Скачиваний:
37
Добавлен:
15.08.2013
Размер:
345.75 Кб
Скачать

9. Acidity and basicity

399

performed using 0.1 0.5 M solutions, the results may have been perturbed by aggregation effects, as pointed out by the authors and by Streitwieser117. Streitwieser proposed an indicator scale of caesium ion pairs in THF. A table of pKa values of organic indicator acids measured in THF, CHA, DMSO and DME has been also reported118. In THF and CHA acidities have been placed on an absolute basis by assigning a pKa of 18.49 to 9-PhFl (9-phenyl fluorene), while pKa values of 17.9 and 17.55 have been determined in DMSO and DME respectively. (In DME the values were referred to 1,1,3,3,-tetraphenylpropene with an assumed pKa value of 25.25, statistically corrected.)

For a number of carbon acids, an excellent correlation exists between the relative pKa values in these four solvents:

pKa(solvent) D slope ð pKa THF C intercept

(24)

Solvent

 

Slope

Intercept

r2

 

CHA

D

1.02

0.307

0.999

 

DMSO

D

0.934

C1.52

0.995

 

DME

D

0.999

0.402

0.997

 

The independence of CH carbon acidities from the solvent can be attributed to the fact that only highly delocalized carbanions have been considered. When aggregation or specific solvent ion-pairing effects are operative, deviations from this behaviour are expected.

Due to the high interest in metalation reactions with lithium amide or alkyllithiums, an indicator scale of lithium ion pairs in THF has been developed119. Aggregation studies have indicated that organolithium species exist predominantly, if not exclusively, as monomers in the 10 3 10 4 M concentration range. Particular attention has been devoted to the lithium and caesium ion-pair acidities of diphenylamine in THF120 that, at 25 °C, have been found to be 19.05 and 24.20, respectively.

A range of more than 10 pKa units is presented by the THF121 acidities of 15 secondary amines having either two alkyl or an alkyl and a silyl, or two silyl substituents of different bulk, as shown in Table 9.

TABLE 9. pKa values for secondary amines in THF121

Amine

pKa

(CH3)3SiNHSi(CH3)3

29.5

(CH3)2CHNHSi(CH3)3

31.4

(CH3)3CCH2NHSi(CH3)3

33.2

(CH3)3CNHSi(CH3)3

33.6

C2H5CH(CH3)NHCH2CH2CH3

35

cis-2,6-Dimethylpiperidine

35.2

(CH3)2CHNHCH(CH3)2

35.7

Dicyclohexylamine

35.7

2,2,6,6-Tetramethylpiperidine

37.3

NH

 

 

37.6

NH

N(CH3 )2

 

37.9

(continued overleaf )

400

 

Silvia Bradamante

 

TABLE 9.

(continued)

 

 

 

 

Amine

pKa

 

 

 

 

 

NHC(CH3 )3

 

 

38.3

 

 

NHC(CH3 )3

 

 

39.1

 

 

NHC(CH3 )3

 

 

39.5

NHCH2 C(CH3 )3

40.0

The first acidity measurement of aniline was reported by McEwen122. He assigned an ion-pair pKa of 27 to aniline in benzene in relation to MeOH in benzene, which was assumed to have an ion-pair pKa of 16. Acidities of a number of substituted anilines were measured by means of the H method in H2O DMSO mixture123, and by NMR124, calorimetric125 and electrochemical126 methods in liquid ammonia at 31 °C.

The equilibrium acidities of aniline and 26 of its derivatives have recently been measured127 in DMSO solution and they cover a range of 15 pKa units. The pKa values of anilines in DMSO are 10 units higher than the ion-pair pKa values obtained in liquid ammonia. This effect has been attributed to the greater capacity of ammonia to solvate the proton. When the pKa values determined in DMSO were compared with those derived in H2O DMSO or EtOH DMSO, it was evident that the latter were appreciably lower. This behaviour is different from that found for the carbon acids. The H2O DMSO mixture can solvate nitranions better than pure DMSO, in line with the fact that nitranions are better HBA than carbanions128 (see Table 10).

A plot of the pKa values of anilines in DMSO versus the pKa values of anilinium ions in water is linear with a slope of 1.8. This allows the extrapolation of 41 š 1 for the pKa of ammonia in DMSO from that of NH4C in water (9.27)127. Alternatively, the pKa of ammonia in DMSO has been extrapolated105 from the intersystem correlation between the DMSO acidities of NH2X and PhNHX as a value of 35.8. Extrapolation of the pKa value of ammonia from the Taft-like dual substituent parameter (DSP) of NH2X DMSO acidities gave a similar value of 36.6.

C. NH Bond Dissociation Energy

Bordwell129 developed a method of estimating relative bond dissociation energies (BDE) for families of acids, HA, by combining equilibrium acidity constants, pKHA, with the oxidation potential of their conjugated bases, A , both measured in DMSO:

BDE kcal mol 1 D 1.37 pKHA C 23.06 Eox A

25

 

 

9. Acidity and basicity

 

 

 

401

 

TABLE 10. Equilibrium acidities of aromatic amines determined by

 

 

 

the H method in H2O/DMSO and pure DMSO

 

 

 

 

 

 

 

pKa (DMSO)

pKa (H )

 

 

 

Anilines

 

 

 

 

 

 

 

3-Me

 

31.0

 

 

 

 

 

H

 

30.7

 

 

 

 

 

3-MeO

 

30.5

 

 

 

 

 

3-Cl

 

28.5

25.6

 

 

 

 

3-CF3

 

28.2

25.4

 

 

 

 

3-Br

 

28.4

 

 

 

 

 

3-CN

 

27.5

24.6

 

 

 

 

4-MeSO2

 

25.6

 

 

 

 

 

4-CN

 

25.3

22.7

 

 

 

 

4-PhSO2

 

24.9

 

 

 

 

 

4-PhCO

 

24.4

 

 

 

 

 

4-NO2

 

20.9

 

 

 

 

 

2-F

 

28.7

 

 

 

 

 

4-F

 

30.7

 

 

 

 

 

2,4-(F)2

 

28.6

 

 

 

 

 

2-Cl

 

27.6

 

 

 

 

 

4-PhS

 

28.2

 

 

 

 

 

4-MeCO

 

25.3

 

 

 

 

 

4-CF3SO2

 

21.8

 

 

 

 

 

4-Cl

 

29.4

 

 

 

 

 

4-CF3

 

27

 

 

 

 

 

4-Br

 

29.1

 

 

 

 

 

2,4-(NO2 )2

 

15.9

15

 

 

 

 

4-NO2-2,5(Cl2)2

 

17.4

 

 

 

 

 

2,6-(Cl)2

 

24.8

 

 

 

 

 

2,4(Cl)2

 

26.3

 

 

 

 

 

Pyridines

 

 

 

 

 

 

 

4-NH2-pyridine

 

26.5

22.3

 

 

 

 

2-NH2-pyridine

 

27.7

23.5

 

 

 

 

Diphenylamines

 

 

 

 

 

 

 

4-NO2C6H4NHPh

 

16.85

15.7

 

 

 

 

3-ClC6H4NHPh

 

23.4

20.7

 

 

 

 

Ph2NH

 

24.95

22.4

 

 

 

TABLE 11. Homolytic bond dissociation energies (BDEs) of some H N bonds133

 

 

 

pKHAa

Eox A b

BDE (kcal mol 1)c

BDE (kcal mol 1)d

PhNH2

30.6

0992

92 š 1

88 š 2e

e

PhNHMe

29.5

1.054

89.0

87.51š 2

 

Ph2NH

24.95

0.856

87.5 š 1

87.3

g

 

Pyrrole

23.05

0.355

97.0

99 š 6

 

 

a Measured in DMSO.

bMeasured by cyclic voltammetry in DMSO. c Calculated.

dLiterature data. eReference 134. fReference 135. gReference 136.

402

Silvia Bradamante

 

Subsequently, he used the equation proposed by Friedrich130 to estimate BDEs:

 

 

BDE kcal mol 1 D 1.37 pKHA C 23.06Eox A C C

26

with C D 55.4 55.9 kcal mol 1 (a value of 56 has also been used, and more recently one of 73131,132). The BDEs estimated in this way agree remarkably well with the values obtained in the gas phase (š2 kcal) (Table 11), although the relationship must be considered empirical133. The cleavage modes of radical anions have also been considered131:

H + A

path A

HA

(27)

H+ A

path B

In the case of anilines, path A will be favoured on the basis of the values of oxidation potentials, in analogy with the gas-phase results.

An extensive study132 has been presented on the polarizability effects of alkyl groups in RX moieties (R D Me, Et, i-Pr and t-Bu; X D CH2, S, SO2, O and N) in families of weak acids and on the stabilities of adjacent anions and radicals in DMSO solution. Some of the results related to the 9-(dialkylamino)fluorenes are given in Table 12. The increases in acidity are believed to be caused by the progressive increases in anion stabilizing

TABLE 12. Acidities and homolytic bond dissociation energies of the acidic C H bonds in 9-(dialkylamino)fluorenes (9-R1R2NFlH)

Substituted amino

pKHAa

Eox A b

BDE

(R1R2N)

 

 

 

 

1

2

D Me

22.5

1.418

71.5

R1

D R2

R1

D R2

D Et

21.4

1.388

70.5

R

D R

D i-Pr

20.8

1.242

73

 

N

 

22.5

1.382

72

 

N

 

21.4

1.348

71.5

 

 

 

 

N

 

19.4

1.198

72

N

18.2

1.166

71

a Equilibrium acidities measured in DMSO: Reference 139.

b Oxidation potentials of the conjugated anions measured in DMSO solution and referenced to the ferrocene/ferrocenium couple.

9. Acidity and basicity

403

polarizability effects of the R group that parallel increases in alkyl size. Polarizability effects stabilize anions but not the analogous radicals.

Taking advantage137 of the outlined method and considering that RSE (radical stabilization energy) is equivalent to the variation in BDE, Bordwell estimated the RSEs of radicals of the type A Cž H D, where A is an electron acceptor and D an electron donor. Comparison of the oxidation potential of ˛-H2N and ˛-R2N carbanions derived from fluorene, acetophenones and malononitrile indicates that N-alkylation has little or no effect on the stability of the corresponding ˛-amino radicals (no captodative effect). Nevertheless, when saturation effects are taken into account there is evidence of a synergistic (captodative) effect. This indicates that the presence of strong donor and acceptor groups attached to a radical centre stabilizes the radicals with synergistic effects, in accord with the previous literature data138.

RSE D BDE

28

D. Carboxamides and Related Compounds

A comparison of acidities of six series of analogous oxygen, nitrogen and carbon acids in DMSO solution and in the gas phase has shown that the nitrogen acids are more acidic than their carbon acid counterparts by 17 š 5 kcal mol 1 (kcal mol 1 D 1.37 pKa unit), and that the oxygen acids are more acidic than the nitrogen acids by a similar amount. Differently, carboxamides have been found only slightly more acidic than their carbon acid analogues, the ketones (1 2 kcal mol 1 in DMSO and 7 8 kcal mol 1 in the gas phase140). This behaviour has been rationalized in terms of destabilization of carboxamide conjugate bases by lone pair lone pair repulsions, together with some resonance stabilization of the carboxamide.

The replacement of the CDO bond in carboxamides by the CDS bond, as in thio-

amides141, causes increases in acidity of the N H bond by 8 10 kcal mol 1

and

 

 

 

 

weakening of the N H bond by 14 17 kcal mol 1, evaluated by using relation 26. This effect has been considered a consequence of large increases in ground state energies and the superior ability of sulphur than oxygen in stabilizing a negative charge or odd electron.

In the gas phase, the acidity of H C bond increases as the s character of carbon increases and the homolytic bond dissociation energy decreases as shown by considering the following compounds: H3CCH3, H2CDCH2 and HC CH with Gacid kcal mol 1 D 413, 401 and 370 and BDE kcal mol 1 D 98, 110 and 132, respectively. A similar trend with the changes in hybridization of nitrogen was not demonstrated, due to the paucity of available data. The problem could be more complicated in the case of substrates containing more acidifying centres as acetamidine, for which the following scheme can be considered:

 

NH

 

 

 

 

 

NH

 

 

 

 

 

NH

H+

 

 

 

NH

 

 

 

 

 

 

 

 

 

 

 

 

 

 

+

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

 

CH3

NH2

 

CH3

+H+

CH3

 

CH3

NH

 

NH2

 

 

NH

 

 

(16)

 

 

(16)

 

 

 

(16a)

 

 

(16a)

Although an analogous scheme can be written for acetamide, recent ab initio calculations142 have questioned the importance of resonance in amides, i.e. 17 $ 170 , indicating that the nitrogen in 17 has a slight negative charge rather than the positive

404

Silvia Bradamante

charge imposed by the resonance contributor 170.

 

O

 

O

 

 

H+

O

 

 

O

 

 

 

 

+

 

C

 

 

C

 

C

 

C

 

 

 

 

 

 

 

 

 

CH3

NH2

 

CH3

NH2 +H+ CH3

NH

 

CH3

NH

 

(17)

 

(17)

 

 

 

 

(17a)

 

 

(17a)

The equilibrium acidities in DMSO have been measured for acetamidine, benzamidine, N,N0 -diphenylbenzamidine, N,N-diethylbenzamidine, diphenylmethanimine, guanidine, N,N0-diphenylguanidine, N,N0-diphenylurea, and N,N0-diphenylthiourea143. Combination of the resulting pKa values with the oxidation potentials of their conjugated bases gave estimates of their BDEs. These values have been compared with those of the corresponding carboxamides and thiocarboxamides. The changes in hybridization of nitrogen between NH3 and Ph2CDNH produce a changes in acidities and BDEs similar to the increases observed for the increase in the s character of carbon. Nevertheless, the BDE

of H

 

N in HN3

is 25 kcal mol 1

lower than that in Ph2C

D

NH, despite the apparent

similarities in hybridization.

 

 

 

Thioacetamide

is more acidic

than acetamide and

acetamidine by 9.6 and

11.8 kcal mol 1 (Table 13). Replacement of the methyl group by an amino group causes a decrease in acidity. The effects of substituting one phenyl group at the acidic sites produce an increase in acidity, not different from the increase obtained by introducing a phenyl group at each of the nitrogen atoms. This is due to the strong steric interaction between the two phenyl groups that allow only one phenyl group to overlap with the negative charge or odd electron.

The equilibrium acidities of aldoximes, amidoximes and ketoximes144 have also been measured in DMSO solution. It has been found that syn- and anti-benzaldoximes have nearly the same acidities, contrary to the finding in aqueous solution where the anti isomers are less acidic. The lower acidity of the anti isomers in aqueous solution is attributable to steric inhibition of solvation of the strongly hydrogen-bonded oxide ion by the phenyl group, a factor that is absent in DMSO. The 8.3 pKa unit difference between the acidity of acetaldoxime (28.5) and benzaldoxime (syn- 20.2, anti- 20.3) in DMSO points to appreciable delocalization of the negative charge in the benzaldoximide ion into the benzene ring.

Ionization equilibria of sulphamide, phenylsulphamide, mono-, diand tri-substituted sulphamides have been considered and pKa values have been determined145 for

TABLE 13. Acidities of amidines and related compounds in DMSO

 

pKa

CH3C(DNH)NH2

27.1

CH3C(DO)NH2

25.5

CH3C(DS)NH2

18.5

PhC(DNH)NH2

26.7

PhC(DO)NH2

23.3

CH3C(DS)NH2

16.9

Ph2CDNH

31.0

(H2N)2CDNH

28.5

(H2N)2CDO

26.9

(H2N)2CDS

21.0

 

 

9. Acidity and basicity

 

405

equilibrium a in 60% v/v EtOH

 

H2O:

 

 

 

 

 

 

 

RNHSO2NHR

a

RN SO2NHR

b

RN SO2N R

 

!

!

(29)

 

 

 

 

 

 

RNHSO2NMe2

 

a

RN SO2NMe2

 

 

 

!

 

 

(30)

 

 

 

 

 

 

R D H, X-aryl, alkyl

It has been found that in the case of equilibrium 30 the corresponding aliphatic series exhibit decreased pKa values by about 5 pKa units146.

Under the same experimental conditions147 the following equilibria have also been considered and pKa values have been determined to be in the range 9 12.

a

X-C6H4N SO2NR1R2

 

X-C6H4NHSO2NR1R2 !

(31)

 

 

 

X D Me, H, Br, Cl, OMe, OEt, NO2

R1 D H, Ac, Me

R2 D cyclo-C6H11, Bu, CH2C6H5

When cyclic sulphamides (18 and 19) were studied148, it was found that the sixmembered rings are more acidic than their acyclic analogues by ca 2.5 pKa units and that the five-membered cyclic sulphamides are ca 4 pKa units more acidic than the model open-chain counterparts.

H

H

 

N

N

SO2

SO2

 

N

H

 

N

 

H

(18)

(19)

VI. HYDROXAMIC ACIDS AS AMPHOTERIC SYSTEMS

Notwithstanding their name, hydroxamic acids (20) are amphoteric systems with behaviours that are both basic and acidic. The chemistry of hydroxamic acids and N- hydroxyimides was reviewed by Bauer and Exner149 in 1974. Since then, the basic properties of hydroxamic acid derivatives have received little attention. Given the similarity of the activity coefficient behaviour of hydroxamic acids and other carbonyl bases (amides), it was originally assumed that the carbonyl oxygen rather than the nitrogen atom was the site of protonation150,151, but some recent IR studies have revealed signs of N-protonation152. This question has been recently addressed by Bagno and colleagues153, who concluded from NMR measurements made in aqueous solution that acetohydroxamic acid undergoes carbonyl protonation as the dominant process. The protonation equilibrium of this compound was determined in aqueous sulphuric acid and the use of the excess acidity method led to mŁ D 0.25 and pKBH D 1.15. The gas-phase basicity of some

406

Silvia Bradamante

hydroxamic acid derivatives (Table 14) has been determined154, the values indicating that they are 7 kcal mol 1 weaker bases than amides. The protonation site cannot be determined from gas-phase results, but a pH-metric study of the dissociation constants of L-˛-alaninehydroxamic and ˇ-alaninehydroxamic acids has shown that the NH3C group is more acidic than the NHOH group for the ˛-derivative, while the reverse is true for the ˇ-isomer155.

 

 

 

 

 

O

 

 

 

 

 

 

 

N

 

 

 

 

 

 

 

 

 

 

 

R

OH

 

 

 

 

 

O

H

 

 

 

 

 

 

N

 

 

 

 

 

 

 

 

 

 

 

R

O

O

 

H

H+

 

 

 

 

 

N

 

 

 

 

R

 

OH

 

 

 

 

 

+H +

 

H O+

H

 

 

 

 

 

 

 

 

 

 

 

N

 

 

 

 

 

 

 

 

 

 

 

R

OH

OH

 

 

 

 

O

H

 

 

 

 

 

 

 

 

+ H

 

 

 

 

 

 

 

N

 

 

 

N

 

 

 

 

 

 

 

 

 

 

 

R

 

OH

 

 

R

OH

 

 

 

 

 

 

(20)

 

 

O

H

 

 

 

 

 

 

 

 

 

 

 

N

 

 

 

 

 

R

O+

 

 

 

 

 

 

H

H

TABLE 14. Gas-phase basicities of some hydroxamic acid derivatives (kcal mol 1)

Compound

GB

 

 

MeCONHOH

197.3

MeCONMeOH

200.9

MeCONHOMe

201.3

 

 

 

9. Acidity and basicity

407

TABLE 15. Gas-phase acidities of acetohydroxamic acid,

acetamides and their methyl derivatives

 

 

 

 

 

 

 

N

Compounds

Gacid°

(kcal mol 1)

1

MeCONHOH

 

339.1

 

2

MeCONHOMe

 

343.7

 

3

MeCONMeOH

 

346.9

 

4

MeCONH2

 

355.0

 

5

MeCONHMe

 

354.5

 

 

 

 

 

 

There has been a longstanding controversy as to whether hydroxamic acids are NH or OH acids. The IR and UV evidence presented by Exner and coworkers149 indicated that they are NH acids in dioxane and aqueous alcohol solvents, and this conclusion has also been supported by a 17O NMR study156. However, other studies157 on meta- and para-substituted benzohydroxamic acids led to the conclusion that, in aqueous solution, RCONHO ions were at least as present as RCON(OH) ions. More recently, Crumbliss and coworkers158 have concluded that, in an aqueous 2M NaNO3 solution, acetoand benzohydroxamic acids act as OH rather than NH acids. Bordwell’s group159 measured the equilibrium acidities in DMSO of acetoand benzohydroxamic acids, as well as their N- and O-alkyl derivatives, and concluded that hydroxamic acids act as NH acids in non-HBD solvents, such as DMSO, DMF, CH3CN etc., but they may act primarily as OH acids in hydroxylic solvents.

The gas-phase acidities of hydroxamic acid and its N- and O-methyl derivatives have been measured using FT ion cyclotron resonance (Table 15)160. The acidity order is the same as that found in DMSO although, in this solvent, the O-methylation of acetohydroxamic acid decreases the acidity by 1 pK unit and the N-methylation by 3.6 pK units, respectively.

VII. HETEROCYCLES

Imidazole is a five-membered heterocycle with both basic and acidic properties. As part of histidine, it plays an important role in proton transfer processes at the active site of certain enzymes, and is thus considered a good model for studying enzymatic modes of action by means of computational approaches. Much work has been done to characterize this molecule and its derivatives. The proton affinity161 of imidazole has been determined experimentally by Taft and coworkers, who also determined the gas-phase basicities of pyrazole, all possible isomeric methylpyrazoles (13 derivatives), and a selected set of methylimidazoles (8 derivatives), using STO-3G fully-optimized geometries of neutral molecules and their corresponding cations. The protonation energies of imidazole, calculated subsequently by Meyer162 using the HF/6-311GŁŁ basis set with additional polarization functions on hydrogen atoms163, are lower than those previously calculated without polarization functions. The substitution of hydrogen with a methyl group causes a small increase in the deprotonation/protonation energies of 1 kcal mol 1. The overall picture indicates that imidazole and 4-methyl imidazole seem to be interchangeable in model studies of the protonation of histidine.

Knowing the equilibrium constants for the protonation of the amino and imidazole nitrogens, and the kinetics of deprotonation of the C(2) carbon of imidazole, it was possible to investigate nitrogen-protonation microequilibria and the C(2)-deprotonation microkinetics of histidine, histamine and other related compounds164.

408

Silvia Bradamante

I s min

14

13

12

2

0

2

4

6

8

4

pKa

FIGURE 7. Correlation between experimentally determined aqueous pKa values and ring nitrogens

Is,min of ten azines and azoles. The least-squares equation of the line is y D 13.2004 0.1468x, with a correlation coefficient of 0.99. Reprinted with permission from Reference 170. Copyright (1991)

American Chemical Society

Elguero’s group reported165 an empirical relationship between pKa values (acid and

basic) for a large number of azoles in water at 25 °C.

 

pKa(NH) D 0.985pKa NHC C 6.95

32

All of the azoles showed a linear variation of these values except the pyrazoles, which belong to a parallel line 4.5 pKa units apart. Fully optimized INDO geometries have been calculated for 12 azoles, as well as their cations and anions, both isolated and specifically solvated by five water molecules166. Evaluation of the protonation and deprotonation energies of the solvated molecules indicates a behaviour similar to that found experimentally in solution. In particular, the difference between pyrazoles (and indazoles) and all the other azoles is a consequence of the protonation of the nitrogen contiguous to NH, that is due to a difference in basicity.

Among the semiempirical methods used to evaluate protonation and deprotonation energies in azoles, only INDO seems to estimate the electrostatic proximity effects167 correctly.

The thermodynamic properties of 3,5-bis(trifluoromethyl)-1,2,4-triazole have been measured and discussed in the light of the conclusions provided by ab initio calculations (MP2/6-31GŁ//6-31GŁ level of accuracy) applied to trifluoromethyl-substituted triazoles, the parent triazole and pyrazole. It is worth noticing that 21 is more stable than 21a by 6.7 kcal mol 1168 .

N

N N

N

 

N

N

H

H

(21)

(21a)

Соседние файлы в папке Patai S., Rappoport Z. 1996 The chemistry of functional groups. The chemistry of amino, nitroso, nitro and related groups