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 15 毫秒
1.
《Inorganica chimica acta》1988,141(2):281-288
The crystal structures and 95Mo NMR spectra of two complexes formed between 2-α-hydroxybenzyl- benzimidazole (C6H5·CHOH·C7H5N2=HOBB), as its sodium salt, and MoO2Cl2 are reported. [MoO2- (OBB)2]·EtOH (OBB=deprotonated HOBB) crystallizes in space group P21/n, with a=12.8441(7), b=15.917(3), c=13.314(2) Å, β=97.163(8)° and Z =4. The structure was determined from 3096 observed reflections and refined to a final R value of 0.030. The complex is a six coordinate cis-dioxo species, the 95Mo spectrum of which shows a single sharp peak at 56 ppm in dimethylformamide (DMF). The second complex, [Mo2O5(OBB)2]·EtOH·H2O, crystallizes in space group Pbca, with a=22.482(4), b=16.442(3), c=18.407(3) Å and Z=8. The structure was determined from 2936 observed reflections and refined to a final R value of 0.061. The complex is a binuclear doubly bridged species in which one metal atom is six coordinate while the other is five coordinate. Its 95Mo NMR spectrum in DMF shows a sharp peak at 124 ppm and a second broader much weaker peak at 51 ppm.  相似文献   

2.
The structures of versatile N-heterocyclic carbene-containing Au(III) complexes in the ground and low-lying excited states have been optimized at the B3LYP functional and the single-excitation configuration interaction (CIS) method, respectively. The spectral properties are predicted with time-dependent density functional theory (TDDFT). In addition, the charge transport quality has been estimated approximately by the predicted reorganization energy (λ). As revealed from the calculations, the introduction of methyl has a bigger influence on the spectral properties than phenyl. Furthermore, we find that changing the auxiliary ligand could tune the charge transfer properties.  相似文献   

3.
We report here the results on N-acetyl-l-proline-N',N'-dimethylamide (Ac-Pro-NMe2) as a model for polyproline at the HF/6-31+G(d) level with the conductor-like polarizable continuum model of self-consistent reaction field methods to figure out the conformational preference and cis-trans isomerization of polyproline in the gas phase, chloroform, methanol, and water. The second methyl substitution at the carboxyl amide end results in different backbone structures and their populations from those of N-acetyl-L-proline-N-methylamide (Ac-Pro-NHMe). In particular, all conformations with the C7 hydrogen bond between acetyl and amide ends, which is the most probable conformations of Ac-Pro-NHMe in the gas phase and in nonpolar solvents, disappeared for Ac-Pro-NMe2 even in the gas phase due to the lack of amide hydrogen. The dominant conformation for Ac-Pro-NMe2 is the polyproline II structure with the trans prolyl peptide bond in the gas phase and in solutions. In methanol, the population of the polyproline I structure with the cis prolyl peptide bond is calculated to be larger than that in water, which is consistent with experiments. It should be noted that Ac-Pro-NMe2 has higher rotational barriers for the cis-trans isomerization of the Ac-Pro peptide bond than Ac-Pro-NHMe in the gas phase and in solutions, which could be due to the lack of the intramolecular hydrogen bond between prolyl nitrogen and carboxyl N-H group for the transition state of Ac-Pro-NMe2. The rotational barriers for Ac-Pro-NMe2 are increased with the increase of solvent polarity, as seen for Ac-Pro-NHMe.  相似文献   

4.
Methylidencyclopropabenzene (MCPB) 1 and Fulvalenes 2–4 are molecules of special interest due to the relation between structure and aromaticity. The aim of this work was to analyze this relation and to quantify the aromaticity in 1–4 using different methods. Magnetic properties are directly related with aromaticity; here we studied the magnetic susceptibility and the anisotropy of the magnetic susceptibility. Nucleus indepedent chemical shift (NICS) and the anisotropy of the induced current density (ACID) were also employed. Tools of very different nature, geometric indexes HOMA and Bird, were determinated too for 1–4. All of these measures were found to be in agreement. Figure Both spatial NICS and ACID plot allow to show the aromaticity/antiaromaticity of a ring  相似文献   

5.
The structure, stability and infrared spectral signatures of triflic acid (TA) with water clusters (Wn) and protonated water clusters (TAH+Wn, n?=?1???6) were computed using DFT and MP2 methods. Our calculations show that a minimum of three water molecules are necessary to stabilize the dissociated zwitterionic form of TA. It can be seen from the results that there is no significant movement of protons in smaller (n?=?1 and 2) and linear (n?=?1 – 6) types of water clusters. Further, the geometries of TAWn clusters first form a neutral pair (NP) to contact ion pair (CIP), then form a solvent separated ion pair (SSIP) in a water hexamer. These findings reveal that proton transfer may take place through NP to CIP and then CIP to SSIP. The calculated binding energies (BEs) of ion pair clusters is always higher than that of NP clusters (i.e., more stable than the NP). Existing excess proton linear chain clusters transfer a proton to adjacent water molecules via a Grotthuss mechanism, whereas the same isomers in the branched motifs do not conduct protons. Examination of geometrical parameters and infrared frequencies reveals hydronium ion (H3O+ also called Eigen cation) formation in both TAWn and protonated TAWn clusters. The stability of Eigen water clusters is three times higher than that of other non-Eigen water clusters. Our study shows clearly that formation of ion pairs in TAWn and TAH+Wn clusters greatly favors proton transfer to neighboring water molecules and also enhances the stability of these complexes.  相似文献   

6.
The structural and mechanical properties of methane and carbon dioxide hydrates were investigated using density functional theory simulations. Well-established equations of state of solids and exchange-correlation functionals were used for fitting the unit lattice total energy as a function of volume, and the full second-order elastic constants of these two gas hydrates were determined by energy–strain analyses. The polycrystalline elastic properties were also calculated from the unit lattice results. The final results for methane hydrate agree well with available experimental data and with other theoretical results. The two gas hydrates were found to be highly elastically isotropic, but they differed significantly in shear properties. The presented results for carbon dioxide hydrates are the first complete set reported so far. The results are a significant contribution to the ab initio material characterisation of gas hydrates required for ongoing fundamental studies and technological applications.  相似文献   

7.
Copper(II) complexes supported by bulky tridentate ligands L1H (N,N-bis(2-quinolylmethyl)-2-phenylethylamine) and L1Ph (N,N-bis(2-quinolylmethyl)-2,2-diphenylethylamine) have been prepared and their crystal structures as well as some physicochemical properties have been explored. Each complex exhibits a square pyramidal structure containing a coordinated solvent molecule at an equatorial position and a weakly coordinated counter anion (or water) at an axial position. The copper(II) complexes reacted readily with H2O2 at a low temperature to give mononuclear hydroperoxo copper(II) complexes. Kinetics and DFT studies have suggested that, in the initial stage of the reaction, deprotonated hydrogen peroxide attacks the cupric ion, presumably at the axial position, to give a hydroperoxo copper(II) complex retaining the coordinated solvent molecule (H R ·S). H R ·S then loses the solvent to give a tetragonal copper(II)-hydroperoxo complex (H R ), in which the –OOH group may occupy an equatorial position. The copper(II)–hydroperoxo complex H R exhibits a relatively high O–O bond stretching vibration at 900 cm−1 compared to other previously reported examples.Electronic Supplementary Material Supplementary material is available for this article at  相似文献   

8.
9.
The molecular structure (bond distances and angles), conformational properties, dipole moment and vibrational spectroscopic data (vibrational frequencies, IR and Raman intensities) of phenyl benzoate were calculated using Hartree–Fock (HF), density functional (DFT), and second order Møller–Plesset perturbation theory (MP2) with basis sets ranging from 6-31G* to 6-311++G**. The theoretical results are discussed mainly in terms of comparisons with available experimental data. For geometric data, good agreement between theory and experiment is obtained for the MP2, B3LYP and B3PW91 levels with basis sets including diffuse functions. The B3LYP/6-31+G* theory level estimates the shape of the experimental functions for phenyl torsion around the Ph–O and Ph–C bonds well, but reproduces the height of the rotational barriers poorly. The B3LYP/6-31+G* harmonic force constants were scaled by applying the scaled quantum mechanical force field (SQM) technique. The calculated vibrational spectra were interpreted and band assignments were reported. They are in excellent agreement with experimental IR and Raman spectra.Figure Calculated and experimental (GED) potential energy functions for torsional motion of phenyl benzoate relative to the minimum value. a The potential function for torsion about the O3–C4 bond. b The potential function for torsion about the C2–C10 bond.  相似文献   

10.
In muscle cells, reactive oxygen species (ROS) are continually generated. It is believed that these molecules have a well-established role as physiological modulators of skeletal muscle functions, ranging from development to metabolism and from blood flow to contractile functions. Moreover, ROS may contribute to the development of muscle fatigue, inflammation, and degeneration, and may be implicated in many muscle diseases. The aim of the present study was to verify the role of short or prolonged exposure to oxidative stress, generated by different concentrations of H(2)O(2), on growth, chromosomal aberrations, and apoptosis induced in cultured L6C5 rat muscle cells used as model for myoblasts. Our results indicate that, in L6C5 cells, reactive oxygen intermediates (ROI) can activate distinct cell pathways leading to cell growth induction and development of resistant phenotype, or to chromosomal aberrations, cell cycle arrest, or cell death. The positive vs. negative effects of H(2)O(2)-altered redox potential in myoblasts are strictly related to the intensity of oxidative stress, likely depending on the types and number of cellular targets involved. Among these, DNA molecules appear to be very sensitive to breakage by H(2)O(2), although DNA damage is not directly responsible for ROI-induced apoptosis in L6C5 rat myoblasts.  相似文献   

11.
The ternary complex [Cu(5′-IMP)(dpa)(H2O)]2 has been prepared and its structure analyzed by x-ray diffraction. It has a dimeric structure in which the 5′-IMP ligands coordinate solely through their phosphate groups. This geometry is in marked contrast to that of another Cu5′-IMP ternary complex, [Cu(5′-IMPH)(bipy)(H2O)2]+, which shows metal binding through the purine base rather than the phosphate group.  相似文献   

12.
Full geometric optimization of endo-tricyclo[3.2.1.02,4]oct-6-ene (endo-TCO) by ab initio and DFT methods allowed us to investigate the structure of the molecule. The double bond is endo-pyramidalized and its two faces are no longer found to be equivalent. The exo face of the double bond has regions with far more electron density (qi,HOMO) and more negative electrostatic potential. The endo-TCO-Br2 system was investigated at the B3LYP/6-311+G** level and the endo-TCO···Br2(exo) molecular complex was found to be relatively more stable than the endo-TCO···Br2(endo) complex. The cationic intermediates of the reaction were studied by ab initio and DFT methods. The bridged exo-bromonium cation(I) is relatively more stable than the endo-bromonium cation(II). An absolute exo-facial selectivity should be observed in the addition reaction of Br2 to endo-TCO, which is caused by steric and electronic factors. The nonclassical rearranged cation IV was found to be the most stable ion among the cationic intermediates and the ionic addition occurs via the formation of this cation. The mechanism of the addition reaction is also discussed.  相似文献   

13.
14.
The conformational preference of alpha-D-Manp-(1-->2)-alpha-D-Manp-(1-->O)-L-Ser has been investigated by one-dimensional (1)H,(1)H T-ROESY experiments and molecular-dynamics simulations with CHARMM22 type of force fields and water as explicit solvent. Proton-proton distances were obtained from the simulations and subsequently experimentally determined distances could be derived. Measurements were performed on the title compound as well as on selectively deuterium-substituted analogues synthesized as part of this study to alleviate possible NMR spectroscopic difficulties. A very good agreement was present between the separate NMR experiments. In the subsequent analysis a key nuclear Overhauser effect between the anomeric protons in the two sugar residues was used to assess the conformational dynamics revealed by the molecular simulations. The combined results support a model in which two states are significantly populated as a result of flexibility around the bond defined by the glycosidic torsion angle psi.  相似文献   

15.
A comparative 270 MHz NMR spectroscopic study on the solution structure of the dimer d(TpT) 1, and its two analogues, namely, d(TpST) 2, and NH2d(TcmT) 4 has been reported. Analysis of chemical shifts and coupling constants indicate that: (i) The sugar moieties of the constituent nucleotides are not affected by modification of the internucleotide linkages and adopt preferentially an S-type conformation. (ii) The C4'-C5' bond in the pT part of the modified dimers 2 and 4 shows a large conformational freedom (gamma+ = 32% and 35%, respectively) compared to 1 (gamma+ = 75%). (iii) The population of the trans conformer about C5'-O5' is less important in d(TpST) 2 compared to d(TpT) 1. (iv) The C3'-O3' bond in 2 adopts a trans conformation as in 1. (v) The glycosidic bonds in the modified dimers 2 and 4 showed preferential syn conformation. UV and CD data show that the modified dimers 2 and 4 have poor tendency to stack intramolecularly, they also base pair less efficiently with d(ApA) as compared to d(TpT) 1.  相似文献   

16.
Nitric oxide (NO) has been shown to both enhance hydrogen peroxide (H(2)O(2)) toxicity and protect cells against H(2)O(2) toxicity. In order to resolve this apparent contradiction, we here studied the effects of NO on H(2)O(2) toxicity in cultured liver endothelial cells over a wide range of NO and H(2)O(2) concentrations. NO was generated by spermine NONOate (SpNO, 0.001-1 mM), H(2)O(2) was generated continuously by glucose/glucose oxidase (GOD, 20-300 U/l), or added as a bolus (200 microM). SpNO concentrations between 0.01 and 0.1 mM provided protection against H(2)O(2)-induced cell death. SpNO concentrations >0.1 mM were injurious with low H(2)O(2) concentrations, but protective at high H(2)O(2) concentrations. Protection appeared to be mainly due to inhibition of lipid peroxidation, for which SpNO concentrations as low as 0.01 mM were sufficient. SpNO in high concentration (1 mM) consistently raised H(2)O(2) steady-state levels in line with inhibition of H(2)O(2) degradation. Thus, the overall effect of NO on H(2)O(2) toxicity can be switched within the same cellular model, with protection being predominant at low NO and high H(2)O(2) levels and enhancement being predominant with high NO and low H(2)O(2) levels.  相似文献   

17.
《Inorganica chimica acta》1988,141(2):275-279
The crystal structure of Mo2O5[SC6H4NHCH2C5H4N]2(C3N7NO)3 is reported and seen to consist of a single oxo-bridged species with each Mo atom bonded to cis dioxo groups and the nitrogen atoms and thiolate group of the tridentate ligand. 95Mo NMR spectra of this and three related complexes are presented and attempts made to interpret them in terms of their crystal structures.  相似文献   

18.
《Carbohydrate research》1986,147(2):247-264
l-(1-13C, 5-2H)Arabinose (6D) and l-(2-13C, 5-2H)arabinose (8D) have been synthesized by degradation of 2,3-O-isopropylidene-α-l-rhamnofuranose (2) to l-(4-2H)erythrose (,D), with subsequent chain elongation to 6D plus l-(1-13C, 5-2H)ribose (7D), the latter being converted into 8D. Intermediates were identified by complete assignment of the 13C chemical shifts employing carbon-carbon and carbon-deuterium coupling constants, deuteration shifts, differential isotope-shifts, and deuterium spectra. The anomeric carbon atoms of 2 and 2,3-O-isopropylidene-l-(1-2H) erythrose (4D) gave only single 13C resonances, suggesting that these two compounds exists in only one major anomeric configuration, clarifying previously reported work. The synthesis of 2,3-O-isopropylidene-l-(1-2H)rhmanitol (3D) facilitated the assignment of the signals in the 13C spectra of the nondeuterated analog. Specific deuterium-enrichment and the observed carbon-deuterium coupling (1JC,D ∼22 Hz) not only served to identify the deuterated carbon atom unambiguously in 3 but also permitted assignment of closely spaced resonances. The deuterium spectrum of 2,3-O-isopropylidene-l-(1-2H)erythrofuranose (4D) showed only a single resonance, indicating preponderance of one anomer, in accord with the observation of a single C-1 resonance in the 13C spectrum.  相似文献   

19.
20.
The photochemistry of the [(CpR)Mo(CO)(3)](2) molecules, where CpR = eta(5)-C(5)H(4)(CH(2))(2)C(O)NCH(3)(CH(2))(n)CH(3) (n = 3, 8, 13, and 18), was examined using femtosecond pump-probe transient absorption spectroscopy. The goal of this study was to investigate the importance of radical size and mass on the dynamics and efficiency of geminate radical-radical recombination. The femtosecond results demonstrated the lack of any size/mass dependence of the recombination efficiency. This finding contrasts with results from a prior study that did find a size/mass dependence using a steady-state photochemical technique. To explain these conflicting results, it is proposed that the femtosecond pump-probe results are a measurement of the efficiency of primary geminate recombination whereas the steady-state method results are a measurement of the sum of primary and secondary geminate recombination efficiencies. The size/mass dependence is evident in the latter because secondary geminate recombination is a slower diffusive recombination process and therefore depends on the steric properties of the radicals. Although the existence of primary and secondary recombination channels is often taken for granted, experimental differentiation of primary and secondary caging has proven to be difficult because it is not possible for a single experimental technique to span the entire timescale for recombination of a radical cage pair and adequately resolve these recombination pathways.  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