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 640 毫秒
1.
Nucleosides and Nucleotides. Part 12. Synthesis of Dinucleoside Monophosphates Containing 1-(2′-Deoxy-β-D -ribofuranosyl)-2(1H)-pyrimidone The connexion of the modified nucleoside 1-(2′-deoxy-β-D -ribofuranosyl)-2(1H)-pyrimidone (Md, 2 ) with the natural nucleotides pTd and pGd is described. The protected dinucleoside monophosphates (MeOTr)MdpTd ( 6 ) and (MeOTr)MdpG ( 9 ) were prepared by the standard phosphodiester method using DCC as condensing agent. MdpTd ( 7 ) was obtained by treatment of 6 with formic acid/methanol 7 : 3 at 0° 9 was converted to the free dinucleoside monophosphate MdpGd ( 11 ) by removing the N-isobutyryl- and p-methoxytrityl protecting group on consecutive treatment with 0.04N CH3ONa in CH3OH and HCOOH/CH3OH 7:3 at 0° respectively. Enzymatic degradation of the free dinucleoside of the free dinucleoside monophosphates 7 and 11 yielded the corresponding nucleosides and nucleotides in the correct ratios.  相似文献   

2.
3.
Dynamic viscoelasticity measurements were carried out for concentrated solutions of linear d-glucans in BmimCl to examine the effect of the linkage between repeating units of glucose on the rheological properties. The values of molecular weight between entanglements (M e) were determined for four d-glucans: curdlan, pullulan, cellulose, and amylose. From the concentration dependence of M e, the value of M e in the molten state (M e,melt) for each d-glucan was estimated as a material constant. The order of M e,melt became cellulose?<?pullulan?<?curdlan?<?amylose, indicating that the linkage is actually influential in M e,melt for the linear d-glucans. The relationship between M e,melt and the molecular structure of the d-glucans were discussed assuming that the values of M e,melt for the d-glucans primarily reflect the chain stiffness such as the characteristic ratio C on the analogy of synthetic polymers. Although the trend was not so clear, it was shown that N unit is a decreasing function of C .  相似文献   

4.
The reaction between formic acid and bromine in strongly acid aqueous media at 298 K was studied by absorption spectrophotometry (λ = 447 nm). Reaction rates, expressed as R = -d[Br2]/dt, depend on the concentrations of HCOOH (0.3–2.4M), Br2[(2.7–13.6) × 10?3M], H+ (0.03–2.0M), and Br? (up to 0.6M). The mechanism with k1 = 20.2 ± 1.2 M?1 sec?1, pK2 = 3.76, pK3 = ?1.20, accounts for all experimental observations. Br3? and HCOOH can be considered unreactive within experimental error. Apparent deviations from the basic mechanism at higher acidities can be quantitatively ascribed to the nonideality of ionic species.  相似文献   

5.
The rate constants k CR of orthointo para-positronium conversion reactions promoted by 3d complexes are linearly correlated with the electron delocalization of unpaired metal electrons caused by ligands, as measured by the ratio between the interelectronic repulsion parameters in complexes and in the free gaseous ions (1 > 0). The values may be evaluated by means of the following empirical equation: = 1 t M h L, where t M and h L are characteristic constants of 3d ions and ligands, respectively. By measuring k CR and applying the correlations mentioned above, h L constants may be derived providing that t M is known. The method was used to measure the h L value of N 3, not yet known, and to check that of Cl ligand.  相似文献   

6.
The reaction of peroxomonophosphoric acid and hydrazinium ion in acid perchlorate solutions occurs as per stoichiometry (i), and the rate law (ii) at large [N2H5 +], where K′d is the first acid dissociation constant of H3PO5 and k 1 and k 2 are rate constants found to be 2.6 × 10?4 s?1 and 5.0 × 10?2 M?1 s?1, respectively, at 35°. The reaction is greatly catalyzed by iodide ions. The mechanism involves a redox cycle I?/I2 and the rate is independent of [N2H5 +] in the presence of iodide ions. K′d was found to be 0.55 M?1 and independent of temperature.  相似文献   

7.
Two chiral fluorescent chemosensors 1 and 2 were synthesized, and the structure characterized by IR, 1H NMR, 13C NMR, MS spectra and elemental analysis. Their recognition ability was studied in aqueous solution (Tris–HCl buffer pH 7.4, MeOH/H2O = 1:1) through fluorescence spectra. Receptors 1 and 2 showed a good binding ability to the copper ion. The host 1-Cu2+ complex showed a chiral recognition ability to mandelate anions with a preferable binding to l-mandelate than d-mandelate anions. The host 1-Cu2+ complex and l- or d-mandelate anions formed 1:1 stoichiometric complex. The binding constant for l-mandelate is 576 M−1, whereas that for d-mandelate is only 38 M−1, which can be distinguished by the different change of fluorescence intensity.  相似文献   

8.
The effect of dissociation on the reactivity and inhibition capacity of p-nitro-phenol /p-NPH/ towards orthopositronium atom has been investigated by performing positron lifetime measurements in methanol and methanol containing 0.5M NaOH solutions, respectively. In the latter case the solute exists in anionic form /p-NP/. It has been found that as a result of dissociation, both the rate constant /k/ and the inhibition coefficient // decrease significantly. Their values are k=/7.5±0.4/×108 s–1 M–1 and =6.8±0.2 M–1 for p-NPH and kd=/0.6±0.2/×108 s–1 M–1 and d=4.0±0.2 M–1 for p-NP. The possible reasons for these effects are discussed.  相似文献   

9.
Extraction of carbazole in heptane was performed at 25±1°C with an aqueous dimethyl sulfoxide (DMSO) medium containing -cyclodextrin (CD) at consecutive concentrations in the range of 0–10 mM. The fluorescence intensity of carbazole remaining in the heptane phase was measured by synchronous scanning fluorimetry. The apparent formation constant (K f) for a 1:1 carbazole: CD inclusion complex in water-DMSO medium was determined by using a linear plot of the distribution ratio calculated from the fluorescence intensities vs. the -CD concentration. The values thus obtained ranged from 477 M–1 in a 10% v/v DMSO medium to 12.1 M–1 in a 60% v/v medium. Good linear relationships were observed between logK f and the DMSO concentration ([DMSO]), and also between logK f and the logarithm of the distribution coefficient (K d) for carbazole. The formation constant in 100% water was estimated to be approximately 1.0×103 M–1 on the basis of the logK f vs. [DMSO] and the logK f vs. logK d correlations.  相似文献   

10.
Summary The calculation of capacity factors, k, from net retention times, tR, and the corresponding dead times, tM, at different temperatures suffers from the limited accuracy of the tM values. If the temperature coefficient racy of the tM values. If the temperature coefficient d ln k/d (1/T) only is required, it is sufficient to determine net retention times (tR)p at constant inlet pressure pi for different temperatures, since the temperature dependence of (tM)p can be assumed as (tM)p=A·eB/T, with B being approximately independent of the column inlet pressure and of the nature of the carrier gas. The extrapolation and interpolation of (tR)p may be either performed by linear regression or graphically with a nomogram for ln (tR)p versus 1/T. The resolution factor, , of two components, e.g. enantiomers which are resolved on a chiral stationary phase, can be treated in a similar way. Examples are given for the resolution of enantiomers of two non-proteinogenic amino acids on the new polysiloxane phase L-Chirasil-CPG.  相似文献   

11.
Glucose oxidase from Aspergillus niger, the specific enzyme for β-d-glucose oxidation, can also oxidize other related saccharides at very slow or negligible rates. The present study aimed to compare the kinetics of d-glucose oxidation using immobilized glucose oxidase on bead cellulose for the oxidation of related saccharides using the same biocatalyst. The significant differences were observed between the reaction rates for d-glucose and other saccharides examined. As a result, k cat/K M ratio for d-glucose was determined to be 42 times higher than d-mannose, 61.6 times higher than d-galactose, 279 times higher than d-xylose, and 254 times higher than for d-fructose and d-cellobiose. On the basis of these differences, the ability of immobilized glucose oxidase to remove d-glucose from d-cellobiose, d-glucose from d-xylose, and d-xylose from d-lyxose was examined. Immobilized catalase on Eupergit and mixed with immobilized glucose oxidase on bead cellulose or co-immobilized with glucose oxidase on bead cellulose was used for elimination of hydrogen peroxide from the reaction mixture. The accelerated elimination of d-glucose and d-xylose in the presence of co-immobilized catalase was observed. The co-immobilized glucose oxidase and catalase were able to decrease d-glucose or d-xylose content to 0–0.005% of their initial concentrations, while a minimum decrease of low oxidized saccharides d-xylose, d-cellobiose, and d-lyxose, respectively, was observed.  相似文献   

12.
The molecular structures of cyclic group 6 transition-metal (M = Cr, Mo, W, Sg) oxides (M 3O 9 0/1?/2? ) species have been optimized at density functional theory (DFT) levels. The photoelectron spectra (PES) of M 3O 9 ? (M = Cr, Mo, W) were calculated at the time-dependent DFT and approximate coupled-cluster singles doubles (CC2) levels and compared with experimental results. The CC2 calculations did not yield any reliable PES, whereas the molecular structures can be identified by comparing PES obtained at the DFT level with experiment. Magnetically induced current densities were calculated at the DFT level using the gauge-including magnetically induced current (gimic) approach. The current strengths and current pathways of the neutral M3O9 and the dianionic M3O 9 2? (M = Cr, Mo) oxides were investigated and analyzed with respect to a previous prediction of d-orbital aromaticity for Mo3O9 anions. Current-density calculations provide ring-current strengths that are used to assess the degree of aromaticity. Comparison of current-density calculations and calculations of nucleus-independent chemical shifts (NICS) shows that NICS calculations are not a reliable tool for determining the degree of aromaticity of the metal oxides.  相似文献   

13.
The irreducible representations consisting of linear combinations of cluster atomic f-orbitals are obtained for f, f,f and f orbitals in M3(D 3h ), M4(D 4h ), M4(T d ), M6(O h ) and M8(O h ) clusters. The charge overlap of any pair of two atoms in a cluster is decomposed in terms of a set of coefficients for the , , , and overlaps, respectively. A vector method is devised for this decomposition which may be extended to any arbitrary orientation and to higher orbitals. The decomposition coefficients represent fundamental geometrical properties of the cluster and are applicable to clusters of arbitrary dimensions.Contribution No. 0004 ITI Basic Research Laboratory. Presented in part at the 187th ACS National Meeting in St. Louis, MO, USA  相似文献   

14.
Cationic polymerizations of trioxane in 1,2‐ethylene dichloride and benzene were heterogeneous and reversible. Phase separation accompanying with crystallization occurred during the polymerization. Three morphological changes were found in the course of the polymerization as were investigated by dilatometry and precipitation method. Based on the findings of morphological changes and three reversible processes for the polymerization, a rate equation was proposed to describe the polymerization. The proposed rate equation was fairly good in describing the experimental data, and kinetics constants including Kp, Kd, Kp′, Kd′, M, M, and Kdis/Kcr for the polymerization at 30, 40, and 50°C in 1,2‐ethylene dichloride and benzene were obtained. Factors that affected the kinetics constants were discussed. © 1999 John Wiley & Sons, Inc. J Polym Sci A: Polym Chem 37: 483–492, 1999  相似文献   

15.
Based on the phenomenon of freezing point depression of a solvent byT, experimental evidence is presented to show that the distance between the junction points can be calculated fromT. Direct measurements of the temperature-time-curve of the cooling network and the Differential Scanning Calorimetry offer the determination ofT. Except the mean distances ¯d c in dependence on cross-linking density, swelling degree, and other network parameters, the distribution of the distance between the junction pointsH(dc) can be determined, which allows conclusions on the course of cross-linking reaction. This paper attempts to give experimental evidence of influences of the breadth ofH(dc) on application-relevant properties.  相似文献   

16.
One-Electron Redox Reactions of Octaphenyl[4]radialene: Generation and ESR/ENDOR Characterisation of Its Radical Anion and Radical Cation The cyclovoltammograms of octaphenyl[4]radialene in DMF or THF at room temperature disclose each two quasireversible reduction and oxidation potentials at ?1.4 V/?1.7 V and +0.7 V/+0.9 V. Accordingly, both the radical anion and the radical cation can be generated: Ph8C by K metal mirror reduction of a [2.2.2]cryptand containing THF solution, and Ph8C by TI3⊕(?OOCCF3)3 oxidation in H2CCl2. Their ESR/ENDOR and General Triple spectra differ considerably in the number of resolved 1H couplings (M·?: 5 and M·⊕: 3) as well as in their spectral widths (M·?: a1H 0.090 to 0.017 mT; M·⊕: a1H 0.066 to 0.023 mT) suggesting different changes in the D2d structure of the neutral molecule on electron uptake or extrusion.  相似文献   

17.
Phosphorous-bridged bisphenoxy titanium complexes were synthesized and their ethylene polymerization behavior was investigated. Bis[3-tert-butyl-5-methyl-2-phenoxy](phenyl)phosphine tetrahydrofuran titanium dichloride (4a) was obtained by treatment of 3 equiv of n-BuLi with bis[3-tert-butyl-2-hydroxy-5-methylphenyl](phenyl)phosphine hydrochloride salt (3a) followed by TiCl4(THF)2 in THF. THF-free complexes 5a-5d were synthesized more conveniently by the direct reaction of MOM-protected ligands (2a-2d) with TiCl4 in toluene. X-ray analysis of 4a revealed that the ligand is bonded to the octahedral titanium (IV) center in a facial fashion and two chlorine atoms possess cis-geometry. Complexes 4a and 5a-5d were utilized as catalyst precursors for ethylene polymerization. Complex 5c gave high molecular weight polyethylene (Mw = 1,170,000, Mw/Mn = 2.0) upon activation with Al(iBu)3/[Ph3C][B(C6F5)4] (TB). Ethylene polymerization activity of 5d activated with Al(iBu)3/TB reached 49.0 × 106 g mol (cat) −1 h−1.  相似文献   

18.
The crystal structures of the two oxides Bi46M8O89 (M=P, V) have been solved from single crystals X-ray data at room temperature. Bi46P8O89 crystallizes in the monoclinic symmetry (space group C2/m) with the cell parameters , , and β=112.14(3)°. The symmetry of Bi46V8O89 is also monoclinic but the space group is P21/c with the unit-cell parameters: , , and β=107.27(3)°. Both structures derive from an oxygen deficient fluorite-type structure where the Bi and M cations (M=P, V) are ordered in the framework. The structures are characterised by isolated MO4 tetrahedra (M=P, V) which contradicts the previous results. The difference between the two structures is only due to a different order of the M atoms (M=P, V) in the fluorite-type superstructure. It will be shown that some oxygen sites are partially occupied in both structures which can explain the ion conduction properties of these phases. A structural building principle will be proposed that can explain the large domain of solid solution related to the fluorite-type observed in both systems.  相似文献   

19.
IR spectroscopy is applied to study polyelectrolyte films of 0.1–5 μm thickness. This approach permits to obtain not only the usual chemical information about polymer conformation, chemical bonds, interacting groups, and molecular/ionic state, but also important structural parameters of the film, such as total thickness, average thickness of each layer, and sparseness of polymer packing. All these information can be obtained from one single sample.

IR cell for thickness measurements of a polyelectrolyte film. The pathlength d of the cell is measured twice: before and after film deposition. The film thickness, Δd = dedf, can be calculated using the interference maxima that determine values de and df.  相似文献   


20.
Three molar mass series were produced by different methods of degradation (namely ultrasonic (seven samples), oxidation (seven samples) and autoclaving (eight samples)) from a methylhydroxyethylcellulose (MHEC) sample with an average degree of substitution (DS) of 1.3, a molar degree of substitution (MS) of 0.46, a radius of gyration of 67 nm and a weight-average molar mass, M w, of 318,000 g/mol. The degraded samples were characterized in terms of their molar mass and particle size together with their respective distributions with a hyphenated apparatus consisting of size exclusion chromatography and multi-angle laser light scattering and concentration detector (SEC/MALLS/DRI) at 25 °C in 0.1 M NaNO3 solution (with 200 ppm NaN3 as antibactericide). The refractive index increment was determined as dn/dc = 0.135 cm3/g. It was possible to reduce the weight-average molar mass down to approximately 10% of the initial molar mass using all the methods. In a comparison of the three degradation methods it was shown that only ultrasonic degradation retains the monomodal distribution, whereas the other two degradation methods yield a bimodal molar mass distribution. Consequently, only ultrasonic degradation represents a suitable method for producing homologous molar mass series. An R GM relationship of R G = 0.0511 × M 0.56 was established for the sample used in this case, and from this it was possible to calculate an []–M relationship of [] = 0.3587 × M 0.68.  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