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.
The 24-h secretory pattern of GH was evaluated in normal women (n = 4) and hyperprolactinemic-amenorrheic women with pituitary microadenoma (n=6). Serum GH concentrations significantly differed according to the time of day, and there were nocturnal rises in normal women and hyperprolactinemic women. However, episodic fluctuations in serum GH levels were almost absent during the day time in hyperprolactinemic women, and the mean 24-h GH concentration was significantly lower in these women than in normal women. Thus, GH secretion was diminished in hyperprolactinemic women with pituitary microadenoma, though a nocturnal increase in GH secretion continued.  相似文献   

2.
To investigate the efficacy of endocrine evaluation in diagnosing and localizing the cause of anterior pituitary failure, 17 patients with suprasellar space-occupying lesions, 4 patients with intrasellar tumors, 8 patients with no detectable anatomical lesion, 1 patient with posttraumatic failure and 1 patient with septooptical dysplasia were investigated. Endocrine evaluation consisted of measuring adrenocorticotropic hormone (ACTH), cortisol, and growth hormone (GH) levels during insulin hypoglycemia test (IHT) and after administration of corticotropin-releasing hormone (CRH) and growth hormone-releasing hormone (GRH). In addition, basal prolactin levels, gonadal and thyroid function were evaluated. The results showed that 4 of 17 patients with suprasellar tumors had normal ACTH and GH responses during IHT and after releasing hormone (RH) administration. Five of these patients had a normal ACTH or cortisol rise but no GH response during IHT. All 5 had a normal ACTH and 3 had normal GH rise after RH. Seven patients with suprasellar tumors had no ACTH or GH response during IHT, but all had an ACTH response to CRH. Only 3 of this group had a GH response to GRH. There was one exception of a patient who showed a GH and ACTH rise during IHT but only a blunted ACTH and no GH rise after RH administration. Four patients with pituitary failure and no demonstrable lesion had an ACTH rise after CRH but no GH rise after GRH, whereas in 3 patients with isolated ACTH deficiency no ACTH rise after CRH was seen. In 4 patients with nonsecreting pituitary tumors normal ACTH responses to IHT and CRH were seen, whereas GH rose during IHT only in 1 patient.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

3.
The response of growth hormone (GH) to acute administration of GH-releasing hormone 1-40 (GHRH) was evaluated in 12 patients with Turner's syndrome and in 12 prepubertal or early pubertal girls. In 7 of 12 patients GHRH induced a definite increase (greater than 10 ng/ml) of plasma GH levels. In 5 patients there was a poor GH rise after GHRH administration (less than 10 ng/ml). Overall, the mean GH response of patients was significantly lower than that of normal girls. Five out of 7 patients with a 45 X,O karyotype had a reduced GH rise after GHRH, while all patients with non X,O karyotype (mosaicism and/or 46 X,iX) had a normal GH response to GHRH. Although the cause of short stature in patients with Turner's syndrome is most likely multifactorial, a reduced pituitary GH reserve, as documented by the reduced GH response to GHRH in some of our patients, may contribute to the growth impairment in this disorder.  相似文献   

4.
To establish a single and reliable test for evaluating growth hormone (GH) secretion, we examined successive GH provocation by two agents with different modes of action, GH releasing-hormone (GHRH) and arginine (Arg) in 60 children of short stature, 6 patients with pituitary dwarfism and 9 normal young adults. Their GH profiles were qualitatively classified into 4 types: 25 children and 7 adults responded to both stimuli with 2 GH peaks (48.7 +/- 4.3 [SEM] micrograms/L for GHRH and 32.2 +/- 2.6 micrograms/L for Arg in children; 25.8 +/- 7.6 micrograms/L and 30.1 +/- 9.2 micrograms/L respectively in adults) (type A). A single peak for GHRH (57.7 +/- 4.6 micrograms/L) without an Arg-induced peak was obtained in 29 younger children (type B), which is considered to be a GHRH-dominant pattern. Two of them were diagnosed as hypothalamic GHRH deficiency based on a low nocturnal plasma GH and good response to GH treatment. Six adolescents and 2 adults showed a blunted response to GHRH (9.0 +/- 1.1 micrograms/L) but a normal response to Arg (40.6 +/- 9.5 micrograms/L) (type C), which appears to be caused by somatostatin (SRIH) hypertonicity. None with pituitary dwarfism responded to both stimuli (4.5 +/- 1.3 and 2.3 +/- 0.5 micrograms/L). Thus, the GHRH-Arg test makes it possible to evaluate the counterbalance between GHRH and SRIH as well as to differentiate pituitary GH deficiency from hypothalamic GHRH dysfunction.  相似文献   

5.
Fifty-seven children with growth hormone deficiency and 15 healthy age-matched controls were studied by magnetic resonance imaging (MRI). Of the patients, 36 (63%) had isolated GH deficiency (IGHD) and 21 (37%) multiple pituitary hormone deficiency (MPHD). MRI studies showed a marked reduction in pituitary volume in all patients in comparison with normal controls. Moreover, a striking morphological abnormality with the apparent absence of the pituitary stalk and an ectopic posterior pituitary lobe was detected in 34 of the patients (59%). This pituitary stalk abnormality was detected in 95% of the MPHD patients and in 39% of the IGHD patients. All but one of the patients with a normal pituitary stalk had IGHD. Endocrine evaluation showed no correlation with MRI data: in particular patients with an apparent anatomical interruption of the hypothalamic-pituitary axis showed a variety of patterns of hormonal responses. In conclusion, our study shows a high frequency of hypothalamic-pituitary anomalies in patients with GH deficiency, particularly related with MPHD. However, further studies are needed to improve our understanding of the relationship between MRI and endocrine data.  相似文献   

6.
Rat anterior pituitary glands were incubated in buffered medium, pH 7.4, containing 32Pi. After incubation the tissue and medium were separated and a post-mitochondrial supernate (PMS) of the tissue homogenate was prepared. Gel filtration of the PMS and medium resulted in a radioactive peak which coincided with the elution volume of authentic rat growth hormone (rGH). Polyacrylamide gel electrophoresis of the radioactive peak under denaturing condition resulted in a protein-staining band having the same mobility as authentic rGH. Autoradiography of the gels revealed radioactivity precisely at the position of growth hormone as well as elsewhere. The specific radioactivity of the PMS [32P]GH was estimated to be 5 to 10 times greater than that of tissue [32P]GH. These results indicate that phosphorylated GH is synthesized and secreted by pituitary glands in vitro.  相似文献   

7.
ObjectiveTo analyze the possible causes of growth hormone (GH) deficiency, whether isolated (GHD) or in combination with other pituitary deficiencies classified as idiopathic.Patients and methodsWe studied patients with idiopathic GHD included in a protocol of recombinant GH treatment in adults attending the outpatient clinic of the Endocrinology and Nutrition Service of the San Cecilio University Hospital. Perinatal history, findings on magnetic resonance imaging (MRI) of the hypothalamic-pituitary axis and diagnosis of GHD and other deficiencies were retrospectively evaluated.ResultsA total of 17 patients were included: 14 men and 3 women with a mean age at diagnosis of 8.4±7.3 years. Perinatal adverse events occurred in 12 patients (69.2%). MRI showed empty sella (2 patients), pituitary hypoplasia or absence of the pituitary stalk (7 patients) and pituitary hypoplasia with ectopic posterior pituitary gland (6 patients); in the remaining 2 patients these data were not available. All had an established diagnosis of GHD: 15 with (88.2%) gonadotropin deficiency, 9 (52.9%) with adrenocorticotropic hormone (ACTH) deficiency and 8 (47.1%) with thyroid-stimulating hormone (TSH) deficiency.ConclusionsIn our patients, adverse events during pregnancy or the perinatal period and the presence of anatomical abnormalities identified by MRI are a marker of pituitary dysfunction and may be important in the pathogenesis of this entity. The clinical spectrum of disease varies from isolated GH deficiency to multiple pituitary hormone deficiencies.  相似文献   

8.
9.
Although the primary control of gonadotropin secretion is by the hypothalamic GnRH and the gonadal function is controlled by the pituitary gonadotropins and prolactin, the emerging evidence suggests a vital role of the somatotropic axis, growth hormone (GH), and insulin-like growth factor-I (IGF-I) in the control of the pituitary and gonadal functions. It has been shown that GH deficiency, GH resistance, and experimental alterations in IGF-I secretion modify folliculogenesis, ovarian maturation, ovulation, and pregnancy, and in the male, GH/IGF-I plays an important role in spermatogenesis and the Leydig cell function. The primary focus of this review is to examine the role of GH/ IGF-I on the onset of puberty, fertility, pituitary, and gonadal endocrine functions. A number of studies have revealed that fertility is affected in GH-deficient dwarf and in IGF-I gene-ablated mice, possibly due to subnormal function of either the pituitary gland or the gonads. In the female GH receptor gene knockout (GHR-KO) mice, there was impairment in follicular development, ovulation rate, sexual maturation, production of and responsiveness to pheromonal signals, and the corpus luteum function. In IGF-I-deficient male GHR-KO mice, puberty is delayed, spermatogenesis is affected, and neuroendocrine-gonadal function is attenuated. Similarly, in some of the human Laron syndrome patients, puberty is delayed due to GH resistance. These data suggest that, in addition to GnRH and gonadotropins, GH/IGF-I influences the pituitary and gonadal functions in animals and humans.  相似文献   

10.
Active acromegaly is almost always the result of a benign growth hormone (GH)-secreting adenoma of the pituitary gland. Because the same pituitary stem cell can produce both GH and prolactin (PRL), many acromegalic patients also have hyperprolactinemia. The advantages of surgical excision of pituitary adenomas associated with acromegaly include: (1) prompt decrease in GH; (2) reliable and immediate relief of the mass effect from the tumor (decompression of the optic nerves and chiasm), and (3) the opportunity to obtain tumor tissue for characterization and investigative study. Currently, more than 97% of operations for removal of pituitary tumors associated with acromegaly are done using the transsphenoidal approach rather than craniotomy. Technical advances to make the surgery safer continue to evolve, and include endoscopic approaches, computer-guided image-based intraoperative visualization, and intraoperative magnetic resonance imaging. Criteria for satisfactory remission of acromegaly after surgery are the same as those used for medical management. They include normal insulin-like growth factor (IGF)-I and suppression of GH to undetectable levels (<1.0 ng/ml) during an oral glucose tolerance test (OGTT). Data from a recent series of 86 patients operated upon for acromegaly at the University of Virginia and followed for more than 1 year have been reviewed. In patients receiving surgery as the initial procedure, 67% had a normal IGF-I, and 52% suppressed to <1.0 ng/ml in an OGTT. There was one true recurrence of disease diagnosed 81 months after surgery. Results are best in patients with noninvasive microadenomas. Gamma knife radiosurgery has been a valuable adjunct in those patients who fail to achieve postoperative remission. Pathological evaluation of the tumors revealed that 16% expressed GH only, 25% stained for GH and glycoprotein hormones (follicle stimulating hormone, thyroid hormone, thyroid stimulating hormone, alpha-subunit), 21% for GH and PRL, and 33% for GH, PRL and glycoprotein hormones. There was one acidophil stem cell tumor and 10% had the mammosomatotroph subtype. This contemporary series was free of mortality or serious complications. One patient had a transient cerebrospinal fluid leak and 3 developed transient SIADH with hyponatremia. Surgical treatment remains an important aspect of the combined management of patients with acromegaly.  相似文献   

11.
Ectopic acromegaly is a very rare clinical entity occurring in less than 1% of acromegalic patients. In most cases it is caused by GHRH or rarely GH-secreting neoplasms. Even rarer are ectopic pituitary adenomas located in the sphenoid sinus or nasopharynx that originate from pituitary remnants in the craniopharyngeal duct. This dissertation presents the difficulties in visualizing GH-secreting adenoma located in the sphenoid sinus. A 55-year-old man had somatic features of acromegaly for several years. MRI imaging revealed a slightly asymmetric pituitary gland (14 yen 4 mm) without focal lesions. Simultaneously, a spherical mass, 10 mm in diameter, corresponding with ectopic microadenoma was demonstrated on the upper wall of the sphenoid sinus. The serum GH level was 4.3 mg/l, IGF-1 = 615 mg/l, and a lack of GH suppression with oral glucose was proven. After preliminary treatment with a long-acting somatostatin analogue, transsphenoidal pituitary tumour removal was performed. Histopathological, electron microscopical and immunohistochemical analysis revealed densely granulated somatotropic pituitary adenoma: GH(+), PRL(-), ACTH(-), TSH(-), FSH(-), LH(-), MIB1 < 1%, SSTR3(+) and SSTR5(+). Post-surgical evaluation showed normal pituitary MRI scans, GH and IGF-1 levels 0.18 mug/l and 140 mg/l, respectively, as well as normal GH suppression with oral glucose. The careful analysis of possible pituitary embryonic malformations points out their significance for proper localization of extrapituitary adenomas.  相似文献   

12.
Somatostatin (SST) inhibits growth hormone (GH) secretion and regulates multiple processes by signaling through its receptors sst1-5. Differential expression of SST/ssts may contribute to sex-specific GH pattern and fasting-induced GH rise. To further delineate the tissue-specific roles of SST and sst1-5 in these processes, their expression patterns were evaluated in hypothalamus, pituitary, and stomach of male and female mice under fed/fasted conditions in the presence (wild type) or absence (SST-knockout) of endogenous SST. Under fed conditions, hypothalamic/stomach SST/ssts expression did not differ between sexes, whereas male pituitary expressed more SST and sst2A/2B/3/5A/5TMD2/5TMD1 and less sst1, and male pituitary cell cultures were more responsive to SST inhibitory actions on GH release compared with females. This suggests that local pituitary SST/ssts can contribute to the sexually dimorphic pattern of GH release. Fasting (48 h) reduced stomach sst2A/B and hypothalamic SST/sst2A expression in both sexes, whereas it caused a generalized downregulation of pituitary sst subtypes in male and of sst2A only in females. Thus, fasting can reduce SST sensitivity across tissues and SST input to the pituitary, thereby jointly contributing to enhance GH release. In SST-knockout mice, lack of SST differentially altered sst subtype expression levels in both sexes, supporting an important role for SST in sex-dependent control of GH axis. Evaluation of SST, IGF-I, and glucocorticoid effects on hypothalamic and pituitary cell cultures revealed that these hormones could directly account for alterations in sst2/5 expression in the physiological states examined. Taken together, these results indicate that changes in SST output and sensitivity can contribute critically to precisely define, in a tissue-dependent manner, the sex-specific metabolic regulation of the GH axis.  相似文献   

13.
Gel filtration of female rat plasma with normal growth hormone (GH) concentrations (less than 100 ng/m1) showed that nearly all the immuno-reactivity was centred on a peak with an apparent molecular weight in the region of 82,000. In contrast, pituitary GH was almost entirely monomeric. The majority of plasma prolactin (PRL) in the same samples had a molecular weight of 23,000 (i.e. monomeric), and was similar in profile to pituitary PRL. Samples from male rats showed some GH immunoreactivity at the 82,000 molecular weight position but more than 65% coeluted with monomeric PRL. In female plasma with GH concentration between 300 and 1,000 ng/ml, immuno-reactivity resolved into peaks at the void volume, the monomeric position, and a peak at 82,000 that decreased, as a percentage of the total, with increasing GH concentration. These results indicate the possible presence of a GH binding factor, with greater activity in female than male rat plasma.  相似文献   

14.
Plasma growth hormone (GH) responses to various stimuli were examined in 21 patients with GH-producing pituitary adenomas, classified into three types by the immunohistochemistry of cytokeratin and the glycoprotein hormone alpha-subunit distribution. Seven type 1 adenomas were exclusively composed of cells in which the cytokeratin formed a dot-like pattern; they were chromophobic to hematoxylin and eosin (H&E), occasionally positive for GH, and almost completely negative for the alpha-subunit. Thirteen type 2 adenomas were composed of cells with cytokeratin that had a perinuclear distribution; they were eosinophilic to H&E, and diffusely positive for both GH and the alpha-subunit. One patient had a type 3 adenoma which had a mixed pattern of intracellular cytokeratin distribution and was chromophobic and eosinophilic to H&E. Clinically, type 1 is characterized by earlier onset, larger tumor size, and more frequent aggressive extension. Paradoxical GH responses to TRH and OGTT were seen in 1 of 6 patients (16.7%) of type 1 and 8 of 9 patients (88.9%) of type 2, and 0% of type 1 and 62.5% of type 2, respectively. Type 2 cases showed higher plasma GH response to GH-releasing hormone, and a tendency to greater suppression of plasma GH by bromocriptine compared with type 1. Octreotide acetate administration revealed that the nadir/basal ratio of plasma GH levels was 42.9 +/- 6.6% in type 1 and 13.5 +/- 5.8% in type 2. These results suggest that there is a pathophysiological difference between these two distinct types of GH-producing pituitary adenomas.  相似文献   

15.
The peptide hormone adropin, encoded by the energy homeostasis-associated (Enho) gene, plays a role in energy homeostasis and the control of vascular function. The aim of this study was to examine the role of adropin in growth hormone (GH) gene expression at the pituitary level in tilapia. As a first step, the antiserum for the tilapia adropin was produced, and its specificity was confirmed by antiserum preabsorption and immunohistochemical staining in the tilapia pituitary. Adropin could be detected immunocytochemically in the proximal pars distalis (PPD) of the tilapia pituitary. In primary cultures of tilapia pituitary cells, tilapia adropin was effective in increasing GH mRNA levels. However, removal of endogenous adropin by immunoneutralization using adropin antiserum inhibited GH gene expression. In parallel experiments, pituitary cells co-treated with ovine pituitary adenylate cyclase activating polypeptide 38 (oPACAP38) and adropin showed a similar increase level compared to those treated with oPACAP38 alone, whereas insulin-like growth factor 1 (IGF1) not only had an inhibitory effect on basal GH mRNA levels, but also could abolish adropin stimulation of GH gene expression. In pituitary cells pretreated with actinomycin D, the half-life of GH mRNA was enhanced by adropin. Taken together, these findings suggest that adropin may serve as a novel local stimulator for GH gene expression in tilapia pituitary.  相似文献   

16.
17.
18.
BACKGROUND/AIMS: The efficacy of transsphenoidal surgery in the treatment of patients with acromegaly is largely dependent on tumour size. A reduction in pituitary tumour volume by medical therapy might therefore improve subsequent surgical cure rates. This study prospectively determined the effects of the depot somatostatin analogue octreotide LAR on pituitary tumour size, GH and IGF-I levels and clinical symptoms in a cohort of previously untreated patients with acromegaly. METHODS: Six patients newly diagnosed with acromegaly (mean age 53 years; range 42-76 years) received intramuscular octreotide LAR every 28 days for 6 months. The initial dose of LAR was 20 mg, but increased to 30 mg after the initial 3 injections if mean GH levels were >5 mU/l. Prior to commencing LAR therapy, each patient received 3 injections of subcutaneous octreotide (50, 100 and 200 mug) in a randomized order on separate days, and the serum GH response was measured. Pituitary tumour volume was calculated from MRI or computed tomography scans at baseline, then 3 and 6 months after initiation of treatment, and assessed by a 'blinded' radiologist in random order. At baseline, 4 patients had a macroadenoma and 2 patients had a microadenoma. For the latter, the whole gland volume was measured. RESULTS: Serum GH levels decreased from 29.6 +/- 19.2 mU/l (mean +/- SD) at baseline to 12.1 +/- 10.5 mU/l at 3 months and 10.4 +/- 9.3 mU/l at 6 months. Three patients achieved a mean serum GH level of <5 mU/l. In these patients, the serum GH had declined to <5 mU/l in response to a single 100 mug subcutaneous octreotide injection. Serum IGF-I levels decreased by a mean of 45 +/- 7.4%. Tumour volume decreased in all patients: mean baseline volume 2,175 mm(3) (range 660-6,998) decreasing to 1,567 mm(3) (range 360-4,522) at 3 months (p < 0.05) and 1,293 mm(3) (range 280-4,104) at 6 months (p < 0.002). The mean percentage decrease in size was 29% (range -54 to +4%) at 3 months (p < 0.02) and 47% (range 21-97%) at 6 months (p < 0.002). There was no statistically significant correlation between GH response and tumour shrinkage. CONCLUSIONS: A single test dose of subcutaneous octreotide may be useful in predicting the subsequent efficacy of octreotide LAR. Octreotide LAR results in significant shrinkage of pituitary tumours of newly diagnosed patients with acromegaly. Whether its administration to such patients for 6-12 months can improve the efficacy of subsequent transsphenoidal surgery will require further study.  相似文献   

19.
The aim of this work was to study the effect of cyclosporine on the somatotropic axis. Accordingly, growth hormone (GH) secretion, circulating insulin-like growth factor I (IGF-I) and IGF binding proteins (IGFBPs) in response to cyclosporin A (CsA) treatment were examined in adult male Wistar rats. Cyclosporine administration (5, 10 or 20 mg/Kg daily) over 8 days did not modify the body weight, but it did decrease serum concentration of corticosterone and increased serum IGF-I and GH levels. Rats treated with 5 and 10 mg/Kg of cyclosporine had similar levels of serum IGFBPs to control rats, but there was an increase in circulating IGFBP-3 and IGFPB-1,2 in the group treated with 20 mg/Kg of CsA. The increase in circulating GH correlates with a decrease in pituitary GH content in CsA treated rats, with no modification in hypothalamic somatostatin content, suggesting an increase in pituitary GH release. In order to test this hypothesis, anterior pituitary cell cultures were exposed to different CsA concentrations during a 4 h incubation period. Cyclosporine increased GH secretion in cultured pituitary cells (p<0.05). These data suggest that cyclosporine increases circulating IGF-I and GH by stimulating pituitary GH release.  相似文献   

20.
Plasma growth hormone (GH) responses to the repetitive administrations of synthetic human pancreatic growth hormone releasing factor (hpGRF-44) were studied in 15 patients with GH deficiency (11 diagnosed as idiopathic and 4 diagnosed as secondary to hypothalamo-pituitary tumor). hpGRF-44 was administered by single iv bolus (2 micrograms/kg), repetitive im (100 micrograms, twice a day), and/or repetitive iv infusion (2.5 micrograms/min for 90 min, once a day) for three to six consecutive days. Three of the eleven idiopathic GH deficient patients had plasma GH responses to both single iv bolus injection and repetitive administrations by im, or iv infusion of hpGRF. In four of the remaining eight, who had not had peak plasma GH levels above 5 ng/ml to a single iv bolus of the peptide, repetitive administrations of hpGRF-44 by im injection and/or iv infusion induced GH responses to the peptide. In the four patients with secondary GH deficiency, three had plasma GH response to hpGRF administration but one patient, who had indications of pituitary disorder, did not show any plasma GH response to either single iv injection or repetitive administrations of hpGRF-44. These data show that repetitive administrations of hpGRF-44 can induce plasma GH responses in some GH deficient patients who do not respond to a single iv bolus of the peptide.  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