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 218 毫秒
1.
OBJECTIVE: To compare the efficacy of two antiandrogens, cyproterone acetate (CPA) and spironolactone, in the treatment of hirsutism. DESIGN: Prospective randomized single-blinded study. SETTING: A tertiary hirsutism clinic. PATIENTS: Forty-two premenopausal patients with hirsutism were selected. INTERVENTIONS: Subjects were randomized to receive either 100 mg spironolactone and an oral contraceptive (OC) containing 150 microg desogestrel and 30 microg ethinyl E2 or 50 mg CPA daily on days 1 to 10 of the menstrual cycle, which was administered with 35 microg ethinyl E2 daily on days 1 to 21. MAIN OUTCOME MEASURES: Hirsutism scores were measured according to Ferriman-Gallwey scoring system and side effects were monitored for 9 months of treatment. Blood samples were taken at each visit for assessment of endocrine, biochemical, and hematologic parameters. RESULTS: Hirsutism scores were decreaded significantly in both groups at the end of 9 months. The percent of change in hirsutism scores in CPA and spironolactone group were as follows: 19.23% +/- 14.77% and 24.48% +/- 14.27% at 3 months; 39.01% +/- 19.77% and 37.46% +/- 16.90% at 6 months; and 51.89% +/- 20.87% and 46.39% +/- 16.10% at 9 months, respectively. There was a trend toward a better response with CPA treatment, which did not achieve significance. None of the patients stopped treatment because of side effects. CONCLUSION: The present data suggest that both spironolactone and CPA were similarly effective in treatment of hirsutism.  相似文献   

2.
OBJECTIVE: To compare the efficacy of finasteride and spironolactone in the treatment of idiopathic hirsutism. DESIGN: Prospective, randomized, single-blind study. SETTING: A tertiary hirsutism clinic. PATIENT(S): Forty women with idiopathic hirsutism were selected. INTERVENTION(S): Patients were assigned randomly to receive either 5 mg of finasteride or 100 mg of spironolactone for 9 months. MAIN OUTCOME MEASURE(S): Hirsutism scores were measured according to the Ferriman-Gallwey scoring system, and side effects were monitored for 9 months of treatment. Blood samples were taken at each visit for assessment of endocrine, biochemical, and hematologic parameters. RESULT(S): Hirsutism scores were decreased significantly in both groups at the end of 9 months. The mean percent change (+/- SD) in hirsutism scores in the finasteride and spironolactone groups was as follows: 5.91% +/- 7.18% and 20.60% +/- 12.59% at 3 months, 10.61% +/- 12.18% and 32.57% +/- 15.68% at 6 months, and 15.15% +/- 15.38% and 42.36% +/- 12.31% at 9 months, respectively. There was a significantly better response with spironolactone treatment at the end of 9 months. Eleven (55%) of 20 patients in the spironolactone group experienced side effects. However, none of them stopped treatment because of side effects. CONCLUSION(S): The present data suggest that both finasteride and spironolactone are effective in the treatment of idiopathic hirsutism. However, it appears that the spironolactone group responded significantly better.  相似文献   

3.
OBJECTIVE: To evaluate the effects of long-term administration of finasteride on hirsutism score, basal gonadotropin, and androgen secretion in women with idiopathic hirsutism. DESIGN: Randomized single-blinded study. PATIENTS: Eighteen patients with moderate-severe hirsutism were recruited for the study. INTERVENTIONS: Nine hirsute patients received 7.5 mg/d oral finasteride for a period of 9 months whereas the other nine were treated with placebo. Hirsutism score, serum basal gonadotropin, androgens, estrogen, and sex hormone-binding globulin (SHBG) levels were evaluated in all patients before treatment and every 3 months during treatment. RESULTS: After 6 and 9 months of treatment, the hirsutism score improved significantly in the patients receiving finasteride, whereas no significant modifications were observed in patients treated with placebo. The side effects observed were headache and depression of modest entity during the 1st month of treatments, whereas libido did not change. Serum levels of LH, FSH, androstenedione, unbound T, DHEAS, E2, 17 alpha-hydroxyprogesterone, and SHBG did not change during therapy. Hirsute patients treated with finasteride exhibited a marked decrease of dihydrotestosterone and a significant increase of T serum levels from the 3rd and 6th months of treatment, respectively. CONCLUSION: Finasteride decreased the hirsutism score of patients affected by idiopathic hirsutism with few side effects during treatment. No modification of libido was observed.  相似文献   

4.
Finasteride, a 5 alpha-reductase inhibitor, decreases prostate size and improves symptoms in men with benign prostatic hyperplasia. However, little is known about prostate histopathology in men taking finasteride. To determine the mechanism by which finasteride reduces prostate size, tissue was collected at the time of prostatectomy from men taking either no medication (n = 10) or 5 mg finasteride daily for 6-18 days (n = 6; group 1), 23-73 days (n = 5; group 2), or 3 months to 4 yr (n = 5; group 3). To assess whether finasteride causes epithelial atrophy, morphometric measurement of epithelial cell and duct width was used. The mean epithelial cell width in control prostates (mean +/- SEM, 21 +/- 0.7 microns) decreased with duration of treatment to 19 +/- 1 microns in group 1, 15 +/- 2 microns in group 2, and 8 +/- 0.3 microns in group 3. Mean duct width decreased from 135 +/- 6 microns in the control prostates to 128 +/- 10 microns in group 1, 103 +/- 3 microns in group 2, and 63 +/- 6 microns in group 3. To assess whether prostate cell death was occurring, sections were in situ end labeled for DNA breaks and immunostained for tissue transglutaminase (tTG), a marker of apoptosis (programmed cell death). The percentage of epithelial cells staining for DNA breaks was 0.4 +/- 0.2 in control prostates, 2.8 +/- 0.9 in group 1, 1.7 +/- 0.5 in group 2, and 0.7 +/- 0.3 microns in group 3. Anti-tTG staining of epithelial cells was graded on a scale of 0-4. In control prostates, 3 +/- 1% of the ducts were grade 3 or 4 (> 50% of epithelial cells staining). In finasteride-treated prostates, 2 +/- 2% of the prostates in group 1, 13 +/- 4% of the prostates in group 2, and 0.5 +/- 0.5% of the prostates in group 3 were grade 3-4. These results indicate that a progressive decrease in epithelial cell size and function occurs during the first several months in the prostates of men treated with finasteride. The staining for DNA breaks and the tTG staining also indicate that an increased rate of apoptosis is occurring transiently in these prostates. We conclude that finasteride causes prostate involution through a combination of atrophy and cell death.  相似文献   

5.
Steroid 5alpha-reductase is a system of two isozymes (5alphaR-1 and 5alphaR-2) which catalyzes the NADPH-dependent reduction of testosterone to dihydrotestosterone in many androgen sensitive tissues and which is related to several human endocrine diseases such as benign prostatic hyperplasia (BPH), prostatic cancer, acne, alopecia, pattern baldness in men and hirsutism in women. The discovery of new potent and selective 5alphaR inhibitors is thus of great interest for pharmaceutical treatment of these diseases. The synthesis of a novel class of inhibitors for human 5alphaR-1 and 5alphaR-2, having the 19-nor-10-azasteroid skeleton, is described. The inhibitory potency of the 19-nor-10-azasteroids was determined in homogenates of human hypertrophic prostates toward 5alphaR-2 and in DU-145 human prostatic adenocarcinoma cells toward 5alphaR-1, in comparison with finasteride (IC50 = 3 nM for 5alphaR-2 and approximately 42 nM for 5alphaR-1), a drug which is currently used for BPH treatment. The inhibition potency was dependent on the type of substituent at position 17 and on the presence and position of the unsaturation in the A and C rings. delta9(11)-19-Nor-10-azaandrost-4-ene-3,17-dione (or 10-azaestra-4,9(11)-diene-3,17-dione) (4a) and 19-nor-10-azaandrost-4-ene-3,17-dione (5) were weak inhibitors of 5alphaR-2 (IC50 = 4.6 and 4.4 microM, respectively) but more potent inhibitors of 5alphaR-1 (IC50 = 263 and 299 nM, respectively), whereas 19-nor-10-aza-5alpha-androstane-3,17-dione (7) was inactive for both the isoenzymes. The best result was achieved with the 9:1 mixture of delta9(11)- and delta8(9)-17beta-(N-tert-butylcarbamoyl)-19-nor-10-aza-4- androsten-3-one (10a,b) which was a good inhibitor of 5alphaR-1 and 5alphaR-2 (IC50 = 127 and 122 nM, respectively), with a potency very close to that of finasteride. The results of ab initio calculations suggest that the inhibition potency of 19-nor-10-azasteroids could be directly related to the nucleophilicity of the carbonyl group in the 3-position.  相似文献   

6.
OBJECTIVE: To develop a pharmacokinetic-pharmacodynamic model that characterizes the conversion of testosterone to dihydrotestosterone (DHT) by 5 alpha-reductase types 1 and 2 and the irreversible inhibition of 5 alpha-reductase by finasteride, a 5 alpha-reductase type 2 inhibitor and by GI198745 (dutasteride), a potent and specific dual 5 alpha-reductase inhibitor. METHODS: Healthy men (n = 48) received doses of 0.1 to 40 mg GI198745 (n = 4 subjects per dose), 5 mg finasteride (n = 8), or placebo (n = 8) in a parallel-group study. Plasma concentrations of GI198745, finasteride, and DHT were measured frequently up to 8 weeks after dosing. Models were fitted with mixed-effects modeling with the NONMEM program. RESULTS: The pharmacodynamics were well described with a model that accounted for the rates of DHT formation and elimination, 5 alpha-reductase turnover, relative capacity of the 2 5 alpha-reductase isozymes, and the rates of irreversible inhibition of one (finasteride) or both (GI198745) types of 5 alpha-reductase. The model indicated that type 2 5 alpha-reductase contributed approximately 80% of plasma DHT. GI198745 was about 3-fold more potent than finasteride on 5 alpha-reductase type 2. Nearly full blockade of both isozymes was achieved at doses of 10 mg or more GI198745, although the potency of this agent on 5 alpha-reductase type 1 was less than on type 2. CONCLUSIONS: A physiologically based model for the turnover and irreversible inhibition of 5 alpha-reductase and for formation and elimination of DHT described the data well. This model helps explain differences in the rates of onset and offset of effect and offers a way to determine the relative potency of the irreversible 5 alpha-reductase inhibitors.  相似文献   

7.
A novel series of indole and benzimidazole derivatives were synthesized and evaluated for their inhibitory activity of rat prostatic 5alpha-reductase. Among these compounds, 4-?2-[1-(4,4'-dipropylbenzhydryl)indole-5-carboxamido]phenoxy?buty ric acid (15) and its benzimidazole analogue 25 showed potent inhibitory activities for rat prostatic 5alpha-reductase (IC50 values of 9.6+/-1.0 and 13+/-1.5 nM, respectively), with the potency very close to that of finasteride. Compound 30, in which the moiety between the benzene ring and amide bond was replaced by quinolin-4-one ring, showed almost equipotent activity (IC50= 19+/-6.2nM) with the correspondent amide derivative 13. This result was consistent with the previous observation that the coplanarity of this moiety might contribute to the potent inhibitory activity.  相似文献   

8.
We studied bone mineral content (BMC), bone mineral density (BMD), cortical thickness/total width (CT/TW) ratio and cortical area/total area (CA/TA) ratio in boys with constitutional delay of puberty and the effect of short-term testosterone treatment on bone mass. Seventeen boys (age 13.1-15.8 years) who met the family history and the clinical criteria of constitutional delay of puberty were selected and enrolled in the study. All subjects were eating a diet assuring an adequate intake of calories and calcium. A subset of 8 boys (group A) was treated with testosterone depot (100 mg/month x 6 months) while 9 boys (group B) were not. At inclusion, BMC and BMD were reduced in the patients according to their chronological age (BMC -4.04 +/- 1.34 standard deviation scores [SDS]; BMD -2.95 +/- 0.56 SDS), statural age (BMC -1.75 +/- 0.79 SDS; BMD -1.69 +/- 0.78 SDS), and bone age (BMC -1.80 +/- 0.65 SDS; BMD -1.86 +/- 0.68 SDS). No significant differences between the groups were found (group A: BMC 0.480 +/- 0.57 g/cm, BMD 0.488 +/- 0.037 g/cm2, CT/TW ratio 0.43 +/- 0.4, CA/TA ratio 0.68 +/- 0.04; group B: BMC 0.476 +/- 0.060, p = NS vs. group A; BMD 0.491 +/- 0.036 g/cm2, p = NS vs. group A). At 12 months of follow-up, BMC, BMD, CT/TW ratio, and CA/TA ratio significantly increased in group A (BMC 0.70 +/- 0.13 g/cm, delta +41.1 +/- 28.8%, p < 0.003 vs. 0 month; BMD 0.617 +/- 0.082 g/cm2, delta +26.2 +/- 13.6%, p < 0.005 vs. 0 month; CT/TW ratio 0.52 +/- 0.05, delta +20.59 +/- 10.65%, p < 0.001 vs. 0 month; CA/TA ratio 0.77 +/- 0.05 vs. 0 month; CT/TW ratio 13.60 +/- 6.65%, p < 0.004 vs 0 month), but not in group B (BMC: 0.48 +/- 0.05 g/cm; delta +5.1 7.8%, p = NS vs. 00 month; BMD: 0.492 +/- 0.037 g/cm2; delta +0.54 +/- 8.7%, p = NS vs. 0 month; CT/TW ratio 0.44 +/- 0.04, delta +4.04 +/- 6.75%, p = NS vs. 0 month; CA/TA ratio 0.68 +/- 0.05, delta +2.39 +/- 5.90%, p = NS vs. 0 month). We conclude that boys with constitutional delay of puberty have reduced BMC and BMD. The delay in statural and bone ages did not totally account for the decreased bone mass. Testosterone treatment for 6 months significantly increased BMC, BMD, CT/TW ratio, and CA/TA ratio in these patients, but definitive conclusions on the efficacy of the treatment in improving adult bone mass can be drawn only when our patients reach early childhood.  相似文献   

9.
OBJECTIVES: Identify factors predicting favorable outcome after medical management of valve ring abscesses in order to propose a surveillance schedule for conservative treatment. METHODS: A multicentric study conducted from July 1989 to February 1996 included 28 patients (mean age 64 +/- 16 years, range 26-83) hospitalized for active endocarditis and valve ring abscesses diagnosed at transthoracic or transesophageal echography. Conservative medical therapy was given because of a decision of the medico-surgical team (n = 9), high surgical risk (n = 12), or patient refusal of surgery (n = 7). Outcome was favourable in 18 patients (Group I) and unfavorable in 10 (Group II) due to death (n = 9) or subsequent surgery (n = 1). Univariate and multivariate analysis were used to determine differences between the groups in terms of clinical and laboratory data. RESULTS: Mean follow-up in Group I was 33 +/- 18 months and 15 +/- 10 months in Group II. Univariate analysis showed significant differences between Group I and II respectively for age (59 +/- 18 yr vs 72 +/- 10, p = 0.04), delay to apyrexia after antibiotics (4.3 +/- 2.8 vs 8.3 +/- 2.4 days, p < 0.0008), heart failure (5% vs 70%, p = 0.003), grade III or IV valvular regurgitation (5% vs 60%, p < 0.04), and mean surface area of the abscess (1.5 +/- 1.2 vs 5.4 +/- 6.4 cm2, p < 0.03). Independent factors at multivariate analysis were by decreasing order: lack of heart failure at admission, delay to apyrexia, abscess surface area, and age. Outcome was favorable (mean follow-up 33 +/- 10 months) in all patients with an abscess surface area < 1.5 cm2, no signs of heart failure, no grade III or IV valvular regurgitation, apyrexia after less than 8 days on antibiotics and no staphylococcus positive blood culture. CONCLUSION: Medical management of valve ring abscesses may be indicated in selected patients in care units with rigorous surveillance facilities. Further studies are needed to precisely identify surveillance and treatment criteria.  相似文献   

10.
PURPOSE: Finasteride therapy for benign prostatic hyperplasia (BPH) results in a marked lowering of serum prostate specific antigen (PSA) levels. However, little is known about the effect of finasteride on unbound or free serum levels of PSA. Such information would be important since percent free PSA may substantially improve the cancer specificity of PSA testing. Thus, we prospectively studied the effect of finasteride therapy on total and free serum PSA levels. MATERIALS AND METHODS: In a randomized, placebo controlled, double-blind trial 40 men with histologically confirmed BPH (age range 52 to 78 years) were treated with either 5 mg. finasteride daily (26 patients) for 9 months or placebo (14) for 6 months. Prostate volume was assessed by transrectal ultrasound. Serum levels of free and total PSA were measured from archived serum samples stored at -70C at baseline and for as long as 9 months of treatment. RESULTS: In the finasteride group mean total PSA levels declined from 3.0 ng./ml. at baseline to 1.5 ng./ml. after 6 months of treatment (50% decrease, p <0.01). In the placebo group, with similar baseline levels, no significant change was observed. PSA density declined significantly in finasteride treated men (p <0.01) but not in men receiving placebo. The mean percent free PSA (13 to 17% at baseline) was not altered significantly by finasteride or placebo. CONCLUSIONS: Total PSA serum levels decreased by an average of 50% during finasteride therapy but percent free PSA did not change significantly. This information is potentially useful in the interpretation of PSA data used for early detection of prostate cancer in men receiving finasteride. However, further studies are required to demonstrate the use of percent free PSA to detect the development of cancer.  相似文献   

11.
Nine healthy male subjects underwent measurement of reflex sympathetic function, pressor responsiveness and baroreflex sensitivity to phenylephrine (PE) and glyceryltrinitrate (GTN) before (C1) and following six days of treatment (E6) with cortisol (F), 200 mg/day. Seven subjects had washout studies (W) performed at least two weeks following the end of treatment. The BP responses to head tilt, isometric exercise and mental arithmetic were unaltered by F, however, there was a significant diminution of the diastolic BP response to cold pressor stimulus (delta DBP: 19 +/- 3 vs 25 +/- 5 vs 27 +/- 5 mmHg; E6 vs C1 vs W, p < 0.05 C1 vs E6 and W). Baroreflex sensitivity to PE was increased (28 +/- 3 vs 19 +/- 2 ms/mmHg, E6 vs C1, p = 0.03). These data demonstrate that increased BP during F treatment is not attributable to increased SNS activity, and suggest that SNS activity may be decreased by F.  相似文献   

12.
The rhesus macaque types 1 and 2 5alpha-reductase (5aR1 and 5aR2) were cloned and expressed in COS cells to facilitate comparison of rhesus and human 5aRs. The deduced protein sequences of the rhesus SaRs shared 94% and 96% identity with the human type 1 and 2 isozymes, respectively. Despite a four amino acid insertion at the N-terminal region of rhesus 5aR1, the biochemical properties of rhesus and human homologs are very similar with respect to pH optimum, Km values for testosterone and progesterone, and inhibition by a variety of inhibitors. As expected, the biochemical properties of the human and rhesus 5aR2 are also very similar. The mechanism of inhibition of the rhesus 5aR1 and 5aR2 by finasteride was investigated in more detail. Finasteride displays time dependent inhibition of the rhesus 5aR1 and 5aR2 with second order rate constants of 4 x 10(3) M(-1) s(-1) and 5.2 x 10(5) M(-1)s(-1). Inhibition of rhesus 5aR2 with 3H-finasteride resulted in 3H bound to the enzyme which is not released by dialysis. Heat denaturation of the [rhesus SaR2:inhibitor] complex releases dihydrofinasteride, a breakdown product presumably related to the NADP+-adduct previously identified with the human SaRs (Bull et al., Mechanism-based inhibition of human steroid 5alpha-reductase by finasteride: Enzyme catalyzed formation of NADP-dihydrofinasteride, a potent bisubstrate analog inhibitor. J. Amer. Chem. Soc., 1996, 118, 2359-2365). Taken together, these results provide good evidence that the rhesus macaque is a suitable model to evaluate the pharmacological properties of finasteride and other 5aR inhibitors.  相似文献   

13.
In the treatment of acne in women, the use of antiandrogens and other hormonal approaches is a valuable alternative to standard therapy. These treatments that are based on physiologically sound principles produce gratifying results in selected women with acne, and are the primary treatment for women with hirsutism. The drugs discussed in this article include spironolactone, cyproterone acetate, flutamide, oral contraceptives, corticosteroids, finasteride, and gonadotropin-releasing hormone agonists. Patient selection, pretreatment evaluation, and case studies also are discussed with an emphasis on practical applications.  相似文献   

14.
In this study we addressed whether the proportion and the function of antigen presenting cells (APC), T and NK lymphocytes are modified in the apheresis product of six healthy donors who received a stem cell mobilizing treatment with glycosylated G-CSF at 10 microg/kg/day x 5 days s.c. Flow cytometry analysis showed comparable percentages of HLA-DR+, CD19+, CD86+, CD80+ and CD1a+ cells in preG-CSF-peripheral blood mononuclear cells (preG-PBMC) and after mobilization in G-PBMC, whereas the proportion of CD14+ monocytes significantly increased in G-PBMC (3+/-1% vs 17+/-8%, P = 0.003). Analysis of lymphocyte subsets in preG-PBMC and G-PBMC showed similar proportions of CD3+, CD4+, CD8+ and CD28+ T cells, but a significantly lower percentage of CD16+ (11+/-7% vs 4+/-1%, P=0.01), CD56+ (15+/-6% vs 5+/-2%, P= 0.008), CD57+ (16+/-9% vs 5+/-2%, P=0.04), CD25+ (19+/-2% vs 9+/-6%, p=0.009) and CD122+ (5+/-2% vs 2+/-1%, P = 0.05) cells in G-PBMC. Unfractionated preG-PBMC and G-PBMC were irradiated and tested in primary mixed leukocyte culture (MLC) with two HLA-incompatible responders and induced efficient alloresponses in four of six cases, whereas G-PBMC stimulated poorly in the remaining two cases. Also, in allo-MLC with irradiated G-PBMC we detected lower amounts of IFN-gamma (P = 0.04) and of IL-2 (P = 0.06) than in allo-MLC with preG-PBMC. Furthermore, freshly isolated preG-PBMC and G-PBMC from each donor exerted comparable allogeneic responses to HLA-incompatible irradiated mononuclear cells in all cases. However, G-PBMC showed no NK activity against K562 target cells at any effector:target ratio tested. These data suggest that normal G-PBMC may prevent Thl alloresponses, maintain efficient alloreactivity to HLA mismatched antigens and have impaired NK activity.  相似文献   

15.
To evaluate whether atherosclerosis may be associated with altered leucocyte rheology, we assessed leucocyte count (by Coulter counter), aggregation (by means of the leukergy test) and expression of adhesion molecules integrin LFA-1 and CD 44 (by means of immunofluorescence staining and flow cytometry) in 9 patients with carotid plus lower limb artery atherosclerosis (group A), 14 patients with carotid atherosclerosis only (group B) and 23 controls without atherosclerosis (group C). The level of LFA-1 (calculated as mean fluorescence channels-MFCs) on neutrophils, lymphocytes and monocytes was significantly higher (p < 0.05) in group A and B patients than in controls (group A-mean +/- SE: 383.77 +/- 9.42 vs 295.45 +/- 5.76; 474.22 +/- 8.86 vs 388.35 +/- 7.84; 457.66 +/- 12.03 vs 396.25 +/- 4.37. Group B: 322.42 +/- 6.36 vs 295.45 +/- 5.76; 421.42 +/- 7.21 vs 388.35 +/- 7.84; 415.71 +/- 7.73 vs 396.25 +/- 4.37, respectively); furthermore, the MFC of LFA-1 on neutrophils was significantly different (p < 0.05) between group A and B patients. The percentage of aggregated leucocytes was significantly higher (p < 0.05) in group A patients (4.46 +/- 1.07) than those in groups B (1.75 +/- 0.38) and C (1.43 +/- 0.25), whereas no significant difference was detected between groups B and C. Leucocyte number and expression of CD44 were not significantly different among the 3 groups. In conclusion, changes in leucocyte rheology are present in patients with atherosclerosis and may contribute to chronic ischaemia.  相似文献   

16.
STUDY OBJECTIVE: To evaluate serial lung function studies, including elastic recoil, in patients with severe emphysema who undergo lung volume reduction surgery (LVRS). To determine mechanism(s) responsible for changes in airflow limitation. METHODS: We studied 12 (10 male) patients aged 68+/-9 years (mean+/-SD) 6 to 12 months prior to and at 6-month intervals for 2 years after thoracoscopic bilateral LVRS for emphysema. RESULTS: At 2 years post-LVRS, relief of dyspnea remained improved in 10 of 12 patients, and partial or full-time oxygen dependency was eliminated in 2 of 7 patients. There was significant reduction in total lung capacity (TLC) compared with pre-LVRS baseline, 7.8+/-0.6 L (mean+/-SEM) (133+/-5% predicted) vs 8.6+/-0.6 L (144+/-5% predicted) (p=0.003); functional residual capacity, 5.6+/-0.5 L (157+/-9% predicted) vs 6.7+/-0.5 L (185+/-10% predicted) (p=0.001); and residual volume, 4.9+/-0.5 L (210+/-16% predicted) vs 6.0+/-0.5 L (260+/-13% predicted) (p=0.000). Increases were noted in FEV1, 0.88+/-0.08 L (37+/-6% predicted) vs 0.72+/-0.05 L (29+/-3% predicted) (p=0.02); diffusing capacity, 8.5+/-1.0 mL/min/mm Hg (43+/-3% predicted) vs 4.2+/-0.7 mL/min/mm Hg (18+/-3% predicted) (p=0.001); static lung elastic recoil pressure at TLC (Pstat), 13.7+/-0.5 cm H2O vs 11.3+/-0.6 cm H2O (p=0.008); and maximum oxygen consumption, 8.7+/-0.8 mL/min/kg vs 6.9+/-1.5 mL/min/kg (p=0.03). Increase in FEV1 correlated with the increase in TLC Pstat/TLC (r=0.75, p=0.03), but not with any baseline parameter. CONCLUSION: Two years post-LVRS, there is variable clinical and physiologic improvement that does not correlate with any baseline parameter. Increased lung elastic recoil appears to be the primary mechanism for improved airflow limitation.  相似文献   

17.
1. The presence of A2 receptors mediating relaxation in the rat isolated aorta has been previously demonstrated. However, agonist dependency of the degree of rightward shift elicited by 8-sulphophenyltheophylline (8-SPT) led to the suggestion that the population of receptors in this tissue is not a homogeneous one. In this study we have re-examined the effects of 8-SPT in the absence and presence of the NO synthase inhibitor L-NAME (NG-nitro-L-arginine methyl ester) and investigated antagonism of responses by the potent A2a receptor ligands PD 115,199 (N-[2-dimethylamino)ethyl]-N-methyl-4-(2,3,6,7-tetrahydro-2,6-dioxo-1,3 dipropyl-1H-purin-8-yl)) benzene sulphonamidexanthine), ZM 241385 (4-(2-[7-amino-2-(2-furyl) [1,2,4]-triazolo[2,3-a][1,3,5]triazin-5-yl amino]ethyl)phenol), and CGS 21680 (2-[p-(2-carboxyethyl)phenylamino]-5'-N-ethylcarboxamidoadenosine). We have also investigated the antagonist effects of BWA1433 (1,3-dipropyl-8-(4-acrylate)phenylxanthine) which has been shown to have affinity at rat A3 receptors. 2. Adenosine, R-PIA (N6-R-phenylisopropyl adenosine), CPA (N6-cyclopentyladenosine) and NECA (5'-N-ethylcarboxamidoadenosine) all elicited relaxant responses in the phenylephrine pre-contracted rat isolated aorta with the following potency order (p[A50] values in parentheses): NECA (7.07 +/- 0.11) > R-PIA (5.65 +/- 0.10) > CPA (5.05 +/- 0.12) > adenosine (4.44 +/- 0.12). 3. 8-SPT (10-100 microM) caused parallel rightward shifts of the E/[A] curves to NECA (pKB = 5.23 +/- 0.16). A smaller rightward shift of E/[A] curves to CPA was observed (pA2 = 4.85 +/- 0.17). However, no significant shifts of E/[A] curves to either adenosine or R-PIA were observed. 4. In the absence of endothelium E/[A] curves to NECA and CPA were right-shifted compared to controls. However, removal of the endothelium did not produce a substantial shift of adenosine E/[A] curves, and E/[A] curves to R-PIA were unaffected by removal of the endothelium. 5. In the presence of L-NAME (100 microM) E/[A] curves to NECA and CPA were right-shifted. However, no further shift of the CPA E/[A] curve was obtained when 8-SPT (50 microM) was administered concomitantly. The locations of curves to R-PIA and adenosine were unaffected by L-NAME (100 microM). 6. In the presence of PD 115,199 (0.1 microM) a parallel rightward shift of NECA E/[A] curves was observed (pA2 = 7.50 +/- 0.19). PD 115,199 (0.1 and 1 microM) gave smaller rightward shifts of E/[A] curves to R-PIA and CPA, but E/[A] curves to adenosine were not significantly shifted in the presence of PD 115,199 (0.1 or 1 microM). 7. The presence of ZM 241385 (3 nM-0.3 microM) caused parallel rightwad shifts of NECA E/[A] curves (pKB = 8.73 +/- 0.11). No significant shifts of E/[A] curves to adenosine, CPA or R-PIA were observed in the presence of 0.1 microM ZM 241385. 8. CGS 21680 (1 microM) elicited a relaxant response equivalent to approximately 40% of the NECA maximum response. In the presence of this concentration of CGS 21680, E/[A] curves to NECA were right-shifted in excess of 2-log units, whereas E/[A] curves to R-PIA were not significantly shifted. 9. BWA1433 (100 microM) caused a small but significant right-shift of the E/[A] curve to R-PIA yielding a pA2 estimate of 4.1 IB-MECA (N6-(3-iodo-benzyl)adenosine-5(1)-N-methyl uronamide) elicited relaxant responses which were resistant to blockade by 8-SPT (p[A]50 = 5.26 +/- 0.13). 10. The results suggest that whereas relaxations to NECA (10 nM-1 microM) are mediated via adenosine A2a receptors, which are located at least in part on the endothelium, R-PIA and CPA may activate A2b receptors on the endothelium and an additional, as yet undefined site, which is likely to be located on the smooth muscle and which is not susceptible to blockade by 8-SPT, PD 115,199 or ZM 241385. This site is unlikely to be an A3 receptor since the very small shift obtained in the presence of BWA1433 (100 microM), and the low potency of IB-MECA is not consistent with the affin  相似文献   

18.
The effects of thapsigargin (Tg) and cyclopiazonic acid (CPA), two selective blockers of the sarcoplasmic reticulum Ca2+-ATPase were studied in rabbit isolated perfused hearts. Tg and CPA were infused into the hearts for 60 min followed by 60 min of wash-out. Left-ventricular developed pressure (LVDP), left-ventricular end diastolic pressure (LVEDP) and the relaxation time constant,tau, were assessed with a fluid-filled LV intraventricular balloon. Both Tg and CPA induced a concentration-dependent reduction in LVDP and dose-dependently altered diastolic function parameters LVEDP and tau. After 60 min of perfusion, both Tg (0.01, 0.1 and 1.0 microM) and CPA (0.1, 1.0 and 10.0 microM) decreased LVDP from 98+/-1 mmHg in control to 83+/-4; 81+/-5 and 55+/-7 mmHg and to 91+/-3, 80+/-5 and 65+/-4 mmHg, respectively. LVEDP increased from 5+/-1 mmHg in controls to 6+/-0.2, 10+/-1 and 29+/-4 mmHg and to 7+/-0.2, 9+/-1 and 11+/- mmHg; while tau elevated from 28+/-1 ms to 32+/-1, 38+/-4 and 99+/-18 ms and to 34+/-1, 38+/-2 and 48+/-4 ms in Tg (0.01, 0.1 and 1.0 microM) and CPA (0.1, 1.0 and 10.0 microM), respectively. The effects of Tg were more pronounced than those of CPA and were modulated by extracellular Ca2+. With 1 mm Ca2+, both agents Tg (0.03 microM) and CPA (0.1 microM) produced a vasodilatation (81.7+/-2. 6 and 89.1+/-3.1% of pre-drug values, respectively). Pretreatment of the hearts with L-NMMA, a specific inhibitor of nitric oxide production, completely abolished the relaxing effect of Tg and CPA as well as the production of cGMP. These data show that the two SR-Ca2+ ATPase inhibitors, Tg and CPA, are negatively inotropic and lusitropic agents and that both Tg and CPA induce a vasodilatation mediated by a NO-dependent mechanism.  相似文献   

19.
Heart rate variability (HRV) (SD of the RR interval), an index of parasympathetic tone, was measured at rest and during exercise in 13 healthy older men (age 60 to 82 years) and 11 healthy young men (age 24 to 32 years) before and after 6 months of aerobic exercise training. Before exercise training, the older subjects had a 47% lower HRV at rest compared with the young subjects (31 +/- 5 ms vs 58 +/- 4 ms, p = 0.0002). During peak exercise, the older subjects had less parasympathetic withdrawal than the young subjects (-45% vs -84%, p = 0.0001). Six months of intensive aerobic exercise training increased maximum oxygen consumption by 21% in the older group and 17% in the young group (analysis of variance: overall training effect, p = 0.0001; training effect in young vs old, p = NS). Training decreased the heart rate at rest in both the older (-9 beats/min) and the young groups (-5 beats/min, before vs after, p = 0.0001). Exercise training increased HRV at rest (p = 0.009) by 68% in the older subjects (31 +/- 5 ms to 52 +/- 8 ms) and by 17% in the young subjects (58 +/- 4 ms to 68 +/- 6 ms). Exercise training increases parasympathetic tone at rest in both the healthy older and young men, which may contribute to the reduction in mortality associated with regular exercise.  相似文献   

20.
OBJECTIVES: Androgen ablation with luteinizing hormone-releasing hormone (LHRH) agonists, orchiectomy, or oral estrogens has significant untoward sexual side effects. We evaluated a combination of finasteride and flutamide as potency-sparing androgen ablative therapy (AAT) for advanced adenocarcinoma of the prostate. In addition, we evaluated whether finasteride provided additional intraprostatic androgen blockade to flutamide. METHODS: Twenty men with advanced prostate cancer were given flutamide, 250 mg orally three times daily. Serum prostate-specific antigen (PSA) values were measured weekly. At a nadir PSA value, finasteride, 5 mg orally every day, was added. PSA values were then measured weekly until a second nadir PSA value was achieved. Sexual function was evaluated at baseline, at the second nadir PSA value, and every 3 months thereafter. Testosterone, dihydrotestosterone (DHT), and dehydroepiandrostenedione (DHEA) levels were measured at baseline and at the first and second nadir PSA values. RESULTS: The median follow-up period was 16.9 months. Therapy failed in 1 patient with Stage D2 disease at 12 months, but an additional response to subsequent LHRH agonist therapy was observed. One patient developed National Cancer Institute grade 3 diarrhea and was withdrawn from the study. Seven of 20 men developed mild gynecomastia, and 3 of 20 developed mild transient liver function test elevations. Mean PSA levels were 94.6 +/- 38.2 ng/mL at baseline and 7.8 +/- 2.7 and 4.7 +/- 2.2 ng/mL at the first and second PSA nadir values, respectively (P = 0.034). Mean percent decline in PSA value from baseline was 87.0 +/- 3.1% with flutamide alone and 94.0 +/- 1.9% with both flutamide and finasteride (P = 0.001). Eleven of 20 men were potent at baseline. At the second nadir PSA value, 9 (82%) of 11 were potent, whereas 2 (18%) of 11 were impotent. With longer follow-up (median 16.4 months), 6 (55%) of 11 men were potent, 2 (18%) of 11 were partially potent, and 3 (27%) of 11 were impotent. With flutamide alone, testosterone rose a mean of 77 +/- 14.7% of baseline (P = 0.0001), DHEA fell a mean of 32.4 +/- 4.6% (P = 0.0001), and DHT was unchanged. With the addition of finasteride, testosterone rose another 14 +/- 6% (P = 0.06, not significant), DHEA was unchanged, and DHT fell a mean of 34.8 +/- 4.7% (P = 0.0009). CONCLUSIONS: Finasteride and flutamide were safe and well tolerated as AAT for advanced prostate cancer. Finasteride provided additional intraprostatic androgen blockade to flutamide, as measured by additional PSA suppression. Sexual potency was preserved initially in most patients, although there was a reduction in potency and libido in some patients on longer follow-up. Further evaluation of this therapy is needed.  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