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 62 毫秒
1.
Raman spectra of Fe3+ and Pd2+ octaethylporphyrin (OEP) and their α′, β′, γ′, and δ′ deutero derivatives were measured with the 5145, 4880 and 4765 Å lines of an Ar ion laser. Raman bands due to methine-bridge stretching vibrations were assigned and their vibrational amplitudes were calculated from the observed frequency shifts on deuterium substitution of methine-bridge hydrogens. These vibrations correspond to the spin-state sensitive Raman bands of heme proteins. On the basis of symmetry considerations and the observed polarizations, vibrational assignments of other Raman bands were made.  相似文献   

2.
BaVSe3 has been synthesized and its crystal structure determined at 293(2)°K. The structure was solved in the hexagonal space group P63/mmc (D46h), with a = 6.9990(11) and c = 5.8621(13) Å. Scans (2 Θ) of a polycrystalline sample revealed that BaVSe3 undergoes a transition to an orthorhombic unit cell (b′ 31/2 a, aa, cc) at 303(5)°K. Magnetic susceptibility measurements between 4 and 300°K indicate that BaVSe3 is paramagnetic down to 41(1)°K, where magnetic ordering occurs, with a magnetic moment in the ordered phase of 0.2 μB per vanadium atom. The orthorhombic lattice distortion may be caused by the “freezing in” of “soft” vibrational modes.  相似文献   

3.
The electrochemical reduction of the black dye photosensitizer [(H3-tctpy)RuII(NCS)3] (H3-tctpy=2,2′:6′,2′′-terpyridine-4,4′,4′′-tricarboxylic acid) used in photovoltaic cells has been found to be a complex process when studied in dimethylformamide. At low temperatures, fast scan rates and at a glassy carbon electrode, the chemically reversible ligand based one-electron reduction process [(H3-tctpy)Ru(NCS)3]+e[(H3-tctpy√)Ru(NCS)3]2− is detected. This process has a reversible half-wave potential (Er1/2) of −1585±20 mV versus Fc/Fc+ at 25°C. Under other conditions, a deprotonation reaction occurs upon reduction, which produces [(H3−x-tctpyx)Ru(NCS)3](1+x)− and hydrogen gas. Mechanistic pathways giving rise to the final products are discussed. The Er1/2-value for the ligand based reductions of the deprotonated complex is 0.70 V more negative than for [(H3-tctpy)Ru(NCS)3]. Consequently, data obtained from molecular orbital calculations are consistent with the reaction [(H3-tctpy)Ru(NCS)3]+e→[(H2-tctpy)Ru(NCS)3]2−+1/2H2 yielding the monodeprotonated complex as the major product obtained after electrochemical reduction of [(H3-tctpy)Ru(NCS)3]. The Er1/2-values for the metal based RuII/III process differ by 0.30 V when data obtained for the protonated and deprotonated forms of the black dye are compared. Electronic spectra obtained during the course of experiments in an optically transparent thin layer electrolysis configuration are consistent with the overall reaction scheme proposed on the basis of voltammetric measurements and molecular orbital calculations. Reduction studies on the free ligand, H3-tcpy, are consistent with results obtained with [(H3-tctpy)Ru(NCS)3].  相似文献   

4.
Oxidatively generated damage to DNA frequently appears in the human genome as an effect of aerobic metabolism or as the result of exposure to exogenous oxidizing agents. Due to these facts, it has been decided to present the structural propriety and charge distribution of 5′,8-cyclo-2′-deoxyadenosine/guanosine (cdA, cdG) in their 5′R and 5′S diastereomeric forms. For all points of quantum mechanics studies presented, the density functional theory (DFT) with B3LYP parameters on 6-311++G** basis set level was used. The 2-deoxyribose moiety of cyclopurines has adopted the 0T1 conformation in their cationic, neutral and anionic forms. The natural population analysis (NPA) of charge distribution between purine/2-deoxyribose moieties exhibited positive/positive value for cations, positive/negative for neutral molecules. NPA data for anionic forms showed negative/negative values in gas (exclude (5′S)cdG) and positive/negative in water. The dipole moments of 5′,8-cyclopurine-2′-deoxynucleosides were found as follows: 7.83(5′R)cdG, 6.86(5′S)cdG, 3.99(5′R)cdA, 1.99(5′S)cdA in the gaseous phase, 11.29(5′R)cdG, 9.99(5′S)cdG, 6.44(5′R)cdA, 4.14(5′S)cdA in the aqueous phase.  相似文献   

5.
2′-Hydroxy-4-methoxy-5′-methyl chalkone oxime (HMMCO) is used for the extraction and spectrophotometric determination of microgram quantities of vanadium. HMMCO forms a green colored complex with vanadium at 9 M HCl extracted into chloroform. The complex absorbs maximum at 630 nm. Beer's law holds good over the concentration range of 2.7 to 94.0 μg of vanadium per ml. The Ringbom plot shows the effective working range of 4.5 to 67.0 μg of vanadium per ml. The extracted species shows the composition to be 1:2 (V:HMMCO). Vanadium can be extracted quantitatively without any serious interference of the foreign ions. It is possible to extract and determine vanadium quantitatively from binary mixtures that contain niobium, tantalum, or titanium. The method is also applicable to some ores.  相似文献   

6.
We search for narrow resonances in cross-sections for electron collisions with HCl, DCl, HBr and DBr molecules, calculated with the nonlocal resonance model. Narrow resonances corresponding to long-lived metastable states of anionic molecules are indeed found in both elastic and vibrational excitation cross-sections. The largest lifetime τ = 0.11 ms is predicted for HBr with rotational quantum number J = 20. For HCl we find maximum τ = 0.6 ns for J = 22 and τ = 0.6 μs for J = 33 in DCl. A surprising isotope effect is found for DBr, where the largest lifetime is τ = 1.5μs, i.e., much smaller than for HBr.  相似文献   

7.
8.
Triplet energy level-dependent decay pathways of excitons populated on iridium (Ir) complexes within π-conjugated polymeric matrices were studied by means of photoluminescence (PL) and photoconduction action spectroscopy. We chose a set of matrices, poly(9-vinylcarbazole) (PVK), poly[9,9-bis(2-ethylhexyl)fluorene-2,7-diyl] (PF2/6), poly [2-(5′-cyano-5′-methyl-hexyloxy)-1,4-phenylene] (CNPPP), and poly [2-(5′-cyano-5′-methyl-hexyloxy)-1,4-phenylene-co-pridine] (CNPPP-py10 and CNPPP-Py20), having triplet energy levels ranging from 2.2 up to 3.0 eV. As Ir-complex dopants, we selected three phosphorescent emitters, iridium(III)bis(2-(2′-benzothienyl) pyridinato-N-acetylacetonate) (Ir(btp)2acac), iridium(III)fac-tris(2-phenylpyridine) (Ir(ppy)3), and iridium(III)bis[(4,6-fluorophenyl)-pyridinato-N,C2′]picolinate (FIrpic), having triplet energy levels of 2.1, 2.5, and 2.7 eV, respectively. It was found that the triplet emission from the dopants, being populated via energy transfer from the matrices, was strongly dependent on the matching of triplet energy levels between matrix polymers and Ir-complexes. Photocurrent action spectra confirm effective exciton confinement at the dopants sites in the case of PVK matrix systems.  相似文献   

9.
The intracellular second messenger deprotonated adenosine 3′,5′-cyclic monophosphate anion (cAMP-H), generated as gaseous species by electrospray ionization (ESI) and stored in a Paul ion-trap mass spectrometer, has been investigated by mass-resolved infrared multiple photon dissociation (IRMPD) spectroscopy in the 900–1800 cm−1 fingerprint wavenumber range, exploiting the powerful and continuously tunable radiation from a free electron laser (FEL) at the Centre Laser Infrarouge d’Orsay (CLIO). The IRMPD features are interpreted by comparison with the IR spectra obtained by quantum chemical calculations for different low-lying conformers, allowing an assignment for the observed IRMPD bands. It is to be noted that the calculated IR spectra for the most stable conformers look all rather similar and do not allow an unambiguous structural assignment, based exclusively on the IRMPD spectrum. However, the positions and intensities of the IRMPD features of isolated (cAMP-H) ions are consistent with a species deprotonated at the phosphate group and compatible with the main equilibrium structures lying within 18 kJ mol−1 from the lowest lying conformation, the anti-chair form with a C3′-endo sugar twist.  相似文献   

10.
α-Carbohydrated pyridinyl sulfones, prepared from commercially available d-(−)-ribose, d-(+)-galactose, and d-(+)-glucose through a five-step sequence, have been employed in the Julia–Lythgoe–Kocienski olefination with aldehydes. This one-pot protocol, using solid KOH at room temperature, affords the corresponding glycosidic enol ethers in moderate to excellent yields and (E)-stereoselectivities. These glycosylated adducts undergo hetero-Diels–Alder reactions with 2-formyl-1-malondialdehyde to afford 2′,5′-dideoxygenated disaccharides in good yields and complete regio- and endo-selectivity. Alternatively, the [2+2]-cycloaddition reaction of the glycosidic enol ethers with chlorosulfonyl isocyanate provided glycosylated β-lactams regioselectively and with only trans-stereoselectivity. The β-lactams could be converted to N-methylthio derivatives which show decent antibacterial activity toward methicillin-resistant strains of Staphylococcus aureus.  相似文献   

11.
A polarographic study has been carried out on the colored product formed in the interaction of iron(III) with 2-hydroxymethyl-6-(2′-hydroxymethyl-5′-hydroxy-4′-pyrone-6′)-pyranyl-3,2[b]pyran-4,8-dione. Measurements using the differential pulse method show that iron(III) exhibits a coordination number of 2 and a reproducible formation constant that varies somewhat depending upon the supporting electrolyte employed.  相似文献   

12.
The new Cs3Mo6Br13O oxybromide, synthesized by solid-state chemistry, crystallizes in the trigonal system (Rc space group; a = 15.5784(2) Å, c = 19.5103(5) Å, V = 4100.5(1) Å3 and Z = 6). It is based on a [Mo6L14] unit that contains an unprecedented μ3 face-capping oxygen. The crystal structure determined by single crystal X-ray diffraction is built up from discrete face-capped [Mo6Bri6Li2Bra6]3– (L = 0.5 O + 0.5 Br) anionic units in which two inner positions are randomly occupied by one bromine and one oxygen whereas the other ligand positions are fully occupied by bromine. The cesium cations randomly occupy two close crystallographic positions generated by the A-B-C-A-B-C′ close-packed stacking of the units. The cesium site occupancy is related to the random distribution of oxygen and bromine on the Li inner positions. To cite this article: K. Kirakci et al., C. R. Chimie 8 (2005).  相似文献   

13.
We have measured Young's modulus (using a vibrating reed technique) and the specific heat (using ac calorimetry) on the same crystals of α′-NaV2O5 at its Tc=34 K magnetic phase transition. Both properties exhibit large, unsymmetrical, and sample-dependent anomalies. While the specific heat results suggest tricritical behavior of the transition, large fluctuation effects are observed in the modulus above Tc. Fits of the modulus in terms of the specific heat, entropy, and free energy suggest that fluctuations are strongly stress- and sample-dependent.  相似文献   

14.
The ground- and excited-state structures for a series of Os(II) diimine complexes [Os(NN)(CO)2I2] (NN = 2,2′-bipyridine (bpy) (1), 4,4′-di-tert-butyl-2,2′-bipyridine (dbubpy) (2), and 4,4′-dichlorine-2,2′-bipyridine (dclbpy) (3)) were optimized by the MP2 and CIS methods, respectively. The spectroscopic properties in dichloromethane solution were predicted at the time-dependent density functional theory (TD-DFT, B3LYP) level associated with the PCM solvent effect model. It was shown that the lowest-energy absorptions at 488, 469 and 539 nm for 13, respectively, were attributed to the admixture of the [dxy (Os) → π*(bpy)] (metal-to-ligand charge transfer, MLCT) and [p(I) → π*(bpy)] (interligand charge transfer, LLCT) transitions; their lowest-energy phosphorescent emissions at 610, 537 and 687 nm also have the 3MLCT/3LLCT transition characters. These results agree well with the experimental reports. The present investigation revealed that the variation of the substituents from H → t-Bu → Cl on the bipyridine ligand changes the emission energies by altering the energy level of HOMO and LUMO but does not change the transition natures.  相似文献   

15.
We have studied the effect of vibrational mode activation in the CF3 radical on the bromine abstraction reaction; CF3 + Br2 → CF3Br + Br. Excess vibrational energy resides in the symmetric modes of the radical after 248 nm photolysis of the parent molecule, CF3I. Our data indicate that the hot radicals react no faster than thermalized CF3, and may actually have a lower cross-section for reaction. Dynamical factors that result in poor coupling of the vibrational energy to the reaction coordinate, as well as other similar considerations, could be responsible for the experimental observations. In addition, we have made an independent determination of the rate for the bromine abstraction reaction of (1.08 ± .13) × 1012 s?1 cm3 mol?1.  相似文献   

16.
The dynamics of the IH + Br heavy-light + heavy system is treated using the Born-Oppenheimer-type separation between the light and the heavy nuclear motions. The hydrogenic wavefunctions, that describe the H-nucleus motion for clamped I and Br nuclei, provide the potential energy curves and couplings that govern the I-Br relative motion. The Hatom exchange cross sections in the reactive collision: IH(v= 0,j) + Br → I + HBr(v′,j′) are found to strongly depend upon j and to be large only for v′ = 1 with j′ = 16–20 and for v′ = 2 with j′ = 8–11. The present results are compared to our previous estimates, to results of classical trajectory calculations and to available experimental data.  相似文献   

17.
The Raman scattering spectrum of 2,2′-cyanine on colloidal silver metal particles is discussed. Preliminary assignments of some of the vibrational Raman bands to the motions of specific chromophoric units are presented and multiplet character of some bands is discussed. Enhanced Raman scattering of 2,2′-cyanine occurs when the laser radiation is tuned to the J-aggregate absorption feature at 575 nm. The enhancement in Raman intensity is the result of a diminution of fluorescence intensity, as well as a quantitative increase in Raman scattering intensity, and is distinct from other types of enhancement phenomena (e.g., resonance Raman of monomeric solution dye, and surface-enhanced Raman scattering (SERS)). The resonance Raman enhancement, due to excitation at the frequency corresponding to the J-aggregate absorption, is found to be 2 × 10+3.  相似文献   

18.
A quasiclassical trajectory analysis of the reaction of O(3P) with HCl at a collision energy of 3.2 kcal/mol is presented. The potential surface used is based on ab initio calculations at the MRCl level. The MRCl energies with the Davidson correction (MRCI + Q) were scaled in a manner similar to the scaling external correlation (SEC) method of Brown and Truhlar and fitted to a simple analytic function. This surface has a reaction barrier of approximately 10.06 kcal/mol. The classical dynamics of the reaction are studied by initiating 20000 trajectories in each of the three initial states HCl (ν = 2, J = 1, 6, 9). The results are compared to the experiments of Zare and coworkers.  相似文献   

19.
The resonance character of Cu/Ag/Au bonding is investigated in B???M?X (M=Cu, Ag, Au; X=F, Cl, Br, CH3, CF3; B=CO, H2O, H2S, C2H2, C2H4) complexes. The natural bond orbital/natural resonance theory results strongly support the general resonance‐type three‐center/four‐electron (3c/4e) picture of Cu/Ag/Au bonding, B:M?X?B+?M:X?, which mainly arises from hyperconjugation interactions. On the basis of such resonance‐type bonding mechanisms, the ligand effects in the more strongly bound OC???M?X series are analyzed, and distinct competition between CO and the axial ligand X is observed. This competitive bonding picture directly explains why CO in OC???Au?CF3 can be readily replaced by a number of other ligands. Additionally, conservation of the bond order indicates that the idealized relationship bB???M+bMX=1 should be suitably generalized for intermolecular bonding, especially if there is additional partial multiple bonding at one end of the 3c/4e hyperbonded triad.  相似文献   

20.
By analyzing the stationary absorption spectrum of 2,2′:5′,2″-terthiophene in dichloromethane at room temperature, the zero–zero S1←S0 transition energy is determined. Time-resolved fluorescence measurements show that the S1→S0 radiative relaxation in various solvents decays exponentially with a common time constant of 170±10 ps. The fluorescence anisotropy of 2,2′:5′,2″-terthiophene in dichloromethane is also studied and the transient dichroism is interpreted in terms of the revolution of a hydrodynamic prolate ellipsoid with the emission dipole aligned parallel to the long molecular axis.  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