99
Construction of Adjacent Quaternary and Tertiary Stereocenters via an Organocatalytic Transformation of Morita-Baylis-Hillman Carbonates Dirk Jan van Steenis

Construction of adjacent quarternary and tertiary … · Web viewTitle Construction of adjacent quarternary and tertiary stereocenters via a organocatalytic transformation of Morita-Baylis-Hillman

  • Upload
    others

  • View
    2

  • Download
    0

Embed Size (px)

Citation preview

Construction of adjacent quarternary and tertiary stereocenters via a organocatalytic transformation of Morita-Baylis-Hillman carbonates

H

O

O

R

H

O

-

O

R

R

3

+

N

P

h

C

H

O

O

O

R

R

3

+

N

P

h

O

-

H

N

R

3

O

O

R

P

h

O

H

H

O

R

3

+

N

P

h

-

O

E

H

H

R

B

A

A

1

)

2

)

3

)

1

2

Construction of Adjacent Quaternary and Tertiary Stereocenters via an Organocatalytic Transformation of Morita-Baylis-Hillman Carbonates

O

H

R

E

W

G

+

R

*

O

H

E

W

G

B

a

s

e

P

h

O

B

o

c

O

O

M

e

b

-

I

C

P

D

1

a

N

O

H

N

O

*

O

M

e

P

h

N

u

N

O

H

N

O

(

E

)

O

M

e

N

O

H

N

O

O

M

e

P

h

N

O

H

N

O

B

o

c

-

C

O

2

t

-

B

u

O

N

u

H

N

u

P

h

*

N

u

O

O

M

e

3

a

2

a

A

B

C

P

h

t

-

B

u

O

H

*

*

C

O

O

M

e

R

C

O

O

E

t

C

N

R

'

N

N

O

O

H

Dirk Jan van Steenis

Report Research project, master Molecular Design Synthesis and Catalysis (MDSC)

What is organocatalysis?

“Organocatalysis is the acceleration of a chemical transformation through addition of a sub-stoichiometric amount of an organic compound which does not contain a metal atom”

- Peter I. Dalko and Lionel Moisan

“Catalytic reactions mediated by small organic molecule in absence of metals or metal ions.”

- Carlos F. Barbas, III

“ A field of chemistry that pays my mortgage and has gotten me many free dinners.”

- David W. C. Macmillan

A part of this report has been submitted for publication:

Dirk Jan V. C. van Steenis, Tommaso Marcelli, Martin Lutz, Anthony L. Spek, Jan H. van Maarseveen, Henk Hiemstra*. 2006.

Table of Contents

4Summary

5List of abbreviations

7Chapter 1; Introduction

71.1 Organocatalysis

91.2 Cinchona Alkaloids

121.2.1 Bi-functional cinchona alkaloids

141.3 Morita-Baylis-Hillman (MBH) Reaction

151.3.1 The Morita-Baylis-Hillman mechanism

161.3.2 Morita-Baylis-Hillman Adducts

191.4 Dynamic Kinetic Asymmetric Transformation (DYKAT) of Morita-Baylis-Hillman Adducts

211.5 Functionalization of Morita-Baylis-Hillman Adducts via a Tandem SN2’-SN2’ Mechanism

251.6 All-carbon Quaternary Stereocenters

Chapter 2; Allylic Nucleophilic Substitution of Morita-Baylis-27Hillman Carbonates

272.1 Goal

292.2 Catalyst Synthesis

322.3 Preliminary Experiments and Optimization

352.4 Results

352.4.1 Substrate scope

382.4.2X-Ray analysis

2.4.3Mechanistic considerations40

412.5 Conclusions

43Samenvatting

44Acknowledgements

45Chapter 3 Experimental

References70

Summary

Organocatalysis, catalytic transformations with small organic molecules, has found renewed interest in both academia and industry. Although it was already known for a century that organic compounds could catalyze asymmetric reactions, it is only since half a decade that the potential of organocatalysis is understood and is starting to be fully explored. In the treatment of diseases, one enantiomer of a medicine is usually more potent, and in the worst scenario, the opposite enantiomer can cause serious side effects or even death. Therefore, the demand from both pharmaceutical and chemical industry for new reliable asymmetric transformations of molecular skeletons is higher then ever. Nowadays, the construction of (highly) functionalized asymmetric skeletons still suffers from drawbacks, especially when there is a quaternary stereocenter involved in the target molecule. Organocatalysis gained popularity in a relatively short time span, because it has led to a large assortment of new asymmetric demanding transformations in the last five years. In some cases, organocatalysts meet the selectivity and efficiency levels of established metal catalyzed organic reactions. Since organocatalysis has a hidden potential, it could provide a solution for challenging alterations in the future. Moreover, organocatalysis can serve as a green alternative for transition metal catalyzed reactions. During this master research project in organocatalysis we found that the organocatalyst β-isocupreidine (β-ICPD) effectively catalyzes a one-step asymmetric transformation of Morita-Baylis-Hillman carbonates. This transformation led to new molecular assemblies with vicinal quaternary and tertiary stereocenters. We observed high chemo-, diastereo-, and enantioselectivities with this reaction catalyzed by β-isocupreidine, in addition the chemical yields of these transformations are excellent. This one-step organocatalytic allylic alkylation is the first, out of six reactions reported, in which the reaction mechanism can not be only explained in terms of a conjugate addition and thereby leading to adjacent quaternary and tertiary stereocenters.

List of abbreviations

H2O: water

CDCl3:

deutero-chloroform

CHCl3: chloroform

DCM: dichloromethaneCH2Cl2:

dichloromethaneEtOAc:

ethyl acetate

EI:

electron impact

Et2O:

diethyl ether

FAB:

fast atom bombardment

1H NMR:

proton1 nuclear magnetic resonance

13C NMR:

carbon13 nuclear magnetic resonance

HRMS:

high resolution mass spectroscopy

J: coupling constant

Hz:

hertz

IR:

infra red

FTIR:

Fourier transform infra red

UV: ultraviolet

K2CO3:

potassium carbonate

MD3OD:

deutero-methanol

MeOH:

methanol

MgSO4:

magnesium sulfate

MHz:

megahertz

MS:

mass spectrometer

nm:

nanometer

PE:

petroleum ether

HCl: hydrochloric acid

KOH: potassium hydroxide

H3PO4: phosphoric acid

N2: nitrogen gas

DBU:

2,3,4,6,7,8-hexahydropyrrolo[1,2-a]pyrimidine

DABCO: 1,4-diaza-bicyclo[2.2.2]octane

DMAP: N,N-dimethylpyridin-4-amine

CO2: carbon dioxide

TLC: thin layer chromatography

Boc2O: tert-butoxycarbonyl anhydride

THF:

tetrahydrofuran

DMSO: dimethylsulfoxide

ee: enantiomeric excess

dr:

diastereomeric ratio

h: hours

min:

minutes

Pd: palladium

Boc: tert-butylcarbonate

OAc: acetate

BINAP:2-(diphenylphosphino)-1-(2-(diphenylphosphino)naphthalen-1-yl)naphthalene

NH3: ammoniac

UV:

ultra violet

AAA:

asymmetric allylic alkylation

DYKAT:

dynamic kinetic asymmetric transformation

KAT

kinetic asymmetric transformation

MBH

Morita-Baylis-Hillman

Å:

Angström

Eq.:

equivalent

KBr:

potassium bromide

Mp:

melting point

H:

proton

λ:

Lambda

HPLC:

high performance liquid chromatography

s:

singlet

d:

doublet

m:

multiplet

dd:

double doublet

t:

triplet

mmol:

millimoles

M:

molarity

Chapter 1; Introduction

1.1 Organocatalysis

Less then a decade ago it was generally accepted that highly efficient asymmetric transformations where only restricted to chiral organometallic complexes and enzymes. Indeed, the levels of regio- and enantioselectivity achieved nowadays by these catalyst systems is impressive. The number of organic reactions reported in the last decades catalyzed by these complexes has enormously contributed to our quality of life. The oil, chemical and pharmaceutical industry and even our economy, are extensively relying on these ways of influencing molecular properties. Nevertheless, the need for new and reliable asymmetric transformations is high. The amount of new chiral drugs introduced on the market is higher then ever and subsequently the demand for new efficient asymmetric transformations. Nature is chiral and does everything with tremendous (asymmetric) efficiency, moreover, in multiple reactions catalyzed by nature, there is no metal involved. Apparently nature does not always need metals for challenging transformations. Scientists nowadays use nature as an inspiration source and try to mimic these reactions in the lab. The concept of organocatalysis can be seen as a blackboard example of this. In recent years asymmetric organocatalysis, is regaining interest and can be considered as a rapidly expanding research field. In addition, organocatalysis has proven in the last years to be a valuable alternative in respect to the traditional asymmetric catalytic methods, a few examples are given below;

· Cheap catalyst sources with respect to organometallic catalysts

· No toxic transition metals in the product (catalyst leaching).

· Environmentally friendly

· Usually less demanding reaction conditions

Evidence has been found that this metal free type of catalysis has played a important-role in the formation of essential key-building blocks for life. The natural amino acids, L-alanine and L-isovaline which can catalyze for instance certain Aldol reactions, have been found in an enantiomeric excess (ee) of 15 % on meteorites1b. Although the first asymmetric transformation with a small organic molecule was reported in 1912, by Bredig and Fiske, it is no more then a few years ago that the potential of organocatalysis has been understood by the scientific community. List and Barbas reported in 2000 a breakthrough in organocatalysis and “pulled the trigger”. They reported that simple proline catalyzes an aldol reaction in good enantiomeric excesses. Since then, organocatalysis is starting to get the attention it deserves and has become a mature concept in the field of (homogenous)catalysis and in some cases organocatalysts meet the selectivity and efficiency levels of established metal catalyzed organic reactions. During the ninety years of absence, only a few scientists understood the potential of metal free catalysis. In the sixties, Pracejus applied cinchona alkaloids in the asymmetric conversion of ketenes to (S)-methyl hydratropate. The seventies brought an other milestone, Hajos and Parrish reported a L-proline catalyzed Robinson annulation in excellent enantioselectivities. Moreover, Wiechert reported an organocatalyic aldol reactions in good enantioselectivities. In the eighties, Wynberg and co-workers reported various 1,2 and 1,4 additions catalyzed by cinchona alkaloids. Besides these notable reports, the field of organocatalysis has been remarkably overlooked. In respect to the traditional transition metal based catalysts, organocatalysts are often inexpensive, environmentally friendly and the reactions can usually be performed under aerobic reaction conditions. Different from enzymes, organocatalysts do not require buffer conditions, and therefore have less solubility problems and have a broad substrate scope. Moreover, organocatalysts can usually simply be separated from reaction mixtures without considerable catalyst leaching and recovered via extraction. An other interesting and attractive feature of organocatalysts is that they usually possess an excellent functional group tolerance. These characteristics can make them practically and interesting counterparts of traditional homogenous catalysts and enzymes, therefore it is justified to continue further exploration of organocatalysis.

N

N

O

O

S

O

O

F

F

F

N

N

O

N

The vast majority of (asymmetric) organic reactions can be explained by a nucleophile and electrophile mechanism. Organocatalysts can be divided into several subclasses namely; Lewis bases, Lewis acids, Brønsted bases and Brønsted acids1d. The majority of the organocatalyst published to date works via an Lewis base mechanism (Fig. 1). For simplicity reasons, only the Lewis base and the Brønsted acid mechanism are taken into account in this report. The simplified Lewis base mechanism can be best described by the following steps; the Lewis base (B), for instance a cinchona alkaloid, starts the catalytic cycle via a nucleophilic addition to, or deprotonation of, the substrate (S). The resulting chiral intermediate undergoes a reaction and the newly formed product (P) is separated from the catalyst. The catalyst is now regenerated and available for a new catalytic cycle. Many organocatalytic reactions described in literature cannot simply be explained by a single Lewis base mechanism. Often there is a second mechanism involved, usually a Brønsted acid mechanism. This type of organocatalysis is called bi-functional organocatalysis. Cinchona bi-functional organocatalysis was first introduced by Wynberg and coworkers7d. They proposed that both a base and an acid can be crucial in activation and orientating the reaction partners, an elegant example being the asymmetric addition of aryl-thiols to conjugated cycloalkenones7c. In a bi-functional catalyzed reaction, the organocatalyst bears besides a Lewis base, typical a nitrogen or phosphorus atom, also a Brønsted acid functionality. These facts have been only recently understood and are only since the last years frequently applied, although it was already known in the eighties. Especially, activation and orientation through thio-urea hydrogen bonding is emerging. Other widely applied Brønsted acids functionalities in organocatalysis are urea and alcohol functional groups. Similar to enzyme catalysis, there is now also hydrogen bonding involved in the transition state. This kind of hydrogen bonding usually makes the electrophile more electron deficient and therefore more prone to nucleophilic attack. Besides activation, the hydrogen bonding motive holds the electrophile also in a kind of a conformational lock (orientation). Via this conformational lock higher enantioselectivities can be obtained. Brønsted acid activation through thio-ureas has been applied in asymmetric reactions such as the Michael (conjugate) addition, (aza)-Henry, Strecker and Friedel-Craft alkylations.

1.2 Cinchona Alkaloids

For practical and economical considerations, an important prerequisite for an organocatalyst is that both enantiomers of the catalyst are readily available. Now, there is access to both enantiomers of a product with usually similar ranges of asymmetric induction. Cinchona alkaloids are a natural class of compounds that exhibit this unique feature and makes them highly attractive candidates for asymmetric catalysis. Quinine (QN) / Quinidine (QD), and Cinchonine (CN) / Cinchonidine (CD) (Fig. 2), are two pairs of so called pseudoenantiomers. These pseudoenatiomers are not enantiomers, because they are not mirror images.

N

N

O

N

H

2

N

N

O

N

H

S

N

H

F

3

C

C

F

3

They are called pseudoenantiomers because they are “almost enantiomeric pairs”. To date there are more then twenty different cinchona alkaloid producing plants known, by far the most important is the “Cinchona pubescens”. To date, more then thirty different kinds of cinchona alkaloids have been isolated and characterised. The tree bark of the Cinchona pubescens can contain over 50 % of the main four cinchona alkaloids which can be easily gathered, via extraction. The cinchona alkaloids family forms a unique class of natural compounds and has found widespread application, such as in the pharmaceutical, chemical and beverage industry and are produced on a multi-ton scale (aprox. 700 t/year) and therefore readily available. Already three-hundred years ago, quinine (QN) received attention because of its biological activity although the structure was not understood yet. Quinine (QN) was officially isolated and reported by Pelletier in 1820 In the past decades quinine (QN) has been extensively studied and used for the treatment of malaria. However, alternatives for the treatment of malaria have been developed due to side effects of quinine. Nowadays cinchona alkaloids are mainly used by the chemical and beverage industry. Quinine (QN) is used by the soft drink industry for the typical bitter flavour. The chemical compositions of cinchona alkaloids differ, but usually they bear two key characteristics; they all have an aromatic quinoline part and usually a basic quinuclidine moiety in the molecular skeleton (Fig.2). The pseudoenantiomers, quinine (QN) / quinidine (QD) have an additional functionality; they bear at the 6’-position of the aromatic quinoline part a methoxy-group. This functionality gives a handgrip for further functionalization. A milestone achieved by making use of the chiral properties of cinchona alkaloids, was the resolution of racemic

C

N

P

h

C

O

O

E

t

C

N

tartaric acid. Pasteur, achieved the separation of the enantiomeric pair by using the cinchona alkaloid derivatives quinicine and cinchonicine (Fig. 3). The first asymmetric reaction catalyzed by a cinchona alkaloid was reported in 1912 by Bredig and Fiske2. The asymmetric addition of hydrogen cyanide to benzaldehyde was reported in an

C

O

O

M

e

P

h

C

O

O

E

t

C

N

B

r

enantiomeric excess (ee) of 20 %. Further development of cinchona alkaloids in asymmetric catalysis was reported in 1960 by Pracejus in the asymmetric conversion of ketenes4. An other cinchona alkaloid pioneer, Wynberg, expanded the reactions catalyzed by cinchona alkaloids to various 1,2 and 1,4 additions.7 In the last two decades cinchona alkaloids are recognized as a privileged class of chiral auxiliaries and have found a variety of applications in asymmetric synthesis.

1.2.1 Bi-functional cinchona alkaloids

C

O

O

M

e

C

O

O

E

t

P

h

C

N

O

2

N

C

O

O

M

e

P

h

C

O

O

E

t

C

N

In order to enlarge to scope of chemical transformations catalyzed by cinchona alkaloids, many efforts have been put into the modification of these natural compounds. In the past two decades modified cinchona alkaloids have led to excellent results in various research fields. In phase transfer catalysis (PTC), quaternary ammonium ions derived from cinchona alkaloids have played an important role. The quinuclidine nitrogen atom can perfectly undergo transformation into a quaternary ammonium ion and could serve as a phase transfer catalyst. Cinchona alkaloids have played a key role in asymmetric dihydroxylation reactions, C9-dimerized Cinchona alkaloids have successfully been applied as chiral ligands in the Sharpless asymmetric dihydroxylation (AD). To stress the importance of these ligands; various analogues of these dimers are now commercially available. Barry K. Sharpless received the Nobel prize in chemistry in 2001 for his contribution to asymmetric synthesis. The cinchona dihydroxylation ligands formed a substantial part of his work. Other alterations, different than at the C9 position of the cinchona alkaloid, usually occur at the C6’-position of quinine (QN) and quinidine (QD). When this position is changed from a methoxy into a Brønsted acid functionality, these organocatalysts become bi-functional (Fig. 4) as discussed in paragraph 1.1. These types of organocatalysts are often referred to as cupreines and cupreidines. In 1999, Hatakeyama proved for the first time that when a C6’ methoxy is transformed into a C6’-OH Brønsted acid function, high enantioselectivities could be obtained via this moiety. Since then bi-functional cinchona organocatalysts are fully explored. In the past few years several bi-functional modified quini(di)ne catalysts have been published A broad spectrum of reactions are catalyzed with excellent (enantio)selectivities. The reactions published in the past few years have been performed with surprisingly simple and accessible organocatalysts. Besides unmodified cinchona alkaloids, three different kind of successful catalysts classes for the creation of chiral skeletons can be distinguished in literature namely;

1) Quinidine derived oxaza-twistanes (i.e. β-isocupreidine)11

2) C6’ hydroxyl and C9 alkylated cupreidine catalysts15

3) a C6 or C9 thiourea equipped quini(di)ne 16

C

O

O

M

e

C

O

O

E

t

C

N

C

l

C

O

O

M

e

P

h

C

O

O

E

t

C

N

C

O

O

M

e

P

h

C

O

O

E

t

C

N

O

C

N

O

O

Constrained oxaza-twistanes (Fig. 5), are a class of catalysts whose the synthesis has been mainly explored by the group of Hofmann. These compounds posses special characteristics, due to the extra constrained ring between C9 and C3, in respect to other cinchona alkaloids. Especially β-isocupreidine, (IUPAC name; 3R, 8R, 9S-10, 11- Dihydro-3,9-epoxy-6’-hydroxycinchonane), derived in one step from quinidine, has proven to be a excellent catalyst in various reactions. The β-isocupreidine quinuclidine nitrogen atom owes, due to the extra cycle, more basic and nucleophilic character. Moreover, β-isocupreidine has less conformations due to the extra ring. β-Isocupreidine has been successfully applied to multiple reactions such as the asymmetric Morita-Baylis-Hillman reaction(see later) and the asymmetric aza-MBH reaction between the strongly activated Michael acceptor 1,1,1,3,3,3-hexa-fluoroisopropyl acrylate and aryl-imines. β-isocupreidine has also been applied in the construction of quaternary stereocenters51f. The enantio-complementary catalyst of β-Isocupreidine has recently been synthesized. Although the synthesis involves 19 steps, now there is access to both enantiomers of a product. A second superior catalyst class has been introduced by Deng and co-workers. C6’ hydroxyl and C9 alkylated cupreidine catalysts, again easily accessible, gave excellent levels of enantio- and diastereo-discrimination in a broad compilation of conjugate additions. In this elegant collection of papers published by Deng, it is demonstrated how various 1,3-dicarbonyl, 1,3-nitro-carbonyl and 1,3 cyano-carbonyl pronucleophiles can react with simple Michael-acceptors. These catalysts reported by Deng, can also simultaneously construct adjacent quaternary and tertiary stereocenters, which is a tremendous challenge in synthetic organic chemistry. These reported asymmetric reactions elegantly show how via organocatalysis highly functionalised skeletons are accessible with relatively simple catalysts. (Fig. 6) The third expanding catalyst category is thiourea modified quinidines15. Two versions have been published recently, the C6´ 9e and C9 thiourea equipped quinidine. Although the catalyst synthesis requires additional steps compared to the catalysts of Deng, they have important advantages. Due to the electron deficient thio-urea moiety of the catalyst, high levels of enantiodiscrimination can be achieved like in a fundamental reaction like the Henry reaction. The organocatalytic Henry reaction, a reaction between an aldehyde and typical nitro-(m)ethane, has recently been elucidated in high enantioselectivities9e, 15. The catalyst of choice was a C6’ thiourea equipped quinidine with a the C9 carbon a benzyl protective group (fig 6,C6’ THQD). In these two papers dealing with cinchona catalyzed Henry reaction, it is verified that the C6’ thiourea moiety is a stronger hydrogen donor compared to the C6’ hydroxyl donor (BnCPD).

1.3 Morita-Baylis-Hillman (MBH) Reaction

O

C

O

O

M

e

O

O

B

r

O

C

O

O

M

e

O

O

The Morita-Baylis-Hillman reaction was first described in 1968 by Morita; he observed a phosphine-derivative catalyzed addition of an aldehyde to an acrylate, yielding densely funtionalised β-hydroxy-α-methylene esters (Fig. 7). He named the reaction the “Carbinol Addition”. In 1972 the reaction was reinvented and patented by two chemists at the Celanese Corporation in New York, Anthony Baylis and Melville E. D. Hillman. The yield described by Morita was poor but Baylis and Hillman found that if 1,4-DiAzaBiCyclo[2.2.2]Octane (DABCO) was used instead of the phosphine derivative, the yield was roughly three times higher. After its discovery, the Morita-Baylis-Hillman reaction was almost “forgotten” for a decade, partly due to the slow reaction rate (typically days). However, the Morita-Baylis-Hillman is recognized now as an important carbon-carbon bond forming reaction like the aldol, Grignard, Friedel-Crafts, Heck and Suzuki reaction. The MBH reaction has become a well-established reaction due to fact that it meets several important standards; the MBH reaction generates multiple functional groups and thereby it provides a simple and atom economic carbon-carbon bond forming process. Moreover, the MBH reaction encloses an immense synthetic potential, almost every aldehyde and activated olefin undergoes the MBH reaction. Supporting evidence for this statement can be found with a simple search in literature. The amount of papers dealing with MBH chemistry is impressive. Up to now the MBH reaction emerged to an important classic “text book” organic reaction but still faces major challenges, illustrative features being the selective formation of the tertiary stereocenter and rate. Recently, Verkade and co-workers have made an impressive step forward. A highly active two catalyst combination, relying on a proazaphosphatrane organocatalyst and titanium-tetra-chloride was reported. Although the products are obtained racemic and the catalysts are air-sensitive, the methodology presented tolerates a broad substrate scope. For more then a decade the selective formation of the tertiary stereocenter has become an enormous challenge. Selective formation of this stereocenter would provide a route to optically enriched β-hydroxy-α-methylene esters, ketones, nitriles etc. The first breakthrough in the asymmetric MBH reaction was reported in 1999 by Hatakeyama and coworkers17. They used β-isocupreidine (Fig. 5) to catalyze the asymmetric MBH reaction between aldehydes and the strongly activated Michael acceptor, 1,1,1,3,3,3-hexa-fluoroisopropyl acrylate. Manifold chiral catalysts have been tried, including Brønsted acids and derivatized binaphthtyls. So far high enantioselectivities have not been reported with simple acceptors such as methyl acrylate and acrylonitrile. A recent trend in the asymmetric (aza)-MBH reaction is the use of catalysts combinations. Usually DABCO and a chiral thiourea are used.

1.3.1 The Morita-Baylis-Hillman mechanism

The Morita-Baylis-Hillman reaction involves a three component, nucleophilic amine or phosphine mediated addition of an aldehyde to an activated olefin. The commonly accepted mechanism involves in the first step 1) a reversible conjugate (Michael-type) addition of the nucleophilic catalyst to activated olefin (acrylate). The resulting zwitterionic enolate (1)(Fig. 8) can subsequently undergo two pathways; elimination of the catalyst or an aldol-like nucleophilic attack on the aldehyde 2) giving the second zwitterionic intermediate (2). Third, the second zwitterionic intermediate can undergo now again two reaction pathways. Reaction pathway 3) yielding the product, comprises an elimination and transfer of a proton. The other pathway is the elimination of benzaldehyde end thereby the intermediate is returning into the first zwitterionic species (1). Direct evidence for this reaction mechanism was never presented; it was only based on assumptions. Only direct evidence for the zwitterionic species (1) was presented by Drewes and co-workers it was assumed that the rate limiting step (RLS) of the MBH reaction is the elimination/transfer of a proton step 3) since a species like (1) could be isolated and analysed by X-ray crystallography. In 2004, the groups of Coelho and Eberlin, did a fundamental study, by electrospray ionization (ESI), detecting the intermediates in the MBH reaction. They confirmed the generally assumed MBH reaction mechanism. In 2005 the group of Aggarwal and co-workers, published a study towards the kinetics and mechanism of the MBH reaction in a-protic solvents. Interesting results have been reported. During the MBH reaction when an excess of starting materials are present, the RLS is the elimination/transfer of the proton in step (3). When significant amounts of products are present, the RLS is 2. This is due to hydrogen bond donor capabilities of the product (Fig. 8, intermediate A), the product can promote, via hydrogen bonding, elimination and hydrogen transfer of step 3. This effect can be described as autocatalysis.

O

C

O

O

M

e

O

O

O

2

N

O

C

O

O

M

e

O

O

C

l

1.3.2 Morita-Baylis-Hillman Adducts

Nowadays the Morita-Baylis-Hillman chemistry can be divided into three major research domains namely;

1) Improvement of rate, yield and elucidation of the reaction mechanism

2) Selective formation of the tertiary stereocenter

3) Chemical development of a various transformations of MBH adducts

O

C

O

O

M

e

O

O

O

C

O

O

M

e

O

O

Key developments of first two domains have been shortly overviewed in paragraph 1.3 and 1.3.1, the third, transformations of MBH products, will be discussed shortly in the following text. So far, in the literature many papers appeared where MBH adducts are used to obtain compounds with certain molecular properties. Therefore various efficient transformations of MBH products have been reported and methodologies have been developed. MBH adducts have proven to be important intermediates in various applications, such as total synthesis. As discussed before the MBH reaction relies on an atom economic process, yielding three functionalities is close proximity. These three functionalities, an alcohol, an alkene and an electron withdrawing group (usually an ester), are essential key functional groups in organic chemistry today. In principle, these key functionalities are susceptible of stereo-, chemo- and enantioselective transformations and therefore MBH adducts posses an enormous synthetic potential all together. Upon modification of MBH alcohols/adducts various known organic compounds classes can be obtained like; (γ-butyro-) lactones, β-lactams, quinolines, indolizines, methylene-dioxanones, pyrrolidines, coumarins, naphthalenes and derivatives thereof. Moreover, MBH adducts undergo (perfectly) a tremendous scope of classic reactions such as: asymmetric dihydroxylation (AD), epoxidation, ring-closing metathesis, aldol condensations, aminohydroxylation, radical cyclizations, Heck, Friedel-Craft and hydrogenation. Many other reductions and photochemical reactions also have been reported25. In figure (Fig. 9), a small selection of successfully applied alterations is presented.

Especially, for the construction of highly functionalized molecular skeletons, like natural compounds, MBH adducts can be interesting building blocks. MBH adducts have been often used multifarious in total synthesis, examples being (-)-mycestericin E by Hatakeyama et al. (Fig. 10), Pinnatoxin A by Kishi et al.(Fig. 11) and Salinosporamide A by Corey et al.(Fig12).

O

H

C

O

O

M

e

B

r

O

H

C

O

O

M

e

O

H

C

O

O

M

e

O

H

C

O

O

M

e

O

2

N

O

H

C

O

O

M

e

C

l

O

H

C

N

1.4 Dynamic kinetic Asymmetric transformation (DYKAT) of Morita-Baylis-Hillman adducts

O

H

C

O

O

M

e

Since the creation of highly optically enriched β-hydroxy-α-methylene esters (and ketones, nitriles etc.) still remains a barrier till to date, Trost and co-workers have introduced in 2000 an alternative strategy, the so called “dynamic kinetic asymmetric transformation” (DYKAT) of Morita-Baylis-Hillman adducts. This deracemization reaction procedure forms one out of the at least five enantiodiscrimination mechanisms in the famous palladium catalyzed asymmetric allylic alkylation (AAA). This strategy consists of a palladium catalyzed dynamic kinetic asymmetric transformation (DYKAT) of racemic Morita-Baylis-Hillman carbonates as depicted in figure 13. Dynamic kinetic asymmetric transformation refers to the conversion of chiral racemic substrates having the potential of being completely converted to one single enantiomer of the product. The DYKAT mechanism can be explained by the next illustrative example. In the first of step of a DYKAT reaction, the oxidative addition of a chiral palladium complex to the olefin occurs. Subsequently ionization of the leaving group happens and a diastereomeric complex is formed. Next, there are two possible options for the formation of a diastereomeric complex. Usually one diastereomeric complex is more favoured for nucleophilic attack (oxygen nucleophile), followed by decomplexation leading to the product. The second diastereomeric complex usually does not react or in much slower rates. In a DYKAT reaction inter conversion of this “slow reacting” diastereo-complex/π-allyl-intermediate to the “fast reacting” diastereo-complex π-allyl-intermediate happens and subsequently reacts. So the theoretically yield in a DYKAT reaction can be 100% compared to 50% yield in a kinetic asymmetric transformation (KAT). A KAT reaction, also known as a kinetic resolution, refers to the transformation of one single enantiomer of racemic starting material into the product. The remaining other enantiomer of a substrate does not react or does it in slower rates. Since the introduction, the DYKAT of MBH-adducts has been frequently used by Trost et al. to overcome difficult key steps in the total synthesis of natural and biological active compounds. A whole array of compounds has been synthesized via this methodology, examples being the total synthesis of (+)-Hippospongic acid A(Fig. 15), morphine, codeine38b (Fig. 14), and Furaquinocin’s A, B, and E (Fig. 16).

N

O

H

R

N

N

N

R

O

H

Q

u

i

n

i

n

e

(

Q

N

)

C

i

n

c

h

o

n

i

d

i

n

e

(

C

D

)

R

=

O

M

e

R

=

H

Q

u

i

n

i

d

i

n

e

(

Q

D

)

C

i

n

c

h

o

n

i

n

e

(

C

N

)

1

1

8

9

3

4

6

´

6

´

1

8

9

3

4

*

*

H

N

N

R

O

Q

u

i

n

i

c

i

n

e

C

i

n

c

h

o

n

i

c

i

n

e

R

=

O

M

e

R

=

H

B

S

B

P

B

P

S

N

O

H

N

O

R

L

e

w

i

s

B

a

s

e

B

r

ø

n

s

t

e

d

a

c

i

d

9

6

´

N

O

H

N

O

9

6

´

3

N

O

H

9

6

´

O

N

N

O

B

M

P

O

M

e

C

O

O

M

e

O

B

n

N

O

B

M

P

C

O

O

M

e

O

B

n

M

e

O

H

H

N

O

C

l

O

M

e

O

H

O

1.5Functionalization of Morita-Baylis-Hillman adducts via a tandem SN2’-SN2’ mechanism

N

N

R

1

R

2

1

8

9

3

6

´

4

B

n

C

P

D

R

1

=

O

H

P

H

N

C

P

D

R

1

=

O

H

C

6

´

T

h

Q

D

R

1

=

s

c

C

9

T

h

Q

D

R

1

=

H

R

2

=

B

n

R

2

=

P

h

n

R

2

=

B

n

R

2

=

s

c

s

c

=

R

1

H

N

H

N

S

C

F

3

C

F

3

C

O

2

-

O

O

O

H

H

+

H

N

O

O

O

H

O

H

H

H

O

O

O

T

B

S

O

O

O

T

E

S

O

H

H

H

N

H

A

l

l

o

c

O

C

O

2

t

B

u

M

s

O

As shown before, simple Morita-Baylis-Hillman adducts possess a huge synthetic potential25. As a result, in recent years a lot of papers have been published dealing with the (enantioselective) modification of the β-position, the alcohol functionality of MBH adducts. A number of strategies have been developed in the recent years focusing on two different targets. One is the expansion of synthetic methodologies in order to obtain various MBH adducts. The other is the development of a highly enantioselective route for obtaining MBH-alcohols, because the direct and highly enantioselective Morita-Baylis-Hillman reaction between a simple aldehyde and an activated olefin has proven to be a daunting challenge. In both processes, usually the MBH-alcohol is transformed into a leaving group. First the MBH reaction is performed without a chiral catalyst. Once the product is obtained, the MBH alcohol is transformed in typically an acetate or carbonate moiety. Now these MBH-adducts are prone to a tandem SN2’-SN2’ mechanism. This mechanism, which is comparable with the DYKAT transformations of MBH adducts, includes several steps. The first step is usually an attack of a nucleophilic (chiral) nitrogen or phosporus atom on the vinylic moiety of the MBH adduct occurs, with subsequent ionization of the leaving group. Once the leaving group has deprotonated the pro-nucleophile, the intermediate (zwitterion) is prone to nucleophilic attack, finally leading to the product. In the last five years a lot of papers appeared making use of this mechanism. The mechanism can be described as an allylic substitution reaction or, when carbon/nitrogen nucleophiles are used, an allylic alkylation or aminaton. In 2002, Kim and co-workers published an organocatalytic methodology which relies on this tandem SN2’-SN2’ mechanism (Fig. 17),. MBH-acetates (racemic) are transformed into MBH-alcohols with optical purities of 54-92% ee and in yields of 25-42%. The process involves a successive SN2’-SN2’ attack of a quinidine derived Sharpless asymmetric dihydroxylation ligand, (DHQD)2PHAL, to the vinyl moiety, followed by attack of a hydroxy anion, leading to the product. The first step of the process occurs via a kinetic resolution step, (DHQD)2PHAL reacts faster with one enantiomer of the starting material. Since the obtained enantioselectivities obtained can not be only explained by a kinetic resolution step, the author proposed that during the

C

6

H

1

4

C

O

O

R

C

6

H

1

4

O

C

O

O

H

O

H

O

H

H

2

N

O

H

O

O

second step there is additional asymmetric induction. Supplementary, Kim and co-workers reported in 2002 similar

N

O

H

N

O

reactions relying on again SN2’-SN2’ mechanism. MBH-acetates were reacted with a sub-

O

B

o

c

C

O

O

M

e

R

C

O

O

E

t

C

N

R

'

+

stoichiometric amount of DABCO and various pro-nucleophiles.

N

N

R

1

R

2

1

8

9

3

6

´

4

1

B

n

C

P

D

R

1

=

O

H

2

P

H

N

C

P

D

R

1

=

O

H

3

C

P

D

R

1

=

O

H

R

2

=

B

n

R

2

=

P

h

n

R

2

=

O

H

R

4

N

O

2

R

3

R

1

R

2

+

N

O

2

R

4

R

1

R

3

R

2

C

a

t

1

,

2

,

3

When pro-nucleophiles, such as tosylamide, methyl-sulfonamide, phthalimide, benzotriazole were used, the allylic substitution (amination) products were obtained in moderate yields (40-91%). Similar work was in that year reported by Orena and co-workers. In this elegant paper is shown the sensitivity of the MBH-adducts towards bases. When N-p-toluenesulfonimide derived MBH where reacted with a sub-stoichiometric amount of DABCO in DCM, the allylic substitution products (Fig. 19, compound a) were obtained via the tandem SN2’-SN2’ mechanism. On the other hand when DABCO was replaced by DBU, the SN2’-SN2’ mechanism does not occur. Now, a SN2’-decarboxylation mechanism occurs leading to product b, as depicted in figure 19. The author proposed that the results can be explained by the

E

W

G

O

H

R

R

B

a

k

e

r

´

s

Y

e

a

s

t

E

p

o

x

i

d

a

t

i

o

n

D

i

h

y

d

r

o

x

y

l

a

t

i

o

n

O

H

O

E

W

G

R

O

O

H

O

H

C

O

M

e

C

9

H

1

9

O

H

O

H

H

O

L

a

c

t

a

m

i

z

a

t

i

o

n

N

O

H

O

P

h

H

y

d

r

o

g

e

n

a

t

i

o

n

N

H

O

R

=

m

-

n

i

t

r

o

-

p

h

e

n

y

l

A

m

i

n

o

h

y

d

r

o

x

y

l

a

t

i

o

n

R

E

W

G

O

H

O

H

N

H

T

s

higher basicity (less nucleophilic) of DBU compared to DABCO. DABCO favours a nucleophilic SN2’-attack to the double bond instead of deprotonation of the N-p-toluenesulfonimide. MBH adduct. Basavaiah co-workers reported a similar methodology in order

R

E

W

G

O

M

o

c

+

N

u

H

R

E

W

G

R

E

W

G

o

r

N

u

N

u

P

d

*

to obtain enantiomerically enriched Morita-Baylis-Hillman adducts. Again the methodology relies on a allylic substitution tandem mechanism (Fig. 20). First the organocatalyst (quinidine) replaces the bromo atom, resulting in a chiral-intermediate, similar of kind intermediate as reported by Kim. Subsequently, the chiral intermediate is prone to attack by the pro-nucleophile, ultimately leading to the product. Unfortunately, the results obtained by Basavaiah are rather poor with yields up to 47% and enantioselectivities up to 40 % ee.

N

N

O

M

e

O

H

9

6

´

K

B

r

-

H

3

P

O

4

1

0

0

o

C

,

5

d

N

N

O

M

e

O

H

H

N

N

O

M

e

O

H

1

,

2

-

h

y

d

r

o

g

e

n

s

h

i

f

t

c

y

c

l

i

z

a

t

i

o

n

d

e

m

e

t

h

y

l

a

t

i

o

n

N

N

O

6

1

%

(

2

)

O

H

(

1

)

In 2004 the groups of Krische and Lu independently reported the first metal free organocatalytic allylic alkylation relying on the tandem SN2’-SN2’ mechanism in excellent yields and regioselectivity’s. The group of Krische reported the highly regioselective formation of γ-butenolides. Various γ-butenolides where created through a phosphine catalyzed substitution of MBH-acetates. The diastereo-selectivity ratios obtained in this reaction were excellent, up to 20:1. The regioselectivities were good, in general higher than 9:1. In addition, the reported yields were excellent, up to 94%. The group of Krische reported as well a phosphine catalyzed allylic amination and dynamic kinetic resolution and of MBH acetates. A protocol for the amination of several MBH-adducts by 4,5-dichlorophthalimide and phthalimide is presented. Besides racemic aminations, there is as well a chiral example shown, commercially available Cl-OMe-BIPHEP promotes the amination of MBH acetates with phthalimide in 56% ee. Lu and co-workers have described a β-IC catalyzed nucleophilic

O

O

O

E

t

O

O

B

e

t

a

-

I

C

T

o

l

u

e

n

e

,

r

t

,

1

2

h

O

O

E

t

N

u

+

O

O

E

t

N

u

substitution of tert-butylcarbonate by various carbon, anime, oxygen and phosphorus nucleophiles (Fig. 21). The tandem SN2-SN2’ nucleophilic substitutions occurs in high regioselectivities and excellent yields. However, in general the enantioselectivities obtained are moderate. This methodology shows a great tolerance towards various nucleophiles including; nitrogen, phosphorus, sulphur, oxygen and carbon. The proposed mechanism includes (Fig. 22), first the addition of the nucleophilic chiral amine to β-methylene functionality (vinyl moiety). After the organocatalyst is added to the α-methylene function, a cascade of steps occurs. First the activating group leaves the starting material, leading to CO2 and tert-butoxide. Sub sequentially, deprotonation of the pronucleophile by the tert-butoxide anion leads to a SN2’ attack of the pronucleophile on the chiral intermediate. During the attack of the pronucleophile, the organocatalyst is displaced from the intermediate giving the product and the regenerated catalyst. Lu and co-workers found that β-isocupreiidine was the catalyst leading to the highest enantioselectivities and reactivity’s. In the report there is one asymmetric example shown with a 1,3-di-carbonyl carbon pro-nucleophile producing the expected product in a moderate enantioselection (51% ee), high SN2-SN2’ ratio (chemoselectivity) and high yield.

N

N

O

N

H

N

H

S

F

3

C

C

F

3

6

´

N

N

O

N

H

N

H

S

F

3

C

C

F

3

6

´

N

N

O

O

S

O

O

C

F

3

6

´

N

N

O

O

H

6

´

N

N

O

N

P

h

P

h

6

´

N

N

O

N

H

2

D

M

A

P

1

0

m

o

l

%

,

D

C

M

,

r

e

f

l

u

x

6

h

P

h

-

N

-

(

O

T

f

)

2

,

1

.

1

e

q

.

P

d

(

O

A

c

)

2

,

1

0

m

o

l

%

H

N

=

C

P

h

2

,

1

.

1

e

q

.

(

+

/

-

)

B

I

N

A

P

,

1

0

m

o

l

%

C

e

s

i

u

m

c

a

r

b

o

n

a

t

e

1

.

5

e

q

.

,

T

H

F

,

r

e

f

l

u

x

1

8

h

T

H

F

/

H

2

O

c

i

t

r

i

c

a

c

i

d

3

,

5

-

C

F

3

P

h

N

C

S

,

1

.

0

5

e

q

.

T

H

F

,

2

0

m

i

n

.

(

2

)

(

3

)

(

4

)

(

5

)

(

6

)

4

7

%

7

6

%

8

1

%

6

9

%

1.6 All-carbon Quaternary stereocenters

Synthetic chemists nowadays can create almost every tertiary stereocenter with excellent levels of enantiocontrol and chemical yields. Various methodologies/series of tailor made ligands have been developed and are used commonly in organic synthesis. However, catalytic enantioselective C-C bond formation of all-carbon quaternary stereocenters, i.e. carbon stereocenters bearing four different carbon substituents, still represents a tremendous challenge for synthetic organic chemists. Moreover, when a carbon stereocenter is situated near a vicinal tertiary or quaternary stereocenter, the construction of these features become even more problematic. The difficulty for the construction of these motives arises often from steric hindrance and a limited amount of reliable reactions. Frequently used quaternary C-C bond formation reactions are; cycloadditions like Diels-Alder, Pd-allylations reactions and Michael additions, also known as conjugate additions. Nevertheless, the assembly of quaternary stereocenters still remains a considerably underdeveloped research area even though they are common motives in natural and pharmaceutical compounds. To overcome these difficulties, additional research is needed in the future to have access to complex molecular skeletons in an economic and straightforward manner. An alternative strategy in order to obtain quaternary stereocenters can be completed by organocatalysis1. As a matter of fact, in the last years a number of organocatalyzed formations of these assemblies have been reported. A variety of organocatalysts like proline and cinchona derivatives effectively catalyzes the formation of quaternary stereocenters. An outstanding example of these assemblies is the cinchona catalyzed Michael addition reported by Deng and co-workers in 200550c. The addition of various pro-chiral tri-substituted carbon nucleophiles to nitro-olefins was reported in high enantioselectivities (up to >99 % ee) and diastereoselectivities (up to >98:2 %) (Fig. 23). The organocatalysts of choice were simple cupreidine derived catalysts. To stress the challenge of simultaneous arrangements of vicinal quaternary and tertiary stereocenters; to date only six examples have been reported. Interestingly, five of them are organocatalytic.

O

B

o

c

+

(

9

)

b

-

I

C

,

2

0

m

o

l

%

1

e

q

1

.

2

e

q

T

o

l

u

e

e

n

0

.

0

5

M

1

2

h

,

r

t

*

*

C

O

O

E

t

P

h

C

O

O

M

e

C

O

O

M

e

(

1

1

)

C

N

(

1

0

)

P

h

*

C

O

O

E

t

C

N

C

N

(

1

0

)

C

O

O

E

t

Chapter 2 Allylic Nucleophilic Substitution of Morita-Baylis-

Hillman Carbonates

2.1 Goal

O

B

o

c

O

O

M

e

(

9

)

O

H

O

E

t

O

O

B

o

c

O

O

E

t

1

)

D

A

B

C

O

5

0

m

o

l

%

M

e

O

H

0

.

7

5

e

q

.

,

4

d

+

2

)

(

B

o

c

)

2

O

1

.

1

e

q

.

D

M

A

P

5

m

o

l

%

D

C

M

,

1

h

.

(

7

)

3

4

%

Enantioselective organocatalysis has revealed a extraordinary number of new methodologies in only half a decade1Although organocatalysis only recently found renewed interest, the hidden potential of organocatalysis does not require further explanation by a simple look a the literature of the past years. Since there is still a vast request for new complex chiral skeletons and many problems chemists have to deal with, are still unsolved, organocatalysis can serve as a real alternative in organic chemistry in academia and industry. The palladium catalyzed asymmetric allylic alkylation (AAA)38, mainly developed by Trost and co-workers, is one of the most prominent and frequently used asymmetric transition metal catalyzed reactions. As mentioned before in paragraph 1.4, the AAA has a wide scope and allows access to an incredible amount of asymmetric skeletons. The organocatalytic AAA, the dynamic kinetic asymmetric transformation (DYKAT) of Morita-Baylis-Hillman adducts, described by Lu et al.48, is the first procedure tolerating a variety of carbon, oxygen, nitrogen and phosporus nucleophiles in excellent regioselectivities. In fact, Lu and co-workers are the first who established an organocatalytic allylic alkylation with carbon nucleophiles, when dimethyl malonate was used a ee of 51% was obtained. This enantioselectivity is moderate, leaving room for improvement. During this master research project we wanted to contribute, and further explore the potent and exciting research area of organocatalysis. We became interested in the organocatalytic allylic nucleophilic substitution of MBH adducts for the reasons described before, and wanted to investigate the further potential of this reaction. Since Lu and co-workers have shown that 1,3-di-carbonyl entities easily can act as pro-nucleophiles, we were interested whereas via C-C bond forming process also quartenary stereocenters could be constructed. We wanted to expand the organocatalytic allylic substitution by using chiral trisubstituted carbon nucleophiles yielding in one step vicinal quaternary and tertiary stereocenters (paragraph 1.6) assembled MBH-adducts. In literature chiral trisubstituted carbon nucleophiles, like ethyl-phenyl-cyanoacetate, are used frequently to construct all-carbon or heteroatom quaternary stereocenters. MBH-adducts are, as discussed before in paragraph 1.3, valuable synthons in various syntheses making this procedure even more attractive. Moreover, we wanted to construct a β-isocupreidine catalyst with a C6’ thiourea motive. This thiourea motive, discussed in paragraph 1.2.1, is a much better hydrogen bond donor compared to a C6’ hydroxyl group and therefore it can lead to higher enantioselectivities in the organocatalytic allylic nucleophilic substitution with carbon nucleophiles.

2.2 Catalyst synthesis

O

B

o

c

+

(

7

)

b

-

I

C

,

2

0

m

o

l

%

O

M

e

O

O

M

e

O

1

e

q

1

.

2

e

q

T

o

l

u

e

e

n

0

.

0

5

M

1

2

h

,

r

t

O

M

e

O

O

M

e

O

C

O

O

E

t

C

O

O

E

t

(

8

)

R

1

E

W

G

L

G

R

1

*

E

W

G

N

u

R

1

E

W

G

R

1

E

W

G

X

R

3

P

d

[

P

d

]

X

R

3

N

u

L

G

(

X

=

P

,

N

)

P

d

-

c

a

t

a

l

y

z

e

d

a

l

l

y

l

i

c

s

u

b

s

t

i

t

u

t

i

o

n

O

r

g

a

n

o

c

a

t

a

l

y

t

i

c

a

l

l

y

l

i

c

s

u

b

s

t

i

t

u

t

i

o

n

N

N

O

(

2

a

)

O

M

e

N

O

R

N

O

P

h

O

B

o

c

C

O

O

M

e

P

h

*

C

O

O

M

e

2

a

,

b

-

I

Q

D

(

2

0

m

o

l

%

)

t

o

l

u

e

n

e

,

r

t

P

h

C

O

O

E

t

b

-

I

C

P

D

b

-

I

Q

D

C

N

1

a

R

=

H

R

=

M

e

<

5

%

c

o

n

v

a

f

t

e

r

1

4

d

P

h

O

B

o

c

C

N

P

h

*

C

N

2

a

,

b

-

I

C

P

D

(

2

0

m

o

l

%

)

t

o

l

u

e

n

e

,

-

2

0

º

C

,

4

8

h

P

h

C

O

O

E

t

C

N

1

4

9

5

%

y

i

e

l

d

,

<

2

%

e

e

6

'

3

a

1

5

To confirm the results described by Lu et al., we first started the synthesis of the catalyst β-isocupreidine. To date, the most convenient and high yielding synthesis of β-isocupreidine (2) was reported by Hoffman11 and Hatakeyama et al.17 β-Isocupreidine is formed in one step from quinidine (1) under highly acidic conditions. This process comprises a acid-induced cyclo-isomerization via a carbocation mechanism (Fig. 24). First, the vinyl moiety is protonated by HBr, leading to a secondary carbocation. This is followed by a 1,2-hydrogen shift, because a tertiary carbocation is more stable then a secondary. Next, attack of the C9 hydroxyl group on the tertiary carbocation takes place, leading to cyclization to an oxo-aza-twistane. In addition, in situ demethylation occurs, leading to a bi-functional catalyst. Similar work-up procedures were used as described by Hatakeyama and co-workers, although for an analytically pure batch of β-isocupreidine, the compound was subjected to re-crystallization and after purification, the batch was once again subjected to column chromatography. The obtained yield was similar to what is described by Hatakeyama et al. (61% yield). This reaction suffers unfortunately from side-product formation, a plausible option is likely to be cyclization via the secondary carbocation intermediate. Next, we started the synthesis of the β-isocupreidine catalyst, equipped with a strong Brøndsted acid thiourea-moiety (Fig. 26). We envisaged to synthesize the target structure, starting from β-isocupreidine via a similar strategy as reported by Hiemstra et al.15. The next envisioned steps included, in chronical order, a C6’OH triflation, a palladium catalyzed Buchwald amination of the triflate and subsequent hydrolysis. Once the synthesis of C6’-NH2 functionality was achieved, we envisioned to react the C6’NH2 with 3,5-bis-trifluorophenyl-isothiocyanate, yielding the target structure. We started our synthesis from pure β-isocupreidine and performed the triflation step of the C6’ hydroxyl group with N-phenyl-bis(trifluormethane sulfonimide) catalyzed by a

R

1

O

B

o

c

C

O

O

M

e

R

*

C

O

O

M

e

b

-

I

C

P

D

(

2

0

m

o

l

%

)

t

o

l

u

e

n

e

*

E

t

O

O

C

C

N

9

a

-

f

R

2

E

t

O

O

C

C

N

R

2

+

1

0

a

-

b

1

1

a

-

f

R

1

P

h

(

9

a

)

2

-

M

e

C

6

H

4

(

9

b

)

3

-

C

l

C

6

H

4

(

9

c

)

4

-

N

O

2

C

6

H

4

(

9

d

)

2

-

n

a

p

h

t

y

l

(

9

e

)

3

-

B

r

C

6

H

4

(

9

f

)

R

2

P

h

(

1

0

a

)

M

e

(

1

0

b

)

catalytic amount (10 mol%) of DMAP. The

(

R

)

C

O

O

R

b

-

I

C

P

D

(

2

0

m

o

l

%

)

t

o

l

u

e

n

e

,

7

2

h

,

-

2

0

º

C

(

R

)

P

h

C

O

O

E

t

9

f

C

N

+

1

0

a

1

1

f

(

2

e

q

u

i

v

.

)

(

1

e

q

u

i

v

.

)

(

S

)

O

B

o

c

C

O

O

M

e

9

f

+

B

r

B

r

8

0

%

e

e

9

0

%

e

e

reaction was stirred for 6h in DCM at reflux. Flash chromatography afforded (3) as a pure compound in 47% yield. Next, we subjected (3) to a palladium catalyzed Buchwald amination in THF. The reaction went smoothly overnight at reflux conditions and the crude reaction mixture was subjected to flash chromatography. Compound (4) was obtained as pure compound in 76 % yield. The C6´-NH2 was obtained via hydrolysis of the imine with citric acid. The reaction went without problems and compound (5) was obtained as a single product in 81% yield. The final step of the synthesis includes a nucleophilic addition of the free amine to the electrophile 3,5-bis-trifluoro-phenyl-isothiocyanate. The substrates were stirred in THF and were allowed to react for 20 min. During the reaction, the product precipitated from the reaction mixture as a white solid. Compound (6) was obtained as a single product in 69% yield. Meanwhile, the synthesis of the catalyst was reported by Deng et al. The results reported by Deng matched with our results and observations. In order to examine the bi-functional character of β-isocupreidine (2), which was shown to be fundamental in the direct Baylis-Hillman reaction, we have decided the synthesize β-isoquinidine (2a). This organocatalyst has a identical molecular skeleton as (2) but now the bi-functional character is removed from the catalyst. The synthesis of the catalyst was accomplished by using sodium hydride and methyl iodide.

Ph

(R)

Nu

COOMe

P

h

O

B

o

c

C

O

O

M

e

P

h

*

C

O

O

M

e

b

-

I

C

P

D

(

2

0

m

o

l

%

)

t

o

l

u

e

n

e

,

r

t

,

4

8

h

.

*

+

O

O

O

E

t

O

C

O

O

E

t

1

2

1

3

2.3 Preliminary experiments and optimization

H

C

O

O

M

e

B

r

+

O

H

Q

D

(

N

R

3

)

D

C

M

,

r

t

,

2

4

h

H

C

O

O

M

e

N

+

R

3

B

r

-

O

M

e

O

O

P

h

C

O

O

M

e

N

H

T

s

P

h

C

O

O

M

e

N

H

T

s

a

b

In order to check the results reported by Lu; we first wanted to start a reaction with dimethyl malonate and Morita-Baylis-Hillman carbonate (7). Therefore we needed to synthesize the MBH carbonate. Since the synthesis of the MBH carbonate was not reported by Lu, we had planned to synthesize the MBH alcohol via an adapted procedure of Aggarwal et al56. and subsequently the Boc protection via procedure Trost et al. (Fig. 28)

P

h

O

C

O

O

M

e

N

H

T

s

O

P

h

*

O

A

c

C

O

O

M

e

(

D

H

Q

D

)

2

P

H

A

L

T

H

F

/

H

2

O

P

h

C

O

O

M

e

N

R

3

N

a

H

C

O

3

P

h

O

A

c

C

O

O

M

e

P

h

O

H

C

O

O

M

e

The procedure of Aggarwal represents a simple neat MBH reaction between benzaldehyde and ethyl acrylate, catalyzed by 50 mol% DABCO. 0.75 eq. of methanol was added as an additive to the reaction mixture, because of the hydrogen bond donor capabilities to zwitterionic species 2 (paragraph 1.3.1) of methanol, resulting in rate enhancement. The MBH alcohol was obtained after flash-chromatography in a yield of 83% (6.4 g.). Next, the Boc protection was performed of the MBH alcohol with Boc-anhydride. Catalytic amount of DMAP (5 mol%) was used and the reaction was left stirring for 1h, the reaction mixture was purified by flash chromatography, yielding compound (7) as a colourless oil in 42% yield. The yield was somewhat disappointing, probably due to a short reaction time (concluded from other experiments) or by attack of DMAP on the product and consequent liberation of the Boc activating group. Subsequently, we repeated the experiment (Fig. 29) reported by Lu to confirm the level of enantioselection and yield. The reaction was performed at a 0.1 mmol scale of MBH carbonate and was left stirring overnight, a drop of the reaction mixture was filtered to remove the catalysts, the filtrate was analyzed by chiral HPLC. And indeed, the results reported by Lu et al. were reproducible, the level of enantioselection was 53% and the chemoselectivity was 97%. Next, we examined the role of the acrylate function in the carbonate. We made a MBH carbonate like (7) but

O

H

C

O

O

E

t

employed with a methyl ester instead of an ethyl ester electron withdrawing group. Carbonate (9) was synthesized in a similar way as (7), although higher in a higher yield, the compound (9) was obtained in a yield of 67% over two steps. The catalysis experiment was repeated with carbonate (9), and HPLC analysis revealed that the level of enantiodiscrimination was changed slightly; we observed an ee of 63% instead of 51% in case of (7). We found that the enantioselection process shows a concentration effect. We found an optimum at 0.05 M, a similar concentration as used by Lu and co-workers. A decrease or increase in concentration of the reaction mixture gave lower levels of asymmetric induction. Employment of diisopropyl malonate as pronucleophile gave slightly lower levels of asymmetric induction, 58% ee. Subsequently, we investigated the Brønsted acid functionality of the catalyst in the organocatalytic allylic alkylation. The thiourea derived β-isocupreidine catalyst (6) was tested in a catalysis experiment under similar conditions as β-isocupreidine itself. The reaction went to completion within 24 h, and had a lower level of asymmetric induction, 41% ee. Employment of other solvents (increased polarity) resulted in lower levels of asymmetric induction and disordered product formation. These results where in comparison as described by Lu, toluene remains the solvent of choice. We decided to lower the catalyst loading of (6) to 10 mol%, leading to a decreased polarity of the reaction medium. We where delighted to observe a significant higher ee of 58%. This result supports the idea that the enantioselection process of the organocatalytic allylic alkylation is favoured by an apolar reaction medium. Hence, we lowered the reaction temperature. The reaction with 10 mol% of (6) at -20 ºC gave almost no asymmetric induction, 14 % ee. Interestingly, β-isocupreidine (2), gave 72% ee., although the reaction rate was not practical and did not go to completion after almost a week. These findings may suggest a reversed temperature dependency of Brønsted acid activation. After we had confirmed the results reported by Lu et al. and examined the role of the thiourea Brønsted acid moiety, we started to investigate the

O

C

O

O

E

t

O

O

chiral compound ethyl-phenyl-cyanoacetate (10) as pro-nucleophile in the organocatalytic allylic alkylation. Employment of this chiral nucleophile should theoretically result in four different compounds, two diastereoisomers with their corresponding enantiomers. These four compounds complicate analysis by chiral HPLC, a prerequisite for enantiomeric excess determination being baseline separation. Therefore, we started first an experiment, by using 1 eq of carbonate (9) and 1 eq. of pro-nucleophile (10) catalyzed by the a-chiral catalyst DABCO (50 mol%). Fortunately, the reaction went to completion in less then 30 min. Removal of the catalyst by a short filtration over silica and analysis of the mixture by chiral HPLC revealed that the 4 different

C

O

O

E

t

M

e

O

O

M

e

O

O

compounds were separable. Next we started a similar catalysis experiment as in the case for dimethyl malonate, but we replaced dimethyl malonate by (10). To our delight the reaction went to completion in 18h. (Fig. 30) Analysis of the crude reaction mixture revealed an enantiomeric excess of the main diastereoisomer of 68%. For the

N

O

O

R

R

=

H

,

M

o

r

p

h

i

n

e

R

=

C

H

3

,

C

o

d

e

i

n

e

determination of the diastereomeric ratio we used 1H-NMR and we observed a ratio of 3:1. Lowering the reaction temperature to -20ºC, gave an improvement of both diastereo- and enantioselectivity, dr 4:1 and 84% ee. However the reaction rate was not practical roughly 30% conversion in 72h. Moreover, there was an other serious drawback; we could not separate the remaining amount of pro-nucleophile (10) from the product (11) by column chromatography. Consequently, we started two similar experiments at -20ºC; one experiment (a) with one equivalent carbonate (9) and 5 equivalents pro-nucleophile (10) and one experiment (b) with reversed ratios of the starting substrates, 5:1 respectively. We were delighted to observe a strong rate enhancement in both cases, experiment (b) went to completion after 22h. The levels of diastereomeric and enantiomeric control were similar for this experiment compared to earlier experiments. There was an important additional advantage observed in experiment (b), we could isolate the product after column chromatography. For further experiments we decided to make a compromise between reaction rate, atom economy and stereo chemical considerations of the organocatalytic allylic alkylation. We fixed the amount of substrate for further experiments, two equivalents of carbonate (9) and one equivalent of pro-nucleophile (10). During the optimization of the organocatalytic allylic alkylation we found that complete removal of Boc-anhydride is essential for obtaining high enantiomeric excesses. This is most likely explainable in terms of Boc protection of the catalyst, catalyst poisoning.

2.4 Results

2.4.1 Substrate scope

M

e

O

O

C

O

O

O

O

M

e

O

O

H

[

s

c

]

R

3

R

2

R

1

[

s

c

]

=

A

=

R

1

=

O

H

,

R

2

=

C

H

3

,

R

3

=

C

H

2

O

H

B

=

R

1

=

O

H

,

R

2

=

C

H

2

O

H

,

R

3

=

C

H

3

E

=

[

s

c

]

=

O

H

We decided to continue our examinations with the employment of different electrophile and pro-nucleophile reaction partners. Therefore we needed first to synthesize various Morita-Baylis-Hillman carbonates (as depicted in table one) via the procedures of Aggarwal and Trost. The synthesis of the MBH-substrates went without significant problems and generally in moderate to good yields (47-83%). The pro-nucleophiles were commercially avaiable. All the reactions were performed as discussed in the end of paragraph 2.3. namely; two equivalents of carbonate, one equivalent of pro-nucleophile, 20 mol% β-isocupreidine and at a concentration of 0.05 M. In general, what be can be concluded from table 1 is that compounds 11a-f were obtained in high yields and good enantioselectivities. A closer look at the table reveals the sensitivity of the pro-nucleophile with respect to the selected carbonate substrate, hindered substrates (10b and 10d, entries 2 and 4) required higher temperatures to obtain satisfying reaction rates; fortunately, the level of enantioselection was not negatively affected. A low enantiomeric excess and diastereomeric excess was observed for substrate 9d, possibly due to its high reactivity towards conjugate addition possibly, the electron withdrawing p-nitro group interferes with the hydroxyl functionality of the catalyst.

Entry

Product

T [ºC]

t [h]

Yield [%]b

drc

ee [%]d

1

COOMe

PhCOOEt

CN

11a

-20

48

94

4:1

83

2

COOMe

PhCOOEt

CN

11b

0

96

95

4:1

79

3

COOMe

MeCOOEt

CN

Cl

11c

-20

72

95

1.4:1

80

4

COOMe

PhCOOEt

CN

O

2

N

11d

-20

48

95

1.1:1

16

5

COOMe

PhCOOEt

CN

11e

20

24

66e

4:1

85

6

COOMe

COOEt

Br

Ph

CN

11f

-20

72

95f

3:1

80

a Conditions: 0.6 mmol 9, 0.3 mmol 10, 0.06 mmol β-ICPD in 6 mL toluene.

b Isolated yield (mixture of diastereomers).

c Determined by 1H-NMR analysis of the crude reaction mixture.

d ee of the major diastereoisomer determined by chiral HPLC.

e Compound 11e was obtained as a pure diasteromer.

f Reaction performed on 9.0 mmol of 9f

An important prerequisite of an asymmetric methodology is the possibility to upgrade the optical purity, if the levels of enantioselection are not excellent. Since some adducts were obtained as crystalline solids after chromatography, we run a reaction on a gram scale with the aim to upgrade the optical purity of the product via recrystallization. Hence, compound 9f (3.3 g, 9.0 mmol) was reacted with cyanoacetate 10a (0.78 mL, 4.5 mmol) to afford, after column chromatography, adduct 11f in 80% ee and 95% yield (entry 7). A single recrystallization from cyclohexane afforded compound 11f (59% yield, calculated on cyanoacetate 10a) with 98% ee as a pure diastereomer. In addition, a larger scale reaction between compound 9a (0.88 g, 6.0 mmol) and 10a (0.52 mL, 3.0 mmol) yielded adduct 11f after column chromatography in 86% ee and 95% yield. A single recrystallization from cyclohexane afforded compound 11f (63% yield, calculated on cyanoacetate 10a) with 99% ee as a pure diastereomer. Instead of employment of a prochiral trisubstituted cyanoacetate as pro-nucleophiles, we investigated the use of prochiral trisubstituted 1,3-dicarbonyl moieties. We decided to use ethyl 2-oxocyclohexanecarboxylate as reaction partner (Fig. 31). The mixture was allowed to react for 48 h at room temperature and the reaction was performed under similar conditions as used in table 1 (Fig. 31). Analysis of the crude reaction mixture revealed full conversion of the pronucleophile, yielding the

C

O

O

M

e

C

O

O

E

t

O

postulated product. We were glad to observe an excellent diastereomeric ratio of >98:2 however the enantiomeric excess was moderate: 61% ee. Purification by column chromatography afforded 13 as a single diastereoisomer. Nevertheless, we were not able to determine the yield since the resulting compound 13 was not stable. The compound was prone to decomposition in CDCl3. Employment of ethyl 2-methyl-3-oxobutanoate as pro-nucleophile gave immediately disappointing results and we were not able to isolate a pure compound.

C

6

H

1

4

C

O

O

R

C

6

H

1

4

O

C

O

O

H

O

H

O

H

H

2

N

O

H

O

O

2.4.2X-Ray analysis

O

H

R

E

W

G

+

R

*

O

H

E

W

G

B

a

s

e

M

e

O

O

C

O

O

H

C

O

O

E

t

N

N

O

N

H

S

N

H

F

3

C

C

F

3

N

N

O

N

H

2

Since we could crystallize two products in pure form, we submitted the crystals for X-ray analysis. From these crystals we could assign unambiguously, the absolute and relative configuration of compound 11f. (Fig. 32) Both stereocenters have the R configuration. From compound 11a it was only possible to assign the relative configuration. The absolute configuration of compound 11a could not be reliably determined because of the lack of a heavy atom. The relative configuration of this compound was analogous to compound 11f, (Fig. 34). During column chromatography of the reaction mixture of compound 11f, we could recover the excess of the starting material used. Interestingly, HPLC analysis of the recovered starting material (9f) gave evidence for a kinetic resolution effect. Compound 9f was isolated in 90% ee. Recrystallization afforded enantiopure crystals (47% yield based on 10a). X-Ray analysis revealed the absolute configuration, S (Fig. 35). From this result we can conclude that the catalyst has a preference for a MBH-carbonate bearing the R configuration. As suggested by Kim and coworkers,42 who observed a similar effect, this can be explained in terms of a kinetic resolution in the first step. This process, however, does not affect the enantioselectivity of the second conjugate addition generating the C-C bond. To confirm this statement we examined a catalytic experiment of 1 equivalent of 9a with 2 equivalents of 10a at room temperature (hence reversing the reagents ratio): after 24 h, 9a was completely converted to 11a with similar stereoselectivity (68% ee, 3:1 dr), showing that both enantiomers of the MBH carbonate undergo Michael addition although with different rates, followed by elimination of the leaving group leading to the same intermediate.

N

N

O

O

S

O

O

F

F

F

N

N

O

O

H

2.4.3 Mechanistic considerations

The importance of employment of bi-functional catalysts in organocatalysis, for obtaining high(er) enantioselectivities and reactivities, has been shown in numerous examples. As well in the asymmetric Morita-Baylis-Hillman reaction, the bi-functional character of β-isocupreidine was shown to be fundamental. We believe that the bi-functional character of the catalyst is also essential in the organocatalytic allylic alkylation of MBH carbonates. In order to confirm this assumption we started a catalytic experiment with β-isoquinidine (2a) (β-IQD) as the catalyst. Indeed, the role of the C6’-OH of β-ICPD is crucial for obtaining high enantioselectivities. We observed a drastically lower enantiomeric excess of 30%. Furthermore, less than 5% conversion was observed after two weeks at room temperature, suggesting that the C6’-OH of β-ICPD is also crucial for activation of the system. Additional evidence was acquired via substrate 14, featuring a cyano group, a poorer hydrogen bond acceptor, instead of a methyl ester. We started a similar experiment as in case of the other allylic substitution experiments (table 1) and MBH electrophile 14 underwent the desired reaction (Fig. 36). As result, product 15 was obtained in high yield but no asymmetric induction was observed (Scheme 4).

N

N

O

N

O

H

C

O

O

M

e

O

H

C

N

On the basis of all these observations, we propose a tentative hypothesis to rationalize our experimental results (Scheme 5). First, substrate 9a undergoes a conjugate addition to form adduct A (fig. 22, paragraph 1.5) . Subsequently, elimination of the OBoc group, leading to the formation of CO2 and tert-butoxide anion provides Michael acceptor (B) (fig. 22, paragraph 1.5). We believe that this irreversible reaction step is responsible for the observed kinetic resolution. As shown in the experiment with β-isoquinidine, the presence of the C6’-OH in the catalyst is required for activation of the ester prior to conjugate addition. As already proposed by Lu, the tert-butoxide anion deprotonates cyanoacetate 10a which, in turn, attacks the -unsaturated intermediate (β-ICPD and MBH carbonate). Probably, an intramolecular hydrogen bond between the phenolic OH and the ester moiety is responsible for the face shielding conferring stereoselectivity to the reaction; a DFT-minimized (Fig. 37) structure B (fig. 22) is supporting this hypothesis. Elimination of the catalyst from intermediate C liberates alkylated product 11a, closing the catalytic cycle.

O

H

C

O

O

M

e

C

l

O

H

C

O

O

M

e

O

2

N

O

H

C

O

O

M

e

B

r

O

H

C

O

O

M

e

O

H

C

O

O

M

e

O

C

O

O

M

e

O

O

O

C

O

O

M

e

O

O

2.5 Conclusions

In conclusion, we have developed new methodology that allows access to highly functionalized MBH adducts, featuring two vicinal stereocenters, one of them bearing four carbon substituents. These MBH-adducts can be created via a one step organocatalytic allylic alkylation of MBH carbonates and are obtained in excellent yields and chemo/enantio/ diastereo-selectivities. Moreover, this is the first, metal free, one-step creation of vicinal quaternary and tertiary stereocenters that is not completely explainable by a conjugate addition mechanism. The methodology could be extended to a broad scope of MBH aryl electrophiles, ranging from hindered to activated substrates. The role of the pro-nucleophiles was also investigated; compound 2a and 2b react with similar levels of asymmetric induction. However compound 2b was obtained in a lower diastereoselectivity, a plausible explanation being that steric-bulk is vital for obtaining high diastereoselectivities. Subsequently, we have shown for two examples that optically pure compound could be obtained, making this reaction practically executable. The crystals obtained for compounds 9f and 11f allowed assignment of the absolute configuration. For compound 11a only the diastereoselectivity could be determined. During our investigations we found evidence for the reaction mechanism. The reaction occurs most likely via two mechanisms, starting with a kinetic resolution followed by enantioselective attack of the chiral nucleophile. In addition, we showed that the carbonate that bears the R configuration, preferentially reacts with the catalyst. We also demonstrated the involvement of the OH-functionality of the catalyst in the catalytic cycle. The carbonyl moiety is vital to obtain highly enantiomerically enriched compounds. Future investigations could be aimed at broadening the scope of this transformation by employing alkyl instead of aryl electrophiles. Further design of catalysts that catalyze this re