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.
Reaction of isocyanates XNCO (X = Ar, i-Pr, t-Bu) with CH(2)(Y)CO(2)CH(2)CCl(3) (Y = CO(2)Me, CO(2)CH(2)CCl(3), CN) gave 15 amides XNHCOCH(Y)CO(2)CH(2)CCl(3) (6) or enols of amides XNHC(OH)=C(Y)CO(2)CH(2)CCl(3) (5) systems. The amide/enol ratios in solution depend strongly on the substituent Y and the solvent and mildly on the substituent X. The percentage of enol for group Y increases according to Y = CN > CO(2)CH(2)CCl(3) > CO(2)Me and decreases with the solvent according to CCl(4) > C(6)D(6) > CDCl(3) > THF-d(8) > CD(3)CN > DMSO-d(6). With the most acidic systems (Y = CN) amide/enol exchange is observed in moderately polar solvents and ionization to the conjugate base is observed in DMSO-d(6). The solid-state structure of the compound with Y = CN, X = i-Pr was found to be that of the enol. The reasons for the stability of the enols were discussed in terms of polar and resonance effects. Intramolecular hydrogen bonds result in a very low delta(OH) and contribute to the stability of the enols and are responsible for the higher percentage of the E-isomers when Y = CO(2)Me and the Z-isomers when Y = CN. The differences in delta(OH), delta(NH), K(enol), and E/Z enol ratios from the analogues with CF(3) instead of CCl(3) are discussed.  相似文献   

2.
The structures of anilido cyano(fluoroalkoxycarbonyl)methanes ArNHCOCH(CN)CO(2)R, where R = CH(2)CF(3) or CH(CF(3))(2), Ar = p-XC(6)H(4), and X = MeO, Me, H, or Br, were investigated. In the solid state, all exist as the enols ArNHC(OH)=C(CN)CO(2)R 7 (R = CH(2)CF(3)) and 9 (R = CH(CF(3))(2)) with cis arrangement of the hydrogen-bonded ROC=O.HO moiety and a long C1=C2 bond. The product composition in solution is solvent dependent. In CDCl(3) solution, only a single enol is observed, whereas in THF-d(8) and CD(3)CN, two enols (E and Z) are the major products, and the amide is the minor product or not observed at all (K(Enol) 1.04-9 (CD(3)CN, 298 K) and 3 to >/=100 (THF, 300 K)). The percentage of the amide and the Z-enol increase upon an increase in temperature. In all solvents, the percent enol is higher for 9 than for 7. In CD(3)CN, more enol is observed when the aryl group is more electron-donating. The spectra in DMSO-d(6) and DMF-d(7) indicate the presence of mostly a single species, whose spectra do not change on addition of a base and is ascribed to the anion of the ionized carbon acid. Comparison with systems where the CN is replaced by a CO(2)R group (R = CH(2)CF(3), CH(CF(3))(2)) shows a higher percentage of enol for the CN-substituted system. Intramolecular (to CO(2)R) and intermolecular hydrogen bonds determine, to a significant extent, the stability of the enols, their Z/E ratios (e.g., Z/E (THF, 240 K) = 3.2-4.0 (7) and 0.9-1.3 (9)), and their delta(OH) in the (1)H spectra. The interconversion of Z- and E-enol by rotation around the C=C bond was studied by DNMR, and DeltaG() values of >/=15.3 and 14.1 +/- 0.4 kcal/mol for Z-7 and Z-9 were determined. Features of the NMR spectra of the enols and their anions are discussed.  相似文献   

3.
Reactions of isocyanates XNCO (e.g., X = p-An, Ph, i-Pr) with (MeO)2P(=O)CH2CO2R [R = Me, CF3CH2, (CF3)2CH] gave 15 formal "amides" (MeO)2P(=O)CH(CO2R)CONHX (6/7), and with (CF3CH2O)2P(=O)CH2CO2R [R = Me, CF3CH2] they gave eight analogous amide/enols 17/18. X-ray crystallography of two 6/7, R = (CF3)2CH systems revealed Z-enols of amides structures (MeO)2P(=O)C(CO2CH(CF3)2)=C(OH)NHX 7 where the OH is cis and hydrogen bonded to the O=P(OMe)2 group. The solid phosphonates with R = Me, CF3CH2 have the amide 6 structure. The structures in solution were investigated by 1H, 13C, 19F, and 31P NMR spectra. They depend strongly on the substituent R and the solvent and slightly on the N-substituent X. All systems displayed signals for the amide and the E- and Z-isomers. The low-field two delta(OH) and two delta(NH) values served as a probe for the stereochemistry of the enols. The lower field delta(OH) is not always that for the more abundant enol. The % enol, presented as K(enol), was determined by 1H, 19F, and 31P NMR spectra, increases according to the order for R, Me < CF3CH2 < (CF3)2CH, and decreases according to the order of solvents, CCl4 > CDCl3 approximately THF-d8 > CD3CN >DMSO-d6. In DMSO-d6, the product is mostly only the amide, but a few enols with fluorinated ester groups were observed. The Z-isomers are more stable for all the enols 7 with E/Z ratios of 0.31-0.75, 0.15-0.33, and 0.047-0.16 when R = Me, CF3CH2, and (CF3)2CH, respectively, and for compounds 18, R = Me, whereas the E-isomers are more stable than the Z-isomers. Comparison with systems where the O=P(OMe)2 is replaced by a CO2R shows mostly higher K(enol) values for the O=P(OMe)2-substituted systems. A linear correlation exists between delta(OH)[Z-enols] activated by two ester groups and delta(OH)[E-enols] activated by phosphonate and ester groups. Compounds (MeO)2P(=O)CH(CN)CONHX show 相似文献   

4.
Reaction of beta-methylglutaconic anhydride with NaOMe followed by reaction with methyl or phenyl chloroformate gave the corresponding O-methoxy (and O-phenoxy) carbonylation derivatives. Reaction of the anhydride with MgCl2/pyridine, followed by methyl chloroformate gave C-methoxycarbonylation at C3 of the anhydride. The product (4) was previously suggested by calculation to be the enol of the anhydride 5 and this is confirmed by X-ray crystallography (bond lengths: C-OH, 1.297 A; C1C2 1.388 A; HO...O=C(OMe) distance 2.479 A) making it the first solid enol of an anhydride. In CDCl3, CD3CN, or C6D6 solution it displays the OH as a broad signal at ca. 15 ppm, suggesting a hydrogen bond with the CO2Me group. NICS calculations indicate that 4 is nonaromatic. With D2O in CDCl3 both the OH and the C5H protons exchange rapidly the H for D. An isomeric anhydride 5a of 5 is formed in equilibrium with 4 in polar solvents. In solution, anhydride(s)/enol equilibria are rapidly established with Kenol of 6.40 (C6D6, 298 K), 0.52 (CD3CN, 298 K), 9.8 (CDCl3, 298 K), 22.8 (CDCl3, 240 K), and decreasing Kenol in CDCl3:CD3CN mixtures with the increase in percent of CD3CN. The percentage of the rearranged anhydride in CDCl3:(CD3)2CO increases with the increased percent of (CD3)2CO. In DMSO-d6 and DMF-d7 the observed species are mainly the conjugated base 4- and 5a. Deuterium effects on the delta(13C) values were determined. An analogous C2-OH enol of anhydride 15 substituted by 3-CO2Me and 4-OCO2Me groups was prepared. Its structure was confirmed by X-ray crystallography (CO bond length 1.298 A, O...O distance 2.513 A); delta(OH) = 12.04-13.22 ppm in CDCl3, THF-d8, and CD3CN, and Kenol = > or = 100, 7.7, and 3.4 respectively. In DMSO-d6 enol 15 ionizes to its conjugate base. Substantial upfield shifts of the apparent delta("OH") proton on diluting the enol solutions are ascribed to the interaction of the H+ formed with the traces of water in the solvent to give H3O+, which gives the alleged "OH proton" signal.  相似文献   

5.
Condensation of phenyl isocyanate substituted by 4-MeO, 4-Me, 4-H, 4-Br, and 2,4-(MeO)(2) with esters CH(2)(CO(2)R)CO(2)R', R = CH(2)CF(3), R' = CH(3), CH(2)CF(3), CH(CF(3))(2), or R = CH(3), R' = CH(CF(3))(2) gave 17 "amides" ArNHCOCH(CO(2)R)CO(2)R' containing three, six, or nine fluorines in the ester groups. X-ray crystallography of six of them revealed that compounds with > or =6 fluorine atoms exist in the solid state as the enols of amides ArNHC(OH)=C(CO(2)R)CO(2)R' whereas the ester with R = R' = CH(3) was shown previously to have the amide structure. In the solid enols, the OH is cis and hydrogen bonded to the better electron-donating (i.e., with fewer fluorine atoms) ester group. X-ray diffraction could not be obtained for compounds with only three fluorine atoms, i.e., R = CH(2)CF(3), R' = CH(3) but the (13)C CP-MAS spectra indicate that they have the amide structure in the solid state, whereas esters with six and nine fluorine atoms display spectra assigned to the enols. The solid enols show unsymmetrical hydrogen bonds and the expected features of push-pull alkenes, e.g., long C(alpha)=C(beta) bonds. The structure in solution depends on the number of fluorine atoms and the solvent, but only slightly on the substituents. The symmetrical systems (R = R' = CH(2)CF(3)) show signals for the amide and the enol, but all systems with R not equal R' displayed signals for the amide and for two enols, presumably the E- and Z-isomers. The [Enol I]/[Enol II] ratio is 1.6-2.9 when R = CH(2)CF(3), R' = CH(3), CH(CF(3))(2) and 4.5-5.3 when R = CH(3), R' = CH(CF(3))(2). The most abundant enol display a lower field delta(OH) and a higher field delta(NH) and assigned the E-structure with a stronger O-H.O=C(OR) hydrogen bond than in the Z-isomer. delta(OH) and delta(NH) values are nearly the same for all systems with the same cis CO(2)R group. The [Enols]/[Amide] ratio in various solvents follows the order CCl(4) > CDCl(3) > CD(3)CN > DMSO-d(6). The enols always predominate in CCl(4) and the amide is the exclusive isomer in DMSO-d(6) and the major one in CD(3)CN. In CDCl(3) the major tautomer depends on the number of fluorines. For example, in CDCl(3,) for Ar = Ph, the % enol (K(Enol)) is 35% (0.54) for R = CH(2)CF(3,) R' = CH(3), 87% (6.7) for R = R' = CH(2)CF(3), 79% (3.8) for R = CH(3), R' = CH(CF(3))(2) and 100% (> or =50) for R = CH(2)CF(3), R' = CH(CF(3))(2). (17)O and (15)N NMR spectra measured for nine of the enols are consistent with the suggested assignments. The data indicate the importance of electron withdrawal at C(beta), of intramolecular hydrogen bonding, and of low polarity solvents in stabilizing the enols. The enols of amides should no longer be regarded as esoteric species.  相似文献   

6.
Twenty open-chain mono-, di-, and trialkyl and aryl-N-substituted cyanomalonamides R2R1NCOCH(CN)CONHR3 were prepared. In solution, signals for both amide and a single enol are mostly observed, despite the potential for E and Z isomeric enols. The equilibrium (KEnol) values between the amides and the enols were determined in different solvents by NMR spectra. They decrease on increasing the polarity of the solvent in the order CDCl3 approximately C6D6>THF-d8>(CD3)2CO>CD3CN>DMF-d7>DMSO-d6. For the R1R2NCOCH(CN)CONHR3 system when R1=R2=H, Me or R1=H, R2=Me, KEnol for R3 follows the order: C6F5>Ph>or=An>or= i-Pr>or= t-Bu, and for R1, R2:H, H>Me, H>Me, Me in all solvents. A unique feature is the appreciable % enol in DMSO-d6 when R1=R2=H, in contrast with enol systems with other electron-withdrawing groups (EWGs). Calculations (B3LYP/6-31G**) corroborate the higher KEnol values for less alkyl-substituted systems, showing that in the most stable conformer when R1=H, R2=R3=Me the N-hydrogens are closer to the CN group. The order of promoting substituents for enol of amide formation is CONH2>CO2CH2CF3>CO2Me>CONHMe. The solid-state structures of the isolated species, determined by X-ray crystallography, were either amides or enols, and a higher KEnol(CDCl3) value does not ensure a solid enol structure. In no system were both solid species isolated. The X-ray structures of the enols were temperature-dependent. In most cases, the difference between the O-H and O...H bond lengths at low temperature were appreciable, but they become closer at the higher temperature. Similar tendency for either the C=C/C-C or the C-O/C=O bonds was observed. This is ascribed to a hydrogen shift between two regioisomeric enols in an asymmetric double-well potential, which becomes faster at a higher temperature. Calculations show that the enol structures are nonsymmetrical, resembling the lower temperature structures, even when they are chemically symmetrical, but the energy differences between the two regioisomers are <1 kcal. The hydrogen bonds in the enol moiety are strong, with O...O distances <2.45 A, and are resonance-assisted hydrogen bonds. IR spectra in solution and the solid state qualitatively corroborate the NMR and X-ray structure determination.  相似文献   

7.
A search for the enol structures of several amides YY'CHCONHPh with Y,Y' = electron-withdrawing groups (EWGs) was conducted. When Y = CN, Y' = CO(2)Me the solid structure is that of the enol (8b) MeO(2)CC(CN)=C(OH)NHPh, whereas in solution the NMR spectrum indicate the presence of both the amide MeO(2)CCH(CN)CONHPh (8a) and 8b. When Y = NO(2), Y' = CO(2)Et the main compound in CDCl(3) is the amide, but <10% of enol(s), presumably EtO(2)CC(NO(2))=C(OH)NHPh (9b), are also present. When Y = COEt, Y' = CO(2)Me or Y = COMe, Y' = CO(2)Et (10 and 11) enolization in solution and of 11 also in the solid state occurs at the carbonyl rather than at the ester site. With Y = Y' = CN a rapid exchange between the amide (NC)(2)CHCONHPh (12a) and a tautomer, presumably the enol, take place in several solvents on the NMR time scale. With YY' = barbituric acid moiety the species in DMSO-d(6) is an enol of an amide although which CONH group enolizes is unknown. B3LYP/6-31G calculations showed that the enol (NC)(2)C=C(OH)NH(2) (13b) is more stable by DeltaG of 0.4 kcal/mol than (NC)(2)CHCONH(2) (13a) due to a combination of stabilization of 13b and destabilization of 13a and both are much more stable than the hydroxyimine and ketene imine tautomers. The effect of Y,Y' and the solvent on the relative stabilization of enols of amides is discussed.  相似文献   

8.
[Structure: see text] The first example of a crystalline amide and its tautomeric enol was obtained for the amide MeNHCSCH(CN)CONHMe (8) and its enol MeNHCSC(CN)=C(OH)NHMe (9). Their X-ray structures were determined, and their structural features resemble those of other related amides and enols. No other example of a similar pair was obtained. In solution, both 8 and 9 and a small percentage of the isomeric enol of thioamide MeNHCOC(CN)=C(SH)NHMe (10) were obtained in solvent-dependent compositions, which are rapidly established.  相似文献   

9.
The gas-phase acidities DeltaG degrees (acid) of some 20 amides/enols of amides RNHCOCHYY'/RNHC(OH)=CYY' [R = Ph, i-Pr; Y, Y' = CO(2)R', CO(2)R' ', or CN, CO(2)R', R', R' ' = Me, CH(2)CF(3), CH(CF(3))(2)], the N-Ph and N-Pr-i amides of Meldrum's acid, 1,3-cyclopentanedione, dimedone, and 1,3-indanedione, and some N-p-BrC(6)H(4) derivatives and of nine CH(2)YY' (Y, Y' = CN, CO(2)R', CO(2)R' '), including the cyclic carbon acids listed above, were determined by ICR. The acidities were calculated at the B3LYP/6-31+G//B3LYP/6-31+G level for both the enol and the amide species or for the carbon acid and the enol on the CO in the CH(2)YY' series. For 12 of the compounds, calculations were also conducted with the larger base sets 6-311+G and G-311+G. The DeltaG degrees (acid) values changed from 341.3 kcal/mol for CH(2)(CO(2)Me)(2) to 301.0 kcal/mol for PhNHC(OH)=C(CN)CH(CF(3))(2). The acidities increased for combinations of Y and Y' based on the order CO(2)Me < CO(2)CH(2)CF(3) < CN, CO(2)CH(CF(3))(2) for a single group and reflect the increased electron-withdrawal ability of Y,Y' coupled with the ability to achieve planarity of the crowded anion. The acidities of corresponding YY'-substituted systems follow the order N-Ph enols > N-Pr-i enols > CH(2)YY'. Better linear relationships between DeltaG degrees (acid) values calculated for the enols and the observed values than those for the values calculated for the amides suggest that the ionization site is the enolic O-H of most of the noncyclic trisubstituted methanes. The experimental DeltaG degrees (acid) value for Meldrum's acid matches the recently reported calculated value. The calculated structures and natural charges of all species are given, and the changes occurring in them on ionization are discussed. Correlations between the DeltaG degrees (acid) values and the pK(enol) values, which are linear for the trisubstituted methanes, excluding YY' = (CN)(2) and nonlinear for the CH(2)YY' systems, are discussed.  相似文献   

10.
The structures of the poly(methoxycarbonyl)cyclopentadienes C(5)H(6)(-)(n)()(CO(2)Me)(n)(), n = 5 (Cp-5), n = 4 (Cp-4), n = 3 (1, 2,4-; Cp-3) and n = 2 (1,2-; 1,2-Cp-2-I) were investigated. The X-ray diffractions of Cp-5 (known), Cp-4, and Cp-3 showed an enol of ester structure in the solid state. The enolic hydrogen forms a symmetrical hydrogen bond to a neighboring ester carbonyl, so that the vicinal "enolic" CO(2)Me groups in the 1, 2-C(=CO(2)Me)-C(CO(2)Me)(4) moiety are identical. The relevant X-ray parameters for the three enols are similar. The CP-MAS spectra of Cp-3-Cp-5 generally resemble their (13)C NMR spectra in CDCl(3) except for some differences of mostly <1 ppm. The (1)H, (13)C, and (17)O NMR spectra of Cp-3-Cp-5 in CDCl(3) are consistent with those of the hydrogen bonded enols. Most characteristic are the (1)H and (17)O signals of the OH groups at 19.7-20.1 and 221-225 ppm, respectively. Proton addition to sodium 1, 2-bis(methoxycarbonyl)cyclopentadienide gave a mixture of four 1, 2-bis(methoxycarbonyl)cyclopentadienes. The isomer (1,2-Cp-2-I) formed in 10-20% displays delta(O(1)H) at 19.3 ppm and is the enol analogue of Cp-5 whereas its main isomer (30-55%) (1,2-Cp-2-IV) has the ester structure. In CD(3)CN and DMSO-d(6) only one signal was observed at room temperature for each type of H, C, or O of Cp-5, suggesting a complete ionization to the symmetrical anion of Cp-5. In contrast, Cp-4 and Cp-3 in CD(3)CN at room temperature display OH signals in both (1)H and (17)O NMR spectra, and Cp-5 shows a broad OH signal in the (1)H spectrum at 240 K. The enol of ester structure is the main species, although exchange with the corresponding anion is possible. On standing in DMF-d(7) at room temperature, new signals are observed for Cp-3 and Cp-4. On raising the temperature in Cl(2)CDCDCl(2), Cp-3-Cp-5 show line broadening and appearance of new signals. These were ascribed to rearrangment and decomposition processes.  相似文献   

11.
The enols R 1 R 2 P(E)(CN)C = CR 3 OH (E = O or S) gave in solutions either neutral metal complexes ML x or M(OH) y L x . The anionic ambidentate ligands are coordinated through E and O atoms in solutions, and O, E, and N atoms in in crystals.  相似文献   

12.
2-Hydroxyoxol-2-ene (C(5)-1), the enol tautomer of gamma-butyrolactone, was generated in the gas phase as the first representative of the hitherto elusive class of lactone enols and shown by neutralization-reionization mass spectrometry to be remarkably stable as an isolated species. Ab initio calculations by QCISD(T)/6-311+G(3df,2p) provided the enthalpies of formation, proton affinities, and gas-phase basicities for gaseous lactone enols with four- (C(4)-1), five- (C(5)-1), and six-membered rings (C(6)-1). The acid-base properties of C(4)-C(6) lactones and enols and reference carboxylic acid enols CH(2)=C(OH)(2) (3) and CH(2)=C(OH)OCH(3) (4) were also calculated in aqueous solution. The C(4)-C(6) lactone enols show gas-phase proton affinities in the range of 933-944 kJ mol(-)(1) and acidities in the range of 1401-1458 kJ mol(-)(1). In aqueous solution, the lactone enols are 15-20 orders of magnitude more acidic than the corresponding lactones, with enol pK(a) values increasing from 5.6 (C(4)-1) to 14.5 (C(6)-1). Lactone enols are moderately weak bases in water with pK(BH) in the range of 3.9-8.1, whereas the lactones are extremely weak bases of pK(BH) in the range of -10.5 to -17.4. The acid-base properties of lactone enols point to their high reactivity in protic solvents and explain why no lactone enols have been detected thus far in solution studies.  相似文献   

13.
(E)-2-(m-methoxymesityl)-1,2-dimesitylethenol (3a) isomerizes in the absence of a catalyst in solution to a 1.0:0.9 E/Z (3a/3b) equilibrium mixture. In CDCl3 the isomerization is first order in 3a within a run, but the plot of the rate constant k(obs) vs the changing [3a]0 in different runs is a half-parabola, indicating self-catalysis by more than one enol molecule. At 0.09 M enol, the isotope effect k(3a)/k(3a)-OD = 2.1. In the presence of 0.025-0.25 M pyridine-d5, the k(obs) vs [pyridine-d5] plot displays a bell-shaped profile. The change in the shape of the OH signals of the 3a/3b mixture at 295-430 K in C6D5NO2 was followed by DNMR. The four signals of the diastereomeric 3a/3b mixture observed at 330 K coalesce at 350 K with barriers of 18.3 and 18.4 kcal x mol(-1) due to the diastereomerization of the vinyl propellers. The resulting two signals observed at >360 K further coalesce at 425 K with a barrier of 22.9 kcal x mol(-1) due either to oxygen-to-oxygen proton exchange or to E/Z isomerization. The estimated upper limit for the rate of proton exchange of k(ex) < or = (2-4) x 10(3) M(-1) x s(-1) at 425 K between the enol molecules is sufficiently slow to be a rate-controlling step in the isomerization. A process in which several enol molecules catalyze the isomerization is suggested, and several mechanistic routes are analyzed.  相似文献   

14.
The literature describing the observation of enols of carboxylic anhydrides and mixed carboxylic‐sulfuric anhydrides was examined. In the phenylbutyric anhydride system, the alleged enol was shown to be ethylphenylketene, and the monoenol EtC(Ph)=C(OH)OC(=O)CH(Ph)Et ( 5 ) and the dienol ( 6 ) should not be observed according to calculations. Calculations also show that the claimed enols H2C=C(OH)OSO2Y, Y=SO, Ac ( 15 ) and the enol of 2H‐pyran‐2,6(3H)‐dione ( 7 ) are too unstable to be observed. The bulky enols of β,β‐ditipylacetic formic ( 35a ) or trifluoroacetic ( 35b ) anhydride were calculated to be unstable with pKEnol=7.7 (6.2). The suggestion that compounds with the 3‐acyl or 3‐aroyl‐2H‐pyran‐2,6(3H)‐dione skeleton are enolic was examined. In the solid state, all the known structures show that enolization takes place on C(5)=O. However, B3LYP/6‐31G** calculations show that, for 3‐acetyl‐4‐methyl‐2H‐pyran‐2,6(3H)‐dione ( 10 , R1=Me, R2=H), which is completely enolic, the enol on the acetyl group (cf. 12 ) is only 0.9 kcal/mol more stable than the enol on the anhydride (cf. 11 ). Calculations also revealed that 3‐(trifluoroacetyl)‐2H‐pyran‐2,6(3H)‐dione ( 28 ) should exist in nearly equal amounts of the enol of anhydride (cf. 30 ) and the enol of the acyl group (cf. 29 ), whereas the enol of anhydride (cf. 32 ) is the only stable species for 3‐(methoxycarbonyl)‐2H‐pyran‐2,6(3H)‐dione ( 31 ). Furan‐2,5‐diol ( 27 ) and 5‐hydroxyfuran‐2‐one ( 26 ) are calculated not to give observable isomers of succinic anhydride ( 25 ) (pKEnol=30 and 18, resp.) in spite of the expected aromatic stabilization of 27 . Surprisingly, the calculations reveal that the enol (NC)2C=C(OH)OCHO ( 38 ) is less stable than its tautomeric anhydride ( 37 ) (pKEnol=1.6). Comparison of calculated pKEnol values for (NC)2CHC(=O)X ( 41 ) and MeC(=O)X indicates that the assumption that substitution by two β‐CN groups affects similarly all the systems regardless of X is incorrect. A pKEnol((NC)2CHC(=O)X) vs. pKEnol(MeC(=O)X) plot is linear for most substituents with severe and mild negative deviations, respectively, for X=NH2 and MeO. Appropriate isodesmic reactions have shown that the β,β‐(CN)2 substitution increases the stabilization of the enol of amide (X=NH2) by 14.6 kcal/mol over that for the anhydride (X=OCHO), whereas the amide form is 7.1 kcal/mol less destabilized than for the anhydride. The pKEnol value for (MeOCO)2CHCOOCHO ( 43 ) is 3.6, i.e., stabilization by these β‐electron‐withdrawing groups is insufficient to make the enols observable.  相似文献   

15.
The energetics of the stationary points of the gas-phase reactions CH(3)X+F(-)-->CH(3)F+X(-) (X=F, Cl, CN, OH, SH, NH(2) and PH(2)) have been definitively computed using focal point analyses. These analyses entailed extrapolation to the one-particle limit for the Hartree-Fock and MP2 energies using basis sets of up to aug-cc-pV5Z quality, inclusion of higher-order electron correlation [CCSD and CCSD(T)] with basis sets of aug-cc-pVTZ quality, and addition of auxiliary terms for core correlation and scalar relativistic effects. The final net activation barriers for the forward reactions are: E (b/F,F)=-0.8, E (b/F, Cl)=-12.2, E (b/F,OH)=+13.6, E b/F,OH=+16.1, E b/F,SH=+2.8, Eb/F, NH=+32.8, and E b/F,PH =+19.7 kcal x mol(-1). For the reverse reactions E b/F,F= -0.8, Eb/Cl,F =+18.3, E b/CN,F=+12.2, E b/OH,F =-1.8, E b/SH,F =+13.2, E b/NH(2),=-1.5, and E b/PH(2) =+9.6 kcal x mol(-1). The change in energetics between the CCSD(T)/aug-cc-pVTZ reference prediction and the final extrapolated focal point value is generally 0.5-1.0 kcal mol(-1). The inclusion of a tight d function in the basis sets for second-row atoms, that is, utilizing the aug-cc-pV(X+d)Z series, appears to change the relative energies by only 0.2 kcal x mol(-1). Additionally, several decomposition schemes have been utilized to partition the ion-molecule complexation energies, namely the Morokuma-Kitaura (MK), reduced variational space (RVS), and symmetry adapted perturbation theory (SAPT) techniques. The reactant complexes fall into two groups, mostly electrostatic complexes (FCH(3).F(-) and ClCH(3).F(-)), and those with substantial covalent character (NCCH(3).F(-), CH(3)OH.F(-), CH(3)SH.F(-), CH(3)NH(2).F(-) and CH(3)PH(2).F(-)). All of the product complexes are of the form FCH(3).X(-) and are primarily electrostatic.  相似文献   

16.
Carbon-acid ionization constants, Q(K)(a)(concentration quotient at ionic strength = 0.10 M), were determined by spectrophotometric titration in aqueous solution for benzo[b]-2,3-dihydrofuran-2-one (3, pQ(K)(a) = 11.87), benzo[b]-2,3-dihydrothiophene-2-one (2, pQ(K)(a) = 8.85), and benzo[b]-2,3-dihydrofuran-2-thione (1, pQ(K)(a) = 2.81). Rates of approach to keto-enol equilibrium were also measured for the latter two substrates in perchloric acid, sodium hydroxide, and buffer solutions, and the rate profiles constructed from these data gave the ionization constants of the enols ionizing as oxygen or sulfur acids pQ(E)(a) = 5.23 for 2 and pQ(E)(a) = 2.69 for 1. Combination of these acidity constants with the carbon-acid ionization constants according to the relationship Q(K)(a)/Q(E)(a) = K(E) then gave the keto-enol equilibrium constants pK(E) = 3.62 for 2 and pK(E) = 0.12 for 1. The fourth, all-sulfur, member of this series, benzo[b]-2,3-dihydrothiophene-2-thione (4), proved to exist solely as the enol in aqueous solution, and only the enol ionization constant pQ(E)(a) = 3.44 could be determined for this substance; the limits pK(E) < 1.3 and pQ(K)(a) < 2.1, however, could be set. The unusually high acidities and enol contents of these substances are discussed, as are also the relative values of the ketonization and enolization rate constants measured; in the latter cases, Marcus rate theory is used to determine intrinsic kinetic reactivities, free of thermodynamic effects.  相似文献   

17.
The reaction of omega-chloroalkyl isocyanates Cl(CH2)nNCO (n = 2 (2), 3 (4)) and isothiocyanate Cl(CH2)2NCS (3) with active methylene compounds CH2YY' 1 in the presence of Et3N or Na give 2-YY'-methylene-1,3-oxazolidines, (E,Z)-1,3-thiazolidines, and 1,3-oxazines from 2, 3, and 4, respectively. 2-(Chloromethyl)phenyl isocyanate 8 gives with 1 the corresponding benzo-oxazines. Ethyl 2-isothiocyanatobenzoate 10 gives the corresponding benzothiazolinone, whereas the analogous isocyanate 12 gives noncyclic enols. Ethoxycarbonyl isothiocyanate 14 gives an open-chain thioenol or an enol-thioamide. The cyanoamides CH2(CN)CONHR, R = H, Me, CHPh2, give with Et3N and 2 the bicyclic imidazopyrimidinediones 16, derived from two molecules of 2, but with their preformed Na salt they give the 1,3-oxazolidines. Reaction of cyanoacetamide with 3 in the presence of Na gave a tricyclic triaza(thia)indacene, derived from two molecules of 3. A reaction mechanism involving an initial attack of the anion 1- on the N=C=X (X = O, S) moiety gives an anion 18, which cyclizes intramolecularly and after tautomerization gives the mono-ring heterocycle. With the cyanoamides, the N- site of the ambident ion 18 attacks another molecule of 2 giving the anion 20, which by intramolecular attack on the CN, followed by expulsion of the Cl- gives the bicyclic 16 after tautomerization.  相似文献   

18.
The diagnostic values of the following three spectral criteria for the configuration of N-acyl-alpha,beta-dehydro-alpha-amino acid esters were examined: (i) the proton at the beta-position at the double bond of a Z-isomer is shielded if compared with the respective E-isomer (delta(beta)Z < delta(beta)E); (ii) the proton at the nitrogen atom is shielded in a Z-isomer in comparison with the corresponding E-isomer (delta(NH)Z < delta(NH)E); and (iii) changing of the solvent from CDCl3 to deuterated trifluoroacetic acid (TFA) causes shielding of the H(beta) vinylic proton of an E-isomer or deshielding of the respective proton of the Z-isomer (delta(CDCl3)E > delta(TFA)E or delta(CDCl3)Z < delta(TFA)Z). The investigations were based on a set of 22 (Z)- and (E)-N-acyl-alpha,beta-dehydro-alpha-amino acid esters of diverse structures, with aliphatic, aromatic and heteroaromatic substituents at the vinylic beta-carbon; most of the examined compounds were hitherto unknown. The application of the substituent effect additivity rule given by Pascual et al. for olefinic protons leads to evidently erroneous configuration assignments of N-acyl-alpha,beta-dehydro-alpha-amino acid esters. The considered criteria were fulfilled for all the examined cases with one exception [the second criterion for the alpha-pivaloylamino-beta-(2-furyl)acrylates]. The comparison of changes in the chemical shifts of H(beta) vinylic protons in CDCl3 and deuterated TFA seems to be the most reliable and useful configuration criterion, as it can be used in the case of a single isomer.  相似文献   

19.
Flash photolysis of diazophenylacetamide in aqueous solution produced phenylcarbamoylcarbene, whose hydration generated a transient species that was identified as the enol isomer of mandelamide. This assignment is based on product identification and the shape of the rate profile for decay of the enol transient, through ketonization to its carbonyl isomer, as well as by the form of acid-base catalysis of and solvent isotope effects on the decay reaction. Rates of enolization of mandelamide were also determined, by monitoring hydrogen exchange at its benzylic position, and these, in combination with the ketonization rate measurements, gave the keto-enol equilibrium constant pK(E) = 15.88, the acidity constant of the enol ionizing as an oxygen acid, pQ(E)(a)= 8.40, and the acidity constant of the amide ionizing as a carbon acid pQ(K)(a)= 24.29. (These acidity constants are concentration quotients applicable at ionic strength = 0.10 M.) These results show the enol content and carbon acid strength of mandelamide, like those of mandelic acid and methyl mandelate, to be orders of magnitude less than those of simple aldehydes and ketones; this difference can be attributed to resonance stabilization of the keto isomers of mandelic acid and its ester and amide derivatives, through electron delocalization into their carbonyl groups from the oxygen and nitrogen substituents adjacent to these groups. The enol of mandelamide, on the other hand, again like the enols of mandelic acid and methyl mandelate, is a substantially stronger acid than the enols of simple aldehydes and ketones. This difference can be attributed to the electronegative nature of the oxygen and nitrogen substituents geminal to the enol hydroxyl group in the enols of mandelic acid and its derivatives; in support of this, the acidity constants of these enols correlate well with field substituent constants of these geminal groups.  相似文献   

20.
The structures of seven gas phase identity S(N)2 reactions of the form CH(3)X + X(-) have been characterized with seven distinct theoretical methods: RHF, B3LYP, BLYP, BP86, MP2, CCSD, and CCSD(T), in conjunction with basis sets of double and triple zeta quality. Additionally, the energetics of said reactions have been definitively computed using focal point analyses utilizing extrapolation to the one-particle limit for the Hartree-Fock and MP2 energies using basis sets of up to aug-cc-pV5Z quality, inclusion of higher order correlation effects [CCSD and CCSD(T)] with basis sets of aug-cc-pVTZ quality, and additional auxiliary terms for core correlation and scalar relativistic effects. Final net activation barriers for the reactions are E(b)(F,F)= -0.8, E(b)(Cl,Cl)= 1.6, E(b)(CN,CN)= 28.7, E(b)(OH,OH)= 14.3, E(b)(SH,SH)= 13.8, E(b)(NH2,NH2)= 28.6, and E(b)(PH2,PH2)= 25.7 kcal mol(-1). General trends in the energetics, specifically the performance of the density functionals, and the component energies of the focal point analyses are discussed. The utility of classic Marcus theory as a technique for barrier predictions has been carefully analyzed. The standard Marcus theory results show disparities of up to 9 kcal mol(-1) with respect to explicitly computed results. However, when alternative approaches to Marcus theory, independent of the well-depths, are considered, excellent performance is achieved, with the largest deviations being under 3 kcal 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