• No results found

University of Groningen Exploring asymmetric catalytic transformations Guduguntla, Sureshbabu

N/A
N/A
Protected

Academic year: 2021

Share "University of Groningen Exploring asymmetric catalytic transformations Guduguntla, Sureshbabu"

Copied!
21
0
0

Bezig met laden.... (Bekijk nu de volledige tekst)

Hele tekst

(1)

Exploring asymmetric catalytic transformations Guduguntla, Sureshbabu

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: 2017

Link to publication in University of Groningen/UMCG research database

Citation for published version (APA):

Guduguntla, S. (2017). Exploring asymmetric catalytic transformations. University of Groningen.

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.

(2)

1

(3)

2

The word chirality is derived from the Greek word for hand, χείρ (kheir). When a mirror image of an object or a system is not superimposable on its original it is called chiral. A chiral object and its mirror image are called enantiomorphs (Greek opposite forms) or, when referring to molecules, enantiomers. Enantiomers are also called optical isomers, which is due to their ability to rotate the plan of polarised light. In 1815,1 Jean-Baptiste Biot observed the rotation of plain polarised light in certain liquids and vapours of organic substances such as turpentine. Depending on which direction a chiral molecule rotate the plain polarised light, it is designated as d (clockwise; dextrorotatory) or l (anti-clockwise; levorotatory). In 1849,2,3 Louis Pasteur performed a famous experiment on the resolution of tartaric acids. He sorted the crystals by hand which gave two forms of the compound: solutions of one form rotate polarized light clockwise, while the other form rotates light anti-clockwise. In 1874,4 Jacobus Henricus van't Hoff and Joseph Achille Le Bel5 independently proposed that this phenomenon of optical activity in carbon compounds is due to the three-dimensional geometry of carbon- based compound when a carbon atom is connected with four different groups in a tetrahedral structure. This theory leads to another notation of chiral molecules based on the Cahn-Ingold-Prelog rules. By using these rules the absolute configuration of a molecule is assigned either R or S (Figure 1).6

Figure 1: (S)-(+)-Lactic acid (left) and (R)-(–)-lactic acid (right) are

non-superimposable mirror images of each other.

One of the most fundamental aspect of all the living organisms is self-replication. Self-replication arises from biologically active molecules like DNA and RNA, which have the ability to self-organize and interact dynamically in complex molecular networks. Therefore, considerable experimental efforts have been reported in attempts to elucidate how life began by mimicking prebiotic systems.7,8,9,10,11,12 Several mechanisms

(4)

3

have been proposed using self-replication as a key process to explain prebiotic chemical evolution of life in terms of pre-“RNA world” and “systems chemistry”, etc.7-12,13

In addition to self-replication, the essential biological molecules possess homochirality, which is also another key aspect of life on the Earth. Sugars such as deoxyribose and ribose are present in the most important biologically active molecules like DNA and RNA, which transfer the genetic information in living organisms, are right-handed. Amino acids that form the proteins that are essential for the structure and chemical transformations in cells are left-handed. The origin of homochirality in biomolecules is still one of the great current mysteries which is directly associated with the ˋorigin of lifeˊ question.14 There are several hypotheses reported in an attempt to explain possible prebiotic chemical pathways towards the asymmetric formation of the smallest chiral biomolecules.15 From a chemical point of view, molecular self-replication processes function via autocatalytic, cross catalytic or collectively catalytic pathways with an additional information transfer (templating) from the product to the reactants (Scheme 1).7-13,15 It was suggested that, in the earliest stages of life, a small enantiomeric excess might be amplified via auto- or cross- catalytic processes.15

Scheme1: a) Autocatalytic reaction, b) Cross catalytic reaction.

Up to now the asymmetric autocatalytic reaction reported by the group of Soai16 is the only reliable example known and its mechanistic aspects are still not fully understood.17 Therefore it would be a real challenge to find other reactions which show asymmetric autocatalytic behaviour.18

1.1 Design of a self-replicating system

In a self-replicating system, a molecule C acts as a template for the starting materials A and B that are bound by noncovalent interactions to

(5)

4

form a termolecular complex [A.C.B]. This complex allows components A and B to organize in such a way that they can react to form a copy of the original template molecule C. In this way an autocatalytic synthesis is accomplished (Scheme 2).

Scheme 2: General representation of a self-replicating system.

In 1986, von Kiedrowski19 reported the first non-enzymatic self-replicating system based on the formation of a hexa eoxynucleotide from two trideoxynucleotides. This system was designed based on the oligonucleotide (template) synthesis from monomers; reported by Orgel

et al.20 Studies21 on the von Kiedrowski system reveals that the highest efficiency based on a template mechanism was observed when the trimers and hexamer are completely complimentary to each other. Orgel

et al.22 reported a similar system by taking advantage of phosphoramidate backbone formation for the synthesis of a tetranucleotide template from dinucleotides. These early reported self-replicating systems suffered from low efficiency due to a strong binding between the product dimer, resulting in catalyst inhibition.23 By using the concept of phosphoramidate formation von Kiedrowski et al.24 reported a modified system (Scheme 3). In the presence of EDC, the trideoxynucleotide 3'-phosphate 1 reacts with the 5'-amino trideoxynucleotide 2 to give the hexameric 3'-5'-phosphoramidate 3. Compound 3 acts as a self-complementary unit for its own formation through a trimeric complex 4. This complex 4 allows components 1 and 2 to form a network of H-bonding with 3 and to organize in such a way that they can react to form a copy of the original template molecule 3. In this way an autocatalytic

(6)

5

process is accomplished. For the first time this system showed a sigmoidal (S-shaped) growth of template molecules with increase in efficiency by 75-folds compared with previously reported systems.19,22

Scheme 3: von Kiedrowski system using phosphoramidate chemistry.24 Achieving complete exponential replication in oligonucleotide based self-replicating systems has proven to be difficult.25 Exponential growth has been gained by attaching the oligonucleotides to a solid surface and

(7)

6

subsequently manually melting and annealing the strands, thereby avoiding product inhibition.26 There are many types of autocatalytic self-replicating systems based on peptides, small molecules etc. which are well reviewed by Bissette and Fletcher.11

It is worth to mention that in all of these synthetic self-replicating systems developed so far, homochirality has been predefined with the precursors employed being homochiral, i.e. no asymmetric centre is formed during the reaction. It was specified earlier that, biologically important molecules such as amino acids and sugars are homochiral in nature. So it is very important in this context to elucidate the origin of homochirality on earth.

1.2 Asymmetric autocatalysis and autoinduction.

For the first time the term asymmetric autocatalysis was introduced by Wynberg, who very early acknowledged the great potential of a system where the chiral product affects the stereoselectivity of that reaction.27 In 1989, Wynberg hypothesized that asymmetric autocatalysis would represent the next generation in asymmetric catalysis.27,28

1.2.1 Asymmetric autoinduction

In a chemical process where a chiral product that does not possess catalytic activity participates in the formation of a chiral complex with the enantioselective catalyst to effect the stereochemical outcome is called asymmetric autoinduction (Scheme 4).27,29

Scheme 4: Asymmetric autoinduction.

In asymmetric autoinduction, the product of the reaction modifies the course of the reaction by changing the nature of the reagent or the

(8)

7

catalyst. The autoinduction can involve the formation of a product different from, or identical to, the catalyst. Asymmetric autocatalysis involves only the latest.

Wynberg et al.27,28 reported the first example of asymmetric autoinduction, in the nucleophilic addition of ethyllithium to benzaldehyde (Scheme 5). The addition of stoichiometric enantioenriched product of the reaction (6, deuterated to distinguish it from the alcohol product) was found to influence the enantioselectivity of the reaction. The ee of the product 6 was found to be 17% in favour of the isomer with the same absolute configuration as that of added (Scheme 5).27,28

Scheme 5: Proof of asymmetric autoinduction in nucleophilic addition of EtLi to

benzaldehyde using stoichiometric amount of product.

However, the same effect was also demonstrated in a catalytic process (Scheme 6).27 The addition of diethylzinc to benzaldehyde is not a spontaneous reaction, but can be catalyzed by orthotitanates. When such a catalyst was prepared from TiCl4 and 6-d1, and used in the addition of

diethylzinc to benzaldehyde, the product was produced in 32% ee again with the same absolute configuration as for the Ti ligands. There are many other reactions which shows asymmetric autoinduction, which have been reviewed by Todd.30

Scheme 6: Asymmetric autoinduction in the addition of Et2Zn to benzaldehyde using catalytic amount of the Ti-complex of product.

(9)

8 1.2.2 Asymmetric autocatalysis

Biologically important molecules such as L-amino acids and D-sugars are the main carriers of homochirality in living organisms.31 One of the widely accepted hypothesis for the origin of homochirality is astrochemically relevant and based on the fact that only an excess of left-handed L-amino acids have been found in carbonaceous meteorites.32 Preferential excess of L-amino acid found in carbonaceous meteorites may indicate that the origin of life on the earth has an extraterrestrial input. But to get homochirality in life there would need to be some mechanism of transferring the single-handedness between different types of amino acids. An important question is how this initial preference was further amplified.

It has also been suggested that a weak electronic forces based on the principle of parity violation could be responsible for the symmetry breaking.33 However, it has often been argued that this force is too small to be effective.34 Another source of symmetry breaking is circular polarized light (CPL). CPL is capable of activating one enantiomer more than the other resulting in faster deterioration of this enantiomer, resulting in a small enantiomeric excess. In the universe large areas exist where CPL is present.35 This might be the origin of a small enantiomeric excess in organic compounds present on meteorites.36

Although these described processes of initial symmetry breaking lead to a small enantiomeric excess in a specific area, this is however not likely to be sufficient to induce homochirality in all living systems. Therefore these small initial enantiomeric excesses need to be amplified. One of the earliest hypothesis is that asymmetric autocatalysis11,16,37 (Scheme 1a) played an important role in the amplification of chirality. A reaction is classified as an autocatalytic reaction if the product acts as a catalyst for its own formation (Scheme 1a). If the product of an autocatalytic reaction is chiral and a selective catalyst, meaning that one enantiomer of the product catalyzes for its own production, an asymmetric autocatalytic reaction is obtained.

(10)

9

Many non-asymmetric autocatalytic reactions are known, one of the familiar example is the acid catalyzed hydrolysis of an ester.11,16,38 As a greater amount of ester gets hydrolyzed more free acid is present in the solution to catalyze the hydrolysis, providing an autocatalytic reaction.38 Additionally there have been a wide range of autocatalytic reactions to amplify an initially small signal.39 In general it was assumed that (asymmetric) autocatalytic reactions have played an important role in the origin of life. Autocatalysis provides a tool to obtain relevant compounds in sufficient amounts. Furthermore an autocatalytic reaction can be viewed as one of the simplest chemical metabolic cycles. Additionally as mentioned above, the asymmetric variant provides a tool to amplify not only the relevant compounds but also its chirality.

In 1953 Frank et al.37a proposed a theoretical model to understand the concept of amplification of chirality by means of asymmetric autocatalysis. According to his model, in order to achieve asymmetric amplification, the process of self-replication of a catalyst must be accompanied by the suppression of the activity of its enantiomer, referred to by Frank as mutual antagonism (Figure 2).37a

Figure 2: Schematic representation of Frank’s model.

R and S are the enantiomers which can act as autocatalysts for their own formation, by reacting with molecules of substrate.40 Together with this autocatalytic process, there is a mutual antagonism between R and S such that, when they react together, they deactivate and lose their self-replicating property. As shown in Figure 2, the self-replication of the enantiomers cause a change in the R:S ratio, which grows as long as there was a small initial imbalance at the beginning of the process. The cooperation of the autocatalysis and mutual antagonism propagate and

(11)

10

amplify the imbalance of the enantiomers. If the substrate pool is large enough, the process can be sustained and the selectivity towards the self-production of one enantiomer will dominate.41

For the first time as a proof of concept, Soai et al.42 (Scheme 7) reported an asymmetric autocatalytic reaction in 1995, which involves the nucleophilic addition of diisopropyl zinc to pyrimidine-5-carbaldehyde. This process satisfied the principles for an asymmetric autocatalytic reaction as proposed by Frank et al.37a In this reaction, the starting mixture was seeded with product that contained as little as 0.00005% ee, which allowed, after several cycles, each time using the slightly more enriched product of the previous round as seed, for the product to be obtained in >99.5% ee (Scheme 7).43

Scheme 7: Soai reaction.

Subsequently Soai et al. have shown that besides an initial enantiomeric excess also other chiral factors such as sodium chlorate salts,44 helicenes,45 quartz crystals,46 CPL,47 single wall carbon nanotubes48 and even isotopes of hydrogen,49 carbon,50 oxygen,51 and nitrogen52 are capable of directing the chirality of the Soai reaction.53 Carreira et al.54 demonstrated the versatility of the Soai reaction in synthetic chemistry, to produce the drug Efivarenze. By using an autocatalytic reaction, the transformation is not only catalyzed but because the catalyst is also the product the purification of the product becomes more efficient.

There are several attempts reported to understand the mechanism of Soai reaction but so far there is no clear explanation.55 Initially it was hypothesized that the Soai reaction operates with very high selectivity due to the formation of homochiral and heterochiral dimers. It was suggested that the homochiral dimers are capable of catalyzing the

(12)

11

reaction and all racemic product is inactive, which leads to the perfect selectivity in the Soai reaction (Scheme 8a).55c,d Later it was shown via DFT calculations that both dimers have similar reactivity, therefore proving that the dimer formation does not sufficiently explain the selectivity.55e Additional kinetic and 1H-NMR experiments suggested the formation of a tetrameric complex 9. During the enantioselective reaction, the association of this complex with the aldehyde forms a trimeric complex 10 (Scheme 8b).55f-i

Scheme 8: Proposed intermediates in Soai reaction.

Use of diisopropyl zinc makes the Soai reaction highly reactive towards water and therefore not very likely to have occurred under prebiotic conditions. Moreover the Soai reaction is in itself very interesting, but this transformation did not lead to any biological relevant molecules. Therefore, it remains very challenging and highly important to develop an autocatalytic reaction that can yield biologically relevant molecules, and to investigate if these reactions show enantioselectivity.

1.3 Cu-catalyzed asymmetric allylic substitution.

Transition metal catalyzed asymmetric transformation is a highly efficient tool for the construction of chiral molecules with control over

(13)

12

the absolute configuration, which is often a key factor determining the activity of pharmaceuticals.56 This strategy presents various advantages such as atom economy, reduction of by-product formation, functional group tolerance and the capability to provide the desired product with high levels of selectivity. Several transition metals have proven their efficiency in catalytic asymmetric processes.57 In particular, the use of copper provides cheaper and readily available catalyst systems while maintaining high activity and selectivity. Among the Cu-catalyzed enantioselective processes for asymmetric C-C bond formation, asymmetric allylic substitution (AAS) represents a useful and powerful methodology for the synthesis of chiral building blocks, in particular for the synthesis of pharmaceuticals and natural products.57,58 This transformation generally leads to the SN2′-functionalised product,

through the reaction of a nucleophile and an allylic system, bearing a suitable leaving group, and provides a carbon-based stereocenter next to a terminal olefin (Scheme 9). This alkene moiety can be further converted into a variety of functionalities while preserving the stereochemical integrity of the starting material.59

Scheme 9: Cu-catalyzed AAS protocols.

Cu-catalyzed AAS represents a highly efficient catalytic process for the use of organometallic reagents, such as organozinc,60 organomagnesium,61 organoaluminium,62 organozirconium,63 organolithium64 compounds, etc.65,66,67 The AAS reactions, with these “hard” C-nucleophiles, are complementary to the well-known Pd-catalyzed asymmetric allylic substitutions which employs soft and stabilized nucleophiles.68 In Cu-catalyzed AAS, the organometallic

(14)

13

compound first undergoes transmetallation to a Cu (I) complex, followed by π-complex formation and subsequent oxidative addition to form a Cu (III) σ-allyl complex. This is also the enantiodiscriminating step. Finally, reductive elimination gives rise to the new C–C bond and the branched (SN2′) product (Scheme 10).69

Scheme 10: General mechanism for the Cu-catalyzed AAS reaction.

Detailed introduction about the various AAS reactions is discussed in Chapters 2, 3 and 4.

1.4 Thesis outline

In these thesis two main areas of research work has been described. The first half of the thesis involves Cu-catalyzed asymmetric allylic substitution reactions with organolithium reagents. The second half of the thesis describes efforts towards the development of new asymmetric autocatalytic reactions.

In Chapter 2, the synthesis of optically active β- or γ-alkyl substituted alcohols through copper-catalyzed asymmetric allylic alkylation with organolithium reagents are reported.

In Chapter 3, a route to chiral diarylmethanes via copper-catalyzed asymmetric allylic arylation with organolithium compounds is described.

In Chapter 4, the enantioselective synthesis of di- and tri-arylated all-carbon quaternary stereocenters via copper-catalyzed allylic arylations with organolithium compounds are presented.

(15)

14

In Chapter 5, efforts towards the development of a new asymmetric autocatalytic reaction based on nucleophilic addition of dialkyl phosphites to aldehydes are described. The goal was to take advantage of the aggregation behavior of metal alkoxides and to use them as catalysts to promote their own synthesis in a nucleophilic addition reaction.

In Chapter 6, efforts towards the development of a new asymmetric autocatalytic reaction based on a metal-ligand approach is discussed.

In Chapter 7, efforts towards the development of new asymmetric autocatalytic reactions based on H-bond donor approach and of an imine based autocatalysis using urea motifs, are discussed.

1.5 References

1) Biot, J. B. Bulletin des Sciences, par la Société Philomatique de Paris, 1815, 190. 2) Pasteur, L. Comp. Rend. Paris 1848, 26, 535.

3) Pasteur, L. Bull. Soc. Chim. Fr. 1884, 41, 215.

4) Van't Hoff, J. H. Arch. Neerl. Sci. Exactes Nat. 1874, 9, 445. 5) Le Bell, J. A. Bull. Soc. Chim. Fr. 1874, 22, 337.

6) Clayden, J.; Greeves, N.; Warren, S.; Wothers, P. Organic Chemistry, Oxford University Press, Oxford, 2001.

7) Eschenmoser, A. Tetrahedron 2007, 63, 12821.

8) Feringa, B. L.; van Delden, R. A. Angew. Chem. Int. Ed. 1999, 38, 3418. 9) Hein, J. E.; Blackmond, D. G. Acc. Chem. Res. 2012, 45, 2045.

10) Ruiz-Mirazo, K.; Briones, C.; de la Escosura, A. Chem. Rev. 2014, 114, 285. 11) Bissette, A. J.; Fletcher, S. P. Angew. Chem. Int. Ed. 2013, 52, 12800.

12) a) Peyralans, J. J.; Otto, S. Curr. Opin. Chem. Biol. 2009, 13, 705. b) Lee, D. H.; Severin, K.; Yokobayashi, Y.; Ghadiri, M. R. Nature 1997, 390, 591. c) Paul, N.; Joyce, G. F. Curr. Opin. Chem. Biol. 2004, 8, 634. d) Joyce, G. F. Nature 2002, 418, 214.

13) a) Kindermann, M.; Stahl, I.; Reimold, M.; Pankau, W. M.; von Kiedrowski, G. Angew. Chem. 2005, 117, 6908. b) Kauffman, S. A. The origins of order: self-organization and selection in evolution; Oxford University Press: New York, 1993. c) Eigen, M. The hypercycle, a principle of natural self-organization; Springer-Verlag: Berlin ; New York, 1979.

(16)

15

15) a) Blackmond, D. G. Phil. Trans. R. Soc. B 2011, 366, 2878. b) Budin, I.; Szostak, J. W. Annu. Rev. Biophys. 2010, 39, 245.

16) a) Amplification of Chirality (Soai, K., Ed.), Springer-Verlag, Berlin, 2008. b) Soai, K.; Shibata, T.; Morioka, H.; Choji, K. Nature 1995, 378, 767. c) Shibata, T.; Morioka, H.; Hayase, T.; Choji, K.; Soai, K. J. Am. Chem. Soc. 1996, 118, 471. 17) a) Buono, F. G.; Blackmond, D. G. J. Am. Chem. Soc. 2003, 125, 8978. b)

Blackmond, D. G.; McMillan, C. R.; Ramdeehul, S.; Schorm, A.; Brown, J. M. J. Am. Chem. Soc. 2001, 123, 10103. c) Blackmond, D. G. Proc. Natl. Acad. Sci. USA

2004, 101, 5732. d) Islas, J. R.; Lavabre, D.; Grevy, J. –M.; Lamoneda, R. H.;

Cabrera, H. R.; Micheau, J. –C.; Buhse, T. Proc. Natl. Acad. Sci. USA 2005, 102, 13743. e) Ercolani, G.; Schiaffino, L. J. Org. Chem. 2011, 76, 2619.

18) a) Mauksch, M.; Tsogoeva, S. B.; Martynova, I. M.; Wei, S. Angew. Chem. Int. Ed.

2007, 46, 393. b) Tsogoeva, S. B. Chem. Commun. 2010, 46, 7662. c) Mauksch,

M.; Tsogoeva, S. B.; Wei, S.; Martynova, I. M. Chirality 2007, 19, 816. d) Blackmond, D. G. Angew. Chem. Int. Ed. 2009, 48, 2648. e) Wang, X.; Zhang, Y.; Tan, H.; Wang, Y.; Han, P.; Wang, D. Z. J. Org. Chem. 2010, 75, 2403. f) Amedjkouh, M.; Brandberg, M. Chem. Commun. 2008, 3043.

19) von Kiedrowski, G. Angew. Chem. Int. Ed. Engl. 1986, 25, 932. 20) Inoue, T.; Orgel, L. E. Science 1983, 219, 859.

21) von Kiedrowski, G.; Wlotzka, B.; Helbing, J. Angew. Chem. Int. Ed. Engl. 1989, 28, 1235.

22) Zielinski, W. S.; Orgel, L. E. Nature 1987, 327, 346. 23) von Kiedrowski, G. Bioorg. Chem. Front. 1993, 3, 113.

24) von Kiedrowski, G.; Wlotzka, B.; Helbing, J.; Matzen, M.; Jordan, S. Angew. Chem. Int. Ed. Engl. 1991, 30, 423.

25) a) Patzke, V.; von Kiedrowski, G. Archive for Organic Chemistry 2007, v, 293. 26) Luther, A.; Brandsch, R.; von Kiedrowski, G. Nature 1998, 396, 245.

27) a) Alberts, A. H.; Wynberg, H. J. Am. Chem. Soc. 1989, 111, 7265. b) H. Wynberg Chimia, 1989, 43, 150.

28) Wynberg, H. J. Macromol. Sci. Part - Chem. 1989, 26, 1033. 29) Wynberg, H.; Feringa, B. Tetrahedron 1976, 32, 2831. 30) Todd, M. H. Chem. Soc. Rev. 2002, 31, 211.

31) Fuß, W. Chirality 2009, 21, 299.

32) a) Sephton, M. A. Nat. Prod. Rep. 2002, 19, 292. b) Pizzarello, S.; Cronin, J. Geochim. Cosmochim. Acta 2000, 64, 329. c) Glavin, D. P.; Dworkin, J. P. Proc. Natl. Acad. Sci. 2009, 106, 5487. d) Pizzarello, S.; Huang, Y.; Alexandre, M. R. Proc. Natl. Acad. Sci. 2008, 105, 3700. e) Miller, S. L.; Urey, H. C. Science 1959, 130, 245. f) Miller, S. L. Science 1953, 117, 528.

(17)

16

33) a) Mason, S. Nature 1984, 311, 19-23. b) Quack, M. Angew. Chem. Int. Ed. 2002, 41, 4618.

34) Bonner, W. Chirality 2000, 12, 114-126.

35) Bailey, J.; Chrysostomou, A.; Hough, J.; Gledhill, T.; McCall, A.; Clark, S.; Menard, F.; Tamura, M. Science 1998, 281, 672.

36) Pizzarello, S.; Cronin, J. R. Geochim. Cosmochim. Acta 2000, 64, 329. b) Cronin, J. R.; Pizzarello, S. Science 1997, 275, 951.

37) a) Frank, F. C. Biochim. Biophys. Acta 1953, 11, 459. b) Wynberg, H. Chimia

1989, 43, 150.

38) Jogunola, O.; Salmi, T.; Eränen, K.; Wärna, J.; Mikkola, J. -P. Chem. Eng. Process

2011, 50, 665.

39) Scrimin, P.; Prins, L. J. Chem. Soc. Rev. 2011, 40, 4488.

40) Blackmond, D. G. Philos. Trans. R. Soc. B Biol. Sci. 2011, 366, 2878.

41) A non-catalytic system which fulfil the theory of Frank’s model: Steendam, R. R. E.; Verkade, J. M. M.; van Benthem, T. J. B.; Meekes, H.; van Enckevort, W. J. P.; Raap, J.; Rutjes, F. P. J. T.; Vlieg, E. Nat. Commun. 2014, 5, 5543.

42) (a) Soai, K.; Shibata, T.; Morioka, H.; Choji, K. Nature 1995, 378, 767. (b) Amplification of Chirality (Soai, K., ed.) Springer-Verlag, Berlin, 2008. (c) Shibata, T.; Morioka, H.; Hayase, T.; Choji, K.; Soai, K. J. Am. Chem. Soc. 1996, 118, 471.

43) Sato, I.; Urabe, H.; Ishiguro, S.; Shibata, T.; Soai, K. Angew. Chem. Int. Ed. 2003, 42, 315.

44) Sato, I.; Kadowaki, K.; Soai, K. Angew. Chem. Int. Ed. 2000, 39, 1510.

45) Sato, I.; Yamashima, R.; Kadowaki, K.; Yamamoto, J.; Shibata, T.; Soai, K. Angew. Chem. Int. Ed. 2001, 40, 1096.

46) Soai, K.; Osanai, S.; Kadowaki, K.; Yonekubo, S.; Shibata, T.; Sato, I. J. Am. Chem. Soc. 1999, 121, 11235.

47) a) Shibata, T.; Yamamoto, J.; Matsumoto, N.; Yonekubo, S.; Osanai, S.; Soai, K. J. Am. Chem. Soc. 1998, 120, 12157. b) Kawasaki, T.; Sato, M.; Ishiguro, S.; Saito, T.; Morishita, Y.; Sato, I.; Nishino, H.; Inoue, Y.; Soai, K. J. Am. Chem. Soc. 2005, 127, 3274.

48) Hitosugi, S.; Matsumoto, A.; Kaimori, Y.; Iizuka, R.; Soai, K.; Isobe, H. Org. Lett.

2014, 16, 645.

49) Kawasaki, T.; Ozawa, H.; Ito, M.; Soai, K. Chem. Lett. 2011, 40, 320.

50) Kawasaki, T.; Matsumura, Y.; Tsutsumi, T.; Suzuki, K.; Ito, M.; Soai, K. Science

2009, 324, 492.

51) Kawasaki, T.; Okano, Y.; Suzuki, E.; Takano, S.; Oji, S.; Soai, K. Angew. Chem. Int. Ed. 2011, 50, 8131.

52) Matsumoto, A.; Ozaki, H.; Harada, S.; Tada, K.; Ayugase, T.; Ozawa, H.; Kawasaki, T.; Soai, K. Angew. Chem. Int. Ed. 2016, 55, 15246.

(18)

17

53) a) Soai, K.; Kawasaki, T.; Matsumoto, A. Acc. Chem. Res. 2014, 47, 3643. b) Matsumoto, A.; Kaimori, Y.; Uchida, M.; Omori, H.; Kawasaki, T.; Soai, K. Angew. Chem. Int. Ed. 2016, 55, 1.

54) Chinkov, N.; Warm, A.; Carreira, E. M. Angew. Chem. Int. Ed. 2011, 50, 2957. 55) a) Sato, I.; Omiya, D.; Igarashi, H.; Kato, K.; Ogi, Y.; Tsukiyama, K.; Soai, K.

Tetrahedron Asymmetry 2003, 14, 975. b) Buhse, T. Tetrahedron Asymmetry 2003, 14, 1055. c) Blackmond, D. G.; McMillan, C. R.; Ramdeehul, S.; Schorm, A.; Brown, J. M. J. Am. Chem. Soc. 2001, 123, 10103. d) Blackmond, D. G. Adv. Synth. Catal. 2002, 344, 156. e) Gridnev, I. D.; Serafimov, J. M.; Brown, J. M. Angew. Chem. Int. Ed. 2004, 43, 4884. f) Buono, F. G.; Blackmond, D. G. J. Am. Chem. Soc. 2003, 125, 8978. g) Blackmond, D. G. Proc. Natl. Acad. Sci. 2004, 101, 5732. h) Quaranta, M.; Gehring, T.; Odell, B.; Brown, J. M.; Blackmond, D. G. J. Am. Chem. Soc.2010, 132, 15104. i) Gehring, T.; Quaranta, M.; Odell, B.; Blackmond, D. G.; Brown, J. M. Angew. Chem. Int. Ed. 2012, 51, 9539. j) Ercolani, G.; Schiaffino, L. J. Org. Chem. 2011, 76, 2619.

56) a) Asymmetric catalysis on industrial scale; Blaser, H. U., Federsel, H. -J., Eds.; Wiley, Weinheim, 2010. b) Catalytic asymmetric synthesis, 2nd Edn; Ojima, I., Ed.; Wiley, New York, 2000. c) Comprehensive asymmetric catalysis I–III; Jacobsen, E. N., Pfaltz, A., Yamamoto, H., Eds.; Springer, Berlin, 1999.

57) Catalytic asymmetric synthesis; Ojima, I., Ed.; Wiley, Weinheim, 2013.

58) a) Karlström, A. S. E.; Bäckvall, J. E. Copper-Mediated Enantioselective Substitution Reactions. In Modern Organocopper Chemistry; Krause, N., Ed.; Wiley, Weinheim, 2002, pp 259–288. b) Yorimitsu, H.; Oshima, K. Angew. Chem. Int. Ed. 2005, 44, 4435. c) Geurts, K.; Fletcher, S. P.; van Zijl, A. W.; Minnaard, A. J.; Feringa, B. L. Pure. Appl. Chem. 2008, 80, 1025. d) Harutyunyan, S. R.; den Hartog, T.; Geurts, K.; Minnaard, A. J.; Feringa, B. L. Chem. Rev. 2008, 108, 2824. e) Alexakis, A.; Bäckvall, J. E.; Krause, N.; Pamies, O.; Dieguez, M. Chem Rev. 2008, 108, 2796. f) Falciola, C. A.; Alexakis, A. Eur. J. Org. Chem. 2008, 22, 3765. g) Langlois, J. B.; Alexakis, A. Top. Organomet. Chem. 2012, 38, 235. h) Copper-catalyzed asymmetric synthesis; Alexakis, A., Krause, N., Woodward, S., Eds.; Wiley, Weinheim, 2014. i) Hornillos, V.; Gualtierotti, J. -B.; Feringa, B. L. Top. Organomet. Chem. 2016, 58, 1.

59) (a) Zhang, A.; Rajanbabu, T. V. Org. Lett. 2004, 6, 3159. (b) Fuganti, C.; Serra, S.; Dulio, A. J. Chem. Soc. Perkin Trans. 1 1999, 279. (c) Grassi, D.; Alexakis, A. Adv. Synth. Catal. 2015, 357, 3171. (d) Huo, S.; Negishi, E. Org. Lett. 2001, 3, 3253. (e) Kondakov, D. Y.; Negishi, E. J. Am. Chem. Soc. 1995, 117, 10771. (f) Kondakov, D. Y.; Negishi, E. J. Am. Chem. Soc. 1996, 118, 1577. (g) Liang, B.; Novak, T.; Tan, Z.; Negishi, E. J. Am. Chem. Soc. 2006, 128, 2770. (h) Pérez, M.; Fañanás-Mastral, M.; Hornillos, V.; Rudolph, A.; Bos, P. H.; Harutyunyan, S.

(19)

18

R.; Feringa, B. L. Chem. Eur. J. 2012, 18, 11880. (i) van Zijl, A. W.; López, F.; Minnaard, A. J.; Feringa, B. L. J. Org. Chem. 2007, 72, 2558. (j) Guduguntla, S.; Fañanás-Mastral M.; Feringa, B. L. J. Org. Chem. 2013, 78, 8274.

60) Selected examples: a) Dübner, F.; Knockel, P. Angew. Chem. Int. Ed. 1999, 38, 379. b) Malda, H.; van Zijl, A. W.; Arnold, L. A.; Feringa, B, L. Org. Lett. 2001, 3, 1169. c) Shi, W. J.; Wang, L. X.; Fu, Y.; Zhu, S. F.; Zhou, Q. L. Tetrahedron: Asymmetry 2003, 14, 3867. d) Van Zijl, A. W.; Arnold, L. A.; Minnaard, A. J.; Feringa, B. L. Adv. Synth. Catal. 2004, 346, 413. e) Luchaco-Cullis, C. A.; Mizutani, H.; Murphy, K. E.; Hoveyda, A. H. Angew. Chem. Int. Ed. 2001, 40, 1456. f) Kacprzynski, M. A.; Hoveyda, A. H. J. Am. Chem. Soc. 2004, 126, 10676. g) Ongeri, S.; Piarulli, U.; Roux, M.; Monti, C.; Gennari, C. Helv. Chim. Acta.

2002, 85, 3388. h) Yoshikai, N.; Miura, K.; Nakamura, E. Adv. Synth. Catal. 2009,

351, 1014. i) Van Veldhuizen, J. J.; Campbell, J. E.; Giudici, R. E.; Hoveyda, A. H. J. Am. Chem. Soc. 2005, 127, 6877. j) Jennequin, T.; Wencel-Delord, J.; Rix, D.; Daubignard, J.; Crévisy, C.; Mauduit, M. Synlett 2010, 1661. k) Kacprzynski, M. A.; May, T. L.; Kazane, S. A.; Hoveyda, A. H. Angew. Chem. Int. Ed. 2007, 46, 4554. l) Piarulli, U.; Daubos, P.; Claverie, C.; Roux, M.; Gennari, C. Angew. Chem. Int. Ed. 2003, 42, 234. m) Piarulli, U.; Claverie, C.; Daubos, P.; Gennari, C.; Minnaard, A. J.; Feringa, B. L. Org. Lett. 2003, 5, 4493. n) Bertozzi, F.; Crotti, P.; Macchia, F.; Pineschi, M.; Feringa, B. L. Angew. Chem. Int. Ed. 2001, 40, 930. o) Pineschi, M.; Del Moro, F.; Crotti, P.; Di Bussolo, V.; Macchia, F. Synthesis 2005, 334.

61) Selected examples: a) Karlstrom, A. S. E.; Huerta, F. F.; Meuzelaar, G. J.; Bäckvall, J. E. Synlett 2001, 923. b) Cotton, H. K.; Norinder, J.; Bäckvall, J. E.

Tetrahedron 2006, 62, 5632. c) Lölsberg, W.; Ye, S.; Schmalz, H. G. Adv. Synth. Catal. 2010, 352, 2023. d) Fang, F.; Zhang, H.; Xie, F.; Yang, G.; Zhang, W. Tetrahedron 2010, 66, 3593. e) Magre, M.; Mazuela, J.; Diéguez, M.; Pàmies, O.; Alexakis, A. Tetrahedron: Asymmetry 2012, 23, 67. f) Magrez, M.; Le Guen, Y.; Baslé, O.; Crévisy, C.; Mauduit, M. Chem. Eur. J. 2013, 19, 1199. g) Hornillos, V.; Pérez, M.; Fañanás-Mastral, M.; Feringa, B. L. J. Am. Chem. Soc. 2013, 135, 2140. h) Selim, K. B.; Matsumoto, Y.; Yamada, K.; Tomioka, K. Angew. Chem. Int. Ed.

2009, 48, 8733. i) Li, H.; Alexakis, A. Angew. Chem. Int. Ed. 2012, 51, 1055. j)

Giannerini, M.; Fañanás-Mastral, M.; Feringa, B. L. J. Am. Chem. Soc. 2012, 134, 4108. k) Geurts, K.; Fletcher, S. P.; Feringa, B. L. J. Am. Chem. Soc. 2006, 128, 15572.

62) Selected examples: a) Dabrowski, J. A.; Gao, F.; Hoveyda, A. H. J. Am. Chem. Soc.

2011, 133, 4778. b) Gao, F.; McGrath, K. P.; Lee, Y.; Hoveyda, A. H. J. Am.

Chem. Soc. 2010, 132, 14315. c) Palais, L.; Bournaud, C.; Micouin, L.; Alexakis, A.; Chem. Eur. J. 2010, 16, 2567. d) Akiyama, K.; Gao, F.; Hoveyda, A. H. Angew.

(20)

19

Chem. Int. Ed.2010, 49, 419. e) Gao, F.; Lee, Y.; Mandai, K.; Hoveyda, A. H. Angew. Chem. Int. Ed. 2010, 49, 8370. f) Dabrowski, J. A.; Haeffner, F.; Hoveyda, A. H. Angew. Chem. Int. Ed. 2013, 52, 7694.

63) a) You, H.; Rideau, E.; Sidera, M.; Fletcher, S. P. Nature 2015, 517, 351. b) Sidera, M.; Fletcher, S. P. Chem. Commun. 2015, 51, 5044. c) Rideau, E.; Fletcher, S. P. Beilstein J. Org. Chem. 2015, 11, 2435.

64) a) Pérez, M.; Fañanás-Mastral, M.; Bos, P. H.; Rudolph, A.; Harutyunyan, S. R.; Feringa, B. L. Nat. Chem. 2011, 3, 377. b) Kiyotsuka, Y.; Kobayashi, Y. Tetrahedron 2010, 66, 676. c) Fañanás-Mastral, M.; Pérez, M.; Bos, P. H.; Rudolph, A.; Harutyunyan, S. R.; Feringa, B. L. Angew. Chem. Int. Ed. 2012, 51, 1922. d) Guduguntla, S.; Fañanás-Mastral, M.; Feringa, B. L. J. Org. Chem.

2013, 78, 8274. e) Fañanás-Mastral, M.; Vitale, R.; Pérez, M.; Feringa, B. L. Chem.

Eur. J. 2015, 21, 4209. f) Pérez, M.; Fañanás-Mastral, M.; Hornillos, V.; Rudolph, A.; Bos, P. H.; Harutyunyan, S. R.; Feringa, B. L. Chem. Eur. J. 2012, 18, 11880. g) Bos, P. H.; Rudolph, A.; Pérez, M.; Fañanás-Mastral, M.; Harutyunyan, S. R.; Feringa, B. L. Chem. Commun. 2012, 48, h) Guduguntla, ornillos, essier, R Fa an s-Mastral, M.; Feringa, B. L. Org. Lett. 2016, 18, 252. i) Guduguntla, S.; Gualtierotti, J. -B.; Goh, S. S.; Feringa, B. L. ACS Catal. 2016, 6, 6591.

65) Selected examples using boronic esters: a) Shintani, R.; Takatsu, K.; Takeda, M.; Hayashi, T. Angew. Chem. Int. Ed. 2011, 50, 8656. b) Takeda, M.; Takatsu, K.; Shintani, R.; Hayashi, T. J. Org. Chem. 2014, 79, 2354.

66) Selected example using ketene silyl acetals: Li, D.; Ohmiya, H.; Sawamura, M. J. Am. Chem. Soc. 2011, 133, 5672.

67) Other examples: a) Ohmiya, H.; Zhang, H.; Shibata, S.; Harada, A.; Sawamura, M.

Angew. Chem. Int. Ed. 2016, 55, 4777. b) Harada, A.; Makida, Y.; Sato, T.;

Ohmiya, H.; Sawamura, M. J. Am. Chem. Soc. 2014, 136, 13932. a) Zhu, S.; Niljianskul, N.; Buchwald, S. L. Nat. Chem. 2016, 8, 144. b) Wang, Y. -M.; Buchwald, S. L. J. Am. Chem. Soc. 2016, 138, 5024. c) Nguyen, T. N. T; Thiel, N. O.; Pape, F.; Teichert, J.F. Org. Lett. 2016, 18, 2455.

68) a) Trost, B. M.; Crawey, M. L. Chem. Rev. 2003, 103, 2921. b) Trost, B. M. J. Org. Chem. 2004, 69, 5813. c) Lu, Z.; Ma, S. M. Angew. Chem. Int. Ed. 2008, 47, 258. d) Milhau, L.; Guiry, P. J.; Top. Organomet. Chem. 2012, 38, 95.

(21)

Referenties

GERELATEERDE DOCUMENTEN

An interesting addition would be the synthesis of more steroid derivatives; either by changing the Grignard reagent in the conjugate addition step, or by having a

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

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

Herein we present a highly enantioselective one pot synthesis of β-alkyl- substituted alcohols through Cu-catalyzed AAA of allyl bromides with various organolithium reagents

organic layer was dried over MgSO 4 , filtered, and concentrated under.. The residue was purified by column chromatography on silica gel using a mixtures of

Here, we report the first regio- and enantioselective Cu(I)-catalyzed asymmetric allylic arylation (Cu-AAAr) of trisubstituted allyl bromides using aryl

The combined organic layers were dried over anhydrous MgSO 4 , filtered and the volatiles were evaporated under vacuum to afford tetrabutylammonium phosphite salt which

The main objective of the research discussed in this chapter was to develop a new asymmetric autocatalytic reaction based on a metal-ligand approach. We choose two