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 46 毫秒
1.
CO adsorption microcalorimetry was employed in the study of γ-Al2O3-supported Pt, Pt-Sn and Pt-Fe catalysts. The results indicated that the initial differential heat of CO adsorption of the Pt/γ-Al2O3 catalyst was 125 kJ/mol. As CO coverage increased, the differential heat of adsorption decreased. At higher coverages, the differential heat of adsorption decreased significantly. 60% of the differential heat of CO adsorption on the Pt/γ-N2O3 catalyst was higher than 100 kJ/mol. No significant effect on the initial differential heat was found after adding Sn and Fe to the Pt/γ-Al2O3 catalyst. The amount of strong CO adsorption sites decreased, while the portion of CO adsorption sites with differential heat of 60–110 kJ/mol increased after increasing the Sn or Fe content. This indicates that the surface adsorption energy was changed by adding Sn or Fe to Pt/γ-N2O3. The distribution of differential heat of CO adsorption on the Pt-Sn(C)/γ-Al2O3 catalyst was broad and homogeneous. Comparison of the dehydrogenation performance of C4 alkanes with the number of CO adsorption sites with differential heat of 60–110 kJ/mol showed a good correlation. These results indicate that the surface Pt centers with differential heats of 60–110 kJ/mol for CO adsorption possess superior activity for the dehydrogenation of alkanes. Project supported by FORD and the National Natural Science Foundation of China (Grant No. 09412302) and the Transcentury Training Program Foundation for the Talents by The State Education Commission of China.  相似文献   

2.
    
The collision-free, room temperature gas-phase photodissociation dynamics of CH3CFCl2 (HCFC-141b) was studied using Lyman-α laser radiation (121.6 nm) by the laser photolysis/laser-induced fluorescence ‘pump/probe’ technique. Lyman-α radiation was used both to photodissociate the parent molecule and to detect the nascent H atom products via (2p 2P → 1s 2S) laser-induced fluorescence. Absolute H atom quantum yield, ϕH = (0.39 ± 0.09) was determined by calibration method in which CH4 photolysis at 121.6 nm was used as a reference source of well-defined H atom concentrations. The line shapes of the measured H atom Doppler profiles indicate a Gaussian velocity distribution suggesting the presence of indirect H atom formation pathways in the Lyman-α photodissociation of CH3CFCl2. The average kinetic energy of H atoms calculated from Doppler profiles was found to be E T(lab) = (50 ± 3) kJ/mol. The nearly statistical translational energy together with the observed Maxwell-Boltzmann velocity distribution indicates that for CH3CFCl2 the H atom forming dissociation process comes closer to the statistical limit.  相似文献   

3.
Experimental results obtained by thermal hydroperoxidation of 1,3-diisopropylbenzene in the temperature region of 102–105°C, using a Vodnár-1 type apparatus with discontinuous operation are presented. Based on the results, the activation energy (E=48.5±2 kJ/mol), enthalpy (ΔH+=47±2 kJ/mol), entropy (ΔS+) and the activation free enthalpy (ΔG+) were calculated. XIII. J. Vodnár, A. Chis, A. Biro, S. Békássy and M. Dragan:Studia Univ. Babes-Bolyai, Chemia, Cluj-Napoca, 1997,42, 2.  相似文献   

4.
Sr+ + H2O ↔ SrOH+ + H equilibrium was studied spectrophotometrically. This reaction occurs in natural gas combustion products. Its enthalpy Δr H (0) = 61.4 ± 2.8 kJ/mol and bond energy D 0(Sr+-OH) = 432.6 ± 2.8 kJ/mol were determined using the third law of thermodynamics. The experimental data on this reaction obtained earlier in hydrogen flames, Δr H (0) = 55.3 ± 10.6 and D 0(Sr+-OH) = 438.7 ± 10.6 kJ/mol, were interpreted anew. The D 0(Sr+-OH) = 432.8 ± 2.7 kJ/mol value was eventually obtained.  相似文献   

5.
The relative enthalpies, ΔHo (0) and ΔHo (298.15), of stationary points (four minimum and three transition structures) on the O3H potential energy surface were calculated with the aid of the G3MP2B3 as well as the CCSD(T)–CBS (W1U) procedures from which we earlier found mean absolute deviations (MAD) of 3.9 kJ mol−1 and 2.3 kJ mol−1, respectively, between experimental and calculated standard enthalpies of the formation of a set of 32 free radicals. For CCSD(T)-CBS (W1U) the well depth from O3 + H to trans-O3H, ΔHowell(298.15) = −339.1 kJ mol−1, as well as the reaction enthalpy of the overall reaction O3 + H→O2 + OH, ΔrHo(298.15) = −333.7 kJ mol−1, and the barrier of bond dissociation of trans-O3H → O2 + OH, ΔHo(298.15) = 22.3 kJ mol−1, affirm the stable short-lived intermediate O3H. In addition, for radicals cis-O3H and trans-O3H, the thermodynamic functions heat capacity Cop(T), entropy So (T), and thermal energy content Ho(T) − Ho(0) are tabulated in the range of 100 − 3000 K. The much debated calculated standard enthalpy of the formation of the trans-O3H resulted to be ΔfHo(298.15) = 31.1 kJ mol −1 and 32.9 kJ mol −1, at the G3MP2B3 and CCSD(T)-CBS (W1U) levels of theory, respectively. In addition, MR-ACPF-CBS calculations were applied to consider possible multiconfiguration effects and yield ΔfHo(298.15) = 21.2 kJ mol −1. The discrepancy between calculated values and the experimental value of −4.2 ± 21 kJ mol−1 is still unresolved. Note added in proof: Yu-Ran Luo and J. Alistair Kerr, based on the discussion in reference 12, recently presented an experimental value of ΔfHo(298.15) = 29.7 ± 8.4 kJ mol−1 in the 85th edition of the CRC Handbook of Chemistry and Physics (in progress).  相似文献   

6.
The active molybdenum sulfide compound Mo2S3, which should be considered as a cathode material for thin-layer rechargeable power source, has been produced by electrolysis. Using impedance spectroscopy and potential relaxation method after current interruption, the kinetic parameters of lithium intercalation in electrolytic Mo2S3 have been obtained. Activation energy of Li+ migration in electrolyte (13.76 kJ/mol), charge transfer through the Mo2S3 electrode/electrolyte interface (38.8 kJ/mol), and Li+ diffusion in a solid phase (57.3 kJ/mol) have also been established. Taking into account the coefficient data of charge mass transfer in a solid phase and the reaction rate coefficient of charge transfer through the interface electrode/electrolyte within the temperature range 20–50 °C, the stage of Li+ transfer in a solid phase has been determined as a limiting stage for lithium intercalation in electrolytic molybdenum sulfide Mo2S3.  相似文献   

7.
Relative enthalpies for low-and high-temperature modifications of Na3FeF6 and for the Na3FeF6 melt have been measured by drop calorimetry in the temperature range 723–1318 K. Enthalpy of modification transition at 920 K, δtrans H(Na3FeF6, 920 K) = (19 ± 3) kJ mol−1 and enthalpy of fusion at the temperature of fusion 1255 K, δfusH(Na3FeF6, 1255 K) = (89 ± 3) kJ mol−1 have been determined from the experimental data. Following heat capacities were obtained for the crystalline phases and for the melt, respectively: C p(Na3FeF6, cr, α) = (294 ± 14) J (mol K)−1, for 723 = T/K ≤ 920, C p(Na3FeF6, cr, β) = (300 ± 11) J (mol K)−1 for 920 ≤ T/K = 1233 and C p(Na3FeF6, melt) = (275 ± 22) J (mol K)−1 for 1258 ≤ T/K ≤ 1318. The obtained enthalpies indicate that melting of Na3FeF6 proceeds through a continuous series of temperature dependent equilibrium states, likely associated with the production of a solid solution.   相似文献   

8.
 Di-tert-butyl(N-pyrrolyl)phosphane (1),-sulfide (2) and-selenide (3) were studied by 1H, 13C, 15N, 31P and 77Se NMR spectroscopy in solution. Restricted rotation about the P–N bond was observed for 1 (ΔG C≈58.5 kJ/mol) and 2, 3 (ΔG C≈45 kJ/mol) by 1H and 13C NMR at variable temperature. The fairly high pyrrolyl rotation barrier in 1 is ascribed to the expected steric effects exerted by the tert-butyl groups and to repulsion between the lone pairs of electrons at the phosphorus and nitrogen atoms. These repulsive forces are reduced in 2 or 3, or in di-tert-butyl(phenyl) phosphane (ΔG C≈44 kJ/mol). Signs of coupling constants nJ(31P, 13C) and n+1J(31P, 1H) (n=2, 3) were determined by two-dimensional (2D) 13C/1H heteronuclear shift correlations. The preferred orientation of the pyrrolyl group in 1 is evident from the coupling constants 2J(31P, 13C(2))=+35.4 Hz and 2J(31P, 13C(5))=−9.3 Hz, typical of C(2) in syn and C(5) in anti position with respect to the assumed axis of the phosphorus lone pair. Hahn-echo extended (HEED) polarization transfer pulse sequences served for measuring 1J(31P, 15N) and isotope induced chemical shifts 1Δ15/14N(31P) from 31P NMR spectra. Received: 4 April 1996/Revised: 6 May 1996/Accepted: 11 May 1996  相似文献   

9.
 Di-tert-butyl(N-pyrrolyl)phosphane (1),-sulfide (2) and-selenide (3) were studied by 1H, 13C, 15N, 31P and 77Se NMR spectroscopy in solution. Restricted rotation about the P–N bond was observed for 1 (ΔG C≈58.5 kJ/mol) and 2, 3 (ΔG C≈45 kJ/mol) by 1H and 13C NMR at variable temperature. The fairly high pyrrolyl rotation barrier in 1 is ascribed to the expected steric effects exerted by the tert-butyl groups and to repulsion between the lone pairs of electrons at the phosphorus and nitrogen atoms. These repulsive forces are reduced in 2 or 3, or in di-tert-butyl(phenyl) phosphane (ΔG C≈44 kJ/mol). Signs of coupling constants nJ(31P, 13C) and n+1J(31P, 1H) (n=2, 3) were determined by two-dimensional (2D) 13C/1H heteronuclear shift correlations. The preferred orientation of the pyrrolyl group in 1 is evident from the coupling constants 2J(31P, 13C(2))=+35.4 Hz and 2J(31P, 13C(5))=−9.3 Hz, typical of C(2) in syn and C(5) in anti position with respect to the assumed axis of the phosphorus lone pair. Hahn-echo extended (HEED) polarization transfer pulse sequences served for measuring 1J(31P, 15N) and isotope induced chemical shifts 1Δ15/14N(31P) from 31P NMR spectra. Received: 4 April 1996/Revised: 6 May 1996/Accepted: 11 May 1996  相似文献   

10.
Summary. Four new organic ammonium tetrathiotungstates (NMeenH2)[WS4] (1), (N,N′-dm-1,3-pnH2)[WS4] (2), (1,4-bnH2)[WS4] (3), and (mipaH)2[WS4] (4), (NMeenH2 = N-methylethylenediammonium, N,N′-dm-1,3-pnH2 = N,N′-dimethyl-1,3-propanediammonium, 1,4-bnH2 = 1,4-butanediammonium, and mipaH = monoisopropylammonium) were synthesized by the base promoted cation exchange reaction and characterized by elemental analysis, infrared, Raman, UV-Vis and 1H NMR spectroscopy as well as single crystal X-ray crystallography. The structures of 14 consist of [WS4]2− tetrahedra which are linked to the organic ammonium cations via N–H⋯S hydrogen bonding. The strength and number of the S⋯H interactions affect the W–S bond lengths as evidenced by distinct short and long W–S bonds. The IR spectra exhibit splitting of the W–S vibrations, which can be attributed to the distortion of the [WS4]2− tetrahedron. From a comparative study of several known tetrathiotungstates it is observed that a difference of more than 0.033 ? between the longest and shortest W–S bonds in a tetrathiotungstate will result in the splitting of the asymmetric stretching vibration of the W–S bond.  相似文献   

11.
Nanotube Li-Ti-O compound with high surface (198.6 m2·g−1) was prepared by a method involving the treatment of nanotube Na2Ti2O5·H2O in molten LiNO3 and characterization by means of transmission electron microscopy (TEM), energy-dispersive spectra (EDS), X-ray diffraction (XRD), X-ray photoelectron spectroscopy (XPS), and thermogravimetry-differential thermal analysis (TG/DTG). Results show that the nanotube Li-Ti-O compound prepared by this method involves two crystal phases: spinel Li2Ti2O4 and anatase LixTiO2 (x < 0.1). Li+ exhibits different Li1s binding energy in the two crystal phases. In ambient air, the Li-Ti-O compound adsorbs water easily, and the chemically adsorbed water is difficult to remove below 400°C. Translated from Chinese Journal of Inorganic Chemistry, 2006, 22(12): 2135–2139 [译自: 无机化学学报]  相似文献   

12.
Li2O–Cr2O3–GeO2–P2O5 based glasses were synthesized by a conventional melt-quenching method and successfully converted into glass-ceramics through heat treatment. Experimental results of DTA, XRD, ac impedance techniques and FESEM indicated that Li1.4Cr0.4Ge1.6(PO4)3 glass-ceramics treated at 900 °C for 12 h in the Li1 + xCrxGe2 − x(PO4)3 (x = 0–0.8) system exhibited the best glass stability against crystallization and the highest ambient conductivity value of 6.81 × 10−4 S/cm with an activation energy as low as 26.9 kJ/mol. In addition, the Li1.4Cr0.4Ge1.6(PO4)3 glass-ceramics displayed good chemical stability against lithium metal at room temperature. The good thermal and chemical stability, excellent conducting property, easy preparation and low cost make it promising to be used as solid-state electrolytes for all-solid-state lithium batteries.  相似文献   

13.
Phase solubility techniques were used to obtain the complexation parameters of cisapride (Cisp) with β-cyclodextrin (β-CD) in aqueous 0.05 M citrate buffer solutions. From the UV absorption spectra and the pH solubility profile, two basic pK as were estimated: pK a(1+) = 8.7 and pK a(2+) < 2. The inherent solubility (S o) of Cisp was found to increase as pH decreases, but is limited by the solubility product of the CispH+·citrate1− salt at low pH (pK sp = 3.0). Cisp forms soluble 1:1 and 1:2 Cisp/β-CD complexes. A quantitative measure of the hydrophobic effect (desolvation) contribution to 1:1 complex formation was obtained from the linear variation of free energy of 1:1 Cisp/β-CD complex formation (ΔG 11 = −RT ln K 11 < 0) with that of the inherent solubility of Cisp . The results show that the hydrophobic character of Cisp contributes about 35% of the total driving force to 1:1 complex formation (slope = −0.35), while other factors, including specific interactions, contribute −10.6 kJ/mol (intercept). Protonated 1:1 Cisp/β-CD complex formation at pH 6.0 is driven by favorable enthalpy (ΔH° = −9 kJ/mol) and entropy (ΔS° = 51 J/mol K) changes. In contrast, inherent Cisp solubility is impeded by unfavorable enthalpy (ΔH° = 12 kJ/mol) and entropy (ΔS° = 90 J/mol K) changes. 1H-NMR spectra in D2O and molecular mechanical studies indicate the formation of inclusion complexes. The dominant driving force for neutral Cisp/β-CD complexation in vacuo was predominantly van der Waals with very little electrostatic contribution.  相似文献   

14.
Metal cage complexes [(Me2N)3MO]4 (M = Nb, 3; Ta, 4) have been prepared from the reactions of M(NMe2)5 (M = Nb, 1; Ta, 2) with water. Single crystal X-ray diffraction studies of 3 and 4 reveal that they adopt cubane-like structures with M–O bridges. Variable-temperature NMR studies of –NMeAMeB rotations in 3 and 4 have been performed to give the following activation parameters for the exchanges: ΔH  = −1.4(1.1) kJ/mol, ΔS  = −209(8) J/mol K, \Updelta G 30 8  \textK 1 = 6 4( 2)  \textkJ/\textmol \Updelta G_{{_{{ 30 8\;{\text{K}}}} }}^{{^{ \ne } }} = 6 4\left( 2\right)\;{\text{kJ}}/{\text{mol}} for 3, and ΔH  = −0.9(1.2) kJ/mol, ΔS  = −2.1(0.2) × 102 J/mol K, \Updelta G 30 8  \textK 1 = 6 3( 6)  \textkJ/\textmol \Updelta G_{{ 30 8\;{\text{K}}}}^{{^{ \ne } }} = 6 3\left( 6\right)\;{\text{kJ}}/{\text{mol}} for 4.  相似文献   

15.
Ab initio UMP2 and UQCISD(T) calculations, with 6-311G** basis sets, were performed for the titled reactions. The results show that the reactions have two product channels: NH2+ HNCO?NH3+NCO (1) and NH2+HNCO?N2H3+CO (2), where reaction (1) is a hydrogen abstraction reaction via an H-bonded complex (HBC), lowering the energy by 32.48 kJ/mol relative to reactants. The calculated QCISD(T)//MP2(full) energy barrier is 29.04 kJ/mol, which is in excellent accordance with the experimental value of 29.09 kJ/mol. In the range of reaction temperature 2300–2700 K, transition theory rate constant for reaction (1) is 1.68×1011–3.29×1011 mL·mol-1·s-1, which is close to the experimental one of 5.0×1011mL·mol-1·s-1or less. However, reaction (2) is a stepwise reaction proceeding via two orientation modes,cis andtrans, and the energy barriers for the rate-control step at our best calculations are 92.79 kJ/mol (forcis-mode) and 147.43 kJ/mol (fortrans-mode), respectively, which is much higher than reaction (1). So reaction (1) is the main channel for the titled reaction.  相似文献   

16.
The structure and stability of classical and bridged C2H 3 + is reinvestigated. The SCF and CEPA-PNO computations performed with flexibles andp basis sets including twod-sets on carbon confirm our previous results. We find the protonated acetylene structure to be more stable than the vinyl cation by 3.5–4 kcal/mol. The energy barrier for the interconversion of these two structures is at most a few tenths of a kcal/mol. The equilibrium SCF geometries of Weberet al. [15] are affected insignificantly by further optimization at the CEPA-PNO level. Several structures for the interaction of C2H 3 + with HF have been investigated at the SCF level. With our largest basis set which includes a complete set of polarization functions we find a remarkable levelling of the stabilities of most of the structures. In these cases the stabilization energy ΔE ranges from −10 to −13 kcal/mol.  相似文献   

17.
Summary The geometric isomerization and the dehydrogenation of HP=PH in the ground and some low-lying excited states are investigated by theoretical calculations. The reaction paths are traced by either the CASSCF or UHF-SCF calculations using the 6-31G(d,p) basis functions, and the accompanying energy changes are calculated by the MRD-CI method employing the [5s3p1d]/[2s1p] basis functions. The barrier heights for the trans-to-cis isomerization, by the planar inversion and the nonplanar twisting, in the ground state are calculated to be 265 and 144 kJ/mol (with the vibrational zero-point energy corrections), respectively. The latter barrier is noticeably lower than the H-P and the P-P bond dissociation energies oftrans-HP=PH (1Ag), which are 304 and 271 kJ/mol, respectively. The ground-state HP2 radical (2A'), which is to be formed by the dehydrogenation of HP=PH, should suffer further decomposition into P2 (1 g + ) and H with an activation energy of 139 kJ/mol. The lowest excited state of HP2 is found to be a hydrogen-bridged 3-electron system (2A2) having an isosceles triangle structure. It has proved to be formed by the dehydrogenation of the lowest excited singlet state (1B) of HP=PH via a transition state which lies 194 kJ/mol above the1B state. The excited HP2 (2A2) is state-correlated with P2 (3u)+H.  相似文献   

18.
Gold nanoparticles (nano Au)/titanium dioxide (TiO2) hollow microsphere membranes were prepared on the carbon paste electrode (CPE) for enhancing the sensitivity of DNA hybridization detection. The immobilization of nano Au and TiO2 microsphere was investigated with cyclic voltammetry (CV) and electrochemical impedance spectroscopy (EIS). The hybridization events were monitored with EIS using [Fe(CN)6]3−/4− as indicator. The sequence-specific DNA of the 35S promoter from cauliflower mosaic virus (CaMV35S) gene was detected with this DNA electrochemical sensor. The dynamic detection range was from 1.0×10−12 to 1.0×10−8 mol/L DNA and a detection limit of 2.3×10−13 mol/L could be obtained. The polymerase chain reaction (PCR) amplification of the terminator of nopaline synthase (NOS) gene from the real sample of a kind of transgenic soybean was also satisfactorily detected. Supported by the National Natural Science Foundation of China (Grant Nos. 20635020 and 20375020), Doctoral Foundation of the Ministry of Education of China (Grant No. 20060426001) and Natural Science Foundation of Qingdao City (Grant No. 04-2-JZP-8)  相似文献   

19.
The existence of a short C–H ⋯ π (alkyl–alkyne) interaction in the structure of a strained and relatively rigid tolanophane is expected to hinder the rotation about the C–C sp3 single bond. Variable-temperature NMR experiments (performed in three solvents, CDCl3, THF-d8, and acetone-d6) and ab initio density functional calculations were carried out to investigate its dynamic nature. An energy barrier of 48.6 kJ/mol is determined at coalescence (210 K) with acetone-d6 which is in good agreement with calculation result (54 kJ/mol). Correspondence: Hossein Reza Darabi, Chemistry and Chemical Engineering Research Center of Iran, Pajoohesh Blvd., km 17, Karaj Hwy, 14968-13151 Tehran, Iran.  相似文献   

20.
The heats of the reaction of sodium with ethyl and methyl alcohol were determined by calorimetry. The difference in the standard heats of the formation of triethylarsenite and arsenic trichloride was obtained by calorimetration of the reaction of arsenic trichloride with sodium ethylate, the value of which was −382.42 ± 3.60 kJ/mol. The standard enthalpies of formation were determined from a critical analysis of all data on thermochemistry of trialkylarsenites for the following compounds: triethylarsenite Δf H 298 [(C2H5O)3As(liquid)] = (−704.38 ± 3.85) kJ/mol; trimethylarsenite Δf H 298 [(CH3O)3As(liquid)] = (−599.36 ± 1.88) kJ/mol. The values of standard enthalpies of formation were not adjusted for the following substances in liquid state: arsenic trichloride (−321.96 ± 3.85 kJ/mol), tris-(diethylamido)arsenic(III) As(NEt2)3(liquid) (−129.81 ± 4.41 kJ/mol), tri-n-propylarsenite (−720.61 ± 4.49 kJ/mol), triisopropylarsenite (−756.11 ± 4.65 kJ/mol), tri-n-butylarsenite (−775.11 ± 4.53 kJ/mol), and triisobutylarsenite (−809.71 ± 4.59 kJ/mol). The use of sodium alcoxide solutions for the calorimetration of halogen anhydrides of various acids was demonstrated.  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