Literature
Home医源资料库在线期刊循环研究杂志2005年第95卷第5期

Sequelae of Insulin-Resistant Adipose Tissue

来源:循环研究杂志
摘要:Adiposetissuewasthesitewhereexcessenergywasstored,intheformoftriglycerides(TGs),andwherethatenergy,whenneededelsewhereinthebody,wasreleasedintheformoffattyacid(FA)。NormalAdiposeTissueRegulatesPartitioningandUtilizationofLipidsbytheBodyAfterameal,......

点击显示 收起

    The Department of Medicine, Columbia University College of Physicians and Surgeons, New York, NY.

    Abstract

    For many years adipose tissue was viewed as the site where excess energy was stored, in the form of triglycerides (TGs), and where that energy, when needed elsewhere in the body, was released in the form of fatty acids (FAs). Recently, it has become clear that when the regulation of the storage and release of energy by adipose tissue is impaired, plasma FA levels become elevated and excessive metabolism of FA, including storage of TGs, occurs in nonadipose tissues. Most recently, work by several laboratories has made it clear that in addition to FA, adipose tissue communicates with the rest of the body by synthesizing and releasing a host of secreted molecules, collectively designated as adipokines. Several recent reviews have described how these molecules, along with FA, significantly effect total body glucose metabolism and insulin sensitivity. Relatively little attention has been paid to the effects of adipokines on lipid metabolism. In this review, we will describe, in detail, the effects of molecules secreted by adipose tissue, including FA, leptin, adiponectin, resistin, TNF-, IL-6, and apolipoproteins, on lipid homeostasis in several nonadipose tissues, including liver, skeletal muscle, and pancreatic  cells.

    Key Words: lipids  fatty acids  adipose tissue  insulin resistance  cytokines

    Introduction

    For many years adipose tissue was viewed as playing a passive role in total body lipid and energy homeostasis. Adipose tissue was the site where excess energy was stored, in the form of triglycerides (TGs), and where that energy, when needed elsewhere in the body, was released in the form of fatty acid (FA). Work by numerous laboratories during the second half of the 20th century made it clear that under normal conditions, the storage and release of TG and FA, respectively, are both coordinated and tightly regulated so that lipid fuels are stored during the immediate postprandial periods and released during periods of fasting. The biochemistry of key molecules critical to coordinating storage and release of energy, such as lipoprotein lipase (LPL) and hormone sensitive lipase (HSL), was characterized in detail during that period. More recently, it has become clear that when the regulation of storage and release of energy by adipose tissue is impaired, particularly when release of FA becomes dissociated from energy requirements in other organs and tissues, plasma FA levels become elevated and excessive metabolism of FA, including storage of TG, occurs in nonadipose tissues.

    Our initial understanding of how fat tissue communicates with other tissues and organs to integrate total body lipid homeostasis was limited to the effects of circulating FA on hepatic lipid and glucose metabolism. During the last 15 years, however, it has become clear that adipose tissueeCderived FA can affect lipid metabolism in several tissues, including muscle and pancreatic  cells. The expanded view of ectopic metabolism or accumulation of FA or TG is that it causes dysfunction, or lipotoxicity, in these organs and tissues. Most recently, work by several laboratories has further expanded our view of the way adipose tissue communicates with, and affects, total body energy homeostasis. We now know that in addition to FA, adipose tissue communicates with the rest of the body by synthesizing and releasing a host of secreted molecules, collectively designated as adipokines. These molecules, along with FA, have significant effects on total body glucose and lipid metabolism, and insulin sensitivity. Several recent articles have described the effects of the adipokines on glucose metabolism, insulin sensitivity, and inflammation;1eC5 little has been written about their effects on lipid metabolism. In this review, we will combine a detailed description of the effects of FA on total body lipid metabolism with an update on the rapidly evolving understanding of the role of adipokines in lipid homeostasis.

    Normal Adipose Tissue Regulates Partitioning and Utilization of Lipids by the Body

    After a meal, dietary TG is digested by pancreatic lipase in the small intestines and absorbed by intestinal enterocytes as FA and sn-2-monoacylglycerols.6 These lipid molecules reassemble as TG and are then exported into the circulation as chylomicrons. In the fed state, adipose tissue is the major site for uptake of dietary TG FA, after LPL-mediated hydrolysis of chylomicron TG. LPL is produced mainly in adipose tissue and muscle, secreted by those tissues, and transported to the surfaces of capillary endothelial cells in these tissues. LPL binds to proteoglycans at the luminal side of the capillary blood vessels, where it interacts with TG-rich lipoprotein particles7,8 and, in the presence apolipoprotein (apo) CII,9 hydrolyzes TG. ApoCIII inhibits LPL-mediated TG hydrolysis, and the ratio of apoCII to apoCIII may be critical in regulating the delivery of dietary TG to adipose tissue.10 Insulin is a major regulator of LPL synthesis and activity in adipose tissue. In the fed state, whereas LPL is downregulated in muscle, it is upregulated in adipose tissue.7,8 The opposite regulation characterizes the fasted state.

    The FAs that are released in high concentrations by lipolysis of TG-enriched lipoproteins have to be efficiently trapped in the local adipose tissue to prevent high FA flux through the plasma to tissues that do not, at that moment, require FA as an energy source.8 Both inhibition of adipocyte HSL-mediated TG lipolysis during the hyperinsulinemic postprandial period and efficient acylation of newly arrived FA inside adipocytes help create a gradient for FA uptake in adjacent adipose tissue. Physical channeling of FA through the endothelial cells to the adipocyte also occurs during lipolysis.11 Then, FA can be rapidly taken up and trapped, most likely, by a combination of passive diffusion12 and the action of fatty acid transport proteins.13

    Acylation-stimulating protein (ASP) is an adipocyte-secreted protein that may also play an important role in the efficient adipocyte trapping of FA released after lipolysis of lipoprotein TG.14 ASP is generated by the interaction of complement C3 with factors B and D (the latter is also called adipsin). The result is cleavage of the N-terminal of C3, producing C3a, which is then cleaved by plasma carboxypeptidase to produce C3desarg or ASP. Studies in adipocytes indicate that ASP increases glucose uptake (a substrate for glycerol formation) and the activity of diacylglycerol acyltransferase (DGAT), thereby facilitating use of FA for TG synthesis. Additionally, ASP has been shown to increase reesterification of FA released from TG by HSL, as well as reduce HSL-mediated lipolysis. In humans, fasting ASP levels correlate with postprandial TG clearance.15 ASP-deficient mice have delayed postprandial TG clearance (males on a mixed C57BL/6x129Sv background)16 and reduced body weight and fat mass (females on the same mixed background),17 even on the ob/ob background.18 Additionally, intraperitoneal injection of ASP facilitated postprandial TG clearance in several mouse models of obesity and insulin resistance.16 However, other investigators, using the same mouse (they had originally generated a complement C3 knockout mouse that would, by definition, also be deficient in ASP), did not find alterations in energy metabolism or lipid metabolism.19 This controversy remains unresolved at present, although the weight of evidence favors some role for ASP in postprandial TG and FA partitioning into fat.

    As a result of these various processes, 90% to 100% of LPL-released FA is trapped by adipocytes after feeding.20 The efficiency of TGFA uptake by adipose tissue in normal humans is evidenced by the fact that postprandial plasma FFA levels are actually lower than fasting levels.20 This reduction in plasma FA results both from inhibition of adipose tissue HSL-mediated TG lipolysis and the facilitated uptake of chylomicron or very low density lipoprotein (VLDL)eCgenerated FA.

    By contrast, during the fasting state, insulin levels are low, HSL is activated, and catecholamine-stimulated adipose tissue lipolysis becomes unopposed. Additionally, adipose tissue LPL activity is downregulated and muscle LPL is upregulated. Thus, during fasting, FA is released from fat, plasma FA levels rise, and muscle uptake of FA increases. FA flux to liver also increases during fasting, contributing energy for hepatic gluconeogenesis and ketogenesis. During longer periods of fasting, FA becomes the primary fuel in organs where glucose is not an obligatory fuel, including skeletal muscle, heart, and the renal cortex. In this manner, glucose is spared for its required utilization by the central nervous system.

    Insulin-Resistant Adipose Tissue and Dysregulation of Lipid Partitioning and Utilization

    Insulin resistance is traditionally assessed by insulin’s ability to promote normal glucose metabolism. The physiological role of insulin is, however, much broader, and includes the metabolism of all 3 macronutrients (carbohydrates, lipids, and proteins) as well as cellular growth. Insulin’s action on lipid metabolism is analogous to its role in glucose metabolism, ie, promoting anabolism and inhibiting catabolism. Specifically, insulin upregulates LPL and stimulates gene expression of intracellular lipogenic enzymes, such as acetyl-CoA carboxylase (ACC) and fatty acid synthase (FAS).21 In addition, insulin inhibits adipocyte HSL through inhibition of its phosphorylation.22

    In the insulin-resistant state, the responses of both LPL and HSL to insulin are blunted. Thus, with insulin resistance, inefficient trapping of dietary energy occurs both because of decreased LPL-mediated lipolysis of chylomicron-TG and ineffective inhibition of HSL-mediated lipolysis in adipose tissue.23 Postprandial lipemia and elevated plasma FA levels are well-recognized abnormalities in obesity and insulin resistance.24 Reduced adipose tissue uptake and storage of TG FA during the postprandial period results in greater partitioning of dietary lipids to nonadipose tissues, including muscle and liver.25,26 Thus, the metabolic consequence of inadequate insulin action at the adipocyte during the fed state is a condition similar to the normal fasting state, when insulin levels are appropriately low. However, in the fed state, this maladaptive response creates a situation in which postprandial FA are directed to various nonadipose tissues and organs at a time when lipid energy is not needed. The specific consequences of increased FA flux to nonadipose tissues are discussed in more detail in the following sections.

    The ultimate example of the loss of physiological lipid partitioning is total lipodystrophy, a disorder in which there is no adipose tissue.27 Several genes have been identified that can cause total lipodystrophy, including 1-acylglycerol-3-phosphate O-acyltransferase 2 (AGPAT2), which converts 1-acyl glycerol phosphate to 1,2 diacylglycerol phosphate (phosphaditic acid); without AGPAT2, TG synthesis and adipogenesis are disrupted.28,29 A second gene, called seipin, has no known function, and the way in which it causes lipodystrophy is unknown.29 There are partial lipodystrophies, some of which have been associated with mutations in the gene for lamins A and C (LMNA), and in the PPAR gene.30 In total lipodystrophy, plasma FA levels are very high, and TG accumulates in muscle, liver, islet cells, and plasma, resulting in insulin resistance, hypertriglyceridemia, and defective insulin secretion with diabetes.31 It is noteworthy that patients with total lipodystrophy have no intra-abdominal fat, including no visceral fat, and yet have severe insulin resistance. In partial lipodystrophy, ectopic fat deposition occurs in several organs, and there is also visceral fat, hypertriglyceridemia, and insulin resistance.

    Rodent models of lipodystrophy32eC34 have provided important insights regarding the effects of a failure to partition FA to adipose tissue. These models all have ectopic fat, particularly fatty livers, and both insulin resistance and increased secretion of VLDL TG.

    Role of Increased Plasma FA Flux to the Liver and Increased VLDL Secretion

    As noted, insulin resistance in adipose tissue will disrupt normal lipid partitioning in both the fed and fasting states, and nonadipose tissues will be faced with increased FA delivery. Based on blood flow, and a very high capacity to take up plasma FA, the liver will be a major recipient of those FA. It is likely that both "flip-flop" diffusion processes12 and one or more fatty acid "receptors" or "transport proteins"13 are involved in the high capacity of the liver to take up FA. In addition, intracellular proteins that bind or chemically modify FA, such as long-chain acyl-CoA synthetase (ACLS),35,36 are very likely to facilitate the net transport of FA from the plasma into hepatocytes.

    Increased plasma FA flux to the liver as a cause of increased VLDL apoB and TG secretion has been observed both in vitro and in vivo.37eC39 Studies of cultured liver cells have often, but not always, shown FA to stimulate apoB secretion.39 Some of the inconsistencies seem to derive from the fact that when studying primary rodent hepatocytes, the composition and caloric content of the prior diet may be important determinants of the response to FA; fast rats, or rats fed high carbohydrate diets, respond to increased FA with a rise in VLDL assembly and secretion. Several in vivo rodent models also support a close link between increased plasma FA flux to the liver and increased apoB secretion. In the sucrose-fed hamster,40 which is a model of insulin resistance and hypertriglyceridemia, increased plasma FA levels are correlated with increased assembly and secretion of apoB-lipoproteins (LPs). Studies with mice having genetic alterations in the fatty acid transporter, CD36 (also called FA translocase or FAT), indicate that plasma FA flux to the liver is a critical stimulus for VLDL secretion. Thus, when CD36 was deleted in all tissues (CD expression is normally low in the liver), plasma FA and TG levels were increased despite an apparent increase in overall insulin sensitivity.41 By contrast, in mice overexpressing CD36 in muscle, lower levels of plasma FA and TG were observed despite concomitant overall insulin resistance. Importantly, mice lacking HSL had significantly reduced levels of plasma FA and TG, and also reduced rates of hepatic TG secretion.42 We recently demonstrated, in vivo, that intravenous delivery of albumin-bound FA could stimulate secretion of apoB from livers of normal mice.43

    Increased apoB secretion was first demonstrated in insulin-resistant and diabetic subjects,44 in whom increased FA flux was linked to both VLDL TG and apoB secretion.45 Obese patients were also shown to incorporate plasma FA efficiently into VLDL TG.46 Increased VLDL apoB and TG secretion has been documented in a patient with lipodystrophy, severe insulin resistance, and high levels of plasma FA.47 Impaired insulin-mediated inhibition of FA release by adipose tissue has been related to increased VLDL secretion in subjects with familial combined hyperlipidemia.48 The most direct evidence was provided by studies in which intravenous infusions of the TG-phospholipid emulsion, intralipid, resulted in both acute elevations in plasma FA levels and increased secretion of both apoB and TG in healthy young men.38

    FA delivery to the liver could also lead to increased secretion of VLDL by inhibiting insulin-stimulated apoB degradation.49 Increased availability of FA, either from endogenous TG stores or as dietary TG, not only causes insulin resistance in skeletal muscle, but also in the liver50,51; this would affect insulin’s ability to inhibit VLDL secretion. Indeed, although hyperinsulinemia has been shown to reduce apoB secretion in normal humans,52,53 insulin is ineffective in obese52 or diabetic54 subjects.

    Diminished LPL activity could reduce partitioning of TGFA to adipose tissue from chylomicron TG (in the fed state) and VLDL TG (in the fasted state). This would leave TG-enriched remnant lipoproteins in the circulation, and the liver would be the major repository for these TGFA. Hepatic uptake of apoB-lipoprotein remnants occurs by a multioption process that involves LDL receptors (the primary mediator of uptake), LDL-receptoreCrelated protein (LRP), hepatic lipase (HL), possibly LpL associated with remnants, proteoglycans, and scavenger receptor-B1 (SR-B1).55,56 It is clear from both in vitro57 and in vivo58 studies that uptake by the liver of TG-enriched LPs can stimulate subsequent secretion of hepatic apoB-LPs. It is not clear, however, if LpL/HL-liberated lipoprotein FA or lysosomal-liberated lipoprotein FA have similar effects on apoB-LP assembly and secretion. Indeed, our recent studies43 demonstrate qualitative differences between FA delivered by albumin and FA derived from the internalization of remnants on the assembly and secretion of VLDL.

    Despite the ability of the liver to assemble and secrete VLDL TG, and thereby "unload" excess FA, fatty liver is a common consequence of inappropriate flux of plasma FA and lipoprotein remnant TGFA to that organ.59,60 Why the liver is unable to efficiently unload the excess FA is unclear, but several possibilities can be considered. First, delivery of albumin-bound FA may have different quantitative and qualitative effects compared with delivery of remnant lipoprotein TGFA, with the latter less potent at stimulating secretion of VLDL.43 Second, the state of lipogenesis in the liver may be important; excessive de novo fatty acid synthesis, driven by hyperinsulinemia61 or hyperglycemia,62 in the face of increased exogenous FA influx, may overwhelm the secretory capacity of the liver. Third, insulin-mediated degradation of apoB will clearly be a major determinant of the liver’s ability to assemble and secrete VLDL. Thus both the degree of hyperinsulinemia, and the degree of hepatic insulin resistance, will be critical determinants of VLDL secretion and, therefore, TG accumulation in the liver. For example, if insulin-mediated degradation of apoB is preserved in the setting of increased FA flux and hepatic lipogenesis, fatty liver would be a likely outcome.

    Role of Increased FA Flux to Muscle and Insulin Resistance

    Beginning with Randle’s publication of the "glucose-fatty acid cycle" in skeletal muscle in 1963,63 it has become increasingly clear that disordered FA metabolism is critical in the pathophysiology of muscle insulin resistance.50 Although this is not a review of insulin action or glucose metabolism, it is important to review the effects of increased FA flux to muscle on FA metabolism, and therefore, insulin/glucose metabolism in that tissue. Recent studies have provided strong evidence that the ability of FA to interfere with insulin signaling and glucose transport into muscle correlates with the generation of FA metabolites, such as fatty acyl CoA molecules or diacylglycerol. Increased generation of these metabolites as a response to inappropriate increases in FA uptake can stimulate serine-phosphorylation of IRS-1, and possibly IRS-2, via the actions of certain protein kinase C isoforms.64,65 Targeted disruption of PKC- in mice protected them from FA-induced insulin resistance.66 Recent evidence also suggests that PKC-mediated serine phosphorylation of inhibitor of  kinase (IKK), leading to its degradation and the unregulated translocation of nuclear factor- (NF-) into the nucleus, may also be important to FA-induced insulin resistance.67 We realize that there are many other potential reasons for muscle to be insulin resistant, but it is clear that simply providing excess FA can induce insulin resistance.50

    Why are potentially harmful intermediates of FA metabolism accumulating in muscle of insulin-resistant individuals Is it just increased FA flux into the muscle, especially when those FA are not needed for energy, or is there a defect in FA utilization Of interest in this regard are the findings that insulin-resistant and diabetic individuals have decreased fractional and, in some cases, absolute uptake of FA by muscle.68,69 However, the combination of normal or only slightly reduced plasma FA levels during the immediate postprandial period in insulin-resistant individuals, a time when plasma FA levels should be dramatically suppressed,20 together with modestly reduced extraction of FA by muscle, would still create a situation where FA supply is out of proportion to the need for FA to supply energy to that tissue. Additionally, there is considerable evidence that mitochondrial oxidation of FA is defective in insulin-resistant individuals,70,71 and that this can be reversed by weight loss and exercise.72 Excess FA trapping in muscle, together with reduced FA oxidation, would result in increased TG synthesis. Indeed, increased intramyocellular TG is highly correlated with insulin resistance.73 Although some investigators have suggested that intramyocellular TG is the lipotoxic substance in muscle, it is more likely only a marker of reduced oxidation of FA relative to FA availability.74

    Role of Increased FA Flux to  Cells and Insulin Secretory Defects

    Plasma FA can be taken up by pancreatic  cells where it can stimulate insulin secretion.75 The latter appears to occur as FAs are used to generate long-chain fatty acyl CoA, which can contribute to the synthesis of signaling molecules such as DAG and phosphatidic acid (PA) or directly stimulate PKC isoforms involved in secretion.76 The ability of FA to stimulate insulin secretion requires the presence of stimulatory glucose; the potential reasons for this interaction include the generation from glucose of (1)  glycerol phosphate, which is needed as the backbone molecule for DAG and PA, and (2) malonyl CoA, which inhibits FA oxidation and stimulates lipid formation. Of particular relevance to insulin resistance is the finding that circulating FA are critical for the response of the pancreas to glucose in the fasting state.77 Thus, the high basal insulin response characteristic of insulin resistance might derive, at least in part, from increased levels of FA before the administration of glucose. High-plasma FA uptake by  cells will also lead to increased storage of FA as TG. The release of TGFA from cytosolic TG droplets by HSL and associated lipases is another source of FA that could stimulate insulin release. Indeed, HSL-deficient mice have reduced insulin response to glucose.78

    There is also clear evidence, however, that continued high-cytosolic FA concentrations lead ultimately to decreased insulin secretion.76 Several mechanisms have been reported, including inhibition of insulin synthesis,79 inhibition of the  cell transcription factor, IDX-1,80 and the combination of reduced expression of acetylCoA carboxylase and increased expression of CPT-1.81 In a scheme that parallels the normal interaction of FA and glucose on insulin secretion, the deleterious effects of sustained high FA levels in the islet may require simultaneously elevated levels of glucose.82 Finally, chronic exposure of  cells to high FA and glucose was shown to cause apoptosis in ZDF rats.75,83

    Role of Adipokines in Lipid Metabolism

    Leptin

    Leptin is a 16-kDa protein synthesized mainly in adipose tissue. The leptin gene was identified as the causative mutation in ob/ob mice.84 Numerous reviews have been published describing its regulation and its role in eating and energy homeostasis.85,86 However, studies from a number of laboratories, using in vitro and in vivo rodent models, indicate a direct role for leptin in lipid metabolism. The effect on lipid metabolism may be mediated both through central and peripheral actions of leptin. For example, central administration of leptin increased resting metabolic rates, resulting in reduced TG content in both adipose and nonadipose tissues, as well as reduced plasma-free FA and TG levels in pair-fed Sprague-Dawley rats.87 Leptin may also have autocrine or paracrine effects on adipocyte fat metabolism: incubation of mouse adipocytes with leptin stimulates lipolysis of intracellular TG, and this effect was not seen in db/db mice lacking leptin receptors.88 Overexpression of leptin in adipocytes also reduced gene expression of acetyl CoA carboxylase (ACC).89 In other tissues, leptin also appears to inhibit lipogenesis and stimulate FA oxidation.83 The mechanism underlying these actions of leptin may be via binding to its receptor with succeeding activation of the Jak/Stat pathway90 or by direct stimulation 5'-AMP activated protein kinase (AMPK),91 which will phosphorylate and thereby inhibit ACC activity and lipogenesis. Leptin also can inhibit the expression of sterol response element binding protein-1c (SREBP-1c) in liver, pancreatic islets, and adipose tissue, thereby inhibiting lipogenesis in those tissues.

    The importance of leptin in regulating lipid metabolism in several tissues is best exemplified by the lipotoxicity that is associated with the complete absence of leptin or leptin action.83 For example, as described earlier in this review, individuals without adipose tissue, who also lack leptin, have ectopic fat deposition and lipotoxicity.30,31 Furthermore, rodent models of either naturally occurring leptin deficiency or unresponsiveness, or transgenically created lipodystrophy,32 have similar lipotoxic phenotypes. Importantly, administration of leptin to people with lipodystrophy92 or to rodent models of leptin deficiency,93 reverses many of the lipotoxic sequelae. This is due partly to direct effects of leptin on metabolism, although some of the effects of administered leptin may be through suppression of hyperphagia. The importance of leptin for FA metabolism in nonadipose tissue of obese people who have, in general, increased levels of leptin, remains to be determined.

    Adiponectin

    Adiponectin (ACRP30, adipoQ, apM1, or GBP28) was identified as the product of a gene induced during adipocyte differentiation.94 It is a 30-kDa protein that is synthesized and secreted from adipocytes. Adiponectin has an N-terminal collagenase domain followed by a C-terminal globular domain that can undergo homotrimerization.95 In plasma, adiponectin circulates as either a trimer, a hexamer (called the low molecular weight or LMW form), or as multimeric forms of 12 to 18 subunits (called the high molecular weight or HMW form).95 Of note, recent studies suggest that the HMW form may undergo cleavage to smaller units, including the trimer, and that the latter may be the active form or be further cleaved to an even smaller fragment that transduces adiponectin’s signal to cells, particularly to hepatocytes.96 On the other hand, the globular domain itself (which has not yet been observed in an in vivo setting) can increase fatty acid oxidation in mouse muscle,97 probably via activation of AMPK.98,99 Indeed, transgenic overexpression of the globular domain of adiponectin in the liver of ob/ob mice improved total body insulin sensitivity and increased fatty acid oxidation in skeletal muscle.100 To make matters more complex, studies have indicated that different forms of circulating adiponectin have affects in specific tissues.99,101 Three putative receptors for adiponectin have been cloned.102,103 One is mainly in liver, one in skeletal muscle, and one endothelial cells and smooth muscle. PPAR agonists increase,104 whereas both tumor necrosis factor- (TNF-) and interleukin (IL)-6105 inhibit, adiponectin gene expression.

    Adiponectin-deficient mice develop insulin resistance on a high-fat diet.106,107 Transgenic expression of adiponectin in adipose tissue resulted in increased plasma levels and inhibition of hepatic glucose production with improved glucose tolerance.108 Administration of recombinant adiponectin to lipodystrophic or obese mice results in reductions in plasma FA and TG levels,109 whereas transgenic mice with increased secretion of full-length adiponectin had improved clearance of an exogenous fat load.108

    Adiponectin levels in plasma, which are very high compared with other adipokines, are reduced in obese, insulin-resistant, diabetic, or dyslipidemic subjects.110 In humans, adiponectin levels are inversely correlated with plasma TG levels and positively correlated with HDL cholesterol concentration, although it is not clear if there are direct links between adiponectin and plasma lipid levels or these findings are related to adiponectin’s effects on insulin sensitivity. However, a recent study suggested a strong relationship between adiponectin levels and LPL activity in humans,111 and increased adipose tissue LPL activity was present in a transgenic mouse that secretes increased quantities of full-length adiponectin from adipose tissue.108 In humans, adiponectin levels are predictive of hepatic steatosis and are inversely related to hepatic fat content and hepatic insulin resistance before and after treatment with a PPAR agonist.112

    Resistin

    Resistin is a 12-kDa protein that is synthesized and secreted from adipose tissue. It was discovered in mice by searching for genes that were suppressed by a PPAR agonist.113 Resistin levels are elevated in both diet-induced obesity and genetic (leptin or leptin receptor deficient) mouse models of obesity/diabetes.113 Recombinant resistin decreases insulin sensitivity in mice, and antibodies against resistin block this effect. Resistin infusions in rodents during euglycemic hyperinsulinemic clamp conditions, caused increased hepatic glucose production.114 Mice treated with Western-type high-fat diets had concomitant increases in resistin levels, and antisense oligodeoxynucleotides that normalized resistin levels also reversed hepatic insulin resistance.115 These results are consistent with the data obtained from resistin knockout mice, which despite having no changes in body weight or fat mass, showed significantly lower fasting plasma glucose levels when fed with normal chow diet. When fed with Western-type high-fat diet, the resistin-deficient mice also showed significantly better glucose tolerance than the wild-type mice.116 Deletion of the resistin gene was associated with increased AMPK activity in hepatocytes, decreases in gluconeogenic enzymes, and decreased hepatic glucose production.116 On the other hand, transgenic overexpression of resistin was associated with increased hepatic glucose production and glucose intolerance.117 Of interest, like adiponectin,95 resistin appears to circulate in mice in both trimeric and hexameric forms, with the smaller form (or possibly the monomer) being the most biologically active molecule.118

    Resistin has more recently been shown to have effects on lipid metabolism in mice. When adenovirus-mediated overexpression of resistin was achieved in normal chow-fed mice, plasma TG levels increased, and this was associated with increased secretion of TG from the liver.119 LDL cholesterol also increased whereas HDL cholesterol fell in the mice, and these changes were associated with reduced expression of the LDL receptor and apoAI genes, respectively.119 Wistar rats that had been made "hyper-resistinemic" by adenoviral administration also had increased plasma TG levels.120 No mechanistic studies of lipid metabolism were performed in the latter study.

    The role of resistin in humans is less certain, however. Clinical studies in humans do not show a consistent link between resistin levels and either obesity or insulin resistance.4,121,122 There is also controversy regarding the importance of adipocytes as a source of resistin in human.123,124 Human resistin is only 64% homologous (53% identical with) with the murine counterpart, and both are members of the family of resistin like molecules, RELM (also called FIZZ), which are C-terminal cysteine-rich proteins.124 Given the diverse roles and tissue specificities of this protein family, and the low homology between the rodent and human forms, it is unclear at present whether human resistin plays a similar role as murine resistin, and if it does, how important human resistin is in the pathogenesis of the human insulin resistance.

    Role of Adipocytokines in Lipid Metabolism

    During acute injury or infection, a host of defense responses are elicited, including changes in the partitioning of lipid-derived energy that are associated with insulin resistance and increases in plasma TG and FFA levels. It is believed that these acute metabolic responses are initiated, at least in part, by proinflammatory cytokines. Further, dyslipidemia, insulin resistance, atherosclerosis, and obesity125 have been linked with chronic inflammation. Cytokines are produced not only by immune cells, but also by adipocytes (hence the term adipocytokine).125,126 In fact, adipose tissue is able to secrete a range of "acute-phase reactants," which have been linked, both epidemiologically and biochemically, to the chronic disorder, insulin resistance. The demonstration of cytokine production in adipose tissue suggests that obesity is either seen by the whole organism as an inflammatory state, or that adipocytokines are playing a physiological role in attempting to maintain normal fuel homeostasis.

    Tissue Necrosis Factor-

    TNF- is a 26-kDa transmembrane protein, which is released into the circulation as a 17-kDa soluble protein after extracellular cleavage by a metalloproteinase.127 TNF- has 2 main receptors, type 1 and 2, which are expressed on many cells, including adipocytes.128 Although circulating TNF- levels are relatively low and have no clear correlation with obesity or insulin resistance, tissue expression levels of TNF- correlate positively with both conditions. In mice, chronic exposure of cells or whole animals to TNF- induces insulin resistance, and treatment with soluble forms of TNF- receptors neutralize this effect. Furthermore, mice with targeted gene deletion of TNF- or its receptors showed increased insulin sensitivity and improved plasma FFA levels.129

    TNF- has multiple effects on lipid metabolism, via both paracrine effects on adipocytes and effects in the liver. In adipose tissue, TNF- promotes lipolysis,130 leading to elevation of plasma FA levels. Although there are differences between murine and human adipocytes regarding mechanistic details, TNF-eCinduced lipolysis is, in general, mediated by p44/42 and JNK. These kinases can phosphorylate perilipin, thereby displacing it from lipid droplets and making TG accessible to HSL.131 Additionally, TNF- causes reductions in the expression of genes involved in adipogenesis and lipogenesis in adipocytes, likely through NF-eCmediated transcription.132 By contrast, in liver, TNF- increases the expression of genes involved in de novo fatty acid synthesis while decreasing expression of those involved in fatty acid oxidation. It is not surprising, therefore, that TNF- acutely stimulates VLDL production from liver.133 Increased VLDL secretion, together with TNF-eCmediated inhibition of LPL in adipose tissue,134 results in significant hypertriglyceridemia. TNF- impairs insulin signaling as well, probably by activating Ser/Thr kinases (PKC, PKC, Raf1, and IKK-) that act on insulin receptor and IRS molecules, making them poor substrates for insulin-mediated tyrosine phosphorylation and signal propagation.135 Hepatic insulin resistance increases apolipoprotein B secretion by reducing intracellular degradation of this protein.49 Finally, TNF- can affect lipid metabolism by altering the secretion of other adipokines. For example, TNF- potently reduces expression of adiponectin (Acrp30) and increases expression of IL-1 and IL-6.132

    Interleukin-6

    IL-6 is a protein of 22 to 27 kDa, with various degrees of glycosylation.136 In contrast to TNF-, IL-6 circulates at relatively high levels, and adipocytes contribute 1/3 of the plasma IL-6.137 Plasma levels of IL-6 correlate positively with fat mass, insulin-resistant syndrome, and plasma FFA levels.138,139 Administration of IL-6 to humans results in elevations of plasma FA and increased FA appearance in plasma, consistent with increased adipocyte lipolysis of TG.140,141 Fatty acid oxidation also appears to increase during IL-6 administration to humans.140 These results are consistent with the demonstration that an IL-6 receptor, homologous to the leptin receptor, is present in adipose tissue.4 IL-6 infusions also produce dose-dependent increases in plasma glucose secondary to the development of insulin resistance or increased levels of glucagon.142 Studies in hepatocytes showed that IL-6 is able to impair insulin signaling by downregulation of IRS and upregulation of SOCS-3 (suppressor of cytokine signaling 3), a negative regulator of insulin signaling. Recently, upregulation of SOCS-3 by IL-6 was affirmed in human skeletal muscle and adipose tissue.143 Moreover, IL-6 inhibits adiponectin production in adipocytes,144 an action that may contribute to IL-6eCinduced hepatic insulin resistance. Finally, IL-6 suppresses LPL activity in adipose tissue.145

    IL-6 signaling is complex, however, because of its apparently opposite role in the central nervous system.4 Central administration of IL-6 in rodents causes increased energy expenditure, resulting in decreased body fat. This is consistent with the negative correlation between IL-6 levels in the central nervous system and body fat mass observed in overweight human subjects.146 The existence of a powerful central action of IL-6 may be used to explain the phenotype of the IL-6 knockout mice. Thus, it was reported that these mice were actually obese, leptin resistant, glucose intolerant, and in the females, hypertriglyceridemic and hypercholesterolemic147: phenotypes opposite to what would be expected from the lack of peripheral action of IL-6. Central replacement of IL-6 in these animals partially reversed obesity, whereas peripheral administration of IL-6 had no effect.147 It should be noted, however, that the obese and insulin-resistant phenotypes in IL-6eCdeficient mice were not confirmed in a recent report,148 and the reason for the different findings was not clear.

    Secretion of Apolipoproteins by Adipocytes

    It is not totally surprising that adipose tissue, which not only contain the large majority of the body’s TG stores, but also the largest store of cholesterol in the body, synthesizes and secretes several apolipoproteins. ApoE is a 34-kDa protein that is well known as a key apolipoprotein involved in both secretion and uptake of lipoproteins by tissues and in cholesterol efflux from cells.149 Expression of apoE by adipocytes, and regulation of that expression by cellular cholesterol, were demonstrated over a decade ago.150 Consistent with the latter finding, and with prior studies in macrophages indicating a role for apoE in cholesterol efflux, recent studies have shown that the apoE gene is a target of adipocyte liver X receptors (LXR).151 Potentially relevant to this review are recent findings indicating that PPAR agonists increased apoE mRNA in human adipose tissue and 3T3-L1 cells.152 Additionally, although incubation of cultured 3T3-L1152 or SW872 liposarcoma cells153 with insulin did not affect apoE mRNA, apoE secretion fell in the liposarcoma cells exposed to increasing insulin concentrations.153 If adipose apoE secretion is responsive to insulin in otherwise insulin-resistant adipocytes, then adipocyte apoE secretion might be reduced in states of whole-body hyperinsulinemia. This could lead to defective cholesterol efflux from fat.

    Two other apolipoproteins, apoC-I and apoD, have also been identified as adipocyte products.153,154 ApoC-I is a 6.6-kDa protein that plays a role in regulating receptor-medicated removal of lipoproteins from plasma.10 In SW872 liposarcoma cells, expression of apoC-I increased with increasing cell lipids, and this was associated with increased apoC-I secretion. Insulin treatment, however, reduced both apoC-I mRNA levels and apoC-I secretion into the media.153 The relevance of these findings remains unclear, but the fact that transgenic mice overexpressing apoC-I have defects in fatty acid uptake by adipose tissue155 and resistance to obesity156 suggests that in hyperinsulinemic states, reduced apoC-I synthesis and secretion could contribute to obesity. ApoD is a 6.6-kDa protein that is a member of the lipocalin family of lipid transporters.154 ApoD has recently been identified as a target for LXR in adipocytes, suggesting a role in cholesterol efflux.157 What role apoD plays in adipocyte function in vivo and whether apoD might be relevant to adipocyte dysfunction in insulin-resistant states remain to be determined.

    Summary

    In this review, we have focused particularly on how unregulated secretion of FA from insulin-resistant adipocytes can affect lipid metabolism in liver, muscle, and pancreatic  cells. It is clear from published literature that dysregulated partitioning of FA between storage sites (adipocytes) and sites of utilization (muscle, liver, and  cells) upsets lipid homeostasis in the latter tissues, with deleterious outcomes, including TG accumulation and insulin resistance. We did not mean to suggest, however, that adipose tissue insulin resistance is the only cause of the insulin resistance present either in those other sites, or in the whole body, in rodents or people with the metabolic syndrome. Although we believe that adipose tissue insulin resistance plays a key role in the "typical" insulin resistance found in obese humans, it is clear from several mouse models that muscle158 and liver51 can be the sites of primary insulin resistance, with the possibility for subsequent secondary insulin resistance in the other tissues, including adipose tissue. Further, the basis of adipocyte insulin resistance is very poorly defined. Indeed, even the roles of insulin receptors and PPAR, 2 key proteins involved in energy storage by adipose tissue, are complex and incompletely understood. For example, adipocyte-specific targeted deletion of the insulin receptor (FIRKO mice) results in smaller mice with less body fat, increased whole-body insulin sensitivity, and resistance to age- and obesity-related insulin resistance, despite a clear lack of insulin stimulated glucose and lipid metabolism in adipocytes.159 However, the FIRKO mice do not have increased plasma FA or TG levels, have normal basal and isoproterenol-stimulated lipolysis, and have increased plasma levels of both leptin and adiponectin; all factors that could protect against systemic insulin resistance in this complicated model. Overall, mouse models of tissue-specific loss of insulin receptors or glucose transporters are indicative of this species’ ability to shift metabolic substrates from an insulin-resistant tissue to unaffected tissues.160 We propose that when humans are unable to store energy normally, whether because of the quantitative or qualitative absence of adequate adipose tissue, they also seem to transfer that energy, in the form of FA, to alternative sites such as liver, muscle, and  cells. Whether the accumulation of FA at those sites is adequate to produce the detrimental metabolic sequelae we described in detail in this review or whether concomitant dysregulation of the secretion of adipokines and adipocytokines is required is not clear. What is clear, however, is that the inability to store energy in adipocytes leads to abnormal lipid metabolism and potentially insulin resistance, inflammation, and cell death in the alternative storage sites.

    Acknowledgments

    This work was supported by the following grants: Ginsberg: R01 HL55638, R01 HL73030; Yu: K08 DK60530.

    References

    Guerre-Millo M. Adipose tissue and adipokines: for better or worse. Diabetes Metab. 2004; 30: 13eC19.

    Lyon CJ, Law RE, Hsueh WA. Minireview: Adiposity, inflammation, and atherogenesis. Endocrinology. 2003; 144: 2195eC2200.

    Meier U, Gressner AM. Endocrine regulation of energy metabolism: review of pathobiochemical and clinical chemical aspects of leptin, ghrelin, adiponectin, and resistin. Clinical Chemistry. 2004; 50: 1511eC1525.

    Kershaw EE, Flier JS. Adipose tissue as an endocrine organ. J Clin Endocrinol Metab. 2004; 89: 2548eC2556.

    Havel PJ. Regulation of energy balance and carbohydrate/lipid metabolism. Diabetes. 2004; 53: S143eCS151.

    Phan CT, Tso P. Intestinal lipid absorption and transport. Front Biosci. 2001; 6: D299eCD319.

    Goldberg IJ. Lipoprotein lipase and lipolysis: central roles in lipoprotein metabolism and atherosclerosis. J Lipid Res. 1996; 37: 693eC707.

    Fielding BA, Frayn KN. Lipoprotein lipase and the disposition of dietary fatty acids. Br J Nutr. 1998; 80: 495eC502.

    LaRosa JC, Levy RI, Herbert P, Lux SE, Fredrickson DS. A specific apoprotein activator for lipoprotein lipase. Biochem Biophy Res Commun. 1970; 41: 57eC62.

    Shachter NS. Apolipoproteins C-I and C-III as important modulators of lipoprotein metabolism. Curr Opin Lipidol. 2001; 12: 297eC304.

    Blanchette-Mackie EJ, Scow RO. Lipid filling and lipolysis in adipose tissue and cells. Int J Obes. 1985; 9: 7eC14.

    Hamilton JA, Kamp F. How are free fatty acids transported in membranes Is it by proteins or by free diffusion through the lipids Diabetes. 1999; 48: 2255eC2269.

    Abumrad NA, Coburn C, Ibrahimi A. Membrane proteins implicated in long-chain fatty acid uptake by mammalian cells: CD36, FATP and FABPm. Biochim Biophys Acta. 1999; 1441: 4eC13.

    Faraj M, Lu HL, Cianflone K. Diabetes, lipids, and adipocyte secretagogues. Biochem Cell Biol. 2004; 82: 170eC190.

    Koisiten HA, Vidal H, Karonen SL, Dusserre E, Vallier P, Koivisto VA, Ebeling P. Plasma acylation stimulating protein concentration and subcutaneous adipose tissue C3 mRNA expression in nondiabetic and type 2 diabetic men. Arterioscler Thromb Vasc Biol. 2001; 21: 1034eC1039.

    Murray I, Sniderman AD, Cianflone K. Mice lacking acylation stimulating protein (ASP) have delayed postprandial triglyceride clearance. J Lipid Res. 1999; 40: 1671eC1676.

    Murray I, Havel PJ, Sniderman AD, Cianflone K. Reduced body weight, adipose tissue, and leptin levels despite increased energy intake in female mice lacking acylation-stimulating protein. Endocrinology. 2000; 141: 1041eC1049.

    Xia Z, Sniderman AD, Cianflone K. Acylation-stimulating protein (ASP) deficiency induces obesity resistance and increased energy expenditure in ob/ob mice. J Biol Chem. 2002; 277: 45874eC45879.

    Wetsel RA, Kildsgaard J, Zsigmond E, Liazo W, Chan L. Genetic deficiency of acylation stimulating protein (ASP(C3ades-Arg)) does not cause hyperapobetalipoproteinemia in mice. J Biol Chem. 1999; 274: 19429eC19433.

    Evans K, Burdge GC, Wootton SA, Clark ML, Frayn KN. Regulation of dietary fatty acid entrapment in subcutaneous adipose tissue and skeletal muscle. Diabetes. 2002; 51: 2684eC2690.

    Kersten S. Mechanisms of nutritional and hormonal regulation of lipogenesis. EMBO Rep. 2001; 2: 282eC286.

    Anthonsen MW, Ronnstrand L, Wernstedt C, Degerman E, Holm C. Identification of novel phosphorylation sites in hormone-sensitive lipase that are phosphorlated in response to isoproterenol and govern activation properties in vitro. J Biol Chem. 1998; 273: 215eC221.

    Coppack SW, Evans RD, Fisher RM, Frayn KN, Gibbons GF, Humphreys SM, Kirk ML, Potts JL, Hockaday TD. Adipose tissue metabolism in obesity: lipase action in vivo before and after a mixed meal. Metabolism. 1992; 41: 264eC272.

    Axelsen M, Smith U, Eriksson JW, Taskinen MR, Jansson PA. Postprandial hypertriglyceridemia and insulin resistance in normoglycemic first-degree relatives of patients with type 2 diabetes. Ann Intern Med. 1999; 131: 27eC31.

    Frayn KN. Adipose tissue as a buffer for daily lipid flux. Diabetologia. 2002; 45: 1201eC1210.

    Frayn KN, Humphreys SM, Coppack SW. Net carbon flux across subcutaneous adipose tissue after a standard meal in normal-weight and insulin-resistant obese mice. Int J Obes Relat Metab Disord. 1996; 20: 795eC800.

    Seip M, Trygstad O. Generalized lipodystrophy, congenital and acquired (lipoatrophy). Acta Paediatri Suppl. 1996; 413: 2eC28.

    Agarwal AK, Arioglu E, de Almeida S, Akkoc N, Taylor SI, Bowcock AM, Barnes RI, Garg A. AGPAT2 is mutated in congenital generalized lipodystrophy linked to chromosome 9q34. Nature Genetics. 2002; 31: 21eC23.

    Fu M, Kazlauskaite R, Baracho MF, Santos MG, Brandao-Neto J, Villares S, Celi FS, Wajchenberg BL, Shuldiner AR. Mutations in Gng31g and AGPAT2 in Berardinell-Seip congenital lipodystrophy and Brunzell syndrome: phenotype variability suggests important modifier effects. J Clin Endocrinol Metab. 2004; 89: 2916eC2922.

    Garg A. Acquired and Inherited Lipodystrophies. N Engl J Med. 2004; 350: 1220eC1234.

    Garg A, Misra A. Lipodystrophies: rare disorders causing metabolic syndrome. Endocrinol Metab Clinic NA. 2004; 33: 305eC331.

    Moitra J, Mason MM, Olive M, Krylov D, Gavrilova O, Marcus-Samuels B, Feigenbaum L, Lee E, Aoyama T, Eckhaus M, Reitman ML, Vinson C. Life without white fat: a transgenic mouse. Genes & Development. 1998; 12: 3168eC3181.

    Shimomura I, Hammer RE, Richardson JA, Ikemoto S, Bashmakov Y, Goldstein JL, Brown MS. Insulin resistance and diabetes mellitus in transgenic mice expressing nuclear SREBP-1c in adipose tissue: model for congenital generalized lipodystrophy. Gene Develop. 1998; 12: 3182eC3194.

    Siri P, Candela N, Ko C, Zhang Y, Eusufzai S, Ginsberg HN, Huang LS. Post-transcriptional stimulation of the assembly and secretion of triglyceride-rich apolipoprotein B-lipoproteins in a mouse with selective deficiency of brown adipose tissue, obesity and insulin resistance. J Biol Chem. 2001; 276: 46064eC46072.

    Gargiulo CE, Stuhlsatz-Krouper SM, Schaffer JE. Localization of adipocyte long-chain fatty acyl-CoA synthetase at the plasma membrane. J Lipid Res. 1999; 40: 881eC892.

    Chiu HC, Kovacs A, Ford DA, Hsu FF, Garcia R, Herrero P, Saffitz JE, Schaffer JE. A novel mouse model of lipotoxic cardiomypathy. J Clin Invest. 2001; 107: 813eC822.

    Dixon JL, Ginsberg HN. Regulation of hepatic secretion of apolipoprotein B-containing lipoproteins: information obtained from cultured liver cells. J Lipid Res. 1993; 34: 167eC179.

    Lewis GF. Fatty acid regulation of very low density lipoprotein (VLDL) production. Curr Opin Lipidol. 1999; 10: 475eC477.

    Fisher EA, Ginsberg HN. Complexity in the secretory pathway: the assembly and secretion of apolipoprotein B-containing lipoproteins. J Biol Chem. 2002; 277: 17377eC17380.

    Taghibiglou C, Carpentier A, Van Iderstine SC, Chen B, Rudy D, Aiton A, Lewis GF, Adeli K. Mechanisms of hepatic very low density lipoprotein overproduction in insulin resistance. Evidence for enhanced lipoprotein assembly, reduced intracellular apoB degradation, and increased microsomal triglyceride transfer protein in a fructose-fed hamster model. J Biol Chem. 2000; 275: 8416eC8425.

    Febbraio M, Abumrad NA, Hajjar DP, Sharma K, Cheng W, Pearce SF, Silverstein RL. A null mutation in murine CD36 reveals an important role in fatty acid and lipoprotein metabolism. J Biol Chem. 1999; 274: 19055eC19062.

    Haemmerle G, Zimmermann R, Hayn M, Theussl C, Waeg G, Wagner E, Sattler W, Magin TM, Wagner EF, Zechner R. Hormone-sensitive lipase deficiency in mice causes diglyceride accumulation in adipose tissue, muscle, and testis. J Biol Chem. 2002; 277: 4806eC4815.

    Zhang Y-L, Hernandez-Ono A, Ko C, Yasunaga K, Huang L-S, Ginsberg HN. Regulation of hepatic apolipoprotein B-lipoprotein assembly and secretion by the availability of fatty acids 1: differential effects of delivering fatty acids via albumin or remnant-like emulsion particles. J Biol Chem. 2004; 279: 19362eC19374.

    Kissebah AH, Alfarsi S, Adams PW, Wynn V. The metabolic fate of plasma lipoproteins in normal subjects and in patients with insulin resistance and endogenous hypertriglyceridaemia. Diabetologia. 1976; 12: 501eC509.

    Kissebah AH, Alfarsi S, Adams PW, Seed M, Folkard J, Wynn V. Transport kinetics of plasma free fatty acid, very low density lipoprotein triglycerides and apoprotein in patients with endogenous hypertriglycerideaemia. Atherosclerosis. 1976; 24: 199eC218.

    Barter PJ, Nestel PJ. Precursors of plasma triglyceride fatty acids in obesity. Metabolism. 1973; 22: 779eC783.

    Franklin B, Ginsberg H, Hague WU, Ginsberg-Fellner F. Very low density lipoprotein metabolism in lipoatrophic diabetes. Metabolism. 1984; 33: 814eC819.

    Castro Cabezas M, de Bruin TW, de Valk HW, Shoulders CC, Jansen H, Willem Erkelens D. Impaired fatty acid metabolism in familial combined hyperlipidemia: a mechanism associating hepatic apolipoprotein B overproduction and insulin resistance. J Clin Invest. 1993; 92: 160eC168.

    Sparks JD, Sparks CE. Insulin regulation of triacylglycerol-rich lipoprotein synthesis and secretion. Biochim Biophys Acta. 1994; 1215: 9eC32.

    Boden G, Shulman GI. Free fatty acids in obesity and type 2 diabetes: defining their role in the development of insulin resistance and -cell dysfunction. European J Clin Invest. 2002; 32: 14eC23.

    Samuel VT, Liu ZX, Qu X, Elder BD, Bilz S, Befroy D, Romanelli AJ, Shulman GI. Mechanism of hepatic insulin resistance in non-alcoholic fatty liver disease. J Biol Chem. 2004; 279: 32345eC32353.

    Lewis GF, Uffelman KD, Szeto LW, Steiner G. Effects of acute hyperinsulinemia on VLDL triglyceride and VLDL apoB production in normal weight and obese individuals. Diabetes. 1993; 42: 833eC842.

    Malmstrom R, Packard CJ, Watson TD, Rannikko S, Caslake M, Bedford D, Stewart P, Yki-Jarvinen H, Shepherd J, Taskinen MR. Metabolic basis of hypotriglyceridemic effects of insulin in normal men. Arterioscler Thromb Vasc Biol. 1997; 17: 1454eC1464.

    Malmstrom R, Packard CJ, Caslake M, Bedford D, Stewart P, Jarvinen H, Shepherd J, Taskinen MR. Defective regulation of triglyceride metabolism by insulin in the liver in NIDDM. Diabetologia. 1997; 40: 454eC462.

    Yu KC, Cooper AD. Postprandial lipoproteins and atherosclerosis. Front Biosci. 2001; 6: D332eCD354.

    Out R, Kruijt JK, Rensen PC, Hldebrand RB, De Vos P, VanEck M, van Berkel TJ. Scavenger receptor BI plays a role in facilitating chylomicron metabolism. J Biol Chem. 2004; 279: 18401eC18406.

    Wu X, Sakata N, Dixon JL, Ginsberg HN. Exogenous VLDL stimulates apolipoprotein B from HepG2 cells by both pre- and post-translational mechanisms. J Lipid Res. 1994; 35: 1200eC1210.

    Cohn JS, Wagner DA, Cohn SD, Millar JS, Schaefer EJ. Measurement of very low density and low density lipoprotein apolipoprotein (Apo) B-100 and high density lipoprotein Apo A-I production in human subjects using deuterated leucine: effect of fasting and feeding. J Clin Invest. 1990; 85: 804eC811.

    Angulo P. Nonalcholic fatty liver disease. N Engl J Med. 2002; 346: 1221eC1231.

    Browning JD, Szczepaniak LS, Dobbins R, Nuremberg P, Horton JD, Cohen JC, Grundy SM, Hobbs HH. Prevalence of hepatic steatosis in an urban population in the United States: impact of ethnicity. Hepatology. 2004; 40: 1387eC1395.

    Shimomura I, Matsuda M, Hammer RE, Bashmakov Y, Brown MS, Goldstein JL. Decreased IRS-2 and increased SREBP-1c lead to mixed insulin resistance and sensitivity in livers of lipodystrophic and ob/ob mice. Mol Cell. 2000; 6: 77eC86.

    Uyeda K, Yamashita H, Kawaguchi T. Carbohydrate responsive element-binding protein (ChREBP): a key regulator of glucose metabolism and fat storage. Biochem Pharmcol. 2002; 63: 2075eC2080.

    Randle PJ, Garland PB, Hales CN, Newsholme EA. The glucose fatty-acid cycle. Its role in insulin sensitivity and the metabolic disturbances of diabetes mellitus. Lancet. 1963; 1: 785eC789.

    Yu C, Chen Y, Cline GW, Zhang D, Zong H, Wang Y, Bergeron R, Kim JK, Cushman SW, Cooney GJ, Atcheson B, White MF, Kraegen EW, Shulman GI. Mechanism by which fatty acids inhibit insulin activation of insulin receptor substrate-1 (IRS-1)-associated phosphatidylinositol 3-kinase activity in muscle. J Biol Chem. 2002; 277: 50230eC50236.

    Itani SI, Ruderman NB, Schmieder F, Boden G. Lipid-induced insulin resistance in human muscle is associated with changes in diacylglycerol, protein kinase C, and IkappaB-alpha. Diabetes. 2002; 51: 2005eC2011.

    Kim JK, Fillmore JJ, Sunshine MJ, Albrecht B, Higashimori T, Kim D-W, Liu Z-X, Soos TJ, Cline GW, O’Brien WR, Littman DR, Shulman GI. PKC-theta knockout mice are protected from fat-induced insulin resistance. J Clin Invest. 2004; 114: 823eC827.

    Hundal RS, Petersen KF, Mayerson AB, Randhawa PS, Inzucchi S, Shoelson SE, Shulman GI. Mechanism by which high-dose aspirin improves glucose metabolism in type 2 diabetes. J Clin Invest. 2002; 109: 1321eC1326.

    Kelley DE, Simoneau JA. Impaired free fatty acid utilization by skeletal muscle in non-insulin-dependent diabetes mellitus. J Clin Invest. 1994; 94: 2349eC2356.

    Blaak EE. Basic disturbances in skeletal muscle fatty acid metabolism in obesity and type 2 diabetes mellitus. Proc Nutr Soc. 2004; 63: 323eC330.

    Ritov VB, Menshikova EV, He J, Ferrell RE, Goodpaster BH, Kelley DE. Deficiency of subsarcolemmal mitochondria in obesity and type 2 diabetes. Diabetes. 2005; 54: 8eC14.

    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: 664eC671.

    Menshikova EV, Ritov VB, Toledo FG, Ferrell RE, Goodpaster BH, Kelley DE. Effects of weight loss and physical activity on skeletal muscle mitochondrial function in obesity. Am J Physiol Endocrinol Metab. 2005; 288: E818eCE825

    Kelley DE, Williams KV, Price JC, McKolanis TM, Goodpaster BH, Thaete FL. Plasma fatty acids, adiposity, and variance of skeletal muscle insulin resistance in type 2 diabetes mellitus. J Clin Endocrinol Metab. 2001; 86: 5412eC5419.

    Goodpaster BH, Kelley DE. Skeletal muscle triglyceride: marker or mediator of obesity-induced insulin resistance in type 2 diabetes mellitus Curr Diab Rep. 2002; 2: 216eC222.

    McGarry JD, Dobbins RL. Fatty acids, lipotoxicity and insulin secretion. Diabetologia. 1999; 42: 128eC138.

    Yaney GC, Corkey BE. Fatty acid metabolism and insulin secretion in pancreatic beta cells. Diabetologia. 2003; 46: 1297eC1312.

    Dobbins RL, Chester MW, Daniels MB, McGarry JD, Stein DT. Circulating fatty acids are essential for efficient glucose-stimulated insulin secretion after prolonged fasting in humans. Diabetes. 1998; 47: 1613eC1618.

    Roduit R, Masiello P, Wang SP, Li H, Mitchell GA, Prentki M. A role for hormone-sensitive lipase in glucose-stimulated insulin secretion: a study in hormone-sensitive lipase-deficient mice. Diabetes. 2001; 50: 1970eC1975.

    Bollheimer LC, Skelly RH, Chester MW, McGarry JD, Rhodes CJ. Chronic exposure to free fatty acid reduces pancreatic beta cells insulin content by increasing basal insulin secretion that is not compensated for by a corresponding increase in proinsulin biosynthesis translation. J Clin Invest. 1998; 101: 1094eC1101.

    Gremlich S, Bonny C, Waeber G, Thorens B. Fatty acids decrease ID-X-1 expression in rat pancreatic islets and reduces GLUT2, glucokinase, insulin, and somatostatin levels. J Biol Chem. 1997; 272: 30261eC30269.

    Brun T, Assimacopoulos-Jeannet F, Corkey BE, Prentki M. Long-chain fatty acids inhibit acetyl-CoA carboxylase gene expression in the pancreatic beta-cell line INS-1. Diabetes. 1997; 46: 393eC400.

    Prentki M, Joly E, El-Assaad W, Roduit R. Malonyl-CoA signaling, lipid partitioning, and glucolipotoxicity. Role in -cell adaptation and failure in the etiology of diabetes. Diabetes. 2002; 51: S405eCS413.

    Unger RH, Orci L. Lipoapoptosis: its mechanism and its diseases. Biochimica et Biophysica Acta. 2002; 1585: 202eC212.

    Zhang Y, Proenca R, Maffei M, Barone M, Leopold L, Friedman JM. Positional cloning of the mouse obese gene and its human homologue. Nature. 1994; 372: 425eC432.

    Leibel RL. The role of leptin in the control of body weight. Nutr Rev. 2002; 60: S15eCS19.

    Friedman JM. The function of leptin in nutrition, weight, and physiology. Nutr Rev. 2002; 60: S1eCS14.

    Dobbins RL, Szczepaniak LS, Zhang W, McGarry JD. Chemical sympathectomy alters regulation of body weight during prolonged ICV leptin infusion. Am J Physiol Endocrinol Metab. 2003; 284: E778eCE787.

    Fruhbeck G, Aguado M, Martinez JA. In vitro lipolytic effect of leptin on mouse adipocytes: evidence for a possible autocrine/paracrine role of leptin. Biochem Biophy Res Commun. 1997; 240: 590eC594.

    Bai Y, Zhang S, Kim KS, Lee JK, Kim KH. Obese gene expression alters the ability of 30A5 preadipocytes to respond to lipogenic hormones. J Biol Chem. 1996; 271: 13939eC13942.

    Unger RH. Minireview: Weapons of lean body mass destruction: the role of ectopic lipids in the metabolic syndrome. Endocrinology. 2003; 144: 5159eC5165.

    Minokoshi Y, Yim YB, Peroni OD, Fryer LG, Muller C, Carling D, Kahn BB. Leptin stimulates fatty-acid oxidation by activating AMP-activated protein kinase. Nature. 2002; 415: 268eC269.

    Oral EA, Simha V, Ruiz E, Andewelt A, Premkumar A, Snell P, Wagner AJ, DePaoli AM, Reitman ML, Taylor SI, Gorden P, Garg A. Leptin-replacement therapy for lipodystrophy. N Engl J Med. 2002; 346: 570eC578.

    Shimomura I, Hammer RE, Ikemoto S, Brown MS, Goldstein JL. Leptin reverses insulin resistance and diabetes mellitus in mice with congenital lipodystrophy. Nature. 1999; 401: 73eC76.

    Pajvani UB, Scherer PE. Adiponectin: systemic contributor to insulin sensitivity. Curr Diab Rep. 2003; 3: 207eC213.

    Pajvani UB, Du X, Combs TP, Berg AH, Rajala MW, Schulthess T, Engel J, Brownlee M, Scherer PE. Structure-function studies of the adipocyte-secreted hormone Acrp30/adiponectin. Implications for metabolic regulation and bioactivity. J Biol Chem. 2003; 278: 9073eC9085.

    Pajvani UB, Hawkins M, Combs TP, Rajala MW, Doebber T, Berger JP, Wagner JA, Wu M, Knopps A, Xiang AH, Utzschneider KM, Kahn SE, Olefsky JM, Buchanan TA, Scherer PE. Complex distribution, not absolute amount of adiponectin, correlates with thiazolidinedione-mediated improvement in insulin sensitivity. J Biol Chem. 2004; 279: 12152eC12162.

    Fruebis J, Tsao TS, Javorschi S, Ebbets-Reed D, Erickson MR, Yen FT, Bihain BE, Lodish HF. Proteolytic cleavage product of 30-kDa adipocyte complement-related protein increases fatty acid oxidation in muscle and causes weight loss in mice. Proc Natl Acad Sci U S A. 2001; 98: 2005eC2010.

    Tomas E, Tsao TS, Saha AK, Murrey HE, Zhang CC, Itani SI, Lodish HF, Ruderman NB. Enhanced muscle fat oxidation and glucose transport by ACRP30 gobular domain: acetyl-CoA carboxylase inhibition and AMP-activated protein kinase activation. Proc Natl Acad Sci U S A. 2002; 99: 16309eC16313.

    Yamauchi T, Kamon J, Minokoshi Y, Ito Y, Waki H, Uchida S, Yamashita S, Noda M, Kita S, Ueki K, Eto K, Akanuma Y, Froguel P, Foufelle F, Ferre P, Carling D, Kimura S, Nagai R, Kahn BB, Kadowaki T. Adiponectin stimulates glucose utilization and fatty-acid oxidation by activating AMP-activated protein kinase. Nat Med. 2002; 8: 1288eC1295.

    Yamauchi T, Kamon J, Waki H, Imai Y, Shimozawa N, Hioki K, Uchida S, Ito Y, Takakuwa K, Matsui J, Takata M, Eto K, Terauchi Y, Komeda K, Tsunoda M, Murakami K, Ohnishi Y, Naitoh T, Yamamura K, Ueyama Y, Froguel P, Kimura S, Nagai R, Kadowaki T. Globular adiponectin protected ob/ob mice from diabetes and ApoE-deficient mice from atherosclerosis. J Biol Chem. 2003; 278: 2461eC2468.

    Tsao TS, Tomas E, Murrey HE, Hug C, Lee DH, Ruderman NB, Heuser JE, Lodish HF. Role of disulfide bonds in Acrp30/adiponectin structure and signaling specificity. Different oligomers activate different signal transduction pathways. J Biol Chem. 2003; 278: 50810eC50817.

    Yamauchi T, Kamon J, Ito Y, Tsuchida A, Yokomizo T, Kita S, Suglyama T, Miyagishi M, Hara K, Tsunoda M, Murakami K, Ohteki T, Uchida S, Takekawa S, Waki H, Tsuno NH, Shibata Y, Terauchi Y, Froguel P, Tobe K, Koyasu S, Taira K, Kitamura T, Shimizu RN, Kadowaki T. Cloning of adiponectin receptors that mediate antidiabetic metabolic effects. Nature. 2003; 423: 762eC769.

    Hug C, Wang J, Ahmad NS, Bogan JS, Tsao TS, Lodish HF. T-cadherin is a receptor for hexameric and high-molecular-weight forms of Acrp30/adiponectin. Proc Natl Acad Sci U S A. 2004; 101: 10308eC10313.

    Maeda N, Takahashi M, Funahashi T, Kihara S, Nishizawa H, Kishida K, Nagaretani H, Matsuda M, Komuro R, Ouchi N, Kuriyama H, Hotta K, Nakamura T, Shimomura I, Matsuzawa Y. PPARgamma ligands increase expression and plasma concentrations of adiponectin, an adipose-derived protein. Diabetes. 2001; 50: 2094eC2099.

    Bruun JM, Lihn AS, Verdich C, Pedersen SB, Toubro S, Astrup A, Richelsen B. Regulation of adiponectin by adipose tissue-derived cytokines: in vivo and in vitro investigations in humans. Am J Physiol Endocrinol Metab. 2003; 285: E527eCE533.

    Maeda N, Shimomura I, Kishida K, Nishizawa H, Matsuda M, Nagaretani H, Furuyama N, Kondo H, Takahashi M, Arita Y, Komuro R, Ouchi N, Kihara S, Tochino Y, Okutomi K, Horie M, Takeda S, Aoyama T, Funahashi T, Matsuzawa Y. Diet-induced insulin resistance in mice lacking adiponectin/ACRP30. Nat Med. 2002; 8: 731eC737.

    Kubota N, Terauchi Y, Yamauchi T, Kubota T, Moroi M, Matsui J, Eto K, Yamashita T, Kamon J, Satoh H, Yano W, Froguel P, Nagai R, Kimura S, Kadowaki T, Noda T. Disruption of adiponectin causes insulin resistance and neointimal formation. J Biol Chem. 2002; 277: 25863eC25866.

    Combs TP, Pajvani UB, Berg AH, Lin Y, Jelicks LA, Laplante M, Nawrocki AR, Rajala MW, Parlow AF, Cheeseboro L, Ding YY, Russell RG, Lindemann D, Hartley A, Bake GR, Obici S, Deshaies Y, Ludgate M, Rossetti L, Scherer PE. A transgenic mouse with a deletion in the collagenous domain of adiponectin displays elevated circulating adiponectin and improved insulin sensitivity. Endocrinology. 2004; 145: 367eC383.

    Yamauchi T, Kamon J, Waki H, Terauchi Y, Kubota N, Hara K, Mori Y, Ide T, Murakami K, Tsuboyama-Kasaoka N, Ezaki O, Akanuma Y, Gavrilova O, Vinson C, Reitman ML, Kagechika H, Shudo K, Yoda M, Nakano Y, Tobe K, Nagai R, Kimura S, Tomita M, Froguel P, Kadowaki T. The fat-derived hormone adiponectin reverses insulin resistance associated with both lipoatrophy and obesity. Nat Med. 2001; 7: 941eC946.

    Kadowaki T, Hara K, Yamauchi T, Terauchi Y, Tobe K, Nagai R. Molecular mechanism of insulin resistance and obesity. Exp Biol Med. 2003; 228: 1111eC1117.

    von Eynatten M, Schneider JG, Humpert PM, Rudofsky G, Schmidt N, Barosch P, Hamann A, Morcos M, Kreuzer J, Bierhaus A, Nawroth PP, Dugi KA. Decreased plasma lipoprotein lipase in hypoadiponectinemia: an association independent of systemic inflammation and insulin resistance. Diabetes Care. 2004; 27: 2925eC2929.

    Bajaj M, Suraamornkul S, Piper P, Hardies LJ, Glass L, Cersosimo E, Pratipanawatr T, Miyazaki Y, Defronzo RA. Decreased plasma adiponectin concentrations are closely related to hepatic fat content and hepatic insulin resistance in pioglitazone-treated type 2 diabetes patients. J Clin Endocrinol Metab. 2004; 89: 200eC206.

    Steppan CM, Bailey ST, Bhat S, Brown EJ, Banerjee RR, Wright CM, Patel HR, Ahima RS, Lazar MA. The hormone resistin links obesity to diabetes. Nature. 2001; 409: 307eC312.

    Rajala MW, Obici S, Scherer PE, Rossetti L. Adipose-derived resistin and gut-derived resistin-like molecule-beta selectively impair insulin action on glucose production. J Clin Invest. 2003; 111: 225eC230.

    Muse ED, Obici S, Bhanot S, Monia BP, McKay RA, Rajala MW, Scherer PE, Rossetti L. Role of resistin in diet-induced hepatic insulin resistance. J Clin Invest. 2004; 114: 232eC239.

    Banerjee RR, Rangwala SM, Shapiro JS, Rich AS, Rhoades B, Qi Y, Wang J, Rajala MW, Pocai A, Scherer PE, Steppan CM, Ahima RS, Obici S, Rossetti L, Lazar MA. Regulation of fasted blood glucose by resistin. Science. 2004; 303: 1195eC1198.

    Rangwala SM, Rich AS, Rhoades B, Shapiro JS, Obici S, Rossetti L, Lazar MA. Abnormal glucose homeostasis due to chronic hyperresistinemia. Diabetes. 2004; 53: 1937eC1941.

    Patel SD, Rajala MW, Rossetti L, Scherer PE, Shapiro L. Disulfide-dependent multimeric assembly of resistin family hormones. Science. 2004; 304: 1154eC1158.

    Sato N, Kobayashi K, Inoguchi T, Sonoda N, Imamura M, Sekiguchi N, Nakashima N, Nawata H. Adenovirus-mediated high expression of resistin causes dyslipidemia in mice. Endocrinology. 2005; 146: 273eC279.

    Satoh H, Nguyen MT, Miles PD, Imamura T, Usui I, Olefsky JM. Adenovirus-mediated chronic "hyper-resistinemia" leads to in vivo insulin resistance in normal rats. J Clin Invest. 2004; 114: 224eC231.

    Banerjee RR, Lazar MA. Resistin: molecular history and prognosis. J Mol Med. 2003; 81: 218eC226.

    Kielstein JT, Becker B, Graf S, Brabant G, Haller H, Fliser D. Increased resistin blood levels are not associated with insulin resistance in patients with renal disease. Am J Kidney Dis. 2003; 42: 62eC66.

    Fain JN, Cheema PS, Bahouth SW, Lloyd Hiler M. Resistin release by human adipose tissue explants in primary culture. Biochem Biophys Res Commun. 2003; 300: 674eC678.

    Steppan CM, Brown EJ, Wright CM, Bhat S, Banerjee RR, Dai CY, Enders GH, Silberg DG, Wen X, Wu GD, Lazar MA. A family of tissue-specific resistin-like molecules. Proc Natl Acad Sci U S A. 2001; 98: 502eC506.

    Weisberg SP, McCann D, Desai M, Rosenbaum M, Leibel RL, Ferrabtem AW, Jr. Obesity is associated with macrophage accumulation in adipose tissue. J Clin Invest. 2003; 112: 1796eC1808.

    Fried SK, Bunkin DA, Greenberg AS. Omental and subcutaneous adipose tissues of obese subjects release interleukin-6: depot difference and regulation by gluccorticoid. J Clin Endocrinol Metab. 1998; 83: 847eC850.

    Gearing AJ, Beckett P, Christodoulou M, Churchill M, Clements J, Davidson AH, Drummond AH, Galloway WA, Gilbert R, Gordon JL. Processing of tumour necrosis factor-alpha percusor by metalloproteinases. Nature. 1994; 370: 555eC557.

    Hotamisligil GS, Arner P, Atkinson RL, Spiegelman BM. Differential regulation of the p80 tumor necrosis factor receptor in human obesity and insulin resistance. Diabetes. 1997; 46: 451eC455.

    Uysal KT, Wiesbrock SM, Marino MW, Hotamisligil GS. Protection from obesity-induced insulin resistance in mice lacking TNF- function. Nature. 1997; 389: 610eC614.

    Green A, Dobias SB, Walters DJ, Brasier AR. Tumor necrosis factor increases the rate of lipolysis in primary cultures of adipocytes without altering levels of hormone-sensitive lipase. Endocrinology. 1994; 134: 2581eC2588.

    Zhang HH, Halbleib M, Ahmad F, Manganiello VC, Greenberg AS. Tumor necrosis factor-alpha stimulates lipolysis in differentiated human adipocytes through activation of extracellular signal-related kinase and elevation of intracellular cAMP. Diabetes. 2002; 51: 2929eC2935.

    Ruan H, Hacohen N, Golub TR, Van Parijs L, Lodish HF. Tumor necrosis factor-alpha suppresses adipocyte-specific genes and activates expression of preadipocyte genes in 3T3eCL1 adipocytes: nuclear factor-kappaB activation by TNF-alpha is obligatory. Diabetes. 2002; 51: 1319eC1336.

    Grunfeld C, Feingold KR. Tumor necrosis factor, interleukin, and interferon induced changes in lipid metabolism as part of host defense. Proc Soc Exp Biol Med. 1992; 200: 224eC227.

    Beutler B, Greenwald D, Hulmes JD, Chang M, Pan YC, Mathison J, Ulevitch R, Cerami A. Identity of tumour necrosis factor and the macrophase-secreted factor cachectin. Nature. 1985; 316: 552eC554.

    Peraldi P, Hotamisligil GS, Buurman WA, White MF, Spiegelman BM. Tumor necrosis factor (TNF)-alpha inhibits insulin signaling through stimulation of the p55 TNF receptor and activation of sphingomyelinase. J Biol Chem. 1996; 271: 13018eC13022.

    Gross V, Andus T, Castell J, Vom Berg D, Heinrich PC, Gerok W. O- and N- glycosylation lead to different molecular mass forms of human monocyte interleukin-6. FEBS Lett. 1989; 247: 323eC326.

    Mohamed-Ali V, Goodrick S, Rawesh A, Katz DR, Miles JM, Yudkin JS, Klein S, Coppack SW. Subcutaneous adipose tissue releases interleukin-6, but not tumor necrosis factor-alpha, in vivo. J Clin Endocrinol Metab. 1997; 82: 4196eC4200.

    Kern PA, Ranganathan S, Li C, Ranganathan G. Adipose tissue tumor necrosis factor and interleukin-6 expression in human obesity and insulin resistance. Am J Physiol Endocrinol Metab. 2001; 280: E745eCE751.

    Vozarova B, Weyer C, Hanson K, Tataranni PA, Borgardus C, Pratley RE. Circulating interleukin-6 in relation to adiposity, insulin action, and insulin secretion. Obes Res. 2001; 9: 414eC417.

    van Hall G, Steensberg A, Sacchetti M, Fischer C, Keller C, Schjerling P, Hiscock N, Moller K, Saltin B, Febbraio MA, Pedersen BK. Interleukin-6 stimulates lipolysis and fat oxidation in humans. J Clin Endocrinol Metab. 2003; 88: 3005eC3010.

    Petersen EW, Carey AL, Sacchetti M, Steinberg GR, Macaulay SL, Febbraio MA, Pedersen BK. Acute IL-6 treatment increases fatty acid turnover in elderly humans in vivo and in tissue culture in vitro. Am J Physiol Endocrinol Metab. 2005; 288: E155eCE162.

    Tsigos C, Papanicolaou DA, Kyrou I, Defensor R, Mitsiadis CS, Chrousos GP. Dose-dependent effects of recombinant human interleukin-6 on glucose regulation. J Clin Endocrinol Metab. 2001; 82: 4167eC4170.

    Rieusset J, Bouzakri K, Chevillotte E, Ricard N, Jacquet D, Bastard JP, Laville M, Vidal H. Suppressor of cytokine signaling 3 expression and insulin resistance in skeletal muscle of obese and type 2 diabetic patients. Diabetes. 2004; 53: 2232eC2241.

    Fasshauer M, Kralisch S, Klier M, Lossner U, Bluher M, Klein J, Paschke R. Adiponectin gene expression and secretion is inhibited by interleukin-6 in 3T3eCL1 adipocytes. Biochem Biophy Res Commun. 2003; 301: 1045eC1050.

    Greenberg AS, Nordan RP, McIntosh J, Calvo JC, Scow RO, Jablons D. Interleukin 6 reduces lipoprotein lipase activity in adipose tissue of mice in vivo and in adipocytes: a possible role for interleukin 6 in cancer cachexia. Cancer Res. 1992; 52: 4113eC4116.

    Stenlof K, Wernstedt I, Fjallman T, Wallenius K, Jansson JO. Interleukin-6 levels in the central nervous system are negatively correlated with fat mass in overweight/obese subjects. J Clin Endocrinol Metab. 2003; 88: 4379eC4383.

    Wallenius V, Wallenius K, Ahren B, Rudling M, Carlsten H, Dickson SL, Ohlsson C, Jansson J-O. Interleukin-6-deficient mice develop mature-onset obesity. Nat Med. 2002; 8: 75eC79.

    Di Gregorio GB, Hensley L, Lu T, Ranganathan G, Kern PA. Lipid and carbohyrdate metabolism in mice with a targeted mutation in the IL-6 gene: absence of development of age-related obesity. Am J Physiol Endocrinol Metab. 2004; 287: E182eCE187.

    Mahley RW. Apolipoprotein E. Cholesterol transport protein with expanding role in cell biology. Science. 1988; 240: 622eC630.

    Zechner R, Moser R, Newman TC, Fried SK, Breslow JL. Apolipoprotein E gene expression in mouse 3T3eCL1 adipocytes and human adipose tissue and its regulation by differentiation and lipid content. J Biol Chem. 1991; 266: 10583eC10588.

    Laffitte BA, Repa JJ, Joseph SB, Wilpitz DC, Kast HR, Mangelsdorf DJ, Tontonoz P. LXRs control lipid-inducible expression of the apolipoprotein E gene in macrophages and adipocytes. Proc Natl Acad Sci U S A. 2001; 98: 507eC512.

    Yue L, Rasouli N, Ranganathan G, Kern PA, Mazzone T. Divergent effects of peroxisome proliferator-activated receptor  agonists and tumor necrosis factor  on adipocyte apoE expression. J Biol Chem. 2004; 279: 47626eC47632.

    Wassef H, Bernier L, Davignon J, Cohn JS. Synthesis and secretion of ApoC-1 and ApoE during maturation of human SW872 liposarcoma cells. J Nutr. 2004; 134: 2935eC2941.

    Rassart E, Bedirian A, Do Carmo S, Guinard O, Sirois J, Terrisse L, Milne R. Apolipoprotein D. Biochim Biophys Acta. 2000; 1482: 185eC198.

    Jong MC, Gijbels MJ, Dahlmans VE, Gorp PJ, Koopman SJ, Ponee M, Hofker MH, Havekes LM. Hyperlipidemia and cutaneous abnormalities in transgenic mice overexpressing human apolipoprotein C1. J Clin Invest. 1998; 101: 145eC152.

    Jong MC, Voshol PJ, Muurling M, Dahlmans VE, Romijn JA, Pijl H, Havekes LM. Protection from obesity and insulin resistance in mice overexpressing human apolipoprotein C1. Diabetes. 2001; 50: 2779eC2785.

    Hummasti S, Laffitte BA, Watson MA, Galardi C, Chao LC, Ramamurthy L, Moore JT, Tontonoz P. Liver X receptors are regulators of adipocyte gene expression but not differentiation: identification of apoD as a direct target. J Lipid Res. 2004; 45: 616eC625.

    Kim JK, Zisman A, Fillmore JJ, Peroni OD, Kotani K, Perret P, Zong H, Dong J, Kahn CR, Kahn BB, Shulman GI. Glucose toxicity and the development of diabetes in mice with muscle-specific inactivation of GLUT4. J Clin Invest. 2001; 108: 153eC160.

    Bluher M, Michael DM, Peroni OD, Ueki K, Carter N, Kahn BB, Kahn RC. Adipose tissue selective insulin receptor knockout protects against obesity and obesity-related glucose intolerance. Dev Cell. 2002; 3: 25eC38.

    Minokoshi Y, Kahn CR, Kahn BB. Tissue-specific ablation of the GLUT4 glucose transporter or the insulin receptor challenges assumptions about insulin action and glucose homeostasis. J Biol Chem. 2003; 278: 33609eC33612.

作者: Yi-Hao Yu, Henry N. Ginsberg 2007-5-18
医学百科App—中西医基础知识学习工具
  • 相关内容
  • 近期更新
  • 热文榜
  • 医学百科App—健康测试工具