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.
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   

2.
1,3-Diaryl-2-propen-1-ones, I, reacted with guanidine hydrochloride (II) in the presence of 3 moles of sodium hydroxide to give the corresponding 2-amino-4,6-diarylpyrimidines, III. The structure and configuration of the products are based on chemical and spectroscopic evidence. The protonation constants of these compounds (series A and series B) have been determined in 50 volume percent ethanol-water medium. Excellent linear correlations are obtained when pKa values of the two series of 2-amino-4,6-diarylpyrimidines, IIIa-j and IIIk-r, are plotted against the substituent constant, σx, and the polar substituent constant, σ* xC6H4, for substituted phenyl groups. The pKa values have also been correlated with the extended Hammett equation. The correlation follows the equations: Series A; pKa = 3.273 - 0.820σI,X - 0.662σR,X Series B; pKa = 3.169 - 0.424σI,X - 0.137σR,X  相似文献   

3.
The synthesis and physicochemical properties of a range of 2‐ and 6‐amido‐3‐hydroxypyridin‐4‐ones are described. All the amido‐substituted 3‐hydroxypyridin‐4‐ones have lower pKa values than 1,2‐dimethyl‐3‐hydroxypyridin‐4‐one (deferiprone). This is due to the inductive effect of the amido group. Furthermore, the pKa values of the 3‐hydroxy group in 1‐nonsubstituted pyridinones are dramatically lower than those of the corresponding 1‐alkyl analogues, indicating that a strong hydrogen bond exists between the 2‐amido function and the 3‐oxygen anion, which stabilises the anion. As a result of the decreased competition with protons, the pFe3+ values of this group of molecules are higher than that of deferiprone. The distribution coefficients of these molecules are also increased despite the lack of a hydrophobic 1‐alkyl substituent and this is ascribed to the intramolecular hydrogen bond. X‐ray diffraction studies confirm the existence of the intramolecular hydrogen bond.  相似文献   

4.
Inductive charge dispersal to the α- β- and γ-positions of the solvated quinuclidinium ion has been examined by comparing the pKa and the derived inductivities ρI of several 2- 3- and 4-substituted quinuclidinium perchlorates 4, 5 , and 6 , respectively. The same inductivity is observed at the practically equidistant β- and γ-positions. It, therefore, appears that polar substituent effects are transmitted directly through the molecule. As expected, inductivity is considerably higher at the α-positions where through-bond and direct induction coincide. The fact that the pKa of all three series of salts correlate linearly with each other points to the common nature of these inductive electron displacements.  相似文献   

5.
Hydrolysis of bicyclo[2.2.2]octylp-nitrobenzenesulfonate ( 14a , X = p-NO2C6H4SO3), and nineteen 4-R-substituted derivatives 14b–14t in 70% aqueous dioxane yield the corresponding bicyclo[2.2.2]octanols 14 (X = OH), exclusively. The 7-center fragmentation to 1,4-dimethylidene-cyclohexane ( 15 ) is not observed. The logarithms of most of the rate constants, measured in 80% ethanol, correlate well with the corresponding inductive substituent constants σ of R. Hence, in these cases ionization rate is controlled by the inductive effect of R only. Poor correlations result when the substituents are potentially electrofugal groups, such as COO?, CH2OH, CH2NH2, CONH2 and H, the deviations from the inductive regression line corresponding to rate enhancements of 1.6 to 8. These exalted substituent effects are tentatively ascribed to extended hyperconjugation involving two σ-bonds. This study corroborates previous evidence that the inductive effect alone does not fully account for the polar effect of some substituents in reactions involving carbocations.  相似文献   

6.
Electronic spectra of 4-substituted pyridine N-oxides and their EDA complexes with iodine were studied. The substituent effect on the near u.v.1A1 intramolecular CT bands of the N-oxides and on the blue shifted iodine bands caused by CT complex formation are discussed in terms of a general equation, theoretically derived in order to describe the substituent effect on electronic spectra by means of substituent constants. The results are quite successful and supported by semi-empirical SCFMO-CI calculations. Based on the results mentioned above, the character of n-σ type N-oxide—iodine CT complexes is also examined. The complex formation constants (log K) and pKa values of the N-oxides correlate especially well, indicating that the CT interaction mechanism cannot be neglected in proton addition reactions such as hydrogen bonding and pKa values.  相似文献   

7.
Physicochemical properties of new reagents, azo-substituted pyrocatechol derivatives and their tin(II) complexes, are studied. The acid-base properties of the hydroxy groups (pKi, pKi), parameters of complex formation reactions (pH, temperature, time), and instability constants of the complexes formed (pK i) are determined. Quantitative correlations between the dissociation constants (pKa) of the functional analytical group, and the electronic Hammett constant σ for a substituent (pKa-pH50 of the complex formation reaction), as well as between pKa and instability constants of the complexes (pK a), are established. The quantitative correlations established allow the prediction of the physicochemical properties of the reagents and tin(II) complexes with new reagents of this class with the same functional analytical group (FAG) but other substituents.  相似文献   

8.
The electronic effect of substituents on the acid-base properties of 6- and 7-substituted 2-thionolepidines and 4-thionoquinaldines were investigated. It is shown that the pK a (–H+) values for 6- and 7-substituted 2- and 4-thioquinolones and the pK a (+H+) values for 7-substituted 2-thionolepidines and 6-substituted 4-methylmercaptoquinaldines correlate with the M substituent constants of Jaffe and Taft. The effect of a substituent is transmitted primarily via an inductive mechanism, regardless of its position.Translated from Khimiya Geterotsiklicheskikh Soedinenii, No. 3, pp. 399–402, March, 1974.  相似文献   

9.
The rate constants (log k) for the solvolysis of 4e-substituted 2e- and 2a-adamantyl p-nitrobenzenesulfonates 14 and 15 , respectively, in 80% EtOH correlate linearly with the respective inductive substituent constants σ. Therefore, relative rates are controlled by the I effect of the substituents at C(4). The derived reaction constants, or inductivities, ρI of −0.80 and −0.64 for the series 14 and 15 , respectively, are far smaller than those previously determined for 6-substituted 2-norbornyl and 2-bicyclo[2.2.2]octyl sulfonates, in which the partial structure containing the substituent and the leaving group is the same. The ratio of the retained and inverted adamantanols obtained upon hydrolysis of the series 14 falls from 2.85 for R = CH3 to ca. 1 for R = CN, i.e. as the substituent at C(4) becomes more electron-attracting. In the 2a-series 15 this ratio is uniformly higher. These findings confirm that the 2-adamantyl cation is weakly bridged and that through-space induction in carbocations involves graded bridging of the cationic center by neighboring C-atoms.  相似文献   

10.
The influence of substitution of amidine group on tautomeric equilibria constants and basicities is discussed. Equations based on correlation analysis methods are derived enabling predictions of both, microscopic pKa, values of individual tautomers, measured macroscopic pKa values of the tautomeric mixture, as well as the tautomeric equilibrium constant (as pKT). It is shown that pKarn values of unsymmetrically N,N'-disubstituted amidines should obey a non-linear relation with σ° constants, and only for symmetrically N,N'-disubstituted amidines obey the linear Hammett equation. Tautomeric equilibrium constants of N,N'-disubstituted amidines correlate withσ° substituent constants. The methods of prediction of pKa value of both tautomers and pKT value are proposed.Derived relations are applied to the series of N,N'-diphenylacetamidines and benzamidines.  相似文献   

11.
The dissociation constant (pKa) of a drug is a key parameter in drug discovery and pharmaceutical formulation. The hydroxy substituent has a significant effect on the acidity of hydroxycinnamic acid. In this work, the acidic constants of coumaric acids are obtained experimentally by spectrophotometry using the chemometric method and calculated theoretically using ab initio quantum mechanical method at the HF/6‐31G* level of theory in combination with the SMD continuum solvation method. Rank annihilation factor analysis (RAFA) is an efficient chemometric technique based on the elimination of the contribution of one of the chemical components from the data matrix. RAFA cannot be performed because the pure spectrum of HA? is not available. So, two‐rank annihilation factor analysis (TRAFA) is proposed for the determination of the pKa OF H2A. A comparison between the pKa values obtained previously by TRAFA for the molecules o‐coumaric acid (4.13, 9.58), m‐coumaric acid (4.48, 10.35), and p‐coumaric acid (4.65, 9.92) makes it clear that there is good agreement between the results obtained by TRAFA and ab initio quantum mechanical method.  相似文献   

12.
We have obtained pKa values of p-nitrophenol–TiO2 by measuring the adsorption equilibrium constants of p-nitrophenol (PNP) on the TiO2 surface at different pH values. These values have been obtained from Langmuir isotherms and from a plot of 1/rate vs. 1/[PNP]o obtained during TiO2 catalyzed solar light photo-degradation of PNP. Two limit equilibrium constants are readily obtained depending on the solution pH: at pH 5 at which the TiO2 surface is mainly positively charged and at pH 8 when it is negatively charged. With these and other adsorption equilibrium constants and the PNP pKa value in solution, thermodynamic cycles are established in order to obtain the PNP pKa when it is adsorbed on positively charged, neutral and negatively charged TiO2 surfaces. From these pKa values useful information on the PNP–TiO2 interaction is readily obtained. For instance, the PNP nitro group interacts with the TiO2 surface via a hydrogen bond, arising from the complex of water molecules with the Ti4+ ions on its surface. The weaker the hydrogen bond donor, the stronger the oxygen nitro group basicity. Therefore, pKa changes on the phenolic hydroxyl group result from these interactions. Linear free energy correlations, maximum PNP adsorption capacity values (QL) and FTIR ATR, spectrum support this proposal. A kobs vs. pH degradation profile of p-nitrophenol is also provided.  相似文献   

13.
In the framework of our studies on acid=nbase equilibria in systems comprisingsubstituted pyridines and nonaqueous solvents, acid dissociation constants havebeen determined potentiometrically for a variety of cationic acids conjugatedwith pyridine and its derivatives in the polar protophobic aprotic solvent nitromethane. The potentiometric method enabled a check as to whether and to whatextent cationic homoconjugation equilibria of the BH+/B type, as well as cationicheteroconjugation equilibria in BH+/B1 systems without proton transfer, are setup in nitromethane. The equilibrium constants were compared with thosedetermined in water and two other polar protophobic aprotic solvents, propylenecarbonate and acetonitrile. The pK a values of acids conjugate to the N-bases innitromethane fall in the pK a range of 5.84 to 17.67, i.e., 6 to 7 pK a units, onaverage, higher than in water, 1 to 2 units higher than in propylene carbonate,and less than 1 unit lower than in acetonitrile. This means that the basicity ofthe pyridine derivatives increases on going from propylene carbonate throughnitromethane to acetonitrile. Further, it was found that the sequence of the pK achanges of the protonated amines was consistent in all three media, thus providingthe basis for establishing linear correlations among these values. In the majorityof the BH+/B systems in nitromethane, cationic homoconjugation equilibria havebeen established. The cationic homoconjugation constants, log K BHB+, arerelatively low, falling in the range 1.60–2.89. A comparison of the homoconjugationconstants in nitromethane with those in propylene carbonate and acetonitrile showsthat nitromethane is a more favorable solvent for the cationic homoconjugationequilibria than the other two solvents. Moreover, results of the potentiometricmeasurements revealed that cationic heteroconjugation equilibria were not presentin the majority of the BH+/B1 systems in nitromethane. The heteroconjugationconstant could be determined in one system only, with logdiK BHB1 + = 2.56.  相似文献   

14.
The ionization potentials of alkyl and hydrogen halides are found to be excellent linear fractions of the polar and inductive substituent constants, indicating that the effect of alkyl substituents on the electron density of the halogen atoms is inductive. The slopes of the four regression lines vary widely in the order RF ? RCl > RBr > RI, which shows that the susceptibility of the halogen atoms to inductive effects varies in the same order. Values of σI and σ* for alkyl groups not previously available are estimated.  相似文献   

15.
N.P. Slabbert 《Tetrahedron》1977,33(7):821-824
Ionisation constants of catechin (1), robinetinidol (2), leuco-fisetinidin (3), fustin (4), dihydrorobinetin (5), dihydroquercetin (6) and a series of polyphenols are reported. The pKa values of these flavonoids are shown to fit the linear relationships between pKa, and substituent σ constant for a series of related phenols.  相似文献   

16.
The relationship between hydrophobicity and RM has been studied for four series of acids, namely arylacetic, cinnamic, β-arylisobutyric, and α-methylcinnamic. The possibility of using pK values or polar constants to correct for dissociation of the acids investigated is discussed. The influence of substituent alkoxy groups on the relationship was investigated, and it was found that polar constants were important, even in systems in which dissociation was inhibited by addition of formic acid. In these circumstances, the polar constants probably compensate for modification of the hydrogen ion donor-acceptor properties of alkoxy derivatives. This phenomenon evidently plays a decisive role in those chromatographic systems markedly different from the reference partition system n-octanol-water.  相似文献   

17.
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   

18.
Capillary electrophoresis (CE) has been applied for determination of the thermodynamic acidity constants (pKa) of the sulfamidoalkyl and sulfonamidoalkyl groups, the actual and limiting ionic mobilities and hydrodynamic radii of important compounds, eight carborane-based inhibitors of carbonic anhydrases, which are potential new anticancer drugs. Two types of carboranes were investigated, (i) icosahedral cobalt bis(dicarbollide)(1-) ion with sulfamidoalkyl moieties, and (ii) 7,8-nido-dicarbaundecaborate with sulfonamidoalkyl side chains. First, the mixed acidity constants, pKamix, of the sulfamidoalkyl and sulfonamidoalkyl groups of the above carboranes and their actual ionic mobilities were determined by nonlinear regression analysis of the pH dependences of their effective electrophoretic mobility measured by capillary electrophoresis in the pH range 8.00−12.25, at constant ionic strength (25 mM), and constant temperature (25°C). Second, the pKamix were recalculated to the thermodynamic pKas using the Debye–Hückel theory. The sulfamidoalkyl and sulfonamidoalkyl groups were found to be very weakly acidic with the pKas in the range 10.78−11.45 depending on the type of carborane cluster and on the position and length of the alkyl chain on the carborane scaffold. These pKas were in a good agreement with the pKas (10.67−11.27) obtained by new program AnglerFish (freeware at https://echmet.natur.cuni.cz ), which provides thermodynamic pKas and limiting ionic mobilities directly from the raw CE data. The absolute values of the limiting ionic mobilities of univalent and divalent carborane anions were in the range 18.3−27.8 TU (Tiselius unit, 1 × 10−9 m2/Vs), and 36.4−45.9 TU, respectively. The Stokes hydrodynamic radii of univalent and divalent carborane anions varied in the range 0.34−0.52 and 0.42−0.52 nm, respectively.  相似文献   

19.
The values of pKams (Kams represents ionization constant of conjugate acid of amine base in mixed water–acetonitrile solvent) for all amines, except for charged amine bases, show a mild decrease (ca. 0.1–0.4 pK units) with the increase in CH3CN content from 2 to ∼60% v/v. However, the pKams values at 70% v/v CH3CN become nearly equal or slightly larger (by ≤0.7 pK units) than the corresponding pKams at 2% v/v CH3CN for all neutral and charged amines. The values of pKams for phenol increase from 10.17 to 13.38 with the increase in the content of CH3CN from 2 to 70% v/v in mixed aqueous solvent. Taft reaction constants, ρ*, obtained from the plots of pKams against ∑σ* for primary and secondary amines decrease by ca. 0.8 ρ* units with the increase in the CH3CN content from 2 to 70% v/v. The values of pKams show an empirical linear relationship with the corresponding values of pKaw (where pKaw represents the pKa obtained in aqueous solvent containing 2% v/v CH3CN), which allows the estimation of a pKa in mixed H2O CH3CN solvents from that in water. © 2000 John Wiley & Sons, Inc. Int J Chem Kinet 32: 146–152, 2000  相似文献   

20.
Grob's basicity data sets of 4-substituted quinuclidines are successfully correlated with the LArSR Eq. 1, the correlation parameters of which indicate that the inductive substituent constants involved in δop just correspond to 0.74°I.  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