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.
A series of octanuclear iodine-bromine interhalides [InBr8−n]2− (n=0, 2, 3, 4) were prepared systematically in two steps. Firstly, addition of a dihalogen (Br2 or IBr) to the triaminocyclopropenium bromide salt [C3(NEt2)3]Br forms the corresponding trihalide salt with Br3 or IBr2 anions, respectively. Secondly, addition to Br3 of half an equivalent of Br2 gives the octabromine polyhalide [Br8]2−, whereas addition to IBr2 of half an equivalent of Br2, IBr or I2 gives the corresponding interhalides: [I2Br6]2−, [I3Br5]2−, and [I4Br4]2−, respectively. The four octahalides were characterized by X-ray crystallography, computational studies, Raman and Far-IR spectroscopies, as well as by TGA and melting point. All of the salts were found to be ionic liquids.  相似文献   

2.
Syntheses, Vibrational Spectra, and Normal Coordinate Analysis of Halogenonitrosylruthenates [Ru(NO)ClnBr5–n]2–, n = 0–5, and the Crystal Structure of (CH2py2)[Ru(NO)ClBr4] By treatment of [Ru(NO)Cl5]2– with anhydrous HBr in dichloromethane a mixture of [Ru(NO)ClnBr5–n]2–, n = 0–5, is formed, from which individual complexes can be separated by ion exchange chromatography on diethylaminoethyl cellulose. The X-Ray structure determination on a single crystal of (CH2py2)[Ru(NO)ClBr4] (monoclinic, space group P21/c, a = 11.480(2), b = 10.175(4), c = 16.025(6) Å, β = 107.40(1)°, Z = 4) reveals, that the chlorine atom is trans positioned to the nitrosyl group. The low temperature IR and Raman spectra have been recorded of six complexes of the series (n-Bu4N)2[Ru(NO)ClnBr5–n], n = 0–5, and are assigned by normal coordinate analysis. A good agreement between observed and calculated frequencies is achieved. The valence force constants are fd(NO) = 13.86–13.93 und fd(RuN) = 5.43–5.49 mdyn/Å.  相似文献   

3.
Crystal Structures, Vibrational Spectra, and Normal Coordinate Analyses of the Chloro-Iodo-Rhenates(IV) (CH2Py2)[ReCl5I], cis -(CH2Py2)[ReCl4I2] · 2 DMSO, trans -(CH2Py2)[ReCl4I2] · 2 DMSO, and fac -(EtPh3P)2[ReCl3I3] [ReCl5I]2–, cis-[ReCl4I2]2–, trans-[ReCl4I2]2–, and fac-[ReCl3I3]2– have been synthesized by ligand exchange reactions of [ReI6]2– with HCl and are separated by ion exchange chromatography on diethylaminoethyl cellulose. X-ray structure determinations have been performed on single crystals of (CH2Py2)[ReCl5I] ( 1 ) (triclinic, space group P1 with a = 7.685(2), b = 9.253(2), c = 12.090(4) Å, α = 90.06(2), β = 101.11(2), γ = 95.07(2)°, Z = 2), cis-(CH2Py2)[ReCl4I2] · 2 DMSO ( 2 ) (triclinic, space group P1 with a = 8.662(2), b = 12.109(2), c = 12.9510(12) Å, a = 97.533(11), β = 96.82(2), γ = 89.90(2)°, Z = 2) , trans-(CH2Py2)[ReCl4I2] · 2 DMSO ( 3 ) (triclinic, space group P1 with a = 9.315(7), b = 9.663(3), c = 15.232(3) Å, α = 80.09(2), β = 81.79(4), γ = 83.99(5)°, Z = 2) and fac-(EtPh3P)2[ReCl3I3] ( 4 ) (monoclinic, space group P21/a with a = 17.453(2), b = 13.366(1), c = 19.420(1) Å, β = 112.132(8)°, Z = 4). The crystal structure of ( 1 ) reveals a positional disorder of the anion sublattice along the asymmetric axis. Due to the stronger trans influence of I compared with Cl on asymmetric axes Cl˙–Re–I′ is caused a mean lenghthening of the Re–Cl˙ distances of 0.020 Å (0.8%) and a shortening of the Re–I′ distances of 0.035 Å (1.3%) with regard to symmetrically coordinated axes Cl–Re–Cl and I–Re–I, respectively. Using the molecular parameters of the X-Ray determinations the low temperature (10 K) IR and Raman spectra of the (n-Bu4N) salts of all four chloro-iodo-rhenates(IV) are assigned by normal coordinate analyses. The weakening of the Re–Cl˙ bonds and the strengthening of the Re–I′ bonds is indicated by a decrease or increase of the valence force constants each by 9%.  相似文献   

4.
Formation of Organosilicon Compounds. LXII. Partial Brominated Carbosilanes The photobromination of 1 leads to compound 2 as well as to C-chlorinated derivatives if the time of reaction is prolonged. Compound 2 is also formed from (Br2Si–CH2)3; Gl. (1) see ?Inhaltsübersicht”?. In a corresponding reaction (Cl3Si–CH2)2SiCl2 gives successively Cl3Si–CHBr–SiCl2–CH2–SiCl3, Cl3Si–CBr2–SiCl2–CH2–SiCl3 and Cl3Si–CCl2–SiCl2–CH2–SiCl3. (Cl3Si)2CBr2 is accessible through the photobromination of (Cl3Si)2CH2. The reactivity of the CBr2-group is quite obvious in the reaction of Cl2Si–CBr2–SiCl2–CH2–SiCl3 with LiAlH4 yielding (H3Si–CH2)2SiCl2 as well as in the reaction of compound 2 with CH3MgCl yielding [(CH3)2Si–CH2]3. By treatment of the SiH groups with bromine the preparation of compounds with the general formulas CH3SiHnBr3?n; (H3?nSiBrn)2CH2; (H3?nSiBrn? CH2)2SiH2?nBrn; (H2?nBrnSi? CH2)3 and (H3?nSiBrn)2CCl2 is possible. Analysis of the nmr spectra shows that 1,3-Dibromo-1,3,5-trisilacyclohexane is formed to 67% in the trans and to 33% in the cis configuration; 1,3,5-Tribromo-1,3,5-trisilacyclohexane is formed to 80–90% in teh cis-trans configuration. The results of 1H and 29Si NMR investigations are reported.  相似文献   

5.
Preparation and Structure of the Compounds Ba2Pb4F10Br2–xIx (x = 0–2) with Related Structure Motifs of the Fluorites and Matlockites Colourless single crystals of Ba2Pb4F10Br2–xIx (x = 0–2) have been obtained under hydrothermal conditions (T = 250 °C, 10 d), starting from stoichiometric amounts of BaF2, PbF2, PbBr2 and PbI2. The compounds crystallize in the tetragonal space group P4/nmm (No. 129). A complete miscibility in the region x = 0–2 has been observed. The mixed crystals follow Vegard's rule. For the compounds with the composition Ba2Pb4F10Br2 (a = 5.9501(2) Å, c = 9.6768(10) Å, R[F2 > 2σ(F2)] = 0.022, wR(F2 all reflections) = 0.059), Ba2Pb4F10Br1.1I0,9 (a = 5.9899(3) Å, c = 9.7848(5) Å, R[F2 > 2σ(F2)] = 0.014, wR(F2 all reflections) = 0.035) and Ba2Pb4F10I2 (a = 6.6417(3) Å, c = 9.9216(10) Å, R[F2 > 2σ(F2)] = 0.023, wR(F2 all reflections) = 0.049) complete structure analyses have been performed on the basis of single crystal diffractometer data. Microcrystalline single phase compounds Ba2Pb4F10Br2–xIx (x = 0–2) have been obtained by coprecipitation from aqueous solutions of KF, KBr (KI) and Ba(CH3COO)2, Pb(NO3)2 in acetic acid medium. For Ba2Pb4F10Br1.5I0.5 and Ba2Pb4F10Br0.5I1.5 powder data of microcrystalline samples were used for the Rietveld analyses (RBragg = 0.077 for Ba2Pb4F10Br1,5I0,5 and RBragg = 0.065 for Ba2Pb4F10Br0.5I1.5). The crystal structure comprises alternating structural features of fluorite related type (CaF2) around Ba and matlockite related type (PbFCl) around Pb1 and Pb2 along the c axis. Barium shows a {8 + 4} cuboctahedral coordination of fluorine. The coordination polyhedron around the two crystallographically independent lead atoms is a monocapped quadratic antiprism built of {4 + 1} fluorine and {4} bromine or iodine atoms, respectively.  相似文献   

6.
To improve our understanding of the electrospray ionization (ESI) process, we have subjected equimolar mixtures of salts A+X (A+ = Li+, NBu4+; X = Br, ClO4, BF4, BPh4) in different solvents (CH3CN, tetrahydrofuran, CH3OH, H2O) to negative‐ion mode ESI and analyzed the relative ESI activity of the different anionic model analytes. The ESI activity of the large and hydrophobic BPh4 ion greatly exceeds that of the smaller and more hydrophilic anions Br, ClO4 and BF4, which we ascribe to its higher surface activity. Moreover, the ESI activity of the anions is modulated by the action of the counter‐ions and their different tendency toward ion pairing. The tendency toward ion pairing can be reduced by the addition of the chelating ligands 12‐crown‐4 and 2.2.1 cryptand and is, although to a smaller degree, further influenced by the variation of the solvent. Complementary electrical conductivity measurements afford additional information on the interactions of the ionic constituents of the sample solutions. Copyright © 2017 John Wiley & Sons, Ltd.  相似文献   

7.
Preparation and Vibrational Spectra of Nonahalogenodirhodates(III), [Rh2ClnBr9-n]3?, n = 0–9 The pure nonahalogenodirhodates(III), A3[Rh2ClnBr9-n], A = K, Cs, (TBA); n = 0–4, 9, have been prepared. They are formed from the monomer chlorobromorhodates(III), [RhClnBr6-n]3?, n = 0–6, which are bridged to confacial bioctahedral complexes by ligand abstraction in less polar organic solvents. From the mixtures the complexions are separated by ion exchange chromatography on DEAE-cellulose. The solid, air-stable, air-stable, K-, Cs- and (TBA)-salts of [Rh2ClnBr9-n]3?, n = 0–4, are green, of [Rh2Cl9]3? are brown. The IR and Raman spectra of [Rh2Br9]3? and [Rh2Cl9]3? are assigned according to the point group D3h. The chlorobromodirhodates exist as mixtures of geometrical and structural isomers, which belong to different point groups. The vibrational spectra exhibit bands in characteristic regions; at high wavenumbers stretching vibrations with terminal ligands v(Rh—Clt): 360–320, v(Rh—Brt): 280–250; in a middle region with bridging ligands v(Rh—Clb): 300–270, v(Rh—Brb): 210–170 cm?1; the deformation bands are observed at distinct lower frequencies. The terminal ligands are fixed very strong, and the distance between v(Rh—Xt) and v(Rh—Xb) increases with decreasing size of the cations.  相似文献   

8.
[Tetrakis(acetonitrile)‐dibromo‐nickel(II)]‐di‐acetonitrile was obtained from a solution of nickel(II) dibromide in acetonitrile at 258 K. The crystal structure [monoclinic, P21/n (no.14), a = 1005.5(5), b = 831.3(5) , c = 1131.7(5) pm, β = 106.263(5)°, V = 908.1(8)·106 pm3, Z = 2, R1 for 1580 reflections with I0>2σ(I0): 0.0505] contains sixfold coordinated NiII atoms. Two trans coordinating bromide anions and four equatorial acetonitrile molecules form an elongated octahedron around the central NiII atom. [Ni(CH3CN)4Br2] octahedra are connected via hydrogen bonds to neighboring octahedra as well as to solvate acetonitrile molecules.  相似文献   

9.
10.
Cyclohexane and methylcyclopentane dimerize into dimethyldecalins on treatment with superelectrophilic systems containing polyhalomethanes (CBr4, CCl4, CHCl3) and aluminum halides (AlBr3, AlCl3). At 20 °C, the yields of C12H22 hydrocarbons reach 140 mol.% based on the superelectrophile. Under the action of CBr4·2AlBr3 at 20 °C in CH2Br2 or without a solvent, n-pentane is converted predominantly into lower alkanes (mainly, isopentane). In addition, higher branched C8--C12 alkanes, small amounts of alkylated cyclohexanes, cycloalkenes, dienes, and alkylbenzenes are formed in a total yield of 20--23 % (w/w) based on pentane.  相似文献   

11.
Synthesis, Crystal Structure, and Vibrational Spectra of cis ‐(CH2Py2)[ReBr4Py2]2 · (CH3)2CO By reaction of (n‐Bu4N)2[ReBr6] with pyridine and (n‐Bu4N)BH4 in dichloromethane halogeno‐pyridine‐rhenium(III)complexes are formed and purified by chromatography. X‐ray structure determination on a single crystal has been performed of cis‐(CH2Py2)[ReBr4Py2]2 · (CH3)2CO (monoclinic, space group P21/c, a = 15.0690(9), b = 8.3337(8), c = 35.588(4) Å, β = 96.409(7), Z = 4). Based on the molecular parameters of the X‐ray structure determination and assuming C2 point symmetry for the anion cis‐[ReBr4Py2] the IR and Raman spectra are assigned by normal coordinate analysis. The valence force constants are in the Br–Re–Br axis fd(ReBr) = 1.49, in the asymmetrically coordinated N′–Re–Br · axes fd(ReBr · ) = 1.03 und fd(ReN′) = 2.52 mdyn/Å.  相似文献   

12.
Geometric, electronic, and energy characteristics of the complexes formed in the CF4 ·nAIF3 (n = I or 2) and CBr4 ·nAIBr3 (n = 1, 2, or 4) systems have been determined by the semiempirical AM I method. Besides the donor-acceptor complexes, the CBr3 +...AIBr4 , CBr3 +...Al2Br7 , CBr22+...(AlBr4 )2, and CBr2 2+...(Al2Br7 )2 ionic complexes can be formed in the CBr4 ·nAlBr3 systems. In the cations and dications of polyhalomethanes (when Hal = Cl, Br, or l) in both the free and bound (included in ionic complexes) states, carbon atoms carry negative charges, the C-Hal bonds are substantially shortened, and the positive charges are located on one-coordinate halogen atoms. These cations and dications can be considered as halenium ions that differ from halenium salts with dicoordinate halogen atoms. In the cationic and dicationic complexes of the CBr4 ·nAlBr3 systems, the maximum positive charges on the Br atoms are 0.39 and 0.94, respectively. Fluorine-containing cations and dications have structures similar to those of carbenium ions, whereas in the CF4 ·nAIF3 systems (n = l or 2), only donor-acceptor complexes are formed.Translated fromIzvestiya Akademii Nauk. Seriya Khimicheskaya. No. 3, pp. 554–560, March, 1996.  相似文献   

13.
An analysis of thermochemical and kinetic data on the bromination of the halomethanes CH4–nXn (X = F, Cl, Br; n = 1–3), the two chlorofluoromethanes, CH2FCl and CHFCl2, and CH4, shows that the recently reported heats of formation of the radicals CH2Cl, CHCl2, CHBr2, and CFCl2, and the C? H bond dissociation energies in the matching halomethanes are not compatible with the activation energies for the corresponding reverse reactions. From the observed trends in CH4 and the other halomethanes, the following revised ΔH°f,298 (R) values have been derived: ΔH°f(CH2Cl) = 29.1 ± 1.0, ΔH°f(CHCl2) = 23.5 ± 1.2, ΔHf(CH2Br) = 40.4 ± 1.0, ΔH°f(CHBr2) = 45.0 ± 2.2, and ΔH°f(CFCl2) = ?21.3 ± 2.4 kcal mol?1. The previously unavailable radical heat of formation, ΔH°f(CHFCl) = ?14.5 ± 2.4 kcal mol?1 has also been deduced. These values are used with the heats of formation of the parent compounds from the literature to evaluate C? H and C? X bond dissociation energies in CH3Cl, CH2Cl2, CH3Br, CH2Br2, CH2FCl, and CHFCl2.  相似文献   

14.
About the Preparation of N-Chloro-N-Methylammonium Salts (CH3)nNCl4–n+MF6? (n = 1–3; M = As, Sb) and (CH3)2NClX+MF6? (X = F, Br) Simple one-step methods for the preparation of the methylated chloroammonium salts (CH3)nNCl4–n+MF6? (n = 1–3; M = As, Sb) and for (CH3)2NClX+MF6? (X = F, Br) are reported. Their vibrational and NMR-spectroscopical data are discussed in comparison.  相似文献   

15.
A new organic–inorganic hybrid compound, catena‐poly[bis(1‐ethyl‐3‐methylimidazolium) [μ5‐bromido‐tri‐μ3‐bromido‐tri‐μ2‐bromido‐pentacuprate(I)]], {(C6H11N2)2[Cu5Br7]}n, has been obtained under ionothermal conditions from a reaction mixture containing Ba(OH)2·8H2O, Cu(OH)2·2H2O, As2O5, 1‐ethyl‐3‐methylimidazolium bromide and distilled water. The crystal structure consists of complex [Cu5Br7]2− anions arranged in sinusoidal {[Cu5Br7]2−}n chains running along the a axis, which are surrounded by 1‐ethyl‐3‐methylimidazolium cations. Three of the five unique Br atoms and one of the three CuI atoms occupy special positions with half‐occupancy (a mirror plane perpendicular to the b axis, site symmetry m). The CuI ions are in a distorted tetrahedral coordination environment, with four Br atoms at distances ranging from 2.3667 (10) to 2.6197 (13) Å, and an outlier at 3.0283 (12) Å, exceptionally elongated and with a small contribution to the bond‐valence sum of only 6.7%. Short C—H...Br contacts build up a three‐dimensional network. The Cu...Cu distances within the chain range from 2.8390 (12) to 3.0805 (17) Å, indicating the existence of weak CuI...CuI cuprophilic interactions.  相似文献   

16.
The Preparation of Methylthio(trihalogeno)phosphonium Salts ClnBr3?nPSCH3+MF6?(n = 0–3; M = As, Sb) and Hal3PSCH3+SbCl6?(Hal = Br, Cl) The methylthio(trihalogeno) phosphonium salts BrnCl3?nPSCH3+MF6? (n = 0–3; M = As, Sb) are prepared by methylation of the corresponding thiophosphorylhalides BrnCl3?nPS in the system SO2/CH3F/MF5. The hexachloroantimonates Hal3PSCH3+SbCl6?(Hal = Br, Cl) are synthesized by thiomethylation of PBr3 and PCl3 with CH3SCl/SbCl5. All salts are characterized by vibrational and NMR spectroscopy.  相似文献   

17.
The new asymmetrical organic ligand 2‐{4‐[(1H‐imidazol‐1‐yl)methyl]phenyl}‐5‐(pyridin‐4‐yl)‐1,3,4‐oxadiazole ( L , C17H13N5O), containing pyridine and imidazole terminal groups, as well as potential oxdiazole coordination sites, was designed and synthesized. The coordination chemistry of L with soft AgI, CuI and CdII metal ions was investigated and three new coordination polymers (CPs), namely, catena‐poly[[silver(I)‐μ‐2‐{4‐[(1H‐imidazol‐1‐yl)methyl]phenyl}‐5‐(pyridin‐4‐yl)‐1,3,4‐oxadiazole] hexafluoridophosphate], {[Ag( L )]PF6}n, catena‐poly[[copper(I)‐di‐μ‐iodido‐copper(I)‐bis(μ‐2‐{4‐[(1H‐imidazol‐1‐yl)methyl]phenyl}‐5‐(pyridin‐4‐yl)‐1,3,4‐oxadiazole)] 1,4‐dioxane monosolvate], {[Cu2I2( L )2]·C4H8O2}n, and catena‐poly[[[dinitratocopper(II)]‐bis(μ‐2‐{4‐[(1H‐imidazol‐1‐yl)methyl]phenyl}‐5‐(pyridin‐4‐yl)‐1,3,4‐oxadiazole)]–methanol–water (1/1/0.65)], {[Cd( L )2(NO3)2]·2CH4O·0.65H2O}n, were obtained. The experimental results show that ligand L coordinates easily with linear AgI, tetrahedral CuI and octahedral CdII metal atoms to form one‐dimensional polymeric structures. The intermediate oxadiazole ring does not participate in the coordination interactions with the metal ions. In all three CPs, weak π–π interactions between the nearly coplanar pyridine, oxadiazole and benzene rings play an important role in the packing of the polymeric chains.  相似文献   

18.
The vibrational (IR and Raman) and photoelectron spectral properties of hydrated iodine‐dimer radical‐anion clusters, I2.? ? n H2O (n=1–10), are presented. Several initial guess structures are considered for each size of cluster to locate the global minimum‐energy structure by applying a Monte Carlo simulated annealing procedure including spin–orbit interaction. In the Raman spectrum, hydration reduces the intensity of the I? I stretching band but enhances the intensity of the O? H stretching band of water. Raman spectra of more highly hydrated clusters appear to be simpler than the corresponding IR spectra. Vibrational bands due to simultaneous stretching vibrations of O? H bonds in a cyclic water network are observed for I2.? ? n H2O clusters with n≥3. The vertical detachment energy (VDE) profile shows stepwise saturation that indicates closing of the geometrical shell in the hydrated clusters on addition of every four water molecules. The calculated VDE of finite‐size small hydrated clusters is extrapolated to evaluate the bulk VDE value of I2.? in aqueous solution as 7.6 eV at the CCSD(T) level of theory. Structure and spectroscopic properties of these hydrated clusters are compared with those of hydrated clusters of Cl2.? and Br2.?.  相似文献   

19.
The data on temperature, solvent, and high hydrostatic pressure influence on the rate of the ene reactions of 4‐phenyl‐1,2,4‐triazoline‐3,5‐dione ( 1 ) with 2‐carene ( 2 ), and β‐pinene ( 4 ) have been obtained. Ene reactions 1 + 2 and 1 + 4 have high heat effects: ∆Hrn ( 1 + 2 ) −158.4, ∆Hrn( 1 + 4 ) −159.2 kJ mol−1, 25°C, 1,2‐dichloroethane. The comparison of the activation volume (∆V( 1 + 2 ) −29.9 cm3 mol−1, toluene; ∆V( 1 + 4 ) −36.0 cm3 mol−1, ethyl acetate) and reaction volume values (∆Vr‐n( 1 + 2 ) −24.0 cm3 mol−1, toluene; ∆Vr‐n( 1 + 4 ) −30.4 cm3 mol−1, ethyl acetate) reveals more compact cyclic transition states in comparison with the acyclic reaction products 3 and 5 . In the series of nine solvents, the reaction rate of 1+2 increases 260‐fold and 1+4 increases 200‐fold, respectively, but not due to the solvent polarity.  相似文献   

20.
The purpose of this article was to calculate the structures and energetics of CH3O(H2O)n and CH3S(H2O)n in the gas phase; the maximum number of water molecules that can directly interact with the O of CH3O; and when n is larger, we asked how the CH3O and CH3S moiety of CH3O(H2O)n and CH3S(H2O)n changes and how we can reproduce experimental ΔH 0n−1, n. Using the ab initio closed-shell self-consistent field method with the energy gradient technique, we carried out full geometry optimizations with the MP2/aug-cc-pVDZ for CH3O(H2O)n (n=0, 1, 2, 3) and the MP2/6–31+G(d,p) (for n=5, 6). The structures of CH3S(H2O)n (n=0, 1, 2, 3) were fully optimized using MP2/6–31++G(2d,2p). It is predicted that the CH3O(H2O)6 does not exist. We also performed vibrational analysis for all clusters [except CH3O(H2O)6] at the optimized structures to confirm that all vibrational frequencies are real. Those clusters have all real vibrational frequencies and correspond to equilibrium structures. The results show that the above maximum number of water molecules for CH3O is five in the gas phase. For CH3O(H2O)n, when n becomes larger, the C—O bond length becomes longer, the C—H bond lengths become smaller, the HCO bond angles become smaller, the charge on the hydrogen of CH3 becomes more positive, and these values of CH3O(H2O)n approach the corresponding values of CH3OH with the n increment. The C—O bond length of CH3O(H2O)3 is longer than the C—O bond length of CH3O in the gas phase by 0.044 Å at the MP2/aug-cc-pVDZ level of theory. The structure of the CH3S moiety in CH3S(H2O)n does not change with the n increment. ©1999 John Wiley & Sons, Inc. J Comput Chem 20: 1138–1144, 1999  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