Open access peer-reviewed chapter

Physiological Functions of Mitochondrial Reactive Oxygen Species

Written By

Tae Gyu Choi and Sung Soo Kim

Submitted: 11 January 2019 Reviewed: 03 July 2019 Published: 13 December 2019

DOI: 10.5772/intechopen.88386

From the Edited Volume

Free Radical Medicine and Biology

Edited by Kusal Das, Swastika Das, Mallanagouda Shivanagouda Biradar, Varaprasad Bobbarala and S. Subba Tata

Chapter metrics overview

1,656 Chapter Downloads

View Full Metrics

Abstract

Mitochondria are the major energy producers within a cell in the form of adenosine triphosphate by oxidative phosphorylation. Normal mitochondrial metabolism inevitably generates reactive oxygen species (ROS), which have been considered to solely cause cellular damage. Increase of oxidative stress has been linked to various pathologies. Thus, mitochondrial ROS (mROS) were basically proposed as byproducts of oxidative metabolism, which undergo normalized by antioxidant enzymes. However, the mROS have extensively been esteemed to function as signalling molecules to regulate a wide variety of physiology. These phenomena are indeed dependent on mitochondrial redox status, which is dynamically altered under different physiological and pathological conditions. The oxidative stress is incurred by which the redox status is inclined to exceeded oxidation or reduction. Here, we attempt to integrate the recent advances in our understanding of the physiological functions of mROS.

Keywords

  • mitochondrial ROS
  • oxidative stress
  • oxidative metabolism
  • redox signaling
  • mitochondrial physiology

1. Introduction

Mitochondria are double-membrane-bound cellular organelles found in most eukaryotic organisms. The number of mitochondria in cell differs widely according to organisms, tissues and cell types, which is determined by the energy demand. Mitochondria occupy around 40% of the cytoplasm in heart muscle cells and 20–25% with ~2000 per cell in liver cells. Mitochondria, as the power plants of the cell, mainly generate energy in forms of adenosine triphosphates (ATPs) by oxidative phosphorylation (OXPHOS) during glucose metabolism [1, 2]. The OXPHOS is coupled with mitochondrial respiration in which mitochondrial transmembrane potential (MMP, ΔΨm) is generated by pumping the protons via mitochondrial complexes I, III and IV of the electron transport chain (ETC) [3].

Molecular oxygen (O2) is essential for the mitochondrial bioenergetic metabolism, which functions as the final electron acceptor for cytochrome c oxidase (complex IV) in the respiratory ETC that catalyses the four-electron reduction of O2 to H2O. Mitochondria are an important source of reactive oxygen species (ROS) within most mammalian cells [4, 5]; mitochondrial ROS (mROS) are basically produced as byproducts of this bioenergetic metabolism during the OXPHOS [6]. Electron leaks at complex I and III from ETC lead to forming partially reduced and highly reactive metabolites of O2, including superoxide anion (O2 ·−) and hydrogen peroxide (H2O2), formed by one- and two-electron reductions of O2, respectively [7]. In the presence of transition metal ions, the more reactive hydroxyl radical (OH·) is formed. The O2 ·− is rapidly dismutated to H2O2 by two dismutases including Cu/Zn-superoxide dismutase (Cu/ZnSOD) in mitochondrial intermembrane space and manganese-dependent superoxide dismutase (MnSOD) in mitochondrial matrix. Unless the dismutation of O2 ·− is catalyzed into H2O2, the radical oxidant promotes DNA damage, protein oxidation and lipid peroxidation in many types of cells. H2O2 is also cell damaging molecule to be degraded to water by catalase [8]. Although the O2 ·− generation by respiratory complexes is a well-established phenomenon, it is still poorly understood in mechanism [9].

Mitochondria have been implicated in the regulation of a number of physiological and pathological processes, including proliferation, differentiation, programmed cell death, innate immunity, autophagy, redox signalling, calcium homeostasis, hypoxic stress responses and stem cell reprogramming [10, 11, 12, 13, 14, 15, 16]. The mROS production contributes to mitochondrial damage in a range of pathologies, which is also is closely related to redox signalling in the cell [4, 17]. However, accumulating evidences show that mROS are not only deleterious molecules derived from the cellular metabolism but also indispensable participants in diverse cellular signalling and regulations [18, 19, 20].

In this chapter, we briefly summarize recent developments in our understanding of the involvement of mROS as signalling mediators in redox biology, rather than pathological stress, underlying physiological conditions.

Advertisement

2. Mitochondrial physiology and ROS production

Mitochondria, cellular organelles in cells of eukaryotic organisms, have a primarily role in the process of pyruvate breakdown and ATP synthesis, generating water and carbon dioxide (CO2) as the end products via aerobic respiration [21]. Mitochondria turned into driving forces in biological evolution after the symbiotic engulfment of aerobic α-proteobacteria by a precursor of the eukaryotic cells around 2 billion years ago [22, 23]. Although mitochondria have preserved the double-membrane shape and ATP production characters of their ancestors, they have attained numerous additional functions in the cell, dramatically altering their structure and composition [24]. Most part of the bacterial genome was rapidly lost or transferred to the nuclear DNA [25]. Mammalian mitochondrial genome is transmitted solely through the female germ line [26]. Human mitochondrial DNA (mtDNA) is a double-stranded, circular molecule of 16,569 bp and contains 37 genes coding for two ribosomal RNAs (rRNAs), 22 transfer RNAs (tRNAs) and 13 proteins [22]. As the major power plants, mitochondria constantly produce reactive radical oxidants as byproducts during OXPHOS. Thus, in response to the metabolic or environmental stresses, mitochondria have accomplished antioxidant defence system [27]. Mitochondria are also highly dynamic to maintain the functions, which form a tubular network that continually changes by fission and fusion [28]. In this section, we concisely discuss overall mitochondrial biology and the ROS generation.

2.1 Mitochondrial structure and genome

A mitochondrion comprises four subcompartments, the outer mitochondrial membrane (OMM) and inner mitochondrial membrane (IMM), and the two soluble compartments intermembrane space (IMS) and matrix, which are architecturally and functionally distinct. The OMM encloses the organelle, which has a protein-to-phospholipid ratio similar to that of the plasma membrane [29]. It contains large numbers of integral membrane protein, porin [30]. Voltage-dependent anion channel (VDAC) is a major trafficking protein that forms a beta barrel spanning the outer membrane, which transports nucleotides, ions and metabolites between cytosol and intermembrane space [31, 32]. The IMM is found inside of the OMM, which encloses mitochondrial matrix, extensively folded and compartmentalized [33]. The IMM is non-permeable to nucleotides, sugars and small ions; thus specific carrier proteins enable the molecules to transport across the membrane [34].

The mitochondrial respiratory complexes I–IV, in which electrochemical gradient is generated for OXPHOS to occur for ATP synthesis, are embedded in the IMM [22]. Mitochondria contain two aqueous compartments: the IMS and matrix. The IMS, existing between the OMM and IMM, relays molecular transport from cytosol to mitochondrial matrix or reversely [35]. The compositions of small molecules such as ions and sugars in IMS are chemically similar to those in cytosol, as OMM is selectively permeable to those molecules [36]. However, in case of large proteins, the specific signalling peptides are required to be transported across the OMM. Thus, the protein composition of the IMS is different to the protein composition of the cytosol (e.g. cytochrome c) [37]. The mitochondrial matrix, enclosed by IMM, contains mitochondrial DNA (mtDNA), RNA and proteins. Especially, a number of proteins in the matrix are involved in diverse biochemical processes such as tricarboxylic acid (TCA) cycle, fatty acid oxidation, amino acid degradation and mitochondrial dynamics (fission and fusion) [27, 38] (Figure 1).

Figure 1.

Mitochondrial structure.

2.2 Mitochondrial genome

Mitochondria contain their own genetic material (mtDNA), which is maternally inherited without DNA recombination and encodes 37 genes that participate in mitochondrial ATP synthesis. Thirteen genes of them are involved in OXPHOS, and the rest two rRNAs and 22 tRNAs. One human cell has hundreds to thousands of mtDNA copies [39, 40]. mtDNA has high rates of mutation and sequence evolution, and mutant and wild-type mtDNA are present in the cell at different proportions [41, 42]. The mtDNA mutations lead to abnormality in OXPHOS activity and ATP synthesis [43]. The mtDNA is exposed to OXPHOS-derived ROS without conventional histone proteins. Moreover, in the lacking repair mechanisms, mtDNA damage further amplifies during DNA replication [44]. Therefore, the mtDNA is susceptible to mutation and damage. Therefore, mtDNA mutations and damage cause mitochondrial dysfunction, including ATP synthesis impediment, intracellular calcium level elevation, phospholipase activation and membrane phospholipids decomposition [45, 46].

2.3 Mitochondrial bioenergetics and dynamics

The most prominent roles of mitochondria are to produce the cellular energy in forms of ATPs via aerobic respiration and to regulate the cellular metabolism. Nutrients such as sugars (mostly glucose), lipids and amino acids are oxidized to primarily produce the energy [47]. Approximately 90% of cellular energy requirements are generated in mitochondria [48]. Glucoses and lipids are glycolysed into pyruvic and fatty acids, respectively, in the cytoplasm. Subsequently, these acids form acetyl coenzyme A (Acetyl CoA) via a series of catabolic reactions and then enter the TCA cycle in the mitochondrial matrix [49]. In the reaction of the TCA cycle, convert three equivalents of nicotinamide adenine dinucleotide (NAD+) into three equivalents of reduced NAD+ (NADH), one equivalent of flavin adenine dinucleotide (FAD) into one equivalent of FADH2 and one equivalent each of guanosine diphosphate (GDP) and inorganic phosphate (Pi) into one equivalent of guanosine triphosphate (GTP). The NADH and FADH2 are, in turn, used by the OXPHOS to generate ATPs. Thus, oxidation of nutrients provides electrons to the mitochondrial ETC in the form of NADH and FADH2. The sequential transport of electrons from complex I or II to III and IV extrudes protons from the matrix to the IMS, generating an electrochemical gradient. In this process, the ETC requires two electron carriers: coenzyme Q 10 (CoQ 10, also known as ubiquinone) and cytochrome c (Cytc) [50]. Along this electrochemical gradient, the protons flow through complex V (ATP synthase) on the IMM to return to the mitochondrial matrix. This reflux alters the conformation of complex V and drives the synthesis of ATP from ADP and Pi [47].

2.4 Mitochondrial dynamics (fission and fusion)

Mitochondria are highly dynamic and interacting organelles. Mitochondria are able to autonomously integrate (fusion) and divide (fission) by remodelling their morphology and moving along cytoskeletal tracks, in response to their metabolic or pathogenic conditions and cellular environment [29]. The mitochondrial lengths and networks are determined by the balance between fission and fusion rates [51]. Mitochondrial fission and fusion processes are mainly mediated by large guanosine triphosphatases (GTPases) in the dynamin family [51].

Mitochondrial fission requires the dividing of mitochondrial proteins and mtDNA thus that each daughter organelle normally functions without significant loss of soluble proteins from the mitochondrial matrix or intermembrane space [52]. Fission is required for the cellular distribution of mitochondria during cell division and embryonic growth [53]. Exceeded mitochondrial fission, not mutually balanced with fusion, leads to glucose oxidation, MMP decrease and hence the downregulation of ATP production [54]. The fission process is coordinated by a set of components in the cytosol, cytoskeleton, as well as mitochondria. Fission is mediated by a cytosolic dynamin family member, dynamin-related protein 1 (Drp1). Drp1 is recruited from the cytosol to form spirals around mitochondria and, subsequently, constricts the membranes at the fission site to split the mitochondrial cluster [29].

Mitochondrial fusion is mediated by a different set of dynamin-related GTPases. Mitochondrial outer membrane fusion is coordinated with inner membrane fusion. Three large GTPases are essential for mitochondrial fusion [55]. The mitofusins (Mfn1 and Mfn2) are transmembrane GTPases embedded in the OMM [56]. OPA1 is a dynamin-related GTPase associated with the IMM or IMS. Mitofusins and OPA1 physically interact to mechanistically mediate OMM and IMM fusion, respectively [29, 57, 58]. Mitochondrial fusion may increase to maximize the fidelity for OXPHOS in cellular energy demands [27].

2.5 Mitochondrial ROS production and antioxidant enzymes

Mitochondria are the major source of ROS generation [9]. In an organism, mitochondria utilize approximately 98% of the total amount of inhaled O2, including 1–2% for ROS generation [59, 60]. Mitochondria actually produce ROS in a number of enzymatic reactions; the vast majority of the free radicals from the mitochondria are formed in the ETC during OXPHOS [61]. In the process of OXPHOS, electron leaks from the ETC combine with O2 molecules to form (O2 ·−). Mitochondrial O2 ·−, primarily generated in complexes I and III, is catalysed by Cu/ZnSOD or Mn SOD to disproportionate into H2O2. Subsequently, H2O2 can be converted to OH· by Fenton reaction. Mitochondrial O2 ·− can also bind with protons to form uncharged HOO· radicals and subsequently react with unsaturated fatty acid of mitochondrial membrane lipids to produce lipid radicals. Mitochondrial nitric oxide (NO) interacts with O2 ·− to form reactive nitrogen oxide species (RNS) such as peroxynitrite (ONOO), which produce cellular dysfunction by S-nitrosylating proteins [62]. Mammalian cells have multiple enzymes to degrade H2O2, including peroxiredoxins (Prxs), glutathione peroxidases (Gpxs), thioredoxins (Trxs) and catalase. Mitochondrial H2O2 is primarily eliminated by the action of Gpx1, Gpx2 and Gpx4, Prx3 and Prx5 and Trx2 systems, which requires glutathione (GSH) [63, 64, 65]. Oxidized GSH (GSSG) is reduced to GSH by glutathione reductase (GR) activity [66]. Similarly, oxidized Trx2 is recycled by Trx reductase (TrxR). These H2O2 scavenging system ultimately depends on reduced nicotinamide adenine dinucleotide phosphate (NADPH) which is regenerated by three mitochondrial matrix-located enzymes: NADP+-linked isocitrate dehydrogenase (IDH), malate dehydrogenase (MDH) and nicotinamide nucleotide transhydrogenase (NNT) [61]. Catalase catalyses the decomposition of hydrogen peroxide to water and oxygen, existing as a tetramer composed of four identical monomers, each of which contains a heme group at the active site. Catalase also requires NADPH as a reducing equivalent to prevent oxidative inactivation of the enzyme [67] (Figure 2).

Figure 2.

Reactions and transformations of mitochondrial ROS. SOD enzymes catalyse the dismutation of superoxide (O2 ·−), generating hydrogen peroxide (H2O2). The catalase (CAT), glutathione peroxidases (Gpxs) and peroxiredoxins (Prxs) convert H2O2 into water. H2O2 reacts with redox-active iron to generate the hydroxyl radical (OH·) through the Fenton reaction. The reaction between O2 ·− and nitric oxide (NO·) produces peroxynitrite (ONOO), whose decomposition in turn gives rise to some highly oxidizing intermediates including NO2 ·, OH·, CO3 · as well as, finally, stable NO3 . Thus, increased O2 ·− levels can also reduce NO· and generate ONOO toxicity. O2 ·− by itself can reduce ferric iron (Fe3+) to ferrous iron (Fe2+) in iron-sulphur centres of proteins, leading to enzyme inactivation and concomitant loss of Fe2+ from the enzymes. The protonation of O2 ·− can form the more reactive hydroperoxyl radical (HO2 ·).

Advertisement

3. Physiological functions of mitochondrial ROS in diverse cellular processes

mROS generation is a ubiquitous phenomenon during life of eukaryotic cells [68]. mROS-induced oxidative stress is considered a main contributor to the aetiology of both normal senescence and severe pathologies. Under normal physiological conditions, mROS emission is accounted for ∼2% of the total O2 consumption, of which the decomposition is well-controlled [2]. Accumulation of mROS, which is an imbalance of neutralization, induces deleterious consequences such as neurodegenerative disease [69], cardiovascular disease [70] and cancers [71]. However, depending on the cellular environment, antioxidant machinery-regulated oxidative stress could initiate diverse cellular responses, involved in cell protection, initiating coordinated activation of mitochondrial fission and autophagy to carry out clearance of abnormal mitochondria and cells, which are to protect spreading the damage to the adjacent cells [72, 73]. H2O2 is the primary molecule of mROS utilized for intracellular signalling, which selectively reacts with cysteine residues in redox-sensitive proteins, altering activities or conformations of the proteins to regulate signal transduction [74, 75, 76]. Mechanistically, H2O2 oxidizes thiol groups (SH) on cysteine residues to form sulphenic acid (SOH), which react with GSH to become glutathionylated (GSSG), with neighbouring thiols to form a disulphide bond (S-S) or with amides to form a sulphenyl amide (S-N) [77, 78]. In this section, we introduce the physiological roles and regulations of mROS in diverse cellular processes such as proliferation, differentiation, autophagy, immunity and aging (Figure 3).

Figure 3.

Physiological regulation by mitochondrial ROS. mROS contribute to the various physiological cellular processes, including proliferation, differentiation autophagy, immunity and aging.

3.1 Proliferation

Accumulation of mitochondria-derived ROS enables to prompt cell proliferation inhibition and cellular senescence [79, 80]. However, the cells essentially utilize mROS for survival and growth via multiple mechanisms in diverse circumstances.

mROS regulate cell proliferation during hypoxia. Under the hypoxic condition (a low O2 environment, generally 0.3–3% of O2), the cells raise transcriptional and non-transcriptional responses to increase O2 supply, simultaneously reducing O2 consumption. These adaptations to hypoxia are enhanced by mROS. The hypoxia-inducible factors (HIFs) such as HIF1, HIF2 and HIF3 orchestrate the transcriptional response to the hypoxia, promoting erythropoietin (EPO) expression to increase erythropoiesis, vascular endothelial growth factor (VEGF) to promote blood vessel formation and glycolysis enzymes to retain ATP levels [81, 82]. HIFs are heterodimers consisting of two basic helix-loop-helix/PAS proteins: a stable β-subunit and one of three unstable labile α-subunits (HIF1α, HIF2α and HIF-3α) [83, 84]. Under normoxic conditions, prolyl hydroxylase domain protein 2 (PHD2) leads to hydroxylation of HIFα at two proline residues, which target via Von Hippel-Lindau (VHL) E3 ubiquitin ligase-dependent proteasomal degradation [85]. However, under the hypoxic condition, HIFα is stabilized, which is then dimerized with HIF-1β and binds HIF-response elements (HRE) to recruit gene transcription [86]. Moreover, mitochondrial DNA lacking ρ° cells are unable to stabilize HIFα proteins under hypoxic condition, which results from failure of mROS production by ETC deficiency. In contrast, MMP reconstitution restores mROS, which leads to HIFα and cell proliferation [87]. Chemical inhibition of mitochondrial ETC also attenuates mROS production in mitochondria-repleted cells, interrupting to stabilize HIFα under hypoxia [88]. Genetic loss of the complex III subunit Rieske iron-sulphur protein (RISP) or Cytc also inhibits mROS production and HIFα stabilization [89, 90, 91]. It is also indicated that mROS are requisite to activate HIFs by non-hypoxic stimulus [92].

mROS are also involved in vascular smooth muscle cell (VSMC) proliferation. Angiotensin II (AngII) is a peptide hormone basically involved in sodium and water homeostasis and vascular contraction, which is also recognized to influence cell growth and proliferation [93]. AngII exerts physiological effects by signalling via interacting with angiotensin type 1 receptors (AT1Rs) [94]. In VSMCs, AngII signalling is required to activate a multitude of mitogenic signalling cascades via crosstalk with growth factor receptors such as epidermal growth factor receptor (EGFR), platelet-derived growth factor receptors (PDGFR) and insulin receptor (IR). Intracellular signalling of VSMC proliferation is stimulated by AngII signalling-triggered mROS production and subsequently induced via mitogenic serine/threonine kinases, including ERK1/2 and p38MAPK [95].

Despite the detrimental effects, mROS function as signal transduction molecules in regulation of stem cells [96]. Depletion of ataxia telangiectasia mutated (ATM) kinase or forkhead box O (FOXO) transcriptional factors increases mROS levels, which impairs hematopoietic stem cell (HSC) proliferation [97, 98, 99]. Although the increased mROS level impairs the differentiation of HSCs, a decreased mROS level also has negative effects for self-renewal in neural and spermatogonia stem cells (SCs) [100, 101].

3.2 Differentiation

mROS function as active signalling molecules for diverse cell differentiation. Stem cells (SCs, embryonic or adult) have potentials to self-renew for maintaining stem cell pool or differentiate to the multicellular organism and supply de novo functional cells to tissues throughout the life of the organism. During differentiation of SCs, the mitochondrial oxidative metabolism is highly stimulated, and thus cellular respiration and mROS production increase [102, 103, 104, 105].

In SCs, generally, mitochondria exhibit immature mitochondrial networks and primitive cristae [106]. As bone marrow-derived human mesenchymal stem cells (MSCs) differentiate to osteoblasts, mitochondrial biosynthesis increases by PGC-1α activation [102]. Mitochondrial mass and oxygen consumption increase during differentiation of human embryonic stem cells (ESCs) [107] or pluripotent stem cells (PSCs) [108]. Knockdown of the complex III protein RISP or mitochondrial-targeted antioxidants inhibited differentiation of human MSCs to adipocytes, indicating that mROS are required for differentiation of MSCs [109]. Furthermore, during differentiation of human PSCs, uncoupling protein 2 (UCP2) expression is repressed, which is required for metabolic transition from glycolysis to mitochondrial glucose oxidation. Knockdown of UCP2 expression facilitates mROS accumulation, which stimulate the PSC differentiation to cardiomyocytes. Ectopic UCP2 expression impairs the differentiation with retardation of mROS accumulation and embryonic body formation [110].

mROS, at least within physiological concentrations, have critical roles in processes of myogenic differentiation and muscle regeneration [111]. mROS could promote mitochondrial biogenesis, which is an essential molecule in myogenic differentiation, via peroxisome proliferator-activated receptor gamma coactivator 1 (PGC1)-activated signalling pathway [112]. Myogenic cells are armed with antioxidant enzymes such as SODs, catalase, Gpxs, Prxs, γ-glutamylcysteine synthetase (γGCS) and heme oxygenase-1 (HO-1) [113, 114, 115, 116, 117, 118, 119, 120]. These antioxidant enzymes could play as critical signalling molecules to maintain muscle homeostasis in company with primarily neutralizing excessive ROS [121]. mROS facilitate myoblast differentiation and hypertrophy via insulin growth factor 1 (IGF1) signalling pathway [111], which enhances phosphorylation of IGF1 receptor (IGF1R) [122]. Mitochondrial complex I-derived H2O2 acts as a signalling molecule to induce cardiac myogenic differentiation. Chemical inhibition of the complex I and treatment of mitochondrial-specific antioxidant exhibits reduction in mROS production and thus impairs the myoblast differentiation [123]. Moreover, mROS induced phosphatase and tensin homolog (PTEN) oxidative inactivation and thereby stimulated phosphoinositide 3-kinase (PI3K)-AKT signalling pathway to express myogenic genes during skeletal myoblast differentiation and muscle regeneration [124]. In differentiation of VSMCs, mROS production also elevates to activate p38 MAPK signalling pathway [125]. However, the complexity of mROS involvement still requires further investigation to elucidate the certain roles of oxidative stress in myogenic differentiation and muscle regeneration.

3.3 Autophagy

Autophagy is a conserved catabolic process that controls cellular degradation of unnecessary or dysfunctional cellular components in the lysosome [126]. Generally, the autophagy continuously occurs to recycle damaged proteins and organelles for cellular homeostasis under normal conditions [127]. The autophagy has at least three different types: (1) Macroautophagy (usually referred to as autophagy): cytosolic contents are delivered to the lysosome by autophagosomes. (2) Microautophagy: the contents are directly introduced into lysosomal membrane. (3) Chaperone-mediated autophagy: the target proteins contain a motif KFERQ , and then the chaperone (KFERQ)-protein complex binds lysosome-associated membrane protein 2A (LAMP2A) receptors on the lysosomal membrane [128]. Autophagy induction results in recruitment of autophagy-related proteins (ATGs) to a punctate structure, phagophore assembly site (PAS), where proteins of the uncoordinated-51-like kinase 1 (ULK1) complex assemble to initiate autophagosome formation [129].

In autophagy signalling, mitochondria are considered as main source of ROS [130]. mROS, especially as H2O2, are required for autophagy induction in response to nutrient starvation and rapamycin, tumour necrosis factor α (TNFα) and nerve growth factor (NGF) deprivation [131, 132, 133, 134]. H2O2 modulates the cysteine protease Atg4, which cleaves c-terminus of Atg8 (or light chain 3, LC3), and thus enables the addition of phosphatidylethanolamine (PE) to Atg8. Subsequently the active Atg8 is conjugated on the autophagosomal membrane, leading to the autophagosome formation [131]. H2O2 also disrupts the MMP to inhibit Akt/mammalian target of rapamycin (mTOR) signalling pathway for autophagy initiation [135, 136]. Furthermore, elevated H2O2 induces autophagy via activation of p38 MAPK signalling pathway in cardiac or skeletal muscle [137, 138].

In physiological energy metabolism, mitochondrial ATP production by OXPHOS induces mROS generation, resulting in a certain degree of constitutive mitochondrial damage and submitochondrial particles. The damaged mitochondria cause ATP depletion and Cytc release, which eventually leads to activation of caspases and then onset of apoptosis [139, 140]. To prevent cell death, the dysfunctional mitochondria are thus sequestered from the mitochondrial network and eliminated by selective autophagy, mitophagy, to properly maintain mitochondrial quantity and quality [130]. Therefore, mitophagy limits further mROS generation, which promotes turnover of mitochondria and avoids accumulation of dysfunctional mitochondria. Mitophagy is mainly controlled by the PTEN-induced kinase 1 (PINK1)-Parkin pathway, which is stimulated upon the MMP depolarization. PINK1 is a Ser/Thr kinase that translocates on the outer mitochondrial membrane, which is stabilized by low MMP, thereby sensing mitochondrial depolarization [141, 142, 143]. Then, PINK1 recruits Parkin that ubiquitylates OMM-located proteins such as VDAC1, resulting in recruitment of autophagic machinery and the selective sequestration of ubiquitylated mitochondria within autophagosomes [130]. Furthermore, the mitochondrial proteins, BCL2/adenovirus E1B 19-kDa-interacting protein 3 (Bnip3) and Bcl-2/adenovirus E1B 19-kDa-interacting protein 3 (Bnip3L/NIX), participate in mitophagy [144]. In response to oxidative stress after ischemia/reperfusion (I/R), Bnip3 is homodimerised, to be activated, resulting in induction of mitophagy [145]. NIX, an atypical BH3 protein, is required for mitophagy in erythrocyte development. It directly recognizes autophagosome-sited GABA receptor-associated protein (GABARAP) that is a functional homolog of LC3 and subsequently induces mitophagy [126, 146]. Bnip3 and NIX directly bind to the autophagy machinery components, differently to PINK1 or Parkin [147]. ULK1 also regulates mitophagy via translocation to mitochondria to phosphorylate FUN14 domain containing 1 (FUNDC1) protein, a mitochondrial outer membrane protein, which is a receptor for hypoxia-induced mitophagy [148].

3.4 Immunity

In immune system, it is well known that ROS contribute to directly eliminate pathogens via the oxidative burst mediated by NADPH oxidases (NOXs) that are plasma membrane-bound enzyme complexes in phagosomes. However, intracellular redox state intervened by mROS has emerged to be essential for innate and adaptive immune responses [149, 150].

mROS are crucial for Toll-like receptor (TLR) signalling pathways [19]. Activation of cell surface TLRs such as TLR1, TLR2 and TLR4 increases in mROS production via TNF receptor-associated factor 6 (TRAF) and evolutionary conserved signalling intermediate in Toll pathways (ECSIT) signalling pathway [151]. The TRAF6 or ECSIT depletion promotes reduction of mROS generation in macrophages and thus impairment of bacterial clearance [151]. Lipopolysaccharide (LPS)-induced pro-inflammatory cytokines such as TNFα and IL-6 are controlled by mROS generation [152]. Innate immune response enhancement in patients with TNF receptor-associated periodic syndrome (TRAPS) that is an autoinflammatory disorder is affected by missense mutations in the type 1 TNF receptor (TNFR1), which might be attributable to mitochondrial ROS generation [152].

mROS control pattern recognition receptors (PRRs) such as nuclear oligomerization domain (NOD)-like receptors (NLRs). NLRs form multisubunit protein complexes termed inflammasomes that activate caspase-1 resulting in proteolytic cleavage and pro-inflammatory cytokine IL-1β maturation [153, 154]. Pathogen-associated molecular patterns (PAMPs) and damage-associated molecular patterns (DAMPs) such as lipopolysaccharide (LPS), asbestos, ATP and uric acid activate NLR family pyrin domain containing 3 (NLRP3) inflammasome via mROS generation [155, 156]. Pharmacologic or genetic inhibition of autophagy elevates mROS concentration, which heightens inflammasome activation [157, 158]. Increase of mROS persuades lysosomal membrane permeabilization, which is required for NLRP3 activation [159]. Activation of NLRP3 inflammasome results in mitochondrial damage, interrupting mitophagic signalling [160]. Notably, calcium influx contributes to mitochondrial damage, which might increase mROS production and mtDNA release to amplify NLRP3 inflammasome activation [161, 162]. However, it remains to be further delineated how PAMPs and DAMPs increase mROS to properly activate NLRP3 inflammasome.

In adaptive immune responses, T cells are functionally crucial in response to the pathogens [150, 156]. In infectious condition, naïve T cells promptly proliferate and differentiate into effector T cells [163]. The activation of T cells requires increase in glycolysis and mitochondrial metabolism for synthesis of macromolecules in process of the proliferation and differentiation [156, 164, 165]. Elevated mROS concentration contributes to the T-cell activation; treatment of antioxidants inhibits cellular proliferation and interleukin-2 (IL-2) production [166]. Similarly, antioxidant administration to mice exhibits their reduced immunity after infection of the virus, suggesting that mROS are indispensable for the T-cell functions in vivo [167, 168]. The T-cell receptor (TCR) stimulation induces mROS production from complex I, which leads to activation of NF-κB and AP1 signalling, and in turn facilitates IL-2 and IL-4 productions that are imperative drivers in T-cell activation [169, 170].

3.5 Aging

Aging is a process that is concomitant with the accumulation of cellular damage over the time of all living organisms. In the 1950s, Denham Harman suggested the ‘free radical theory of aging’ as a molecular explanation for aging [171], in which free radicals, as byproducts of energy metabolism, develop cumulative cellular damage resulting in loss of organismal ability over time. The theory has been revised that the mitochondria-derived free radicals are causative of aging [172]. Mitochondrial dysfunction and consequent excessive ROS production result in inevitable cellular damage and subsequent cell death [173]. Oxidative damage to genomes, proteins and lipids has been associated with mitochondrial dysfunction and ultimately cellular senescence or death [174]. Consistently, overexpression of antioxidant enzymes reduces ROS production and subsequently protects DNA, which is interconnected to a prolonged life span in Drosophila melanogaster [175, 176].

Despite numerous evidences underpinning the detrimental roles of mROS in aging, the discoveries are questioning a direct correlation between oxidative stress and the lifespan. A mitochondrial enzyme, doublecortin-like kinase 1 (MCLK1), reduction induces mitochondrial dysfunction that displays the regression of electron transport in mitochondrial respiratory chain and decline of TCA cycle activity [177]. In Drosophila melanogaster, mROS levels elevate along with age, but do not intervene with life span [178]. Furthermore, moderate ROS levels have been associated with an extension of longevity in Drosophila melanogaster and in young mice [179, 180, 181]. Therefore, physiologically controlled mROS might activate adaptive responses that are beneficial to the organism and extend life span.

Advertisement

4. Conclusion

Mitochondria are primary energy producers to generate ATPs via oxidative phosphorylation. For a long time, mROS have been considered as byproducts of biological energy metabolism during the ATP generation or by cellular redox system imbalance, which are highly aggressive and detrimental to the neighbouring cells and tissues. However, the roles of mROS have been extensively substantiated to understand normal physiology and pathology over the past decades. Mitochondria-derived H2O2 have been unequivocally recognized as essential molecules in a range of physiological processes in cells.

In this chapter, we have provided a brief discussion of current understanding of physiological roles of mROS by which mitochondria indeed contribute to the implementation of cellular proliferation, differentiation autophagy, innate and adaptive immunity and aging. In understanding the mechanisms regulating mitochondrial physiology and homeostasis, mROS production might provide a significant potential for the development of novel therapeutic strategy for the treatment of a wide range of human pathologies.

Advertisement

Acknowledgments

This work was supported by the National Research Foundation of Korea (NRF) grant funded by the Korean government (MEST) (NRF-2013R1A1A2061214, NRF-2018R1D1A1B07048909 and NRF-2018R1A6A1A03025124).

Advertisement

Conflict of interest

The authors declare no conflict of interest.

References

  1. 1. Henze K, Martin W. Evolutionary biology: Essence of mitochondria. Nature. 2003;426:127-128
  2. 2. Zorov DB, Juhaszova M, Sollott SJ. Mitochondrial reactive oxygen species (ROS) and ROS-induced ROS release. Physiological Reviews. 2014;94:909-950
  3. 3. Brand MD, Nicholls DG. Assessing mitochondrial dysfunction in cells. The Biochemical Journal. 2011;435:297-312
  4. 4. Balaban RS, Nemoto S, Finkel T. Mitochondria, oxidants, and aging. Cell. 2005;120:483-495
  5. 5. Brand MD. Mitochondrial generation of superoxide and hydrogen peroxide as the source of mitochondrial redox signaling. Free Radical Biology & Medicine. 2016;100:14-31
  6. 6. Thannickal VJ, Fanburg BL. Reactive oxygen species in cell signaling. American Journal of Physiology. Lung Cellular and Molecular Physiology. 2000;279:L1005-L1028
  7. 7. Li X, Fang P, Mai J, Choi ET, Wang H, Yang XF. Targeting mitochondrial reactive oxygen species as novel therapy for inflammatory diseases and cancers. Journal of Hematology & Oncology. 2013;6:19
  8. 8. Hayyan M, Hashim MA, AlNashef IM. Superoxide ion: Generation and chemical implications. Chemical Reviews. 2016;116:3029-3085
  9. 9. Murphy MP. How mitochondria produce reactive oxygen species. The Biochemical Journal. 2009;417:1-13
  10. 10. Finkel T. Signal transduction by reactive oxygen species. The Journal of Cell Biology. 2011;194:7-15
  11. 11. Chandel NS. Mitochondrial regulation of oxygen sensing. Advances in Experimental Medicine and Biology. 2010;661:339-354
  12. 12. Antico Arciuch VG, Elguero ME, Poderoso JJ, Carreras MC. Mitochondrial regulation of cell cycle and proliferation. Antioxidants & Redox Signaling. 2012;16:1150-1180
  13. 13. Rambold AS, Pearce EL. Mitochondrial dynamics at the Interface of immune cell metabolism and function. Trends in Immunology. 2018;39:6-18
  14. 14. Gunter TE, Yule DI, Gunter KK, Eliseev RA, Salter JD. Calcium and mitochondria. FEBS Letters. 2004;567:96-102
  15. 15. Nikoletopoulou V, Markaki M, Palikaras K, Tavernarakis N. Crosstalk between apoptosis, necrosis and autophagy. Biochimica et Biophysica Acta. 2013;1833:3448-3459
  16. 16. Kamer KJ, Mootha VK. The molecular era of the mitochondrial calcium uniporter. Nature Reviews. Molecular Cell Biology. 2015;16:545-553
  17. 17. Droge W. Free radicals in the physiological control of cell function. Physiological Reviews. 2002;82:47-95
  18. 18. Finkel T. Oxygen radicals and signaling. Current Opinion in Cell Biology. 1998;10:248-253
  19. 19. Sena LA, Chandel NS. Physiological roles of mitochondrial reactive oxygen species. Molecular Cell. 2012;48:158-167
  20. 20. Schieber M, Chandel NS. ROS function in redox signaling and oxidative stress. Current Biology. 2014;24:R453-R462
  21. 21. McCommis KS, Finck BN. Mitochondrial pyruvate transport: A historical perspective and future research directions. The Biochemical Journal. 2015;466:443-454
  22. 22. Friedman JR, Nunnari J. Mitochondrial form and function. Nature. 2014;505:335-343
  23. 23. Lane N, Martin W. The energetics of genome complexity. Nature. 2010;467:929-934
  24. 24. Burki F. Mitochondrial evolution: Going, going, gone. Current Biology. 2016;26:R410-R412
  25. 25. Gabaldon T, Huynen MA. Shaping the mitochondrial proteome. Biochimica et Biophysica Acta. 2004;1659:212-220
  26. 26. Taanman JW. The mitochondrial genome: Structure, transcription, translation and replication. Biochimica et Biophysica Acta. 1999;1410:103-123
  27. 27. Youle RJ, van der Bliek AM. Mitochondrial fission, fusion, and stress. Science. 2012;337:1062-1065
  28. 28. Lackner LL. Shaping the dynamic mitochondrial network. BMC Biology. 2014;12:35
  29. 29. Pernas L, Scorrano L. Mito-morphosis: Mitochondrial fusion fission, and cristae remodeling as key mediators of cellular function. Annual Review of Physiology. 2016;78:505-531
  30. 30. Jap BK, Walian PJ. Structure and functional mechanism of porins. Physiological Reviews. 1996;76:1073-1088
  31. 31. Hoogenboom BW, Suda K, Engel A, Fotiadis D. The supramolecular assemblies of voltage-dependent anion channels in the native membrane. Journal of Molecular Biology. 2007;370:246-255
  32. 32. Zeth K. Structure and evolution of mitochondrial outer membrane proteins of beta-barrel topology. Biochimica et Biophysica Acta. 2010;1797:1292-1299
  33. 33. Mannella CA. Structure and dynamics of the mitochondrial inner membrane cristae. Biochimica et Biophysica Acta. 2006;1763:542-548
  34. 34. Wohlrab H. Transport proteins (carriers) of mitochondria. IUBMB Life. 2009;61:40-46
  35. 35. Herrmann JM, Riemer J. The intermembrane space of mitochondria. Antioxidants & Redox Signaling. 2010;13:1341-1358
  36. 36. O’Rourke B. Mitochondrial ion channels. Annual Review of Physiology. 2007;69:19-49
  37. 37. Backes S, Herrmann JM. Protein translocation into the intermembrane space and matrix of mitochondria: Mechanisms and driving forces. Frontiers in Molecular Biosciences. 2017;4:83
  38. 38. Picard M, Taivassalo T, Gouspillou G, Hepple RT. Mitochondria: Isolation, structure and function. The Journal of Physiology. 2011;589:4413-4421
  39. 39. Boore JL. Animal mitochondrial genomes. Nucleic Acids Research. 1999;27:1767-1780
  40. 40. Blanco FJ, Valdes AM, Rego-Perez I. Mitochondrial DNA variation and the pathogenesis of osteoarthritis phenotypes. Nature Reviews Rheumatology. 2018;14:327-340
  41. 41. Anderson S, Bankier AT, Barrell BG, de Bruijn MH, Coulson AR, Drouin J, et al. Sequence and organization of the human mitochondrial genome. Nature. 1981;290:457-465
  42. 42. Wallace DC. Maternal genes: Mitochondrial diseases. Birth Defects Original Article Series. 1987;23:137-190
  43. 43. Dautant A, Meier T, Hahn A, Tribouillard-Tanvier D, di Rago JP, Kucharczyk R. ATP synthase diseases of mitochondrial genetic origin. Frontiers in Physiology. 2018;9:329
  44. 44. Park CB, Larsson NG. Mitochondrial DNA mutations in disease and aging. The Journal of Cell Biology. 2011;193:809-818
  45. 45. Madamanchi NR, Runge MS. Mitochondrial dysfunction in atherosclerosis. Circulation Research. 2007;100:460-473
  46. 46. Terzioglu M, Larsson NG. Mitochondrial dysfunction in mammalian ageing. Novartis Foundation Symposium. 2007;287:197-208; discussion 208-113
  47. 47. Liesa M, Shirihai OS. Mitochondrial dynamics in the regulation of nutrient utilization and energy expenditure. Cell Metabolism. 2013;17:491-506
  48. 48. Herrera AS, Del CAEM, Md Ashraf G, Zamyatnin AA, Aliev G. Beyond mitochondria, what would be the energy source of the cell? Central Nervous System Agents in Medicinal Chemistry. 2015;15:32-41
  49. 49. Anderson NM, Mucka P, Kern JG, Feng H. The emerging role and targetability of the TCA cycle in cancer metabolism. Protein & Cell. 2018;9:216-237
  50. 50. Cogliati S, Enriquez JA, Scorrano L. Mitochondrial cristae: Where beauty meets functionality. Trends in Biochemical Sciences. 2016;41:261-273
  51. 51. van der Bliek AM, Shen Q , Kawajiri S. Mechanisms of mitochondrial fission and fusion. Cold Spring Harbor Perspectives in Biology. 2013;5:30-33
  52. 52. Bartolak-Suki E, Imsirovic J, Nishibori Y, Krishnan R, Suki B. Regulation of mitochondrial structure and dynamics by the cytoskeleton and mechanical factors. International Journal of Molecular Sciences. 2017;18:34-36
  53. 53. Yaffe MP. The machinery of mitochondrial inheritance and behavior. Science. 1999;283:1493-1497
  54. 54. Bach D, Pich S, Soriano FX, Vega N, Baumgartner B, Oriola J, et al. Mitofusin-2 determines mitochondrial network architecture and mitochondrial metabolism. A novel regulatory mechanism altered in obesity. The Journal of Biological Chemistry. 2003;278:17190-17197
  55. 55. Scott I, Youle RJ. Mitochondrial fission and fusion. Essays in Biochemistry. 2010;47:85-98
  56. 56. Santel A, Fuller MT. Control of mitochondrial morphology by a human mitofusin. Journal of Cell Science. 2001;114:867-874
  57. 57. Chan DC. Fusion and fission: Interlinked processes critical for mitochondrial health. Annual Review of Genetics. 2012;46:265-287
  58. 58. Westermann B. Mitochondrial fusion and fission in cell life and death. Nature Reviews. Molecular Cell Biology. 2010;11:872-884
  59. 59. Cadenas E, Davies KJ. Mitochondrial free radical generation, oxidative stress, and aging. Free Radical Biology & Medicine. 2000;29:222-230
  60. 60. Kausar S, Wang F, Cui H. The role of mitochondria in reactive oxygen species generation and its implications for neurodegenerative diseases. Cells. 2018;7:5-7
  61. 61. Starkov AA. The role of mitochondria in reactive oxygen species metabolism and signaling. Annals of the New York Academy of Sciences. 2008;1147:37-52
  62. 62. Kalogeris T, Bao Y, Korthuis RJ. Mitochondrial reactive oxygen species: A double edged sword in ischemia/reperfusion vs preconditioning. Redox Biology. 2014;2:702-714
  63. 63. Cox AG, Winterbourn CC, Hampton MB. Mitochondrial peroxiredoxin involvement in antioxidant defence and redox signalling. The Biochemical Journal. 2009;425:313-325
  64. 64. Andreyev AY, Kushnareva YE, Murphy AN, Starkov AA. Mitochondrial ROS metabolism: 10 years later. Biochemistry (Mosc). 2015;80:517-531
  65. 65. Liemburg-Apers DC, Willems PH, Koopman WJ, Grefte S. Interactions between mitochondrial reactive oxygen species and cellular glucose metabolism. Archives of Toxicology. 2015;89:1209-1226
  66. 66. Birben E, Sahiner UM, Sackesen C, Erzurum S, Kalayci O. Oxidative stress and antioxidant defense. World Allergy Organization Journal. 2012;5:9-19
  67. 67. Kirkman HN, Rolfo M, Ferraris AM, Gaetani GF. Mechanisms of protection of catalase by NADPH. Kinetics and stoichiometry. The Journal of Biological Chemistry. 1999;274:13908-13914
  68. 68. Stairs CW, Leger MM, Roger AJ. Diversity and origins of anaerobic metabolism in mitochondria and related organelles. Philosophical Transactions of the Royal Society of London. Series B, Biological Sciences. 2015;370:20140326
  69. 69. Angelova PR, Abramov AY. Role of mitochondrial ROS in the brain: From physiology to neurodegeneration. FEBS Letters. 2018;592:692-702
  70. 70. Kornfeld OS, Hwang S, Disatnik MH, Chen CH, Qvit N, Mochly-Rosen D. Mitochondrial reactive oxygen species at the heart of the matter: New therapeutic approaches for cardiovascular diseases. Circulation Research. 2015;116:1783-1799
  71. 71. Nazarewicz RR, Dikalova A, Bikineyeva A, Ivanov S, Kirilyuk IA, Grigor’ev IA, et al. Does scavenging of mitochondrial superoxide attenuate cancer prosurvival signaling pathways? Antioxidants & Redox Signaling. 2013;19:344-349
  72. 72. Zorov DB, Bannikova SY, Belousov VV, Vyssokikh MY, Zorova LD, Isaev NK, et al. Reactive oxygen and nitrogen species: Friends or foes? Biochemistry (Mosc). 2005;70:215-221
  73. 73. Runkel ED, Baumeister R, Schulze E. Mitochondrial stress: Balancing friend and foe. Experimental Gerontology. 2014;56:194-201
  74. 74. Janssen-Heininger YM, Mossman BT, Heintz NH, Forman HJ, Kalyanaraman B, Finkel T, et al. Redox-based regulation of signal transduction: Principles, pitfalls, and promises. Free Radical Biology & Medicine. 2008;45:1-17
  75. 75. D’Autreaux B, Toledano MB. ROS as signalling molecules: Mechanisms that generate specificity in ROS homeostasis. Nature Reviews. Molecular Cell Biology. 2007;8:813-824
  76. 76. Diebold L, Chandel NS. Mitochondrial ROS regulation of proliferating cells. Free Radical Biology & Medicine. 2016;100:86-93
  77. 77. Brandes N, Schmitt S, Jakob U. Thiol-based redox switches in eukaryotic proteins. Antioxidants & Redox Signaling. 2009;11:997-1014
  78. 78. Finkel T. From sulfenylation to sulfhydration: What a thiolate needs to tolerate. Science Signaling. 2012;5:pe10
  79. 79. Nagiec EE, Wu L, Swaney SM, Chosay JG, Ross DE, Brieland JK, et al. Oxazolidinones inhibit cellular proliferation via inhibition of mitochondrial protein synthesis. Antimicrobial Agents and Chemotherapy. 2005;49:3896-3902
  80. 80. Onyango IG, Khan SM, Bennett JP Jr. Mitochondria in the pathophysiology of Alzheimer’s and Parkinson’s diseases. Frontiers in Bioscience (Landmark Ed). 2017;22:854-872
  81. 81. Haase VH. Hypoxic regulation of erythropoiesis and iron metabolism. American Journal of Physiology. Renal Physiology. 2010;299:F1-F13
  82. 82. Beleslin-Cokic BB, Cokic VP, Yu X, Weksler BB, Schechter AN, Noguchi CT. Erythropoietin and hypoxia stimulate erythropoietin receptor and nitric oxide production by endothelial cells. Blood. 2004;104:2073-2080
  83. 83. Weidemann A, Johnson RS. Biology of HIF-1alpha. Cell Death and Differentiation. 2008;15:621-627
  84. 84. Kaelin WG Jr, Ratcliffe PJ. Oxygen sensing by metazoans: The central role of the HIF hydroxylase pathway. Molecular Cell. 2008;30:393-402
  85. 85. Semenza GL. Hypoxia-inducible factor 1 (HIF-1) pathway. Science’s STKE. 2007;2007:cm8
  86. 86. Semenza GL. Hypoxia-inducible factors in physiology and medicine. Cell. 2012;148:399-408
  87. 87. Martinez-Reyes I, Diebold LP, Kong H, Schieber M, Huang H, Hensley CT, et al. TCA cycle and mitochondrial membrane potential are necessary for diverse biological functions. Molecular Cell. 2016;61:199-209
  88. 88. Chandel NS, McClintock DS, Feliciano CE, Wood TM, Melendez JA, Rodriguez AM, et al. Reactive oxygen species generated at mitochondrial complex III stabilize hypoxia-inducible factor-1alpha during hypoxia: A mechanism of O2 sensing. The Journal of Biological Chemistry. 2000;275:25130-25138
  89. 89. Brunelle JK, Bell EL, Quesada NM, Vercauteren K, Tiranti V, Zeviani M, et al. Oxygen sensing requires mitochondrial ROS but not oxidative phosphorylation. Cell Metabolism. 2005;1:409-414
  90. 90. Guzy RD, Hoyos B, Robin E, Chen H, Liu L, Mansfield KD, et al. Mitochondrial complex III is required for hypoxia-induced ROS production and cellular oxygen sensing. Cell Metabolism. 2005;1:401-408
  91. 91. Mansfield KD, Guzy RD, Pan Y, Young RM, Cash TP, Schumacker PT, et al. Mitochondrial dysfunction resulting from loss of cytochrome c impairs cellular oxygen sensing and hypoxic HIF-alpha activation. Cell Metabolism. 2005;1:393-399
  92. 92. Patten DA, Lafleur VN, Robitaille GA, Chan DA, Giaccia AJ, Richard DE. Hypoxia-inducible factor-1 activation in nonhypoxic conditions: The essential role of mitochondrial-derived reactive oxygen species. Molecular Biology of the Cell. 2010;21:3247-3257
  93. 93. Wolf G, Wenzel UO. Angiotensin II and cell cycle regulation. Hypertension. 2004;43:693-698
  94. 94. Kim S, Zingler M, Harrison JK, Scott EW, Cogle CR, Luo D, et al. Angiotensin II regulation of proliferation, differentiation, and engraftment of hematopoietic stem cells. Hypertension. 2016;67:574-584
  95. 95. Mehta PK, Griendling KK. Angiotensin II cell signaling: Physiological and pathological effects in the cardiovascular system. American Journal of Physiology. Cell Physiology. 2007;292:C82-C97
  96. 96. Holmstrom KM, Finkel T. Cellular mechanisms and physiological consequences of redox-dependent signalling. Nature Reviews. Molecular Cell Biology. 2014;15:411-421
  97. 97. Ito K, Hirao A, Arai F, Matsuoka S, Takubo K, Hamaguchi I, et al. Regulation of oxidative stress by ATM is required for self-renewal of haematopoietic stem cells. Nature. 2004;431:997-1002
  98. 98. Miyamoto K, Araki KY, Naka K, Arai F, Takubo K, Yamazaki S, et al. Foxo3a is essential for maintenance of the hematopoietic stem cell pool. Cell Stem Cell. 2007;1:101-112
  99. 99. Tothova Z, Kollipara R, Huntly BJ, Lee BH, Castrillon DH, Cullen DE, et al. FoxOs are critical mediators of hematopoietic stem cell resistance to physiologic oxidative stress. Cell. 2007;128:325-339
  100. 100. Morimoto H, Iwata K, Ogonuki N, Inoue K, Atsuo O, Kanatsu-Shinohara M, et al. ROS are required for mouse spermatogonial stem cell self-renewal. Cell Stem Cell. 2013;12:774-786
  101. 101. Le Belle JE, Orozco NM, Paucar AA, Saxe JP, Mottahedeh J, Pyle AD, et al. Proliferative neural stem cells have high endogenous ROS levels that regulate self-renewal and neurogenesis in a PI3K/Akt-dependant manner. Cell Stem Cell. 2011;8:59-71
  102. 102. Chen CT, Shih YR, Kuo TK, Lee OK, Wei YH. Coordinated changes of mitochondrial biogenesis and antioxidant enzymes during osteogenic differentiation of human mesenchymal stem cells. Stem Cells. 2008;26:960-968
  103. 103. Chung S, Dzeja PP, Faustino RS, Perez-Terzic C, Behfar A, Terzic A. Mitochondrial oxidative metabolism is required for the cardiac differentiation of stem cells. Nature Clinical Practice. Cardiovascular Medicine. 2007;4(Suppl 1):S60-S67
  104. 104. Lonergan T, Brenner C, Bavister B. Differentiation-related changes in mitochondrial properties as indicators of stem cell competence. Journal of Cellular Physiology. 2006;208:149-153
  105. 105. Khacho M, Clark A, Svoboda DS, Azzi J, MacLaurin JG, Meghaizel C, et al. Mitochondrial dynamics impacts stem cell identity and fate decisions by regulating a nuclear transcriptional program. Cell Stem Cell. 2016;19:232-247
  106. 106. Papa L, Djedaini M, Hoffman R. Mitochondrial role in stemness and differentiation of hematopoietic stem cells. Stem Cells International. 2019;2019:4067162
  107. 107. Cho YM, Kwon S, Pak YK, Seol HW, Choi YM, Park DJ, et al. Dynamic changes in mitochondrial biogenesis and antioxidant enzymes during the spontaneous differentiation of human embryonic stem cells. Biochemical and Biophysical Research Communications. 2006;348:1472-1478
  108. 108. Prigione A, Adjaye J. Modulation of mitochondrial biogenesis and bioenergetic metabolism upon in vitro and in vivo differentiation of human ES and iPS cells. The International Journal of Developmental Biology. 2010;54:1729-1741
  109. 109. Tormos KV, Anso E, Hamanaka RB, Eisenbart J, Joseph J, Kalyanaraman B, et al. Mitochondrial complex III ROS regulate adipocyte differentiation. Cell Metabolism. 2011;14:537-544
  110. 110. Zhang J, Khvorostov I, Hong JS, Oktay Y, Vergnes L, Nuebel E, et al. UCP2 regulates energy metabolism and differentiation potential of human pluripotent stem cells. The EMBO Journal. 2011;30:4860-4873
  111. 111. Barbieri E, Sestili P. Reactive oxygen species in skeletal muscle signaling. Journal of Signal Transduction. 2012;2012:982794
  112. 112. Adhihetty PJ, Irrcher I, Joseph AM, Ljubicic V, Hood DA. Plasticity of skeletal muscle mitochondria in response to contractile activity. Experimental Physiology. 2003;88:99-107
  113. 113. Powers SK, Criswell D, Lawler J, Ji LL, Martin D, Herb RA, et al. Influence of exercise and fiber type on antioxidant enzyme activity in rat skeletal muscle. The American Journal of Physiology. 1994;266:R375-R380
  114. 114. Manabe Y, Takagi M, Nakamura-Yamada M, Goto-Inoue N, Taoka M, Isobe T, et al. Redox proteins are constitutively secreted by skeletal muscle. The Journal of Physiological Sciences. 2014;64:401-409
  115. 115. Tidball JG. Mechanisms of muscle injury, repair, and regeneration. Comprehensive Physiology. 2011;1:2029-2062
  116. 116. El Haddad M, Jean E, Turki A, Hugon G, Vernus B, Bonnieu A, et al. Glutathione peroxidase 3, a new retinoid target gene, is crucial for human skeletal muscle precursor cell survival. Journal of Cell Science. 2012;125:6147-6156
  117. 117. Kozakowska M, Ciesla M, Stefanska A, Skrzypek K, Was H, Jazwa A, et al. Heme oxygenase-1 inhibits myoblast differentiation by targeting myomirs. Antioxidants & Redox Signaling. 2012;16:113-127
  118. 118. Ding Y, Choi KJ, Kim JH, Han X, Piao Y, Jeong JH, et al. Endogenous hydrogen peroxide regulates glutathione redox via nuclear factor erythroid 2-related factor 2 downstream of phosphatidylinositol 3-kinase during muscle differentiation. The American Journal of Pathology. 2008;172:1529-1541
  119. 119. Catani MV, Savini I, Duranti G, Caporossi D, Ceci R, Sabatini S, et al. Nuclear factor kappaB and activating protein 1 are involved in differentiation-related resistance to oxidative stress in skeletal muscle cells. Free Radical Biology & Medicine. 2004;37:1024-1036
  120. 120. Won H, Lim S, Jang M, Kim Y, Rashid MA, Jyothi KR, et al. Peroxiredoxin-2 upregulated by NF-kappaB attenuates oxidative stress during the differentiation of muscle-derived C2C12 cells. Antioxidants & Redox Signaling. 2012;16:245-261
  121. 121. Kozakowska M, Pietraszek-Gremplewicz K, Jozkowicz A, Dulak J. The role of oxidative stress in skeletal muscle injury and regeneration: Focus on antioxidant enzymes. Journal of Muscle Research and Cell Motility. 2015;36:377-393
  122. 122. Handayaningsih A-E, et al. Reactive Oxygen Species Play an Essential Role in IGF-I Signaling and IGF-I-Induced Myocyte Hypertrophy in C2C12 Myocytes. Endocrinology. 2011;152(3). DOI: 10.1210/en.2010-0981
  123. 123. Lee S, Tak E, Lee J, Rashid MA, Murphy MP, Ha J, et al. Mitochondrial H2O2 generated from electron transport chain complex I stimulates muscle differentiation. Cell Research. 2011;21:817-834
  124. 124. Kim JH, Choi TG, Park S, Yun HR, Nguyen NNY, Jo YH, et al. Mitochondrial ROS-derived PTEN oxidation activates PI3K pathway for mTOR-induced myogenic autophagy. Cell Death and Differentiation. 2018;25:1921-1937
  125. 125. Clempus RE, Griendling KK. Reactive oxygen species signaling in vascular smooth muscle cells. Cardiovascular Research. 2006;71:216-225
  126. 126. Filomeni G, De Zio D, Cecconi F. Oxidative stress and autophagy: The clash between damage and metabolic needs. Cell Death and Differentiation. 2015;22:377-388
  127. 127. Lee J, Giordano S, Zhang J. Autophagy, mitochondria and oxidative stress: Cross-talk and redox signalling. The Biochemical Journal. 2012;441:523-540
  128. 128. Parzych KR, Klionsky DJ. An overview of autophagy: Morphology, mechanism, and regulation. Antioxidants & Redox Signaling. 2014;20:460-473
  129. 129. Hurley JH, Young LN. Mechanisms of autophagy initiation. Annual Review of Biochemistry. 2017;86:225-244
  130. 130. Dikic I, Elazar Z. Mechanism and medical implications of mammalian autophagy. Nature Reviews. Molecular Cell Biology. 2018;19:349-364
  131. 131. Scherz-Shouval R, Shvets E, Fass E, Shorer H, Gil L, Elazar Z. Reactive oxygen species are essential for autophagy and specifically regulate the activity of Atg4. The EMBO Journal. 2007;26:1749-1760
  132. 132. Djavaheri-Mergny M, Amelotti M, Mathieu J, Besancon F, Bauvy C, Souquere S, et al. NF-kappaB activation represses tumor necrosis factor-alpha-induced autophagy. The Journal of Biological Chemistry. 2006;281:30373-30382
  133. 133. Kirkland RA, Saavedra GM, Franklin JL. Rapid activation of antioxidant defenses by nerve growth factor suppresses reactive oxygen species during neuronal apoptosis: Evidence for a role in cytochrome c redistribution. The Journal of Neuroscience. 2007;27:11315-11326
  134. 134. Kirkland RA, Adibhatla RM, Hatcher JF, Franklin JL. Loss of cardiolipin and mitochondria during programmed neuronal death: Evidence of a role for lipid peroxidation and autophagy. Neuroscience. 2002;115:587-602
  135. 135. Byun YJ, Kim SK, Kim YM, Chae GT, Jeong SW, Lee SB. Hydrogen peroxide induces autophagic cell death in C6 glioma cells via BNIP3-mediated suppression of the mTOR pathway. Neuroscience Letters. 2009;461:131-135
  136. 136. Zhang H, Kong X, Kang J, Su J, Li Y, Zhong J, et al. Oxidative stress induces parallel autophagy and mitochondria dysfunction in human glioma U251 cells. Toxicological Sciences. 2009;110:376-388
  137. 137. Yuan H, Perry CN, Huang C, Iwai-Kanai E, Carreira RS, Glembotski CC, et al. LPS-induced autophagy is mediated by oxidative signaling in cardiomyocytes and is associated with cytoprotection. American Journal of Physiology. Heart and Circulatory Physiology. 2009;296:H470-H479
  138. 138. McClung JM, Judge AR, Powers SK, Yan Z. p38 MAPK links oxidative stress to autophagy-related gene expression in cachectic muscle wasting. American Journal of Physiology. Cell Physiology. 2010;298:C542-C549
  139. 139. Mao K, Klionsky DJ. Participation of mitochondrial fission during mitophagy. Cell Cycle. 2013;12:3131-3132
  140. 140. Bolisetty S, Jaimes EA. Mitochondria and reactive oxygen species: Physiology and pathophysiology. International Journal of Molecular Sciences. 2013;14:6306-6344
  141. 141. Ordureau A, Sarraf SA, Duda DM, Heo JM, Jedrychowski MP, Sviderskiy VO, et al. Quantitative proteomics reveal a feedforward mechanism for mitochondrial PARKIN translocation and ubiquitin chain synthesis. Molecular Cell. 2014;56:360-375
  142. 142. Shiba-Fukushima K, Arano T, Matsumoto G, Inoshita T, Yoshida S, Ishihama Y, et al. Phosphorylation of mitochondrial polyubiquitin by PINK1 promotes Parkin mitochondrial tethering. PLoS Genetics. 2014;10:e1004861
  143. 143. Lazarou M, Sliter DA, Kane LA, Sarraf SA, Wang C, Burman JL, et al. The ubiquitin kinase PINK1 recruits autophagy receptors to induce mitophagy. Nature. 2015;524:309-314
  144. 144. Ney PA. Mitochondrial autophagy: Origins, significance, and role of BNIP3 and NIX. Biochimica et Biophysica Acta. 2015;1853:2775-2783
  145. 145. Saito T, Sadoshima J. Molecular mechanisms of mitochondrial autophagy/mitophagy in the heart. Circulation Research. 2015;116:1477-1490
  146. 146. Novak I, Kirkin V, McEwan DG, Zhang J, Wild P, Rozenknop A, et al. Nix is a selective autophagy receptor for mitochondrial clearance. EMBO Reports. 2010;11:45-51
  147. 147. Matsuda N. Phospho-ubiquitin: Upending the PINK-Parkin-ubiquitin cascade. Journal of Biochemistry. 2016;159:379-385
  148. 148. Wu W, Tian W, Hu Z, Chen G, Huang L, Li W, et al. ULK1 translocates to mitochondria and phosphorylates FUNDC1 to regulate mitophagy. EMBO Reports. 2014;15:566-575
  149. 149. Chen Y, Zhou Z, Min W. Mitochondria, oxidative stress and innate immunity. Frontiers in Physiology. 2018;9:1487
  150. 150. Yang Y, Bazhin AV, Werner J, Karakhanova S. Reactive oxygen species in the immune system. International Reviews of Immunology. 2013;32:249-270
  151. 151. West AP, Brodsky IE, Rahner C, Woo DK, Erdjument-Bromage H, Tempst P, et al. TLR signalling augments macrophage bactericidal activity through mitochondrial ROS. Nature. 2011;472:476-480
  152. 152. Bulua AC, Simon A, Maddipati R, Pelletier M, Park H, Kim KY, et al. Mitochondrial reactive oxygen species promote production of proinflammatory cytokines and are elevated in TNFR1-associated periodic syndrome (TRAPS). The Journal of Experimental Medicine. 2011;208:519-533
  153. 153. Schroder K, Tschopp J. The inflammasomes. Cell. 2010;140:821-832
  154. 154. Broz P, Dixit VM. Inflammasomes: Mechanism of assembly, regulation and signalling. Nature Reviews. Immunology. 2016;16:407-420
  155. 155. Li F, Xu M, Wang M, Wang L, Wang H, Zhang H, et al. Roles of mitochondrial ROS and NLRP3 inflammasome in multiple ozone-induced lung inflammation and emphysema. Respiratory Research. 2018;19:230
  156. 156. Weinberg SE, Sena LA, Chandel NS. Mitochondria in the regulation of innate and adaptive immunity. Immunity. 2015;42:406-417
  157. 157. Saitoh T, Fujita N, Jang MH, Uematsu S, Yang BG, Satoh T, et al. Loss of the autophagy protein Atg16L1 enhances endotoxin-induced IL-1beta production. Nature. 2008;456:264-268
  158. 158. Liu Q , Zhang D, Hu D, Zhou X, Zhou Y. The role of mitochondria in NLRP3 inflammasome activation. Molecular Immunology. 2018;103:115-124
  159. 159. Heid ME, Keyel PA, Kamga C, Shiva S, Watkins SC, Salter RD. Mitochondrial reactive oxygen species induces NLRP3-dependent lysosomal damage and inflammasome activation. Journal of Immunology. 2013;191:5230-5238
  160. 160. Yu J, Nagasu H, Murakami T, Hoang H, Broderick L, Hoffman HM, et al. Inflammasome activation leads to Caspase-1-dependent mitochondrial damage and block of mitophagy. Proceedings of the National Academy of Sciences of the United States of America. 2014;111:15514-15519
  161. 161. Gurung P, Lukens JR, Kanneganti TD. Mitochondria: Diversity in the regulation of the NLRP3 inflammasome. Trends in Molecular Medicine. 2015;21:193-201
  162. 162. Murakami T, Ockinger J, Yu J, Byles V, McColl A, Hofer AM, et al. Critical role for calcium mobilization in activation of the NLRP3 inflammasome. Proceedings of the National Academy of Sciences of the United States of America. 2012;109:11282-11287
  163. 163. Smith-Garvin JE, Koretzky GA, Jordan MS. T cell activation. Annual Review of Immunology. 2009;27:591-619
  164. 164. Carr EL, Kelman A, Wu GS, Gopaul R, Senkevitch E, Aghvanyan A, et al. Glutamine uptake and metabolism are coordinately regulated by ERK/MAPK during T lymphocyte activation. Journal of Immunology. 2010;185:1037-1044
  165. 165. Sinclair LV, Rolf J, Emslie E, Shi YB, Taylor PM, Cantrell DA. Control of amino-acid transport by antigen receptors coordinates the metabolic reprogramming essential for T cell differentiation. Nature Immunology. 2013;14:500-508
  166. 166. Chaudhri G, Clark IA, Hunt NH, Cowden WB, Ceredig R. Effect of antioxidants on primary alloantigen-induced T cell activation and proliferation. Journal of Immunology. 1986;137:2646-2652
  167. 167. Laniewski NG, Grayson JM. Antioxidant treatment reduces expansion and contraction of antigen-specific CD8+ T cells during primary but not secondary viral infection. Journal of Virology. 2004;78:11246-11257
  168. 168. Sena LA, Li S, Jairaman A, Prakriya M, Ezponda T, Hildeman DA, et al. Mitochondria are required for antigen-specific T cell activation through reactive oxygen species signaling. Immunity. 2013;38:225-236
  169. 169. Gill T, Levine AD. Mitochondria-derived hydrogen peroxide selectively enhances T cell receptor-initiated signal transduction. The Journal of Biological Chemistry. 2013;288:26246-26255
  170. 170. Kaminski MM, Sauer SW, Klemke CD, Suss D, Okun JG, Krammer PH, et al. Mitochondrial reactive oxygen species control T cell activation by regulating IL-2 and IL-4 expression: Mechanism of ciprofloxacin-mediated immunosuppression. Journal of Immunology. 2010;184:4827-4841
  171. 171. Harman D. Aging: A theory based on free radical and radiation chemistry. Journal of Gerontology. 1956;11:298-300
  172. 172. Harman D. The biologic clock: The mitochondria? Journal of the American Geriatrics Society. 1972;20:145-147
  173. 173. Ziegler DV, Wiley CD, Velarde MC. Mitochondrial effectors of cellular senescence: Beyond the free radical theory of aging. Aging Cell. 2015;14:1-7
  174. 174. Bokov A, Chaudhuri A, Richardson A. The role of oxidative damage and stress in aging. Mechanisms of Ageing and Development. 2004;125:811-826
  175. 175. Orr WC, Sohal RS. Extension of life-span by overexpression of superoxide dismutase and catalase in Drosophila melanogaster. Science. 1994;263:1128-1130
  176. 176. Schriner SE, Linford NJ, Martin GM, Treuting P, Ogburn CE, Emond M, et al. Extension of murine life span by overexpression of catalase targeted to mitochondria. Science. 2005;308:1909-1911
  177. 177. Lapointe J, Hekimi S. Early mitochondrial dysfunction in long-lived Mclk1+/− mice. The Journal of Biological Chemistry. 2008;283:26217-26227
  178. 178. Cocheme HM, Quin C, McQuaker SJ, Cabreiro F, Logan A, Prime TA, et al. Measurement of H2O2 within living Drosophila during aging using a ratiometric mass spectrometry probe targeted to the mitochondrial matrix. Cell Metabolism. 2011;13:340-350
  179. 179. Copeland JM, Cho J, Lo T Jr, Hur JH, Bahadorani S, Arabyan T, et al. Extension of Drosophila life span by RNAi of the mitochondrial respiratory chain. Current Biology. 2009;19:1591-1598
  180. 180. Basisty N, Dai DF, Gagnidze A, Gitari L, Fredrickson J, Maina Y, et al. Mitochondrial-targeted catalase is good for the old mouse proteome, but not for the young: ‘Reverse’ antagonistic pleiotropy? Aging Cell. 2016;15:634-645
  181. 181. Csiszar A, Labinskyy N, Perez V, Recchia FA, Podlutsky A, Mukhopadhyay P, et al. Endothelial function and vascular oxidative stress in long-lived GH/IGF-deficient Ames dwarf mice. American Journal of Physiology. Heart and Circulatory Physiology. 2008;295:H1882-H1894

Written By

Tae Gyu Choi and Sung Soo Kim

Submitted: 11 January 2019 Reviewed: 03 July 2019 Published: 13 December 2019