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.
Dissociation of singly protonated leucine enkephalin (YGGFL) was studied using surface-induced dissociation (SID) in a Fourier transform ion cyclotron resonance mass spectrometer (FT-ICR MS) specially configured for studying ion activation by collisions with surfaces. The energetics and dynamics of seven primary dissociation channels were deduced from modeling the time- and energy-resolved fragmentation efficiency curves for different fragment ions using an RRKM-based approach developed in our laboratory. The following threshold energies and activation entropies were determined in this study: E(0) = 1.20 eV and DeltaS++ = -20 eu(1) (MH(+)-->b(5)); E(0) = 1.14 eV and DeltaS++ = -14.7 eu (MH(+)-->b(4)); E(0) = 1.42 eV and DeltaS++ = -2.5 eu (MH(+)-->b(3)); E(0) = 1.30 eV and DeltaS++ = -4.1 eu (MH(+)-->a(4)); E(0) = 1.37 eV and DeltaS++ = -5.2 eu (MH(+)-->y ions); E(0) = 1.50 eV and DeltaS++ = 1.6 eu (MH(+)-->internal fragments); E(0) = 1.62 eV and DeltaS++ = 5.2 eu (MH(+)-->F). Comparison with Arrhenius activation energies reported in the literature demonstrated for the first time the reversal of the order of activation energies as compared to threshold energies for dissociation.  相似文献   

2.
The three molal dissociation quotients for citric acid were measured potentiometrically with a hydrogen-electrode concentration cell from 5 to 150°C in NaCl solutions at ionic strengths of 0.1, 0.3, 0.6, and 1 molal. The molal dissociation quotients and available literature data at infinite dilution were fitted by empirical equations in the all-anionic form involving an extended Debye-Hückel term and up to five adjustable parameters involving functions of temperature and ionic strength. This treatment yielded the following thermodynamic quantitites for the first dissociation equilibrium at 25°C: logK 1a=−3.127±0.002, ΔH 1a o =4.1±0.2 kJ-mol−1, ΔS 1a o =−46.3±0.7 J-K−1-mol−1, and ΔCp 1a o =−162±7 J-K−1-mol−1; for the second acid dissociation equilibrium at 25°C: logK 2a =−4.759±0.001, ΔH 2a o =2.2±0.1, ΔS 2a o =−83.8±0.4, and ΔCp 2a o =−192±15, and for the third dissociation equilibrium at 25°C: logK 3a=−6.397±0.002, ΔH 3a o =−3.6±0.2, ΔS 3a o =−134.5±0.7, and ΔCp 3a o =−231±7.  相似文献   

3.
The dissociation pathways, kinetics, and energetics of protonated oligosaccharides in the gas phase were investigated using blackbody infrared radiative dissociation (BIRD). Time-resolved BIRD measurements were performed on singly protonated ions of cellohexaose (Cel6), which is composed of β-(1 → 4)-linked glucopyranose rings, and five malto-oligosaccharides (Malx, where x = 4–8), which are composed of α-(1 → 4)-linked glucopyranose units. At the temperatures investigated (85–160 °C), the oligosaccharides dissociate at the glycosidic linkages or by the loss of a water molecule to produce B- or Y-type ions. The Y ions dissociate to smaller Y or B ions, while the B ions yield exclusively smaller B ions. The sequential loss of water molecules from the smallest B ions (B1 and B2) also occurs. Rate constants for dissociation of the protonated oligosaccharides and the corresponding Arrhenius activation parameters (Ea and A) were determined. The Ea and A-factors measured for protonated Malx (x > 4) are indistinguishable within error (~19 kcal mol−1, 1010 s−1), which is consistent with the ions being in the rapid energy exchange limit. In contrast, the Arrhenius parameters for protonated Cel6 (24 kcal mol−1, 1012 s−1) are significantly larger. These results indicate that both the energy and entropy changes associated with the glycosidic bond cleavage are sensitive to the anomeric configuration. Based on the results of this study, it is proposed that formation of B and Y ions occurs through a common dissociation mechanism, with the position of the proton establishing whether a B or Y ion is formed upon glycosidic bond cleavage.  相似文献   

4.
The content of oxygen in Ca0.6 − y Sr0.4La y MnO3 − δ, where y = 0 and 0.05, was determined by coulometric titration over the temperature range 650–950°C at oxygen partial pressure in the gas phase varied from 10−4 to 1 atm. The results were used to calculate the partial molar enthalpy, Δ$ \bar H $ \bar H O(δ), and entropy, Δ$ \bar S $ \bar S O(δ), of oxygen in manganites. Changes in the Δ$ \bar H $ \bar H O(δ) and Δ$ \bar S $ \bar S O(δ) dependences caused by the introduction of lanthanum are evidence of the formation of local clusters of the double perovskite type in the Ca0.6Sr0.4MnO3 − δ matrix.  相似文献   

5.
In an effort to spectroscopically determine the structures of solvated ions composed of nucleic acid bases and amino acids, methods for their gas-phase synthesis have been studied. Ions were electrosprayed and solvated in the accumulation cell of a hybrid Q-FTICR filled with methanol or water vapor at ∼10−2 bar. There were subsequently transferred to the FTICR cell at 10−10 mbar. Following their isolation in the FTICR, they can be investigated by studying their unimolecular blackbody infrared radiative dissociation (BIRD) or infrared multiple photon dissociation (IRMPD) spectroscopy. The IRMPD spectra for (Ade)2Li+ and (Ade)2Li(H2O)+ are reported and compared as well as BIRD rate constants for multiply solvated and metalated adenine ions.  相似文献   

6.
Various types of isokinetic (isoparametric) relationships in heterolytic reactions were summarized and critically analyzed. It was presumed that the series of substrate reactivity is reversed after passing the isoparametric point, and the bimolecular reaction mechanism changes to unimolecular: SN2-SN1, SN2-E1, SE2-SE1, SE2-SN1, and SN2(SSIP)-SN2(C+). Three particular cases of isoparametric relationships are discussed: (1) isoentropy (ΔS = const) which reflects formation of contact ion pair; (2) isoenthalpy (ΔH = const) which reflects formation of space-separated ion pair; and (3) isoenergy (ΔG = const), when ΔH = ΔG = ΔE r. The rate of heterolysis in cyclohexane does not depend on the substrate nature, and a universal minimal rate of heterolysis exists, k25 ≈ 10−10 s−1, τ1/2 = 220 years. There is no nucleophilic assistance by the solvent in unimolecular heterolysis.  相似文献   

7.
Acidobasic properties of purine and pyrimidine bases (adenine, cytosine) and relevant nucleosides (adenosine, cytidine) were studied by means of glass-electrode potentiometry and the respective dissociation constants were determined under given experimental conditions (I = 0.1 M (NaCl), t = (25.0 ± 0.1) °C): adenine (pK HL = 9.65 ± 0.04, pK H2L = 4.18 ± 0.04), adenosine (pK H2L = 3.59 ± 0.05), cytosine (pK H2L = 4.56 ± 0.01), cytidine (pK H2L = 4.16 ± 0.02). In addition, thermodynamic parameters for bases: adenine (ΔH 0 = (−17 ± 4) kJ mol−1, ΔS 0 = (23 ± 13) J K−1 mol−1), cytosine (ΔH 0 = (−22 ± 1) kJ mol−1, ΔS 0 = (13 ± 5) J K−1 mol−1) were calculated. Acidobasic behavior of oligonucleotides (5′CAC-CAC-CAC3′ = (CAC)3, 5′AAA-CCC-CCC3′ = A3C6, 5′CCC-AAA-CCC3′ = C3A3C3) was studied under the same experimental conditions by molecular absorption spectroscopy. pH-dependent spectral datasets were analyzed by means of advanced chemometric techniques (EFA, MCR-ALS) and the presence of hemiprotonated species concerning (C+-C) a non-canonical pair (i-motif) in titled oligonucleotides was proposed in order to explain experimental data obtained according to literature.  相似文献   

8.
Summary Kinetic and mechanistic studies on the anation of cis-[Ru(tap)2(H2O)2]2+ (where tap = 2-(m-tolylazo)pyridine) by pyridine-2-aldoxime (LH) have been made spectrophotometrically at different temperatures (35–50° C) in aqueous medium and various EtOH-H2O mixtures. The following rate law has been established at pH 5.6: k obs = k 1 k 2[LH]/(k −1 + k 2[LH]) where k 1 is the H2O dissociation rate constant of the substrate complex and k -1 and k 2 are the aquation and ligand capturing rate constants of the pentacoordinate intermediate, [Ru(tap)2(H2O)]2+. The rate constants are independent of ionic strength. The reaction rate increases with increasing pH. Activation parameters (ΔH‡ and ΔS‡) have been calculated for media of four different dielectric constants and compared with other substitution reactions in aqueous medium. A dissociative mechanism is proposed.  相似文献   

9.
Titanium dioxide nanoparticles were employed for the sorption of Ge(IV) ions from aqueous solution. The process was studied in detail by varying the sorption time, pH, and temperature. The sorption process was found to be fast, equilibrium was reached within 3 min. A maximum sorption could be achieved from solution when the pH ranges between 4.0 and 11.0. Sorbed Ge(IV) ions can be completely desorbed with 2 mL of 0.3 mol L−1 K3PO4-1.0 mol L−1 H2SO4 mixture solution. The kinetic experimental data properly correlate with the second-order kinetic model (k 2 = 0.88 g mg−1 min−1 (25°C)), Reichenberg equation and Morris-Weber model. The estimated E a for Ge(IV) adsorption on nano-TiO2 was 19.66 kJ mol−1. The overall rate process appears to be influenced by intra-particle diffusion. The sorption data could be well interpreted with the Langmuir and Dubinin-Radushkevich (D-R) type sorption isotherms. The D-R parameters were calculated to be K = −0.00321 mol2 kJ−2, q m = 0.59 mmol g−1 and E = 12.48 kJ mol−1 at room temperature. Furthermore, the thermodynamic parameters were also determined, and the ΔH 0 and ΔG 0 values indicated a spontaneous exothermic behavior.  相似文献   

10.
CO activation in the [Ru(NH3)5CO]2+ ion has been demonstrated under nucleophilic conditions in pyridine or 2-ethoxyethanol solution at 100 °C. In the presence of Me3NO the observed pseudo-first order rate constants were found to be sensitive only to auxiliary ligand concentration (pyridine or methyl pyridines), but with a tendency towards rate saturation and the same limiting rate at large excess of each entering ligand. A mechanism is proposed in which the rate-limiting step is viewed as an auxiliary ligand-assisted CO2 elimination, preceded by a fast reversible addition of Me3NO. This reaction pathway is also supported by the values determined for ΔH (81 ± 13 kJ mol−1) and ΔS (−114 ± 36 J mol−1 K−1). This revised version was published online in June 2006 with corrections to the Cover Date.  相似文献   

11.
The CEPA-PNO method is used for calculating the energy difference ΔE ST between the3 and the1Δ states of diatomic molecules in electronic π2 configurations. An analysis of the contribution of electron correlation to ΔE ST is performed in terms of physically understandable effects such as direct correlation, dynamic spin polarization, semiinternal and internal excitations. It is shown that these effects are of completely different importance for the molecules treated in this study: For C2 the direct correlation between the two singly occupied π-orbitals is the dominant correlation contribution to ΔE ST; for O2, S2, SO the internal excitation π u 2 → π g 2 is predominant, whereas for NH and PH there is a close competition between the direct correlation and the spin polarization of the underlying σ-orbitals. The basis set dependence of these effects is investigated, in particular for NH. Our final results reproduce experimental values of ΔE ST within 0.05–0.10 eV.  相似文献   

12.
The complexation of terfenadine (Terf) with β-cyclodextrin (β-CD) in solution and solid state has been investigated by phase solubility diagram (PSD), differential scanning calorimetry (DSC), powder X-ray diffractometry (PXD) and proton nuclear magnetic resonance (1H-NMR). The PSD results indicated that the salt saturation with the buffer counter ion (citrate−2, H2PO4−1 and Cl−1 ions) of Terf (pK a = 9.5) and the hydrophobic effect play in tandem to increase the value of the complex formation constant (K11) measured at different conditions of pH, ionic strength, buffer type and buffer concentration. The correlation of the free energy of complex formation (ΔG11) with the free energy of inherent solubility of Terf (ΔGSo) obtained by changing the pH, ionic strength and buffer concentration was used to measure the contribution of the hydrophobic effect (desolvation) to complex formation. The hydrophobic effect was found to constitute 57.8% of the driving force for complex stability, while other factors including specific interactions contribute −13.4 kJ/mol. 1H-NMR spectra of Terf–citrate and Terf–HCl salts gave identical chemical shift displacements (ΔΔ) upon complexation, thus indicating that the counter anions are positioned somewhere outside of the β-CD cavity. DSC, XRPD and 1H-NMR proved the formation of solid Terf/acid/β-CD ternary complexes.  相似文献   

13.
Electrochemistry of hydrofullerene C60H36 was studied by cyclic voltammetry in THF and CH2Cl2 in the −47–14 °C temperature range. Hydrofullerene undergoes reversible one-electron reduction to form a radical anion in THF (E 0=−3.18 V (Fc0/Fc+), Fc=ferrocene) and irreversible one-electron oxidation in CH2Cl2 (E p a =1.22 V (Fc0/Fc+)). The reduction potential was used to estimate electron affinity of hydrofullerene as EA=−0.33 eV. It was suggested that C60H36 is an isomer withT-symmetry in which 12 double bonds form four isolated benzenoid rings located in vertices of an imaginary inscribed tetrahedron on the molecular surface. For hydrofullerene, the “electrochemical gap” is an analog of the energy gap (HOMO−LUMO), equal to (E OxE Red)=4.4 V, and indicates that C60H36 is a sufficiently “hard” molecule with a low reactivity in redox reactions. Translated fromIzvestiya Akademii Nauk. Seriya Khimicheskaya, No. 11, pp. 2083–2087, November, 1999.  相似文献   

14.
The NMR spectra of [2.2]paracyclophane with β- or γ-cyclodextrin in DMF-d7 at room temperature do not show significant complexation, while HPLC of the complexes in mixed H2O:alcohol solvents demonstrate complexation with different stoichiometries. At 243 K in DMF solution the H3 and H5 NMR signals of γ-cyclodextrin (but not β) exhibit complexation-induced chemical shifts denoting complex formation. According to HPLC, at room temperature the [2.2]paracyclophane complex with β-cyclodextrin in 20% H2O:EtOH exhibits 1:2 stoichiometry with K 1 = 1×102 ± 2, K 2 = 9.0×104 ± 2×103 (K = 9×106) while that with γ-cyclodextrin in 50% H2O:MeOH exhibits 1:1 stoichiometry with K 1 = 4×103 ± 150 M−1. Thermodynamic parameters for both complexes have been estimated from the retention time temperature dependence. For the β-cyclodextrin complexation at 25°C ΔG 0 CD is −39.7 kJ mol−1 while ΔH 0 CD and ΔS 0 CD are −88.2 kJ mol−1 and −0.16 kJ mol−1 K−1. For γ-cyclodextrin, the corresponding values are ΔG 0 CD = −20.5 kJ mol−1, ΔH 0 CD = −33.5 kJ mol−1 and ΔS 0 CD = −0.04 kJ mol−1 K−1.   相似文献   

15.
A radiometric study of the kinetics of the displacement reaction between nickel(II) and65Zn-labeled zinc salt of ethylenediaminetetraacetic acid, which was previously used by the authors for the analysis of trace quantities of nickel, has been carried out under varying conditions of temperature, pH etc. The above reaction was confirmed to be first order with respect to both Ni2+ and to*ZnEDTA. The overall reaction rate constant, kf, has been shown to be inversely proportional to the concentration of Zn2+ and directly proportional to the concentration of H+. From the dependence of the rate constant on the concentration of Zn2+ and H+ a three-step mechanism is proposed for the above reaction. The values for Ea, ΔH, ΔG and ΔS for the overall reaction have been computed from the experimental data.  相似文献   

16.
The kinetics of the intra-molecular electron transfer of an adduct of l-ascorbic acid and the [Fe3IIIO(CH3COO)6(H2O)3]+ cation in aqueous acetate buffer was studied spectrophotometrically, over the ranges 2.55 ≤ pH ≤ 3.74, 20.0 ≤ θ ≤ 35.0 °C, at an ionic strength of 0.50 and 1.0 mol dm−3 (NaClO4). The reaction of l-ascorbic acid and the complex cation involves the rapid formation of an adduct species followed by a slower reduction in the iron centres through consecutive one-electron transfer processes. The final product of the reaction is aqueous iron(II) in acetate buffer. The proposed mechanism involves the triaqua and diaqua-hydroxo species of the complex cation, both of which form adducts with l-ascorbic acid. At 25 °C, the equilibrium constant for the adduct formation was found to be 86 ± 15 and 5.8 ± 0.2 dm3 mol−1 for the triaqua and diaqua-hydroxo species, respectively. The kinetic parameters derived from the rate expression have been found to be: k 0 = (1.12 ± 0.02) × 10−2 s−1 for the combined spontaneous decomposition and k 1 = (4.47 ± 0.06) × 10−2 s−1H 1 = 51.0 ± 2.3 kJ mol−1, ΔS 1 = −100 ± 8 J K−1 mol−1), k 2 = (4.79 ± 0.38) × 10−1 s−1H 2 = 76.5 ± 0.8 kJ mol−1, ΔS 2 = 6 ± 3 J K−1 mol−1) for the triaqua and diaqua-hydoxo species, respectively.  相似文献   

17.
The basic kinetic parameters of thermal polymerization of hexafluoropropylene, namely, general rate constants, degree of polymerization, and their temperature and pressure dependences in the range of 230–290 °C and 2–12 kbar (200–1200 MPa) were determined. The activation energy (E act = 132±4 kJ mol−1) and activation volume (ΔV 0 = −27±1 cm3 mol−1) were calculated. The activation energy of thermal initiation of polymerization was estimated. The reaction scheme based on the assumption about a biradical mechanism of polymerization initiation was proposed.  相似文献   

18.
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   

19.
The temperature dependency of the saturated vapor pressure of Ir(acac)3 has been measured by the method of calibrated volume (MCV), the Knudsen method, the flow transpiration method, and the membrane method. The thermodynamic parameters of phase transition of a crystal to gas were calculated using each of these methods, and the following values of ΔH T 0 (kJ mol−1) and ΔS T 0 (J mol−1K−1), respectively, were obtained: MCV: 101.59, 156.70; Knudsen: 130.54, 224.40; Flow transpiration: 129.34, 212.23; Membrane: 95.45, 149.44 Coprocessing of obtaining data (MCV, flow transportation method and Knudsen method) at temperature ranges 110−200°C as also conducted:ΔH T 0 =127.9±2.1 (kJ mol−1 ); ΔS T 0 =215.2±5.0 (J mol−1 K−1 ). This revised version was published online in August 2006 with corrections to the Cover Date.  相似文献   

20.
Nucleophilic substitution of Pd(RaaiR′)Cl2 [(RaaiR′ = 1-alkyl-2-(arylazo)imidazole, p-R-C6H4-N=N-C3H2NN-1-R′; where R = H(a)/ Me(b)/ Cl(c) and R′ = Et(1)/Bz(2)] with 2-Mercaptopyridine (2-SH-Py) in acetonitrile (MeCN) at 298 K, to form [Pd2(2-S-Py)4], has been studied spectrophotometrically under pseudo-first-order conditions and the analyses support the nucleophilic association path. The reaction follows the rate law, Rate = {k 0 + k [2-SH-Py] 0 2 }[Pd(RaaiR′)Cl2]: first order in Pd(RaaiR′)Cl2 and second order in 2-SH-Py. The rate of the reaction follows the order: Pd(RaaiEt)Cl2 (1) < Pd(RaaiBz)Cl2 (2) and Pd(MeaaiR′)Cl2 (b) < Pd(HaaiR′)Cl2 (a) < Pd(ClaaiR′)Cl2 (c). External addition of Cl (LiCl) and HCl suppresses the rate (Rate ∝ 1/[Cl]0 & ∝1/[HCl]0). The reactions have been studied at different temperatures (293–308 K) and activation parameters (Δ H° and Δ S°) of the reactions were calculated from the Eyring plot and support the proposed mechanism.  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