Download Phosphine-Catalyzed Additions of Nucleophiles and Electrophiles to

Survey
yes no Was this document useful for you?
   Thank you for your participation!

* Your assessment is very important for improving the workof artificial intelligence, which forms the content of this project

Document related concepts

Physical organic chemistry wikipedia , lookup

Marcus theory wikipedia , lookup

Woodward–Hoffmann rules wikipedia , lookup

Kinetic resolution wikipedia , lookup

George S. Hammond wikipedia , lookup

Fischer–Tropsch process wikipedia , lookup

Organosulfur compounds wikipedia , lookup

Vinylcyclopropane rearrangement wikipedia , lookup

Enantioselective synthesis wikipedia , lookup

Discodermolide wikipedia , lookup

Haloalkane wikipedia , lookup

Wolff rearrangement wikipedia , lookup

Ene reaction wikipedia , lookup

Strychnine total synthesis wikipedia , lookup

Alkene wikipedia , lookup

Petasis reaction wikipedia , lookup

Stille reaction wikipedia , lookup

Aldol reaction wikipedia , lookup

Elias James Corey wikipedia , lookup

Ring-closing metathesis wikipedia , lookup

Baylis–Hillman reaction wikipedia , lookup

Nucleophilic acyl substitution wikipedia , lookup

Wolff–Kishner reduction wikipedia , lookup

1,3-Dipolar cycloaddition wikipedia , lookup

Asymmetric induction wikipedia , lookup

Hydroformylation wikipedia , lookup

Transcript
PHOSPHINE-CATALYZED ADDITIONS OF NUCLEOPHILES AND ELECTROPHILES TO
α,β–UNSATURATED CARBONYL COMPOUNDS
Reported by Michael Scott Bultman
November 4, 2004
INTRODUCTION
Organophosphorous compounds are becoming increasingly important in organic synthesis.
Phosphines serve as precursors to phosphonium ylides in the Wittig reaction,1 and as nucleophilic
triggers in the Mitsunobu2 and Staudinger3 reactions.
In these processes, the phosphine is
stoichiometrically consumed and converted into a phosphine oxide. Phosphines are also commonly used
as ligands for transition metal-catalyzed reactions, to modulate reactivity and stereocontrol.4 On the
other hand, the use of phosphines as nucleophilic catalysts for organic reactions has only gained
attention in the last ten years. First reported by Rauhut and Currier in 1963,5 phosphine catalysis has
since been reinvestigated after the phosphine ligands in some transition-metal-catalyzed reactions were
found to be better catalysts than the metal/phosphine complexes alone!6 Phosphines are well suited for
catalyzing the addition of both nucleophiles and electrophiles to electron deficient alkenes, alkynes, and
allenes. Activation of these α,β-unsaturated carbonyl systems with the phosphine enables the formation
of new bonds at the α-, β-, and γ-positions. This report will highlight these different modes of addition
to α,β-unsaturated carbonyl systems under phosphine catalysis that allow for the formation of a wide
array of products from a single class of substrates.
GENERAL REACTIVITY OF PHOSPHINES
Key characteristics required for successful nucleophilic catalysis lie in the balance of leaving
group ability, nucleophilicity, and ease of ylid formation. Increasing leaving group ability can often be
correlated with decreasing basicity. Whereas phosphines are less basic than amines (pKa values: HPEt3+
(8.7), HNEt3+ (10.7) in H2O), and therefore better leaving groups, they have strikingly different
nucleophilic profiles. Tertiary phosphines are much stronger nucleophiles than corresponding amines.
For example, the rate of the SN2 substitution of C2H5I with Et2PhP is more than 500 times faster than
that of Et2PhN.7 Greater nucleophilicity is observed with electron-donating alkyl substituents than with
aryl groups, resulting in an increased nucleophilic reactivity of trialkyl phosphines. Lastly, the ability to
stabilize an adjacent carbanion to form an ylide is an important characteristic of phosphines. This
unique combination of properties allows phosphines to act as strong nucleophiles for conjugate addition,
stabilize high energy intermediates, and still possess the ability to serve as an efficient leaving group.
Copyright © 2004 by Michael S. Bultman
49
RAUHUT-CURRIER AND MORITA-BAYLIS-HILMAN REACTIONS
Formation of new bonds at the α-position Scheme 1
O
of α,β-unsaturated carbonyl compounds using
O
2.5 mol% PPh
CH
3
O
3
phosphine catalysis was first observed by Rauhut H C
3
t-BuOH, 118 oC
H3C
and Currier5 during the course of their studies on
60%
1
2
8
the dimerization of activated olefins (Scheme 1). Later work by Morita, Baylis, and Hillman expanded
the scope of the reaction to include aldehydes as the electrophilic component. Unfortunately, long
reaction times and limited substrate dependence precluded broad application of this method in synthesis.
Scheme 2 O
O
As is the case with all reactions catalyzed by
O
R'
phosphines, the proposed catalytic cycle for the
R'
4
R'
Rauhut-Currier
PR3
3
8
H
begins
with
the
conjugate addition of phosphine 3 to enone 4 to
provide β-phosphonium enolate 5 (Scheme 2).
O
O
reaction
O
R'
Subsequent conjugate addition of 5 to another
R'
R'
7
molecule of enone 4 affords δ-phosphonium
5
PR3
PR3
enolate 6. Finally, a proton transfer provides β-
O
O
R'
H
phosphonium enolate 7, which undergoes βO
R'
6
elimination to regenerate catalyst 3 and form the
R'
PR3
4
dimerized enone product 8.
Krische9 and Roush10 have developed intramolecular variants of the Rauhut-Currier reaction,
which eliminate the problem of homo-dimerization found in the coupling of different alkenes (Scheme
3). In cases where enones of similar electrophilicities are used, no selectivity is observed and mixtures
are obtained. Instead, the use of enones with sufficiently different electrophilicity ensures the generation
of a single constitutional isomer. As shown from the product ratio the catalyst favors addition to the
(more electrophilic) α,β-unsaturated thioester. The (less electrophilic) α,β-unsaturated ester then serves
as the conjugate acceptor for the ring-closing step, strongly favoring the production of 10.9
Scheme 3
O
O
EtS
9
( )n
O
OEt
20 mol% PMe3
t-BuOH
30 oC
O
O
O
EtS
10
( )n
OEt
EtS
n = 1, 89% yield
(98:1 10:11)
11 ( )n
n = 2, 82% yield
(99:1 10:11)
OEt
50
CONJUGATE ADDITION TO ACTIVATED ALKENES AND ALKYNES
Phosphines are very effective catalysts for the addition of nucleophiles to α,β-unsaturated
Phosphines mediate the conjugate addition of nucleophiles such as alcohols,11
carbonyl compounds.
oximes,12 and carbon acids7,13 to
activated
alkenes
(Scheme
Scheme 4
O
O
4).
NuH, PR3
R'
Whereas these reactions are generally
R'
4
catalyzed by a strong alkoxide base,
12
O
the use of phosphines allows the H3C
addition to occur under much milder
O
C6H4p-NO2
EtO
13 H3C
conditions.
Nu
O
OMe
14
85%
O
EtO
N
NO2
15
CH3
CH
3
82%
85%
The mechanism for conjugate addition differs significantly from the Rauhut-Currier reaction in
that
Nu-H
Scheme 5
O
O
16
4
catalyze the addition of the nucleophile to
5
PR3
3
β-
phosphine-generated
phosphonium enolate 5 acts as a base to
R'
R'
the
enone 4 (Scheme 5).
PR3
Addition of the
phosphine 3 to enone 4, generates βO
O
R'
R'
17
PR3
19
O
O
H
Nu
18
R'
17
PR3
Nu
O
R'
20
Nu
of enone 4 to generate enolate 19.
phosphonium enolate 5, which serves as
the active catalyst by deprotonating
pronucleophile
16
to
generate
β-
phosphonium ketone 17. The resulting
activated nucleophile 18 then undergoes
R'
4
conjugate addition into another molecule
Subsequent deprotonation of phosphonium ketone 17 then
regenerates 5 and produces the β-substituted product 20.
The analogous conjugate addition of alcohols activated to terminal alkynes has also been
achieved. Inanaga described the addition of benzyl alcohol to ynones to form vinylogous benzyl esters
in excellent yields and olefin selectivities (Scheme 6).14
Carbon acids, saturated secondary and tertiary alcohols
failed to add under similar conditions.
O
These results
indicate that nucleophiles with minimal steric hinderance
are essential for successful addition.
Scheme 6
21
10 mol% PBu3
BnOH
THF, 10 min
98%
O
22
OBn
Not surprisingly, tributylphosphine is more effective than
triphenylphosphine because of its increased nucleophilicity and reduced steric hindrance.
51
Under certain conditions with specific nucleophiles attack will take place at the α-position of an
activated alkyne (Scheme 7).15 The addition at both the a- and b-carbons of activated alkynes can be
understood by the unified mechanism shown in Scheme 8. The conjugate addition of phosphine 3 to
alkyne
26
protonation
followed
with
by
pronucleophile (Nu-H) affords
phosphonium salt 28.
This
intermediate
two
possesses
Scheme 7
the
10 mol% PPh3
O
50 mol% HOAc
O
50 mol% NaOAc
N
toluene
EtO
O
105 oC
Ph
25
82%
O
O
HN
EtO
Ph
O
23
24
electrophilic centers; one at the β-position (from conjugation with the carbonyl group) and another at the
α-position (from electrostatic activation by the phosphonium ion). Under buffered acidic conditions
with weak nucleophiles, nucleophilic attack occurs at the α-position. Alternatively, attack occurs at the
β-position under neutral conditions with strong alkoxide nucleophiles. Subsequent proton transfers and
phosphine elimination afford the addition product 29 and 30 and regenerate the catalyst.
O
Scheme 8
α−addition
O
PR3
3
O Nu
O
R'
Nu-H
R'
R''
26
Nu
R'
R''
27
PR3
29
R''
R'
R''
28
O
PR3
R'
β−addition
R''
30
Nu
γ-ADDITION AND ISOMERIZATION OF ACTIVATED ALKYNES AND ALLENES
A novel mode of reactivity is available when activated alkynes bear acidic protons at the γposition. This variant was first demonstrated by Trost16 in the addition of carbon nucleophiles (Scheme
9). A contemporaneous report by Lu17 employed activated allenes as the starting materials under milder
conditions to create the same product 33. This avenue is made possible by isomerization of electron
deficient alkynes to allenes under the buffered acidic reaction conditions. In general, better yields are
obtained with activated alkynes than with the corresponding allenes. Heteroatom nucleophiles (e.g.
alcohols and imides), and intramolecular additions have also been employed.18
52
Scheme 9
O
O
MeO
CH3
31
MeO
10 mol% PPh3
50 mol% HOAc
50 mol% NaOAc
toluene
105 oC
63%
O
OMe
32
O
O
MeO
MeO
34
O
32
O
O
MeO
OMe
33
5 mol% PPh3
benzene, rt
65%
OMe
OMe
O
If hydrogens are present on the δ-position of the substrate, yet another mode of reactivity can be
accessed. For example, in the absence of a nucleophile activated alkynoate 35 will isomerize to 2,4dienoate 36 (Scheme 10). Trost,6 Lu,19a and Rychnovsky19b have all used this method synthetically for
the production of polyenes.
As shown in Scheme 10, only electron deficient alkynes undergo
isomerization under these conditions. Differentiation of activated alkynes of different electrophilicities
and the use of activated allenes have also been explored under isomerization conditions.6
Scheme 10
10-40 mol% PPh3
50 mol% AcOH
O
H3C
O
CH3
35
xylene, 60 oC, 5 h
79%
O
H3C
O
CH3
36
A mechanism proposed for these competitive processes is shown in Scheme 11. After conjugate
addition of phosphine 3 to the electron poor alkyne 37, a proton transfer forms extended β-phosphonium
enolate 39. Protonation by the pronucleophile (Nu-H) results in phosphonium salt 40, which is uniquely
electrophilic at the γ-position. If the nucleophile is sufficiently strong and conditions allow, attack
occurs at the γ-position; otherwise deprotonation at the δ-position predominates. Subsequent proton
transfers and elimination of the phosphine affords the respective products 41 and 42 with regeneration of
the catalyst.
Scheme 11
37
PR3
3
CH3
NuH R'
R'
R'
CH3
R3P
38
R3P
R'
O
O
O
O
R'
O
γ-addition
CH3
39
α Nu
β δ
R3P
40
γ
41
O
H
CH3
Nu
R'
isomerization
42
In specific cases, the product distribution can be controlled to favor isomerization or the γaddition product.
For example, use of a polar coordinating solvents such as DMSO with
triphenylphosphine as a catalyst strongly favors formation of the isomerized dienoate 45 (Scheme 12).18b
53
On the other hand, switching from a monodentate phosphine to the bidentate 1,3-bis(diphenylphosphino)propane (dppp, 46) strongly favors formation of the γ-addition product 44. The
bidentate phosphine is thought to act as a nucleophile and a general base catalyst in order to mediate
formation of γ-addition product.
Scheme 12
O
O
O
conditions
MeO
MeO
MeO
OH
O
43
PPh3, HOAc, DMSO, 110 oC
dppp, HOAc, PhCH3, 110 oC
OH
44
45
9%
97%
91%
3%
PPh2 PPh2
dppp
46
ANNULATIONS - Bifunctional Nucleophiles
The potential for addition to adjacent positions in α,β-unsaturated carbonyl compounds using
phosphine catalysis allows for annulation reactions.
Tethered bifunctional nucleophiles allow for
nonconcerted annulations. Lu has constructed seven member rings, as well as dihydrofuran systems,
with this method (Scheme 13).20 The reaction proceeds through a phosphine-catalyzed nucleophilic
addition to the α-position of the activated alkyne, followed by a separate phosphine-catalyzed conjugate
addition with the second tethered nucleophile.
Scheme 13
[3 + 2] Cycloadditions
Electron
alkenes
are
OEt
deficient
used
as
TsHN
NHTs
47
48 O
NTs
10 mol% PPh3
toluene
80 oC, 72 h
TsN
OEt
49 88% O
dipolarophiles in phosphinecatalyzed [3+2] cycloadditions to form substituted cyclopentenes from allenes.21 Poor regioselectivity
in the cycloaddition limits the utility of this process. However, alkynes are found to give better yields
and selectivities (Scheme 14).
The proposed mechanism begins with conjugate addition of the
phosphine to generate phosphonium enolate 52 followed by [3+2] cycloaddition with the activated
alkene. Subsequent proton transfers allow for the elimination of the phosphine and formation of isomers
53 and 54.
The observation that diethyl fumarate and maleate react stereospefically to afford
exclusively trans and cis cycloadducts, respectively, provides support for a concerted reaction.21
54
Scheme 14
O
O
CN
EtO
50
Me
51
10% PBu3
O
EtO
benzene
rt
CN
O
EtO
Bu3P
CN
EtO
80% yield 54
(93:7 53:54)
53
52
CN
The scope of dipolarophiles has been expanded to include imines for the synthesis of pyrrole
derivatives,22a exocyclic alkenes to form spirocycles.22b
22c
have been explored as starting materials.
In addition, traditional phosphonium ylides
Intramolecular and enantioselective variants of the [3+2]
cycloaddition have also been devised.23
[4 + 2] Cycloadditions
Stepwise [4+2] cycloadditions of α-substituted allenoates with a wide variety of imines have
been developed by Kwon.24 This cyclization allows access to highly substituted tetrahydropyridines in
excellent yields and diastereoselectivities (Scheme 15).
Scheme 15
O
O
EtO
Ph
NTs 20 mol% PBu3
Ph CH2Cl2, rt
56
57
Ph
EtO
O
NTs
Bu3P
Ph
58
O
Ph
EtO
NTs
Bu3P
Ph
EtO
NTs
Ph
59
Ph
60 98%
Substitution at the α-position of 2,3-butadienoate 56 with an alkyl group blocks the α-attack and
instead leads to γ-addition of the zwitterionic intermediate 58 into the dipolarophile 57. Two proton
transfers allow the new zwitterionic intermediate 59 to undergo 6-endo cyclization; expulsion of the
catalyst generates tetrahydropyridine 60.
CONCLUSION
Trialkyl and triaryl phosphines catalyze the formation of adducts from a variety of nucleophiles
and electrophiles with α,β-unsaturated carbonyl compounds. Although some of the reactions are limited
by substrate specificity and long reaction times, this class of reactions has become a useful tool in
organic synthesis. This is illustrated by applications in the production of polymers25 and the synthesis of
natural products.26 The versatility of this methodology relies in the ability to form a wide array of
products from a single substrate class.
Limitations result from the scope of reaction partners, and
further work is needed to expand this field. Future work is also required to elucidate the factors that
dictate product distribution and develop more effective enantioselective methods.
55
REFERENCES
(1)
(2)
(3)
(4)
(5)
(6)
(7)
(8)
(9)
(10)
(11)
(12)
(13)
(14)
(15)
(16)
(17)
(18)
(19)
(20)
(21)
(22)
(23)
(24)
(25)
(26)
Rein, T.; Pedersen, T. M. Synthesis 2002, 5, 579.
Dembinkski, R. Eur. J. Org. Chem. 2004, 13, 3130.
Gololobov, Y. G.; Kashukin, L. F. Tetrahedron. 1992, 48, 1353.
Miura, M. Angew. Chem., Int. Ed. 2004, 43, 2201.
Rauhut, M.; Currier, H. (American Cyanamide Co.). U.S. Patent 3,074,999, 1963; Chem. Abstr.
1963, 58, 11224a.
Trost, B. M.; Kazmaier, U. J. Am. Chem. Soc. 1992, 114, 7933.
White, D. A.; Baizer, M. M. Tetrahedron Lett. 1973, 37, 3597.
(a) Morita, K.; Suzuki, Z.; Hirose. H. Bull. Chem. Soc. Jpn. 1968, 2815. (b) Baylis, A. B.;
Hillman, M. E. D. German Patent 2,155,113, 1972; Chem. Abstr. 1972, 77, 34174q.
Wang, L. C.; Luis, A. L.; Agapiou, K.; Krische, M. J. J. Am. Chem. Soc. 2002, 124, 2402.
(a) Frank, S. A.; Mergott, D. J.; Roush, W. R. J. Am. Chem. Soc. 2002, 124, 2404. (b) Mergott,
D. J.; Frank, S. A.; Roush, W. R. Org. Lett. 2002, 4, 3157.
Stewart, I. C.; Bergman, R. G.; Toste, F. D. J. Am. Chem. Soc. 2003, 125, 8696.
Bhuniya, D.; Mohan, S.; Narayanan, S. Synthesis 2003, 1018.
Gomez-Bengoa, E.; Cuerva, J. M.; Mateo, C.; Echavarren, M. J. Am. Chem. Soc. 1996, 118,
8553.
Inanaga, J.; Baba, Y.; Hanamoto, T. Chem. Lett. 1993, 241.
Trost, B. M.; Drake, G. R. J. Am. Chem. Soc. 1997, 119, 7595.
Trost, B.; Li, C. J. J. Am. Chem. Soc. 1994, 116, 3167.
Zhang, C.; Lu, X. Synlett 1995, 645.
(a) Trost, B. M.; Dake, G. R. J. Org. Chem. 1997, 62, 5670. (b) Trost. B. M.; Li, C. J. J. Am.
Chem. Soc. 1994, 116, 10819.
(a) Gou, C.; Lu, X. J. Chem. Soc. Perkin Trans. 1 1993, 1921. (b) Rychnovsky, S. D.; Kim, J. J.
Org. Chem. 1994, 59, 2659.
Lu, C.; Lu, X. Org. Lett. 2002, 4, 4677.
Zhang, C.; Lu, X. J. Org. Chem. 1995, 60, 2906.
(a) Xu, Z.; Lu, X. Tetrahedron Lett. 1997, 38, 3461. (b) Du, Y.; Lu, X.; Yu, Y. J. Org. Chem.
2002, 67, 8901. (c) Du, Y.; Lu, X.; Zhang, C. Angew. Chem., Int. Ed. 2003, 42, 1035.
(a) Wang, J. C.; Krische, M. J. Angew. Chem., Int. Ed . 2003, 42, 5855. (b) Guoxin, Z.; Chen,
Z.; Qiongzhong, J.; Dengming, X.; Cao, P.; Zhang, X. J. Am. Chem. Soc. 1997, 119, 3836.
Zhu, X. F.; Lan, J.; Kwon, O. J. Am. Chem. Soc. 2003, 125, 4716.
Kuroda, H.; Tomita, I.; Endo, T. Macromolecules 1995, 28, 433.
(a) Agapiou, K.; Krische, M. J. Org. Lett. 2003, 5, 1737. (b) Du, Y.; Lu, X. J. Org. Chem. 2003,
68, 6463. (c) Methot, J. L.; Roush, W. R. Adv. Synth. Cat. 2004, 346, 1035.
56