首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 31 毫秒
1.
CGS 20267 is a new non-steroidal compound which potently inhibits aromatase in vitro (IC50 of 11.5 nM) and in vivo (ED50 of 1–3 μg/kg p.o.). CGS 20267 maximally inhibits estradiol production in vitro in LH-stimulated hamster ovarian tissue at 0.1 μM with an IC50 of 0.02 μM and does not significantly affect progesterone production up to 350 μM. In ACTH-stimulated rat adrenal tissue in vitro, aldosterone production was inhibited with an IC50 of 210 μM (10,000 times higher than the IC50 for estradiol production); no significant effect on corticosterone production was seen at 350 μM. In vivo, in ACTH-treated rats, CGS 20267 does not affect plasma levels of corticosterone or aldosterone at a dose of 4 mg/kg p.o. (1000 times higher than the ED50 for aromatase inhibition in vivo). In adult female rats, a 14-day treatment with 1 mg/kg p.o. daily, completely interrupts ovarian cyclicity and suppresses uterine weight to that seen 14 days after ovariectomy. In adult female rats bearing estrogen-dependent DMBA-induced mammary tumors, 0.1 mg/kg p.o. given daily for 42 days caused almost complete regression of tumors present at the start of treatment. Thus compared to each other, CGS 16949A and CGS 20267 are both highly potent in inhibiting estrogen biosynthesis in vitro and in vivo. The striking difference between them is that unlike CGS 16949A, CGS 20267 does not affect adrenal steroidogenesis in vitro or in vivo, at concentrations and doses several orders of magnitude higher than those required to inhibit estrogen biosynthesis.  相似文献   

2.
The in vitro metabolism of cortisol in human liver fractions is highly complex and variable. Cytosolic metabolism proceeds predominantly via A-ring reduction (to give 3,5β-tetrahydrocortisol; 3,5β-THF), while microsomal incubations generate upto 7 metabolites, including 6β-hydroxycortisol (6β-OHF), and 6β-hydroxycortisone (6β-OHE), products of the cytochrome P450 (CYP) 3A subfamily. The aim of the present study was, therefore, to examine two of the main enzymes involved in cortisol metabolism, namely, microsomal 6β-hydroxylase and cytosolic 4-ene-reductase. In particular, we wished to assess the substrate specificity of these enzymes and identify compounds with inhibitory potential. Incubations for 30 min containing [3H]cortisol, potential inhibitors, microsomal or cytosolic protein (3 mg), and co-factors were followed by radiometric HPLC analysis. The Km value for 6β-OHF and 6β-OHE formation was 15.2 ± 2.1 μM (mean ± SD; n = 4) and the Vmax value 6.43 ± 0.45 pmol/min/mg microsomal protein. The most potent inhibitor of cortisol 6β-hydroxylase was ketoconazole (Ki = 0.9 ± 0.4 μM; N = 4), followed by gestodene (Ki = 5.6 ± 0.6 μM) and cyclosporine (Ki = 6.8 ± 1.4 μM). Both betamethasone and dexamethasone produced some inhibition (Ki = 31.3 and 54.5 μ, respectively). However, substrates for CYP2C (tolbutamide), CYP2D (quinidine), and CYP1A (theophylline) were essentially non-inhibitory. The Km value for cortisol 4-ene-reductase was 26.5 ± 11.2 μM (n = 4) and the Vmax value 107.7 ± 46.0 pmol/min/mg cytosolic protein. The most potent inhibitors were androstendione (Ki = 17.8 ± 3.3 μM) and gestodene (Ki = 23.8 ± 3.8 μM). Although both compounds have identical A-rings to cortisol, and undergo reduction, inhibition was non-competitive.  相似文献   

3.
Cytokines produced by immune-activated testicular interstitial macrophages (TIMs) may play a fundamental role in the local control mechanisms of testosterone biosynthesis in Leydig cells. We investigated whether in vivo immune-activation of TIMs can modulate Leydig cell steroidogenesis. To immune activate TIMs in vivo, mice were injected intraperitoneally (i.p.) with lipopolysaccharide (LPS, 6 mg/kg). TIMs and Leydig cells were purified for RNA analysis. LPS treatment resulted in a 47-fold increase in interleukin-1β (IL-1β) mRNA in TIMs. P450c17 mRNA levels in the Leydig cells from the same animals, decreased to less than 10% compared to control. The effect of LPS on IL-1β and P450c17 mRNA levels was reversible on both TIMs and Leydig cells, respectively. To determine if the effect of LPS on P450c17 was mediated by a possible decrease in pituitary LH secretion, mice were co-injected with LPS and hCG. Treatment with hCG did not change the effect observed with LPS alone, in TIMs or in Leydig cells. In vitro, LPS treatment of TIMs resulted in marked induction of IL-1β mRNA expression. In parallel, in vitro treatment of Leydig cells with recombinant IL-1 resulted in a dose-dependent inhibition of P450c17 mRNA expression and testosterone production. These data demonstrate that LPS treatment, in vivo and in vitro, induced IL-1 gene expression in TIMs, and that IL-1 inhibits P450c17 mRNA in vitro. Therefore, we suggest that immune-activation of TIMs might have caused the observed inhibition of P450c17 gene expression in Leydig cells in vivo.  相似文献   

4.
A series of dihydroxamic acid ligands of the formula [RN(OH)C(O)]2(CH2)n, (n = 2, 4, 6, 7, 8; R = CH3, H) has been studied in 2.0 M aqueous sodium perchlorate at 25.0 °C. These ligands may be considered as synthetic analogs to the siderophore rhodotorulic acid. Acid dissociation constants (pKa) have been determined for the ligands and for N-methylacetohydroxamic acid (NMHA). The pKa1 and pKa2 values are: n = 2, R = CH3 (8.72, 9.37); N = 4, R = CH3 (8.79, 9.37); N = 6, R = CH3; N = 7, R = CH3 (8.95, 9.47); N = 8, R = CH3 (8.93, 9.45); N = 8, R = H (9.05, 9.58). Equilibrium constants for the hydrolysis of coordinated water (log K) have been estimated for the 1:1 feeric complexes of the ligands n = 2, 4, 8; R = CH3. The N = 8 ligand forms a monomeric complex with Fe(III) while the n = 2 and 4 ligands form dimeric complexes. For hydrolysis of the n = 8 monomeric complex, log K1 = −6.36 and log K2 = −9.84. Analysis of the spectrophotometric data for the dimeric complexes indicates deprotonation of all four coordinated waters. The successive hydrolysis constants, log K1–4, for the dimeric complexes are as follows: n = 2 (−6.37, −5.77, −10.73, −11.8); n = 4 (−5.54, −5.07, −11.57, −10.17). The log K2 values for the dimers are unexpectedly high, higher in fact than log K1, inconsistent with the formation of simple ternary hydroxo complexes. A scheme is proposed for the hydrolysis of the ferric dihydroxamate dimers, which includes the possible formation of μ-hydroxo and μ-oxo bridges.  相似文献   

5.
The metabolism of dihydrotachysterol (DHT), a hydrogenated analogue of vitamin D, has been studied in vivo using man and rat and in vitro using the perfused rat kidney, and hepatoma (3B) and osteosarcoma (UMR-106) cell lines. In vivo a large number of metabolites appeared in the plasma of rats given DHT2 and DHT3. Of particular interest was a compound more polar than 25-hydroxy-DHT, which has been designated compound H. Further study of this compound showed that it was composed of two components, one (Ha) being in much lower concentration than the other (Hb). The production of T2/H (peak H from DHT2) was demonstrated in human plasma after administration of oral DHT2. Comparison of the metabolites formed in vivo with those isolated from the rat kidney perfused with 25-hydroxy-DHT3 in vitro showed that 25-hydroxy-DHT3 was metabolized along two metabolic pathways previously described for vitamin D, culminating in the production of 25-hydroxy-DHT3-23,26-lactone and 23,25-dihydroxy-24-oxo-DHT3. The osteosarcoma cell line metabolized 25-OH-DHT3 in vitro along the same two metabolic pathways already demonstrated in the perfused rat kidney. More polar metabolites than compound H seen in rat plasma in vivo were shown to be metabolites of compound H and similar metabolites were also produced in the osteosarcoma cell line from chemically synthesized 1,25-dihydroxy-DHT3. The hepatoma cell line 25-hydroxylated DHT and no feed-back inhibition was observed. Use of the hepatoma cell to 25-hydroxylate a number of chemically synthesized 1-hydroxy-DHTs indicated that compound Ha was indistinguishable from 1,25-dihydroxy-DHT whereas compound Hb is possibly 1β,25-dihydroxy-DHT. Studies with the VDR in both chick gut and calf thymus indicated that 1,25-dihydroxy-DHT is very effective in displacing radiolabelled 1,25-dihydroxyvitamin-D3 and is thus most likely to be the calcaemic metabolite of DHT.  相似文献   

6.
Complexes of type A4[VO(tart)]2·nH2O, where A = Rb or Cs and tart =d,l-tartrate(4−) (n = 2) or d,d-tartrate(4−) (n = 2 for Rb and n = 3 for Cs), were prepared from an aqueous mixture of V2O5, AOH and H4tart. These complexes were studied by single-crystal X-ray diffraction methods: Rb4[VO(d,l-tart)]2·2H2O, space group P1 with a = 8.156(1),b = 8.246(1),c = 8.719(1)Å, = 66.09(1)°, β = 65.07(1)°, γ = 82.40(1)°,Z = 2, 1917 observed reflections, and final Rw = 0.035; Cs4[VO(d,l-tart)]2·2H2O, space group P21/c with a = 9.350(1),b = 13.728(2),c = 8.479(1)Å, β = 106.77(1)°,Z = 4, 2235 observed reflections, and final Rw = 0.054; Rb4[VO(d,d-tart)]2·2H2O, space group P4122 with a = 8.072(1),c = 32.006(3)Å,Z = 8, 1014 observed reflections and final Rw = 0.038; Cs4[VO(d,d-tart)]2·3H2O, space group P122 with a = 8.184(1),c = 33.680(5)Å,Z = 8, 1310 observed reflections, and final Rw = 0.063. Bulk magnetic susceptibility data (1.5–300 K) for these compounds and A4[VOl,l-tart)]2·nH2O (A = Rb, Cs) were obtained on polycrystalline samples. These data were analyzed in terms of a Van Vleck exchange coupled S = 1/2 model which was modified to include an interdimer exchange parameters Θ. Analysis of the low-temperature (1.5–20 K) susceptibility data gave 2J = +1.30 cm−1 and Θ = −1.86 K for Rb4[VO(d,l-tart)]2·2H2O, 2J = +1.16 cm−1 and Θ = −1.69 K for Cs4[VO(d,l-tart)]2·2H2O, 2J = +1.90 cm−1 and Θ = −0.82 K for Rb4[VO(d,d-tart)]2·2H2O, 2J = +2.04 cm−1 and Θ = −0.80 K for Rb4[VO(l,l-tart)]2·2H2O, 2J = +1.52 cm−1 and Θ = −0.25 K for Cs4[VO(d,d-tart)]2·3H2O, and 2J = +1.64 cm−1 and Θ = −0.31 K for Cs4[VO(l,l-tart)]2·3H2O. These results suggest the magnitudes of intradimer (ferromagnetic and interdimer (antiferromagnetic) exchange interactions are similar in these complexes, as observed for the analogous Na salts.  相似文献   

7.
Substituted 1-[(benzofuran-2-yl)-phenylmethyl]-imidazoles are a new class of potent aromatase inhibitor with in vitro IC50 values < 10 nM for certain members using human placental enzyme. At a dose of 2 mg/kg in PMSG-stimulated rats, selected compounds effectively reduce the oestradiol levels by 82–98%.  相似文献   

8.
J. B. Thomas  H. H. Nijhuis 《BBA》1968,153(4):868-877
The time course of aerobic photobleaching of various chlorophyll-protein complexes in vivo at high light intensities was studied with isolated Aspidistra elatior chloroplasts.

1. 1. Ca680 bleaching starts with the onset of irradiation and, initially, proceeds linearly with time. Washing the chloroplasts causes a nearly constant increase of the bleaching rate throughout the experiment.

2. 2. Ca670 does not appreciably, if at all, bleach initially; subsequently, bleaching proceeds linearly with time and at a slightly higher rate than that for Ca680. Washing makes Ca670 bleach concomitantly with the onset of illumination, and at a nearly constant rate.

3. 3. Bleaching at 665 nm is likely to start only after a relatively long period of illumination. Washing shows no effects during this period. Once bleaching has started, washing causes its rate to increase.

4. 4. No indication of the occurrence of “short-wave” chlorophyll a forms other than Ca670 and Ca665 was obtained.

5. 5. Cb bleaching starts concomitantly with illumination at a low rate. The rate increases more or less exponentially with time. Washing enhances bleaching in two steps.

6. 6. The importance of the results is discussed.

Abbreviations: Ca700,Ca695, Ca680, Ca670, Ca665, chlorophyll a-protein complexes in vivo with absorption maxima around 700, 695, 680, 670, and 665 nm, respectively; Cb; chlorophyll b-protein complex in vivo

Abbreviations: DCIP, 2,6-dichlorophenolindophenol  相似文献   


9.
J. B. Thomas  F. Bretschneider 《BBA》1970,205(3):390-400
1. The absorption spectrum of chlorophyll b in vivo at 77°K is presented as the difference spectrum between preparations of spinach and chlorophyll b-free Vischeria stellata chloroplasts.

2. A shoulder on this spectrum around 662 nm is due to a component different from chlorophyll b. This component may well be identical with the chlorophyll a form, chlorophyll a (665).

3. The 77°K chlorophyll b absorption spectra in the nonfractionated photosyn-thetic pigment apparatus and in fractions mainly representing Photosystems 1 or 2 are not significantly different.

4. The aerobic irreversible photobleaching of chlorophyll b was studied in the intact pigment complex as well as in fractions mainly consisting of Photosystem 1 or 2. A two-step photobleaching was observed in all cases. The time-course of this bleaching was not significantly different for chlorophyll b in both fractions.

5. These results do not indicate that more than a single chlorophyll b complex occurs in vivo.  相似文献   


10.
目的: 自编程定量分析不同电刺激方式对在体和离体蛙腓肠肌单收缩的影响。方法: 实验分为四个组:间接刺激在体标本组(n=12),直接刺激在体标本组(n=8),间接刺激离体标本组(n=12),直接刺激离体标本组(n=8)。分别制备在体和离体蛙腓肠肌标本, 施以间接电刺激(经坐骨神经)或直接电刺激(细针灸针刺激电极直接刺入腓肠肌中):强度从0 V开始,周期3 s,增量0.02 V,刺激150次;用生物机能实验系统(BL-420F)实时记录不同强度刺激对肌肉收缩的影响。后续通过自编程辅助处理分析肌肉收缩数据,对单收缩特征参数进行定量比较分析。 结果: ①对在体标本,与直接刺激比较:间接刺激的阈强度、半高强度和最适强度均更小(P<0.05);最大单收缩幅度更大,收缩期更长,上升斜率更小(P<0.05)。②对离体标本,与直接刺激比较:间接刺激的阈强度、半高强度和最适强度也均更小(P<0.05);最大单收缩幅度更大,收缩期更长,上升斜率更小(P<0.05)。③均采用间接刺激或直接刺激时,与离体标本比较,在体标本单收缩的各项参数均无差异(P>0.05)。 结论: 不论使用间接刺激或是直接刺激,在体标本和离体标本的单收缩功能特点均无显著差异; 但间接刺激比直接刺激更容易触发腓肠肌产生单收缩,且单收缩幅度更高。  相似文献   

11.
1. 1. The development of thermotolerance has been shown to protect blowfly flight muscle mitochondrial function from damage resulting from an LD50 in vivo heat dose.
2. 2. The principal sites of the damage have been studied using specific inhibitors of the respiratory chain, rotenone and antimycin A, together with substrates that stimulate respiration through the different complexes.
3. 3. Complex I was identified as the primary site for heat damage. State III respiration was inhibited following the LD50 in vivo heat dose, and uncoupling with FCCP did not restore respiration to control levels, indicating that the respiratory enzymes were inactivated. The development of thermotolerance protected this site from heat damage.
4. 4. In contrast, G3-P stimulated respiration was the same in control, LD50 in vivo treated controls and LD50, in vivo treated thermotolerant mitochondria, and significantly higher than state III respiration of LD50 in vivo treated controls. This suggested that respiration through G3-P dehydrogenase, Co enzyme Q and Complex III is not damaged. However, as G3-P stimulated respiration of coupled mitochondria from LD50 in-vivo treated flies was markedly reduced (El-Wadawi and Bowler, 1995. J. exp. Biol. 198: 2413–2421), phosphorylation at complex III may be inhibited also.
5. 5. Ferrocyanide stimulated respiration through cytochrome c-Complex IV was also inhibited in LD50 in vivo treated flies, as compared with unheated control mitochondria. However, thermotolerance protected this site also from heat damage.
  相似文献   

12.
13.
Esenbeckia febrifuga (Rutaceae) is a plant traditionally used to treat malaria in the Brazilian Amazon region. Ethanol extract of stems displayed a good antiplasmodial activity against Plasmodium falciparum strains W-2 (IC50 15.5±0.71 μg/ml) and 3 D7 (IC50 21.0±1.4 μg/ml). Two coumarins (bergaptene 1 and isopimpinellin 2), five alkaloids (flindersiamine 3, kokusaginine 4, skimmiamine 5, γ-fagarine 6 and 1-hydroxy-3-methoxy-N-methylacridone, 7), besides a limonoid (rutaevine 8), have been isolated for the first time from this species. Antiplasmodial activity of compounds 3, 5–8 has been evaluated in vitro against P. falciparum strains (W-2 and 3D7) and the furoquinolines 5 and 6 were the most potent displaying IC50 values <50 μg/ml; flindersiamine (3) showed a weak activity while alkaloid 7 and rutaevine (8) were inactive (IC50>100 μg/ml).  相似文献   

14.
Aldosterone was isolated from hamster adrenal cells and was identified by high performance liquid chromatography and thermospray mass spectroscopy analysis. Basal outputs from adrenal cell suspensions were of the same order of magnitude, 8.4 ± 1.9 ng and 8.0 ± 0.7 ng/2 h/50,000 cells, for aldosterone and corticosteroid, respectively. The outputs of aldosterone and corticosteroid increased with K+ concentrations to reach maxima of 3.3- and 1.6-fold at 10 meq/l of K+. AngiotensinII (AII) produced dose-dependent increases in aldosterone and corticosteroid outputs with maxima of 3- and 4-fold, respectively. In contrast, ACTH induced relatively no changes in aldosterone output, whereas dose-dependent increases in corticosteroid output were found. In time study experiments, with 10−8 M AII, aldosterone and corticosteroid outputs were maximally increased after 1 h (6-fold) and 3 h (1.8-fold), respectively. At 10−8 M, ACTH had a small stimulatory effect on aldosterone output after 6 h, whereas it provoked a gradual increase in corticosteroid output (up to 7-fold after 8 h of incubation). The effects of AII and ACTH on adrenal cytochrome P-45011β involved in the last steps of aldosterone formation were evaluated by c combined in vivo andin vitro experiments. The P-45011β mRNA level was increased by a low sodium intake but not by a 24 h ACTH stimulus. These results taken together indicate that ACTH and AII differentially regulate P-45011β. It is postulated that these two regulatory peptides regulate the hamster adrenal steroidogenesis by different P-450 genes.  相似文献   

15.
The crystal structures of Li[Fe(trtda)]·3H2O and Na[Fe(eddda)]·5H2O (trtda = trimethylenediaminetetraacetate and eddda = ethylenediamine-N,N′-diacetate-N,N′-di-3-propionate) have been determined by single crystal X-ray diffraction techniques. The former crystal was monoclinic with the space group P21/n,a = 17.775(3),b = 10.261(1),c = 8.883(2)Å, β = 95.86(4)° and Z = 4. The latter was also monoclinic with the space group P21/n,a = 6.894(2),b = 20.710(6),c = 13.966(3)Å, β = 101.44(2)° and Z = 4. Both complex anions were found to adopt an octahedral six-coordinated structure with all of six ligand atoms of trdta4− or eddda4− coordinated to the Fe(III) ion, unlike the corresponding edta4− complex which is usually seven-coordinate with the seventh coordination site occupied by H2O. Of the three geometrical isomers possible for the eddda complex, the trans(O5) isomer was actually found in the latter crystal. Factors determining the structural types of metal–edta complexes are discussed in detail.  相似文献   

16.
Cobalt(III) complexes with a thiolate or thioether ligand, t-[Co(mp)(tren)]+ (2), t-[Co(mtp)(tren)]2+ (1Me) and t-[Co(mta)(tren)]2+ (2Me), (mp = 3-mercaptopropionate, MA = 3-(methylthio)propionate and MTA = 2-(methylthio)acetate) have been prepared in aqueous solutions. The crystal structures of 1, 2, 1Me and 2Me were determined by X-ray diffraction methods. The crystal data are as follows, t-[Co(mp)(tren)]ClO4 (1CIO4): monoclinic, P21/n, A = 10.877(8), B = 11.570(4), c = 12.173(7) Å, β = 92.20(5)°, V = 1531(1) Å3, Z = 4 and R = 0.060; t-[Co(ma)(tren)]Cl·3H2O (2Cl·3H2O): monoclinic, P21/n, a = 7.7688(8), B = 27.128(2), C = 7.858(1) Å, β = 100.63(1)°, V = 1627.7(3) Å3, Z = 4 and R = 0.066; (+)465CD-t-[Co(mtp)(tren)](ClO4)2 ((+)465CD-1Me(ClO4)2): orthorhombic, P212121, A = 10.6610(7), B = 11.746(1), C = 15.555(1) Å, V = 1947.9(3) Å3, Z = 4 and R = 0.068; (+)465CD-t-[Co(mta)(tren)](ClO4)2 ((+)465CD-2Me(ClO4)2): orthorhombic, P212121, a = 10.564(1), B = 11.375(1), C = 15.434(2) Å, V = 1854.7(4) Å3, Z = 4 and R = 0.047. All central Co(III) atoms have approximately octahedral geometry, coordinated by four N, one O, and one S atoms. All of the complexes are only isomer, of which the sulfur atom in the didentate-O,S ligands are located at the trans position to the tertiary amine nitrogen atom of tren. 1 and 1Me contain six-membered chelate ring, and 2 and 2Me do five-membered chelate ring in the didentate ligand. The chirality of the asymmetric sulfur donor atom in (+)465CD-1Me is the S configuration and that in (+)465CD-2Me is the R one. The 1H NMR, 13C NMR and electronic absorption spectral behaviors and electrochemical properties of the present complexes are discussed in relation to their stereochemistries.  相似文献   

17.
Two new spin-crossover complexes, [Fe(Medpq)(py)2(NCS)2] · py · 0.5H2O (1) and [Fe(Medpq)(py)2(NCSe)2] · py (2) (Medpq = 2-methyldipyrido[3,2-f:2′,3′-h]-quinoxaline, py = pyridine), have been synthesized. The crystal structures were determined at both room temperature (298 K) and low temperature (110 K). Complexes 1 and 2 crystallize in the orthorhombic space group Pbca and monoclinic space group P21/n, respectively. In both complexes, the distorted [FeN6] octahedron is formed by six nitrogen atoms from Medpq, the trans pyridine molecules and the cis NCX groups. The thermal spin transition is accompanied by the shortening of the mean Fe–N distances by 0.194 Å for 2. The mononuclear [Fe(Medpq)(py)2(NCS)2] and [Fe(Medpq)(py)2(NCSe)2] neutral species interact each other via π-stacking, resulting in a one-dimensional extended structure for both 1 and 2. There exist C–HX (X = S, Se) hydrogen bonds for both complexes. Variable-temperature magnetic susceptibility measurements and Mössbauer spectroscopy reveal the occurrence of a gradual spin transition. The transitions are centered at T1/2 = 120 K for 1 and T1/2 = 180 K for 2, respectively.  相似文献   

18.
An improved high-performance liquid chromatographic (HPLC) method using electrochemical detection (ED) is described capable of routinely measuring the low levels of acetylcholine (ACh) typically found in rat brain microdialysis samples. Microdialysis was performed in the striatum of the urethane anesthetized rat using a 4-mm membrane length, high recovery (40% at 1.0 μl/min; ambient conditions), loop-design probe perfused with an artificial cerebrospinal fluid (aCSF) solution containing physiologically normal calcium levels (1.2 mM). The HPLC method utilizes a polymeric stationary phase to resolve choline (Ch) from ACh. These analytes are then converted to hydrogen peroxide (H2O2) by a solid-phase reactor (containing immobilized choline oxidase and acetylcholinesterase enzymes). The H2O2 is detected amperometrically and quantitated on a platinum (Pt) working electrode (+300 mV; with a unique analytical cell featuring a solid-state palladium reference electrode). Two designs of the Pt working electrode were examined, differing only in the support material used (Kel-F or PEEK). The Kel-F/Pt electrode had a limit of detection (LOD) for both analytes of <30 fmol per 10 μl with a signal-to-noise ratio of 3:1. Striatal microdialysis perfusates were monitored for ACh and Ch over a 0–1000 nM range of neostigmine (NEO) in the CSF perfusion medium. Using the 4-mm probe, basal ACh and Ch levels were detected with a NEO level as low as 10 nM and were found to be 37 ± 3 fmol and 22 ± 1 pmol per 10 μl (mean ± S.E.M., n = 6 replicates) respectively. In similar experiments using 3-mm concentric probes comparable (lower) levels of ACh were found with the 50 and 1000 nM NEO doses (n = 4–21 animals). ACh could not be reliably quantitated when animals were perfused with the 10 nM dose of NEO (n = 4). The PEEK/Pt electrode had an improved LOD of < 20 fmol per 10 μl due to a two- to three-fold decrease in the background noise component. Basal striatal levels of ACh in the absence of NEO approached the LOD and were found to be 15 ± 2 fmol per 10 μl; Ch was 5 ± 1 pmol per 10 μl (n = 2, mean of five basal samples). The analytical system requires very little maintenance; a simple electrochemical electrode cleaning step eliminates the need for routine polishing of the Pt electrode and the mobile phase is stable for up to one week. Both ACh and Ch are resolved in under 7 min making this method highly suitable for analysis of microdialysis samples.  相似文献   

19.
In order to better understand the function of aromatase, we carried out kinetic analyses to asses the ability of natural estrogens, estrone (E1), estradiol (E2), 16-OHE1, and estriol (E3), to inhibit aromatization. Human placental microsomes (50 μg protein) were incubated for 5 min at 37°C with [1β-3H]testosterone (1.24 × 103 dpm 3H/ng, 35–150 nM) or [1β-3H,4-14C]androstenedione (3.05 × 103 dpm 3H/ng, 3H/14C = 19.3, 7–65 nM) as substrate in the presence of NADPH, with and without natural estrogens as putative inhibitors. Aromatase activity was assessed by tritium released to water from the 1β-position of the substrates. Natural estrogens showed competitive product inhibition against androgen aromatization. The Ki of E1, E2, 16-OHE1, and E3 for testosterone aromatization was 1.5, 2.2, 95, and 162 μM, respectively, where the Km of aromatase was 61.8 ± 2.0 nM (n = 5) for testosterone. The Ki of E1, E2, 16-OHE1, and E3 for androstenedione aromatization was 10.6, 5.5, 252, and 1182 μM, respectively, where the Km of aromatase was 35.4 ± 4.1 nM (n = 4) for androstenedione. These results show that estrogens inhibit the process of andrigen aromatization and indicate that natural estrogens regulate their own synthesis by the product inhibition mechanism in vivo. Since natural estrogens bind to the active site of human placental aromatase P-450 complex as competitive inhibitors, natural estrogens might be further metabolized by aromatase. This suggests that human placental estrogen 2-hydroxylase activity is catalyzed by the active site of aromatase cytochrome P-450 and also agrees with the fact that the level of catecholestrogens in maternal plasma increases during pregnancy. The relative affinities and concentration of androgens and estrogens would control estrogen and catecholestrogen biosynthesis by aromatase.  相似文献   

20.
Exposure of endothelium to a nominally uniform flow field in vivo and in vitrofrequently results in a heterogeneous distribution of individual cell responses. Extremes in response levels are often noted in neighboring cells. Such variations are important for the spatial interpretation of vascular responses to flow and for an understanding of mechanotransduction mechanisms at the level of single cells. We propose that variations of local forces defined by the cell surface geometry contribute to these differences. Atomic force microscopy measurements of cell surface topography in living endothelium both in vitro and in situ combined with computational fluid dynamics demonstrated large cell-to-cell variations in the distribution of flow-generated shear stresses at the endothelial luminal surface. The distribution of forces throughout the surface of individual cells of the monolayer was also found to vary considerably and to be defined by the surface geometry. We conclude that the endothelial three-dimensional surface geometry defines the detailed distribution of shear stresses and gradients at the single cell level, and that there are large variations in force magnitude and distribution between neighboring cells. The measurements support a topographic basis for differential endothelial responses to flow observed in vivo and in vitro. Included in these studies are the first preliminary measurements of the living endothelial cell surface in an intact artery.  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