Vous êtes sur la page 1sur 6

Highlights

Organocatalysis

Nucleophilic Catalysis by 4-(Dialkylamino)pyridines


Revisited—The Search for Optimal Reactivity and
Selectivity
Alan C. Spivey* and Stellios Arseniyadis
Keywords:
asymmetric catalysis · esterification · N ligands ·
nucleophilic catalysis · organocatalysis

4-(Dimethylamino)pyridine (4-DMAP, associated with this apparently straight- acetylation of 1-methylcyclohexanol


1) is well known as a catalyst for the forward process is highlighted. with Ac2O (Scheme 3).[17]
esterification of alcohols by acid anhy- That pyridine and 4-substituted de- Hassner et al. noted the lack of
drides and for various other syntheti- rivatives act primarily as nucleophilic correlation between the pKa value and
cally useful transformations involving rather than general base catalysts for the catalytic activity; they suggested that
acyl transfer.[1, 2] Its catalytic potential alcohol esterification follows from the the relative efficiencies of the various
was first discovered by the groups of dramatic loss of activity that accompa- catalysts reflected the stabilities of the
Litvinenko and Steglich in the late nies 2-alkyl substitution despite the respective derived acyl pyridinium in-
1960s[3, 4] and its synthetic utility and relatively marginal effect that this sub- termediates in a mechanistic scenario
that of its congeners, including polymer- stitution has on the pKa value of these involving equilibrium formation of these
ic variants,[5] have been reviewed.[6–9] derivatives. Such steric inhibition of salts followed by rate-determining reac-
Recently, attention has been focused catalysis was first shown to be character- tion with the alcohol (Scheme 4).
on the development of enantiomerically istic of nucleophilic catalysis in work by Additionally, Hassner et al. noted
pure chiral 4-(dialkylamino)pyridines Gold and Jefferson in the early 1950s on that the order of catalytic activity of 4-
for the kinetic resolution of alcohols the hydrolysis of Ac2O with a series of aminopyridine derivatives [4-pyrrolidi-
and related enantioselective transfor- methyl-substituted pyridines.[14, 15] The no (2) > 4-dimethylamino (1) > 4-piper-
mations.[10, 11] As a result of this interest, effect was quantified by Litvinenko idino (5) > 4-morpholino (7)] mirrored
the detailed mechanism of catalysis by and co-workers in 1981 for the catalysis the order of reactivity of cyclohexa-
4-(dialkylamino)pyridines and the fac- of benzoylation of benzyl alcohol with none-derived enamines towards electro-
tors that influence their reactivity have BzCl.[16] In addition to confirming the philes.[18–24] This order has been ration-
come under renewed scrutiny. In partic- nucleophilic nature of the catalysis, this alized as a balance of stereoelectronic
ular, Steglich and co-workers[12] report- work also highlighted the particularly (nN !p*C=C) and steric effects (A1,3
ed pyridonaphthyridine 3 as being the high catalytic activity of 1, which exhib- strain) which dictates the efficiency with
most catalytically active 4-DMAP ana- its a rate of 3.4 ; 108 relative to the which the lone pair of electrons on the
logue yet prepared for the acetylation of uncatalyzed reaction (Scheme 2). enamine nitrogen atom interacts with
tertiary alcohols, and work by Kattnig The high catalytic reactivity of 1 had the CC double bond. By analogy with
and Albert[13] has illustrated the key role previously been noted by Litvinenko the enamine series,[19] Hassner et al.
of the anion and general base catalysis in and co-workers in the benzoylation of 3- noted that there was a qualitative corre-
regulating the rate and regioselectivity chloroaniline[3] and subsequently, but lation between the degree of shielding of
of polyol acetylation by 1 (Scheme 1). independently, 1 was shown by Steglich the pyridyl b-hydrogen atoms in the
1
Herein, the detailed mechanism of and co-workers to enable esterification H NMR spectra of the catalytically
catalysis of esterification by 4-(dialkyl- of even hindered tertiary alcohols with active 4-aminopyridine derivatives and
amino)pyridines is reexamined in light Ac2O.[4] Esterification reactions of ter- their catalytic efficiency (see
of these findings, and the complexity tiary alcohols are relatively slow and Scheme 3).[17] They inferred that the
particularly susceptible to steric factors extent of electronic communication be-
and therefore proved to be useful for tween the lone pair of electrons of the
[*] Dr. A. C. Spivey, S. Arseniyadis exploring structure–activity relations for exocyclic nitrogen atom and the carbon-
Department of Chemistry catalysis by 4-DMAP analogues. Ac- yl function through the pyridyl ring was
South Kensington Campus
cordingly, Hassner et al. found that 4- a key factor in stabilizing the acyl
Imperial College
London SW7 2AZ (UK) pyrrolidinopyridine (4-PPY, 2) was the pyridinium intermediate.
Fax: (+ 44) 20-7594-5841 most efficient of a series of 4-amino- The design of pyridonaphthyridine 3
E-mail: a.c.spivey@imperial.ac.uk pyridine derivatives, including 1, for the (Scheme 1), recently disclosed by Steg-

5436  2004 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim DOI: 10.1002/anie.200460373 Angew. Chem. Int. Ed. 2004, 43, 5436 –5441
Angewandte
Chemie

Scheme 1. Relative rates of acetylation of 1-ethynylcyclohexanol with Ac2O catalyzed by 4-DMAP (1), 4-pyrrolidinopyridine (4-PPY, 2), or pyrido-
naphthyridine 3;[12] and the regioselectivity of acetylation of octyl b-d-glucopyranoside with AcCl/pyridine versus Ac2O/pyridine catalyzed by 4-
DMAP (1).[13]

dimethylamino or pyrrolidi-
no analogues (i.e. 21 or 22)
as judged by their relative
reactivities towards, for ex-
ample, the Danishefsky di-
ene (Scheme 5).[25, 26]
As the tricyclic scaffold
found in derivative 24 allows
Scheme 2. Catalysis of benzoylation of benzyl alcohol with the most efficient electronic
BzCl by substituted pyridines.[16] communication between the
lone pair of electrons of the
amine and the cation, pyri-
donaphthyridine 3 was pre-
lich and co-workers,[12] evolved from this dicted to form highly stabilized acyl Scheme 4. Mechanism of nucleophilic cataly-
analysis. Prior studies by Mayr et al. had pyridinium salts and consequently to sis of alcohol esterification according to Hass-
ner et al.[17] rds = rate-determining step.
established a quantitative framework be a particularly efficacious nucleophilic
for comparing the influence of 4-dial- acylation catalyst. Moreover, density
kylamino substitution on benzhydryl functional theory calculations of the
cation reactivity and specifically had reaction enthalpies for acetyl transfer be the most catalytically active 4-
shown that “conformationally fixed” from pyridine to 3 corroborated the DMAP analogue yet prepared, display-
cations such as 23 and 24 are signifi- expected stability of the acetylpyridini- ing a rate of  6 relative to 1 for the
cantly more stabilized than either the um salt of 3. In the event, 3 was shown to acetylation of 1-ethynylcyclohexanol

Scheme 3. Catalysis of acetylation of 1-methylcyclohexanol with Ac2O by substituted pyridines and related azines.[17]

Angew. Chem. Int. Ed. 2004, 43, 5436 –5441 www.angewandte.org  2004 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim 5437
Highlights
a more detailed catalytic cycle is pre- (e.g. H2O > hexane). This is partic-
sented below (Scheme 6). ularly pronounced because in sol-
In this four-step cycle, reversible vents of low polarity a solvated ion
formation of N-acylpyridinium salt II pair is expected, with a high propen-
by the addition of 1 to the acyl donor sity to return in a quasi-unimolecular
via the tetrahedral intermediate I fashion to starting materials, where-

Scheme 5. Relative reactivities of various


benzhydryl cations towards the Danishefsky
diene as a measure of the ability of the 4-dial-
kylamino group to participate in nN !p*C=C
overlap.[26]
Scheme 6. Proposed catalytic cycle for alcohol esterification with 4-DMAP (1).

with Ac2O and with Et3N as auxiliary (steps 1 and 2) is followed by irreversi- as in highly polar solvents the dis-
base (Scheme 1). ble nucleophilic addition of the alcohol sociated ions will experience a sig-
These results clearly have practical to salt II (step 3, !III) with concomi- nificant activation entropy for re-
synthetic significance for the esterifica- tant proton transfer (through transition turn.
tion of highly hindered alcohols and state III#) and finally elimination to
lend support to the notion that the regenerate 1 (step 4). Evidence to support these expect-
efficiency of nucleophilic catalysis by The position of the equilibrium to ations has been summarized previous-
this class of compound is closely related form salt II is dictated by the relative ly.[9] Of particular note were VT-NMR
to the stability of the intermediate affinities of the anion X and 1 for the spectroscopy experiments (VT = varia-
acylpyridinium salt. However, the mech- acyl group in the reaction solvent. ble temperature), which showed that
anistic framework represented in Because of the aforementioned influ- mixtures of 2 and Ac2O (1:1.5) in CDCl3
Scheme 4 is simplistic. Although the ence of steric effects and conjugation consist of 5–10 % acetylpyridinium ace-
kinetics of the esterification by 1 under (IIa$IIb$IIc, Scheme 6)[28, 29] these tate at room temperature, whereas
standard conditions in a low-polarity “acyl affinities” will not mirror pKa equivalent mixtures of 2 and AcCl
solvent are broadly consistent with the values,[30] but the equilibrium will be consist of  100 % acetylpyridinium
steady-state formation of N-acylpyridi- shifted in favor of N-acylpyridinium salt chloride. Parallel experiments with pyr-
nium salt followed by the rate-determin- formation for: idine resulted in no detectable forma-
ing reaction of this salt with the alco- * pyridines that have high nucleophi- tion of N-acetylpyridinium acetate and
hol,[27] this analysis fails, at least explic- licity and impart stabilization of the precipitation of insoluble N-acetylpyri-
itly, to account for a number of impor- acyl group by conjugation (e.g. 1 > dinium chloride respectively. Solubility
tant features of the catalysis, such as the pyridine), is an important factor that affects the
significant differences in the rate of * anions that have high nucleofugacity concentration of reactive N-acylpyridi-
catalysis when employing acid chlorides (e.g. Cl > OAc), and nium salts in solution: N-Acyl 4-(di-
versus anhydrides and when varying the * polar solvents that can solvate the alkylamino)pyridinium salts II tend to
auxiliary base.[9, 12, 13] To facilitate a dis- charge-separated salt more efficient- be significantly more soluble in non-
cussion that encompasses these features, ly than the neutral starting materials polar organic solvents than the corre-

5438  2004 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim www.angewandte.org Angew. Chem. Int. Ed. 2004, 43, 5436 –5441
Angewandte
Chemie

sponding unsubstituted pyridinium salts. strong evidence for the case of 4-DMAP
Loss to competing reaction manifolds, (1) catalysis. They found that when using
such as elimination to form ketenes, also K2CO3 (4 equiv) as auxiliary base, the 4-
has an impact on the concentration of DMAP-catalyzed (5 mol %) acetylation
salt.[31, 32] Ketene formation is only pos- of 2-propanol in CDCl3 at room temper-
sible when the acyl function has b- ature with Ac2O (2 equiv) proceeded
hydrogen atoms and is retarded by about 10 times faster than the analogous
conjugation between the carbonyl group reaction with AcCl (t1/2 = 18 min vs.
and the 4-dialkylamino substituent. Sus- 200 min). In contrast, with pyridine
ceptibility to ketene formation increases (2 equiv) as auxiliary base the relative
with the basicity of the auxiliary base. Scheme 7. Relative rates of 4-DMAP-mediated rates were dramatically reversed
The rate of esterification through this acetylation of 1-ethynylcyclohexanol with AcCl ( 700 times rate difference: t1/2 =
pathway is slow relative to that through and Ac2O.[9] 120 min vs. < 10 s). Similar results were
nucleophilic catalysis.[33] obtained with 1-propanol (Scheme 8).
The overall rate of the catalyzed
reaction clearly depends not only on the mented: The highest rates are found in
concentration of the N-acylpyridinium nonpolar solvents[9] which appears para-
salt II (and of the alcohol) in solution doxical given that such solvents accom-
but also the reactivity of the salt towards modate only low concentrations of N-
addition under the reaction conditions. acylpyridinium salts (see above)!
That a favorable balance between these How then can we account for these
two factors is required is apparent from reactivity trends? The key is to examine
the comparison between AcCl-mediat- solvation and general base catalysis by
ed acetylation of alcohols catalyzed by the anion as crucial factors in dictating
pyridine and 1 in nonpolar solvents: the the reactivity of N-acylpyridinium salts.
carbonyl group in N-acetylpyridinium Strong solvation by polar solvents leads
chloride itself is “intrinsically” more to dissociated ions with low reactivity
activated than in the corresponding 4- (whose reactivities do parallel intrinsic
DMAP-derived salt (IR: ñ(C=O) 1800 carbonyl activation), whereas weak sol-
vs. 1755 cm1, respectively)[9] but pyri- vation by nonpolar solvents leads to
dine fails to mediate efficient catalysis highly reactive ion pairs (whose reactiv-
primarily because its N-acylpyridinium ities are structure-dependent). Conse-
salt is not present in a kinetically mean- quently, in aqueous solution N-acetyl-
ingful concentration. pyridinium chloride hydrolyzes about
A confounding factor in the above 2000 times faster than N-acetyl-4-(dime-
analysis, however, is that the observed thylamino)pyridinium chloride[34] (i.e.
Scheme 8. Relative rates of 4-DMAP-catalyzed
reactivities of N-acylpyridinium salts do the reverse of their relative reactivities acetylation of 1- and 2-propanol with various
not correlate well with their intrinsic towards alcohols in nonpolar solvents as AcCl/Ac2O and K2CO3/pyridine combina-
carbonyl activation as expected from described above). The important struc- tions.[13]
resonance and spectrochemical proper- tural parameters that influence the rel-
ties. In particular, their reactivity is ative reactivity of N-acylpyridinium ion
highly anion- and solvent-dependent.[9] pairs towards alcohols in nonpolar sol- They concluded that when using
The anion dependence is illustrated by vents are ion mobility and the efficacy of insoluble K2CO3, the acetate or chloride
the threefold-greater rate of 4-DMAP- general base catalysis by the anion. anion must act as a general base, where-
mediated (3 equiv) acetylation of 1- The mobility of the constituent ions as when using pyridine deprotonation
ethynylcyclohexanol when using Ac2O dictates the ease of access of the alcohol can also be carried out by this auxiliary
(2 equiv) relative to AcCl (2 equiv) to the reactive carbonyl carbon. base.[36, 37] Their results certainly militate
(Scheme 7).[9] “Loose” delocalized ion pairs (e.g. ace- in favor of deprotonation at the transi-
Considering that < 10 % of the tate/4-(dialkylamino)pyridinium) are tion state III# (Scheme 6) being a critical
Ac2O will be present as N-acetyl-4- more reactive than “tight” ion pairs component of the rate-determining step.
(dimethylamino)pyridinium acetate in (e.g. chloride/pyridinium).[9] The ability This is also consistent with the observa-
one case in contrast to  100 % of the of the anion to deprotonate the alcohol tion by Steglich and co-workers that t1/2
AcCl as N-acetyl-4-(dimethylamino)- nucleophile in the rate-determining for acetylation of 1-ethynylcyclohexanol
pyridinium chloride in the other (see transition state III# will mirror its basic- with Ac2O catalyzed by 1 decreased
above), this rate difference equates to a ity (e.g. acetate > chloride). This general from 465 to 151 min when triethylamine
significantly higher reactivity for the base catalysis has long been mooted as was used as auxiliary base rather than
acetate relative to the chloride salt.[9] being important during nucleophilic pyridine.[12] Moreover, the results shown
The solvent dependence of 4- catalysis,[35] but the recent experiments in Scheme 7 may be partially accounted
DMAP-catalyzed reactions is well docu- of Kattnig and Albert[13] provide the first for by an approximately threefold great-

Angew. Chem. Int. Ed. 2004, 43, 5436 –5441 www.angewandte.org  2004 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim 5439
Highlights
er steady-state concentration of free 4- (Scheme 1). In line with previous work [14] V. Gold, E. G. Jefferson, J. Chem. Soc.
DMAP (1) available to act as a general by Yoshida and co-workers,[40] they at- 1953, 1409 – 1415.
base in the reaction with Ac2O relative tributed the regioselectivity under the [15] A. C. Spivey, A. Maddaford, A. J. Red-
grave, J. Chem. Soc. Perkin Trans. 1
to that with AcCl. Ac2O–pyridine conditions to noncova-
2001, 1785 – 1794.
However, since Kattnig and Albert lent (hydrogen bonding) interactions of [16] L. I. Bondarenko, A. I. Kirichenko,
used just 5 mol % of 4-DMAP (1) it is the acetate ion with the substrate. L. M. Litvinenko, I. N. Dmitrenko,
unclear why their AcCl–pyridine condi- In summary, the search for optimal V. D. Kobets, J. Org. Chem. USSR (Engl.
tions result in much faster rates than catalytic activity for 4-(dialkylamino)- Transl.) 1981, 2310 – 2316; Zh. Org.
their Ac2O–pyridine conditions. The pyridine-catalyzed esterification de- Khim. 1981, 17, 2588 – 2594.
concentrations of free pyridine present mands balancing a delicate raft of fac- [17] A. Hassner, L. R. Krepski, V. Alexani-
an, Tetrahedron 1978, 34, 2069 – 2076.
to act as general base must have been tors. It is hoped that the foregoing
[18] B. Kempf, N. Hampel, A. R. Ofial, H.
comparable, therefore presumably a discussion has illuminated the complex Mayr, Chem. Eur. J. 2003, 9, 2209 – 2218.
much greater concentration of N-ace- interplay between catalyst structure, [19] P. W. Hickmott, Tetrahedron 1982, 38,
tyl-4-(dimethylamino)pyridinium chlo- acylating agent, auxiliary base, and sol- 1975 – 2050.
ride relative to the corresponding acetate vent that conspire to set this balance and [20] W. H. Daly, J. G. Underwood, S. C. Kuo,
must be responsible. Given that the that future studies in this area will shed Tetrahedron Lett. 1971, 12, 4375 – 4378.
former ion pair is less reactive than the additional light. In particular, develop- [21] M. E. Kuehne, T. Garrbacik, J. Org.
latter (see above),[38] the equilibrium ments in asymmetric organocatalysis Chem. 1970, 35, 1555 – 1558.
[22] F. Johnson, Chem. Rev. 1968, 68, 375 –
between 4-DMAP (1) and salt II alone with chiral 4-(dialkylamino)pyridines
413.
(Scheme 6) could not be responsible for and related systems can be expected to [23] G. Opitz, A. Griesinger, Liebigs Ann.
such an extreme concentration differ- build upon our understanding and to Chem 1963, 665, 101 – 113.
ence. A plausible explanation is that the provide further information on this [24] G. Stork, A. Brizzolara, H. Landesman,
equilibrium for neutralization of the catalytic manifold. J. Szmuszkovicz, R. Terrell, J. Am.
acid generated during the reaction con- Chem. Soc. 1963, 85, 207 – 222.
tributes to this situation. Usually at least Published Online: September 17, 2004 [25] H. Mayr, B. Kempf, A. R. Ofial, Acc.
Chem. Res. 2003, 36, 66 – 77.
1 equivalent of a tertiary amine such as
[26] H. Mayr, T. Bug, M. F. Gotta, N. Hering,
Et3N is used as an auxiliary base for this B. Irrgang, B. Janker, B. Kempf, R. Loos,
purpose and under these conditions 4- A. R. Ofial, G. Remennikov, H. Schim-
[1] C. Grondal, Synlett 2003, 10, 1568 – 1569.
DMAP-catalyzed esterifications are mel, J. Am. Chem. Soc. 2001, 123, 9500 –
[2] A. Hassner in Encyclopedia of Reagents
usually faster with Ac2O than with for Organic Synthesis, Vol. 3 (Ed.: L. A. 9512.
AcCl.[17] Being slightly more basic than Paquette), Wiley, New York, 1995, [27] E. M. Cherkasova, S. V. Bogatkov, Z. P.
4-DMAP (1) (pKa  11 vs.  10) and p. 2022 – 2024. Golovina, Russ. Chem. Rev. 1977, 46,
[3] L. M. Litvinenko, A. I. Kirichenko, 246 – 263.
present in excess, the tertiary amine
Dokl. Chem. 1967, 763 – 766; Dokl. [28] P. J. Battye, E. M. Ihsan, R. B. Moodie, J.
effectively sequesters the acid (HCl or
Akad. Nauk SSSR Ser. Khim. 1967, Chem. Soc. Perkin Trans. 2 1980, 741 –
HOAc) and maintains the 4-DMAP (1) 748.
176, 97 – 100.
in the unprotonated form and available [29] M. L. Bender in Mechanisms of Homo-
[4] W. Steglich, G. HQfle, Angew. Chem.
for nucleophilic catalysis. However, 1969, 81, 1001; Angew. Chem. Int. Ed. geneous Catalysis from Protons to Pro-
when pyridine (pKa  5) is employed as Engl. 1969, 8, 981. teins (Ed.: M. L.Bender), Wiley, New
the auxiliary base, sequestration of the [5] M. Benaglia, A. Puglisi, F. Cozzi, Chem. York, 1971.
acid will be less efficient, and significant Rev. 2003, 103, 3401 – 3429. [30] For a discussion of the pKa values of
[6] R. Murugan, E. F. V. Scriven, Aldrichi- pyridines in various solvents see: L. J.
protonation of 4-DMAP (1) may occur.
mica Acta 2003, 36, 21 – 27. Chmurzynski, Heterocycl. Chem 2000,
The findings of Kattnig and Albert 37, 71 – 74.
[7] U. Ragnarsson, L. Grehn, Acc. Chem.
would then be consistent with the extent Res. 1998, 31, 494 – 501. [31] A. K. Sheinkman, S. I. Suminov, A. N.
of protonation of 4-DMAP (1) versus [8] F. V. Scriven, Chem. Soc. Rev. 1983, 12, Kost, Russ. Chem. Rev. 1973, 42, 642 –
pyridine by HOAc exceeding that by 129 – 162. 661.
HCl, or proton transfer between pyridi- [9] G. HQfle, W. Steglich, H. Vorbruggen, [32] Unimolecular elimination of the N-acyl-
ne·HOAc and 4-DMAP·HOAc being Angew. Chem. 1978, 90, 602 – 615; An- pyridinium salt to form an acylium ion is
gew. Chem. Int. Ed. Engl. 1978, 17, 569 – also conceivable but this process would
slow relative to that between pyridi-
583. be expected to be more significant for
ne·HCl and 4-DMAP·HCl.[39] the initial acyl donor (Scheme 6).
[10] A. C. Spivey, A. Maddaford, A. Red-
Kattnig and Albert went on to show grave, Org. Prep. Proced. Int. 2000, 32, [33] The intervention of a ketene pathway
that synthetically useful levels of regio- 331 – 365. during 4-(dialkylamino)pyridine-cata-
control can be exerted in 4-DMAP- [11] S. France, D. J. Guerin, S. J. Miller, T. lyzed esterification has been proposed
catalyzed polyol acetylation by the ap- Lectka, Chem. Rev. 2003, 103, 2985 – but not substantiated.[15, 17]
propriate choice of conditions. For ex- 3012. [34] M. Wakselman, E. Guibe-Jampel, Tetra-
[12] M. R. Heinrich, H. S. Klisa, H. Mayr, W. hedron Lett. 1970, 11, 1521 – 1525.
ample, the primary alcohol group at C6
Steglich, H. Zipse, Angew. Chem. 2003, [35] A. R. Butler, I. H. Robertson, J. Chem.
of octyl b-d-glucopyranoside is esteri- Soc. Perkin Trans. 2 1975, 660 – 663.
115, 4975 – 4977; Angew . Chem. Int. Ed.
fied by using AcCl–pyridine, whereas 2003, 42, 4826 – 4828. [36] Kattnig and Albert obtained values in
the secondary alcohols at C3/C4 are [13] E. Kattnig, M. Albert, Org. Lett. 2004, 6, the range 1.3–1.8 for primary kinetic
esterified by using Ac2O–pyridine 945 – 948. isotope effects (KIEs, kH/kD) for the

5440  2004 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim www.angewandte.org Angew. Chem. Int. Ed. 2004, 43, 5436 –5441
Angewandte
Chemie

reactions shown in Scheme 8. These are (Scheme 7),[9] but also from studies in cies involved in proton transfer: Z.
not consistent with a primary KIE of which isolated N-acetyl-4-(dimethylami- Dega-Szafran, M. Szafran, Heterocycles
0.81 for 4-DMAP-mediated acetylation no)pyridinium acetate but NOT the 1994, 37, 627 – 659 and Z. Dega-Szafran,
of tBuOH with Ac2O in CDCl3[37] and do corresponding chloride, tosylate, or tet- M. Gdaniec, M. Grundwald-Wyspian-
not provide clear-cut evidence in favor rafluoroborate was shown to react with ska, Z. Kosturkiewicz, J. Koput, P. Krzy-
of general base catalysis. tBuOH in CHCl3.[37] zanowski, M. Szafran, J. Mol. Struct.
[37] E. Guibe-Jampel, G. Le Corre, M. Wak- [39] Ion pair “complexes” of 4-DMAP (1) 1992, 270, 99 – 124.
selman, Tetrahedron Lett. 1975, 16, with carboxylic acids are relatively sta- [40] T. Kurahashi, T. Mizutani, J.-i. Yoshida,
1157 – 1160. ble and have been characterized by X- J. Chem. Soc. Perkin Trans. 1 1999, 465 –
[38] Evidence for this comes not only from ray diffraction. Higher aggregates (e.g. 473.
the results of Steglich and co-workers 1H+·OAc·HOAc) may be the key spe-

Angew. Chem. Int. Ed. 2004, 43, 5436 –5441 www.angewandte.org  2004 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim 5441

Vous aimerez peut-être aussi