Next Article in Journal
Casticin Inhibits A375.S2 Human Melanoma Cell Migration/Invasion through Downregulating NF-κB and Matrix Metalloproteinase-2 and -1
Next Article in Special Issue
Recent Advances in the Analysis of Phenolic Compounds in Unifloral Honeys
Previous Article in Journal
Production of Fusaric Acid by Fusarium spp. in Pure Culture and in Solid Medium Co-Cultures
Previous Article in Special Issue
Targeting Reactive Carbonyl Species with Natural Sequestering Agents
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Coenzyme Q and Its Role in the Dietary Therapy against Aging

by
Alfonso Varela-López
1,
Francesca Giampieri
2,
Maurizio Battino
2,3,* and
José L. Quiles
1,*
1
Department of Physiology, Institute of Nutrition and Food Technology “José Mataix”, Biomedical Research Center (CIBM), University of Granada, Avda. del Conocimiento s.n., Armilla, Granada 18100, Spain
2
Dipartimento di Scienze Cliniche Specialistiche ed Odontostomatologiche (DISCO), Facoltà di Medicina, Università Politecnica delle Marche, Ancona 60131, Italy
3
Centre for Nutrition & Health, Universidad Europea del Atlantico (UEA), Santander 39011, Spain
*
Authors to whom correspondence should be addressed.
Molecules 2016, 21(3), 373; https://doi.org/10.3390/molecules21030373
Submission received: 9 February 2016 / Revised: 10 March 2016 / Accepted: 11 March 2016 / Published: 18 March 2016
(This article belongs to the Special Issue 20th Anniversary of Molecules—Recent Advances in Natural Products)

Abstract

:
Coenzyme Q (CoQ) is a naturally occurring molecule located in the hydrophobic domain of the phospholipid bilayer of all biological membranes. Shortly after being discovered, it was recognized as an essential electron transport chain component in mitochondria where it is particularly abundant. Since then, more additional roles in cell physiology have been reported, including antioxidant, signaling, death prevention, and others. It is known that all cells are able to synthesize functionally sufficient amounts of CoQ under normal physiological conditions. However, CoQ is a molecule found in different dietary sources, which can be taken up and incorporated into biological membranes. It is known that mitochondria have a close relationship with the aging process. Additionally, delaying the aging process through diet has aroused the interest of scientists for many years. These observations have stimulated investigation of the anti-aging potential of CoQ and its possible use in dietary therapies to alleviate the effects of aging. In this context, the present review focus on the current knowledge and evidence the roles of CoQ cells, its relationship with aging, and possible implications of dietary CoQ in relation to aging, lifespan or age-related diseases.

1. Coenzyme Q: Structure, Localization and Forms

Coenzyme Q (CoQ) is a naturally occurring molecule formed from the conjugation of a benzoquinone ring with a hydrophobic isoprenoid chain of varying chain length, depending on the species [1,2,3]. It is mainly located in the hydrophobic domain of the phospholipid bilayer of the inner membrane system of the mitochondria, but it is also present in all the other biological membranes at significant levels [4,5,6,7,8,9], as well as in plasma lipoproteins [10]. Moreover, CoQ is found in every plant and animal cell [2,11]. Due to its ubiquitous presence in nature and its quinone structure, it is also called ubiquinone [2,12]. Ubiquinone is referred to as “coenzyme” because of its unique ability to participate in chemical reactions but it remains at steady-state levels in the cell [3,13]. There are different ubiquinone molecules that are classified based on the length of their isoprenoid side chain with a subscript indicating the number of carbons in the chain (CoQn) [1,3]. CoQ9 (2,3-dimethoxy-5-methyl-6-noneprenyl-1,4-benzoquinone) is the predominant form in rats and mice, whereas in humans and other long-living mammals, the major homologue is CoQ10 (2,3-dimethoxy-5-methyl-6-decaprenyl-1,4-benzoquinone) [2,3,14]. The importance of the length of the polyisoprenyl chain is related to the stability of the molecule within the hydrophobic lipid bilayer. In addition, this feature seems to affect other properties, such as mobility, intermolecular interaction with membrane proteins, and autoxidizability [3,15,16].
The benzoquinone ring of CoQ can assume three alternate redox states due to the existence of different possible levels of protonation yielding three alternative CoQ forms (Figure 1): the fully oxidized (CoQ) or ubiquinone, the fully reduced (CoQH2) or ubiquinol and the partially reduced (CoQH) or ubisemiquinone [17,18]. Because of its extreme hydrophobicity, it is possible to find natural CoQ in three physical states: dissolved in lipid bilayers, forming micellar aggregates, or bound to proteins. In cells, CoQ is distributed between the two first states [18], whereas the importance of the last one is only experimental [17]. All cells are able to synthesize functionally sufficient amounts of this molecule under normal physiological conditions [19]. However, the content of CoQ as well as the ratios between its forms are different depending on the analyzed species, tissue or even organelle. For instance, murine kidney and heart show higher CoQ levels than brain or liver homogenates [20,21]. Likewise, it has been reported that lysosomes and Golgi membranes contain relatively higher concentrations of CoQ than mitochondrial membranes or microsomes [22,23]. In non-mitochondrial biological membranes, CoQ continuously cycles between reduced and oxidized states thanks to different enzymes with CoQ reductase activity [24]. These enzymes are NAD(P)H dehydrogenases which form part of the plasma membrane redox system, an electron transport where CoQ acts as a mediator accepting electrons from cytosolic NAD(P)H [6]. This system has been related to the maintenance of intracellular redox homeostasis, membrane antioxidant protection, regulation of cell signaling and other functions [6] that will be discussed below.

2. Physiological Roles of CoQ

2.1. Mitochondrial CoQ Is an Essential Factor for Cell Bioenergetics

CoQ was isolated and characterized in 1955 and only two years later it was shown to be an essential mitochondrial electron transport chain (mtETC) component [25] (Figure 2). Thus, CoQ is an essential factor for cell bioenergetics [6]. It acts as carrier of electrons from respiratory complexes I and II to complex III [23,26,27,28,29]. Additionally, CoQ also accepts electrons from other donors including dihydroorotate dehydrogenase, and acyl-CoA dehydrogenase. Since it has several electron donors but only transfers them to complex III, it may be considered a crossroad in the mitochondrial electron transport activity. Therefore, CoQ continuously cycles between the reduced and the oxidized forms in mitochondria, but it is also bound to proteins and in a membrane free form. The sum of CoQ in any of these possible states constitutes the mitochondrial CoQ pool [23]. In relation to this, Lopez-LLuch et al. [6] have proposed that an equilibrated CoQ pool may perform a better electron flow adaptation than a higher or lower CoQ pool by keeping a better mitochondrial homeostasis control. In addition, it has been reported that protein-bound CoQ participates in and is needed for the maintenance of complex III stability in yeast [30]. According to this finding, CoQ levels in mitochondria would not only affect cell bioenergetics by its participation as coenzyme in mtETC, but also could affect protein complex activity and structure [6]. Moreover, by means of plasma membrane redox system, CoQ is involved in balancing the NAD+:NADH ratio in cells, which also regulates cell bioenergetics [31].

2.2. CoQ Acts as Lipid-Soluble Antioxidant in All Biological Membranes

One of the main functions of CoQ in biological membranes is as an antioxidant [6] (Figure 2). Furthermore, it constitutes the only lipid-soluble antioxidant endogenously synthesized that has shown to efficiently prevent oxidation of proteins, lipids and DNA [32]. In this role, CoQ can act as antioxidant by itself or as free radical quencher [11,16,33]. It also participates in the regeneration of other antioxidants such as ascorbate [6,34,35] and α-tocopherol [33,36]. In particular, it has been proposed that the antioxidant effect of quinones in mitochondrial membranes is mediated by α-tocopherol recycling [6,37,38]. By itself, reduced CoQ10 has also demonstrated to be an important physiological lipid-soluble antioxidant [39]. In contrast to other antioxidants, ubiquinol can inhibit both the initiation and propagation of lipid peroxidation by reacting with perferryl radical and radicals generating ubisemiquinone and a non-radical lipid hydroperoxide [19,25]. Physiological ubisemiquinones having long side chains do not react with oxygen, except in complex III under the very special conditions of the CoQ-cycle [18]. On the other hand, short-chain ubiquinones have a pro-oxidant effect in complex I. Prevention of the lipid peroxidation propagation by the quenching of the initiating perferryl radicals also protects proteins from oxidation. In relation to this, it is also important to point out that ubiquinol effectively regenerates vitamin E from the α-tocopheroxyl radical which additionally contributes to slow the propagation step of lipid peroxidation. Lastly, CoQ has also been shown to prevent DNA base oxidation, which is particularly important in the case of mtDNA [19]. Furthermore, deficiency in some nutrients considered as antioxidants like vitamin E and selenium could be compensated by the induction of the CoQ-dependent antioxidant system [6]. This has been observed in hepatocytes where both, CoQ levels and CoQ reductases, have been shown to be affected by this compensatory mechanism [40]. The latter reinforces the role of CoQ as antioxidant particularly under some physiological situations such as nutritional deficiencies, which is particularly interesting because it is endogenously synthesized.
Explanations for the exceptionally high efficiency of CoQ as antioxidant include its intramembranous localization, its general and abundant distribution and its effective reduction/reactivation by a number of cellular systems [19,25]. Firstly, its localization is of central significance since ·OH and O2· generated in the membrane would otherwise rapidly react with neighboring lipid and protein molecules that necessitate the availability of effective protective agents close to the site of radical production [19]. Secondly, several effective enzymatic systems catalyze CoQ reduction to achieve its active (reduced) form in eukaryotic cells [32]. In mitochondria, the reduced antioxidant form of CoQ is regenerated by the respiratory chain [19]. Besides the mitochondrial respiratory chain, other enzymes with such function have been found. These include NADH-cytochrome b5 reductase that can reduce CoQ through a one-electron reaction mechanism [40,41] and the soluble enzyme NAD(P)H-quinone oxidoreductase 1 (NQO1) that can reduce quinones by a two-electron reaction [18,42]. This enzyme is induced under oxidative challenge and can maintain the electron flow through the plasma membrane redox system to CoQ when this system is working mainly as antioxidant [43,44]. In addition, a distinct cytosolic NADPH-CoQ reductase different from NQO1 has also been described [38], which seems to be a main factor for non-mitochondrial CoQ reduction [45,46,47]. However, more studies are needed to clarify its role and regulation under oxidative stress [6].
Lastly, it is important to note that ubiquinol is also present in lipoproteins, where it exerts its antioxidant activity. In particular, it has been reported that ubiquinol is the most efficient antioxidant in low-density lipoproteins (LDL) [48] which also contain α-tocopherol [19,48]. As a consequence of LDL protection from oxidation, CoQ would also have anti-atherosclerotic properties.

2.3. CoQ Collaborates in Redox State Regulation by Plasma Membrane Redox System Activity

It has been proposed that plasma membrane redox system activity becomes essential in the maintenance of bioenergetics in cells when activity in mitochondria decreases, as occurs in aging [6,49] (Figure 2). Actually, the plasma membrane redox system is up-regulated in cells lacking functional mitochondria [31,50,51]. In more detail, it has been reported that the amount of CoQ in plasma membrane increases and CoQ-dependent reductases are induced after mtDNA removal that induces the accumulation of cytosolic NADH. This mechanism allows cells to maintain the cytosolic NAD+:NADH ratio in response to increase in cytosolic NADH [31,51,52,53] which is necessary for glycolysis to occur correctly. CoQ is also able to act as an oxidant by autoxidation of ubisemiquinone form [11,16,33]. Based on its apparently paradoxical property to potentially act both as a pro-oxidant and an antioxidant, Sohal and Forster [23] suggest that CoQ may also be a modulator of the cellular redox state under physiological and/or pathological conditions affecting the aging process in this way.

2.4. CoQ Participates in Cell Signaling by Plasma Membrane Redox System

The activity of plasma membrane redox system also plays an important role in growth and development of organisms [6,33]. In this sense, CoQ-dependent NADH oxidases of plasma membrane have been shown to be involved in regulation of cell growth and differentiation. In particular, Gómez-Díaz et al. [34] found that the activity of CoQ reductases in plasma membrane is modulated during erythrocyte differentiation. Likewise, the vitamin D3-induced differentiation of myeloid cells to monocytes is enhanced when this system is activated by ascorbate [53,54]. Similarly, in serum-free cultures, addition of CoQ induces cell growth [55]. In support of the role of plasma membrane redox system in cell signaling, it has been reported that extracellular signaling molecule growth factors, insulin and pituitary extracts activated NADH oxidoreductases of plasma membrane [56]. At least in the case of myeloid cell differentiation, it has been observed that modulation of the differentiation program by plasma membrane redox system activation includes modulation of intracellular second messengers and regulation of the activity of transcription factors [54,57]. Findings from studies on this topic suggest different mechanisms by means of which CoQ might exert these effects on cell signaling, which include the following.
  • Reactive oxygen species (ROS): As previously stated, NAD(P)H dehydrogenases found in plasma membrane, cytochrome b5-reductase and other NAD(P)H dehydrogenases found in plasma membrane, reduce CoQ in a one-electron mechanism yielding semiquinone forms [58]. These CoQ forms have pro-oxidant activity of generating O2· or H2O2 that would act as second messengers on cell signaling mechanisms. Consequently, these would modulate different cell responses affecting cell growth and differentiation processes [59].
  • Tyrosine kinases: It has been suggested that the redox state of CoQ in plasma membrane could control the activity of tyrosine kinases indirectly by generation of H2O2 and further inactivation of protein phosphatases, or directly by induction of redox-dependent conformational changes [6,60].
  • Voltage-dependent anion channel (VDAC) proteins: As in mitochondria, plasma membrane also contains proteins belonging to the VDAC protein family. One of the components of this family, VDAC1, can function as NADH-ferricyanide reductase, an activity associated to the plasma membrane redox system [61]. Taken into consideration that CoQ is involved in the regulation of VDAC/permeability transition pore in mitochondria [62], López-Lluch et al. [6] have suggested a putative relationship of the activity of CoQ in plasma membrane and cell signaling linked to VDAC1 to be considered in future research.
  • NAD+-dependent deacetylases: This group of proteins implicated in genetic expression, such as sirtuins, could be affected and regulated in some manner by the activity of NADH-dependent reductase in plasma membrane [63]. It has been proposed that the variations in activity of CoQ-dependent NADH oxidoreductases in the different biological membranes could also regulate sirtuins because of the effects on redox state [6].
  • Mg2+-dependent neutral sphingomyelinase: It has been reported that CoQ-dependent plasma membrane redox system is involved in inhibition of Mg2+-dependent neutral sphingomyelinase after oxidative stress [64,65,66]. This is an integral plasma membrane protein involved in the release of ceramide from plasma membrane sphingomyelin and participates in cell signaling, apoptosis, and the modulation of cell responses [6,67].

2.5. CoQ Exerts Anti-Inflammatory Effects through Its Antioxidant Activity

CoQ exerts multiple anti-inflammatory effects by influencing the expression of Nuclear factor κB (NF-κB)1-dependent genes [68]. Likewise, H2O2 has been identified as an activator of the pro-inflammatory trancription factor NF-κB [69]. In view of the antioxidant properties of the reduced form of CoQ10 and the effective enzymatic conversion of oxidized CoQ10 into its reduced form, CoQ10 might mediate its observed anti-inflammatory effects via gene expression (Figure 2). For the study of the expression of genes implicated in inflammatory processes, in vitro models based on cells triggered with lipopolysaccharide (LPS) have aroused interest. This is because this bacterial product induces downstream signaling cascades of the transcription factor NF-κB, which in turn leads to the induction of inflammatory genes. In this context, it has been reported that CoQ10 down-regulates LPS-inducible genes in the monocytic cell line THP-1, presumably due to its antioxidant impact on gene expression [70]. Thus, as Genova and Lenaz [18] indicated, it is likely that all effects of CoQ at the genetic level may be mediated by its antioxidant effect.

2.6. Mitochondrial CoQ Prevents Events Leading to Programmed Cell Death

Different studies have confirmed the protective role of CoQ10 against apoptosis by inducing the inhibition of cell death independently from its free radical scavenging properties or antioxidant effects [62,71,72]. Presumably, this occurs by direct inhibition [62] of the opening of the permeability transition pores a high-conductance protein channel located in the internal mitochondrial membrane [73] (Figure 2). Permeability transition pore opening by its inducers has shown to lead to cell death [74]. Opening of the mitochondrial membrane transition pore depolarizes the mitochondrion [18], but also allows the translocation of molecules as large as 1500 Da in size [62] leading to the release of proteins present in mitochondrial intermembrane space into the cytosol. These last constitute different factors that trigger the process of programmed cell death like cytochrome c [18]. Thus, CoQ would counteract other downstream apoptotic events such as ATP depletion, caspase-9 activation, and DNA fragmentation [75,76]. Moreover, the increase of mitochondrial CoQ levels has been related to the lack of induction of permeability transition pore opening in diabetic rats [77].
To elucidate the possible mechanism under this effect, it is important to point out that permeability transition pore shows an ubiquinone binding site where different (natural and artificial) forms of CoQ interact stabilizing the pore in the closed conformation [75,76]. In particular, quinones have also been shown to exert a direct effect on permeability transition pores. These compounds are able to modulate the permeability transition pore through a common binding site rather than through oxidation-reduction reactions. Occupancy of this site can modulate the permeability transition pore open-closed transitions, possibly through secondary changes of the Ca2+-binding affinity for the pore [18]. Based on their relation with permeability transition pores, Walter et al. [76] distinguish three functional classes of quinone analogs:
  • permeability transition pores inhibitors, like CoQ0 ,CoQ2, and decylubiquinone.
  • permeability transition pores inducers, like idebenone 2,3-dimethoxy-5-methyl-6-(10-hydroxydecyl)-1,4-benzoquinone).
  • permeability transition pores-inactive quinones, which counteract the effects of both inhibitors and inducers, such as CoQ1.
An indirect study [78] suggests that CoQ10 may be a permeability transition pore inhibitor. These authors exposed SHSY5Y neuroblastoma cells to neurotoxic β-amyloid peptides and oxygen-glucose deprivation to investigate the neuroprotective effect of CoQ10. In these neuronal cells CoQ10 increased resistance against β-amyloid peptides-induced cell death that was related to the regulation of permeability transition pore opening. In addition, a decrease of O2· production was also noted [78]. Similar studies indicated a protective effect of CoQ10 on permeability transition pore opening against amitriptyline toxicity [79]. However, because ROS signaling could be implicated in permeability transition pore opening, it is not yet fully clear whether these CoQ10 effects on the transition pore depend on a direct mechanism or are mediated by the antioxidant effect [18].

2.7. CoQ Present in Lysosomal Membrane Participates in pH Maintenance

The NADH-dependent CoQ reductase present in lysosomal membrane is implicated in H+ translocation from cytosol to lysosomal lumen by means of an ATP-independent mechanism having O2 as the final electron acceptor [80] (Figure 2). As expected, this activity contributes to the maintenance of acidic pH into lysosomes. However, the exact role of this enzyme on lysosomal activity is currently under study [6,81].

3. Endogenous and Exogenous Sources of CoQ

3.1. CoQ Biosynthesis

CoQ endogenous bionsynthesis requires the synthesis of both a benzoquinone ring and an isoprenoid side chain. The precursor of the benzoquinone ring is 4-hydroxybenzoate that is synthesized from tyrosine or, at least theoretically, from phenylalanine [19]. Meanwhile, the isoprenoid side chain is synthesized by a series of reactions starting from acetyl-CoA and ending up with farnesyl pyrophosphate (farnesyl-PP), which comprises the mevalonate pathway [24,82,83]. Consequently, some reactions involved in CoQ bionsynthesis are shared with cholesterol and other lipid compounds [82]. Both the final product of this pathway, farnesyl-pyrophosphate (farnesyl-PP), and the intermediary of this pathway, isopentenyl pyrophosphate (isopentenyl-PP) are utilized for the synthesis of the isoprenoid side-chain of CoQ [19]. Then, the long isoprenoid side-chain of CoQ (which contains 6–10 isoprene units in different species) is synthesized by trans-prenyltransferase, which condenses farnesyl-PP with several molecules of isopentenyl-PP, all in the trans configuration [84].
Then, the 4-hydroxybenzoate-polyprenyl transferase, which is encoded by the gene Coq-2 in humans [85], acts by catalyzing the condensation of the isoprenoid side-chain with 4-hydroxybenzoate. After condensation, the benzoquinone ring undergoes a sequence of modifications including C-hydroxylations, decarboxylation, O-methylation and C-methylation to synthesize CoQ. This sequence has been studied mainly in bacteria and yeast. Meanwhile, in mammals, various functions of genes possibly implicated in these modifications have been established through the complementary recognition in yeast [19]. At the moment, six genes encoding different enzymes that catalyze reactions in this sequence have been identified in humans (termed from Coq-3 to Coq-8) [11,12]. However, full details of the synthesis of CoQ in animal tissues have not yet been clarified [19,86].
Endogenous CoQ10, synthesized using the radiolabelled precursor 14C-para-hydroxybenzoate (pHB), was first detected in mitochondria and later incorporated into mitochondria-associated membranes and endoplasmic reticulum, from where it is transported to other membranes in the cell [87]. This suggests that CoQ biosynthesis mostly occurs in mitochondria but it is redistributed to other organelles. Although this process is largely unknown, as mentioned, variations in the amounts present in different cellular organelles have been observed, with mitochondria, lysosomes and Golgi vesicles showing the highest concentrations [11]. Despite the fact that all cells are able to produce CoQ, its distribution is not uniform among different organs, being present at the highest concentrations in heart, kidney and liver [88]. Likewise, CoQ amounts can also vary among different tissues and structures of the same organ as occurs in bovine brain where CoQ10 levels range from 3 pg/g in the white matter to 25 pg/g in the striatum [89]. Moreover, interindividual variations in total CoQ10 have been described, as well as a significant difference in total CoQ10 between healthy males and females [90]. All these variations would indicate that CoQ10 concentrations are tightly distributed around a homeostatic set point in both organs and individuals. In any case, all cells should functionally synthesize sufficient amounts of CoQ under normal physiological conditions. Thus, in contrast to cholesterol, no redistribution via uptake from the circulation is required. As already mentioned, the liver releases a certain amount of CoQ that associates with very low-density lipoproteins (VLDLs), but this pool is not redistributed to other organs [19,91].
Under normal circumstances, it has been assumed that there is a lack of dependence on exogenous sources of CoQ, since all tissues are able to synthesize CoQ. However, situations may arise in which some cell synthetic capacity is insufficient to reach its CoQ10 requirements. Metabolically active cells presumably have the highest requirements for CoQ10 which explains that CoQ10 deficiency susceptibility appears to be greatest in them [3,12]. For this reason, deepening on our understanding about CoQ synthesis regulation, it is interesting to deal with different pathologies. Hydroxybenzoate is generally present in excess, so that the rate of this reaction is determined by the availability of the polyisoprenoid chain [19]. Farnesyl-PP produced by mevalonate pathway is precursor for cholesterol, CoQ, dolichol and isoprenylated proteins [82]. Moreover, the intermediary isopentenyl-PP is utilized for the synthesis of the isoprenoid side-chain of CoQ, but also for the synthesis of dolichol. It has been suggested that synthesis of all end-products of the mevalonate pathway is co-regulated since the initial sequence of reactions leading to such lipids is identical [19]. However, terminal regulation must also occur [19], which would explain greatly varying synthesis rates and different amounts found for these lipids [92]. The terminal points of regulation probably involve the branch-point enzymes that have farnesyl-PP as substrate which are considered to be rate-limiting for the terminal portion of the biosynthetic sequences [93]. These are squalene synthase, trans-prenyltransferase, cis-prenyltransferase and farnesyl- or geranylgeraniol-protein transferases for cholesterol, CoQ, dolichol and isoprenylated proteins. Traditionally, a central role in the regulation of the mevalonate pathway has been attributed to 3-hydroxy-3-metylglutaryl-CoA (HMG-CoA) reductase, but it seems that its primary regulatory role falls to cholesterol biosynthesis [19]. Since squalene synthase Michaellis constant (KM) for farnesyl-PP is high, it is expected that the total amount of farnesyl-PP in the cell exerts its major influence on cholesterol synthesis. Therefore, when this substrate concentration decreases, the enzyme is not saturated and the cholesterol synthesis rate is reduced [94]. In contrast, all of the other branch-point enzymes, i.e., trans- and cis-prenyltransferases, farnesyl- and geranylgeraniol-protein transferases, exhibit low KMs and remain saturated even when the farnesyl-PP pool is smaller.
As in other metabolic pathways, it seems that endogenous compounds also play regulatory roles on CoQ biosynthesis [19]. Epoxidated derivatives of certain all-trans polyisoprenoids, solanesol, tocotrienols, vitamin K2 and CoQ up-regulate CoQ synthesis in HepG2 cells by themselves. Additionally, it has been reported that several of these compounds inhibit the biosynthesis of cholesterol in the same cultures [95]. However, none of the epoxidated poly-cis polyisoprenoids, which occur naturally without an epoxy group in large numbers, exhibit any effect on products of the mevalonate pathway in this same system [19]. Tocotrienols exert the most profound effects on this biosynthetic pathway. It has been observed that dolichol synthesis is also up-regulated but the importance of this phenomenon has not been clarified since the function for non-phosphorylated form in animals has not been established despite its abundance. In general, improved CoQ biosynthesis is achieved by regulating entire biosynthetic machinery, while decreasing cholesterol synthesis specifically involves the inhibition of oxidosqualene cyclase. Therefore, it has been suggested that small amounts of mono- and diepoxides polyisoprenoids, as those having many organisms, could be biological regulators of the mevalonate pathway [19]. At translational level, some molecules have been shown to stimulate CoQ10 synthesis by Coq-7 expression regulation [96,97]. In particular, it has been found that CoQ10 biosynthesis is dependent on NF-κB, which binds specifically to two binding sites present in the 5’-flanking region of the Coq-7 gene, inducing both the Coq-7 expression and CoQ10 biosynthesis [98].

3.2. Dietary CoQ Has Shown to Increase CoQ Levels in Different Body Compartments

CoQ is a molecule naturally found in different dietary sources, which can be taken up from intestinal lumen in a similar way to other lipids [83]. According to results from different studies in rats [96,97,98,99,100,101,102,103], it has been historically considered that approximately 6% of orally administered CoQ permeates the gastrointestinal tract into the blood and is transferred to liver and spleen. Consequently, uptake in the whole body ranged between 2% and 3% of the total dose. Moreover, CoQ10 has been found in plasma, largely in the reduced form [100]. In vitro studies with CaCo-2 models of absorption have suggested that CoQ10 is reduced to ubiquinol either during or following absorption [104]. In turn, CoQ uptake by other tissues such as heart, kidney, brain and skeletal muscle was considered low or completely absent [100,102,103], unless endogenous levels fall below a critical threshold.
In humans, different supplementation studies have been carried out [105,106,107,108,109,110,111,112,113,114] varying in duration (20 days, 3 or even 16 months) and dosages (100–2400 mg/day), as well as in experimental group’s characteristics. Despite the differences, all showed an increase of CoQ10 levels in blood. Further, such increase seems to be a dose-dependent one according to a study where the effects of supplementation with 90, 150, and 300 mg/day of the reduced form of CoQ10 were compared [108]. In elderly women it was reported that the change was inversely associated with the baseline concentration [107], which suggests that CoQ uptake is also affected by levels present in the body. However, when CoQ10 concentration was assessed in other samples (homogenized muscle and mitochondria), no increase was noted [105]. Regarding how duration of dietary treatment affects CoQ body levels, it was observed that plasma ubiquinol concentration nearly reached steady-state by two weeks after the start of treatment, and plasma ubiquinol levels returned to those before administration six months after completion of treatment [108]. In another study [109], maximum plasma CoQ10 concentration was reached in two weeks of supplementation with all-trans form of CoQ10 and then decreased to basal level after withdrawal, but CoQ10 levels increased by a third compared to the study by Hosoe et al. using ubiquinol [108].
Results from more recent studies in mice and rats [20,21,110,111,115,116,117,118,119,120,121,122] firmly refute this long-held notion that CoQ content of tissues other than plasma, liver or spleen cannot be significantly augmented by dietary administration of CoQ10. It was first shown by Matthews et al. [110] that CoQ10 intake in diets for two months by 12- or 24-months-old rats increased CoQ content in brain mitochondria. Subsequently, a series of studies conducted by Sohal et al. demonstrated that CoQ10 administration (from 0.7 to 370 mg/Kg per day) via food to mice [20,21,116,120,121] and rats [113,115,116,117,118,119] caused an increase in amounts of CoQ10 in plasma and in tissue homogenates and mitochondria of brain, heart, kidney, skeletal muscle and liver. In all tissues, the amount of CoQ augmentation was greater in mitochondria than in the homogenate, suggesting its preferential sequestration in mitochondria [6]. In brain, the increase was of a lesser magnitude and occurred primarily in mitochondria and not in the homogenate. With this background, it seems more clear that different tissues tend to vary in their capacity for CoQ accretion, with liver and skeletal muscle exhibiting the highest elevations, and brain showing the lowest [23]. The mentioned studies in animals were performed with relatively high daily dietary amounts of CoQ suggesting that higher plasma CoQ10 concentrations are necessary to facilitate uptake by peripheral tissues [83,116], which could explains differences with earlier studies. However, it has also been observed that feeding rats with very low-dosages of CoQ10 from weaning leads to increase of mitochondrial CoQ levels of heart and liver in 12- and 24-months-old rats. Moreover, it has been observed that this difference observed between rats increases as the animals aged [118,123]. Thus, age at beginning and duration of the treatment must be taken into account too. Importantly, dietary CoQ10 also leads to an increase of the rat specific form CoQ9 in all tissues where it was measured [20,21,118].
Notwithstanding, as Sohal and Forster [23] pointed out, the uptake of CoQ and other lipid substances is a complex process dependent upon a considerable number of different factors. In that sense, it has been reported that intestinal absorption is threefold faster if CoQ10 is administrated with food intake in rats [124]. In mice, the increase of plasma levels per 100 mg values was remarkably higher with the reduced form of CoQ10 when compared to the results obtained with the oxidized form CoQ10 [22]. Additionally, other aspects of CoQ form also seem to be important. Villalba et al. [83] reviewed the efficacy of a variety of commercial formulations that have been developed to solubilize CoQ10 and promote its better absorption in vivo, and its use in the therapy of pathologies associated with low CoQ levels with emphasis in the results of the clinical trials. They concluded that the relative bioavailability of CoQ10 is dependent on the type and amounts of oil in the formulations as well as on its delivery system, the order of decreasing bioavailability being: nanoparticulated, solubilized, oil-emulsioned and powder. Some additional dietary factors have been shown to influence CoQ contents in the body and they could also influence the magnitude of dietary CoQ effect. In that sense, it has been reported that monounsarturated fatty acids (MUFA)-rich dietary fats increased CoQ mitochondrial contents, whereas these decreased in diets rich in polyunsarturated fatty acids (PUFA) [125]. In the same sense, it has been observed that a diet rich in saturated fat for 50 days decreased CoQ levels in plasma and liver mitochondria of rabbits, which was restored by oral administration of soluble CoQ10 (25 mg/kg per day) over 30 days after such period [126].
Regarding CoQ supplements safety, different assessments in human and animals reviewed by Hidaka et al. [127] concluded that the endogenous biosynthesis of CoQ10 is not influenced by exogenous inputs. Moreover, it does not accumulate into plasma or tissues when supplementation ends. In rats, a 52-week chronic toxicity study has indicated that the acceptable daily intake is 12 mg/kg per day, calculated from the no-observed-adverse-effect level of 1200 mg/kg per day. In humans, clinical trials suggest an observed safety level for CoQ10 was 1200 mg/day per person. In addition, doses up to 3000 mg/day of CoQ10 did not cause serious adverse effects in humans. However, several moderate adverse effects as nausea and other adverse gastrointestinal effects have been reported, although these effects were not causally related to the active ingredient because there was no dose-response relationship [127]. At very high amounts (2.6 mg/g), prolonged intake of CoQ10 has shown to exacerbated cognitive and sensory impairments in old mice. In turn, intake at lower amounts had no discernable negative impact on these functions [128].

4. Evidence for CoQ as Anti-Aging Compound

4.1. CoQ Levels Are Affected by Aging

It has been proposed that along the entire life of organisms, the synthesis of CoQ changes and is significantly reduced from the initial phases of aging [10,129]. This statement seems to contradict findings from C. elegans where older individuals showed increases in CoQ content [130]. However, measures of CoQ amount in mammals, mostly rodents, have differed. In mice, age-related decreases of CoQ levels have been observed in liver homogenates [130] as well as in mitochondrial CoQ levels of skeletal muscle, but not of kidney, brain or heart [120]. Like in mice, liver CoQ levels also decrease with age in pigs (Sus scrofa domestica) [130]. However, mitochondria of liver, heart and kidney from 19-month-old ad libitum fed rats showed lower CoQ9 than those from 4-month-old animals, although this CoQ depletion was not detectable in tissue homogenates [131]. Authors attributed this fact to the preferential sequestration of CoQ in the mitochondrial fraction. Similarly, no changes in the levels of CoQ in rat brain after an increase during the first months of life have been reported [132], but in this organ no changes were observed when only mitochondria were evaluated [133]. On the other hand, an age-associated decrease of CoQ concentration was detected in pancreas and adrenal gland, brain, heart and lung after an early increase like in brain reaching highest values at 30 days of age [134]. In an old study, CoQ increased between 2 and 18 months and decreased significantly at 25 months in the heart and kidney, and the gastrocnemius, oblique and deep aspect (red) of vastus lateralis muscles. On the other hand, CoQ concentration of rat liver increased over the life span, while it remained relatively constant in brain, lung, and the superficial aspect (white) of the vastus lateralis muscle [135]. These results suggest that decrease could appear relatively later in life explaining why some studies do not find clear age-associated changes.
Results of the various studies described do not support the existence of a common trend for all living organisms, although age-associated changes in CoQ content are most evident in mitochondria as was suggested by Sohal and Forster [23]. In mammals, CoQ levels tend to decrease with aging, but this depends on tissues and organisms and probably many other factors. This could explain, at least in part, the differences found in aging processes among different organisms, and also why some tissues are more susceptible to aging or aging-related diseases. Studies evaluating the association between age and CoQ concentration in humans have also supplied contradictory results. In elderly women, no significant correlations were found between CoQ10 plasma levels and age [107]. In human pancreas and adrenal gland, CoQ10 levels were highest at one year of age, and then they decrease, whereas in the brain, heart and lung, the corresponding peak value was at 20 years of age and was followed by a continuous decrease upon further aging [134]. Age-associated decrease in brain has been confirmed by later studies [136,137]. In fact, it has been suggested that the mevalonate pathway is affected in the aged brain [137].

4.2. Several Age-Related Pathologies Are Associated with Low Levels of CoQ

Diseases such as cardiovascular diseases, neuropathies, inflammation, metabolic syndrome, arthritis, carcinogenesis, diabetes or hypercholesterolemia worsen during aging and are considered as major age-related diseases [6]. Different studies have reported a beneficial effect of CoQ on them [138,139,140,141,142], almost all in animal models. This fact has been taken into account to suggest that CoQ becomes an essential factor in the maintenance of the normal activity of cells in such conditions [6]. However, it is important to note that some interventions in humans in relation to Parkinson disease [112] and Amyotrophic Lateral Sclerosis [113] have not offered clear benefits for CoQ on disease progression.
These and other neurodegenerative diseases have been widely studied in this respect. Case-control studies have reported lower CoQ10 levels in serum of Lewy’s body disease patients [143] and plasma of subjects with amyotrophic lateral sclerosis [144] compared with healthy controls. Likewise, lower levels of this molecule have been found in cerebrospinal fluid of Alzheimer´s disease patients [145]. In addition, a negative correlation has been reported between CoQ levels and the duration of illness in Amyotrophic Lateral Sclerosis [145], Alzheimer’s [146] and Parkinson’s disease case studies [147]. These observations, together with the decrease of mitochondrial activity and the increase of free radical levels found in neuroneurodegenerative diseases [147,148,149,150], highlight the importance that CoQ has in these processes as suggested by Lopez-Lluch et al. [6]. Actually, it has been reported that NQO1 expression increases during the initial steps of Alzheimer’s disease, indicating a higher lipid peroxidation coupled to a higher necessity for CoQ-dependent antioxidant activity [6,151].
Some authors have also focused their studies on a possible relationship between CoQ amount and some metabolic syndrome components as risk factors of cardiovascular disease. An important component associated to aging is hypercholesterolemia since it affects a significant part of the aged population. In elderly women, no differences were seen between hyperlipidemic and normolipidemic subjects in relation to serum CoQ10 levels. Similarly, no significant correlations were found with body mass index, another risk factor for cardiovascular disease [108]. However, as indicated above, CoQ is the main antioxidant in LDL [48]. In relation to this, it has been shown that total CoQ10: total cholesterol ratio was reduced in diabetes mellitus patients, another component of metabolic syndrome, compared with subjects with normal glucose tolerance and impaired fasting glucose [152].
The development of certain types of cancer also could be related with low CoQ levels. Epidemiological studies in humans have shown negative associations between CoQ levels in blood and breast cancer [153,154], myeloma [153], melanoma [155], Graves’ disease or follicular and papillary thyroid carcinomas [156]. Some studies have also suggested that tumoral cells have lower CoQ levels than normal cells. In women presenting carcinomas and non-malignant-breast cancer, CoQ10 concentrations in breast tumor tissues significantly decreased when compared to the surrounding normal tissues [157]. Similarly, melanoma cell lines also presented low concentrations of CoQ10 [155]. In relation to these groups of diseases, it has been found that CoQ treatment decreases cell growth of a prostate cancer line (PC3 line) but not of non cancer cells (PNT2) and was associated with a high production of ROS [158], suggesting an interesting use for this molecule in cancer therapy.

4.3. Aging, Development and Lifespan Are Associated to Changes in CoQ Bionsynthesis

In mammals, there are at least 10 different proteins participating in the biosynthesis of CoQ that are encoded by Coq genes. These have both catalytic and regulatory activities. The first group is encoded by the genes Coq-1, Coq-2, Coq-3, and Coq-7, whereas it has been suggested that Coq-4, Coq-8 and Coq-9 gene products have regulatory functions [6,84]. Different studies using animal model mutants for these genes have been carried out to elucidate the role of endogenous CoQ synthesis on aging. C. elegans models have been extensively used in this way. clk-1 (mammals Coq-7 orthologue gene) mutant nematodes that produce a very low amount of CoQ [159] showed an increased lifespan compared to wild-type animals [160,161]. However, analysis of mutants in Coq genes, other than clk-1, provided different results. In most of cases, Coq genes knockout showed deleterious defects leading to early developmental arrest [162,163,164]. This has been reported for Coq-1, Coq-3, Coq-3 and Coq-8. Extension of longevity by silencing Coq genes [165] occurs upon moderate low levels of global CoQ content (up to 50%) but not in the case of more severe CoQ depletions as observed in Coq-8 [162] and Coq-3 [164] mutants. To explain this difference it has been proposed that moderated CoQ depletion is associated to lower ROS production extending lifespan [165], whereas a higher CoQ depletion would lead to developmental and reproductive inefficiency [162,164].
According to findings from the studies with C. elegans, it has also been suggested that Coq-7 and CoQ synthesis is also related to aging in humans [166] and other mammals. However, experimental evidence suggests that important differences could exist between mammals and C. elegans in this sense. In contrast to observations in nematodes, murine Coq-7 knockout embryos arrest development at midgestation [167]. Similarly, it has been reported that some deletions in Coq-7 affect mitochondrial integrity and neurogenesis [168], which is comparable to some effects of CoQ deficiency found in humans [169,170]. However, heterozygous mice for this gene carrying a single functional copy show a notable increase in lifespan [171].
Mutations in genes implicated in CoQ biosynthesis have been identified as a cause of different pathologies associated with CoQ deficiencies in humans, most of which appear early in life [172,173,174,175,176]. The most severe human CoQ10 deficiencies are due to autosomal recessive mutations and can be classified as primary deficiencies when mutations affect CoQ10 biosynthetic genes, or secondary if the cause is related to other genetic defects [177]. Hereditary CoQ10 deficiencies caused by these mutations usually lead to cardiomyopathies and degenerative muscle and neuronal diseases [169,173,178,179,180]. The major phenotypes provoked by CoQ10 deficiencies are encephalomyopathy, severe infantile multisystemic disease, cerebellar ataxia, Leigh syndrome with growth retardation, ataxia and deafness, and isolated myopathy [181].

5. Studies on Dietary Therapies with CoQ on Aging

According to previous observations, dietary supplementation with CoQ10 could constitute an anti-aging strategy. In humans, there is evidence, mainly indirect, that exogenous orally administered CoQ10 may be incorporated into mitochondria, at least in conditions of partial CoQ tissue deficiency, where it may enhance electron transfer and ATP synthesis with improvement of pathological situations such as cardiac failure [182,183], Parkinson’s disease [115,184,185,186], Alzheimer’s disease [187,188,189] and Friedreich’s ataxia [190].
Results from animal studies are not clear about the dietary CoQ effects on longevity. In C. elegans, it has been reported that dietary CoQ prolonged lifespan [162], but this was also noted with a CoQ-deficient diet [191]. One possible explanation for poor diet effect could be in the adaptability of these nematodes to stressful conditions. C. elegans life span is extended by the intake of antimycin A, an mtETC inhibitor, whereas this molecule is toxic to most other aerobic species [192]. According to this observation, CoQ deficiency might induce a hypometabolic or a dauer-like state, which would facilitate survival under adverse conditions. However, a study focused on features of bacteria used to feed worms has evidenced that this phenomenon may be more complex. In this referred study, a diet based on respiratory incompetent E. coli, regardless if they were CoQ-less or CoQ-replete, produced a robust life extension in wild-type C. elegans [193]. An explanation for these observations was that the fermentation-based metabolism of the E. coli diet is an important parameter of C. elegans longevity [193].
As in invertebrates, simple dietary CoQ supplementation has shown no direct concluding results on lifespan extension in rodents. Information from this model indicates that CoQ10 supplementation with daily dosages ranged from 10 to 370 mg/Kg has no effect on longevity [117,119,129,194]. Despite the absence of evidence supporting that dietary CoQ can increase lifespan in animals, some interventions in the same way have been found to retard certain aging detrimental aspects in different animal models for aging or age-related diseases. In SAMP mice, a mouse model for accelerated senescence and severe senile amyloidosis, life-long supplementation with CoQH2 substantially decreased the senescence grading scores at different ages, although it did not alter some age-associated features of the model like the senile amyloid deposition rate. Again, this intervention did not have an effect on the lifespan [195]. In older mice with clear cognitive and psychomotor impairments, short-time (15 days) CoQ-supplementation improved spatial learning [196]. A cardinal event of diabetes like diabetic neuropathy has also been reported to be positively modified upon CoQ administration in diabetic rats [140]. In the hypercholestolemic ApoE knockout mouse, dietary CoQ had an anti-atherogenic effect preventing the accumulation of lipid peroxides in aorta [139]. In turn, in most of cases beneficial effects of CoQ over mitochondrial function and oxidative stress have been demonstrated [139,143,196,197]. However, long-term CoQ10 intake in healthy mice fed a standard diet failed to modulate mitochondrial respiratory capacity in liver or levels of oxidative stress in liver, kidney, skeletal muscle or brain [23,119]. According to different findings, López-LLuch et al. [6] suggested that supplementation with CoQ is not needed when the organism is young and healthy because cell membranes seem to be nearly saturated at the functional level. However, this supplementation becomes necessary when the organism shows deficiency, as in aging. This could explain why CoQ effects are clearer in animal models of disease than on lifespan of healthy animals if they are related to low CoQ levels.
On the other hand, the combination of dietary treatments using CoQ supplements in certain nutritional conditions associated with higher oxidative stress levels and age-related detrimental effects offers interesting expectations. From this standpoint, studies comparing CoQ effects between isocaloric diets with different lipid profile are particularly interesting. The effects of long-term supplementation with daily CoQ10 at 0.7 mg/kg on rats fed on MUFA-rich diets have been compared with those found in n-6 PUFA-rich diets [118,123,198,199,200,201]. One of the most interesting findings from such studies was that dietary CoQ10 produced significant increases of mean and maximum lifespan in rats fed a diet rich in n-6 PUFA [123,199,200]. At the histopathological level, when sunflower oil was the main fat in the diet, CoQ supplementation seemed to improve endocrine pancreas structure and in particular β-cell mass resembling positive effects of virgin olive oil [201]. Similar effects were noted in rat alveolar bone loss associated to aging [202]. Dietary CoQ treatments have also been shown to be effective in counteracting many of the high-fat diet consequences in animals [129,203,204,205,206,207,208]. In other mouse models, post-weaning dietary supplementation with CoQ10 rescued many of the detrimental effects of nutritional programming on cardiac aging by low birth-weight and catch-up growth [197].
The biochemical basis of potential beneficial effects of CoQ on lifespan or other aging detrimental effects may include enhancement of the cellular antioxidant protection systems in cell membranes, where CoQ sustains lipids in its reduced redox state preventing lipid peroxidation, particularly the unstable PUFA [6,209]. In previous studies in rats, diets containing CoQ were associated with lower lipid peroxidation markers [118,199], as well as with lower oxidative damage of other macromolecules such as DNA or proteins. A higher antioxidant capacity [199,200] compared to that found in animals maintained on the same diet without additional CoQ10 has been reported [118,122,198,202]. In addition, a lower impairment in mitochondrial function was also observed in CoQ-fed animals [118]. All these findings would indicate that dietary CoQ10 avoids, at least in part, oxidative stress linked to aging under certain conditions. Furthermore, it has been shown that life-long dietary supplementation with CoQ10 attenuated a variety of changes in enzymatic activities associated with aging in rats [209,210]. These include increases in the hepatic activities of Mg2+-dependent sphingomielinase [209] and of cytosolic and membrane-bound NQO1 activities [210], as well as decreases in cytosolic glutathione-S-transferase and microsomal Se-independent glutathione peroxidase in liver plasma membrane [209]. Proteomic analysis in rats under similar conditions has shown that serum albumin, which decreases with age in the rat, was significantly increased by CoQ10 supplementation. Additionally, it induced significant modifications of several proteins in plasma. These modifications support the beneficial role of dietary CoQ10 decreasing both oxidative stress and cardiovascular risk, and modulating inflammation and osteogenesis during aging [211].
In humans, some studies have suggested similar effects for dietary CoQ in relation to oxidative stress, at least in combination with certain dietary patterns. Short-term (4 weeks) dietary CoQ effects on aging have been tested in combination with the Mediterranean diet. In this regard, elderly subjects ingested a Western diet rich in saturated fatty acids (SFA), a Mediterranean diet (rich in MUFA), and a Mediterranean diet supplemented with CoQ following a cross-over design [212,213,214,215,216]. CoQ10 addition to MUFA-rich diet reduced some postprandial oxidative stress marker levels when subjects took a breakfast with a lipid profile similar to their experimental diets [212], which correlated with a lower expression of antioxidant enzyme components [212,215]. Moreover, dietary CoQ has also been shown to improve DNA repair systems [213,214] and modulate inflammatory signaling cascade as well as to reduce endoplasmic reticulum stress [214].
All these results suggest that although CoQ supplementation does not directly extend lifespan, it may help to prevent life span shortening due to oxidative insults [6] as it has been suggested by its effect in all aspect related to mitochondrial function, oxidative stress and antioxidant defenses both in animals [139,196,197] and humans [114,143]. However, despite animal studies have shown certain beneficial effects on the health of different disease models, there are clinical trials reported no significant effects on the progression of some nervous central system disease [113,114]. As has been stated, brain CoQ uptake is very low compared to other organs [23] and animal studies reported beneficial effects on this organ [195,196] have used very high daily dosages in relation to body weight compared with those used in clinical trials [113,114]. Future trials in humans focused on diseases affecting tissues and organs that have shown low CoQ uptake capacity should be more careful at this respect and to use higher CoQ dosages and/or chemical formulation with higher bioavailability such as those nanoparticulated or solubilized [83]. In the same sense, possible differences in bioavailability and efficacy between ubiquinol and ubiquinone should also be taken into account. On the other hand, CoQ effectivity for different disease treatment could depend on many other conditions not considered in clinical trials: among others, possible differences in etiology and pathophysiology among animal models and human diseases.
Due to differences in dosage, duration, chemical formulation and subject age at the beginning of the treatment, it is very difficult to establish “ideal conditions” for CoQ use as “anti-aging therapy”. Notwithstanding, according some to animal studies [118,122,198,202], life-long dietary interventions could result useful to prevent negative consequences of different health insults throughout life, particularly in relation to lifestyle. However, ubiquinone dosages used (0.7 mg/kg) already seem very high to take them. In this sense, supplementing with some foods, particularly those that are more “pro-oxidant”, could result useful.

Acknowledgments

F.G. was supported by a Fondazione Umberto Veronesi Fellowship. A.V.-L. is recipient of a predoctoral FPU contract from the Spanish Ministry of Education, Culture and Sport and member of the research group CTS-627 supported by Junta de Andalucia and University of Granada.

Author Contributions

A.V-L and F.G. wrote the review. Figures were drawn by A.V.-L. M.B. and J.L.Q. proposed the scheme, supervised and corrected the review.

Conflicts of Interest

The authors declare no conflict of interest.

Abbreviations

The following abbreviations are used in this manuscript:
CoQCoenzyme Q
CoQHpartially reduced coenzyme Q
CoQH2fully reduced coenzyme Q
Cytb5RNADH: cytochrome b5 reductase
FAD+reduced flavine adenine dinucleotide
FADH2reduced flavine adenine dinucleotide
farnesyl-PPfarnesyl-pyrophosphate(isopentenyl-PP)
HMG-CoA3-hydroxy-3-metylglutaryl-CoA
LDLlow-density lipoprotein
LOOlipid hydroperoxyl radical
LOOHlipid hydroperoxyde
LPSlipopolysaccharide
MAPKsmitogen-activated protein kinases
mPTPmitochondrial permeability transition pore
mtDNAmitochondrial DNA
mtETCmitochondrial electron transport chain
MUFAmonounsarturated fatty acid
NQO1NAD(P)H-quinone oxidoreductase 1
pHBpara-hydroxybenzoate
PUFApolyunsarturated fatty acids
ROSReactive oxygen species
SFAsaturated fatty acids
SIRT1sirtuin1
VDACVoltage-dependent anion channelNuclear factor κB (NF-κB)
VLDLvery low-density lipoprotein
α-tpα-tocopherol
α-tp*α-tocopheryl anion radical

References

  1. Cluis, C.P.; Burja, A.M.; Martin, V.J. Current prospects for the production of coenzyme Q10 in microbes. J. Trends Biotechnol. 2007, 25, 514–521. [Google Scholar]
  2. Prakash, S.; Sunitha, J.; Hans, M. Role of coenzyme Q10 as an antioxidant and bioenergizer in periodontal diseases. Indian J. Pharmacol. 2010, 42, 334. [Google Scholar] [CrossRef] [PubMed]
  3. Varela-López, A.; Bullón, P.; Giampieri, F.; Quiles, J.L. Non-Nutrient, Naturally Occurring Phenolic Compounds with Antioxidant Activity for the Prevention and Treatment of Periodontal Diseases. Antioxidants 2015, 4, 447–481. [Google Scholar] [CrossRef] [PubMed]
  4. Battino, M.; Ferri, E.; Gorini, A.; Villa, R.F.; Rodriguez Huertas, J.F.; Fiorella, P.; Genova, M.L.; Lenaz, G.; Marchetti, M. Natural distribution and occurrence of coenzyme Q homologues. Membr. Biochem. 1990, 9, 179–190. [Google Scholar] [CrossRef] [PubMed]
  5. Lenaz, G.; Fato, R.; Di Bernardo, S.; Jarreta, D.; Costa, A.; Genova, M.L.; Parenti Castelli, G. Localization and mobility of coenzyme Q in lipid bilayers and membranes. Biofactors 1999, 9, 87–93. [Google Scholar] [CrossRef] [PubMed]
  6. López-Lluch, G.; Rodríguez-Aguilera, J.C.; Santos-Ocaña, C.; Navas, P. Is coenzyme Q a key factor in aging? Mech. Ageing Dev. 2010, 131, 225–235. [Google Scholar] [CrossRef] [PubMed]
  7. Samorì, B.; Lenaz, G.; Battino, M.; Marconi, G.; Domini, I. On coenzyme Q orientation in membranes: A linear dichroism study of ubiquinones in a model bilayer. J. Membr. Biol. 1992, 128, 193–203. [Google Scholar] [CrossRef] [PubMed]
  8. Fato, R.; Battino, M.; Castelli, G.P.; Lenaz, G. Measurement of the lateral diffusion coefficients of ubiquinones in lipid vesicles by fluorescence quenching of 12-(9-anthroyl) stearate. FEBS Lett. 1985, 179, 238–242. [Google Scholar] [CrossRef]
  9. Lenaz, G.; Samori, B.; Fato, R.; Battino, M.; Castelli, G.P.; Domini, I. Localization and preferred orientations of ubiquinone homologs in model bilayers. Biochem. Cell Biol. 1992, 70, 504–514. [Google Scholar] [CrossRef] [PubMed]
  10. Crane, F.L. Biochemical functions of coenzyme Q10. J. Am. Coll. Nutr. 2001, 20, 591–598. [Google Scholar] [CrossRef] [PubMed]
  11. Turunen, M.; Olsson, J.; Dallner, G. Metabolism and function of coenzyme Q. Biochim. Biophys. Acta 2004, 1660, 171–199. [Google Scholar] [CrossRef] [PubMed]
  12. Gaby, A.R. The role of coenzyme Q10 in clinical medicine: Part I. Altern. Med. Rev. 1996, 1, 11–17. [Google Scholar]
  13. Littarru, G.P.; Tiano, L. Bioenergetic and antioxidant properties of coenzyme Q10: Recent developments. Mol. Biotechnol. 2007, 37, 31–37. [Google Scholar] [CrossRef] [PubMed]
  14. Bhagavan, H.N.; Chopra, R.K. Coenzyme Q10: Absorption, tissue uptake, metabolism and pharmacokinetics. Free Radic. Res. 2006, 40, 445–453. [Google Scholar] [CrossRef] [PubMed]
  15. Matsura, T.; Yamada, K.; Kawasaki, T. Difference in antioxidant activity between reduced coenzyme Q9 and reduced coenzyme Q10 in the cell: Studies with isolated rat and guinea pig hepatocytes treated with a water-soluble radical initiator. Biochim. Biophys. Acta 1992, 1123, 30–315. [Google Scholar] [CrossRef]
  16. James, A.M.; Smith, R.A.J.; Murphy, M.P. Antioxidant and prooxidant properties of mitochondrial Coenzyme Q. Arch. Biochem. Biophys. 2004, 423, 47–56. [Google Scholar] [CrossRef] [PubMed]
  17. Fato, R.; Battino, M.; Degli Esposti, M.; Parenti Castelli, G.; Lenaz, G. Determination of partition and lateral diffusion coefficients of ubiquinones by fluorescence quenching of n-(9-anthroyloxy) stearic acids in phospholipid vesicles and mitochondrial membranes. Biochemistry 1986, 25, 3378–3390. [Google Scholar] [CrossRef] [PubMed]
  18. Genova, M.L.; Lenaz, G. New developments on the functions of coenzyme Q in mitochondria. Biofactors 2011, 37, 330–354. [Google Scholar] [CrossRef] [PubMed]
  19. Bentinger, M.; Tekle, M.; Dallner, G. Coenzyme Q–biosynthesis and functions. Biochem. Biophys. Res. Commun. 2010, 396, 74–79. [Google Scholar] [CrossRef] [PubMed]
  20. Lass, A.; Forster, M.J.; Sohal, R.S. Effects of coenzyme Q10 and alpha-tocopherol administration on their tissue levels in the mouse: Elevation of mitochondrial alpha-tocopherol by coenzyme Q10. Free Radic. Biol. Med. 1999, 26, 1375–1382. [Google Scholar] [CrossRef]
  21. Lass, A.; Sohal, R.S. Effects of coenzyme Q10 and alphatocopherol administration on their tissue levels in the mouse: Elevation of mitochondrial alpha-tocopherol by coenzyme Q10. FASEB J. 2000, 14, 87–94. [Google Scholar] [PubMed]
  22. Dallner, G.; Sindelar, P.J. Regulation of ubiquinone metabolism. Free Radic. Biol. Med. 2000, 29, 285–294. [Google Scholar] [CrossRef]
  23. Sohal, R.S.; Forster, M.J. Coenzyme Q, oxidative stress and aging. Mitochondrion 2007, 7, S103–S111. [Google Scholar] [CrossRef] [PubMed]
  24. Ernster, L.; Dallner, G. Biochemical, physiological and medical aspects of ubiquinone function. Biochim. Biophys. Acta 1995, 1271, 195–204. [Google Scholar] [CrossRef]
  25. Mitchell, P. Protonmotive redox mechanism of the cytochrome b-c1 complex in the respiratory chain: Protonmotive ubiquinone cycle. FEBS Lett. 1975, 56, 1–6. [Google Scholar] [CrossRef]
  26. Bentinger, M.; Brismar, K.; Dallner, G. The antioxidant role of coenzyme Q. Mitochondrion 2007, 7, S41–S50. [Google Scholar] [CrossRef] [PubMed]
  27. Battino, M.; Fato, R.; Parenti-Castelli, G.; Lenaz, G. Coenzyme Q can control the efficiency of oxidative phosphorylation. Int. J. Tissue React. 1989, 12, 137–144. [Google Scholar]
  28. Rauchová, H.; Battino, M.; Fato, R.; Lenaz, G.; Drahota, Z. Coenzyme Q-pool function in glycerol-3-phosphate oxidation in hamster brown adipose tissue mitochondria. J. Bioenerg. Biomembr. 1992, 24, 235–241. [Google Scholar] [CrossRef] [PubMed]
  29. Forbes-Hernández, T.Y.; Giampieri, F.; Gasparrini, M.; Mazzoni, L.; Quiles, J.L.; Alvarez-Suarez, J.M.; Battino, M. The effects of bioactive compounds from plant foods on mitochondrial function: A focus on apoptotic mechanisms. Food Chem. Toxicol. 2014, 68, 154–182. [Google Scholar] [CrossRef] [PubMed]
  30. Santos-Ocaña, C.; Do, T.Q.; Padilla, S.; Navas, P.; Clarke, C.F.J. Uptake of exogenous coenzyme Q and transport to mitochondria is required for bc1 complex stability in yeast coq mutants. Biol. Chem. 2002, 277, 10973–10981. [Google Scholar] [CrossRef] [PubMed]
  31. Larm, J.A.; Vaillant, F.; Linnane, A.W.; Lawen, A.J. Up-regulation of the plasma membrane oxidoreductase as a prerequisite for the viability of human Namalwa rho 0 cells. Biol. Chem. 1994, 269, 30097–30100. [Google Scholar]
  32. Ernster, L. Lipid peroxidation in biological membranes: Mechanisms and implications. In Active Oxygens, Lipid Peroxides, and Antioxidants; Yagi, K., Ed.; CRC Press: Boca Raton, FA, USA, 1993; pp. 1–38. [Google Scholar]
  33. Crane, F.L.; Navas, P. The diversity of coenzyme Q function. Mol. Aspects Med. 1997, 18, S1–S6. [Google Scholar] [CrossRef]
  34. Gómez-Díaz, C.; Rodríguez-Aguilera, J.C.; Barroso, M.P.; Villalba, J.M.; Navarro, F.; Crane, F.L.; Navas, P. Antioxidant ascorbate is stabilized by NADH-coenzyme Q10 reductase in the plasma membrane. J. Bioenerg. Biomembr. 1997, 29, 251–257. [Google Scholar] [CrossRef] [PubMed]
  35. Santos-Ocaña, C.; Villalba, J.M.; Córdoba, F.; Padilla, S.; Crane, F.L.; Clarke, C.F.; Navas, P.J. Genetic evidence for coenzyme Q requirement in plasma membrane electron transport. Bioenerg. Biomembr. 1998, 30, 465–475. [Google Scholar] [CrossRef]
  36. Bello, R.I.; Kagan, V.E.; Tyurin, V.; Navarro, F.; Alcaín, F.J.; Villalba, J.M. Regeneration of lipophil antioxidants by NAD(P)H:quinone oxidoreductase 1. Protoplasma 2003, 221, 129–135. [Google Scholar] [CrossRef] [PubMed]
  37. Kagan, V.; Serbinova, E.; Packer, L. Antioxidant effects of ubiquinones in microsomes and mitochondria are mediated by tocopherol recycling. Biochem. Biophys. Res. Commun. 1990, 169, 851–857. [Google Scholar] [CrossRef]
  38. Bindoli, A.; Valente, M.; Cavallini, L. Inhibition of lipid peroxidation by alpha tocopherolquinone and alpha tocopherolhydroquinone. Biochem. Int. 1985, 10, 753–761. [Google Scholar] [PubMed]
  39. Frei, B.; Kim, M.C.; Ames, B.N. Ubiquinol-10 is an effective lipid-soluble antioxidant at physiological concentrations. Proc. Natl. Acad. Sci. USA 1990, 87, 4879–4883. [Google Scholar] [CrossRef] [PubMed]
  40. Navarro, F.; Navas, P.; Burgess, J.R.; Bello, R.I.; de Cabo, R.; Arroyo, A.; Villalba, J.M. Vitamin E and selenium deficiency induces expression of the ubiquinone-dependent antioxidant system at the plasma membrane. FASEB J. 1998, 12, 1665–1673. [Google Scholar] [PubMed]
  41. Arroyo, A.; Navarro, F.; Navas, P.; Villalba, J.M. Ubiquinol regeneration by plasma membrane ubiquinone reductase. Protoplasma 1998, 205, 107–113. [Google Scholar] [CrossRef]
  42. Beyer, R.E.; Segura-Aguilar, J.; di Bernardo, S.; Cavazzoni, M.; Fato, R.; Fiorentini, D.; Galli, M.C.; Setti, M.; Landi, L.; Lenaz, G. The role of DT-diaphorase in the maintenance of the reduced antioxidant form of coenzyme Q in membrane systems. Proc. Natl. Acad. Sci. USA 1996, 93, 2528–2532. [Google Scholar] [CrossRef] [PubMed]
  43. Rushmore, T.H.; Morton, M.R.; Pickett, C.B.J. The antioxidant responsive element. Activation by oxidative stress and identification of the DNA consensus sequence required for functional activity. Biol. Chem. 1991, 266, 11632–11639. [Google Scholar]
  44. Takahashi, T.; Yamaguchi, T.; Shitashige, M.; Okamoto, T.; Kishi, T. Reduction of ubiquinone in membrane lipids by rat liver cytosol and its involvement in the cellular defence system against lipid peroxidation. Biochem. J. 1995, 309, 883–890. [Google Scholar] [CrossRef] [PubMed]
  45. Kishi, T.; Takahashi, T.; Usui, A.; Hashizume, N.; Okamoto, T. Cytosolic NADPH-UQ reductase, the enzyme responsible for cellular ubiquinone redox cycle as an endogenous antioxidant in the rat liver. Biofactors 1999, 9, 189–197. [Google Scholar] [CrossRef] [PubMed]
  46. Takahashi, T.; Hohda, T.; Sugimoto, N.; Mizobuchi, S.; Okamoto, T.; Mori, K.; Kishi, T. Antioxidant roles of cellular ubiquinone and related redox cycles: Potentiated resistance of rat hepatocytes having stimulated NADPH-dependent ubiquinone reductase against hydrogen peroxide toxicity. Biol. Pharm. Bull. 1999, 22, 1226–1233. [Google Scholar] [CrossRef] [PubMed]
  47. Takahashi, T.; Okuno, M.; Okamoto, T.; Kishi, T. NADPH-dependent coenzyme Q reductase is the main enzyme responsible for the reduction of non-mitochondrial CoQ in cells. Biofactors 2008, 32, 59–70. [Google Scholar] [CrossRef] [PubMed]
  48. Stocker, R.; Bowry, V.W.; Frei, B. Ubiquinol-10 protects human low density lipoprotein more efficient against lipid peroxidation than does alpha-tocopherol. Proc. Natl. Acad. Sci. USA 1991, 88, 1646–1650. [Google Scholar] [CrossRef] [PubMed]
  49. De Grey, A.D. A proposed mechanism for the lowering of mitochondrial electron leak by caloric restriction. Mitochondrion 2001, 1, 129–139. [Google Scholar] [CrossRef]
  50. Gómez-Díaz, C.; Villalba, J.M.; Pérez-Vicente, R.; Crane, F.L.; Navas, P. Ascorbate stabilization is stimulated in rho(0)HL-60 cells by CoQ10 increase at the plasma membrane. Biochem. Biophys. Res. Commun. 1997, 234, 79–81. [Google Scholar] [CrossRef]
  51. Hyun, D.-H.; Emerson, S.S.; Jo, D.-G.; Mattson, M.P.; de Cabo, R. Calorie restriction up-regulates the plasma membrane redox system in brain cells and suppresses oxidative stress during aging. Proc. Natl. Acad. Sci. USA 2006, 103, 19908–19912. [Google Scholar] [CrossRef] [PubMed]
  52. Hyun, D.-H.; Hunt, N.D.; Emerson, S.S.; Hernandez, J.O.; Mattson, M.P.; de Cabo, R. Up-regulation of plasma membrane-associated redox activities in neuronal cells lacking functional mitochondria. J. Neurochem. 2007, 100, 1364–1374. [Google Scholar] [CrossRef] [PubMed]
  53. Quesada, J.M.; Lopez-LLuch, G.; Buron, M.I.; Alcain, F.J.; Borrego, F.; Velde, J.P.; Blanco, I.; Bouillon, R.; Navas, P. Ascorbate increases the 1,25 dihydroxyvitamin D3-induced monocytic differentiation of HL-60 cells. Calcif. Tissue Int. 1996, 59, 277–282. [Google Scholar] [CrossRef] [PubMed]
  54. López-Lluch, G.; Burón, M.I.; Alcaín, F.J.; Quesada, J.M.; Navas, P. Redox regulation of cAMP levels by ascorbate in 1,25-dihydroxy-vitamin D3-induced differentiation of HL-60 cells. Biochem. J. 1998, 331, 21–27. [Google Scholar] [CrossRef] [PubMed]
  55. Sun, I.L.; Sun, E.E.; Crane, F.L. Stimulation of serum-free cell proliferation by coenzyme Q. Biochem. Biophys. Res. Commun. 1992, 189, 8–13. [Google Scholar] [CrossRef]
  56. Brightman, A.O.; Wang, J.; Miu, R.K.; Sun, I.L.; Barr, R.; Crane, F.L.; Morré, D.J. A growth factor- and hormone-stimulated NADH oxidase from rat liver plasma membrane. Biochim. Biophys. Acta 1992, 1105, 109–117. [Google Scholar] [CrossRef]
  57. López-Lluch, G.; Blázquez, M.V.; Pérez-Vicente, R.; Macho, A.; Burón, M.I.; Alcaín, F.J.; Muñoz, E.; Navas, P. Cellular redox state and activating protein-1 are involved in ascorbate effect on calcitriol-induced differentiation. Protoplasma 2001, 217, 129–136. [Google Scholar] [CrossRef] [PubMed]
  58. Villalba, J.M.; Navarro, F.; Córdoba, F.; Serrano, A.; Arroyo, A.; Crane, F.L.; Navas, P. Coenzyme Q reductase from liver plasma membrane: Purification and role in trans-plasma-membrane electron transport. Proc. Natl. Acad. Sci. USA 1995, 92, 4887–4891. [Google Scholar] [CrossRef] [PubMed]
  59. Linnane, A.W.; Kios, M.; Vitetta, L. Coenzyme Q(10)–its role as a prooxidant in the formation of superoxide anion/hydrogen peroxide and the regulation of the metabolome. Mitochondrion 2007, 7, S51–S61. [Google Scholar] [CrossRef] [PubMed]
  60. Crane, F.L.; Sun, I.L.; Crowe, R.A.; Alcain, F.J.; Löw, H. Coenzyme Q10, plasma membrane oxidase and growth control. Mol. Aspects Med. 1994, 15, S1–S11. [Google Scholar] [CrossRef]
  61. Baker, M.A.; Lane, D.J.R.; Ly, J.D.; De Pinto, V.; Lawen, A.J. VDAC1 is a transplasma membrane NADH-ferricyanide reductase. Biol. Chem. 2004, 279, 4811–4819. [Google Scholar] [CrossRef] [PubMed]
  62. Papucci, L.; Schiavone, N.; Witort, E.; Donnini, M.; Lapucci, A.; Tempestini, A.; Formigli, L.; Zecchi-Orlandini, S.; Orlandini, G.; Carella, G.; et al. Coenzyme Q10 prevents apoptosis by inhibiting mitochondrial depolarization independently of its free radical scavenging property. J. Biol. Chem. 2003, 278, 28220–28228. [Google Scholar] [CrossRef] [PubMed]
  63. Smith, J.S.; Brachmann, C.B.; Celic, I.; Kenna, M.A.; Muhammad, S.; Starai, V.J.; Avalos, J.L.; Escalante-Semerena, J.C.; Grubmeyer, C.; Wolberger, C.; et al. A phylogenetically conserved NAD+-dependent protein deacetylase activity in the Sir2 protein family. Proc. Natl. Acad. Sci. USA 2000, 97, 6658–6663. [Google Scholar] [CrossRef] [PubMed]
  64. Fernández-Ayala, D.J.; Martín, S.F.; Barroso, M.P.; Gómez-Díaz, C.; Villalba, J.M.; Rodríguez-Aguilera, J.C.; López-Lluch, G.; Navas, P. Coenzyme Q protects cells against serum withdrawal-induced apoptosis by inhibition of ceramide release and caspase-3 activation. Antioxid. Redox. Signal. 2000, 2, 263–275. [Google Scholar] [CrossRef] [PubMed]
  65. Martín, S.F.; Gómez-Díaz, C.; Bello, R.I.; Navas, P.; Villalba, J.M. Inhibition of neutral Mg2+-dependent sphingomyelinase by ubiquinol-mediated plasma membrane electron transport. Protoplasma 2003, 221, 109–116. [Google Scholar] [CrossRef] [PubMed]
  66. Villalba, J.M.; Navas, P. Plasma membrane redox system in the control of stress-induced apoptosis. Antioxid. Redox. Signal. 2000, 2, 213–230. [Google Scholar] [CrossRef] [PubMed]
  67. Chatterjee, S.; Han, H.; Rollins, S.; Cleveland, T.J. Molecular cloning, characterization, and expression of a novel human neutral sphingomyelinase. Biol. Chem. 1999, 274, 37407–37412. [Google Scholar] [CrossRef]
  68. Schmelzer, C.; Lindner, I.; Vock, C.; Fujii, K.; Döring, F. Functional connections and pathways of coenzyme Q10-inducible genes: An in-silico study. IUBMB Life 2007, 59, 628–633. [Google Scholar] [PubMed]
  69. Kaltschmidt, B.; Sparna, T.; Kaltschmidt, C. Activation of NF-kappa B by reactive oxygen intermediates in the nervous system. Antioxid. Redox. Signal. 1999, 1, 129–144. [Google Scholar] [CrossRef] [PubMed]
  70. Schmelzer, C.; Döring, F. Identification of LPS-inducible genes downregulated by ubiquinone in human THP-1 monocytes. Biofactors 2010, 36, 222–228. [Google Scholar] [CrossRef] [PubMed]
  71. Naderi, J.; Somayajulu-Nitu, M.; Mukerji, A.; Sharda, P.; Sikorska, M.; Borowy-Borowski, H.; Antonsson, B.; Pandey, S. Water-soluble formulation of Coenzyme Q10 inhibits Bax-induced destabilization of mitochondria in mammalian cells. Apoptosis 2006, 11, 1359–1369. [Google Scholar] [CrossRef] [PubMed]
  72. Yamamura, T.; Otani, H.; Nakao, Y.; Hattori, R.; Osako, M.; Imamura, H.; Das, D.K. Dual involvement of coenzyme Q10 in redox signaling and inhibition of death signaling in the rat heart mitochondria. Antioxid. Redox. Signal. 2001, 3, 103–112. [Google Scholar] [CrossRef] [PubMed]
  73. Bernardi, P.; Forte, M. The mitochondrial permeability transition pore. Novartis Found Symp. 2007, 287, 157–164; discussion 164–169. [Google Scholar] [PubMed]
  74. Devun, F.; Walter, L.; Belliere, J.; Cottet-Rousselle, C.; Leverve, X.; Fontaine, E. Ubiquinone analogs: A mitochondrial permeability transition poredependent pathway to selective cell death. PLoS ONE 2010, 5, e11792. [Google Scholar] [CrossRef] [PubMed]
  75. Fontaine, E.; Ichas, F.; Bernardi, P.J. A ubiquinone-binding site regulates the mitochondrial permeability transition pore. Biol. Chem. 1998, 273, 25734–25740. [Google Scholar] [CrossRef]
  76. Walter, L.; Nogueira, V.; Leverve, X.; Heitz, M.P.; Bernardi, P.; Fontaine, E. Three classes of ubiquinone analogs regulate the mitochondrial permeability transition pore through a common site. J. Biol. Chem. 2000, 275, 29521–29527. [Google Scholar] [CrossRef] [PubMed]
  77. Ferreira, F.M.; Seiça, R.; Oliveira, P.J.; Coxito, P.M.; Moreno, A.J.; Palmeira, C.M.; Santos, M.S. Diabetes induces metabolic adaptations in rat liver mitochondria: Role of coenzyme Q and cardiolipin contents. Biochim. Biophys. Acta 2003, 1639, 113–120. [Google Scholar] [CrossRef] [PubMed]
  78. Li, G.; Zou, L.-Y.; Cao, C.-M.; Yang, E.S. Coenzyme Q10 protects SHSY5Y neuronal cells from beta amyloid toxicity and oxygen-glucose deprivation by inhibiting the opening of the mitochondrial permeability transition pore. Biofactors 2005, 25, 97–107. [Google Scholar] [CrossRef] [PubMed]
  79. Cordero, M.D.; Moreno-Fernández, A.M.; Gomez-Skarmeta, J.L.; de Miguel, M.; Garrido-Maraver, J.; Oropesa-Avila, M.; Rodríguez-Hernández, A.; Navas, P.; Sánchez-Alcázar, J.A. Coenzyme Q10 and alpha-tocopherol protect against amitriptyline toxicity. Toxicol. Appl. Pharmacol. 2009, 235, 329–337. [Google Scholar] [CrossRef] [PubMed]
  80. Gille, L.; Nohl, H. The existence of a lysosomal redox chain and the role of ubiquinone. Biochem. Biophys. 2000, 375, 347–354. [Google Scholar] [CrossRef] [PubMed]
  81. Crane, F.L. The evolution of coenzyme Q. Biofactors 2008, 32, 5–11. [Google Scholar] [CrossRef] [PubMed]
  82. Grünler, J.; Ericsson, J.; Dallner, G. Branch-point reactions in the biosynthesis of cholesterol, dolichol, ubiquinone and prenylated proteins. Biochim. Biophys. Acta 1994, 1212, 259–277. [Google Scholar] [CrossRef]
  83. Villalba, J.M.; Parrado, C.; Santos-Gonzalez, M.; Alcain, F.J. Therapeutic use of coenzyme Q10 and coenzyme Q10-related compounds and formulations. Expert Opin. Investig. Drugs 2010, 19, 535–554. [Google Scholar] [CrossRef] [PubMed]
  84. Tran, U.C.; Clarke, C.F. Endogenous synthesis of coenzyme Q in eukaryotes. Mitochondrion 2007, 7, S62–S71. [Google Scholar] [CrossRef] [PubMed]
  85. Forsgren, M.; Attersand, A.; Lake, S.; Grunler, J.; Swiezewska, E.; Dallner, G.; Climent, I. Isolation and functional expression of human COQ2, a gene encoding a polyprenyl transferase involved in the synthesis of CoQ. Biochem. J. 2004, 382, 519–526. [Google Scholar] [CrossRef] [PubMed]
  86. Quinzii, C.M.; DiMauro, S.; Hirano, M. Human coenzyme Q10 deficiency. Neurochem. Res. 2007, 32, 723–727. [Google Scholar] [CrossRef] [PubMed]
  87. Fernández-Ayala, D.J.M.; Brea-Calvo, G.; López-Lluch, G.; Navas, P. Coenzyme Q distribution in HL-60 human cells depends on the endomembrane system. Biochim. Biophys. Acta 2005, 1713, 129–137. [Google Scholar] [CrossRef] [PubMed]
  88. Aberg, F.; Appelkvist, E.L.; Dallner, G.; Ernster, L. Distribution and redox state of ubiquinones in rat and human tissues. Biochem. Biophys. 1992, 295, 230–234. [Google Scholar] [CrossRef]
  89. Runquist, M.; Parmryd, I.; Thelin, A.; Chojnacki, T.; Dallner, G.J. Distribution of branch point prenyltransferases in regions of bovine brain. Neurochem 1995, 65, 2299–2306. [Google Scholar] [CrossRef]
  90. Molyneux, S.L.; Florkowski, C.M.; Lever, M.; George, P.M. Biological variation of coenzyme Q10. Clin. Chem. 2005, 51, 455–457. [Google Scholar] [CrossRef] [PubMed]
  91. Elmberger, P.G.; Kal`en, A.; Appelkvist, E.L.; Dallner, G. In vitro and in vivo synthesis of dolichol and other main mevalonate products in various organs of the rat. Eur. J. Biochem. 1987, 168, 1–11. [Google Scholar] [CrossRef] [PubMed]
  92. Andersson, M.; Elmberger, P.G.; Edlund, C.; Kristensson, K.; Dallner, G. Rates of cholesterol, ubiquinone, dolichol and dolichyl-P biosynthesis in rat brain slices. FEBS Lett. 1990, 269, 15–18. [Google Scholar] [CrossRef]
  93. Goldstein, J.L.; Brown, M.S. Regulation of the mevalonate pathway. Nature 1990, 343, 425–430. [Google Scholar] [CrossRef] [PubMed]
  94. Brown, M.S.; Goldstein, J.L. Multivalent feedback regulation of HMG CoA reductase, a control mechanism coordinating isoprenoid synthesis and cell growth. Lipid Res. 1980, 21, 505–517. [Google Scholar]
  95. Bentinger, M.; Tekle, M.; Brismar, K.; Chojnacki, T.; Swiezewska, E.; Dallner, G.J. Polyisoprenoid epoxides stimulate the biosynthesis of coenzyme Q and inhibit cholesterol synthesis. Biol. Chem. 2008, 283, 14645–14653. [Google Scholar] [CrossRef] [PubMed]
  96. Aberg, F.; Zhang, Y.; Teclebrhan, H.; Appelkvist, E.L.; Dallner, G. Increases in tissue levels of ubiquinone in association with peroxisome proliferation. Chem. Biol. Interact. 1996, 99, 205–218. [Google Scholar] [CrossRef]
  97. Brea-Calvo, G.; Rodríguez-Hernández, A.; Fernández-Ayala, D.J.M.; Navas, P.; Sánchez-Alcázar, J.A. Chemotherapy induces an increase in coenzyme Q10 levels in cancer cell lines. Free Radic. Biol. Med. 2006, 40, 1293–1302. [Google Scholar] [PubMed]
  98. Brea-Calvo, G.; Siendones, E.; Sánchez-Alcázar, J.A.; de Cabo, R.; Navas, P. Cell survival from chemotherapy depends on NF-kappaB transcriptional upregulation of coenzyme Q biosynthesis. PLoS ONE 2009, 4, e5301. [Google Scholar] [CrossRef] [PubMed]
  99. Reahal, S.; Wrigglesworth, J. Tissue concentrations of coenzyme Q10 in the rat following its oral and intraperitoneal administration. Drug Metab. Dispos. 1992, 20, 423–427. [Google Scholar] [PubMed]
  100. Zhang, Y.; Aberg, F.; Appelkvist, E.L.; Dallner, G.; Ernster, L. Uptake of dietary coenzyme Q supplement is limited in rats. J. Nutr. 1995, 125, 446–453. [Google Scholar] [PubMed]
  101. Zhang, Y.; Turunen, M.; Appelkvist, E.L. Restricted uptake of dietary coenzyme Q is in contrast to the unrestricted uptake of alpha-tocopherol into rat organs and cells. J. Nutr. 1996, 126, 2089–2097. [Google Scholar] [PubMed]
  102. Turunen, M.; Appelkvist, E.L.; Sindelar, P.; Dallner, G. Blood concentration of coenzyme Q(10) increases in rats when esterified forms are administered. J. Nutr. 1999, 129, 2113–2118. [Google Scholar] [PubMed]
  103. Bentinger, M.; Dallner, G.; Chojnacki, T.; Swiezewska, E. Distribution and breakdown of labeled coenzyme Q10 in rat. Free Radic. Biol. Med. 2003, 34, 563–575. [Google Scholar] [CrossRef]
  104. Bhagavan, H.N.; Chopra, R.K.; Craft, N.E.; Chitchumroonchokchai, C.; Failla, M.L. Assessment of coenzyme Q10 absorption using an in vitro digestion-Caco-2 cell model. Int. J. Pharm. 2007, 333, 112–117. [Google Scholar] [CrossRef] [PubMed]
  105. Svensson, M.; Malm, C.; Tonkonogi, M.; Ekblom, B.; Sjödin, B.; Sahlin, K. Effect of Q10 supplementation on tissue Q10 levels and adenine nucleotide catabolism during high-intensity exercise. Int. J. Sport Nutr. 1999, 9, 166–180. [Google Scholar] [PubMed]
  106. Watts, G.F.; Playford, D.A.; Croft, K.D.; Ward, N.C.; Mori, T.A.; Burke, V. Coenzyme Q(10) improves endothelial dysfunction of the brachial artery in Type II diabetes mellitus. Diabetologia 2002, 45, 420–426. [Google Scholar] [CrossRef] [PubMed]
  107. Wolters, M.; Hahn, A. Plasma ubiquinone status and response to six-month supplementation combined with multivitamins in healthy elderly women–results of a randomized, double-blind, placebo-controlled study. Int. J. Vitam. Nutr. Res. 2003, 73, 207–214. [Google Scholar] [CrossRef] [PubMed]
  108. Hosoe, K.; Kitano, M.; Kishida, H.; Kubo, H.; Fujii, K.; Kitahara, M. Study on safety and bioavailability of ubiquinol (Kaneka QH) after single and 4-week multiple oral administration to healthy volunteers. Regul. Toxicol. Pharmacol. 2007, 47, 19–28. [Google Scholar] [CrossRef] [PubMed]
  109. Ikematsu, H.; Nakamura, K.; Harashima, S.-I.; Fujii, K.; Fukutomi, N. Safety assessment of coenzyme Q10 (Kaneka Q10) in healthy subjects: A double-blind, randomized, placebo-controlled trial. Regul. Toxicol. Pharmacol. 2006, 44, 212–218. [Google Scholar] [CrossRef] [PubMed]
  110. Matthews, R.T.; Yang, L.; Browne, S.; Baik, M.; Beal, M.F. Coenzyme Q10 administration increases brain mitochondrial concentrations and exerts neuroprotective effects. Proc. Natl. Acad. Sci. USA 1998, 95, 8892–8897. [Google Scholar] [CrossRef] [PubMed]
  111. Kwong, L.K.; Kamzalov, S.; Rebrin, I.; Bayne, A.-C.V.; Jana, C.K.; Morris, P.; Forster, M.J.; Sohal, R.S. Effects of coenzyme Q(10) administration on its tissue concentrations, mitochondrial oxidant generation, and oxidative stress in the rat. Free Radic. Biol. Med. 2002, 33, 627–638. [Google Scholar] [CrossRef]
  112. The Parkinson Study Group QE3 Investigators. A randomized clinical trial of high-dosage coenzyme Q10 in early Parkinson disease. JAMA Neurol. 2014, 71, 543–552. [Google Scholar]
  113. Kaufmann, P.; Thompson, J.L.P.; Levy, G.; Buchsbaum, S.J.; Krivickas, L.S.; Katz, J.; Rollins, Y.; Barohn, R.J.; Jackson, C.E.; Tiryaki, E.; et al. Phase II trial of CoQ10 for ALS finds insufficient evidence to justify phase III. Ann. Neurol. 2009, 66, 235–244. [Google Scholar] [CrossRef] [PubMed]
  114. Lee, B.J.; Huang, Y.C.; Chen, S.J.; Lin, P.T. Coenzyme Q10 supplementation reduces oxidative stress and increases antioxidant enzyme activity in patients with coronary artery disease. Nutrition 2012, 28, 250–255. [Google Scholar] [CrossRef] [PubMed]
  115. Bargossi, A.M.; Grossi, G.; Fiorella, P.L.; Gaddi, A.; di Giulio, R.; Battino, M. Exogenous CoQ10 supplementation prevents plasma ubiquinone reduction induced by HMG-CoA reductase inhibitors. Mol. Aspects Med. 1994, 15, s187–s193. [Google Scholar] [CrossRef]
  116. Bargossi, A.M.; Battino, M.; Gaddi, A.; Fiorella, P.L.; Grossi, G.; Barozzi, G.; Di Giulio, R.; Descovich, G.; Sassi, S.; Genova, M.L. Exogenous CoQ10 preserves plasma ubiquinone levels in patients treated with 3-hydroxy-3-methylglutaryl coenzyme A reductase inhibitors. Int. J. Clin. Lab. Res. 1994, 24, 171–176. [Google Scholar] [CrossRef] [PubMed]
  117. Lee, C.-K.; Pugh, T.D.; Klopp, R.G.; Edwards, J.; Allison, D.B.; Weindruch, R.; Prolla, T.A. The impact of alpha-lipoic acid, coenzyme Q10 and caloric restriction on life span and gene expression patterns in mice. Free Radic. Biol. Med. 2004, 36, 1043–1057. [Google Scholar] [CrossRef] [PubMed]
  118. Ochoa, J.J.; Quiles, J.L.; Huertas, J.R.; Mataix, J. Coenzyme Q10 protects from aging-related oxidative stress and improves mitochondrial function in heart of rats fed a polyunsaturated fatty acid (PUFA)-rich diet. J. Gerontol. A Biol. Sci. Med. Sci. 2005, 60, 970–975. [Google Scholar] [CrossRef] [PubMed]
  119. Sohal, R.S.; Kamzalov, S.; Sumien, N.; Ferguson, M.; Rebrin, I.; Heinrich, K.R.; Forster, M.J. Effect of coenzyme Q10 intake on endogenous coenzyme Q content, mitochondrial electron transport chain, antioxidative defenses, and life span of mice. Free Radic. Biol. Med. 2006, 40, 480–487. [Google Scholar] [CrossRef] [PubMed]
  120. Lass, A.; Kwong, L.; Sohal, R.S. Mitochondrial coenzyme Q content and aging. Biofactors 1999, 9, 199–205. [Google Scholar] [CrossRef] [PubMed]
  121. Kamzalov, S.; Sumien, N.; Forster, M.J.; Sohal, R.S. Coenzyme Q intake elevates the mitochondrial and tissue levels of Coenzyme Q and alpha-tocopherol in young mice. J. Nutr. 2003, 133, 3175–3180. [Google Scholar] [PubMed]
  122. Huertas, J.R.; Martinez-Velasco, E.; Ibánñez, S.; López-Frias, M.; Ochoa, J.J.; Quiles, J.; Parenti Castelli, G.; Mataix, J.; Lenaz, G. Virgin olive oil and coenzyme Q10 protect heart mitochondria from peroxidative damage during aging. Biofactors 1999, 9, 337–343. [Google Scholar] [CrossRef] [PubMed]
  123. Quiles, J.L.; Ochoa, J.J.; Huertas, J.R.; Mataix, J. Coenzyme Q supplementation protects from age-related DNA double-strand breaks and increases lifespan in rats fed on a PUFA-rich diet. Exp. Gerontol. 2004, 39, 189–194. [Google Scholar] [CrossRef] [PubMed]
  124. Ochiai, A.; Itagaki, S.; Kurokawa, T.; Kobayashi, M.; Hirano, T.; Iseki, K. Improvement in intestinal coenzyme q10 absorption by food intake. Yakugaku Zasshi 2007, 127, 1251–1254. [Google Scholar] [CrossRef] [PubMed]
  125. Mataix, J.; Mañas, M.; Quiles, J.; Battino, M.; Cassinello, M.; Lopez-Frias, M.; Huertas, J.R. Coenzyme Q content depends upon oxidative stress and dietary fat unsaturation. Mol. Aspect Med. 1997, 18, 129–135. [Google Scholar] [CrossRef]
  126. Ramirez-Tortosa, M.C.; Granados, S.; Ramirez-Tortosa, C.L.; Ochoa, J.J.; Camacho, P.; García-Valdós, L.; Battino, M.; Quiles, J.L. Oxidative stress status in liver mitochondria and lymphocyte DNA damage of atherosclerotic rabbits supplemented with water soluble coenzyme Q10. Biofactors 2008, 32, 263–273. [Google Scholar] [CrossRef] [PubMed]
  127. Hidaka, T.; Fujii, K.; Funahashi, I.; Fukotomi, M.; Hosoe, K. Safety assesment of coenzyme Q10 (CoQ). Biofactors 2008, 32, 199–208. [Google Scholar] [CrossRef] [PubMed]
  128. Sumien, N.; Heinrich, K.R.; Shetty, R.A.; Sohal, R.S.; Foster, M.J. Prolonged intake of coenzyme Q10 impairs cognitive functions in mice. J. Nutr. 2009, 139, 1926–1932. [Google Scholar] [CrossRef] [PubMed]
  129. Alho, H.; Lonnrot, K. Coenzyme Q supplementation and longevity. In Coenzyme Q: Molecular Mechanisms in Health and Disease, 1st ed.; Kagan, V.E., Quinn, P.J., Eds.; CRC Press: Boca Raton, FL, USA, 2000; pp. 371–380. [Google Scholar]
  130. Onur, S.; Niklowitz, P.; Fischer, A.; Metges, C.C.; Grune, T.; Menke, T.; Rimbach, G.; Döring, F. A comparative study into alterations of coenzyme Q redox status in ageing pigs, mice, and worms. Biofactors 2014, 40, 346–354. [Google Scholar] [CrossRef] [PubMed]
  131. Kamzalov, S.; Sohal, R.S. Effect of age and caloric restriction on coenzyme Q and alpha-tocopherol levels in the rat. Exp. Gerontol. 2004, 39, 1199–1205. [Google Scholar] [CrossRef] [PubMed]
  132. Zhang, Y.; Appelkvist, E.L.; Kristensson, K.; Dallner, G. The lipid compositions of different regions of rat brain during development and aging. Neurobiol. Aging 1996, 17, 869–875. [Google Scholar] [CrossRef]
  133. Battino, M.; Gorini, A.; Villa, R.F.; Genova, M.L.; Bovina, C.; Sassi, S.; Littarru, G.P.; Lenaz, G. Coenzyme Q content in synaptic and non-synaptic mitochondria from different brain regions in the ageing rat. Mech. Ageing Dev. 1995, 78, 173–187. [Google Scholar] [CrossRef]
  134. Kalén, A.; Appelkvist, E.L.; Dallner, G. Age-related changes in the lipid compositions of rat and human tissues. Lipids 1989, 24, 579–584. [Google Scholar] [CrossRef] [PubMed]
  135. Beyer, R.E.; Burnett, B.A.; Cartwright, K.J.; Edington, D.W.; Falzon, M.J.; Kreitman, K.R.; Kuhn, T.W.; Ramp, B.J.; Rhee, S.Y.; Rosenwasser, M.J. Tissue coenzyme Q (ubiquinone) and protein concentrations over the life span of the laboratory rat. Mech. Ageing Dev. 1985, 32, 267–281. [Google Scholar] [CrossRef]
  136. Söderberg, M.; Edlund, C.; Kristensson, K.; Dallner, G. Lipid compositions of different regions of the human brain during aging. J. Neurochem. 1990, 54, 415–423. [Google Scholar] [CrossRef] [PubMed]
  137. Edlund, C.; Söderberg, M.; Kristensson, K.; Dallner, G. Ubiquinone, dolichol, and cholesterol metabolism in aging and Alzheimer’s disease. Biochem. Cell Biol. 1992, 70, 422–428. [Google Scholar] [CrossRef] [PubMed]
  138. Kagan, V.E.; Quinn, P.J. Coenzyme Q: Molecular Mechanisms in Health and Disease, 1st ed.; CRC Press: Boca Ratón, FL, USA, 2000; p. 408. [Google Scholar]
  139. Witting, P.K.; Pettersson, K.; Letters, J.; Stocker, R. Anti-atherogenic effect of coenzyme Q10 in apolipoprotein E gene knockout mice. Free Radic. Biol. Med. 2000, 29, 295–305. [Google Scholar] [CrossRef]
  140. Ayaz, M.; Tuncer, S.; Okudan, N.; Gokbel, H. Coenzyme Q(10) and alphalipoic acid supplementation in diabetic rats: Conduction velocity distributions. Methods Find Exp. Clin. Pharmacol. 2008, 30, 367–374. [Google Scholar] [CrossRef] [PubMed]
  141. Chew, G.T.; Watts, G.F.; Davis, T.M.E.; Stuckey, B.G.A.; Beilin, L.J.; Thompson, P.L.; Burke, V.; Currie, P. Hemodynamic effects of fenofibrate and coenzyme Q10 in type 2 diabetic subjects with left ventricular diastolic dysfunction. J. Diabetes Care 2008, 31, 1502–1509. [Google Scholar] [CrossRef] [PubMed]
  142. Sena, C.M.; Nunes, E.; Gomes, A.; Santos, M.S.; Proença, T.; Martins, M.I.; Seiça, R.M. Supplementation of coenzyme Q10 and alpha-tocopherol lowers glycated hemoglobin level and lipid peroxidation in pancreas of diabetic rats. Nutr. Res. 2008, 28, 113–121. [Google Scholar] [CrossRef] [PubMed]
  143. Molina, J.A.; de Bustos, F.; Ortiz, S.; Del Ser, T.; Seijo, M.; Benito-Léon, J.; Oliva, J.M.; Pérez, S.; Manzanares, J. Serum levels of coenzyme Q in patients with Lewy body disease. J. Neural Transm. 2002, 109, 1195–1201. [Google Scholar] [CrossRef] [PubMed]
  144. Sohmiya, M.; Tanaka, M.; Suzuki, Y.; Tanino, Y.; Okamoto, K.; Yamamoto, Y. An increase of oxidized coenzyme Q-10 occurs in the plasma of sporadic ALS patients. J. Neurol. Sci. 2005, 228, 49–53. [Google Scholar] [CrossRef] [PubMed]
  145. Isobe, C.; Abe, T.; Terayama, Y. Levels of reduced and oxidized coenzyme Q-10 and 8-hydroxy-2’-deoxyguanosine in the CSF of patients with Alzheimer’s disease demonstrate that mitochondrial oxidative damage and/or oxidative DNA damage contributes to the neurodegenerative process. J. Neurol. 2010, 257, 399–404. [Google Scholar] [CrossRef] [PubMed]
  146. Isobe, C.; Abe, T.; Terayama, Y. Increase in the oxidized/total coenzyme Q-10 ratio in the cerebrospinal fluid of Alzheimer’s disease patients. Dement Geriatr. Cogn. Disord. 2009, 28, 449–454. [Google Scholar] [CrossRef] [PubMed]
  147. Coyle, J.T.; Puttfarcken, P. Oxidative stress, glutamate, and neurodegenerative disorders. Science 1993, 262, 689–695. [Google Scholar] [CrossRef] [PubMed]
  148. Jenner, P.; Schapira, A.H.; Marsden, C.D. New insights into the cause of Parkinson’s disease. Neurology 1992, 42, 2241–2250. [Google Scholar] [CrossRef] [PubMed]
  149. Olanow, C.W. A radical hypothesis for neurodegeneration. Trends Neurosci. 1993, 16, 439–444. [Google Scholar] [CrossRef]
  150. Schapira, A.H.; Hartley, A.; Cleeter, M.W.; Cooper, J.M. Free radicals and mitochondrial dysfunction in Parkinson’s disease. Biochem. Soc. Trans. 1993, 21, 367–370. [Google Scholar] [CrossRef] [PubMed]
  151. Isobe, C.; Murata, T.; Sato, C.; Terayama, Y. Increase of oxidized/total coenzyme Q-10 ratio in cerebrospinal fluid in patients with Parkinson’s disease. J. Clin. Neurosci. 2007, 14, 340–343. [Google Scholar] [CrossRef] [PubMed]
  152. Lim, S.C.; Tan, H.H.; Goh, S.K.; Subramaniam, T.; Sum, C.F.; Tan, I.K.; Lee, B.L.; Ong, C.N. Oxidative burden in prediabetic and diabetic individuals: Evidence from plasma coenzyme Q(10). Diabetes Med. 2006, 23, 1344–1349. [Google Scholar] [CrossRef] [PubMed]
  153. Folkers, K.; Osterborg, A.; Nylander, M.; Morita, M.; Mellstedt, H. Activities of vitamin Q10 in animal models and a serious deficiency in patients with cancer. Biochem. Biophys. Res. Commun. 1997, 234, 296–299. [Google Scholar] [CrossRef] [PubMed]
  154. Jolliet, P.; Simon, N.; Barré, J.; Pons, J.Y.; Boukef, M.; Paniel, B.J.; Tillement, J.P. Plasma coenzyme Q10 concentrations in breast cancer: Prognosis and therapeutic consequences. Int. J. Clin. Pharmacol. Ther. 1998, 36, 506–509. [Google Scholar] [PubMed]
  155. Rusciani, L.; Proietti, I.; Paradisi, A.; Rusciani, A.; Guerriero, G.; Mammone, A.; de Gaetano, A.; Lippa, S. Recombinant interferon alpha-2b and coenzyme Q10 as a postsurgical adjuvant therapy for melanoma: A 3-year trial with recombinant interferon-alpha and 5-year follow-up. Melanoma Res. 2007, 17, 177–183. [Google Scholar] [CrossRef] [PubMed]
  156. Mano, T.; Iwase, K.; Hayashi, R.; Hayakawa, N.; Uchimura, K.; Makino, M.; Nagata, M.; Sawai, Y.; Oda, N.; Hamada, M.; et al. Vitamin E and coenzyme Q concentrations in the thyroid tissues of patients with various thyroid disorders. Am. J. Med. Sci. 1998, 315, 230–232. [Google Scholar] [PubMed]
  157. Portakal, O.; Ozkaya, O.; Erden Inal, M.; Bozan, B.; Koşan, M.; Sayek, I. Coenzyme Q10 concentrations and antioxidant status in tissues of breast cancer patients. Clin. Biochem. 2000, 33, 279–284. [Google Scholar] [CrossRef]
  158. Arroyo, A.; Santos-Ocaña, C.; Ruiz-Ferrer, M.; Padilla, S.; Gavilán, A.; Rodríguez-Aguilera, J.C.; Navas, P. Coenzyme Q is irreplaceable by demethoxy-coenzyme Q in plasma membrane of Caenorhabditis elegans. FEBS Lett. 2006, 580, 1740–1746. [Google Scholar] [CrossRef] [PubMed]
  159. Quiles, J.L.; Farquharson, A.J.; Ramírez-Tortosa, M.C.; Grant, I.; Milne, L.; Huertas, J.R.; Battino, M.; Mataix, J.; Wahle, K.W.J. Coenzyme Q differentially modulates phospholipid hydroperoxide glutathione peroxidase gene expression and free radicals production in malignant and non-malignant prostate cells. Biofactors 2003, 18, 265–270. [Google Scholar] [CrossRef] [PubMed]
  160. Jonassen, T.; Davis, D.E.; Larsen, P.L.; Clarke, C.F. Reproductive fitness and quinone content of Caenorhabditis elegans clk-1 mutants fed coenzyme Q isoforms of varying length. J. Biol. Chem. 2003, 278, 51735–51742. [Google Scholar] [CrossRef] [PubMed]
  161. Wong, A.; Boutis, P.; Hekimi, S. Mutations in the clk-1 gene of Caenorhabditis elegans affect developmental and behavioral timing. Genetics 1995, 139, 1247–1259. [Google Scholar] [PubMed]
  162. Asencio, C.; Navas, P.; Cabello, J.; Schnabel, R.; Cypser, J.R.; Johnson, T.E.; Rodríguez-Aguilera, J.C. Coenzyme Q supports distinct developmental processes in Caenorhabditis elegans. Mech. Ageing Dev. 2009, 130, 145–153. [Google Scholar] [CrossRef] [PubMed]
  163. Gavilán, A.; Asencio, C.; Cabello, J.; Rodríguez-Aguilera, J.C.; Schnabel, R.; Navas, P. C. elegans knockouts in ubiquinone biosynthesis genes result in different phenotypes during larval development. Biofactors 2005, 25, 21–29. [Google Scholar] [CrossRef] [PubMed]
  164. Hihi, A.K.; Gao, Y.; Hekimi, S.J. Ubiquinone is necessary for Caenorhabditis elegans development at mitochondrial and non-mitochondrial sites. Biol. Chem. 2002, 277, 2202–2206. [Google Scholar] [CrossRef] [PubMed]
  165. Asencio, C.; Rodríguez-Aguilera, J.C.; Ruiz-Ferrer, M.; Vela, J.; Navas, P. C. elegans knockouts in ubiquinone biosynthesis genes result in different phenotypes during larval development. FASEB J. 2003, 17, 1135–1137. [Google Scholar] [PubMed]
  166. Vajo, Z.; King, L.M.; Jonassen, T.; Wilkin, D.J.; Ho, N.; Munnich, A.; Clarke, C.F.; Francomano, C.A. Conservation of the Caenorhabditis elegans timing gene clk-1 from yeast to human: A gene required for ubiquinone biosynthesis with potential implications for aging. Mamm. Genome 1999, 10, 1000–1004. [Google Scholar] [CrossRef] [PubMed]
  167. Levavasseur, F.; Miyadera, H.; Sirois, J.; Tremblay, M.L.; Kita, K.; Shoubridge, E.; Hekimi, S. Ubiquinone is necessary for mouse embryonic development but is not essential for mitochondrial respiration. J. Biol. Chem. 2001, 276, 46160–46164. [Google Scholar] [CrossRef] [PubMed]
  168. Nakai, D.; Yuasa, S.; Takahashi, M.; Shimizu, T.; Asaumi, S.; Isono, K.; Takao, T.; Suzuki, Y.; Kuroyanagi, H.; Hirokawa, K.; et al. Mouse homologue of coq7/clk-1, longevity gene in Caenorhabditis elegans, is essential for coenzyme Q synthesis, maintenance of mitochondrial integrity, and neurogenesis. Biochem. Biophys. Res. Commun. 2001, 289, 463–471. [Google Scholar] [CrossRef] [PubMed]
  169. Boitier, E.; Degoul, F.; Desguerre, I.; Charpentier, C.; François, D.; Ponsot, G.; Diry, M.; Rustin, P.; Marsac, C. A case of mitochondrial encephalomyopathy associated with a muscle coenzyme Q10 deficiency. J. Neurol. Sci. 1998, 156, 41–46. [Google Scholar] [CrossRef]
  170. Musumeci, O.; Naini, A.; Slonim, A.E.; Skavin, N.; Hadjigeorgiou, G.L.; Krawiecki, N.; Weissman, B.M.; Tsao, C.Y.; Mendell, J.R.; Shanske, S.; et al. Familial cerebellar ataxia with muscle coenzyme Q10 deficiency. Neurology 2001, 56, 849–855. [Google Scholar] [CrossRef] [PubMed]
  171. Liu, X.; Jiang, N.; Hughes, B.; Bigras, E.; Shoubridge, E.; Hekimi, S. Evolutionary conservation of the clk-1-dependent mechanism of longevity: Loss of mclk1 increases cellular fitness and lifespan in mice. Genes Dev. 2005, 19, 2424–2434. [Google Scholar] [CrossRef] [PubMed]
  172. López, L.C.; Schuelke, M.; Quinzii, C.M.; Kanki, T.; Rodenburg, R.J.; Naini, A.; Di Mauro, S.; Hirano, M. Leigh syndrome with nephropathy and CoQ 10 deficiency due to decaprenyl diphosphate synthase subunit 2 (PDSS2) mutations. Am. J. Hum. Genet. 2006, 79, 1125–1129. [Google Scholar] [CrossRef] [PubMed]
  173. Quinzii, C.; Naini, A.; Salviati, L.; Trevisson, E.; Navas, P.; DiMauro, S.; Hirano, M. A mutation in para-hydroxybenzoate-polyprenyl transferase (COQ2) causes primary coenzyme Q 10 deficiency. Am. J. Hum. Genet. 2006, 78, 345–349. [Google Scholar] [CrossRef] [PubMed]
  174. Mollet, J.; Giurgea, I.; Schlemmer, D.; Dallner, G.; Chretien, D.; Delahodde, A.; Bacq, D.; de Lonlay, P.; Munnich, A.; Rötig, A. Prenyldiphosphate synthase, subunit 1 (PDSS1) and OH-benzoate polyprenyltransferase (COQ2) mutations in ubiquinone deficiency and oxidative phosphorylation disorders. J. Clin. Investig. 2007, 117, 765. [Google Scholar] [CrossRef] [PubMed]
  175. Lagier-Tourenne, C.; Tazir, M.; López, L.C.; Quinzii, C.M.; Assoum, M.; Drouot, N.; Busso, C.; Makri, S.; Ali-Pacha, L.; Benhassine, T.; et al. ADCK3, an ancestral kinase, is mutated in a form of recessive ataxia associated with coenzyme Q 10 deficiency. Am. J. Hum. Genet. 2008, 82, 661–672. [Google Scholar] [CrossRef] [PubMed]
  176. Duncan, A.J.; Bitner-Glindzicz, M.; Meunier, B.; Costello, H.; Hargreaves, I.P.; López, L.C.; Hirano, M.; Quinzii, C.M.; Sadowski, M.I.; Hardy, J.; Singleton, A.; Clayton, P.T.; Rahman, S. A nonsense mutation in COQ9 causes autosomal-recessive neonatal-onset primary coenzyme Q 10 deficiency: A potentially treatable form of mitochondrial disease. Am. J. Hum. Genet. 2009, 84, 558–566. [Google Scholar] [CrossRef] [PubMed]
  177. Di Mauro, S.; Quinzii, C.M.; Hirano, M. Mutations in coenzyme Q10 biosynthetic genes. J. Clin. Investig. 2007, 117, 587–589. [Google Scholar] [CrossRef] [PubMed]
  178. Ogasahara, S.; Engel, A.G.; Frens, D.; Mack, D. Muscle coenzyme Q deficiency in familial mitochondrial encephalomyopathy. Proc. Natl. Acad. Sci. USA 1989, 86, 2379–2382. [Google Scholar] [CrossRef] [PubMed]
  179. Rötig, A.; Appelkvist, E.L.; Geromel, V.; Chretien, D.; Kadhom, N.; Edery, P.; Lebideau, M.; Dallner, G.; Munnich, A.; Ernster, L.; et al. Quinone-responsive multiple respiratory-chain dysfunction due to widespread coenzyme Q10 deficiency. Lancet 2000, 356, 391–395. [Google Scholar] [CrossRef]
  180. Artuch, R.; Brea-Calvo, G.; Briones, P.; Aracil, A.; Galván, M.; Espinós, C.; Corral, J.; Volpini, V.; Ribes, A.; Andreu, A.L.; et al. Cerebellar ataxia with coenzyme Q10 deficiency: Diagnosis and follow-up after coenzyme Q10 supplementation. J. Neurol. Sci. 2006, 246, 153–158. [Google Scholar] [CrossRef] [PubMed]
  181. Quinzii, C.M.; López, L.C.; Naini, A.; di Mauro, S.; Hirano, M. Human CoQ10 deficiencies. Biofactors 2008, 32, 113–118. [Google Scholar] [CrossRef] [PubMed]
  182. Molyneux, S.L.; Florkowski, C.M.; Richards, A.M.; Lever, M.; Young, J.M.; George, P.M. Coenzyme Q10; an adjunctive therapy for congestive heart failure? N. Z. Med. J. 2009, 122, 74–79. [Google Scholar] [PubMed]
  183. Rosenfeldt, F.; Marasco, S.; Lyon, W.; Wowk, M.; Sheeran, F.; Bailey, M.; Esmore, D.; Davis, B.; Pick, A.; Rabinov, M.; et al. Coenzyme Q10 therapy before cardiac surgery improves mitochondrial function and in vitro contractility of myocardial tissue. J. Thorac. Cardiovasc. Surg. 2005, 129, 25–32. [Google Scholar] [CrossRef] [PubMed]
  184. Beal, M.F. Mitochondrial dysfunction and oxidative damage in Alzheimer’s and Parkinson’s diseases and coenzyme Q10 as a potential treatment. J. Bioenerg. Biomembr. 2004, 36, 381–386. [Google Scholar] [CrossRef] [PubMed]
  185. Shults, C.W.; Oakes, D.; Kieburtz, K.; Beal, M.F.; Haas, R.; Plumb, S.; Juncos, J.L.; Nutt, J.; Shoulson, I.; Carter, J.; et al. Effects of coenzyme Q10 in early Parkinson disease: Evidence of slowing of the functional decline. Arch. Neuro 2002, 59, 154–1550. [Google Scholar] [CrossRef]
  186. Thomas, B.; Beal, M.F. Mitochondrial therapies for Parkinson’s disease. Mov. Disord. 2010, 25, S155–S160. [Google Scholar] [CrossRef] [PubMed]
  187. Dumont, M.; Lin, M.T.; Beal, M.F. Neuroprotective strategies involving ROS in Alzheimer disease. J. Alzheimers Dis. 2010, 20, S633–S643. [Google Scholar] [CrossRef] [PubMed]
  188. Dumont, M.; Beal, M.F. Neuroprotective strategies involving ROS in Alzheimer disease. Free Radic. Biol. Med. 2011, 51, 1014–1026. [Google Scholar] [CrossRef] [PubMed]
  189. Yang, X.; Dai, G.; Li, G.; Yang, E.S. Coenzyme Q10 reduces beta-amyloid plaque in an APP/PS1 transgenic mouse model of Alzheimer’s disease. J. Mol. Neurosci. 2010, 41, 110–113. [Google Scholar] [CrossRef] [PubMed]
  190. Hart, P.E.; Lodi, R.; Rajagopalan, B.; Bradley, J.L.; Crilley, J.G.; Turner, C.; Blamire, A.M.; Manners, D.; Styles, P.; Schapira, A.H.V.; et al. Antioxidant treatment of patients with Friedreich ataxia: Four-year follow-up. Arch. Neurol. 2005, 62, 621–626. [Google Scholar] [CrossRef] [PubMed]
  191. Larsen, P.L.; Clarke, C.F. Extension of life-span in Caenorhabditis elegans by a diet lacking coenzyme Q. Science 2002, 295, 120–123. [Google Scholar] [CrossRef] [PubMed]
  192. Dillin, A.; Hsu, A.-L.; Arantes-Oliveira, N.; Lehrer-Graiwer, J.; Hsin, H.; Fraser, A.G.; Kamath, R.S.; Ahringer, J.; Kenyon, C. Rates of behavior and aging specified by mitochondrial function during development. Science 2002, 298, 2398–2401. [Google Scholar] [CrossRef] [PubMed]
  193. Saiki, R.; Lunceford, A.L.; Bixler, T.; Dang, P.; Lee, W.; Furukawa, S.; Larsen, P.L.; Clarke, C.F. Altered bacterial metabolism, not coenzyme Q content, is responsible for the lifespan extension in Caenorhabditis elegans fed an Escherichia coli diet lacking coenzyme Q. Aging Cell 2008, 7, 291–304. [Google Scholar] [CrossRef] [PubMed]
  194. Lönnrot, K.; Holm, P.; Lagerstedt, A.; Huhtala, H.; Alho, H. The effects of lifelong ubiquinone Q10 supplementation on the Q9 and Q10 tissue concentrations and life span of male rats and mice. Biochem. Mol. Biol. Int. 1998, 44, 727–737. [Google Scholar] [CrossRef] [PubMed]
  195. Yan, J.; Fujii, K.; Yao, J.; Kishida, H.; Hosoe, K.; Sawashita, J.; Takeda, T.; Mori, M.; Higuchi, K. Reduced coenzyme Q10 supplementation decelerates senescence in SAMP1 mice. Exp. Gerontol. 2006, 41, 130–140. [Google Scholar] [CrossRef] [PubMed]
  196. Shetty, R.A.; Forster, M.J.; Sumien, N. Coenzyme Q10 supplementation reverses age-related impairments in spatial learning and lowers protein oxidation. Age 2012, 35, 1821–1834. [Google Scholar] [CrossRef] [PubMed]
  197. Tarry-Adkins, J.L.; Blackmore, H.L.; Martin-Gronert, M.S.; Fernandez-Twinn, D.S.; McConnell, J.M.; Hargreaves, I.P.; Giussani, D.A.; Ozanne, S.E. Coenzyme Q10 prevents accelerated cardiac aging in a rat model of poor maternal nutrition and accelerated postnatal growth. Mol. Metabol. 2013, 2, 480–490. [Google Scholar] [CrossRef] [PubMed]
  198. Ochoa, J.J.; Quiles, J.L.; López-Frías, M.; Huertas, J.R.; Mataix, J. Effect of lifelong coenzyme Q10 supplementation on age-related oxidative stress and mitochondrial function in liver and skeletal muscle of rats fed on a polyunsaturated fatty acid (PUFA)-rich diet. J. Gerontol. A Biol. Sci. Med. Sci. 2007, 62, 1211–1218. [Google Scholar] [CrossRef] [PubMed]
  199. Quiles, J.L.; Pamplona, R.; Ramirez-Tortosa, M.C.; Naudí, A.; Portero-Otin, M.; Araujo-Nepomuceno, E.; López-Frías, M.; Battino, M.; Ochoa, J.J. Coenzyme Q addition to an n-6 PUFA-rich diet resembles benefits on age-related mitochondrial DNA deletion and oxidative stress of a MUFA-rich diet in rat heart. Mech. Ageing Dev. 2010, 131, 38–47. [Google Scholar] [CrossRef] [PubMed]
  200. Ochoa, J.J.; Pamplona, R.; Ramirez-Tortosa, M.C.; Granados-Principal, S.; Perez-Lopez, P.; Naudí, A.; Portero-Otin, M.; López-Frías, M.; Battino, M.; Quiles, J.L. Age-related changes in brain mitochondrial DNA deletion and oxidative stress are differentially modulated by dietary fat type and coenzyme Q 10. Free Radic. Biol. Med. 2011, 50, 1053–1064. [Google Scholar] [CrossRef] [PubMed]
  201. González-Alonso, A.; Ramírez-Tortosa, C.L.; Varela-López, A.; Roche, E.; Arribas, M.I.; Ramírez-Tortosa, M.C.; Giampieri, F.; Ochoa, J.J.; Quiles, J.L. Sunflower oil but not fish oil resembles positive effects of virgin olive oil on aged pancreas after life-long coenzyme Q addition. Int. J. Mol. Sci. 2015, 16, 23425–23445. [Google Scholar] [CrossRef] [PubMed]
  202. Varela-Lopez, A.; Bullon, P.; Battino, M.; Ramirez-Tortosa, M.C.; Ochoa, J.J.; Cordero, M.D.; Ramirez-Tortosa, C.L.; Rubini, C.; Zizzi, A.; Quiles, J.L. Coenzyme Q Protects Against Age-Related Alveolar Bone Loss Associated to n-6 Polyunsaturated Fatty Acid Rich-Diets by Modulating Mitochondrial Mechanisms. J. Gerontol. A Biol. Sci. Med. Sci. 2015, glv063. [Google Scholar] [CrossRef] [PubMed]
  203. Cano, A.; Ciaffoni, F.; Safwat, G.M.; Aspichueta, P.; Ochoa, B.; Bravo, E.; Botham, K.M. Hepatic VLDL assembly is disturbed in a rat model of nonalcoholic fatty liver disease: Is there a role for dietary coenzyme Q? J. Appl. Physiol. 2009, 107, 707–717. [Google Scholar] [CrossRef] [PubMed]
  204. Ratnam, D.V.; Chandraiah, G.; Meena, A.K.; Ramarao, P.; Kumar, M.N.V.R. The co-encapsulated antioxidant nanoparticles of ellagic acid and coenzyme Q10 ameliorates hyperlipidemia in high fat diet fed rats. J. Nanosci. Nanotechnol. 2009, 9, 6741–6746. [Google Scholar] [CrossRef] [PubMed]
  205. Safwat, G.M.; Pisanò, S.; D’Amore, E.; Borioni, G.; Napolitano, M.; Kamal, A.A.; Ballanti, P.; Botham, K.M.; Bravo, E. Induction of non-alcoholic fatty liver disease and insulin resistance by feeding a high-fat diet in rats: Does coenzyme Q monomethyl ether have a modulatory effect? Nutrition 2009, 25, 1157–1168. [Google Scholar] [CrossRef] [PubMed]
  206. Sohet, F.M.; Neyrinck, A.M.; Pachikian, B.D.; de Backer, F.C.; Bindels, L.B.; Niklowitz, P.; Menke, T.; Cani, P.D.; Delzenne, N.M. Coenzyme Q10 supplementation lowers hepatic oxidative stress and inflammation associated with diet-induced obesity in mice. Biochem. Pharmacol. 2009, 78, 1391–1400. [Google Scholar] [CrossRef] [PubMed]
  207. Jiménez-Santos, M.A.; Juárez-Rojop, I.E.; Tovilla-Zárate, C.A.; Espinosa-García, M.T.; Juárez-Oropeza, M.A.; Ramón-Frías, T.; Bermúdez-Ocaña, D.Y.; Díaz-Zagoya, J.C. Coenzyme Q10 supplementation improves metabolic parameters, liver function and mitochondrial respiration in rats with high doses of atorvastatin and a cholesterol-rich diet. Lipids Health Dis. 2014, 13, 22. [Google Scholar] [CrossRef] [PubMed]
  208. Orlando, P.; Silvestri, S.; Brugè, F.; Tiano, L.; Kloting, I.; Falcioni, G.; Polidori, C. High-fat diet-induced met-hemoglobin formation in rats prone (WOKW) or resistant (DA) to the metabolic syndrome: Effect of CoQ10 supplementation. Biofactors 2014, 40, 603–609. [Google Scholar] [CrossRef] [PubMed]
  209. Bello, R.I.; Gómez-Díaz, C.; Burón, M.I.; Alcaín, F.J.; Navas, P.; Villalba, J.M. Enhanced anti-oxidant protection of liver membranes in long-lived rats fed on a coenzyme Q10-supplemented diet. Exp. Gerontol. 2005, 40, 694–706. [Google Scholar] [CrossRef] [PubMed]
  210. Gómez-Díaz, C.; Burón, M.I.; Alcaín, F.J.; González-Ojeda, R.; González-Reyes, J.A.; Bello, R.I.; Herman, M.D.; Navas, P.; Villalba, J.M. Effect of dietary coenzyme Q and fatty acids on the antioxidant status of rat tissues. Protoplasma 2003, 221, 11–17. [Google Scholar] [CrossRef] [PubMed]
  211. Santos-González, M.; Gómez Díaaz, C.; Navas, P.; Villalba, J.M. Modifications of plasma proteome in long-lived rats fed on a coenzyme Q10-supplemented diet. Exp. Gerontol. 2007, 42, 798–806. [Google Scholar] [CrossRef] [PubMed]
  212. Yubero-Serrano, E.M.; Delgado-Casado, N.; Delgado-Lista, J.; Perez-Martinez, P.; Tasset-Cuevas, I.; Santos-Gonzalez, M.; Caballero, J.; Garcia-Rios, A.; Marin, C.; Gutierrez-Mariscal, F.M.; et al. Postprandial antioxidant effect of the Mediterranean diet supplemented with coenzyme Q10 in elderly men and women. Age 2010, 33, 579–590. [Google Scholar] [CrossRef] [PubMed]
  213. Gutierrez-Mariscal, F.M.; Perez-Martinez, P.; Delgado-Lista, J.; Yubero-Serrano, E.M.; Camargo, A.; Delgado-Casado, N.; Cruz-Teno, C.; Santos-Gonzalez, M.; Rodriguez-Cantalejo, F.; Castaño, J.P.; et al. Mediterranean diet supplemented with coenzyme Q10 induces postprandial changes in p53 in response to oxidative DNA damage in elderly subjects. Age 2012, 34, 389–403. [Google Scholar] [CrossRef] [PubMed]
  214. Yubero-Serrano, E.M.; Gonzalez-Guardia, L.; Rangel-Zuñiga, O.; Delgado-Lista, J.; Gutierrez-Mariscal, F.M.; Perez-Martinez, P.; Delgado-Casado, N.; Cruz-Teno, C.; Tinahones, F.J.; Villalba, J.M.; et al. Mediterranean diet supplemented with coenzyme Q10 modifies the expression of proinflammatory and endoplasmic reticulum stress-related genes in elderly men and women. J. Gerontol. A Biol. Sci. Med. Sci. 2012, 67A, 3–10. [Google Scholar] [CrossRef] [PubMed]
  215. González-Guardia, L.; Yubero-Serrano, E.M.; Delgado-Lista, J.; Perez-Martinez, P.; Garcia-Rios, A.; Marin, C.; Camargo, A.; Delgado-Casado, N.; Roche, H.M.; Perez- Jimenez, F.; et al. Effects of the mediterranean diet supplemented with coenzyme Q10 on metabolomic profiles in elderly men and women. J. Gerontol. A Biol. Sci. Med. Sci. 2015, 70, 78–84. [Google Scholar] [CrossRef] [PubMed]
  216. Gutierrez-Mariscal, F.M.; Yubero-Serrano, E.M.; Rangel-Zuñiga, O.A.; Marín, C.; García-Rios, A.; Perez-Martinez, P.; Delgado-Lista, J.; Malagón, M.M.; Tinahones, F.J.; Pérez-Jimenez, F.; et al. Postprandial activation of p53-dependent DNA repair is modified by mediterranean diet supplemented with coenzyme Q10 in elderly subjects. J. Gerontol. A Biol. Sci. Med. Sci. 2014, 69, 886–893. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Alternative redox forms of Coenzyme Q.
Figure 1. Alternative redox forms of Coenzyme Q.
Molecules 21 00373 g001
Figure 2. Different activities exerted by coenzyme Q in cells: (a) coenzyme Q has a central role in mitochondrial electron transport chain; (b) plasma membrane redox system components; (c) antioxidant roles of coenzyme Q (this can occurs in all membranes where ubiquinol is present); (d) coenzyme Q prevent permeability transition pore opening; (e) coenzyme Q participates in lysosomal pH maintenance; (f) coenzyme Q stimulatory ( Molecules 21 00373 i001) or inhibitory ( Molecules 21 00373 i002) effects on gene expression. Abbreviations: α-tp: α-tocopherol; α-tp*: α-tocopheryl anion radical; CoQ: oxidized coenzyme Q; CoQ·H: partially reduced coenzyme Q; CoQH2: fully reduced coenzyme Q; Cytb5R: NADH: cytochrome b5 reductase; e: electron; FAD+: reduced flavine adenine dinucelotide; FADH2: reduced flavine adenine dinucleotide; H2O2: hydrogen peroxide; LOO*: lipid hydroperoxyl radical; LOOH: lipid hydroperoxyde; MAPKs: mitogen-activated protein kinases; mPTP: mitochondrial permeability transition pore; NAD+: oxidized nicotine adenine dinucleotide; NADH: reduced nicotine adenine dinucleotide; NQO1: NAD(P)H:Ubiquinone oxidase 1; O2·: superoxide anion; ROS: reactive oxygen species; SIRT1: sirtuin1.
Figure 2. Different activities exerted by coenzyme Q in cells: (a) coenzyme Q has a central role in mitochondrial electron transport chain; (b) plasma membrane redox system components; (c) antioxidant roles of coenzyme Q (this can occurs in all membranes where ubiquinol is present); (d) coenzyme Q prevent permeability transition pore opening; (e) coenzyme Q participates in lysosomal pH maintenance; (f) coenzyme Q stimulatory ( Molecules 21 00373 i001) or inhibitory ( Molecules 21 00373 i002) effects on gene expression. Abbreviations: α-tp: α-tocopherol; α-tp*: α-tocopheryl anion radical; CoQ: oxidized coenzyme Q; CoQ·H: partially reduced coenzyme Q; CoQH2: fully reduced coenzyme Q; Cytb5R: NADH: cytochrome b5 reductase; e: electron; FAD+: reduced flavine adenine dinucelotide; FADH2: reduced flavine adenine dinucleotide; H2O2: hydrogen peroxide; LOO*: lipid hydroperoxyl radical; LOOH: lipid hydroperoxyde; MAPKs: mitogen-activated protein kinases; mPTP: mitochondrial permeability transition pore; NAD+: oxidized nicotine adenine dinucleotide; NADH: reduced nicotine adenine dinucleotide; NQO1: NAD(P)H:Ubiquinone oxidase 1; O2·: superoxide anion; ROS: reactive oxygen species; SIRT1: sirtuin1.
Molecules 21 00373 g002

Share and Cite

MDPI and ACS Style

Varela-López, A.; Giampieri, F.; Battino, M.; Quiles, J.L. Coenzyme Q and Its Role in the Dietary Therapy against Aging. Molecules 2016, 21, 373. https://doi.org/10.3390/molecules21030373

AMA Style

Varela-López A, Giampieri F, Battino M, Quiles JL. Coenzyme Q and Its Role in the Dietary Therapy against Aging. Molecules. 2016; 21(3):373. https://doi.org/10.3390/molecules21030373

Chicago/Turabian Style

Varela-López, Alfonso, Francesca Giampieri, Maurizio Battino, and José L. Quiles. 2016. "Coenzyme Q and Its Role in the Dietary Therapy against Aging" Molecules 21, no. 3: 373. https://doi.org/10.3390/molecules21030373

Article Metrics

Back to TopTop