Skip to main content

PERSPECTIVE article

Front. Cell Dev. Biol., 07 September 2021
Sec. Cellular Biochemistry
Volume 9 - 2021 | https://doi.org/10.3389/fcell.2021.716406

Reactive Oxygen Species: Not Omnipresent but Important in Many Locations

  • Institute for Medical Microbiology, Immunology and Hygiene, Cologne, Germany

Reactive oxygen species (ROS), such as the superoxide anion or hydrogen peroxide, have been established over decades of research as, on the one hand, important and versatile molecules involved in a plethora of homeostatic processes and, on the other hand, as inducers of damage, pathologies and diseases. Which effects ROS induce, strongly depends on the cell type and the source, amount, duration and location of ROS production. Similar to cellular pH and calcium levels, which are both strictly regulated and only altered by the cell when necessary, the redox balance of the cell is also tightly regulated, not only on the level of the whole cell but in every cellular compartment. However, a still widespread view present in the scientific community is that the location of ROS production is of no major importance and that ROS randomly diffuse from their cellular source of production throughout the whole cell and hit their redox-sensitive targets when passing by. Yet, evidence is growing that cells regulate ROS production and therefore their redox balance by strictly controlling ROS source activation as well as localization, amount and duration of ROS production. Hopefully, future studies in the field of redox biology will consider these factors and analyze cellular ROS more specifically in order to revise the view of ROS as freely flowing through the cell.

Reactive Oxygen Species: Handle With Care!

Reactive oxygen species (ROS) are molecules with higher reactivity than molecular oxygen (O2) (Fenton, 1894; Haber et al., 1997; D’Autreaux and Toledano, 2007; Hatz et al., 2007; Prousek, 2007; Das and Roychoudhury, 2014; Odobasic et al., 2016; Aratani, 2018; Juan et al., 2021) and include highly reactive free radicals, such as the superoxide anion (O2) (Hayyan et al., 2016), and non-radical species, such as hydrogen peroxide (H2O2) (Chavez et al., 2009; Odobasic et al., 2016; Aratani, 2018).

A plethora of different ROS sources, such as NADPH oxidases (Nox) (Bedard and Krause, 2007; Nakano et al., 2008; Lambeth and Neish, 2014; Gluschko et al., 2018; Wolf et al., 2020), mitochondria (Afanas’ev et al., 1999; Han et al., 2003; Muller et al., 2004; Murphy, 2009; Brand, 2010; Shoshan-Barmatz et al., 2010; West et al., 2011a,b; Bezawork-Geleta et al., 2017; Zhao et al., 2019; Hadrava Vanova et al., 2020), xanthine oxidase (Nomura et al., 2014; Al-Shehri et al., 2020), peroxisomes (Lismont et al., 2015; Fransen et al., 2017; Shai et al., 2018) and cytochrome P450 oxidases (Omura and Sato, 1962; Zhang et al., 2020) can be responsible for cellular ROS production and it highly depends on the stimulus and the cell type whether a single or multiple ROS sources are activated, for how long this occurs and for what purpose (Banfi et al., 2004; Gavazzi et al., 2006; Bedard and Krause, 2007; Aguirre and Lambeth, 2010; Brand, 2010; Dikalova et al., 2010; Donko et al., 2010; Carnesecchi et al., 2011; Al-Mehdi et al., 2012; Lanciano et al., 2013; Kim et al., 2014; Lambeth and Neish, 2014; Ives et al., 2015; Xu et al., 2017; Herb et al., 2019b; Hernansanz-Agustin et al., 2020; Herb and Schramm, 2021).

O2 is the precursor of all cellular ROS (Chavez et al., 2009; Niethammer et al., 2009) and, under physiological pH, cannot diffuse over cell membranes due to its negative charge (Takahashi and Asada, 1983; Shoshan-Barmatz et al., 2010; Cordeiro, 2014). It has a half-life of a couple of seconds or even less (Marklund, 1976; Rapoport et al., 1994; D’Autreaux and Toledano, 2007; Taverne et al., 2013). Despite being not a strong oxidizing substance by itself, O2 readily oxidizes iron-sulfur structures of proteins (Liochev and Fridovich, 1999; Imlay, 2003), which can lead to protein malfunction and iron release from proteins. The released iron reacts with H2O2 to form the highly reactive and toxic OH radical (Liochev and Fridovich, 1999). Therefore, under healthy conditions cells keep O2 levels low (∼ 10–11-10–12 M) by compartmentalization and quick removal of O2by superoxide dismutates (D’Autreaux and Toledano, 2007; Kehrer et al., 2010; Wang et al., 2018), which are expressed in all cellular compartments (Melov et al., 1999; Lambeth and Neish, 2014; Wang et al., 2018). A sustained increase in cellular O2 levels is associated with damage to cellular structures (Davies, 2016; Pizzino et al., 2017; Sies et al., 2017; Gutteridge and Halliwell, 2018; Su et al., 2019; Juan et al., 2021). However, O2 can also contribute to cellular signaling (Chen et al., 2009; Xu et al., 2017; Ren et al., 2021).

O2 quickly dismutates to H2O2, which is more, although not freely, diffusible for cellular membranes (Bienert et al., 2006; Wang et al., 2020; Chauvigne et al., 2021), which questions saturation of the cell with H2O2 to fulfill signaling functions in compartments, which are not in direct proximity to the ROS source (Beretta et al., 2020; Sies, 2021). Communication between cellular compartments can be achieved by aquaporins, which facilitate a controlled passage of H2O2 over membranes (Bienert and Chaumont, 2014; Wang et al., 2020). H2O2 has a longer cellular half-life (∼1 ms) with concentrations of ∼10–7 M under cellular homoeostatic conditions (D’Autreaux and Toledano, 2007). Because of these properties, it functions as an important signaling molecule involved in many different cellular processes (Kamata et al., 2005; Tonks, 2005; Rhee, 2006; Marinho et al., 2013; Holmstrom and Finkel, 2014; Romero et al., 2014; Jones et al., 2016; Short et al., 2016; Zhang et al., 2016; Herb et al., 2019b; Sies and Jones, 2020; Chauvigne et al., 2021). H2O2-mediated signaling is mainly based on the oxidation of cysteine residues of proteins (Chiarugi et al., 2001; Rhee, 2006; Herscovitch et al., 2008; Romero et al., 2014; Jones et al., 2016; Short et al., 2016; Herb et al., 2019b). These cysteine residues have a low pKa, are exposed to the cytosol and deprotonated to thiolate groups (Finkel, 2011; Poole, 2015). An increase to nanomolar concentrations (∼100 nM) of H2O2 is sufficient to induce reversible oxidation. This can lead to allosteric protein changes that alter the enzymatic function of the target proteins in many ways (Lee et al., 1998; Meng et al., 2002; Kamata et al., 2005; Tonks, 2005). ROS-mediated oxidation can also lead to covalent linkage of cysteine residues by disulfide bonds (Herscovitch et al., 2008; Zhou et al., 2014; Herb et al., 2019b). Since these H2O2-mediated protein oxidations can be reversed by the antioxidant defense system, they represent important redox switches involved in various cellular processes (Barford, 2004; Holmstrom and Finkel, 2014). Excessive H2O2 production, however, leads to further oxidation of the oxidized cysteines, which is an irreversible process and results in protein malfunction (Winterbourn and Hampton, 2008).

Growing evidence indicates that the redox status in different cellular compartments varies greatly (Fransen et al., 2017; Beretta et al., 2020; Chauvigne et al., 2021; Sies, 2021; Wang et al., 2021b), is tightly regulated (Nakamura, 2005; Kirkman and Gaetani, 2007; Kelley et al., 2010; Brigelius-Flohe and Maiorino, 2013; Poljsak et al., 2013; Couto et al., 2016; Jones et al., 2016; Chauvigne et al., 2021) and every elevation of ROS levels is controlled by the cell in various ways (Babior, 2002; Bulua et al., 2011; Hoeven et al., 2011; West et al., 2011a; Gluschko et al., 2018; Herb et al., 2019b; Chauvigne et al., 2021). The condition of ROS levels exceeding the capacity of cellular antioxidant defense systems is termed oxidative stress (Niki, 2016; Sies and Jones, 2020). Oxidative stress can be further divided into two subforms: (1) Oxidative distress, which represents excessive and prolonged oxidative stress, causes damage to cellular components and results in a number of different pathologies (Cross et al., 1987; Rangasamy et al., 2005; Thimmulappa et al., 2006). Notably, excessive oxidative distress is not always detrimental, if produced at the right place. Exceeded generation of oxidative stress in pathogen-containing phagosomes of phagocytes, for example, is an important factor of antimicrobial immunity (West et al., 2011a; Winterbourn and Kettle, 2013; Gluschko et al., 2018; Herb and Schramm, 2021). (2) Oxidative eustress represents a tightly controlled increase in cellular ROS levels (Niki, 2016; Sies and Jones, 2020), which are sufficient to fulfill important cellular processes, but do not induce critical damage to cellular structures (Tai et al., 2009; Finkel, 2011; Nathan and Cunningham-Bussel, 2013; Reczek and Chandel, 2014; Herb et al., 2019b).

Unfortunately, a lot of studies, which show the important role of ROS in various cellular processes, often suggest that ROS are produced in excess, saturate the cell and react randomly with redox-sensitive targets. This is mainly due to experimental setups that might lead to misinterpretation of the location of ROS production in cells.

Many studies use only one type of ROS probe, but do not provide an explanation for the choice, such as specificity for a cellular compartment or a defined type of ROS subspecies. Often probes are used that show neither a specificity for a ROS subspecies nor a defined cellular compartment, which leads to the frequently used terms “intracellular ROS” or “total cellular ROS,” which implicate that ROS once produced are equally distributed in the cell. Common examples for diffusible ROS probes are luminol (Caldefie-Chezet et al., 2002; Pavelkova and Kubala, 2004), 2′,7′-dichlordihydrofluorescein-diacetat (H2DCF-DA) (Ushijima et al., 1997; Hempel et al., 1999; Kim and Xue, 2020; Kim et al., 2021; Wang et al., 2021a) or dihydroethidium (DHE) (Gatliff et al., 2017; Wang and Zou, 2018; Zeller et al., 2021), which are regarded as compartment-specific but in fact they are not (Lundqvist and Dahlgren, 1996; Ushijima et al., 1997; Hempel et al., 1999; Wang and Zou, 2018). There are compartment-specific derivates available for these ROS probes, namely Isoluminol (Lundqvist and Dahlgren, 1996; Dahlgren and Karlsson, 1999; Caldefie-Chezet et al., 2002; Gluschko et al., 2018; Herb et al., 2019b; Wolf et al., 2020), 5-(and −6)-carboxy-2′,7′-dihydrochlorofluorescein-diacetat (5/6-Carboxy-DCF) (Hempel et al., 1999; Mak et al., 2017; Herb et al., 2019b; Wolf et al., 2020) and MitoSOX Red (MitoSOX) (Robinson et al., 2006; Mukhopadhyay et al., 2007) as alternatives, whose combined usage gives a much more confluent picture of the cellular ROS production. The preferable option for most precise ROS measurements concerning compatibility and specificity for ROS subspecies is represented by genetically modified cells, which express the ROS probe of choice in the cellular compartment of choice, like the HyPer family reporters and roGFP2-Orp1 (Belousov et al., 2006; Gutscher et al., 2009; Markvicheva et al., 2011; Bilan et al., 2013; Hernández-Barrera et al., 2013; Wang et al., 2021b). Another precise approach for compartment-specific ROS measurements is the coupling of ROS probes to cargo/particles, which can be engulfed by cells. This technique is especially useful in phagocytes like macrophages, to determine ROS levels in the phagosome (Geng et al., 2015; Ligeon et al., 2021a,b). For further reading on topics regarding ROS detection methods we want to point out to other reviews (Ermakova et al., 2014; Herb and Schramm, 2021).

Also the combined use of only globally working ROS scavengers in combination with ROS probes that detect total cellular ROS can lead to results, which suggest that ROS are present in the whole cell after diffusion from the location of their production. With NAC as most prominent globally working ROS scavenger (Patriarca et al., 2005; Aldini et al., 2018; Ezerina et al., 2018) only the general involvement of ROS in the cellular process of interest can be investigated, but no compartment-specific ROS production can be analyzed. More examples of globally working ROS scavengers are Tempol (4-Hydroxy-Tempo) (dismutation of O2 into H2O2) (Bernardy et al., 2017; Herb et al., 2019b), Tiron (a global O2 scavenger) (Krishna et al., 1992; Hein and Kuo, 1998; Manzano et al., 2000), Trolox (globally scavenges OOH and OOR) (Davies et al., 1988; Dugas et al., 2000) and ebselen (effectively removes H2O2 and ONOO) (Nakamura et al., 2002; Matsushita et al., 2004; Mugesh, 2013). All of the scavengers mentioned above are diffusible (Davies et al., 1988; Krishna et al., 1992; Hein and Kuo, 1998; Haj-Yehia et al., 1999; Dugas et al., 2000; Manzano et al., 2000; Rak et al., 2000; Nakamura et al., 2002; Matsushita et al., 2004; Mugesh, 2013; Herb and Schramm, 2021). Assessment of specific removal of ROS subspecies and therefore their involvement in cellular processes is possible with these substances, but they cannot be used to identify the specific compartment in which the ROS exert their function.

Not only the location of ROS production, but also their various sources and their activation, regulation and termination is of major importance for the understanding of the complex redox maintenance in cells. For the identification of ROS sources it is not always possible to provide genetic evidence with a knock-out system or by siRNA usage. ROS source inhibitors are in these cases an option to block ROS production and analyze possible ROS sources. There are a lot of specific ROS source inhibitors commercially available and the choice is continuously expanded (Murphy, 2009; Wind et al., 2010; Altenhofer et al., 2015; Herb and Schramm, 2021).

For Nox enzymes, as one of the most prominent ROS sources in many cell types, the well-validated general Nox inhibitors VAS2870 (Leusen et al., 1995; ten Freyhaus et al., 2006; Wind et al., 2010; Altenhofer et al., 2012, 2015) or GKT 137831 (Laleu et al., 2010; Sedeek et al., 2010; Gaggini et al., 2011; Aoyama et al., 2012; Strengert et al., 2014; Kim et al., 2021) can be used. Both inhibitors show no intrinsic antioxidant activity and do not inhibit other flavoproteins (Wind et al., 2010; Altenhofer et al., 2012; Teixeira et al., 2017). Both also inhibit Nox-derived ROS production in vitro and in vivo (Carnesecchi et al., 2009, 2011; Aoyama et al., 2012; Green et al., 2012; Bettaieb et al., 2015; Gorin et al., 2015). In sharp contrast to VAS2870 and GKT 137831, the substances apocynin and DPI are still used and falsely addressed as specific Nox inhibitors in many otherwise convincing and excellent studies (Barbieri et al., 2003; Kiritoshi et al., 2003; Dostert et al., 2008; Choi et al., 2011; Abuaita et al., 2015; Gatliff et al., 2017; Alonso et al., 2019; Fan et al., 2019; Damiano et al., 2020; Geng et al., 2020; Inomata et al., 2020; Prestes et al., 2020; Ahmad et al., 2021; Ligeon et al., 2021b; Martinez et al., 2021 #1039; Troia et al., 2021). Several studies have shown that apocynin directly scavenges ROS due to its antioxidant capacities (Aldieri et al., 2008; Heumuller et al., 2008; Mora-Pale et al., 2009; Wingler et al., 2011; Trevelin et al., 2016), while DPI inhibits flavoproteins in general (O’Donnell et al., 1993; Wind et al., 2010; Altenhofer et al., 2015) including Nox2 (Reis et al., 2020), but also various other targets, such as complex I of the mitochondrial electron transport chain (Bloxham, 1979; Lambeth et al., 2008; Bulua et al., 2011), iNOS (Stuehr et al., 1991; Geyer et al., 1997) or xanthine oxidase (O’Donnell et al., 1993; Wind et al., 2010) as well as calcium transporters (Tazzeo et al., 2009). Since genetic knock-out models with the CRISPR-Cas9 technology (Ledford, 2015; Anzalone et al., 2020; Carlson-Stevermer et al., 2020), either for cell lines, ex vivo cells or mice, as well as knock-down via siRNA (Han, 2018) are readily available tools for analyzing possible roles of Nox enzymes in cellular processes, the use of apocynin or DPI, especially in combination with diffusible ROS probes, should not be recommended, since it may lead to false interpretations of results regarding Nox enzyme involvement and the location of ROS production.

In mitochondria, the complexes of the ETC not only are essential for energy generation of the cell, but are also ROS production sites (Nohl et al., 2003; Lambeth and Neish, 2014). Inhibition of the complexes for analysis of ROS production might also result in energy deprivation and the energy status of the cell has to be checked every time these inhibitors are used. Typically used inhibitors are rotenone (Stowe and Camara, 2009; Heinz et al., 2017; Scialo et al., 2017), which inhibits complex I and increases ROS production inside the mitochondrial matrix (St-Pierre et al., 2002; Lambert and Brand, 2004; Panov et al., 2005; Stowe and Camara, 2009; Sena et al., 2013) and antimycin A (Murphy, 2009; Bleier and Drose, 2013), which inhibits complex III and increases ROS production into the intermembrane space (IMS) (Chen et al., 2003; Han et al., 2003; Al-Mehdi et al., 2012; Quinlan et al., 2012; Herb et al., 2019a). The most commonly used ROS probe for detection of mitochondrial ROS is MitoSOX, which measures O2 exclusively inside the mitochondrial matrix (Robinson et al., 2006; Mukhopadhyay et al., 2007; Ernst et al., 2021). However, since the ETC complexes show compartment-specific differences concerning ROS production (Fridovich, 1997; Murphy, 2009; Brand, 2010; West et al., 2011b; Herb and Schramm, 2021), this probe can only be used to measure ROS production inside mitochondria and, therefore, other cellular compartments should always be analyzed in addition. In healthy, undamaged mitochondria, ROS cannot escape the mitochondrial matrix because of the very effective antioxidative defense system (Roca and Ramakrishnan, 2013; Briston et al., 2017; Hos et al., 2017; Hernansanz-Agustin et al., 2020; Lin et al., 2020; Wang et al., 2020; Zhao et al., 2020). Only after prolonged overproduction or when the structure of the mitochondrial membranes is ruptured, either by opening of the mitochondrial permeability transition pore or direct damage, e.g., by pathogenic toxins, ROS can escape from the matrix into the cytosol (Koterski et al., 2005; Stavru et al., 2011; Roca and Ramakrishnan, 2013; Briston et al., 2017; Hos et al., 2017; Zhang Y. et al., 2019; Zhao et al., 2020). Nevertheless, the general term “mitochondrial ROS” is used in many studies, which often is synonymous for matrix-located mitochondrial ROS production measured by MitoSOX. ROS measurements in other cellular compartments as well as an explanation if and how the mitochondrial ROS escape from the matrix and fulfill their role in the cell, with a few exceptions (Koterski et al., 2005; Zhou et al., 2011; Roca and Ramakrishnan, 2013; Briston et al., 2017; Hos et al., 2017; Herb et al., 2019b; Zhao et al., 2020), are often not provided. Additionally, the usage of inhibitors of the ETC, like rotenone or antimycin A, which have compartment-specific effects on ROS production in combination with diffusible ROS probes can also lead to misinterpretations of the performed ROS measurements. For further reading concerning ROS scavengers and inhibitors, we like to point to other reviews (Wind et al., 2010; Altenhofer et al., 2015; Herb and Schramm, 2021).

ROS Production: The Dose Makes the Poison

A model that involves an uncontrolled increase in total cellular ROS levels implies that cells take into account the collateral damage that ROS can inflict while enroute to their redox-sensitive target, that can be at a completely different cellular location (Bulua et al., 2011; Nazarewicz et al., 2013; Kelly et al., 2015; Garaude et al., 2016; Kim et al., 2017; Figure 1). But oxidative distress (Buczynski et al., 2013; Li et al., 2013; Bhattacharyya et al., 2014; Niki, 2016; Sweeney and McAuley, 2016; Yang et al., 2016; Sies and Jones, 2020) is a situation for the healthy cell that has to be avoided. Tightly controlled production of ROS in direct vicinity of a redox-sensitive target (Tai et al., 2009; Finkel, 2011; Nathan and Cunningham-Bussel, 2013; Reczek and Chandel, 2014; Herb et al., 2019b) requires much less ROS production and hence results in much less collateral damage, while fulfilling important cellular functions, i.e., oxidative eustress (Niki, 2016; Sies and Jones, 2020; Figure 2).

FIGURE 1
www.frontiersin.org

Figure 1. Several studies suggested that ROS are produced in excess, saturate the cell and find their redox-sensitive targets at random. Usage of diffusible ROS probes, globally working ROS scavengers and unspecific inhibitors often place the suggested ROS source at a completely different location than the redox-sensitive target, which might lead to the interpretation that cells “take into account” the damage that ROS can inflict on their way to the target molecule. The importance of ROS in general for various cellular processes was shown by many excellent studies (Bulua et al., 2011; Nazarewicz et al., 2013; Kelly et al., 2015; Garaude et al., 2016; Kim et al., 2017), however, diffusible ROS probes or only one ROS probe are often used to determine ROS production in cells, which might lead to the suggestions, e.g., that (1) ROS escape from the mitochondrial matrix and regulate expression and secretion of cytokines (Bulua et al., 2011; Kelly et al., 2015), (2) extracellular Nox2-derived ROS modify enzyme activity in the mitochondrial matrix or matrix-derived ROS modulate Nox2 activity (Nazarewicz et al., 2013; Garaude et al., 2016) or (3) ROS produced by ER-located Nox4 reach the phagosome for inactivation of phagocytosed parasites (Kim et al., 2017).

FIGURE 2
www.frontiersin.org

Figure 2. Growing evidence supports the hypothesis that cellular compartments show big differences and tight regulation of their redox status. The induction of ROS production is controlled by the cell in terms of location, source, duration and amount. The localized and timely controlled ROS production in the direct vicinity of the redox-sensitive target reduces the induced damage to cellular components and results in beneficial consequences for the cell, a condition termed as oxidative eustress (Sies, 2021). Examples of localized ROS production are (1) the production of antimicrobial ROS by Nox2 (Craig and Slauch, 2009; Gluschko et al., 2018) or mitochondria, which are recruited to pathogen-containing phagosomes (West et al., 2011a; Geng et al., 2015), (2) the recruitment of the redox-regulated target to ROS-producing mitochondria for NLRP3 inflammasome activation (Zhou et al., 2011) or the relocation of mitochondria to the nucleus for ROS-mediated nuclear signaling (Al-Mehdi et al., 2012) and (3) ROS production by ER-localized Nox4 during formation of mitochondria-associated membranes for regulation of calcium signaling (Beretta et al., 2020).

Of note, to reach levels at which the ROS can fulfill their important cellular roles, ROS production has to overcome the highly effective antioxidative defense systems of one or more cellular compartments (Nakamura, 2005; Kirkman and Gaetani, 2007; Kelley et al., 2010; Brigelius-Flohe and Maiorino, 2013; Poljsak et al., 2013; Lismont et al., 2015; Couto et al., 2016; Jones et al., 2016; Shai et al., 2018; Chauvigne et al., 2021; Herb and Schramm, 2021). Also H2O2-detoxifying enzymes, such as glutathione peroxidases or catalase, which can quickly decrease cellular H2O2 concentrations (Winterbourn and Hampton, 2008) as well as the inactivation of these scavenger enzymes by H2O2 itself broaden the regulatory potential of cells for selective and localized ROS signaling (Wood et al., 2003; Woo et al., 2010). This further places the model of ROS molecules, which “flood” the whole cell from a single location of ROS production in rather unrealistic light. Nevertheless, the model of ROS as “flooding” the cell is still popular in the scientific community, often referred to as “total cellular ROS levels” (Wang et al., 2013, 2019; Dinakar et al., 2016; Kim and Xue, 2020; Loth et al., 2020; Thorne et al., 2021), “intracellular ROS levels” (Tepel et al., 2000; Bensaad et al., 2009; Choi et al., 2011; Lee et al., 2017; Wang et al., 2019, 2021a; Wei et al., 2019; Zaidieh et al., 2019; Zhang W. et al., 2019; Mendiola et al., 2020; Winitchaikul et al., 2021; Zhong et al., 2021) or simply “ROS levels” (Chen et al., 2012, 2021; Wei et al., 2019; Agarwal and Ganesh, 2020; Kim et al., 2021; Knight et al., 2021; Zeller et al., 2021) in many studies, mainly because of the usage of diffusible ROS probes, which suggest free diffusion of ROS through the cell without regard of the location of ROS production. Mutations, e.g., in cancer cells (Schumacker, 2006; Liou and Storz, 2010; Reczek and Chandel, 2017; Zaidieh et al., 2019; Perillo et al., 2020), pathogenic invasion (West et al., 2011a; Abuaita et al., 2018; Gluschko et al., 2018; Roca et al., 2019) or metabolic disbalance (Li et al., 2016; Mak et al., 2017; Peng et al., 2021) are prominent examples, in which the ROS production of the cell can enter an uncontrolled stage and quickly overcome the antioxidative defense system leading to rapidly increased ROS levels in nearly every compartment of the cell with often detrimental consequences. In this context, cells can be regarded as “overflowing with ROS,” however, in healthy cells a redox balance between all producing and eliminating ROS sources, mediated by the antioxidant defense system is crucial for cellular functioning (Nathan and Cunningham-Bussel, 2013; Deshmukh et al., 2017; Suzuki et al., 2019; Wei et al., 2019; Baird and Yamamoto, 2020; Saito and Kimura, 2021). Importantly, in the extracellular milieu H2O2 can travel much further than inside the cell and fulfills important signaling (Levine et al., 1994; Sharma et al., 2012; Das and Roychoudhury, 2014; Hervera et al., 2018; Huang et al., 2019; Janku et al., 2019; Deng et al., 2020) and chemotactic functions (Niethammer et al., 2009; Rieger and Sagasti, 2011).

Concluding Remarks

In recent years, more and more studies supported the model–and highlighted the importance–of localized cellular ROS production in direct vicinity of the redox target (Meinhard and Grill, 2001; Veal et al., 2007; Go and Jones, 2008; Craig and Slauch, 2009; West et al., 2011a; Wink et al., 2011; Zhou et al., 2011; Al-Mehdi et al., 2012; Naviaux, 2012; Allan et al., 2014; Romero et al., 2014; Geng et al., 2015; To et al., 2017; Gluschko et al., 2018; Herb et al., 2019b; Acin-Perez et al., 2020; Beretta et al., 2020; Chanin et al., 2020; Miller et al., 2020; Sies and Jones, 2020; Herb and Schramm, 2021; Ligeon et al., 2021b; Sies, 2021; Wong et al., 2021). Therefore, the model of ROS molecules as “omnipresent and freely diffusing throughout the cell” should always be interpreted carefully in the context of research and highly depends on the proper use of ROS probes, scavengers and inhibitors. In healthy cells, ROS should be considered as molecules, whose production is tightly controlled in terms of stimulus, source, location, duration and amount.

Data Availability Statement

The original contributions presented in the study are included in the article/supplementary material, further inquiries can be directed to the corresponding author/s.

Author Contributions

MH and MS: conceptualization. MH: writing–original draft preparation and visualization. MH, AG, and MS: writing–review and editing and funding acquisition. MS: supervision and project administration. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by the Deutsche Forschungsgemeins- chaft (DFG) grant SCHR 1627/2-1 to MS, by the Köln Fortune grant 278/2019 to AG, and by the Köln Fortune grant 302/2020 to MH.

Conflict of Interest

The authors declare that the research was conducted in the absence of any commercial or financial relationships that could be construed as a potential conflict of interest.

Publisher’s Note

All claims expressed in this article are solely those of the authors and do not necessarily represent those of their affiliated organizations, or those of the publisher, the editors and the reviewers. Any product that may be evaluated in this article, or claim that may be made by its manufacturer, is not guaranteed or endorsed by the publisher.

References

Abuaita, B. H., Burkholder, K. M., Boles, B. R., and O’Riordan, M. X. (2015). The endoplasmic reticulum stress sensor inositol-requiring enzyme 1alpha augments bacterial killing through sustained oxidant production. mBio 6:e00705.

Google Scholar

Abuaita, B. H., Schultz, T. L., and O’Riordan, M. X. (2018). Mitochondria-derived vesicles deliver antimicrobial reactive oxygen species to control phagosome-localized Staphylococcus aureus. Cell Host Microbe 24, 625–636.e5.

Google Scholar

Acin-Perez, R., Iborra, S., Marti-Mateos, Y., Cook, E. C. L., Conde-Garrosa, R., Petcherski, A., et al. (2020). Fgr kinase is required for proinflammatory macrophage activation during diet-induced obesity. Nat. Metab. 2, 974–988. doi: 10.1038/s42255-020-00273-8

PubMed Abstract | CrossRef Full Text | Google Scholar

Afanas’ev, I. B., Ostrachovitch, E. A., and Korkina, L. G. (1999). Lucigenin is a mediator of cytochrome C reduction but not of superoxide production. Arch. Biochem. Biophys. 366, 267–274. doi: 10.1006/abbi.1999.1215

PubMed Abstract | CrossRef Full Text | Google Scholar

Agarwal, S., and Ganesh, S. (2020). Perinuclear mitochondrial clustering, increased ROS levels, and HIF1 are required for the activation of HSF1 by heat stress. J. Cell Sci. 133:jcs245589.

Google Scholar

Aguirre, J., and Lambeth, J. D. (2010). Nox enzymes from fungus to fly to fish and what they tell us about Nox function in mammals. Free Radic. Biol. Med. 49, 1342–1353. doi: 10.1016/j.freeradbiomed.2010.07.027

PubMed Abstract | CrossRef Full Text | Google Scholar

Ahmad, A., Nawaz, M. I., Siddiquei, M. M., and Abu El-Asrar, A. M. (2021). Apocynin ameliorates NADPH oxidase 4 (NOX4) induced oxidative damage in the hypoxic human retinal Muller cells and diabetic rat retina. Mol. Cell. Biochem. 476, 2099–2109. doi: 10.1007/s11010-021-04071-y

PubMed Abstract | CrossRef Full Text | Google Scholar

Aldieri, E., Riganti, C., Polimeni, M., Gazzano, E., Lussiana, C., Campia, I., et al. (2008). Classical inhibitors of NOX NAD(P)H oxidases are not specific. Curr. Drug Metab. 9, 686–696. doi: 10.2174/138920008786049285

PubMed Abstract | CrossRef Full Text | Google Scholar

Aldini, G., Altomare, A., Baron, G., Vistoli, G., Carini, M., Borsani, L., et al. (2018). N-Acetylcysteine as an antioxidant and disulphide breaking agent: the reasons why. Free Radic. Res. 52, 751–762. doi: 10.1080/10715762.2018.1468564

PubMed Abstract | CrossRef Full Text | Google Scholar

Allan, E. R., Tailor, P., Balce, D. R., Pirzadeh, P., Mckenna, N. T., Renaux, B., et al. (2014). NADPH oxidase modifies patterns of MHC class II-restricted epitopic repertoires through redox control of antigen processing. J. Immunol. 192, 4989–5001. doi: 10.4049/jimmunol.1302896

PubMed Abstract | CrossRef Full Text | Google Scholar

Al-Mehdi, A. B., Pastukh, V. M., Swiger, B. M., Reed, D. J., Patel, M. R., Bardwell, G. C., et al. (2012). Perinuclear mitochondrial clustering creates an oxidant-rich nuclear domain required for hypoxia-induced transcription. Sci. Signal. 5:ra47. doi: 10.1126/scisignal.2002712

PubMed Abstract | CrossRef Full Text | Google Scholar

Alonso, D., Serrano, E., Bermejo, F. J., and Corral, R. S. (2019). HIF-1alpha-regulated MIF activation and Nox2-dependent ROS generation promote Leishmania amazonensis killing by macrophages under hypoxia. Cell. Immunol. 335, 15–21. doi: 10.1016/j.cellimm.2018.10.007

PubMed Abstract | CrossRef Full Text | Google Scholar

Al-Shehri, S. S., Duley, J. A., and Bansal, N. (2020). Xanthine oxidase-lactoperoxidase system and innate immunity: biochemical actions and physiological roles. Redox Biol. 34:101524. doi: 10.1016/j.redox.2020.101524

PubMed Abstract | CrossRef Full Text | Google Scholar

Altenhofer, S., Kleikers, P. W., Radermacher, K. A., Scheurer, P., Rob Hermans, J. J., Schiffers, P., et al. (2012). The NOX toolbox: validating the role of NADPH oxidases in physiology and disease. Cell. Mol. Life Sci. 69, 2327–2343. doi: 10.1007/s00018-012-1010-9

PubMed Abstract | CrossRef Full Text | Google Scholar

Altenhofer, S., Radermacher, K. A., Kleikers, P. W., Wingler, K., and Schmidt, H. H. (2015). Evolution of NADPH oxidase inhibitors: selectivity and mechanisms for target engagement. Antioxid. Redox Signal. 23, 406–427. doi: 10.1089/ars.2013.5814

PubMed Abstract | CrossRef Full Text | Google Scholar

Anzalone, A. V., Koblan, L. W., and Liu, D. R. (2020). Genome editing with CRISPR-Cas nucleases, base editors, transposases and prime editors. Nat. Biotechnol. 38, 824–844. doi: 10.1038/s41587-020-0561-9

PubMed Abstract | CrossRef Full Text | Google Scholar

Aoyama, T., Paik, Y. H., Watanabe, S., Laleu, B., Gaggini, F., Fioraso-Cartier, L., et al. (2012). Nicotinamide adenine dinucleotide phosphate oxidase in experimental liver fibrosis: GKT137831 as a novel potential therapeutic agent. Hepatology 56, 2316–2327. doi: 10.1002/hep.25938

PubMed Abstract | CrossRef Full Text | Google Scholar

Aratani, Y. (2018). Myeloperoxidase: its role for host defense, inflammation, and neutrophil function. Arch. Biochem. Biophys. 640, 47–52. doi: 10.1016/j.abb.2018.01.004

PubMed Abstract | CrossRef Full Text | Google Scholar

Babior, B. M. (2002). The activity of leukocyte NADPH oxidase: regulation by p47PHOX cysteine and serine residues. Antioxid. Redox Signal. 4, 35–38. doi: 10.1089/152308602753625834

PubMed Abstract | CrossRef Full Text | Google Scholar

Baird, L., and Yamamoto, M. (2020). The molecular mechanisms regulating the KEAP1-NRF2 Pathway. Mol. Cell. Biol. 40:e00099-20.

Google Scholar

Banfi, B., Malgrange, B., Knisz, J., Steger, K., Dubois-Dauphin, M., and Krause, K. H. (2004). NOX3, a superoxide-generating NADPH oxidase of the inner ear. J. Biol. Chem. 279, 46065–46072. doi: 10.1074/jbc.m403046200

PubMed Abstract | CrossRef Full Text | Google Scholar

Barbieri, S. S., Eligini, S., Brambilla, M., Tremoli, E., and Colli, S. (2003). Reactive oxygen species mediate cyclooxygenase-2 induction during monocyte to macrophage differentiation: critical role of NADPH oxidase. Cardiovasc. Res. 60, 187–197. doi: 10.1016/s0008-6363(03)00365-1

CrossRef Full Text | Google Scholar

Barford, D. (2004). The role of cysteine residues as redox-sensitive regulatory switches. Curr. Opin. Struct. Biol. 14, 679–686. doi: 10.1016/j.sbi.2004.09.012

PubMed Abstract | CrossRef Full Text | Google Scholar

Bedard, K., and Krause, K. H. (2007). The NOX family of ROS-generating NADPH oxidases: physiology and pathophysiology. Physiol. Rev. 87, 245–313. doi: 10.1152/physrev.00044.2005

PubMed Abstract | CrossRef Full Text | Google Scholar

Belousov, V. V., Fradkov, A. F., Lukyanov, K. A., Staroverov, D. B., Shakhbazov, K. S., Terskikh, A. V., et al. (2006). Genetically encoded fluorescent indicator for intracellular hydrogen peroxide. Nat. Methods 3, 281–286. doi: 10.1038/nmeth866

PubMed Abstract | CrossRef Full Text | Google Scholar

Bensaad, K., Cheung, E. C., and Vousden, K. H. (2009). Modulation of intracellular ROS levels by TIGAR controls autophagy. EMBO J. 28, 3015–3026. doi: 10.1038/emboj.2009.242

PubMed Abstract | CrossRef Full Text | Google Scholar

Beretta, M., Santos, C. X., Molenaar, C., Hafstad, A. D., Miller, C. C., Revazian, A., et al. (2020). Nox4 regulates InsP3 receptor-dependent Ca(2+) release into mitochondria to promote cell survival. EMBO J. 39:e103530.

Google Scholar

Bernardy, C. C. F., Zarpelon, A. C., Pinho-Ribeiro, F. A., Calixto-Campos, C., Carvalho, T. T., Fattori, V., et al. (2017). Tempol, a superoxide dismutase mimetic agent, inhibits superoxide anion-induced inflammatory pain in mice. Biomed Res. Int. 2017:9584819.

Google Scholar

Bettaieb, A., Jiang, J. X., Sasaki, Y., Chao, T. I., Kiss, Z., Chen, X., et al. (2015). Hepatocyte Nicotinamide Adenine Dinucleotide Phosphate Reduced Oxidase 4 regulates stress signaling, fibrosis, and insulin sensitivity during development of steatohepatitis in mice. Gastroenterology 149, 468–480.e10.

Google Scholar

Bezawork-Geleta, A., Rohlena, J., Dong, L., Pacak, K., and Neuzil, J. (2017). Mitochondrial complex II: at the crossroads. Trends Biochem. Sci. 42, 312–325. doi: 10.1016/j.tibs.2017.01.003

PubMed Abstract | CrossRef Full Text | Google Scholar

Bhattacharyya, S., Sinha, K., and Sil, P. C. (2014). Cytochrome P450s: mechanisms and biological implications in drug metabolism and its interaction with oxidative stress. Curr. Drug 15, 719–742. doi: 10.2174/1389200215666141125121659

PubMed Abstract | CrossRef Full Text | Google Scholar

Bienert, G. P., and Chaumont, F. (2014). Aquaporin-facilitated transmembrane diffusion of hydrogen peroxide. Biochim. Biophys. Acta 1840, 1596–1604. doi: 10.1016/j.bbagen.2013.09.017

PubMed Abstract | CrossRef Full Text | Google Scholar

Bienert, G. P., Schjoerring, J. K., and Jahn, T. P. (2006). Membrane transport of hydrogen peroxide. Biochim. Biophys. Acta 1758, 994–1003.

Google Scholar

Bilan, D. S., Pase, L., Joosen, L., Gorokhovatsky, A. Y., Ermakova, Y. G., Gadella, T. W., et al. (2013). HyPer-3: a genetically encoded H(2)O(2) probe with improved performance for ratiometric and fluorescence lifetime imaging. ACS Chem. Biol. 8, 535–542. doi: 10.1021/cb300625g

PubMed Abstract | CrossRef Full Text | Google Scholar

Bleier, L., and Drose, S. (2013). Superoxide generation by complex III: from mechanistic rationales to functional consequences. Biochim. Biophys. Acta 1827, 1320–1331. doi: 10.1016/j.bbabio.2012.12.002

PubMed Abstract | CrossRef Full Text | Google Scholar

Bloxham, D. P. (1979). The relationship of diphenyleneiodonium-induced hypoglycaemia to the specific covalent modification of NADH-ubiquinone oxidoreductase. Biochem. Soc. Trans. 7, 103–106. doi: 10.1042/bst0070103

PubMed Abstract | CrossRef Full Text | Google Scholar

Brand, M. D. (2010). The sites and topology of mitochondrial superoxide production. Exp. Gerontol. 45, 466–472. doi: 10.1016/j.exger.2010.01.003

PubMed Abstract | CrossRef Full Text | Google Scholar

Brigelius-Flohe, R., and Maiorino, M. (2013). Glutathione peroxidases. Biochim. Biophys. Acta 1830, 3289–3303.

Google Scholar

Briston, T., Roberts, M., Lewis, S., Powney, B., Staddon, J. M., Szabadkai, G., et al. (2017). Mitochondrial permeability transition pore: sensitivity to opening and mechanistic dependence on substrate availability. Sci. Rep. 7:10492.

Google Scholar

Buczynski, B. W., Maduekwe, E. T., and O’reilly, M. A. (2013). The role of hyperoxia in the pathogenesis of experimental BPD. Semin. Perinatol. 37, 69–78. doi: 10.1053/j.semperi.2013.01.002

PubMed Abstract | CrossRef Full Text | Google Scholar

Bulua, A. C., Simon, A., Maddipati, R., Pelletier, M., Park, H., Kim, K. Y., et al. (2011). Mitochondrial reactive oxygen species promote production of proinflammatory cytokines and are elevated in TNFR1-associated periodic syndrome (TRAPS). J. Exp. Med. 208, 519–533. doi: 10.1084/jem.20102049

PubMed Abstract | CrossRef Full Text | Google Scholar

Caldefie-Chezet, F., Walrand, S., Moinard, C., Tridon, A., Chassagne, J., and Vasson, M. P. (2002). Is the neutrophil reactive oxygen species production measured by luminol and lucigenin chemiluminescence intra or extracellular? Comparison with DCFH-DA flow cytometry and cytochrome c reduction. Clin. Chim. Acta 319, 9–17. doi: 10.1016/s0009-8981(02)00015-3

CrossRef Full Text | Google Scholar

Carlson-Stevermer, J., Kelso, R., Kadina, A., Joshi, S., Rossi, N., Walker, J., et al. (2020). CRISPRoff enables spatio-temporal control of CRISPR editing. Nat. Commun. 11:5041.

Google Scholar

Carnesecchi, S., Deffert, C., Donati, Y., Basset, O., Hinz, B., Preynat-Seauve, O., et al. (2011). A key role for NOX4 in epithelial cell death during development of lung fibrosis. Antioxid. Redox Signal. 15, 607–619. doi: 10.1089/ars.2010.3829

PubMed Abstract | CrossRef Full Text | Google Scholar

Carnesecchi, S., Deffert, C., Pagano, A., Garrido-Urbani, S., Metrailler-Ruchonnet, I., Schappi, M., et al. (2009). NADPH oxidase-1 plays a crucial role in hyperoxia-induced acute lung injury in mice. Am. J. Respir. Crit. Care Med. 180, 972–981. doi: 10.1164/rccm.200902-0296oc

PubMed Abstract | CrossRef Full Text | Google Scholar

Chanin, R. B., Winter, M. G., Spiga, L., Hughes, E. R., Zhu, W., Taylor, S. J., et al. (2020). Epithelial-derived reactive Oxygen Species enable AppBCX-mediated aerobic respiration of Escherichia coli during intestinal inflammation. Cell Host Microbe 28, 780–788.e5.

Google Scholar

Chauvigne, F., Ducat, C., Ferre, A., Hansen, T., Carrascal, M., Abian, J., et al. (2021). A multiplier peroxiporin signal transduction pathway powers piscine spermatozoa. Proc. Natl. Acad. Sci. U.S.A. 118:e2019346118. doi: 10.1073/pnas.2019346118

PubMed Abstract | CrossRef Full Text | Google Scholar

Chavez, V., Mohri-Shiomi, A., and Garsin, D. A. (2009). Ce-Duox1/BLI-3 generates reactive oxygen species as a protective innate immune mechanism in Caenorhabditis elegans. Infect. Immun. 77, 4983–4989. doi: 10.1128/iai.00627-09

PubMed Abstract | CrossRef Full Text | Google Scholar

Chen, C., Mahar, R., Merritt, M. E., Denlinger, D. L., and Hahn, D. A. (2021). ROS and hypoxia signaling regulate periodic metabolic arousal during insect dormancy to coordinate glucose, amino acid, and lipid metabolism. Proc. Natl. Acad. Sci. U.S.A. 118:e2017603118. doi: 10.1073/pnas.2017603118

PubMed Abstract | CrossRef Full Text | Google Scholar

Chen, H. C., Su, L. T., Gonzalez-Pagan, O., Overton, J. D., and Runnels, L. W. (2012). A key role for Mg(2+) in TRPM7’s control of ROS levels during cell stress. Biochem. J. 445, 441–448. doi: 10.1042/bj20120248

PubMed Abstract | CrossRef Full Text | Google Scholar

Chen, Q., Vazquez, E. J., Moghaddas, S., Hoppel, C. L., and Lesnefsky, E. J. (2003). Production of reactive oxygen species by mitochondria: central role of complex III. J. Biol. Chem. 278, 36027–36031.

Google Scholar

Chen, Y., Azad, M. B., and Gibson, S. B. (2009). Superoxide is the major reactive oxygen species regulating autophagy. Cell Death Differ. 16, 1040–1052. doi: 10.1038/cdd.2009.49

PubMed Abstract | CrossRef Full Text | Google Scholar

Chiarugi, P., Fiaschi, T., Taddei, M. L., Talini, D., Giannoni, E., Raugei, G., et al. (2001). Two vicinal cysteines confer a peculiar redox regulation to low molecular weight protein tyrosine phosphatase in response to platelet-derived growth factor receptor stimulation. J. Biol. Chem. 276, 33478–33487. doi: 10.1074/jbc.m102302200

PubMed Abstract | CrossRef Full Text | Google Scholar

Choi, J., Ifuku, M., Noda, M., and Guilarte, T. R. (2011). Translocator protein (18 kDa)/peripheral benzodiazepine receptor specific ligands induce microglia functions consistent with an activated state. Glia 59, 219–230. doi: 10.1002/glia.21091

PubMed Abstract | CrossRef Full Text | Google Scholar

Cordeiro, R. M. (2014). Reactive oxygen species at phospholipid bilayers: distribution, mobility and permeation. Biochim. Biophys. Acta 1838, 438–444. doi: 10.1016/j.bbamem.2013.09.016

PubMed Abstract | CrossRef Full Text | Google Scholar

Couto, N., Wood, J., and Barber, J. (2016). The role of glutathione reductase and related enzymes on cellular redox homoeostasis network. Free Radic. Biol. Med. 95, 27–42. doi: 10.1016/j.freeradbiomed.2016.02.028

PubMed Abstract | CrossRef Full Text | Google Scholar

Craig, M., and Slauch, J. M. (2009). Phagocytic superoxide specifically damages an extracytoplasmic target to inhibit or kill Salmonella. PLoS One 4:e4975. doi: 10.1371/journal.pone.0004975

PubMed Abstract | CrossRef Full Text | Google Scholar

Cross, C. E., Halliwell, B., Borish, E. T., Pryor, W. A., Ames, B. N., Saul, R. L., et al. (1987). Oxygen radicals and human disease. Ann. Intern. Med. 107, 526–545. doi: 10.7326/0003-4819-107-4-526

PubMed Abstract | CrossRef Full Text | Google Scholar

Dahlgren, C., and Karlsson, A. (1999). Respiratory burst in human neutrophils. J. Immunol. Methods 232, 3–14. doi: 10.1016/s0022-1759(99)00146-5

CrossRef Full Text | Google Scholar

Damiano, S., Sozio, C., La Rosa, G., and Santillo, M. (2020). NOX-dependent signaling dysregulation in severe COVID-19: clues to effective treatments. Front. Cell. Infect. Microbiol. 10:608435. doi: 10.3389/fcimb.2020.608435

PubMed Abstract | CrossRef Full Text | Google Scholar

Das, K., and Roychoudhury, A. (2014). Reactive oxygen species (ROS) and response of antioxidants as ROS-scavengers during environmental stress in plants. Front. Environ. Sci. 2:53. doi: 10.3389/fenvs.2014.00053

CrossRef Full Text | Google Scholar

D’Autreaux, B., and Toledano, M. B. (2007). ROS as signalling molecules: mechanisms that generate specificity in ROS homeostasis. Nat. Rev. Mol. Cell Biol. 8, 813–824. doi: 10.1038/nrm2256

PubMed Abstract | CrossRef Full Text | Google Scholar

Davies, K. J. (2016). Adaptive homeostasis. Mol. Aspects Med. 49, 1–7.

Google Scholar

Davies, M. J., Forni, L. G., and Willson, R. L. (1988). Vitamin E analogue Trolox C. E.s.r. and pulse-radiolysis studies of free-radical reactions. Biochem. J. 255, 513–522.

Google Scholar

Deng, H., Yang, W., Zhou, Z., Tian, R., Lin, L., Ma, Y., et al. (2020). Targeted scavenging of extracellular ROS relieves suppressive immunogenic cell death. Nat. Commun. 11:4951.

Google Scholar

Deshmukh, P., Unni, S., Krishnappa, G., and Padmanabhan, B. (2017). The Keap1-Nrf2 pathway: promising therapeutic target to counteract ROS-mediated damage in cancers and neurodegenerative diseases. Biophys. Rev. 9, 41–56. doi: 10.1007/s12551-016-0244-4

PubMed Abstract | CrossRef Full Text | Google Scholar

Dikalova, A. E., Gongora, M. C., Harrison, D. G., Lambeth, J. D., Dikalov, S., and Griendling, K. K. (2010). Upregulation of Nox1 in vascular smooth muscle leads to impaired endothelium-dependent relaxation via eNOS uncoupling. Am. J. Physiol. Heart Circ. Physiol. 299, H673–H679.

Google Scholar

Dinakar, C., Vishwakarma, A., Raghavendra, A. S., and Padmasree, K. (2016). Alternative oxidase pathway optimizes photosynthesis during osmotic and temperature stress by regulating cellular ROS, malate valve and antioxidative systems. Front. Plant Sci. 7:68. doi: 10.3389/fpls.2016.00068

PubMed Abstract | CrossRef Full Text | Google Scholar

Donko, A., Ruisanchez, E., Orient, A., Enyedi, B., Kapui, R., Peterfi, Z., et al. (2010). Urothelial cells produce hydrogen peroxide through the activation of Duox1. Free Radic. Biol. Med. 49, 2040–2048. doi: 10.1016/j.freeradbiomed.2010.09.027

PubMed Abstract | CrossRef Full Text | Google Scholar

Dostert, C., Petrilli, V., Van Bruggen, R., Steele, C., Mossman, B. T., and Tschopp, J. (2008). Innate immune activation through Nalp3 inflammasome sensing of asbestos and silica. Science 320, 674–677. doi: 10.1126/science.1156995

PubMed Abstract | CrossRef Full Text | Google Scholar

Dugas, A. J. Jr., Castaneda-Acosta, J., Bonin, G. C., Price, K. L., Fischer, N. H., and Winston, G. W. (2000). Evaluation of the total peroxyl radical-scavenging capacity of flavonoids: structure-activity relationships. J. Nat. Prod. 63, 327–331. doi: 10.1021/np990352n

PubMed Abstract | CrossRef Full Text | Google Scholar

Ermakova, Y. G., Bilan, D. S., Matlashov, M. E., Mishina, N. M., Markvicheva, K. N., Subach, O. M., et al. (2014). Red fluorescent genetically encoded indicator for intracellular hydrogen peroxide. Nat. Commun. 5:5222.

Google Scholar

Ernst, O., Sun, J., Lin, B., Banoth, B., Dorrington, M. G., Liang, J., et al. (2021). A genome-wide screen uncovers multiple roles for mitochondrial nucleoside diphosphate kinase D in inflammasome activation. Sci. Signal. 14:eabe0387. doi: 10.1126/scisignal.abe0387

PubMed Abstract | CrossRef Full Text | Google Scholar

Ezerina, D., Takano, Y., Hanaoka, K., Urano, Y., and Dick, T. P. (2018). N-Acetyl cysteine functions as a fast-acting antioxidant by triggering intracellular H2S and Sulfane Sulfur Production. Cell Chem. Biol. 25, 447–459.e4.

Google Scholar

Fan, L. M., Geng, L., Cahill-Smith, S., Liu, F., Douglas, G., Mckenzie, C. A., et al. (2019). Nox2 contributes to age-related oxidative damage to neurons and the cerebral vasculature. J. Clin. Invest. 129, 3374–3386. doi: 10.1172/jci125173

PubMed Abstract | CrossRef Full Text | Google Scholar

Fenton, H. J. H. (1894). LXXIII.—Oxidation of tartaric acid in presence of iron. J. Chem. Soc. Trans. 65, 899–910. doi: 10.1039/ct8946500899

CrossRef Full Text | Google Scholar

Finkel, T. (2011). Signal transduction by reactive oxygen species. J. Cell Biol. 194, 7–15.

Google Scholar

Fransen, M., Lismont, C., and Walton, P. (2017). The peroxisome-mitochondria connection: How and Why? Int. J. Mol. Sci. 18:1126. doi: 10.3390/ijms18061126

PubMed Abstract | CrossRef Full Text | Google Scholar

Fridovich, I. (1997). Superoxide anion radical (O2-.), superoxide dismutases, and related matters. J. Biol. Chem. 272, 18515–18517. doi: 10.1074/jbc.272.30.18515

PubMed Abstract | CrossRef Full Text | Google Scholar

Gaggini, F., Laleu, B., Orchard, M., Fioraso-Cartier, L., Cagnon, L., Houngninou-Molango, S., et al. (2011). Design, synthesis and biological activity of original pyrazolo-pyrido-diazepine, -pyrazine and -oxazine dione derivatives as novel dual Nox4/Nox1 inhibitors. Bioorg. Med. Chem. 19, 6989–6999. doi: 10.1016/j.bmc.2011.10.016

PubMed Abstract | CrossRef Full Text | Google Scholar

Garaude, J., Acin-Perez, R., Martinez-Cano, S., Enamorado, M., Ugolini, M., Nistal-Villan, E., et al. (2016). Mitochondrial respiratory-chain adaptations in macrophages contribute to antibacterial host defense. Nat. Immunol. 17, 1037–1045. doi: 10.1038/ni.3509

PubMed Abstract | CrossRef Full Text | Google Scholar

Gatliff, J., East, D. A., Singh, A., Alvarez, M. S., Frison, M., Matic, I., et al. (2017). A role for TSPO in mitochondrial Ca(2+) homeostasis and redox stress signaling. Cell Death Dis. 8:e2896. doi: 10.1038/cddis.2017.186

PubMed Abstract | CrossRef Full Text | Google Scholar

Gavazzi, G., Banfi, B., Deffert, C., Fiette, L., Schappi, M., Herrmann, F., et al. (2006). Decreased blood pressure in NOX1-deficient mice. FEBS Lett. 580, 497–504. doi: 10.1016/j.febslet.2005.12.049

PubMed Abstract | CrossRef Full Text | Google Scholar

Geng, J., Sun, X., Wang, P., Zhang, S., Wang, X., Wu, H., et al. (2015). Kinases Mst1 and Mst2 positively regulate phagocytic induction of reactive oxygen species and bactericidal activity. Nat. Immunol. 16, 1142–1152. doi: 10.1038/ni.3268

PubMed Abstract | CrossRef Full Text | Google Scholar

Geng, L., Fan, L. M., Liu, F., Smith, C., and Li, J. (2020). Nox2 dependent redox-regulation of microglial response to amyloid-beta stimulation and microgliosis in aging. Sci. Rep. 10:1582.

Google Scholar

Geyer, O., Podos, S. M., and Mittag, T. (1997). Nitric oxide synthase activity in tissues of the bovine eye. Graefes Arch. Clin. Exp. Ophthalmol. 235, 786–793. doi: 10.1007/bf02332864

PubMed Abstract | CrossRef Full Text | Google Scholar

Gluschko, A., Herb, M., Wiegmann, K., Krut, O., Neiss, W. F., Utermohlen, O., et al. (2018). The beta2 Integrin Mac-1 induces protective LC3-associated phagocytosis of Listeria monocytogenes. Cell Host Microbe 23, 324–337.e5.

Google Scholar

Go, Y. M., and Jones, D. P. (2008). Redox compartmentalization in eukaryotic cells. Biochim. Biophys. Acta 1780, 1273–1290. doi: 10.1016/j.bbagen.2008.01.011

PubMed Abstract | CrossRef Full Text | Google Scholar

Gorin, Y., Cavaglieri, R. C., Khazim, K., Lee, D. Y., Bruno, F., Thakur, S., et al. (2015). Targeting NADPH oxidase with a novel dual Nox1/Nox4 inhibitor attenuates renal pathology in type 1 diabetes. Am. J. Physiol. Renal Physiol. 308, F1276–F1287.

Google Scholar

Green, D. E., Murphy, T. C., Kang, B. Y., Kleinhenz, J. M., Szyndralewiez, C., Page, P., et al. (2012). The Nox4 inhibitor GKT137831 attenuates hypoxia-induced pulmonary vascular cell proliferation. Am. J. Respir. Cell Mol. Biol. 47, 718–726. doi: 10.1165/rcmb.2011-0418oc

PubMed Abstract | CrossRef Full Text | Google Scholar

Gutscher, M., Sobotta, M. C., Wabnitz, G. H., Ballikaya, S., Meyer, A. J., Samstag, Y., et al. (2009). Proximity-based protein thiol oxidation by H2O2-scavenging peroxidases. J. Biol. Chem. 284, 31532–31540. doi: 10.1074/jbc.m109.059246

PubMed Abstract | CrossRef Full Text | Google Scholar

Gutteridge, J. M. C., and Halliwell, B. (2018). Mini-Review: Oxidative stress, redox stress or redox success? Biochem. Biophys. Res. Commun. 502, 183–186. doi: 10.1016/j.bbrc.2018.05.045

PubMed Abstract | CrossRef Full Text | Google Scholar

Haber, F., Weiss, J., and Pope, W. J. (1997). The catalytic decomposition of hydrogen peroxide by iron salts. Proc. R. Soc. Lond. Ser. A Math. Phys. Sci. 147, 332–351. doi: 10.1098/rspa.1934.0221

CrossRef Full Text | Google Scholar

Hadrava Vanova, K., Kraus, M., Neuzil, J., and Rohlena, J. (2020). Mitochondrial complex II and reactive oxygen species in disease and therapy. Redox Rep. 25, 26–32. doi: 10.1080/13510002.2020.1752002

PubMed Abstract | CrossRef Full Text | Google Scholar

Haj-Yehia, A. I., Nassar, T., Assaf, P., Nassar, H., and Anggard, E. E. (1999). Effects of the superoxide dismutase-mimic compound TEMPOL on oxidant stress-mediated endothelial dysfunction. Antioxid. Redox Signal. 1, 221–232. doi: 10.1089/ars.1999.1.2-221

PubMed Abstract | CrossRef Full Text | Google Scholar

Han, D., Antunes, F., Canali, R., Rettori, D., and Cadenas, E. (2003). Voltage-dependent anion channels control the release of the superoxide anion from mitochondria to cytosol. J. Biol. Chem. 278, 5557–5563. doi: 10.1074/jbc.m210269200

PubMed Abstract | CrossRef Full Text | Google Scholar

Han, H. (2018). RNA interference to knock down gene expression. Methods Mol. Biol. 1706, 293–302. doi: 10.1007/978-1-4939-7471-9_16

CrossRef Full Text | Google Scholar

Hatz, S., Lambert, J. D., and Ogilby, P. R. (2007). Measuring the lifetime of singlet oxygen in a single cell: addressing the issue of cell viability. Photochem. Photobiol. Sci. 6, 1106–1116. doi: 10.1039/b707313e

PubMed Abstract | CrossRef Full Text | Google Scholar

Hayyan, M., Hashim, M. A., and Alnashef, I. M. (2016). Superoxide ion: generation and chemical implications. Chem. Rev. 116, 3029–3085. doi: 10.1021/acs.chemrev.5b00407

PubMed Abstract | CrossRef Full Text | Google Scholar

Hein, T. W., and Kuo, L. (1998). LDLs impair vasomotor function of the coronary microcirculation: role of superoxide anions. Circ. Res. 83, 404–414. doi: 10.1161/01.res.83.4.404

CrossRef Full Text | Google Scholar

Heinz, S., Freyberger, A., Lawrenz, B., Schladt, L., Schmuck, G., and Ellinger-Ziegelbauer, H. (2017). Mechanistic investigations of the mitochondrial complex I inhibitor rotenone in the context of pharmacological and safety evaluation. Sci. Rep. 7:45465.

Google Scholar

Hempel, S. L., Buettner, G. R., O’malley, Y. Q., Wessels, D. A., and Flaherty, D. M. (1999). Dihydrofluorescein diacetate is superior for detecting intracellular oxidants: comparison with 2′,7′-dichlorodihydrofluorescein diacetate, 5(and 6)-carboxy-2′,7′-dichlorodihydrofluorescein diacetate, and dihydrorhodamine 123. Free Radic. Biol. Med. 27, 146–159.

Google Scholar

Herb, M., and Schramm, M. (2021). Functions of ROS in macrophages and antimicrobial immunity. Antioxidants10:313. doi: 10.3390/antiox10020313

PubMed Abstract | CrossRef Full Text | Google Scholar

Herb, M., Farid, A., Gluschko, A., Kronke, M., and Schramm, M. (2019a). Highly efficient transfection of primary macrophages with in vitro transcribed mRNA. J. Vis. Exp. 153.

Google Scholar

Herb, M., Gluschko, A., Wiegmann, K., Farid, A., Wolf, A., Utermohlen, O., et al. (2019b). Mitochondrial reactive oxygen species enable proinflammatory signaling through disulfide linkage of NEMO. Sci. Signal. 12:eaar5926. doi: 10.1126/scisignal.aar5926

PubMed Abstract | CrossRef Full Text | Google Scholar

Hernández-Barrera, A., Quinto, C., Johnson, E. A., Wu, H.-M., Cheung, A. Y., and Cárdenas, L. (2013). Using hyper as a molecular probe to visualize hydrogen peroxide in living plant cells: a method with virtually unlimited potential in plant biology. Methods Enzymol. 527, 275–290. doi: 10.1016/b978-0-12-405882-8.00015-5

PubMed Abstract | CrossRef Full Text | Google Scholar

Hernansanz-Agustin, P., Choya-Foces, C., Carregal-Romero, S., Ramos, E., Oliva, T., Villa-Pina, T., et al. (2020). Na(+) controls hypoxic signalling by the mitochondrial respiratory chain. Nature 586, 287–291.

Google Scholar

Herscovitch, M., Comb, W., Ennis, T., Coleman, K., Yong, S., Armstead, B., et al. (2008). Intermolecular disulfide bond formation in the NEMO dimer requires Cys54 and Cys347. Biochem. Biophys. Res. Commun. 367, 103–108. doi: 10.1016/j.bbrc.2007.12.123

PubMed Abstract | CrossRef Full Text | Google Scholar

Hervera, A., De Virgiliis, F., Palmisano, I., Zhou, L., Tantardini, E., Kong, G., et al. (2018). Reactive oxygen species regulate axonal regeneration through the release of exosomal NADPH oxidase 2 complexes into injured axons. Nat. Cell Biol. 20, 307–319. doi: 10.1038/s41556-018-0039-x

PubMed Abstract | CrossRef Full Text | Google Scholar

Heumuller, S., Wind, S., Barbosa-Sicard, E., Schmidt, H. H., Busse, R., Schroder, K., et al. (2008). Apocynin is not an inhibitor of vascular NADPH oxidases but an antioxidant. Hypertension 51, 211–217. doi: 10.1161/hypertensionaha.107.100214

PubMed Abstract | CrossRef Full Text | Google Scholar

Hoeven, R., Mccallum, K. C., Cruz, M. R., and Garsin, D. A. (2011). Ce-Duox1/BLI-3 generated reactive oxygen species trigger protective SKN-1 activity via p38 MAPK signaling during infection in C. elegans. PLoS Pathog. 7:e1002453. doi: 10.1371/journal.ppat.1002453

PubMed Abstract | CrossRef Full Text | Google Scholar

Holmstrom, K. M., and Finkel, T. (2014). Cellular mechanisms and physiological consequences of redox-dependent signalling. Nat. Rev. Mol. Cell Biol. 15, 411–421. doi: 10.1038/nrm3801

PubMed Abstract | CrossRef Full Text | Google Scholar

Hos, N. J., Ganesan, R., Gutierrez, S., Hos, D., Klimek, J., Abdullah, Z., et al. (2017). Type I interferon enhances necroptosis of Salmonella Typhimurium-infected macrophages by impairing antioxidative stress responses. J. Cell Biol. 216, 4107–4121.

Google Scholar

Huang, H., Ullah, F., Zhou, D. X., Yi, M., and Zhao, Y. (2019). Mechanisms of ROS regulation of plant development and stress responses. Front. Plant Sci. 10:800. doi: 10.3389/fpls.2019.00800

PubMed Abstract | CrossRef Full Text | Google Scholar

Imlay, J. A. (2003). Pathways of oxidative damage. Annu. Rev. Microbiol. 57, 395–418. doi: 10.1146/annurev.micro.57.030502.090938

PubMed Abstract | CrossRef Full Text | Google Scholar

Inomata, M., Xu, S., Chandra, P., Meydani, S. N., Takemura, G., Philips, J. A., et al. (2020). Macrophage LC3-associated phagocytosis is an immune defense against Streptococcus pneumoniae that diminishes with host aging. Proc. Natl. Acad. Sci. U.S.A. 117, 33561–33569. doi: 10.1073/pnas.2015368117

PubMed Abstract | CrossRef Full Text | Google Scholar

Ives, A., Nomura, J., Martinon, F., Roger, T., Leroy, D., Miner, J. N., et al. (2015). Xanthine oxidoreductase regulates macrophage IL1beta secretion upon NLRP3 inflammasome activation. Nat. Commun. 6:6555.

Google Scholar

Janku, M., Luhova, L., and Petrivalsky, M. (2019). On the Origin and Fate of Reactive Oxygen Species in Plant Cell Compartments. Antioxidants 8:105. doi: 10.3390/antiox8040105

PubMed Abstract | CrossRef Full Text | Google Scholar

Jones, A. I., Meshulam, T., Oliveira, M. F., Burritt, N., and Corkey, B. E. (2016). Extracellular redox regulation of intracellular reactive oxygen generation, mitochondrial function and lipid turnover in cultured human Adipocytes. PLoS One 11:e0164011. doi: 10.1371/journal.pone.0164011

PubMed Abstract | CrossRef Full Text | Google Scholar

Juan, C. A., De La Lastra, J. M. P., Plou, F. J., and Perez-Lebena, E. (2021). 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. 22, 4642.

Google Scholar

Kamata, H., Honda, S., Maeda, S., Chang, L., Hirata, H., and Karin, M. (2005). Reactive oxygen species promote TNFalpha-induced death and sustained JNK activation by inhibiting MAP kinase phosphatases. Cell 120, 649–661. doi: 10.1016/j.cell.2004.12.041

PubMed Abstract | CrossRef Full Text | Google Scholar

Kehrer, J. P., Robertson, J. D., and Smith, C. V. (2010). “1.14 - Free radicals and reactive oxygen species,” in Comprehensive Toxicology, 2nd Edn, ed. C. A. Mcqueen (Oxford: Elsevier), 277–307.

Google Scholar

Kelley, E. E., Khoo, N. K., Hundley, N. J., Malik, U. Z., Freeman, B. A., and Tarpey, M. M. (2010). Hydrogen peroxide is the major oxidant product of xanthine oxidase. Free Radic. Biol. Med. 48, 493–498. doi: 10.1016/j.freeradbiomed.2009.11.012

PubMed Abstract | CrossRef Full Text | Google Scholar

Kelly, B., Tannahill, G. M., Murphy, M. P., and O’neill, L. A. (2015). Metformin Inhibits the Production of Reactive Oxygen Species from NADH:Ubiquinone Oxidoreductase to Limit Induction of Interleukin-1beta (IL-1beta) and Boosts Interleukin-10 (IL-10) in Lipopolysaccharide (LPS)-activated Macrophages. J. Biol. Chem. 290, 20348–20359. doi: 10.1074/jbc.m115.662114

PubMed Abstract | CrossRef Full Text | Google Scholar

Kim, H. K., Lee, H. Y., Riaz, T. A., Bhattarai, K. R., Chaudhary, M., Ahn, J. H., et al. (2021). Chalcone suppresses tumor growth through NOX4-IRE1alpha sulfonation-RIDD-miR-23b axis. Redox Biol. 40:101853. doi: 10.1016/j.redox.2021.101853

PubMed Abstract | CrossRef Full Text | Google Scholar

Kim, H., and Xue, X. (2020). Detection of total reactive Oxygen Species in Adherent Cells by 2′,7′-Dichlorodihydrofluorescein Diacetate Staining. J. Vis. Exp. 160:10.3791/60682. doi: 10.3791/60682

PubMed Abstract | CrossRef Full Text | Google Scholar

Kim, J. H., Lee, J., Bae, S. J., Kim, Y., Park, B. J., Choi, J. W., et al. (2017). NADPH oxidase 4 is required for the generation of macrophage migration inhibitory factor and host defense against Toxoplasma gondii infection. Sci. Rep. 7:6361.

Google Scholar

Kim, N. H., Choi, S., Han, E. J., Hong, B. K., Choi, S. Y., Kwon, H. M., et al. (2014). The xanthine oxidase-NFAT5 pathway regulates macrophage activation and TLR-induced inflammatory arthritis. Eur. J. Immunol. 44, 2721–2736. doi: 10.1002/eji.201343669

PubMed Abstract | CrossRef Full Text | Google Scholar

Kiritoshi, S., Nishikawa, T., Sonoda, K., Kukidome, D., Senokuchi, T., Matsuo, T., et al. (2003). Reactive oxygen species from mitochondria induce cyclooxygenase-2 gene expression in human mesangial cells: potential role in diabetic nephropathy. Diabetes Metab. Res. Rev. 52, 2570–2577. doi: 10.2337/diabetes.52.10.2570

PubMed Abstract | CrossRef Full Text | Google Scholar

Kirkman, H. N., and Gaetani, G. F. (2007). Mammalian catalase: a venerable enzyme with new mysteries. Trends Biochem. Sci. 32, 44–50. doi: 10.1016/j.tibs.2006.11.003

PubMed Abstract | CrossRef Full Text | Google Scholar

Knight, L. J., Martis, R. M., Acosta, M. L., Donaldson, P. J., and Lim, J. C. (2021). Detection of reduced mitochondrial ROS production but increased ROS levels and oxidative damage in the young xCT knockout mouse retina. Invest. Ophthalmol. Vis. Sci. 62, 2230–2230.

Google Scholar

Koterski, J. F., Nahvi, M., Venkatesan, M. M., and Haimovich, B. (2005). Virulent Shigella flexneri causes damage to mitochondria and triggers necrosis in infected human monocyte-derived macrophages. Infect. Immun. 73, 504–513. doi: 10.1128/iai.73.1.504-513.2005

PubMed Abstract | CrossRef Full Text | Google Scholar

Krishna, C. M., Liebmann, J. E., Kaufman, D., Degraff, W., Hahn, S. M., Mcmurry, T., et al. (1992). The catecholic metal sequestering agent 1,2-dihydroxybenzene-3,5-disulfonate confers protection against oxidative cell damage. Arch. Biochem. Biophys. 294, 98–106. doi: 10.1016/0003-9861(92)90142-j

CrossRef Full Text | Google Scholar

Laleu, B., Gaggini, F., Orchard, M., Fioraso-Cartier, L., Cagnon, L., Houngninou-Molango, S., et al. (2010). First in class, potent, and orally bioavailable NADPH oxidase isoform 4 (Nox4) inhibitors for the treatment of idiopathic pulmonary fibrosis. J. Med. Chem. 53, 7715–7730. doi: 10.1021/jm100773e

PubMed Abstract | CrossRef Full Text | Google Scholar

Lambert, A. J., and Brand, M. D. (2004). Inhibitors of the quinone-binding site allow rapid superoxide production from mitochondrial NADH:ubiquinone oxidoreductase (complex I). J. Biol. Chem. 279, 39414–39420. doi: 10.1074/jbc.m406576200

PubMed Abstract | CrossRef Full Text | Google Scholar

Lambeth, J. D., and Neish, A. S. (2014). Nox enzymes and new thinking on reactive oxygen: a double-edged sword revisited. Annu. Rev. Pathol. 9, 119–145. doi: 10.1146/annurev-pathol-012513-104651

PubMed Abstract | CrossRef Full Text | Google Scholar

Lambeth, J. D., Krause, K. H., and Clark, R. A. (2008). NOX enzymes as novel targets for drug development. Semin. Immunopathol. 30, 339–363. doi: 10.1007/s00281-008-0123-6

PubMed Abstract | CrossRef Full Text | Google Scholar

Lanciano, P., Khalfaoui-Hassani, B., Selamoglu, N., Ghelli, A., Rugolo, M., and Daldal, F. (2013). Molecular mechanisms of superoxide production by complex III: a bacterial versus human mitochondrial comparative case study. Biochim. Biophys. Acta 1827, 1332–1339. doi: 10.1016/j.bbabio.2013.03.009

PubMed Abstract | CrossRef Full Text | Google Scholar

Ledford, H. (2015). CRISPR, the disruptor. Nature 522, 20–24. doi: 10.1038/522020a

PubMed Abstract | CrossRef Full Text | Google Scholar

Lee, E. Y., Bae, H. C., Lee, H., Jang, Y., Park, Y. H., Kim, J. H., et al. (2017). Intracellular ROS levels determine the apoptotic potential of keratinocyte by Quantum Dot via blockade of AKT Phosphorylation. Exp. Dermatol. 26, 1046–1052. doi: 10.1111/exd.13365

PubMed Abstract | CrossRef Full Text | Google Scholar

Lee, S. R., Kwon, K. S., Kim, S. R., and Rhee, S. G. (1998). Reversible inactivation of protein-tyrosine phosphatase 1B in A431 cells stimulated with epidermal growth factor. J. Biol. Chem. 273, 15366–15372. doi: 10.1074/jbc.273.25.15366

PubMed Abstract | CrossRef Full Text | Google Scholar

Leusen, J. H., Fluiter, K., Hilarius, P. M., Roos, D., Verhoeven, A. J., and Bolscher, B. G. (1995). Interactions between the cytosolic components p47phox and p67phox of the human neutrophil NADPH oxidase that are not required for activation in the cell-free system. J. Biol. Chem. 270, 11216–11221.

Google Scholar

Levine, A., Tenhaken, R., Dixon, R., and Lamb, C. (1994). H2O2 from the oxidative burst orchestrates the plant hypersensitive disease resistance response. Cell 79, 583–593. doi: 10.1016/0092-8674(94)90544-4

CrossRef Full Text | Google Scholar

Li, T. Y., Sun, Y., Liang, Y., Liu, Q., Shi, Y., Zhang, C. S., et al. (2016). ULK1/2 constitute a bifurcate node controlling glucose metabolic fluxes in addition to autophagy. Mol. Cell 62, 359–370. doi: 10.1016/j.molcel.2016.04.009

PubMed Abstract | CrossRef Full Text | Google Scholar

Li, X., Fang, P., Mai, J., Choi, E. T., Wang, H., and Yang, X. F. (2013). Targeting mitochondrial reactive oxygen species as novel therapy for inflammatory diseases and cancers. J. Hematol. Oncol. 6:19. doi: 10.1186/1756-8722-6-19

PubMed Abstract | CrossRef Full Text | Google Scholar

Ligeon, L. A., Pena-Francesch, M., and Munz, C. (2021a). Measuring oxidation within LC3-associated phagosomes that optimizes MHC class II restricted antigen presentation. Methods Cell Biol. 164, 187–200. doi: 10.1016/bs.mcb.2021.02.003

PubMed Abstract | CrossRef Full Text | Google Scholar

Ligeon, L. A., Pena-Francesch, M., Vanoaica, L. D., Nunez, N. G., Talwar, D., Dick, T. P., et al. (2021b). Oxidation inhibits autophagy protein deconjugation from phagosomes to sustain MHC class II restricted antigen presentation. Nat. Commun. 12:1508.

Google Scholar

Lin, T. K., Lin, K. J., Lin, K. L., Liou, C. W., Chen, S. D., Chuang, Y. C., et al. (2020). When friendship turns sour: effective communication between mitochondria and intracellular organelles in Parkinson’s disease. Front. Cell. Dev. Biol. 8:607392. doi: 10.3389/fcell.2020.607392

PubMed Abstract | CrossRef Full Text | Google Scholar

Liochev, S. I., and Fridovich, I. (1999). Superoxide and iron: partners in crime. IUBMB Life 48, 157–161. doi: 10.1080/713803492

PubMed Abstract | CrossRef Full Text | Google Scholar

Liou, G. Y., and Storz, P. (2010). Reactive oxygen species in cancer. Free Radic. Res. 44, 479–496.

Google Scholar

Lismont, C., Nordgren, M., Van Veldhoven, P. P., and Fransen, M. (2015). Redox interplay between mitochondria and peroxisomes. Front. Cell Dev. Biol. 3:35. doi: 10.3389/fcell.2015.00035

PubMed Abstract | CrossRef Full Text | Google Scholar

Loth, M. K., Guariglia, S. R., Re, D. B., Perez, J., De Paiva, V. N., Dziedzic, J. L., et al. (2020). A novel interaction of translocator protein 18 kDa (TSPO) with NADPH Oxidase in Microglia. Mol. Neurobiol. 57, 4467–4487. doi: 10.1007/s12035-020-02042-w

PubMed Abstract | CrossRef Full Text | Google Scholar

Lundqvist, H., and Dahlgren, C. (1996). Isoluminol-enhanced chemiluminescence: a sensitive method to study the release of superoxide anion from human neutrophils. Free Radic. Biol. Med. 20, 785–792. doi: 10.1016/0891-5849(95)02189-2

CrossRef Full Text | Google Scholar

Mak, T. W., Grusdat, M., Duncan, G. S., Dostert, C., Nonnenmacher, Y., Cox, M., et al. (2017). Glutathione primes T cell metabolism for inflammation. Immunity 46, 675–689. doi: 10.1016/j.immuni.2017.03.019

PubMed Abstract | CrossRef Full Text | Google Scholar

Manzano, V. M., Munoz, J. C., Jimenez, J. R., Puyol, M. R., Puyol, D. R., Kitamura, M., et al. (2000). Human renal mesangial cells are a target for the anti-inflammatory action of 9-cis retinoic acid. Br. J. Pharmacol. 131, 1673–1683. doi: 10.1038/sj.bjp.0703728

PubMed Abstract | CrossRef Full Text | Google Scholar

Marinho, H. S., Cyrne, L., Cadenas, E., and Antunes, F. (2013). The cellular steady-state of H2O2: latency concepts and gradients. Methods Enzymol. 527, 3–19. doi: 10.1016/b978-0-12-405882-8.00001-5

PubMed Abstract | CrossRef Full Text | Google Scholar

Marklund, S. (1976). Spectrophotometric study of spontaneous disproportionation of superoxide anion radical and sensitive direct assay for superoxide dismutase. J. Biol. Chem. 251, 7504–7507. doi: 10.1016/s0021-9258(17)32878-8

CrossRef Full Text | Google Scholar

Markvicheva, K. N., Bilan, D. S., Mishina, N. M., Gorokhovatsky, A. Y., Vinokurov, L. M., Lukyanov, S., et al. (2011). A genetically encoded sensor for H2O2 with expanded dynamic range. Bioorg. Med. Chem. 19, 1079–1084. doi: 10.1016/j.bmc.2010.07.014

PubMed Abstract | CrossRef Full Text | Google Scholar

Martinez, M. A., Ubeda, A., and Trillo, M. A. (2021). Role of NADPH oxidase in MAPK signaling activation by a 50 Hz magnetic field in human neuroblastoma cells. Electromagn. Biol. Med. 40, 103–116. doi: 10.1080/15368378.2020.1851250

PubMed Abstract | CrossRef Full Text | Google Scholar

Matsushita, T., Fukuda, K., Yamamoto, H., Yamazaki, K., Tomiyama, T., Oh, M., et al. (2004). Effect of ebselen, a scavenger of reactive oxygen species, on chondrocyte metabolism. Mod. Rheumatol. 14, 25–30. doi: 10.3109/s10165-003-0261-6

CrossRef Full Text | Google Scholar

Meinhard, M., and Grill, E. (2001). Hydrogen peroxide is a regulator of ABI1, a protein phosphatase 2C from Arabidopsis. FEBS Lett. 508, 443–446. doi: 10.1016/s0014-5793(01)03106-4

CrossRef Full Text | Google Scholar

Melov, S., Coskun, P., Patel, M., Tuinstra, R., Cottrell, B., Jun, A. S., et al. (1999). Mitochondrial disease in superoxide dismutase 2 mutant mice. Proc. Natl. Acad. Sci. U.S.A. 96, 846–851.

Google Scholar

Mendiola, A. S., Ryu, J. K., Bardehle, S., Meyer-Franke, A., Ang, K. K., Wilson, C., et al. (2020). Transcriptional profiling and therapeutic targeting of oxidative stress in neuroinflammation. Nat. Immunol. 21, 513–524.

Google Scholar

Meng, T. C., Fukada, T., and Tonks, N. K. (2002). Reversible oxidation and inactivation of protein tyrosine phosphatases in vivo. Mol. Cell 9, 387–399. doi: 10.1016/s1097-2765(02)00445-8

CrossRef Full Text | Google Scholar

Miller, B. M., Liou, M. J., Zhang, L. F., Nguyen, H., Litvak, Y., Schorr, E. M., et al. (2020). Anaerobic respiration of NOX1-derived hydrogen peroxide licenses bacterial growth at the colonic surface. Cell Host Microbe 28, 789–797.e5.

Google Scholar

Mora-Pale, M., Weiwer, M., Yu, J., Linhardt, R. J., and Dordick, J. S. (2009). Inhibition of human vascular NADPH oxidase by apocynin derived oligophenols. Bioorg. Med. Chem. 17, 5146–5152. doi: 10.1016/j.bmc.2009.05.061

PubMed Abstract | CrossRef Full Text | Google Scholar

Mugesh, G. (2013). Glutathione peroxidase activity of ebselen and its analogues: some insights into the complex chemical mechanisms underlying the antioxidant activity. Curr. Chem. Biol. 7, 47–56. doi: 10.2174/2212796811307010005

CrossRef Full Text | Google Scholar

Mukhopadhyay, P., Rajesh, M., Hasko, G., Hawkins, B. J., Madesh, M., and Pacher, P. (2007). Simultaneous detection of apoptosis and mitochondrial superoxide production in live cells by flow cytometry and confocal microscopy. Nat. Protoc. 2, 2295–2301. doi: 10.1038/nprot.2007.327

PubMed Abstract | CrossRef Full Text | Google Scholar

Muller, F. L., Liu, Y., and Van Remmen, H. (2004). Complex III releases superoxide to both sides of the inner mitochondrial membrane. J. Biol. Chem. 279, 49064–49073. doi: 10.1074/jbc.m407715200

PubMed Abstract | CrossRef Full Text | Google Scholar

Murphy, M. P. (2009). How mitochondria produce reactive oxygen species. Biochem. J. 417, 1–13. doi: 10.1042/bj20081386

PubMed Abstract | CrossRef Full Text | Google Scholar

Nakamura, H. (2005). Thioredoxin and its related molecules: update 2005. Antioxid. Redox Signal. 7, 823–828. doi: 10.1089/ars.2005.7.823

PubMed Abstract | CrossRef Full Text | Google Scholar

Nakamura, Y., Feng, Q., Kumagai, T., Torikai, K., Ohigashi, H., Osawa, T., et al. (2002). Ebselen, a glutathione peroxidase mimetic seleno-organic compound, as a multifunctional antioxidant. Implication for inflammation-associated carcinogenesis. J. Biol. Chem. 277, 2687–2694. doi: 10.1074/jbc.m109641200

PubMed Abstract | CrossRef Full Text | Google Scholar

Nakano, Y., Longo-Guess, C. M., Bergstrom, D. E., Nauseef, W. M., Jones, S. M., and Banfi, B. (2008). Mutation of the Cyba gene encoding p22phox causes vestibular and immune defects in mice. J. Clin. Invest. 118, 1176–1185.

Google Scholar

Nathan, C., and Cunningham-Bussel, A. (2013). Beyond oxidative stress: an immunologist’s guide to reactive oxygen species. Nat. Rev. Immunol. 13, 349–361. doi: 10.1038/nri3423

PubMed Abstract | CrossRef Full Text | Google Scholar

Naviaux, R. K. (2012). Oxidative shielding or oxidative stress? J. Pharmacol. Exp. Ther. 342, 608–618. doi: 10.1124/jpet.112.192120

PubMed Abstract | CrossRef Full Text | Google Scholar

Nazarewicz, R. R., Dikalova, A. E., Bikineyeva, A., and Dikalov, S. I. (2013). Nox2 as a potential target of mitochondrial superoxide and its role in endothelial oxidative stress. Am. J. Physiol. Heart Circ. Physiol. 305, H1131– H1140.

Google Scholar

Niethammer, P., Grabher, C., Look, A. T., and Mitchison, T. J. (2009). A tissue-scale gradient of hydrogen peroxide mediates rapid wound detection in Zebrafish. Nature 459, 996–999. doi: 10.1038/nature08119

PubMed Abstract | CrossRef Full Text | Google Scholar

Niki, E. (2016). Oxidative stress and antioxidants: Distress or eustress? Arch. Biochem. Biophys. 595, 19–24. doi: 10.1016/j.abb.2015.11.017

PubMed Abstract | CrossRef Full Text | Google Scholar

Nohl, H., Kozlov, A. V., Gille, L., and Staniek, K. (2003). Cell respiration and formation of reactive oxygen species: facts and artefacts. Biochem. Soc. Trans. 31, 1308–1311. doi: 10.1042/bst0311308

PubMed Abstract | CrossRef Full Text | Google Scholar

Nomura, J., Busso, N., Ives, A., Matsui, C., Tsujimoto, S., Shirakura, T., et al. (2014). Xanthine oxidase inhibition by febuxostat attenuates experimental atherosclerosis in mice. Sci. Rep. 4:4554.

Google Scholar

Odobasic, D., Kitching, A. R., and Holdsworth, S. R. (2016). Neutrophil-mediated regulation of innate and adaptive immunity: the role of Myeloperoxidase. J. Immunol. Res. 2016:2349817.

Google Scholar

O’Donnell, B. V., Tew, D. G., Jones, O. T., and England, P. J. (1993). Studies on the inhibitory mechanism of iodonium compounds with special reference to neutrophil NADPH oxidase. Biochem. J. 290(Pt 1), 41–49. doi: 10.1042/bj2900041

PubMed Abstract | CrossRef Full Text | Google Scholar

Omura, T., and Sato, R. (1962). A new cytochrome in liver microsomes. J. Biol. Chem. 237, 1375–1376.

Google Scholar

Panov, A., Dikalov, S., Shalbuyeva, N., Taylor, G., Sherer, T., and Greenamyre, J. T. (2005). Rotenone model of Parkinson disease: multiple brain mitochondria dysfunctions after short term systemic rotenone intoxication. J. Biol. Chem. 280, 42026–42035.

Google Scholar

Patriarca, S., Furfaro, A. L., Domenicotti, C., Odetti, P., Cottalasso, D., Marinari, U. M., et al. (2005). Supplementation with N-acetylcysteine and taurine failed to restore glutathione content in liver of streptozotocin-induced diabetics rats but protected from oxidative stress. Biochim. Biophys. Acta 1741, 48–54. doi: 10.1016/j.bbadis.2005.04.003

PubMed Abstract | CrossRef Full Text | Google Scholar

Pavelkova, M., and Kubala, L. (2004). Luminol-, isoluminol- and lucigenin-enhanced chemiluminescence of rat blood phagocytes stimulated with different activators. Luminescence 19, 37–42. doi: 10.1002/bio.754

PubMed Abstract | CrossRef Full Text | Google Scholar

Peng, H. Y., Lucavs, J., Ballard, D., Das, J. K., Kumar, A., Wang, L., et al. (2021). Metabolic reprogramming and reactive Oxygen Species in T cell immunity. Front. Immunol. 12:652687. doi: 10.3389/fimmu.2021.652687

PubMed Abstract | CrossRef Full Text | Google Scholar

Perillo, B., Di Donato, M., Pezone, A., Di Zazzo, E., Giovannelli, P., Galasso, G., et al. (2020). ROS in cancer therapy: the bright side of the moon. Exp. Mol. Med. 52, 192–203. doi: 10.1038/s12276-020-0384-2

PubMed Abstract | CrossRef Full Text | Google Scholar

Pizzino, G., Irrera, N., Cucinotta, M., Pallio, G., Mannino, F., Arcoraci, V., et al. (2017). Oxidative stress: harms and benefits for human health. Oxid. Med. Cell. Longev. 2017:8416763.

Google Scholar

Poljsak, B., Suput, D., and Milisav, I. (2013). Achieving the balance between ROS and antioxidants: when to use the synthetic antioxidants. Oxid. Med. Cell. Longev. 2013:956792.

Google Scholar

Poole, L. B. (2015). The basics of thiols and cysteines in redox biology and chemistry. Free Radic. Biol. Med. 80, 148–157. doi: 10.1016/j.freeradbiomed.2014.11.013

PubMed Abstract | CrossRef Full Text | Google Scholar

Prestes, E. B., Alves, L. S., Rodrigues, D. A. S., Dutra, F. F., Fernandez, P. L., Paiva, C. N., et al. (2020). Mitochondrial reactive oxygen species participate in signaling triggered by Heme in macrophages and upon Hemolysis. J. Immunol. 205, 2795–2805. doi: 10.4049/jimmunol.1900886

PubMed Abstract | CrossRef Full Text | Google Scholar

Prousek, J. (2007). Fenton chemistry in biology and medicine. Pure Appl. Chem. 79, 2325–2338. doi: 10.1351/pac200779122325

CrossRef Full Text | Google Scholar

Quinlan, C. L., Orr, A. L., Perevoshchikova, I. V., Treberg, J. R., Ackrell, B. A., and Brand, M. D. (2012). Mitochondrial complex II can generate reactive oxygen species at high rates in both the forward and reverse reactions. J. Biol. Chem. 287, 27255–27264. doi: 10.1074/jbc.m112.374629

PubMed Abstract | CrossRef Full Text | Google Scholar

Rak, R., Chao, D. L., Pluta, R. M., Mitchell, J. B., Oldfield, E. H., and Watson, J. C. (2000). Neuroprotection by the stable nitroxide Tempol during reperfusion in a rat model of transient focal ischemia. J. Neurosurg. 92, 646–651. doi: 10.3171/jns.2000.92.4.0646

PubMed Abstract | CrossRef Full Text | Google Scholar

Rangasamy, T., Guo, J., Mitzner, W. A., Roman, J., Singh, A., Fryer, A. D., et al. (2005). Disruption of Nrf2 enhances susceptibility to severe airway inflammation and asthma in mice. J. Exp. Med. 202, 47–59. doi: 10.1084/jem.20050538

PubMed Abstract | CrossRef Full Text | Google Scholar

Rapoport, R., Hanukoglu, I., and Sklan, D. (1994). A fluorimetric assay for hydrogen peroxide, suitable for NAD(P)H-dependent superoxide generating redox systems. Anal. Biochem. 218, 309–313. doi: 10.1006/abio.1994.1183

PubMed Abstract | CrossRef Full Text | Google Scholar

Reczek, C. R., and Chandel, N. S. (2014). ROS-dependent signal transduction. Curr. Opin. Cell Biol. 33C, 8–13. doi: 10.1016/j.ceb.2014.09.010

PubMed Abstract | CrossRef Full Text | Google Scholar

Reczek, C. R., and Chandel, N. S. (2017). The two faces of reactive oxygen species in cancer. Annu. Rev. Cancer Biol. 1, 79–98. doi: 10.1146/annurev-cancerbio-041916-065808

CrossRef Full Text | Google Scholar

Reis, J., Massari, M., Marchese, S., Ceccon, M., Aalbers, F. S., Corana, F., et al. (2020). A closer look into NADPH oxidase inhibitors: validation and insight into their mechanism of action. Redox Biol. 32:101466. doi: 10.1016/j.redox.2020.101466

PubMed Abstract | CrossRef Full Text | Google Scholar

Ren, X., Wang, M., Wang, Y., and Huang, A. (2021). Superoxide anion generation response to wound in Arabidopsis hypocotyl cutting. Plant Signal. Behav. 16:1848086. doi: 10.1080/15592324.2020.1848086

PubMed Abstract | CrossRef Full Text | Google Scholar

Rhee, S. G. (2006). Cell signaling. H2O2, a necessary evil for cell signaling. Science 312, 1882–1883. doi: 10.1126/science.1130481

PubMed Abstract | CrossRef Full Text | Google Scholar

Rieger, S., and Sagasti, A. (2011). Hydrogen peroxide promotes injury-induced peripheral sensory axon regeneration in the Zebrafish skin. PLoS Biol. 9:e1000621. doi: 10.1371/journal.pbio.1000621

PubMed Abstract | CrossRef Full Text | Google Scholar

Robinson, K. M., Janes, M. S., Pehar, M., Monette, J. S., Ross, M. F., Hagen, T. M., et al. (2006). Selective fluorescent imaging of superoxide in vivo using ethidium-based probes. Proc. Natl. Acad. Sci. U.S.A. 103, 15038–15043. doi: 10.1073/pnas.0601945103

PubMed Abstract | CrossRef Full Text | Google Scholar

Roca, F. J., and Ramakrishnan, L. (2013). TNF dually mediates resistance and susceptibility to mycobacteria via mitochondrial reactive oxygen species. Cell 153, 521–534. doi: 10.1016/j.cell.2013.03.022

PubMed Abstract | CrossRef Full Text | Google Scholar

Roca, F. J., Whitworth, L. J., Redmond, S., Jones, A. A., and Ramakrishnan, L. (2019). TNF induces pathogenic programmed macrophage necrosis in tuberculosis through a mitochondrial-lysosomal-endoplasmic reticulum circuit. Cell 178, 1344–1361.e11.

Google Scholar

Romero, L. C., Aroca, M. A., Laureano-Marin, A. M., Moreno, I., Garcia, I., and Gotor, C. (2014). Cysteine and cysteine-related signaling pathways in Arabidopsis thaliana. Mol. Plant 7, 264–276. doi: 10.1093/mp/sst168

PubMed Abstract | CrossRef Full Text | Google Scholar

Saito, Y., and Kimura, W. (2021). Roles of phase separation for cellular redox maintenance. Front. Genet. 12:691946. doi: 10.3389/fgene.2021.691946

PubMed Abstract | CrossRef Full Text | Google Scholar

Schumacker, P. T. (2006). Reactive oxygen species in cancer cells: live by the sword, die by the sword. Cancer Cell 10, 175–176. doi: 10.1016/j.ccr.2006.08.015

PubMed Abstract | CrossRef Full Text | Google Scholar

Scialo, F., Fernandez-Ayala, D. J., and Sanz, A. (2017). Role of mitochondrial reverse electron transport in ROS signaling: potential roles in health and disease. Front. Physiol. 8:428. doi: 10.3389/fphys.2017.00428

PubMed Abstract | CrossRef Full Text | Google Scholar

Sedeek, M., Callera, G., Montezano, A., Gutsol, A., Heitz, F., Szyndralewiez, C., et al. (2010). Critical role of Nox4-based NADPH oxidase in glucose-induced oxidative stress in the kidney: implications in type 2 diabetic nephropathy. Am. J. Physiol. Renal Physiol. 299, F1348–F1358.

Google Scholar

Sena, L. A., Li, S., Jairaman, A., Prakriya, M., Ezponda, T., Hildeman, D. A., et al. (2013). Mitochondria are required for antigen-specific T cell activation through reactive oxygen species signaling. Immunity 38, 225–236. doi: 10.1016/j.immuni.2012.10.020

PubMed Abstract | CrossRef Full Text | Google Scholar

Shai, N., Yifrach, E., Van Roermund, C. W. T., Cohen, N., Bibi, C., IJlst, L., et al. (2018). Systematic mapping of contact sites reveals tethers and a function for the peroxisome-mitochondria contact. Nat. Commun. 9:1761.

Google Scholar

Sharma, P., Jha, A. B., Dubey, R. S., and Pessarakli, M. (2012). Reactive oxygen Species, oxidative damage, and antioxidative defense mechanism in plants under stressful conditions. J. Bot. 2012:217037.

Google Scholar

Short, J. D., Downs, K., Tavakoli, S., and Asmis, R. (2016). Protein Thiol Redox signaling in monocytes and macrophages. Antioxid. Redox Signal. 25, 816–835. doi: 10.1089/ars.2016.6697

PubMed Abstract | CrossRef Full Text | Google Scholar

Shoshan-Barmatz, V., De Pinto, V., Zweckstetter, M., Raviv, Z., Keinan, N., and Arbel, N. (2010). VDAC, a multi-functional mitochondrial protein regulating cell life and death. Mol. Aspects Med. 31, 227–285. doi: 10.1016/j.mam.2010.03.002

PubMed Abstract | CrossRef Full Text | Google Scholar

Sies, H. (2021). Oxidative eustress: on constant alert for redox homeostasis. Redox Biol. 41:101867. doi: 10.1016/j.redox.2021.101867

PubMed Abstract | CrossRef Full Text | Google Scholar

Sies, H., and Jones, D. P. (2020). Reactive oxygen species (ROS) as pleiotropic physiological signalling agents. Nat. Rev. Mol. Cell Biol. 21, 363–383. doi: 10.1038/s41580-020-0230-3

PubMed Abstract | CrossRef Full Text | Google Scholar

Sies, H., Berndt, C., and Jones, D. P. (2017). Oxidative stress. Annu. Rev. Biochem. 86, 715–748.

Google Scholar

Stavru, F., Bouillaud, F., Sartori, A., Ricquier, D., and Cossart, P. (2011). Listeria monocytogenes transiently alters mitochondrial dynamics during infection. Proc. Natl. Acad. Sci. U.S.A. 108, 3612–3617. doi: 10.1073/pnas.1100126108

PubMed Abstract | CrossRef Full Text | Google Scholar

Stowe, D. F., and Camara, A. K. (2009). Mitochondrial reactive oxygen species production in excitable cells: modulators of mitochondrial and cell function. Antioxid. Redox Signal. 11, 1373–1414. doi: 10.1089/ars.2008.2331

PubMed Abstract | CrossRef Full Text | Google Scholar

St-Pierre, J., Buckingham, J. A., Roebuck, S. J., and Brand, M. D. (2002). Topology of superoxide production from different sites in the mitochondrial electron transport chain. J. Biol. Chem. 277, 44784–44790. doi: 10.1074/jbc.m207217200

PubMed Abstract | CrossRef Full Text | Google Scholar

Strengert, M., Jennings, R., Davanture, S., Hayes, P., Gabriel, G., and Knaus, U. G. (2014). Mucosal reactive oxygen species are required for antiviral response: role of Duox in influenza a virus infection. Antioxid. Redox Signal. 20, 2695–2709. doi: 10.1089/ars.2013.5353

PubMed Abstract | CrossRef Full Text | Google Scholar

Stuehr, D. J., Fasehun, O. A., Kwon, N. S., Gross, S. S., Gonzalez, J. A., Levi, R., et al. (1991). Inhibition of macrophage and endothelial cell nitric oxide synthase by diphenyleneiodonium and its analogs. FASEB J. 5, 98–103. doi: 10.1096/fasebj.5.1.1703974

PubMed Abstract | CrossRef Full Text | Google Scholar

Su, L. J., Zhang, J. H., Gomez, H., Murugan, R., Hong, X., Xu, D., et al. (2019). Reactive Oxygen Species-induced lipid peroxidation in apoptosis, autophagy, and ferroptosis. Oxid. Med. Cell. Longev. 2019:5080843.

Google Scholar

Suzuki, T., Muramatsu, A., Saito, R., Iso, T., Shibata, T., Kuwata, K., et al. (2019). Molecular mechanism of cellular oxidative stress sensing by Keap1. Cell Rep. 28, 746–758.e4.

Google Scholar

Sweeney, R. M., and McAuley, D. F. (2016). Acute respiratory distress syndrome. Lancet 388, 2416–2430.

Google Scholar

Tai, T. C., Wong-Faull, D. C., Claycomb, R., and Wong, D. L. (2009). Hypoxic stress-induced changes in adrenergic function: role of HIF1 alpha. J. Neurochem. 109, 513–524. doi: 10.1111/j.1471-4159.2009.05978.x

PubMed Abstract | CrossRef Full Text | Google Scholar

Takahashi, M. A., and Asada, K. (1983). Superoxide anion permeability of phospholipid membranes and chloroplast thylakoids. Arch. Biochem. Biophys. 226, 558–566. doi: 10.1016/0003-9861(83)90325-9

CrossRef Full Text | Google Scholar

Taverne, Y. J., Bogers, A. J., Duncker, D. J., and Merkus, D. (2013). Reactive oxygen species and the cardiovascular system. Oxid. Med. Cell. Longev. 2013: 862423.

Google Scholar

Tazzeo, T., Worek, F., and Janssen, L. (2009). The NADPH oxidase inhibitor diphenyleneiodonium is also a potent inhibitor of cholinesterases and the internal Ca(2+) pump. Br. J. Pharmacol. 158, 790–796. doi: 10.1111/j.1476-5381.2009.00394.x

PubMed Abstract | CrossRef Full Text | Google Scholar

Teixeira, G., Szyndralewiez, C., Molango, S., Carnesecchi, S., Heitz, F., Wiesel, P., et al. (2017). Therapeutic potential of NADPH oxidase 1/4 inhibitors. Br. J. Pharmacol. 174, 1647–1669.

Google Scholar

ten Freyhaus, H., Huntgeburth, M., Wingler, K., Schnitker, J., Baumer, A. T., Vantler, M., et al. (2006). Novel Nox inhibitor VAS2870 attenuates PDGF-dependent smooth muscle cell chemotaxis, but not proliferation. Cardiovasc. Res. 71, 331–341. doi: 10.1016/j.cardiores.2006.01.022

PubMed Abstract | CrossRef Full Text | Google Scholar

Tepel, M., Echelmeyer, M., Orie, N. N., and Zidek, W. (2000). Increased intracellular reactive oxygen species in patients with end-stage renal failure: effect of hemodialysis. Kidney Int. 58, 867–872. doi: 10.1046/j.1523-1755.2000.00236.x

PubMed Abstract | CrossRef Full Text | Google Scholar

Thimmulappa, R. K., Lee, H., Rangasamy, T., Reddy, S. P., Yamamoto, M., Kensler, T. W., et al. (2006). Nrf2 is a critical regulator of the innate immune response and survival during experimental sepsis. J. Clin. Invest. 116, 984–995. doi: 10.1172/jci25790

PubMed Abstract | CrossRef Full Text | Google Scholar

Thorne, L. S., Rochford, G., Williams, T. D., Southam, A. D., Rodriguez-Blanco, G., Dunn, W. B., et al. (2021). Cytoglobin protects cancer cells from apoptosis by regulation of mitochondrial cardiolipin. Sci. Rep. 11:985.

Google Scholar

To, E. E., Vlahos, R., Luong, R., Halls, M. L., Reading, P. C., King, P. T., et al. (2017). Endosomal NOX2 oxidase exacerbates virus pathogenicity and is a target for antiviral therapy. Nat. Commun. 8:69.

Google Scholar

Tonks, N. K. (2005). Redox redux: revisiting PTPs and the control of cell signaling. Cell 121, 667–670. doi: 10.1016/j.cell.2005.05.016

PubMed Abstract | CrossRef Full Text | Google Scholar

Trevelin, S. C., Dos Santos, C. X., Ferreira, R. G., De Sa Lima, L., Silva, R. L., Scavone, C., et al. (2016). Apocynin and Nox2 regulate NF-kappaB by modifying thioredoxin-1 redox-state. Sci. Rep. 6:34581.

Google Scholar

Troia, A., Knutsen, R. H., Halabi, C. M., Malide, D., Yu, Z. X., Wardlaw-Pickett, A., et al. (2021). Inhibition of NOX1 Mitigates Blood Pressure Increases in Elastin Insufficiency. Function (Oxf) 2:zqab015.

Google Scholar

Ushijima, Y., Totsune, H., Nishida, A., and Nakano, M. (1997). Chemiluminescence from human polymorphonuclear leukocytes activated with opsonized zymosan. Free Radic. Biol. Med. 22, 401–409. doi: 10.1016/s0891-5849(96)00329-2

CrossRef Full Text | Google Scholar

Veal, E. A., Day, A. M., and Morgan, B. A. (2007). Hydrogen peroxide sensing and signaling. Mol. Cell 26, 1–14. doi: 10.1016/j.molcel.2007.03.016

PubMed Abstract | CrossRef Full Text | Google Scholar

Wang, H., Schoebel, S., Schmitz, F., Dong, H., and Hedfalk, K. (2020). Characterization of aquaporin-driven hydrogen peroxide transport. Biochim. Biophys. Acta Biomembr. 1862:183065. doi: 10.1016/j.bbamem.2019.183065

PubMed Abstract | CrossRef Full Text | Google Scholar

Wang, J., Zhang, W., Wang, S., Wang, Y., Chu, X., Ji, H., et al. (2021a). Lactobacillus plantarum exhibits antioxidant and cytoprotective activities in porcine intestinal epithelial cells exposed to hydrogen peroxide. Oxid. Med. Cell. Longev. 2021:8936907.

Google Scholar

Wang, X., Jiang, Z., Yue, N., Jin, X., Zhang, X., Li, Z., et al. (2021b). Barley stripe mosaic virus gammab protein disrupts chloroplast antioxidant defenses to optimize viral replication. EMBO J 40:e107660.

Google Scholar

Wang, P., Geng, J., Gao, J., Zhao, H., Li, J., Shi, Y., et al. (2019). Macrophage achieves self-protection against oxidative stress-induced ageing through the Mst-Nrf2 axis. Nat. Commun. 10:755.

Google Scholar

Wang, Q., and Zou, M. H. (2018). Measurement of reactive oxygen Species (ROS) and Mitochondrial ROS in AMPK Knockout Mice Blood Vessels. Methods Mol. Biol. 1732, 507–517. doi: 10.1007/978-1-4939-7598-3_32

CrossRef Full Text | Google Scholar

Wang, Y., Branicky, R., Noe, A., and Hekimi, S. (2018). Superoxide dismutases: dual roles in controlling ROS damage and regulating ROS signaling. J. Cell Biol. 217, 1915–1928. doi: 10.1083/jcb.201708007

PubMed Abstract | CrossRef Full Text | Google Scholar

Wang, Z., Cai, F., Chen, X., Luo, M., Hu, L., and Lu, Y. (2013). The role of mitochondria-derived reactive oxygen species in hyperthermia-induced platelet apoptosis. PLoS One 8:e75044. doi: 10.1371/journal.pone.0075044

PubMed Abstract | CrossRef Full Text | Google Scholar

Wei, R., Enaka, M., and Muragaki, Y. (2019). Activation of KEAP1/NRF2/P62 signaling alleviates high phosphate-induced calcification of vascular smooth muscle cells by suppressing reactive oxygen species production. Sci. Rep. 9:10366.

Google Scholar

West, A. P., Brodsky, I. E., Rahner, C., Woo, D. K., Erdjument-Bromage, H., Tempst, P., et al. (2011a). TLR signalling augments macrophage bactericidal activity through mitochondrial ROS. Nature 472, 476–480. doi: 10.1038/nature09973

PubMed Abstract | CrossRef Full Text | Google Scholar

West, A. P., Shadel, G. S., and Ghosh, S. (2011b). Mitochondria in innate immune responses. Nat. Rev. Immunol. 11, 389–402.

Google Scholar

Wind, S., Beuerlein, K., Eucker, T., Muller, H., Scheurer, P., Armitage, M. E., et al. (2010). Comparative pharmacology of chemically distinct NADPH oxidase inhibitors. Br. J. Pharmacol. 161, 885–898. doi: 10.1111/j.1476-5381.2010.00920.x

PubMed Abstract | CrossRef Full Text | Google Scholar

Wingler, K., Hermans, J. J., Schiffers, P., Moens, A., Paul, M., and Schmidt, H. H. (2011). NOX1, 2, 4, 5: counting out oxidative stress. Br. J. Pharmacol. 164, 866–883. doi: 10.1111/j.1476-5381.2011.01249.x

PubMed Abstract | CrossRef Full Text | Google Scholar

Winitchaikul, T., Sawong, S., Surangkul, D., Srikummool, M., Somran, J., Pekthong, D., et al. (2021). Calotropis gigantea stem bark extract induced apoptosis related to ROS and ATP production in colon cancer cells. PLoS One 16:e0254392. doi: 10.1371/journal.pone.0254392

PubMed Abstract | CrossRef Full Text | Google Scholar

Wink, D. A., Hines, H. B., Cheng, R. Y., Switzer, C. H., Flores-Santana, W., Vitek, M. P., et al. (2011). Nitric oxide and redox mechanisms in the immune response. J. Leukoc. Biol. 89, 873–891.

Google Scholar

Winterbourn, C. C., and Hampton, M. B. (2008). Thiol chemistry and specificity in redox signaling. Free Radic. Biol. Med. 45, 549–561. doi: 10.1016/j.freeradbiomed.2008.05.004

PubMed Abstract | CrossRef Full Text | Google Scholar

Winterbourn, C. C., and Kettle, A. J. (2013). Redox reactions and microbial killing in the neutrophil phagosome. Antioxid. Redox Signal. 18, 642–660. doi: 10.1089/ars.2012.4827

PubMed Abstract | CrossRef Full Text | Google Scholar

Wolf, A., Herb, M., Schramm, M., and Langmann, T. (2020). The TSPO-NOX1 axis controls phagocyte-triggered pathological angiogenesis in the eye. Nat. Commun. 11:2709.

Google Scholar

Wong, H. S., Mezera, V., Dighe, P., Melov, S., Gerencser, A. A., Sweis, R. F., et al. (2021). Superoxide produced by mitochondrial site IQ inactivates cardiac succinate dehydrogenase and induces hepatic steatosis in Sod2 knockout mice. Free Radic. Biol. Med. 164, 223–232. doi: 10.1016/j.freeradbiomed.2020.12.447

PubMed Abstract | CrossRef Full Text | Google Scholar

Woo, H. A., Yim, S. H., Shin, D. H., Kang, D., Yu, D. Y., and Rhee, S. G. (2010). Inactivation of peroxiredoxin I by phosphorylation allows localized H(2)O(2) accumulation for cell signaling. Cell 140, 517–528. doi: 10.1016/j.cell.2010.01.009

PubMed Abstract | CrossRef Full Text | Google Scholar

Wood, Z. A., Poole, L. B., and Karplus, P. A. (2003). Peroxiredoxin evolution and the regulation of hydrogen peroxide signaling. Science 300, 650–653. doi: 10.1126/science.1080405

PubMed Abstract | CrossRef Full Text | Google Scholar

Xu, J., Tran, T., Padilla Marcia, C. S., Braun, D. M., and Goggin, F. L. (2017). Superoxide-responsive gene expression in Arabidopsis thaliana and Zea mays. Plant Physiol. Biochem. 117, 51–60. doi: 10.1016/j.plaphy.2017.05.018

PubMed Abstract | CrossRef Full Text | Google Scholar

Yang, Y., Karakhanova, S., Hartwig, W., D’haese, J. G., Philippov, P. P., Werner, J., et al. (2016). Mitochondria and mitochondrial ROS in cancer: novel targets for anticancer therapy. J. Cell. Physiol. 231, 2570–2581. doi: 10.1002/jcp.25349

PubMed Abstract | CrossRef Full Text | Google Scholar

Zaidieh, T., Smith, J. R., Ball, K. E., and An, Q. (2019). ROS as a novel indicator to predict anticancer drug efficacy. BMC Cancer 19:1224. doi: 10.1186/s12885-019-6438-y

PubMed Abstract | CrossRef Full Text | Google Scholar

Zeller, J., Bogner, B., Kiefer, J., Braig, D., Winninger, O., Fricke, M., et al. (2021). CRP Enhances the Innate Killing Mechanisms Phagocytosis and ROS formation in a conformation and complement-dependent manner. Front. Immunol. 12:754. doi: 10.3389/fimmu.2018.00754

PubMed Abstract | CrossRef Full Text | Google Scholar

Zhang, J., Wang, X., Vikash, V., Ye, Q., Wu, D., Liu, Y., et al. (2016). ROS and ROS-mediated cellular signaling. Oxid. Med. Cell. Longev. 2016:4350965.

Google Scholar

Zhang, W., Hu, X., Shen, Q., and Xing, D. (2019). Mitochondria-specific drug release and reactive oxygen species burst induced by polyprodrug nanoreactors can enhance chemotherapy. Nat. Commun. 10:1704.

Google Scholar

Zhang, X. D., Peng, Y. Q., Zhao, J., Li, Q., Yu, X. J., Acevedo-Rocha, C. G., et al. (2020). Bacterial cytochrome P450-catalyzed regio- and stereoselective steroid hydroxylation enabled by directed evolution and rational design. Bioresour. Bioprocess. 7:2.

Google Scholar

Zhang, Y., Yao, Y., Qiu, X., Wang, G., Hu, Z., Chen, S., et al. (2019). Listeria hijacks host mitophagy through a novel mitophagy receptor to evade killing. Nat. Immunol. 20, 433–446. doi: 10.1038/s41590-019-0324-2

PubMed Abstract | CrossRef Full Text | Google Scholar

Zhao, Q., Liu, J., Deng, H., Ma, R., Liao, J. Y., Liang, H., et al. (2020). Targeting mitochondria-located circRNA SCAR Alleviates NASH via Reducing mROS Output. Cell 183, 76–93.e22.

Google Scholar

Zhao, R. Z., Jiang, S., Zhang, L., and Yu, Z. B. (2019). Mitochondrial electron transport chain, ROS generation and uncoupling (Review). Int. J. Mol. Med. 44, 3–15.

Google Scholar

Zhong, J., Li, Q., Luo, H., and Holmdahl, R. (2021). Neutrophil-derived reactive oxygen species promote tumor colonization. Commun. Biol. 4:865.

Google Scholar

Zhou, L., Yeo, A. T., Ballarano, C., Weber, U., Allen, K. N., Gilmore, T. D., et al. (2014). Disulfide-mediated stabilization of the IkappaB kinase binding domain of NF-kappaB essential modulator (NEMO). Biochemistry 53, 7929–7944. doi: 10.1021/bi500920n

PubMed Abstract | CrossRef Full Text | Google Scholar

Zhou, R., Yazdi, A. S., Menu, P., and Tschopp, J. (2011). A role for mitochondria in NLRP3 inflammasome activation. Nature 469, 221–225.

Google Scholar

Keywords: mitochondria, NADPH oxidases, ROS sources, reactive oxygen species, redox balance, oxidative stress, ROS probes, ROS inhibitors

Citation: Herb M, Gluschko A and Schramm M (2021) Reactive Oxygen Species: Not Omnipresent but Important in Many Locations. Front. Cell Dev. Biol. 9:716406. doi: 10.3389/fcell.2021.716406

Received: 28 May 2021; Accepted: 19 August 2021;
Published: 07 September 2021.

Edited by:

Bechan Sharma, Allahabad University, India

Reviewed by:

Takuya Sakaguchi, Cleveland Clinic, United States
Andrea N. Beltrán González, Molecular Biology and Neurosciences (IFIBYNE), Argentina
Roberto Giovannoni, University of Pisa, Italy

Copyright © 2021 Herb, Gluschko and Schramm. This is an open-access article distributed under the terms of the Creative Commons Attribution License (CC BY). The use, distribution or reproduction in other forums is permitted, provided the original author(s) and the copyright owner(s) are credited and that the original publication in this journal is cited, in accordance with accepted academic practice. No use, distribution or reproduction is permitted which does not comply with these terms.

*Correspondence: Marc Herb, marc.herb@uk-koeln.de

Download