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 46 毫秒
1.
Bark, wood and leaves of Ocotea catharinensis contain respectively 10 (average yield 0.7%.), 15 (average yield 0.004%.) and one (yield 0.4%.) neolignans of the bicyclo[3.2.1]octanoid and the hydrobenzofuranoid structural types, including the new rel-(7S,8R,1′R,4′S,5′R,6′R)-Δ8′-4′,6′-dihydroxy-5′-methoxy-3,4-methylenedioxy-3′-oxo-8.1′,7.5′-neolignan, (7S,8S)-Δ1′,3′,5′,8′-5,3′,5′-trimethoxy-3,4-methylenedioxy-8.1′,7.O.6′,4.O.7′-neolignan, (7R,8S,1′R,3′R)-Δ5′,8′-3,4,3′,5′-tetramethoxy-4′-oxo-8.1′,7.O.6′-neolignan and rel-(7R,8S,1′R,2′S)-Δ4′,8′-2′-hydroxy-3,4-dimethoxy-3′-oxo-8.1′,7.O.2′-neolignan.  相似文献   

2.
Isochrysis galbana, a marine prymnesiophyte microalga, is rich in long chain polyunsaturated fatty acids such as docosahexaenoic acid (C22:6n-3, Δ4,7,10,13,16,19). We used a polymerase chain reaction-based strategy to isolate a cDNA, designated IgASE1, encoding a polyunsaturated fatty acid-elongating activity from I. galbana. The coding region of 263 amino acids predicts a protein of 30 kDa that shares only limited homology to animal and fungal proteins with elongating activity. Functional analysis of IgASE1, by expression in Saccharomyces cerevisiae, was used to determine its activity and substrate specificity. Transformed yeast cells specifically elongated the C18-Δ9 polyunsaturated fatty acids, linoleic acid (C18:2n-6, Δ9,12) and -linolenic acid (C18:3n-3, Δ9,12,15), to eicosadienoic acid (C20:2n-6, Δ11,14) and eicosatrienoic acid (C20:3n-3, Δ11,14,17), respectively. To our knowledge this is the first time such an elongating activity has been functionally characterised. The results also suggest that a major route for eicosapentaenoic acid (C20:5n-3, Δ5,8,11,14,17) and docosahexaenoic acid syntheses in I. galbana may involve a Δ8 desaturation pathway.  相似文献   

3.
Sterol compositions of the cold water ophiuroids Ophioplocus januarii and Ophionotus victoriae and of the tropical ophiuroids Ophiocoma echinata and Ophiocoma wendtii are reported. The four sterol mixtures contain Δ5 mono- and di-unsaturated common 3β-hydroxy-sterols. Ophioplocus januarii and O. victoriae contain 24-methylcholesta-5,24(28)-dien-3β-ol and 24ζ-ethylcholesta-5,24(28)-dien-3β-ol in higher abundance than in O. echinata. These sterols were not found in O. wendtii. An interesting finding is the presence of Δ5,24(28)-24-n-propylidenecholesterol in 7.6% in Ophionotus victoriae.  相似文献   

4.
Three series of new cannabinoids were prepared and their affinities for the CB1 and CB2 cannabinoid recptors were determined. These are the 1-methoxy-3-(1′,1′-dimethylalkyl)-, 1-deoxy-11-hydroxy-3-(1′,1′-dimethylalkyl)- and 11-hydroxy-1-methoxy-3-(1′,1′-dimethylalkyl)-Δ8-tetrahydrocannabinols, which contain alkyl chains from dimethylethyl to dimethylheptyl appended to C-3 of the cannabinoid. All of these compounds have greater affinity for the CB2 receptor than for the CB1 receptor, however only 1-methoxy-3-(1′,1′-dimethylhexyl)-Δ8-THC (JWH-229, 6e) has effectively no affinity for the CB1 receptor (Ki=3134±110 nM) and high affinity for CB2 (Ki=18±2 nM).  相似文献   

5.
Magnetic field-dependent recombination measurements together with magnetic field-dependent triplet lifetimes (Chidsey, E.D., Takiff, L., Goldstein, R.A. and Boxer, S.G. (1985) Proc. Natl. Acad. Sci USA 82, 6850–6854) yield a free energy change ΔG(P+H3P*) = 0.165 eV ±0.008 at 290 K. This does not depend on whether nuclear spin relaxation in the state 3P* is assumed to be fast or slow compared to the lifetime of this state. This value, being (almost) temperature independent, indicates ΔG(P+H3P*) ΔH(P+H3P*) and is consistent with ΔG(1P* − P+H) and ΔH(1P* − 3P*) from previous delayed fluorescence and phosphorescence data, implying ΔG ΔH for all combinations of these states.  相似文献   

6.
A reduction of previously reported 2-methoxyethyl and 2-methylthioethyl functionalized zirconocenedichlorides (η5-C5Me4CH2CH2EMe)(η5-C5Me5)(ZrCl2 (E = O, S) and (η5-C5Me4CH2CH2EMe)(η5-C5Me4CH2CH2E′Me)ZrCl2 (E = O, S; E′ = O, S) with Mg/Hg in THF leads unexpectedly to the products of O---Me and S---Me bond cleavage (η5,σ-C5Me4CH2CH2E)(η5-C5Me5)ZrMe (E = O, S), (η5,σ-C5Me4CH2CH2E)(η5-C5Me4CH2CH2E′Me)ZrMe (E = O, S; E′ = O), and (η5,σ-C5Me4CH2CH2S)2Zr respectively. The crystal structure of (η5,σ-C5Me4CH2CH2S)2Zr was established by X-ray analysis. At that same time the reduction of (ηsu5-C5Me4CH2CH2EMe)(η5-C5Me5)ZrCl2 (E> = O, S) under 1 atm of CO gives either only the dicarbonyl derivative (η5-C5Me4CH2CH2EMe) (η5-C6Me5)Zr(CO)2 (E = O) or a complex mixture of products (E = S).  相似文献   

7.
Previous in vitro experiments showed that both, Taenia crassiceps and Taenia solium cysticerci have the ability to metabolize exogenous androstenedione to testosterone. Here we evaluate on the capacity of both cysticerci to synthesize several sex steroid hormones, using different hormonal precursors. Experiments using thin layer chromatography (TLC) showed that both cysticerci were able to produce 3H-hydroxyprogesterone, 3H-androstenedione and 3H-testosterone when 3H-progesterone was used as the precursor. They also synthesized 3H-androstenediol and 3H-testosterone when 3H-dehydroepiandrosterone was the precursor. In addition, both cysticerci interconverted 3H-estradiol and 3H-estrone. These results, strongly suggest the presence and activity of the Δ4 and Δ5 steroid pathway enzymes, 3β-hydroxysteroid dehydrogenase/Δ5-4 isomerase-like enzyme (3β-HSD), that converts androstenediol into testosterone; and the 17β-hydroxysteroid dehydrogenase that interconverts estradiol and estrone, in both types of cysticerci.  相似文献   

8.
The synthesis of the 3-heptyl, and the eleven isomeric 3-methylheptyl-Δ8-tetrahydrocannabinols (3–7, R and S methyl epimers, and 8) has been carried out. The synthetic approach entailed the synthesis of substituted resorcinols, which were subjected to acid catalyzed condensation with trans-para-menthadienol to provide the Δ8-THC analogue. The 1′-, 2′- and 3′-methylheptyl analogues (3–5) are considerably more potent than Δ8-THC. The 4′-, 5′- and 6′-methylheptyl isomers (6–8) are approximately equal in potency to Δ8-THC.  相似文献   

9.
In this study, we assessed the effects of tibolone and its metabolites on the production of a progesterone sensitive parameter, prolactin, in human endometrium stroma cells in vitro. In addition, the metabolism of the compounds by isolated stromal and epithelial cells was evaluated.

The reference compounds, progesterone, Org 2058, and DHT all induced prolactin production. Oestradiol also slightly induced prolactin production and enhanced the response to Org 2058. Tibolone and Δ4-tibolone were similar with regard to potency to induce prolactin levels in the culture supernatant. Their potency was lower than that of Org 2058, similar to that of progesterone and higher than that of DHT. The efficacies of tibolone, Δ4-tibolone and Org 2058 were similar (200-fold induction). The estrogenic tibolone metabolites 3- and 3β-OH tibolone also significantly stimulated prolactin production. Their potency, however, was low since significance was reached only at the highest concentrations tested.

The PR antagonist Org 31710 inhibited both tibolone- and Δ4-tibolone-induced prolactin production. The responses of tibolone and Δ4-tibolone were not affected by co-incubation with the androgen receptor antagonist OH-flutamide. The effect of tibolone, but not Δ4-tibolone, was antagonized approximately 50% in combination with the highest dose (1 μM) estrogen receptor antagonist, ICI 164384. The induction of prolactin by 3- and 3β-OH tibolone was antagonized most potently by Org 31710, but also by ICI 164384 and OH-flutamide.

Tibolone is metabolized differently in epithelial and stromal cells of the human endometrium. The epithelial cells mostly produce the progestagenic/androgenic Δ4-tibolone. The stromal cells produce predominantly the 3β-OH tibolone, and some Δ4-tibolone, but the net effect observed with regard to prolactin production is progestagenic. When the metabolites 3-OH, 3β-OH, and Δ4-tibolone were added to the cultures no conversions were observed. The HPLC analyses showed no evidence for the production of sulfated metabolites.

In conclusion, the net effects on endometrial stromal cells are predominantly progestagenic. Tibolone is converted by epithelial cells into Δ4-tibolone which displays progestagenic and androgenic activities, whereas in stromal cells also the estrogenic metabolites 3- and 3β-OH tibolone are formed.  相似文献   


10.
The reactivity, towards nucleophiles and electrophiles, of dimolybdenum allenylidene complexes of the type [Cp2Mo2(CO)4(μ,η2(4e)-C=C=CR1R2)] (Cp=η5-C5H5) has been investigated. The nucleophilic attacks occur at the Cγ carbon atom, while electrophiles affec the C atom. Variable temperature solution 1H NMR studies show a dynamic behavior of these complexes consisting of an equilibrium between two enantiomers with a symmetrical [Cp2Mo2(CO)4(μ-σ,σ(2e)-C=C=CR1R2)] transition state. Extended Hückel MO calculations have been carried out on the model [Cp2Mo2(CO)4(μ,η2-C=C=CH2]. The calculated charges of the allenylidene carbon atoms suggest that the electrophilic attacks are under charge control, while the nucleophilic attacks are rather under orbital control.  相似文献   

11.
The kinetics of substitution reactions of [η-CpFe(CO)3]PF6 with PPh3 in the presence of R-PyOs have been studied. For all the R-PyOs (R = 4-OMe, 4-Me, 3,4-(CH)4, 4-Ph, 3-Me, 2,3-(CH)4, 2,6-Me2, 2-Me), the reactions yeild the same product [η5-CpFe(CO)2PPh3]PF6, according to a second-order rate law that is first order in concentrations of [η5-CpFe(CO)3]PF6 and of R-PyO but zero order in PPh3 concentration. These results, along with the dependence of the reaction rate on the nature of R-PyO, are consistent with an associative mechanism. Activation parameters further support the bimmolecular nature of the reactions: ΔH = 13.4 ± 0.4 kcal mol−1, ΔS = −19.1 ± 1.3 cal k−1 mol−1 for 4-PhPyO; ΔH = 12.3 ± 0.3 kcal mol−1, ΔS = 24.7 ±1.0 cal K−1 mol−1 for 2-MePyO. For the various substituted pyridine N-oxides studied in this paper, the rates of reaction increase with the increasing electron-donating abilities of the substituents on the pyridine ring or N-oxide basicities, but decrease with increasing 17O chemical shifts of the N-oxides. Electronic and steric factors contributing to the reactivity of pyridine N-oxides have been quantitatively assessed.  相似文献   

12.
Nitrogen dioxide (NO2) is a key biological oxidant. It can be derived from peroxynitrite via the interaction of nitric oxide with superoxide, from nitrite with peroxidases, or from autoxidation of nitric oxide. In this study, submicromolar concentrations of NO2 were generated in < 1 μs using pulse radiolysis, and the kinetics of scavenging NO2 by glutathione, cysteine, or uric acid were monitored by spectrophotometry. The formation of the urate radical was observed directly, while the production of the oxidizing radical obtained on reaction of NO2 with the thiols (the thiyl radical) was monitored via oxidation of 2,2′-azino-bis-(3-ethylthiazoline-6-sulfonic acid). At pH 7.4, rate constants for reaction of NO2 with glutathione, cysteine, and urate were estimated as 2 × 107, 5 × 107, and 2 × 107 M−1 s−1, respectively. The variation of these rate constants with pH indicated that thiolate reacted much faster than undissociated thiol. The dissociation of urate also accelerated reaction with NO2 at pH > 8. The thiyl radical from GSH reacted with urate with a rate constant of 3 × 107 M−1 s−1. The implications of these values are: (i) the lifetime of NO2 in cytosol is < 10 μs; (ii) thiols are the dominant ‘sink’ for NO2 in cells/tissue, whereas urate is also a major scavenger in plasma; (iii) the diffusion distance of NO2 is 0.2 μm in the cytoplasm and < 0.8 μm in plasma; (iv) urate protects GSH against depletion on oxidative challenge from NO2; and (v) reactions between NO2 and thiols/urate severely limit the likelihood of reaction of NO2 with NO• to form N2O3 in the cytoplasm.  相似文献   

13.
A series of cationic nickel complexes [(η3-methally)Ni(PP(O))]SbF6 (1–4) [PP(O) = Ph2P(CH2)P(O)Ph2 (dppmO) (1), Ph2P(CH2)2P(O)Ph2 (dppeO) (2), Ph2P(CH2)3P(O)Ph2 (dpppO) (3), pTol2P(CH2)P(O)pTol2 (dtolpmO) (4)] has been synthesized in good yields by treatment of [(η3-methally)NiBr]2 with biphosphine monoxides and AgSbF6. The ligands are coordinated in a bidentate way. Starting from [(η3-all)PdI]2 the cationic complexes [(η3-all)PP(O))]Y (8–14). [PP(O) = dppmO, dppeO, dpppO, dtolpmO;Y = BF4, SbF6, CF3SO3, pTolSO3] were synthesized in good yields. The coordination mode of the ligand is dependent on the backbone and the anion, revealing a monodentate coordination with dppmO for stronger coordinating anions. The intermediates [(η3-all)Pd(I)(PP(O)-κ1-P)] (5–7) [PP(O) = dppmO (5), dppeO (6), dtolpmO (7)] were isolated and characterized. Neutral methyl complexes [(Cl)(Me)Pd(PP(O))] (15–18). [PP(O) = dppmO (15), dppeO (16), dpppO (17), dtolpmO (18)] can easily be obtained in high yields starting from [(cod)PdCl2]. For dppmO two different routes are presented. The structure of [(Me)(Cl)Pd{;Ph2P(CH2-P(O)Ph22-P,O};] · CH2Cl2 (15) with the chlorine atom trans to phosphorus was determined by X-ray diffraction.  相似文献   

14.
The time course of effects following i.p. injections of 0.3–10.0 mg/kg Δ9-tetrahydrocannabinol (δ9-THC) and 0.1–3.0 mg/kg 11-OH-Δ9-THC were determined in rats during 6 hour sessions under a fixed-interval 90-second schedule of food reinforcement. In addition, the acute and chronic effects of these two drugs were tested in different rats trained on a differential reinforcement of low rate 15-second schedule. The onset of activity of 11-OH-Δ9-THC was faster, and usually abruptly suppressed all responding, while Δ9-THC's onset was slower and often resulted in a decreased and steady pattern of responding. 11-OH-Δ9-THC was about 3 times more potent, had a shorter duration of effect and when responding resumed, it returned to control rates within a shorter time than for Δ9-THC. Tolerance and cross-tolerance developed at the same rate to equipotent doses of the two drugs. The time course for plasma and brain levels of radioactivity were studied in other rats after i.p. administration of H39-THC and H3-11-OH-Δ9-THC. 11-OH-Δ9-THC was absorbed more quickly than Δ9-THC and reached peak plasma and brain levels earlier. In addition, higher plasma and brain levels, and larger brain to plasma ratio of radioactivity were attained after 11-OH-Δ9-THC. Therefore, differences in behavioral effects produced by Δ9-THC and it's 11-hydroxy metabolite were accompanied by differences in absorption and disposition.  相似文献   

15.
Lewis acid adducts of the hydrides cis- and trans-Re(CO)(PMe3)4H (1) and (2), mer-Re(CO)2(PMe3)3H (3), fac-Re(CO)2(PMe3)3H (4) and trans-Re(CO)3(PMe3)2H (5) were studied with BH3 and 9-borabicyclo[3,3,1] norbonane (BBNH). Using BH3·THF and (BBNH)2 1 and 2 afforded Re(CO)(PMe3)32-BH4) (6) and Re(CO)(PMe3)32-BBNH2) (7) as stable and isolable products. VT IR studies established for the reaction to 7 that BBNH first attaches in a pre-equilibrium to the OCO atom of 1 or 2. At higher temperatures ReH adduct formation occurs with instantaneous transformation to 7 and elimination of PMe3·BBNH. In a similar way, the hydrides 3 and 4 were converted with BH3·THF and (BBNH)2 to yield the stable complexes Re(CO)2(PMe3)22-BH4) (8) and Re(CO)2(PMe3)22-BBNH2) (9). The intermediacy of the η1-BH4 adducts mer-/fac-Re(CO)2(PMe3)31-BH4) was confirmed by VT 1H, 31P NMR and VT IR experiments. The conversion of 5 with BH3·THF led to equilibria with adducts at the OCO terminus in trans position to H and with HRe as revealed by VT IR studies. Temperature dependent 31P equilibrium studies allowed to calculate ΔH=−4.9 kcal mol−1 and ΔS=+0.034 e.u. for this reaction. These adducts could not be isolated. Compound 5 does not react with (BBNH)2 even at elevated temperatures. DFT calculations were carried out to support the structures of the BH3 adducts of 5. In addition a vibrational analysis helped to unravel the IR band assignments of the involved compounds. DFT calculations on 8 confirmed its C2v structure. X-ray diffraction studies were carried out on single crystals of 6 and 7.  相似文献   

16.
Kinetic results are reported for intramolecular PPh3 substitution reactions of Mo(CO)21-L)(PPh3)2(SO2) to form Mo(CO)22-L)(PPh3)(SO2) (L = DMPE = (Me)2PC2H4P(Me)2 and dppe=Ph2PC2H4PPh2) in THF solvent, and for intermolecular SO2 substitutions in Mo(CO)32-L)(η2-SO2) (L = 2,2′-bipyridine, dppe) with phosphorus ligands in CH2Cl2 solvent. Activation parameters for intramolecular PPh3 substitution reactions: ΔH values are 12.3 kcal/mol for dmpe and 16.7 kcal/mol for dppe; ΔS values are −30.3 cal/mol K for dmpe and −16.4 cal/mol K for dppe. These results are consistent with an intramolecular associative mechanism. Substitutions of SO2 in MO(CO)32-L)(η2-SO2) complexes proceed by both dissociative and associative mechanisms. The facile associative pathways for the reactions are discussed in terms of the ability of SO2 to accept a pair of electrons from the metal, with its bonding transformations of η2-SO2 to η1-pyramidal SO2, maintaining a stable 18-e count for the complex in its reaction transition state. The structure of Mo(CO)2(dmpe)(PPh3)(SO2) was determined crystallographically: P21/c, A=9.311(1), B = 16.344(2), C = 18.830(2) Å, ß=91.04(1)°, V=2865.1(7) Å3, Z=4, R(F)=3.49%.  相似文献   

17.
The reaction of benzyl 2,6,6′-tri-O-benzyl-3′,4′-O-isopropylidene-β-lactoside with 1,11-ditosyloxy-3,6,9-trioxaundecane gave benzyl 2,6,6′-tri-O-benzyl-3′,4′-O-isopropylidene-3,2′-O--(3,6,9-trioxaundecane-1,11-diyl)-β-lactoside (2, 47%). Acid hydrolysis of 2 and condensation of the product with 1,14-ditosyloxy-3,6,9,12-tetra-oxatetradecane afforded benzyl 2,6,6′-tri-O-benzyl-3′,4′-O-(3,6,9,12-tetraoxa-tetradecane-1,14-diyl)-3,2′-O-(3,6,9-trioxaundecane-1,11-diyl)-β-lactoside (29%). Similarly, the reaction of benzyl 2,6,2′,4′,6′-penta-O-benzyl-β-lactoside with Ts[OCH2CH2]4OTs gave benzyl 2,6,2′,4′,6′-penta-O-benzyl-3,3′-O-(3,6,9-trioxaundecane-1,11-diyl)-β-lactoside (78%). 1H-N.m.r. spectroscopy has been used to study the formation of host-guest complexes with some of these macrocyclic compounds and benzyl ammonium thiocyanate.  相似文献   

18.
Desaturation of fatty acids is a key reaction in the biosynthesis of moth sex pheromones. The main component of Spodoptera littoralis sex pheromone blend is produced by the action of Δ11 and Δ9 desaturases. In this article, we report on the cloning of four desaturase-like genes in this species: one from the fat body (Sls-FL1) and three (Sls-FL2, Sls-FL3 and Sls-FL4) from the pheromone gland. By means of a computational/phylogenetic method, as well as functional assays, the desaturase gene products have been characterized. The fat body gene expressed a Δ9 desaturase that produced (Z)-9-hexadecenoic and (Z)-9-octadecenoic acids in a (1:4.5) ratio, whereas the pheromone gland Sls-FL2 expressed a Δ9 desaturase that produced (Z)-9-hexadecenoic and (Z)-9-octadecenoic acids in a (1.5:1) ratio. Although both Δ9 desaturases produced (Z)-9-tetradecenoic acid from myristic acid, transformed yeast grown in the presence of a mixture of myristic and (E)-11-tetradecenoic acids produced (Z,E)-9,11-tetradecadienoic acid, but not (Z)-9-tetradecenoic acid. The Sls-FL3 gene expressed a protein that produced a mixture of (E)-11-tetradecenoic, (Z)-11-tetradecenoic, (Z)-11-hexadecenoic and (Z)-11-octadecenoic acids in a 5:4:60:31 ratio. Despite having all the characteristics of a desaturase gene, no function could be found for Sls-FL4.  相似文献   

19.
Uncatalyzed reaction between cyclopentadiene and (E)-3,4,5,6,7-pentaacetoxy-1-nitrohept-1-enes having the -manno, -galacto, and -gluco configurations at C-3—C-7 led, in each case, to the four stereoisomeric 5-nitro-6-(1,2,3,4,5-penta-O-acetylpentitol-1-yl)bicyclo[2.2.1]hept-2-enes. Face selectivity is discussed in terms of the sugar-chain configuration. The structures assigned the adducts are based on their n.m.r. spectra, and, in the case of the -manno compounds, on X-ray data. Also described are the 13C-n.m.r. spectra of the starting nitroalkenes. The crystal structures of (5S,6S)1,2,3,4,5-penta-O-acetyl-1-C-(5-exo-nitrobicyclo[2.2.1]hept-2-en-6-endo-yl- -manno-pentitol (3a) and (5S,6S)1,2,3,4,5-penta-O-acetyl-1-C-(5-endo-nitrobicyclo[2.2.1]hept-2-en-6-exo-yl- -manno-pentitol (5a) were determined from three-dimensional, X-ray data. Crystals of 3a are monoclinic, space group P21, with two molecules in a cell of dimensions a = 9.054(3), b = 15.580(11), c = 10.138(4) Å, β = 116.27(3)°. The structure was refined to an R-factor of 0.050 on the basis of 1485 observations. Crystals of 5a are triclinic, space group P1, with one molecule in a cell of dimensions a = 8.680(4), b = 9.760(4), c = 8.695(7) Å, = 98.69(5), β = 103.13(5), γ = 112.09(3)°. The structure was refined to an R-factor of 0.074 based on 970 observations.  相似文献   

20.
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