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.
In the synthesis of the disordered lyotropic liquid crystalline L3 sponge phase prepared with the cosurfactants cetylpyridinium chloride and hexanol, aqueous NaCl solution is used as the solvent. When this sponge phase is used as the template for L3 silica-phase processing, we replace NaCl with HCl to facilitate the acid catalysis of tetramethoxysilane in forming a templated silica gel, assuming that changing the solvent from NaCl(aq) to HCl(aq) of equivalent ionic strength does not affect the stability range of the L3 phase. In this work, we confirm that changing the pH of the solvent from neutral to acidic (with HCl) has negligible effect on the L3 phase region. Equivalent ionic strength is provided by either NaCl(aq) or HCl(aq) solvent; therefore, a similar phase behavior is observed regardless of which aqueous solvent is used.  相似文献   

2.
根据体积排除色谱(SEC)研究高分子溶质优先溶剂化的基本原理,证明在SEC色谱图中被束缚溶剂产生的面积与自由溶剂产生的面积大小相等方向相反,在研究优先溶剂化时从高分子峰入手和从自由溶剂峰入手在理论上具有等价性.分析表明,通过对溶剂化高分子峰的研究,还可以得到另一个重要物理参数——恒化学位时高分子溶液的折光指数增量.  相似文献   

3.
《Polyhedron》1987,6(6):1337-1342
The composition of the solvent cage of chloropentaamminechromium(III) ion was determined in water—dimethyl sulfoxide media using proton NMR line-broadening methods and the approach of Covington and coworkers. The number of solvent molecules in the solvent cage was found to be 10. The stepwise formation constants for the substitution of 10 water molecules by 10 dimethylsulfoxide molecules in this solvent cage were calculated. After the first such substitution each successive substitution becomes 1027 J mol−1 more difficult, exclusive of statistical factors, than the preceding substitution. The solvent cage composition was assumed to apply to the chloropentaamminecobalt(III) ion. Mercury(II)-assisted removal of chloride ion from the latter complex gave [Co(NH3)5{OSMe2}]3+/[Co(NH3)5(H2O)]3+ product ratios which did not correlate with either the solvent cage composition or the activity ratio of the two solvent components in the bulk phase of the solvent.  相似文献   

4.
With the aim of improving the field-effect mobilities in poly(3-hexylthiophene) (P3HT) thin film transistors, we controlled the nanostructures of P3HT thin film by changing the solvent vapor pressure in a spin-coating chamber during solidification. The transistors with P3HT thin films spin-coated under a high solvent vapor pressure (56.5 KPa), showing the one-dimensional nanowire morphologies, resulted in the relatively high field-effect mobilities (0.02 cm2/(V.s)) that are typically more than 1 order of magnitude higher than those prepared under ambient conditions, showing the featureless morphologies. This can be attributed to the higher solvent vapor pressure during film formation, providing the solvent is allowed to evaporate slowly and the degree of ordering within the P3HT crystalline domains is dramatically improved.  相似文献   

5.
The conformation of a polymer chain in solution is intrinsically coupled to the thermodynamic and structural properties of the solvent. Here we study such solvent effects in a system consisting of a flexible interaction-site n-mer chain immersed in a monomeric solvent. Chain conformation is described with a set of intramolecular site-site probability functions. We derive an exact density expansion for these intramolecular probability functions and give a diagrammatic representation of the terms contributing at each order of the expansion. The expansion is tested for a short hard-sphere chain (n=3 or 4) with site diameter sigma in a hard-sphere solvent with solvent diameter D. In comparison with Monte Carlo simulation results for 0.2< or =D/sigma< or =100, the expansion (taken to second order) is found to be quantitatively accurate for low to moderate solvent volume fractions for all size ratios. Average chain dimensions are predicted accurately up to liquidlike solvent densities. The hard-sphere chains are compressed with both increasing solvent density and decreasing solvent size. For small solvent (D相似文献   

6.
The infrared spectroscopy studies of the C3 and C20 carbonyl stretching vibrations (upsilon(C=O)) of progesterone in CHCl3/cyclo-C6H12 binary solvent systems were undertaken to investigate the solute-solvent interactions. With the mole fraction of CHC13 in the binary solvent mixtures increase, three types of C3 and C20 carbonyl stretching vibration band of progesterone are observed, respectively. The assignments of upsilon(C=O) of progesterone are discussed in detail. In the CHCl3-rich binary solvent systems or pure CHCl3 solvent, two kinds of solute-solvent hydrogen bonding interactions coexist for C20 C=O. Comparisons are drawn for the solvent sensitivities of upsilon(C=O) for acetophenone and 5alpha-androstan-3,17-dione, respectively.  相似文献   

7.
以混合溶剂作淋洗剂的体积排除色谱(SEC)中高分子样品一般会出现2个峰, 分别是溶剂化高分子峰和自由溶剂峰. 理论分析结果表明, 溶剂化高分子峰面积( A3eff )是“裸高分子”峰面积(A3)和被束缚溶剂峰面积(A1*)的加和, 被束缚溶剂峰面积(A1*)和自由溶剂峰面积(A2*)大小相等符号相反. 以聚苯乙烯(3)-氯仿(1)-甲醇(2)体系为研究对象, 分别对A3eff, A3及A2*进行了实验测定, 证实了理论推断的正确性, 表明由高分子峰和由溶剂峰来求算优先吸附系数是等价的.  相似文献   

8.
Summary Demixing effects in thin-layer chromatography have been investigated with NH2-modified silica gel precoated plates and for frequently used hydroorganic binary solvents of various compositions with salt addition.The position of the solvent demixing front depends on: (1) solvent composition, (2) nature of the organic modifier and (3) salt concentration. Whatever the organic modifier, for a given water percentage in the developing solvent, solvent demixing is found to occur at the same concentration of NaCl.  相似文献   

9.
[reaction: see text] The reaction of CH(4) with CO(2) has been performed in anhydrous acids using VO(acac)(2) and K(2)S(2)O(8) as promoters. NMR analysis establishes that the primary product is a mixed anhydride of acetic acid and the acid solvent. In sulfuric acid, the overall reaction is CH(4) + CO(2) + SO(3) --> CH(3)C(O)-O-SO(3)H. Hydrolysis of the mixed anhydride produces acetic acid and the solvent acid. When trifluoroacetic acid is the solvent, acetic acid is primarily formed via the reaction CH(4) + CF(3)COOH --> CH(3)COOH + CHF(3).  相似文献   

10.
韩艳春 《高分子科学》2013,31(7):1029-1037
The surface composition of poly(3-hexylthiophene-2,5-diyl) and fullerene derivative [6,6]-phenyl-C61-butyric acid methyl ester (P3HT/PCBM) blend films could be changed by controlling the film formation process via using mixed solvents with different evaporation rates. The second solvent, with a higher boiling point than that of the first solvent and much better solubility for PCBM than P3HT, is chosen to mix with the first solvent with a lower boiling point and good solubility for both PCBM and P3HT. The slow evaporation rate of the second solvent provides enough time for PCBM to diffuse upwards during the solvent evaporation. Thus, the weight ratio of PCBM and P3HT (m PCBM/m P3HT) at surface of the blend films was varied from ca. 0.1 to ca. 0.72, i.e., it increases about seven times by changing from single solvent to mixed solvents. Meanwhile, the mixed solvents were in favor to form P3HT naonofiber network and enhance phase separation of P3HT/PCBM blend films. As a result, the power conversion efficiency of the device from mixed solvents with slow evaporation process was about 1.5 times of the one from single solvents.  相似文献   

11.
We study via lattice Monte Carlo simulation and Flory theory the properties of g=1-6 dendrimers in variable solvent quality. For all the generations studied, we find that the radius of gyration R(g) collapses significantly (factor of 2) going from athermal to extreme poor solvent conditions, indicating that varying solvent quality is an effective means of controlling dendrimer size. We also find that in athermal, theta, and extreme poor solvent conditions, the radius of gyration of dendrimers scales with the total number of monomers roughly as R(g) approximately N(1/3). However, a more careful analysis shows that in athermal and theta solvents, there is, in fact, a small but systematic deviation of R(g) from R(g) approximately N(1/3) scaling and the simulation data is described better by the Flory theory prediction of R(g) approximately N(1/5)[(g+1)m](2/5) in athermal solvents and R(g) approximately N(1/4)[(g+1)m](1/4) in theta solvents. We also find for our simulation data that stronger deviations from constant density scaling are possible, with scaling behavior as shallow as R(g) approximately N(0.26) possible for solvent conditions in between theta and the completely collapsed state. It is evident therefore that dendrimers do not obey (or even approximately obey) R(g) approximately N(1/3) scaling under all solvent conditions. Under all solvent conditions, we find that the intramolecular density is dense corelike (i.e., the density maximum is in the interior of the dendrimer) and terminal groups are delocalized throughout the dendrimer.  相似文献   

12.
The behavior of protonated binary solvents injected into deuterated binary mobile phases in capillary LC is studied with NMR. Specifically, the solvent elution is followed on-flow with a capillary LC coupled to a 900 nL volume microcoil NMR probe. A range of identical composition 5% protonated (and 95% deuterated) solvents is injected into composition-matched deuterated mobile phases of CD(3)CN/D(2)O and CD(3)OD/D(2)O. The protonated components separate for all solvent combinations except at 80% CD(3)CN/20% D(2)O and similar to 72% CD(3)OD/28% D(2)O where only a single retention time is observed. The more hydrophilic protonated component, HOD, elutes first with higher percentages of hydrophilic solvent, D(2)O, in the mobile phase whereas retention is reversed with the higher percentage of the more hydrophobic solvent (CD(3)CN and CD(3)OD) in the mobile phase. The hydrophilic/hydrophobic nature of the chromatographic system as a function of mobile phase composition is characterized by following the retention times of protonated solvents.  相似文献   

13.
用毛细管气相色谱、色-质谱和旋光色散及圆二色性谱仪等方法对乙酸乙烯酯在非极性和极性溶剂中的不对称氢甲酰化反应产物进行分离和鉴定。实验结果表明,在非极性溶剂中反应的收率、选择性、光学收率e.e.值(enantiomericexcess)均比极性溶剂中的结果为好。由此探讨了不对称氢甲酰化反应中溶剂的影响。  相似文献   

14.
Raman spectra of water+N,N-dimethylformamide (DMF) mixtures and their solutions with NaNCS, KNCS and NH(4)NCS were obtained. The bands of nu(CO) stretching, delta(OCN) bending, r(CH(3)) rocking and nu(N-CH)(3)) stretching of the DMF molecule with and without salts were studied. The dependence of the vibration frequencies and Raman intensities of the bands on the composition of the mixed solvent was discussed. The change of the band frequencies as a result of the presence of the salts and the solvation of the cations by the solvent molecules was examined. The stronger cation solvation by the aprotic solvent molecules instead of the water molecules in DMF concentrated solutions was discussed. The nu(CN) and nu(CS) vibrations of the SCN(-) ions were observed as a function of the cation present and the solvent composition. The presence of the SCN(-) ions as "free", contact ion pairs, or solvent separated pairs, was discussed.  相似文献   

15.
The gelation behaviours of low molecular weight gelators 1,3:2,5:4,6-tris(3,4-dichlorobenzylidene)-D-mannitol(G1)and 2,4-(3,4-dichlorobenzylidene)-N-(3-aminopropyl)-D-gluconamide(G2)in 34 solvents have been studied.We found that sample dissolved at low concentrations may become a gel or precipitate at higher concentrations.The Hansen solubility parameters(HSPs)and a Teas plot were employed to correlate the gelation behaviours with solvent properties,but with no success if the concentration of the tests was not maintained constant.Instead,on the basis of the gelation results obtained for the G1 and G2 in single solvents,we studied the gelation behaviours of G1 and G2 in 23 solvent mixtures and found that the tendency of a gelator to form a gel in mixed solvents is strongly correlated with its gelation behaviours in good solvents.If the gelation occurs in a good solvent at higher concentrations,it will take place as well in a mixed solvent(the good solvent plus a poor solvent)at a certain volume ratio.In contrast,if the gelator forms a precipitate in a good solvent at higher concentrations,no gelation is to be observed in the mixed solvents.A gelation rule for mixed solvents is thus proposed,which may facilitate decision making with regard to solvent selection for gel formation in the solvent mixtures in practical applications.  相似文献   

16.
The electrogenerated chemiluminescence (ECL) of 9,10-diphenylanthracene (DPA), rubrene, and anthracene has been studied in fluorinated aromatic solvents. Mixed annihilation ECL between aromatic luminophores and quinones was observed in solvent systems containing acetonitrile and either benzene, benzotrifluoride, 3-fluorobenzotrifluoride, or 1,3-bis(trifluoromethyl)benzene. Increases in ECL efficiency (phi ecl, photons generated per redox event) correlated with decreasing solvent polarity when 1,4-benzoquinone was used as a nonemitting ECL partner. However, opposite results were observed using 1,4-naphthaoquinone (NQ) as a nonemitting partner. phi ecl also correlated with radical anion stability of NQ in these solvent systems, as indicated by reverse/forward current ratios ( I r/ I f), suggesting noncovalent interactions between the solvent and the nonemitting ECL partner. Specifically, the reaction of an aromatic luminophore with 1,4-naphthoquinone in acetonitrile/benzotrifluoride showed a 1.03-1.63-fold increases in ECL efficiency over that of acetonitrile/benzene. Slight blue shifts ( approximately 3 nm) in photoluminescence and ECL emissions were seen as solvent polarity increased. Reaction enthalpies of each system were estimated using half-wave potentials of oxidation and reduction and were found to correlate well with emission energy.  相似文献   

17.
The fluorescence quenching of 4-aminodiphenyl (4ADP) with chloromethanes (CH2Cl2, CHCl3 and CCl4) have been studied in solvents of different polarity and viscosity. The quenching rate constants (kq) have been determined in all solvents. For CCl4 and CHCl3 quenching, the kq depends on solvent viscosity whereas for CH2Cl2, the kq values show a mixed trend with no clear-cut variation with either solvent polarity or solvent viscosity. Quenching mechanism involving an intermediate donor-acceptor complex formation is proposed for CH2Cl2 quenching. A positive deviation was observed in the Stern-Volmer (SV) plot for CCl4 quenching in hexane. The static-dynamic model could explain this.  相似文献   

18.
A novel corrole-type macrocycle, oxocorrologen (2), substituted with hemiquinone groups, has been synthesized. It was found to undergo multiple tautomerism of its exchangeable protons between electronegative atom sites at the macrocyclic core (nitrogen atoms) and periphery (phenol oxygen atoms). Alkylation at one macrocyclic nitrogen atom with a 4-nitrobenzyl group gave 3, which can exist in only two tautomeric forms depending on the solvent. Tautomerism has been studied by means of (1)H NMR spectroscopy in a variety of solvents and solvent mixtures. Tautomer structure assignments have been supported by DFT calculations of the relative energies of the tautomers. X-ray crystallography of the N-nitrobenzyl derivative has revealed that intramolecular hydrogen bonding may be responsible for stabilizing the observed tautomers. The solvent dependence of the tautomerism of 2 and 3 confers solvatochromism. Electrochemical measurements on 2 and 3 in their respective quinone forms have revealed irreversible processes, but indicate that they are both electron-deficient with a small HOMO-LUMO gap and first reduction potentials close to those of fullerene electron acceptors.  相似文献   

19.
The standard for chemical shift is dilute tetramethylsilane (TMS) in CDCl3, but many measurements are made relative to TMS in other solvents, the proton resonance of the solvent peak or relative to the lock frequency. Here, the chemical shifts of TMS and the proton and deuterium chemical shifts of the solvent signals of several solvents are measured over a wide temperature range. This allows for the use of TMS or the solvent and lock signal as a secondary reference for other NMR signals, as compared with dilute TMS in CDCl3 at a chosen temperature; 25 degrees C is chosen here. An accuracy of 0.02 ppm is achievable for dilute solutions, provided that the interaction with the solvent is not very strong. The proton chemical shift of residual water is also reported where appropriate.  相似文献   

20.
The effects of the use of three generalized Born (GB) implicit solvent models on the thermodynamics of a simple polyalanine peptide are studied via comparing several hundred nanoseconds of well-converged replica exchange molecular dynamics (REMD) simulations using explicit TIP3P solvent to REMD simulations with the GB solvent models. It is found that when compared to REMD simulations using TIP3P the GB REMD simulations contain significant differences in secondary structure populations, most notably an overabundance of alpha-helical secondary structure. This discrepancy is explored via comparison of the differences in the electrostatic component of the free energy of solvation (DeltaDeltaG(pol)) between TIP3P (via thermodynamic Integration calculations), the GB models, and an implicit solvent model based on the Poisson equation (PE). The electrostatic components of the solvation free energies are calculated using each solvent model for four representative conformations of Ala10. Since the PE model is found to have the best performance with respect to reproducing TIP3P DeltaDeltaG(pol) values, effective Born radii from the GB models are compared to effective Born radii calculated with PE (so-called perfect radii), and significant and numerous deviations in GB radii from perfect radii are found in all GB models. The effect of these deviations on the solvation free energy is discussed, and it is shown that even when perfect radii are used the agreement of GB with TIP3P DeltaDeltaG(pol) values does not improve. This suggests a limit to the optimization of the effective Born radius calculation and that future efforts to improve the accuracy of GB models must extend beyond such optimizations.  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