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 93 毫秒
1.
Variable-temperature and -pressure 13C-NMR studies of the 1,3,5-trithiane-capped triruthenium clusters [Ru3(CO)93-(η3-1,3,5-trithiane)}] ( 1 ) and [Ru3(t-BuNC)(CO)83-(η3-1,3,5-trithiane)}] ( 2 ) revealed that CO site exchanges occur via an intramolecular merry-go-round process, involving a transition state mostly dissociative in character.  相似文献   

2.
Vibrational Spectra of Trimethylphosphonium Cations (CH3)3PX+ (X = H, D) and Crystal Structures of (CH3)3PD+SbCl6? and (CH3)3PCl+SbCl6? The trimethylphosphonium salts (CH3)3PX+SbCl6? (X = H, D) and (CH3)3PH+MF6? (M = As, Sb) are prepared and characterized by vibrational and NMR spectroscopy (1H, 31P, 13C). In addition the crystal structures of (CH3)3PD+SbCl6? and (CH3)3PCl+SbCl6? are reported. (CH3)3PD+SbCl6? crystallizes in the orthorhombic space group Pnma with a = 1555(1) pm, b = 753.1(8) pm, c = 1166(1) pm Z = 4. (CH3)3PCl+SbCl6? crystallizes triclinic in the space group P1 with a = 704.6(4) pm, b = 729.5(3) pm, c = 1391.1(7) pm, α = 89.57(4)°, b? = 88.04(4)°, γ = 74.98(4)° and Z = 2.  相似文献   

3.
Complex fac‐[Fe(CO)3(TePh)3]? was employed as a “metallo chelating” ligand to synthesize the neutral (CO)3Mn(μ‐TePh)3Fe(CO)3 obtained in a one‐step synthesis by treating fac‐[Fe(CO)3(TePh)3]? with fac‐[Mn‐(CO)3(CH3CN)3]+. It seems reasonable to conclude that the d6 Fe(II) [(CO)3Fe(TePh)3]? fragment is isolobal with the d6 Mn(I) [(CO)3Mn(TePh)3]2? fragment in complex (CO)3Mn(μ‐TePh)3Fe(CO)3. Addition of fac‐[Fe(CO)3(TePh)3]? to the CpNi(I)(PPh3) in THF resulted in formation of the neutral CpNi(TePh)(PPh3) also obtained from reaction of CpNi(I)(PPh3) and [Na][TePh] in MeOH. This investigation shows that fac‐[Fe(CO)3(TePh)3]? serves as a tridentate metallo ligand and tellurolate ligand‐transfer reagent. The study also indicated that the fac‐[Fe(CO)3(SePh)3]? may serve as a better tridentate metallo ligand and chalcogenolate ligand‐transfer reagent than fac‐[Fe(CO)3(TePh)3]? in the syntheses of heterometallic chalcogenolate complexes.  相似文献   

4.
H2Ru33-S)(CO)9 is deprotonated by K[HBBus3] to give cluster anions which react with [O{Au(PPh3)}3]+ or with AuCl(PPh3)/T1+ to give HRu3Au(μ3-S)(CO)9(PPh3) (1) and Ru3Au23-S)(CO)9(PPh3)2 (3). A similar sequence with HRu33-SBut)(CO)9 leads to Ru3Au(μ3-SBut)(CO)9(PPh3) (2) as the main product although some 1 also forms, indicating SC cleavage competes with deprotonation of HRu33-SBut)(CO)9 by [HBBus3]?. The X-ray crystal structures of 1, 2 and 3 are described; (1) and (2) have “butterfly” AuRu3 cores with markedly different hinge angles of 119 and 148° respectively, while 3 has a trigonal-bipyramidal Au2Ru3 skeleton. All three clusters have the sulphur atom symmetrically bridging the Ru3 triangular face.  相似文献   

5.
Bi3+ and lanthanide ions have been codoped in metal oxides as optical sensitizers and emitters. But such codoping is not known in typical semiconductors such as Si, GaAs, and CdSe. Metal halide perovskite with coordination number 6 provides an opportunity to codope Bi3+ and lanthanide ions. Codoping of Bi3+ and Ln3+ (Ln=Er and Yb) in Cs2AgInCl6 double perovskite is presented. Bi3+-Er3+ codoped Cs2AgInCl6 shows Er3+ f-electron emission at 1540 nm (suitable for low-loss optical communication). Bi3+ codoping decreases the excitation (absorption) energy, such that the samples can be excited with ca. 370 nm light. At that excitation, Bi3+-Er3+ codoped Cs2AgInCl6 shows ca. 45 times higher emission intensity compared to the Er3+ doped Cs2AgInCl6. Similar results are also observed in Bi3+-Yb3+ codoped sample emitting at 994 nm. A combination of temperature-dependent (5.7 K to 423 K) photoluminescence and calculations is used to understand the optical sensitization and emission processes.  相似文献   

6.
Five polymer-type new compounds-(η3-cyclopentadienyl)palladiumchloride ( 6 ), (η3-indenyl) palladiumchloride ( 7 ), (η3-cycloheptatrienyl)palladiumchloride ( 8 ), (η3-phenalenyl)palladiumchloride ( 9 ) and (1,2,3-η3-4,5,6,7-η4-cyclopheptatrienyl)(palladium-chloride)(tricarboryl Iron) ( 10 ) have been prepared from the reaction of Na2PdCl4 with 1-trimethylsilylcyclopentadiene ( 1 ), 1-trimethylsilylindene ( 2 ), 1-trimethylsilyl cycloheptatriene ( 3 ), 1-trimethylsilylphenalene ( 4 ) and 1-trimethylsilylcycloheptatriene tricarbonyl Iron ( 5 ) respectively. All the complexes( 6 )-(10) are obtained in excellent yield using the improved preparation route. Furthermore, a reaction mechanism is proposed.  相似文献   

7.
Two Fluoride Borates of Gadolinium: Gd2F3[BO3] and Gd3F3[BO3]2 By flux‐supported solid‐state reaction of Gd2O3 and GdF3 with B2O3 (flux: CsCl, molar ratio: 1 : 1 : 1 : 6, sealed tantalum capsule, 700 °C, 7 d) the new gadolinium fluoride borate Gd2F3[BO3] (monoclinic, P21/c; a = 1637.2(1), b = 624.78(4), c = 838.04(6) pm, β = 93.341(8)°; Vm = 64.418(6) cm3/mol, Z = 8) was obtained as colourless, prismatic, face‐rich single crystals. The four crystallographically different Gd3+ cations (CN = 9) are all capped square‐antiprismatically surrounded by fluoride and oxide anions, in which the latter represent always components of isolated trigonal planar [BO3]3— anions. The six crystallographically independent F anions all reside in more or less planar coordination of three Gd3+ cations. Thus the constitution of Gd2F3[BO3] can be described as a sequence of alternating layers each of the composition Gd[BO3] and GdF3 parallel (100), respectively. The crystal structures of Gd2F3[BO3] and the shortly published Gd3F3[BO3]2 (monoclinic, C2/c; a = 1253.4(1), b = 623.7(1), c = 836.0(1) pm, β = 97.404(6)°; Vm = 97.571(9) cm3/mol, Z = 4) are compared with each other. Due to the structural analogies between these two gadolinium fluoride borates, a disorder model of the boron atoms frequently found for Gd2F3[BO3] is able to be transferred to Gd3F3[BO3]2 as well.  相似文献   

8.
The Sr3Y(PO4)3:0.05Sm3+, Sr3Y(PO4)3:0.005Tb3+, and Sr3Y(PO4)3:0.005Tb3+, 0.05Sm3+ phosphors were synthesized using a conventional solid-state reaction technique at high temperature and their photoluminescence properties under ultraviolet (UV) excitation were studied. We observed the UV sensitization of Sm3+ emission (565, 600, and 648 nm) by Tb3+ in Sr3Y(PO4)3:0.005Tb3+, 0.05Sm3+, that leads to a white light emission with the CIE coordinate (0.367, 0.312) of Sr3Y(PO4)3:0.005Tb3+, 0.05Sm3+ phosphor under UV excitation. The emission is a result of partial energy transfer from Tb3+ to Sm3+, which is discussed in detail in terms of the corresponding excitation and emission spectra.  相似文献   

9.
A new cluster compound {[Mo3(μ3-O)(μ-S)3(dtp)3(py)3][CdI(dtp)2]} (dtp=S2P(O-was obtained from the reaction of (Mo3OS3(dtp)4(H2O)] with a CdI2-Bu4NI mixture.The molecular structure is composed of a cluster cation [Mo3OS3(dtp)3(py)3]+ and the complex anion [CdI(dtp)2]-Crystal data:Triclinic,space group P1 with cell parameters a=1.4672(7),b=1.5356(5),c=1.6806(5) nm,α=74.59(3),β=67.89(4),7=78.86(3)°,V=3.364(2) nm3,and Z=2,least-squares refinement of 8941 reflections gives a final agreement factor of R=0.052,Rw=0.065.  相似文献   

10.
The reaction IO + CH3SCH3 → products (3) was studied at room temperature and near 1 Torr pressure of He, using the discharge flow mass spectrometric technique. The rate constant was found to be k3 = (1.5 ± 0.5) × 10?11 cm3 molecule?1 s?1. CH3S(O)CH3 was detected as a product suggesting the following channel: IO + CH3SCH3 → CH3S(O)CH3 + I. The rate constant of the reaction IO + IO → products (1) was also measured: k1 = (3 ± 0.5) × 10?11 at 298 K and 1 Torr pressure. The atmospheric implication of reaction (3) is discussed. The results indicate that this reaction could be a potential important sink of CH3SCH3 in marine atmosphere.  相似文献   

11.
The title compound is formed together with Li[Tsi‐InI3] (the main product), (Tsi‐In(Me)I)2 and (Tsi)2InMe in very low yield in the reaction of InI3 with Tsi‐Li (Tsi = ‐C(Si(CH3)3)3)in toluene. According to the X‐ray structure determination this compound crystallizes in the triclinic space group P&1marc; and consists of slightly associated dimers of [Tsi‐InI3] anions via weak InI···I(‐InIII) contact‐bonds of 345.6(1) to 373.8(2) pm. Additionally the InI atom is capped by a D6‐benzene molecule.  相似文献   

12.
The intercombination transition 3s2 1S0 ?3s3p3P 1 0 in Ni16+ and Cu17+ has been studied by beam-foil spectroscopic methods. Decay curve analysis yields lifetime values of (12.0±1.0)ns and (8.8±0.6)ns for Ni and Cu in agreement with various predictions.  相似文献   

13.
The novel metalloid germanium cluster [Ge9(Hyp)2HypGe] ( 1 ) was synthesized, exhibiting two different bulky groups [Hyp = Si(SiMe3)3; HypGe = Ge(SiMe3)3]. Further reaction of 1 with ZnCl2 gives the derivative [ZnGe18(Hyp)4(HypGe)2] ( 2 ) in good yield, showing that the substitution of Si(SiMe3)3 by Ge(SiMe3)3 within a metalloid Ge9R3 compound leads to a comparable reactivity. 1 and 2 are characterized by NMR spectroscopy, mass spectrometry ( 1 ) and single crystal structure analyses ( 2 ). 1 and 2 are the first metalloid germanium clusters bearing germyl groups.  相似文献   

14.
Five new quaternary chalcogenides of the 1113 family, namely BaAgTbS3, BaCuGdTe3, BaCuTbTe3, BaAgTbTe3, and CsAgUTe3, were synthesized by the reactions of the elements at 1173–1273 K. For CsAgUTe3 CsCl flux was used. Their crystal structures were determined by single‐crystal X‐ray diffraction studies. The sulfide BaAgTbS3 crystallizes in the BaAgErS3 structure type in the monoclinic space group C3,2hC2/m, whereas the tellurides BaCuGdTe3, BaCuTbTe3, BaAgTbTe3, and CsAgUTe3 crystallize in the KCuZrS3 structure type in the orthorhombic space group D1,27,hCmcm. The BaAgTbS3 structure consists of edge‐sharing [TbS69–] octahedra and [AgS59–] trigonal pyramids. The connectivity of these polyhedra creates channels that are occupied by Ba atoms. The telluride structure features 2[MLnTe32–] layers for BaCuGdTe3, BaCuTbTe3, BaAgTbTe3, and 2[AgUTe31–] layers for CsAgUTe3. These layers comprise [MTe4] tetrahedra and [LnTe6] or [UTe6] octahedra. Ba or Cs atoms separate these layers. As there are no short Q ··· Q (Q = S or Te) interactions these compounds achieve charge balance as Ba2+M+Ln3+(Q2–)3 (Q = S and Te) and Cs+Ag+U4+(Te2–)3.  相似文献   

15.
The complexes [Au3(dcmp)2][X]3 {dcmp=bis(dicyclohexylphosphinomethyl)cyclohexylphosphine; X=Cl? ( 1 ), ClO4? ( 2 ), OTf? ( 3 ), PF6? ( 4 ), SCN?( 5 )}, [Ag3(dcmp)2][ClO4]3 ( 6 ), and [Ag3(dcmp)2Cl2][ClO4] ( 7 ) were prepared and their structures were determined by X‐ray crystallography. Complexes 2 – 4 display a high‐energy emission band with λmax at 442–452 nm, whereas 1 and 5 display a low‐energy emission with λmax at 558–634 nm in both solid state and in dichloromethane at 298 K. The former is assigned to the 3[5dσ*6pσ] excited state of [Au3(dcmp)2]3+, whereas the latter is attributed to an exciplex formed between the 3[5dσ*6pσ] excited state of [Au3(dcmp)2]3+ and the counterions. In solid state, complex [Ag3(dcmp)2][ClO4]3 ( 6 ) displays an intense emission band at 375 nm with a Stokes shift of ≈7200 cm?1 from the 1[4dσ*→5pσ] absorption band at 295 nm. The 375 nm emission band is assigned to the emission directly from the 3[4dσ*5pσ] excited state of 6 . Density functional theory (DFT) calculations revealed that the absorption and emission energies are inversely proportional to the number of metal ions (n) in polynuclear AuI and AgI linear chain complexes without close metal???anion contacts. The emission energies are extrapolated to be 715 and 446 nm for the infinite linear AuI and AgI chains, respectively, at metal???metal distances of about 2.93–3.02 Å. A QM/MM calculation on the model [Au3(dcmp)2Cl2]+ system, with Au???Cl contacts of 2.90–3.10 Å, gave optimized Au???Au distances of 2.99–3.11 Å in its lowest triplet excited state and the emission energies were calculated to be at approximately 600–690 nm, which are assigned to a three‐coordinate AuI site with its spectroscopic properties affected by AuI???AuI interactions.  相似文献   

16.
By sulfurization of phosphaalkenes ( a ) either (σ35)‐phosphoranes ( b ) or (σ33)‐thiaphosphiranes ( c ) are formed. In this study, Density Functional Theory (DFT) and coupled cluster (CCSD(T)) calculations have been carried out for model and experimental structures of (σ35)‐phosphoranes and (σ33)‐thiaphosphiranes to elucidate the factors influencing relative stabilities of b and c . According to the results of quantum chemical calculations, sterically bulky substituents make the phosphorane form more favored. Conversely, electronic effects of the most substituents provide higher stability for thiaphosphirane isomers. The only exception has been found in the cases where the substituent at the phosphorus atom possesses π‐donor and σ‐acceptor properties (e.g., in the case of amino group) and the substituents at carbon atom exhibit σ‐donor/π‐acceptor effects (e.g., silyl groups). The stability of the cyclic form c decreases further, if the substituents at the carbon atom are amino groups. In this case, a quite unusual structure has been theoretically predicted, which is considerably different from those of the hitherto known phosphoranes. It indicates a pyramidal configuration at the phosphorus atom and can be conventionally presented as a donor–acceptor adduct of diaminocarbene with thioxophosphine. © 2012 Wiley Periodicals, Inc.  相似文献   

17.
New Research of Reaction Behaviour of Triorganylcyclotriphosphines. The Crystal Structures of [(PPh3)2Pt(PtBu)3], [(PPh3)2Pd(PtBu)2], [(CO)4Cr{(PiPr)3}2], [RhCl(PPh3)(PtBu)3], [(NiCO)62-CO)3{(PtBu)2}2], and [(CpFeCO)2(μ-CO)(μ-PHtBu)]+ · [FeCl3(thf)] By the reaction of triorganylcyclotriphosphines with transition metal complexes single- and polynuclear compounds are formed, in which the cyclophosphines are bonded in different ways to the metal, the ring either preserving structure or under going ring opening. Depending on the reaction conditions the following compounds can be characterized: [(PPh3)2Pt(PtBu)3] ( 1 ), [(PPh3)2Pd(PtBu)2] ( 2 ), [(CO)4Cr{(PiPr)3}2] ( 3 ), [RhCl(PPh3)(PtBu)3] ( 4 ), [(NiCO)62-CO)3{(PtBu)2}2] ( 5 ) and [(CpFeCO)2(μ-CO)(μ-PHtBu)]+ · [FeCl3(thf)] ( 6 ). The structures of 1 – 6 were obtained by X-ray single crystal structure analysis ( 1 : space group P21/n (No. 14), Z = 4, a = 1279.6(3) pm, b = 1733.1(4) pm, c = 2079.1(4) pm, β = 90.20(3)°; 2 : space group P21/c (No. 14), Z = 4, a = 1053.3(2) pm, b = 2085.2(4) pm, c = 1855.7(4) pm, β = 98.77(3)°; 3 : space group P 1 (No. 2), Z = 2, a = 1022.6(2) pm, b = 1026.4(2) pm, c = 1706.0(3) pm, α = 82.36(3)°, β = 86.10(3)°, γ = 64.40(3)°; 4 : space group P 1 (No. 2), Z = 2, a = 980.2(2) pm, b = 1309.5(3) pm, c = 1573.4(3) pm, α = 99.09(3)°, β = 99.46(3)°, γ= 111.87(3)°; 5 : space group P21/c (No. 14), Z = 4, a = 1804.0(5) pm, b = 2261.2(6) pm, c = 1830.1(7) pm, β = 96.99(3)°; 6 : space group P21/c (No. 14), Z = 4, a = 943.2(3) pm, b = 2510.6(7) pm, c = 1325.1(6) pm, β = 98.21(3)°).  相似文献   

18.
Rate constants for the gas‐phase reactions of CH3OCH2CF3 (k1), CH3OCH3 (k2), CH3OCH2CH3 (k3), and CH3CH2OCH2CH3 (k4) with NO3 radicals were determined by means of a relative rate method at 298 K. NO3 radicals were prepared by thermal decomposition of N2O5 in a 700–750 Torr N2O5/NO2/NO3/air gas mixture in a 1‐m3 temperature‐controlled chamber. The measured rate constants at 298 K were k1 = (5.3 ± 0.9) × 10?18, k2 = (1.07 ± 0.10) × 10?16, k3 = (7.81 ± 0.36) × 10?16, and k4 = (2.80 ± 0.10) × 10?15 cm3 molecule?1 s?1. Potential energy surfaces for the NO3 radical reactions were computationally explored, and the rate constants of k1k5 were calculated according to the transition state theory. The calculated values of rate constants k1k4 were in reasonable agreement with the experimentally determined values. The calculated value of k5 was compared with the estimate (k5 < 5.3 × 10?21 cm3 molecule?1 s?1) derived from the correlation between the rate constants for reactions with NO3 radicals (k1k4) and the corresponding rate constants for reactions with OH radicals. We estimated the tropospheric lifetimes of CH3OCH2CF3 and CHF2CF2OCH2CF3 to be 240 and >2.4 × 105 years, respectively, with respect to reaction with NO3 radicals. The tropospheric lifetimes of these compounds are much shorter with respect to the OH reaction. © 2009 Wiley Periodicals, Inc. Int J Chem Kinet 41: 490–497, 2009  相似文献   

19.
The NMR Spectra of CF3I, CF3IF2, and CF3IF4 The 19F-NMR and 13C-NMR spectra of CF3I, CF3IF2 and CF3IF4 were recorded in acetonitrile solution. The chemical shifts of the CF3-groups are strongly dependent on the oxidation state of the iodine atom. With increasing oxidation state the resonances of the CF3-groups in the 19F-NMR spectra are characteristically shifted to high field, whereas in the 13C-NMR spectra a characteristic shift to low field is measured. The absolute value of the coupling constants 1J(19F? 13C) increases with increasing oxidation state from 344 Hz (CF3I) via 354 Hz (CF3IF2) to 359 Hz (CF3IF4).  相似文献   

20.
A series of Eu3+ ions co-doped (Gd0.9Y0.1)3Al5O12:Bi3+, Tb3+ (GYAG) phosphors have been synthesized by means of solvothermal reaction method. The XRD pattern of GYAG phosphor sintered at 1500 °C confirms their garnet phase. The luminescence properties of these phosphors have been explored by analyzing their excitation and emission spectra along with their decay curves. The excitation spectra of the GYAG:Bi3+, Tb3+, Eu3+ phosphors consists of broad bands in the shorter wavelength region due to 4f8 → 4f75d1 transition of Tb3+ ions overlapped with 6s2 → 6s16p1 (1S0 → 3P1) transition of Bi3+ ions and the charge transfer band of Eu3+–O2?. The present phosphors exhibit green and red colors due to 5D4 → 7F5 transition of Tb3+ ions and 5D0 → 7F1 transition of Eu3+ ions, respectively. The emission was shifted from green to red color by co-doping with Eu3+ ions, which indicate that the energy transfer probability from Tb3+ to Eu3+ ions are dependent strongly on the concentration of Eu3+ ions.  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