Synthetic applications of enantioselective organotransition-metal

Dec 8, 1989 - plexes to effect the kinetic resolution of racemic allylic alcohols (eq 1). OH .... TABLE 6. Asymmetric Catalytic Hydroformylation42. 0...
0 downloads 0 Views 2MB Size
Chemical Reviews Volume 89, Number 8

December 1989

Synthetic Applications of Enantioselective Organotransition-Metal-Mediated Reactions SHERI L. BLYSTONE Oak Rldge Naibnal Laboretory, P.O. Box 2008, Oak Rldge, Tennessee 378314022

Received March 15, 1988 (Revissd MatlUS@t Recalved A@ 25, 1989)

contents I. Introduction 11. Asymmetric Homogeneous Catalysis A. Hydrogenations and Hydrosilylations B. Asymmetric Coupling Reactions 111. Asymmetric Alkylations Involving the Use of Chirai Auxiliaries A. Palladium E. Copper and Zinc C. Gold IV. Metal-Centered Asymmetry A. Chirai Rhenium Complexes B. Chirai Iron Complexes C. Chirai Molytdenum Complexes V. Chiral Olefin-Metal Templates VI. Use of Chiral Nucleophiles VII. Concluding Remarks

1663 1664 1664 1666 1667 1667 1670 1670 1671 1671 1671 1673 1674 1675 1676

I . Introductlon The field of organometallic chemistry has, in recent years, developed into a large and diversified discipline and has been the subject of many texts.' The application of organotransition metals to organic synthesis has been a very active area of this discipline, particularly the use of iron,= molybdenum,*'2 mangane~e,'>~5 chromium,'+'* cobalt," and palladium2'-" complexes to effect the regio- and/or stereospecific construction of organic molecules. The application of these procedures toward enantioselective organic synthesis has become an ever growing concern. This review is intended to highlight the achievements that have been accomplished in this field thus far. It is concerned with those transitionmetal reactions in which a new bond (i.e., carbon-carbon, carbon-hydrogen, or carbon-halogen) is formed and does not deal with such areas as transitionmetal-mediated asymmetric isomerizations.rJ Asymmetric homogeneous catalysis (primarily hydrogenations) is a large part of thii field with extensive coverage in the literature.3O It is included again here primarily for its importance to the field as well as to introduce 0009-2665/89/0789-1663$06.50/0

S M L. ElvJtone received her B.S. w e e from Gannon U n k d y in 1983. She then attended Case Western Resew Univerlty in Cleveland. OH, where in 1988 she received a Ph.D. degree in organic chemistry under Professor Anthony J. Peersm Cwrently she is working wfih Dr. F. F. Knapp, Jr., in the Nuclear Medicine group at the Oak Ridge National Laboratory under a Department of Energy Postgraduate Fellowship. While attending CWRU. she received ! hStandard oil of Ohio claduate F&owship h 1985. Her research interests include the development of new methodology for the construction of organic molecules of medical or biological

interest

conceptsvital to the other topics discussed. In addition, some newer references in this area are included. This review covers references from the seventies through early 1988, focusing primarily on the group 6-10 transition metals and their application to asymmetric organic synthesis. It is intended to be a reference point for those organometallic methodologies that have proven to be, or are potentially, synthetically valuable. The field of asymmetric oxidations, particularly the Sharpless epoxidations, suitable for a separate review, is not specifically addressed. The studies in the area covered by this review can be grouped into three general categories. The fmt involves diastereoselective reactions involving a complex in which a chiral auxiliary has been incorporated. The chual auxiliary in these cases may be either the organic 0 1989 American Chemical SocW

1664 Chemical Reviews, 1989, Vol. 89, No. 8

Blystone

TABLE 1. Asymmetric Catalytic Hydrogenation of Olefins

RH H2 R,

Rh i

3 Prophos

MexMe

Ph,P

PPh,

G CI H 2 P P h , COZtB" 5 BPPM

4

Chiraehos

R i Rz

Ra

L'

%e.e. (config)

L'

X

Ref.

hPPh2 PPh,

6 Norehos

q

PPh,

aYPPh2 9 BlNAP

7

89

(S)

(b)

100

(S)

(C)

87

(9

(C)

H

H C Q H

N-IcoMe

H

Ph CQH

NHWh

9 (R)

Ph

H

NHCOPh

9 (S)

H

Ph C Q b

WCWt

9

(S)

93 (A)

(C)

H

Ph CQH

"?e

9

(9

84 (R)

(C)

H

H

NHCOPh

9 (SI

98 (R)

(C)

(.)

DIPHOL Me

,p Ph,PO

A

NPPh,

I

Me 10 A'aNOP

CQH

,ppr

nNPPh,

PhzPO

I

Ph,PO

NPPh2

I

Me 11 VaiNOP

Me 12

EFKE

CQH

oPracejua, G.; Pracejua, H. J. Mol. Catal. 1984, 24, 227. bDobler, C.; Kreuzfeld, H.-J.; Pracejus, H. J. Organomet. Chem. 1988, 344, 89. 'Miyashita, A.; Takaya, H.;Souchi, T.; Noyori, R. Tetrahedron 1984,40, 1245.

H%,.

QCH,OPPh, I

[(Diene) M (diphos)]'

PPh,

I

13 ProNOP

Figure 1. Some chiral phosphine ligands.

ligand of interest itself or other ligands on the metal. In the latter case, this is generally a chiral phosphine. Alternatively, the metal itself may be considered chiral by virtue of having different ligands attached. The second category involves the enantioselective addition of a nucleophile to an organometallic complex where the olefii, by having the metal complexed to one face of the n system, becomes the center of chirality. Finally, in a new extension of this area, a chiral nucleophile can be added diastereoselectively to an achiral organometallic complex.

I I . Asymmetrlc Homogeneous Catalysls

[H M (diphos) (RH)]'

[M (diphos) alkene]+

dtaslereomers

diarlereomsrs

[H2 M (diphos) alkene]+

H,

Figure 2. Catalytic cycle for hydrogenations."

i. (eq)

ph2Y

(eq)

....

h-

A. Hydrogenations and Hydrosllylations One of the most successful applications of transition metals in asymmetric organic synthesis has been in hydrohomogeneous catalysis-hydrogenations, silylations, hydroformylations, and coupling reactions. In 1968, the first instance of enantioselectivitywas seen during the hydrogenation of a prochiral olefin with a Wilkinson-type rhodium catalyst that incorporated a chiral phosphorus monodentate ligand.31 It was soon discovered, however, that the degree of optical induction could be greatly enhanced by the use of chiral phosphine bidentate ligands.32 With these types of ligands, some of which are shown in Figure 1, hydrogenated

i?

dt6slereOmei$

Ph2

(ax)

Figure 3. Conformation of complex intermediate in catalytic hydrogenation.=

products have been obtained in >95% optical purity. Some recent examples of catalytic asymmetric hydrogenation of olefins are shown in Table I. Bidentate ligands were found to impart a conformational rigidity to the intermediate catalyst-substrate complexes (Figure 2). This appears to be a major factor involved in the increased selectivity seen in these case ~ . Diphosphines ~ ~ , ~ DIOP (1) and BPPM (5) form seven-membered chelate rings whereas DIPAMP (2)) Prophos (3))Chiraphos (4), and Norphos (6) form more

Chemical Reviews, 1989, Vol. 89, No. 8 1665

Organotransltlon-MetakMediated Reactions

TABLE 2. Asymmetric Catalytic Hydrogenation of Ketones"

H

Ph

H

H

C0,Et

RCO :lH P :h

H %

Y

R R Ho c,R z(z-l)

Figure 4. Substrates most successful in asymmetric catalytic hydrogenation.u

M

CHZNME~

S

'PI Ph

CHzNMe2

S S

CHZNMez

M M

M M Me

A Figure 5. Second avenue of chelation in h y d r o g e n a t i ~ n . ~ ~

rigid, puckered, five-membered-ring chelates (Figure 3). With the smaller ring size, substituents on the phosphine tend to orient themselves in a pseudoequatorial position. It was thought that the fixed fivemembered-ring conformation, by reducing the number of possible diastereomeric interactions with the enantiotopic faces of the substrate (Figure 3), would more readily allow one conformation to predominate, thereby leading to a single enantiomer of the hydrogenated product being preferentially formed.33 Halpernss has in fact shown that, while indeed a major diastereomeric catalyst-substrate complex is formed, the major product produced, and therefore the selectivity of the reaction, are derived from the minor diastereomer. Apparently, the minor diastereomer has a greatly enhanced reactivity over the more stable major component, and this reactivity then determines the configuration of the hydrogenated product. Noyori3' has recently employed the difference in reactivity of the diastereomeric catalyst-substrate complexes to effect the kinetic resolution of racemic allylic alcohols (eq 1).

-1( + CO,Me

(9 >99O/& e

&COzMe

85 100

95 (9 92 (R)

97 100 100 100

83 ( R ) 98 ( R ) >99 (R) 96 (S)

R

42 100

93 (R) 100 (R,R)

S

100

c4b

R

CHZCOzEt CHzCONM+

R

W

S R

CHzCOSEt CYCH(CH3)COMe

99

(S,S)

TABLE 3. Asymmetric Catalytic Hydrogenation of &Diketones& Rul 9

0

0

OH

OH

an,,

Ri

Rz

ant1:syn:mono

OH

OH

OH

W

0

mono

%e.e. (anti)

_...__.________._________

Me

Me

Me

Et

99.1.94.6:-

>99 94

hk

'P r

97:3:-

96

Me Et

'BU

Et

91:9.98:2'-

96 96

Me

Ph

89:9:2

99

TABLE 4. Asymmetric Catalytic Hydrogenation of a&Unsaturated Ketones Substrate

Product

H

M

OH

9 RuiOCOMDl~

%e.e. (COnllg)

96 (R) 95 (S)

CHzCHzCH

C

b

cope

yield

72 83

R R

b b

I

5

config. of ligand

E

M

e

D

H

Catalyst

%yield

%e.e. ref.

20

62 (R) (a)

H

H

HRUCI(TBPC)ZC 40

22 (R) (a)

H

H

36

26 (S) (a)

(1)

(2R.3R) 3 7 % ~e

The role of the substrate in determining the stereochemical outcome is not to be discounted. In general, olefins that produce the highest degree of enantiomeric excess (ee) upon hydrogenation are those that are capable of a secondary interaction with the metal center in addition to the primary coordination of the These general types are shown in Figure 4. It should be noted that these olefins all possess an appendant oxygen functionality that can coordinate to the "ligand-deficient" metal center and help to direct the reaction (Figure 5). This secondary interaction helps to increase the rigidity of the catalyst-substrate system, which, as discussed earlier, helps to increase the observed enantioselectivity. Olefins that are incapable of this secondary interaction generally result in lower asymmetric induction. Catalytic asymmetric homogeneous hydrogenation has been extensively studied and is the subject of many

HRuCI(TBPC);

26

46.5

(b)

RuZCI,(DIOP),

35

40

(b)

a LeMaux, P.; Simonneaux, G. J. Organomet. Chem. 1987,327, 269. LeMaux, P.; Massonneau, V.; Simonneaux, G. Tetrahedron 1988, 44, 1409. (-)-trans-l,2-Bis[(dipheny1phosphino)methyl)l-

cyc1obutan.e.

reviews.*36a It is still a popular field of research today as new chiral ligands, catalyst systems, and substrates are explored for their effect on the enantioselectivity of the reaction. For example, Noyori and co-workersm have applied a chiral ruthenium-BINAP (9) catalyst for the hydrogenation of ketones with excellent results (Table 2).

1666 Chemical Reviews, 1989, Voi. 89, No. 8 R 2 c H a H F - o s ' r

[L, Rhl

Blystone R-M

T S l H

M [L, Rh (R,C-OSi) H]

L

=

Mg,

Catalyst

[Ln Rh (St)HI

R'-X

+

=

catalyst

Zn,AI, Zr, Sn,B, Hg, Lt Pd or Ni

h

2

c

LnNi

a

+

chiral ligand

' 7-i R-X

[L, Rh (Si) (R,C=O) H]

R-R

\t

'I7

.

X LnNi'

.

M-X

Figure 6. Catalytic cycle for hydroeilylation.%

R-R'

'R

R-M

Figure 7. Asymmetric catalytic cross-c0upling.4~ TABLE 5. Asymmetric Catalytic Hydrosilylation of Ketones

TABLE 6. Asymmetric Catalytic Hydroformylati~n~~

-

L'

PW

Ph-

co

CHO

H2

Ri

Rz

L'

-/&.e. (config.)

Ref.

._._._...___

Me

Ph

Me

Ph

L'

%e.e. (config.)

10 11

23 9

12 13 (R R)-1

Me

%yield

(S)

1 4 2 (S)

95 96

36 3 (S) 48 1 ( S ) ' 8 (SI

92 5 97 85 5

Ph

Me

Ph

Me

Ph

Me

Ph

Me

Ph

Me

'Bu

Me

'Bu

Me

'Bu

Me

PhCH2

Me

PhCH2

Me

PhCHzCH,

'Mokhleaur, A. F. M.; Wild, S. B. J . Mol. Catal. 1987, 39, 155. bBrunner, H.; Backer, R.; Riepl, G. Organometallics 1984,s. 1354.

Additionally, 1,3-diketones40(Table 3) and a,@-unsaturated ketones (Table 4) have been examined. As with hydrogenations, asymmetric catalytic hydrosilylation of prochiral carbonyl compounds, imines, and olefins is promoted by transition-metal catalysts possessing chiral phosphine ligands. A reasonable catalytic cycle for the reaction is shown in Figure 6.34 Generally, the asymmetric induction involved in hydrosilylation is somewhat lower than that achieved in hydrogenation. Interestingly, as opposed to hydrogenation, both uni- and bidentate phosphine ligands appear to work equally well. The nature of the silane, however, has a marked effect on the selectivity. For example, hydrosilylation of phenyl tert-butyl ketone with dimethylphenylsilane results in the stereoselective formation of the S enantiomer of the corresponding alcohol in 62% ee, whereas the use of trimethylsilane results ~~ hyin a 28% ee of the R e n a n t i ~ m e r .Asymmetric drosilylation of a-keto esters and amides and y-keto esters gives somewhat better results, up to 7 5 8 5 % ee.41 As with hydrogenation, selectivity appears to be increased when a possibility for a secondary interaction

of an ancillary carbonyl functionality with the metal center exists. Recent examples are shown in Table 5. In a similar reaction, transiton-metal catalysts will also promote an asymmetric one-carbon homologation of a prochiral olefin in the presence of an alcohol and carbon monoxide to give a saturated ester in low to moderate enantiomeric excess.41 An example, using a palladium catalyst, is shown in eq 2. Catalytic asymmetric hydroformylation gives chiral aldehydes in moderate enantiomeric excess (Table 6).42

A

Ph

lpdl

17-

PhL C O , t ,

(2)

CO 1238 a m )

69%e e

(S)

B. Asymmetric Coupllng Reactions

Chiral transition-metal catalysts have also been used to mediate asymmetric cross-coupling reactions of alkenyl or aryl halides or allylic compounds with some metalated species. Generally, catalysts of nickel or palladium bearing chiral phosphine ligands are employed.43 The actual mechanism of this coupling is not known; however, an unsymmetrical diorganometallic complex, L,Mn(R)(R'), is most likely the key intermediate (Figure 7). Chiral phosphine ligands such as have been previously discussed (Figure 1) are also employed in crosscoupling reactions. In addition, a large number of ferrocene and amino acid derivatives have also been studied (Figure 8).43 The first examples of this procedure gave disappointing results ( 4 7 % ee), but more recent experiments have yielded higher selectivities (Table 7). It has been noted from studies involving the coupling of (1-phenylethy1)magnesium chloride with vinyl bromide (entry 2, Table 7) that higher stereoselectivities are obtained with those ligands possessing an amine functionality and that this stereoselectivity is strongly affected by changing the steric bulk of this amine in the phosphinoferrocenyl ligands 20a-g. The optical purity is most likely determined during the transmetalation of the alkyl group from the Grignard reagent to the catalyst (Figure 7). The amine grouping is then able

Chemical Reviews, 1989, Vol. 89, No. 8 1667

Organotrans%ion-MetaI-MedlatedReactions

aPPh2 wCHZNMe2 aPPh2 Fa

Fa

a P P h , (S)-(R)-PPFA

(R)-(R)-PPFA

14

&Et

(S)-FcPN

(R)-PPEF

15

16

14

L

Figure 9. Intermediatein asymmetric catalytic cross-coupling."

J

X

-

OAc. OPh. OH, NR,. S0,Ph. epoxide, halide

Figure 10. Palladium-catalyzed allylic alkylation.

23

Figure 8. Chiral phosphine ligands used in cross-coupling and hydr~formylation.'~ TABLE 7. Asymmetric Catalytic Cross-Coupling

L

24

Figure 11. Alkylation of palladium chloride dimers?'* R,

X,

X,

L'

M

%ea. Ret.

ENTRY

R,

R,

1

Me

Et

Ph

Br

CI

NI

1

17

44

2

Me

Ph

CH-CH,

CI

Br

NI

14

68 (S)

43

Me

Ph

CH-CH,

CI

Br

NI

20a

35 (R)

Me

Ph

CH-CH,

CI

Br

NI

20c

15 (S)

Me

Ph

CH=CH,

CI

Er

NI

20d

62 (R)

Me

Ph

CH=CH,

CI

Er

NI

201

17 (R)

Me

Ph

CH.CH,

CI

Br

NI

20g

65 (R)

Me

Ph

CH=CH,

CI

Er

NI

21c

71 (S)

Me

Ph

CH-CH,

CI

Br

NI

211

81 (S)

Me

Ph

CH-CH,

CI

Br

NI

21h

94 (R)

Other applications of transition-metal catalysts containing a chiral ligand include asymmetric codimerization of olefins (eq 3)," an asymmetric version of the Felkin reaction (eq 4 and 5),48 and asymmetric dihydroxylation of olefins (Table 8).M)

70°/a.e.

14.9%e.e.

3

Me

p-MeC8H,

CH.CH,

Br

Er

NI

218

83 (S)

45

4

Me

"C~H,J

CHdb

CI

Br

NI

14

37 (S)

46

5

Ph

SiMe:,

CH-C.CH,

Br

Sr

Pd

95 (R)

47

6

Ns

Ph

CI

Br

Pd

15*/&.e.

tnmPhCH-CH

OPPh, 13

I I I . Asymmetric Alkyiatlons Involving the Use of Chiral AuxlHaries

48

(s)

A. Palladlum

NMePPh,

to coordinate to the magnesium, forming an intermediate or transition state such as is shown in Figure 9.43 This is analogous to the secondary coordination of appendant carbonyl functionalities of the substrate and their effect on the observed stereoselectivity during hydrogenations and hydrosilylations.

One of the most useful reactions developed in the organometallic field is the palladium(0)-catalyzed alkylation of allylic acetates, alcohols, halides, amines, etc. to give a new allylic species (Figure Complementary stoichiometric reactions involving (7dlyl)palladium dimers are also known where alkylation of 25 will occur in the presence of added phos-

1668 Chemical Reviews, 1989, Vol. 89, No. 8

3,C:Et

l#gand

/-++I -., Me

NaCHICOIE1)2

A -

\

Me

Me

Blystone

26

dC02H

Me 27

I

\

H

Me

28

%e.e.

Ligand

L

L

Figure 13. Paths of nucleophile addition to meso palladium complexes."

..--.__.__.___ ~

(+)-DIOP

12-22

(t)-ACMP

18-24

(-)-Sparteine

20

(-1-DMIP

2

H 29

U

Figure 12. Asymmetric alkylation of a palladium chloride dimer.62

V

CHIRAPHOS

DlOPiBlNAP

TABLE 8. Oemium-Catalyzed Asymmetric Di hydroxylationM OSO,

(Calj

R3

L'

Olefin

0

%e.e. (config)

BINAPO

Figure 14. "Chiral pocket" concept for chelating ligands on allyl-palladium complex."

Ph-

R'

Ph

=

p-chlorobenzoyl

p.g4YH '&-

d

R,R diastereomer of 31 following the desulfonylation of the alkylation product (eq 6).

A

P

h

w

b , A c -

A

30

Ph

h

w

O

A

\/\\/\ Ph-Ph

racemic

(+]-I

I

Pd L' 'L 30a

I

S0,Ph

INmg

(6)

A

cHex-

P

I

PdlOl

B A

C0,Me

C0,Me

c

CO,Me I

A A

A

62% (R.R) 38% (S,S) 31

B

In this case, a symmetrical palladium complex intermediate, 30a, is formed with achiral ligands. The use of chiral ligands, then, causes both allylic termini to become diastereotopic. Preferred attack of a nucleophile at one of these termini then results in asymmetric phines (Figure 11). These phosphine ligands presuminduction. Alternatively, this induction may be due to ably displace chlorine, generating a cationic allyl-pala preference for the formation of a more stable interladium complex (24) which is subsequently alkylated.51a mediate olefin-palladium complex (Figure 10). In recent years, this chemistry has been extended into T r ~ s has t ~ suggested ~ a kinetic approach to the the field of asymmetric organic synthesis. Trost requestion in which the chud ligands of a meso palladium ported in 197362the asymmetric addition of sodium dimethyl malonate to (syn,syn-l,3-dimethyl-a-allyl)- complex such as 30a form a local asymmetric environment, a "chiral pocket", which then directs the nupalladium chloride dimer (26) in the presence of a vacleophile preferentially to one end of the allyl moiety riety of chiral phosphine ligands (Figure 12). By fur(Figure 13). The larger the ring size formed in the ther chemical transformation to compounds of known bidentate phosphine-palladium complex, the more the rotation, 28 or 29, he was able to determine the amount "arms" of the ligand are forced around the allyl moiety of asymmetric induction in the original addition adduct and the incoming nucleophile, and the higher the ob27. served selectivity (Figure 14)." Indeed, alkylation of Better results have been achieved in analogous cat32 with bis(phenylsulfony1)methane using a variety of alytic reactions. T r ~ sfirst t ~ reported ~ that the use of ligands that give different chelate ring sizes evidenced a chiral phosphine ligand in the palladium-catalyzed alkylation of cis-3-acetoxy-5-carbomethoxycyclohexane this general trend (Table 9). Of particular note, the introduction of meta substitution on the phosphine aryl (30) resulted in a 24% diastereomeric excess (de) of the A

Chemical Reviews, 1989, Voi. 89, No. 8 188Q

Orgnotransftion-MetaCMediatedReactions

TABLE 9. Asymmetric Palladium-Catalyzed Alkylation of Lactone" 0

C02Me 33a

I

32 racemic

S02Ph

CHELATE RING SIZE

L'

33

%d.e.

7

(+)-I

16

82

7

(-)-9

31

92

9

(-)-BINAPO

36

66

69

82

9

L'

g..; :

NaCHIOCYllel,

C02Me

NaCHiC02MB)WPh) Ph

TPh;::;::;;:z:;: phY=+h NaCHIQ M e h

NUC

N.CH(S0gPh12

NaCHICOMeh

%.e.

%yield

37 16

Camphos

37

Camphos

39

1

46

CDCI, CMF

1.7 1.6

M

1.7

CDCI,

1 .8

CDCI, DVlF

1 1

CDCI,

CMF

6 75

53 53 53

CDCI, CMF THF

5.7 6 3.7

53 53

CDCI,

4 6

50 68 55 66 90

iR) (Si IS)

(SI (SI

Yc6H4::co*

81 75

87 85 97

54 54 54 54 (b)

92

ibl

THF

(&Ph

L?NAC4M]L''

[

;;pN;C02MJR..

-Nut

1 6

39

(0)

45

IC)

t

62

66

84

3.1 2.7

CDCI,

5 7.4

Ph

IC)

Ph

OAfter decarboxylation or desulfonylation. Hayashi, T.; Yamamoto, A.; Hagihara, T.; Ito, Y. Tetrahedron Lett. 1986, 27, 191. Hayashi, T.; Yamamoto, A.; Ito, Y. Chem. Lett. 1987,177. CGenet, J.-P.; Juge, S.; Montez, J. R.; Gaudin, J.-M. J. Chem. SOC.,Chem. Commun. 1988,718.

rings, by nature of their propeller-like arrangement, creates the most specific "chiral pocket", resulting in quite high diastereomeric excess. Alkylation of allylic acetates has also been examined (Table 10). In cases where unsymmetrical allyl-palladium complexes are formed as intermediates (i.e., entry 1, Table lo), equal amounts of complexes 33 and 34 should be formed regardless of whether the ligands are chiral if one assumes that the oxidative addition of Pd(0) to the allylic acetate is stereospecific. It is known that exo attack of a nucleophile is stereospecific and therefore, since 33 and 34 are formed in equal amounts, one would expect that 33a and 34a would also be formed in equal amounts (Figure 15). Bosnich and Mackenzies6 have suggested that the asymmetric induction that is in fact observed in these cases depends upon an equilibrium established between the two diastereomeric wallyl-palladium intermediates. They have measured this equilibrium by 31PNMR using (Sa-Chiraphos (4) as the ligand. The magnitude of this equilibrium constant could then be a measure

86

62

CDCI, CMF

M

NYC

-0Ac

%.e..

Ref

t w 9 BlMW BINAPO BINAPO A

Equil. Ratio

1 1

TABLE 10. Asymmetric Palladium-Catalyzed Alkylation of Allylic Acetates Produel

Solvent

CDCI, CMF

TMS

NucICOphib

34a

TABLE 11. Equilibrium Studies of (S,S)-4-Palladium Complexes of Various Allylic Fragments&

%yield

Allyl Ligand

Ammc

34

Figure 15. Modes of nucleophile addition to diastereomeric allyl-palladium complex intermediatmM

J3

Ph

4

CDCI,

W

CDCI,

W

A b+hP

CDCi, W

1.5 1.6

1.3 1.2

14 12 12

Of Droduct after alkylation using dimethyl sodiomalonate.

of any asymmetric induction involved in the reaction. A study of this equilibrium with a variety of allyls is shown in Table 11. It should be noted that the observed 90 ee's from subsequent alkylation do not correspond directly with the measured diastereomeric ratio of the two intermediate palladium complexes, indicating that the asymmetric induction involves more than simple thermodynamic control.

1670 Chemical Reviews, 1989, Voi. 89, No. 8

Blystone

TABLE 12. Cuprate Addition to ad-Unsaturated Carbonyls in the Presence of Asymmetric Chelating Ligands Substrate

Cupraie

%e.e. (con!lg)

L'

%Yield

Re!.

TABLE 13. Alkylation of Carbonyl Compounds with Alkylzinc Reagents in the Presence of Chiral Ligands Subilrai@

Zinc Rgl.

L'

Product

XVleld

%.e.

"0UlC"LI

$0

H

Me

P

h

R

P

t;;yI2 O H

98

91 IS) 51 (SI 65 is)

loo

71 (R)

100

70 (R) 62 IR)

92

IRI 45 (RI

72

12 [R)

53

44 (R)

78

40

+

h

MeHph

Ref.

92 (Si

100 91

97

75

MeCuLnl

B

'Soai, K.; Ookawa, A.; Ogawa, K.; Kaba, T. J. Chem. SOC., Chem. Commun. 1987, 467. *Soai, K.; Yokoyama, S.; Hayasaka, T.; Ebihara, K. J. Org.Chem. 1988,53, 4148.

MeCuLiL

TABLE 14. Gold-Catalyzed Asymmetric Aldol Condensations of a-Isooyano Carboxylates*

~0"C"LIL'

*

CNcYCqMs

7J O

; \N p

+

O

w

e

trans NR,

R

cis

1rsns:cis

% yield

e

%e.& (trans)

______-_________________________________-----_NZO

Leyendecker, F.; Jeeser, F.; Ruhland, B. Tetrahedron Lett. 1981,22,3601. *Langer,W.;Seebach, D. Helv. Chim. Acta 1979, 62,1710. 'Corey, E. J.; Naef, R.; Hannon, F. J. J. Am. Chem. SOC. 1986, 108, 7114. dDieter, R. K.; Tokles, M. J. Am. Chem. SOC. 1987,109,2040. eLeyendecker,F.; Laucher, D. Tetrahedron Lett. 1983, 24, 3517. fImamoto, T.; Mukaiyama, T. Chem. Lett. 1980, 45.

Other applications of asymmetric palladium catalysts include cyclizations (eq 7),56allylation of l,&diketones (eq 8),67and intramolecular cyclization of dicarbonates (eq 9)58and dicarbamates (eq 0

0

89 (45, 5R)

100

99 : 1

92 (4S, 5R)

Ph

93

95 : 5

95 (4s. 5R)

86

95 : 5

96 (45, 5R)

gocHo

TABLE 15. Gold-Catalyzed Intramolecular Asymmetric Aldol Condensations of a-Isocyano Carboxylateswb C W M e

I

H 48%ee ( R i

89 : 11

'Pr

R

0

99

AYI

L'

NRz NMe,

r

O

&NI"NR2

L' .=

&p;z2

w

%yield

99

Me

100

Et 'P r Ph

89 99 75

%e.e. (config)

52

(s)

64 ( S ) 70 (SI 71 (SI 67

B. Copper and Zinc UP

Chiral ligands have also been employed to effect the asymmetric Michael addition of organocuprates to a,0-unsatured carbonyl compounds (Table 12). Similarly, alkylzinc reagents have been utilized for asymmetric alkylation in the presence of chiral ligands (Table 13).

lo 73%e e

C. Gold 0

up 10

73% e.e. wilh

Hayashi and Ito have recently employed the phosphinoferrocenyl chiral ligands used in catalytic crosscoupling reactions to effect asymmetric gold-catalyzed aldol reactions of a-isocyano carboxylates (Tables 14 and 15).@' This methodology has been employed for the synthesis of threo- and erythro-sphingosines.61

Chemical Reviews, 1989, Vol. 89, No. 8 1671

F'

-

0 Clr,Fe Ph,P'

0-Menthyl

CD

Ph'

CP 0 Ce,. ;e R'Ph,P'

yH

R'

Me (R,S)-37

-

(S)-2-methylbutyl

Figure 17. Asymmetric cyclopropanation using chiral iron c0mp1exes.B~

Gladysz has been primarily interested in determining the mechanistic aspects of stereospecific hydride abstraction and various asymmetric rhenium to carbon inductions that occur in this system.@ Some interesting synthetic applications of his mechanistic work involving CP T deuterium include the development of a synthon for a I 0' I D-C-H (R) chiral, pyramidal methyl carbenium ion (Figure 16LS7 0 N\'X'Ph, T' RI Additionally, chiral rhenium-ketone complexes can H be used to give optically active alcohols in high enanFigure 16. Chiral rhenium complexes as chiral methyl synthon~.~ tiomeric excess (eq 13).68 I V. Metabcentered Asymmetry 0 TP FP CD

T

1)

21

A. Chlrai Rhenium Complexes

7:

M

N"Re#, PPh,

In this section we will be dealing with stoichiometric organometallic reactions where the metal itself is a chiral center. These types of complexes can be resolved analogously to normal organic compounds into optically enriched (or pure) material. For instance, Gladysz6, synthesized optically active CpRe(N0)(PPh,)(X) complexes to allow methodology developed on racemic compounds62to be utilized in an asymmetric fashion. These complexes can then be transformed into optically active CpRe(N0)(PPh,) (CH,), [CpRe(N0) (PPh,) (= CHJ]+PF;, etc. For example, reaction of optically pure 35 with a chiral nitrile proceeds with overall retention of configuration at rhenium (eq 11).64Alkylation of alkylidene and acetylide complexes is also readily observed (eq

(+)-(S)-35

CD

CO

A

"'y:

ON"R$,pPh,

CP

;,do

Mei,~,..(~

(+W)

(+)-(R.S)

1"

>98% 0.0.

(13) CP

I

+

, Y O H M i R

-

N"y:#pph, O f 0 c F3

R Et. 95% 8.8. . I

R

Ph, ~ 9 9 % 8.8.

B. Chlral Iron Complexes Chiral-at-iron complexes are one of the most active areas of study in the asymmetric organometallic field. Generally, chiral Fp* complexes are employed [Fp* = carbonyl(q5-cyclopentadienyl)(tripheny1phosphine)ironl. Chiral iron-alkyl complexes (36) and chiral ironcarbene complexes (37) have been used to effect asymmetric cyclopropanation of olefins (Figure 17).69 Alkylation of halo compounds 38 and olefin derivatives 39 can proceed with a significant degree of asymmetric induction, especially when the steric bulk of the iron center is increased by using tri( 0-biphenyl) phosphite (OBP) rather than triphenylphosphine (Table 16).70 Chiral iron-acyl complexes such as 40 have been synthesized by Brunner'l from the corresponding menthyl esters by the addition of methyllithium (eq 14). CP Ph3Pi,.&e

CP

0 N"?"PPh,

H

__c

0 C'

'C02Menthyl

CP

0

octl*.ie
lo0 1 1 5.4 75 1 111.4 82:1 >lo0 1 111.0 3.7 : 1 20 ' 1 1-12 >IO0 1

RCHO

73 74' 73 73 73 73 74' 73 73 74' 73 74a

MecH3 hW2-0 ECHO ECHO 'PrCHO 'PrCHO 'BUCHO 'BUCHO

Ftcw

OThis disparity may be a function of the excess of aluminum reeaent emoloved bv Davies.

by Davies. These complexes can be deprotonated to form enolate 41 and reacted with a variety of electrop h i l e ~ . ~In~ the first cases, with a simple lithium counterion, aldol condensations with aldehydes proceeded in high yield, but with little or no selectivity. It was soon discovered by both groups, however, that changing the counterion of 41 produces significant changes in stereoselectivity (Table 17). Liebeskind also noticed this effect of counterion during the addition of imines to 41 (Table 18). D a ~ i e has s ~ ~employed the remarkable stereoselectivity of these reactions in the synthesis of diastereomerically pure l-hydoxypyrrolizidin-3-onesafter a single recrystallization (eq 15).

!R.S) (15)

The enolate derived from Fp*-COCH2CH3(42) has also been found to undergo highly selective aldol con-

M

AIEt2 CSN AIEt, CLCN AIEt2 CLCN AIEt2 CdN CtCN

Temp(%)

-100 -78 -1 00 -78 -100 -78 -1 0 0 -78 -78

a : b : c : d

100 14 100 10 100 7 100 4 4

. 100

. 100 100 100 100

7