University of Groningen Monodentate secondary phosphine ... · Asymmetric catalysis This chapter...

16
University of Groningen Monodentate secondary phosphine oxides (SPO's), synthesis and application in asymmetric catalysis Jiang, Xiaobin IMPORTANT NOTE: You are advised to consult the publisher's version (publisher's PDF) if you wish to cite from it. Please check the document version below. Document Version Publisher's PDF, also known as Version of record Publication date: 2004 Link to publication in University of Groningen/UMCG research database Citation for published version (APA): Jiang, X. (2004). Monodentate secondary phosphine oxides (SPO's), synthesis and application in asymmetric catalysis. [Groningen]: s.n. Copyright Other than for strictly personal use, it is not permitted to download or to forward/distribute the text or part of it without the consent of the author(s) and/or copyright holder(s), unless the work is under an open content license (like Creative Commons). Take-down policy If you believe that this document breaches copyright please contact us providing details, and we will remove access to the work immediately and investigate your claim. Downloaded from the University of Groningen/UMCG research database (Pure): http://www.rug.nl/research/portal. For technical reasons the number of authors shown on this cover page is limited to 10 maximum. Download date: 19-05-2020

Transcript of University of Groningen Monodentate secondary phosphine ... · Asymmetric catalysis This chapter...

Page 1: University of Groningen Monodentate secondary phosphine ... · Asymmetric catalysis This chapter introduces asymmetric catalysis, especially asymmetric hydrogenation of C=C and C=O

University of Groningen

Monodentate secondary phosphine oxides (SPO's), synthesis and application in asymmetriccatalysisJiang, Xiaobin

IMPORTANT NOTE: You are advised to consult the publisher's version (publisher's PDF) if you wish to cite fromit. Please check the document version below.

Document VersionPublisher's PDF, also known as Version of record

Publication date:2004

Link to publication in University of Groningen/UMCG research database

Citation for published version (APA):Jiang, X. (2004). Monodentate secondary phosphine oxides (SPO's), synthesis and application inasymmetric catalysis. [Groningen]: s.n.

CopyrightOther than for strictly personal use, it is not permitted to download or to forward/distribute the text or part of it without the consent of theauthor(s) and/or copyright holder(s), unless the work is under an open content license (like Creative Commons).

Take-down policyIf you believe that this document breaches copyright please contact us providing details, and we will remove access to the work immediatelyand investigate your claim.

Downloaded from the University of Groningen/UMCG research database (Pure): http://www.rug.nl/research/portal. For technical reasons thenumber of authors shown on this cover page is limited to 10 maximum.

Download date: 19-05-2020

Page 2: University of Groningen Monodentate secondary phosphine ... · Asymmetric catalysis This chapter introduces asymmetric catalysis, especially asymmetric hydrogenation of C=C and C=O

Chapter 1

Chapter 1

Asymmetric catalysis This chapter introduces asymmetric catalysis, especially asymmetric hydrogenation of

C=C and C=O bonds in prochiral substrates. Several parameters that influence rate and

enantioselectivity in asymmetric hydrogenation are also discussed. Contents

1.1 Selectivity in organic synthesis 2

1.2 Asymmetric catalysis in organic synthesis 3

1.3 Introduction to asymmetric hydrogenation 4

1.4 Important parameters in asymmetric hydrogenation 7

1.5 References and notes 11

1

Page 3: University of Groningen Monodentate secondary phosphine ... · Asymmetric catalysis This chapter introduces asymmetric catalysis, especially asymmetric hydrogenation of C=C and C=O

Asymmetric catalysis

1.1 Selectivity in organic synthesis Over the years, a large number of organic reactions has been developed which provide us tools for organic synthesis. However, many of these reactions lack selectivity. This imperfect control leads to a reduced yield of the desired product and the necessity of separation from side products.1 This purification is often time consuming, expensive and less efficient with regard to material use. The advantages of complete control and atom economy transformations are obvious.2 We can identify three major types of control in organic synthesis, all dealing with the selectivity issue: chemoselectivity; regioselectivity and stereoselectivity.3

Chemoselectivity implies the preferential reaction of one functional group over another under the same reaction conditions. For example, a ketone can usually be reduced to a secondary alcohol by NaBH4 in the presence of ester group (Scheme 1.1).

R2O R1

OONaBH4

R2O R1

OHO

H

Scheme 1.1 Example of chemoselectivity Regioselectivity is the preferential formation of one isomer of the product over another. In the reaction of epoxides ring opening, nucleophiles can attack from different position to form different products depending on the nucleophiles, reaction mechanism and conditions (Scheme 1.2). 3

R2

Nu

OH

R1 R2

OHR1

Nu

O

R1 R2or+ Nucleophiles

Scheme 1.2 Regioselective nucleophilic ring opening of epoxides The most important and difficult control to achieve pertains to stereoselectivity, which can be divided into enantioselectivity and diastereoselectivity. Enantiomers are a special kind of stereoisomers where the two molecules have a mirror image relationship. They have the same physical properties but different arrangements or orientation of their atoms in space. Enantioselectivity is the preferential formation of one enantiomer of the product over the other and it is normally expressed as the enantiomeric excess (e.e.).

e.e. = R(%) -S(%)

R(%) +S(%)100%X

Diastereomers exist of compounds that contain at least two stereogenic centers or stereo elements. Some achiral compounds (alkenes) can also exist as diastereomers: the cis and trans isomers (the necessary and sufficient condition is that one substituent at each end of

2

Page 4: University of Groningen Monodentate secondary phosphine ... · Asymmetric catalysis This chapter introduces asymmetric catalysis, especially asymmetric hydrogenation of C=C and C=O

Chapter 1

the double bond differ from the other, also called E/Z isomers) or geometrical isomers are not enantiomers but diastereomers. Diastereoselectivity is the preferential formation of one diastereoisomer of the product over the others and it is usually given as the diastereoisomeric excess (d.e.) according to:

d.e. = D1(%) -D2(%)

D1(%) +D2(%)100%X

Most chiral organic compounds found in nature are enantiopure. In most cases, the two enantiomers exhibit different biological activities.4,5 For example, the L-enantiomer of some α-amino acids such as leucine, phenylalanine, tyrosine and tryptophane tastes bitter, whereas their corresponding D-enantiomers are sweet.5

The active molecules of many drugs are single enantiomers of chiral compounds. In some cases, the other enantiomer may behave antagonistically and/or cause undesired side effects.6 Organic synthesis that leads to a single isomer of the products is therefore highly valuable. The design and development of new methods for stereoselective synthesis will continue to be one of the major challenges in the future. 1.2 Asymmetric catalysis in organic synthesis Ideally, the synthesis of optical active compounds induces complete control of the product and the intermediate stereochemistry to the extent that they are formed as a single or major stereoisomer. These enantiopure compounds can be made by optical resolution 3a,7 or asymmetric synthesis,7-9 which can be achieved in several ways: (a) from chiral natural products (“chiral pool” approach);3a,7 (b) use of a chiral auxiliary;8 (c) asymmetric catalysis.9 The chiral pool method is limited to natural-occurring chiral compounds and often only one of the enantiomers is available, e.g. D-sugars, α-amino acids. The major disadvantage of optical resolution is that half of the product is the undesired enantiomer, although via racemization of the undesired enantiomer, the yield can be improved. The chiral auxiliary in method c has to be used in stoichiometric amounts. After reaction, removing the chiral auxiliary requires additional steps. By contrast, an asymmetric catalytic reaction with a turnover number of 1000 means that 1000 new chiral molecules can be generated with one molecule of a chiral catalyst, which leads to higher economy and efficiency. The work-up is often relatively simple, as only a small amount of the catalyst is used. In the past 20 years, asymmetric catalysis has become one of the most important areas of research and major breakthroughs have been achieved.6-10 Some of the catalytic methods developed have been successfully applied in the synthesis of enantiopure drugs on industrial scale.6, , 12, 22 11

For instance, asymmetric hydrogenation was applied in the production of a key intermediate 1.1 for L-DOPA, which was one of the first industrial asymmetric syntheses

3

Page 5: University of Groningen Monodentate secondary phosphine ... · Asymmetric catalysis This chapter introduces asymmetric catalysis, especially asymmetric hydrogenation of C=C and C=O

Asymmetric catalysis

(Scheme 1.3). By using rhodium and (R, R)-DIPAMP, the prochiral substrate could be hydrogenated with excellent e.e..12

OH

OH

NH2

CO2H

L-DOPA

OAc

NHCOMe

OMe

HO2C

OAc

NHCOMe

OMe

HO2C

H2, MeOH

Rh(COD)2BF4

96% e.e., 100% e.e after recrystalization

1.1

3 bar, 50oC

sub./cat.>10000

(R, R)-DMAP

Scheme 1.3 Asymmetric hydrogenation in the synthesis of L-DOPA

1.3 Introduction to asymmetric hydrogenation Asymmetric hydrogenation may well be the most powerful and successful method in asymmetric catalysis and a number of examples with high enantiomeric excess have been reported.13 In 1966 Wilkinson and co-workers found that Rh(PPh3)3Cl can be used efficiently as a hydrogenation catalyst in most apolar solvents such as benzene, toluene, DCM, etc.14 Soon, the idea of replacing PPh3 in the Wilkinson catalyst by a chiral phosphine (methylpropylphenylphosphine) was reported independently by the Knowles15 and Horner16 groups. However, the e.e. of only 15% they reported for the hydrogenation of non-functionalized prochiral olefins was very low. Major breakthroughs were realized by the development of a number of chiral bidentate ligands with different backbones such as DIOP by Kagan;17 DIPAMP by Knowles;12 Chiraphos by Bosnich;18 BINAP by Noyori;19 DuPhos and BPE by Burk;20 Ir-P^N L1.1 by Pfaltz21; xyliphos L1.5 (Ar = 3,5-xylyl) by Blaser22 as well as other ligands L1.2-L1.4, L1.6 (Figure 1.1). Their rhodium or ruthenium complexes were successfully applied in asymmetric hydrogenation.9,16-26

P

PPh

Ph MeO

MeO

P

PR

R

R

R

O

O PPh2

PPh2

PPh2

PPh2

Chiraphos (R,R)-DIOP (R,R)-DIPAMP DuPHOS, R =Me, Et

P

P

R

R

R

RPPh2

PPh2

O

NP

t-Bu(o-Tol) (o-Tol)

P(C6H11)2

P(C6H11)2

(R,R)-BPE (R)-BINAP P^N L1.1 L1.223

4

Page 6: University of Groningen Monodentate secondary phosphine ... · Asymmetric catalysis This chapter introduces asymmetric catalysis, especially asymmetric hydrogenation of C=C and C=O

Chapter 1

PAr2

H

PPh2FeOO PPh2

PPh2

H

H

PPh2

PPh2

CHEt2

PPh2

CHEt2

PPh2

Fe

L1.324 L1.425 L1.5 L1.626

Figure 1.1 Examples of bidentate ligands in asymmetric hydrogenation

Numerous prochiral substrates with C=C, C=O bonds can be hydrogenated by the catalysts mentioned above with excellent enantioselectivities, especially the α-dehydroamino acids (esters) 1.2 to 1.3 (Scheme 1.4).27

NHAcPh

COOR

NHAcPh

COOR

H2, chiral catalysts

R = H, Me, Et ee>90% 1.2 1.3

Scheme 1.4 Asymmetric hydrogenation of α-dehydroamino acids and esters 1.2

Ru-based catalysts [mainly for the hydrogenation of ketones (C=O)] are nicely complimentary to Rh-based catalysts [for the hydrogenation of imines (C=N) and functionalized alkenes (C=C)] in asymmetric hydrogenation. Ir-based catalysts are used in the hydrogenation of imines (C=N) and non-functionalized alkenes (see scheme 1.5). Catalysts have been reported that can hydrogenate substrates 1.4,27c and 1.628 selectively at only one C=C or C=O bond to the corresponding products 1.5 and 1.7 with excellent enantio- and chemo-selectivity. Simple N-acetyl enamides 1.8, as mixtures of E/Z isomers can be hydrogenated to 1.9 with excellent e.e.29 The non-functionalized substrate 1.10 can be hydrogenated by Ir-P^N L1.1 complex to alkane 1.11 with excellent e.e. The catalyst based on Ir and L1.5 resulted in very fast imine hydrogenation. 22 With this catalyst, imine 1.12 was successfully hydrogenated to chiral amine 1.13, 22 which is a key intermediate for the synthesis of the herbicide (S)-Metolachlor, with extremely high TON = 2×106, TOF ≥ 2×105 h-1 (Scheme 1.5)

NHAc

CO2Me

NHAc

CO2Me

99% ee, 200:1 chemoselectivity

Rh-(R,R)-Et-DuPhos

6bar H2, MeOH

1.4 1.5

5

Page 7: University of Groningen Monodentate secondary phosphine ... · Asymmetric catalysis This chapter introduces asymmetric catalysis, especially asymmetric hydrogenation of C=C and C=O

Asymmetric catalysis

O O

OMe

OH O

OMe99% ee

H2

RuCl2-(R)-BINAP

1.6 1.7

NHAc

R

NHAc

R

E/Z mixtures 95-97% ee

4 bar H2, i-PrOH, -10oC

Rh-(R,R)-Me-BPE

1.8 1.9

MeO MeO

98% ee

Ir-P^N complex

50 bar H2, RT

1.10 1.11

N

O

ClCH2

OMe

NO

NH

OClCH2COCl

(S)-Metolachlor80% ee

Ir-xyliphos L1.5

H2 80bar, 50oC

1.12 1.13

Scheme 1.5 Examples of asymmetric hydrogenation with difficult substrates Most of these successful bidentate phosphine ligands in asymmetric hydrogenation were C2-symmetric. They are quite difficult to make, frequently requiring long and tedious synthetic routes. Nevertheless they were believed to be the only class of effective ligands for a long time as they confer rigidity to the metal-complex, which was assumed to be a prerequisite for high enantioselectivity. In recent years, some monodentate ligands have proven to be equally effective in asymmetric hydrogenation.30-34 A recent breakthrough is the development of monodentate phosphoramidites for the highly enantioselective hydrogenation of α− or β-dehydro- amino acids 30 and N-acetyl enamides. 31 Other monodentate ligands such as phosphonites,32 phosphites33 and phosphines34 have been developed and performed well in asymmetric hydrogenation (Figure 1.2).

6

Page 8: University of Groningen Monodentate secondary phosphine ... · Asymmetric catalysis This chapter introduces asymmetric catalysis, especially asymmetric hydrogenation of C=C and C=O

Chapter 1

O

OP X R

X=O, (S)-Phosphites

X=C, (S)-Phosphonites

O

OP N

R

R

(S)-Phosphoramidites

R = Me, MonoPhos

P

Ph

PhPhP R

PhospholanePhosphines

Figure 1.2 Structures of monodentate ligands in asymmetric hydrogenation

Compared to bidentate ligands, monodentate phosphoramidites and phosphites ligands can be easily made from relatively cheap BINOL, PCl3 with different nucleophiles such as amines or alcohols in 1-2 steps in high yields.35 As both enantiomers of BINOL are available, the (R)- and (S)-enantiomers of ligands can be made equally well. When they are used as ligands in asymmetric hydrogenation, two enantiomers of the desired products can be obtained simply by changing the configuration of ligands.30, 31 The competition between monodentate and bidentate ligands in asymmetric hydrogenation is still going on. It would seem that the monodentate ligands are now favored for rhodium-catalyzed olefin hydrogenation, 30-34 however, in the hydrogenation of ketones bidentate phosphine ligands give the best results.36

1.4 Important parameters in asymmetric hydrogenation In general, the results of asymmetric hydrogenation are strongly dependent on a variety of parameters, such as solvents, metal precursors, pressure, temperature, substrates and ligands. 7 How these parameters influence the results is often not very clear as they may have opposing effects on different steps of the catalytic cycle, which makes structure activity relationship studies near impossible. (a) Solvent effect One of the reasons why homogeneous catalysts are very active is that they are quite soluble in most organic solvents.37 Alcohols (mainly methanol, 2-propanol), benzene, benzene-alcohol mixtures, toluene, THF, DCM, EtOAc and even aqueous alcohols have been used as solvents. Solvent effects are difficult to predict. Sometimes, the solvent effect is not large, but generally this effect is dramatic. Many examples are known in which the solvent molecules are integral parts of the catalysts.38

(b) Pressure effect The hydrogen pressure is important for the hydrogenation process.39 High pressure means a high concentration of hydrogen in the solution. Since most hydrogenations are first order in hydrogen, an increase of the pressure usually has an accelerating effect. However,

7

Page 9: University of Groningen Monodentate secondary phosphine ... · Asymmetric catalysis This chapter introduces asymmetric catalysis, especially asymmetric hydrogenation of C=C and C=O

Asymmetric catalysis

this process is often accompanied by a decrease of enantioselectivity. This is not always the case. Recently, it was found that in Rh(I)-MonoPhos- catalyzed hydrogenation of prochiral substrates, the reaction rate can be accelerated by increasing hydrogen pressure without loss of any enantioselectivity, which might suggest a different mechanism. (c) Temperature effect The reaction temperature is also an important parameter in the hydrogenation.40-41 In general, the enantioselectivity can be increased by lowering the temperature, which is usually accompanied by a sharp decrease in the reaction rate.40 However, this is not always the case. Several studies show that an increase in temperature may result in increasing enantioselectivities. This is a result of the different kinetic equations for the formation of the two enantiomers which each has a different temperature dependence. In fact the chance could be 50/50 if the temperature effect is positive or negative.41

(d) Substrate effect Although rhodium-phosphine catalyzed hydrogenation is one of the most powerful methods to prepare chiral compounds with high enantioselectivity, the scope of substrates is limited to those possessing a second group that can bind to the metal. Examples are α- or β-dehydroamino acids (esters), itaconic acid (ester) and N-acetyl enamides (Figure 1.3).

COOR

COOR

NHAc

RR1 OR2

NHAc

O

R1 OR2

ONHAc

Figure 1.3 Standard substrates in asymmetric hydrogenation

The configuration of the C=C bond in the substrates is also important, although recent examples show that the substrates with a mixture of E/Z isomers can be hydrogenated in high e.e. (>99%).42 Generally, Z isomers of substrates give higher selectivities than the E isomers. For example, in the Rh-DIPAMP catalyzed hydrogenation of 2-acetyl aminocinnamic acids (esters), Z isomers react much faster (16-100 times) and more selective than the E isomers (94% e.e. and 47%, respectively). One of the reasons why the E isomers are inferior to the Z ones may be their different complexation with the metal center. For example, (E)-2-benzamidocinnamic acid 1.14 is shown to bind with the carboxyl C=O to rhodium, while in the case of Z isomers the amide C=O is coordinated (Figure 1.4).43

P*Rh

P* H

HNH

O PhH

Ph

O OH

NHO

P*Rh

P*

PhH

H

HCO2H

Ph Rh-(E)-1.14 complex Rh-(Z)-1.14 complex

Figure 1.4 E and Z-2-benzamidocinnamic acids 1.14 coordinated to rhodium

8

Page 10: University of Groningen Monodentate secondary phosphine ... · Asymmetric catalysis This chapter introduces asymmetric catalysis, especially asymmetric hydrogenation of C=C and C=O

Chapter 1

With beta-dehydroaminoacid derivatives it is exactly the other way around: The Z is the difficult one because of the intramolecular hydrogen bond.30d

It is known that rhodium-phosphine catalysts can also promote the E→Z isomerization44 of the substrates;45this isomerization process remained unclear until it could be monitored. This was accomplished by Knowles and coworkers who reduced E and Z isomer of 1.14 with D2, creating a second stereogenic center at the C-3 position. The different diastereomers came from the E and Z isomers, which could be identified by D-NMR (Scheme 1.6).46

Rh-DIPAMP

D2D

D NHBzPh

CO2H

NHBz

H CO2H

Ph

D

D NHBzPh

CO2H

NHBzH

CO2HPh D2

Rh-DIPAMP

(Z)-1.14 >90% e.e. (E)-1.14 30% e.e.

Scheme 1.6 D2 experiments of hydrogenation of E- and Z-1.14 In general, it is hard to predict the stereoselectivity in a hydrogenation for a specific substrate. In addition, the reaction conditions need to be optimized separately. For this reason a high throughput screening approach is particularly useful. (e) Ligand effect As already discussed in section 1.3, it is hard to predict the efficiency of monodentate and bidentate ligands in the hydrogenation of an unknown substrate. Many phosphorus ligands are quite sensitive to oxygen; in particular phosphines are readily oxidized to phosphine oxides. On the other hand, monodentate phosphoramidites (MonoPhos),30,31 and phosphit- es30, 33 are not sensitive to oxygen, though depending on their structure, they may react with water. In general, monodentate ligands are much cheaper and easier to prepare and thus are more amenable to tune activity and (enantio-) selectivity than bidentate ones. (f) Mechanism of enantioselective hydrogenation of N-acetyl-dehydroaminoacids Few mechanistic studies have been performed that rationalize the parameters (vide supra) that influence asymmetric hydrogenation.47-48 There are two major studies in the field of rhodium-catalyzed olefin hydrogenation. One was performed by Halpern and co-workers;47 another one by Imamoto has led to the postulation of the so-called “dihydride mechanism”.48 The Halpern mechanism assumes that chiral recognition (enantioselective determining step) takes place in the step of olefin complexation. The major isomer of the diastereomeric substrate-catalyst complexes gives the minor product; whereas the minor one gives the major product (Figure 1.5). The rate-determining step in this mechanism is the oxidative addition of hydrogen to form a dihydride species.

9

Page 11: University of Groningen Monodentate secondary phosphine ... · Asymmetric catalysis This chapter introduces asymmetric catalysis, especially asymmetric hydrogenation of C=C and C=O

Asymmetric catalysis

NHCOMePh

CO2MeNH

MeO

Ph

MeOOCL*

RhL*

NH

MeO

Ph

CO2Me

L*

L*

Rh

NH

MeO

Ph

MeOOCH

Rh

L*

HL*

NH

MeO

Ph

CO2Me

L*

HL*

H

Rh

NHO

Me

CO2Me

Ph

L*

L*H

RhS' NH

OMe

Ph

L*

L*

MeOOC

H

S'Rh

L*

S'L*

Rh

S'

O

NH

MeH

Ph

CO2Me

O

NH

MeH

Ph

MeOOCmajor productminor product

H2

S'

H2

S'

"minor"manifold

"major"manifold

(R) (S)

Figure 1.5 Halpern mechanism of asymmetric hydrogenation In the dihydride mechanism, the oxidative addition of H2 to form a dihydride intermediate is the enantioselectivity determining step. According to this mechanism, the major isomer of the rhodium-alkane species gives the major product and the minor one gives the minor product (Figure 1.6). The rate-determining step in this mechanism is the olefin coordination to the dihydride species.

NHCOMePh

CO2Me

NH

MeO

Ph

MeOOCH

Rh

L*

HL*

NHO

Me

CO2Me

Ph

L*

L*H

RhS'

O

NH

MeH

Ph

CO2Me

O

NH

MeH

Ph

MeOOC

H

RhS'

L*

H

S'

L*MeNHCO Ph

MeOOC

NH

MeO

Ph

CO2MeH

Rh

L*

H L*

NHO

Me

Ph

L*

L* H

RhS'

MeOOC

L*

S' L*

Rh

S'

H

RhS'

L*

H

S'

L*

minor productmajor product

H2

S'

"minor"manifold

"major"manifold

(R) (S)

S'

S'

Figure 1.6 Dihydride mechanism of asymmetric hydrogenation In conclusion, asymmetric hydrogenation is a very useful, powerful and important method in asymmetric synthesis. In total synthesis, the number of synthetic steps can often be

10

Page 12: University of Groningen Monodentate secondary phosphine ... · Asymmetric catalysis This chapter introduces asymmetric catalysis, especially asymmetric hydrogenation of C=C and C=O

Chapter 1

reduced by preparing the key intermediates using asymmetric catalysis. A main challenge for the future is the preparation of cheap and effective (chiral) catalysts that can meet the economic demands for industrial application. 1.5 References and notes 1 (a). Series of “Organic Synthesis Collective”, Vol. 1 (1967) to 80 (2003), John Wiley & Sons, New York. (b). Series of “Organic Reactions Collective”, Vol. 1 (1942) to 63 (2004), Wiley, New York. (c). Barton, D.; Ollis, W. D. eds., Comprehensive Organic Chemistry, Vol. 1-6, Pergamon Press, Oxford, 1979. (d). Trost, B. M.; Fleming, I. Comprehensive Organic Synthesis, Vol. 1-9, Pregamon Press, Oxford, 1991. (e). Carey, F. A.; Sundberg, R. J. eds., Advanced Organic Chemistry, Part A & B, Kluwer Academic/Plenum Publishers, New York, 2000. (f). Warren, S. ed., Organic Synthesis-The Disconnection Approach, Wiley, 1982. 2 (a). Trost, B. M. Science 1991, 254, 1471. (b). Trost, B. M. Acc. Chem. Res. 2002, 35, 695. (c). Trost, B. M. Angew. Chem., Int. Ed. 1995, 34, 259. (d). Trost, B. M.; Toste, F. D.; Greenman, K. J. Am. Chem. Soc. 2003, 125, 4518. (e). Trost, B. M.; Jonasson, C.; Wuchrer, M. J. Am. Chem. Soc. 2001, 123, 12736. (f). Trost, B. M.; Oi, S. J. Am. Chem. Soc. 2001, 123, 1230. (g). Trost, B. M.; Shi, Y. J. Am. Chem. Soc. 1992, 114, 791. 3 (a). Eliel, E. L.; Wilen, S. H.; Mander, L. N. eds., Stereochemistry of Organic Compounds, Wiley, New York, 1994. (b). Atkinson, R. S. ed., Stereoselective Synthesis, Wiley, 1995. (c). Kagan, H. B. ed., Stereochemistry: Fundamentals and Methods, Vol. 1-4, Georg Thieme, Stuttgart, 1977. (d). Vogtle, F.; Weber, E. eds., Stereochemistry, Springer-Verlag, Berlin, 1984. (e). Nasipuri, D. ed., Stereochemistry of Organic Compounds: Principles and Applications, Wiley, New York, 1991. (f). Robinson, M. J. T. ed., Organic Stereochemistry, Oxford University Press, Oxford, 2000. 4 A review of different biological activities of chiral compounds, see: Bertil, W. Chirality 1993, 5, 350. Other examples, see: (a). Maliepaard, M.; Groot, S. E.; de Mol, N. J.; Janssen, L. H. M.; Freiks, M.; Verboon, W.; Reinhoudt, D. N.; Stephens, M.; Stratford, I. J. Anti-cancer Drug Design 1996, 11, 403. (b). Chae, K.; Gibson, M. K.; Korach, K. S. Mol. Pharm. 1991, 40, 806. (c). Korach, K. S.; Chae, K.; Levy, L. A.; Duax, W. L.; Sarver, P. J. J. Bio. Chem. 1989, 264, 5642. (d). Attwood, M. R.; Brown, B. S.; Dunsdon, R. M.; Hurst, D. N.; Jones, P. S.; Kay, P. B. Bioorg. & Med. Chem. Lett. 1992, 2, 229. 5 (a). Friedman, L.; Miller, J. G. Science 1971, 172, 1044 (b). Solms, J.; Vuataz, L.; Egli, R. H. Experientia 1965, 21, 692. 6 (a). Collins, A. N.; Sheldrake, G. N.; Crosby, J. eds., Chirality in Industry, 1992; Chirality in Industry II: Developments in the Commercial Manufacture of Optically- Active Compounds, Wiley, Chichester, 1996 and 1997. (b). Aronson, J. K. eds., Side Effects of Drugs, Elsevier, Amsterdam, 1997. (c). Makins, R.; Ballingger, A. Expert Opinion on Drug Safety 2003, 2, 421. (d). Keller, T. H.; Bray-French, K.; Demnitz, F. W. J.; Muller, T.; Pombo-villar, E.; Walker, C. Chem. Pharm. Bull. 2001, 49, 1009. (e). Eriksen, J. L.; Sagi, S. A.; Smith, T. E.; Weggen, S.; Das, P.; McLendon, D. C.; Ozolos, V.

11

Page 13: University of Groningen Monodentate secondary phosphine ... · Asymmetric catalysis This chapter introduces asymmetric catalysis, especially asymmetric hydrogenation of C=C and C=O

Asymmetric catalysis

V.; Jessing, K. W.; Golde, T. E. J. Clini. Invest. 2003, 112, 440. (f). Patent, Li, Z.; Lu, X.; Liao, C.; Shi, L.; Liu, Z.; Ma, B.; Ning, Z.; Shan, S.; Deng, T. WO2004048333 A1, 2004. 7 (a). Nogradi, M. ed., Stereoselective Synthesis, A Practical Approach, 2nd edition, VCH, 1995. (b). Finn, M. G.; Sharpless, K. B. eds., Asymmetric Synthesis, Academic Press, New York, 1985. (c). Coppola, G. M.; Schuster, H. F. eds., Asymmetric Synthesis (Construction of Chiral Molecules using Amino Acids), Wiley, 1987. 8 (a). Atiken, R. A.; Kilenyi, S. N. eds., Asymmetric Synthesis, Blackie Academic & Professional, Chapman & Hall, 1992. (b). Morrison, J. D.; Mosher, H. S. eds., Asymmetric Organic Reactions, Prentice-Hall, Inc., Englewood Cliffs, New Jersey, 1971. 9 (a). Jacoben, E. N.; Pfaltz, A.; Yamamoto, H. eds., Comprehensive Asymmetric Catalysis, I-III, Springer, 1999. (b). Ojima, I. ed., Catalytic Asymmetric Synthesis, VCH, New York, 1993. 10 (a). Helmchen, G.; Hoffmann, R. W.; Mulzer, J.; Schaumann, E. eds., Methods of Organic Chemistry, Volume E 21a, Stereoselective Synthesis, Thieme, Stuttgart, 1995, p 75. (b). Morrison, J. D. ed., Asymmetric Synthesis, 5 volumes, Academic Press, Orlando, 1983-1985. 11 For recent examples of asymmetric synthesis in industry, see: (a). Sheldon, R. A.; Blaser, H. –U. eds., Fine Chemicals Issue, Wiley-VCH, Germany, 2003. (b). Hawkins, J. M.; Watson, T. J. N. Angew. Chem. Int. Ed. 2004, 43, 3224. (c). Stuermer, R. Angew. Chem. Int. Ed. 2004, 43, 2600. (d). Spargo, P. Org. Process Res. Dev. 2004, 8, 298. (e). Genet, J. –P. Acc. Chem. Res. 2003, 36, 908. (f). Calvisi, G.; Dell, N.; De Angelis, F.; Dejas, R.; Giannessi, F.; Tinti, M. O. Eur. J. Org. Chem. 2003, 23, 4501. (g). Sheldon, R. A. ed., Chirotechnology: Industrial Synthesis of Optically Active Compounds, Dekker, New York, 1993. 12 Vineyard, B. D.; Knowles, W. S.; Sabacky, M. J.; Bachman, G. L.; Weinkauff, D. J. J. Am. Chem. Soc. 1977, 99, 5946. 13 For a recent review about asymmetric hydrogenation, see: Tang, W.; Zhang, X. Chem. Rev. 2003, 103, 3029. 14 Osborne, J. A.; Jardine, F. H.; Young, J. F.; Wilkinson, G. J. Chem. Soc. A 1966, 1711. 15 Knowles, W. S.; Sabacky, M. J. J. Chem. Soc., Chem. Comm. 1968, 1445. 16 Horner, L.; Siegel, H.; Buthe, H. Angew. Chem. Int. Ed. 1968, 7, 942. 17 (a). Kagan, H. B.; Dang, T.-P. J. Chem. Soc., Chem. Comm. 1971, 481. (b). Dang, T.-P.; Kagan, H. B. J. Am. Chem. Soc. 1972, 94, 6429. 18 Fryzuck, M. D.; Bosnich, B. J. Am. Chem. Soc. 1977, 99, 6262. 19 (a). Takaya, H.; Ohta, T.; Sayo, N.; Kumobayashi, H.; Akutagawa, S.; Inoue, S.; Kasahara, I.; Noyori, R. J. Am. Chem. Soc. 1987, 109, 1596. (b). Kitamura, M.; Ohkuma, T.; Inoue, S.; Sayo, N.; Kumobayashi, H.; Akutagawa, S.; Ohta, T.; Takaya, H.; Noyori, R. J. Am. Chem. Soc. 1988, 110, 629. (c). Noyori, R.; Ohta, M.; Hsiao, Y.; Kitamura, M.; Ohta, T.; Rakaya, H. J. Am. Chem. Soc. 1986, 108, 7117. (d). Kitamura, M.; Tokunaga, M.; Noyori, R. J. Org. Chem. 1992, 57, 4053. (e). Kitamura, M.; Tokunaga, M.; Ohkuma, T.; Noyori, R. Tetrahedron Lett. 1991, 32, 4163. (f). Shao, L.; Seki, T.; Kawano, H.; Saburi,

12

Page 14: University of Groningen Monodentate secondary phosphine ... · Asymmetric catalysis This chapter introduces asymmetric catalysis, especially asymmetric hydrogenation of C=C and C=O

Chapter 1

M. Tetrahedron Lett. 1991, 32, 7699. (g). Heiser, B.; Broger, E. A.; Crameri, Y. Tetrahedron: Asymm. 1991, 2, 51. 20 Burk, M. J.; Gross, M. F.; Martinez, J. P. J. Am. Chem. Soc. 1995, 117, 9375. 21 (a). Schnider, P.; Koch, G.; Pretôt, R.; Wang, G.; Bohnen, F. M.; Kruger, C.; Pfaltz, A. Chem. Eur. J. 1997, 3, 887. (b). Lightfoot, A.; Schnider, P.; Pfaltz, A. Angew. Chem. Int. Ed. 1998, 37, 2897. 22 (a). Bader, R.R.; Baumeister, P.; Blaser, H.-U. Chimia 1996, 50, 99. (b). Spindler, F.; Pugin, B.; Jalett, H.-P.; Buser, H.-P.; Pittelkow, U.; Blaser, H.-U. in Catalysis of Organic Reactions, Malz, R.E. Jr. ed., Chemical Industries 68, 1996, p 153. (c). Spindler, F.; Blaser, H. -U. Adv. Synth. Catal. 2001, 343, 68. 23 (a). Miyashita, A.; Karino, H.; Shimamura, J.; Chiba, T.; Nagano, K.; Nohira, H.; Takaya, H. Chem. Lett. 1989, 1849. (b). Chiba, T.; Miyashita, A.; Nohira, H.; Takaya, H. Tetrahedron Lett. 1991, 32, 4745. 24 Chan, A.; Hu, W. H.; Pai, C. C.; Lau, C. P.; Jiang, Y. Z.; Mi, A. Q.; Yan, M.; Sun, J.; Lou, R. L.; Deng, J. G. J. Am. Chem. Soc. 1997, 119, 9570. 25 Zhu, G. X.; Cao, P.; Jiang, Q. Z.; Zhang, X. M. J. Am. Chem. Soc. 1997, 119, 1799. 26 Kang, J.; Lee, J. H.; Ahn, S. H.; Choi, J. S. Tetrahedron Lett. 1998, 39, 5523. 27 (a). Travins, J. M.; Etzkorn, F. A. J. Org. Chem. 1997, 62, 8387. (b). Doi, T.; Kokubo, M.;Yamamoto, K.; Takahashi, T. J. Org. Chem. 1998, 63, 428. (c). Burk, M. J.; Allen, J. G.; Kiesman, W. F. J. Am. Chem. Soc. 1998, 120, 657. (d). Rossen, K.; Weissman, S. A.; Sager, J.; Reamer, R. A.; Askin, D.; Volante, R. P.; Reider, P. J. Tetrahedron Lett. 1995, 36, 6419. (e). Rajanbabu, T. V.; Ayers, T. A.; Halliday, G. A.; You, K. K.; Calabrese, J. C. J. Org. Chem. 1997, 62, 6012. (f). Corey, E. J.; Link, J. O. Tetrahedron Lett. 1992, 33, 4141. 28 Noyori, R.; Ohkuma, T.; Kitamura, M.; Takaya, H.; Sayo, N.; Kumobayashi, H. ; Akutagawa, S. J. Am. Chem. Soc. 1987, 109, 5856. 29 Burk, M. J.; Wang, Y. M.; Lee, J. R. J. Am. Chem. Soc. 1996, 118, 5142. 30 (a). van den Berg, M.; Minnaard, A. J.; Schudde, E. P.; van Esch, J.; de Vries, A. H. M.; de Vries, J. G.; Feringa, B. L. J. Am. Chem. Soc. 2000, 122, 11539. (b). van den Berg, M.; Minnaard, A. J.; de Vries, J. G.; Feringa, B. L. (DSM N. V.), WO 02/04466, 2002. (c). van den Berg, M.; Minnaard, A. J.; Haak, R. M.; Leeman, M.; Schudde, E. P.; Meetsma, A.; Feringa, B. L.; de Vries, A. H. M.; Maljaars, C. E. P.; Willans, C. E.; Hyett, D.; Boogers, J. A. F.; Henderickx, H. J. W.; de Vries, J. G. Adv. Synth. Catal. 2003, 345, 308. (d). Peña, D.; Minnaard, A. J.; de Vries, J. G.; Feringa, B. L. J. Am. Chem. Soc. 2002, 124, 14552. (e). Zeng, Q.; Liu, H.; Mi, A.; Jiang, Y.; Li, X.; Choi, M. C. K.; Chan, A. S. C. Tetrahedron 2002, 58, 8799. 31 (a). van den Berg, M.; Haak, R. M.; Minnaard, A. J.; de Vries, A. H. M.; de Vries, J. G.; Feringa, B. L. Adv. Synth. Catal. 2002, 344, 1003. (b). Jia, X.; Guo, R.; Li, X.; Yao, X.; Chan, A. S. C. Tetrahedron Lett. 2002, 43, 5541. 32 Claver, C.; Fernandez, E.; Gillon, A.; Heslop, K.; Hyett, D. J.; Martorell, A.; Orpen, A. G.; Pringle, P. G. Chem. Commun. 2000, 961. 33 Reetz, M. T.; Mehler, G. Angew. Chem. Int. Ed. 2000, 39, 3889.

13

Page 15: University of Groningen Monodentate secondary phosphine ... · Asymmetric catalysis This chapter introduces asymmetric catalysis, especially asymmetric hydrogenation of C=C and C=O

Asymmetric catalysis

34 (a). Guillen, F.; Fiaud, J.-C. Tetrahedron Lett. 1999, 40, 2939. (b). Guillen, F.; Rivard, M.; Toffano, M.; Legros, J. -Y.; Daran, J. -C.; Fiaud, J. -C. Tetrahedron 2002, 58, 5895. (c). Junge, K.; Oehme, G.; Monsees, A.; Riermeier, T.; Dingerdissen, U.; Beller, M. Tetrahedron Lett. 2002, 43, 4977. 35 For MonoPhos, 1 g, € 53 from Strem, 2003. 36 Ohkuma, T.; Noyori, R. in Jacobsen, E. N.; Pfaltz, A.; Yamamoto, H. eds., Comprehensive Asymmetric Catalysis I-III, Vol I, chap. 6.1, p 199, Springer, 1999. 37 (a). James, B. R. in Wilkinson, G.; Stone, F. G. A.; Abel, E. W. eds., Comprehensive Organometallic Chemistry, Pergamon Press, Oxford, 1982. (b). James, B. R. ed., Homogeneous Hydrogenation, Wiley Interscience, New York, 1974. 38 (a). Longley, C. J.; Goodwin, T. J.; Wilkinson, G. Polyhedron 1986, 5, 1625. (b). Breuer, D.; Goen, T.; Haupt, H. J. J. Mol. Catal. 1990, 61, 149. (c). Matteoli, U.; Menchi, G.; Bianchi, M.; Piacenti, F. J. Organomet. Chem. 1986, 299, 233. (d). Strohmeier, W.; Hitzel, E. J. Organomet. Chem. 1975, 87, 257. (e). Horner, L.; Buethe, H.; Siegel, H. Tetrahedron Lett. 1968, 37, 4023. (f). James, B. R. Catalysis Today 1997, 37, 209. (g). Becalski, A. G.; Cullen, W. R.; Fryzuk, M. D.; James, B. R.; Kang, G.-J.; Rettig, S. J. Inorg. Chem. 1991, 30, 5002. (h). Cheong, C. Y. N.; Osborn, J. A. J. Am. Chem. Soc. 1990, 112, 9400. 39 (a). Ojima, I.; Kogure, T.; Yoda, N. J. Org. Chem. 1980, 45, 4728. (b). Ojima, I.; Kogure, T. Chem. Lett. 1979, 495. (c). Li, R.-X.; Cheng, P.-M.; Li, D.-W.; Chen, H.; Li, X.-J.; Wessman, C.; Wong, N.-B.; Tin, K.-C. J. Mol. Catal. A 2000, 159, 179. (d). Sun, Y.; Landau, R. N.; Wang, J.; LeBlond, C.; Blackmond, D. G. J. Am. Chem. Soc. 1996, 118, 1348. (e). Sinou, D. Tetrahedron Lett. 1981, 2987. 40 (a). Cullen, W. R.; Sugi, Y. Tetrahedeon Lett. 1978, 1635. (b). Uehara, A.; Kubota, T. ; Tsuchiya, R. Chem. Lett., 1983, 441 (c). Halpern, J. Pure Appl. Chem. 1983, 55, 99. (d). Landis, C. R.; Halpern, J. J. Am. Chem. Soc. 1987, 109, 1746. (e). Halpern, J. Science 1982, 217, 401. (f). Riley, D. P.; Shumate, R. E. J. Org. Chem. 1980, 45, 5187. (g). Heller, D.; Buschmann, H.; Scharf, H.-D. Angew. Chem. Int. Ed. 1996, 35, 1852. 41 (a). Sun, Y.; Wang, J.; LeBlond, C.; Landau, R. N.; Laquidara, J.; Sowa, J. R.; Blackmond, D. G. J. Mol. Catal. A 1997, 115, 495. (b). Sun, Y.; Wang, J.; LeBlond, C.; Reamer, R. A.; Laquidara, J.; Sowa, J. R.; Blackmond, D. G. J. Organomet. Chem. 1997, 548, 65. 42 (a). Tang, W.; Zhang, X. Org. Lett. 2002, 4, 4159. (b). Zhou, Y.; Tang, W.; Wang, W.; Li, W.; Zhang, X. J. Am. Chem. Soc. 2002, 124, 4952. 43 (a). Brown, J. M.; Chaloner, P. A. J. Chem. Soc., Chem. Comm. 1979, 613. (b). Brown, J. M.; Chaloner, P. A. Tetrahedron Lett. 1978, 21, 1877. 44 For general reviews about rhodium catalyzed E/Z isomerization of olefins, see: (a). Davies, N. R. Rev. of Pure and Appl. Chem. 1967, 17, 83. (b). Akutagawa, S. in Jacobsen, E. N.; Pfaltz, A.; Yamamoto, H. eds., Comprehensive Asymmetric Catalysis I-III, Vol II, chap. 23, p 813, Springer, 1999. 45 (a). Breit, B.; Fuchs, E. Chem. Commun. 2004, 6, 694. (b). Morrill, T. C.; D’Souza, C. A. Organometallics 2003, 22, 1626. (c). Yoshihiro, S.; Yoshihiro, O.; Miwako, M.

14

Page 16: University of Groningen Monodentate secondary phosphine ... · Asymmetric catalysis This chapter introduces asymmetric catalysis, especially asymmetric hydrogenation of C=C and C=O

Chapter 1

Organometallics 2003, 22, 30. (d). Tan, K. L.; Bergman, R. G.; Ellman, J. A. J. Am. Chem. Soc. 2002, 124, 13964. (e). Buriak, J. M.; Klein, J. C.; Herrington, D. G.; Osborn, J. A. Chem. Eru. J. 2000, 6, 139. (f). Yu, J.; Spencer, J. B. J. Am. Chem. Soc. 1997, 119, 5257. (g). Moasser, B.; Gladfelter, W. L.; Roe, D. C. Organometallics 1995, 14, 3832. (h). Trzeciak, A. M.; Szterenberg, L.; Wolszczak, E.; Ziolkowski, J. J. J. Mol. Catal. A 1995, 99, 23. (i). Matsuda, I.; Kato, T.; Sato, S.; Izumi, Y. Tetrahedron Lett. 1986, 27, 5747. 46 Koenig, K. E.; Knowles, W. S. J. Am. Chem. Soc. 1978, 100, 7561. 47 (a). Chua, P. S.; Roberts, N. K.; Bosnich, B.; Okrasinski, S. J.; Halpern, J. J. Chem. Soc., Chem. Comm. 1981, 1278. (b). Halpern, J.; Riley, D. P.; Chan, A. S. C.; Pluth, J. J. J. Am. Chem. Soc. 1977, 99, 8055. (c). Chan, A. S. C.; Halpern, J. J. Am. Chem. Soc. 1980, 102, 838. (d). Chan, A. S. C.; Pluth, J. J.; Halpern, J. Inorg. Chim. Acta. 1979, 37, 2477. 48 (a). Gridnev, I. D.; Higashi, N.; Asakura, K.; Imamoto, T. J. Am. Chem. Soc. 2000, 122, 7183. (b). Gridnev, I. D.; Yamanoi, Y.; Higashi, N.; Tsuruta, H.; Yasutake, M.; Imamoto, T. Adv. Synth. Catal. 2001, 343, 118.

15