[acs symposium series] selectivity in catalysis volume 517 || catalytic asymmetric heck reaction

11
Chapter 6 Catalytic Asymmetric Heck Reaction Fumiyuki Ozawa, Akihiko Kubo, and Tamio Hayashi Catalysis Research Center and Graduate School of Pharmaceutical Science, Hokkaido University, Sapporo 060, Japan Catalytic asymmetric arylation of 2,3-dihydrofuran with phenyl triflate in benzene in the presence of 1,8-bis(dimethylamino)naphthalene (proton sponge) and a chiral palladium catalyst, generated in situ from Pd(OAc)2 and 2 equiv/Pd of (R)-BINAP, at 40 °C gave optically active (R)-2-phenyl-2,3-dihydrofuran in extremely high enantioselectivity (>96% ee), together with a small amount of the regioisomer (S)-2- phenyl-2,5-dihydrofuran. Under similar conditions, various aryl triflates (ArOTf; Ar = p-ClC 6 H 4 , m-ClC 6 H 4 , o-ClC 6 H 4 , p-AcC 6 H 4 , p- NCC6H4, p-FC6H4, p-MeOC6H4, p-MeC6H4, and 2-naphthyl) reacted with 2,3-dihydrofuran to yield the corresponding (R)-2-aryl-2,3- dihydrofurans of >96-87% ee. The mechanism and factors governing this novel catalytic asymmetric reaction are discussed in detail. The palladium-catalyzed arylation and alkenylation of olefins, which were first discovered in the 1970's by Heck (7,2) and Mizoroki (3) and have been often called the "Heck reaction", are versatile synthetic means for making a carbon-carbon bond. These reactions have been extensively used for organic synthesis during the past two decades (4-7). However, no reports on the "asymmetric Heck reaction" have been appeared until very recently. Shibasaki reported an asymmetric intramolecular cyclization of alkenyl iodides to give ris-decalin derivatives of 80-91% ee (8-10). Overman reported an intramolecular cyclization of alkenyl triflate, giving a chiral quaternary carbon center of 45% ee (77). We report herein the first example of intermolecular asymmetric Heck-type arylation of cyclic olefins catalyzed by (R)- BINAP-coordinated palladium complexes (Scheme 1) (12,13). Catalytic Asymmetric Arylation of Olefins Scheme 2 shows the mechanism generally accepted for the catalytic arylation of olefins with aryl iodides in the presence of a tertiary phosphine-coordinated palladium catalyst and a base (4). Oxidative addition of aryl iodide (Arl) to a Pd(0) species (A), which is most commonly generated from palladium diacetate and a tertiary phosphine ligand, forms an arylpalladium iodide complex (B). Coordination of olefin on Β followed by insertion of the coordinated olefin into the Pd-Ar bond forms a σ-alkylpalladium species (C), which undergoes β-hydrogen elimination reaction to give the arylation 0097-6156/93/0517-0075$06.00/0 © 1993 American Chemical Society Downloaded by JOHNS HOPKINS UNIV on October 2, 2013 | http://pubs.acs.org Publication Date: May 5, 1993 | doi: 10.1021/bk-1993-0517.ch006 In Selectivity in Catalysis; Davis, M., et al.; ACS Symposium Series; American Chemical Society: Washington, DC, 1993.

Upload: steven-l

Post on 11-Dec-2016

213 views

Category:

Documents


0 download

TRANSCRIPT

Chapter 6

Catalytic Asymmetric Heck Reaction

Fumiyuki Ozawa, Akihiko Kubo, and Tamio Hayashi

Catalysis Research Center and Graduate School of Pharmaceutical Science, Hokkaido University, Sapporo 060, Japan

Catalytic asymmetric arylation of 2,3-dihydrofuran with phenyl triflate in benzene in the presence of 1,8-bis(dimethylamino)naphthalene (proton sponge) and a chiral palladium catalyst, generated in situ from Pd(OAc)2 and 2 equiv/Pd of (R)-BINAP, at 40 °C gave optically active (R)-2-phenyl-2,3-dihydrofuran in extremely high enantioselectivity (>96% ee), together with a small amount of the regioisomer (S)-2-phenyl-2,5-dihydrofuran. Under similar conditions, various aryl triflates (ArOTf; Ar = p-ClC6H4, m-ClC6H4, o-ClC6H4, p-AcC6H4, p-NCC6H4, p-FC6H4, p-MeOC6H4, p-MeC6H4, and 2-naphthyl) reacted with 2,3-dihydrofuran to yield the corresponding (R)-2-aryl-2,3-dihydrofurans of >96-87% ee. The mechanism and factors governing this novel catalytic asymmetric reaction are discussed in detail.

The palladium-catalyzed arylation and alkenylation of olefins, which were first discovered in the 1970's by Heck (7,2) and Mizoroki (3) and have been often called the "Heck reaction", are versatile synthetic means for making a carbon-carbon bond. These reactions have been extensively used for organic synthesis during the past two decades (4-7). However, no reports on the "asymmetric Heck reaction" have been appeared until very recently. Shibasaki reported an asymmetric intramolecular cyclization of alkenyl iodides to give ris-decalin derivatives of 80-91% ee (8-10). Overman reported an intramolecular cyclization of alkenyl triflate, giving a chiral quaternary carbon center of 45% ee (77). We report herein the first example of intermolecular asymmetric Heck-type arylation of cyclic olefins catalyzed by (R)-BINAP-coordinated palladium complexes (Scheme 1) (12,13).

Catalytic Asymmetric Arylation of Olefins

Scheme 2 shows the mechanism generally accepted for the catalytic arylation of olefins with aryl iodides in the presence of a tertiary phosphine-coordinated palladium catalyst and a base (4). Oxidative addition of aryl iodide (Arl) to a Pd(0) species (A), which is most commonly generated from palladium diacetate and a tertiary phosphine ligand, forms an arylpalladium iodide complex (B). Coordination of olefin on Β followed by insertion of the coordinated olefin into the Pd-Ar bond forms a σ-alkylpalladium species (C), which undergoes β-hydrogen elimination reaction to give the arylation

0097-6156/93/0517-0075$06.00/0 © 1993 American Chemical Society

Dow

nloa

ded

by J

OH

NS

HO

PKIN

S U

NIV

on

Oct

ober

2, 2

013

| http

://pu

bs.a

cs.o

rg

Pub

licat

ion

Dat

e: M

ay 5

, 199

3 | d

oi: 1

0.10

21/b

k-19

93-0

517.

ch00

6

In Selectivity in Catalysis; Davis, M., et al.; ACS Symposium Series; American Chemical Society: Washington, DC, 1993.

76 SELECTIVITY IN CATALYSIS

product and a hydridopalladium complex (D). Complex D thus formed further interacts with a base to regenerate the Pd(0) species (A) which further carries the catalytic cycle.

The main reason why the Heck reaction has not been applied to asymmetric synthesis until very recently is that the Heck reaction generally does not create new chiral carbon centers in the products. The chiral carbon center formed by the olefin-insertion is lost at the subsequent β-hydrogen elimination step. An exception to this was found in the reaction of a cyclic olefin such as 2,3-dihydrofuran where the chiral carbon center remains in the arylation product (Scheme 3). The reaction of 2,3-dihydrofuran with an arylpalladium species via the common cis insertion process forms alkylpalladium intermediate E. In this intermediate, the hydrogen bound to the arylated carbon, that is trans to the palladium, can not participate in the β-hydrogen elimination. Alternatively, the cis β-hydrogen at another β-carbon may be abstracted by palladium. Dissociation of the coordinated olefin from the resulting hydrido-olefin complex F gives 2-aryl-2,5-dihydrofuran (3), whereas insertion of the olefin into the Pd-H bond in F followed by β-hydrogen elimination forms the thermodynamically more stable 2-aryl-2,3-dihydrofuran (2), in which the arylated carbon center remains chiral.

We therefore studied asymmetric arylation of 2,3-dihydrofuran. In our initial attempts, we applied chiral diphosphine ligands to a common arylation system using phenyl iodide as the arylating reagent, but no or little asymmetric induction has been observed For example, the reaction of 2,3-dihydrofuran with phenyl iodide in THF in the presence of triethylamine and 3 mol% of a palladium catalyst, prepared in situ from Pd(OAc)2 and (fl)-BINAP, at 40 °C for 44 h gave racemic 2-phenyl-2,3-dihydrofuran (2a) and 2-phenyl-2,5-dihydrofuran (3a) in 23% and 2% yields, respectively. This reaction must involve a [PdPh(I)(BINAP)] intermediate formed by oxidative addition of phenyl iodide to a [Pd(0)(BINAP)] species generated in the system. Recent mechanistic studies on olefin-insertion into the related acylpalladium halide complexes [Pd(COR)(X)L2] having two tertiary phosphine ligands (L) revealed that the insertion process involves displacement of one of the phosphine ligands with olefin to give a [Pd(COR)(olefin)X(L)] type intermediate (74,75). Therefore, we reasoned that the absence of asymmetric induction in the arylation system with phenyl iodide is attributable to partial dissociation of BINAP ligand during the olefin-insertion process.

On the other hand, we recently found that a cationic acetylpalladium complex bearing l,2-bis(diphenylphosphino)ethane (dppe) as a potent bidentate ligand [Pd(COMe)(NCMe)(dppe)]+BF4~ (4) possesses extremely high reactivity toward olefin-insertion (Scheme 4) (16). Complex 4 reacted with cyclic olefins, including cyclopentene and cycloheptene, under very mild conditions (-40 °C) to give the olefin-inserted complexes. Cationic complex 4 has a coordination site that is weakly coordinated with acetonitrile and easily accessible for coordination of olefins. The presence of bidentate ligand allows the acetyl group and the coordinated olefin to come to the adjacent sites (cis) required for the subsequent olefin-insertion. The overall situation results in facile olefin-insertion without involving the partial dissociation of phosphine ligand.

Taking the above mechanistic information into account, we next sought the catalytic asymmetric arylation system that involves a cationic arylpalladium intermediate. First, we examined the reaction of 2,3-dihydrofuran with phenyl iodide and Et3N in the presence of AgC104 and a palladium catalyst coordinated with (R)-BINAP ligand. The silver salt employed may react with the [PdPh(I)(BIANP)] intermediate to generate a desirable cationic phenylpalladium species. The reaction carried out in THF at 40 °C for 44 h actually gave optically active (7?)-2-phenyl-2,3-dihydrofuran (2a) of 65% ee (25% yield) together with a small amount of its regioisomer (3a) (1% yield). Secondly, we tried to employ arylpalladium triflate intermediate [Pd(Ar)(OTf)(BINAP)] (5) which may be generated in the catalytic system by oxidative addition of aryl triflate to a Pd(0) species (72). Complex 5 having the triflate ligand as a good leaving group was expected to form a cationic olefin-

Dow

nloa

ded

by J

OH

NS

HO

PKIN

S U

NIV

on

Oct

ober

2, 2

013

| http

://pu

bs.a

cs.o

rg

Pub

licat

ion

Dat

e: M

ay 5

, 199

3 | d

oi: 1

0.10

21/b

k-19

93-0

517.

ch00

6

In Selectivity in Catalysis; Davis, M., et al.; ACS Symposium Series; American Chemical Society: Washington, DC, 1993.

OZAWA ET AL. Catalytic Asymmetric Heck Reaction

ArOTf

1

[Pd(OAc) 2/2(fl)-BINAP]

base

in benzene

Ο ( *CL ""Ar I N / ^ A r Ό

(R)-2

(>96-87% ee)

'Cf

(S)-3

P P h 2

P P h 2

(fl)-BINAP

Ar= Ph, p-CIC 6 H 4 , m-CIC 6 H 4 , o-CIC 6 H 4 , p - A c C 6 H 4 . p - N C C 6 H 4 , p - F C 6 H 4 , p -MeOC 6 H 4 , p -MeC 6 H 4 , 2-naphthyl

Base-HI

Base

Scheme 1.

PdL n

(A)

PdH(l)L2 (D)

Arl

ArPd(l)L2 (B)

RCH=CHAr

R

L CH > / N C H 2 A r

/ \ (O

Scheme 2.

RCH=CH 2

(E) {trans )

-[Pd-H]

^Ar

. u Ar p . d ~ H Ar . .

(2) (F) (3)

Scheme 3.

Dow

nloa

ded

by J

OH

NS

HO

PKIN

S U

NIV

on

Oct

ober

2, 2

013

| http

://pu

bs.a

cs.o

rg

Pub

licat

ion

Dat

e: M

ay 5

, 199

3 | d

oi: 1

0.10

21/b

k-19

93-0

517.

ch00

6

In Selectivity in Catalysis; Davis, M., et al.; ACS Symposium Series; American Chemical Society: Washington, DC, 1993.

78 SELECTIVITY IN CATALYSIS

coordinated species [Pd(Ar)(olefin)(BINAP)]+OTf- (6) with retention of the chelate coordination of the BINAP ligand, leading to the highly enantioselective Heck reaction.

Reaction of 2,3-dihydrofuran with aryl triflate in benzene in the presence of a base and a palladium catalyst, prepared in situ from Pd(OAc)2 and 2 equivalents of (/?)-BINAP, gave (/?)-2-aryl-2,3-dihydrofuran (2) and a small amount of the regioisomer (S)-2-aryl-2,3-dihydrofuran (3) (Scheme 1). Table I demonstrates that the arylation reaction proceeds in extremely high enantioselectivity with a variety of aryl triflates. Particularly, almost enantiomerically pure (R)-2 was obtained with 2-naphthyl triflate and phenyl triflates bearing an electron-withdrawing substituent at the para or meta position.

The regioselectivity and the enantiomeric purities of the arylation products are strongly affected by the base employed (Ozawa, F.; Kubo, Α.; Hayashi, T., Tetrahedron Lett., in press). Table II lists the results of catalytic reactions of 2,3-dihydrofuran with phenyl triflate in the presence of a variety of bases. Extremely high enantioselectivity for (J?)-2a (>96% ee) was obtained with highly basic and sterically demanding l,8-bis(dimethylamino)naphthalene (proton sponge) (run 1). On the other hand, the enantiomeric purity of (fl)-2a was lower in the reactions using more compact aliphatic amines and less basic amines including pyridine derivatives (runs 2-6). Inorganic bases such as sodium acetate and sodium carbonate also gave lower enantioselectivities (runs 7 and 8). It is interesting to note that increasing product ratio of the minor product (5)-3a tends to increase the enantiomeric purity of major product (/?)-2a.

Under similar reaction conditions, N-substituted pyrrolines were arylated in 68-83% ee (Scheme 5) (Ozawa, F.; Hayashi, T., / . Organomet. Chem., in press).

Reaction Mechanism

Figure 1 depicts a molecular view of [PdPh{(/?)-(BINAP)}]+ species which was drawn with the Chem 3D computer program using the partial X-ray structure of [PdCl2{(/?)-BINAP}]. It is seen that the lower part of the empty coordination site widely opens whereas the upper part is blocked by one of the phenyl groups of the BINAP ligand.

Scheme 6 shows the two possible modes for coordination of 2,3-dihydrofuran to the [PdPh{(/?)-BINAP}]+ species. As suggested from the molecular model in Figure 1, dihydrofuran may coordinate to the palladium center more easily in mode (a) than (b). The olefin-coordination in mode (a) followed by olefin-insertion and β-hydrogen elimination reactions forms the phenylation product having (/^-configuration, that is the observed configuration in major regioisomer 2 in the actual catalytic reactions.

An interesting feature of this asymmetric reaction is that the arylation products 2 and 3 have the opposite configurations to each other, indicating that a kinetic resolution process enhances the enantiomeric purity of major product (R)-2 by selective elimination of (S)-isomer as minor product 3 from the catalytic cycle (Scheme 7). Coordination of dihydrofuran on arylpalladium intermediate 5 forms two types of olefin-coordinated complexes 6R and 65, depending on the selection of enantiofaces of olefin. Complexes 6R and 65 thus formed undergo the subsequent olefin-insertion and β-hydrogen elimination reactions to give a pair of diastereomers of hydrido-olefin complexes having (7?)-2-aryl-2,5-dihydrofuran and (5)-2-aryl-2,5-dihydrofuran ligands, SR and 85, respectively. Molecular inspection using C P K models has suggested that SR has a preferable structure for further olefin-insertion and β-hydrogen elimination reactions to give major product (R)-2 (see Scheme 3). On the other hand, the diastereomer 85 suffers considerable steric repulsion between the coordinated olefin

Dow

nloa

ded

by J

OH

NS

HO

PKIN

S U

NIV

on

Oct

ober

2, 2

013

| http

://pu

bs.a

cs.o

rg

Pub

licat

ion

Dat

e: M

ay 5

, 199

3 | d

oi: 1

0.10

21/b

k-19

93-0

517.

ch00

6

In Selectivity in Catalysis; Davis, M., et al.; ACS Symposium Series; American Chemical Society: Washington, DC, 1993.

6. OZAWA ET AL. Catalytic Asymmetric Heck Reaction 79

c P h 2

^ O M e

Pd

P / N N C M e P h 2

(4)

B F 4

Ο - M e C N

inCH 2CI 2 , a t - ^ 0 ° C

BF 4 ~

(n = 1,3)

Scheme 4.

Table I. Catalytic Asymmetric Arylation of 2,3-Dihydrofuran with Aryl Triflates a

run AiOTf

(Ar)

9 ?>eeb

2/3 d

yield ( %)C run AiOTf

(Ar) (R)-2 (5)-3 2/3 d (Λ)-2 (5)-3

1 Ph (la) >96 17 71/29 46 24

2 P-CIQ5H4 (lb) >96 6 70/30 54 21

3 OT-CIC6H4 (1c) >96 39 72/28 66 22

4 0-CIC6H4 (Id) 92 53 73/27 53 21

5 p-AcQsH4 (le) >96 7 e 50 24

6 /7-NCC6H4 (If) >96 4(Λ) 66/34 52 26

7 f P-FC6H4 (lg) >96 24 71/29 46 19

8 g p-MeOC6H4 (lh) 87 11 W 63/37 42 25

9 f p-MeC6H4 (li) 88 42 77/23 52 15

10 2-naphthyl d j ) >96 10 69/31 52 26

a The reaction was carried out in benzene at 40 °C under a nitrogen atmosphere in the presence of l,8-bis(dimethylamino)naphthalene (proton sponge) as the base for 2-9 days unless otherwise noted. Initial conditions: ArOTf/2,3-dihydrofuran/base/ Pd(OAc)2/(/?)-BINAP = 1/5/3/0.03/0.06. b Determined by *H N M R analysis using an optically active shift reagent Eu(hfc)3. c Isolated yield by silica gel column chromato­graphy at 100% conversion of aryl triflate. ^ Product ratio in the catalytic solution as determined by G L C . e Not measured. f The reaction was carried out at 60 °C. g The reaction was carried out at 50 °C.

Dow

nloa

ded

by J

OH

NS

HO

PKIN

S U

NIV

on

Oct

ober

2, 2

013

| http

://pu

bs.a

cs.o

rg

Pub

licat

ion

Dat

e: M

ay 5

, 199

3 | d

oi: 1

0.10

21/b

k-19

93-0

517.

ch00

6

In Selectivity in Catalysis; Davis, M., et al.; ACS Symposium Series; American Chemical Society: Washington, DC, 1993.

80 SELECTIVITY IN CATALYSIS

Table II. Effect of Base on the Catalytic Asymmetric Arylation of 2,3-Dihydrofuran with Phenyl Triflate (la) a

run baseb reaction conversion c 2a/3ac 9 bee d

time of la (%) (/?)-2a (5)-3a

Me^N N M e 2

1 9 d 100 71/29 >96 17

2 C y 2 N H 17 h 100 86/14 82 43

3 /-Pr 2NEt 24 h 100 92/8 82 60

4 E t 3 N 26 h 100 98/2 75 9

5 5 d 88 100/0 77 —

6 9 d 38 100/0 69 —

7 e MeOOONa 2 d 82 96/4 71 3(R) 8 e N a 2 C 0 3 I d 100 97/3 75 93 (R)

a The reaction was carried out in benzene under a nitrogen atmosphere at 40 °C unless otherwise noted. Initial conditions: la/2,3-a^hydrofin*an/base/Pd(OAc)2/(/?)-BINAP = 1/5/3/0.03/0.06. b C y 2 N H = dicyclohexylamine. c Determined by G L C . d Determined by lR N M R analysis using an optically active shift reagent Eu(hfc)3. e The reaction was carried out at 60 °C.

Γ\ [Pd(OA0) 2/2(H)-B,NAP ] ΓΛ f+C\ \

base "'Ar X

N ' / ^ A r

COOR in benzene C Q O R ^ COOR J

Ar = Ph, p-CIC 6 H 4 , />AcC 6 H 4 ) p - N C C 6 H 4

R = Me, Ph (68-83% ee)

Scheme 5.

Dow

nloa

ded

by J

OH

NS

HO

PKIN

S U

NIV

on

Oct

ober

2, 2

013

| http

://pu

bs.a

cs.o

rg

Pub

licat

ion

Dat

e: M

ay 5

, 199

3 | d

oi: 1

0.10

21/b

k-19

93-0

517.

ch00

6

In Selectivity in Catalysis; Davis, M., et al.; ACS Symposium Series; American Chemical Society: Washington, DC, 1993.

6. OZAWA ET AL. Catalytic Asymmetric Heck Reaction 81

Mode (b) W I " - ^ - ^ p ^ , p

+

d r - p ^ J f L ^ ^ P ^ p ^ P .

Mode (a) ^ - — Q

Scheme 6. A schematic process for the enantioselective insertion of 2,3-dihydro­furan into the Pd-Ph bond in [PdPh{(fl)-BINAP}] + species, showing the two possible modes (a) and (b) for olefin-coordination. The binaphthyl group in (/?)-BINAP ligand is omitted for clarity.

Dow

nloa

ded

by J

OH

NS

HO

PKIN

S U

NIV

on

Oct

ober

2, 2

013

| http

://pu

bs.a

cs.o

rg

Pub

licat

ion

Dat

e: M

ay 5

, 199

3 | d

oi: 1

0.10

21/b

k-19

93-0

517.

ch00

6

In Selectivity in Catalysis; Davis, M., et al.; ACS Symposium Series; American Chemical Society: Washington, DC, 1993.

Sche

me

7. P

ropo

sed

mec

hani

sm f

or th

e ki

netic

reso

lutio

n pr

oces

s.

Dow

nloa

ded

by J

OH

NS

HO

PKIN

S U

NIV

on

Oct

ober

2, 2

013

| http

://pu

bs.a

cs.o

rg

Pub

licat

ion

Dat

e: M

ay 5

, 199

3 | d

oi: 1

0.10

21/b

k-19

93-0

517.

ch00

6

In Selectivity in Catalysis; Davis, M., et al.; ACS Symposium Series; American Chemical Society: Washington, DC, 1993.

6. OZAWA ET AL. Catalytic Asymmetric Heck Reaction 83

and one of the phenyl groups in the BINAP ligand. Consequently, 85 is prone to release the coordinated olefin (S)-3. The overall process gives rise to the enhancement in the enantiomeric purity of major product (R)-2.

For example, based on the product ratio of (#)-2a to (S)-3a and the enantiomeric purities of both compounds, the relative ratio of SR to 85 in run 1 in Table I may be calculated as 83 : 17. Therefore, if both diastereomers gave the same regioisomer 2a, the catalytic reaction should give (/?)-2a of 66% ee; the value being much lower than that observed in the actual catalytic system (>96% ee).

The initial step of the catalytic cycle is oxidative addition of aryl triflate to a BINAP-coordinated Pd(0) species. Since, in the actual catalytic system, Pd(OAc)2 and (/?)-BINAP are used as the precursors of the Pd(0) species, reduction of Pd(OAc)2 into the BINAP-coordinated Pd(0) species should be operative prior to the catalytic reaction. Although Pd(OAc)2 is the most commonly used precursor of a Pd(0) species in many palladium-catalyzed organic reactions, no direct information has been reported so far on its reduction process. In this study, we confirmed for the first time that the reduction proceeds according to the process involving a combination of tertiary phosphine (BINAP) and water as the reducing reagent (Scheme 8) (Ozawa, F.; Kubo, Α.; Hayashi, T., submitted for publication).

Treatment of a benzene solution of Pd(OAc)2 and 3 equivalents of (fl)-BINAP with an excess amount of Et3N resulted in change in the color of the solution from yellow to red. 3 1 P{ lH] NMR study of the resulting solution revealed the formation of [Pd{(/?)-BINAP} 2] and (tf)-BINAP monoxide ((fl)-BINAP(O)) in a 1 : 1 ratio. Labeling experiments showed the participation of 1 equivalent of residual water in the reduction system. Thus, the reaction carried out in solution saturated with 1 8 OH2 gave ^O-labeled (/?)-BINAP(l8o).

On the basis of the well-defined information on the catalytic system, our attention was directed toward further details of the kinetic resolution process. Thus, the catalytic reaction was examined with isolated [Pd{(/?)-BINAP))2l catalyst and i-PrçNEt as the base (Table III). Surprisingly, 2a was exclusively formed in the catalytic reaction using the [Pd{(/?)-BINAP)}2] catalyst (run 1), indicating that no kinetic resolution takes place. Addition of BINAP(O) did not cause the kinetic resolution (run 2). In contrast, the kinetic resolution occurred in the presence of added acetic acid (runs 3 and 4). The larger amount of acetic acid gave rise to the higher ratio of 3a and the higher enantiomeric purity of (/?)-2a. In the presence of 5 equiv/Pd of acetic acid, almost enantiomerically pure (#)-2a was obtained.

Since the catalytic system contains an excess amount of amine, acetic acid added to the system forms the ammonium acetate salt, which is also formed in the catalytic system starting with Pd(OAc)2 as the catalyst precursor during the conversion of Pd(OAc)2 into a Pd(0) species (Scheme 8). The results described above demonstrate that the acetate anion thus formed serves as the promoter of the kinetic resolution process. Since the hydrido-olefin complex 85 in Scheme 8 has a 16e square planar structure, dissociation of the coordinated olefin should proceed via an associative mechanism involving an 18e transition state formed by nucleophilic attack of an incoming ligand to the palladium center. The acetate anion may possess the high nucleophilicity toward the cationic palladium center in 85 to cause the dissociation of the olefin. The nucleophilicity of acetate anion should be defined by the counter cation particularly in nonpolar media such as benzene. The marked difference in effect of bases in the kinetic resolution process is attributable to the difference in the nucleophilicity of acetate anion that is controlled by the ammonium cation. Proton sponge as one of the most strong bases for trapping proton in organic solvents may generate the highly reactive acetate anion that promotes the efficient kinetic resolution giving (R)-2 of extremely high enantiomeric purity.

Dow

nloa

ded

by J

OH

NS

HO

PKIN

S U

NIV

on

Oct

ober

2, 2

013

| http

://pu

bs.a

cs.o

rg

Pub

licat

ion

Dat

e: M

ay 5

, 199

3 | d

oi: 1

0.10

21/b

k-19

93-0

517.

ch00

6

In Selectivity in Catalysis; Davis, M., et al.; ACS Symposium Series; American Chemical Society: Washington, DC, 1993.

84 SELECTIVITY IN CATALYSIS

Pd(OAc) 2 + H 2 0 + 3 (fl)-BINAP + 2 Et 3 N benzene

[Pd{(fl)-BINAP}2] + (fl)-BINAP(O) + 2 E t 3 N H O A c

*P(0)Ph 2

. P P h 2

s A ^ 1 (fl)-BINAP(O)

Scheme 8.

Table III. Catalytic Asymmetric Arylation of 2,3-Dihydrofuran with Phenyl Triflate (la) Promoted by [Pd{(fl)-BINAP}2] Catalysta

run additive % ee^

(equiv./Pd) 2a/3ac (/?)-2a (5)-3a

1 — 100/0 77 — 2 BINAP(0)(1) 100/0 76 — 3 AcOH (2) 89/11 83 64 4 AcOH (5) 84/16 >96 60

a The reaction was carried out in benzene at 40 °C under a nitrogen atmosphere in the presence of /-PrçNEt as the base for 1-2 days. Initial conditions: la/2,3-dihydrofuran/ /-Pr2NEt/catalyst = 1/5/3/0.03. b Determined by lH NMR analysis using an optically active shift reagent Eu(hfc)3. c Product ratio in the catalytic solution as determined by GLC. D

ownl

oade

d by

JO

HN

S H

OPK

INS

UN

IV o

n O

ctob

er 2

, 201

3 | h

ttp://

pubs

.acs

.org

P

ublic

atio

n D

ate:

May

5, 1

993

| doi

: 10.

1021

/bk-

1993

-051

7.ch

006

In Selectivity in Catalysis; Davis, M., et al.; ACS Symposium Series; American Chemical Society: Washington, DC, 1993.

6. OZAWA ET AL. Catalytic Asymmetric Heck Reaction 85

Literature Cited

1. Heck, R. R; Nolley, Jr., J. P. J. Org. Chem. 1972, 37, 2320. 2. Dieck, Η. Α.; Heck, R. F. J. Am. Chem. Soc. 1974, 96, 1133. 3. Mori, K.; Mizoroki, T.; Ozaki, A. Bull. Chem. Soc. Jpn. 1973, 46, 1505. 4. Heck, R. F. Org. React. 1982, 27, 345. 5. Heck, R. F. Acc. Chem. Res. 1979, 12, 146. 6. Heck, R. F. In Palladium Reagents in Organic Synthesis; Academic Press: New

York, 1985. 7. Daves, Jr., G. D.; Hallberg, A. Chem. Rev. 1989, 89, 1433. 8. Sato, Y.; Sodeoka M.; Shibasaki, M. J. Org. Chem. 1989, 54, 4738. 9. Sato, Y.; Sodeoka M.; Shibasaki, M. Chem. Lett. 1990, 1953.

10. Kagechika, K.; Shibasaki, M. J. Org. Chem. 1991, 56, 4093. 11. Carpenter, N. E.; Kucera, D. J.; Overman, L. E. J. Org. Chem. 1989, 54, 5846. 12. Ozawa, F.; Kubo, Α.; Hayashi, T. J. Am. Chem. Soc. 1991, 113, 1417. 13. (R)-BINAP: (R)-2,2'-bis(diphenylphosphino)-1,1'-binaphthyl. Takaya, T.;

Mashima, K.; Koyano, K.; Yagi, M.; Kumobayashi, H.; Taketomi, T.; Akutagawa, S.; Noyori, R. J. Org. Chem. 1986, 51, 629, and references cited therein.

14. Brumbaugh, J. S.; Whittle, R. R.; Parvez, M.; Sen, A. Organometallics 1990, 9, 1735.

15. Samsel, E. G.; Norton, J. R. J. Am. Chem. Soc. 1984, 106, 5505. 16. Ozawa, F.; Hayashi, T.; Koide, H.; Yamamoto, A. J. Chem. Soc., Chem.

Commun. 1991, 1469.

RECEIVED July 9, 1992

Dow

nloa

ded

by J

OH

NS

HO

PKIN

S U

NIV

on

Oct

ober

2, 2

013

| http

://pu

bs.a

cs.o

rg

Pub

licat

ion

Dat

e: M

ay 5

, 199

3 | d

oi: 1

0.10

21/b

k-19

93-0

517.

ch00

6

In Selectivity in Catalysis; Davis, M., et al.; ACS Symposium Series; American Chemical Society: Washington, DC, 1993.