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 46 毫秒
1.
This article reports new square‐planar Fe(CO)4 D4h structures that are optimized, using the Hartree–Fock (HF) approach, and multiconfiguration self‐consistent field (MCSCF) theory in active space [2b2g2ega1ga2u]8, and which energy increased in sequence: 3B2g TS < 1A1g TS < 1A1g GS. A triple ζ valence basis set supplemented with 4f for Fe and 3d for C and O polarization shells [TZV (DF)] was used. At the HF/TZV (DF) level, 1A1g TS and 3B2g TS (3B2g TS energetically more favorable), there are transition states of tetrahedral inversion (defining stereochemical flexibility of Fe(CO)4) between known equivalent 1A1 and 3B2 Jahn–Teller distorted tetrahedron C2v structures with activation energy at ~0.96 kcal/mol according to the experimental data. 1A1g TS differs from 1A1g GS in electronic configuration by occupation of a1g and a2u MOs. At the MCSCF/ TZV (DF) level, 1A1g TS and 1A1g GS are optimized as near‐pure states in different potential energy surfaces (PES) avoided conical intersection with near‐equal interatomic distances, and define electronic flexibility of Fe(CO)4. Estimation of the energy separation in a two‐level system that avoids a conical intersection from vibrational spectrum is based on the effective Hamiltonian of the perturbation theory. The energy gap between two square‐planar Fe(CO)4 D4h 1A1g TS < 1A1g GS is 0.27 kcal/mol. The energy gap between 1A1g GS and 1A1 is 1.28 kcal/mol. It is possible to observe 3B2, 1A1 and 1A1g GS separately in the course of the experiment. © 2005 Wiley Periodicals, Inc. Int J Quantum Chem, 2006  相似文献   

2.
In a dissociation attachment experiment of water, three peaks were observed at 7,9, and 12 eV. The origin of the third peak has been believed to be 2B2. However, the calculated energy of this state is 0.6 eV higher than the experimental value. This discrepancy is quite large compared with the case of the lower two peaks. In this study we propose new candidates for resonant states responsible for the third peak. The configurations considered are (3a1)?1(3pa1)2, (3a1)?1(3pb1)2, (3a1)?1(3pb2)2, (3a1)?1(3pa1)1(3pb1)1, (3a1)?1(3pb2)1(3pa1)1, and (3a1)?1(3pb2)1(3pb1)1 which have the parent state (3a1)?1(3pa1)1, (3a1)?1(3pb1)1, or (3a1)?1(3pb2)1. The energy levels arising from these configurations are calculated by a method of configuration interaction. A Few resonance states, which could be responsible for the third peak, are found. New decay process of these states are proposed.  相似文献   

3.
A simple and effective synthetic route to homo‐ and heteroleptic rare‐earth (Ln = Y, La and Nd) complexes with a tridentate Schiff base anion has been demonstrated using exchange reactions of rare‐earth chlorides with in‐situ‐generated sodium (E)‐2‐{[(2‐methoxyphenyl)imino]methyl}phenoxide in different molar ratios in absolute methanol. Five crystal structures have been determined and studied, namely tris(2‐{[(2‐methoxyphenyl)imino]methyl}phenolato‐κ3O1,N,O2)lanthanum, [La(C14H12NO2)3], ( 1 ), tris(2‐{[(2‐methoxyphenyl)imino]methyl}phenolato‐κ3O1,N,O2)neodymium tetrahydrofuran disolvate, [La(C14H12NO2)3]·2C4H8O, ( 2 )·2THF, tris(2‐{[(2‐methoxyphenyl)imino]methyl}phenolato)‐κ3O1,N,O23O1,N,O22N,O1‐yttrium, [Y(C14H12NO2)3], ( 3 ), dichlorido‐1κCl,2κCl‐μ‐methanolato‐1:2κ2O:O‐methanol‐2κO‐(μ‐2‐{[(2‐methoxyphenyl)imino]methyl}phenolato‐1κ3O1,N,O2:2κO1)bis(2‐{[(2‐methoxyphenyl)imino]methyl}phenolato)‐1κ3O1,N,O2;2κ3O1,N,O2‐diyttrium–tetrahydrofuran–methanol (1/1/1), [Y2(C14H12NO2)3(CH3O)Cl2(CH4O)]·CH4O·C4H8O, ( 4 )·MeOH·THF, and bis(μ‐2‐{[(2‐methoxyphenyl)imino]methyl}phenolato‐1κ3O1,N,O2:2κO1)bis(2‐{[(2‐methoxyphenyl)imino]methyl}phenolato‐2κ3O1,N,O2)sodiumyttrium chloroform disolvate, [NaY(C14H12NO2)4]·2CHCl3, ( 5 )·2CHCl3. Structural peculiarities of homoleptic tris(iminophenoxide)s ( 1 )–( 3 ), binuclear tris(iminophenoxide) ( 4 ) and homoleptic ate tetrakis(iminophenoxide) ( 5 ) are discussed. The nonflat Schiff base ligand displays μ2‐κ3O1,N,O2O1 bridging, and κ3O1,N,O2 and κ2N,O1 terminal coordination modes, depending on steric congestion, which in turn depends on the ionic radii of the rare‐earth metals and the number of coordinated ligands. It has been demonstrated that interligand dihedral angles of the phenoxide ligand are convenient for comparing steric hindrance in complexes. ( 4 )·MeOH has a flat Y2O2 rhomboid core and exhibits both inter‐ and intramolecular MeO—H…Cl hydrogen bonding. Catalytic systems based on complexes ( 1 )–( 3 ) and ( 5 ) have demonstrated medium catalytic performance in acrylonitrile polymerization, providing polyacrylonitrile samples with narrow polydispersity.  相似文献   

4.
The relativistic CI method is used to determine N-electron wavefunctions for the 1s 2 2s 2 2p 2 (3 P 0,3 P 2,1 D 2), 1s 2 2p 4 3 P 2 even levels, and the 1s 22s2p 3 (3 D 1,3 P 1,3 S 1,1 P 1), 1s 22s 22p3s (3 P 1 and1 P 1), 1s 22s 22p3d (3 D 1,3 P 1,1 P 1)J=1 OIII levels. Excitation energies and emission probabilities between these levels are reported in the electric dipole approximation, both for the Coulomb and the Babushkin gauges.ns, p,np,nd- andnd (n17) numerical basis functions have been used for the construction of CSF's entering the CI expansion for the ASF's of these levels. Radiative matrix elements of the type calculated here within the framework of the relativistic CI method, may be used in laser assisted spectroscopic studies of atoms and ions.  相似文献   

5.
Ligand substitution kinetics for the reaction [PtIVMe3(X)(NN)]+NaY=[PtIVMe3(Y)(NN)]+NaX, where NN=bipy or phen, X=MeO, CH3COO, or HCOO, and Y=SCN or N3, has been studied in methanol at various temperatures. The kinetic parameters for the reaction are as follows. The reaction of [PtMe3(OMe)(phen)] with NaSCN: k1=36.1±10.0 s−1; ΔH1=65.9±14.2 kJ mol−1; ΔS1=6±47 J mol−1 K−1; k−2=0.0355±0.0034 s−1; ΔH−2=63.8±1.1 kJ mol−1; ΔS−2=−58.8±3.6 J mol−1 K−1; and k−1/k2=148±19. The reaction of [PtMe3(OAc)(bipy)] with NaN3: k1=26.2±0.1 s−1; ΔH1=60.5±6.6 kJ mol−1; ΔS1=−14±22 J mol−1K−1; k−2=0.134±0.081 s−1; ΔH−2=74.1±24.3 kJ mol−1; ΔS−2=−10±82 J mol−1K−1; and k−1/k2=0.479±0.012. The reaction of [PtMe3(OAc)(bipy)] with NaSCN: k1=26.4±0.3 s−1; ΔH1=59.6±6.7 kJ mol−1; ΔS1=−17±23 J mol−1K−1; k−2=0.174±0.200 s−1; ΔH−2=62.7±10.3 kJ mol−1; ΔS−2=−48±35 J mol−1K−1; and k−1/k2=1.01±0.08. The reaction of [PtMe3(OOCH)(bipy)] with NaN3: k1=36.8±0.3 s−1; ΔH1=66.4±4.7 kJ mol−1; ΔS1=7±16 J mol−1K−1; k−2=0.164±0.076 s−1; ΔH−2=47.0±18.1 kJ mol−1; ΔS−2=−101±61 J mol−1 K−1; and k−1/k2=5.90±0.18. The reaction of [PtMe3(OOCH)(bipy)] with NaSCN: k1 =33.5±0.2 s−1; ΔH1=58.0±0.4 kJ mol−1; ΔS1=−20.5±1.6 J mol−1 K−1; k−2=0.222±0.083 s−1; ΔH−2=54.9±6.3 kJ mol−1; ΔS−2=−73.0±21.3 J mol−1 K−1; and k−1/k2=12.0±0.3. Conditional pseudo-first-order rate constant k0 increased linearly with the concentration of NaY, while it decreased drastically with the concentration of NaX. Some plausible mechanisms were examined, and the following mechanism was proposed. [Note to reader: Please see article pdf to view this scheme.] © 1998 John Wiley & Sons, Inc. Int J Chem Kinet 30: 523–532, 1998  相似文献   

6.
The vibrational, rotational, and centrifugal constants are calculated for the B 1Π u , C 1Π u , (1) 1Π g , and (2) electronic states of a 85Rb2 molecule. The calculations are based on the semi-empirical potential curves obtained in this work. The results from calculating the molecular constants are compared with experimental data. The Franck-Condon factors and R v′v″ centroids are calculated for the electronic transitions B 1Π u -X 1Σ g +, C 1Π u -X 1Σ g +, C 1Π u -(1) 1Π g , and C 1Π u -(2) 1Σ g +.  相似文献   

7.
Ab initio electronic structure calculations are reported for low-lying electronic states, 1A1, 1A2, 3A2, 1B1, 3B1, 1B2, and 3B2 of the FNO2 molecule. Geometric parameters for the ground state 1A1 are predicted by MRSDCI calculations with a double-zeta plus polarization basis set. The vertical excitation energies for these electronic states are determined using MRSDCI/DZ+P calculations at the ground-state equilibrium conformation. The oscillator strengths and radiative lifetimes for some electronic states are calculated based on the MRSDCI wave functions. © 1993 John Wiley & Sons, Inc.  相似文献   

8.
Abstract

The EPR spectra of single crystals of 63Cu(II) doped N, N'-bis(salicylidene)ethylenediimine Ni(II), [Ni(sal)2en] and 7-methyl-N, N'-bis(salicylidene)ethylenediimine Ni(II), [Ni(7-me sal)2en] have been studied. The usual doublet spin-Hamiltonian parameters for the complexes have been found to be: Cu(II)[(sal)2en]; g z =2.192 ± 0.002; g x =2.046 ± 0.004; g y =2.049 ± 0.004; A z =201.0 × 10?4 cm?1; A x =29.3 × 10?4 cm?1; A y =31.3 × 10?4 cm?1; AN z =12.6 × 10?4 cm?1; A N x =14.5 × 10?4 cm?1; A N y =15.7 × 10?4 cm?1; A H z =6.3 × 10?4 cm?1; A H x =7.3 × 10?4 cm?1; A H y =7.9 × 10?4 cm?1; Cu(II)[(7-me sal)2en]; g z =2.189 ± 0.002; g x =2.037 ± 0.004; g y =2.046 ± 0.004; A z =203.0 × 10?4 cm?1; A x =36.9 × 10?4 cm?1; A y =22.7 × 10?4 cm?1; A N z =12.6 × 10?4 cm?1; A N x =13.3 × 10?4 cm?1; A N y =14.0 × 10?4 cm?1. Values of molecular orbital coefficients calculated for these complexes show that their bonding properties are similar to those of other compounds of this type. There is considerable covalency in the metal-ligand [sgrave]-bonds, and significant in-plane pi-bonding is present.  相似文献   

9.
A series of [Mn6O2(R1OH)4(sao)6(R2COO)2] complexes with terminal functional groups ( 1 : R1 = CH3, R2 = HO‐C6H4, 2 : R1 = C2H5, R2 = H2N‐C6H4, 3 : R1 = CH3, R2 = Cl‐C6H4, 4 : R1 = CH3, R2 = CH3S‐C6H4, 5 : R1 = CH3, R2 = I‐C6H4, 6 : R1 = CH3, R2 = pymSCH2, 7 : R1 = CH3, R2 = ortho‐pyr‐SCH3, 8 : R1 = C2H5, R2 = (CH3)3OOCNHCH2C6H4; sao = doubly deprotonated salicylaldoxime ligand, pym = pyrimidyl, pyr = pyridyl) have been obtained in a reaction of a ligand R2C6H4COOH, salicylaldoxime, manganese(II) perchlorate and [NEt4](OH) in methanol or a 1:1 mixture of ethanol and dichloromethane. In this report, structural aspects as well as preliminary studies of magnetic and thermal properties are presented. Compounds 1 , 3 , 6 , 8 exhibit an antiferromagnetic coupling of the Mn2+ ions, whereas 4 and 7 show ferromagnetic interactions. The title compounds may act as starting materials for further derivatization addressing the functional groups.  相似文献   

10.
This work demonstrates a new nonconventional ligand design, imidazole/pyridine‐based nonsymmetrical ditopic ligands ( 1 and 1 S ), to construct a dynamic open coordination cage from nonsymmetrical building blocks. Upon complex formation with Pd2+ at a 1:4 molar ratio, 1 and 1 S initially form mononuclear PdL4 complexes (Pd2+( 1 )4 and Pd2+( 1 S )4) without formation of a cage. The PdL4 complexes undergo a stoichiometrically controlled structural transition to Pd2L4 open cages ((Pd2+)2( 1 )4 and (Pd2+)2( 1 S )4) capable of anion binding, leading to turn‐on anion binding. The structural transitions between the Pd2L4 open cage and the PdL4 complex are reversible. Thus, stoichiometric addition (2 equiv) of free 1 S to the (Pd2+)2( 1 S )4 open cage holding a guest anion ((Pd2+)2( 1 S )4?G?) enables the structural transition to the Pd2+( 1 S )4 complex, which does not have a cage and thus causes the release of the guest anion (Pd2+( 1 S )4+G?).  相似文献   

11.
A one‐pot template condensation of 2‐(2‐(dicyanomethylene)hydrazinyl)benzenesulfonic acid (H2L1, 1 ) or 2‐(2‐(dicyanomethylene)hydrazinyl)benzoic acid (H2L2, 2 ) with methanol (a), ethylenediamine (b), ethanol (c) or water (d) on copper(II), led to a variety of metal complexes, that is, mononuclear [Cu(H2O)2O1N2 L1a] ( 3 ) and [Cu(H2O)(κO1N3 L1b)] ( 4 ), tetranuclear [Cu4(1 κO1N2:2 κO1 L2a)3‐(1 κO1, κN2:2 κO2 L2a)] ( 5 ), [Cu2(H2O)(1 κO1, κN2:2 κO1 L2c)‐(1 κO1,1 κN2:2 κO1,2 κN1‐ L2c)]2 ( 6 ) and [Cu2(H2O)2O1N2‐ L1dd)‐(1 κO1N2:2 κO1 L1dd)(μ‐H2O)]2 ? 2 H2O ( 7? 2 H2O), as well as polymer‐ ic [Cu(H2O)(κO1,1 κN2:2 κN1 L1c)]n ( 8 ) and [Cu(NH2C2H5)(κO1,1 κN2:2 κN1L2a)]n ( 9 ). The ligands 2‐SO3H‐C6H4‐(NH)N?C{(CN)[C(NH2)‐(?NCH2CH2NH2)]} (H2L1b, 10 ), 2‐CO2H‐C6H4‐(NH)N?{C(CN)[C(OCH3)‐(?NH)]} (H2L2a, 11 ) and 2‐SO3H‐C6H4‐(NH)N?C{C(?O)‐(NH2)}2 (H2L1dd, 12 ) were easily liberated upon respective treatment of 4 , 5 and 7 with HCl, whereas the formation of cyclic zwitterionic amidine 2‐(SO3?)? C6H4? N?NC(? C?(NH+)CH2CH2NH)(?CNHCH2CH2NH) ( 13 ) was observed when 1 was treated with ethylenediamine. The hydrogen bond‐induced E/Z isomerization of the (HL1d)? ligand occurs upon conversion of [{Na(H2O)2(μ‐H2O)2}(HL1d)]n ( 14 ) to [Cu(H2O)6][HL1d]2 ? 2 H2O ( 15 ) and [{CuNa(H2O)‐(κN1,1 κO2:2 κO1 L1d)2}K0.5(μ‐O)2]n ? H2O ( 16 ). The synthesized complexes 3 – 9 are catalyst precursors for both the selective oxidation of primary and secondary alcohols (to the corresponding carbonyl compounds) and the following diastereoselective nitroaldol (Henry) reaction, with typical yields of 80–99 %.  相似文献   

12.
Multireference perturbation theory with complete active space self-consistent field (CASSCF) reference functions is applied to the study of the valence π→π* excited states of 1,3-butadiene, 1,3,5-hexatriene, 1,3,5,7-octatetraene, and 1,3,5,7,9-decapentaene. Our focus was put on determining the nature of the two lowest-lying singlet excited states, 11Bu+ and 21Ag, and their ordering. The 11Bu+ state is a singly excited state with an ionic nature originating from the HOMO→LUMO one-electron transition while the covalent 21Ag state is the doubly excited state which comes mainly from the (HOMO)2→(LUMO)2 transition. The active-space and basis-set effects are taken into account to estimate the excitation energies of larger polyenes. For butadiene, the 11Bu+ state is calculated to be slightly lower by 0.1 eV than the doubly excited 21Ag state at the ground-state equilibrium geometry. For hexatriene, our calculations predict the two states to be virtually degenerate. Octatetraene is the first polyene for which we predict that the 21Ag state is the lowest excited singlet state at the ground-state geometry. The present theory also indicates that the 21Ag state lies clearly below the 11Bu+ state in decapentaene with the energy gap of 0.4 eV. The 0–0 transition and the emission energies are also calculated using the planar C2h relaxed excited-state geometries. The covalent 21Ag state is much more sensitive to the geometry variation than is the ionic 11Bu+ state, which places the 21Ag state significantly below the 11Bu+ state at the relaxed geometry. © 1998 John Wiley & Sons, Inc. Int J Quant Chem 66 : 157–175, 1998  相似文献   

13.
Abstract— Semimethylene blue was generated by reductive quenching of triplet methylene blue, 3MBH2+, with diphenylamine at pH 0.62–3.4. A Q-switched ruby laser flash-photolysis-kinetic spectro-photometric apparatus was used to characterize the absorption spectrum of semimethylene blue from 350 to 900 nm and a number of physical constants at 25°C with μ= 0.4 M and Cl? as the anion. The specific rate of quenching of 3MBH2+ by DPA is 2.8 × 109M?1 s?1 in 5% EtOH-95% water and 1.2×109M?1 s?1 in 50 v/v% aq. CH3CN. Corresponding efficiencies of net electron transfer are, respectively, 0.15 and 0.62. Spectral characteristics in 5% EtOH are, for MBH22±, λmax= 375 nm, ε375= 9000 M?1 cm?1; λmax= 880 nm, ε880= 12700 M?1 cm?1; for MBH±, λmax= 410 nm, ε410= 9800 M?1 cm?1, λmax= 880 nm, ε880= 33000 M?1 cm?1; for MBH± in 50 v/v% AN, λmax= 400 nm, ε400= 11000 M?1 cm?1 and λmax= 880 nm,ε880= 39000 M?1 cm?1. The pKa of MBH22ε calculated from the pH dependence of the absorption spectrum is 1.86 × 0.04 in 5% EtOH and 1.15 in 50 v/v% AN. Rate constants, kdecay, for reaction DPAH±+ with MBH22ε and MBH± in 5% EtOH are, respectively, 3.9 × 109 and 9.5 × 109M?1 s?1. The value of pKa of MBH22ε calculated from the dependence of kdecay on pH is 1.75 in 5% EtOH.  相似文献   

14.
Two new coordination polymers (CPs) formed from 5‐iodobenzene‐1,3‐dicarboxylic acid (H2iip) in the presence of the flexible 1,4‐bis(1H‐imidazol‐1‐yl)butane (bimb) auxiliary ligand, namely poly[[μ2‐1,4‐bis(1H‐imidazol‐1‐yl)butane‐κ2N3:N3′](μ3‐5‐iodobenzene‐1,3‐dicarboxylato‐κ4O1,O1′:O3:O3′)cobalt(II)], [Co(C8H3IO4)(C10H14N4)]n or [Co(iip)(bimb)]n, (1), and poly[[[μ2‐1,4‐bis(1H‐imidazol‐1‐yl)butane‐κ2N3:N3′](μ2‐5‐iodobenzene‐1,3‐dicarboxylato‐κ2O1:O3)zinc(II)] trihydrate], {[Zn(C8H3IO4)(C10H14N4)]·3H2O}n or {[Zn(iip)(bimb)]·3H2O}n, (2), were synthesized and characterized by FT–IR spectroscopy, thermogravimetric analysis (TGA), solid‐state UV–Vis spectroscopy, single‐crystal X‐ray diffraction analysis and powder X‐ray diffraction analysis (PXRD). The iip2− ligand in (1) adopts the (κ11‐μ2)(κ1, κ1‐μ1)‐μ3 coordination mode, linking adjacent secondary building units into a ladder‐like chain. These chains are further connected by the flexible bimb ligand in a transtranstrans conformation. As a result, a twofold three‐dimensional interpenetrating α‐Po network is formed. Complex (2) exhibits a two‐dimensional (4,4) topological network architecture in which the iip2− ligand shows the (κ1)(κ1)‐μ2 coordination mode. The solid‐state UV–Vis spectra of (1) and (2) were investigated, together with the fluorescence properties of (2) in the solid state.  相似文献   

15.
An entirely new class of heterobimetallic homoleptic glycolate complexes of the type Nb(OGO)3{Ta(OGO)2} [where G=CMe2CH2CH2CMe2 (G1) (3); CMe2CH2 CHMe(G2) (4); CHMeCHMe (G3) (5); CH2CMe2CH2 (G4) (6); CMe2CMe2(G5) (7); CH2CHMeCH2 (G6) (8); CH2CEt2CH2 (G7) (9); CH2CMe(Prn)CH2 (G8) (10)] have been prepared by the reactions of Nb(OGO)2(OGOH) [G=G1 (1a); G2 (1b); G3 (1c); G4 (1d); G5 (1e); G6 (1f); G7 (1g); G8 (1h)] with Ta(OGO)2 (OPri) (G=G1 (2a); G2 (2b); G3 (2c); G4 (2d); G5 (2e) G6 (2f); G7 (2g); G8 (2h). In addition to the novel derivatives (2)(10), our earlier investigations on heterobimetallic glycolate-alkoxide derivatives have been extended to derivatives of the type Nb(OGO) [where M=A1 n=3, G=G3 (11);G4 (12); G6 (13) G7 (14); Gs (15); G9=CH2CH2CH2 (16) and M=Ti (n=4, G=G4) (17), Zr(n=4,G=G4) (18)], which are conveniently prepared by the reactions of metalloligands Nb(OGO)2(OGOH) [G=G3 (1c); G4 (1d); G6 (1f); G7 (1g); G8 (1h); G9 (1i)] with different metal alkoxides. All of these new complexes have been characterized by elemental analyses, molecular weight determinations, and spectroscopic (I.r. and 1H, 27Al-n.m.r.) studies. Structural features of the new derivatives have been elucidated on the basis of molecular weight and spectroscopic data.  相似文献   

16.
Monodisperse metal clusters provide a unique platform for investigating magnetic exchange within molecular magnets. Herein, the core–shell structure of the monodisperse molecule magnet of [Gd52Ni56(IDA)48(OH)154(H2O)38]@SiO2 ( 1 a @SiO2) was prepared by encapsulating one high‐nuclearity lanthanide–transition‐metal compound of [Gd52Ni56(IDA)48(OH)154(H2O)38]?(NO3)18?164 H2O ( 1 ) (IDA=iminodiacetate) into one silica nanosphere through a facile one‐pot microemulsion method. 1 a @SiO2 was characterized using transmission electron microscopy, N2 adsorption–desorption isotherms, and inductively coupled plasma‐atomic emission spectrometry. Magnetic investigation of 1 and 1 a revealed J1=0.25 cm?1, J2=?0.060 cm?1, J3=?0.22 cm?1, J4=?8.63 cm?1, g=1.95, and z J=?2.0×10?3 cm?1 for 1 , and J1=0.26 cm?1, J2=?0.065 cm?1, J3=?0.23 cm?1, J4=?8.40 cm?1 g=1.99, and z J=0.000 cm?1 for 1 a @SiO2. The z J=0 in 1 a @SiO2 suggests that weak antiferromagnetic coupling between the compounds is shielded by silica nanospheres.  相似文献   

17.
The collision-induced dissociation (CID) spectra of five alkylmethyleneimmonium ions (H2C-N+R1R2, (a) R1 = R2 = C2H5, (b) R1 = n-C3H7, R2 = H, (c) R1 = n-C3H7, R2 = CH3, (d) R1 = n-C3H7, R2 = C2H5, (e) R1 = R2 = n-C3H7) are reported and discussed in terms of the mechanism of alkane loss. The most abundant alkane losses result from 2-azaallylic bond cleavages within R1 and R2 leading to daughter ions of m/z 84. Ion d (R1 = n-C3H7, R2 = C2H5) was chosen for a deuterium-labelling study because it exhibited methane loss nearly free from interferences with other fragmentations. The methane lost consists to a great extent (95%) of the methyl moiety of R2. Whereas the methyl moiety obviously stays intact during the fragmentation process, the hydrogen additionally needed originates from all positions of R1 and the double-bonded methylene in an approximately random distribution, suggesting extensive hydrogen migrations preceding the transfer step.  相似文献   

18.
Accurate lower and upper bounds for the nonrelativistic lowest energies1 E 0 and3 E 0 of the singlet and triplet-system of the4He-Isotop are calculated with the linearized method of variance minimization. The same was done for1 E 1 the energy of the first excitedS-state 21 S. The results especially for1 E 0 and3 E 0 in a.u. are −2.903307699751 E 0 ≤ −2.90330769218 −2.174932426373 E 0 ≤ −2.17493242459 i.e. the values are determined with an absolute error smaller than 0.00167 cm−1 for1 E 0 and 0.00039 cm−1 for3 E 0.  相似文献   

19.
2‐Phenylethanol, racemic 1‐phenyl‐2‐propanol, and 2‐methyl‐1‐phenyl‐2‐propanol have been pyrolyzed in a static system over the temperature range 449.3–490.6°C and pressure range 65–198 torr. The decomposition reactions of these alcohols in seasoned vessels are homogeneous, unimolecular, and follow a first‐order rate law. The Arrhenius equations for the overall decomposition and partial rates of products formation were found as follows: for 2‐phenylethanol, overall rate log k1(s−1)=12.43−228.1 kJ mol−1 (2.303 RT)−1, toluene formation log k1(s−1)=12.97−249.2 kJ mol−1 (2.303 RT)−1, styrene formation log k1(s−1)=12.40−229.2 kJ mol−1(2.303 RT)−1, ethylbenzene formation log k1(s−1)=12.96−253.2 kJ mol−1(2.303 RT)−1; for 1‐phenyl‐2‐propanol, overall rate log k1(s−1)=13.03−233.5 kJ mol−1(2.303 RT)−1, toluene formation log k1(s−1)=13.04−240.1 kJ mol−1(2.303 RT)−1, unsaturated hydrocarbons+indene formation log k1(s−1)=12.19−224.3 kJ mol−1(2.303 RT)−1; for 2‐methyl‐1‐phenyl‐2‐propanol, overall rate log k1(s−1)=12.68−222.1 kJ mol−1(2.303 RT)−1, toluene formation log k1(s−1)=12.65−222.9 kJ mol−1(2.303 RT)−1, phenylpropenes formation log k1(s−1)=12.27−226.2 kJ mol−1(2.303 RT)−1. The overall decomposition rates of the 2‐hydroxyalkylbenzenes show a small but significant increase from primary to tertiary alcohol reactant. Two competitive eliminations are shown by each of the substrates: the dehydration process tends to decrease in relative importance from the primary to the tertiary alcohol substrate, while toluene formation increases. © 1999 John Wiley & Sons, Inc. Int J Chem Kinet 31: 401–407, 1999  相似文献   

20.
The conformational changes in a sugar moiety along the hydrolytic pathway are key to understand the mechanism of glycoside hydrolases (GHs) and to design new inhibitors. The two predominant itineraries for mannosidases go via OS2B2,51S5 and 3S13H41C4. For the CAZy family 92, the conformational itinerary was unknown. Published complexes of Bacteroides thetaiotaomicron GH92 catalyst with a S-glycoside and mannoimidazole indicate a 4C14H5/1S51S5 mechanism. However, as observed with the GH125 family, S-glycosides may not act always as good mimics of GH's natural substrate. Here we present a cooperative study between computations and experiments where our results predict the E5B2,5/1S51S5 pathway for GH92 enzymes. Furthermore, we demonstrate the Michaelis complex mimicry of a new kind of C-disaccharides, whose biochemical applicability was still a chimera.  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