首页 | 官方网站   微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 343 毫秒
1.
The electronic and geometrical structures of the low-energy states of 1,4,5,8-naphthalenetetracarboxylic dianhydride parent diimide (1) are studied in terms of the complete active space self-consistent field (CASSCF) method employed at different level with respect to the size and the quality of the active space. In the framework of the vibronic model based on the Franck–Condon (FC) effect the absorption and magnetic circular dichroism (MCD) spectra are studied in the excitation region corresponding to two low-energy 11Ag → 11B2u and 11Ag → 11B3u electronic transitions in diimides. In that (visible) excitation region the CASSCF computations with the 5π[4n]5π active space (i.e., the naphthalene-like π orbitals enriched by the four lone pair orbitals of the oxygen atoms) were found to reproduce very well the empirical absorption and the MCD spectra measured for the dicyclohexyl-N,N-substituted diimide (2). At the same CASSCF/5π[4n]5π level, the electronic absorption of diimides in the near UV excitation region were attributed to the 11Ag → 21B1u, 11Ag → 21B3u and 11Ag → 21B2u electronic transitions; the latter two are mostly localized on the “diimide chromophore”. For these transitions the calculated magneto-optical characteristics, such as sign pattern and intensity distribution in the MCD spectrum, were found to be consistent with that experimentally observed for the diimide 2 compound.  相似文献   

2.
The photophysical properties, which vary as R is varied, of a series of [Pt(N2O2)] complexes bearing bis(phenoxy)bipyridine auxiliaries with different substituents R=H (Pt-H) (1), 4,4′-2NH2 (Pt-NH2) (2), 4,4′-2tBu (Pt-tBu) (3), 4,4′-2CN (Pt-CN) (4), and 4,4′-2NO2 (Pt-NO2) (5) are investigated using density functional theory (DFT) and time-dependent density functional theory (TDDFT). The solvent effects are discussed in CH2Cl2, CH3CN and CH3OH solutions, respectively, by polarizable continuum model (PCM). It is anticipated that compared with σ-donor substituents, π-acceptors have more dramatic effects on the electronic and optical properties in this series of complexes. Introduction of π-electron withdrawing substituents on bipyridine ligand will benefit the LLCT (or MLCT) and prohibit the non-radiative pathways via d–d transitions by increasing the energy gap between the HOMO–LUMO and d–d transitions. The results also reveal that the lowest-energy excitations of all complexes show blue-shifts in the polarized solution and when the polarity of the solvent increases from CH2Cl2, CH3CN and CH3OH, the low-energy broad absorption band exhibit blue-shifts. The lowest-energy excitations and photoluminescence of all complexes are dominated by π(phenoxy)→π*(bpy/NO2) (LLCT) excited state mixed with some energetically dπ (Pt)→π*(bpy/NO2) (MLCT) transition.  相似文献   

3.
The optical spectroscopy of YAG:Ni, Zr(Si) and GSGG:Ni, Zr(Si) crystals is presented. Absorption bands of Ni2+ ions in octahedral and tetrahedral coordinations are observed. Comparison of experimentally observed and predicted energy levels for Ni2+ in both sites is made and the parameters Dq, B and C are determined. Luminescence transitions in the near-infrared, green and red are investigated and assigned to the 3T23A2, 1T23A2 and 1T23T2 transitions of octahedral Ni2+ ions. Photocoloration of the crystals indicate that part of the dopant tetrahedral nickel ions turn into the trivalent state. The absorption spectrum of Ni3+ ions in GSGG:Ni is analyzed.  相似文献   

4.
Zeeman spectral data are presented for the 2Π3/2: J = 7/2J= 9/2,2Π3/2: J= 7/22Π1/2:J= 5/2 and2Π3/2 J= 3/2J = 5/2 transitions in OD. Data for the 2Π3/2. J=3/2→ J= 5/2 and 2Π3/2 J= 5/22Πl/2 : J= 3/2 transitions in OH, taken under similar conditions, are included.  相似文献   

5.
The equilibrium geometries, excitation energies, force constants and vibrational frequencies for seven low-lying electronic states X 1A1, 1B1, 3B1, 1A2, 3A2, 1B2 and 3B2 of dichlorocarbene CCl2 have been calculated at the MRSDCI level with a double-zeta plus polarization basis set. Our calculated equilibrium geometry for the X 1A1 state, excitation energy for X 1A11B1 and vibrational frequencies for the X 1A1 and 1B1 states are in good agreement with experimental data. The electronic transition dipole moments, oscillator strengths for the 1B1 → X 1A1 and 1B2 → X 1A1 transitions, radiative lifetimes for the 1B1 and 1B1 states are calculated using MRSDCI wavefunctions, predicting results in reasonable agreement with experiment.  相似文献   

6.
CdII complexes with glycine (gly) and sarcosine (sar) were studied by glass electrode potentiometry, direct current polarography, virtual potentiometry, and molecular modelling. The electrochemically reversible CdII–glycine–OH labile system was best described by a model consisting of M(HL), ML, ML2, ML3, ML(OH) and ML2(OH) (M = CdII, L = gly) with the overall stability constants, as log β, determined to be 10.30 ± 0.05, 4.21 ± 0.03, 7.30 ± 0.05, 9.84 ± 0.04, 8.9 ± 0.1, and 10.75 ± 0.10, respectively. In case of the electrochemically quasi-reversible CdII–sarcosine–OH labile system, only ML, ML2 and ML3 (M = CdII, L = sar) were found and their stability constants, as log β, were determined to be 3.80 ± 0.03, 6.91 ± 0.07, and 8.9 ± 0.4, respectively. Stability constants for the ML complexes, the prime focus of this work, were thus established with an uncertainty smaller than 0.05 log units. The observed departure from electrochemical reversibility for the Cd–sarcosine–OH system was attributed mainly to the decrease in the transfer coefficient . The MM2 force field, supplemented by additional parameters, reproduced the reported crystal structures of diaqua-bis(glycinato-O,N)nickel(II) and fac-tri(glycinato)-nickelate(II) very well. These parameters were used to predict structures of all possible isomers of (i) [Ni(H2O)4(gly)]+ and [Ni(H2O)4(sar)]+; and (ii) [Ni(H2O)3(IDA)] and [Ni(H2O)3(MIDA)] (IDA = iminodiacetic acid, MIDA = N-methyl iminodiacetic acid) by molecular mechanics/simulated annealing methods. The change in strain energy, ΔUstr, that accompanies the substitution of one ligand by another (ML + L′ → ML′ + L), was computed and a strain energy ΔUstr = +0.28 kcal mol−1 for the reaction [Ni(H2O)4(gly)]+ + sar → [Ni(H2O)4(sar)]+ + gly was found. This predicts the monoglycine complex to be marginally more stable. By contrast, for the reaction [Ni(H2O)3IDA] + MIDA → [Ni(H2O)3MIDA] + IDA, ΔUstr = −0.64 kcal mol−1, and the monoMIDA complex is predicted to be more stable. This correlates well with (i) stability constants for Cd–gly and Cd–sar reported here; and (ii) known stability constants of ML complex for glycine, sarcosine, IDA, and MIDA.  相似文献   

7.
Reactions of [(η6-arene)RuCl2]2 (1) (η6-arene=p-cymene (1a), 1,3,5-Me3C6H3 (1b), 1,2,3-Me3C6H3 (1c) 1,2,3,4-Me4C6H2(1d), 1,2,3,5-Me4C6H2 (1e) and C6Me6 (1f)) or [Cp*MCl2]2 (M=Rh (2), Ir (3); Cp*=C5Me5) with 4-isocyanoazobenzene (RNC) and 4,4′-diisocyanoazobenzene (CN–R–NC) gave mononuclear and dinuclear complexes, [(η6-arene)Ru(CNC6H4N=NC6H5)Cl2] (4a–f), [Cp*M(CNC6H4N=NC6H5)Cl2] (5: M=Rh; 6: M=Ir), [{(η6-arene)RuCl2}2{μ-CNC6H4N=NC6H4NC}] (8a–f) and [(Cp*MCl2)2(μ-CNC6H4N=NC6H4NC)}] (9: M=Rh; 10: M=Ir), respectively. It was confirmed by X-ray analyses of 4a and 5 that these complexes have trans-forms for the ---N=N--- moieties. Reaction of [Cp*Rh(dppf)(MeCN)](PF6)2 (dppf=1,1′-bis (diphenylphosphino)ferrocene) with 4-isocyanoazobenzene gave [Cp*Rh(dppf)(CNC6H4N=NC6H5)](PF6)2 (7), confirmed by X-ray analysis. Complex 8b reacted with Ag(CF3SO3), giving a rectangular tetranuclear complex 11b, [{(η6-1,3,5-Me3C6H3)Ru(μ-Cl}4(μ-CNC6H4N=NC6H4NC)2](CF3SO3)4 bridged by four Cl atoms and two μ-diisocyanoazobenzene ligands. Photochemical reactions of the ruthenium complexes (4 and 8) led to the decomposition of the complexes, whereas those of 5, 7, 9 and 10 underwent a trans-to-cis isomerization. In the electrochemical reactions the reductive waves about −1.50 V for 4 and −1.44 V for 8 are due to the reduction of azo group, [---N=N---]→[---N=N---]2−. The irreversible oxidative waves at ca. 0.87 V for the 4 and at ca. 0.85 V for 8 came from the oxidation of Ru(II)→Ru(III).  相似文献   

8.
The reaction of 2-arylpyridinecarboxaldimine [RH4C6NC(H)Py, L (1)] with hydrated RuX3 (X = Cl, Br) in boiling C2H5OH affords dark crystals of RuX2L2. Two geometrical isomers of the compound have been isolated and characterized by analytical and spectroscopic data. The trans isomer of RuCl2L2 shows a single sharp band for ν(Ru---Cl), whereas two bands are observed for the corresponding cis isomer. The highresolution 1H NMR spectra of the isolated complexes are reported and completely assigned. All the complexes have multiple t2→π*(L) transitions in the visible region. Each of the complexes display a quasi-reversible oxidative response due to an RuIII/RuII couple in the range 0.25–0.40 V vs S.C.E. at a platinum working electrode. The formal potentials of this couple obey the Hammett relationship. The ligand-based irreversible oxidations are also briefly noted.  相似文献   

9.
The reaction of RuII(PPh3)3X2 (X = Cl, Br) with o-(OH)C6H4C(H)=N-CH2C6H5 (HL) under aerobic conditions affords RuII(L)2(PPh3)2, 1, in which both the ligands (L) are bound to the metal center at the phenolic oxygen (deprotonated) and azomethine nitrogen and RuIII(L1)(L2)(PPh3), 2, in which one L is in bidentate N,O form like in complex 1 and the other ligand is in tridentate C,N,O mode where cyclometallation takes place from the ortho carbon atom (deprotonated) of the benzyl amine fragment. The complex 1 is unstable in solution, and undergoes spontaneous oxidative internal transformation to complex 2. In solid state upon heating, 1 initially converts to 2 quantitatively and further heating causes the rearrangement of complex 2 to the stable RuL3 complex. The presence of symmetry in the diamagnetic, electrically neutral complex 1 is confirmed by 1H and 31P NMR spectroscopy. It exhibits an RuII → L, MLCT transition at 460 nm and a ligand based transition at 340 nm. The complex 1 undergoes quasi-reversible ruthenium(II)—ruthenium(III) oxidation at 1.27V vs. SCE. The one-electron paramagnetic cyclometallated ruthenium(III) complex 2 displays an L → RuIII, LMCT transition at 658 nm. The ligand based transition is observed to take place at 343 nm. The complex 2 shows reversible ruthenium(III)—ruthenium(IV) oxidation at 0.875V and irreversible ruthenium(III)—ruthenium(II) reduction at −0.68V vs. SCE. It exhibits a rhombic EPR spectrum, that has been analysed to furnish values of axial (6560 cm−1) and rhombic (5630 cm−1) distortion parameters as well as the energies of the two expected ligand field transitions (3877 cm−1 and 9540 cm−1) within the t2 shell. One of the transitions has been experimentally observed in the predicted region (9090 cm−1). The first order rate constants at different temperatures and the activation parameter ΔH#S# values of the conversion process of 1 → 2 have been determined spectrophotometrically in chloroform solution.  相似文献   

10.
The collisional behaviour of Ba[6s5d(3DJ)], 1.151 eV above the 6s2(1S0) electronic ground state, in the presence of atomic strontium, has been investigated in the ‘long-time domain' (ca. 100 μs–1 ms) following the pulsed dye-laser excitation of barium vapour at elevated temperature at λ = 553.5 nm (Ba[6s6p(1P1)] ← Ba[6s2(1S0)]. Ba(3DJ) is subsequently produced from the short-lived 1P1 state (τe = 8.37 ± 0.38 ns) by a number of radiative and collisional processes. It may then be monitored in the ‘long-time domain' by atomic spectroscopic marker methods involving either collisional activation of Ba(3DJ) by Ba(1S0) and He buffer gas to yield Ba[6s6p(3PJ)] with subsequent emission from the 3P1 state (τe = 1.2 ± 0.1 μs): Ba[6s6p(3P1)] → Ba[6s2(1S0)] + hv (λ = 791.1 nm). Alternatively, emission from Ba(1P1) may be monitored at long times following the generation of this short-lived state by energy pooling following self-annihilation of Ba(3DJ) + Ba(3DJ) from Ba[6s6p(1P1)] → Ba[6s2(1S0)] + hv (λ = 553.5 nm). The generation of Ba(3DJ) in the presence of atomic strontium yields emission in the long-time domain from Sr[5s5p(3P1)] (τe = 19.6 μs): Sr[5s5p(3P1)] → Sr[5s2(1S0)]  + hv (λ = 689.3 nm). Whilst the decay profiles at short times are complex in form, at long times all these atomic profiles show first-order kinetic removal with the decay coefficients for λ = 791.1 nm, 689.3 nm and 553.5 nm emissions in the ratio 1 : 2 : 2, consistent with overall third-order activation of the form: Ba(3DJ) + Ba(3DJ) + Sr(1S0) → Sr(3PJ) + 2Ba(1S0). The mechanism is modelled in detail, including measurement of integrated emission intensities, yielding kinetic data for fundamental collisional processes. The overall rate constant for the third-order collisional activation of Sr[5s5p(3PJ])from 2Ba[6s5d(3DJ)] + Sr[5s2(1S0)] takes the upper limit of 5.8 × 10−27 cm6 atom−2 s−1 (T = 900 K). The rate constant for the two body collisional quenching of Ba[6s5d(3DJ)] by ground state atomic strontium, Sr[5s2(1S0)], is found to be (2.0 ± 0.1) × 10−12 cm3 atom−1 s−1 (T = 900 K).  相似文献   

11.
Polarized absorption spectra of Ba(MnO4)2·3H2O/Ba(ClO4)2·3H2O mixed single crystals are reported at 4.2°K. Previous 1T21A1 assignments for the 5200 Å and 3000 Å absorption bands of MnO4 are substantiated; further support is provided for the 1T11A1 assignment of the 3600 Å absorption band of MnO4. The site-splitting of the 5200 Å 1T2 state is E(1E)−E(1A) ≈ −150 cm−1; that of the 3000 Å 1T2 state is E(1E)−E(1A) ≈ 300 cm−1. A significant e vibronic intensity component is observed in the 5200 Å 1T2 state.  相似文献   

12.
The and -benzyl derivatives (1 and 2, respectively) of (+)-camphor have been synthesized and are found to exert a strong influence on the circular dichroism n→π* Cotton effects: 1: Δε301max -0.36 (n- heptane) and 2: Δε302max +3.22, relative to camphor: Δε304max +1.8 (n-heptane). Evidence for electric dipole transition moment coupling in these γ, δ -unsaturated systems is found in the n→π* UV: 1: ε291max 84 (n-heptane) and 2: ε285max 303, relative to camphor: ε290max 25.  相似文献   

13.
The complexes formed from copper(II) and 2-(5-bromo-2-pyridylazo)-5-diethylaminophenol (5-Br-PADAP or HL) in aqueous methanol solution was studied by electrospray ionization mass spectrometry. The solution of a 1:1 complex of Cu(II) with 5-Br-PADAP showed five peaks assignable to a binuclear complex [Cu2L2(AcO)]+ and mononuclear complexes [CuL]+, [CuL(H2O)]+, [CuL(AcOH)]+ and [CuL(HL)]+ (AcO=acetate). Collision activated dissociation revealed the relative order of bonding strengths; Cu–L>Cu–HL>CuL–AcOH>CuL–H2O. The peak intensities of the binuclear complex showed second-order dependency on those of the mono complex. As for the solution of Ni(II)–5-Br-PADAP, no binuclear complex was observed in the mass spectra. Thus, it was suggested that [Cu2L2(AcO)]+ was formed by the fast gas phase reaction: 2[CuL]++AcO[Cu2L2(AcO)]+.  相似文献   

14.
The synthesis of the new (η2-dppe)(η5-C5Me5)Fe---CC---1,3-(C6H4X) (m-2a/2b; X=F/Br) and (η2-dppe)(η5-C5Me5)Fe---CC---1,4-(C6H4I) (2c) complexes, as well as the solid-state structure of the known (η2-dppe)(η5-C5Me5)Fe---CC---1,4-(C6H4F) (2a) complex are described. The catalytic coupling reactions of the bromo complexes with various alkynes were next investigated. Starting from the known (η2-dppe)(η5-C5Me5)Fe---CC---1,4-(C6H4Br) complex (2b), the synthesis of the (η2-dppe)(η5-C5Me5)Fe---CC---1,4-(C6H4)---CC---H complex (6d) and of the corresponding silyl-protected precursors (η2-dppe)(η5-C5Me5)Fe---CC---1,4-(C6H4)CC---SiR3 (6b/6c; R=iPr/Me) are reported. By use of lithium---bromine exchange reactions on 2b, the silyl- (7a; E=Si; R=Me) and tin- (7b–7d; E=Sn; R=Me, Bu, Ph) substituted analogues (η2-dppe)(η5-C5Me5)Fe---CC---1,4-(C6H4)ER3 are also isolated. The spectroscopic and electrochemical characterisations of all these new Fe(II)/Fe(III) redox-active building blocks are presented and the electronic substituent parameters for the “(η2-dppe)(η5-C5Me5)Fe---CC” group are determined by means of 19F-NMR.  相似文献   

15.
Four novel mixed ligand complexes of Cu(II), Co(II), Ni(II) and Zn(II) with saccharin and nicotinamide were synthesised and characterised on the basis of elemental analysis, FT-IR spectroscopic study, UV–Vis spectrometric and magnetic susceptibility data. The structure of the Cu (II) complex is completely different from those of the Co(II), Ni(II) and Zn(II) complexes. From the frequencies of the saccharinato CO and SO2 modes, it has been proven that the saccharinato ligands in the structure of the Cu complex are coordinated to the metal ion ([Cu(NA)2(Sac)2(H2O)], where NA — nicotinamide, Sac — saccharinato ligand or ion), whilst in the Co(II), Ni(II) and Zn(II) complexes are uncoordinated and exist as ions ([M(NA)2(H2O)4](Sac)2).  相似文献   

16.
The one-pot reaction between the novel proton transfer compound (pydaH2)2+(phendc)2−, LH2, and Cu(II) afforded the compounds (pydaH)2[Cu(phendc)2]·10H2O, 1, and (pydaH)2[Cu(phendc)(phendcH)]2·5H2O, 2, where pyda=2,6-diaminopyridine, and phendcH2=1,10-phenanthroline-2,9-dicarboxylic acid. The single crystal X-ray diffraction analysis of 1 and 2 revealed that these are two novel self-assembled 3D Cu(II) complex-organo-networks, in which (pydaH)+ ions and [Cu(phendc)2]2− or complex units are held together by ion pairing, H-bonding, and π–π interactions. Magnetic measurements over the temperature range 1.8–310 K revealed no significant magnetic coupling between Cu(II) centers in 1 or 2.  相似文献   

17.
Saddle point geometries and barrier heights have been calculated for the H abstraction reaction HO2(2A″)+H(2S) → H2(1Σ+g)+O2(3Σg) and the concerted H approach-O removing reaction HO2 (2A″)+H(2S) → H2O(1A1)+O(3P) by using SDCI wavefunctions with a valence double-zeta plus polarization basis set. The saddle points are found to be of Cs symmetry and the barrier heights are respectively 5.3 and 19.8 kcal by including size consistent correction. Moreoever kinetic parameters have been evaluated within the framework of the TST theory. So activation energies and the rate constants are estimated to be respectively 2.3 kcal and 0.4×109 ℓ mol−1 s−1 for the first reaction, 20.0 kcal and 5.4.10−5 ℓ mol−1 s−1 for the second. Comparison of these results with experimental determinations shows that hydrogen abstraction on HO2 is an efficient mechanism for the formation of H2 + O2, while the concerted mechanism envisaged for the formation of H2O + O is highly unlikely.  相似文献   

18.
Mg+—Ar ion—molecule complexes are produced in a pulsed supersonic nozzle cluster source. The complexes are mass selected and studied with laser photodissociation spectroscopy in a reflectron time-of-flight mass spectrometer system. An electronic transition assigned as X 2Σ+2Π is observed with an origin at 31387 cm−1 (vac) for 24Mg+—Ar. The 24Mg+—Ar spectrum is characterized by a 15 member progression with a frequency (ω′e) of 272 cm−1. An extrapolation of this progression fixes the excited state dissociation energy (Do) at 5552 cm−1. The corresponding ground-state value (Do) is 1270 cm−1 (3.6 kcal/mol). The 2Π , spin—orbit splitting is 76 cm.  相似文献   

19.
In addition to the red phosphorescence (T1(3 A2n, π*) → S0) xanthione exhibits in solution an emission with a maximum at ≈ 23 000 cm−1 and φf(298°) = 5 × 10−3. It is shown that this emission is fluorescence from the second excited singlet state (S2 (1A1 π, π*) → S0).  相似文献   

20.
The ground-state and low-lying excited electronic states in mononuclear, {H2B[3,5-(CF3)2Pz]2M(2,4,6-Cn)} (M1) and dinuclear {[3,5-(CF3)2Pz]M(2,4,6-Cn)}2, (M2) (Pz = pyrazole, Cn = collidine and M = Cu, Ag), are studied using DFT approach. Electronic properties are calculated using B3LYP, while excited singlet and triplet-states are examined using TD-B3LYP. All the calculated low-lying transitions are categorized as 1MLCT transitions. A good agreement was found between experimental spectra and predicted emission wavelengths (λem), the corresponding emissive states being assigned as 3MLCT for Cu1 and Ag2, 3MLLCT for Ag1 and 3LLCT for Cu2.  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-23

京公网安备 11010802026262号