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 31 毫秒
1.
Surface enhancement mechanism of Raman scattering from molecules adsorbed on silver oxide colloids is reported. Absorption spectra and Raman spectra of the cyanine dye D266 and pyridine molecules adsorbed on Ag2O colloids, and the influences of S2O32− and OH on the SERS are studied respectively. The results indicate that ‘chemical' enhancement is dominant in Ag2O colloidal solution. Surface complexes of adsorbed molecules and small silver ion clusters Agn+ as the SERS active sites make an important contribution to surface enhanced Raman scattering (SERS). At these active sites, charge transfer between the adsorbed molecules and the small silver ion clusters is the main enhancement origin. The enhancement factor of D266 adsorbed on Ag2O colloids is theoretically estimated with the excited-state charge transfer model, which is roughly in accordance with the experiments.  相似文献   

2.
In trans‐bis(5‐n‐butyl­pyridine‐2‐carboxyl­ato‐κ2N,O)­bis­(methanol‐κO)copper(II), [Cu(C10H12NO2)2(CH4O)2], the Cu atom lies on a centre of symmetry and has a distorted octahedral coordination. The Cu—O(methanol) bond length in the axial direction is 2.596 (3) Å, which is much longer than the Cu—­O(carboxylate) and Cu—N distances in the equatorial plane [1.952 (2) and 1.977 (2) Å, respectively]. In mer‐tris(5‐n‐bu­tyl­pyridine‐2‐carboxyl­ato‐κ2N,O)­iron(III), [Fe(C10H12NO2)3], the Fe atom also has a distorted octahedral geometry, with Fe—O and Fe—N bond‐length ranges of 1.949 (4)–1.970 (4) and 2.116 (5)–2.161 (5) Å, respectively. Both crystals are stabilized by stacking interactions of the 5‐n‐butyl­pyridine‐2‐carboxyl­ate ligand, although hydrogen bonds also contribute to the stabilization of the copper(II) complex.  相似文献   

3.
High oxidation potential perfluorinated zinc phthalocyanines (ZnFnPcs) are synthesised and their spectroscopic, redox, and light‐induced electron‐transfer properties investigated systematically by forming donor–acceptor dyads through metal–ligand axial coordination of fullerene (C60) derivatives. Absorption and fluorescence spectral studies reveal efficient binding of the pyridine‐ (Py) and phenylimidazole‐functionalised fullerene (C60Im) derivatives to the zinc centre of the FnPcs. The determined binding constants, K, in o‐dichlorobenzene for the 1:1 complexes are in the order of 104 to 105 M ?1; nearly an order of magnitude higher than that observed for the dyad formed from zinc phthalocyanine (ZnPc) lacking fluorine substituents. The geometry and electronic structure of the dyads are determined by using the B3LYP/6‐31G* method. The HOMO and LUMO levels are located on the Pc and C60 entities, respectively; this suggests the formation of ZnFnPc.+–C60Im.? and ZnFnPc.+–C60Py.? (n=0, 8 or 16) intra‐supramolecular charge‐separated states during electron transfer. Electrochemical studies on the ZnPc–C60 dyads enable accurate determination of their oxidation and reduction potentials and the energy of the charge‐separated states. The energy of the charge‐separated state for dyads composed of ZnFnPc is higher than that of normal ZnPc–C60 dyads and reveals their significance in harvesting higher amounts of light energy. Evidence for charge separation in the dyads is secured from femtosecond transient absorption studies in nonpolar toluene. Kinetic evaluation of the cation and anion radical ion peaks reveals ultrafast charge separation and charge recombination in dyads composed of perfluorinated phthalocyanine and fullerene; this implies their significance in solar‐energy harvesting and optoelectronic device building applications.  相似文献   

4.
Both of the title compounds, catena‐poly­[[[tetra­aqua­magnesium(I)]‐μ‐4,4′‐bi­pyridine‐κ2N:N′] diiodide bis(4,4′‐bi­pyridine) solvate], {[Mg(C10H8N2)(H2O)4]I2·2C10H8N2}n, (I), and catena‐poly­[[[μ‐4,4′‐bi­pyridine‐bis­[di­iodo­bis­(propan‐1‐ol)­strontium(I)]]‐di‐μ‐4,4′‐bi­pyridine‐κ4N:N′] bis(4,4′‐bi­pyri­dine) solvate], {[Sr2I4(C10H8N2)3(C3H8O)4]·2C10H8N2}n, (II), are one‐dimensional polymers which are single‐ and double‐stranded, respectively, the metal atoms being linked by the 4,4′‐bi­pyridine moieties. The Mg complex, (I), is [cis‐{(H2O)4Mg(N‐4,4′‐bi­pyridine‐N′)(2/2)}](∞|∞)I2·4,4′‐bi­pyridine and Mg has a six‐coordinate quasi‐octahedral coordination environment. The Sr complex, (II), is isomorphous with its previously defined Ba counterpart [Kepert, Waters & White (1996). Aust. J. Chem. 49 , 117–135], being [(propan‐1‐ol)2I2Sr(N‐4,4′‐bi­pyridine‐N′)(3/2)](∞|∞)·4,4′‐bi­pyridine, with the I atoms trans‐axial in a seven‐coordinate pentagonal–bipyramidal Sr environment.  相似文献   

5.
Surface-enhanced Raman scattering (SERS) spectroscopy and density functional theory (DFT) calculations were used to investigate the nature of the charge-transfer (CT) process between nitrothiophenol (NTP) isomers and the n-type semiconductor, TiO2. The Raman signals of p-NTP and m-NTP that were chemisorbed onto TiO2 were significantly enhanced with respect to their corresponding neat compounds. In particular, an enhancement factor (EF) of 102–103 was observed for both p-NTP and m-NTP, with m-NTP displaying a larger EF compared to p-NTP. The Raman signal of o-NTP on TiO2 was not detectable, owing to interference from fluorescence emissions. A molecule-to-TiO2 charge-transfer mechanism was responsible for the enhanced Raman signals observed in p-NTP and m-NTP. This transfer was due to a strong coupling between the adsorbate and the metal oxide, which led to an optically driven CT transition from the HOMO of NTP into the conduction band of TiO2. Based on the mesomeric effect, the NO2 group para to the thiol had a stronger electron-withdrawing ability than the NO2 group at the meta position. A less-efficient CT transition from p-NTP to TiO2 in the surface complex resulted in a weaker Raman-signal enhancement for p-NTP compared to m-NTP. The DFT calculation determined that the HOMO and the LUMO of NTP bound to TiO2 were located entirely on the adsorbate and the semiconductor, respectively, thereby supporting the experimental findings that a molecule-to-TiO2 mechanism was the driving force behind the observed SERS effect.  相似文献   

6.
Synthesis, Crystal Structures, Vibrational Spectra, and Normal Coordinate Analyses of the Tetrahalogeno‐bis‐Pyridine‐Osmium(III) Complexes cis ‐( n ‐Bu4N)[OsCl4Py2] and trans ‐( n ‐Bu4N)[OsX4Py2], X = Cl, Br By reaction of (n‐Bu4N)2[OsX6], X = Cl, Br, with pyridine and (n‐Bu4N)[BH4] tetrahalogeno‐bis‐pyridine‐osmium(III) complexes are formed and purified by chromatography. X‐ray structure determinations on single crystals have been performed of cis‐(n‐Bu4N)[OsCl4Py2] ( 1 ) (triclinic, space group P1, a = 9.4047(9), b = 10.8424(18), c = 17.007(2) Å, α = 71.833(2), β = 81.249(10), γ = 67.209(12)°, Z = 2), trans‐(n‐Bu4N)[OsCl4Py2] ( 2 ) (orthorhombic, space group P212121, a = 8.7709(12), b = 20.551(4), c = 17.174(4) Å, Z = 4) and trans‐(n‐Bu4N)[OsBr4Py2] ( 3 ) (triclinic, space group P1, a = 9.132(3), b = 12.053(3), c = 15.398(2) Å, α = 95.551(18), β = 94.12(2), γ = 106.529(19)°, Z = 2). Based on the molecular parameters of the X‐ray structure determinations and assuming C2 point symmetry for the anion of 1 and D2h point symmetry for the anions of 2 and 3 the IR and Raman spectra are assigned by normal coordinate analysis. The valence force constants of 1 are in the Cl–Os–Cl axis fd(OsCl) = 1.58, in the asymmetrically coordinated N′–Os–Cl · axes fd(OsCl · ) = 1.45, fd(OsN′) = 2.48, of 2 fd(OsCl) = 1.62, fd(OsN) = 2.42 and of 3 fd(OsBr) = 1.39 and fd(OsN) = 2.34 mdyn/Å.  相似文献   

7.
In the title compound, poly­[[(2,2′‐bi­pyridine‐κ2N,N′)­manganese(II)]‐μ3N‐tosyl‐l ‐glutamato‐κ4O,O′:O′′:O′′′], [Mn(tsgluo)(bipy)]n, where tsgluo is N‐tosyl‐l ‐glutamate (C12H13NO6S) and bipy is 2,2′‐bi­pyridine (C10H8N2), the Mn atoms are octahedrally coordinated by two N atoms of one bipy ligand and by four O atoms of three tsgluo2− anions. The γ‐carboxyl group coordinates to the MnII atom in a chelating mode, while the α‐carboxyl group coordinates in a bidentate–bridging mode. The complex displays a one‐dimensional double‐chain structure.  相似文献   

8.
The crystal structures of three first‐row transition metal–pyridine–sulfate complexes, namely catena‐poly[[tetrakis(pyridine‐κN)nickel(II)]‐μ‐sulfato‐κ2O:O′], [Ni(SO4)(C5H5N)4]n, (1), di‐μ‐sulfato‐κ4O:O‐bis[tris(pyridine‐κN)copper(II)], [Cu2(SO4)2(C5H5N)6], (2), and catena‐poly[[tetrakis(pyridine‐κN)zinc(II)]‐μ‐sulfato‐κ2O:O′‐[bis(pyridine‐κN)zinc(II)]‐μ‐sulfato‐κ2O:O′], [Zn2(SO4)2(C5H5N)6]n, (3), are reported. Ni compound (1) displays a polymeric crystal structure, with infinite chains of NiII atoms adopting an octahedral N4O2 coordination environment that involves four pyridine ligands and two bridging sulfate ligands. Cu compound (2) features a dimeric molecular structure, with the CuII atoms possessing square‐pyramidal N3O2 coordination environments that contain three pyridine ligands and two bridging sulfate ligands. Zn compound (3) exhibits a polymeric crystal structure of infinite chains, with two alternating zinc coordination environments, i.e. octahedral N4O2 coordination involving four pyridine ligands and two bridging sulfate ligands, and tetrahedral N2O2 coordination containing two pyridine ligands and two bridging sulfate ligands. The observed coordination environments are consistent with those predicted by crystal field theory.  相似文献   

9.
In the two ruthenium(II)–porphyrin–carbene complexes ­(di­benzoyl­carbenyl‐κC)(pyridine‐κN)(5,10,15,20‐tetra‐p‐tolyl­porphyrinato‐κ4N)­ruthenium(II), [Ru(C15H10O2)(C5H5N)(C48H36N4)], (I), and (pyridine‐κN)(5,10,15,20‐tetra‐p‐tolyl­porphyrinato‐κ4N)[bis(3‐tri­fluoro­methyl­phenyl)­carbenyl‐κC]­ruthenium(II), [Ru(C15H8F6)(C5H5N)(C48H36N4)], (II), the pyridine ligand coordinates to the octahedral Ru atom trans with respect to the carbene ligand. The C(carbene)—Ru—N(pyridine) bonds in (I) coincide with a crystallographic twofold axis. The Ru—C bond lengths of 1.877 (8) and 1.868 (3) Å in (I) and (II), respectively, are slightly longer than those of other ruthenium(II)–porphyrin–carbene complexes, owing to the trans influence of the pyridine ligands.  相似文献   

10.
Mononitrosyl and trans ‐Dinitrosyl Complexes of Phthalocyaninates of Manganese and Rhenium Tetra(n‐butyl)ammonium or di(triphenylphosphane)iminium nitrosylacidophthalocyaninato(2–)manganate, (cat)[Mn(NO)(X)pc2–] (X = ONO, NCO, N3; cat = nBu4N, PNP) is prepared from acidophthalocyaninato(2–)manganese, [Mn(X)pc2–], (cat)NO2 and (nBu4N)BH4 in CH2Cl2 or from nitrosylphthalocyaninato(2–)manganese, [Mn(NO)pc2–] and (nBu4N)X (X = ONO, NCO, N3, NCS) at T < 120 °C, respectively. [Mn(NO)(X)pc2–] dissociates in methanol, and [Mn(NO)pc2–] precipitates. Nitrito(O)phthalocyaninato(2–)manganese, (cat)NO2 and hydrogensulfide yield trans‐di(nitrosyl)phthalocyaninato(2–)manganate, trans[Mn(NO)2pc2–], isolated as red violet (PNP) and (nBu4N) complex salt. Nitrosyl(triphenylphosphane oxide)phthalocyaninato(2–)manganese, [Mn(NO)(OPPh3)pc2–] is obtained by addition of OPPh3 to [Mn(NO)pc2–] at 200 °C. Di(triphenylphosphane)phthalocyaninato(2–)rhenium(II) and (PNP)NO2 in CH2Cl2 or in molten (PNP)NO2 and PPh3 at 100 °C yields green blue l‐di(triphenylphosphane)iminium nitrosylnitrito(O)phthalocyaninato(2–)rhenate, l(PNP)[Re(NO)(ONO)pc2–]. Similarly, but with (nBu4N)NO2 red plates of tetra‐(n‐butyl)ammonium trans‐di(nitrosyl)phthalocyaninato(2–)rhenate, (nBu4N)trans[Re(NO)2pc2–] is isolated. Addition of (PNP)Br or (PNP)PF6 to a concentrated solution of (nBu4N)trans[Re(NO)2pc2–] in pyridine precipitates l(PNP)trans[Re(NO)2pc2–]. (nBu4N)trans[Re(NO)2pc2–] and PPh3 at 300 °C yield blue green nitrosyl(triphenylphosphane oxide)phthalocyaninato(2–)‐ rhenium, [Re(NO)(OPPh3)pc2–], that is oxidised with iodine precipitating nitrosyl(triphenylphosphane oxide)phthalocyaninato(2–)rhenium triiodide, [Re(NO)(OPPh3)pc2–]I3. The crystal structures of l(PNP)[Mn(NO)(ONO)pc2–] ( 1 ), l(PNP)‐ [Mn(NO)(NCO)pc2–] ( 2 ), l(PNP)trans[Mn(NO)2pc2–] ( 3 ), l(PNP)trans[Re(NO)2pc2–] ( 4 ) [Mn(NO)(OPPh3)pc2–] ( 5 ), [Re(NO)(OPPh3)pc2–] ( 6 ), and [Re(NO)(OPPh3)pc2–]I3 · CH2Cl2 ( 7 ) have been determined. The M–N(NO) distance varies between 1.623(12) Å in 5 and 1.846(3) Å in 3 . The M–N–O moiety is almost linear. The UV‐Vis spectra with the B band at ca. 14500 cm–1and the Q band at 30400 cm–1 do not dependent significantly on the axial ligand and the metal atom and its oxidation state. N–O stretching vibrations are observed in the IR spectra between 1701 cm–1 in 3 and 1753 cm–1 in [Mn(NO)pc2–] or for the Re series between 1571 cm–1 in 4 and 1724 cm–1 in 7 . M–N(NO) stretching and M–N–O deformation vibrations are assigned in the IR spectra and resonance Raman spectra between 486 cm–1 in 4 and 620 cm–1 in 1 .  相似文献   

11.
Three distinct AgI‐DMAP [DMAP = 4‐(dimethylamino)pyridine] coordination polymers [Ag2I2(DMAP)2]n ( 1 ), [Ag2(CN)2(DMAP)2.5 · DMAP]n ( 2 ), and [Ag(SCN)(DMAP)]n ( 3 ) were constructed by monatomic I, diatomic CN, and triatomic SCN bridges, respectively. 1 – 3 were determined by FT‐IR spectroscopy, elemental analyses, TGA, powder and single‐crystal X‐ray diffraction. 1 exhibits a 1D wavelike chain structure, sustained by 3‐connected I bridges, whereas 2 shows a unique 1D single‐ and double‐strand alternating chain, supported by 3‐connected CN bridges. Compound 3 has a 2D 3‐connected network architecture, fabricated by 3‐connected SCN bridges, and exhibits a (4 · 82) topology. The luminescence and nitrobenzene sensing properties of 1 – 3 were explored in 2‐propanol suspensions, which revealed that compounds 1 – 3 exhibit DMAP originated luminescence emissions and are highly sensitive for nitrobenzene detection.  相似文献   

12.
A new series of platinum(II) complexes with tridentate ligands 2,6‐bis(1‐alkyl‐1,2,3‐triazol‐4‐yl)pyridine and 2,6‐bis(1‐aryl‐1,2,3‐triazol‐4‐yl)pyridine (N7R), [Pt(N7R)Cl]X ( 1 – 7 ) and [Pt(N7R)(C?CR′)]X ( 8 – 17 ; R=n‐C4H9, n‐C8H17, n‐C12H25, n‐C14H29, n‐C18H37, C6H5, and CH2‐C6H5; R′=C6H5, C6H4‐CH3p, C6H4‐CF3p, C6H4‐N(CH3)2p, and cholesteryl 2‐propyn‐1‐yl carbonate; X=OTf?, PF6?, and Cl?), has been synthesized and characterized. Their electrochemical and photophysical properties have also been studied. Two amphiphilic platinum(II)? 2,6‐bis(1‐dodecyl‐1,2,3‐triazol‐4‐yl)pyridine complexes ( 3‐Cl and 8 ) were found to form stable and reproducible Langmuir–Blodgett (LB) films at the air/water interface. These LB films were characterized by the study of their surface‐pressure–molecular‐area (π–A) isotherms, XRD, and IR and polarized‐IR spectroscopy.  相似文献   

13.
Uniform and dense Au nanoparticles grown on Ge (Au/Ge) were fabricated by a facile galvanic displacement method and employed as surface‐enhanced Raman scattering (SERS) substrates. The substrates exhibited excellent reproducibility in the detection of rhodamine 6G aqueous solution with a relative standard deviation of <20%. The substrate showed a high Raman enhancement factor of 3.44 × 106. This superior SERS sensitivity was numerical confirmed by the three‐dimensional finite‐difference time‐domain method, which demonstrated a stronger electric field intensity (|E/E0|2) distribution around the Au nanoparticles grown on Ge. This facile and low‐cost prepared Au/Ge substrate with high SERS sensitivity and reproducibility might have potential applications in monitoring in situ reaction in aqueous solution. Copyright © 2014 John Wiley & Sons, Ltd.  相似文献   

14.
Nucleophilic incorporation of [18F]F? under aqueous conditions holds several advantages in radiopharmaceutical development, especially with the advent of complex biological pharmacophores. Sulfonyl fluorides can be prepared in water at room temperature, yet they have not been assayed as a potential means to 18F‐labelled biomarkers for PET chemistry. We developed a general route to prepare bifunctional 4‐formyl‐, 3‐formyl‐, 4‐maleimido‐ and 4‐oxylalkynl‐arylsulfonyl [18F]fluorides from their sulfonyl chloride analogues in 1:1 mixtures of acetonitrile, THF, or tBuOH and Cs[18F]F/Cs2CO3(aq.) in a reaction time of 15 min at room temperature. With the exception of 4‐N‐maleimide‐benzenesulfonyl fluoride ( 3 ), pyridine could be used to simplify radiotracer purification by selectively degrading the precursor without significantly affecting observed yields. The addition of pyridine at the start of [18F]fluorination (1:1:0.8 tBuOH/Cs2CO3(aq.)/pyridine) did not negatively affect yields of 3‐formyl‐2,4,6‐trimethylbenzenesulfonyl [18F]fluoride ( 2 ) and dramatically improved the yields of 4‐(prop‐2‐ynyloxy)benzenesulfonyl [18F]fluoride ( 4 ). The N‐arylsulfonyl‐4‐dimethylaminopyridinium derivative of 4 ( 14 ) can be prepared and incorporates 18F efficiently in solutions of 100 % aqueous Cs2CO3 (10 mg mL?1). As proof‐of‐principle, [18F] 2 was synthesised in a preparative fashion [88(±8) % decay corrected (n=6) from start‐of‐synthesis] and used to radioactively label an oxyamino‐modified bombesin(6–14) analogue [35(±6) % decay corrected (n=4) from start‐of‐synthesis]. Total preparation time was 105–109 min from start‐of‐synthesis. Although the 18F‐peptide exhibited evidence of proteolytic defluorination and modification, our study is the first step in developing an aqueous, room temperature 18F labelling strategy.  相似文献   

15.
Surface‐enhanced Raman scattering (SERS) is a process with origins, electromagnetic and chemical. The electromagnetic enhancement consists of the excitation of surface plasmons in the metallic support of the thin film. With only the electromagnetic enhancement mechanism, the surface spectra should not differ from volume Raman spectra. However, between SERS and volume Raman spectra, there are differences resulting from the chemical reactions taking place at the polymer/metal interface, intermediated by solvent molecules, that finally depend on the types of polymers and metallic supports. Polyaniline (PAN) is an excellent material to emphasize the chemical component of SERS. This is due to its particular structure with a repeating unit that contains two entities at different weights—a reduced state and an oxidized state–that, in turn, react differently with a metallic substrate. SERS spectra depend on the oxidizing properties of the metal surface, which involves an intermediate compound of the types Ag2O and Au2O3 when N‐methyl‐2‐pyrrolidinone is used as the solvent. This article presents new results concerning the surface chemical effects that produce variations of the PAN SERS spectra. The SERS spectra of the PAN emeraldine base (PAN‐EB) layered on Au support are characterized by a semiquinoid structure that we believe is induced on the intermediate compound Au2O3. In the presence of H2SO4, the SERS spectra change gradually as the degree of acid protonation doping increases. The SERS spectra of the fully protonated PAN‐EB are identical to those obtained on PAN emeraldine salt (PAN‐ES) synthesized by cyclic voltammetry in an acid medium and are invariable with the type of metallic support. The SERS spectra show that the emeraldine salt can be partially or totally deprotonated with water or NH4OH. The deprotonation is complete for the Ag support and partial for the Au support. The SERS spectra of the fully protonated PAN‐EB are characterized by a double band with maxima at about 1330 and 1370 cm−1. Although the generation process of positive charge on the macromolecular chain of PAN‐EB doped in the presence of (C4H9)4NBF4 is similar to that due to protonic acid doping, involving cation addition (C4H or H+ ions, respectively) in SERS spectra, the complex band situated at about 1330–1370 cm−1 no longer appears. The doping of PAN‐EB with FeCl3 produces two polymer forms: a salt type characterized by a protonated structure similar to that found for PAN‐ES and a base type similar to the leucoemeraldine form. © 2000 John Wiley & Sons, Inc. J Polym Sci B: Polym Phys 38: 2599–2609, 2000  相似文献   

16.
Hydrocarbon‐bridged Metal Complexes. L Dicarbonyl Cyclopentadienyl Pyridoyl Iron Complexes as Ligands Dicarbonyl‐cyclopentadienyl‐2‐ and 3‐pyridoyl‐iron (L1, L2) and 2,6‐dicarbonyl‐pyridine‐bis(dicarbonyl‐cyclopentadienyl‐iron) (L3) function as ligands in metal complexes and the N,O‐chelates [(OC)4M(L1)] (M = Mo, W, 8 a, b ) and [(Ph3P)2Cu(L1)]+BF4 ( 9 ) were prepared. Monodentate coordination of L1 and L2 through the pyridine N‐atom occurs in the palladium(II) complexes [Cl2Pd(PnBu3)(L1)] ( 10 ), [Cl2Pd(PnBu3)(L2)] ( 11 ) and [Cl2Pd(L2)2] ( 12 ). Ligand L3 forms the O,N,O‐bis(chelate) [Cl2Zn(L3)] ( 13 ). The crystal and molecular structures of L1, 8 b (M = W), 9–11 and 13 were determined by X‐ray diffraction.  相似文献   

17.
Synthesis, Crystal Structures, Vibrational Spectra, and Normal Coordinate Analyses of the mer ‐Trihalogeno‐tris‐Pyridine‐Osmium(III) Complexes mer‐[OsX3Py3], X = Cl, Br, I By reaction of the hexahalogenoosmates(IV) with pyridine and iso‐amylalcohol mer‐trihalogeno‐tris‐pyridine‐osmium(III) complexes are formed and purified by chromatography. X‐ray structure determinations on single crystals have been performed of mer‐[OsBr3Py3] (monoclinic, space group P21/n, a = 9.098(5), b = 12.864(5), c = 15.632(5) Å, β = 90.216(5)°, Z = 4) and mer‐[OsI3Py3] (monoclinic, space group P21/n, a = 9.0952(17), b = 13.461(4), c = 15.891(10), β = 91.569(5)°, Z = 4). The pyridine rings are twisted propeller‐like against the N3 meridional plane with mean angles of 49° (Cl), 46° (Br), 44° (I). Based on the molecular parameters of the X‐ray structure determinations and assuming C2 point symmetry, the IR and Raman spectra are assigned by normal coordinate analysis. Due to the stronger trans influence of pyridine as compared with the halide ligands for N'–Os–X · axes significantly different valence force constants are observed in comparison with symmetrically coordinated octahedron axes: fd(OsCl) = 1.74, fd(OsCl·) = 1.49, fd(OsBr) = 1.43, fd(OsBr · ) = 1.18, fd(OsI) = 0.99, fd(OsI · ) = 0.96, fd(OsN) between 1.96 and 2.07 and fd(OsN') between 2.13 and 2.32 mdyn/Å.  相似文献   

18.
Gold phosphides show unique optical or semiconductor properties and there are extensive high technology applications, e.g. in laser diodes, etc. In spite of the various AuP structures known, the search for new materials is wide. Laser ablation synthesis is a promising screening and synthetic method. Generation of gold phosphides via laser ablation of red phosphorus and nanogold mixtures was studied using laser desorption ionisation time‐of‐flight mass spectrometry (LDI TOFMS). Gold clusters Aum+ (m = 1 to ~35) were observed with a difference of one gold atom and their intensities were in decreasing order with respect to m. For Pn+ (n = 2 to ~111) clusters, the intensities of odd‐numbered phosphorus clusters are much higher than those for even‐numbered phosphorus clusters. During ablation of P‐nanogold mixtures, clusters Aum+ (m = 1‐12), Pn+ (n = 2‐7, 9, 11, 13–33, 35–95 (odd numbers)), AuPn+ (n = 1, 2–88 (even numbers)), Au2Pn+ (n = 1‐7, 14–16, 21–51 (odd numbers)), Au3Pn+ (n = 1‐6, 8, 9, 14), Au4Pn+ (n = 1‐9, 14–16), Au5Pn+ (n = 1‐6, 14, 16), Au6Pn+ (n = 1‐6), Au7Pn+ (n = 1‐7), Au8Pn+ (n = 1‐6, 8), Au9Pn+ (n = 1‐10), Au10Pn+ (n = 1‐8, 15), Au11Pn+ (n = 1‐6), and Au12Pn+ (n = 1, 2, 4) were detected in positive ion mode. In negative ion mode, Aum (m = 1–5), Pn (n = 2, 3, 5–11, 13–19, 21–35, 39, 41, 47, 49, 55 (odd numbers)), AuPn (n = 4–6, 8–26, 30–36 (even numbers), 48), Au2Pn (n = 2–5, 8, 11, 13, 15, 17), Au3Pn (n = 6–11, 32), Au4Pn (n = 1, 2, 4, 6, 10), Au6P5, and Au7P8 clusters were observed. In both modes, phosphorus‐rich AumPn clusters prevailed. The first experimental evidence for formation of AuP60 and gold‐covered phosphorus Au12Pn (n = 1, 2, 4) clusters is given. The new gold phosphides generated might inspire synthesis of new Au‐P materials with specific properties. Copyright © 2012 John Wiley & Sons, Ltd.  相似文献   

19.
We report on the structural, electronic, and magnetic properties of manganese‐doped silicon clusters cations, SinMn+ with n=6–10, 12–14, and 16, using mass spectrometry and infrared spectroscopy in combination with density functional theory computations. This combined experimental and theoretical study allows several structures to be identified. All the exohedral SinMn+ (n=6–10) clusters are found to be substitutive derivatives of the bare Sin+1+ cations, while the endohedral SinMn+ (n=12–14 and 16) clusters adopt fullerene‐like structures. The hybrid B3P86 functional is shown to be appropriate in predicting the ground electronic states of the clusters and in reproducing their infrared spectra. The clusters turn out to have high magnetic moments localized on Mn. In particular the Mn atoms in the exohedral SinMn+ (n=6–10) clusters have local magnetic moments of 4 μB or 6 μB and can be considered as magnetic copies of the silicon atoms. Opposed to other 3d transition‐metal dopants, the local magnetic moment of the Mn atom is not completely quenched when encapsulated in a silicon cage.  相似文献   

20.
The anionic polymerization behaviors of ethynylstyrene derivatives containing isomeric pyridine moieties, 2‐(2‐(4‐vinylphenyl)ethynyl)pyridine ( A ), 3‐(2‐(4‐vinylphenyl)ethynyl)pyridine ( B ), and 4‐(2‐(4‐vinylphenyl)ethynyl)pyridine ( C ), were investigated in the identical conditions. The anionic polymerization of A – C was performed with (diphenylmethyl)potassium (Ph2CHK) in tetrahydrofuran (THF) at ?78 °C. The polymerization of A proceeded quantitatively at –78 °C for 4 h, and the resulting poly( A ) possessed predictable molecular weights (Mn = 3300–68,500) and narrow molecular weight distributions (MWDs) (Mw/Mn = 1.04–1.11). In contrast, the anionic polymerization of B was not performed at –78 °C for 4 h due to the occurrence of side reactions. The monomer B was quantitatively recovered after the reaction. In the polymerization of C performed at –78 °C for 6 h, observed Mn values of the resulting poly( C ) were in good agreement with calculated molecular weights based on monomer to initiator ratios, but the MWDs were somewhat broad (Mw/Mn = 1.23–1.31). To estimate the reactivity of A and to characterize its living nature, the block copolymerization of A with 2‐vinylpyridine (2VP) and methyl methacrylate (MMA) was performed. The well‐defined block copolymers, poly(2VP)‐b‐poly( A ) and poly( A )‐b‐poly(MMA), were successfully synthesized without any additives. © 2011 Wiley Periodicals, Inc. J Polym Sci Part A: Polym Chem, 2011  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