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.
The investigations were made on photoinduced electron transfer (ET) from the singlet excited state of rubrene (1RU*) to p-benzoquinone derivatives (duroquinone, 2,5-dimethyl-p-benzoquinone, p-benzoquinone, 2,5-dichloro-p-benzoquinone, and p-chloranil) in benzonitrile (PhCN) by using the steady state and time-resolved spectroscopies. The photoinduced ET produces solvent-separated type charge-separated (CS) species and the charge-recombination (CR) process between RU radical cation and semiquinone radical anions obeys second-order kinetics. Not only the CS species but also the triplet excited state of RU (3RU*) is seen in the transient absorption spectra upon laser excitation of a PhCN solution of RU and p-benzoquinone derivatives. The comparison of their time profiles clearly suggests that the CR process between RU radical cation and semiquinone radical anions to the ground state is independent from the deactivation of 3RU*. This indicates that the CR in a highly exergonic ET occurs at a longer distance with a large solvent reorganization energy, which results in faster ET to the ground state than to the triplet excited state that is lower in energy than the CS state. Photoinduced ET from 3RU* in addition from 1RU* also occurs when p-benzoquinone derivatives with electron-withdrawing substituents were employed as electron acceptors.  相似文献   

2.
The optimized geometries, adiabatic electron affinities, and IR-active vibrational frequencies have been predicted for the long linear carbon chains HC(2n)H. The B3LYP density functional combined with the DZP basis set was used in this theoretical study. The computed physical properties are discussed. The predicted electron affinities form a remarkably regular sequence: 1.78 (HC(12)H), 2.08 (HC(14)H), 2.32 (HC(16)H), 2.53 (HC(18)H), 2.69 (HC(20)H), 2.83 (HC(22)H), and 2.95 eV (HC(24)H). The predicted structures display an alternating triple and very short single bond pattern, with the degree of bond alternation significantly less for the radical anions.  相似文献   

3.
《Chemical physics letters》2001,331(3-4):313-317
The dissociation energies of Fe(CO)n (n=2–4) are computed using correlation consistent basis sets and the CCSD(T) approach. The dissociation energies are extrapolated to the CBS limit and are corrected for core–valence (CV), scalar relativistic, spin–orbit, zero-point, and thermal effects. Our iron carbonyl bond strengths agree with experiment within the respective error bars. We use our dissociations energies at 298 K to obtain the heats of formation of Fe(CO)n (n=1–4).  相似文献   

4.
In recent experimental studies of infrared (IR) spectra of hydrated phenol clusters, the Mikami and Ebata group at Tohoku University attributed apparently two very different spectra to PhOH(H2O)4. The first spectrum has a wide transparent region between 3500 and 3700 cm-1, which they named a window region. The second spectrum has several absorption bands in this window region. Our previous study revealed that the first spectrum was assigned to the isomers which have a single-ring structure of a hydrogen-bonding network of OH's of phenol and waters. The experimentalists suggested that the second spectrum might be identified to a proton-transferred phenol(SINGLEBOND)water cluster. In the present article, the geometrical structures of proton-transferred hydrated phenol clusters were determined with the ab initio molecular orbital method and their IR spectra were calculated. The proton-transferred phenol(SINGLEBOND)water cluster is at a stable local minimum, but the energy is much higher than that of the most stable nontransferred cluster. The calculated IR spectrum has a still wider window region and is far different from the experimental spectrum of the second type. © 1996 John Wiley & Sons, Inc.  相似文献   

5.
The gas-phase acidity of 3,3-dimethylcyclopropene (1) has been measured by bracketing and equilibrium techniques. Consistent with simple hybridization arguments, our value (deltaH degrees (acid) = 382.7 +/- 1.3 kcal mol(-)(1)) is indistinguishable from that for methylacetylene (i.e., deltadeltaH degrees (acid)(1 - CH(3)Ctbd1;CH) = 1.6 +/- 2.5 kcal mol(-)(1)). The electron affinity of 3,3-dimethylcyclopropenyl radical (1r) was also determined (EA = 37.6 +/- 3.5 kcal mol(-)(1)), and these quantities were combined in a thermodynamic cycle to afford the homolytic C-H bond dissociation energy. To our surprise, the latter quantity (107 +/- 4 kcal mol(-)(1)) is the same as that for methane, which cannot be explained in terms of the s-character in the C-H bonds. An orbital explanation (delocalization) is proposed to account for the extra stability of 1r. All of the results are supplemented with G3 and B3LYP computations, and both approaches are in good accord with the experimental values. We also note that for simple hydrocarbons which give localized carbanions upon deprotonation there is an apparent linear correlation between any two of the following three quantities: deltaH degrees (acid), BDE, and EA. This observation could be of considerable value in many diverse areas of chemistry.  相似文献   

6.
The reactions of [Bt3NH][μ-CO)(μ-RS)Fe2(CO)6] (1a-d) (R=nPr,nBu,tBu, Ph) with p-MeC6H4C=CPh or m-NO2C6H4C=CPh were studied and products of the vinyl type (μ-σ,π-p-MeC6H4C=CHPh)(μ-RS)Fe2(CO)6 (2) and (μ-σ,π-phC=CHC6H4Me-p) (μ-RS)Fe2(CO)6 (3) (R=nPr, nBu, tBu, Ph) were obtained. While the structures of all the compounds 2a-d and 3a-d were characterized by elemental analysis, IR,1H NMR and MS spectroscopies, the single-crystal structure of 2c(R=tBu) was determined by X-ray diffraction method. The X-ray dim-action result of 2c showed that the substituted vinyl ligand p-MeC6H4C=CHPh bridges the two uon centers, being σ-bonded to one metal through the olefinic carbon attached to p-MeC6H44 group, while π-bonded to the other via olefinic linkage; the Fe2(CO)6 and proton units are in a cis fashion and the tBu group is bonded to sulfur atom through an e-type of bond.  相似文献   

7.
We test several binning techniques to obtain mode-specific final-state distributions for polyatomic reactions. Normal mode analysis is done after an exact transformation to the Eckart frame. Standard histogram binning (HB) and three different variants of the energy-based Gaussian binning (1GB) are employed to obtain the probabilities of the vibrational states. We consider the two major issues of the polyatomic quasiclassical product analysis, i.e., (1) rounding the classical action to the nearest integer can result in unphysical states and (2) the normal-mode analysis can break down for highly distorted geometries. We show that 1GB can handle issue 1 when the total vibrational energy is evaluated in the normal mode space using the harmonic approximation and both issues 1 and 2 can be solved when the total vibrational energy is calculated exactly in the Cartesian space. We found that anharmonicity in the quantized energy levels does not have a significant effect on the final-state distributions. Quasiclassical trajectory calculations are performed for the reactant ground-state and bending-excited Cl((2)P(3/2)) + CH(4)(v(4/2) = 0, 1) → H + CH(3)Cl reactions using an ab initio potential energy surface. The product analysis techniques are successfully applied to the CH(3)Cl product molecules and some qualitative features of the results are discussed.  相似文献   

8.
A series of molecular rhenium sulfide clusters [Re6S8(OH)6?n (H2O) n ](4?n)? (n = 0, 2, 4, 6) catalyze dehydrogenation of alcohols, and hydrogenation of ketones and olefins in a hydrogen stream at 350 °C. The catalytic activities of the dianionic and neutral clusters (n = 2, 4) are lower than those of tetraanionic and dicationic clusters (n = 0, 6) for all the reactions. When 1,4-butanediol is allowed to react over K4[Re6S8(OH)6], dehydrogenation proceeds to yield 2-hydroxytetrahydrofuran and successively γ-butyrolactone above 300 °C. Over [Re6S8(H2O)6]SO4 dehydration proceeds to yield tetrahydrofuran above 250 °C. The thermal activation mechanisms of these clusters were studied by powder X-ray diffraction analyses, Raman spectrometry, extended X-ray absorption fine structure spectrometry, thermogravimetry, and differential thermal analyses. The catalytically active site of K4[Re6S8(OH)6] is an uncoordinated metal site (Lewis acid site) developed by the loss of a water molecule from two hydroxo ligands. The active site of [Re6S8(H2O)6]SO4 is a Brønsted acid site; the anhydrous aqua cluster dication disproportionates to a hydroxo cluster monocation and a proton. Both of the octahedral cluster frameworks are retained up to 500 °C.  相似文献   

9.
A series of cycloalkylidene-bridged biscyclopentadienyldiiron complexes, C(CH2) n [(5-C5H4)Fe(CO)]2(-CO)2 (n = 4, 5 and 6) have been synthesized by the reacting C(CH2) n Cp2 (Cp = C5H5) with Fe(CO)5 in refluxing xylene. The molecular structures of C(CH2)5[(5-C5H4)Fe(CO)]2(-CO)2 (2) and trans-C(CH2)4[(5-t-BuC5H3)Fe(CO)]2(-CO)2 (4t) have been determined by X-ray diffraction. The Fe—Fe bond distance [2.466 Å], in (2) is the shortest reported to date for bridged biscyclopentadienyldiiron complexes.  相似文献   

10.
Shen  Xiaoping  Li  Baolong  Zou  Jianzhong  Xu  Zheng  Yu  Yunpeng  Liu  Shixiong 《Transition Metal Chemistry》2002,27(4):372-376
K3[Cr(CN)6] reacts with the mononuclear MnIII complex Mn(salen)ClO4 · 2H2O [salen: N,N-ethylenebis(salicylideneiminato)dianion] to give a bimetallic heptanuclear complex cation salt [Cr{(CN)Mn(salen · H2O)}6][Cr(CN)6]6H2O. In the complex anion, [Cr{(CN)Mn(salen · H2O)}6]3+, six MnIII ions coordinate to a CrIII center via cyano bridges, forming a spherical species with 3 symmetry. A study of magnetic properties shows the presence of antiferromagnetic interaction through the cyanide bridge between CrIII (S = 3/2) and MnIII (S = 4/2) and results in a ground state S = 21/2.  相似文献   

11.
《Chemical physics letters》1987,135(6):525-528
The pressure shifts of the first three bands appearing in the visible spectra of [Pt3(CO)6]2−n (n = 3–5) have been measured in solution over the range 0–10 kbar. Previous electronic calculations performed on the dimer in conjunction with these results afford a possible set of assignments for the first three bands appearing in the visible spectrum for the dimer.  相似文献   

12.
13.
The state of d-elements uranogermanates MII(HGeUO6)2·6H2O in aqueous salt solutions in a wide range of ionic strength, ionic composition, and acidity has been investigated. The pH ranges of the uranogermanates stability have been determined, and products of their transformation have been identified. Solubility, solubility equilibrium constant, and Gibbs energy of formation have been determined for the studied uranogermanates. Diagrams of uranium(VI), germanium(IV), and M(II) state in aqueous solutions and in equilibrium solid phases have been plotted.  相似文献   

14.
使用B3LYP/TZVP//B3LYP/aug-cc-pVTZ方法系统研究了饱和烷烃分子CnH2n+2(n=4-6)的轨道电子动量光谱,比较了同分异构体CnH2n+2(n=4-6)对轨道动量分布的影响.结合二维空间分析方法对电子在坐标空间中的密度分布进行了系统的研究.计算结果表明,最内价壳层电荷分布主要由s电子贡献,第二近邻芯价壳层则主要由p电子贡献,而其余的价壳层则为sp杂化.最内价轨道表现出最大的谱线强度并且远大于其它轨道的谱线强度,而且正烷烃的谱线强度要大于异烷烃等同分异构体的谱线强度,表现出了明显的与甲基移动的个数有关的性质.  相似文献   

15.
《Polyhedron》2001,20(9-10):1107-1113
The reactions of dipropargyl manolate and terephthalate, respectively, with Co2(CO)8 in THF at room temperature gave four new compounds [R(CO2CH2C2H-μ)2][Co2(CO)6]2 (R=CH2, 1a; R=C6H4, 1b) and [(HC2CH2OCO)R(CO2CH2C2H-μ)][Co2(CO)6] (R=CH2, 2a; R=C6H4-1,4-, 2b), and compounds 2a and b reacted with RuCo2(CO)11 to form two new linked clusters [R(CO2CH2C2H-μ)2][Co2(CO)6][RuCo2(CO)9] (R=CH2, 3a; R=C6H4-1,4-, 3b). The treatment of two dipropargyl esters, respectively, with RuCo2(CO)11 afforded another two new clusters [R(CO2CH2C2H-μ)2][RuCo2(CO)9]2 (R=CH2, 4a; R=C6H4-1,4-, 4b). The reactions of dipropargyl manolate, terephalate with Mo2Cp2(CO)4 gave rise to the formation of dinuclear complexes [(HC2CH2OCO)R(CO2CH2C2H-μ)][Mo2Cp2(CO)4] (R=CH2, 5a; R=C6H4-1,4-, 5b), compound 5a reacted with Co2(CO)8 to produce the cluster [CH2(CO2CH2C2H-μ)2][Co2(CO)6][Mo2Cp2(CO)4] 6a. All the new clusters have been characterized by C/H elemental analysis, IR and 1H NMR spectroscopies. The structure of [CH2(CO2CH2C2H-μ)2][Co2(CO)6]2 1a and [p-(HC2CH2OCO)C6H4(CO2CH2C2H-μ)][Co2(CO)6] 2b have been determined by single crystal X-ray diffraction methods.  相似文献   

16.
Summary Quantum chemical calculations based on density functional theory have been performed on Cr(CO)6, (6-C6H6)Cr(CO)3 and (6-C6H6)Cr(CO)2(CS) at the local and nonlocal level of theory using different functionals. Good agreement is obtained with experiment for both optimized geometries and metal-ligand binding energies. In particular, a comparison of metal-arene bond energies calculated for the (6-C6H6)Cr(CO)3 and (6-C6H6)Cr(CO)2(CS) complexes correlates well with kinetic data demonstrating that substitution of one CO group by CS leads to an important labilizing effect of this bond, which may be primarily attributed to a larger -backbonding charge transfer to the CS ligand as compared with CO.  相似文献   

17.
《Tetrahedron: Asymmetry》1999,10(14):2665-2674
Reactions of (menthyl)PH2 and H2CCHRf6 (menthyl=1R,3R,4S; Rfn=(CF2)n−1CF3) or H2CCHRf8 (AIBN, refluxing THF) give (menthyl)PH(CH2CH2Rfn) and then (menthyl)P(CH2CH2Rfn)2 (n=6, 7; n=8, 8), but with purification or other difficulties at each stage. Reactions of (menthyl)PCl2 with IMgCH2CH2Rfn give, under careful conditions, analytically pure 7 or 8 in 28–32% yields after distillation. Some Rfn(CH2)4Rfn also form. These represent the first chiral (and non-racemic) fluorous phosphines. Reactions of 7 or 8 with [Ir(COD)Cl]2 and CO give trans-[(menthyl)P(CH2CH2Rfn)2]2Ir(Cl)(CO) (n=6, 71%; 8, 51%) as analytically pure yellow oils. Their IR νCO values show the donor/acceptor properties of 7 and 8 to be intermediate between those of P((CH2)3Rf8)3 and P((CH2)4Rf8)3. The CF3C6F11:toluene partition coefficients of 7 and 8 (27°C, 78.4:21.6 and 93.7:6.3) are distinctly lower than those of P((CH2)2Rfn)3 (n=6, 98.8:1.2; n=8, >99.7:<0.3), reflecting the replacement of a linear C8–C10 group that is ca. 75–80% fluorinated by a cyclic C10 terpenyl group. Reactions of 7 or 8 with [Rh(COD)Cl]2 give [(menthyl)P(CH2CH2Rfn)2]Rh(Cl)(COD) (n=6, 69%; 8, 70%) as orange crystallizable oils.  相似文献   

18.
Complexes of nickel atoms and small clusters with acetylene molecules are studied within the density functional theory. A trend toward the predominant formation of structures with bridge hydrogen atoms is observed in reactions between Ni n and acetylene with rising n.  相似文献   

19.
Low-energy collision induced dissociation has been used to investigate the structure and stability of microsolvated clusters of the prototypical, aprotic multiply charged anion, Pt(CN)(4)(2-), i.e. Pt(CN)(4)(2-)·(H(2)O)(n) n = 1-4, Pt(CN)(4)(2-)·(MeCN)(m) m =1, 2, and Pt(CN)(4)(2-)·(H(2)O)(3)·MeCN. For all of the systems studied, the lowest energy fragmentation pathway was found to correspond to decay of the cluster with loss of the entire solvent ensemble. No sequential solvent evaporation was observed. These observations suggest that the Pt(CN)(4)(2-) solvent clusters studied here form hydrogen-bonded "surface solvated" structures. Electronic structure calculations are presented to support the experimental results. In addition, the detailed fragmentation patterns observed are interpreted with reference to the differential solvation of the ionic fragmentation and electron detachment potential energy surfaces of the core Pt(CN)(4)(2-) dianion. The results described represent some of the first experiments to probe the microsolvation of this important class of multiply charged anions.  相似文献   

20.
We present an extension of a previously published work (J. Solid State Chem. 181 (2008) 3229) concerning Metal-Organic Frameworks (MOFs) of general formula Ni5(OH)6(CnH2n−4O4)2. A modified synthesis procedure comprising a room-temperature step prior to the hydrothermal treatment was employed. This preliminary step made use of peristaltic pumps allowing slow mixing of the reactants at a constant pH value. Samples of better purity and crystallinity were consequently obtained. In particular, the better crystallinity allowed us to work on two other members of the series, n = 10 and n = 12, which were characterized using synchrotron powder X-ray powder diffraction. These two compounds are isoreticular with the n = 6 and n = 8 compounds previously reported. The crystal structure incorporates the long alkane dioic acid molecules as pillars between complex inorganic layers. Samples of better purity for n = 6 and 8, as well as those of the new compounds with n = 10 and 12, gave us the opportunity to revise the magnetic properties of these MOFs. We found similar magnetic behaviors, independently of the interlayer spacing. We show that, below 19 K, these materials most probably enter a spin-glass or cluster spin-glass state rather than a three-dimensionally long-range ordered state. We link this behavior to the complex topology of the magnetic exchange interactions within the inorganic layers which is very likely to be source of magnetic frustration.  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