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.
3D computer models of Fe–Ni–Co, Fe–Ni–FeS–NiS, Fe–Co–FeS–CoS, Ni–Co–NiS–CoS Txy diagrams have been designed. The geometric structure (35 surfaces, two-phase surface of the reaction type change, 17 phase regions) of the Fe–Ni–FeS–NiS Txy diagram is investigated in detail. The liquidus hypersurfaces prediction of the Fe–Ni–Co–FeS–NiS–CoS subsystem is represented.  相似文献   

2.
In the present work, a series of eight new imidazole, 4,5–dichloroimidazole, 4,5–diphenylimidazole and benzimidazole based nitro–functionalized mono–N –heterocyclic carbene (NHC)–silver(I) acetate ( 7a–d ) and bis–NHC–silver(I) hexafluorophosphate complexes ( 8a–d ) were synthesised by the reaction of the corresponding azolium hexafluorophosphate salts ( 6a–d ) with silver(I) acetate and silver(I) oxide in methanol and acetonitrile, respectively. All the synthesised compounds were fully characterized by various spectroscopic techniques and elemental analyses. Additionally, the structure of bis–(1–benzyl–3–(p –nitrobenzyl)–4,5–dichloroimidazole–2–ylidene)silver(I) hexafluorophosphate complex ( 8b ) was confirmed by single crystal X–ray diffraction analysis. Preliminary in vitro antibacterial evaluation was conducted for all the compounds ( 6a–d) , ( 7a–d) , and ( 8a–d) by Kirby–Bauer's disc diffusion method followed by the determination of Minimum Inhibitory Concentration (MIC) from broth macrodilution method against five standard bacteria; two Gram–positive bacterial strains (Staphylococcus aureus and Bacillus subtilis) and three Gram–negative bacterial strains ( Escherichia coli , Shigella sonnei, and Salmonella typhi). All the hexafluorophosphate salts ( 6a – d) were found inactive against the tested bacterial strains and their corresponding mono– and bis–NHC–silver(I) complexes ( 7a–d and 8a–d ) exhibited moderate to high antibacterial activity with MIC value in the range 8–128 μg/mL. In addition, preliminary in vitro anticancer potential of all the silver(I) complexes ( 7a–d and 8a–d ) was determined against the human derived breast adenocarcinoma cells (MCF 7) by MTT assay. All the mono– and bis–NHC–silver(I) complexes ( 7a–d and 8a–d ) orchestrated high anticancer potential with IC50 values ranging from 10.39 to 59.56 nM. In comparison, mono– NHC–silver(I) complexes performed better than the bis–NHC–silver(I) complexes.  相似文献   

3.
1H and 13C NMR spectra of AC–PSer–Gly, Ala–PSer–Gly and Gly–PSer–Phe have been measured and analysed as a function of pD. The NMR parameters of the PSeryl side chain are a function of the sequence. The second titration step of the phosphate group (pK2 = 5.7) is much more difficult to detect in Ac–PSer–Gly and Ala–PSer–Gly than in Gly–PSer–Phe. The conformation in which H-α? C-α? C-β? O? P forms a planar W-type arrangement predominates only for Ala–PSer–Gly. In the other two phosphopeptides the gauche conformations contribute increasingly, in particular for Gly–PSer–Phe.  相似文献   

4.
The fragmentation chemistry of protonated H–Val–Asn–OH, H–Val–Gln–OH and H–Val–Lys–OH is investigated in this work by means of modeling and density functional theory calculations. Former experimental studies indicate that the ratio of a 1 and y 1 ions cannot be explained by considering the proton affinities of the corresponding dissociating species on the a 1y 1 pathway, while the fragmentation of other dipeptides can be understood in this way. We demonstrate that considering the correct PA value for H–Asn–OH eliminates the deviation observed for H–Val–Asn–OH. The larger than expected a 1/y 1 ratio of H–Val–Gln–OH is explained by considering the dissociation kinetics of the proton-bound dimers formed on the a 1y 1 pathway and competition of the deamidation and a 1y 1 channels. For H–Val–Lys–OH, it is proposed that a 1 ions are indeed formed from one of the primary products, protonated H–Val–Cap–OH.  相似文献   

5.
Thermodynamic Calculations on the Transport of Tungsten in Systems Containing Fluorine Compounds Information on the transport behaviour of tungsten in systems containing fluorine is obtained by a thermodynamic analysis of the systems W(s)–F, W(s)–F–O, W(s)–F–Br, W(s)–F–Br–O, W(s)–Si–F, W(s)–B–F, W(s)–Si–F–Br, and W(s)–B–F–Br. The results are compared with experimental findings. The influence of bromine, oxygen, silicon, and boron on the transport of tungsten is investigated. The nonblackening condition (dλW/dT < 0) is fulfilled in systems containing fluorine as well as bromine.  相似文献   

6.
Two novel compounds, 8–[2–(2–thienyl)vinyl]–10,10–dimethyl–10H–pyrido[1,2–a] indolium perchlorate ( 3a ) and 8–[2–(5–phenyl–2–thienyl)vinyl]–10,10–dimethyl–10H–pyrido[1,2–a]indolium perchlorate ( 3b ) were synthesized and characterized by IR, 1H–NMR, elemental analyses, and X–ray diffraction. Crystal structural analysis suggested that either 3a or 3b exhibited good coplanarity and rings and vinyl in the target molecule could make up a large conjugated system. Ultraviolet–visible absorption analysis indicated both 3a and 3b possessed large maximum absorptions, and 3b underwent a significant redshift (43.0 nm) in comparison with 3a .  相似文献   

7.
This articles studied and determined the viscosities of the binary mixtures of water–methanol, water–ethanol, water–propanol, water–acetone, acetone–ethanol, methanol–ethanol, and acetone–hexane and the ternary mixtures of water–methanol–ethanol and water–ethanol–acetone at 20°C. It is shown that the mixing of water with the alcohols and acetone resulted in a positive deviation of viscosity, which reached the maximum value at the water mole fraction x 1 ~ 0.7 for water–methanol, x 1 ~ 0.72 for water–ethanol, x 1 ~ 0.74 for water–propanol, and x 1 ~ 0.83 for water–acetone binary mixture. This viscosity deviation can be mainly attributed to the formation of micelles of alcohol or acetone molecules in water because of the hydrophobic attraction between the hydrocarbon chains. The micelle surfaces are surrounded by hydration layers, leading to the positive viscosity deviation in the liquid mixtures because the water in hydration layers has a much higher viscosity than bulk water. Also, the contrary observation was found in the binary mixtures of acetone–ethanol and acetone–hexane, having a negative viscosity deviation.  相似文献   

8.
Photocatalytic decolorization properties of cobalt doped-ZrO2-multiwalled carbon nanotubes (Co–ZrO2–MWCNTs) and chitosan–sodium alginate encapsulated Co–ZrO2–MWCNTs (CS/Alg–Co–ZrO2–MWCNTs) with varying weight percentage of Co–ZrO2–MWCNTs are presented in this research paper. The Co–ZrO2–MWCNTs was first synthesized through homogenous co-precipitation method and introduced into the chitosan–sodium sodium alginate (CS/Alg) biopolymer matrix. The bio-nanocomposites were characterized using X-ray powder diffraction, Fourier transform infrared spectroscopy, transmission electron microscopy, (UV–Vis)-spectroscopy and energy dispersive spectroscopy to obtain information on their structure, formation, morphology, size and elemental analysis. The photodecolorization efficiency of the samples was determined through their decolorization of trypan blue dye aqueous solution in 180 min. Recyclability of the catalysts was also assessed. The bio-nanocomposites experienced reduced band gap values with subsequent improvement in visible light activity compared to the uncapped Co–ZrO2–MWCNTs. All the CS/Alg–Co–ZrO2–MWCNTs exhibited higher photodecolorization activities than the uncapped Co–ZrO2–MWCNTs. The most efficient catalyst (CS/Alg–40 % Co–ZrO2–MWCNTs) with a band gap of 2.56 eV displayed 94 % decolorization efficiency of the dye. Though reusability of the catalyst is significant, its efficiency diminished consistently after each cycle.  相似文献   

9.
8,13–Epoxy–3β–hydroxylabd–14–en–2–one (1) from the wood of Lagarostrobos colensoi has been converted in 25% overall yield into 8α–acetoxy–2β,3β–dihydroxy–13,14,15,16–tetranorlabdan–12–oic acid (2), a useful intermediate for synthesis.  相似文献   

10.
Synthesis and radical polymerization of a novel optically active methacrylate, (S)–2–tert–butoxycarbonylamino–3–phenylpropyl methacrylate (MA–F–BOC), were examined. MA–F–BOC was synthesized from methacrylic acid and N–protected (L)–phenylalaninol. Radical polymerization of MA–F–BOC quantitatively afforded the corresponding polymethacrylate with a relatively high molecular weight. Radical copolymerizations of MA–F–BOC were carried out with styrene and acrylamide to afford the copolymers. Radical polymerization of MA–F–BOC in the presence of n–butanethiol afforded the oligomers, whose degrees of polymerizations were 3.3–8.0. The BOC group was completely cloven with HBr to afford the corresponding optically active polymeric amine quantitatively. © 1998 John Wiley & Sons, Inc. J. Polym. Sci. A Polym. Chem. 36: 1981–1986, 1998  相似文献   

11.
12.
Liquid crystal trimers and tetramers containing two kinds of flexible spacers, namely O(CH2) m O and COO(CH2) n O, were divided into four classes according to the odd/even nature of the number of atoms in the flexible spacers: specifically, even–even, odd–odd, even–odd, and odd–even trimers, and even–even–even, odd–odd–odd, odd–even–odd, and even–odd–even tetramers. The transition properties of the four types of trimers and of tetramers were compared. Although the nematic–isotropic transition temperature and the associated entropy changes were primarily related to the number of the even-membered flexible spacers in these molecules, the different combinations of the flexible spacers significantly affected their transition properties.  相似文献   

13.
We present the principles of stoichiography and a reference-free stoichiographic differential (separating) dissolution method used to study the composition and structure of thin films and nanostructured systems: HTS films with 123 different compositions, Al–Au–Sn–Co–Mn, Si/SiO2/Ni(Cr)–Cu–Cu2S, Cr–Cu–S, and Cu–S multilayer films, Bi–Ti–O films on Ru/SiO2/Si, Mn1–xZn x S, and ZnS–EuS supports, and also nanostructured manganese ferrite in borate glass matrices, nanodisperse composite sorbents and the Co–Si–Pt–O/Al2O3 catalyst modified by Pt nanoparticles, and oxide catalyst precursor Fe2Co/Al2O3 for the synthesis of carbon nanotubes.  相似文献   

14.
The double-layer characteristics of liquid renewable Cd–Ga (0.3 at % Cd) and In–Ga (14.2 at % In) electrodes in the gamma-butyrolactone (GBL) solutions of various electrolytes are studied by measuring the differential capacitance and using the method of open-circuit jet electrode. For the (Cd–Ga)/GBL and (In–Ga)/GBL interfaces, the zero-charge potentials, which are not distorted by the specific adsorption of ions, and the chemisorption potential drops of solvent are determined. It is shown that, in spite of the fact that the work function decreases as we pass from (In–Ga) to (Cd–Ga), the chemisorption potential drops of solvent on both electrodes are close. This behavior is explained by a closer approach of GBL dipoles to the surface of (Cd-Ga) electrode providing more effective overlapping of donor–acceptor levels of metal and solvent. It is shown that, in GBL, the adsorption parameters of halide ions and their surface activity series depend on the metal nature. On the (Cd–Ga) and (In–Ga) electrodes, the reversed surface activity series of halide ions is observed: on the Hg electrode in various solvents, the surface activity increases in the series Cl < Br < I, whereas on the (Cd–Ga) and (In–Ga) electrodes in GBL, it varies in the reverse series I < Br < Cl.  相似文献   

15.
The title compound, 2,2′‐(oxalyldiimino)bis(3‐methylbutanoic acid), C12H20N2O6, possesses a centre of symmetry. In the crystal, mol­ecules are connected by hydrogen bonds between ox­amide and carboxyl groups, similar to the pattern of the monoclinic forms of HO–Gly–CO–CO–Gly–OH and HO–Aib–CO–CO–Aib–OH (Gly is glycine and Aib is 2‐amino­isobutyric acid). The characteristic torsion angles in the title compound are close to those in peptide α‐helices.  相似文献   

16.
Enzymes that act on substrates R–O–PO3H2 often work on substrate analogues R–O–AsO3H2; such substrates are unstable, since esters of H3AsO4 hydrolyse easily. They also form easily, so that an enzyme that acts on R–O–PO3H2 often acts on a mixture of R–OH and arsenate via an ester that forms at the active site. Similarly coenzyme analogues may be formed; for example, a stable and active aspartate aminotransferase forms from the apoenzyme with free pyridoxal and arsenate. Enzymes that convert R–O–PO3H2 into a diester often act on R–CH2–AsO3H2, a stable substrate analogue; then the product is unstable and hydrolyses to re-form the analogue, giving a futile cycle. For example, RNA polymerase acquires exonuclease activity in the presence of H2O3P–CH2–AsO3H2; adenylate kinase acquires ATPase activity in the presence of the arsonomethyl analogue of AMP. A recent observation is that HO–CH2–CHOH–CH2–CH2–AsO3H2 is a good substrate for glycerol-3-phosphate dehydrogenase. The product is unstable and eliminates arsenite, sharing this ability with other 3-oxoalkylarsonates. Thus this enzyme–catalysed oxidation is a lethal synthesis, in view of the toxicity of arsenite. Another unusual biochemical reaction of an arsonic acid is seen in the ability of a bacterium to use arsonoacetate as its sole source of carbon and energy. In contrast with the elimination of arsenite by 3-oxoalylarsonic acids, 3-oxoalkylphosphonic acids, R–CO–CH2–CH2–PO3H2, are stable. 2-Oxoalkylphosphonic acids, R–CO–CH2–PO3H2, however, are moderately unstable to hydrolysis, yielding phosphate and R–CO–CH3. 2-Oxoalkylarsonic acids, R–CO–CH2–AsO3H2, decompose in the same way, but much more readily, yielding arsenate. © 1997 by John Wiley & Sons, Ltd.  相似文献   

17.
The synthesis of six insulin fragments is described, in which various sequences of the two chains are linked by the disulfide bridge between A20 and B19. The fragments in question are: A20–21–B19–21, A20–21–B18–21, A20–21–B17–21, A19–21–B19–21, A16–21–B18–21 and A20–21–B12–21. In order to build up the simpler fragments the disulfide bridge was established by oxidation with iodine of two S-trityl cysteine peptides in which the carboxyl and amino groups were protected by the t-butyl and t-butyloxycarbonyl residue. From the mixture obtained the unsymmetrical cystine peptide was separated in all cases from the two symmetrical ones by counter-current distribution. In the synthesis of the more complex fragments advantageous use was made of smaller unsymmetrical fragments prepared as above but having one amino group protected by the N-trityl residue. After selective elimination of this group it was possible to lengthen the peptide chain at this position. The free peptides were obtained by removal of the protecting groups with strong acids, in particular concentrated hydrochloric acid. While in this deprotecting step the disulfide bond was stable, conditions are discussed under which disproportionation was observed. None of the six synthetic insulin fragments showed activity in stimulating rat adipose tissue to convert 14C-labelled glucose to CO2 in vitro.  相似文献   

18.
The neutral part of the acetone extract from the bark of Pinus luchuensis Mayer has been investigated and found to consist of alkanes (C22–C34) and triterpenes of serratene type. The triterpenes are 3β–methoxyserrat–14–en–21–one, serrat–14–en–3, 21–dione, 3β–hydroxyserrat–14–en–21–one, 3β–21α–dimethoxyserrat–l4–ene and 3β–methoxyserrat–14–en–21α–ol.  相似文献   

19.
The Ga–Al eutectic melt saturated with praseodymium was studied in the temperature range 572–1076 K by electromotive force (emf) method relative to the reference electrode (InL + PrIn3, where L is the liquid phase) in a LiCl–KCl–CsCl eutectic electrolyte. The partial molar thermodynamic functions (enthalpy, entropy, and Gibbs energy) of praseo dymium in a Ga–Al eutectic melt were calculated. According to the emf measurements of the two-phase Pr–Ga–AlL + intermetallic compound alloys equilibrated with the Ga–Al eutectic melt saturated with praseodymium, there are intermetallic compounds PrGa6, PrGa4, and Pr0.22Ga0.78(PrGa2) in the temperature ranges 572–741, 741–883, and 883–1076 K, respectively.  相似文献   

20.
On electrochemical initiation of alternating copolymerizations of styrene–acrylonitrile (AN) and styrene–diethyl fumarate (DEF) in the presence of ZnCl2, radical anions of AN–ZnCl2 and DEF–ZnCl2 complexes produced at the cathode were assumed to initiate copolymerization. In analogy with the cathode-initiated copolymerization, the radical anions of AN–ZnCl2 and DEF–ZnCl2, generated with the carbanions such as sodium naphthalene, disodium α-methylstyrene tetramer dianion, and butyllithium, were also found to produce alternating copolymers of styrene–AN and styrene–DEF. On the contrary, no polymers were obtained from methyl methacrylate (MMA)–styrene and methacrylonitrile (MAN)–styrene in the presence of ZnCl2 either with carbanions or by electrochemical reduction. Styrene–MAN–ZnCl2 yielded an alternating copolymer with carbanions upon introduction of oxygen.  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