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 105 毫秒
1.
N. Xaba  D. Jaganyi 《Polyhedron》2009,28(6):1145-1149
Hydroboration reactions of 4-octene with HBBr2 · SMe2, HBCl2 · SMe2 and H2BBr · SMe2 in CH2Cl2 were studied as function of concentration and temperature and compared with those of 1-octene. On average, hydroboration with dihaloborane proceeded 16 times slower for 4-octene than for 1-octene. In the case of the reactions with the monohaloborane, this factor is halved. This can be explained by the difference in the relative rates of dissociates of Me2S from the dihaloborane and a monohaloborane complex, respectively. The reactions involving H2BBr · SMe2 also exhibited a k?2 value, an indication of the presence of a parallel reaction, most likely a rearrangement process facilitating isomerization by way of a π-complex. The moderate ΔH values accompanied by small ΔS values (94 ± 4 kJ mol?1, ?3 ± 13 J K?1 mol?1 for HBBr2 · SMe2; 93 ± 1 kJ mol?1, ?17 ± 4 J K?1 mol?1 for HBCl2 · SMe2 and in the case of H2BBr · SMe2, 90 ± 13 kJ mol?1, +12 ± 44 J K?1 mol?1 and 83 ± 13 kJ mol?1, ?24 ± 45 J K?1 mol?1, respectively, for the k2 and k?2 processes) imply a process that is dissociatively dominated, with the overall mode of activation being interchange dissociative (Id).  相似文献   

2.
Two pure hydrated lead borates, Pb(BO2)2·H2O and PbB4O7·4H2O, have been characterized by XRD, FT-IR, DTA-TG techniques and chemical analysis. The molar enthalpies of solution of Pb(BO2)2·H2O and PbB4O7·4H2O in 1 mol dm?3 HNO3(aq) were measured to be (?35.00 ± 0.18) kJ mol?1 and (35.37 ± 0.14) kJ mol?1, respectively. The molar enthalpy of solution of H3BO3(s) in 1 mol dm?3 HNO3(aq) was measured to be (21.19 ± 0.18) kJ mol?1. The molar enthalpy of solution of PbO(s) in (HNO3 + H3BO3)(aq) was measured to be ?(61.84 ± 0.10) kJ mol?1. From these data and with incorporation of the enthalpies of formation of PbO(s), H3BO3(s) and H2O(l), the standard molar enthalpies of formation of ?(1820.5 ± 1.8) kJ mol?1 for Pb(BO2)2·H2O and ?(4038.1 ± 3.4) kJ mol?1 for PbB4O7·4H2O were obtained on the basis of the appropriate thermochemical cycles.  相似文献   

3.
The solubility measurements of sodium dicarboxylate salts; sodium oxalate, malonate, succinate, glutarate, and adipate in water at temperatures from (278.15 to 358.15 K) were determined. The molar enthalpies of solution at T = 298.15 K were derived: ΔsolHm (m = 2.11 mol · kg?1) = 13.86 kJ · mol?1 for sodium oxalate; ΔsolHm (m = 3.99 mol · kg?1) = 14.83 kJ · mol?1 for sodium malonate; ΔsolHm (m = 2.45 mol · kg?1) = 14.83 kJ · mol?1 for sodium succinate; ΔsolHm (m = 4.53 mol · kg?1) = 16.55 kJ · mol?1 for sodium glutarate, and ΔsolHm (m = 3.52 mol · kg?1) = 15.70 kJ · mol?1 for sodium adipate. The solubility value exhibits a prominent odd–even effect with respect to terms with odd number of sodium dicarboxylate carbon numbers showing much higher solubility. This odd–even effect may have implications for the relative abundance of these compounds in industrial applications and also in the atmospheric aerosols.  相似文献   

4.
The mobility of uranium under oxidizing conditions can only be modeled if the thermodynamic stabilities of the secondary uranyl minerals are known. Toward this end, we synthesized metaschoepite (UO3(H2O)2), becquerelite (Ca(UO2)6O4(OH)6(H2O)8), compreignacite (K2(UO2)6O4(OH)6(H2O)7), sodium compreignacite (Na2(UO2)6O4(OH)6(H2O)7), and clarkeite (Na(UO2)O(OH)) and performed solubility measurements from both undersaturation and supersaturation under controlled-pH conditions. The solubility measurements rigorously constrain the values of the solubility products for these synthetic phases, and consequently the standard-state Gibbs free energies of formation of the phases. The calculated lg solubility product values (lg Ksp), with associated 1σ uncertainties, for metaschoepite, becquerelite, compreignacite, sodium compreignacite, and clarkeite are (5.6 ?0.2/+0.1), (40.5 ?1.4/+0.2), (35.8 ?0.5/+0.3), (39.4 ?1.1/+0.7), and (9.4 ?0.9/+0.6), respectively. The standard-state Gibbs free energies of formation, with their 2σ uncertainties, for these same phases are (?1632.2 ± 7.4) kJ · mol?1, (?10305.6 ± 26.5) kJ · mol?1, (?10107.3 ± 21.8) kJ · mol?1, (?10045.6 ±24.5) kJ · mol?1, and (?1635.1 ± 23.4) kJ · mol?1, respectively. Combining our data with previously measured standard-state enthalpies of formation for metaschoepite, becquerelite, sodium compreignacite, and clarkeite yields calculated standard-state entropies of formation, with associated 2σ uncertainties, of (?532.5 ± 8.1) J · mol?1 · K?1, (?3634.5 ± 29.7) J · mol?1 · K?1, ( ?2987.6 ± 28.5) J · mol?1 · K?1, and (?300.5 ± 23.9) J · mol?1 · K?1, respectively. The measurements and associated calculated thermodynamic properties from this study not only describe the stability and solubility at T = 298 K, but also can be used in predictions of uranium mobility through extrapolation of these properties to temperatures and pressures of geologic and environmental interest.  相似文献   

5.
The relative hydrophobicity of the phases of several {polyethylene glycol (PEG) 8000 + sodium sulfate (Na2SO4)} aqueous two-phase systems (ATPSs), all containing 0.01 mol · L?1 sodium phosphate buffer (NaPB, pH 7.4) and increasing concentration of a salt additive, NaCl or KCl, up to 1.0 mol · L?1, was measured by the free energy of transfer of a methylene group between the phases, ΔG(CH2). The ΔG(CH2) of the systems was determined by partitioning of a homologous series of five sodium salts of dinitrophenylated (DNP) – amino acids with aliphatic side chains in three different tie-lines of each biphasic system. The relative hydrophobicity of the phases ranged from ?0.125 to ?0.183 kcal · mol?1, being the NaCl salt the one to provide the more effective changes. The results show that, within each system, there is a linear relationship between the ΔG(CH2) and the tie-line length (TLL), and biphasic systems with high salt additive concentration present the most negative ΔG(CH2) values. Therefore, the feasibility of establishing a relationship between the relative hydrophobicity of the phases in a given TLL and the ionic strength of the salt additive was investigated and a satisfactory correlation was found for each salt.  相似文献   

6.
The solubility of anthracene was measured in pure water and in sodium chloride aqueous solution (salt concentration, m/mol · kg?1 = 0.1006, 0.5056, and 0.6082) at temperatures between (278 and 333) K. Solubility of anthracene in pure water agrees fairly well with values reported in earlier similar studies. Solubility of anthracene in sodium chloride aqueous solutions ranged from (6 · 10?8 to 143 · 10?8) mol · kg?1. Sodium chloride had a salting-out effect on the solubility of anthracene. The salting-out coefficients did not vary significantly with temperature over the range studied. The average salting-out coefficient for anthracene was 0.256 kg · mol?1.The standard molar Gibbs free energies, ΔtrG°, enthalpies, ΔtrH°, and entropies, ΔtrS°, for the transfer of anthracene from pure water to sodium chloride aqueous solutions were also estimated. Most of the estimated ΔtrG° values were positive [(20 to 1230) J · mol?1]. The analysis of the thermodynamic parameters shows that the transfer of anthracene from pure water to sodium chloride aqueous solution is thermodynamically unfavorable, and that this unfavorable condition is caused by a decrease in entropy.  相似文献   

7.
Thermodynamic properties of Mg(NH2)2 and LiNH2 were investigated by measurements of NH3 pressure-composition isotherms (PCI). Van’t Hoff plot of plateau pressures of PCI for decomposition of Mg(NH2)2 indicated the standard enthalpy and entropy change of the reactions were ΔH° = (120 ± 11) kJ · mol?1 (per unit amount of NH3) and ΔS° = (182 ± 19) J · mol?1 · K?1 for the reaction: Mg(NH2)2  MgNH + NH3, and ΔH° = 112 kJ · mol?1 and ΔSo = 157 J · mol?1 · K?1 for the reaction: MgNH  (1/3)Mg3N2 + (1/3)NH3. PCI measurements for formation of LiNH2 were carried out, and temperature dependence of plateau pressures indicated ΔH° = (?108 ± 15) kJ · mol?1 and ΔS° = (?143 ± 25) J · mol?1 · K?1 for the reaction: Li2NH + NH3  2LiNH2.  相似文献   

8.
The enthalpy of formation of zinc acetate dihydrate (Zn(CH3COO)2 · 2H2O) was measured with respect to crystalline zinc oxide (ZnO), glacial acetic acid (CH3COOH) and liquid water by room temperature solution calorimetry. The enthalpy of formation was verified by utilizing two independent thermodynamic cycles, using enthalpy of solution measurements in 5 mol · L?1 sodium hydroxide (NaOH) and in 5 mol · L?1 hydrochloric acid (HCl) solutions. The enthalpy of the reaction ZnO (cr) + 2CH3COOH (l) + H2O (l) to form Zn(CH3COO)2 · 2H2O (cr) is –(65.78 ± 0.36) kJ · mol?1 for measurements in 5 mol · L?1 NaOH and –(66.25 ± 0.17) kJ · mol?1 for measurements in 5 mol · L?1 HCl. The standard enthalpy of formation of Zn(CH3COO)2 · 2H2O from the elements is –(1669.35 ± 1.30) kJ · mol?1. This work provides the first calorimetric measurement of the enthalpy of formation of Zn(CH3COO)2 · 2H2O.  相似文献   

9.
The low-temperature heat capacity of NiAl2O4 and CoAl2O4 was measured between T = (4 and 400) K and thermodynamic functions were derived from the results. The measured heat-capacity curves show sharp anomalies peaking at around T = 7.5 K for NiAl2O4 and at T = 9 K for CoAl2O4. The exact cause of these anomalies is unknown. From our results, we suggest a standard entropy for NiAl2O4 at T = 298.15 K of (97.1 ± 0.2) J · mol?1 · K?1 and for CoAl2O4 of (100.3 ± 0.2) J · mol?1 · K?1.  相似文献   

10.
In reply to “Comment on the possible role of reaction H+H2O→H2+OH in the radiolysis of water at high temperatures” (Bartels, 2009 Comment on the possible role of the reaction H+H2O→H2+OH in the radiolysis of water at high temperatures. Radiat. Phys. Chem. 78, 191–194) we present an alternative thermodynamic estimation of the reaction rate constant k. Based on the non-symmetric standard state convention we have calculated that the Gibbs energy of reaction ΔrG=57.26 kJ mol?1 and the reaction rate constant k=7.23×10?5 M?1 s?1 at ambient temperature. Re-analysis of the thermodynamic estimation (Bartels, 2009 Comment on the possible role of the reaction H+H2O→H2+OH in the radiolysis of water at high temperatures. Radiat. Phys. Chem. 78, 191–194) showed that the upper limit for the rate constant at 573 K is k=1.75×104 M?1 s?1 compared to the value predicted by the diffusion-kinetic modelling (3.18±1.25)×104 M?1 s?1 (Swiatla-Wojcik, D., Buxton, G.V., 2005. On the possible role of the reaction H+H2O→H2+OH in the radiolysis of water at high temperatures. Radiat. Phys. Chem. 74(3–4), 210–219). The presented thermodynamic evaluation of k(573) is based on the assumption that k can be calculated from ΔrG and the rate constant of the reverse reaction which, as discussed, are both uncertain at high temperatures.  相似文献   

11.
The pure hydrated metalloborophosphate sample, Na2[CuB3P2O11(OH)]·0.67H2O, has been synthesized and characterized by XRD, FT-IR, DTA-TG techniques, and chemical analysis. The molar enthalpies of solution of Na2[CuB3P2O11(OH)]·0.67H2O(s) in 1 mol · dm?3 HCl (aq), of Cu(OH)2 (s) in (HCl + H3BO3) (aq), and of NaH2PO4·2H2O (s) in (HCl + H3BO3 + Cu(OH)2) (aq) were measured, respectively. With the incorporation of the previously determined enthalpy of solution of H3BO3 (s) in 1 mol · dm?3 HCl (aq), together with the use of the standard molar enthalpies of formation for NaH2PO4·2H2O (s), Cu(OH)2 (s), H3BO3 (s), and H2O (l), the standard molar enthalpy of formation of ?(4988.4 ± 2.5) kJ · mol?1 for Na2[CuB3P2O11(OH)]·0.67H2O at T = 298.15 K was obtained on the basis of the appropriate thermochemical cycle.  相似文献   

12.
Unless the radiolytic reducing species are neutralised or converted into oxidising species, an EB remediation system cannot be considered a true Advanced Oxidation Processes (AOP). A water/H2O2 system irradiated by UVC mercury lamps constitutes a widely used OH production method. Employing H2O2 in radiolysis as well, an enhancement of the oxidative efficiency of an EB treatment can be obtained. Pulse radiolysis measurements of an aerated aqueous/H2O2/KSCN system have been systematically undertaken to assess the optimal H2O2 concentration. By linearly fitting a competition kinetics relationship, it is found that the scavengeable extra-yield of OH is ΔG(OH)=0.24 μmol J?1 (R=0,9958), while the maximum experimental yield is measured G(OH)max=(0.52±0.02) μmol J?1 when [H2O2]=5–10 mM. Exceeding these concentrations the OH yield drops off.  相似文献   

13.
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   

14.
We determined apparent molar volumes V? at 298.15 ? (T/K) ? 368.15 and apparent molar heat capacities Cp,? at 298.15 ? (T/K) ? 393.15 for aqueous solutions of HIO3 at molalities m from (0.015 to 1.0) mol · kg?1, and of aqueous KIO3 at molalities m from (0.01 to 0.2) mol · kg?1 at p = 0.35 MPa. We also determined V? at the same p and at 298.15 ? (T/K) ? 368.15 for aqueous solutions of KI at m from (0.015 to 7.5) mol · kg?1. We determined Cp,? at the same p and at 298.15 ? (T/K) ? 393.15 for aqueous solutions of KI at m from (0.015 to 5.5) mol · kg?1, and for aqueous solutions of NaIO3 at m from (0.02 to 0.15) 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 were fitted to our results for each compound. Values of Ka, ΔrHm, and ΔrCp,m for the proton ionization reaction of aqueous HIO3 are calculated and discussed.  相似文献   

15.
Molecular interactions of five thiazine dyes with increasing alkyl substitution have been studied in aqueous and microemulsion media at 303 K within a concentration range of (1.35–7.00) × 10?4 M. The dimerization constant (Kd) values for the five dyes are ranged between 1.761 and 6.258 × 103 l mol?1 in bulk water media, where as in microemulsion media, Kd's are ranged between 1.760 and 4.110 × 103 l mol?1. Thionine (with no methyl substitution) and azure A (with two methyl substitution) displayed slightly larger Kd values in microemulsion water pools compared to bulk water while other dyes recorded significant drop in Kd values. The influence of microemulsion media on the molecular interaction of dyes has been explained in terms of electrostatic and hydrophobic factors. The monomer and the dimer spectra are explained in terms of molecular exciton model and the optical absorption parameters of both the species are reported in bulk and confined media.  相似文献   

16.
The heat capacity of polycrystalline germanium disulfide α-GeS2 has been measured by relaxation calorimetry, adiabatic calorimetry, DSC and heat flux calorimetry from T = (2 to 1240) K. Values of the molar heat capacity, standard molar entropy and standard molar enthalpy are 66.191 J · K?1 · mol?1, 87.935 J · K?1 · mol?1 and 12.642 kJ · mol?1. The temperature of fusion and its enthalpy change are 1116 K and 23 kJ · mol?1, respectively. The thermodynamic functions of α-GeS2 were calculated over the range (0 ? T/K ? 1250).  相似文献   

17.
Activity coefficients of CaCl2 in disaccharide {(maltose, lactose) + water} mixtures at 298.15 K were determined by cell potentials. The molalities of CaCl2 ranged from about 0.01 mol · kg?1 to 0.20 mol · kg?1, the mass fractions of maltose from 0.05 to 0.25, and those of lactose from 0.025 to 0.125. The cell potentials were analyzed by using the Debye–Hückel extended equation and the Pitzer equation. The activity coefficients obtained from the two theoretical models are in good agreement with each other. Gibbs free energy interaction parameters (gES) and salting constants (kS) were also obtained. These were discussed in terms of the stereo-chemistry of saccharide molecules and the structural interaction model.  相似文献   

18.
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   

19.
20.
Two pure zinc borates with microporous structure 3ZnO·3B2O3·3.5H2O and 6ZnO·5B2O3·3H2O have been synthesized and characterized by XRD, FT-IR, TG techniques and chemical analysis. The molar enthalpies of solution of 3ZnO·3B2O3·3.5H2O(s) and 6ZnO·5B2O3·3H2O(s) in 1 mol · dm−3 HCl(aq) were measured by microcalorimeter at T = 298.15 K, respectively. The molar enthalpies of solution of ZnO(s) in the mixture solvent of 2.00 cm3 of 1 mol · dm−3 HCl(aq) in which 5.30 mg of H3BO3 were added were also measured. With the incorporation of the previously determined enthalpy of solution of H3BO3(s) in 1 mol · dm−3 HCl(aq), together with the use of the standard molar enthalpies of formation for ZnO(s), H3BO3(s), and H2O(l), the standard molar enthalpies of formation of −(6115.3 ± 5.0) kJ · mol−1 for 3ZnO·3B2O3·3.5H2O and −(9606.6 ± 8.5) kJ · mol−1 for 6ZnO·5B2O3·3H2O at T = 298.15 K were obtained on the basis of the appropriate thermochemical cycles.  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