Next Article in Journal
Enteral Nutrition in Adult Crohn’s Disease: Toward a Paradigm Shift
Next Article in Special Issue
Heart Histopathology and Mitochondrial Ultrastructure in Aged Rats Fed for 24 Months on Different Unsaturated Fats (Virgin Olive Oil, Sunflower Oil or Fish Oil) and Affected by Different Longevity
Previous Article in Journal
Impact of FODMAP Content Restrictions on the Quality of Diet for Patients with Celiac Disease on a Gluten-Free Diet
Previous Article in Special Issue
Sugar Beet Pectin Supplementation Did Not Alter Profiles of Fecal Microbiota and Exhaled Breath in Healthy Young Adults and Healthy Elderly
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

The Paradox of Coenzyme Q10 in Aging

by
M. Elena Díaz-Casado
1,2,*,†,
José L. Quiles
3,
Eliana Barriocanal-Casado
1,2,
Pilar González-García
1,2,
Maurizio Battino
4,5,6,
Luis C. López
1,2 and
Alfonso Varela-López
3,*,†
1
Institute of Biotechnology, Department of Physiology, Biomedical Research Center, University of Granada, Avda del Conocimiento sn, 18016 Granada, Spain
2
Centro de Investigación Biomédica en Red de Fragilidad y Envejecimiento Saludable (CIBERFES), 18016 Granada, Spain
3
Institute of Nutrition and Food Technology “José Mataix Verdú”, Department of Physiology, Biomedical Research Center, University of Granada, Avda del Conocimiento sn, 18016 Granada, Spain
4
Department of Clinical Sicences, Università Politecnica delle Marche, 60131 Ancona, Italy
5
Nutrition and Food Science Group, Department of Analytical and Food Chemistry, CITACA, CACTI, University of Vigo, 36310 Vigo, Spain
6
International Research Center for Food Nutrition and Safety, Jiangsu University, Zhenjiang 212013, China
*
Authors to whom correspondence should be addressed.
The two authors contributed equally.
Nutrients 2019, 11(9), 2221; https://doi.org/10.3390/nu11092221
Submission received: 9 August 2019 / Revised: 6 September 2019 / Accepted: 8 September 2019 / Published: 14 September 2019
(This article belongs to the Special Issue Nutrition, Diet and Longevity)

Abstract

:
Coenzyme Q (CoQ) is an essential endogenously synthesized molecule that links different metabolic pathways to mitochondrial energy production thanks to its location in the mitochondrial inner membrane and its redox capacity, which also provide it with the capability to work as an antioxidant. Although defects in CoQ biosynthesis in human and mouse models cause CoQ deficiency syndrome, some animals models with particular defects in the CoQ biosynthetic pathway have shown an increase in life span, a fact that has been attributed to the concept of mitohormesis. Paradoxically, CoQ levels decline in some tissues in human and rodents during aging and coenzyme Q10 (CoQ10) supplementation has shown benefits as an anti-aging agent, especially under certain conditions associated with increased oxidative stress. Also, CoQ10 has shown therapeutic benefits in aging-related disorders, particularly in cardiovascular and metabolic diseases. Thus, we discuss the paradox of health benefits due to a defect in the CoQ biosynthetic pathway or exogenous supplementation of CoQ10.

1. Introduction

1.1. CoQ Biosynthesis

Coenzyme Q (CoQ), ubiquinone or 2,3-dimethoxy-5-methyl-6-polyprenyl-1,4-benzoquinone is a two-part molecule composed of a benzoquinone ring, which has redox active sites, and a long polyisoprenoid lipid chain that positions the molecule in the mid-plane of a membrane bilayer. The length of the chain depends on the species, i.e., 10 isoprene units (CoQ10) in humans and S. pombe, eight units (CoQ8) in Escherichia coli, six units (CoQ6) in Saccharomyces cerevisiae and nine units (CoQ9) in rodents. Nevertheless, some species have more than one CoQ form, e.g., human cells and tissues also contain CoQ9 and rodent cells and tissues also contain CoQ10 as minor forms.
The synthesis of CoQ in eukaryotes mainly occurs in mitochondria by a set of nuclear-encoded CoQ proteins through a biochemical pathway that has not been fully defined. Moreover, some authors have suggested that CoQ may also be produced outside of mitochondria, mainly in the endoplasmic reticulum-Golgi [1,2]. On the other hand, recent studies suggest that the endoplasmic reticulum mitochondria encounter structure (ERMES) might promote CoQ biosynthesis and the movement of CoQ and its precursors and intermediates between these organelles [3]. Thus, the mitochondrial production of CoQ could contribute to extramitochondrial pools of CoQ. The known steps required for CoQ production were first identified in S. cerevisiae and later confirmed in mammals. Specifically, S. cerevisiae CoQ (CoQ1–CoQ9) mutants, but also E. coli ubi (ubiA-J and ubiX) mutants, have allowed identifying the CoQ genes and biosynthetic intermediaries of the biosynthetic pathway of CoQ [4,5,6,7,8,9]. At least 11 genes (ubiA, B, C, D, E, F, G, H, I, J, and X) in E. coli and 13 genes (YAH1, ARH1, and CoQ1–CoQ11) in S. cerevisiae are involved in CoQ biosynthesis. Human and mouse orthologues for almost all of these genes have already been identified (with the possible exception of CoQ11). In the cases of CoQ8 and CoQ10, humans have two orthologues for each one: ADCK3 (CoQ8A) and ADCK4 (CoQ8B), and CoQ10A and CoQ10B, respectively. In Homo sapiens the polypernyl diphosphate synthase is a heterotetramer formed by the dimer of PDSS1 and PDSS2 proteins [10,11]. The expression of the human homologs (CoQ1–CoQ10) for yeast genes restores CoQ6 biosynthesis in the corresponding S. cerevisiae CoQ mutants, indicating the profound functional conservation in human cells [12].
Structurally, CoQ biosynthesis could be divided into four stages: quinone synthesis, isoprenoid tail synthesis, both molecules condensation and ring modification (Figure 1a).

1.1.1. Quinone Synthesis

The primary aromatic precursor of the benzoquinone ring is 4-hydroxybenzoic acid (4HB). For the biosynthesis of 4HB, human cells utilize phenylalanine or tyrosine as ring precursors. Many of the steps involved in the generation of 4HB from tyrosine are not completely understood and the responsible enzymes are unidentified [13,14]. Current studies have identified the first and last reactions of this pathway and the enzymes involved in these reactions in yeast. Thus, 4HB synthesis starts with the deamination of tyrosine to 4-hydroxyphenylpyruvate by Aro8 and Aro9 and finishes with the oxidation of 4-hydroxybenzaldehyde to 4HB by Hfd1. Human Hdf1 homolog ALDH3A1 plays a role in CoQ biosynthesis and is able to rescue CoQ deficiency in yeast with the inactivated HFD1 gene [15,16]. Aro8 and Aro9 human homolog remain unidentified, but α-aminoadipate aminotransferase (AADAT) and the tyrosine aminotransferase (TAT) have been postulated as candidates [17].
Although 4HB is depicted as the main precursor of CoQ, other aromatic ring precursors may be incorporated into CoQ biosynthesis. The 2,4-dihydroxybenzoic acid (2,4-diHB), 3,4-dihydroxybenzoic acid (3,4-diHB) and vanillic acid are analogs of 4HB that may allow the stimulation and/or bypass of CoQ biosynthesis defects in yeast CoQ6 and CoQ7 mutants, as well as in CoQ7 conditional knockout mice, CoQ9R239X mice and human skin fibroblasts from patients with mutations in CoQ7 and CoQ9 [18,19,20,21,22,23]. Para-coumarate and resveratrol also can be used as CoQ ring precursors in E. coli, S. cerevisiae, and human and mouse cells [24]. Moreover, in mouse and human kidney cells, kaempferol is able to increase CoQ levels. This phenolic compound apparently acts as a biosynthetic ring precursor competing with 4HB [25].

1.1.2. Polyisoprenyl Tail Synthesis

The lipophilic polyprenyl tail is synthetized by the addiction of dimethylally pyrophosphate (DMAPP) and isopentenyl pyrophosphate (IPP) to form decaprenyl diphosphate and nonaprenyl diphosphate via PDSS1/PDSS2 in humans and mice. Consequently, the heterotetramer PDSS1/PDSS2 is responsible for the length of the polyprenyl tail. The polyprenyl precursors come from the mevalonate pathway and are generated in the cytosol. The mechanism by which these precursors are transported from the cytosol to mitochondria is unknown, and it has been suggested that some transporters may exist in the inner mitochondrial membrane [17].

1.1.3. Attachment of the Ring/Chain

Furthermore, 4HB and polyprenyl tail are condensed in a reaction catalyzed by human or mice CoQ2 to form 3-decaprenyl-4HB and 3-nonaprenyl-4HB (Figure 1a).

1.1.4. Next Steps of CoQ Biosynthesis-Ring Modifications

Then, 3-decaprenyl-4HB or 3-nonaprenyl-4HB undergoes subsequent modifications of the ring through a set or reactions of methylations, decarboxylation and hydroxylations to finally generate the molecule of CoQ10 or CoQ9.
The first reaction is the C5-hydroxylation, catalyzed by CoQ6, to produce 3-decaprenyl-4,5-dihydroxybenzoic acid and 3-nonaprenyl-4,5-dihydroxybenzoic acid. In yeast, this step requires two enzymes, ferredoxin Yah1p and ferredoxin reductase Arh1p, which provide electrons for CoQ6 activity. It is unknown whether the mammalian orthologs, ferredoxin reductase (FDXR) and ferredoxin 2 (FDX2), respectively, act in mammalian CoQ biosynthesis. CoQ6 is characterized as a flavin-dependent monooxygenase peripherally associated with inner mitochondrial membrane on the matrix face [26,27].
C5-hydroxylation is followed by O-methylation catalyzed by CoQ3 to form 3-decaprenyl-4-hydroxy-5-methoxybenzoic acid and 3-nonaprenyl-4-hydroxy-5-methoxybenzoic acid. CoQ3 is an S-adenosylmethionine-dependent methyltransferase that is involved in two O-methylation reactions in the CoQ biosynthetic pathway. This polypeptide is also peripherally associated with the mitochondrial inner membrane on the matrix face [28,29].
The proteins involved in the subsequent steps, C1-decarboxylation and C1-hydroxylation, have not been identified in the eukaryotes. However, in prokaryotes the decarboxylation and hydroxylation in prokaryotes are catalyzed by the 3-Octaprenyl-4-hydroxybenzoate decarboxylase UbiD and the 2-octaprenyl-6-methoxyphenol hydroxylase UbiH, respectively [30,31]. In prokaryotic decarboxylation, the flavin prenyltransferase UbiX is important for synthesizing prenylated flavin, which is used as a cofactor for UbiD [32]. Nevertheless, there is no sequence homolog for UbiD or UbiX in humans, suggesting that the C1-hydroxylation could be part of the decarboxylation mechanism.
The ring is then further modified by C2-methylation in a reaction catalyzed by CoQ5. In this reaction, 2-decaprenyl-6-methoxy-1,4-benzenediol and 2-nonaprenyl-6-methoxy-1,4-benzenediol is converted to 2-decaprenyl-3methyl-6-methoxy-1,4-benzenediol and 2-nonaprenyl-3methyl-6-methoxy-1,4-benzenediol (DMQH2) [33]. CoQ5 is an S-adenosyl methionine (SAM)-dependent methyltransferase (SAM-MTase) that is peripherally associated with the mitochondrial inner membrane on the matrix face [34]. DMQH2 is converted to 2-decaprenyl-3methyl-6methoxy-1,4,5-benzenetriol and 2-nonaprenyl-3methyl-6methoxy-1,4,5-benzenetriol (DMeQH2) in a C6-hydroxylation catalyzed by CoQ7, which is a carboxylate bridged diiron hydroxylase peripherally associated with the mitochondrial inner membrane on the matrix side [35].
The final step of the CoQ biosynthetic pathway is an O-methylation also catalyzed by CoQ3 (Figure 1a).

1.1.5. Other Mitochondrial Proteins Involved in CoQ Biosynthesis

There are additional proteins needed for CoQ biosynthesis since their dysfunctions induce a decline in the levels of CoQ. However, their molecular exact functions are yet unclear. These are the cases of CoQ8A, CoQ8B, CoQ9, CoQ10A, CoQ10B and CoQ4.
CoQ8A (=ADCK3) and CoQ8B (=ADCK4) have been related with CoQ10 biosynthesis and mutations in these proteins are associated with CoQ10 deficiency [36,37,38,39]. These two proteins are considered the human orthologues of yeast CoQ8p because CoQ6 biosynthesis is rescued by the expression of human CoQ8A and CoQQB in CoQ8-mutant yeast strains [39,40]. However, the precise biochemical activity and the role of CoQ8 in CoQ biosynthesis remains unknown. Although initially CoQ8 was hypothesized to be a protein kinase that may phosphorylate other CoQ biosynthetic proteins, recent studies have demonstrated that CoQ8A shows ATPase activity that is essential for CoQ biosynthesis. In any case, it seems that the activity of CoQ8A could stabilize Complex Q [41].
CoQ9 is a lipid-binding protein associated with the inner mitochondrial membrane, on the matrix face [42]. This protein binds aromatic isoprenes with high specificity, including CoQ intermediates that likely reside within the bilayer. Therefore, CoQ9 seems to present the intermediates of the CoQ biosynthetic pathway directly to CoQ enzymes [43]. CoQ9 is also necessary for the stability and activity of CoQ7, so CoQ9 mutant mice and human cells show reduced levels of CoQ7 and accumulation of DMQH2, the substrate of the reaction catalyzed by CoQ7 [23,27,44,45,46].
CoQ10A and CoQ10B in humans [47] are homologues of the yeast CoQ10, a protein that contains a lipid-binding domain that binds CoQ and CoQ intermediates, and that is required for efficient biosynthesis of CoQ6 and electron transport in the mitochondrial respiratory chain. The molecular basis for these effects is unknown, but CoQ10 probably directs to CoQ from synthesis site to function sites [48,49,50].
Finally, the function of the CoQ4 is still unknown, although some studies suggest that it is required for the assembly and stability of Complex Q, as it is associated with CoQ3, CoQ5, CoQ6 and CoQ9 [28,51,52]. Also, it is speculated that CoQ4 could bind to CoQ and CoQ intermediates and it could organize the enzymes that perform the ring modifications. In humans, there are two different isoforms of the CoQ4 polypeptide but only one is localized in mitochondria [53].

1.1.6. Complex Q

CoQ biosynthesis depends on a multi-subunit CoQ polypeptide complex formed by several CoQ biosynthetic proteins and by non-proteinaceous components, including some CoQ intermediates and CoQ itself. It is widely accepted that CoQ3–CoQ9 assembly into a high molecular mass complex called Complex Q. These complexes are peripherally associated with the matrix-side of the inner mitochondrial membrane and are essential feature for the de novo biosynthesis of CoQ [17,51,54,55]. In fact, the lack of a single CoQ polypeptide in yeasts, mice and human cells results in the degradation of several CoQ proteins [23,40,42,45,56].
The functional role of Complex Q is to enhance the catalytic efficiency, optimizing the orientation of the substrates and active sites of the enzymes, and to minimize the escape of intermediates, which could be toxic due to their redox or electrophilic properties [17,45,54]. However, many questions about Complex Q remain unknown, including its regulation and the key components involved in its formation and stability.

1.2. Functions of CoQ

The lipophilic characteristic of CoQ together with its redox capacity allows this molecule to participate in multiple cellular pathways and functions.

1.2.1. The Role of CoQ in the Mitochondrial Respiratory Chain

The best-known role of CoQ is its function as an essential electron carrier in the mitochondrial respiratory chain (MRC). CoQ passes electrons between complex I (NADH-ubiquinone oxidoreductase) or Complex II (succinate-ubiquinone oxidoreductase) and Complex III (succinate-cytochrome c oxidoreductase) [57] (Figure 1b). Thus, CoQ can be found in both oxidized (CoQ or ubiquinone) and reduced forms (CoQH2 or ubiquinol), and the conversion between these oxidized-reduced states allows it to act as a cofactor of enzymatic reactions transferring electrons to substrates. This oxidation–reduction cycle may occur by a two-step transfer of electrons producing a semiquinone (CoQH·) intermediate.
In mitochondria, CoQ also accepts electrons from sulfide quinone oxidoreductase (SQOR) during sulfide detoxification [58,59,60,61]; proline dehydrogenase 1 (PDH), an enzyme required for proline and arginine metabolism [62]; coline dehydrogenase (CHDH), required for glycine metabolism; mitochondrial glycerol-3-phosphate dehydrogenase (G3PDH), which connects oxidative phosphorylation, fatty acid metabolism and glycolysis [63]; dihydroorotate dehydrogenase (DHOH), an enzyme involved in pyrimidine biosynthesis [64,65,66]; electron transport flavoprotein dehydrogenase (ETFDH), a key enzyme involved in the fatty acid β-oxidation [67]; and hydroxylproline dehydrogenase, involved in the glyoxylate metabolism [68] (Figure 1b). All these processes generate CoQH2, which is re-oxidized by complex III. Therefore, CoQ is an excellent redox carrier involved in different cellular processes since it has the ability to sustain continuous oxidation–reduction cycles.

1.2.2. The Antioxidant Capacity

Reactive Oxygen Species (ROS) formed in the cell are able to damage lipids, proteins and DNA. The mitochondrion is the compartment considered to be the major source of ROS production in the cell and, therefore, this organelle is highly susceptible to suffer oxidative damage. To avoid this harmful effect, the mitochondrion possesses antioxidant compounds and systems, being CoQ one of them. CoQ is one of the most powerful endogenously synthesized membrane antioxidants, being present in all membranes. Its antioxidant function efficiently protects lipids from harmful oxidative damage, but also DNA and proteins [69,70].
CoQ is able to prevent lipid peroxidation in most subcellular membranes and its effectiveness to inhibit lipid peroxidation depends on its complex interaction during the peroxidation process. CoQH2 has the ability to both prevent production and eliminate directly lipid peroxyl radicals (LOO), while vitamin E acts by quenching these radicals. Also, CoQH2 regenerates vitamin E from the α-tocopheroxyl radical [71]. CoQ is considered an efficient antioxidant against radicals produced in membranes for its regeneration capacity, its lipid solubility and because of its involvement in the initiation and propagation of lipid peroxidation [72].
CoQ is effective in the prevention of DNA damage as shown by the decreased strand breaks in DNA and other oxidative DNA damage markers in mitochondria [73], which is important for mitochondrial DNA since its oxidative damage is approximately 10-fold higher than nuclear DNA and is not easily reparable. These data are supported by Tomasetti and colleagues [74,75] who showed that CoQ10 supplementation inhibits DNA strand break formation and oxidative damage and enhances DNA repair enzyme activities in lymphocytes. Both the antioxidant activity of CoQ10 and the role that CoQ10 may play in the redox mechanism implicated in the transactivation of DNA repair enzymes could explain the capability of CoQ10 to protect DNA from oxidative damage. However, the precise mechanism under these effects remains unclear.
The high efficiency of CoQ as an antioxidant is related to its effective reactivation redox, its ubiquitous distribution, its localization in membranes, its relatively high concentration and the large capacity of the cell to regenerate CoQ at all its locations [76,77].

1.2.3. Other Controversial Functions

It has been suggested that CoQ is one of the compounds influencing the mitochondrial permeability transition pore (PTP). Studies about the regulation of the PTP by ubiquinone analogues show that ubiquinone inhibits the Ca2+ dependent PTP opening, which is mediated through a ubiquinone-binding site directly involved in PTP regulation rather than through redox reactions [78]. However, further details about the organization, structure, components and regulation of the PTP are required to understand whether CoQ has a primary function in the PTP. Other mitochondrial components susceptible to being regulated by redox reactions are the uncoupling proteins (UCPs). However, controversial results have been reported regarding to the involvement of CoQ in the regulation of the UCPs [79,80,81,82,83].

1.3. Pathologic Conditions with Decreased Levels of CoQ

The levels of CoQ are quite stable in cells. However, CoQ levels can be severely reduced in a group of mitochondrial diseases called CoQ deficiencies, which are clinically and genetically heterogeneous disorders characterized by a decrease in the levels of CoQ in tissues or cells. If the deficiency is caused by pathogenic mutations in the genes required for CoQ10 biosynthesis, it is classified as primary CoQ10 deficiencies. Nevertheless, secondary CoQ10 deficiencies are caused by mutations in genes unrelated to CoQ biosynthesis or are derived from other physiological processes or pharmacological treatments. The secondary forms are probably more frequent than the primary defects.
CoQ deficiencies cause the inhibition of oxidative phosphorylation and ATP production, since CoQ is an essential component of the mitochondrial respiratory chain. Because CoQ has other roles unrelated to ATP production, CoQ deficiency can impair other vital cellular functions, such as sulfide metabolism [58,59,60], and may induce oxidative damage [46,84,85,86].

1.3.1. Primary CoQ10 Deficiencies

Primary ubiquinone deficiency is a subset of mitochondrial diseases caused by autosomal recessive mutations in CoQ genes. Ogasahara et al. [87] described the first patients with primary CoQ10 deficiency as two sisters with progressive muscle weakness, fatigability and central nervous system dysfunction, learning disability, and seizures in one and cerebellar syndrome in the other [87]. Since then, approximately 200 patients have been described with pathogenic mutations in PDSS1 [88], PDSS2 [89,90], CoQ2 [88,89,91], CoQ6 [89,92], CoQ7 [22], CoQ4 [93,94,95], CoQ5 [96], CoQ8A [36,37,38], CoQ8B [89,97] and CoQ9 [98] genes. The clinical manifestations are very heterogeneous among different genes and among patients with mutations in individual genes. Multiple organs may be affected and lead to different symptomatology; when the central nervous system (CNS) is affected the clinical manifestations are variable, including encephalopathy, seizures, cerebellar ataxia, epilepsy, intellectual disability, hypotonia, dystonia, spasticity; in kidney, the most common manifestations is steroid-resistant nephrotic syndrome (SRNS); myopathy in muscle and hypertrophic cardiomyopathy in heart. Many symptoms are common to other mitochondrial diseases [99].

1.3.2. Secondary CoQ10 Deficiencies

Secondary CoQ deficiencies may affect any of the multiple CoQ biological functions and its connection to other metabolic pathways [100,101]. Low CoQ levels have been observed in a wide spectrum of diseases, such as oxidative phosphorylation (OXPHOS) diseases due to defects in nDNA-encoded proteins, OXPHOS diseases due to defects in mtDNA-encoded proteins and non-OXPHOS diseases (mitochondrial energy metabolism disorders and other non-mitochondrial diseases). Examples of diseases that may display secondary CoQ deficiencies include mitochondrial myopathies, mitochondrial DNA depletion syndrome, multiple acyl-CoA dehydrogenase deficiency (by mutations in ETFDH), ataxia-oculomotor apraxia syndrome (by mutations in APTX), cardiofaciocutaneous syndrome (by mutations in BRAF), methylmalonic aciduria, GLUT1 deficiency syndrome, mucopolysaccharidosys type III, spinocerebellar ataxia-10 (by mutations in ANO10) or multisystem atrophy (MSA) [102,103,104,105,106]. The symptoms are highly dependent on the original pathology and the most common symptoms are muscular and central nervous system manifestations. The exact mechanisms by which CoQ deficiency occurs are still unknown but it has been speculated that the low CoQ values in patients with mitochondrial diseases may be explained by a combination of different mechanisms, such as an incomplete or abnormal respiratory chain that may affect the formation of Complex Q linked to the respiratory chain (resulting in alterations in CoQ biosynthesis); an increase in CoQ degradation as a consequence of the oxidative stress derived from OXPHOS dysfunction; an interference with signaling pathways that potentially regulate CoQ biosynthesis; or a general deterioration of mitochondrial function [55,107].
Other diseases such as liver cirrhosis, phenylketonuria, fibromyalgia and cardiomyopathies also present a decrease in plasma CoQ levels [101,108], although this parameter is not reliable for a diagnostic purpose [109]. The use of statins to treat hypercholesterolemia can cause secondary CoQ deficiency because statins inhibit the 3-hydroxy-3-methyl-glutaryl-coenzyme A reductase (HMGCR) in the mevalonate pathway, which is required to produce farnesyl pyrophosphate, an intermediate in CoQ biosynthesis [110]. Also, a decrease in the levels of the components of Complex Q has recently been related to secondary CoQ deficiency in various mouse models of mitochondrial diseases [111], as well as in muscle and adipose tissue of patients and a mouse model with insulin resistance [112].

2. Long-Life Animal Models with CoQ Deficiency: Potential Mechanisms

The multiple metabolic pathways that require the function of CoQ can be altered during aging. Also, the function of CoQ as an antioxidant may be important in aging since it is widely reported that there is an increase in oxidative stress with aging. Derived from the free radical theory on aging introduced by Harman et al. [113], oxidative stress resistance and longevity are closely related [113]. Later, Harman [114] published an extension of this theory, showing the mitochondria as a primary source and target of ROS [114]. For that reason, it was unexpected and surprising that some models with defects in the CoQ biosynthetic pathway have increased life expectancy. An explanation for such data has been attempted by introducing the concept of mitochondrial hormesis or mitohormesis, which means that mild mitochondrial stress in early stages of life may induces physiological responses and adaptations in the organism, resulting in a longer life span [115].

2.1. Worm Models

Caenorhabditis elegans has been used extensively as a model organism in different fields of research, including aging. Clk-1 (CoQ7 homologous)-deficient worms are characterized by an increased life span together with a deregulation and slowing down of a number of physiological processes, developmental processes and behavioral patterns, e.g., embryonic and postembryonic development, reproduction, periods of defecation, swimming and pumping cycles [116,117,118]. Because the Clk-1 gene encodes demethoxyubiquinone hydrosylase (CoQ7), then clk-1 mutant worms are deficient in CoQ7 and, therefore, they cannot produce CoQ9 and consequently, accumulate large quantities of demethoxyubiquinone (DMQ9) [119].
The mechanism underlying the increased longevity in clk-1 mutants is not clear. Different hypotheses have been proposed. Initially, it was proposed that the low metabolic rate of the clk-1 mutants could result in a slower production of ROS, which could explain the prolonged life span in these worms [120,121]. Low levels of ROS on the mutants may affect the ras pathway and reduce the levels of oxidized LDL-like lipoproteins. These changes may alter the oxidative modifications of cellular constituents in clk-1 mutants during germline development [122]. However, Braeckman and colleagues [123] did not find significant changes either in metabolic rate or in the activity of two antioxidant enzymes, superoxide dismutase and catalase, in clk-1 mutants. Also, the same authors reported a mild reduction in oxygen consumption rates and elevated ATP levels in clk-1 mutants, suggesting an increase in the efficiency of the OXPHOS system [123,124,125,126].
Later, it was hypothesized that the clk-1 mutants prolong life span by a decrease in CoQ levels whether this is directly due to a clk-1 mutation or a diet lacking CoQ [127]. Also, Asencio and colleagues [128] showed that repressing different CoQ genes in C. elegans also extends life span [128], suggesting that the extension of life span could be related to the reduction in the levels of CoQ, rather than to a specific CoQ gene or to the accumulation of DMQ. The authors attributed the extension of life span to the maintenance of the efficiency of respiration together with a reduction in the production of superoxide anion. Nevertheless, Kayser and colleagues [129] showed a profound defect in complex I+III activity in clk-1 C. elegans mutants, a fact that may contribute to ROS production and its link to life span. Also, we need to take into account that clk-1 mutant worms obtain significant amounts of CoQ8 from diet, which would help to stabilize complex III. In fact, when clk-1 mutants are fed with E. coli mutants that lack CoQ, they show developmental arrest and sterility [119,130], while clk-1 worms fed with genetically engineered bacteria that produce CoQ6, CoQ7, CoQ8, CoQ9 or CoQ10 show increased life span, decreased mitochondrial respiration and decreased oxidative damage [118,131]. Similarly, mutants in mev-1, a gene that encodes cytochrome b, show elevated superoxide anion production in mitochondria and the exogenous administration of CoQ10 reduced superoxide anion levels and extended life span [132]. In addition, exogenous CoQ10 supply partially restores the expression of a cluster of genes important for growth and oxidation reactions, which are down and up regulation, respectively, in clk-1 mutants [133].
On the other hand, DMQ could be a non-functional competitor of CoQ, because it could bind the Q sites in the mitochondrial complexes but it cannot functionally replace CoQ9 [118,129,134]. In fact, clk-1 mutants cannot grow on bacteria producing only DMQ and those fed with a CoQ-less diet die as larvae [130,134]. However, another assay showed that DMQ9 would support a sufficient level of respiration and support respiration in later development [129]. The antioxidant properties of DMQ has also been proposed to be responsible for the extension of life span in clk-1 mutants, although studies in yeast have showed that DMQ is a less effective antioxidant than CoQ [135,136]. Probably, rhodoquinone (RQ9) and bacterial CoQ8 may also have an influence on life span [127]. All these data support an important role of the CoQ in the aging process in C. elegans, minimum levels of CoQ being necessary for adequate embryo development and fertility.
In summary, the reduction in the electron flow on the respiratory chain in clk-1 mutants worms may explain the low levels of ROS, a physiological adaptation and the increase in life span [127,129] (Figure 2a). Furthermore, the reduced respiration of clk-1 mutants suggests that the reduction in mitochondrial function determines longevity by a defect specific to complex I-dependent substrates.

2.2. Mouse Models

The relationship between a dysfunctional CoQ7 protein or the reduction in the CoQ levels from the embryonic development and the extension of life span has also been observed in two different mouse models, as reported in heterozygous mclk1 (=CoQ7) mice [137] and CoQ9Q95X mice [138].
Heterozygous mclk1 mice have some common characteristics with clk-1 mutant worms. Although these mice have normal CoQ levels in most of the tissues, they show a reduced level of MCLK1 protein (approximately 2-fold), reduced ROS levels, decreased oxidative stress and decreased oxidative damage, and their life span is 15–30% longer than wild-type mice [137,139,140].
In livers of old mclk1+/ mice, a loss of mclk1 expression in most of the cells of the entire hepatic lobules was observed. These cells are more resistant to age-dependent oxidative stress and apoptosis, and they can be propagated in old livers. Presumably, for this reason, the CoQ levels in the liver of old mclk1+/ mice are lower than in the liver of old mclk1+/+ mice. However, DMQ is not detectable in any organ of mclk1+/ mice, in contrast to what was observed in worms, suggesting that the increased life span of these mice cannot be due to the presence of DMQ. Moreover, the levels of CoQ in the tissues of young mclk1+/ mice are similar to those in the tissues of mclk1+/+ mice. Therefore, these results suggest that the reduction in the levels of MCLK1 and/or the mild decrease in CoQ levels in the liver of old mclk1+/ mice could be responsible for the increased resistance to hepatic oxidative stress and, consequently, for the increased life span in mclk1+/ mice [137,140]. Similar connections have been found in others genetic models, e.g., the p66sh/ mouse model, the igf1r+/ mouse model and the long-lived dwarf mouse model. In all these cases, a relationship between an increased resistance to oxidative stress and an extension of life span has been identified [141,142,143,144].
More recent studies have suggested that CLK-1/MCLK1 could have other roles in mitochondrial function and that CoQ levels are not totally responsible for the extension of life span. Lapointe and Hekimi [145] showed that the reduction in MCLK1 expression does not affect the levels of CoQ in young mclk1+/ mice. However, the electron transport chain, ATP synthesis and the total nicotinamide adenine dinucleotide pool were reduced, and the tricarboxylic acid cycle enzymes were altered. These changes lead to early mitochondrial dysfunction and an increase in mitochondrial oxidative stress, but a reduction in the levels of cytosolic oxidative damage and in the levels of systemic biomarkers of aging [145]. From those data, the authors suggested that although the mclk1/ clones have only been found in old mice, the anti-aging effects propitiated by low MCLK1 levels already starts at a young age. These data support the existence of a link between aging and mitochondrial energy metabolism, but they seem to be incompatible with the mitochondrial oxidative stress theory of aging. Studies in the long-lived C. elegans daf-2 mutants, in different Drosophila lines and in C. elegans under caloric restriction have also found a relationship between elevated mitochondrial ROS production in early stages of life and increase in life span [146,147,148]. This link could be explained by the fact that the early increase in the mitochondrial oxidative stress in the liver induces a reduction in oxidative stress in other cellular compartments, then reducing the systemic oxidative damage, which contributes to the increase in life span. The low levels of oxidized nicotinamide adenine dinucleotide (NAD+) would condition a reduced activity of the NADH/NADPH oxidases, the main generators of extra-mitochondrial ROS [140]. Thus, this hypothesis would be compatible with the concept of mitohormesis [33,115] (Figure 2a).
The CoQ9Q95X mouse model has a mutation in the CoQ9 gene that causes a premature termination in the CoQ9 protein. As a result, CoQ9Q95X mice have undetectable levels of CoQ9 protein and a reduction (approximately 50%) in the levels of CoQ9 and CoQ10 in kidneys, brain, skeletal muscle and heart, leading to a very mild late-onset mitochondrial myopathy, especially evident in females, and to an increase of 15% in life span [23,46,138]. Because the CoQ9 protein is required for the stability and activity of CoQ7, the lack of CoQ9 in CoQ9Q95X mice determines a severe reduction in the CoQ7 protein levels in brain, kidney, skeletal muscle, heart and liver. These alterations lead to an impairment in mitochondrial bioenergetic function that is characterized by a reduction in mitochondrial complex I+III activity and oxidative phosphorylation phosphorylating respiration in both, muscle and kidney [23,46,138]. Unlike mclk1+/ mice, CoQ9Q95X mice have neither a decrease in the CoQ levels in liver, DMQ accumulation nor mitochondrial dysfunction in this tissue. Therefore, there is no evidence that the increased life span in CoQ9Q95X mice is due to early mitochondrial dysfunction in liver, as proposed in mclk1+/ mice. However, the increased life span in CoQ9Q95X mice could be attributed to mitochondrial dysfunction in other tissues, such as brain, kidney or muscle, or to unknown mechanisms [138].

3. Dietary Supplementation of CoQ10 during Aging

3.1. Changes in CoQ Biosynthesis during Aging

The levels of CoQ are quite stable in cells, but its concentration varies among different tissues and organs. Also, the amount of CoQ varies depending on dietary conditions, as the cells can incorporate CoQ from dietary sources [149], and age. During the aging process, as well as in some particular aging-related diseases, a significant reduction in the rate of CoQ biosynthesis seems to occur. Beyer et al. [150] reported decreased CoQ levels in heart, kidney, gastrocnemius and oblique muscles in rats at 25 months of age, although CoQ levels in liver increased life span and other tissues like brain and lung had constant CoQ levels [150]. The decrease in CoQ levels with aging has also been reported in some human tissues, where the highest values have been found at 20 years of age [151]. Contrary to those data, Battino and colleagues [152] showed a direct correlation on CoQ and aging in three brain areas of rats [153]. The results showing a reduction in CoQ levels in some tissues during aging would support potential benefits of its exogenous supplementation during aging (Figure 2b).

3.2. CoQ10 Supplementation in Aging: Effects on Life span and Longevity

The possibility of increasing CoQ10 levels in different organs or tissues through dietary supplementation has been widely explored in recent decades. Studies in rodents [153,154,155,156,157] suggest that CoQ10 administration is able to increase the amounts of CoQ10 in plasma and liver significantly, and in heart, kidney and skeletal muscle moderately. Similarly, different authors have reported increased systemic levels of CoQ10 in humans after supplementing with CoQ10 at different daily doses (100 to 2400 mg) and duration (20 days, 3 or even 16 months) in multiple trials [158,159,160,161,162]. Regarding the safety of CoQ10 supplements, different assessments in human and animals (reviewed by Hidaka et al. [163]) 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. Based on the absence of adverse effects, the use of a daily dosage of 12 mg/kg of body weight per day in rats following a long-term feeding regimen has been suggested. In humans, a safety level of 1200 mg/day per person has been proposed, although doses of up to 3000 mg/day do not cause serious adverse effects and have been used in shorter clinical trials [164]. Despite no serious adverse effects being found, moderate adverse effects such as nausea and other adverse gastrointestinal effects have been reported. However, these effects were not causally related to the active ingredient because there was no dose–response relationship [163]. More recently, it has been reported that CoQ10 is generally safe and well-tolerated at a dose of 2400 mg/day in patients suffering from early-stage Huntington disease [160].
From a biochemical standpoint, CoQ10 benefits in relation to aging have been traditionally attributed to their antioxidant properties and to its role in MRC, which would influence mitochondrial functionality and ROS production. Supporting the protective role of CoQ10 against oxidative stress, some studies in animals have indicated that CoQ10 supplementation can reduce oxidative damage accumulation in certain tissues at least during some stages of the life [156,165]. In mice, 3 weeks of CoQ10 supplementation at a dose of 2.81 mg/g of diet was able to attenuate oxidative damage to proteins in liver in aged mice [165]. A similar effect on protein oxidative damage was found in skeletal muscle mitochondria of 14-week-old male rats after 13 weeks of CoQ10 supplementation. In the same study, a reductive shift was found in plasma aminothiol status [156]. This could imply an increase in the activity of antioxidant enzymes, higher levels of ROS scavengers or a decreased production of ROS. In humans, supplementation with high amounts of CoQ10 (1200 mg/day) to 65 patients undergoing hemodialysis has been shown to reduce plasma levels of the lipid peroxidative damage marker F2-isoprostane, which correlates with an increase in plasma CoQ10 concentrations [166]. On the contrary, in 55-year-old men, peroxidative markers and total thiols or total antioxidant capacity were also not modified by administrating CoQ10H2 in spite of the fact that the plasma levels of both total CoQ10 and CoQ10H2 were increased and a higher concentration was reached—notwithstanding that the treatment only had a duration of two weeks [162]. In addition, many of the mentioned studies on animals found effects only in some organs, cell types or organelles; despite some benefits against aging, detrimental effects on health or organ function were observed [156,165]. This might be explained by the different capacity of uptake of exogenous CoQ10 of the tissues [167].
In cases in which treatments with CoQ10 supplements have been able to reverse or delay some age-related changes, health improvement that would result in lower mortality rates and and extension of life span is expected. The potential of CoQ10 supplementation in increasing life span and longevity has been evaluated in different models (Table 1). In C. elegans, exogenous CoQ10 prolonged life span [168]. Also, the addition of CoQ10 to the diet has been shown to increase life span in rodents, at least under certain circumstances—notwithstanding that in most of the studies in both rats and mice, CoQ10 supplementation was ineffective for increasing life span [157,169,170]. Importantly, according to the variety of used CoQ10 doses (ranged from 10 to 370 mg/Kg per day) and the duration of the different studies, the lack of effects on longevity seems, in many of these studies, not to depend on these conditions. Instead, CoQ10 supplementation would be effective in increasing median life span when it is combined with certain nutritional conditions associated with elevated oxidative stress and age-related detrimental effects. From this standpoint, a study on rats comparing CoQ10 effects between isocaloric diets with different lipid profiles by using virgin olive, sunflower oil or fish oil as a dietary fat source is particularly interesting [171]. In this study, supplementation of the fat with a low-dose of CoQ10 from weaning was able to improve survival in rats receiving a diet with sunflower oil, increasing median life span values. However, no effects were observed in those fed on diets based on virgin olive or fish oil. These effects were observed when sunflower oil was administered in a proportion of 8% w/w in the diet, which is the double of current recommendations for rodents [172]. This is in concordance with other findings in different tissues and organs of old and young rats fed on similar diets [173,174,175,176,177,178,179,180,181,182,183,184]. Alternatively, it has been suggested that CoQ10 supplementation does not truly extend life span, but that it could prevent life span shortening due to oxidative insults [161], as has been suggested by its effect in all aspects related to mitochondrial function, oxidative stress and antioxidant defenses both in animals and humans (Figure 2b). On the other hand, the different oils used could modulate CoQ10 absorption since fatty acids significantly help in the absorption of CoQ10 via bile acid production [185,186]. In animals fed on similar CoQ-supplemented diets, plasma CoQ10 concentration values were 39.24 ± 9.25 and 50.71 ± 5.22 μM when they received a diet based on virgin olive oil and sunflower oil, respectively [174]. Another study reported even lower values in those animals fed on a fish oil-based diet [182]. This seems in concordance with the proposed hypothesis, but the values found in animals receiving different oils were not statistically compared in these studies.

3.3. Reversal of Age-Related Changes by CoQ10 Supplementation

The modification of CoQ10 levels in different tissues as a consequence of increasing dietary CoQ10 intake would be on the basis of the improvement in conditions related to aging observed in both humans and animals. These include obesity and metabolic disorders [158,189,190,191,192], cardiovascular diseases [193,194] or skin aging by exposure to sun radiation. In this sense, dietary CoQ10 has been able to reduce insulin resistance (IR) in adults with prediabetes [158], as well as overweight and obese patients with coronary heart disease and type 2 diabetes mellitus (DM2) [190]—notwithstanding that data confirming the effect of CoQ10 treatment on blood levels were only provided in the first case. In the last group, the treatment also reduced the serum insulin levels and β-cell function [190], although it did not affect insulin levels in patients with prediabetes [158]. Likewise, it has been reported that the consumption of a Mediterranean diet for 4 weeks led to improvement in parameters related to the mentioned metabolic alterations in old persons, including a reduction in postprandial levels of advanced glycation end products (AGEs) and an increase in AGE receptor-1 and glyoxalase I gene expression. Such effects were accentuated by CoQ10 supplementation, which correlated with increased fasting and postprandial CoQ levels in plasma [195]. Also, CoQ10 intake during 2 months led to a significant reduction in serum protein oxidative damage markers in type 2 diabetic patients with coronary heart disease, although it did not affect the systemic levels of lipid peroxidative damage markers or thiol concentrations [196]. Still, there were no available data on CoQ levels after the treatment. The therapeutic effects of CoQ10 supplementation against metabolic disorders have been confirmed by a meta-analysis on 14 randomized-controlled trials (RCTs) examining CoQ10 effects on fasting blood glucose, fasting insulin and HbA1c. This study indicated that CoQ10 supplementation slightly but significantly reduced fasting blood glucose, although it does not support the existence of significant effects on fasting insulin and glycosylated hemoglobin [189]. Moreover, results from animal studies suggest that dietary CoQ10 would prevent or mitigate diabetes complications [197,198]. In relation to other metabolic alterations, supplementation with CoQ10 has also shown beneficial effects on the treatment of hypercholesterolemia and hypertriglyceridemia by modifying blood lipid concentration. In atherosclerosis-prone apolipoprotein E (ApoE) knockout (Apoe/) mice, which is a well-established mouse model for the study of human atherosclerosis, CoQ10 had an anti-atherogenic effect [193]. Supplementation with the reduced form of CoQ10 (CoQ10H2) modulated adipocyte differentiation in female KKAy mice, a model of obesity and DM2, reducing white adipose tissue content and improving brown adipose tissue function. Moreover, the expression of lipid metabolism-related factors was modified, indicating that it regulated lipid metabolism. Among other effects, the decomposition of lipids was enhanced, and the de novo synthesis of fatty acids was inhibited. Thus, the development and progression of obesity could be stopped or at least slowed down [192]. Mechanistically, it has been reported that a seven-day treatment with CoQ10H2 (250 mg/kg of body weight per day) was able to induce changes in the hepatic expression of lipid metabolism genes functionally connected by the peroxisome proliferator-activated receptor (PPAR)α signaling pathway in C57BL6J mice [199]. However, a meta-analysis of RCTs evaluating the effect of treatment with CoQ10 on DM2 patients did not find clear evidence of the capability of CoQ10 to alter low-density lipoprotein cholesterol (LDL-C), high-density lipoprotein cholesterol (HDL-C) and blood pressure, although it reduced triacylglyceride (TAG) levels. Moreover, evidence of improved glycemic control by this molecule was not enough [191]. Important conclusions arise from a systematic review compiling RCTs in healthy adults or those at high risk of cardiovascular disease investigating the effects of CoQ10 alone as a single supplement in the absence of lifestyle intervention. This systemic review does not support the efficacy of CoQ10 supplements in reducing systolic blood pressure, total cholesterol, LDL- C, or HDL-C [200]. Some of the reported effects on systemic markers and metabolic alterations might contribute to cardiovascular disease prevention. In this sense, dietary CoQ10 combined with a Mediterranean diet in elderly patients led to increased plasma levels and improved several markers of endothelial function that is also a known risk factor for important cardiovascular diseases [201]. Additional cardiovascular benefits have been found in patients suffering from heart failure, where a short-term (12 weeks or less) daily treatment with oral CoQ10 (100 mg per person) improved left ventricular ejection fraction [202]. In contrast, no effect was observed on exercise capacity [203].
CoQ10 therapy has also been tested in non-metabolic or cardiovascular diseases. A promising protective effect of ubiquinol (600 mg/kg of body weight per day of ubiquinol for 4 weeks) for kidneys in three-week-old heminephrectomized male Sprague–Dawley rats fed on high salt (8%) diets has been reported [204]. However, a meta-analysis revealed that CoQ10 did not have reliable effects against the initiation and progression of diabetic kidney disease [205]. Lastly, the beneficial effects of CoQ10 on skin in middle-aged healthy women with Fitzpatrick skin phototypes II and III, where dietary treatment with a water-soluble form of CoQ10 limited the seasonal deterioration of viscoelasticity and reduced some visible signs of aging, has been reported [206]. In patients with Alzheimer disease, no changes in F2 isoprostanes—neither in biomarkers of amyloid or tau pathology in the cerebrospinal fluid—has been reported after supplementation with 400 mg of CoQ10, three times a day for 16 weeks [207]. However, systemic or cerebral CoQ levels were not assessed, so this lack of effects might be related to the previously reported low uptake of CoQ10 by the brain [167]. There are also studies supporting the idea that CoQ10 exerts anti-apoptotic and anti-inflammatories activities, which could be a consequence of redox-dependent mechanisms. A three-month treatment with CoQ10 supplements reduced calcitonin gene-related peptide and tumor necrosis factor (TNF)-α in middle-aged women, although there were no significant differences in serum IL-6 and IL-10 [159]. However, CoQ10 intake during two months led to a significant increase in CoQ10 plasma levels and a reduction in IL-6 in DM2 patients with coronary heart disease, although no data on CoQ10 levels were provided [196]. A meta-analysis of studies in patients with cardio cerebral vascular disease, multiple sclerosis, obesity, renal failure, rheumatoid arthritis, diabetes, and fatty liver disease supports the efficacy of CoQ10 in doses ranging from 60 to 500 mg/day, as they showed a decrease in plasma levels of CRP, IL-6 and TNF-α [208], which seems to confirm the anti-inflammatory potential of CoQ10 supplements, at least under certain conditions. Likewise, an earlier meta-analysis found that CoQ10 in doses ranging from 60 to 300 mg/day significantly reduced the levels of IL-6, although not of CRP [209].
Chronic inflammation and oxidative stress play essential roles in the pathogenesis of many age-related diseases in which low CoQ levels may be a pathophysiological factor. Thus, the reported health benefits of CoQ10 in previous studies could be a consequence of homeostasis recovery by directly increasing CoQ10 or indirectly improving cell antioxidant defenses and/or mitochondrial function. In addition, CoQ10 might be useful for good management of the aging process or by preventing age-related alterations that finally lead to disease and death. In this sense, many researchers have investigated the possible effects of long-term interventions with dietary CoQ10 on the senescence process itself at different levels (Table 2). Long-life dietary supplementation with CoQ10H2 slows aging in the senescence-accelerated mouse prone 1 (SAMP1) model in different studies [210,211]. The mechanism of action under this effect has been associated with the deceleration of the normal decline in the expression of Sirt1, Sirt3, Pgc-1a, and Ppara genes [210]. In contrast, oxidized CoQ10 had no effect on senescence in the same model [211]. On the other hand, Sohal et al. [157] did not find effects on the activity of different antioxidant enzymes in the liver, heart, kidney, skeletal muscle, and brain of mice. In male Wistar rats, long-term supplementation with low doses of CoQ10 has been shown to attenuate age-related alveolar bone loss [173,212], prevent age-related decline in BMD [174] and reduce histological alterations in endocrine pancreas, which mainly affected β-cell mass and insulin levels in aged animals [175]. In parallel, an aging-associated increase in urinary F2-isoprostanes was prevented by the addition of CoQ10, which suggests that CoQ benefits could be a consequence of a reduction in oxidative stress [174]. In gums, an age-related increase in the expression of genes involved in mitochondrial biogenesis and antioxidantdefense has also been found [173]. Such changes could also occur in other tissues, contributing to reducing oxidative stress. However, many of these effects were present only when animals were fed on diets using sunflower oil as a unique dietary fat source. The protective effects of CoQ10 have also been observed when this molecule was added to similar diets but with a high-fat content (8% w/w) [176,177,178,179,180,181,213]. In contrast, the effects of CoQ10 on animals receiving diets with other dietary fats as unique fat sources were not so clear. In animals fed a fish oil-based diet, CoQ10 has shown a clear benefit on bone health. Namely, animals supplemented with CoQ10 showed a higher value of BMD than their younger counterparts, which correlates with lower levels of urinary F2-isoprostanes and DNA strand breaks [182]. However, this molecule had no effect on periodontal tissues [173] and pancreas alterations found in aged animals that mainly affected the exocrine gland [175]. The lack of effect exerted by CoQ10 was accentuated when it was added to a diet using virgin olive oil since there were practically no changes in markers related to aging in different tissues and organs or improvements in health [173,174].
Overall, studies comparing the effects of the different diets with different fatty acid profiles or the prevalence of a particular dietary fat on aging or certain age-related diseases support that n-6 polyunsaturated fatty acids (PUFA) would be detrimental for health. This negative effect has usually been related to oxidative stress situations or pro-inflammatory environments [214,215]. Thus, it seems that CoQ10 is useful to counteract the consequences of unhealthy diets that “accelerate” the aging process, but it has no additional effects under more favorable conditions. In this sense, positive effects against health detriments and aging have been reported for short-term supplementation with CoQ10 in animals fed on different high-fat diets [216,217,218]. CoQ10 prevented methemoglobin formation and CoQ oxidation induced by feeding young Wistar rats on a high-fat diet for four weeks [216]. CoQ10 decreased inflammation and metabolic stress markers in mice fed on a high-fat diet (72%) consisting of corn oil and lard and fructose during 8 weeks [217]. However, the treatment failed in decreasing obesity and tissue lipid peroxides [217]. Likewise, in another study on rats with hyperlipidemia, with insulin resistance and non-alcoholic fatty liver disease (NAFLD) induced by a 4-week high-fat diet (57% of energy), CoQ did not ameliorate the effects of the diet [219]. Similarly, in rats with hyperlipidemia, insulin resistance and NAFLD induced by a 10-week high-fat diet (57% energy from fat), plasma levels of insulin, alanine aminotransferase and HOMA-IR, TG, VLDL and LDL increased and liver lipid accumulation and TG levels were not ameliorated or prevented [219,220]—notwithstanding that apolipoprotein B mRNA and microsomal TG levels were increased, and the phospholipid content of microsomal membranes was altered. In the same sense, CoQ10 effects seem to depend on the presence of risk factors or conditions in many cases related to inadequate diets. In that sense, CoQ10 reduced protein carbonyl levels and IL6 serum levels [196] and increased plasma glutathione in diabetic patients with coronary heart disease [190]. Other studies in diabetic patients presenting nephropathy suggested that these effects would be mediated, at least in part, by gene expression changes, since CoQ10 administration upregulated the gene expression of PPAR-γ and downregulated the gene expression of IL-1 and TNF-α in PBMCs [221].
Finally, there are also some studies supporting the possible use of CoQ10 against the decline in reproductive success with increasing age (i.e., reproductive senescence) and particularly against fertility decline. Female reproductive capacity declines with age as a consequence of an age-related decrease in oocyte quality and quantity, which is accompanied by mitochondrial dysfunction associated with decreased oxidative phosphorylation and reduced ATP level [222,223,224]. Moreover, there is an increase in meiotic spindle abnormalities favoring oocyte aneuploidy that leads to a reduced embryo quality, as well as an increased incidence of miscarriages and birth defects [225]. That the impaired mitochondrial performance associated with reproductive senescence could be a consequence of suboptimal CoQ10 availability, which will lead to age-associated oocyte deficits, has been proposed [223]. Actually, a diminished expression of genes encoding two enzymes responsible for CoQ production, Pdss2 and CoQ6, in oocytes of older females in both mouse and human has been reported [223]. Likewise, oocyte-specific Pdss2-deficient animals have also shown a diminished ovarian reserve, a condition prevented by maternal dietary administration of CoQ10 [223]. A similar intervention also reversed age-related decline in oocyte quality and quantity in an aging mouse model [223]. In addition, cumulus granulosa cells also display a decline in cell function linked to mitochondrial activity accompanied by a decreased expression of Pdss2 and CoQ6 genes in both human and mouse models [225]. Concerning these cells, supplementation with CoQ10 restored cumulus cell number, stimulated glucose uptake, and increased progesterone production in an aged mouse model. These preclinical studies suggest that CoQ10 might improve fertility in females of advanced maternal age by improving mitochondrial metabolism in oocyte and cumulus cells, which would result in increased quantity and quality [225]. However, there are no clinical studies in aged women confirming this. On the other hand, male infertility has also been associated with oxidative stress and healthy sperm has been correlated with CoQ10 levels in seminal fluid [226]. In addition, a meta-analysis of 15 randomized clinical trials suggests that some dietary CoQ10 could beneficially modulate sperm quality parameters, improving male fertility. Namely, these interventions have been shown to increase total sperm concentrations and sperm count, and enhanced sperm total motility was enhanced by CoQ10 [227]. Therefore, dietary CoQ10 could also be useful in reducing male fertility decline. However, the limited sample size of the studies included in the meta-analysis and the inter-study heterogeneity are important factors for carefully interpreting these results [227].

4. Conclusions

Some defects in CoQ biosynthesis induce an increase in life span in animal models, most likely due to either a mild reduction in the flow of electrons in the mitochondrial respiratory chain and the subsequent decrease in ROS production or early mitochondrial dysfunction and physiological adaptation to this condition—the latter phenomenon being compatible with the concept of mitohormesis. Moreover, CoQ10 supplementation has also shown therapeutic benefits in animal models of disease and human studies, especially in conditions associated with oxidative stress. Those benefits are dependent on CoQ10 bioavailability and tissue uptake and, consequently, liver, adipose tissue and circulating cells have a good response to CoQ10 supplementation. Therefore, both CoQ biosynthesis defects and CoQ10 supplementation are therapeutically relevant depending on the moment and the context of the intervention. However, further studies are required to understand the mechanisms of CoQ10 therapy in aging-related diseases, especially in the context of the multiple biological functions of CoQ—an essential molecule for mammalian cells.

Author Contributions

M.E.D.-C. and A.V.-L. wrote and edited the manuscript; J.L.Q. reviewed and edited the manuscript; E.B.-C. reviewed the manuscript; P.G.-G. reviewed the manuscript; M.B. reviewed the manuscript; L.C.L. reviewed and edited the manuscript.

Funding

This work was supported by grants from Ministerio de Ciencia, Innovación y Universidades, Spain, and the ERDF (Grant Number RTI2018-093503-B-I00), from the Muscular Dystrophy Association (grant number 602322) and from the University of Granada (grant reference “UNETE”, UCE-PP2017-06). E.B.-C. is supported by the Junta de Andalucía. P.G.-G. is a “FPU fellow” from the Ministerio de Educación Cultura y Deporte, Spain.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Kalén, A.; Appelkvist, E.L.; Chojnacki, T.; Dallner, G. Nonaprenyl-4-hydroxybenzoate transferase, an enzyme involved in ubiquinone biosynthesis, in the endoplasmic reticulum-Golgi system of rat liver. J. Biol. Chem. 1990, 265, 1158–1164. [Google Scholar] [PubMed]
  2. Kalén, A.; Norling, B.; Appelkvist, E.; Dallner, G. Ubiquinone biosynthesis by the microsomal fraction from rat liver. Biochim. Biophys. Acta 1987, 926, 70–78. [Google Scholar] [CrossRef]
  3. Eisenberg-Bord, M.; Tsui, H.S.; Antunes, D.; Fernandez-Del-Rio, L.; Bradley, M.C.; Dunn, C.D.; Nguyen, T.P.T.; Rapaport, D.; Clarke, C.F.; Schuldiner, M. The Endoplasmic Reticulum-Mitochondria Encounter Structure Complex Coordinates Coenzyme Q Biosynthesis. Contact 2019, 2. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  4. Turunen, M.; Olsson, J.; Dallner, G. Metabolism and function of coenzyme Q. Biochim. Biophys. Acta 2004, 1660, 171–199. [Google Scholar] [CrossRef] [PubMed]
  5. Tzagoloff, A.; Akai, A.; Needleman, R.B.; Zulch, G. Assembly of the mitochondrial membrane system. Cytoplasmic mutants of Saccharomyces cerevisiae with lesions in enzymes of the respiratory chain and in the mitochondrial ATPase. J. Biol. Chem. 1975, 250, 8236–8242. [Google Scholar] [PubMed]
  6. Tzagoloff, A.; Dieckmann, C.L. PET genes of Saccharomyces cerevisiae. Microbiol. Rev. 1990, 54, 211–225. [Google Scholar] [PubMed]
  7. Meganathan, R. Ubiquinone biosynthesis in microorganisms. FEMS Microbiol. Lett. 2001, 203, 131–139. [Google Scholar] [CrossRef] [PubMed]
  8. Hajj Chehade, M.; Loiseau, L.; Lombard, M.; Pecqueur, L.; Ismail, A.; Smadja, M.; Golinelli-Pimpaneau, B.; Mellot-Draznieks, C.; Hamelin, O.; Aussel, L.; et al. ubiI, a new gene in Escherichia coli coenzyme Q biosynthesis, is involved in aerobic C5-hydroxylation. J. Biol. Chem. 2013, 288, 20085–20092. [Google Scholar] [CrossRef]
  9. Aussel, L.; Loiseau, L.; Hajj Chehade, M.; Pocachard, B.; Fontecave, M.; Pierrel, F.; Barras, F. ubiJ, a new gene required for aerobic growth and proliferation in macrophage, is involved in coenzyme Q biosynthesis in Escherichia coli and Salmonella enterica serovar Typhimurium. J. Bacteriol. 2014, 196, 70–79. [Google Scholar] [CrossRef]
  10. Awad, A.M.; Bradley, M.C.; Fernández-Del-Río, L.; Nag, A.; Tsui, H.S.; Clarke, C.F. Coenzyme Q10 deficiencies: Pathways in yeast and humans. Essays Biochem. 2018, 62, 361–376. [Google Scholar] [CrossRef]
  11. Saiki, R.; Nagata, A.; Kainou, T.; Matsuda, H.; Kawamukai, M. Characterization of solanesyl and decaprenyl diphosphate synthases in mice and humans. FEBS J. 2005, 272, 5606–5622. [Google Scholar] [CrossRef] [PubMed]
  12. Hayashi, K.; Ogiyama, Y.; Yokomi, K.; Nakagawa, T.; Kaino, T.; Kawamukai, M. Functional Conservation of Coenzyme Q Biosynthetic Genes among Yeasts, Plants, and Humans. PLoS ONE 2014, 9, e99038. [Google Scholar] [CrossRef] [PubMed]
  13. Kawamukai, M. Biosynthesis of coenzyme Q in eukaryotes. Biosci. Biotechnol. Biochem. 2016, 80, 23–33. [Google Scholar] [CrossRef] [PubMed]
  14. Marbois, B.; Xie, L.X.; Choi, S.; Hirano, K.; Hyman, K.; Clarke, C.F. para-Aminobenzoic Acid Is a Precursor in Coenzyme Q6 Biosynthesis in Saccharomyces cerevisiae. J. Biol. Chem. 2010, 285, 27827–27838. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  15. Payet, L.-A.; Leroux, M.; Willison, J.C.; Kihara, A.; Pelosi, L.; Pierrel, F. Mechanistic Details of Early Steps in Coenzyme Q Biosynthesis Pathway in Yeast. Cell Chem. Biol. 2016, 23, 1241–1250. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  16. Stefely, J.A.; Kwiecien, N.W.; Freiberger, E.C.; Richards, A.L.; Jochem, A.; Rush, M.J.P.; Ulbrich, A.; Robinson, K.P.; Hutchins, P.D.; Veling, M.T.; et al. Mitochondrial protein functions elucidated by multi-omic mass spectrometry profiling. Nat. Biotechnol. 2016, 34, 1191–1197. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  17. Stefely, J.A.; Pagliarini, D.J. Biochemistry of Mitochondrial Coenzyme Q Biosynthesis. Trends Biochem. Sci. 2017, 42, 824–843. [Google Scholar] [CrossRef]
  18. Doimo, M.; Trevisson, E.; Airik, R.; Bergdoll, M.; Santos-Ocaña, C.; Hildebrandt, F.; Navas, P.; Pierrel, F.; Salviati, L. Effect of vanillic acid on COQ6 mutants identified in patients with coenzyme Q10 deficiency. Biochim. Biophys. Acta 2014, 1842, 1–6. [Google Scholar] [CrossRef] [PubMed]
  19. Hidalgo-Gutierrez, A.; Barriocanal-Casado, E.; Bakkali, M.; Diaz-Casado, M.E.; Sanchez-Maldonado, L.; Romero, M.; Sayed, R.K.; Prehn, C.; Escames, G.; Duarte, J.; et al. beta-RA reduces DMQ/CoQ ratio and rescues the encephalopathic phenotype in Coq9 (R239X) mice. EMBO Mol. Med. 2019, 11, e9466. [Google Scholar] [CrossRef]
  20. Pierrel, F. Impact of Chemical Analogs of 4-Hydroxybenzoic Acid on Coenzyme Q Biosynthesis: From Inhibition to Bypass of Coenzyme Q Deficiency. Front. Physiol. 2017, 8, 436. [Google Scholar] [CrossRef]
  21. Wang, Y.; Oxer, D.; Hekimi, S. Mitochondrial function and lifespan of mice with controlled ubiquinone biosynthesis. Nat. Commun. 2015, 6, 6393. [Google Scholar] [CrossRef] [PubMed]
  22. Wang, Y.; Smith, C.; Parboosingh, J.S.; Khan, A.; Innes, M.; Hekimi, S. Pathogenicity of two COQ7 mutations and responses to 2,4-dihydroxybenzoate bypass treatment. J. Cell. Mol. Med. 2017, 21, 2329–2343. [Google Scholar] [CrossRef] [PubMed]
  23. Luna-Sánchez, M.; Díaz-Casado, E.; Barca, E.; Ángel, T.M.; Ángeles, M.-G.; Cobos, E.J.; Escames, G.; Acuña-Castroviejo, D.; Quinzii, C.M.; López, L.C. The clinical heterogeneity of coenzyme Q10 deficiency results from genotypic differences in the Coq9 gene. EMBO Mol. Med. 2015, 7, 670–687. [Google Scholar]
  24. Xie, L.X.; Williams, K.J.; He, C.H.; Weng, E.; Khong, S.; Rose, T.E.; Kwon, O.; Bensinger, S.J.; Marbois, B.N.; Clarke, C.F. Resveratrol and para-coumarate serve as ring precursors for coenzyme Q biosynthesis[S]. J. Lipid Res. 2015, 56, 909–919. [Google Scholar] [CrossRef] [PubMed]
  25. Fernandez-Del-Rio, L.; Nag, A.; Gutierrez Casado, E.; Ariza, J.; Awad, A.M.; Joseph, A.I.; Kwon, O.; Verdin, E.; de Cabo, R.; Schneider, C.; et al. Kaempferol increases levels of coenzyme Q in kidney cells and serves as a biosynthetic ring precursor. Free Radic. Biol. Med. 2017, 110, 176–187. [Google Scholar] [CrossRef] [PubMed]
  26. Gin, P.; Hsu, A.Y.; Rothman, S.C.; Jonassen, T.; Tzagoloff, A.; Clarke, C.F.; Lee, P.T. TheSaccharomyces cerevisiae COQ6Gene Encodes a Mitochondrial Flavin-dependent Monooxygenase Required for Coenzyme Q Biosynthesis. J. Biol. Chem. 2003, 278, 25308–25316. [Google Scholar] [CrossRef] [PubMed]
  27. Ozeir, M.; Mühlenhoff, U.; Webert, H.; Lill, R.; Fontecave, M.; Pierrel, F. Coenzyme Q Biosynthesis: Coq6 Is Required for the C5-Hydroxylation Reaction and Substrate Analogs Rescue Coq6 Deficiency. Chem. Biol. 2011, 18, 1134–1142. [Google Scholar] [CrossRef] [PubMed]
  28. Marbois, B.; Gin, P.; Faull, K.F.; Poon, W.W.; Strahan, J.; Clarke, C.F.; Lee, P.T.; Shepherd, J.N. Coq3 and Coq4 Define a Polypeptide Complex in Yeast Mitochondria for the Biosynthesis of Coenzyme Q. J. Biol. Chem. 2005, 280, 20231–20238. [Google Scholar] [CrossRef] [Green Version]
  29. Zhu, Y.; Wu, B.; Zhang, X.; Fan, X.; Niu, L.; Li, X.; Wang, J.; Teng, M. Structural and biochemical studies reveal UbiG/Coq3 as a class of novel membrane-binding proteins. Biochem. J. 2015, 470, 105–114. [Google Scholar] [CrossRef]
  30. Cox, G.B.; Young, I.G.; McCann, L.M.; Gibson, F. Biosynthesis of Ubiquinone in Escherichia coli K-12: Location of Genes Affecting the Metabolism of 3-Octaprenyl-4-hydroxybenzoic Acid and 2-Octaprenylphenol. J. Bacteriol. 1969, 99, 450–458. [Google Scholar] [Green Version]
  31. Young, I.G.; Stroobant, P.; Macdonald, C.G.; Gibson, F. Pathway for Ubiquinone Biosynthesis in Escherichia coli K-12: Gene-Enzyme Relationships and Intermediates. J. Bacteriol. 1973, 114, 42–52. [Google Scholar] [PubMed]
  32. White, M.D.; Payne, K.A.P.; Fisher, K.; Marshall, S.A.; Parker, D.; Rattray, N.J.W.; Trivedi, D.K.; Goodacre, R.; Rigby, S.E.J.; Scrutton, N.S.; et al. UbiX is a flavin prenyltransferase required for bacterial ubiquinone biosynthesis. Nature 2015, 522, 502–506. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  33. Dai, Y.N.; Zhou, K.; Cao, D.D.; Jiang, Y.L.; Meng, F.; Chi, C.B.; Ren, Y.M.; Chen, Y.; Zhou, C.Z. Crystal structures and catalytic mechanism of the C-methyltransferase Coq5 provide insights into a key step of the yeast coenzyme Q synthesis pathway. Acta Crystallogr. D Biol. Crystallogr. 2014, 70, 2085–2092. [Google Scholar] [CrossRef] [PubMed]
  34. Nguyen, T.P.; Casarin, A.; Desbats, M.A.; Doimo, M.; Trevisson, E.; Santos-Ocaña, C.; Navas, P.; Clarke, C.F.; Salviati, L. Molecular characterization of the human COQ5 C-methyltransferase in coenzyme Q10 biosynthesis. Biochim. Biophys. Acta 2014, 1841, 1628–1638. [Google Scholar] [CrossRef] [PubMed]
  35. Stenmark, P.; Grunler, J.; Mattsson, J.; Sindelar, P.J.; Nordlund, P.; Berthold, D.A. A new member of the family of di-iron carboxylate proteins. Coq7 (clk-1), a membrane-bound hydroxylase involved in ubiquinone biosynthesis. J. Biol. Chem. 2001, 276, 33297–33300. [Google Scholar] [CrossRef] [PubMed]
  36. Gerards, M.; Bosch, B.V.D.; Calis, C.; Schoonderwoerd, K.; Van Engelen, K.; Tijssen, M.; De Coo, R.; Van Der Kooi, A.; Smeets, H. Nonsense mutations in CABC1/ADCK3 cause progressive cerebellar ataxia and atrophy. Mitochondrion 2010, 10, 510–515. [Google Scholar] [CrossRef] [PubMed]
  37. Horváth, P.O.; Czermin, B.; Gulati, S.; Pyle, A.; Hassani, A.; Foley, C.; Taylor, R.W.; Chinnery, P.F. 003 Adult-onset cerebellar ataxia due to mutations in the CABC1/ADCK3 gene. J. Neurol. Neurosurg. Psychiatry 2012, 83. [Google Scholar] [CrossRef]
  38. 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 Q10 Deficiency. Am. J. Hum. Genet. 2008, 82, 661–672. [Google Scholar] [CrossRef]
  39. Vazquez Fonseca, L.; Doimo, M.; Calderan, C.; Desbats, M.A.; Acosta, M.J.; Cerqua, C.; Cassina, M.; Ashraf, S.; Hildebrandt, F.; Sartori, G.; et al. Mutations in COQ8B (ADCK4) found in patients with steroid-resistant nephrotic syndrome alter COQ8B function. Hum. Mutat. 2018, 39, 406–414. [Google Scholar] [CrossRef] [PubMed]
  40. Xie, L.X.; Ozeir, M.; Tang, J.Y.; Chen, J.Y.; Jaquinod, S.-K.; Fontecave, M.; Clarke, C.F.; Pierrel, F. Overexpression of the Coq8 Kinase in Saccharomyces cerevisiae coq Null Mutants Allows for Accumulation of Diagnostic Intermediates of the Coenzyme Q6 Biosynthetic Pathway. J. Biol. Chem. 2012, 287, 23571–23581. [Google Scholar] [CrossRef] [Green Version]
  41. Reidenbach, A.G.; Kemmerer, Z.A.; Aydin, D.; Jochem, A.; McDevitt, M.T.; Hutchins, P.D.; Stark, J.L.; Stefely, J.A.; Reddy, T.; Hebert, A.S.; et al. Conserved Lipid and Small-Molecule Modulation of COQ8 Reveals Regulation of the Ancient Kinase-like UbiB Family. Cell Chem. Biol. 2018, 25, 154–165 e111. [Google Scholar] [CrossRef] [PubMed]
  42. Hsieh, E.J.; Gin, P.; Gulmezian, M.; Tran, U.C.; Saiki, R.; Marbois, B.N.; Clarke, C.F. Saccharomyces cerevisiae Coq9 Polypeptide is a Subunit of the Mitochondrial Coenzyme Q Biosynthetic Complex. Arch. Biochem. Biophys. 2007, 463, 19–26. [Google Scholar] [CrossRef] [PubMed]
  43. Lohman, D.C.; Aydin, D.; Von Bank, H.C.; Smith, R.W.; Linke, V.; Weisenhorn, E.; McDevitt, M.T.; Hutchins, P.; Wilkerson, E.M.; Wancewicz, B.; et al. An Isoprene Lipid-Binding Protein Promotes Eukaryotic Coenzyme Q Biosynthesis. Mol. Cell 2019, 73, 763–774.e10. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  44. He, C.W.; Black, D.S.; Nguyen, T.P.T.; Wang, C.; Srinivasan, C.; Clarke, C.F. Yeast Coq9 controls deamination of coenzyme Q intermediates that derive from para-aminobenzoic acid. Biochim. Biophys. Acta 2015, 1851, 1227–1239. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  45. Lohman, D.C.; Forouhar, F.; Beebe, E.T.; Stefely, M.S.; Minogue, C.E.; Ulbrich, A.; Stefely, J.A.; Sukumar, S.; Luna-Sánchez, M.; Jochem, A.; et al. Mitochondrial COQ9 is a lipid-binding protein that associates with COQ7 to enable coenzyme Q biosynthesis. Proc. Natl. Acad. Sci. USA 2014, 111, E4697–E4705. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  46. Garcia-Corzo, L.; Luna-Sanchez, M.; Doerrier, C.; Garcia, J.A.; Guaras, A.; Acin-Perez, R.; Bullejos-Peregrin, J.; Lopez, A.; Escames, G.; Enriquez, J.A.; et al. Dysfunctional Coq9 protein causes predominant encephalomyopathy associated with CoQ deficiency. Hum. Mol. Genet. 2013, 22, 1233–1248. [Google Scholar] [CrossRef] [PubMed]
  47. Wang, Y.; Hekimi, S. Molecular genetics of ubiquinone biosynthesis in animals. Crit. Rev. Biochem. Mol. Biol. 2013, 48, 69–88. [Google Scholar] [CrossRef]
  48. Allan, C.M.; Hill, S.; Morvaridi, S.; Saiki, R.; Johnson, J.S.; Liau, W.S.; Hirano, K.; Kawashima, T.; Ji, Z.; Loo, J.A.; et al. A conserved START domain coenzyme Q-binding polypeptide is required for efficient Q biosynthesis, respiratory electron transport, and antioxidant function in Saccharomyces cerevisiae. Biochim. Biophys. Acta 2013, 1831, 776–791. [Google Scholar] [CrossRef]
  49. Cui, T.Z.; Kawamukai, M. Coq10, a mitochondrial coenzyme Q binding protein, is required for proper respiration in Schizosaccharomyces pombe. FEBS J. 2009, 276, 748–759. [Google Scholar] [CrossRef]
  50. Tsui, H.S.; Pham, N.V.B.; Amer, B.R.; Bradley, M.C.; Gosschalk, J.E.; Gallagher-Jones, M.; Ibarra, H.; Clubb, R.T.; Blaby-Haas, C.E.; Clarke, C.F. Human COQ10A and COQ10B are distinct lipid-binding START domain proteins required for coenzyme Q function. J. Lipid Res. 2019, 60, 1293–1310. [Google Scholar] [CrossRef]
  51. Marbois, B.; Gin, P.; Gulmezian, M.; Clarke, C.F. The yeast Coq4 polypeptide organizes a mitochondrial protein complex essential for coenzyme Q biosynthesis. Biochim. Biophys. Acta 2009, 1791, 69–75. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  52. Belogrudov, G.I.; Lee, P.T.; Jonassen, T.; Hsu, A.Y.; Gin, P.; Clarke, C.F. Yeast COQ4 Encodes a Mitochondrial Protein Required for Coenzyme Q Synthesis. Arch. Biochem. Biophys. 2001, 392, 48–58. [Google Scholar] [CrossRef] [PubMed]
  53. Casarin, A.; Jimenez-Ortega, J.C.; Trevisson, E.; Pertegato, V.; Doimo, M.; Ferrero-Gomez, M.L.; Abbadi, S.; Artuch, R.; Quinzii, C.; Hirano, M.; et al. Functional characterization of human COQ4, a gene required for Coenzyme Q10 biosynthesis. Biochem. Biophys. Res. Commun. 2008, 372, 35–39. [Google Scholar] [CrossRef] [PubMed]
  54. Allan, C.M.; Awad, A.M.; Johnson, J.S.; Shirasaki, D.I.; Wang, C.; Blaby-Haas, C.E.; Merchant, S.S.; Loo, J.A.; Clarke, C.F. Identification of Coq11, a New Coenzyme Q Biosynthetic Protein in the CoQ-Synthome in Saccharomyces cerevisiae. J. Biol. Chem. 2015, 290, 7517–7534. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  55. He, C.H.; Xie, L.X.; Allan, C.M.; Tran, U.C.; Clarke, C.F. Coenzyme Q supplementation or over-expression of the yeast Coq8 putative kinase stabilizes multi-subunit Coq polypeptide complexes in yeast coq null mutants. Biochim. Biophys. Acta 2014, 1841, 630–644. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  56. Stefely, J.A.; Licitra, F.; Laredj, L.; Reidenbach, A.G.; Kemmerer, Z.A.; Grangeray, A.; Jaeg-Ehret, T.; Minogue, C.E.; Ulbrich, A.; Hutchins, P.D.; et al. Cerebellar Ataxia and Coenzyme Q Deficiency through Loss of Unorthodox Kinase Activity. Mol. Cell 2016, 63, 608–620. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  57. Crane, F.; Hatefi, Y.; Lester, R.; Widmer, C. Isolation of a quinone from beef heart mitochondria. Biochim. Biophys. Acta 1957, 25, 220–221. [Google Scholar] [CrossRef]
  58. Quinzii, C.M.; Luna-Sanchez, M.; Ziosi, M.; Hidalgo-Gutierrez, A.; Kleiner, G.; Lopez, L.C. The Role of Sulfide Oxidation Impairment in the Pathogenesis of Primary CoQ Deficiency. Front. Physiol. 2017, 8, 525. [Google Scholar] [CrossRef]
  59. Ziosi, M.; Di Meo, I.; Kleiner, G.; Gao, X.H.; Barca, E.; Sanchez-Quintero, M.J.; Tadesse, S.; Jiang, H.; Qiao, C.; Rodenburg, R.J.; et al. Coenzyme Q deficiency causes impairment of the sulfide oxidation pathway. EMBO Mol. Med. 2017, 9, 96–111. [Google Scholar] [CrossRef] [PubMed]
  60. Luna-Sanchez, M.; Hidalgo-Gutierrez, A.; Hildebrandt, T.M.; Chaves-Serrano, J.; Barriocanal-Casado, E.; Santos-Fandila, A.; Romero, M.; Sayed, R.K.; Duarte, J.; Prokisch, H.; et al. CoQ deficiency causes disruption of mitochondrial sulfide oxidation, a new pathomechanism associated with this syndrome. EMBO Mol. Med. 2017, 9, 78–95. [Google Scholar] [CrossRef] [PubMed]
  61. Zhang, M.; Wakitani, S.; Hayashi, K.; Miki, R.; Kawamukai, M. High production of sulfide in coenzyme Q deficient fission yeast. BioFactors 2008, 32, 91–98. [Google Scholar] [CrossRef] [PubMed]
  62. Blake, R.L.; Hall, J.G.; Russell, E.S. Mitochondrial proline dehydrogenase deficiency in hyperprolinemic PRO/Re mice: Genetic and enzymatic analyses. Biochem. Genet. 1976, 14, 739–757. [Google Scholar] [CrossRef] [PubMed]
  63. 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]
  64. Evans, D.R.; Guy, H.I. Mammalian Pyrimidine Biosynthesis: Fresh Insights into an Ancient Pathway. J. Biol. Chem. 2004, 279, 33035–33038. [Google Scholar] [CrossRef] [Green Version]
  65. Jones, E.M. Pyrimidine Nucleotide Biosynthesis in Animals: Genes, Enzymes, and Regulation of UMP Biosynthesis. Annu. Rev. Biochem. 1980, 49, 253–279. [Google Scholar] [CrossRef] [PubMed]
  66. López-Martín, J.M.; Salviati, L.; Trevisson, E.; Montini, G.; DiMauro, S.; Quinzii, C.; Hirano, M.; Rodriguez-Hernandez, A.; Cordero, M.D.; Sánchez-Alcázar, J.A.; et al. Missense mutation of the COQ2 gene causes defects of bioenergetics and de novo pyrimidine synthesis. Hum. Mol. Genet. 2007, 16, 1091–1097. [Google Scholar] [CrossRef] [PubMed]
  67. Watmough, N.J.; Frerman, F.E. The electron transfer flavoprotein: Ubiquinone oxidoreductases. Biochim. Biophys. Acta 2010, 1797, 1910–1916. [Google Scholar] [CrossRef] [PubMed]
  68. Summitt, C.B.; Johnson, L.C.; Jönsson, T.J.; Parsonage, D.; Holmes, R.P.; Lowther, W.T. Proline dehydrogenase 2 (PRODH2) is a hydroxyproline dehydrogenase (HYPDH) and molecular target for treating primary hyperoxaluria. Biochem. J. 2015, 466, 273–281. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  69. Forsmark-Andrée, P.; Dallner, G.; Ernster, L. Endogenous ubiquinol prevents protein modification accompanying lipid peroxidation in beef heart submitochondrial particles. Free Radic. Biol. Med. 1995, 19, 749–757. [Google Scholar] [CrossRef]
  70. Godic, A.; Poljsak, B.; Adamič, M.; Dahmane, R. The Role of Antioxidants in Skin Cancer Prevention and Treatment. Oxidative Med. Cell. Longev. 2014, 2014, 1–6. [Google Scholar] [CrossRef]
  71. Mukai, K.; Kikuchi, S.; Urano, S. Stopped-flow kinetic study of the regeneration reaction of tocopheroxyl radical by reduced ubiquinone-10 in solution. Biochim. Biophys. Acta 1990, 1035, 77–82. [Google Scholar] [CrossRef]
  72. 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]
  73. Forsmark-Andree, P.; Ernster, L. Evidence for a protective effect of endogenous ubiquinol against oxidative damage to mitochondrial protein and DNA during lipid peroxidation. Mol. Asp. Med. 1994, 15, S73–S81. [Google Scholar] [CrossRef]
  74. Tomasetti, M.; Littarru, G.; Stocker, R.; Alleva, R. Coenzyme Q10 enrichment decreases oxidative DNA damage in human lymphocytes. Free Radic. Biol. Med. 1999, 27, 1027–1032. [Google Scholar] [CrossRef]
  75. Tomasetti, M.; Alleva, R.; Collins, A.R. In vivo supplementation with coenzyme Q10 enhances the recovery of human lymphocytes from oxidative DNA damage. FASEB J. 2001, 15, 1425–1427. [Google Scholar] [CrossRef] [PubMed]
  76. Bentinger, M.; Brismar, K.; Dallner, G. The antioxidant role of coenzyme Q. Mitochondrion 2007, 7, S41–S50. [Google Scholar] [CrossRef] [PubMed]
  77. 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]
  78. Fontaine, E.; Ichas, F.; Bernardi, P. A Ubiquinone-binding Site Regulates the Mitochondrial Permeability Transition Pore. J. Biol. Chem. 1998, 273, 25734–25740. [Google Scholar] [CrossRef] [Green Version]
  79. Echtay, K.S.; Winkler, E.; Klingenberg, M.; Echtay, K.S.; Winkler, E.; Echtay, K.S.; Winkler, E.; Klingenberg, M. Coenzyme Q is an obligatory cofactor for uncoupling protein function. Nature 2000, 408, 609–613. [Google Scholar] [CrossRef]
  80. Echtay, K.S.; Winkler, E.; Frischmuth, K.; Klingenberg, M. Uncoupling proteins 2 and 3 are highly active H(+) transporters and highly nucleotide sensitive when activated by coenzyme Q (ubiquinone). Proc. Natl. Acad. Sci. USA 2001, 98, 1416–1421. [Google Scholar] [CrossRef]
  81. Jaburek, M.; Garlid, K.D. Reconstitution of recombinant uncoupling proteins: UCP1, -2, and -3 have similar affinities for ATP and are unaffected by coenzyme Q10. J. Biol. Chem. 2003, 278, 25825–25831. [Google Scholar] [CrossRef] [PubMed]
  82. Esteves, T.C.; Echtay, K.S.; Jonassen, T.; Clarke, C.F.; Brand, M.D. Ubiquinone is not required for proton conductance by uncoupling protein 1 in yeast mitochondria. Biochem. J. 2004, 379, 309–315. [Google Scholar] [CrossRef] [PubMed]
  83. Sluse, F.E.; Jarmuszkiewicz, W.; Navet, R.; Douette, P.; Mathy, G.; Sluse-Goffart, C.M. Mitochondrial UCPs: New insights into regulation and impact. Biochim. Biophys. Acta 2006, 1757, 480–485. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  84. Quinzii, C.M.; Lopez, L.C.; Von-Moltke, J.; Naini, A.; Krishna, S.; Schuelke, M.; Salviati, L.; Navas, P.; DiMauro, S.; Hirano, M. Respiratory chain dysfunction and oxidative stress correlate with severity of primary CoQ10 deficiency. FASEB J. 2008, 22, 1874–1885. [Google Scholar] [CrossRef] [PubMed]
  85. Quinzii, C.M.; Lopez, L.C.; Gilkerson, R.W.; Dorado, B.; Coku, J.; Naini, A.B.; Lagier-Tourenne, C.; Schuelke, M.; Salviati, L.; Carrozzo, R.; et al. Reactive oxygen species, oxidative stress, and cell death correlate with level of CoQ10 deficiency. FASEB J. 2010, 24, 3733–3743. [Google Scholar] [CrossRef] [PubMed]
  86. Quinzii, C.M.; Garone, C.; Emmanuele, V.; Tadesse, S.; Krishna, S.; Dorado, B.; Hirano, M. Tissue-specific oxidative stress and loss of mitochondria in CoQ-deficient Pdss2 mutant mice. FASEB J. 2013, 27, 612–621. [Google Scholar] [CrossRef] [PubMed]
  87. 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]
  88. Mollet, J.; Giurgea, I.; Schlemmer, D.; Dallner, G.; Chrétien, 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, 765–772. [Google Scholar] [CrossRef] [Green Version]
  89. Sadowski, C.E.; Lovric, S.; Ashraf, S.; Pabst, W.L.; Gee, H.Y.; Kohl, S.; Engelmann, S.; Vega-Warner, V.; Fang, H.; Halbritter, J.; et al. A single-gene cause in 29.5% of cases of steroid-resistant nephrotic syndrome. J. Am. Soc. Nephrol. 2015, 26, 1279–1289. [Google Scholar]
  90. López, L.C.; Schuelke, M.; Quinzii, C.M.; Kanki, T.; Rodenburg, R.J.T.; Naini, A.; DiMauro, S.; Hirano, M. Leigh Syndrome with Nephropathy and CoQ10 Deficiency Due to decaprenyl diphosphate synthase subunit 2 (PDSS2) Mutations. Am. J. Hum. Genet. 2006, 79, 1125–1129. [Google Scholar] [CrossRef] [PubMed]
  91. 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 Q10 deficiency. Am. J. Hum. Genet. 2006, 78, 345–349. [Google Scholar] [CrossRef] [PubMed]
  92. Heeringa, S.F.; Chernin, G.; Chaki, M.; Zhou, W.; Sloan, A.J.; Ji, Z.; Xie, L.X.; Salviati, L.; Hurd, T.W.; Vega-Warner, V.; et al. COQ6 mutations in human patients produce nephrotic syndrome with sensorineural deafness. J. Clin. Investig. 2011, 121, 2013–2024. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  93. Sondheimer, N.; Hewson, S.; Cameron, J.M.; Somers, G.R.; Broadbent, J.D.; Ziosi, M.; Quinzii, C.M.; Naini, A.B. Novel recessive mutations in COQ4 cause severe infantile cardiomyopathy and encephalopathy associated with CoQ10 deficiency. Mol. Genet. Metab. Rep. 2017, 12, 23–27. [Google Scholar] [CrossRef] [PubMed]
  94. Chung, W.K.; Martin, K.; Jalas, C.; Braddock, S.R.; Juusola, J.; Monaghan, K.G.; Warner, B.; Franks, S.; Yudkoff, M.; Lulis, L.; et al. Mutations in COQ4, an essential component of coenzyme Q biosynthesis, cause lethal neonatal mitochondrial encephalomyopathy. J. Med. Genet. 2015, 52, 627–635. [Google Scholar] [CrossRef] [PubMed]
  95. Brea-Calvo, G.; Haack, T.B.; Karall, D.; Ohtake, A.; Invernizzi, F.; Carrozzo, R.; Kremer, L.; Dusi, S.; Fauth, C.; Scholl-Burgi, S.; et al. COQ4 mutations cause a broad spectrum of mitochondrial disorders associated with CoQ10 deficiency. Am. J. Hum. Genet. 2015, 96, 309–317. [Google Scholar] [CrossRef]
  96. Malicdan, M.C.V.; Vilboux, T.; Ben-Zeev, B.; Guo, J.; Eliyahu, A.; Pode-Shakked, B.; Dori, A.; Kakani, S.; Chandrasekharappa, S.C.; Ferreira, C.R.; et al. A novel inborn error of the coenzyme Q10 biosynthesis pathway: Cerebellar ataxia and static encephalomyopathy due to COQ5 C-methyltransferase deficiency. Hum. Mutat. 2018, 39, 69–79. [Google Scholar] [CrossRef]
  97. Ashraf, S.; Gee, H.Y.; Woerner, S.; Xie, L.X.; Vega-Warner, V.; Lovric, S.; Fang, H.; Song, X.; Cattran, D.C.; Avila-Casado, C.; et al. ADCK4 mutations promote steroid-resistant nephrotic syndrome through CoQ10 biosynthesis disruption. J. Clin. Investig. 2013, 123, 5179–5189. [Google Scholar] [CrossRef]
  98. 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.; et al. A Nonsense Mutation in COQ9 Causes Autosomal-Recessive Neonatal-Onset Primary Coenzyme Q10 Deficiency: A Potentially Treatable Form of Mitochondrial Disease. Am. J. Hum. Genet. 2009, 84, 558–566. [Google Scholar] [CrossRef]
  99. Alcázar-Fabra, M.; Trevisson, E.; Brea-Calvo, G. Clinical syndromes associated with Coenzyme Q10 deficiency. Essays Biochem. 2018, 62, 377–398. [Google Scholar] [CrossRef]
  100. Padilla, S.; González-Mariscal, I.; Martín-Montalvo, A.; Gonzalez-Mariscal, I.; García-Testón, E.; Martin-Montalvo, A.; Pomares-Viciana, T.; Vazquez-Fonseca, L.; Gandolfo-Domínguez, P.; Santos-Ocaña, C.; et al. Regulation of coenzyme Q biosynthesis in yeast: A new complex in the block. IUBMB Life 2014, 66, 63–70. [Google Scholar]
  101. Yubero, D.; Montero, R.; Martin, M.A.; Montoya, J.; Ribes, A.; Grazina, M.; Trevisson, E.; Rodriguez-Aguilera, J.C.; Hargreaves, I.P.; Salviati, L.; et al. Secondary coenzyme Q10 deficiencies in oxidative phosphorylation (OXPHOS) and non-OXPHOS disorders. Mitochondrion 2016, 30, 51–58. [Google Scholar] [CrossRef] [PubMed]
  102. Montero, R.; Grazina, M.; López-Gallardo, E.; Montoya, J.; Briones, P.; Navarro-Sastre, A.; Land, J.M.; Hargreaves, I.P.; Artuch, R.; O’Callaghan, M.D.M.; et al. Coenzyme Q10 deficiency in mitochondrial DNA depletion syndromes. Mitochondrion 2013, 13, 337–341. [Google Scholar] [CrossRef] [PubMed]
  103. Quinzii, C.M.; Kattah, A.G.; Naini, A.; Akman, H.O.; Mootha, V.K.; DiMauro, S.; Hirano, M. Coenzyme Q deficiency and cerebellar ataxia associated with an aprataxin mutation. Neurology 2005, 64, 539–541. [Google Scholar] [CrossRef] [PubMed]
  104. Gempel, K.; Topaloglu, H.; Talim, B.; Schneiderat, P.; Schoser, B.G.H.; Hans, V.H.; Pálmafy, B.; Kale, G.; Tokatli, A.; Quinzii, C.; et al. The myopathic form of coenzyme Q10 deficiency is caused by mutations in the electron-transferring-flavoprotein dehydrogenase (ETFDH) gene. Brain 2007, 130, 2037–2044. [Google Scholar] [CrossRef] [PubMed]
  105. Aeby, A.; Sznajer, Y.; Cave, H.; Rebuffat, E.; Van Coster, R.; Rigal, O.; Van Bogaert, P.; Coster, R.; Bogaert, P. Cardiofaciocutaneous (CFC) syndrome associated with muscular coenzyme Q10 deficiency. J. Inherit. Metab. Dis. 2007, 30, 827. [Google Scholar] [CrossRef] [PubMed]
  106. Balreira, A.; Boczonadi, V.; Barca, E.; Pyle, A.; Bánsági, B.; Appleton, M.; Graham, C.; Hargreaves, I.P.; Rasic, V.M.; Lochmüller, H.; et al. ANO10 mutations cause ataxia and coenzyme Q10 deficiency. J. Neurol. 2014, 261, 2192–2198. [Google Scholar] [CrossRef] [PubMed]
  107. Miranda, S.; Foncea, R.; Guerrero, J.; Leighton, F. Oxidative Stress and Upregulation of Mitochondrial Biogenesis Genes in Mitochondrial DNA-Depleted HeLa Cells. Biochem. Biophys. Res. Commun. 1999, 258, 44–49. [Google Scholar] [CrossRef] [PubMed]
  108. Desbats, M.A.; Lunardi, G.; Doimo, M.; Trevisson, E.; Salviati, L. Genetic bases and clinical manifestations of coenzyme Q10 (CoQ10) deficiency. J. Inherit. Metab. Dis. 2015, 38, 145–156. [Google Scholar] [CrossRef] [PubMed]
  109. Rahman, S.; Clarke, C.F.; Hirano, M. 176th ENMC International Workshop: Diagnosis and treatment of coenzyme Q10 deficiency. Neuromuscul. Disord. 2012, 22, 76–86. [Google Scholar] [CrossRef] [PubMed]
  110. Marcoff, L.; Thompson, P.D. The role of coenzyme Q10 in statin-associated myopathy: A systematic review. J. Am. Coll. Cardiol. 2007, 49, 2231–2237. [Google Scholar] [CrossRef] [PubMed]
  111. Kühl, I.; Miranda, M.; Atanassov, I.; Kuznetsova, I.; Hinze, Y.; Mourier, A.; Filipovska, A.; Larsson, N.-G. Transcriptomic and proteomic landscape of mitochondrial dysfunction reveals secondary coenzyme Q deficiency in mammals. eLife 2017, 6, 6. [Google Scholar] [CrossRef] [PubMed]
  112. Fazakerley, D.J.; Chaudhuri, R.; Yang, P.; Maghzal, G.J.; Thomas, K.C.; Krycer, J.R.; Humphrey, S.J.; Parker, B.L.; Fisher-Wellman, K.H.; Meoli, C.C.; et al. Mitochondrial CoQ deficiency is a common driver of mitochondrial oxidants and insulin resistance. eLife 2018, 7, 7. [Google Scholar] [CrossRef] [PubMed]
  113. Harman, D. Aging: A Theory Based on Free Radical and Radiation Chemistry. J. Gerontol. 1956, 11, 298–300. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  114. Harman, D. The Biologic Clock: The Mitochondria? J. Am. Geriatr. Soc. 1972, 20, 145–147. [Google Scholar] [CrossRef] [PubMed]
  115. Yun, J.; Finkel, T. Mitohormesis. Cell Metab. 2014, 19, 757–766. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  116. 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]
  117. Cristina, D.; Cary, M.; Lunceford, A.; Clarke, C.; Kenyon, C. A Regulated Response to Impaired Respiration Slows Behavioral Rates and Increases Lifespan in Caenorhabditis elegans. PLoS Genet. 2009, 5, e1000450. [Google Scholar] [CrossRef] [PubMed]
  118. Yang, Y.-Y.; Gangoiti, J.A.; Sedensky, M.M.; Morgan, P.G. The effect of different ubiquinones on lifespan in Caenorhabditis elegans. Mech. Ageing Dev. 2009, 130, 370–376. [Google Scholar] [CrossRef] [Green Version]
  119. Jonassen, T.; Larsen, P.L.; Clarke, C.F. A dietary source of coenzyme Q is essential for growth of long-lived Caenorhabditis elegans clk-1 mutants. Proc. Natl. Acad. Sci. USA 2001, 98, 421–426. [Google Scholar] [CrossRef]
  120. Lakowski, B.; Hekimi, S. Determination of Life-Span in Caenorhabditis elegans by Four Clock Genes. Science 1996, 272, 1010–1013. [Google Scholar] [CrossRef]
  121. Barnes, T.M.; Lakowski, B.; Ewbank, J.J.; Lussier, M.; Bussey, H.; Hekimi, S. Structural and Functional Conservation of the Caenorhabditis elegans Timing Gene clk-1. Science 1997, 275, 980–983. [Google Scholar]
  122. Shibata, Y.; Branicky, R.; Landaverde, I.O.; Hekimi, S. Redox Regulation of Germline and Vulval Development in Caenorhabditis elegans. Science 2003, 302, 1779–1782. [Google Scholar] [CrossRef] [PubMed]
  123. Braeckman, B.P.; Houthoofd, K.; De Vreese, A.; Vanfleteren, J.R. Apparent uncoupling of energy production and consumption in long-lived Clk mutants of Caenorhabditis elegans. Curr. Biol. 1999, 9, 493–497. [Google Scholar] [CrossRef] [Green Version]
  124. Braeckman, B.P.; Houthoofd, K.; Brys, K.; Lenaerts, I.; De Vreese, A.; Van Eygen, S.; Raes, H.; Vanfleteren, J.R. No reduction of energy metabolism in Clk mutants. Mech. Ageing Dev. 2002, 123, 1447–1456. [Google Scholar] [CrossRef]
  125. Van Voorhies, W.A. The influence of metabolic rate on longevity in the nematode Caenorhabditis elegans. Aging Cell 2002, 1, 91–101. [Google Scholar] [CrossRef] [PubMed]
  126. Braeckman, B.P.; Houthoofd, K.; Vanfleteren, J.R. Assessing metabolic activity in aging Caenorhabditis elegans: Concepts and controversies. Aging Cell 2002, 1, 82–88. [Google Scholar] [CrossRef] [PubMed]
  127. Larsen, P.L. Extension of Life-Span in Caenorhabditis elegans by a Diet Lacking Coenzyme Q. Science 2002, 295, 120–123. [Google Scholar] [CrossRef] [PubMed]
  128. Asencio, C.; Rodríguez-Aguilera, J.C.; Ruiz-Ferrer, M.; Vela, J.; Navas, P. Silencing of ubiquinone biosynthesis genes extends life span in Caenorhabditis elegans. FASEB J. 2003, 17, 1135–1137. [Google Scholar] [CrossRef] [PubMed]
  129. Kayser, E.B.; Sedensky, M.M.; Morgan, P.G.; Hoppel, C.L. Mitochondrial oxidative phosphorylation is defective in the long-lived mutant clk-1. J. Biol. Chem. 2004, 279, 54479–54486. [Google Scholar] [CrossRef] [PubMed]
  130. 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 ofCaenorhabditis elegans. FEBS Lett. 2006, 580, 1740–1746. [Google Scholar] [CrossRef] [PubMed]
  131. Gomez, F.; Saiki, R.; Chin, R.; Srinivasan, C.; Clarke, C.F. Restoring de novo Coenzyme Q biosynthesis in Caenorhabditis elegans coq-3 mutants yields profound rescue compared to exogenous Coenzyme Q supplementation. Gene 2012, 506, 106–116. [Google Scholar] [CrossRef] [PubMed]
  132. Ishii, N.; Senoo-Matsuda, N.; Miyake, K.; Yasuda, K.; Ishii, T.; Hartman, P.S.; Furukawa, S. Coenzyme Q10 can prolong C. elegans lifespan by lowering oxidative stress. Mech. Ageing Dev. 2004, 125, 41–46. [Google Scholar] [CrossRef] [PubMed]
  133. Fischer, A.; Niklowitz, P.; Menke, T.; Döring, F. Promotion of growth by Coenzyme Q10 is linked to gene expression in C. elegans. Biochem. Biophys. Res. Commun. 2014, 452, 920–927. [Google Scholar] [CrossRef] [PubMed]
  134. Hihi, A.K.; Gao, Y.; Hekimi, S. Ubiquinone is necessary for Caenorhabditis elegans development at mitochondrial and non-mitochondrial sites. J. Biol. Chem. 2002, 277, 2202–2206. [Google Scholar] [CrossRef] [PubMed]
  135. Padilla, S.; Jonassen, T.; Jiménez-Hidalgo, M.A.; Fernández-Ayala, D.J.M.; López-Lluch, G.; Marbois, B.; Navas, P.; Clarke, C.F.; Santos-Ocaña, C.; Fernández-Ayala, D.J.M. Demethoxy-Q, An Intermediate of Coenzyme Q Biosynthesis, Fails to Support Respiration in Saccharomyces cerevisiae and Lacks Antioxidant Activity. J. Biol. Chem. 2004, 279, 25995–26004. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  136. Miyadera, H.; Kano, K.; Miyoshi, H.; Ishii, N.; Hekimi, S.; Kita, K. Quinones in long-lived clk-1 mutants of Caenorhabditis elegans. FEBS Lett. 2002, 512, 33–37. [Google Scholar] [CrossRef]
  137. 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. Genome Res. 2005, 19, 2424–2434. [Google Scholar] [CrossRef] [PubMed]
  138. Rodríguez-Hidalgo, M.; Luna-Sánchez, M.; Hidalgo-Gutiérrez, A.; Barriocanal-Casado, E.; Mascaraque, C.; Acuña-Castroviejo, D.; Rivera, M.; Escames, G.; López, L.C. Reduction in the levels of CoQ biosynthetic proteins is related to an increase in lifespan without evidence of hepatic mitohormesis. Sci. Rep. 2018, 8, 14013. [Google Scholar] [CrossRef]
  139. Takahashi, M.; Shimizu, T.; Moriizumi, E.; Shirasawa, T. Clk-1 deficiency induces apoptosis associated with mitochondrial dysfunction in mouse embryos. Mech. Ageing Dev. 2008, 129, 291–298. [Google Scholar] [CrossRef]
  140. 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] [Green Version]
  141. Holzenberger, M.; Dupont, J.; Ducos, B.; Leneuve, P.; Geloen, A.; Even, P.C.; Cervera, P.; Le Bouc, Y. IGF-1 receptor regulates lifespan and resistance to oxidative stress in mice. Nature 2003, 421, 182–187. [Google Scholar] [CrossRef] [PubMed]
  142. Baba, T.; Shimizu, T.; Suzuki, Y.-I.; Ogawara, M.; Isono, K.-I.; Koseki, H.; Kurosawa, H.; Shirasawa, T. Estrogen, Insulin, and Dietary Signals Cooperatively Regulate Longevity Signals to Enhance Resistance to Oxidative Stress in Mice. J. Biol. Chem. 2005, 280, 16417–16426. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  143. Napoli, C.; Martin-Padura, I.; De Nigris, F.; Giorgio, M.; Mansueto, G.; Somma, P.; Condorelli, M.; Sica, G.; De Rosa, G.; Pelicci, P. Deletion of the p66Shc longevity gene reduces systemic and tissue oxidative stress, vascular cell apoptosis, and early atherogenesis in mice fed a high-fat diet. Proc. Natl. Acad. Sci. USA 2003, 100, 2112–2116. [Google Scholar] [CrossRef] [PubMed]
  144. Hauck, S.J.; Aaron, J.M.; Wright, C.; Kopchick, J.J.; Bartke, A. Antioxidant Enzymes, Free-Radical Damage, and Response to Paraquat in Liver and Kidney of Long-Living Growth Hormone Receptor/Binding Protein Gene-Disrupted Mice. Horm. Metab. Res. 2002, 34, 481–486. [Google Scholar] [CrossRef] [PubMed]
  145. Lapointe, J.; Hekimi, S. Early mitochondrial dysfunction in long-lived Mclk1+/- mice. J. Biol. Chem. 2008, 283, 26217–26227. [Google Scholar] [CrossRef] [PubMed]
  146. Schulz, T.J.; Zarse, K.; Voigt, A.; Urban, N.; Birringer, M.; Ristow, M. Glucose Restriction Extends Caenorhabditis elegans Life Span by Inducing Mitochondrial Respiration and Increasing Oxidative Stress. Cell Metab. 2007, 6, 280–293. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  147. Ballard, J.W.O.; Katewa, S.D.; Melvin, R.G.; Chan, G.; Ballard, J.W.O. Comparative Analysis of Mitochondrial Genotype and Aging. Ann. N. Y. Acad. Sci. 2007, 1114, 93–106. [Google Scholar] [CrossRef]
  148. Brys, K.; Vanfleteren, J.R.; Braeckman, B.P. Testing the rate-of-living/oxidative damage theory of aging in the nematode model Caenorhabditis elegans. Exp. Gerontol. 2007, 42, 845–851. [Google Scholar] [CrossRef]
  149. Miles, M.V. The uptake and distribution of coenzyme Q(10). Mitochondrion 2007, 7, S72–S77. [Google Scholar] [CrossRef]
  150. 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.S.; Rosenwasser, M.J.; et al. 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] [Green Version]
  151. 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]
  152. Battino, M.; Gorini, A.; Villa, R.; Genova, M.; Bovina, C.; Sassi, S.; Littarru, G.; 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]
  153. Åberg, F.; Zhang, Y.; Teclebrhan, H.; Appelkvist, E.-L.; Dallner, G. Increases in tissue levels of ubiquinone in association with peroxisome proliferation. Chem. Interact. 1996, 99, 205–218. [Google Scholar] [CrossRef]
  154. Lass, A.; Forster, M.J.; Sohal, R.S. Effects of coenzyme Q10 and α-tocopherol administration on their tissue levels in the mouse: Elevation of mitochondrial α-tocopherol by coenzyme Q10. Free Radic. Biol. Med. 1999, 26, 1375–1382. [Google Scholar] [CrossRef]
  155. Lass, A.; Sohal, R.S. Effect of coenzyme Q(10) and alpha-tocopherol content of mitochondria on the production of superoxide anion radicals. FASEB J. 2000, 14, 87–94. [Google Scholar] [CrossRef] [PubMed]
  156. 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 Q10 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]
  157. 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]
  158. Yoo, J.-Y.; Yum, K.-S. Effect of Coenzyme Q10 on Insulin Resistance in Korean Patients with Prediabetes: A Pilot Single-Center, Randomized, Double-Blind, Placebo-Controlled Study. BioMed Res. Int. 2018, 2018, 1–6. [Google Scholar] [CrossRef] [PubMed]
  159. Dahri, M.; Tarighat-Esfanjani, A.; Asghari-Jafarabadi, M.; Hashemilar, M. Oral coenzyme Q10 supplementation in patients with migraine: Effects on clinical features and inflammatory markers. Nutr. Neurosci. 2019, 22, 607–615. [Google Scholar] [CrossRef] [PubMed]
  160. McGarry, A.; McDermott, M.; Kieburtz, K.; de Blieck, E.A.; Beal, F.; Marder, K.; Ross, C.; Shoulson, I.; Gilbert, P.; Mallonee, W.M.; et al. A randomized, double-blind, placebo-controlled trial of coenzyme Q10 in Huntington disease. Neurology 2017, 88, 152–159. [Google Scholar] [CrossRef]
  161. 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. [Google Scholar] [CrossRef] [PubMed]
  162. Zhang, Y.; Liu, J.; Chen, X.-Q.; Chen, C.-Y.O. Ubiquinol is superior to ubiquinone to enhance Coenzyme Q10 status in older men. Food Funct. 2018, 9, 5653–5659. [Google Scholar] [CrossRef] [PubMed]
  163. Hidaka, T.; Fujii, K.; Funahashi, I.; Fukutomi, N.; Hosoe, K. Safety assessment of coenzyme Q10 (CoQ10). BioFactors 2008, 32, 199–208. [Google Scholar] [CrossRef] [PubMed]
  164. Hathcock, J.N.; Shao, A. Risk assessment for coenzyme Q10 (Ubiquinone). Regul. Toxicol. Pharmacol. 2006, 45, 282–288. [Google Scholar] [CrossRef] [PubMed]
  165. Shetty, R.A.; Forster, M.J.; Sumien, N. Coenzyme Q10 supplementation reverses age-related impairments in spatial learning and lowers protein oxidation. AGE 2013, 35, 1821–1834. [Google Scholar] [CrossRef] [PubMed]
  166. Rivara, M.B.; Yeung, C.K.; Robinson-Cohen, C.; Phillips, B.R.; Ruzinski, J.; Rock, D.; Linke, L.; Shen, D.D.; Ikizler, T.A.; Himmelfarb, J. Effect of Coenzyme Q10 on Biomarkers of Oxidative Stress and Cardiac Function in Hemodialysis Patients: The CoQ10 Biomarker Trial. Am. J. Kidney Dis. 2017, 69, 389–399. [Google Scholar] [CrossRef] [PubMed]
  167. García-Corzo, L.; Luna-Sánchez, M.; Doerrier, C.; Ortiz, F.; Escames, G.; Acuña-Castroviejo, D.; López, L.C. Ubiquinol-10 ameliorates mitochondrial encephalopathy associated with CoQ deficiency. Biochim. Biophys. Acta 2014, 1842, 893–901. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  168. 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]
  169. Lee, C.-K.; Pugh, T.D.; Klopp, R.G.; Edwards, J.; Allison, D.B.; Weindruch, R.; Prolla, A.T. The impact of α-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]
  170. Lönnrot, K.; Alho, H.; Holm, P.; Lagerstedt, A.; Huhtala, H. The effects of lifelong ubiquinone Q10 supplementation on the Q9 and Q10 tissue concentrations and life span of male rats and mice. IUBMB Life 1998, 44, 727–737. [Google Scholar] [CrossRef]
  171. Ramirez-Tortosa, C.L.; Varela-López, A.; Navarro-Hortal, M.D.; Ramos-Pleguezuelos, F.M.; Márquez-Lobo, B.; Ramirez-Tortosa, M.C.; Ochoa, J.J.; Battino, M.; Quiles, J.L. Longevity and cause of death in male Wistar rats fed lifelong diets based on virgin olive oil, sunflower oil or fish oil. J. Gerontol. Ser. A 2019. [Google Scholar] [CrossRef] [PubMed]
  172. Reeves, P.G. Components of the AIN-93 Diets as Improvements in the AIN-76A Diet. J. Nutr. 1997, 127, 838S–841S. [Google Scholar] [CrossRef] [PubMed]
  173. Varela-Lopez, A.; Bullon, P.; Battino, M.; Ramirez-Tortosa, M.; 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. Ser. A Biomed. Sci. Med. Sci. 2015, 71, 593–600. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  174. Varela-Lopez, A.; Ochoa, J.J.; Llamas-Elvira, J.M.; Lopez-Frias, M.; Planells, E.; Speranza, L.; Battino, M.; Quiles, J.L. Loss of Bone Mineral Density Associated with Age in Male Rats Fed on Sunflower Oil Is Avoided by Virgin Olive Oil Intake or Coenzyme Q Supplementation. Int. J. Mol. Sci. 2017, 18, 1397. [Google Scholar] [CrossRef] [PubMed]
  175. Gonzalez-Alonso, A.; Ramirez-Tortosa, C.L.; Varela-López, A.; Roche, E.; Arribas, M.I.; Ramirez-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] [Green Version]
  176. Huertas, J.R.; Ibáñez, S.; López-Frias, M.; Ochoa, J.J.; Quiles, J.L.; Castelli, G.P.; Mataix, J.; Lenaz, G.; Huertas, J.F.R.; Martinez-Velasco, E.; et al. Virgin olive oil and coenzyme Q10protect heart mitochondria from peroxidative damage during aging. BioFactors 1999, 9, 337–343. [Google Scholar] [CrossRef] [PubMed]
  177. 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 Q10. Free Radic. Biol. Med. 2011, 50, 1053–1064. [Google Scholar] [CrossRef] [PubMed]
  178. Ochoa, J.J.; Quiles, J.L.; López-Frías, M.; Huertas, J.R.; Mataix, J.; Huertas, J.F.R. 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. Ser. A Biol. Sci. Med. Sci. 2007, 62, 1211–1218. [Google Scholar] [CrossRef] [PubMed]
  179. Quiles, J.L.; Ochoa, J.J.; Battino, M.; Nepomuceno, E.A.; Frías, M.L.; Huertas, J.R.; Mataix, J.; Gutierrez-Rios, P.; Huertas, J.F.R.; Gutierrez-Rios, P. Life-long supplementation with a low dosage of coenzyme Q10 in the rat: Effects on antioxidant status and DNA damage. BioFactors 2005, 25, 73–86. [Google Scholar] [CrossRef]
  180. Quiles, J.L.; Ochoa, J.J.; Huertas, J.R.; Mataix, J.; Huertas, J.F.R. 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]
  181. 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]
  182. Varela-Lopez, A.; Ochoa, J.J.; Llamas-Elvira, J.M.; Lopez-Frias, M.; Planells, E.; Ramirez-Tortosa, M.; Ramirez-Tortosa, C.L.; Giampieri, F.; Battino, M.; Quiles, J.L. Age-Related Loss in Bone Mineral Density of Rats Fed Lifelong on a Fish Oil-Based Diet Is Avoided by Coenzyme Q10 Addition. Nutrients 2017, 9, 176. [Google Scholar] [CrossRef] [PubMed]
  183. Bello, R.I.; Gómez-Díaz, C.; Burón, M.I.; Alcain, 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]
  184. Bello, R.I.; Herman, M.D.; Navas, P.; Villalba, J.M.; Gómez-Díaz, C.; Burón, M.I.; Alcain, F.J.; González-Ojeda, R.; González-Reyes, J.A. Effect of dietary coenzyme Q and fatty acids on the antioxidant status of rat tissues. Protoplasma 2003, 221, 11–17. [Google Scholar]
  185. Dash, S.; Xiao, C.; Morgantini, C.; Lewis, G.F. New Insights into the Regulation of Chylomicron Production. Annu. Rev. Nutr. 2015, 35, 265–294. [Google Scholar] [CrossRef] [PubMed]
  186. Birru, W.A.; Warren, D.B.; Williams, H.D.; Benameur, H.; Porter, C.J.H.; Chalmers, D.K.; Ibrahim, A.; Pouton, C.W. Digestion of Phospholipids after Secretion of Bile into the Duodenum Changes the Phase Behavior of Bile Components. Mol. Pharm. 2014, 11, 2825–2834. [Google Scholar] [CrossRef] [PubMed]
  187. Takahashi, M.; Ogawara, M.; Shimizu, T.; Shirasawa, T. Restoration of the behavioral rates and lifespan in clk-1 mutant nematodes in response to exogenous coenzyme Q10. Exp. Gerontol. 2012, 47, 276–279. [Google Scholar] [CrossRef] [PubMed]
  188. 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]
  189. Moradi, M.; Haghighatdoost, F.; Feizi, A.; Larijani, B.; Azadbakht, L. Effect of Coenzyme Q10 Supplementation on Diabetes Biomarkers: A Systematic Review and Meta-analysis of Randomized Controlled Clinical Trials. Arch. Iran. Med. 2016, 19, 588–596. [Google Scholar] [PubMed]
  190. Raygan, F.; Rezavandi, Z.; Dadkhah Tehrani, S.; Farrokhian, A.; Asemi, Z. The effects of coenzyme Q10 administration on glucose homeostasis parameters, lipid profiles, biomarkers of inflammation and oxidative stress in patients with metabolic syndrome. Eur. J. Nutr. 2016, 55, 2357–2364. [Google Scholar] [CrossRef]
  191. Suksomboon, N.; Poolsup, N.; Juanak, N. Effects of coenzyme Q10supplementation on metabolic profile in diabetes: A systematic review and meta-analysis. J. Clin. Pharm. Ther. 2015, 40, 413–418. [Google Scholar] [CrossRef]
  192. Xu, Z.; Huo, J.; Ding, X.; Yang, M.; Li, L.; Dai, J.; Hosoe, K.; Kubo, H.; Mori, M.; Higuchi, K.; et al. Coenzyme Q10 Improves Lipid Metabolism and Ameliorates Obesity by Regulating CaMKII-Mediated PDE4 Inhibition. Sci. Rep. 2017, 7, 8253. [Google Scholar] [CrossRef]
  193. 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]
  194. Zhang, X.; Liu, H.; Hao, Y.; Xu, L.; Zhang, T.; Liu, Y.; Guo, L.; Zhu, L.; Pei, Z. Coenzyme Q10 protects against hyperlipidemia-induced cardiac damage in apolipoprotein E-deficient mice. Lipids Heal. Dis. 2018, 17, 279. [Google Scholar] [CrossRef] [PubMed]
  195. Lopez-Moreno, J.; Quintana-Navarro, G.M.; Delgado-Lista, J.; Garcia-Rios, A.; Alcala-Diaz, J.F.; Gomez-Delgado, F.; Camargo, A.; Perez-Martinez, P.; Tinahones, F.J.; Striker, G.E.; et al. Mediterranean Diet Supplemented With Coenzyme Q10 Modulates the Postprandial Metabolism of Advanced Glycation End Products in Elderly Men and Women. J. Gerontol. Ser. A Biol. Sci. Med. Sci. 2018, 73, 340–346. [Google Scholar]
  196. Mirhashemi, S.M.; Najafi, V.; Raygan, F.; Asemi, Z. The effects of coenzyme Q10 supplementation on cardiometabolic markers in overweight type 2 diabetic patients with stable myocardial infarction: A randomized, double-blind, placebo-controlled trial. ARYA Atheroscler 2016, 12, 158–165. [Google Scholar] [PubMed]
  197. Ayaz, M.; Tuncer, S.; Okudan, N.; Gokbel, H. Coenzyme Q(10) and alpha-lipoic acid supplementation in diabetic rats: Conduction velocity distributions. Methods Find. Exp. Clin. Pharmacol. 2008, 30, 367. [Google Scholar] [CrossRef] [PubMed]
  198. Monsef, A.; Shahidi, S.; Komaki, A. Influence of Chronic Coenzyme Q10 Supplementation on Cognitive Function, Learning, and Memory in Healthy and Diabetic Middle-Aged Rats. Neuropsychobiology 2019, 77, 92–100. [Google Scholar] [CrossRef]
  199. Schmelzer, C.; Kubo, H.; Mori, M.; Sawashita, J.; Kitano, M.; Hosoe, K.; Boomgaarden, I.; Doring, F.; Higuchi, K. Supplementation with the reduced form of Coenzyme Q10 decelerates phenotypic characteristics of senescence and induces a peroxisome proliferator-activated receptor-alpha gene expression signature in SAMP1 mice. Mol. Nutr. Food Res. 2010, 54, 805–815. [Google Scholar] [CrossRef]
  200. Flowers, N.; Hartley, L.; Todkill, D.; Stranges, S.; Rees, K. Co-enzyme Q10 supplementation for the primary prevention of cardiovascular disease. Cochrane Database Syst. Rev. 2014. [Google Scholar] [CrossRef]
  201. 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. Ser. A 2012, 67A, 3–10. [Google Scholar] [CrossRef] [PubMed]
  202. Fotino, A.D.; Thompson-Paul, A.M.; Bazzano, L.A. Effect of coenzyme Q10 supplementation on heart failure: A meta-analysis. Am. J. Clin. Nutr. 2013, 97, 268–275. [Google Scholar] [CrossRef] [PubMed]
  203. Madmani, M.E.; Yusuf Solaiman, A.; Tamr Agha, K.; Madmani, Y.; Shahrour, Y.; Essali, A.; Kadro, W. Coenzyme Q10 for heart failure. Cochrane Database Syst. Rev. 2014. [Google Scholar] [CrossRef] [PubMed]
  204. Ishikawa, A.; Kawarazaki, H.; Ando, K.; Fujita, M.; Fujita, T.; Homma, Y. Renal preservation effect of ubiquinol, the reduced form of coenzyme Q10. Clin. Exp. Nephrol. 2011, 15, 30–33. [Google Scholar] [CrossRef] [PubMed]
  205. Bolignano, D.; Cernaro, V.; Gembillo, G.; Baggetta, R.; Buemi, M.; D’Arrigo, G. Antioxidant agents for delaying diabetic kidney disease progression: A systematic review and meta-analysis. PLoS ONE 2017, 12, 0178699. [Google Scholar] [CrossRef] [PubMed]
  206. žmitek, J.; Šmidovnik, A.; Fir, M.; Prošek, M.; Zmitek, K.; Walczak, J.; Pravst, I. Relative Bioavailability of Two Forms of a Novel Water-Soluble Coenzyme Q10. Ann. Nutr. Metab. 2008, 52, 281–287. [Google Scholar]
  207. Galasko, D.R.; Peskind, E.; Clark, C.M.; Quinn, J.F.; Ringman, J.M.; Jicha, G.A.; Cotman, C.; Cottrell, B.; Montine, T.J.; Thomas, R.G.; et al. Antioxidants for Alzheimer disease: A randomized clinical trial with cerebrospinal fluid biomarker measures. Arch. Neurol. 2012, 69, 836–841. [Google Scholar] [CrossRef]
  208. Fan, L.; Feng, Y.; Chen, G.-C.; Qin, L.-Q.; Fu, C.-L.; Chen, L.-H. Effects of coenzyme Q10 supplementation on inflammatory markers: A systematic review and meta-analysis of randomized controlled trials. Pharmacol. Res. 2017, 119, 128–136. [Google Scholar] [CrossRef]
  209. Mazidi, M.; Kengne, A.P.; Banach, M. Effects of coenzyme Q10 supplementation on plasma C-reactive protein concentrations: A systematic review and meta-analysis of randomized controlled trials. Pharmacol. Res. 2018, 128, 130–136. [Google Scholar] [CrossRef]
  210. Tian, G.; Sawashita, J.; Kubo, H.; Nishio, S.-Y.; Hashimoto, S.; Suzuki, N.; Yoshimura, H.; Tsuruoka, M.; Wang, Y.; Liu, Y.; et al. Ubiquinol-10 Supplementation Activates Mitochondria Functions to Decelerate Senescence in Senescence-Accelerated Mice. Antioxid. Redox Signal. 2014, 20, 2606–2620. [Google Scholar] [CrossRef]
  211. 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]
  212. Bullon, P.; Battino, M.; Varela-Lopez, A.; Perez-Lopez, P.; Granados-Principal, S.; Ramirez-Tortosa, M.C.; Ochoa, J.J.; Cordero, M.D.; Gonzalez-Alonso, A.; Ramirez-Tortosa, C.L.; et al. Diets based on virgin olive oil or fish oil but not on sunflower oil prevent age-related alveolar bone resorption by mitochondrial-related mechanisms. PLoS ONE 2013, 8, e74234. [Google Scholar] [CrossRef] [PubMed]
  213. Ochoa, J.J.; Quiles, J.L.; Huertas, J.R.; Mataix, J.; Huertas, J.F.R. 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. Ser. A Biol. Sci. Med. Sci. 2005, 60, 970–975. [Google Scholar] [CrossRef] [PubMed]
  214. Ochoa, J.J.; Quiles, J.L.; Ibáñez, S.; Martínez, E.; López-Frías, M.; Huertas, J.F.R.; Mataix, J. Aging-Related Oxidative Stress Depends on Dietary Lipid Source in Rat Postmitotic Tissues. J. Bioenerg. Biomembr. 2003, 35, 267–275. [Google Scholar] [CrossRef] [PubMed]
  215. Quiles, J.L.; Martínez, E.; Ibáñez, S.; Ochoa, J.J.; Martín, Y.; López-Frías, M.; Huertas, J.F.R.; Mataix, J. Ageing-Related Tissue-Specific Alterations in Mitochondrial Composition and Function Are Modulated by Dietary Fat Type in the Rat. J. Bioenerg. Biomembr. 2002, 34, 517–524. [Google Scholar] [CrossRef] [PubMed]
  216. Orlando, P.; Silvestri, S.; Brugè, F.; Tiano, L.; Klöting, I.; Falcioni, G.; Polidori, C.; Sonia, S. High-fat diet-induced met-hemoglobin formation in rats prone (WOKW) or resistant (DA) to the metabolic syndrome: Effect of CoQ10supplementation. BioFactors 2014, 40, 603–609. [Google Scholar] [CrossRef] [PubMed]
  217. 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]
  218. Salehpour, F.; Farajdokht, F.; Mahmoudi, J.; Erfani, M.; Farhoudi, M.; Karimi, P.; Rasta, S.H.; Sadigh-Eteghad, S.; Hamblin, M.R.; Gjedde, A. Photobiomodulation and Coenzyme Q10 Treatments Attenuate Cognitive Impairment Associated With Model of Transient Global Brain Ischemia in Artificially Aged Mice. Front. Cell. Neurosci. 2019, 13, 74. [Google Scholar] [CrossRef] [PubMed]
  219. 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]
  220. Cano, A.; Ciaffoni, F.; Safwat, G.M.; Aspichueta, P.; Ochoa, B.; Bravo, E.; Botham, K.M.; Olascoaga, B.O. 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]
  221. Heidari, A.; Hamidi, G.; Soleimani, A.; Aghadavod, E.; Asemi, Z. Effects of Coenzyme Q10 Supplementation on Gene Expressions Related to Insulin, Lipid, and Inflammation Pathways in Patients With Diabetic Nephropathy. Iran. J. Kidney Dis. 2018, 12, 14–21. [Google Scholar] [PubMed]
  222. Kujjo, L.L.; Acton, B.M.; Perkins, G.A.; Ellisman, M.H.; D’Estaing, S.G.; Casper, R.F.; Jurisicova, A.; Perez, G.I. Ceramide and its transport protein (CERT) contribute to deterioration of mitochondrial structure and function in aging oocytes. Mech. Ageing Dev. 2013, 134, 43–52. [Google Scholar] [CrossRef] [PubMed]
  223. Ben-Meir, A.; Burstein, E.; Borrego-Alvarez, A.; Chong, J.; Wong, E.; Yavorska, T.; Naranian, T.; Chi, M.; Wang, Y.; Bentov, Y.; et al. Coenzyme Q10 restores oocyte mitochondrial function and fertility during reproductive aging. Aging Cell 2015, 14, 887–895. [Google Scholar] [CrossRef] [PubMed]
  224. Bentov, Y.; Yavorska, T.; Esfandiari, N.; Jurisicova, A.; Casper, R.F. The contribution of mitochondrial function to reproductive aging. J. Assist. Reprod. Genet. 2011, 28, 773–783. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  225. Ben-Meir, A.; Kim, K.; McQuaid, R.; Esfandiari, N.; Bentov, Y.; Casper, R.F.; Jurisicova, A. Co-Enzyme Q10 Supplementation Rescues Cumulus Cells Dysfunction in a Maternal Aging Model. Antioxidants 2019, 8, 58. [Google Scholar] [CrossRef] [PubMed]
  226. Gvozdjakova, A.; Kucharská, J.; Dubravicky, J.; Mojto, V.; Singh, R.B. Coenzyme Q10, α-Tocopherol, and Oxidative Stress Could Be Important Metabolic Biomarkers of Male Infertility. Dis. Markers 2015, 2015, 1–6. [Google Scholar] [CrossRef] [PubMed]
  227. Salas-Huetos, A.; Rosique-Esteban, N.; Becerra-Tomás, N.; Vizmanos, B.; Bulló, M.; Salas-Salvadó, J. The Effect of Nutrients and Dietary Supplements on Sperm Quality Parameters: A Systematic Review and Meta-Analysis of Randomized Clinical Trials. Adv. Nutr. 2018, 9, 833–848. [Google Scholar] [CrossRef] [PubMed]
  228. Yang, E.J.; Kim, S.; Kim, J.S.; Choi, I.H. Inflammasome formation and IL-1beta release by human blood monocytes in response to silver nanoparticles. Biomaterials 2012, 33, 6858–6867. [Google Scholar] [CrossRef] [PubMed]
  229. Takahashi, K.; Ohsawa, I.; Shirasawa, T.; Takahashi, M. Early-onset motor impairment and increased accumulation of phosphorylated alpha-synuclein in the motor cortex of normal aging mice are ameliorated by coenzyme Q. Exp. Gerontol. 2016, 81, 65–75. [Google Scholar] [CrossRef] [PubMed]
Figure 1. The coenzyme Q (CoQ) biosynthesis pathway and functions of CoQ. (a) Schematic model of mammalian cells CoQ biosynthesis pathway: 4-hydroxybenzoic acid (4HB) is the ring precursor, dimethylally pyrophosphate (DMAPP) and isopentenyl pyrophosphate (IPP) are precursors to form polyprenyl diphosphate via prenyldiphosphate synthase subunit (PDSS)1/PDSS2. CoQ2 attaches the polyisoprenyl tail to 4HB to form 3-hexaprenyl-4HB (HHB). The next intermediates are: 3-hexaprenyl-4,5-dihydroxybenzoic acid (DHHB), 3-hexaprenyl-4-hydroxy-5-methoxybenzoic acid (HMHB), 2-hexaprenyl-6-methoxy-1,4-benzenediol (DDMQH2), 2-hexaprenyl-3-methyl-6-methoxy-1,4-benzenediol (DMQH2), and 2-hexaprenyl-3-methyl-6-methoxy-1,4,5-benzenetriol (DMeQH2) to ultimately produce reduced coenzyme Q (CoQnH2). (b) CoQ functions in the mitochondria. CoQ accepts electrons from complex I and complex II, sulfide quinone oxidoreductase (SQOR), proline dehydrogenase 1 (PDH), coline dehydrogenase (CHDH), mitochondrial glycerol-3-phosphate dehydrogenase (G3PDH), dihydroorotate dehydrogenase (DHOH) and electron transport flavoprotein dehydrogenase (ETFDH).
Figure 1. The coenzyme Q (CoQ) biosynthesis pathway and functions of CoQ. (a) Schematic model of mammalian cells CoQ biosynthesis pathway: 4-hydroxybenzoic acid (4HB) is the ring precursor, dimethylally pyrophosphate (DMAPP) and isopentenyl pyrophosphate (IPP) are precursors to form polyprenyl diphosphate via prenyldiphosphate synthase subunit (PDSS)1/PDSS2. CoQ2 attaches the polyisoprenyl tail to 4HB to form 3-hexaprenyl-4HB (HHB). The next intermediates are: 3-hexaprenyl-4,5-dihydroxybenzoic acid (DHHB), 3-hexaprenyl-4-hydroxy-5-methoxybenzoic acid (HMHB), 2-hexaprenyl-6-methoxy-1,4-benzenediol (DDMQH2), 2-hexaprenyl-3-methyl-6-methoxy-1,4-benzenediol (DMQH2), and 2-hexaprenyl-3-methyl-6-methoxy-1,4,5-benzenetriol (DMeQH2) to ultimately produce reduced coenzyme Q (CoQnH2). (b) CoQ functions in the mitochondria. CoQ accepts electrons from complex I and complex II, sulfide quinone oxidoreductase (SQOR), proline dehydrogenase 1 (PDH), coline dehydrogenase (CHDH), mitochondrial glycerol-3-phosphate dehydrogenase (G3PDH), dihydroorotate dehydrogenase (DHOH) and electron transport flavoprotein dehydrogenase (ETFDH).
Nutrients 11 02221 g001
Figure 2. The role of coenzyme Q in aging: (a) Defects in CoQ biosynthesis cause decreased CoQ levels and increased life span, a fact that may be due to the changes in mitochondrial function and oxidative stress. (b) The increase in CoQ levels through CoQ10 dietary supplementation prevents some age-related damages that are associated with changes in redox states and mitochondrial function.
Figure 2. The role of coenzyme Q in aging: (a) Defects in CoQ biosynthesis cause decreased CoQ levels and increased life span, a fact that may be due to the changes in mitochondrial function and oxidative stress. (b) The increase in CoQ levels through CoQ10 dietary supplementation prevents some age-related damages that are associated with changes in redox states and mitochondrial function.
Nutrients 11 02221 g002
Table 1. Coenzyme CoQ treatment effects on life span of different animal models.
Table 1. Coenzyme CoQ treatment effects on life span of different animal models.
ModelStrainAge, Gender, nCoQ form (Daily Dose/Conc), Treatment DurationDiet/FoodEffect on LongevityRef
Caenorhabditis elegansN2 Bristol wild-typeEgg, n = 88–105Water-soluble CoQ10 1 (dose)Nematode growth medium (NGM) with E. coli OP50No effect[187]
L1, n = 96–98CoQ10 (50g/mL)NGM with E. coli OP50Average life span extended by 6%[188]
CoQ10 (150g/mL)NGM with E. coli OP50Average life span extended by 18%[188]
Clk-1 mutantEgg, n = 88–105Water-soluble CoQ10 1 (dose), 24 hNGM with E. coli OP50No effect[187]
Eggs, n = 100Different engineered Escherichia coli strains producing either CoQ6 to CoQ10NGM with different engineered E. coli strains or E. coli OP50 (which produce CoQ8)Median adult life span increased by 19% but only with CoQ10-producing bacteria[118]
Mev-1 (kn1) mutantL1, n = 96–98CoQ10 (50g/mL)NGM with E. coli OP50Average life span increased by 13%[132]
CoQ10 (150g/mL)NGM with E. coli OP50Average life span increased by 19%
Mouse (Mus musculus)C57BL/63.5 months old, n = 50CoQ10 (93 mg/kg of bw)ad libitum Purina diet 5001No effect[157]
CoQ10 (371 mg/kg of bw)ad libitum Purina diet 5001No effect[157]
C57/B172 m old, male, n = 43CoQ10 (10 mg/kg of bw)normal animal dietNo effect[170]
C57BL/6 C3H (B6C3F1)14 m old, male n = 60CoQ10 (100 mg/kg)AIN93 dietNo effect[169]
Rat (Rattus norvegicus)Sprague–DawleyFrom pregnancy, male, n = 75CoQ10 (10 mg/kg)normal animal dietNo effect[170]
Wistar28 d old (weaning), male, n = 43CoQ10 (2.5 mg/kg of bw)AIN93 diet but with 8% sunflower oil or virgin olive oil as unique dietary fatsMedian life span increased by 11.7%.[180]
28 d old (weaning), male, n = 22–25CoQ10 (2.5 mg/kg of bw)AIN93 diet with sunflower oil as unique dietary fat (4%)Increased median life span by 25.5%[171]
AIN93 diet with fish oil as unique dietary fat (4%)No effect[171]
AIN93 diet with virgin olive oil as unique dietary fat (4%)No effect[171]
1 A 122 mM emulsion with a mean diameter of 52 nm in 20% glycerol-fatty acid ester and 50% high-fructose corn syrup. Abbreviations: bw: body weight, CoQ: coenzyme Q, d: days, m: months, n: sample size, and h: hours.
Table 2. Studies on long-term coenzyme CoQ treatment effects on age-related changes in disease-free animal models.
Table 2. Studies on long-term coenzyme CoQ treatment effects on age-related changes in disease-free animal models.
ModelStrain, Age, SexCoQ form (Dose), DurationDiet/FoodTissues, Organs or Systems (Sample Size Per Group)Consequences in Age-Related ChangesRef
C. elegansN2 Bristol wild-type, EggWater-soluble CoQ10 1NGM containing 200 μg/mL streptomycin with E. coli OP50, 24 hNervous systemNo effect on pharyngeal contraction and defecation rate.[187]
N2 Bristol wild-type, L1CoQ10 (50 or 150 g/mL)With a lawn of E. coli OP50-Reduced O2 in the presence of succinate, although in a slight manner.[131]
Clk-1 mutant, EggWater-soluble CoQ10 1 (0.1, 1, or 10 μM)NGM containing 200 μg/mL streptomycin with E. coli OP50, 24 hNervous systemIncreased pharyngeal pumping rate and defecation rate slowed in a dose-dependent manner, with comparable values found in wild-type strains.[187]
Different engineered E. coli strains producing CoQ6 to CoQ10NGM with different engineered strains or the OP50 strain of E. coli-
(100)
Bacteria containing CoQ6, CoQ7 or CoQ10 decreased complex I-dependent respiration rates compared to those containing CoQ8 or CoQ9.
Bacteria containing CoQ7 or CoQ10 decreased complex II-dependent respiration rates compared to those fed on bacteria containing vitamin K12, CoQ6 or CoQ8.
[228]
Mouse (Mus musculus)C57BL/6, 3.5 m old, male,CoQ10 (0.072 or 0.281%, w/w), 16/22 m of agead libitum Purina diet 5001Liver, heart, skeletal muscle and brain (2–6)No effect on enzymatic antioxidant defenses (SOD, catalase and GPX activities).
No effect on protein oxidative damage (carbonyl content).
No effect on mitochondrial Reactive Oxygen Species (ROS) production.
No effect on glutathione redox state.
No effect on mitochondrial function (mtETC complexes activities).
[157]
C57BL/6NCr, 15/6 m old, male,Water-soluble CoQ10 1 (150 μM) via drinking waterStandard chow dietBrain (motor cortex) (6–9)Restored aging-associated decreases in mitochondrial function (OCR).
Restored aging-associated motor function,
phosphorylated α-synuclein and glutamate transporter 1 levels.
[229]
SAMP1, 4 wk old, male and femaleCoQ10 (0.2%, w/w), 10/12/14/16 mStandard laboratory mouse dietGeneral (9–11)No effect on senescence evaluated by the grading system by Hosokawa et al., 1984.[211]
Urine (9–11)No effect on oxidative damage (acrolein-lysine adduct and OhdG).
Brain (9–11)No effect on senile amyloid deposition rate.[211]
CoQ10 H2 (0.2%, w/w), 10/12/14/16 mGeneral (10)Slowed senescence evaluated by the grading system.[211]
Urine (10)No effect on oxidative damage (acrolein-lysine adduct and OhdG).[211]
Brain (10)No effect on senile amyloid deposition rate.[211]
SAMP1/Sku Slc, 4 wk old, femaleCoQ10 H2 (0.3%, w/w),
2/7/13/ and 19 m of age
CE-2General (11–20)Slowed degree of senescence 2.
Slowed the rate of age-related hearing loss.
[210]
Liver (11–20)Prevented age-related decreases in the expression of sirtuin gene family members and increased intracellular cyclic AMP (cAMP) levels.
Maintained mitochondrial biogenesis and oxidative metabolism by maintaining PPARγ coactivator (PGC)-1α activated.
Maintained enzymatic antioxidant defenses (SOD and isocitrate dehydrogenase [IDH]2).
Increased mitochondrial function (complex I activity).
Decreased protein, lipid and DNA oxidative damage (protein carbonyls, MDA and apurinic/apyrimidinic sites).
Increased the GSH:GSSG ratio.
[210]
Rat (Rattus novergicus)Sprague–Dawley, 14 m old, maleCoQ10 (0.324%, w/w), 13 wkNIH-31 dietBlood (8)Increased the GSH:GSSG ratio.[156]
Liver (8)No effect on protein oxidative damage (protein carbonyls).
No effect on enzymatic antioxidant defense (catalase, SOD and GPX).
[156]
Heart and BrainNo effect on lipid oxidative damage (hydroperoxides).
No effect on mitochondrial ROS (H2O2) production.
[156]
Skeletal muscleNo effect on lipid oxidative damage (hydroperoxides).
Decreased protein oxidative damage (protein carbonyls) at mitochondria.
No effect on mitochondrial ROS (H2O2) production.
No effect on enzymatic antioxidant defense (catalase, SOD and GPX).
[156]
Wistar, 28 d old (weaning) maleCoQ10 (0.005%, w/w), 6/24 m of ageAIN93 diet with sunflower oil as unique fat source (4%)Urine (6)Reduced aging-associated increase in urinary F2-isoprostanes.[174]
Pancreas (6)Improved endocrine pancreas structure, in particular β-cell mass.[175]
Bone (6)Prevented aging-associated bone mass loss decline.[174]
Alveolar bone (6)Attenuated aging-associated alveolar bone loss.[173]
Gingivae (6)Increased antioxidant enzymatic defenses (antioxidant enzyme gene expression).
Increased mitochondrial biogenesis markers.
[173]
AIN93 diet with fish oil as unique fat source (4%)PBMCsReduced DNA oxidative damage markers (DNA strand breaks) in 24-m-old rats.[182]
UrineReduced lipid oxidative damage markers (F2-isoprostanes) in 6-m-old rats.[182]
PancreasNo effect on structural alterations in exocrine compartment.[175]
BoneIncreased bone mass density in 24-m-old rats.[182]
Alveolar boneNo effect on aging-associated alveolar bone loss.[173]
GingivaeNo effect on mitochondrial biogenesis markers.[173]
AIN93 diet with virgin olive oil as unique fat source (4%)UrineNo effect on lipid oxidative damage (F2-isoprostanes) markers.[174]
PancreasNo effect on histopathological alterations.[175]
BoneNo effect on aging-associated bone mass density loss.[174]
Alveolar boneNo effect on aging-associated alveolar bone loss.[173]
GingivaeNo effect on mitochondrial biogenesis markers.
No effect on enzymatic antioxidant defense
[173]
CoQ10 (0.005%, w/w), 6/12 mAIN93 diet but with 8% of sunflower oil as fat sourceHeart (8)Attenuated an aging-associated increase in lipid oxidative damage (hydroperoxides).[176]
CoQ10 (0.062%, w/w), 6/12 m of ageAIN93 diet but with 8% of sunflower oil as fat sourceLiver (8)Decreased cytosolic and membrane-bound NQO1 activity.[184]
Brain (8)Decreased cytosolic and membrane-bound NQO1 activity[184]
AIN93 diet but with 8% of virgin olive oil as fat sourceLiver (8)Decreased cytosolic and membrane-bound NQO1 activity.[184]
Brain (8)Decreased cytosolic and membrane-bound NQO1 activity.[184]
CoQ10 (0.005%, w/w), 6/12/18/24 m of ageAIN93 diet but with 8% of sunflower oil as fat sourceBlood (8)Decreased DNA oxidative damage markers (DNA strand breaks) in PBMCs in 18- and 24-m-old rats.[180]
CoQ10 (0.005%, w/w), 6/12/24 m of ageAIN93 diet but with 8% of sunflower oil as fat sourceHeart (20)Decreased lipid oxidative damage (hydroperoxides) in 12- and 14-m-old rats.[213]
Liver (8)Prevented an aging-associated decrease in glutathione-S-transferase (GST) activity but Se-dep GPX was not clearly affected.
No effect on basal lipid oxidative damage markers (hydroperoxides) but attenuated formation against AAPH in old rats.
[183]
CoQ10 (0.005%, w/w) 6/24 m of ageAIN93 diet with 8% of sunflower oil as fat sourceBlood (20)Increased non-enzymatic antioxidant defenses (α-tocopherol and retinol) and total antioxidant capacity in aged rats.
Decreased DNA oxidative damage markers (DNA strand breaks) in PBMCs in young rats.
[179]
LiverPrevented an aging-associated increase in lipid oxidative damage markers (hydroperoxides).
Increased non-enzymatic antioxidant defenses (α-Tocopherol).
Prevented an aging-associated decrease in enzymatic antioxidant defenses (catalase activity).
[178]
Skeletal muscleIncreased non-enzymatic antioxidant defenses (α-Tocopherol) in young rats but attenuated its aging-associated increase.
Reduced lipid oxidative damage markers (hydroperoxides) at any age.
Prevented an aging-associated increase in enzymatic antioxidant defenses (catalase activity).
[178]
HeartIncreased non-enzymatic antioxidant defenses.
Partially prevented an age-associated mitochondrial function (mtETC II and III and COX were decreased).
Prevented an age-associated increase in lipid and DNA oxidative damage in mitochondria (deleted mtDNA and hydroperoxides).
Improved mitochondrial ultrastructure (area, perimeter, cristae density) in aged rats.
[181]
BrainIncreased non-enzymatic antioxidants (α-tocopherol) at mitochondria.
Decreased mitochondrial ROS production.
Decreased enzymatic antioxidant defenses (GPX content) in cytosol.
Increased mitochondrial function (mtETC complex I, IV and III activities) in young rats, but this decreased (complex I activity) in aged rats.
Decreased oxidative DNA and lipid damage markers at mitochondria (hydroperoxides and deleted mtDNA).
[178]
1 A 122 mM emulsion with a mean diameter of 52 nm in 20% glycerol-fatty acid ester and 50% high-fructose corn syrup. 2 The degree of senescence was evaluated by a grading system using eleven categories of behavioral activity and gross appearance of the skin, eyes, and spine were considered to be associated with the aging process: passivity, reactivity; glossiness, coarseness, hair loss, skin ulcers; periophthalmic lesions, corneal opacity, corneal ulcer, cataracts, and lordokyphosis. Abbreviations: AAPH: 2,2′-Azobis(2-amidinopropane) dihydrochloride, bw: body weight, cAMP: cyclic AMP, CoQ: coenzyme Q, CoQ10H2: reduced CoQ10, COX: cytochrome C oxidase, d: days, m: months, mETC: mitochondrial electron transport chain, n: sample size, h: hours, IDH2: isocitrate dehydrogenase 2, mtDNA: mitochondrial DNA, MDA: malondialdehyde, NQO1: NQO1-NAD(P)H dehydrogenase [quinone] 1 reductase, OCR: oxygen consumption rate; OhdG; 8-hydroxydeoxyguanosine; PBMCs: peripheral blood mononuclear cells, GPX: Glutathione peroxidase; GSH: reduced glutathione, GSSG: oxidized glutathione, GST: glutathione-S-transferase, ROS: reactive oxygen species, Se-dep: selenium-dependent, and SOD: superoxide dismutase.

Share and Cite

MDPI and ACS Style

Díaz-Casado, M.E.; Quiles, J.L.; Barriocanal-Casado, E.; González-García, P.; Battino, M.; López, L.C.; Varela-López, A. The Paradox of Coenzyme Q10 in Aging. Nutrients 2019, 11, 2221. https://doi.org/10.3390/nu11092221

AMA Style

Díaz-Casado ME, Quiles JL, Barriocanal-Casado E, González-García P, Battino M, López LC, Varela-López A. The Paradox of Coenzyme Q10 in Aging. Nutrients. 2019; 11(9):2221. https://doi.org/10.3390/nu11092221

Chicago/Turabian Style

Díaz-Casado, M. Elena, José L. Quiles, Eliana Barriocanal-Casado, Pilar González-García, Maurizio Battino, Luis C. López, and Alfonso Varela-López. 2019. "The Paradox of Coenzyme Q10 in Aging" Nutrients 11, no. 9: 2221. https://doi.org/10.3390/nu11092221

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop