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 578 毫秒
1.
On the Lewis Acidity of Fluorinated Sulfonium Ions NMR investigations show, that sulfonium salts [(CF3)nSF3?n]+ AsF6? ( 1–3 , n = 0–2) add CH3CN under formation of ψ-pentacoordinated sulfuranonium ions [(CF3)nSF3?n · NCCH3]+ ( 1a – 3a ,) with the donor in an axial position. In solution NSF3 ( 4 ,) forms similar salts [(CF3)nSF3?n · NSF3]+ AsF6? ( 1b-3b ,) with weaker donor-acceptor interactions. With NSF2NMe2 ( 5 ,) the step of the primary addition products is passed very quickly, by fluoride-migration from 1 , and 2 , persulfuranonium ions [(CH3)2NSF3NSF2]+ ( 6 ,) and [(CH3)2NSF3NSFCF3]+ ( 7 ,), respectively, are formed, while from 3 , only decomposition products (Me2NSF2+, CF3SSCF3, CF4) were obtained.  相似文献   

2.
4-(3-Alkylureido)-2, 2, 6, 6-tetramethylpiperidine-1-oxyls are rapidly oxidized by N2O4 or NOCl to 4-(3-alkylureido)-2, 2, 6, 6-tetramethyl-1-oxopiperidinium nitrates and chlorides, which are then nitrosated to 4-(3-alkyl-3-nitrosoureido)-2, 2, 6, 6-tetramethyl-1-oxopiperidinium salts. The perchlorates of the latter were prepared by an exchange reaction with HClO4. The nitrosation of alkylureidooxoammonium salts is the first example of chemical modification of oxoammonium derivatives in which the highly reactive >N+=O group is inert toward the reagent.Translated fromIzvestiya Akademii Nauk. Seriya Khimicheskaya, No. 3, pp. 542–547, March, 1993.  相似文献   

3.
The crystal structures of six halobismuth(III) salts of variously substituted aminopyridinium cations display discrete mononuclear [BiCl6]3? and dinuclear [Bi2X10]4? anions (X = Cl or Br), and polymeric cis‐double‐halo‐bridged [BinX4n]n? anionic chains (X = Br or I). Bis(2‐amino‐3‐ammoniopyridinium) hexachloridobismuth(III) chloride monohydrate, (C5H9N3)2[BiCl6]Cl·H2O, (1), contains discrete mononuclear [BiCl6]3? and chloride anions. Tetrakis(2‐amino‐3‐methylpyridinium) di‐μ‐chlorido‐bis[tetrachloridobismuth(III)], (C6H9N2)4[Bi2Cl10], (2), tetrakis(2‐amino‐3‐methylpyridinium) di‐μ‐bromido‐bis[tetrabromidobismuth(III)], (C6H9N2)4[Bi2Br10], (3), and bis(4‐amino‐3‐ammoniopyridinium) di‐μ‐chlorido‐bis[tetrachloridobismuth(III)] dihydrate, (C5H9N3)2[Bi2Cl10]·2H2O, (4), incorporate discrete [Bi2X10]4? anions (X = Cl or Br), while catena‐poly[2,6‐diaminopyridinium [[cis‐diiodidobismuth(III)]‐di‐μ‐iodido]], {(C5H8N3)[BiI4]}n, (5), and catena‐poly[2,6‐diaminopyridinium [[cis‐dibromidobismuth(III)]‐di‐μ‐bromido]], {(C5H7N2)[BiBr4]}n, (6), include [BinX4n]n? anionic chains (X = Br or I). Structures (2) and (3) are isostructural, while that of (5) is a pseudomerohedral twin. There is no discernible correlation between the type of anionic species obtained and the cation or halide ligand used. The BiIII centres always have a slightly distorted octahedral geometry and there is a correlation between the Bi—X bond lengths and the number of classic N—H…X hydrogen bonds that the X ligand accepts, with a greater number of interactions corresponding with slightly longer Bi—X distances. The supramolecular networks formed by classic N—H…X hydrogen bonds include ladders, bilayers and three‐dimensional frameworks.  相似文献   

4.
Simple pentafluorobenzyl‐substituted ammonium and pyridinium salts with different anions can be easily obtained by treatment of the parent amine or pyridine with the respective pentafluorobenzyl halide. Hexafluorophosphate is introduced as the anion by salt metathesis. In the case of the ammonium salt 4 , water co‐crystallisation seems to suppress effective anion–π interactions of bromide with the electron‐deficient aromatic system, whereas with salts 5 and 6 such interactions are observed despite the presence of water. However, due to asymmetric hydrogen‐bonding interactions with ammonium side chains, the anion of 5 is located close to the rim of the pentafluorophenyl group (η1 interaction). In 6 the CH–anion hydrogen bonding is more symmetric and fixes the anion on top of the ring (η6). A similar structure‐controlling effect is observed in case of the 1,4‐diazabicyclo[2.2.2]octane derivatives 7 . Here the position of the anion (Cl, Br, I) is shifted according to the length of the weak CH–halide interaction. The hexafluorophosphate 7 d reveals that this “non‐coordinating” anion can be located on top of an aromatic π system. In the methyl‐substituted pyridinium salts 9 and 10 different locations of the bromide anions with respect to the π system are observed. This is due to different conformations of the mono‐ versus disubstituted pyridine, which leads to different directions of the weak, but structurally important, HMe? Br bonds.  相似文献   

5.
6.
Ethyl 3-aminocrotonate, when reacted with hydroxy(tosyloxy)iodobenzene, forms the tosylate of ethyl 3-amino-2-phenyliodoniocrotonate which crystallizes well in up to 80% yield. X-ray analysis confirms the structure of the phenyliodonium salt, revealing intramolecular and unusual intermolecular hydrogen bonds, stabilizing the compound in the crystalline state. Reaction with pyridine, its 4-substituted derivatives, and 4,4-bipyridine yields tosylates of 2-pyridinio-substituted ethyl 3-aminocrotonates.1H NMR and IR spectra support formation of an intramolecular hydrogen bond for the E-isomer, and iodonium salts in the case of pyridinium salts for the Z-isomer. The UV spectra of the pyridinium salts show an intramolecular charge transfer band.Translated from Khimiya Geterotsiklicheskikh Soedinenii, No. 6, pp. 770–777, June, 2000.  相似文献   

7.
A number of salts of 2,2′:6′,2″ ‐terpyridyl (‘tpy’) with univalent anions (halides : X = Cl, Br, I; oxyanions of increasing basicity: ClO4, NO3, ‘tfa’ = trifluoroacetate, ‘tca’ = trichloroacetate), variously solvated, have been structurally characterized by single crystal X‐ray studies. In all cases the tpy moieties are found to be doubly protonated [tpyH2]2+, the hydrogen atoms being associated with the nitrogen atoms of the peripheral rings, these together with the central nitrogen atom being directed towards a common focus, in most cases ‘chelating’ one of the counter‐ion components in diverse ways. Thus the chloride and bromide compounds are isomorphous [(tpyH2)X]+X·H2O arrays; a second dihydrate phase is also described for the chloride, the two forms having the unchelated anion and water molecules engaged in hydrogen‐bonded networks essentially independent of [(tpyH2)X]+. The iodide is anhydrous, and of a different structural type, the anions, presumably too large for chelation, lying out of plane to either side, and linking different cations into a one‐dimensional polymer; in the perchlorate, the unsolvated aggregate is now discrete [(tpyH2)X2], a pair of perchlorate ions disposed to either side of the tpy plane, lying each with one oxygen atom interacting with both of the two protonating hydrogen atoms. In the anhydrous X = NO3, tfa, tca arrays, the lattices are solvated by the parent acids; one oxygen atom of each anion is chelated by the [tpyH2]2+ as in the chlorides, the other anion, with the acid, forming an independent ‘acid salt’ counterion [XHX] in each case, retaining the additional protonic hydrogen rather than further protonating the central ring, all being of the form [(tpyH2)X]X·HX = [(tpyH2)X][X(HX)].  相似文献   

8.
The Reaction of Bis(hydroxyphenyl)methanes with Phosphorus Pentachloride — Phosphocins with Phosphonium and Phosphorane Structure [1] Bis(2‐hydroxyphenyl)methane derivatives react with phosphorus pentachloride to give cyclic dioxaphosphocinium salts, as well as trichloro‐dioxaphosphocins. Chlorination of 6‐chloro‐dioxaphosphocins also leads to trichloro‐dioxaphosphocins. Hydrolysis of the dioxaphosphocin derivatives gives stable cyclic phosphates.  相似文献   

9.
Bromonium salts [(RF)2Br]Y with perfluorinated groups RFC6F5, CF3CFCF, C2F5CFCF, and CF3C≡C were isolated from reactions of BrF3 with RFBF2 in weakly coordinating solvents (wcs) like CF3CH2CHF2 (PFP) or CF3CH2CF2CH3 (PFB) in 30-90% yields. C6F5BF2 formed independent of the stoichiometry only [(C6F5)2Br][BF4]. 1:2 reactions of BrF3 and silanes C6F5SiY3 (Y = F, Me) ended with different products - C6F5BrF2 or [(C6F5)2Br][SiF5] - as pure individuals, depending on Y and on the reaction temperature (Y = F). With C6F5SiF3 at ≥−30 °C [(C6F5)2Br][SiF5] resulted in 92% yield whereas the reaction with less Lewis acidic C6F5SiMe3 only led to C6F5BrF2 (58%). The interaction of K[C6F5BF3] with BrF3 or [BrF2][SbF6] in anhydrous HF gave [(C6F5)2Br][SbF6]. Attempts to obtain a bis(perfluoroalkyl)bromonium salt by reactions of C6F13BF2 with BrF3 or of K[C6F13BF3] with [BrF2][SbF6] failed. The 3:2 reactions of BrF3 with (C6F5)3B in CH2Cl2 gave [(C6F5)2Br][(C6F5)nBF4−n] salts (n = 0-3). The mixture of anions could be converted to pure [BF4] salts by treatment with BF3·base.  相似文献   

10.
近年来,黄耀曾等应用有机鉮盐原地生成叶立德并将它用于有机合成,使叶立德化学有了更大的发展,我们将原地生成的胂叶立德与羰基化合物反应,产物为环氧化合物或烯烃,影响反应产物的因素除胂叶立德的结构外,溶剂效应往往起决定作用,为了进一步研究半稳定胂叶立德的反应性能,寻求有利的反应条件,探索溶剂效应。  相似文献   

11.
ESRStudiesonN-AlkylphenothiazineRadicalCationSaltsGUOQing-xiang ̄(1,2),LIUBo ̄1andLIUYou-chengi ̄(1,2).( ̄*NatioiialLaboratoryofA...  相似文献   

12.
Powder and coatings of metal-like refractory compounds (MLRC) can be produced by electrochemical synthesis from molten salts. The stoichiometry of the deposited MLRC was found to correlate with the molar ratio of MLRC component ions in the melts. The system Ti-Si-B is of particular interest in terms of electrochemical synthesis since the titanium, silicon and boron potentials in the melt are close together. The electrochemical synthesis has been investigated in the system NaCl-KCl-NaF-K2TiF6-K2SiF6-KBF4 at 700°C. The opportunity of deposition of new ternary compound in the system Ti-Si-B is shown by the electrochemical synthesis from molten salts.  相似文献   

13.
A very short synthesis of indolizidines, quinolizidines and some higher homologues was developed by alkylation of 2-methyl-1-pyrroline or 6-methyl-2,3,4,5-tetrahydropyridine with 1,3- or 1,4-dihaloalkanes, followed by reduction of the intermediate iminium salts, resulting in the desired 1-azabicyclo[m.n.0]alkanes in good yields.  相似文献   

14.
Syntheses and single crystal X‐ray structure characterisations are recorded for some novel series of crystalline complexes formed between salts of univalent anions of the ethane‐1,2‐diaminium cation, [enH2]X2, and 1,10‐phenanthroline (‘phen’), variously hydrated, thus: [enH2]X2·mphen(·nH2O), for m = 2, 4 and 10 (one example), n various. In all cases, the motifs constituting the arrays comprise columns of [enH2]2+ cations, carrying the protonic array but linked in a second dimension by hydrogen‐bonding to associated anions and water molecules (where present), expanding the column in some cases to form a sheet, different degrees of hydration compensating for changing anion bulk. In a third dimension the protonic hydrogen complement also links to the nitrogenous component of phen stacks which surround the column. Thus, for the m = 2 array, in a triclinic cell, a, b, c broadly 10‐11 (x2), 7Å, α, β, γ 80, 70° (x2), the cation and phen columns lie parallel to c; in the unsolvated trichloroacetate compound, the cation column is associated with anions to either side, these linking into a sheet with water molecules in the more highly hydrated trifluoroacetate (‘tfa’) and nitrate (n = 2), and chloride and bromide (n = 4) arrays (the tfa adduct a superlattice doubled in c). In the m = 4 arrays, an additional phen stack is inserted, forming a sheet with the first in the second dimension for the perchlorate tetrahydrate array, the iodide pentahydrate counterpart being a 2 x c superlattice. A second nitrate salt, m = 10, n = 4, is also described, a complex array of multiple networks of the above type. Single crystal X‐ray structure determinations are also recorded for salts [phenH](PF6)·phen and [2,9‐Me2phenH](PF6). In the phen adduct, the protonic hydrogen atom is closely associated (N···H 0.90(4) Å) with one of the two independent phen moieties, these disposed alternately in a stack up b close to the 21 screw axis, so that the hydrogen bridges to the unprotonated moiety (H···N′ 2.36(4) Å) pitched at an angle of 47° to it in the screw‐related stack. In the Me2phen salt, the phen moieties lie in crystallographic mirror planes, normal to and stacked up b, with the protonic hydrogen atom contacting a PF6 fluorine atom (H···F 1.96(3) Å). The structure of unsolvated Me2phen is also recorded.  相似文献   

15.
Dichloromethylenammonium salts, particularly dichloromethylenedimethylammonium chloride, occupy a unique place as stable but reactive building blocks for synthesis. They contain three mobile chlorine atoms activated by an amino group on the same carbon atom. These salts can be regarded as chlorinated Vilsmeier or Mannich reagents and are thus at a higher oxidation level. As in the Mannich or Vilsmeier reaction, the carbon condenses here as an electrophile with formation of C? C or C? hetero atom bonds in a variety that is still far from being exhausted.  相似文献   

16.
Mono-, bis- and tris-(1,3,2,4-dithiadiazolium) salts [R-(CNSNS +)n]n+[AsF-6]n (R = aryl, n = 1, 2, 3) were found to initiate the cationic ring-opening polymerization of tetrahydrofuran (THF) at room temperature to give clear gels from which the pure polymer was precipitated. 1,3,2,4-Dithiadiazolium cations associated with the hard [AsF6]- anion thus constitute a new class of cationic polymerization initiators. The poly(THF) formed by initiation with 1,3,2,4-dithiadiazolium cation was characterized by gel permeation chromatography, infrared spectrophotometry, and 13C-NMR spectroscopy. Number-average molecular weights of 198 700 g mol-1 (polydispersity 1.96) and 190 000 g mol-1 (polydispersity 1.61) were obtained using [PhCNSNS ] [AsF6] and [C6H3-1,3,5-(CNSNS )3][AsF6]3, respectively, as initiators. The use of multifunctional dithiadiazolium salts as initiators suggests that they may be useful in the preparation of starburst and dendritic polymers. © 1992 John Wiley & Sons, Inc.  相似文献   

17.
On the Preparation of Dimercapto(methyl)Sulfonium Salts [CH3S(SH)2]+ AsF6? and [CH3S(SH)2]+SbCl6? and the Bis(chlorothio)methylsulfonium Salts [CH3S(SCI)2]+ AsF6? and [CH3S(SCI)2]+ SbCl6? The preparation of the dimercapto(methyl)sulfonium salts [CH3S(SH)2]+ AsF6? and [CH3S(SH)2]+SbCl6? from [CH3SCl2]+ salts and H2S at 195 K is reported. The salts are stable below 210 K. They are characterized by additional Raman spektroscopic measurements of the isotopic labelled cations [CH3S(SD)2]+, [CH3S(34SH)2]+ and [CH3S(34SD)2]+. The dimercapto(methyl)sulfonium salts are transfered into bis(chlorthio)methylsulfonium salts by reaction with Cl2 at 195 K.  相似文献   

18.
TeX4 (X = Cl, Br) react in HCl/HBr with [Ph(CH3)2Te]X (X = Cl, Br) to give [PhTe(CH3)2]2[TeCl6] (1) and [PhTe(CH3)2]2[TeBr6] (2). The reaction of PhTeX3 (X = Cl, Br, I) in cooled methanol with [(Ph)3Te]X (X = Cl, Br, I) leads to [Ph3Te][PhTeCl4] (3), [Ph3Te][PhTeBr4] (4) and [Ph3Te][PhTeI4] (5). In the lattices of the telluronium tellurolate salts 1 and 2, octahedral TeCl6 and TeBr6 dianions are linked by telluronium cations through Te?Cl and Te?Br secondary bonds, attaining bidimensional (1) and three-dimensional (2) assemblies. The complexes 3, 4 and 5 show two kinds of Te?halogen secondary interactions: the anion-anion interactions, which form centrosymmetric dimers, and two identical sets of three telluronium-tellurolate interactions, which accomplish the centrosymmetric fundamental moiety of the supramolecular arrays of the three compounds, with the tellurium atoms attaining distorted octahedral geometries. Also phenyl C-H?halogen secondary interactions are structure forming forces in the crystalline structures of compounds 3, 4 and 5.  相似文献   

19.
A series of beta-cyclodextrin (beta-CD) dimers containing fluorescent 2,2'-oxamidobisbenzoyl and 4,4'-oxamidobisbenzoyl linkers--that is, 6,6'-[2,2'-oxamidobis(benzoylamino)]ethyleneamino-6,6'-deoxy-bis(beta-CD) (2), 6,6'-[2,2'-oxamidobis(benzoylamino)]diethylenediamino-6,6'-deoxy-bis(beta-CD) (3), 6,6'-[4,4'-oxamidobis(benzoylamino)]ethyleneamino-6,6'-deoxy-bis(beta-CD) (4), and 6,6'-[4,4'-oxamidobis(benzoylamino)]diethylenediamino-6,6'-deoxy- bis(beta-CD) (5)--were synthesized from the corresponding oxamidobis(benzoic acid)s through treatment with mono[6-aminoethyleneamino-6-deoxy]-beta-CD or mono[6-diethylenetriamino-6-deoxy]-beta-CD. Further treatment of 2-5 with copper perchlorate gave their Cu(II) complexes 6-9 in satisfactory yields. The conformation and binding behavior of 2-9 towards two bile salt guests--sodium cholate (CA) and sodium deoxycholate (DCA)--was comprehensively investigated by circular dichroism, 2D NMR spectroscopy, and fluorescence spectroscopy in Tris-HCl buffer solution (pH 7.2) at 25 degrees C. Thanks to the cooperative host-linker-guest binding mode, the stoichiometric 1:1 complexes formed by bis(beta-CD)s 2-5 with bile salts gave high stability constants (KS values) of up to 10(3)-10(4) M(-1). Significantly, benefiting from the intramolecular 1:2 or 2:4 binding stoichiometry, the resulting complexes of metallobis(beta-CD)s 6-9 with bile salts gave much higher KS values of up to 10(6)-10(7) M(-2). The enhanced binding abilities of bis(beta-CD)s and metallobridged bis(beta-CD)s are discussed from the viewpoints of induced-fit interactions and multiple recognition between host and guest.  相似文献   

20.
A novel type of transformations of the pyrimidine ring under the action of C-nucleophiles was found and a new method was developed for the synthesis of 3-hetarylquinolines from quinazoline derivatives and quaternary heterocyclic salts. An independent synthesis was carried out and transformations of one of the probable intermediates were studied. By-products were isolated. The effects of the nature of the heterocycle and substituents on the course of the ring transformation reaction were found, and the mechanism of the reaction was suggested. Translated fromIzvestiya Akademii Nauk. Seriya Khimicheskaya, No. 6, pp. 1210–1215, June, 1998.  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