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 765 毫秒
1.
Under Ammonia chemical Ionization conditions the source decompositions of [M + NH4]+ ions formed from epimeric tertiary steroid alchols 14 OHβ, 17OHα or 17 OHβ substituted at position 17 have been studied. They give rise to formation of [M + NH4? H2O]+ dentoed as [MHsH]+, [MsH? H2O]+, [MsH? NH3]+ and [MsH? NH3? H2O]+ ions. Stereochemical effects are observed in the ratios [MsH? H2O]+/[MsH? NH3]+. These effects are significant among metastable ions. In particular, only the [MsH]+ ions produced from trans-diol isomers lose a water molecule. The favoured loss of water can be accounted for by an SN2 mechanism in which the insertion of NH3 gives [MsH]+ with Walden inversion occurring during the ion-molecule reaction between [M + NH4]+ + NH3. The SN1 and SNi pathways have been rejected.  相似文献   

2.
Reaction of [RhCl2Cp*]2 (Cp* = η-C5Me5) with salicyloxazolines in the presence of NaOMe gives complexes [RhCl(R-saloxaz)Cp*] (1-4) which have been fully characterised. The diastereoselectivity of complexation depends on the substituents and the absolute configuration at the metal centre is unstable in solution. Treatment of 2 with 4-methylpyridine and NaSbF6 in methanol at reflux gave [Rh(4-Mepy){(S)-iPr-saloxaz}Cp*][SbF6] (5) whilst [Rh(OH2)(Me2-saloxaz)Cp*][SbF6] (6) was prepared by reaction of 1 with AgSbF6. Three complexes, [RhCl(Me2-saloxaz)Cp*] (1), [RhCl{(S)-iPr-saloxaz}Cp*] (2), and [Rh(OH2)(Me2-saloxaz)Cp*][SbF6] (6) have been characterised by X-ray crystallography. Some of the complexes, after treatment with AgSbF6, have been tested as enantioselective catalysts for the Diels-Alder reaction of methacrolein with cyclopentadiene.  相似文献   

3.
The reaction behavior of [XeF][AsF6] in solution toward hydrogen iodide, HI, was investigated, and Xe, HF, and [I4][AsF6]2 were identified as the final reaction products. The reaction enthalpy of the gas-phase reaction ([XeF]+ + HI → [XeI]+ (1Σ) + HF) was calculated at the optimized MP4(SDQ) geometries at the QCISD(TQ) level to be: ΔH0298 [QCISD(TQ)/LANL2DZ//MP4(SDQ)/LANL2DZ] = −63.3 kcal mol−1. The [XeI]+ cation is bound only in the 1Σ singlet state, and the triplet state (3Π) was shown to be essentially unbound at all levels of theory applied and very close in energy to the singlet state at equilibrium structure. According to the ab initio calculations, [XeI]+ can react with HI in a thermodynamically and spin-symmetry allowed reaction to yield the [XeI]+ (1Σ) cation that may, after interconversion into the unbound triplet state, immediately dissociate into xenon (1S) and I+ (3P). © 1997 John Wiley & Sons, Inc. Heteroatom Chem 8: 473–478, 1997  相似文献   

4.
The ammonia chemical ionization (CI/[NH4+]) mass spectra of a series of diastereomeric methyl and benzyl ethers derived from 3-hydroxy steroids (unsaturated in position 5 and saturated) have been studied. The adduct ions [M+NH4]+ and [MH]+ and the substitution product ions [M+NH4? ROH]+ (thereafter called [MsH]+) are characterized by an inversion in their relative stabilites in relation to their initial configuration. [M+NH4]α+ and [MH]α+ formed from the α-Δ5-steroid isomers are stabilized by the presence of a hydrogen bond which is not possible for the β-isomers. This stereochemical effect has also been observed in the mass analysed ion kinetic energy (MIKE) spectra of [M+NH4]+ and [MH]+. The MIKE spectra of [MsH]+ indicate that those issued from the β-isomers are more stable than the one originating from the α-isomers. This behavior is also observed in the first field free region (HV scan spectra) for [MH]+, [MsH]+ and [M+NH4]+ which are precursors of the ethylenic carbocations (base peak in the conventional CI/[NH4]+ spectra). Mechanisms, such as SN1 and SNi, have been ruled out for the formation of [MsH]+, but instead the data support an SN2 mechanism during the ion-molecule reaction between [M+NH4]+ and NH3.  相似文献   

5.
The reaction efficiency and enantioselectivity of an asymmetric Pauson-Khand-type reaction catalyzed by cationic rhodium are heavily dependent on the solvent. Coordinating solvents, such as THF, provide a faster reaction and better stereoselectivity than non-coordinating solvents, such as toluene. These beneficial effects can be attributed to a significant increase in the more reactive catalytic species of [Rh(bisphosphane ligand)(solvent)n]+ (3) than of [Rh(bisphosphine ligand)CO(solvent)]+(4) and [Rh(bisphosphine ligand)∗(CO)2]+ (5) in a coordinating solvent.  相似文献   

6.
The paper reports new data evidencing for a high electrophilicity of the positively charged titanium atom in the previously described zwitterionic titanocene monochloride Cp[η5-C5H4B(C6F5)3]TiCl (1) and titanocene monobromide Cp[η5-C5H4B(C6F5)3]TiBr (2), containing a B(C6F5)3 group in one of the C5 rings. It has been established that on a contact of a toluene solution of these zwitterions with water vapour at 20 °C under Ar, a rapid protolytic cleavage of the otherwise inert B-C6F5 bond in the tris(pentafluorophenyl)borane moiety occurs to afford pentafluorobenzene and the corresponding halogenide hydroxide complex of titanocene Cp[η5-C5H4B(C6F5)2]TiX(μ-OH), where X = Cl (3), Br (4). An X-ray diffraction study of the complexes has shown that the hydroxide group in 3 and 4 is bonded via the oxygen atom both to the titanium and boron atoms. Under similar conditions, the interaction of zwitterion 1 with methanol gives rise to pentafluorobenzene and the chloride methoxide complex of titanocene Cp[η5-C5H4B(C6F5)2]TiCl(μ-OCH3). It has been suggested that the driving force of the protolysis of the B-C6F5 bond in 1 and 2 is a sharp increase in the acidity of water or methanol molecule as a result of their complexation with the positively charged titanium centre in the starting zwitterion.  相似文献   

7.
The reaction of N-nitro-O-(4-nitrophenyl)hydroxylamine (1) with conc. H2SO4 affords 4-nitropyrocatechol and that with conc. sulfonic acids (RSO3H where R = Me, CF3) affords 2-hydroxy-5-nitrophenyl-R-sulfonates in yields of 80?C85%. These reactions are assumed to proceed through an intermediate (phenoxy)oxodiazonium ion [NO2C6H4O-N=N=O]+, which eliminates the N2O molecule to form the aryloxenium ion [NO2C6H4O]+. The latter reacts with acid anions at the ortho-carbon atom of the phenyl ring. The thermodynamical parameters of the elementary reactions resulting in the formation of the (phenoxy)oxodiazonium ion [NO2C6H4O-N=N=O]+ and aryloxenium ion [NO2C6H4O]+ were calculated in the B3LYP/6?311+G(d) study of the combined molecular system (nitrohydroxylamine 1 + [H3SO4]+). The reaction of nitrohydroxylamine 1 with aqueous solutions of strong acids (??70% H2SO4, CF3SO3H) affords mainly 4-nitrophenol. It appears that the mechanism of this reaction does not involve the formation of the aryloxenium ion.  相似文献   

8.
N-mesityl-N′-pyridyl-imidazolium chloride 1a and the corresponding bromide salt 1b have been deprotonated with NaH in THF giving the free N-heterocyclic carbene N-mesityl-N′-pyridyl-imidazolin-2-ylidene 2 in 80% yield (starting from 1a). Imidazolium salt 1a reacts with RuCl3 · xH2O to give a racemic mixture of dinuclear di-μ-chloro bridged ruthenium complexes [(κ2-2)2Ru(μ-Cl)2Ru(κ2-2)2]2+ [3a]2+. The carbene carbon atoms as well as the halides are arranged in cis-positions to each other whereas the nitrogen atoms adopt a trans-configuration. The di-μ-bromo bridged derivative [(κ2-2)2Ru(μ-Br)2Ru(κ2-2)2]2+ [3b]2+ was obtained from RuCl3 · xH2O and 1b. The bridging halide ligands can be removed by the reaction with silver or sodium salts of bidentate Lewis acids. Complex [3a]2+ reacts with silver pyridylcarboxylate to give a racemic mixture of the mononuclear complex [4]+. Reaction of [3a]2+ with the sodium salt of l-proline resulted in a diastereomeric mixture of complexes [5]+. The free N-heterocyclic carbene 2 reacts with [FeCl2(PPh3)2] to give after anion exchange with NaBPh4 cis/cis/trans coordinated [Fe(κ2-2)2(MeCN)2](BPh4)2 [6](BPh4)2. The molecular structures of [3b](PF6)2, [4]PF6 and [6](BPh4)2 · H2O are reported.  相似文献   

9.
Addition of BH3·thf to 1-alkylimidazoles (alkyl=methyl, butyl) and 1-methylbenzimidazole leads to BH3 adducts, which are deprotonated by BuLi to yield the organolithium compounds (L)Li+(1bd). In the solid state (thf)Li+1b is dimeric. The acyl–iron complexes (thf)3Li+(3b,d) are formed from (thf)Li+(1b,d) and Fe(CO)5. (L)Li+(1ac) react with [CpFe(CO)2X], however, the only complex obtained is [CpFe(CO)21a] (5a). The analogous reaction of (L)Li+1a with the pentadienyl complex [(C7H11)Fe(CO)2Br] yields the corresponding iron compound 6a. Their compositions follow from spectroscopic data. Treatment of Cp2TiCl with (L)Li+1a leads to [Cp2Ti1a] (7a), which could not be oxidized with PbCl2 to give the corresponding Ti(IV) complex. The compounds [Li(py)4]+9a and [Li(L)4]+(10bd) are obtained when (L)Li+1 are reacted with VCl3 and ScCl3. The X-ray structure analysis of the vanadium complex reveals a distorted tetrahedron of the anion [V(1a)4] with two smaller and four larger CVC angles. The scandium compound [Li(dme)2+10c] has a different structure: the distorted tetrahedron of the anion [Sc(1c)4] contains two larger (140.2 and 142.9°) and four smaller CScC angles (93.9–98.7°). This arrangement allows the formation of four bridging BHSc 3c,2e bonds to give an eight-fold coordination. The anion 10c is formally a 16e complex.  相似文献   

10.
The reaction of (η5-C5H4Me)4Fe4(HCCH)2 (1) with 1 equiv. of N-bromosuccinimide (NBS) gives the one-electron oxidized form in 83% yield. Further treatment of [1]+ with NBS results in the stepwise bromination of four acetylenic protons to give [(η5-C5H4Me)4Fe4(HCCH)(HCCBr)]+ ([2]+), [(η5-C5H4Me)4Fe4(HCCBr)2]+ ([3a]+), [(η5-C5H4Me)4Fe4(HCCBr)(BrCCBr)]+ ([4]+), and [(η5-C5H4Me)4Fe4(BrCCBr)2]+ ([5]+) in moderate yields, with the isomer of [3a]+, [(η5-C5H4Me)4Fe4(HCCH)(BrCCBr)]+ ([3b]+), formed as a minor product. These compounds are characterized by analytical and spectroscopic techniques, and the molecular structures of [2](PF6), [4](TFPB), and [5](TFPB) are established by X-ray diffraction analysis [TFPB = tetrakis{bis(3,5-trifluoromethyl)phenyl}borate]. The compounds are confirmed to retain the butterfly core of four iron atoms as in [1](TFPB). The bromoacetylene part in [2]+ exhibits high reactivity toward various nucleophiles: Cluster[2]+ is moisture-sensitive and is converted to a mixture of [(η5-C5H4Me)4Fe4(HCCH)(μ3-CH)(μ3-CO)]+ ([6]+) and [1]+. Reactions of [2]+ with ZnR2 (R = Me, Et) give [(η5-C5H4Me)4Fe4(HCCH)(HCC-R)]+ in good yields (R = Me ([9]+, 88%), Et ([10]+, 91%)). Accordingly, treatment of [2]+ with HC CMgBr and LiSpTol leads to the introduction of the ethynyl and thiolate groups to give [(η5-C5H4Me)4Fe4(HCCH)(HCC-CCH)]+ ([11]+, 95%) and [(η5-C5H4Me)4Fe4(HCCH)(HCC-SpTol)]+ ([12]+, 78%), respectively. Substitution of the bromo group in [2]+ with pyridine affords [(η5-C5H4Me)4Fe4(HCCH)(HCC-Py)]2+ ([13]2+) in 90% yield. The reaction with 4,4′-bipyridyl (bpy) requires the severer conditions (70 °C, 2 days), probably due to the relative low basicity of bpy, giving [(η5-C5H4Me)4Fe4(HCCH)(HCC-bpy)]2+ ([14]2+) in 54% yield. The substitution reaction with 4,4′-bipyridyl is strongly accelerated by treatment with silver salt to give [14]2+ in 90% yield. The products derived from [2]+ and nucleophiles are unequivocally determined by elemental, spectroscopic, and X-ray diffraction analyses.  相似文献   

11.
Novel mixed [Cu(4a–c)2]CF3SO3 pentanuclear copper(I)/chromium(0) complexes containing tweezer ligands that contain, within the same molecule, a 2,9-disubstituted-1,10-phenanthroline and two arene-chromium subunits. The mononuclear copper(I) complexes [Cu(5a–c)2] CF3SO3 have been prepared from the chromium free ligand species 5a–c, obtained by quantitative photochemical decomplexation of the parent species 4a–c. Cyclic voltammetry showed that in these complexes copper(I) can be reversibly oxidized to copper(II) in CH2Cl2. The CuII/I couples are more positive in the presence of the four surrounding redox-active arenechromium subunits, revealing an electron-withdrawing effect. Complexes [Cu(4c)2]+, [Cu(5c)2]+ form new species (1 : 1 ligand copper stoichiometry) upon addition of excess [Cu(CH3CN)4]+, which were identified electrochemically and by FAB+ mass spectroscopy.  相似文献   

12.
《Polyhedron》1999,18(23):2981-2985
The reaction of [{Ru(η6-C6H6)Cl(μ-Cl)}2] with Py3COH in ethanol results in the formation of the cation [Ru(η6-C6H6)(N,N′,O,-(C5H4N)3CO)]+ which is isolated as its hexafluorphosphate salt 1. The cation acts as a ligand towards other transition metal ions. With Ag+ the hetero-trinuclear complex [{Ru(η6-C6H6)((C5H4N)3CO)}2Ag][PF6]3 2 is formed, while reaction with [Pd(PhCN)2Cl2] gives the bimetallic [Ru(η6-C6H6)((C5H4N)3CO)PdCl2][PF6] 3. Both compounds were fully characterised by spectroscopic methods and the trinuclear complex was additionally characterised by X-ray diffraction.  相似文献   

13.
[Re(CO)6][BF4] reacts with HMPA to form [Re(CO)3(HMPA)3][BF4] (4), whose structure was determined by X-ray crystallography and proves to be a key intermediate in the ligand exchange reaction between three CO and Cp; and may be related to other cations such as [Re(CO)3(H2O)3]+, [Re(CO)3(CH3CN)3]+, [Re(CO)3(DMSO)3]+, obtained by different ways, and important in the field of organometallic radiopharmaceuticals.  相似文献   

14.
The cationic rearrangement of four homocubane bridgehead carbinols viz dimethyl 4-(1-bromopentacyclo[4.3.0.02,5.03,8.04,7]nonyl-9-one ethylene ketal) carbinol 2, diphenyl 4-(1-bromopentacyclo[4.3.0.02,5.03,8.04,7]nonyl-9-one ethylene ketal) carbinol 3, 4-(1-bromopentacyclo[4.3.0.02,5.03,8.04,7] nonyl-9-one ethylene ketal) carbinol 4 and 4-(1-bromopentacyclo[4.3.0.02,5.03,8.04,7]nonyl) carbinol 16, has been studied undervarious conditions.Exclusive migration of the C4C7 (or the equivalent C3C4 bond) in the homocubane skeleton was observed leading to 1,3-bishomocubane bridgehead alcohols. Relief of cage constraint governs the selective course of these cage expansions.  相似文献   

15.
《Tetrahedron》1986,42(22):6101-6109
The crystal structures of 2,7-diazatetracyclo[6.2.2.23,6.02.7]tetradec-4-ene, 2, its cation radical nitrate salt 2+, NO3-, 2,7-diazatetracyclo[6.2.2.23,6.03,7]tetradecane, 3, its dication dihexafluorophosphate salt 32+(PF6-)2, and a low quality structure of the monocation radical tosylate salt of 3 are reported and compared with MNDO calculations of these structures. Cations 2+ and 3+ are found to be significantly syn bent at nitrogen, and the dication 32+ has a longer N-N distance than its azo analogue, 2,3-diazabicyclo [2.2.2]oct-2-ene (11).  相似文献   

16.
Shin-ichi Naya 《Tetrahedron》2004,60(41):9139-9148
Ring transformation of 7,9-dimethylcyclohepta[b]pyrimido[5,4-d]furan- 8(7H),10(9H)-dionylium tetrafluoroborate 4+·BF4 to 7,9-dimethylcyclohepta[b]pyrimido[5,4-d]pyrrrole-8(7H),10(9H)-dionylium tetrafluoroborate 6a-d+·BF4 consists of the reaction of 4+·BF4 with amines and subsequent exchange of the counter-ion using aq. HBF4. Reactions of 4+·BF4 with aniline and 4-substituted anilines afforded the corresponding pyrrole derivatives 6a-c+·BF4 directly in good yields. On the other hand, reaction of 4+·BF4 with benzylamine gave the troponeimine intermediate 9, which was not converted to 6d+·BF4 and reverted to 4+·BF4 by adding HBF4; however, it was converted to 6d+·BF4 upon treatment with (COCl)2 or SOCl2, followed by exchange of the counter-ion. In a search for the characteristics of 9, inspection and comparison of the X-ray crystal analyses, NMR and UV-vis spectra, and CV measurement of 9 and N,N-disubstituted troponeimine derivatives 12 were carried out to suggest the remarkable structure of 12 having ionic C-O bonding between the imine-carbon atom and the oxygen atom of the barbituric acid moiety in the solid state. Thus, characteristics of 9 were ascribed to the sterically hindered and favorable conformation of N-protonated troponeimine intermediates. Furthermore, novel photo-induced oxidation reactions of a series of 4+·BF4, 5+·BF4, and 6a,e+·BF4 towards some amines under aerobic conditions were carried out to give the corresponding imines in 455-8362% yields [based on compounds 4+, 5+, and 6a,e+], suggesting the oxidation reaction occurs in an autorecycling process. Mechanistic aspects of the amine-oxidation reaction are also postulated.  相似文献   

17.
The reaction of H2[OsBr6] with DMSO in ethanol solution resulted in DMSO complex [H(dmso-O)2][OsIII(dmso-S)2Br4] (1) described previously as an intermediate product in the reaction of K2[OsBr6] with DMSO and characterized by EAS and ESR spectra. The coordination of DMSO molecules was established by IR and 1H and 13C NMR spectroscopy. The oxidation state of osmium and trans arrangement of DMSO molecules in the anion were established by ESR. The behavior of complex 1 in solutions was studied by EAS, ESR, and mass-spectrometry: a displacement of Br? ions accompanied by the reduction of osmium to oxidation state +2 occurs in DMSO, a solvation with displacement of DMSO molecules is observed at the first stage in water and methanol (rate constants 2.3 × 10?4 and 1.7 × 10?3 s?1, respectively), the sequential substitution of DMSO molecules and osmium oxidation to form [OsIVBr6]2? ions takes place in 4 mol/L HBr.  相似文献   

18.
The ammonia desorption chemical ionization (NH3-DCI) mass spectra of peracetylated gentiobiose (1) and two isotopically labelled gentiobioses (2 and 3) were examined. Compound 2 is labelled with trideuteroacetyl groups in the non-reducing moiety and 3 with trideuteroacetyl groups in the reducing moiety. It is shown that the [M + NH4 – 42]+ ion is not formed direct from [M + NH4]+ by loss of ketene but appears to be formed by way of a nucleophilic acyl substitution reaction resulting in a neutral species which complexes with NH4+. The disaccharides undergo cleavage at either side of the glycosidic oxygen joining the two sugar residues, a process which is accompanied by addition of H or CH3CO to afford neutral species which complex with NH4+. The structures of the ions resulting from H transfer have been inferred by comparison of their mass-analysed ion kinetic energy (MIKE) spectra with MIKE spectra of the [M + NH4]+ ions of compounds of established structure. A ring fragmentation reaction of 1, 2 and 3 is reported.  相似文献   

19.
Reaction of [2,6-(MeOCH2)2C6H3]Li (1) with SbCl3 in 1:1 molar ratio yielded except the intended product [2,6-(MeOCH2)2C6H3]SbCl2 (2) unexpected complex 3 consisting of antimony anion [Sb6Cl22]4− compensated by four intramolecularly coordinated organoantimony cations [2,6-(MeOCH2)2C6H3]2Sb+. Compound 3 is labile in CH2Cl2(CHCl3) solution and decomposes to compound 2 and SbCl3. Both compounds were characterized by the help of 1H and 13C NMR spectroscopy, ESI-mass spectrometry and in the case of 3 by single crystal X-ray diffraction techniques.  相似文献   

20.
Ab initio molecular orbital calculations with moderately large polarization basis sets and including valence-electron correlation have been used to examine the structure and dissociation mechanisms of protonated methanol [CH3OH2]+. Stable isomers and transition structures have been characterized using gradient techniques. Protonated methanol is found to be the only stable isomer in the [CH5O]+ potential surface. There is no evidence for a tightly-bound complex, [HOCH2]+…?H2, analogous to the preferred structure [CH3]+…?H2 of [CH5]+. Protonated methanol is found to possess a pyramidal arrangement of bonds at the oxygen atom with a barrier to inversion of 8kJ mol?1. The lowest energy fragmentation pathways are dissociation into methyl cation and water (predicted to require 284 kJ mol?1 with zero reverse activation energy) and loss of molecular hydrogen (endothermic by 138 kJ mol?1 but with a reverse activation barrier of 149 kJ mol?1). The results offer a possible explanation as to why production of [CH2OH]+ from the reaction of methyl cation with water is not observed. Other dissociation processes examined include loss of a hydrogen atom to yield the methylenoxonium radical cation or methanol radical cation (requiring 441 and 490 kJ mol?1, respectively) and loss of a proton to yield neutral methanol (requiring 784 kJ mol?1).  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