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.
Abstract— Inactivation of tobacco mosaic virus RNA (TMV-RNA) by u.v. radiation is slower in D2O than in H2O, and TMV-RNA which has been inactivated in D2O is photoreactivated faster (on Pinto bean) than TMV-RNA which has been inactivated in H2O. The maximum amount of photoreactivation is unaffected by the solvent, H2O or D2O, present during irradiation. These deuterium isotope effects for inactivation and photoreactivation suggest that pyrimidine hydrates are photoreactivable lesions on Pinto bean.  相似文献   

2.
Diaquacobinamide (H2O)2Cbi2+ or its conjugate base hydroxyaquacobinamide (OH(H2O)Cbi+)) can bind up to two cyanide ions, making dicyanocobinamide. This transition is accompanied by a significant change in color, previously exploited for cyanide determination. The reagent OH(H2O)Cbi+ is used in excess; when trace amounts of cyanide are added, CN(H2O)Cbi+ should be formed. But the spectral absorption of CN(H2O)Cbi+ is virtually the same as that of OH(H2O)Cbi+. It has been inexplicable how trace amounts of cyanide are sensitively measured by this reaction. It is shown here that even with excess OH(H2O)Cbi+, (CN)2Cbi is formed first due to kinetic reasons; this only slowly forms CN(H2O)Cbi+. This understanding implies that CN(H2O)Cbi+ will itself be a better reagent.  相似文献   

3.
On the basis of DFT calculations (B3LYP/6‐311+G**), the possibility to include solvent effects is considered in the investigation of the H2O‐exchange mechanism on [Be(H2O)4]2+ within the widely used cluster approach. The smallest system in the gas phase, [Be(H2O)4(H2O)]2+, shows the highest activation barrier of +15.6 kcal/mol, whereas the explicit addition of five H‐bonded H2O molecules in [{Be(H2O)4(H2O)}(H2O)5]2+ reduces the barrier to +13.5 kcal/mol. Single‐point calculations applying CPCM (B3LYP(CPCM:H2O)/6‐311+G**//B3LYP/6‐311+G**) on [Be(H2O)4(H2O)]2+ lower the barrier to +9.6 kcal/mol. Optimization of the precursor and transition state of [Be(H2O)4(H2O)]2+ within an implicit model (B3LYP(CPCM:H2O)/6‐311+G** or B3LYP(PCM:H2O)/6‐311+G**) reduces the activation energy further to +8.3 kcal/mol but does not lead to any local minimum for the precursor and is, therefore, unfavorable.  相似文献   

4.
Viscosities were measured for the ternary aqueous systems NaCl–mannitol(C6H14O6)–H2O, NaBr–mannitol–H2O, KCl–mannitol–H2O, KCl–glycine(NH2CH2COOH)–H2O, KCl–CdCl2–H2O, and their binary subsystems NaCl–H2O, KCl–H2O, NaBr–H2O, CdCl2–H2O, mannitol–H2O, and glycine–H2O at 298.15 K. A powerful new approach is presented for theoretical modeling of the viscosity of multicomponent solutions in terms of the properties of their binary solutions. In this modeling, the semi-ideal solution theory was used to associate the solvation structure formed by each ion and its first solvation shell in a binary solution with the solvation structure of the same ion and its first solvation shell in multicomponent solutions. Then, the novel mechanism proposed by Omta et al. (Science, 301:347–349, 2003) for the effect of a single electrolyte on the viscosity of water was extended to describe the influence of solute mixtures on the viscosity of water, including electrolyte mixtures, nonelectrolyte mixtures, and mixtures of electrolytes with nonelectrolytes. The established simple equation was verified by comparison with measured viscosities and viscosities reported in literature. The agreements are very impressive. This formulation provides a powerful new approach for modeling this transport property in solutions. It can stimulate further research in establishing a dynamical analogue to that formulated for the thermodynamics of multicomponent solutions. It is also very important for the study of hydration of ions.  相似文献   

5.
Ab initio calculations for the interacting system of lower excited states of planar and bent H2CO with H2O have been carried out with a minimum basis set, using the recently proposed electron-hole potential method. The blue shifts of the η-π* transition are evaluated as 1100 and 1420 cm?1 for the singlet and triplet transitions, respectively. In the η-π* states of bent H2CO, the most stable geometry is one in which an H2O hydrogen atom is coordinated to the carbon atom.  相似文献   

6.
A study for AlF3 crystallization from water solution was performed in the temperature range 100 to 200°C.Four solid phases were found to be precipitated, AlF3.3H2O (up to ca.120°C, cubic α-AlF3.H2O (decomposition of AlF3.3H2O in suspension), hexagonal β -AlF3.H2O (direct from solution) and the hydroxyfluoride Al(OH,F)3.H2O with an F/Al ratio of ca. 2.5 (hydrolysis of AlF3). The extent of hydrolysis was established as a function of the initial AlF3 concentration.X-ray diffraction and thermogravimetric data for the monohydrates were given and differences between the two indicated.  相似文献   

7.
The H2 and CH4 chemical ionization mass spectra of a series of series of substituted benzoic acids and substituted benzyl alcohols have been determined. For the benzoic acids the major fragmentation reactions of the protonated molecule involve elimination of H2O or elimination of CO2, the latter reaction involving migration of the carboxylic hydrogen to the aromatic ring. For the benzyl alcohols the major fragmentation reactions of [MH]+ involve loss of H2O or CH2O, analogous to the CO2 elimination reaction for the benzoic acids. It is shown that the CO2 and CH2O elimination reactions occur only when a conjugated aromatic ring system is present, and that for the carboxylic acid systems, methyl groups and, to a lesser extent, phenyl groups are capable of migrating. The only discernible effect of substituents on the fragmentation of [MH]+ is an enhancement of the H2O loss reaction in the benzoic acid system when an amino, hydroxyl, or halogen substituent is ortho to the carboxyl function. This ‘ortho’ effect, which differs in scope from that observed in electron impact mass spectra, is attributed to an intramolecular catalysis by the ortho substituent of the 1,3 hydrogen migration in the carbonyl protonated acid followed by H2O elimination. Apparently, this route is favoured over the direct elimination of H2O from the carbonyl protonated acid, since the latter has a high activation energy barrier because of unfavourable orbital symmetry restrictions.  相似文献   

8.
9.
The dynamic viscoelastic behavior of a concentrated solution of silk fibroin dissolved in the “MU” solvent is measured. The dynamic viscosity η′ and dynamic elasticity G′ increase with increasing concentration of silk fibroin at constant frequency; however, the increasing frequency decreases η′ and G′ at a constant concentration of silk fibroin. When the mixing ratio of C2H5OH/H2O in the “MU” solvent is increased at a constant concentration of LiBr·H2O, η′ and G′ sharply increase at constant frequency. If the LiBr·H2O concentration is varied in the “MU” solvents whose ratio of C2H5OH/H2O is kept constant at 100 : 0, both η′ and G′ are greater for LiBr·H2O concentrations of 50% by weight compared to concentrations of 40% by weight. The dependence of η′ on the temperature of the solution can be predicted by Andrade's viscosity equation. Spinnability improves when the SF concentration is increased. © John Wiley & Sons, Inc. J Polym Sci A: Polym Chem 35 : 1955–1959, 1997  相似文献   

10.
The ground and low-lying excited states of H3O(H2O) k radicals are studied. The character of the unpaired electron localization in the systems is analyzed, and the relative probability of the radical dissociation onto a water cluster and atomic hydrogen is estimated. Reaction coordinates of the dissociation are constructed and conditions of metastable existence of an H3O radical are determined. Structures, in which H3O can spontaneously dissociate, are found. Lifetimes of H3O(H2O) k clusters before the hydrogen atom detachment at the initial conditions of two kinds, namely, upon the vertical attachment of an electron to H3O+(H2O) k cations and upon the vibrational excitation of metastable neutral H3O(H2O) k systems, are estimated.  相似文献   

11.
Zusammenfassung Es wurden die Verbindungen HYT *·4 H2O, Y4 T 3·14 H2O, LiYT·4 H2O, NaYT·5 H2O, KYT·3 H2O, RbYT·4 H2O, CsYT·4 H2O, NH4YT·3 H2O, K2YTOH·4 H2O, K3YT(OH)2·4 H2O, K4YT(OH)3·3 H2O, K5YT(OH)4·3 H2O, KYH4 T 2·3 H2O, K2YH3 T 2·5 H2O, K3YH2 T 2·4 H2O, KY2 T(OH)3·5 H2O, K2Y2 T(OH)4·5 H2O isoliert. Die Präparate wurden mit Hilfe von Thermoanalyse, IR-Absorptionsspektren und Röntgenstreuung näher charakterisiert und ihre Löslichkeit in Wasser untersucht.
Some complexes of Yttrium with tartrates were isolated and the compounds characterised by thermogravimetric analysis, IR-spectroscopy and X-ray diffraction. Solubility in water was examined.
  相似文献   

12.
Crystal structure of 2,3-dihydroxypyridine (H2L) is determined. Mn(HL)Cl · H2O, Co(HL)Cl · 2H2O, Cu(HL)Cl, Ni(HL)OH · H2O, and Zn(HL)OH · H2O complexes are synthesized by reacting Mn(II), Co(II), Ni(II), Cu(II), and Zn(II) chlorides with H2L in ethanol solutions and identified. In these complexes, 2,3-dihydroxypyridine is involved in coordination as a monoanion. Spectral parameters of neutral and anionic forms of a ligand are determined and the acidity and complex formation constants are calculated. The compositions of complexes are established.  相似文献   

13.
Crystal Structures of Sr(OH)2 · H2O, Ba(OH)2 · H2O (o.-rh. and mon.), and Ba(OH)2 · 3 H2O The crystal structures of Ba(OH)2 · 3 H2O (Pnma, Z = 4), γ-Ba(OH)2 · H2O (P21/m, Z = 2) and the isotypic Sr(OH)2 · H2O and β-Ba(OH)2 · H2O (Pmc21, Z = 2) were determined using X-ray single crystal data. Ba(OH)2 · 3 H2O and Ba(OH)2 · H2O mon. crystallize in hitherto unknown structure types. The structure of Ba(OH)2 · H2O mon. is strongly related to that of rare earth hydroxides M(OH)3 with space group P63/m (super group of P21/m). The metal-oxygen distances are significantly shorter for OH? ions (mean Ba—O bond lengths of all hydroxides under investigation 278.1 pm) than for H2O molecules (289.9 pm). Corresponding to other hydrates of ionic hydroxides, the water molecules form strong hydrogen bonds to adjacent OH? ions whereas the hydroxide are not H-bonded.  相似文献   

14.
Interactions of dimethyl sulfoxide with carbon dioxide and water molecules which induce 18 significantly stable complexes are thoroughly investigated. An addition of CO2 or H2O molecules into the DMSO⋯1CO2 and DMSO⋯1H2O systems leads to an increase in the stability of the resulting complexes, in which it is larger for a H2O addition than a CO2. The overall stabilization energy of the DMSO⋯1,2CO2 is mainly contributed by the S=O⋯C Lewis acid–base interaction, whereas the O − H⋯O hydrogen bond plays a significant role in stabilizing complexes of DMSO⋯1,2H2O and DMSO⋯1CO2⋯1H2O. Remarkably, the complexes of DMSO⋯2H2O are found to be more stable than DMSO⋯1CO2⋯1H2O and DMSO⋯2CO2. The level of the cooperativity of multiple interactions in ternary complexes tends to decrease in going from DMSO⋯2H2O to DMSO⋯1CO2⋯1H2O and finally to DMSO⋯2CO2. It is generally found that the red shift of the O − H bond involved in an O − H⋯O hydrogen bond increases while the blue shift of a C − H bond in a C − H⋯O hydrogen bond decreases when a cooperative effect occurs in ternary complexes as compared to those of the corresponding binary complexes. © 2018 Wiley Periodicals, Inc.  相似文献   

15.
Five fluoride structures differing in the character of linkage of cationic polyhedra and in relative water contents are considered. The cationic sublattice is four-layered (ABAC) in K2Cu(ZrF6)2·6H2O, face-centered (F-type, ABC) in CuZrF6·4H2O primitive cubic (P-type) in Cu3(ZrF7)2·16H2O, hexagonal one-layered (AA) in Cu2ZrF8·12H2O and hexagonal two-layered (AB) in Zr2F8·6H2O. Some structures have considerably distorted symmetries of cationic sublattices compared to anhydrous structures because of the anisotropy of cation surroundings in the first, second, and third spheres, but al characteristic close-packed cation planes of the corresponding structural types are retained. In structures with cation-cation distances of ~3.5 Å, dimeric cations are arranged as single fragments for planes with dhkl>3.5 Å or as separate fragments for planes with smaller dhkl.  相似文献   

16.
The FOGO method is used to calculate absolute proton affinities of the molecules H2, HF, NH3, H2O, CH3OH, C2H5OH, H2O2, CH2O, CO, and CH2CO. Comparison with experimental values demonstrates that the geometrical and energetical data resulting from this type of ab initio calculation are of chemical accuracy. Predictive data for higher energy isomers, such as hydroxymethylene and ethynol are given as possible aid for the identification of these species.  相似文献   

17.
The TG, DTG, DTA, DSC, EC and dilatometric curves of SrCi2-6H2O are presented. Intermediate hydrates having the composition, SrCl2-2H2O and SrCl2-H2O, were detected while there was evidence for SrCl2-0.5H2O in the high pressure DSC curves. DSC curves for this compound were obtained at pressures from 0.2 to 50 atm.  相似文献   

18.
Hydrates of Barium Chloride. X-ray, Thermoanalytical, Raman, and I.R. Data In the system BaCl2? H2O the hydrates BaCl2 · 2 H2O, BaCl2 · 1 H2O, BaCl2 · 1/2 H2O, and BaCl2 · uH2O were obtained. X-ray powder data, i.r. and Raman spectra, as well as thermoanalytical measurements (TG, DTA) are reported. BaCl2 · 1 H2O and BaCl2 · 1/2 H2O, which are both isotype with the corresponding hydrates of SrCl2, were prepared by dehydration of BaCl2 · 2 H2O or by back hydration of anhydrous BaCl2 with the calculated amounts of water. BaCl2 · uH2O (u ≈? 1) is formed as the primary product by the reaction of anhydrous BaCl2 with water vapour at room temperature. Preparation methods of salt hydrates by controlled back hydration of the anhydrous salts are reported.  相似文献   

19.
Electrophilic trisubstituted ethylenes, phenoxy ring-substituted methyl 2-cyano-3-phenyl-2-propenoates, RPhCH=C(CN)CO2CH3, where R is 4-(4-BrC6H5O), 2-(4-ClC6H5O), 3-(4-ClC6H5O), 4-(3-ClC6H5O), 4-(4-ClC6H5O), 4-(4-FC6H5O), 2-(3-CH3OC6H5O), 2-(4-CH3OC6H5O), 3-(4-CH3OC6H5O), 4-(4-CH3OC6H5O), 3-(4-CH3C6H5O) were prepared and copolymerized with styrene. The monomers were synthesized by the piperidine catalyzed Knoevenagel condensation of phenoxy ring-substituted benzaldehydes and methyl cyanoacetate, and characterized by CHN analysis, IR, 1H and 13C-NMR. All the ethylenes were copolymerized with styrene (M1) in solution with radical initiation (ABCN) at 70°C. The compositions of the copolymers were calculated from nitrogen analysis and the structures were analyzed by IR, 1H and 13C-NMR. The order of relative reactivity (1/r1) for the monomers is 4-(4-CH3OC6H5O) (6.07) > 3-(4-ClC6H5O) (3.38) > 3-(4-CH3OC6H5O) (2.78) > 4-(3-ClC6H5O) (2.77) > 2-(4-ClC6H5O) (2.29) > 3-(4-CH3C6H5O) (1.98) > 4-(4-FC6H5O) (1.92) > 4-(4-ClC6H5O) (1.89) > 2-(3-CH3OC6H5O) (1.39) > 2-(4-CH3OC6H5O) (0.90) > 4-(4-BrC6H5O) (0.77). Relatively high Tg of the copolymers in comparison with that of polystyrene indicates a decrease in chain mobility of the copolymer due to the high dipolar character of the trisubstituted ethylene monomer unit. Decomposition of the copolymers in nitrogen occurred in two steps, first in the 200-500°C range with residue (2.5-8.0% wt), which then decomposed in the 500-800°C range.  相似文献   

20.
The synthesis, reduction, optical and e.p.r. spectral properties of a series of new binuclear copper(II) complexes, containing bridging moieties (OH, MeCO2 , NO2 , and N3 ), with new proline-based binuclear pentadentate Mannich base ligands is described. The ligands are: 2,6-bis[(prolin-1-yl)methyl]4-bromophenol [H3L1], 2,6-bis[(prolin-1-yl)methyl]4-t-butylphenol [H3L2] and 2,6-bis[(prolin-1-yl)methyl]4-methoxyphenol [H3L3]. The exogenous bridging complexes thus prepared were hydroxo: [Cu2L1(OH)(H2O)2] · H2O (1a), [Cu2L2(OH)(H2O)2] · H2O (1b), [Cu2L3(OH)(H2O)2] · H2O (1c), acetato [Cu2L1(OAc)] · H2O (2a), [Cu2L2(OAc)] · H2O (2b), [Cu2L3(OAc)] · H2O (2c), nitrito [Cu2L1(NO2)(H2O)2] · H2O (3a), [Cu2L2(NO2)(H2O)2] · H2O (3b), [Cu2L3(NO2)(H2O)2] · H2O (3c) and azido [Cu2L1(N3)(H2O)2] · H2O (4a), [Cu2L2(N3)(H2O)2] · H2O (4b) and [Cu2L3(N3)(H2O)2] · H2O (4c). The complexes were characterized by elemental analysis and by spectroscopy. They exhibit resolved copper hyperfine e.p.r. spectra at room temperature, indicating the presence of weak antiferromagnetic coupling between the copper atoms. The strength of the antiferromagnetic coupling lies in the order: NO2 N3 OH OAc. Cyclic voltammetry revealed the presence of two redox couples CuIICuII CuIICuI CuICuI. The conproportionality constant K con for the mixed valent CuIICuI species for all the complexes have been determined electrochemically.  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