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.
Four of the most stable conformers of 2-amino-pyridine betaine (1-carboxymethyl-2-amino-pyridinium inner salt) monohydrates, 2-NH2PB·H2O, and one anhydrous were analyzed by the B3LYP/6-31G(d,p) calculations and compared with the X-ray data. Two types of optimized conformers can be distinguished: (a) with NH2 and COO groups and (b) an imino tautomer with NH and COOH groups. A common feature of the optimized molecules are intramolecular hydrogen bonds between the COO and H2N or COOH and HN groups. In the crystal both NH2 and COO groups participate in intermolecular hydrogen bonds. The probable assignments of the anharmonic experimental solid state vibrational frequencies of 2-NH2PB·H2O and 2-ND2PB·D2O (conformer 2) based on the calculated B3LYP/6-31G(d,p) harmonic frequencies have been made. Correlations between experimental chemical shifts for 2-NH2PB, its hydrochloride and 1-carboxyethyl-2-amino-pyridinium inner salt (13C and 1H in D2O) and GIAO/B3LYP/6-31G(d,p) calculated isotropic shielding constants, δexp=a+calc, are reported. Good linear regression between experimental and theoretical results for 13C was obtained. Only in 2-NH2PB the hydrogen at -position is outside the linear correlation.  相似文献   

2.
J. K. G. Watson   《Chemical physics》1995,190(2-3):291-300
Some qualitative effects of anharmonicity on the spectra of H3+ and D3+ between low vibrational levels are described. Using large-basis vibration-rotation calculations with a Morse-based discrete variable representation for the vibrations and a symmetric-top basis for the rotations, new spectra of H3+ and D3+ have been assigned. This procedure was assisted by adjusting eight coefficients for H3+ and six coefficients for D3+ in the Meyer-Botschwina-Burton ab initio potential function, and eventually 621 new and old lines of H3+ to levels up to 3ν2 and 529 new and old lines of D3+ to levels up to 2ν2 have been fitted with standard deviations of 0.118 and 0.059 cm−1, respectively. An attempt is made to compare five different potential energy functions for the H3+ system, two ab initio and three adjusted to fit spectra of H3+ or D3+, by expanding them by the same procedure in the same variables. For extension of the present work to higher vibrational levels, more accurate boundary behaviour at linear configurations will be required, and some aspects of the use of hyperspherical coordinates are discussed.  相似文献   

3.
Saddle point geometries and barrier heights have been calculated for the H abstraction reaction HO2(2A″)+H(2S) → H2(1Σ+g)+O2(3Σg) and the concerted H approach-O removing reaction HO2 (2A″)+H(2S) → H2O(1A1)+O(3P) by using SDCI wavefunctions with a valence double-zeta plus polarization basis set. The saddle points are found to be of Cs symmetry and the barrier heights are respectively 5.3 and 19.8 kcal by including size consistent correction. Moreoever kinetic parameters have been evaluated within the framework of the TST theory. So activation energies and the rate constants are estimated to be respectively 2.3 kcal and 0.4×109 ℓ mol−1 s−1 for the first reaction, 20.0 kcal and 5.4.10−5 ℓ mol−1 s−1 for the second. Comparison of these results with experimental determinations shows that hydrogen abstraction on HO2 is an efficient mechanism for the formation of H2 + O2, while the concerted mechanism envisaged for the formation of H2O + O is highly unlikely.  相似文献   

4.
Cross sections and rate constants for vibrational deexcitation of H2 and D2 colliding with 4He are calculated using the semiclassical coupl  相似文献   

5.
Peter C. Junk  Jonathan W. Steed   《Polyhedron》1999,18(27):4646-3597
[Co(η2-CO3)(NH3)4](NO3)·0.5H2O and [(NH3)3Co(μ-OH)2(μ-CO3)Co(NH3)3][NO3]2·H2O were prepared by prolonged aerial oxidation of a solution of Co(NO3)2·6H2O and ammonium carbonate in aqueous ammonia. The formation of these side products highlights the richness of the chemistry of these systems and the possibility of by products if methods are not strictly adhered to. The X-ray crystal structures of [Co(η2-CO3)(NH3)4][NO3]·0.5H2O and [(NH3)3Co(μ-OH)2(μ-CO3)Co(NH3)3][NO3]2·H2O reveal a monomeric octahedral cobalt center with η2-bound CO32− in the former, while the latter consists of a dimeric array where the two cobalt centers are bridged by two OH and one μ2-CO32− groups with three terminal NH3 ligands for each Co center. In both complexes extensive hydrogen bonding interactions are evident.  相似文献   

6.
On the basis of ab initio MP2/6–31 + + G(2d,2p) calculations, we examined the potential energy surfaces of the water·hydrocarbon complexes H2O·CH4, H2O·C2H2 and H2O·C2H2 to locate all the minimum energy structures and estimate the hydrogen bond energies and vibrational frequencies associated with the C(spn)---H·O and the O---H·C(spn) bonds (n = 1−3). Our calculations show that H2O·C2H2, H2O·C2H4 and H2O·CH4 have two minimum energy structures (i.e., the C---H·O and O---H·C hydrogen bond forms), but H2O·C2H4 has only one when the vibrational motion is taken into account, the O---H·C hydrogen bond form. We have also computed the barrier for the interconversion from one minimum to the other. The fully optimized geometries of H2O·CH4, H2O·C2H4 and H2O·C2H2 as well as the vibrational shifts of the C---H stretching frequencies in their C---H·O hydrogen-bonded forms are in good agreement with the available experimental data. The calculated hydrogen bond energies show that the C(spn---H·O bond strengths decrease in the order C(sp)---H·O>C(sp2)---H·O>C(sp3)---O>C(sp3---H·O, which is also consistent with the available experimental data.  相似文献   

7.
We report here an ab initio investigation of the cluster effect (i.e., the formation of four-member groups of nearly degenerate rotation-vibration energy levels at higher J and Ka values) in the H2Te molecule. The potential energy function has been calculated ab initio at a total of 334 molecular geometries by means of the CCSD (T) method where the (1s-4f) core electrons of the Te atom were described by an effective core potential. The values of the potential energy function obtained cover the region up to around 10 000 cm−1 above the equilibrium energy. On the basis of the ab initio potential, the rotation-vibration energy spectra of H2 130Te and its deuterated isotopomers have been calculated with the MORBID (Morse oscillator rigid bender internal dynamics) Hamiltonian and computer program. In particular, we have calculated the rotational energy manifolds for J40 in the vibrational ground state, the ν2 state, the “first triad” (the ν13/2 ν2 interacting vibrational states), and the “second triad” (the (ν1 + ν2)/(ν2 + ν3)/3 ν2 states) of H2130Te. We have also investigated the cluster formation in the vibrational ground state of H2 130Te by first fitting the rotational data available from experiment with a modified Watson-type effective Hamiltonian and then using the optimized ground state constants to extrapolate the rotational structure to higher J values. Both the ab initio calculation and the prediction with the effective Hamiltonian show that the cluster formation in H2Te is very similar to that in H2Se and H2S, which we have studied previously. However, contrary to semiclassical predictions, we do not determine any significant displacement of the clusters towards lower J values relative to H2Se. Hence the experimental observation of the cluster states in H2Te will be at least as difficult as in H2Se.  相似文献   

8.
The compound [Zn(H2O)4]2[H2As6V15O42(H2O)]·2H2O (1) has been synthesized and characterized by elemental analysis, IR, ESR, magnetic measurement, third-order nonlinear property study and single crystal X-ray diffraction analysis. The compound 1 crystallizes in trigonal space group R3, a=b=12.0601(17) Å, c=33.970(7) Å, γ=120°, V=4278.8(12) Å3, Z=3 and R1(wR2)=0.0512 (0.1171). The crystal structure is constructed from [H2As6V15O42(H2O)]4− anions and [Zn(H2O)4]2+ cations linked through hydrogen bonds into a network. The [H2As6V15O42(H2O)]6− cluster consists of 15 VO5 square pyramids linked by three As2O5 handle-like units.  相似文献   

9.
The H2O2-based epoxidation of bridged cyclic alkenes in a monophasic system containing low concentrations (<2 mM) of [Bu4nN]4[Pr2iNH3]2H[P{Ti(O2)}2W10O38]·H2O (1) (with two η2-peroxotitanium sites in the anion) has been studied in search of the catalytically active species involved. 31P NMR spectra of 1, measured under a variety of conditions, revealed that the active species was not hydroperoxotitanium complex [P{Ti(OOH)}2W10O38]7−or [P{Ti(OOH)}Ti(O2)W10O38]7−. The reaction pathways for the alkene epoxidation are discussed to understand the kinetics (especially the initial [H2O2] dependence). It was concluded that the net catalytic reaction for the epoxidation occurred through the two-electron oxidation at the hydroperoxotitanium site in the catalyst.  相似文献   

10.
Rate constants for the tunneling reaction (HD + D → h + D2) in solid HD increase steeply with increasing temperature above 5 K, while they are almost constant below 4.2 K. The apparent activation energy for the tunneling reaction above 5 K is 95 K, which is consistent with the energy (91–112 K) for vacancy formation in solid hydrogen. The results above 5 K were explained by the model that the tunneling reaction was accelerated by a local motion of hydrogen molecules and hydrogen atoms. The model of the tunneling reaction assisted by the local motion of the reactans and products was applied to the temperature dependence of the proton-transfer tunneling reaction (C6H6 + C2H5OH → C6H7 + C2H5O) in solid ethanol, the tunneling elimination of H2 molecule of H2 molecule ((CH3)2 CHCH(CH3)2+ → (CH3)2 C = C(CH3)2+ + H2) in solid 2,3-dimethylbutane, and the selective tunneling reaction of H atoms in solid neo-C5H12-alkane mixtures.  相似文献   

11.
The epoxidation of cyclohexene with hydrogen peroxide in a biphase medium (H2O/CHCl3) was carried out with the reaction-controlled phase transfer catalyst composed of quaternary ammonium heteropolyoxotungstates [π-C5H5N(CH2)15CH3]3[PW4O16]. A conversion of about 90% and a selectivity of over 90% were obtained for epoxidation of cyclohexene on the catalyst. The fresh catalyst, the catalyst under reaction conditions and the used catalysts were characterized by FT-IR, Raman and 31P NMR spectroscopy. It appears that the insoluble catalyst could degrade into smaller species, [(PO4){WO(O2)2}4]3−, [(PO4){WO(O2)2}2{WO(O2)2(H2O)}]3−, and [(PO3(OH)){WO(O2)2}2]2− after the reaction with hydrogen peroxide and becomes soluble in the CHCl3 solvent. The active oxygen in the [W2O2(O2)4] structure unit of these soluble species reacts with olefins to form the epoxides and consequently the corresponding W---Ob---W (corner-sharing) and W---Oc---W (edge-sharing) bonds are formed. The peroxo group [W2O2(O2)4] can be regenerated when the W---Ob---W and W---Oc---W bonds react with hydrogen peroxide again. These soluble species lose active oxygen and then polymerize into larger compounds with the W---Ob---W and W---Oc---W bonds and then precipitate from the reaction solution after the hydrogen peroxide is consumed up. Part of the used catalyst seems to form more stable compounds with Keggin structure under the reaction conditions.  相似文献   

12.
The epoxidation of cyclopentene with hydrogen peroxide catalyzed by 12-heteropolyacids of molybdenum and tungsten (H3PMo12−nWnO40, n = 1–11), 12-tungstophosphoric acid and 12-molybdophosphoric acid combined with cetylpyridinium bromide as a phase transfer reagent was carried out in acetonitrile. Among 13 heteropolyacids investigated, catalyst of H3PMo6W6O40 showed the highest activity, giving a conversion of 60% and a selectivity of 95% in the epoxidation of cyclopentene. The fresh catalysts and the catalysts under reaction condition were characterized by UV–vis, FT-IR and 31P NMR spectroscopy, which has revealed that all of the molybdotungstophosphoric acids were degraded in the presence of hydrogen peroxide to form a considerable amount of phosphorus-containing species. The active species resulted from H3PMo6W6O40 are new kinds of phosphorus-containing species, which is different from {PO4[WO(O2)2]4}3−.  相似文献   

13.
以H3PW12O40和H4SiW12O40(简写为PW12和SiW12)为效应物, 测定其对酪氨酸酶活力的抑制作用. 通过非变性聚丙烯凝胶(Native-PAGE)电泳确定酪氨酸酶是多家族基因编码, 其分子量为3×104~3.4×104, 4.2×104~4.6×104, 6.4×104~6.8×104, 且均具有活性, 测定PW12和SiW12对酪氨酸酶的抑制效果, 并结合酶动力学法研究其抑制机理. 结果表明, 当PW12和SiW12浓度分别达到13和25 mmol/L时, 酪氨酸酶的活力完全被抑制, 即PW12和SiW12对酪氨酸酶二酚酶具有不同程度的抑制效果. 当所加酶量为0.0173 mg/mL时, PW12和SiW12对酪氨酸酶二酚酶活力的半抑制率 (IC50)分别为1.57和2.31 mmol/L, 它们对酪氨酸酶二酚酶的抑制均为可逆过程. 其中, PW12对二酚酶的抑制类型为混合型, 其KIKIS分别为0.34和0.43 mmol/L, SiW12对二酚酶的抑制类型表现为竞争型, 其KI为0.59 mmol/L. 综合考虑IC50值和抑制常数等参数, PW12对酪氨酸酶二酚酶的抑制能力优于SiW12.  相似文献   

14.
We utilized gas phase hydrogen/deuterium (H/D) exchange reactions and ab initio calculations to investigate the complexation between a model peptide (Arg-Gly-AspRGD) with various alkali metal ions. The peptide conformation is drastically altered upon alkali metal ion complexation. The associated conformational changes depend on both the number and type of complexing alkali metal ions. Sodium has a smaller ionic diameter and prefers a multidentate interaction that involves all three amino acids of the peptide. Conversely, potassium and cesium form different types of complexes with the RGD. The [RGD + 2Cs − H]+ species exhibit the slowest H/D exchange reactivity (reaction rate constant of 6 × 10−13 cm3molecule−1s−1 for the fastest exchanging labile hydrogen with ND3). The reaction rate constant of the protonated RGD is two orders of magnitude faster than that of the [RGD + 2Cs − H]+. Addition of the first cesium to the RGD reduces the H/D exchange reaction rate constant (i.e., D0) by a factor of seven whereas sodium reduces this value by a factor of thirty. Conversely, addition of the second alkali metal ions has the opposite effect; the rate of D0 disappearance for all [RGD + 2Met − H]+ species (MetNa, K, and Cs) decreases with the alkali metal ion size.  相似文献   

15.
Quasiclassical trajectory calculations have been performed to determine the effect of reactant collision energy on product state distributions in the reaction O(1D) + H2 → OH(2Π) + H. The product vibrational distribution becomes more excited as the collision energy is increased. This is not due to an increase in the cross section for collinear abstraction. A detailed analysis has shown that strong O---H2 repulsion, which occurs during the insertion of the O into the H---H bond, converts the kinetic energy of the reacting system to vibrational motion of the intermediate.  相似文献   

16.
Infrared spectroscopy has been coupled with the matrix isolation technique, firstly for a study of the molecular complexes between ozone and water, secondly to investigate the mechanism of the water photooxidation by ozone at 15 K by UV light. The 1:1, 1:2 and 2:1 complexes are isolated and mainly characterized by a shift in the infrared absorption of the H2O (D2O) submolecule. Force field calculations involving anharmonicity corrections allows us to conclude into the non-equivalence of the two OH oscillators within the 1:1 complex. This result suggests the formation of a very weak hydrogen bond of the type HOH...OO2. By photolysis of the water-ozone mixture in solid argon (λ < 310 nm), formation of H2O2 with very small amounts of O2H and OH is observed. This process occurs through the reaction between the O(1D) atom generated by photodissociation of O3 belonging to the one H2O: O3 pair and the water molecule of the same pair, without diffusion of the oxygen atom.  相似文献   

17.
[Eu(ABA)(phen)2(H2O)3](ClO4)3·3phen·4.5H2O (1) and [Eu(Val)(phen)2(H2O)3](ClO4)3·3phen·2H2O (2) are two new europium complexes with amino acids and 1,10-phenanthroline (phen=1,10-phenanthroline, ABA=-amino butyl acid, Val= -valine). Their crystal structures were measured by X-ray crystallography. Europium atoms in both complexes are nine-coordinated with bidentate 1,10-phenanthroline and carboxylate anion of amino acids, and water molecules. In the solid state, 1 and 2 have a structure involving aromatic stacking of the coordinated and non-coordinated 1,10-phenanthroline and the oxygen atoms of non-coordinated perchlorate anions being H-bond acceptors connect [Eu(ABA)(phen)2(H2O)3]3+·3phen·4.5H2O or [Eu(Val)(phen)2(H2O)3]3+·3phen·2H2O in their structures. In their interactions, several C–HO bonds play an important role. Owing to their different amino acid ligands and the number of lattice water molecules, there is some difference in their hydrogen bond patterns in 1 and 2. The side chain of -valine is involved in the formation of C–HO bonds. Hydrogen bond and π–π interactions determine the supramolecular formation of three-dimensional net works of both complexes.  相似文献   

18.
The bimetallic [Pt(NH3)4]2[W(CN)8][NO3]·2H2O is characterised by single-crystal X-ray diffraction [S.G.P21/m(11), a=8.0418(7), b=19.122(2), c=9.0812(6) Å, Z=2]. All platinum centres have the square-plane D4h geometry with average dimensions Pt(1)–N 2.042(2) and Pt(2)–N 2.037(10) Å. The octacyanotungstate anion has the square-antiprismatic D4d configuration with average dimensions W(1)–C 2.164(13), C–N 1.140(12), W(1)–N 3.303(5) Å. The structure exhibits two different mutual orientations of Pt versus W units resulting in Pt(2)–W(1), W(1)* separations of 4.77(2), 4.55(2)* and Pt(1)–W(1) of 6.331(8) Å. A centrosymmetric structure reveals groups of two distinct columns: the first is formed by intercalated NO3 between parallel [Pt(1)(NH3)4]2+ planes and the second consists of [W(CN)8]3− interlayered by, parallel to square faces of W-antiprisms, [Pt(2)(NH3)4]2+. The structure is stabilised through a three-dimensional hydrogen bond network via nitrogen atoms of cyanide ligands, hydrogen atoms of NH3 ligands, water molecules and oxygen atoms of NO3 counteranions. The vibrational pattern and the range of ν(CN) frequencies attributable to the electronic environment of W(V) and W(IV) are consistent with the ground state Pt(II)↔W(V) charge transfer.  相似文献   

19.
The following structural peculiarities of the agostic acyl structure 2R) (R = H, SiMe3) and some characteristic chemical reactivity of the M-η2-acyl and iminoacyl linkage are described. (i) A structural comparison of the bonding parameters within three agostic acetyl Mo complexes containing the dithioacid ligand, indicates that the agostic interaction strengthens upon increasing the electron-releasing properties of the S-chelating ligand. (ii) The acyl-xanthate complex Mo(C(O)Me)(S2COR)(CO)(PMe3)2 undergoes loss of a sulfur atom from the coordinated xanthate and coupling with the acyl ligand to form complexes containing coordinated alkoxythiocarbonyl and monothioacetate ligands. The latter can be metathetically replaced by KS2COR. (iii) Upon heating at 70°C η2-acyl-dicarbonyl bispirazolilborate complexes of molybdenum of the type Mo(H2B(pz*)2)(η2-C(O)Me)(CO)2(PMe3) (pz* = 3,5-dimethyl-pyrazol-1-yl) yield functionalized acyl ligands derived from the stereo- and regioselective intramolecular addition of one of the B---H bonds of the H2B(pz*)2 group across the C=O moiety of the η2-acyl group. (iv) The η2-acyl-isocyanide complexes {Mo}(η2-C(O)R)(CNR′) ({Mo} = Mo(H2B(pz*)2)(CO)(PMe3)) undergo irreversible thermal isomerization to the corresponding η2-iminoacyl-carbonyl derivatives {MO}(η2-C(NR′)R)(CO). This isomerization reaction follows first-order kinetics.  相似文献   

20.
Pulses from a mechanically chopped CO-laser were used to optically pump the first vibrational level of NO molecules in their fundamental band near 5.3 μm. The population of NO (υ = 1) was followed by measuring the resonance fluorescence of NO-γ-bands from a microwave discharge lamp in the UV region. Analysis of the first order decays of NO(υ = 1) following the excitation pulses yielded rate constants for V---T and V---V energy transfer processes in collisions of NO(υ = 1) with ground state NO and added gas molecules He, Ne, Ar, Kr, Xe, H2, HD, D2, N2, O2 and N2O.  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