IMR Press / FBL / Volume 28 / Issue 3 / DOI: 10.31083/j.fbl2803061
Open Access Review
Mitochondrial Coenzyme Q Redox Homeostasis and Reactive Oxygen Species Production
Show Less
1 Laboratory of Mitochondrial Biochemistry, Department of Bioenergetics, Adam Mickiewicz University, 61-614 Poznan, Poland
*Correspondence: wieslawa.jarmuszkiewicz@amu.edu.pl (Wieslawa Jarmuszkiewicz)
These authors contributed equally.
Front. Biosci. (Landmark Ed) 2023, 28(3), 61; https://doi.org/10.31083/j.fbl2803061
Submitted: 4 January 2023 | Revised: 22 February 2023 | Accepted: 3 March 2023 | Published: 23 March 2023
(This article belongs to the Special Issue Mitochondrial Biology in Health and Disease)
Copyright: © 2023 The Author(s). Published by IMR Press.
This is an open access article under the CC BY 4.0 license.
Abstract

Mitochondrial coenzyme Q (mtQ) of the inner mitochondrial membrane is a redox active mobile carrier in the respiratory chain that transfers electrons between reducing dehydrogenases and oxidizing pathway(s). mtQ is also involved in mitochondrial reactive oxygen species (mtROS) formation through the mitochondrial respiratory chain. Some mtQ-binding sites related to the respiratory chain can directly form the superoxide anion from semiubiquinone radicals. On the other hand, reduced mtQ (ubiquinol, mtQH2) recycles other antioxidants and directly acts on free radicals, preventing oxidative modifications. The redox state of the mtQ pool is a central bioenergetic patameter that alters in response to changes in mitochondrial function. It reflects mitochondrial bioenergetic activity and mtROS formation level, and thus the oxidative stress associated with the mitochondria. Surprisingly, there are few studies describing a direct relationship between the mtQ redox state and mtROS production under physiological and pathological conditions. Here, we provide a first overview of what is known about the factors affecting mtQ redox homeostasis and its relationship to mtROS production. We have proposed that the level of reduction (the endogenous redox state) of mtQ may be a useful indirect marker to assess total mtROS formation. A higher mtQ reduction level (mtQH2/mtQtotal) indicates greater mtROS formation. The mtQ reduction level, and thus the mtROS formation, depends on the size of the mtQ pool and the activity of the mtQ-reducing and mtQH2-oxidizing pathway(s) of respiratory chain. We focus on a number of physiological and pathophysiological factors affecting the amount of mtQ and thus its redox homeostasis and mtROS production level.

Keywords
mitochondrial coenzyme Q (mtQ)
mtQ redox state
mtQ pool size
mtQ homeostasis
mtQ-reducing and QH2-oxidizing pathways
reactive oxygen species formation
1. Introduction

Mitochondria are multifunctional organelles crucial for cellular energy production and the formation of reactive oxygen species (ROS) [1]. In addition, mitochondria are involved in many other processes critical to cell function and dysfunction, including maintaining metabolic and ionic homeostasis, calcium signaling, cell differentiation and growth, cell cycle control, and cell death.

Mitochondria generate a significant amount of energy in the form of the adenosine 5′-triphosphate (ATP) molecule needed to power the cell’s biochemical reactions [2]. The tricarboxylic acid cycle (the TCA cycle) and oxidative phosphorylation are the mitochondrial stages of cellular respiration leading to the oxidation of reduced carbon compounds. The four electron transport complexes of the respiratory chain (complexes I–IV) and ATP synthase (complex V), which directly catalyzes the formation of ATP, form the oxidative phosphorylation system located in the inner mitochondrial membrane (Fig. 1). The mitochondrial coenzyme Q (mtQ)-reducing pathway of the respiratory chain consists of complex I (NADH:mtQ oxidoreductase) and complex II (succinate oxidoreductase:mtQ), which transfer electrons from reduced nucleotides, nicotinamide adenine dinucleotide (NADH) and dinucleotide adenine (FADH2), respectively. Complex III (mtQH2:cytochrome c oxidoreductase, bc1 complex), cytochrome c, and finally complex IV constitute the mtQH2-oxidizing pathway in the respiratory chain. Electron transport in the respiratory chain and ATP production are coupled by a proton electrochemical gradient formed by proton pumps (complexes I, III and IV) across the inner mitochondrial membrane. In addition to ATP synthesis by ADP phosphorylation, the proton electrochemical gradient can drive proton leak pathways, which do not generate ATP but may protect mitochondria from oxidative damage by decreasing mitochondrial ROS (mtROS) formation [3]. Moreover, the protonmotive force is coupled directly to the transport of several metabolites and calcium ions across the inner mitochondrial membrane and to the nicotinamide nucleotide transhydrogenase that regulates mitochondrial NADPH levels and mitochondrial redox balance.

Fig. 1.

The oxidative phosphorylation system in the inner mitochondrial membrane. Electron (e¯) transfer cofactors: FMN, flavin mononucleotide; FeS, iron-sulfur cluster; heme bH and heme bL of cytochrome b; heme c1 of cytochrome c1; heme c of cytochrome c; mtQ, oxidized mitochondrial coenzyme Q; mtQH2, reduced mtQ, ubiquinol. Qo, mtQH2 binding site of complex III; Qi, mtQ binding site of complex III. The mtQ cycle is marked in the box.

1.1 Structure and Redox States of Q

Coenzyme Q (Q, ubiquinone or 2,3-dimethoxy-5-methyl-6-polyprenyl-1,4-benzoquinone) is a ubiquitous fat-soluble vitamin-like compound distributed in all cell membrane systems, including the plasma membrane and all intracellular membranes, primarily the inner mitochondrial membrane [4, 5, 6]. Q consists of a redox active benzoquinone ring with two methoxy groups, one methyl group, and an isoprenoid side chain with a species-specific length (6 to 10 units) (Fig. 2). For example, ten isoprenoid units (Q10) have been described as the most commonly found Q in humans and nine units (Q9) in rodents (Q6) [7]. Some species have more than one Q isoform, e.g., rodent tissues contain different proportions of Q9 and Q10. In cells, Q is present in three different redox states, i.e., fully oxidized (ubiquinone, Q), intermediate semioxidized (semiubiquinone radical, QH), and fully reduced (ubiquinol, QH2) (Fig. 2) [8, 9, 10]. Thus, Q has the ability to accept and transfer one or two electrons, which is crucial for its role as a one-electron carrier in the mitochondrial electron transport chain because of the iron-sulfur (FeS) clusters, which can only accept one electron at a time [11], and for its role as a powerful antioxidant scavenging free radicals [12, 13].

Fig. 2.

Q redox states. Oxidized form of Q (ubiquinone) can be reduced to ubiquinol (QH2) by two one-electron reactions through semiubiquinone (QH) intermediate, or by one reaction of two electrons. Superoxide (O2¯) and hydrogen peroxide (H2O2) can be produced by the leakage of electrons from the redox group of QH with unpaired electron.

1.2 Q Biosynthesis

Unlike other lipophilic antioxidants such as vitamin A, carotenoids, and vitamin E, which are usually obtained from dietary sources, Q can also be produced endogenously [14]. The synthesis of Q takes place in two main steps, one of which provides the quinone ring, while the other produces the isoprenoid tail (for reviews see [15, 16, 17, 18, 19, 20]). The 4-benzoquinone ring is synthesized by converting tyrosine (or phenylalanine) to 4-hydroxybenzoate, while the polyprenyl tail is provided via the mevalonate pathway, which also synthesizes cholesterol and dolichols, i.e., large isoprenoid chains comprised of multiple isoprene units. The synthesis of Q precursors (4-hydroxybenzoate and polyprenylpyro-phosphate) takes place in the cytosol, and the last stages of Q biosynthesis take place on the matrix side of the inner mitochondrial membrane [21, 22]. The quinone head is attached to a polyprenylated tail by a prenyl transferase (COQ2) which, together with other enzymes, forms a multiprotein Q biosynthetic complex on the matrix side of the inner mitochondrial membrane [7, 21]. After multiple modifications of the aromatic ring, including methylation, decarboxylation, hydroxylation and deamination (catalyzed by COQ3, COQ5, COQ6, and COQ7, respectively), the final Q molecule is formed. Defects in Q synthesis, including the highly regulated mitochondrial biosynthetic complex, can lead to a reduction in mtQ levels and thus to disturbance of mtQ redox homeostasis and oxidative stress [23].

Despite the vast number of publications on both cellular and mitochondrial Q, its antioxidant and bioenergetic functions, role in health and disease, the literature is surprisingly limited regarding the relationship between Q redox state and ROS production in mitochondria under physiological and pathological conditions. In addition, the role of the mtQ redox state (mQH2/mtQ ratio) is poorly understood because determining its value is technically difficult. In this review, we focus on factors affecting mtQ redox homeostasis and its relationship to mtROS production. We first describe the mtQ-reducing and mtQ-oxidizing pathways of the mitochondrial electron transport chain and the factors influencing the redox state or reduction level of the mtQ pool. We then discuss the relationship between mtQ and mtROS formation, with an emphasis on the mtQ-biding sites of mtROS production, the dependence of mtROS formation on mtQ reduction level, and the relationship between mtQ content and mtROS production. Finally, we provided an introduction to the role of reduced mtQ as an antioxidant. In conclusions, we describe the technical progress in determining the redox state of Q in cells, tissues and isolated mitochondria.

2. mtQ as an Electron Carrier in the Mitochondrial Respiratory Chain

mtQ forms the mtQ pool in the lipid phase of the inner mitochondrial membrane. mtQ plays a central role as an electron carrier in the mitochondrial respiratory chain [24, 25, 26]. Dehydrogenases that oxidize respiratory substrates reduce mtQ to mtQH2, and mQH2-oxidizing pathway(s) convert mtQH2 to mtQ (Fig. 1). mtQ regulates the flow of electrons in the respiratory chain and the building of a proton electrochemical gradient (including the mitochondrial membrane potential, mtΔΨ) across the inner mitochondrial membrane [3, 18, 27, 28]. Therefore, mtQ is the only mobile nonprotein electron carrier in the respiratory chain involved in the formation of the protonmotive force and, consequently, the oxidative phosphorylation process leading to ATP synthesis.

2.1 mtQ-Reducing Pathways

mtQ is reduced to mtQH2 by several dehydrogenases that oxidize reducing equivalents derived from various food fuels, i.e., glucose, fats and amino acids [29]. Thus, mtQ is a common reservoir of electrons that contribute to oxidative oxidative phosphorylation, regardless of source.

Complex I and complex II of the mitochondrial respiratory chain reduce mtQ pool with electrons from the reduced nucleotides (NADH and FADH2, respectively) derived mainly from the TCA cycle and thus from the oxidation of major food molecules. In the rotenone-sensitive complex I, two electrons are transferred from NADH to the flavin mononucleotide (FMN) and then, through the iron-sulfur centers to mtQ, which is reduced to mtQH2 [30, 31, 32]. The malonate-sensitive complex II is an enzyme involved in the TCA cycle and oxidizes succinate to fumarate. Complex II transfers two electrons from FADH2 through the iron-sulfur centers to mtQ, thereby also introducing two electrons into the mitochondrial respiratory chain [33, 34, 35].

In addition to the basic dehydrogenase complexes of the respiratory chain, i.e., complex I and complex II, mtQ receives electrons from a number of electron transporting flavoproteins that supply electrons to the mtQ pool [36, 37, 38]. mtQ is reduced by electron transfer flavoprotein (ETF):mtQ oxidoreductase (ETF:mtQ oxidoreductase) that oxidizes FADH2 derived from β-oxidation of fatty acids in the mitochondrial matrix [39, 40]. Electrons are also delivered to the mtQ pool from the oxidation of dihydroorotate to orotate by dihydroorotate dehydrogenase, a key mitochondrial enzyme in pyrimidine nucleotide biosynthesis [41, 42]. Another mtQ-reducing dehydrogenase located on the outer surface of the inner mitochondrial membrane is FAD-linked glycerol-3-phosphate dehydrogenase, which catalyzes the conversion of glycerol-3-phosphate to dihydroxyacetone [43, 44]. The oxidation of glycerol-3-phosphate by mitochondrial glycerol 3-phosphate dehydrogenase is a major pathway for the transfer of cytosolic reducing equivalents from lipid and carbohydrate metabolism to the mitochondrial electron transport chain. In addition, mtQ accepts electrons from FAD-linked proline dehydrogenase, which is involved in the metabolism of amino acids [45, 46]. In the branched mitochondrial respiratory chain of plants, fungi, and some protists, mtQ receives electrons from alternative rotenone-resistant NAD(P)H dehydrogenases located on the outer and inner surfaces of the inner mitochondrial membrane [47, 48]. Among the mtQ-reducing dehydrogenases, only complex I of mitochondrial respiratory chain pumps protons and therefore participates in the generation of the proton electrochemical gradient.

2.2 mtQH2-Oxidizing Pathway(s)

mtQ metabolism depends on mtQH2 reoxidation by the mitochondrial mtQH2-oxidizing pathway(s). In mammalian mitochondria, the cytochrome pathway (complex III, cytochrome c, and finally complex IV) is the only mtQH2-oxidizing pathway in the respiratory chain [2, 27]. The antimycin A- and myxothiazol-sensitive complex III and the cyanide-sensitive complex IV pump protons through the inner mitochondrial membrane to the intermembrane space and therefore participates in the generation of the protonmotive force. In the branched mitochondrial respiratory chain of plants, fungi, and some protists, there is an additional antimycin A- and cyanide-resistant alternative quinol oxidase (AOX) that oxidizes mtQ, bypassing the cytochrome pathway [49, 50]. Transfer of electrons in AOX is not accompanied by proton pumping and thus mitochondrial ATP production.

Complex III (mtQH2:cytochrome c oxidoreductase, cytochrome bc1 complex) of the cytochrome pathway oxidizes the mtQH2 pool directly, catalyzing a series of electron transfer reactions from mtQH2 to cytochrome c, called the mtQ cycle (Fig. 1, marked box) [51, 52, 53]. Thus, complex III connects two mobile electron carriers of the respiratory chain, mtQ and cytochrome c. The reducing force of the mtQH2 pool and the oxidizing force of the cytochrome c pool influence the redox state of the four metal redox centers in complex III: hemes bH and bL in cytochrome b, and heme c1 and the Rieske FeS center in cytochrome c1. mtQ reduction and mtQH2 oxidation reactions take place at two separate catalytic centers. The myxothiazol-sensitive Qo site is the site of binding and oxidation of QH2 located toward the intermembrane space, while the antimycin A-sensitive Qi site is the site of binding and reduction of mtQ located on the matrix side.

The mtQ cycle of complex III involves all three forms of Q (fully oxidized, semioxidized, and fully reduced) (Fig. 3) and participates in the generation of the protonmotive force as during mtQH2 oxidation protons are released into intermembrane space [51, 52, 53, 54]. The mtQ cycle involves a series of one-electron redox reactions as a result of bifurcation of electrons at the Qo site (Fig. 1, marked box). At this site, one electron from mtQH2 is transferred to the Rieske protein, then to cytochrome c1, and finally to cytochrome c; the second electron is translocated through heme bH and heme bL to the catalytic site Qi, where mtQ is reduced to a semiubiquinone radical (QH) [51, 52, 53, 54, 55]. The oxidation of two mtQH2 molecules involves the reduction of two cytochrome c molecules and completes the reduction of semiubiquinone to mtQH2 at the Qi site. Thus, a complete enzymatic cycle of complex III gives the oxidation of one mtQH2 with the reduction of two molecules of cytochrome c and the return of half of the electrons within the mtQ pool.

Fig. 3.

Factors influencing the mtQ reduction level. mtQ, oxidized mitochondrial coenzyme Q; mtQH2, reduced mtQ, quinol; mtQtot, total mtQ (mtQ + mtQH2) in the inner mitochondrial membrane.

Although one of the main functions of the mtQ-reducing and mtQH2-oxidizing pathways is to maintain the mtQ redox homeostasis, there are few studies describing a direct relationship between their activity and the level of mtQ reduction (or mtQ redox state) and mtROS production under physiological and pathological conditions.

3. Redox State or Reduction Level of the mtQ Pool

mtQ levels, its redox state and its interaction with mtQ-dependent enzymes may play a key role in modulating mitochondrial homeostasis. One of the main properties of the mtQ pool in the inner membrane is its redox state or mtQ reduction level, defined as the mtQH2/Q ratio or the ratio of mtQH2 to the total mtQ content in the membrane (mtQH2/mtQtotal ratio), respectively. The redox state of mtQ depends on the several factors, including the activity of mtQ-reducing pathways (respiratory substrate-oxidizing dehydrogenases), the activity of mtQH2-oxidizing pathway(s) (the cytochrome pathway and AOX), and the size of the mtQ pool of the inner mitochondrial membrane (Fig. 3). The mtQH2/mtQ ratio is sensitive to an imbalance between the supply of electrons from reducing equivalents and the need for ATP. It changes rapidly and reversibly in response to changes in the function of mitochondria. Interestingly, it has recently been proposed that succinate evolved as a signaling modality because its concentration reflects the mtQ redox state [56]. By balancing with the mtQ pool, succinate enables the transfer of the mtQ redox state from the mitochondria to the rest of the cell, to the circulation and other cells.

The redox state of the mtQ pool reflects the bioenergetic mitochondrial activity of cells and is considered to be one of the markers of oxidative stress [29, 52, 57, 58, 59, 60, 61]. It represents the level of mtROS formation, and thus mitochondrial oxidative stress, and the need for protective antioxidants in mitochondria. It should be noted that mitochondrial reactive nitrogen species are also key contributors to oxidative stress [62, 63]. Nitric oxide (NO) can react with ubiquinol (QH2) to form nitric oxide radical (NO), which can react with O2 to form peroxynitrite (ONOO¯), and semiubiquinone (QH), which can react with O2 to form superoxide (O2¯) [64].

The size of the mtQ pool together with cellular Q may be modified by various factors, including endurance training [65, 66, 67], statin-induced mtQ synthesis impairment [68], cell age and oxygen level [69, 70, 71] or certain diseases, including cardiovascular and neurodegenerative disease or cancer [22, 60, 72, 73]. However, the size of the mtQ pool is rarely determined. Similarly, alteration in the redox state of the mtQ pool has so far been determined for only a few physiological and pathophysiological conditions, including endurance training [65, 66, 67] and statin treatment [68]. Interestingly, the size of the reduced nonoxidizable pools of mtQ may reflect the cell’s need for Q as an antioxidant [74]. Namely, we have recently shown that in the tissues of rats with different energy requirements (heart, brain, liver and lungs), the production of ROS at the tissue and mitochondrial levels was strongly related to the available reduced nonoxidizable Q and mtQ pools, i.e., the cellular QH2 and mtQH2 content measured in the absence or depletion of oxidative metabolism substrates [74].

The mtQ reduction level adapts dynamically to variable metabolic conditions in the cell. Therefore, another possible function of the mtQ reduction level is related to the action of mitochondrial uncoupling proteins (UCPs), which mediate regulated proton leakage across the inner mitochondrial membrane, leading to the dissipation of energy stored in the proton electrochemical gradient [75, 76]. First, it was observed that mtQ is an obligatory cofactor for UCP function [77], and the regulation of uncoupling activity by mtQ depends on its reduction level [78]. Moreover, we have reported that mtQH2 but not mtQ acts as a negative modulator for the UCP inhibition by purine nucleotides [79].

4. mtQ and mtROS Formation

Mitochondria are an important source of ROS (mtROS) in eukaryotic cells. Superoxide (O2¯) and hydrogen peroxide (H2O2) are produced by the leakage of electrons from the redox centers of the mitochondrial respiratory chain and substrate oxidation-associated enzymes, leading to a one-electron or two-electron reduction of oxygen [36, 80, 81]. mtQ participates in the generation of O2¯ from semiubiquinone radicals by the leakage of electrons from its redox group with unpaired electron (Fig. 2).

mtROS are formed as a by-product of normal aerobic metabolism or under conditions of metabolic stress leading to oxidative stress. Under physiological conditions, mtROS can act as beneficial signaling molecules sent from the mitochondria to other parts of the cell that is essential for cellular adaptive responses [36, 82, 83, 84]. Dysfunction of the mitochondrial respiratory chain leads to decreased energy production and mtROS overproduction. Under oxidative stress conditions, excessive mtROS production can lead to a range of oxidative damage, including oxidative modification of cellular macromolecules such as lipids, DNA and proteins, which is the cause of aging and many diseases. In addition, damaged mitochondria release apoptotic factors that act as signals inducing cell death.

In mammalian mitochondria, at least eleven sites can generate mtROS (O2¯ and/or H2O2) in the electron transport chain (seven sites) and related enzymes (four sites) involved in the oxidation of the reduced substrates, including the TCA cycle and β-fatty acid oxidation enzymes (Fig. 4) [80, 85]. The sites of O2¯/H2O2 production are present in respiratory chain complexes (complexes I, II and III) and dehydrogenases (mitochondrial glycerol-3-phosphate dehydrogenase, dihydroorotate dehydrogenase, and ETF:mtQ oxidoreductase), the redox cofactor of which is mtQ. mtQ is directly involved in the formation of O2¯/H2O2 at four mtQ-binding sites, i.e., the mtQ-reducing sites of complex I (IQ) [86], mitochondrial glycerol-3-phosphate dehydrogenase (GQ) [87], dihydroorotate dehydrogenase (DQ) [88], and the mtQH2-oxidizing site of complex III (IIIQo) [89]. Other dehydrogenases linked to mtQ, such as proline dehydrogenase, feed electrons into the mtQ pool and then reduce other sites that produce O2¯/H2O2 [90]. However, it is possible that these dehydrogenases also produce O2¯/H2O2 at mtQ-reducing sites, but this production is too small to be detected [80]. Changes in the bioenergetic states of mitochondria, as a consequence of alterations of mitochondrial substrate availability and the reduction states of mitochondrial redox centers, including mtQ, affect the rate of O2¯/H2O2 production in various mitochondrial sites [80, 85]. There are seven characterized O2¯/H2O2 production sites that are not directly related to mtQ reduction but contain redox-active flavin (FMN or FAD): IF of complex I [86], IIF of complex II [91], EF of ETF/ETF:mtQ system [92], and OF, BF, AF, and PF sites of 2-oxoacid dehydrogenase complexes, i.e., 2-oxoglutarate dehydrogenase (the TCA cycle enzyme), branched-chain 2-oxoacid dehydrogenase, 2-oxoadipate dehydrogenase, and pyruvate dehydrogenase complexes, respectively [93]. The 2-oxoacid dehydrogenase complexes contain a dihydrolipoamide dehydrogenase subunit with redox-active FAD, which acts as an O2¯/H2O2 production site. They feed electrons to NAD+. Other NAD-linked dehydrogenases, such as malate dehydrogenase, isocitrate dehydrogenase (TCA cycle enzymes), and glutamate dehydrogenase reduce other sites that produce O2¯/H2O2 [80]. Electrons from NADH formed in all NAD-linked dehydrogenases are transferred to the flavin (FMN) in the NADH-oxidizing site of complex I (IF) and then to the mtQ-reducing site of complex I (IQ).

Fig. 4.

Eleven mitochondrial sites of O2¯/H2O2 formation. mtQ, ubiquinone; mtQH2, ubiquinol; ETF, electron transfer flavoprotein; ETF:Q, ETF:Q oxidoreductase; ACADS, acetyl-CoA dehydrogenase; IMM, inner mitochondrial membrane. Flavin-dependent (F, purple stars) or mtQ-binding sites (Q, green stars) of the 2-oxoglutarate dehydrogenase (OGDh) (OF), branched chain 2-oxoacid dehydrogenase (BCODh) (BF), 2-oxoadipate dehydrogenase (OADh) (AF), and pyruvate dehydrogenase (PDh) (PF) complexes, of complex I (IF, IQ), complex II (IIF), the outer mtQ site of complex III (IIIQo), mitochondrial glycerol-3-phosphate dehydrogenase (mtGPDh) (GQ), the ETF/ETF:mtQ oxidoreductase system (ETF/ETF:mtQ) (EF), and dihydroorotate dehydrogenase (DHODh) (DQ). Based on [85].

Site IIIQo of complex III and site GQ of glycerol-3-phosphate dehydrogenase generate O2¯ in both the mitochondrial matrix and intermembrane space [80, 87, 89]. Other mtROS formation sites produce O2¯/H2O2 in the mitochondrial matrix. These different O2¯/H2O2 formation topologies are important for intracellular signaling using mtROS as signaling molecules.

4.1 mtQ-Biding Sites of mtROS Production

In the inner mitochondrial membrane, mtQ is a molecule directly involved in the production of mtROS due to the formation of O2¯/H2O2 by the leakage of electrons from the semiubiquinone radical. All four mtQ-binding sites (IQ, IIIQo, GQ, DQ) produce O2¯ and only one of them (DQ) is likely to produce H2O2 as well [80, 85].

At IQ of complex I, O2¯ is mainly generated during reverse electron transport (RET), i.e., a transfer of electrons against the redox potential gradient supported by a high proton electrochemical gradient (a high protonmotive force) across the inner mitochondrial membrane and a high mtQ redox state (highly reduced mtQ pool) [86]. Under these conditions, mtQH2 is oxidized to mtQ at the mtQ-binding site (IQ) to drive the reduction of NAD+ to NADH at the FMN site of complex I (IF). Inhibition of mtQ-binding with rotenone inhibits RET and lowers mtROS production.

At the outer mtQ-binding site of complex III (IIIQo), O2¯ production is low in the absence of Qi site inhibitors [94]. In the presence of antimycin A, which binds to the Qi site and blocks the oxidation of cytochrome bL heme, the mtQ cycle is disrupted. The resulting accumulation of reduced cytochrome bL heme limits semiubiquinone oxidation at site Qo. Semiubiquinone interacts with oxygen to form O2¯ on both sides of the inner mitochondrial membrane. In the absence of antimycin A, a high protonmotive force and highly reduced mtQ pool decreases the electron transfer from the cytochrome bL heme to the cytochrome bH heme, leading to more reduced bL heme and thus increased O2¯ production [95].

Site GQ of mitochondrial glycerol-3-phosphate dehydrogenase produces O2¯ in approximately equal amounts on each side of the mitochondrial inner membrane, suggesting that the mtQ-binding site of the enzyme is the major site of O2¯ formation during glycerol-3-phosphate oxidation [87]. Dihydroorotate dehydrogenase can generate O2¯ and/or H2O2 directly at its mtQ-binding site (DQ) at low rates [88]. The enzyme is also capable of indirect production at higher rates from other sites through its ability to reduce the mtQ pool.

In a cell, the net O2¯/H2O2formation rate at mitochondrial mtQ-binding sites depends on the substrates being oxidized and the antioxidant systems acting on both sides of the inner mitochondrial membrane. Among the mtQ-binding sites, IIIQo and IQ, which is mainly active during RET, dominate in O2¯ production and play an important role in redox signaling and the induction of oxidative damage [80].

4.2 mtQ Reduction Level and mtROS Formation

As a mobile component of the mitochondrial electron transport chain, mtQ acts as a pro-oxidant in the semiubiquinone state. It is generally accepted that the production of mtROS depends on the level of reduction of the mtQ pool, which determines the formation of O2¯ from semiubiquinone. However, there are only a few data relating the mtQ redox state, i.e., mtQH2/Q ratio [96] and mtQH2/mtQtotal ratio [59, 66, 68, 95], with mtROS generation. The formation of mtROS related to the activity of the respiratory chain, and thus the mtQ redox state, depends on the activity of all mtQ-reducing dehydrogenases and mtQH2-oxidazing pathway(s), and on the activity of the oxidative phosphorylation process in which ATP is formed [59]. We have shown that under ADP phosphorylating conditions (state 3), a low mtQ reduction level and a low mΔΨ are accompanied by decreased mtROS production [59]. In a nonphosphorylating state (state 4), a highly reduced mtQ pool (a high mtQ reduction level) and a high mΔΨ are accompanied by increased mtROS production. Thus, the mtQH2/mtQ ratio is sensitive to changes in electron supply from reducing equivalents and to ATP demand. The activity of uncoupling proteins (UCPs), which mediate proton leakage, can lower mΔΨ and the mtQ reduction level, leading to a decrease in mtROS production [76, 84].

The relationship between mtROS production and mtΔΨ was investigated more often than the interplay between mtROS formation and the mtQ reduction level. For the first time, the dependence of mtROS formation on mtΔΨ was described in rat heart mitochondria [97]. Namely, a very steep dependence of H2O2 formation on mtΔΨ was observed in nonphosphorylating rat heart mitochondria after exceeding a threshold value near mtΔΨ of ADP phosphorylating mitochondria. However, not under all conditions lowering mtΔΨ leads to a decrease in mtROS production (e.g., in the presence of inhibitors of complexes III and IV), indicating that mtROS is not a direct function of the mtΔΨ level (Fig. 5A) [59]. For example, we have shown that in the mitochondria of the amoeba Acanthamoeba castellanii, with inhibition of the Qi site of complex III by antimycin A or complex IV by cyanide, mΔΨ decreases while the mtQ reduction and H2O2 production increase [59]. Marphy’s group has shown that the formation of O2¯ at the Qo site is associated with the accumulation of reduced cytochrome bL when the Qi site is blocked by antimycin A, and this phenomenon is not dependent on mtΔΨ [95].

Fig. 5.

The relationship between mtROS formation versus mtΔΨ and the mtQ reduction level. The relationship between H2O2 formation versus mtΔΨ (A) for the cytochrome pathway (CII + CIII + CIV) fueled by complex II (CII) and the relationship between H2O2 formation versus the mtQ reduction level (mtQH2/mtQtot) (B) for the CII + CIII + CIV pathway (blue curves and lines) and the CII-fueled alternative oxidase (AOX) (CII + AOX) (purple line). Based on results described in A. castellanii mitochondria [59]. The mtΔΨ level (A) and mtQ reduction level (B) of the CII + CIII + CIV pathway was varied with the ATP synthesis and CII inhibitors, with the uncoupler (blue solid curves), and with the inhibitor of CIII (antimycin A, AA) (blue dashed lines) when AOX inhibitor was present. Measurements were carried out under nonphosphorylating (state 4) and phosphorylating (state 3) conditions. (B) For the CII + AOX pathway, the mtQ reduction level was titrated with an activator of AOX (GMP) and an inhibitor of CII (malonate) in the presence of AA to block the cytochrome pathway.

We have kinetically described a direct dependence of mtROS production on the mtQ reduction level in A. castellanii mitochondria [59]. For both mtQH2-oxidizing pathways (the cytochrome and AOX pathways), during the oxidation of succinate fueling the respiratory chain at the level of complex II, a higher level of mtQ reduction leads to greater mtROS formation (Fig. 5B) [59]. This relationship is also observed for the cytochrome pathway in the presence of antimycin A, which by inhibiting the Qi site lowers mΔΨ and increases mtQ reduction level. In the case of the cytochrome pathway, H2O2 generation depends nonlinearly on the reduction level of mtQ pool. At levels of mtQ reduction higher than that of the phosphorylating state (above 35%) there is a steep relationship between H2O2 generation and the mtQ reduction level. In contrast, AOX is active only when the mtQ reduction level is high (above 40%) and the dependence between H2O2 generation and mtQ reduction level is approximately linear (Fig. 5B). We have proposed that the mtQ pool reduction level (the endogenous redox state of mtQ) may be a useful endogenous marker that allows for the assessment of total mitochondrial mtROS production [59].

High levels of mtROS are generated in response to stress via RET through complex I [61, 98]. RET occurs when the mtQ pool it is strongly reduced by electrons from the respiratory complex II (succinate dehydrogenase) [95, 98]. In various cell lines, it has been observed how the respiratory chain is optimized for better oxidation of various reducing respiratory fuels by the redox state of mtQ (mtQH2/Q ratio) acting as a metabolic detector and mtROS formed by complex I (via RET) acting as an executor [96]. In the rat heart mitochondria, the formation of mtROS by RET at complex I has been shown to be dependent on the mtQ reduction level and mtΔΨ that indicates the sensitivity of O2¯ formation by the mtQ-binding site of complex I to these two metabolic variables [95].

4.3 mtQ Content and mtROS Production

It seems that the production of mtROS does not directly depend on the amount of mtQ but rather on the reduction level of the mtQ pool (mtQH2/mtQtotal ratio) as a function of the activity of mtQ-reducing pathways (mtQ-reducing dehydrogenases) and the mtQH2-oxidizing cytochrome pathway(s). In the heart mitochondria of MCLK1 (CoQ7) mutant mice, which contain only ~10% of normal Q content, lower rates of mtROS production from all known mtROS production sites were detected at respiration levels similar to those in control mitochondria [99]. One possible explanation for how low mtQ levels reduce mtROS formation is that low mtQ induces the formation of supercomplexes that have been proposed to increase electron transport efficiency from one redox component to another, thus minimizing electron leakage and mtROS production [100, 101]. In cultured fibroblasts from Q-deficient patients, increased mtROS production was observed during intermediate Q deficiency (50–70%) possibly as a result of increased semiubiquinone generation from increased redox cycling of the restricted mtQ pool, while a more severe loss of Q (>85%) was not accompanied by significant mtROS production [102]. Interestingly, we have shown that in lung mitochondria of endurance-trained rats, reduced mtQ content (mtQ9 by ~16% and mtQ10 by ~42%) and decreased activity of complex I as well as increased activity of cytochrome pathway (complex III plus complex IV) may account for the observed diminished mtQ reduction level during succinate and/or malate oxidation, resulting in a general decrease of mtROS formation by mitochondria [66]. Moreover, endurance training downregulates complex I in supercomplexes and upregulates complex III in the CIII2 + CIV supercomplex in the inner membrane of lung mitochondria. In addition, comparison of the effects of endurance training on mtQ amount and mtROS generation in rat tissues with high energy demand, i.e., in the brain, liver and heart, indicates that endurance training may induce various mitochondrial and tissue responses related to mtQ acting as an electron carrier in the respiratory chain and as an antioxidant [67]. Changes in the formation of mtROS observed in the mitochondria of individual rat tissues may result from changes in the size of the pool of oxidized mtQ, which acts as an electron carrier, as well as the amount and activity of individual complexes of the oxidative phosphorylation system and its molecular organization.

A decrease in the mtQ pool size and mtQ reduction level may results from drug treatment. We have shown that in cultured endothelial cells, chronic exposure to atorvastatin, an inhibitor of HMG-CoA reductase led to a significant decrease in mtQ10 content (~23%) [68]. The activity of complex III and its amount in a supercomplex with complex IV was also diminished. The statin-induced changes at the respiratory chain level of the endothelial mitochondria led to a decrease in succinate and/or malate oxidation accompanied by a decreased mtΔΨ, as well as to an increase in the mtQ10 reduction level and consequently increased mtROS production.

Thus, mtQ redox homeostasis is a key factor in modulating mtROS production. As mentioned above, depending on the activity of mtQ-reducing pathways and mtQH2-oxidizing pathways, a decreased amount of mtQ can lead to either an increased reduction level of mtQ (a higher mtQH2/mtQtotal ratio) and thus to an increased production of mtROS or to a decreased reduction level of mtQ (a lower mtQH2/mtQtotal ratio) and hence decreased production of mtROS. Superphysiological mtQ levels could possibly have the same effect by altering the level of mtQ reduction. Thus, changing the size of the mtQ pool and its redox state is an important physiological adaptation. The metabolic response to the altered mtQ pool appears to depend on whether it is more important to maintain an efficient oxidative phosphorylation process or whether it is more important to defend against oxidative stress.

The level of Q, including mtQ, varies with the type of membrane, tissue and organism [4, 103]. It is highest in organs with a high metabolic rate and hence a high energy requirement, such as the heart, muscles, and liver. However, despite high metabolic activity, total Q levels and the QH2/Q ratio in the human brain are very low compared to other tissues [4]. The mtQ pool is in excess compared to other components of the respiratory chain [100]. However, with age and in many disease states, including primary and secondary Q deficiencies, associated with increased production and action of mtROS, the level of cellular and thus mitochondrial Q decreases [23, 60, 71, 72, 104]. mtQ deficiency may impair oxidative phosphorylation and therefore respiratory chain activity and ATP synthesis [105, 106, 107, 108], and may induce oxidative stress (mtROS overproduction) leading to oxidative modification of macromolecules, and ultimately to apoptosis. However, the question of how Q deficiency due to disease or aging affect the redox state of mtQ has not yet been thoroughly investigated.

Increasing mtQ levels must also result in a change in mtQ redox homeostasis and mtROS production. For example, we have observed an increase in mtQ9 content (by ~20%) in skeletal muscle mitochondria after eight-week endurance training [54]. In these mitochondria, H2O2 production and mtQ reduction level were elevated under nonphosphorylating conditions, and decreased under phosphorylating conditions, while the efficiency of oxidative phosphorylation increased with unchanged components of the respiratory chain. Recently, it has been shown that elevated levels of mtQ caused by Q10 delivery (via by BPM31510) lead to an increase in ROS generation and consequently, an imbalance in cellular redox homeostasis that activates mtROS-mediated cell death pathways in pancreatic cancer cells [109]. Similarly, cell death activation due to increased mtQ pool was observed in vitro in pancreatic cancer cells, and both human patient-derived organoids and tumor xenografts. Interestingly, in other studies, chemotherapy induced increased production of ROS and increased levels of Q10 (especially its reduced form Q10H2) in all cancer cell lines tested [110]. But again, we do not know how mtQ redox state changes under these conditions and what role it plays in mtROS production.

5. mtQ as an Antioxidant

The redox state of the mtQ pool is an important indicator of the bioenergetic and antioxidant status of the mitochondria. The fully reduced form of mtQ (mtQH2) can act as a nonprotein lipophilic antioxidant [111, 112]. The level of mtROS depends not only on their production in the mitochondria and thus on the amount of mtQ and its redox state in the inner mitochondrial membrane, but also on the mtROS scavenging by reduced forms of mtQ. Additionally, by acting as an antioxidant, mtQ protects against oxidative damage caused by mtROS. mtQH2 regenerates other antioxidants (vitamin C or vitamin E) as well as reduces and neutralizes oxidants or free radicals, including mtROS [11, 22, 112]. As an effective free radical scavenger, mtQH2 prevents oxidative modifications of mitochondrial proteins and DNA and inhibits the initiation and propagation of membrane lipid peroxidation. In lipid peroxidation reactions, one-electron oxidation of mtQH2 to semiubiquinone is linked to the reduction of perferryl radicals or lipid peroxyl radicals [12]. Reactive semiubiquinone radicals are neutralized by reduction to QH2 by α-tocopherol (vitamin E). However, mtQH2 is still a capable antioxidant molecule when vitamin E is absent [113]. On the other hand, the indirect antioxidant effect of mtQH2 can regenerate α-tocopherol from α-tocopheroxyl radicals (α-TO) [114]. Interestingly, in addition to the plasma membrane and intracellular membranes, Q is also a component of low-density proteins (LDL), suggesting its potential antioxidative role along with α-tocopherol in the prevention of atherosclerosis [115].

Thus, as a molecule with antioxidant and prooxidative properties, mtQ contributes both to oxidative damage to mitochondria and to their antioxidant defense. The antioxidant action of mtQH2 is associated with deprotonation (and electron donation), leading to the formation of semiubiquinone radicals and ultimately oxidized mtQ, which is then reduced back to mtQH2 by the mitochondrial respiratory chain [112, 116].

It is widely believed that the antioxidant properties of Q may have beneficial effects on health and length of life [117]. Q has been widely introduced as an antiaging supplement as well as a potential treatment for several human diseases related to mitochondrial dysfunction-induced oxidative stress. However, understanding the complex effect of Q on the human body and its relationship to lifespan is still incomplete in view of inconsistent literature reports on Q supplementation, describing its beneficial or nil effects [23, 29, 60, 71, 73]. Paradoxically, some Q biosynthesis defects result in an increase in life span in animal models, most likely due to a mild reduction in respiratory chain activity and thus a reduction in mtROS production or due to a physiological adaptation to early mitochondrial dysfunction [118]. On the other hand, Q10 supplementation may have positive effects in the treatment of certain cancers in patients with reduced plasma and tissue Q levels [109]. Similarly, Q10 supplementation may have beneficial effects on aging-related disorders accompanied by a reduced Q content, especially in cardiovascular and metabolic diseases [119, 120, 121, 122]. However, effect of Q10 supplementation on reversing functional decline of mitochondrial bioenergetics is unclear. Moreover, it is unclear if this improvement is always related to the increase in mtQ levels and the change in mtQ redox state after Q10 supplementation. Supplementation with high doses of Q10 may increase both circulating and intracellular levels, but there are conflicting results regarding bioavailability [71, 123, 124]. Although it has been shown that endogenous levels of mtQ in the mitochondria of the rodent brain, skeletal muscle, and heart can be increased after Q10 administration [125, 126].

Nonoxidizable mtQH2 Pool as an Mitochondrial Antioxidant

We have proposed that the antioxidant effect may also be exerted by the reduced mtQH2 pool in the inner mitochondrial membrane, which is not oxidized by the respiratory chain [74]. Interestingly, in mitochondria isolated from different rat tissues with different energy requirements and different mitochondrial oxidative activity heart, brain, liver and lungs, the size of the mtQH2 pool that was not oxidized by the respiratory chain (measured in the absence of respiratory substrates) was associated with the mtROS production capacity during the oxidation of respiratory substrates [74]. Thus, the size of the reduced nonoxidizable pools of mtQ (nonoxidizable mtQH2 pool) may reflect the need for mtQ as a mitochondrial antioxidant. Therefore, it is possible that both the mtQH2 pool, which participates in the mtQ cycle of the electron transport chain, and the mtQH2 pool, which is not oxidized by the chain, contribute to the overall antioxidative activity of mitochondria.

6. Conclusions and Perspectives

As a redox active cofactor of mitochondrial respiratory chain complexes and related enzymes, mtQ influences mtROS formation and contributes to oxidative damage to mitochondria. On the other hand, thanks to its redox properties, mtQ acts as an antioxidant that protects the mitochondria from oxidative modifications. Mitochondria contain the most Q in a cell but are also the most dependent on its content. A number of physiological and pathophysiological factors affect the amount of mtQ and thus its redox state and mtROS production level (Fig. 6), indicating that mtQ homeostasis is essential for metabolic adaptation and the maintenance of mitochondrial function.

Fig. 6.

Factors modulating the mtQ pool content which influences the redox state of mtQ and thus the formation of mtROS. mtQ, mitochondrial coenzyme Q; mtROS, mitochondrial reactive oxygen species.

The purpose of our review was to highlight the rather little-studied and discussed crucial role of mtQ redox homeostasis in modulating mtROS production. As we proposed, the level of reduction (the endogenous redox state) of mtQ may be a useful marker to assess total mtROS generation. However, there are few studies describing a direct relationship between the mtQ redox state and mtROS production under physiological and pathological conditions. mtQ redox ratio has rarely been measured, especially under conditions of Q deficiency. Undoubtedly, measuring the Q redox state in cells, tissues or isolated mitochondria is technically demanding and not as popular as measuring mtROS formation. Determination of the QH2/Q ratio is more complicated than determining the total Q pool because ubiquinol (QH2) is particularly sensitive to oxidation. The difficulty in determining the redox state of mtQ in biological samples and in vivo has been a significant obstacle in characterizing mtQ function in mitochondria under physiological and pathophysiological conditions. An extraction technique followed by high performance liquid chromatography (HPLC) detection [127, 128] is usually used to determine the concentration of cellular and mitochondrial Q and QH2, and to calculate the level of Q reduction (QH2/Qtotal) or Q redox state (QH2/Q) in cells/tissues and isolated mitochondria, respectively. However, this technical approach requires a large amount of biological material, is time consuming and, importantly, results in a non-kinetic point measurement. Presumably, the recent advancement in the technique of simultaneous kinetic measurement of oxygen consumption and mtQ redox state by combining a Clark-type oxygen electrode and a Q-type electrode [28, 129] will contribute to the widespread measurement of the redox state of mtQ in isolated respiring mitochondria. In addition, a sensitive approach to assess the Q redox state by liquid chromatography-tandem mass spectrometry (LC-MS/MS) in isolated mitochondria, cells and tissues in vivo has recently been developed [130]. This technical approach allows the analysis of both the redox state of Q (non-kinetic point measurements) and the size of the Q pool with negligible changes in the redox state due to isolation and extraction from small amounts of biological material. Understanding metabolism linked to the mtQ pool is challenging but new experimental models and methodologies will undoubtedly help. Determining the redox state of Q under physiological and pathological conditions is of great importance for better understanding how Q contributes to the maintenance of cellular and mitochondrial homeostasis.

Author Contributions

WJ, KD, AB, KW and LG wrote the manuscript. All authors contributed to editorial changes in the manuscript. All authors read and approved the final manuscript.

Ethics Approval and Consent to Participate

Not applicable.

Acknowledgment

Not applicable.

Funding

This research was funded by National Science Centre, Poland, OPUS 2020/37/B/NZ1/01188. KD was funded by PRELUDIUM 2019/N/NZ1/01366.

Conflict of Interest

The authors declare no conflict of interest.

References
[1]
Chakrabarty RP, Chandel NS. Beyond ATP, new roles of mitochondria. The Biochemist. 2022; 44: 2–8.
[2]
Nicholls DG, Fergusson S. Bioenergetics. 4th edn. Academic Press: Netherlands, USA. 2013.
[3]
Brand MD, Orr AL, Perevoshchikova IV, Quinlan CL. The role of mitochondrial function and cellular bioenergetics in ageing and disease. The British Journal of Dermatology. 2013; 169 Suppl 2: 1–8.
[4]
Aberg F, Appelkvist EL, Dallner G, Ernster L. Distribution and redox state of ubiquinones in rat and human tissues. Archives of Biochemistry and Biophysics. 1992; 295: 230–234.
[5]
Crane FL. Discovery of ubiquinone (coenzyme Q) and an overview of function. Mitochondrion. 2007; 7: S2–S7.
[6]
Ericsson J, Dallner G. Distribution, biosynthesis, and function of mevalonate pathway lipids. Sub-cellular Biochemistry. 1993; 21: 229–272.
[7]
Bentinger M, Tekle M, Dallner G. Coenzyme Q–biosynthesis and functions. Biochemical and Biophysical Research Communications. 2010; 396: 74–79.
[8]
Song Y, Buettner GR. Thermodynamic and kinetic considerations for the reaction of semiquinone radicals to form superoxide and hydrogen peroxide. Free Radical Biology & Medicine. 2010; 49: 919–962.
[9]
Ozawa T. Formation of free radicals in the electron transfer chain and antioxidant properties of coenzyme Q. In Lenaz G. (eds), Coenzyme Q (pp. 441–456). Wiley and Sons: USA. 1985.
[10]
Crane FL. Biochemical functions of coenzyme Q10. Journal of the American College of Nutrition. 2001; 20: 591–598.
[11]
Genova ML, Lenaz G. New developments on the functions of coenzyme Q in mitochondria. BioFactors. 2011; 37: 330–354.
[12]
Kagan V, Serbinova E, Packer L. Antioxidant effects of ubiquinones in microsomes and mitochondria are mediated by tocopherol recycling. Biochemical and Biophysical Research Communications. 1990; 169: 851–857.
[13]
Bentinger M, Brismar K, Dallner G. The antioxidant role of coenzyme Q. Mitochondrion. 2007; 7: S41–S50.
[14]
Rizzo AM, Berselli P, Zava S, Montorfano G, Negroni M, Corsetto P, et al. Endogenous antioxidants and radical scavengers. Advances in Experimental Medicine and Biology. 2010; 698: 52–67.
[15]
Baschiera E, Sorrentino U, Calderan C, Desbats MA, Salviati L. The multiple roles of coenzyme Q in cellular homeostasis and their relevance for the pathogenesis of coenzyme Q deficiency. Free Radical Biology & Medicine. 2021; 166: 277–286.
[16]
Villalba JM, Navas P. Regulation of coenzyme Q biosynthesis pathway in eukaryotes. Free Radical Biology & Medicine. 2021; 165: 312–323.
[17]
Acosta MJ, Vazquez Fonseca L, Desbats MA, Cerqua C, Zordan R, Trevisson E, et al. Coenzyme Q biosynthesis in health and disease. Biochimica et Biophysica Acta. 2016; 1857: 1079–1085.
[18]
Alcázar-Fabra M, Navas P, Brea-Calvo G. Coenzyme Q biosynthesis and its role in the respiratory chain structure. Biochimica et Biophysica Acta. 2016; 1857: 1073–1078.
[19]
Lee SQE, Tan TS, Kawamukai M, Chen ES. Cellular factories for coenzyme Q10 production. Microbial Cell Factories. 2017; 16: 39.
[20]
Fernández-Del-Río L, Clarke CF. Coenzyme Q Biosynthesis: An Update on the Origins of the Benzenoid Ring and Discovery of New Ring Precursors. Metabolites. 2021; 11: 385.
[21]
Kawamukai M. Biosynthesis and bioproduction of coenzyme Q10 by yeasts and other organisms. Biotechnology and Applied Biochemistry. 2009; 53: 217–226.
[22]
Turunen M, Olsson J, Dallner G. Metabolism and function of coenzyme Q. Biochimica et Biophysica Acta. 2004; 1660: 171–199.
[23]
Hernández-Camacho JD, Bernier M, López-Lluch G, Navas P. Coenzyme Q10 Supplementation in Aging and Disease. Frontiers in Physiology. 2018; 9: 44.
[24]
Kröger A, Klingenberg M. The kinetics of the redox reactions of ubiquinone related to the electron-transport activity in the respiratory chain. European Journal of Biochemistry. 1973; 34: 358–368.
[25]
Crane FL, Widmer C, Lester RL, Hatefi Y. Studies on the electron transport system. XV. Coenzyme Q (Q275) and the succinoxidase activity of the electron transport particle. Biochimica et Biophysica Acta. 1959; 31: 476–489.
[26]
Hatefi Y, Lester RL, Crane FL, Widmer C. Studies on the electron transport system. XVI. Enzymic oxidoreduction reactions of coenzyme Q. Biochimica et Biophysica Acta. 1959; 31: 490–501.
[27]
Banerjee R, Purhonen J, Kallijärvi J. The mitochondrial coenzyme Q junction and complex III: biochemistry and pathophysiology. The FEBS Journal. 2022; 289: 6936–6958.
[28]
Komlódi T, Cardoso L, Doerrier C, Moore A, Rich P, Gnaiger E. Coupling and pathway control of coenzyme Q redox state and respiration in isolated mitochondria. Bioenerg Commun. 2021; 2021: 3.
[29]
Wang Y, Hekimi S. Understanding Ubiquinone. Trends in Cell Biology. 2016; 26: 367–378.
[30]
Galante YM, Hatefi Y. Resolution of complex I and isolation of NADH dehydrogenase and an iron–sulfur protein. Methods in Enzymology. 1978; 53: 15–21.
[31]
Sazanov LA. A giant molecular proton pump: structure and mechanism of respiratory complex I. Nature Reviews. Molecular Cell Biology. 2015; 16: 375–388.
[32]
Wikström M, Hummer G. Stoichiometry of proton translocation by respiratory complex I and its mechanistic implications. Proceedings of the National Academy of Sciences of the United States of America. 2012; 109: 4431–4436.
[33]
Iverson TM. Catalytic mechanisms of complex II enzymes: a structural perspective. Biochimica et Biophysica Acta. 2013; 1827: 648–657.
[34]
Yankovskaya V, Horsefield R, Törnroth S, Luna-Chavez C, Miyoshi H, Léger C, et al. Architecture of succinate dehydrogenase and reactive oxygen species generation. Science. 2003; 299: 700–704.
[35]
Anderson RF, Shinde SS, Hille R, Rothery RA, Weiner JH, Rajagukguk S, et al. Electron-transfer pathways in the heme and quinone-binding domain of complex II (succinate dehydrogenase). Biochemistry. 2014; 53: 1637–1646.
[36]
Murphy MP. How mitochondria produce reactive oxygen species. The Biochemical Journal. 2009; 417: 1–13.
[37]
Quinlan CL, Perevoshchikova IV, Hey-Mogensen M, Orr AL, Brand MD. Sites of reactive oxygen species generation by mitochondria oxidizing different substrates. Redox Biology. 2013; 1: 304–312.
[38]
Enriquez JA, Lenaz G. Coenzyme q and the respiratory chain: coenzyme q pool and mitochondrial supercomplexes. Molecular Syndromology. 2014; 5: 119–140.
[39]
Ruzicka FJ, Beinert H. A new iron-sulfur flavoprotein of the respiratory chain. A component of the fatty acid beta oxidation pathway. The Journal of Biological Chemistry. 1977; 252: 8440–8445.
[40]
Watmough NJ, Frerman FE. The electron transfer flavoprotein: ubiquinone oxidoreductases. Biochimica et Biophysica Acta. 2010; 1797: 1910–1916.
[41]
Lieberman I, Kornberg A. Enzymic synthesis and breakdown of a pyrimidine, orotic acid. I. Dihydro-orotic dehydrogenase. Biochimica et Biophysica Acta. 1953; 12: 223–234.
[42]
Reis RAG, Calil FA, Feliciano PR, Pinheiro MP, Nonato MC. The dihydroorotate dehydrogenases: Past and present. Archives of Biochemistry and Biophysics. 2017; 632: 175–191.
[43]
Green DE. alpha-Glycerophosphate dehydrogenase. The Biochemical Journal. 1936; 30: 629–644.
[44]
Mráček T, Drahota Z, Houštěk J. The function and the role of the mitochondrial glycerol-3-phosphate dehydrogenase in mammalian tissues. Biochimica et Biophysica Acta. 2013; 1827: 401–410.
[45]
Taggart JV, Krakaur RB. Studies on the cyclophorase system; the oxidation of proline and hydroxyproline. The Journal of Biological Chemistry. 1949; 177: 641–653.
[46]
Servet C, Ghelis T, Richard L, Zilberstein A, Savoure A. Proline dehydrogenase: a key enzyme in controlling cellular homeostasis. Frontiers in Bioscience-Landmark. 2012; 17: 607–620.
[47]
Rasmusson AG, Geisler DA, Møller IM. The multiplicity of dehydrogenases in the electron transport chain of plant mitochondria. Mitochondrion. 2008; 8: 47–60.
[48]
Antos-Krzeminska N, Jarmuszkiewicz W. Alternative Type II NAD(P)H Dehydrogenases in the Mitochondria of Protists and Fungi. Protist. 2019; 170: 21–37.
[49]
Vanlerberghe GC. Alternative oxidase: a mitochondrial respiratory pathway to maintain metabolic and signaling homeostasis during abiotic and biotic stress in plants. International Journal of Molecular Sciences. 2013; 14: 6805–6847.
[50]
Sluse FE, Jarmuszkiewicz W. Alternative oxidase in the branched mitochondrial respiratory network: an overview on structure, function, regulation, and role. Brazilian Journal of Medical and Biological Research. 1998; 31: 733–747.
[51]
Mitchell P. The protonmotive Q cycle: a general formulation. FEBS Letters. 1975; 59: 137–139.
[52]
Sarewicz M, Osyczka A. Electronic connection between the quinone and cytochrome C redox pools and its role in regulation of mitochondrial electron transport and redox signaling. Physiological Reviews. 2015; 95: 219–243.
[53]
Crofts AR, Holland JT, Victoria D, Kolling DRJ, Dikanov SA, Gilbreth R, et al. The Q-cycle reviewed: How well does a monomeric mechanism of the bc(1) complex account for the function of a dimeric complex? Biochimica et Biophysica Acta. 2008; 1777: 1001–1019.
[54]
Cramer WA, Hasan SS, Yamashita E. The Q cycle of cytochrome bc complexes: a structure perspective. Biochimica et Biophysica Acta. 2011; 1807: 788–802.
[55]
Wikström M, Krab K. The semiquinone cycle. A hypothesis of electron transfer and proton translocation in cytochrome bc-type complexes. Journal of Bioenergetics and Biomembranes. 1986; 18: 181–193.
[56]
Murphy MP, Chouchani ET. Why succinate? Physiological regulation by a mitochondrial coenzyme Q sentinel. Nature Chemical Biology. 2022; 18: 461–469.
[57]
Acworth IN, Ullucci PA, Gamache PH. Determination of oxidized and reduced CoQ10 and CoQ9 in human plasma/serum using HPLC-ECD. Methods in Molecular Biology. 2008; 477: 245–258.
[58]
Scialo F, Sanz A. Coenzyme Q redox signalling and longevity. Free Radical Biology & Medicine. 2021; 164: 187–205.
[59]
Dominiak K, Koziel A, Jarmuszkiewicz W. The interplay between mitochondrial reactive oxygen species formation and the coenzyme Q reduction level. Redox Biology. 2018; 18: 256–265.
[60]
Gutierrez-Mariscal FM, Arenas-de Larriva AP, Limia-Perez L, Romero-Cabrera JL, Yubero-Serrano EM, López-Miranda J. Coenzyme Q10 Supplementation for the Reduction of Oxidative Stress: Clinical Implications in the Treatment of Chronic Diseases. International Journal of Molecular Sciences. 2020; 21: 7870.
[61]
Scialò F, Fernández-Ayala DJ, Sanz A. Role of Mitochondrial Reverse Electron Transport in ROS Signaling: Potential Roles in Health and Disease. Frontiers in Physiology. 2017; 8: 428.
[62]
Ghafourifar P, Richter C. Nitric oxide synthase activity in mitochondria. FEBS Letters. 1997; 418: 291–296.
[63]
Aranda-Rivera A, Cruz-Gregorio A, Arancibia-Hernández Y, Hernández-Cruz E, Pedraza-Chaverri J. RONS and oxidative stress: an overview of basic concepts. Oxygen. 2022; 2: 437–478.
[64]
Poderoso JJ, Carreras MC, Schöpfer F, Lisdero CL, Riobó NA, Giulivi C, et al. The reaction of nitric oxide with ubiquinol: kinetic properties and biological significance. Free Radical Biology & Medicine. 1999; 26: 925–935.
[65]
Zoladz JA, Koziel A, Woyda-Ploszczyca A, Celichowski J, Jarmuszkiewicz W. Endurance training increases the efficiency of rat skeletal muscle mitochondria. Pflugers Archiv: European Journal of Physiology. 2016; 468: 1709–1724.
[66]
Jarmuszkiewicz W, Dominiak K, Galganski L, Galganska H, Kicinska A, Majerczak J, et al. Lung mitochondria adaptation to endurance training in rats. Free Radical Biology & Medicine. 2020; 161: 163–174.
[67]
Dominiak K, Galganski L, Budzinska A, Woyda-Ploszczyca A, Zoladz JA, Jarmuszkiewicz W. Effects of Endurance Training on the Coenzyme Q Redox State in Rat Heart, Liver, and Brain at the Tissue and Mitochondrial Levels: Implications for Reactive Oxygen Species Formation and Respiratory Chain Remodeling. International Journal of Molecular Sciences. 2022; 23: 896.
[68]
Broniarek I, Dominiak K, Galganski L, Jarmuszkiewicz W. The Influence of Statins on the Aerobic Metabolism of Endothelial Cells. International Journal of Molecular Sciences. 2020; 21: 1485.
[69]
Lass A, Kwong L, Sohal RS. Mitochondrial coenzyme Q content and aging. BioFactors. 1999; 9: 199–205.
[70]
Kalén A, Appelkvist EL, Dallner G. Age-related changes in the lipid compositions of rat and human tissues. Lipids. 1989; 24: 579–584.
[71]
Aaseth J, Alexander J, Alehagen U. Coenzyme Q10 supplementation - In ageing and disease. Mechanisms of Ageing and Development. 2021; 197: 111521.
[72]
Thapa M, Dallmann G. Role of coenzymes in cancer metabolism. Seminars in Cell & Developmental Biology. 2020; 98: 44–53.
[73]
López-Lluch G. Coenzyme Q homeostasis in aging: Response to non-genetic interventions. Free Radical Biology & Medicine. 2021; 164: 285–302.
[74]
Dominiak K, Jarmuszkiewicz W. The Relationship between Mitochondrial Reactive Oxygen Species Production and Mitochondrial Energetics in Rat Tissues with Different Contents of Reduced Coenzyme Q. Antioxidants. 2021; 10: 533.
[75]
Nicholls DG, Bernson VS, Heaton GM. The identification of the component in the inner membrane of brown adipose tissue mitochondria responsible for regulating energy dissipation. Experientia. Supplementum. 1978; 32: 89–93.
[76]
Woyda-Ploszczyca AM, Jarmuszkiewicz W. The conserved regulation of mitochondrial uncoupling proteins: From unicellular eukaryotes to mammals. Biochimica et Biophysica Acta. Bioenergetics. 2017; 1858: 21–33.
[77]
Echtay KS, Winkler E, Klingenberg M. Coenzyme Q is an obligatory cofactor for uncoupling protein function. Nature. 2000; 408: 609–613.
[78]
Jarmuszkiewicz W, Navet R, Alberici LC, Douette P, Sluse-Goffart CM, Sluse FE, et al. Redox state of endogenous coenzyme q modulates the inhibition of linoleic acid-induced uncoupling by guanosine triphosphate in isolated skeletal muscle mitochondria. Journal of Bioenergetics and Biomembranes. 2004; 36: 493–502.
[79]
Woyda-Ploszczyca A, Jarmuszkiewicz W. Ubiquinol (QH(2)) functions as a negative regulator of purine nucleotide inhibition of Acanthamoeba castellanii mitochondrial uncoupling protein. Biochimica et Biophysica Acta. 2011; 1807: 42–52.
[80]
Brand MD. Mitochondrial generation of superoxide and hydrogen peroxide as the source of mitochondrial redox signaling. Free Radical Biology & Medicine. 2016; 100: 14–31.
[81]
Fridovich I. Fundamental aspects of reactive oxygen species, or what’s the matter with oxygen? Annals of the New York Academy of Sciences. 1999; 893: 13–18.
[82]
D’Autréaux B, Toledano MB. ROS as signalling molecules: mechanisms that generate specificity in ROS homeostasis. Nature Reviews. Molecular Cell Biology. 2007; 8: 813–824.
[83]
Kowaltowski AJ, de Souza-Pinto NC, Castilho RF, Vercesi AE. Mitochondria and reactive oxygen species. Free Radical Biology & Medicine. 2009; 47: 333–343.
[84]
Zhao RZ, Jiang S, Zhang L, Yu ZB. Mitochondrial electron transport chain, ROS generation and uncoupling (Review). International Journal of Molecular Medicine. 2019; 44: 3–15.
[85]
Wong HS, Dighe PA, Mezera V, Monternier PA, Brand MD. Production of superoxide and hydrogen peroxide from specific mitochondrial sites under different bioenergetic conditions. The Journal of Biological Chemistry. 2017; 292: 16804–16809.
[86]
Treberg JR, Quinlan CL, Brand MD. Evidence for two sites of superoxide production by mitochondrial NADH-ubiquinone oxidoreductase (complex I). The Journal of Biological Chemistry. 2011; 286: 27103–27110.
[87]
Orr AL, Quinlan CL, Perevoshchikova IV, Brand MD. A refined analysis of superoxide production by mitochondrial sn-glycerol 3-phosphate dehydrogenase. The Journal of Biological Chemistry. 2012; 287: 42921–42935.
[88]
Hey-Mogensen M, Goncalves RLS, Orr AL, Brand MD. Production of superoxide/H2O2 by dihydroorotate dehydrogenase in rat skeletal muscle mitochondria. Free Radical Biology & Medicine. 2014; 72: 149–155.
[89]
Quinlan CL, Gerencser AA, Treberg JR, Brand MD. The mechanism of superoxide production by the antimycin-inhibited mitochondrial Q-cycle. The Journal of Biological Chemistry. 2011; 286: 31361–31372.
[90]
Goncalves RLS, Rothschild DE, Quinlan CL, Scott GK, Benz CC, Brand MD. Sources of superoxide/H2O2 during mitochondrial proline oxidation. Redox Biology. 2014; 2: 901–909.
[91]
Quinlan CL, Orr AL, Perevoshchikova IV, Treberg JR, Ackrell BA, Brand MD. Mitochondrial complex II can generate reactive oxygen species at high rates in both the forward and reverse reactions. The Journal of Biological Chemistry. 2012; 287: 27255–27264.
[92]
Perevoshchikova IV, Quinlan CL, Orr AL, Gerencser AA, Brand MD. Sites of superoxide and hydrogen peroxide production during fatty acid oxidation in rat skeletal muscle mitochondria. Free Radical Biology & Medicine. 2013; 61: 298–309.
[93]
Quinlan CL, Goncalves RLS, Hey-Mogensen M, Yadava N, Bunik VI, Brand MD. The 2-oxoacid dehydrogenase complexes in mitochondria can produce superoxide/hydrogen peroxide at much higher rates than complex I. The Journal of Biological Chemistry. 2014; 289: 8312–8325.
[94]
Bleier L, Dröse S. Superoxide generation by complex III: from mechanistic rationales to functional consequences. Biochimica et Biophysica Acta. 2013; 1827: 1320–1331.
[95]
Robb EL, Hall AR, Prime TA, Eaton S, Szibor M, Viscomi C, et al. Control of mitochondrial superoxide production by reverse electron transport at complex I. The Journal of Biological Chemistry. 2018; 293: 9869–9879.
[96]
Guarás A, Perales-Clemente E, Calvo E, Acín-Pérez R, Loureiro-Lopez M, Pujol C, et al. The CoQH2/CoQ Ratio Serves as a Sensor of Respiratory Chain Efficiency. Cell Reports. 2016; 15: 197–209.
[97]
Korshunov SS, Skulachev VP, Starkov AA. High protonic potential actuates a mechanism of production of reactive oxygen species in mitochondria. FEBS Letters. 1997; 416: 15–18.
[98]
Chouchani ET, Pell VR, Gaude E, Aksentijević D, Sundier SY, Robb EL, et al. Ischaemic accumulation of succinate controls reperfusion injury through mitochondrial ROS. Nature. 2014; 515: 431–435.
[99]
Wang Y, Oxer D, Hekimi S. Mitochondrial function and lifespan of mice with controlled ubiquinone biosynthesis. Nature Communications. 2015; 6: 6393.
[100]
Genova ML, Lenaz G. Functional role of mitochondrial respiratory supercomplexes. Biochimica et Biophysica Acta. 2014; 1837: 427–443.
[101]
Maranzana E, Barbero G, Falasca AI, Lenaz G, Genova ML. Mitochondrial respiratory supercomplex association limits production of reactive oxygen species from complex I. Antioxidants & Redox Signaling. 2013; 19: 1469–1480.
[102]
Quinzii CM, López LC, Von-Moltke J, Naini A, Krishna S, Schuelke M, et al. Respiratory chain dysfunction and oxidative stress correlate with severity of primary CoQ10 deficiency. FASEB Journal: Official Publication of the Federation of American Societies for Experimental Biology. 2008; 22: 1874–1885.
[103]
Lenaz G, Genova ML. Mobility and function of coenzyme Q (ubiquinone) in the mitochondrial respiratory chain. Biochimica et Biophysica Acta. 2009; 1787: 563–573.
[104]
Hargreaves I, Heaton RA, Mantle D. Disorders of Human Coenzyme Q10 Metabolism: An Overview. International Journal of Molecular Sciences. 2020; 21: 6695.
[105]
Quinzii CM, Hirano M. Coenzyme Q and mitochondrial disease. Developmental Disabilities Research Reviews. 2010; 16: 183–188.
[106]
Rustin P, Munnich A, Rötig A. Mitochondrial respiratory chain dysfunction caused by coenzyme Q deficiency. Methods in Enzymology. 2004; 382: 81–88.
[107]
Trevisson E, DiMauro S, Navas P, Salviati L. Coenzyme Q deficiency in muscle. Current Opinion in Neurology. 2011; 24: 449–456.
[108]
DiMauro S, Schon EA, Carelli V, Hirano M. The clinical maze of mitochondrial neurology. Nature Reviews. Neurology. 2013; 9: 429–444.
[109]
Dadali T, Diers AR, Kazerounian S, Muthuswamy SK, Awate P, Ng R, et al. Elevated levels of mitochondrial CoQ10 induce ROS-mediated apoptosis in pancreatic cancer. Scientific Reports. 2021; 11: 5749.
[110]
Brea-Calvo G, Rodríguez-Hernández A, Fernández-Ayala DJM, Navas P, Sánchez-Alcázar JA. Chemotherapy induces an increase in coenzyme Q10 levels in cancer cell lines. Free Radical Biology & Medicine. 2006; 40: 1293–1302.
[111]
Mellors A, Tappel AL. The inhibition of mitochondrial peroxidation by ubiquinone and ubiquinol. The Journal of Biological Chemistry. 1966; 241: 4353–4356.
[112]
James AM, Smith RAJ, Murphy MP. Antioxidant and prooxidant properties of mitochondrial Coenzyme Q. Archives of Biochemistry and Biophysics. 2004; 423: 47–56.
[113]
Forsmark P, Aberg F, Norling B, Nordenbrand K, Dallner G, Ernster L. Inhibition of lipid peroxidation by ubiquinol in submitochondrial particles in the absence of vitamin E. FEBS Letters. 1991; 285: 39–43.
[114]
Lass A, Sohal RS. Electron transport-linked ubiquinone-dependent recycling of alpha-tocopherol inhibits autooxidation of mitochondrial membranes. Archives of Biochemistry and Biophysics. 1998; 352: 229–236.
[115]
Stocker R, Bowry VW, Frei B. Ubiquinol-10 protects human low density lipoprotein more efficiently against lipid peroxidation than does alpha-tocopherol. Proceedings of the National Academy of Sciences of the United States of America. 1991; 88: 1646–1650.
[116]
Rich PR. Electron and proton transfers through quinones and cytochrome bc complexes. Biochimica et Biophysica Acta. 1984; 768: 53–79.
[117]
Cirilli I, Damiani E, Dludla PV, Hargreaves I, Marcheggiani F, Millichap LE, et al. Role of Coenzyme Q10 in Health and Disease: An Update on the Last 10 Years (2010-2020). Antioxidants. 2021; 10: 1325.
[118]
Díaz-Casado ME, Quiles JL, Barriocanal-Casado E, González-García P, Battino M, López LC, et al. The Paradox of Coenzyme Q10 in Aging. Nutrients. 2019; 11: 2221.
[119]
Hernández-Camacho JD, García-Corzo L, Fernández-Ayala DJM, Navas P, López-Lluch G. Coenzyme Q at the Hinge of Health and Metabolic Diseases. Antioxidants. 2021; 10: 1785.
[120]
Langsjoen P, Langsjoen P, Willis R, Folkers K. Treatment of essential hypertension with coenzyme Q10. Molecular Aspects of Medicine. 1994; 15: S265–S272.
[121]
Zozina VI, Covantev S, Goroshko OA, Krasnykh LM, Kukes VG. Coenzyme Q10 in Cardiovascular and Metabolic Diseases: Current State of the Problem. Current Cardiology Reviews. 2018; 14: 164–174.
[122]
Samimi M, Zarezade Mehrizi M, Foroozanfard F, Akbari H, Jamilian M, Ahmadi S, et al. The effects of coenzyme Q10 supplementation on glucose metabolism and lipid profiles in women with polycystic ovary syndrome: a randomized, double-blind, placebo-controlled trial. Clinical Endocrinology. 2017; 86: 560–566.
[123]
Bentinger M, Dallner G, Chojnacki T, Swiezewska E. Distribution and breakdown of labeled coenzyme Q10 in rat. Free Radical Biology & Medicine. 2003; 34: 563–575.
[124]
Zhang Y, Turunen M, Appelkvist EL. Restricted uptake of dietary coenzyme Q is in contrast to the unrestricted uptake of alpha-tocopherol into rat organs and cells. The Journal of Nutrition. 1996; 126: 2089–2097.
[125]
Kamzalov S, Sumien N, Forster MJ, Sohal RS. Coenzyme Q intake elevates the mitochondrial and tissue levels of Coenzyme Q and alpha-tocopherol in young mice. The Journal of Nutrition. 2003; 133: 3175–3180.
[126]
Matthews RT, Yang L, Browne S, Baik M, Beal MF. Coenzyme Q10 administration increases brain mitochondrial concentrations and exerts neuroprotective effects. Proceedings of the National Academy of Sciences of the United States of America. 1998; 95: 8892–8897.
[127]
Van den Bergen CW, Wagner AM, Krab K, Moore AL. The relationship between electron flux and the redox poise of the quinone pool in plant mitochondria. Interplay between quinol-oxidizing and quinone-reducing pathways. European Journal of Biochemistry. 1994; 226: 1071–1078.
[128]
Galinier A, Carrière A, Fernandez Y, Bessac AM, Caspar-Bauguil S, Periquet B, et al. Biological validation of coenzyme Q redox state by HPLC-EC measurement: relationship between coenzyme Q redox state and coenzyme Q content in rat tissues. FEBS Letters. 2004; 578: 53–57.
[129]
Szibor M, Heyne E, Viscomi C, Moore AL. Measuring the Mitochondrial Ubiquinone (Q) Pool Redox State in Isolated Respiring Mitochondria. Methods in Molecular Biology. 2022; 2497: 291–299.
[130]
Burger N, Logan A, Prime TA, Mottahedin A, Caldwell ST, Krieg T, et al. A sensitive mass spectrometric assay for mitochondrial CoQ pool redox state in vivo. Free Radical Biology & Medicine. 2020; 147: 37–47.

Publisher’s Note: IMR Press stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share
Back to top