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.
This work describes the atropisomeric relationships of 3-methyl-5-(3-methyl-5-phenyl-1H-pyrazol-1-yl)-1-phenyl-1H-pyrazol-4-amine (2d), which belongs to series 4-aminobipyrazole derivatives designed as anti-inflammatory agents. The (1)H nuclear magnetic resonance spectra obtained in the presence of a chiral lanthanide shift salt associated to chiral high-performance liquid chromatography analysis, X-ray diffraction, and molecular modeling tools confirmed that ortho bis-functionalized bipyrazole 2d exists as a mixture of aR,aS-atropisomers. These results provide useful information to understand the pharmacological profile of this derivative and of other 4-aminobipyrazole analogs.  相似文献   

2.
A series of N-[4-(4-nitrophenoxy)phenyl]-4-(substituted)-1,3-thiazol-2-amines was synthesized. Structural elucidation was accomplished by 1H NMR, 13C NMR, IR, and elemental analyses of synthesized compounds. The title compounds were derived from 4-(4-nitrophenoxy)phenyl thiourea, which is the key intermediate in the synthesis of nitroscanate, an anthelmintic drug. Among the synthesized compounds, N-[4-(4-nitrophenoxy)phenyl]-4-(4-fluorophenyl)-1,3-thiazol-2-amine and N-[4-(4-nitrophenoxy)phenyl]-4-(4-methoxyphenyl)-1,3-thiazol-2-amine exhibited potent anthelmintic and antibacterial activities.  相似文献   

3.
To avoid problems related to the discharge of sulfidic spent caustics, a biotechnological process is developed for the treatment of gases containing both hydrogen sulfide and methanethiol. The process operates at natron-alkaline conditions (>1 mol L−1 of sodium- and potassium carbonates and a pH of 8.5–10) to enable the treatment of gases with a high partial CO2 pressure. In the process, methanethiol reacts with biologically produced sulfur particles to form a complex mixture predominantly consisting of inorganic polysulfides, dimethyl disulfide (DMDS), and dimethyl trisulfide (DMTS). The effect of these organic sulfur compounds on the biological oxidation of sulfide to elemental sulfur was studied with natron-alkaliphilic bacteria belonging to the genus Thioalkalivibrio. Biological oxidation rates were reduced by 50% at 0.05 mM methanethiol, while for DMDS and DMTS, this was estimated to occur at 1.5 and 1.0 mM, respectively. The inhibiting effect of methanethiol on biological sulfide oxidation diminished due to its reaction with biologically produced sulfur particles. This reaction increases the feasibility of biotechnological treatment of gases containing both hydrogen sulfide and methanethiol at natron-alkaline conditions.  相似文献   

4.
An efficient one-pot three-component synthesis of thiadiazole N-nucleosides with high atom economy from β-D-ribosylhydrazine, aryl thiamide, and aromatic aldehyde promoted by (diacetoxyiodo)benzene under microwave irradiation is reported. The strategy involves formation of thiourea derivatives by microwave-assisted addition of an arylisothiocyanate formed in situ by (diacetoxyiodo)benzene-promoted oxidative rearrangement of an aryl thiamide with Schiff bases of β-D-ribosylhydrazine and a substituted/unsubstituted aromatic aldehyde. The thiourea intermediate on intramolecular heterocyclization yielded thiadiazole N-nucleosides, 2-(arylimino)-3-(β-D-ribosyl)-5-aryl-1,3,4-thiadiazoles. The whole reaction sequence proceeded with quantitative transformation of reactants into thiadiazole N-nucleosides at an ambient temperature. The reaction sequence is supported by isolation of Schiff bases and their quantitative conversion into final product by reaction with arylisothiocyanate under the same reaction conditions.  相似文献   

5.
Addition of KO2 in dimethyl sulfoxide (DMSO) to the in vitro bacterial luciferase reaction subsequent to its initiation resulted in a biphasic decay of light emission. The first and more rapid phase is attributed to quenching by DMSO. With DMSO alone the continuing decay is kinetically the same as in a control reaction. With KO2 added the second decay phase is more rapid and dependent on the KO2 concentration. The enhanced decay is attributed to superoxide anion generated from KO2 reacting without light emission with an enzyme peroxy intermediate, breaking down of the peroxide bond through intermolecular electron transfer from the superoxide anion, in competiton with an intramolecular electron transfer from the N(5) position of the flavin ring, which normally leads to the production of the excited luciferase-dihydroflavin-4a-hydroxide. © 1997 John Wiley & Sons, Ltd.  相似文献   

6.
Epoxide hydrolase activity of Streptomyces strains   总被引:2,自引:0,他引:2  
The discovery of epoxide hydrolases within a Streptomyces sp. strain collection is described. Screening was performed in 96 well microtiter plates using a modified 4-(p-nitrobenzyl)pyridine assay with styrene oxide, 1,2-epoxy-hexane or 3-phenyl ethylglycidate (3-PEG) as substrates. Out of 120 strains investigated, S. antibioticus Tü4, S. arenae Tü495 and S. fradiae Tü27 exhibited epoxide hydrolase activity. These strains were further investigated by performing laboratory-scale biotransformations utilizing styrene oxide, 1,2-epoxy-hexane and 3-PEG followed by subsequent quantitative analysis employing chiral gas chromatography. The highest conversions were achieved with whole cells from S. antibioticus Tü4 in the presence of 10% (v/v) DMSO. However, enantioselectivity was only satisfying (E = 31) in the presence of 5% (v/v) acetone, which allowed isolation of optically pure non-hydrolyzed (R)-styrene oxide (99% enantiomeric excess (ee)) and (S)-phenyl-1,2-ethandiol (72% ee) at 55% conversion after 24 h. The resolution of 3-PEG proceeded with slightly lower enantioselectivity albeit higher reaction rates. With S. fradiae Tü27 and S. arenae Tü495 enantioselectivity towards styrene oxide was only E = 3-4.  相似文献   

7.
Purified naphthalene dioxygenase (NDO) from Pseudomonas sp. strain NCIB 9816-4 oxidized toluene to benzyl alcohol and benzaldehyde by reactions involving benzylic monooxygenation and dioxygen-dependent alcohol oxidation, respectively. Xylene and nitrotoluene isomers were also oxidized to substituted benzyl alcohol and benzaldehyde derivatives. NDO oxidized ethylbenzene sequentially through (S)-1-phenethyl alcohol (77% enantiomeric excess) and acetophenone to 2-hydroxyacetophenone. In addition, NDO also oxidized ethylbenzene through styrene to (R)-1-phenyl-1,2-ethanediol (74% enantiomeric excess) by reactions involving desaturation and dihydroxylation, respectively. Isotope experiments with 18O2, H2 18O, and D2O suggest that 1-phenethyl alcohol is oxidized to acetophenone by a minor reaction involving desaturation followed by tautomerization. The major reaction in the conversion of 1-phenethyl alcohol and benzyl alcohol to acetophenone and benzaldehyde, respectively, probably involves monohydroxylation to form a gem-diol intermediate which stereospecifically loses the incoming hydroxyl group to leave the carbonyl product. These results are compared with similar reactions catalyzed by cytochrome P-450.  相似文献   

8.
A homologous Ni–Co based nanowire system, consisting of both nickel cobalt oxide and nickel cobalt sulfide nanowires, is developed for efficient, complementary water splitting. The spinel‐type nickel cobalt oxide (NiCo2O4) nanowires are hydrothermally synthesized and can serve as an excellent oxygen evolution reaction catalyst. Subsequent sulfurization of the NiCo2O4 nanowires leads to the formation of pyrite‐type nickel cobalt sulfide (Ni0.33Co0.67S2) nanowires. Due to the 1D nanowire morphology and enhanced charge transport capability, the Ni0.33Co0.67S2 nanowires function as an efficient, stable, and robust nonnoble metal electrocatalyst for hydrogen evolution reaction (HER), substantially exceeding CoS2 or NiS2 nanostructures synthesized under similar methods. The Ni0.33Co0.67S2 nanowires exhibit low onset potential of ?65, ?39, and ?50 mV versus reversible hydrogen electrode, Tafel slopes of 44, 68, and 118 mV dec?1 at acidic, neutral, and basic conditions, respectively, and excellent stability, comparable to the best reported non‐noble metal‐based HER catalysts. Furthermore, the homologous Ni0.33Co0.67S2 nanowires and NiCo2O4 nanowires are assembled into an all‐nanowire based water splitting electrolyzer with a current density of 5 mA cm?2 at a voltage as 1.65 V, thus suggesting a unique homologous, earth abundant material system for water splitting.  相似文献   

9.
This study investigated whether slow-releasing organic hydrogen sulfide donors act through the same mechanisms as those of inorganic donors to protect neurons from oxidative stress. By inducing oxidative stress in a neuronal cell line HT22 with glutamate, we investigated the protective mechanisms of the organic donors: ADT-OH [5-(4-hydroxyphenyl)-3H-1,2-dithiole-3-thione], the most widely used moiety for synthesizing slow-releasing hydrogen sulfide donors, and ADT, a methyl derivative of ADT-OH. The organic donors were more potent than the inorganic donor sodium hydrogensulfide (NaHS) in protecting HT22 cells against glutamate toxicity. Consistent with previous publications, NaHS partially restored glutamate-depleted glutathione (GSH) levels, protected HT22 from direct free radical damage induced by hydrogen peroxide (H2O2), and NaHS protection was abolished by a KATP channel blocker glibenclamide. However, neither ADT nor ADT-OH enhanced glutamate-depleted GSH levels or protected HT22 from H2O2-induced oxidative stress. Glibenclamide, which abolished NaHS neuroprotection against oxidative stress, did not block ADT and ADT-OH neuroprotection against glutamate-induced oxidative stress. Unexpectedly, we found that glutamate induced AMPK activation and that compound C, a well-established AMPK inhibitor, remarkably protected HT22 from glutamate-induced oxidative stress, suggesting that AMPK activation contributed to oxidative glutamate toxicity. Interestingly, all hydrogen sulfide donors, including NaHS, remarkably attenuated glutamate-induced AMPK activation. However, under oxidative glutamate toxicity, compound C only increased the viability of HT22 cells treated with NaHS, but did not further increase ADT and ADT-OH neuroprotection. Thus, suppressing AMPK activation likely contributed to ADT and ADT-OH neuroprotection. In conclusion, hydrogen sulfide donors acted through differential mechanisms to confer neuroprotection against oxidative toxicity and suppressing AMPK activation was a possible mechanism underlying neuroprotection of organic hydrogen sulfide donors against oxidative toxicity.  相似文献   

10.
[3+2] Cycloaddition of 5-azido-5-deoxy-1,2-O-isopropylidene-α-d-xylofuranose with 1,3-diphenyl-prop-3-enones, followed by oxidation of the intermediate triazolines in a tandem manner, led to the regioselective formation of 4-benzoyl-1-(5-deoxy-1,2-O-isopropylidene-α-d-xylofuranos-5-yl)-5-phenyl-1H-1,2,3-triazoles in moderate to good yields.  相似文献   

11.
Sodium butyrate (butyrate), 5-azacytidine (5Aza-C), dimethyl sulfoxide (DMSO), and dimethyl formamide (DMF) were applied to a human melanoma cell line for the purpose of inducing pigmentation and terminal differentiation. The results are summarized as follows: 1) butyrate, DMSO, and DMF had a strong cytostatic effect, arresting cells in the G1 phase of the cycle; 2) butyrate caused a morphological change to spindle shape whereas DMSO and DMF produced rounded cells, without affecting the levels of vimentin and intermediate filaments; 3) tyrosinase activity and melanization were stimulated by DMSO and DMF but not by butyrate; 4) butyrate induced several membrane-bound enzyme activities (alkaline phosphatase and -γ-glutamyl transpeptidase); 5) changes in the expression of antigens related to tyrosinase activity (2B7 and 5C12) only partly corresponded to the changes in enzyme activity; 6) expression of the melanosomal B863 antigen was decreased by butyrate, DMSO, and DMF; and 7) the action of DMF resembled that of DMSO whereas 5Aza-C had little effect. The results indicate that these differentiating agents activate different sets of genes, the melanogenic pathway being activated independently of -γ-glutamyltranspeptidase. The down regulation of B8G3 antigen by these agents may provide a common focus for understanding the essential action of differentiation inducers in melanoma cells.  相似文献   

12.
Bioactive sucrose esters from Bidens parviflora   总被引:4,自引:0,他引:4  
Wang N  Yao X  Ishii R  Kitanaka S 《Phytochemistry》2003,62(5):741-746
An investigation on Bidens parviflora led to the isolation of three sucrose esters and a substituted truxillate. Their structures were elucidated as (6-O-(E)-p-coumaroyl)-beta-D-fructofuranosyl-(2-->1)-alpha-D-glucopyranoside, (6-O-(E)-p-coumaroyl)-beta-D-fructofuranosyl-(2-->1)-(6-O-(E)-p-coumaroyl)-alpha-D-glucopyranoside II, 6,6'-sucrose ester of (1alpha,2alpha,3beta,4beta)-3,4-bis(4-hydroxyphenyl)-1,2-cyclobutanedicarboxylic acid, dimethyl ester of (1alpha,2alpha,3alpha,4alpha)-2,4-bis(3,4-dihydroxyphenyl)-1,3-cyclobutanedicarboxylic acid on the basis of spectral and chemical evidence. These compounds were subjected to the following bioassays: the histamine release inhibition of rat mast cells induced by antigen-antibody reaction and the inhibitory activity of PGE(2) production by macrophages.  相似文献   

13.
The anomeric composition of d-fructose in dimethyl sulfoxide changes when the solution is heated from room temperature to 150 °C, with a small increase in the α-furanose form at the expense of the β-pyranose tautomer. Additionally, a small amount of α-pyranose form was also observed at 150 °C. A mechanism is proposed for the dehydration of d-fructose to 5-hydroxymethylfurfural in DMSO at 150 °C, where the solvent acts as the catalyst. A key intermediate in the reaction was identified as (4R,5R)-4-hydroxy-5-hydroxymethyl-4,5-dihydrofuran-2-carbaldehyde by using 1H and 13C NMR spectra of the sample during the reaction.  相似文献   

14.
Hemoglobin I (HbI) from Lucina pectinata reacts with hydrogen sulfide to form the ferric sulfide complex needed to transport H2S to the bacterial endosymbiont. To further study HbI, expression studies of this protein were performed in Escherichia coli. This is the first time that the recombinant HbI was produced using a recombinant DNA expression system. Hemoglobin I cDNA was amplified and cloned into the TOPO-PBAD expression vector, which contains a fusion tag of six histidine residues (6XHis tag). Plasmid clone sequence analysis was carried out in order to ensure that the insert was in the correct reading frame for proper protein expression in E. coli. The expression of recombinant HbI was optimal when induced for 5 hr with 0.002% of l-arabinose as detected by Western blot analysis. The proto-porphyrin group was inserted into the recombinant HbI. Purification of the heme-bound recombinant protein was performed under native conditions by affinity chromatography using Ni-NTA and Probond resins. The sodium dithionite-reduced recombinant protein presented a shift from the Soret band at 413-435 nm, indicating the presence of the heme group in the adequate amino acid environment of HbI. These results indicate that recombinant HbI from Lucina pectinata can be successfully expressed in a prokaryotic system retaining its activity toward reduction, oxidation, and ligand binding.  相似文献   

15.
Synthesis and catalyst performance of 2,3‐ (types B and C ) and 2,8‐disubstituted (type D ) thiourea bifunctional organocatalysts was attempted. The synthesis of catalyst of type B has, so far, not been realized, while catalysts of type C , i.e., the 2,3‐exo‐ and the 2‐endo‐3‐exo‐thiourea catalysts, were prepared in six steps starting from (+)‐camphor. The catalysts of type D were prepared from (+)‐camphor in eight steps. All the potential catalysts as well as most of the intermediate products were carefully structurally characterized. The thiourea bifunctional organocatalysts were tested in a model reaction of Michael addition of dimethyl malonate to trans‐β‐nitrostyrene. Chirality 27:39–52, 2015. © 2014 Wiley Periodicals, Inc.  相似文献   

16.
Treatment of 4-(d-manno-pentitol-1-yl)-2-phenyl-2H-1,2,3-triazole with one molar equivalent of 2,4,6-triisopropylbenzenesulfonyl chloride (TIBSCl) in pyridine solution afforded the homo-C-nucleoside analog; 4-(2,5-anhydro-d-manno-pentitol-1-yl)-2-phenyl-2H-1,2,3-triazole in 54% yield and 4-(α-d-arabinopyranosyl)-2-phenyl-2H1,2,3-triazole analog in 3% yield. The 4-(5-O-triisopropylbenzenesulfonyl)-d-manno-pentitol-1-yl)-2-phenyl-2H-1,2,3-triazole analog was isolated as an intermediate and identified as its tetra-O-acetyl derivative. The 4-(5-chloro-5-deoxy-d-manno-pentitol-1-yl)-2-phenyl-2H-1,2,3-triazole analog was isolated as a byproduct. The structure and anomeric configuration of the products were determined by acylation, NMR spectroscopy, and mass spectrometry.  相似文献   

17.
A study on fluorescence quenching was carried out for the coumarin derivative 2‐acetyl‐3H‐benzo[f]chromen‐3‐one (2AHBC) with aniline at room temperature. Efficient fluorescence quenching was observed and Stern–Volmer (S–V) plots showed upward curves from linearity in all solvents of different polarities. For the solute 2AHBC, ground state complex formation does not hold in our study. The kinetic distance (r) value was found to be greater than the encounter distance (R) and indicated that the quenching reaction was held within the sphere of action. Diffusion‐limited reactions were found to be more prominent in high polarity solvents, namely dimethyl sulfoxide (DMSO), DMF, ACN, methanol, ethanol, propanol and DCM. The relationships between quenching constant (KSV) and dielectric constants (ε) of the different solvents were studied.  相似文献   

18.
Methionine (Met) sulfoxide reductase A (MsrA) is a key endogenous antioxidative enzyme with longevity benefits in animals. Only very few approaches have been reported to enhance MsrA function. Recent reports have indicated that the antioxidant capability of MsrA may involve a Met oxidase activity that facilities the reaction of Met with reactive oxygen species (ROS). Herein, we used a homology modeling approach to search the substrates for the oxidase activity of MsrA. We found that dimethyl sulfide (DMS), a main metabolite that produced by marine algae, emerged as a good substrate for MsrA‐catalytic antioxidation. MsrA bounds to DMS and promoted its antioxidant capacity via facilitating the reaction of DMS with ROS through a sulfonium intermediate at residues Cys72, Tyr103, and Glu115, followed by the release of dimethyl sulfoxide (DMSO). DMS reduced the antimycin A‐induced ROS generation in cultured PC12 cells and alleviated oxidative stress. Supplement of DMS exhibited cytoprotection and extended longevity in both Caenorhabditis elegans and Drosophila. MsrA knockdown abolished the cytoprotective effect and the longevity benefits of DMS. Furthermore, we found that the level of physiologic DMS was at the low micromolar range in different tissues of mammals and its level decreased after aging. This study opened a new window to elucidate the biological role of DMS and other low‐molecular sulfides in the cytoprotection and aging.  相似文献   

19.
The model ultimate carcinogen, trans-4-N-acetoxy-N-acetylaminostilbene (N-acetoxy-AAS), was reacted with guanosine (Guo) and deoxyguanosine (d-Guo) and the resulting adducts were purified by Sephadex LH-20 chromatography and HPLC for structure identification. A number of new adducts was identified by mass and 1H-NMR spectroscopy. The generation of all known adducts can now be explained by a common mechanism. The electrophile formed from the hydroxamic acid ester at C-beta reacts in a first step predominantly with N2 of guanine (Gua). The resulting quinone-imide intermediate reacts in a second step with either one of three nucleophiles: (1) predominantly with N3 of Gua to yield the previously described angular cyclic adducts ((5R,6R)/(5S,6S)-9-oxo-5,6,7,9-tetrahydro-imidazo(2,1-b)purines); (2) with N1 of Gua to yield linear cyclic adducts ((6R,7R)/(6S,7S)-9-oxo-5,6,7,9-tetrahydro-imidazo(1,2-a)purines); (3) with water to yield the open ring (1R,2R)/(1S,2S)-2-(N2'-guanyl)-1-hydroxyethanes. To some minor extent (1:8-1:9) the electrophile reacts first with N1 or N3 of guanine which leads to the formation of two pairs of the corresponding regioisomeric cyclic adducts. This reaction mechanism may also explain the formation of cross-links between different bases.  相似文献   

20.
New tin and thallium reagents capable of transferring the 1,2-di-tert-butylcyclopentadienyl moiety are easily prepared and utilized in furthering the transition metal organometallic chemistry of this intersting ligand. Lithium 1,2-di-tert-butylcyclopentadienide (1) reacts cleanly and selectively with SnClMe3 to give 2,3-di-tert-butyl-5-trimethylstannyl-1,3-cyclopentadiene (2), which in turn reacts with Re(CO)5Br to form the half-sandwich complex [Re(η5-C5H3(1,2-But)2)(CO)3] (3). The reaction between thallium ethoxide and 1,5-ditert-butyl-1,3-cyclopentadiene in hexane affords the excellent cyclopentadienyl transfer reagent, thallium 1,2-di-tert-butylcyclopentadienide (4). The thallium salt reacts with [Ru(COD)Cl2]n to give the sandwich complex [Ru(η5-C5H3(1,2-Bu2t)2)] (5).  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