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 15 毫秒
1.
BACKGROUND: Although an insulin tolerance test (ITT) is the most commonly used method for detecting growth hormone (GH) deficiency (GHD) in adults, measurements of serum insulin-like growth factor-I (IGF-I) may also be of value. OBJECTIVE: To validate the use of serum IGF-I concentration in the diagnosis of GHD in adults. DESIGN: A cross-sectional study. PATIENTS: One hundred and four patients, 60 men and 44 women, with known pituitary disease and verified GHD based on ITT. MEASUREMENTS: Serum IGF-I was determined by radioimmunoassay after acid-ethanol extraction. Body composition was estimated with total body potassium combined with total body water assessments. RESULTS: According to age- and sex-adjusted population-based references values, 51 patients had serum IGF-I concentrations below -2 SD of the predicted values and 53 had concentrations within 2 SD. Fifty-seven per cent of the patients aged 41 years (25th percentile) or below and 39% of the patients aged 57 years (75th percentile) or above had serum IGF-I concentrations below -2 SD. Women had lower mean IGF-I SD scores than men (P < 0.01). Serum IGF-I was correlated with peak GH response during ITT (r = 0.40; P < 0.001), age (r = -0.27; P < 0.01), duration of hypopituitarism (r = -0.52; P < 0.001), number of pituitary hormonal deficiencies (r = -0.35; P < 0.001), body cell mass (r = 0.30; P < 0.01) and serum insulin (r = 0.21; P < 0.05). The peak GH response during ITT correlated with spontaneous GH secretion, duration (P = -0.48; P < 0.001) and number of deficiencies (r = -0.50; P 0.001). CONCLUSION: The measurement of serum IGF-I concentrations is not suitable as a single diagnostic test for growth hormone deficiency in adults. Even as a screening test, its use appears to be limited, especially in elderly subjects. The serum level of IGF-I was influenced by several factors in addition to GH, such as age, gender, anthropodometry and serum insulin level. The peak GH response during the insulin tolerance test appears to be influenced to a lesser degree by these factors.  相似文献   

2.
Due to the use of various, and mostly indirect, methods to estimate total body water (TBW) and extracellular water (ECW), there is no agreement about whether body water distribution, i.e. the ECW to TBW ratio, is normal in GH-deficient (GHD) subjects at baseline and during recombinant human GH (rhGH) treatment. We studied body water distribution in 14 patients with adult-onset GHD and in 28 healthy controls. We also investigated the effect of GH replacement therapy for 4 and 52 weeks on body water distribution. All patients started with a dose of 0.6 IU rhGH/day for the first 4 weeks. After 52 weeks, the dose varied between 0.6-1.8 IU/day. TBW and ECW were measured by dilution of deuterium and bromide, respectively. Both parameters were also estimated using multifrequency bioelectrical impedance (BIA). Patients with GHD had significantly lower ECW and TBW than healthy controls. In addition, the ECW to TBW ratio was significantly lower in GHD patients than in healthy controls. Four weeks of GH treatment significantly increased body weight, TBW, ECW, and ECW/TBW. A further increase in TBW, but not ECW, was found after 52 weeks of treatment. The mean increases in TBW and ECW from the baselines were 2.5 +/- 0.3 and 2.0 +/- 0.3 L, respectively. The correlation coefficient and the estimated reliability between measured and estimated TBW and ECW at any time point were all high (> 0.91 and > 0.95, respectively). In general, both ECW and TBW were overestimated by multifrequency BIA in GHD adults. During treatment, the overestimation of both ECW and TBW diminished. The estimation error was correlated with the level of the body water compartment and the ratio of ECW to TBW. The estimated change in ECW with rhGH treatment was underestimated by multifrequency BIA. We conclude that GHD adults have lower ECW and TBW and a lower ECW to TBW ratio, as measured by dilution techniques. The ECW to TBW ratio can be normalized within 4 weeks of rhGH treatment at a dose of 0.6 IU/day. Finally, we conclude that multifrequency impedance measurements do not give valid estimates of body water compartments in the follow-up of patients with GHD.  相似文献   

3.
The secretion of GH changes during the menstrual cycle, exhibiting high levels during the periovulatory phase (PO). Previous studies have not investigated whether this difference in GH status is due to increased secretion or reduced clearance of pituitary GH and amplified pulsatile vs. basal GH secretion. It is also unclear whether the PO phase is accompanied by changes in circulating insulin-like growth factor I (IGF-I). In this study we investigated the 24-h GH release patterns in the early follicular (EF) vs. the periovulatory menstrual phase in the same individuals. Ten young (aged 24-34 yr) healthy women with regular menses were studied with deconvolution analysis of GH profiles obtained by blood sampling every 20 min for 24 h, followed by an arginine stimulation test. A high sensitivity immunofluorometric GH assay was used. All women were studied in both the EF and PO phases in random order. There were no differences in the basal GH secretion rate or GH half-life during the two phases. The number of GH secretory bursts identified during the 24-h sampling period was significantly increased during the PO (13.3 +/- 0.5) compared to the EF (10.3 +/- 0.6) phase (P = 0.002); conversely, the mean interburst interval was shorter in the PO (107 +/- 5 min) than in the EF (134 +/- 8 min) phase (P = 0.004). There was no difference in GH pulse mass (P = 0.13) or amplitude (P = 0.21) between the two phases. The pulsatile GH production rate (milligrams per L/24 h) was significantly elevated during the PO (61 +/- 6) compared to that during the EF (37 +/- 8; P = 0.004). Increased total GH pulse area was confirmed by Cluster analysis (P = 0.027). Furthermore, the 24-h mean serum GH concentration was significantly increased in the PO (1.4 +/- 0.1 mg/L) vs. that in the EF (0.9 +/- 0.1 mg/L; P = 0.002). There was a positive correlation between estradiol (E2) and GH secretory pulse amplitude, frequency, and mean 24-h serum GH concentration in the PO cycle phase, indicating E2 to be a major statistical determinant of GH secretion. Serum GH increased significantly after arginine infusion in both phases (P < 0.001), whereas there was no difference between the two cycle phases (P = 0.20). Serum IGF-I levels were increased during the PO phase (253 +/- 20 mg/L) compared to those during the EF phase (210 +/- 16 mg/L; P = 0.03), whereas serum IGF-binding protein-3, IGF-II, and GH-binding protein were similar during the two phases. This study unequivocally documents elevated GH levels during the PO phase of the menstrual cycle, mediated by increased GH production rate and burst frequency. The concomitant increase in serum IGF-I suggests a central stimulation of the GH-IGF-I axis, which may be mediated by endogenous E2 levels.  相似文献   

4.
The use of growth hormone (GH) as an anabolic agent is limited by its tendency to cause hyperglycemia and by its inability to reverse nitrogen wasting in some catabolic conditions. In a previous study comparing the anabolic actions of GH and IGF-I (insulin-like growth factor I), we observed that intravenous infusions of IGF-I (12 micrograms/kg ideal body wt [IBW]/h) attenuated nitrogen wasting to a degree comparable to GH given subcutaneously at a standard dose of 0.05 mg/kg IBW per d. IGF-I, however, had a tendency to cause hypoglycemia. In the present study, we treated seven calorically restricted (20 kcal/kg IBW per d) normal volunteers with a combination of GH and IGF-I (using the same doses as in the previous study) and compared its effects on anabolism and carbohydrate metabolism to treatment with IGF-I alone. The GH/IGF-I combination caused significantly greater nitrogen retention (262 +/- 43 mmol/d, mean +/- SD) compared to IGF-I alone (108 +/- 29 mmol/d; P < 0.001). GH/IGF-I treatment resulted in substantial urinary potassium conservation (34 +/- 3 mmol/d, mean +/- SE; P < 0.001), suggesting that most protein accretion occurred in muscle and connective tissue. GH attenuated the hypoglycemia induced by IGF-I as indicated by fewer hypoglycemic episodes and higher capillary blood glucose concentrations on GH/IGF-I (4.3 +/- 1.0 mmol/liter, mean +/- SD) compared to IGF-I alone (3.8 +/- 0.8 mmol/liter; P < 0.001). IGF-I caused a marked decline in C-peptide (1,165 +/- 341 pmol/liter; mean +/- SD) compared to the GH/IGF-I combination (2,280 +/- 612 pmol/liter; P < 0.001), suggesting maintenance of normal carbohydrate metabolism with the latter regimen. GH/IGF-I produced higher serum IGF-I concentrations (1,854 +/- 708 micrograms/liter; mean +/- SD) compared to IGF-I only treatment (1,092 +/- 503 micrograms/liter; P < 0.001). This observation was associated with increased concentrations of IGF binding protein 3 and acid-labile subunit on GH/IGF-I treatment and decreased concentrations on IGF-I alone. These results suggest that the combination of GH and IGF-I treatment is substantially more anabolic than either IGF-I or GH alone. GH/IGF-I treatment also attenuates the hypoglycemia caused by IGF-I alone. GH/IGF-I treatment could have important applications in diseases associated with catabolism.  相似文献   

5.
The gene product from the ob gene, leptin, has recently been characterized in humans. The circulating level of leptin is related to body mass index (BMI) and more closely to estimates of total body fat, whereas visceral fat has been reported to be of minor importance. However, it is unknown if leptin is directly regulated by hormones that influence substrate metabolism and body composition. We studied leptin in adult growth hormone (GH)-deficient (GHD) patients substituted with GH treatment for 12 months in a parallel double-blind, placebo-controlled study. Twenty-seven GHD adults aged 44.9 +/- 1.9 years underwent anthropometric measurements for determination of regional and total body fat (BMI, waist to hip ratio [WHR], computed tomographic [CT] scan, dual-energy x-ray absorptiometry [DEXA] scan, and bioimpedance analysis [BIA]) before and after 12 months of placebo-controlled GH substitution (2 IU/m2) in a parallel design. The same measurements were performed in 42 healthy adults aged 39.1 +/- 1.7 years. The logarithm of serum leptin levels correlated positively with abdominal subcutaneous fat and total body fat (BIA and DEXA) in untreated GHD patients and healthy subjects. Fasting insulin did not correlate with leptin levels in either of the groups. After 12 months of GH administration, the body composition of GHD patients was significantly changed with respect to a marked decrease in body fat. The relations of leptin to the estimates of body fat were maintained, and leptin was furthermore related to BMI and fasting insulin. In multiple linear regression analyses, additional estimates of visceral adiposity (intraabdominal fat and maximal anterior-posterior diameter determined by CT scan) were significant determinants of leptin in the healthy subjects. The increase in fasting insulin levels during GH substitution correlated negatively with the reduction in leptin levels (r = -.823, P = .003). At baseline, leptin levels were increased in the patients compared with controls in both sexes (women, 21.8 +/- 3.3 v 11.3 +/- 1.4 ng/mL, P = .002; men, 8.1 +/- 1.2 v 4.7 +/- 0.7 ng/mL, P = .008). Leptin levels were similar in GHD patients treated for 12 months compared with healthy controls for both women and men (women, 15.9 +/- 2.3 and 11.3 +/- 1.4 ng/mL, P = .163; men, 7.1 +/- 2.8 and 4.7 +/- 0.7 ng/mL, P = .759). In healthy adults and in GHD patients, leptin levels were significantly higher in women than in men (11.3 +/- 1.4 v 4.7 +/- 0.7 ng/mL, P < .001; 21.8 +/- 3.3 v 8.1 +/- 1.2 ng/mL, P < .001). Gender remained a significant determinant of leptin levels in several models of multiple linear regression analysis also including age, estradiol levels, insulin, and estimates of body fat. We conclude that leptin is increased but not differently regulated in GHD patients compared with normal subjects, and that leptin levels are closely related to estimates of body fat. This relationship is maintained during a decrease in body fat due to GH substitution.  相似文献   

6.
OBJECTIVE: Recent studies have demonstrated the restoration of a normal 24 h GH profile induced by a reduction of insulinaemia after weight loss, suggesting a reciprocal relationship between plasma insulin and GH concentrations. We aimed to clarify if an opiate-induced reduction in plasma insulin could affect GH secretion in obesity. DESIGN: We have studied the insulin response to an oral glucose tolerance test (OGTT) and the GH response to GHRH before and after prolonged treatment with Naltrexone (NTX). C-peptide, IGF-I, IGFBP-3 plasma levels and the IGF-I/IGFBP-3 molar ratio were also determined. SUBJECTS: Twelve obese women (aged 25-41 y; Body mass index (BMI): 31-39 kg/m2) and six lean normal women (aged 25-38; BMI: 19.8-23.1 kg/m2). MEASUREMENT: GH was determined by the IRMA method; insulin, C-peptide, IGF-I and IGFBP-3 were assayed by the RIA method. For molar comparison between IGF-I and IGFBP-3 we have considered 30.5 kDa the molar weight of IGFBP-3. Results are expressed as mean +/- s.e.m. RESULTS: We observed a significant decrease in basal concentration of both insulin (230.1 +/- 34.9 vs 133.2 +/- 16.9 pmol/L; P < 0.005) and C-peptide (3.7 +/- 0.3 vs 2.4 +/- 0.1 micrograms/L; P < 0.02). No modifications in the insulin secretory response to the OGTT were observed. A significant increase of the GHRH-induced GH peak response (7.7 +/- 1.4 vs 19.7 +/- 3.1 micrograms/L; P < 0.01) and GH-AUC (533 +/- 151 vs 1415 +/- 339 micrograms/L/120 min; P < 0.01) was found after NTX treatment. A negative correlation was found between basal insulin and GH peak values, both before (r = -0.641, P = 0.027) and after NTX (r = -0.714, P = 0.013). No modifications were found in IGF-I, IGFBP-3 and IGF-I/IGFBP-3 molar ratio. Moreover, NTX affected neither the insulin response to OGTT or IGF-I, IGFBP-3 and IGF-I/IGFBP-3 molar ratio in a group of six lean controls. Conversely, NTX significantly reduced the GH response to GHRH, when expressed as both peak and AUC values. CONCLUSIONS: The opiate antagonist significantly reduced basal insulin concentrations and augmented the GH response to GHRH in obese subjects. In the absence of modifications in IGF-I and IGFBP-3 plasma levels and their molar ratio, we propose that insulin may exert a negative feedback on GH secretion.  相似文献   

7.
Treatment with insulin-like growth factor I (IGF-I) alone failed to affect glucocorticoid-induced protein catabolism in a previous study from our laboratory. To assess the effects of the combination of IGF-I and GH in a similar protocol, 24 normal subjects received (in a double-blind, randomized, placebo-controlled manner) s.c. injections of either GH alone (0.3 IU/kg.day), the combination of IGF-I (80 micrograms/kg.day) and GH (0.3 IU/kg.day), or placebo for a period of 6 days during which they were treated with methylprednisolone (0.5 mg/kg.day). Whole-body protein kinetics measured, using the [1-13C]-leucine infusion technique, demonstrated that leucine flux (a parameter of protein breakdown) increased during administration of glucocorticoids alone (placebo group) and during GH-treatment, whereas the glucocorticoid-induced increase was abolished during IGF-I plus GH (P < 0.03 vs. GH). Leucine oxidation (a parameter of irreversible protein catabolism) increased in the placebo group (+60 +/- 14.5%, P < 0.005, day 7 vs. day 1), remained unchanged in the GH group (+2.5 +/- 10%), and decreased in the combination group (-17.7 +/- 3.3%, P < 0.002, day 7 vs. day 1). Glucose MCR decreased in the group receiving placebo (P < 0.05) and remained unchanged during combined treatment with IGF-I plus GH. It is concluded that glucocorticoid-induced protein, catabolism (leucine oxidation) is abolished during coadministration of GH (anticatabolic effect), whereas treatment with IGF-I and GH results in a net anabolic effect without adverse effects on peripheral glucose clearance.  相似文献   

8.
To determine the effect of aging on the suppression of GH secretion by insulin-like growth factor (IGF)-I, we studied 11 healthy young adults (6 men, 5 women, mean +/- SD: 25.2 +/- 4.6 yr old; body mass index 23.7 +/- 1.8 kg/m2) and 11 older adults (6 men, 5 women, 69.5 +/- 5.8 yr old; body mass index 24.2 +/- 2.5 kg/m2). Saline (control) or recombinant human IGF-I (rhIGF-I) (2 h baseline then, in sequence, 2.5 h each of 1, 3, and 10 micrograms/kg.h) was infused iv during the last 9.5 h of a 40.5-h fast; serum glucose was clamped within 15% of baseline. Baseline serum GH concentrations (mean +/- SE: 3.3 +/- 0.7 vs. 1.9 +/- 0.5 micrograms/L, P = 0.02) and total IGF-I concentrations (219 +/- 15 vs. 103 +/- 19 micrograms/L, P < 0.01) were higher in the younger subjects. In both age groups, GH concentrations were significantly decreased by 3 and 10 micrograms/kg.h, but not by 1 microgram/kg.h rhIGF-I. The absolute decrease in GH concentrations was greater in young than in older subjects during the 3 and 10 micrograms/kg.h rhIGF-I infusion periods, but both young and older subjects suppressed to a similar GH level during the last hour of the rhIGF-I infusion (0.78 +/- 0.24 microgram/L and 0.61 +/- 0.16 microgram/L, respectively). The older subjects had a greater increase above baseline in serum concentrations of both total (306 +/- 24 vs. 244 +/- 14 micrograms/L, P = 0.04) and free IGF-I (8.5 +/- 1.4 vs. 4.2 +/- 0.6 micrograms/L, P = 0.01) than the young subjects during rhIGF-I infusion, and their GH suppression expressed in relation to increases in both total and free serum IGF-I concentrations was significantly less than in the young subjects. We conclude that the ability of exogenous rhIGF-I to suppress serum GH concentrations declines with increasing age. This suggests that increased sensitivity to endogenous IGF-I negative feedback is not a cause of the decline in GH secretion that occurs with aging.  相似文献   

9.
OBJECTIVE: Studies with GH substitution in GH-deficient (GHD) adults lasting more than 6 months have so far been uncontrolled. End-points such as physical fitness and body composition may be subject to a considerable placebo effect which weakens the validity of open studies. We therefore tested GH (2 IU/m2 per day) versus placebo treatment for 12 months. DESIGN: Twenty-nine patients (mean age 45.5 +/- 2.0 years) with adult-onset GHD were studied in a double-blind, parallel design. Measurements of body composition by means of conventional anthropometry, bioelectrical impedance (BIA), CT scan and DEXA scan, exercise capacity, and isometric muscle strength were performed at baseline and after 12 months treatment. For body composition measurements a control group of 39 healthy, age and sex-matched subjects was included. RESULTS: Sum of skinfolds (SKF) at 4 sites decreased significantly after GH treatment. Total body fat (TBF) as assessed by DEXA and BIA was elevated at baseline but normalized after GH. TBF assessed by SKF revealed significantly higher levels compared to DEXA and BIA, although all estimates intercorrelated closely. Visceral and subcutaneous abdominal fat decreased by 25 and 17%, respectively after GH (P < 0.01) to levels no longer different from the control group. CT of the mid thigh revealed a significant reduction in fat tissue and a significant increase in muscle volume after GH treatment, both of which resulted in a normalization of the muscle: fat ratio (%) (placebo: 58:42 (baseline) vs 58:42 (12 months); GH: 66:34 (baseline) vs 72:28 (12 months) (P = 0.002); normal subjects: 67:33 (P < 0.05 when compared to 12 months placebo data)). Total body resistance and resistance relative to muscle volume decreased significantly after GH treatment suggesting over-hydration as compared to normal subjects. Exercise capacity (kJ) increased significantly after GH treatment (placebo: 54.7 +/- 9.8 (baseline) vs 51.6 +/- 8.2 (12 months); GH: 64.9 +/- 13.3 (baseline) vs 73.5 +/- 13.6 (12 months) (P < 0.05)). Isometric quadriceps strength increased after GH but no treatment effect could be detected owing to a small increase in the placebo group. Serum IGF-I levels (microgram/l) were low baseline and increased markedly after GH treatment to a level exceeding that of normal subjects (270 +/- 31 (12 months GH) vs 156 +/- 8 (normal subjects (P < 0.01)). The levels of serum electrolytes and HbA1c remained unchanged. The number of adverse effects were higher in the GH group after 3 months, but not after 6 and 12 months. CONCLUSIONS: (1) The reduction in excess visceral fat during GH substitution is pronounced and sustained; (2) beneficial effects on total body fat, muscle volume and physical fitness can be reproduced during prolonged placebo-controlled conditions; (3) uncontrolled data on muscle strength must be interpreted with caution; (4) a daily GH substitution dose of 2 IU/m2 seems too high in many adult patients.  相似文献   

10.
11.
To evaluate the role of serum free or unbound insulin-like growth factor I (IGF-I) on bone growth, we measured serum free IGF-I levels in 354 healthy children and adults (193 males and 161 females, aged 0-40 yr) and in 21 prepubertal GH-deficient (GHD) children (complete GHD, n = 5; partial GHD, n = 16) using a recently developed immunoradiometric assay. We obtained the following results. 1) In the normal children, the serum free IGF-I levels were low in infancy (<1 yr of age; males, 0.71 +/- 0.26 microg/L, mean +/- SD; females, 1.05 +/- 0.49 microg/L), increased during puberty (males, 5.84 +/- 2.18 microg/L; females, 5.80 +/- 1.49 microg/L), and declined thereafter. 2) Free IGF-I in the serum occupied about 0.95-2.02% of the total IGF-I values, with the highest ratio occurring in infancy (males, 1.77 +/- 0.60%; females, 2.02 +/- 0.87%). 3) The SD scores of serum free IGF-I in the 21 GHD children ranged from -3.30 to 0.30, and the 5 complete GHD children had free IGF-I values more than -2 SD below those of age-matched normal subjects. 4) There was a significant correlation between the SD scores of free IGF-I and those of total IGF-I (r = 0.715; P < 0.0005) in the GHD children. 5) In the 16 partial GHD children receiving GH treatment, the serum free IGF-I levels were elevated to 209% of pretreatment levels after 1 month of GH treatment and remained high during GH therapy. The GH-induced increase in the serum free IGF-I levels was significantly higher than those of the total IGF-I and IGF binding protein-3 levels. 6) The percent increase in the serum free IGF-I level after 1 month of GH treatment showed a significant positive correlation with that of the GH-induced improvement in the percent increase in the height velocity during 1 yr of GH therapy (r = 0.526; P < 0.05). These results show that free IGF-I in the serum has an essential role in bone formation because the higher free IGF-I levels were observed when the growth rate accelerated. The measurement of serum free IGF-I may become a useful tool for both diagnosing GH deficiency and predicting growth responses to long term GH therapy.  相似文献   

12.
Catch-up growth was analyzed in 20 prepubertal children with primary hypothyroidism (PH) starting treatment at an age of 4.4 (1.2-10.1) years and a height (HT) SD score (HT SDS) of -3.1 (+/-0.8). All patients were followed for at least 3 prepubertal years. HT velocity was 12.3 +/- 2.3, 9.0 +/- 1.8 and 7.5 +/- 2.2 cm/year, and change in HT SDS was 1.60 +/- 0.56, 0.57 +/- 0.33 and 0.28 +/- 0.38 during the 1st, 2nd and 3rd year, respectively. The 11 children followed to adult height reached a HT SDS of -0.11 +/- 1.1, all within their target HT range. HT gain (DeltaHT SDS) during the 1st year was correlated with the degree of catch-up growth (r2 = 0.78, p < 0.001). While catch-up growth in childhood-onset PH is complete, this is not the case in GH deficiency (GHD). Based on the auxological characteristics of the patients with PH, HT velocities during the first 2 years were predicted applying prediction models devised for prepubertal children with idiopathic GHD. The modalities of GH treatment observed in the models were used to calculate predicted HT velocities of the PH patients. Observed HT velocities in PH were higher than predicted HT velocities during the 1st (10.67 +/- 1.37 cm/year, p < 0.01) and 2nd (8.35 +/- 0.86 cm/year, p = 0.128) year. The data show that catch-up potential in idiopathic GHD of childhood onset is reduced compared to PH. Since early catch-up as well as total HT recovery in children with GHD are often not reached by present treatment modalities, catch-up growth in PH may serve as a model towards optimizing GH treatment. The data suggest that initial GH doses of 1.0 IU/kg/week, rather than the presently recommended 0. 6 IU/kg/week, need to be given in GHD in order to achieve the degree of early catch-up observed in PH and to consequently improve the final outcome.  相似文献   

13.
The effect of recombinant GH on strength, body composition and endocrine parameters in power athletes was investigated in a controlled study. Twenty-two healthy, non-obese males (age 23.4 +/- 0.5 years; ideal body weight 122 +/- 3.1%, body fat 10.1 +/- 1.0%; mean +/- SEM) were included. Probands were assigned in a double-blind manner to either GH treatment (0.09U (kg BW)-1 day-1 sc) or placebo for a period of six weeks. To exclude concurrent treatment with androgenic-anabolic steroids urine specimens were tested at regular intervals for these substances. Serum was assayed for GH, IGF-I, IGF-binding proteins, insulin and thyroxine before the onset of the study and at two-weekly intervals thereafter. Maximal voluntary strength of the biceps and quadriceps muscles was measured on a strength training apparatus. Fat mass and lean body mass were derived from measurements of skinfolds at ten sites with a caliper. For final evaluation only data of those 8 and 10 subjects in the two groups who completed the study were analyzed. GH, IGF-I and IGF-binding protein were in the normal range before therapy and increased significantly in the GH-treated group. Fasting insulin concentrations increased insignificantly and thyroxine levels decreased significantly in the GH-treated probands. There was no effect of GH treatment on maximal strength during concentric contraction of the biceps and quadriceps muscles. Body weight and body fat were not changed significantly during treatment. We conclude that the anabolic, lipolytic effect of GH therapy in adults depends on the degree of fat mass and GH deficiency.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

14.
The diagnosis of GH deficiency (GHD) in the elderly is based at present on the peak GH concentration during a stimulation test. We have now evaluated the performance of urinary GH (uGH), urinary insulin-like growth factor I (uIGF-I), and urinary IGF-binding protein-3 (uIGFBP-3) in the diagnosis of GHD in this group. Twenty GHD elderly patients with a history of pituitary disease and a peak GH response to arginine stimulation of less than 3 ng/mL (15 men and 5 women; age, 61.1-83.4 yr) and 19 controls (12 men and 7 women; age, 60.8-87.5 yr) were studied. GH secretion was assessed by 24-h profile and expressed as the area under the curve (AUCGH). Serum (s) IGF-I and sIGFBP-3 were measured in a single morning, fasted sample. Urinary GH, uIGF-I, and uIGFBP-3 were measured in a 24-h urine sample collected over the same interval as the GH profile, and results were expressed as total amount excreted in 24 h (tuGH24, nanograms; tuIGF-I24, nanograms; tuIGFBP-3(24), micrograms). Data are presented as the mean +/- SD, except for AUCGH, tuGH24, and tuIGFBP-3(24), which are presented as the geometric mean (-1, +1 tolerance factor). AUCGH, sIGF-I, and sIGFBP-3 were significantly lower in GHD subjects than in controls. Total uGH24 was lower in GHD subjects, but tuIGF-I24 and tuIGFBP-3(24) excretion were not different in the two groups. AUCGH provided the best separation between GHD and control subjects, whereas there was substantial overlap for sIGF-I, sIGFBP-3, and tuGH24. In both groups sIGF-I was correlated to sIGFBP-3 (GHD, r = 0.75; controls, r = 0.65; both P < 0.01), whereas tuIGF-I24 was not correlated to tuIGFBP-3(24) in either group. Moreover, tuIGF-I24 and tuIGFBP-3(24) were not related to their respective serum concentrations in either group. Total uGH24 was correlated with AUCGH only in controls (r = 0.54; P < 0.05). These data demonstrate that urinary GH and urinary and serum IGF-I and IGFBP-3 are not suitable diagnostic markers for GHD in elderly subjects.  相似文献   

15.
It is known that GH stimulates bone turnover and that GH-deficient adults have a lower bone mass than healthy controls. In order to evaluate the influences of GH replacement therapy on markers of bone turnover and on bone mineral density (BMD) in patients with adult onset GH deficiency, a double-blind placebo-controlled study of treatment with recombinant human GH (rhGH; mean dose 2.4 IU daily) in 20 patients for 6 months and an extended open study of 6 to 12 months were conducted. Eighteen patients, fourteen men and four women, with a mean age of 44 years with adult onset GH deficiency were evaluated in the study. Compared with placebo, after 6 months serum calcium (2.39 +/- 0.02 vs 2.32 +/- 0.02 mmol/l, P = 0.037) and phosphate (0.97 +/- 0.06 vs 0.75 +/- 0.05 mmol/l, P = 0.011) increased and the index of phosphate excretion (0.03 +/- 0.03 vs 0.19 +/- 0.02, P < 0.001) decreased significantly, and there was a significant increase in the markers of bone formation (osteocalcin, 64.8 +/- 11.8 vs 17.4 +/- 1.8 ng/ml, P < 0.001; procollagen type I carboxyterminal propeptide (PICP), 195.3 +/- 26.4 vs 124.0 +/- 15.5 ng/ml, P = 0.026) as well as those of bone resorption (type I collagen carboxyterminal telopeptide (ICTP), 8.9 +/- 1.2 vs 3.3 +/- 0.5 ng/ml, P < 0.001; urinary hydroxyproline, 0.035 +/- 0.006 vs 0.018 +/- 0.002 mg/100 ml glomerular filtration rate, P = 0.009). BMD did not change during this period of time. IGF-I was significantly higher in treated patients (306 +/- 45.3 vs 88.7 +/- 22.5 ng/ml, P < 0.001). An analysis of the data compiled from 18 patients treated with rhGH for 12 months revealed similar significant increases in serum calcium and phosphate, and the markers of bone turnover (osteocalcin, PICP, ICTP, urinary hydroxyproline). Dual energy x-ray absorptiometry (DXA)-measured BMD in the lumbar spine (1.194 +/- 0.058 vs 1.133 +/- 0.046 g/cm2, P = 0.015), femoral neck (1.009 +/- 0.051 vs 0.936 +/- 0.034 g/cm2, P = 0.004), Ward's triangle (0.881 +/- 0.055 vs 0.816 +/- 0.04 g/cm2, P = 0.019) and the trochanteric region (0.869 +/- 0.046 vs 0.801 +/- 0.033 g/cm2, P = 0.005) increased significantly linearly (compared with the individual baseline values). At 12 months, BMD in patients with low bone mass (T-score < -1.0 S.D.) increased more than in those with normal bone mass (lumbar spine 11.5 vs 2.1%, P = 0.030, and femoral neck 9.7 vs 4.2%, P = 0.055). IGF-I increased significantly in all treated patients. In conclusion, treatment of GH-deficient adults with rhGH increases bone turnover for at least 12 months. BMD in the lumbar spine and the proximal femur increases continuously in this time (open study) and the benefit is greater in patients with low bone mass. Therefore, GH-deficient patients exhibiting osteopenia or osteoporosis should be considered candidates for GH supplementation. However, long-term studies are needed to establish that the positive effects on BMD are persistent and are associated with a reduction in fracture risk.  相似文献   

16.
Administration of GH complexed with monoclonal antibodies (MABs) potentiates the in vivo actions of the hormone. In particular, growth and serum IGF-I concentrations of GH-treated hypophysectomized rats are increased by concomitant injection of anti-GH MABs. Among 37 anti-bovine GH (bGH) MABs, we selected one MAB with the most potentiating effects to investigate the mechanisms responsible for this phenomenon. Hypophysectomized rats were killed 18 h after a single s.c. injection of bGH (100 micrograms/rat), alone or complexed with increasing doses of MAB (4, 40, 400 micrograms/rat; MAB:bGH molar ratio: 0.005, 0.05, 0.5). IGF-I was measured by radioimmunoassay in acid-extracted sera and livers, whereas liver IGF-I mRNA was quantified by Northern blot hybridization. The in vivo occupancy of liver somatogenic (GH) receptors was derived from the determinations of total and free 125I-labelled bGH binding to liver homogenates treated with 4 mol MgCl2/l or water. Injection of MAB-bGH complexes enhanced body weight gain and raised serum IGF-I, liver IGF-I and liver IGF-I mRNA more than bGH alone (1.6-, 6-, 10- and 7-fold increases at the highest dose of MAB, compared with bGH alone; P < 0.001). These potentiating effects of the MAB were dose-dependent and significant potentiation of the growth response was already observed with the lowest dose of MAB. In vivo occupancy of liver GH receptors was markedly higher 18 h after injection of MAB-bGH complexes than after bGH alone, and this effect was also dose-dependent (receptor occupancy of 28%, 37% and 83% after 4, 40 and 400 micrograms of MAB respectively compared with 6% after bGH alone; P < 0.05, 0.05 and 0.001 respectively). In contrast, the in vitro binding of 125I-labelled bGH to liver homogenates was decreased in the presence of high doses of MAB. We conclude that low amounts of MABs complexed with bGH potentiate the stimulation by the hormone of liver IGF-I synthesis and secretion in a dose-dependent manner. These effects are mediated, at least in part, through changes in hormone-receptor interaction in vivo, leading to enhanced and/or prolonged binding of bGH to its somatogenic receptors.  相似文献   

17.
To determine the time course of recovery of GH release from insulin-like growth factor I (IGF-I) suppression, 11 healthy adults (18-29 yr) received, in randomized order, 4-h i.v. infusions of recombinant human IGF-I (rhIGF-I; 3 microg/kg-h) or saline (control) from 25.5-29.5 h of a 47.5-h fast. Serum GH was maximally suppressed within 2 h and remained suppressed for 2 h after the rhIGF-I infusion; during this 4-h period, GH concentrations were approximately 25% of control day levels [median (interquartile range), 1.2 (0.4-4.0) vs. 4.8 (2.8-7.9) microg/L; P < 0.05]. A rebound increase in GH concentrations occurred 5-7 h after the end of rhIGF-I infusion [7.6 (4.6 -11.7) vs. 4.3 (2.5-6.0) microg/L; P < 0.05]. Thereafter, serum GH concentrations were similar on both days. Total IGF-I concentrations peaked at the end of the rhIGF-I infusion (432 +/- 43 vs. 263 +/- 44 microg/L; P < 0.0001) and remained elevated 18 h after the rhIGF-I infusion (360 +/- 36 vs. 202 +/- 23 microg/L; P = 0.001). Free IGF-I concentrations were approximately 140% above control day values at the end of the infusion (2.1 +/- 0.4 vs. 0.88 +/- 0.3 microg/L; P = 0.001), but declined to baseline within 2 h after the infusion. The close temporal association between the resolution of GH suppression and the fall of free IGF-I concentrations, and the lack of any association with total IGF-I concentrations suggest that unbound (free), not protein-bound, IGF-I is the major IGF-I component responsible for this suppression. The rebound increase in GH concentrations after the end of rhIGF-I infusion is consistent with cessation of an inhibitory effect of free IGF-I on GH release.  相似文献   

18.
It is unknown whether hypogonadism contributes to decreased insulin-like growth factor I (IGF-I) production and/or how testosterone administration may effect the GH-IGF-I axis in human immunodeficiency virus (HIV)-infected men with the acquired immunodeficiency syndrome (AIDS) wasting syndrome (AWS). In this study, we investigate the GH-IGF-I axis in men with the AWS and determine the effects of testosterone on GH secretory dynamics, pulse characteristics determined from overnight frequent sampling, arginine stimulation, and total and free IGF-I levels. Baseline GH-IGF-I parameters in hypogonadal men with AWS (n=51) were compared before testosterone administration (300 mg, im, every 3 weeks vs. placebo for 6 months) with cross-sectional data obtained in two age-matched control groups: eugonadal men with AIDS wasting (n=10) and healthy age-matched normal men (n=15). The changes in GH-IGF-I parameters were then compared prospectively in testosterone- and placebo-treated patients. Mean overnight GH levels [1.8+/-0.3 and 2.4+/-0.3 vs. 0.90+/-0.1 microg/L (P=0.04 and P=0.003 vs. healthy controls)] and pulse frequency [0.35+/-0.06 and 0.37+/-0.02 vs. 0.22+/-0.03 pulses/h (P=0.06 and P=0.002 vs. healthy controls)] were comparably elevated in the eugonadal and hypogonadal HIV-positive groups, respectively, compared to those in the healthy control group. No significant differences in pulse amplitude, interpulse interval, or maximal GH stimulation to arginine administration (0.5 g/kg, i.v.) were seen between either the eugonadal and hypogonadal HIV-positive or healthy control patients. In contrast, IGF-I levels were comparably decreased in both HIV-positive groups compared to the healthy control group [143+/-16 and 165+/-14 vs. 216+/-14 microg/L (P=0.004 and P=0.02 vs. healthy controls)]. At baseline, before treatment with testosterone, overnight GH levels were inversely correlated with IGF-I (r=-0.42; P=0.003), percent ideal body weight (r=-0.36; P=0.012), albumin (r=-0.37; P=0.012), and fat mass (r=-0.52; P=0.0002), whereas IGF-I levels correlated with free testosterone (r=0.35; P=0.011) and caloric intake (r=0.32; P= 0.023) in the hypogonadal HIV-positive men. In a stepwise regression model, albumin (P=0.003) and testosterone (P=0.011) were the only significant predictors of GH [mean GH (microg/L)=-1.82 x albumin (g/dL) + 0.003 x total testosterone (microg/L) + 6.5], accounting for 49% of the variation in GH. Mean overnight GH levels decreased significantly in the testosterone-treated patients compared to those in the placebo-treated hypogonadal patients (0.9+/-0.3 vs. 0.2+/-0.4 microg/L; P=0.020). In contrast, no differences in IGF-I or free IGF-I were observed in response to testosterone administration. The decrement in mean overnight GH in response to testosterone treatment was inversely associated with increased fat-free mass (r=-0.49; P= 0.024), which was the only significant variable in a stepwise regression model for change in GH [change in mean GH (microg/L)=-0.197 x kg fat-free mass - 0.53] and accounted for 27% of the variation in the change in GH. In this study, we demonstrate increased basal GH secretion and pulse frequency in association with reduced IGF-I concentrations, consistent with GH resistance, among both hypogonadal and eugonadal men with AIDS wasting. Testosterone administration decreases GH in hypogonadal men with AIDS wasting. The change in GH is best predicted by and is inversely related to the magnitude of the change in lean body mass in response to testosterone administration. These data demonstrate that among hypogonadal men with the AWS, testosterone administration has a significant effect on the GH axis.  相似文献   

19.
In obesity, there is a markedly decreased GH secretion. The diagnosis of GH deficiency (GHD) in adults is based on peak GH responses to stimulation tests. In the severely obese, peak GH levels after pharmacological stimulation are often in the range that is observed in hypopituitary patients. To distinguish obese subjects from GHD patients, it will be necessary to demonstrate that reduced GH responsiveness to a given test is reversible in the former, but not in the latter, group. Recent studies have shown that reduction of plasma free fatty acids (FFA) with acipimox in obese patients restores their somatotrope responsiveness. There are no data evaluating GH responsiveness to acipimox plus GHRH in obese adults with hypopituitarism. The aim of the present study was to evaluate the effect of acute pharmacological reduction of plasma FFA on GHRH-mediated GH secretion in obese normal subjects and obese adults with hypopituitarism. Eight obese patients with a body mass index of 34.2+/-1.2; eight obese adults with hypopituitarism, with a body mass index of 35.5+/-1.9; and six control subjects were studied. All the patients showed an impaired response to an insulin-tolerance test (0.15 U/kg, i.v.), with a peak GH secretion of less than 3 microg/L. Two tests were carried out. On one day, they were given GHRH (100 microg, i.v., 0 min), preceded by placebo; and blood samples were taken every 15 min for 60 min. On the second day, they were given GHRH (100 microg, i.v., 0 min), preceded by acipimox (250 mg, orally, at -270 min and -60 min); and blood samples were taken every 15 min for 60 min. The administration of acipimox induced a FFA reduction during the entire test. Normal control subjects had a mean peak (microg/L) of 23.8+/-4.8 after GHRH-induced GH secretion; previous acipimox administration increased GHRH-induced GH secretion, with a mean peak of 54.7+/-14.5. In obese patients, GHRH-induced GH secretion was markedly reduced, with a mean peak (microg/L) of 3.9+/-1; previous administration of acipimox markedly increased GHRH-mediated GH secretion, with a mean peak of 16.0+/-3.2 (P < 0.05). In obese adults with hypopituitarism, GHRH-induced GH secretion was markedly reduced, with a mean peak (microg/L) of 2+/-0.7; previous acipimox administration did not significantly modify GHRH-mediated GH secretion, with a mean peak of 3.3+/-1.1 (P < 0.05). The GH response of obese patients and obese adults with hypopituitarism was similar after GHRH alone. In contrast, the GH response after GHRH plus acipimox, was markedly decreased in obese adults with hypopituitarism (mean peak, 3.3+/-1.1), compared with obese patients (mean peak, 16.0+/-3.2) (P < 0.05) and control subjects (mean peak, 54.7+/-14.5) (P < 0.01). In conclusion, GH secretion, after GHRH-plus-acipimox administration, is reduced in obese adults with hypopituitarism patients, when compared with obese normal patients. Testing with GHRH plus acipimox is safe and is free from side effects and could be used for the diagnosis of GHD in adults.  相似文献   

20.
In children and adolescents with type 1 diabetes, we have reported an association between duration of puberty and the prevalence of nephromegaly and microalbuminuria (MA), which are early markers of diabetic nephropathy. Growth hormone (GH), IGF-I, testosterone, and prorenin are potential mediators of this effect. This study examined the relationship of these hormonal factors to kidney volume (KV) and MA in 155 subjects (78 males, age 13.2 +/- 3.5 years [mean +/- SD]) with similar diabetes duration (6.83 +/- 1.6 years) but varying pubertal experience (0-10 years). KV (by ultrasound), plasma IGF-I, testosterone, prorenin, and NaLi countertransport, and urinary albumin, urinary GH, and urinary IGF-I from three 24-h collections were measured. Multiple regression analysis showed that BSA (P < 0.0001) and urinary IGF-I (P = 0.001) were significantly associated with KV. MA subjects (albumin excretion rate 15-200 microg/min) had higher urinary IGF-I (P = 0.005) and urinary GH (P = 0.05) compared with normoalbuminuric subjects. Only 9% of the variance in urinary IGF-I could be attributed to plasma IGF-I (r = 0.30, P < 0.0001). Testosterone and prorenin were not associated with MA, but they were associated with KV in univariate analyses. The strong association of urinary IGF-I with KV, a marker for glomerular hypertrophy, and of both urinary IGF-I and urinary GH with MA suggests a role for these growth factors in the development of human diabetic nephropathy. Together, these data support animal studies that have shown that renal GH and IGF-I may contribute significantly to the pathogenesis of early diabetic nephropathy.  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