HOME JOURNALS CONTACT

International Journal of Cancer Research

Year: 2006 | Volume: 2 | Issue: 3 | Page No.: 290-298
DOI: 10.3923/ijcr.2006.290.298
Coenzyme Q in Cancer Therapy
Zhu Liancai, Wang Bochu, Tan Jrm and Luo Min

Abstract: The presentation is a brief review of the oxidative damage mechanism of carcinogenesis, the anti-oxidative properties of Coenzyme Q (CoQ) and the expectation of CoQ applying to cancer treatment. The oxidative damage to DNA, lipid and protein has been suggested to contribute to initiation and progression of carcinogenesis. Nevertheless CoQ, known as the electon carrier in mitochondrial respiratory chain, has shown potent anti-oxidative activity, which is more efficient than vitamin E by about four mechanisms. Thought few researches have been reported on the pharmacology of CoQ on cancer or tumor, CoQ deficiency in human or animals bearing tumor has been long observed and CoQ administration has shown a benefit for cancer patients. But more researches should be launched to explore the relationship between CoQ and cancers, especially the dynamic relation and to illustrate the anti-carcinogenesis mechanisim of CoQ in order to enable a wider application of CoQ to cancer therapeutics.

Fulltext PDF Fulltext HTML

How to cite this article
Zhu Liancai, Wang Bochu, Tan Jrm and Luo Min , 2006. Coenzyme Q in Cancer Therapy. International Journal of Cancer Research, 2: 290-298.

Keywords: coenzyme Q, anti-oxidation, Oxidative damage and carcinogenesis

Oxidative Damage and Cancer

Cancer is the number one killer in diverse diseases and is significant in human mortality. WHO reported that each year over nine million cancer cases occurred and five million persons died of cancer, which would soar up to twenty million and ten million by 2020, respectively. How does cancer occur? Accompanying with oxidative phosphorylation and ATP’s production, Reactive Oxygen Species (ROS) including free radical come into being. ROS is any species capable of independent existence that contains one or more unpaired electrons. In the presence of metal ions, ROS may cause an oxidative damage by reacting with macromolecules including proteins, lipids and DNA in the cells (Halliwell and Gutterage, 1989a; Halliwell, 1991), which is called as oxidative stress. Oxidative damage has been suggested to contribute to the initiation and progression of carcinogenesis (Halliwell and Gutterage, 1989b) and causes of oxidative stress including transition metals (e.g., iron and copper), ultraviolet radiation, inflammation, some chemicals or drugs (e.g., carbon tetrachloride) are all associated with carcinogenesis or tumor biology (Halliwell and Gutterage, 1989a).

Since free radicals are usually generated near membranes, lipid is the first target of free radicals and lipid peroxidation is the first reaction to occur. It has been reported that products of lipid peroxidation may cause DNA damage (Halliwell and Gutterage, 1989b; Esterbauer, 1990; Marinari et al., 1984). Lipid hydroperoxides may directly induce DNA chain breaking (Marinari et al., 1984) and lipid peroxyl and alkoxyl radicals may cause base oxidation in DNA (Cochrane, 1991). Peroxide and hydroperoxides have also demonstrated tumor promoting activity in vivo (Park, 1992). It is known that the epoxy derivative of 4-hydroxynonenal (HNE) has a tumor causing activity (Chung et al., 1993; Sodum, 1991; Wang et al., 1996). Reaction of MDA with DNA is a hot subject, which has drawn considerable attention because of its mutagenicity. An endogenous MDA-deoxyguanosine adduct that has been implicated in the induction of G to T transversions was detected by mass spectrometry in healthy human liver (Imlay et al., 1988).

Protein oxidation usually occurs at certain amino acid residues of a particular protein. Reactive oxygen radicals react with amino residues in proteins and give rise to produce of carbonyl products, changing protein structure or converting sulphydryl (thiol) groups of proteins into disulfide groups (Butterfield et al., 1998; Dean et al., 1997). Peroxynitrite generated after the reaction between superoxide radicals and nitric oxide damages protein by binding nitro groups with protein tyrosine residues to form nitrotyrosine (Kaur and Halliwell, 1994; Halliwell, 1997). Oxidation of cellular proteins might result in the inactivity of anti-oxidative enzymes and especially DNA repair enzymes.

DNA is another major object targeted by free radicals. Oxidative damage of DNA could result in strand breaks,base modifications and DNA-protein cross-links. H2O2 is known to cause DNA breaks in intact cells and purified DNA (Imlay et al., 1988; Baker and He, 1991). Though there are a variety of modified DNA bases produced by free radical reactions, only the modifications of C-8 position of guanine has been studied in many aspects because of the sensitivity detection method of 8-oxoguanine (8-oxoG) by high performance liquid chromatography and the electrochemical detector developed in 1986 (Baker and He, 1991). It has been revealed that C-8 position of guanine is hydroxylated to produce 8-oxoGs. This hydroxylation completely changes the stereographic charge mapping of the molecule to allow guanine to pair adenine as well as guanine to pair cytosine. Accordingly, 8-oxoG induces G:C to T:A transversions in DNA replication (Kuchino et al., 1987; Shibutani et al., 1991), which appears to be important in carcinogenesis and tumor biology.

In fact, oxidative stress has been founded in human or animals bearing diversity cancer or tumor (Gibanananda et al., 2000; Ilker et al., 2003). In plasma or cancer tissue, reactive oxygen species (e.g., superoxide anion, hydrogen peroxide and hydroxyl radical) and products of lipid peroxidation and protein oxidation increase, while the activities of some antioxidant enzymes activity decrease.

Under normal physiological conditions oxidation and anti-oxidation are in equilibrium. Changes in this balance in favor of free radical formation would result in oxidative stress (Rizzo et al., 1992). Anti-oxidants are enzymes and nonenzymatic agents that eliminate ROS or prevent their formation. Antioxidant enzymes include superoxide dismutase and various peroxidases such as glutathione peroxidase, catalase, thioredoxin reductase and peroxiredoxin (Holmgren and Bjornstedt, 1995; Wood et al., 2003); vitamins C and E, carotenoids, glutathione, a-lipoic acid, flavinoids, the reduced form of CoQ (CoQH2) are recognized as nonenzymatic agents of anti-oxidative activities. Antioxidants have been used as radioprotectors, antimutagens and anticarcinogens (Ames and Gold, 1990). Vitamin E (α-tocopherol) is a biological lipid antioxidant that prevents the formation of free radicals from lipid peroxidation and has proved an antimutagen or anticarcinogen in Salmonella tester strains (Tavan et al., 1997) as well as in human leucocytes in vitro (Bolkenius, 1991). Of note here is the fact that the anti-oxidative activity of CoQ is more efficient than Vitamin E (Frei et al., 1990; Shi et al., 1999), implying that CoQ might be used as antimutagen or anticarcinogen.

What Is Coenzyme Q?
In 1957, CoQ was first isolated and purified from beef hearts by Fred Crane. Now, CoQ has been found in all cells, tissues and membranes (Dallner and Sindelar, 2000). For animals and humans, CoQ distributes in subcellular organelles, blood plasma and serum lipoproteins. Its molecular formula is shown in Fig. 1 CoQ consists of a quinonoid head group attached to a long, hydrophobic tail of 5-carbon isoprene units numbering from 6 to 12 in different species (Battino et al., 1990) (e.g., 6, 8, 9, 10 in Saccharomyces cerevisiae, Escherichia coli, rodent, human, respectively). In vivo, the quinonoid head mainly exists alternately in three different redox states: ubiquinone, the fully oxidized form; ubisemiquinone (●QH), the partially reduced form, also a free radical; and ubiquinol (CoQH2), the fully reduced form (Turrens et al., 1985; Kozlov et al., 1998). CoQ plays multiple functional roles in living cells in many aspects. Among those functions, three have been well characterized.

Fig. 1: Structure of ubiquinone, n=6,7,....,12

First, as an endogenous enzyme cofactor, CoQ is an essential component of the mitochondrial respiratory chain and adenosine triphosphate (ATP) production, carrying electrons from complexes I or II to complex III (Ernster and Dallner, 1995). Second, CoQ functions as pro-oxidant. Superoxide radicals are generated during the redox process associated with ●QH formation, which accounts for the major part of superoxide anion and hydrogen peroxide physiologically generated in the mitochondria (James et al., 2004). Third, as the only endogenous lipophilic antioxidant, ubiquinol exerts strong anti-oxidant power on plasma and cellular membranes (e.g. scavenging reactive oxygen species or lipid radicals and regenerating α-tocopherol from a-tocopheroxyl radical) (Crane, 2001; Hargreaves, 2003; Genova et al., 2003). Anthony and Hayden ( 2004) supposed on the base of these three roles that CoQ could regulate gene expression and metabolism.

The Anti-oxidation and Pro-oxidation of Coenzyme Q
Literature on the anti-oxidation of CoQ is exhaustive. Ubiquinol is a potent antioxidant and its anti-oxidative activity is munch stronger than that of α-tocopherol (a-T) (Frei et al., 1990). When phosphatidylcholine liposomes are oxidized with a water-soluble radical initiator in the presence of ascorbate, a-T and CoQ, the antioxidants are consumed in the order: Ascorbate-CoQ-a-T (Shi et al., 1999); while with a lipidsoluble radical initiator, the order is CoQ-ascorbate-a-T. It has been reported that ubiquinol inhibits lipid peroxidation in vitro using submitochondrial particles (Mellors and Tappel, 1966; Takayanagi et al., 1980). Simultaneously ubiquinone appears workable for anti-oxidation even without appropriate reducing systems (Tomasetti et al., 1999). It is now well established that CoQ prevents lipid peroxidation in most subcellular membranes (Ernster and Dallner, 1995) and acts as antioxidant also in the circulation (Romagnoli et al ., 1994; Alleva et al., 1995). Investigations suggested that the endogenous content of CoQ might prevent membrane proteins and DNA from oxidative damage mediated by lipid peroxidation (Forsmark, et al., 1995; Forsmark and Ernster, 1994). DNA oxidation measured by 8-oxoG formation in rat liver mitochondria and DNA strand breaks in human lymphocytes are also prevented by CoQ administration in vitro or in vivo (Tomasetti et al., 1999; Atroshi et al., 1997; Tomasetti et al., 2001). CoQ supplementation increases CoQ homologues in tissues and their mitochondria, decreases selectively protein oxidative damage andincreases anti-oxidative potential in rat (Kwong et al., 2002). CoQ prevents the hydrazine- and chloramphenicol-induced changes in the membrane potential of mitochondria and decreased ROS generation rate in mitochondria. Furthermore, CoQ accelerates the repair of damage induced by chloramphenicol in mitochondrial structure and functions (Teranishi et al., 1999). Various studies have reported the beneficial effects of CoQ supplementation in animal experimentation (Rauscher et al., 2001; Kwong et al ., 2002) or human therapy on different diseases related to oxidation damage (Geromel et al., 2002; Hodgson et al., 2002).

Fig. 2: The redox process of CoQ

CoQ in vivo undergoes the redox process shown in Fig. 2. Ubiquinone is reduced by ubiquinone reductasesfirst to ●QH and then to ubiquinol, each step with an additional electron and an additional proton (Rich and Harper, 1990). Because cells have effective systems to reduce ubiquinone at all intracellular locations, CoQ mainly exists as ubiquinol andthe ratio of ubiquinol and ubiquinone varies from one tissue to another (Aberg et al., 1992). The cellular pool of ubiquinol in some tissues was reported to be elevated following the external administration of ubiquinone in order to increase the antioxidant capacity (Crane et al., 1984). Various NADH-dehydrogenases associated with cell membranes are supposed to transform ubiquinone to ubiquinol (Kishi et al., 1999). Although the antioxidant activity of CoQ in systems is mainly attributed to ubiquinol, the role of ubiquinone should not be neglected. According to reported researches, the probable mechanisms of CoQ as an anti-oxidant could be as the follows: (1) ubiquinol (CoQH2) prevents the initiation and/or propagation of free radical chain reaction. CoQH2 acts by affecting the initiation process and preventing the formation of lipid peroxyl radicals (LOO) (Ernster and Forsmark, 1993) by reaction (i). It is also possible that CoQH2 eliminated LOO directly by reactions (ii) and (iii) where a CoQH2 molecule scavenges two free radicals eventually.

(i)
(ii)
(iii)

(2) CoQ spares and regenerates α-tocopherol. It is established that CoQH2 regenerates α-tocopherol from a-tocopheroxyl radical. In reactions (iv) and (v) α-tocopherol was transferred to α-tocopheroxyl radical by reacting with lipid radical andin reaction(vi) CoQH2 reacts with α-tocopheroxyl radical to produce α-tocopherol. When coexisting with α-tocopherol, it is CoQH2 that is first oxidized by radical initiator, which thus spares α-tocopherol (Shi et al., 1999). This sparing effect of CoQH2 on α-tocopherol has also been observed in low-density lipoprotein (LDL) (Stocker et al., 1991; Thomas et al., 1995). Vitamin E level does not decrease, but CoQH2 is oxidized to ubiquinone at an early stage in the oxidation of human plasma (Yamamoto et al., 1991). CoQ administration enhanced α-tocopherol level in mitochondria (Kagan et al., 1990; Maguire et al., 1992; Stoyanovsky et al ., 1995) However, it is evident that the antioxidant function of CoQH2 is not dependent on the presence of α-tocopherol. An investigation suggested that submitochondria particles containing CoQ were protected against lipid peroxidation even in the absence of α-tocopherol (Forsmark et al., 1991).

(iv)
(v)
(vi)

(3) CoQ is a cofactor for uncoupling protein (UCP) function of performing anti-oxidative activity. UCP, situated in the inner mitochondria membrane, transfers H+ from the outside to the inside of the mitochondria. UCP could be involved in suppressing the generation of oxygen radicals. Recently Echtay et al. (2000, 2001) testified that CoQ was an obligatory cofactor for UCP function by using bacterial overexpressed UCP1, -2 and -3 in liposomes. (4) CoQ stabilizes membrane structure so as to block the propagation reaction in phospholipid bilayer (Landi et al., 1987). The mechanism suggests that such antioxidant activity is related to the localization of ubiquinone within the lipid bilayer, where CoQ prevents autocatalytic free radical reaction by stacking among phospholipid molecules and keeping the quinone ring in the nonpolar phase (Fato et al., 1986). In fact, fluorescence studies have demonstrated that ubiquinone homologues possess a strong ordering effect on the lipid bilayer, which is much higher than that produced by ubiquinol (Jemola et al., 1996). The ordering effect decreases the rate of free radicals leaking from mitochondria in the cell when they consume oxygen to generate adenosine triphosphate.

On the other hand, CoQ plays key roles in physiology as pro-oxidant. During the redox process associated with CoQ semiquinone formation as shown in Fig. 2 and Reaction ii, superoxide radicals are generated. Some superoxide radicals are converted to hydrogen peroxide (H2O2) by superoxide dismutase. More should not be re-visited here. But it should be emphasized that the pro-oxidative activity of CoQ is not evil because intact mitochondria produces only a small amount of H2O2, and this the very low level of H2O2 is crucial to both cell function and the role of H2O2 as a second messenger molecule (Gille and Nohl, 2000; St- Pierre et al., 2002).

It is emphasized that CoQ pro-oxidant and anti-oxidant roles are not mutually exclusive (Anthony and Hayden, 2004), but essential to cells. On the one hand, the action of CoQ leads to the cellular bioenergy modulation through superoxide formation and the synthesis of the mitogen H2O2, which is accompanied by the formation of ROS and the damage of some lipid, protein and DNA. On the other hand, as an anti-oxidant, CoQ contributes to alleviating the oxidative damage derived from ROS.

Coenzyme Q and Cancer
CoQ has certain relation with the occurrence and development of cancer or tumor even though researches on this subject have not been bountiful. CoQ in tissue or plasm is decreased in subjects with cancer (Kozlov et al., 1998; Quiles et al., 1994). For breast cancer patients, the following conditions were observed: CoQ concentrations in tumor tissues significantly decreased compared with the surrounding normal tissues; higher MDA levels were observed in tumor tissues than noncancerous tissues; The activities of MnSOD, total SOD, GSH-Px and catalase in tumor tissues significantly increased compared with the controls. These findings may support that reactive oxygen species increased in malignant cells andcaused overexpression of antioxidant enzymes and the consumption of CoQ. Increased antioxidant activities may be related to the susceptibility of cells to carcinogenic agents and the response of tumor cells to the chemotherapeutic agents. Hyperplastic noduli is the first stage during development of chemically induced hepatocellular cancer in rat and during this stage the amount of CoQ increases (Olsson et al., 1991, 1995). Interestingly, the level of CoQ in the mitochondria is stable andthe change of CoQ concentrations is attributed to the extra-mitochondrial compartments. During the first stage of the disease, there is increasing oxidative stress, which has been suggested to induce an adaptive response, resulting in the cell protecting itself by raising the concentrations of antioxidants (Droge, 2002). Upon progression of the disease, manifest cancer develops and the amount of CoQ decreases to only 40% in human and 76% in rat (Eggens et al., 1989).

Because of the reasons presented above and CoQ preventing lipid, protein and DNA from oxidation damage, CoQ is considered to be feasible to treat cancer. Several small studies have shown a benefit for people with breast or prostate cancer. There are a number of reports on the use of CoQ10 in the treatment of cancer, specially breast cancer (Lockwood et al., 1994; 1995).

Additionally, Other findings have indicated that CoQ boosts the immune system (Folkers et al., 1993), possibly helping to limit the spread of cancerous tissue. Karl Folkers published that complete biochemistry relating to biosyntheses of CoQ and the DNA bases was a rationale for the therapy of cancer with CoQ (Folkers, 1996).

Conclusions

Attributed to its anti-oxidative activity and other intrinsic functions, CoQ has shown promise for cancer. Nevertheless, researches on pharmacology of CoQ on cancer are so rare that the anti-cancer mechanisms of CoQ and the dynamic relation between CoQ concentration in tissue or plasma and carcinogenesis are still unclear. More researches should be lunched to explore the anti-carcinogenesis properties and the mechanisms in order to enable CoQ to benefit human health widely.

Acknowlegement

Supportted by the Graduate Innovation Foundation of Chongqing University (Number:200504Y1A0120119)

REFERENCES

  • Aberg, F., E.L. Appelkvist, G. Dallner and L. Ernster, 1992. Distribution and redox state of ubiquinones in rat and human tissues. Arch. Biochem. Biophys., 295: 230-234.
    PubMed    Direct Link    


  • Alleva, R., M. Tomasetti and M. Battino, 1995. The roles of coenzyme Q10 and vitamin E on the peroxidation of human low density lipoprotein subfractions. Proc. Natl. Acad. Sci. USA., 92: 9388-9391.
    PubMed    


  • Ames, B.N. and L.S. Gold, 1990. Chemical carcinogenesis too many rodent carcinogens. Proc. Natl. Acad. Sci. USA, 87: 7772-7776.
    Direct Link    


  • Anthony, W.L. and E. Hayden, 2004. Cellular redox poise modulation: The role of coenzyme Q10, gene and metabolic regulation. Mitochondrion, 4: 779-789.
    CrossRef    


  • Atroshi, F., A. Rizzo and I. Biese, 1997. T-2 Toxin-induced DNA damage in mouse livers-the effect of pretreatment with coenzyme Q10 and α-tocopherol. Mol. Aspects Med., 18: 255-258.
    CrossRef    


  • Baker, M.A. and S. He, 1991. Elaboration of cellular DNA breaks by hydroperoxides. Free Rad. Biol. Med., 11: 563-572.
    Direct Link    


  • Battino, M., E. Ferri and A. Gorini, 1990. Natural distribution and occurrence of coenzyme Q homologues. Membr. Biochem., 9: 179-190.


  • Bolkenius, F.N., 1991. Leukocyte-mediated inactivation of α-1-proteinase inhibitor is inhibited by amino analogues of α-tocopherol. Biochem. Biophys. Acta, 1095: 23-29.


  • Butterfield, D.A., T. Koppal, B. Howard, R. Subramaniam and N. Hall et al., 1998. Structural and functional changes in proteins induced by free radical-mediated oxidative stress and protective action of the antioxidants N-tert-butyl-α-phenylnitrone and vitamin E. Ann. N. Y. Acad. Sci., 854: 448-462.
    CrossRef    PubMed    Direct Link    


  • Chung, F.L., H.J.C. Chen and J.B. Guttenplan, 1993. 2,3-Epoxy-4hydroxynonanal as a potential tumor-initiating agent of lipid peroxidaion. Carcinogenesis, 14: 2073-2077.


  • Cochrane, C.G., 1991. Cellular injury by oxidants. Am. J. Med., 91: 23-30.
    CrossRef    


  • Crane, F.L., I.L. Sun and R. Barr, 1984. Coenzyme Q in golgi apparatus membrane redox activity and proton uptake. Amsterdam Elsevier, 4: 77-86.


  • Dallner, G. and P. Sindelar, 2000. Regulation of ubiquinone metabolism. Free Radical Biol. Med., 29: 285-294.


  • Dean, R.T., S. Fu, R. Stocker and M.J. Davies, 1997. Biochemistry and pathology of radical-mediated protein oxidation. Biochem. J., 324: 1-18.
    CrossRef    PubMed    Direct Link    


  • Droge, W., 2002. Free radicals in the physiological control of cell function. Phsyol. Rev., 82: 47-95.
    CrossRef    PubMed    Direct Link    


  • Echtay, K.S., E. Winkler and M. Klingenberg, 2000. Coenzyme Q is an obligatory cofactor for uncoupling protein function. Nature, 408: 609-613.
    CrossRef    


  • Echtay, K.S., E. Winkler and K. Frischmuth, 2001. Uncoupling proteins 2 and 3 are highly active H+ transporters and highly nucleotide sensitive when activated by coenzyme. Proc. Natl. Acad. Sci. USA., 98: 1416-1421.


  • Ernster, L. and A.P. Forsmark, 1993. Ubiquinol an endogenous antioxidant in aerobic organisms. Clin. Invest., 71: 60-65.


  • Ernster, L. and G. Dallner, 1995. Biochemical, physiological and medical aspects of ubiquinone function. Biochimica Biophysica Acta (BBA)-Mol. Basis Dis., 1271: 195-204.
    CrossRef    Direct Link    


  • Esterbauer, H., 1990. Possible mutagens derived from lipids and lipid procursors. Mutat. Res., 238: 223-233.


  • Fato, R., M. Battino and E.M. Degli, 1986. Determination of partition and lateral diffusion coefficients of ubiquinones by fluorescence quenching of n-(9-anthroyloxy) stearic acids in phospholipid vesicles and mitochondrial membranes. Biochemistry, 25: 3378-3390.


  • Folkers, K., M. Morita and J.J. McRee, 1993. The activities of coenzyme Q10 and vitamin B6 for immune responses. Biochem. Biophys. Res. Commun., 193: 88-92.
    CrossRef    Direct Link    


  • Folkers, K., 1996. Relevance of the biosynthesis of coenzyme Q10 and of the four bases of DNA as a rationale for the molecular causes of cancer and a therapy. Biochem. Biophys. Res. Commun., 224: 358-361.


  • Forsmark, A.P., F. Aberg and B. Norling, 1991. Inhibition of lipid peroxidation by ubiquinol in submitochondrial particles in the absence of vitamin E. FEBS Lett., 285: 39-43.


  • Forsmark, A.P. and L. Ernster, 1994. Evidence for a protective effect of endogenous ubiquinol against oxidative damage to mitochondrial protein and DNA during lipid peroxidation. Mol. Aspects Med., 15: S73-S81.


  • Forsmark, A.P., G. Dallner and L. Ernster, 1995. Endogenous ubiquinol prevents protein modification accompanying lipid peroxidation in heart submitochondrial particles. Free. Radical Biol. Med., 9: 749-756.


  • Frei, B., M.C. Kim and B.N. Ames, 1990. Ubiquinol-10 is an effective lipidsoluble antioxidant at physiological concentrations. Proc. Natl. Acad. Sci. USA., 87: 4879-4883.


  • Genova, M.L., M.M. Pich and A. Biondi, 2003. Mitochondrial production of oxygen radical species and the role of coenzyme Q as an antioxidant. Exp. Biol. Med., 228: 506-513.


  • Geromel, V., N. Darin and D. Chretien, 2002. Coenzyme Q10 and idebenone in the therapy of respiratory chain diseases: Rationale and comparative benefits. Mole. Gene Metab., 77: 21-30.


  • Gibanananda, R., S. Batra and N.K. Shukla, 2000. Lipid peroxidation free radical production and antioxidant status in breast cancer. Breast Cancer Res. Treat., 59: 163-170.


  • Gille, L. and H. Nohl, 2000. Where do the electrons go. Nature, 401: 30-31.


  • Halliwell, B. and J.M.C. Gutteridge, 1990. Role of free radicals and catalytic metal ions in human disease: An overview. Methods Enzymol., 186: 1-85.
    CrossRef    Direct Link    


  • Halliwell, B. and J.M.C. Gutteridge, 1989. Free Radicals in Biology and Medicine. Clarendon Press, Oxford, pp: 33-46


  • Halliwell, B., 1991. Reactive oxygen species in living systems source biochemistry and role in human disease. Am. J. Med., 91: 14-21.
    Direct Link    


  • Halliwell, B., 1997. What nitrates tyrosine Is nitrotyrosine specific as a biomarker of peroxynitrite formation in vivo. FEBS Lett., 411: 157-160.


  • Hargreaves, I.P., 2003. Ubiquinone cholesterols reclusive cousin. Ann. Clin. Biochem., 40: 207-218.


  • Hodgson, J.M., G.F. Watts, D.A. Playford, V. Burke and K.D. Croft, 2002. Coenzyme Q10 improves blood pressure and glycaemic control: A controlled trial in subjects with type 2 diabetes. Eur. J. Clin. Nutr., 56: 1137-1142.
    PubMed    Direct Link    


  • Holmgren, A. and M. Bjornstedt, 1995. Thioredoxin and thioredoxin reductase. Methods Enzymol., 252: 199-208.
    Direct Link    


  • Ilker, A.Y., T. Akcay and U. Cakatay, 2003. Relation between bladder cancer and protein oxidation. Int. Urol. Nephrol., 35: 345-350.
    Direct Link    


  • Imlay, J.A., S.M. Chin and S. Linn, 1988. Toxic DNA damages by hydrogen peroxide through the Fanton reaction in vivo and in vitro. Science, 240: 640-642.
    CrossRef    


  • James, A.M., R.A.J. Smith and M.P. Murphy, 2004. Antioxidant and prooxidant properties of mitochondrial Coenzyme Q. Arch. Biochem. Biophys., 423: 47-56.


  • Jemola, R.M., J. Kruk and M. Skowronek, 1996. Location of ubiquinone homologues in liposome membranes studied by fluorescence anisotropy of diphenyl-hexatriene and trymethylammonium-diphenyl-hexatriene. Chem. Phys. Lipids, 79: 55-63.


  • Kagan, V., E. Serbinova and L. Packer, 1990. Antioxidant effects of ubiquinones in microsomes and mitochondria are mediated by tocopherol recycling. Biochem. Biophys. Res. Commun., 169: 851-857.
    Direct Link    


  • Kaur, H. and B. Halliwell, 1994. Evidence for nitric oxide-mediated oxidative damage in chronic inflammation nitrotyrosine in serum and synovial fluid from rheumatoid patients. FEBS Lett., 350: 9-12.
    Direct Link    


  • Kishi, T., T. Takahashi and A. Usui, 1999. Cytosolic Nadph-UQ reductase the enzyme responsible for cellular ubiquinone redox cycle as an endogenous antioxidant in the rat liver. Biofactors, 9: 189-197.
    Direct Link    


  • Kozlov, A.V., H. Nohl and L. Gille, 1998. Are reduced ubiquinones oxygen radical generators. Bioorg. Chem., 26: 334-344.
    CrossRef    


  • Kuchino, Y., F. Mori and H. Kasai, 1987. DNA templates containing 8-hydroxydeoxyguanosine are misread both at the modified base and at adjacent residues. Nature, 327: 77-79.


  • Kwong, L., S. Kamzalov, I. Rebrin, A.C.V. Bayne and C.K. Jana et al., 2002. Effects of coenzyme Q10 administration on its tissue concentrations mitochondrial oxidant generation and oxidative stress in the rat. Free Radical Biol. Med., 33: 627-638.
    CrossRef    


  • Landi, L., P. Pasquali and P. Bassi, 1987. Effect of oxygen free radicals on ubiquinone in aqueous solution and phospholipid vesicles. Biochem. Biophys. Acta, 902: 200-206.
    Direct Link    


  • Lockwood, K., S. Moesgaard and T. Hanioka, 1994. Apparent partial remission of breast cancer in high risk patients supplemented with nutritional antioxidants, essential fatty acids and coenzyme Q10. Mol. Aspects Med., 15: s231-s231.


  • Lockwood, K., S. Moesgaard and T. Yamamoto, 1995. Progress on therapy of breast cancer with vitamin Q10 and the regression of metastases. Biochem. Biophys. Res. Commun., 212: 172-177.


  • Maguire, J.J., V. Kagan and A.C.B. Ackrell, 1992. Succinate-ubiquinone reductase linked recycling of α-tocopherol in reconstituted systems and mitochondria requirement for reduced ubiquinone. Arch. Biochem. Biophys., 292: 47-53.


  • Marinari, U.M., M. Ferro and L. Sciaba, 1984. DNA-damagingactivity of biotic and xenobiotic aldehydes in Chinese hamster ovary cells. Cell Biochem. Funct., 2: 243-248.


  • Mellors, A. and A.L. Tappel, 1966. The inhibition of mitochondrial peroxidation by ubiquinone and ubiquinol. J. Biol. Chem., 241: 4353-4356.
    Direct Link    


  • Olsson, J.M., L.C. Eriksson and G. Dallner, 1991. Lipid compositions of intracellular membranes isolated from rat liver nodules in Wistar rats. Cancer Res., 51: 3774-3780.


  • Olsson, J.M., S. Schedin and H. Teclebrhan, 1995. Enzymes of the mevalonate pathway in rat liver nodules induced by 2-acetylaminofluorene treatment. Carcinogenesis, 16: 599-605.


  • Park, D., 1992. Peroxyl and alxocyl radicals cause DNA base modifications. Cancer Lett., 28: 1235-1240.


  • Quiles, J.L., J.R. Huertas and M. Mana, 1994. Mataix peroxidative extent and coenzyme Q levels in the rat influence of physical training and dietary fats. Mol. Aspects Med., 15: 89-95.


  • Rauscher, F.M., R.A. Sanders and J.B. Watkins, 2001. Effects of coenzyme Q10 treatment on antioxidant pathways in normal and streptozotocin-induced diabetic rats. J. Biochem. Mol. Toxicol., 15: 41-46.


  • Rich, P.R. and R. Harper, 1990. Partition coefficients of quinones and hydroquinones and their relation to biochemical reactivity. FEBS Lett., 269: 139-144.


  • Rizzo, A., F. Atroshi and T. Hirvi, 1992. The hemolytic activity of deoxynivalenol and T-2 toxin. Nat. Toxins, 1: 106-110.


  • Romagnoli, A., A. Oradei, C.A. Destito, I. Marin and G.P. Littarru, 1994. Protective role in vivo of coenzyme Q10 during reperfusion of ischemic limbs. Mol. Aspects Med., 15: 177-185.
    PubMed    


  • Shi, H., N. Noguchi and E. Niki, 1999. Comparative study on dynamics of antioxidative action of alpha-tocopheryl hydroquinone, ubiquinol and α-tocopherol against lipid peroxidation. Free Radical Biol. Med., 27: 334-346.


  • Shibutani, S., M. Takeshita and A.P. Grollman, 1991. Insertion of specific bases during DNA synthesis past the oxidation-damaged base 8-oxodG. Nature, 349: 431-434.
    CrossRef    


  • Sodum, R.S., 1991. Steroselective formation of in vitro nucleic acid adducts by 2,3 epoxy-4-hidroxynonenal. Cancer Res., 51: 137-143.


  • St-Pierre, J., J.A. Buckingham S.J. Roebuck and M.D. Brand, 2002. Topology of superoxide production from different sites in the mitochondrial electron transport chain. J. Biol. Chem., 277: 44784-44790.
    CrossRef    Direct Link    


  • Stocker, R., V.W. Bowry and B. Frei, 1991. Ubiquinol-10 protects human low density lipoprotein more efficiently against lipid peroxidation than does α-tocopherol. Proc. Natl. Acad. Sci. USA., 88: 1646-1650.
    Direct Link    


  • Stoyanovsky, D.A., A.N. Osipov, P.J. Quinn and V.E. Kagan, 1995. Ubiquinone-dependent recycling of vitamin E radicals by superoxide. Arch. Biochem. Biophys., 323: 343-351.
    Direct Link    


  • Takayanagi, R., K. Takeshige and S. Minakami, 1980. Nadh-and Nadph dependent lipid peroxidation in bovine heart submitochondrial particles. Dependence on the rate of electron flow in the respiratory chain and an antioxidant role of ubiquinol. J. Biochem., 192: 853-860.


  • Gulkac, M.D., G. Akpinar, H. Ustun and A.O. Kanli, 2004. Effects of vitamin A on doxorubicin-induced chromosomal aberrations in bone marrow cells of rats. Mutagenesis, 19: 231-236.
    Direct Link    


  • Teranishi, M., M. Karbowski C. Kurono Y. Nishizawa, J. Usukura, T. Soji and T. Wakabayashi, 1999. Effects of coenzyme Q10 on changes in the membrane potential and rate of generation of reactive oxygen species in hydrazine-and chloramphenicol-treated rat liver mitochondria. Arch. Biochem. Biophys., 366: 157-167.
    Direct Link    


  • Thomas, S.R., J. Neuzil, D. Mohr and R. Stocker, 1995. Coantioxidants make α-tocopherol an efficient antioxidant for low-density lipoprotein. Am. J. Clin. Nutr., 62: 1357-1364.
    Direct Link    


  • Tomasetti, M., G.P. Littarru, R. Stocker and R. Alleva, 1999. Coenzyme Q10 enrichment decreases oxidative DNA damage in human lymphocytes. Free Radical Biol. Med., 27: 1027-1032.
    Direct Link    


  • Tomasetti, M., R. Alleva and B. Borghi, 2001. In vivo supplementation with coenzyme Q10 enhances the recovery of human lymphocytes from oxidative DNA damage. J. FASEB, 15: 1425-1427.


  • Turrens, J.F., A. Alexandre and A.L. Lehninger, 1985. Ubisemiquinone is the electron donor for superoxide formation by complex III of heart mitochondria. Arch. Biochem. Biophys., 237: 408-414.
    PubMed    


  • Wang, M., K. Dhingra and W.N. Hittelman, 1996. Lipid peroxidation-induced putative malondialdehyde-DNA adducts in human breast cancer tissue. Cancer Epi. Biom. Prev., 5: 705-710.
    Direct Link    


  • Wood, Z.A., L.B. Poole and P.A. Karplus, 2003. Peroxiredoxin evolution and the regulation of hydrogen peroxide signaling. Science, 300: 650-653.
    CrossRef    


  • Yamamoto, Y., M. Kawamura and K. Tatsuno, 1991. Formation of Lipid Hydroperoxides in the Cupric Ion-induced Oxidation of Plasma and Low Density Lipoprotein. Oxidative Damage and Repair. 1st Edn., Pergamon Press, London, pp: 287-291


  • Crane, F.L., 2001. Biochemical functions of coenzyme Q10. J. Am. Coll. Nutr., 20: 591-598.
    CrossRef    Direct Link    


  • Eggens, I., P.G. Elmberger and P. Low, 1989. Polyisoprenoid cholesterol and ubiquinone levels in human hepatocellular carcinomas. Br. J. Exp. Pathol., 70: 83-92.

  • © Science Alert. All Rights Reserved