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 34 毫秒
1.
The physical interactions of polymers with neighboring molecules are determined by only two kinds of interactions: London dispersion forces and Lewis acid–base interactions. These two kinds of attractive energies (together with certain steric restrictions) determine solubility, solvent retention, plasticizer action, wettability, adsorption, adhesion, reinforcement, crystallinity, and mechanical properties. The London dispersion force interaction energies of polymers have been quantified by the dispersion force contribution to cohesive energy density (δ2d) and the dispersion force contribution to surface energy (δd). The Lewis acid–base interactions, often referred to as “polar” interactions, can be best quantified by Drago's CA and EA constants for acid sites and CB and EB constants for basic sites. In this article infrared spectral shifts are featured as a method of determining enthalpies of acid–base interaction, and the C and E constants for polymers, plasticizers, and solvents. Examples are given where acid–base complexation of polymers with solvents dominate solubility and swelling phenomena. Enthalpies of acid–base complexation in polymer blends are determined from spectral shifts.  相似文献   

2.
The electronic spectra of carbazone in a number of organic solvents of different properties have been examined. It is deduced that in dilute solutions of basic solvents, carbazone (H2L) exists in acid—base equilibrium (H2L ⇋ HL + H+). The proportional concentration of the base form (HL) in a medium is largely dependent on the carbazone concentration, basicity of the organic solvent and the tendency of stabilization of the HL form through H-bond interaction with protic solvent molecules. The observed absorption bands belonging to absorption of both H2L and HL forms are assigned to a transition involving the whole solute associated with intramolecular CT transition. The longer wavelength band appearing in the spectra of the H2L form in the weak basic solvents (acetonitrile, acetone, dioxane and CCl4) is suggested to be due to absorption of the solvated polar complex resulting from interaction of solvent polarity with the polar solute carbonyl group.  相似文献   

3.
Adsorption of an anionic disazo dye, Food Black 2 on coating pigments from polar solvents was studied. Pigments (kaolin, precipitated CaCO3, amorphous SiO2, TiO2, Al2O3, and talc) were chosen to be representative of those used in paper for ink‐jet printing. The predominating driving force for the adsorption was electrostatic interactions for all the pigments except talc, for which hydrophobic interactions were dominant. Lewis acid‐base properties of the pigments were studied by measuring adsorption of Lewis acid‐probe molecules from both polar and nonpolar media. The pigments studied were found to be amphoteric, containing both acidic and basic surface groups. However, in the presence of polar solvents, the adsorption of acid‐base probe molecules was greatly reduced, which indicates that Lewis acid‐base interactions have only a small influence on the adsorption of azo dyes from polar media such as water.  相似文献   

4.
The rate constant for the Menschutkin reaction of 1,2‐dimethylimidazole with benzyl bromide to produce 3‐benzyl‐1,2‐dimethylimidazolium bromide was determined in a number of ionic liquids and molecular organic solvents. The rate constants in 12 ionic liquids are in the range of (1.0–3.2) × 10?3 L mol?1 s?1 and vary with the solvent anion in the order (CF3SO2)2 N? < PF6? < BF4?. Variations with the solvent cation (butylmethylimidazolium, octylmethylimidazolium, butyldimethylimidazolium, octyldimethylimidazolium, butylmethylpyrrolidinium, and hexyltributylammonium) are minimal. The rate constants in the ionic liquids are comparable to those in polar aprotic molecular solvents (acetonitrile, propylene carbonate) but much higher than those in weakly polar organic solvents and in alcohols. Correlation of the rate constants with the solvatochromic parameter E T(30) is reasonable within each group of similar solvents but very poor when all the solvents are correlated together. Better correlation is obtained for the organic solvents by using a combination of two parameters, π* (dipolarity/polarizibility) and α (hydrogen bond acidity), while additional parameters such as δ (cohesive energy density) do not provide any further improvement. © 2004 Wiley Periodicals, Inc. *
  • 1 This article is a US Government work and, as such, is in the public domain of the United States of America.
  • Int J Chem Kinet 36: 253–258, 2004  相似文献   

    5.
    The solvatochromic behavior of a penta‐tert‐butyl prydinium N‐phenolate betaine dye was studied using UV‐visible spectrophotometry in several binary mixture solvents. The solvent polarity parameter, ET (1) (kcal. mol?1) was calculated from the position of the longest‐wavelength intramolecular charge transfer absorption band of this penta‐tert‐butyl betaine dye. For binary solvent mixtures, all plots of ET (1) versus the mole fraction of a more polar component are nonlinear owing to preferential solvation of the probe by one component of the binary solvent mixture. In the computation of ET (1) it was assumed that the two solvents mixed interact to form a common structure with an ET (1) value not always intermediate between those of the two solvents mixed. The results obtained are explained by the strong synergism observed for some of the binary mixtures with strong hydrogen bond donors (HBD) solvents such as alcohols.  相似文献   

    6.
    Relative quantum yields and time constants for the fluorescence from pyrene (S1) stimulated by UV light have been measured for dispersions of the aromatic in several liquid solvents and aqueous surfactant micelles. Values of the relative radiative decay parameter, kF, were extracted and its medium dependence tabulated. This parameter was found to vary with medium polarity, being higher in more polar media. This effect, characterized for homogeneous liquid phases, was used to demonstrate that pyrene in surfactant micelles is strongly affected by the polar influences of water molecules which deeply penetrate the micelle in the region of the probe.  相似文献   

    7.
    A statistical treatment of data for the standard molar Gibbs free energies of transfer of monovalent ions from water to polar solvents has been made in terms of properties of the solvents and the ions. A common multiple regression equation with seven fitting constants, for almost 200 data points, has been found to describe the data in terms of four solvent properties: their electron donor and acceptor abilities, dielectric constant, and cohesive energy density, and three ionic properties: charge, size, and softness. For the ions Na+, K+, Rb+, Cs+, Tl+, (Ph)4As+, Cl?, Br? and N 3 ? the predictions of the equation are within acceptable error limits of the data, and encourage its application to solvents beyond the thirteen used for the data base. For other ions, e.g. H+, Ag+, and the larger anions, further interactions must be taken into account.  相似文献   

    8.
    A new heat-resistant polymer has been synthesized from N-(4-carboxy phenyl) trimellitimide and benzidine. The polymer was characterized by nitrogen analysis, IR spectroscopy, density, viscosity, and x-ray measurements. The polymer was found to be insoluble in all polar organic solvents and only soluble in formic acid and concentrated H2SO4. The solubility parameter value has been determined. Thermal studies show that the polymer is thermostable up to 300°C. The electrical properties have also been measured.  相似文献   

    9.
    Experimental results on various photophysical properties of coumarin‐30 (C30) dye, namely, Stokes' shift (Δv), fluorescence quantum yield (τf), fluorescence lifetime (τf), radiative rate constant (kf) and nonradiative rate constant (knr), as obtained using absorption and fluorescence measurements have been reported. Though in most of the solvents the properties of C30 show more or less linear correlation with the solvent polarity function, Δf= [(ε ‐ 1)/(2ε+ 1) ‐ (n2 ‐ 1)/ (2n2+ l)], they show unusual deviations in nonpolar solvents at one end and in high‐polarity protic solvents at the other end. From the solvent polarity and temperature effect on the photophysical properties of the dye, following inferences have been drawn: ( 1 ) in nonpolar solvents, the dye exists in a nonpolar structure, where its 7‐NEt2 substituent adopts a pyramidal configuration and the amino lone pair is out of resonance with the benzopyrone π cloud; ( 2 ) in medium to higher polarity solvents, the dye exists in a polar intra‐molecular charge transfer structure, where the 7‐NEt2 group and the 1,2‐benzopyrone moiety are in the same plane and the amino lone pair is in resonance with the benzopyrone π cloud; ( 3 ) in protic solvents, the dye‐solvent intermolecular hydrogen bonding influences the photophysical properties of the dye; and ( 4 ) in high‐polarity protic solvents, the excited C30 undergoes a new activation‐controlled nonradiative deexcitation process because of the involvement of a twisted intra‐molecular charge transfer (TICT) state. Contrary to most other TICT molecules, the activation barrier for this deexcitation process in C30 is observed to increase with solvent polarity. A rational for this unusual behavior has been given on the basis of the solvent polarity‐dependent stabilization and crossing of relevant electronic states and the relative propensity of interconversion among these states.  相似文献   

    10.
    Acid dissociation constants in methanol for eight substituted pyridine N-oxides having a wide range of acid-base properties, [quinoline N-oxide (bi-cyclic amine N-oxide) and pyridine (heterocyclic amine)] have been determined using the potentiometric titration method. A linear correlation between ourmethanol data and aqueous pK a values from the literature has been found. As in polar aprotic solvents cationic homoconjugation phenomenon has been found to be present for sufficiently basic N-oxides. The tendency of substituted pyridine N-oxides towards cationic homoconjugation in methanol is weaker than in polar aprotic solvents and increases with increasing basicity of N-oxides. It has also been found that, in contrast to polar aprotic solvents, the cationic homoconjugation phenomenon in methanol is much more pronounced for heterocyclic amines than their N-oxides.  相似文献   

    11.
    《Chemical physics letters》1999,291(5-6):363-370
    Time-resolved emission is used to explore the intramolecular excited-state electron transfer in a donor-substituted acridinium dye in a variety of solvents at room temperature. Emission decays are predominantly bi-exponential, with the faster of the two time constants being a good measure of the charge transfer time. In 14 high-polarity solvents this time is correlated to the solvation times previously measured with the solvation probe Coumarin 153 via an apparent power law, τLE→CT≈8.4〈τsolv0.65. In solvents less polar than acetone the times observed deviate widely from this relationship, probably due to the effects of ion pairing.  相似文献   

    12.
    Growing interest is being shown in the role of the solvent (particularly polar solvents) in chemical reactions[1–4]. The most remarkable polar aprotic solvent appears to us to be the hexamethylated triamide of orthophosphoric acid, OP(N(CH3)2)3, which is known as hexamethylphosphoramide or tris(dimethylamino)phosphine oxide. Hexamethylphosphoramide[5–7] has recently been studied with respect to its solvent properties for gases[5], for many organic and inorganic salts[7], and for polymers[5], as well as its use as a polymerization co-catalyst. However, these aspects are not considered in the present paper, which deals with the physical and chemical properties of hexamethylphosphoramide, and in particular with its use as a reaction medium. Our own work on hexamethylphosphoramide began in 1961, and until then very little work had been done on this compound. Even in a paper published by Parker[1b] in 1965, very little reference is made to hexamethylphosphoramide.  相似文献   

    13.
    The complexation of electron donor–acceptor complexes of 8-hydroxyquinoline (8HQ) and metadinitrobenzene (MNB) have been studied spectrophotometrically and thermodynamically in different polar solvent at room temperature. A new absorption band due to charge transfer (CT) transition is observed in the visible region. A new theoretical model has been developed which take into account the interaction between electronic subsystem of 8HQ and MNB. The results indicate the extent of charge transfer complexes (CTCs) formation to be more in less polar solvents. Stoichiometry of the complex was found to be 1:1 by straight line method and 1H NMR between donor and acceptor at the maximum absorption bands. Ionization potential (ID) and resonance energy (RN) were determined from the CT transition energy in different solvents. The formation constants of the complexes were determined in different polar solvents from which ΔG° formation of the complexes was estimated and also extinction coefficient of the charge transfer complex (CTC) was calculated. Oscillator strength, transition dipole strengths and maximum wavelength of the CTC (λCT) in various solvents and IR spectra of the CTC have also been discussed. It has been observed that all parameters described above changed with change in polarity and concentration of donor.  相似文献   

    14.
    By using the potentiometric method, acidity constants have been determined in systems of tri- and tetra-substituted pyridine N-oxides. The potentiometric measurements in systems of four 4-chloropyridine N-oxide derivatives containing the chlorine atom at position 4 to the NO2 group and four bromine counterparts were carried out in polar non-aqueous solvents, viz. amphiprotic methanol (MeOH) and aprotic protophilic dimethyl sulfoxide (DMSO). It was found that in all the systems studied the pKa values were readily determinable (as indicated by small standard deviations) in MeOH, whereas in DMSO large standard deviations were obtained making the pKa values either hardly determinable or indeterminable from potentiometric measurements. Furthermore, it was demonstrated that the acidity constants of protonated N-oxides studied in MeOH changed according to the sequence of their acidity constants in water. It was also found that in the polar solvents studied, i.e. in the amphiprotic methanol and the highly basic aprotic dimethyl sulfoxide, the cationic homo-conjugation equlibrium constants could not be determined using potentiometric method. Also, by using ab initio methods at the RHF and MP2 levels and the PCM model, utilizing the Gaussian 6-31++G∗∗ basis set, energies and Gibbs free energies of the protonation reactions of the N-oxides have been determined. The energy parameters have been compared with acidity constants of the protonated N-oxides determined by potentiometric titration in methanol to establish a correlation between these approaches.  相似文献   

    15.
    The solubility of CO2 in polar solvents is poorly predicted by all estimation methods that use only properties of the pure components. It is thought that this is because CO2 molecules, although they do not have a permanent dipole moment, behave like electrical multipoles as a consequence of their strongly polar bonds. An equation, proposed in a previous paper, for estimating the activity coefficient of a nonpolar gas dissolved in a polar liquid is modified by adding a term containing the quadrupole moment of the gas molecule. Errors in estimating gas solubilities on that basis are less than 20%, except for the solvents in which specific interactions (e.g., acid–base interactions) are present.  相似文献   

    16.
    Carbon dots (CDs) have attracted increasing attention in applications such as bio‐imaging, sensors, catalysis, and drug delivery. However, unlike metallic and semiconductor nanoparticles, the transfer of CDs between polar and non‐polar phases is little understood. A class of amine‐terminated CDs is developed and their phase transfer behavior has been investigated. It is found that these CDs can reversibly transfer between aqueous and organic solvents by alternatively bubbling and removing CO2 at atmospheric pressure. The mechanism of such CO2‐switched phase transfer involves reversible acid–base reaction of amine‐terminated CDs with CO2 and the reversible formation of hydrophilic ammonium salts. By using the CDs as catalysts, the phase transfer is applied in the Knoevenagel reaction for efficient homogeneous reaction, heterogeneous separation, and recycling of the catalysts.  相似文献   

    17.
    Abstract

    A new redox system, dioxane-ascorbic acid, has been investigated for the homopolymerization of vinyl monomers. Detailed kinetic studies on the aqueous polymerization of acrylamide by this initiating system have been done iodometrically at 35 ± 0.2°C. The effect of various additives, such as organic solvents, inorganic salts, surfactants, etc., on the rate of polymerization has been studied. The retardation constants for organic solvents have been evaluated by the “intercept method.” The overall energy of activation has been found to be 8.75 kcal/deg/mol, within the temperature range 25–45°C. A suitable mechanism has been suggested. The following rate expression: Rp α [acrylamide]1.0 [dioxane]1.0 [ascorbic acid]0, has been observed.  相似文献   

    18.
    Coordination equilibria in the Co(II)–Ni(II)–2-aminopropanoic acid (HAla)–EDTA system have been studied spectrophotometrically at different molar ratios of the reagents in a wide pH range. It has been found that, when metal ions are in excess with respect to EDTA at pH 5–9, polyheteronuclear complexonates [(CoAla)Edta(NiAla)]2–, [(CoAla2)Edta(NiAla2)]4–, [(NiAla2)Edta(CoAla2)2]4–, [(CoAla2)Edta(NiAla2)2]4–, and [(NiAla2)2Edta(CoAla2)2]4– form in a solution. The equilibrium constants of formation of these complexes and their overall stability constants have been calculated. Possible structures of the polynuclear complexonates are discussed.  相似文献   

    19.
    The effect of the macrocyclic host, cucurbit[7]uril (CB7), on the photophysical properties of the 2‐(2′‐hydroxyphenyl)benzimidazole (HPBI) dye have been investigated in aqueous solution by using ground‐state absorption and steady‐state and time‐resolved fluorescence measurements. All three prototropic forms of the dye (cationic, neutral, and anionic) form inclusion complexes with CB7, with the largest binding constant found for the cationic form (K≈2.4×106 M ?1). At pH≈4, the appearance of a blue emission band upon excitation of the HPBI cation in the presence of CB7 indicates that encapsulation into the CB7 cavity retards the deprotonation process of the excited cation, and hence reduces its subsequent conversion to the keto form. Excitation of the neutral form (pH≈8.5), however, leads to an increase in the keto form fluorescence, indicating an enhanced excited‐state intramolecular proton‐transfer process for the encapsulated dye. In both the ground and excited states, the two pKa values of the HPBI dye show upward shifts in the presence of CB7. The prototropic equilibrium of the CB7‐complexed dye is represented by a six‐state model, and the pH‐dependent changes in the binding constants have been analyzed accordingly. It has been observed that the calculated pKa values using this six‐state model match well with the values obtained experimentally. The changes in the pKa values in the presence of CB7 have been corroborated with the modulation of the proton‐transfer process of the dye within the host cavity.  相似文献   

    20.
    Dissociation constants (KD) and solvatochromic shifts of the Soret bands of the dicationic form of hematoporphyrin dichloride were measured against the solvatochromic parameters of Taft, Dimroth, Gutmann and Swan. Taft's parameter representing solvent capacity of proton acceptance from the solutes was accepted as the best cybotactic parameter. Soret bands of the spectrum undergo blue shifts upon changing nonpolar to polar solvents. The difference in hydrogen bond energy calculated between the pyrrolic proton and Cl in hexane and isopropanol is 2.8 kcal/mol.  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