Next Article in Journal
Long-Term Vaptan Treatment of Idiopathic SIADH in an Octogenarian
Next Article in Special Issue
The Relationship between Mitochondrial Respiratory Chain Activities in Muscle and Metabolites in Plasma and Urine: A Retrospective Study
Previous Article in Journal
Portal Vein Embolization: Impact of Chemotherapy and Genetic Mutations
Previous Article in Special Issue
Mitochondrial Modification Techniques and Ethical Issues
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Biochemical Assessment of Coenzyme Q10 Deficiency

by
Juan Carlos Rodríguez-Aguilera
1,2,
Ana Belén Cortés
1,2,
Daniel J. M. Fernández-Ayala
2,3 and
Plácido Navas
2,3,*
1
Laboratorio de Fisiopatología Celular y Bioenergética, 41013 Sevilla, Spain
2
Centro de Investigación Biomédica en Red de Enfermedades Raras, Instituto de Salud Carlos III, Universidad Pablo de Olavide-CISC, 41013 Sevilla, Spain
3
Centro Andaluz de Biología del Desarrollo, 41013 Sevilla, Spain
*
Author to whom correspondence should be addressed.
J. Clin. Med. 2017, 6(3), 27; https://doi.org/10.3390/jcm6030027
Submission received: 18 January 2017 / Revised: 25 February 2017 / Accepted: 28 February 2017 / Published: 5 March 2017

Abstract

:
Coenzyme Q10 (CoQ10) deficiency syndrome includes clinically heterogeneous mitochondrial diseases that show a variety of severe and debilitating symptoms. A multiprotein complex encoded by nuclear genes carries out CoQ10 biosynthesis. Mutations in any of these genes are responsible for the primary CoQ10 deficiency, but there are also different conditions that induce secondary CoQ10 deficiency including mitochondrial DNA (mtDNA) depletion and mutations in genes involved in the fatty acid β-oxidation pathway. The diagnosis of CoQ10 deficiencies is determined by the decrease of its content in skeletal muscle and/or dermal skin fibroblasts. Dietary CoQ10 supplementation is the only available treatment for these deficiencies that require a rapid and distinct diagnosis. Here we review methods for determining CoQ10 content by HPLC separation and identification using alternative approaches including electrochemical detection and mass spectrometry. Also, we review procedures to determine the CoQ10 biosynthesis rate using labeled precursors.

1. Introduction

The mitochondrial respiratory chain (MRC) generates most of the cellular ATP and is comprised of five multi-subunit enzyme complexes. Both the mitochondrial DNA (mtDNA) and the nuclear DNA (nDNA) encode for polypeptides of these complexes and also proteins involved in mitochondrial function. Besides MRC enzyme complexes, two electron carriers, coenzyme Q (CoQ) and cytochrome c, are vital for mitochondrial synthesis of ATP. Mutations in genes of either genome may cause mitochondrial diseases, which are common among inherited metabolic and neurological disorders [1].
CoQ is a lipid-soluble component of virtually all cell membranes. It is composed of a benzoquinone ring with a polyprenyl side chain, the number of isoprene units being a characteristic of given specie, e.g., 10 in humans (CoQ10). CoQ10 transports electrons from MRC Complexes I and II to Complex III. These electrons come from either NADH or succinate [2] although CoQ10 can be alternatively reduced with electrons provided by different redox reactions in mitochondria [3]. Consequently, CoQ10 is essential for ATP production inside mitochondria, although it is also an indispensible antioxidant in extramitochondrial membranes and a key factor for pyrimidine nucleotide synthesis [4].
CoQ biosynthesis depends on a pathway that involves at least 11 genes (COQ genes), showing a high degree of conservation among species, and is carried out by a putative multi-subunit enzyme complex [5]. Most of the information about the CoQ biosynthesis pathway comes from yeast, and maintains a high homology with mammal gene components (Table 1) [6]. The CoQ10 biosynthesis pathway is highly regulated by transcription factors PPARα and NFκB [7,8,9]. HuR and hnRNP C1/C2 binding proteins stabilize COQ7 mRNA as another CoQ10 biosynthesis regulatory mechanism [10].
Coq7p is post-translationally regulated in yeast that involves mitochondrial phosphatase Ptc7 [11,12]. Ptc7 human orthologue (PPTC7) is related to cellular bioenergetics and stress resistance [13]. Coq7p activity is a key regulator of the CoQ biosynthesis complex [6,14], which may depend on the interaction with Coq9p contributing to the stabilization of the biosynthesis complex [15,16,17,18]. The level of CoQ is highly regulated inside cells and tissues but its concentration is different in each tissue and organ, and depends on dietary conditions and age [19,20]. CoQ also varies greatly in human diseases such as Alzheimer’s disease, cardiomyopathy, Niemann-Pick and diabetes.

2. CoQ10 Deficiency Syndrome

CoQ10 deficiency syndrome includes diverse inherited pathological diseases defined by the decrease of CoQ10 content in muscle and/or cultured skin fibroblasts. CoQ10 deficiency impairs oxidative phosphorylation and causes clinically heterogeneous mitochondrial diseases [21,22]. When the decrease in CoQ10 content is due to mutations in genes encoding proteins of the CoQ biosynthesis pathway or its regulation (COQ genes), it causes primary CoQ10 deficiency [23,24]. Secondary CoQ10 deficiencies may be due to defects in genes unrelated to the CoQ10 biosynthetic pathway. Secondary CoQ10 deficiency is a common finding in oxidative phosphorylation (OXPHOS) and non-OXPHOS disorders [25]. A low mitochondrial CoQ10 content is described in mtDNA depletion [26], mutations in the DNA repairing aprataxin [27], mutations of the enzyme ETFDH of the β-oxidation of fatty acids [28], recurrent food intolerance and allergies [29], methylmalonic aciduria [30], myalgic encephalomyelitis chronic fatigue syndrome [31], and propionic acidemia [32]. We propose that cases of secondary CoQ10 deficiency associated with OXPHOS defects could be adaptive mechanisms to maintain a balanced OXPHOS which is required to keep cells alive, although the mechanisms explaining these deficiencies and the pathophysiological role in the disease are unknown.
The clinical phenotypes of primary CoQ10-deficient patients are broader than initially reported in 1989 [33], including (i) a multisystem disorder with steroid-resistant nephrotic syndrome as the main clinical manifestation (COQ1-PDSS2) [34], (COQ2) [35], (COQ6) [36] and (ADCK4) [37]; (ii) a multisystem disorder without nephrotic syndrome (COQ1-PDSS1) [38], (COQ9) [39] and (COQ7) [40]; (iii) cerebellar ataxia (COQ8-ADCK3) [41,42,43,44,45,46,47]; and (iv) myopathy and encephalopathy (COQ4) [48,49,50].

3. Primary CoQ10 Deficiency Therapy

Primary CoQ10 deficiency is unique among mitochondrial diseases because an effective therapy is available for patients, which is the supplementation of CoQ10. Ubiquinol, the reduced form of CoQ10, was recently approved as an orphan drug for primary CoQ10 deficiency [51]
While this approach is quite successful in some patients, with a clear improvement of the pathological phenotype [52], some cases do not show any clinical relief as would be expected [53], probably because they are suffering secondary CoQ10 deficiency. High-dose oral CoQ10 supplementation can stop the progression of the encephalopathy and allows the recovery of renal damage [52]. High-dose CoQ10 supplementation was also able to prevent the onset of renal symptoms in PDSS2-deficient mice [54]. Furthermore, CoQ10 but not other quinones can restore mitochondrial function in deficient human fibroblasts [55]. Due to the therapeutic possibility of CoQ10 supplementation for these patients, a rapid and unequivocal diagnosis of the deficiency is essential.

4. CoQ10 Determination in Cells and Tissues

Content of CoQ10 has been determined in plasma, white blood cells, skin fibroblasts and skeletal muscle biopsies to assess a deficiency diagnosis [56,57,58], and recently useful determination in the urine of pediatric patients was demonstrated [59]. Although CoQ can be measured in plasma and white blood cells, you cannot use it for the diagnosis of mitochondrial diseases since CoQ content in plasma and white blood cells is often not decreased in these conditions.
CoQ10 content is mainly analyzed by the injection of lipid extracts in HPLC and detected by either electrochemical and/or UV-vis detectors, or mass spectrometry. Electrochemical detection has significant advantages compared to UV-vis detection; these include higher sensitivity and also the ability to measure oxidized and reduced forms of CoQ, either separately or combined, according to differential positioning of the conditioning cell (before or after the injector valve, respectively).
CoQ10 extraction from biological samples (0.5 mg protein) requires the disruption of hydrophobic elements (lipid bilayers and lipoproteins) by adding SDS (1% final concentration). Lipids are dispersed with an alcohol cocktail (2-propanol 5% in ethanol) mixed with the disrupted biological sample (ratio 1:2 v/v), and they undergo subsequent triplicated hexane extraction (dispersed sample:hexane ratio 3:5 v/v). Hexane fractions are mixed and dried under vacuum, and then reconstituted in ethanol prior to HPLC analysis. To estimate CoQ10 recovery, 100 pmol CoQ9 was included in the alcohol cocktail (2-propanol 5% in ethanol). Trace amounts of CoQ9 may have eventually been found in human tissues (probably from dietary uptake), but this does not interfere with the significant amount of internal standard added.
For convenience in high-throughput analysis, volumes are scaled down for extraction and vortex in 1.5 mL polypropylene tubes or 2 mL cryo vials.
Separation in C18 RP-HPLC columns (5 µm, 150 × 4.6 mm) requires 20 mM AcNH4 pH 4.4 in methanol (solvent A) and 20 mM AcNH4 pH 4.4 in propanol (solvent B). A gradient method with a 85:15 solvent mixture (A:B ratio), and a flow rate of 1.2 mL/min, is regularly used as the starting conditions. The mobile phase turns to a 50:50 A:B ratio starting in minute 6 and completed in minute 8, as the flow rate decreases to 1.1 mL/min. After 20 min (run time) at 40 °C, the columns are re-equilibrated to the initial conditions for three additional minutes.
The detection of total CoQ10 can be achieved either by UV-vis (set to 275 nm) or electrochemical (ECD) detectors (channel 1 set to −700 mV and channel 2 set to +500 mV, conditioning guard cell after injection valve). For complex samples including many peaks, the CoQ10 peak is confirmed by spectral information (UV-vis) or by the redox area ratio (ECD detector, −700/+500 area ratio), compared to pure CoQ10. Figure 1 illustrates two chromatograms that correspond to normal age-matched human dermal fibroblasts (black plot) compared to patient dermal fibroblasts with CoQ10 deficiency (red plot).

5. Analysis of CoQ10 Biosynthesis

Another important approach to assess CoQ10 deficiency in cells is to determine the rate of biosynthesis by the level of incorporation of labeled of CoQ10 precursors such as para-hydroxybenzoate (p-HB) labeled with either 13C-p-HB or 14C-p-HB, which is the precursor of the benzoquinone ring, or 2H-mevalonate, which is the precursor of the isoprenyl side chain [10,60].
Polyprenyl-pHB transferase activity was assayed by measuring the incorporation of 14C-p-HB into nonaprenyl-4-hydroxybenzoate [35]. Isolated mitochondria (0.1–1 mg protein) were mixed with assay buffer (50 mM phosphate buffer, pH 7.5, 10 mM MgCl2, 5 mM EGTA containing 1 mM PMSF, 20 μg/mL each of the protease inhibitors chymostatin, leupeptin, antipain, and pepstatin A, 5 μM solanesyl pyrophosphate solubilized in detergent solution (1% in water), and 105 DPM of 14C-p-HB). A sufficient volume of a 10% detergent stock solution was also added to the reaction medium to achieve a final detergent concentration of 1%. The following detergents were tested: Triton X-100, Chaps, sodium cholate, sodium deoxycholate, lysophosphatidyl choline, and octylglucoside. After incubation for 30 min at 37 °C with gentle stirring, the reaction was stopped by chilling samples to 4 °C. Prenylated 14C-p-HB was separated by organic extraction with hexane and then measured using a liquid scintillation counter. Specific activity was expressed as disintegrations per minute (DPM) min−1·mg·protein−1.
Biosynthesis of 14C-CoQ10 has been quantified in any type of cell culture, such as cancer cells, human skin fibroblasts, and murine embryonic fibroblast and stem cells [10,61]. Previously, cultures were incubated with 4.5 nM 14C-p-HB for one to three days, depending on the cell-specific rate of growth. The 14C-p-HB was chemically synthesized in our laboratory from 14C-thyrosine [61]. Labeled-CoQ10 content is analyzed by lipid extract injection in HPLC and detected by the radio-flow detector LB 509 with a solid cell YG 150 Al-U4D (Berthold Technologies, Bad Wildbad, Germany) in parallel with either electrochemical or UV-vis detectors. Lipid extraction is done as we described above for CoQ10 determination in cells and tissues, but isocratic HPLC analysis lipid separation is performed with methanol:propanol (65:35) plus 20 mM AcNH4 pH 4.4 at a constant flow rate of 1 mL/min (Figure 2).
Alternatively, a non-radioactive protocol to analyze CoQ10 biosynthesis was developed using either 2H-mevalonate or 13C-phydroxybenzoate as CoQ10 precursors as described by Buján et al. (2014) [60]. Human fibroblasts at 60%–70% were incubated with these precursors for 24–72 h at different concentrations. After incubation, cells were trypsinized and washed twice with isotonic buffer. Pelleted cells were resuspended with 300 μL of a buffer solution containing 0.25 mmol/L sucrose, 2 mmol/L EDTA, 10 mmol/L Tris and 100 UI/mL heparin, pH 7.4, and sonicated twice for 5 s. These homogenates were used to determine CoQ10 biosynthesis measuring by HPLC-MS/MS, as described in Arias et al. (2012) [62]. Briefly, HPLC separation was as indicated above and extracted peaks were analyzed by MS/MS in a Micromass Quattro micro™ (Waters/Micromass, Manchester, UK). The MS/MS was operated in the electrospray positive ion mode with a cone voltage (CV), and collision energy (CE) of 15 V and 20 eV, respectively. The following multiple-reaction monitoring transitions were selected: m/z 900 > 203 and 897>197 for 13C-CoQ10 or 2H-CoQ10, respectively, 894 > 197 for the physiological CoQ10 and 826 > 197 for CoQ9 (internal standard). The dwell time for each transition was 200 ms and the run-time was 16 min. Nitrogen (at a flow rate of 50 L/h) and argon (adjusted to obtain a vacuum of 3°—10−3 bar) were used as the nebulizing and collision gas, respectively.

6. Concluding Remarks

Coenzyme Q10 deficiency syndrome includes a group of mitochondrial diseases showing diverse inherited pathological phenotypes. The common aspect of them is the lower content of CoQ10 in tissues and organs. Primary deficiency is caused by defects in proteins encoded by COQ genes, which are components of the biosynthesis pathway or its regulation. CoQ10 supplementation is the current treatment of primary CoQ10 deficiency, which highly improves symptoms. A rapid and distinct characterization of the deficiency is important, and it is mainly determined in skeletal muscle and/or skin dermal fibroblasts. The main approach is to analyze the total content of CoQ10 in lipid extracts by HPLC and UV and/or electrochemical detection. Alternatively, the CoQ10 biosynthesis rate in cultured cells can be determined by incubation with radiolabeled precursors.

Acknowledgments

This work has been funded by the Instituto de Salud Carlos III FIS PI14-01962 grant. Authors were also funded by the Junta de Andalucía BIO177 research group.

Author Contributions

J.C.R.-A., A.B.C., D.J.M.F.-A., and P.N. have contributed to writing and editing the review and figures.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Gorman, G.S.; Chinnery, P.F.; DiMauro, S.; Hirano, M.; Koga, Y.; McFarland, R.; Suomalainen, A.; Thorburn, D.R.; Zeviani, M.; Turnbull, D.M. Mitochondrial diseases, Nature reviews. Dis. Prim. 2016, 2, 16080. [Google Scholar] [CrossRef] [PubMed]
  2. Lapuente-Brun, E.; Moreno-Loshuertos, R.; Acin-Perez, R.; Latorre-Pellicer, A.; Colas, C.; Balsa, E.; Perales-Clemente, E.; Quiros, P.M.; Calvo, E.; Rodriguez-Hernandez, M.A.; et al. Supercomplex assembly determines electron flux in the mitochondrial electron transport chain. Science 2013, 340, 1567–1570. [Google Scholar] [CrossRef] [PubMed]
  3. Alcazar-Fabra, M.; Navas, P.; Brea-Calvo, G. Coenzyme Q biosynthesis and its role in the respiratory chain structure. Biochim. Biophys. Acta 2016, 1857, 1073–1078. [Google Scholar] [CrossRef] [PubMed]
  4. Lopez-Lluch, G.; Rodriguez-Aguilera, J.C.; Santos-Ocana, C.; Navas, P. Is coenzyme Q a key factor in aging? Mech. Ageing Dev. 2010, 131, 225–235. [Google Scholar] [CrossRef] [PubMed]
  5. Bentinger, M.; Tekle, M.; Dallner, G. Coenzyme Q—Biosynthesis and functions. Biochem. Biophys. Res. Commun. 2010, 396, 74–79. [Google Scholar] [CrossRef] [PubMed]
  6. Gonzalez-Mariscal, I.; Garcia-Teston, E.; Padilla, S.; Martin-Montalvo, A.; Viciana, T.P.; Vazquez-Fonseca, L.; Dominguez, P.G.; Santos-Ocana, C. The regulation of coenzyme q biosynthesis in eukaryotic cells: All that yeast can tell us. Mol. Syndromol. 2014, 5, 107–118. [Google Scholar] [CrossRef] [PubMed]
  7. Turunen, M.; Peters, J.M.; Gonzalez, F.J.; Schedin, S.; Dallner, G. Influence of peroxisome proliferator-activated receptor alpha on ubiquinone biosynthesis. J. Mol. Biol. 2000, 297, 607–614. [Google Scholar] [CrossRef] [PubMed]
  8. Bentinger, M.; Tekle, M.; Brismar, K.; Chojnacki, T.; Swiezewska, E.; Dallner, G. Stimulation of coenzyme Q synthesis. Biofactors 2008, 32, 99–111. [Google Scholar] [CrossRef] [PubMed]
  9. Brea-Calvo, G.; Siendones, E.; Sanchez-Alcazar, J.A.; de Cabo, R.; Navas, P. Cell survival from chemotherapy depends on NF-kappaB transcriptional up-regulation of coenzyme Q biosynthesis. PLoS ONE 2009, 4, e5301. [Google Scholar] [CrossRef] [PubMed]
  10. Cascajo, M.V.; Abdelmohsen, K.; Noh, J.H.; Fernandez-Ayala, D.J.; Willers, I.M.; Brea, G.; Lopez-Lluch, G.; Valenzuela-Villatoro, M.; Cuezva, J.M.; Gorospe, M.; et al. RNA-binding proteins regulate cell respiration and coenzyme Q biosynthesis by post-transcriptional regulation of COQ7. RNA Biol. 2016, 13, 622–634. [Google Scholar] [CrossRef] [PubMed]
  11. Martin-Montalvo, A.; Gonzalez-Mariscal, I.; Padilla, S.; Ballesteros, M.; Brautigan, D.L.; Navas, P.; Santos-Ocana, C. Respiratory-induced coenzyme Q biosynthesis is regulated by a phosphorylation cycle of Cat5p/Coq7p. Biochem. J. 2011, 440, 107–114. [Google Scholar] [CrossRef] [PubMed]
  12. Martín-Montalvo, A.; González-Mariscal, I.; Pomares-Viciana, T.; Padilla-López, S.; Ballesteros, M.; Vazquez-Fonseca, L.; Gandolfo, P.; Brautigan, D.L.; Navas, P.; Santos-Ocaña, C. The phosphatase Ptc7 induces coenzyme Q biosynthesis by activating the hydroxylase Coq7 in yeast. J. Biol. Chem. 2013, 288, 28126–28137. [Google Scholar] [CrossRef] [PubMed]
  13. Lanning, N.J.; Looyenga, B.D.; Kauffman, A.L.; Niemi, N.M.; Sudderth, J.; DeBerardinis, R.J.; MacKeigan, J.P. A mitochondrial RNAi screen defines cellular bioenergetic determinants and identifies an adenylate kinase as a key regulator of ATP levels. Cell Rep. 2014, 7, 907–917. [Google Scholar] [CrossRef] [PubMed]
  14. Gonzalez-Mariscal, I.; Garcia-Teston, E.; Padilla, S.; Martin-Montalvo, A.; Pomares-Viciana, T.; Vazquez-Fonseca, L.; Gandolfo-Dominguez, P.; Santos-Ocana, C. Regulation of coenzyme Q biosynthesis in yeast: A new complex in the block. IUBMB Life 2014, 66, 63–70. [Google Scholar] [CrossRef] [PubMed]
  15. Padilla, S.; Tran, U.C.; Jimenez-Hidalgo, M.; Lopez-Martin, J.M.; Martin-Montalvo, A.; Clarke, C.F.; Navas, P.; Santos-Ocana, C. Hydroxylation of demethoxy-Q6 constitutes a control point in yeast coenzyme Q6 biosynthesis. Cell. Mol. Life Sci. 2009, 66, 173–186. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  16. Lohman, D.C.; Forouhar, F.; Beebe, E.T.; Stefely, M.S.; Minogue, C.E.; Ulbrich, A.; Stefely, J.A.; Sukumar, S.; Luna-Sanchez, 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]
  17. 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]
  18. Stefely, J.A.; Reidenbach, A.G.; Ulbrich, A.; Oruganty, K.; Floyd, B.J.; Jochem, A.; Saunders, J.M.; Johnson, I.E.; Minogue, C.E.; Wrobel, R.L.; et al. Mitochondrial ADCK3 Employs an Atypical Protein Kinase-like Fold to Enable Coenzyme Q Biosynthesis. Mol. Cell 2015, 57, 83–94. [Google Scholar] [CrossRef] [PubMed]
  19. Parrado-Fernandez, C.; Lopez-Lluch, G.; Rodriguez-Bies, E.; Santa-Cruz, S.; Navas, P.; Ramsey, J.J.; Villalba, J.M. Calorie restriction modifies ubiquinone and COQ transcript levels in mouse tissues. Free Radic. Biol. Med. 2011, 50, 1728–1736. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  20. Lopez-Lluch, G.; Santos-Ocana, C.; Sanchez-Alcazar, J.A.; Fernandez-Ayala, D.J.; Asencio-Salcedo, C.; Rodriguez-Aguilera, J.C.; Navas, P. Mitochondrial responsibility in ageing process: Innocent, suspect or guilty. Biogerontology 2015, 16, 599–620. [Google Scholar] [CrossRef] [PubMed]
  21. Schon, E.A.; DiMauro, S.; Hirano, M.; Gilkerson, R.W. Therapeutic prospects for mitochondrial disease. Trends Mol. Med. 2010, 16, 268–276. [Google Scholar] [CrossRef] [PubMed]
  22. Quinzii, C.M.; Emmanuele, V.; Hirano, M. Clinical presentations of coenzyme q10 deficiency syndrome. Mol. Syndromol. 2014, 5, 141–146. [Google Scholar] [CrossRef] [PubMed]
  23. Doimo, M.; Desbats, M.A.; Cerqua, C.; Cassina, M.; Trevisson, E.; Salviati, L. Genetics of coenzyme q10 deficiency. Mol. Syndromol. 2014, 5, 156–162. [Google Scholar] [CrossRef] [PubMed]
  24. Desbats, M.A.; Lunardi, G.; Doimo, M.; Trevisson, E.; Salviati, L. Genetic bases and clinical manifestations of coenzyme Q10 (CoQ 10) deficiency. J. Inherit. Metab. Dis. 2015, 38, 145–156. [Google Scholar] [CrossRef] [PubMed]
  25. 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]
  26. Montero, R.; Sanchez-Alcazar, J.A.; Briones, P.; Navarro-Sastre, A.; Gallardo, E.; Bornstein, B.; Herrero-Martin, D.; Rivera, H.; Martin, M.A.; Marti, R.; et al. Coenzyme Q10 deficiency associated with a mitochondrial DNA depletion syndrome: A case report. Clin. Biochem. 2009, 42, 742–745. [Google Scholar] [CrossRef] [PubMed]
  27. 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]
  28. Gempel, K.; Topaloglu, H.; Talim, B.; Schneiderat, P.; Schoser, B.G.; Hans, V.H.; Palmafy, 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]
  29. Miles, M.V.; Putnam, P.E.; Miles, L.; Tang, P.H.; DeGrauw, A.J.; Wong, B.L.; Horn, P.S.; Foote, H.L.; Rothenberg, M.E. Acquired coenzyme Q10 deficiency in children with recurrent food intolerance and allergies. Mitochondrion 2011, 11, 127–135. [Google Scholar] [CrossRef] [PubMed]
  30. Haas, D.; Niklowitz, P.; Horster, F.; Baumgartner, E.R.; Prasad, C.; Rodenburg, R.J.; Hoffmann, G.F.; Menke, T.; Okun, J.G. Coenzyme Q(10) is decreased in fibroblasts of patients with methylmalonic aciduria but not in mevalonic aciduria. J. Inherit. Metab. Dis. 2009, 32, 570–575. [Google Scholar] [CrossRef] [PubMed]
  31. Maes, M.; Mihaylova, I.; Kubera, M.; Uytterhoeven, M.; Vrydags, N.; Bosmans, E. Coenzyme Q10 deficiency in myalgic encephalomyelitis/chronic fatigue syndrome (ME/CFS) is related to fatigue, autonomic and neurocognitive symptoms and is another risk factor explaining the early mortality in ME/CFS due to cardiovascular disorder. Neuro Endocrinol. Lett. 2009, 30, 470–476. [Google Scholar] [PubMed]
  32. Fragaki, K.; Cano, A.; Benoist, J.F.; Rigal, O.; Chaussenot, A.; Rouzier, C.; Bannwarth, S.; Caruba, C.; Chabrol, B.; Paquis-Flucklinger, V. Fatal heart failure associated with CoQ10 and multiple OXPHOS deficiency in a child with propionic acidemia. Mitochondrion 2011, 11, 533–536. [Google Scholar] [CrossRef] [PubMed]
  33. 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]
  34. Lopez, L.C.; Schuelke, M.; Quinzii, C.M.; Kanki, T.; Rodenburg, R.J.; 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]
  35. 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]
  36. 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]
  37. 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] [PubMed]
  38. Mollet, J.; Giurgea, I.; Schlemmer, D.; Dallner, G.; Chretien, D.; Delahodde, A.; Bacq, D.; de Lonlay, P.; Munnich, A.; Rotig, 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] [PubMed]
  39. Duncan, A.J.; Bitner-Glindzicz, M.; Meunier, B.; Costello, H.; Hargreaves, I.P.; Lopez, 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] [PubMed]
  40. Freyer, C.; Stranneheim, H.; Naess, K.; Mourier, A.; Felser, A.; Maffezzini, C.; Lesko, N.; Bruhn, H.; Engvall, M.; Wibom, R.; et al. Rescue of primary ubiquinone deficiency due to a novel COQ7 defect using 2,4-dihydroxybensoic acid. J. Med. Genet. 2015, 52, 779–783. [Google Scholar] [CrossRef]
  41. Mollet, J.; Delahodde, A.; Serre, V.; Chretien, D.; Schlemmer, D.; Lombes, A.; Boddaert, N.; Desguerre, I.; de Lonlay, P.; de Baulny, H.O.; et al. CABC1 gene mutations cause ubiquinone deficiency with cerebellar ataxia and seizures. Am. J. Hum. Genet. 2008, 82, 623–630. [Google Scholar] [CrossRef] [PubMed]
  42. Lagier-Tourenne, C.; Tazir, M.; Lopez, 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] [PubMed]
  43. Horvath, R.; Czermin, B.; Gulati, S.; Demuth, S.; Houge, G.; Pyle, A.; Dineiger, C.; Blakely, E.L.; Hassani, A.; Foley, C.; et al. Adult-onset cerebellar ataxia due to mutations in CABC1/ADCK3. J. Neurol. Neurosurg. Psychiatry 2012, 83, 174–178. [Google Scholar] [CrossRef] [PubMed]
  44. Mignot, C.; Apartis, E.; Durr, A.; Lourenco, C.M.; Charles, P.; Devos, D.; Moreau, C.; de Lonlay, P.; Drouot, N.; Burglen, L.; et al. Phenotypic variability in ARCA2 and identification of a core ataxic phenotype with slow progression. Orphanet J. Rare Dis. 2013, 8, 173. [Google Scholar] [CrossRef] [Green Version]
  45. Liu, Y.T.; Hersheson, J.; Plagnol, V.; Fawcett, K.; Duberley, K.E.; Preza, E.; Hargreaves, I.P.; Chalasani, A.; Laura, M.; Wood, N.W.; et al. Autosomal-recessive cerebellar ataxia caused by a novel ADCK3 mutation that elongates the protein: Clinical, genetic and biochemical characterisation. J. Neurol. Neurosurg. Psychiatry 2014, 85, 493–498. [Google Scholar] [CrossRef]
  46. Blumkin, L.; Leshinsky-Silver, E.; Zerem, A.; Yosovich, K.; Lerman-Sagie, T.; Lev, D. Heterozygous Mutations in the ADCK3 Gene in Siblings with Cerebellar Atrophy and Extreme Phenotypic Variability. JIMD Rep. 2014, 12, 103–107. [Google Scholar] [PubMed]
  47. Hikmat, O.; Tzoulis, C.; Knappskog, P.M.; Johansson, S.; Boman, H.; Sztromwasser, P.; Lien, E.; Brodtkorb, E.; Ghezzi, D.; Bindoff, L.A. ADCK3 mutations with epilepsy, stroke-like episodes and ataxia: A POLG mimic? Eur. J. Neurol. 2016, 23, 1188–1194. [Google Scholar] [CrossRef] [PubMed]
  48. Salviati, L.; Trevisson, E.; Hernandez, M.A.R.; Casarin, A.; Pertegato, V.; Doimo, M.; Cassina, M.; Agosto, C.; Desbats, M.A.; Sartori, G.; et al. Haploinsufficiency of COQ4 causes coenzyme Q10 deficiency. J. Med. Genet. 2012, 49, 187–191. [Google Scholar] [CrossRef] [PubMed]
  49. 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] [PubMed]
  50. 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]
  51. Public Health. Community Register of Orphan Medicinal Products. Available online: http://ec.europa.eu/health/documents/community-register/html/o1765.htm (accessed on 4 March 2017).
  52. Montini, G.; Malaventura, C.; Salviati, L. Early coenzyme Q10 supplementation in primary coenzyme Q10 deficiency. N. Engl. J. Med. 2008, 358, 2849–2850. [Google Scholar] [CrossRef] [PubMed]
  53. Pineda, M.; Montero, R.; Aracil, A.; O’Callaghan, M.M.; Mas, A.; Espinos, C.; Martinez-Rubio, D.; Palau, F.; Navas, P.; Briones, P.; et al. Coenzyme Q(10)-responsive ataxia: 2-year-treatment follow-up. Mov. Disord. 2010, 25, 1262–1268. [Google Scholar] [CrossRef] [PubMed]
  54. Saiki, R.; Lunceford, A.L.; Shi, Y.; Marbois, B.; King, R.; Pachuski, J.; Kawamukai, M.; Gasser, D.L.; Clarke, C.F. Coenzyme Q10 supplementation rescues renal disease in Pdss2kd/kd mice with mutations in prenyl diphosphate synthase subunit 2, American journal of physiology. Ren. Physiol. 2008, 295, F1535–F1544. [Google Scholar] [CrossRef] [PubMed]
  55. Lopez, L.C.; Quinzii, C.M.; Area, E.; Naini, A.; Rahman, S.; Schuelke, M.; Salviati, L.; Dimauro, S.; Hirano, M. Treatment of CoQ(10) deficient fibroblasts with ubiquinone, CoQ analogs, and vitamin C: Time- and compound-dependent effects. PLoS ONE 2010, 5, e11897. [Google Scholar] [CrossRef] [PubMed]
  56. Yubero, D.; Montero, R.; Armstrong, J.; Espinos, C.; Palau, F.; Santos-Ocana, C.; Salviati, L.; Navas, P.; Artuch, R. Molecular diagnosis of coenzyme Q10 deficiency. Expert Rev. Mol. Diagn. 2015, 15, 1049–1059. [Google Scholar] [CrossRef] [PubMed]
  57. Yubero, D.; Montero, R.; Artuch, R.; Land, J.M.; Heales, S.J.; Hargreaves, I.P. Biochemical diagnosis of coenzyme q10 deficiency. Mol. Syndromol. 2014, 5, 147–155. [Google Scholar]
  58. Trevisson, E.; DiMauro, S.; Navas, P.; Salviati, L. Coenzyme Q deficiency in muscle. Curr. Opin. Neurol. 2011, 24, 449–456. [Google Scholar] [CrossRef] [PubMed]
  59. Yubero, D.; Montero, R.; Ramos, M.; Neergheen, V.; Navas, P.; Artuch, R.; Hargreaves, I. Determination of urinary coenzyme Q10 by HPLC with electrochemical detection: Reference values for a paediatric population. Biofactors 2015, 41, 424–430. [Google Scholar] [CrossRef] [PubMed]
  60. Bujan, N.; Arias, A.; Montero, R.; Garcia-Villoria, J.; Lissens, W.; Seneca, S.; Espinos, C.; Navas, P.; de Meirleir, L.; Artuch, R.; et al. Characterization of CoQ(1)(0) biosynthesis in fibroblasts of patients with primary and secondary CoQ(1)(0) deficiency. J. Inherit. Metab. Dis. 2014, 37, 53–62. [Google Scholar] [CrossRef] [PubMed]
  61. Fernandez-Ayala, D.J.; Lopez-Lluch, G.; Garcia-Valdes, M.; Arroyo, A.; Navas, P. Specificity of coenzyme Q10 for a balanced function of respiratory chain and endogenous ubiquinone biosynthesis in human cells. Biochim. Biophys. Acta 2005, 1706, 174–183. [Google Scholar] [CrossRef] [PubMed]
  62. Arias, A.; Garcia-Villoria, J.; Rojo, A.; Bujan, N.; Briones, P.; Ribes, A. Analysis of coenzyme Q(10) in lymphocytes by HPLC-MS/MS. J. Chromatogr. B Anal. Technol. Biomed. Life Sci. 2012, 908, 23–26. [Google Scholar] [CrossRef] [PubMed]
Figure 1. HPLC elution profile of lipid extracts from human skeletal muscular tissue. Patient pathological profile (red plot) shows that CoQ10 is clearly diminished compared to healthy control volunteers (black plot). CoQ9 is used as internal standard for normalization.
Figure 1. HPLC elution profile of lipid extracts from human skeletal muscular tissue. Patient pathological profile (red plot) shows that CoQ10 is clearly diminished compared to healthy control volunteers (black plot). CoQ9 is used as internal standard for normalization.
Jcm 06 00027 g001
Figure 2. HPLC elution profile of lipid extracts from human fibroblasts cultured with the radiolabeled precursor 14C-p-HB. Patient pathological profile (red plot) shows that CoQ10 is clearly diminished compared to control cells from healthy humans (blue plot). Left Y-axis shows the radio-flow detector scale (volts). Right Y-axis shows the UV-detector scale (absorbance units) for a standard pool of CoQ10 and CoQ9 (black plot). Notice that the only peak detected in this analysis corresponded with CoQ10.
Figure 2. HPLC elution profile of lipid extracts from human fibroblasts cultured with the radiolabeled precursor 14C-p-HB. Patient pathological profile (red plot) shows that CoQ10 is clearly diminished compared to control cells from healthy humans (blue plot). Left Y-axis shows the radio-flow detector scale (volts). Right Y-axis shows the UV-detector scale (absorbance units) for a standard pool of CoQ10 and CoQ9 (black plot). Notice that the only peak detected in this analysis corresponded with CoQ10.
Jcm 06 00027 g002
Table 1. Yeast COQ genes and their characterized human homologues.
Table 1. Yeast COQ genes and their characterized human homologues.
YeastHumanFunction
COQ1PDSS1 */PDSS2 *Synthesis of polyprenyl-diphosphate
COQ2COQ2 *pHB-prenyl-transferase
COQ3COQ3 *Methyltransferase
COQ4COQ4 *Organization of the multi-enzyme complex
COQ5COQ5Methyltransferase
COQ6COQ6 *Mono-oxygenase
COQ7COQ7 *Hydroxylase
COQ8ADCK3 */ADCK4 *Unorthodox kinase (regulatory)
COQ9COQ9 *Lipid binding protein
COQ10COQ10A/COQ10BCoQ chaperone
PTC7PPTC7Phosphatase (regulatory)
* These genes were mutated in human causing primary CoQ10 deficiency.

Share and Cite

MDPI and ACS Style

Rodríguez-Aguilera, J.C.; Cortés, A.B.; Fernández-Ayala, D.J.M.; Navas, P. Biochemical Assessment of Coenzyme Q10 Deficiency. J. Clin. Med. 2017, 6, 27. https://doi.org/10.3390/jcm6030027

AMA Style

Rodríguez-Aguilera JC, Cortés AB, Fernández-Ayala DJM, Navas P. Biochemical Assessment of Coenzyme Q10 Deficiency. Journal of Clinical Medicine. 2017; 6(3):27. https://doi.org/10.3390/jcm6030027

Chicago/Turabian Style

Rodríguez-Aguilera, Juan Carlos, Ana Belén Cortés, Daniel J. M. Fernández-Ayala, and Plácido Navas. 2017. "Biochemical Assessment of Coenzyme Q10 Deficiency" Journal of Clinical Medicine 6, no. 3: 27. https://doi.org/10.3390/jcm6030027

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