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 813 毫秒
1.
The standard molar energies of combustion, at T = 298.15 K, of crystalline 1,4-benzodioxan-2-carboxylic acid and 1,4-benzodioxan-2-hydroxymethyl were measured by static bomb calorimetry in an oxygen atmosphere. The standard molar enthalpies of sublimation, at T = 298.15 K, were obtained by Calvet microcalorimetry. These values were used to derive the standard molar enthalpies of formation of the compounds in the gas phase at T = 298.15 K: 1,4-benzodioxan-2-carboxylic acid ?(547.7 ± 3.0) kJ · mol?1 and 1,4-benzodioxan-2-hydroxymethyl ?(374.2 ± 2.3) kJ · mol?1.In addition, density functional theory calculations using the B3LYP hybrid exchange–correlation energy functional with extended basis sets, 6-311G7 and cc-pVTZ, have been performed for the compounds studied. We have also tested two more accurate computational procedures involving multiple levels of electron structure theory in order to get reliable estimates of the thermochemical parameters of the compounds studied. The agreement between experiment and theory gives confidence to estimate the enthalpies of formation of other 2-R derivatives of 1,4-benzodioxan (R = –CH2COOH, –OH, –COCH3, –CHO, –CH3, –CN, and –NO2).  相似文献   

2.
A new bis-quinolylimine ligand containing an azadiene moiety, 1,4-bis(2-quinolyl)-2,3-diaza-1,3-buthadiene (1), was synthesized by one-step facile condensation. This simple ligand, when dissolved in acetonitrile, shows a Cu2+-selective fluorescence enhancement. Coordination of 1 with Cu2+ produces two kinds of complexes with 1:1 and 1:2 stoichiometries. The 1:2 complex shows a strong fluorescence (ΦF = 0.37), while the 1:1 complex does not (ΦF < 0.01). Ab initio molecular orbital calculation reveals that the 1:1 complex has a distorted structure, while the 1:2 complex has a planar structure. The planar configuration of the 1:2 complex, therefore, allows an extended π-conjugation over the entire molecule and, hence, results in fluorescence enhancement.  相似文献   

3.
《Polyhedron》2007,26(9-11):2247-2251
The reaction of Vo(CO)6 and representative quinones, A (A = benzoquinone, chloranil, 2,3-dicyano-1,4-naphthoquinone, and dihydroxy-1,4-benzoquinone), form materials of V(A)2 · zCH2Cl2 (z < 0.1) composition, which exhibits antiferromagnetic coupling and do not magnetically order above 5 K.  相似文献   

4.
The interaction between imidacloprid (IMI) and human serum albumin (HSA) was investigated using fluorescence and UV/vis absorption spectroscopy. The experimental results showed that the fluorescence quenching of HSA by IMI was a result of the formation of IMI–HSA complex; static quenching was confirmed to result in the fluorescence quenching. The apparent binding constant KA between IMI and HSA at three differences were obtained to be 1.51 × 104, 1.58 × 104, and 2.19 × 104 L mol?1, respectively. The thermodynamic parameters, Δ and Δ were estimated to be 28.44 kJ mol?1, 174.76 J mol?1 K?1 according to the van’t Hoff equation. Hydrophobic interactions played a major role in stabilizing the complex. The distance r between donor (HSA) and acceptor (IMI) was obtained according to fluorescence resonance energy transfer. The effect of IMI on the conformation of HSA was analyzed using synchronous fluorescence spectroscopy CD and three-dimensional fluorescence spectra, the environment around Trp and Tyr residues were altered.  相似文献   

5.
The thermodynamic parameters, ΔBG, ΔBH, ΔBS, and ΔBCp, of the drugs flurbiprofen (FLP), nabumetone (NAB), and naproxen (NPX) binding to β-cyclodextrin (βCD) and to γ-cyclodextrin (γCD) in 0.10 M sodium phosphate buffer were determined from isothermal titration calorimetry (ITC) measurements over the temperature range from 293.15 K to 313.15 K. The heat capacity changes for the binding reactions ranged from −(362 ± 48) J · mol−1 · K−1 for FLP and −(238 ± 90) J · mol−1 · K−1 for NAB binding in the βCD cavity to 0 for FLP and −(25.1 ± 9.2) J · mol−1 · K−1 for NPX binding in the larger γCD cavity, implying that the structure of water is reorganized in the βCD binding reactions but not reorganized in the γCD binding reactions. Comparison of the fluorescence enhancements of FLP and NAB upon transferring from the aqueous buffer to isopropanol with the maximum fluorescence enhancements observed for their βCD binding reactions indicated that some localized water was retained in the FLP–βCD complex and almost none in the NAB–βCD complex. No fluorescence change occurs with drug binding in the larger γCD cavity, indicating the retention of the bulk water environment in the drug–γCD complex. Since the specific drug binding interactions are essentially the same for βCD and γCD, these differences in the retention of bulk water may account for the enthalpically driven nature of the βCD binding reactions and the entropically driven nature of the γCD binding reactions.  相似文献   

6.
The enthalpies of solution of 1,4-dioxane in {(1  x)F + xH2O}, {(1  x)NMF + xH2O}, and {(1  x)DMF + xH2O} have been measured within the whole mole fraction range at T = 298.15 K. Based on the obtained data, the effect of substituting methyl groups at the nitrogen atom in formamide on the preferential solvation of 1,4-dioxane has been analyzed. A simple model has been proposed to describe the influence of structural and energetic properties of the mixed solvent on the energetic effect of hydrophobic hydration and preferential solvation of 1,4-dioxane by the components of the examined mixture.  相似文献   

7.
8.
Benzhydryl protection by diphenyldiazomethane of an alcohol in enantiomeric base-sensitive ribonolactones allows short efficient syntheses of 1,4-dideoxy-1,4-imino-d-lyxitol (DIL) and of 1,4-dideoxy-1,4-imino-l-lyxitol (LIL). DIL showed potent [Ki = 0.13 μM]—and LIL showed weak [Ki = 113 μM]—competitive inhibition of α-d-galactosidase. Both enantiomers N-benzyl-DIL [Ki = 64 μM] and N-benzyl-LIL [Ki = 13 μM] were moderate competitive inhibitors of naringinase, an α-l-rhamnosidase.  相似文献   

9.
《Chemical physics letters》1999,291(1-2):75-81
The fluorescence spectrum of all-trans-β-carotene was recorded at 170 K. The 1Bu+  1Ag fluorescence exhibited clear vibrational structures constituting a mirror image with those of the 1Bu+  1Ag absorption, and the deconvolution of the entire spectrum identified the 2Ag(0)  1Ag(0) transition at 14 500 cm−1. The displacements of the 1Bu+ and 2Ag potential minima along ν1 and ν2 (the CC stretching and C–C stretching normal coordinates, respectively) were determined to be 1.2 and 0.9, and 1.6 and 1.5, respectively. Thus, much larger potential displacements in the 2Ag state than in the 1Bu+ state have been shown.  相似文献   

10.
We determined apparent molar volumes V? at 278.15 ? (T/K) ? 368.15 and apparent molar heat capacities Cp,? at 278.15 ? (T/K) ? 393.15 at p = 0.35 MPa for aqueous solutions of tetrahydrofuran at m from (0.016 to 2.5) mol · kg?1, dimethyl sulfoxide at m from (0.02 to 3.0) mol · kg?1, 1,4-dioxane at m from (0.015 to 2.0) mol · kg?1, and 1,2-dimethoxyethane at m from (0.01 to 2.0) mol · kg?1. Values of V? were determined from densities measured with a vibrating-tube densimeter, and values of Cp,? were determined with a twin fixed-cell, differential, temperature-scanning calorimeter. Empirical functions of m and T for each compound were fitted to our V? and Cp,? results.  相似文献   

11.
Electrochemical and spectroscopic properties of Tb(III) in molten LiCl–KCl eutectic at high temperature were investigated by cyclic voltammetry and time-resolved laser-induced fluorescence spectroscopy (TRLFS). The diffusion coefficient of Tb(III) and the formal standard potential of Tb(III)/Tb0 were determined to be 2.06 ± 0.4 × 10? 5 cm2 s? 1 and ? 2.83 ± 0.03 V vs. Cl2/Cl? at 887 K, respectively. Additionally, visible fluorescence of Tb(III) due to the electronic transitions from 5D3 and 5D4 to 7FJ was observed and measured by TRLFS for the first time. These results provide the first fluorescence spectroscopic evidence for a direct in situ quantification of Tb(III) in the high temperature molten salt system.  相似文献   

12.
Excess molar enthalpies HmEand excess molar volumesVmE of (1,3-dimethyl-2-imidazolidinone  +  benzene, or methylbenzene, or 1,2-dimethylbenzene, or 1,3-dimethylbenzene, or 1,4-dimethylbenzene, or 1,3,5-trimethylbenzene, or ethylbenzene) over the whole range of compositions have been measured at T =  298.15 K. The excess molar enthalpy values were positive for five of the seven systems studied and the excess molar volume values were negative for six of the seven systems studied. The excess enthalpy ranged from a maximum of 435 J · mol  1for (1,3-dimethyl-2-imidazoline  +  1,3,5-trimethylbenzene) to a minimum of   308 J · mol  1for (1,3-dimethyl-2-imidazoline  +  benzene). The excess molar volume values ranged from a maximum of 0.95cm3mol  1 for (1,3-dimethyl-2-imidazoline  +  ethylbenzene) and a minimum of   1.41 cm3mol  1for (1,3-dimethyl-2-imidazoline  +  methylbenzene). The Redlich–Kister polynomial was used to correlate both the excess molar enthalpy and the excess molar volume data and the NRTL and UNIQUAC models were used to correlate the enthalpy of mixing data. The NRTL equation was found to be more suitable than the UNIQUAC equation for these systems. The results are discussed in terms of the polarizability of the aromatic compound and the effect of methyl substituents on the benzene ring.  相似文献   

13.
Excess molar volumes VmE of binary mixtures of 2,2,2-trifluoroethanol with water, or acetone, or methanol, or ethanol, or 1-alcholos, or 1,4-difluorobenzene, or 4-fluorotoluene or α,α,α-trifluorotoluene were measured in a vibrating tube densimeter at temperature 298.15 K and pressure of 101 kPa. The VmE extrema are: 1.540 cm3 · mol−1 for (2,2,2-trifluoroethanol + 1-heptanol); 1.452 cm3 · mol−1 for (2,2,2-trifluoroethanol + 1-hexanol); 1.238 cm3 · mol−1 for (2,2,2-trifluoroethanol + 1-butanol); 0.821 cm3 · mol−1 for (2,2,2-trifluoroethanol + 4-fluorotoluene); 0.817 cm3 · mol−1 for (2,2,2-trifluoroethanol + ethanol); 0.647 cm3 · mol−1 for (2,2,2-trifluoroethanol + methanol); 0.618 cm3 · mol−1 for (2,2,2-trifluoroethanol + acetone); 0.605 cm3 · mol−1 for (2,2,2-trifluoroethanol + α,α,α-trifluorotoluene); 0.485 cm3 · mol−1 for (2,2,2-trifluoroethanol + 1,4-difluorobenzene); and −0.656 cm3 · mol−1 for (2,2,2-trifluoroethanol + water). The limiting excess partial molar volumes are estimated.  相似文献   

14.
(Solid + liquid) equilibria (SLE) and (liquid + liquid) equilibria (LLE) for the binary systems: {ionic liquid (IL) N-butyl-4-methylpyridinium tosylate (p-toluenesulfonate) [BM4Py][TOS], or N-butyl-3-methylpyridinium tosylate [BM3Py][TOS], or N-hexyl-3-methylpyridinium tosylate [HM3Py][TOS], or N-butyl-4-methylpyridinium bis{(trifluoromethyl)sulfonyl}imide [BM4Py][NTf2], or 1,4-dimethylpyridinium tosylate [M1,4Py][TOS], or 2,4,6-collidine tosylate [M2,4,6Py][TOS], or 1-ethyl-3-methylimidazolium thiocyanate [EMIM][SCN], or 1-butyl-3-methylimidazolium thiocyanate [BMIM][SCN], or 1-hexyl-3-methylimidazolium thiocyanate [HMIM][SCN], or triethylsulphonium bis(trifluoromethylsulfonyl)imide [Et3S][NTf2] + thiophene} have been determined at ambient pressure. A dynamic method was used over a broad range of mole fractions and temperatures from (270 to 390) K. In the case of systems (pyridinium IL, or sulphonium IL + thiophene) the mutual immiscibility with an upper critical solution temperature (UCST) was detected at the very narrow and low mole fraction of the IL. For the binary systems containing (imidazolium thiocyanate IL + thiophene), the mutual immiscibility with the lower critical solution temperature (LCST) was detected at the higher mole fraction range of the IL. The basic thermal properties of the pure ILs, i.e. melting and glass-transition temperatures as well as the enthalpy of fusion have been measured using a differential scanning microcalorimetry technique (DSC). The well-known NRTL equation has been used to correlate experimental SLE/LLE data sets.  相似文献   

15.
Interaction of 1,1,1,3,3,3-hexafluoroisopropanol (HFIP) and isopropanol in the presence of equimolar quantities of guanidine thiocyanate (GndSCN) with bovine α-lactalbumin (α-LA) has been investigated by using a combination of isothermal titration calorimetry, circular dichroism, fluorescence, and ultra-violet spectroscopies at in 20 · 10?3 mol · dm?3 phosphate buffer pH 7.0. All the thermal unfolding transitions, in the presence of both the (alcohol + salt) mixtures were found to be reversible as judged by the same values of absorbance observed at different temperatures during cooling after the completion of thermal unfolding. In the presence of the 0.25 mol · dm?3 (HFIP + GndSCN) equimolar mixture and 0.85 mol · dm?3 (isopropanol + GndSCN) equimolar mixture, α-lactalbumin was observed to be in the partially folded state with significant loss of native tertiary structure. Intrinsic fluorescence results, acrylamide and potassium iodide quenching, 8-anilino-1-naphthalenesulfonic acid (ANS) binding, and energy transfer results also corroborate the presence of partially folded states of α-lactalbumin. Apart from the generation of the partially folded states, it was also observed that destabilizing action of GndSCN is reduced in the presence of isopropanol compared to that in HFIP. Isothermal titration calorimetry has been used to characterize the energetics of ANS binding to the partially folded states of the protein. ITC results indicate that ANS binds to these partially folded states at pH 7.0 due to the presence of two sequentially binding sites on the protein under the solvent conditions employed. For example, ANS binds to the 0.25 mol · dm?3 (HFIP + GndSCN) induced partially folded state with affinity constants K1 = (858 ± 220), K2 = (1.12 ± 0.25) · 103; enthalpies of binding ΔH1 = (4.4 ± 1.0) kJ · mol?1, ΔH2 = (2.1 ± 0.2) kJ · mol?1; and entropies of binding ΔS1 = 70 J · K?1 · mol?1 and ΔS2 = 65 J · K?1 · mol?1, respectively at these two sequential binding sites. In light of the fluorescence results, possible binding sites where ANS can bind to the protein have also been suggested.  相似文献   

16.
Two new hybrid materials, (C4H14N2)[MII(H2O)6](SO4)2·4H2O (MII: Co (I), Ni (II)), have been synthesised by slow evaporation method at room temperature and crystallographically characterized. They crystallise isotypically in the monoclinic system, space group P21/n, with the following unit-cell parameters a = 9.2285(3), b = 11.3333(4), c = 10.6693(4) Å, β = 109.004(2)°, Z = 2 and V = 1055.07(6) Å3 for I and a = 9.2127(2), b = 11.3182(2), c = 10.6434(2) Å, β = 109.094(1)°, Z = 2 and V = 1048.74(4) Å3 for II. The structure of the two supramolecular compounds consists of metallic cation octahedrally coordinated to six water molecules, sulfate anions, 1,4-butanediammonium cation and water molecules linked together via two types of hydrogen bonds, O–H?O and N–H?O. The two compounds are not stable at room temperature and their partial dehydration depends on the humidity of the environment. The thermal decomposition of precursors, studied by thermogravimetric analysis (TG) and temperature-dependent X-ray diffraction (TDXD), shows successive intermediate hydrates and crystalline anhydrous compounds upon dehydration.  相似文献   

17.
In order to add to the existing knowledge of aqueous solution behavior of bile salts in presence of amino acids, the micellization properties of sodium cholate (NaC) (1 to 20) mmol · kg−1, and sodium deoxycholate (NaDC) (0.5 to 10) mmol · kg−1 in 0.1 mol · kg−1 aqueous solution of glycine, leucine, methionine, and histidine have been investigated at different temperatures (293.15 to 318.15) K at intervals of T = 5 K by using conductivity and fluorescence probe studies. The critical micelle concentration (CMC) values have been determined and elucidated in terms of hydrophobicity as well as hydrophilicity of NaC and NaDC in aqueous solution of these additives. Thermodynamic parameters of micellization viz. standard Gibbs free energy (ΔmicGo), standard enthalpy (ΔmicHo), and standard entropy (ΔmicSo) have also been calculated to extract information regarding the nature of micellization of bile salts in aqueous solutions. The (enthalpy + entropy) compensation plots have been interpreted to the contribution of chemical part towards micellization or stability of the micelle formed.  相似文献   

18.
Recent years have witnessed burgeoning interest in plant flavonoids as novel therapeutic drugs targeting cellular membranes and proteins. Motivated by this scenario, we explored the binding of robinetin (3,7,3′,4′,5′-pentahydroxyflavone, a bioflavonoid with remarkable ‘two color’ intrinsic fluorescence properties), with egg yolk phosphatidylcholine (EYPC) liposomes and normal human hemoglobin (HbA), using steady state and time resolved fluorescence spectroscopy. Distinctive fluorescence signatures obtained for robinetin indicate its partitioning (Kp = 8.65 × 104) into the hydrophobic core of the membrane lipid bilayer. HbA–robinetin interaction was examined using both robinetin fluorescence and flavonoid-induced quenching of the protein tryptophan fluorescence. Specific interaction with HbA was confirmed from three lines of evidence: (a) bimolecular quenching constant Kq ? diffusion controlled limit; (b) closely matched values of Stern–Volmer quenching constant and binding constant; (c) τ0/τ = 1 (where τ0 and τ are the unquenched and quenched tryptophan fluorescence lifetimes, respectively). Absorption spectrophotometric assays reveal that robinetin inhibits EYPC membrane lipid peroxidation and HbA glycosylation with high efficiency.  相似文献   

19.
Seven Cd(II)–ferrocenesuccinate coordination complexes with the formulas [Cd(η2-FcCOC2H4COO)2(pbbbm)]2 (1), [Cd(η2-FcCOC2H4COO)(pbbbm)Cl]2 (2), [Cd(η2-FcCOC2H4COO)(pbbbm)I]2 (3), {[Cd(η2-FcCOC2H4COO)2(btx)2]2(CH3OH)0.5} (4), [Cd(η2-FcCOC2H4COO)2(bix)]2(H2O) (5), {[Cd(η2-FcCOC2H4COO)(bbbm)1.5Cl] · (CH3OH)0.5}n (6), and {[Cd(η2-FcCOC2H4COO)(mbbbm)Cl] · (H2O)2.75}n (7) [pbbbm = 1,4-Bis(benzimidazole-1-ylmethyl)benzene), btx = 1,4-bis(triazol-1-ylmethyl)benzene), mbbbm = 1,3-bis(benzimidazole-1-ylmethyl)benzene), bix = 1,4-bis(imidazol-1-ylmethyl)benzene, bbbm = 1,1-(1,4-Butanediyl)bis-1H-benzimidazole)] have been synthesized and characterized. Single-crystal X-ray analysis reveals that complexes 15 are all dimers and bridged by pbbbm, btx and bix, respectively. But the five complexes present some differences in their dimeric conformations, which can be ascribed to the impacts of adjuvant ligands and counter anions. In contrast to complexes 1–5, both 6 and 7 are of 1-D structures (with the same counter anions), and the former is double ladder-like structure only bridged by bbbm, while the latter is chain-like structure bridged by chlorine anions and adjuvant ligand mbbbm. Notably, various π–π interactions are found in complexes 17, and they have significant contributions to molecular self-assembly processes. The electrochemical studies of complexes 17 in DMF solution display irreversible redox waves and indicate that the half-wave potentials of the ferrocenyl moieties in these complexes are all shifted to positive potential compared with that of ferrocenesuccinate.  相似文献   

20.
The vapour pressures of {ethanediamine (EDA) + water}, {1,2-diaminopropane (1,2-DAP) + water}, {1,3-diaminopropane (1,3-DAP) + water} or {1,4-diaminobutane (1,4-DAB) + water} binary mixtures, and of pure EDA, 1,2-DAP, 1,3-DAP, 1,4-DAB, and water components were measured by means of two static devices at temperatures between (293 and 363) K. The data were correlated with the Antoine equation. From these data, the excess Gibbs function (GE) was calculated for several constant temperatures and fitted to a fourth-order Redlich–Kister equation using the Barker’s method. The {ethanediamine (EDA) + water}, and {1,2-diaminopropane (1,2-DAP) + water} binary systems show negative azeotropic behaviour. The aqueous solutions of EDA, 1,2-DAP, or 1,3-DAP exhibit negative deviations in GE for all investigated temperatures over the whole composition range whereas the (1,4-DAB + water) binary mixture shows negative GE for temperatures (293.15 < T/K < 353.15) and a sinusoidal shape for GE at T = 363.15 K.  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