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 750 毫秒
1.
The metallacarborane [3,3′‐Co(1,2‐closo‐C2B9H11)2]? has been synthesized. This species allows the formation of redox couples in which both partners are negatively charged. The E1/2 potential can be tuned by adjusting the nature and number of substituents on B and C. The octaiodinated species [3,3′‐Co(1,2‐closo‐C2B9H7I4)2]? is the most favorable, as it is isolatable and stable in air. A DFT study on stability and redox potentials of complexes has been performed.  相似文献   

2.
Zwitterionic 4,8,8′-exo-{Ph3PCu}-4,8,8′-(μ-H)3-commo-3,3′-Co(1,2-C2B9H9)-(1′,2′-C2B9H10) and ionic [(PPh3)3Cu][commo-3,3′-Co(1,2-C2B9H11)2 complexes were synthesized in moderate yields by the reaction of anionic commo-complex [Cs][commo-3,3′-Co-(1,2-C2B9H11)2]) in a CH2Cl2 solution with anhydrous CuCl2 or CuCl in the presence of PPh3. The complexes were also synthesized by alternative methods and characterized by NMR and X-ray diffraction methods.  相似文献   

3.
In the title compound, [Rh(C2H11B9)(NO3)(C18H15P)2]·2.2CH2Cl2, studied as a 2.2‐solvate of what was assumed to be dichloromethane, the nitrate ligand lies cis with respect to both cage C atoms. Accordingly, the compound displays a pronounced preferred exopolyhedral ligand orientation (ELO) which is traced to both the greater trans influence of the cage B over the cage C atoms and the greater trans influence of the triphenylphosphane ligands over the nitrate ligand. The overall molecular architecture therefore agrees with that of a number of similar 3‐L‐3,3‐L2‐3,1,2‐closoMC2B9H11 species in the literature.  相似文献   

4.
The analysis of complex mixtures of chlorinated paraffins (CPs) with short (SCCPs, C10–C13) and medium (MCCPs, C14–C17) chain lengths can be disturbed by mass overlap, if low resolution mass spectrometry (LRMS) in the electron capture negative ionization mode is employed. This is caused by CP congeners with the same nominal mass, but with five carbon atoms more and two chlorine atoms less; for example C11H1737Cl35Cl6 (m/z 395.9) and C16H2935Cl5 (m/z 396.1). This can lead to an overestimation of congener group quantity and/or of total CP concentration. The magnitude of this interference was studied by evaluating the change after mixing a SCCP standard and a MCCP standard 1+1 (S+MCCP mixture) and comparing it to the single standards. A quantification of the less abundant C16 and C17 congeners present in the MCCP standard was not possible due to interference from the major C11 and C12 congeners in the SCCPs. Also, signals for SCCPs (C10–C12) with nine and ten chlorine atoms were mimicked by MCCPs (C15–C17) with seven and eight chlorine atoms (for instance C10H12Cl10 by C15H24Cl8). A similar observation was made for signals from C15–C17 CPs with four and five chlorine atoms resulting from SCCPs (C10–C12) with six and seven chlorine atoms (such as C15H28Cl4 by C10H16Cl6) in the S+MCCP mixture. It could be shown that the quantification of the most abundant congeners (C11–C14) is not affected by any interference. The determination of C10 and C15 congeners is partly disturbed, but this can be detected by investigating isotope ratios, retention time ranges and the shapes of the CP signals. Also, lower chlorinated compounds forming [M+Cl] as the most abundant ion instead of [M-Cl] are especially sensitive to systematic errors caused by superposition of ions of different composition and the same nominal mass.  相似文献   

5.
Two‐electron reduction of 1,1′‐bis(o‐carborane) followed by reaction with [Ru(η‐mes)Cl2]2 affords [8‐(1′‐1′,2′‐closo‐C2B10H11)‐4‐(η‐mes)‐4,1,8‐closo‐RuC2B10H11]. Subsequent two‐electron reduction of this species and treatment with [Ru(η‐arene)Cl2]2 results in the 14‐vertex/12‐vertex species [1‐(η‐mes)‐9‐(1′‐1′,2′‐closo‐C2B10H11)‐13‐(η‐arene)‐1,13,2,9‐closo‐Ru2C2B10H11] by direct electrophilic insertion, promoted by the carborane substituent in the 13‐vertex/12‐vertex precursor. When arene=mesitylene (mes), the diruthenium species is fluxional in solution at room temperature in a process that makes the metal–ligand fragments equivalent. A unique mechanism for this fluxionality is proposed and is shown to be fully consistent with the observed fluxionality or nonfluxionality of a series of previously reported 14‐vertex dicobaltacarboranes.  相似文献   

6.
A novel compound, [MnPhen3][(B9C2H11)Co(B8C2H10)Co(B9C2H11)]· CH3CN (Phen = 1,10-phenantroline), comprising a Co(III) dicobaltacarborane cluster anion has been prepared and characterized by single crystal X-ray diffraction. Crystal data are the following: C44H59B26N7Co2Mn, M = 1139.84, triclinic, space group , unit cell parameters: a = 13.2465(11) Å, b = 14.521(2) Å, c = 15.2536(15) Å; α = 77.027(9)°, β = 88.500(8)°, γ = 77.274(9)°; V = 2788.5(5) Å3, Z = 2, d calc = 1.358 g/cm3, T = 295 K, F(000) = 1162, μ = 0.853 mm−1. The structure was solved by the direct and Fourier methods and refined anisotropically (isotropically for hydrogen atoms) using the full-matrix technique to final factors R 1 = 0.0374, wR 2 = 0.0915 for 7397 I hkl ≥2σI of 9779 I hkl measured (diffractometer Enraf-Nonius CAD-4, λMoK α , graphite monochromator, θ/2θ-scanning). The structure is formed from [MnPhen3]2+ cations, [(B9C2H11)×Co(B8C2H10)Co(B9C2H11)]2− anions, and acetonitrile molecules CH3CN. Central Mn atom in the cation has a distorted octahedral coordination environment formed by six nitrogen atoms of three bi-dentate Phen ligands, average Mn-N bond length being 2.263(2) Å. The anion has a chain-like structure built from three icosahedra sharing common vertices occupied by the cobalt atoms. The central icosahedron including ten light atoms (8B, 2C) provides two vertices for the cobalt atoms shared with the other icosahedra having 11 light atoms (9B, 2C). The arrangement of-C2-groups in the anion corresponds to a quasi-gauche-configuration of asymmetric sandwich complexes of both cobalt atoms. Original Russian Text Copyright ? 2005 by T. M. Polyanskaya, V. V. Volkov, and M. K. Drozdova __________ Translated from Zhurnal Strukturnoi Khimii, Vol. 46, No. 4, pp.730–740, July–August, 2005.  相似文献   

7.
A simple method for the functionalization of closo‐borates [closo‐B10H10]2? ( 1 ), [closo‐1‐CB9H10]? ( 2 ), [closo‐B12H12]2? ( 3 ), [closo‐1‐CB11H12]? ( 4 ), and [3,3′‐Co(1,2‐C2B9H11)2]? ( 5 ) is described. Treatment of the anions and their derivatives with ArI(OAc)2 gave aryliodonium zwitterions, which were sufficiently stable for chromatographic purification. The reactions of these zwitterions with nucleophiles provided facile access to pyridinium, sulfonium, thiol, carbonitrile, acetoxy, and amino derivatives. The synthetic results are augmented by mechanistic considerations.  相似文献   

8.
Halogenation of nido-B10H14 with C2H2Cl4, C2Cl6, Br2, or I2, produces by cluster degradation the (2 n)-closo-clusters B9X9 (X = Cl, Br, I). The synthesis of salts of the perhalogenated radical anions of the type (2 n + 1)-closo-[B9X9]· – and of the corresponding dianions (2 n + 2)-closo-[B9X9]2– from neutral B9X9 is described [n is the number of cluster atoms; (2 n), (2 n + 1), and (2 n + 2) is the number of cluster electrons]. Molecular and crystal structures of B9Cl9, B9Br9, [(C6H5)4P][B9Br9] · CH2Cl2, and [(C4H9)4N]2[B9Br9] · CH2Cl2 have been determined via X-ray diffraction. All three oxidation states of the cluster retain the tricapped trigonal prism. The reduction of the clusters B9X9 was shown by cyclic voltammetry in CH2Cl2 to proceed via two successive one-electron reversible steps, separated by at least 0.4 V. The paramagnetic radical anions [B9X9]· – (X = Cl, Br) were further characterized by magnetic susceptibility measurements of [Cp2Fe][B9X9] and [Cp2Co][B9X9], respectively. The EPR spectra of [B9X9]· – (X = Cl, Br, I) in glassy frozen CH2Cl2 solutions showed increasing g anisotropy for the heavier halogen derivatives, illustrating significant halogen participation at the singly occupied MO. The 11B NMR spectra of CD2Cl2 solutions of the neutral clusters B9X9 exhibit only one sharp resonance, indicating that the boron atoms are highly fluxional in solution. In contrast, two different boron resonances as expected for a rigid tricapped trigonal prism are clearly observed for the [B9X9]2– dianions in solutions and for solid B9Br9 in the 11B MAS NMR spectra. Temperature dependent 11B MAS NMR experiments on B9Br9 and [B9Br9]2– in the solid state show a reversible coalescence of the two resonances at higher temperature. 11B MAS NMR spectra and DTA measurements of [B9Br9]2– showed a phase transition.  相似文献   

9.
o‐Carborane (C2B10H12) was adapted to perform as the core of globular macromolecules, dendrons or dendrimers. To meet this objective, precisely defined substitution patterns of terminal olefin groups on the carborane framework were subjected to Heck cross‐coupling reactions or hydroboration leading to hydroxyl terminated arms. These led to new terminal groups (chloro, bromo, and tosyl leaving groups, organic acid, and azide) that permitted ester production, click chemistry, and oxonium ring opening to be performed as examples of reactions that demonstrate the wide possibilities of the globular icosahedral carboranes to produce new dendritic or dendrimer‐like structures. Polyanionic species were obtained in high yield through the ring‐opening reaction of cyclic oxonium compound [3,3′‐Co(8‐C4H8O2‐1,2‐C2B9H10)(1′,2′‐C2B9H11)] by using terminal hydroxyl groups as nucleophiles. These new polyanionic compounds that contain multiple metallacarborane clusters at their periphery may prove useful as new classes of compounds for boron neutron capture therapy with enhanced water solubility and as cores to make a new class of high‐boron globular macromolecules.  相似文献   

10.
Crystallization of chloro­(2,2′:6′,2′′‐terpyridine)platinum(II) chloride from dimethyl sulfoxide yields a red polymorph, [PtCl(C15H11N3)]Cl·C2H6OS, (I), which exhibits stacking along the a axis through pairs of Pt⋯Pt(−x, −y, −z) inter­actions of 3.3155 (8) Å. The cations are further associated through close Pt⋯Pt(1 − x, −y, −z) distances of 3.4360 (8) Å. Recrystallization from water gives a mero­hedrally twinned yellow–orange dihydrate form, [PtCl(C15H11N3)]Cl·2H2O, (II), with pairwise short Pt⋯Pt(1 − x, 2 − y, −z) contacts of 3.3903 (5) Å but no long‐range stacking through the crystals. Inter­pair Pt⋯Pt(−x, 2 − y, −z) distances between cation pairs in the hydrate are 4.3269 (5) Å.  相似文献   

11.
A series of cobaltacarborane complexes including Cs[(1,2-C2B9H11)2Co], Cs[(8,9,12-Br3-C2B9H8)2Co], Cs[(8-I-C2B9H10)2Co], and polynuclear chain cobaltacarborane complexes of the general formula Csn[(1,2-C2B9H11)2Con(1,2-C2B8H10)n-1], where n=2,3,4,5,6,7, were studied by Raman spectroscopy. In the Raman spectra of these complexes, the frequencies and intensities of the bands of the ligand-metal-ligand totally symmetric stretching vibrations are characteristic; therefore, these data may be used as analytical characteristics. Institute of Inorganic Chemistry, Siberian Branch, Russian Academy of Sciences. Translated fromZhurnal Strukturnoi Khimii, Vol. 37, No. 4, pp. 716–720, July–August, 1996. Translated by L. Smolina  相似文献   

12.
Hydridorhodacarboranes 3,3-(Ph2RP)2-3-H-3,1,2-RhC2B9H11−n F n (R=Ph, Me;n=1, 2, 4) containing F atoms at the B atoms of the π-carborane ligand were synthesized from (Ph3P)3RhCl or (Ph2MeP)3RhCl andnido-7,8-C2B9H12−n F n (n=1, 2, 4) salts. Hydridorhodacarboranes 3,3-(Ph2MeP)2-3-H-3,1,2-RhC2B9H11−n F n readily exchange the H atom at the Rh atom for the Cl atom under the action of CH2Cl2 to give 3,3-(Ph2MeP)2-3-Cl-3,1,2-RhC2B9H11−n F n . The structures of the 3,3-(Ph3P)2-3-H-3,1,2-RhC2B9H7F4 and 3,3-(Ph2MeP)2-3-Cl-3,1,2-RhC2B9H9F2 complexes were determined by X-ray diffraction analysis. Catalytic properties of the rhodacarbonanes obtained in hydrosilylation of styrene and phenylacetylene by PhMe2SiH were studied. Translated fromIzvestiya Akademii Nauk. Seriya Khimicheskaya, No. 3, pp. 570–578, March, 1997.  相似文献   

13.
The 35Cl-NQR spectra of 45 chlorosubsitituted acetanilides, ClxC6H5?xNHCOCH3?yCly, were investigated and the temperature dependence of some spectra, especially of monochloroacetic acid derivatives, was measured. A preliminary assignment of the NQR frequencies is given. A correlation between NQR frequencies and substituent parameters permits the study of the substituent effect of the acetamido group, ? NHCOCH3?yCly. The chloro-substitution in the side chain of the acetanilides seems to have no noticeable influence on the 35Cl-NQR frequencies of the chlorine atoms at the benzene ring. The NQR frequencies of the chlorine atoms in the chloroacetamido group are, on the other hand, insensitive to substitutions at the benzene nucleus. The possibility of steric influences on the NQR spectrum of ortho-chloro-substituted acetanilides is discussed. The investigation further confirms that a crystal field effect of about ±500 kHz must be considered in the interpretation of NQR spectra of chlorobenzene derivatives.  相似文献   

14.
Icosahedral metallacarboranes are θ-shaped anionic molecules in which two icosahedra share one vertex that is a metal center. The most remarkable of these compounds is the anionic cobalt-based metallacarborane [Co(C2B9H11)2], whose oxidation-reduction processes occur via an outer sphere electron process. This, along with its low density negative charge, makes [Co(C2B9H11)2] very appealing to participate in electron-transfer processes. In this work, [Co(C2B9H11)2] is tethered to a perylenediimide dye to produce the first examples of switchable luminescent molecules and materials based on metallacarboranes. In particular, the electronic communication of [Co(C2B9H11)2] with the appended chromophore unit in these compounds can be regulated upon application of redox stimuli, which allows the reversible modulation of the emitted fluorescence. As such, they behave as electrochemically-controlled fluorescent molecular switches in solution, which surpass the performance of previous systems based on conjugates of perylendiimides with ferrocene. Remarkably, they can form gels by treatment with appropriate mixtures of organic solvents, which result from the self-assembly of the cobaltabisdicarbollide-perylendiimide conjugates into 1D nanostructures. The interplay between dye π-stacking and metallacarborane electronic and steric interactions ultimately governs the supramolecular arrangement in these materials, which for one of the compounds prepared allows preserving the luminescent behavior in the gel state.  相似文献   

15.
The effect of an alkyl substituted in the aromatic ring of the salen ligand on the polymerization of butadiene (Bd) with (salen)Co(II) complexes in combination with methylaluminoxane (MAO) was investigated. The activity for the polymerization of Bd was influenced significantly by the introduction of alkyl groups at the 3,3′,5,5′‐positions in the aromatic ring of the salen ligand, and both the polymerization rate and 1,4‐cis contents increased in the following order with respect to the alkyl group: H < CH3 < t‐C4H9. This is in good agreement with the bulkiness of the alkyl groups. The activity for the polymerization of the (salen)Co(II) complex possessing t‐C4H9 at the 3,3′‐positions was higher than that of the (salen)Co(II) bearing t‐C4H9 at the 5,5′‐positions. Thus, the introduction of bulky substituents at the 3,3′‐positions of the salen ligand was an important factor in achieving both high activity and high 1,4‐cis selectivity in the polymerization of Bd with (salen)Co(II) complexes in combination with MAO. © 2006 Wiley Periodicals, Inc. J Polym Sci Part A: Polym Chem 44: 4088–4094, 2006  相似文献   

16.
The title compound, [Co(C12H6N2O4)(H2O)2]n, has been hydro­thermally synthesized and structurally characterized. It consists of polymeric chains of [Co{μ‐(2,2′‐bipyridyl‐3,3′‐di­carboxyl­ato‐κ4N,N′:O,O′)}(H2O)2] units, in which each CoII cation is octahedrally coordinated by two chelating pyridyl N atoms, two chelating carboxyl O atoms from different carboxylate groups of another bipyridyl ligand, and two water mol­ecules as terminal ligands. A crystallographic twofold axis parallel to the chain axis, passes through the Co atom.  相似文献   

17.
In the title coordination polymer, {[Cd(C6H8O4S)(C13H14N2)]·H2O}n, the CdII atom displays a distorted octahedral coordination, formed by three carboxylate O atoms and one S atom from three different 3,3′‐thiodipropionate ligands, and two N atoms from two different 4,4′‐(propane‐1,3‐diyl)dipyridine ligands. The CdII centres are bridged through carboxylate O atoms of 3,3′‐thiodipropionate ligands and through N atoms of 4,4′‐(propane‐1,3‐diyl)dipyridine ligands to form two different one‐dimensional chains, which intersect to form a two‐dimensional layer. These two‐dimensional layers are linked by S atoms of 3,3′‐thiodipropionate ligands from adjacent layers to form a three‐dimensional network.  相似文献   

18.
7‐Ethyl‐10‐hydroxycamptothecin [systematic name: (4S)‐4,11‐diethyl‐4,9‐dihydroxy‐1H‐pyrano[3′,4′:6,7]indolizino[1,2‐b]quinoline‐3,14(4H,12H)‐dione, SN‐38] is an antitumour drug which exerts activity through the inhibition of topoisomerase I. The crystal structure of SN‐38 as the monohydrate, C22H20N2O5·H2O, reveals that it is a monoclinic crystal, with one SN‐38 molecule and one water molecule in the asymmetric unit. When the crystal is heated to 473 K, approximately 30% of SN‐38 is hydrolyzed at its lactone ring, resulting in the formation of the inactive carboxylate form. The molecular arrangement around the water molecule and the lactone ring of SN‐38 in the crystal structure suggests that SN‐38 is hydrolyzed by the water molecule at (x, y, z) nucleophilically attacking the carbonyl C atom of the lactone ring at (x − 1, y, z − 1). Hydrogen bonding around the water molecules and the lactone ring appears to promote this hydrolysis reaction: two carbonyl O atoms, which are hydrogen bonded as hydrogen‐bond acceptors to the water molecule at (x, y, z), might enhance the nucleophilicity of this water molecule, while the water molecule at (−x, y + , −z), which is hydrogen bonded as a hydrogen‐bond donor to the carbonyl O atom at (x − 1, y, z − 1), might enhance the electrophilicity of the carbonyl C atom.  相似文献   

19.
The crystal structures of 3,3‐di­methyl‐3‐(tri­chloro­germyl)­propionic acid, [Ge(C5H9O2)Cl3], 3,3‐di­methyl‐3‐(tri­phenyl­germyl)­propionic acid, [Ge(C6H5)3(C5H9O2)], and 3,3‐di­methyl‐3‐(tri‐p‐toly­lgermyl)­propionic acid, [Ge(C7H7)3(C5H9O2)], have slightly distorted tetrahedral geometries about the Ge atoms. All the structures form dimers via strong O—H·O hydrogen bonds, resulting in eight‐membered rings that can be best described in terms of graph‐set notation (8).  相似文献   

20.
The redox aptitude of a series of cobalt(III) or cobalt(I) sandwich complexes bearing a charge compensated dicarbollide ligand ([9-L-7,8-C2B9H10]) as a constant unit and different counterparts (varying from classical [7,8-C2B9H11]2− to charge-compensated [9-L-7,8-C2B9H10] dicarbollides, from cyclopentadienyl [C5R5] (R = Me, H) to cyclobutadiene [C4Me4]0 ligands) has been studied. All the Co(III) complexes display the reversible sequence Co(III)/Co(II)/Co(I). In contrast, the Co(I) complexes (namely, those capped by tetramethylcyclobutadiene) accede reversibly only to the Co(II) oxidation state, the passage to Co(III) being irreversible. When possible, the Co(II) intermediates have been characterized by EPR spectroscopy. The molecular structures of the monocation [Co(η-9-SMe2-7,8-C2B9H10)2]+ in its DD/LL and meso diastereomeric forms as well as that of heteroleptic (η-7,8-C2B9H11)Co(η-9-SMe2-7,8-C2B9H10) have been obtained by single-crystal diffraction. Presented at the 3rd Chianti Electrochemistry Meetings July 3−9, 2004, Certosa di Pontignano, Italy  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