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.
Cadmium acetate reacts with sulfamethoxazole (5‐methyl‐3‐isoxazolyl sulfanilamide) and with DMF / DMSO / pyridine to give the crystalline polymers [Cd(sulfamethoxazolato)2(L2)]n {L = DMF ( 1 ), DMSO ( 2 )} and [Cd(sulfamethoxazolato)2(Py)2]n·n(Py) ( 3 ). Complexes 1 , 2 and 3 confirm the tectonic character of the [Cd(sulfamethoxazolato)2(L)2] moieties and the remarkable ability of the {Cd(sulfamethoxazolato)4} fragments to be non selectively stabilized by monodentate ligands. In the polymeric assemblies of the title complexes the cadmium(II) atoms are linked through sulfamethoxazolato anions which alternate in their coordination with the isoxazolic N‐atoms and the aromatic amino groups. The chains of vicinal rings build tunnels along the crystallographic c axis.  相似文献   

2.
The kinetics of the reaction of substituted anilines with 2‐chloro‐5‐nitropyridine were studied in dimethyl sulfonide (DMSO) and dimethyl formamide (DMF) at different amine concentrations and temperatures in the range 45–60°C. In both solvents the reaction was not a base‐catalyzed one. A plot of ΔH# versus ΔS# for the reaction in DMSO and DMF gave good straight lines with isokinetic temperatures 128°C and 105°C, respectively. Good linear relationships were obtained from the plots of log k1 against σ° values at all temperatures with negative ρ values (?1.63 to ?1.28 in DMSO) and (?1.26 to ?0.90 in DMF). © 2002 Wiley Periodicals, Inc. Int J Chem Kinet 34: 645–650, 2002  相似文献   

3.
Silver nanoparticles (NPs) were synthesized in organic solvents. Spontaneous reduction of silver salts takes place in N,N′-dimethyl formamide (DMF) and dimethyl sulfoxide (DMSO) at room temperature. The formed colloids are not stable without a stabilizing agent, hence rarely used, and inexpensive organic molecules (β-cyclodextrin and cholic acid) were used as surface modifiers in DMF. The stabilization was successful; the Ag NPs remained stable for more than 3 months. Additionally, Ag NPs were prepared using Ag-2-ethylhexanoate and Na-citrate as capping agent in DMSO. The resulting NPs are stable, of 4.4 nm average size, and at the same time reactive for catalytic purposes. The interaction of Ag NPs with pollutant atmospheric gases (NO and SO2) was studied. UV–visible spectra show the oxidation of silver and the very efficient reduction of NO at room temperature. SO2 molecules are adsorbed on the NPs surface, causing their aggregation and precipitation.  相似文献   

4.
The spin-lattice relaxation time (T 1) of 23 Na was measured in solutions of NaClO 4 and (or) NaBr in formamide,N-methylformamide,N,N-dimethylformamide (DMF), MeCN, Me2CO, tetrahydrofuran (THF), and dimethyl sulfoxide (DMSO), and 133 Cs in a solution of CsCl in formamide. The values of (1/T 1) 0 obtained by extrapolation are discussed in terms of current theories of quadrupolar magnetic relaxation of ionic nuclei. A correlation was found between (1/T 1) 0 for 23 Na and Gutmann's donor numbers.For Part I, see ref. 1.  相似文献   

5.
Excess enthalpy, excess isobaric heat capacity, density, and speed of sound in mixtures of heavy water (D2O) + dimethylsulfoxide (DMSO), and D2O + dimethylformamide (DMF) were measured at 25‡C. The same properties of the mixtures of normal water + DMSO, and H2O + DMF were also measured to estimate isotope effects on the thermodynamic excess functions. Both DMSO and DMF are proton acceptors and thus form hydrogen bonds with water. Large negative excess enthalpies and volumes of mixing and excess isentropic compressibilities show that the hydrogen bonding structures of DMSO and DMF with water are stronger and more compact than those in pure water. The excess heat capacity of DMSO-containing mixtures changes sign from negative to positive with increasing water content. The deviations of the excess properties of D2O systems from those of H2O systems indicate that the hydrogen bonding structure with D2O is stronger and more compact.  相似文献   

6.
The intermolecular interaction between poly(methyl methacrylate) (PMMA) and polystyrene (PS) intetrahydrofuran (THF) and N,N'-dimethyl formamide (DMF) solvents was studied at 28℃ using a dilute solution viscometrymethod. Solvent is believed to play a key role in characterizing the viscosity behavior of the polymer solution. The intrinsicviscosity and viscosity interaction parameter were experimentally measured for the binary (solvent/polymer) and for theternary systems in two solvents. The compatibility of the polymer mixture was discussed in terms of the sign of △b_m. Theresults show that the compatibility of PMMA/PS blend in DMF is larger than that in THF.  相似文献   

7.
The enthalpies of solution of tetra-n-butylammonium bromide have been measured in mixtures of formamide (F) with water, N-methylformamide (NMF) with water, N,N-dimethylformamide (DMF) with water, F with NMF, DMF with F, and NMF with DMF at 25°C in the whole mole-fraction range. The enthalpies of solution vs composition profiles show a maximum value in the DMF-H2O and in the DMF-F systems. In the F-NMF and NMF-DMF mixturesn-Bu4NBr displays a nearly ideal behavior, whereas in the other solvent systems the excess enthalpy of solution deviates substantially from zero.  相似文献   

8.
RuX2(DMSO)4 (X=Cl,cis; Br,trans) undergoes ligand substitution in N,N-dimethylformamide (DMF) to give RuX2(DMSO)3DMF, which catalyzes the oxidation of secondary alcohols by NMO to ketones. Kinetics of the reaction catalyzed bytrans-RuBr2(DMSO)4 differed from that ofcis-RuCl2(DMSO)4. A mechanism is proposed involving the formation of Ru(IV)oxo species as the active intermediate and a rate expression is derived.  相似文献   

9.
The poly(o‐phenylenediamine) (PoPD) was synthesized from the monomer o‐phenylenediamine in various organic solvent medium viz. dimethyl sulfoxide (DMSO), N,N‐dimethyl formamide (DMF) and methanol using ammonium per sulfate as a radical initiator. The structure just like polyaniline derivative with free ?NH functional groups of the synthesized polymers confirmed by various standard characterizations was explained from the proposed polymerization mechanism. All the synthesized polymers were completely soluble in common organic solvent like DMSO and DMF because of the presence of polar free ?NH functional groups in its structure. The formation of polymer nanofiber by reverse salting‐out process was confirmed, and the synthesized polymer in DMSO medium was the best polymer in terms of nano‐morphology as well as conducting properties. Interestingly, the average DC conductivity of undoped polymer film was recorded as 2.21 × 10?6 Scm?1 because of induced doping through self charge separation. Moreover, the conductivity of the polymer film was further increased to 1.16 × 10?3 Scm?1 after doping by sulfuric acid. Copyright © 2016 John Wiley & Sons, Ltd.  相似文献   

10.
Styrene–acrylonitrile (St–AN) copolymers of three compositions—27.4 mole-% (SA1); 38.5 mole-% (SA2); and 47.5 mole-% (SA3) acrylonitrile—and styrene–methyl methacrylate (St–MMA) copolymer (SM) of 46.5 mole-% methyl methacrylate were prepared by bulk polymerization at 60°C with benzoyl peroxide as the initiator, and were then fractionated. The molecular weights of unfractionated and fractionated samples were determined by light scattering in a number of solvents. The [η] versus M?w relations at 30°C were established for SA1, SA2, SM, and polystyrene (PSt) in ethyl acetate (EAc), dimethyl formamide (DMF), and γ-butyrolactone (γ-BL), and for SA3 in methyl ethyl ketone (MEK), DMF, and γ-BL. Second virial coefficients A2 and the Huggins constant were determined. From values of A2 and the exponent a of the Mark–Houwink relation it is seen that the solvent power for samples SA1, SA2, and PSt is in the order EAc < γ-BL < DMF, while for sample SA3 the solvent power is in the order MEK < γ-BL < DMF. The solvent power decreases with an increase in AN content. The solvent power of the three solvents used for SM copolymer sample is practically the same within experimental errors. From the a values it is concluded that in a given solvent the copolymer chains are more extended than the corresponding homopolymers.  相似文献   

11.
The kinetics of reactions of acrylamide derivatives (acrylamidotrihydroxymethylmethane (TA), sodium 4-acrylamidobutanoate (AA3), sodium 6-acrylamidohexanoate (AA5), and sodium 11-acrylamido-undecanoate (AA10)) with piperidine and morpholine in water (for TA, also in DMF, DMSO, and formamide) is studied at 293 K. These compounds are weak surfactant monomers. The critical concentrations of micelle formation (CCM) for them are determined. The self-association of AA3, AA5, and AA10 producing micelles results in a decrease in their reactivity compared to the monomeric forms. The rates of the reactions of surfactant monomers (SM) with morpholine and piperidine are described by the second-order rate law w 0 = k[SM]0[Amine]0. An empirical equation is derived that relates the CCM values to the rate constant for the reaction of a surfactant monomer with a secondary amine with charges on the -carbon and oxygen atoms of the amide group of a surfactant monomer. The rates of the reactions of TA with piperidine and morpholine are determined by the electrophilicity (acidity) of the medium, which is favorable for the Michael reaction.  相似文献   

12.
Agarose sols have been seen for long as solutions of flexible chains that, on cooling, produce thermoreversible gels through double-helix formation. Investigations of the chain conformation in the sol state by small-angle neutron scattering reveals instead a rigid chain with a very large persistence length (lp > 9 nm). The chain cross-section radius and mass per unit length correspond to characteristics of helices as those described by Foord and Atkins. These results lead one to a reappraisal of the occurrence of double helices in the gelation process, as they rather suggest a transition of the type loose-single helixtight single helix. Studies of gels from agarose/water/cosolvent where the cosolvent is Dimethyl Sulfoxide (DMSO), Dimethyl Formamide (DMF), and Methyl Formamide (MF) have led one to conclude on the formation of agarose/water/ cosolvent ternary complexes. The contrast variation method by neutron scattering gives further support to this assumption. Finally, determination of the gel nanostructure allows one to account for the two regimes observed for the variation of the elastic modulus vs concentration.  相似文献   

13.
The inverse Kirkwood-Buff integral (IKBI) approach is applied, as far as the relevant data exist in the literature, to the preferential solvation of Li+, Na+, K+, Rb+, Cs+, Ag+, and Tl+ in aqueous mixtures of methanol, ethanol, 1,2-ethanediol (EG), acetonitrile, formamide, N,N,-dimethylformamide (DMF), N,N,N′,N′,N″,N″-hexamethyl phosphoric triamide (HMPT), and dimethyl sulfoxide (DMSO). In aqueous EG and formamide the preferential solvation is very small. Water is the preferred solvent in the solvation shells in aqueous methanol, ethanol and acetonitrile (except for Ag+ in the latter solvent) but the co-solvent is preferred in aqueous mixtures of DMF, HMPT, and DMSO over most or all of the composition range. For the latter three mixtures the larger donicity of the co-solvents causes their preference, whereas where water is preferred over other protic solvents, it is the small size of the water molecules that appears to be the cause.  相似文献   

14.
Bacillus sp. HR-08 screened from soil samples of Iran, is capable of producing proteolytic enzymes. 16S rDNA analysis showed that this strain is closely related to Bacillus subtilis, Bacillus licheniformis, Bacillus pumilus, Bacillus mojavensis, and Bacillus atrophaeus. The zymogram analysis of the crude extract revealed the presence of five extracellular proteases. One of the proteases was purified in three steps procedure involving ammonium sulfate precipitation, DEAE-Sepharose ionic exchange and Sephacryl S-200 gel filtration chromatography. The molecular mass of the enzyme on SDS-PAGE was estimated to be 29 kDa. The protease exhibited maximum activity at pH 10.0 and 60 °C and was inhibited by PMSF but it was not affected by cysteine inhibitors, suggesting that the enzyme is a serine alkaline protease. Irreversible thermoinactivation of enzyme was examined at 50, 60, and 70 °C in the presence of 10 mM CaCl2. Results showed that the protease activity retains more than 80% and 50% of its initial activity after incubation for 30 min at 60 and 70 °C, respectively. This enzyme had good stability in the presence of H2O2, nonionic surfactant, and local detergents and its activity was enhanced in the presence of 20% of dimethyl sulfoxide (DMSO), dimethyl formamide (DMF) and isopropanol. The enzyme retained more than 90% of its initial activity after pre-incubation 1 h at room temperature in the presence of 20% of these solvents. Also, activation can be seen for the enzyme at high concentration (50%, v/v) of DMF and DMSO.  相似文献   

15.
Summary The existence of methyl sulfite under KF-conditions was proved by IR spectroscopy. Methyl sulfate, the expected reaction product of methyl sulfite in the KF-reaction, has been isolated from used KF-solutions. Photometry served as an independent method to check the electrochemical work of Verhoef and Barendrecht [17, 18]. Their results about the pH-dependence of the reaction rate constant k 3 and the interdependence of k 3 and the iodide concentration have been verified. Substitution of H2O by D2O showed that water is not involved in the rate controlling step of the KF-reaction. The cause for the increase of KF-reaction speed, when using chloroform and formamide in mixture with methanol, has been clarified. In chloroform/methanol the stability constant of triiodide is decreased and the equilibrium shifted towards the more reactive iodine species. In contrast to this indirect effect of chloroform formamide seems to be directly involved in the KF-reaction. In 50% formamide the reaction rates of both I2 and I 3 are increased. Presumably a formamide-SO2 analogon of methyl sulfite is oxidized and hydrolyzed faster than the pure methyl esters.  相似文献   

16.
A series of Co(II), Ni(II), and Cu(II) complexes ML?·?3H2O have been synthesized with Schiff bases derived from 3-substituted-4-amino-5-mercapto-1,2,4-triazole and 5-formyl-6-hydroxy coumarin. The complexes are insoluble in common organic solvents but soluble in DMF and DMSO. The measured molar conductance values in DMF indicate that the complexes are non-electrolytes. In view of analytical, spectral (infrared, UV-Vis, ESR, TG, and FAB-mass), and magnetic studies, it has been concluded that all the metal complexes possess octahedral geometry in which ligand is coordinated to metal through azomethine nitrogen, phenolic oxygen, and sulfur via deprotonation. The Schiff bases and their complexes have been screened for antibacterial (Escherichia coli, Staphylococcus aureus, Pseudomonas aeruginosa, and Salmonella typhi) and antifungal activities (Aspergillus niger, Aspergillus flavus, and Cladosporium) by the minimum inhibitory concentration method. DNA cleavage is studied by agarose gel electrophoresis.  相似文献   

17.
The adducts formed by Eu(DK)2(CDK) and Eu(DK)(CDK)2 [DK = achiral β-diketone ligand, and CDK = chiral β-diketone ligand] with either dimethyl sulfoxide (DMSO) or N,N-dimethyl formamide (DMF) have been studied by means of circularly polarized luminescence spectroscopy. While extremely strong stereoselective effects had been observed in earlier studies involving Eu(CDK)3 adducts, analogous phenomena were found to be much smaller in the mixed-ligand compounds. These observations imply that steric interactions involving bulky camphorato ligands are significantly greater than those involving the achiral (and planar) β-diketone ligands.  相似文献   

18.
The copolymerization of styrene with methyl methacrylate (S/MMA = 4/1) or acrylonitrile (S/AN = 1/1) in the presence of ethylaluminum sesquichloride (EASC) yields 1/1 copolymer in toluene or chlorobenzene. In chloroform the S-MMA-EASC polymerization yields 60/40 copolymer while the S-AN-EASC polymerization yields 1/1 copolymer. In the presence of EASC, styrene-α-chloroacrylonitrile yields 1/1 copolymer (DMF or DMSO), S-AN yields 1/1 copolymer (DMSO) or radical copolymer (DMF), S-MMA yields radical copolymer (DMF or DMSO), α-methylstyrene-AN yields radical copolymer (DMSO) or traces of copolymer (DMF), and α-MS-methacrylo-nitrile yields traces of copolymer (DMSO) or no copolymer (DMF). When zinc chloride is used as complexing agent in DMF or DMSO, none of the monomer pairs undergoes polymerization. However, radical catalyzed polymerization of isoprene-AN-ZnCl2 in DMF yields 1/1 alternating copolymer. The copolymerization of S/MMA in the presence of EASC yields 1/1 alternating copolymer up to 100°C, while the copolymerization of S/AN deviates from 1/1 alternating copolymer above 50°C. The copolymerization of S/MMA deviates from 1/1 copolymer at MMA/EASC mole ratios above 20 while the copolymerization of S/AN deviates from 1/1 copolymer at MMA/EASC ratios above 50.  相似文献   

19.
Co(II), Ni(II), and Cu(II) complexes, ML2 · 2H2O have been synthesized with Schiff bases derived from m-substituted thiosemicarbazides and 2-methoxy benzaldehyde. The complexes are soluble in DMF/DMSO and non-electrolytes. From analytical, spectral (IR, UV-Vis, ESR, and FAB-mass), magnetic and thermal studies octahedral geometry is proposed for the complexes. The redox behavior of the complexes was investigated using cyclic voltammetry. The Schiff bases and their metal complexes have been screened for antibacterial (Escherichia coli, Staphylococcus aureus, Pseudomonas aeruginosa, and Salmonella typhi) and antifungal activities (Aspergillus niger, Aspergillus flavus, and Cladosporium) by Minimum Inhibitory Concentration method. DNA cleavage is studied by agarose gel electrophoresis method.  相似文献   

20.
Hydroxyl-terminated hyperbranched polyesters (HBPEs) with aromatic/aliphatic structure were synthesized at three different monomer mole ratios (A2/B3=1, 0.9, and 0.6, respectively) by melt polycondensation of isophthalic acid and trimethylolpropane via A2+ B3 approach. Fourier transform infrared (FTIR) spectroscopy indicated that the expected HBPEs, which showed excellent solubility in a variety of polar solvents such as N, N-dimethyl formamide (DMF), N-methyl-2-pyrrolidone (NMP), dimethyl sulfoxide (DMSO), and terahydrofuran (THF), were produced without gelation during the polymerization. The weight-average molecular weight of HBPEs ranged from 7014 to 8306 and their inherent viscosities were varied from 0.17 to 0.34 dL/g. The degree of branching of the HBPEs was estimated to be 0.39–0.49 by1H-NMR and 13C-NMR measurement. Their glass transition temperature measured by differential scanning calorimetry (DSC) was between 64 and 86°C. The thermogravimetric analysis (TGA) measurement revealed that HBPEs had ca 8.5% weight-loss at 300°C in N2.  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