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.
Aureobasidium pullulans, originally introduced as an inadvertent contaminant in solutions used for evaluating the stability of prostaglandins, proved to lead to the rapid disappearance of the cyclopentenone unit of PGA2 (as monitored by circular dichroic spectroscopy). The cyclopentenone unit is converted, in various metabolites, to a 9-keto, 9α or 9β-hydroxy group lacking the ring unsaturation. The major EtoAc-soluble 9-hydroxy metabolite (Compound-I) was shown to be 9α, 15α-dihydroxy-2,3,4,5-tetranor-13-trans-prostenoic acid. Similar tetranor 9-hydroxy metabolites with one additional degree of unsaturation, and with a 9β-hydroxy group, also occur but these have not been fully characterized. Only two of the wide range of 9-keto metabolites are fully characterized by mass spectral (MS) data: 9,15-oxo-2,3,4,5-tetranorprostanoic acid and 9,15-oxo-2,3,4,5-tetranor-13-trans-prostenoic acid. The water soluble metabolites have not been characterized further.The fully characterized metabolites together with MS data from mixtures of minor metabolites indicate that A. pullulans can perform the following transformations: β-oxidation, dehydrogenation at C-15, reduction of the enone carbon-carbon double bonds (both Δ10,11 and Δ13,14), reduction of the 9-ketone, and possibly migration of the cyclopentyl double bond (Δ10,11 → Δ11,12). A. pullulans metabolizes 15-epimeric PGA2 equally readily with the production of similar products. PGA1 affords less 9-keto metabolites with compound I constituting 33% of the product by HPLC analysis. A. pullulans displays some enantioselectivity, PGA2 and 15-epi-PGA2 are each metabolized more rapidly than their enantiomers. Other prostaglandins appear to be less readily metabolized.  相似文献   

2.
R Simantov 《Life sciences》1978,23(25):2503-2508
Mouse pituitary tumor cells grown in tissue culture release endorphins spontaneously to the culture medium. Depolarization of these cells by incubation with high K+ concentration (56 mM) increased 2–3 folds the release of endorphins. The K+ evoked release was Ca++ dependent by that: a, removal of Ca++ ions inhibited 90% of K+ stimulated release. b, ethyleneglycol-bis (β-aminoethyl ether) N,N′-tetraacetic acid (EGTA) inhibited release of endorphins in the presence of high K+ and Ca++. It is suggested that dual regulatory system inhibit and/or stimulate in-vivo release of endorphins from the pituitary glands.  相似文献   

3.
The cytoplasmic and outer membranes containing either trans9-octadecenoate, trans9-hexadecenoate or cis9-octadecenoate as predominant unsaturated fatty acid residues in the phospholipids were prepared from a fatty acid auxotroph, Escherichia coli strain K1062. Order-disorder transitions of the phospholipids were revealed in both fractions of the cell envelope by fluorescent probing or wide angle X-ray diffraction. The mid-transition temperatures, Tt, and the range of the transition, ΔT, are similar in the outer and cytoplasmic membrane. Relative to the corresponding extracted lipids, 60–80% of the hydrocarbon chains take part in the transition in the cytoplasmic membrane whereas in the outer membrane only 25–40% of the chains become ordered. The results suggest that in the outer membrane part of the lipids form fluid domains in the form of mono- and/or bilayers.  相似文献   

4.
Cytokinin-autonomous tobacco callus was incubated in defined mineral medium containing 3H-adenine for 60 minutes. Radioactivity was incorporated into the four predominant free cytokinins, ribosyl-trans-zeatin, trans-zeatin, N6-(Δ2-isopentenyl) adenosine and N6-(Δ2-isopentenyl) adenine. The bases were more abundant than their respective ribosides, N6-(Δ2-isopentenyl) adenine being the most abundant cytokinin. No discrete peaks of radioactivity could be detected on the HPLC column eluate corresponding to the elution volumes of cis-zeatin and ribosyl-cis-zeatin.  相似文献   

5.
A slow reacting substance, produced by murine mastocytoma cells, has been shown to have the structure 5(S)-hydroxy-6(R)-S-glutathionyl-7,9,11-trans-14-cis-eicosatetraenoic acid (11-trans leukotriene C, previously referred to as leukotriene C-2) by ultraviolet spectroscopy, amino acid analyses, lipoxygenase conversion and comparisions with a synthetic compound of known structure and stereochemistry.  相似文献   

6.
Theodore Dashman 《Life sciences》1980,27(15):1415-1422
The enol-ether amino acid, L-2-amino-4-methoxy-trans-butenoic acid (AMTB) is an inhibitor of porphobilinogen synthase (PBG synthase) when added prior to the addition of the substrate δ-aminolevulinic acid. The inhibition of PBG synthase by several stereoisomers and analogues of AMTB was investigated to determine those structural features of AMTB which may be necessary for inhibition. The D-trans isomer was also an inhibitor after preincubation, whereas the L-cis isomer inhibited with or without preincubation. The amino acid analogues, DL-vinylglycine, DL-2-aminobutanoic acid, the reduced form of L-2-amino-4-methoxy-trans-3-butenoic acid, L-2-amino-4-(2-aminoethoxy)-trans-3-butenoic acid and its reduced congener did not inhibit PBG synthase even with preincubation. This structure activity relationship indicates that the trans double bond and methoxy moiety of L-2-amino-4-methoxy-trans-3-butenoic acid are probably required for inhibition.Heme, when preincubated with PBG synthase, was an inactivator of the enzyme. However, when both L-2-amino-4-methoxy-trans-3-butenoic acid and heme were simulatneously preincubated with PBG synthase, inactivation of the enzyme was greater than with either compound separately. The possibility of multiple catalytic sites was suggested by the use of multiple inhibition kinetics in the presence of heme and L-2-amino-4-methoxy-trans-3-butenoic acid.  相似文献   

7.
Behavioral comparisons of the stereoisomers of tetrahydrocannabinols   总被引:1,自引:0,他引:1  
The potencies of (?)-trans9-THC, (+)-trans9-THC, (+)-cis9-THC, (?)-trans8-THC and (+)-trans8-THC were compared in several different species. (?)-trans9-THC was 100 times more potent than (+)-trans9-THC in depressing schedule-controlled responding in monkeys. The (+)-trans isomers were less effective than their corresponding (?)-trans isomers in the dog static-ataxia test, but potency ratios could not be determined due to a lack of dose-responsiveness of the (+)-trans isomers. However, it appeared that their potency differed by at least ten fold. The potency of (+)-cis9-THC in the dog static-ataxia test was comparable to that of (+)-trans9-THC. The hypothermia in mice produced by the (?) isomers of trans9-THC and trans8-THC were 9.1 and 30.4 times greater than that produced by their respective (+)-isomers. Also, the potency ratio of the (+)- and (?)-trans9-THC was 5.6 as measured by depression of spontaneous activity in mice. The magnitude of the potency ratios of the THC stereo-isomers is dependent upon the species and the pharmacological test used.  相似文献   

8.
S S Tobe  G E Pratt 《Life sciences》1975,17(3):417-422
We have compared, on an individual basis, the volume of the corpora allata with their ability to synthesize and release juvenile hormone (JH) using glands taken at daily intervals throughout the period of sexual maturation and the first two ovarian cycles in Schistocerca gregaria. A standard in vitro radiochemical assay was used to measure the rates of both spontaneous JH biosynthesis from [methyl-14C]-methionine, and of JH biosynthesis stimulated by optimal concentrations of [C-2 3H]-farnesenic acid. Computation of results showed that there are, during this period, changes of up to 250-fold in the rate of spontaneous JH biosynthesis per unit volume corpora allata. It is concluded that the volume of the corpora allata is of no value as an indicator of the spontaneous synthetic activity of the glands in this species, and that the overall rate of JH synthesis is regulated by mechanisms that do not involve large changes in the volume of the gland cells. However, in the presence of farnesenic acid, there is a corelation between stimulated JH synthesis and glandular volume, suggesting that the volume of the gland reflects the maximum activity of the final two stages in JH biosynthesis.  相似文献   

9.
The induction of freezing tolerance in bromegrass (Bromus inermis Leyss) cell culture was used to investigate the activity of absisic acid (ABA) analogs. Analogs were either part of an array of 32 derived from systematic alterations to four regions of the ABA molecule or related, pure optical isomers. Alterations were made to the functional group at C-1 (acid replaced with methyl ester, aldehyde, or alcohol), the configuration at C-2, C-3 (cis double bond replaced with trans double bond), the bond order at C-4, C-5 (trans double bond replaced with a triple bond), and ring saturation (C-2′, C-3′ double bond replaced with a single bond so that the C-2′ methyl and side chain were cis). All deviations in structure from ABA reduced activity. A cis C-2, C-3 double bond was the only substituent absolutely required for activity. Overall, acids and esters were more active than aldehydes and alcohols, cyclohexenones were more active than cyclohexanones, and dienoic and acetylenic analogs were equally active. The activity associated with any one substituent was, however, markedly influenced by the presence of other substituents. cis, trans analogs were more active than their corresponding acetylenic analogs unless the C-1 was an ester. Cyclohexenones were more active than cyclohexanones regardless of oxidation level at C-1. An acetylenic side chain decreased the activity of cyclohexenones but increased the activity of cyclohexanones relative to their cis, trans counterparts. Trends suggested that for activity the configuration at C-1′ has to be the same as in (S)-ABA, in dihydro analogs the C-2′-methyl and the side chain must be cis, small positional changes of the 7′-methyl are tolerable, and the C-1 has to be at the acid oxidation level.  相似文献   

10.
Leukotriene C-1, a “Slow Reacting Substance” (SRS), has been shown to possess the molecular Structure depicted by V (5(S)-hydroxy-6(R)-S-glutathionyl-7,9-trans-11,14-cis-eicosatetraenoic acid) by its identity with a totally synthetic product of known structure and stereochemistry.  相似文献   

11.
Supplementation of culture medium with elaidic acid (400 μg/flask) in L-M cells results in the formation of an otherwise undetected lipid component. We have identified this lipid component to be a mixture of free fatty alcohols containing primarily elaidyl alcohol with cetyl, stearoyl, and oleoyl alcohols as minor constituents. Formation of fatty alcohols by fatty acid supplementation seems to be specific with trans fatty acids (i.e., elaidate, trans vaccenate, and linolelaidate); addition of stearate and oleate to the L-M cells does not produce fatty alcohols. The fatty alcohols accumulated by the trans fatty acid supplementation are associated with both the particulate and supernatant fractions of the cells.  相似文献   

12.
We wish to report here the syntheses of (5S, 6R)-5-hydroxy-, (5R, 6R)-5-hydroxy-, (5R, 6S)-5-hydroxy-, and (5S, 6S)-5-hydroxy-PGI1 and their methyl ester derivatives. Treatment of (5R, 6S)-5-epoxy- and (5S, 6R)-5-epoxy-PFG methyl esters with acid washed silica gel afforded (5R, 6R)-5-hydroxy- and (5S, 6S)-5-hydroxy-PGI1 methyl esters; correspondingly, silica promoted cyclization of (5S, 6S)-epoxy- and (5S, 6R)-5-epoxy- PGF1 methyl esters yielded (5S, 6R)-5-hydroxy- and (5R, 6S)-5-hydroxy- PGI1 methyl esters. Alternatively, the 5-hydroxyl group was introduced into the PGI1 skeleton via reaction of the 5-mercuric halides with sodium borohydride in the presence of oxygen. Stereochemical assignments were based on their mode of synthesis and 1H nmr shift differences.  相似文献   

13.
The properties of carnitine transport were studied in rat kidney cortex slices. Tissue: medium concentration gradients of 7.9 for L-[methyl-14C]carnitine were attained after 60-min incubation at 37°C in 40 μM substrate. L- and D-carnitine uptake showed saturability. The concentration curves appeared to consist of (1) a high-affinity component, and (2) a lower affinity site. When corrected for the latter components, the estimated Km for L-carnitine was 90 μM and V = 22nmol/min per ml intracellular fluid; for D-carnitine, Km = 166 μM and V = 15 nmol/min per ml intracellular fluid. The system was stereospecific for L-carnitine. The uptake of L-carnitine was inhibited by (1) D-carnitine, γ-butyrobetaine, and (2) acetyl-L-carnitine. γ-Butyrobetaine and acetyl-L-carnitine were competitive inhibitors of L-carnitine uptake. Carnitine transport was not significantly reduced by choline, betaine, lysine or γ-aminobutyric acid. Carnitine uptake was inhibited by 2,4-dinitrophenol, carbonyl cyanide m-chlorophenylhydrazone, N2 atmosphere, KCN, N-ethylmaleimide, low temperature (4°C) and ouabain. Complete replacement of Na+ in the medium by Li+ reduced L- and D-carnitine uptake by 75 and 60%, respectively. Complete replacement of K+ or Ca2+ in the medium also significantly reduces carnitine uptake. Two roles for the carnitine transport system in kidney are proposed: (1) a renal tubule reabsorption system for the steady-state maintenance of plasma carnitine; and (2) maintenance of normal carnitine levels in kidney cells, which is required for fatty acid oxidation.  相似文献   

14.
Ac-[Pro1, D-Phe2, D-Trp3, D-Trp6]-LH-RH completely inhibited ovulation in cycling rats at 200μg/rat and is comparable in activity to the corresponding D-1-analogue. This Ac-Pro1-analogue is the most potent antiovulatory peptide yet known having an L-amino acid residue in position 1. This result shows that for the design of potent inhibitors of ovulation, a D-amino acid residue is not essential in position 1. The corresponding Ac-D-Pro1- and Kic1-analogues completely inhibited ovulation at 750μg/rat, but not at 200μg/rat, and the Cpc1-analogue was inactive at these dosages.  相似文献   

15.
The enantiomer selection in the nucleophilic addition reaction of optically active amines such as α-amino acid esters to phenylalanine and N-methylphenylalanine N-carboxyanhydride in m-dimethoxybenzene as a solvent has been investigated. Stereoselectivity between the amines and the N-carboxyanhydrides was found to change markedly according to the reaction conditions. This experimental finding is in contrast to the idea hitherto accepted that in the nucleophilic addition-type polymerization of α-amino acid N-carboxyanhydride the growing chain end reacts preferentially with one of the enantiomorphic N-carboxyanhydrides having the same configuration, and indicates the importance of the investigation of stereoselectivity in the N-carboxyanhydride polymerization using suitable model reactions. Most (S)-α-amino acid esters reacted preferentially with (R)-phenylalanine N-carboxyanhydride, and this type of stereoselectivity increased with the N-methylation of N-carboxyanhydride and with increasing bulkiness of the Cα substituent of α-amino acid esters (alanine < norleucine < leucine < valine). The relationship observed between the stereoselectivity and the structures of amines and N-carboxyanhydrides was explained satisfactorily in terms of the transition state model in which the interaction of N-carboxyanhydride nitrogen and α-amino acid ester carbonyl as well as the interaction of N-carboxyanhydride carbonyl and α-amino acid ester nitrogen was taken into account. (S)-Proline ethyl ester did not show enantiomer selectivity toward phenylalanine N-carboxyanhydride, but reacted preferentially with (S)-(N)-methylphenylalanine N-carboxyanhydride. for the reaction of proline ester with N-carboxyanhydride a transition-state model was proposed, which was different from the transition state model proposed for other α-amino acid esters. Some experiments were carried out to examine the transition-state models proposed. The implications of the present investigation in stereoselectivity in the nucleophilic addition-type polymerization of N-carboxyanhydride hitherto reported are discussed.  相似文献   

16.
The peptide t-butyloxycarbonyl-α-aminoisobutyryl-L-prolyl-L-prolyl-N-methylamide has been shown to adopt an extended structure in the solid state. The Pro-Pro segment occurs in the poly-proline II conformation. On dissolution of single crystals at ~ 233°K, a single species corresponding to the all trans peptide backbone is observed by 270 MHz 1H NMR. On warming, trans to cis isomerization about the Pro-Pro bond is facilitated. Both cis' (ψ ~?50°) and trans' (ψ ~ 130°) rotamers about the Pro3 CαCO bond are detectable in the Pro-Pro cis conformer, at low temperature. These observations demonstrate unambiguously the large differences in the solid state and solution conformations of a Pro-Pro sequence.  相似文献   

17.
A proton magnetic resonance procedure with tri(3-heptafluorobutyryl-d-camphorato)praseodymium (III) as a chiral shift eagent has been developed to determine the enantimeric purity of monoglycerides 1,2-diglycerides and triglycerides with one mono-unsaturated fatty acid at position sn-1 or sn-3 and two saturated fatty acids at the two other glycerol positions. A model compound, 1-oleoyl-2,3-dipalmitoyl-sn-glycerol, was converted ito the trimethylsilyl either of 2,3-dipalmitoyl-an-glycerol by epoxidation of the double bond, followed by pancreatic hydrolysis and separation and trimethylsilylation of the resulting sn-1,2, and sn-2,3-diglycerides. This separation becomes feasible by the contribution of the epoxy group to the polarity of the diglyceride. The protons of the trimethysilyl ether group were used for determining the enantiomeric ratio. The addition of a chira shift reagent induces a useful enantiomeric splitting which allows the accurate determination of the ratio of both enantiomers. The trimethylsilyl emers of 1,2-diglycerides are better suited for this purpose than the acetyl compounds. For monoglycetides, the earlier published method with the diaceltates gives a better line separation in 1H-NMR spectra.  相似文献   

18.
A series of unsaturated and polyunsaturated fatty acids with a sulfur atom substituting for a methylene unit of the chain has been prepared and characterized. The syntheses were accomplished by the Wittig coupling of the ylid derived from the triphenylphosphonium salt of 9-bromononanoic acid with aldehydes containing sulfur. The newly formed double bond had predominately the natural Z geometry even when the starting aldehyde was conjugated with the sulfur atom. The sulfides 13-thia-9(Z)-octadecenoic acid (2), 13-thia-9(Z), 11(E)-octadecadienoic acid (5) and 13-thia-9(E), 11(E)-octadecadienoic acid (6) were readily converted into their sulfoxide derivatives by treatment with an equivalent amount of m-chloroperoxybenzoic acid. The structures of the novel compounds were confirmed by the application of ir, uv, 1H-nmr, 13C-nmr, and (as methyl esters) chemical ionization mass spectrometry. Two members of this new family of fatty acids (5 and 6) were found to inhibit the catalysis of the oxygenation of linoleic acid by soybean type-1 lipoxygenase. The analysis of the kinetic data for compound 5 indicated that the type of inhibition was reversible competitive with an inhibition constant of 30 μM.  相似文献   

19.
3-Methylcholanthrene and five related dihydrodiols have been tested for microsome-mediated mutagenicity towards Salmonellatyphimurium TA100 and for the induction of mutation to 8-azaguanine resistance in V79 Chinese hamster cells and malignant transformation in M2 mouse fibroblasts. In both mutagenicity test systems, the 9,10-diol was considerably more active than either the parent hydrocarbon, the related cis-2α,3-diol, the trans-4,5-, the trans-7,8- or the trans-11,12-dihydrodiols. At a non-toxic concentration (1μg/ml medium), the 9,10-diol induced the formation of more transformed malignant foci in cultures of M2 cells than 3-methylcholanthrene and the other diols were either inactive or only weakly active in this test system. The results obtained indicate that the 9,10-dihydrodiol derived from 3-methylcholanthrene is involved, presumably following conversion into the corresponding vicinal diol-epoxide, 9,10-dihydro-9,10-dihydroxy-3-methylcholanthrene 7,8-oxide, in the metabolic activation of this carcinogenic polycyclic hydrocarbon.  相似文献   

20.
[4-14C]Cholesterol was incubated with an adrenocortical preparation in the presence of 16O2 and 18O2 devoid of significant 16O18O. Isolated (20R,22R)-20,22-dihydroxycholesterol was converted to a trimethylsilyl derivative and analyzed by gas chromatography - mass spectrometry to determine the isotope distribution of the oxygen atoms at C-20 and C-22. The ions of me 289, 291, and 293 (comprising the C8 C-20 to C-27 side-chain and containing, respectively, 16O2, 16O18O, and 18O2) exhibited a binomial distribution indicating that the oxygen atoms of the vicinal glycol were drawn at random from the atomic pool of the oxygen molecules. If both side-chain hydroxyl groups had originated from the atoms of the same oxygen molecule, the ion of me 291 would have been absent.  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