29
University of Groningen Novel asymmetric copper-catalysed transformations Bos, Pieter Harm 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: 2012 Link to publication in University of Groningen/UMCG research database Citation for published version (APA): Bos, P. H. (2012). Novel asymmetric copper-catalysed transformations. 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: 04-07-2020

University of Groningen Novel asymmetric copper-catalysed ... · 14 Chapter2-Final-nocode.docx Chapter 2 (Scheme 2). In this case simple organolithium reagents gave the best results

  • Upload
    others

  • View
    1

  • Download
    0

Embed Size (px)

Citation preview

Page 1: University of Groningen Novel asymmetric copper-catalysed ... · 14 Chapter2-Final-nocode.docx Chapter 2 (Scheme 2). In this case simple organolithium reagents gave the best results

University of Groningen

Novel asymmetric copper-catalysed transformationsBos, Pieter Harm

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:2012

Link to publication in University of Groningen/UMCG research database

Citation for published version (APA):Bos, P. H. (2012). Novel asymmetric copper-catalysed transformations. 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: 04-07-2020

Page 2: University of Groningen Novel asymmetric copper-catalysed ... · 14 Chapter2-Final-nocode.docx Chapter 2 (Scheme 2). In this case simple organolithium reagents gave the best results

Chapter 2 Catalytic Asymmetric Conjugate Addition of Grignard Reagents to α,β-Unsaturated Sulfones

In this chapter a highly efficient method is reported for the asymmetric conjugate addition of Grignard reagents to ,-unsaturated 2-pyridylsulfones. Using a Cu/TolBinap complex, excellent enantioselectivities and high yields are obtained for a wide variety of aliphatic substrates.*

* Parts of this chapter have been published: Bos, P. H.; Minnaard, A. J.; Feringa, B.L. Org. Lett. 2008, 10, 4219.

Page 3: University of Groningen Novel asymmetric copper-catalysed ... · 14 Chapter2-Final-nocode.docx Chapter 2 (Scheme 2). In this case simple organolithium reagents gave the best results

12

Chapter2-Final-nocode.docx

Chapter 2

2.1 Introduction

As already described in the introductory chapter of this thesis, the conjugate addition of

organometallic reagents to ,-unsaturated compounds is one of the most versatile methods for the formation of C-C bonds.1, 2 This transformation is used as a key step in the synthesis of numerous natural products and biologically active compounds and has been the subject of intensive research over the past decades.3-12 The development of a catalytic method for

the enantioselective conjugate addition reaction of organometallic reagents to ,-unsaturated sulfones is an important goal in extending the current methodology.

2.1.1 The use of sulfones in organic chemistry

The utility of sulfones for organic synthesis was recognized in the late 1970’s13 and because of their duality of functioning both as nucleophiles in basic media and electrophiles in Lewis acidic media they have been dubbed “chemical cameleons”.14 Sulfonyl-containing intermediates have frequently been used in the total synthesis of a large number of biologically active natural compounds.15 As a result, methods for their synthesis have been well developed.13, 16, 17

Sulfones bearing a stereocenter at the -position are highly versatile intermediates in organic chemistry due to the ease of derivatization and by providing access to a wide range of building blocks, including aldehydes and ketones, alkynes, alkenes, alkanes, and haloalkanes.15, 18 This versatility was nicely demonstrated by Carretero et al. in an article

describing the catalytic asymmetric conjugate reduction of ,-disubstituted ,-

unsaturated sulfones (see also section 2.1.3).19 The resulting -substituted highly enantioenriched sulfones were converted into four differently functionalized chiral compounds without compromising the enantiomeric excess (see Scheme 1).

Page 4: University of Groningen Novel asymmetric copper-catalysed ... · 14 Chapter2-Final-nocode.docx Chapter 2 (Scheme 2). In this case simple organolithium reagents gave the best results

13

Chapter2-Final-nocode.docx

Catalytic Asymmetric Conjugate Addition of Grignard Reagents to α,β-Unsaturated Sulfones

Me

Ph

SO2Py

(S)-1 (94% ee)originating from

conjugate reduction

Me

Ph

Me

Ph

Me

Ph

Ph

Ph

OEt

O

Me

Ph

Ph

O

KHMDS, DME

-78 oC

then PhCHO(R)-5, 87%

1. n-BuLi, then BnBr2. Na(Hg), Na2HPO4, MeOH

(R)-2, 75%

1. n-BuLi, then ClCO2Et

2. Zn, aq. NH4Cl(R)-3, 80%

1. n-BuLi, then ClCOPh2. Zn, aq. NH4Cl

(R)-4, 69% Scheme 1 Examples of synthetic applications of chiral enantioenriched 2-pyridylsulfones. HMDS: hexamethyldisilazide; DME: 1,2-dimethoxyethane.19

The first three transformations (Scheme 1, compounds (R)-2, (R)-3 and (R)-4) are based on the generation of the highly nucleophilic sulfonyl carbanion followed by carbon-carbon bond formation by reaction with an appropriate carbon electrophile (benzyl bromide, ethyl chloroformate, or benzoyl chloride respectively). After subsequent desulfonylation the products were isolated in good yields without compromising the enantiomeric excess. In the last example, Julia-Kocienski olefination of (S)-1 with benzaldehyde afforded alkene (R)-5 directly in 87% yield, with complete selectivity for the E isomer and no loss of enantiomeric excess was observed. The Julia-Kocienski olefination occurs without racemization at the allylic stereogenic center.20

2.1.2 Conjugate addition of organometallic reagents to ,-unsaturated sulfones

A vast number of diastereoselective conjugate addition reactions to ,-unsaturated sulfones have been reported.21, 22 Groundbreaking work of Fuchs et al. is especially

noteworthy.21 Using ,-unsaturated sulfones as substrates together with either organolithium reagents or organocuprates interesting molecular structures were synthesized. In the first example a number of functionalized cyclooctane structures were built up using the addition of organometallic reagents to cyclooctenyl phenyl sulfone 6

Page 5: University of Groningen Novel asymmetric copper-catalysed ... · 14 Chapter2-Final-nocode.docx Chapter 2 (Scheme 2). In this case simple organolithium reagents gave the best results

14

Chapter2-Final-nocode.docx

Chapter 2

(Scheme 2). In this case simple organolithium reagents gave the best results giving predominantly the syn diastereomers (ratio 7a:7b, up to 100:0) with yields up to 93%.23

Scheme 2 Conjugate addition (6) and allylic alkylation (8) of organometallic reagents.23

Allylic alkylation of methyllithium to epoxy cyclooctenyl phenyl sulfone 8 gave the product (9) in 90% isolated yield. Unfortunately, the authors could not determine the syn:anti ratio in this case. The total synthesis of (+)-carbacyclin is another example in which the utility of the

,-unsaturated sulfone group is demonstrated elegantly.24 In this case the presence of the

,-unsaturated sulfone group allows for the regio- and stereoselective introduction of carbon substituents onto a preformed ring (Scheme 3).

Scheme 3 Total synthesis of (+)-carbacyclin 15.24

Page 6: University of Groningen Novel asymmetric copper-catalysed ... · 14 Chapter2-Final-nocode.docx Chapter 2 (Scheme 2). In this case simple organolithium reagents gave the best results

15

Chapter2-Final-nocode.docx

Catalytic Asymmetric Conjugate Addition of Grignard Reagents to α,β-Unsaturated Sulfones

Reaction of bromocuprate 11 with optically active 10 afforded the allylic alkylation product. This compound was subsequently converted into 12 in two steps. Conjugate addition of chiral vinyllithium reagent 13 to 12 and subsequent intramolecular trapping of the resulting carbanion followed by desilylation with TBAF provided 14. Treatment with lithium in liquid ammonia afforded the desulfonylated, debenzylated product and completion of the synthesis was achieved by selective oxidation of the primary alcohol to give (+)-carbacyclin 15.24 Another method reported in literature is based on a stereoselective conjugate addition

of methyllithium to enantiomerically pure γ-alkoxy-,-unsaturated phenyl sulfone 16 for the stereoselective construction of polypropionate chains using an iterative approach (Scheme 4).25

Scheme 4 Iterative construction of polypropionate chains.25

After the conjugate addition the syn isomer 17 was isolated exclusively. By a three step

protocol, ,-unsaturated phenyl sulfone 18 can be generated, which could be employed as a substrate in a subsequent conjugate addition reaction with methyllithium. In this way polypropionate segments with up to four consecutive stereocenters were constructed.

2.1.3 Catalytic asymmetric conjugate reduction of ,-disubstituted ,-unsaturated sulfones

A complementary approach to the catalytic asymmetric conjugate addition is the catalytic

asymmetric conjugate reduction of ,-disubstituted Michael acceptors. This method is a useful and practical alternative for the preparation of enantioenriched carbonyl compounds

and related systems bearing a stereocenter at the -position. Pioneering work on the copper

hydride catalyzed asymmetric conjugate reduction of ,-disubstituted ,-unsaturated esters by Buchwald et al. led to the development of a vast number of conjugate reduction procedures using various Michael acceptors.26-36 Remarkably, despite the great chemical versatility of sulfones in synthesis37, the

catalytic asymmetric conjugate reduction of ,-disubstituted ,-unsaturated sulfones was only developed recently. In 2007, the group of Carretero et al. reported the catalytic

asymmetric conjugate reduction of ,-unsaturated 2-pyridylsulfones using PhSiH3 as the hydride source and CuCl/t-BuONa/(R)-Binap as the chiral catalytic system (see Scheme 5).19 This methodology has quite a broad scope with regard to the substitution of the vinyl

Page 7: University of Groningen Novel asymmetric copper-catalysed ... · 14 Chapter2-Final-nocode.docx Chapter 2 (Scheme 2). In this case simple organolithium reagents gave the best results

16

Chapter2-Final-nocode.docx

Chapter 2

Scheme 5 Enantioselective conjugate reduction developed by Carretero et al.19

sulfone 19 and the resulting -substituted 2-pyridylsulfones were obtained with high enantiomeric excess and in excellent isolated yields. The authors noted that the use of the 2-pyridylsulfonyl group in the substrate was absolutely necessary in order to get satisfactory results. If an ordinary phenylsulfonyl group was used no conjugate reduction was observed under the reaction conditions. This effect was also noted for the rhodium-catalyzed

conjugate addition of boronic acids to ,-unsaturated sulfones (see section 2.1.4) and it is believed that the possible coordination between copper and the 2-pyridylsulfone group can result in a strong rate acceleration in the conjugate reduction reaction.38, 39 In the same year an extension to this methodology was published by Charette et al. circumventing the necessity of the 2-pyridylsulfonyl group.40 In their paper a general procedure is reported for the enantioselective reduction of simple vinyl phenyl sulfones catalyzed by a copper-phosphine complex (see Scheme 6).

Scheme 6 Enantioselective conjugate reduction developed by Charette et al.40

Absolutely crucial to the success of this procedure is the use of the hemilabile bidentate ligand Me-DuPhos monoxide L1. The addition of aqueous sodium hydroxide was necessary to obtain reproducible conversions. With the optimized procedure in hand a variety of vinyl phenyl sulfones were shown to give the desired product 22 with excellent yield and enantiomeric excess. It must be noted that when using an aliphatic acyclic substrate a bulkier ligand had to be used in order to reach a satisfactory enantiomeric excess.

Page 8: University of Groningen Novel asymmetric copper-catalysed ... · 14 Chapter2-Final-nocode.docx Chapter 2 (Scheme 2). In this case simple organolithium reagents gave the best results

17

Chapter2-Final-nocode.docx

Catalytic Asymmetric Conjugate Addition of Grignard Reagents to α,β-Unsaturated Sulfones

2.1.4 Rhodium-catalyzed asymmetric conjugate addition of organoboronic acids to ,-unsaturated sulfones

In 2004, Carretero et al. developed a general method for the catalytic asymmetric conjugate

addition of arylboronic acids to acyclic ,-unsaturated sulfones.41 As a model reaction the behavior of a variety of propenyl sulfones 23, with a different substitution pattern at the sulfur atom, were evaluated using the standard conditions described for the rhodium-catalyzed conjugate addition of phenylboronic acids to enones (Table 1).42 Table 1 Influence of sulfone substitution on the Rhodium-catalyzed conjugate addition.41

23 R Conversion (%)a

23a-e

, , , ,

<2

23f

>98 (74%)

23g

>98 (98%)

23h N

<2

a Isolated yield in parentheses.

Not only was phenyl sulfone 23a inert under these reaction conditions, the same was also observed for sulfones with a 2-(dimethylamino)phenyl, 1,3-pyrimidinyl, tetrazoyl and benzimidazoyl moiety (23b-e). Both imidazoyl (23f) as well as 2-pyridyl (23g) sulfones gave the desired product, but 23g gave the highest reactivity and isolated yield. The high reactivity of the 2-pyridyl substituted sulfone was attributed to a coordination effect of the rhodium catalyst with the appropriately placed nitrogen atom.43 This hypothesis was further supported by studying the behavior of the 4-pyridyl isomer 23h. In this case an intramolecular chelation with the rhodium catalyst is impossible and this is reflected in the lack of conversion.

Page 9: University of Groningen Novel asymmetric copper-catalysed ... · 14 Chapter2-Final-nocode.docx Chapter 2 (Scheme 2). In this case simple organolithium reagents gave the best results

18

Chapter2-Final-nocode.docx

Chapter 2

After establishing the crucial role of the 2-pyridyl sulfone group different chiral ligands were screened for the enantioselective addition of arylboronic acids to 23g. Using Chiraphos as the chiral ligand, the desired products could be isolated in excellent yields (up to 98%) with good to excellent enantiomeric excess (77-92%) (see Scheme 7). The major drawback of this method is that it is limited to the introduction of aryl groups.

Scheme 7 Catalytic asymmetric conjugate addition of arylboronic acids.41

The same catalytic system was also used for the formation of all-carbon quaternary centers

upon addition of alkenylboronic acids to ,-disubstituted ,-unsaturated sulfones (Scheme 8).44 Although lower conversions were achieved compared to the conjugate addition using arylboronic acids, the products were obtained with excellent enantiomeric excess (88-99%).

Scheme 8 Catalytic asymmetric conjugate addition of alkenylboronic acids.44

2.1.5 Catalytic asymmetric conjugate addition of diorganozinc reagents to ,-unsaturated sulfones

An alternative to the rhodium-catalyzed conjugate addition and the copper-catalyzed conjugate reduction described earlier in this chapter was reported by Charette et al.45 Shortly before our results on this topic were published (vide infra) a method was reported for the catalytic asymmetric conjugate addition of diorganozinc reagents to vinyl sulfones (Scheme 9).

Page 10: University of Groningen Novel asymmetric copper-catalysed ... · 14 Chapter2-Final-nocode.docx Chapter 2 (Scheme 2). In this case simple organolithium reagents gave the best results

19

Chapter2-Final-nocode.docx

Catalytic Asymmetric Conjugate Addition of Grignard Reagents to α,β-Unsaturated Sulfones

Scheme 9 Catalytic asymmetric conjugate addition of diorganozinc reagents.45

Using a copper/(R)-Binap complex, optically active sulfones 31 were obtained with enantiomeric excess up to 98%. Several diorganozinc reagents were reported, but the system is limited to primary diorganozinc reagents and yields are modest to excellent (52-93%). Again, the 2-pyridyl sulfone group was essential in order to get satisfactory results.

2.2 Goal

The aim of this research project was to develop methodology for the catalytic asymmetric

conjugate addition of Grignard reagents to ,-unsaturated sulfones. The resulting optically

active sulfones with a stereocenter at the -position have been shown to be highly versatile intermediates in organic chemistry due to the ease of derivatization and provide access to a wide range of building blocks. These products are not accessible via the rhodium-catalyzed conjugate addition of boronic acids, since this methodology is limited to the introduction of arylboronic acids. Major advantage of the conjugate addition of Grignard reagents, compared to the related conjugate reduction, is that this approach is more modular and thus circumvents the necessity to introduce the substituents at the stereogenic center in the early stages of the synthesis.

2.3 Results and Discussion

2.3.1 Catalyst screening

Initially, we studied the addition of ethylmagnesium bromide to ,-unsaturated sulfone 32a using bidentate phosphine ligands (L2-L5) (Table 2). All reactions gave full conversion overnight, but the best results were obtained using binaphthyl-type phosphine ligands L2 and L5, whereas ferrocenyl-type ligands L3 and L4 gave negligible enantioselectivity. Tol-Binap L2 provided a slightly higher enantiomeric excess compared to Binap (L5) and was used for further screening.

Page 11: University of Groningen Novel asymmetric copper-catalysed ... · 14 Chapter2-Final-nocode.docx Chapter 2 (Scheme 2). In this case simple organolithium reagents gave the best results

20

Chapter2-Final-nocode.docx

Chapter 2

Table 2 Copper/ligand catalyzed addition of EtMgBr to ,-unsaturated sulfone 32a.a, b

Entry Ligand eec,d (%)

1 (R)-Tol-Binap (L2) 47 (R) 2 (R,SFc)-Josiphos (L3) 6 (S) 3 (R,RFc)-Taniaphos (L4) 3 (S) 4 (R)-Binap (L5) 46 (R)

a Conditions: 32a (1 eq., 0.1 mmol in DCM) was added to a solution of EtMgBr (1.2 eq.), CuI (with

L2/L5) or CuBr·Me2S (with L3/L4) (5 mol %) and L2-L5 (5 mol %) in t-BuOMe at 40 oC, 16 h. b Full conversion after 16 h, determined by GC-MS. c Enantiomeric excess determined by chiral HPLC (see Experimental Section). d Determined by comparison with literature data based on the sign of the optical rotation.

2.3.2 Solvent screening

We observed that the use of an alkyl substituted substrate gave higher ee. Therefore, we switched to aliphatic substrates and applying the Cu/Tol-Binap catalytic system, the

addition to ,-unsaturated sulfone 34a in several solvents was examined (Table 3). In all

cases full conversion was obtained overnight at 40 oC. Running the reaction in DCM or t-BuOMe resulted in similar enantioselectivities (Table 3, entries 1 and 3). Using toluene, Et2O or CPME as a solvent provided a slightly lower ee. However, slow addition of the substrate over five hours to the reaction mixture in t-BuOMe increased the enantiomeric excess significantly (Table 3, entry 4). Notably, the use of THF resulted in a very low enantiomeric excess, probably due to coordination with the copper-catalyst or a shift in the Schlenk equilibrium to monomeric EtMgBr species. This dependence is in contrast to that reported by Charette et al. for the conjugate addition of organozinc reagents in which an

Page 12: University of Groningen Novel asymmetric copper-catalysed ... · 14 Chapter2-Final-nocode.docx Chapter 2 (Scheme 2). In this case simple organolithium reagents gave the best results

21

Chapter2-Final-nocode.docx

Catalytic Asymmetric Conjugate Addition of Grignard Reagents to α,β-Unsaturated Sulfones

increase in enantioselectivity was observed with THF as the solvent using a different catalytic system.45

Table 3 Solvent dependence in the addition of EtMgBr to sulfone 34a.a, b

Entry Solvent eec,d (%)

1 DCM 87 (+) 2 Toluene 80 (+) 3 t-BuOMe 88 (+) 4 t-BuOMee 92 (+) 5 THF 8 (−) 6 Et2O 76 (+) 7 CPMEf 68 (+)

a Conditions: 34a (1 eq., 0.1 mmol) was added to EtMgBr (1.2 eq.), CuI (5 mol %), L2 (6 mol %) in

solvent at 40 oC, 16 h.b Full conversion after 16 h, determined by GC-MS. c Determined by chiral HPLC (see Experimental Section).d The absolute configuration of the product is not known. e Slow addition of the substrate solution over 5 h. f CPME = cyclopentyl methyl ether.

2.3.3 Optimization of the copper salt

With the exception of copper(I)cyanide, which gave a lower enantiomeric excess, all copper(I)- and copper(II)-salts tested provided similar results in the conjugate addition reaction of EtMgBr to sulfone 34a (Table 4). In all cases full conversion was obtained after 16 h and no significant effect of the

change in counterion (except for CN) was observed. Slow addition of the substrate solution to the reaction mixture increased the enantioselectivity in the case of CuI (Table 4, entries 4 and 5) while copper(I)chloride gave rise to quantitative conversion and excellent enantiomeric excess (93% ee) even with faster addition times (Table 4, entries 6 and 7).

Page 13: University of Groningen Novel asymmetric copper-catalysed ... · 14 Chapter2-Final-nocode.docx Chapter 2 (Scheme 2). In this case simple organolithium reagents gave the best results

22

Chapter2-Final-nocode.docx

Chapter 2

Table 4 Influence of the copper salt on the addition of EtMgBr to sulfone 34a.a, b

Entry Copper salt eec,d (%)

1 CuCN 69 (+) 2 CuBr·Me2S 89 (+) 3 Cu(OTf)2 87 (+) 4 CuI 88 (+) 5 CuIe 92 (+) 6 CuCl 93 (+) 7 CuCle 93 (+)

a Conditions: 34a (1 eq., 0.1 mmol in t-BuOMe) added directly to EtMgBr (1.2 eq.), Cu salt (5 mol

%), L2 (6 mol %) in t-BuOMe at 40 oC, 16 h.b Full conversion after 16 h, determined by GC-MS. c Determined by chiral HPLC (see Experimental Section).d The absolute configuration of the product is not known. e Slow addition of substrate over 5 h.

2.3.4 Optimization of copper/ligand ratio

Increasing the metal to ligand ratio from 1:1 to 2:1 results in a decrease in enantioselectivity (Table 5). We attribute this to the fact that not all of the copper is bound to the ligand, giving rise to a significant amount of ligand-free copper mediated reaction. It was found that a small excess of ligand with respect to the copper gave the best result (Table 5, entry 4). Increasing the ligand to metal ratio further (Table 5, entry 3) did not improve the enantioselectivity. Table 5 Influence of the copper to ligand ratio on the addition of EtMgBr to sulfone 34a.a, b

Entry CuCl (mol%) L2 (mol%) [Cu]:L2 eec,d (%)

1 5 5 1:1 83 (+) 2 10 5 2:1 62 (+) 3 5 10 1:2 85 (+) 4 5 6 1:1.2 93 (+)

a Conditions: 34a (1 eq., 0.1 mmol in t-BuOMe) was added directly to EtMgBr (1.2 eq.), CuCl and L2

in t-BuOMe at 40 oC, 16 h.b Full conversion after 16 h, determined by GC-MS. c Determined by chiral HPLC (see Experimental Section).d The absolute stereochemistry of the product is not known.

Page 14: University of Groningen Novel asymmetric copper-catalysed ... · 14 Chapter2-Final-nocode.docx Chapter 2 (Scheme 2). In this case simple organolithium reagents gave the best results

23

Chapter2-Final-nocode.docx

Catalytic Asymmetric Conjugate Addition of Grignard Reagents to α,β-Unsaturated Sulfones

2.3.5 Scope of Grignard reagents

Several Grignard reagents were examined using the conditions optimized for EtMgBr (Table 6). In all cases full conversion was observed after 16 h and high isolated yields were obtained. Excellent ee’s were obtained for alkyl Grignard reagents (Table 6, entries 1-5). With MeMgBr a slightly lower enantiomeric excess and yield were attained. Both n-BuMgBr and C6H5C2H4MgBr gave similar enantioselectivities and a slightly lower yield (Table 6, entries 3 and 4). Furthermore, the use of but-3-enylmagnesium bromide also resulted in excellent yield and enantioselectivity (Table 6, entry 5). This functionalized Grignard reagent provides an additional handle for further functionalization.46 Notably, the reaction using PhMgBr did not proceed in an enantioselective manner (Table 6, entry 6). Table 6 Asymmetric conjugate addition of various Grignard reagents to sulfone 34a.a, b

Entry R Product Yieldc (%) eed, e (%)

1 Et 35a 97 93 (+) 2 Me 35b 80 89 (−) 3 n-Bu 35c 88 93 (+) 4 C6H5C2H4 35d 87 87 (−) 5 But-3-enyl 35e 95 94 (+) 6 Ph 35f 80 0

a Conditions: 34a (1 eq., 0.4 M in t-BuOMe) added over 5 h to RMgBr (1.2 eq.), CuCl (5 mol%), L2

(6 mol%) in t-BuOMe at 40 oC, 16 h.b Full conversion after 16 h, determined by GC-MS. c Isolated yields. d Determined by chiral HPLC (see Experimental Section).e The absolute configuration of the product is not known.

2.3.6 Scope of ,-unsaturated sulfones

The synthetic applicability of this highly enantioselective procedure was extended using a

set of ,-unsaturated substrates under the optimized conditions (Table 7). All sulfones provided the desired products in excellent yields (88-97%) and excellent

enantioselectivities (88-94%). Substitution of the ,-unsaturated sulfone at the - or -position did not influence the enantioselectivities or yields (Table 7, entry 1 to 5) and the reactions proceed with both excellent yields and enantiomeric excesses. The presence of a

protected alcohol group (Table 7, entry 6) or phenyl group at the -position in the substrate (Table 7, entry 7) did not affect this enantioselective transformation either and the reactions

Page 15: University of Groningen Novel asymmetric copper-catalysed ... · 14 Chapter2-Final-nocode.docx Chapter 2 (Scheme 2). In this case simple organolithium reagents gave the best results

24

Chapter2-Final-nocode.docx

Chapter 2

resulted in high isolated yields of optically active -disubstituted sulfones with excellent ee’s. Table 7 Asymmetric conjugate addition of EtMgBr to ,-unsaturated sulfones.a, b

Entry 34 R Product Yieldc (%) eed, e (%)

1 34a n-Pent 35a 97 93 (+) 2 34b n-Oct 36 90 92 (+) 3 34c i-Bu 37 88 94 (−) 4 34d i-Pr 38 93 88 (−) 5 34e c-Hex 39 94 94 (−) 6 34f TBDPSOC2H4 40 91 92 (+) 7 34g C6H5C2H4 41 91 93 (+)

a Conditions: 34 (0.2 mmol, 1 eq., 0.4 M in t-BuOMe) added over 5 h to EtMgBr (1.2 eq.), CuCl (5

mol%), L2 (6 mol%) in t-BuOMe at 40 oC, 16 h.b Full conversion after 16 h, determined by GC-MS. c Isolated yields. d Determined by chiral HPLC (see Experimental Section).e The absolute configuration of the product is not known.

2.3.7 The influence of the 2-pyridyl group and limitations of the system

The influence of the 2-pyridyl group was examined by applying the asymmetric conjugate

addition to the corresponding p-tolyl substituted ,-unsaturated sulfone 42 instead of 2-pyridyl substituted sulfone 34a (Scheme 10).

Scheme 10 Asymmetric conjugate addition of EtMgBr to sulfone 42.

In this experiment the reaction rate and enantiomeric excess decreased dramatically. This effect of the 2-pyridyl group has also been noted by Carretero19, 41, 44, 47 and Charette45 and co-workers for related systems. As already mentioned in section 2.1.3-2.1.5, the 2-pyridyl group seems to be necessary both in terms of enantioselectivity and reactivity.

Page 16: University of Groningen Novel asymmetric copper-catalysed ... · 14 Chapter2-Final-nocode.docx Chapter 2 (Scheme 2). In this case simple organolithium reagents gave the best results

25

Chapter2-Final-nocode.docx

Catalytic Asymmetric Conjugate Addition of Grignard Reagents to α,β-Unsaturated Sulfones

A limitation of this methodology is that ,-unsaturated sulfones substituted with a phenyl group at the β-position unfortunately give only moderate ee’s under the optimized conditions (Table 8). Table 8 Asymmetric conjugate addition of EtMgBr to vinyl phenyl sulfones.a, b

Entry R Product Yieldc (%) eed (%)

1 H 33a 75 47 (R) 2 CF3 33b 65 51e 3 Br 33c 76 70e

a Conditions: 34a (1 eq.,0.4 M in t-BuOMe) added over 5 h to EtMgBr (1.2 eq.), CuCl (5 mol%), L2

(6 mol%) in t-BuOMe at 40 oC, 16 h.b Full conversion after 16 h, determined by GC-MS. c Isolated yields. d Determined by chiral HPLC.e The absolute configuration of the product is not known.

2.4 Conclusion

In summary, we have developed a novel copper-catalyzed asymmetric conjugate addition

reaction of Grignard reagents to a range of aliphatic ,-unsaturated sulfones. This

procedure has a broad scope for aliphatic substrates and provides -substituted 2-pyridylsulfones in both excellent yields (88-97%) and enantioselectivities (88-94%). These enantioenriched sulfones are versatile intermediates in the preparation of a wide variety of functionalized chiral building blocks.

Page 17: University of Groningen Novel asymmetric copper-catalysed ... · 14 Chapter2-Final-nocode.docx Chapter 2 (Scheme 2). In this case simple organolithium reagents gave the best results

26

Chapter2-Final-nocode.docx

Chapter 2

2.5 Experimental section

General Chromatography: Merck silica gel type 9385 230-400 mesh, TLC: Merck silica gel 60, 0.25 mm. Components were visualized by staining with a solution of a mixture of KMnO4 (10 g) and K2CO3 (10 g) in H2O (500 mL). Progress and conversion of the reaction were determined by GC-MS (GC, HP6890: MS HP5973) with an HP1 or HP5 column (Agilent Technologies, Palo Alto, CA). Mass spectra were recorded on a AEI-MS-902 mass spectrometer (EI+) or a LTQ Orbitrap XL (ESI+). 1H- and 13C-NMR were recorded on a Varian AMX400 (400 and 100.59 MHz, respectively) or a Varian VXR300 (300 and 75 MHz, respectively) using CDCl3 as solvent. Chemical shift values are reported in ppm with

the solvent resonance as the internal standard (CHCl3: 7.26 for 1H, 77.0 for 13C). Data are reported as follows: chemical shifts, multiplicity (s = singlet, d = doublet, t = triplet, q = quartet, br = broad, m = multiplet), coupling constants (Hz), and integration. Carbon assignments are based on APT 13C-NMR experiments. Optical rotations were measured on a Schmidt + Haensch polarimeter (Polartronic MH8) with a 10 cm cell (c given in g/100 mL). Enantioselectivities were determined by HPLC analysis using a Shimadzu LC-10ADVP HPLC equipped with a Shimadzu SPD-M10AVP diode array detector. Elemental analysis was performed on a EuroVector Euro EA-3000 Elemental Analyzer. All reactions were carried out under a nitrogen atmosphere using flame dried glassware. t-BuOMe was purchased as anhydrous grade, stored on 4Å MS and used without further purification. All copper-salts and chiral ligands (L2-L5) were purchased from Aldrich or Acros and used without further purification. Grignard reagents were purchased from Aldrich (MeMgBr, EtMgBr) or prepared from the corresponding alkylbromides and magnesium turnings in Et2O following standard procedures. Grignard reagents were titrated using s-BuOH and catalytic amounts of 1,10-phenanthroline. Racemic 1,4-addition

products were synthesized by reaction of the ,-unsaturated sulfones (32a-c, 34a-g) with the corresponding Grignard reagent at –40 °C in THF in the presence of a stoichiometric amount of CuI.

Page 18: University of Groningen Novel asymmetric copper-catalysed ... · 14 Chapter2-Final-nocode.docx Chapter 2 (Scheme 2). In this case simple organolithium reagents gave the best results

27

Chapter2-Final-nocode.docx

Catalytic Asymmetric Conjugate Addition of Grignard Reagents to α,β-Unsaturated Sulfones

Synthesis of 2-(methylsulfonyl)pyridine19

To a solution of 2-mercaptopyridine (11.11 g, 100 mmol) in dry THF (200 mL) and acetonitrile (20 mL), cooled to 0 oC, DBU (16.75 g, 110 mmol) was added. The resulting mixture was stirred at 0 oC for 5 min before methyl iodide (15.61 g, 110 mmol) was added slowly. The ice bath was removed and the mixture was stirred overnight. The reaction mixture was washed with water (100 mL) and the aqueous layer was extracted with ethyl acetate (3 x 100 mL). The combined organic layers were dried (MgSO4), filtered and concentrated. The residue was purified by flash chromatography (n-hexane/ethyl acetate, 2:1) to afford the methyl 2-pyridyl sulfide (A) as a colorless oil; yield: 12.26 g (98%). 1H-

NMR (300 MHz): 8.42 (m, 1H), 7.47 (m, 1H), 7.16 (m, 1H), 6.96 (m, 1H), 2.55 (s, 3H). 13C-NMR (75 MHz): 159.9, 149.3, 135.6, 121.3, 119.0, 13.1. To a solution of A (12.26 g, 98 mmol) in ethyl acetate (125 mL) were added H2O (15 mL) and Na2WO4·2H2O (3.23 g, 9.8 mmol). The resulting mixture was cooled to 0 ºC before an aqueous solution of H2O2 (3 eq, 30%, 30 mL, 294 mmol) was added dropwise. The reaction mixture was stirred at 0 oC for 30 min and at rt for 1 h, cooled to 0 oC and saturated aqueous NaHSO3 (25 mL) was added slowly. The organic layer was separated and the aqueous layer was extracted with ethyl acetate (2 x 50 mL). The combined organic layers were dried (NaSO4), filtered and concentrated. The residue was purified by flash chromatography (n-hexane/ ethyl acetate, 2:1) to afford the sulfone as a colorless oil; yield:

14.35 g (93%). 1H-NMR (300 MHz): 8.77-8.70 (m, 1H), 8.12-7.89 (m, 2H), 7.62-7.49 (m,

1H), 3.21 (s, 3H). 13C-NMR (75 MHz): 157.8, 149.9, 138.3, 127.4, 121.0, 39.9. General procedure for the synthesis of ,-unsaturated sulfones (34a-g) To a solution of 20 mmol of 2-(methylsulfonyl)pyridine (3.14 g) in dry THF (40 mL),

cooled to 78 oC, a 1.6 M solution of n-BuLi in hexane (13.75 mL, 22 mmol, 1.1 eq) was

added. The mixture was stirred at 78 oC for 30 min followed by addition of the aldehyde

(22 mmol, 1.1 eq.) at 78 oC and the reaction mixture was slowly warmed to room temperature. The reaction mixture was quenched with saturated aqueous NH4Cl (25 mL). The organic layer was separated and the aqueous layer was extracted with ethyl acetate (3 x 50 mL). The combined organic layers were dried with Na2SO4, filtered and concentrated in vacuo. Generally, the crude alcohol can be used without further purification for the next dehydration step. If the resulting 2-pyridylsulfonylalcohol was not completely pure, a simple flash chromatography on silica gel (pentane/Et2O) of this intermediate was performed before the dehydration step. The crude alcohol was dissolved in dry DCM (150

Page 19: University of Groningen Novel asymmetric copper-catalysed ... · 14 Chapter2-Final-nocode.docx Chapter 2 (Scheme 2). In this case simple organolithium reagents gave the best results

28

Chapter2-Final-nocode.docx

Chapter 2

mL) under nitrogen atmosphere and 80 mmol of DMAP (9.77 g, 4 eq.) was added and the reaction mixture was cooled down to 0 oC. Subsequently, 40 mmol of methanesulfonyl chloride (4.58 g, 2 eq.) was added and the mixture was stirred while slowly warming to room temperature overnight. The reaction mixture was quenched with saturated aqueous NH4Cl (100 mL). The layers were separated and the aqueous layer was extracted with DCM (3 x 50 mL). The combined organic layers were dried with Na2SO4, filtered and the solvent evaporated in vacuo. The crude product was purified by flash chromatography on

silica gel (Pentane:Et2O 2:1-1:1) to yield the pure ,-unsaturated sulfone. (E)-2-(Styrylsulfonyl)pyridine (32a)45

White solid, yield: 67%. Mp: 101.6 oC. 1H NMR (400 MHz,

CDCl3) 8.75 (br d, J = 4.7 Hz, 1H), 8.15 (dt, J= 7.7, 0.9 Hz, 1H), 7.96 (dt, J = 7.7, 7.7, 1.5 Hz, 1H), 7.79 (d, J = 15.5 Hz, 1H), 7.63-7.58 (m, 3H), 7.46 – 7.38 (m, 3H), 7.12 (d, J = 15.5

Hz, 1H). 13C NMR (100 MHz, CDCl3): 158.5 (s), 150.4 (d), 145.1 (d), 138.2 (d), 132.3 (s), 131.4 (d), 129.1 (d), 128.8 (d),

127.1 (d), 124.5 (d), 121.9 (d). HRMS (EI+, m/z): calcd. for C13H11NO2S [M]+: 245.0510; found: 245.0502. Anal. Calcd for C13H11NO2S: C, 63.65; H, 4.52; N, 5.71; S, 13.04. Found: C, 63.43; H, 4.57; N, 5.50; S, 13.05. (E)-2-(Hept-1-enylsulfonyl)pyridine (34a)47, 48

Colorless oil, yield: 75%. 1H NMR (400 MHz, CDCl3):

8.73 (d, J = 4.1 Hz, 1H), 8.08 (d, J = 7.8 Hz, 1H), 7.94 (dt, J = 7.8, 7.8, 1.7 Hz, 1H), 7.51 (ddd, J = 7.6, 4.7, 0.9 Hz, 1H), 7.12 (td, J = 15.1, 6.8, 6.8 Hz, 1H), 6.53 (td, J = 15.2, 1.6, 1.6 Hz, 1H) 2.28 (dq, J = 6.8, 1.6 Hz, 2H), 1.54-1.43 (m, 2H), 1.36-1.20 (m, 4H), 0.9-0.8 (m, 3H).

13C NMR (100 MHz, CDCl3): 158.4, 150.6, 150.2, 138.2, 127.4, 127.1, 121.9, 31.8, 31.1, 27.1, 22.3, 13.9. The physical and spectroscopic properties were in accordance with those described in literature. (E)-2-(Dec-1-enylsulfonyl)pyridine (34b)

Colorless oil, yield: 50%. 1H NMR (400 MHz,

CDCl3): 8.71 (d, J = 4.6 Hz, 1H), 8.06 (dd, J = 7.9, 0.7 Hz, 1H), 7.86 ( td, J = 7.6, 1.6 Hz, 1H), 7.50 (dd, J = 7.6, 4.7 Hz, 1H), 7.10 (dt, J = 14.8, 6.8 Hz, 1H), 6.52 (dd, J = 15.2, 0.6 Hz, 1H), 2.27 (dd, J = 14.0, 6.9 Hz, 2H), 1.50-1.40

(m, 2H), 1.35-1.12 (m 10H), 0.84 (m, 3H). 13C NMR (100 MHz, CDCl3): 158.5 (s), 150.3

Page 20: University of Groningen Novel asymmetric copper-catalysed ... · 14 Chapter2-Final-nocode.docx Chapter 2 (Scheme 2). In this case simple organolithium reagents gave the best results

29

Chapter2-Final-nocode.docx

Catalytic Asymmetric Conjugate Addition of Grignard Reagents to α,β-Unsaturated Sulfones

(d), 150.2 (d), 138.1 (d), 127.5 (d), 127.0 (d), 121.7 (d), 31.7 (t), 31.7 (t), 29.1 (t), 29.0 (t), 28.9 (t), 27.4 (t), 22.5 (t), 14.0 (q). HRMS (ESI+, m/z): calcd. for C15H24NO2S [M+H]+, 282.15223; found, 282.15232. (E)-2-(4-Methylpent-1-enylsulfonyl)pyridine (34c)

Colorless oil, yield: 53%. 1H NMR (400 MHz, CDCl3): 8.70 (dd, J = 2.8, 1.9 Hz, 1H), 8.05 (dd, J = 7.9, 1.0 Hz, 1H), 7.92 (tt, J = 7.8, 7.8, 1.9, 1.9 Hz, 1H), 7.49 (dddd, J = 7.8, 4.7, 2.0, 1.2 Hz, 1H), 7.05 (dtd, J = 9.8, 7.5, 7.5, 2.3 Hz, 1H), 6.50 (m, 1H), 2.2-2.1 (m, 2H), 1.79 (m, 1H), 0.89 (dd, J = 6.7, 2.3 Hz,

6H). 13C NMR (100 MHz, CDCl3): 158.4 (s), 150.2 (d), 149.1 (d), 138.1 (d), 128.5 (d), 127.0 (d), 121.7 (d), 40.7 (t), 27.5 (d), 22.1 (q). HRMS (EI+, m/z): calcd. for C11H15NO2S [M]+, 224.0745; found, 224.0734. (E)-2-(3-Methylbut-1-enylsulfonyl)pyridine (34d)47, 48

Colorless oil, yield: 62%. 1H NMR (400 MHz, CDCl3): 8.72 (dd, J = 4.7, 0.8 Hz, 1H), 8.07 (dd, J = 7.9, 0.8 Hz, 1H), 7.93 (tt, J = 7.8, 7.8, 1.5, 1.5 Hz, 1H), 7.60-7.42 (m, 1H), 7.09 (ddd, J = 15.3, 6.2, 1.5 Hz, 1H), 6.49 (td, J = 15.3, 1.5, 1.5 Hz, 1H), 2.55 (m, 1H), 1.08 (dd, J = 6.8, 1.5 Hz, 6H). 13C NMR (100 MHz,

CDCl3): 158.5 (s), 155.8 (d), 150.2 (d), 138.1 (d), 127.0 (d), 125.4 (d), 121.8 (d), 30.9 (d), 20.7 (q). HRMS (EI+, m/z): calcd. for C9H10NO2S [M – CH3]

+: 196.0432; found: 196.0422. (E)-2-(2-Cyclohexylvinylsulfonyl)pyridine (34e)

Colorless oil, yield: 41%. 1H NMR (400 MHz, CDCl3): 8.70 (br d, J = 4.1 Hz, 1H), 8.05 (td, J = 7.9, 0.8, 0.8 Hz, 1H), 7.92 (dt, J = 7.8, 7.7, 1.7 Hz, 1H), 7.49 (ddd, J = 7.7, 4.7, 0.9 Hz, 1H), 7.04 (dd, J = 15.3, 6.3 Hz, 1H), 6.5 (dd, J = 15.3, 1.4 Hz, 1H), 2.21 (m, 1H), 1.86-1.68 (m, 4H), 1.68-1.59 (m, 1H), 1.33-

1.06 (m, 5H). 13C NMR (100 MHz, CDCl3): 158.5 (s), 154.5 (d), 150.2 (d), 138.1 (d), 126.9 (d), 125.6 (d), 121.7 (d), 40.1 (d), 31.1 (t), 25.6 (t), 25.5 (t). HRMS (EI+, m/z): calcd. for C13H16NO2S [M – H]+: 250.0902; found: 250.0915.

Page 21: University of Groningen Novel asymmetric copper-catalysed ... · 14 Chapter2-Final-nocode.docx Chapter 2 (Scheme 2). In this case simple organolithium reagents gave the best results

30

Chapter2-Final-nocode.docx

Chapter 2

(E)-2-(4-(tert-Butyldiphenylsilyloxy)but-1-enylsulfonyl)pyridine (34f) Colorless oil, yield: 58%. 1H NMR (400 MHz,

CDCl3): 8.70 (m, 1H), 8.09 (br d, J = 7.8 Hz, 1H), 7.91 (dt, J = 7.9, 7.8, 1.8 Hz, 1H), 7.61 (dd, J = 7.8, 1.4 Hz, 4H), 7.54-7.45 (m, 1H), 7.46-7.29 (m, 6H), 7.15 (td, J = 15.2, 6.8, 6.8 Hz, 1H), 6.68 (td, J = 15.2, 1.4, 1.4 Hz, 1H), 3.79 (t, J = 6.0, 6.0 Hz, 2H), 2.51 (dq, J = 6.3, 6.2, 6.2, 1.3 Hz, 2H), 0.96 (s, 9H).

13C NMR (100 MHz, CDCl3): 158.4 (s), 150.3 (d), 146.8 (d), 138.0 (d), 135.4 (d), 133.2 (s), 129.7 (d), 129.3 (d), 127.7 (d), 127.0 (d), 121.8 (d), 61.4 (t), 34.8 (t), 26.6 (q), 19.0 (s). HRMS (EI+, m/z): calcd. for C21H20NO3SiS [M–C4H9]

+: 394.0933; found: 394.0945. (E)-2-(4-Phenylbut-1-enylsulfonyl)pyridine (34g)49

White solid, yield: 48%. Mp: 106.6 oC (Lit: 96-98 oC).49 1H NMR (400 MHz, CDCl3): 8.72 (d, J = 4.6 Hz, 1H), 8.05 (dd, J = 7.8, 0.7 Hz, 1H), 7.93 (dt, J = 7.7, 7.7, 1.5 Hz, 1H), 7.69-7.37 (m, 1H), 7.28-7.13 (m, 6H), 6.68-6.41 (m, 1H), 2.8 (t, J = 7.6 Hz, 2H), 2.61 (q, J = 7.6 Hz, 2H). 13C NMR (100 MHz, CDCl3): 158.4 (s), 150.2 (d),

148.8 (d), 139.9 (s), 138.1 (d), 128.5 (d), 128.4 (d), 128.3 (d), 127.0 (d), 126.3 (d), 121.8 (d), 33.7 (t), 33.4 (t). HRMS (ESI+, m/z): calcd. for C15H16NO2S [M+H]+: 274.08963; found: 274.08962. Anal. Calcd for C15H15NO2S: C, 65.91; H, 5.53; N, 5.12. Found: C, 65.65; H, 5.63; N, 4.92. General procedure for the asymmetric conjugate addition of Grignard reagents to ,-unsaturated sulfones

CuCl (12.5 mol, 1.24 mg, 5 mol%) and (R)-(+)-Tol-Binap (L2, 15.0 mol, 10.18 mg, 6 mol%) were dissolved in 4 mL of dry t-BuOMe under a nitrogen atmosphere. The mixture

was stirred for 15 min and cooled down to 40 oC and the Grignard reagent (1.2 eq) was added. After stirring for 15 min, the substrate (0.25 mmol, dissolved in 0.5 mL of t-BuOMe) was added over 5 h and the mixture was stirred overnight. Aqueous saturated NH4Cl solution (2 mL) was added and the mixture was warmed up to room temperature. The mixture was diluted with Et2O and the layers were separated. The aqueous layer was extracted with DCM (3 x 10 mL) and the combined organic layers were dried with anhydrous Na2SO4, filtered and the solvent evaporated in vacuo. The crude product was purified by flash chromatography on silica (Pentane: Et2O, 4:1-1:1) to afford the products as colorless to pale yellow oils.

Page 22: University of Groningen Novel asymmetric copper-catalysed ... · 14 Chapter2-Final-nocode.docx Chapter 2 (Scheme 2). In this case simple organolithium reagents gave the best results

31

Chapter2-Final-nocode.docx

Catalytic Asymmetric Conjugate Addition of Grignard Reagents to α,β-Unsaturated Sulfones

(+)-2-(2-Ethylheptylsulfonyl)pyridine (35a) Following the general procedure for the asymmetric Cu-catalyzed conjugate addition, 35a was isolated in 97% yield. Enantiomeric excess: 93% determined by chiral HPLC analysis, Chiralpak AD column, 1.0 mL/min, n-heptane: i-PrOH 99:1, 40 oC, 210 nm, retention times

(min): 32.5 (minor) and 38.5 (major). []D = +2.8 (c 1.0,

CHCl3). 1H NMR (400 MHz, CDCl3): 8.74 (ddd, J = 4.7, 1.7, 0.9 Hz, 1H), 8.10 (td, J =

7.8, 1.0, 1.0 Hz, 1H), 7.96 (td, J = 7.8, 1.0, 1.0 Hz, 1H), 7.54 (ddd, J = 7.7, 4.7, 1.2 Hz, 1H), 3.33 (d, J = 6.1 Hz, 2H), 2.02-1.92 (m, 1H), 1.56-1.42 (m, 2H), 1.43-1.34 (m, 2H),

1.31-1.10 (m, 6H), 0.84 (dt, J = 7.4, 7.2, 5.9 Hz, 6H). 13C NMR (100 MHz, CDCl3): 157.9 (s), 150.2 (d), 138.1 (d), 127.2 (d), 121.9 (d), 55.2 (t), 34.0 (d), 32.5 (t), 31.8 (t), 25.6 (t) 25.5 (t), 22.5 (t), 14.0 (q), 10.1 (q). HRMS (EI+, m/z): calcd. for C12H18NO2S [M - C2H5]

+: 240.1058; found: 240.1068.

()-2-(2-Methylheptylsulfonyl)pyridine (35b) Following the general procedure for the asymmetric Cu-catalyzed conjugate addition, 35b was isolated in 80% yield. Enantiomeric excess: 89% determined by chiral HPLC analysis, Chiralcel OD-H 0.5 mL/min, n-heptane: i-PrOH 98:2, 40 oC, 210 nm, retention times (min): 46.9

(minor) and 49.1 (major). []D = 1.4 (c 1.0, CHCl3). 1H

NMR (400 MHz, CDCl3): 8.75 (d, J = 4.6 Hz, 1H), 8.10 (dd, J = 7.8, 0.9 Hz, 1H), 7.96 (tt, J = 7.6, 1.4Hz, 1H), 7.55 (dd, J = 6.9, 4.7 Hz, 1H), 3.43 (dd, J = 14.2, 4.6 Hz, 1H), 3.20 (dd, J = 14.2, 8.0 Hz, 1H), 2.13 (qt, J = 13.7, 13.7, 6.9, 6.8, 6.8 Hz, 1H), 1, 1.48-1.12 (m,

8H), 1.06 (d, J = 6.7 Hz, 3H), 0.85 (t, J = 7.0, 7.0 Hz, 3H). 13C NMR (100 MHz, CDCl3): 158.0 (s), 105.2 (d), 138.1 (d), 127.2 (d), 121.9 (d), 57.9 (t), 36.6 (t), 31.6 (t), 28.2 (d), 26.0 (t), 22.5 (t), 19.8 (q), 14.0 (q). HRMS (EI+, m/z): calcd. for C12H18NO2S [M – CH3]

+: 240.1058; found: 240.1070. (+)-2-(2-Butylheptylsulfonyl)pyridine (35c)

Following the general procedure for the asymmetric Cu-catalyzed conjugate addition, 35c was isolated in 88% yield. Enantiomeric excess: 93% determined by chiral HPLC analysis, Chiralpak AD-H 0.5 mL/min, n-heptane: i-PrOH 98:2, 40 oC, 210 nm, retention times (min): 35.0

(minor) and 36.7 (major). []D = +9.2 (c 1.0, CHCl3). 1H

NMR (400 MHz, CDCl3): 8.74 (ddd, J = 4.6, 1.5, 0.8 Hz, 1H), 8.09 (td, J = 7.8, 1.0, 1.0

Page 23: University of Groningen Novel asymmetric copper-catalysed ... · 14 Chapter2-Final-nocode.docx Chapter 2 (Scheme 2). In this case simple organolithium reagents gave the best results

32

Chapter2-Final-nocode.docx

Chapter 2

Hz, 1H), 7.95 (dt, J = 7.8, 7.8, 1.7 Hz, 1H), 7.54 (ddd, J = 7.6, 4.7, 1.2 Hz, 1H), 3.33 (d, J = 6.1 Hz, 2H), 2.06-1.95 (m, 1H), 1.50-1.30 (m, 4H), 1.30-1.10 (m, 10H), 0.84 (t, J = 7.0,

7.0 Hz, 6H). 13C NMR (100 MHz, CDCl3): 158.0 (s), 150.1 (d), 138.1 (d), 127.2 (d), 122.0 (d), 55.6 (t), 33.1 (t), 32.8 (t), 32.7 (d), 31.8 (t), 28.0 (t) 25.5 (t), 22.6 (t), 22.5 (t), 14.0 (q), 13.9 (q). HRMS (EI+, m/z): calcd. for C12H18NO2S [M–C4H9]

+: 240.1058; found: 240.1069.

()-2-(2-Phenethylheptylsulfonyl)pyridine (35d) Following the general procedure for the asymmetric Cu-catalyzed conjugate addition, 35d was isolated in 87% yield. Enantiomeric excess: 87% determined by chiral HPLC analysis, Chiralcel OD-H, 0.5 mL/min, n-heptane: i-PrOH 97:3, 40 oC, 210 nm, retention times (min): 53.6

(major) and 57.1 (minor). []D = 9.8 (c 1.0, CHCl3). 1H

NMR (400 MHz, CDCl3): 8.76-8.73 (m, 1H), 8.04 (dq, J = 7.6, 0.8 Hz, 1H), 7.93 (tdd, J = 7.8, 7.8, 1.7, 0.8 Hz,

1H), 7.53 (ddt, J = 7.6, 4.8, 1.0 Hz, 1H), 7.31-7.21 (m, 2H), 7.21-7.07 (m, 3H), 3.40 (t, J = 5.6, 5.6 Hz, 2H), 2.58 (t, J = 8.1, 8.1 Hz, 2H), 2.12-1.96 (m, 1H), 1.92-1.66 (m, 2H), 1.54-1.38 (m, 2H), 1.36-1.05 (m, 6H), 0.86 (t, J = 7.1, 7.1 Hz, 3H). 13C NMR (100 MHz,

CDCl3): 157.1 (s), 150.2 (d), 141.7 (s), 138.1 (d), 128.3 (d), 127.2 (d), 125.8 (d), 122.1 (d), 55.4 (t), 34.9 (t), 33.0 (t), 32.4 (d), 32.3 (t), 31.7 (t), 25.4 (t), 22.5 (t), 14.0 (q). HRMS (ESI+, m/z): calcd. for C20H28NO2S [M+H]+: 346.18353; found: 346.18350. (+)-2-(2-(But-3-enyl)heptylsulfonyl)pyridine (35e)

Following the general procedure for the asymmetric Cu-catalyzed conjugate addition, 35e was isolated in 95% yield. Enantiomeric excess: 94% determined by chiral HPLC analysis, Chiralpak AD-H, 0.5 mL/min, n-heptane: i-PrOH 98.5:1.5, 40 oC, 210 nm, retention times

(min): 53.0 (minor) and 54.4 (major). []D = +3.0 (c 1.0,

CHCl3). 1H NMR (400 MHz, CDCl3): 8.76-8.73 (m,

1H), 8.09 (dq, J = 7.8, 1.9, 1.0 Hz, 1H), 7.96 (ddt, J = 7.8, 7.8, 1.7, 0.9 Hz, 1H), 7.54 (tdd, J = 7.6, 4.7, 0.9, 0.9 Hz, 1H), 5.71 (tdd, J = 16.9, 10.2, 6.6, 6.6 Hz, 1H), 5.06-4.80 (m, 2H), 3.35 (dd, J = 6.0, 2.4 Hz, 2H), 2.15-1.91 (m, 3H), 1.62-1.33 (m, 4H), 1.32-1.09 (m, 6H),

0.84 (t, J = 7.1, 7.1 Hz, 3H). 13C NMR (100 MHz, CDCl3): 157.8 (s), 150.2 (d), 138.1 (d), 138.0 (d), 127.2 (d), 122.0 (d), 114.9 (t), 55.4 (t), 32.9 (t), 32.3 (t), 32.2 (d), 31.7 (t), 30.2 (t), 25.4 (t), 22.5 (t), 14.0 (q). HRMS (ESI+, m/z): calcd. for C16H26NO2S [M+H]+: 296.16788; found: 296.16782.

SO

ON

35e

Page 24: University of Groningen Novel asymmetric copper-catalysed ... · 14 Chapter2-Final-nocode.docx Chapter 2 (Scheme 2). In this case simple organolithium reagents gave the best results

33

Chapter2-Final-nocode.docx

Catalytic Asymmetric Conjugate Addition of Grignard Reagents to α,β-Unsaturated Sulfones

2-(2-Phenylheptylsulfonyl)pyridine (35f)45 Following the general procedure for the asymmetric Cu-catalyzed conjugate addition, 35f was isolated as a white solid in 80% yield. Enantiomeric excess: 0% determined by chiral HPLC analysis, Chiralpak AD, 1.0 mL/min, n-heptane: i-PrOH 97:3, 40 oC, 210 nm, retention times (min): 23.8 and 25.7. M.p.: 85-87 oC. 1H NMR (400

MHz, CDCl3): 8.63-8.48 (m, 1H), 7.76-7.63 (m, 2H), 7.40-7.31 (m, 1H), 7.11-6.94 (m, 5H), 3.95 (dd, J = 14.6, 8.8 Hz, 1H), 3.56 (dd, J = 14.6, 5.3 Hz, 1H), 3.30-3.19 (m, 1H), 1.88-1.72 (m, 1H), 1.69-1.55 (m, 1H), 1.39-0.84 (m, 6H),

0.80 (t, J = 6.6, 6.6 Hz, 3H). 13C NMR (100 MHz, CDCl3): 157.5 (s), 149.8 (d), 141.4 (s), 137.6 (d), 128.3 (d), 127.8 (d), 126.7 (d), 126.6 (d), 122.1 (d), 57.8 (t), 40.7 (d), 36.4 (t), 31.4 (t), 26.6 (t), 22.4 (t), 13.9 (q). HRMS (ESI+, m/z): calcd. for C18H24NO2S [M+H]+: 318.15223; found: 318.15210. (+)-2-(2-Ethyldecylsulfonyl)pyridine (36)

Following the general procedure for the asymmetric Cu-catalyzed conjugate addition, 36 was isolated in 90% yield. Enantiomeric excess: 92% determined by chiral HPLC analysis, Chiralpak AD, 1.0 mL/min, n-heptane: i-PrOH 99:1, 40 oC, 210 nm, retention times (min): 25.8

(minor) and 30.1 (major). []D = +4.6 (c 1.0, CHCl3). 1H NMR (400 MHz, CDCl3): 8.73

(d, J = 4.6 Hz, 1H), 8.08 (d, J = 7.8 Hz, 1H), 7.95 (dt, J = 7.8, 7.8, 1.7 Hz, 1H), 7.53 (ddd, J = 7.6, 4.7, 1.0 Hz, 1H), 3.32 (d, J = 6.1 Hz, 2H), 2.02-1.90 (m, 1H), 1.57-1.41 (m, 2H),

1.42-1.31 (m, 2H), 1.30-1.11 (m, 12H), 0.94-0.74 (m, 6H). 13C NMR (100 MHz, CDCl3): 157.9 (s), 150.1 (d), 138.1 (d), 127.2 (d), 121.9 (d), 55.2 (t), 33.9 (d), 32.6 (t), 31.8 (t), 29.6 (t), 29.4 (t), 29.2 (t), 25.8 (t), 25.6 (t), 22.6 (t), 14.0 (q), 10.1 (q). HRMS (EI+, m/z): calcd. for C15H24NO2S [M–C2H5]

+: 282.1528; found: 282.1535.

()-2-(2-Ethyl-4-methylpentylsulfonyl)pyridine (37) Following the general procedure for the asymmetric Cu-catalyzed conjugate addition, 37 was isolated in 88% yield. Enantiomeric excess: 94% determined by chiral HPLC analysis, Chiralcel OD-H, 0.5 mL/min, n-heptane: i-PrOH 98:2, 40 oC, 210 nm, retention times (min): 40.8 (major) and 46.7

(minor). []D = 6.2 (c 1.0, CHCl3). 1H NMR (400 MHz, CDCl3): 8.74 (br d, J = 4.5 Hz,

1H), 8.09 (br d, J = 7.8 Hz, 1H), 7.95 (dt, J = 7.8, 7.8, 1.7 Hz, 1H), 7.54 (ddd, J = 7.6, 4.7, 1.1 Hz, 1H), 3.34-3.27 (m, 2H), 2.07-1.95 (m, 1H), 1.62-1.40 (m, 3H), 1.22 (t, J = 7.1, 7.1

SO

ON

37

Page 25: University of Groningen Novel asymmetric copper-catalysed ... · 14 Chapter2-Final-nocode.docx Chapter 2 (Scheme 2). In this case simple organolithium reagents gave the best results

34

Chapter2-Final-nocode.docx

Chapter 2

Hz, 2H), 0.85-0.75 (m, 9H). 13C NMR (100 MHz, CDCl3): 157.9 (s), 150.1 (d), 138.1 (d), 127.2 (d), 122.0 (d), 55.4 (t), 42.5 (t), 31.7 (d), 25.8 (t), 24.9 (d), 22.6 (q), 22.3 (q), 9.7 (q). HRMS (EI+, m/z): calcd. for C12H18NO2S [M–CH3]

+: 240.1058; found: 240.1065.

()-2-(2-Ethyl-3-methylbutylsulfonyl)pyridine (38) Following the general procedure for the asymmetric Cu-catalyzed conjugate addition, 38 was isolated in 93% yield. Enantiomeric excess: 88% determined by chiral HPLC analysis, Chiralpak AD, 1.0 mL/min, n-heptane: i-PrOH 98:2, 40 oC, 210 nm, retention

times (min): 19.1 (major) and 30.3 (minor). []D = 9.0 (c 1.0,

CHCl3). 1H NMR (400 MHz, CDCl3): 8.75 (br d, J = 4.0 Hz,

1H), 8.09 (br d, J = 7.8 Hz, 1H), 7.96 (dt, J = 7.7, 7.7, 1.4 Hz, 1H), 7.54 (br dd, J = 7.0, 4.8 Hz, 1H), 3.39 (dd, J = 14.5, 4.5 Hz, 1H), 3.16 (dd, J = 14.5, 7.1 Hz, 1H), 1.97-1.86 (m, 1H), 1.84-1.76 (m, 1H), 1.56-1.34 (m, 2H), 0.84 (t, J = 7.4, 7.4 Hz, 3H), 0.80 (d, J = 6.8

Hz, 6H). 13C NMR (100 MHz, CDCl3): 157.7 (s), 150.2 (d), 138.0 (d), 127.2 (d), 122.1 (d), 52.8 (t), 40.0 (d), 28.5 (d), 23.3 (t), 19.0 (q), 18.0 (q), 11.3 (q). HRMS (EI+, m/z): calcd. for C9H12NO2S [M–C3H7]

+: 198.0589; found: 198.0592.

()-2-(2-Cyclohexylbutylsulfonyl)pyridine (39) Following the general procedure for the asymmetric Cu-catalyzed conjugate addition, 39 was isolated in 94% yield. Enantiomeric excess: 94% determined by chiral HPLC analysis, Chiralpak AS, 1.0 mL/min, n-heptane: i-PrOH 95:5, 40 oC, 210 nm, retention times (min): 12.6 (minor) and 14.1

(major). []D = 3.4 (c 1.0, CHCl3). 1H NMR (400 MHz,

CDCl3): 8.74 (br d, J = 3.1 Hz, 1H), 8.15-8.03 (m, 1H), 8.01-7.89 (m, 1H), 7.61-7.46 (m, 1H), 3.46 (td, J = 8.8, 4.2, 4.2 Hz, 1H), 3.16 (ddd, J = 14.5, 7.1, 4.5 Hz, 1H), 1.87-1.36 (m,

8H), 1.30-0.80 (m, 9H). 13C NMR (100 MHz, CDCl3): 157.8 (s), 150.1 (d), 138.0 (d), 127.2 (d), 122.1 (d), 53.2 (t), 39.6 (d), 39.1 (d), 29.4 (t), 28.8 (t), 26.4 (t), 23.3 (t), 11.4 (q). HRMS (EI+, m/z): calcd. for C13H18NO2S [M – C2H5]

+: 252.1058; found: 252.1060. (+)-2-(4-(tert-Butyldiphenylsilyloxy)-2-ethylbutylsulfonyl)pyridine (40)

Following the general procedure for the asymmetric Cu-catalyzed conjugate addition, 40 was isolated in 91% yield. Enantiomeric excess: 92% determined by chiral HPLC analysis, Chiralcel OD-H, 0.5 mL/min, n-heptane: i-PrOH 97:3, 40 oC, 210 nm, retention

times (min): 31.4 (major) and 33.6 (minor). []D =

SO

ON

39

Page 26: University of Groningen Novel asymmetric copper-catalysed ... · 14 Chapter2-Final-nocode.docx Chapter 2 (Scheme 2). In this case simple organolithium reagents gave the best results

35

Chapter2-Final-nocode.docx

Catalytic Asymmetric Conjugate Addition of Grignard Reagents to α,β-Unsaturated Sulfones

+2.5 (c 1.0, CHCl3). 1H NMR (400 MHz, CDCl3): 8.68 (ddd, J = 4.6, 1.5, 0.7 Hz, 1H),

8.08 (br d, J = 7.8 Hz, 1H), 7.89 (dt, J = 7.8, 7.8, 1.7 Hz, 1H), 7.72-7.55 (m, 4H), 7.49 (ddd, J = 7.7, 4.7, 1.0 Hz, 1H), 7.46-7.31 (m, 6H), 3.66 (t, J = 6.3, 6.3 Hz, 2H), 3.42 (dq, J = 14.5, 14.5, 14.5, 6.1 Hz, 2H), 2.26-2.11 (m, 1H), 1.82-1.60 (m, 2H), 1.58-1.42 (m, 2H),

1.00 (s, 9H), 0.82 (t, J = 7.4 Hz, 3H). 13C NMR (100 MHz, CDCl3): 157.8 (s), 150.1 (d), 138.0 (d), 135.5 (d), 133.5 (s), 129.6 (d), 127.6 (d), 127.1 (d), 121.9 (d), 61.2 (t), 55.2 (t), 35.1 (t), 31.5 (d), 26.7 (q), 25.5 (t), 19.0 (s), 10.1 (q). HRMS (EI+, m/z): calcd. for C23H26NO3SiS [M–C4H9]

+: 424.1403; found: 424.1399. (+)-2-(2-Ethyl-4-phenylbutylsulfonyl)pyridine (41)

Following the general procedure for the asymmetric Cu-catalyzed conjugate addition, 41 was isolated in 91% yield. Enantiomeric excess: 93% determined by chiral HPLC analysis, Chiralpak AD-H, 0.5 mL/min, n-heptane: i-PrOH 97:3, 40 oC, 210 nm, retention times

(min): 60.1 (minor) and 62.4 (major). []D = +12.4 (c

1.0, CHCl3). 1H NMR (400 MHz, CDCl3): 8.74 (br d, J = 4.7 Hz, 1H), 8.05 (br d, J = 7.8

Hz, 1H), 7.93 (dt, J = 7.6, 7.6, 1.4 Hz, 1H), 7.53 (ddd, J = 7.6, 4.7, 0.8 Hz, 1H), 7.33-7.20 (m, 2H), 7.19-7.08 (m, 3H), 3.40 (dd, J = 6.1, 1.8 Hz, 2H), 2.60-2.54 (m, 2H), 2.08-1.97 (m, 1H), 1.90-1.65 (m, 2H), 1.66-1.46 (m, 2H), 0.87 (t, J = 7.4 Hz, 3H). 13C NMR (100

MHz, CDCl3): 157.8 (s), 150.2 (d), 141.7 (s), 138.1 (d), 128.3 (d), 128.3 (d), 127.2 (d), 125.8 (d), 122.0 (d), 55.0 (t), 34.4 (t), 33.7 (d), 32.3 (t), 25.6 (t), 10.0 (q). HRMS (ESI+, m/z): calcd. for C17H23NO2S [M + H]+: 304.13658; found: 304.13639.

Page 27: University of Groningen Novel asymmetric copper-catalysed ... · 14 Chapter2-Final-nocode.docx Chapter 2 (Scheme 2). In this case simple organolithium reagents gave the best results

36

Chapter2-Final-nocode.docx

Chapter 2

2.6 References

(1) Perlmutter, P. In Conjugate Addition Reactions in Organic Synthesis; Tetrahedron Organic Chemistry Series 9; Pergamon Press, Oxford, U.K., 1992.

(2) Rossiter, B.; Swingle, N. Chem. Rev. 1992, 92, 771. (3) Tomioka, K.; Nagaoka, Y. In Comprehensive Asymmetric Catalysis; Jacobsen, E. N., Pfaltz, A.

and Yamamoto, H., Eds.; Springer: New York, 1999, Vol. 3, 1105. (4) Feringa, B. L. Acc. Chem. Res. 2000, 33, 346. (5) Krause, N.; Hoffmann-Roder, A. Synthesis 2001, 171. (6) Feringa, B. L.; Naasz, R.; Imbos, R.; Arnold, L. A. In Modern Organocopper Chemistry;

Krause, N., Ed; VCH: Weinheim, Germany, 2002, 224. (7) Alexakis, A.; Benhaim, C. Eur. J. Org. Chem. 2002, 3221. (8) Hayashi, T.; Yamasaki, K. Chem. Rev. 2003, 103, 2829. (9) Woodward, S. Angew. Chem. Int. Ed. 2005, 44, 5560. (10) López, F.; Minnaard, A. J.; Feringa, B. L. Acc. Chem. Res. 2007, 40, 179. (11) López, F.; Minnaard, A. J.; Feringa, B. L. In The Chemistry of Organomagnesium Compounds;

Rappoport, Z.; Marek, I., Eds.; Wiley: Chichester, U.K., 2008; Part 2, Chapter 17. (12) Harutyunyan, S. R.; den Hartog, T.; Geurts, K.; Minnaard, A. J.; Feringa, B. L. Chem. Rev.

2008, 108, 2824. (13) Barton, D.; Ollis, W. D. Comprehensive Organic Chemistry; Pergamon Press: Oxford, U.K.,

1979. (14) Trost, B. M. Bull. Chem. Soc. Jpn. 1988, 61, 107. (15) Prilezhaeva, E. N. Russ. Chem. Rev. 2000, 69, 367. (16) Patai, S.; Rappoport, L.; Stirling, C. J. M., Eds.; The Chemistry of Sulfoxides and Sulfones; John

Wiley & Sons: Chichester, UK, 1988. (17) Kresze, G. Methoden der Organischen Chemie (Houben-Weyl) 1985, 669. (18) Kelly, S. E. In Comprehensive Organic Synthesis; Trost, B. M., Fleming, I., Eds.; Pergamon

Press: Oxford, U.K., 1991; Vol. 1, 792. (19) Llamas, T.; Gómez Arrayás, R.; Carretero, J. C. Angew. Chem. Int. Ed. 2007, 46, 3329. (20) Llamas, T.; Gómez Arrayás, R.; Carretero, J. C. Angew. Chem. 2007, 119, 3393. (21) Fuchs, P. L.; Braish, T. F. Chem. Rev. 1986, 86, 903. (22) Nigel S., S. Tetrahedron 1990, 46, 6951. (23) Hardinger, S. A.; Fuchs, P. L. J. Org. Chem. 1987, 52, 2739. (24) Hutchinson, D. K.; Fuchs, P. L. J. Am. Chem. Soc. 1987, 109, 4755. (25) Carretero, J. C.; Dominguez, E. J. Org. Chem. 1993, 58, 1596. (26) Rendler, S.; Oestreich, M. Angew. Chem. 2007, 119, 504. (27) Lipshutz, B. H.; Servesko, J. M.; Taft, B. R. J. Am. Chem. Soc. 2004, 126, 8352. (28) Lipshutz, B. H.; Tanaka, N.; Taft, B. R.; Lee, C. -T. Org. Lett. 2006, 8, 1963. (29) Rainka, M. P.; Aye, Y.; Buchwald, S. L. Proc. Natl. Acad. Sci. U. S. A. 2004, 101, 5821. (30) Moritani, Y.; Appella, D. H.; Jurkauskas, V.; Buchwald, S. L. J. Am. Chem. Soc. 2000, 122,

6797. (31) Yun, J.; Buchwald, S. L. Org. Lett. 2001, 3, 1129.

Page 28: University of Groningen Novel asymmetric copper-catalysed ... · 14 Chapter2-Final-nocode.docx Chapter 2 (Scheme 2). In this case simple organolithium reagents gave the best results

37

Chapter2-Final-nocode.docx

Catalytic Asymmetric Conjugate Addition of Grignard Reagents to α,β-Unsaturated Sulfones

(32) Lipshutz, B. H.; Servesko, J. M.; Petersen, T. B.; Papa, P. P.; Lover, A. A. Org. Lett. 2004, 6, 1273.

(33) Hughes, G.; Kimura, M.; Buchwald, S. L. J. Am. Chem. Soc. 2003, 125, 11253. (34) Lee, D.; Kim, D.; Yun, J. Angew. Chem. 2006, 118, 2851. (35) Czekelius, C.; Carreira, E. M. Org. Lett. 2004, 6, 4575. (36) Czekelius, C.; Carreira, E. M. Angew. Chem. 2003, 115, 4941. (37) Simpkins, N. S. In Sulphones in Organic Chemistry; Tetrahedron Organic Chemistry Series 10;

Pergamon Press: Oxford, U.K., 1993. (38) Mauleón, P.; Carretero, J. C. Org. Lett. 2004, 6, 3195. (39) Mauleón, P.; Carretero, J. C. Chem. Commun. 2005, 4961. (40) Desrosiers, J. –N.; Charette, A. B. Angew. Chem. Int. Ed. 2007, 46, 5955. (41) Mauleón, P.; Carretero, J. C. Org. Lett. 2004, 6, 3195. (42) Hayashi, T.; Yamasaki, K. Chem. Rev. 2003, 103, 2829. (43) Itami, K.; Mitsudo, K.; Nokami, T.; Kamei, T.; Koike, T.; Yoshida, J. -I. J. Organomet. Chem.

2002, 653, 105. (44) Mauleón, P.; Carretero, J. C. Chem. Commun. 2005, 4961. (45) Desrosiers, J. -N; Bechara, W. S.; Charette, A. B. Org. Lett. 2008, 10, 2315. (46) Mao, B.; Geurts, K.; Fañanás-Mastral, M.; Van Zijl, A. W.; Fletcher, S. P.; Minnaard, A. J.;

Feringa, B. L. Org. Lett. 2011, 13, 948. (47) Mauleón, P.; Alonso, I.; Rivero, M. R.; Carretero, J. C. J. Org. Chem. 2007, 72, 9924. (48) Mauleón, P.; Carretero, J. C. Org. Lett. 2004, 6, 3195. (49) Wnuk, S. F.; Garcia, P. I.; Wang, Z. Org. Lett. 2004, 6, 2047.

Page 29: University of Groningen Novel asymmetric copper-catalysed ... · 14 Chapter2-Final-nocode.docx Chapter 2 (Scheme 2). In this case simple organolithium reagents gave the best results

38

Chapter2-Final-nocode.docx

Chapter 2