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 31 毫秒
1.
The gas‐phase nucleophilic substitution reactions at saturated oxygen X? + CH3OY (X, Y = Cl, Br, I) have been investigated at the level of CCSD(T)/6‐311+G(2df,p)//B3LYP/6‐311+G(2df,p). The calculated results indicate that X? preferably attacks oxygen atom of CH3OY via a SN2 pathway. The central barriers and overall barriers are respectively in good agreement with both the predictions of Marcus equation and its modification, respectively. Central barrier heights (ΔH and ΔH) correlate well with the charges (Q) of the leaving groups (Y), Wiberg bond orders (BO) and the elongation of the bonds (O? Y and O? X) in the transition structures. © 2007 Wiley Periodicals, Inc. Int J Quantum Chem, 2007  相似文献   

2.
Vibrational Spectra of the Cluster Compounds (M6X12i) · 8H2O, M = Nb, Ta; Xi = Cl, Br; Xa = Cl, Br, I IR and, for the first time, Raman spectra at 80 K of the cluster compounds (M6X)X · 8H2O; M = Nb, Ta; Xi = Cl, Br; Xa = Cl, Br, I, have been recorded, characterized by typical frequencies of the (M6X) unit, which are only slightly influenced by the terminal Xa ligands. The most intense line with the depolarisation ≈? 0.2 in all Raman spectra is caused by inphase movement of all atoms and assigned to the symmetric metal-metal vibration v1, observed for the clusters (Nb6Cl) at 233–234, for (Nb6Br) at 186–187, for (Ta6Cl) at 199–203, and for (Ta6Br) at 176–179 cm?1. The IR spectra exhibit in the same series intense bands at 233, 204, 207, and 179 cm?1, assigned to the antisymmetric metal-metal vibration. The metal-metal frequencies are significantly higher than discussed before. The tantalum clusters show on excitation with the krypton line 647.1 nm in the region of a d–d transition at 645 nm a resonance Raman effect with series of overtones and combination bands. In case of (Ta6Br) another polarisized band is observed at 229 cm?1 and assigned to the Ta? Bri vibration v2. From the progressions of v1 and v2 anharmonicity constants of about ?3 cm?1 are calculated indicating a strong distortion of the potential curves.  相似文献   

3.
The haloacetyl ions [XCH2CO+], X = Cl, Br, I, have been produced by the dissociative ionization of selected precursor molecules and identified via their fragmentation characteristics. Their heats of formation, Δ, were 708 ± 8, 750 ± 25 and 784 ± 10 kJ mol?1, respectively, all appreciably above ΔH for [CH3CO+], 653 kJ mol?1. The effect of halogen substitution α to carbonyl is discussed for neutral molecules and their molecular ions.  相似文献   

4.
Relativistic and electron correlation effects in thallium halides TlX and TlX3 (X?F, Cl, Br, and I) are investigated by extensive ab initio configuration interaction calculations. Spin–orbit coupling is included at the Hartree–Fock level for the diatomic TlBr and TlI. At the best level of treatment of electron correlation (quadratic configuration interaction), the calculated molecular properties are in good agreement with experimental results, i.e., for the diatomic thallium halides deviations from experimental values are <0.06 Å for bond distances, <0.14 mdyn/Å for force constants, <35 kJ/mol for dissociation energies, and <0.3 D for dipole moments. The convergence of the Møller–Plesset series up to the fourth order is discussed. Two alternative structures of TlI3 are compared. At the Møller–Plesset level of theory, the trigonal planar structure with thallium in the oxidation state + 3 is the preferred gas phase arrangement compared with the bent arrangement containing a linear I unit and thallium in the oxidation state + 1, the difference being ca. 95 kJ/mol. Vibrational frequencies are predicted for all trigonal planar thallium(III) halides. © 1993 John Wiley & Sons, Inc.  相似文献   

5.
Divalent Samarium: AISm X5 (AI = K, Rb; X = Cl, Br, I) Ternary halides with divalent Samarium ASm2X5 were prepared and investigated by X-ray techniques. The paramagnetic susceptibility of Sm2+ has been measured with a Faraday balance and calculated theoretically.  相似文献   

6.
On the chemistry of the elements niobium and tantalum. 84. The niobium and tantalum complexes [Me6X]X · n H2O with Me = Nb, Ta; X1 = Cl, Br; Xa = Cl, Br, J The known and unknown compounds mentioned in the title were prepared. In this group of compounds four different crystal structures (A, B, C, D) occur. Lattice constants are given of the six compounds with structure C which crystallize in the hexagonal system and are isotypic with Ba2[Nb6Cl12]Cl6. Regarding the IR-spectra and the thermal behaviour, possible principles of structure are discussed.  相似文献   

7.
Synthesis, Crystal Structure and Spectroscopic Properties of the Cluster Anions [(Mo6Br )X ]2? with Xa = F, Cl, Br, I The tetrabutylammonium (TBA), tetraphenylphosphonium (TPP) and tetraphenylarsonium (TPAs) salts of the octa-μ3-bromo-hexahalogeno-octahedro-hexamolybdate(2?) anions [(Mo6Br)X]2? (Xa = F, Cl, Br, I) are synthesized from solutions of the free acids H2[(Mo6Br)X] · 8 H2O with Xa = Cl, Br, I. The crystal structures show systematic stretchings in the Mo? Mo bond length and a slight compression of the Bri8 cube in the Fa to Ia series. The cations do not change much. The i.r. and Raman spectra show at 10 K almost constant frequencies of the (Mo6Bri8) cluster vibrations, whereas all modes with Xa ligand contribution are characteristically shifted. The most important bands are assigned by polarization measurements and the force constants are derived from normal coordinate analysis. The 95Mo nmr signals are shifted to lower field with increasing electronegativity of the Xa ligands. The fluorine compound shows a sharp 19F nmr singlet at ?184.5 ppm.  相似文献   

8.
The equilibrium constants of the reactions MBr2(s) + Al2Br6(sln) ? MAl2Br8(sln) M = Cr, Mn, Co, Ni, Zn, Cd have been measured at 298 K in toluene. Ni: 0.017 ± 0.0024, Co: 0.54 ± 0.07, Zn: 1.5 ± 0.2, Mn: 2.1 ± 0, 7, Cr: 2.2 ± 1, Cd: 7 ± 5. They are compared with literature values of the equilibrium constants of analogous reactions in the gas phase MX2(s) + Al2X6(g) ? MAl2X8(g), X = Cl, Br. For CoAl2Br8(sln) the temperature dependence of the equilibrium constant yielded ΔfH = ?9.4 ± 1 kJ mol?1 and ΔfS = ?39.5 ± 3 J mol?1 K?1 while literature values for CoAl2Br8(g) are ΔfH = 42.4 ± 2 kJ mol?1 and ΔfS = 42.9 ± 2 J mol?1 K?1. The solubility of Al2Br6 in toluene as well as its enthalpy of dissolution have been measured in order to evaluate ΔH° and ΔS° of the solvation of Al2Br6(g) and CoAl2Br8(g) in toluene by a thermodynamic cycle. Solvation of Al2Br6(g): ΔH = ?72.7 ± 1 kJ mol?1, ΔS = ?139.6 ± 4 J mol?1 K?1, solvation of CoAl2Br8(g): ΔH = ?124.5 ± 4kJ mol?1, ΔS = ?222 ± 9J mol?1 K?1. Thus, CoAl2Br8 interacts more strongly with the solvent toluene than Al2Br6 does.  相似文献   

9.
Ab initio calculations at the CCSD(T)/6‐311++G(2d,p)//B3LYP/6‐311++G(d,p) level of theory have been carried out for three prototypical rearrangement processes of organosilicon anion systems. The first two are reactions of enolate ions which involve oxygen–silicon bond formation via three‐ and four‐membered states, respectively. The overall reactions are: The ΔG (reaction) values for the two processes are +175 and +51 kJ mol?1, with maximum barriers (to the highest transition state) of +55 and +159 kJ mol?1, respectively. The third studied process is the following: (CH3O)C(?CH2)Si(CH3)2CH → (CH3)2(C2H5)Si? + CH2CO, a process involving an SNi reaction between ‐CH and CH3O‐ followed by silicon–carbon bond cleavage. The reaction is favourable [ΔG(reaction) = ?39 kJ mol?1] with the barrier for the SNi process being 175 kJ mol?1. The previous experimental and the current theoretical data are complementary and in agreement. Copyright © 2009 John Wiley & Sons, Ltd.  相似文献   

10.
Vibrational Spectra and Normal Coordinate Analysis of 92Mo, 100Mo, 35Cl, and 37Cl Isotopomers of the Cluster Anions [(Mo6X )Y ]2?; Xi = Cl, Br; Ya = F, Cl, Br, I The tetrabutylammonium (TBA) salts of the octa-μ3-halogeno-hexahalogeno-octahedro-hexamolybdate(2 –) anions [(Mo6X)Y]2?; Xi = Cl, Br; Ya = F, Cl, Br, I; have been synthesized using 92Mo, 100Mo, 35Cl, and 37Cl. The 10 K IR and Raman spectra reveal significant frequency shifts due to the isotopic labelling of the Mo6 cage, the inner sphere halides X8i or the outer sphere ligands Y, respectively. The normal coordinate analysis yields (Mo? Mo) valence force constants of about 1.3 to 1.5 mdyn/Å. For the μ3-bonded halogenes Cli and Bri valence force constants of 1.1 resp. 1.0 mdyn/Å are calculated. The values for (Mo? Ya) bonds are found in the usual halide range. The observed isotopic shifts are verified very well by the calculations, allowing detailed assignment of the IR and Raman spectra of these compounds for the first time.  相似文献   

11.
The interaction of the palladium(II) complex [Pd(hzpy)(H2O)2]2+, where hzpy is 2‐hydrazinopyridine, with purine nucleoside adenosine 5′‐monophosphate (5′‐AMP) was studied kinetically under pseudo‐first‐order conditions, using stopped‐flow techniques. The reaction was found to take place in two consecutive reaction steps, which are both dependent on the actual 5′‐AMP concentration. The activation parameters for the two reaction steps, i.e. ΔH = 32 ±2 kJ mol?1, ΔS = ?168 ±7 J K?1 mol?1, and ΔH = 28 ± 1 kJ mol?1, ΔS = ?126 ± 5 J K?1 mol?1, respectively, were evaluated and suggested an associative mode of activation for both substitution processes. The stability constants and the associated speciation diagram of the complexes were also determined potentiometrically. The isolated solid complex was characterized by C, H, and N elemental analyses, IR, magnetic, and molar conductance measurements. © 2009 Wiley Periodicals, Inc. Int J Chem Kinet 42: 132–142, 2010  相似文献   

12.
The kinetics of formation and dissociation of [V(H2O)5NCS]2+ have been studied, as a function of excess metal-ion concentration, temperature, and pressure, by the stopped-flow technique. The thermodynamic stability of the complex was also determined spectrophotometrically. The kinetic and equilibrium data were submitted to a combined analysis. The rate constants and activation parameters for the formation (f) and dissociation (r) of the complex are: k/M ?1 · S?1 = 126.4, k/s?1 = 0.82; ΔH /kJ · mol?1 = 49.1, ΔH/kJ · mol?1 = 60.6; ΔS/ J·K?1·mol?1= ?39.8, ΔSJ·K?1·mol?1 = ?43.4; ΔV/cm3·mol?1 = ?9.4, and ΔV/cm3 · mol?1 =?17.9. The equilibrium constant for the formation of the monoisothiocynato complex is K298/M ?1 = 152.9, and the enthalpy and entropy of reaction are ΔH0/kJ · mol?1 = ? 11.4 and ΔS0/J. K?1mol?1 = +3.6. The reaction volume is ΔV0/cm3· mol?1 = +8.5. The activation parameters for the complex-formation step are similar to those for the water exchange on [V(H2O)6]3+ obtained previously by NMR techniques. The activation volumes for the two processes are consistent with an associative interchange, Ia, mechanism.  相似文献   

13.
Photoelectron Spectra and Molecular Properties. 132. Trifluoromethylsulfane and Derivatives F3CSX (X ? CF3, Cl, Br, I) The He(I) photoelectron spectra of trifluoromethylsulfane F3CSH and its derivatives F3CSX (X ? CF3, Cl, Br, I) are assigned by Koopmans' correlations, IE = ?ε, with MNDO eigenvalues, by radical cation state comparison and based on resolved vibrational fine structures, which can' be discussed by MNDO FORCE calculations. The spin/orbit splitting in F3CSI can be approximated by additional ITEREX-85 calculations. Gasphase thermolysis of the trifluoromethylhalogensulfanes F3CSX at 10?4 mbar yields decomposition temperatures, which decrease from X ? Cl to I, and as fragmentation products of presumably radical intermediates, in addition to the respective halogens X2 and F2C?S, also F3CX as well as S2 and CS2 (X ?Cl, Br) are PE spectroscopically detected.  相似文献   

14.
Kinetics of the complex formation of chromium(III) with alanine in aqueous medium has been studied at 45, 50, and 55°C, pH 3.3–4.4, and μ = 1 M (KNO3). Under pseudo first-order conditions the observed rate constant (kobs) was found to follow the rate equation: Values of the rate parameters (kan, k, KIP, and K) were calculated. Activation parameters for anation rate constants, ΔH(kan) = 25 ± 1 kJ mol?1, ΔH(k) = 91 ± 3 kJ mol?1, and ΔS(kan) = ?244 ± 3 JK?1 mol?1, ΔS(k) = ?30 ± 10 JK?1 mol?1 are indicative of an (Ia) mechanism for kan and (Id) mechanism for k routes (‥substrate Cr(H2O) is involved in the k route whereas Cr(H2O)5OH2+ is involved in k′ route). Thermodynamic parameters for ion-pair formation constants are found to be ΔH°(KIP) = 12 ± 1 kJ mol?1, ΔH°(K) = ?13 ± 3 kJ mol?1 and ΔS°(KIP) = 47 ± 2 JK?1 mol?1, and ΔS°(K) = 20 ± 9 JK?1 mol?1.  相似文献   

15.
The reactions indicated in the title have been studied in terms of direct processes and complex formation. Quantum-chemical methods have been applied to the passage of an acid (H+, CH, X+) from CH3X to CH3X, and the abstraction of a radical (H· CH, X·) from CH3X by CH3X. It has been shown that a complex represented by a dimer of a methyl-halide radical cation, (CH3X), with a two-center three-electron bond X? X, has fairly high stability. These investigations were based on non-empirical quantum-chemical calculations, the results being systematically compared with experimental determinations. Some calculations included all electrons (X=F, Cl, Br), others were based on relativistic pseudopotentials (X=F through At). The two sets of calculations agree qualitatively with each other and with experimental observations.  相似文献   

16.
Ab initio molecular orbital (MO) calculations are carried out on the nonidentity allyl transfer processes, X? + CH2CHCH2Y ? CH2CHCH2 X + Y?, with X? = H, F, and Cl and Y = H, NH2, OH, F, PH2, SH, and Cl. The Marcus equation applies well to the allyl transfer reactions. The transition state (TS) position along the reaction coordinate and the TS structure are strongly influenced by the thermodynamic driving force, whereas the TS looseness is originated from the intrinsic barrier. The intrinsic barrier, ΔE, looseness, %L?, and absolute asymmetry, %AS?, are well correlated with the percentage bond elongation, %CY? = [(d ? d)/d] × 100 and/or %CX?. The %CY? and the bond orders indicate that a stronger nucleophile and/or a stronger nucleofuge (or a better leaving group) leads to an earlier TS on the reaction coordinate with a lesser degree of bond making as well as bond breaking. These are consistent with the Bell-Evans-Polanyi principle and the Leffler-Hammond postulate. © 1995 by John Wiley & Sons, Inc.  相似文献   

17.
Geometry, thermodynamic, and electric properties of the π‐EDA complex between hexamethylbenzene (HMB) and tetracyanoethylene (TCNE) are investigated at the MP2/6‐31G* and, partly, DFT‐D/6‐31G* levels. Solvent effects on the properties are evaluated using the PCM model. Fully optimized HMB–TCNE geometry in gas phase is a stacking complex with an interplanar distance 2.87 × 10?10 m and the corresponding BSSE corrected interaction energy is ?51.3 kJ mol?1. As expected, the interplanar distance is much shorter in comparison with HF and DFT results. However the crystal structures of both (HMB)2–TCNE and HMB–TCNE complexes have interplanar distances somewhat larger (3.18 and 3.28 × 10?10 m, respectively) than our MP2 gas phase value. Our estimate of the distance in CCl4 on the basis of PCM solvent effect study is also larger (3.06–3.16 × 10?10 m). The calculated enthalpy, entropy, Gibbs energy, and equilibrium constant of HMB–TCNE complex formation in gas phase are: ΔH0 = ?61.59 kJ mol?1, ΔS = ?143 J mol?1 K?1, ΔG0 = ?18.97 kJ mol?1, and K = 2,100 dm3 mol?1. Experimental data, however, measured in CCl4 are significantly lower: ΔH0 = ?34 kJ mol?1, ΔS = ?70.4 J mol?1 K?1, ΔG0 = ?13.01 kJ mol?1, and K = 190 dm3 mol?1. The differences are caused by solvation effects which stabilize more the isolated components than the complex. The total solvent destabilization of Gibbs energy of the complex relatively to that of components is equal to 5.9 kJ mol?1 which is very close to our PCM value 6.5 kJ mol?1. MP2/6‐31G* dipole moment and polarizabilities are in reasonable agreement with experiment (3.56 D versus 2.8 D for dipole moment). The difference here is due to solvent effect which enlarges interplanar distance and thus decreases dipole moment value. The MP2/6‐31G* study supplemented by DFT‐D parameterization for enthalpy calculation, and by the PCM approach to include solvent effect seems to be proper tools to elucidate the properties of π‐EDA complexes. © 2008 Wiley Periodicals, Inc. Int J Quantum Chem, 2008  相似文献   

18.
Open sheet and framework structures [CuX{cyclo-(MeAsO)4}] (X=Cl, Br, I) 1 – 3 and [Cu3X3{cyclo-(MeAsO)4}2] (X=Cl, Br) 4 and 5 may be prepared by self-assembly from CuX and methylcycloarsoxane (MeAsO)n in acetonitrile solution. 1 – 3 exhibit 44 nets in which (CuX)2 units are connected through μ-1 KAs1 : 2 KAs3 coordinated (MeAsO)4 ligands into large 28-membered rings. In contrast, adjacent [CuX] chains in 4 and 5 are connected into sheets by μ4-K4 As coordinated (MeAsO)4 building blocks, with μ-1 KAs1 : 2 KAs3 bridging of these layers by independent (MeAsO)4 cyclotetramers leading to the generation of a porous framework structure. 1 – 5 were characterised by X-ray structural analysis.  相似文献   

19.
A systematic investigation on the SN2 displacement reactions of nine carbene radical anions toward the substrate CH3Cl has been theoretically carried out using the popular density functional theory functional BHandHLYP level with different basis sets 6‐31+G (d, p)/relativistic effective core potential (RECP), 6‐311++G (d, p)/RECP, and aug‐cc‐pVTZ/RECP. The studied models are CX1X2?? + CH3Cl → X2X1CH3C? + Cl?, with CX1X2?? = CH2??, CHF??, CHCl??, CHBr??, CHI??, CF2??, CCl2??, CBr2??, and CI2??. The main results are proposed as follows: (a) Based on natural bond orbital (NBO), proton affinity (PA), and ionization energy (IE) analysis, reactant CH2?? should be a strongest base among the anion‐containing species (CX1X2??) and so more favorable nucleophile. (b) Regardless of frontside attacking pathway or backside one, the SN2 reaction starts at an identical precomplex whose formation with no barrier. (c) The back‐SN2 pathway is much more preferred than the front‐SN2 one in terms of the energy gaps [ΔE(front)?ΔE(back)], steric demand, NBO population analysis. Thus, the back‐SN2 reaction was discussed in detail. On the one hand, based on the energy barriers (ΔE and ΔE) analysis, we have strongly affirmed that the stabilization of back attacking transition states (b‐TSs) presents increase in the order: b‐TS‐CI2 < b‐TS‐CBr2 < b‐TS‐CCl2 < b‐TS‐CHI < b‐TS‐CHBr < b‐TS‐CHCl < b‐TS‐CF2 < b‐TS‐CHF < b‐TS‐CH2. On the other hand, depended on discussions of the correlations of ΔE with influence factors (PA, IE, bond order, and ΔE), we have explored how and to what extent they affect the reactions. Moreover, we have predicted that the less size of substitution (α‐atom) required for the gas‐phase reaction with α‐nucleophile is related to the α‐effect and estimated that the reaction with the stronger PA nucleophile, holding the lighter substituted atom, corresponds to the greater exothermicity given out from reactants to products. © 2012 Wiley Periodicals, Inc. J Comput Chem, 2012  相似文献   

20.
The enthalpy of formation of tris(methylidene)-cyclopropane (“[3]radialene”, 1 ) has been determined as ΔH = 396 ± 12 kJ mol?1 from three fragmentation reactions of its molecular ion 1 + formed from 1 by photoionisation using synchrotron radiation. Comparative electron impact measurements using conventional mass spectrometry were also performed. A treatment of the latter data is described which leads to satisfactory agreement with the photoionization data. The experimental value of ΔH( 1 ) is compared with various theoretical estimates. The strain energy of 1 is calculated to be 226.3 kJ mol?1. Linear extrapolation of this quantity from the increase of strain in passing from cyclopropane to methylidenecyclopropane yields 282.4 kJ mol?1. The discrepancy between these values, already predicted by Dewar and Baird ten years ago from theoretical calculations, is discussed on the basis of maximum overlap considerations. The enthalpy of formation of bis(methylidene)cyclopropane is predicted to be ΔH= 309 kJ mol?1.  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