro and three-dimensional (3D) in vivo models from skin, the lung, and the liver [190,191], or on-chip placenta models [192]. Despite advanced organ on-chip models, a number of concerns have to be solved to ensure the comparability to living systems in obtained outcomes [193].

5. Conclusions

Currently, nanotechnology is considered to be one of the most attractive research topics due to its huge application potential and commercial impact. Due to the large number of newly manufactured nanomaterials, it is necessary to evaluate their possible cytotoxic effects in men. At present, there is a large request to investigate thee potential acute and chronic effects of nanomaterials especially in vitro in cells. Those studies can provide a mechanistic view on nanomaterial cellular acting. However, the use of proper and relevant bioanalytical methods for evaluating the nanomaterials effects in cells is necessary.
In this study, we aimed to provide a recent and detailed view on ROS production induced by nanomaterials, especially considering the metalic nanoparticles. In cells, the nanotoxicity can be mediated by a number of substances including ROS. Depending on the composition and shape of a nanomaterial, a variety of ROS can be formed in cells, i.e., O2●−, 1O2, OH, and H2O2. Thus, the importance of the present review can be recognized in the mechanistic description of a relation of nanomaterials of different chemical compositions and ROS production. We provided the current knowledge of ROS-mediated cellular nanotoxicity together with the possibilities of ROS detection in cells using specific fluorescent probes. In addition, we summarized the detailed description of the relationship between nanomaterials-mediated ROS production and glutathione depletion. Altogether, the prooxidative action of nanomaterials can ultimately lead to the activation of cellular signaling pathways, causing a change in cellular metabolism, cell damage, or even cell death.

Funding

Financial support was received from the Ministry of Education, Youth, and Sports of the Czech Republic via project NANOBIO (Reg. No. CZ.02.1.01/0.0/0.0/17_048/0007421).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Hayyan, M.; Hashim, M.A.; AlNashef, I.M. Superoxide Ion: Generation and Chemical Implications. Chem. Rev. 2016, 116, 3029–3085. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  2. Lushchak, V.I. Free radicals, reactive oxygen species, oxidative stress and its classification. Chem. Biol. Interact. 2014, 224, 164–175. [Google Scholar] [CrossRef] [PubMed]
  3. Juan, C.A.; Perez de la Lastra, J.M.; Plou, F.J.; Perez-Lebena, E. The Chemistry of Reactive Oxygen Species (ROS) Revisited: Outlining Their Role in Biological Macromolecules (DNA, Lipids and Proteins) and Induced Pathologies. Int. J. Mol. Sci. 2021, 22, 4642. [Google Scholar] [CrossRef] [PubMed]
  4. Simon, H.U.; Haj-Yehia, A.; Levi-Schaffer, F. Role of reactive oxygen species (ROS) in apoptosis induction. Apoptosis 2000, 5, 415–418. [Google Scholar] [CrossRef] [PubMed]
  5. Wagner, H.; Cheng, J.W.; Ko, E.Y. Role of reactive oxygen species in male infertility: An updated review of literature. Arab. J. Urol. 2018, 16, 35–43. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  6. Ott, M.; Gogvadze, V.; Orrenius, S.; Zhivotovsky, B. Mitochondria, oxidative stress and cell death. Apoptosis 2007, 12, 913–922. [Google Scholar] [CrossRef] [PubMed]
  7. Zhao, R.Z.; Jiang, S.; Zhang, L.; Yu, Z.B. Mitochondrial electron transport chain, ROS generation and uncoupling (Review). Int. J. Mol. Med. 2019, 44, 3–15. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  8. Suski, J.; Lebiedzinska, M.; Bonora, M.; Pinton, P.; Duszynski, J.; Wieckowski, M.R. Relation Between Mitochondrial Membrane Potential and ROS Formation. Methods Mol. Biol. 2018, 1782, 357–381. [Google Scholar] [PubMed]
  9. Mazat, J.P.; Devin, A.; Ransac, S. Modelling mitochondrial ROS production by the respiratory chain. Cell Mol. Life Sci. 2020, 77, 455–465. [Google Scholar] [CrossRef] [PubMed]
  10. Parey, K.; Wirth, C.; Vonck, J.; Zickermann, V. Respiratory complex I—structure, mechanism and evolution. Curr. Opin. Struct. Biol. 2020, 63, 1–9. [Google Scholar] [CrossRef]
  11. Husen, P.; Nielsen, C.; Martino, C.F.; Solov’yov, I.A. Molecular Oxygen Binding in the Mitochondrial Electron Transfer Flavoprotein. J. Chem. Inf. Model. 2019, 59, 4868–4879. [Google Scholar] [CrossRef] [PubMed]
  12. Mailloux, R.J. An Update on Mitochondrial Reactive Oxygen Species Production. Antioxidants 2020, 9, 472. [Google Scholar] [CrossRef] [PubMed]
  13. Papa, S.; Skulachev, V.P. Reactive oxygen species, mitochondria, apoptosis and aging. Mol. Cell BioChem. 1997, 174, 305–319. [Google Scholar] [CrossRef] [PubMed]
  14. Ventura, J.J.; Cogswell, P.; Flavell, R.A.; Baldwin, A.S., Jr.; Davis, R.J. JNK potentiates TNF-stimulated necrosis by increasing the production of cytotoxic reactive oxygen species. Genes Dev. 2004, 18, 2905–2915. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  15. Zhang, M.; Harashima, N.; Moritani, T.; Huang, W.; Harada, M. The Roles of ROS and Caspases in TRAIL-Induced Apoptosis and Necroptosis in Human Pancreatic Cancer Cells. PLoS ONE 2015, 10, e0127386. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  16. Yang, J.; Zhao, X.; Tang, M.; Li, L.; Lei, Y.; Cheng, P.; Guo, W.; Zheng, Y.; Wang, W.; Luo, N.; et al. The role of ROS and subsequent DNA-damage response in PUMA-induced apoptosis of ovarian cancer cells. Oncotarget 2017, 8, 23492–23506. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  17. Liou, G.Y.; Storz, P. Reactive oxygen species in cancer. Free Radic. Res. 2010, 44, 479–496. [Google Scholar] [CrossRef] [Green Version]
  18. Storz, P. Reactive oxygen species in tumor progression. Front. BioSci. 2005, 10, 1881–1896. [Google Scholar] [CrossRef] [Green Version]
  19. Zhou, B.; Guo, X.; Yang, N.; Huang, Z.; Huang, L.; Fang, Z.; Zhang, C.; Li, L.; Yu, C. Surface engineering strategies of gold nanomaterials and their applications in biomedicine and detection. J. Mater. Chem. B 2021, 9, 5583–5598. [Google Scholar] [CrossRef]
  20. Sakr, T.M.; Korany, M.; Katti, K.V. Selenium nanomaterials in biomedicine—An overview of new opportunities in nanomedicine of selenium. J. Drug Deliv. Sci. Technol. 2018, 46, 223–233. [Google Scholar] [CrossRef]
  21. Mehlenbacher, R.D.; Kolbl, R.; Lay, A.; Dionne, J.A. Nanomaterials for in vivo imaging of mechanical forces and electrical fields. Nat. Rev. Mater. 2017, 3, 1–17. [Google Scholar] [CrossRef]
  22. Musial, J.; Krakowiak, R.; Mlynarczyk, D.T.; Goslinski, T.; Stanisz, B.J. Titanium Dioxide Nanoparticles in Food and Personal Care Products-What Do We Know about Their Safety? Nanomaterials 2020, 10, 1110. [Google Scholar] [CrossRef]
  23. Holmila, R.J.; Vance, S.A.; King, S.B.; Tsang, A.W.; Singh, R.; Furdui, C.M. Silver Nanoparticles Induce Mitochondrial Protein Oxidation in Lung Cells Impacting Cell Cycle and Proliferation. Antioxidants 2019, 8, 552. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  24. Drasler, B.; Sayre, P.; Steinhäuser, K.G.; Petri-Fink, A.; Rothen-Rutishauser, B. In vitro approaches to assess the hazard of nanomaterials. NanoImpact 2017, 8, 99–116. [Google Scholar] [CrossRef]
  25. Yin, J.J.; Liu, J.; Ehrenshaft, M.; Roberts, J.E.; Fu, P.P.; Mason, R.P.; Zhao, B. Phototoxicity of nano titanium dioxides in HaCaT keratinocytes--generation of reactive oxygen species and cell damage. Toxicol. Appl. Pharmacol. 2012, 263, 81–88. [Google Scholar] [CrossRef] [Green Version]
  26. Ray, P.C.; Yu, H.T.; Fu, P.P. Toxicity and Environmental Risks of Nanomaterials: Challenges and Future Needs. J. Environ. Sci. Health C 2009, 27, 1–35. [Google Scholar] [CrossRef] [Green Version]
  27. Daimon, T.; Nosaka, Y. Formation and behavior of singlet molecular oxygen in TiO2 photocatalysis studied by detection of near-infrared phosphorescence. J. Phys. Chem. C 2007, 111, 4420–4424. [Google Scholar] [CrossRef]
  28. Phaniendra, A.; Jestadi, D.B.; Periyasamy, L. Free Radicals: Properties, Sources, Targets, and Their Implication in Various Diseases. Indian J. Clin. Biochem. 2015, 30, 11–26. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  29. Yang, H.; Liu, C.; Yang, D.; Zhang, H.; Xi, Z. Comparative study of cytotoxicity, oxidative stress and genotoxicity induced by four typical nanomaterials: The role of particle size, shape and composition. J. Appl. Toxicol. 2009, 29, 69–78. [Google Scholar] [CrossRef]
  30. He, X.; Sanders, S.; Aker, W.G.; Lin, Y.; Douglas, J.; Hwang, H.M. Assessing the effects of surface-bound humic acid on the phototoxicity of anatase and rutile TiO2 nanoparticles in vitro. J. Environ. Sci. 2016, 42, 50–60. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  31. Zhang, J.; Wang, B.; Wang, H.; He, H.; Wu, Q.; Qin, X.; Yang, X.; Chen, L.; Xu, G.; Yuan, Z.; et al. Disruption of the superoxide anions-mitophagy regulation axis mediates copper oxide nanoparticles-induced vascular endothelial cell death. Free Radic. Biol. Med. 2018, 129, 268–278. [Google Scholar] [CrossRef] [PubMed]
  32. Onodera, A.; Nishiumi, F.; Kakiguchi, K.; Tanaka, A.; Tanabe, N.; Honma, A.; Yayama, K.; Yoshioka, Y.; Nakahira, K.; Yonemura, S.; et al. Short-term changes in intracellular ROS localisation after the silver nanoparticles exposure depending on particle size. Toxicol. Rep. 2015, 2, 574–579. [Google Scholar] [CrossRef] [Green Version]
  33. Ahamed, M.; Akhtar, M.J.; Raja, M.; Ahmad, I.; Siddiqui, M.K.; AlSalhi, M.S.; Alrokayan, S.A. ZnO nanorod-induced apoptosis in human alveolar adenocarcinoma cells via p53, survivin and bax/bcl-2 pathways: Role of oxidative stress. Nanomedicine 2011, 7, 904–913. [Google Scholar] [CrossRef] [PubMed]
  34. Jimenez-Relinque, E.; Castellote, M. Hydroxyl radical and free and shallowly trapped electron generation and electron/hole recombination rates in TiO2 photocatalysis using different combinations of anatase and rutile. Appl. Catal. A Gen. 2018, 565, 20–25. [Google Scholar] [CrossRef]
  35. Thit, A.; Selck, H.; Bjerregaard, H.F. Toxic mechanisms of copper oxide nanoparticles in epithelial kidney cells. Toxicol. In Vitro 2015, 29, 1053–1059. [Google Scholar] [CrossRef]
  36. Thubagere, A.; Reinhard, B.M. Nanoparticle-induced apoptosis propagates through hydrogen-peroxide-mediated bystander killing: Insights from a human intestinal epithelium in vitro model. ACS Nano 2010, 4, 3611–3622. [Google Scholar] [CrossRef]
  37. Gao, W.; Xu, K.; Ji, L.; Tang, B. Effect of gold nanoparticles on glutathione depletion-induced hydrogen peroxide generation and apoptosis in HL7702 cells. Toxicol. Lett. 2011, 205, 86–95. [Google Scholar] [CrossRef] [PubMed]
  38. Wang, J.X.; Fan, Y.B.; Gao, Y.; Hu, Q.H.; Wang, T.C. TiO2 nanoparticles translocation and potential toxicological effect in rats after intraarticular injection. Biomaterials 2009, 30, 4590–4600. [Google Scholar] [CrossRef] [PubMed]
  39. Guo, D.; Bi, H.; Liu, B.; Wu, Q.; Wang, D.; Cui, Y. Reactive oxygen species-induced cytotoxic effects of zinc oxide nanoparticles in rat retinal ganglion cells. Toxicol. In Vitro 2013, 27, 731–738. [Google Scholar] [CrossRef]
  40. Yang, E.J.; Kim, S.; Kim, J.S.; Choi, I.H. Inflammasome formation and IL-1beta release by human blood monocytes in response to silver nanoparticles. Biomaterials 2012, 33, 6858–6867. [Google Scholar] [CrossRef] [PubMed]
  41. Hirakawa, K.; Hirano, T. Singlet oxygen generation photocatalyzed by TiO2 particles and its contribution to biomolecule damage. Chem. Lett. 2006, 35, 832–833. [Google Scholar] [CrossRef]
  42. Lee, S.H.; Jun, B.H. Silver Nanoparticles: Synthesis and Application for Nanomedicine. Int. J. Mol. Sci. 2019, 20, 865. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  43. Zhou, Z.; Song, J.; Tian, R.; Yang, Z.; Yu, G.; Lin, L.; Zhang, G.; Fan, W.; Zhang, F.; Niu, G.; et al. Activatable Singlet Oxygen Generation from Lipid Hydroperoxide Nanoparticles for Cancer Therapy. Angew. Chem. Int. Ed. Engl. 2017, 56, 6492–6496. [Google Scholar] [CrossRef]
  44. Abdal Dayem, A.; Hossain, M.K.; Lee, S.B.; Kim, K.; Saha, S.K.; Yang, G.M.; Choi, H.Y.; Cho, S.G. The Role of Reactive Oxygen Species (ROS) in the Biological Activities of Metallic Nanoparticles. Int. J. Mol. Sci. 2017, 18, 120. [Google Scholar] [CrossRef] [Green Version]
  45. Kermanizadeh, A.; Jantzen, K.; Ward, M.B.; Durhuus, J.A.; Juel Rasmussen, L.; Loft, S.; Moller, P. Nanomaterial-induced cell death in pulmonary and hepatic cells following exposure to three different metallic materials: The role of autophagy and apoptosis. Nanotoxicology 2017, 11, 184–200. [Google Scholar] [CrossRef] [PubMed]
  46. Ge, D.; Du, Q.; Ran, B.; Liu, X.; Wang, X.; Ma, X.; Cheng, F.; Sun, B. The neurotoxicity induced by engineered nanomaterials. Int. J. Nanomed. 2019, 14, 4167–4186. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  47. Fridovich, I. Biological effects of the superoxide radical. Arch. BioChem. Biophys. 1986, 247, 1–11. [Google Scholar] [CrossRef]
  48. Apel, K.; Hirt, H. Reactive oxygen species: Metabolism, oxidative stress, and signal transduction. Annu. Rev. Plant. Biol. 2004, 55, 373–399. [Google Scholar] [CrossRef] [Green Version]
  49. McIntyre, M.; Bohr, D.F.; Dominiczak, A.F. Endothelial function in hypertension: The role of superoxide anion. Hypertension 1999, 34, 539–545. [Google Scholar] [CrossRef]
  50. Bielski, B.H.J.; Cabelli, D.E. Superoxide and Hydroxyl Radical Chemistry in Aqueous Solution. Act. Oxyg. Chem. 1995, 2, 66–104. [Google Scholar]
  51. Ahsan, H.; Ali, A.; Ali, R. Oxygen free radicals and systemic autoimmunity. Clin. Exp. Immunol. 2003, 131, 398–404. [Google Scholar] [CrossRef]
  52. Perry, J.J.P.; Shin, D.S.; Getzoff, E.D.; Tainer, J.A. The structural biochemistry of the superoxide dismutases. Bba-Proteins Proteom. 2010, 1804, 245–262. [Google Scholar] [CrossRef] [Green Version]
  53. Borgstahl, G.E.O.; Oberley-Deegan, R.E. Superoxide Dismutases (SODs) and SOD Mimetics. Antioxidants 2018, 7, 156. [Google Scholar] [CrossRef] [Green Version]
  54. Landis, G.N.; Tower, J. Superoxide dismutase evolution and life span regulation. Mech. Ageing Dev. 2005, 126, 365–379. [Google Scholar] [CrossRef] [PubMed]
  55. Loschen, G.; Flohe, L.; Chance, B. Respiratory Chain Linked H2o2 Production in Pigeon Heart Mitochondria. FEBS Lett. 1971, 18, 261–264. [Google Scholar] [CrossRef] [Green Version]
  56. Loschen, G.; Azzi, A.; Richter, C.; Flohe, L. Superoxide Radicals as Precursors of Mitochondrial Hydrogen-Peroxide. FEBS Lett. 1974, 42, 68–72. [Google Scholar] [CrossRef] [Green Version]
  57. Wilson, D.F.; Erecinska, M.; Dutton, P.L. Thermodynamic Relationships in Mitochondrial Oxidative-Phosphorylation. Annu. Rev. Biophys. Bio 1974, 3, 203–230. [Google Scholar] [CrossRef] [PubMed]
  58. Ballard, J.W.; Youngson, N.A. Review: Can diet influence the selective advantage of mitochondrial DNA haplotypes? Biosci. Rep. 2015, 35. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  59. Liu, Y.B.; Fiskum, G.; Schubert, D. Generation of reactive oxygen species by the mitochondrial electron transport chain. J. NeuroChem. 2002, 80, 780–787. [Google Scholar] [CrossRef] [PubMed]
  60. Kushnareva, Y.; Murphy, A.N.; Andreyev, A. Complex I-mediated reactive oxygen species generation: Modulation by cytochrome c and NAD(P)+ oxidation-reduction state. Biochem. J. 2002, 368, 545–553. [Google Scholar] [CrossRef] [Green Version]
  61. Murphy, M.P. How mitochondria produce reactive oxygen species. Biochem. J. 2009, 417, 1–13. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  62. Wikstrom, M.K.; Berden, J.A. Oxidoreduction of cytochrome b in the presence of antimycin. Biochim. Biophys. Acta 1972, 283, 403–420. [Google Scholar] [CrossRef]
  63. Muller, F.; Crofts, A.R.; Kramer, D.M. Multiple Q-cycle bypass reactions at the Qo site of the cytochiome bc1 complex. Biochemistry 2002, 41, 7866–7874. [Google Scholar] [CrossRef] [PubMed]
  64. Bleier, L.; Drose, S. Superoxide generation by complex III: From mechanistic rationales to functional consequences. Biochim. Biophys. Acta 2013, 1827, 1320–1331. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  65. Grzelak, A.; Wojewodzka, M.; Meczynska-Wielgosz, S.; Zuberek, M.; Wojciechowska, D.; Kruszewski, M. Crucial role of chelatable iron in silver nanoparticles induced DNA damage and cytotoxicity. Redox Biol. 2018, 15, 435–440. [Google Scholar] [CrossRef] [PubMed]
  66. Jayaram, D.T.; Payne, C.K. Intracellular Generation of Superoxide by TiO2 Nanoparticles Decreases Histone Deacetylase 9 (HDAC9), an Epigenetic Modifier. Bioconjug. Chem. 2020, 31, 1354–1361. [Google Scholar] [CrossRef]
  67. Masoud, R.; Bizouarn, T.; Trepout, S.; Wien, F.; Baciou, L.; Marco, S.; Houee Levin, C. Titanium Dioxide Nanoparticles Increase Superoxide Anion Production by Acting on NADPH Oxidase. PLoS ONE 2015, 10, e0144829. [Google Scholar] [CrossRef] [Green Version]
  68. Akhtar, M.J.; Kumar, S.; Alhadlaq, H.A.; Alrokayan, S.A.; Abu-Salah, K.M.; Ahamed, M. Dose-dependent genotoxicity of copper oxide nanoparticles stimulated by reactive oxygen species in human lung epithelial cells. Toxicol. Ind. Health 2016, 32, 809–821. [Google Scholar] [CrossRef] [PubMed]
  69. Piret, J.P.; Jacques, D.; Audinot, J.N.; Mejia, J.; Boilan, E.; Noel, F.; Fransolet, M.; Demazy, C.; Lucas, S.; Saout, C.; et al. Copper (II) oxide nanoparticles penetrate into HepG2 cells, exert cytotoxicity via oxidative stress and induce pro-inflammatory response. Nanoscale 2012, 4, 7168–7184. [Google Scholar] [CrossRef]
  70. Piao, M.J.; Kang, K.A.; Lee, I.K.; Kim, H.S.; Kim, S.; Choi, J.Y.; Choi, J.; Hyun, J.W. Silver nanoparticles induce oxidative cell damage in human liver cells through inhibition of reduced glutathione and induction of mitochondria-involved apoptosis. Toxicol. Lett. 2011, 201, 92–100. [Google Scholar] [CrossRef] [PubMed]
  71. Zielonka, J.; Srinivasan, S.; Hardy, M.; Ouari, O.; Lopez, M.; Vasquez-Vivar, J.; Avadhani, N.G.; Kalyanaraman, B. Cytochrome c-mediated oxidation of hydroethidine and mito-hydroethidine in mitochondria: Identification of homo- and heterodimers. Free Radic. Biol. Med. 2008, 44, 835–846. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  72. Ross, M.F.; Kelso, G.F.; Blaikie, F.H.; James, A.M.; Cocheme, H.M.; Filipovska, A.; Da Ros, T.; Hurd, T.R.; Smith, R.A.J.; Murphy, M.P. Lipophilic triphenylphosphonium cations as tools in mitochondrial bioenergetics and free radical biology. Biochemistry 2005, 70, 222–230. [Google Scholar] [CrossRef] [PubMed]
  73. Robinson, K.M.; Janes, M.S.; Pehar, M.; Monette, J.S.; Ross, M.F.; Hagen, T.M.; Murphy, M.P.; Beckman, J.S. Selective fluorescent imaging of superoxide in vivo using ethidium-based probes. Proc. Natl. Acad. Sci. USA 2006, 103, 15038–15043. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  74. Kauffman, M.E.; Kauffman, M.K.; Traore, K.; Zhu, H.; Trush, M.A.; Jia, Z.; Li, Y.R. MitoSOX-Based Flow Cytometry for Detecting Mitochondrial ROS. React. Oxyg. Species 2016, 2, 361–370. [Google Scholar] [CrossRef] [Green Version]
  75. Mukhopadhyay, P.; Rajesh, M.; Yoshihiro, K.; Hasko, G.; Pacher, P. Simple quantitative detection of mitochondrial superoxide production in live cells. BioChem. Biophys. Res. Commun. 2007, 358, 203–208. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  76. Roelofs, B.A.; Ge, S.X.; Studlack, P.E.; Polster, B.M. Low micromolar concentrations of the superoxide probe MitoSOX uncouple neural mitochondria and inhibit complex IV. Free Radic. Biol. Med. 2015, 86, 250–258. [Google Scholar] [CrossRef] [Green Version]
  77. Ohyashiki, T.; Nunomura, M.; Katoh, T. Detection of superoxide anion radical in phospholipid liposomal membrane by fluorescence quenching method using 1,3-diphenylisobenzofuran. Bba-Biomembranes 1999, 1421, 131–139. [Google Scholar] [CrossRef] [Green Version]
  78. Krieg, M. Determination of singlet oxygen quantum yields with 1,3-diphenylisobenzofuran in model membrane systems. J. BioChem. Biophys. Methods 1993, 27, 143–149. [Google Scholar] [CrossRef]
  79. Zamojc, K.; Zdrowowicz, M.; Rudnicki-Velasquez, P.B.; Krzyminski, K.; Zaborowski, B.; Niedzialkowski, P.; Jacewicz, D.; Chmurzynski, L. The development of 1,3-diphenylisobenzofuran as a highly selective probe for the detection and quantitative determination of hydrogen peroxide. Free Radic. Res. 2017, 51, 38–46. [Google Scholar] [CrossRef]
  80. Andresen, M.; Regueira, T.; Bruhn, A.; Perez, D.; Strobel, P.; Dougnac, A.; Marshall, G.; Leighton, F. Lipoperoxidation and protein oxidative damage exhibit different kinetics during septic shock. Mediat. Inflamm. 2008. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  81. Wu, D.F.; Cederbaum, A.I. Alcohol, oxidative stress, and free radical damage. Alcohol. Res. Health 2003, 27, 277–284. [Google Scholar] [PubMed]
  82. Halliwell, B.; Chirico, S. Lipid-Peroxidation ― Its Mechanism, Measurement, and Significance. Am. J. Clin. Nutr. 1993, 57, 715–725. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  83. Wang, X.; Zhang, L. Kinetic study of hydroxyl radical formation in a continuous hydroxyl generation system. RSC Adv. 2018, 8, 40632–40638. [Google Scholar] [CrossRef] [Green Version]
  84. Kehrer, J.P. The Haber-Weiss reaction and mechanisms of toxicity. Toxicology 2000, 149, 43–50. [Google Scholar] [CrossRef]
  85. Weinstein, J.; Bielski, B.H.J. Kinetics of the Interaction of Ho2 and O2-Radicals with Hydrogen-Peroxide—Haber-Weiss Reaction. J. Am. Chem. Soc. 1979, 101, 58–62. [Google Scholar] [CrossRef]
  86. Koppenol, W.H. The Haber-Weiss cycle—70 years later. Redox Rep. 2001, 6, 229–234. [Google Scholar] [CrossRef]
  87. Fischbacher, A.; von Sonntag, C.; Schmidt, T.C. Hydroxyl radical yields in the Fenton process under various pH, ligand concentrations and hydrogen peroxide/Fe (II) ratios. Chemosphere 2017, 182, 738–744. [Google Scholar] [CrossRef]
  88. Wang, T.; Zhang, H.; Liu, H.; Yuan, Q.; Ren, F.; Han, Y.; Sun, Q.; Li, Z.; Gao, M. Boosting H2O2-Guided Chemodynamic Therapy of Cancer by Enhancing Reaction Kinetics through Versatile Biomimetic Fenton Nanocatalysts and the Second Near-Infrared Light Irradiation. Adv. Funct. Mater. 2019, 30. [Google Scholar] [CrossRef]
  89. Li, X.; Hao, S.J.; Han, A.L.; Yang, Y.Y.; Fang, G.Z.; Liu, J.F.; Wang, S. Intracellular Fenton reaction based on mitochondria-targeted copper (II)-peptide complex for induced apoptosis. J. Mater. Chem. B 2019, 7, 4008–4016. [Google Scholar] [CrossRef]
  90. Hackenberg, S.; Scherzed, A.; Technau, A.; Kessler, M.; Froelich, K.; Ginzkey, C.; Koehler, C.; Burghartz, M.; Hagen, R.; Kleinsasser, N. Cytotoxic, genotoxic and pro-inflammatory effects of zinc oxide nanoparticles in human nasal mucosa cells in vitro. Toxicol. In Vitro 2011, 25, 657–663. [Google Scholar] [CrossRef]
  91. Ekici, S.; Turkarslan, S.; Pawlik, G.; Dancis, A.; Baliga, N.S.; Koch, H.G.; Daldal, F. Intracytoplasmic copper homeostasis controls cytochrome c oxidase production. mBio 2014, 5. [Google Scholar] [CrossRef] [Green Version]
  92. Huang, G.; Chen, H.; Dong, Y.; Luo, X.; Yu, H.; Moore, Z.; Bey, E.A.; Boothman, D.A.; Gao, J. Superparamagnetic iron oxide nanoparticles: Amplifying ROS stress to improve anticancer drug efficacy. Theranostics 2013, 3, 116–126. [Google Scholar] [CrossRef] [PubMed]
  93. Lehman, S.E.; Morris, A.S.; Mueller, P.S.; Salem, A.K.; Grassian, V.H.; Larsen, S.C. Silica nanoparticle-generated ROS as a predictor of cellular toxicity: Mechanistic insights and safety by design. Environ. Sci-Nano 2016, 3, 56–66. [Google Scholar] [CrossRef] [Green Version]
  94. Chairuangkitti, P.; Lawanprasert, S.; Roytrakul, S.; Aueviriyavit, S.; Phummiratch, D.; Kulthong, K.; Chanvorachote, P.; Maniratanachote, R. Silver nanoparticles induce toxicity in A549 cells via ROS-dependent and ROS-independent pathways. Toxicol. In Vitro 2013, 27, 330–338. [Google Scholar] [CrossRef] [PubMed]
  95. Fang, X.W.; Mark, G.; von Sonntag, C. OH radical formation by ultrasound in aqueous solutions Part І: The chemistry underlying the terephthalate dosimeter. Ultrason. SonoChem. 1996, 3, 57–63. [Google Scholar] [CrossRef]
  96. Yan, E.B.; Unthank, J.K.; Castillo-Melendez, M.; Miller, S.L.; Langford, S.J.; Walker, D.W. Novel method for in vivo hydroxyl radical measurement by microdialysis in fetal sheep brain in utero. J. Appl. Physiol. 2005, 98, 2304–2310. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  97. Yapici, N.B.; Jockusch, S.; Moscatelli, A.; Mandalapu, S.R.; Itagaki, Y.; Bates, D.K.; Wiseman, S.; Gibson, K.M.; Turro, N.J.; Bi, L.R. New Rhodamine Nitroxide Based Fluorescent Probes for Intracellular Hydroxyl Radical Identification in Living Cells. Org. Lett. 2012, 14, 50–53. [Google Scholar] [CrossRef]
  98. Bai, X.Y.; Huang, Y.Y.; Lu, M.Y.; Yang, D. HKOH-1: A Highly Sensitive and Selective Fluorescent Probe for Detecting Endogenous Hydroxyl Radicals in Living Cells. Angew. Chem. Int. Ed. 2017, 56, 12873–12877. [Google Scholar] [CrossRef]
  99. Cheng, X.; Guo, H.; Zhang, Y.; Wu, X.; Liu, Y. Non-photochemical production of singlet oxygen via activation of persulfate by carbon nanotubes. Water Res. 2017, 113, 80–88. [Google Scholar] [CrossRef] [PubMed]
  100. Cadenas, E. Biochemistry of Oxygen-Toxicity. Annu. Rev. BioChem. 1989, 58, 79–110. [Google Scholar] [CrossRef] [PubMed]
  101. Agnez-Lima, L.F.; Melo, J.T.A.; Silva, A.E.; Oliveira, A.H.S.; Timoteo, A.R.S.; Lima-Bessa, K.M.; Martinez, G.R.; Medeiros, M.H.G.; Di Mascio, P.; Galhardo, R.S.; et al. DNA damage by singlet oxygen and cellular protective mechanisms. Mutat. Res. Rev. Mutat. 2012, 751, 15–28. [Google Scholar] [CrossRef] [PubMed]
  102. Hampton, M.B.; Kettle, A.J.; Winterbourn, C.C. Inside the neutrophil phagosome: Oxidants, myeloperoxidase, and bacterial killing. Blood 1998, 92, 3007–3017. [Google Scholar] [CrossRef] [PubMed]
  103. Bigot, E.; Bataille, R.; Patrice, T. Increased singlet oxygen-induced secondary ROS production in the serum of cancer patients. J. PhotoChem. PhotoBiol. B 2012, 107, 14–19. [Google Scholar] [CrossRef]
  104. Sies, H.; Menck, C.F.M. Singlet Oxygen Induced DNA Damage. Mutat Res. 1992, 275, 367–375. [Google Scholar] [CrossRef]
  105. Kanofsky, J.R. Singlet Oxygen Production by Biological-Systems. Chem. Biol. Interact. 1989, 70, 1–28. [Google Scholar] [CrossRef]
  106. Dumont, E.; Gruber, R.; Bignon, E.; Morell, C.; Moreau, Y.; Monari, A.; Ravanat, J.L. Probing the reactivity of singlet oxygen with purines. Nucleic Acids Res. 2016, 44, 56–62. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  107. Davies, M.J. Singlet oxygen-mediated damage to proteins and its consequences. BioChem. Biophys. Res. Commun. 2003, 305, 761–770. [Google Scholar] [CrossRef]
  108. Gracanin, M.; Hawkins, C.L.; Pattison, D.I.; Davies, M.J. Singlet-oxygen-mediated amino acid and protein oxidation: Formation of tryptophan peroxides and decomposition products. Free Radic. Bio Med. 2009, 47, 92–102. [Google Scholar] [CrossRef]
  109. Hirakawa, T.; Nosaka, Y. Properties of O2.− and OH center dot formed in TiO2 aqueous suspensions by photocatalytic reaction and the influence of H2O2 and some ions. Langmuir 2002, 18, 3247–3254. [Google Scholar] [CrossRef]
  110. Kim, S.Y.; Lee, S.M.; Park, J.W. Antioxidant enzyme inhibitors enhance singlet oxygen-induced cell death in HL-60 cells. Free Radic. Res. 2006, 40, 1190–1197. [Google Scholar] [CrossRef] [PubMed]
  111. Deng, J.; Liu, F.; Wang, L.; An, Y.; Gao, M.; Wang, Z.; Zhao, Y. Hypoxia- and singlet oxygen-responsive chemo-photodynamic Micelles featured with glutathione depletion and aldehyde production. Biomater. Sci. 2018, 7, 429–441. [Google Scholar] [CrossRef]
  112. Kim, S.Y.; Lee, S.M.; Tak, J.K.; Choi, K.S.; Kwon, T.K.; Park, J.W. Regulation of singlet oxygen-induced apoptosis by cytosolic NADP+-dependent isocitrate dehydrogenase. Mol. Cell Biochem. 2007, 302, 27–34. [Google Scholar] [CrossRef]
  113. Umezawa, N.; Tanaka, K.; Urano, Y.; Kikuchi, K.; Higuchi, T.; Nagano, T. Novel Fluorescent Probes for Singlet Oxygen. Angew. Chem. Int. Ed. Engl. 1999, 38, 2899–2901. [Google Scholar] [CrossRef]
  114. Brega, V.; Yan, Y.; Thomas, S.W., 3rd. Acenes beyond organic electronics: Sensing of singlet oxygen and stimuli-responsive materials. Org. Biomol. Chem. 2020, 18, 9191–9209. [Google Scholar] [CrossRef] [PubMed]
  115. Ruiz-Gonzalez, R.; Bresoli-Obach, R.; Gulias, O.; Agut, M.; Savoie, H.; Boyle, R.W.; Nonell, S.; Giuntini, F. NanoSOSG: A Nanostructured Fluorescent Probe for the Detection of Intracellular Singlet Oxygen. Angew. Chem. Int. Ed. Engl. 2017, 56, 2885–2888. [Google Scholar] [CrossRef]
  116. Lennicke, C.; Rahn, J.; Lichtenfels, R.; Wessjohann, L.A.; Seliger, B. Hydrogen peroxide—Production, fate and role in redox signaling of tumor cells. Cell Commun. Signal. 2015, 13, 39. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  117. Hossain, M.A.; Bhattacharjee, S.; Armin, S.M.; Qian, P.; Xin, W.; Li, H.Y.; Burritt, D.J.; Fujita, M.; Tran, L.S. Hydrogen peroxide priming modulates abiotic oxidative stress tolerance: Insights from ROS detoxification and scavenging. Front. Plant. Sci. 2015, 6, 420. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  118. Halliwell, B.; Clement, M.V.; Long, L.H. Hydrogen peroxide in the human body. FEBS Lett. 2000, 486, 10–13. [Google Scholar] [CrossRef] [Green Version]
  119. Mates, J.M.; Perez-Gomez, C.; De Castro, I.N. Antioxidant enzymes and human diseases. Clin. Biochem. 1999, 32, 595–603. [Google Scholar] [CrossRef]
  120. Angermuller, S.; Islinger, M.; Volkl, A. Peroxisomes and reactive oxygen species, a lasting challenge. HistoChem. Cell Biol. 2009, 131, 459–463. [Google Scholar] [CrossRef]
  121. Topo, E.; Fisher, G.; Sorricelli, A.; Errico, F.; Usiello, A.; D’Aniello, A. Thyroid hormones and D-aspartic acid, D-aspartate oxidase, D-aspartate racemase, H2O2, and ROS in rats and mice. Chem. Biodivers. 2010, 7, 1467–1478. [Google Scholar] [CrossRef] [PubMed]
  122. Royall, J.A.; Ischiropoulos, H. Evaluation of 2’,7’-Dichlorofluorescin and Dihydrorhodamine 123 as Fluorescent-Probes for Intracellular H2o2 in Cultured Endothelial-Cells. Arch. Biochem. Biophys. 1993, 302, 348–355. [Google Scholar] [CrossRef] [PubMed]
  123. Rastogi, R.P.; Singh, S.P.; Hader, D.P.; Sinha, R.P. Detection of reactive oxygen species (ROS) by the oxidant-sensing probe 2’,7’-dichlorodihydrofluorescein diacetate in the cyanobacterium Anabaena variabilis PCC 7937. Biochem. Biophys. Res. Commun. 2010, 397, 603–607. [Google Scholar] [CrossRef] [PubMed]
  124. Gomes, A.; Fernandes, E.; Lima, J.L.F.C. Fluorescence probes used for detection of reactive oxygen species. J. Biochem. Bioph Meth. 2005, 65, 45–80. [Google Scholar] [CrossRef]
  125. Crow, J.P. Dichlorodihydrofluorescein and dihydrorhodamine 123 are sensitive indicators of peroxynitrite in vitro: Implications for intracellular measurement of reactive nitrogen and oxygen species. Nitric Oxide 1997, 1, 145–157. [Google Scholar] [CrossRef]
  126. Chignell, C.F.; Sik, R.H. A photochemical study of cells loaded with 2’,7’-dichlorofluorescin: Implications for the detection of reactive oxygen species generated during UVA irradiation. Free Radic. Biol. Med. 2003, 34, 1029–1034. [Google Scholar] [PubMed]
  127. Zhu, H.; Bannenberg, G.L.; Moldeus, P.; Shertzer, H.G. Oxidation pathways for the intracellular probe 2’,7’-dichlorofluorescein. Arch. Toxicol. 1994, 68, 582–587. [Google Scholar] [CrossRef]
  128. Lebel, C.P.; Ischiropoulos, H.; Bondy, S.C. Evaluation of the Probe 2’,7’-Dichlorofluorescin as an Indicator of Reactive Oxygen Species Formation and Oxidative Stress. Chem. Res. Toxicol. 1992, 5, 227–231. [Google Scholar] [CrossRef] [Green Version]
  129. Hsiao, I.L.; Huang, Y.J. Titanium Oxide Shell Coatings Decrease the Cytotoxicity of ZnO Nanoparticles. Chem. Res. Toxicol. 2011, 24, 303–313. [Google Scholar] [CrossRef] [PubMed]
  130. Akhtar, M.J.; Ahamed, M.; Kumar, S.; Khan, M.M.; Ahmad, J.; Alrokayan, S.A. Zinc oxide nanoparticles selectively induce apoptosis in human cancer cells through reactive oxygen species. Int. J. Nanomed. 2012, 7, 845–857. [Google Scholar]
  131. Sharma, V.; Anderson, D.; Dhawan, A. Zinc oxide nanoparticles induce oxidative DNA damage and ROS-triggered mitochondria mediated apoptosis in human liver cells (HepG2). Apoptosis 2012, 17, 852–870. [Google Scholar] [CrossRef] [PubMed]
  132. Setyawati, M.I.; Tay, C.Y.; Leong, D.T. Effect of zinc oxide nanomaterials-induced oxidative stress on the p53 pathway. Biomaterials 2013, 34, 10133–10142. [Google Scholar] [CrossRef]
  133. Aliakbari, F.; Haji Hosseinali, S.; Khalili Sarokhalil, Z.; Shahpasand, K.; Akbar Saboury, A.; Akhtari, K.; Falahati, M. Reactive oxygen species generated by titanium oxide nanoparticles stimulate the hemoglobin denaturation and cytotoxicity against human lymphocyte cell. J. Biomol. Struct. Dyn. 2019, 37, 4875–4881. [Google Scholar] [CrossRef] [PubMed]
  134. Bhattacharya, K.; Davoren, M.; Boertz, J.; Schins, R.P.; Hoffmann, E.; Dopp, E. Titanium dioxide nanoparticles induce oxidative stress and DNA-adduct formation but not DNA-breakage in human lung cells. Part. Fibre Toxicol. 2009, 6, 17. [Google Scholar] [CrossRef] [Green Version]
  135. Liu, S.; Xu, L.; Zhang, T.; Ren, G.; Yang, Z. Oxidative stress and apoptosis induced by nanosized titanium dioxide in PC12 cells. Toxicology 2010, 267, 172–177. [Google Scholar] [CrossRef] [PubMed]
  136. Park, E.J.; Yi, J.; Chung, K.H.; Ryu, D.Y.; Choi, J.; Park, K. Oxidative stress and apoptosis induced by titanium dioxide nanoparticles in cultured BEAS-2B cells. Toxicol. Lett. 2008, 180, 222–229. [Google Scholar] [CrossRef] [PubMed]
  137. Miwa, S.; Treumann, A.; Bell, A.; Vistoli, G.; Nelson, G.; Hay, S.; von Zglinicki, T. Carboxylesterase converts Amplex red to resorufin: Implications for mitochondrial H2O2 release assays. Free Radic. Biol. Med. 2016, 90, 173–183. [Google Scholar] [CrossRef] [PubMed]
  138. Zhu, A.; Romero, R.; Petty, H.R. A sensitive fluorimetric assay for pyruvate. Anal. BioChem. 2010, 396, 146–151. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  139. Towne, V.; Will, M.; Oswald, B.; Zhao, Q.J. Complexities in horseradish peroxidase-catalyzed oxidation of dihydroxyphenoxazine derivatives: Appropriate ranges for pH values and hydrogen peroxide concentrations in quantitative analysis. Anal. Biochem. 2004, 334, 290–296. [Google Scholar] [CrossRef] [PubMed]
  140. Debski, D.; Smulik, R.; Zielonka, J.; Michalowski, B.; Jakubowska, M.; Debowska, K.; Adamus, J.; Marcinek, A.; Kalyanaraman, B.; Sikora, A. Mechanism of oxidative conversion of Amplex (R) Red to resorufin: Pulse radiolysis and enzymatic studies. Free Radic. Bio Med. 2016, 95, 323–332. [Google Scholar] [CrossRef] [Green Version]
  141. Niethammer, P.; Grabher, C.; Look, A.T.; Mitchison, T.J. A tissue-scale gradient of hydrogen peroxide mediates rapid wound detection in zebrafish. Nature 2009, 459, 996–999. [Google Scholar] [CrossRef] [PubMed]
  142. Maeda, H.; Fukuyasu, Y.; Yoshida, S.; Fukuda, M.; Saeki, K.; Matsuno, H.; Yamauchi, Y.; Yoshida, K.; Hirata, K.; Miyamoto, K. Fluorescent probes for hydrogen peroxide based on a non-oxidative mechanism. Angew. Chem. Int. Ed. Engl. 2004, 43, 2389–2391. [Google Scholar] [CrossRef] [PubMed]
  143. Wolfbeis, O.S.; Durkop, A.; Wu, M.; Lin, Z.H. A europium-ion-based luminescent sensing probe for hydrogen peroxide. Angew. Chem. Int. Ed. 2002, 41, 4495–4498. [Google Scholar] [CrossRef]
  144. Staniek, K.; Nohl, H. H2O2 detection from intact mitochondria as a measure for one-electron reduction of dioxygen requires a non-invasive assay system. Bba-Bioenergetics 1999, 1413, 70–80. [Google Scholar] [CrossRef] [Green Version]
  145. Bartosz, G. Use of spectroscopic probes for detection of reactive oxygen species. Clin. Chim. Acta 2006, 368, 53–76. [Google Scholar] [CrossRef] [PubMed]
  146. Mohammadinejad, R.; Moosavi, M.A.; Tavakol, S.; Vardar, D.O.; Hosseini, A.; Rahmati, M.; Dini, L.; Hussain, S.; Mandegary, A.; Klionsky, D.J. Necrotic, apoptotic and autophagic cell fates triggered by nanoparticles. Autophagy 2019, 15, 4–33. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  147. Mittler, R.; Vanderauwera, S.; Suzuki, N.; Miller, G.; Tognetti, V.B.; Vandepoele, K.; Gollery, M.; Shulaev, V.; Van Breusegem, F. ROS signaling: The new wave? Trends Plant. Sci. 2011, 16, 300–309. [Google Scholar] [CrossRef] [PubMed]
  148. Zhang, J.; Wang, X.; Vikash, V.; Ye, Q.; Wu, D.; Liu, Y.; Dong, W. ROS and ROS-Mediated Cellular Signaling. Oxid Med. Cell Longev. 2016. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  149. Bae, Y.S.; Oh, H.; Rhee, S.G.; Yoo, Y.D. Regulation of reactive oxygen species generation in cell signaling. Mol. Cells 2011, 32, 491–509. [Google Scholar] [CrossRef] [Green Version]
  150. Bonizzi, G.; Karin, M. The two NF-kappaB activation pathways and their role in innate and adaptive immunity. Trends Immunol. 2004, 25, 280–288. [Google Scholar] [CrossRef]
  151. Kaul, N.; Gopalakrishna, R.; Gundimeda, U.; Choi, J.; Forman, H.J. Role of protein kinase C in basal and hydrogen peroxide-stimulated NF-kappa B activation in the murine macrophage J774A.1 cell line. Arch. Biochem. Biophys. 1998, 350, 79–86. [Google Scholar] [CrossRef] [PubMed]
  152. Schmidt, K.N.; Amstad, P.; Cerutti, P.; Baeuerle, P.A. The roles of hydrogen peroxide and superoxide as messengers in the activation of transcription factor NF-kappa B. Chem. Biol. 1995, 2, 13–22. [Google Scholar] [CrossRef] [Green Version]
  153. Schoonbroodt, S.; Ferreira, V.; Best-Belpomme, M.; Boelaert, J.R.; Legrand-Poels, S.; Korner, M.; Piette, J. Crucial role of the amino-terminal tyrosine residue 42 and the carboxyl-terminal PEST domain of I kappa B alpha in NF-kappa B activation by an oxidative stress. J. Immunol. 2000, 164, 4292–4300. [Google Scholar] [CrossRef] [PubMed]
  154. Takada, Y.; Mukhopadhyay, A.; Kundu, G.C.; Mahabeleshwar, G.H.; Singh, S.; Aggarwal, B.B. Hydrogen peroxide activates NF-kappa B through tyrosine phosphorylation of I kappa B alpha and serine phosphorylation of p65: Evidence for the involvement of I kappa B alpha kinase and Syk protein-tyrosine kinase. J. Biol. Chem. 2003, 278, 24233–24241. [Google Scholar] [CrossRef] [Green Version]
  155. Liu, X.; Lu, B.; Fu, J.; Zhu, X.; Song, E.; Song, Y. Amorphous silica nanoparticles induce inflammation via activation of NLRP3 inflammasome and HMGB1/TLR4/MYD88/NF-kb signaling pathway in HUVEC cells. J. Hazard. Mater. 2021, 404, 124050. [Google Scholar] [CrossRef] [PubMed]
  156. Kyriakis, J.M.; Avruch, J. Sounding the alarm: Protein kinase cascades activated by stress and inflammation. J. Biol. Chem. 1996, 271, 24313–24316. [Google Scholar] [CrossRef] [Green Version]
  157. Nakano, H.; Nakajima, A.; Sakon-Komazawa, S.; Piao, J.H.; Xue, X.; Okumura, K. Reactive oxygen species mediate crosstalk between NF-kappaB and JNK. Cell Death Differ. 2006, 13, 730–737. [Google Scholar] [CrossRef]
  158. Torres, M.; Forman, H.J. Redox signaling and the MAP kinase pathways. Biofactors 2003, 17, 287–296. [Google Scholar] [CrossRef] [PubMed]
  159. Dabrowski, A.; Boguslowicz, C.; Dabrowska, M.; Tribillo, I.; Gabryelewicz, A. Reactive oxygen species activate mitogen-activated protein kinases in pancreatic acinar cells. Pancreas 2000, 21, 376–384. [Google Scholar] [CrossRef]
  160. Guyton, K.Z.; Liu, Y.; Gorospe, M.; Xu, Q.; Holbrook, N.J. Activation of mitogen-activated protein kinase by H2O2. Role in cell survival following oxidant injury. J. Biol. Chem. 1996, 271, 4138–4142. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  161. Hou, N.; Torii, S.; Saito, N.; Hosaka, M.; Takeuchi, T. Reactive oxygen species-mediated pancreatic beta-cell death is regulated by interactions between stress-activated protein kinases, p38 and c-Jun N-terminal kinase, and mitogen-activated protein kinase phosphatases. Endocrinology 2008, 149, 1654–1665. [Google Scholar] [CrossRef] [PubMed]
  162. Choi, B.H.; Hur, E.M.; Lee, J.H.; Jun, D.J.; Kim, K.T. Protein kinase Cdelta-mediated proteasomal degradation of MAP kinase phosphatase-1 contributes to glutamate-induced neuronal cell death. J. Cell Sci. 2006, 119, 1329–1340. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  163. Matsuzawa, A.; Saegusa, K.; Noguchi, T.; Sadamitsu, C.; Nishitoh, H.; Nagai, S.; Koyasu, S.; Matsumoto, K.; Takeda, K.; Ichijo, H. ROS-dependent activation of the TRAF6-ASK1-p38 pathway is selectively required for TLR4-mediated innate immunity. Nat. Immunol. 2005, 6, 587–592. [Google Scholar] [CrossRef] [PubMed]
  164. Pitzschke, A.; Djamei, A.; Bitton, F.; Hirt, H. A Major Role of the MEKK1-MKK1/2-MPK4 Pathway in ROS Signalling. Mol. Plant. 2009, 2, 120–137. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  165. Lluis, J.M.; Buricchi, F.; Chiarugi, P.; Morales, A.; Fernandez-Checa, J.C. Dual role of mitochondrial reactive oxygen species in hypoxia signaling: Activation of nuclear factor-kB via c-SRC and oxidant-dependent cell death. Cancer Res. 2007, 67, 7368–7377. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  166. Dong, J.; Ramachandiran, S.; Tikoo, K.; Jia, Z.; Lau, S.S.; Monks, T.J. EGFR-independent activation of p38 MAPK and EGFR-dependent activation of ERK1/2 are required for ROS-induced renal cell death. Am. J. Physiol. Renal Physiol. 2004, 287, 1049–1058. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  167. Forman, H.J.; Torres, M. Reactive oxygen species and cell signaling: Respiratory burst in macrophage signaling. Am. J. Respir. Crit. Care Med. 2002, 166. [Google Scholar] [CrossRef]
  168. Cheng, G.; Guo, W.; Han, L.; Chen, E.; Kong, L.; Wang, L.; Ai, W.; Song, N.; Li, H.; Chen, H. Cerium oxide nanoparticles induce cytotoxicity in human hepatoma SMMC-7721 cells via oxidative stress and the activation of MAPK signaling pathways. Toxicol. In Vitro 2013, 27, 1082–1088. [Google Scholar] [CrossRef] [PubMed]
  169. Guo, C.; Xia, Y.; Niu, P.; Jiang, L.; Duan, J.; Yu, Y.; Zhou, X.; Li, Y.; Sun, Z. Silica nanoparticles induce oxidative stress, inflammation, and endothelial dysfunction in vitro via activation of the MAPK/Nrf2 pathway and nuclear factor-kappaB signaling. Int. J. Nanomed. 2015, 10, 1463–1477. [Google Scholar] [CrossRef] [Green Version]
  170. You, R.; Ho, Y.S.; Hung, C.H.; Liu, Y.; Huang, C.X.; Chan, H.N.; Ho, S.L.; Lui, S.Y.; Li, H.W.; Chang, R.C. Silica nanoparticles induce neurodegeneration-like changes in behavior, neuropathology, and affect synapse through MAPK activation. Part. Fibre Toxicol. 2018, 15, 28. [Google Scholar] [CrossRef] [Green Version]
  171. Hu, Q.; Wang, H.; He, C.; Jin, Y.; Fu, Z. Polystyrene nanoparticles trigger the activation of p38 MAPK and apoptosis via inducing oxidative stress in zebrafish and macrophage cells. Environ. Pollut. 2021, 269, 116075. [Google Scholar] [CrossRef] [PubMed]
  172. Zhou, Y.; Ji, J.; Ji, L.; Wang, L.; Hong, F. Respiratory exposure to nano-TiO2 induces pulmonary toxicity in mice involving reactive free radical-activated TGF-beta/Smad/p38MAPK/Wnt pathways. J. Biomed. Mater. Res. A 2019, 107, 2567–2575. [Google Scholar] [CrossRef] [PubMed]
  173. Yi, C.; Liu, D.; Fong, C.C.; Zhang, J.; Yang, M. Gold nanoparticles promote osteogenic differentiation of mesenchymal stem cells through p38 MAPK pathway. ACS Nano 2010, 4, 6439–6448. [Google Scholar] [CrossRef] [PubMed]
  174. Wang, Q.; Chen, B.; Cao, M.; Sun, J.; Wu, H.; Zhao, P.; Xing, J.; Yang, Y.; Zhang, X.; Ji, M.; et al. Response of MAPK pathway to iron oxide nanoparticles in vitro treatment promotes osteogenic differentiation of hBMSCs. Biomaterials 2016, 86, 11–20. [Google Scholar] [CrossRef] [PubMed]
  175. Vousden, K.H.; Lu, X. Live or let die: The cell’s response to p53. Nat. Rev. Cancer 2002, 2, 594–604. [Google Scholar] [CrossRef] [Green Version]
  176. Liu, B.; Chen, Y.; St Clair, D.K. ROS and p53: A versatile partnership. Free Radic. Biol. Med. 2008, 44, 1529–1535. [Google Scholar] [CrossRef] [Green Version]
  177. Song, Y.X.; Li, X.W.; Li, Y.; Li, N.; Shi, X.X.; Ding, H.Y.; Zhang, Y.H.; Li, X.B.; Liu, G.W.; Wang, Z. Non-esterified fatty acids activate the ROS-p38-p53/Nrf2 signaling pathway to induce bovine hepatocyte apoptosis in vitro. Apoptosis 2014, 19, 984–997. [Google Scholar] [CrossRef] [PubMed]
  178. Nakano, K.; Vousden, K.H. PUMA, a novel proapoptotic gene, is induced by p53. Mol. Cell 2001, 7, 683–694. [Google Scholar] [CrossRef]
  179. Liu, B.R.; Yuan, B.; Zhang, L.; Mu, W.M.; Wang, C.M. ROS/p38/p53/Puma signaling pathway is involved in emodin-induced apoptosis of human colorectal cancer cells. Int. J. Clin. Exp. Med. 2015, 8, 15413–15422. [Google Scholar] [PubMed]
  180. Yu, J.; Zhang, L. PUMA, a potent killer with or without p53. Oncogene 2008, 27, 71–83. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  181. Samuelsen, J.T.; Dahl, J.E.; Karlsson, S.; Morisbak, E.; Becher, R. Apoptosis induced by the monomers HEMA and TEGDMA involves formation of ROS and differential activation of the MAP-kinases p38, JNK and ERK. Dent. Mater. 2007, 23, 34–39. [Google Scholar] [CrossRef]
  182. Sakon, S.; Xue, X.; Takekawa, M.; Sasazuki, T.; Okazaki, T.; Kojima, Y.; Piao, J.H.; Yagita, H.; Okumura, K.; Doi, T.; et al. NF-kappaB inhibits TNF-induced accumulation of ROS that mediate prolonged MAPK activation and necrotic cell death. Embo J. 2003, 22, 3898–3909. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  183. Kamata, H.; Honda, S.; Maeda, S.; Chang, L.; Hirata, H.; Karin, M. Reactive oxygen species promote TNFalpha-induced death and sustained JNK activation by inhibiting MAP kinase phosphatases. Cell 2005, 120, 649–661. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  184. Akcan, R.; Aydogan, H.C.; Yildirim, M.S.; Tastekin, B.; Saglam, N. Nanotoxicity: A challenge for future medicine. Turk. J. Med. Sci. 2020, 50, 1180–1196. [Google Scholar] [CrossRef] [PubMed]
  185. Graham, U.M.; Dozier, A.K.; Oberdorster, G.; Yokel, R.A.; Molina, R.; Brain, J.D.; Pinto, J.M.; Weuve, J.; Bennett, D.A. Tissue Specific Fate of Nanomaterials by Advanced Analytical Imaging Techniques—A Review. Chem. Res. Toxicol. 2020, 33, 1145–1162. [Google Scholar] [CrossRef] [PubMed]
  186. Pu, S.; Gong, C.; Robertson, A.W. Liquid cell transmission electron microscopy and its applications. R. Soc. Open Sci. 2020, 7, 191204. [Google Scholar] [CrossRef] [Green Version]
  187. Kiio, T.M.; Park, S. Nano-scientific Application of Atomic Force Microscopy in Pathology: From Molecules to Tissues. Int. J. Med. Sci. 2020, 17, 844–858. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  188. Erofeev, A.; Gorelkin, P.; Garanina, A.; Alova, A.; Efremova, M.; Vorobyeva, N.; Edwards, C.; Korchev, Y.; Majouga, A. Novel method for rapid toxicity screening of magnetic nanoparticles. Sci. Rep. 2018, 8, 7462. [Google Scholar] [CrossRef] [Green Version]
  189. Rahman, L.; Williams, A.; Gelda, K.; Nikota, J.; Wu, D.; Vogel, U.; Halappanavar, S. 21st Century Tools for Nanotoxicology: Transcriptomic Biomarker Panel and Precision-Cut Lung Slice Organ Mimic System for the Assessment of Nanomaterial-Induced Lung Fibrosis. Small 2020, 16. [Google Scholar] [CrossRef]
  190. Kohl, Y.; Runden-Pran, E.; Mariussen, E.; Hesler, M.; El Yamani, N.; Longhin, E.M.; Dusinska, M. Genotoxicity of Nanomaterials: Advanced In Vitro Models and High Throughput Methods for Human Hazard Assessment-A Review. Nanomaterials (Basel) 2020, 10, 1911. [Google Scholar] [CrossRef] [PubMed]
  191. Zhang, M.; Xu, C.; Jiang, L.; Qin, J. A 3D human lung-on-a-chip model for nanotoxicity testing. Toxicol. Res. 2018, 7, 1048–1060. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  192. Yin, F.; Zhu, Y.; Zhang, M.; Yu, H.; Chen, W.; Qin, J. A 3D human placenta-on-a-chip model to probe nanoparticle exposure at the placental barrier. Toxicol. In Vitro 2019, 54, 105–113. [Google Scholar] [CrossRef] [PubMed]
  193. van Duinen, V.; Trietsch, S.J.; Joore, J.; Vulto, P.; Hankemeier, T. Microfluidic 3D cell culture: From tools to tissue models. Curr. Opin. Biotechnol. 2015, 35, 118–126. [Google Scholar] [CrossRef] [PubMed] [Green Version]
Figure 1. Formation of reactive oxygen species. Abbreviations: SOD = superoxide dismutase; MPO = myeloperoxidase; O2 = oxygen; 1O2 = singlet oxygen; O2 = superoxide; H2O2 = hydrogen peroxide; OH = hydroxyl radical; HOCl = hypochlorous acid; and = radiation. ROS colored in red are free oxygen radicals.
Figure 1. Formation of reactive oxygen species. Abbreviations: SOD = superoxide dismutase; MPO = myeloperoxidase; O2 = oxygen; 1O2 = singlet oxygen; O2 = superoxide; H2O2 = hydrogen peroxide; OH = hydroxyl radical; HOCl = hypochlorous acid; and = radiation. ROS colored in red are free oxygen radicals.
Molecules 26 04710 g001
Figure 2. Detection of superoxide using MitoSox fluorescent probe. Abbreviation: O2 = superoxide.
Figure 2. Detection of superoxide using MitoSox fluorescent probe. Abbreviation: O2 = superoxide.
Molecules 26 04710 g002
Figure 3. Generation of 1O2 in a photocatalytic reaction on the TiO2 surface. Abbreviations: 3O2 = molecular oxygen; 1O2 = singlet oxygen; and hν = radiation.
Figure 3. Generation of 1O2 in a photocatalytic reaction on the TiO2 surface. Abbreviations: 3O2 = molecular oxygen; 1O2 = singlet oxygen; and hν = radiation.
Molecules 26 04710 g003
Figure 4. Detection of hydrogen peroxide using a probe DCFH-DA. Abbreviations: DCFH-DA = 2′,7′-dichlorodihydrofluorescein diacetate; DCFH = 2′,7′-dichlorodihydrofluorescein; DCF = 2′,7′-dichlorofluorescein; and ROS = reactive oxygen species.
Figure 4. Detection of hydrogen peroxide using a probe DCFH-DA. Abbreviations: DCFH-DA = 2′,7′-dichlorodihydrofluorescein diacetate; DCFH = 2′,7′-dichlorodihydrofluorescein; DCF = 2′,7′-dichlorofluorescein; and ROS = reactive oxygen species.
Molecules 26 04710 g004
Figure 5. Oxidation of Amplex Red to a fluorescent (resorufin) and non-fluorescent (resazurin) product. Abbreviation: HRP = horseradish peroxidase.
Figure 5. Oxidation of Amplex Red to a fluorescent (resorufin) and non-fluorescent (resazurin) product. Abbreviation: HRP = horseradish peroxidase.
Molecules 26 04710 g005
Figure 6. Possible pathways of induction of apoptosis by nanomaterials in cells. Abbreviations: ER = endoplasmatic reticulum and FADD = FAS-associated death domain protein.
Figure 6. Possible pathways of induction of apoptosis by nanomaterials in cells. Abbreviations: ER = endoplasmatic reticulum and FADD = FAS-associated death domain protein.
Molecules 26 04710 g006
Table 1. Overview of nanomaterials capable of ROS production [28].
Table 1. Overview of nanomaterials capable of ROS production [28].
NanomaterialProduced ROSROS Half-Life
ZnO [29], SiO2 [29], TiO2 [30], CuO [31], Ag NPs [32]SuperoxideO210−6 s
ZnO [33], TiO2 [34], CuO [35]Hydroxyl radicalOH10−10 s
Polystyrene NPs [36], Au NPs [37], TiO2 [38], ZnO [39], Ag NPs [40]Hydrogen peroxideH2O2Stable (x.s, min)
TiO2 [41], Ag NPs [42], FeO [43]Singlet oxygen1O210−6 s
Table 2. Overview of fluorescent probes for the detection of ROS [79,95,97,98,114,115,137,141,142,143,145].
Table 2. Overview of fluorescent probes for the detection of ROS [79,95,97,98,114,115,137,141,142,143,145].
Type of ROSFluorescent Probe Excitation/Emission Wavelengths
SuperoxideMitoSox535/610 nm
1,3–diphenylisobenzofuran410/455 nm
Hydroxyl radicalTerephthalic acid310/420 nm
Rhodamine nitroxide560/588 nm
HKOH-1500/520 nm
Singlet oxygenDPAX-1495/515 nm
DMAX495/515 nm
Singlet Oxygen Sensor Green®504/525 nm
Hydrogen peroxide2′,7′-dichlorodihydrofluorescein498/522 nm
Amplex Red563/587 nm
HyPer ratiometric sensor485/516 nm
Pentafluorobenzenesulfonyl fluoresceins485/530 nm
Europium ion400/616 nm
Homovanilic acid312/420 nm
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Čapek, J.; Roušar, T. Detection of Oxidative Stress Induced by Nanomaterials in Cells—The Roles of Reactive Oxygen Species and Glutathione. Molecules 2021, 26, 4710. https://doi.org/10.3390/molecules26164710

AMA Style

Čapek J, Roušar T. Detection of Oxidative Stress Induced by Nanomaterials in Cells—The Roles of Reactive Oxygen Species and Glutathione. Molecules. 2021; 26(16):4710. https://doi.org/10.3390/molecules26164710

Chicago/Turabian Style

Čapek, Jan, and Tomáš Roušar. 2021. "Detection of Oxidative Stress Induced by Nanomaterials in Cells—The Roles of Reactive Oxygen Species and Glutathione" Molecules 26, no. 16: 4710. https://doi.org/10.3390/molecules26164710

Article Metrics

Back to TopTop