Effects of metformin and other biguanides on oxidative phosphorylation in mitochondria

Hannah R Bridges, Andrew J Y Jones, Michael N Pollak, Judy Hirst, Hannah R Bridges, Andrew J Y Jones, Michael N Pollak, Judy Hirst

Abstract

The biguanide metformin is widely prescribed for Type II diabetes and has anti-neoplastic activity in laboratory models. Despite evidence that inhibition of mitochondrial respiratory complex I by metformin is the primary cause of its cell-lineage-specific actions and therapeutic effects, the molecular interaction(s) between metformin and complex I remain uncharacterized. In the present paper, we describe the effects of five pharmacologically relevant biguanides on oxidative phosphorylation in mammalian mitochondria. We report that biguanides inhibit complex I by inhibiting ubiquinone reduction (but not competitively) and, independently, stimulate reactive oxygen species production by the complex I flavin. Biguanides also inhibit mitochondrial ATP synthase, and two of them inhibit only ATP hydrolysis, not synthesis. Thus we identify biguanides as a new class of complex I and ATP synthase inhibitor. By comparing biguanide effects on isolated complex I and cultured cells, we distinguish three anti-diabetic and potentially anti-neoplastic biguanides (metformin, buformin and phenformin) from two anti-malarial biguanides (cycloguanil and proguanil): the former are accumulated into mammalian mitochondria and affect oxidative phosphorylation, whereas the latter are excluded so act only on the parasite. Our mechanistic and pharmacokinetic insights are relevant to understanding and developing the role of biguanides in new and existing therapeutic applications, including cancer, diabetes and malaria.

Figures

Figure 1. Effects of biguanides on isolated…
Figure 1. Effects of biguanides on isolated bovine complex I
(A) Dependence of NADH oxidation on biguanide concentration, relative to a biguanide-free control. Colours are as in (B), and the IC50 values (in mM) are noted. Data points are means±S.E.M. (n=3–5). (B) Relationship between the inhibition IC50 values and octanol:PBS distribution coefficient (D) values of the biguanides. IC50 values are in mM with 95% confidence intervals; log D values are means±S.E.M. (n=3). Linear fit with R2=0.963. Biguanide structures shown are for the neutral forms. (C) 12 K EPR spectra of the FeS clusters in complex I in the presence and absence of biguanides. Ph, phenyl; PhCl, para-chlorophenyl. (D) The effect of 25 mM metformin on NADH:decylubiquinone oxidoreduction, presented as the measurement of KM for decylubiquinone. The data (means±S.E.M.; n=3) were fit using the Michaelis–Menten equation. (E) Metformin inhibition of NADH oxidation by SMPs (4.5 μg protein/ml). Trace 1, 100 mM metformin added before initiation of catalysis by 200 μM NADH at t=0. Trace 2, catalysis initiated by 200 μM NADH 10 min before the addition of 100 mM metformin at t=0. Traces 3 and 4, controls for 1 and 2, with NaCl instead of metformin. Four traces are overlaid for each condition.
Figure 2. Biguanide effects on the flavin…
Figure 2. Biguanide effects on the flavin site of complex I
(A) Effects of metformin on flavin site reactions that require nucleotide binding to the reduced flavin [NADH:HAR (blue) and NADH:paraquat (green) oxidoreduction] [27], and nucleotide-free reduced flavin [NADH:FeCN (purple) and NADH:O2 (red) oxidoreduction] [26]. NADH:O2 oxidoreduction (H2O2 production) was detected directly (by NADH oxidation, circles) or as H2O2 by the Amplex Red assay (squares) [15], with/without (open/closed symbols respectively) 1 μM rotenone. Data for H2O2 production by subcomplex Iλ are shown as open diamonds (◇). Data points are means±S.E.M., n=3–5. (B) Stimulation of flavin-site reactions, measured at the inhibitory IC50 concentrations (black, H2O2 production measured using Amplex Red; grey, NADH:FeCN oxidoreduction). Gu, 25 mM guanidinium; Me, 25 mM metformin; Bu, 5 mM buformin; Ph, 0.5 mM phenformin; Cy, 0.7 mM cycloguanil; Pr, 0.05 mM proguanil. Data are presented as means±S.E.M., n=3–5. (C) NADH-dependence of NADH:FeCN catalysis by isolated bovine complex I. ●, 100 mM metformin; ○, control. Data points are means±S.E.M., n=3–5 and data were fit as described previously [26] with KMNADH=220 μM, kcatNADH=1500 s−1, kcatFeCN=3.7×107 M−1·s−1 and KNADHRed (the dissociation constant for NADH binding to the reduced flavin)=KNADHSemi=11 μM (control) or 181 μM (metformin). (D) Dependence of H2O2 production on the NAD+ potential, set by the Nernst equation with 30 μM NADH and variable NAD+ [15]. Control [with (●) and without (○) 200 mM NaCl], 20 mM phenformin (green) and 200 mM metformin (purple). Data points are means±S.E.M., n=3–5. The stimulation of H2O2 production by metformin (red) is also represented as the non-normalized percentage activity relative to the control (see A for the effect of metformin in NADH only). (E) Simplified scheme illustrating how metformin affects the different flavin site reactions differently. The boxes represent different states of the complex I flavin site, with the flavin oxidized or reduced, and with or without NADH or NAD+ bound.
Figure 3. Biguanide interactions with respiratory complexes…
Figure 3. Biguanide interactions with respiratory complexes II, III and IV
(A) Rates of complex II + III + IV activity in bovine SMPs measured by a coupled enzyme assay [28] in the presence of biguanide concentrations equivalent to IC50 for NADH:decylubiquinone catalysis. Results are means±S.E.M. as a percentage of the biguanide-free control, n=3–4. (B) Rates of respiratory complex activity in the presence of 0.7 mM cycloguanil as a percentage of their respective cycloguanil-free controls. ****P<0.0001.
Figure 4. Biguanide inhibition of F 1…
Figure 4. Biguanide inhibition of F1F0-ATP synthase in bovine SMPs
(A) Inhibition of ATP hydrolysis by SMPs (colours as in B). Data points are means±S.E.M., n=3–5, and IC50 values are indicated in mM. (B) Relationship between the IC50 values for ATP hydrolysis and the log D values. IC50 values are in mM with 95% confidence intervals; log D values are means±S.E.M., n=3. Linear fit with R2=0.948. (C) Relative inhibition of ATP production in SMPs in the presence of piericidin A (○) or metformin (●). ATP synthesis was driven using 200 μM NADH; NADH oxidation was measured spectroscopically, and the concentration of ATP determined after 3.5 min (during this time a linear rate is observed). NADH oxidation rates were adjusted using 0–50 nM piericidin A or 0–250 mM metformin, with the ionic strength kept constant at 250 mM using NaCl. (DF) Dose-dependent effects of biguanides on succinate-driven ATP production. Data points are means±S.E.M., n=3, as a percentage of ATP production in the absence of biguanide. ATP concentrations were determined after 3.5 min (○) then corrected for the rate of succinate oxidation, detected spectroscopically using a coupled assay system [28] (●). The IC50 values for hydrolysis are marked with vertical lines.
Figure 5. Selective uptake of biguanides into…
Figure 5. Selective uptake of biguanides into mitochondria
(A) Effects of biguanides on the rotenone-sensitive OCR of Hep G2 cells. The traces are the means±S.D. of multiple traces. The biguanide concentrations are equal to the complex I IC50/10 values. Purple, 1.9 mM metformin; green, 0.04 mM phenformin; orange, 0.07 mM cycloguanil added in DMSO; red, 0.007 mM proguanil in DMSO; black, control (no biguanide); cyan, control (DMSO). (B) The effects of biguanides on the ECAR of Hep G2 cells. Conditions and colours as in (A). (C) Mitochondrial respiratory coupling test on 143B cells permeabilized with 2 nM PMP, respiring on pyruvate and malate and using biguanides at concentrations equal to their complex I IC50/10 values. Black, control; purple, 2 mM metformin; orange, 0.07 mM cycloguanil; red, 0.007 mM proguanil. Data points are means±S.E.M., n=5–6. (D) Rates of rotenone-sensitive ADP-stimulated NADH-linked respiration in permeabilized 143B cells (light grey) and rat liver mitochondria (dark grey) after 15 min of incubation with 2 mM metformin (Met.), 0.07 mM cycloguanil (Cyclo.) or 0.007 mM proguanil (Pro.). Values are mean percentage of the control±S.E.M., n=5–6. ****P< 0.01. Cont. control.
Figure 6. Biguanide effects on cells in…
Figure 6. Biguanide effects on cells in culture: accumulation and reversibility
(A) The normalized rotenone-sensitive OCR 6 h after the addition of metformin or phenformin to Hep G2 (black) or 143B (grey) cells. Results are means±S.E.M., n=3–4. The IC50 values (with 95% confidence intervals) are metformin, 240±10 μM (143B) and 330±20 μM (Hep G2); phenformin, 3.9±1.0 μM (143B) and 3.8±0.4 μM (Hep G2). (B) Metformin inhibition of the OCR by Hep G2 cells. Results are means of multiple traces±S.E.M., n=3–4. Rotenone (200 nM) was added to half of the samples part way through the experiment; pairs of traces with and without rotenone are coloured the same. Then, the assay medium in each experiment was exchanged for a metformin and rotenone-free medium. All data were measured using a Seahorse Extracellular Flux Analyzer at 37°C.

References

    1. Shaw R. J., Lamia K. A., Vasquez D., Koo S.-H., Bardeesy N., DePinho R. A., Montminy M., Cantley L. C. The kinase LKB1 mediates glucose homeostasis in liver and therapeutic effects of metformin. Science. 2005;310:1642–1646. doi: 10.1126/science.1120781.
    1. Miller R. A., Chu Q., Xie J., Foretz M., Viollet B., Birnbaum M. J. Biguanides suppress hepatic glucagon signalling by decreasing production of cyclic AMP. Nature. 2013;494:256–260. doi: 10.1038/nature11808.
    1. Madiraju A. K., Erion D. M., Rahimi Y., Zhang X.-M., Braddock D. T., Albright R. A., Prigaro B. J., Wood J. L., Bhanot S., MacDonald M. J., et al. Metformin suppresses gluconeogenesis by inhibiting mitochondrial glycerophosphate dehydrogenase. Nature. 2014;510:542–546. doi: 10.1038/nature13270.
    1. Pollak M. N. Investigating metformin for cancer prevention and treatment: the end of the beginning. Cancer Discov. 2012;2:778–790. doi: 10.1158/-12-0263.
    1. Pernicova I., Korbonits M. Metformin-mode of action and clinical implications for diabetes and cancer. Nat. Rev. Endocrinol. 2014;10:143–156. doi: 10.1038/nrendo.2013.256.
    1. Yin M., van der Horst I. C. C., van Melle J. P., Qian C., van Gilst W. H., Silljé H. H. W., de Boer R. A. Metformin improves cardiac function in a nondiabetic rat model of post-MI heart failure. Am. J. Physiol. Heart Circ. Physiol. 2011;301:H459–H468. doi: 10.1152/ajpheart.00054.2011.
    1. Owen M. R., Doran E., Halestrap A. P. Evidence that metformin exerts its anti-diabetic effects through inhibition of complex I of the mitochondrial respiratory chain. Biochem. J. 2000;348:607–614. doi: 10.1042/0264-6021:3480607.
    1. Dykens J. A., Jamieson J., Marroquin L., Nadanaciva S., Billis P. A., Will Y. Biguanide-induced mitochondrial dysfunction yields increased lactate production and cytotoxicity of aerobically-poised HepG2 cells and human hepatocytes in vitro. Toxicol. Appl. Pharm. 2008;233:203–210. doi: 10.1016/j.taap.2008.08.013.
    1. Wheaton W. W., Weinberg S. E., Hamanaka R. B., Soberanes S., Sullivan L. B., Anso E., Glasauer A., Dufour E., Mutlu G. M., Budigner G. R. S., Chandel N. S. Metformin inhibits mitochondrial complex I of cancer cells to reduce tumorigenesis. eLife. 2014;3:e02242. doi: 10.7554/eLife.02242.
    1. Hawley S. A., Ross F. A., Chevtzoff C., Green K. A., Evans A., Fogarty S., Towler M. C., Brown L. J., Ogunbayo O. A., Evans A. M., Hardie D. G. Use of cells expressing γ-subunit variants to identify diverse mechanisms of AMPK activation. Cell Metab. 2010;11:554–565. doi: 10.1016/j.cmet.2010.04.001.
    1. Green B. D., Irwin N., Duffy N. A., Gault V. A., O’Harte F. P. M., Flatt P. R. Inhibition of dipeptidyl peptidase-IV activity by metformin enhances the antidiabetic effects of glucagon-like peptide-1. Eur. J. Pharmacol. 2006;547:192–199. doi: 10.1016/j.ejphar.2006.07.043.
    1. Ouyang J., Parakhia R. A., Ochs R. S. Metformin activates AMP kinase through inhibition of AMP deaminase. J. Biol. Chem. 2011;286:1–11. doi: 10.1074/jbc.M110.121806.
    1. Salani B., Marini C., del Rio A., Ravera S., Massollo M., Orengo A. M., Amaro A., Passalacqua M., Maffioli S., Pfeffer U., et al. Metformin impairs glucose consumption and survival in Calu-1 cells by direct inhibition of hexokinase-II. Sci. Rep. 2013;3:2070. doi: 10.1038/srep02070.
    1. Hirst J. Mitochondrial complex I. Annu. Rev. Biochem. 2013;82:551–575. doi: 10.1146/annurev-biochem-070511-103700.
    1. Kussmaul L., Hirst J. The mechanism of superoxide production by NADH:ubiquinone oxidoreductase (complex I) from bovine heart mitochondria. Proc. Natl. Acad. Sci. U.S.A. 2006;103:7607–7612. doi: 10.1073/pnas.0510977103.
    1. Murphy M. P. How mitochondria produce reactive oxygen species. Biochem. J. 2009;417:1–13. doi: 10.1042/BJ20081386.
    1. Nattrass M., Alberti K. G. M. M. Biguanides. Diabetologia. 1978;14:71–74. doi: 10.1007/BF01263443.
    1. Painter H. J., Morrisey J. M., Mather M. W., Vaidya A. B. Specific role of mitochondrial electron transport in blood-stage Plasmodium falciparum. Nature. 2007;446:88–91. doi: 10.1038/nature05572.
    1. Vanichtanankul J., Taweechai S., Uttamapinant C., Chitnumsub P., Vilaivan T., Yuthavong Y., Kamchonwongpaisan S. Combined spatial limitation around residues 16 and 108 of Plasmodium falciparum dihydrofolate reductase explains resistance to cycloguanil. Antimicrob. Agents Chemother. 2012;56:3928–3935. doi: 10.1128/AAC.00301-12.
    1. Sharpley M. S., Shannon R. J., Draghi F., Hirst J. Interactions between phospholipids and NADH:ubiquinone oxidoreductase (complex I) from bovine mitochondria. Biochemistry. 2006;45:241–248. doi: 10.1021/bi051809x.
    1. Bridges H. R., Grgic L., Harbour M. E., Hirst J. The respiratory complexes I from the mitochondria of two Pichia species. Biochem. J. 2009;422:151–159. doi: 10.1042/BJ20090492.
    1. Sazanov L. A., Carroll J., Holt P., Toime L., Fearnley I. M. A role for native lipids in the stabilization and two-dimensional crystallization of the Escherichia coli NADH:ubiquinone oxidoreductase (complex I) J. Biol. Chem. 2003;278:19483–19491. doi: 10.1074/jbc.M208959200.
    1. Pryde K. R., Hirst J. Superoxide is produced by the reduced flavin in mitochondrial complex I: a single, unified mechanism that applies during both forward and reverse electron transfer. J. Biol. Chem. 2011;286:18056–18065. doi: 10.1074/jbc.M110.186841.
    1. Chappell J. B., Hansford R. G. Preparation of mitochondria from animal tissues and yeasts. In: Birnie G. D., editor. Subcellular Components: Preparation and Fractionation. 2nd ed. London: Butterworths; 1972. pp. 77–91.
    1. Lutter R., Saraste M., van Walraven H. S., Runswick M. J., Finel M., Deatherage J. F., Walker J. E. F1F0-ATP synthase from bovine heart mitochondria: development of the purification of a monodisperse oligomycin-sensitive ATPase. Biochem. J. 1993;295:799–806.
    1. Birrell J. A., Yakovlev G., Hirst J. Reactions of the flavin mononucleotide in complex I: a combined mechanism describes NADH oxidation coupled to the reduction of APAD+, ferricyanide or molecular oxygen. Biochemistry. 2009;48:12005–12013. doi: 10.1021/bi901706w.
    1. Birrell J. A., King M. S., Hirst J. A ternary mechanism for NADH oxidation by positively charged electron acceptors, catalyzed at the flavin site in respiratory complex I. FEBS Lett. 2011;585:2318–2322. doi: 10.1016/j.febslet.2011.05.065.
    1. Jones A. J. Y., Hirst J. A spectrophotometric, coupled enzyme assay to measure the activity of succinate dehydrogenase. Anal. Biochem. 2013;442:19–23. doi: 10.1016/j.ab.2013.07.018.
    1. Pullman M. E., Penefsky H. S., Datta A., Racker E. Partial resolution of the enzymes catalyzing oxidative phosphorylation. J. Biol. Chem. 1960;235:3322–3329.
    1. Danielsson L.-G., Zhang Y.-H. Methods for determining n-octanol-water partition coefficients. Trends Anal. Chem. 1996;15:188–196.
    1. Ray P. Complex compounds of biguanides and guanylureas with metallic elements. Chem. Rev. 1961;61:313–359. doi: 10.1021/cr60212a001.
    1. Forneris F., Orru R., Bonivento D., Chiarelli L. R., Mattevi A. ThermoFAD, a Thermofluor-adapted flavin ad hoc detection system for protein folding and ligand binding. FEBS J. 2009;276:2833–2840. doi: 10.1111/j.1742-4658.2009.07006.x.
    1. Birrell J. A., Morina K., Bridges H. R., Friedrich T., Hirst J. Investigating the function of [2Fe–2S] cluster N1a, the off-pathway cluster in complex I, by manipulating its reduction potential. Biochem. J. 2013;456:139–146. doi: 10.1042/BJ20130606.
    1. Holford N. H. G., Sheiner L. B. Understanding the dose-effect relationship: clinical application of pharmacokinetic-pharmacodynamic models. Clin. Pharmacokinet. 1981;6:429–453. doi: 10.2165/00003088-198106060-00002.
    1. Grivennikova V. G., Vinogradov A. D. Partitioning of superoxide and hydrogen peroxide production by mitochondrial respiratory complex I. Biochim. Biophys. Acta. 2013;1827:446–454. doi: 10.1016/j.bbabio.2013.01.002.
    1. Reda T., Barker C. D., Hirst J. Reduction of the iron-sulfur clusters in mitochondrial NADH:ubiquinone oxidoreductase (complex I) by EuII-DTPA, a very low potential reductant. Biochemistry. 2008;47:8885–8893. doi: 10.1021/bi800437g.
    1. Briggs G. E., Haldane J. B. S. A note on the kinetics of enzyme action. Biochem. J. 1925;19:338–339.
    1. Baradaran R., Berrisford J. M., Minhas G. S., Sazanov L. A. Crystal structure of the entire respiratory complex I. Nature. 2013;494:443–448. doi: 10.1038/nature11871.
    1. Galkin A., Meyer B., Wittig I., Karas M., Schägger H., Vinogradov A., Brandt U. Identification of the mitochondrial ND3 subunit as a structural component involved in the active/deactive enzyme transition of respiratory complex I. J. Biol. Chem. 2008;283:20907–20913. doi: 10.1074/jbc.M803190200.
    1. Roberts P. G., Hirst J. The deactive form of respiratory complex I from mammalian mitochondria is a Na+/H+ antiporter. J. Biol. Chem. 2012;287:34743–34751. doi: 10.1074/jbc.M112.384560.
    1. Kalia J., Swartz K. J. Elucidating the molecular basis of action of a classic drug: guanidine compounds as inhibitors of voltage-gated potassium channels. Mol. Pharmacol. 2011;80:1085–1095. doi: 10.1124/mol.111.074989.
    1. Hatefi Y., Stempel K. E., Hanstein W. G. Inhibitors and activators of the mitochondrial reduced diphosphopyridine nucleotide dehydrogenase. J. Biol. Chem. 1969;244:2358–2365.
    1. Hirst J., Carroll J., Fearnley I. M., Shannon R. J., Walker J. E. The nuclear encoded subunits of complex I from bovine heart mitochondria. Biochim. Biophys. Acta. 2003;1604:135–150. doi: 10.1016/S0005-2728(03)00059-8.
    1. Sled V. D., Rudnitzky N. I., Hatefi Y., Ohnishi T. Thermodynamic analysis of flavin in mitochondrial NADH:ubiquinone oxidoreductase (complex I) Biochemistry. 1994;33:10069–10075. doi: 10.1021/bi00199a034.
    1. Carvalho C., Correia S., Santos M. S., Seiça R., Oliveira C. R., Moreira P. I. Metformin promotes isolated rat liver mitochondria impairment. Mol. Cell. Biochem. 2008;308:75–83. doi: 10.1007/s11010-007-9614-3.
    1. Zou M.-H., Kirkpatrick S. S., Davis B. J., Nelson J. S., Wiles W. G., Schlattner U., Neumann D., Brownlee M., Freeman M. B., Goldman M. H. Activation of the AMP-activated protein kinase by the anti-diabetic drug metformin in vivo: role of mitochondrial reactive nitrogen species. J. Biol. Chem. 2004;279:43940–43951. doi: 10.1074/jbc.M404421200.
    1. MacKenzie R. M., Salt I. P., Miller W. H., Logan A., Ibrahim H. A., Degasper A., Dymott J. A., Hamilton C. A., Murphy M. P., Delles C., Dominiczak A. F. Mitochondrial reactive oxygen species enhance AMP-activated protein kinase activation in the endothelium of patients with coronary artery disease and diabetes. Clin. Sci. 2013;124:403–411. doi: 10.1042/CS20120239.
    1. Algire C., Moiseeva O., Deschênes-Simard X., Amrein L., Petruccelli L., Birman E., Viollet B., Ferbeyre G., Pollak M. N. Metformin reduces endogenous reactive oxygen species and associated DNA damage. Cancer Prev. Res. 2012;5:536–543. doi: 10.1158/1940-6207.CAPR-11-0536.
    1. El-Mir M.-Y., Nogueira V., Fontaine E., Avéret N., Rigoulet M., Leverve X. Dimethylbiguanide inhibits cell respiration via an indirect effect targeted on the respiratory chain complex I. J. Biol. Chem. 2000;275:223–228. doi: 10.1074/jbc.275.1.223.
    1. Drahota Z., Palenickova E., Endlicher R., Milerova M., Brejchova J., Vosahlikova M., Svoboda P., Kazdova L., Kalous M., Cervinkova Z., Cahova M. Biguanides inhibit complex I, II and IV of rat liver mitochondria and modify their functional properties. Physiol. Res. 2013;63:1–11.
    1. Corominas-Faja B., Quirantes-Piné R., Oliveras-Ferraros C., Vazquez-Martin A., Cufí S., Martin-Castillo B., Micol V., Joven J., Segura-Carretero A., Menendez J. A. Metabolomic fingerprint reveals that metformin impairs one-carbon metabolism in a manner similar to the antifolate class of chemotherapy drugs. Aging. 2012;4:480–496.
    1. Cabreiro F., Au C., Leung K.-Y., Vergara-Irigaray N., Cochemé H. M., Noori T., Weinkove D., Schuster E., Greene N. D. E., Gems D. Metformin retards aging in C. elegans by altering microbial folate and methionine metabolism. Cell. 2013;153:228–239. doi: 10.1016/j.cell.2013.02.035.
    1. Monera O. D., Kay C. M., Hodges R. S. Protein denaturation with guanidine hydrochloride or urea provides a different estimate of stability depending on the contributions of electrostatic interactions. Protein Sci. 1994;3:1984–1991. doi: 10.1002/pro.5560031110.
    1. England J. L., Haran G. Role of solvation effects in protein denaturation: from thermodynamics to single molecules and back. Annu. Rev. Phys. Chem. 2011;62:257–277. doi: 10.1146/annurev-physchem-032210-103531.
    1. Gledhill J. R., Montgomery M. G., Leslie A. G. W., Walker J. E. How the regulatory protein, IF1, inhibits F1-ATPase from bovine mitochondria. Proc. Natl. Acad. Sci. U.S.A. 2007;104:15671–15676. doi: 10.1073/pnas.0707326104.
    1. Atwal K. S., Ahmad S., Ding C. Z., Stein P. D., Lloyd J., Hamann L. G., Green D. W., Ferrara F. N., Wang P., Rogers W. L., et al. N-[1-aryl-2-(1-imidazolo)ethyl]-guanidine derivatives as potent inhibitors of the bovine mitochondrial F1FO ATP hydrolase. Bioorg. Med. Chem. Lett. 2004;14:1027–1030. doi: 10.1016/j.bmcl.2003.11.077.
    1. Schulz M., Schmoldt A. Therapeutic and toxic blood concentrations of more than 800 drugs and other xenobiotics. Pharmazie. 2003;58:447–474.
    1. Helsby N. A., Edwards G., Breckenridge A. M., Ward S. A. The multiple dose pharmacokinetics of proguanil. Br. J. Clin. Pharmacol. 1993;35:653–656. doi: 10.1111/j.1365-2125.1993.tb04197.x.
    1. Kerb R., Fux R., Mörike K., Kremsner P. G., Gil J. P., Gleiter C. H., Schwab M. Pharmacogenetics of antimalarial drugs: effect on metabolism and transport. Lancet Infect. Dis. 2009;9:760–764. doi: 10.1016/S1473-3099(09)70320-2.
    1. Wang D.-S., Jonker J. W., Kato Y., Kusuhara H., Schinkel A. H., Sugiyama Y. Involvement of organic cation transporter 1 in hepatic and intestinal distribution of metformin. J. Pharmacol. Exp. Ther. 2002;302:510–515. doi: 10.1124/jpet.102.034140.
    1. Segal E. D., Yasmeen A., Beauchamp M.-C., Rosenblatt J., Pollak M., Gotlieb W. H. Relevance of the OCT1 transporter to the antineoplastic effect of biguanides. Biochem. Biophys. Res. Commun. 2011;414:694–699. doi: 10.1016/j.bbrc.2011.09.134.
    1. Lamhonwah A.-M., Tein I. Novel localization of OCTN1, an organic cation/carnitine transporter, to mammalian mitochondria. Biochem. Biophys. Res. Commun. 2006;345:1315–1325. doi: 10.1016/j.bbrc.2006.05.026.
    1. Shitara Y., Nakamichi N., Norioka M., Shima H., Kato Y., Horie T. Role of organic cation/carnitine transporter 1 in uptake of phenformin and inhibitory effect on complex I respiration in mitochondria. Toxicol. Sci. 2012;132:32–42. doi: 10.1093/toxsci/kfs330.
    1. Appleby R. D., Porteous W. K., Hughes G., James A. M., Shannon D., Wei Y.-H., Murphy M. P. Quantitation and origin of the mitochondrial membrane potential in human cells lacking mitochondrial DNA. Eur. J. Biochem. 1999;262:108–116. doi: 10.1046/j.1432-1327.1999.00350.x.
    1. Galkin A., Abramov A. Y., Frakich N., Duchen M. R., Moncada S. Lack of oxygen deactivates mitochondrial complex I: implications for ischemic injury? J. Biol. Chem. 2009;284:36055–36061. doi: 10.1074/jbc.M109.054346.
    1. Campanella M., Casswell E., Chong S., Farah Z., Wieckowski M. R., Abramov A. Y., Tinker A., Duchen M. R. Regulation of mitochondrial structure and function by the F1F0-ATPase inhibitor protein, IF1. Cell Metab. 2008;8:13–25. doi: 10.1016/j.cmet.2008.06.001.
    1. Chen Q., Camara A. K. S., Stowe D. F., Hoppel C. L., Lesnefsky E. J. Modulation of electron transport protects cardiac mitochondria and decreases myocardial injury during ischemia and reperfusion. Am. J. Physiol. Cell. Physiol. 2007;292:C137–C147. doi: 10.1152/ajpcell.00270.2006.
    1. Chouchani E. T., Methner C., Nadtochiy S. M., Logan A., Pell V. R., Ding S., James A. M., Cochemé H. M., Reinhold J., Lilley K. S., et al. Cardioprotection by S-nitrosation of a cysteine switch on mitochondrial complex I. Nat. Med. 2013;19:753–759. doi: 10.1038/nm.3212.
    1. Grover G. J., Atwal K. S., Sleph P. G., Wang F.-L., Monshizadegan H., Monticello T., Green D. W. Excessive ATP hydrolysis in ischemic myocardium by mitochondrial F1FO-ATPase: effect of selective pharmacological inhibition of mitochondrial ATPase hydrolase activity. Am. J. Physiol. Heart Circ. Physiol. 2004;287:H1747–H1755. doi: 10.1152/ajpheart.01019.2003.
    1. Batandier C., Guigas B., Detaille D., El-Mir M.-Y., Fontaine E., Rigoulet M., Leverve X. M. The ROS production induced by a reverse-electron flux at respiratory chain complex I is hampered by metformin. J. Bioenerg. Biomembr. 2006;38:33–42. doi: 10.1007/s10863-006-9003-8.
    1. Paiva M. A., Rutter-Locher Z., Gonçalves L. M., Goncalves L. M., Providência L. A., Davidson S. M., Yellon D. M., Mocanu M. M. Enhancing AMPK activation during ischemia protects the diabetic heart against reperfusion injury. Am. J. Physiol. Heart Circ. Physiol. 2011;300:H2123–H2134. doi: 10.1152/ajpheart.00707.2010.
    1. Suissa S., Azoulay L. Metformin and the risk of cancer: time-related biases in observational studies. Diabetes Care. 2012;35:2665–2673. doi: 10.2337/dc12-0788.
    1. Vander Heiden M. G., Cantley L. C., Thompson C. B. Understanding the Warburg effect: the metabolic requirements of cell proliferation. Science. 2009;324:1029–1033. doi: 10.1126/science.1160809.
    1. Costello L. C., Franklin R. B., Feng P. Mitochondrial function, zinc, and intermediary metabolism relationships in normal prostate and prostate cancer. Mitochondrion. 2005;5:143–153. doi: 10.1016/j.mito.2005.02.001.
    1. Birsoy K., Possemato R., Lorbeer F. K., Bayraktar E. C., Thiru P., Yucel B., Wang T., Chen W. W., Clish C. B., Sabatini D. M. Metabolic determinants of cancer cell sensitivity to glucose limitation and biguanides. Nature. 2014;508:108–112. doi: 10.1038/nature13110.

Source: PubMed

3
Sottoscrivi