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.
The stabilizing neighboring effect of halo substituents on silyl cations was tested for a series of peri-halo substituted acenaphthyl-based silyl cations 3 . The chloro- ( 3 b ), bromo- ( 3 c ), and iodo- ( 3 d ) stabilized cations were synthesized by the Corey protocol. Structural and NMR spectroscopic investigations for cations 3 b – d supported by the results of density functional calculations, which indicate their halonium ion nature. According to the fluorobenzonitrile (FBN) method, the silyl Lewis acidity decreases along the series of halonium ions 3 , the fluoronium ion 3 a being a very strong and the iodonium ion 3 d a moderate Lewis acid. Halonium ions 3 b and 3 c react with starting silanes in a substituent redistribution reaction and form siliconium ions 4 b and 4 c . The structure of siliconium borate 4 c 2[B12Br12] reveals the trigonal bipyramidal coordination environment of the silicon atom with the two bromo substituents in the apical positions.  相似文献   

2.
On the Photochemistry of 1, 2-Benzisoxazoles in Strongly Acidic Solution The 1, 2-benzisoxazoles 1a, 1b and 1d when dissolved in 96% sulfuric acid and irradiated through a quartz filter with a mercury high-presure lamp yield, after work-up, mixtures of 2, 5- and 2, 3-dihydroxy-acylbenzenes ( 2 and 3 , respectively; cf. Schemes 1 and 3 and Table 1). Irradiation of 3, 5-dimethyl-1, 2-benzisoxazole ( 1c ) in 96% sulfuric acid leads to the formation of 2, 3-dihydroxy-5-methyl-acetophenone ( 3c ) in only 6% yield (cf. Table 1). It is assumed that the 1, 2-benzisoxazolium ions react in the excited singlet state by heterolytic cleavage of the N, O-bond to yield the corresponding aryl oxenium ions 7 in the singlet ground state (see Scheme 5). Reaction of 7 with HSO 4 ? ions, present in 96% sulfuric acid, yields, after hydrolysis, the dihydroxy compounds 2 and 3 . Photolysis of 3-methyl-1, 2-benzisoxazole ( 1b ) in diluted sulfuric acid (0,5 to 9 M ) in methanol or water leads only to the formation of 2-amino-phenol ( 6 ; see Scheme 3), presumable via photo-isomerization of 1b to 2-methylbenzoxazole ( 5b ) which then is hydrolyzed to give 6 .  相似文献   

3.
The gas phase chemistry of C3H6Br+ cations generated via low energy electron impact on various dibromopropanes has been studied by using Fourier transform ion cyclotron resonance mass spectrometry. Neutral substrate molecules that have been selected to probe the bimolecular reactivity of the C3H6Br+ isomers are ammonia, methylamine, trimethylamine, cis-butene, and 2, 3-dimethyl-2-butene. At least three different isomers are characterized on the basis of their different reactivity toward the various substrate molecules. It is suggested that these isomers have (a) the 2-bromo-2-propyl cation structure, (b) the propylenebromomum ion structure, and (c) the cyclic four-membered trimethylenebromonium ion structure. The 2-bromo-2-propyl cations react predominantely via proton transfer. This reaction is hampered for the propylenebromonium ions, which react mainly as electrophiles or bromanyl cation donors. Cyclic trimethylenebromoruum ions react predominantly via adduct formation, even under low pressure conditions, which implies that tturd body collisions are not the only stabilization mechanism.  相似文献   

4.
Electrodialytic behavior of cation exchange membrane having cationic polyelectrolyte layer on its surface (relative transport number between two cations, P, current efficiency, and electric resistance of the membrane during the electrodialysis) was measured using various cations, and the mechanism of selective permeation of lower-valent cations than the higher and of larger hydrated cations than the smaller was investigated. The following conclusions were obtained: (1) the cationic polyelectrolyte layer narrowed the pathway for ions; (2) when the pathway was made narrower by other methods, the remarkable change of the permselectivity of the membrane did not occur; (3) the effectiveness of the cationic polyelectrolyte layer was depressed by the treatment with anionic polyelectrolytes to neutralize the cationic charge on the membrane surface; and (4) the cation with larger hydrated diameter (potassium ions) more easily permeated through the membrane having the polyelectrolyte layer than that with the smaller hydrated diameter (sodium ions). It is concluded that the change of P by the cationic charge layer on the membrane surface is based on the difference in the strength of the electrostatic repulsion between two cations against the cationic charge layer.  相似文献   

5.
Donor‐substituted diarylcarbenium ions Ar2CH+ react with ynamides to give 1‐amido‐substituted allyl cations (α,β‐unsaturated iminium ions). Kinetic studies show that these adducts, which correspond to the addition of a C? H bond across the C?C bond, are formed stepwise with initial formation of keteniminium ions and subsequent 1,3‐hydride shifts. The linear correlations between the second‐order rate constants (lg k2, 20 °C) with the electrophilicity parameters E of the diarylcarbenium ions allow us to include ynamides in our comprehensive nucleophilicity scale and thus predict potential electrophilic reaction partners.  相似文献   

6.
Pyrrole ( 2a ) and 2,5-dimethylpyrrole ( 2b ) react with the dialkoxycarbenium tetrafluoroborates 1a-1c under kinetic control to yield the corresponding acylpyrrole derivatives. 5H-Dibenz[b,f]azepine ( 9a ) and the 10,11-dihydro derivative 9b react only with the most electrophilic of the series of electrophiles tested, namely, diethoxycarbenium tetrafluoroborate ( 1a ), to furnish the corresponding formyl derivatives. Similarly, in arene chemistry, the highly electron-rich N,N-dimethylaniline ( 13a ) and 1,3,5-trimethoxybenzene ( 13b ) are formylated by reaction with 1a .  相似文献   

7.
Results of polymerization and laser flash photolysis studies concerning the direct and indirect photoinitiation of cationic polymerizations using N-alkoxy pyridinium and N-alkoxy quinolinium salts are presented. The indirect action can be based on (a) the generation of carbocations via the oxidation of photochemically produced free radicals and (b) the generation of radical cations via the reaction of electronically excited sensitizers with pyridinium or quinolinium ions. With respect to (a) substituted vinyl bromides were found be an effective source of oxidizable free radicals (vinyl radicals), and regarding (b) singlet or triplet excited states of thioxanthone, anthracene, perylene and phenothiazine, which, in the ground state, strongly absorb light in the 300 to 400 nm range, were found to react rapidly with pyridinium salts.  相似文献   

8.
The epimeric bromoghydrins 13/14 as well as their ethers 5/6 rearrange in the presence of silver ions to the 5-hydroxypyridones 15 or their ketals 9 respectively. with the bromohydrin ethers 5/6a and b this transformation is effected also with sodium methoxide. Under these reaction conditions the ring expansion is presumend ot proceed via bicyclic aziridines of which 7 could be issolated, The N-methylated compounds 5c/6c which cannot react likewise, form with sodium methoxide the 6-methoxypyridone 12 .  相似文献   

9.
A direct MS/MS method for the ortho, meta or para configuration assignment of any single molecule that forms reference ions upon ionization and dissociation is demonstrated. Gas-phase structure diagnostic ion–molecule reactions with acetonitrile are shown to distinguish the isomeric 2-, 3- and 4-hydroxybenzoyl cations and the 2- from the 3- and 4-aminobenzoyl cations. These reference ions, which display indistinguishable 15 eV collision-induced dissociation product ion mass spectra, react with acetonitrile to yield characteristic ratios of product ions, most particularly for the 2-isomers. The reactivity of the 2-benzoyl cations is the most characteristic since the ortho configuration allows for [4+ + 2] polar cycloaddition that yields relatively stable heterocycles in N-protonated forms. Distinction of the reference isomeric 2-, 3- and 4-hydroxy- and aminobenzoyl cations permits, therefore, partially or completely, direct ‘MS-only’ positional assignment of either ortho, meta or para configuration for any single molecule that forms such reference ions upon ionization and dissociation. This “class-universal” method for direct MS assignment of a single positional isomer should therefore be applicable to many members of the homologous series of isomeric ortho, meta and para acyl and amidyl anilines and phenols and derivatives. Such molecules dissociate, or are likely to dissociate, after or during ionization processes to form the reference and structurally diagnostic ortho, meta or para hydroxy- or aminobenzoyl cations. Copyright © 2004 John Wiley & Sons, Ltd.  相似文献   

10.
Counter-cations are essential components of polyoxometalates (POMs), which have a distinct influence on the solubility, stabilization, self-assembly, and functionality of POMs. To investigate the roles of cations in the packing of POMs, as a systematic investigation, herein, a series of triol-ligand covalently modified Cu-centered Anderson-Evans POMs with different counter ions were prepared in an aqueous solution and characterized by various techniques including single-crystal X-ray diffraction. Using the strategy of controlling Mo sources, in the presence of triol ligand, NH4+, Cu2+ and Na+ were introduced successfully into POMs. When (NH4)6Mo7O24 was selected, the counter cations of the produced POMs were ammonium ions, which resulted in the existence of clusters in the discrete state. Additionally, with the modulation of the pH of the solutions, the modified sites of triol ligands on the cluster can be controlled to form δ- or χ-isomers. By applying MoO3 in the same reaction, Cu2+ ions served as linkers to connect triol-ligand modified polyanions into chains. When Na4Mo8O26 was employed as the Mo source to react with triol ligands in the presence of CuCl2, two 2-D networks were obtained with {Na4(H2O)14} or {{Na2(H2O)4} sub-clusters as linkers, where the building blocks were δ/δ- and χ/χ-isomers, respectively. The present investigation reveals that the charges, sizes and coordination manners of the counter cations have an obvious influence on the assembled structure of polyanions.  相似文献   

11.
The binding properties of three p-tert-butyldihomooxacalix[4]arene tetraketone derivatives (tert-butyl 2b, adamantyl 2c and phenyl 2d) in the cone conformation and one derivative (methyl 2a) in a partial cone conformation, towards alkali and alkaline earth metal cations have been established by extraction studies of metal picrates from water into dichloromethane, stability constant measurements in methanol and acetonitrile, and by 1H NMR spectrometry. Transport experiments of metal picrates through a dichloromethane membrane were also performed. The results are compared to those obtained with closely-related calix[n]arene derivatives (n = 4 and 5) and discussed in terms of the substituents, size and conformational effects. Methylketone 2a is a poor binder for all the cations studied, due to its partial cone conformation. Ketones 2b, 2c and 2d show high extraction and complexation levels for the alkali cations, with similar profiles and preference for K+ and Na+ (plateau selectivity). Towards alkaline earth cations, these ketones show a strong peak selectivity for Ba2+ in extraction, but a plateau selectivity for Ca2+, Sr2+ and Ba2+ in complexation. The nature of the substituent attached to the ketone function has some influence on their binding properties, with phenylketone 2d being a slightly weaker binder than ketones 2b and 2c. 1H NMR titrations confirm the formation of 1:1 complexes between the ketones and the cations studied, also indicating that they should be located inside the cavity defined by the phenoxy and carbonyl oxygen atoms. Ketones 2b, 2c and 2d show transport rates that do not follow, in general, the same trends observed in extraction and complexation.  相似文献   

12.
The ion/molecule reactions of the molecular ion, the C3H ion, and the C3H ion obtained from 3-chloropropene. 1-bromopropene, 2-bromopropene, 3-bromopropene, bromocyclopropane. and 3-iodopropene have been studied with their neutral precursor in a Fourier-transform mass spectrometer (FT/ICR). The molecular ions react to yield primarily C6H except for the ion derived from 1-bromopropene that is unreactive. The kinetics of the 3-bromopropene molecular ion reveals that 18% of these ions must possess a different structure which is unreactive. The fact that C3H ions obtained from 2-bromopropene are the only ones to undergo proton transfer is taken as evidence that only this parent compound gives rise to 2-propenyl cations by low-energy electron impact. The C3H ions generated in these systems are shown to be roughly an equal mixture of propargylium ions that react to yield C6H and unreactive cyclopropenium ions.  相似文献   

13.
Peptides adducted with different divalent Group IIB metal ions (Zn2+, Cd2+, and Hg2+) were found to give very different ECD mass spectra. ECD of Zn2+ adducted peptides gave series of c-/z-type fragment ions with and without metal ions. ECD of Cd2+ and Hg2+ adducted model peptides gave mostly a-type fragment ions with M+• and fragment ions corresponding to losses of neutral side chain from M+•. No detectable a-ions could be observed in ECD spectra of Zn2+ adducted peptides. We rationalized the present findings by invoking both proton-electron recombination and metal-ion reduction processes. As previously postulated, divalent metal-ions adducted peptides could adopt several forms, including (a) [M + Cat]2+, (b) [(M + Cat – H) + H]2+, and (c) [(M + Cat – 2H) + 2H]2+. The relative population of these precursor ions depends largely on the acidity of the metal–ion peptide complexes. Peptides adducted with divalent metal-ions of small ionic radii (i.e., Zn2+) would form predominantly species (b) and (c); whereas peptides adducted with metal ions of larger ionic radii (i.e., Hg2+) would adopt predominantly species (a). Species (b) and (c) are believed to be essential for proton-electron recombination process to give c-/z-type fragments via the labile ketylamino radical intermediates. Species (c) is particularly important for the formation of non-metalated c-/z-type fragments. Without any mobile protons, species (a) are believed to undergo metal ion reduction and subsequently induce spontaneous electron transfer from the peptide moiety to the charge-reduced metal ions. Depending on the exothermicity of the electron transfer reaction, the peptide radical cations might be formed with substantial internal energy and might undergo further dissociation to give structural related fragment ions.  相似文献   

14.
Reactions of 5H,2λ5-Azaphospholes with Arylazocarbonitriles and Dialkyl Acetylenedicarboxylates Azaphospholes 1a – c react with activated arylazocarbonitriles to 1,5,2λ5-diazaphosphorines 2a – c and 3a – c . The reaction of 1a – c with diethyl or dimethyl acetylenedicarboxyiates yields 7H-1,4λ5-azaphosphepines 4a – c . The structures of 2b , 3a , and 4a are established by an X-ray diffraction analysis.  相似文献   

15.
16.
Aminophosphonium salts [Ph3PN(H)R]BPh4 ( 1 ) [R = C6H5CH2 ( 1a ), 4‐CH3C6H4CH2 ( 1b ), C6H5 ( 1c )] were obtained by allowing hydride IrHCl2(PPh3)2{P(OEt)3} to react first with triflic acid and then with the organic azide RN3. The compounds were characterized spectroscopically and by X‐ray crystal structure determination of [Ph3PN(H)CH2C6H4‐4‐CH3]BPh4 ( 1b ). A reaction path for the formation of aminophosphonium cations is also proposed.  相似文献   

17.
Synthesis of the title compounds was achieved using the anils 2a , 2b , 2c , 2d , 2e and 5a , 5b , 5c derived from the 4‐aminopyrazole 1 as starting materials. These compounds were allowed to react with mercaptoacetic acid in boiling dry benzene to afford the corresponding thiazolidinones and spiro‐thiazolidinones 3a , 3b , 3c , 3d , 3e and 6a , 6b , 6c , respectively. Pictet—Spengler reaction of the 4‐aminopyrazole hydrochloride 7 with aromatic aldehydes and cyclic ketones resulted in the formation of new pyrazolo[4,3‐e]pyrrolo[1,2‐a]pyrazines 8a , 8b , 8c , 8d , 8e and 9a , 9b , respectively. Other derivatives of pyrazolo pyrrolopyrazines 10 and 11 were obtained via the reaction of the amino derivative 1 with 1,1′‐carbonyldiimidazol and CS2, respectively. J. Heterocyclic Chem., (2011).  相似文献   

18.
The intrinsic gas-phase reactivity of cyclic N-alkyl- and N-acyliminium ions toward addition of allyltrimethylsilane (ATMS) has been compared using MS(2) and MS(3) pentaquadrupole mass spectrometric experiments. An order of electrophilic reactivity has been derived and found to agree with orders of overall reactivity in solution. The prototype five-membered ring N-alkyliminium ion 1a and its N-CH(3) analogue 1b, as well as their six-membered ring analogues 1c and 1d, lack N-acyl activation and they are, accordingly, inert toward ATMS addition. The five- and six-membered ring N-acyliminium ions with N-COCH(3) exocycclic groups, 3a and 3b, respectively, are also not very reactive. The N-acyliminium ions 2a and 2c, with s-trans locked endocyclic N-carbonyl groups, are the most reactive followed closely by 3c and 3d with exocyclic (and unlocked) N-CO(2)CH(3) groups. The five-membered ring N-acyliminium ions are more reactive than their six-membered ring analogues, that is: 2a > 2c and 3c > 3d. In contrast with the high reactivity of 2a, its N-CH(3) analogue 2b is inert toward ATMS addition. For the first time, the transient intermediates of a Mannich-type condensation reaction were isolated-the beta-silyl cations formed by ATMS addition to N-acyliminium ions-and their intrinsic gas-phase behavior toward dissociation and reaction with a nucleophile investigated. When collisionally activated, the beta-silyl cations dissociate preferentially by Grob fragmentation, that is, by retro-addition. With pyridine, they react competitively and to variable extents by proton transfer and by trimethylsilylium ion abstraction-the final and key step postulated for alpha-amidoalkylation. Becke3LYP/6-311G(d,p) reaction energetics, charge densities on the electrophilic C-2 site, and AM1 LUMO energies have been used to rationalize the order of intrinsic gas-phase electrophilic reactivity of cyclic iminium and N-acyliminium ions.  相似文献   

19.
On ultraviolet irradiation in toluene, cyclopent-2-en-one ( 1a ) and 3-methylcyclopent-2-en-one ( 1b ) undergo dimerisation (→ 3a, b, 4b ), whereas 3-t-butyl-cyclopent-2-en-one ( 1e ) and the bicyclo [3.2.1]octenone 2a only react with the solvent (→ 5e–7e, 8–10 ). The 3-ethyl-( 1c ) and 3-isopropyl-cyclopent-2-en-ones ( 1d ) afford by dimerisation products 3c, d, 4c and by reaction with the solvent 5c, d, 6c, d, 7d . The β-methoxy-enones 1f and 2b are unreactive under comparable irradiation conditions. The head-to-head cyclobutane dimer 3b (HH) by separate irradiation at < 3400 Å in toluene, is reconverted to 1b . By similar irradiation each of the head-to-tail dimers 3b (anti- and syn-HT) and also the unsaturated dimer 4b give a new isomer of unknown structure besides small amounts of monomer 1b and relatively large amounts of insoluble material. The photoreactions of 1b-e can be quenched by naphthalene. Stern-Volmer plots for the quenching of the dimers of 1b, 1c , and 1d , and of 3-t-butylcyclopentanone ( 5e ) and the dihydro dimer 7e are linear and within each experiment, the slopes satisfactorily coincide, whereas the slope for the α-benzyl ketone 6c is distinctly greater than that for the dimers 3c . A similar differentiation, although less pronounced, is found between the benzyl ketones 6d and 6e and the respective accompanying products.  相似文献   

20.
The enaminones 1b,d,f react with 4‐phenyl‐3‐methyl‐5‐pyrazoleamine 3a to yield the pyrazole derivatives 4a‐c that cyclised readily on reflux in pyridine solution in presence of hydrochloric acid to yield the pyrazolo[1,5‐a]pyrimidines 5a‐c. Similarly 3(5)‐amino‐1H‐triazole (3b) reacted with 1b,d,f to yield the triazolo[1,5‐a]pyrimidines 5d‐f. In contrast attempted condensation of the 5‐tetrazoloamine (3c) with 1a,d,e resulted in its trimerisation and only triaroylbenzene 8a,d,e was isolated. The reaction of 1a,b,d with anthranilonitrile 9a and the reaction of 1a‐c with the 2‐aminocyclohexene thiophene‐3‐nitrile 10a afforded the cis enaminones 11a‐c and 12a‐c. Similarly, reaction of 1a‐c with the methylanthranilate 9b and reaction of 1b,e with ethyl 2‐aminocyclohexene thiophene‐3‐carboxylate 10b afforded the cis enaminones 11d‐f and 12d,e respectively. Attempted cyclization of 11a‐c into quinoline failed. Successful cyclization of 11d into the quinolinone 13 could be affected, on heating for five minutes in a domestic microwave oven at full power. The reaction of 1a‐c,f with piperidine afforded the trans enaminones 14a‐d. Similarly, trans 14e was formed from the reaction of 1b with morpholine. The coupling reaction of 1b with excess of benzene diazonium chloride afforded the formazane 16. The enaminone 2 reacted with heterocyclic amines to yield the pyridones 17,18.  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