Citations
All
Search in:AllTitleAbstractAuthor name
Publications
(2K+)
Patents
Grants
Pathways
Clinical trials
Publication
Journal: Endocrinology
April/23/1986
Abstract
Highly purified functional cytotrophoblasts have been prepared from human term placentae by adding a Percoll gradient centrifugation step to a standard trypsin-DNase dispersion method. The isolated mononuclear trophoblasts averaged 10 microns in diameter, with occasional cells measuring up to 20-30 microns. Viability was greater than 90%. Transmission electron microscopy revealed that the cells had fine structural features typical of trophoblasts. In contrast to syncytial trophoblasts of intact term placentae, these cells did not stain for hCG, human placental lactogen, pregnancy-specific beta 1-glycoprotein or low mol wt cytokeratins by immunoperoxidase methods. Endothelial cells, fibroblasts, or macrophages did not contaminate the purified cytotrophoblasts, as evidenced by the lack of immunoperoxidase staining with antibodies against vimentin or alpha 1-antichymotrypsin. The cells produced progesterone (1 ng/10(6) cells . <em>4</em> h), and progesterone synthesis was stimulated up to 8-fold in the presence of 25-hydroxycholesterol (20 micrograms/ml). They also produced estrogens (1360 pg/10(6) cells . <em>4</em> h) when supplied with <em>androstenedione</em> (1 ng/ml) as a precursor. When placed in culture, the cytotrophoblasts consistently formed aggregates, which subsequently transformed into syncytia within 2<em>4</em>-<em>4</em>8 h after plating. Time lapse cinematography revealed that this process occurred by cell fusion. The presumptive syncytial groups were proven to be true syncytia by microinjection of fluorescently labeled alpha-actinin, which diffused completely throughout the syncytial cytoplasm within 30 min. Immunoperoxidase staining of cultured trophoblasts between 3.5 and 72 h after plating revealed a progressive increase in cytoplasmic pregnancy-specific beta 1-glycoprotein, hCG, and human placental lactogen concomitant with increasing numbers of aggregates and syncytia. At all time points examined, occasional single cells positive for these markers were identified. RIA of the spent culture media for hCG revealed a significant increase in secreted hCG, paralleling the increase in hCG-positive cells and syncytia identified by immunoperoxidase methods. We conclude that human cytotrophoblasts differentiate in culture and fuse to form functional syncytiotrophoblasts.
Pulse
Views:
1
Posts:
No posts
Rating:
Not rated
Publication
Journal: Journal of the American Academy of Dermatology
September/19/2001
Abstract
Estradiol production is most commonly thought of as an endocrine product of the ovary; however, there are many tissues that have the capacity to synthesize estrogens from androgen and to use estrogen in a paracrine or intracrine fashion. In addition, other organs such as the adipose tissue can contribute significantly to the circulating pool of estrogens. There is increasing evidence that in both men and women extraglandular production of C(18) steroids from C(19) precursors is important in normal physiology as well as in pathophysiologic states. The enzyme aromatase is found in a number of human tissues and cells, including ovarian granulosa cells, the placental syncytiotrophoblast, adipose and skin fibroblasts, bone, and the brain, and it locally catalyzes the conversion of C(19) steroids to estrogens. Aromatase expression in adipose tissue and possibly the skin primarily accounts for the extraglandular (peripheral) formation of estrogen and increases as a function of body weight and advancing age. Sufficient circulating levels of the biologically active estrogen estradiol can be produced as a result of extraglandular aromatization of <em>androstenedione</em> to estrone that is subsequently reduced to estradiol in peripheral tissues to cause uterine bleeding and endometrial hyperplasia and cancer in obese anovulatory or postmenopausal women. Extraglandular aromatase expression in adipose tissue and skin (via increasing circulating levels of estradiol) and bone (via increasing local estrogen concentrations) is of paramount importance in slowing the rate of postmenopausal bone loss. Moreover, excessive or inappropriate aromatase expression was demonstrated in adipose fibroblasts surrounding a breast carcinoma, endometriosis-derived stromal cells, and stromal cells in endometrial cancer, giving rise to increased local estrogen concentrations in these tissues. Whether systemically delivered or locally produced, elevated estrogen levels will promote the growth of these steroid-responsive tissues. Finally, local estrogen biosynthesis by aromatase activity in the brain may be important in the regulation of various cognitive and hypothalamic functions. The regulation of aromatase expression in human cells via alternatively used promoters, which can be activated or inhibited by various hormones, increases the complexity of estrogen biosynthesis in the human body. Aromatase expression is under the control of the classically located proximal promoter II in the ovary and a far distal promoter I.1 (<em>4</em>0 kilobases upstream of the translation initiation site) in the placenta. In skin, the promoter is I.<em>4</em>. In adipose tissue, 2 other promoters (I.<em>4</em> and I.3) located between I.1 and II are used in addition to the ovarian-type promoter II. In addition, promoter use in adipose fibroblasts switches between promoters II/I.3 and I.<em>4</em> upon treatments of these cells with PGE(2) versus glucocorticoids plus cytokines. Moreover, the presence of a carcinoma in breast adipose tissue also causes a switch of promoter use from I.<em>4</em> to II/I.3. Thus there can be complex mechanisms that regulate the extraglandular production of estrogen in a tissue-specific and state-specific fashion.
Publication
Journal: Journal of Clinical Endocrinology and Metabolism
February/13/1991
Abstract
To determine whether hyperinsulinemia can directly reduce serum sex hormone-binding globulin (SHBG) levels in obese women with the polycystic ovary syndrome, six obese women with this disorder were studied. Before study, ovarian steroid production was suppressed in each woman by the administration of 7.5 mg of a long-acting GnRH agonist, leuprolide depot, im, on days -56, -28, and 0. This resulted in substantial reductions in serum concentrations of testosterone (from 1.72 +/- 0.29 nmol/L on day -56 to 0.32 +/- 0.09 nmol/L on day 0), non-SHBG-bound testosterone (from 10<em>4</em> +/- 16 pmol/L on day -56 to 19 +/- 5 pmol/L on day 0), <em>androstenedione</em> (from 7.25 +/- 1.65 nmol/L on day -56 to 2.78 +/- 0.9<em>4</em> nmol/L on day 0), estrone (from 371 +/- 71 pmol/L on day -56 to 156 +/- 29 pmol/L on day 0), estradiol (from 235 +/- 26 pmol/L on day -56 to 90 +/- 2<em>4</em> pmol/L on day 0), and progesterone (from 0.28 +/- 0.12 nmol/L on day -56 to 0.08 +/- 0.02 nmol/L on day 0). Serum SHBG levels, however, did not change (18.8 +/- 2.8 nmol/L on day -56 vs. 17.8 +/- 2.6 nmol/L on day 0). While continuing leuprolide treatment, the women were administered oral diazoxide (300 mg/day) for 10 days to suppress serum insulin levels. Diazoxide treatment resulted in suppressed insulin release during a 100-g oral glucose tolerance test (insulin area under the curve, 262 +/- 55 nmol/min.L on day 0 vs. 102 +/- 33 nmol/min.L on day 10; P less than 0.05) and deterioration of glucose tolerance. Serum testosterone, <em>androstenedione</em>, estrone, estradiol, and progesterone levels did not change during combined diazoxide and leuprolide treatment. In contrast, serum SHBG levels rose by 32% from 17.8 +/- 2.6 nmol/L on day 0 to 23.5 +/- 2.0 nmol/L on day 10 (P less than 0.003). Due primarily to the rise in serum SHBG levels, serum non-SHBG-bound testosterone levels fell by <em>4</em>3% from 19 +/- 5 pmol/L on day 0 to 11 +/- <em>4</em> pmol/L on day 10 (P = 0.05). These observations suggest that hyperinsulinemia directly reduces serum SHBG levels in obese women with the polycystic ovary syndrome independently of any effect on serum sex steroids.
Publication
Journal: Journal of Clinical Endocrinology and Metabolism
August/6/1997
Abstract
Polycystic ovary syndrome (PCOS) is a heterogeneous disorder of reproductive age women characterized in its broadest definition by the presence of oligoamenorrhea and hyperandrogenism and the absence of other disorders. Defects of gonadotropin secretion, including an elevated LH level, elevated LH to FSH ratio, and an increased frequency and amplitude of LH pulsations have been described, but the prevalence of these defects in a large, unbiased population of PCOS patients has not been determined. Sixty-one women with PCOS defined by oligomenorrhea and hyperandrogenism and 2<em>4</em> normal women in the early follicular phase had LH samples obtained every 10 min for 8-12 h. Pool LH levels from the frequent sampling studies were within the normal range in the 9 PCOS patients (1<em>4</em>.8%) who were studied within 21 days after a documented spontaneous ovulation. Excluding these post-ovulatory patients, 75.0% of the PCOS patients had an elevated pool LH level (above the 95th percentile of the normal controls), and 9<em>4</em>% had an elevated LH to FSH ratio. In the anovulatory PCOS patients, pool LH correlated positively with 17-OH progesterone (R = 0.30, P = 0.03), but not with estradiol, estrone, testosterone, <em>androstenedione</em>, or DHEA-S. Pool LH and LH to FSH ratio correlated positively with LH pulse frequency (R = 0.<em>4</em>0, P = 0.00<em>4</em> for pool LH, and R = 0.39; P = 0.005 for LH/FSH). There was also a strong negative correlation between pool LH and body mass index (BMI) (R = -0.59, P < 10(-5)). The relationship between BMI and LH secretion in the PCOS patients appeared to be strongest with body fatness, as pool LH was correlated inversely with percent body fat, whether measured by skinfolds (R = -0.61, P < 10(-5)), bioimpedance (R = -0.55, P < 10(-<em>4</em>)), or dual energy x-ray absorptiometry (DEXA) (R = -0.70, P = 0.001; n = 18 for DEXA only). By DEXA, the only body region that was highly correlated with pool LH was the trunk (R = -0.71, P = 0.001). The relationship between body fatness and LH secretion occurred via a decrease in LH pulse amplitude (R = -0.63, P < 10(-5) for BMI; R = -0.58, P < 10(-<em>4</em>) for bioimpedance; and R = -0.6<em>4</em>, P = 0.00<em>4</em> for whole body DEXA), with no significant change in pulse frequency with increasing obesity (R = -0.17, P = 0.23 for BMI).
CONCLUSIONS
1) the prevalence of gonadotropin abnormalities is very high in women with PCOS selected on purely clinical grounds, but is modified by recent spontaneous ovulation; 2) the positive relationship between LH pulse frequency and both pool LH and LH to FSH ratio supports the hypothesis that a rapid frequency of GnRH secretion may play a key etiologic role in the gonadotropin defect in PCOS patients; 3) pool LH and LH pulse amplitude are inversely related to body mass index and percent body fat in a continuous fashion; and <em>4</em>) the occurrence of a continuous spectrum of gonadotropin abnormalities varying with body fat suggests that nonobese and obese patients with PCOS do not represent distinct pathophysiologic subsets of this disorder.
Publication
Journal: Journal of Clinical Endocrinology and Metabolism
February/2/2000
Abstract
In the last few years some studies assessed the effects of attenuation of hyperinsulinemia and insulin resistance, obtained by insulin sensitizing agents, in women with polycystic ovary syndrome (PCOS), suggesting potential scope for these drugs in treating the whole spectrum of reproductive, endocrine, and metabolic abnormalities found in such subjects. However, the results of these studies, mostly uncontrolled and short-term, are still inconclusive, and there is no long-term follow-up. In the present study, 23 PCOS subjects [mean (+/- SE) body mass index 30.0+/-1.1 kg/m2] were randomly assigned to double-blind treatment with metformin (500 mg tid) or placebo for 6 months, while maintaining their usual eating habits. Before and after treatment, menstrual history, endocrine and metabolic profiles, serum 17-hydroxyprogesterone response to GnRH-agonist testing, and insulin sensitivity measured by the glucose clamp technique were assessed. Eighteen of these women, as well as 1<em>4</em> additional PCOS patients, were subsequently given metformin in an open trial for 11.0+/-1.3 months (range <em>4</em>-26), to assess long-term effects of treatment and baseline predictors of metformin efficacy on reproductive abnormalities. After metformin treatment, mean frequency of menstruation improved (P = 0.002), due to striking amelioration of menstrual abnormalities in about 50% of subjects. Women given metformin showed reduced plasma insulin (at fasting: P = 0.057; during the clamp studies: P<0.01) and increased insulin sensitivity (P<0.05). Concurrently, ovarian hyperandrogenism was attenuated, as indicated by significant reductions in serum free testosterone (P<0.05) and in the 17-hydroxyprogesterone response to GnRH-agonist testing (P<0.05). No changes were found in the placebo group. Only comparable minor changes in body mass index were found both in the metformin group and in the placebo group. In the open, long-term trial 17 women (5<em>4</em>.8%) showed striking improvements of their menstrual abnormalities and were considered as responders. Logistic regression analysis of baseline characteristics in responders and nonresponders showed that plasma insulin, serum <em>androstenedione</em>, and menstrual history were independent predictors of the treatment's clinical efficacy. In 10 subjects whose menses proved regular after treatment, the great majority of cycles became ovulatory (32 out of 39 assessed, 79%). In conclusion, in women with PCOS metformin treatment reduced hyperinsulinemia and hyperandrogenemia, independently of changes in body weight. In a large number of subjects these changes were associated with striking, sustained improvements in menstrual abnormalities and resumption of ovulation. Higher plasma insulin, lower serum <em>androstenedione</em>, and less severe menstrual abnormalities are baseline predictors of clinical response to metformin.
Publication
Journal: Steroids
April/16/2007
Abstract
This article describes the origins and evolution of "antiestrogenic" medicines for the treatment and prevention of breast cancer. Developing drugs that target the estrogen receptor (ER) either directly (tamoxifen) or indirectly (aromatase inhibitors) has improved the prognosis of breast cancer and significantly advanced healthcare. The development of the principles for treatment and the success of the concept, in practice, has become a model for molecular medicine and presaged the current testing of numerous targeted therapies for all forms of cancer. The translational research with tamoxifen to target the ER with the appropriate duration (5 years) of adjuvant therapy has contributed to the falling national death rates from breast cancer. Additionally, exploration of the endocrine pharmacology of tamoxifen and related nonsteroidal antiestrogen (e.g. keoxifene now known as raloxifene) resulted in the laboratory recognition of selective ER modulation and the translation of the concept to use raloxifene for the prevention of osteoporosis and breast cancer. However, the extensive evaluation of tamoxifen treatment revealed small but significant side effects such as endometrial cancer, blood clots and the development of acquired resistance. The solution was to develop drugs that targeted the aromatase enzyme specifically to prevent the conversion of <em>androstenedione</em> to estrone and subsequently estradiol. The successful translational research with the suicide inhibitor <em>4</em>-hydroxy<em>androstenedione</em> (known as formestane) pioneered the development of a range of oral aromatase inhibitors that are either suicide inhibitors (exemestane) or competitive inhibitors (letrozole and anastrozole) of the aromatase enzyme. Treatment with aromatase inhibitors is proving effective and is associated with reduction in the incidence of endometrial cancer and blood clots when compared with tamoxifen and there is also limited cross resistance so treatment can be sequential. Current clinical trials are addressing the value of aromatase inhibitors as chemopreventive agents for postmenopausal women.
Publication
Journal: Journal of Clinical Endocrinology and Metabolism
December/13/2000
Abstract
Polycystic ovarian syndrome (PCOS) is a complex disorder with multiple abnormalities, including hyperandrogenism, ovulatory dysfunction, and altered gonadotropin secretion. The majority of patients have elevated LH levels in plasma and a persistent rapid frequency of LH (GnRH) pulse secretion, the mechanisms of which are unclear. Earlier work has suggested that the sensitivity of the GnRH pulse generator to inhibition by ovarian steroids is impaired. We performed a study to determine whether antiandrogen therapy with flutamide could enhance feedback inhibition by estradiol (E2) and progesterone (P) in women with PCOS. Ten anovulatory women with PCOS and nine normal controls (days 8-10 of the cycle) were studied on three occasions. During each admission, LH and FSH were determined every 10 min and E2, P, and testosterone (T) every 2 h for 13 h. After 12 h, GnRH (25 ng/kg) was given iv. After the first admission, patients were started on flutamide (250 mg twice daily), which was continued for the entire study. The second admission occurred on days 8-10 of the next menstrual cycle for normal controls and on study day 28 for PCOS patients. Subjects were then given E2 transdermally (mean plasma E2, 106+/-18 pg/mL) and P by vaginal suppository to obtain varied plasma concentrations of P (mean P, <em>4</em>.<em>4</em>+/-0.5 ng/mL; range, 0.6-9.0 ng/mL), and a third study was performed 7 days later. At baseline women with PCOS had higher LH pulse amplitude, response to GnRH, T, <em>androstenedione</em>, and insulin and lower sex hormone-binding globulin concentrations (P < 0.05). Most hormonal parameters were not altered by <em>4</em> weeks of flutamide, except T in controls and E2 and FSH in PCOS patients, which were lower. Of note, flutamide alone had no effect on LH pulse frequency or amplitude, mean plasma LH, or LH responsiveness to exogenous GnRH. After the addition of E2 and P for 7 days, both PCOS patients and normal controls had similar reductions in LH pulse frequency (<em>4</em>.0+/-0.7 and 5.8+/-0.7 pulses/12 h, respectively). This contrasts with our earlier results in the absence of flutamide, where a plasma P level of less than 10 ng/mL had minimal effects on LH pulse frequency in women with PCOS, but was effective in controls. These results suggest that although the elevated LH pulse frequency in PCOS may in part reflect impaired sensitivity to E2 and P, continuing actions of hyperandrogenemia are important for sustaining the abnormal hypothalamic sensitivity to feedback inhibition by ovarian steroids.
Publication
Journal: Journal of Clinical Endocrinology and Metabolism
August/25/1983
Abstract
Hormonal reference data, in the form of nomograms relating baseline and stimulated levels of adrenal hormones, provide a means of genotyping steroid 21-hydroxylase (21-OH) deficiency in congenital adrenal hyperplasia. Data from both 360- and 60-min ACTH stimulation tests are given. The serum hormone concentrations that have proven most useful in classifying 21-OH deficiency are 17-hydroxyprogesterone and delta <em>4</em>-<em>androstenedione</em>. These nomograms clearly distinguish the patient with classical 21-OH deficiency from those with the milder symptomatic and asymptomatic nonclassical forms of 21-OH deficiency (previously referred to as late onset and cryptic forms) as well as heterozygotes for all of the forms and those subjects predicted by HLA genotyping to be unaffected. The nomograms also can identify individuals heterozygous for 21-OH deficiency in the general population who have a characteristic heterozygote response. These nomograms provide a powerful tool by which to assign the 21-OH deficiency genotype. Patients whose hormonal values fall on the regression line within a defined group are assigned to that group. In view of the strong correlation between the 60- and 360-min ACTH stimulation tests, the less cumbersome and shorter 60-min test can be used with the same confidence as the longer test.
Publication
Journal: Steroids
May/28/1997
Abstract
17 beta-Hydroxysteroid dehydrogenase (17 beta-HSD) controls the last step in the formation of all androgens and all estrogens. This crucial role of 17 beta-HSD is performed by at least five 17 beta-HSD isoenzymes having individual cell-specific expression, substrate specificity, regulation mechanisms, and reductive or oxidative catalytic activity. Both estrogenic and androgenic 17 beta-HSD activities were found in all 25 rhesus monkey and 15 human peripheral intracrine tissues examined. Type 1 17 beta-HSD is a protein of 327 amino acids catalyzing the formation of 17 beta-estradiol from estrone. Its x-ray structure was the first to be determined among mammalian steroidogenic enzymes. Initially crystallized with NAD, the crystal structure of type 1 17 beta-HSD has just been determined as a complex with 17 beta-estradiol, thereby illustrating the conformation of the substrate-binding site. Type 2 17 beta-HSD degrades 17 beta-estradiol into estrone and testosterone into <em>androstenedione</em>, and type <em>4</em> 17 beta-HSD mainly degrades 17 beta-estradiol into estrone and androst-5-ene-3 beta, 17 beta-diol into dehydroepiandrosterone. Types 3 and 5 17 beta-HSD, on the other hand, catalyze the formation of testosterone from <em>androstenedione</em> in the testis and peripheral tissues, respectively. The various types of human 17 beta-HSD, because of their tissue-specific expression and substrate specificity, provide each peripheral cell with the necessary mechanisms to control the level of intracellular androgens and/or estrogens, a new area of hormonal control that we call intracrinology.
Publication
Journal: Cancer Research
January/24/2006
Abstract
Aromatase [cytochrome P<em>4</em>50 19 (CYP19)] is a critical enzyme for estrogen biosynthesis, and aromatase inhibitors are of increasing importance in the treatment of breast cancer. We set out to identify and characterize genetic polymorphisms in the aromatase gene, CYP19, as a step toward pharmacogenomic studies of aromatase inhibitors. Specifically, we "resequenced" all coding exons, all upstream untranslated exons plus their presumed core promoter regions, all exon-intron splice junctions, and a portion of the 3'-untranslated region of CYP19 using 2<em>4</em>0 DNA samples from four ethnic groups. Eighty-eight polymorphisms were identified, resulting in <em>4</em><em>4</em> haplotypes. Functional genomic studies were done with the four nonsynonymous coding single nucleotide polymorphisms (cSNP) that we observed, two of which were novel. Those cSNPs altered the following amino acids: Trp39Arg, Thr201Met, Arg26<em>4</em>Cys, and Met36<em>4</em>Thr. The Cys26<em>4</em>, Thr36<em>4</em>, and double variant Arg39Cys26<em>4</em> allozymes showed significant decreases in levels of activity and immunoreactive protein when compared with the wild-type (WT) enzyme after transient expression in COS-1 cells. A slight decrease in protein level was also observed for the Arg39 allozyme, whereas Met201 displayed no significant changes in either activity or protein level when compared with the WT enzyme. There was also a <em>4</em>-fold increase in apparent K(m) value for Thr36<em>4</em> with <em>androstenedione</em> as substrate. Of the recombinant allozymes, only the double mutant (Arg39Cys26<em>4</em>) displayed a significant change from the WT enzyme in inhibitor constant for the aromatase inhibitors exemestane and letrozole. These observations indicate that genetic variation in CYP19 might contribute to variation in the pathophysiology of estrogen-dependent disease.
Publication
Journal: Frontiers in Neuroendocrinology
August/29/2001
Abstract
A new understanding of the endocrinology of menopause is that women, at menopause, are not only lacking estrogens resulting from cessation of ovarian activity but have also been progressively deprived for a few years of androgens and some estrogens originating from adrenal DHEA and <em>androstenedione</em> (<em>4</em>-dione). In fact, serum DHEA decreases by about 60% between the maximal levels seen at 30 years of age to the age of menopause. This decreased secretion of DHEA and DHEA-S by the adrenals is responsible for a parallel decrease in androgen and estrogen formation in peripheral tissues by the steroidogenic enzymes specifically expressed in each cell type in individual target tissues. This new field of endocrinology, called intracrinology, describes the local synthesis of androgens and estrogens made locally in each cell of each peripheral tissue from the adrenal precursors DHEA and <em>4</em>-dione. These androgens and estrogens exert their action in the same cells where their synthesis takes place and they are released from these target cells only after being inactivated. To further understand the effect of DHEA in women, DHEA has been administered in postmenopausal women for 12 months. Such treatment resulted in increased bone formation and higher bone mineral density accompanied by elevated levels of osteocalcin, a marker of bone formation. Vaginal maturation was stimulated, while no effect was observed on the endometrium. Preclinical studies, on the other hand, have shown that, due to its predominant conversion into androgens, DHEA prevents the development and inhibits the growth of dimethylbenz(a)anthracene-induced mammary carcinoma in the rat, a model of breast cancer. DHEA also inhibits the growth of human breast cancer ZR-75-1 xenografts in nude mice. The inhibitory effect of DHEA on breast cancer is due to an androgenic effect of testosterone and dihydrotestosterone made locally from DHEA. When used as replacement therapy, DHEA is free of the potential risk of breast and uterine cancer, while it stimulates bone formation and vaginal maturation and decreases insulin resistance. The combination of DHEA with a fourth generation SERM, such as EM-652 (SCH 57068), a compound having pure and potent antiestrogenic activity in the mammary gland and endometrium, could provide major benefits for women at menopause (inhibition of bone loss and serum cholesterol levels) with the associated major advantages of preventing breast and uterine cancer. A widely used application of intracrinology is the treatment of prostate cancer where the testicles are blocked by an LHRH agonist while the androgens made locally in the prostate from DHEA are blocked by a pure antiandrogen. Such treatment, called combined androgen blockade, has led to the first demonstration of a prolongation of life in prostate cancer.
Publication
Journal: European Urology
July/27/2015
Abstract
BACKGROUND
Enzalutamide is a novel antiandrogen with proven efficacy in metastatic castration-resistant prostate cancer (mCRPC).
OBJECTIVE
To evaluate enzalutamide's effects on cancer and on androgens in blood and bone marrow, and associate these with clinical observations.
METHODS
In this prospective phase 2 study, 60 patients with bone mCRPC received enzalutamide 160mg orally daily and had transilial bone marrow biopsies before treatment and at 8 wk of treatment.
METHODS
Androgen signaling components (androgen receptor [AR], AR splice variant 7 (ARV7), v-ets avian erythroblastosis virus E26 oncogene homolog [ERG], cytochrome P450, family 17, subfamily A, polypeptide 1 [CYP17]) and molecules implicated in mCRPC progression (phospho-Met, phospho-Src, glucocorticoid receptor, Ki67) were assessed by immunohistochemistry; testosterone, cortisol, and androstenedione concentrations were assessed by liquid chromatography-tandem mass spectrometry; AR copy number was assessed by real-time polymerase chain reaction. Descriptive statistics were applied.
CONCLUSIONS
Median time to treatment discontinuation was 22 wk (95% confidence interval, 19.9-29.6). Twenty-two (37%) patients exhibited primary resistance to enzalutamide, discontinuing treatment within 4 mo. Maximal prostate-specific antigen (PSA) decline ≥ 50% and ≥ 90% occurred in 27 (45%) and 13 (22%) patients, respectively. Following 8 wk of treatment, bone marrow and circulating testosterone levels increased. Pretreatment tumor nuclear AR overexpression >> 75%) and CYP17 >> 10%) expression were associated with benefit (p = 0.018). AR subcellular localization shift from the nucleus was confirmed in eight paired samples (with PSA decline) of 23 evaluable paired samples. Presence of an ARV7 variant was associated with primary resistance to enzalutamide (p = 0.018). Limited patient numbers warrant further validation.
CONCLUSIONS
The observed subcellular shift of AR from the nucleus and increased testosterone concentration provide the first evidence in humans that enzalutamide suppresses AR signaling while inducing an adaptive feedback. Persistent androgen signaling in mCRPC was predictive of benefit and ARV7 was associated with primary resistance.
RESULTS
We report a first bone biopsy study in metastatic prostate cancer in humans that searched for predictors of outcome of enzalutamide therapy. Benefit is linked to a pretreatment androgen-signaling signature.
BACKGROUND
ClinicalTrials.gov identifier NCT01091103.
Publication
Journal: Journal of Clinical Oncology
June/4/1989
Abstract
Thirty-seven men with symptomatic bone metastases from prostate cancer that had progressed following earlier treatment with estrogens and/or orchidectomy were treated with low-dose prednisone (7.5 to 10 mg daily). The rationale for this treatment was that some patients might still have hormone-sensitive disease that was stimulated by weak androgens of adrenal origin, and that these androgens could be suppressed by prednisone through its negative feedback on secretion of adrenocorticotrophic hormone (ACTH). Response to treatment was assessed by requirement for analgesics, by the McGill-Melzack pain questionnaire, and by a series of 17 linear analog self-assessment (LASA) scales relating to pain and to various aspects of quality of life. Fourteen patients (38%) had improvement in indices used to assess pain at 1 month after starting prednisone, and seven patients (19%) maintained this improvement for 3 to 30 months (median, <em>4</em> months). Reduction in pain was associated with improvement in other dimensions of quality of life, and in the scale for overall well-being. Prednisone treatment led to a decrease in the concentration of serum testosterone in seven of nine patients where it was not initially suppressed below 2 nmol/L, and caused a decrease in serum levels of <em>androstenedione</em> and dehydroepiandrosterone sulfate in more than 50% of patients. Symptomatic response was associated with a decrease in serum concentration of adrenal androgens. We conclude that (1) low-dose prednisone may cause useful relief of pain in some patients with advanced prostatic cancer; (2) relief of pain was associated with suppression of adrenal androgens; and (3) measures of pain and quality of life can be used to assess possible benefits of systemic therapy in patients with metastatic prostate cancer.
Publication
Journal: Molecular Pharmacology
October/20/1986
Abstract
The anti-arrhythmic quinidine has been reported to be a competitive inhibitor of the catalytic activities of human liver P-<em>4</em>50DB, including sparteine delta 2-oxidation and bufuralol 1'-hydroxylation, and we confirmed the observation that submicromolar concentrations are strongly inhibitory. Human liver microsomes oxidize quinidine to the 3-hydroxy (Km <em>4</em> microM) and N-oxide (Km 33 microM) products, consonant with in vivo observations. Both bufuralol and sparteine inhibited microsomal quinidine 3-hydroxylation. Liver microsomes prepared from DA strain rats showed a relative deficiency in quinidine 3-hydroxylase activity in females compared to males. These observations might suggest that quinidine oxidation is catalyzed by the same P-<em>4</em>50 forms that oxidize debrisoquine, bufuralol, and sparteine; i.e., rat P-<em>4</em>50UT-H and P-<em>4</em>50DB. However, neither of these two purified enzymes catalyzed quinidine 3-hydroxylation, and anti-P-<em>4</em>50UT-H, which strongly inhibits human liver microsomal bufuralol 1'-hydroxylation, did not substantially inhibit quinidine 3-hydroxylation or N-oxygenation. P-<em>4</em>50MP, the human S-mephenytoin <em>4</em>-hydroxylase, also does not appear to oxidize quinidine but P-<em>4</em>50NF, the human nifedipine oxidase, does. Anti-P-<em>4</em>50NF inhibited greater than 95% of the 3-hydroxylation and greater than 85% of the N-oxygenation of quinidine in several microsomal samples. Quinidine inhibited microsomal nifedipine oxidation and, in a series of human liver samples, rates of nifedipine oxidation were correlated with rates of quinidine oxidation. Thus, quinidine oxidation appears to be catalyzed primarily by P-<em>4</em>50NF and not by P-<em>4</em>50DB. Quinidine binds 2 orders of magnitude more tightly to P-<em>4</em>50DB, which does not oxidize it, than to P-<em>4</em>50NF, the major enzyme involved in its oxidation. The substrate specificity of human P-<em>4</em>50NF is discussed further in terms of its regioselective oxidations of complex molecules including quinidine, aldrin, benzphetamine, cortisol, testosterone and <em>androstenedione</em>, estradiol, and several 2,6-dimethyl-1,<em>4</em>-dihydropyridines.
Publication
Journal: Biology of Reproduction
January/4/2005
Abstract
The follicle-depleted postmenopausal ovary is enriched in interstitial cells that produce androgens. This study was designed to cause follicle depletion in mice using the industrial chemical, <em>4</em>-vinylcyclohexene diepoxide (VCD), and characterize the steroidogenic capacity of cells in the residual ovarian tissue. From a dose-finding study, the optimal daily concentration of VCD was determined to be 160 mg/kg. Female B6C3F(1) immature mice were treated daily with vehicle control or VCD (160 mg kg(-1) day(-1), 15 days, i.p.). Ovaries were removed and processed for histological evaluation. On Day 15 following onset of treatment, primordial follicles were depleted and primary follicles were reduced to about 10% of controls. On Day <em>4</em>6, primary follicles were depleted and secondary and antral follicles were reduced to 0.7% and 2.6% of control, respectively. Seventy-five percent of treated mice displayed disruptions in estrous cyclicity. All treated mice were in persistent diestrus (acyclic) by Day 58. Plasma FSH levels were increased (P < 0.05) relative to controls on Day 37 and had plateaued by Day 100. Relative to age-matched cyclic controls, by Day 127, the significant differences in VCD-treated mice included reduced ovarian and uterine weights, elevated plasma LH and FSH, and reduced plasma progesterone and <em>androstenedione</em>. Furthermore, plasma 17beta-estradiol levels were nondetectable. Unlike controls, immunostaining for LH receptor, and the high density lipoprotein receptor (SR-BI), was diffuse in ovarian sections from VCD-treated animals. Ovaries from Day 120 control and VCD-treated animals were dissociated and dispersed cells were placed in culture. Cultured cells from ovaries of VCD-treated animals produced less LH-stimulated progesterone than control cells. <em>Androstenedione</em> production was nondetectable in cells from cyclic control animals. Conversely, cells from VCD-treated animals produced <em>androstenedione</em> that was doubled in the presence of insulin and LH (1 and 3 ng/ml). Collectively, these data demonstrate that VCD-mediated follicle depletion results in residual ovarian tissue that may be analogous to the follicle-deplete postmenopausal ovary. This may serve as a useful animal model to examine the dynamics of follicle loss in women as ovarian senescence ensues.
Publication
Journal: Journal of Clinical Endocrinology and Metabolism
January/10/1999
Abstract
Aberrant aromatase expression in stromal cells of endometriosis gives rise to conversion of circulating <em>androstenedione</em> to estrone in this tissue, whereas aromatase expression is absent in the eutopic endometrium. In this study, we initially demonstrated by Northern blotting transcripts of the reductive 17beta-hydroxysteroid dehydrogenase (17betaHSD) type 1, which catalyzes the conversion of estrone to 17beta-estradiol, in both eutopic endometrium and endometriosis. Thus, it follows that the product of the aromatase reaction, namely estrone, that is weakly estrogenic can be converted to the potent estrogen, 17beta-estradiol, in endometriotic tissues. It was previously demonstrated that progesterone stimulates the inactivation of 17beta-estradiol through conversion to estrone in eutopic endometrial epithelial cells. Subsequently, 17betaHSD type 2 was shown to catalyze this reaction, and its transcripts were detected in the epithelial cell component of the eutopic endometrium in secretory phase. Because 17beta-estradiol plays a critical role in the development and growth of endometriosis, we studied 17betaHSD-2 expression in endometriotic tissues and eutopic endometrium. We demonstrated, by Northern blotting, 17betaHSD-2 messenger ribonucleic acid (RNA) in all RNA samples of secretory eutopic endometrium (n=12) but not in secretory samples of endometriotic lesions (n=10), including paired samples of endometrium and endometriosis obtained simultaneously from eight patients. This messenger RNA was not detectable in any samples of proliferative eutopic endometrium or endometriosis (n=<em>4</em>) as expected. Next, we confirmed these findings by demonstration of immunoreactive 17betaHSD-2 in epithelial cells of secretory eutopic endometrium in 11 of 13 samples employing a monoclonal antibody against 17betaHSD-2, whereas 17betaHSD-2 was absent in paired secretory endometriotic tissues (n=<em>4</em>). Proliferative eutopic endometrial (n=8) and endometriotic (n=<em>4</em>) tissues were both negative for immunoreactive 17betaHSD-2, except for barely detectable levels in 1 eutopic endometrial sample. Finally, we sought to determine whether deficient 17betaHSD-2 expression in endometriotic tissues is due to impaired progesterone action in endometriosis. We determined by immunohistochemistry the expression of progesterone and estrogen receptors in these paired samples of secretory (n=<em>4</em>) and proliferative (n=<em>4</em>) eutopic endometrium and endometriosis, and no differences could be demonstrated. In conclusion, inactivation of 17beta-estradiol is impaired in endometriotic tissues due to deficient expression of 17betaHSD-2, which is normally expressed in eutopic endometrium in response to progesterone. The lack of 17betaHSD-2 expression in endometriosis is not due to alterations in the levels of immunoreactive progesterone or estrogen receptors in this tissue and may be related to an inhibitory aberration in the signaling pathway that regulates 17betaHSD-2 expression.
Publication
Journal: Journal of Clinical Endocrinology and Metabolism
August/25/1983
Abstract
A possible role for increased androgenic/estrogenic activity in the pathogenesis of upper body fat localization and its accompanying cellular and metabolic characteristics was examined. Eighty healthy, nonhirsute, premenopausal, caucasian women with a wide range of body fat topography [waist to hips girth ratio (WHR), 0.6<em>4</em> to 1.02] and obesity level (percentage of ideal body weight, 92-251%) were studied. Increasing androgenicity, as reflected by a decrease in plasma sex hormone-binding globulin capacity and an increase in the percentage of free testosterone, was accompanied by 1) increasing WHR, this relationship being independent of and additive to that of obesity level; 2) increasing size of abdominal, but not femoral, adipocytes; 3) increasing plasma glucose and insulin levels, both basally and in response to oral glucose loading; and <em>4</em>) diminished in vivo insulin sensitivity, as revealed by increasing steady state plasma glucose levels at comparable plasma insulin levels, attained by the infusion of somatostatin, insulin, and glucose. No association was found between total plasma testosterone, <em>androstenedione</em>, dehydroepiandrosterone sulfate, or estradiol concentrations and WHR, fat cell size, or metabolic profiles. We, therefore, propose that in premenopausal women, a relative increase in tissue exposure to unbound androgens may be responsible in part for localization of fat in the upper body, enlargement of abdominal adipocytes, and the accompanying imbalance in glucose-insulin homeostasis.
Publication
Journal: Clinical Endocrinology
October/15/1997
Abstract
OBJECTIVE
Polycystic ovary syndrome (PCOS) is a common endocrinopathy of unknown aetiology. The aims of this study were to identify whether ovarian thecal cell steroidogenesis is abnormally regulated in PCOS by measuring steroid responses to a single dose of hCG before, and during, suppression of endogenous LH levels by GnRH analogue (GnRHa).
METHODS
Serum levels of LH, FSH, 17 alpha hydroxyprogesterone (17OHP), <em>androstenedione</em>, testosterone and dehydroepiandrosterone sulphate were measured before, and <em>4</em>8 hours after, a single intramuscular injection of 10,000 IU hCG. The test was repeated <em>4</em> weeks after suppression of endogenous LH levels by GnRHa.
METHODS
The ovarian responses to hCG were compared in three groups of women. Eleven women had normal ovaries and regular cycles, eight had polycystic ovaries but not clinical features of the syndrome (PCO group) and eight had polycystic ovaries, anovulation and either severe hirsutism or alopecia (PCOS group).
RESULTS
Before GnRHa treatment, LH levels were significantly higher in the PCOS group but hCG stimulated a similar rise in 17OHP in all three groups. Following analogue, LH levels were suppressed in all three groups but the 17OHP responses to hCG were significantly higher in both the PCO and PCOS groups compared with normal controls.
CONCLUSIONS
These findings provide further evidence in favour of an intrinsic abnormality of thecal cell steroidogenesis in the polycystic ovary.
Publication
Journal: Journal of Clinical Investigation
April/2/1974
Abstract
Total and unbound testosterone and Delta(<em>4</em>)-<em>androstenedione</em> have been determined in 10<em>4</em> cord blood samples. The same sexual steroids and pituitary gonadotropins have been measured in <em>4</em>6 normal male infants aged 27-3<em>4</em>8 days and 3<em>4</em> normal female infants aged 19-332 days. In cord blood of female neonates mean total and unbound testosterone was 29.6+/-7.5 and 0.89+/-0.<em>4</em> ng/100 ml, respectively (mean+/-1 SD); Delta(<em>4</em>)-<em>androstenedione</em> was 93+/-38 ng/100 ml. In male neonates mean plasma total and unbound testosterone was 38.9+/-10.8 and 1.12+/-0.<em>4</em> ng/100 ml; Delta(<em>4</em>)-<em>androstenedione</em> was 85+/-27 ng/100 ml. In female infants testosterone concentrations remained constant during the 1st yr of life with a mean concentration of 7+/-3 ng/100 ml. Mean unbound testosterone and Delta(<em>4</em>)-<em>androstenedione</em> concentrations were 0.05+/-0.03 and 16.7+/-8.3 ng/100 ml, respectively. Mean plasma levels of follicle-stimulating hormone and luteinizing hormone were 8.7+/-3.3 and 12.9+/-7.7 mU/ml. In male infants mean plasma total testosterone concentration increased to 208+/-68 ng/100 ml from birth to 1-3 mo of age, decreasing thereafter to 95+/-53 ng/100 ml at 3-5 mo, 23.2+/-18 ng/100 ml at 5-7 mo, and reached prepubertal levels (6.6+/-<em>4</em>.6 ng/100 ml) at 7-12 mo. Mean unbound testosterone concentration plateaued from birth to 1-3 mo of age (1.3+/-0.2 ng/100 ml) decreasing to prepubertal values very rapidly. Mean Delta(<em>4</em>)-<em>androstenedione</em> concentration, although progressively decreasing during the 1st yr of life to 11.7+/-<em>4</em>.5 ng/100 ml, was higher than in the female at 1-3 mo of life (3<em>4</em>+/-11 ng/100 ml). Mean plasma level of follicle-stimulating hormone was 6.7+/-2.9 mU/ml, and that of luteinizing hormone was 19.7+/-13.5 mU/ml, significantly higher than in the female. There was no correlation between gonadotropin and age or testosterone. The present data demonstrate that the testes are active during the first natal period. It is tempting to correlate this phenomenon to a progressive maturation of the hypothalamo-pituitary-gonadal axis. It is possible that the surge in testosterone occurring the first 3 mo could play a role in the future life pattern of the male human being.
Publication
Journal: Endocrinology
September/21/1992
Abstract
We have investigated the effects of steroids on the intracellular calcium ion concentration [Ca2+]i in chicken granulosa cells obtained from the two largest preovulatory follicles of laying hens. [Ca2+]i was measured in cells loaded with the Ca(2+)-responsive fluorescent dye fura-2. The resting [Ca2+]i in these cells was 100 +/- 5 nM. There was an immediate (i.e. less than 5 sec) <em>4</em>- to 8-fold increase in [Ca2+]i in all of the 76 cells examined after the addition of 10(-7) M estradiol-17 bdta. Estradiol-17 beta was effective between 10(-10)-10(-6) M. Estradiol-17 alpha, estrone, and estriol (10(-8)-10(-6) M) were as effective as estradiol-17 beta, but the progestins, pregnenolone, and progesterone, and the androgens, testosterone, <em>androstenedione</em>, or 5 alpha-dihydrotestosterone were ineffective at concentrations up to 10(-5) M. The prompt estradiol-17 beta-induced [Ca2+]i spike was not affected by incubating the cells in Ca(2+)-free medium containing 2 mM EGTA or by pretreating them with the Ca2+ channel blockers lanthanum (1 mM), cobalt (5 mM), methoxyverapamil (D600; 50 microM), or nifedipine (20 microM). The estrogen-triggered [Ca2+]i surge was also not affected by pretreating the cells with the conventional estrogen receptor antagonist tamoxifen (10(-5) M), or the RNA and protein synthesis inhibitors actinomycin D (1 microgram/ml) and cycloheximide (1 microgram/ml), but was abolished by pretreating the cells with inhibitors of inositol phospholipid hydrolysis, neomycin (1.5 mM) and U-73,122 (2.5 microM). The closely related, but inactive, compound U-73,3<em>4</em>3 (1 microM) did not affect the estrogen-triggered [Ca2+]i surge. Estradiol-17 beta (10(-7) M), but not progesterone (10(-5) M), also triggered a large [Ca2+]i surge in pig granulosa cells, which, like the [Ca2+]i surge in chicken granulosa cells, was almost immediate, transient, and unaffected by incubation in Ca(2+)-free medium or pretreatment with methoxyverapamil (D600; 50 microM), lanthanum (1 mM), or tamoxifen (10(-5)M). However, granulosa cells from immature rats primed with diethylstilbestrol or PMSG did not respond to estradiol-17 beta, even at concentrations as high as 10(-5) M, although they promptly generated a [Ca2+]i transient upon exposure to LHRH (10(-5) M). These results suggest that estrogens almost instantaneously trigger the release of Ca2+ from intracellular stores which may be mediated through phosphoinositide breakdown. The striking rapidity of this estrogen-induced internal Ca2+ mobilization is consistent with the activation of a cell surface receptor which is different from the conventional slowly acting, gene-stimulating nuclear estrogen receptor.
Publication
Journal: Maturitas
June/19/1995
Abstract
From a longitudinal prospective study, 160 women with spontaneous menopause and without steroid medication were followed during the transition from pre- to postmenopause. After 12 years 152 women were still participating in the study. Blood samples were drawn every 6 months until 1 year after the menopause and every 12 months thereafter. Measurements of bone mineral density (BMD) on the forearm were performed every second year. All women routinely completed a questionnaire concerning symptoms frequently attributed to the climacteric period. All data were grouped around the onset of the menopause, thereby allowing longitudinal evaluation of the changes in the variables from the premenopausal to the postmenopausal period. The beginning of the perimenopausal period was characterized by transitory elevations of follicle-stimulating hormone (FSH). A significant increase in serum levels of gonadotropins was observed for both FSH and luteinizing hormone (LH) from about 5 years before the menopause. Within the 6 month period around the menopause there was a further increase which culminated within the first postmenopausal year for LH and 2-3 years postmenopause for FSH. Thereafter, a continuous decrease in LH occurred over the following 8 years. With respect to FSH, there was a slight decline starting about <em>4</em> years postmenopause. During the premenopausal period an increasing frequency of inadequate luteal function or anovulation occurred and, in the postmenopausal years, the serum levels of progesterone (P) were invariably low. Gradually, the ratio between estrone (E1) and 17-beta-estradiol (E2) increased, reflecting the declining follicular steroidogenesis. A marked decrease in estrogen levels occurred during the 6 month period around the menopause, most pronounced in E2. During the next 3 years, the levels of E2 and E1 showed an essentially parallel, moderate decline. Around the menopause, serum levels of testosterone (T), delta <em>4</em>-<em>androstenedione</em> (A) and sex hormone-binding globulin (SHBG) showed small but significant decreases. From about 3 years postmenopause, the levels were relatively constant over the following 5 years. A decrease in BMD was observed in the postmenopause, and from about 3 years postmenopause, estradiol correlated positively with BMD. Before, as well as after the menopause, body mass index (BMI) showed an inverse correlation with SHBG. Postmenopausal <em>androstenedione</em> correlated positively with E1, E2 and T. BMI correlated positively with E1 and E2. The concentrations of the free fraction of E2 and T are dependent on the levels of SHBG, which in turn has a negative correlation with BMI. The impact of this will influence the severity of symptoms, the degree of bone loss and the need for supplementary therapy.
Publication
Journal: Environmental Science & Technology
January/24/2006
Abstract
The potential occurrence of endocrine-disrupting compounds (EDCs) as well as pharmaceuticals and personal care products (PPCPs) in drinking water supplies raises concern over the removal of these compounds by common drinking water treatment processes. Three drinking water supplies were spiked with 10 to 250 ng/L of 62 different EDC/ PPCPs; one model water containing an NOM isolate was spiked with <em>4</em>9 different EDC/PPCPs. Compounds were detected by LC/MS/MS or GC/MS/MS. These test waters were subjected to bench-scale experimentation to simulate individual treatment processes in a water treatment plant (WTP). Aluminum sulfate and ferric chloride coagulants or chemical lime softening removed some polyaromatic hydrocarbons (PAHs) but removed <25% of most other EDC/ PPCPs. Addition of 5 mg/L of powder activated carbon (PAC) with a <em>4</em>-h contact time removed 50% to >98% of GC/ MS/MS compounds (more volatile) and 10% to >95% of LC/ MS/MS compounds (more polar); higher PAC dosages improved EDC/PPCP removal. EDC/PPCP percentage removal was independent of the initial compound concentration. Octanol-water partition coefficients served as a reasonable indicator of compound removal under controlled PAC test conditions, except for EDC/PPCPs that were protonated or deprotonated at the test pH and some that contained heterocyclic or aromatic nitrogen. Separate chlorine or ozone experiments decreased the EDC/PPCP initial concentration by <10% to >90%; EDC/PPCPs were likely transformed to oxidation byproducts. Ozone oxidized steroids containing phenolic moieties (estradiol, ethynylestradiol, or estrone) more efficiently than those without aromatic or phenolic moieties (<em>androstenedione</em>, progesterone, and testosterone). EDC/PPCP reactivity with oxidants were separated into three general groups: (1) compounds easily oxidized (>80% reacted) by chlorine are always oxidized at least as efficiently by ozone; (2) 6 of the -60 compounds (TCEP, BHC, chlordane, dieldrin, heptachlor epoxide, musk ketone) were poorly oxidized (<20% reacted) by chlorine or ozone; (3) compounds (2<em>4</em> of 60) reacting preferentially (higher removals) with ozone rather than chlorine. Conventional treatment (coagulation plus chlorination) would have low removal of many EDC/PPCPs, while addition of PAC and/or ozone could substantially improve their removals. Existing strategies that predict relative removals of herbicides, pesticides, and other organic pollutants by activated carbon or oxidation can be directly applied for the removal of many EDC/PPCPs, but these strategies need to be modified to account for charged (protonated bases or deprotonated acids) and aliphatic species. Some compounds (e.g., DEET, ibuprofen, gemfibrozil) had low removals unless ozonation was used. Other compounds had low removals by all the WTP processes considered (atrazine, iopromide, meprobamate, TCEP), and removal processes capable of removing these types of compounds should be investigated.
Pulse
Views:
1
Posts:
No posts
Rating:
Not rated
Publication
Journal: Progress in Neuro-Psychopharmacology and Biological Psychiatry
May/8/2005
Abstract
The term "neurosteroid" (NS) was introduced by Baulieu in 1981 to name a steroid hormone, dehydroepiandrosterone sulfate (DHEAS), that was found at high levels in the brain long after gonadectomy and adrenalectomy, and shown later to be synthetized by the brain. Later, <em>androstenedione</em>, pregnenolone and their sulfates and lipid derivatives as well as tetrahydrometabolites of progesterone (P) and deoxycorticosterone (DOC) were identified as neurosteroids. The term "neuroactive steroid" (NAS) refers to steroids which, independent of their origin, are capable of modifying neural activities. NASs bind and modulate different types of membrane receptors. The GABA and sigma receptor complexes have been the most extensively studied, while glycine-activated chloride channels, nicotinic acetylcholine receptors, voltage-activated calcium channels, although less explored, are also modulated by NASs. Within the glutamate receptor family, N-methyl-d-aspartate (NMDA) receptors, alpha-amino-3-hydroxy-5-methyl-<em>4</em>-isoxazolepropionic acid (AMPA) receptors and kainate receptors have also been demonstrated to be a target for steroid modulation. Besides their membrane effects, once inside the neuron oxidation of Ring A reduced pregnanes, THP and THDOC, bind to the progesterone intracellular receptor and regulate gene expression through this path. The involvement of NASs on depression syndromes, anxiety disorders, stress responses to different stress stimuli, memory processes and related phenomena such as long-term potentiation are reviewed and critically evaluated. The importance of context for the interpretation of behavioral effects of hormones as well as for hormonal levels in body fluids is emphasized. Some suggestions for further research are given.
Publication
Journal: Journal of Clinical Oncology
November/11/2012
Abstract
OBJECTIVE
Estrogens and androgens are elevated in obesity and associated with increased postmenopausal breast cancer risk, but the effect of weight loss on these biomarkers is unknown. We evaluated the individual and combined effects of a reduced-calorie weight loss diet and exercise on serum sex hormones in overweight and obese postmenopausal women.
METHODS
We conducted a single-blind, 12-month, randomized controlled trial from 2005 to 2009. Participants (age 50 to 75 years; body mass index>> 25.0 kg/m(2), exercising < 100 minutes/wk) were randomly assigned using a computer-generated sequence to (1) reduced-calorie weight loss diet ("diet"; n = 118), (2) moderate- to vigorous-intensity aerobic exercise ("exercise"; n = 117), (3) combined reduced-calorie weight loss diet and moderate- to vigorous-intensity aerobic exercise ("diet + exercise"; n = 117), or (<em>4</em>) control (n = 87). Outcomes were estrone concentration (primary) and estradiol, free estradiol, total testosterone, free testosterone, <em>androstenedione</em>, and sex hormone-binding globulin (SHBG) concentrations (secondary).
RESULTS
Mean age and body mass index were 58 years and 30.9 kg/m(2), respectively. Compared with controls, estrone decreased 9.6% (P = .001) with diet, 5.5% (P = .01) with exercise, and 11.1% (P < .001) with diet + exercise. Estradiol decreased 16.2% (P < .001) with diet, <em>4</em>.9% (P = .10) with exercise, and 20.3% (P < .001) with diet + exercise. SHBG increased 22.<em>4</em>% (P < .001) with diet and 25.8% (P < .001) with diet + exercise. Free estradiol decreased 21.<em>4</em>% (P < .001) with diet and 26.0% (P < .001) with diet + exercise. Free testosterone decreased 10.0% (P < .001) with diet and 15.6% (P < .001) with diet + exercise. Greater weight loss produced stronger effects on estrogens and SHBG.
CONCLUSIONS
Weight loss significantly lowered serum estrogens and free testosterone, supporting weight loss for risk reduction through lowering exposure to breast cancer biomarkers.
load more...