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 15 毫秒
1.
A fluorescent compound was isolated and purified from rice cultures of Fusarium equiseti (Alaska 2-2). Mass spectrometry and nuclear magnetic resonance data indicated that its structure is 2,2-dimethyl-5-amino-6-(3′-hydroxyl-4′-methoxyl-butyryl)-4-chromone. It is an analog of the mycotoxin fusarochromanone, in which the amino group on C-3′ is replaced by a hydroxyl group and the hydroxyl group on C-4′ is replaced by a methoxyl group.  相似文献   

2.
The rate constants for hydrolysis of the enantiomers of amino acid p-nitrophenyl esters catalyzed by bifunctional comicellar catalysts containing the imidazolyl and hydroxyl groups have been determined at pH 7.30, 0.02 m phosphate buffer, and 25°C. The kinetic analysis suggests a reaction scheme which involves acylation followed by deacylation at the imidazolyl group. Although no appreciable cooperative catalytic efficiencies are observed between the bifunctional groups in the acylation step, it is found that the deacylation rates are thus accelerated by surfactant hydroxyl groups, and some of the stereoselective acyl transfer reaction occurs from the imidazolyl to the hydroxyl group in optically active comicellar systems.  相似文献   

3.
A new series of 2,4-diphenyl-6-aryl pyridines containing hydroxyl group(s) at the ortho, meta, or para position of the phenyl ring were synthesized, and evaluated for topoisomerase I and II inhibitory activity and cytotoxicity against several human cancer cell lines for the development of novel anticancer agents. Structure–activity relationship study revealed that the substitution of hydroxyl group(s) increased topoisomerase I and II inhibitory activity in the order of meta > para > ortho position. Substitution of hydroxyl group on the para position showed better cytotoxicity.  相似文献   

4.
The reaction of some 2,3-anhydroaldo-hexo- and -pento-pyranoside derivatives with MgBr2-etherate was found to afford bromodeoxy products in high yield. In the absence of any free hydroxyl group in the molecule, rigid bicyclic, and flexible monocyclic, 2,3-anhydro-α-d-aldopyranoside derivatives, mainly yielded 3-bromo-3-deoxy products through an unusual, diequatorial opening of the oxirane ring. In contrast, similar 2,3-anhydro derivatives having a free hydroxyl group in the molecule underwent the usual, diaxial opening of the oxirane ring, affording the 2-bromo-2-deoxy product. However, methyl 2,3-anhydro-4-O-methyl-β-d-ribopyranoside, despite the absence of a free hydroxyl group, underwent trans-diaxial opening of the oxirane ring.  相似文献   

5.
Hydroxylamino aromatic compounds are converted to either the corresponding aminophenols or protocatechuate during the bacterial degradation of nitroaromatic compounds. The origin of the hydroxyl group of the products could be the substrate itself (intramolecular transfer mechanism) or the solvent water (intermolecular transfer mechanism). The conversion of hydroxylaminobenzene to 2-aminophenol catalyzed by a mutase from Pseudomonas pseudoalcaligenes JS45 proceeds by an intramolecular hydroxyl transfer. The conversions of hydroxylaminobenzene to 2- and 4-aminophenol by a mutase from Ralstonia eutropha JMP134 and to 4-hydroxylaminobenzoate to protocatechuate by a lyase from Comamonas acidovorans NBA-10 and Pseudomonas sp. strain 4NT were proposed, but not experimentally proved, to proceed by the intermolecular transfer mechanism. GC-MS analysis of the reaction products formed in H218O did not indicate any 18O-label incorporation during the conversion of hydroxylaminobenzene to 2- and 4-aminophenols catalyzed by the mutase from R. eutropha JMP134. During the conversion of 4-hydroxylaminobenzoate catalyzed by the hydroxylaminolyase from Pseudomonas sp. strain 4NT, only one of the two hydroxyl groups in the product, protocatechuate, was 18O labeled. The other hydroxyl group in the product must have come from the substrate. The mutase in strain JS45 converted 4-hydroxylaminobenzoate to 4-amino-3-hydroxybenzoate, and the lyase in Pseudomonas strain 4NT converted hydroxylaminobenzene to aniline and 2-aminophenol but not to catechol. The results indicate that all three types of enzyme-catalyzed rearrangements of hydroxylamino aromatic compounds proceed via intramolecular transfer of hydroxyl groups.  相似文献   

6.
O-Alkylation of myo-inositol derivatives containing more than one hydroxyl group via their alkali metal alkoxides (sodium or lithium) preferentially occurs at a hydroxyl group having a vicinal cis-oxygen atom. In general the observed selectivity is relatively higher for lithium alkoxides than for the corresponding sodium alkoxide. The observed regioselectivity is also dependent on other factors such as the solvent and reaction temperature. A perusal of the results presented in this article as well as those available in the literature suggests that chelation of metal ions by inositol derivatives plays a significant role in the observed regioselectivity. Steric factors associated with the axial or equatorial disposition of the reacting hydroxyl groups do not contribute much to the outcome of these O-alkylation reactions. These results could serve as guidelines in planning synthetic strategies involving other carbohydrates and their derivatives.  相似文献   

7.
《Phytomedicine》2015,22(4):498-503
Our previous results elucidated that the leaves of Eriobotrya japonica possessed the potential to suppress ovariectomy-induced bone mineral density deterioration, and ursolic acid, the major bioactive component in these leaves, suppressed the osteoclast differentiation. The aim of this study was to discover more candidates for development of novel antiosteoclastogenesis agents from the leaves of E. japonica. Phytochemical analysis following a cell-based osteoclastic tartrate-resistant acid phosphatase (TRAP) activity assay revealed 11 more compounds with a potent antiosteoclastogenesis effect. The potency of ursane-type triterpenoids from the leaves of E. japonica prompted us to investigate the structure–activity relationships underlying their antiosteoclastogenesis. The results revealed that both the hydroxyl group at C-3 and the carboxylic group at C-17 played indispensable roles in the antiosteoclastogenesis activity of ursane-type triterpenoids. The configuration at C-3 (a beta-form of the hydroxyl group) was found to be important for this activity. While introducing a hydroxyl group at C-19 increased the inhibitory activity of ursane-type triterpenoids carrying an alpha-form hydroxyl group at C-3. The bioactivity analyses of ursolic acid and oleanolic acid demonstrated that the antiosteoclastogenesis effect of ursolic acid may be related to different positions of the C-29 and C-30 methyl groups on the E-ring, since oleanolic acid showed limited activity. The addition of a hydroxyl group at C-2 would dramatically improve the inhibition of oleanane-type triterpenoids. Collectively, these findings could provide important clues for the improvement of multi-targeted antiosteoclastogenesis agents from the leaves of E. japonica.  相似文献   

8.
Rhodococcus chlorophenolicus PCP-I, a degrader of polychlorinated phenols, guaiacols (2-methoxyphenols), and syringols (2,6-dimethoxyphenols), was shown to O-methylate the degradation intermediate, a chlorinated para-hydroquinone, into 4-methoxyphenol. O-methylation was constitutively expressed, whereas the degradation of chlorophenols and chlorohydroquinones was inducible in R. chlorophenolicus. The O-methylating reaction required two hydroxyl groups in positions para to each other. R. chlorophenolicus selectively methylated the hydroxyl group flanked by two chlorine substituents. Tetrachlorohydroquinone, trichlorohydroquinone, and 2,6-dichlorohydroquinone were methylated into tetrachloro-4-methoxyphenol, 2,3,5-trichloro-4-methoxyphenol, and 3,5-dichloro-4-methoxyphenol, respectively. Chlorohydroquinones with only one chlorine adjacent to a hydroxyl group were methylated only in trace amounts, and no metabolite was formed from hydroquinone. The degradation intermediates formed in hydroxylation of tetrachloroguaiacol and trichlorosyringol by R. chlorophenolicus were O-methylated into two isomeric trichlorodimethoxyphenols and two isomeric dichlorotrimethoxyphenols, respectively. R. chlorophenolicus also degraded the polychlorinated methylation products (tetrachlorinated and trichlorinated 4-methoxyphenols), but not mono- and dichlorinated 4-methoxyphenols.  相似文献   

9.
New sulfoquinovosyldiacylglycerols derived from 2-O-β-d-glucopyranosyl-sn-glycerol, carrying acyl chains of various length on the glycerol moiety, were prepared through a convenient synthetic procedure in which a sulfonate is introduced at the C-6 position of glucose by oxidation of a thioacetate in the presence of the unprotected secondary hydroxyl groups, and tested for their anti-tumor-promoting activity using a short-term in vitro assay for Epstein-Barr virus early antigen (EBV-EA) activation. Our study has allowed to ascertain the role of the 6′-sulfonate group and the need of a free hydroxyl group on the glycerol moiety in inhibiting the EBV activation promoted by the tumor promoter 12-O-tetradecanoylphorbol-13-acetate (TPA).  相似文献   

10.
Palladium complexes with a salicylaldiminato ligand bearing a hydroxyl group (1a and 1b) have been synthesized and characterized. The structures of these complexes were confirmed by X-ray crystallography. A reversible deprotonation/protonation of the hydroxyl moiety on 1b was observed, while such behaviour was impossible with a related palladium complex (1c) bearing a methoxyl group in place of the hydroxyl group. The deprotonation affected its catalytic behaviour: the activity for polymerization of methyl acrylate catalyzed by 1b considerably decreased in the presence of 1 equiv. of tBuOK.  相似文献   

11.
The formation of borate complexes contributes significantly to the distribution coefficients obtained in the separation of polyhydroxy acids in borate medium. A study of a large number of straight-chain hydroxy acids shows that these contributions tend to increase with the number of vicinal hydroxyl groups and with their distance from the carboxylate group. Vicinal hydroxyl groups in gauche conformation have the ability to form strong complexes, provided that the steric conditions are favourable in other respects. The contributions of vicinal hydroxyl groups in anti conformation are much less.  相似文献   

12.
The behavior of a variety of N-acetyl- and N-(benzyloxycarbonyl)-aminocyclitols with 2,2-dimethoxypropane-N,N-dimethylformamide-p-toluenesulfonic acid has been examined. Both cis- and trans-vicinal hydroxyl groups are readily bridged to give 1,3-dioxolanes. In one (sterically favorable) case, the reagent linked the nitrogen atom of an acetamido group with a vicinal hydroxyl group to give an N-acetyl-2,2- dimethyloxazolidine; this heterocyclic system is less labile to acid than are the 1,3-dioxolanes.  相似文献   

13.
In a previous study, we demonstrated that the anticancer synthetic retinoid N-(4-hydroxyphenyl)retinamide (4HPR) redox cycles at the mitochondrial enzyme dihydroorotate dehydrogenase to trigger anomalous reactive oxygen species (ROS) production and attendant apoptosis in transformed human epithelial cells. Furthermore, we speculated that the hydroxyl functional group of 4HPR was required for this pro-oxidant property. In this study, we investigated the role of the hydroxyl functional group in the in vitro cytotoxicity of 4HPR. Using 4HPR, its primary in vivo metabolite N-(4-methoxyphenyl)retinamide (4MPR), and the synthetic derivative N-(4-trifluoromethylphenyl)retinamide (4TPR), we examined the pro-oxidant and apoptotic effects, as well as the cellular uptake, of these three N-(4-substituted-phenyl)retinamides in premalignant and malignant human skin, prostate, and breast epithelial cells. Compared to 4HPR, both 4MPR and 4TPR were ineffective in promoting conspicuous cellular ROS production, mitochondrial disruption, or DNA fragmentation in these transformed cells. Interestingly, both 4MPR and 4TPR were not particularly cell permeative relative to 4HPR in skin or breast epithelial cells, which implied an additional role for the hydroxyl functional group in the cellular uptake of 4HPR. Moreover, the short-term uptake of 4HPR was directly proportional to cell size, but this characteristic, in obvious contrast to cellular bioenergetic status and/or dihydroorotate dehydrogenase expression, was not fundamentally influential in the overall sensitivity to the promotion of cellular ROS production and apoptosis induction by this agent. Together, these results strongly implicate the hydroxyl functional group in the cytotoxic effects of 4HPR.  相似文献   

14.
On hydrazinolysis in 1:4 acetic acid-pyridine, and in pyridine, partial O-deacylation of fully acylated methyl glycosides and some other glycosyl compounds (23 compounds) was found to be induced, to give, in good yields, products bearing one free hydroxyl group; the results obtained indicated that, among the primary and secondary O-acyl groups, the 2-O-acyl groups were, in general, the most labile toward the nucleophile (hydrazine). Hydrazinolysis of 1,2-O-isopropylidenealdofuranose acylates (3 compounds), on the other hand, gave, in high yield, the corresponding monoacyl derivatives having the protecting group on their primary hydroxyl group. The factors possibly involved in the regioselectivity of the hydrazinolysis were discussed.  相似文献   

15.
In order to clarify whether arginine has a promoting effect on the acylation of hydroxyl groups of serine, threonine, or tyrosine by homobifunctional cross-linking agents in aqueous solution, we carried out systematic experiments with model peptides, comparing relative reaction yields with covalently protected and unprotected arginines by MALDI-MS. The guanidinium group could be demonstrated to contribute to the reactivity of hydroxyl groups toward N-hydroxysuccinimide esters and catalyze the nucleophilic substitution, probably via hydrogen bonds.  相似文献   

16.
Low levels of reactive oxygen species (ROS) act as important signaling molecules, but in excess they can damage biomolecules. ROS regulation is therefore of key importance. Several polyphenols in general and flavonoids in particular have the potential to generate hydroxyl radicals, the most hazardous among all ROS. However, the generation of a hydroxyl radical and subsequent ROS formation can be prevented by methylation of the hydroxyl group of the flavonoids. O-Methylation is performed by O-methyltransferases, members of the S-adenosyl-l-methionine (SAM)-dependent O-methyltransferase superfamily involved in the secondary metabolism of many species across all kingdoms. In the filamentous fungus Podospora anserina, a well established aging model, the O-methyltransferase (PaMTH1) was reported to accumulate in total and mitochondrial protein extracts during aging. In vitro functional studies revealed flavonoids and in particular myricetin as its potential substrate. The molecular architecture of PaMTH1 and the mechanism of the methyl transfer reaction remain unknown. Here, we report the crystal structures of PaMTH1 apoenzyme, PaMTH1-SAM (co-factor), and PaMTH1-S-adenosyl homocysteine (by-product) co-complexes refined to 2.0, 1.9, and 1.9 Å, respectively. PaMTH1 forms a tight dimer through swapping of the N termini. Each monomer adopts the Rossmann fold typical for many SAM-binding methyltransferases. Structural comparisons between different O-methyltransferases reveal a strikingly similar co-factor binding pocket but differences in the substrate binding pocket, indicating specific molecular determinants required for substrate selection. Furthermore, using NMR, mass spectrometry, and site-directed active site mutagenesis, we show that PaMTH1 catalyzes the transfer of the methyl group from SAM to one hydroxyl group of the myricetin in a cation-dependent manner.  相似文献   

17.
As a part of ongoing studies in developing novel anticancer agents, a series of modified 2,4-diaryl-5H-indeno[1,2-b]pyridines were designed, and synthesized by introducing hydroxyl and chlorine moieties. They were evaluated for topoisomerase inhibitory activity and cytotoxicity against HCT15, T47D, and HeLa cancer cell lines. This modification allowed us to demonstrate structure–activity relationship (SAR) study with respect to the non-substituted 2,4-diaryl-5H-indeno[1,2-b]pyridines. Compounds (2, 3, 4, 5, 8, and 9) with meta or para hydroxyl group on 2 or 4-phenyl ring have enhanced topo I and II inhibitory activity and cytotoxicity. However, additional substitution of chlorine group on furyl or thienyl ring (11, 12, 14, 16–18) generally reduced topo I and II inhibitory activity but improved cytotoxicity. The observation of cytotoxic properties and SAR study according to the position of hydroxyl and chlorine group will provide valuable insight for further study of development of novel anticancer agents with related scaffolds.  相似文献   

18.
Oxymercuration-demercuration of hydroxy alkenes follows an intramolecular pathway to furnish 1,4-epoxides (tetrahydrofurans) when the hydroxyl group is β (trans only) or γ to a double bond and 1,5-epoxides (tetrahydropyrans) when the hydroxyl group is δ to the double bond. The cis and trans isomers of methyl ricinoleate and methyl 9-hydroxyoctadec-12-enoate, and a series of cis and trans octadecenols (Δ2–Δ6) are used to establish these relationships.1,4- and 1,5-Epoxides are also formed during the oxymercuration of methyl densipolate and methyl 12,13-dihydroxyoleate and during the hydroxymercuration of methyl octadeca-9,12 and 8,12-dienoates.  相似文献   

19.
Microbial transformation of some steroid 19-hydroxy compounds and 1(10),5-dienes, substituted by a hydroxyl group in position 3, by the action ofProactinomyces globerulus, yielded substances with aromatic A ring. Substrates with 17-keto group underwent simultaneously a partial reduction resulting in the formation of 17β-hydroxyl.  相似文献   

20.
Peng F  Ren JL  Peng B  Xu F  Sun RC  Sun JX 《Carbohydrate research》2008,343(17):2956-2962
Hemicellulose-based hydrophobic biomaterials with degrees of substitution ranging from 0.46 to 1.54 were synthesized under mild conditions in homogeneous media (N,N-dimethylformamide-lithium chloride) by reacting the native wheat straw hemicellulosic polymers with lauroyl chloride using 4-dimethylaminopyridine as a catalyst. Other catalysts such as N-bromosuccinimide, N-methyl pyrrolidine, N-methyl pyrrolidinone, and pyridine were also investigated. Under optimum reaction conditions (2 equiv of lauroyl chloride and triethylamine per hydroxyl group, 5% 4-dimethylaminopyridine, 40 °C, 35 min), a high DS value of 1.54 was obtained. The biomaterials were characterized by FT-IR spectroscopy and 13C NMR spectroscopy as well as by thermal analysis. The results showed that the lauroylation occurred preferably at the C-3 hydroxyl group of β-d-Xylp units in the hemicelluloses, and the thermal stability of the hydrophobic polymers increased by esterification.  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