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 312 毫秒
1.
Thermal behavior, relative stability, and enthalpy of formation of α (pink phase), β (blue phase), and red NaCoPO4 are studied by differential scanning calorimetry, X-ray diffraction, and high-temperature oxide melt drop solution calorimetry. Red NaCoPO4 with cobalt in trigonal bipyramidal coordination is metastable, irreversibly changing to α NaCoPO4 at 827 K with an enthalpy of phase transition of −17.4±6.9 kJ mol−1. α NaCoPO4 with cobalt in octahedral coordination is the most stable phase at room temperature. It undergoes a reversible phase transition to the β phase (cobalt in tetrahedra) at 1006 K with an enthalpy of phase transition of 17.6±1.3 kJ mol−1. Enthalpy of formation from oxides of α, β, and red NaCoPO4 are −349.7±2.3, −332.1±2.5, and −332.3±7.2 kJ mol−1; standard enthalpy of formation of α, β, and red NaCoPO4 are −1547.5±2.7, −1529.9±2.8, and −1530.0±7.3 kJ mol−1, respectively. The more exothermic enthalpy of formation from oxides of β NaCoPO4 compared to a structurally related aluminosilicate, NaAlSiO4 nepheline, results from the stronger acid-base interaction of oxides in β NaCoPO4 (Na2O, CoO, P2O5) than in NaAlSiO4 nepheline (Na2O, Al2O3, SiO2).  相似文献   

2.
Enthalpies for the two proton ionizations of glycine, N,N-bis(2-hyroxyethyl)glycine (“bicine”) and N-tris(hydroxymethyl)methylglycine (“tricine”) were obtained in water-methanol mixtures with methanol mole fraction (Xm) from 0 to 0.360. With increasing methanol the ionization enthalpy for the first proton (ΔH1) of glycine increased from 4.4 to 9.4 kJ mol−1 with a minimum of 4.1 kJ mol−1 at Xm = 0.059. The ionization enthalpy of the second proton (ΔH2) for glycine decreased from 46.3 to 38.1 kJ mol−1. ΔH1 of bicine increased from 3.5 to 7.6 kJ mol−1 at Xm = 0.273 before dropping to 4.1 kJ mol−1 at Xm = 0.360. ΔH2 of bicine increased from 24.9 to 29.4 kJ mol−1. For tricine, ΔH1 increased from 6.7 to 9.8 kJ mol−1 at Xm = 0.194 then dropped to 7.4 kJ mol−1 at Xm = 0.360. ΔH2 for tricine first dropped from 30.8 to 28.5 kJ mol−1 at Xm = 0.059 before increasing to 33.3 kJ mol−1 at Xm = 0.273. The solvent composition was selected so as to include the region of maximum structure enhancement of water by methanol. The results were interpreted in terms of solvent-solvent and solvent-solute interactions.  相似文献   

3.
Enthalpies for the two proton ionizations of the biochemical buffers N-[2-hydroxyethyl]piperazine-N′-[2-ethane sulfonic acid] (HEPES) and N-[2-hydroxyethyl]piperazine-N′-[2-hydroxypropane sulfonic acid] (HEPPSO) were obtained in water-methanol mixtures with methanol mole fraction (Xm) from 0 to 0.360. With increasing methanol, the ionization enthalpy for the first proton (ΔH1) of HEPES increased steadily from 8.4 to 15.3 kJ mol−1 whereas that for HEPPSO rose to a maximum of 21.0 kJ mol−1 at Xm = 0.123 before dropping to 18.4 kJ mol−1 at Xm = 0.360. The ionization enthalpy for the second proton (ΔH2) of HEPES varied from 20.8 kJ mol−1 in water to 13.6 kJ mol−1 at Xm = 0.360 with a maximum of 24.8 kJ mol−1 at Xm = 0.194. For HEPPSO, ΔH2 increased steadily from 23.4 to 29.2 kJ mol−1. The solvent composition was selected so as to include the region of maximum structure enhancement of water by methanol. The results were interpreted in terms of solvent-solvent and solvent-solute interactions.  相似文献   

4.
This contributions shows with a series of ab initio MP2 and DFT (BP86 and B3-LYP) computations with large basis sets up to cc-pVQZ quality that the literature value of the standard enthalpy of depolymerization of Sb4F20(g) to give SbF5(g) (+18.5 kJ mol−1) [J. Fawcett, J.H. Holloway, R.D. Peacock, D.R. Russell, J. Fluorine Chem. 20 (1982) 9] is by about 50 kJ mol−1 in error and that the correct value of (Sb4F20(g)) is +68 ± 10 kJ mol−1. We assign , , and values for SbnF5n with n = 2-4 and compare the results to available experimental gas phase data. Especially the MP2/TZVPP values obtained in an indirect procedure that rely on isodesmic reactions or the highly accurate compound methods G2 and CBS-Q are in excellent agreement with the experimental data, and reproduce also the fine experimental details at temperatures of 423 and 498 K. With these data and the additional calculation of [SbnF5n+1] (n = 1-4), we then assessed the fluoride ion affinities (FIAs) of SbnF5n(g), nSbF5(g), nSbF5(l) and the standard enthalpies of formation of SbnF5n(g) and [SbnF5n+1](g): FIA(SbnF5n(g)) = 514 (n = 1), 559 (n = 2), 572 (n = 3) and 580 (n = 4) kJ mol−1; FIA(nSbF5(g)) = 667 (n = 2), 767 (n = 3) and 855 (n = 4) kJ mol−1; FIA(nSbF5(l)) = 434 (n = 1), 506 (n = 2), 528 (n = 3) and 534 (n = 4) kJ mol−1. Error bars are approximately ±10 kJ mol−1. Also the related Gibbs energies were derived. ΔfH°([SbnF5n+1](g)) = −2064 ± 18 (n = 1), −3516 ± 25 (n = 2), −4919 ± 31 (n = 3) and −6305 ± 36 (n = 4) kJ mol−1.  相似文献   

5.
In the present study, the stability of gaseous barium silicates was confirmed by the high temperature mass spectrometry. On the basis of equilibrium constants measured for gas-phase reactions, the standard formation enthalpies were determined for gaseous barium silicates as (−510 ± 15) kJ · mol−1 and (−884 ± 18) kJ · mol−1 at 298 K; standard atomization enthalpies as (1637 ± 17) kJ · mol−1 and (2318 ± 20) kJ · mol−1 at 298 K for BaSiO2 and BaSiO3, respectively. Based on the results obtained the critical analysis of the literature data was carried out.  相似文献   

6.
The kinetics and mechanism of the hydroboration reactions of 1-octene with HBBr2 · SMe2 and HBCl2 · SMe2, in CH2Cl2 as a solvent, were studied. Rates of hydroboration were monitored using 11B NMR spectroscopy. The reactions exhibited simple second-order kinetics of the form . The HBCl2 · SMe2 was found to be 20 times more reactive than the HBBr2 · SMe2. The overall activation parameters (ΔH, ΔS) for the reaction of HBBr2 · SMe2 with 1-octene were found to be 82 ± 1 kJ mol−1, −18 ± 4 J K−1 mol−1 and with 1-hexyne were 78 ± 4 kJ mol−1 −34 ± 12 J K−1 mol−1. For the reaction of HBCl2 · SMe2 with 1-octene, ΔH and ΔS were 104 ± 5 kJ mol−1 and 43 ± 16 J K−1 mol−1, respectively. The activation parameters (ΔH, ΔS) for the dissociation of Me2S from HBBr2 · SMe2 were found to be 104 ± 2 kJ mol−1, +33 ± 8 J K−1 mol−1, respectively. Based on the activation parameters, it was concluded that the detaching of Me2S from the boron centre follows a dissociative mechanism, while the hydroboration process follows an associative pathway. It was also concluded that the dissociation of Me2S from the boron centre is the rate determining step.  相似文献   

7.
Low-temperature heat capacities of the compound Na(C4H7O5)·H2O(s) have been measured with an automated adiabatic calorimeter. A solid-solid phase transition and dehydration occur at 290-318 K and 367-373 K, respectively. The enthalpy and entropy of the solid-solid transition are ΔtransHm = (5.75 ± 0.01) kJ mol−1 and ΔtransSm = (18.47 ± 0.02) J K−1 mol−1. The enthalpy and entropy of the dehydration are ΔdHm = (15.35 ± 0.03) kJ mol−1 and ΔdSm = (41.35 ± 0.08) J K−1 mol−1. Experimental values of heat capacities for the solids (I and II) and the solid-liquid mixture (III) have been fitted to polynomial equations.  相似文献   

8.
Two pure strontium borates SrB2O4·4H2O and SrB2O4 have been synthesized and characterized by means of chemical analysis and XRD, FT-IR, DTA-TG techniques. The molar enthalpies of solution of SrB2O4·4H2O and SrB2O4 in 1 mol dm−3 HCl(aq) were measured to be −(9.92 ± 0.20) kJ mol−1 and −(81.27 ± 0.30) kJ mol−1, respectively. The molar enthalpy of solution of Sr(OH)2·8H2O in (HCl + H3BO3)(aq) were determined to be −(51.69 ± 0.15) kJ mol−1. With the use of the enthalpy of solution of H3BO3 in 1 mol dm−3 HCl(aq), and the standard molar enthalpies of formation for Sr(OH)2·8H2O(s), H3BO3(s), and H2O(l), the standard molar enthalpies of formation of −(3253.1 ± 1.7) kJ mol−1 for SrB2O4·4H2O, and of −(2038.4 ± 1.7) kJ mol−1 for SrB2O4 were obtained.  相似文献   

9.
A complex of holmium perchlorate coordinated with l-glutamic acid, [Ho2(l-Glu)2(H2O)8](ClO4)4·H2O, was prepared with a purity of 98.96%. The compound was characterized by chemical, elemental and thermal analysis. Heat capacities of the compound were determined by automated adiabatic calorimetry from 78 to 370 K. The dehydration temperature is 350 K. The dehydration enthalpy and entropy are 16.34 kJ mol−1 and 16.67 J K−1 mol−1, respectively. The standard enthalpy of formation is −6474.6 kJ mol−1 from reaction calorimetry at 298.15 K.  相似文献   

10.
The XRD, SEM, isothermal oxidation-weight loss and non-isothermal thermogravimetry (TG)-differential thermogravimetry (DTG) were used to study the oxidation properties and oxidation decomposition kinetics of three-dimensional (3-D) braided carbon fiber (abbreviated as fiber). The results showed that the non-isothermal oxidation process of fiber exhibited self-catalytic characteristic. The kinetic parameters and oxidation mechanism of fiber were studied through analyzing the TG and DTG data by differential and integral methods. The oxidation mechanism was random nucleation, the kinetic parameters were: lg A=10.299 min−1; Ea=156.29 kJ mol−1.  相似文献   

11.
Isothermal depolymerization of the two polymers of C60, i.e. of 1D orthorhombic phase (O) and of “dimer state” (DS) have been studied by means of Infra-red spectroscopy in the temperature ranges 383-423 and 453-503 K, respectively. Differential Scanning Calorimetry (DSC) has been used to obtained depolymerization polytherms for O-phase and DS. Standard set of reaction models have been applied to describe depolymerization behavior of O-phase and DS. The choice of the reaction models was based primarily on the isotherms. Several models however demonstrated almost equal goodness of fit and were statistically indistinguishable. In this case we looked for simpler/more realistic mechanistic model of the reaction. For DS the first-order expression (Mampel equation) with the activation energy Ea = 134 ± 7 kJ mol−1 and preexponential factor ln(A/s−1) = 30.6 ± 2.1, fitted the isothermal data. This activation energy was nearly the same as the activation energy of the solid-state reaction of dimerization of C60 reported in the literature. This made the enthalpy of depolymerization close to zero in accord with the DSC data on depolymerization of DS. Mampel equation gave the best fit to the polythermal data with Ea = 153 kJ mol−1 and preexponential factor ln(A/s−1) = 35.8. For O-phase two reasonable reaction models, i.e. Mampel equation and “contracting spheres” model equally fitted to the isothermal data with Ea = 196 ± 2 and 194 ± 8 kJ mol−1, respectively and ln(A/s−1) = 39.1 ± 0.5 and 37.4 ± 0.2, respectively and to polythermal data with Ea = 163 and 170 kJ mol−1, respectively and ln(A/s−1) = 32.5 and 29.5, respectively.  相似文献   

12.
Synthetic Na-magadiite sample was used for organofunctionalization process with N-propyldiethylenetrimethoxysilane and bis[3-(triethoxysilyl)propyl]tetrasulfide, after expanding the interlayer distance with polar organic solvents such as dimethylsulfoxide (DMSO). The resulted materials were submitted to process of adsorption with arsenic solution at pH 2.0 and 298±1 K. The adsorption isotherms were adjusted using a modified Langmuir equation with regression nonlinear; the net thermal effects obtained from calorimetric titration measurements were adjusted to a modified Langmuir equation. The adsorption process was exothermic (ΔintH=−4.15-5.98 kJ mol−1) accompanied by increase in entropy (ΔintS=41.32-62.20 J k−1 mol−1) and Gibbs energy (ΔintG=−22.44−24.56 kJ mol−1). The favorable values corroborate with the arsenic (III)/basic reactive centers interaction at the solid-liquid interface in the spontaneous process.  相似文献   

13.
Bispropargyl ether of bisphenol-A (BPEBPA), 4,4′-bismaleimido diphenyl ether (BMIE) and a blend consisting 60 mol% of BPEBPA and 40 mol% of BMIE are prepared. The materials are structurally characterized by FTIR. The curing characteristics of the monomers are measured by FTIR and DSC. The results indicated that BPEBPA-BMIE blend has low ΔHcure (J g−1) for the thermal polymerization and the whole temperature window for the exothermic curing reaction is shifted to lower temperature compared to BPEBPA. Borchardt and Daniels method is used to study the cure kinetics of the materials. The thermal curing of BMIE requires activation energy of 156.0 kJ mol−1 whereas BPEBPA needs slightly higher activation energy (177.2 kJ mol−1). From the TG studies, it can be concluded that the cured BPEBPA exhibits higher thermal stability than the cured BMIE due to the more complex network structure that are formed during thermal polymerization of BPEBPA. Dharwadkar and Kharkhanavala equation is employed to calculate the activation energy needed for the thermal degradation of the thermally cured materials. BPEBPA shows much higher activation energy (65.5 kJ mol−1) for thermal degradation indicating the higher thermal stability over the other two materials (BMIE: 42.5 kJ mol−1 and BPEBPA-BMIE blend: 46.9 kJ mol−1). The isothermal degradation of cured materials is effected in nitrogen atmosphere for constant time interval (10 min). The detailed analysis of the degradation products by GC-MS revealed the formation of phenols and several substituted phenols. This finding hints that the competitive C-C and C-O scissions of the chromene ring units formed via the Claisen rearrangement of the aryl propargyl ether system present in BPEBPA is operative.  相似文献   

14.
The heat capacity of LuPO4 was measured in the temperature range 6.51-318.03 K. Smoothed experimental values of the heat capacity were used to calculate the entropy, enthalpy and Gibbs free energy from 0 to 320 K. Under standard conditions these thermodynamic values are: (298.15 K) = 100.0 ± 0.1 J K−1 mol−1, S0(298.15 K) = 99.74 ± 0.32 J K−1 mol−1, H0(298.15 K) − H0(0) = 16.43 ± 0.02 kJ mol−1, −[G0(298.15 K) − H0(0)]/T = 44.62 ± 0.33 J K−1 mol−1. The standard Gibbs free energy of formation of LuPO4 from elements ΔfG0(298.15 K) = −1835.4 ± 4.2 kJ mol−1 was calculated based on obtained and literature data.  相似文献   

15.
A series of new C2-symmetric 2,2′-bipyridine-contaning crown macrocycles 1-4 has been developed for enantiomeric recognition of amino acid derivatives. These new macrocycles have been showed to be strong complexing agents for primary organic ammonium salts (with K up to 4.83×105 M−1 and −ΔG0 up to 32.4 kJ mol−1) and also useful chromophores for UV-vis titration studies. These macrocyclic hosts exhibited enantioselective binding towards the (S)-enantiomer of phenylglycine methyl ester hydrochloride (Am1) with K(S)/K(R) up to 2.10 (ΔΔG0=−1.84 kJ mol−1) in CH2Cl2 with 0.25% CH3OH. The structure-binding relationship studies showed that the aromatic subunit and the ester group of the ammonium guests are both important for good enantioselectivity. In addition, the host-guest complexes have been studied using various NMR experiments.  相似文献   

16.
Chlorophyll derivative possessing a trifluoroacetyl group at the 3-position was synthesized as a new chemosensor for alcohols and amines. Intense Qy peak of the trifluoroacetyl-chlorin (701 nm in CHCl3) showed blue shifts to 667 nm in MeOH and 665 nm in n-BuNH2 due to the formation of the corresponding hemiacetal and hemiaminal with visible color changes. Thermodynamic parameters for the complexation between trifluoroacetyl-chlorin and n-BuNH2 in CDCl3 were determined to be ΔH = −48 kJ mol−1 and ΔS = −147 J K−1 mol−1. Ratiometric fluorescence sensing of n-BuNH2 in THF was also demonstrated.  相似文献   

17.
The enthalpies of proton ionization of the biochemical buffers N,N-bis[2-hydroxyethyl]-2-aminoethanesulfonic acid (BES) and N-tris[hydroxymethyl]methyl-2-aminoethanesulfonic acid (TES) were obtained in water-methanol mixtures in which the methanol mole fraction (Xm) varied in the range 0-0.36. For both buffers, ionization enthalpy for the first proton (ΔH1) was small in all solvent media. However, upon addition of methanol, ΔH2 increased steadily from 22.2 to a maximum of 27.2 kJ mol−1 for BES, whereas for TES it varied from 30.0 to 32.4, with a minimum of 28.6 kJ mol−1 at Xm=0.123. It is noteworthy that this solvent composition lies within the region of maximum structure enhancement of water by methanol. The results were interpreted in terms of methanol-water interactions.  相似文献   

18.
The standard molar heat capacity C°p,m of adenine(cr) has been measured using adiabatic calorimetry over the range 6<(T/K)<310 and the results used to derive thermodynamic functions for adenine(cr) at smoothed temperatures. At T=298.15 K, C°p,m=(142.67±0.29) J · K−1 · mol−1 and the third law entropy S°m=(145.62±0.29) J · K−1 · mol−1. The standard molar Gibbs free energy of formation ΔfG°m at T=298.15 K for crystalline adenine was calculated, using the standard molar enthalpy of formation for the compound and entropies of the elements from the literature, and found to be ΔfG°m=(301.4±1.0) kJ · mol−1. The results were combined with solution calorimetry and solubility measurements from the literature to yield revised values for the standard molar thermodynamic properties of aqueous adenine at T=298.15 K: ΔfG°m=(313.4±1.0) kJ · mol−1, ΔfH°m=(129.5±1.4) kJ · mol−1, and Sm°=(217.68±0.44) J · K−1 · mol−1.  相似文献   

19.
The vaporization of DyI3(s) was investigated in the temperature range between 833 and 1053 K by the use of Knudsen effusion mass spectrometry. The ions DyI2+, DyI3+, Dy2I4+, Dy2I5+, Dy3I7+, and Dy3I8+ were detected in the mass spectrum of the equilibrium vapor. The gaseous species DyI3, (DyI3)2, and (DyI3)3 were identified and their partial pressures determined. Enthalpies and entropies of sublimation resulted according to the second- and third-law methods. The following sublimation enthalpies at 298 K were determined for the gaseous species given in brackets: 274.8±8.2 kJ mol−1 [DyI3], 356.0±11.3 kJ mol−1 [(DyI3)2], and 436.6±14.6 kJ mol−1 [(DyI3)3]. The enthalpy changes of the dissociation reactions (DyI3)2=2 DyI3 and (DyI3)3=3 DyI3 were obtained as ΔdH°(298)=193.3±5.6 and 390.3±13.0 kJ mol−1, respectively.  相似文献   

20.
Two new quaternary delafossite type oxides with the general formula Ag(Li1/3M2/3)O2, M=Rh, Ir, have been synthesized, and their structures characterized. Based on X-ray and electron diffraction analyses the structural similarity with AgRhO2 delafossite, has been evidenced. The real structures of the quaternary delafossites have been revealed, which has allowed to fully explain the diffuse scattering as observed in X-ray powder diffraction. AgRhO2 is thermally stable up to 1173 K, the behavior of the two quaternary compounds AgLi1/3Rh2/3O2 and AgLi1/3Ir2/3O2 is comparable, and they decompose above 950 and 800 K, respectively. AgRhO2 shows temperature independent paramagnetism, while for the other two an effective magnetic moment of 1.77μB for Ir, and 1.70μB for Rh were determined, applying the Curie-Weiss law. All compounds are semiconducting with activation energies of 4.97 kJ mol−1 (AgLi1/3Rh2/3O2), 11.42 kJ mol−1 (AgLi1/3Ir2/3O2) and 17.58 kJ mol−1 (AgRhO2).  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