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 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   

2.
A new optically active ONNO-type tetradentate ligand, ethylenediamine-N,N′- di-S-isobutylacetate (SS-eniba), has been synthesized. During the preparation of diaqua cobalt(III) complexes of SS-eniba, [Co(SS-eniba)(H2O)2]+, the title ligand has coordinated stereospecifically to the cobalt(III) ion to give three isomers, Δ-s-cis, Δ-uns-cis and Λ-uns-cis, which have been isolated and characterized via electronic absorption, circular dichroism (CD), and 1H NMR spectroscopy, along with elemental analysis data. The preparation of Δ-s-cis-[Co(SS-eniba)Cl2]+ and Δ-s-cis-[Co(SS-eniba)CO3]+ are also reported.  相似文献   

3.
The fraction FΣ of excited-state oxygen formed as b 1Σg+ was determined for a series of triplet-state photosensitizers in CCl4 solutions. FΣ was determined by monitoring the intensities of (a) O2(b 1Σg+) fluorescence at 1926 nm (O2(b 1Σg+)→O2(a 1Δg) and (b) O2(a 1 Δg) phosphorescence at 1270 nm (O2(a 1Δg) → O2(X3Σg)). Oxygen excited states were formed by energy transfer from substituted benzophenones and acetophenones. The data indicate that FΣ depends on several variables including the orbital configuration of the lowest triplet state and the triplet-state energy. The available data indicate that the sensitizer-oxygen charge transfer (CT) state is not likely to influence FΣ strongly by CT-mediated mixing of various sensitizer-oxygen states.  相似文献   

4.
Large-scale MRD CI calculations assign to AlP the ground state X 3Σ (9σ22) and a close-lying state 1 3Π (9σ3π3) (Te = 0.08 eV). Up to transition energies of 2.0 eV, other states are described by the configurations 9σ3π3 (11Π), 8σ24 (1 1Σ+), 9σ22 (1 1Δ and 2 1Σ+) and 9σ3π24π (1 5Π). The 2 3Π state, located at ≈ 2.30 eV, shows a shallow double minimum. Numerous perturbations are expected to induce predissociation upon 2 3Π. Multiplets arising from the occupation 8σ234π are clustered in the 3.25–3.50 eV region. Quintet states with the configuration 8σ9σ3π34π are bound, with Te values (in eV) of 3.80 (1 5Σ+), 4.44 (1 5Δ) and 4.88 (3 5Σ), respectively. The 9σ → 4s Rydberg members 5Σ and 3Σ lie in the 4.58–4.72 eV energy region. The first ionization potential (ionization to X4Σ of AlP+, 9σ → ∞) is estimated to be 7.65 eV. Ionization to the 1 2Σ and 1 2Π states of AlP+ is suggested to occur between 8.0 and 8.8 eV. The dipole moments of X 3Σ, 1 1Δ and 2 1Σ+ are close to 1.0 D, whereas the 1 1Σ+ state has μ = 3.49 D; 1 3Π and 1 1Π have dipole moments from 2.45 to 2.91 D. All low-lying states show a polarity Al+P. Finally, the electronic structure and transition energies of AlP are compared with those of the isoelectronic species BN, AIN, and SiP+.  相似文献   

5.
Twenty-two isomers/conformers of C3H6S+√ radical cations have been identified and their heats of formation (ΔHf) at 0 and 298 K have been calculated using the Gaussian-3 (G3) method. Seven of these isomers are known and their ΔHf data are available in the literature for comparison. The least energy isomer is found to be the thioacetone radical cation (4+) with C2v symmetry. In contrast, the least energy C3H6O+√ isomer is the 1-propen-2-ol radical cation. The G3 ΔHf298 of 4+ is calculated to be 859.4 kJ mol−1, ca. 38 kJ mol−1 higher than the literature value, ≤821 kJ mol−1. For allyl mercaptan radical cation (7+), the G3 ΔHf298 is calculated to be 927.8 kJ mol−1, also not in good agreement with the experimental estimate, 956 kJ mol−1. Upon examining the experimental data and carrying out further calculations, it is shown that the G3 ΔHf298 values for 4+ and 7+ should be more reliable than the compiled values. For the five remaining cations with available experimental thermal data, the agreement between the experimental and G3 results ranges from fair to excellent.

Cation CH3CHSCH2+√ (10+) has the least energy among the eleven distonic radical cations identified. Their ΔHf298 values range from 918 to 1151 kJ mol−1. Nevertheless, only one of them, CH2=SCH2CH2+√ (12+), has been observed. Its G3 ΔHf298 value is 980.9 kJ mol−1, in fair agreement with the experimental result, 990 kJ mol−1.

A couple of reactions involving C3H6S+√ isomers CH2=SCH2CH2+√ (12+) and trimethylene sulfide radical cation (13+) have also been studied with the G3 method and the results are consistent with experimental findings.  相似文献   


6.
The radiative lifetimes of the b1Σ+ and a1Δ states have been evaluated by perturbation expansions including X3Σ, a1Δ, b1Σ+, 13,1Π, 23,1Π, 23Σ and 21Σ+ states. All wavefunctions result from large MRD CI calculations. The b—X transition is dominated by the parallel transition moment; it is found to be much stronger than the a—X transition. The calculated radiative lifetimes of τ(1Σ+)=18 ms, τ(1Δ)=2.2 s for NF and τ(1Σ+)=2.5–3.5 ms for NCl are in good accord with corresponding experimentally deduced values. The lifetime for the a1Δ state in NCl is found to be τ(1Δ)=1.1 s, ie. much longer than derived from a recent experiment. Its magnitude is consistent with the τ(b1Σ+)/τ(a1Δ) ratio of similar systems and with the decrease in lifetime from NF to NCl and is thus believed to be quite reliable. A detailed analysis of all contributions of the perturber states to the transition mechanism is made and comparison with the related data in SO, O2 and S2 is undertaken. The b-a transition probability dominated by the quadrupole transition is fairly constant in all the systems in the order of A = 0.013 (NF) - 0.0013 (S2) s−1.  相似文献   

7.
Medium-resolution spectra of the N2 b1Πu-X1Σg+ band system were recorded by 1 + 1 multiphoton ionization. In the spectra we found different linewidths for transitions to different vibrational levels in the b 1Πu state: Δν0 = 0.50 ± 0.05 cm−1, Δν1 = 0.28 ± 0.02 cm−1, Δν2 = 0.65 ± 0.06 cm−1, Δν3 = 3.2 ± 0.5 cm−1, Δν4 = 0.60 ± 0.07 cm−1, and Δν5 = 0.28 ± 0.02 cm−1. From these linewidths, predissociation lifetimes τν were obtained: τ0 = 16 ± 3 ps, τ1 > 150 ps, τ2 = 10 ± 2 ps, τ3 = 1.6 ± 0.3 ps, τ4 = 9 ± 2 ps, and τ5 > 150 ps. Band origins and rotational constants for the b 1Πuν = 0 and 1 levels were determined for the 14N2 and 14N15N molecules.  相似文献   

8.
Gao J  Zha F  Chen H  Kang J 《Talanta》1995,42(12):1897-1903
The reaction behaviour of the β-type chelates of lanthanide ions (Ln3+) with p-bromochloroarsenazo (4-CAsA-pB) in 0.01 mol l−1 HClO4 solution has been studied systematically by a spectrophotometric method. All the lanthanide ions can form β-type chelates with p-bromochloroarsenazo. The maximum absorption wavelength is in the range 727–731 nm, the molar absorptivities are about 6.0 × 104 – 9.0 × 104 cm2 mol−1, the composition ratio of Ln3+ ions with 4-CAsA-pB is 1:2 and the actual combining ratio is 2:4. The optimum acidity range (ΔpH value) of the formation of β-type chelates has been obtained. Kinetic parameters, such as the reaction order and rate constants, have also been studied and a formation mechanism for the β-type chelates has been proposed.  相似文献   

9.
The reactions of the diruthenium carbonyl complexes [Ru2(μ-dppm)2(CO)4(μ,η2-O2CMe)]X (X=BF4 (1a) or PF6 (1b)) with neutral or anionic bidentate ligands (L,L) afford a series of the diruthenium bridging carbonyl complexes [Ru2(μ-dppm)2(μ-CO)22-(L,L))2]Xn ((L,L)=acetate (O2CMe), 2,2′-bipyridine (bpy), acetylacetonate (acac), 8-quinolinolate (quin); n=0, 1, 2). Apparently with coordination of the bidentate ligands, the bound acetate ligand of [Ru2(μ-dppm)2(CO)4(μ,η2-O2CMe)]+ either migrates within the same complex or into a different one, or is simply replaced. The reaction of [Ru2(μ-dppm)2(CO)4(μ,η2-O2CMe)]+ (1) with 2,2′-bipyridine produces [Ru2(μ-dppm)2(μ-CO)22-O2CMe)2] (2), [Ru2(μ-dppm)2(μ-CO)22-O2CMe)(η2-bpy)]+ (3), and [Ru2(μ-dppm)2(μ-CO)22-bpy)2]2+ (4). Alternatively compound 2 can be prepared from the reaction of 1a with MeCO2H–Et3N, while compound 4 can be obtained from the reaction of 3 with bpy. The reaction of 1b with acetylacetone–Et3N produces [Ru2(μ-dppm)2(μ-CO)22-O2CMe)(η2-acac)] (5) and [Ru2(μ-dppm)2(μ-CO)22-acac)2] (6). Compound 2 can also react with acetylacetone–Et3N to produce 6. Surprisingly [Ru2(μ-dppm)2(μ-CO)22-quin)2] (7) was obtained stereospecifically as the only one product from the reaction of 1b with 8-quinolinol–Et3N. The structure of 7 has been established by X-ray crystallography and found to adopt a cis geometry. Further, the stereospecific reaction is probably caused by the second-sphere π–π face-to-face stacking interactions between the phenyl rings of dppm and the electron-deficient six-membered ring moiety of the bound quinolinate (i.e. the N-included six-membered ring) in 7. The presence of such interactions is indeed supported by an observed charge-transfer band in a UV–vis spectrum.  相似文献   

10.
Condensation of thiosemicarbazide or N(4)-ethylthiosemicarbazide with 1,2,8,9-tetraphenyl-3,7-diazanona-1,9-dione in the presence of copper(II) acetate in 96% ethanol leads to Δ6-5,6-diphenyl-5-methoxy-1,2,4-triazacyclohexene-3-thione, C16H15N3OS, or Δ6-4-methyl-5,6-diphenyl-5-ethoxy-1,2,4-triazacyclohexene-3-thione, C18H19N3OS. For C16H15N3OS the crystal data are monoclinic, P21/c, a=9.7780(7), b=8.5120(3), c=18.2210(13) Å, β=100.958(3)°, V=1488.89(16) Å3, and Z=4 in agreement with an earlier report. For C18H19N3OS the crystal data are orthorhombic, P212121, a=8.6940(3), b=12.9946(3), c=15.5139(5) Å, V=1752.68(9) Å3, and Z=4.  相似文献   

11.
CpIr(η4-C6H6) (2) has been obtained in high yield by a four-step synthesis. Thermal reaction of 2 with CpCO(C2H4)2 and photochemical reaction of 2 with CpRh(C2H4)2 or CpRh(C2H4)2 give the compounds μ-(η3: η3-C6H6)CoIrCp2 (3), μ-(η3: η3-C6H6)RhIrCp2 (4), and μ-(η3: η3-C6H6)(RhCp)(IrCp) (5), respectively. The X-ray crystallography data of 3 and 4 reveal a boat-shaped conformation of the synfacially bridging benzene ligand with a rather long Co---Ir bond distance in 3 and a relatively short Rh---Ir bond length in 4 which are caused by almost constant folding angles of the benzene unit. The dynamic behaviour of the benzene bridge was investigated by NMR spectrometry.  相似文献   

12.
A mixture of NF3 and Ar is passed through an rf discharge in a flow-system to produce, among other species, F and NF2. When H2, D2, or CH4 are added downstream, reactions with F atoms produce vibrationally excited HF or DF together with H, D, or CH3. The latter free radicals can react with NF2, probably by an elimination reaction to produce electronically excited NF: NF2(2B1) + H(D, CH3) → HF*(DF* + NF(a1Δ). A vibrational-to-electronic energy transfer process between the products of this reaction then produces the next higher state of NF: HF(ν 2) + NF(a1Δ) → HF(ν−2) + NF(b1Σ+). A similar transfer process has also been found between the electronically excited a1Δ states of O2 and NF: O2(a1Δ) + NF(a1Δ) → O2(X3Σ) + NF(b1Σ+). The H or D atoms but not the CH3 radicals are then found to react with either NF(a1Δ) or NF(X3Σ) to produce electronically excited N(2D) atoms, which in turn react with the NF(a1Δ) molecules to produce N2(B3Πg). The observed nitrogen first positive radiation has been demonstrated to be produced entirely by this reaction mechanism rather than by the N(4S) recombination that accounts for the Rayleigh afterglow. In addition, the occurrence of the reaction N(2D) + N2O → NO(B2Πr) + N2 (X1Σ+g) has been verified. Finally we have observed emission at 3344 Å, which we attribute to the NF(A3Π), which has not been previously reported.  相似文献   

13.
The crystal structures of propionaldehyde complex (RS,SR)-(η5-C5H5)Re(NO)(PPh3)(η2-O=CHCH2CH3)]+ PF6 (1b+ PF6s−; monoclinic, P21/c (No. 14), a = 10.166 (1) Å, b = 18.316(1) Å, c = 14.872(2) Å, β = 100.51(1)°, Z = 4) and butyraldehyde complex (RS,SR)-[(η5-C5H5)Re(NO)(PPh3)(η2-O=CHCH2CH2CH3)]+ PF6 (1c+PF6; monoclinic, P21/a (No. 14), a = 14.851(1) Å, b = 18.623(3) Å, c = 10.026(2) Å, β = 102.95(1)°, Z = 4) have been determined at 22°C and −125°C, respectively. These exhibit C O bond lengths (1.35(1), 1.338(5) Å) that are intermediate between those of propionaldehyde (1.209(4) Å) and 1-propanol (1.41 Å). Other geometric features are analyzed. Reaction of [(η5-C5H5)Re(NO)(PPh3)(ClCH2Cl)]+ BF4 and pivalaldehyde gives [(η5-C5H5)Re(NO)(PPh3)(η2-O=CHC(CH3)3)]+BF4 (81%), the spectroscopic properties of which establish a π C O binding mode.  相似文献   

14.
1,2-Bis(dimethylamino)-1,2-dibora-[2]ferrocenophane (1) was prepared by the reaction of 1,1′-dilithioferrocene with 1,2-dichlorobis(dimethylamino)diborane(4). In addition to hindered rotation about the B-N bond (ΔG > 80 kJ mol−1), another dynamic process was revealed by 1H and 13C NMR in solution at low temperature, and interpreted as motion of the cyclopentadienyl rings between staggered and eclipsed conformations (ΔG(233 K) = 44 ± 1 kJ mol−1).  相似文献   

15.
The bis(μ3-ethylidyne) tricobalt cluster [(CpCo)33-CCH3)2] (1b) is protonated by trifluoroacetic acid to give the dicobalt edge-protonated cation [H(CpCo)33-CCH3)2]+ [lb + H]+. Protonation of the μ3-ethylidyne tetracobalt cluster hydride [H(CpCo)43-CCH3)] (3) takes place in two consecutive steps. At low temperature [H2(CpCo)43-CCH3)]+ [3 + H]+ is formed first, and is then slowly converted into [H3(CpCo)43-CCH3)]2+ [3 + 2H]2+ by an excess of acid. As judged by the 1H NMR data and the crystal structure of [3 + X]+[(CF3COO)2X] (X = H or D) the endo hydrogens in [3 + H]+ and [3 + 2H]2+ occupy μ3-(Co3) face capping hydridic positions. The cations [1b + H]+ and [3 + H]+ show hydride fluxionality in solution, which in the case of [3 + H]+ can be frozen out on the NMR timescale at low temperature (ΔG (203 K) = 40.8 kJ/mol). The structure of [3 + X]+ [(CF3COO)2X] (X = H or D) was determined by X-ray crystallography. One of the hydrides/deuterides is located on the crystallographic mirror plane, capping a tricobalt face of the cluster cation. The other endo hydrogen atom is believed to be disordered between the other two μ3-(Co3) sites, which are related by space group symmetry. Deuteronation of 3 shows a strong normal kinetic deuterium isotope effect. From the temperature independence of the 1H NMR spectrum of [3 + 2D]2+ a non-fluxional solution structure can be inferred. In all the systems studied, hydridic (μ2- or μ3-) sites are thermodynamically preferred to possible isomeric agostic CoHC or Co2HC sites for the endo hydrogens. Agostic interactions cannot, however, be ruled out in transient intermediates during the course of the protonations.  相似文献   

16.
The crystal structure of bis(trifluoroacetato)-(N-methyl-meso-tetraphenylporphyrinato) thallium(III), Tl(N---Me---tpp)(CF3CO2)2 (2), was established and the coordination sphere around the Tl3+ ion is described as 4:3 tetragonal base–trigonal base piano stool seven-coordinate geometry in which the two cis CF3CO2 − groups occupy two apical sites. The plane of the three pyrrole nitrogen atoms [i.e. N(2), N(3) and N(4)] strongly bonded to Tl3+ is adopted as the reference plane 3N. The pyrrole N(1) ring bearing the methyl group [i.e. C(45)H3] is the most deviated one from the 3N plane making a dihedral angle of 23.3° whereas smaller angles of 9.9, 2.7 and 4.7° occur with pyrroles N(2), N(3), and N(4), respectively. Because of the larger size of the thallium(III) ion, Tl is considerably out of the 3N plane; its displacement of 1.02 Å is in the same direction as that of the two apical CF3CO2 − ligands. The intermolecular trifluoroacetate exchange process for 2 in CD2Cl2 solvent is examined through 19F and 13C NMR temperature-dependent measurements. In the slow-exchange region, the CF3 and carbonyl (CO) carbons of the CF3CO2 − groups in 2 are separately located at δ 114.3 [1J(C–F)=290 Hz, 3J(Tl–C)=411 Hz] and 155.1 [2J(C–F)=37 Hz, 2J(Tl–C)=204 Hz], respectively, at −106 °C. In the same slow-exchange region, the fluorine atoms of 2, Tl(N---Me---tpp)(CF3CO2)+ and the free CF3CO2 − are located at δ −73.76 [4J(Tl–F)=44 Hz], −73.30 [4J(Tl–F)=22 Hz], and −76.15 ppm at −97 °C, respectively.  相似文献   

17.
Some (η5-cyclopentadienyl)(1,2-bis(diarylphosphino)ethane)(diorganosulfide)ruthenium complexes, [Ru(η5-C5H5)(Ar2PCH2CH2-PAr2)(R1R 2S)]BF4 (Ar = Ph, p-Tol; R1, R2 = Ph, Et) were prepared. Variable temperature NMR spectra of these complexes showed the existence of two fluxional processes; inversion at the sulfur atom and δ-λ interconversion of the chelate ring. The former process was slower, and its barriers in these complexes were calculated as ca. 7 kcal mol−1. The spectral features of ethyl phenyl sulfide complexes suggested that substantiation of the new chiral center at sulfur induces a significant conformational rigidity at the chelate ring.  相似文献   

18.
The perphenylmetallocene complexes (η5-C5Ph5)2W (1), [(η5-C5Ph5)2W]+I3 (1+I3), (η5-C5Ph5)2Mo (2) and [(η5-C5Ph5)2Mo]+I3 (2+I3) have been prepared. Hydrogenation of 1 in THF produces (η5-C5Ph5)2WH2 (4), while (η5-C5Ph5)2WHCl (3) is afforded in 1,2-dichloroethane solvent. Carbonylation of 1 produces (η5-C5Ph5)2W(CO) (5). Treatment of 1 with the strong acid CF3SO3H leads to the dicationic species [(η5-C5Ph5)2W]+2[CF3SO3]2 (1+2Tf2) after crystallization. The structures of 2+I3 and 1+2Tf2 have been determined by an X-ray diffraction study. The magnetic susceptibility study indicates a 3E2g ground-state for 1 and 2, and a 4A2g ground-state for 1+ and 2+.  相似文献   

19.
D. Kupfer 《Tetrahedron》1961,15(1-4):193-196
The reduction of steroidal ketones in different solvents yielded different products. These conditions which altered the normal path, yielded a selective reduction of the Δ4-3 carbonyl without the concomitant reduction of the C-17 or C-20 ketones; this permitted a one step partial synthesis of 3β-hydroxy-Δ4-pregnen-20-one (I), 3β-hydroxy-5-pregnan-20-one (II), and of 3β,11β-dihydroxyandrostan-17-one (IV).  相似文献   

20.
In order to understand the nature of the putative cationic 12-electron species [M(η51-C5R4SiMe2NR′)R″]+ of titanium catalysts supported by a linked amido-cyclopentadienyl ligand, several derivatives with different cyclopentadienyl C5R4 and amido substituents R′ were studied systematically. The use of tridentate variants (C5R4SiMe2NCH2CH2X)2− (C5R4=C5Me4, C5H4, C5H3tBu; X=OMe, SMe, NMe2) allowed the NMR spectroscopic observation of the titanium benzyl cations [Ti(η51-C5Me4SiMe2NCH2CH2X)(CH2Ph)]+. Isoelectronic neutral rare earth metal complexes [Ln(η51-C5R4SiMe2NR′)R″] can be expected to be active for polymerization. To arrive at neutral 12-electron hydride and alkyl species of the rare earth metals, we employed a lanthanide tris(alkyl) complex [Ln(CH2SiMe3)3(THF)2] (Ln=Y, Lu, Yb, Er, Tb), which allows the facile synthesis of the linked amido-cyclopentadienyl complex [Ln(η51-C5Me4SiMe2NCMe3)(CH2SiMe3)(THF)]. Hydrogenolysis of the linked amido-cyclopentadienyl alkyl complex leads to the dimeric hydrido complex [Ln(η51-C5Me4SiMe2NCMe3)(THF)(μ-H)]2. These complexes are single-site, single-component catalysts for the polymerization of ethylene and a variety of polar monomers such as acrylates and acrylonitrile. Nonpolar monomers such as -olefins and styrene, in contrast, give isolable mono-insertion products which allow detailed studies of the initiation process.  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