• No results found

University of Groningen Copper-catalyzed enantioselective conjugate addition of Grignard reagent to non-activated acceptors Yan, Xingchen

N/A
N/A
Protected

Academic year: 2021

Share "University of Groningen Copper-catalyzed enantioselective conjugate addition of Grignard reagent to non-activated acceptors Yan, Xingchen"

Copied!
47
0
0

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

Hele tekst

(1)

University of Groningen

Copper-catalyzed enantioselective conjugate addition of Grignard reagent to non-activated acceptors

Yan, Xingchen

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: 2019

Link to publication in University of Groningen/UMCG research database

Citation for published version (APA):

Yan, X. (2019). Copper-catalyzed enantioselective conjugate addition of Grignard reagent to non-activated acceptors. 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)

103

Chapter 4: Copper-catalyzed enantioselective direct

conjugate addition of Grignard reagents to α,β-unsaturated

carboxylic acids enabled by Lewis acid

This chapter describes a strategy that allows highly efficient catalytic asymmetric additions of organomagnesium reagents to unprotected α,β-unsaturated carboxylic acids. The strategy behind this success is based on preventing the formation of unreactive carboxylate salts by means of formation of a reactive intermediate, allowing modifications of the carbon chain to proceed unhindered, while the stereochemistry is controlled with a chiral copper catalyst. A wide variety of β-chiral carboxylic acids, obtained with excellent enantioselectivities and yields, can be further transformed into highly valuable molecules through for instance catalytic decarboxylative cross-coupling reactions.

(3)

104

4.1 Introduction

Unprotected carboxylic acids are important constituents of biologically active compounds and essential precursors in the synthesis of numerous useful derivatives. As such, they are often produced industrially and used in the production of polymers, pharmaceuticals, solvents, and food additives.1-3 Furthermore, carboxylic acids possess advantageous characteristics in terms of easy handling, storage, stability and toxicity and they are the main precursors of the majority of carbonyl based Michael acceptors (Figure 1a) for which conjugate additions have been developed.4 More importantly, carboxylic acids (mainly achiral) also serve as immediate starting materials in various decarboxylative coupling reactions, a class of reactions that has witnessed enormous progress in the last 10 years and allows straightforward access to many structural motives.5-8 Some noteworthy examples are decarboxylative alkylation, alkenylation, azidation, borylation, amination, halogenation and alkynylation (Figure 1a).9-16 Thus, a methodology for direct synthesis of

β-substituted carboxylic acids also implies efficient and straightforward access to the chiral

analogues of those structural motives.

Figure 1: importance and fundamental problem of carboxylic acids. a, Overview of transformations of the

carboxylic acid functional group leading to its derivatives and new structural motives: carboxylic acids can undergo straightforward functionalization and decarboxylative cross-coupling reactions. Carrying out these transformations with acids bearing a chiral carbon chain (R) will lead to formation of the chiral analogues of the presented products. b, Fundamental problem that prevents development of conjugate additions to unprotected unsaturated carboxylic acids: mixing of organometallics with carboxylic acids leads to an acid-base reaction resulting in a carboxylate salt A nearly unreactive towards further reactions.

So far, chiral β-substituted carboxylic acids are mainly obtained by kinetic resolution or asymmetric hydrogenation reactions.17-20 However these methods are often limited to aryl substituents in the β-position of the substrate, require precious transition metal catalysts and make use of molecular frameworks in which all the carbons have already been preinstalled. Thus, all C−C bonds must be formed in preceding reactions. Another common, indirect way is through asymmetric additions to premade α,β-unsaturated esters,

(4)

105

thioesters or amides using chiral catalysts or chiral auxiliaries, followed by a hydrolysis step.21-25

One of the simplest ways to generate target carboxylic acids would exist of taking simple, readily available variants and modifying the carbon chain by for example introducing functional groups, forming additional carbon-carbon bonds and introducing chirality. This implies (asymmetric) conjugate addition of organometallics, a highly important and widely used method that allows the introduction of carbon-carbon (C–C) bonds and chirality in a single step.26-30 However, applying this method to unprotected α,β-unsaturated carboxylic acids is inhibited by a fundamental problem, namely that upon mixing with any organometallics the reactivity of the acid invariably leads to deprotonation and the formation of carboxylate salts (metal carboxylates), as the organometallic functions primarily as a base instead of as a nucleophile (Figure 1b). Once the salt is formed, further organic reactions become unfeasible because of the inherent low reactivity and insolubility of salts in organic solvents. This issue can be partially overcome for racemic reactions.31-36 However, the necessity to use high reaction temperatures or an excess of highly flammable organolithium reagents, leading to complex reaction mixtures, low yields and a limited scope of products, prevented further developments in this direction. Consequently, even though the first effort to do this, in a non-enantioselective manner, dates back to 195331 and despite many further attempts,32-36 to our knowledge no examples of direct applications of unsaturated carboxylic acids in either catalytic or stoichiometric enantioselective reactions with hard organometallics, nor with organoboron or organosilicon reagents, are known. Solving this last piece of the puzzle would allow direct access to β-chiral carboxylic acids from simple substrates and without any derivatizations or protecting and deprotecting steps.

Thus, our goal was to develop a general platform for direct catalytic synthesis of enantiopure β-substituted carboxylic acids from simple carboxylic acid building blocks via C–C bond forming reactions. Making this possible via additions of organometallics directly to unprotected α,β-unsaturated carboxylic acids would present a unique and important step forward in organic synthesis, but requires circumventing the fundamental issue of the acid-base reactions hindering the desired carbon-carbon bond forming process.

4.2 Results and discussions

Racemic reaction development. We realized that the fundamental issue of the acid-base reactions leading to the formation of unreactive carboxylate salts, could be circumvented by in situ formation of a reactive intermediate B from or instead of the carboxylate salt A (Figure 2a). This would require an intermediate that is reactive towards enantioselective conjugate additions, is easily formed under organometallic addition reaction conditions and would lead to the final carboxylic acid product without additional chemical reactions, simply upon reaction quenching or product isolation. We set out to identify a compound that fulfils all of these stringent requirements, drawing on our past experience with combining Lewis acids and highly reactive organometallics in conjugate additions.25,37,38 We speculated that using common Lewis acids such as TMSOTf or BF3·Et2O as electrophiles

(5)

106

might lead to the formation of soluble and unstable silyl- or boron- intermediates, with anticipated high reactivity towards conjugate addition of Grignard reagents and can lead directly to the final unprotected carboxylic acid product. This particular choice of Lewis acids was based on our previous studies, where these Lewis acids were found to be compatible with Grignard reagents and successfully used to enhance reaction rates with various electrophiles.25,37,38

Figure 2: reaction development. a, Our approach, based on the use of Lewis acid to promote in-situ formation

of a reactive intermediate B that can undergo conjugate addition of organometallic reagent. b, Racemic conjugate addition of EtMgBr to substrate 1a with varying conditions. c, Rationalization of the experimental data in entries 1-7 obtained for conjugate addition of EtMgBr in various conditions.

To address the feasibility of this strategy we first investigated the racemic addition of EtMgBr to trans-2-hexenoic acid 1a (Figure 2b). At 0 °C a complex mixture of products (including 35% of 2a) was observed in the absence of any catalyst or additives, indicating that the system is dominated by various side reactions at such a relatively high temperature (see Supplementary Table 1). As expected, no conversion of the substrate 1a was seen at −55 °C (Figure 2b, entry 1), because only the Mg carboxylate A-Mg was formed and precipitated out of the reaction mixture (and hydrolyzed back to the substrate during quenching of the reaction). We proceeded by adding the reactive Lewis acids TMSOTf or BF3·Et2O to the mixture, hoping they would react with the Mg-carboxylate A-Mg to form a more reactive boron or silyl intermediate (B-Si or B-B depending on the Lewis acid used),

(6)

107

but unfortunately this only had a minor effect (Figure 2b and c, entries 2 and 3). Anticipating poor reactivity of the Mg-carboxylate A-Mg towards the silyl and boron electrophiles we decided to add the latter first and EtMgBr second, and to our delight this yielded the addition product 2a with 10% conversion with BF3·Et2O and 52 % conversion with TMSOTf (Figure 2b, entries 4 and 5 respectively). The much higher conversion towards the addition product 2a obtained with TMSOTf can be rationalized by the higher electrophilicity of the latter in comparison with BF3·Et2O.

Figure 3: 1H NMR studies (at −55 °C in CD2Cl2) to detect the formation of silyl intermediate carried out with

TBSOTf. a, Free carboxylic acid 1b. b, Combination of free carboxylic acid 1b with 2.2 equiv. of TBSOTf. c, Premade and isolated TBS-ester for the reference. d, Combination of free carboxylic acid 1b with 2.2 equiv. of TBSOTf and 1.0 equiv. of MeMgBr.

We believe that the sequence of addition of the reagents is important, because the nucleophilicity of the Mg carboxylate A-Mg is insufficient to react with the boron or silyl electrophile (presumably due to aggregates formation), and thus nearly no reactive intermediates are generated (Figure 2c, entries 2 and 3). Conversely, when either Lewis acid is added first, it combines with the carboxylic acid 1a to form an initial complex, after which addition of EtMgBr leads to abstraction of the proton and concomitant formation of intermediates B-B or B-Si (depending on the Lewis acid used), that are reactive towards conjugate addition (Figures 2b and 2c, entries 4 and 5 respectively). Hence, the formation of the Mg-carboxylate A-Mg is avoided and conjugate addition to form the product 2a can proceed. This rationalization was confirmed by 1H NMR spectroscopy for the silyl intermediate (Figure 3). Initial 1H NMR experiments to detect the TMS-ester formation failed due to its instability. To circumvent this problem, we decided to carry out these

(7)

108

studies using TBSOTf that can form a more stable Si-ester. For this purpose we decided to use substrate 1b instead of 1a in order to obtain simpler for interpretation spectra. We first recorded the reference spectra of substrate 1b and the isolated pure TBS-ester of 1b (Figures 3a and 3c respectively). Next we recorded the spectrum of the mixture of 1b with TBSOTf and found that a new compound was formed that corresponds to the complex between these two species (Figures 3b). Finally, addition of MeMgBr to the mixture of 1b and TBSOTf resulted in the formation of the TBS-ester of 1b, as can be seen from the comparison of the 1H NMR signals observed in the spectra of Figures 3c and 3d. As expected, changing the order of addition to first adding MeMgBr to 1b followed by the addition of TBSOTf did not afford the Si-ester and only peaks corresponding to Mg-carboxylate salt of 1b were observed.

To investigate whether carboxylate salts are in general unreactive towards the formation of silyl intermediate or whether this depends on the nature of the metal ion, Li and Na carboxylates A-Li and A-Na were prepared from n-BuLi and NaH, respectively, and subjected to the reaction with TMSOTf and EtMgBr (Figures 2b and 2c, entries 6 and 7). Now the silyl intermediate B-Si was formed in both cases, and addition of 1 equivalent of EtMgBr resulted in the formation of product 2a with 74 and 66% conversions respectively (Figures 2b and 2c, entries 6 and 7). These results indicate that once the intermediate B-Si is formed, conjugate addition works well, but that when the metal carboxylate A is formed first, the success of the reaction depends on the counter ions (Figure 2c). Since metal carboxylates A-Mg, A-Li, and A-Na are all poorly soluble and precipitate in the reaction solvent, the differing reactivities could also originate from solubility differences. However, we found that the solubility of Li-carboxylate A-Li and Na-carboxylate A-Na is lower than that of Mg-carboxylate B (see Supplementary Figures 1−3), and thus the difference in reactivity can only be attributed to a higher nucleophilicity of the Li- and Na-carboxylate. Development of catalytic asymmetric reaction. Having established that the boron and silyl intermediates B-B and especially B-Si are indeed formed under the right conditions and lead to the racemic product 2a, we shifted our attention to the question whether this reaction system would be susceptible to asymmetric catalysis in order to both accelerate the conjugate addition towards higher yields and achieve enantiocontrol. As copper is known to be an efficient catalyst for asymmetric conjugate addition reactions,26-30 we started our investigation by selecting various chiral ligands that can bind to Cu(I). As expected initial experiments in DCM showed no addition of the highly reactive EtMgBr to 1a when performing the reaction in the presence of 5 mol% of L1/Cu(I)-catalyst at −78 °C. Raising the temperature to 0 °C resulted in 64% conversion with only 34% towards racemic addition product 2a and many byproducts (Table 1, entries 1 and 2). At this point we decided to investigate catalytic reactions in the presence of TMSOTf (via the formation of the most reactive intermediate B-Si) and copper complexes with various chiral diphosphine ligands in DCM at −78 °C. Several chiral catalytic systems result in both acceleration of the conjugate addition and in significant enantiodiscrimination (Table 1, entries 3−7). The superior yield and enantioselectivity obtained with diphosphine Tol-BINAP ligand (R)-L4 (entry 6) prompted us to select it as the optimal ligand for this reaction. Subsequently, a thorough optimization process was executed using the catalytic

(8)

109

system derived from L4/CuBr·SMe2 involving the evaluation of various parameters and reaction conditions (for complete set of data see Supplementary Tables 1−5). In particular the effect of different solvents was studied. With the exception of THF, all solvents tested were effectively tolerated, providing 2a with good yields and ee (entries 8−11). We were especially pleased to find exceptionally high yield (95%) and enantioselectivity (92%) in MTBE (entry 11).

Table 1: development of catalytic system for direct asymmetric conjugate addition of EtMgBr to carboxylic acid 1aa

Entry L/Cu(I) LA Solvent T [°C ] Conv. [%]b ee [%]c

1 L1/Cu(I) – DCM –78 0 – 2 L1/Cu(I) – DCM 0 79d Rac 3 L1/Cu(I) TMSOTf DCM –78 74 47 4 L2/Cu(I) TMSOTf DCM –78 70 9 5 L3/Cu(I) TMSOTf DCM –78 72 47 6 L4/Cu(I) TMSOTf DCM –78 87 56 7 L5/Cu(I) TMSOTf DCM –78 75 47 8 L4/Cu(I) TMSOTf THF –78 100 Rac 9 L4/Cu(I) TMSOTf Toluene –78 62 80 10 L4/Cu(I) TMSOTf Et2O –78 91 88

11 L4/Cu(I) TMSOTf MTBE –78 95 92 12e L4/Cu(I) TBSOTf MTBE –78 95 95

13 L4/Cu(I) BF3·Et2O MTBE –78 19 92

14f L4/Cu(I) BF3·Et2O MTBE –78 77 97

15f L4/Cu(I) TMSOTf MTBE –78 99 97

16g L4/Cu(I) TMSOTf MTBE –78 100 95

17 L4/Cu(I) TMSOTf MTBE 0 95 88

18 L4/Cu(I) TMSOTf MTBE –20 97 97

aReaction conditions: 0.1 M of 1a, 5 mol% of CuBr·SMe2, 6 mol% of L and 2−3 equiv. of LA followed by the

addition of 2−3 equiv. of EtMgBr. bConversion was determined by NMR of reaction crude. cEnantiomeric

excess was determined by chiral HPLC after transforming 2a to the corresponding N,N-dimethyl amide derivative. dLess than 20 % of 2a formed with many other byroducts. eThe product was obtained as a mixture

of silyl ester and free carboxylic acid in the ratio of 62:38 respectively. fThe reaction was performed by first

forming Li-carboxylate with n-BuLi followed by addition of corresponding LA and EtMgBr. gThe reaction was

performed by first forming Na-carboxylate with NaH followed by the additions of TMSOTf and EtMgBr.

In order to connect catalytic asymmetric addition with the results of our 1H NMR spectroscopic studies using TBSOTf, we tested the addition of EtMgBr to substrate 1a using this Lewis acid. As expected, we found that the reaction proceeds with excellent conversion

(9)

110

towards the addition product, providing the final product as a mixture of TBS-ester and free carboxylic acid 2a in a ratio of 62:38 respectively and high enantioselectivity (95% ee, entry 12). This composition of the product mixture is not surprising as under these reaction conditions TBS-ester is expected to be relatively stable.

The catalytic reaction with L4/CuBr·SMe2 in MTBE was also investigated using BF3·Et2O as a Lewis acid. Although high levels of enantioselectivity (92%) could be obtained, a lower reactivity towards conjugate addition of Grignard reagent (only 19% of conversion to 2a) was found in this case, similar to that observed earlier for the racemic reaction (entry 13). In contrast, high yields and enantioselectivities can be obtained when using BF3·Et2O in combination with Li-carboxylate (formed by n-BuLi in-situ) followed by addition of EtMgBr (Table 1, entries 14). Similar excellent results were found when using TMSOTf in combination with Li- and Na-carboxylates (formed by n-BuLi and NaH respectively) followed by addition of EtMgBr (Table 1, entries 15 and 16). However, TMSOTf was selected as the Lewis acid of choice for further studies because of the convenience of a procedure using only one organometallic reagent and because it gives the highest conversion (entry 11). Since a temperature of −78 °C is not practical, particularly for large scale synthesis, we evaluated the temperature as well (entries 17 and 18), finding that −20 °C is the optimal temperature for the reaction with Grignard reagent, the chiral L4/Cu(I) catalyst system and TMSOTf. Under these optimized conditions the reaction is completed in 2h providing the final product 2a with 97% of conversion and an enantioselectivity of 97% (entry 18).

Scope of the reaction. With the optimized conditions in hand, initial efforts to explore the scope of this transformation focused on investigating the effect of varying the carboxylic acid substitution at the β-position (Figure 4). A wide variety of substrates allows efficient transformation to the corresponding chiral β-substituted carboxylic acids. The substrates with linear and branched aliphatic chains (including cyclohexyl and cyclopropyl) gave the corresponding addition products 2a−2d with high yields and excellent enantioselectivities. However, when we applied this condition to the aromatic substrate 1e, only 57% ee was obtained for product 2e. Further optimization (see Supplementary Table 5) revealed a different catalytic system to be optimal for aromatic substrates, based on diphosphine ligand (R,R)-L5 in combination with copper salt and lower temperatures. Using 10 mol% of L5/Cu(I) as the catalyst at −40 °C provided the product 2e with 91% enantiopurity and 74% isolated yield (Figure 4).

An aromatic ring with an electron-donating (methoxy, 2f) or electron-withdrawing group (Br, 2g), as well as a heteroaromatic ring (2h, 2i), are well tolerated, but including a m-Br-substituent in the aromatic ring led to the addition product 2g with lower yield (54%) and

ee (86%). When the aromatic ring is at the γ-position, the substrates behave as aliphatic

substrates, and the highest levels of ee and conversion are obtained with catalyst L4/Cu(I) (products 2j and 2k). Finally, our catalytic system tolerates the presence of functional groups in the substrate, providing the corresponding products (2l, 2m) with high yields and ees above 96%. Next we examined the nucleophile scope, starting our investigation with the smallest and least reactive of all Grignard reagents, MeMgBr (Figure 4).

(10)

111

Methylations are highly relevant for building chiral polymethylated arrays commonly found in natural products, but they pose difficulties because of the low reactivity of methylating reagents in general.23-25,28 Currently, asymmetric addition of MeMgBr to α,β-unsaturated esters is only successful with aliphatic substrates. 23-25

Figure 4: scope of the substrate and Grignard reagent. aFor details see Experimental section. Isolated yields

for all the products are shown. The absolute configuration of the products obtained with (R)-L4 or (R,R)-L5 as the ligands are opposite. bReaction conditions: 0.1 M of substrate in MTBE with 5 mol% (R)-L4/CuBr·SMe2 or

in MTBE /Toluene = 1/1 with 10 mol% (R,R)-L5/CuBr·SMe2, 2−3 equiv. of TMSOTf and RMgBr. cUsing 5 mol%

(R,R)-L5/CuBr·SMe2 as a catalyst in the same condition led to 1a with 93% ee. dUsing 5 mol%

(R)-L4/CuBr·SMe2 as a catalyst led to 2e with 57% ee. eReaction performed using 10 mol% of (R)-L4/CuBr·SMe2.

fReaction performed at −40 °C.

To our delight, our catalytic system solves this longstanding problem as it works with similar efficiency for both aliphatic and aromatic substrates, leading to final methylated carboxylic acids with excellent yields and ees (3a, 3b and 3c). All alkyl Grignard reagents afforded addition products with excellent results, independent of the chain length and

(11)

112

branching. The sterically demanding α-, β-, and γ-branched Grignard reagents are tolerated, providing products 3d−3f with high yields and enantioselectivities exceeding 95%. Grignard reagents bearing olefinic substituents also function well, affording the corresponding product 3g with excellent ee and yield. Products 3h and 3i, derived from additions of linear Grignard reagent (n-HexMgBr) to 1a and crotonic acid, respectively, were obtained with high enantiopurities and yields as well.

A few important practical aspects of this chemistry deserve to be highlighted (Table 2). First is the possibility to recycle the catalyst, which can be recovered from the reaction mixture with 83% isolated yield in the form of chiral Cu-complex and reused in the next reaction with similar performance. Furthermore, the catalyst loading can be decreased from 5 to 1 mol% for addition reactions to aliphatic substrates, and the reaction can be carried out with similar outcome in 1 g scale of the substrate. Finally, an additional benefit of this catalytic system is that when substrate conversion exceeds 97% most products can be obtained by simple acid-base extraction rather than time consuming column chromatography, which is important for large-scale industrial application.28

Table 2: practical aspects of the Cu-catalyzed asymmetric conjugate addition of EtMgBr to 1aa

Entry 1a [mmol] MTBE [mL] L4/Cu(I) [mol%] Yield [%]b ee [%]c

1 0.2 2 1 84 94

2 10 50 5d 83 97

3 0.2 2 5e 86 95

aReaction conditions: −20 °C, TMSOTf (2.2 equiv.), EtMgBr (2.5 equiv.) for 2 h. bIsolated yields for 2a are

shown. Work-up was performed by acid base extraction. cEnantiomeric excess were determined by HPLC on a

chiral stationary phase after transforming to the corresponding N,N-dimethyl amide. d83% catalyst can be

recovered. eThe reaction was performed with recovered catalyst.

Application of the catalytic methodology. To showcase the potential of our catalytic protocol for future applications we demonstrate that β-chiral substituted carboxylic acids can easily be transformed into a variety of valuable molecules (Figure 5). The first applications that illustrate utility pertain to the use of our chiral products in stereoselective decarboxylative cross-coupling reactions. Decarboxylative cross-coupling reactions are developing very rapidly and various catalytic systems utilizing aliphatic carboxylic acids (mainly achiral) leading to diverse structural motives have been established over the past decade.6-8 Here we demonstrate how useful chiral analogues of those structural motives can be obtained by combining our methodology and decarboxylative couplings (Figures 5a−5c). Nickel-catalyzed decarboxylative alkylation39 and borylation12 of product 2k afforded chiral alkane 4a and chiral β-substituted boronate ester 4b, maintaining the original enantiopurity of the starting material through the process (96% ee, Figure 5a). Silver-catalyzed decarboxylative bromination of carboxylic acid 3b led to the β-chiral alkyl bromide 4c with an ee of 99% (Figure 5b),40 while Ag-catalyzed decarboxylative azidation

(12)

113

of carboxylic acid 3h provided chiral β-substituted azide 4d, which was further transformed into chiral triazole 4e via click reaction, once again without any racemization (98% ee) (Figure 5c).16 Although some similar molecules can be obtained via other catalytic asymmetric methodologies, these are often limited to specific structures and feature varying levels of enantioselectivities.41,42 For example, chiral β-substituted boronate esters can also be obtained through transition metal catalyzed hydroboration of 1,1-disubstituted alkenes. However, these methods are only effective when an aryl group or directing carbonyl groups are present in the substrate. Applying β-substituted carboxylic acids obtained by our methodology in decarboxylative borylation offers an attractive alternative for accessing a wide range of chiral aliphatic β-substituted boronate esters.

Figure 5: synthetic utility of the process. a, Ni-catalyzed decarboxylative alkylation and borylation of chiral

acid 2k. b, Ag-catalyzed decarboxylative bromination of product 3b. c, Ag-catalyzed decarboxylative azidation of chiral acid 3h followed by click reaction. d, Late stage functionalization of a RXR antagonist 5a. e, Effect of the different procedures on the structure of the final asymmetric conjugate addition of EtMgBr to carboxylic acid substrate 1l. f, Reported synthetic route to a potent 15-lipoxygenase-1-inhibitor 7a. g, Synthesis of the derivative 7b in 2 steps using current methodology. h, Synthesis of chiral acid 3c which is a key intermediate of several natural products. Reaction conditions: (i) HATU, NEt3, NiCl2·glyme,

4,4'-di-t-butyl-2,2'-dipyridyl, ZnEt2, in DMF at RT; (ii) N-hydroxyphthalimide, DCC, in DCM at RT, 2h, then MgBr2·OEt2,

NiCl2·6H2O, 4,4'-dimethoxy-2,2'-bipyridyl, [B2pin2Me]Li, in DMF, THF, at 0 °C 1 h − RT 1 h; (iii) Ag(Phen)2OTf,

dibromoisocyanuric acid, in 1,2-dichloroethane at 60 °C; (iv) AgF, K2S2O8, MesSO2N3 in CH3CN, H2O, at 55 °C;

(13)

114

1:1, at −20 °C; (vii) CuBr·SMe2, (R)-L1, TMSOTf, EtMgBr, in MTBE at −20 °C; (viii) CuBr·SMe2, (R)-L1, n-BuLi,

TMSOTf, EtMgBr, in MTBE at −78 °C 2 h − RT 16 h; (ix) CuBr·SMe2, (R)-L1, TMSOTf, n-HexMgBr, in MTBE at

−78 °C; (x) SOCl2, DMF (1 drop), in DCM at RT, 1 h, then ethyl 6-chloro-1H-indole-2-carboxylate, SnCl4, in

DCM, reflux; (xi) CuBr·SMe2, (S,S)-L2, TMSOTf, MeMgBr, in MTBE:Toluene = 1:1 at −20 °C.

Our methodology is also sufficiently mild and robust to be applied in more complex molecules. For example, UVI3003 5a, a full antagonist of RXR (one of the retinoid receptors involved in the control of various physiological and pathological processes, including cancer and metabolic diseases) that demonstrates potent, nanomolar binding affinity,43 can be functionalized successfully with our strategy without prior protection of the hydroxyl group and afford the product 5b with 76% yield and 99% ee (Figure 5d). Another synthetically useful transformation available with our methodology is the trapping of enolate intermediates formed upon conjugate addition (Figure 5e). For instance, while conjugate addition product 2l can be obtained with 88% yield and 96% ee, modifying the original procedure by using Li-carboxylate for asymmetric conjugate addition at −78 °C, followed by warming up to room temperature and stirring overnight, leads to intramolecular enolate trapping, affording cyclic product 6 with contiguous stereocenters as a single diastereoisomer (70% yield, 91% ee). Recently, chiral indole derivative 7a, synthesized in four steps from the commercially available (S)-citronellal, was reported (Figure 5f) to exhibit in vitro and ex vivo anti-inflammatory properties as a potent 15-lipoxygenase-1 inhibitor.44 However, as such chiral aldehydes or carboxylic acids are rarely commercially available, synthesis of similar chiral compounds with variations of the alkyl chain is difficult, thus limiting the number of molecules available for bioactivity screening. With our methodology, a library of this type of compounds with different substituents at the β-position of the acyl group can be straightforwardly accessed in just two steps, as exemplified by the synthesis of 7b (Figure 5g). Finally, our methodology allows us to effortlessly obtain the aromatic chiral β-substituted carboxylic acid 3c (Figure 5h), which is a key intermediate for the synthesis of several natural products like (S)-(+)-ar-tumerone, (+)-bisacumol and (S)-ar-himachalene.45

4.3 Conclusions

We have shown that a wide range of β-chiral carboxylic acids are now synthetically accessible from their unsaturated analogues in one simple step and under mild conditions with high yields and enantioselectivities. Our strategy is based on activation of carboxylic acids via the formation of reactive and fragile silyl or boron intermediates, which is crucial for overcoming the fundamental problem of carboxylate salt formation during the conjugate addition of organometallics to unsaturated carboxylic acid. Thus, this new approach allows highly enantioselective catalytic C–C bond-forming reactions between organometallics and carboxylic acids without the use of separate protection/deprotection steps.

(14)

115

4.4 Experimental section

4.4.1 General experimental information

All reactions using oxygen- and/or moisture-sensitive materials were carried out with anhydrous solvents under a nitrogen atmosphere using standard Schlenk techniques. Flash column chromatography was performed using Merck 60 Å 230−400 mesh silica gel. Thin layer chromatography was performed using 0.25 mm E. Merck silica plates (60F-254). Unless otherwise indicated, the saturated carboxylic acids were visualized by bromocresol green staining, and other products were visualized by UV and KMnO4 staining. NMR data was collected on Varian VXR400 (1H at 400.0 MHz; 13C at 100.58 MHz) equipped with a 5 mm z-gradient broadband probe. Chemical shifts are reported in parts per million (ppm) relative to residual solvent peak (CDCl3, 1H: 7.26 ppm; 13C: 77.16 ppm; DMSO-d6, 1H: 2.50 ppm; CD2Cl2, 1H: 5.32 ppm;). Coupling constants are reported in Hertz. Multiplicity is reported with the usual abbreviations (s: singlet, d: doublet, t: triplet, q: quartet, m: multiplet). Exact mass spectra were recorded on a LTQ Orbitrap XL apparatus with ESI ionization. Enantiomeric excess (ee) were determined by chiral HPLC analysis using a Shimadzu LC-10ADVP HPLC equipped with a Shimadzu SPD-M10AVP diode array detector. 4.4.2. Chemicals

Unless otherwise indicated, reagents and substrates were purchased from commercial sources and used as received. Unsaturated carboxylic acids 1a−1i, 1n are commercially available, the rest were synthesized according to the literature procedures (see below). Solvents not required to be dry were purchased as technical grade and used as received. Dry DMF and THF used for decarboxylative cross-coupling reactions were purchased from Sigma-Aldrich, and the dry THF is inhibitor-free. Other dry solvents were freshly collected from a dry solvent purification system prior to use. Inert atmosphere experiments were performed with standard Schlenk techniques with dried (P2O5) nitrogen gas. Pent-4-en-1-ylMgBr (2.0 M in MTBE) was prepared from the corresponding alkyl bromides and Mg activated with I2 in MTBE. Organolithium reagents and Grignard reagents were purchased from Sigma-Aldrich: n-BuLi (2.5 M in hexane); EtMgBr, MeMgBr (3.0 M in Et2O); i-BuMgBr ,

i-PentMgBr, n-HexMgBr, c-PentMgBr (2.0 M in Et2O). Chiral ligands (L1−L5) were purchased from Sigma-Aldrich and Solvias. All reported compounds were characterized by 1H and 13C NMR and compared with literature data. All new compounds were fully characterized by 1H and 13C NMR and HRMS techniques.

4.4.3 Determination of absolute configuration and enantiomeric excess

The absolute configuration of 2a−2c, 2e−2i, 2l, 3a, 3d−3h was determined by comparing with reported data after transforming to the corresponding N,N-dimethyl amides.25 The absolute configurations of other compounds were assigned by analogy. The ee of all the products was determined from the corresponding N,N-dimethyl amide derivatives.

(15)

116

4.4.4 Supplementary tables: optimization of reaction conditions

Supplementary Table 1: conjugate addition of EtMgBr to unsaturated carboxylic acids in the absence of chiral catalyst and in the presence of various bases and silyl electrophiles (investigating the reactivity of different metal carboxylates towards conjugate addition of EtMgBr by changing addition

sequence of the reagents).a

Entry T [°C] 1 Addition sequence of reagents2 b 3 Conv. [%]

1 0 EtMgBr (2.5) – – 68c 2 −55 EtMgBr (2.5) – – 1 3 −55 EtMgBr (2.5) BF3·Et2O (2.2) – 1 4 −55 EtMgBr (2.5) TMSOTf (2.2) – 6 5 −55 BF3·Et2O (2.2) EtMgBr (2.5) – 10 6 −55 TMSOTf (2.2) EtMgBr (2.5) – 52

7 −55 EtMgBr (1.0) TMSOTf (2.2) EtMgBr (1.5) 14

8 −55 n-BuLi (1.0) TMSOTf (2.2) EtMgBr (1.5) 74

9 −55 NaH (1.0) TMSOTf (2.2) EtMgBr (1.5) 66

10 −78 EtMgBr (2.5) – – 1

11 −78 EtMgBr (2.5) TMSOTf (2.2) – 3

12 −78 TMSOTf (2.2) EtMgBr (2.5) – 10

13 −78 EtMgBr (1.0) TMSOTf (2.2) EtMgBr (1.5) 1

14 −78 n-BuLi (1.0) TMSOTf (2.2) EtMgBr (1.5) 41

15 −78 NaH (1.0) TMSOTf (2.2) EtMgBr (1.5) 42

aReaction conditions: 0.1 M of 1a in MTBE, 16 h. The reaction was quenched with 1.0 M HCl aqueous solution

and extracted with DCM, conversion was determined by 1H NMR of reaction crude. bValue in brackets

corresponds to the equivalents of the reagents used with respect to 1a. c1,3,5-trimethoxbenzene was used as

the internal standard. About 35% NMR yield of product 2a was obtained with the complex mixture of side products.

Supplementary Table 2: optimization data for the Cu-catalyzed asymmetric conjugate addition of EtMgBr to 1aa

Entry Cu-L LA T [°C] Solvent Conv. [%]b ee [%]c

1 – – 0 DCM 65d

2 L1 – 0 DCM 64d Rac

3 – – −78 DCM 1 –

4 L1 – −78 DCM 1 –

(16)

117 6 L1 TMSOTf −78 DCM 74 47 7 L2 TMSOTf −78 DCM 70 9 8 L3 TMSOTf −78 DCM 72 47 9 L4 TMSOTf −78 DCM 87 56 10 L5 TMSOTf −78 DCM 75 47 11 L4 TMSOTf −78 THF 100 Rac 12 L4 TMSOTf −78 Toluene 62 80 13 L4 TMSOTf −78 Et2O 91 88 14 L4 TMSOTf −78 MTBE 95 92 15 L4 TBSOTf −78 MTBE 95e 95 16 L4 TMSBr −78 MTBE 12 89 17 L4 TMSCl −78 MTBE 1 – 18 L4 BF3·Et2O −78 MTBE 19 92 19 L4 BCl3 −78 MTBE 17 20 20 L4 BBr3 −78 MTBE 16 13

aReaction conditions: 0.1 M of 1a, CuBr·SMe2 (5 mol%), L (6 mol%), LA (3.0 equiv.), EtMgBr (3.0 equiv.). bThe

reaction was quenched with 1.0 M HCl aqueous solution and extracted with DCM. Conversion was determined by 1H NMR of reaction crude. cEnantiomeric excess were determined by HPLC on a chiral stationary phase

after transforming to the corresponding N,N-dimethyl amide. d1,3,5-trimethoxbenzene was used as the

internal standard. About 34% NMR yield of product 2a was obtained with the complex mixture of side products. eTotal conversion of the product and the TBS-protected product, with the ratio of 62:38.

Supplementary Table 3: the reactivity of different metal carboxylates (formed by deptrotonation of 1a using different bases) towards various TMSX and BF3·Et2O in the presence of chiral copper catalyst

Cu-L4a

Entry Base LA LA [equiv.] EtMgBr [equiv.] T [°C] Conv. [%]b ee [%]c

1 EtMgBr TMSOTf 2.2 1.5 −78 1 – 2 n-BuLi TMSOTf 2.2 1.5 −78 99 97 3 n-BuLi TMSOTf 1.2 1.5 −78 99 98 4 NaH TMSOTf 2.2 1.5 −78 100 95 5 n-BuLi TMSBr 1.2 2.5 −78 60 98 6 n-BuLi TMSBr 2.2 2.5 −78 83 99 7 n-BuLi TMSBr 3.0 2.5 −78 90 99 8 n-BuLi BF3·Et2O 3.0 2.5 −78 77 97 9 n-BuLi TMSCl 3.0 2.5 −78 10 –

aReaction conditions: 0.1 M of 1a in MTBE, CuBr·SMe2 (5 mol%), L4 (6 mol%), base (1.0 equiv.). bThe reaction

was quenched with 1.0 M HCl aqueous solution and extracted with DCM, conversion was determined by 1H

NMR of reaction crude. cEnantiomeric excess were determined by HPLC on a chiral stationary phase after

(17)

118

Supplementary Table 4: optimization of the equivalents of TMSOTf and EtMgBr, and temperaturea

Entry TMSOTf [equiv.] EtMgBr [equiv.] T [°C] Conv. [%]b ee [%]c

1 4.0 4.0 −78 94 80 2 3.0 3.0 −78 95 92 3 3.0 2.2 −78 83 87 4 2.5 2.5 −78 92 97 5 2.2 3.0 −78 94 98 6 2.2 2.5 −78 94 98 7 2.2 2.2 −78 91 97 8 1.5 3.0 −78 88 98 9 1.5 2.2 −78 86 98 10 1.5 1.5 −78 40 – 11 2.2 2.5 0 95 88 12 2.2 2.5 −20 97 97

aReaction conditions: 0.1 M of 1a in MTBE, CuBr·SMe2 (5 mol%), L1 (6 mol%). bThe reaction was quenched

with 1.0 M HCl aqueous solution and extracted with DCM. Conversion was determined by 1H NMR of reaction

crude. cEnantiomeric excess were determined by HPLC on a chiral stationary phase after transforming to the

corresponding N,N-dimethyl amide.

Supplementary Table 5: optimization data for the Cu-catalyzed asymmetric conjugate addition of EtMgBr to 1ea

Entry L Cu-L [mol%] T [°C] Solvent [2 mL] Conv. [%]b ee [%]c

1 L4 5 −20 MTBE 94 57 2 L4 5 −78 MTBE 67 87 3 L5 5 −78 MTBE 54 89 4 L5 5 −78 DCM 41 72 5 L5 5 −78 Toluene 45 90 6 L5 5 −20 MTBE 96 64 7 L5 5 −20 MTBE 65 80d 8 L5 10 −20 MTBE 97 80 9 L5 10 −40 MTBE 95 86 10 L5 10 −40 Toluene 89 85 11 L5 10 −40 MTBE:Toluene = 1:1e 98 91

aReaction conditions: 0.1 M of 1e in the solvent. bThe reaction was quenched with 1.0 M HCl aqueous solution

and extracted with DCM, conversion was determined by 1H NMR of reaction crude. cEnantiomeric excess were

determined by HPLC on a chiral stationary phase after transforming to the corresponding N,N-dimethyl amide. dEtMgBr was diluted with MTBE to 1.0 mL and slowly added with syringe pump in 1 h. eCatalyst

(18)

119

L5/Cu(I) is not fully soluble in MTBE, but completely dissolved in toluene. Although the reaction outcome in

MTBE and toluene is similar, MTBE is a still better solvent.

4.4.5 Unsaturated silyl ester formed with TBSOTf: formations and conjugate addition 4.4.5.1 Synthesis of TBS-esters as the references

t-Butyldimethylsilyl 3-ethylhexanoate

In a flame-dried Schlenk tube equipped with septum and magnetic stirring bar, 1a (22.8 mg, 0.2 mmol, 1.0 equiv.), CuBr·SMe2 (2.06 mg, 0.01 mmol, 5 mol%) and THF (2.0 mL) were added. The mixture was cooled to −20 °C and TBSOTf (138 μL, 0.6 mmol, 3.0 equiv.) was added. After 20 min., EtMgBr (0.6 mmol, 3.0 M in Et2O, 3.0 equiv.) was added dropwise, and the reaction mixture was allowed to stir for 2 h. The reaction was quenched with saturated NaHCO3 aqueous solution (3.0 mL) and warmed to RT. The mixture was extracted with Et2O (10.0 mL × 3). The combined organic phase was dried over MgSO4, filtered and evaporated on rotary evaporator. TBS-ester of 2a was obtained as a colorless oil after column chromatography (SiO2, pentane:Et2O = 100:1) [39% yield]. 1H NMR (CDCl3, 400 MHz): δ 2.23 (d, J = 6.8 Hz, 2H, CH2CO2Si), 1.84-1.72 (m, 1H, CH), 1.43-1.18 (m, 6H, CH2), 0.93 (s, 9H, SiCCH3), 0.91-0.85 (m, 6H, CH3CH2), 0.26 (s, 6H, SiCH3). 13C NMR (CDCl3, 100 MHz): δ 174.4, 40.8, 36.6, 35.9, 26.5, 25.7, 19.9, 17.7, 14.5, 11.0, −4.7. HRMS (ESI, m/Z): calcd. for C14H31O2Si [M+H]+: 259.2088, found: 259.2094.

t-Butyldimethylsilyl hex-2-enoate

In a flame-dried Schlenk tube equipped with septum and magnetic stirring bar, 1a (22.8 mg, 0.2 mmol, 1.0 equiv.) was dissolved in MTBE (2.0 mL) and cooled down to −20 °C. TBSOTf (101 μL, 0.44 mmol, 2.2 equiv.) was added. After 20 min., EtMgBr (0.2 mmol, 3.0 M in Et2O, 1.0 equiv.) was added dropwise, and the reaction mixture was allowed to stir for 5 min under nitrogen atmosphere. The reaction was quenched with saturated NaHCO3 aqueous solution (3.0 mL) and warmed to RT. The mixture was extracted with Et2O (10.0 mL × 3). The combined organic phase was dried over MgSO4, filtered and evaporated on rotary evaporator. TBS-ester of 1a was obtained as a colorless oil after column chromatography (SiO2, pentane:Et2O = 100:1) [19% yield]. 1H NMR (CDCl3, 400 MHz): δ 6.91 (dt, J = 15.5, 6.9 Hz, 1H, CH2CH=CH), 5.78 (dt, J = 15.5, 1.6 Hz, 1H, CH2CH=CH), 2.17 (m,

J = 7.2, 1.6 Hz, 2H, CH2CH=CH), 1.49 (m, J = 7.4 Hz, 2H, CH3CH2), 0.95 (s, 9H, SiCCH3), 0.94 (t,

J = 7.3, 3H, CH3CH2), 0.29 (s, 6H, SiCH3). 13C NMR (CDCl3, 100 MHz): δ 166.8, 149.8, 123.4, 34.2, 25.8, 21.5, 17.9, 13.8, −4.6. HRMS (ESI, m/Z): calcd. for C12H25O2Si [M+H]+: 299.1618, found: 299.1620.

(19)

120

t-Butyldimethylsilyl but-2-enoate46

In a flame-dried Schlenk tube equipped with septum and magnetic stirring bar, 1b (17.2 mg, 0.2 mmol, 1.0 equiv.) was dissolved in MTBE (2.0 mL) and cooled to −20 °C. TBSOTf (101 μL, 0.44 mmol, 2.2 equiv.) was added. After 20 min., EtMgBr (0.2 mmol, 3.0 M in Et2O, 1.0 equiv.) was added dropwise, and the reaction mixture was allowed to stir for 5 min under nitrogen atmosphere. The reaction was quenched with saturated NaHCO3 aqueous solution (3.0 mL) and warmed to RT. The mixture was extracted with Et2O (10.0 mL × 3). The combined organic phase was dried over MgSO4, filtered and evaporated on rotary evaporator. TBS-ester of 1b was obtained as a colorless oil after column chromatography (SiO2, pentane:Et2O = 100:1) [13% yield]. 1H NMR (CDCl3, 400 MHz): δ 6.92 (dq, J = 15.4, 6.9 Hz, 1H, CH3CH=CH), 5.81 (dt, J = 15.4, 1.7 Hz, 1H, CH3CH=CH), 1.87 (dd, J = 6.9, 1.7 Hz, 3H, CH3CH=CH), 0.95 (s, 9H, SiCCH3), 0.28 (s, 6H, SiCH3). 13C NMR (CDCl3, 100 MHz): δ 166.6, 145.1, 124.8, 25.8, 18.0, 17.9, −4.6.

4.4.5.2 Isolation of the intermediate TBS-ester and testing of its reactivity in Cu-catalyzed asymmetric conjugate addition

In a flame-dried Schlenk tube equipped with septum and magnetic stirring bar, 1a (22.8 mg, 0.2 mmol, 1.0 equiv.) was dissolved in MTBE (2.0 mL) and cooled down to −78 °C. TBSOTf (101 μL, 0.44 mmol, 2.2 equiv.) was added. After 20 min., EtMgBr (0.2 mmol, 3.0 M in Et2O, 1.0 equiv.) was added dropwise, and the reaction mixture was allowed to stir for 5 min under nitrogen atmosphere. The reaction was quenched with HCl aqueous solution (2.0 mL, 1.0 M) and warmed to RT. The mixture was extracted with Et2O (10.0 mL × 3). The combined organic phase was dried over MgSO4, filtered and evaporated on rotary evaporator. The TBS-ester of 1a was obtained as a colorless oil after column chromatography (SiO2, pentane:Et2O = 100:1) [22% conversion, 10% yield].

In a flame-dried Schlenk tube equipped with septum and magnetic stirring bar, TBS-ester of 1a (22.8 mg, 0.1 mmol, 1.0 equiv.), CuBr·SMe2 (1.03 mg, 0.005 mmol, 5 mol%) and ligand (R)-L4 (4.07 mg, 0.006 mmol, 6 mol%) were dissolved in MTBE (1.0 mL) and stirred under nitrogen atmosphere for 20 min. at RT. The mixture was cooled to −78 °C and EtMgBr (0.15 mmol, 3.0 M in Et2O, 1.5 equiv.) was added dropwise. The reaction mixture was allowed to stir for 16 h. The reaction was quenched with HCl aqueous solution (1.0 mL, 1.0 M) and warmed to RT. The mixture was extracted with DCM(5.0 mL × 3). The combined organic phase was dried over MgSO4, filtered and evaporated on rotary evaporator. [2a:TBS-ester of 2a = 12:88, 99% total conversion, 98% ee].

(20)

121 4.4.5.3 1H NMR spectroscopic studies

Formation of 1b-TBSOTf complex

1b (4.3 mg, 0.05 mmol, 1.0 equiv.) was dissolved in CD2Cl2 (0.5 mL) in a dry NMR tube under nitrogen atmosphere and cooled down to −78 °C. TBSOTf (25 μL, 0.11 mmol, 2.2 equiv.) was added and the resulting mixture was measured by 1H NMR spectroscopy at −55 °C (see Figure 3).

In situ formation of TBS ester of 1b

1b (4.3 mg, 0.05 mmol, 1.0 equiv.) was dissolved in CD2Cl2 (0.5 mL) in in a dry NMR tube under nitrogen atmosphere. The solution was cooled down to −78 °C and TBSOTf (25 μL, 0.11 mmol, 2.2 equiv.) was added. After 20 min., MeMgBr (0.05 mmol, 3.0 M in Et2O, 1.0 equiv.) was added dropwise and the resulting mixture was measured by 1H NMR spectroscopy at −55 °C (see Figure 3).

4.4.6 General procedure for Cu-catalyzed asymmetric conjugate addition of EtMgBr to carboxylate salts formed first by deprotonation of 1a by base

In a flame-dried Schlenk tube equipped with septum and magnetic stirring bar, 1a (22.8 mg, 0.2 mmol, 1.0 equiv.), CuBr·SMe2 (2.06 mg, 0.01 mmol, 5 mol%) and ligand (R)-L4 (8.15 mg, 0.012 mmol, 6 mol%) were dissolved in MTBE (2.0 mL) and stirred under nitrogen atmosphere for 20 min. at RT. The mixture was cooled to −78 °C and the base (0.2 mmol, 1.0 equiv.) was added (when NaH was used as the base, it should be added and stirred for 1 h at RT before the mixture is cooled to −78 °C because of its low solubility in MTBE). After 5 min., TMSOTf was added, then the mixture was allowed to stir for 5 min before EtMgBr was added dropwise. The reaction mixture was allowed to stir for 16 h. The reaction was quenched with HCl aqueous solution (2.0 mL, 1.0 M) and warmed to RT. The mixture was extracted with DCM (10.0 mL × 3). The combined organic phase was dried over MgSO4, filtered and evaporated on rotary evaporator (see Supplementary Table 3).

4.4.7 General procedure for Cu-catalyzed asymmetric conjugate addition of EtMgBr to Li carboxylate A-Li with different electrophiles

In a flame-dried Schlenk tube equipped with septum and magnetic stirring bar, 1a (22.8 mg, 0.2 mmol, 1.0 equiv.), CuBr·SMe2 (2.06 mg, 0.01 mmol, 5 mol%) and ligand (R)-L4 (8.15 mg, 0.012 mmol, 6 mol%) were dissolved in MTBE (2.0 mL) and stirred under nitrogen atmosphere for 20 min. at RT. The mixture was cooled to −78 °C and n-BuLi (0.2 mmol, 2.5 M in hexane, 1.0 equiv.) was added. After 5 min., the electrophile was added, and the mixture was allowed to stir for 5 min before EtMgBr (0.5 mmol, 3.0 M in Et2O, 2.5 equiv.) was added dropwise. The reaction mixture was allowed to stir for 16 h. The reaction was quenched with HCl aqueous solution (2.0 mL, 1.0 M) and warmed to RT. The mixture was extracted with DCM (10.0 mL × 3). The combined organic phase was dried over MgSO4, filtered and evaporated on rotary evaporator (see Supplementary Table 3)

(21)

122

4.4.8 Isolation and measurement of the solubility of Mg, Li and Na carboxylates in MTBE (A-Mg, A-Li and A-Na respectively)

In a flame-dried Schlenk tube equipped with septum and magnetic stirring bar, 1a (144.1 mg, 1.0 mmol, 1.0 equiv.) was dissolved in MTBE (10.0 mL) at RT. EtMgBr (1.0 mmol, 3.0 M in Et2O, 1.0 equiv.) or n-BuLi (1.0 mmol, 2.5 M in hexane, 1.0 equiv.) or NaH (1.0 mmol, 60% in mineral oil, 1.0 equiv.) were added. The reaction mixture was allowed to stir for 16 h at RT and the metal carboxylate has precipitated. The precipitate was centrifuged and washed with MTBE (10.0 mL × 3). The precipitate was dried in vacuo during overnight to give the metal carboxylates A-Mg or A-Li or A-Na as the white solid respectively.

The metal carboxylates A-Mg or A-Li or A-Na (0.1 mmol) was added to 50 mL MTBE, and the mixture was refluxed for 3 h, followed by addition of 1,3,5-trimethoxybenzene (8.4 mg, 0.05 mmol) as the internal standard. The mixture was filtered and the filtrate was evaporated on rotary evaporator. DMSO-d6 was added and the corresponding 1H NMR spectra of the samples were recorded (see Supplementary Figures 1−3). The solubility of A-Mg: 0.4318 mM, A-Na: 0.1225 mM. No peaks of A-Li were observed because the solubility is under NMR detection limit.

Magnesium bromide hex-2-enoate (A-Mg)

1H NMR (DMSO-d6, 400 MHz): δ 6.60 (dt, J = 15.5, 7.0 Hz, 1H, CH2CH=CH), 5.76-5.69 (m, 1H, CH2CH=CH), 2.13-2.04 (m, 2H, CH2CH=CH), 1.46-1.34 (m, 2H, CH3CH2), 0.88 (t, J = 7.4 Hz, 3H, CH3).

Lithium hex-2-enoate (A-Li)

1H NMR (DMSO-d6, 400 MHz): δ 6.20 (dt, J = 14.6, 6.9 Hz, 1H, CH2CH=CH), 5.58 (dt, J = 15.3, 1.5 Hz, 1H, CH2CH=CH), 2.02-1.94 (m, 2H, CH2CH=CH), 1.42-1.30 (m, 2H, CH3CH2), 0.86 (t, J = 7.4 Hz, 3H, CH3).

Sodium hex-2-enoate (A-Na)

1H NMR (DMSO-d6, 400 MHz): δ 6.24-6.13 (m, 1H, CH2CH=CH), 5.58 (dt, J = 15.4, 1.4 Hz, 1H, CH2CH=CH), 2.02-1.93 (m, 2H, CH2CH=CH), 1.42-1.30 (m, 2H, CH3CH2), 0.86 (t, J = 7.3 Hz, 3H, CH3).

(22)

123

Supplementary Figure 1: a, Measurement of the solubility of Mg carboxylate A-Mg in MTBE. 1H NMR spectra

were obtained after evaporating MTBE and dissolving the residue in DMSO-d6 using 1,3,5-trimethoxybenzene

as the internal standard. b, 1H NMR spectrum of Mg carboxylate A-Mg in DMSO-d6 as the reference.

Supplementary Figure 2: a, Measurement of the solubility of Li carboxylate A-Li in MTBE. 1H NMR spectra

were obtained after evaporating MTBE and dissolving the residue in DMSO-d6 using 1,3,5-trimethoxybenzene

(23)

124

Supplementary Figure 3: a, Measurement of the solubility of Na carboxylate A-Na in MTBE. 1H NMR spectra

were obtained after evaporating MTBE and dissolving the residue in DMSO-d6 using 1,3,5-trimethoxybenzene

as the internal standard. b, 1H NMR spectrum of Na carboxylate A-Na in DMSO-d6 as the reference. 4.4.9 General procedure for the synthesis of α,β-unsaturated carboxylic acids

The reactions were performed according to the literature procedure.47 The corresponding alkene (10.0 mmol, 1.0 equiv.) and acrylic acid (20.0 mmol, 2.0 equiv.) were added simultaneously to a stirred solution of 1 mol% of M2 catalyst in DCM (10.0 mL) at RT. The reaction was refluxed under nitrogen for 16 h. The solvent and the remaining acrylic acid were removed under reduced pressure and the corresponding α,β-unsaturated carboxylic acids was purified by column chromatography and rinsed with pentane.

(E)-4-Phenylbut-2-enoic acid (1j)48

The crude product was purified by column chromatography on silica gel (SiO2, pentane:Et2O = 4:1) and rinsed with pentane to afford product 1j as a white solid [49%

(24)

125

yield]. 1H NMR (CDCl3, 400 MHz): δ 7.35-7.29 (m, 2H, CHAr), 7.28-7.15 (m, 4H, CHAr, CH2CH=CH), 5.82 (dt, J = 15.5, 1.7 Hz, 1H, CH2CH=CH), 3.55 (dd, J = 6.9, 1.6 Hz, 2H, CH2). 13C NMR (CDCl3, 100 MHz): δ 172.2, 150.4, 137.4, 128.9, 128.8, 126.9, 121.8, 38.67.

(E)-4-(4-Methoxyphenyl)but-2-enoic acid (1k)48

The crude product was purified by column chromatography on silica gel (SiO2, pentane:Et2O = 4:1) and rinsed with pentane to afford product 1k as a white solid [45% yield]. 1H NMR (CDCl3, 400 MHz): δ 7.19 (dt, J = 15.5, 6.7 Hz, 1H, CH2CH=CH), 7.11-7.06 (m, 2H, CHAr), 6.89-6.83 (m, 2H, CHAr), 5.79 (dt, J = 15.5, 1.6 Hz, 1H, CH2CH=CH), 3.80 (s, 3H, CH3O), 3.49 (dd, J = 6.8, 1.6 Hz, 2H, CH2). 13C NMR (CDCl3, 100 MHz): δ 172.2, 158.6, 150.8, 129.9, 129.4, 121.5, 114.3, 55.4, 37.8.

(E)-6-Bromohex-2-enoic acid (1l)

The crude product was purified by column chromatography on silica gel (SiO2, pentane:Et2O = 4:1) and rinsed with pentane to afford product 1l as a white solid [46% yield]. 1H NMR (CDCl3, 400 MHz): δ 7.04 (dt, J = 15.7, 7.0 Hz, 1H, CH2CH=CH), 5.89 (dt, J = 15.6, 1.6 Hz, 1H, CH2CH=CH), 3.42 (t, J = 6.5 Hz, 2H, BrCH2), 2.42 (m, J = 7.2, 1.6 Hz, 2H, CH2CH=CH), 2.08-2.00 (m, 2H, BrCH2CH2). 13C NMR (CDCl3, 100 MHz): δ 172.1, 150.0, 122.0, 32.6, 30.7, 30.6. HRMS (ESI, m/Z): calcd. for C6H8BrO2 [M−H]−: 190.9713, found: 190.9716.

(E)-6-(Benzyloxy)hex-2-enoic acid (1m)

The crude product was purified by column chromatography on silica gel (SiO2, pentane:Et2O = 3:1) and rinsed with pentane to afford product 1m as a white solid [39% yield]. 1H NMR (CDCl3, 400 MHz): δ 7.38-7.24 (m, 5H, CHAr), 7.09 (dt, J = 15.6, 7.0 Hz, 1H, CH2CH=CH), 5.84 (dt, J = 15.6, 1.6 Hz, 1H, CH2CH=CH), 4.50 (s, 2H, PhCH2), 3.50 (t, J = 6.2 Hz, 2H, BnOCH2), 2.38-2.30 (m, 2H, CH2CH=CH), 1.83-1.74 (m, 2H, BnOCH2CH2). 13C NMR (CDCl3, 100 MHz): δ 172.0, 151.6, 138.4, 128.5, 127.8, 127.7, 121.2, 73.1, 69.3, 29.2, 28.1. HRMS (ESI, m/Z): calcd. for C13H15O3 [M−H]−: 219.1027, found: 219.1032.

(25)

126

4.4.10 Cu-catalyzed asymmetric conjugate addition of Grignards to α,β-unsaturated carboxylic acids

4.4.10.1 General procedures

In a flame-dried Schlenk tube equipped with septum and magnetic stirring bar, the substrate (0.2 mmol, 1.0 equiv.), CuBr·SMe2 (0.01 mmol, 5 mol%), and ligand L (0.012 mmol, 6 mol%) were dissolved in the solvent (2.0 mL) and stirred under nitrogen atmosphere for 20 min. at RT. The mixture was cooled to −20 or −40 °C and TMSOTf (0.44 mmol, 2.2 equiv.) was added. After 20 min., RMgBr (0.5 mmol, 2.5 equiv.) was added dropwise by hand in 10 min. (syringe pump use is also an option), and the reaction mixture was allowed to stir for 2 h. (Note: TMSOTf should be a new bottle, and dry solvents should be freshly collected from a dry solvent purification system and used immediately. Syringe pump can be used to add Grignards in the big scale reaction, but the Grignards cannot be diluted by the solvent. Otherwise the conversion will decrease.)

4.4.10.2 General Work-up A

The reaction was quenched with HCl aqueous solution (2.0 mL, 1.0 M) and warmed to RT. The mixture was extracted with DCM (10.0 mL × 3). The combined organic phase was dried over MgSO4, filtered and evaporated on rotary evaporator. Pentane (1.0 mL × 3) was added to the residue and the mixture was filtered with a small piece of cotton in glass pipette to remove most of the catalyst. The crude was purified by flash chromatography on silica gel. 4.4.10.3 General Work-up B

The reaction was quenched with saturated NaHCO3 aqueous solution (2.0 mL), warmed to room temperature and the organic phase was extracted. The organic phase was further extracted with saturated NaHCO3 aqueous solution (2.0 mL) for another two times. The combined aqueous phase was acidified with HCl aqueous solution (1.5 mL, 12.0 M), and extracted with DCM (10.0 mL × 3). The combined organic phase was dried over MgSO4, filtered and evaporated on rotary evaporator.

4.4.10.4 General Work-up C

The reaction was quenched with saturated NaHCO3 aqueous solution (2.0 mL), warmed to room temperature and the organic phase was extracted. The organic phase was further extracted with saturated Na2CO3 aqueous solution (2.0 mL) for another three times. The combined aqueous phase was acidified with HCl aqueous solution (3.0 mL, 12.0 M), and extracted with DCM (10.0 mL × 3). The combined organic phase was dried over MgSO4, filtered and evaporated on rotary evaporator.

4.4.10.5 General procedure for the synthesis of racemic conjugate addition products In a flame-dried Schlenk tube equipped with septum and magnetic stirring bar, the substrate (0.2 mmol, 1.0 equiv.), CuBr·SMe2 (0.01 mmol, 5 mol%) and THF (2.0 mL) were added. The mixture was cooled to −20 °C and TMSOTf (0.6 mmol, 3.0 equiv.) was added. After 20 min., RMgBr (0.6 mmol, 3.0 equiv.) was added dropwise, and the reaction mixture

(26)

127

was allowed to stir for 2 h. The reaction was quenched with HCl aqueous solution (2.0 mL, 1.0 M) and warmed to RT. The mixture was extracted with DCM (10.0 mL × 3). The combined organic phase was dried over MgSO4, filtered and evaporated on rotary evaporator. The crude was purified by flash column chromatography on silica gel.

4.4.10.6 Procedure for the preparative scale (10 mmol) Cu-catalyzed asymmetric conjugate addition of EtMgBr to 1a and the recovery of the chiral catalyst L4/Cu In a flame-dried three-neck round-bottom flask equipped with septum and mechanistic stirring bar, the substrate 1a (1.14g, 10.0 mmol, 1.0 equiv.), CuBr·SMe2 (102.8 mg, 0.5 mmol, 5 mol%) and ligand (R)-L4 (407.3 mg, 0.6 mmol, 6 mol%) were dissolved in MTBE (50 mL) and stirred under nitrogen atmosphere for 20 min. at RT. The mixture was cooled to −20 °C and TMSOTf (3.98 mL, 22 mmol, 2.2 equiv.) was added. After 20 min., EtMgBr (8.33 mL, 25 mmol, 2.5 equiv.) was added with syringe pump in 20 min., and the reaction mixture was allowed to stir for 2 h. The reaction was quenched with water (10.0 mL) and warmed to RT. The aqueous phase was discarded and the organic phase was extracted with saturated NaHCO3 aqueous solution (50.0 mL × 3). In this step, the chiral catalyst L4/Cu(I) is in the organic phase while the ACA product 2a is in the aqueous phase. The organic phase was washed with HCl aqueous solution (10.0 mL, 1.0 M), dried over MgSO4, filtered and evaporated on rotary evaporator. The residue was rinsed with a little pentane and dried in vacuo for overnight to afford the recovered chiral catalyst L4/Cu(I) as a light yellow powder [83% yield]. The combined aqueous phase was acidified with HCl aqueous solution (50.0 mL, 12.0 M), and extracted with DCM (100.0 mL × 3). The combined organic phase was dried over MgSO4, filtered and evaporated on rotary evaporator to yield the product 2a as a colorless oil [83% yield, 97% ee].

4.4.10.7 Specific experimental details and product characterization (R)-3-Ethylhexanoic acid (2a)49

The reaction was performed with 1a (22.8 mg, 0.2 mmol, 1.0 equiv.), CuBr·SMe2 (2.06 mg, 0.01 mmol, 5 mol%), ligand (R)-L4 (8.15 mg, 0.012 mmol, 6 mol%), TMSOTf (80 μL, 0.44 mmol, 2.2 equiv), EtMgBr (0.5 mmol, 3.0 M in Et2O, 2.5 equiv.), MTBE (2.0 mL) at −20 °C, and following General Work-up B. Product 2a was obtained as a colorless oil without further purification [97% conversion, 91% yield, 97% ee]. 1H NMR (CDCl3, 400 MHz): δ 2.28 (d, J = 6.9 Hz, 2H, CH2CO2H), 1.88-1.77 (m, 1H, CH), 1.45-1.22 (m, 6H, CH2), 0.93-0.85 (m, 6H, CH3). 13C NMR (CDCl3, 100 MHz): δ 180.4, 38.7, 36.2, 35.8, 26.4, 19.8, 14.4, 10.9. The

ee of this compound was determined from the corresponding N,N-dimethyl amide

derivative. HPLC: Chiracel-OBH, n-heptane/i-PrOH 98:2, 0.5 mL/min., 40 °C, detection at 206 nm. Retention time (min): 14.7 (minor) and 16.7 (major).

(27)

128

(S)-3-Methylpentanoic acid (2b)49

The reaction was performed with 1b (17.2 mg, 0.2 mmol, 1.0 equiv.), CuBr·SMe2 (2.06 mg, 0.01 mmol, 5 mol%), ligand (R)-L4 (8.15 mg, 0.012 mmol, 6 mol%), TMSOTf (80 μL, 0.44 mmol, 2.2 equiv), EtMgBr (0.5 mmol, 3.0 M in Et2O, 2.5 equiv.), MTBE (2.0 mL) at −20 °C, and following General Work-up B. Product 2b was obtained as a colorless oil after column chromatography (SiO2, pentane:Et2O = 5:1) [98% conversion, 74% yield, 95% ee]. 1H NMR (CDCl3, 400 MHz): δ 2.36 (dd, J = 15.0, 6.0 Hz, 1H, CHHCO2H), 2.15 (dd, J = 15.0, 8.1 Hz, 1H, CHHCO2H), 1.96-1.82 (m, 1H, CH), 1.46-1.33 (m, 1H, CH3CHH), 1.32-1.19 (m, 1H, CH3CHH), 0.97 (d, J = 6.7 Hz, 3H, CH3CH), 0.90 (t, J = 7.4 Hz, 3H, CH3CH2). 13C NMR (CDCl3, 100 MHz): δ 180.1, 41.4, 31.9, 29.4, 19.4, 11.4. The ee of this compound was determined from the corresponding N,N-dimethyl amide derivate. HPLC: Chiracel-OBH, n-heptane/i-PrOH 90:10, 0.5 mL/min., 40 °C, detection at 206 nm. Retention time (min): 10.7 (major) and 13.1 (minor).

(S)-3-Cyclohexylpentanoic acid (2c)50

The reaction was performed with 1c (30.8 mg, 0.2 mmol, 1.0 equiv.), CuBr·SMe2 (2.06 mg, 0.01 mmol, 5 mol%), ligand (R)-L4 (8.15 mg, 0.012 mmol, 6 mol%), TMSOTf (80 μL, 0.44 mmol, 2.2 equiv), EtMgBr (0.5 mmol, 3.0 M in Et2O, 2.5 equiv.), MTBE (2.0 mL) at −20 °C, and following General Work-up A. Product 2c was obtained as a colorless oil after column chromatography (SiO2, pentane:Et2O = 15:1) [98% conversion, 83% yield, 98% ee]. 1H NMR (CDCl3, 400 MHz): δ 2.36 (dd, J = 15.4, 6.0 Hz, 1H, CHHCO2H), 2.19 (dd, J = 15.4, 7.6 Hz, 1H, CHHCO2H), 1.79-1.56 (m, 6H, CH2, CH), 1.49-0.93 (m, 8H, CH2), 0.89 (t, J = 7.4 Hz, 3H, CH3). 13C NMR (CDCl3, 100 MHz): δ 181.0, 42.0, 40.2, 36.1, 30.2, 29.3, 26.9, 26.9, 26.8, 24.0, 11.8. The ee of this compound was determined from the corresponding N,N-dimethyl amide derivative. HPLC: Chiracel-OZH, n-heptane/i-PrOH 95:5, 0.5 mL/min., 40 °C, detection at 207 nm. Retention time (min): 17.0 (major) and 18.4 (minor).

(S)-3-Cyclopropylpentanoic acid (2d)

The reaction was performed with 1d (22.4 mg, 0.2 mmol, 1.0 equiv.), CuBr·SMe2 (2.06 mg, 0.01 mmol, 5 mol%), ligand (R)-L4 (8.15 mg, 0.012 mmol, 6 mol%), TMSOTf (80 μL, 0.44 mmol, 2.2 equiv), EtMgBr (0.5 mmol, 3.0 M in Et2O, 2.5 equiv.), MTBE (2.0 mL) at −20 °C, and following General Work-up B. Product 2d was obtained as a colorless oil without further purification [99% conversion, 83% yield, 93% ee]. 1H NMR (CDCl3, 400 MHz): δ

Referenties

GERELATEERDE DOCUMENTEN

Our approach to synthesize them by asymmetric addition of Grignard reagents to ketones and ketimines is based on the discovery made by our group in 2012 of a Cu(I)-chiral

The same trend found earlier for the addition to enones repeated again for aryl alkyl ketones: β-branched Grignard reagents were required in order to obtain products

In summary, we have demonstrated that it is possible to achieve catalytic asymmetric addition of organometallic reagents to stereochemically challenging

an amenable substrate and product 5c was obtained with full retention of enantioselectivity as a mixture of diastereoisomers (3:1). Activated ketones were

In summary, in this chapter we have explored the Brook rearrangement of simple, chiral tertiary benzylic α-hydroxysilanes. Brook rearrangement can be followed by trapping of methyl

I want to thank Pieter, Johan, Theodora and Monique for the help with NMR, HRMS, HPLC and GC-MS.. Annete, I would like to thank you for the

Activated by Lewis acid and controlled by a chiral copper catalyst, addition of EtMgBr to quinoline, followed by trapping with acetyl chloride was successful,

Copper-catalyzed enantioselective conjugate addition of Grignard reagent to non-activated acceptors enabled by Lewis acid.. van