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 875 毫秒
1.
Reactions of (R)Ph2PN? S3N3 heterocycles with olefins such as norbornadiene, norbornene and dicyclopentadiene have yielded different results. Like Ph3PN? S3N3, (R)Ph2PN? S3N3 (R = C4H8N? , C5H10N? , C6H12N? , CH3NC4H8N? and OC4H8N? ) compounds have given the cycloaddition products (R)Ph2PN? S3N3 · C7H8 (yield 41 - 62%) with norbornadiene, while norbornene and dicyclopentadiene have not produced the corresponding adducts under identical conditions. With Ph3PN? S3N3 both norbornene and dicyclopentadiene have given the ring expanded heterocycle, 1,5-[Ph3PN]2S4N4 in ca. 65% yield. The solution phase decomposition studies of (secondary amino) Ph2PN? S3N3 derivatives have lead to the formation of (R)Ph2PNH2+X?, while (primary amino) Ph2PN? S3N3 has given Ph2PS2N3 heterocycle in ca. 80% yield. Acid hydrolysis of (R)Ph2PN? S3N3 derivatives has resulted in the isolation of Ph2P(O)OH in all the cases, where R is an amino group.  相似文献   

2.
A new method for the synthesis of benzotetrazine-1,3-dioxides was developed from the 2-(tert-butyl-NNO-azoxy)-N-nitroaniline O-alkyl derivatives upon the action of strong acids (H2SO4, MeSO3H, CF3CO2H) or BF3·Et2O. A mechanism suggested for these reactions includes transformation of the N=N(O)OR group (R = Me, Pri) to the oxodiazonium ion (N=N=O)+, which intramolecularly reacts with the neighboring tert-butyl-NNO-azoxy group, furnishing the tetrazine-1,3-dioxide ring.  相似文献   

3.
Organometallic Lewis Acids. XLII. Carbonyl- and Nitrosyl Complexes of Manganese and Rhenium of Weakly Coordinated Anions (Ph3P)2(ON)2MnX, (Ph3P)n(OC)5–nMX (M = Mn, Re; n = 1, 2; X = FBF3, OSO2CF3, OSO2F, OCORf) The complexes (Ph3P)2(ON)2MnX (X = FBF3, OSO2CF3, OSO2F, OCOCF3, OCOC3F7) and (Ph3P)n(OC)5–nMX (M = Mn, Re; n = 1, 2; X = FBF3, OSO2CF3) have been obtained by reaction of (Ph3P)2(ON)2MnH and (Ph3P)n(OC)5–nMeMe with the corresponding acids HX or from (Ph3P)n(OC)5–nReBr (n = 1, 2) with silver salts AgX, respectively. The compounds have been characterized by their IR and partially by 19F-NMR data. An efficient method for the preparation of the hydride (Ph3P)2(ON)2MnH is reported.  相似文献   

4.
Summary The SCF method is applied to determine the (gas phase) structure of [(CF3)2PN]2NVCl2, which agrees with the solid-state X-ray structure within typical errors of 2 pm and 2° in bond distances and angles. The electronic structure of atoms forming the ring is best described in terms of divalent N and tetravlent P+ with appreciable declocalization of nitrogen lone pairs into low-lying empty orbitals of neighbouring atoms P and V. No evidence for aromaticity of the ring system is found.  相似文献   

5.
Bis(cyclopentadienyl)mercury readily undergoes Diels—Alder reactions with RCCR (R = CO2Me or CF3), CF3CFCFCF3, CF3CFCF2, (CF3)2CC(CN)2, C2(CN)4 and PhNCONNCO to give stable adducts characterised by1H, 19F and 13C NMR, spectroscopy. Similar reactions of CF3CCCF3 and CF3CFCFCF3 with the cyclopentadiene derivatives Me3MC5H5 and (Me3M)2C5H4 (M = Si, Sn) are also described.  相似文献   

6.
Sodium dithionite initiated addition of CF2Br2, CF3I and (CF3)2CFI to the terminal double bond of allylbenzenes and of (CF3)2CFI to allylpyridines in a MeCN/H2O system were investigated. The reactions of CF2Br2 with allylbenzenes gave comparable amounts of adducts, 1-(2,4-dibromo-4,4-difluorobutyl)benzenes, debrominated products,1-(4-bromo-4,4-difluorobutyl)benzenes, and dimeric compounds in total yields 40-66%. Treatment of the adducts with DBU resulted in double dehydrohalogenation affording 4-aryl-1,1-difluorobutadienes which undergo Diels-Alder condensation with nitrogen dienophiles to give N-heterocycles with difluoromethylene group in the ring. The reactions of CF3I and (CF3)2CFI with allylbenzenes gave the respective adducts, (4,4,4-trifluoro-2-iodobutyl)benzenes and 1-(4,5,5,5-tetrafluoro-4-(trifluoromethyl)-2-iodopentyl)benzenes as the main products. Dehydrohalogenation of these adducts resulted, respectively, in (4,4,4-trifluoro-but-1-enyl)benzenes and 4-aryl-1,1-bis(trifluoromethyl)butadienes in high yields. (CF3)2CFI reacted rapidly with allylpyridines to give mixtures from which, after treatment with DBU, 4-pyridyl-1,1-bis(trifluoromethyl)butadienes were isolated in a ca. 60% yield.  相似文献   

7.
The geminal frustrated Lewis pair tBu2PCH2B(Fxyl)2 ( 1 ; Fxyl=3,5‐(CF3)2C6H3) is accessible in 65 % yield from tBu2PCH2Li and (Fxyl)2BF. According to NMR spectroscopy and X‐ray crystallography, 1 is monomeric both in solution and in the solid state. The intramolecular P ??? B distance of 2.900(5) Å and the full planarity of the borane site exclude any significant P/B interaction. Compound 1 readily activates a broad variety of substrates including H2, EtMe2SiH, CO2/CS2, Ph2CO, and H3CCN. Terminal alkynes react with heterolysis of the C?H bond. Haloboranes give cyclic adducts with strong P?BX3 and weak R3B?X bonds. Unprecedented transformations leading to zwitterionic XP/BCX3 adducts occur on treatment of 1 with CCl4 or CBr4 in Et2O. In less polar solvents (C6H6, n‐pentane), XP/BCX3 adduct formation is accompanied by the generation of significant amounts of XP/BX adducts. FLP 1 catalyzes the hydrogenation of PhCH=NtBu and the hydrosilylation of Ph2CO with EtMe2SiH.  相似文献   

8.
The electronic structures of lanthanide tris(β-diketonate) complexes and their adducts, being of interest as luminophores, were studied by UV photoelectron spectroscopy of vapors and X-ray photoelectron spectroscopy. The spectral regularities identified by the density functional theory revealed the influence of the substitution of the Me groups in the ligands by But, CF3, and neutral ligand in the adducts on the electronic structure of the complexing agent from Pr to Lu.  相似文献   

9.
The synthesis, X‐ray crystal structures, electrochemical, and spectroscopic studies of a series of hexanuclear gold(I) μ3‐ferrocenylmethylphosphido complexes stabilized by bridging phosphine ligands, [Au6(P?P)n(Fc‐CH2‐P)2][PF6]2 (n=3, P?P=dppm (bis(diphenylphosphino)methane) ( 1 ), dppe (1,2‐bis(diphenylphosphino)ethane) ( 2 ), dppp (1,3‐bis(diphenylphosphino)propane) ( 3 ), Ph2PN(C3H7)‐PPh2 ( 4 ), Ph2PN(Ph‐CH3p)PPh2 ( 5 ), dppf (1,1′‐bis(diphenylphosphino)ferrocene) ( 6 ); n=2, P?P=dpepp (bis(2‐diphenylphosphinoethyl)phenylphosphine) ( 7 )), as platforms for multiple redox‐active ferrocenyl units, are reported. The investigation of the structural changes of the clusters has been probed by introducing different bridging phosphine ligands. This class of gold(I) μ3‐ferrocenylmethylphosphido complexes has been found to exhibit one reversible oxidation couple, suggestive of the absence of electronic communication between the ferrocene units through the Au6P2 cluster core, providing an understanding of the electronic properties of the hexanuclear AuI cluster linkage. The present complexes also serve as an ideal system for the design of multi‐electron reservoir and molecular battery systems.  相似文献   

10.
Abstract

The reactions of chiral diphosphazanes. Ph2PN((S)-*CHMePh)PPhY (Y =Ph, N2C3HMe2-3,5) with [CpRu (PPh3)2Cl] and those of the monosulfides, Ph2PN(R)P(S)Ph2 (R = (S)-*CHMePh or CHMe2) with Ru3(CO)12. [RhCl(cod)]2 and [RhCI(CO)2]2 have been investigated. Molybdenum-palladium heterometallic complexes of the diphosphazanes, MeN(P(OR)2)2 (R = CH2CF3 or Ph) have been synthesised. Some unusual complexes have been obtained by the reductive carbonylation of cobalt and ruthenium halides in the presence of diphosphazanes, RN(PX2)2 (R = Me, X = OCH2CS or OPh; R = CHMe2, X = Ph). The structures of the products have been elucidated by NMR spectoscopy and in some cases confirmed by X-ray crystallography (e.g., 1–4).  相似文献   

11.
Hetero-Diels–Alder reactions of perfluoroalkyl thioamides with electron-rich 1,3-dienes such as 2,3-dimethylbutadiene, isoprene or penta-1,3-diene gave a simple and efficient access to new 2-aminosubstituted-3,6-dihydro-2H-thiopyrans. Three different procedures were used depending on the nature of the polyfluoroalkyl chains (RF=CF3, (CF2)nCF3, (CF2)4H) and on the nitrogen substituents of the thioamides (R1, R2=H, p-Tol, morpholino, Ac). Moreover, cycloadditions of silyloxydienes (1- or 2-trimethylsilyloxy-1,3-butadiene and Danishefsky's diene) with N-acyl,N-tolyl trifluoromethylthioamides afforded in almost all cases the corresponding 3,6-dihydro-2H-thiopyrans or 3-oxo-tetrahydrothiopyrans. For non-symmetrical 1,3-dienes, the regio- and stereochemistry of the reactions were studied (especially using X-ray diffraction analysis) indicating a strong similarity with those reported for fluorinated thiocarboxyl derivatives. Finally, two silylated 3,6-dihydro-2H-thiopyrans underwent an unexpected base-induced ring contraction to give new 1,3-thiazolidin-4-ones.  相似文献   

12.
The different courses of the interactions of cyclopentadienyl and arene derivatives of Group VI and VII transition metal carbonyl and transition complexes with Lewis acids, in solutions, have been studied by IR spectroscopy.The information of adducts involving the metal atom was observed for CpRe(CO)2L (L = CO, PR3) with SnCl4, SnBr4, TiCl4; AreneM(CO)3 (M = Cr, Mo, W) with SnCl4, TiCl4; and Ph3PC5H4M(CO)3 (M = Cr, Mo, W) with TiCl4 and AlCl3. Complexes CpM(CO)2NO and CpM(CO(NO)PPh3, depending on thier donor and acceptor nature, form adducts involving the oxygen atoms of CO or NO groups or the metal atom. CpCr(NO)2Cl reacts with Lewis acids via the chlorine atom. The relative basicity of the different sites in the complexes investigated is discussed.  相似文献   

13.
A novel scandium(III) complex with disulfonates as counter anions, [Sc(μ-OH)(H2O)5]2[O3S(CF2)3SO3]2 (5), was prepared from scandium oxides (Sc2O3) and perfluoropropane-1,3-disulfonic acid (1, HO3SCF2CF2CF2SO3H). By X-ray analysis, 5 was found to be a μ-OH-bridged dimeric structure bearing two perfluoropropane-1,3-disulfonates without bonding to scandium(III) centers. A series of lanthanide(III) complexes were also prepared from 1 and lanthanide oxides (Ln2O3; Ln = La, Nd, Sm, and Gd). In sharp contrast to the dimeric scandium(III) complex, the corresponding lanthanide(III) complexes had monomeric structures. Interestingly, the dimeric scandium(III) complex, but not the monomeric lanthanide complexes, with perfluoropropane-1,3-disulfonates served as an efficient Lewis acid catalyst for the hydrolysis of esters.  相似文献   

14.
The following reactions have been accomplished: 2(CF3)2NO· + Ph2CCO → Ph2C[ON(CF3)2]CO2N(CF3)2 → (on hydrolysis) Ph2C[ON(CF3)2]CO2H; 2 (CF3)2NO· + Ph2CHCOX → (CF3)2NOH + Ph2C[ON(CF3)2]COX (X  OH, Cl,NH2).  相似文献   

15.
The reactions of bromophenylacetylene with the diphenylphosphino amines (Ph2P)2NH and (Ph2P)3N result in the 1,1,3,3,4-pentaphenyl-1,2,5-azadiphospholium bromide ( 5 ) and its 4-diphenylphosphino derivative ( 7 ). Their X-ray structure analyses show planar rings for the cations. The small endocyclic angle at the nitrogen ring member is associated with PN bonds that are longer than in the acyclic cation (Ph3P)2N+. The data fit to a negative linear relationship of PN bond lengths and PNP bond angles. © 1997 John Wiley & Sons, Inc.  相似文献   

16.
Organoboranes carrying electron‐withdrawing substituents are commonly used as Lewis acidic catalysts or cocatalysts in a variety of organic processes. These Lewis acids also became popular through their application in `frustrated Lewis pairs', i.e. combinations of Lewis acids and bases that are unable to fully neutralize each other due to steric or electronic effects. We have determined the crystal and molecular structures of four heteroleptic arylboranes carrying 2‐(trifluoromethyl)phenyl, 2,6‐bis(trifluoromethyl)phenyl, 3,5‐bis(trifluoromethyl)phenyl or mesityl substituents. [3,5‐Bis(trifluoromethyl)phenyl]bis[2‐(trifluoromethyl)phenyl]borane, C22H11BF12, (I), crystallizes with two molecules in the asymmetric unit which show very similar geometric parameters. In one of the two molecules, both trifluoromethyl groups of the 3,5‐bis(trifluoromethyl)phenyl substituent are disordered over two positions. In [3,5‐bis(trifluoromethyl)phenyl]bis[2,6‐bis(trifluoromethyl)phenyl]borane, C24H9BF18, (II), only one of the two meta‐trifluoromethyl groups is disordered. In [2,6‐bis(trifluoromethyl)phenyl]bis[3,5‐bis(trifluoromethyl)phenyl]borane, C24H9BF18, (III), both meta‐trifluoromethyl groups of only one 3,5‐bis(trifluoromethyl)phenyl ring are disordered. [3,5‐Bis(trifluoromethyl)phenyl]dimesitylborane, C26H25BF6, (IV), carries only one meta‐trifluoromethyl‐substituted phenyl ring, with one of the two trifluoromethyl groups disordered over two positions. In addition to compounds (I)–(IV), the structure of bis[2,6‐bis(trifluoromethyl)phenyl]fluoroborane, C16H6BF13, (V), is presented. None of the ortho‐trifluoromethyl groups is disordered in any of the five compounds. In all the structures, the boron centre is in a trigonal planar coordination. Nevertheless, the bond angles around this atom vary according to the bulkiness and mutual repulsion of the substituents of the phenyl rings. Also, the ortho‐trifluoromethyl‐substituted phenyl rings usually show longer B—C bonds and tend to be tilted out of the BC3 plane by a higher degree than the phenyl rings carrying ortho H atoms. A comparison with related structures corroborates the conclusions regarding the geometric parameters of the boron centre drawn from the five structures in this paper. On the other hand, CF3 groups in meta positions do not seem to have a marked effect on the geometry involving the boron centre. Furthermore, it has been observed for the structures reported here and those reported previously that for CF3 groups in ortho positions of the aromatic ring, disorder of the F atoms is less probable than for CF3 groups in meta or para positions of the ring.  相似文献   

17.
1,3-Dimethyl-2-imidazolidinone (dimethylethylene urea, DMEU) and 1,3-di- methyl-3,4,5,6-tetrahydro-2(IH)-pyrimidinone (dimethylpropylene urea, DMPU) adducts of the type Ph3SnX·L (X = Cl, Br and I), Ph3PbX·L (X = Br, I), 3Ph3PbCl·2DMEU and 2Ph3PbCl · DMPU have been prepared and characterized. Assignments are made for ν(CO) and ν(CN) frequencies in the IR, and for skeletal frequencies observed in both the IR and Raman spectra in the range 400 to 100 cm?1 Infrared measurements show that the adducts are bound through the carbonyl oxygen, and are highly dissociated in dichloromethane solution. 1H and 119Sn or 207Pb NMR measurements reveal that ligand exchange, fast on the NMR time scale, occurs in solution. Coordination of the ligand causes a large upfield shift in the 119Sn or 207Pb resonances, but Ph3MI · L have shifts similar to those for the parent iodides, indicating almost complete dissociation. Thermodynamic parameters are reported for the dissociation of Ph3SnX · DMPU (X = Cl, Br) in CH2Cl2 solution.  相似文献   

18.
Thermodynamic data have been obtained by calorimetric titration in benzene solution at 30° for reaction of organotin compounds with Lewis bases; data are reported for forty acid/base systems.Ph3SnCl forms 11 adducts of low stability with pyridine (py) or 4-methyl-pyridine (4-mepy). Ph2SnCl2, Me2SnCl2, Bu2SnCl2 and Bu2Sn(NCS)2 form simultaneously 11 and 12 adducts with py or 4-mepy and 11 adducts with 2,2′-bipyridine or 1,10-phenanthroline (phen); the enthalpies of formation of the phen adducts are similar to those of 12 adducts with 4-mepy. With BuSnCl3 and PhSnCl3 it was not possible to obtain data for each step in addition of pyridine or 4-mepy. Adduct stabilities increase with increasing chloride substitution and in the order Bu < Me < Ph; adducts of Bu2Sn(NCS)2 are more stable than those of Bu2SnCl2.Tributylphosphine does not react with Ph3SnCl but gives 11 adducts with the other tin compounds; only PhSnCl3 adds a second molecule of this base. The 11 adducts are more stable than those with heterocyclic bases. Tributylamine brings about disproportionation of the compounds R2SnX2 to R4Sn and SnX4NBu3.  相似文献   

19.
Stannylation Experiments with NH-functional Aminoiminophosphoranes. Synthesis and Structure of the Tricyclic Stannaphosphazenes [Me2Sn(tBu2PN)NH]2 and [nBu2Sn(Ph2PN)2NH]2 Aminoiminophosphoranes tBu2P(NH)NH2 ( 1 ) and (H2NPPh2)N(Ph2PNH) ( 2 ) react with diaminostannanes R2Sn(NEt2)2 by cyclocondensation to give cyclostannaphosphazenes [Me2Sn(tBu2PN)NH]2 ( 3 ) and [R2Sn(Ph2PN)2NH]2 ( 4 a , b ) ( a : R = Me, b : R = nBu). With 2 and Me3SnNEt2 the ring compound Me2Sn(Ph2PN)2NSnMe3 ( 5 ) besides Me4Sn is formed by per-N-stannylation and Sn-methyl group transfer. The crystal structures of 3 and 4 b were determined by X-ray structure analysis. 3 forms a planar heterotricyclus containing three four-membered rings with two pentacoordinated tin atoms (space group P 1 (No. 2); Z = 1). 4 b consists of a tricyclic molecule with two puckered six-membered rings and one planar four membered tin-nitrogen ring with two pentacoordinated tin atoms (space group P 1 (No. 2); Z = 1).  相似文献   

20.
We report herein the synthesis and full characterization of the donor‐free Lewis superacids Al(ORF)3 with ORF=OC(CF3)3 ( 1 ) and OC(C5F10)C6F5 ( 2 ), the stabilization of 1 as adducts with the very weak Lewis bases PhF, 1,2‐F2C6H4, and SO2, as well as the internal C? F activation pathway of 1 leading to Al2(F)(ORF)5 ( 4 ) and trimeric [FAl(ORF)2]3 ( 5 , ORF=OC(CF3)3). Insights have been gained from NMR studies, single‐crystal structure determinations, and DFT calculations. The usefulness of these Lewis acids for halide abstractions has been demonstrated by reactions with trityl chloride (NMR; crystal structures). The trityl salts allow the introduction of new, heteroleptic weakly coordinating [Cl‐Al(ORF)3]? anions, for example, by hydride or alkyl abstraction reactions.  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