You are on page 1of 4

Skeletal muscle mitochondrial deciency does not mediate insulin resistance13

John O Holloszy
ABSTRACT Patients with type 2 diabetes, insulin-resistant obese individuals, and insulin-resistant offspring of patients with diabetes have 30% less mitochondria in their skeletal muscles than age-matched healthy controls. It has been hypothesized that this deciency of mitochondria mediates insulin resistance by impairing the ability of muscle to oxidize fatty acids (FAs). However, a 30% decrease in mitochondria should not impair the ability of muscle to oxidize FAs because the capacity of muscle to oxidize substrate is far in excess of what is needed to supply energy in the basal state, ie, in resting muscle. In pathologic states in which mitochondrial content/ function is so severely impaired as to limit substrate oxidation in resting muscle, glucose uptake and insulin action are actually enhanced. Recent studies have shown that feeding rodents high-fat diets and raising FA concentrations results in muscle insulin resistance despite an increase muscle mitochondria that enhances the capacity for fat oxidation. Furthermore, it was recently shown that skeletal muscle mitochondrial capacity for oxidative phosphorylation in Asian Indians with type 2 diabetes is the same as in nondiabetic Indians and higher than in healthy European Americans. In light of this evidence, it seems highly unlikely that mitochondrial deciency causes muscle insulin resistance. Am J Clin Nutr 2009;89(suppl):463S6S.

30% less mitochondria than normal was conrmed and extended to obese insulin-resistant individuals and to insulinresistant offspring of diabetic parents in subsequent studies (39). As a result of these ndings, the concept that mitochondrial deciency/dysfunction causes muscle insulin resistance appears to have been widely accepted (1012). The mechanism that has been proposed to mediate the insulin resistance is a decrease in the capacity to oxidize fat, resulting in accumulation of intramuscular lipids (2, 10, 11). In support of this concept, it has been reported that both substrate oxidation, determined noninvasively by using [13C]magnetic resonance spectroscopy to measure incorporation 13C from [13C]acetate into glutamate, and the rate of mitochondrial oxidative phosphorylation is 30% lower in resting muscle of insulin-resistant individuals than in normal individuals (6, 13). In keeping with the concept that mitochondrial deciency causes insulin resistance, it has also been reported that feeding rodents high-fat diets of the type that result in muscle insulin resistance results in decreases in muscle peroxisome proliferator activated receptor c coactivator 1a (PGC1a) mRNA (1416) and protein (14) concentrations, as well as in the levels of a range of mitochondrial enzymes (14).
REGULATION OF SUBSTRATE OXIDATION/OXIDATIVE PHOSPHORYLATION

Downloaded from ajcn.nutrition.org by guest on May 1, 2013

INTRODUCTION

The incidence of obesity, metabolic syndrome, and type 2 diabetes has attained epidemic proportions and is, arguably, our most serious current public health problem. Because insulin resistance is the primary cause of type 2 diabetes, the mechanism responsible for decreased insulin action in prediabetes and type 2 diabetes has been the subject of intense investigation. One line of research on this topic has led to the hypothesis that a deciency and/or dysfunction of muscle mitochondria is responsible for the insulin resistance that leads to type 2 diabetes. The purpose of this review is to marshal the evidence against this concept.
BACKGROUND

The rate of substrate oxidation in muscle is regulated by the rate of ATP breakdown/ADP production, which is determined by the cells energy requirement (17). In resting muscle cells, the demand for energy/ATP is determined by housekeeping functions, such as protein synthesis and maintenance of membrane potential, and is very low relative to the maximal capacity of muscle to oxidize substrate. Increasing the supply of FAs or glucose to resting muscle does not increase the rate of substrate oxidation above what is needed for ATP resynthesis, although it can change the relative proportions of these substrates that are oxidized.
1

Kelley and coworkers (1, 2) found that the muscles of patients with type 2 diabetes contain less mitochondria than those of age-matched insulin-sensitive individuals. They hypothesized that this reduction in mitochondrial content impairs the ability of muscle to oxidize fatty acids (FAs), and that this results in intramuscular lipid accumulation and insulin resistance (1, 2). The nding that muscles of patients with type 2 diabetes contain

From the Division of Geriatrics and Nutritional Sciences, Washington University School of Medicine, St. Louis, MO. 2 Presented at the symposium The Emerging Interplay among Muscle Mitochondrial Function, Nutrition, and Disease, held at Experimental Biology 2008, San Diego, CA, 5 April 2008. 3 Reprints not available. Address correspondence to JO Holloszy, Washington University School of Medicine, 4566 Scott Avenue, Campus Box 8113, St Louis, MO 63110. E-mail: jhollosz@dom.wustl.edu. First published online December 3, 2008; doi: 10.3945/ajcn.2008.26717C.

Am J Clin Nutr 2009;89(suppl):463S6S. Printed in USA. 2009 American Society for Nutrition

463S

464S

HOLLOSZY

Skeletal muscle contains sufcient mitochondria to make possible a 150-fold increase in oxygen uptake per kilogram of muscle during strenuous exercise in well-trained young individuals (18). Middle-aged patients with type 2 diabetes are usually in poor physical condition with a low maximal oxygen uptake capacity (4). However, even if one makes the very conservative assumption that the maximal oxygen uptake capacity of the muscles of insulin-resistant patients is only 25% as great as that of young trained men, this still represents a 3040-fold increase above resting values. Because the capacity of muscle to oxidize substrate is so far in excess of what is needed to supply the energy needs of resting muscle, it seems clear that a 30% decrease in mitochondrial content should have no effect on the ability of resting muscle to oxidize fat. If oxygen uptake/substrate oxidation is measured under state 3 conditions, ie, when the availability of ADP, inorganic phosphate (Pi), and substrate is not limiting, in permeabilized muscle bers, muscle homogenates, or the mitochondrial fraction of muscle, one would expect to nd a 30% lower value in diabetic than in control muscle. In other words, maximal oxidative capacity should be decreased in proportion to the lower content of mitochondria. This expected relation is what was found by Boushel et al (9) in muscle from patients with type 2 diabetes, and by Holloway et al (19) in muscle from obese women, compared with normal controls. These investigators evaluated whether there is mitochondrial dysfunction in insulin-resistant muscle and found that, although there was a decrease in mitochondrial content, the remaining mitochondria functioned normally (9, 19). It has also been reported that the increase in oxidative phosphorylation in muscle in response to insulin is lower in muscle from patients with diabetes than in normal muscle (20). The increase in ATP turnover in response to insulin is mediated by the increase in the rate of ADP production that results from increased glucose uptake and ATP utilization in the reactions catalyzed by hexokinase and nucleoside diphosphate kinase, ie, 2 ATPs for each glucose molecule taken up. The smaller increase in oxidative phosphorylation in diabetic muscle in response to insulin is the result of a lower rate of glucose transport into the insulinresistant muscles. On the other hand, the rates of substrate oxidation and oxidative phosphorylation in muscles not stimulated by insulin or exercise are determined by the low rate of ATP utilization/ADP production required for housekeeping functions, not by mitochondrial content, except when a pathologic process has destroyed or damaged most of the mitochondria. In the above context, the reports (6, 13) that substrate oxidation and ATP production rates in muscle are 30% lower in insulinresistant individuals than in normal individuals under basal conditions are puzzling. For such a decrease in resting oxidative metabolism to occur, the energy requirement for housekeeping functions would have to be 30% lower than normal in insulinresistant muscle. Alternatively, if mitochondria are more tightly coupled in insulin-resistant muscle, so that more ATP is formed per oxygen molecule, this could explain a lower rate of substrate oxidation but not a slower rate of ATP formation. If the ability of muscle to oxidize fat in the basal state is not impaired in insulin-resistant individuals, what is the explanation for the well-documented elevation of their intramyocellular lipid (IMCL) content? Clearly, impaired oxidative capacity is not the only mechanism for an increase in IMCL lipid stores, as evi-

denced by the nding that highly trained athletes also frequently have high IMCL despite an increase in muscle mitochondria (2123). The likely cause is increased delivery and/or uptake of FAs in excess of the muscle cells energy requirement.
DOES MITOCHONDRIAL DEFICIENCY/DYSFUNCTION CAUSE MUSCLE INSULIN RESISTANCE?

Although insulin-resistant individuals usually have 30% less mitochondria in their muscles than average, their capacity for aerobic metabolism is still within the normal range (4). Unless they have some additional pathology, patients with type 2 diabetes and insulin-resistant offspring of diabetic parents are not limited in their everyday physical activities. In contrast, a decrease in muscle mitochondrial number, or a mitochondrial pathology/dysfunction sufciently severe to limit the rate of ATP production at rest, will result in such extreme muscle fatigue and weakness that normal physical activity becomes impossible. Contrary to the mitochondrial deciency causes insulin resistance hypothesis, a number of studies have shown that in pathologic states in which mitochondrial deciency is so severe that it limits the rate of substrate oxidation in resting muscle, both basal and insulin-stimulated glucose transport are increased (24, 25) despite a large accumulation of IMCL (24). Hypoxia, which represents the ultimate degree of mitochondrial dysfunction or nonfunction, also increases basal and insulinstimulated glucose transport (26, 27). These responses to extreme mitochondrial dysfunction are mediated by a decrease in the ATP:AMP ratio with activation of AMP-dependent protein kinase (24, 25, 27), which stimulates glucose transport, increases muscle insulin sensitivity, and induces increased expression of the GLUT4 glucose transporter (27, 28). Rodents fed high-fat diets develop muscle insulin resistance within a few weeks. If the high-fat diet is continued long enough, the insulin resistance progresses to the rodent equivalent of metabolic syndrome or type 2 diabetes, depending on the genetic background (2932). The recent nding that, in rats fed high-fat diets, muscle insulin resistance develops concomitantly with an increase in muscle mitochondria and oxidative capacity provides direct evidence against the concept that insulin resistance is mediated by a deciency of muscle mitochondria (33, 34). A recent study showing that skeletal muscle mitochondrial capacity for oxidative phosphorylation in insulin-resistant Asian Indians with type 2 diabetes is the same as in nondiabetic Indians and higher than in healthy Americans of northern European descent also demonstrates that insulin resistance is not due to mitochondrial deciency (35).
HIGH-FAT DIETS CAUSE INSULIN RESISTANCE DESPITE INCREASED MITOCHONDRIAL BIOGENESIS

Downloaded from ajcn.nutrition.org by guest on May 1, 2013

It has been reported that raising FA levels in humans (14) and feeding rodents high-fat diets (1416) cause a decrease in PGC1 mRNA in muscle. In a study by Sparks et al (14), feeding mice a high-fat diet for 3 weeks was reported to result in 90% decreases in the mRNA levels of PGC1a and a number of mitochondrial enzymes, and 40% decreases in the protein levels of PGC1a and cytochrome c. On the other hand, earlier studies showed that high-fat diets induce increases in mitochondrial enzymes (3639). Other, more recent studies have shown that

MITOCHONDRIAL DEFICIENCY AND INSULIN RESISTANCE

465S

raising serum FAs to very high levels with a high-fat diet plus heparin (40), or a more modest increase in FA induced by a highfat diet alone (33, 34), induces an increase in mitochondrial biogenesis as evidenced by increases in the protein levels of a range of mitochondrial enzymes, in the capacity to oxidize FAs, and in mitochondrial DNA copy number. Additional evidence that FAs stimulate mitochondrial biogenesis comes from a study showing that overexpression of lipoprotein lipase in muscle results in a large increase in mitochondria (41) and from the observation that lowering serum free FA with the use of acipimox results in a decrease in expression of a number of mitochondrial markers (42). The FA-induced increase in mitochondrial biogenesis is mediated by an increase in PGC1a protein (33, 34, 43) that occurs in the absence of an increase in PGC1a mRNA (34, 40). In contrast to the transcriptionally mediated increases in PGC1a expression induced by stimuli such as exercise (4447), calcium (48), and cold (49), which occur rapidly (within hours), the posttranscriptionally mediated increase in PGC1a induced by an increase in FAs occurs very slowly (4 wk) (34). Previous studies documented that activation or overexpression of peroxisome proliferator activated receptor d (PPARd) in skeletal muscle induces an increase in mitochondrial biogenesis (50, 51). The authors concluded that this effect was mediated directly by PPARd, because increasing PPARd activity or expression did not result in an increase in PGC1a mRNA, and PGC1a protein was not measured (50, 51). FAs are natural ligands/activators of PPARd (52), and raising serum FAs results in activation of PPARd (40). Therefore, it is probable that activation of PPARd is the initial step in the pathway by which raising FA concentrations causes an increase in mitochondrial biogenesis. However, PPARd regulates the expression of only a subset of mitochondrial proteins. An increase in mitochondrial biogenesis requires the coordinated expression of a large number of genes encoded in the nucleus, as well as the genes encoded in the mitochondrial genome. This process requires the activation of various transcription factors, including nuclear respiratory factor 1 (NRF-1), NRF-2, mitochondrial transcription factor A (TFAM), estrogen-related receptors (ERRs), and myocyte enhancer factors (MEFs) (5355) in addition to PPARd. The transcription coactivator PGC1a, which activates the NRFs, ERRs, PPARs, and MEFs, mediates the coordinated increase in transcription of the genes encoding mitochondrial proteins (54, 55). The interpretation that FAs induce an increase in mitochondrial biogenesis by inducing a posttranscriptional increase in PGC1a is supported by the nding that overexpression of PPARd in skeletal muscle induces an increase in PGC1a expression in the absence of an increase in PGC1a mRNA (34). The mechanism by which PPARd increases PGC1 protein expression is still unknown. If the decrease in muscle mitochondria associated with insulin resistance and its consequences is not mediated by FAs, what is responsible? One possibility raised by Asmann and coworkers (8) is that muscle mitochondrial dysfunction in type 2 diabetes is not an intrinsic defect but instead a functional defect related to impaired response to insulin. Support for this possibility is provided by a study on mice that developed insulin resistance and diabetes in response to a high-fat, high-sucrose diet (56). In that study, insulin resistance developed within 1 mo on the diet, whereas mitochondria remained normal. However, an extended

diet intervention that induced diabetes resulted in decreased mitochondrial biogenesis, structure, and function, which the authors attributed to oxidative stress. Further support is provided by a study reporting that rosiglitazone administration in patients with type 2 diabetes restored muscle PGC1a and PPARd gene expression and succinate dehydrogenase activity to normal levels (57). A second possibility, that the susceptibility to development of type 2 diabetes and low muscle content of mitochondria are genetically linked, is suggested by the nding that insulinresistant offspring of parents with type 2 diabetes have or show a 30% reduction in muscle mitochondria before developing hyperglycemia or other metabolic abnormalities (other than insulin resistance) (6, 7). A third possibility is that exercise deciency is the cause of the mitochondrial deciency, because obesity, insulin resistance, and type 2 diabetes are to a large extent the result of exercise deciency in genetically predisposed individuals whose energy intake is not balanced by energy expenditure. In support of this possibility, exercise training has been shown to normalize muscle mitochondrial content in patients with type 2 diabetes (58). In conclusion, although the mechanism responsible for the nding that insulin-resistant individuals have 30% less mitochondria in their muscles than normal individuals has not yet been determined, it seems clear from the information reviewed above that this mitochondrial deciency is not responsible for muscle insulin resistance. (Other articles in this supplement to the Journal include references 59 and 60.)
The author had no conicts of interest.

Downloaded from ajcn.nutrition.org by guest on May 1, 2013

REFERENCES
1. He J, Watkins S, Kelley DE. Skeletal muscle lipid content and oxidative enzyme activity in relation to muscle ber type in type 2 diabetes and obesity. Diabetes 2001;50:81723. 2. Kelley DE, He J, Menshikova EV, Ritov VB. Dysfunction of mitochondria in human skeletal muscle in type 2 diabetes. Diabetes 2002;51: 294450. 3. Patti ME, Butte AJ, Crunkhorn S, et al. Coordinated reduction of genes of oxidative metabolism in humans with insulin resistance and diabetes: potential role of PGC1 and NRF1. Proc Natl Acad Sci USA 2003;100: 846671. 4. Mootha VK, Lindgren CM, Eriksson K-F, et al. PGC-1a-responsive genes involved in oxidative phosphorylation are coordinately downregulated in human diabetes. Nat Genet 2003;34:26773. 5. Sreekumar R, Halvatsiotis P, Schimke JC, Nair KS. Gene expression prole in skeletal muscle of type 2 diabetes and the effect of insulin treatment. Diabetes 2002;51:191320. 6. Petersen KF, Dufour S, Befroy D, Garcia R, Shulman GI. Impaired mitochondrial activity in the insulin-resistant offspring of patients with type 2 diabetes. N Engl J Med 2004;350:66471. 7. Morino K, Petersen KF, Dufour S, et al. Reduced mitochondrial density and increased IRS-1 serine phosphorylation in muscle of insulin-resistant offspring of type 2 diabetic parents. J Clin Invest 2005;115:358793. 8. Asmann YW, Stump CS, Short KR, et al. Skeletal muscle mitochondrial functions, mitochondrial DNA copy numbers, and gene transcript proles in type 2 diabetic and nondiabetic subjects at equal levels of low or high insulin and euglycemia. Diabetes 2006;55:330919. 9. Boushel R, Gnaiger E, Schjerling P, Skovbro M, Kraunsoe R, Dela F. Patients with type 2 diabetes have normal mitochondrial function in skeletal muscle. Diabetologia 2007;50:7906. 10. Lowell BB, Shulman GI. Mitochondrial dysfunction and type 2 diabetes. Science 2005;307:3847. 11. Morino K, Petersen KF, Shulman GI. Molecular mechanisms of insulin resistance in humans and their potential links with mitochondrial dysfunction. Diabetes 2006;55:S915. 12. Taylor R. Causation of type 2 diabetes: the Gordian knot unravels. N Engl J Med 2004;350:63941.

466S

HOLLOSZY
36. Miller WC, Bryce GR, Conlee RK. Adaptations to a high-fat diet that increase exercise endurance in male rats. J Appl Physiol 1984;56:7883. 37. Nemeth PM, Rosser BWC, Choksi RM, Norris BJ, Baker KM. Metabolic response to a high-fat diet in neonatal and adult rat muscle. Am J Physiol 1992;262:C2826. 38. McAinch AJ, Lee J-S, Bruce CR, Tunstall RJ, Hawley JA, CameronSmith D. Dietary regulation of fat oxidative gene expression in different skeletal muscle ber types. Obes Res 2003;11:14719. 39. Simi B, Sempore B, Mayet M-H, Favier RJ. Additive effects of training and high-fat diet on energy metabolism during exercise. J Appl Physiol 1991;71:197203. 40. Garcia-Roves P, Huss JM, Han D-H, et al. Raising plasma fatty acid concentration induces increased biogenesis of mitochondria in skeletal muscle. Proc Natl Acad Sci USA 2007;104:1070913. 41. Levak-Frank S, Radner H, Walsh A, et al. Muscle-specic overexpression of lipoprotein lipase causes a severe myopathy characterized by proliferation of mitochondria and peroxisomes in transgenic mice. J Clin Invest 1995;96:97686. 42. Bajaj M, Medina-Navarro R, Suraamornkul S, Meyer C, DeFronzo RA, Mandarino LJ. Paradoxical changes in muscle gene expression in insulin-resistant subjects after sustained reduction in plasma free fatty acid concentration. Diabetes 2007;56:74352. 43. Hoeks J, Briede JJ, de Vogel J, et al. Mitochondrial function, content and ROS production in rat skeletal muscle: effect of high-fat feeding. FEBS Lett 2008;582:5106. 44. Baar K, Song Z, Semenkovich CF, et al. Skeletal muscle overexpression of nuclear respiratory factor 1 increases glucose transport capacity. FASEB J 2003;17:166673. 45. Pilegaard H, Saltin B, Neufer PD. Exercise induces transient transcriptional activation of the PGC-1a gene in human skeletal muscle. J Physiol 2003;546:8518. 46. Akimoto T, Pohnert SC, Li P, et al. Exercise stimulates Pgc-1a transcription in skeletal muscle through activation of the p38 MAPK pathway. J Biol Chem 2005;280:1958793. 47. Wright DC, Han D-H, Garcia-Roves PM, Geiger PC, Jones TE, Holloszy JO. Exercise-induced mitochondrial biogenesis begins before the increase in muscle PGC-1a expression. J Biol Chem 2007;282:1949. 48. Wright DC, Geiger PC, Han D-H, Jones TE, Holloszy JO. Calcium induces increases in peroxisome proliferator-activated receptor c coactivator-1a and mitochondrial biogenesis by a pathway leading to p38 mitogen-activated protein kinase activation. J Biol Chem 2007;282:187939. 49. Cao W, Daniel KW, Robidoux J, et al. p38 Mitogen-activated protein kinase is the central regulator of cyclic AMP-dependent transcription of the brown fat uncoupling protein 1 gene. Mol Cell Biol 2004;24:305767. 50. Luquet S, Lopez-Soriano J, Holst D, et al. Peroxisome proliferator-activated receptor d controls muscle development and oxydative capability. FASEB J 2003;17:2299301. 51. Wang Y-X, Zhang C-L, Yu RT, et al. Regulation of muscle ber type and running endurance by PPARd. PLoS Biol 2004;2:15329. 52. Forman BM, Chen J, Evans RM. Hypolipidemic drugs, polyunsaturated fatty acids, and eicosanoids are ligands for peroxisome proliferatoractivated receptors a and d. Proc Natl Acad Sci USA 1997;94:43127. 53. Scarpulla RC. Nuclear control of respiratory gene expression in mammalian cells. J Cell Biochem 2006;97:67383. 54. Lin J, Handschin C, Spiegelman BM. Metabolic control through the PGC-1 family of transcription coactivators. Cell Metab 2005;1:36170. 55. Spiegelman BM, Heinrich R. Biological control through regulated transcriptional coactivators. Cell 2004;119:15767. 56. Bonnard C, Durand A, Peyrol S, et al. Mitochondrial dysfunction results from oxidative stress in the skeletal muscle of diet-induced insulinresistant mice. J Clin Invest 2008;118:789800. 57. Mensink M, Hesselink MK, Russell AP, Schaart G, Sels JP, Schrauwen P. Improved skeletal muscle oxidative enzyme activity and restoration of PGC-1 alpha and PPAR beta/delta gene expression upon rosiglitazone treatment in obese patients with type 2 diabetes mellitus. Int J Obes 2007;31:130210. 58. Toledo FG, Menshikova EV, Ritov VB, et al. Effects of physical activity and weight loss on skeletal muscle mitochondria and relationship with glucose control in type 2 diabetes. Diabetes 2007;56:21427. 59. Holloway GP, Bonen A, Spriet LL. Regulation of skeletal muscle mitochondrial fatty acid metabolism in lean and obese individuals. Am J Clin Nutr 2009;89(suppl):455S62S. 60. Lanza IR, Nair KS. Muscle mitochondrial changes with aging and exercise. Am J Clin Nutr 2009;89(suppl):467S71S.

13. Befroy DE, Petersen KF, Dufour S, et al. Impaired mitochondrial substrate oxidation in muscle of insulin-resistant offspring of type 2 diabetic patients. Diabetes 2007;56:137681. 14. Sparks LM, Xie H, Koza RA, et al. A high-fat diet coordinately downregulates genes required for mitochondrial oxidative phosphorylation in skeletal muscle. Diabetes 2005;54:192633. 15. Crunkhorn S, Dearie F, Mantzoros C, et al. PGC-1 expression is reduced in obesity: potential pathogenic role of saturated fatty acids and p38 MAP kinase activation. J Biol Chem 2007;282:1543950. 16. Koves TR, Li P, An J, et al. Peroxisome proliferator-activated receptor-c co-activator 1a-mediated metabolic remodeling of skeletal myocytes mimics exercise training and reverses lipid-induced mitochondrial inefciency. J Biol Chem 2005;280:3358898. 17. McGilvery RW. Biochemistry: a functional approach. Philadelphia, PA: WB Saunders, 1970. 18. Andersen P, Saltin B. Maximal perfusion of skeletal muscle in man. J Physiol 1985;366:23349. 19. Holloway GP, Thrush AB, Heigenhauser GJF, et al. Skeletal muscle mitochondrial FAT/CD36 content and palmitate oxidation are not decreased in obese women. Am J Physiol Endocrinol Metab 2007;292: E17829. 20. Petersen KF, Dufour S, Shulman GI. Decreased insulin-stimulated ATP synthesis and phosphate transport in muscle of insulin-resistant offspring of type 2 diabetic parents. PLoS Med 2005;2:e233. 21. van Loon LJC, Koopman R, Manders R, van der Weegen W, van Kranenburg GP, Keizer HA. Intramyocellular lipid content in type 2 diabetes patients compared with overweight sedentary men and highly trained endurance athletes. Am J Physiol Endocrinol Metab 2004;287: E55865. 22. Goodpaster BH, He J, Watkins S, Kelley DE. Skeletal muscle lipid content and insulin resistance: evidence for a paradox in endurancetrained athletes. J Clin Endocrinol Metab 2001;86:575561. 23. Hoppeler H, Howard H, Conley K, et al. Endurance training in humans: aerobic capacity and structure of skeletal muscle. J Appl Physiol 1985; 59:3207. 24. Han D-H, Nolte LA, Ju J-S, Coleman T, Holloszy JO, Semenkovich CF. UCP-mediated energy depletion in skeletal muscle increases glucose transport despite lipid accumulation and mitochondrial dysfunction. Am J Physiol 2004;286:E34753. 25. Wredenberg A, Freyer C, Sandstrom ME, et al. Respiratory chain dysfunction in skeletal muscle does not cause insulin resistance. Biochem Biophys Res Commun 2006;350:2027. 26. Cartee GD, Douen AG, Ramlal T, Klip A, Holloszy JO. Stimulation of glucose transport in skeletal muscle by hypoxia. J Appl Physiol 1991;70: 1593600. 27. Fisher JS, Gao J, Han D-H, Holloszy JO, Nolte LA. Activation of AMP kinase enhances sensitivity of muscle glucose transport to insulin. Am J Physiol 2002;282:E1823. 28. Winder WW, Hardie DG. AMP-activated protein kinase, a metabolic master switch: possible roles in type 2 diabetes. Am J Physiol 1999;277: E110. 29. Storlien LH, James DE, Burleigh KM, Chisholm DJ, Kraegen EW. Fat-feeding causes widespread in vivo insulin resistance, decreased energy expenditure, and obesity in rats. Am J Physiol 1986;251: E57683. 30. Rosholt MN, King PA, Horton ES. High-fat diet reduces glucose transporter responses to both insulin and exercise. Am J Physiol 1994; 266:R95101. 31. Han D-H, Hansen PA, Host HH, Holloszy JO. Insulin resistance of muscle glucose transport in rats fed a high-fat diet: a reevaluation. Diabetes 1997;46:17617. 32. Hansen PA, Han D-H, Marshall BA, et al. A high fat diet impairs stimulation of glucose transport in muscle: functional evaluation of potential mechanisms. J Biol Chem 1998;273:2615763. 33. Turner N, Bruce CR, Beale SM, et al. Excess lipid availability increases mitochondrial fatty acid oxidative capacity in muscle: evidence against a role for reduced fatty acid oxidation in lipid-induced insulin resistance in rodents. Diabetes 2007;56:208592. 34. Hancock CR, Han D-H, Chen M, et al. High fat diets cause insulin resistance despite an increase in muscle mitochondria. Proc Natl Acad Sci USA 2008;105:781520. 35. Nair KS, Bigelow ML, Asmann YW, et al. Asian Indians have enhanced skeletal muscle mitochondrial capacity to produce ATP in association with severe insulin resistance. Diabetes 2008;57:116675.

Downloaded from ajcn.nutrition.org by guest on May 1, 2013

You might also like