Acetylation turns leucine into a drug by membrane transporter switching

Grant C Churchill, Michael Strupp, Cailley Factor, Tatiana Bremova-Ertl, Mallory Factor, Marc C Patterson, Frances M Platt, Antony Galione, Grant C Churchill, Michael Strupp, Cailley Factor, Tatiana Bremova-Ertl, Mallory Factor, Marc C Patterson, Frances M Platt, Antony Galione

Abstract

Small changes to molecules can have profound effects on their pharmacological activity as exemplified by the addition of the two-carbon acetyl group to make drugs more effective by enhancing their pharmacokinetic or pharmacodynamic properties. N-acetyl-D,L-leucine is approved in France for vertigo and its L-enantiomer is being developed as a drug for rare and common neurological disorders. However, the precise mechanistic details of how acetylation converts leucine into a drug are unknown. Here we show that acetylation of leucine switches its uptake into cells from the L-type amino acid transporter (LAT1) used by leucine to organic anion transporters (OAT1 and OAT3) and the monocarboxylate transporter type 1 (MCT1). Both the kinetics of MCT1 (lower affinity compared to LAT1) and the ubiquitous tissue expression of MCT1 make it well suited for uptake and distribution of N-acetyl-L-leucine. MCT1-mediated uptake of a N-acetyl-L-leucine as a prodrug of leucine bypasses LAT1, the rate-limiting step in activation of leucine-mediated signalling and metabolic process inside cells such as mTOR. Converting an amino acid into an anion through acetylation reveals a way for the rational design of drugs to target anion transporters.

Trial registration: ClinicalTrials.gov NCT03759639 NCT03759665 NCT03759678.

Conflict of interest statement

I have read the journal's policy and the authors of this manuscript have the following competing interests: MS is Joint Chief Editor of the Journal of Neurology, Editor in Chief of Frontiers of Neuro-otology and Section Editor of F1000. He has received speaker’s honoraria from Abbott, Actelion, Auris Medical, Biogen, Eisai, Grünenthal, GSK, Henning Pharma, Interacoustics, Merck, MSD, Otometrics, Pierre-Fabre, TEVA, UCB. He is a shareholder of IntraBio. He acts as a consultant for Abbott, Actelion, AurisMedical, Heel, IntraBio and Sensorion. MF is a co-founder, shareholder, and Chairman of IntraBio. MP is a shareholder of IntraBio and has received consulting fees, honoraria, and research grants from Actelion Pharmaceuticals Ltd. and Biomarin. TBE received honoraria for lecturing from Actelion and Sanofi Genzyme. CF is a consultant to IntraBio. GCC, AG and FP are cofounders, shareholders and consultants to IntraBio. IntraBio Ltd is the applicant for patents WO2018229738 (Treatment For Migraine), WO2017182802 (Acetyl-Leucine Or A Pharmaceutically Acceptable Salt Thereof For Improved Mobility And Cognitive Function), WO2019078915 and WO2018029658 (Therapeutic Agents For Neurodegenerative Diseases), WO2018029657 (Pharmaceutical Compositions And Uses Directed To Lysosomal Storage Disorders), and WO2019079536 (Therapeutic Agents For Improved Mobility And Cognitive Function And For Treating Neurodegenerative Diseases And Lysosomal Storage Disorders).

© 2021. The Author(s).

Figures

Figure 1
Figure 1
The effects of N-acetylation on the chemical properties and pharmacological consequences of the drug N-acetyl-l-leucine. (a) Mechanisms of absorption illustrated by crossing a membrane by passive diffusion or carrier-mediated uptake. In general, hydrophobic uncharged molecules can cross lipid bilayer membranes through simple passive diffusion, whereas charged molecules including zwitterions (a positive and negative charge within the same molecule), anions and cations cannot cross membranes without a transporter. Over 400 Solute Carrier (SLC) transporters are known with broad but overlapping selectivities for substrate. l-leucine as an obligate zwitterion at all biologically relevant pH values has an absolute requirement for its carrier, the high-affinity l-type amino Acid Transporter (LAT1). In contrast, N-acetyl-l-leucine can exist as a neutral species and passively cross membranes at low pH, or as an anion recognized by other transporters. (b) Comparison of the physicochemical properties of l-leucine and N-acetyl-l-leucine relating to oral bioavailability. (c) Speciation curves for the protonation states of l-leucine and N-acetyl-l-leucine. The gross charge distribution of a molecule as a function of pH is calculated as well. The dominant species is indicated in several tissues relevant to drug absorption and distribution. (d) Chemical structures showing charge at pH 7 with the pKa of the amino and carboxylic acid groups labelled. (e) Lewis structures illustrating the effect of N-acetylation on the pKa of nitrogen atom. Resonance delocalization of the lone pair electrons in the amide greatly decreases the basicity of the nitrogen relative to the amine, making the molecule neutral or charged, respectively.
Figure 2
Figure 2
N-acetyl-l-leucine is not transported by either the l-type amino acid transporter (LAT1) nor the peptide transporter (PepT1). Concentration–response curves for the uptake of N-acetyl-l-leucine by (a) LAT1 and (b) PepT1. Concentration-inhibition curves for the inhibition of uptake of known substrates (c) gabapentin (10 µM) for LAT1 and (d) dipeptide Gly-Sar (50 µM) for PepT1. DMSO (0.5%) was the solvent control (SC) and the known inhibitor for the positive control (PC) was JPH203 (10 µM) for LAT1 and losartan (200 µM) for PepT1. Data were fit to either the Michaelis–Menten equation for uptake or the Hill equation for inhibition using the solvent control to define the top and the positive control inhibitor to define the bottom. Symbols represent the mean ± SEM, n = 3. When the error bars are smaller than the symbol, they are not visible.
Figure 3
Figure 3
N-acetyl-l-leucine is transported by organic anion transporters (OAT). Concentration–response curves for the uptake of N-acetyl-l-leucine by (a) OAT1 and (b) OAT3. Concentration-inhibition curves for the inhibition of uptake of known substrates (c) chlorothiazide for OAT1 (3 µM) and (d) estrone-3-sulfate for OAT3 (2 µM). DMSO (0.5%) was the solvent control (SC) and the known inhibitor diclofenac (100 µM) was the positive control (PC). Data were fit to either the Michalis-Menten equation for uptake or the Hill equation for inhibition using the solvent control to define the top and the positive control inhibitor to define the bottom. Symbols represent the mean ± SEM, n = 3. When the error bars are smaller than the symbol, they are not visible.
Figure 4
Figure 4
Both enantiomers of N-acetyl-leucine are transported by the monocarboxylate transporter (MCT1) but only the l-enantiomer is metabolized. Concentration–response curves for the uptake of (a) N-acetyl-l-leucine and (b) N-acetyl-d-leucine. Concentration-inhibition curves for the inhibition of uptake of the known substrate (c, d) 2-thiophene glyoxylate (500 µM). DMSO (0.5%) was the solvent control (SC) and the known inhibitor phloretin (500 µM) was the positive control (PC). (e) Chemical structure of deuterated N-acetyl-d,l-leucine incubated with cellular fraction S9 from liver to determine metabolism using liquid chromatography and mass spectrometry. (f, g) Time courses for loss of deuterated N-acetyl-d,l-leucine (1 µM initial concentration) and appearance of deuterated L-leucine for extracts derived from human and mouse livers. Data are colour-coded according to the chemical structures and names shown in e with deuterated N-acetyl-d,l-leucine in blue and deuterated l-leucine in red. (h) Concentration versus initial velocities relationship for metabolism yielded a Km values of 216 and 69 µM and Vmax values of 6.8 and 2.6 µmol/min/mg protein for human and mouse, respectively. Data were fit to the Michalis–Menten equation for transporter uptake and metabolism or the Hill equation for transporter inhibition using the solvent control to define the top and the positive control inhibitor to define the bottom. Symbols represent the mean ± SEM, n = 3. When the error bars are smaller than the symbol, they are not visible.

References

    1. Neuzil E, Ravaine S, Cousse H. La N-acétyl-DL-leucine, médicament symptomatique de vertigineux. Bull. Soc. Pharm. Bordeaux. 2002;141:15–38.
    1. Strupp M, et al. Effects of acetyl-DL-leucine in patients with cerebellar ataxia: A case series. J. Neurol. 2013;260:2556–2561. doi: 10.1007/s00415-013-7016-x.
    1. Schniepp R, et al. Acetyl-DL-leucine improves gait variability in patients with cerebellar ataxia-a case series. Cerebellum Ataxias. 2016;3:8. doi: 10.1186/s40673-016-0046-2.
    1. Kalla R, Strupp M. Aminopyridines and acetyl-DL-leucine: new therapies in cerebellar disorders. Curr. Neuropharmacol. 2019;17:7–13. doi: 10.2174/1570159X16666180905093535.
    1. Platt F, Strupp M. An anecdotal report by an Oxford basic neuroscientist: Effects of acetyl-DL-leucine on cognitive function and mobility in the elderly. J. Neurol. 2016;263:1239–1240. doi: 10.1007/s00415-016-8048-9.
    1. Bremova T, et al. Acetyl-DL-leucine in Niemann-Pick type C: A case series. Neurology. 2015;85:1368–1375. doi: 10.1212/WNL.0000000000002041.
    1. Cortina-Borja M, et al. Annual severity increment score as a tool for stratifying patients with Niemann-Pick disease type C and for recruitment to clinical trials. Orphanet. J. Rare Dis. 2018;13:143. doi: 10.1186/s13023-018-0880-9.
    1. Kaya E, et al. Acetyl-leucine slows disease progression in lysosomal storage disorders. Brain Commun. 2020 doi: 10.1093/braincomms/fcaa148.
    1. Kaya E, et al. Beneficial effects of acetyl-DL-leucine (ADLL) in a mouse model of Sandhoff disease. J. Clin. Med. 2020;9:1050. doi: 10.3390/jcm9041050.
    1. Strupp M, Bayer O, Feil K, Straube A. Prophylactic treatment of migraine with and without aura with acetyl-DL-leucine: A case series. J. Neurol. 2019;266:525–529. doi: 10.1007/s00415-018-9155-6.
    1. Schoser B, Schnautzer F, Bremova T, Strupp M. Treatment of restless legs syndrome with acetyl-DL-leucine: Accidental findings and a small case series. Eur. J. Neurol. 2019;26:694. doi: 10.1111/ene.13776.
    1. Fields T, et al. A master protocol to investigate a novel therapy acetyl-L-leucine for three ultra-rare neurodegenerative diseases: Niemann-Pick type C, the GM2 gangliosidoses, and ataxia telangiectasia. Trials. 2021;22:84. doi: 10.1186/s13063-020-05009-3.
    1. Churchill GC, Strupp M, Galione A, Platt FM. Unexpected differences in the pharmacokinetics of N-acetyl-DL-leucine enantiomers after oral dosing and their clinical relevance. PLoS ONE. 2020;15:e0229585. doi: 10.1371/journal.pone.0229585.
    1. International Transporter Consortium et al. Membrane transporters in drug development. Nat. Rev. Drug. Discov. 2010;9:215–236. doi: 10.1038/nrd3028.
    1. Keogh JP. Membrane transporters in drug development. Adv. Pharmacol. 2012;63:1–42. doi: 10.1016/B978-0-12-398339-8.00001-X.
    1. Missner A, Pohl P. 110 years of the Meyer-Overton rule: Predicting membrane permeability of gases and other small compounds. ChemPhysChem. 2009;10:1405–1414. doi: 10.1002/cphc.200900270.
    1. Lipinski CA. Drug-like properties and the causes of poor solubility and poor permeability. J. Pharmacol. Toxicol. Methods. 2000;44:235–249. doi: 10.1016/S1056-8719(00)00107-6.
    1. Camenisch G, Folkers G, van de Waterbeemd H. Review of theoretical passive drug absorption models: Historical background, recent developments and limitations. Pharm. Acta Helv. 1996;71:309–327. doi: 10.1016/S0031-6865(96)00031-3.
    1. Gleeson MP, Hersey A, Montanari D, Overington J. Probing the links between in vitro potency, ADMET and physicochemical parameters. Nat. Rev. Drug Discov. 2011;10:197–208. doi: 10.1038/nrd3367.
    1. Leeson PD, Springthorpe B. The influence of drug-like concepts on decision-making in medicinal chemistry. Nat. Rev. Drug. Discov. 2007;6:881–890. doi: 10.1038/nrd2445.
    1. Waring MJ. Defining optimum lipophilicity and molecular weight ranges for drug candidates: Molecular weight dependent lower logD limits based on permeability. Bioorg. Med. Chem. Lett. 2009;19:2844–2851. doi: 10.1016/j.bmcl.2009.03.109.
    1. van de Waterbeemd H, Camenisch G, Folkers G, Chretien JR, Raevsky OA. Estimation of blood-brain barrier crossing of drugs using molecular size and shape, and H-bonding descriptors. J. Drug Target. 1998;6:151–165. doi: 10.3109/10611869808997889.
    1. Walter A, Gutknecht J. Monocarboxylic acid permeation through lipid bilayer membranes. J. Membr. Biol. 1984;77:255–264. doi: 10.1007/BF01870573.
    1. Neuhoff S, Ungell A-L, Zamora I, Artursson P. pH-Dependent passive and active transport of acidic drugs across Caco-2 cell monolayers. Eur. J. Pharm. Sci. 2005;25:211–220. doi: 10.1016/j.ejps.2005.02.009.
    1. Benard P, Cousse H, Bengone T, Germain C. Autoradiography in brain of Macaca fascicularis monkeys after injection of acetyl-DL-leucine [2-14C] (Tanganil) Eur. J. Drug Metab. Pharmacokinet. 2001;26:71–76. doi: 10.1007/BF03190379.
    1. Sugano K, et al. Coexistence of passive and carrier-mediated processes in drug transport. Nat. Rev. Drug Discov. 2010;9:597–614. doi: 10.1038/nrd3187.
    1. Kanai Y, et al. Expression cloning and characterization of a transporter for large neutral amino acids activated by the heavy chain of 4F2 antigen (CD98) J. Biol. Chem. 1998;273:23629–23632. doi: 10.1074/jbc.273.37.23629.
    1. Krehbiel CR, Matthews JC. Absorption of amino acids and peptides. In: D’Mello JPF, editor. Amino Acids in Animal Nutrition. CABI Publishing; 2003. pp. 41–70.
    1. Scalise M, Galluccio M, Console L, Pochini L, Indiveri C. The human SLC7A5 (LAT1): The intriguing histidine/large neutral amino acid transporter and its relevance to human health. Front. Chem. 2018;6:243. doi: 10.3389/fchem.2018.00243.
    1. Soares-da-Silva P, Serrão MP. High- and low-affinity transport of L-leucine and L-DOPA by the hetero amino acid exchangers LAT1 and LAT2 in LLC-PK1 renal cells. Am. J. Physiol. Renal. Physiol. 2004;287:F252–261. doi: 10.1152/ajprenal.00030.2004.
    1. del Amo EM, Urtti A, Yliperttula M. Pharmacokinetic role of L-type amino acid transporters LAT1 and LAT2. Eur. J. Pharm. Sci. 2008;35:161–174. doi: 10.1016/j.ejps.2008.06.015.
    1. Fagerberg L, et al. Analysis of the human tissue-specific expression by genome-wide integration of transcriptomics and antibody-based proteomics. Mol. Cell Proteomics. 2014;13:397–406. doi: 10.1074/mcp.M113.035600.
    1. Nicklin P, et al. Bidirectional transport of amino acids regulates mTOR and autophagy. Cell. 2009;136:521–534. doi: 10.1016/j.cell.2008.11.044.
    1. Bröer S, Fairweather SJ. Amino acid transport across the mammalian intestine. Compr. Physiol. 2018;9:343–373. doi: 10.1002/cphy.c170041.
    1. Pochini L, Scalise M, Galluccio M, Indiveri C. Membrane transporters for the special amino acid glutamine: Structure/function relationships and relevance to human health. Front. Chem. 2014;2:61. doi: 10.3389/fchem.2014.00061.
    1. Nagamori S, et al. Structure-activity relations of leucine derivatives reveal critical moieties for cellular uptake and activation of mTORC1-mediated signaling. Amino Acids. 2016;48:1045–1058. doi: 10.1007/s00726-015-2158-z.
    1. Sawada K, Terada T, Saito H, Hashimoto Y, Inui KI. Recognition of L-amino acid ester compounds by rat peptide transporters PEPT1 and PEPT2. J. Pharmacol. Exp. Ther. 1999;291:705–709.
    1. Thompson BR, Shi J, Zhu H-J, Smith DE. Pharmacokinetics of gemcitabine and its amino acid ester prodrug following intravenous and oral administrations in mice. Biochem. Pharmacol. 2020;180:114127. doi: 10.1016/j.bcp.2020.114127.
    1. Brandsch M, Knütter I, Leibach FH. The intestinal H+/peptide symporter PEPT1: Structure-affinity relationships. Eur. J. Pharm. Sci. 2004;21:53–60. doi: 10.1016/S0928-0987(03)00142-8.
    1. Rubio-Aliaga I, Daniel H. Peptide transporters and their roles in physiological processes and drug disposition. Xenobiotica. 2008;38:1022–1042. doi: 10.1080/00498250701875254.
    1. Bloch K, Borek E. Biological acetylation of natural amino acids. J. Biol. Chem. 1946;164:483. doi: 10.1016/S0021-9258(18)43087-6.
    1. Koepsell H, Endou H. The SLC22 drug transporter family. Pflugers Arch. 2004;447:666–676. doi: 10.1007/s00424-003-1089-9.
    1. Lin L, Yee SW, Kim RB, Giacomini KM. SLC transporters as therapeutic targets: Emerging opportunities. Nat. Rev. Drug Discov. 2015;14:543–560. doi: 10.1038/nrd4626.
    1. Nigam SK, et al. The organic anion transporter (OAT) family: A systems biology perspective. Physiol. Rev. 2015;95:83–123. doi: 10.1152/physrev.00025.2013.
    1. Halestrap AP, Wilson MC. The monocarboxylate transporter family: Role and regulation. IUBMB Life. 2012;64:109–119. doi: 10.1002/iub.572.
    1. Enerson BE, Drewes LR. Molecular features, regulation, and function of monocarboxylate transporters: Implications for drug delivery. J. Pharm. Sci. 2003;92:1531–1544. doi: 10.1002/jps.10389.
    1. Puri S, Juvale K. Monocarboxylate transporter 1 and 4 inhibitors as potential therapeutics for treating solid tumours: A review with structure-activity relationship insights. Eur. J. Med. Chem. 2020;199:112393. doi: 10.1016/j.ejmech.2020.112393.
    1. Wang G, et al. Intestinal OCTN2- and MCT1-targeted drug delivery to improve oral bioavailability. Asian J. Pharm. Sci. 2020;15:158–173. doi: 10.1016/j.ajps.2020.02.002.
    1. Cheng Y, Prusoff WH. Relationship between the inhibition constant (Ki) and the concentration of inhibitor which causes 50 per cent inhibition (I50) of an enzymatic reaction. Biochem. Pharmacol. 1973;22:3099–3108. doi: 10.1016/0006-2952(73)90196-2.
    1. Bloch K, Rittenberg D. The metabolism of acetylamino acids. J. Biol. Chem. 1947;169:467–476. doi: 10.1016/S0021-9258(17)30863-3.
    1. Sheffner AL, et al. Metabolic studies with acetylcysteine. Biochem. Pharmacol. 1966;15:1523–1535. doi: 10.1016/0006-2952(66)90197-3.
    1. Neuhäuser M, Wandira JA, Göttmann U, Bässler KH, Langer K. Utilization of N-acetyl-L-tyrosine and glycyl-L-tyrosine during long-term parenteral nutrition in the growing rat. Am. J. Clin. Nutr. 1985;42:585–596. doi: 10.1093/ajcn/42.4.585.
    1. ImHaesook A, MeyerPaul D, Stegink LD. N-acetyl-L-tyrosine as a tyrosine source during total parenteral nutrition in adult rats. Pediatr. Res. 1985;19:514–518. doi: 10.1203/00006450-198506000-00002.
    1. Birnbaum SM, Levintow L, Kingsley RB, Greenstein JP. Specificity of amino acid acylases. J. Biol. Chem. 1952;194:455–470. doi: 10.1016/S0021-9258(18)55898-1.
    1. Gregori-Puigjané E, et al. Identifying mechanism-of-action targets for drugs and probes. Proc. Natl. Acad. Sci. USA. 2012;109:11178–11183. doi: 10.1073/pnas.1204524109.
    1. Lee AJ, et al. A (14)C-leucine absorption, distribution, metabolism and excretion (ADME) study in adult Sprague-Dawley rat reveals β-hydroxy-β-methylbutyrate as a metabolite. Amino Acids. 2015;47:917–924. doi: 10.1007/s00726-015-1920-6.
    1. Kalogeropoulou D, Lafave L, Schweim K, Gannon MC, Nuttall FQ. Leucine, when ingested with glucose, synergistically stimulates insulin secretion and lowers blood glucose. Metabolism. 2008;57:1747–1752. doi: 10.1016/j.metabol.2008.09.001.
    1. Eriksson T, Björkman S, Höglund P. Clinical pharmacology of thalidomide. Eur. J. Clin. Pharmacol. 2001;57:365–376. doi: 10.1007/s002280100320.
    1. Smith QR, Momma S, Aoyagi M, Rapoport SI. Kinetics of neutral amino acid transport across the blood-brain barrier. J. Neurochem. 1987;49:1651–1658. doi: 10.1111/j.1471-4159.1987.tb01039.x.
    1. Ananieva EA, Powell JD, Hutson SM. Leucine metabolism in T cell activation: mTOR signaling and beyond. Adv. Nutr. 2016;7:798S–805S. doi: 10.3945/an.115.011221.
    1. Tighilet B, Leonard J, Bernard-Demanze L, Lacour M. Comparative analysis of pharmacological treatments with N-acetyl-DL-leucine (Tanganil) and its two isomers (N-acetyl-L-leucine and N-acetyl-D-leucine) on vestibular compensation: Behavioral investigation in the cat. Eur. J. Pharmacol. 2015;769:342–349. doi: 10.1016/j.ejphar.2015.11.041.
    1. te Vruchte D, Galione A, Strupp M, Mann M. Effects of N-acetyl-leucine and its enantiomers in Niemann-Pick disease type C cells. BioRxiv. 2019 doi: 10.1101/826222.
    1. Brandsch M, Knütter I, Bosse-Doenecke E. Pharmaceutical and pharmacological importance of peptide transporters. J. Pharm. Pharmacol. 2008;60:543–585. doi: 10.1211/jpp.60.5.0002.
    1. Poole RC, Halestrap AP. Transport of lactate and other monocarboxylates across mammalian plasma membranes. Am. J. Physiol. 1993;264:C761–782. doi: 10.1152/ajpcell.1993.264.4.C761.
    1. Levy S, Kafri M, Carmi M, Barkai N. The competitive advantage of a dual-transporter system. Science. 2011;334:1408–1412. doi: 10.1126/science.1207154.
    1. Pade V, Stavchansky S. Link between drug absorption solubility and permeability measurements in Caco-2 cells. J. Pharm. Sci. 1998;87:1604–1607. doi: 10.1021/js980111k.
    1. VanWert AL, Gionfriddo MR, Sweet DH. Organic anion transporters: Discovery, pharmacology, regulation and roles in pathophysiology. Biopharm. Drug Dispos. 2010;31:1–71.
    1. Vibert N, Vidal PP. In vitro effects of acetyl-DL-leucine (Tanganil) on central vestibular neurons and vestibulo-ocular networks of the guinea-pig. Eur. J. Neurosci. 2001;13:735–748. doi: 10.1046/j.0953-816x.2000.01447.x.
    1. Kennedy BE, et al. Adaptations of energy metabolism associated with increased levels of mitochondrial cholesterol in Niemann-Pick type C1-deficient cells. J. Biol. Chem. 2014;289:16278–16289. doi: 10.1074/jbc.M114.559914.
    1. Jha MK, et al. Metabolic Connection of inflammatory pain: pivotal role of a pyruvate dehydrogenase kinase-pyruvate dehydrogenase-lactic acid axis. J. Neurosci. 2015;35:14353–14369. doi: 10.1523/JNEUROSCI.1910-15.2015.
    1. Oláh J, et al. Increased glucose metabolism and ATP level in brain tissue of Huntington’s disease transgenic mice. FEBS J. 2008;275:4740–4755. doi: 10.1111/j.1742-4658.2008.06612.x.
    1. Sun S, Li H, Chen J, Qian Q. Lactic acid: No longer an inert and end-product of glycolysis. Physiology. 2017;32:453–463. doi: 10.1152/physiol.00016.2017.
    1. Patet C, Suys T, Carteron L, Oddo M. Cerebral lactate metabolism after traumatic brain injury. Curr. Neurol. Neurosci. Rep. 2016;16:31. doi: 10.1007/s11910-016-0638-5.
    1. Hashimoto T, Hussien R, Oommen S, Gohil K, Brooks GA. Lactate sensitive transcription factor network in L6 cells: Activation of MCT1 and mitochondrial biogenesis. FASEB J. 2007;21:2602–2612. doi: 10.1096/fj.07-8174com.
    1. Weimer M, et al. The impact of data transformations on concentration-response modeling. Toxicol. Lett. 2012;213:292–298. doi: 10.1016/j.toxlet.2012.07.012.

Source: PubMed

3
Se inscrever