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.
A predominantly localized electron pair scheme is outlined for describing the electron distribution and bonding in closo borane anions BnHn2− and related electron deficient deltahedral clusters, in which a skeletal electron pair is assigned to each vertex, one pair being regarded as delocalized just inside the roughly spherical surface on which the skeletal atoms lie. The scheme gives a clearer picture of the electron distribution than is conveyed by resonating 2- and 3-centre bonds in the polyhedron edges and faces, and allows the bond orders of the polyhedron edge links to be calculated readily. The consequence of formal removal of BH2+ units from closo species BnHn2− to generate nido species Bn−1Hn−14− and arachno species Bn−2Hn−26− is explored, and seen to allow rationalization of two features of such deltahedral-fragment clusters: (i) why a high-connectivity vertex is left vacant and (ii) why the frontier orbitals of such species concentrate electronic charge around their open faces. Moreover, in the case of D4‘h B4H46− (cf. C4H42−) and D5h B5H56− (cf. C5H5), the approach leads directly to the familiar picture for aromatic ring systems in which the highest filled, doubly degenerate π-bonding molecular orbital concentrates electronic charge in rings above and below the polygon on which the skeletal nuclei lie. It also leads to the expectation that arachno clusters with non-adjacent vacant vertices will be more stable than those with adjacent vacant vertices.  相似文献   

2.
A new approach in the synthesis of water-soluble boron-rich compounds was proposed. The closo-dodecaborate cage is used as a hydrophilic substitutent providing for the water-solubility of the molecule whereas the carborane cage can be used for attachment to biomolecules using earlier developed methods. The double-cage molecules [o-, m-, and p-CB10H10C(CH2)4OB12H11]2− were prepared by the reaction of the tetramethylene oxonium derivative of the closo-dodecaborate anion, [B12H11O(CH2)4], with the corresponding lithiated carboranes. The compounds obtained have doubled the boron contents and could serve for the synthesis of agents for boron neutron capture therapy (BNCT).  相似文献   

3.
Ultra-soft X-ray fluorescence spectra of ortho- and meta-carborane C2B10H12 were obtained. Ab initio self-consistent field (SCF) quantum-chemical calculations of these molecules were performed to interpret BK and CK spectra. Distinctions between electronic structure of closo-carboranes 1,2- and 1,7-C2B10H12 are caused by different efficiency in the interaction of carbon and boron atoms. Location of boron atom between carbon atoms leads to stronger delocalization of electron density in meta-carborane molecule. The correlation between molecular orbitals (MOs) of the anion B12H122− and the closo-carboranes was carried out.  相似文献   

4.
Thermolysis of [arachno-4-SB8H12] (1) in boiling cyclohexane gives two isomers 2 and 3 of 18-vertex [S2B16H16], together with known 12-vertex [closo-1-SB11H11] (4) and known 11-vertex [nido-7-SB10H12] (5). Compounds 2 and 3 are characterised by single-crystal X-ray diffraction analyses and single- and double-resonance 11B- and 1H-NMR spectroscopy. The [n-S2B16H16] isomer 2 takes the form of nido ten-vertex: nido ten-vertex [anti-B18H22] with the 9 and 9′ positions occupied by S vertices, whereas the [iso-S2B16H16] isomer 3 takes the form of a nido 11-vertex {SB10} subcluster fused via a common two-boron edge to a nido-type {B8} subcluster that is additionally linked exo to the {SB10} subcluster by a bridging S atom that is held endo to the {B8} unit. Isomer 2 is readily deprotonated and its monoanion 6 is characterised by NMR spectroscopy and by a single-crystal X-ray diffraction analysis of its [tmndH]+[n-S2B16H15] salt 6b; deprotonation has occurred from an open-face B---H---B bridging site.  相似文献   

5.
Preparation and characterization of the first derivatives of the fused macropolyhedral anion [B22H22]2− are reported. The species [B22H21OH]2− (1) and [B22H21OEt]2− (2) are obtained from workup of the products of the reaction between HgBr2 and [NBzlEt3]2[B22H22]; a cluster involving the conjoining of a closo-B12 icosahedron with a nido-B10 cluster. Washing the products with ethanol followed by thin-layer chromatography allows the isolation of 1 and 2, reproducibly, in yields of 27 and 20%, respectively. The species were characterized by NMR spectroscopy, elemental analysis and X-ray diffraction studies. The crystal structure determinations of the two species identify novel features. Apparently the influence of the O atoms in the ions [B22H21OH]2− and [B22H21OEt]2− results in the lengthening of what was a gunwale B---B connection adjacent to the junction of the two cages such that the distances are 2.180 and 2.230 Å, respectively. These latter are longer than the corresponding distance in the parent species [B22H22]2−, which is 2.09 Å; quite long for a normal B---B distance. Thus it is assumed that these B atoms, in 1 and 2, one of which bears the substituent, are not bonded to each other.  相似文献   

6.
Closo-BnHn−2(CO)2 (n = 5–12), isolobal analogues of closo-C2Bn−2Hn, have been investigated at the B3LYP/6-311+G**density functional level of theory. The most stable isomers of closo-BnHn−2(CO)2 are similar to those of closo-C2Bn−2Hn in geometric patterns apart from closo-B6H4(CO)2, and closo-BnHn−2(CO)2 is much less strained than closo-C2Bn−2Hn. Energetic analysis identifies closo-B6H4(CO)2, closo-B12H10(CO)2 and closo-B10H8(CO)2 to be most stable, of which the latter two cages have been prepared experimentally. On the basis of the negative and rather large nucleus independent chemical shifts (NICS), closo-BnHn−2(CO)2 are aromatic. To aid further experimental study, the CO stretching frequencies have been computed.  相似文献   

7.
The infrared spectra of solid samples of C4H7K and C4D7K have been investigated in the 4000 to 30 cm−1 range. A complete assignment of intramolecular fundamentals of C4H7 and C4D7 ions and of potassium-allyl vibrations is proposed and the intramolecular force constants are calculated. The C(CH2)32− anion has been identified spectroscopically. Structures of C3H5, C4H7 and C(CH3)32− are discussed and compared with those optimised by the MINDO/3 method.  相似文献   

8.
The new 11-vertex nido-diphosphaborane, 7,9-Ph2-nido-7,9-P2B9H9, has been synthesized by the reaction of Me4N+[nido-B9H12] with PhPCl2 in the presence of NaH. A single crystal X-ray diffraction determination and DFT/GIAO/NMR methods have both established that the compound has an open cage structure containing the phosphorus atoms in non-adjacent positions on the open face.  相似文献   

9.
The molecular and crystal structure of the nido-6-tungstadecaborane [6,6,6,6-(CO)2(PPh3)2-nido-6-WB9H13] (1) has been determined showing that the tungsten atom is incorporated into the 6-position of a nido 10-vertex (WB9) cage. The tungsten atom has a seven-coordinate capped trigonal prismatic environment and is bonded to two hydrogen and three boron atoms of the {B9H13} cage, in addition to two CO groups and two PPh3 ligands. Variable-temperature (−90°C to +50°C) 31P{1H} NMR spectroscopy of 1 reveals that the exo-polyhedral ligands about the tungsten atom are fluxional with respect to PPh3 site exchange with an activation energy (ΔG‡), at the coalescence temperature (−73°C), of <38 kJ mol−1.  相似文献   

10.
In situ reaction of Li[closo-1-Ph-1,2-C2B10H10] with 7-azabicyclo [4.1.0] heptane results in the formation of the disubstituted carborane, closo-1-Ph-2-(2′-aminocyclohexyl)-1,2-C2B10H10 (1), in 63% yield. Decapitation of (1) with potassium hydroxide in refluxing ethanol produces the cage-opened nido-carborane, K[nido-7-Ph-8-(2′-aminocyclohexyl)-7,8-C2B9H10] (2), in 80% yield. Deprotonation of the above monoanion with two equivalents of n-butyllithium followed by reaction with anhydrous MCl4 · 2THF (M = Zr, Ti) provides d0-half-sandwich metallocarboranes, closo-1-M(Cl)-2-Ph-3-(2′-σ-(H)N-cyclohexyl)-2,3-η5-C2B9H9 (3 M = Zr; 4 M = Ti) in 53% and 42% yields, respectively. The reaction of Li[closo-1,2-C2B10H11] with 7-azabicyclo [4.1.0] heptane in THF affords closo-1-(2′-aminocyclohexyl)-1,2-C2B10H10 (5) in 59% yield. Immobilization of the carboranyl amino ligand (1) to an organic support, Merrifield’s peptide resin (1%), has been achieved by the reaction of the sodium salt of (5) with polystyryl chloride in THF to produce closo-1-(2′-aminocyclohexyl)-2-polystyryl-1,2-C2B10H10 (6) in 87% yield. Further reaction of the dianion derived from (6) with anhydrous ZrCl4 · 2THF led to the formation of the organic polystyryl supported d0-half-sandwich metallocarborane, closo-1-Zr(Cl)-2-(2′-σ-(H)N-cyclohexyl)-3-polystyryl-2,3-η5-C2B9H9 (7), in 38% yield. These new compounds have been characterized by elemental analyses, NMR, and IR spectra. Polymerizations of both ethylene and vinyl chloride with (3) and (7) have been performed in toluene using MMAO-7 (13% ISOPAR-E) as the co-catalyst. Molecular weights up to 32.8 × 103 (Mw/Mn = 1.8) and 9.5 × 103 (Mw/Mn = 2.1) were obtained for PE and PVC, respectively.  相似文献   

11.
By using ab initio methods of all-electron or effective core potential calculations, the electronic structures and the possible aromaticity of some 10π-electron systems, C6H64− (1), N64− (2), P64− (3), S62− (4), Te62− (5) and S3N3 (6), have been studied at the SCF levels using 4-31G//4-31G and 6-31G*//6-31G* basis sets. The bonding characteristics of these systems are analysed in terms of the canonical molecular orbital and the Foster-Boys localized molecular orbital results. The application of the second-order Jahn-Teller theorem to the stability of these diamagnetical planar species is presented.  相似文献   

12.
The new iodoammonium salts o-C6H4(NH2)2I+I (1) and o-C6H4(NH2)2I+ AsF6 (2) were prepared by reaction of o-phenylene diamine with I2 or I3+AsF6, respectively. Compound 1 reacts with AlI3 yielding quantitatively the corresponding tetraiodoaluminate o-C6H4(NH2)2I+AlI4 (3). The species were characterized by chemical analysis, vibrational (IR and Raman) and temperature-dependent 1H NMR spectropscopy. Direct evidence for a N---I bond was found in the Raman spectra of 1, 2 and 3 (ν(NI) = 599–600 cm−1).  相似文献   

13.
The generality of a two-electron reduction process involving an mechanism has been established for M3(CO)12 and M3(CO)12n(PPh3)n (M = Ru, Os) clusters in all solvents. Detailed coulometric and spectral studies in CH2Cl2 provide strong evidence for the formation of an ‘opened’ M3(CO)122− species the triangulo radical anions M3(CO)12−· having a half-life of < 10−6 s in CH2Cl2. However, the electrochemical response is sensitive to the presence of water and is concentration dependent. An electrochemical response for “opened” M3(CO)122− is only detected at low concentrations < 5 × 10−4 mol dm−3 and under drybox conditions. The electroactive species ground at higher concentrations and in the presence of water M3(CO)112− and M6(CO)182− were confirmed by a study of the electrochemistry of these anions in CH2Cl2; HM3(CO)11 is not a product. The couple [M6(CO)18]−/2− is chemically reversible under certain conditions but oxidation of HM3(CO)11 is chemically irreversible. Different electrochemical behaviour for Ru3(CO)12 is found when [PPN][X] (X = OAc, Cl) salts are supporting electrolytes. In these solutions formation of the ultimate electroactive species [μ-C(O)XRu3(CO)10] at the electrode is stopped under CO or at low temperatures but Ru3(CO)12−· is still trapped by reversible attack by X presumably as [η1-C(O)XRu3(CO)11]. It is shown that electrode-initiated electron catalysed substitution of M3(CO)12 only takes place on the electrochemical timescale when M = Ru, but it is slow, inefficient and non-selective, whereas BPK-initiated nucleophilic substitution of Ru3(CO)12 is only specific and fast in ether solvents particulary THF. Metal---metal bond cleavage is the most important influence on the rate and specificity of catalytic substitution by electron or [PPN]-initiation. The redox chemistry of M3(CO)12 clusters (M = Fe, Ru, Os) is a consequence of the relative rates of metal---metal bond dissociation, metal-metal bond strength and ligand dissociation and in many aspects resembles their photochemistry.  相似文献   

14.
When heated under reflux in CH2Cl2 solution with [Os(CO)3Cl2]2, two nido-[B9H12] units edge-fuse to form anti-[B18H21].  相似文献   

15.
The reactions of RNHSi(Me)2Cl (1, R=t-Bu; 2, R=2,6-(Me2CH)2C6H3) with the carborane ligands, nido-1-Na(C4H8O)-2,3-(SiMe3)2-2,3-C2B4H5 (3) and Li[closo-1-R′-1,2-C2B10H10] (4), produced two kinds of neutral ligand precursors, nido-5-[Si(Me)2N(H)R]-2,3-(SiMe3)2-2,3-C2B4H5, (5, R=t-Bu) and closo-1-R′-2-[Si(Me)2N(H)R]-1,2-C2B10H10 (6, R=t-Bu, R′=Ph; 7, R=2,6-(Me2CH)2C6H3, R′=H), in 85, 92, and 95% yields, respectively. Treatment of closo-2-[Si(Me)2NH(2,6-(Me2CH)2C6H3)]-1,2-C2B10H11 (7) with three equivalents of freshly cut sodium metal in the presence of naphthalene produced the corresponding cage-opened sodium salt of the “carbons apart” carborane trianion, [nido-3-{Si(Me)2N(2,6-(Me2CH)2C6H3)}-1,3-C2B10H11]3− (8) in almost quantitative yield. The reaction of the trianion, 8, with anhydrous MCl4 (M=Ti and Zr) in 1:1 molar ratio in dry tetrahydrofuran (THF) at −78 °C, resulted in the formation of the corresponding half-sandwich neutral d0-metallacarborane, closo-1-M[(Cl)(THF)n]-2-[1′-η1σ-N(2,6-(Me2CH)2C6H3)(Me)2Si]-2,4-η6-C2B10H11 (M=Ti (9), n=0; M=Zr (10), n=1) in 47 and 36% yields, respectively. All compounds were characterized by elemental analysis, 1H-, 11B-, and 13C-NMR spectra and IR spectra. The carborane ligand, 7, was also characterized by single crystal X-ray diffraction. Compound 7 crystallizes in the monoclinic space group P21/c with a=8.2357(19) Å, b=28.686(7) Å, c=9.921(2) Å; β=93.482(4)°; V=2339.5(9) Å3, and Z=4. The final refinements of 7 converged at R=0.0736; wR=0.1494; GOF=1.372 for observed reflections.  相似文献   

16.
Rate constants for the tunneling reaction (HD + D → h + D2) in solid HD increase steeply with increasing temperature above 5 K, while they are almost constant below 4.2 K. The apparent activation energy for the tunneling reaction above 5 K is 95 K, which is consistent with the energy (91–112 K) for vacancy formation in solid hydrogen. The results above 5 K were explained by the model that the tunneling reaction was accelerated by a local motion of hydrogen molecules and hydrogen atoms. The model of the tunneling reaction assisted by the local motion of the reactans and products was applied to the temperature dependence of the proton-transfer tunneling reaction (C6H6 + C2H5OH → C6H7 + C2H5O) in solid ethanol, the tunneling elimination of H2 molecule of H2 molecule ((CH3)2 CHCH(CH3)2+ → (CH3)2 C = C(CH3)2+ + H2) in solid 2,3-dimethylbutane, and the selective tunneling reaction of H atoms in solid neo-C5H12-alkane mixtures.  相似文献   

17.
The three cyanocuprate(I) complexes, Cu(CN)2, Cu(CN)32−, and Cu(CN)43−, photoeject electrons with high efficiency when excited in aqueous solution by 266 nm laser pulses of 7 ns duration with quantum yields of 0.37±0.06, 0.224±0.021, and 0.240±0.005, for Cu(CN)2 (at 2 M ionic strength), Cu(CN)32−, and Cu(CN)43− (both measured at 1 M ionic strength). Along with hydrated electrons, two transient intermediates, absorbing at 460 and 340 nm, respectively, form consecutively after excitation through bimolecular reactions with ground-state Cu(I) in solutions of Cu(CN)2, and Cu(CN)32−, but not in Cu(CN)43−. All photoprocesses are essentially monophotonic. A mechanism is proposed that suggests the formation of a dinuclear excited-state complex such as an excimer.  相似文献   

18.
A new method has been developed for ion-interaction chromatography with suppressed conductivity detection and a new graphitized carbon packing, which is sintered from carbonic material at a high temperature. Combinations of various eluting agents, tetrabutylammonium hydroxide (TBA) and acetonitrile have been investigated to optimize the separation of eight common anions (F, Cl, NO2, Br, NO3, SO42−, HPO42− and I). Calibration curves were linear from 0.5 to 10 μg/ml for F, from 1.0 to 20 μg/ml for Cl, NO2 and NO3, from 2.5 to 50 μg/ml for Br and SO42− and from 5.0 to 100 μg/ml for HPO42− and I with a correlation coefficient (r) of 0.999 or better. The relative standard deviations (R.S.D.s) of peak areas were between 0.2 and 0.9% for 10 repeated measurements. The application of this newly developed method was demonstrated by the determination of chloride, bromide and sulfate in pharmaceutical compounds using the direct injection method. The analytical results were within ±2% (relative) of the theoretical value, and thus in good agreement with the theoretical value for each sample.  相似文献   

19.
This paper describes an electrostatic ion chromatographic system in which the separation selectivity for inorganic anions, especially for sulfate and phosphate, could be manipulated by altering the molar ratio of the zwitterionic and cationic surfactants in the column coating solution used to prepare the stationary phase. The zwitterionic surfactant used for this study was 3-(N,N-dimethyltetradecylammonio)propanesulfonate (Zwittergent-3-14) and the cationic surfactant was tetradecyltrimethylammonium (TTA). Using a reversed-phase C18 column (250×4.6 mm I.D.) coated with 10/10 (mM/mM) of TTA/Zwittergent-3-14 mixed micelles as the stationary phase and either NaHCO3 or Na2CO3 aqueous solution as the eluent, together with suppressed conductivity detection, baseline separation of seven model inorganic anions was obtained. The elution order for those anions was found to be F42−42−23. Under the same conditions but using 1/10 (mM/mM) of TTA/Zwittergent-3-14 mixed micelles as the column coating solution, the elution order for these model ions was F42−42−23. The early elution of phosphate and sulfate is a unique attribute of this system. Detection limits for F, HPO42−, Cl, SO42−, NO2, Br and NO3 (S/N=3, sample injection volume 100 μl) were 0.11, 0.12, 0.12, 0.18, 0.49, 0.49, 0.52 μM, respectively.  相似文献   

20.
Organic-rich natural waters from peat bogs in continental (Switzerland) and maritime (Shetland Islands, Scotland) areas were analysed for Cl, NO2, Br, NO3, HPO42−, SO42− and oxalate using ion chromatography. These anions can be determined simultaneously in the surface and pore water samples from the continental bogs using a 250-μl injection loop. Using this loop, the detection limits were ca. 5 ng/g for the monovalent anions and SO42− and 10 ng/g for HPO42− and oxalate. An organics-removal cartridge (Dionex OnGuard P) was used to remove humic materials. These cartridges did not significantly affect the measured concentrations of anions in blind standards. Analyses of deionized water treated with these cartridges are not significantly different from those for untreated deionized water. For the maritime bogs, the relatively high concentrations of Cl (more than 100μ/g in many samples) and SO42− (up to 50 μg/g) require two separate determinations for complete analyses. A 10-μl injection loop was used to determine Cl, Br and SO42−. A 250-μl injection loop was used to measure NO2, NO3, HPO 42− and oxalate. In each instance a Dionex OnGuard P cartridge was used to remove humic materials. In addition, a chloride-removal cartridge (Dionex OnGuard AG) was used to remove Cl when the larger injection loop was used. This cartridge has no significant effect on the measurement of HPO4-2− at concentrations of 20 ng/g. In each of the bog water chromatograms there were usually a number of unknown peaks. These are probably due mainly to organic anions.  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