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 31 毫秒
1.
Familial hypobetalipoproteinemia (FHBL) subjects may develop fatty liver. Liver fat was assessed in 21 FHBL with six different apolipoprotein B (apoB) truncations (apoB-4 to apoB-89) and 14 controls by magnetic resonance spectroscopy (MRS). Liver fat percentages were 16.7 +/- 11.5 and 3.3 +/- 2.9 (mean +/- SD) (P = 0.001). Liver fat percentage was positively correlated with body mass index, waist circumference, and areas under the insulin curves of 2 h glucose tolerance tests, suggesting that obesity may affect the severity of liver fat accumulation in both groups. Despite 5-fold differences in liver fat percentage, mean values for obesity and insulin indexes were similar. Thus, for similar degrees of obesity, FHBL subjects have more hepatic fat. VLDL-triglyceride (TG)-fatty acids arise from plasma and nonplasma sources (liver and splanchnic tissues). To assess the relative contributions of each, [2H2]palmitate was infused over 12 h in 13 FHBL subjects and 11 controls. Isotopic enrichment of plasma free palmitate and VLDL-TG-palmitate was determined by mass spectrometry. Non-plasma sources contributed 51 +/- 15% in FHBL and 37 +/- 13% in controls (P = 0.02). Correlations of liver fat percentage and percent VLDL-TG-palmitate from liver were r = 0.89 (P = 0.0001) for FHBL subjects and r = 0.69 (P = 0.01) for controls. Thus, apoB truncation-producing mutations result in fatty liver and in altered assembly of VLDL-TG.  相似文献   

2.
Dyslipidemia is common in patients with HIV infection. In this study, a two-stage euglycemic hyperinsulinemic clamp, with infusion of stable isotopically labeled tracers, was used to evaluate insulin action in skeletal muscle, liver, and adipose tissue in HIV-infected men with dyslipidemia (HIV-DL; plasma triglyceride >250 mg/dl and HDL <45 mg/dl; n=12), HIV-infected men without dyslipidemia (HIV w/o DL; n=12), and healthy men (n=6). Basal rates of glucose production (glucose R(a)), glucose disposal (glucose R(d)), and lipolysis (palmitate R(a)) were similar between groups. The relative suppression of glucose R(a) (63+/- 4, 77+/- 2, and 78+/- 3%, P=0.008) and palmitate R(a) (49+/-4, 63+/-3, and 68+/-3%, P=0.005) during ow-dose insulin infusion (plasma insulin approximately 30 microU/ml), and the relative stimulation of glucose R(d) (214+/-21, 390+/-25, and 393+/-46%, P=0.001) during high-dose insulin infusion (plasma insulin approximately 75 microU/ml) were lower in HIV-DL than in HIV w/o DL and healthy volunteers, respectively. Suppression of basal glucose R(a) correlated with plasma adiponectin (r=0.44, P=0.02) and inversely with plasma IL-6 (r=-0.49, P<0.001). Stimulation of glucose R(d) correlated directly with adiponectin (r=0.48, P<0.01) and inversely with IL-6 (r=-0.49, P=0.02). We conclude that dyslipidemia in HIV-infected men is indicative of multiorgan insulin resistance, and circulating adipokines may be important in the pathogenesis of impaired insulin action.  相似文献   

3.
Liver and skeletal muscle triglyceride stores are elevated in type 2 diabetes and correlate with insulin resistance. As postprandial handling of dietary fat may be a critical determinant of tissue triglyceride levels, we quantified postprandial fat storage in normal and type 2 diabetes subjects. Healthy volunteers (n = 8) and diet-controlled type 2 diabetes subjects (n = 12) were studied using a novel 13C magnetic resonance spectroscopy protocol to measure the postprandial increment in liver and skeletal muscle triglyceride following ingestion of 13C-labeled fatty acids given with a standard mixed meal. The postprandial increment in hepatic triglyceride was rapid in both groups (peak increment controls: +7.3 +/- 1.5 mmol/l at 6 h, P = 0.002; peak increment diabetics: +10.8 +/- 3.4 mmol/l at 4 h, P = 0.009). The mean postprandial incremental AUC of hepatic 13C enrichment between the first and second meals (0 and 4 h) was significantly higher in the diabetes group (6.1 +/- 1.4 vs. 1.7 +/- 0.6 mmol x l(-1) x h(-1), P = 0.019). Postprandial increment in skeletal muscle triglyceride in the control group was small compared with the diabetic group, the mean 24-h postprandial incremental AUC being 0.2 +/- 0.3 vs. 1.7 +/- 0.4 mmol x l(-1) x h(-1) (P = 0.009). We conclude that the postprandial uptake of fatty acids by liver and skeletal muscle is increased in type 2 diabetes and may underlie the elevated tissue triglyceride stores and consequent insulin resistance.  相似文献   

4.
Polymorphisms in the fat mass- and obesity-associated (FTO) gene have been identified to be associated with obesity and diabetes in large genome-wide association studies. We hypothesized that variation in the FTO gene has an impact on whole body fat distribution and insulin sensitivity, and influences weight change during lifestyle intervention. To test this hypothesis, we genotyped 1,466 German subjects, with increased risk for type 2 diabetes, for single-nucleotide polymorphism rs8050136 in the FTO gene and estimated glucose tolerance and insulin sensitivity from an oral glucose tolerance test (OGTT). Distribution of fat depots was quantified using whole body magnetic resonance (MR) imaging and spectroscopy in 298 subjects. Two-hundred and four subjects participated in a lifestyle intervention program and were examined after a follow-up of 9 months. In the cross-sectional analysis, the A allele of rs8050136 in FTO was associated with a higher BMI, body fat, and lean body mass (all P < 0.001). There was a significant effect of variation in the FTO gene on subcutaneous fat (P < or = 0.05) and a trend for liver fat content, nonvisceral adipose tissue, and visceral fat (all P < or = 0.1). However, the single-nucleotide polymorphism was not associated with insulin sensitivity or secretion independent of BMI (all P > 0.05). During lifestyle intervention, there was also no influence of the FTO polymorphism on changes in body weight or fat distribution. In conclusion, despite an association with BMI and whole body fat distribution, variation in the FTO locus has no effect on the success of a lifestyle intervention program.  相似文献   

5.
The human immunodeficiency virus (HIV)-lipodystrophy syndrome is associated with fat redistribution and metabolic abnormalities, including insulin resistance. Increased intramyocellular lipid (IMCL) concentrations are thought to contribute to insulin resistance, being linked to metabolic and body composition variables. We examined 46 women: HIV infected with fat redistribution (n = 25), and age- and body mass index-matched HIV-negative controls (n = 21). IMCL was measured by 1H-magnetic resonance spectroscopy, and body composition was assessed with computed tomography, dual-energy X-ray absorptiometry (DEXA), and magnetic resonance imaging. Plasma lipid profile and markers of glucose homeostasis were obtained. IMCL was significantly increased in tibialis anterior [135.0 +/- 11.5 vs. 85.1 +/- 13.2 institutional units (IU); P = 0.007] and soleus [643.7 +/- 61.0 vs. 443.6 +/- 47.2 IU, P = 0.017] of HIV-infected subjects compared with controls. Among HIV-infected subjects, calf subcutaneous fat area (17.8 +/- 2.3 vs. 35.0 +/- 2.5 cm2, P < 0.0001) and extremity fat by DEXA (11.8 +/- 1.1 vs. 15.6 +/- 1.2 kg, P = 0.024) were reduced, whereas visceral abdominal fat (125.2 +/- 11.3 vs. 74.4 +/- 12.3 cm2, P = 0.004), triglycerides (131.1 +/- 11.0 vs. 66.3 +/- 12.3 mg/dl, P = 0.0003), and fasting insulin (10.8 +/- 0.9 vs. 7.0 +/- 0.9 microIU/ml, P = 0.004) were increased compared with control subjects. Triglycerides (r = 0.39, P = 0.05) and extremity fat as percentage of whole body fat by DEXA (r = -0.51, P = 0.01) correlated significantly with IMCL in the HIV but not the control group. Extremity fat (beta = -633.53, P = 0.03) remained significantly associated with IMCL among HIV-infected patients, controlling for visceral abdominal fat, abdominal subcutaneous fat, and antiretroviral medications in a regression model. These data demonstrate increased IMCL in HIV-infected women with a mixed lipodystrophy pattern, being most significantly associated with reduced extremity fat. Further studies are necessary to determine the relationship between extremity fat loss and increased IMCL in HIV-infected women.  相似文献   

6.
The current study was undertaken to examine metabolic and body composition correlates of fatty liver in type 2 diabetes mellitus (DM). Eighty-three men and women with type 2 DM [mean body mass index (BMI): 34 +/- 0.5 kg/m2] and without clinical or laboratory evidence of liver dysfunction had body composition assessments of fat mass (FM), visceral adipose tissue (VAT), liver and spleen computed tomography (CT) attenuation (ratio of liver to spleen), muscle CT attenuation, and thigh adiposity; these assessments were also performed in 12 lean and 15 obese nondiabetic volunteers. Insulin sensitivity was measured with a euglycemic insulin infusion (40 mU. m-2. min-1) combined with systemic indirect calorimetry to assess glucose and lipid oxidation, and with infusions of [2H2]glucose for assessment of endogenous glucose production. A majority of those with type 2 DM (63%) met CT criteria for fatty liver, compared with 20% of obese and none of the lean nondiabetic volunteers. Fatty liver was most strongly correlated with VAT (r = -0.57, P < 0.0001) and less strongly but significantly associated with BMI (r = -0.42, P < 0.001) and FM (r = -0.37, P < 0.001), but only weakly associated with subcutaneous adiposity (r = -0.29; P < 0.01). Fatty liver was also correlated with subfascial adiposity of skeletal muscle (r = -0.44; P < 0.01). Volunteers with type 2 DM and fatty liver were substantially more insulin resistant those with type 2 DM but without fatty liver (P < 0.001) and had higher levels of plasma free fatty acids (P < 0.01) and more severe dyslipidemia (P < 0.01), a pattern observed in both genders. Plasma levels of cytokines were increased in relation to fatty liver (r = -0.34; P < 0.01). In summary, fatty liver is relatively common in overweight and obese volunteers with type 2 DM and is an aspect of body composition related to severity of insulin resistance, dyslipidemia, and inflammatory markers.  相似文献   

7.
Patients with cirrhosis of the liver often have insulin resistance and elevated circulating growth hormone levels. This study was undertaken (a) to evaluate glucose intolerance, insulin resistance and abnormal growth hormone secretion and (b) to determine if GH suppression improves insulin resistance. Glucose tolerance tests (GTT), intravenous insulin tolerance tests (IVITT), arginine stimulation tests (AST) and glucose clamp studies before and during GH suppression with somatostatin were performed in a group of patients with alcohol-induced liver cirrhosis. During GTT cirrhotic subjects had a 2-hour plasma glucose of 200 +/- 9.8 ng/dl (N = 14) compared to 128 +/- 8.0 ng/dl in normal controls (N = 15), P less than 0.001. Basal GH was elevated in cirrhotic patients and in response to arginine stimulation reached a peak of 17.0 +/- 5.4 ng/ml (N = 7), compared to a peak of 11.3 +/- 1.8 ng/ml in 5 normal controls (P = NS). During IVITT patients with cirrhosis had a glucose nadir of 60.0 +/- 4.0 mg/dl (N = 9), compared to 29.0 +/- 7.0 mg/dl in controls (N = 5), P less than 0.001. Peak GH levels during IVITT were not significantly different in cirrhotics and controls. Glucose utilization rates in 4 patients with cirrhosis of the liver before somatostatin mediated GH suppression was 3.1 +/- 0.5 mg/kg/min and 6.5 +/- 1.5 mg/kg/min during somatostatin infusion, P less than 0.025. We conclude that patients with alcohol induced cirrhosis have sustained GH elevations resulting in insulin resistance which improves after GH suppression.  相似文献   

8.
A fatty liver is associated with fasting hyperinsulinemia, which could reflect either impaired insulin clearance or hepatic insulin action. We determined the effect of liver fat on insulin clearance and hepatic insulin sensitivity in 80 nondiabetic subjects [age 43 +/- 1 yr, body mass index (BMI) 26.3 +/- 0.5 kg/m(2)]. Insulin clearance and hepatic insulin resistance were measured by the euglycemic hyperinsulinemic (insulin infusion rate 0.3 mU.kg(-1).min(-1) for 240 min) clamp technique combined with the infusion of [3-(3)H]glucose and liver fat by proton magnetic resonance spectroscopy. During hyperinsulinemia, both serum insulin concentrations and increments above basal remained approximately 40% higher (P < 0.0001) in the high (15.0 +/- 1.5%) compared with the low (1.8 +/- 0.2%) liver fat group, independent of age, sex, and BMI. Insulin clearance (ml.kg fat free mass(-1).min(-1)) was inversely related to liver fat content (r = -0.52, P < 0.0001), independent of age, sex, and BMI (r = -0.37, P = 0.001). The variation in insulin clearance due to that in liver fat (range 0-41%) explained on the average 27% of the variation in fasting serum (fS)-insulin concentrations. The contribution of impaired insulin clearance to fS-insulin concentrations increased as a function of liver fat. This implies that indirect indexes of insulin sensitivity, such as homeostatic model assessment, overestimate insulin resistance in subjects with high liver fat content. Liver fat content correlated significantly with fS-insulin concentrations adjusted for insulin clearance (r = 0.43, P < 0.0001) and with directly measured hepatic insulin sensitivity (r = -0.40, P = 0.0002). We conclude that increased liver fat is associated with both impaired insulin clearance and hepatic insulin resistance. Hepatic insulin sensitivity associates with liver fat content, independent of insulin clearance.  相似文献   

9.
A fatty meal induces vasodilatation (of both resting and stimulated forearm flow) in healthy young adults, an effect most likely mediated by the vasodilator actions of insulin. We therefore hypothesized that an impaired meal-related vascular response might be an in vivo marker of vascular insulin resistance, related to the presence of diabetes and/or higher age. Postprandial vascular responses were assessed in three groups of subjects: 15 Type 2 diabetic subjects (age 58 +/- 8 yr), 15 age-, gender-, and body mass index (BMI)-matched older control subjects (age 57 +/- 9 yr), and 15 healthy young control subjects (age 33 +/- 7 yr). Studies were carried out before and 3 and 6 h after a standardized high-fat meal (1,030 kcal, 61 g fat). Forearm microvascular flows were measured by strain gauge plethysmography and large-artery function by ultrasound. Resting blood flow and hyperemic area under curve (AUC) flow were not significantly different in diabetic subjects (resting 117 +/- 42% and AUC 134 +/- 46% of premeal values) compared with age-matched controls (resting 131 +/- 39% and AUC 134 +/- 47%); however, the response in diabetic subjects was blunted compared with young controls (resting 171 +/- 67% and AUC 173 +/- 99% of premeal values; P = 0.02 and P = 0.18, respectively). On multiple regression analysis, we found that increasing age (but not BMI or diabetes) was significantly associated with impaired postprandial vascular responses (resting: r = -0.4, P = 0.002; AUC: r = -0.4, P = 0.006). Therefore, meal ingestion results in impaired vasodilator responses in older nondiabetic and diabetic adults, related to aging rather than insulin resistance.  相似文献   

10.
Insulin resistance, impaired insulin secretion, and low adiponectin levels have been shown to be predictors for type 2 diabetes. However, it is not yet clear whether these associations (1) are independent of changes in body weight, or (2) are valid for changes in glucose tolerance in the prediabetic state. Sixty-two non-diabetics (50 with normal glucose tolerance) aged 41 +/- 11 years, BMI 30.5 +/- 5.3 kg/m2 (mean +/- SD) were studied twice with a standard oral glucose tolerance test (oGTT, mean follow-up time 3.0 +/- 1.8 years (mean +/- SD) [range 0.5-6.5 years]). Insulin sensitivity and insulin secretion were estimated from oGTT using validated indices. Two-hour blood glucose during oGTT deteriorated over time (baseline 2 h glucose 6.32 +/- 0.21 VS. follow-up 2 h glucose 7.14 +/- 0.22 mM, p < 0.001) while the percentage body fat did not change (32.7 +/- 1.2 VS. 32.6 +/- 1.2%, p = 0.46). Follow-up 2 h blood glucose was predicted by adiponectin (p = 0.01), baseline insulin sensitivity (p = 0.02) and baseline insulin secretion relative to insulin sensitivity (p = 0.03) independent of sex, age, baseline 2 h blood glucose or change in percentage body fat. Our results suggest that low adiponectin levels, insulin resistance and low beta cell function predict the continuous deterioration of glucose tolerance in early prediabetic states, independent of changes in adiposity. Therefore, the early influence of these parameters should be the subject of future prevention programs to prevent deterioration of glucose tolerance.  相似文献   

11.
Two experiments were conducted to examine reproductive and endocrine responses of ram lambs to exogenous glucose. In Experiment 1, three ram lambs (6 mo of age) received 100 ml ip of saline (0.9%) daily and three animals received 50 g glucose (100 ml 50% dextrose) daily for 18 d. In Experiment 2, ten lambs (5 per group) were treated similarly for 10 d. Serum samples were collected intensively before and after GnRH treatment on the last day of both experiments. After 15 d of glucose treatment in Experiment 1, treated rams weighed 58 kg compared with 68 kg for the controls (P = 0.08). A similar numerical trend was observed in Experiment 2, suggesting that intraperitoneal glucose decreases feed intake. In both experiments, 50 g of glucose induced a rapid elevation in serum glucose to greater than 120 mg/dl compared with 70 to 80 mg/dl for the controls (P < 0.05). Serum insulin rose to over 6 ng/ml in both trials in lambs receiving glucose compared with values of about 2 ng/ml for the controls (P < 0.01). Serum growth hormone was not altered (P > 0.10) by glucose in either experiment and IGF-1 was similar (P > 0.20) between groups in Experiment 2. Although serum concentrations of prolactin tended (P = 0.14) to be reduced by glucose treatment (64 +/- 21 ng/ml) compared with that of the controls (120 +/- 21 ng/ml) in Experiment 1, the opposite trend (P = 0.20) was observed in Experiment 2. Serum thyroxine was elevated (P = 0.08) in glucose-treated rams compared with that in controls in Experiment 2 but triiodothyronine concentrations were similar (P > 0.80) between groups. In Experiment 1, area under the curve (AUC) for LH after a GnRH challenge tended (P = 0.14) to be greater in glucose-treated (1,351 units) than in control (999 +/- 139 units) animals. The AUC for FSH (Experiment 1) did not differ (P = 0.30) between groups. The LH AUC in Experiment 2 was about 2,500 units for both groups (P = 0.80). The AUC for testosterone in Experiment 1, was 5,452 and 2,597 (+/- 1051) units for rams treated with 0 and 50 g glucose/d (P = 0.13), but testosterone AUC in Experiment 2 was similar between groups (P > 0.70). No effect of exogenous glucose was evident in either experiment for semen traits. Results suggest that 50 g ip glucose daily for 10 or 18 d induced large increases in serum insulin but other metabolic and reproductive hormones were not greatly influenced.  相似文献   

12.
Eleven Merino sheep fetuses were supplemented with glucose by direct continuous intravenous infusion of 50% dextrose into the fetus from day 115 of gestation until spontaneous delivery. Infusion rates of 15 or 25 g/day per kg were used and equivalent volumes of saline were infused into 11 control fetuses. Infusion periods approximated 27 days in both groups. Fetal plasma glucose concentrations were significantly (P less than 0.001) elevated throughout glucose infusion and resulted in variable but consistently higher plasma insulin concentrations in the glucose than in the saline-infused fetuses. Glucose-infused fetuses were significantly heavier than controls (mean +/- SEM; 3.86 +/- 0.16 vs 3.28 +/- 0.24 kg, P less than 0.05) and body fat depots (in g/kg body wt.) were larger in glucose-infused than control fetuses (9.91 +/- 0.65 vs 6.73 +/- 0.37, P less than 0.005, for internal brown fat depots; 1.25 +/- 0.44 vs 0.27 + 0.13, P less than 0.05, for subcutaneous white adipose tissue). The results indicate that growth and lipid deposition in the sheep fetus are responsive to increased glucose supply, an effect which may be mediated through the actions of insulin. Mean gestation length was 146.60 +/- 1.45 days for controls and 144.18 +/- 1.23 days for glucose-infused animals (normal term 150 days).  相似文献   

13.
We evaluated the acute effects of OXM on glucose metabolism in diet-induced insulin-resistant male C57Bl/6 mice. To determine the effects on glucose tolerance, mice were intraperitoneally injected with OXM (0.75, 2.5, or 7.5 nmol) or vehicle prior to an ip glucose tolerance test. OXM (0.75 nmol/h) or vehicle was infused during a hyperinsulinemic euglycemic clamp to quantify insulin action on glucose production and disposal. OXM dose-dependently improved glucose tolerance as estimated by AUC for glucose (OXM: 7.5 nmol, 1,564 +/- 460, P < 0.01; 2.5 nmol, 1,828 +/- 684, P < 0.01; 0.75 nmol, 2,322 +/- 303, P < 0.05; control: 2,790 +/- 222 mmol.l(-1).120 min). Insulin levels in response to glucose administration were higher in 7.5 nmol OXM-treated animals compared with controls. In basal clamp conditions, OXM increased EGP (82.2 +/- 14.7 vs. 39.9 +/- 5.7 micromol.min(-1).kg(-1), P < 0.001). During insulin infusion, insulin levels were twice as high in OXM-treated mice compared with controls (10.6 +/- 2.8 vs. 4.4 +/- 2.2 ng/ml, P < 0.01). Consequently, glucose infusion rate (118.6 +/- 30.8 vs. 38.8 +/- 26.4 microl/h, P < 0.001) and glucose disposal (88.1 +/- 13.0 vs. 45.2 +/- 6.9 micromol.min(-1).kg(-1), P < 0.001) were enhanced in mice that received OXM. In addition, glucose production was more suppressed during OXM infusion (35.7 +/- 15.5 vs. 15.8 +/- 11.4% inhibition, P < 0.05). However, if these data were expressed per unit concentration of circulating insulin, OXM did not affect insulin action on glucose disposal and production. These results indicate that OXM beneficially affects glucose metabolism in diet-induced insulin-resistant C57Bl/6 mice. It ameliorates glucose intolerance, most likely because it elevates glucose-induced plasma insulin concentrations. OXM does not appear to impact on insulin action.  相似文献   

14.
Nonalcoholic fatty liver (NAFL) is a common comorbidity in patients with type 2 diabetes and links to the risk of coronary syndromes. The aim was to determine the manifestations of metabolic syndrome in different organs in patients with liver steatosis. We studied 55 type 2 diabetic patients with coronary artery disease using positron emission tomography. Myocardial perfusion was measured with [15O]H2O and myocardial and skeletal muscle glucose uptake with 2-deoxy-2-[18F]fluoro-D-glucose during hyperinsulinemic euglycemia. Liver fat content was determined by magnetic resonance proton spectroscopy. Patients were divided on the basis of their median (8%) into two groups with low (4.6 +/- 2.0%) and high (17.4 +/- 8.0%) liver fat content. The groups were well matched for age, BMI, and fasting plasma glucose. In addition to insulin resistance at the whole body level (P = 0.012) and muscle (P = 0.002), the high liver fat group had lower insulin-stimulated myocardial glucose uptake (P = 0.040) and glucose extraction rate (P = 0.0006) compared with the low liver fat group. In multiple regression analysis, liver fat content was the most significant explanatory variable for myocardial insulin resistance. In addition, the high liver fat group had increased concentrations of high sensitivity C-reactive protein, soluble forms of E-selectin, vascular adhesion protein-1, and intercellular adhesion molecule-1 (P < 0.05) and lower coronary flow reserve (P = 0.02) compared with the low liver fat group. In conclusion, in patients with type 2 diabetes and coronary artery disease, liver fat content is a novel independent indicator of myocardial insulin resistance and reduced coronary functional capacity. Further studies will reveal the effect of hepatic fat reduction on myocardial metabolism and coronary function.  相似文献   

15.
The objective of this study was to determine the change of plasma endothelin (ET)-1 concentrations and insulin resistance index after therapy for hyperthyroidism. We studied 20 patients with hyperthyroidism (15 women and 5 men; age, 34.0 +/- 2.8 years), and 31 patients with euthyroid goiters as controls (27 women, 4 men; age, 37.0 +/- 2.4 years). All hyperthyroid patients were treated with antithyroid drugs. The patients received evaluations before and after normalization of thyroid function. The evaluations included body mass index (BMI), body fat, and measurement of circulating concentrations of thyroid hormones, glucose, insulin, and ET-1. Hyperthyroid subjects had higher plasma ET-1 concentrations than the control group (P < 0.001). No significant differences in serum glucose and insulin concentrations or insulin resistance index estimated by the R value of the homeostasis model assessment (HOMA-R) were noted between the groups. Plasma ET-1 concentrations decreased after correction of hyperthyroidism compared with pretreatment (P = 0.006). Serum glucose concentrations decreased after correction of hyperthyroidism (P = 0.005). Moreover, both body weight-adjusted insulin concentrations and the HOMA-R index were also decreased after correction of hyperthyroidism compared with pretreatment (P = 0.026 and P = 0.019, respectively). Pearson's correlation revealed that plasma ET-1 levels positively correlated with serum triiodothyronine (T3) and free thyroxine (FT4) levels. Serum insulin levels and the HOMA-R index positively correlated with BMI and body fat. The HOMA-R index also positively correlated with serum T3 and FT4 levels. Neither insulin levels nor the HOMA-R index correlated with ET-1 levels. Hyperthyroidism is associated with higher plasma ET-1 concentrations. In addition, correction of hyperthyroidism is also associated with a decrease of plasma ET-1 levels as well as the insulin resistance index calculated by HOMA-R.  相似文献   

16.
We determined whether insulin therapy changes liver fat content (LFAT) or hepatic insulin sensitivity in type 2 diabetes. Fourteen patients with type 2 diabetes (age 51+/-2 yr, body mass index 33.1+/-1.4 kg/m2) treated with metformin alone received additional basal insulin for 7 mo. Liver fat (proton magnetic resonance spectroscopy), fat distribution (MRI), fat-free and fat mass, and whole body and hepatic insulin sensitivity (6-h euglycemic hyperinsulinemic clamp combined with infusion of [3-(3)H]glucose) were measured. The insulin dose averaged 75+/-10 IU/day (0.69+/-0.08 IU/kg, range 24-132 IU/day). Glycosylated hemoglobin A1c (Hb A1c) decreased from 8.9+/-0.3 to 7.4+/-0.2% (P<0.001). Whole body insulin sensitivity increased from 2.21+/-0.38 to 3.08+/-0.40 mg/kg fat-free mass (FFM).min (P<0.05). This improvement could be attributed to enhanced suppression of hepatic glucose production (HGP) by insulin (HGP 1.04+/-0.28 vs. 0.21+/-0.19 mg/kg FFM.min, P<0.01). The percent suppression of HGP by insulin increased from 72+/-8 to 105+/-11% (P<0.01). LFAT decreased from 17+/-3 to 14+/-3% (P<0.05). The change in LFAT was significantly correlated with that in hepatic insulin sensitivity (r=0.56, P<0.05). Body weight increased by 3.0+/-1.1 kg (P<0.05). Of this, 83% was due to an increase in fat-free mass (P<0.01). Fat distribution and serum adiponectin concentrations remained unchanged while serum free fatty acids decreased significantly. Conclusions: insulin therapy improves hepatic insulin sensitivity and slightly but significantly reduces liver fat content, independent of serum adiponectin.  相似文献   

17.
Plasma apolipoprotein A-IV (apoA-IV) levels are found elevated in hypertriglyceridemic patients. However, the relationship between plasma apoA-IV level and postprandial lipemia is not well known and remains to be elucidated. Thus, our objective was to study the relationship between plasma apoA-IV and postprandial TG after an oral fat load test (OFLT). Plasma apoA-IV was measured at fast and during an OFLT in 16 normotriglyceridemic, normoglucose-tolerant android obese subjects (BMI = 34.6 +/- 2.9 kg/m(2)) and 30 normal weight controls (BMI = 22.2 +/- 2.3 kg/m(2)). In spite of not statistically different fasting plasma TG levels in controls and obese patients, the former group showed an altered TG response after OFLT, featuring increased nonchylomicron TG area under the curve (AUC) compared with controls (516 +/- 138 vs. 426 +/- 119 mmol/l x min, P < 0.05). As compared to controls, obese patients showed increased apoA-IV levels both at fast (138.5 +/- 22.4 vs. 124.0 +/- 22.8 mg/l, P < 0.05) and during the OFLT (apoA-IV AUC: 79,833 +/- 14,281 vs. 68,176 +/- 17,463 mg/l x min, P < 0.05). Among the whole population studied, as among the control and obese subgroups, fasting plasma apoA-IV correlated significantly with AUC of plasma TG (r = 0.60, P < 0.001), AUC of chymomicron TG (r = 0.45, P < 0.01), and AUC of nonchylomicron TG (r = 0.62, P < 0.001). In the multivariate analysis, fasting apoA-IV level constituted an independent and highly significant determinant of AUC of plasma TG, AUC of chymomicron TG, AUC of nonchylomicron TG, and incremental AUC of plasma TG. In conclusion, we show a strong link between fasting apoA-IV and postprandial TG metabolism. Plasma fasting apoA-IV is shown to be a good marker of TG response after an OFLT, providing additional information on post-load TG response in conjunction with other known factors such as fasting TGs.  相似文献   

18.
We determined whether acquired obesity is associated with increases in liver or intra-abdominal fat or impaired insulin sensitivity by studying monozygotic (MZ) twin pairs discordant and concordant for obesity. We studied nineteen 24- to 27-yr-old MZ twin pairs, with intrapair differences in body weight ranging from 0.1 to 24.7 kg [body mass index (BMI) range 20.0-33.9 kg/m2], identified from a population-based FinnTwin16 sample. Fat distribution was determined by magnetic resonance imaging, percent body fat by dual-energy X-ray absorptiometry, liver fat by proton spectroscopy, insulin sensitivity by measuring the fasting insulin concentration, and whole body insulin sensitivity by the euglycemic insulin clamp technique. Intrapair differences in BMI were significantly correlated with those in intra-abdominal fat (r = 0.82, P < 0.001) and liver fat (r = 0.57, P = 0.010). Intrapair differences in fasting insulin correlated with those in subcutaneous abdominal (r = 0.60, P = 0.008), intra-abdominal (r = 0.75, P = 0.0001) and liver (r = 0.49, P = 0.048) fat. Intrapair differences in whole body insulin sensitivity correlated with those in subcutaneous abdominal (r = -0.72, P = 0.001) and intra-abdominal (r = -0.55, P = 0.015) but not liver (r = -0.20, P = 0.20) fat. We conclude that acquired obesity is associated with increases in intra-abdominal and liver fat and insulin resistance, independent of genetic factors.  相似文献   

19.
Lecithin:cholesterol acyltransferase (LCAT), phospholipid transfer protein (PLTP), and cholesteryl ester transfer protein (CETP) are key factors in remodeling of high density lipoproteins (HDL) and triglyceride-rich lipoproteins. We examined the effect of a large, 24 h intravenous fat load on plasma lipids and free fatty acids (FFA) as well as on plasma LCAT, PLTP, and CETP activity levels in 8 healthy men. The effect of concomitant insulin infusion was also studied, with 1 week between the study days. During Lipofundin(R) infusion, plasma triglycerides and FFA strongly increased after 8 and 24 h (P < 0.001), whereas HDL cholesterol decreased (P < 0.01). The increase in triglycerides was mitigated with concomitant insulin infusion (P < 0.05 from without insulin). Plasma LCAT activity increased by 17.7 +/- 7.7% after 8 h (P < 0.001) and by 26.1 +/- 11. 1% after 24 h (P < 0.001), PLTP activity increased by 19.7 +/- 15.6% after 24 h (P < 0.001), but CETP activity remained unchanged. Concomitant insulin infusion blunted the increase in plasma LCAT activity (P < 0.05 from without insulin), but not that in PLTP activity. One week after the first fat load, plasma non-HDL cholesterol (P < 0.02), and triglycerides (P = 0.05) were increased, whereas HDL cholesterol was decreased (P < 0.02). Plasma CETP and PLTP activity levels were increased by 34.8 +/- 30.4% (P < 0.02) and by 15.9 +/- 6.4% (P < 0.02), respectively, but LCAT activity was then unaltered. In summary, plasma LCAT, PLTP, and CETP activity levels are stimulated by a large intravenous fat load, but the time course of their responses and the effects of insulin coadministration are different. Changes in plasma LCAT and PLTP activities may be implicated in HDL and triglyceride-rich lipoprotein remodeling under the present experimental conditions.  相似文献   

20.
The development of insulin resistance in the obese individual could impair the ability to appropriately adjust metabolism to perturbations in energy balance. We investigated a 12- vs. 48-h fast on hepatic glucose production (R(a)), peripheral glucose uptake (R(d)), and skeletal muscle insulin signaling in lean and obese subjects. Healthy lean [n = 14; age = 28.0 +/- 1.4 yr; body mass index (BMI) = 22.8 +/- 0.42] and nondiabetic obese (n = 11; age = 34.6 +/- 2.3 yr; BMI = 36.1 +/- 1.5) subjects were studied following a 12- and 48-h fast during 2 h of rest and a 3-h 40 mUxm(-2)xmin(-1) hyperinsulinemic-euglycemic clamp (HEC). Basal glucose R(a) decreased significantly from the 12- to 48-h fast (lean 1.96 +/- 0.23 to 1.63 +/- 0.15; obese 1.23 +/- 0.07 to 1.07 +/- 0.07 mgxkg(-1)xmin(-1); P = 0.004) and was equally suppressed during the HEC after both fasts. The increase in glucose R(d) during the HEC after the 12-h fast was significantly decreased in lean and obese subjects after the 48-h fast (lean 9.03 +/- 1.17 to 4.16 +/- 0.34, obese 6.10 +/- 0.77 to 3.56 +/- 0.30 mgxkg FFM(-1)xmin(-1); P < 0.001). After the 12- but not the 48-h fast, insulin-stimulated AKT Ser(473) phosphorylation was greater in lean than obese subjects. We conclude that 1) 48 h of fasting produces a marked decline in peripheral insulin action, while suppression of hepatic glucose production is maintained in lean and obese men and women; and 2) the magnitude of this decline is greater in lean vs. obese subjects.  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