首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 46 毫秒
1.
J Fowelin  S Attvall  H Von Schenck  U Smith  I Lager 《Diabetes》1989,38(11):1357-1364
The occurrence and mechanisms for late (6.5- to 7.5-h) posthypoglycemic insulin resistance were studied with the euglycemic clamp in 19 healthy subjects. Comparisons were made with a control study with the same insulin infusion rate but where hypoglycemia was prevented by glucose infusion. Glucose production and utilization were studied with D-[3-3H] glucose infusions. Hypoglycemia induced marked insulin resistance shown by lower glucose infusion rates compared with the control study 3.1 +/- 0.3 vs. 6.0 +/- 0.7 mg.kg-1.min-1, P less than .001). This late posthypoglycemic insulin resistance was mainly due to a decreased insulin effect on glucose utilization. Infusion of propranolol did not prevent insulin resistance, whereas somatostatin partially prevented its appearance. Somatostatin plus metyrapone completely normalized posthypoglycemic insulin resistance. A positive correlation (r = .72, P less than .001) was found between initial insulin sensitivity and percent reduction of the insulin effect after hypoglycemia. Thus, hypoglycemia is followed by prolonged (6- to 8-h) insulin resistance. In contrast to early-phase (2- to 3-h) resistance, long-term resistance is not due to beta-adrenergic stimulation but to the combined effect of growth hormone and cortisol. This resistance is also more pronounced in subjects with initially high insulin sensitivity.  相似文献   

2.
Obesity is associated with insulin resistance, particularly when body fat has a central distribution. However, insulin resistance also frequently occurs in apparently lean individuals. It has been proposed that these lean insulin-resistant individuals have greater amounts of body fat than lean insulin-sensitive subjects. Alternatively, their body fat distribution may be different. Obesity is associated with elevated plasma leptin levels, but some studies have suggested that insulin sensitivity is an additional determinant of circulating leptin concentrations. To examine how body fat distribution contributes to insulin sensitivity and how these variables are related to leptin levels, we studied 174 individuals (73 men, 101 women), a priori classified as lean insulin-sensitive (LIS, n = 56), lean insulin-resistant (LIR, n = 61), and obese insulin-resistant (OIR, n = 57) based on their BMI and insulin sensitivity index (S(I)). Whereas the BMI of the two lean groups did not differ, the S(I) of the LIR subjects was less than half that of the LIS group. The subcutaneous and intra-abdominal fat areas, determined by computed tomography, were 45 and 70% greater in the LIR subjects (P < 0.001) and 2.5- and 3-fold greater in the OIR group, as compared with the LIS group. Fasting plasma leptin levels were moderately increased in LIR subjects (10.8 +/- 7.1 vs. 8.1 +/- 6.4 ng/ml in LIS subjects; P < 0.001) and doubled in OIR subjects (21.9 +/- 15.5 ng/ml; P < 0.001). Because of the confounding effect of body fat, we examined the relationships between adiposity, insulin sensitivity, and leptin concentrations by multiple regression analysis. Intra-abdominal fat was the best variable predicting insulin sensitivity in both genders and explained 54% of the variance in S(I). This inverse relationship was nonlinear (r = -0.688). On the other hand, in both genders, fasting leptin levels were strongly associated with subcutaneous fat area (r = 0.760) but not with intra-abdominal fat. In line with these analyses, when LIS and LIR subjects were matched for subcutaneous fat area, age, and gender, they had similar leptin levels, whereas their intra-abdominal fat and insulin sensitivity remained different. Thus, accumulation of intra-abdominal fat correlates with insulin resistance, whereas subcutaneous fat deposition correlates with circulating leptin levels. We conclude that the concurrent increase in these two metabolically distinct fat compartments is a major explanation for the association between insulin resistance and elevated circulating leptin concentrations in lean and obese subjects.  相似文献   

3.
The objective of this work was to evaluate the relationship between sex steroid hormones, sex hormone-binding-globulin, leptin, insulin and insulin resistance in obese men. Anthropometrical indexes, total testosterone (Tt), free testosterone (fT), estradiol (E), sex hormone-binding-globulin (SHBG), glucemia, insulin and leptin were measured in 77 men, with ages between 20 and 60 years. According to their body mass index (BMI), subjects were grouped into three categories: normal body weight (< 24.9 kg/m2), overweight (25-29.9 kg/m2) and obese group (> 30 kg/m2). Insulin resistance index was obtained by the homeostasis assessment model for insulin resistance (HOMA-IR). Total testosterone and SHBG concentrations were lower in the obese group compared with normal and overweight subjects (p < 0.05). The mean insulin concentration was significantly higher in the obese group compared with the other groups (p < 0.05). T was negatively correlated with the BMI (r = -0.447; p < .01), WC (r = -0.464); p < .01, leptin (r = -0.382; p < .01), insulin (r = -0.391; p < 0.01) and also with the HOMA-IR (r = -0.416; p < 0.01). The SHBG negatively and significantly correlated with BMI (r = -0.334; p < 0.01) and WC index (= -0.322; p < 0.01), as well with insulin levels (r = -0.313; p < 0.01) and insulin resistance (= -0.266; p < 0.05). Our results shows that in a sample of men, Tt and SHBG concentrations proportionally diminished with both the increase of BMI and insulin resistance index.  相似文献   

4.
Adiponectin is a 29-kDa adipocyte protein that has been linked to the insulin resistance of obesity and lipodystrophy. To better understand the regulation of adiponectin expression, we measured plasma adiponectin and adipose tissue adiponectin mRNA levels in nondiabetic subjects with varying degrees of obesity and insulin resistance. Plasma adiponectin and adiponectin mRNA levels were highly correlated with each other (r = 0.80, P < 0.001), and obese subjects expressed significantly lower levels of adiponectin. However, a significant sex difference in adiponectin expression was observed, especially in relatively lean subjects. When men and women with a BMI <30 kg/m(2) were compared, women had a twofold higher percent body fat, yet their plasma adiponectin levels were 65% higher (8.6 +/- 1.1 and 14.2 +/- 1.6 micro g/ml in men and women, respectively; P < 0.02). Plasma adiponectin had a strong association with insulin sensitivity index (S(I)) (r = 0.67, P < 0.0001, n = 51) that was not affected by sex, but no relation with insulin secretion. To separate the effects of obesity (BMI) from S(I), subjects who were discordant for S(I) were matched for BMI, age, and sex. Using this approach, insulin-sensitive subjects demonstrated a twofold higher plasma level of adiponectin (5.6 +/- 0.6 and 11.2 +/- 1.1 micro g/ml in insulin-resistant and insulin-sensitive subjects, respectively; P < 0.0005). Adiponectin expression was not related to plasma levels of leptin or interleukin-6. However, there was a significant inverse correlation between plasma adiponectin and tumor necrosis factor (TNF)-alpha mRNA expression (r = -0.47, P < 0.005), and subjects with the highest levels of adiponectin mRNA expression secreted the lowest levels of TNF-alpha from their adipose tissue in vitro. Thus, adiponectin expression from adipose tissue is higher in lean subjects and women, and is associated with higher degrees of insulin sensitivity and lower TNF-alpha expression.  相似文献   

5.
In vivo studies have shown that insulin resistance in late pregnancy results from a decreased sensitivity of liver and peripheral tissues. In the present study, measurements of the rates of glucose utilization by skeletal muscles (soleus, extensor digitorum longus, epitrochlearis, and diaphragm), white adipose tissue, and brain of virgin and 19-day pregnant rats were performed in the basal condition and during a euglycemic, hyperinsulinemic (400 microU/ml) clamp to quantify the partition of glucose utilization and to identify the tissues other than liver responsible for insulin resistance. Fetal and placental glucose utilization rates were also measured in pregnant rats. The fetal glucose utilization rate (22 mg/min/kg) was very high and was not stimulated by physiologic maternal hyperinsulinemia. By contrast, the placental glucose utilization rate (29 mg/min/kg) was increased by 30% during hyperinsulinemia. The glucose utilization rate of the conceptus represented 23% of the maternal glucose utilization rate in the basal state. Glucose utilization rates in the basal condition were not statistically altered by pregnancy in brain, skeletal muscles, and white adipose tissue. During hyperinsulinemia (400 microU/ml), glucose utilization rates in extensor digitorum longus, epitrochlearis, and white adipose tissue were 30-70% lower in pregnant than in virgin rats. Insulin sensitivity of glucose metabolism in all the tissues tested other than brain was 50% lower in pregnant than in virgin rats. We conclude that skeletal muscles and, to a smaller extent, adipose tissue are involved in the insulin resistance of late pregnancy.  相似文献   

6.
Ng TW  Watts GF  Farvid MS  Chan DC  Barrett PH 《Diabetes》2005,54(3):795-802
We investigated the relationship of plasma adipocytokine concentrations with VLDL apolipoprotein B (apoB)-100 kinetics in men. Plasma adiponectin, leptin, resistin, interleukin-6 (IL-6), and tumor necrosis factor-alpha (TNF-alpha) concentrations were measured using enzyme immunoassays and insulin resistance by homeostasis model assessment (HOMA) score in 41 men with BMI of 22-35 kg/m(2). VLDL apoB kinetics were determined using an intravenous infusion of 1-[(13)C]leucine, gas chromatography-mass spectrometry, and compartmental modeling. Visceral and subcutaneous adipose tissue mass (ATM) were determined using magnetic resonance imaging, and total ATM was measured by bioelectrical impedance. In univariate regression, plasma adiponectin and leptin concentrations were inversely and directly associated, respectively, with plasma triglyceride; HOMA score; and visceral, subcutaneous, and total ATMs. Conversely, adiponectin and leptin were directly and inversely correlated, respectively, with VLDL apoB catabolism and HDL cholesterol concentration (P < 0.05). Resistin, IL-6, and TNF-alpha were not significantly associated with any of these variables. In multivariate regression, adiponectin was the most significant predictor of plasma VLDL apoB concentration (P = 0.001) and, together with total or subcutaneous ATM, was an independent predictor of VLDL apoB catabolism (P < 0.001); HOMA score was the most significant predictor of VLDL apoB hepatic secretion (P < 0.05). Leptin was not an independent predictor of VLDL apoB kinetics. In conclusion, plasma VLDL apoB kinetics may be differentially controlled by adiponectin and insulin resistance, with adiponectin regulating catabolism and insulin resistance regulating hepatic secretion in men. Total body fat may also independently determine the rate of VLDL catabolism, but leptin, resistin, IL-6, and TNF-alpha do not have a significant effect in regulating apoB kinetics.  相似文献   

7.
Cancer cachexia may be mediated by endogenous peptides such as tumour necrosis factor alpha (TNF-alpha). Insulin resistance occurs in these patients, and is also seen experimentally with TNF-alpha administration. In this study, insulin sensitivity and energy metabolism were measured in 11 patients with gastrointestinal cancer and ten controls, using the euglycaemic glucose clamp and indirect calorimetry. Patients with cancer were significantly more insulin resistant than controls (P < 0.01) and in such patients insulin resistance correlated with serum TNF-alpha level (rs = 0.74, P < 0.01). Fasting insulin levels also correlated inversely with insulin sensitivity (rs = -0.62, P = 0.003). These results suggest a possible association between endogenous TNF-alpha production and insulin resistance in patients with gastrointestinal cancer.  相似文献   

8.
Some animal models suggest that tumor necrosis factor (TNF)-alpha is a key component in obesity-linked insulin resistance because it inhibits insulin receptor signaling and glucose transport in insulin-sensitive tissues. However, in vivo data in humans have given conflicting results regarding the relationship between circulating TNF-alpha levels and insulin sensitivity. In the present study, the potential local role of TNF-alpha on insulin action in human subcutaneous adipose tissue was studied in 42 obese women (BMI 39+/-10 kg/m2). We found a strong inverse correlation between adipose TNF-alpha secretion and maximum insulin-stimulated glucose transport in adipocytes that was independent of fat cell volume, age, and BMI (P < 0.001, r = 0.58). As much as one-third of the variation in insulin-stimulated glucose transport could be accounted for by variations in TNF-alpha secretion. There was no significant correlation (r = 0.11) between secretion of adipose plasminogen activator inhibitor 1 and glucose transport. Furthermore, subcutaneous adipose tissue of 4 obese women (BMI 40+/-4) incubated with TNF-A for 24 h showed a one-third concentration-dependent inhibition of insulin-stimulated glucose transport (P < 0.01). In conclusion, adipose TNF-alpha may be an important specific and local factor in adipose tissue that influences the ability of insulin to stimulate glucose transport in human fat cells, at least in obese women.  相似文献   

9.
The endocrine system coordinates development of the mammary gland with reproductive development and the demand of the offspring for milk. Three categories of hormones are involved. The levels of the reproductive hormones, estrogen, progesterone, placental lactogen, prolactin, and oxytocin, change during reproductive development or function and act directly on the mammary gland to bring about developmental changes or coordinate milk delivery to the offspring. Metabolic hormones, whose main role is to regulate metabolic responses to nutrient intake or stress, often have direct effects on the mammary gland as well. The important hormones in this regard are growth hormone, corticosteroids, thyroid hormone, and insulin. A third category of hormones has recently been recognized, mammary hormones. It currently includes growth hormone, prolactin, PTHrP, and leptin. Because a full-term pregnancy in early life is associated with a reduction in breast carcinogenesis, an understanding of the mechanisms by which these hormones bring about secretory differentiation may offer clues to the prevention of breast cancer.  相似文献   

10.
Bernard-Kargar C  Ktorza A 《Diabetes》2001,50(Z1):S30-S35
Endocrine pancreas plasticity may be defined as the ability of the organ to adapt the beta-cell mass to the variations in insulin demand. For example, during late pregnancy and obesity, the increase of the beta-cell mass, in association with beta-cell hyperactivity, contributes to insulin oversecretion in response to insulin resistance. There is increasing evidence that the ability of the beta-cell mass to expand in adult mammals is much higher than previously thought. During pregnancy, placental hormones, especially placental lactogens, are mainly responsible for the changes in beta-cell mass. The factors involved in beta-cell growth in obesity are far from clear, although increased free fatty acids seem to be the main candidate. Many data suggest that the impairment of insulin secretion in type 2 diabetes is partly related to reduction of beta-cell mass, at least relative to prevailing insulin demand. This defect may originate from genetic predisposition, but the situation is likely worsened by environmental factors such as hyperglycemia (glucotoxicity) and hyperlipidemia (lipotoxicity). Better understanding of beta-cell growth and regeneration mechanisms may allow new strategies in the treatment of type 2 diabetes based on early limitation of beta-cell damage and/or restoration of a functional beta-cell mass.  相似文献   

11.
Altered glucocorticoid hormone action may contribute to the etiology of the metabolic syndrome, but the molecular mechanisms are poorly defined. Tissue sensitivity to glucocorticoid is regulated by expression of the glucocorticoid receptor (GR)-alpha and 11beta-hydroxysteroid dehydrogenase type I (11beta-HSD1)-mediated intracellular synthesis of active cortisol from inactive cortisone. We have analyzed GRalpha and 11beta-HSD1 expression in skeletal myoblasts from men (n = 14) with contrasting levels of insulin sensitivity (euglycemic clamp measurements of insulin-dependent glucose disposal rate), blood pressure, and adiposity. Positive associations were evident between myoblast expression of GRalpha under basal conditions and levels of insulin resistance (r(2) = 0.34, P < 0.05), BMI (r(2) = 0.49, P < 0.01), percent body fat (r(2) = 0.34, P < 0.02), and blood pressure (r(2) = 0.86, P < 0.001). Similar associations were evident when myoblasts were incubated with physiological levels of cortisol (P < 0.01 for all). Importantly, GRalpha expression was unaffected by variations in in vivo concentrations of insulin, IGF-1, or glucose concentrations. In common with the GR, 11beta-HSD1 expression in myoblasts incubated with physiological concentrations of cortisol in vitro was positively associated with levels of insulin resistance (r(2) = 0.68, P < 0.001), BMI (r(2) = 0.63, P < 0.005), and blood pressure (r(2) = 0.27, P < 0.05). Regulation of GRalpha and 11beta-HSD1 by cortisol was abolished by the GR antagonist RU38486. In summary, our data suggest that raised skeletal muscle cell expression of GRalpha and 11beta -HSD1-mediated regulation of intracellular cortisol may play a fundamental role in mechanisms contributing to the pathogenesis of the metabolic syndrome.  相似文献   

12.
We investigated whether a polymorphism in codons 22 and 23 of the glucocorticoid (GC) receptor gene [GAGAGG(GluArg) --> GAAAAG(GluLys)] is associated with altered GC sensitivity, anthropometric parameters, cardiovascular risk factors, and sex steroid hormones. In a subgroup of 202 healthy elderly subjects of the Rotterdam Study, we identified 18 heterozygotes (8.9%) for the 22/23EK allele (ER22/23EK carriers). In the highest age group, the number of ER22/23EK carriers was higher (67-82 years, 12.9%) than in the youngest age group (53-67 years, 4.9%; P < 0.05). Two dexamethasone (DEX) suppression tests with 1 and 0.25 mg DEX were performed, and serum cortisol and insulin concentrations were compared between ER22/23EK carriers and noncarriers. After administration of 1 mg DEX, the ER22/23EK group had higher serum cortisol concentrations (54.8 +/- 18.3 vs. 26.4 +/- 1.4 nmol/l, P < 0.0001), as well as a smaller decrease in cortisol (467.0 +/- 31.7 vs. 484.5 +/- 10.3 nmol/l, P < 0.0001). ER22/23EK carriers had lower fasting insulin concentrations (P < 0.001), homeostasis model assessment- insulin resistance (IR) (index of IR, P < 0.05), and total (P < 0.02) and LDL cholesterol concentrations (P < 0.01). Our data suggest that carriers of the 22/23EK allele are relatively more resistant to the effects of GCs with respect to the sensitivity of the adrenal feedback mechanism than noncarriers, resulting in a better metabolic health profile.  相似文献   

13.
Resistin is an adipokine with putative prodiabetogenic properties. Like other hormones secreted by adipose tissue, resistin is being investigated as a possible etiologic link between excessive adiposity and insulin resistance. Although there is growing evidence that circulating levels of this adipokine are proportional to the degree of adiposity, an effect on insulin resistance in humans remains unproven. To evaluate the relations among resistin, obesity, and insulin resistance, we measured fasting serum resistin levels in 113 nondiabetic (75-g oral glucose tolerance test) Pima Indians (ages 29 +/- 7 years, body fat 31 +/- 8%, resistin 3.7 +/- 1.1 ng/ml [means +/- SD]), who were characterized for body composition (assessed by hydrodensitometry or dual-energy X-ray absorptiometry), whole-body insulin sensitivity (M; assessed by hyperinsulinemic clamp), basal hepatic glucose output (BHGO) and hepatic glucose output during low-dosage insulin infusion of a hyperinsulinemic clamp (HGO; a measure of hepatic insulin resistance), and acute insulin secretory response (AIR; assessed by 25-g intravenous glucose tolerance test). Follow-up measurements of M, BHGO, HGO, and AIR were available for 34 subjects who had normal glucose tolerance at baseline and remained nondiabetic at follow-up. The average time to follow-up was 4.5 +/- 2.7 years. In cross-sectional analyses, serum resistin levels were positively associated with percent body fat (r = 0.37, P = 0.0001) and 2-h glucose (r = 0.19, P = 0.04), respectively. Serum resistin levels were not associated with fasting glucose and insulin levels, M, BHGO, HGO, or AIR (r = 0.17, 0.12, -0.13, -0.06, -0.03, and -0.04, respectively; all P > 0.05). After adjusting for percent body fat, there was no association between serum resistin levels and 2-h glucose (r = 0.06, P = 0.6). In prospective analyses, high serum resistin levels at baseline were not associated with a decline in M (r = -0.1, P > 0.5). Resistin levels were, however, associated with increases in percent body fat, fasting plasma insulin, and HGO (r = 0.34, 0.36, and 0.37; all P < 0.05) after adjusting for sex, age, and time to follow-up. After additional adjustment for the change in percent body fat, there was no association between baseline serum resistin levels and changes in plasma insulin or HGO (r = 0.26 and 0.23; both P > 0.1). We conclude that in Pima Indians, like other human populations, circulating resistin levels are proportional to the degree of adiposity, but not the degree of insulin resistance. We unexpectedly found that high serum resistin levels do predict future increases in percent body fat. Our data suggest that resistin promotes obesity but not obesity-associated insulin resistance in humans.  相似文献   

14.
Mice lacking the p85alpha regulatory subunit of phosphoinositide (PI) 3-kinase (Pik3r1(-/-)) showed increased glucose uptake in white adipose tissue (WAT) and skeletal muscle due to increased phosphatidylinositol (3,4,5)-triphosphate [PtdIns(3,4,5)P3] production and on a normal diet had a body weight and fat mass similar to wild-type mice. After 3 months on a high-fat diet, Pik3r1(-/-) mice still had increased insulin sensitivity and better glucose tolerance than wild-type mice, but showed markedly greater increases in body weight and WAT mass than wild-type mice. On the normal diet, serum leptin levels of Pik3r1(-/-) mice were significantly higher than in wild-type mice as a result of increased leptin secretion from adipocytes, presumably due to the increased PtdIns(3,4,5)P3 production in adipocytes. Leptin (5 microg/g body wt per day) caused a reduction in food intake and decrease in body weight by the wild-type mice as well as Pik3r1(-/-) mice, suggesting Pik3r1(-/-) mice having leptin sensitivity similar to wild-type mice. The slightly increased serum leptin compensated for the increased glucose uptake by adipocytes in Pik3r1(-/-) mice, thereby preventing adiposity on the normal diet. On the high-fat diet, leptin (5 microg/g body wt per day) failed to decrease food intake or body weight in either genotype, indicating that both genotypes had indeed become severely leptin resistant. Consequently, leptin secretion was unable to sufficiently compensate for the severe leptin resistance caused by the high-fat diet, thereby failing to prevent obesity in Pik3r1(-/-) mice. Our findings suggest that primary increase in serum leptin on the normal diet play a role in the protection from adiposity in Pik3r1(-/-) mice.  相似文献   

15.
AIM: The aim of the present study is: (i) to evaluate the serum concentrations of leptin and resistin in the paediatric patients with chronic renal impairment (CRI), on haemodialysis (HD) and on peritoneal dialysis (PD) treatment; (ii) to examine the relationship between these hormones; and (iii) to investigate the possible influence of these hormones on the insulin resistance and sensitivity indexes as well as on serum insulin-like growth factor-1 (IGF-1) and insulin-like growth factor binding protein-3 (IGFBP-3) levels. METHODS: In total, 52 patients (15 patients with CRI, 24 PD patients and 13 HD patients) and 23 healthy age- and sex-matched control subjects were included in the present study. RESULTS: Homeostasis model assessment of insulin resistance (HOMA-IR) was higher than 2.5 in 47.1% of the patients. IGF-1 levels of patients with CRI, PD and HD patients were significantly lower than those in the controls (P < 0.001, P < 0.001, P < 0.001, respectively). The leptin levels of patients with CRI and on PD and HD treatment were significantly higher than the control group (P = 0.038, P = 0.002, P = 0.006, respectively). Similarly, serum resistin levels of patients with CRI and those of PD and HD patients were higher when compared with healthy controls (P = 0.037, P < 0.001, P = 0.005, respectively). CONCLUSION: Leptin and resistin levels were increased in the children with CRF; however, this elevation was not found to be associated with hyperinsulinism. Further studies to explain the mechanisms and consequences of the accumulation of these hormones in CRF may provide the therapeutical approach aiming to normalize their circulating levels.  相似文献   

16.
Obesity-related glucose intolerance is a function of hepatic (homeostatic model assessment-insulin resistance [HOMA-IR]) and peripheral insulin resistance (S(i)) and beta-cell dysfunction. We determined relationships between changes in these measures, visceral (VAT) and subcutaneous (SAT) adipose tissue, and systemic adipocytokine biomarkers 1 and 6 months after surgical weight loss. HOMA-IR decreased significantly (-50%) from baseline by 1 month and decreased further (-67%) by 6 months, and S(i) was improved by 6 months (2.3-fold) weight loss. Plasma concentrations of leptin decreased and adiponectin increased significantly by 1 month, and decreases in interleukin-6, C-reactive protein (CRP), and tumor necrosis factor-alpha were observed at 6 months of weight loss. Longitudinal decreases in CRP (r = -0.53, P < 0.05) were associated with increases in S(i), and decreases in HOMA-IR were related to increases in adiponectin (r = -0.37, P < 0.05). Decreases in VAT were more strongly related to increases in adiponectin and decreases in CRP than were changes in general adiposity or SAT. Thus, in severely obese women, specific loss of VAT leads to acute improvements in hepatic insulin sensitivity mediated by increases in adiponectin and in peripheral insulin sensitivity mediated by decreases in CRP.  相似文献   

17.
In vitro studies revealed that insulin resistance might be associated with the intracellular formation of ceramide, the second messenger in the sphingomyelin signaling pathway. The aim of the present study was to examine the content and composition of fatty acids in ceramide and sphingomyelin in human muscle and to evaluate their relationships with insulin sensitivity. The study was conducted on 27 male subjects with normal glucose tolerance. Euglycemic-hyperinsulinemic clamps and biopsies of vastus lateralis muscle were performed. In 10 subjects, additional biopsies were taken after a 4-h clamp and after a clamp with concurrent Intralipid/heparin infusion. We identified 13 ceramides and sphingomyelins according to fatty acid residues. Insulin sensitivity was related to total ceramide content (r = -0.49, P = 0.01) and to ceramide consisting of palmitic (r = -0.48, P = 0.011), palmitoleic (r = -0.45, P = 0.019), mirystic (r = -0.42, P = 0.028), and nervonic acid (r = -0.39, P = 0.047). Hyperinsulinemia did not affect estimated muscle parameters. Intralipid/heparin infusion resulted in a 24.73% decrease in insulin sensitivity (P = 0.007) and a 47.81% increase in ceramide content (P = 0.005). These changes were significantly related to each other (r = -0.64, P = 0.046). A relationship with the decrease in insulin sensitivity was also observed for ceramides consisting of palmitic (r = -0.68, P = 0.03) and linoleic (r = -0.66, P = 0.038) acid. Our data indicate that the sphingomyelin signaling pathway in muscle might be an important factor determining the development of insulin resistance in humans.  相似文献   

18.
INTRODUCTION: The ob gene product leptin is secreted by fat cells and the serum leptin levels reflects the body fat content. Markedly elevated serum leptin levels have been reported in patients with chronic renal failure. The aim of the present study was to assess if the dialysate leptin levels in peritoneal dialysate are similar to what can be expected from passive diffusion or if intraperitoneal synthesis of leptin may occur. METHODS: We studied 39 patients (20 males), mean age 54+/-12 years, who had been treated with peritoneal dialysis for 17+/-12 months. Ten of the patients were diabetics of which seven used intraperitoneal insulin. A 24-h collection of dialysate was performed and dialysate and fasting blood samples were analysed for leptin, albumin and beta2-microglobulin, and the peritoneal clearances (PCl) were calculated for these solutes. RESULTS: Serum leptin (mean 47+/-76, range 3-350 ng/ml) was related to body mass index (r=0.35, P<0.05). In multiple regression analysis, serum leptin also correlated to serum TNF-alpha. Although dialysate leptin levels correlated to serum leptin, they were higher than expected from the molecular weight of 16 kD. PCl of leptin was 1.3 ml/min (range 0.2-5.9 ml/min), which was 1.6 times higher than expected from the molecular weight of leptin and PCl for albumin and beta2-microglobulin, not taking the protein binding of leptin into account. A strong correlation was found between PCI for albumin and beta2-microglobulin (r = 0.68, P < 0.0001) but neither PCl albumin, nor PCl beta2-microglobulin correlated to PCI leptin. The PCl of leptin was markedly higher in diabetics using intraperitoneal insulin (n = 7) compared to the other 32 patients (2.6+/-2.0 vs 1.1+/-0.7 ml/min, P<0.05). CONCLUSION: Serum leptin is locally produced in the peritoneal cavity, and intraperitoneal insulin enhances local production of leptin.  相似文献   

19.
Glucocorticoids induce insulin resistance in humans, whereas thiazolidinediones enhance insulin sensitivity. Although the effects of glucocorticoids and thiazolidinediones have been assessed in isolation, interaction between these drugs, which both act as ligands for nuclear receptors, has been less well studied. Therefore, we examined the metabolic effects of dexamethasone and troglitazone, alone and in combination, for the first time in humans. A total of 10 healthy individuals with normal glucose tolerance (age 40 +/- 11 years, BMI 31 +/- 6.1 kg/m(2)) were sequentially studied at baseline, after 4 days of dexamethasone (4 mg/day), after 4-6 weeks on troglitazone alone (400 mg/day), and again after 4 days of dexamethasone added to troglitazone. Key metabolic variables included glucose tolerance assessed by blood glucose and insulin responses to an oral glucose tolerance test (OGTT), insulin sensitivity evaluated via hyperinsulinemic-euglycemic clamp, free fatty acids (FFAs) and FFA suppressibility by insulin during the clamp study, and fasting serum leptin. Dexamethasone drastically impaired glucose tolerance, with fasting and 2-h OGTT insulin values increasing by 2.3-fold (P < 0.001) and 4.4-fold (P < 0.001) over baseline values, respectively. The glucocorticoid also induced a profound state of insulin resistance, with a 34% reduction in maximal glucose disposal rates (GDRs; P < 0.001). Troglitazone alone increased GDRs by 20% over baseline (P = 0.007) and completely prevented the deleterious effects of dexamethasone on glucose tolerance and insulin sensitivity, as illustrated by a return of OGTT glucose and insulin values and maximal GDR to near-baseline levels. Insulin-mediated FFA suppressibility (FFA decline at 30 min during clamp/FFA at time 0) was also markedly reduced by dexamethasone (P = 0.002). Troglitazone had no effect per se, but it was able to normalize FFA suppressibility in subjects coadministered dexamethasone. Futhermore, the magnitudes of response of FFA suppressibility and GDR to dexamethasone were proportionate. The same was true for the reversal of dexamethasone-induced insulin resistance by troglitazone, but not in response to troglitazone alone. Leptin levels were increased 2.2-fold above baseline by dexamethasone. Again, troglitazone had no effect per se but blocked the dexamethasone-induced increase in leptin. Subjects experienced a 1.7-kg weight gain while taking troglitazone but no other untoward effects. We conclude that in healthy humans, thiazolidinediones antagonize the action of dexamethasone with respect to multiple metabolic effects. Specifically, troglitazone reverses both glucocorticoid-induced insulin resistance and impairment of glucose tolerance, prevents dexamethasone from impairing the antilipolytic action of insulin, and blocks the increase in leptin levels induced by dexamethasone. Even though changes in FFA suppressibility were correlated with dexamethasone-induced insulin resistance and its reversal by troglitazone, a cause-and-effect relationship cannot be established. However, the data suggest that glucocorticoids and thiazolidinediones exert fundamentally antagonistic effects on human metabolism in both adipose and muscle tissues. By preventing or reversing insulin resistance, troglitazone may prove to be a valuable therapeutic agent in the difficult clinical task of controlling diabetes in patients receiving glucocorticoids.  相似文献   

20.
BACKGROUND: Both ghrelin and leptin are important signals in the regulation of food intake and energy balance. Leptin concentrations are elevated in the majority of obese individuals, and its levels usually correlate with adiposity and body mass index. Ghrelin as a new growth hormone (GH)-releasing peptide was discovered in 1999. Ghrelin stimulates food intake and exhibits gastroprotective properties. Many other regulatory effects of both ghrelin and leptin involving cardiovascular, gastrointestinal, renal, and endocrine systems were revealed. New experimental studies show both hormones as new acute phase reactants in animal models of inflammatory reaction. The aim of this study was to characterize the levels of circulating ghrelin and leptin in relation to systemic inflammatory response. We used a postoperative bacterial sepsis after large abdominal surgery as a model of cytokine network hyperstimulation. PATIENTS AND METHODS:The prospective study was performed on 25 surgical patients with proven postoperative intra-abdominal sepsis after large abdominal surgery. Plasma levels of ghrelin (RIA), leptin, TNF-alpha, IL-1beta, sIL-2R, IL-6 (ELISA analysis), CRP and alpha1-antitrypsin (nephelometric analysis) were analyzed. RESULTS: Authors demonstrate statistically significant elevation of plasma ghrelin (492.3+/-70.6 ng/l) and leptin (31.6+/-12.2 microg/l) compared with the control group (336.5+/-46,1, p<0.01 for ghrelin, 3.5+/-1.2 microg/l, p<0.001 for leptin). The regression coefficient was the highest for ghrelin and IL-6 (r=0,44, p<0.05), and for ghrelin and TNF (r=0.43, p<0.05) in the sepsis group. In regard to leptin, the regression coefficient was the highest for IL-6 and leptin (r=0.53, p<0.05) and for leptin and CRP (r=0.51, p<0.05). There was no significant correlation between ghrelin and IL-1beta, ghrelin and sIL-2R, and leptin and IL-1beta. CONCLUSIONS: During postoperative intra-abdominal sepsis, both ghrelin and leptin plasma levels are elevated and positively correlate with both inflammatory cytokines (TNF-alpha, IL-6) and main APP member (CRP). It supports experimental finding that TNF-alpha and IL-6 can be important regulatory factors of their synthesis. This hormonal reaction is not specific to sepsis--the significant increase of both ghrelin and leptin occurs during an uncomplicated postoperative response, although in a lesser extent than was shown in sepsis.  相似文献   

设为首页 | 免责声明 | 关于勤云 | 加入收藏

Copyright©北京勤云科技发展有限公司  京ICP备09084417号