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.
A quasi-classical trajectory method (QCT) running on the 1A′ and 1A″ potential energy surfaces (PESs) given by Dobbyn and Knowles [A.J. Dobbyn, P.J. Knowles, Mol. Phys. 91 (1997) 1107] has been employed to study the dynamical stereochemistry of the chemical reaction O(1D) + D2 → OD + D, especially the vector correlations between products and reagents. The results indicate that product rotational angular momentum j′ is not only aligned, but also oriented along the direction perpendicular to the scattering plane on both PESs, with different rotational polarization behaviors of product OD for the two PESs and for different collision energies. Calculations show that the alignment effect of products become weaker with an increase of the collision energy on the 1A′ PES but is not sensitive to the collision energy on the 1A″ PES. When the collision energy increases, the product OD mainly tends to the forward scattering on the 1A′ PES and displays a switch from the backward scattering to the forward one on the 1A″ PES. These differences are probably attributed to the different characteristics of the two PESs.  相似文献   

2.
The reaction dynamics of ground state boron atoms, B(2Pj), with acetylene, was reinvestigated and combined with novel electronic structure calculations. Our study suggests that the boron atom adds to the carbon–carbon triple bond of the acetylene molecule to yield initially a cyclic intermediate undergoing two successive hydrogen atom migrations to form ultimately an intermediate i3. The latter was found to decompose predominantly to the c-BC2H(X2A′) isomer plus atomic hydrogen via a tight exit transition state. To a minor amount, an isomerization of i3i4 prior to a hydrogen atom ejection forming the linear structure, HBCC(X1Σ+), has to be taken into account. Since the c-BC2H(X2A′) and HBCC(X1Σ+) isomers are separated by an isomerization barrier to ring closure of only 3 kJ mol−1, internally excited HBCC(X1Σ+) products can isomerize to the c-BC2H(X2A′) structure and vice versa.  相似文献   

3.
The complex [Pt(5,5′-dmbipy)Cl4] (1) (5,5′-dmbipy is 5,5′-dimethyl-2,2′-bipyridine) was prepared from the reaction of H2PtCl6·6H2O with 5,5′-dimethyl-2,2′-bipyridine in methanol. The same method was employed to make [Pt(6-mbipy)Cl4] (2) (6-mbipy is 6-methyl-2,2′-bipyridine). Both complexes were characterized by elemental analysis, IR, UV–Vis, 1H NMR, 13C NMR and 195Pt NMR spectroscopy. Their solid state structures were determined by the X-ray diffraction method.  相似文献   

4.
A DFT computational study is performed on different Cp2TiIV(L,L′-BID) complexes with L,L′-BID = dioxolene, dithiolene or diselenolene. A fragment analysis of the titanocene-ligand bonding in the DFT optimized geometries showed that out of plane folding for maximum Ti ← L π donation increases Cp2TiIV(O,O′-BID) (35°) < Cp2TiIV(S,S′-BID) (43–49°) < Cp2TiIV(Se,Se′-BID) (48–53°).  相似文献   

5.
Three new Cu(II)–Ni(II) heterodinuclear complexes: [Cu(PMoxd)Ni(phen)2](ClO4)2 (1), [Cu(PMoxd)Ni(NO2-phen)2](ClO4)2 (2), [Cu(PEoxd)Ni(Me2-bpy)2](ClO4)2 (3), [where Cu(PMoxd)=N,N′-bis(pyridyl-methyl)oxamidatocopper(II), Cu(PExod)=N,N′-bis(2-pyridyl-ethyl)oxamidatocopper(II), phen=1,10-phenanthroline and NO2-phen=5-nitro-1,10-phenanthroline and bpy=2,2′-bipyridine] were prepared and characterized by i.r. and electronic spectra, and by magnetic properties. The magnetic analysis was carried out by means of the theoretical expression of the magnetic susceptibility deduced from the spin Hamiltonian H=−2JS1S2, leading to J=−70.83 cm−1 (1); −56.23 cm−1 (2); −57.30 cm−1 (3), indicating a weak antiferromagnetic spin–exchange interaction between Cu(II) and Ni(II) ions within three complexes.  相似文献   

6.
[2′,3′,5′,6′-2H4]-2-Hydroxynaringenin is synthesised and incubated with commercially available UDP-glucose and the crude protein extract from Desmoduim uncinatum leaves. The organic extract produces isotopically labelled [2′,3′,5′,6′-2H4]-vitexin and [2′,3′,5′,6′-2H4]-isovitexin. Repeating the experiment with denatured protein or replacing the 2-hydroxynaringenin with [2′,3′,5′,6′-2H4]-apigenin or [2′,3′,5′,6′-2H4]-naringenin results in no observable incorporation. 2-Hydroxynaringenin is therefore the substrate for C-glucosylflavonoid biosynthesis in D. uncinatum.  相似文献   

7.
A series of new manganese(I) and ruthenium(II) monometallic and bimetallic complexes made of 2,2′-bipyridine and 1,10-phenanthroline ligands, [Mn(CO)3(NN)(4,4′-bpy)]+, [{(CO)3(NN)Mn}2(4,4′-bpy)]2+ and [(CO)3(NN)Mn(4,4′-bpy)Ru(NN)2Cl]2+ (NN = 2,2′-bipyridine, 1,10-phenanthroline; 4,4′-bpy = 4,4′-bipyridine) are synthesized and characterized, in addition to already known ruthenium(II) complexes [Ru(NN)2Cl(4,4′-bpy)]+ and [Cl(NN)2Ru(4,4′-bpy)Ru(NN)2Cl]2+. The electrochemical properties show that there is a weak interaction between two metal centers in Mn–Ru heterobimetallic complexes. The photophysical behavior of all the complexes is studied. The Mn(I) monometallic and homobimetallic complexes have no detectable emission. In Mn–Ru heterobimetallic complexes, the attachment of Mn(I) with Ru(II) provides interesting photophysical properties.  相似文献   

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.
Force field calculations have been carried out for the planar and non-planar modes of pyrazine-N,N′-O2 using the observed vibrational frequencies obtained from the IR and Raman spectral studies on pyrazine-N,N′-O2-h4 and pyrazine-N,N′-O2-d4 reported in the literature [D.A. Thornton, P.F.M. Verhoeven, G.M. Watkins, Herman O. Desseyn, Benjamin J. Van der Veken, Spectrochim. Acta 46A (1990) 1439]. The purpose of the present work is to determine force fields for the pyrazine-N,N′-O2 molecule and to present vibrational assignments for the observed IR and Raman frequencies to the fundamental modes, combination bands and overtones. The planar force field determined in the present case is expected to be better than that reported earlier [S. Szöke, G. Varsanyi, E. Baitz, Acta Chim. 53 (1967) 345] because of the inclusion of the observed frequencies due to pyrazine-N,N′-O2-d4 isotopomer. In addition, the non-planar force field for this molecule is reported for the first time.  相似文献   

10.
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   

11.
This paper presents examples of mixed-ligand Co(II), Cu(II), Ni(II) and Mn(II) complexes, with a distorted octahedral coordination geometry, with 2,2′-dipyridyl or 1,10-phenanthroline and phosphortriamide ligands. The complexes of the general type ML2·Lig (where M = Co(II), Cu(II), Ni(II), Mn(II); L = {Cl3C(O)NP(O)R2} (R = NHBz, NHCH2CHCH2, NEt2); Lig = 2,2′-dipyridyl or 1,10-phenanthroline) were synthesised and characterised by means of X-ray diffraction, IR and UV–Vis spectroscopy. The phosphortriamide ligands are coordinated via oxygen atoms of phosphoryl and carbonyl groups involved in six-membered metal cycles. The additional ligands 2,2′-dipyridyl or 1,10-phenanthroline are coordinated to the central atom, forming five-membered cycles.  相似文献   

12.
The reaction of [Pt(RaaiR′)(solvent)2]2+ with nucleobases (NB), adenine and guanine was studied and the products [Pt(RaaiR′)(NB)(H2O)](PF6)2RaaiR′ = 1-alkyl-2-(arylazo)imidazole, R′ = Me (1), Et (2); R = H (a),OMe (b), NO2(c)] characterized by i.r, u.v.–vis. and 1H-n.m.r spectroscopy. The solution spectra exhibit metal-to-ligand charge-transfer transitions (MLCT) and intra-ligand charge-transfer transitions. The position and symmetry of the bands depend on the nucleobase and arylazoimidazole. The coordination of the ligand is supported by 1H-n.m.r. spectral data. The redox property has been examined by cyclic voltammetric technique and shows the involvement of the azo group of the chelated ligand, RaaiR′ in the reduction process along with the irreversible reduction of the coordinated nucleobase. Binding with Bovine Serum Albumin (BSA) and Calf Thymus DNA was performed spectrophotometrically in the physiological buffer medium and spectral profile was recorded in 24 h.  相似文献   

13.
Ag+-assisted dechlorination of blue cis-trans-cis Ru(R-aai-R′)2Cl2 followed by the reaction with chloranilic acid (H2CA) in the presence of Et3N, gives a neutral mononuclear violet complex [Ru(R-aai-R′)2(CA)]. [R-aai-R′=p-R-C6H4—N=N—C3H2—NN, abbreviated as an N,N′ chelator where N(imidazole) and N(azo) represent N and N′, respectively; R = H (a), OMe (b), NO2 (c) and R′= Me (4), Et(5), Bz(6)]. All the complexes exhibit strong intense MLCT transitions in the visible region and weak broad bands at higher wavelength (>700 nm). Visible transitions (580–595 nm) show a negative solvatochromic effect. The cyclic voltammograms show two quasireversible to irreversible couples positive to SCE and are due to CA/CA2− (1.2–1.35 V) and Ru(III)/Ru(II) (1.6–1.8 V) redox processes. Three couples, negative to SCE, are assigned to CA2−/CA3− (−0.2 to −0.3 V), and azo reductions (−0.5 to −0.7, −0.8 to −0.9 V) of the chelated R-aai-R′.  相似文献   

14.
Topologically challenging, the protocovalent NO bond, which belongs to a distinct class called charge-shift bonds, has been identified in the HONO (cis, trans) molecules on the basis of topological analysis of the ELF and ELI-D functions obtained from the B3LYP and CASSCF(12,10) calculations. The presence of the protocovalent NO bond is associated with energetically possible dissociation channel: HONO(1A′) → OH(2π) + NO(2π).  相似文献   

15.
Nucleophilic substitution of Pd(RaaiR′)Cl2 [RaaiR′=1-alkyl-2-(arylazo)imidazole, p-R—C6H4— N=N—C3H2NN-1-R′; where R= H(a)/Me(b)/Cl(c) and R′ = Et(1)/Bz(2)] with adenine (A) in MeCN–water (1:1) at 298 K, to form [Pd(A)2]Cl2, has been studied spectrophotometrically under pseudo-first-order conditions and the analyses support a nucleophilic association path. The reaction follows the rate law, rate = {a+k [A] 02[Pd(RaaiR′)Cl2]: first-order in Pd(RaaiR′)Cl2 and second-order in A. The rate increases as follows: Pd(RaaiEt)Cl2(1) < Pd(RaaiBz)Cl2(2) and Pd(MeaaiR′)Cl2(b) < Pd(HaaiR′)Cl2(a) < Pd(ClaaiR′)Cl2(c). External addition of Cl (LiCl) suppresses the rate (rate 1/[Cl]). The activation parameters, H0 and S0 of the reactions were calculated from the Eyring plot and support the proposed mechanism.  相似文献   

16.
The excited states of the HNO radical have been studied using the equations of motion method. These calculations confirm the presence of a low-lying 3A″ state at 5485 cm−I, which lies between the IA′ ground state and IA″ excited state.  相似文献   

17.
Complexes [Ir(C^N)2(G1-bpy)]PF6, where C^N is a cyclometallating ligand derived from 2-(2′-thienyl)pyridine and 2-phenylpyridine, and G1-bpy is a dendritic bipyridine ligand of the first generation, 4,4′-bis[3″,5″-bis(benzyloxy)phenylethyl]-2,2′-bipyridine, were prepared and characterized by 1H NMR, electronic absorption, and emission spectroscopy. The polyether dendritic substituents exert a “ soft” effect on the spectral and luminescence properties of the complexes, manifested as slight destabilization of the electronically excited charge-transfer state involving the bipyridine ligand, as compared to the model complexes [Ir(C^N)2(bpy)]PF6.__________Translated from Zhurnal Obshchei Khimii, Vol. 75, No. 5, 2005, pp. 705–711.Original Russian Text Copyright © 2005 by Kulikova, McClenaghan, Balashev.  相似文献   

18.
The reaction of the heteroleptic Nd(III) iodide, [Nd(L′)(N″)(μ-I)] with the potassium salts of primary aryl amides [KN(H)Ar′] or [KN(H)Ar*] affords heteroleptic, structurally characterised, low-coordinate neodymium amides [Nd(L′)(N″)(N(H)Ar′)] and [Nd(L′)(N″)(N(H)Ar*)] cleanly (L′ = t-BuNCH2CH2[C{NC(SiMe3)CHNt-Bu}], N″ = N(SiMe3)2, Ar′ = 2,6-Dipp2C6H3, Dipp = 2,6-Pri2C6H3, Ar* = 2,6-(2,4,6-Pri3C6H2)2C6H3). The potassium terphenyl primary amide [KN(H)Ar*] is readily prepared and isolated, and structurally characterised. Treatment of these primary amide-containing compounds with alkali metal alkyl salts results in ligand exchange to give alkali metal primary amides and intractable heteroleptic Nd(III) alkyl compounds of the form [Nd(L′)(N″)(R)] (R = CH2SiMe3, Me). Attempted deprotonation of the Nd-bound primary amide in [Nd(L′)(N″)(N(H)Ar*)] with the less nucleophilic phosphazene superbase ButNP{NP(NMe2)3}3 resulted in indiscriminate deprotonations of peripheral ligand CH groups.  相似文献   

19.
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   

20.
The imidazolium salts 1,1′-dibenzyl-3,3′-propylenediimidazolium dichloride and 1,1′-bis(1-naphthalenemethyl)-3,3′-propylenediimidazolium dichloride have been synthesized and transformed into the corresponding bis(NHC) ligands 1,1′-dibenzyl-3,3′-propylenediimidazol-2-ylidene (L1) and 1,1′-bis(1-naphthalenemethyl)-3,3′-propylenediimidazol-2-ylidene (L2) that have been employed to stabilize the PdII complexes PdCl22-C,C-L1) (2a) and PdCl22-C,C-L2) (2b). Both latter complexes together with their known homologous counterparts PdCl22-C,C-L3) (1a) (L3 = 1,1′-dibenzyl-3,3′-ethylenediimidazol-2-ylidene) and PdCl22-C,C-L4) (1b) (L4 = 1,1′-bis(1-naphthalenemethyl)-3,3′-ethylenediimidazol-2-ylidene) have been straightforwardly converted into the corresponding palladium acetate compounds Pd(κ1-O-OAc)22-C,C-L3) (3a) (OAc = acetate), Pd(κ1-O-OAc)22-C,C-L4) (3b), Pd(κ1-O-OAc)22-C,C-L1) (4a), and Pd(κ1-O-OAc)22-C,C-L2) (4b). In addition, the phosphanyl-NHC-modified palladium acetate complex Pd(κ1-O-OAc)22-P,C-L5) (6) (L5 = 1-((2-diphenylphosphanyl)methylphenyl)-3-methyl-imidazol-2-ylidene) has been synthesized from corresponding palladium iodide complex PdI22-P,C-L5) (5). The reaction of the former complex with p-toluenesulfonic acid (p-TsOH) gave the corresponding bis-tosylate complex Pd(OTs)22-P,C-L5) (7). All new complexes have been characterized by multinuclear NMR spectroscopy and elemental analyses. In addition the solid-state structures of 1b·DMF, 2b·2DMF, 3a, 3b·DMF, 4a, 4b, and 6·CHCl3·2H2O have been determined by single crystal X-ray structure analyses. The palladium acetate complexes 3a/b, 4a/b, and 6 have been employed to catalyze the oxidative homocoupling reaction of terminal alkynes in acetonitrile chemoselectively yielding the corresponding 1,4-di-substituted 1,3-diyne in the presence of p-benzoquinone (BQ). The highest catalytic activity in the presence of BQ has been obtained with 6, while within the series of palladium-bis(NHC) complexes, 4b, featured with a n-propylene-bridge and the bulky N-1-naphthalenemethyl substituents, revealed as the most active compound. Hence, this latter precursor has been employed for analogous coupling reaction carried out in the presence of air pressure instead of BQ, yielding lower substrate conversion when compared to reaction performed in the presence of BQ. The important role of the ancillary ligand acetate in the course of the catalytic coupling reaction has been proved by variable-temperature NMR studies carried out with 6 and 7′ under catalytic reaction conditions.  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