Skip to main content
Molecules and Cells logoLink to Molecules and Cells
. 2011 Dec 31;32(6):491–509. doi: 10.1007/s10059-011-0276-3

Regulation of Reactive Oxygen Species Generation in Cell Signaling

Yun Soo Bae 1,*, Hyunjin Oh 1, Sue Goo Rhee 1,*, Young Do Yoo 2,*
PMCID: PMC3887685  PMID: 22207195

Abstract

Reactive oxygen species (ROS) including superoxide anion and hydrogen peroxide (H2O2) are thought to be byproducts of aerobic respiration with damaging effects on DNA, protein, and lipid. A growing body of evidence indicates, however, that ROS are involved in the maintenance of redox homeostasis and various cellular signaling pathways. ROS are generated from diverse sources including mitochondrial respiratory chain, enzymatic activation of cytochrome p450, and NADPH oxidases further suggesting involvement in a complex array of cellular processes. This review summarizes the production and function of ROS. In particular, how cytosolic and membrane proteins regulate ROS generation for intracellular redox signaling will be detailed.

Keywords: apoptosis, cell signaling, cytochrome p450, mitochondria, NADPH oxidase, reactive oxygen species

INTRODUCTION

Generation of cellular Reactive Oxygen Species (ROS) is induced by both endogenous and exogenous stimuli. Recent evidences convincingly argue that these ROS play an important role in numerous physiological and pathological processes (Droge, 2002; Martindale and Holbrook, 2002). Although high levels of ROS contribute to carcinogenesis and other diseases related to oxidative damage, appropriate levels of ROS have been shown to be indispensible for cell survival, apotosis, and differentiation (Fig. 1). Consistently, various signaling proteins including NF-κB, PI3-K, MAPK, and p53 perform their respective cellular roles in responses to ROS generation (Burdon and Rice-Evans, 1989; Finkel, 1998; Martindale and Holbrook, 2002). Moreover, the relationship between hydrogen peroxide (H2O2) and mammalian cell proliferation has been studied in different cell types (Caporossi et al., 2003; Chung et al., 2009; Liu et al., 2002a; Na et al., 2008; Sigaud et al., 2005; Stone and Collins, 2002; Stone and Yang, 2006). For example, reduction of endogenous ROS levels by addition or overexpression of antioxidant proteins inhibit the proliferation of both vascular smooth muscle cells and tumor cells (Caporossi et al., 2003; Liu et al., 2002a; Sigaud et al., 2005; Stone and Collins, 2002). Importantly, inhibition of endogenous ROS generation causes cell-cycle arrest in the G1 phase, demonstrating that ROS steady-state levels are required for entry into the S phase (Chung et al., 2009; Lee et al., 2011; Sekharam et al., 1998). ROS production required for redox signaling is mainly induced by NADPH oxidase, and various growth factors and cytokines stimulate ROS generation by activating this enzyme (Droge, 2002). Although whether ROS originating from the mitochondria contribute to redox signaling is not much known, it is increasingly evident that H2O2 released to the cytosol participates in various signaling networks, such as cell-cycle transition and redox balance (Kow-altowski et al., 2009). The objective of this review is to discuss the regulation of ROS production from mitochondria, cytochrome p450, and NADPH oxidase and the implications in cellular signaling.

Fig. 1. Mitochondrial ROS accumulation in response to various stimuli. Mitochondria sense various external signals and stresses to induce mitochondrial ROS release to the cytosol. Mitochondrial ROS are indispensible for normal cellular function. However, dysregulation of mitochondrial ROS production or release to the cytosol is implicated in many diseases, especially those involving inflammation.

Fig. 1.

REGULATION OF MITOCHONDRIAL REACTIVE OXYGEN SPECIES (ROS) GENERATION

Most ROS produced intracellularly originate in the mitochondrial respiratory chain and subsequently produce toxic metabolic byproducts. However, mitochondrial ROS have been increasingly implicated in cellular signaling pathways, including those for survival and cell death. In these processes, many proteins modulate the release of mitochondrial ROS to the cytosol and they are tightly controlled. This ROS release is implicated in the maintenance of redox homeostasis and various cellular signaling pathways. This section summarizes the sources of ROS in the mitochondria and the recent findings regarding how the nuclear encoded proteins regulate ROS release to the cytosol for intracellular redox signaling.

ROS generation in the mitochondrial respiratory chain

One of the main functions of the mitochondria is oxidative ATP production, in which oxygen (O2) is reduced to water, and consequently the major source of intracellular ROS generation is the mitochondrial respiratory chain (Brand, 2010; Loschen et al., 1973). The respiratory chain consists of five multi-subunits protein complexes located in the mitochondrial IMS (respiratory complexes I-IV and the F1Fo-ATP synthase) and two factors (cytochrome c, Cyt c; coenzyme Q10) (Galluzzi et al., 2010). During respiration, electrons released from the mitochondrial electron transport chain incompletely reduce O2 to form supeoxide (Chance et al., 1979). Superoxide is converted into H2O2 by manganese superoxide dismutase (Mn-SOD) in the mitochondrial matrix or by Cu, Zn-SOD in the IMS of mitochondria (Okado-Matsumoto and Fridovich, 2001; Sturtz et al., 2001; Weisiger and Fridovich, 1973a; 1973b). The 1-2% of O2 consumed during respiration is estimated to be incompletely reduced to O2 to produce superoxide in isolated mitochondria treated with respiratory chain inhibitors (Chance et al., 1979; Kudin et al., 2004). However, 0.12-0.15% of O2 produces H2O2 when palmitoyl-CoA or glutamate/malate are used as substrates (Kudin et al., 2004; Murphy, 2009; St-Pierre et al., 2002). Recently, Brand suggested that mitochondrial superoxide is produced at seven major sites of mitochondria and all sites release it into the matrix (Brand, 2010). Out of the seven sites, complex III (site IIIQo) and glycerol 3-phosphate dehydrogenase also liberate superoxide into the IMS (Chen et al., 2003b; Miwa and Brand, 2005; Muller et al., 2004). Since the two major sites for superoxide production are known to be complexes I and III of the mitochondrial respiratory chain (Chen et al., 2003b; Poyton et al., 2009), these two sites are discussed in this review.

Complex III

Complex III (ubiquinol-cytochrome c oxidoreductase) consists of 11 polypeptides, three hemes and a Fe-S center (Iwata et al., 1998). This complex receives electrons from coenzyme Q (ubiquinol, QH2) and transfers them to Cyt c. Mitochondrial superoxide is generated through the Q-cycle, as documented in recent papers (Andreyev et al., 2005; Hamanaka and Chandel, 2010; Turrens et al., 1985). Coenzyme Q (ubiquinone, Q) accepts two electrons transferred from mitochondrial complexes I or II to form reduced QH2. Complex III binds to QH2 in the intermembrane space-proximal Qo site and transfers these electrons into Cyt c. Because Cyt c can only accept one electron, the first electron is transferred into Cyt c1 through the Rieske iron-sulfur protein (RISP), resulting in the transient formation of ubisemiquinone (QH˙). This electron reduces Cyt c and is transferred to complex IV. The other electron of QH˙ is immediately transferred to the cytochrome bH center through the cytochrome bL center of complex III. The oxidized coenzyme Q or QH˙ can be reduced by cytochrome bH center at the matrix-proximal Qi site and the resulting QH2 then transfers the electron into Cyt c1 in the intermembrane space-proximal Qo site. The superoxide is mainly generated by reaction electron of QH˙ with O2 at the Qo site of complex III, liberating it into the IMS (St-Pierre et al., 2002). Antimycin increases superoxide production by blocking electron transfer from the cytochrome b center to Q at the Qi site (Chen et al., 2003b). In contrast, myxothiazol and stigmatellin inhibit superoxide production at the Qo site of complex III (Ksenzenko et al., 1983; Turrens et al., 1985).

Complex I

Complex I (NADH-ubiquinone oxidoreductase) transfers electrons received from NADH to complex III through Coenzyme Q. The mitochondrial superoxide produced by complex I is released into the matrix (St-Pierre et al., 2002). The superoxide at complex I is produced by two mechanisms: NADH-linked forward electron transport and succinate-linked reverse electron transport (Hansford et al., 1997; Kudin et al., 2004; Kushnareva et al., 2002; Kwong and Sohal, 1998; Lambert and Brand, 2004; Liu et al., 2002b; Votyakova and Reynolds, 2001). Higher mitochondrial superoxide generation is observed when the matrix NADH/NAD+ ratio is high or during reverse electron transport from succinate to NAD+ in isolated mitochondria. However, the superoxide production at complex I under the physiological conditions is controversial. Furthermore, many potential sites of mitochondrial ROS production have been suggested but whether they contribute to mitochondrial ROS production is unclear (Brand, 2010).

Intracellular ROS increase in mitochondria in response to external stimuli

Tumor necrosis factor-α (TNF-α)

TNF-α plays an important role in proliferation, differentiation, and cell death for tissue remodeling during inflammation (Locksley et al., 2001; Wajant et al., 2003). TNF-α signaling is tightly controlled; however, deregulated TNF-α production causes chronic inflammation, which is directly associated with a variety of diseases such as cancer, inflammatory bowel diseases, arthritis, sclerosis, and Alzheimer’s disease (Balkwill and Coussens, 2004; Chen and Goeddel, 2002). There are two main pathways for cell survival and death in TNF-α signaling. One is the pro-survival/pro-inflammatory pathway through NF- κB and MAPK activation, which is mediated by TNF-α-induced signaling complex I (Micheau and Tschopp, 2003). After activation of the cell surface receptor (TNF-R1) by binding of TNF-α, the silencer of death domain (SODD) dissociates from TNF-R1 complexes, and the TNF receptor-associated protein with death domain (TRADD) interacts with the cytoplasmic domain of the receptor (Hsu et al., 1995). TRADD recruits TNF receptor-associated factor 2 (TRAF2), receptor-interacting protein 1 (RIP1), and cellular inhibitor of apoptosis proteins to form TNF-α- induced signaling complex I (Chen and Goeddel, 2002; Karin and Lin, 2002). TRAF2 then recruits the IKK complex to the TNF-R1 signaling complex to activate the kinases of the IKK complex via RIP1 (Devin et al., 2000). The activated IKKs phosphorylate IκB to induce the proteolytic degradation of IκB, thereby liberating NF-κB and allowing nuclear translocation (Verma et al., 1995). NF-κB transcription factor regulates expression of the pro-survival proteins, cFLIP, Gadd45b, A20, XIAP, and cIAPs (Amanullah et al., 2003; Papa et al., 2004). NF-κB also induces expression of FHC and Mn-SOD to prevent TNF-R-induced cell death by eliminating ROS (Pham et al., 2004; Sakon et al., 2003). Subsequently, TNF-α induced signaling complex II is formed to activate the other TNF-α signaling pathway that is implicated in cell death, in which ROS, a caspase cascade, and the mitochondria function as downstream mediators. After TRADD, TRAF2, and RIP1 dissociate from the receptor, they interact with Fas-associated death domain protein (FADD) and pro-caspase-8 to form TNF-α-induced signaling complex II (Hsu et al., 1996; Rothe et al., 1994).

ROS have been implicated as mediators of cell survival and cell death triggered by TNF-α signaling. ROS play an important role in TNF-induced cell survival by NF-κB activation (Sulciner et al., 1996). TNF-α-induced ROS production has been also reported to play an important role in TNF-α-induced prolonged JNK activation and cell death (De Smaele et al., 2001; Ichijo et al., 1997; Matsuzawa and Ichijo, 2005; Tang et al., 2001). The generation of ROS in response to signals for TNF-α-induced cell death occurs in the mitochondria (Corda et al., 2001; Lo and Cruz, 1995; Locksley et al., 2001; Matsuzawa and Ichijo, 2005; Meier et al., 1989; Shoji et al., 1995). However, ROS generation originating from NADPH oxidase after TNF-α treatment is also implicated in TNF-α-induced cell death (Kim et al., 2007b). Although many reports show that the main source of TNF-α-triggered ROS is the mitochondria, the mitochondrial origin of ROS formation triggered by TNF-α is not clear. Recently, we demonstrated that Romo1 is a key mediator between TNF-α signaling and the mitochondria for ROS production, as it recruits B-cell lymphoma-extra large (Bcl-XL) to reduce the mitochondrial membrane potential, resulting in ROS production and apoptosis (Kim et al., 2010). More discussion about Romo1 can be found later in this paper.

Hypoxia

Cells often have to respond and adapt to the unfavorable condition of low oxygen levels (hypoxia). The stability of the hypoxia inducible factors (HIFs) is increased in response to hypoxia and can play a central role in the generation of new vasculature to increase the oxygen supply. Although the exact hypoxiasensing mechanism is controversial, the mitochondrial electron transport chain is suggested to play an important role in oxygen sensing (Galluzzi et al., 2010). In early studies, it was suggested that mitochondrial production of ROS is involved in hypoxia signaling (Chandel et al., 1998; 2000). Mitochondrial production of ROS is increased in response to hypoxia and these ROS act as upstream mediators of hypoxia-induced signaling process. Treatment with some inhibitors of the electron transport chain inhibits stabilization of HIF-1α in response to hypoxia. In later experiments, it was suggested that mitochondrial complex III is responsible for hypoxia-triggered ROS production (Brunelle et al., 2005; Guzy et al., 2005). These experiments were performed by suppressing the expression of one factor of complex III, by using mitochondrial inhibitors and by using antioxidants. Down-regulation of the RISP blocks hypoxiatriggered HIF-1α stabilization. Hypoxic HIF-1α activation was also suppressed in cells lacking Cyt c, and exogenous treatment with H2O2 recovered HIF-1α stability (Mansfield et al., 2005). These studies demonstrate that ROS required for hypoxia signaling are released from mitochondria to cytosol and their origin is mitochondrial complex III. Although there have been many studies about hypoxia-induced ROS generation, how mitochondrial complex III senses hypoxia remains to be investigated, as well as the exact mechanism by which hypoxia induces ROS production at complex III.

Serum deprivation

Serum provides cell growth factors, hormones, and various other factors for most in vitro cell cultures (van der Valk et al., 2004). Serum withdrawal triggers cell death and is correlated with ROS production in many cell lines (Charles et al., 2005; Greene, 1978; King et al., 2003; Pandey et al., 2003; Satoh et al., 1996; Zhuge and Cederbaum, 2006). Because serum deprivation is one component of ischemia, there have been several reports about ischemia (Bialik et al., 1999). In serum-deprived neuronal cells, extensive cell death is observed and addition of nerve growth factor (NGF) inhibits the cell death (Ferrari et al., 1995; Greene, 1978). Antioxidant also suppresses cell death, though it cannot recover cell proliferation compared to cells grown in complete media (Ferrari et al., 1995). Although there have been many reports demonstrating that serum deprivation triggers increases in ROS, the mechanism by which serum deprivation induces ROS generation and the source inside cells are not fully elucidated. In our recent paper, we showed that serum deprivation-induced ROS generation, which was measured by MitoSOX, a mitochondrial superoxide indicator, was derived from complex III of the mitochondrial respiratory chain (Lee et al., 2010). Serum deprivation-induced ROS production was inhibited by mitochondrial complex III inhibitors (myxothizol and stigmatellin). However, other mitochondrial complex inhibitors (rotenone, malonate and sodium azide) failed to block the increase in ROS production in serum-starved cells. We also showed that the serum deprivation-triggered increase in ROS was mediated by Romo1 (Lee et al., 2010). However, the exact mechanism by which Romo1 increases ROS levels in the mitochondria of serum-starved cells requires more extensive investigation.

Other stimuli

Various stresses also stimulate mitochondrial ROS production. Oxidative stress-induced ROS production among external stimuli is significantly implicated in neighboring mitochondrial damage and cell death and this phenomenon is called “ROS-induced ROS release (RIRR)” (Zorov et al., 2000; 2006). The oxidative stress causes the collapse of the mitochondrial membrane potential (ΔΨm), resulting in an increase in ROS. The amplified ROS can injure neighboring mitochondria in a positive feedback loop (Kim et al., 2006; Zorov et al., 2000; 2006). Other stresses, including phorbol ester (TPA) and irradiation, also stimulate mitochondria ROS production. TPA-induced ROS production is known to be mediated by NADPH oxidase and/or mitochondria and ROS induced by TPA are associated with tumor cell invasion (Frost et al., 1994; Wu, 2006). 5-FU, arsenic trioxide, cisplatin, paclitaxel, bleomycin, and adriamycin have also been reported to trigger ROS generation (Adler et al., 1999; Huang et al., 2000; Hwang et al., 2001; Marcillat et al., 1989; Miyajima et al., 1997; Pelicano et al., 2003; 2004; Tan et al., 1998).

Regulation of mitochondrial ROS release to the cytosol

Mitochondrially generated ROS are released to the cytosol and participate in a variety of intracellular functions including cell proliferation, cell death, cell cycle, and redox homeostasis (Droge, 2002). H2O2 produced by mitochondrial superoxide dismutase can migrate into the cytosol across the membrane by simple diffusion. However, cells more efficiently use mitochondrial ROS by function of the proteins encoded by nuclear genes. In this section, the nuclear encoded proteins that regulate the ROS release to the cytosol will be discussed.

p53

The tumor suppressor protein p53 is a key regulator of cell cycle, senescence, or apoptosis. The main action of the p53 protein is as a transcription factor that induces the expression of target genes and causes cell-cycle arrest, DNA repair, or apoptosis in response to cellular genotoxic stresses (Giaccia and Kastan, 1998; Levine, 1997). Although p53 participates in many cellular activities, only the control of intracellular ROS by p53 will be discussed in this review. p53 differentially regulates the intracellular ROS according to the levels and types of stresses. p53 exerts transcription-dependent pro-apoptotic effects by induction of the pro-oxidant genes. Polyak et al. showed that enhanced p53 activity increased the oxidative stress by inducing its target genes, including two ROS-generating enzymes, NQO1 (quinone oxidoreductase, PIG3) and proline oxidase (POX, PIG6), and resulting in apoptosis (Polyak et al., 1997; Rivera and Maxwell, 2005). BAX, PUMA and p66Shc are also downstream targets of the p53 protein and they induce mitochondrial ROS elevation, which stimulates the opening of the mitochondrial permeability transition pore (PTP) to trigger apoptosis (Giorgio et al., 2005; Liu et al., 2005; Sablina et al., 2005; Trinei et al., 2002). A microarray analysis of H2O2-treated human cells showed that one-third of the 48 highly H2O2-reponsive genes are related to p53-mediated transcriptional effects (Desaint et al., 2004). Interestingly, p53 plays a role in the maintenance of redox homeostasis by controlling the expression of antioxidant genes, such as SESN1 (mammalian sestrin homologue), SESN2, glutathione peroxidase-1 (GPX1), and TP53INP1 (Cano et al., 2009; Sablina et al., 2005; Stambolsky et al., 2006; Tomko et al., 2006). This antioxidant function of p53 is suggested to be one of the tumor-suppressing mechanisms of p53 (Sablina et al., 2005).

The cellular regulation of p53 mainly occurs by p53-mediated transcriptional activity in the nucleus. However, many studies have reported that a fraction of cellular p53 migrates into the outer mitochondrial membrane in response to a variety of stress signals, including DNA damage and hypoxic stress, and promotes transcription-independent p53-mediated cell death (Chipuk et al., 2004; Dumont et al., 2003; Mihara et al., 2003; Pim and Banks, 2004; Sansome et al., 2001; Yoo et al., 2005). This mitochondrial migration seems to be promoted by monoubiquitylation of p53 (Marchenko et al., 2007). In the mitochondria, p53 interacts with the Bcl-2 family members present at the mitochondrial outer membrane, including Bcl-2 and Bcl-XL, thereby inducing mitochondrial outer membrane permeabilization (MOMP). MOMP is an upstream event of programmed cell death and is controlled by the Bcl-2 family (Green and Kroemer, 2004; Reed, 2006). p53 appears to bind Bcl-2 or Bcl-XL to release Bax from their inhibitory complexes to trigger MOMP (Chipuk et al., 2004; Mihara et al., 2003). p53 also interacts with Bak, Bax, and PUMA to induce apoptosis (Chipuk et al., 2004; 2005; Leu et al., 2004). Thus, p53 up-regulates or down-regulates the ROS levels in transcription-dependent or -independent mechanisms.

p66Shc

The Shc adaptor protein family encodes the three isoforms of Shc protein (p46Shc, p52Shc, and p66Shc) that transmit mitogenic signals to RAS (Pelicci et al., 1992). p66Shc is serine phosphorylated in response to external stimuli such as ultraviolet light or ROS (Migliaccio et al., 1999). p66Shc is a downstream mediator of p53 for the elevation of intracellular ROS, Cyt c release and apoptosis (Trinei et al., 2002). Cells lacking p66Shc are resistant to apoptosis triggered by stresses, and p66Shc-/- mice show a 30% increase in life span (Migliaccio et al., 1999). In response to external stresses, p66Shc is phosphorylated on serine 36 residue by the stress-activated kinase PKC-beta and translocated into the mitochondrial IMS (Gertz et al., 2008; Nemoto et al., 2006; Pinton et al., 2007). p66Shc then oxidizes reduced cytochrome c to induce H2O2 production, resulting in PTP opening that triggers apoptosis (Giorgio et al., 2005).

The Bcl-2 family

The Bcl-2 family contains proteins with the Bcl-2 homology (BH) region and these proteins are classified into anti-apoptotic and pro-apoptotic proteins. The anti-apoptotic proteins, Bcl-2 and Bcl-XL, include four BH domains (BH1234). In contrast, the proapoptotic proteins Bax and Bak have three BH domains (BH123), and other pro-apoptotic proteins Bad and Bid contain only the BH3 domain (Reed, 2006; Rong and Distelhorst, 2008). The anti-apoptotic and pro-apoptotic functions of the Bcl-2 family are well investigated. However, redox control of Bcl-2 at the outer mitochondrial membrane is not well identified. Although the mechanism of Bcl-2 as an antioxidant is not well elucidated, Bcl-2 has been reported to have an antioxidant effect and this effect is correlated with the anti-apoptotic action of Bcl-2 (Ellerby et al., 1996; Hockenbery et al., 1993; Kane et al., 1993). The antioxidant function of Bcl-2 also regulates cell-cycle arrest at G1 (Deng et al., 2003). Another well-known anti-apoptotic protein Bcl-XL is reported to have antioxidant action in several papers. It plays a role in mitochondrial homeostasis by stabilizing ΔΨm (Vander Heiden et al., 1997). Bcl-XL expression prevents the early decrease in ΔΨm and the subsequent ROS production stimulated by TNF-α (Gottlieb et al., 2000). Bcl-XL also inhibits TNF-α-induced ROS production and cell death mediated by Romo1 (Kim et al., 2010). Bax localized in the cytoplasm translocates into the outer mitochondrial membrane to induce MOMP, causing Cyt c release in stressful conditions (Wolter et al., 1997). During apoptotic cell death, ROS levels are increased and these are blocked in cells lacking Bax (Kirkland and Franklin, 2001; Kirkland et al., 2002). The mechanism of Bax-triggered increase in ROS levels remains to be studied.

Romo1

ROS produced by the mitochondrial electron transport chain are released to cytosol and are indispensible for cell survival and sometimes the cause of many pathological disorders. Although H2O2 is membrane permeable and can be transferred into the cytosol, cells regulate ROS release to the cytosol with some proteins encoded by the nucleus. One excellent example of these is the Romo1 protein. Romo1 acts as a key regulator of mitochondrial ROS release to the cytosol, and its exact mechanisms have begun to be uncovered in the past few years. Romo1 was first identified as the gene expressed differentially in the tumor tissue of a patient who was initially sensitive to chemotherapy but became resistant after a recurrence (Chung et al., 2006). In a subsequent experiment, enforced expression of Romo1 through transfection of the Romo1 gene into several cell lines enhanced ROS levels, as measured by DCF-DA, MitoSOX, or Amplex Red (Chung et al., 2006; 2008). Therefore, it was suggested that Romo1 induces mitochondrial ROS production and that Romo1-triggered ROS production originates in mitochondrial complex III (Chung et al., 2008). It is unlikely that Romo1 generates ROS in response to various stimuli, similar to NADPH oxidase and other enzymes such as xanthine oxidase, cyclooxygenases, and lipoxygenases. Romo1 is located in the outer mitochondrial membrane and its molecular weight is approximately 8.9 kDa, with a single transmembrane domain and no homology domain when compared with these other enzymes (Chung et al., 2006; Kim et al., 2010). Therefore, it was suggested that Romo1 is located in the mitochondrial membrane and that Romo1 acts as a modulator to supply ROS from complex III of the mitochondrial respiratory chain into the cytosol (Chung et al., 2006). If so, how does Romo1 modulate mitochondria ROS release to the cytosol? A recent report showed that TNF-α-induced ROS are associated with Romo1 located in the outer mitochondrial membrane, demonstrating the possible mechanism of ROS release to the cytosol by Romo1 (Kim et al., 2010). In response to TNF-α, TNF complex II interacts with the C-terminus of Romo1. Simultaneously, Romo1 binds to Bcl-XL to reduce ΔΨm, resulting in ROS production and apoptotic cell death. Interestingly, the interaction between Romo1 and Bcl-XL after TNF-α treatment liberates Bax from Bcl-XL (Fig. 2). Bax was also reported to induce the mitochondrial production of ROS (Kirkland et al., 2002).

Fig. 2. TNF-α-induced ROS production through Romo1 and Bcl-XL. In response to TNF-α, TNF-α signaling complex II is translocated into the mitochondria and interacts with the C-terminal region of Romo1. Simultaneously, Romo1 recruits Bcl-XL and mediates the early TNF-α- induced decrease of ΔΨm, probably by binding to Bcl-XL and preventing its function. Subsequently, TNF-α-induced decrease of ΔΨm triggers ROS production, which leads to the late decrease of ΔΨm and ROS production. Eventually, the amplified ROS generation induces prolonged JNK activation and apoptotic cell death.

Fig. 2.

What is the physiological role of Romo1? Although the role of Romo1 has not been fully identified, several reports demonstrate possible roles of Romo1 in the cells. The first possibility is that Romo1 maintains the redox homeostasis for cell survival (Na et al., 2008). Romo1 knockdown decreased the basal levels of ROS and inhibited the cell growth of various normal and tumor cells. The steady-state ROS levels generated from the endogenous Romo1 protein are also indispensable for both normal and cancer cell proliferation, because Romo1 knockdown inhibits cell-cycle transition through inhibition of Erk activation and p27Kip1 expression (Chung et al., 2009; Na et al., 2008). Therefore, Romo1 may play an important role in redox signaling during normal cell proliferation. The second possibility is that Romo1 is a molecular bridge between TNF-α signaling and the mitochondria for ROS production that triggers TNF-α- mediated apoptosis (Kim et al., 2010). However, further studies are needed to elucidate the precise role of Romo1 in the TNF-α signaling pathway. The third possibility is that Romo1 expression is up-regulated by various stresses, including serum deprivation, 5-FU, and replicative senescence (Chung et al., 2008; Hwang et al., 2007; Lee et al., 2010). It has been reported that serum withdrawal stimulates mitochondrial ROS production (Ferrari et al., 1995; Satoh et al., 1996). This ROS generation in serum-starved cells is mediated by Romo1 expression (Lee et al., 2010). The last possibility is that Romo1 plays an important role in c-Myc turnover (Lee et al., 2011). The c-Myc protein is up-regulated in response to growth-stimulatory signals to trigger the cell-cycle progression. Subsequently, c- Myc stimulates Romo1 expression. ROS derived from Romo1 expression then trigger Skp2-mediated c-Myc degradation in a negative feedback mechanism.

Romo1 expression is up-regulated in most cancer cell lines, suggesting that increased Romo1 expression might confer chronic oxidative stress to tumor cells (Chung et al., 2006; Hwang et al., 2007). There have been some reported experimental data about Romo1’s contribution to carcinogenesis. The enhanced Romo1 expression might induce the endogenous production of ROS, which could cause many DNA mutations in tumor cells, activate some oncogenes or suppress some of the tumor suppressor genes. Indeed, enforced Romo1 expression increases intracellular ROS levels and nuclear DNA damage (Chung et al., 2008). This finding demonstrates that ROS produced by Romo1 modulation move into the nucleus. There is still no direct evidence that Romo1 contributes to human carcinogenesis in vivo. Therefore, further studies will be needed to elucidate the precise role of Romo1 during carcinogenesis.

Romo1 could be a promising drug target, and drug development targeting Romo1 is possible in TNF-α pathobiology. The balance between survival and death after TNF-α stimulation is well controlled by crosstalk between the pro-survival pathway and the pro-apoptotic pathway. Deregulated activation of TNF- α signaling has been reported to contribute to a variety of diseases (Chen and Goeddel, 2002). So far, many agents that block TNF-α function, for example antibodies specific for TNF-α such as Infliximab, Adalimumab, Golimumab, and CDP571 or antibodies specific for TNF-R2 including Etanercept, have been developed, and the clinical trials were successful (Zidi et al., 2010). Since these agents bind to TNF-α or TNF-R to inhibit the physiological function, they may elicit unwanted outcomes including lymphoma development and infection (Bongartz et al., 2006). Romo1 expression is increased in pathological conditions, such as in senescent cells and cancer cells, and TNF-α- induced ROS production is involved in inflammatory diseases (Chung et al., 2006; 2008; Han et al., 2009). Therefore, we suggest that Romo1 is a novel and promising target for the development of new anti-inflammatory drugs that target TNF-α signaling. These agents targeting Romo1 can block the TNF-α- derived ROS production but not the complex I-mediated pathway of TNF-α signaling. The agents can also be utilized as inflammatory agents that do not remove the already-produced ROS but rather block the origin of ROS production. We expect that a Romo1 antagonist that inhibits the cell death pathway of TNF-α signaling will be more effective than current antioxidants for curing diseases associated with TNF-α.

REGULATION OF CYTOCHTOME P450

The cytochrome P450 (CYP) superfamily is a large group of enzymes that catalyze the oxidation of hydrophobic organic molecules. CYP enzymes have been identified in all branches of life (Danielson, 2002). The substrates of mammalian CYP enzymes include metabolic intermediates such as lipids and steroidal hormones, as well as xenobiotic substances such as drugs and other toxic chemicals. The most common reaction catalyzed by mammalian CYPs is a monooxygenase reaction in which CYPs bind oxygen and use one atom of oxygen to hydroxylate their substrate (RH) and the other to produce H2O:

  • RH + O2 + 2H+ + 2e → ROH + H2O

The two electrons required for the conversion of oxygen atom to H2O are generally furnished by NADPH. Two types of routes are known for the passage of the electrons from NADPH to mammalian CYPs, which are present in both endoplasmic reticulum (microsomes) and mitochondria. In the monooxygenation reaction catalyzed by microsomal CYP, electrons are passed to CYP via cytochrome p450 reductase which is comprised of two coenzymes, FAD and FMN (Degtyarenko and Kulikova, 2001). In contrast, the electron transfer to mitochondrial CYPs depends on adrenodoxin reductase, which is an FAD-containing flavoprotein and adrenodoxin, which is a ferredoxin type iron-sulfur protein (Ziegler et al., 1999).

CYP enzymes share a common structural topology with a heme group in their active site and derive their name from the absorption maximum of their reduced carbon monoxide complex at 450 nm (Coon, 2005).The heme iron is tethered to the CYP protein via a thiolate ligand derived from a cysteine residue. This cysteine and several flanking residues are highly conserved in CYPs. In the substrate-free form, CYPs have their heme iron predominantly in its six-coordinated low spin state (Fig. 3). The substrate (RH) binds to the active site of the enzyme, in close proximity to the heme group, on the side opposite to the cysteine. The bound substrate induces a change in the conformation of the active site, displacing a water molecule from the heme iron, shifting the spin equilibrium towards to the five-coordinated high spin form (step 1). The change in the spin state favors the transfer of an electron from NADPH via cytochrome P450 reductase or adrenodoxin, reducing the ferric heme iron to the ferrous state (step 2). The ferrous CYP then binds an oxygen molecule to form an oxy-complex of CYP (step 3). A second electron is transferred via either cytochrome P450 reductase or adrenodoxin to the oxy-complex of CYP to give a peroxy-complex (step 4). The peroxo group formed in step 4 is rapidly protonated twice and results in the heterolytic cleavage of the O-O bond, producing one water molecule, and forming a highly reactive species commonly referred to as Compound I (Rittle and Green, 2010) (Step 5). The activated oxygen atom of compound I is then inserted into the substrate molecule to produce ROH (step 6).

Fig. 3. The catalytic cycle of cytochrome p450 (CYP) with the main cycle that leads to substrate hydroxylation (shown in black arrows) and with the leaky branches that lead to ROS production (shown in red arrows). See text for details.

Fig. 3.

During the catalytic cycle, a significant portion of the activated oxygen is released from the catalytic intermediate to produce superoxide or H2O2 (Gorsky et al., 1984; Hanukoglu, 2006; Zangar et al., 2004). One such intermediate is the one-electron reduced oxy-complex (produced in step 3), which decays to produce superoxide (branch 1). The other intermediate is the two-electron-reduced peroxy-complex, which is protonated with the formation of H2O2 (branch 2). Continuous production of such ROS is an inevitable result of NADPH consumption by CYP both in the presence and in the absence of substrates. However, the partition of these branches depends on many conditions, such as the types of CYP and substrate, pH, ionic strength, oxygen concentration, and other factors (Gorsky et al., 1984; Hanukoglu, 2006; Zangar et al., 2004). The efficiency of electron transfer from NADPH through the electron carriers to the CYP for monooxygenation of substrate is called the degree of coupling. If a portion of electrons are transferred to other acceptors, such as O2 leading to production of ROS, then the process is referred to as “leaky” electron transfer. Although the degree of coupling (or leakiness) varies for different CYP species, it is usually less than 50% (Gorsky et al., 1984; Hanukoglu, 2006; Zangar et al., 2004; Zhukov and Archakov, 1982). Due to such leakiness, the rate of NADPH and oxygen consumption by CYPs is only weakly dependent on the presence of substrates. Even in the absence of any substrates, the CYP electrontransfer chain continues to oxidize NADPH and produce ROS.

The microsomal CPYs and their reductase are bound to the outer surface of the ER membrane by transmembrane anchors, whereas the components of mitochondrial CYP system (CYP, adrenodoxin, and adrenodoxin reductase) are located on the matrix side of the inner membrane mitochondria membrane (Gorsky et al., 1984; Hanukoglu, 2006; Hanukoglu et al., 1990; Zangar et al., 2004). The amount of microsomal CPY system components and the intensity of electron transfer processes in the ER are often underestimated. Despite the prevailing impression that most electron transfer processes in the cell are localized in the mitochondria, more than 60% of the electrontransfer hemoproteins and about 20-30% of membrane-bound flavoproteins are localized in the ER of the liver cell (Venditti et al., 1998). Although the amount of CYPs in other tissues is usually lower than in liver (McKinnon and McManus, 1996), it commonly remains comparable with the content of mitochondrial electron carriers. In steroidogenic cells, the amounts of mitochondrial CYP system components are as much as 10 times higher than other electron-transfer chain proteins (Hanukoglu, 2006). Thus, even if the CYP systems leak electrons at low rates, their total capacity for ROS generation could be high, especially in the liver and steroidogenic tissues.

ROS produced by CYP systems can potentially cause lipid peroxidation, cell toxicity, and death. Mitochondria are known to have a rich repertoire of antioxidant enzymes such as manganes superoxide dismutase (Mn-SOD), and glutathione peroxodase 1 and 5, as well as small antioxidant molecules such as ascorbic acid, tocopherol, and carotene to prevent the toxic effect of ROS (Hanukoglu, 2006). Peroxiredoxins (Prxs) are peroxidases that catalyze the reduction of reactive oxygen species (ROS) (Rhee and Woo, 2011). There are six different mammalian isoforms Prx I to VI), each with different subcellular localization. Prx III is found exclusively in mitochondria and plays an essential role in mitochondrial homeostasis (Chang et al., 2004). The active site cysteine residue of members of the 2- Cys Prx subgroup (Prx I to IV) of Prxs is hyperoxidized to cysteine sulfinic acid (Cys-SO2) during catalysis with concomitant loss of peroxidase activity. Reactivation of the hyperoxidized Prx is catalyzed by sulfiredoxin (Srx) (Rhee and Woo, 2011).

It is well known that ethanol consumption induces the accumulation of cytochrome P450 2E1 (CYP2E1), a major contributor to ethanol-induced ROS production in the liver (Gong and Cederbaum, 2006). It was shown recently that chronic ethanol feeding markedly increased the expression of Srx in the liver of mice in a largely Nrf2-dependent manner (Bae et al., 2010). Among Prx I to IV, only Prx I was found to be hyperoxidized in the liver of ethanol-fed wild-type mice because a large fraction of Prx I is located together with CYP2E1 at the cytosolic side of the ER membrane (Bae et al., 2010). The selective role of Prx I in ROS removal is thus likely attributable to the proximity of Prx I and CYP2E1. The pivotal functions of Srx and Prx I in protection of the liver in ethanol-fed mice were evident from the severe oxidative damage observed in mice lacking either Srx or Prx I (Bae et al., 2010).

REGULATION OF NADPH OXIDASE

General structure of NADPH oxidase

Growing evidences indicate that transient ROS can be made by receptor-mediated enzymatic processes, but not simple electron- leaking from mitochondrial membrane. Agonist-mediated ROS generation for host defense has been extensively studied in gp91phox/NADPH oxidase 2 (Nox2) of phagocytic cell. However, after identification of homologs of gp91phox (Nox1, Nox3, Nox4, Nox5, Duox1 and Duox2) in non-phagocytic cells the function of generated ROS has been extended into understanding of various cellular events including cell growth, differentiation, apoptosis, and immune responses. Since gp91phox (renamed as Nox2) was identified in phagocytic cells, six homologues of Nox2 in non-phagocytic cells to date have been reported. Seven Nox isozymes are all single polypeptides and can be divided into three types based on their functional domains. The catalytic protein Nox2 serves as the core protein of NADPH oxidase complex in phagocytic cells (Segal, 2005). The carboxyl terminal region of Nox2 contains NADPH- and FAD-binding sites homologues to ferredoxin-NADP+ reductase (FNR) (Segal et al., 1992; Sumimoto et al., 1992) and NH2- terminal region of Nox2 bears six transmembrane α-helical domains containing tandem heme groups homologues to ferric reductase (Fre) (Finegold et al., 1996) (Fig. 4). Active form of Nox2 mediates an electron transport from NADPH (Km for NADPH is 40 μM; Km for NADH is 2.5 mM) through FAD and two hemes to oxygen finally to generate superoxide anion (Clark et al., 1987). Nox1, Nox3, and Nox4 have similar structure with Nox2 and then categorized into prototype. Nox5 and Duox1/2 contain EF hand structure and appear to be activated by intracellular calcium (Figs. 4D and 4E).

Fig. 4. Structure and activation of NADPH oxidase (Nox) isozymes. Nox isozymes contain six transmembrane α-helical domains containing tandem heme binding sites in NH2-terminal region and NADPH- and FAD-binding sites in long carboxyl terminal region. Nox1, Nox3, and Nox4 have similar structure with Nox2 and then categorized into prototype. Nox5 and Duox1/2 bear EF hand structure in NH2-terminal region. Additional peroxidasge-like domain of Duox is located in extracellular NH2-terminal region. p22phox protein contains proline rich region (PRR) providing a binding site for p47phox and NoxO1. See text for details.

Fig. 4.

Oxidase domain of Nox5 constitutes prototype structure and additional NH2-terminal region bearing four EF-hands (three canonical motifs and one uncanonical motif) as regulatory domain. Increasing intracellular calcium binding to EF-hands in Nox5 induces conformational change leading to intramolecular interaction of EFhands with oxidase domain in NH2-terminal region (Banfi et al., 2004). Because calcium affinity in EF-hands of Nox5 is relatively low (EC50 of calcium is 1 μM), additional regulatory mechanism is required for calcium sensitivity. Either phosphorylation of FAD-binding domain or association of calmodulin with NADPH-binding site in Nox5 increases calcium sensitivity (Jagnandan et al., 2007; Pandey et al., 2011; Tirone and Cox, 2007). Duox isozymes contain two EF-hands in NH2-terminal region (De Deken et al., 2000). Intracellular calcium binds to EF-hands and induces conformational changes resulting in Duox activation. Partial digestion of EF-hands by protease treatment shows increased Duox activity suggesting that EFhands of Duox act as negative regulator (Dupuy et al., 1992). Peroxidase-like domain of Duox is located in extracellular NH2- terminal region. Although peroxidase activity generally requires for heme group, peroxidase-like domain of Duox does not contain to heme group. Because critical histidine residues in conserved region were replaced by serine in peroxidase-like ectodomain of Duox, it appears to be inactive form (Meitzler and Ortiz de Montellano, 2009). It has been well established that Nox enzyme catalyzes NADPH-dependent one electron reduction of oxygen to generate superoxide anion (O2-) that could be dismutated by spontaneously or enzymatic reaction to generate H2O2 (Nakamura et al., 1989; 1991). However, Duox in thyroid and lung airway cells appears to be source of H2O2 (Dupuy et al., 1991; Forteza et al., 2005; Leseney et al., 1999). EPR (Electron paramagnetic resonance) experiment with spin trap BMPO demonstrated that Duox generates H2O2 directly through unknown mechanism (Ameziane-El-Hassani et al., 2005).

Regulation of gp91phox/Nox2

The catalytic protein gp91phox (renamed as Nox2) serves as the core protein of NADPH oxidase complex in phagocytic cells (Segal, 2005). Catalytic core protein, gp91phox/Nox2 needs to form a complex with one membrane-integrated protein, p22phox, three cytosolic proteins- p47phox, p40phox, and p67phox- and small G-protein Rac. Integral protein p22phox tightly associates with catalytic protein Nox2 to form cytochrome b558 (Borregaard et al., 1983; Huang et al., 1995) providing for the stabilization of Nox2 and recruitment of cytosolic protein p47phox. Expression of Nox2 is undetectable in phagocytes from chronic granulomatous disease patients with defect of p22phox suggesting heterodimerization of Nox2 with p22phox is critical for the stabilization of Nox2 expression (Dinauer et al., 1990; Parkos et al., 1989; Stasia et al., 2002) (Fig. 4A). Moreover, p22phox protein seems to be important integral partner for other Nox isozymes such as Nox1, Nox3, and Nox4 (Ambasta et al., 2004; Kawahara et al., 2005; Martyn et al., 2006; Takeya et al., 2003; Ueno et al., 2005). The p22phox protein contains proline-rich region (PRR) in carboxyl terminal region providing an interacting site for tandem Src homology 3 (SH3) domains of p47phox (DeLeo et al., 1995; Leto et al., 1994).

The p47phox protein acts as adaptor molecule for the formation of NADPH oxidase complex. It contains phagocyte oxidase (PX) domain in NH2-terminal region for targeting phosphoinositide lipids in membrane, tandem SH3 domains in central region for binding with p22phox, autoinhibitory region (AIR) next by SH3 domain, and proline-rich region (PRR) in carboxyl terminal region for recruiting SH3 domain of p67phox. In resting stage, tandem SH3 domains of p47phox protein form intra-molecular interaction with AIR interfering interaction with other proteins (Groemping et al., 2003; Yuzawa et al., 2004a; 2004b). Upon response to stimuli, extensive phosphorylation of AIR by various protein kinases (PKC, Erk, p38MAPK, Pak, and Akt) (Chen et al., 2003a; El Benna et al., 1994; 1996; Fontayne et al., 2002; Martyn et al., 2005) leads to conformational change in p47phox which renders interaction of tandem SH3 domain of p47phox with p22phox and PRR in carboxyl terminal region to SH3 domain of p67phox protein 2003(Ago et al., 2003). Conformational change by phosphorylation also allows PX domain accessible to phospho inositide lipid such as phosphtidylinositol (3,4)-bisphosphate [PI(3,4)P2], a product of phosphatidylinositol 3-kinase (PI3K), in membrane (Ago et al., 2003). PX domain also has a binding activity with acidic phospholipids including phosphatidic acid (PA) and phosphatidylserine (PS) which support membrane binding of p47phox (Karathanassis et al., 2002; Stahelin et al., 2003).

Small G-protein, Rac protein is a essential molecule for the action of Nox2 (Abo et al., 1991; Knaus et al., 1991; Mizuno et al., 1992). The p67phox protein contains four tetratricopeptide repeats (TPR) for a binding site of G-protein Rac (Koga et al., 1999; Lapouge et al., 2000). TPR domains consist of 34 amino acid repeat (Lamb et al., 1995; Ponting, 1996) and two hairpin loop connecting TPR1-TPR2 and TPR2-TPR3 provide an interacting site for GTP-bound form of Rac (Lapouge et al., 2000). Truncated domain of p67phox (amino acids 1-199) is completely loss of Nox2 activation, whereas longer stretch domain of p67phox (amino acids 1-212) restores their activity (Hata et al., 1997). Short stretch region (amino acids 201-210) of p67phox was named as activation domain (AD) (Han et al., 1998). Moreover, mutant V204A of p67phox shows null function of Nox2 activation (Han et al., 1998; Price et al., 2002). Thus, TPR domain in NH2-terminal region and short stretch of activation domain are essential for p67phox function. Two SH3 domains are located into carboxyl terminal region. Although the function of NH2-terminal SH3 is unclear, carboxyl terminal SH3 domain can specifically interact with PRR of p47phox leading to translocation of Nox2 protein complex (de Mendez et al., 1997; Hata et al., 1997). The p67phox protein contains PB1 domain in between two SH3 domains mediating molecular interaction with PB1 domain of p40phox (Kuribayashi et al., 2002). In crystal structure of p67phox and p40phox, interaction of Lys355 of p67phox with Asp289, Glu291, and Asp293 of p40phox contributes intermolecular association (Wilson et al., 2003). Translocation of p40phox is mediated by PX domain located in NH2-terminal region (Ueyama et al., 2007). PX domain in p40phox can specifically bind to phosphatidylinositol 3-phosphate [PI(3)P], highly enriched in the phagosomal membrane (Bravo et al., 2001; Kanai et al., 2001; Stahelin et al., 2003). However, the function of SH3 domain in p40phox remains to be elucidated.

Regulation of Nox1

Nox1 is firstly identified as a homologue of gp91phox/Nox2 from non-phagocytic cells, colon epithelium (Suh et al., 1999) and requires for one integral protein, p22phox and two cytosolic proteins, Nox organizer 1 (NoxO1) as homologue of p47phox, Nox activator 1 (NoxA1) as homologue of p67phox, and small Gprotein Rac (Cheng et al., 2006; Geiszt et al., 2003a; Kawahara et al., 2005). Function of p22phox for Nox1 complex is similar to that for Nox2 (Fig. 4B). The p22phox protein interacts with catalytic core protein Nox1 and recruits NoxO1 protein through the interaction of PRR of p22phox with SH3 domain of NoxO1 (Takeya et al., 2003). NoxO1 protein acts as central scaffolding molecule in Nox1 activation. Structure of NoxO1 is very similar with that of p47phox protein except lacking of AIR domain. It contains PX domain in NH2-terminal region for binding of phosphoinositide lipid, tandem SH3 domains for PRR of p22phox, and PRR in carboxyl terminal region for the interaction with SH3 domain of NoxA1. Because NoxO1 protein lacks of AIR, free tandem SH3 domains of NoxO1 interact with PRR of p22phox providing translocation of NoxO1 to membrane (Cheng and Lambeth, 2004). Therefore, Nox1 appears to be constitutively active. NoxA1 protein contains four TPR domains in NH2- terminal region for Rac binding sites and one SH3 domain located in central region for interaction with PRR of NoxO1 as similar as p67phox (Cheng et al., 2006; Takeya et al., 2003). However, there is no supporting evidence that PB1 domain of NoxA1 can interact with p40phox or its homologue.

The multiprotein complex formation seems to be necessary not only for the activation but also for the regulation of Nox activity. The c-Src substrate protein Tks4 (tyrosine kinase substrate with four SH3 domains) is functional member of a p47phox-related organizer superfamily and interact with PRR of NoxA1 protein through SH3 domain-mediated interaction resulting in Nox1 activation (Gianni et al., 2011). Interestingly, Src-mediated phosphorylation of Tyr110 on NoxA1 and of Tyr508 on Tks4 stimulate their association and induce Nox1-dependent ROS generation (Gianni et al., 2010). However, Tks proteins do not support Nox2 and Nox4 activities. It has been shown that PKA phosphorylates Ser172 and Ser461 of NoxA1 providing a binding site for 14-3-3 protein (Kim et al., 2007a). Association of phosphorylated NoxA1 with 14-3-3 results in dissociation of NoxA1 from Nox1 protein complex and cessation of superoxide anion production. In addition, recent two reports showed that the action of Erk leads to phosphorylation of Ser282 of NoxA1 and that phosphorylation of Ser282 in turn results in negative regulation of Nox1 activity (Kroviarski et al., 2010; Oh et al., 2010).

Regulation of Nox3 and Nox4

Since Nox3 gene was identified from fetal tissues, substantial evidence indicated that the Nox3 protein is expressed in inner ear. Positional cloning experiment demonstrated that defect of Nox3 results in head tilt mice (Nox3het) through impairment of otoconia formation in inner ear (Paffenholz et al., 2004). After head-tilt mice were due to non-function of Nox3 in inner ear, these mice models (nmf333 mice with defect of p22phox, hstl mice with defect of NoxO1) allow understanding the regulatory mechanism in which p22phox and NoxO1 are required for activation of Nox3 (Kiss et al., 2006; Nakano et al., 2008) (Fig. 4B). Although the transfection with p67phox, p47phox, and Rac1 protein stimulates Nox3 (Cheng et al., 2004; Miyano et al., 2009; Ueyama et al., 2006), the regulatory mechanism of Nox3 remains to be elucidated.

Nox4 protein is typical catalytic subunit which requires for only p22phox and seems to be constitutively active (Ambasta et al., 2004; Martyn et al., 2006) (Fig. 4C). Substantial evidence indicated that the expression of Nox4 is inducible and the activity is dependent on its expression level. Although molecular partner for Nox4 to date has not been reported, polymerase delta-interacting protein (Poldip2) as a p22phox-interacting protein stimulates Nox4 activity in SMC (Lyle et al., 2009). However, the molecular mechanism and function of Nox4 by Poldip2 remains to be determined.

Regulation of Nox5 and Duox1/2

No binding protein for Nox5 has been reported. Integral protein p22phox can not be interacted with Nox5 and knockdown of p22phox does not affect Nox5 activation (Brar et al., 2003; Kawahara et al., 2005). Activation of Nox5 protein only requires for intracellular calcium because it contains four EF-hands in NH2-terminal region (Banfi et al., 2001; 2004) (Fig. 4D). Moreover, PMA-dependent phosphorylation of Nox5 on Thr494 and Ser498 allows increasing calcium sensitivity (Jagnandan et al., 2007). It is likely that concerted action of phosphorylation and calcium binding is involved in the regulation of Nox5 activity. However, studies of molecular function of Nox5 are limited because Nox5 does not express in mouse.

As like as Nox5, Duox isozymes requires intracellular calcium mobilization for their activation through calcium binding to EF-hands (Ameziane-El-Hassani et al., 2005). The association induces conformational changes resulting in Duox activation. Recently, Duox activators, DuoxA1 and DuoxA2, identified as maturation factor are essential for the delivery of Duox from endoplasmic recticulum to plasma membrane (Grasberger and Refetoff, 2006) (Fig. 4E). DuoxA1 and DuoxA2 are located on chromosome 15 in a head to head arrangement with Duox1 and Duox2, respectively, suggesting that the expression of Duox1/DuoxA1 and Duox2/DuoxA2 are coordinated by a common transcription (Pachucki et al., 2004). DuoxA provides Duox translocation from ER-Golgi to plasma membrane as well as stimulates glycosylation of Duox protein during subcellular translocation of Duox providing stable conformation between Duox and DuoxA (Grasberger and Refetoff, 2006). Moreover, Leto and his colleagues showed that functional pairing of Duox1/DuoxA1 and Duox2/DuoxA2 contributes H2O2 production in reconstitution experiments (Morand et al., 2009).

Function of Nox in host defense

The first described microbicidal function of Nox was in phagocytic cells such as macrophages and neutrophils highly expressing Nox2. Superoxide anion generated by NOX isozymes is not much contributed in microbial killing because it is less reactive. Instead, phagocytic cells utilize more relevant mode in which HOCl is generated in sufficient concentrations to kill bacteria by combined action of H2O2 and secrected myeloperoxidase (MPO) in phagosome (Jiang et al., 1997). However, genetic deficiency in MPO does not able to show antimicrobial defect as like as shown in the defect of gp91phox at an increased risk of infection (Kutter et al., 2000). While MPO has a contribution to Nox-mediated antimicrobial mechanisms, several lines of evidence suggest that there is MPO-independent microbicidal pathway. In addition, Nox2 activation leads to rise in phagosomal pH through the generation of hydroxyl ion from superoxide anion, leading to a charge build-up that requires compensation (Henderson et al., 1987; Schrenzel et al., 1998). Most of the charge compensation not only occurs through H+ channels opening but also K+ influx (Ahluwalia et al., 2004; Reeves et al., 2002; Segal, 2005). Moreover, K+ influx could contribute to bacterial killing through changes in phagosomal osmolarity and activation of cationic proteases such as elastase and cathepsin G (Reeves et al., 2002). The host defense concept in macrophage suggested that both of HOCl generated by concerted action of Nox2 with MPO and the charge compensation by K+ influx are essential for microbial killing. Based on this mechanism, it is clear that chronic granumatous disease (CGD) is caused by the genetic mutations of gp91phox, p22phox, p47phox, and p67phox and fails to generate H2O2 leading to a failure to kill microbes properly. The CGD patients provide a clear role of Nox2 and its accessory proteins in phagocytic cells.

Since Duox isozymes were identified as thyroid NADPH oxidases producing H2O2 for biosynthesis of thyroid hormone (De Deken et al., 2000; Dupuy et al., 1999), it has been reported that Duox isozymes are found in the ducts of the salivary gland (Geiszt et al., 2003b), airway epithelia, rectal mucosa, and intestinal colon epithelia (Dupuy et al., 2000; El Hassani et al., 2005). Conceptually, epithelia of airway and intestinal colon are front line to contact many bacterial species and require appropriate machinery to controlling microbes. Substantial evidences indicate that both of Duox and lactoperoxidase (LPO) secreted in airway and mucosal epithelia appear to be an important role in microbial killing. LPO generates hypothiocyatnate (OSCN-) from the oxidation of SCN- by Duox-dependent H2O2 (Geiszt et al., 2003b). In airway cells, amount of H2O2 (about 90 nM) produced in exhale breath appears to be sufficient for maintaining LPO activity (Schleiss et al., 2000). ATP binding cassette (ABC) family protein, cystic fibrosis transmembrane conductance regulator (CFTR) was shown to transport Cl- (Tabcharani et al., 1993; Verkman et al., 2003) as well as SCN- in apical membrane of airway epithelium (Moskwa et al., 2007). It is shown that Duox-induced H2O2 serves as fuel for LPO-mediated hypothiocyanate production, providing robust antimicrobial defense network in epithelial cells.

Because Duox contains intracellular EF hand Ca2+ binding motif, agonists triggering intracellular Ca2+ mobilization appear to activate Duox isozymes. Extracellular ATP binds to purinergic receptor expressed on surface of airway epithelia cells and then mobilizes intracellular Ca2+ through classical G-protein signaling cascade, leading to H2O2 production followed by Duox activation on apical membrane (Forteza et al., 2005). It has recently reported that calcium-dependent Duox activity was regulated by protein kinases (Rigutto et al., 2009) and interaction with NoxA1 (Pacquelet et al., 2008). These experimental results provide an insight to regulation mechanism of Duox activity. Since genetic defect of Duox gene associates to hypothyroidsm (Moreno et al., 2002), a few clinical reports indicate correlation of Duox with lung and mucosal diseases. Knaus and collegues reported that silencing Duox gene by hypermethylation on CpG-rich promoter region results in increasing incidence of lung cancer suggesting that Duox plays an important role in mucosal homeostasis and wound healing (Luxen et al., 2008).

Receptor-mediated Nox activation

Recently, many reports have demonstrated that intracellular ROS, produced in mammalian cells in response to the activation of various receptors, serve as important second messengers in cell signaling. Cysteine residue in active center has low pKa and is very susceptible to oxidant. Transient produced ROS oxidized sulfhydryl group (-SH) of cysteine to sulfenic (-SOH). Several reports indicate that cysteine residue of protein tyrosine phosphatase is oxidized by agonist-induced ROS and converts to inactive form leading to increasing tyrosine phosphorylation in cells. Because uncontrolled ROS generation is very toxic, the activity of Nox isozymes in cells is tightly regulated by receptor and its signaling cascades. In following part, we will discuss Nox activation and ROS generation through receptor-mediated cell signaling.

Growth factor receptor

Upon ligation of growth factor to their receptor on cell surface, the ligand stimulates receptor dimerization resulting in transactivation of intracellular protein tyrosine kinase (PTK). Tyrosine phosphorylation in COOH-terminal region by PTK serves as binding site for cytosolic signaling proteins containing Srchomology domain 2 (SH2). It activates phospholipase Cγ (PLCγ)-PKC, phophatidylinositol 3-kinase (PI3K)-Akt, and Grb2- Sos-Ras-Erk pathway. Since the novel finding of PDGFdependent H2O2 production in SMC by Finkel and colleagues (Sundaresan et al., 1995), many reports demonstrated that PDGF stimulates ROS generation in various cell types such as endothelial cells, heaptocytes, pancreatic satellite cells, and fibroblast. Although the expression of Nox2 is low in the nonphagocytic cells, substantial reports showed that Nox2 and their accessory proteins are involved in ROS generation in response to PDGF in endothelial cells, smooth muscle cells, and pancreatic satellite cells. The p47phox was phosphorylated by PKC and translocated to membrane in response to PDGF in EC and SMC. Moreover, vascular SMC from p47phox+/+ mice was activated and produced ROS in response to PDGF or PMA, whereas p47phox-/- SMCs failed to generate ROS. PI3K generates PI(3)P, PI(3,4)P2, and PI(3,4,5)P3 from PI, PI(4)P and PI(4,5)P2, respectively. The phosphorylated PI serves as binding site for Phox homology (PX) domain in cytosolic proteins of Nox complex and pleckstrin homology (PH) domain in RacGEF protein. Interaction of PI(3,4)P2 or PI(3,5)P2 with p47phox or NoxO1 stimulated membrane translocation and activated Nox2 or Nox1 complex. Rac is essential component for Nox activity and is regulated by Guanine nucleotide exchange factors (GEF) as Rac activator and GTPase-activating proteins (GAP) as Rac inhibitor. Activation of βPix as RacGEF through the interaction of PI(3,4,5)P3 or PI(3,4)P2 with PH domain contributes to stimulate Nox1 activity (Park et al., 2004). Moreover, βPix binds to COOH-terminal region of Nox1 providing a complex formation of Nox1, βPix, and NoxO1 (Park et al., 2006b). It has been reported that basic fibroblast growth factor (bFGF) stimulates PI3K-Rac pathway resulting in Nox1 activation in SMCs (Schroder et al., 2007). Insulin induced ROS production through the regulation of Nox4 expression resulting in adipocyte differenttiation and depletion of Nox4 by the transfection of Nox4 siRNA inhibited Erk phosphorylation (Mouche et al., 2007). However, the report did not provide underlying molecular mechanism between insulin signaling pathway and epigenetic regulation for Nox4 expression.

G-protein coupled receptor

G-protein coupled receptor contains seven transmembrane domains and cytosolic long carboxyl terminal region providing heterotrimeric G-proteins. G-proteins are consisted of three subunits, α, β, and γ. Alpha subunit of G protein (Gα) can bind GDP in inactive state and become exchanged GTP for GDP leading to dissociation of Gβγ complex and in turns to active state. Binding of ligand induced conformational change of receptor resulting in the activation of G-protein. Angiotensin II (Ang II) binds directly to the angiotensin type 1 receptor (AT1R) which belongs to the G protein-coupled receptor family. Engagement of AT1R elicits a complex network of signaling cascades that result in short-term vascular effects such as contraction, as well as long-term effects such as cell growth, migration, and adhesion, that eventually lead to pathophysiological vascular remodeling and cardiac hypertrophy (de Gasparo et al., 2000; Spat and Hunyady, 2004). Ang II-AT1R complex elicited a biphasic ROS production in SMCs. Early ROS production was responsible for Gαq-mediated activation of phospholipase C (PLC)-β, which is followed by Ca2+ mobilization and activation of protein kinase C (PKC). In reconstituted cells (Nox1-NoxO1- NoxA1 or Nox2-p47phox-p67phox) stimulated with Ang II, Nox2 and Nox1 are activated by Ang II-AT1R signaling cascade. However, the activity of Nox2-p47phox-p67phox complex is required for the activation of PKC, PI3K and Rac1, rather than that of Nox1-NoxO1-NoxA1 complex. Second peak for ROS generation by Ang II was resulting from up-regulation of Nox1, p22phox, and p47phox (Choi et al., 2008). Transcriptional factor Ets1 plays a critical for Ang II-mediated ROS generation by regulating the expression of p47phox (Ni et al., 2007).

Toll-like receptor

In human, 10 members of the Toll-like receptor (TLR) family have been identified and shown to be involved in innate immunity and inflammation responses. TLRs are belong to TLR/ interleukin 1 receptor superfamily containing leucine-rich-repeat (LRR) domain for ligand binding site and conserved carboxyl terminal region (Toll/IL-1R, TIR domain) for recruitment of intracellular signaling molecules. Extracellular domain of TLRs recognizes pathogen-associated molecular patterns (PAMPs) on the surface of pathogens. Most of TLRs constitute homodimer formation for recognition of pathogens (Beutler, 2009; Kawai and Akira, 2010). However, TLR2 forms with heterodimer with TLR1 or TLR6. TLR2-TLR1 and TLR2-TLR6 recognize triacety-lated and diacetylated lipoprotein, respectively. TLR4 binds to lipo-polysaccaride (LPS), an integral component of the outer membrane of Gram-negative bacteria, and TLR5 recognizes bacterial flagellin. TLR7/8 and TLR9 are involved in recognition of viral single strand RNA and bacterial DNA containing CpG motifs. Ligation of TLRs by ligand leads to stimulation of signaling networks conferring an increased expression of host defense proteins and inflammatory cytokines. Upon binding of ligand with TLRs, the cytoplasmic region of TLRs recruits adaptor protein myeloid differentiation factor 88 (MyD88) or TIR domain-containing adaptor protein-inducing IFNβ (TRIF) (Baccala et al., 2007). The MyD88-dependent pathway appears to be involved in the recruitment of IL-1R associated kinase (IRAK) associated with TRAF6 leading to NF-κB activation. Activated cells by TLR3 agonist also operate MyD88-independent TRIF-dependent pathway stimulating IRF3 for INFβ production and delayed NF-κB activation through PKC, PI3-Kinase, and Akt.

It has been reported that Nox1 is involved in TLR-mediated innate immunity and host defense in gastric and colon epithelial cells (Rokutan et al., 2006). LPS from Helicobacter pylori stimulated ROS generation through the induction of Nox1 and NoxO1 expression in guinea pig gastric pit cells resulting from TLR4-dependent NF-κB activation pathway. LPS-dependent ROS contributed the overexpression of tumor necrosis factor α (TNFα) and cyclooxygenase in gastric pit cells suggesting that it might be involved in Helicobacter pylori-dependent inflammation and pathogenesis. Flagellin from Salmonella enteritides and recombinant flagellin (rFliC) were shown to induce ROS generation in guinea pig large intestinal epithelial cells (LIECs) and T84 human colon epithelium co-transfected with NoxO1 and NoxA1 leading to triggering inflammatory cytokine IL-8 production through TGFβ-activated kinase 1 and TGFβ-activated kinase 1-binding protein 1 and induction of Nox1 through NF-κB and INFγ. Moreover, INFγ-stimulated transcription factor appears to be involved in Nox1 expression leading to mucosal host defense and pathogenic inflammation in the colon.

It has been reported that binding of LPS to TLR4 triggers various cell signaling pathways including NF-κB activation and ROS production. LPS-induced ROS generation and NF-κB activation was mediated by a direct interaction of TLR4 with Nox4 in HEK293T cells (Park et al., 2006a). Yeast two hybrid and GST pull-down assays indicated that the carboxyl terminal region of Nox4 interactes with the cytoplasmic region of TLR4. Knockdown of Nox4 using Nox4 siRNA in HEK293T cells expressing TLR4 along with MD2 and CD14 resulted in inhibition of LPS-induced ROS generation and NF-κB activation. Moreover, it has been demonstrated that Nox4-dependent ROS generation plays an important role in LPS-induced proinflammatory cytokine production by EC and activation, leading to adhesion molecule expression. Downregulation of Nox4 in HAECs resulted in a failure to induce ROS generation and subsequent expression of intercellular adhesion molecule-1 (ICAM-1) and chemokines such as IL-8 and monocyte chemo-attractant protein- 1 (MCP-1) in response to LPS. These results suggest that generated ROS from EC may play an important role in the process of atherogenesis because infiltration of blood vessel intima by leukocytes, aided by the expression of chemokines and adhesion molecules, is an initial and rate-limiting step in the development of atherosclerotic lesions. Recent report suggested that Nox4 is shown to be involved in chemokine CXCR6 expression through TLR4-Nox4-AP1 cascade (Patel et al., 2006). However, molecular connection between TLR4 and Nox4 in vascular cells is needed further studies.

TLRs also recognize endogenous ligands which are involved in inflammation process and are found in inflammation sites. The endogenous ligands appear to be damage-associated molecular patterns (DAMPs) which are derived from damage tissues and triggers inflammation cascades in injured sites. Apoptotic cells contain Oxidation of molecules generates oxidation- specific epitopes serving as danger signals to trigger inflammation reactions. It has been reported that minimally oxidized LDL (mmLDL) interacts with CD14 and induces secretion of certain cytokines through TLR4 (Miller et al., 2003; 2005). mmLDL stimulates ROS generation in macrophages via activation of TLR4, and a subsequent signaling cascade involving Syk, PLCγ, PKC and Nox2 (Bae et al., 2009). Moreover, the production and functional role of specific pro-inflammatory cytokines in response to mmLDL requires the) presence of Nox2. mmLDL stimulated intracellular reactive oxygen species (ROS) generation in macrophages through Nox2 through TLR4. ROS generation by mmLDL required the recruitment and activation of spleen tyrosine kinase (Syk) and that mmLDL also induced PLCγ1 phosphorylation and PKC membrane translocation resulting in Nox2 activation. Generation of ROS by Nox2 modulated mmLDL activation of macrophages by regulating the expression of proinflammatory cytokines IL-1β, IL-6 and RANTES. RANTES was able to stimulate migration of mouse aortic smooth muscle cells (MASMC). These data explained mechanisms by which endogenous ligands mmLDL for TLR4 can induce proatherogenic activation of macrophages. However, active molecule in mmLDL is still unclear. Oxidized form of 1-palmitoyl-2-arachidonyl-sn-glycerol-3-phosphocholine (Ox- PAPC) stimulated Nox activity resulting in induction of inflammatory and sterol regulatory genes in EC (Rouhanizadeh et al., 2005).

CONCLUSIONS

While moderate levels of ROS are useful for cellular proliferation and host defense, excessive amounts of ROS can be harmful to cellular metabolism and cause unwanted events, such as damage to DNA, RNA, and proteins, and induction of apoptosis or necrosis (Valko et al., 2006). Mitochondrial superoxide is continuously produced from the mitochondrial respiratory chain and the majority of ROS are generated by leakage of electrons in the mitochondrial respiratory chain. This ROS production contributes to mitochondrial and cellular damages, and ROS play a critical role in many diseases including cancer, arthritis, aging, neurodegenerative disorders, and diabetes (Droge, 2002). Since persistent accumulation of DNA damage induced by ROS is proposed to be a major contributor of cancer and aging, the possibility of ROS as therapeutic targets in anti-cancer therapy has been frequently raised. The levels of ROS in cancer cells are close to the threshold that causes cell death and the source of ROS production in tumor cells is different from that of normal cells (Laurent et al., 2005; Szatrowski and Nathan, 1991). Therefore, Kong et al. have suggested that up-regulation of ROS by treatment with antioxidant enzyme inhibitors or ROS inducers can cause further oxidative stress that preferentially triggers tumor cell death (Kong et al., 2000).

ROS generation is tightly controlled by various proteins including mitochondrial proteins, cytochrome p450 (CYP), and Nox isozymes. Although cells regulate ROS generation for the maintenance of redox homeostasis and cellular signaling pathways, the molecular mechanism by which the proteins located in sub-cellular organelles such as mitochondria, endoplasmic reticulum, endosome and plasma membrane regulate ROS generation still remains to be elucidated. If the expression and the activation of these proteins are dysregulated in cells, normal redox signaling will be disturbed and the subcellular organelles will sometimes face oxidative damages. Therefore, understanding and verifying mechanism of ROS generation will provide important information for drug development for diseases related to redox homeostasis.

Acknowledgments

This research was supported by Basic Science Research Program (2011-0025760 to YDY), by National Nuclear R&D Program (2011-0006304 to YDY), by the National Core Research Center (grant R15-2006-020-00000-0 to Y.S.B), by World Class University program (R31-2008-000-10010-0 to Y.S.B), by National Research Laboratory program (grant number ROA-2007- 000-20004-00 to Y.S.B) of the Ministry of Education, Science and Technology/National Research Foundation of Korea (NRF), and by a grant from the National R&D Program for Cancer Control, Ministry for Health, Welfare and Family Affairs (1020180 to YDY). H.J.O is recipient of a Brain Korea 21 scholarship.

References

  • 1.Abo A., Pick E., Hall A., Totty N., Teahan C.G., Segal A.W. Activation of the NADPH oxidase involves the small GTP-binding protein p21rac1. Nature. (1991);353:668–670. doi: 10.1038/353668a0. [DOI] [PubMed] [Google Scholar]
  • 2.Adler V., Yin Z., Tew K.D., Ronai Z. Role of redox potential and reactive oxygen species in stress signaling. Oncogene. (1999);18:6104–6111. doi: 10.1038/sj.onc.1203128. [DOI] [PubMed] [Google Scholar]
  • 3.Ago T., Kuribayashi F., Hiroaki H., Takeya R., Ito T., Kohda D., Sumimoto H. Phosphorylation of p47phox directs phox homology domain from SH3 domain toward phosphoinositides, leading to phagocyte NADPH oxidase activation. Proc. Natl. Acad. Sci. USA. (2003);100:4474–4479. doi: 10.1073/pnas.0735712100. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 4.Ahluwalia J., Tinker A., Clapp L.H., Duchen M.R., Abramov A.Y., Pope S., Nobles M., Segal A.W. The large-conductance Ca2+-activated K+ channel is essential for innate immunity. Nature. (2004);427:853–858. doi: 10.1038/nature02356. [DOI] [PMC free article] [PubMed] [Google Scholar] [Retracted]
  • 5.Amanullah A., Azam N., Balliet A., Hollander C., Hoffman B., Fornace A., Liebermann D. Cell signalling: cell survival and a Gadd45-factor deficiency. Nature. (2003);424:741. doi: 10.1038/424741b. discussion 742. [DOI] [PubMed] [Google Scholar]
  • 6.Ambasta R.K., Kumar P., Griendling K.K., Schmidt H.H., Busse R., Brandes R.P. Direct interaction of the novel Nox proteins with p22phox is required for the formation of a functionally active NADPH oxidase. J. Biol. Chem. (2004);279:45935–45941. doi: 10.1074/jbc.M406486200. [DOI] [PubMed] [Google Scholar]
  • 7.Ameziane-El-Hassani R., Morand S., Boucher J.L., Frapart Y.M., Apostolou D., Agnandji D., Gnidehou S., Ohayon R., Noel- Hudson M.S., Francon J., et al. Dual oxidase-2 has an intrinsic Ca2+-dependent H2O2-generating activity. J. Biol. Chem. (2005);280:30046–30054. doi: 10.1074/jbc.M500516200. [DOI] [PubMed] [Google Scholar]
  • 8.Andreyev A.Y., Kushnareva Y.E., Starkov A.A. Mitochondrial metabolism of reactive oxygen species. Biochemistry (Mosc) (2005);70:200–214. doi: 10.1007/s10541-005-0102-7. [DOI] [PubMed] [Google Scholar]
  • 9.Baccala R., Hoebe K., Kono D.H., Beutler B., Theofilopoulos A.N. TLR-dependent and TLR-independent pathways of type I interferon induction in systemic autoimmunity. Nat. Med. (2007);13:543–551. doi: 10.1038/nm1590. [DOI] [PubMed] [Google Scholar]
  • 10.Bae Y.S., Lee J.H., Choi S.H., Kim S., Almazan F., Witztum J.L., Miller Y.I. Macrophages generate reactive oxygen species in response to minimally oxidized low-density lipoprotein: toll-like receptor 4- and spleen tyrosine kinase-dependent activation of NADPH oxidase 2. Circ. Res. (2009);104:210–218. doi: 10.1161/CIRCRESAHA.108.181040. 221p following 218. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 11.Bae S.H., Sung S.H., Cho E.J., Lee S.K., Lee H.E., Woo H.A., Yu D.Y., Kil I.S., Rhee S.G. Concerted action of sulfiredoxin and peroxiredoxin I protects against alcohol-induced oxidative injury in mouse liver. Hepatology. (2010);53:945–953. doi: 10.1002/hep.24104. [DOI] [PubMed] [Google Scholar]
  • 12.Balkwill F., Coussens L.M. Cancer: an inflammatory link. Nature. (2004);431:405–406. doi: 10.1038/431405a. [DOI] [PubMed] [Google Scholar]
  • 13.Banfi B., Molnar G., Maturana A., Steger K., Hegedus B., Demaurex N., Krause K.H. A Ca(2+)-activated NADPH oxidase in testis, spleen, and lymph nodes. J. Biol. Chem. (2001);276:37594–37601. doi: 10.1074/jbc.M103034200. [DOI] [PubMed] [Google Scholar]
  • 14.Banfi B., Tirone F., Durussel I., Knisz J., Moskwa P., Molnar G.Z., Krause K.H., Cox J.A. Mechanism of Ca2+ activation of the NADPH oxidase 5 (NOX5). J. Biol. Chem. (2004);279:18583–18591. doi: 10.1074/jbc.M310268200. [DOI] [PubMed] [Google Scholar]
  • 15.Beutler B. Microbe sensing, positive feedback loops, and the pathogenesis of inflammatory diseases. Immunol. Rev. (2009);227:248–263. doi: 10.1111/j.1600-065X.2008.00733.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 16.Bialik S., Cryns V.L., Drincic A., Miyata S., Wollowick A.L., Srinivasan A., Kitsis R.N. The mitochondrial apoptotic pathway is activated by serum and glucose deprivation in cardiac myocytes. Circ. Res. (1999);85:403–414. doi: 10.1161/01.res.85.5.403. [DOI] [PubMed] [Google Scholar]
  • 17.Bongartz T., Sutton A.J., Sweeting M.J., Buchan I., Matteson E.L., Montori V. Anti-TNF antibody therapy in rheumatoid arthritis and the risk of serious infections and malignancies: systematic review and meta-analysis of rare harmful effects in randomized controlled trials. JAMA. (2006);295:2275–2285. doi: 10.1001/jama.295.19.2275. [DOI] [PubMed] [Google Scholar]
  • 18.Borregaard N., Heiple J.M., Simons E.R., Clark R.A. Subcellular localization of the b-cytochrome component of the human neutrophil microbicidal oxidase: translocation during activation. J. Cell Biol. (1983);97:52–61. doi: 10.1083/jcb.97.1.52. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 19.Brand M.D. The sites and topology of mitochondrial superoxide production. Exp. Gerontol. (2010);45:466–472. doi: 10.1016/j.exger.2010.01.003. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 20.Brar S.S., Corbin Z., Kennedy T.P., Hemendinger R., Thornton L., Bommarius B., Arnold R.S., Whorton A.R., Sturrock A.B., Huecksteadt T.P., et al. NOX5 NAD(P)H oxidase regulates growth and apoptosis in DU 145 prostate cancer cells. Am. J. Physiol. Cell Physiol. (2003);285:C353–369. doi: 10.1152/ajpcell.00525.2002. [DOI] [PubMed] [Google Scholar]
  • 21.Bravo J., Karathanassis D., Pacold C.M., Pacold M.E., Ellson C.D., Anderson K.E., Butler P.J., Lavenir I., Perisic O., Hawkins P.T., et al. The crystal structure of the PX domain from p40(phox) bound to phosphatidylinositol 3-phosphate. Mol. Cell. (2001);8:829–839. doi: 10.1016/s1097-2765(01)00372-0. [DOI] [PubMed] [Google Scholar]
  • 22.Brunelle J.K., Bell E.L., Quesada N.M., Vercauteren K., Tiranti V., Zeviani M., Scarpulla R.C., Chandel N.S. Oxygen sensing requires mitochondrial ROS but not oxidative phosphorylation. Cell Metab. (2005);1:409–414. doi: 10.1016/j.cmet.2005.05.002. [DOI] [PubMed] [Google Scholar]
  • 23.Burdon R.H., Rice-Evans C. Free radicals and the regulation of mammalian cell proliferation. Free Radic. Res. Commun. (1989);6:345–358. doi: 10.3109/10715768909087918. [DOI] [PubMed] [Google Scholar]
  • 24.Cano C.E., Gommeaux J., Pietri S., Culcasi M., Garcia S., Seux M., Barelier S., Vasseur S., Spoto R.P., Pebusque M.J., et al. Tumor protein 53-induced nuclear protein 1 is a major mediator of p53 antioxidant function. Cancer Res. (2009);69:219–226. doi: 10.1158/0008-5472.CAN-08-2320. [DOI] [PubMed] [Google Scholar]
  • 25.Caporossi D., Ciafre S.A., Pittaluga M., Savini I., Farace M.G. Cellular responses to H(2)O(2) and bleomycininduced oxidative stress in L6C5 rat myoblasts. Free Radic. Biol. Med. (2003);35:1355–1364. doi: 10.1016/j.freeradbiomed.2003.08.008. [DOI] [PubMed] [Google Scholar]
  • 26.Chance B., Sies H., Boveris A. Hydroperoxide metabolism in mammalian organs. Physiol. Rev. (1979);59:527–605. doi: 10.1152/physrev.1979.59.3.527. [DOI] [PubMed] [Google Scholar]
  • 27.Chandel N.S., Maltepe E., Goldwasser E., Mathieu C.E., Simon M.C., Schumacker P.T. Mitochondrial reactive oxygen species trigger hypoxia-induced transcription. Proc. Natl. Acad. Sci. USA. (1998);95:11715–11720. doi: 10.1073/pnas.95.20.11715. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 28.Chandel N.S., McClintock D.S., Feliciano C.E., Wood T.M., Melendez J.A., Rodriguez A.M., Schumacker P.T. Reactive oxygen species generated at mitochondrial complex III stabilize hypoxia-inducible factor-1alpha during hypoxia: a mechanism of O2 sensing. J. Biol. Chem. (2000);275:25130–25138. doi: 10.1074/jbc.M001914200. [DOI] [PubMed] [Google Scholar]
  • 29.Chang T.S., Cho C.S., Park S., Yu S., Kang S.W., Rhee S.G. Peroxiredoxin III, a mitochondrion-specific peroxidase, regulates apoptotic signaling by mitochondria. J. Biol. Chem. (2004);279:41975–41984. doi: 10.1074/jbc.M407707200. [DOI] [PubMed] [Google Scholar]
  • 30.Charles I., Khalyfa A., Kumar D.M., Krishnamoorthy R.R., Roque R.S., Cooper N., Agarwal N. Serum deprivation induces apoptotic cell death of transformed rat retinal ganglion cells via mitochondrial signaling pathways. Invest. Ophthalmol. Vis. Sci. (2005);46:1330–1338. doi: 10.1167/iovs.04-0363. [DOI] [PubMed] [Google Scholar]
  • 31.Chen G., Goeddel D.V. TNF-R1 signaling: a beautiful pathway. Science. (2002);296:1634–1635. doi: 10.1126/science.1071924. [DOI] [PubMed] [Google Scholar]
  • 32.Chen Q., Powell D.W., Rane M.J., Singh S., Butt W., Klein J.B., McLeish K.R. Akt phosphorylates p47phox and mediates respiratory burst activity in human neutrophils. J. Immunol. (2003a);170:5302–5308. doi: 10.4049/jimmunol.170.10.5302. [DOI] [PubMed] [Google Scholar]
  • 33.Chen Q., Vazquez E.J., Moghaddas S., Hoppel C.L., Lesnefsky E.J. Production of reactive oxygen species by mitochondria: central role of complex III. J. Biol. Chem. (2003b);278:36027–36031. doi: 10.1074/jbc.M304854200. [DOI] [PubMed] [Google Scholar]
  • 34.Cheng G., Diebold B.A., Hughes Y., Lambeth J.D. Nox1-dependent reactive oxygen generation is regulated by Rac1. J. Biol. Chem. (2006);281:17718–17726. doi: 10.1074/jbc.M512751200. [DOI] [PubMed] [Google Scholar]
  • 35.Cheng G., Lambeth J.D. NOXO1, regulation of lipid binding, localization, and activation of Nox1 by the Phox homology (PX) domain. J. Biol. Chem. (2004);279:4737–4742. doi: 10.1074/jbc.M305968200. [DOI] [PubMed] [Google Scholar]
  • 36.Cheng G., Ritsick D., Lambeth J.D. Nox3 regulation by NOXO1, p47phox, and p67phox. J. Biol. Chem. (2004);279:34250–34255. doi: 10.1074/jbc.M400660200. [DOI] [PubMed] [Google Scholar]
  • 37.Chipuk J.E., Kuwana T., Bouchier-Hayes L., Droin N.M., Newmeyer D.D., Schuler M., Green D.R. Direct activation of Bax by p53 mediates mitochondrial membrane permeabilization and apoptosis. Science. (2004);303:1010–1014. doi: 10.1126/science.1092734. [DOI] [PubMed] [Google Scholar]
  • 38.Chipuk J.E., Bouchier-Hayes L., Kuwana T., Newmeyer D.D., Green D.R. PUMA couples the nuclear and cytoplasmic proapoptotic function of p53. Science. (2005);309:1732–1735. doi: 10.1126/science.1114297. [DOI] [PubMed] [Google Scholar]
  • 39.Choi H., Leto T.L., Hunyady L., Catt K.J., Bae Y.S., Rhee S.G. Mechanism of angiotensin II-induced superoxide production in cells reconstituted with angiotensin type 1 receptor and the components of NADPH oxidase. J. Biol. Chem. (2008);283:255–267. doi: 10.1074/jbc.M708000200. [DOI] [PubMed] [Google Scholar]
  • 40.Chung Y.M., Kim J.S., Yoo Y.D. A novel protein, Romo1, induces ROS production in the mitochondria. Biochem. Biophys. Res. Commun. (2006);347:649–655. doi: 10.1016/j.bbrc.2006.06.140. [DOI] [PubMed] [Google Scholar]
  • 41.Chung Y.M., Lee S.B., Kim H.J., Park S.H., Kim J.J., Chung J.S., Yoo Y.D. Replicative senescence induced by Romo1-derived reactive oxygen species. J. Biol. Chem. (2008);283:33763–33771. doi: 10.1074/jbc.M805334200. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 42.Chung J.S., Lee S.B., Park S.H., Kang S.T., Na A.R., Chang T.S., Kim H.J., Yoo Y.D. Mitochondrial reactive oxygen species originating from Romo1 exert an important role in normal cell cycle progression by regulating p27(Kip1) expression. Free Radic. Res. (2009);43:729–737. doi: 10.1080/10715760903038432. [DOI] [PubMed] [Google Scholar]
  • 43.Clark R.A., Leidal K.G., Pearson D.W., Nauseef W.M. NADPH oxidase of human neutrophils. Subcellular localization and characterization of an arachidonate-activatable superoxidegenerating system. J. Biol. Chem. (1987);262:4065–4074. [PubMed] [Google Scholar]
  • 44.Coon M.J. Cytochrome P450: nature’s most versatile biological catalyst. Annu. Rev. Pharmacol. Toxicol. (2005);45:1–25. doi: 10.1146/annurev.pharmtox.45.120403.100030. [DOI] [PubMed] [Google Scholar]
  • 45.Corda S., Laplace C., Vicaut E., Duranteau J. Rapid reactive oxygen species production by mitochondria in endothelial cells exposed to tumor necrosis factor-alpha is mediated by ceramide. Am. J. Respir. Cell Mol. Biol. (2001);24:762–768. doi: 10.1165/ajrcmb.24.6.4228. [DOI] [PubMed] [Google Scholar]
  • 46.Danielson P.B. The cytochrome P450 superfamily: biochemistry, evolution and drug metabolism in humans. Curr. Drug Metab. (2002);3:561–597. doi: 10.2174/1389200023337054. [DOI] [PubMed] [Google Scholar]
  • 47.De Deken X., Wang D., Many M.C., Costagliola S., Libert F., Vassart G., Dumont J.E., Miot F. Cloning of two human thyroid cDNAs encoding new members of the NADPH oxidase family. J. Biol. Chem. (2000);275:23227–23233. doi: 10.1074/jbc.M000916200. [DOI] [PubMed] [Google Scholar]
  • 48.de Gasparo M., Catt K.J., Inagami T., Wright J.W., Unger T. International union of pharmacology. XXIII. The angiotensin II receptors. Pharmacol. Rev. (2000);52:415–472. [PubMed] [Google Scholar]
  • 49.de Mendez I., Homayounpour N., Leto T.L. Specificity of p47phox SH3 domain interactions in NADPH oxidase assembly and activation. Mol. Cell. Biol. (1997);17:2177–2185. doi: 10.1128/mcb.17.4.2177. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 50.De Smaele E., Zazzeroni F., Papa S., Nguyen D.U., Jin R., Jones J., Cong R., Franzoso G. Induction of gadd45beta by NF-kappaB downregulates pro-apoptotic JNK signalling. Nature. (2001);414:308–313. doi: 10.1038/35104560. [DOI] [PubMed] [Google Scholar]
  • 51.Degtyarenko K.N., Kulikova T.A. Evolution of bioinorganic motifs in P450-containing systems. Biochem. Soc. Trans. (2001);29:139–147. doi: 10.1042/0300-5127:0290139. [DOI] [PubMed] [Google Scholar]
  • 52.DeLeo F.R., Yu L., Burritt J.B., Loetterle L.R., Bond C.W., Jesaitis A.J., Quinn M.T. Mapping sites of interaction of p47-phox and flavocytochrome b with random-sequence peptide phage display libraries. Proc. Natl. Acad. Sci. USA. (1995);92:7110–7114. doi: 10.1073/pnas.92.15.7110. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 53.Deng X., Gao F., May W.S., Jr. Bcl2 retards G1/S cell cycle transition by regulating intracellular ROS. Blood. (2003);102:3179–3185. doi: 10.1182/blood-2003-04-1027. [DOI] [PubMed] [Google Scholar]
  • 54.Desaint S., Luriau S., Aude J.C., Rousselet G., Toledano M.B. Mammalian antioxidant defenses are not inducible by H2O2. J. Biol. Chem. (2004);279:31157–31163. doi: 10.1074/jbc.M401888200. [DOI] [PubMed] [Google Scholar]
  • 55.Devin A., Cook A., Lin Y., Rodriguez Y., Kelliher M., Liu Z. The distinct roles of TRAF2 and RIP in IKK activation by TNF-R1: TRAF2 recruits IKK to TNF-R1 while RIP mediates IKK activation. Immunity. (2000);12:419–429. doi: 10.1016/s1074-7613(00)80194-6. [DOI] [PubMed] [Google Scholar]
  • 56.Dinauer M.C., Pierce E.A., Bruns G.A., Curnutte J.T., Orkin S.H. Human neutrophil cytochrome b light chain (p22- phox). Gene structure, chromosomal location, and mutations in cytochrome-negative autosomal recessive chronic granulomatous disease. J. Clin. Invest. (1990);86:1729–1737. doi: 10.1172/JCI114898. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 57.Droge W. Free radicals in the physiological control of cell function. Physiol. Rev. (2002);82:47–95. doi: 10.1152/physrev.00018.2001. [DOI] [PubMed] [Google Scholar]
  • 58.Dumont P., Leu J.I., Della Pietra A.C., 3rd, George D.L., Murphy M. The codon 72 polymorphic variants of p53 have markedly different apoptotic potential. Nat. Genet. (2003);33:357–365. doi: 10.1038/ng1093. [DOI] [PubMed] [Google Scholar]
  • 59.Dupuy C., Virion A., Ohayon R., Kaniewski J., Deme D., Pommier J. Mechanism of hydrogen peroxide formation catalyzed by NADPH oxidase in thyroid plasma membrane. J. Biol. Chem. (1991);266:3739–3743. [PubMed] [Google Scholar]
  • 60.Dupuy C., Virion A., De Sandro V., Ohayon R., Kaniewski J., Pommier J., Deme D. Activation of the NADPHdependent H2O2-generating system in pig thyroid particulate fraction by limited proteolysis and Zn2+ treatment. Biochem. J. (1992);283:591–595. doi: 10.1042/bj2830591. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 61.Dupuy C., Ohayon R., Valent A., Noel-Hudson M.S., Deme D., Virion A. Purification of a novel flavoprotein involved in the thyroid NADPH oxidase. Cloning of the porcine and human cdnas. J. Biol. Chem. (1999);274:37265–37269. doi: 10.1074/jbc.274.52.37265. [DOI] [PubMed] [Google Scholar]
  • 62.Dupuy C., Pomerance M., Ohayon R., Noel-Hudson M.S., Deme D., Chaaraoui M., Francon J., Virion A. Thyroid oxidase (THOX2) gene expression in the rat thyroid cell line FRTL-5. Biochem. Biophys. Res. Commun. (2000);277:287–292. doi: 10.1006/bbrc.2000.3671. [DOI] [PubMed] [Google Scholar]
  • 63.El Benna J., Faust L.P., Babior B.M. The phosphorylation of the respiratory burst oxidase component p47phox during neutrophil activation. Phosphorylation of sites recognized by protein kinase C and by proline-directed kinases. J. Biol. Chem. (1994);269:23431–23436. [PubMed] [Google Scholar]
  • 64.El Benna J., Faust R.P., Johnson J.L., Babior B.M. Phosphorylation of the respiratory burst oxidase subunit p47- phox as determined by two-dimensional phosphopeptide mapping. Phosphorylation by protein kinase C, protein kinase A, and a mitogen-activated protein kinase. J. Biol. Chem. (1996);271:6374–6378. doi: 10.1074/jbc.271.11.6374. [DOI] [PubMed] [Google Scholar]
  • 65.El Hassani R.A., Benfares N., Caillou B., Talbot M., Sabourin J.C., Belotte V., Morand S., Gnidehou S., Agnandji D., Ohayon R., et al. Dual oxidase2 is expressed all along the digestive tract. Am. J. Physiol. Gastrointest Liver Physiol. (2005);288:G933–942. doi: 10.1152/ajpgi.00198.2004. [DOI] [PubMed] [Google Scholar]
  • 66.Ellerby L.M., Ellerby H.M., Park S.M., Holleran A.L., Murphy A.N., Fiskum G., Kane D.J., Testa M.P., Kayalar C., Bredesen D.E. Shift of the cellular oxidation-reduction potential in neural cells expressing Bcl-2. J. Neurochem. (1996);67:1259–1267. doi: 10.1046/j.1471-4159.1996.67031259.x. [DOI] [PubMed] [Google Scholar]
  • 67.Ferrari G., Yan C.Y., Greene L.A. N-acetylcysteine (D- and L-stereoisomers) prevents apoptotic death of neuronal cells. J. Neurosci. (1995);15:2857–2866. doi: 10.1523/JNEUROSCI.15-04-02857.1995. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 68.Finegold A.A., Shatwell K.P., Segal A.W., Klausner R.D., Dancis A. Intramembrane bis-heme motif for transmembrane electron transport conserved in a yeast iron reductase and the human NADPH oxidase. J. Biol. Chem. (1996);271:31021–31024. doi: 10.1074/jbc.271.49.31021. [DOI] [PubMed] [Google Scholar]
  • 69.Finkel T. Oxygen radicals and signaling. Curr. Opin. Cell Biol. (1998);10:248–253. doi: 10.1016/s0955-0674(98)80147-6. [DOI] [PubMed] [Google Scholar]
  • 70.Fontayne A., Dang P.M., Gougerot-Pocidalo M.A., El-Benna J. Phosphorylation of p47phox sites by PKC alpha, beta II, delta, and zeta: effect on binding to p22phox and on NADPH oxidase activation. Biochemistry. (2002);41:7743–7750. doi: 10.1021/bi011953s. [DOI] [PubMed] [Google Scholar]
  • 71.Forteza R., Salathe M., Miot F., Conner G.E. Regulated hydrogen peroxide production by Duox in human airway epithelial cells. Am. J. Respir. Cell Mol. Biol. (2005);32:462–469. doi: 10.1165/rcmb.2004-0302OC. [DOI] [PubMed] [Google Scholar]
  • 72.Frost J.A., Geppert T.D., Cobb M.H., Feramisco J.R. A requirement for extracellular signal-regulated kinase (ERK) function in the activation of AP-1 by Ha-Ras, phorbol 12- myristate 13-acetate, and serum. Proc. Natl. Acad. Sci. USA. (1994);91:3844–3848. doi: 10.1073/pnas.91.9.3844. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 73.Galluzzi L., Morselli E., Kepp O., Vitale I., Rigoni A., Vacchelli E., Michaud M., Zischka H., Castedo M., Kroemer G. Mitochondrial gateways to cancer. Mol. Aspects Med. (2010);31:1–20. doi: 10.1016/j.mam.2009.08.002. [DOI] [PubMed] [Google Scholar]
  • 74.Geiszt M., Lekstrom K., Witta J., Leto T.L. Proteins homologous to p47phox and p67phox support superoxide production by NAD(P)H oxidase 1 in colon epithelial cells. J. Biol. Chem. (2003a);278:20006–20012. doi: 10.1074/jbc.M301289200. [DOI] [PubMed] [Google Scholar]
  • 75.Geiszt M., Witta J., Baffi J., Lekstrom K., Leto T.L. Dual oxidases represent novel hydrogen peroxide sources supporting mucosal surface host defense. FASEB J. (2003b);17:1502–1504. doi: 10.1096/fj.02-1104fje. [DOI] [PubMed] [Google Scholar]
  • 76.Gertz M., Fischer F., Wolters D., Steegborn C. Activation of the lifespan regulator p66Shc through reversible disulfide bond formation. Proc. Natl. Acad. Sci. USA. (2008);105:5705–5709. doi: 10.1073/pnas.0800691105. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 77.Giaccia A.J., Kastan M.B. The complexity of p53 modulation: emerging patterns from divergent signals. Genes Dev. (1998);12:2973–2983. doi: 10.1101/gad.12.19.2973. [DOI] [PubMed] [Google Scholar]
  • 78.Gianni D., Taulet N., DerMardirossian C., Bokoch G.M. c-Src-mediated phosphorylation of NoxA1 and Tks4 induces the reactive oxygen species (ROS)-dependent formation of functional invadopodia in human colon cancer cells. Mol. Biol. Cell. (2010);21:4287–4298. doi: 10.1091/mbc.E10-08-0685. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 79.Gianni D., DerMardirossian C., Bokoch G.M. Direct interaction between Tks proteins and the N-terminal proline-rich region (PRR) of NoxA1 mediates Nox1-dependent ROS generation. Eur. J. Cell Biol. (2011);90:164–171. doi: 10.1016/j.ejcb.2010.05.007. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 80.Giorgio M., Migliaccio E., Orsini F., Paolucci D., Moroni M., Contursi C., Pelliccia G., Luzi L., Minucci S., Marcaccio M., et al. Electron transfer between cytochrome c and p66Shc generates reactive oxygen species that trigger mitochondrial apoptosis. Cell. (2005);122:221–233. doi: 10.1016/j.cell.2005.05.011. [DOI] [PubMed] [Google Scholar]
  • 81.Gong P., Cederbaum A.I. Nrf2 is increased by CYP2E1 in rodent liver and HepG2 cells and protects against oxidative stress caused by CYP2E1. Hepatology. (2006);43:144–153. doi: 10.1002/hep.21004. [DOI] [PubMed] [Google Scholar]
  • 82.Gorsky L.D., Koop D.R., Coon M.J. On the stoichiometry of the oxidase and monooxygenase reactions catalyzed by liver microsomal cytochrome P-450. Products of oxygen reduction. J. Biol. Chem. (1984);259:6812–6817. [PubMed] [Google Scholar]
  • 83.Gottlieb E., Vander Heiden M.G., Thompson C.B. Bclx(L) prevents the initial decrease in mitochondrial membrane potential and subsequent reactive oxygen species production during tumor necrosis factor alpha-induced apoptosis. Mol. Cell. Biol. (2000);20:5680–5689. doi: 10.1128/mcb.20.15.5680-5689.2000. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 84.Grasberger H., Refetoff S. Identification of the maturation factor for dual oxidase. Evolution of an eukaryotic operon equivalent. J. Biol. Chem. (2006);281:18269–18272. doi: 10.1074/jbc.C600095200. [DOI] [PubMed] [Google Scholar]
  • 85.Green D.R., Kroemer G. The pathophysiology of mitochondrial cell death. Science. (2004);305:626–629. doi: 10.1126/science.1099320. [DOI] [PubMed] [Google Scholar]
  • 86.Greene L.A. Nerve growth factor prevents the death and stimulates the neuronal differentiation of clonal PC12 pheochromocytoma cells in serum-free medium. J. Cell Biol. (1978);78:747–755. doi: 10.1083/jcb.78.3.747. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 87.Groemping Y., Lapouge K., Smerdon S.J., Rittinger K. Molecular basis of phosphorylation-induced activation of the NADPH oxidase. Cell. (2003);113:343–355. doi: 10.1016/s0092-8674(03)00314-3. [DOI] [PubMed] [Google Scholar]
  • 88.Guzy R.D., Hoyos B., Robin E., Chen H., Liu L., Mansfield K.D., Simon M.C., Hammerling U., Schumacker P.T. Mitochondrial complex III is required for hypoxia-induced ROS production and cellular oxygen sensing. Cell Metab. (2005);1:401–408. doi: 10.1016/j.cmet.2005.05.001. [DOI] [PubMed] [Google Scholar]
  • 89.Hamanaka R.B., Chandel N.S. Mitochondrial reactive oxygen species regulate cellular signaling and dictate biological outcomes. Trends Biochem. Sci. (2010);35:505–513. doi: 10.1016/j.tibs.2010.04.002. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 90.Han C.H., Freeman J.L., Lee T., Motalebi S.A., Lambeth J.D. Regulation of the neutrophil respiratory burst oxidase. Identification of an activation domain in p67(phox). J. Biol. Chem. (1998);273:16663–16668. doi: 10.1074/jbc.273.27.16663. [DOI] [PubMed] [Google Scholar]
  • 91.Han D., Ybanez M.D., Ahmadi S., Yeh K., Kaplowitz N. Redox regulation of tumor necrosis factor signaling. Antioxid Redox Signal. (2009);11:2245–2263. doi: 10.1089/ars.2009.2611. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 92.Hansford R.G., Hogue B.A., Mildaziene V. Dependence of H2O2 formation by rat heart mitochondria on substrate availability and donor age. J. Bioenerg. Biomembr. (1997);29:89–95. doi: 10.1023/a:1022420007908. [DOI] [PubMed] [Google Scholar]
  • 93.Hanukoglu I. Antioxidant protective mechanisms against reactive oxygen species (ROS) generated by mitochondrial P450 systems in steroidogenic cells. Drug Metab. Rev. (2006);38:171–196. doi: 10.1080/03602530600570040. [DOI] [PubMed] [Google Scholar]
  • 94.Hanukoglu I., Suh B.S., Himmelhoch S., Amsterdam A. Induction and mitochondrial localization of cytochrome P450scc system enzymes in normal and transformed ovarian granulosa cells. J. Cell Biol. (1990);111:1373–1381. doi: 10.1083/jcb.111.4.1373. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 95.Hata K., Takeshige K., Sumimoto H. Roles for prolinerich regions of p47phox and p67phox in the phagocyte NADPH oxidase activation in vitro. Biochem. Biophys. Res. Commun. (1997);241:226–231. doi: 10.1006/bbrc.1997.7807. [DOI] [PubMed] [Google Scholar]
  • 96.Henderson L.M., Chappell J.B., Jones O.T. The superoxide- generating NADPH oxidase of human neutrophils is electrogenic and associated with an H+ channel. Biochem. J. (1987);246:325–329. doi: 10.1042/bj2460325. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 97.Hockenbery D.M., Oltvai Z.N., Yin X.M., Milliman C.L., Korsmeyer S.J. Bcl-2 functions in an antioxidant pathway to prevent apoptosis. Cell. (1993);75:241–251. doi: 10.1016/0092-8674(93)80066-n. [DOI] [PubMed] [Google Scholar]
  • 98.Hsu H., Xiong J., Goeddel D.V. The TNF receptor 1- associated protein TRADD signals cell death and NF-kappa B activation. Cell. (1995);81:495–504. doi: 10.1016/0092-8674(95)90070-5. [DOI] [PubMed] [Google Scholar]
  • 99.Hsu H., Shu H.B., Pan M.G., Goeddel D.V. TRADDTRAF2 and TRADD-FADD interactions define two distinct TNF receptor 1 signal transduction pathways. Cell. (1996);84:299–308. doi: 10.1016/s0092-8674(00)80984-8. [DOI] [PubMed] [Google Scholar]
  • 100.Huang J., Hitt N.D., Kleinberg M.E. Stoichiometry of p22-phox and gp91-phox in phagocyte cytochrome b558. Biochemistry. (1995);34:16753–16757. doi: 10.1021/bi00051a024. [DOI] [PubMed] [Google Scholar]
  • 101.Huang P., Feng L., Oldham E.A., Keating M.J., Plunkett W. Superoxide dismutase as a target for the selective killing of cancer cells. Nature. (2000);407:390–395. doi: 10.1038/35030140. [DOI] [PubMed] [Google Scholar]
  • 102.Hwang P.M., Bunz F., Yu J., Rago C., Chan T.A., Murphy M.P., Kelso G.F., Smith R.A., Kinzler K.W., Vogelstein B. Ferredoxin reductase affects p53-dependent, 5-fluorouracil- induced apoptosis in colorectal cancer cells. Nat. Med. (2001);7:1111–1117. doi: 10.1038/nm1001-1111. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 103.Hwang I.T., Chung Y.M., Kim J.J., Chung J.S., Kim B.S., Kim H.J., Kim J.S., Yoo Y.D. Drug resistance to 5-FU linked to reactive oxygen species modulator 1. Biochem. Biophys. Res. Commun. (2007);359:304–310. doi: 10.1016/j.bbrc.2007.05.088. [DOI] [PubMed] [Google Scholar]
  • 104.Ichijo H., Nishida E., Irie K., ten Dijke P., Saitoh M., Moriguchi T., Takagi M., Matsumoto K., Miyazono K., Gotoh Y. Induction of apoptosis by ASK1, a mammalian MAPKKK that activates SAPK/JNK and p38 signaling pathways. Science. (1997);275:90–94. doi: 10.1126/science.275.5296.90. [DOI] [PubMed] [Google Scholar]
  • 105.Iwata S., Lee J.W., Okada K., Lee J.K., Iwata M., Rasmussen B., Link T.A., Ramaswamy S., Jap B.K. Complete structure of the 11-subunit bovine mitochondrial cytochrome bc1 complex. Science. (1998);281:64–71. doi: 10.1126/science.281.5373.64. [DOI] [PubMed] [Google Scholar]
  • 106.Jagnandan D., Church J.E., Banfi B., Stuehr D.J., Marrero M.B., Fulton D.J. Novel mechanism of activation of NADPH oxidase 5. calcium sensitization via phosphorylation. J. Biol. Chem. (2007);282:6494–6507. doi: 10.1074/jbc.M608966200. [DOI] [PubMed] [Google Scholar]
  • 107.Jiang Q., Griffin D.A., Barofsky D.F., Hurst J.K. Intraphagosomal chlorination dynamics and yields determined using unique fluorescent bacterial mimics. Chem. Res. Toxicol. (1997);10:1080–1089. doi: 10.1021/tx9700984. [DOI] [PubMed] [Google Scholar]
  • 108.Kanai F., Liu H., Field S.J., Akbary H., Matsuo T., Brown G.E., Cantley L.C., Yaffe M.B. The PX domains of p47phox and p40phox bind to lipid products of PI(3)K. Nat. Cell Biol. (2001);3:675–678. doi: 10.1038/35083070. [DOI] [PubMed] [Google Scholar]
  • 109.Kane D.J., Sarafian T.A., Anton R., Hahn H., Gralla E.B., Valentine J.S., Ord T., Bredesen D.E. Bcl-2 inhibition of neural death: decreased generation of reactive oxygen species. Science. (1993);262:1274–1277. doi: 10.1126/science.8235659. [DOI] [PubMed] [Google Scholar]
  • 110.Karathanassis D., Stahelin R.V., Bravo J., Perisic O., Pacold C.M., Cho W., Williams R.L. Binding of the PX domain of p47(phox) to phosphatidylinositol 3,4-bisphosphate and phosphatidic acid is masked by an intramolecular interaction. EMBO J. (2002);21:5057–5068. doi: 10.1093/emboj/cdf519. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 111.Karin M., Lin A. NF-kappaB at the crossroads of life and death. Nat. Immunol. (2002);3:221–227. doi: 10.1038/ni0302-221. [DOI] [PubMed] [Google Scholar]
  • 112.Kawahara T., Ritsick D., Cheng G., Lambeth J.D. Point mutations in the proline-rich region of p22phox are dominant inhibitors of Nox1- and Nox2-dependent reactive oxygen generation. J. Biol. Chem. (2005);280:31859–31869. doi: 10.1074/jbc.M501882200. [DOI] [PubMed] [Google Scholar]
  • 113.Kawai T., Akira S. The role of pattern-recognition receptors in innate immunity: update on Toll-like receptors. Nat. Immunol. (2010);11:373–384. doi: 10.1038/ni.1863. [DOI] [PubMed] [Google Scholar]
  • 114.Kim R., Emi M., Tanabe K., Murakami S., Uchida Y., Arihiro K. Regulation and interplay of apoptotic and nonapoptotic cell death. J. Pathol. (2006);208:319–326. doi: 10.1002/path.1885. [DOI] [PubMed] [Google Scholar]
  • 115.Kim J.S., Diebold B.A., Babior B.M., Knaus U.G., Bokoch G.M. Regulation of Nox1 activity via protein kinase Amediated phosphorylation of NoxA1 and 14-3-3 binding. J. Biol. Chem. (2007a);282:34787–34800. doi: 10.1074/jbc.M704754200. [DOI] [PubMed] [Google Scholar]
  • 116.Kim Y.S., Morgan M.J., Choksi S., Liu Z.G. TNFinduced activation of the Nox1 NADPH oxidase and its role in the induction of necrotic cell death. Mol. Cell. (2007b);26:675–687. doi: 10.1016/j.molcel.2007.04.021. [DOI] [PubMed] [Google Scholar]
  • 117.Kim J.J., Lee S.B., Park J.K., Yoo Y.D. TNF-alpha-induced ROS production triggering apoptosis is directly linked to Romo1 and Bcl-X(L). Cell Death Differ. (2010);17:1420–1434. doi: 10.1038/cdd.2010.19. [DOI] [PubMed] [Google Scholar]
  • 118.King A.R., Francis S.E., Bridgeman C.J., Bird H., Whyte M.K., Crossman D.C. A role for caspase-1 in serum withdrawal- induced apoptosis of endothelial cells. Lab. Invest. (2003);83:1497–1508. doi: 10.1097/01.lab.0000093096.62765.85. [DOI] [PubMed] [Google Scholar]
  • 119.Kirkland R.A., Franklin J.L. Evidence for redox regulation of cytochrome C release during programmed neuronal death: antioxidant effects of protein synthesis and caspase inhibition. J. Neurosci. (2001);21:1949–1963. doi: 10.1523/JNEUROSCI.21-06-01949.2001. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 120.Kirkland R.A., Windelborn J.A., Kasprzak J.M., Franklin J.L. A Bax-induced pro-oxidant state is critical for cytochrome c release during programmed neuronal death. J. Neurosci. (2002);22:6480–6490. doi: 10.1523/JNEUROSCI.22-15-06480.2002. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 121.Kiss P.J., Knisz J., Zhang Y., Baltrusaitis J., Sigmund C.D., Thalmann R., Smith R.J., Verpy E., Banfi B. Inactivation of NADPH oxidase organizer 1 results in severe imbalance. Curr. Biol. (2006);16:208–213. doi: 10.1016/j.cub.2005.12.025. [DOI] [PubMed] [Google Scholar]
  • 122.Knaus U.G., Heyworth P.G., Evans T., Curnutte J.T., Bokoch G.M. Regulation of phagocyte oxygen radical production by the GTP-binding protein Rac 2. Science. (1991);254:1512–1515. doi: 10.1126/science.1660188. [DOI] [PubMed] [Google Scholar]
  • 123.Koga H., Terasawa H., Nunoi H., Takeshige K., Inagaki F., Sumimoto H. Tetratricopeptide repeat (TPR) motifs of p67(phox) participate in interaction with the small GTPase Rac and activation of the phagocyte NADPH oxidase. J. Biol. Chem. (1999);274:25051–25060. doi: 10.1074/jbc.274.35.25051. [DOI] [PubMed] [Google Scholar]
  • 124.Kong Q., Beel J.A., Lillehei K.O. A threshold concept for cancer therapy. Med. Hypotheses. (2000);55:29–35. doi: 10.1054/mehy.1999.0982. [DOI] [PubMed] [Google Scholar]
  • 125.Kowaltowski A.J., de Souza-Pinto N.C., Castilho R.F., Vercesi A.E. Mitochondria and reactive oxygen species. Free Radic. Biol. Med. (2009);47:333–343. doi: 10.1016/j.freeradbiomed.2009.05.004. [DOI] [PubMed] [Google Scholar]
  • 126.Kroviarski Y., Debbabi M., Bachoual R., Perianin A., Gougerot- Pocidalo M.A., El-Benna J., Dang P.M. Phosphorylation of NADPH oxidase activator 1 (NOXA1) on serine 282 by MAP kinases and on serine 172 by protein kinase C and protein kinase A prevents NOX1 hyperactivation. FASEB J. (2010);24:2077–2092. doi: 10.1096/fj.09-147629. [DOI] [PubMed] [Google Scholar]
  • 127.Ksenzenko M., Konstantinov A.A., Khomutov G.B., Tikhonov A.N., Ruuge E.K. Effect of electron transfer inhibitors on superoxide generation in the cytochrome bc1 site of the mitochondrial respiratory chain. FEBS Lett. (1983);155:19–24. doi: 10.1016/0014-5793(83)80200-2. [DOI] [PubMed] [Google Scholar]
  • 128.Kudin A.P., Bimpong-Buta N.Y., Vielhaber S., Elger C.E., Kunz W.S. Characterization of superoxide-producing sites in isolated brain mitochondria. J. Biol. Chem. (2004);279:4127–4135. doi: 10.1074/jbc.M310341200. [DOI] [PubMed] [Google Scholar]
  • 129.Kuribayashi F., Nunoi H., Wakamatsu K., Tsunawaki S., Sato K., Ito T., Sumimoto H. The adaptor protein p40(phox) as a positive regulator of the superoxide-producing phagocyte oxidase. EMBO J. (2002);21:6312–6320. doi: 10.1093/emboj/cdf642. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 130.Kushnareva Y., Murphy A.N., Andreyev A. Complex Imediated reactive oxygen species generation: modulation by cytochrome c and NAD(P)+ oxidation-reduction state. Biochem. J. (2002);368:545–553. doi: 10.1042/BJ20021121. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 131.Kutter D., Devaquet P., Vanderstocken G., Paulus J.M., Marchal V., Gothot A. Consequences of total and subtotal myeloperoxidase deficiency: risk or benefit? Acta Haematol. (2000);104:10–15. doi: 10.1159/000041062. [DOI] [PubMed] [Google Scholar]
  • 132.Kwong L.K., Sohal R.S. Substrate and site specificity of hydrogen peroxide generation in mouse mitochondria. Arch. Biochem. Biophys. (1998);350:118–126. doi: 10.1006/abbi.1997.0489. [DOI] [PubMed] [Google Scholar]
  • 133.Lamb J.R., Tugendreich S., Hieter P. Tetratrico peptide repeat interactions: to TPR or not to TPR? Trends Biochem. Sci. (1995);20:257–259. doi: 10.1016/s0968-0004(00)89037-4. [DOI] [PubMed] [Google Scholar]
  • 134.Lambert A.J., Brand M.D. Superoxide production by NADH:ubiquinone oxidoreductase (complex I) depends on the pH gradient across the mitochondrial inner membrane. Biochem. J. (2004);382:511–517. doi: 10.1042/BJ20040485. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 135.Lapouge K., Smith S.J., Walker P.A., Gamblin S.J., Smerdon S.J., Rittinger K. Structure of the TPR domain of p67phox in complex with Rac. GTP. Mol. Cell. (2000);6:899–907. doi: 10.1016/s1097-2765(05)00091-2. [DOI] [PubMed] [Google Scholar]
  • 136.Laurent A., Nicco C., Chereau C., Goulvestre C., Alexandre J., Alves A., Levy E., Goldwasser F., Panis Y., Soubrane O., et al. Controlling tumor growth by modulating endogenous production of reactive oxygen species. Cancer Res. (2005);65:948–956. [PubMed] [Google Scholar]
  • 137.Lee S.B., Kim J.J., Kim T.W., Kim B.S., Lee M.S., Yoo Y.D. Serum deprivation-induced reactive oxygen species production is mediated by Romo1. Apoptosis. (2010);15:204–218. doi: 10.1007/s10495-009-0411-1. [DOI] [PubMed] [Google Scholar]
  • 138.Lee S.B., Kim J.J., Chung J.S., Lee M.S., Lee K.H., Kim B.S., Yoo Y.D. Romo1 is a negative-feedback regulator of Myc. J. Cell Sci. (2011);124:1911–1924. doi: 10.1242/jcs.079996. [DOI] [PubMed] [Google Scholar]
  • 139.Leseney A.M., Deme D., Legue O., Ohayon R., Chanson P., Sales J.P., Carvalho D.P., Dupuy C., Virion A. Biochemical characterization of a Ca2+/NAD(P)H-dependent H2O2 generator in human thyroid tissue. Biochimie. (1999);81:373–380. doi: 10.1016/s0300-9084(99)80084-4. [DOI] [PubMed] [Google Scholar]
  • 140.Leto T.L., Adams A.G., de Mendez I. Assembly of the phagocyte NADPH oxidase: binding of Src homology 3 domains to proline-rich targets. Proc. Natl. Acad. Sci. USA. (1994);91:10650–10654. doi: 10.1073/pnas.91.22.10650. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 141.Leu J.I., Dumont P., Hafey M., Murphy M.E., George D.L. Mitochondrial p53 activates Bak and causes disruption of a Bak-Mcl1 complex. Nat. Cell Biol. (2004);6:443–450. doi: 10.1038/ncb1123. [DOI] [PubMed] [Google Scholar]
  • 142.Levine A.J. p53, the cellular gatekeeper for growth and division. Cell. (1997);88:323–331. doi: 10.1016/s0092-8674(00)81871-1. [DOI] [PubMed] [Google Scholar]
  • 143.Liu S.L., Lin X., Shi D.Y., Cheng J., Wu C.Q., Zhang Y.D. Reactive oxygen species stimulated human hepatoma cell proliferation via cross-talk between PI3-K/PKB and JNK signaling pathways. Arch. Biochem. Biophys. (2002a);406:173–182. doi: 10.1016/s0003-9861(02)00430-7. [DOI] [PubMed] [Google Scholar]
  • 144.Liu Y., Fiskum G., Schubert D. Generation of reactive oxygen species by the mitochondrial electron transport chain. J. Neurochem. (2002b);80:780–787. doi: 10.1046/j.0022-3042.2002.00744.x. [DOI] [PubMed] [Google Scholar]
  • 145.Liu Z., Lu H., Shi H., Du Y., Yu J., Gu S., Chen X., Liu K.J., Hu C.A. PUMA overexpression induces reactive oxygen species generation and proteasome-mediated stathmin degradation in colorectal cancer cells. Cancer Res. (2005);65:1647–1654. doi: 10.1158/0008-5472.CAN-04-1754. [DOI] [PubMed] [Google Scholar]
  • 146.Lo Y.Y., Cruz T.F. Involvement of reactive oxygen species in cytokine and growth factor induction of c-fos expression in chondrocytes. J. Biol. Chem. (1995);270:11727–11730. doi: 10.1074/jbc.270.20.11727. [DOI] [PubMed] [Google Scholar]
  • 147.Locksley R.M., Killeen N., Lenardo M.J. The TNF and TNF receptor superfamilies: integrating mammalian biology. Cell. (2001);104:487–501. doi: 10.1016/s0092-8674(01)00237-9. [DOI] [PubMed] [Google Scholar]
  • 148.Loschen G., Azzi A., Flohe L. Mitochondrial H2O2 formation: relationship with energy conservation. FEBS Lett. (1973);33:84–87. doi: 10.1016/0014-5793(73)80165-6. [DOI] [PubMed] [Google Scholar]
  • 149.Luxen S., Belinsky S.A., Knaus U.G. Silencing of DUOX NADPH oxidases by promoter hypermethylation in lung cancer. Cancer Res. (2008);68:1037–1045. doi: 10.1158/0008-5472.CAN-07-5782. [DOI] [PubMed] [Google Scholar]
  • 150.Lyle A.N., Deshpande N.N., Taniyama Y., Seidel-Rogol B., Pounkova L., Du P., Papaharalambus C., Lassegue B., Griendling K.K. Poldip2, a novel regulator of Nox4 and cytoskeletal integrity in vascular smooth muscle cells. Circ. Res. (2009);105:249–259. doi: 10.1161/CIRCRESAHA.109.193722. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 151.Mansfield K.D., Guzy R.D., Pan Y., Young R.M., Cash T.P., Schumacker P.T., Simon M.C. Mitochondrial dysfunction resulting from loss of cytochrome c impairs cellular oxygen sensing and hypoxic HIF-alpha activation. Cell Metab. (2005);1:393–399. doi: 10.1016/j.cmet.2005.05.003. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 152.Marchenko N.D., Wolff S., Erster S., Becker K., Moll U.M. Monoubiquitylation promotes mitochondrial p53 translocation. EMBO J. (2007);26:923–934. doi: 10.1038/sj.emboj.7601560. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 153.Marcillat O., Zhang Y., Davies K.J. Oxidative and non-oxidative mechanisms in the inactivation of cardiac mitochondrial electron transport chain components by doxorubicin. Biochem. J. (1989);259:181–189. doi: 10.1042/bj2590181. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 154.Martindale J.L., Holbrook N.J. Cellular response to oxidative stress: signaling for suicide and survival. J. Cell Physiol. (2002);192:1–15. doi: 10.1002/jcp.10119. [DOI] [PubMed] [Google Scholar]
  • 155.Martyn K.D., Kim M.J., Quinn M.T., Dinauer M.C., Knaus U.G. p21-activated kinase (Pak) regulates NADPH oxidase activation in human neutrophils. Blood. (2005);106:3962–3969. doi: 10.1182/blood-2005-03-0859. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 156.Martyn K.D., Frederick L.M., von Loehneysen K., Dinauer M.C., Knaus U.G. Functional analysis of Nox4 reveals unique characteristics compared to other NADPH oxidases. Cell. Signal. (2006);18:69–82. doi: 10.1016/j.cellsig.2005.03.023. [DOI] [PubMed] [Google Scholar]
  • 157.Matsuzawa A., Ichijo H. Stress-responsive protein kinases in redox-regulated apoptosis signaling. Antioxid Redox Signal. (2005);7:472–481. doi: 10.1089/ars.2005.7.472. [DOI] [PubMed] [Google Scholar]
  • 158.McKinnon R.A., McManus M.E. Localization of cytochromes P450 in human tissues: implications for chemical toxicity. Pathology. (1996);28:148–155. doi: 10.1080/00313029600169783. [DOI] [PubMed] [Google Scholar]
  • 159.Meier B., Radeke H.H., Selle S., Younes M., Sies H., Resch K., Habermehl G.G. Human fibroblasts release reactive oxygen species in response to interleukin-1 or tumour necrosis factor-alpha. Biochem. J. (1989);263:539–545. doi: 10.1042/bj2630539. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 160.Meitzler J.L., Ortiz de Montellano P.R. Caenorhabditis elegans and human dual oxidase 1 (DUOX1) “peroxidase” domains: insights into heme binding and catalytic activity. J. Biol. Chem. (2009);284:18634–18643. doi: 10.1074/jbc.M109.013581. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 161.Micheau O., Tschopp J. Induction of TNF receptor Imediated apoptosis via two sequential signaling complexes. Cell. (2003);114:181–190. doi: 10.1016/s0092-8674(03)00521-x. [DOI] [PubMed] [Google Scholar]
  • 162.Migliaccio E., Giorgio M., Mele S., Pelicci G., Reboldi P., Pandolfi P.P., Lanfrancone L., Pelicci P.G. The p66shc adaptor protein controls oxidative stress response and life span in mammals. Nature. (1999);402:309–313. doi: 10.1038/46311. [DOI] [PubMed] [Google Scholar]
  • 163.Mihara M., Erster S., Zaika A., Petrenko O., Chittenden T., Pancoska P., Moll U.M. p53 has a direct apoptogenic role at the mitochondria. Mol. Cell. (2003);11:577–590. doi: 10.1016/s1097-2765(03)00050-9. [DOI] [PubMed] [Google Scholar]
  • 164.Miller Y.I., Viriyakosol S., Binder C.J., Feramisco J.R., Kirkland T.N., Witztum J.L. Minimally modified LDL binds to CD14, induces macrophage spreading via TLR4/MD-2, and inhibits phagocytosis of apoptotic cells. J. Biol. Chem. (2003);278:1561–1568. doi: 10.1074/jbc.M209634200. [DOI] [PubMed] [Google Scholar]
  • 165.Miller Y.I., Viriyakosol S., Worrall D.S., Boullier A., Butler S., Witztum J.L. Toll-like receptor 4-dependent and -independent cytokine secretion induced by minimally oxidized lowdensity lipoprotein in macrophages. Arterioscler Thromb. Vasc. Biol. (2005);25:1213–1219. doi: 10.1161/01.ATV.0000159891.73193.31. [DOI] [PubMed] [Google Scholar]
  • 166.Miwa S., Brand M.D. The topology of superoxide production by complex III and glycerol 3-phosphate dehydrogenase in Drosophila mitochondria. Biochim. Biophys. Acta. (2005);1709:214–219. doi: 10.1016/j.bbabio.2005.08.003. [DOI] [PubMed] [Google Scholar]
  • 167.Miyajima A., Nakashima J., Yoshioka K., Tachibana M., Tazaki H., Murai M. Role of reactive oxygen species in cisdichlorodiammineplatinum- induced cytotoxicity on bladder cancer cells. Br. J. Cancer. (1997);76:206–210. doi: 10.1038/bjc.1997.363. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 168.Miyano K., Koga H., Minakami R., Sumimoto H. The insert region of the Rac GTPases is dispensable for activation of superoxide-producing NADPH oxidases. Biochem. J. (2009);422:373–382. doi: 10.1042/BJ20082182. [DOI] [PubMed] [Google Scholar]
  • 169.Mizuno T., Kaibuchi K., Ando S., Musha T., Hiraoka K., Takaishi K., Asada M., Nunoi H., Matsuda I., Takai Y. Regulation of the superoxide-generating NADPH oxidase by a small GTP-binding protein and its stimulatory and inhibitory GDP/GTP exchange proteins. J. Biol. Chem. (1992);267:10215–10218. [PubMed] [Google Scholar]
  • 170.Morand S., Ueyama T., Tsujibe S., Saito N., Korzeniowska A., Leto T.L. Duox maturation factors form cell surface complexes with Duox affecting the specificity of reactive oxygen species generation. FASEB J. (2009);23:1205–1218. doi: 10.1096/fj.08-120006. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 171.Moreno J.C., Bikker H., Kempers M.J., van Trotsenburg A.S., Baas F., de Vijlder J.J., Vulsma T., Ris-Stalpers C. Inactivating mutations in the gene for thyroid oxidase 2 (THOX2) and congenital hypothyroidism. N. Engl. J. Med. (2002);347:95–102. doi: 10.1056/NEJMoa012752. [DOI] [PubMed] [Google Scholar]
  • 172.Moskwa P., Lorentzen D., Excoffon K.J., Zabner J., McCray P.B., Jr., Nauseef W.M., Dupuy C., Banfi B. A novel host defense system of airways is defective in cystic fibrosis. Am. J. Respir. Crit. Care Med. (2007);175:174–183. doi: 10.1164/rccm.200607-1029OC. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 173.Mouche S., Mkaddem S.B., Wang W., Katic M., Tseng Y.H., Carnesecchi S., Steger K., Foti M., Meier C.A., Muzzin P., et al. Reduced expression of the NADPH oxidase NOX4 is a hallmark of adipocyte differentiation. Biochim. Biophys. Acta. (2007);1773:1015–1027. doi: 10.1016/j.bbamcr.2007.03.003. [DOI] [PubMed] [Google Scholar]
  • 174.Muller F.L., Liu Y., Van Remmen H. Complex III releases superoxide to both sides of the inner mitochondrial membrane. J. Biol. Chem. (2004);279:49064–49073. doi: 10.1074/jbc.M407715200. [DOI] [PubMed] [Google Scholar]
  • 175.Murphy M.P. How mitochondria produce reactive oxygen species. Biochem. J. (2009);417:1–13. doi: 10.1042/BJ20081386. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 176.Na A.R., Chung Y.M., Lee S.B., Park S.H., Lee M.S., Yoo Y.D. A critical role for Romo1-derived ROS in cell proliferation. Biochem. Biophys. Res. Commun. (2008);369:672–678. doi: 10.1016/j.bbrc.2008.02.098. [DOI] [PubMed] [Google Scholar]
  • 177.Nakamura Y., Ohtaki S., Makino R., Tanaka T., Ishimura Y. Superoxide anion is the initial product in the hydrogen peroxide formation catalyzed by NADPH oxidase in porcine thyroid plasma membrane. J. Biol. Chem. (1989);264:4759–4761. [PubMed] [Google Scholar]
  • 178.Nakamura Y., Makino R., Tanaka T., Ishimura Y., Ohtaki S. Mechanism of H2O2 production in porcine thyroid cells: evidence for intermediary formation of superoxide anion by NADPH-dependent H2O2-generating machinery. Biochemistry. (1991);30:4880–4886. doi: 10.1021/bi00234a007. [DOI] [PubMed] [Google Scholar]
  • 179.Nakano Y., Longo-Guess C.M., Bergstrom D.E., Nauseef W.M., Jones S.M., Banfi B. Mutation of the Cyba gene encoding p22phox causes vestibular and immune defects in mice. J. Clin. Invest. (2008);118:1176–1185. doi: 10.1172/JCI33835. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 180.Nemoto S., Combs C.A., French S., Ahn B.H., Fergusson M.M., Balaban R.S., Finkel T. The mammalian longevityassociated gene product p66shc regulates mitochondrial metabolism. J. Biol. Chem. (2006);281:10555–10560. doi: 10.1074/jbc.M511626200. [DOI] [PubMed] [Google Scholar]
  • 181.Ni W., Zhan Y., He H., Maynard E., Balschi J.A., Oettgen P. Ets-1 is a critical transcriptional regulator of reactive oxygen species and p47(phox) gene expression in response to angiotensin II. Circ. Res. (2007);101:985–994. doi: 10.1161/CIRCRESAHA.107.152439. [DOI] [PubMed] [Google Scholar]
  • 182.Oh H., Jung H.Y., Kim J., Bae Y.S. Phosphorylation of serine282 in NADPH oxidase activator 1 by Erk desensitizes EGF-induced ROS generation. Biochem. Biophys. Res. Commun. (2010);394:691–696. doi: 10.1016/j.bbrc.2010.03.053. [DOI] [PubMed] [Google Scholar]
  • 183.Okado-Matsumoto A., Fridovich I. Subcellular distribution of superoxide dismutases (SOD) in rat liver: Cu,Zn-SOD in mitochondria. J. Biol. Chem. (2001);276:38388–38393. doi: 10.1074/jbc.M105395200. [DOI] [PubMed] [Google Scholar]
  • 184.Pachucki J., Wang D., Christophe D., Miot F. Structural and functional characterization of the two human ThOX/ Duox genes and their 5′-flanking regions. Mol. Cell. Endocrinol. (2004);214:53–62. doi: 10.1016/j.mce.2003.11.026. [DOI] [PubMed] [Google Scholar]
  • 185.Pacquelet S., Lehmann M., Luxen S., Regazzoni K., Frausto M., Noack D., Knaus U.G. Inhibitory action of NoxA1 on dual oxidase activity in airway cells. J. Biol. Chem. (2008);283:24649–24658. doi: 10.1074/jbc.M709108200. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 186.Paffenholz R., Bergstrom R.A., Pasutto F., Wabnitz P., Munroe R.J., Jagla W., Heinzmann U., Marquardt A., Bareiss A., Laufs J., et al. Vestibular defects in head-tilt mice result from mutations in Nox3, encoding an NADPH oxidase. Genes Dev. (2004);18:486–491. doi: 10.1101/gad.1172504. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 187.Pandey S., Lopez C., Jammu A. Oxidative stress and activation of proteasome protease during serum deprivationinduced apoptosis in rat hepatoma cells; inhibition of cell death by melatonin. Apoptosis. (2003);8:497–508. doi: 10.1023/a:1025542424986. [DOI] [PubMed] [Google Scholar]
  • 188.Pandey D., Gratton J.P., Rafikov R., Black S.M., Fulton D.J. Calcium/calmodulin-dependent kinase II mediates the phosphorylation and activation of NADPH oxidase 5. Mol. Pharmacol. (2011);80:407–415. doi: 10.1124/mol.110.070193. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 189.Papa S., Zazzeroni F., Bubici C., Jayawardena S., Alvarez K., Matsuda S., Nguyen D.U., Pham C.G., Nelsbach A.H., Melis T., et al. Gadd45 beta mediates the NF-kappa B suppression of JNK signalling by targeting MKK7/JNKK2. Nat. Cell Biol. (2004);6:146–153. doi: 10.1038/ncb1093. [DOI] [PubMed] [Google Scholar]
  • 190.Park H.S., Lee S.H., Park D., Lee J.S., Ryu S.H., Lee W.J., Rhee S.G., Bae Y.S. Sequential activation of phosphatidylinositol 3-kinase, beta Pix, Rac1, and Nox1 in growth factor-induced production of H2O2. Mol. Cell. Biol. (2004);24:4384–4394. doi: 10.1128/MCB.24.10.4384-4394.2004. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 191.Park H.S., Chun J.N., Jung H.Y., Choi C., Bae Y.S. Role of NADPH oxidase 4 in lipopolysaccharide-induced proinflammatory responses by human aortic endothelial cells. Cardiovasc. Res. (2006a);72:447–455. doi: 10.1016/j.cardiores.2006.09.012. [DOI] [PubMed] [Google Scholar]
  • 192.Park H.S., Park D., Bae Y.S. Molecular interaction of NADPH oxidase 1 with betaPix and Nox organizer 1. Biochem. Biophys. Res. Commun. (2006b);339:985–990. doi: 10.1016/j.bbrc.2005.11.108. [DOI] [PubMed] [Google Scholar]
  • 193.Parkos C.A., Dinauer M.C., Jesaitis A.J., Orkin S.H., Curnutte J.T. Absence of both the 91kD and 22kD subunits of human neutrophil cytochrome b in two genetic forms of chronic granulomatous disease. Blood. (1989);73:1416–1420. [PubMed] [Google Scholar]
  • 194.Patel D.N., Bailey S.R., Gresham J.K., Schuchman D.B., Shelhamer J.H., Goldstein B.J., Foxwell B.M., Stemerman M.B., Maranchie J.K., Valente A.J., et al. TLR4-NOX4-AP-1 signaling mediates lipopolysaccharide-induced CXCR6 expression in human aortic smooth muscle cells. Biochem. Biophys. Res. Commun. (2006);347:1113–1120. doi: 10.1016/j.bbrc.2006.07.015. [DOI] [PubMed] [Google Scholar]
  • 195.Pelicano H., Feng L., Zhou Y., Carew J.S., Hileman E.O., Plunkett W., Keating M.J., Huang P. Inhibition of mitochondrial respiration: a novel strategy to enhance drug-induced apoptosis in human leukemia cells by a reactive oxygen species-mediated mechanism. J. Biol. Chem. (2003);278:37832–37839. doi: 10.1074/jbc.M301546200. [DOI] [PubMed] [Google Scholar]
  • 196.Pelicano H., Carney D., Huang P. ROS stress in cancer cells and therapeutic implications. Drug Resist. Updat. (2004);7:97–110. doi: 10.1016/j.drup.2004.01.004. [DOI] [PubMed] [Google Scholar]
  • 197.Pelicci G., Lanfrancone L., Grignani F., McGlade J., Cavallo F., Forni G., Nicoletti I., Pawson T., Pelicci P.G. A novel transforming protein (SHC) with an SH2 domain is implicated in mitogenic signal transduction. Cell. (1992);70:93–104. doi: 10.1016/0092-8674(92)90536-l. [DOI] [PubMed] [Google Scholar]
  • 198.Pham C.G., Bubici C., Zazzeroni F., Papa S., Jones J., Alvarez K., Jayawardena S., De Smaele E., Cong R., Beaumont C., et al. Ferritin heavy chain upregulation by NF-kappaB inhibits TNFalpha-induced apoptosis by suppressing reactive oxygen species. Cell. (2004);119:529–542. doi: 10.1016/j.cell.2004.10.017. [DOI] [PubMed] [Google Scholar]
  • 199.Pim D., Banks L. p53 polymorphic variants at codon 72 exert different effects on cell cycle progression. Int. J. Cancer. (2004);108:196–199. doi: 10.1002/ijc.11548. [DOI] [PubMed] [Google Scholar]
  • 200.Pinton P., Rimessi A., Marchi S., Orsini F., Migliaccio E., Giorgio M., Contursi C., Minucci S., Mantovani F., Wieckowski M.R., et al. Protein kinase C beta and prolyl isomerase 1 regulate mitochondrial effects of the life-span determinant p66Shc. Science. (2007);315:659–663. doi: 10.1126/science.1135380. [DOI] [PubMed] [Google Scholar]
  • 201.Polyak K., Xia Y., Zweier J.L., Kinzler K.W., Vogelstein B. A model for p53-induced apoptosis. Nature. (1997);389:300–305. doi: 10.1038/38525. [DOI] [PubMed] [Google Scholar]
  • 202.Ponting C.P. Novel domains in NADPH oxidase subunits, sorting nexins, and PtdIns 3-kinases: binding partners of SH3 domains? Protein Sci. (1996);5:2353–2357. doi: 10.1002/pro.5560051122. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 203.Poyton R.O., Ball K.A., Castello P.R. Mitochondrial generation of free radicals and hypoxic signaling. Trends Endocrinol. Metab. (2009);20:332–340. doi: 10.1016/j.tem.2009.04.001. [DOI] [PubMed] [Google Scholar]
  • 204.Price M.O., McPhail L.C., Lambeth J.D., Han C.H., Knaus U.G., Dinauer M.C. Creation of a genetic system for analysis of the phagocyte respiratory burst: high-level reconstitution of the NADPH oxidase in a nonhematopoietic system. Blood. (2002);99:2653–2661. doi: 10.1182/blood.v99.8.2653. [DOI] [PubMed] [Google Scholar]
  • 205.Reed J.C. Proapoptotic multidomain Bcl-2/Bax-family proteins: mechanisms, physiological roles, and therapeutic opportunities. Cell Death Differ. (2006);13:1378–1386. doi: 10.1038/sj.cdd.4401975. [DOI] [PubMed] [Google Scholar]
  • 206.Reeves E.P., Lu H., Jacobs H.L., Messina C.G., Bolsover S., Gabella G., Potma E.O., Warley A., Roes J., Segal A.W. Killing activity of neutrophils is mediated through activation of proteases by K+ flux. Nature. (2002);416:291–297. doi: 10.1038/416291a. [DOI] [PubMed] [Google Scholar]
  • 207.Rhee S.G., Woo H.A. Multiple functions of peroxiredoxins: peroxidases, sensors and regulators of the intracellular messenger HO, and protein chaperones. Antioxid Redox Signal. (2011);15:781–794. doi: 10.1089/ars.2010.3393. [DOI] [PubMed] [Google Scholar]
  • 208.Rigutto S., Hoste C., Grasberger H., Milenkovic M., Communi D., Dumont J.E., Corvilain B., Miot F., De Deken X. Activation of dual oxidases Duox1 and Duox2: differential regulation mediated by camp-dependent protein kinase and protein kinase C-dependent phosphorylation. J. Biol. Chem. (2009);284:6725–6734. doi: 10.1074/jbc.M806893200. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 209.Rittle J., Green M.T. Cytochrome P450 compound I: capture, characterization, and C-H bond activation kinetics. Science. (2010);330:933–937. doi: 10.1126/science.1193478. [DOI] [PubMed] [Google Scholar]
  • 210.Rivera A., Maxwell S.A. The p53-induced gene-6 (proline oxidase) mediates apoptosis through a calcineurin-dependent pathway. J. Biol. Chem. (2005);280:29346–29354. doi: 10.1074/jbc.M504852200. [DOI] [PubMed] [Google Scholar]
  • 211.Rokutan K., Kawahara T., Kuwano Y., Tominaga K., Sekiyama A., Teshima-Kondo S. NADPH oxidases in the gastrointestinal tract: a potential role of Nox1 in innate immune response and carcinogenesis. Antioxid Redox Signal. (2006);8:1573–1582. doi: 10.1089/ars.2006.8.1573. [DOI] [PubMed] [Google Scholar]
  • 212.Rong Y., Distelhorst C.W. Bcl-2 protein family members: versatile regulators of calcium signaling in cell survival and apoptosis. Annu. Rev. Physiol. (2008);70:73–91. doi: 10.1146/annurev.physiol.70.021507.105852. [DOI] [PubMed] [Google Scholar]
  • 213.Rothe M., Wong S.C., Henzel W.J., Goeddel D.V. A novel family of putative signal transducers associated with the cytoplasmic domain of the 75 kDa tumor necrosis factor receptor. Cell. (1994);78:681–692. doi: 10.1016/0092-8674(94)90532-0. [DOI] [PubMed] [Google Scholar]
  • 214.Rouhanizadeh M., Hwang J., Clempus R.E., Marcu L., Lassegue B., Sevanian A., Hsiai T.K. Oxidized-1-palmitoyl-2- arachidonoyl-sn-glycero-3-phosphorylcholine induces vascular endothelial superoxide production: implication of NADPH oxidase. Free Radic. Biol. Med. (2005);39:1512–1522. doi: 10.1016/j.freeradbiomed.2005.07.013. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 215.Sablina A.A., Budanov A.V., Ilyinskaya G.V., Agapova L.S., Kravchenko J.E., Chumakov P.M. The antioxidant function of the p53 tumor suppressor. Nat. Med. (2005);11:1306–1313. doi: 10.1038/nm1320. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 216.Sakon S., Xue X., Takekawa M., Sasazuki T., Okazaki T., Kojima Y., Piao J.H., Yagita H., Okumura K., Doi T., et al. NF-kappaB inhibits TNF-induced accumulation of ROS that mediate prolonged MAPK activation and necrotic cell death. EMBO J. (2003);22:3898–3909. doi: 10.1093/emboj/cdg379. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 217.Sansome C., Zaika A., Marchenko N.D., Moll U.M. Hypoxia death stimulus induces translocation of p53 protein to mitochondria. Detection by immunofluorescence on whole cells. FEBS Lett. (2001);488:110–115. doi: 10.1016/s0014-5793(00)02368-1. [DOI] [PubMed] [Google Scholar]
  • 218.Satoh T., Sakai N., Enokido Y., Uchiyama Y., Hatanaka H. Survival factor-insensitive generation of reactive oxygen species induced by serum deprivation in neuronal cells. Brain Res. (1996);733:9–14. doi: 10.1016/0006-8993(96)00527-6. [DOI] [PubMed] [Google Scholar]
  • 219.Schleiss M.B., Holz O., Behnke M., Richter K., Magnussen H., Jorres R.A. The concentration of hydrogen peroxide in exhaled air depends on expiratory flow rate. Eur. Respir. J. (2000);16:1115–1118. doi: 10.1034/j.1399-3003.2000.16f16.x. [DOI] [PubMed] [Google Scholar]
  • 220.Schrenzel J., Serrander L., Banfi B., Nusse O., Fouyouzi R., Lew D.P., Demaurex N., Krause K.H. Electron currents generated by the human phagocyte NADPH oxidase. Nature. (1998);392:734–737. doi: 10.1038/33725. [DOI] [PubMed] [Google Scholar]
  • 221.Schroder K., Helmcke I., Palfi K., Krause K.H., Busse R., Brandes R.P. Nox1 mediates basic fibroblast growth factor-induced migration of vascular smooth muscle cells. Arterioscler. Thromb. Vasc. Biol. (2007);27:1736–1743. doi: 10.1161/ATVBAHA.107.142117. [DOI] [PubMed] [Google Scholar]
  • 222.Segal A.W. How neutrophils kill microbes. Annu. Rev. Immunol. (2005);23:197–223. doi: 10.1146/annurev.immunol.23.021704.115653. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 223.Segal A.W., West I., Wientjes F., Nugent J.H., Chavan A.J., Haley B., Garcia R.C., Rosen H., Scrace G. Cytochrome b-245 is a flavocytochrome containing FAD and the NADPH-binding site of the microbicidal oxidase of phagocytes. Biochem. J. (1992);284(Pt 3):781–788. doi: 10.1042/bj2840781. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 224.Sekharam M., Trotti A., Cunnick J.M., Wu J. Suppression of fibroblast cell cycle progression in G1 phase by Nacetylcysteine. Toxicol. Appl. Pharmacol. (1998);149:210–216. doi: 10.1006/taap.1997.8361. [DOI] [PubMed] [Google Scholar]
  • 225.Shoji Y., Uedono Y., Ishikura H., Takeyama N., Tanaka T. DNA damage induced by tumour necrosis factor-alpha in L929 cells is mediated by mitochondrial oxygen radical formation. Immunology. (1995);84:543–548. [PMC free article] [PubMed] [Google Scholar]
  • 226.Sigaud S., Evelson P., Gonzalez-Flecha B. H2O2- induced proliferation of primary alveolar epithelial cells is mediated by MAP kinases. Antioxid Redox Signal. (2005);7:6–13. doi: 10.1089/ars.2005.7.6. [DOI] [PubMed] [Google Scholar]
  • 227.Spat A., Hunyady L. Control of aldosterone secretion: a model for convergence in cellular signaling pathways. Physiol. Rev. (2004);84:489–539. doi: 10.1152/physrev.00030.2003. [DOI] [PubMed] [Google Scholar]
  • 228.St-Pierre J., Buckingham J.A., Roebuck S.J., Brand M.D. Topology of superoxide production from different sites in the mitochondrial electron transport chain. J. Biol. Chem. (2002);277:44784–44790. doi: 10.1074/jbc.M207217200. [DOI] [PubMed] [Google Scholar]
  • 229.Stahelin R.V., Burian A., Bruzik K.S., Murray D., Cho W. Membrane binding mechanisms of the PX domains of NADPH oxidase p40phox and p47phox. J. Biol. Chem. (2003);278:14469–14479. doi: 10.1074/jbc.M212579200. [DOI] [PubMed] [Google Scholar]
  • 230.Stambolsky P., Weisz L., Shats I., Klein Y., Goldfinger N., Oren M., Rotter V. Regulation of AIF expression by p53. Cell Death Differ. (2006);13:2140–2149. doi: 10.1038/sj.cdd.4401965. [DOI] [PubMed] [Google Scholar]
  • 231.Stasia M.J., Bordigoni P., Martel C., Morel F. A novel and unusual case of chronic granulomatous disease in a child with a homozygous 36-bp deletion in the CYBA gene (A22(0)) leading to the activation of a cryptic splice site in intron 4. Hum. Genet. (2002);110:444–450. doi: 10.1007/s00439-002-0720-8. [DOI] [PubMed] [Google Scholar]
  • 232.Stone J.R., Collins T. The role of hydrogen peroxide in endothelial proliferative responses. Endothelium. (2002);9:231–238. doi: 10.1080/10623320214733. [DOI] [PubMed] [Google Scholar]
  • 233.Stone J.R., Yang S. Hydrogen peroxide: a signaling messenger. Antioxid. Redox Signal. (2006);8:243–270. doi: 10.1089/ars.2006.8.243. [DOI] [PubMed] [Google Scholar]
  • 234.Sturtz L.A., Diekert K., Jensen L.T., Lill R., Culotta V.C. A fraction of yeast Cu,Zn-superoxide dismutase and its metallochaperone, CCS, localize to the intermembrane space of mitochondria. A physiological role for SOD1 in guarding against mitochondrial oxidative damage. J. Biol. Chem. (2001);276:38084–38089. doi: 10.1074/jbc.M105296200. [DOI] [PubMed] [Google Scholar]
  • 235.Suh Y.A., Arnold R.S., Lassegue B., Shi J., Xu X., Sorescu D., Chung A.B., Griendling K.K., Lambeth J.D. Cell transformation by the superoxide-generating oxidase Mox1. Nature. (1999);401:79–82. doi: 10.1038/43459. [DOI] [PubMed] [Google Scholar]
  • 236.Sulciner D.J., Irani K., Yu Z.X., Ferrans V.J., Goldschmidt-Clermont P., Finkel T. rac1 regulates a cytokinestimulated, redox-dependent pathway necessary for NF-kappaB activation. Mol. Cell. Biol. (1996);16:7115–7121. doi: 10.1128/mcb.16.12.7115. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 237.Sumimoto H., Sakamoto N., Nozaki M., Sakaki Y., Takeshige K., Minakami S. Cytochrome b558, a component of the phagocyte NADPH oxidase, is a flavoprotein. Biochem. Biophys. Res. Commun. (1992);186:1368–1375. doi: 10.1016/s0006-291x(05)81557-8. [DOI] [PubMed] [Google Scholar]
  • 238.Sundaresan M., Yu Z.X., Ferrans V.J., Irani K., Finkel T. Requirement for generation of H2O2 for platelet-derived growth factor signal transduction. Science. (1995);270:296–299. doi: 10.1126/science.270.5234.296. [DOI] [PubMed] [Google Scholar]
  • 239.Szatrowski T.P., Nathan C.F. Production of large amounts of hydrogen peroxide by human tumor cells. Cancer Res. (1991);51:794–798. [PubMed] [Google Scholar]
  • 240.Tabcharani J.A., Rommens J.M., Hou Y.X., Chang X.B., Tsui L.C., Riordan J.R., Hanrahan J.W. Multi-ion pore behaviour in the CFTR chloride channel. Nature. (1993);366:79–82. doi: 10.1038/366079a0. [DOI] [PubMed] [Google Scholar]
  • 241.Takeya R., Ueno N., Kami K., Taura M., Kohjima M., Izaki T., Nunoi H., Sumimoto H. Novel human homologues of p47phox and p67phox participate in activation of superoxideproducing NADPH oxidases. J. Biol. Chem. (2003);278:25234–25246. doi: 10.1074/jbc.M212856200. [DOI] [PubMed] [Google Scholar]
  • 242.Tan S., Sagara Y., Liu Y., Maher P., Schubert D. The regulation of reactive oxygen species production during programmed cell death. J. Cell Biol. (1998);141:1423–1432. doi: 10.1083/jcb.141.6.1423. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 243.Tang G., Minemoto Y., Dibling B., Purcell N.H., Li Z., Karin M., Lin A. Inhibition of JNK activation through NF-kappaB target genes. Nature. (2001);414:313–317. doi: 10.1038/35104568. [DOI] [PubMed] [Google Scholar]
  • 244.Tirone F., Cox J.A. NADPH oxidase 5 (NOX5) interacts with and is regulated by calmodulin. FEBS Lett. (2007);581:1202–1208. doi: 10.1016/j.febslet.2007.02.047. [DOI] [PubMed] [Google Scholar]
  • 245.Tomko R.J., Jr., Bansal P., Lazo J.S. Airing out an antioxidant role for the tumor suppressor p53. Mol. Interv. (2006);6:23–25. doi: 10.1124/mi.6.1.5. 22. [DOI] [PubMed] [Google Scholar]
  • 246.Trinei M., Giorgio M., Cicalese A., Barozzi S., Ventura A., Migliaccio E., Milia E., Padura I.M., Raker V.A., Maccarana M., et al. A p53-p66Shc signalling pathway controls intracellular redox status, levels of oxidation-damaged DNA and oxidative stress-induced apoptosis. Oncogene. (2002);21:3872–3878. doi: 10.1038/sj.onc.1205513. [DOI] [PubMed] [Google Scholar]
  • 247.Turrens J.F., Alexandre A., Lehninger A.L. Ubisemiquinone is the electron donor for superoxide formation by complex III of heart mitochondria. Arch. Biochem. Biophys. (1985);237:408–414. doi: 10.1016/0003-9861(85)90293-0. [DOI] [PubMed] [Google Scholar]
  • 248.Ueno N., Takeya R., Miyano K., Kikuchi H., Sumimoto H. The NADPH oxidase Nox3 constitutively produces superoxide in a p22phox-dependent manner: its regulation by oxidase organizers and activators. J. Biol. Chem. (2005);280:23328–23339. doi: 10.1074/jbc.M414548200. [DOI] [PubMed] [Google Scholar]
  • 249.Ueyama T., Geiszt M., Leto T.L. Involvement of Rac1 in activation of multicomponent Nox1- and Nox3-based NADPH oxidases. Mol. Cell. Biol. (2006);26:2160–2174. doi: 10.1128/MCB.26.6.2160-2174.2006. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 250.Ueyama T., Tatsuno T., Kawasaki T., Tsujibe S., Shirai Y., Sumimoto H., Leto T.L., Saito N. A regulated adaptor function of p40phox: distinct p67phox membrane targeting by p40phox and by p47phox. Mol. Biol. Cell. (2007);18:441–454. doi: 10.1091/mbc.E06-08-0731. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 251.Valko M., Rhodes C.J., Moncol J., Izakovic M., Mazur M. Free radicals, metals and antioxidants in oxidative stressinduced cancer. Chem. Biol. Interact. (2006);160:1–40. doi: 10.1016/j.cbi.2005.12.009. [DOI] [PubMed] [Google Scholar]
  • 252.van der Valk J., Mellor D., Brands R., Fischer R., Gruber F., Gstraunthaler G., Hellebrekers L., Hyllner J., Jonker F.H., Prieto P., et al. The humane collection of fetal bovine serum and possibilities for serum-free cell and tissue culture. Toxicol. In Vitro. (2004);18:1–12. doi: 10.1016/j.tiv.2003.08.009. [DOI] [PubMed] [Google Scholar]
  • 253.Vander Heiden M.G., Chandel N.S., Williamson E.K., Schumacker P.T., Thompson C.B. Bcl-xL regulates the membrane potential and volume homeostasis of mitochondria. Cell. (1997);91:627–637. doi: 10.1016/s0092-8674(00)80450-x. [DOI] [PubMed] [Google Scholar]
  • 254.Venditti P., Daniele C.M., De Leo T., Di Meo S. Effect of phenobarbital treatment on characteristics determining susceptibility to oxidants of homogenates, mitochondria and microsomes from rat liver. Cell Physiol. Biochem. (1998);8:328–338. doi: 10.1159/000016294. [DOI] [PubMed] [Google Scholar]
  • 255.Verkman A.S., Song Y., Thiagarajah J.R. Role of airway surface liquid and submucosal glands in cystic fibrosis lung disease. Am. J. Physiol. Cell Physiol. (2003);284:C2–15. doi: 10.1152/ajpcell.00417.2002. [DOI] [PubMed] [Google Scholar]
  • 256.Verma I.M., Stevenson J.K., Schwarz E.M., Van Antwerp D., Miyamoto S. Rel/NF-kappa B/I kappa B family: intimate tales of association and dissociation. Genes Dev. (1995);9:2723–2735. doi: 10.1101/gad.9.22.2723. [DOI] [PubMed] [Google Scholar]
  • 257.Votyakova T.V., Reynolds I.J. DeltaPsi(m)-Dependent and -independent production of reactive oxygen species by rat brain mitochondria. J. Neurochem. (2001);79:266–277. doi: 10.1046/j.1471-4159.2001.00548.x. [DOI] [PubMed] [Google Scholar]
  • 258.Wajant H., Pfizenmaier K., Scheurich P. Tumor necrosis factor signaling. Cell Death Differ. (2003);10:45–65. doi: 10.1038/sj.cdd.4401189. [DOI] [PubMed] [Google Scholar]
  • 259.Weisiger R.A., Fridovich I. Mitochondrial superoxide simutase. Site of synthesis and intramitochondrial localization. J. Biol. Chem. (1973a);248:4793–4796. [PubMed] [Google Scholar]
  • 260.Weisiger R.A., Fridovich I. Superoxide dismutase. Organelle specificity. J. Biol. Chem. (1973b);248:3582–3592. [PubMed] [Google Scholar]
  • 261.Wilson M.I., Gill D.J., Perisic O., Quinn M.T., Williams R.L. PB1 domain-mediated heterodimerization in NADPH oxidase and signaling complexes of atypical protein kinase C with Par6 and p62. Mol. Cell. (2003);12:39–50. doi: 10.1016/s1097-2765(03)00246-6. [DOI] [PubMed] [Google Scholar]
  • 262.Wolter K.G., Hsu Y.T., Smith C.L., Nechushtan A., Xi X.G., Youle R.J. Movement of Bax from the cytosol to mitochondria during apoptosis. J. Cell Biol. (1997);139:1281–1292. doi: 10.1083/jcb.139.5.1281. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 263.Wu W.S. The signaling mechanism of ROS in tumor progression. Cancer Metastasis Rev. (2006);25:695–705. doi: 10.1007/s10555-006-9037-8. [DOI] [PubMed] [Google Scholar]
  • 264.Yoo Y.A., Kim M.J., Park J.K., Chung Y.M., Lee J.H., Chi S.G., Kim J.S., Yoo Y.D. Mitochondrial ribosomal protein L41 suppresses cell growth in association with p53 and p27Kip1. Mol. Cell. Biol. (2005);25:6603–6616. doi: 10.1128/MCB.25.15.6603-6616.2005. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 265.Yuzawa S., Ogura K., Horiuchi M., Suzuki N.N., Fujioka Y., Kataoka M., Sumimoto H., Inagaki F. Solution structure of the tandem Src homology 3 domains of p47phox in an autoinhibited form. J. Biol. Chem. (2004a);279:29752–29760. doi: 10.1074/jbc.M401457200. [DOI] [PubMed] [Google Scholar]
  • 266.Yuzawa S., Suzuki N.N., Fujioka Y., Ogura K., Sumimoto H., Inagaki F. A molecular mechanism for autoinhibition of the tandem SH3 domains of p47phox, the regulatory subunit of the phagocyte NADPH oxidase. Genes Cells. (2004b);9:443–456. doi: 10.1111/j.1356-9597.2004.00733.x. [DOI] [PubMed] [Google Scholar]
  • 267.Zangar R.C., Davydov D.R., Verma S. Mechanisms that regulate production of reactive oxygen species by cytochrome P450. Toxicol. Appl. Pharmacol. (2004);199:316–331. doi: 10.1016/j.taap.2004.01.018. [DOI] [PubMed] [Google Scholar]
  • 268.Zhuge J., Cederbaum A.I. Serum deprivation-induced HepG2 cell death is potentiated by CYP2E1. Free Radic. Biol. Med. (2006);40:63–74. doi: 10.1016/j.freeradbiomed.2005.08.012. [DOI] [PubMed] [Google Scholar]
  • 269.Zhukov A.A., Archakov A.I. Complete stoichiometry of free NADPH oxidation in liver microsomes. Biochem. Biophys. Res. Commun. (1982);109:813–818. doi: 10.1016/0006-291x(82)92012-5. [DOI] [PubMed] [Google Scholar]
  • 270.Zidi I., Mestiri S., Bartegi A., Amor N.B. TNF-alpha and its inhibitors in cancer. Med. Oncol. (2010);27:185–198. doi: 10.1007/s12032-009-9190-3. [DOI] [PubMed] [Google Scholar]
  • 271.Ziegler G.A., Vonrhein C., Hanukoglu I., Schulz G.E. The structure of adrenodoxin reductase of mitochondrial P450 systems: electron transfer for steroid biosynthesis. J. Mol. Biol. (1999);289:981–990. doi: 10.1006/jmbi.1999.2807. [DOI] [PubMed] [Google Scholar]
  • 272.Zorov D.B., Filburn C.R., Klotz L.O., Zweier J.L., Sollott S.J. Reactive oxygen species (ROS)-induced ROS release: a new phenomenon accompanying induction of the mitochondrial permeability transition in cardiac myocytes. J. Exp. Med. (2000);192:1001–1014. doi: 10.1084/jem.192.7.1001. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 273.Zorov D.B., Juhaszova M., Sollott S.J. Mitochondrial ROS-induced ROS release: an update and review. Biochim. Biophys. Acta. (2006);1757:509–517. doi: 10.1016/j.bbabio.2006.04.029. [DOI] [PubMed] [Google Scholar]

Articles from Molecules and Cells are provided here courtesy of Korean Society for Molecular and Cellular Biology

RESOURCES