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.
Semiempirical molecular orbital calculations (AM1) were used to model several possible reaction mechanisms for the third oxidation of the aromatase-catalyzed conversion of androgens to estrogens. The reaction mechanisms considered are based on the assumption that the third oxidation is initiated by 1 beta-hydrogen atom abstraction. Homolytic cleavage of the C10-C19 bond was modeled for both the 3-keto and 2-en-3-ol forms of the androgen 1-radicals. The addition of a protein nucleophile to the 19-oxo intermediate was also considered, and -OCH3, -SCH3, and -NHCH3 were used to represent the Ser, Cys, and Lys adducts. The transition states were estimated and optimized from the reaction coordinates obtained by constraining and increasing the C10-C19 bond lengths. The enthalpies of activation range from 14 to 21 kcal and are approximately 2 kcal lower for cleavage of the enol form. Given the tendency for AM1 to overestimate activation energies, all reactions may be energetically accessible. Other reactions modeled include a homolytic cleavage reaction from a thioether radical cation and the direct additions of oxygen radical compounds to the carbonyl of the 1-radical-2-en-3-ol-19-oxo androgen. A mechanism is proposed in which the 19-oxo intermediate is subject to initial nucleophilic attack by the protein. Since rotation of the 19-carbonyl can bring the oxygen within 2.1 A of the 2 beta-hydrogen, the formation of a tetrahedral intermediate can occur with concomitant removal of the 2 beta-proton. Enolization activates the C1-position for hydrogen atom abstraction, since the resulting radical is resonance stabilized.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

2.
The article summarizes the results of recent studies on the metabolism of 10-ethylestr-4-ene-3,17-dione, 10-[(1R)-1-hydroxyethyl]-,and 10-[(1S)-1-hydroxyethyl]estr-4-ene-3, 17-dione, in placenta. These compounds are the 19-methyl analogs of androstenedione, 19-hydroxyandrostenedione, and 19-oxoandrostenedione, respectively. No conversion of 10-ethylestr-4-ene-3,17-dione to either estrogens or oxygenated metabolites was detected. Both 10-[(1R)-1-hydroxyethyl]- and 10-[(1S)-1-hydroxyethyl]estr-4-ene-3, 17-dione were oxygenated to 10-(1,1-dihydroxyethyl)estr-4-ene-3,17-dione and isolated following in situ dehydration as 10-acetylestr-4-ene-3,17-dione. Evidence for the involvement of aromatase in these conversions is discussed. No conversion of 10-acetylestr-4-ene-3,17-dione to either estrogens or other oxygenated products was detected. These results lead us to propose a new mechanism for the third aromatase monooxygenation. We propose that the third oxygenation is initiated by 1β-hydrogen abstraction at C1 of 19,19-dihydroxyandrostenedione, followed by homolytic cleavage of the C10−C19 bond with concurrent formation of a Δ1(10),4−3-ketosteroid and a C19 carbon radical, and terminated by oxygen rebound at C19.  相似文献   

3.
To explore a stereochemistry of hydrogen removal at C-1 of the powerful aromatase inhibitor 2-methyleneandrostenedione (1), of which the A-ring conformation is markedly different from that of the natural substrate androstenedione (AD), in the course of the aromatase-catalyzed A-ring aromatization producing 2-methylestrone (2), we synthesized [1-2H]labeled steroid 1 and its [1β-2H]stereoisomer, and the metabolic fate of the C-1 deuterium in aromatization was analyzed by gas chromatography–mass spectrometry (GC–MS) in each. Parallel experiments with the natural substrates [1-2H] and [1β-2H]ADs were also carried out. The GC–MS analysis indicated that 2-methyl estrogen 2 produced from [1-2H]labeled substrate 1 retained completely the 1-deuterium (1β-H elimination), while product 2 obtained from [1β-2H]isomer 1 lost completely the 1β-deuterium. Stereospecific 1β-hydrogen elimination was also observed in the parallel experiments with the labeled ADs as established previously. The results indicate that biochemical aromatization of the 2-methylene steroid 1 proceeds through the 1β-hydrogen removal concomitant with cleavage of the C10–C19 bond, yielding 1(10),4-dienone 9, in a similar manner to that involved in AD aromatization. This would give additional evidence for the stereomechanisms for the last step of aromatization of AD, requiring the stereospecific 1β-hydrogen abstraction and cleavage of the C10–C19 bond, and for the enolization of a carbonyl group at C-3 in the A-ring aromatization.  相似文献   

4.
C L Kuo  G M Raner  A D Vaz  M J Coon 《Biochemistry》1999,38(32):10511-10518
Aldehydes are known to inactivate cytochrome P450 in the reconstituted enzyme system containing NADPH and NADPH-cytochrome P450 reductase under aerobic conditions in a mechanism-based reaction involving heme adduct formation [Raner, G. M., Chiang, E. W. , Vaz, A. D. N., and Coon, M. J. (1997) Biochemistry 36, 4895-4902]. In the study presented here, artificial oxidants were used to examine the mechanism of aldehyde activation by purified P450 2B4 in the absence of the usual O(2)-reducing system, and the adducts that were formed were isolated and characterized. With hydrogen peroxide as the oxidant, 3-phenylpropionaldehyde gives an adduct with a mass corresponding to that of native heme modified by a phenylethyl group, presumably arising from the reaction of a peroxy-iron species with the aldehyde to give a peroxyhemiacetal, which upon deformylation yields the alkyl radical. NMR analysis indicated that the substitution is specifically at the gamma-meso position. In contrast, with m-chloroperbenzoic acid as the oxidant, an adduct is formed from 3-phenylpropionaldehyde with a mass that is consistent with the addition of a phenylpropionyl group, apparently arising by hydrogen abstraction from the aldehyde to give the carbonyl carbon radical. m-Chloroperbenzoic acid by itself forms a heme adduct with a mass corresponding to the addition of a chlorobenzoyloxy group apparently derived from homolytic oxygen-oxygen bond cleavage. These and other results with nonanal and 2-trans-nonenal support the concept that this versatile enzyme utilizes discrete oxidizing species in heme adduct formation from aldehydes.  相似文献   

5.
Mechanical treatment of solid K2PtX6 (X=Cl, Br) salts under air or argon leads to the formation of paramagnetic platinum(III) complexes via homolytic cleavage of Pt---X bonds. Lewis acid sites (LASs) were also detected on the surface of mechanically activated K2PtCl6 using a paramagnetic probe. The latter species can be attributed to coordinatively unsaturated platinum(IV) complexes formed as a result of heterolysis of Pt---Cl bonds. Mechanical treatment of solid K2PtCl4, on the contrary, does not lead to homolytic Pt---Cl bond cleavage. In this case only heterolysis of the Pt---Cl bond takes place, leading to the formation of coordinatively unsaturated platinum(II) complexes.  相似文献   

6.
Two Bacillus strains were isolated from the foregut of the water beetle Agabus affinis (Payk.) and tested for their steroid transforming ability. After incubation with androst-4-en-3,17-dione (AD), 13 different transformation products were detected. AD was hydroxylated at C6, C7, C11 and C14, resulting in formation of 6β-, 7-, 11- and 14-hydroxy-AD. One strain also produced small amounts of 6β,14-dihydroxy-AD. Partly, the 6β-hydroxy group was further oxidized to the corresponding 6-oxo steroids. In addition, a specific reduction of the Δ4-double bond was observed, leading to the formation of 5-androstane derivatives. In minor yields the carbonyl functions at C3 and C17 were reduced leading to the formation of 3ξ-OH or 17β-OH steroids. EI mass spectra of the trimethylsilyl and O-methyloxime trimethylsilyl ether derivatives of some transformation products are presented for the first time.  相似文献   

7.
Adult whiteflies are characterized by the presence of copious amounts of wax particles covering all surfaces of the body except the eyes. The lipid composition was determined for wax particles removed from the surfaces of the sweetpotato whitefly, Bemisia tabaci (Gennadius), and the greenhouse whitefly, Trialeurodes vaporariorum (Westwood). The lipid components in the wax particles of both species were mostly mixtures of long-chain aldehydes and long-chain primary alcohols. The major wax particle components for B. tabaci were C34 aldehyde and C34 alcohol and small amounts of C32 aldhyde and alcohol. For the wax particles from T. vaporariorum, C32 aldehyde and C32 alcohol were the major components with lesser amounts of the C30 components. These findings were compared to the surface lipids of fully-waxed B. tabaci and T. vaporariorum adults that contained, in addition to the major amounts of long-chain aldehydes and alcohols, quantities of long-chain wax esters. Wax esters were not present in lipid extracts from the surface of B. tabaci whiteflies at the time of adult emergence (prior to deposition of wax particles). Thus, the appearance of wax esters on the cuticular surfaces occurred during the period of deposition of wax particles. The quantities of wax esters in the surface lipid extracts of wing tissues separated from the bodies of adult whiteflies indicated that the wing surfaces were a major site of wax ester deposition.  相似文献   

8.
Aromatase is a cytochrome P-450 enzyme that catalyzes the conversion of androgens into oestrogens via sequential oxidations at the 19-methyl group. Despite intensive investigation, the mechanism of the third step, conversion of the 19-aldehydes into oestrogens, has remained unsolved. We have previously found that a pre-enolized 19-al derivative undergoes smooth aromatization in non-enzymic model studies, but the role of enolization by the enzyme in transformations of 19-oxoandrogens has not been previously investigated. The compounds 19-oxo[2 beta-2H]testosterone and 19-oxo[2 beta-2H]androstenedione have now been synthesized. Exposure of either of these compounds to microsomal aromatase, in the absence of NADPH, for an extended period led to no significant 2H loss or epimerization at C-2, leaving open the importance of an active-site base. However, in the presence of NADPH there was an unexpected substrate-dependent difference in the stereoselectivity of H loss at C-2 in the enzyme-induced aromatization of 19-oxo[2 beta-2H]-testosterone versus 19-oxo[2 beta-2H]androstenedione. The aromatization results for 17 beta-ol derivative 19-oxo[2 beta-2H]-testosterone correspond to about 1.2:1 2 beta-H/2 alpha-H loss from unlabelled 19-oxotestosterone. In contrast, aromatization results for 19-oxo[2 beta-2H]androstenedione correspond to at least 11:1 2 beta-H/2 alpha-H loss from unlabelled 19-oxoandrostenedione. This substrate-dependent stereoselectivity implies a direct role for an enzyme active-site base in 2-H removal. Furthermore, these results argue against the proposal that 2 beta-hydroxylation is the obligatory third step in aromatase action.  相似文献   

9.
Hexacoordination of the neutral phosphorus compounds 4–6 is evidenced by their high field 31P NMR chemical shifts and is further substantiated by the crystal structure of 5 and 6.5 contains the potentially bis-chelating ligand Ar = (C6H3(CH2NMe2)2-2,6) and 6 the same ligand with a protonated amino group. In both cases the compounds exhibit slightly distorted octahedral geometry. In compound 5, only one NMe2 group is coordinated to the phosphorus atom with an N → P bond of 2.063 Å. In compound 6, the NMe2 group is coordinated to the phosphorus atom with an N → P bond of 2.007 Å while the dimethylammonium substituent is pointing away from the phosphorus atom forming a hydrogen bridge with two oxygen atoms. The fluxional behavior of these three novel six-coordinate phosphorus compounds was studied by dynamic 1H NMR spectroscopy.  相似文献   

10.
The reactions of complex (C5Me5)Ir(Cl) (CO) (Me) (1a) with cyclohexylisocyanide and phosphines (L=CyNC, PHPh2, PMePh2, PMe2Ph) give the products of alkyl migratory insertion (C5Me5Ir(Cl) (COMe) (L), in toluence or tetrahydrofuran at 323 K or higher temperature. The phenyl analogue (C5Me5)Ir(Cl)(CO)(Ph) or the iodide complexes (C5Me5)Ir(I) (CO) (R) (R=Me, Ph_are not reactive under the same conditions. The reaction of (C5Me5)Ir(Cl)(CO)(Me) with PMePh2 and PMe2Ph in acetonitrile yields the chloride substitution product [(C5Me5)Ir(CO)(L)(Me)]+Cl. Kinetic measurements for the reactions of (C5Me5)Ir(Cl)(CO)(Me) in toluene are first order in the iridium complex and exhibit a saturation dependence on the incoming donors L. Analysis of the data suggests a two-step process involving (i) rapid formation of a molecular complex [(C5Me5)Ir(Cl)(CO)(Me), (L)], in which the structure of 1a is unperturbed within the limits of spectroscopic analysis, and (ii) rate determining methyl migration. The reaction parameters are K for the pre-equilibrium step (K = 1.5 (CyNC), 7.3 (PHPh2), 7.1 (PMePh2) dm3 mol−1 at 323 K) and k2 for the slow carbon---carbon bond formation (k2 (105) = 6.9 (CyNC), 1.2 (PHPh2), 1.0 (PMePh2) s−1 at 323 K). The activation parameters for the methyl migration step in the reaction with PMePh2 obtained between 308 and 338 K, are ΔH = 106±16 kJ mol−1 and ΔS = − 14±5 J K−1 mol−1. The reaction of 1a with PMePh2 proceeds at similar rates in tetrahydrofuran (K = 3.7 dm3 mol−1, k2 (105) = 1.2 s−1, 323 K). The crystal structure of (C5Me5)Ir(Cl)(COMe) (PMe2Ph) has been determined by X-ray diffraction. C20H29ClOPIr: Mr = 544.1, monoclinic, P21/n, A = 8.084 (2), B = 9.030(2), C = 28.715 (3) Å, β = 91.41 (3)°, Z = 4, Dc = 1.71 g cm−3, V = 2095.5 Å3, room temperatyre, Mo K, γ = 0.71069, μ = 65.55 cm−1, F(000) = 1044, R = 0.037 for 2453 independent observed reflections. The complex shows a deformed tetrahedral coordination assuming the η5-C5Me5 molecular fragment as a single coordination site. The iridium-chlorine bond is staggered with respect to two adjacent C(ring)-methyl bonds, while the Ir---P and the Ir---COMe bonds are eclipsed with respect to C(ring)-methyl bonds.  相似文献   

11.
Qualitative estimates of the relative stability of hypothetical heterofullerenes C55Y5 (Y=Si, Ge, Sn, B, Al, N, P, SiH, GeH, SnH) and some η5-π-complexes LiC55Y5 were carried out by the MNDO method. Atoms Y (or groups XH) are assumed to substitute those C atoms in fullerene C60 which are located at the -positions of a separated pentagonal face (pent*) of this polyhedral molecule. It is shown that the spin densities in radicals C55Y5 (Y=SiH, GeH, SnH, B, Al, N, P) are localized on the separated pentagon atoms and the Li-pentagonal face (Li-pent*) bonds in η5-π-complexes of these radicals with the Li atom are considerably stronger than Li-pent* bonds in complexes [η5-π-LiC60]+ and [η5-π-LiC60] of unsubstituted C60. In addition, it is established that the Li-pent* bond energies in η5-π-complexes LiC55B5 and LiC55Al5 exceed the energy of the Li-pent* bond in the η5-π-complex LiC60H5 studied earlier. In contrast, the energies of similar bonds for Y=N, P are close to the energy of the Li-pent* bond in the η5-π-complex LiC60H5.  相似文献   

12.
The characteristic surface lipid compositions of several C3 and C4 plants are discussed. C4 plants produce surface lipids (epicuticular waxes) made up of the ubiquitous classes of aliphatic compounds: free fatty acids, aldehydes, primary alcohols, alkanes and aliphatic linear esters. C3 plants synthesize surface lipids comprising the ubiquitous classes and either of the two following groups of compound: (i) lβ-diketones, hydroxy lβ-diketones, alkan-2-ol esters; (il) ketones and secondary alcohols with the functional group in the middle of the hydrocarbon chain. These features are suggested to represent physioIogical characteristics of the plant and to be related to ecological adaptations. Wax class compositions might also be an ancillary method for defining the C3 or C4 mechanism of CO2 assimilation in cases where uncertainty exists.  相似文献   

13.
Highly purified liver microsomal cytochrome P-450 acts as a peroxygenase in catalyzing the reaction, RH+ XOOH→ROH+XOH, Where RH represents any of a large variety of foreign or physiological substrates and ROH the corresponding product, and XOOH represents any of a series of peroxy compounds such as hydroperoxides or peracids serving as the oxygen donor and XOH the resulting alcohol or acid. Several experimental approaches in this and other laboratories have yielded results compatible with a homolytic mechanism of oxygen-oxygen bond cleavage but not with the heterolytic formation of a common iron-oxo intermediate from the various peroxides. Recently, we have found a new reaction, catalyzed by the reconstituted system containing the phenobarbital-inducible form of P-450, which catalyzes the reductive cleavage of hydroperoxides: XRR’C-OOH+ NADPH+H+→ XR’CO + R’H+H2O + NADP+ Thus, cumyl hydroperoxide yields acetophenone and methane, and 13-hydroperoxyoctadeca-9, 11-dienoic acid yields pentane and an as yet unidentified additional product. Since hydroperoxide reduction does not produce the corresponding alcohol, it is concluded that homolytic cleavage of the oxygen-oxygen bond occurs with rearrangement of the resulting alkoxy radical. Studies are in progress to determine how broad a role the new hydroperoxide cleavage reaction plays in the biological peroxidation of lipids.  相似文献   

14.
A sensitive assay of aromatization of 16 alpha-hydroxylated androgens, 16 alpha-hydroxyandrostenedione (16 alpha-OHA), 16 alpha,19-dihydroxyandrostenedione [16 alpha,19-(OH)2A], and 16 alpha-hydroxy-19-oxo androstenedione (16 alpha-OH-19-oxo A), was developed using reversed phase high-performance liquid chromatography with a coulometric detector. The estrogens, estriol and 16 alpha-hydroxyestrone, were simultaneously detected in quantities as low as 300 pg of the estrogens formed in an assay by an internal standard method. Apparent Km and Vmax of the microsomal aromatase for 16 alpha-OHA, 16 alpha,19-(OH)2A or 16 alpha-OH-19-oxo A were 1.06, 4.00 or 571 microM and 0.014, 0.087 or 1.67 pmol/min/micrograms protein, respectively. The results show that the 19-oxo steroid has extremely low affinity for aromatase relative to the other substrates.  相似文献   

15.
A new approach leading to 3-deoxy-D-manno-2-octulosonic acid (KDO) and 4-deoxy-KDO is described. The key step is the formation of the C5---C6 bond catalysed by fructose-1,6-bisphosphate aldolase, which controls the stereochemistry of these two centres. The important step of the aldehyde substrates (1a, 1b, 1c) synthesis is the Barbier reaction of ethyl-bromomethylacrylate or bromomethylacrylonitrile with the monoacetal of glyoxal in the presence of indium.  相似文献   

16.
Aromatase and cyclooxygenases: enzymes in breast cancer   总被引:8,自引:0,他引:8  
Aromatase (estrogen synthase) is the cytochrome P450 enzyme complex that converts C19 androgens to C18 estrogens. Aromatase activity has been demonstrated in breast tissue in vitro, and expression of aromatase is highest in or near breast tumor sites. Thus, local regulation of aromatase by both endogenous factors as well as exogenous medicinal agents will influence the levels of estrogen available for breast cancer growth. The prostaglandin PGE2 increases intracellular cAMP levels and stimulates estrogen biosynthesis, and previous studies in our laboratories have shown a strong linear association between aromatase (CYP19) expression and expression of the cyclooxygenases (COX-1 and COX-2) in breast cancer specimens. To further investigate the pathways regulating COX and CYP19 gene expression, studies were performed in normal breast stromal cells, in breast cancer cells from patients, and in breast cancer cell lines using selective pharmacological agents. Enhanced COX enzyme levels results in increased production of prostaglandins, such as PGE2. This prostaglandin increased aromatase activity in breast stromal cells, and studies with selective agonists and antagonists showed that this regulation of signaling pathways occurs through the EP1 and EP2 receptor subtypes. COX-2 gene expression was enhanced in breast cancer cell lines by ligands for the various peroxisome proliferator-activated receptors (PPARs), and differential regulation was observed between hormone-dependent and -independent breast cancer cells. Thus, the regulation of both enzymes in breast cancer involves complex paracrine interactions, resulting in significant consequences on the pathogenesis of breast cancer.  相似文献   

17.
In Bufo arenarum, androgen biosynthesis occurs through a complete 5-ene pathway, including 5-androstane-3β,17β-diol as the immediate precursor of testosterone. Besides, steroidogenesis changes during the breeding period, turning from androgens to C21-steroids such as 5-pregnan-3,20-diol, 3-hydroxy-5-pregnan-20-one and 5-pregnan-3,20-dione. In B. arenarum, steroid hormones are not involved in hCG-induced spermiation, suggesting that the steroidogenic shift to C21-steroids during the breeding be not related to spermiation. The activity of 17-hydroxylase-C17–20 lyase (CypP450c17) decreases during the reproductive season, suggesting that this enzyme would represent a key enzyme in the regulation of seasonal changes. However, the increase in the affinity for pregnenolone of 3β-hydroxysteroid dehydrogenase (3HSD)/isomerase could also be involved. Moreover, the reduction in CypP450c17 leading to a reduction in C19-steroids, among them dehydroepiandrosterone (DHE), would contribute to the conversion of pregnenolone into progesterone, avoiding the non-competitive inhibition exerted by DHE on this transformation. Additionally, CypP450c17 possesses a higher affinity for pregnenolone than for progesterone, explaining the predominance of the 5-ene pathway for testosterone biosynthesis. Animals in reproductive condition showed a significant reduction in circulating androgens, enhancing the physiological relevance of all the in vitro results. The in vitro effects of mGnRH and hrFSH on testicular steroidogenesis revealed that both hormones inhibited CypP450c17 activity. In summary, these results demonstrate that, in B. arenarum, the change in testicular steroidogenesis during the reproductive period could be partially due to an FSH and GnRH-induced decrease in CypP450c17 activity.  相似文献   

18.
Addition of (Me3SiNHCH2CH2)2NH (H3[N3(TMS)]) or (Me3SiNH-o-C6H4)2NH (H3[ArN3(TMS)]) to a solution of TaMe5 yields [N3(TMS)]TaMe2 or [ArN3(TMS)]TaMe2, respectively. An X-ray study of [ArN3(TMS)]TaMe2 showed it to have an approximate trigonal bipyramidal structure in which the two methyl groups are in equatorial positions and the triamido ligand is approximately planar. Addition of (C6F5NHCH2CH2)2NH (H3[N3(C6F5)]) to TaMe5 yields first [(C6F5NCH2CH2)2NH]TaMe3, which then decomposes to [(C6F5NCH2CH2)2N]TaMe2. An X-ray study of [(C6F5NCH2CH2)2N]TaMe2 shows it to be approximately a trigonal bipyramid, but the C6F5 rings are oriented so that they lie approximately in the TaN3 plane and two ortho fluorines interact weakly with the metal. Trimethylaluminum attacks the central nitrogen atom in [N3(TMS)]TaMe2 to give [(Me3SiNCH2CH2)2NAlMe3]TaMe2, an X-ray study of which shows it to be a trigonal bipyramidal species similar to the first two structures, except that the C-Ta-C bond angle is approximately 30° smaller (106.6(12)°). Addition of B(C6F5)3 to [(C6F5NCH2CH2)2NH]TaMe3 yields {[(C6F5NCH2CH2)2NH]TaMe2}+ {B(C6F5)3Me}, the structure of which most closely resembles that of [(Me3SiNCH2CH2)2NAlMe3]TaMe2 in that the C-Ta-C angle is 102.0(6)°. The C6F5 rings in {[(C6F5NCH2CH2)2NH]TaMe2}+ are turned roughly perpendicular to the TaN3 plane, i.e. ortho fluorines do not interact with the metal in this molecule.  相似文献   

19.
The question addressed in this study was the nature of the enzyme required to remove the side-chain of 17-hydroxycorticosteroids, leading in the case of cortisol to the excretion of 11β-hydroxyandrosterone, 11-oxo-androsterone and the corresponding etiocholanolones. We questioned whether it could be CYP17, the 17-hydroxylase/17,20-lyase utilized in androgen synthesis. The conversion of exogenous cortisol to C19 steroids in patients with complete 17-hydroxylase deficiency (17HD) was studied rationalizing that if CYP17 was involved no C19 steroids would be formed. The urinary excretion of the four 11-oxy-C19 steroids as well as many of the major C21 cortisol metabolites were measured by GC/MS. Our results showed that the conversion of cortisol to C19 steroids was normal in 17HD indicating that a currently unidentified enzyme must be responsible for this transformation.

A secondary goal was to determine to what extent 11-oxy-C19 steroids were metabolites of cortisol or adrenal synthesized 11β-hydroxyandrostenedione. Since cortisol-treated 17HD patients cannot produce androstenedione, all C19 11-oxy-metabolites excreted must be derived from exogenous cortisol. The extent to which 17HD patients have lower relative excretion of C19 steroids should reflect the absence of 11β-hydroxyandrostenedione metabolites. Our results showed almost all of 11-oxo-etiocholanolone and 11β-hydroxyetiocholanolone were cortisol metabolites, but in contrast the excretion of 11β-hydroxyandrosterone was less than 10% that of normal individuals, indicating that in excess of 90% must be a metabolite of 11β-hydroxyandrostenedione.  相似文献   


20.
The reactions of silver perchlorate and tetraiodoethylene in different solvents, namely, benzene and toluene, isolated two silver(I)–iodocarbon complexes, [Ag(C2I4)(C6H6)2(ClO4)] (1) and [Ag(C2I4)(ClO4)] (2). Both compounds contain intact iodoalkenes which coordinate via σ-donation of a halogen lone pair and retain their carbon–iodine bonds. Owing to the participation of the benzene molecules in coordination, complex 1 is found to be a discrete monomer in which the five-coordinate geometry of the silver ion is comprised of two benzene molecules, one C2I4 group and one perchlorate ion. In contrast, the unsaturated coordination environment of the metal ion in 2 is filled by the second iodocarbon group leading to a two-dimensional framework. The coordinated tetraiodoethylene molecules involve severe twisting of the C=C double bond, causing the C=C stretching band to move to a lower frequency.  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