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 171 毫秒
1.
The rate constants (log k) for solvolysis of 2-, 3-, and 4-substituted bicyclo[2.2.2]octyl p-nitrobenzenesulfonates 10, 11 and 12 , respectively, correlate linearly with the corresponding inductive substituent constants s?. The formation of the ion pairs 9 is, therefore, controlled by the I effect of neighboring substituents. It follows from the corresponding reaction constants ρ1 of ?1.54, ?1.12, and ?1.22 that inductivity is highest at the positions α to C(1). It is lower and practically equal at the β- and γ-positions. Therefore, charge dispersal is similar to that previously observed in the quinuclidinium ion 7 .  相似文献   

2.
The present paper describes the preparation by conventional methods (when not available commercially) and the pKa-determination of the α-, β- and γ-isomers of pyridylethanamide, 3-pyridylpropanamide. 4-pyridylbutanamide, 5-pyridyl-pentanamide, pyridylmethanol, 2-pyridylethanol, 3-pyridylpropanol, 4-pyridyl-butanol, 5-pyridylpentanol, pyridylmethylamine, 2-pyridylethylamine, 3-pyridyl-propylamine, 4-pyridylbutylamine, and 5-pyridylpentylamine. While a field effect accounts for many variations in pKa as a function of chain length, marked inductive effects are operative in some methyl and ethyl homologs. The pKa-decreasing influence of an intramolecular H-bond is also apparent in some lower homologs belonging to the α-series.  相似文献   

3.
pKa Values of 42 quinuclidinium perchlorates I have been measured in 0.1 M aqueous KCl-solution. In a few cases small corrections of earlier thermodynamic pKa values are indicated. These measurements, in conjunction with recent X-ray structure determinations, confirm the reliability of the inductive substituent constants σIq derived from them.  相似文献   

4.
The non-isothermal behavior of two binary CdO—persulfate systems has been investigated. The molar ratios and TiTf are established. The temperatures for the α- to β-Na2SO4 phase transition, as well as for α- to β- to γ-CdSO4 of the CdONa2S2O8 system have been fixed. No evidence for the occurrence of the β- to γ-CdSO4 polymorphic trans-formation has been obtained from the reaction of the CdOK2S2O8 system. This is because of the formation of a CdSO4/K2SO4 eutectic mixture which melts at 653°C, i.e., before the β- to γ-phase change transition, which usually occurs later. No basic cadmium sulfate has been identified. The excess cadmium oxide acts as a p-type semiconductor which accelerates the thermal decomposition of pyrosulfates.  相似文献   

5.
Thermodynamic pKa-values have been determined for 38 4-substituted quinuclidinium perchlorates. They are remarkably sensitive to the polar effect of the substituent and cover a range of 3.63 pKa units. Furthermore, they vary linearly and almost equally with temperature since the contribution of the TΔS° term to the free energy of ionization is relatively small and constant. The magnitude of the polar effect of the 4-cyano group varies with the solvent and appears to depend on its ability to form hydrogen bonds to the substituent rather than its dielectric constant. New inductive substituent constants σIq are derived from the pKa values. Their correlation with known inductive constants is only fair or unsatisfactory, especially as regards the relative order of hydrogen and the alkyl groups. The discrepancies can be ascribed mainly to the different models used to derive the substituent constants.  相似文献   

6.
《Tetrahedron letters》2017,58(26):2525-2529
An efficient, direct nucleophilic allylic substitution of α-, β- and γ-substituted alcohols with enamines, using the Pd(OAc)2/PPh3 catalyst system and ZnBr2 as a promoter in CH2Cl2 at reflux, is reported. The reaction course was dependent on the steric hindrance at the α- or γ-positions with respect to the functionalized α-carbon, selectively affording in moderate to good yields, α- or γ-homoallyl ketones, the so-called “linear” and “branched” products, respectively.  相似文献   

7.
2, 3, 4-Tri-O-benzyl-6-O trimethylsilyl-α-D-glucopyranosyl bromide reacted with RSH compounds (R = Et, t-Bu, PhCH2, p-PhBut, Ph, 2-benzoxazole, 2-benzothiazole), with diisopropylethylamine as base, to give S-D-glucosides. The α- and β- anomeric ratio was dependent on the pKa. value of the thiol compound. If pKa value was less than or equal to 7, β-anomer was obtained. If pKa value was equal to or more than 11, α anomer was obtained. If pKa value was between 9 and 10.6, a mixture of α- and β-anomer was obtained. The probable mechanism was discussed. The chemical shifts of proton and carbon in these S-D-glucosides were measured and verified by 2D NMR (cosy and hetcor).  相似文献   

8.
The aim of this work was to gain a deeper understanding of salt effects in the inclusion complex formation of cyclodextrins. For this purpose, thermodynamic study of complex formation of α-, β- and γ-cyclodextrins with p-aminobenzoic acid was carried out in water and solutions of KCl, KBr, KH2PO4 and K2SO4 (0.2 mol/kg). Stability constants were calculated from the binding isotherms obtained on the basis of 1H NMR measurements. Enthalpy and entropy of complex formation were estimated from the van’t Hoff plots. It was found that effects of KCl, KH2PO4 and K2SO4 are insignificant, while the influence of KBr on complex formation of cyclodextrins with p-aminobenzoic acid is more pronounced and results in a decrease of the stability constants. Specific action of Br is caused by the ability of these anions to penetrate into macrocyclic cavity. Coexistence of two complexation equilibria in KBr solution is accompanied by significant solvent reorganization originated from more intensive dehydration of the interacting species. This results in an increase of the enthalpy and entropy of complex formation. Manifestation of Br effect was found to be the same in the binding of p-aminobenzoic acid with α-, β- and γ-cyclodextrins.  相似文献   

9.
The reactions of quinuclidines with phenyl, 4-methylphenyl, and 4-chlorophenyl 2,4-dinitrophenyl carbonates are kinetically evaluated in aqueous solution. The Brønsted-type plots (log kN vs pKa of quinuclidinium ions) are linear. The magnitude of the slopes and validated theoretical scales of electrophilicity and nucleophilicity confirm the concerted nature of these reactions.  相似文献   

10.
Bromophilic attack by the transition metal carbonyl anion, [Re(CO)5]Na (pKa = 21.1), on 2-(1-bromoalkylidene)thiazolidin-4-ones is significantly faster than abstraction of an acidic lactam hydrogen (pKa ∼17-18), when the generated carbanion is stabilized by an α-CN or α-PhCO group. The bromophilic reaction of 2-(1-bromoalkylidene)thiazolidin-4-one, having an α-CN electron-withdrawing group, resulted in formation of a new metallacyclic anionic complex. With less reactive vinyl bromides, containing an α-CONHPh or α-CO2Et group, only deprotonation is observed. The role of the metal carbonyl anion is highlighted by a comparison with the 9-methylfluorenide carbanion (pKa of 9-methylfluorene is 22.3), which reacts exclusively via a deprotonation pathway.  相似文献   

11.
The epoxidation of alkenylphosphorus compounds with hydrogen peroxide was systematically studied, revealing that while alkenylphosphine oxides failed to produce the corresponding epoxides, alkenylphosphonates, or phosphinates having a phenyl group at α-position reacted with H2O2/K2CO3 or alkenylphosphonic acids or phosphinic acids having an aliphatic group at α- or β-positions reacted with H2O2/Na2WO4/Et3N to produce high yields of the corresponding epoxides.  相似文献   

12.
Several isomers are possible when N4-tetradentate ligands coordinate to form metal complexes. Here we report the synthesis and structural analysis of cis-β-{[1,6-di(2′-pyridyl)(2,5-dibenzyl-2,5-diazahexane)(1,2-benzoquinone diimine)]ruthenium(II)} formed exclusively from the β-precusor, β-{[1,6-di(2′-pyridyl)(2,5-dibenzyl-2,5-diazahexane) (dimethylsulfoxide)chloride] ruthenium(II)} hexaflourophosphate. Ruthenium(II) complexes synthesised from 1,6-di(2′-pyridyl)-2,5-dibenzyl-2,5-diazahexane, produce only two isomers which can be separated by recrystallisation into α- and β-[Ru(picenbz2)(dmso)Cl]PF6 (where picenbz2 is 1,6-di(2′-pyridyl)-2,5-dibenzyl-2,5-diazahexane). The distinctively different proton NMR spectra of the isomers are an especially convenient feature with which to assess separation. Isomeric structure of the precursor, α or β, is conserved upon coordination of a bidentate ligand, such as benzene-1,2-diamine, 4,5-dimethyl-benzene-1,2-diamine, naphthalene-2,3-diamine, 2,2′-bipyridine, 1,10-phenanthroline or dipyrido[3,2-d:2′3′-f] quinoxaline, to produce complexes of the type α- or β-[Ru(picenbz2)(bidentate)](PF6)2. The synthesis, separation and characterisation of the α- and β-precursors and the α- and β-[Ru(picenbz2)(bidentate)](PF6)2 complexes are reported. Moreover, the crystal structures have been determined for β-[Ru(picenbz2)(dmso)Cl]PF6.0.5H2O (C30H37N4O1.5F6PSClRu); it is triclinic, space group P 1, a?=?9.987, b?=?12.883, c?=?14.287?Å, α?=?72.11, β?=?78.65, γ?=?88.39° and Z?=?2 and β-[Ru(picenbz2)(bqdi)](PF6)2, (C34H38N6F12P2Ru) which is triclinic, space group P 1, with a?=?10.129, b?=?10.338, c?=?19.587?Å, α?=?104.42, β?=?93.36, γ?=?92.00° and Z?=?2. The structures were determined at room temperature and refined by least-squares methods to R?=?0.044 for 5109 and R?=?0.075 for 3057 non-zero diffractometer data, respectively, for the dmso and bqdi species above.  相似文献   

13.
In the cobalt oxide-molybdenum trioxide-alumina system with a molar ratio of 1:1:1, the amounts of the high-temperature modification (a) of CoMoO4 formed during heating from 500 to 800°C and the low-temperature modification (b) formed by phase transition during the subsequent cooling to room temperature are influenced by the kinds of alumina used, such as α-, γ- and calcined γ-aluminas. Powder X-ray diffraction analysis revealed that in an α-alumina system formation of a-CoMoO4 is most favorable at a calcination temperature of 500°C and phase transition from a- to b-CoMoO4 during cooling is enhanced by higher calcination temperatures. In the γ-alumina system, formation of a-CoMoO4 is slight at 500°C but increases with increase in the calcination temperature, as does slightly the degree of phase transition from a- to b-CoMoO4 upon cooling. In a system containing calcined γ-alumina, formation of a-CoMoO4 similar to α- and γ-alumina systems was observed to occur at 500°C and 800°C, respectively, together with phase transition to b-CoMoO4 during cooling. The degree of dispersion in the CoOMoO3 coexistent system is affected by the particle size of aluminas, such as coarse α-, fine amorphous γ- and calcined γ-alumina consisting of both sizes, as observed with electron microscopy. Presence of finer γ-alumina is considered to suppress or retard the solid state reaction and phase transition.  相似文献   

14.
A simple three-step procedure converted the readily accessible (2-bromo-1,1-difluoroethyl)arenes (2) into α-aryl-α,α-difluoroacetaldehydes (1). Subsequent hydrocyanation, hydrolysis, oxidation and again hydrolysis afforded β-aryl-β,β-difluoro-α-oxopropionic acids (3). Reductive amination transformed the oxoacids 3 into a separable mixture of α-hydroxyacids 11 and racemic β,β-difluoro-β-phenylalanine derivatives (4). Enantiomerically pure β,β-difluorophenylalanine (l-4a) was obtained when α,α-difluoro-α-phenylacet-aldehyde (1a) was condensed with homochiral 1-phenylethylamine, hydrogen cyanide added to the resulting imine, the diastereomeric mixture thus produced hydrolyzed to the carboxamides (15) which were found to be separable by fractional crystallization or chromatography. The pKa values of the β-aryl-β,β-difluoroalanines (4) were measured and biological profile of the latter probed. 3-(4-Chlorophenyl)-3,3-difluoro-2-oxopropionic acid (4c) proved to be a potent (Ki 27 μM) and selective inhibitor of arogenate dehydratase, a key enzyme catalyzing the last step of the phenylalanine biosynthesis.  相似文献   

15.
In exploring the capability of nuclear magnetic resonance (NMR) spectroscopy for pomegranate juice analysis, the eight aromatic singlet resonances of α- and β-punicalagin were clearly identified in the 1H NMR spectra of juice samples. The four downfield resonances were found to be sensitive to small pH changes around pH 3.50 where the NMR spectra of the juice samples were recorded. To understand this unusual behavior, the 1H and 13C resonance assignments of the punicalagin anomers were determined in aqueous solution and pH titrations with UV and 1H NMR detection carried out to characterize the acid–base properties of punicalagin over the pH range 2–8. Simultaneous fitting of all of the pH-sensitive 1H NMR signals produced similar but significantly different pK a values for the first two deprotonation equilibria of the gallagic acid moiety of the punicalagin α- (pK a1?=?4.57?±?0.02, pK a2?=?5.63?±?0.03) and β- (pK a1?=?4.36?±?0.01, pK a2?=?5.47?±?0.02) anomers. Equivalent pK a values, (α?:?6.64?±?0.01, β?:?6.63±?0.01) were measured for the third deprotonation step involving the ellagic acid group, in good agreement with a prior literature report. The punicalagin anomer equilibrium readjusts in parallel with the proton dissociation steps as the pH is raised such that β-punicalagin becomes the most abundant anomer at neutral pH. The unusual upfield shifts observed for the glucose H3 and H5 resonances with increasing pH along with the shift in the α/β anomer equilibrium are likely the consequence of a conformational rearrangement.
Figure
Titration of the punicalagin phenolate protons over the pH range 2–8 results in changes in the aromatic proton chemical shifts and a readjustment of the anomer equilibrium.  相似文献   

16.
Series of photochromic 1,2-bis(thienyl)ethenes possessing perfluorocyclopentene backbones, either hydrogen or methyl groups at the β-positions of the thiophenes, and a variety of substituents in their α'-positions were prepared, which cover the range from electron-donating to electron-withdrawing (Me, −CH2OH, −OTBS, −TMS, −Br, 1,3-dioxan-2-yl, pyridin-4-yl, −CH2OH, −COOH). As a linear free energy relationship the spectroscopic Hammett equation gives fair to excellent fits to the excitation energy of the absorption maxima of the ring-opened as well as the ring-closed forms of the BTEs, when Hammett substituent constants σp were replaced by Brown's modified substituent constants σp+ and σp. Vice versa, hitherto unknown Hammett-Brown substituent constants can be estimated from the UV spectra. Furthermore, we compared the experimentally measured absorption maxima with values which we calculated by three different methods (DFT STEOM-DLPNO-CCSD/def2-TZVPP, TD-DFT ωB97X-D3/6-31G*, TD-DFT ωB97X-D3/6-311++G**).  相似文献   

17.
Methyl β-methoxyacrylate (1) can be lithiated successively in β- and α-positions. Reaction with two electrophiles (at first a carbonyl compound) leads to α, γ-substituted methyl tetronates in a two step synthesis. Application of this method to the synthesis of methyl tetronate whose structure was assigned as that of gregatin B indicates that the gregatins and aspertetronins have the isomeric structures 17.  相似文献   

18.
Nine new polynuclear 2-(5-tetrazolyl)ethyl podands have been obtained by the azidation of the corresponding nitriles. Using Bjerrum distribution functions, the values of pK a 1, pK a 2, pK a 3, and pK a 4 have been determined by a potentiometric method for 14 polynuclear tetrazoles in aqueous and aqueous methanolic solution. The found values lie in the range from 3.5 to 7.5 pH units. The overall rules and the sequence of the ionization of the spatially separated tetrazole fragments in these podand systems are discussed.  相似文献   

19.
Abstract— The lowest excited singlet-state dissociation constants (pKSa) of bromosubstituted pyridines, quinolines, and isoquinolines were determined from the pH-dependent shifts in their electronic absorption spectra. The lowest excited triplet-state dissociation constants (pKTa) of bromosubstituted quinolines and 4-bromoisoquinoline were obtained from the shifts of the 0–0 phosphorescence bands measured in rigid aqueous solution at 77 K. The pKSa values indicate that the basicity of these brominated nitrogen heterocycles is increased in the lowest excited singlet state by 2 to 10 orders of magnitude as compared with the ground state. The pKTa values are found to be significantly different from the corresponding ground-state pKa values, indicating that the basicity of bromoquinolines is increased in the lowest excited triplet state by 1.7 to 3.0 pK units. The enhancement of the excited singlet-and triplet-state basicity of brominated nitrogen heterocycle derivatives as compared with the unsuhstituted parent compounds is attributed to the increased electron-donor conjugative interactions of the bromine atom pπ orbitals with π orbitals in the lowest excited singlet and triplet state.  相似文献   

20.
Heterylation of 3-R1-5-R2-1'2'4-triazoles (pK a 3-12) with N-alkyl-, N-alkenyl-, N-alkoxy-carbonyl-, N-oxoalkyl-, N-nitroxyalkyl, N-nitroaminoalkyl-3'5-dinitro-1'2'4-triazoles results insubstitution of a nitro group in 5 position of the dinitro compound yielding 1-R-methyl-3-nitro-5-(3-R1-5-R2-1,2,4-triazolyl)-1,2,4-triazoles. The side processes: Hydroxide-ion attack on C5 and (or) N1 of the ring both in the substrate and in the target compound afford 1-R-methyl3-nitro-1,2,4-triazol-5-ones, 3,5-dinitro-1,2,4-triazole and NH-acids of N-C-bitriazole series. Optimal reaction media are aprotic dipolar substances, and for compounds prone to heterolysis ethyl acetate-water systems. The azole pK a is the decisive factor controlling the composition and the ratio of reaction products. The process is promising for azoles with pK a > 5, and the optimal range of pK a is 8-10.  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