biocatalysis & multicomponent reactions: the ideal …...biocatalysis & multicomponent...

182
Biocatalysis & Multicomponent Reactions: The Ideal Synergy Asymmetric Synthesis of Substituted Proline Derivatives Anass Znabet 2012

Upload: others

Post on 13-Jul-2020

27 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: Biocatalysis & Multicomponent Reactions: The Ideal …...Biocatalysis & Multicomponent Reactions Chapter 2 Monoamine Oxidase N: 41 A Promising Biocatalyst for Asymmetric Synthesis

Biocatalysis & Multicomponent Reactions: The Ideal Synergy

Asymmetric Synthesis of Substituted Proline Derivatives

Anass Znabet

2012

Page 2: Biocatalysis & Multicomponent Reactions: The Ideal …...Biocatalysis & Multicomponent Reactions Chapter 2 Monoamine Oxidase N: 41 A Promising Biocatalyst for Asymmetric Synthesis

This Research was supported by the Netherlands Organisation for Scientific Research (NWO)

under project number: 017.004.008

Printed by: Ridderprint BV, Ridderkerk, the Netherlands

Lay out: Simone Vinke, Ridderprint BV, Ridderkerk, the Netherlands

Cover Design: Nikki Vermeulen, Ridderprint BV, Ridderkerk, the Netherlands

ISBN: 978-90-5335-497-1

Page 3: Biocatalysis & Multicomponent Reactions: The Ideal …...Biocatalysis & Multicomponent Reactions Chapter 2 Monoamine Oxidase N: 41 A Promising Biocatalyst for Asymmetric Synthesis

VRIJE UNIVERSITEIT

Biocatalysis & Multicomponent Reactions: The Ideal Synergy

Asymmetric Synthesis of Substituted Proline Derivatives

ACADEMISCH PROEFSCHRIFT

ter verkrijging van de graad Doctor aan

de Vrije Universiteit Amsterdam,

op gezag van de rector magnificus

prof.dr. L.M. Bouter,

in het openbaar te verdedigen

ten overstaan van de promotiecommissie

van de faculteit der Exacte Wetenschappen

op donderdag 26 januari 2012 om 13.45 uur

in de aula van de universiteit,

De Boelelaan 1105

door

Anass Znabet

geboren te Amsterdam

Page 4: Biocatalysis & Multicomponent Reactions: The Ideal …...Biocatalysis & Multicomponent Reactions Chapter 2 Monoamine Oxidase N: 41 A Promising Biocatalyst for Asymmetric Synthesis

promotoren: prof.dr. ir. R.V.A. Orru

prof.dr. M.B. Groen

copromotor: dr. E. Ruijter

Page 5: Biocatalysis & Multicomponent Reactions: The Ideal …...Biocatalysis & Multicomponent Reactions Chapter 2 Monoamine Oxidase N: 41 A Promising Biocatalyst for Asymmetric Synthesis

إ وادي

Page 6: Biocatalysis & Multicomponent Reactions: The Ideal …...Biocatalysis & Multicomponent Reactions Chapter 2 Monoamine Oxidase N: 41 A Promising Biocatalyst for Asymmetric Synthesis
Page 7: Biocatalysis & Multicomponent Reactions: The Ideal …...Biocatalysis & Multicomponent Reactions Chapter 2 Monoamine Oxidase N: 41 A Promising Biocatalyst for Asymmetric Synthesis

Table of Contents

Chapter 1 General Introduction: 9

Biocatalysis & Multicomponent Reactions

Chapter 2 Monoamine Oxidase N: 41

A Promising Biocatalyst for Asymmetric Synthesis

Chapter 3 Highly Stereoselective Synthesis of Substituted Prolyl Peptides 53

Using a Combination of Biocatalytic Desymmetrization and

Multicomponent Reactions

Chapter 4 Asymmetric Synthesis of Synthetic Alkaloids by a Tandem 77

Biocatalysis/Ugi/Pictet–Spengler-Type Cyclization Sequence

Chapter 5 A Highly Efficient Synthesis of Telaprevir® by Strategic use of 99

Biocatalysis and Multicomponent Reactions

Chapter 6 Stereoselective Synthesis of Substituted N-Aryl Proline Amides 125

by Biotransformation/Ugi-Smiles Sequence

Chapter 7 Reflections & Outlook 145

Summary 159

Samenvatting (Summary in Dutch) 165

Dankwoord 173

List of Publications/Patents 179

Page 8: Biocatalysis & Multicomponent Reactions: The Ideal …...Biocatalysis & Multicomponent Reactions Chapter 2 Monoamine Oxidase N: 41 A Promising Biocatalyst for Asymmetric Synthesis
Page 9: Biocatalysis & Multicomponent Reactions: The Ideal …...Biocatalysis & Multicomponent Reactions Chapter 2 Monoamine Oxidase N: 41 A Promising Biocatalyst for Asymmetric Synthesis

General Introduction:Biocatalysis & Multicomponent Reactions

Chapter 1

Page 10: Biocatalysis & Multicomponent Reactions: The Ideal …...Biocatalysis & Multicomponent Reactions Chapter 2 Monoamine Oxidase N: 41 A Promising Biocatalyst for Asymmetric Synthesis
Page 11: Biocatalysis & Multicomponent Reactions: The Ideal …...Biocatalysis & Multicomponent Reactions Chapter 2 Monoamine Oxidase N: 41 A Promising Biocatalyst for Asymmetric Synthesis

General Introduction

11

1.1 Introduction

The chemical and pharmaceutical industry provides us with a myriad of useful products

without which our standard of living would not be what it is now. However, the industry is

also one of the major contributors to environmental pollution, due to the use of hazardous

chemicals and in particular large amounts of flammable, volatile and often toxic organic

solvents and reagents. For the production of fine chemicals, the waste/product ratio ranges

between 5 and 50, while for pharmaceuticals this ratio may even be as high as 100.[1] The

problems posed by this, including the inefficient use of resources, energy and capital,

together with the risk to welfare and the environment are widely recognized throughout

society.

Although we have reaped many benefits from our fossil fuel-based economies, man faces an

urgent environmental crisis.

In recent decades, a growing consensus has risen about the negative influences of the

increase of various gases on the global climate, such as CO2 and CH4. For example, since

the start of the industrial revolution in the 18th century, the CO2-concentration in the air

has increased from roughly 100 ppm to more or less 400 ppm.[2-3] These gases, also called

greenhouse gases, share a common feature that they tend to absorb heat and keep earth’s

atmosphere at a comfortable average temperature of 15 °C. Without these greenhouse

gases, the earth would lose too much heat to space and would be too cold to be habitable.

But the increasing amount of these gases in the atmosphere will isolate the earth too much,

resulting in elevated temperatures and the melting of vast amounts of ice on both poles

and various high mountain ranges. Beside ecological destruction of these areas, the oceans

will also rise and flood low-lying areas around the globe. Since many large cities, such as

Amsterdam, New York City, harbors, such as Rotterdam, Singapore, and historical treasures,

such as Venice, are situated at sea level, these will be lost if the sea level rises substantial due

to melt water.

These issues were emphasized when Al Gore’s documentary film “An inconvenient truth” was

aired drawing the attention of politicians as well as that of the general public, which has put

global warming and environmental issues high on the political and socio-economic agenda.

In order to fight environmental decay, rising sea levels and increasing toxic waste piles

development of new technologies for the production of energy, chemicals and products

is vital.

Among others, synthetic chemists are challenged to find solutions that maintain our

standard of living but spare earth’s resources. The focus is set on developing novel, clean,

atom-and step-efficient procedures for sustainable production for valuable fine chemicals

and pharmaceuticals. The “ideal synthesis” should lead to the desired product from readily

available starting materials in one or two reaction steps, in good overall yield and using

environmentally benign reagents.[4] This minimizes energy consumption and waste

Page 12: Biocatalysis & Multicomponent Reactions: The Ideal …...Biocatalysis & Multicomponent Reactions Chapter 2 Monoamine Oxidase N: 41 A Promising Biocatalyst for Asymmetric Synthesis

Chapter 1

12

production. A powerful strategy would be combining two methodologies which have

proven to be efficient and environmentally benign: (i) biocatalysis and (ii) multicomponent

reaction (MCR) methodology.

1.2 Biocatalysis

1.2.1 Enzymes as Catalysts

In chemistry, a catalyst is a substance that decreases the activation energy of a chemical

reaction without itself being changed at the end of the reaction. Catalysts participate in

reactions but are neither reactants nor products of the reaction they catalyze (a strange

‘exception’ is the process of autocatalysis). They work by providing an alternative pathway

for the reaction to occur, thus reducing the activation energy and increasing the reaction

rate (Figure 1).

Figure 1: Generic graph showing the effect of a catalyst in a hypothetical exothermic chemical reaction.

The catalyzed pathway, despite having a lower activation energy, produces the same final result.

In biocatalysis, natural catalysts, mostly enzymes, are used to perform chemical transformation

of organic compounds. Biocatalysis is one of the oldest chemical transformations known to

man; 6000 years ago it was already used for e.g. brewing beverages or cheese making. A

brief historical background is depicted in Table 1.

Page 13: Biocatalysis & Multicomponent Reactions: The Ideal …...Biocatalysis & Multicomponent Reactions Chapter 2 Monoamine Oxidase N: 41 A Promising Biocatalyst for Asymmetric Synthesis

General Introduction

13

Table 1: Brief history of enzyme engineering and their application.

Year Milestones Discoverer

6000 B. C.Chymosin from the stomach of cattle employed for the

production of cheese

1783 Hydrolysis of meat by gastric juice (digestion) demonstrated Spallazani

1846 Invertase activity observed Dubonfout

1893 Definition of a catalyst including enzymes is postulated Ostwald

1894Discovery of enzyme stereospecificity. “Lock-and-key”

model was proposedE. Fischer[5]

1897Cell free extract form yeast was employed for the

conversion of glucose to ethanolBüchner[6]

1908 Application of pancreatic enzymes in the leather industry Röhm

1913-1915Application of pancreatic enzymes to clean laundry.

Commercialized as “Burnus”Röhm

1926 Enzymes are proven to be proteins Sumner[7]

1953The first amino acid sequence of a protein (Insulin)

established, proving the chemical identity of proteinsSanger[8]

1965 “Allosteric model” of enzyme was proposed Monod[9]

After 1980Protein engineering developed for the improvement of

enzyme production and propertiesMany

Since the pioneering work of Büchner[6] (Table 1), it has been demonstrated that enzymes

do not require the environment of a living cell to perform catalysis. From those findings, the

use of enzymes has been increasing in importance and has been employed by the industry

in several applications in food technology, for example in bread, beer, wine, cheese, yoghurt.

Last but not least, also in the production of washing powder, textiles and paper.

Biocatalysis can offer several advantages over conventional chemical catalysis since enzymes

show great chemoselectivity, regioselectivity and enantioselectivity. Furthermore, enzymes

show higher substrate selectivity, milder reaction conditions, lower energy requirements

and fewer side reactions, such as isomerization, racemization and rearrangement than

conventional chemical catalysis. In addition, biocatalytic processes often provide products

of high purity in one reaction step. As a result, biocatalysts can be used in both simple and

complex transformations without the need for the tedious protecting and deprotecting

steps that are common in enantio- and regioselective organic synthesis.

Although biocatalysis offers a great number of advantages, there are some drawbacks that

should be taken into consideration. In particular, the high cost of enzyme isolation and

purification still discourages their use, especially in areas which currently have an established

alternative procedure. The reaction rates for substrates that are not natural to an enzyme are

usually slower than those observed for native substrates or conventional chemical catalysis.

Page 14: Biocatalysis & Multicomponent Reactions: The Ideal …...Biocatalysis & Multicomponent Reactions Chapter 2 Monoamine Oxidase N: 41 A Promising Biocatalyst for Asymmetric Synthesis

Chapter 1

14

The generally unstable nature of enzymes, when removed from their natural environment,

is also a major drawback to their more extensive use. Thus, high concentrations of reagents

can negatively affect enzyme stability and enzymes are often unstable in organic or biphasic

organic/aqueous media and can easily be inactivated. Another drawback is that some

enzymes require cofactors in order to function, such as ATP or NAD(P)H (Figure 2).

Figure 2: Cofactors may be required for biocatalysis.

These cofactors are expensive and often require recycling, since stoichiometric amounts of

ATP or NAD(P)H are extremely expensive. To circumvent these issues, often another set of

enzymes are included that are used to regenerate these cofactors. In the case of NAD(P)H,

these cofactor regeneration enzymes often use a cheap source of hydrogen to replenish the

NAD(P)H, such as ethanol or isopropanol; or use molecular oxygen as an oxidant.

Different strategies exist to obtain enantiomerically pure compounds. As such, resolution

procedures play an important role. A very old resolution procedure which is widely used

is the kinetic resolution (KR). Kinetic resolution is defined as “The achievement of partial or

complete resolution by virtue of unequal rates of reaction of the enantiomers in a racemate with

a chiral agent (reagent, catalyst, solvent, etc.).” This means that two enantiomers (in a racemic

mixture) show different reaction rates in a chemical reaction thereby creating an excess of

the less reactive enantiomer. The maximum yield in kinetic resolutions is 50% (Scheme 1).

Scheme 1: A kinetic resolution (KR).

Page 15: Biocatalysis & Multicomponent Reactions: The Ideal …...Biocatalysis & Multicomponent Reactions Chapter 2 Monoamine Oxidase N: 41 A Promising Biocatalyst for Asymmetric Synthesis

General Introduction

15

The obvious drawback of a kinetic resolution is that the maximum achievable yield is only

50% and the product has to be separated from the reactants. This has led to an important

development called the dynamic kinetic resolution (DKR). Because the reactants form a

chemical equilibrium and exchange, a theoretical conversion of 100% can be achieved. The

enantioselective reaction is much slower than the racemization of the racemic substrate.

This eventually leads to replenishment of the faster reacting enantiomer at the expense of

the slower reacting one (Scheme 2).

Scheme 2: A dynamic kinetic resolution (DKR).

Finally, enzymes possess the ability to differentiate between enantiotopic groups of prochiral

and meso compounds (Scheme 3). In contrast to a kinetic resolution, the theoretical yield

of these conversions is 100%. The enzymatic desymmetrization of meso compounds has

gained popularity in recent years and constitutes an elegant approach for the synthesis of

enantiomerically pure compounds.

Scheme 3: An enzymatic desymmetrization (ED).

Although biocatalysis offers a great number of advantages, there are some objections that

cannot be ignored. Consequently, these objections are currently being addressed and

overcome by protein engineering.

1.2.2 Protein Engineering

The “ideal enzyme” should have the desired activity, stability, specificity and so on in

each individual case, i.e. should meet the demands of any user. One way to achieve this

objective is via protein engineering.[10-16] This approach increases the basic understanding

of how enzymes function and have evolved, and it is the key method of improving enzyme

properties for applications in pharmaceuticals, green chemistry and biofuels. Protein

Page 16: Biocatalysis & Multicomponent Reactions: The Ideal …...Biocatalysis & Multicomponent Reactions Chapter 2 Monoamine Oxidase N: 41 A Promising Biocatalyst for Asymmetric Synthesis

Chapter 1

16

engineering has already proven itself and has been employed industrially with great

success.[17-20] Two main strategies are employed to adjust enzyme properties towards the

desired application: rational design and directed (molecular) evolution.[21]

Rational design[22-24] uses detailed knowledge of the structure and function for the prediction

of changes in the protein structure in order to alter or induce the desired properties. This

generally has the advantage of being economically favorable and technically easy, since

site-directed mutagenesis techniques are well-developed.[25-27] The advances in computing

have helped in creating better protein models to improve predictions for rational design.

However, its major drawback is that detailed structural knowledge of a protein is often

unavailable and structure-activity relationships are still not trivial.

In directed evolution,[28-30] mutant libraries are created by random changes, and a selection

regime is used to pick out variants that have the desired properties. The variants (showing

improved properties) are then subjected to further rounds of evolution (mutation) until the

mutant meets the desired necessities. This selection procedure mimics natural evolution.

In contrast to rational design, prerequisite knowledge of the enzyme structure and how

the structure determines function is not required when performing directed evolution.

Not surprisingly, desired changes are often caused by mutations that were not anticipated.

However, major concerns might arise when the libraries of mutants are not simple enough

to screen for desired function and complex enough to contain rare, beneficial mutations

and also if there is no rapid and cost-effective screen or selection that reflects the desired

function.

Nowadays, more researchers realize that rational design and directed evolution go hand

in hand and both are applied in engineering new biocatalysts. Information available from

related protein structures, families and mutants already identified is combined and then

used for targeted randomization of certain areas of the protein. Table 2 provides a brief

overview of enzymes engineered by rational design and/or directed evolution that are

important for organic synthesis.

1.2.2 Optically Active Synthons generated by Biocatalysis

As mentioned previously in this chapter, enzymes have been employed in numerous

applications. Among others, enzymes have been employed in organic synthesis and

especially in the synthesis of enantiopure compounds. The synthesis of chiral amines,

alcohols, aldehydes/ketones and carboxylic acids is very important as these compounds

are frequently used as synthons for various pharmaceutically active substances and

agrochemicals. The selected examples given below will show some biocatalytic methods to

generate these highly valued synthons.

Page 17: Biocatalysis & Multicomponent Reactions: The Ideal …...Biocatalysis & Multicomponent Reactions Chapter 2 Monoamine Oxidase N: 41 A Promising Biocatalyst for Asymmetric Synthesis

General Introduction

17

Table 2: Brief overview of engineered enzymes. An extensive survey can be found in reference.[41]

Enzyme Property Method Milestone Ref

Dioxygenase Thermostability Rational DesignTmax for enzymatic activity

increase by 15 ˚C[31]

Lipase Thermostability Rational DesignResidual activity after 6 h

increased by 50%[32]

Endoglucanase General stability Rational DesignImproved stability at pH

3-4[33]

Monooxygenase Substrate specificity Rational DesignReaction rate for

2-phenylethanol increased 190-fold

[34]

Galactose Oxidase Substrate specificityDirected Evolution

Oxidation of secondary alcohols

[35]

Alcohol dehydrogenase

Cofactor specificity Rational DesignImproved preference for

NAD(P)H over NAD(H)[36]

Lipase EnantioselectivityDirected Evolution

E-value of 111 for the kinetic resolution of e.g. an

axially chiral allene[37]

Phenylalanine ammonium lyase

Increased activityDirected Evolution

Reaction rate increased 15-fold

[38]

O-Glycosyltransferase Promiscuous activity Rational DesignPreferences switched

from O-Glycosylation to N- Glycosylation

[39]

β-Lactamase MiscellaneousDirected Evolution

Allosteric regulation by metal ions

[40]

1.2.2.1 Optically Active Amines

In principle, various enzymatic routes can be employed to generate optically active

amines using several different enzymes like hydrolases, oxidoreductases and transferases.

Hydrolases are the most frequently used enzymes for the production of optically active

amines by kinetic resolution of racemic starting material, for example, N-acyl amides by

peptidases, amidases or lipases.[42] For instance, Burkholderia plantarii lipase has been used

to generate optically active amines 4 via kinetic resolution (Scheme 1).[43] Most important is

the use of highly activated methoxyacetate 2 as an acyl donor since this donor increases the

carbonyl activity by a factor >100 compared to analogous reactions.

Scheme 4: Kinetic resolution of racemic amines with Burkholderia plantarii lipase.

Page 18: Biocatalysis & Multicomponent Reactions: The Ideal …...Biocatalysis & Multicomponent Reactions Chapter 2 Monoamine Oxidase N: 41 A Promising Biocatalyst for Asymmetric Synthesis

Chapter 1

18

Although this process is very efficient, the yield of a kinetic resolution is limited to a

maximum of 50% unless a racemization step is incorporated to improve the yields. However,

racemization of amines has proven to be fairly complicated and harsh conditions are

typically required.[44]

Alternatively, transaminases can be used to synthesize chiral amines. Transaminases transfer

an amino group from a donor substrate to an acceptor compound. However, it is essential to

add stoichiometric quantities of an amine acceptor (e.g. pyruvate) which in addition to the

cost can also result in inhibition of the transaminase. Turner and coworkers have designed

an indigenous system to circumvent this problem by introducing a catalytic amount of

amine acceptor (pyruvate; 8) together with an amino acid oxidase (Scheme 5) generating

chiral amines 7 in very selective fashion.[45]

Scheme 5: Kinetic resolution of a racemic amine using a transaminase in combination with a catalytic amount of pyruvate and amino acid oxidase.

Transaminases can also be used in an asymmetric fashion (theoretical yield of 100%)

generating chiral amines from ketones using a suitable amine donor (e.g. L-alanine). However,

there is a significant disadvantage to these type of reactions; the unfavorable conversion of

a ketone to an amine. The equilibrium of the reactions is mostly lies on the ketone substrate.

Bornscheuer and coworkers developed a highly efficient methodology (Scheme 6) for

the asymmetric synthesis of optically pure amines 11 by combining a transaminase and a

pyruvate decarboxylase shifting the equilibrium to the amine.[46]

In principle, far more biocatalytic methods exist to synthesis optically active amines which

were not discussed in the preceding part, including, asymmetric synthesis with amine

hydrogenase,[47] asymmetric reduction of aryl amines[48] and monoamine oxidase (discussed

in chapter 2).

Page 19: Biocatalysis & Multicomponent Reactions: The Ideal …...Biocatalysis & Multicomponent Reactions Chapter 2 Monoamine Oxidase N: 41 A Promising Biocatalyst for Asymmetric Synthesis

General Introduction

19

Scheme 6: Asymmetric synthesis of a ketone using transaminase in combination with alanine and pyruvate-decarboxylase.

1.2.2.2 Optically Active Alcohols

Optically pure alcohols are widely used in the agrochemical and pharmaceutical industry,

flavors and fragrances, food applications and in material sciences, such as liquid crystals.

Despite the fact that there are several methods to generate these optically pure alcohols

chemically mostly by ruthenium catalysis,[49-52] biocatalysis offers a very efficient and green

approach to obtain these synthons.[53] Racemic alcohols are attractive starting compounds,

since they are often more easily accessible than the corresponding ketones, although

ketones are frequently reduced in an asymmetric fashion. An example of both strategies

will be discussed.

Kroutil and coworkers have demonstrated an efficient deracemization of racemic alcohols

15 by the stereoinversion of one enantiomer in a tandem reaction sequence making use

of an oxidase from Alcaligenes faecalis and an alcohol dehydrogenase (ADH) in which only

glucose and molecular oxygen are required as co-factors (Scheme 4).[54]

Scheme 7: Tandem oxidation and reduction for the deracemization of secondary alcohols.

Page 20: Biocatalysis & Multicomponent Reactions: The Ideal …...Biocatalysis & Multicomponent Reactions Chapter 2 Monoamine Oxidase N: 41 A Promising Biocatalyst for Asymmetric Synthesis

Chapter 1

20

ADH from Rhodococcus ruber DSM 44541 was also demonstrated to be a highly effective

reducing agent of ketones (19). High concentrations of isopropanol (up to 50% v/v) using

whole cells not only shift the equilibrium toward the product side, but also enhance the

solubility of lipophilic substrates in the monophasic aqueous/organic medium. As a result,

this proved to be a highly regio- and enantioselective method giving the desired products

(20) in excellent yield (Scheme 8).[55]

Scheme 8. Biocatalytic oxidation/reduction of alcohols/ketones by ADH andd cosubstrate (20% (v/v) acetone/50% (v/v) isopropanol).

1.2.2.2 Optically Active Aldehydes/Ketones

Needless to say, aldehydes and ketones present interesting classes of compounds in the

synthesis of complex products, especially when these compounds are used as optically pure

inputs. An appealing synthon in this case is an α-hydroxy ketone (acyloin). Its bifunctional

nature and the presence of chiral information make it a highly valuable building block for

further synthetic transformations. Because of its significance, chemical approaches aiming

to prepare them have been reported.[56-57]

Several representatives of thiamine diphosphate (ThDP)-dependent enzymes have proven

highly efficient biocatalysts in the synthesis of acyloins, i.e. benzaldehyde lyase from

Pseudomonas fluorescens (BAL) and benzoylformate decarboxylase from Pseudomonas

putida (BFD). For example, Trauthwein and coworkers developed an efficient biphasic whole

cell system to generate acyloins 23 in high yield and enantiomeric excess (Scheme 9). No

external cofactors (ThDP and Mg2+) were necessary in combination with these whole cells.

MTBE proved to be the best organic solvent for the whole cells but also for the isolated

enzymes.[58]

Page 21: Biocatalysis & Multicomponent Reactions: The Ideal …...Biocatalysis & Multicomponent Reactions Chapter 2 Monoamine Oxidase N: 41 A Promising Biocatalyst for Asymmetric Synthesis

General Introduction

21

Scheme 9: Acyloin formation using whole cells expressing BAL and BFD with a range of aldehydes in a biphasic system.

α-Substituted chiral aldehydes are widely used as versatile building blocks for the asymmetric

synthesis of complex natural products and drug like molecules, therefore a wide range of

methodologies has been developed for their asymmetric synthesis.[59-63] Unfortunately,

few methods relating to biocatalysis have been developed. A rare example was reported

by Smonou and coworkers where a kinetic resolution by a lipase of corresponding acylal

24 resulted in the corresponding optically active α-substituted aldehyde 26 in modest

conversion and enantiomeric excess (Scheme 10).[64]

Scheme 10: Synthesis of chiral aldehydes through lipase-catalyzed resolution of acylal.

1.2.2.3 Optically Active Carboxylic Acids

Chiral carboxylic acids can be derived from chiral aldehydes and alcohols (see previous

sections) by classical oxidation chemistry. Nevertheless, they can also be formed from

ester hydrolysis as demonstrated by the group of Kazlauskas (Scheme 11).[65] In this kinetic

resolution the crude lipase from Candida rugosa (CRL) was first dissolved in 50% isopropanol

followed by dialysis to remove the isopropanol. This treatment enhanced the activity and

enantioselectivity dramatically. This led to 2-substituted carboxylic acids 29 in moderate

yield and excellent enantiomeric excess.

Scheme 11: Kinetic resolution of 2-substituted carboxylic acids esters by Candida rugosa.

Page 22: Biocatalysis & Multicomponent Reactions: The Ideal …...Biocatalysis & Multicomponent Reactions Chapter 2 Monoamine Oxidase N: 41 A Promising Biocatalyst for Asymmetric Synthesis

Chapter 1

22

Nitrilases are also capable of generating carboxylic acids by hydrolyzing the corresponding

nitriles. This is illustrated by the group of Burk who performed a dynamic kinetic resolution

(DKR) of aromatic α-hydroxy nitriles 32 and 33 (Scheme 12). A wide range of (R)-mandelic

acid derivatives (35) and analogues were generated by nitrilase B and a series of (S)-

phenyllactic acid (34) and other aryllactic acid derivatives were generated by nitrilase A in

high yield and enantiomeric excess.[66]

Scheme 12: Nitrilase catalyzed DKR of α-hydroxy nitriles in production of valuable chiral hydroxy carboxylic acid derivatives.

1.2.2.4 Optically Active Nitriles

Cyanohydrins have proven to be versatile building blocks in organic synthesis.[67-68] Therefore,

the synthesis of chiral cyanohydrins 36 has attracted considerable attention (scheme 13).[69-

71] Nowadays, the majority of cyanohydrin syntheses is performed by hydroxynitrile lyases

(HNLs) owing to their broad applicability, availability, excellent yield and enantiomeric excess

(Scheme 13).[69-70, 72, 73] The HNLs from Prunus amygdalus (PaHNL), Sorghum bicolor (SbHNL),

Manihot esculenta (MeHNL) and Linum usitatissimum (LuHNL) are currently the four major

HNLs that are used in the synthesis of chiral cyanohydrins with ee’s up to >99%.[72] When

starting with aldehydes, the equilibrium usually lies to the product side. Nevertheless, when

ketones are employed, yields are lower due to the considerable steric bulk in the product,

a tertiary alcohol with three other bulky substituents. This can be circumvented by using a

large excess of HCN. Furthermore, chemical synthesis of these cyanohydrins shows lower

yield and enantioselectivity compared to synthesis via HNLs.[70]

Page 23: Biocatalysis & Multicomponent Reactions: The Ideal …...Biocatalysis & Multicomponent Reactions Chapter 2 Monoamine Oxidase N: 41 A Promising Biocatalyst for Asymmetric Synthesis

General Introduction

23

Scheme 13: HNL-catalyzed synthesis of cyanohydrins and possible applications.

1.3 Multicomponent Reactions

The field of organic chemistry has reached a high level of sophistication in the past few

decades; the syntheses of complex natural products have become more and more efficient

and the stereoselectivity has significantly improved. These improvements however, have

not yet led to the ideal synthesis (Figure 3).[4, 74] This proposed ideal synthesis consists of

different factors; both economical and environmental parameters are considered. An

important factor, both economically and environmentally, is one-pot synthesis.

Figure 3: Ideal synthesis described by Wender and coworkers.[4]

Page 24: Biocatalysis & Multicomponent Reactions: The Ideal …...Biocatalysis & Multicomponent Reactions Chapter 2 Monoamine Oxidase N: 41 A Promising Biocatalyst for Asymmetric Synthesis

Chapter 1

24

Multicomponent reactions (MCRs) combine essentially all the atoms of at least three

components to form a single product in one pot for atom and process efficiency with

reduced waste. MCRs are resource effective and environmentally acceptable and thus

greener compared to classical multistep synthesis (Figure 4). Therefore, they contribute to

different aspects of the ideal synthesis. MCRs are not only resource effective, but also more

environmentally friendly than a linear synthesis. Also, because of the versatility and the

simple experimental set-up of MCRs they have proven to be convenient tools for easy access

to large collections of structurally and functionally diverse organic compounds.[75-79] Most

notable, they play an important role in drug discovery.[80-81]

CA

B

D

+ CA

B

D

MCR

A

B

A

B

CCA

B

D

Classical multistep synthesisFigure 4: Classical multistep synthesis versus multicomponent reactions.

In the ideal synthesis a compound should be obtained in 100% yield. Therefore controlling

the stereochemistry of a reaction is of significant importance. Stereochemistry is an

important factor in medicinal chemistry, since often only one enantiomer has the desired

biological activity. Biologically active (natural) products often occur as a single enantiomer,

while the other enantiomer is inactive or in some cases even harmful. A good example

of this is thalidomide (tradename: Softenon®, Figure 5), a drug which was prescribed to

many pregnant women in the 1950s and 1960s to prevent morning sickness. The drug,

which contains one stereocenter, was dosed to the pregnant women as a racemic mixture,

of which the (R)-enantiomer is effective against morning sickness the (S)-isomer is

teratogenic.[82] Due to this drug, over ten thousand children were born with malformations.

It is therefore highly desirable to synthesize enantiomerically pure compounds.

Page 25: Biocatalysis & Multicomponent Reactions: The Ideal …...Biocatalysis & Multicomponent Reactions Chapter 2 Monoamine Oxidase N: 41 A Promising Biocatalyst for Asymmetric Synthesis

General Introduction

25

Figure 5: Structural representation of thalidomide.

In 1850, Strecker reported what is nowadays considered to be the first MCR: a reaction

between an aldehyde or ketone, ammonium chloride and potassium cyanide to form an

α-aminonitrile, which can be hydrolyzed to give an amino acid (Scheme 14).[83] First, iminium

formation (38) takes place, followed by addition of the cyanide. The amino nitrile (39) is then

hydrolyzed to give desired amino acid 40.

Scheme 14: Strecker amino acid synthesis.

In 1891, Biginelli reported a MCR between urea, an aldehyde and a 1,3-diketone, yielding

dihydropyrimidone (Scheme 15).[84] The mechanism starts with condensation of the aldehyde

and urea to form iminium species 43 as an intermediate. This is followed by nucleophilic

addition of the ketoester (44), after which condensation takes place to give the product

(46) containing a newly formed stereocenter. These heterocyclic systems are known to be

of interest in the pharmaceutical industry, since the partly reduced pyrimidone derivatives

show biological effects such as antiviral, antitumor, antibacterial and anti-inflammatory

activity.[85]

Scheme 15: The Biginelli pyrimidone synthesis.

The Mannich reaction, discovered in 1912,[86] is a three-component reaction between a non-

enolizable aldehyde, a primary or secondary amine and an enolizable carbonyl compound

(Scheme 16), which has many useful applications in polymer chemistry and the agrochemical

and pharmaceutical industry.[87] The reaction starts with formation of an iminium ion (49),

Page 26: Biocatalysis & Multicomponent Reactions: The Ideal …...Biocatalysis & Multicomponent Reactions Chapter 2 Monoamine Oxidase N: 41 A Promising Biocatalyst for Asymmetric Synthesis

Chapter 1

26

which is then attacked by the enolized aldehyde (50) to form a β-amino-carbonyl (also

known as a Mannich base; 51) with two stereocenters.

Scheme 16: The Mannich reaction.

In 1921, Passerini described the very first isocyanide-based MCR between a carboxylic acid

(53), a carbonyl compound (52) and an isocyanide (53) giving α-hydroxy carboxamide

55 containing a newly formed stereocenter as a product (Scheme 17).[88] There have

been disputes about the actual mechanism of the Passerini reaction.[89-90] Two different

mechanisms have been postulated; the ionic and the concerted mechanism.

In contrast to the Ugi reaction (see below), the Passerini reaction is accelerated in aprotic

solvents indicating a concerted mechanism.[91-92] The concerted mechanism proceeds

through intermediate 57, which rearranges to the desired α-hydroxy carboxamide 55

(Scheme 17).

Scheme 17: The Passerini reaction via the concerted the mechanism.

Although Passerini reported his isocyanide based MCR as early as 1921, it was not until 1959

that the popularity of isocyanide chemistry began with the first 4-component reaction (4-

CR) with isocyanides reported by Ugi and coworkers (Scheme 18).[93-94] This reaction between

a ketone or aldehyde (52), an amine (58), an isocyanide (54) and a carboxylic acid (53) is a

very versatile reaction, and is widely used in the fields of modern combinatorial and medical

chemistry.[75-81] The proposed mechanism starts with imine formation by condensation of

the amine and the aldehyde/ketone, followed by formation of an iminium ion (59) by proton

exchange with the carboxylic acid. Subsequent addition of the isocyanide to the iminium

Page 27: Biocatalysis & Multicomponent Reactions: The Ideal …...Biocatalysis & Multicomponent Reactions Chapter 2 Monoamine Oxidase N: 41 A Promising Biocatalyst for Asymmetric Synthesis

General Introduction

27

ion followed by nucleophilic addition of the carboxylic acid forms adduct 61. In the final

step, an irreversible Mumm rearrangement takes place, transferring the acyl group from the

oxygen to the nitrogen and generating α-aminoacyl amide derivatives 62 (Scheme 18).

Scheme 18. Proposed mechanism of the Ugi four-component reaction (U-4CR).

1.4 Biocatalysis & Multicomponent Reactions

In spite of the fact that MCRs are very efficient by their nature, the stereocontrol in these

reactions is mostly not straightforward.[95-100] For most MCRs, catalytic asymmetric methods

to control the stereochemical outcome of the reaction are so far not available. Since

asymmetric induction is typically achieved by using optically pure inputs, it is important

to utilize methods which can efficiently and environmentally friendly generate these

inputs. The broad repertoire of stereospecific conversions by biocatalysts presents a unique

opportunity to address the stereoselectivity issue of certain MCRs.

The combination of (i) MCRs with (ii) biocatalysis presents an ideal approach (solution),

regarding selectivity (stereocontrol), flexibility, and sustainability. Several combinations of

biocatalysis with MCRs are conceivable (Figure 6):

(A) Kinetic resolution of a racemic MCR product.

(B) Dynamic kinetic resolution in which the optically pure product reacts in a MCR.

(C) Resolution of a racemic MCR product to generate a chiral input for a second MCR

(combination of A and B).

(D) Biotransformation to generate optically pure inputs for diastereoselective MCRs.

Page 28: Biocatalysis & Multicomponent Reactions: The Ideal …...Biocatalysis & Multicomponent Reactions Chapter 2 Monoamine Oxidase N: 41 A Promising Biocatalyst for Asymmetric Synthesis

Chapter 1

28

Figure 6a: Combination (A-B) of MCRs and biocatalysis.

Figure 6b: Combination (C-D) of MCRs and biocatalysis.

Thus, combining MCRs and biocatalysis can result in enantioselective complex molecules

from readily available compounds, possibly providing molecules having major potential in

medicinal chemistry. However, the combination of these two methods also addresses more

fundamental issues in the stereoselective construction of complex molecules.

Despite the great potential of biocatalysis and MCRs, the combination of these two

methodologies to generate optically pure complex compounds has hardly been described

in literature (see next section).

1.5 Applications of Biocatalysis & Multicomponent Reactions

1.5.1 Multicomponent Reactions/Enzymatic Kinetic Resolution

The first combination of multicomponent chemistry and biocatalysis was reported by

Dordick and coworkers. They combined the selectivity of biocatalysis with the strength of

multicomponent chemistry in the synthesis of a Ugi four-component condensation (U-4CR)

library of α-(acylamino)amides of type 71 (Scheme 19).[101] β-hydroxy acid 63 and amino

Page 29: Biocatalysis & Multicomponent Reactions: The Ideal …...Biocatalysis & Multicomponent Reactions Chapter 2 Monoamine Oxidase N: 41 A Promising Biocatalyst for Asymmetric Synthesis

General Introduction

29

alcohol 66 underwent enzyme-catalyzed acylation prior to being employed in the U-4CR as

these inputs would require protecting and deprotecting steps when chemically synthesized

(Scheme 19). Porcine pancreatic lipase (PPL) was employed since lipases can catalyze the

acylation of both hydroxy acids and amino alcohols.[102-103] PPL proved to be an efficient

biocatalyst and gave acylated carboxylic acid 65 and amine 68 in yields over >90%. The

yield for the acylated carboxylic acid (>90%) implies that the corresponding lipase was non-

stereoselective. With these inputs in hand, the U-4CR was performed using acetaldehyde

69 and methyl isocyanoacetate 70 and resulted in a 9-membered library of α-(acylamino)

amides 71. Unfortunately, no enantiomeric and diastereomeric ratios were reported.

Scheme 19: Lipase-catalyzed acylation of hydroxy acid 68 and amino alcohol 67 followed by a U-4CR generating α-(acylamino)amides 71.

In the same year, enantiomerically pure dihydropyrimidones (DHPMs) were generated

by a Biginelli multicomponent reaction followed by a lipase-catalyzed kinetic resolution

of the corresponding MCR products.[104] DHPMs play an important role as calcium

channel modulators.[105-107] Moreover, different enantiomers of dihydropyrimidones show

differentiation in pharmacological activity (calcium entry blocking versus activation).

Access to enantiomerically pure DHPMs via chemical synthesis showed to be not that

straightforward.[105, 107] So, introducing stereoselectivity using biocatalysis seems appealing.

To show the feasibility of MCRs (Biginelli) and biocatalysis (lipase) as a combination, SQ

32926 (82) a potent orally active antihypertensive agent, was synthesized (Scheme 20).[105-

106] After generating DHPM scaffold 75 via a Biginelli MCR, a cleavable ester functionality

at the N3 position was introduced allowing a kinetic resolution and obtaining the desired

enantiomer (78) in excellent enantiomeric excess (98%). Compound 78 was subsequently

manipulated providing the desired target compound (R)-SQ 32926 (82, Scheme 20).

Page 30: Biocatalysis & Multicomponent Reactions: The Ideal …...Biocatalysis & Multicomponent Reactions Chapter 2 Monoamine Oxidase N: 41 A Promising Biocatalyst for Asymmetric Synthesis

Chapter 1

30

Scheme 20: Synthesis of (R)-SQ 32926 (82) via a Biginelli MCR and a lipase-catalyzed kinetic resolution.

A modified strategy in obtaining enantiomerically pure DHPMs was developed by Prasad

and coworkers.[108] The Biginelli reaction was not performed under conventional heating

but was microwave assisted. This method gave short reaction times and high yields for the

resulting compounds (90, Scheme 21). Also, instead of having a cleavable ester functionality

at the N3 position, they introduced an acetoxy group on the aldehyde‘s aromatic ring

system (87, Scheme 21). This acetoxyl group was cleaved of by Novozyme 435 generating

optically pure DHPMs 88 and 89. Compound 88 was subsequently deacetylated generating

90. Unfortunately, this group was not able to determine the ee’s of their compounds but just

mentioned that the ee’s were moderate to high.

Page 31: Biocatalysis & Multicomponent Reactions: The Ideal …...Biocatalysis & Multicomponent Reactions Chapter 2 Monoamine Oxidase N: 41 A Promising Biocatalyst for Asymmetric Synthesis

General Introduction

31

Scheme 21: Alternative strategy to enantiomerically pure DHPMs by Prasad and coworkers.[108]

Two groups have employed a similar strategy to generate enantiomerically pure α,α-dialkyl-

α-hydroxylic acid esters 96 and 102 where a hydrolase selectively cleaves an ester moiety

from a Passerini multicomponent product (94 and 100, Scheme 22 and 23). α,α-Dialkyl-

α-hydroxylic acid esters are important intermediates in organic chemistry. For example,

this moiety is present in the natural product clerodendrin A[109] but also found use as

cyclooxygenase inhibitor.[110] Ostaszewski and coworkers obtained enantiomerically pure

secondary alcohol 96 by first performing a Passerini MCR of acetic acid, phenylacetaldehyde

and 4-methoxybenxyl isocyanide generating α-acetoxyamide 94 subsequently followed by

a hydrolysis with Wheat Germ lipase obtaining the desired compound (96) in 47% yield and

93% ee (Scheme 22).[111]

Page 32: Biocatalysis & Multicomponent Reactions: The Ideal …...Biocatalysis & Multicomponent Reactions Chapter 2 Monoamine Oxidase N: 41 A Promising Biocatalyst for Asymmetric Synthesis

Chapter 1

32

Scheme 22: Synthesis of optically pure secondary alcohol 96 using a Passerini MCR and enzyme-catalyzed hydrolysis.

Bornscheuer and coworkers obtained enantiomerically pure tertiary alcohol 102 by first

executing a Passerini MCR using 1,1,1-trifluoro-2-butanone, acetic acid and tert-butyl

isocyanide to obtain Passerini product 100 (Scheme 23). Subsequent hydrolysis using an

esterase of metagenome origin (genetic material recovered directly from environmental

samples, esterase 8) led to the desired product in a 46% yield and 85% ee.[112]

Scheme 23: Synthesis of optically pure tertiary alcohol 102 using a Passerini MCR and enzyme-catalyzed hydrolysis.

Another MCR combined with a kinetic resolution has been employed by Beller and

coworkers.[113] This MCR generates O-acyl-substituted cyclohexenes 108 from anhydrides

103, aldehydes 104 and dienophiles 106 and involves a Diels-Alder reaction (Scheme 24).

Page 33: Biocatalysis & Multicomponent Reactions: The Ideal …...Biocatalysis & Multicomponent Reactions Chapter 2 Monoamine Oxidase N: 41 A Promising Biocatalyst for Asymmetric Synthesis

General Introduction

33

The addition of the dienophile was very selective and only the endo-addition product was

observed (with one exception).

Since this MCR is performed at relatively high temperatures (110 °C), the development

of an enantioselective variant of this MCR is difficult. Therefore, a lipase-catalyzed kinetic

resolution was employed in order to obtain enantiomerically pure cyclohexenols. For each

individual ester the optimal enzyme was determined giving high yields (48–50%) with ee’s

ranging from 91–99%. The stereoselectivity could be varied by using different lipases.

Scheme 24: A MCR of anhydrides, aldehydes, and dienophiles involving a Diels-Alder reaction with a subsequent lipase-catalyzed kinetic resolution generating enantiomerically pure cyclohexanols.

Besides anhydrides, also acyl chlorides, alcohols or amides could be used as inputs in

the generation of acylamino-1,3-butadiene 110 (Scheme 25).[114] Furthermore, various

dienophiles could be used, which increases the versatility of this reaction quite significantly

(Scheme 25).[114] The resulting compounds were obtained in comparable yields and

diastereoselectivity as described previously.[113] Unfortunately, cleaving the ester moiety

with a lipase-catalyzed kinetic resolution proved to be quite difficult. The ee’s obtained

from the kinetic resolution of Diels-Alder product 111 and 114 varied significantly (no

conversion-99% ee), whereas Diels-Alder products 112 and 113 did not give any conversion.

Page 34: Biocatalysis & Multicomponent Reactions: The Ideal …...Biocatalysis & Multicomponent Reactions Chapter 2 Monoamine Oxidase N: 41 A Promising Biocatalyst for Asymmetric Synthesis

Chapter 1

34

Scheme 25: Synthesis of a wide range of enantiomerically pure cyclohexenols derivatives by combining MCRs and lipases.

1.5.2 Multicomponent Reactions/Enzymatic Dynamic Kinetic Resolution

The first combination of a MCR and an enzymatic desymmetrization was reported

by Ostaszewski and coworkers in 2005.[115] They developed a one-pot enzymatic

desymmetrization of prochiral 3-phenylglutaric anhydrides 115 coupled to an Ugi 4CR

generating a small library of chiral peptidomimetics (Scheme 26). Various lipases were

screened and Chirazyme L was found to be the most suitable catalyst. Moderate to

reasonable ee’s and excellent yields were obtained of desired carboxylic acids 116. The

subsequent Ugi 4CR reaction tolerated aliphatic and aromatic aldehydes and amines, but

also benzyl isocyanide and ethyl isocyanoacetate. Most notably, all compounds 120 were

obtained as diastereoisomeric mixtures, in approximately 1:1 ratio. This is not very surprising

as it is known that chiral carboxylic acid give poor induction.[76]

Scheme 26: Enzyme-catalyzed desymmetrization combined with an Ugi 4CR generating chiral peptidomimetics 120.

Page 35: Biocatalysis & Multicomponent Reactions: The Ideal …...Biocatalysis & Multicomponent Reactions Chapter 2 Monoamine Oxidase N: 41 A Promising Biocatalyst for Asymmetric Synthesis

General Introduction

35

1.5.3 Enzymes as Catalysts for Multicomponent Reactions

Besides using biocatalysis as a tool for introducing chirality into a compound, enzymes can

also be used to make certain chemical conversions possible. These conversions might be

virtually impossible using traditional methods which could be attributed to the fact that

these conversions require, for example, an extremely expensive reactant, high stereo- or

regio-selectivity, or have a very high energy barrier.

Since substituted aromatic and aliphatic aldehydes lower the yield of the Biginelli

multicomponent reaction considerably,[116-117] bakers’ yeast (Saccharomyces cerevisiae) has

been employed as an efficient catalyst in the Biginelli reaction.[118] With only glucose as

additive, several activated and deactivated aromatic aldehydes (121), β-keto esters 122 and

urea/thiourea (123) were employed and gave excellent yield of the desired DHPMs (124,

Scheme 27). However, no stereoselectivity could be observed when employing this strategy.

Scheme 27: Biginelli reaction catalyzed by bakers’ yeast (Saccharomyces cerevisiae).

Another interesting example of enzymes catalyzing MCRs was shown by Yu and coworkers,

who have shown that lipases are able to catalyze Mannich reactions efficiently. [119-120] The

Mannich reaction is one of most important reaction in generating C-C bonds. The resulting

β-amino carbonyl compounds are important intermediates for several pharmaceuticals and

natural products.[121] Unfortunately, direct Mannich reactions often suffer from selectivity

problems since both the aldehyde and the CH-acidic substrate can often act as either

nucleophile or electrophile. Therefore, preformed electrophiles, such as imines, iminium

salts and hydrazones or nucleophiles such as enolates, enol ethers and enamines, or both

are regularly used (also called an indirect Mannich). These pretreatments may lead to several

drawbacks caused by preparation, isolation and purification. Yu and coworkers developed

a one-pot CRL (Candida rugosa lipase) catalyzed direct Mannich reaction resulting in

the corresponding β-amino carbonyl compounds 128 in high yields but disappointing

diastereomeric ratios (Scheme 28). A wide range of aromatic aldehydes (125) were employed

and electron-withdrawing substituents on the aldehyde were favored. Unfortunately, using

aliphatic aldehydes as inputs did not lead to the desired β-amino carbonyl compounds.

Page 36: Biocatalysis & Multicomponent Reactions: The Ideal …...Biocatalysis & Multicomponent Reactions Chapter 2 Monoamine Oxidase N: 41 A Promising Biocatalyst for Asymmetric Synthesis

Chapter 1

36

Scheme 28: Lipase-catalyzed direct Mannich reactions.

1.6 Conclusions

Multicomponent chemistry is a convenient tool for easy access to large collections

of structurally and functionally diverse organic compounds that can be screened for

pharmacological activity or as ligands for novel transition metal catalysis. In spite of the fact

that MCRs are very efficient by their nature, the stereocontrol in these reactions is mostly

not straightforward. So it is evident that the combination of MCRs and biocatalysis can be

very useful in generating enantiomerically pure complex products. Most of the procedures

that have appeared in the literature rely on an enzymatic kinetic resolution which inherently

leads to 50% loss of the desired optically pure compound. To our knowledge only one

precedent methodology combining an enzymatic dynamic resolution with an MCR is

available (see section 1.5.2). This opens up possibilities to explore the combination of MCRs

and DKR/desymmetrization to generate high added value compounds in excellent yields

and enantiomeric excesses.

1.7 Scope and Outline of this Thesis

The research described in this thesis involves the generation of highly complex and optically

pure molecules based on the combination of optically pure 3,4-disubstituted 1-pyrrolines

and several multicomponent reactions.

Chapter 2 will describe the main biocatalyst (monoamine oxidase N) used during this

research and illustrate how to gain access to optically pure 3,4-disubstituted 1-pyrrolines

efficiently.

In Chapter 3 3,4-disubstituted 1-pyrrolines are employed in a Joullié-Ugi MCR generating

pharmaceutically relevant optically active substituted prolyl peptides. These substituted

prolyl peptides also showed to be useful as highly active organocatalysts for stereoselective

conjugate addition of aldehydes to nitroolefins.

In Chapter 4 the prolyl peptides scaffold could be further complexed and diversified

Page 37: Biocatalysis & Multicomponent Reactions: The Ideal …...Biocatalysis & Multicomponent Reactions Chapter 2 Monoamine Oxidase N: 41 A Promising Biocatalyst for Asymmetric Synthesis

General Introduction

37

by introducing cyclization reactions. Carefully selected prolyl peptides will undergo

Pictet-Spengler cyclization to obtain highly complex alkaloid-like polycyclic compounds

(2,5-diketopiperazines) with high diversity. To the best of our knowledge, it also constitutes the

first example of MCR chemistry to synthesize 5-membered ring-fused 2,5-diketopiperazines.

The efficiency and generality of the approach and the resulting molecular diversity and

complexity makes our monoamine oxidase N and Joullié-Ugi MCR combination highly

interesting for medicinal chemistry. Most notably, a very short and efficient synthesis of

important hepatitis C drug Telaprevir (Incivek™), featuring a biocatalytic desymmetrization

and two multicomponent reactions (Passerini and Joullié-Ugi 3CR) as the key steps is

described in Chapter 5.

In Chapter 6 3,4-disubstituted 1-pyrrolines are combined with a Ugi-Smiles MCR generating

enantiomerically pure N-aryl proline (thio)amides. Remarkably, some Ugi-Smiles products

show unusually high optical rotations.

Finally, Chapter 7 summarizes the main conclusions of this work and provides a further

outlook on the combination of biocatalysis and multicomponent chemistry.

1.8 References & Notes[1] R. A. Sheldon, I. Arends and U. Hanefeld, Green Chemistry and Catalysis, Wiley-VCH, Weinheim,

2007.[2] M. Meinshausen, B. Hare, T. M. L. Wigley, D. Van Vuuren, M. G. J. Den Elzen and R. Swart, Climatic

Change 2006, 75, 151-194.[3] B. Hare and M. Meinshausen, Climatic Change 2006, 75, 111-149.[4] P. A. Wender, S. T. Handy and D. L. Wright, Chem. Ind. 1997, 765-769.[5] E. Fischer, Ber. Dtsch. Chem. Ges. 1894, 27, 2985–2993.[6] E. Buchner, Ber. Dtsch. Chem. Ges. 1897, 30, 117-124.[7] J. B. Sumner, J. Biol. Chem. 1926, 69, 435-441.[8] F. Sanger and E. O. P. Thompson, Biochem. J. 1953, 53, 353-366.[9] J. Monod, J. Wyman and J. P. Changeux, J. Mol. Biol. 1965, 12, 88-118.[10] P. Grunwald, Biocatalysis: biochemical fundamentals and applications, London: Imperial College

Press, Singapore, 2009.[11] H. C. Hailes, P. A. Dalby and J. M. Woodley, J. Chem. Technol. Biotechnol. 2007, 82, 1063-1066.[12] K. E. Jaeger and T. Eggert, Curr. Opin. Biotechnol. 2004, 15, 305-313.[13] A. Schmid, J. S. Dordick, B. Hauer, A. Kiener, M. Wubbolts and B. Witholt, Nature 2001, 409, 258-

268.[14] C. S. Chen and C. J. Sih, Angew. Chem., Int. Ed. 1989, 28, 695-707.[15] S. G. Burton, D. A. Cowan and J. M. Woodley, Nat. Biotechnol. 2002, 20, 37-45.[16] K. M. Ulmer, Science 1983, 219, 666-671.[17] D. A. Estell, T. P. Graycar and J. A. Wells, J. Biol. Chem. 1985, 260, 6518-6521.[18] R. J. Fox, S. C. Davis, E. C. Mundorff, L. M. Newman, V. Gavrilovic, S. K. Ma, L. M. Chung, C. Ching,

S. Tam, S. Muley, J. Grate, J. Gruber, J. C. Whitman, R. A. Sheldon and G. W. Huisman, Nat Biotech 2007, 25, 338-344.

[19] K. A. Gray, L. Zhao and M. Emptage, Curr. Opin. Chem. Biol. 2006, 10, 141-146.[20] A. McCarthy, Chem. Biol. 2003, 10, 893-894.

Page 38: Biocatalysis & Multicomponent Reactions: The Ideal …...Biocatalysis & Multicomponent Reactions Chapter 2 Monoamine Oxidase N: 41 A Promising Biocatalyst for Asymmetric Synthesis

Chapter 1

38

[21] R. J. Kazlauskas and U. T. Bornscheuer, Nat. Chem. Biol. 2009, 5, 526-529.[22] V. G. H. Eijsink, A. Bjørk, S. Gåseidnes, R. Sirevåg, B. Synstad, B. v. d. Burg and G. Vriend, J. Biotechnol.

2004, 113, 105-120.[23] B. van den Burg and V. G. H. Eijsink, Curr. Opin. Biotechnol. 2002, 13, 333-337.[24] C. Ó’Fágáin, Enzyme Microb. Technol. 2003, 33, 137-149.[25] C. A. Hutchison, S. Phillips, M. H. Edgell, S. Gillam, P. Jahnke and M. Smith, J. Biol. Chem. 1978, 253,

6551-6560.[26] J. M. Chalker and B. G. Davis, Curr. Opin. Chem. Biol. 2010, 14, 781-789.[27] L. Wang and P. G. Schultz, Angew. Chem., Int. Ed. 2005, 44, 34-66.[28] N. J. Turner, Nat. Chem. Biol. 2009, 5, 567-573.[29] S. Lutz and W. M. Patrick, Curr. Opin. Biotechnol. 2004, 15, 291-297.[30] D. E. Robertson and B. A. Steer, Curr. Opin. Chem. Biol. 2004, 8, 141-149.[31] J. Wei, Y. Zhou, T. Xu and B. Lu, Appl. Biochem. Biotechnol. 2010, 162, 116-126.[32] Z.-l. Han, S.-y. Han, S.-p. Zheng and Y. Lin, Appl. Microbiol. Biotechnol. 2009, 85, 117-126.[33] T. Belien, I. J. Joye, J. A. Delcour and C. M. Courtin, Protein Eng. Des. Sel. 2009, 22, 587-596.[34] M. Brouk, Y. Nov and A. Fishman, Appl. Environ. Microbiol. 2010, 76, 6397-6403.[35] F. Escalettes and N. J. Turner, ChemBioChem 2008, 9, 857-860.[36] E. Campbell, I. R. Wheeldon and S. Banta, Biotechnol. Bioeng. 2010, 107, 763-774.

[37] J. D. Carballeira, P. Krumlinde, M. Bocola, A. Vogel, M. T. Reetz and J.-E. Baeckvall, Chem. Commun. 2007, 1913-1915.

[38] S. Bartsch and U. T. Bornscheuer, Protein Eng. Des. Sel. 2010, 23, 929-933.[39] J. Haerle, S. Guenther, B. Lauinger, M. Weber, B. Kammerer, D. L. Zechel, A. Luzhetskyy and A.

Bechthold, Chem. Biol. 2011, 18, 520-530.[40] V. Mathieu, J. Fastrez and P. Soumillion, Protein Eng. Des. Sel. 2010, 23, 699-709.[41] G. A. Behrens, A. Hummel, S. K. Padhi, S. Schätzle and U. T. Bornscheuer, Adv. Synth. Catal. 2011, 13,

2191-2215.[42] U. T. Bornscheuer and R. J. Kazlauskas, Hydrolases in Organic Synthesis, Wiley-VCH, Weinheim,

2006.[43] F. Balkenhohl, K. Ditrich, B. Hauer and W. Ladner, J. Prakt. Chem.-Chem. Ztg. 1997, 339, 381-384.[44] Y. Ahn, S.-B. Ko, M.-J. Kim and J. Park, Coord. Chem. Rev. 2008, 252, 647-658.[45] M. D. Truppo, N. J. Turner and J. D. Rozzell, Chem. Commun. 2009, 2127-2129.[46] M. Hoehne, S. Kuehl, R. Karen and U. T. Bornscheuer, ChemBioChem 2008, 9, 363-365.[47] N. Itoh, C. Yachi and T. Kudome, J. Mol. Catal. B: Enzym. 2000, 10, 281-290.[48] T. Vaijayanthi and A. Chadha, Tetrahedron: Asymmetry 2008, 19, 93-96.[49] R. Noyori and S. Hashiguchi, Acc. Chem. Res. 1997, 30, 97-102.[50] S. Hashiguchi, A. Fujii, K. J. Haack, K. Matsumura, T. Ikariya and R. Noyori, Angew. Chem., Int. Ed.

1997, 36, 288-290.[51] G. R. A. Adair and J. M. J. Williams, Chem. Commun. 2007, 2608-2609.[52] Y. Shimada, Y. Miyake, H. Matsuzawa and Y. Nishibayashi, Chem. Asian J. 2007, 2, 393-396.[53] W. Kroutil, H. Mang, K. Edegger and K. Faber, Curr. Opin. Chem. Biol. 2004, 8, 120-126.[54] C. V. Voss, C. C. Gruber and W. Kroutil, Angew. Chem., Int. Ed. 2008, 47, 741-745.[55] W. Stampfer, B. Kosjek, C. Moitzi, W. Kroutil and K. Faber, Angew. Chem., Int. Ed. 2002, 41, 1014-

1017.[56] B. Plietker, Tetrahedron: Asymmetry 2005, 16, 3453-3459.[57] P. Hoyos, J.-V. Sinisterra, F. Molinari, A. R. Alcantara and P. D. De Maria, Acc. Chem. Res. 2010, 43,

288-299.

Page 39: Biocatalysis & Multicomponent Reactions: The Ideal …...Biocatalysis & Multicomponent Reactions Chapter 2 Monoamine Oxidase N: 41 A Promising Biocatalyst for Asymmetric Synthesis

General Introduction

39

[58] P. D. de Maria, T. Stillger, M. Pohl, M. Kiesel, A. Liese, H. Groeger and H. Trauthwein, Adv. Synth. Catal. 2008, 350, 165-173.

[59] A. G. Myers, B. H. Yang, H. Chen, L. McKinstry, D. J. Kopecky and J. L. Gleason, J. Am. Chem. Soc. 1997, 119, 6496-6511.

[60] I. Paterson, K. S. Yeung, C. Watson, R. A. Ward and P. A. Wallace, Tetrahedron 1998, 54, 11935-11954.

[61] S. D. Bull, S. G. Davies, R. L. Nicholson, H. J. Sanganee and A. D. Smith, Tetrahedron: Asymmetry 2000, 11, 3475-3479.

[62] D. A. Evans, P. Nagorny, K. J. McRae, D. J. Reynolds, L.-S. Sonntag, F. Vounatsos and R. Xu, Angew. Chem., Int. Ed. 2007, 46, 537-540.

[63] M. Cheeseman, I. R. Davies, P. Axe, A. L. Johnson and S. D. Bull, Org. Biomol. Chem. 2009, 7, 3537-3548.

[64] Y. S. Angelis and I. Smonou, Tetrahedron Lett. 1997, 38, 8109-8112.[65] I. J. Colton, S. N. Ahmed and R. J. Kazlauskas, J. Org. Chem. 1995, 60, 212-217.[66] G. DeSantis, Z. L. Zhu, W. A. Greenberg, K. V. Wong, J. Chaplin, S. R. Hanson, B. Farwell, L. W.

Nicholson, C. L. Rand, D. P. Weiner, D. E. Robertson and M. J. Burk, J. Am. Chem. Soc. 2002, 124, 9024-9025.

[67] P. Pochlauer, Chim. Oggi 1998, 16, 15-19.[68] D. V. Johnson and H. Griengl, Chim. Oggi 1997, 15, 9-13.[69] R. J. H. Gregory, Chem. Rev. 1999, 99, 3649-3682.[70] M. North, Tetrahedron: Asymmetry 2003, 14, 147-176.[71] J. M. Brunel and I. P. Holmes, Angew. Chem., Int. Ed. 2004, 43, 2752-2778.[72] J. Holt and U. Hanefeld, Curr. Org. Synth. 2009, 6, 15-37.[73] J. Sukumaran and U. Hanefeld, Chem. Soc. Rev. 2005, 34, 530-542.[74] T. Gaich and P. S. Baran, J. Org. Chem.2010, 75, 4657-4673.[75] A. Domling and I. Ugi, Angew. Chem., Int. Ed. 2000, 39, 3169-3210.[76] J. Zhu and H. Bienayme, Multicomponent Reactions, Wiley-VCH, Weinheim, 2005.[77] R. V. A. Orru and M. de Greef, Synthesis 2003, 1471-1499.[78] A. Domling, Chem. Rev. 2006, 106, 17-89.[79] B. Ganem, Acc. Chem. Res. 2009, 42, 463-472.[80] C. Hulme and V. Gore, Curr. Med. Chem. 2003, 10, 51-80.[81] I. Akritopoulou-Zanze, Curr. Opin. Chem. Biol. 2008, 12, 324-331.[82] J. H. Kim and A. R. Scialli, Toxicol. Sci. 2011, 122, 1-6.[83] A. Strecker, Justus Liebigs Ann. Chem. 1850, 75, 27-51.[84] P. Biginelli, Ber. Dtsch. Chem. Ges. 1891, 24, 1317-1319.[85] C. O. Kappe, Eur. J. Med. Chem. 2000, 35, 1043-1052.[86] C. Mannich and W. Krosche, Arch. Pharm. 1912, 250, 647-667.[87] M. Arend, B. Westermann and N. Risch, Angew. Chem., Int. Ed. 1998, 37, 1044-1070.[88] M. Passerini, Gazz. Chim. Ital. 1921, 51, 126-129.[89] D. Seebach, G. Adam, T. Gees, M. Schiess and W. Weigand, Chem. Ber. Recl. 1988, 121, 507-517.[90] T. Carofiglio, P. G. Cozzi, C. Floriani, A. Chiesivilla and C. Rizzoli, Organometallics 1993, 12, 2726-

2736.[91] I. Ugi and R. Meyr, Chem. Ber. Recl. 1961, 94, 2229-2233.[92] S. Maeda, S. Komagawa, M. Uchiyama and K. Morokuma, Angew. Chem., Int. Ed. 2011, 50, 644-649.[93] I. Ugi and C. Steinbrückner, Angew. Chem. 1960, 72, 267-268.[94] I. Ugi, U. Meyr, C. Fetzer and C. Steinbruckner, Angew. Chem. 1959, 71, 386-388.[95] P. R. Andreana, C. C. Liu and S. L. Schreiber, Org. Lett. 2004, 6, 4231-4233.

Page 40: Biocatalysis & Multicomponent Reactions: The Ideal …...Biocatalysis & Multicomponent Reactions Chapter 2 Monoamine Oxidase N: 41 A Promising Biocatalyst for Asymmetric Synthesis

Chapter 1

40

[96] S. E. Denmark and Y. Fan, J. Am. Chem. Soc. 2003, 125, 7825-7827.[97] U. Kusebauch, B. Beck, K. Messer, E. Herdtweck and A. Domling, Org. Lett. 2003, 5, 4021-4024.[98] S. C. Pan and B. List, Angew. Chem., Int. Ed. 2008, 47, 3622-3625.[99] D. J. Ramon and M. Yus, Angew. Chem., Int. Ed. 2005, 44, 1602-1634.[100] T. Yue, M. X. Wang, D. X. Wang, G. Masson and J. P. Zhu, J. Org. Chem. 2009, 74, 8396-8399.[101] X. C. Liu, D. S. Clark and J. S. Dordick, Biotechnol. Bioeng. 2000, 69, 457-460.[102] T. Sugai and H. Ohta, Tetrahedron Lett. 1991, 32, 7063-7064.[103] N. Chinsky, A. L. Margolin and A. M. Klibanov, J. Am. Chem. Soc. 1989, 111, 386-388.[104] B. Schnell, W. Krenn, K. Faber and C. O. Kappe, J. Chem. Soc.-Perkin Trans. 1 2000, 4382-4389.[105] K. S. Atwal, B. N. Swanson, S. E. Unger, D. M. Floyd, S. Moreland, A. Hedberg and B. C. Oreilly, J. Med.

Chem. 1991, 34, 806-811.[106] G. J. Grover, S. Dzwonczyk, D. M. McMullen, D. E. Normandin, C. S. Parham, P. G. Sleph and S.

Moreland, J. Cardiovasc. Pharmacol. 1995, 26, 289-294.[107] G. C. Rovnyak, K. S. Atwal, A. Hedberg, S. D. Kimball, S. Moreland, J. Z. Gougoutas, B. C. Oreilly, J.

Schwartz and M. F. Malley, J. Med. Chem. 1992, 35, 3254-3263.[108] A. K. Prasad, C. Mukherjee, S. K. Singh, R. Brahma, R. Singh, R. K. Saxena, C. E. Olsen and V. S.

Parmar, J. Mol. Catal. B: Enzym. 2006, 40, 93-100.[109] N. Kato, Shibayam.S, K. Munakata and C. Katayama, J. Chem. Soc. D, Chem. Commun. 1971, 1632-

1633.[110] L. Tan, C. Y. Chen, W. R. Chen, L. Frey, A. O. King, R. D. Tillyer, F. Xu, D. Zhao, E. J. J. Grabowski, P. J.

Reider, P. O’Shea, P. Dagneau and X. Wang, Tetrahedron 2002, 58, 7403-7410.[111] W. Szymanski and R. Ostaszewski, J. Mol. Catal. B: Enzym. 2007, 47, 125-128.[112] R. Kourist, G.-S. Nguyen, D. Struebing, D. Boettcher, K. Liebeton, C. Naumer, J. Eck and U. T.

Bornscheuer, Tetrahedron: Asymmetry 2008, 19, 1839-1843.[113] D. Strubing, A. Kirschner, H. Neumann, S. Hubner, S. Klaus, U. T. Bornscheuer and M. Beller, Chem.-

Eur. J. 2005, 11, 4210-4218.[114] D. Strubing, H. Neumann, S. Klaus, A. Jacobi von Wangelin, D. Gordes, M. Beller, P. Braiuca, C. Ebert,

L. Gardossi and U. Kragl, Tetrahedron 2004, 60, 683-691.[115] A. Fryszkowska, J. Frelek and R. Ostaszewski, Tetrahedron 2005, 61, 6064-6072.[116] K. Folkers, H. J. Harwood and T. B. Johnson, J. Am. Chem. Soc. 1932, 54, 3751-3758.[117] P. Wipf and A. Cunningham, Tetrahedron Lett. 1995, 36, 7819-7822.[118] A. Kumar and R. A. Maurya, Tetrahedron Lett. 2007, 48, 4569-4571.[119] K. Li, T. He, C. Li, X.-W. Feng, N. Wang and X.-Q. Yu, Green Chem. 2009, 11, 777-779.[120] T. He, K. Li, M.-Y. Wu, X.-W. Feng, N. Wang, H.-Y. Wang, C. Li and X.-Q. Yu, J. Mol. Catal. B: Enzym.

2010, 67, 189-194.[121] R. Muller, H. Goesmann and H. Waldmann, Angew. Chem., Int. Ed. 1999, 38, 184-187.

Page 41: Biocatalysis & Multicomponent Reactions: The Ideal …...Biocatalysis & Multicomponent Reactions Chapter 2 Monoamine Oxidase N: 41 A Promising Biocatalyst for Asymmetric Synthesis

Monoamine Oxidase N: A Promising Biocatalyst for Asymmetric Synthesis

Partly Published in: Angew. Chem. 2010, 122, 2228-2230;

Angew. Chem. Int. Ed. 2010, 49, 2182-2184

Chapter 2

Page 42: Biocatalysis & Multicomponent Reactions: The Ideal …...Biocatalysis & Multicomponent Reactions Chapter 2 Monoamine Oxidase N: 41 A Promising Biocatalyst for Asymmetric Synthesis
Page 43: Biocatalysis & Multicomponent Reactions: The Ideal …...Biocatalysis & Multicomponent Reactions Chapter 2 Monoamine Oxidase N: 41 A Promising Biocatalyst for Asymmetric Synthesis

Monoamine Oxidase N

43

2.1 Introduction

Amines are ubiquitous in Nature. For example they play an important role in cell growth

and differentiation,[1] neoplastic cell proliferation[2-3] and are also a very important source

for carbon and nitrogen for a number of yeast and bacterial species. It is therefore not

surprising that amine oxidation is an omnipresent process in biology, participating in the

transformation of neurotransmitters via oxidative deamination[4] as well as the detoxification

of xenobiotics.[5]

Enzymes capable of oxidizing amines (amine oxidases; AO) are found in numerous

organisms ranging from bacteria, yeasts, and plants to mammals and can be grouped into

the flavoprotein and quinoprotein families. The mechanism of amine oxidation catalyzed by

the quinoprotein amine oxidases which comprises e.g. diamine oxidases (E. C. 1.4.3.6) has

been well characterized and occurs through the formation of an enzyme–substrate covalent

adduct with either topaquinone (TPQ), tryptophan tryptophylquinone (TTQ), cysteine

tryptophylquinone (CTQ) or lysine tyrosyl quinone (LTQ) as a cofactor (Scheme 1).[6-7]

Scheme 1: Mechanism for the oxidation of amines by a quinoprotein amine oxidase.

The intermediates in Scheme 1 have been characterized by a combination of chemical,

kinetic and spectroscopic techniques. First, TPQox reacts with an amine substrate giving

quinone ketimine A, which, after a formal proton shift, gives rise to quinolaldimine B. After

hydrolysis and release of the aldehyde, TPQred is formed. In a mechanism which is not fully

understood, the reduced cofactor is reoxidized by molecular oxygen back to TPQox to start

a new catalytic cycle.

Page 44: Biocatalysis & Multicomponent Reactions: The Ideal …...Biocatalysis & Multicomponent Reactions Chapter 2 Monoamine Oxidase N: 41 A Promising Biocatalyst for Asymmetric Synthesis

Chapter 2

44

Scheme 2: Proposed mechanism by of amine oxidation by flavoproteins.

In contrast to amine oxidation by the quinoprotein, the mechanism of amine oxidation by

flavoproteins comprising e.g. monoamine oxidases (E. V. 1.4.3.4), amino acid oxidases (E. C.

1.4.3.2 and E. C. 1.4.3.3) and polyamine oxidases (EC 1.5.3.11) is still a topic of debate. The

uncertainties regarding the mechanism of amine oxidation by flavoenzymes are due to the

high chemical versatility of the isoalloxazine system. Because the fully oxidized cofactor is

readily reduced by either one or two electrons, flavoproteins are capable of participating in

several reactions resulting in several proposed mechanisms. While this range of reactivities

is clearly of advantage to biological systems, it complicates the mechanistic analyses of

Page 45: Biocatalysis & Multicomponent Reactions: The Ideal …...Biocatalysis & Multicomponent Reactions Chapter 2 Monoamine Oxidase N: 41 A Promising Biocatalyst for Asymmetric Synthesis

Monoamine Oxidase N

45

flavoenzymes, in that several viable mechanisms must be considered (scheme 2). These

include (i) formation of a substrate carbanion by loss of a proton due to an active site base,[8]

(ii) single electron transfer reaction generating an aminium radical,[9] (iii) nucleophilic attack

by the amine on the flavin C4α followed by proton abstraction by an active site base.[10]

Despite these proposals, the mechanism of oxidation of amines by flavoproteins has been

controversial for decades and a consensus view (although the single electron transfer

mechanism is the one most widely accepted) has not emerged following a period of

intensive research activity over several years.

2.2 Monoamine Oxidase

2.2.1 Mammalian Monoamine Oxidase

In 1928, Mary Bernheim (née Mary Hare) was the first to describe a new oxidative

pathway in the liver and named the enzyme involved tyramine oxidase.[11] This enzyme

was eventually renamed monoamine oxidase (MAO). MAO oxidizes primary aliphatic and

aromatic amines as well as some secondary and tertiary amines. The monoamine oxidases

from Homo sapiens are located intracellularly in the outer membrane of mitochondria

and comprise two isoforms, MAO-A and MAO-B. MAO-A preferentially oxidizes serotonin

and norepinephrine,[12] while MAO-B has a higher affinity for phenylethylamine and

benzylamine.[13] MAO-A and MAO-B are co-expressed in most human tissues and are mostly

found in the intestines.[14] However, fibroblasts and placenta express MAO-A exclusively.[15]

On the other hand, platelets and lymphocytes only express MAO-B.[16] In the human brain

MAO-A is confined to catecholaminergic neurons whereas MAO-B is found in serotonergic

neurons and astrocytes.[17] Although MAO-A and MAO-B are distinctly different in substrate

preference and tissue and cell distribution, they share a covalently bound flavin and 70%

sequence identity. [18]

MAO-A and MAO-B started to attract the attention from the biochemical and pharmaceutical

communities from the 1950s when Zeller and coworkers showed that these enzymes served

as targets for antidepressant drugs.[19] Since those findings more than 14000 papers have

been published relating to MAO inhibitors and structural and functional properties of MAO.

MAO dysfunction (high or low levels of MAO) is thought to be responsible for a number of

psychiatric and neurological disorders. For example, unusually high or low levels of MAOs

in the body have been associated with depression,[20] schizophrenia,[21] attention deficit

disorder (ADD),[22] and migraines.[23] Actually, MAO-A inhibitors (e.g. moclobemide) act as

antidepressant and antianxiety agents (Table 1), while MAO-B inhibitors (e.g. selegiline,

rasagiline) are used alone or in combination to treat Alzheimer’s and Parkinson’s diseases

(Table 1).[24] Also, studies have shown that MAO-B is heavily depleted (40% decrease) by use

of tobacco cigarettes.[25] In fact, MAO inhibitors have recently been reported to facilitate

stopping smoking in highly dependent smokers.[26]

Page 46: Biocatalysis & Multicomponent Reactions: The Ideal …...Biocatalysis & Multicomponent Reactions Chapter 2 Monoamine Oxidase N: 41 A Promising Biocatalyst for Asymmetric Synthesis

Chapter 2

46

Table 1: Clinically used specific inhibitors of MAO-A and MAO-B

Compound MAO Selectivity Application

Moclobemide

A Antidepressant

Befloxatone

A Antidepressant

Clorgyline

A Antidepressant

Selegiline

B Antiparkinson

Rasagiline

B Antiparkinson

Lazabemide

B Antiparkinson

2.2.2 Monoamine Oxidase N

Amine oxidases are also found in lower eukaryotes. In the fungus Apsergillus niger, two types

of amine oxidase have been described. Type I belongs to the quinoprotein family,[27] while

Type II is a flavin-dependent MAO (designated MAO-N).[28] Unlike the mammalian MAOs,

flavin-dependent MAO-N is soluble, peroxisomal, readily extractable, could be isolated

extensively by conventional procedures and showed no tendency to aggregate. Most

interestingly, the flavin cofactor FAD is non-covalently linked, in contrast to MAO-A and -B

from mammalians. Despite the lack of a pentapeptide sequence which covalently anchors

FAD to the mammalian MAO, MAO-N shows a significant sequence similarity (49%) and

sequence identity (24%) with human MAO-A and -B. Also, the substrate specificity of MAO-N

in some respects resembles that of MAO-A and in others that of MAO-B, while it also exhibits

some unique properties.[28] All these observations have led to the suggestion that MAO-N

may be an evolutionary precursor of vertebrate MAOs.

Page 47: Biocatalysis & Multicomponent Reactions: The Ideal …...Biocatalysis & Multicomponent Reactions Chapter 2 Monoamine Oxidase N: 41 A Promising Biocatalyst for Asymmetric Synthesis

Monoamine Oxidase N

47

2.3 Directed Evolution of Monoamine Oxidase N

As shown previously, amine oxidases have been classified into two groups, i.e., Type I

(quinoprotein family) and Type II (flavin-dependent family). In the catalytic cycle, Type II

MAO generate “free” imines (scheme 2, ii), while in Type I MAO the imines stay covalently

bound to the protein (Scheme 1). This has led to the conclusion that Type II MAOs were more

suitable for generating optically active imines.

The possibility of cloning and expressing of MAO-N from Aspergillus niger resulted in

substrate-specificity and kinetic studies. These studies showed MAO-N to have high activity

towards simple aliphatic amines but a much lower activity towards benzylamine.[28-29]

Turner and coworkers chose α-methylbenzylamine as a feasible starting point substrate

for the directed evolution of MAO-N due to the importance of this substrate as chiral

amine[30] but also because L-α-methylbenzylamine although possessing very low activity

for MAO-N, had a significantly higher activity compared to D-α-methylbenzylamine.[31]

First, an effective high-throughput assay was developed were the hydrogen peroxide that is

liberated during MAO-N oxidation is oxidized by a peroxidase in the presence of a substrate

(3,3’-diaminobenzidine) that yields dark products (Figure 1).

Figure 1: High-throughput screening of libraries of MAO-N

A library of MAO-N mutants was generated by random mutagenesis. By using the

colorimetric assay described in Figure 1, Turner and coworkers were able to identify clones

that possessed higher activity towards L-α-methylbenzylamine compared to wild-type

MAO-N. Eventually, a single mutant was identified which showed a 47-fold higher activity

towards L -α-methylbenzylamine than the wild-type. Also, the selectivity towards L-α-

methylbenzylamine versus D -α-methylbenzylamine was increased 5.8-fold relative to the

wild-type.[31] This mutant showed to have a single point mutation where asparagine 336

is replaced by serine (Asn336Ser). Subsequently, an additional mutation was introduced

(Met348Lys) resulting in higher specific activity and expression levels but leaving the

substrate specificity unchanged.[32] This enzyme with two mutations showed the highest

activity towards substrates (S enantiomer) containing a primary amine group flanked by

a methyl group and a bulky alkyl/aryl group, for which the wild-type MAO-N showed no or

very poor activity (Scheme 3).[32]

Page 48: Biocatalysis & Multicomponent Reactions: The Ideal …...Biocatalysis & Multicomponent Reactions Chapter 2 Monoamine Oxidase N: 41 A Promising Biocatalyst for Asymmetric Synthesis

Chapter 2

48

Scheme 3: Schematic overview of substrate specificity of wild-type and mutant MAO-N.

Because the activity towards oxidation of secondary amines was still very poor, Turner and

coworkers performed a new round of random mutagenesis to identify mutants capable of

oxidizing secondary amines.[33] The starting point for the second round was a MAO-N mutant

with four point mutations, i.e., Asn336Ser, Met348Lys (Scheme 3), Arg259Lys and Arg260Lys

(improved expression).

This mutant was subjected to a new round of random mutagenesis with

1-methyltetrahydroisoquinoline as the substrate. This resulted in a clone which showed

significantly more activity (5-fold) than the parent. Interestingly, this mutant also showed

enhanced activity and very high (S)-selectivity (E>100) towards a wide variety of secondary

amines compared to the parent mutant. Analysis showed that this mutant possessed an

additional point mutation (Ile246Met) and it was termed MAO-N-D3.[33]

Turner and coworkers were also interested in mutants capable of oxidizing tertiary amines

as these are difficult to obtain in enantiomerically pure form by alternative methods.[34-35]

Another mutant, with an additional two point mutations from MAO-N-D3 (Thr384Asn and

Asp385Ser, mutant designated MAO-N-D5) showed activity towards tertiary amines whereas

the wild-type and MAO-N-D3 did not show any activity.[36-37] Recently, the structure of MAO-N

was solved using X-ray crystallography and data showed that the Asn336Ser mutation

reduces the steric bulk at the active site, Ile246Met gives more flexibility to the substrate

binding site, and Thr384Asn and Asp385Ser seem to influence the tertiary structure of the

enzyme resulting in change in the active-site environment.[37]

This greatly improved catalyst shows that random mutagenesis with no prior knowledge

of the structure is a powerful tool for the improvement of enzymes to achieve specific

requirements (e.g. improved activity and selectivity) to obtain a highly attractive biocatalyst.

Page 49: Biocatalysis & Multicomponent Reactions: The Ideal …...Biocatalysis & Multicomponent Reactions Chapter 2 Monoamine Oxidase N: 41 A Promising Biocatalyst for Asymmetric Synthesis

Monoamine Oxidase N

49

2.4 Application of Monoamine Oxidase N

Optically pure chiral amines have found widespread use in the agrochemical, pharmaceutical

and chemical industries. Consequently, there is a need for efficient methods to obtain these

compounds.

Turner and coworkers designed an ingenious catalytic method for the preparation of

chiral amines by deracemization of the corresponding racemic mixture (Scheme 4). The

deracemization approach depends upon coupling of an enantioselective amine oxidase

with a nonselective reducing agent (ammonia borane) where stereoinversion from the S

to the R enantiomer via the achiral imine is achieved.[32-33, 36] Mutant monoamine oxidase N

(Asn336Ser/Met348Lys) was used to generate a range of enantiomerically enriched primary

R amines (substrate specificity is depicted in Scheme 3) in high yields and ee’s (scheme 4A,

R3 = H).[32] A second-generation biocatalysts, MAO-N-D3, was used to produce a series of

enantiomerically enriched secondary R amines in high yields and ee’s (scheme 4A, R3 ≠ H).[33]

MAON mutant

NH3� BH3

MAOND5 mutant

NH3� BH

3

A B

NH3 � BH

3

NH3 � BH

3

+

+

+

+

Scheme 4: A. Deracemization of primary and secondary amines using mutant amine oxidases in combination with ammonia-borane. B. Deracemization of tertiary amines using mutant amine

oxidases in combination with ammonia-borane.

In addition, MAO-N-D5 was used to obtain enantiomerically pure cyclic tertiary amines

(scheme 4B) in good yields and excellent ee’s.[36] For example, the complete deracemization

of N-methyl-2-phenylpyrrolidine was achieved within 24 h and yielded (R)-N-methyl-2-

phenylpyrrolidine in 75% isolated yield and an ee of 99%.

Page 50: Biocatalysis & Multicomponent Reactions: The Ideal …...Biocatalysis & Multicomponent Reactions Chapter 2 Monoamine Oxidase N: 41 A Promising Biocatalyst for Asymmetric Synthesis

Chapter 2

50

NH3 � BH

3

()nicotine

Pseudooxynicotine

nicotine

MAOND5

NH3� BH 3

Scheme 5: Synthesis of (R)-nicotine from pseudooxynicotine via reductive amination.

This practical procedure is also useful as a means of enantioselective intramolecular

reductive amination. Exemplary is the synthesis of (R)-nicotine from pseudooxynicotine

(Scheme 5). As pseudooxynicotine predominantly exists as iminium ion rather than the

enamine or hemiaminal in aqueous solution, it would be expected to be reduced much

faster than the pseudooxynicotine and enamine and enter into the deracemization cycle.

As expected, (R)-nicotine was formed after 24 h with no other detectable species in 99% ee.

Moreover, MAO-N-D5 also exhibits high activity towards a wide range of substituted meso-

pyrrolidines (a selection is presented in Figure 2) resulting in a method for the preparation

of substituted 1-pyrrolines in high enantiomeric excess.[38]

Page 51: Biocatalysis & Multicomponent Reactions: The Ideal …...Biocatalysis & Multicomponent Reactions Chapter 2 Monoamine Oxidase N: 41 A Promising Biocatalyst for Asymmetric Synthesis

Monoamine Oxidase N

51

Figure 2: A. Meso-pyrrolidines used as substrates for amine oxidation by MAO-N-D5. B. 1-pyrrolines resulting from oxidation with MAO-N-D5.

Unfortunately, not all imines (e.g. 1, 3 and 4) were readily isolated and had to be derivatized

in situ to the trifluoroacetamide derivative from the corresponding amino nitrile, presumably

due to decomposition of the corresponding imines. The applicability of these substituted

1-pyrrolines was demonstrated by addition of HCN to these imines to yield the corresponding

amino nitriles (Scheme 6). Subsequent hydrolysis of the amino nitrile furnished the free

amino acid in 94 % ee and 51% overall yield. Finally, the absolute configuration of the imines

resulting from the oxidation by MAO-N-D5 was established. By reacting D,L-8 with D-amino

acid oxidase and observing oxidation of D-8 assigned the product derived from the MAO-

N-D5 oxidation as L-8.

Scheme 6: Diastereoselective synthesis of amino acid derivative 8.

2.5 Conclusion

Directed evolution can optimize specific requirements like substrate specificity and selectivity

of enzymes in a highly effective way. Furthermore, the resulting mutants often possess a

broader substrate specificity compared to the wild-type enzyme. Monoamine oxidase N from

Aspergillus niger has proven particularly suitable for directed evolution. Mutant monoamine

oxidases N tolerate primary, secondary, and tertiary amines as substrates. In addition, 3,4

substituted meso-pyrrolidines were shown to be excellent substrates for oxidation. Moreover,

the stereoselectivity of the biocatalyst is very high (ee > 94%) and no additional cofactor is

required. This makes MAO-N one of the leading biocatalysts for amine oxidation.

Page 52: Biocatalysis & Multicomponent Reactions: The Ideal …...Biocatalysis & Multicomponent Reactions Chapter 2 Monoamine Oxidase N: 41 A Promising Biocatalyst for Asymmetric Synthesis

Chapter 2

52

2.5 References & Notes[1] T. Thomas and T. J. Thomas, Cell. Mol. Life Sci. 2001, 58, 244-258.[2] A. E. Pegg, Cancer Res. 1988, 48, 759-774.[3] N. Seiler, C. L. Atanassov and F. Raul, Int. J. Oncol. 1998, 13, 993-1006.[4] D. E. Edmondson, A. Mattevi, C. Binda, M. Li and F. Hubalek, Curr. Med. Chem. 2004, 11, 1983-1993.[5] W. Weyler, Y. P. P. Hsu and X. O. Breakefield, Pharmacol. Ther. 1990, 47, 391-417.[6] S. X. Wang, M. Mure, K. F. Medzihradszky, A. L. Burlingame, D. E. Brown, D. M. Dooley, A. J. Smith,

H. M. Kagan and J. P. Klinman, Science 1996, 273, 1078-1084.[7] M. Mure, S. A. Mills and J. P. Klinman, Biochemistry 2002, 41, 9269-9278.[8] R. J. Rohlfs and R. Hille, J. Biol. Chem. 1994, 269, 30869-30879.[9] R. B. Silverman, Prog. Brain. Res. 1995, 106, 23-31.[10] J. M. Kim, M. A. Bogdan and P. S. Mariano, J. Am. Chem. Soc. 1993, 115, 10591-10595.[11] M. L. C. Hare, Biochem. J. 1928, 22, 968-979.[12] P. A. Shore, J. A. R. Mead, R. G. Kuntzman, S. Spector and B. B. Brodie, Science 1957, 126, 1063-1064.[13] Y. Ochiai, K. Itoh, E. Sakurai, M. Adachi and Y. Tanaka, Biol. Pharm. Bull. 2006, 29, 2362-2366.[14] P. Levitt, J. E. Pintar and X. O. Breakefield, Proc. Natl. Acad. Sci. U. S. A. Biol. Sci. 1982, 79, 6385-6389.[15] Q. S. Zhu, K. Chen and J. C. Shih, J. Neurosci. 1994, 14, 7393-7403.[16] W. K. Wong, K. Chen and J. C. Shih, J. Biol. Chem. 2003, 278, 36227-36235.[17] W. K. Wong, X. M. Ou, K. Chen and J. C. Shih, J. Biol. Chem. 2002, 277, 22222-22230.[18] A. W. J. Bach, N. C. Lan, D. L. Johnson, C. W. Abell, M. E. Bembenek, S. W. Kwan, P. H. Seeburg and J.

C. Shih, Proc. Natl. Acad. Sci. U. S. A. 1988, 85, 4934-4938.[19] J. Barsky, W. L. Pacha, S. Sarkar and E. A. Zeller, J. Biol. Chem. 1959, 234, 389-391.[20] J. H. Meyer, N. Ginovart, A. Boovariwala, S. Sagrati, D. Hussey, A. Garcia, T. Young, N. Praschak-

Rieder, A. A. Wilson and S. Houle, Arch. Gen. Psychiatry 2006, 63, 1209-1216.[21] J. J. Schildkraut, J. M. Herzog, P. J. Orsulak, S. E. Edelman, H. M. Shein and S. H. Frazier, Am. J.

Psychiatry 1976, 133, 438-440.[22] S. D. Jiang, R. Xin, S. C. Lin, Y. P. Qian, G. M. Tang, D. X. Wang and X. D. Wu, Am. J. Med. Genet. 2001,

105, 783-788.[23] O. Meienberg and F. Amsler, Praxis (Bern 1994) 1997, 86, 1107-1112.[24] P. Riederer, L. Lachenmayer and G. Laux, Curr. Med. Chem. 2004, 11, 2033-2043.[25] J. S. Fowler, N. D. Volkow, G. J. Wang, N. Pappas, J. Logan, R. MacGregor, D. Alexoff, C. Shea, D.

Schlyer, A. P. Wolf, D. Warner, I. Zezulkova and R. Cilento, Nature 1996, 379, 733-736.[26] I. Berlin, S. Said, O. Spreuxvaroquaux, J. M. Launay, R. Olivares, V. Millet, Y. Lecrubier and A. J.

Puech, Clin. Pharmacol. Ther. 1995, 58, 444-452.[27] H. Yamada, O. Adachi and K. Ogata, Agric. Biol. Chem. 1965, 29, 117-123.[28] B. Schilling and K. Lerch, Biochim. Biophys. Acta 1995, 1243, 529-537.[29] S. O. Sablin, V. Yankovskaya, S. Bernard, C. N. Cronin and T. P. Singer, Eur. J. Biochem. 1998, 253,

270-279.[30] G. Hieber and K. Ditrich, Chim. Oggi 2001, 19, 16-20.[31] M. Alexeeva, A. Enright, M. J. Dawson, M. Mahmoudian and N. J. Turner, Angew. Chem. Int. Ed.

2002, 41, 3177-3180.[32] R. Carr, M. Alexeeva, A. Enright, T. S. C. Eve, M. J. Dawson and N. J. Turner, Angew. Chem. Int. Ed.

2003, 42, 4807-4810.[33] R. Carr, M. Alexeeva, M. J. Dawson, V. Gotor-Fernandez, C. E. Humphrey and N. J. Turner,

ChemBioChem 2005, 6, 637-639.[34] N. E. Lee and S. L. Buchwald, J. Am. Chem. Soc. 1994, 116, 5985-5986.[35] S. Miyano, L. D. L. Lu, S. M. Viti and K. B. Sharpless, J. Org. Chem. 1985, 50, 4350-4360.[36] C. J. Dunsmore, R. Carr, T. Fleming and N. J. Turner, J. Am. Chem. Soc. 2006, 128, 2224-2225.[37] K. E. Atkin, R. Reiss, V. Koehler, K. R. Bailey, S. Hart, J. P. Turkenburg, N. J. Turner, A. M. Brzozowski

and G. Grogan, J. Mol. Biol. 2008, 384, 1218-1231.[38] V. Kohler, K. R. Bailey, A. Znabet, J. Raftery, M. Helliwell and N. J. Turner, Angew. Chem. Int. Ed. 2010,

49, 2182-2184.

Page 53: Biocatalysis & Multicomponent Reactions: The Ideal …...Biocatalysis & Multicomponent Reactions Chapter 2 Monoamine Oxidase N: 41 A Promising Biocatalyst for Asymmetric Synthesis

Highly Stereoselective Synthesis of Substituted Prolyl Peptides Using

a Combination of Biocatalytic Desymmetrization and

Multicomponent Reactions

Published in: Angew. Chem. 2010, 122, 5417-5420;

Angew. Chem. Int. Ed. 2010, 49, 5289-5292

Chapter 3

Page 54: Biocatalysis & Multicomponent Reactions: The Ideal …...Biocatalysis & Multicomponent Reactions Chapter 2 Monoamine Oxidase N: 41 A Promising Biocatalyst for Asymmetric Synthesis

Chapter 3

54

Abstract: Optically pure 3,4-disubstituted 1-pyrrolines, generated from the corresponding

meso-pyrrolidines by biocatalytic desymmetrization (MAO-N=monoamine oxidase N),

react with carboxylic acids and isocyanides in a highly diastereoselective Ugi-type

multicomponent reaction to give pharmaceutically relevant substituted prolyl peptides

that would require lengthy reaction sequences using other methods. The resulting prolyl

peptides can also be used as highly active organocatalysts for stereoselective conjugate

addition reactions.

Page 55: Biocatalysis & Multicomponent Reactions: The Ideal …...Biocatalysis & Multicomponent Reactions Chapter 2 Monoamine Oxidase N: 41 A Promising Biocatalyst for Asymmetric Synthesis

Highly Stereoselective Synthesis of Substituted Prolyl Peptides

55

3.1 Introduction

3.1.1 Biocatalytic amine oxidation combined with Ugi chemistry

Multicomponent reactions (MCRs) offer the ability to rapidly and efficiently generate

collections of structurally and functionally diverse organic compounds.[1-5] MCRs are

important tools for both combinatorial chemistry and diversity-oriented synthesis, and thus

play a significant role in the development of synthetic methodology for drug discovery.[6, 7]

Although MCRs are very efficient by their nature, the stereocontrol in these reactions is mostly

not trivial.[8-13] For most MCRs, catalytic asymmetric methods to control the stereochemical

outcome of the reaction are so far not available. The Ugi reaction is undoubtedly one of the

most widely applied MCRs.[14] It is of considerable interest owing to its exceptional synthetic

efficiency and is widely used in the fields of modern combinatorial and medicinal chemistry.[1-7]

The Ugi reaction involves a one-pot condensation of an aldehyde (1), an amine (2), a

carboxylic acid (3) and an isocyanide (4) to produce chiral α-acylaminoamides 7 (Scheme 1).

However, as in most MCRs, controlling the newly formed stereocenter is not straightforward.

Scheme 1: Proposed mechanism of the Ugi four-component reaction (U-4CR).

In 1982, Nutt and Joullié reported the use of an Ugi-type three-component reaction (U-3CR)

that employed substituted 1-pyrrolines 10 instead of the amine and aldehyde components

to produce substituted prolyl peptides 13 (scheme 2).[15] Commercially available

3-hydroxypyrrolidine 8 was converted into the corresponding aryl ether 9 via a Mitsunobu

reaction. After deprotection of the Boc group, the N-chloro derivative was synthesized, which

was dehydrohalogenated to form the imine species (10). The authors did not comment on

the regioselectivity of the formed imine. Subsequently, a U-3CR was performed leading to

the desired substituted 1-pyrrolines 13.

Scheme 2. First application of substituted 1-pyrrolines as input for U-3CR.

Page 56: Biocatalysis & Multicomponent Reactions: The Ideal …...Biocatalysis & Multicomponent Reactions Chapter 2 Monoamine Oxidase N: 41 A Promising Biocatalyst for Asymmetric Synthesis

Chapter 3

56

In this case and in later applications,[16-23] the (dia)stereoselectivities were poor or

unpredictable at best, and the routes to the required substituted 1-pyrrolines were tedious

and/or low-yielding. A representative overview of the compounds (14-18) prepared in this

way is given in Figure 1.

Figure 1: U-3CR with substituted 1-pyrrolines. OP = protected alcohol.

Recently, Turner and co-workers reported the biocatalytic desymmetrization of

3,4-substituted meso-pyrrolidines with engineered monoamine oxidase N (MAO-N) from

Aspergillus niger[24, 25] to yield optically active 1-pyrrolines in excellent yields and ee values.[26, 27]

As imines are intermediates for many common multicomponent reactions, the use of

these optically active 1-pyrrolines in a MCR would be highly attractive given the excellent

diastereoselectivity that can be achieved from the addition of nucleophiles to the imine,

owing to its steric bulk (Scheme 3).

Scheme 3: Pyrrolidine derivatives accessible via amine oxidation/MCR strategy.

Page 57: Biocatalysis & Multicomponent Reactions: The Ideal …...Biocatalysis & Multicomponent Reactions Chapter 2 Monoamine Oxidase N: 41 A Promising Biocatalyst for Asymmetric Synthesis

Highly Stereoselective Synthesis of Substituted Prolyl Peptides

57

We envisioned that cyclic imines would give excellent selectivity in the Ugi-type reaction

previously described by Nutt and Joullié due to their steric bulk. Also, these 3,4-substituted

prolyl peptides, with generic structure 22, are of considerable interest to applications in

organocatalysis and medicinal chemistry.

3.1.2 3,4-Disubstituted bisamide prolyl peptides in medicinal chemistry

Compounds with a cis-3,4-substituted bisamide pyrrolidine scaffold (e.g. compounds with

generic structure 22) show a broad range of interesting biological activities. In addition, this

scaffold is quite common in several classes of natural products.

For example, the echinocandins (Figure 2A), are a well-known class of 21-membered cyclic

lipopeptides comprising a cis-4-methyl-3-hydroxy-pyrrolidine moiety. These macrocycles

are isolated from Aspergillus rugulosus and Aspergillus nidulans and are characterized by their

potent antifungal activity[28-30] as they effectively disrupt fungal cell wall development.

Further, the astins A–C (Figure 2B), are a family of cyclopentapeptides isolated from the roots

of Aster tataricus (Compositae) thatare characterized by a 16-membered ring containing a

distinctive cis-3,4-dichloro- pyrrolidine moiety. These astins exhibit potent antitumour

activity both in vitro and in vivo.[31-33] Noteworthy, astins A-C are significantly more potent

than astins D-I which lack the dichloroproline showing the importance of the substituents

on the proline ring.

Also a designed fucodipeptide containing a cis-3,4-dihydroxy-pyrrolidine ring (Figure 2C)

demonstrates excellent inhibitory effects on the binding of sialyl Lewis X (an essential

carbohydrate for activity) to E-selectin.[34] Previous studies have shown that inhibition of

this E-selectin result in anti-inflammatory activity in certain diseases such as reperfusion

injury.[35-37]

Page 58: Biocatalysis & Multicomponent Reactions: The Ideal …...Biocatalysis & Multicomponent Reactions Chapter 2 Monoamine Oxidase N: 41 A Promising Biocatalyst for Asymmetric Synthesis

Chapter 3

58

Figure 2: 3,4-subsituted bisamide pyrrolidines showing a broad range of biological activity. A. Echinocandins. B. Astins. C. Fucopeptide.

Finally, these 3,4-substituted prolyl peptides are also key structural elements of Telaprevir[38,

39] (Vertex and Johnson & Johnson) and Boceprevir[40, 41] (Merck & Co.), which inhibit the

hepatitis C virus NS3 protease (Figure 3). Comprising a bicyclic proline core, these HCV

protease inhibitors have passed clinical trials and have been approved for the treatment

of HCV disease. These drugs are currently on the market under the names Incivek™ and

Victrelis™.

Figure 3: HCV NS3 protease inhibitors.

3.1.3 Bisamide pyrrroldines in organocatalysis

Organocatalysis has become an efficient tool to complement transition-metal complexes

and enzymes for the synthesis of chiral building blocks. Recently, Wennemers and co-

Page 59: Biocatalysis & Multicomponent Reactions: The Ideal …...Biocatalysis & Multicomponent Reactions Chapter 2 Monoamine Oxidase N: 41 A Promising Biocatalyst for Asymmetric Synthesis

Highly Stereoselective Synthesis of Substituted Prolyl Peptides

59

workers showed that small peptides can be used as very efficient catalysts. The H-D-Pro-

Pro-Glu-NH2 tripeptide 27 showed to be a very efficient catalyst in 1,4-addition reactions of

aldehydes 25 to nitro-olefins 26 (Scheme 4).[42-44]

Scheme 4: Conjugate addition reactions catalyzed by tripeptide H-D-Pro-Pro-Glu-NH2.

The generally accepted catalytic cycle of this reaction involves enamine formation (B)

followed by reaction with the nitro-olefin and hydrolysis of the resulting imine (C) which

generates the desired nitro-aldehyde and recycles the catalyst (scheme 5).

Scheme 5: Generally accepted catalytic cycle for the conjugate addition of aldehydes to nitro-olefins.

This H-D-Pro-Pro-Glu-NH2 tripeptide contains the same generic structure as these

3,4-substituted prolyl peptides 22 which are accessible via the MAO-N/U-3CR sequence

making this combination of biocatalysis and multicomponent chemistry very interesting

not only in the synthesis of biological interesting compounds but also in the synthesis of

potential organocatalysts.

Page 60: Biocatalysis & Multicomponent Reactions: The Ideal …...Biocatalysis & Multicomponent Reactions Chapter 2 Monoamine Oxidase N: 41 A Promising Biocatalyst for Asymmetric Synthesis

Chapter 3

60

3.2 Results & Discussion

3.2.1 Solvent study

First, we turned our attention to finding the most suitable conditions for the Ugi-type

MCR. As methanol is usually the solvent of choice in the Ugi reaction, we decided to do a

solvent screen to determine if there would be any solvent effect on the diastereomeric ratio

(d.r.). The reaction of racemic 3-azabicyclo[3.3.0]oct-2-ene (rac-29, synthesized according

to a literature procedure),[45] benzoic acid, and tert-butyl isocyanide was selected as the

model reaction. Various solvents were screened at room temperature (Table 1). Based on

GC analysis, reactions performed in dichloromethane (Figure 4) and toluene gave the best

diastereomeric ratios. In these solvents the influence of the temperature on the d.r. of the

U-3CR was studied. Table 1 shows that dichloromethane at 4 °C gave the best diastereomeric

ratio. The yields were comparable in dichloromethane and methanol in contrast to toluene

where the yields of the U-3CR were lower (data not presented). Finally, we performed the

MCR at -40 °C, but to our surprise the diastereomeric ratio decreased. Because of the only

marginal improvement in diastereomeric ratio at 4 °C we decided to perform our MAO-N/

MCR sequence in dichloromethane at room temperature.

Table 1. Solvent and temperature dependence of the d.r. of the U-3CR of rac-29, benzoic acid, and tert-butyl isocyanide.[a]

Entry Solvent d.r. [b] (RT) d.r. (4 °C) d.r. (-40 °C)

1 H2O 87:13 ---[c] ---[c]

2 buffer[d] 87:13 ---[c] ---[c]

3 MeOH 90:10 92:8 91:9

4 CH2Cl2 92:8 93:7 90:10

5 DMSO 87:13 ---[c] ---[c]

6 DMF 89:11 ---[c] ---[c]

7 Toluene 92:8 93:7 ---[e]

8 TFE 90:10 ---[c] ---[c]

[a] All reactions were performed with 0.073 mmol of imine rac-29 and 0.1 mmol of benzoic acid and tert-butyl isocyanide. Reactions were stirred for 24 h at the appropriate temperature. [b] Based on GC analysis. [c] Not tested. [d] 100 mM KPO4 buffer, pH = 8.0. [e] Reaction too slow for accurate determination.

Page 61: Biocatalysis & Multicomponent Reactions: The Ideal …...Biocatalysis & Multicomponent Reactions Chapter 2 Monoamine Oxidase N: 41 A Promising Biocatalyst for Asymmetric Synthesis

Highly Stereoselective Synthesis of Substituted Prolyl Peptides

61

min

pA

Figure 4: GC-FID trace of the U-3CR of rac-29, benzoic acid and tert-butyl isocyanide in CH2Cl2. Diastereomers are depicted as peak 1 and 2.

3.2.2 Two-pot MAO-N/ U-3CR sequence

Enantiomerically enriched cyclic imine (3S,7R)-29 was prepared by MAO-N catalyzed

desymmetrization of the corresponding commercially available pyrrolidine derivative[27, 46] in

very good yield and ee (85%, 94 % ee). The ee could be improved to 97% by recrystallization

during workup.

With the chiral imine 29 in hand, we turned our attention to the Ugi-type 3CR. Different

carboxylic acids and isocyanides were used to generate substituted prolyl peptides 30a-g

in good yield and diastereoselectivity with very good ee (Table 2). Notably, as the ee’s of

the desired compounds were >94 % we could conclude that no epimerization has occured

during the U-3CR. In illustration of these high ee’s, a chiral HPLC trace (95% ee) of the

benchmark reaction (compound 30b) is shown in Figure 5.

Page 62: Biocatalysis & Multicomponent Reactions: The Ideal …...Biocatalysis & Multicomponent Reactions Chapter 2 Monoamine Oxidase N: 41 A Promising Biocatalyst for Asymmetric Synthesis

Chapter 3

62

Table 2. Scope of U-3CR using optically enriched 29.

Entry Product R1 R2 t (h) Yield (%) d.r.[a] ee (%)[b]

1 30a Me t-Bu 48 73 93:7 95[c]

2 30b Ph t-Bu 24 80 93:7 94

3 30c 2-Furyl i-Pr 48 75 92:8 94

4 30d Ph i-Pr 24 78 92:8 94

5 30e Me Bn 48 71 92:8 94

6 30f Ph Bn 24 81 92:8 97[c]

7 30g i-Pr t-Bu 48 83 93:7 97[c]

[a] d.r. determined by GC analysis. [b] ee determined by HPLC and GC analysis. [c] Partial crystallization of imine.

Figure 5: HPLC trace of prolyl peptide 30b with an ee of 95%. Enantiomers of the major diastereomer are depicted as peak 1 and 2.

Crystallographic analysis of 30f (Figure 6) allowed unambiguous assignment of the

absolute configuration (as the stereochemistry at the C3 and C4 positions, resulting from

the biotransformation, has been reported previously[26]), and showed that attack by the

isocyanide occured from the sterically less hindered face. The 2,3-trans relationship is

in agreement with the generally accepted mechanism of the Ugi reaction (scheme 1), in

which the stereodetermining step is the direct nucleophilic attack of the isocyanide on the

imine (or iminium) carbon. The extraordinarily high selectivity for the 2,3-trans isomer is in

sharp contrast with other reports, where stereoinduction is poor[16-20, 23] or the 2,3-cis isomer

is preferentially formed.[21-22] All other pyrrolidines 30 were assigned the same absolute

stereochemistry as 30f, based on analogy of the 1H NMR spectroscopic data.

Page 63: Biocatalysis & Multicomponent Reactions: The Ideal …...Biocatalysis & Multicomponent Reactions Chapter 2 Monoamine Oxidase N: 41 A Promising Biocatalyst for Asymmetric Synthesis

Highly Stereoselective Synthesis of Substituted Prolyl Peptides

63

Figure 6. Molecular structure of 30f in the crystal. Displacement ellipsoids are drawn at 50% probability.

Subsequently, the sterically demanding imine 35 was prepared by MAO-N-catalyzed

desymmetrization in 84% yield and >99% ee,[46] (Scheme 6). First cyclopentadiene 31 and

maleimide 32 underwent a Diels-Alder reaction forming imide 33. Subsequent reduction

with lithium aluminum hydride furnished the sterically demanding amine 34 which was

used as substrate for the biotransformation with MAO-N and the corresponding imine (35)

was applied in a series of Ugi-type 3CRs. To our delight, substituted prolyl peptides 36 were

obtained as single diastereomers in >99% ee (Table 3, entries 1–8).

Scheme 6: Synthesis of bridged imine 35.

Page 64: Biocatalysis & Multicomponent Reactions: The Ideal …...Biocatalysis & Multicomponent Reactions Chapter 2 Monoamine Oxidase N: 41 A Promising Biocatalyst for Asymmetric Synthesis

Chapter 3

64

Table 3. Scope of U-3CR using optically pure 35.

Entry Product R1 R2 t (h) Yield (%) d.r.[a] ee (%)[b]

1 36a Me t-Bu 48 83 >99:1 >99

2 36b Ph t-Bu 24 82 >99:1 >99

3 36c Furyl i-Pr 48 75 >99:1 >99

4 36d Ph i-Pr 24 78 >99:1 >99

5 36e Me Bn 48 78 >99:1 >99

6 36f Ph Bn 24 80 >99:1 >99

7 36g i-Pr t-Bu 48 81 >99:1 >99

[a] d.r. determined by GC analysis. [b] ee determined by HPLC and GC analysis.

To confirm that the stereochemical outcome of the reaction was solely determined by

the starting chiral imine, we reacted 35 with tert-butyl isocyanide and either Fmoc-L-Pro-

OH or Fmoc-D-Pro-OH to give 36i and 36j, respectively (Figure 7). In both cases, only one

diastereomer was formed.[47] Likewise, we reacted 35 with benzoic acid and either (S)- or (R)-

α-methylbenzyl isocyanide[48] to give 36k and 36l, respectively, as single diastereomers. 1H

NMR analysis indicated that the shown 2,3-trans isomers were selectively formed in all cases.

Figure 7:. Substituted prolyl peptides from optically pure acid or isocyanide inputs.

3.2.3 One-pot MAO-N/U-3CR sequence

With the MAO-N/U-3CR sequence established, attention was focused on combining it in a

one-pot procedure. Earlier results (Table 1, entry 2) show that the U-3CR can be performed

under the conditions of the enzymatic desymmetrization.

First we executed the enzymatic desymmetrization to give (3S,7R)-29. The crude suspension

was then centrifuged to discard the cells and benzoic acid and tert-butyl isocyanide were

Page 65: Biocatalysis & Multicomponent Reactions: The Ideal …...Biocatalysis & Multicomponent Reactions Chapter 2 Monoamine Oxidase N: 41 A Promising Biocatalyst for Asymmetric Synthesis

Highly Stereoselective Synthesis of Substituted Prolyl Peptides

65

added to the supernatant (Scheme 7). After full conversion (no imine was present any more

based on GC analysis) a disappointing isolated yield for 30b of 19% was obtained with a d.r.

of 88:12. This low yield could be due to the stability and solubility of the isocyanide under

aqueous conditions. Therefore we decided to perform the subsequent U-3CR in a biphasic

system. So, after discarding the cells, toluene was added to the supernatant followed by

benzoic acid and tert-butyl isocyanide. Unfortunately, after full conversion of imine (3S,7R)-

29 still rather low yields of 30b were observed (28% based on GC analysis). These low yields

could also be caused by the pH of the reaction mixture. The proposed mechanism of the

Ugi reaction involves a protonation of the imine by the carboxylic acid (Scheme 1). This

reactive iminium species undergoes α-addition by the carbenoid carbon of the isocyanide

to finally rearrange into the desired α-acylaminoamide. However, the pH of the reaction is

8, implicating that most of the imine would not be protonated resulting in low conversion.

Scheme 7: One-pot MAO-N/U-3CR sequence.

3.2.4 3,4-substituted prolyl peptides as organocatalyst

Finally, we reacted imine 35 with Fmoc-D-Pro-OH 37 and methyl 3-isocyanopropionate 38,

which, after treatment with NaOH in methanol/dichloromethane,[49] afforded 40 as a single

stereoisomer (Scheme 8). Compound 40 strongly resembles H-D-Pro-Pro-Glu-NH2 (27),

which was described by Wennemers and co-workers (Scheme 4).

+

39

40

37 38

35

Scheme 8: Synthesis of prolyl peptide 40 as organocatalyst.

Page 66: Biocatalysis & Multicomponent Reactions: The Ideal …...Biocatalysis & Multicomponent Reactions Chapter 2 Monoamine Oxidase N: 41 A Promising Biocatalyst for Asymmetric Synthesis

Chapter 3

66

To our delight, peptide 40 catalyzed the reaction between propanal and nitrostyrene to give

43 (Scheme 9) in 91% yield, 87:13 syn/anti ratio, and 86% ee (compared to 90:10 syn/anti

and 91% ee using H-D-Pro-Pro-Glu-NH2 tripeptide[44]). Thus, our MAO-MCR sequence allows

efficient asymmetric synthesis of proline derivatives containing all structural requirements

for catalytic activity.

Scheme 9: Organocatalytic asymmetric conjugate addition using 40

3.3 Conclusions & Outlook

In conclusion, we have developed a highly efficient combination of MAO-N-catalyzed

desymmetrization of cyclic meso-amines with the Ugi-type 3CR. This procedure is

characterized by mild conditions, simple experiment procedures, and excellent yields and

d.r. and ee values. We also demonstrated that a one-pot biotransformastion/MCR process is

possible although the yield and diastereomeric ratios are superior in the two-pot procedure.

Further optimization is recommended as access to less stable imines would be possible.

We expect this methodology to be applicable to a wide variety of 3,4-cis-substituted

1-pyrrolines and therefore of considerable synthetic value in the construction of arrays

of otherwise hard-to-access 3,4-substituted prolyl peptides, for example, as Wennemers-

type organocatalysts. Moreover, our methodology holds great promise for applications in

medicinal chemistry, especially in the synthesis of novel hepatitis C drugs.

3.4 Experimental section

General Information

Starting materials and solvents were purchased from ABCR and Sigma-Aldrich and were

used without further purification. 3-Azabicylo[3,3,0]octane hydrochloride was purchased

from AK Scientific. (1R,2S,6R,7S)-4-methyl-4-azatricyclo[5.2.1.02.6]dec-8-ene was prepared

according to literature procedure.[45] Column chromatography was performed on silica gel.1H and 13C NMR spectra were recorded on a Bruker Avance 400 (400.13 MHz for 1H and 100.61

MHz for 13C) or Bruker Avance 500 (500.23 MHz for 1H and 125.78 MHz for 13C) in CDCl3 and

DMSO. Chemical shifts are reported in δ values (ppm) downfield from tetramethylsilane.

Page 67: Biocatalysis & Multicomponent Reactions: The Ideal …...Biocatalysis & Multicomponent Reactions Chapter 2 Monoamine Oxidase N: 41 A Promising Biocatalyst for Asymmetric Synthesis

Highly Stereoselective Synthesis of Substituted Prolyl Peptides

67

Electrospray Ionization (ESI) mass spectrometry was carried out using a Bruker micrOTOF-Q

instrument in positive ion mode (capillary potential of 4500 V). GC-MS spectra were recorded

on a Hewlett Packard HP 6890 equipped with a J & W Scientific; HP-1MS; 30 m × 0.32 mm

× 0.25 μm column (injector temp. 300 °C, oven temp. 100 °C to 280 °C at 5 °C/min, hold for

10 min., He 1.6 ml/min. and detector temp. 275 °C) and a HP 5973 Mass Selective Detector.

GC-FID analyses were performed on Agilent 6850 GC with a J & W Scientific; HP-1; 30 m ×

0.32 mm × 0.25 μm (injector temp. 300 °C, oven temp. 100 °C to 280 °C at 5 °C/min, hold

for 10 min., He 1.6 ml/min. and detector temp. 275 °C) and a Varian Chirasil-Dex CB; 25 m

× 0.25 mm × 0.26 μm column (inj. 250 °C, oven temp. 100 °C to 180 °C at 5 °C/min, hold for

10 min., He 1.6 ml/min. and detector temp. 275 °C) equipped with a Gerstel Multipurpose

sampler MPS2L. Normal phase HPLC was performed on Agilent systems. Normal phase

HPLC system was equipped with a G1322A degasser, a G1311A quaternary pump, a G1329

autosampler unit, a G1315B diode array detector and a G1316A temperature controlled

column compartment.

Infrared (IR) spectra were recorded neat, and wavelengths are reported in cm-1. Optical

rotations were measured with a sodium lamp and are reported as follows: [α]D20 (c =g/100

mL, solvent). Methyl 3-isocyanopropionate was synthesized as reported previously.[50]

General Procedure 1: Preparation of optically active imines (3S,7R)-29 and 35.

Unless stated otherwise: imines were synthesised according to literature procedure[46] with

minor adjustments. 2.5 g of freeze-dried MAO-N D5 E.Coli were rehydrated for 30 min. in 20

ml of KPO4 buffer (100 mM, pH = 8.0) at 37 °C. Subsequently 1 mmol amine ((3S,7R)-29 or 35)

in 30 ml of KPO4 buffer (100 mM, pH = 8.0) was prepared. The pH of the solution was adjusted

to 8,0 by addition of NaOH and then added to the rehydrated cells. After 16-17 h the reaction

was stopped (conversions were > 95 %) and worked up. For workup the reaction mixture

was centrifuged at 4000 rpm and 4°C until the supernatant had clarified (40 – 60 min.).

The pH of the supernatant was then adjusted to 10-11 by addition of aq. NaOH and the

supernatant was subsequently extracted with t-butyl methyl ether or dichloromethane (4

x 70 mL). The combined organic phases were dried with Na2SO4 and concentrated at the

rotary evaporator.

General procedure 2: Screening of reaction conditions

Unless stated otherwise: 3-azabicyclo[3,3,0]oct-2-ene (3S,7R)-29 (0.073 mmol) was dissolved

in 3 ml of the appropriate solvent followed by the addition of benzoic acid (0.1 mmol) and

tert-butyl isocyanide (0.1 mmol). The reaction mixture was stirred for 16 h at RT. Samples

were dried before analysis was carried out using GC-MS and GC-FID.

Page 68: Biocatalysis & Multicomponent Reactions: The Ideal …...Biocatalysis & Multicomponent Reactions Chapter 2 Monoamine Oxidase N: 41 A Promising Biocatalyst for Asymmetric Synthesis

Chapter 3

68

General procedure 3: Preparation of optically active Ugi products 30a-g and 36a-g

Unless stated otherwise: Imine (0.70 mmol) was dissolved in 2 ml of DCM followed by the

addition of carboxylic acid (0.93 mmol) and isocyanide (0.93 mmol). The reaction mixture

was stirred for 24 h at RT. DCM (8 mL) was added and the resulting mixture was washed

with Na2CO3 (2x10 mL), dried (MgSO4), filtered, and concentrated. Note: rotamers could be

observed in the NMR data.

General Procedure 4: Determination of enantiomeric excess (ee) and diastereomeric

ratio (d.r.)

Racemic imines were synthesized according to literature procedure.[46] Racemic Ugi-

type products were prepared according to general procedure 3. Diastereomers could be

separated by GC-MS and GC-FID. Enantiomers could be separated by normal phase HPLC

and GC-FID.

Compound 30a: General procedure 3 was followed using 3-azabicyclo[3,3,0]oct-2-ene ((3S,7R)-29, 76 mg, 0.70 mmol), acetic acid (55 mg, 52 μl, 0.91 mmol) and tert-butyl isocyanide (76 mg, 103 μl, 0.91 mmol) giving 30a as a white solid, yield 73%.93:7 d.r. [HP-1, t (major) = 14.852 min, t (minor) = 16.773 min]; 95% ee [CP Chirasil DEX CB, t (minor) = 20.449 min, t (major) = 20.860 min]; [α]D

20 = -47.8 o (c = 0.34, MeCN). 1H NMR (400.1 MHz, CDCl3): d 6.58 (bs, 1H), 4.28 (d, J = 2.1 Hz, 1H), 3.70 (dd, J = 8.3, 10.6

Hz, 1H), 3.24 (dd, J = 4.5, 10.6 Hz, 1H), 2.96-2.93 (m, 1H), 2.91-2.82 (m, 1H), 2.01 (s, 3H), 1.93-1.78 (m, 2H), 1.71-1.42 (m, 2H), 1.41-1.31 (m, 2H), 1.25 (s, 9H); 13C NMR (100.6 MHz, CDCl3) d 170.5, 170.0, 66.8, 54.4, 51.0, 45.0, 42.7, 32.5, 32.3, 28.7, 25.7, 22.6; IR (neat): νmax (cm-1) = 3277 (m), 2957 (m), 1668 (s), 1630 (s), 1549 (s), 1447 (s), 1420 (s), 1223 (s), 667 (m), 606 (m); HRMS (ESI+) calcd for C14H24N2O2 (MH+) 253.1916, found 253.1925.2

Compound 30b: General procedure 3 was followed using 3-azabicyclo[3,3,0]oct-2-ene ((3S,7R)-29, 76 mg, 0.70 mmol), benzoic acid (111 mg, 0.91 mmol) and tert-butyl isocyanide (76 mg, 103 μl, 0.91 mmol) giving 30b as a white solid, yield 73%. 93:7 d.r. [HP-1, t (major) = 23.672 min, t (minor) = 25.601 min]; 95% ee [Daicel Chiralpak AD H, hexane/iso = 96/4, v = 1.0 mL.min1, λ = 254 nm, t (minor) = 10.698 min, t (major) = 11.620 min]; [α]D

20 = -53.7 o (c = 0.34, MeCN). 1H NMR (400.1 MHz, CDCl3): d 7.49-7.38 (m, 5H), 6.66 (bs, 1H), 4.54 (d, J = 2.8 Hz), 3.72 (dd, J =m 11.4, 7.8

Hz, 1H), 3.23 (d, J = 11.0, 1H), 3.15-3.10 (m, 1H), 2.73-2.58 (m, 1H), 1.96-1.82 (m, 1H), 1.82-1.69 (m, 1H), 1.68-1.41 (m, 3H), 3.15-3.10 (m, 1H), 1.28 (s, 9H), 1.24-1.06 (m, 1H); 13C NMR (100.6 MHz, CDCl3) d 170.3, 170.1, 136.3, 130.1, 128.4, 126.9, 67.1, 60.4, 55.9, 51.1, 44.2, 43.3, 33.0, 32.7, 28.7, 26.2; ; IR (neat): νmax (cm-1) = 3310 (m), 2961 (m), 1674 (s), 1618 (s), 1416 (s), 1223 (s), 698 (s); HRMS (ESI+) calcd for C19H26N2O2 (MH+) 315.2073, found 315.2077.

Compound 30c: General procedure 3 was followed using 3-azabicyclo[3,3,0]oct-2-ene ((3S,7R)-29, 76 mg, 0.70 mmol), 3-furoic acid (102 mg, 0.91 mmol) and isopropyl isocyanide (63 mg, 86 μl, 0.91 mmol) giving 30c as a white solid, yield 75%.92:8 d.r. [HP-1, t (major) = 21.290 min, t (minor) = 23.012 min] 94% ee [Daicel Chiralpak AD-H, hexane/isopropanol = 90/10, v = 1.0 mL.min1, λ = 254 nm, t (minor) = 7.417 min, t (major) = 12.039 min]; [α]D

20 = -33.3 o (c = 0.30, MeCN). 1H NMR (400.1 MHz, CDCl3): d 7.80 (bs, 1H), 7.43 (bs, 1H), 6.72 (bs, 1H), 6.51 (d, J = 6.3 Hz, 1H), 4.56

(d, J = 2.3 Hz, 1H), 4.03 (oct, J = 7.1 1H), 3.88 (dd, J = 10.4, 8.3 Hz, 1H), 3.53 (dd, J = 10.4, 3.8 Hz, 1H), 3.09-3.01 (m, 1H), 2.95-2.84 (m, 1H), 2.00-1.84 (m, 2H), 1.74-1.65 (m, 1H), 1.64-1.54 (m, 1H), 1.53-1.43 (m, 1H),

Page 69: Biocatalysis & Multicomponent Reactions: The Ideal …...Biocatalysis & Multicomponent Reactions Chapter 2 Monoamine Oxidase N: 41 A Promising Biocatalyst for Asymmetric Synthesis

Highly Stereoselective Synthesis of Substituted Prolyl Peptides

69

1.43-1.33(m, 1H), 1.17 (d, J = 6.3 Hz, 3H) 1.13 (d, J = 6.3 Hz, 3H); 13C NMR (100.6 MHz, CDCl3) d 170.1, 163.2, 144.3, 142.8, 121.8, 110.4, 66.8, 54.8, 44.4, 43.3, 41.3, 32.4, 32.2, 25.6, 22.5, 22.4; IR (neat): νmax (cm-1) = 3281 (w), 2949 (w), 1647 (m), 1609 (s), 1547 (s), 1427 (s), 1159 (s), 737 (s), 598 (s); HRMS (ESI+) calcd for C16H22N2O3 (MH+) 291.1709, found 291.1721.

Compound 30d: General procedure 3 was followed using 3-azabicyclo[3,3,0]oct-2-ene ((3S,7R)-29, 76 mg, 0.70 mmol), benzoic acid (111 mg, 0.91 mmol) and isopropyl isocyanide (63 mg, 86 μl, 0.91 mmol) giving 30d as a white solid, yield 78%.92:8 d.r. [HP-1, t (major) = 23.809 min, t (minor) = 25.563 min]; 94% ee [Daicel Chiralpak AD-H, hexane/isopropanol = 96/4, v = 1.0 mL.min1, λ = 254 nm, t (minor) = 16.613 min, t (major) = 21.363 min]; [α]D

20 = -52.4 o (c = 0.42, MeCN). 1H NMR (400.1 MHz, CDCl3): d 7.46-7.36 (m, 1H), 6.63 (bs, 1H), 4.59 (d, J = 2.0 Hz, 1H), 4.10-4.01 (m, 1H), 3.73 (dd, J = 11.4,. 7.8 Hz, 1H), 3.73 (dd, J = 11.4,. 7.8 Hz, 1H), 3.32-3.29 (m, 1H), 3.23-3.17(m, 1H), 2.76-2.71 (m, 1H), 2.02-1.94 (m, 1H), 1.88-1.78 (m, 1H), 1.75-1.63 (m, 1H), 1.63-1.50 (m, 1H), 1.16 (d, J = 6.6 Hz, 3H) 1.13 (d, J = 6.6 Hz, 3H); 13C NMR (100.6 MHz, CDCl3) d 170.4, 170.0, 136.2, 130.1, 128.4, 126.9, 126.6, 66.5, 55.9, 44.3, 43.3, 41.5, 32.9, 32.6, 26.1, 22.7, 22.6; IR (neat): νmax (cm-1) = 3300 (m), 2959 (m), 1615 (s), 1545 (s), 1416 (s), 1229 (m), 700 (m);HRMS (ESI+) calcd for C18H24N2O2 (MH+) 301.1916, found 301.1914.

Compound 30e: General procedure 3 was followed using 3-azabicyclo[3,3,0]oct-2-ene ((3S,7R)-29, 76 mg, 0.70 mmol), acetic acid (55 mg, 52 μl, 0.91 mmol) and benzyl isocyanide (107 mg, 111 μl, 0.91 mmol) giving 30e as a white solid, yield 71%.92:8 d.r. [HP-1, t (major) = 25.098 min, t (minor) = 26.457 min]; 94% ee [Daicel Chiralpak OJ-H, hexane/isopropanol = 93/7, v = 1.0 mL.min1, λ = 254 nm, t (minor)

= 9.948 min, t (major) = 10.718 min]; [α]D20 = -18.8 o (c = 0.32, MeCN). 1H NMR (400.1 MHz, CDCl3): d 7.36-

7.25 (m, 5H), 7.16 (bs, 1H), 4.47 (d, J = 2.0 Hz, 1H), 4.25 (dd, J = 15.2, 5.8 Hz, 2H), 3.73 (dd, J = 10.6, 8.3 Hz, 1H), 3.28 (dd, J = 10.5, 4.5 Hz, 1H), 3.10-3.03 (m, 1H), 2.96-2.88 (m, 1H), 2.10 (s, 3H), 2.01-1.85 (m, 1H), 1.82-1.55 (m, 2H), 1.52-1.39 (m, 2H); 13C NMR (100.6 MHz, CDCl3) d 171.3, 170.1, 138.3, 128.5, 128.4, 127.9, 127.4, 127.0, 65.9, 54.3, 45.4, 43.2, 42.7, 32.6, 32.2, 25.5, 22.1; IR (neat): νmax (cm-1) = 3267 (m), 2951 (w), 1626 (s), 1537 (m), 1418 (s), 1231 (s), 1030 (w), 748 (s); HRMS (ESI+) calcd for C17H22N2O2 (MH+) 287.1760, found 281.1765.

Compound 30f: General procedure 3 was followed using 3-azabicyclo[3,3,0]oct-2-ene ((3S,7R)-29, 76 mg, 0.70 mmol), benzoic acid (111 mg, 0.91 mmol) and benzyl isocyanide (107 mg, 111 μl, 0.91 mmol) giving 30f as a white solid, yield 81%.92:8 d.r. [HP-1, t (major) = 33.333 min, t (minor) = 35.085 min]; 97% ee [Daicel Chiralpak AD-H, hexane/isopropanol = 96/4, v = 1.0 mL.min1, λ = 254 nm, t (minor) = 18.134 min, t (major) = 23.440 min]; [α]D

20 = -52.6 o (c = 0.38, MeCN). 1H NMR (400.1 MHz, CDCl3): d 7.34-7.11 (m, 10H), 6.73 (bs, 1H), 4.61 (d, J = 2.8 Hz, 1H). 4.37 (dd, J = 5.3, 2.8 Hz, 2H), 3.66 (dd, J = 11.1, 7.6 Hz, 1H), 3.24 (dd, J = 10.9, 1.8 Hz, 1H), 3.18-3.11 (m, 1H), 2.72-2.64 (m, 1H), 1.92-1.82 (m, 1H), 1.82-1.62 (m, 1H), 1.25-1.13 (m, 1H); 13C NMR (100.6 MHz, CDCl3) d 171.1, 170.3, 138.3, 135.9, 132.6, 130.0, 129.7, 128.4, 128.21, 128.0, 127.7, 127.3, 127.0, 126.9, 126.4, 66.3, 55.8, 44.9, 43.2, 43.1, 32.8, 32.4, 25.9; IR (neat): νmax (cm-1) = 3262 (m), 2928 (m), 1674 (s), 1613 (s), 1545 (s), 1423 (s), 1223 (m), 698 (s), 669 (s); HRMS (ESI+) calcd for C22H24N2O2 (MH+) 349.1916, found 349.1924.

Compound 30g: General procedure 3 was followed using 3-azabicyclo[3,3,0]oct-2-ene ((3S,7R)-29, 76 mg, 0.70 mmol), isobutyric acid (80 mg, 84 μl, 0.91 mmol) and tert-butyl isocyanide (76 mg, 103 μl, 0.91 mmol) giving 30g as a white solid, yield 83%.93:7 d.r. [HP-1, t (major) = 17.165 min, t (minor) = 18.750 min]; 97% ee [CP Chirasil-DEX CB, t (minor) = 21.439 min, t (major) = 21.846 min]; [α]D

20 = -47.8 o (c = 0.34,

Page 70: Biocatalysis & Multicomponent Reactions: The Ideal …...Biocatalysis & Multicomponent Reactions Chapter 2 Monoamine Oxidase N: 41 A Promising Biocatalyst for Asymmetric Synthesis

Chapter 3

70

MeCN). 1H NMR (400.1 MHz, CDCl3): d 6.71 (bs, 1H), 4.37 (d, J = 1.8 Hz, 1H), 3.65 (dd, J = 10.6, 8.3 Hz, 1H), 3.34 (dd, J = 4.3, 10.9 Hz, 1H), 3.04-2.98 (m, 1H), 2.89-2.81 (m, 1H), 2.65 (sep, J = 6.8 Hz, 1H), 2.01-1.83 (m, 2H), 1.70-1.48 (m, 2H), 1.45-1.35 (m, 2H), 1.29(s, 9H), 1.13 (d, J = 6.6 Hz, 3H), 1.10 (d, J = 6.8 Hz, 3H); 13C NMR (100.6 MHz, CDCl3) d 176.5, 170.4, 66.5, 53.0, 43.7, 42.7, 32.8, 32.3, 32.0, 30.8, 24.9, 18.9, 18.5; IR (neat): νmax (cm-1) = 3291 (m), 2963 (m), 2870 (w), 1684 (s), 1618 (s), 1551 (s), 1433 (s), 1225 (s), 1090 (m), 658 (m); HRMS (ESI+) calcd for C14H24N2O2 (MH+) 281.2229, found 281.2235.

Compound 36a: General procedure 3 was followed using (1R,2S,6R,7S)-4-methyl-4-azatricyclo[5.2.1.02,6]dec-8-ene (35, 93 mg, 0.70 mmol), acetic acid (55 mg, 52 μl, 0.91 mmol) and tert-butyl isocyanide (76 mg, 103 μl, 0.91 mmol) giving 36a as a white solid, yield 83%.>99:1 d.r. (t (major) = 18.179 min); >99% ee [Daicel Chiralpak ADH, hexane/isopropanol = 92/8, v = 1.0 mL.min1, λ = 220 nm, t (major) = 5.319 min, t (minor) =

6.587 min]; [α]D20 = -24.0 o (c = 0.25, MeCN). 1H NMR (400.1 MHz, CDCl3): d 6.65 (bs, 1H), 6.14-6.13 (m, 2H),

4.09 (d, J = 2.0 Hz, 1H), 3.47 (dd, J = 11.4, 8.6 Hz, 1H), 3.36-3.32 (m, 1H), 3.15 (dd, J = 11.4, 2.0 Hz, 1H), 2.98-2.92 (m, 3H), 1.95 (s, 3H), 1.51-1.41 (m, 2H), 1.30 (s, 9H); 13C NMR (100.6 MHz, CDCl3) d 170.6, 169.0, 135.4, 134.0, 62.9, 51.7, 51.0, 50.3, 47.0, 46.6, 46.0, 45.1, 28.7, 22.8; IR (neat): νmax (cm-1) = 3283 (w), 2970 (w), 2942 (w), 1647 (s), 1634 (s), 1553 (s), 1414 (s), 1223 (s), 733 (s); HRMS (ESI+) calcd for C16H24N2O2 (MH+) 277.1916, found 277.1922.

Compound 36b: General procedure 3 was followed using 3-azabicyclo[3,3,0]oct-2-ene (35, 76 mg, 0.70 mmol), benzoic acid (111 mg, 0.91 mmol) and tert-butyl isocyanide (76 mg, 103 μl, 0.91 mmol) giving 36b as a white solid, yield 82%.>99:1 d.r. (t (major) = 26.830 min); HP-1, >99% ee [Daicel Chiralpak AD-H, hexane/isopropanol = 96/4, v = 1.0 mL.min1, λ = 220 nm, t (minor) = 9.712 min, t (major) = 11.741 min]; [α]D

20 = -43.1 o (c = 0.33, MeCN). 1H NMR (400.1 MHz, CDCl3): d 7.43-7.34 (m, 1H), 6.63 (bs, 1H), 6.20 (dd, J = 5.8, 3.0 Hz, 1H), 5.91 (dd, J = 5.6, 2.6 Hz, 1H), 4.43

(d, J = 2.0 Hz, 1H), 3.52 (dd, J = 11.9, 8.6 Hz, 1H), 3.44-3.39 (m, 1H), 3.05-3.00 (m, 2H), 2.91-2.85 (m, 1H), 2.78-2.76 (m, 1H), 1.48-1.45 (m, 1H), 1.40-1.37 (m, 1H), 1.32 (s, 9H); 13C NMR (100.6 MHz, CDCl3) d 168.6, 168.0, 135.0, 133.2, 132.7, 128.4, 126.8, 124.9, 61.3, 50.4, 49.9, 49.5, 45.4, 44.9, 43.9, 43.3, 27.1; IR (neat): νmax (cm-1) = 3283 (m), 2970 (m), 2942 (m), 1647 (s), 1634 (s), 1553 (s), 1414 (s), 1223 (s), 733 (s); HRMS (ESI+) calcd for C21H26N2O2 (MH+) 339.2073, found 339.2082.

Compound 36c: General procedure 3 was followed using 1R,2S,6R,7S)-4-methyl-4-azatricyclo[5.2.1.02,6]dec-8-ene (35, 93 mg, 0.70 mmol), 3-furoic acid (102 mg, 0.91 mmol) and isopropyl isocyanide (63 mg, 86 μl, 0.91 mmol) giving 36c as a white solid, yield 75%. >99:1 d.r. (t (major) = 24.364 min); >99% ee [Daicel Chiralpak ADH, hexane/isopropanol = 90/10, v = 1.0 mL.min1, λ = 220 nm, t (minor) = 8.404 min, t (major) = 9.968 min]; [α]D

20 = -35.7 o (c = 0.28, MeCN). 1H NMR (400.1 MHz, CDCl3): d 7.71 (dd, J = 1.5, 0.8 Hz, 1H), 7.42 (dd, J = 2.0, 1.5 Hz, 1H), 6.65 (dd, J = 1.8, 0.8 Hz, 1H),

6.50 (d, J = 6.6 Hz, 1H), 6.19-6.17 (m, 1H), 5.98-5.96 (m, 1H), 4.42 (d, J = 2.0 Hz, 1H), 4.06-3.94 (m, 1H), 3.63 (dd, J =11.4, 8.8 Hz, 1H), 3.43-3.39 (m, 2H), 3-06-3.02 (m, 2H), 2.90-2.88 (m, 1H), 1.51-1.43 (m, 2H), 1.13 (d, J = 6.6 Hz, 3H), 1.10 (d, J = 6.6 Hz, 3H); 13C NMR (100.6 MHz, CDCl3) d 170.2, 162.7, 144.1, 142.9, 135.1, 134.4, 121.9, 110.4, 62.8, 51.6, 51.2, 47.1, 46.5, 45.8, 45.2, 41.5, 22.6, 22.7, 22.6; IR (neat): νmax (cm-1) = 3275 (m), 2970 (m), 2934 (m), 1678 (s), 1594 (s), 1545 (s), 1437 (s), 1219 (s), 1153 (m), 1018 (m)874 (m0, 754 (s); HRMS (ESI+) calcd for C18H22N2O2 (MH+) 315.1709, found 315.1725.

Compound 36d: General procedure 3 was followed using 1R,2S,6R,7S)-4-methyl-4-azatricyclo[5.2.1.02,6]dec-8-ene (35, 93 mg, 0.70 mmol), benzoic acid (111 mg, 0.91 mmol) and isopropyl isocyanide (63 mg, 86 μl, 0.91 mmol) giving 36d as a white solid, yield 78%.>99:1 d.r. (t (major) = 27.054 min); >99% ee [Daicel Chiralpak ADH, hexane/isopropanol = 96/4, v = 1.0 mL.min1, λ = 220 nm, t (minor) = 17.354 min, t (major)

Page 71: Biocatalysis & Multicomponent Reactions: The Ideal …...Biocatalysis & Multicomponent Reactions Chapter 2 Monoamine Oxidase N: 41 A Promising Biocatalyst for Asymmetric Synthesis

Highly Stereoselective Synthesis of Substituted Prolyl Peptides

71

= 29.404 min]; [α]D20 = -38.7 o (c = 0.31, MeCN). 1H NMR (400.1 MHz, CDCl3): d 7.43-7.35 (m, 5H), 6.59 (d, J

= 7.6 Hz, 1H), 6.23-6.21 (m, 1H), 5.93-5.91 (m, 1H), 4.48 (d, J = 1.7 Hz, 1H), 4.11-3.98 (m, 1H), 3.55(dd, J = 11.9, 8.8 Hz, 1H), 3.48-3.45 (m, 1H), 3.07-3.00 (m, 2H), 2.92-2.87 (m, 1H), 2.81-2.77 (m, 1H), 1.48-1.39 (m, 2H), 1.14 (d, J = 6.6 Hz, 3H), 1.10 (d, J = 6.6 Hz, 3H); 13C NMR (100.6 MHz, CDCl3) d 170.1, 169.8, 136.5, 134.8, 134.4, 130.1, 128.4, 126.6, 62.3, 52.0, 51.5, 47.0, 46.5, 45.6, 44.9, 41.5, 22.7, 22.6; IR (neat): νmax (cm-1) = 3287 (m), 2967 (m), 2940 (m), 1682 (s), 1601 (s), 1539 (s), 1424 (s), 1217 (s), 739 (s), 664 (m);HRMS (ESI+) calcd for C20H24N2O2 (MH+) 325.1916, found 325.1919.

Compound 36e: General procedure 3 was followed using 1R,2S,6R,7S)-4-methyl-4-azatricyclo[5.2.1.02,6]dec-8-ene (35, 93 mg, 0.70 mmol), acetic acid (55 mg, 52 μl, 0.91 mmol) and benzyl isocyanide (107 mg, 111 μl, 0.91 mmol) giving 36e as a white solid, yield 78%.>99:1 d.r. (t (major) = 28.213 min); >99% ee [Daicel Chiralpak OJ-H, hexane/isopropanol = 92/8, v = 1.0 mL.min1, λ = 220 nm, t (minor) = 9.100 min, t (major)

= 10.760 min]; [α]D20 = -21.4 o (c = 0.28, MeCN). 1H NMR (400.1 MHz, CDCl3): d 7.31-7.21 (m, 5H), 6.13-6.08

(m, 2H), 4.46 (dd, J = 14.9, 6.1 Hz, 1H), 4.32 (dd, J = 15.2, 5.8 Hz, 1H), 4.25 (d. J = 2.0 Hz, 1H), 3.47-3.43 (m, 2H), 3.18 (dd, J = 11.4, 2.0 Hz, 1H), 3.01-2.95 (m, 3H), 2.0 (s, 3H), 1.54-1.43 (m, 2H); 13C NMR (100.6 MHz, CDCl3) d 171.3, 169.3, 138.3, 135.4, 134.1, 128.6, 127.4, 127.3, 62.2, 51.7, 50.3, 47.1, 46.7, 46.0, 45.1, 43.4, 22.7; IR (neat): νmax (cm-1) = 3314 (w), 3082 (w), 2970 (w), 2932 (w), 1553 (s), 1433 (s), 1360 (m), 1317 (m), 1233 (m), 745 (s), 696 (s); HRMS (ESI+) calcd for C19H22N2O2 (MH+) 311.1760, found 311.1745.

Compound 36f: General procedure 3 was followed using 31R,2S,6R,7S)-4-methyl-4-azatricyclo[5.2.1.02,6]dec-8-ene (35, 93 mg, 0.70 mmol), benzoic acid (111 mg, 0.91 mmol) and benzyl isocyanide (107 mg, 111 μl, 0.91 mmol) giving 36f as a white solid, yield 80%.>99:1 d.r. (t (major) = 36.331 min); >99% ee [Daicel Chiralpak OD-H, hexane/isopropanol = 92/8, v = 1.0 mL.min1, λ = 220 nm, t (major) = 11.489 min, t (minor) = 13.626 min]; [α]D

20 = -35.1 o (c = 0.29, MeCN). 1H NMR (400.1 MHz, CDCl3): d 7.44-7.17 (m, 10H), 6.24-6.18 (m, 1H), 5.95-5.93 (m, 1H), 4.60 (d, J = 1.8 Hz, 1H), 4.45 (d, J = 5.8 Hz, 2H), 3.58-3.50 (m, 2H), 3.10-3.04 (m, 2H), 2.95-2.89 (m, 1H), 2.83-2.79 (m, 1H), 1.50-1.41 (m, 2H); 13C NMR (100.6 MHz, CDCl3) d 171.1, 169.9, 138.4, 136.4, 134.4, 130.1, 128.6, 128.4, 127.4, 127.3, 126.4, 62.2, 52.1, 51.6, 47.0, 46.6, 45.6, 45.0, 43.5; IR (neat): νmax (cm-1) = 3268 (m), 3077 (w), 2972 (w), 2872 (w), 1684 (s), 1597 (s), 1560 (s), 1495 (m), 1431 (s), 1221 (s), 731 (s), 696 (s); HRMS (ESI+) calcd for C24H24N2O2 (MH+) 373.1916, found 373.1901.

Compound 36g: General procedure 3 was followed using 1R,2S,6R,7S)-4-methyl-4-azatricyclo[5.2.1.02,6]dec-8-ene (35, 93 mg, 0.70 mmol), isobutyric acid (80 mg, 84 μl, 0.91 mmol) and tert-butyl isocyanide (76 mg, 103 μl, 0.91 mmol) giving 36g as a white solid, yield 81%.>99:1 d.r. (t (major) = 19.912 min); >99% ee [Daicel Chiralpak AD-H, hexane/isopropanol = 95/5, v = 1.0 mL.min1, λ = 220 nm, t (minor) = 5.037 min, t (major) = 6.877 min]; [α]D

20 = -35.3 o (c = 0.34, MeCN). 1H NMR (400.1 MHz, CDCl3): d 6.13-6.08 (m, 2H), 4.15 (d, J = 1.8 Hz, 1H), 3.43 (dd, J = 11.4, 8.6 Hz, 1H), 3.37-3.33 (m, 1H), 3.26 (dd, J = 11.4, 2.0 Hz, 1H), 2.99-2.91 (m, 3H), 2.50 (sep, J = 6.8 Hz, 1H), 1.54-1.38 (m, 2H), 1.27 (s, 9H), 1.06 (d, J = 6.8 Hz, 3H), 1.03(d, J = 6.8 Hz, 3H); 13C NMR (100.6 MHz, CDCl3) d 175.6, 170.7, 135.3, 134.2, 62.7, 51.8, 50.8, 49.1, 47.01, 46.5, 45.2, 32.2, 28.7, 19.2, 18.3; IR (neat): νmax (cm-1) = 3325 (m), 2966 (m), 1678 (m), 1624 (s), 1553 (s), 1435 (s), 1315 (w), 1231 (m), 1088 (w) ; HRMS (ESI+) calcd for C18H28N2O2 (MH+) 305.2229, found 305.222

Page 72: Biocatalysis & Multicomponent Reactions: The Ideal …...Biocatalysis & Multicomponent Reactions Chapter 2 Monoamine Oxidase N: 41 A Promising Biocatalyst for Asymmetric Synthesis

Chapter 3

72

Compound 36i: General procedure 3 was followed using 1R,2S,6R,7S)-4-methyl-4-azatricyclo[5.2.1.02,6]dec-8-ene (35, 93 mg, 0.70 mmol), Fmoc-D-Pro-OH (307 mg, 0.91 mmol) and tert-butyl isocyanide (76 mg, 103 μl, 0.91 mmol). The crude product 36i was subjected using column chromatography (SiO2, EtOAc (1): cyclohexane (2)). Fmoc deprotection using 25 % piperidine in DMF followed by column chromatography (DCM/MeOH 9:1) gave 36i as a white solid in 66% yield over two steps.

[α]D20 = -75.0 o (c = 0.16, MeCN). 1H NMR (500.2 MHz, CDCl3): d 6.75 (bs, 1H), 6.11 (d, J = 5.7 Hz, 2H), 5.01

(bs, 1H), 4.20 (bs, 1H), 3.92-3.83 (m, 1H), 3.45-3.41 (m, 1H), 3.24-3.22 (m, 1H), 3.20-3.10 (m, 2H), 3.03-2.98 (m, 2H), 2.95-2.89 (m, 1H), 2.19-2.09 (m, 1H), 1.94- 1.57 (m, 3H), 1.57-1.39 (m, 2H), 1.28 (s, 9H); 13C NMR (125.8 MHz, CDCl3) d 170.5, 135.4, 134.4, 63.7, 61.4, 51.6, 51.0, 49.01, 47.2, 46.9, 46.5, 46.3, 45.4, 30.0, 28.7, 25.9; IR (neat): νmax (cm-1) = 2960 (w), 1668 (s), 1622 (s), 1566 (m), 1414 (s), 1234 (m), 1094 (w), 853 (s); HRMS (ESI+) calcd for C19H29N3O2 (MH+) 332.2338, found 332.2342.

Compound 36j: General procedure 3 was followed using 1R,2S,6R,7S)-4-methyl-4-azatricyclo[5.2.1.02,6]dec-8-ene (35, 93 mg, 0.70 mmol), Fmoc-L-Pro-OH (307 mg, 0.91 mmol) and tert-butyl isocyanide (76 mg, 103 μl, 0.91 mmol) giving 36j as a white solid, yield 76%. [α]D

20 = -6.7 o (c = 0.60, MeCN). 1H NMR (500.2 MHz, CDCl3): d 7.76 (d, J = 7.6 Hz, 2H), d 7.59 (d, J = 7.5 Hz, 1H), 7.55 (d, J = 7.5 Hz, 1H), 7.40 (d, J = 7.4 Hz, 2H), 7.40 (d, J = 7.4 Hz, 2H), 7.31 (d, J = 7.4 Hz, 2H), ), 6.62 (bs, 1H), 6.19 (dd, J = 5.6, 2.9 Hz, 1H), 6.11 (dd, J = 5.6, 2.5 Hz, 1H), 4.35-4.33 (m, 1H), 4.30-4.28 (m, 2H), 4.24-4.22 (m, 1H), 4.20

(d, J = 1.9 Hz, 1H), 3.72-3.58 (m, 2H), 3.35-3.32 (m, 1H), 3.25-3.22 (m, 2H), 2.93-2.90 (m, 2H), 2.18-2.13 (m, 2H), 1.95-1.91 (m, 2H), 1.52-1.40 (m, 2H), 1.29 (s, 9H) ; ); 13C NMR (125.8 MHz, CDCl3) d 170.6, 170.4, 154.8, 143.9, 141.3, 135.9, 134.2, 127.8, 127.7, 127.1, 127.0, 125.1, 125.0, 120.0, 67.5, 63.3, 58.3, 51.8, 51.2, 49.4, 47.7, 47.4, 47.2, 47.1, 46.9, 44.4, 28.7, 28.6, 25.0; IR (neat): νmax (cm-1) = 2965 (w), 1643 (s), 1520 (w), 1449 (s), 1418 (s), 1358 (m), 1123 (m), 758 (m), 739 (s); HRMS (ESI+) calcd for C34H39N3O4 (MH+) 554.3019, found 554.3019.

Compound 36k: General procedure 3 was followed using 31R,2S,6R,7S)-4-methyl-4-azatricyclo[5.2.1.02,6]dec-8-ene (35, 93 mg, 0.70 mmol), benzoic acid (111 mg, 0.91 mmol) and (R)-(+)-methylbenzyl isocyanide (119 mg, 123 μl, 0.91 mmol) giving 36k as a white solid, yield 73%. [α]D

20 = -24.0 o (c = 0.25, MeCN). 1H NMR (500.2 MHz, CDCl3): d 7.45-7.19 (m, 11H), 6.24 dd, J = 5.7, 2.9 Hz, 1H), 5.93 dd, J = 5.7, 2.9 Hz, 1H), 5.11-5.03 (m, 1H), 4.59 d, J = 1.8 Hz, 1H), 3.53-3.49 (m, 1H), 3.40-3.35 (m, 1H), 3.02-2.97 (m, 2H), 2.88-2.81 (m, 1H), 2.77-2.75 (m, 1H), 1.48-1.39 (m, 2H), 1.45 (d, J = 7.0 Hz, 2H); 13C NMR (125.8

MHz, CDCl3): d 169.8, 169.8, 144.0, 136.4, 134.8, 134.3, 130.0, 128.5, 128.4, 126.9, 126.4, 125.7, 62.0, 51.8, 51.5, 49.0, 47.0, 46.4, 45.5, 44.4, 22.7; IR (neat): νmax (cm-1) = 3302 (w), 3239 (w), 3059 (w), 2665 (w), 2929 (w), 1664 (s), 1559 (s), 1558 (s), 1427 (s), 1248 (m), 1020 (m), 698 (s), 667 (m); HRMS (ESI+) calcd for C25H26N2O2 (MH+) 387.2073, found 387.2067.

Compound 36l: General procedure 3 was followed using 31R,2S,6R,7S)-4-methyl-4-azatricyclo[5.2.1.02,6]dec-8-ene (35, 93 mg, 0.70 mmol), benzoic acid (111 mg, 0.91 mmol) and (S)-(−)-methylbenzyl isocyanide (119 mg, 123 μl, 0.91 mmol) giving 36l as a white solid, yield 70%. [α]D

20 = -78.6 o (c = 0.28, MeCN). 1H NMR (400.1 MHz, CDCl3): d 7.43-7.16 (m, 11H), 6.24-6.19 (m, 1H), 5.97-5.90 (m, 1H), 5.10-5.01 (m, 1H), 4.52 (bs, 1H), 3.61-3.56 (m, 1H), 3.47-3.44 (m, 1H), 3.09-3.06 (m, 1H), 2.99 (bs, 1H), 2.93-2.87 (m, 1H), 2.80(bs, 1H), 1.50-1.39 (m, 2H), 1.43 (d, J = 6.8 Hz, 3H); 13C NMR (125.8MHz, CDCl3) d 170.1, 169.9, 143.3, 136.5, 134.8, 134.4, 130.1, 128.7, 128.5, 127.2,

126.5, 126.1, 62.4, 52.1, 51.6, 49.1, 47.0, 46.5, 45.6, 44.9, 22.4; IR (neat): νmax (cm-1) = 3281 (m), 3281 (m), 2955 (m), 1638 (s), 1528 (s), 1397 (s), 1343 (m), 1227 (m), 1117 (m), 700 (s); HRMS (ESI+) calcd for C25H26N2O2 (MH+) 387.2073, found 387.2067

Page 73: Biocatalysis & Multicomponent Reactions: The Ideal …...Biocatalysis & Multicomponent Reactions Chapter 2 Monoamine Oxidase N: 41 A Promising Biocatalyst for Asymmetric Synthesis

Highly Stereoselective Synthesis of Substituted Prolyl Peptides

73

Compound 40: General procedure 3 was followed using 1R,2S,6R,7S)-4-methyl-4-azatricyclo[5.2.1.02,6]dec-8-ene (35, 93 mg, 0.70 mmol), Fmoc-L-Pro-OH (307 mg, 0.91 mmol) and tert-butyl isocyanide (76 mg, 103 μl, 0.91 mmol). The crude product was purified by column chromatography (SiO2, EtOAc (1): cyclohexane (2)). Simultaneous Fmoc deprotection and saponification according to literature procedure[49] followed by addition of 1.1 eq. TFA and purification using reversed phase chromatography (C18, H2O (1): EtOH (1)) gave 40 as a colorless solid, in 62% yield over two steps.

[α]D20 = -6.8 o (c = 0.30, MeCN). 1H NMR (500.2 MHz, DMSO): d 12.43 (bs, 1H), 9.68-9.44 (m, 1H), 8.54 (bs,

1H), 8.19-8.14 (m, 1H), 6.29-6.25 (m, 1H), 6.13 (dd, J = 5.7, 2.9 Hz, 1H), 3.95 (bs, 1H), 3.66-3.59 (m, 1H), 3.37-3.18 (m, 5H), 2.96 (bs, 2H), 2.79-2.73(m, 1H), 2.50-2.38 (m, 4H), 1.97-1.85 (m, 2H), 1.69 1.57 (m, 2H), 1.46-1.35 (m, 2H); 13C NMR (125.8 MHz, DMSO) d 172.7, 171.2, 165.8, 158.1, 157.9, 135.6, 134.8, 117.9, 115.5, 62.9, 58.1, 50.9, 49.6, 49.0, 46.5, 46.4, 45.8, 44.1, 34.6, 33.8, 27.9, 23.6; IR (neat): νmax (cm-1) = 2949 (w), 1640 (s), 1175 (s) 1130 (s), 833 (m), 719 (m); HRMS (ESI+) calcd for C20H26F3N3O3 (MH+) 347.1845,

found 348.1929.

X-ray crystal structure determination of 30f

3.5 References & Notes[1] A. Domling and I. Ugi, Angew. Chem. Int. Ed. 2000, 39, 3169-3210.[2] J. Zhu and H. Bienayme, Multicomponent Reactions, Wiley-VCH, Weinheim, 2005.[3] R. V. A. Orru and M. de Greef, Synthesis 2003, 1471-1499.[4] A. Domling, Chem. Rev. 2006, 106, 17-89.[5] B. Ganem, Accounts Chem. Res. 2009, 42, 463-472.[6] C. Hulme and V. Gore, Curr. Med. Chem. 2003, 10, 51-80.[7] I. Akritopoulou-Zanze, Curr. Opin. Chem. Biol. 2008, 12, 324-331.[8] U. Kusebauch, B. Beck, K. Messer, E. Herdtweck and A. Domling, Org. Lett. 2003, 5, 4021-4024.

Page 74: Biocatalysis & Multicomponent Reactions: The Ideal …...Biocatalysis & Multicomponent Reactions Chapter 2 Monoamine Oxidase N: 41 A Promising Biocatalyst for Asymmetric Synthesis

Chapter 3

74

[9] S. E. Denmark and Y. Fan, J. Am. Chem. Soc. 2003, 125, 7825-7827.[10] P. R. Andreana, C. C. Liu and S. L. Schreiber, Org. Lett. 2004, 6, 4231-4233.[11] S. C. Pan and B. List, Angew. Chem. Int. Ed. 2008, 47, 3622-3625.[12] T. Yue, M. X. Wang, D. X. Wang, G. Masson and J. P. Zhu, J. Org. Chem. 2009, 74, 8396-8399.[13] D. J. Ramon and M. Yus, Angew. Chem. Int. Ed. 2005, 44, 1602-1634.[14] I. Ugi, Angew. Chem. Int. Ed. 1959, 71, 386-386.[15] R. F. Nutt and M. M. Joullie, J. Am. Chem. Soc. 1982, 104, 5852-5853.[16] L. Banfi, A. Basso, G. Guanti, S. Merlo, C. Repetto and R. Riva, Tetrahedron 2008, 64, 1114-1134.[17] M. M. Bowers, P. Carroll and M. M. Joullie, J. Chem. Soc.-Perkin Trans. 1 1989, 857-865.[18] D. M. Flanagan and M. M. Joullie, Synth. Commun. 1989, 19, 1-12.[19] L. Banfi, A. Basso, G. Guanti and R. Riva, Tetrahedron Lett. 2004, 45, 6637-6640.[20] T. M. Chapman, I. G. Davies, B. Gu, T. M. Block, D. I. C. Scopes, P. A. Hay, S. M. Courtney, L. A. McNeill,

C. J. Schofield and B. G. Davis, J. Am. Chem. Soc. 2005, 127, 506-507.[21] K. M. Bonger, T. Wennekes, S. V. P. de Lavoir, D. Esposito, R. den Berg, R. Litjens and G. A. van der

Marel, Qsar Comb. Sci. 2006, 25, 491-503.[22] K. M. Bonger, T. Wennekes, D. V. Filippov, G. Lodder, G. A. van der Marel and H. S. Overkleeft, Eur. J.

Org. Chem. 2008, 3678-3688.[23] T. Ngouansavanh and J. Zhu, Angew. Chem. Int. Ed. 2007, 46, 5775-5778.[24] M. Alexeeva, A. Enright, M. J. Dawson, M. Mahmoudian and N. J. Turner, Angew. Chem. Int. Ed.

2002, 41, 3177-3180.[25] R. Carr, M. Alexeeva, A. Enright, T. S. C. Eve, M. J. Dawson and N. J. Turner, Angew. Chem. Int. Ed.

2003, 42, 4807-4810.[26] V. Kohler, K. R. Bailey, A. Znabet, J. Raftery, M. Helliwell and N. J. Turner, Angew. Chem. Int. Ed. 2010,

49, 2182-2184.[27] M. D. Clay, J. Colbeck, J. M. Gruber, J. Lalonde, J. Liang, J. Mavinhalli, B. Mijts, S. Muley, J. D. Munger,

L. M. Newman, R. Sheldon, X. Zhang, J. Zhu, J. Mavinahalli, L. Newman, M. Clay, J. Gruber and J. Munger in Vol. CODEXIS INC (CODE-Non-standard), 2009, pp. 2307419-A2307412:.

[28] R. S. Gordee, D. J. Zeckner, L. F. Ellis, A. L. Thakkar and L. C. Howard, J. Antibiot. 1984, 37, 1054-1065.

[29] R. F. Hector, Clin. Microbiol. Rev. 1993, 6, 1-21.[30] M. I. Morris and M. Villmann, Am. J. Health-Syst. Pharm. 2006, 63, 1693-1703.[31] H. Morita, S. Nagashima, K. Takeya, H. Itokawa and Y. Iitaka, Tetrahedron 1995, 51, 1121-1132.[32] H. Morita, S. Nagashima, K. Takeya and H. Itokawa, Chem. Pharm. Bull. 1993, 41, 992-993.[33] R. Cozzolino, P. Palladino, F. Rossi, G. Cali, E. Benedetti and P. Laccetti, Carcinogenesis 2005, 26,

733-739.[34] C. C. Lin, T. Kimura, S. H. Wu, G. WeitzSchmidt and C. H. Wong, Bioorg. Med. Chem. Lett. 1996, 6,

2755-2760.[35] M. Buerke, A. S. Weyrich, Z. L. Zheng, F. C. A. Gaeta, M. J. Forrest and A. M. Lefer, J. Clin. Invest. 1994,

93, 1140-1148.[36] M. S. Mulligan, J. C. Paulson, S. Defrees, Z. L. Zheng, J. B. Lowe and P. A. Ward, Nature 1993, 364,

149-151.[37] T. Murohara, J. Margiotta, L. M. Phillips, J. C. Paulson, S. DeFrees, S. Zalipsky, L. S. S. Guo and A. M.

Lefer, Cardiovasc. Res. 1995, 30, 965-974.[38] P. Revill, N. Serradell, J. Bolos and E. Rosa, Drugs Fut. 2007, 32, 788-798.[39] G. J. Tanoury, M. Chen, J. E. Cochran, V. Jurkauskas and A. Looker in Vol. Vertex Pharm Inc; 2007.[40] F. G. Njoroge, K. X. Chen, N. Y. Shih and J. J. Piwinski, Acc. Chem. Res. 2008, 41, 50-59.[41] G. Wu, F. X. Chen, P. Rashatasakhon, J. M. Eckert, G. S. Wong, H. Lee, N. C. Erickson, J. A. Vance, P. C.

Nirchio, J. Weber, D. J. Tsai, Z. Nanfei, G. S. K. Wong and N. Zou in Vol. Schering Corp; Wu G; Chen F X, 2008.

Page 75: Biocatalysis & Multicomponent Reactions: The Ideal …...Biocatalysis & Multicomponent Reactions Chapter 2 Monoamine Oxidase N: 41 A Promising Biocatalyst for Asymmetric Synthesis

Highly Stereoselective Synthesis of Substituted Prolyl Peptides

75

[42] M. Wiesner, G. Upert, G. Angelici and H. Wennemers, J. Am. Chem. Soc. 2010, 132, 6-7.[43] M. Wiesner, M. Neuburger and H. Wennemers, Chem.-Eur. J. 2009, 15, 10103-10109.[44] M. Wiesner, J. D. Revell and H. Wennemers, Angew. Chem. Int. Ed. 2008, 47, 1871-1874.[45] S. Michaelis and S. Blechert, Chem.-Eur. J. 2007, 13, 2358-2368.[46] V. Kohler, K. R. Bailey, A. Znabet, J. Raftery, M. Helliwell and N. J. Turner, Angew. Chem. Int. Ed. 2010,

49, 2182-2184.[47] In case of 36h, the presence of rotameric signals prevented straightforward determination of the

d.r. After Fmoc deprotection, it became evident that a single diastereomer was formed.[48] J. L. Zhu, X. Y. Wu and S. J. Danishefsky, Tetrahedron Lett. 2009, 50, 577-579.[49] V. Theodorou, K. Skobridis, A. G. Tzakos and V. Ragoussis, Tetrahedron Lett. 2007, 48, 8230-8233.[50] N. Elders, E. Ruijter, F. J. J. de Kanter, E. Janssen, M. Lutz, A. L. Spek and R. V. A. Orru, Chem.-Eur. J.

2009, 15, 6096-6099.

Page 76: Biocatalysis & Multicomponent Reactions: The Ideal …...Biocatalysis & Multicomponent Reactions Chapter 2 Monoamine Oxidase N: 41 A Promising Biocatalyst for Asymmetric Synthesis
Page 77: Biocatalysis & Multicomponent Reactions: The Ideal …...Biocatalysis & Multicomponent Reactions Chapter 2 Monoamine Oxidase N: 41 A Promising Biocatalyst for Asymmetric Synthesis

Asymmetric Synthesis of Synthetic Alkaloids by a Tandem Biocatalysis/Ugi/

Pictet–Spengler-Type Cyclization Sequence

Published in: Chem. Commun. 2010, 46, 7706-7708

Chapter 4

Page 78: Biocatalysis & Multicomponent Reactions: The Ideal …...Biocatalysis & Multicomponent Reactions Chapter 2 Monoamine Oxidase N: 41 A Promising Biocatalyst for Asymmetric Synthesis

Chapter 4

78

Abstract: A highly efficient combination of a biocatalytic desymmetrization of 3,4-cis-

substituted meso-pyrrolidines with an Ugi-type multicomponent reaction followed by a

Pictet–Spengler-type cyclization reaction was developed. This sequence generated highly

complex alkaloid-like polycyclic compounds.

Page 79: Biocatalysis & Multicomponent Reactions: The Ideal …...Biocatalysis & Multicomponent Reactions Chapter 2 Monoamine Oxidase N: 41 A Promising Biocatalyst for Asymmetric Synthesis

Synthetic Alkaloids by Tandem Biocatalysis/Ugi/Pictet-Spengler

79

4.1 Introduction

Multicomponent reactions (MCRs)[1] are powerful tools for the synthesis of complex,

biologically relevant molecules. The atom economy of MCRs, their convergent character,

operational simplicity, and the structural diversity and complexity of the resulting molecules

make this chemistry exceptionally useful for discovery and optimization processes in

the pharmaceutical industry.[2-3] However, catalytic asymmetric methods to control the

stereochemical outcome of MCRs are scarce.[4-6]

We recently reported a novel method combining the biocatalytic desymmetrization of

3,4-cis-substituted meso-pyrrolidines[7] using engineered monoamine oxidase N (MAO-N)

from Aspergillus niger and an Ugi-type three-component reaction (MAO-N oxidation/MCR

sequence) to generate highly functionalized and optically pure 3,4-cis-substituted prolyl

peptides.[8] These prolyl peptides are of considerable interest in organocatalysis[9-12] but

also in medicinal chemistry specifically as key structural elements of clinically important

hepatitis C virus NS3 protease inhibitors such as Telaprevir (Figure 1B). We demonstrated the

utility of our method by the development of a highly efficient and convergent synthesis of

this important drug (see Chapter 5).[13]

We realized that combination of our MCR approach with cyclization reactions[14-15] could

further increase the resulting molecular complexity and diversity considerably.

Figure 1: A. Biocatalytic desymmetrization and diastereoselective Ugi-type 3CR towards optically pure 3,4-substituted prolyl peptides; B. Hepatitis C NS3 protease inhibitor telaprevir with 3,4-substituted

proline moiety highlighted.

Recently, El Kaim and coworkers described an interesting approach involving an Ugi four-

component reaction (U-4CR) between amines, aldehydes, a-ketocarboxylic acids and

homoveratryl isocyanide to afford the Ugi intermediate 5, which subsequently undergoes a

Pictet–Spengler cyclization to obtain 2,5-diketopiperazines 6 (DKPs; Scheme 1).[16]

Page 80: Biocatalysis & Multicomponent Reactions: The Ideal …...Biocatalysis & Multicomponent Reactions Chapter 2 Monoamine Oxidase N: 41 A Promising Biocatalyst for Asymmetric Synthesis

Chapter 4

80

Scheme 1: Ugi/Pictet–Spengler formation of 2,5-diketopiperazine 6 from aldehyde/ketone 1, amine 2, α-keto acid 3 and isocyanide 4a.

For medicinal purposes, DKPs (Figure 2) containing natural products are highly interesting.

The six membered ring structure of DKPs provides an increased stability in enzymatic

cleavage and allows the compound to overcome the weak metabolic properties of peptides.

Their important and wide range of physiological activities makes them even more attractive

as building block in medicinal chemistry. The broad range of bioactivities of these cyclic

dipeptides includes antitumor,[17] antiviral,[18] antifungal[19] and antibacterial activities.[20]

These scaffolds can also serve as potent oxytocin antagonists (involved in preterm birth).[21-23]

Figure 2: General structure of diketopiperazines (DKPs).

The Pictet-Spengler reaction (Scheme 2) in which β-arylethylamines undergo ringclosure

after condensation with an aldehyde or ketone is the most important method for the

synthesis of alkaloid scaffolds.[24]

Scheme 2: The mechanism of the Pictet-Spengler reaction.

Page 81: Biocatalysis & Multicomponent Reactions: The Ideal …...Biocatalysis & Multicomponent Reactions Chapter 2 Monoamine Oxidase N: 41 A Promising Biocatalyst for Asymmetric Synthesis

Synthetic Alkaloids by Tandem Biocatalysis/Ugi/Pictet-Spengler

81

The Pictet-Spengler reaction is a very efficient method which is emphasized by the fact that

nature also uses this reaction to synthesize alkaloids.[25] The first Pictet–Spenglerase, named

strictosidine synthase (STR; EC 4.3.3.2), was already discovered in 1970’s.[26-27] Different

alkaloids synthesized biosynthetically via a Pictet-Spengler were extracted from nature and

showed to exhibit therapeutically interesting properties such as ajmalicine (a hypertensive

drug),[28] ajmaline (antiarrhythmic agent, Figure 3),[29] camptothecin (anticancer, Figure 3),[30]

vinblastine (anticancer),[28] vincristine (anticancer),[28] and toxiferines (pesticide).[31]

The great potential of the Pictet-Spengler reaction has been compellingly proven in the

construction of stereochemically and structurally complex alkaloids.[24, 32-34]

Figure 3: Selection of alkaloids synthesized in nature by means of a Pictet-Spengler reaction.

We soon realized that a successive Pictet–Spengler-type cyclization could present an

efficient approach to expand the structural diversity accessible through our MAO-N

oxidation/MCR sequence.[8] Here we describe the stereoselective generation of polycyclic

DKP derivatives 10 by a novel MAO-N oxidation/Ugi MCR/ Pictet–Spengler-type cyclization

(MUPS) sequence (Scheme 3). We also envisioned using 3-(2-isocyanoethyl)-1H-indole as

isocyanide input generating alkaloids with an indole moiety.

Scheme 3: General scheme of the MUPS sequence.

Page 82: Biocatalysis & Multicomponent Reactions: The Ideal …...Biocatalysis & Multicomponent Reactions Chapter 2 Monoamine Oxidase N: 41 A Promising Biocatalyst for Asymmetric Synthesis

Chapter 4

82

The compounds (10) generated from the MUPs sequence might posses interesting

biological activity, as the core structure resemble certain mycotoxins (e.g. fumitremorgin

C). Fumitremorgin C (Figure 4) has shown to be a potent and specific inhibitor of the breast

cancer resistance protein (BCRP/ABCG2) multidrug transporter.[35]

Figure 4: Fumitremorgin C.

4.2 Results & Discussion

First, suitable conditions for the Pictet–Spengler-type cyclization in the MUPS sequence

towards 15 were developed. The cyclization of Ugi product 14, synthesized from bridged

imine 11, phenylglyoxylic acid 12 and homoveratryl isocyanide 13 was selected as the

benchmark reaction (Table 1). Based on the standard reaction conditions for the Pictet–

Spengler cyclization reported by El Kaim and coworkers,[16] we started the screening

of reaction conditions using 33% TFA in CH2Cl2. Under these conditions, no reaction

was observed after stirring at room temperature for 3 h (entry 1, Table 1). Also when the

reaction was performed in TFA as the solvent at RT (4 h; entry 2, Table 1) or under microwave

conditions at elevated temperatures (entries 3 and 4, Table 1) no DKP 15 was observed and

the starting material 14 was recovered. In another attempt, the reaction mixture was heated

to reflux in pure TFA. After 16 h, conversion to 15 was observed, but was accompanied by

significant side product formation (entry 5, Table 1).

Page 83: Biocatalysis & Multicomponent Reactions: The Ideal …...Biocatalysis & Multicomponent Reactions Chapter 2 Monoamine Oxidase N: 41 A Promising Biocatalyst for Asymmetric Synthesis

Synthetic Alkaloids by Tandem Biocatalysis/Ugi/Pictet-Spengler

83

Table 1: Optimization of the Pictet–Spengler-type cyclization of Ugi product 14 towards DKP 15a

Entry Solvent AdditiveT (°C)/Time

(min)Conversion (%)

Isolated yield (%)

1TFA/CH2Cl2

(1 : 2)- RT/180 - -

2 TFA - RT/240 - -

3 TFA - MW 150/5 - -

4 TFA - MW 200/5 - -

5 TFA - ∆/960 n.d.c, d n.d.c, d

6 CH2Cl2 TMSOTfb 10/960 100 72

a Reactions were performed with 0.073 mmol of Ugi product 14; b 1.1 equivalent of TMSOTf was added in one portion at the appropriate temperature; c Not determined; d Unidentified side products detected.

This prompted us to use trimethylsilyl triflate (TMSOTf) instead of TFA. With TMSOTf as

the Lewis acid mediator the reaction indeed took place. Stirring at -5°C did not lead to

any conversion based on HPLC-analysis (A, Figure 5), but letting the mixture slowly warm

up to room temperature led to conversion to the desired DKP 15 (B-E, Figure 5). After 23

hours of stirring HPLC-analysis showed that all of starting material was consumed and the

formation of DKP 15 had taken place. DKP 15 could be isolated in 72% yield (entry 6, Table

1). We later determined that full conversion can be reached after 16 hours at 10 ˚C (data

not presented). Remarkably, we could also observe the hemi-aminal formed prior to Pictet-

Spengler cyclization on HPLC (F, Figure 5).

Page 84: Biocatalysis & Multicomponent Reactions: The Ideal …...Biocatalysis & Multicomponent Reactions Chapter 2 Monoamine Oxidase N: 41 A Promising Biocatalyst for Asymmetric Synthesis

Chapter 4

84

HA

HA

HA HA

Ugi

Ugi

Ugi

Ugi

PS

B A

C

PS PS

PS PS

E

D

1 °C 2 h

7 °C 3 h

11 °C 4 h

15 °C 5 h

18 °C 23 h

Figure 5: A-E. HPLC analysis of the Pictet-Spengler (PS) reaction progress with respect to time and temperature. F. Heminal formation of Ugi 14 prior to Pictet-Spengler cyclization. HA = Hemiaminal

With this procedure in hand, we prepared a small set of different substituted prolyl peptides

using several α-keto acids, imines, and isocyanides (Figure 6). Thus, pyrrolidine derivatives

14, 16-21 were obtained in good yields and excellent diastereomeric ratios (Figure 6).

Page 85: Biocatalysis & Multicomponent Reactions: The Ideal …...Biocatalysis & Multicomponent Reactions Chapter 2 Monoamine Oxidase N: 41 A Promising Biocatalyst for Asymmetric Synthesis

Synthetic Alkaloids by Tandem Biocatalysis/Ugi/Pictet-Spengler

85

Figure 6: Overview of synthesized substituted prolyl peptides using optically enriched 11, α-keto acids and isocyanides.

Although these products were obtained relatively easy, not all of the desired Ugi-like

products were collected (Figure 7). Prolyl peptide 23 was not obtained due to solubility

issues of the α-keto acid and prolyl peptide 24 was not obtained probably due to side

product formation (e.g. aldol like condensations).

Figure 7: Not obtainable prolyl peptides.

The synthesized prolyl peptides were then subjected to the optimized Pictet–Spengler

cyclization conditions to provide, in most cases, the desired DKPs 15, 25-28 in moderate

to excellent yields and diastereomeric ratios (Scheme 4A). DKP 29 could not be obtained

probably due to the conjugation between the α-keto acid component and the indole group

making the keto group less electrophilic (Scheme 4B).

Page 86: Biocatalysis & Multicomponent Reactions: The Ideal …...Biocatalysis & Multicomponent Reactions Chapter 2 Monoamine Oxidase N: 41 A Promising Biocatalyst for Asymmetric Synthesis

Chapter 4

86

Scheme 4: A. Structures of fused tricyclic DKPs 15, 25–28. B. Possible explanation of failed synthesis of DKP 29.

Strikingly, using homoveratryl isocyanide as input in the synthesis of DKP 15 resulted in a

single diastereomer whereas using 3-(2-isocyanoethyl)-1H-indole in the synthesis of DKP 25

as input resulted in a diastereomeric ratio of 57:43. These results may be explained by the

fact that the mechanism of these compounds proceed via different pathways (Scheme 5).

The mechanism of 15 proceeds via a direct attack on the N-acyliminium ion (Scheme 5A)

in contrary to the mechanism of 25 which is thought to proceed via a spiro intermediate

(Scheme 5B),[36-38] although it has been postulated that cyclization can occur by direct attack.[39] This may explain the difference in diastereomeric ratios.

Scheme 5: A. Proposed mechanism of the DKP 15. B. Proposed mechanism of DKP 25.

The optimized Pictet–Spengler conditions did not produce DKPs of prolyl peptides 17

and 18 (Figure 6). In these examples the starting Ugi products 17 and 18 were recovered.

Such 5-membered ring-fused DKPs have rarely been described in literature,[40] which is

not surprising since they result from a disfavored 5-endo-trig cyclization. This explains

Page 87: Biocatalysis & Multicomponent Reactions: The Ideal …...Biocatalysis & Multicomponent Reactions Chapter 2 Monoamine Oxidase N: 41 A Promising Biocatalyst for Asymmetric Synthesis

Synthetic Alkaloids by Tandem Biocatalysis/Ugi/Pictet-Spengler

87

why harsher conditions are required to obtain such 5-membered ring-fused DKPs. Full

conversion of 17 was achieved using 1 eq. of trifluoroacetic anhydride in TFA/CH2Cl2 (1 : 1)

allowing isolation of the desired Pictet–Spengler-type product 30 in 60% yield as a single

diastereomer (Scheme 6A). Under identical conditions, however, reaction of 18 (Scheme 4)

gave the condensation product 32 instead of the desired 31. In this case, the disfavored

5-endo-trig cyclization does not occur because loss of a proton to give 32 is available as an

alternative reaction pathway for the intermediate N-acyliminium ion (Scheme 6B).

Scheme 6: A. 5-membered ring-fused DKP 30. B. Condensation product 32 forms in favor of DKP 31.

Interestingly, the use of phenylglyoxylic acid and homoveratryl isocyanide (15; Scheme

6A) as components for the Ugi reaction resulted in a diastereoselective Pictet–Spengler-

type cyclization. However, when 2-oxoisocaproic acid was employed, the corresponding

DKPs were obtained as mixtures of diastereoisomers (26-27, Scheme 6A). The X-ray crystal

structure of 15 (Figure 8) allowed the relative stereochemistry of the six asymmetric carbon

atoms to be determined. Since the absolute stereochemistry at C2, C3 and C4 of the

substituted proline residue is known and was previously reported by us,[7-8] the absolute

configuration of the molecule could be deduced.

Page 88: Biocatalysis & Multicomponent Reactions: The Ideal …...Biocatalysis & Multicomponent Reactions Chapter 2 Monoamine Oxidase N: 41 A Promising Biocatalyst for Asymmetric Synthesis

Chapter 4

88

Figure 8: Molecular structure of 15 in the crystal. Displacement ellipsoids are drawn at 50% probability level.

Subsequently, the sterically less demanding optically active 3-azabicyclo-[3,3,0]oct-2-ene

(also prepared by MAO-N catalyzed desymmetrization[7]) was used as an input for a series of

Pictet–Spengler-type substrates 33-35 (Figure 9A). The desired DKPs 36–38 were obtained

in moderate to very good yields and diastereomeric ratios (Figure 9B).

Figure 9: A. Structures of Ugi products 33–35. B. Fused tricyclic DKPs 36–38.

Page 89: Biocatalysis & Multicomponent Reactions: The Ideal …...Biocatalysis & Multicomponent Reactions Chapter 2 Monoamine Oxidase N: 41 A Promising Biocatalyst for Asymmetric Synthesis

Synthetic Alkaloids by Tandem Biocatalysis/Ugi/Pictet-Spengler

89

4.3 Conclusions & Outlook

In summary, a highly efficient synthesis of fused DKPs has been demonstrated. The

combination of MAO-N catalyzed desymmetrization of meso-pyrrolidines with an Ugi-type

3CR followed by Pictet–Spengler-type cyclization led to highly complex structures with

high diversity. In addition, the experimental procedure is simple and very efficient. To the

best of our knowledge, it also constitutes the first example of MCR chemistry to synthesize

5-membered ring-fused DKPs. The efficiency and generality of the approach and the

resulting molecular diversity and complexity makes our methodology highly interesting for

medicinal chemistry.

4.4 Experimental Section

General Information

Starting materials and solvents were purchased from ABCR and Sigma-Aldrich and were

used without treatment. 3-Azabicylo[3,3,0]octane hydrochloride was purchased from AK

Scientific. (1R,2S,6R,7S)-4-methyl-4-azatricyclo[5.2.1.02,6]dec-8-ene was prepared according

to literature procedure.[41] Column chromatography was performed on silica gel.1H and 13C NMR spectra were recorded on a Bruker Avance 400 (400.13 MHz for 1H and 100.61

MHz for 13C) or Bruker Avance 500 (500.23 MHz for 1H and 125.78 MHz for 13C) in CDCl3.

Chemical shifts are reported in d values (ppm) downfield from tetramethylsilane.

Electrospray Ionisation (ESI) mass spectrometry was carried out using a Bruker micrOTOF-Q

instrument in positive ion mode (capillary potential of 4500 V). .

Infrared (IR) spectra were recorded neat, and wavelengths are reported in cm-1. Optical

rotations were measured with a sodium lamp and are reported as follows: [α]D20 (c = g/100

mL, solvent).

2-(2-isocyanoethyl)-1H-indole, 4-(isocyanomethyl)-1,2-dimethoxybenzene and

4-(2-isocyanoethyl)-1,2-dimethoxy-benzene were synthesized according to literature

procedures.[42]

General Procedure 1: Preparation of optically active imines (3S,7R)-11 and

azabicyclo-[3,3,0]oct-2-ene

Unless stated otherwise: imines were synthesised according to literature procedure[7] with

minor adjustments. 0.7 g of freeze-dried MAO-N D5 E. coli were rehydrated for 30 min. in 20

ml of KPO4 buffer (100 mM, pH = 8.0) at 37 °C. Subsequently 1 mmol amine ((3S,7R)-11 or

azabicyclo-[3,3,0]oct-2-ene) in 30 ml of KPO4 buffer (100 mM, pH = 8.0) was prepared. The pH

of the solution was adjusted to 8.0 by addition of NaOH and then added to the rehydrated

cells. After 16-17 h the reaction was stopped (conversions were > 95 %) and worked up. For

workup the reaction mixture was centrifuged at 4000 rpm and 4 °C until the supernatant

Page 90: Biocatalysis & Multicomponent Reactions: The Ideal …...Biocatalysis & Multicomponent Reactions Chapter 2 Monoamine Oxidase N: 41 A Promising Biocatalyst for Asymmetric Synthesis

Chapter 4

90

had clarified (40 – 60 minutes). The pH of the supernatant was then adjusted to 10-11 by

addition of aq. NaOH and the supernatant was subsequently extracted with t-butyl methyl

ether or dichloromethane (4 x 70 mL). The combined organic phases were dried with Na2SO4

and concentrated at the rotary evaporator.

General procedure 2: Preparation of optically active Ugi derivatives 14, 16-22 and 34-36Unless stated otherwise: Ugi derivatives were synthesised according to literature procedure.8

Imine (0.70 mmol) was dissolved in 2 ml of CH2Cl2 followed by the addition of carboxylic

acid (0.93 mmol) and isocyanide (0.93 mmol). The reaction mixture was stirred for 24 h at RT.

CH2Cl2 (8 mL) was added and the resulting mixture was washed with Na2CO3 (2x10 mL), dried

(MgSO4), filtered, and concentrated in vacuo. Subsequently the crude product was subjected

to column chromatography (SiO2, EtOAc (1): Cyclohexane (1)). After concentration in vacuo,

the pure oily compound was dissolved in a CH2Cl2/hexane mixture and concentrated to give

the product as a solid. Note: Rotamer formation and traces of hexane might be observed in the

NMR data.

General procedure 3: Preparation of DKP derivatives 25-29 and 37-39 via a Pictet-Spengler cyclizationUnless stated otherwise: A dry 500 ml flask with activated 4 Å molecular sieves was prepared.

Ugi derivative (0.25 mmol) was dissolved in 300 ml dry CH2Cl2 and cooled down to -10°C. 1.3

eq. (0.325 mmol) of TMSOTf was dissolved in 5 ml dry CH2Cl2 and dropwise added in 5 h to

the mixture while stirring the flask. After complete addition of the TMSOTf the mixture was

allowed to warm up to room temperature. The reaction mixture was stirred for another 11

h. The resulting mixture was filtered and washed with NaHCO3 (2x20 mL), dried (MgSO4),

filtered, and concentrated in vacuo. Subsequently the crude product was subjected to

column chromatography (SiO2, EtOAc (1): Cyclohexane (1)). After concentration in vacuo, the

pure oily compound was dissolved in a CH2Cl2/hexane mixture and concentrated to give the

product as a solid. Note: Traces of hexane might be observed in the NMR data.

General procedure 4: Preparation of DKP derivatives 31 and 33 via a Pictet-Spengler cyclizationUnless stated otherwise: A dry 50 ml flask with activated 4 Å molecular sieves was prepared.

Ugi derivative (0.25 mmol) was dissolved in 10 ml dry CH2Cl2 and 10 ml TFA. 1.0 eq. (0.25

mmol) of trifluoroacetic anhydride was added in and subsequently stirred for 16 h. The

resulting mixture was filtered and washed with NaHCO3 (2x20 mL), dried (MgSO4), filtered,

and concentrated in vacuo. Subsequently the crude product was subjected to column

chromatography (SiO2, EtOAc (1): Cyclohexane (1)). After concentration in vacuo, the pure

oily compound was dissolved in a CH2Cl2/hexane mixture and concentrated to give the

product as a solid. Note: Traces of hexane might be observed in the NMR data.

Page 91: Biocatalysis & Multicomponent Reactions: The Ideal …...Biocatalysis & Multicomponent Reactions Chapter 2 Monoamine Oxidase N: 41 A Promising Biocatalyst for Asymmetric Synthesis

Synthetic Alkaloids by Tandem Biocatalysis/Ugi/Pictet-Spengler

91

Compound 14: General procedure 2 was followed using imine (94.4 mg, 0.709 mmol), phenylglyoxylic acid (138.6 mg, 0.923 mmol) and 4-(2-isocyanoethyl)-1,2-dimethoxybenzene (160.9 mg, 0.842 mmol) giving 14 as a pale yellow solid, yield 72%. [α]D

20 = -18.2 (c = 0.440, MeCN). 1H NMR (500.23 MHz, CDCl3), δ 7.91 (dd, J = 8.2, 1.1 Hz, 2H), 7.67-7.64 (m, 1H), 7.53-7.49 (m, 2H), 6.80 (d, J = 8.3 Hz, 1H), 6.75 (d, J = 1.9 Hz, 1H), 6.72-6.69 (m, 1H), 6.64 (bs, 1H), 6.25 (dd, J = 5.7, 3.0 Hz, 1H), 6.00 (dd, J = 5.7, 3.0 Hz, 1H), 4.32 (d, J = 2.1 Hz, 1H), 3.86 (s, 3H), δ 3.83 (s, 3H), 3.81-3.80 (m, 1H), 3.54-3.40 (m, 3H), 3.10 (m, 2H), 2.94-2.89 (m, 1H), 2.87-2.83 (m, 1H), 2.77 (t, J = 7.2 Hz, 2H), 1.53-1.51 (m, 1H), 1.43-1.40 (m, 1H); 13C NMR (125.78 MHz, CDCl3), δ = 191.0, 170.2, 164.8, 149.0, 147.7, 135.4, 134.9, 134.7, 132.5, 131.2, 129.9, 129.1,

120.7, 111.9, 111.4, 64.5, 62.2, 55.9, 55.8, 51.7, 49.8, 47.2, 46.6, 45.0, 41.2, 35.0; IR (neat): νmax (cm-1) = 3320 (w), 2930 (w), 1676 (m), 1628 (s), 1514 (m), 1443 (m), 1260 (s), 1234 (s), 1140 (m), 1026 (s), 714 (s), 665 (s); HRMS (ESI+) calcd for C28H31N2O5 (MH+) 475.2155, found 475.2213.

Compound 16: General procedure 2 was followed using imine (70.5 mg, 0.529 mmol), 4-methyl-2-oxopentanoic acid (89.6 mg, 85 µL, 0.688 mmol) and 4-(2-isocyanoethyl)-1,2-dimethoxybenzene (140.6 mg, 0.735 mmol) giving 16 as a yellow wax, yield 77%. 1H NMR (500.23 MHz, CDCl3), δ 6.78 (d, J = 7.9 Hz, 1H), 6.72-6.67 (m, 2H), 6.46 (bs, 1H), 6.12 (dd, J = 5.6, 3.0 Hz, 1H), 6.08 (dd, J = 5.6, 2.7 Hz, 1H), 4.16 (d, J = 1.9 Hz, 1H), δ 3.87 (s, 3H), δ 3.85 (s, 3H), 3.49-3.41 (m, 3H), 3.35 (m, 2H), 3.03-2.98 (m, 1H), 2.95-2.90 (m, 2H), 2.72-2.54 (m, 4H), 2.14 (sep, J = 6.7 Hz, 1H), δ 1.52-1.49 (m, 1H), δ 1.43-1.40 (m, 1H), 0.95 (d, J = 6.7 Hz, 3H), 0.94 (d, J = 6.7 Hz, 3H); 13C NMR (125.78 MHz, CDCl3), δ = 200.1, 170.2, 163.8, 148.9, 147.6, 134.8, 134.7, 131.2, 120.6, 111.8,

111.2, 62.8, 55.9, 55.8, 51.7, 49.8, 48.0, 47.0, 46.6, 45.7, 45.3, 40.9. 35.2, 23.9, 22.6, 22.4; IR (neat): νmax (cm-

1) = 3310 (w), 2957 (w), 1624 (s), 1530 (m), 1454 (m), 1341 (m), 1221 (m), 739 (s); HRMS (ESI+) calcd for C26H35N2O5 (MH+) 455.2468, found 455.2537.

Compound 17: General procedure 2 was followed using imine (133.0 mg, 1.0 mmol), phenylglyoxylic acid (195 mg, 1.3 mmol) and 4-(2-isocyanomethyl)-1,2-dimethoxybenzene (230 mg, 1.3 mmol) giving 17 as a yellow solid, yield 79%. [α]D

20 = +16.5 (c = 0.121, MeCN). 1H NMR (500.23 MHz, CDCl3), δ 7.86-7.84 (m, 1H), 7.65-7.62 (m, 1H), 7.51-7.45 (m, 3H), 7.00 (bs, 1H), 6.83-6.79 (m, 3H), 6.27 (dd, J = 5.7, 2.9 Hz, 1H), 6.04 (dd, J = 5.5, 2.9 Hz, 1H), 4.45 (dd, J = 14.7, 6.0 Hz, 1H), 4.36 (dd, J = 14.7, 5.6 Hz, 1H), 4.40 (d, J = 1.7 Hz, 1H), 3.86 (s, 3H), 3.84 (s, 3H), 3.52-3.47 (m, 1H), 3.12 (dd, J = 11.9, 1.9 Hz, 1H), 3.09-3.06 (m, 1H), 3.01-2.94 (m, 2H), 2.89-2.86 (m, 1H), 1.54-1.51 (m, 1H), 1.46-1.43 (m, 1H); 13C NMR (125.78 MHz, CDCl3), δ 190.7, 170.0, 165.3, 149.1, 148.3, 135.4, 134.9, 134.7, 132.4, 130.6, 129.9, 129.0, 119.8,

111.2, 110.8, 62.3, 55.9, 55.9, 51.7, 49.8, 47.0, 46.6, 46.1, 45.2, 43.5, ; IR (neat): νmax (cm-1) = 3306 (w), 2940 (w), 1632 (s), 1512 (s), 1443 (s), 1234 (s) 1138 (s), 1022 (s), 718 (s), 667 (m), 459 (m); HRMS (ESI+) calcd for C27H29N2O5 (MH+) 461.2076, found 461.2067.

Compound 18: General procedure 2 was followed using imine (70.7 mg, 0.531 mmol), 4-methyl-2-oxopentanoic acid (89.8 mg, 85 µL, 0.690 mmol) and 4-(isocyanomethyl)-1,2-dimethoxybenzene (122.2 mg, 0.690 mmol) giving 18 as yellow solid, yield 44%. 1H NMR (500.23 MHz, CDCl3): 6.84-6.75 (m, 3H), 6.13 (dd, J = 5.7, 3.0 Hz, 1H), 6.11 (dd, J = 5.5, 2.8 Hz, 1H), 4.34 (d, J = 5.8 Hz, 2H), 4.25 (d, J = 1.9 Hz, 1H), 3.86 (s, 3H), 3.85 (s, 3H), 3.55 (dd, J = 12.5, 8.8 Hz, 1H), 3.42-3.30 (m, 2H), 3.02-2.93 (m, 3H), 2.71-2.52 (m, 2H), 2.13-2.09 (m, 1H), 1.53-1.51 (m, 1H), 1.45-1.39 (m, 1H), 0.94 (d, J = 6.9 Hz, 3H), 0.92 (d, J = 6.9 Hz, 3H); 13C NMR (125.78 MHz, CDCl3), δ = 200.0, 170.1,

163.9, 149.1, 148.3, 134.9, 134.8, 130.6, 119.6, 111.1, 110.7, 62.9, 55.9, 55.8, 51.7, 51.7, 49.9, 48.1, 47.0, 46.6, 45.6, 45.4, 43.3, 23.9, 22.6, 22.5; IR (neat): νmax (cm-1) = 3308 (w), 2959 (w), 2928 (w), 1711 (w), 1630

Page 92: Biocatalysis & Multicomponent Reactions: The Ideal …...Biocatalysis & Multicomponent Reactions Chapter 2 Monoamine Oxidase N: 41 A Promising Biocatalyst for Asymmetric Synthesis

Chapter 4

92

(s), 1451 (m), 1358 (m), 1261 (s), 1234 (s), 1138 (s), 1026 (s), 729 (m); HRMS (ESI+) calcd for C25H33N2O5 (MH+) 441.2311, found 441.2387.

Compound 19: General procedure 2 was followed using imine (66.8 mg, 0.501 mmol), phenylglyoxylic acid (98.0 mg, 0.653 mmol) and 2-(2-isocyanoethyl)-1H-indole (110.8 mg, 0.651 mmol) giving 19 as a pale yellow solid, yield 77%. [α]D

20 = -18.00 (c = 0.445, MeCN). 1H NMR (500. 23MHz, CDCl3), δ 8.05 (bs, 1H), 7.89 (dd, J = 1.3, 8.3 Hz, 2H), 7.65-7.08 (m, 7H), 6.60 (bs, 1H), 6.23 (dd, J = 5.7, 3.0 Hz, 1H), 5.99 (dd, J = 5.7, 3.0 Hz, 1H), 4.31 (d, J = 2.1 Hz, 1H), 3.64-3.56 (m, 2H), 3.43-3.38 (m, 2H), 3.01-2.98 (m, 3H), 2.96-2.87 (m, 3H), 2.85-2.82 (m, 1H), 2.80-2.78 (m, 1H), 1.51-1.48 (m, 1H), 1.41-1.39 (m, 1H); 13C NMR (125.78 MHz, CDCl3), δ = 191.0, 170.1, 165.2, 136.4, 135.3, 134.9, 134.7, 132.4, 129.9, 129.0, 127.2,

122.5, 122.1, 119.5, 118.7, 112.6, 111.2, 64.5, 62.2, 51.7, 49.8, 47.0, 46.6, 46.5, 45.0, 39.8, 25.2; IR (neat): νmax (cm-1) = 3312 (w), 2930 (w), 1624 (s), 1530 (m), 1447 (m), 1343 (m), 1215 (s), 739 (s), 714 (s), 667 (s); HRMS (ESI+) calcd for C28H28N3O3 (MH+) 454.2052, found 454.2121.

Compound 20: General procedure 3 was followed using imine (70.2 mg, 0.527 mmol), 2-(furan-2-yl)-2-oxoacetic acid (96.0 mg, 0.685 mmol) and 2-(2-isocyanoethyl)-1H-indole (116.6 mg, 0.685 mmol) giving 20 as a yellow solid, yield 71%. [α]D

20 = -12.5 (c = 0.320, MeCN). 1H NMR (500.23 MHz, CDCl3): δ 8.04 (bs, 1H), 7.64 (d, J = 1.0 Hz, 1H), 7.61 (d, J = 7.9 Hz, 1H), 7.34 (d, J = 8.1 Hz, 1H), 7.24 (dd, J = 0.5, 3.7 Hz, 1H), 7.19 (t, J = 7.0 Hz, 1H), 7.14-7.10 (m, 1H), 7.06 (d, J = 2.2 Hz, 1H), 6.53 (dd, J = 1.7, 3.7 Hz, 1H), 6.46 (bs, 1H), 6.17 (dd, J = 3.0, 5.7 Hz, 1H), 6.02 (dd, J = 2.6, 5.7 Hz, 1H), 4.29 (d, J = 2.1 Hz, 1H), 3.60-3.53 (m, 2H), 3.44-3.35 (m, 2H), 3.24 (dd, J = 1.3, 11.9 Hz, 2H), 3.19-3.15 (m, 1H), 3.03-2.93 (m, 3H), 2.88-2.85 (m, 2H) , 1.49 (m, 1H). 1.41 (m, 1H); 13C NMR (125.78 MHz, CDCl3), δ = 177.6, 169.9, 163.5, 149.6, 149.1, 136.4, 135.0, 134.6,

127.2, 122.5, 122.1, 119.4, 118.6, 113.0, 112.6, 111.3, 62.5, 51.6, 49.9, 47.0, 46.7, 46.2, 45.0, 39.6, 25.1; IR (neat): νmax (cm-1) = 3314 (w), 2924 (w), 1628 (s), 1535 (w), 1452 (s), 1389 (m), 1011 (m), 741 (s), 590 (m); HRMS (ESI+) calcd for C26H26N3O4 (MH+) 444.1845, found 444.1897.

Compound 21: General procedure 3 was followed using imine (74.1 mg, 0.556 mmol), 4-methyl-2-oxopentanoic acid (94.1 mg, 89 µL, 0.723 mmol) and 2-(2-isocyanoethyl)-1H-indole (124.0 mg, 0.729 mmol) giving 21 as a yellow solid, yield 48%. [α]D

20 = -21.3 (c = 0.470, MeCN). 1H NMR (500.23 MHz, CDCl3), δ 8.04 (bs, 1H), 7.59 (d, J = 9.7 Hz, 1H), 7.35 (d, J = 9.8 Hz, 1H), 7.22-7.10 (m, 2H), 7.01 (d, J = 2.3 Hz, 1H), 6.42 (bs, 1H), 6.11-6.08 (m, 2H), 4.16 (d, J = 1.9 Hz, 1H), 3.59-3.51 (m, 2H), 3.46-3.43 (m, 1H), 3.33-3.29 (m, 2H), 3.00-2.89 (m, 4H), 2.63-2.49 (m, 2H), 2.16-2.06 (m, 1H), 1.51-1.48 (m, 1H), 1.41-1.39 (m, 1H), 0.95 (d, J = 6.7 Hz, 3H), 0.94 (d, J = 6.7 Hz, 3H); 13C NMR (125.78 MHz, CDCl3), δ = 200.2, 170.1, 163.8, 136.3, 134.8, 127.2, 122.2, 122.1, 119.4, 118.7,

112.7, 111.2, 62.8, 51.7, 49.8, 48.0, 47.0, 46.6, 45.8, 45.2, 39.7, 25.1, 23.8, 22.6, 22.4; IR (neat): νmax (cm-1) =; HRMS (ESI+) calcd for C26H32N3O3 (MH+) 434.2365, found 434.2438.

Compound 22: General procedure 2 was followed using imine (68.3 mg, 0.513 mmol), 2-(1H-indol-2-yl)-2-oxoacetic acid (130.0 mg, 0.687 mmol) and 4-(2-isocyanomethyl)-1,2-dimethoxybenzene (135.1 mg, 0.707 mmol) giving 22 as a yellow solid, yield 67%. 1H NMR (500.23 MHz, CDCl3), δ 9.43-9.27 (m, 1H), 8.40-8.33 (m, 1H), 7.69-7.67 (m, 1H), 7.48-7.41 (m, 1H), 7.37-7.29 (m, 2H), 6.83-6.67 (m, 3H), 6.65-6.59 (m, 1H), 6.17-6.13 (m, 1H), 6.09-6.05 (m, 1H), 4.31 (d, J = 1.1 Hz, 1H), 3.86 (s, 3H), 3.78 (s, 3H), 3.59-3.38 (m, 4H), 3.36-3.31 (m, 1H), 3.02 (bs, 1H), 2.96-2.83 (m, 2H), 2.83-2.69 (m, 2H), 1.50-1.46 (m, 1H), 1.40-1.37 (m, 1H); 13C NMR (125.78 MHz, CDCl3), δ 185.0, 170.9, 165.6, 148.9, 147.5, 136.4, 1361, 135.4, 134.3, 131.4, 125.4, 124.4, 122.2, 121.0, 62.4,

Page 93: Biocatalysis & Multicomponent Reactions: The Ideal …...Biocatalysis & Multicomponent Reactions Chapter 2 Monoamine Oxidase N: 41 A Promising Biocatalyst for Asymmetric Synthesis

Synthetic Alkaloids by Tandem Biocatalysis/Ugi/Pictet-Spengler

93

55.9, 55.9, 51.5, 50.0, 47.0, 46.7, 46.4, 45.1, 42.1, 41.0, 35.0; IR (neat): νmax (cm-1) = 3296 (w), 2926 (w), 1610 (s), 1514 (s), 1418 (s), 1236 (s) 1138 (s), 1026 (s), 752 (s), 711 (s), 642 (m); HRMS (ESI+) calcd for C30H31N3O5 (MH+) 513.2344, found 461.2067.

Compound 15: General procedure 3 was followed using 14 (138.3 mg, 0.292 mmol) and TMSOTf (71.3 mg, 58 µL, 0.321 mmol) giving 25 as a white solid, yield 72%. [α]D

20 = +292.0 (c = 0.210, MeCN). 1H NMR (500.23 MHz, CDCl3), δ 7.36-7.26 (m, 5H), 6.61 (s, 1H), 6.59 (s, 1H), 6.23 (dd, J = 5.7, 3.0 Hz, 1H), 5.90 (dd, J = 5.7, 3.0 Hz, 1H), 4.96 (ddd, J = 12.9, 2.5, 2.3 Hz, 1H), 3.86-3.82 (m, 1H), 3.84 (s, 3H), 3.65 (s, 3H), 3.53-3.49 (m, 1H), 3.32 (ddd, J = 25.1, 12.5, 3.3 Hz, 1H), 3.10-3.07 (m, 1H), δ 3.05-3.03 (m, 2H), δ 3.01-2.96 (m, 2H), δ 2.87-2.83 (m, 1H), δ 2.79-2.73 (m, 1H), δ 1.70-1.67 (m, 1H), 1.61-1.58 (m, 1H); 13C NMR (125.78 MHz, CDCl3), δ = 169.5, 165.7, 148.3, 146.8, 141.6, 136.9, 136.7, 128.9, 128.3, 127.2, 126.3, 125.4, 115.1, 110.3, 69.6, 61.5, 55.9,

55.8, 52.6, 50.0, 48.7, 45.7, 45.6, 43.8, 40.2, 29.0; IR (neat): νmax (cm-1) = (w), 1665 (s), 1542 (m), 1398 (s), 1261 (s), 1219 (m), 743 (s), 706 (m); HRMS (ESI+) calcd for C28H29N2O4 (MH+) 457.2049, found 457.2107.

Compound 25: General procedure 4 was followed using 19 (137.8 mg, 0.304 mmol) and TMSOTf (92.3 mg, 75 µL, 0.415 mmol) giving 25 as a pale yellow solid, yield 92%. (Note: Minor diastereomer is given in italic). 1H NMR (500.23 MHz, CDCl3): δ 9.28 (bs, 1H), 7.51-7.08 (m, 9H), 6.32-6.28 (m, 1H), 6.24 (dd, J = 5.7, 3.0 Hz, 1H), δ 4.79 (dd, J = 12.5, 5.8 Hz, 1H), 4.14-4.02 (m, 1H), 3.75 (d, J = 8.9 Hz, 1H), 3.39-3.34 (m, 1H), 3.23-3.20 (m, 1H), 3.04-2.87 (m, 3H), 2.80-2.58 (m, 3H), 1.89-1.86 (m, 1H), 1.76-1.73 (m, 1H); 1H NMR (500.23 MHz, CDCl3): δ 9.70 (bs, 1H), 7.51-7.08 (m, 9H), 6.34 (dd, J = 5.7, 3.0 Hz, 1H), 6.32-6.28 (m, 1H), δ 4.89 (dd, J = 12.6, 5.6 Hz, 1H), 4.14-4.02 (m, 1H), 3.86 (d, J = 7.9 Hz, 1H), 3.39-3.34 (m, 1H), 3.18-3.15 (m, 1H), 3.04-2.87 (m, 4H), 2.80-2.58 (m, 2H), 1.84-1.81 (m, 1H), 1.68-1.65 (m, 1H); 13C NMR (125.78 MHz, CDCl3), δ = 164.5, 163.4, 142.5, 137.7,

137.3, 126.6, 126.4, 126.2, 122.7, 119.8, 118.6, 111.4, 110.3, 60.9, 53.5, 53.1, 47.4, 45.0, 44.3, 44.0, 37.2, 20.7; 13C NMR (125.78 MHz, CDCl3), δ = 165.6, 164.0, 143.3, 137.6, 137.4, 126.4, 126.2, 126.0, 122.7, 119.6, 118.5, 111.6, 111.0, 60.6, 53.4, 53.0, 48.0, 45.3, 44.7, 43.3, 37.5, 20.6; IR (neat): νmax (cm-1) = 3318 (w), 2928 (w), 1651 (s), 1422 (s), 1300 (m), 1233 (m), 733 (s), 505 (s); HRMS (ESI+) calcd for C28H26N3O2 (MH+) 436.1947, found 436.2005.

Compound 26: General procedure 3 was followed using 16 (109.6 mg, 0.241 mmol) and TMSOTf (70.1 mg, 57 µL, 0.315 mmol) giving 26 as a white solid, yield 86%. (Note: Minor diastereomer is given in italic). 1H NMR (400.13 MHz, CDCl3), δ 7.44 (s, 1H), 6.47 (s, 1H), 6.18 (dd, J = 5.7, 3.2 Hz, 1H), 6.13 (dd, J = 5.1, 2.7 Hz, 1H), 4.74 (ddd, J = 20.4, 13.3, 7.2 Hz, 1H), 4.13 (dd, J = 12.3, 9.0 Hz, 1H), 3.78 (s, 3H), 3.76 (s, 3H), 3.43 (d, J = 8.2 Hz, 1H), 3.18-3.10 (m, 1H), 3.08-2.98 (m, 3H), 2.91-2.82 (m, 2H), 2.57-2.46 (m, 3H), 1.86 (dd, J = 4.9, 14.8 Hz, 1H), 1.76-1.74 (m, 1H), 1.62-1.60 (m, 1H), 1.50 (sep, J = 6.4 Hz, 1H), 0.82 (d, J = 6.7 Hz, 3H), 0.78 (d, J = 6.6 Hz, 3H); 1H NMR (400.13 MHz, CDCl3): δ 7.53 (s, 1H), 6.45 (s, 1H), 6.27 (dd, J = 5.7, 3.0 Hz, 1H), 6.22 (dd, J = 4.8, 2.0 Hz, 1H), 4.85-4.81 (m, 1H), 3.91-3.86 (m, 1H), 3.84 (s, 3H), 3.77 (s, 3H), 3.62 (d, J = 7.6 Hz, 1H), 3.25-3.21 (m, 1H),

3.10-3.06 (m, 1H), 2.87-2.79 (m, 4H), 2.65 (dd, J = 12.3, 7.0 Hz, 1H), 2.56-2.50 (m, 1H), 2.11 (dd, J = 14.8, 6.3 Hz, 1H), 2.00-1.94 (m, 1H), 1.77-1.72 (m, 1H), 1.62-1.53 (m, 2H), 0.84 (d, J = 6.7 Hz, 3H), 0.74 (d, J = 6.6 Hz, 3H); 13C NMR (100.61 MHz, CDCl3), δ = 167.5, 154.0, 147.3, 137.5, 137.2, 129.8, 124.9, 111.9, 108.8, 66.5, 60.0, 55.9, 55.8, 53.5, 53.4, 49.4, 45.8, 45.3, 44.5, 43.2, 36.9, 27.2, 25.2, 23.6, 21.8; 13C NMR (100.61 MHz, CDCl3), δ = 166.9, 166.2, 148.2, 146.8, 137.6, 137.3, 127.7, 127.3, 113.1, 110.9, 67.0, 60.4, 56.1, 55.8, 53.4, 52.6, 49.3, 48.3, 45.3, 44.8, 43.3, 37.8, 29.0, 24.8, 24.3, 23.7; IR (neat): νmax (cm-1) = 2953 (w), 1651 (s), 1514 (m), 1408 (s), 1256 (s), 1219 (s), 1096 (m), 737 (m); HRMS (ESI+) calcd for C26H33N2O4 (MH+) 437.2362, found 437.2419.

Page 94: Biocatalysis & Multicomponent Reactions: The Ideal …...Biocatalysis & Multicomponent Reactions Chapter 2 Monoamine Oxidase N: 41 A Promising Biocatalyst for Asymmetric Synthesis

Chapter 4

94

Compound 27: General procedure 3 was followed using 21 (110.0 mg, 0.254 mmol) and TMSOTf (73.4 mg, 60 µL, 0.330 mmol) giving 27 as a white solid, yield 90%. (Note: Minor diastereomer is given in italic).1H NMR (400.13 MHz, CDCl3), δ 9.33 (bs, 1H), 7.46 (d, J = 7.6 Hz, 1H), 7.33 (d, J = 8.1 Hz, 1H), 7.08 (t, J = 7.17 Hz, 1H), 7.06 (t, J = 7.8 Hz, 1H), 6.35 (dd, J = 5.7, 3.0 Hz, 1H), 6.31-6.29 (m, 1H), 5.03 (dd, J = 13.0, 4.5 Hz, 1H), 4.04 (dd, J = 12.4, 9.0 Hz, 1H), 3.72 (d, J = 8.0 Hz, 1H), 3.28-3.21 (m, 1H), 3.19-3.15 (m, 2H), 2.92-2.88 (3H), 2.80-2.70 (m, 2H), 2.34 (dd, J = 14.8, 6.2 Hz, 1H), δ 2.10 (dd, J = 14.7, 5.6 Hz, 1H), 1.85-1.83 (m, 1H), 1.73-1.66 (m, 2H); 1H NMR (400.13 MHz, CDCl3), δ 8.91 (bs, 1H), 7.46-7.08 (m, 1H), 6.35 (dd, J = 5.7, 3.0 Hz, 1H), 6.21 (dd, J = 5.6, 2.8 Hz, 1H), 5.03 (dd, J = 13.0, 4.5 Hz, 1H), 4.21 (dd, J = 12.3, 9.0 Hz, 1H), 3.65 (d, J = 8.5 Hz, 1H), 3.28-3.21 (m, 1H), 3.19-3.15 (m,

2H), 2.92-2.88 (3H), 2.80-2.70 (m, 2H), 2.59 (dd, J = 12.3, 8.1 Hz, 1H), δ 2.49 (dd, J = 14.7, 6.5 Hz, 1H), 1.85-1.83 (m, 1H), 1.73-1.66 (m, 2H); 13C NMR (100.61 MHz, CDCl3), δ = 165.8, 165.3, 137.6, 137.5, 136.1, 133.5, 126.3, 122.4, 119.5, 118.3, 111.5, 108.6, 65.2, 60.4, 53.5, 52.9, 48.8, 47.9, 45.2, 44.7, 43.3, 37.0, 25.0, 23.9, 23.5, 20.6; (ESI+) calcd for C26H30N3O2 (MH+) 416.2360, found 416.2313. 13C NMR (100.61 MHz, CDCl3), δ = 165.6, 163.9, 137.7, 137.1, 135.6, 133.1, 126.6, 122.4, 119.6, 118.3, 111.3, 107.3, 64.5, 60.8, 53.5, 53.1, 48.8, 47.2, 45.0, 44.3, 43.9, 36.8, 25.1, 23.6, 22.8, 20.6; IR (neat): νmax (cm-1) = 3329 (w), 2953 (w), 1649 (s), 1416 (s), 1300 (m), 1233 (m), 733 (s); HRMS (ESI+) calcd for C26H30N3O2 (MH+) 416.2360, found 416.2313.

Compound 28: General procedure 4 was followed using 20 (137.8 mg, 0.304 mmol) and TMSOTf (92.3 mg, 75 µL, 0.415 mmol) giving 28 as a white solid, yield 92%. (Note: Minor diastereomer is given in italic). 1H NMR (500.23 MHz, CDCl3): δ 9.57, 7.54-7.12 (m, 5H), 6.36-6.33 (m, 1H), 6.27-6.24 (m, 2H), 6.08 (dd, J = 3.3, 0.72 Hz, 1H), 4.84-4.81 (m, 1H), 4.27 (dd, J = 12.2, 9.0 Hz, 1H), 3.75 (d, J = 9.0 Hz, 1H), 3.40-3.33 (m, 1H), 3.21-3.19 (m, 1H), 3.10-2.75 (m, 3H), 2.61 (dd, J = 12.2, 8.3 Hz, 1H), 1.91-1.89 (m, 1H), 1.77-1.75 (m, 1H); 1H NMR (500.23 MHz, CDCl3): δ 9.28, 7.54-7.12 (m, 5H), 6.39 (dd, J = 5.8, 3.1 Hz, 1H), 6.36-6.33 (m, 1H), 6.27-6.24 (m, 1H), 5.99 (dd, J = 3.3, 0.72 Hz, 1H), 4.97-4.92 (m, 1H), 4.02-3.97 (m, 2H), 3.40-3.33 (m, 1H), 3.21-3.19 (m, 1H), 3.10-2.75 (m, 4H), 1.86-1.84 (m, 1H), 1.71-1.69 (m, 1H); 13C NMR (125.78 MHz, CDCl3): δ = 164.7, 161.4,

152.5, 142.9, 137.7, 137.1, 135.817, 127.8, 126.3, 122.8, 119.7, 118.6, 111.6, 111.3, 110.5, 110.3, 61.2, 53.6, 52.7, 47.6, 44.8, 44.2, 44.2, 36.8, 29.7, 20.6; 13C NMR (125.78 MHz, CDCl3), δ =166.9, 162.4, 151.7, 143.3, 137.6, 137.4, 136.4, 128.3, 125.9, 122.8, 119.5, 118.6, 111.7, 111.4, 110.9, 110.4, 60.6, 53.4, 52.5, 48.2, 45.4, 44.9, 43.4, 37.2, 31.6, 22.7; IR (neat): νmax (cm-1) = 3318 (w), 2928 (w), 1651 (s), 1422 (s), 1300 (m), 1233 (m), 733 (s), 505 (s); HRMS (ESI+) calcd for C26H24N3O3 (MH+) 427.1739, found 427.1904.

Compound 30: General procedure 4 was followed using 17 (110.0 mg, 0.255 mmol) and trifluoroacetic anhydride (33.7 mg, 21 µL, 0.255 mmol) giving 30 as a white solid, 60% yield. [α]D

20 = +247.6 (c = 0.210, MeCN). 1H NMR (500.23 MHz, CDCl3), δ 7.36-7.32 (m, 5H), 7.22 (s, 1H), 6.76 (s, 1H), 6.23 (dd, J = 7.1, 3.0 Hz, 1H), 5.94 (dd, J = 5.0, 2.6 Hz, 1H), 4.97 (d, J = 15.0 Hz, 1H), 4.80 (d, J = 15.0 Hz, 1H), 3.86 (s, 3H), 3.85 (s, 3H), 3.85-3.82 (m, 1H), 3.58-3.54 (m, 1H), 3.28 (d, J = 5.9 Hz, 1H), 3.09-3.03 (m, 1H), 3.00-2.97 (m, 1H), 2.94-2.91 (m, 1H), 2.87 (dd, J = 6.1, 12.5 Hz, 1H), 1.72-1.68 (m, 1H), 1.63-1.58 (m, 1H); 13C NMR (125.78 MHz, CDCl3), δ 167.7, 165.9, 150.0, 149.3, 139.8, 137.2, 136.6,

132.0, 129.2, 128.3, 125.9, 124.6, 108.3, 104.8, 75.5, 61.7, 56.1, 56.1 , 52.8, 50.9, 49.1, 48.3, 45.7, 45.5, 44.3; ; IR (neat): νmax (cm-1) = 2924 (w), 2855 (w), 1667 (s), 1408 (s), 1327 (m), 849 (m), 741 (s), 702 (m), 606 (m), 467 (w); HRMS (ESI+) calcd for C27H27N2O4 (MH+) 443.1971, found 443.1972.

Page 95: Biocatalysis & Multicomponent Reactions: The Ideal …...Biocatalysis & Multicomponent Reactions Chapter 2 Monoamine Oxidase N: 41 A Promising Biocatalyst for Asymmetric Synthesis

Synthetic Alkaloids by Tandem Biocatalysis/Ugi/Pictet-Spengler

95

Compound 32: General procedure 4 was followed using 18 (69.9 mg, 0.159 mmol) and trifluoroacetic anhydride (34.4 mg, 21 µL, 0.260 mmol) giving 32 as a pale yellow solid, 60% yield. [α]D

20 = -126.8 (c = 0.205, MeCN). 1H NMR (500.23MHz, CDCl3): δ = 6.80-6.70 (m, 3H), 6.31-6.28 (m, 2H), 5.42 (d, J = 10.0 Hz, 1H), 4.90 (d, J = 15.4, 1H), 4.74 (d, J = 15.4, 1H), 3.90 (dd, J = 9.6, 12.6 Hz, 1H), 3.67 (d, J = 6.5 Hz, 1H), 3.52-3.45 (m, 1H), 3.12- 3.09 (m, 1H), 3.08-3.02 (m, 1), 2.98-2.94 (m, 1H), 2.84 (dd, J = 6.5, 12.7 Hz, 1H), 1.79-1.77 (m, 1H), 1.68- 1.66 (m, 1H), 0.98 (d, J = 6.6, 3H), 0.90 (d, J = 6.6 Hz, 3H); 13C NMR (125.78 MHz, CDCl3), δ = 167.0, 159.2, 149.1, 148.3, 137.6, 137.0, 134.7, 129.6, 129.1, 119.4, 111.2, 110.6, 60.8, 55.9, 55.9, 53.1, 50.7, 47.7, 47.7,

45.5, 45.4, 43.8, 26.7, 23.2, 23.0; IR (neat): νmax (cm-1) = 2963 (w), 1674 (s), 1624 (s), 1516 (s), 1452 (m), 1393 (s), 1258 (s), 1138 (s), 1026 (s), 733 (s); HRMS (ESI+) calcd for C25H31N2O4 (MH+) 423.2206, found 423.2269.

Compound 33: General procedure 2 was followed using imine (54.4 mg, 0.498 mmol), phenylglyoxylic acid (123.9 mg, 0.648 mmol) and 4-(2-isocyanoethyl)-1,2-dimethoxybenzene (97.3 mg, 0.648 mmol) giving 33 as a white solid, yield 75%. [α]D

20 = +25.0 (c = 0.240, MeCN). 1H NMR (500 .23MHz, CDCl3), δ 7.96 (d, J = 7.4 Hz, 2H), δ 7.64 (t, J = 7.0 Hz, 1H), δ 7.51 (t, J = 7.6 Hz, 2H), δ 6.78-6.67 (m, 3H), δ 6.59-6.55 (m, 1H), δ 4.44 (s, 1H), δ 3.85 (s, 3H), δ 3.83 (s, 3H), δ 3.65-3.51 (m, 3H), δ 3.26 (dd, J = 2.7, 11.0 Hz, 2H), δ 3.02-2.96 (m, 1H), δ 2.81-2.71 (m, 3H), δ 2.01-1.88 (m, 1H), 1.87-1.77 (m, 1H), 1.76-1.65 (m, 1H), 1.63-1.45 (m, 2H), 1.36-1.27 (m, 1H); 13C NMR (125.78 MHz, CDCl3), δ = 190.9, 170.2, 165.9, 148.9, 147.6, 134.9, 132.6, 131.2,

129.8, 129.1, 120.7, 111.9, 111.3, 66.1, 55.8, 55.8, 53.5, 45.7, 42.8, 41.1, 40.7, 35.3, 32.8, 32.1, 25.8; IR (neat): νmax (cm-1) = 3312 (w), 2930 (w), 1676 (m), 1632 (s), 1514 (s), 1443 (s), 1260 (s), 1234 (s), 1140 (s), 1024 (s), 802 (m), 718 (s); HRMS (ESI+) calcd for C26H31N2O5 (MH+) 451.2155, found 451.2217.

Compound 34: General procedure 2 was followed using imine (54.2 mg, 0.496 mmol), phenylglyoxylic acid (97.2 mg, 0.647 mmol) and 2-(2-isocyanoethyl)-1H-indole (110.3 mg, 0.648 mmol) giving 34 as a yellow solid, yield 75%. [α]D

20 = -9.3 (c = 0.215, MeCN). 1H NMR (500.23 MHz, CDCl3), δ 8.07 (bs, 1H), 7.96-7.93 (m, 2H), 7.65-7.62 (m, 2H), 7.48-7.45 (m, 2H), 7.38-7.35 (m, 1H), 7.22-7.12 (m, 3H), 6.49 (bs, 1H), 4.44 (d, J = 2.6 Hz, 1H), 3.78-3.58 (m, 3H), 3.25 (dd, J = 11.2, 3.6 Hz, 1H) 3.05-3.01 (m, 2H), 2.88-2.84 (m, 1H), 2.77-2.74 (m, 1H), 1.99-1.23 (m, 6H); 13C NMR (125.78 MHz, CDCl3), δ = 191.1, 170.2, 166.0, 136.4, 134.9, 132.7, 129.8, 129.1, 127.3, 122.6, 122.1, 119.4, 118.7, 112.6, 111.3, 66.2, 53.5, 45.7, 42.8, 39.9, 32.9, 32.2,

25.9, 25.2; IR (neat): νmax (cm-1) = 3304 (w), 2945 (w), 1670 (m), 1628 (s), 1530 (m), 1447 (s), 1227 (s), 741 (s), 718 (s), 665 (s); HRMS (ESI+) calcd for C26H28N3O3 (MH+) 430.2052, found 430.2116.

Compound 35: General procedure 2 was followed using imine (74.7 mg, 0.684 mmol), 4-methyl-2-oxopentanoic acid (115.8 mg, 110 µL, 0.890 mmol) and 2-(2-isocyanoethyl)-1H-indole (151.4 mg, 0.890 mmol) giving 36 as an orange solid, yield 56%.[α]D

20 = -19.7 (c = 0.305, MeCN). 1H NMR (500.23MHz, CDCl3): δ = 8.10 (t, J = 21.7 Hz, 1H), 7.60 (d, J = 7.8 Hz, 1H), 7.34 (q, J = 8.0 Hz, 1H), 7.19 (p, J = 7.7 Hz, 1H), 7.12 (q, J = 7.3 Hz, 1H), 7.06-7.04 (m, 1H), 6.43-6.33 (m, 1H), 4.31 (d, J = 2.2 Hz, 1H), 3.76-3.68 (m, 1H), 3.62-3.49 (m, 2H), 3.47-3.44 (m, 1H), 3.01-2.93 (m, 3H), 2.81-2.73 (m, 1H), 2.64- 2.58 (m, 2H), 2.20-1.29 (m, 7H), 0.97-0.90 (m, 6H); 13C NMR (125.78 MHz, CDCl3), δ = 199.6, 169.2, 163.3, 135.3, 126.3, 121.2, 121.1, 118.4, 117.7, 111.7, 110.2,

65.9, 52.8, 48.4, 47.1, 43.9, 42.0, 38.7, 31.7, 31.3, 24.8, 24.1, 23.0, 21.6, 21.4. IR (neat): νmax (cm-1) = 3314 (w), 2955 (w), 1624 (s), 1532 (w), 1454 (m), 1358 (m), 1227 (m), 741 (s), 424 (s).; HRMS (ESI+) calcd for C24H32N3O3 (MH+) 410.2438, found 410.2431.

Page 96: Biocatalysis & Multicomponent Reactions: The Ideal …...Biocatalysis & Multicomponent Reactions Chapter 2 Monoamine Oxidase N: 41 A Promising Biocatalyst for Asymmetric Synthesis

Chapter 4

96

Compound 36: General procedure 3 was followed using 33 (90.0 mg, 0.200 mmol) and TMSOTf (57.8 mg, 47 µL, 0.260 mmol) giving 36 as a yellow solid, yield 62%. (Note: Minor diastereomer is given in italic). [α]D

20 = -186.7 (c = 0.225, MeCN). 1H NMR (500.23 MHz, CDCl3), δ 7.30-7.20 (m, 5H), 6.73 (s, 1H), 6.53 (s, 1H), 4.88 (dq, J = 2.4 Hz, 1H), 4.01 (dd, J = 8.6, 12.4 Hz, 1H), 3.78 (s, 3H), 3.63 (s, 3H), 3.24 (td, J = 3.3, 12.3 Hz, 1H), 3.19 (d, J = 7.7 Hz), 3.13 (dd, J = 5.2 Hz, 1H), 3.04-2.98 (m, 1H), 2.87-2.80 (m, 1H), 2.69 (dt, J = 2.7, 15.6 Hz, 1H), 2.64-2.57 (m, 1H), 1.85-1.71 (m, 2H), 1.63-1.49 (m, 2H), 1.36-1.28 (m, 1H), 1.25-1.16 (m, 1H); 1H NMR (500.23 MHz, CDCl3), δ 7.55 (s, 1H), 7.31-7.26 (m, 3H), 7.06-7.03 (m, 2H), 6.65 (s, 1H), 4.48 (ddd, J = 2.0, 7.3, 20.5 Hz, 1H), 4.42 (dd, J = 9.0,

12.3 Hz, 1H), 3.89 (s, 3H), 3.88 (s, 3H), 3.65 (d, J = 9.0 Hz, 1H), 3.15-3.07 (m, 1H), 3.07-3.00 (m, 1H), 2.83 (dd, J = 6.9, 12.4 Hz, 1H), 2.80-2.72 (m, 2H), 2.47 (ddd, J = 1.7, 5.5, 16.5 Hz, 1H), 2.03-1.97 (m, 1H), 1.91-1.82 (m, 1H), 1.79-1.71 (m, 2H), 1.69-1.59 (m, 1H), 1.53-1.46 (m, 1H); 13C NMR (125.78 MHz, CDCl3), δ = 169.2, 165.4, 148.4, 146.8, 142.0, 129.0, 128.3, 127.4, 126.3, 125.5, 115.0, 110.4, 69.5, 64.0, 55.9, 55.8, 51.6, 47.5, 40.5, 40.0, 32.1, 31.4, 29.0, 25.0; 13C NMR (125.78 MHz, CDCl3), δ = 167.2, 163.8, 148.9, 146.9, 143.2, 128.4, 127.9, 127.3, 127.2, 125.8, 112.1, 110.4, 70.2, 62.6, 56.1, 55.9, 51.4, 49.9, 39.9, 37.9, 32.1, 31.5, 31.0, 26.7, 24.7; IR (neat): νmax (cm-1) = 2934 (w), 1655 (s), 1512 (m), 1400 (s), 1258 (s), 1221 (s), 1028 (m), 750 (m), 714 (m), 698 (m); HRMS (ESI+) calcd for C26H29N2O4 (MH+) 433.2049, found 433.2109

Compound 37: General procedure 3 was followed using 34 (75.0 mg, 0.175 mmol) and TMSOTf (50.5 mg, 41 µL, 0.227 mmol) giving 37 as a white solid, yield 83%. (Note: Minor diastereomer is given in italic). 1H NMR (500.23 MHz, CDCl3): δ 9.32 (bs, 1H), δ 7.54-7.12 (m, 9H), 4.82 (dd, J = 12.6, 6.0 Hz, 1H), 4.40 (dd, J = 12.2, 9.1 Hz, 2H), 3.80 (d, J = 6.7 Hz, 1H), 3.11-2.65 (m, 6H), δ 2.16-1.50 (m, 6H); 1H NMR (500.23 MHz, CDCl3): δ 9.74 (bs, 1H), 7.54-7.12 (m, 9H), 4.92 (dd, J = 13.1, 5.1 Hz, 1H), 4.31 (dd, J = 12.7, 9.3 Hz, 1H), 3.97 (d, J = 8.9 Hz, 1H), 3.11-2.65 (m, 6H), δ 2.16-1.50 (m, 6H). 13C NMR (125.78 MHz, CDCl3), δ = 164.4, 163.4, 142.3, 135.7, 129.4, 128.7, 126.6, 126.3, 122.8, 119.8, 118.6, 111.5, 110.3, 67.1, 62.9, 50.8, 49.9, 40.5, 37.2, 31.8, 31.2, 24.6, 20.7; 13C NMR (125.78

MHz, CDCl3), δ =165.4, 163.9, 140.6, 136.2, 130.3, 129.0, 128.6, 126.4, 126.2, 119.7, 118.6, 111.6, 111.1, 67.1, 63.1, 51.4, 49.4, 39.8, 37.4, 32.2, 31.6, 24.9, 20.6; IR (neat): νmax (cm-1) = 3331 (w), 2922 (w), 1649 (s), 1422 (s), 1298 (m), 1234 (m), 739 (s), 694 (s), 577 (m), 509 (m); HRMS (ESI+) calcd for C26H26N3O2 (MH+) 412.1947, found 412.2008.

Compound 38: General procedure 3 was followed using 35 (96.8 mg, 0.236 mmol) and TMSOTf (68.3 mg, 56 µL, 0.307 mmol) giving 38 as a white solid, yield 83%. (Note: Minor diastereomer is given in italic). 1H NMR (500.23 MHz, CDCl3), δ 9.35 (bs, 1H), 7.48-7.09 (m, 4H), δ 5.06 (dd, J = 13.1, 4.7 Hz, 1H), 4.27 (dd, J = 9.3, 12.5 Hz, 1H), 3.76 (d, J = 9.0 Hz, 1H), 3.21-3.12 (m, 1H), 3.00-2.60 (m, 4H), 2.50 (dd, J = 14.7, 6.2, 1H), 2.38 (dd, J = 14.8, 6.1 Hz, 1H), 2.13 (dd, J = 14.8, 5.7 Hz, 1H), 2.00-1.48 (m, 6H), 0.94 (d, J = 4.7 Hz, 3H), 0.93 (d, J = 4.7 Hz, 3H); 1H NMR (500.23 MHz, CDCl3), δ 8.97 (bs, 1H), 7.48-7.09 (m, 4H), 5.06 (dd, J = 13.1, 4.7 Hz, 1H), 4.46 (dd, J = 12.0, 8.8 Hz, 1H), 3.67 (d, J = 9.6 Hz, 1H), 3.21-3.12 (m, 1H), ), 3.00-2.60 (m, 4H), 2.50 (dd, J = 14.7, 6.2, 1H), 2.08 (dd, J = 14.7, 5.7 Hz,

1H), 2.00-1.48 (m, 6H), 0.91 (d, J = 6.4 Hz, 3H), 0.89 (d, J = 3.6 Hz, 3H); 13C NMR (125.78 MHz, CDCl3), δ = 165.6, 165.0, 136.0, 134.0, 126.2, 122.4, 119.5, 118.3, 111.4, 108.5, 65.1, 62.8, 51.2, 49.3, 48.8, 39.6, 36.9, 32.2, 31.5, 25.0, 24.8, 23.9, 23.6, 20.6; 13C NMR (125.78 MHz, CDCl3), δ = 165.4, 163.8, 135.6, 133.1, 126.6, 122.4, 119.6, 118.3, 111.3, 107.4, 64.5, 62.7, 50.6, 49.6, 48.6, 40.4, 36.7, 31.8, 31.2, 25.0, 24.6, 23.6, 23.0, 20.6; IR (neat): νmax (cm-1) = 3349 (w), 2953 (w), 1651 (s), 1418 (s), 1298 (m), 1235 (m), 737 (s), 511 (m); HRMS (ESI+) calcd for C24H30N3O2 (MH+) 392.2260, found 392.2327.

Page 97: Biocatalysis & Multicomponent Reactions: The Ideal …...Biocatalysis & Multicomponent Reactions Chapter 2 Monoamine Oxidase N: 41 A Promising Biocatalyst for Asymmetric Synthesis

Synthetic Alkaloids by Tandem Biocatalysis/Ugi/Pictet-Spengler

97

X-ray crystal structure determination of 15

4.5 References & Notes[1] J. Zhu and H. Bienayme, Multicomponent Reactions, Wiley-VCH, Weinheim, 2005.[2] C. Hulme and V. Gore, Curr. Med. Chem. 2003, 10, 51-80.[3] I. Akritopoulou-Zanze, Curr. Opin. Chem. Biol. 2008, 12, 324-331.[4] D. J. Ramon and M. Yus, Angew. Chem. Int. Ed. 2005, 44, 1602-1634.[5] J. E. Biggs-Houck, A. Younai and J. T. Shaw, Curr. Opin. Chem. Biol. 2010, 14, 371-382.[6] P. R. Andreana, C. C. Liu and S. L. Schreiber, Org. Lett. 2004, 6, 4231-4233.[7] V. Kohler, K. R. Bailey, A. Znabet, J. Raftery, M. Helliwell and N. J. Turner, Angew. Chem. Int. Ed. 2010,

49, 2182-2184.[8] A. Znabet, E. Ruijter, F. J. J. de Kanter, V. Kohler, M. Helliwell, N. J. Turner and R. V. A. Orru, Angew.

Chem. Int. Ed. 2010, 49, 5289-5292.[9] M. Wiesner, J. D. Revell and H. Wennemers, Angew. Chem. Int. Ed. 2008, 47, 1871-1874.[10] M. Wiesner, M. Neuburger and H. Wennemers, Chem.-Eur. J. 2009, 15, 10103-10109.[11] M. Wiesner, G. Upert, G. Angelici and H. Wennemers, J. Am. Chem. Soc. 2010, 132, 6-7.[12] T. M. Chapman, I. G. Davies, B. Gu, T. M. Block, D. I. C. Scopes, P. A. Hay, S. M. Courtney, L. A. McNeill,

C. J. Schofield and B. G. Davis, J. Am. Chem. Soc. 2005, 127, 506-507.[13] A. Znabet, M. M. Polak, E. Janssen, F. J. J. de Kanter, N. J. Turner, R. V. A. Orru and E. Ruijter, Chem.

Commun. 2010, 46, 7918-7920.[14] J. D. Sunderhaus and S. E. Martin, Chem.-Eur. J. 2009, 15, 1300-1308.[15] B. Groenendaal, E. Ruijter and R. V. A. Orru, Chem. Commun. 2008, 5474-5489.

Page 98: Biocatalysis & Multicomponent Reactions: The Ideal …...Biocatalysis & Multicomponent Reactions Chapter 2 Monoamine Oxidase N: 41 A Promising Biocatalyst for Asymmetric Synthesis

Chapter 4

98

[16] L. El Kaim, M. Gageat, L. Gaultier and L. Grimaud, Synlett 2007, 500-502.[17] B. Nicholson, G. K. Lloyd, B. R. Miller, M. A. Palladino, Y. Kiso, Y. Hayashi and S. T. C. Neuteboom,

Anti-Cancer Drugs 2006, 17, 25-31.[18] S. Sinha, R. Srivastava, E. De Clereq and R. K. Singh, Nucleos Nucleot Nucl. 2004, 23, 1815-1824.[19] D. R. Houston, B. Synstad, V. G. H. Eijsink, M. J. R. Stark, I. M. Eggleston and D. M. F. van Aalten, J.

Med. Chem. 2004, 47, 5713-5720.[20] O. S. Kwon, S. H. Park, B. S. Yun, Y. R. Pyun and C. J. Kim, J. Antibiot. 2000, 53, 954-958.[21] P. G. Wyatt, M. J. Allen, A. D. Borthwick, D. E. Davies, A. M. Exall, R. J. D. Hatley, W. R. Irving, D. G.

Livermore, N. D. Miller, F. Nerozzi, S. L. Sollis and A. K. Szardenings, Bioorg. Med. Chem. Lett. 2005, 15, 2579-2582.

[22] A. D. Borthwick, D. E. Davies, A. M. Exall, R. J. D. Hatley, J. A. Hughes, W. R. Irving, D. G. Livermore, S. L. Sollis, F. Nerozzi, K. L. Valko, M. J. Allen, M. Perren, S. S. Shabbir, P. M. Woollard and M. A. Price, J. Med. Chem. 2006, 49, 4159-4170.

[23] A. D. Borthwick and J. Liddle, Med. Res. Rev. 2011, 31, 576-604.[24] W. M. Whaley and T. R. Govindachari, Org. React. 1951, 6, 151-190.[25] J. Stöckigt, A. P. Antonchick, F. Wu and H. Waldmann, Angew. Chem. Int. Ed. 2011, 50, 8538-8564.[26] J. Stockigt and M. H. Zenk, FEBS Lett. 1977, 79, 233-237.[27] J. Stockigt and M. H. Zenk, J. Chem. Soc., Chem. Commun. 1977, 646-648.[28] R. van der Heijden, D. I. Jacobs, W. Snoeijer, D. Hallared and R. Verpoorte, Curr. Med. Chem. 2004,

11, 607-628.[29] D. L. D. Keefe, R. E. Kates and D. C. Harrison, Drugs 1981, 22, 363-400.[30] Y. H. Hsiang, M. G. Lihou and L. F. Liu, Cancer Res. 1989, 49, 5077-5082.[31] W. D. M. Paton and W. L. M. Perry, Br. J. Pharmacol. Chemother. 1951, 6, 299-310.[32] M. Chrzanowska and M. D. Rozwadowska, Chem. Rev. 2004, 104, 3341-3370.[33] E. D. Cox and J. M. Cook, Chem. Rev. 1995, 95, 1797-1842.[34] B. E. Maryanoff, H. C. Zhang, J. H. Cohen, I. J. Turchi and C. A. Maryanoff, Chem. Rev. 2004, 104,

1431-1628.[35] J. D. Allen, A. van Loevezijn, J. M. Lakhai, M. van der Valk, O. van Tellingen, G. Reid, J. H. M. Schellens,

G. J. Koomen and A. H. Schinkel, Mol. Cancer Ther. 2002, 1, 417-425.[36] P. D. Bailey, S. P. Hollinshead, N. R. McLay, K. Morgan, S. J. Palmer, S. N. Prince, C. D. Reynolds and S.

D. Wood, J. Chem. Soc.-Perkin Trans. 1 1993, 431-439.[37] F. Ungemach and J. M. Cook, Heterocycles 1978, 9, 1089-1119.[38] A. H. Jackson, B. Naidoo and P. Smith, Tetrahedron 1968, 24, 6119-6129.[39] G. Casnati, A. Dossena and A. Pochini, Tetrahedron Lett. 1972, 5277-5280.[40] H. C. J. Ottenheijm, J. D. M. Herscheid, G. P. C. Kerkhoff and T. F. Spande, J. Org. Chem. 1976, 41,

3433-3438.[41] S. Michaelis and S. Blechert, Chem.-Eur. J. 2007, 13, 2358-2368.[42] N. Elders, E. Ruijter, F. J. J. de Kanter, E. Janssen, M. Lutz, A. L. Spek and R. V. A. Orru, Chem.-Eur. J.

2009, 15, 6096-6099.

Page 99: Biocatalysis & Multicomponent Reactions: The Ideal …...Biocatalysis & Multicomponent Reactions Chapter 2 Monoamine Oxidase N: 41 A Promising Biocatalyst for Asymmetric Synthesis

A Highly Effi cient Synthesis of Telaprevir® by Strategic use of Biocatalysis and

Multicomponent Reactions

Published in: Chem. Commun. 2010, 46, 7918-7920

Patents: WO2011/103932A1, WO2011/103933A1

Chapter 5

Page 100: Biocatalysis & Multicomponent Reactions: The Ideal …...Biocatalysis & Multicomponent Reactions Chapter 2 Monoamine Oxidase N: 41 A Promising Biocatalyst for Asymmetric Synthesis

Chapter 5

100

Abstract: A very short and efficient synthesis of the important drug Telaprevir, featuring

a biocatalytic desymmetrization and two multicomponent reactions as the key steps, is

presented. The classical issue of lack of stereoselectivity in Ugi- and Passerini-type reactions

is circumvented. The atom economic and convergent nature of the synthetic strategy

require only very limited use of protective groups.

Page 101: Biocatalysis & Multicomponent Reactions: The Ideal …...Biocatalysis & Multicomponent Reactions Chapter 2 Monoamine Oxidase N: 41 A Promising Biocatalyst for Asymmetric Synthesis

Telaprevir©

101

5.1 Introduction

5.1.1 Hepatitis C worldwide

Hepatitis C is one of the major diseases in the world; about 3% of the world’s population is

infected by the Hepatitis C virus (HCV), over 170 million people.

Figure 1: Global prevalence of Hepatitis C according to the world health organization (WHO).

Hepatitis C (discovered in 1989) is a virus infection spread through contact with infected

blood. Common routes of infection include sharing of needles and blood transfusions.

Consequently, from 1992 blood transfusions are always checked for HCV. Because of its

slow progression and mild symptoms, hepatitis C is hard to detect and most cases remain

unnoticed. HCV can cause liver infections, which in turn can lead to cirrhosis, liver failure and

even liver cancer. Most of the infected patients never come to this phase, but suffer from

illness and jaundice.[2]

The HCV is a member of the Flaviviridea, which are a family of viruses that are mainly

spread by arthropod vectors (like mosquitoes and ticks). The Flaviviridea family comprises

many different viruses, which can be subdivided in three groups: Genus Flavivirus, Genus

Hepacivirus and Genus Pestivirus. The Genus Flavivirus contains many identified human and

animal viruses, like yellow fever, West Nile virus and Dengue fever. Genus Pestivirus contains

viruses that infect non-human mammals. Bovine virus diarrhea is classified in this genus. HCV

is the single member of the Genus Hepacivirus, which infects mammals, including humans.

HCV is not a single organism, but is in fact a range of viruses, which are divided in different

genotypes. At present, at least six genotypes are known, which have evolved over a period

of several thousands of years. This evolution would explain the current general global

patterns of the different genotypes (Figure 2). Genotype 1 is the major genotype of the HCV.

It can be found in North & South America, Europe, Asia and Australia. Genotype 2 can also be

Page 102: Biocatalysis & Multicomponent Reactions: The Ideal …...Biocatalysis & Multicomponent Reactions Chapter 2 Monoamine Oxidase N: 41 A Promising Biocatalyst for Asymmetric Synthesis

Chapter 5

102

found in these parts of the world. Genotype 3 is most common in Australia and South Asia.

Genotypes 4 and 5 are highly prevalent in Africa and genotype 6 can be found in different

countries in Asia. These genotypes differ in their genome. The different genotypes are mostly

identified by the sequence which codes for the envelope protein (E1) and the non-structural

protein (NS5).[7] These differences can have huge influence on the sensitivity to treatment.

The boundaries and names shown and the designations used on this map do not imply the expression of any opinion whatsoever on the part of the World Health Organization concerning the legal status of any country, territory, city or area or of its authorities, or concerning the delimitation of its frontiers or boundaries. Dotted lines on maps represent approximate border lines for which there may not yet be full agreement. © WHO 2009. All rights reserved

key

Genotype 1

Genotype 2

Genotype 3

Genotype 4

Genotype 5

Genotype 6

Global distribution of HCV genotypes

Figure 2: Global distribution of different genotypes of HCV, confined to genotypes 1 to 6.

5.1.2 Life cycle of Hepatitis C

HCV is a small enveloped positive stranded RNA virus comprising of a genome of 9600

nucleotides, which encodes a single polyprotein that is co- and post-translationally

cleaved to produce different viral proteins. The N-terminal part of the polyprotein contains

the structural proteins, the core protein (C) a highly basic, nonglycosylated nucleocapsid

protein and two envelope E1 and E2 transmembrane glycoproteins. The nonstructural part

of the RNA comprises of the p7 protein which forms the ion channel, followed by the non-

structural proteins NS2, NS3, NS4A, NS4B, NS5A, and NS5B (Figure 3).

Page 103: Biocatalysis & Multicomponent Reactions: The Ideal …...Biocatalysis & Multicomponent Reactions Chapter 2 Monoamine Oxidase N: 41 A Promising Biocatalyst for Asymmetric Synthesis

Telaprevir©

103

Figure 3: The structure of the viral genome of HCV. Figure courtesy of reference [8].

A viral particle needs to undergo a multistep entry process to infect a target cell.[9-11] During

this progression every step is tightly regulated in time and space. The first step in this process

is virus entry which is initiated by the binding of a protein present at the surface of the

virion to the cell surface of the host cell (step 1, Figure 4). After initial attachment, the viral

particle is then internalized by endocytosis (step 2, Figure 4) which leads to release of the

viral genome into the cytosol (step 3, Figure 4). Then decapsidation of viral nucleocapsids

liberates free positive-strand genomic RNAs in the cell cytoplasm, where they are translated

and processed to generate the HCV polyprotein (step 4, Figure 4). The polyprotein is then

processed to generate viral proteins (step 5, Figure 4), like non-structural proteins which

assemble the replication complex (step 6, Figure 4). Subsequently, new genomic RNA is

formed and (step 7-9, Figure 4) and is released from the cell through the secretory pathway

(step 10-12, Figure 4).

Page 104: Biocatalysis & Multicomponent Reactions: The Ideal …...Biocatalysis & Multicomponent Reactions Chapter 2 Monoamine Oxidase N: 41 A Promising Biocatalyst for Asymmetric Synthesis

Chapter 5

104

Figure 4: Schematic representation of the HCV replication cycle. Figure courtesy of reference [9].

5.1.2 HCV treatment and recovery

Until very recently, the standard treatment for Hepatitis C is a cocktail of pegylated

interferon-α-2a or pegylated interferon-α-2b and the antiviral drug ribavirin for a period of

24 or 48 weeks, depending on hepatitis C virus genotype.[12] In 1986, interferon-α gave the

first effects on patients with Hepatitis C, even without the discovery of HCV. Further research

showed that pegylated interferon, in which a large molecule of poly(ethylene glycol) (PEG)

is covalently attached to interferon-α, resulted in an active molecule with a longer half-life,

better pharmacokinetic profile and higher rate of virological response.[13-14]

There are three general patterns concerning the response to antiviral therapy of Hepatitis

C (Figure 5).[12] These three patterns are: (1) sustained virological response (SVR), (2) end-of

Page 105: Biocatalysis & Multicomponent Reactions: The Ideal …...Biocatalysis & Multicomponent Reactions Chapter 2 Monoamine Oxidase N: 41 A Promising Biocatalyst for Asymmetric Synthesis

Telaprevir©

105

treatment response and relapse and (3) non-response. First, when a SVR is reached, no HCV

RNA is detected during the treatment and for at least 6 months after the treatment. The SVR

state continues in 95% of the patients after those 6 months.[15] In 10-25% of the patients

with a perfect lifestyle, a temporarily response and relapse occurs. These patients will not

experience a long-term cure of Hepatitis C. The relapse is dependent on short treatments

and inadequate or absent doses. Patients with a relapse can be treated to reach SVR, but

will need a longer treatment with higher doses. Last, non-response to treatment is shown

in one-third of the patients with Hepatitis C. These patients will finally never reach the SVR

state, and will always have HCV RNA in their bodies.

Figure 5: Virological responses during Hepatitis C therapy using pegylated interferon and ribavirin.

A 6-month course with Interferon-α leads to sustained response rates of 6-12%, and

extending treatment to 12 months raised this rate to only 16-20%.[16] Better response rates

were obtained, up to 35-40%, by combining the Interferon-α treatment with ribavirin,

which is a broad-spectrum antiviral agent.[17] Pegylated interferon-α in combination with

ribavirin gives response rates of 54-56%, which is around 15% higher than the response

rates obtained previously.[18] These results are quite encouraging, but they also mean that

40-50% of the chronic Hepatitis C patients are not cured over the years. The current therapy

is very expensive and the side effects, like depressions, make this not the ideal treatment of

patients with Hepatitis C. For many different groups of patients the results are even lower, for

example Africans, patients co-infected with HIV and patients with HCV genotype 1.

Page 106: Biocatalysis & Multicomponent Reactions: The Ideal …...Biocatalysis & Multicomponent Reactions Chapter 2 Monoamine Oxidase N: 41 A Promising Biocatalyst for Asymmetric Synthesis

Chapter 5

106

5.1.3 NS3 serine protease

The NS3 protein is an essential virally encoded protease/helicase which is responsible for

the degeneration of the nonstructural part of the protein (Figure 3 and 4). The protein that is

degenerated in this case takes part in the replication of the RNA of HCV. The NS3 protease is

a typical serine protease, similar to chrymotrypsin (Figure 6). The typical chrymotrypsin-like

fold has three catalytic active site residues positioned. The active site of the NS3 protease is

located at the N-terminus. The main function of this protease is the cleavage of the NS4A-

NS4B, NS4B-NS5A and NS5A-NS5B junctions (Figure 3).[8] These junctions self-assemble on

the endoplasmatic reticulum to generate a replicative complex. Nowadays, focus has been

set on finding potent inhibitors blocking NS3 serine protease activity, NS protein processing

and HCV RNA replication.

Figure 6: Structural representation of the NS3 serine protease.

5.1.4 NS3 serine protease inhibitors

In the past few years, a lot of research has been done to identify possible inhibitors of

the NS3 protease. During extensive research efforts different and structurally diverse

compounds have been considered. Because the NS3 protease structure contains a β-barrel

serine protease (including a canonical Asp-His-Ser catalytic triad), substrate specificity is

exclusive because of an extended polydentate substrate binding cleft, which is unique in

Page 107: Biocatalysis & Multicomponent Reactions: The Ideal …...Biocatalysis & Multicomponent Reactions Chapter 2 Monoamine Oxidase N: 41 A Promising Biocatalyst for Asymmetric Synthesis

Telaprevir©

107

NS3 protease. The important protease-substrate interactions are hydrogen bonds with the

substrate backbone and electrostatic or hydrophobic contacts along the binding site. In the

last decade, several NS3 inhibitors have reached clinical trials with mixed success (Figure 7).

Figure 7: Different inhibitors of NS3 protease; Telaprevir (Vertex), Boceprevir (Merck) and BILN 2061 (Boehringer-Ingelheim).

5.1.4.1 BILN2061

BILN 2061 was developed by Boehringer Ingelheim GmbH via an approach involving

classical medicinal chemistry, parallel synthesis and structural data for the optimization

of binding properties.[19] This was on account of the observation of an inhibitory effect on

NS3 by hexameric N-terminal cleavage products of dodecapeptide substrates.[20] BILN 2061

was the first compound of its class to reach clinical trials. It has shown oral bioavailability

and antiviral effect on patients with HCV.[21] In October 2003, Boehringer Ingelheim GmbH

released a statement on their website regarding BILN 2061: “Routine chronic safety testing

of high, supra-therapeutic doses in animals did, however, show relevant side effects which need

further analysis. Boehringer Ingelheim is currently studying the available pre-clinical data in

order to decide on their impact on the clinical development of this substance. There are currently

no trials ongoing with BILN 2061 and decisions about future trials will be made after thorough

evaluation of toxicity findings in animal studies.”[22] Probably the clinical trials of BILN 2061 are

suspended due to cardiac issues shown by monkeys. Boehringer Ingelheim GmbH stopped

reporting on BILN 2061 and the clinical trials after 2003.

5.1.4.2 Boceprevir (Victrelis™)

Schering-Plough (nowadays Merck & Co.) also developed a very interesting NS3 protease

inhibitor, Boceprevir. This drug shows great activity as an anti HCV drug in combination

with ribavirin and pegylated interferon-α with SVR response rates of 63-66% in 24-48

weeks. The search for this compound started with the screening of several compound

libraries, which did not lead to a potential lead candidate. Nevertheless, Schering-Plough

switched their strategy and focused on structure-based drug design, starting from the

Page 108: Biocatalysis & Multicomponent Reactions: The Ideal …...Biocatalysis & Multicomponent Reactions Chapter 2 Monoamine Oxidase N: 41 A Promising Biocatalyst for Asymmetric Synthesis

Chapter 5

108

three dimensional structure of the NS3 protease.[23] The focus was on trapping the catalytic

serine of the NS3 protein with conventional electrophiles such as aldehydes, ketones and

ketoamides. With most of these traps, the desired activity was not acquired. However, a

ketoamide containing an undecapeptide showed excellent activity in the inhibition of

the NS3 protease and was designated as the starting point for further drug development.

After extensive modifications, a Boceprevir-like compound was developed which did not

have the appreciable oral bioavailability in rats, dogs, and monkeys.[24] Optimization and

modifications of this compound resulted in boceprevir. This drug was FDA (Food and Drug

Administration) approved on May 13th, 2011 and will be marketed as Victrelis™.

5.1.4.3 Telaprevir (Incivek™)

Telaprevir was developed by Vertex® pharmaceuticals and Eli Lilly & Co.[25] Telaprevir in

combination with ribavirin and pegylated interferon-α shows a SVR response rate of 69-

75% in 8-12 weeks. Like with Boceprevir, Telaprevir started from a lead compound which

was modified several times to obtain a compound with very good in vitro potency, toxicity

profiles and adequate liver exposure upon oral administration. Further optimization led to

Telaprevir which showed to be the best candidate. This drug was FDA approved on May 23th,

2011 and will be market as Incivek™.

The reported synthesis of Telaprevir involves a lengthy, highly linear strategy relying heavily

on standard peptide chemistry. Yip and coworkers[25] have reported a synthesis of Telaprevir,

based on the fact that Telaprevir can be split up in three parts. These three parts can be

made independently, and combined at the end via peptide coupling chemistry to furnish

Telaprevir (Scheme 1).

Scheme 1: Reported synthesis of Telaprevir.

Page 109: Biocatalysis & Multicomponent Reactions: The Ideal …...Biocatalysis & Multicomponent Reactions Chapter 2 Monoamine Oxidase N: 41 A Promising Biocatalyst for Asymmetric Synthesis

Telaprevir©

109

Dipeptidyl acid unit 2 is readily available by standard peptide chemistry (Scheme 2).[25]

The synthesis starts with the coupling of Boc-protected L-cyclohexylglycine 5 and L-tert-

leucine methyl ester 6 to afford 7. Subsequent N-Boc deprotection and coupling with

pyrazinecarboxylic acid yielded the desired peptide 2. The overall yield of 2 over four steps

was 11%.

Scheme 2: Reported synthesis of Dipeptidyl acid 2.

The central bicyclic proline derivative was synthesized via a ten-step sequence (Scheme

3).[26-27] The synthesis starts with a tandem cycloaddition-cyclization of thiazolium ylide

11 (prepared in situ from thiazolium bromide and triethylamine) with 2-cyclopentenone,

giving a diastereomeric mixture of tetracycles 12a and 12b (d.r. 87:13). Seperation of the

diastereomers was not needed, in view of the fact that the 3-thiotetrahydrofuranyl portions

of 12a and 12b are deleted from the remainder of the molecule. Subsequent reductive

cleavage of the thiazoline C-S bond with tri-n-butyltin hydride followed by in situ hydrolysis

of the resulting hemiaminal and Cbz protection furnished 13. Chiral HPLC separation was

performed to obtain the desired enantiomer 14. Reduction of the ketone moiety with NaBH4

in ethanol afforded hydroxyl intermediate 15. Finally, removal of the hydroxy functionality

was achieved using Barton-McCombie deoxygenation reaction affording the desired

bicylic proline core 3. Because the yield for the last 2 steps (reduction of the carbonyl and

the Barton-McCombie deoxygenation) in the synthesis of this proline derivative was not

reported we made an assessment of 80% yield per step so we could for comparison reasons,

Page 110: Biocatalysis & Multicomponent Reactions: The Ideal …...Biocatalysis & Multicomponent Reactions Chapter 2 Monoamine Oxidase N: 41 A Promising Biocatalyst for Asymmetric Synthesis

Chapter 5

110

establish the overall yield. We determined the overall yield of the bicyclic proline derivative

3 synthesis to be 12% over ten steps.

Scheme 3: Reported synthesis of bicyclic proline derivative 3.

β-Amino-α-hydroxy amide 4 was synthesized[28] (Scheme 4) starting from trans-2-hexanoic

acid 17, which undergoes amide formation with cyclopropylamine in the presence of

isobutyl chloroformate and N-methyl morpholine to form compound 18. Subsequent

epoxidation of 18 with carbamide hydrogen peroxide and regiospecific ring-opening with

sodium azide led to intermediate 20. Reduction of the azide in the presence of palladium

on carbon, gave the racemic mixture of β-Amino-α-hydroxy amide 21. Chiral resolution

of amide 21 with deoxycholic acid followed by anion-exchange with hydrochloric acid,

afforded enantiomerically pure key intermediate 4. The overall yield of the synthesis of

β-Amino-α-hydroxy amide 4 was 12% over 6 steps.

Page 111: Biocatalysis & Multicomponent Reactions: The Ideal …...Biocatalysis & Multicomponent Reactions Chapter 2 Monoamine Oxidase N: 41 A Promising Biocatalyst for Asymmetric Synthesis

Telaprevir©

111

Scheme 4: Reported synthesis of hydroxyl amine 4.

Finally, condensation of dipeptidyl acid unit 2 and bicyclic proline derivative 3 furnished

tripeptide ester 23, which, after hydrolysis, underwent a coupling with amino alcohol 4

yielding the expected adduct 24. Ultimately, oxidation with Dess–Martin periodinane

afforded Telaprevir (Scheme 5).

Scheme 5: Assembly of dipeptidyl acid 2, bicyclic proline ester 3 and amino alcohol 4 into Telaprevir.

Page 112: Biocatalysis & Multicomponent Reactions: The Ideal …...Biocatalysis & Multicomponent Reactions Chapter 2 Monoamine Oxidase N: 41 A Promising Biocatalyst for Asymmetric Synthesis

Chapter 5

112

To summarize, the reported synthesis of Telaprevir is a 24 step synthesis with the longest

sequence comprising nine steps with an overall yield of about 6%.

Optimization of the synthesis of Telaprevir could significantly lower the production costs,

thereby making this drug available to an increased proportion of the world population in

the future.

5.1.4.4 Biocatalysis & Multicomponent Chemistry

Multicomponent reactions (MCRs)[29-32] have become essential tools in the efficient

generation of molecular diversity and complexity. Although MCRs have been used in

medicinal chemistry,[33] issues related to the infamously poor stereoselectivity seriously

hamper wider application of this promising methodology. The broad repertoire of

stereospecific conversions by biocatalysts[34-35] presents a unique opportunity to address the

stereoselectivity issue of certain MCRs. We recently reported the enzymatic desymmetrization

of meso-pyrrolidines by means of a monoamine oxidase N (MAO-N) from Aspergillus niger

optimized by directed evolution[36] and its combination with a highly diastereoselective

Ugi-type three-component reaction (3CR).[37] We soon realized that this methodology could

represent an efficient approach to the synthesis of Telaprevir. A retrosynthetic analysis of 1

exploiting this biocatalysis/MCR sequence is presented in Scheme 6.

The required inputs for the key Ugi-type 3CR are carboxylic acid 2, cyclic imine 25, and

isocyanide 26. The known acid 2[25] is readily available by standard peptide chemistry.

Imine 25 can be generated in situ from commercially available 27 by MAO-N catalyzed

oxidation.[36] Isocyanide 26 is accessible via β-formamido-α-ketoamide 28 which is

available via oxidative cleavage of 29 followed by in situ trapping by cyclopropyl amine.[38-39] Cyanoketophosphorane derivative 29 is prepared by coupling of the corresponding

carboxylic acid 30 which in turn is prepared from commercially available L-norvaline 31.

5.2 Results & Discussion

Thus, coupling of L-cyclohexylglycine methyl ester and pyrazinecarboxylic acid and

subsequent saponification afforded 34 in excellent yield (Scheme 7). Subsequent coupling

with L-tert-leucine methyl ester and saponification furnished the required optically pure acid

2. Compared to the reported synthesis of 2,[25] we have significantly increased both the atom

and step economy and the overall yield (74% vs. 11% over four steps). The use of carbamate

protective groups is avoided by essentially using the N-terminal pyrazine carboxylate as an

amine protective group.

Page 113: Biocatalysis & Multicomponent Reactions: The Ideal …...Biocatalysis & Multicomponent Reactions Chapter 2 Monoamine Oxidase N: 41 A Promising Biocatalyst for Asymmetric Synthesis

Telaprevir©

113

Scheme 6: Retrosynthetic analysis of Telaprevir.

Scheme 7: Synthesis of dipeptidyl acid fragment 2.

We then turned our attention to the construction of the isocyanide fragment 26a (Scheme 8).

Commercial L-norvaline 35 was N-formylated to compound 36 in 90% yield. Subsequently,

coupling of carboxylic acid 36 and (cyanomethylene)phosphorane did not result in

the formation of acylcyanomethylene(tripehenyl)phosphorane 37. We presumed the

Page 114: Biocatalysis & Multicomponent Reactions: The Ideal …...Biocatalysis & Multicomponent Reactions Chapter 2 Monoamine Oxidase N: 41 A Promising Biocatalyst for Asymmetric Synthesis

Chapter 5

114

formamide functionality interfered with the coupling of carboxylic acid 36, so we decided

to protect the amine part of L-norvaline 35 with a Boc-group to give compound 38 in 86%

yield. The coupling of carboxylic acid 38 and (cyanomethylene)phosphorane went smoothly

and resulted in Boc-protected cyanoketophosphorane 39 in a 84% yield. Subsequently,

we attempted the oxidative cleavage of Boc-protected cyanoketophosphorane 40 and

subsequent trapping with cyclopropylamine to form Boc-protected ketoamide 41.

Unfortunately, the desired product was not formed and above all, we could not determine

whether the oxidative cleavage or the nucleophilic trapping was the impeding step.

Scheme 8: Synthesis of isocyanide fragment 26a.

Therefore, an alternative procedure for the construction of isocyanide 26 was required. We

envisioned a strategy making isocyanide 26b accessible via a Passerini three-component

reaction (P-3CR)[30, 40-41] of 43, cyclopropyl isocyanide and acetic acid. Formamido aldehyde

43 can be derived from commercial (S)-2-amino-1-pentanol 44 (Scheme 9).

Page 115: Biocatalysis & Multicomponent Reactions: The Ideal …...Biocatalysis & Multicomponent Reactions Chapter 2 Monoamine Oxidase N: 41 A Promising Biocatalyst for Asymmetric Synthesis

Telaprevir©

115

Scheme 9: Retrosynthetic analysis of isocyanide fragment 26.

Commercial (S)-2-amino-1-pentanol 44 was transformed to aldehyde 43 by N-formylation

and subsequent Dess–Martin oxidation (Scheme 10). A P-3CR between 43, cyclopropyl

isocyanide and acetic acid afforded 42 as a 78 : 22 mixture of diastereomers. However,

aldehyde 43 proved difficult to isolate due to partial dimerization and the P-3CR proceeded

in disappointing yield. We then realized that these two steps might be combined in a one-

pot process. Ngouansavanh and Zhu have previously shown that the P-3CR is compatible

with hypervalent iodine oxidants such as IBX.[42] In our case, both the Dess–Martin oxidation

and the Passerini reaction are performed in CH2Cl2, and the acetic acid formed as a by-

product in the Dess–Martin oxidation can be used as the carboxylic acid input in the P-3CR.

Thus, one-pot Dess– Martin oxidation/Passerini reaction of 45 furnished 42 (60% compared

to 46% in the two-step procedure). Dehydration then afforded the required isocyanide 26 in

very good yield (87%). No racemization of the C3 stereocenter was observed.[43] This crucial

fragment is thus accessible in only three steps from commercial starting materials.

Although the P-3CR has been used to construct Telaprevir fragments similar to 26,[44] our

approach offers significant advantages compared to other approaches in terms of atom

and step economy. We replaced the carbamate groups by a formamide that is actually

incorporated in the final product.

Scheme 10: Synthesis of isocyanide fragment 26b via a P-3CR.

Page 116: Biocatalysis & Multicomponent Reactions: The Ideal …...Biocatalysis & Multicomponent Reactions Chapter 2 Monoamine Oxidase N: 41 A Promising Biocatalyst for Asymmetric Synthesis

Chapter 5

116

The next task at hand was the convergent three-component Ugi-type coupling (Scheme

11). Thus, commercial amine 27 was oxidized to imine 25 (94% ee) by MAO-N as described

previously,[36] and then combined with dipeptidyl acid 2 and isocyanide 26 to give 46. Finally,

cleavage of the acetate followed by Dess–Martin oxidation[25] gave Telaprevir (1) as a 83 :

13 : 4 mixture of diastereomers, with one minor diastereomer derived from the incomplete

stereoinduction of the Ugi-type 3CR, and the other from the minor enantiomer of imine 25.

Flash chromatography allowed straightforward separation of the diastereomers to afford

pure Telaprevir (1) in 80% yield over the last two steps.

Scheme 11: Multicomponent coupling of the fragments and completion of the synthesis of Telaprevir.

5.3 Conclusions & Outlook

We have developed a very short and efficient synthesis of the important drug candidate

Telaprevir using a biotransformation and two multicomponent reactions as the key steps. The

classical issue of lack of stereoselectivity in Ugi- and Passerini-type reactions is circumvented.

The stereogenic carbon formed in the P-3CR is later oxidized to the corresponding ketone

and the stereoselectivity of the Ugi-type 3CR is controlled by the absolute configuration of

the cyclic imine, which is in turn derived from the highly stereoselective biotransformation.

Page 117: Biocatalysis & Multicomponent Reactions: The Ideal …...Biocatalysis & Multicomponent Reactions Chapter 2 Monoamine Oxidase N: 41 A Promising Biocatalyst for Asymmetric Synthesis

Telaprevir©

117

The use of protective groups is limited to two intermediate methyl esters and an acetate.

The use of carbamate protective groups is avoided altogether. The synthesis comprises only

eleven steps in total (seven steps in the longest linear sequence) compared to twenty-four

in the originally reported procedure. The total yield (over the longest linear sequence) is

45% starting from L-cyclohexylglycine methyl ester. Our approach is general and will be

applicable to many other HCV NS3 protease inhibitors that can be derived from meso-

pyrrolidines, such as e.g. boceprevir[45-46] and narlaprevir.[44] The combination of synthetic

efficiency and convergence in our approach allows both faster development of second

generation inhibitors and a more economical production of Telaprevir. We strongly believe

that biocatalysis and MCRs will be increasingly important tools for the improvement

of atom and step economy towards the sustainable production of fine chemicals and

pharmaceuticals.

5.4 Experimental section

General Information

Commercially available starting materials and solvents were used as received. Dry CH2Cl2

was dried and distilled from CaH2 prior to use. 1H and 13C NMR spectra measured at 250.13, 400.13, or 500.23 MHz for 1H and at 62.90,

100.61, or 125.78 MHz for 13C in CDCl3 or DMSO-d6. Chemical shifts are reported in d values

(ppm) downfield from tetramethylsilane.

Column chromatography was performed on silica gel. Chromatographic purification refers

to flash chromatography using the indicated solvent (mixture) and silica (40-63 μm, 60 Å).

Thin Layer Chromatography was performed using silica plates (silica on aluminum with

fluorescence indicator). Compounds on TLC were visualized by UV detection unless stated

otherwise.

Electrospray Ionisation (ESI) mass spectrometry was carried out using a TOF-Quadrupole

instrument in positive ion mode (capillary potential of 4500 V). Infrared (IR) spectra were

recorded neat, and wavelengths are reported in cm-1. Optical rotations were measured with

a sodium lamp and are reported as follows: [α]D20 (c = g/100 mL, solvent).

Bicyclic imine 25. Imine 25 was synthesized according to literature procedure[36] with minor adjustments as follows: 2.5 g of freeze-dried MAO-N D5 E. coli were rehydrated for 30 min. in 20 ml of KPO4 buffer (100 mM, pH = 8.0) at 37 °C. Subsequently 1 mmol amine 27 in 30 ml of KPO4 buffer (100 mM, pH = 8.0) was prepared. The pH of the solution was adjusted to 8.0 by addition

of NaOH and then added to the rehydrated cells. After 16-17 h the reaction was stopped (conversion was > 95 %) and worked up. For workup the reaction mixture was centrifuged at 4000 rpm and 4°C until the supernatant had clarified (40 – 60 min.). The pH of the supernatant was then adjusted to 10-11 by addition of aq. NaOH and the supernatant was subsequently extracted with t-butyl methyl ether (4 x 70 mL). The combined organic phases were dried with Na2SO4 and concentrated at the rotary evaporator.

Page 118: Biocatalysis & Multicomponent Reactions: The Ideal …...Biocatalysis & Multicomponent Reactions Chapter 2 Monoamine Oxidase N: 41 A Promising Biocatalyst for Asymmetric Synthesis

Chapter 5

118

(S)-Methyl 2-cyclohexyl-2-(pyrazine-2-carboxamido)acetate (33). Pyrazinecarboxylic acid (2.72 g, 21.9 mmol) was added to a solution of l-cyclohexylglycine methyl ester (4.13 g, 19.9 mmol) in CH2Cl2 (100 ml) at room temperature under N2, forming a white suspension. Triethylamine (6.33 ml, 4.62 g, 45.8 mmol) was added, followed by Benzotriazol-1-yloxy-tris-(dimethylamino)-phosphonium hexafluorophosphate (BOP; 9.69 g, 21.9 mmol), which turned the

reaction mixture from purple to an orange solution. After two days of stirring at room temperature the reaction mixture was washed two times with 50 ml saturated Na2CO3, followed by the washing of the aqueous layers with CH2Cl2 (2 x 50 ml). The organic layers were collected and dried with MgSO4, followed by concentration in vacuo. Purification by silica gel flash chromatography (c-Hex:EtOAc = 2:1 with 0.5% triethylamine) afforded 33 (5.28 g, 19.03 mmol, 96%) as a yellow oil (that solidified upon standing to give a white solid). [α]20D = +42.5 (c= 1.13, CHCl3); 1H NMR (250.13 MHz, CDCl3) d = 9.39 (d, J= 1.25 Hz, 1H), 8.76 (d, J = 2.5 Hz, 1H), 8.57 (t, J = 1.5 Hz, 1H), 8.25 (d, J = 8.8 Hz, 1H), 4.74 (dd, J = 5.5, 9.3 Hz, 1H), 3.78 (s, 3H), 1.96 (m, 1H), 1.77 (m, 5H), 1.24 (m, 5H); 13C NMR (62.90 MHz, CDCl3): d= 172.0 (C), 147.4 (CH), 144.6 (CH), 142.7(CH), 57.0 (CH), 52.3 (CH3), 41.2 (CH), 29.7 (CH2), 28.4 (CH2), 26.0 (CH2); IR (neat): νmax (cm-1) = 3374 (m), 2920 (s), 2845 (w), 1740 (s), 1665 (s); HRMS (ESI, 4500 V): m/z calcd. for C14H19N3O3Na+ ([M + Na]+) 300.1319, found 300.1319.

(S)-2-cyclohexyl-2-(pyrazine-2-carboxamido)acetic acid (34). A solution of 1 M NaOH (12 ml, 12 mmol) was added to a solution of 33 (2.77 g, 10 mmol) in THF (25 ml) at 0˚C. MeOH was added to the formed suspension, to give a clear, colorless solution. The reaction mixture was stirred overnight at room temperature, followed by concentration in vacuo. The pH of the aqueous layer was set on 3.5 with a 1 M KHSO4 solution and was extracted with EtOAc (2 x 25 ml). The mixture was dried

with Na2SO4, filtered, and concentrated in vacuo, to give 34 (2.49 g, 9.45 mmol, 95%) as a white solid. [α]20D = +50.9 (c= 1.06, CHCl3); 1H NMR (250.13 MHz, CDCl3): d = 9.38 (d, J = 1.5 Hz, 1H), 8.78 (d, J = 2.5 Hz, 1H), 8.58 (dd, J = 1.5, 2.5 Hz, 1H), 8.27 (d, J = 9.0, 1H), 4.77 (dd, J = 4.3, 5.0 Hz, 1H), 2.00 (m, 1H), 1.76 (m, 5H), 1.37 (m, 5H); 13C NMR (62.90 MHz, CDCl3): d = 175.7 (C), 163.0 (C), 147.2 (CH), 144.3 (CH), 144.2 (C), 142.0 (CH), 56.9 (CH), 40.9 (CH), 29.7 (CH2), 28.1 (CH2), 25.9 (CH2); IR (neat): νmax (cm-1) = 3383 (m), 2928 (s), 2852 (w), 1713 (m), 1676 (s), 1518 (s); HRMS (ESI, 4500 V): m/z calcd. For C13H17N3O3Na+ ([M + Na]+) 286.1162, found 286.1158.

(S)-methyl-2-((S)-2-cyclohex yl-2-(pyrazine -2-carboxamido)-acetamido)-3,3-dimethyl-butanoate (9). 34 (0.653 g, 4.5 mmol) was added to a solution of H-Tle-OMe (0.653 g, 4.5 mmol) in DMF (40 ml). 1-Ethyl-3-(3-dimethylaminopropyl)-carbodiimide-HCl (EDC•HCl; 0.919 g, 6.75 mmol) was added to this colorless solution followed by 1-hydroxy-7-azabenzotriazole (HOAt; 1.035 g, 5.4 mmol) giving a bright yellow solution.

The reaction mixture was stirred for 3 days and afterwards concentrated in vacuo. The formed yellow solid was dissolved in EtOAc, washed with 40 ml saturated aqueous ammonium chloride solution and 40 ml of saturated aqueous NaHCO3 solution. The organic layers were collected, dried with MgSO4 and concentrated in vacuo to give 35 (1.48 g, 3.78 mmol, 84%) as a white solid. [α]20D = -2.0 (c= 1.0, CHCl3); 1H NMR (250.13 MHz, CDCl3): d = 9.39 (d, J = 1.5 Hz, 1H), 8.76 (d, J = 2.3 Hz, 1H), 8.55 (dd, J = 2.4, 1.8 Hz, 1H), 8.29 (d, J = 8.1, 1H), 6.40 (d, J = 9.3 Hz, 1H), 4.46 (m, 2H), 3.74 (s, 3H), 1.81 (m, 1H), 1.76 (m, 4H), 1.24 (m, 6H), 0.96 (s, 12H); 13C NMR (62.90 MHz, CDCl3): d = 171.7 (C) , 170.4 (C), 163.0 (C), 147.5 (CH), 144.5 (CH), 144.2 (C), 142.7 (CH), 60.2 (CH3), 58.4 (CH), 51.9 (CH), 40.5 (CH), 31.7 (C), 29.7 (CH2), 28.7 (CH2), 26.6 (CH3), 25.9 (CH2); IR (neat): νmax (cm-1) = 3350 (m), 2928 (m), 2853 (w), 1738 (s), 1686 (s), 1640 (s), 1520 (s); HRMS (ESI, 4500 V): m/z calcd. for C20H30N4O4Na+ ([M + Na]+) 413.2159, found 413.2169.

Page 119: Biocatalysis & Multicomponent Reactions: The Ideal …...Biocatalysis & Multicomponent Reactions Chapter 2 Monoamine Oxidase N: 41 A Promising Biocatalyst for Asymmetric Synthesis

Telaprevir©

119

(S)-2-((S)-2-cyclohexyl-2-(pyrazine-2-carboxamido)acetamido)-3,3-dimethylbutanoic acid (2). A solution of 1 M NaOH (0.94 ml, 0.94 mmol) was added to a solution of 9 (0.31 g, 0.78 mmol) in THF (3 ml) at 0˚C. MeOH was added to the formed suspension, to give a clear and colourless solution. The reaction mixture was stirred overnight at room temperature, followed by concentration in vacuo. The pH of this aqueous layer was set to 3.5 with 1

M KHSO4 and subsequently extracted with EtOAc (2 x 10ml). The mixture was dried with Na2SO4, filtered, and concentrated in vacuo, to give 2 (0.28 g, 0.75 mmol, 95%) as a white solid. [α]20D = +21.7 (c= 1.015, CHCl3); 1H NMR (250.13 MHz, CDCl3): d = 9.39 (d, J = 1.3 Hz, 1H), 8.77 (d, J = 2.5 Hz, 1H), 8.57 (dd, J = 1.5, 2.5 Hz, 1H), 8.35 (d, J = 9 Hz, 1H), 6.70 (d, J = 9.0 Hz, 1H), 4.45 (t, J = 8.8 Hz, 1H), 4.42 (d, J = 9.2 Hz, 1H), 1.94 (m, 1H), 1.71 (m, 5H), 1.20 (m, 5H), 1.01 (s, 9H); 13C NMR (62.90 MHz, CDCl3): d = 173.4 (C), 170.5 (C), 163.3 (C), 147.4 (CH), 144.4 (CH), 144.2 (C), 142.8 (CH), 58.4 (CH), 51.9 (CH), 40.4 (CH), 34.7 (C), 29.8 (CH2), 28.6 (CH2), 26.6 (CH3), 25.8 (CH2); IR (neat): νmax (cm-1) = 3335 (w), 2930 (m), 1726 (m), 1663 (s), 1514 (s); HRMS (ESI, 4500 V): m/z calc. for C19H29N4O4Na+ ([M + Na]+) 399.2003, found 399.2013.

(S)-2-formamidopentanoic acid (36). Acetic anhydride (33 ml, 325 mmol) was added

dropwise added to a solution of L-norvaline (1.17 g, 10 mmol) in formic acid (100 ml) at room temperature. After 2 hours of stirring, the formic acid was evaporated, followed by co-evaporation with toluene (twice). After that, the reaction mixture was washed with saturated Na2CO3 solution. After drying with Na2SO4 and concentration in

vacuo, 37 was obtained in quantative yield. 1H-NMR (250.13 MHz, DMSO-d6): d = 8.33 (d, J = 9.0 Hz, 1H), 8.04 (s, 1H), 4.26 (m, 1H), 1.64 (m, 2H), 1.40 (m, 2H), 0.88 (t, J = 7.4 Hz, 3H).

(S)-2-(tert-butoxycarbonylamino)pentanoic acid (38). L-norvaline (1.17 g, 10 mmol) was dissolved in a mixture of 1 M NaOH aqueous solution (10 ml) and MeCN (10 ml), resulting a white suspension. Boc2O (3.6 g, 16 mmol) in a solution of MCN (10 ml) was added dropwise to the basic reaction mixture. The MeCN was evaporated and the solution acidified to pH 2 with KHSO4 solution. The product was extracted with EtOAc

(2 x 50 ml), the organic layers were collected and after drying with MgSO4 concentrated in vacuo gave 37 as a colourless oil in a yield of 86%. 1H-NMR (250.13 MHz,CDCl3): d = 6.41 (m, 1H), 5.08 (m, 1H), 1.76-1.44 (m, 4H), 1.48 (s, 9H), 0.98 (t, J = 7.3 Hz, 3H).

Compound 39. EDC•HCl (0.21 g, 1.1 mmol) and DMAP (0.012 g, 0.1 mmol)) were added to a solution of 39 (0.22 g, 1 mmol) in CH2Cl2 (70 ml) at room temperature under N2 atmosphere. Next, cyanomethylene)phosphorane (0.60 g, 2 mmol) was added and a light yellow solution was formed. The reaction mixture was stirred overnight at room temperature, followed by concentration in vacuo. After silica gel

flash chromatography (c-hex:EtOAc = 1:1) a yield of 84%. 1H-NMR (250.13 MHz,CDCl3): d = 7.71 – 7.53 (m, 15H), 4.88 (m, 1H), 1.74-1.41 (m, 4H), 1.49 (s, 9H), 1.01 (t, J = 7.3 Hz, 3H); 13C-NMR (62.90 MHz, CDCl3): d= 190.0 (C), 150.0 (C), 133.6 (CH), 129.2 (CH), 123.1 (C), 122.3 (C), 120.9 (C), 85.4 (C), 66.0 (CH), 46.0 (C), 31.3 (CH2), 27.7 (CH3), 19.8 (CH2), 13.7 (CH3).

(S)-N-(1-hydroxypentan-2-yl)formamide (45). (S)-2-amino-1-pentanol (1.00 g, 9.7 mmol) was dissolved in ethylformate (7.84 ml, 7.19 g, 97 mmol). This reaction mixture was refluxed at 80 ˚C for 4 hours, followed by stirring overnight at room temperature. The colourless solution was concentrated in vacuo and stirred for 1 hour in a 10 mol% K2CO3 in MeOH (25 ml). Afterwards, the pH was set to 7 with

DOWEX 50wx8, followed by filtration and concentration in vacuo, which gave 45 (1.26 g, 9.61 mmol, 99%).[α]20D = -29.6 (c= 1.15, CHCl3); 1H NMR (250.13 MHz, CDCl3): d = 8.20 (s, 1H), 5.81 (b, 1H), 4.04 (m, 1H), 2.11 (b, 1H), 1.47 (m, 4H), 0.94 (t, J = 7.0 Hz, 3H); 13C NMR (62.90 MHz, CDCl3): 161.8 (C), 65.1 (CH2), 50.6 (CH), 33.2 (CH2), 19.2 (CH2), 13.9 (CH3); IR (neat): νmax (cm-1) = 3248 (s), 2957 (m), 1651 (s), 1528 (m), 1381 (m); HRMS (ESI, 4500 V): m/z calcd. for C6H13NO2Na+ ([M + Na]+) 154.0838, found 154.0835.

Page 120: Biocatalysis & Multicomponent Reactions: The Ideal …...Biocatalysis & Multicomponent Reactions Chapter 2 Monoamine Oxidase N: 41 A Promising Biocatalyst for Asymmetric Synthesis

Chapter 5

120

(S)-N-(1-oxopentan-2-yl)formamide (43). Dess-Martin periodinane (5.514 g, 13 mmol) was added to a solution of 45 (1.31 g, 10 mmol) in CH2Cl2 (100 ml) at room temperature. The white suspension was stirred for 2 days and subsequently 35 ml MeOH was added and stirred for 30 minutes. The resulting suspension was filtrated and the filtrate was concentrated in vacuo. The crude was purified by silica gel flash chromatography (c-hex:EtOAc = 1:4) to give 43 (1.08 g, 8.29 mmol, 83%)

as a white solid. NMR analysis indicates that 43 is in equilibrium with its cyclic dimer.[α]20D = +37.6 (c= 0.745, CHCl3); 1H NMR assigned to the monomer (250.13 MHz, CDCl3): d = 8.22 (s, 1H), 7.84 (s, 1H), 7.10 (m, 1H), 5.31 (m, 1H), 1.52 (m, 4H), 0.95 (m, 3H); 13C NMR assigned to the monomer (100.61 MHz, CDCl3): 198.8 (CH), 161.7 (CH), 57.4 (CH), 30.8 (CH2), 18.4 (CH2), 13.7 (CH3); 1H NMR assigned to the dimer (400.13 MHz, CDCl3) 8.22 (s, 2H), 5.26 (m, 2H), 3.72 (m, 2H) 1.52 (m, 8H), 0.95 (m, 6H;) 13C NMR (100.61 MHz, CDCl3) assigned to the dimer: 161.7 (CH), 89.8 (CH), 63.1 (CH), 30.8 (CH2), 18.4 (CH2), 13.7 (CH3); IR (neat): νmax (cm-1): 3325 (s), 2959 (s), 1649 (s), 1530 (s), 1381 (m), 1123 (w); HRMS (ESI, 4500 V): m/z calc. for C6H12NO2

+ ([M + H]+) 130.0863, found 130.0858.

(S)-1-(cyclopropylamino)-3-formamido-1-oxohexan-2-yl acetate (42). From 43: (0.892 g, 6.91 mmol) was added to a solution of cyclopropyl isocyanide (0.410 g, 6.12 mmol) in CH2Cl2 (110 ml) and stirred for 5 minutes at room temperature. Acetic acid (0.711 ml, 0.747 g, 12.44 mmol) was added and the yellow reaction was stirred for 3 days at room temperature. The reaction mixture was washed twice with 100 ml saturated Na2CO3, followed by drying with Na2SO4 and concentration in vacuo. The crude was purified by silica gel

flash chromatography (5% MeOH in CH2Cl2, 1% triethylamine). (S)-1-(Cyclopropylamino)-3-formamido-1-oxohexan-2-yl acetate 42 (0.99 g, 3.87 mmol, 56%) was obtained as a white solid as a 78:22 mixture of diastereomers. From 45: Dess Martin periodinane (5.66 g, 12.3 mmol) was added to a solution of (S)-N-(1-hydroxypentan-2-yl)formamide 45 (1.15 g, 8.8 mmol) in CH2Cl2 (12 ml) at room temperature. The white suspension was stirred for 60 minutes and subsequently cyclopropyl isocyanide (0.74 g, 10.0 mmol) was added and stirred for 48 hours. The resulting suspension was filtrated and washed twice with 10 ml saturated Na2CO3, followed by drying with Na2SO4 and concentration in vacuo. The crude product was purified by silica gel flash chromatography (5% MeOH in CH2Cl2, 1% triethylamine) to give 42 (1.34 g, 5.22 mmol, 60%) as a pale yellow solid as a 78:22 mixture of diastereomers.1H NMR (130 °C, 400.13 MHz, DMSO-d6): d = 8.03 (s, 1H), 7.52 (m, 1H), 7.30 (m, 1H), 4.89 (d, J = 4.4, 1H), 4.28 (m, 1H), 2.65 (m, 1H), 2.17(s, 3H), 1.27-1.47 (m, 4H), 0.89(t, J = 7.2, 3H), 0.63 (m, 2H), 0.48 (m, 2H); 13C NMR (125.78 MHz, DMSO-d6): d = 169.8 (C), 168.5 (C), 160.6 (CH), 74.4 (CH), 47.5 (CH), 22.2 (CH), 18.4 (CH3), 13.6 (CH3), 5.7 (CH2); IR (neat): νmax (cm-1) 3283 (s), 2961 (w), 1744 (m), 1661 (s), 1530 (s), 1238 (s); HRMS (ESI, 4500 V): m/z calcd. for C12H20N2O4Na+ ([M + Na]+) 279.1315, found 279.1325.

(S)-1-(cyclopropylamino)-3-isocyano-1-oxohexan-2-yl acetate (26b). N-methylmorpholine (0.66 ml, 0.607 g, 6.00 mmol) was added to a solution of (S)-1-(cyclopropylamino)-3-formamido-1-oxohexan-2-yl acetate 42 (0.769 g, 3.00 mmol) in CH2Cl2 (40 ml) at room temperature. The reaction mixture was cooled to -78 ˚C and triphosgene (0.312 g, 1.05 mmol) was quickly added and stirred for 5 minutes at this temperature. The resulting yellow solution was warmed up to -30 ˚C and was stirred for another 3 h. Subsequently, the reaction was quenched with

water and extracted twice with CH2Cl2 (40 ml). The organic layers were collected, dried with Na2SO4 and concentrated in vacuo. The crude product was purified by silica gel flash chromatography (2% MeOH in DCM) to give 26b (0.23 g, 0.96 mmol, 69%) as a white solid. 1H NMR (250.13 MHz, CDCl3): d= 6.28 (s, 1H), 5.25 (d, J = 2.5 Hz, 1H), 4.2 (m, 1H), 2.74 (m, 1H), 2.24 (s, 3H), 1.55 (m, 4H), 0.96 (m, 3H), 0.84 (m, 2H), 0.60 (m, 2H); 13C NMR (62.90 MHz, CDCl3): d= 169.7 (C), 168.3 (C),

Page 121: Biocatalysis & Multicomponent Reactions: The Ideal …...Biocatalysis & Multicomponent Reactions Chapter 2 Monoamine Oxidase N: 41 A Promising Biocatalyst for Asymmetric Synthesis

Telaprevir©

121

74.4 (CH), 47.5 (CH), 22.0 (CH), 20.6 (CH3), 18.5 (CH2), 13.5 (CH3), 5.5 (CH 2); IR (neat): νmax (cm-1): 3267 (s), 2959 (m), 1745 (m), 1643 (s), 1512 (m), 1221 (s); HRMS (ESI, 4500 V): m/z calcd. for C12H18N2O3Na+ ([M + Na]+) 261.1210, found 261.1214.

Compound 46. Isocyanide 26b (0.119 g, 0.50 mmol) was dropwise added to a solution of imine 25 (0.0949 g, 0.65 mmol) and carboxylic acid 2 (0.245 g, 0.65 mmol) in CH2Cl2 (5 ml) at room temperature. This yellow solution was stirred overnight and afterwards diluted with 5 ml CH2Cl2. The reaction mixture was washed twice with saturated Na2CO3 solution (10 ml) and twice with

saturated NH4Cl. The organic layers were collected, dried with MgSO4 and concentrated in vacuo. The crude product was purified by silica gel flash chromatography (5% MeOH in DCM) to give 46 (0.18 g, 0.25 mmol, 50%) as a mixture of diastereomers. 1H NMR (500.23 MHz, CDCl3): d= 9.50 (s, 1H), 8.75 (d, J = 2.5, 1H), 8.59 (s, 1H), 8.35 (d, J = 9.0, 1H), 6.84 (d, J = 9.0, 1H), 6.44 (s, 1H), 5.20 (d, J = 3.0, 1H), 4.74 (d, J= 9.5, 1H), 4.58 (t, J = 7.5, 1H), 4.38 (m, 1H), 3.37 (d, J= 6.0, 1H), 2.82 (m, 1H), 2.69 (m, 1H), 2.11 (s, 3H), 1.26 (s, 2H), 0.97 (s, 9H), 0.86 (m, 3H), 0.84-2.00 (m, 21H), 0.76 (m, 2H), 0.51 (m, 2H); 13C NMR (125.78 MHz, CDCl3): d = 170.5 (C), 169.3 (C), 162.9 (C), 147.4 (CH), 144.6 (CH), 144.2 (C), 142.8 (CH), 74.4 (CH), 66.6 (CH), 58.3 (CH), 56.6 (CH), 54.5 (CH2), 44.9 (CH), 43.0 (CH), 41.3 (CH), 35.5 (C), 26.4 (CH3), 20.8 (CH3), 19.1 (CH2), 13.8 (CH3), 6.6 (CH2); νmax (cm-1): 3306 (m), 2928 (m), 2931 (m), 1743 (w), 1655 (s), 1520 (m), 1219 (m); HRMS (ESI, 4500 V): m/z calcd. for C38H57N7O7Na+ ([M + Na]+) 746.4212, found 746.4107.

Telaprevir (1). K2CO3 (0.003 g, 0.023 mmol) was added to a solution of 46 (0.166 g, 0.23 mmol) in MeOH (2 ml) at room temperature. The reaction mixture was stirred for 2 hours at room temperature resulting in a pale yellow suspension. After full conversion, the reaction mixture was washed with 2 ml Brine, the aqueous layer was washed again with 3 ml CH2Cl2 (2x). The organic layers were collected,

dried with MgSO4 and concentrated in vacuo, to yield a pale yellow solid. The yellow solid was dissolved in CH2Cl2 (2 ml) and Dess-Martin periodinane (0.105 g, 0.25 mmol) was added at room temperature. The reaction mixture was stirred overnight before adding saturated NaHCO3 solution (2 ml) and saturated Na2S2O3 solution (2 ml). This mixture was stirred for 10 minutes, separated and the aqueous layers were washed with EtOAc (2 x 3 ml). The organic layers were collected, dried with MgSO4 and concentrated in vacuo to give the crude product as an 83:13:4 mixture of diastereomers. After silica gel flash chromatography (1% MeOH in DCM), 1 (65 mg, 0.01 mmol, 80%) was obtained as a white solid.1H-NMR (500.23 MHz, DMSO-d6): d = 9.19 (d, J = 1.4 Hz, 1H), 8.91 (d, J = 24.5 Hz, 1H), 8.76 (dd, J = 1.5, 2.5 Hz, 1H), 8.71 (d, J = 5.3 Hz, 1H), 8.49 (d, J = 9.2 Hz, 1H), 8.25 (d, J = 6.8 Hz, 1H), 8.21 (d, J = 8.9 Hz, 1H), 4.94 (m, 1H), 4.68 (dd, J = 6.5, 9.0 Hz, 1H), 4.53 (d, J = 9.0 Hz, 1H), 4.27 (d, J = 3.5 Hz, 1H), 3.74 (dd, J = 8.0, 10 Hz, 1H), 2.74 (m, 1H), 3.64 (d, J = 3.5 Hz, 1H), 0.92 (s, 9H), 0.87 (t, 3H), 0.84-1.40 (m, 23H), 0.65 (m, 2H), 0.56 (m, 2H); 13C NMR (125.78 MHz, CDCl3): d = 197.0 (C), 171.8 (C), 170.4 (C), 169.0 (C), 162.1 (C), 161.9 (C), 147.9 (CH), 144.0 (C), 143.4 (CH), 56.4 (CH), 56.3 (CH), 54.2 (CH), 53.4 (CH), 42.3 (CH), 41.3 (CH), 32.1 (CH), 31.8 (CH), 31.6 (CH), 29.1 (CH), 28.0 (CH), 26.4 (CH3); νmax (cm-1): 3302 (m), 2928 (m), 2858 (w), 1658 (s), 1620 (s), 1561 (s), 1442 (m); HRMS (ESI, 4500 V): m/z calcd. for C36H53N7O6Na+ ([M + Na]+) 702.3950, found 702.3941.

Page 122: Biocatalysis & Multicomponent Reactions: The Ideal …...Biocatalysis & Multicomponent Reactions Chapter 2 Monoamine Oxidase N: 41 A Promising Biocatalyst for Asymmetric Synthesis

Chapter 5

122

5.5 References & Notes[1] D. Lavanchy, Best. Pract. Res. Clin. Gastroenterol. 2008, 22, 991-1008.[2] J. H. Hoofnagle, Hepatology 2002, 36, 21-29.[3] N. N. Zein, Clin. Microbiol. Rev. 2000, 13, 223-235.[4] P. Simmonds, J. Gen. Virol. 2001, 82, 693-712.[5] P. Simmonds, J. Gen. Virol. 2004, 85, 3173-3188.[6] N. Kato, Acta Med. Okayama 2001, 55, 133-159.[7] P. Simmonds, D. B. Smith, F. McOmish, P. L. Yap, J. Kolberg, M. S. Urdea and E. C. Holmes, J. Gen.

Virol. 1994, 75, 1053-1061.[8] B. D. Lindenbach and C. M. Rice, Nature 2005, 436, 933-938.[9] J. M. Pawlotsky, S. Chevaliez and J. G. McHutchison, Gastroenterology 2007, 132, 1979-1998.[10] B. A. Tews, C. I. Popescu and J. Dubuisson, Viruses-Basel 2010, 2, 1782-1803.[11] J. Dubuisson, F. Helle and L. Cocquerel, Cell. Microbiol. 2008, 10, 821-827.[12] J. J. Feld and J. H. Hoofnagle, Nature 2005, 436, 967-972.[13] P. Glue, J. W. S. Fang, R. Rouzier-Panis, C. Raffanel, R. Sabo, S. K. Gupta, M. Salfi, S. Jacobs and C. I. T.

G. Hepatitis, Clin. Pharmacol. Ther. 2000, 68, 556-567.[14] K. L. Lindsay, C. Trepo, T. Heintges, M. L. Shiffman, S. C. Gordon, J. C. Hoefs, E. R. Schiff, Z. D.

Goodman, M. Laughlin, R. J. Yao and J. K. Albrecht, Hepatology 2001, 34, 395-403.[15] D. T. Y. Lau, D. E. Kleiner, M. G. Ghany, Y. Park, P. Schmid and J. H. Hoofnagle, Hepatology 1998, 28,

1121-1127.[16] A. M. Di Bisceglie and J. H. Hoofnagle, Hepatology 2002, 36, 121-127.[17] J. G. McHutchison and T. Poynard, Semin. Liver Dis. 1999, 19, 353-353.[18] M. P. Manns, J. G. McHutchison, S. C. Gordon, V. K. Rustgi, M. Shiffman, R. Reindollar, Z. D. Goodman,

K. Koury, M. H. Ling and J. K. Albrecht, Lancet 2001, 358, 958-965.[19] A. M. Faucher, M. D. Bailey, P. L. Beaulieu, C. Brochu, J. S. Duceppe, J. M. Ferland, E. Ghiro, V. Gorys,

T. Halmos, S. H. Kawai, M. Poirier, B. Simoneau, Y. S. Tsantrizos and M. Llinas-Brunet, Org. Lett. 2004, 6, 2901-2904.

[20] M. Llinas-Brunet, M. Bailey, G. Fazal, S. Goulet, T. Halmos, S. Laplante, R. Maurice, M. Poirier, M. A. Poupart, D. Thibeault, D. Wernic and D. Lamarre, Bioorg. Med. Chem. Lett. 1998, 8, 1713-1718.

[21] D. Lamarre, P. C. Anderson, M. Bailey, P. Beaulieu, G. Bolger, P. Bonneau, M. Bos, D. R. Cameron, M. Cartier, M. G. Cordingley, A. M. Faucher, N. Goudreau, S. H. Kawai, G. Kukolj, L. Lagace, S. R. LaPlante, H. Narjes, M. A. Poupart, J. Rancourt, R. E. Sentjens, R. St George, B. Simoneau, G. Steinmann, D. Thibeault, Y. S. Tsantrizos, S. M. Weldon, C. L. Yong and M. Llinas-Brunet, Nature 2003, 426, 186-189.

[22] http://us.boehringer-ingelheim.com/about/pressreleases/102803_hepatitis.html, this informa-tion was later removed from the website.

[23] S. Venkatraman, S. L. Bogen, A. Arasappan, F. Bennett, K. Chen, E. Jao, Y. T. Liu, R. Lovey, S. Hendrata, Y. H. Huang, W. D. Pan, T. Parekh, P. Pinto, V. Popov, R. Pike, S. Ruan, B. Santhanam, B. Vibulbhan, W. L. Wu, W. Y. Yang, J. S. Kong, X. Liang, J. Wong, R. Liu, N. Butkiewicz, R. Chase, A. Hart, S. Agrawal, P. Ingravallo, J. Pichardo, R. Kong, B. Baroudy, B. Malcolm, Z. Y. Guo, A. Prongay, V. Madison, L. Broske, X. M. Cui, K. C. Cheng, Y. S. Hsieh, J. M. Brisson, D. Prelusky, W. Korfmacher, R. White, S. Bogdanowich-Knipp, A. Pavlovsky, P. Bradley, A. K. Saksena, A. Ganguly, J. Piwinski, V. Girijavallabhan and F. G. Njoroge, J. Med. Chem. 2006, 49, 6074-6086.

[24] S. L. Bogen, A. Arasappan, F. Bennett, K. Chen, E. Jao, Y. T. Liu, R. G. Lovey, S. Venkatraman, W. D. Pan, T. Parekh, R. E. Pike, S. Ruan, R. Liu, B. Baroudy, S. Agrawal, R. Chase, P. Ingravallo, J. Pichardo, A. Prongay, J. M. Brisson, T. Y. Hsieh, K. C. Cheng, S. J. Kemp, O. E. Levy, M. Lim-Wilby, S. Y. Tamura, A. K. Saksena, V. Girijavallabhan and F. G. Njoroge, J. Med. Chem. 2006, 49, 2750-2757.

[25] Y. Yip, F. Victor, J. Lamar, R. Johnson, Q. M. Wang, J. I. Glass, N. Yumibe, M. Wakulchik, J. Munroe and S. H. Chen, Bioorg. Med. Chem. Lett. 2004, 14, 5007-5011.

[26] J. A. Monn and M. J. Valli, J. Org. Chem. 1994, 59, 2773-2778.

Page 123: Biocatalysis & Multicomponent Reactions: The Ideal …...Biocatalysis & Multicomponent Reactions Chapter 2 Monoamine Oxidase N: 41 A Promising Biocatalyst for Asymmetric Synthesis

Telaprevir©

123

[27] Y. Yip, F. Victor, J. Lamar, R. Johnson, Q. M. Wang, D. Barket, J. Glass, L. Jin, L. F. Liu, D. Venable, M. Wakulchik, C. P. Xie, B. Heinz, E. Villarreal, J. Colacino, N. Yumibe, M. Tebbe, J. Munroe and S. H. Chen, Bioorg. Med. Chem. Lett. 2004, 14, 251-256.

[28] F. Maltais, Y. C. Jung, M. Z. Chen, J. Tanoury, R. B. Perni, N. Mani, L. Laitinen, H. Huang, S. K. Liao, H. Y. Gao, H. Tsao, E. Block, C. Ma, R. S. Shawgo, C. Town, C. L. Brummel, D. Howe, S. Pazhanisamy, S. Raybuck, M. Namchuk and Y. L. Bennani, J. Med. Chem. 2009, 52, 7993-8001.

[29] A. Domling, Chem. Rev. 2006, 106, 17-89.[30] C. Hulme and V. Gore, Curr. Med. Chem. 2003, 10, 51-80.[31] R. V. A. Orru and M. de Greef, Synthesis 2003, 1471-1499.[32] D. J. Ramon and M. Yus, Angew. Chem. Int. Ed. 2005, 44, 1602-1634.[33] I. Akritopoulou-Zanze, Curr. Opin. Chem. Biol. 2008, 12, 324-331.[34] N. J. Turner, Nat. Chem. Biol. 2009, 5, 568-574.[35] K. Faber, Biotransformations in Organic Chemistry, Springer, Berlin, Heidelberg, New York, 1997.[36] V. Kohler, K. R. Bailey, A. Znabet, J. Raftery, M. Helliwell and N. J. Turner, Angew. Chem. Int. Ed. 2010,

49, 2182-2184.[37] A. Znabet, E. Ruijter, F. J. J. de Kanter, V. Kohler, M. Helliwell, N. J. Turner and R. V. A. Orru, Angew.

Chem. Int. Ed. 2010, 49, 5289-5292.[38] H. H. Wasserman and W. B. Ho, J. Org. Chem. 1994, 59, 4364-4366.[39] M. K. Wong, C. W. Yu, W. H. Yuen and D. Yang, J. Org. Chem. 2001, 66, 3606-3609.[40] A. Domling, Curr. Opin. Chem. Biol. 2002, 6, 306-313.[41] A. Domling and I. Ugi, Angew. Chem. Int. Ed. 2000, 39, 3169-3210.[42] T. Ngouansavanh and J. P. Zhu, Angew. Chem. Int. Ed. 2006, 45, 3495-3497.[43] J. L. Zhu, X. Y. Wu and S. J. Danishefsky, Tetrahedron Lett. 2009, 50, 577-579.[44] A. Arasappan, F. Bennett, S. L. Bogen, S. Venkatraman, M. Blackman, K. X. Chen, S. Hendrata, Y.

H. Huang, R. M. Huelgas, L. Nair, A. I. Padilla, W. D. Pan, R. Pike, P. Pinto, S. M. Ruan, M. Sannigrahi, F. Velazquez, B. Vibulbhan, W. L. Wu, W. Y. Yang, A. K. Saksena, V. Girijavallabhan, N. Y. Shih, J. S. Kong, T. Meng, Y. Jin, J. Wong, P. McNamara, A. Prongay, V. Madison, J. J. Piwinski, K. C. Cheng, R. Morrison, B. Malcolm, X. A. Tong, R. Ralston and F. G. Njoroge, ACS Med. Chem. Lett. 2010, 1, 64-69.

[45] F. G. Njoroge, K. X. Chen, N. Y. Shih and J. J. Piwinski, Acc. Chem. Res. 2008, 41, 50-59.[46] G. Wu, J. Xie, P. Rashatasakhon, F. X. Chen, M. Poirier and V. M. Sprague in Vol. US2007149459-A1

Schering Corp, 2007.

Page 124: Biocatalysis & Multicomponent Reactions: The Ideal …...Biocatalysis & Multicomponent Reactions Chapter 2 Monoamine Oxidase N: 41 A Promising Biocatalyst for Asymmetric Synthesis
Page 125: Biocatalysis & Multicomponent Reactions: The Ideal …...Biocatalysis & Multicomponent Reactions Chapter 2 Monoamine Oxidase N: 41 A Promising Biocatalyst for Asymmetric Synthesis

Chapter 6Stereoselective Synthesis of

Substituted N-Aryl Proline Amides by Biotransformation/Ugi-Smiles Sequence

Published in: Org. Biomol. Chem. accepted

Page 126: Biocatalysis & Multicomponent Reactions: The Ideal …...Biocatalysis & Multicomponent Reactions Chapter 2 Monoamine Oxidase N: 41 A Promising Biocatalyst for Asymmetric Synthesis

Chapter 6

126

Abstract: An efficient combination of MAO-N-catalyzed desymmetrization of cyclic meso-

amines with Ugi-Smiles multicomponent chemistry produced optically pure N-aryl proline

amides. This method represents the first report of a fully asymmetric Ugi-Smiles process.

Page 127: Biocatalysis & Multicomponent Reactions: The Ideal …...Biocatalysis & Multicomponent Reactions Chapter 2 Monoamine Oxidase N: 41 A Promising Biocatalyst for Asymmetric Synthesis

Stereoselective Synthesis of N-aryl Amides

127

6.1 Introduction

The synthesis of complex and challenging target molecules requires sophisticated synthetic

concepts and methodologies. Owing to their high flexibility, selectivity, and convergence,

multicomponent reactions[1-5] (MCRs) have emerged as powerful tools in the synthesis of

complex, biologically relevant molecules.[6] Undoubtedly the most widely used MCR is the

Ugi four-component reaction (U-4CR). This reaction between amines, aldehydes, carbox-

ylic acids and isocyanides (Scheme 1) affords a wide range of peptidic and peptidomimetic

products.

Scheme 1. Proposed mechanism of the Ugi four-component reaction (U-4CR).

However, no catalytic asymmetric version of the U-4CR has been described to date and, as

many MCRs, it typically suffers from very poor or unpredictable diastereoselectivity.[7-12]

Recently, we developed a highly stereoselective synthesis of substituted prolyl peptides 1

(see Chapter 3) by combining the biocatalytic desymmetrization of 3,4-cis-substituted meso-

pyrrolidines by using engineered monoamine oxidase N (MAO-N) from Aspergillus niger[13]

and an Ugi–type three-component reaction (MAO-N oxidation/MCR sequence, Scheme 2).[14]

Scheme 2: Desymmetrization and diastereoselective Ugi-type 3CR towards optically pure 3,4-substituted prolyl peptides.[14]

These prolyl peptides are of considerable interest in medicinal chemistry and organocataly-

sis. Also, this method has proven very efficient in the synthesis of alkaloid-type compounds

(3, Figure 1A, Chapter 4)[15] and Hepatitis C drug Telaprevir (Figure 1B, Chapter 5).[16]

Page 128: Biocatalysis & Multicomponent Reactions: The Ideal …...Biocatalysis & Multicomponent Reactions Chapter 2 Monoamine Oxidase N: 41 A Promising Biocatalyst for Asymmetric Synthesis

Chapter 6

128

Figure 1: A. Diastereoselective Ugi-type 3CR followed by a Pictet-Spengler reaction resulting in synthetic alkaloids.[15] B. Hepatitis C NS3 protease inhibitor Telaprevir with 3,4-substituted

proline moiety highlighted.[16]

We then decided to investigate the possibility of extending this methodology to other

MCRs. The Ugi-Smiles reaction, a variation of the Ugi reaction in which the carboxylic acid

component is substituted by an electron-deficient phenol derivative (or a heteroaromatic

analog),[17] is a promising candidate for combination with our biocatalytic oxidation due to

its mechanistic similarity with the Ugi reaction.

The Smiles rearrangement[18] represents a class of intramolecular nucleophilic aromatic sub-

stitution reaction discovered in the 1930’s incorporating a heteroatom X as the nucleophilic

component and displacing an aromatic electrophile, Y, forming an intermediate spirocyclic

anion also known as the Meisenheimer complex[19] (Scheme 3). A limitation of the Smiles

rearrangement is the requirement of an electron-deficient centre.

Scheme 3: The Smiles rearrangement.

The mechanism of the Ugi-Smiles reaction is very similar to the classic Ugi reaction. After

formation of the iminium species, which is induced by the acidic nitrophenols (pKa values

from 7 to 9) the isocyanide and nitrophenol react to form the intermediate imidate 4. This

imidate then undergoes a Smiles rearangement to form N-aryl amine 5 (Scheme 4).

Page 129: Biocatalysis & Multicomponent Reactions: The Ideal …...Biocatalysis & Multicomponent Reactions Chapter 2 Monoamine Oxidase N: 41 A Promising Biocatalyst for Asymmetric Synthesis

Stereoselective Synthesis of N-aryl Amides

129

Scheme 4: Proposed mechanism for the 4-nitrophenol Ugi–Smiles reaction.

Recently, El Kaïm and coworkers described a three-component Ugi–Smiles coupling of race-

mic five- and six-membered cyclic imines.[20]

Because the optically active 1-pyrrolines generated by MAO-N oxidation are versatile

building blocks for a diverse set of MCRs, we envisioned that combining these substituted

1-pyrrolines with the Ugi-Smiles MCR would generate highly functionalized, optically pure

3,4-substituted N-aryl proline (thio)amides (Scheme 5).

Scheme 5: Asymmetric synthesis N-aryl proline amides by MAO-N oxidation and Ugi-Smiles 3CR.

Compounds of this type have been reported to be potent VLA-4 antagonists (very late

antigen, α4β1, CD49d/CD29). This cell adhesion molecule is thought to be involved in

cell trafficking, activation, and development during normal and/or pathophysiological

processes. Therefore, these antagonists maybe useful in the treatment of diseases like

asthma, atherosclerosis, rheumatoid arthritis, inflammatory bowel disease, and multiple

sclerosis.[21-23]

Page 130: Biocatalysis & Multicomponent Reactions: The Ideal …...Biocatalysis & Multicomponent Reactions Chapter 2 Monoamine Oxidase N: 41 A Promising Biocatalyst for Asymmetric Synthesis

Chapter 6

130

6.2 Results & Discussion

We started our investigation by finding the most suitable conditions for the Ugi-Smiles MCR.

The reaction of bridged imine 10, 2-nitrophenol (11) and tert-butyl isocyanide (12) was des-

ignated as benchmark reaction (Table 1). We began by using the conditions reported by El

Kaïm et al.[24] Unfortunately, these conditions afforded 13 in moderate yield at best (entries 1

and 2, Table 1). Therefore, we opted to extend the reaction time and raise the temperature.

However, this did not have any effect and the isolated yield remained more or less the same

(entries 3-6, Table 1). Neither microwave irradiation of the reaction mixture (entries 7-9, Ta-

ble 1) nor the use of Sc(OTf )3 as additive (entries 10 and 11, Table 1) led to an improved yield.

In a final attempt to improve the yield we used a two-fold excess of the imine (entries 12-14,

Table 1). Performing the reaction at room temperature afforded the product in low yield

(entry 12, Table 1). To our delight increasing the reaction temperature to 40 oC increased

the yield to a satisfactory 73% (entry 13, Table 1). Elevating the temperature to 60 oC led to

a significant drop in yield most likely due to the instability of the imine (entry 14, Table 1) at

this temperature.[25]

Table 1: Optimization of the Ugi-Smiles 3CR of bridged imine 10, 2-nitrophenol (11) and tert-butyl isocyanide (12).

Entry SolventTemperature

(oC)Reaction

TimeAdditives 10 : 11 : 12a Yield (%)

1 MeOH 60 24 h - 1.0 : 1.3 : 1.3 53b

2 Toluene 80 24 h - 1.0 : 1.3 : 1.3 26b

3 MeOH 60 48 h - 1.0 : 1.3 : 1.3 55b

4 Toluene 80 48 h - 1.0 : 1.3 : 1.3 25b

5 MeOH RT 72 h - 1.0 : 1.3 : 1.3 39b

6 MeOH 40 24 h - 1.0 : 1.3 : 1.3 43b

7 MeOH 125 30 mind - 1.0 : 1.3 : 1.3 40c

8 MeOH 150 30 mind - 1.0 : 1.3 : 1.3 40c

9 TFE 135 30 mind - 1.0 : 1.3 : 1.3 <10c

10 MeOH RT 24 h Sc(OTf)3 1.0 : 1.3 : 1.3 34b

11 MeOH 60 24 h Sc(OTf)3 1.0 : 1.3 : 1.3 24b

12 MeOH RT 24 h - 2.0 : 1.0 : 1.5 33b

13 MeOH 40 24 h - 2.0 : 1.0 : 1.5 73b

14 MeOH 60 24 h - 2.0 : 1.0 : 1.5 44b

a Molar ratios of the reactants; b Isolated yield; c Yield estimated by HPLC analysis; d Microwave irradiation in closed vessel; MeOH = methanol; TFE = 2,2,2-trifluoroethanol; Sc(OTf)3 = scandium triflate; RT = room temperature.

Page 131: Biocatalysis & Multicomponent Reactions: The Ideal …...Biocatalysis & Multicomponent Reactions Chapter 2 Monoamine Oxidase N: 41 A Promising Biocatalyst for Asymmetric Synthesis

Stereoselective Synthesis of N-aryl Amides

131

With this procedure in hand we turned our attention to the synthesis of a small set of Ugi-

Smiles products. Starting from bridged imine 10, various phenols and isocyanides were em-

ployed to generate the corresponding Ugi-Smiles products 14-24 (Figure 2).

As reported in Figure 1, the combination of ortho-nitrophenol and para-nitrophenol with

different isocyanides gave the desired Ugi-Smiles products in moderate to good yields as

single diastereomers (13-17, Figure 2). 3-Nitro-2-pyridinol and 2-mercaptopyrimidine were

tolerated as inputs and gave the desired products in moderate to good yields as single dia-

stereomers as well (18-21, Figure 2). 2-Pyridylmethyl isocyanide was also tolerated as the

isocyanide input affording the corresponding Ugi-Smiles product 22 in fair yield (Figure 2).

In addition, the use of an optically pure chiral isocyanide input ((R)-α-methylbenzyl isocya-

nide led to the formation of a single diastereomer with a 2,3-trans configuration, confirming

that the starting chiral imine is responsible for the diastereoselectivity of the reaction (23,

Figure 2).

Figure 2: Scope of Ugi-Smiles 3CR using optically pure 8.

Page 132: Biocatalysis & Multicomponent Reactions: The Ideal …...Biocatalysis & Multicomponent Reactions Chapter 2 Monoamine Oxidase N: 41 A Promising Biocatalyst for Asymmetric Synthesis

Chapter 6

132

X-Ray crystallographic analysis of 13 (Figure 3) unequivocally established the absolute con-

figuration since the absolute stereochemistry at C3 and C4 positions, resulting from the bio-

transformation, has been reported previously.[13]

Figure 3. Single-crystal structure of 13. Displacement ellipsoids are drawn at the 50% probability level.

Next, the sterically less demanding enantiomerically enriched (1S,4R)-3-azabicyclo[3.3.0]

oct-2-ene (34)[13] was used in a series of Ugi-Smiles 3CRs (24-33, Figure 4). Very good diaste-

reomeric ratios were obtained for compounds 24-33 and considerably higher yields were

observed compared to reactions using the bulky bridged imine 10 as an input (cf. Figure 2).

This observation can be rationalized by steric considerations or by the difference in stability

between the used cyclic imines.[25]

Page 133: Biocatalysis & Multicomponent Reactions: The Ideal …...Biocatalysis & Multicomponent Reactions Chapter 2 Monoamine Oxidase N: 41 A Promising Biocatalyst for Asymmetric Synthesis

Stereoselective Synthesis of N-aryl Amides

133

Figure 4. Scope of Ugi-Smiles 3CR using enantiomerically enriched 3-azabicyclo[3.3.0]oct-2-ene 34.

Subsequently, we synthesized a set of Ugi-Smiles products (Figure 5) that can undergo

additional complexity-generating reactions, e.g. olefin metathesis,[26] cycloaddition

reactions (Diels–Alder[27] or azide-alkyne cycloaddition[28]) or Pd-catalyzed cross-coupling

reactions[29-30] to provide rapid access to highly complex scaffolds with pharmaceutically

interesting properties.

We chose to introduce an azide and allyl functionality into our Ugi-Smiles products by us-

ing 2-azidoethyl isocyanide[31] and commercially available allyl isocyanide as an input. The

presence of the azide and allyl functionalities in the final adducts allows interesting applica-

tions to heterocycle synthesis. The desired azide compound 35 was obtained in good yield

and diastereoselectivity (72%, d.r. = 88:12). Next, we secured Ugi-Smiles adducts 36 (55%,

d.r. = 89:11) and 37 (34%, d.r. > 99:1) with the desired allyl functionality. In the latter case,

we opted to use the saturated bridged imine 38[13] as the double bond in 10 could possibly

interfere in subsequent follow-up reactions.

Page 134: Biocatalysis & Multicomponent Reactions: The Ideal …...Biocatalysis & Multicomponent Reactions Chapter 2 Monoamine Oxidase N: 41 A Promising Biocatalyst for Asymmetric Synthesis

Chapter 6

134

Figure 5. Ugi-Smiles products with reactivity handles.

Remarkably, Ugi-Smiles products derived from 2-nitrophenol or 3-nitro-2-pyridinol as inputs

showed unusually high αD values (see the Experimental Section). For example, compound

13 has a specific rotation of -1043.1˚. This is the highest specific rotation ever reported for

a multicomponent product. This phenomenon was only observed for compounds with the

nitro substituent in ortho position to the newly formed N-aryl bond. These compounds are

likely highly rigid due to restricted rotation of the N-aryl bond. Further research regarding

factors influencing these extraordinarily high specific rotations is ongoing.

6.3 Conclusions & Outlook

In conclusion, we have developed an efficient combination of MAO-N-catalyzed desym-

metrization of cyclic meso-amines with Ugi-Smiles multicomponent chemistry to generate

optically pure N-aryl proline amides. This method represents the first report of a fully asym-

metric Ugi-Smiles process. The simple procedure, broad substrate scope and the presence of

diverse (heterocyclic) ring systems make these compounds highly attractive. Especially the

possibility to add more diversity and complexity to these N-aryl proline amides by introduc-

ing strategic functional groups makes these products appealing for the design of synthesis

of combinatorial libraries (e.g. better coverage of chemical space) on highly functionalized

heterocyclic small molecules for the pharmaceutical industry.

Page 135: Biocatalysis & Multicomponent Reactions: The Ideal …...Biocatalysis & Multicomponent Reactions Chapter 2 Monoamine Oxidase N: 41 A Promising Biocatalyst for Asymmetric Synthesis

Stereoselective Synthesis of N-aryl Amides

135

6.4 Experimental section

General Information

Starting materials and solvents were purchased from ABCR and Sigma-Aldrich and were

used without treatment. 3-Azabicylo[3,3,0]octane hydrochloride was purchased from AK

Scientific. (1R,2S,6R,7S)-4-methyl-4-azatricyclo[5.2.1.02,6]dec-8-ene was prepared according

to literature procedure.[32] Column chromatography was performed on silica gel.1H and 13C NMR spectra were recorded on a Bruker Avance 400 (400.13 MHz for 1H and 100.61

MHz for 13C) or Bruker Avance 500 (500.23 MHz for 1H and 125.78 MHz for 13C) in CDCl3.

Chemical shifts are reported in d values (ppm) downfield from tetramethylsilane.

Electrospray Ionisation (ESI) mass spectrometry was carried out using a Bruker micrOTOF-Q

instrument in positive ion mode (capillary potential of 4500 V).

Thin Layer Chromatography was performed using TLC plates from Merck (SiO2, Kieselgel 60)

and was visualized by UV detection. Flash chromatography was performed using the indi-

cated solvents and Silicycle Silia-P Flash Silica Gel (particle size 40-63 μm, pore diameter 60 Å).

Infrared (IR) spectra were recorded neat, and wavelengths are reported in cm-1. Optical rota-

tions were measured with a sodium lamp and are reported as follows: [α]D20 (c = g/100 mL,

solvent). Optical rotations were measured on an Optical Activity AA-10 and a Perkin Elmer

241 polarimeter with a sodium lamp. 1-Azido-2-isocyanoethane was synthesized according

to literature procedure,[31] (isocyanomethyl)benzene and 3-(isocyanomethyl)pyridine were

synthesized according to literature procedure[33] and 2-Bromo-4-nitrophenol was synthe-

sized according to literature procedure.[34]

General Procedure 1: Preparation of optically active imines (3S,7R)-10,

azabicyclo-[3,3,0]oct-2-ene 34 and (3S,7R)-38

Unless stated otherwise: imines were synthesized according to literature procedure[13] with

minor adjustments. 0.7 g of freeze-dried MAO-N D5 E. coli were rehydrated for 30 min. in 20

ml of KPO4 buffer (100 mM, pH = 8,0) at 37 °C. Subsequently 1 mmol of bridged amine or

azabicyclo-[3,3,0]oct-2-ene) in 30 ml of KPO4 buffer (100 mM, pH = 8.0) was prepared. The pH

of the solution was adjusted to 8.0 by addition of NaOH and then added to the rehydrated

cells. After 16-17 h the reaction was stopped (conversions were > 95 %) and worked up. For

workup the reaction mixture was centrifuged at 4000 rpm and 4 °C until the supernatant

had clarified (40 – 60 minutes). The pH of the supernatant was then adjusted to 10-11 by

addition of aq. NaOH and the supernatant was subsequently extracted with t-butyl methyl

ether or dichloromethane (4 × 70 mL). The combined organic phases were dried with Na2SO4

and concentrated at the rotary evaporator.

Page 136: Biocatalysis & Multicomponent Reactions: The Ideal …...Biocatalysis & Multicomponent Reactions Chapter 2 Monoamine Oxidase N: 41 A Promising Biocatalyst for Asymmetric Synthesis

Chapter 6

136

General procedure 2: Preparation of optically active Ugi-Smiles derivatives 13-37

Unless stated otherwise: To a solution of optically imine (0.5 mmol, 2.0 equiv.) in 2 ml

methanol nitrophenol or thiol (0.25 mmol, 1.0 equiv.) was added, followed by the isocyanide

(0.375 mmol, 1.5 equiv.). The reaction mixture was stirred at 40°C for 24 h. The mixture was

concentrated in vacuo and the crude product was purified by column chromatography (SiO2,

CHCl2). After concentration in vacuo a bright colored oil or solid was obtained. Diastereomeric

ratios were determined by 1H NMR.

Compound 13: General procedure 2 was followed using imine (3S,7R)-10 (60 mg, 0.45 mmol, 2.0 equiv.), 2-nitrophenol (31 mg, 0.22 mmol, 1.0 equiv.) and tert-butyl isocyanide (28 mg, 0.038 ml, 0.34 mmol, 1.5 equiv.) giving a yield of 57 mg (0.16 mmol, 73%) of a bright orange solid. = -1043.1° (c = 0.51, CHCl3); 1H NMR (500.23 MHz, CDCl3): δ = 7.65 (dd, J = 8.5 Hz, J = 1.5 Hz, 1H), 7.29 (m, 1H), 6.91 (d, J = 8.0 Hz), 6.78 (m 1H), 6.67 (bs, 1H), 6.14 (m, 1H), 5.79 (m, 1H) 4.00 (d, J = 3.0 Hz, 1H), 3.52 (dd, J = 10.8 Hz, J = 9.0, 1H),

3.17 (m, 1H), 3.02 (m, 1H), 2.89 (m, 1H), 2.79 (m, 1H) 2.20 (dd, J = 3.0 Hz, J = 11.0 Hz, 1H), 1.49 (d, J = 8.5 Hz, 1H), 1.37 (d, J = 8.5 Hz, 1H), 1.19 (s, 9H); 13C NMR (125.78 MHz, CDCl3): δ = 171.8, 141.3, 139.3, 135.5, 135.3, 133.3, 126.4, 118.4, 117.4, 65.9, 54.7, 51.9, 51.1, 51.0, 50.8, 48.2, 47.7, 46.6, 54.4, 45.2, 45.0, 28.5, 28.5, 28.5; IR (neat): νmax (cm-1) = 3383, 2959, 2868, 1668, 1504, 852, 737; HR-MS (ESI, 4500 V): m/z calcd for C20H26N3O3

([M+H]+): 356.1969, found 356.1959;

Compound 14: General procedure 2 was followed using imine (3S,7R)-10 (67 mg, 0.5 mmol, 2.0 equiv.), 2-nitrophenol (35 mg, 0.25 mmol, 1.0 equiv.) and isopropyl isocyanide (26 mg, 0.035 ml, 0.375 mmol, 1.5 equiv.) giving a yield of 51 mg (0.15 mmol, 60%) of a bright sticky orange oil. = -828.8° (c = 0.56, CHCl3); 1H NMR (500.23 MHz, CDCl3): δ = 7.65 (dd, J = 3.5 Hz, J = 9.75 Hz, 1H), 7.29 (m, 1H), 6.91 (d, J = 9.0 Hz, 1H), 6.78 (m, 2H), 6.14 (m, 1H), 5.79 (m, 1H), 4.10 (m, 1H), 3.86 (m, 1H), 3.48 (m, 1H), 3.19 (m, 1H), 3.03 (m, 1H), 2.86

(m, 1H), 2.79 (m, 1H), 2.20 (dd, J = 4.0 Hz, J = 11.0 Hz, 1H), 1.48 (m, 1H), 1.36 (m, 1H), 1.07 (dd, J = 2.0 Hz, J = 6.5 Hz, 3H), 0.89 (dd, J = 2.0 Hz, J = 6.5 Hz, 3H); 13C NMR (125.78 MHz, CDCl3): δ = 171.6, 141.2, 139.4, 135.6, 135.2, 133.3, 126.4, 118.5, 117.7, 65.4, 54.6, 51.8, 50.6, 47.7, 46.6, 45.1, 41.3, 22.4, 22.4; IR (neat): νmax (cm-1) = 3369, 2968, 2868, 1657, 1508, 741, 633, 493; HR-MS (ESI, 4500 V): m/z calcd for C19H24N3O3

([M+H]+): 342.1812, found 342.1799.

Compound 15: General procedure 2 was followed using imine (3S,7R)-10 (67 mg, 0.5 mmol, 2.0 equiv.), 2-nitrophenol (35 mmol, 0.25 mmol, 1.0 equiv.) and benzyl isocyanide (44 mg, 0.046 ml, 0.375 mmol, 1.5 equiv.) giving a yield of 49 mg (01.26 mmol, 51%) of a sticky orange oil. = -942.3° (c = 0.52, CHCl3); 1H NMR (500.23 MHz, CDCl3): δ = 7.61 (dd, J = 8.0 Hz, J = 1.5 Hz, 1H), 7.30 (m, 1H), 7.16 (m, 4H), 7.14 (m, 2H), 6.94 (m, 1H), 6.17 (m, 1H), 5.84 (m, 1H), 4.35 (d, J = 6.0 Hz, 1H), 4.28 (d, 5.5 Hz, 1H), 4.17 (d, J = 3.0 Hz, 1H),

3.52 (dd, J = 10.8 J = 8.5, 1H), 3.22 (m, 1H), 3.07 (m, 1H), 2.90 (m, 1H), 2.80 (m, 1H), 2.22 (dd, J = 11.0 Hz J = 3.0, 1H), 1.51 (d, J = 8.5 Hz, 1H), 1.38 (d, J = 8.5 Hz, 1H); 13C NMR (125.78 MHz, CDCl3): δ = 172.7, 141.1, 141.1, 137.9, 135.8, 135.3, 133.3, 128.6, 127.2, 126.4, 118.8, 117.7, 65.6, 54.6, 51.9, 50.9, 47.6, 46.6, 45.2, 43.2; IR (neat): νmax (cm-1) = 3377, 2962, 2866, 1664, 1506, 1259, 1014, 796; HR-MS (ESI, 4500 V): m/z calcd for C23H24N3O3

([M+H]+): 390.1812, found 390.1794.

Page 137: Biocatalysis & Multicomponent Reactions: The Ideal …...Biocatalysis & Multicomponent Reactions Chapter 2 Monoamine Oxidase N: 41 A Promising Biocatalyst for Asymmetric Synthesis

Stereoselective Synthesis of N-aryl Amides

137

Compound 16: General procedure 2 was followed using imine (3S,7R)-10 (67 mg, 0.5 mmol, 2.0 equiv.), 4-nitrophenol (35 mg, 0.25 mmol, 1.0 equiv.) and tert-butyl isocyanide (31 mg, 0.042 ml, 0.375 mmol 1.5 equiv.) giving a yield of 45 mg (0.13 mmol, 51%) of a yellow solid. = -42.7° (c = 0.52, CHCl3); 1H NMR (500.23 MHz, CDCl3): δ = 8.03 (d, J = 9.0 Hz, 2H), 6.34 (d, J = 9.0 Hz, 2H), 6.08 (t, J = 4.0 Hz, 2H), 5.36 (bs, 1H), 3.56 (t, J = 2.0 Hz, 1H), 3.53 (s, 1H), 3.15 (s, 1H), 3.03 (m, 3H), 2.95 (s, 1H), 1.58 (d, J = 8.5 Hz, 1H), 1.44 (d, J = 8.5 Hz, 1H) 1.19 (s, 9H); 13C NMR (100.61 MHz, CDCl3): δ = 171.6, 150.5, 138.3, 136.0,

135.9, 126.1, 126.1, 111.6, 111.6, 67.4, 52.7, 52.4, 52.0, 51.3, 47.4, 46.7, 44.1, 28.6, 28.6, 28.6; IR (neat): νmax (cm-1) = 3298, 2966, 2858, 1595, 1302, 1109, 823, 656; HR-MS (ESI, 4500 V): m/z calcd for C20H26N3O3

([M+H]+): 356.1969, found 356.1956.

Compound 17: General procedure 2 was followed using imine (3S,7R)-10 (67 mg, 0.5 mmol, 2.0 equiv.), 4-nitrophenol (35 mg, 0.25 mmol, 1.0 equiv.) and benzyl isocyanide (44 mg, 0.046 ml, 0.375 mmol and 1.5 equiv.) giving a yield of 56 mg (0.14 mmol, 58%) of a yellow solid. = -49.0° (c = 0.49, CHCl3); 1H NMR (500.23 MHz, CDCl3): δ = 8.00 (d, J = 9.5 Hz, 2H), 7.18 (m, 3H), 7.02 (m, 2H), 6.34 (d, J = 9.0 Hz, 2H), 6.10 (m, 2H), 5.92 (s, 1H), 4.33 (d, J = 6.0 Hz, 2H), 3.74 (d, J = 3.0 Hz, 1H), 3.58 (m, 1H), 3.19 (s, 1H), 3.06 (m, 3H), 2.96 (s, 1H), 1.59 (d, J = 8.5 Hz, 1H), 1.45 (d, J = 8.5 Hz, 1H); 13C NMR (125.78 MHz, CDCl3): δ =172.4, 150.4, 138.4, 137.7, 136.0, 135.9, 128.7, 128.7, 127.7, 127.5, 127.5, 126.1,

126.1, 111.7, 111.7 66.8, 52.8, 52.5, 52.0, 47.4, 45.8, 44.1, 43.3; IR (neat): νmax (cm-1) = 3312, 2962, 2869, 1651, 1597, 1300, 1109, 825, 690; HR-MS (ESI, 4500 V): m/z calcd for C23H24N3O3 ([M+H]+): 390.1812, found 390.1794.

Compound 18: General procedure 2 was followed using imine (3S,7R)-10 (67 mg, 0.5 mmol, 2.0 equiv.), 3-nitropyridin-2-ol (35 mg, 0.25 mmol, 1.0 equiv.) and tert-butyl isocyanide (31 mg, 0.042 ml, 0.375 mmol, 1.5 equiv.) giving a yield of 36 mg (0.10 mmol, 40%) of a dark yellow solid. = -474.6° (c = 0.51, CHCl3); 1H NMR (500.23 MHz, CDCl3): δ = 8.23 (dd, J = 2.0 Hz, J = 4.5 Hz, 1H), 8.03 (dd, J = 1.5 Hz, J = 8.0 1H), 6.69 (dd, J = 4.5 Hz, J = 8.5 1H), 6.33 (bs, 1H), 5.94 (m, 1H), 5.56 (m, 1H), 5.03 (s, 1H), 3.49 (dd, J = 8.5, J = 12 Hz, 1H), 3.13

(m, 1H), 2.76 (m, 1H), 2.16 (dd, J = 2.0, J = 12.0 Hz, 1H), 1.39 (d, J = 8.0 Hz, 1H), 1.31 (d, J = 8.5 Hz, 1H), 1.22 (s, 9H); 13C NMR (125.78 MHz, CDCl3): δ = 171.5, 151.7, 150.1, 150.1, 135.8, 134.7, 133.6, 113.1, 63.6, 53.8, 51.3, 51.0, 48.5, 47.8, 46.8, 44.8, 28.6; IR (neat): νmax (cm-1) = 3398, 2959, 2872, 1668, 1448, 716, 557; HR-MS (ESI, 4500 V): m/z calcd for C19H24N4O3

([M+H]+): 357.1921, found 357.1905.

Compound 19: General procedure 2 was followed using imine (3S,7R)-10 (67 mg, 0.5 mmol, 2.0 equiv.), 2-mercaptopyrimidine (28 mg, 0.25 mmol, 1.0 equiv.) and tert-butyl isocyanide (31 mg, 0.042 ml, 0.375 mmol, 1.5 equiv.) giving a yield of 50 mg (0.15 mmol, 61%) of a light yellow solid. = -80.8° (c = 0.495, CHCl3); 1H NMR (500.23 MHz, CDCl3): δ = 8.23 (d, J = 4.5 Hz, 2H), 7.99 (bs, 1H), 6.48 (t, J = 5.0 Hz, 1H), 5.98 (t, J = 7.0 Hz, 2H), 4.39 (d, J = 2.5 Hz, 1H), 3.60 (dd, J = 3.5 Hz, J = 11.5 Hz, 1H), 3.50 (dd, J = 8.5 Hz, J = 11.8 Hz, 1H), 3.37 (m,

1H), 3.10 (m, 1H), 2.92 (m, 1H), 2.88 (m, 1H), 1.46 (d, J = 8.5 Hz, 1H), 1.38 (m, 10H); 13C NMR (125.78 MHz, CDCl3): δ = 202.6, 159.8, 157.7, 157.7, 135.6, 135.4, 110.6, 72.9, 55.0, 53.4, 51.7, 50.1, 47.5), 46.9, 44.2, 27.6, 27.6, 27.6; IR (neat): νmax (cm-1) = 3303, 2964, 2856, 1680, 1377, 797, 719, 627; HR-MS (ESI, 4500 V): m/z calcd for C18H25N4S ([M+H]+): 329.1794, found 329.1783.

Page 138: Biocatalysis & Multicomponent Reactions: The Ideal …...Biocatalysis & Multicomponent Reactions Chapter 2 Monoamine Oxidase N: 41 A Promising Biocatalyst for Asymmetric Synthesis

Chapter 6

138

Compound 20: General procedure 2 was followed using imine (3S,7R)-10 (67 mg, 0.5 mmol, 2.0 equiv.), 2-mercaptopyrimidine (28 mg, 0.25 mmol, 1.0 equiv.) and benzyl isocyanide (44 mg, 0.046 ml, 0.375 mmol, 1.5 equiv.) giving a yield of 57 mg (0.16 mmol, 63%) of a off-white solid. = -60.0° (c = 0.6, CHCl3); 1H NMR (500.23 MHz, CDCl3): δ = 8.38 (bs, 1H), 8.17 (d, J = 5.0 Hz, 2H), 7.21 (m, 3H), 7.10 (t, J = 5.5 Hz, 2H) 6.45 (t, J = 5.0 Hz, 1H), 5.98 (s, 1H), 4.82 (m, 1H), 4.67 (m, 1H), 4.61 (s, 1H), 3.64 (m, 1H), 3.46 (m, 2H), 3.13 (s, 1H),

2.96 (m, 1H), 2.89 (s, 1H), 1.49 (d, J = 8.5 Hz, 1H), 1.41 (d, J = 8.0 Hz, 1H); 13C NMR (125.78 MHz, CDCl3): δ = 203.7, 158.8, 156.7, 156.7, 135.2, 134.5, 134.4, 134.4, 127.7, 127.7, 127.7, 126.7, 126.7, 109.7, 70.2, 52.5, 50.6, 49.0, 48.4, 46.4, 45.9, 43.2; IR (neat): νmax (cm-1) = 3175, 2966, 2870, 1585, 1493, 795, 696; HR-MS (ESI, 4500 V): m/z calcd for C21H23N4S ([M+H]+): 363.1638, found 363.1632.

Compound 21: General procedure 2 was followed using imine (3S,7R)-10 (67 mg, 0.5 mmol, 2.0 equiv.), 2-mercaptopyrimidine (28 mg, 0.25 mmol, 1.0 equiv.) and isopropyl isocyanide (26 mg, 0.035 ml, 0.375 mmol, 1.5 equiv.) giving a yield of 54 mg (0.17 mmol, 68%) of light orange solid. = -173.7° (c = 0.50, CHCl3); 1H NMR (500.23 MHz, CDCl3): δ = 8.51 (dd, J = 5.0 Hz, J = 10.0 Hz, 1 H), 8.22 (d, J = 5.0 Hz, 2H), 7.92 (bs, 1H), 6.47 (t, J = 4.5 Hz, 1H), 5.98 (m, 2H), 4.53 (m, 1H), 4.49 (d, J = 2.5 Hz, 1H), 3.64 (dd, J = 3.0 Hz, J = 12 Hz, 1H),

3.48 (dd, J = 3.0 Hz, J = 12 Hz, 1H), 3.40 (m, 1H), 3.11 (m, 1H), 2.94 (m, 1H), 2.89 (m, 1H), 1.47 (d, J = 8.5 Hz, 1H), 1.39 (d, J = 8.5 H, 1H), 1.07 (m, 6H); 13C NMR (125.78 MHz, CDCl3): δ = 202.5, 159.9, 157.7, 157.7, 135.4, 135.4, 110.6, 71.4, 53.3, 51.7, 50.0, 47.4, 46.9, 46.7, 44.2, 21.3, 21.1; IR (neat): νmax (cm-1) = 3329, 2966, 2868, 1666, 1580, 1445, 804, 721; HR-MS (ESI, 4500 V): m/z calcd for C17H23N4S ([M+H]+): 315.1638, found 315.1634.

Compound 22: General procedure A was followed using imine (3S,7R)-10 (67 mg, 0.5 mmol, 2.0 equiv.), 2-mercaptopyrimidine (28 mg, 0.25 mmol, 1.0 equiv.) and SBL086 (44 mg, 0.375 mmol, 1.5 equiv.) giving a yield of 34 mg (0.09 mmol, 37%) of an orange solid. = -106.1° (c = 0.51, CHCl3); 1H NMR (400.16 MHz, CDCl3): δ = 8.59 (bs, 1H), 8.41 (m, 2H), 8.16 (d, J = 4.4 Hz, 2H), 7.47 (d, J = 7.6 Hz, 1H), 7.11 (dd, J = 4.8 Hz, J = 7.8 Hz, 1H), 6.46 (t, J = 4.8 Hz, 1H), 5.98 (m, 2H), 4.89 (dd, J = 6.0 Hz, J = 15.2 Hz, 1H), 4.70

(dd, J = 6.0 Hz, J = 15.2 Hz, 1H), 4.58 (d, J = 2.4 Hz, 1H), 3.63 (m, 1H), 3.42 (m, 2H), 3.11 (m, 1H), 2.94 (m, 2H), 1.48 (d, J = 8.4 Hz, 1H), 1.40 (d, J = 8.4 Hz, 1H); 13C NMR (100.62 MHz, CDCl3): δ = 204.7, 158.9, 156.7, 148.1, 134.5, 134.3, 131.1, 122.5, 109.8, 70.3, 52.6, 50.7, 49.1, 46.4, 46.0, 43.2, 41.2; IR (neat): νmax (cm-1) = 3173, 2976, 2204, 1499, 922, 725, 642; HR-MS (ESI, 4500 V): m/z calcd for C20H22N5S ([M+H]+): 364.1590, found 364.1575.

Compound 23: General procedure A was followed using (3S,7R)-10 (67 mg, 0.5 mmol, 2.0 equiv.), 2-nitrophenol (35 mg, 0.25 mmol, 1.0 equiv.) and (S)-(-)-methylbenzyl isocyanide (49 mg, 0.051 ml, 0.375 mmol, 1.5 equiv giving a yield of 53 mg (0.13 mmol, 53%) of an orange solid. = -756.4° (c = 0.55, CHCl3); 1H NMR (400.13 MHz, CDCl3): δ = 7.63 (d, J = 8.0 Hz, 1H), 7.26 (m, 2H), 7.02 (m, 3H), 6.88 (m, 3H), 6.79 (t, J = 7.5 Hz, 1H), 6.15 (m, 1H), 5.84 (m, 1H), 4.92 (m, 1H), 4.09 (s, 1H), 3.56 (dd, J = 8.5 Hz, J = 10.5 Hz, 1H),

3.20 (s, 1H), 3.09 (m, 1H), 2.95 (m, 1H), 2.81 (s, 1H), 2.26 (dd, J = 3.0, J = 10.8, 1H), 1.51 (m, 1H), 1.39 (m, 1H), 1.37 (d, J = 7.0 Hz); 13C NMR (100.61 MHz, CDCl3): δ = 171.7, 143.4, 141.1, 139.7, 135.8, 135.3, 133.3, 128.3 , 128.3, 126.9, 126.3, 125.7, 125.7, 118.9, 117.9, 65.5, 54.5, 52.0, 50.7, 48.8, 47.6, 46.6, 45.2, 22.5; IR (neat): νmax (cm-1) = 3379, 3312, 2966, 1655, 1502, 739, 698; HR-MS (ESI, 4500 V): m/z calcd for C20H26N3O3

([M+H]+): 404.1969, found 404.1949.

Page 139: Biocatalysis & Multicomponent Reactions: The Ideal …...Biocatalysis & Multicomponent Reactions Chapter 2 Monoamine Oxidase N: 41 A Promising Biocatalyst for Asymmetric Synthesis

Stereoselective Synthesis of N-aryl Amides

139

Compound 24: General procedure 2 was followed using 3-azabicyclo[3.3.0]oct-2-ene 34 (55 mg, 0.5 mmol, 2.0 equiv.), 2-nitrophenol (35 mg, 0.25 mmol, 1.0 equiv.) and tert-butyl isocyanide (31 mg, 0.042 ml, 0.375 mmol, 1.5 equiv.) giving a yield of 61 mg (0.18 mmol, 72%) of a bright sticky orange oil. (Note: Minor diastereomer is given in italic). = -734° (c = 0.52, CHCl3);[35] 1H NMR (500.23 MHz, CDCl3): δ = 7.67 (m, 2H), 7.33 (m, 1H), 7.00 (d, J = 8.0 Hz, 1H), 6.86 (m, 1H), 6.56 (bs, 1H), 3.88 (d, J = 6.0 Hz, 1H), 3.71 (m, 1H), 2.73 (m, 1H), 2.62 (m,

1H), 2.49 (dd, J = 3.0 Hz, J = 10.3 Hz, 1H), 1.88 (m, 2H), 1.70 (m, 2H), 1.66 (m, 1H), 1.21 (m, 1H), 1.18 (s, 9H); 13C NMR (125.78 MHz, CDCl3): δ = 171.7, 142.1, 140.4, 135.5, 126.1, 119.5, 117.8, 69.6), 58.4, 50.8), 49.8), 43.2, 32.2, 31.7, 28.5, 28.5, 28.5, 24.9; 1H NMR (500.23 MHz, CDCl3): δ = 7.67 (m, 2H), 7.33 (m, 1H), 7.00 (d, J = 8.0 Hz, 1H), 6.86 (m, 1H), 6.48 (bs, 1H), 4.36 (d, J = 7.6 Hz, 1H), 4.00 (m, 1H), 2.73 (m, 1H), 2.62 (m, 1H), 2.49 (dd, J = 3.0 Hz, J = 10.3 Hz, 1H), 1.88 (m, 2H), 1.70 (m, 2H), 1.66 (m, 1H), 1.21 (m, 1H), 1.18 (s, 9H); 13C NMR (125.78 MHz, CDCl3): δ = 171.7, 142.1, 140.4, 135.2, 125.8, 119.8, 117.7, 70.1, 59.2), 50.8, 46.9, 42.8, 34.6, 29.7, 28.5, 28.5, 28.5, 25.9; IR (neat): νmax (cm-1) = 3689, 3352, 2957, 1655, 1506, 1265, 746, 710; HR-MS (ESI, 4500 V): m/z calcd for C18H26N3O3

([M+H]+): 332.1969, found 332.1958;

Compound 25: General procedure 2 was followed using 3-azabicyclo[3.3.0]oct-2-ene 34 (55 mg, 0.5 mmol, 2.0 equiv.), 2-nitrophenol (35 mg, 0.25 mmol, 1.0 equiv.) and isopropyl isocyanide (26 mg, 0.035 ml, 0.375 mmol, 1.5 equiv.) giving a yield of 56 mg (0.18 mmol, 71%) of a sticky orange oil. (Note: Minor diastereomer is given in italic). = -722.9° (c = 0.55, CHCl3);[35] 1H NMR (400.13 MHz, CDCl3): δ = 7.62 (dd, J = 1.5 Hz, J = 8.0 Hz, 1H), 7.33 (m, 1H), 7.10 (m, 4H), 7.02 (d, J = 8.5 Hz, 1H), 6.96 (t, J = 3.8 Hz, 2H), 6.87 (t, J = 7.5 Hz, 1H), 4.30

(d, J = 6.0 Hz, 2H), 4.05 (d, J = 6.5 Hz, 1H), 3.72 (dd, J = 7.5 Hz, J = 10.3 Hz, 1H), 2.77 (m, 1H), 2.63 (m, 1H), 2.51 (dd, J = 3.5 Hz, J = 10.0 Hz, 1H), 1.91 (m, 2H), 1.72 (m, 2H), 1.61 (m, 1H), 1.21 (m, 1H); 13C NMR (100.62 MHz, CDCl3): δ = 170.5, 141.0, 139.7, 132.5, 125.0, 119.1, 117.1, 68.2, 57.5, 48.7, 42.2, 40.0, 31.2, 30.8, 24.0, 21.4, 21.3; 1H NMR (400.13 MHz, CDCl3): δ = 7.66 (dd, J = 1.5 Hz, J = 8.0 Hz, 1H), 7.33 (m, 1H), 7.00 (d, J = 8.5 Hz, 1H), 6.86 (t, J = 7.5 Hz, 1H), 6.50 (bs, 1H), 4.45 (d, J = 7.6 Hz, 2H), 3.88 (m, 1H), 3.70 (dd, J = 7.0 Hz, J = 10.0 Hz, 1H), 3.49 (m, 1H), 3.72 (m, 1H), 2.99 (m, 1H), 2.81 (m, 2H), 2.50 (dd, J = 3.5 Hz, J = 10.0 Hz, 1H), 1.88 (m, 1H), 1.62 (m, 1H), 1.21 (m, 1H), 1.12 (m, 3H), 0.79 (m, 3H); 13C NMR (100.62 MHz, CDCl3): δ = 170.5, 141.0, 139.7, 132.2, 124.7, 118.7, 117.0, 65.6, 58.3, 45.8, 42.0, 40.0, 33.6, 29.0, 25.0, 21.7, 21.3; IR (neat): νmax (cm-1) = 3371, 2962, 2860, 1651, 1506, 1277, 773, 741; HR-MS (ESI, 4500 V): m/z calcd for C17H24N3O3

([M+H]+): 318.1812, found 318.1799.

Compound 26: General procedure 2 was followed using 3-azabicyclo[3.3.0]oct-2-ene 34 (55 mg, 0.5 mmol, 2.0 equiv.), 2-nitrophenol (35 mg, 0.25 mmol, 1.0 equiv.) and benzylisocyanide (44 mg, 0.046 ml, 0.375 mmol, 1.5 equiv.) giving a yield of 69 mg (0.19 mmol, 76%) of a sticky orange oil. (Note: Minor diastereomer is given in italic). = -650.9° (c = 0.55, CHCl3);[35] 1H NMR (500.23 MHz, CDCl3): δ = 7.62 (dd, J = 1.5 Hz, J = 8.0 Hz, 1H), 7.33 (m, 1H), 7.10 (m, 4H), 7.02 (d, J = 8.5 Hz, 1H), 6.96 (t, J = 3.8 Hz, 2H), 6.87 (t, J = 7.5 Hz, 1H), 4.30

(d, J = 6.0 Hz, 2H), 4.05 (d, J = 6.5 Hz, 1H), 3.72 (dd, J = 7.5 Hz, J = 10.3 Hz, 1H), 2.77 (m, 1H), 2.63 (m, 1H), 2.51 (dd, J = 3.5 Hz, J = 10.0 Hz, 1H), 1.91 (m, 2H), 1.72 (m, 2H), 1.61 (m, 1H), 1.21 (m, 1H); 13C NMR (125.78 MHz, CDCl3): δ = 171.5, 140.9, 139.7, 137.0, 132.5, 127.5, 127.5, 126.3, 126.3, 126.2, 125.1, 118.9, 117.1), 68.2, 57.4, 48.9, 42.3, 41.9, 30.9, 30.6, 23.8; 1H NMR (500.23 MHz, CDCl3): δ = 7.62 (dd, J = 1.5 Hz, J = 8.0 Hz, 1H), 7.33 (m, 1H), 7.10 (m, 4H), 7.02 (d, J = 8.5 Hz, 1H), 6.96 (t, J = 3.8 Hz, 2H), 6.87 (t, J = 7.5 Hz, 1H), 4.55 (d, J = 7.6 Hz, 2H), 4.05 (d, J = 6.5 Hz, 1H), 3.72 (dd, J = 7.5 Hz, J = 10.3 Hz, 1H), 3.17 (m, 1H), 3.01 (m, 1H), 2.51 (dd, J = 3.5 Hz, J = 10.0 Hz, 1H), 1.91 (m, 2H), 1.72 (m, 2H), 1.61 (m, 1H), 1.21 (m, 1H); 13C NMR (125.78 MHz, CDCl3): δ = 169.7, 141.1, 139.7, 136.8, 132.3, 127.4, 127.4, 126.2, 126.2, 126.1, 124.8, 119.2, 116.9, 65.7, 58.3, 45.9, 42.3, 33.6, 28.9, 26.7, 24.9; IR (neat): νmax (cm-1) = 3373, 2945, 2868, 1655, 1508, 1277, 740, 598; HR-MS (ESI, 4500 V): m/z calcd for C21H24N3O3

([M+H]+): 366.1812, found 366.1801.

Page 140: Biocatalysis & Multicomponent Reactions: The Ideal …...Biocatalysis & Multicomponent Reactions Chapter 2 Monoamine Oxidase N: 41 A Promising Biocatalyst for Asymmetric Synthesis

Chapter 6

140

Compound 27: General procedure 2 was followed using 3-azabicyclo[3.3.0]oct-2-ene 34 (55 mg, 0.5 mmol, 2.0 equiv.), 4-nitrophenol (35 mg, 0.25 mmol, 1.0 equiv.) and tert-butyl isocyanide (31 mg, 0.042 ml, 0.375 mmol, 1.5 equiv.) giving a yield of 80 mg (0.23 mmol, 90%) of a yellow solid. (Note: Minor diastereomer is given in italic). = -67.4° (c = 0.48, CHCl3);[35] 1H NMR (400.13 MHz, CDCl3): δ = 8.04 (m, 2H), 6.46 (m, 2H), 5.61 (bs, 1H), 3.81 (m, 2H), 3.12 (dd, J = 6.8 Hz, J = 10.4 Hz, 1H), 2.83 (m, 1H), 2.73 (m, 1H), 2.01 (m, 1H), 1.84 (m, 1H), 1.73 (m, 1H), 1.63 (m, 1H), 1.53 (m, 2H), 1.21 (s, 9H); 13C NMR (100.62 MHz, CDCl3): δ = 170.3, 150.6,

137.6, 124.8, 124.8, 110.7, 110.7, 69.9, 54.7, 50.4, 49.2, 40.7, 31.6, 30.5, 27.6, 27.6, 23.8; 1H NMR (400.13 MHz, CDCl3): δ = 8.04 (m, 2H), 6.87 (m, 2H), 5.61 (bs, 1H), 4.12 (m, 2H), 3.34 (dd, J = 6.8 Hz, J = 10.4 Hz, 1H), 3.66 (m, 1H), 3.35 (m, 1H), 2.01 (m, 1H), 1.84 (m, 1H), 1.73 (m, 1H), 1.63 (m, 1H), 1.53 (m, 2H), 1.21 (s, 9H); 13C NMR (100.62 MHz, CDCl3): δ = 168.3, 150.6, 137.6, 124.6, 124.6, 111.4, 111.4, 66.7, 55.4, 50.4, 46.4, 41.5, 30.0, 28.7, 26.9, 26.9, 25.2; IR (neat): νmax (cm-1) = 3298, 2957, 2864, 1597, 1296, 1109, 825, 752; HR-MS (ESI, 4500 V): m/z calcd for C18H26N3O3

([M+H]+): 332.1969, found 332.1957.

Compound 28: General procedure 2 was followed using 3-azabicyclo[3.3.0]oct-2-ene 34 (55 mg, 0.5 mmol, 2.0 equiv.), 4-nitrophenol (35 mg, 0.25 mmol, 1.0 equiv.) and benzylisocyanide (44 mg, 0.046 ml, 0.375 mmol, 1.5 equiv.) giving a yield of 69 mg (0.19 mmol, 76%) of a sticky yellow oil. (Note: Minor diastereomer is given in italic). = -90.9° (c = 0.55, CHCl3);[35] 1H NMR (400.13 MHz, CDCl3): δ = 7.97 (d, J = 2.0 Hz, 2H), 7.19 (m, 3H), 7.06 (t, J = 5.6 Hz, 2H), 6.42 (m, 2H), 6.25 (bs, 1H), 4.34 (m, 2H), 3.99 (d, J = 2.4 Hz, 1H), 3.80 (dd, J = 8.4 Hz, J = 10.0 Hz, 1H), 3.11 (dd, J = 6.8 Hz, J = 10.0 Hz, 1H), 2.80 (m, 1H), 2.04 (m, 1H), 1.83 (m,

1H), 1.75 (m, 1H), 1.64 (m, 1H), 1.55 (m, 3H); 13C NMR (100.62 MHz, CDCl3): δ = 172.0, 151.5, 138.6, 137.9, 128.7, 128.7, 127.9, 127.9, 127.7, 126.2, 126.2, 112.1, 112.1, 70.2, 55.6, 50.3, 43.3, 38.6, 32.4, 30.8, 24.8; 1H NMR (400.13 MHz, CDCl3): δ = 7.97 (d, J = 2.0 Hz, 2H), 7.19 (m, 3H), 7.06 (t, J = 5.6 Hz, 2H), 6.79 (m, 2H), 6.25 (bs, 1H), 4.34 (m, 2H), 3.99 (d, J = 2.4 Hz, 1H), 3.64 (dd, J = 8.4 Hz, J = 10.0 Hz, 1H), 3.32 (dd, J = 6.8 Hz, J = 10.0 Hz, 1H), 3.46 (m, 1H), 2.04 (m, 1H), 1.83 (m, 1H), 1.75 (m, 1H), 1.64 (m, 1H), 1.55 (m, 3H); 13C NMR (100.62 MHz, CDCl3): δ = 170.4, 152.0, 138.6, 137.7, 128.7, 127.7, 127.7, 127.5, 126.0, 126.0, 112.4, 112.4, 67.3, 56.3, 47.4), 43.4, 42.7, 30.8, 26.1, 24.9; IR (neat): νmax (cm-1) = 3283, 2949, 2866, 1653, 1597, 1294, 1109, 750; HR-MS (ESI, 4500 V): m/z calcd for C21H24N3O3

([M+H]+): 366.1812, found 366.1804.

Compound 29: General procedure 2 was followed using 3-azabicyclo[3.3.0]oct-2-ene 34 (55 mg, 0.5 mmol, 2.0 equiv.), 3-nitropyridin-2-ol (35 mg, 0.25 mmol, 1.0 equiv.) and tert-butyl isocyanide (31 mg, 0.042 ml, 0.375 mmol, 1.5 equiv.) giving a yield of 61 mg (0.18 mmol, 73%) of a sticky orange oil. (Note: Minor diastereomer is given in italic). = -652.4° (c = 0.52, CHCl3);[35] 1H NMR (400.13 MHz, CDCl3): δ = 8.25 (dd, J = 2.0 Hz, J = 4.6, 1H), 8.04 (dd, J = 1.6 Hz, J = 8.0 Hz, 1H), 6.24 (bs, 1H), 4.86 (d, J = 4.4 Hz, 1H), 3.70 (dd, J = 7.6 Hz, J = 11.2 Hz,

1H), 2.82 (m, 1H), 2.68 (m, 1H), 2.54 (dd, J = 2.8 Hz, J = 11.2 Hz, 1H), 1.94 (m, 1H), 1.72 (m, 2H), 1.57 (m, 2H), 1.23 (s, 9H), 1.14 (m, 1H); 13C NMR (100.62 MHz, CDCl3): δ = 170.5, 150.9, 149.8, 133.7, 131.5, 112.5, 67.0), 56.1, 49.3, 46.7, 42.3, 31.7, 31.4, 31.1, 24.1, 24.1, 24.1; 1H NMR (400.13 MHz, CDCl3): δ = 8.25 (dd, J = 2.0 Hz, J = 4.6, 1H), 8.04 (dd, J = 1.6 Hz, J = 8.0 Hz, 1H), 6.07 (bs, 1H), 5.04 (d, J = 7.6 Hz, 1H), 3.70 (dd, J = 7.6 Hz, J = 11.2 Hz, 1H), 3.24 (m, 1H), 2.95 (m, 1H), 2.54 (dd, J = 2.8 Hz, J = 11.2 Hz, 1H), 1.94 (m, 1H), 1.72 (m, 2H), 1.57 (m, 2H), 1.23 (s, 9H), 1.14 (m, 1H); 13C NMR (100.62 MHz, CDCl3): δ = 169.0, 150.9, 149.8, 133.4, 131.5, 112.7, 64.6, 56.2, 49.3, 45.1, 42.6, 31.4, 28.7, 25.9, 25.9, 25.9, 24.32; IR (neat): νmax (cm-1) = 3396, 2953, 2868, 1664, 1456, 763, 540; HR-MS (ESI, 4500 V): m/z calcd for C17H25N4O3

([M+H]+): 333.1921, found 333.1907.

Page 141: Biocatalysis & Multicomponent Reactions: The Ideal …...Biocatalysis & Multicomponent Reactions Chapter 2 Monoamine Oxidase N: 41 A Promising Biocatalyst for Asymmetric Synthesis

Stereoselective Synthesis of N-aryl Amides

141

Compound 30: General procedure 2 was followed using 3-azabicyclo[3.3.0]oct-2-ene 34 (55 mg, 0.5 mmol, 2.0 equiv.), 2-mercaptopyrimidine (28 mg, 0.25 mmol, 1.0 equiv.) and tert-butyl isocyanide (31 mg, 0.042 ml, 0.375 mmol, 1.5 equiv.) giving a yield of 61 mg (0.20 mmol, 80%) of a off-white solid. (Note: Minor diastereomer is given in italic). = -129.4° (c = 0.51, CHCl3);[35] 1H NMR (400.13 MHz, CDCl3): δ = 8.28 (d, J = 4.8 Hz, 2H), 8.07 (bs, 1H), 6.53 (t, J = 4.8 Hz, 1H), 4.67 (d, J = 2.8 Hz, 1H), 3.97 (dd, J = 8.4 Hz, J = 11.4 Hz, 1H), 3.58 (dd, J = 6.4 Hz, J = 11.4 Hz, 1H) 3.03

(m, 1H), 2.84 (m, 1H) 2.05 (m, 1H), 1.85 (m, 1H), 1.73 (m, 1H), 1.60 (m, 2H), 1.51 (m, 1H), 1.24 (s, 9H); 13C NMR (100.62 MHz, CDCl3): δ = 202.3, 160.9, 157.7, 157.7, 111.1, 46.5, 54.1, 51.7, 41.6, 32.8, 31.5, 27.6, 27.6, 27.6, 25.1; 1H NMR (400.13 MHz, CDCl3): δ = 8.28 (d, J = 4.8 Hz, 2H), 7.50 (bs, 1H), 7.08 (t, J = 4.8 Hz, 1H), 4.97 (d, J = 2.8 Hz, 1H), 4.28 (dd, J = 8.4 Hz, J = 11.4 Hz, 1H), 3.58 (dd, J = 6.4 Hz, J = 11.4 Hz, 1H) 3.30 (m, 1H), 3.16 (m, 1H) 2.05 (m, 1H), 1.85 (m, 1H), 1.73 (m, 1H), 1.60 (m, 2H), 1.51 (m, 1H), 1.24 (s, 9H); 13C NMR (100.62 MHz, CDCl3): δ = 199.5, 160.9, 157.7, 157.7, 111.7, 75.3, 54.9, 49.5, 41.7, 29.7, 27.2, 27.6, 27.6, 27.6, 25.8; IR (neat): νmax (cm-1) = 3312, 2961, 2872, 1580, 1460, 1076, 798, 635; HR-MS (ESI, 4500 V): m/z calcd for C16H25N4S

([M+H]+): 305.1794, found 305.1780.

Compound 31: General procedure 2 was followed using 3-azabicyclo[3.3.0]oct-2-ene 34 (55 mg, 0.5 mmol, 2.0 equiv.), 2-mercaptopyrimidine (28 mg, 0.25 mmol, 1.0 equiv.) and benzyl isocyanide (44 mg, 0.046 ml, 0.375 mmol, 1.5 equiv.) giving a yield of 49 mg (0.15 mmol, 58%) of a white solid. (Note: Minor diastereomer is given in italic). = -66.7° (c = 0.48, CHCl3);[35] 1H NMR (400.13 MHz, CDCl3): δ = 8.46 (bs, 1H), 8.21 (d, J = 4.8 Hz, 2H), 7.18 (m, 3H), 7.11 (m, 2H), 6.50 (t, J = 4.4 Hz, 1H), 4.83 (m, 2H), 4.70 (dd, J = 5.2 Hz, J = 15.2, 1H), 3.79

(dd, J = 8.4 Hz, J = 11.4 Hz, 1H), 3.58 (dd, J = 6.4 Hz, J = 11.4 Hz, 1H) 3.04 (m, 1H), 2.81 (m, 1H) 2.05 (m, 1H), 1.80 (m, 1H), 1.59 (m, 2H), 1.50 (m, 2H); 13C NMR (100.62 MHz, CDCl3): δ = 204.4, 160.9, 157.8, 157.8, 136.4, 128.7, 128.0, 127.8, 111.1, 74.8, 54.1, 51.8, 49.3, 41.7, 32.7, 31.4, 25.1; 1H NMR (400.13 MHz, CDCl3): δ = 7.87 (bs, 1H), 8.27 (d, J = 4.8 Hz, 2H), 7.18 (m, 3H), 7.11 (m, 2H), 6.50 (t, J = 4.4 Hz, 1H), 5.15 (d, 1H), 4.83 (m, 1H), 4.98 (dd, J = 5.2 Hz, J = 15.2, 1H), 4.21 (dd, J = 8.4 Hz, J = 11.4 Hz, 1H), 4.04 (dd, J = 6.4 Hz, J = 11.4 Hz, 1H) 3.16 (m, 1H), 2.67 (m, 1H) 2.05 (m, 1H), 1.80 (m, 1H), 1.59 (m, 2H), 1.50 (m, 2H); 13C NMR (100.62 MHz, CDCl3): δ = 202.0, 161.5, 157.7, 157.7, 136.4, 128.6, 127.8, 127.8, 111.7, 73.9, 54.8, 49.6, 48.8, 42.1, 29.5, 27.4, 25.7; IR (neat): νmax (cm-1) = 3196, 2947, 2862, 1491, 964, 744, 696; HR-MS (ESI, 4500 V): m/z calcd for C19H23N4S

([M+H]+): 339.1638, found 339.1626.

Compound 32: General procedure 2 was followed using 3-azabicyclo[3.3.0]oct-2-ene 34 (55 mg, 0.5 mmol, 2.0 equiv.), 2-mercaptopyrimidine (28 mg, 0.25 mmol, 1.0 equiv.) and isopropyl isocyanide (26 mg, 0.035 ml, 0.375 mmol, 1.5 equiv.) giving a yield of 62 mg (0.21 mmol, 85%) of an off-white sticky oil. (Note: Minor diastereomer is given in italic). = -102.9° (c = 0.51, CHCl3);[35] 1H NMR (400.13 MHz, CDCl3): δ = 8.26 (d, J = 4.8 Hz, 2H), 7.95 (bs, 1H), 6.53 (t, J = 4.8 Hz, 1H), 4.69 (d, J = 2.8 Hz, 1H), 4.57 (m, 1H), 3.82 (dd, J = 8.4 Hz, J = 11.4 Hz, 1H), 3.52 (dd, J = 6.0 Hz, J =

11.4 Hz, 1H) 2.98 (m, 1H), 2.79 (m, 1H) 2.00 (m, 1H), 1.79 (m, 1H), 1.57 (m, 2H), 1.50 (m, 2H), 1.05 (m, 6H); 13C NMR (100.62 MHz, CDCl3): δ = 202.2, 160.9, 157.7, 157.7, 111.1, 75.0, 54.1, 51.6, 46.5, 41.7, 32.7, 31.4), 25.1, 20.9; 1H NMR (400.13 MHz, CDCl3): δ = 8.26 (d, J = 4.8 Hz, 2H), 7.39 (bs, 1H), 7.02 (t, J = 4.8 Hz, 1H), 4.69 (d, J = 2.8 Hz, 1H), 4.24 (m, 1H), 3.82 (dd, J = 8.4 Hz, J = 11.4 Hz, 1H), 3.52 (dd, J = 6.0 Hz, J = 11.4 Hz, 1H) 3.24 (m, 1H), 3.11 (m, 1H) 2.00 (m, 1H), 1.79 (m, 1H), 1.57 (m, 2H), 1.50 (m, 2H), 1.05 (m, 6H); 13C NMR (100.62 MHz, CDCl3): δ = 199.6, 161.5, 157.9, 157.9, 111.7, 73.8, 54.9, 49.5, 46.2, 41.9, 29.6, 27.1, 25.7, 21.5, 21.1; IR (neat): νmax (cm-1) ν = 3190, 2961, 2866, 1583, 1499, 795; HR-MS (ESI, 4500 V): m/z calcd for C15H23N4S ([M+H]+): 291.1638, found 291.1628.

Page 142: Biocatalysis & Multicomponent Reactions: The Ideal …...Biocatalysis & Multicomponent Reactions Chapter 2 Monoamine Oxidase N: 41 A Promising Biocatalyst for Asymmetric Synthesis

Chapter 6

142

Compound 33: General procedure 2 was followed using 3-azabicyclo[3.3.0]oct-2-ene 34 (55 mg, 0.5 mmol, 2.0 equiv.), 2-nitrophenol (35 mg, 0.25 mmol, 1.0 equiv.) and (S)-(-)-a -methylbenzyl isocyanide (49 mg, 0.051 ml, 0.375 mmol, 1.5 equiv.) giving a yield of 80 mg (0.23 mmol, 90%) of a sticky orange oil. (Note: Minor diastereomer is given in italic). = -650.0° (c = 0.72, CHCl3);[35] 1H NMR (400.13 MHz, CDCl3): δ = 7.61 (d, J = 1.6 Hz, 1H), 7.22 (m, 2H), 6.98 (m, 1H), 6.85

(m, 5H), 4.91 (m, 1H), 3.96 (d, J = 6.0 Hz, 1H) 3.73 (dd, J = 7.2 Hz, J = 10.0 Hz, 1H), 2.77 (m, 1H), 2.65 (m, 1H), 2.51 (dd, J = 3.6 Hz, J = 10.0 Hz, 1H), 1.88 (m, 2H), 1.66 (m, 3H), 1.34 (d, J = 7.2 Hz, 3H), 1.21 (m, 2H); 13C NMR (100.62 MHz, CDCl3): δ = 170.6, 142.3, 140.8, 140.1, 132.5, 127.2, 125.8, 124.8, 124.6, 119.0, 117.4, 68.2, 57.4, 48.8, 47.4, 42.3, 31.0, 30.7, 23.9, 21.2; 1H NMR (400.13 MHz, CDCl3): δ = 7.61 (d, J = 1.6 Hz, 1H), 7.36 (m, 2H), 6.98 (m, 1H), 6.85 (m, 5H), 4.91 (m, 1H), 4.49 (d, J = 6.0 Hz, 1H) 3.73 (dd, J = 7.2 Hz, J = 10.0 Hz, 1H), 3.16 (m, 1H), 2.94 (m, 1H), 2.51 (dd, J = 3.6 Hz, J = 10.0 Hz, 1H), 1.88 (m, 2H), 1.66 (m, 3H), 1.34 (d, J = 7.2 Hz, 3H), 1.21 (m, 2H); 13C NMR (100.62 MHz, CDCl3): δ = 168.7, 142.3, 140.8, 140.1, 132.4, 127.7, 126.3, 125.1, 124.7, 119.3, 117.1, 65.6, 58.3, 47.4, 45.9, 42.0, 29.0, 28.7, 26.5, 20.1; IR (neat): νmax (cm-1) = 3367, 2941, 2866, 1657, 1508, 741, 698; HR-MS (ESI, 4500 V): m/z calcd for C22H26N3O3

([M+H]+): 380.1969, found 380.1953.

Compound 35: General procedure 2 was followed using 3-azabicyclo[3.3.0]oct-2-ene 34 (55 mg, 0.5 mmol, 1.3 equiv.), 2-nitrophenol (70 mg, 0.5 mmol, 1.3 equiv.) and 1-Azido-2-isocyanoethane (36 mg, 0.375 mmol, 1.0 equiv.) giving a yield of 93 mg (0.27 mmol, 72%) of an orange solid. (Note: Minor diastereomer is given in italic). = -575.8° (c = 0.66, CHCl3);[35] 1H NMR (400.13 MHz, CDCl3): δ = 7.66 (d, J = 8.0 Hz, 1H), 7.33 (m, 1H), 7.02 (m, 1H), 6.87 (m, 1H), 4.04 (d, J = 5.6 Hz, 1H), 3.74 (m, 1H), 3.25 (m, 4H), 2.75 (m, 1H), 2.65 (m,

1H), 2.52 (dd, J = 3.6 Hz, J = 10.2 Hz, 1H), 1.89 (m, 2H), 1.65 (m, 3H), 1.26 (m, 2H); 13C NMR (100.62 MHz, CDCl3): δ = 172.1, 140.9, 139.9, 132.6, 125.2, 119.1, 117.0, 68.2, 57.6, 49.6, 48.8, 42.5, 37.7, 30.9, 30.5, 24.7; 1H NMR (400.13 MHz, CDCl3): δ = 7.66 (d, J = 8.0 Hz, 1H), 7.21 (m, 1H), 7.02 (m, 1H), 6.87 (m, 1H), 4.52 (d, J = 5.6 Hz, 1H), 3.74 (m, 1H), 3.25 (m, 4H), 3.00 (m, 1H), 2.65 (m, 1H), 2.52 (dd, J = 3.6 Hz, J = 10.2 Hz, 1H), 1.89 (m, 2H), 1.65 (m, 3H), 1.26 (m, 2H); 13C NMR (100.62 MHz, CDCl3): δ = 170.3, 141.0, 139.9, 132.3, 124.9, 119.4, 116.8, 65.6, 58.5), 49.7, 45.9, 42.0, 37.3, 28.9, 26.7, 25.0; IR (neat): νmax (cm-1) = 3367, 2951, 2870, 2098, 1664, 1510, 1277, 729, 541; HR-MS (ESI, 4500 V): m/z calcd for C16H21N6O3

([M+H]+): 345.1670, found 345.1654.

Compound 36: General procedure 2 was followed using 3-azabicyclo[3.3.0]oct-2-ene 34 (55 mg, 0.5 mmol, 2.0 equiv.), 4-nitro-3-bromophenol (54 mg, 0.25 mmol, 1.0 equiv.) and allyl isocyanide (27 mg, 0.4 mmol, 1.6 equiv.) giving a yield of 54 mg (0.14 mmol, 55%) of a sticky orange oil. (Note: Minor diastereomer is given in italic). = -122.1° (c = 0.48, CHCl3);[35] 1H NMR (400.13 MHz, CDCl3): δ = 7.91 (d, J = 9.2 Hz, 1H), 6.75 (d, J = 2.8 Hz, 1H), 6.40 (dd, J = 2.8 Hz, J = 9.2 Hz, 1H), 6.03 (bs, 1H), 5.69 (m, 1H), 5.03 (m, 2H), 3.94 (d, J = 2.4 Hz, 1H), 3.80 (m, 3H), 3.12 (dd, J = 6.8 Hz, J = 10.2 Hz, 1H), 2.75 (m, 2H), 2.04 (m,

1H), 1.75 (m, 2H), 1.60 (m, 3H); 13C NMR (100.62 MHz, CDCl3): δ = 170.5, 149.5, 137.7, 132.5), 127.4, 117.2, 116.9, 115.7, 110.3, 69.1, 54.6, 49.3, 40.7, 40.7, 31.4, 30.5, 23.8; 1H NMR (400.13 MHz, CDCl3): δ = 7.85 (d, J = 9.2 Hz, 1H), 7.09 (d, J = 2.8 Hz, 1H), 6.40 (dd, J = 2.8 Hz, J = 9.2 Hz, 1H), 7.30 (bs, 1H), 5.69 (m, 1H), 5.03 (m, 2H), 4.27 (d, J = 2.4 Hz, 1H), 3.66 (m, 3H), 3.12 (dd, J = 6.8 Hz, J = 10.2 Hz, 1H), 3.33 (m, 2H), 2.04 (m, 1H), 1.75 (m, 2H), 1.60 (m, 3H); 13C NMR (100.62 MHz, CDCl3): δ = 169.0, 149.9, 137.7, 132.6, 127.3, 117.6, 116.7, 116.0, 110.5, 66.0, 55.4, 46.4, 41.7, 33.6, 29.8, 27.1, 24.5; IR (neat): νmax (cm-1) = 3286, 2951, 2866, 1589, 1302, 727, 646; HR-MS (ESI, 4500 V): m/z calcd for C17H21BrN3O3

([M+H]+): 394.0761, found 394.0754.

Page 143: Biocatalysis & Multicomponent Reactions: The Ideal …...Biocatalysis & Multicomponent Reactions Chapter 2 Monoamine Oxidase N: 41 A Promising Biocatalyst for Asymmetric Synthesis

Stereoselective Synthesis of N-aryl Amides

143

Compound 37: General procedure 2 was followed using saturated bridged (3S,7R)-38 (68 mg, 0.5 mmol, 2.0 equiv.), 3-bromo-4-nitrophenol (54 mg, 0.25 mmol, 1.0 equiv.) and allyl isocyanide (27 mg, 0.040 ml, 0.4 mmol, 1.5 equiv.), giving a yield of 35 mg (0.08 mmol, 34%) of a yellow solid. = -90.6° (c = 0.53, CHCl3); 1H NMR (400.13 MHz, CDCl3): δ = 7.95 (d, J = 9.2 Hz, 1H), 6.83 (d, J = 2.8 Hz, 1H), 6.49 (dd, J = 2.8 Hz, J = 9.2 Hz, 1H), 5.70 (m, 2H), 5.03 (m, 3H), 4.14 (s, 1H), 3.78 (m, 2H), 3.56 (d, J = 1.6 Hz, 1H), 3.47 (m, 1H), 2.74 (m, 1H), 2.41 (m, 1H), 2.24 (m, 1H), 1.51 (m, 2H), 1.14 (m, 4H); 13C NMR (100.62 MHz, CDCl3):

δ = 172.0, 149.6, 138.3, 133.6, 128.9, 118.3, 117.8, 116.6, 110.9, 63.5, 50.5, 49.3, 42.5, 42.0, 41.82), 41.8, 41.2, 22.8, 22.1; IR (neat): νmax (cm-1) = 3283, 2953, 2869, 1651, 1589, 1304, 837, 744; HR-MS (ESI, 4500 V): m/z calcd for C19H23BrN3O3

([M+H]+): 420.0917, found 420.0897.

6.5 References & Notes[1] A. Domling, Chem. Rev. 2006, 106, 17-89.[2] A. Domling and I. Ugi, Angew. Chem. Int. Ed. 2000, 39, 3169-3210.[3] B. Ganem, Acc. Chem. Res. 2009, 42, 463-472.[4] R. V. A. Orru and M. de Greef, Synthesis 2003, 1471-1499.[5] J. Zhu and H. Bienayme, Multicomponent Reactions, Wiley-VCH, Weinheim, 2005.[6] I. Akritopoulou-Zanze, Curr. Opin. Chem. Biol. 2008, 12, 324-331.[7] P. R. Andreana, C. C. Liu and S. L. Schreiber, Org. Lett. 2004, 6, 4231-4233.[8] S. E. Denmark and Y. Fan, J. Am. Chem. Soc. 2003, 125, 7825-7827.[9] U. Kusebauch, B. Beck, K. Messer, E. Herdtweck and A. Domling, Org. Lett. 2003, 5, 4021-4024.[10] S. C. Pan and B. List, Angew. Chem. Int. Ed. 2008, 47, 3622-3625.[11] D. J. Ramon and M. Yus, Angew. Chem. Int. Ed. 2005, 44, 1602-1634.[12] T. Yue, M. X. Wang, D. X. Wang, G. Masson and J. P. Zhu, J. Org. Chem. 2009, 74, 8396-8399.[13] V. Kohler, K. R. Bailey, A. Znabet, J. Raftery, M. Helliwell and N. J. Turner, Angew. Chem. Int. Ed. 2010,

49, 2182-2184.[14] A. Znabet, E. Ruijter, F. J. J. de Kanter, V. Kohler, M. Helliwell, N. J. Turner and R. V. A. Orru, Angew.

Chem. Int. Ed. 2010, 49, 5289-5292.[15] A. Znabet, J. Zonneveld, E. Janssen, F. J. J. De Kanter, M. Helliwell, N. J. Turner, E. Ruijter and R. V. A.

Orru, Chem. Commun. 2010, 46, 7706-7708.[16] A. Znabet, M. M. Polak, E. Janssen, F. J. J. de Kanter, N. J. Turner, R. V. A. Orru and E. Ruijter, Chem.

Commun. 2010, 46, 7918-7920.[17] L. El Kaim, L. Grimaud and J. Oble, Angew. Chem. Int. Ed. 2005, 44, 7961-7964.[18] W. J. Evans and S. Smiles, J. Chem. Soc. 1935, 181-188.[19] F. Terrier, Chem. Rev. 1982, 82, 77-152.[20] L. El Kaim, L. Grimaud, J. Oble and S. Wagschal, Tetrahedron Lett. 2009, 50, 1741-1743.[21] S. P. Adams and R. R. Lobb, Annu. Rep. Med. Chem. 1999, 34, 179-188.[22] L. Chen, J. W. Tilley, R. W. Guthrie, F. Mennona, T. N. Huang, G. Kaplan, R. Trilles, D. Miklowski, N.

Huby, V. Schwinge, B. Wolitzky and K. Rowan, Bioorg. Med. Chem. Lett. 2000, 10, 729-733.[23] L. Chen, J. W. Tilley, T. N. Huang, D. Miklowski, R. Trilles, R. W. Guthrie, K. Luk, A. Hanglow, K. Rowan,

V. Schwinge and B. Wolitzky, Bioorg. Med. Chem. Lett. 2000, 10, 725-727.[24] L. El Kaim, M. Gizolme, L. Grimaud and J. Oble, Org. Lett. 2006, 8, 4019-4021.[25] Compound 10 can (theoretically) decompose via a retro-Diels-Alder reaction. We experimentally

observed that 10 partially decomposed during the reaction, especially at elevated temperature.[26] A. H. Hoveyda and A. R. Zhugralin, Nature 2007, 450, 243-251.

Page 144: Biocatalysis & Multicomponent Reactions: The Ideal …...Biocatalysis & Multicomponent Reactions Chapter 2 Monoamine Oxidase N: 41 A Promising Biocatalyst for Asymmetric Synthesis

Chapter 6

144

[27] K. C. Nicolaou, S. A. Snyder, T. Montagnon and G. Vassilikogiannakis, Angew. Chem. Int. Ed. 2002, 41, 1668-1698.

[28] S. Dedola, S. A. Nepogodiev and R. A. Field, Org. Biomol. Chem. 2007, 5, 1006-1017.[29] J. Hassan, M. Sevignon, C. Gozzi, E. Schulz and M. Lemaire, Chem. Rev. 2002, 102, 1359-1469.[30] T. Vlaar, E. Ruijter and R. V. A. Orru, Adv. Synth. Catal. 2011, 353, 809-841.[31] F. E. Hahn, V. Langenhahn and T. Pape, Chem. Commun. 2005, 5390-5392.[32] S. Michaelis and S. Blechert, Chem.-Eur. J. 2007, 13, 2358-2368.[33] J. L. Zhu, X. Y. Wu and S. J. Danishefsky, Tetrahedron Lett. 2009, 50, 577-579.[34] K. Hornberger, M. Cheung, M. A. Pobanz, K. A. Emmitte, K. W. Kuntz and J. G. Badiang in Vol.

Smithkline Beecham Corp, 2007.[35] Optical rotation determined for diastereomeric mixture. Value is reported for illustrative

purposes.

Page 145: Biocatalysis & Multicomponent Reactions: The Ideal …...Biocatalysis & Multicomponent Reactions Chapter 2 Monoamine Oxidase N: 41 A Promising Biocatalyst for Asymmetric Synthesis

Refl ections & Outlook

Chapter 7

Page 146: Biocatalysis & Multicomponent Reactions: The Ideal …...Biocatalysis & Multicomponent Reactions Chapter 2 Monoamine Oxidase N: 41 A Promising Biocatalyst for Asymmetric Synthesis
Page 147: Biocatalysis & Multicomponent Reactions: The Ideal …...Biocatalysis & Multicomponent Reactions Chapter 2 Monoamine Oxidase N: 41 A Promising Biocatalyst for Asymmetric Synthesis

Reflections & Outlook

147

7.1 MAO-N & Ugi-type MCRs

The demand for fast, clean, atom and step efficient chemical processes is increasing by the

day. Sustainable synthesis is not only a key feature for the chemical industry but also for

the pharmaceutical world. The key objective of the research in this thesis is to combine

two methodologies that have proven to be efficient and environmentally benign: (i)

multicomponent reactions (MCRs) and (ii) biocatalysis. These methods offer significant

advantages by reducing time, saving money, energy and raw materials, thus resulting in

both economical and environmental benefits. The structurally complex products formed

in MCRs often contain new stereocenters which are difficult to control.[1-6] For most MCRs,

catalytic asymmetric methods to control the stereochemical outcome of the reaction are

so far not available. The broad repertoire of stereospecific conversions by biocatalysts[7-8]

presents a unique opportunity to address the stereoselectivity issue of certain MCRs.

We have developed several methods (Scheme 1) making use of the enzymatic

desymmetrization of meso-pyrrolidines by means of a monoamine oxidase N (MAO-N)

from Aspergillus niger optimized by directed evolution[9] and its combination with highly

diastereoselective Ugi-type three-component (Chapter 3) and Ugi-Smiles reactions

(Chapter 6).[10-12] We were able to synthesize two distinct scaffolds; N-aryl proline amides 2

and 3,4-disubstituted prolyl peptides 3. The latter showed to be a very versatile scaffold

providing access to organocatalyst of type 4. This catalyst possessed the ability to efficiently

catalyze asymmetric 1,4-addition reactions of aldehydes to nitro-olefins.[10] Also, synthetic

alkaloids 5 could be obtained by introducing a subsequent Pictet-Spengler cyclization

(Chapter 4). It also constitutes the first example of MCR chemistry to synthesize 5-membered

ring-fused diketopiperazines.[11] The efficiency and generality of the approach and the

resulting molecular diversity and complexity makes our methodology highly interesting for

medicinal chemistry.

Page 148: Biocatalysis & Multicomponent Reactions: The Ideal …...Biocatalysis & Multicomponent Reactions Chapter 2 Monoamine Oxidase N: 41 A Promising Biocatalyst for Asymmetric Synthesis

Chapter 7

148

Scheme 1: Access to different scaffolds and follow-up chemistry using MAO-N and MCRs.

Most notably, a very short and efficient synthesis of important drug Telaprevir (Incivek™) 6,

featuring a biocatalytic desymmetrization and two multicomponent reactions as the key

steps was developed. The classical issue of lack of stereoselectivity in Ugi- and Passerini-

type reactions was overcome. The atom economic and convergent nature of the synthetic

strategy requires only very limited use of protective groups.[12]

These results show the unrivaled complexity generation of MCRs is very well complemented

by the (stereo)selectivity of enzymes. That is why the combination of biocatalysis and MCRs

is a powerful tool leading to innovative synthetic strategies of high added value compounds,

especially in a pharmaceutical context.

Future directions in this field will focus on exploring diverse combinations of biocatalysis and

complexity generating reactions and application of this methodology to make significant

contributions to future developments in combinatorial chemistry and drug discovery.

7.2 Future Directions

7.2.1 Cyanopyrrolidine derivatives

Protein activity and lifetime in organisms are extremely dependent on their processing by

proteolytic enzymes known as proteases. These proteases are responsible for various tasks

including post-translational modifications, regulation of peptide functions and digestion

Page 149: Biocatalysis & Multicomponent Reactions: The Ideal …...Biocatalysis & Multicomponent Reactions Chapter 2 Monoamine Oxidase N: 41 A Promising Biocatalyst for Asymmetric Synthesis

Reflections & Outlook

149

of proteins in smaller peptides. Most proteases execute highly specific processes and thus,

cleave a limited number of substrates. Also, some substrates are only processed by a very

small number of proteases. Over the past decades, researchers have linked proteases to

certain diseases. For example, dipeptidyl peptidase IV has been linked to type 2 diabetes,[13-15]

prolyl oligopeptidase has been linked to neurological disorders[16-18] and cathepsin K has

been linked to osteoporosis.[19]

Several compound classes are known to exert an inhibitory effect on these proteases, one of

the most prominent being the cyanopyrrolidines 7 (Figure 1A).[20-22] These cyanopyrrolidines

form a reversible covalent bond with the active site cysteine or serine moiety of the

corresponding protease and thus, inhibit the active enzyme (Figure 1B).

Figure 1. A. General structure of cyanopyrrolidines. B. Inhibitory effect of cyanopyrrolidines is the results of reversible covalent bonding with the active site serine or cysteine of the protease.

We envision synthesizing these potential protease inhibitors by making use of the enzymatic

desymmetrization of meso-pyrrolidines by means of MAO-N followed by a Strecker-type

reaction[9] and in situ addition to acyl chlorides (Scheme 2).

Scheme 2: Approach towards asymmetric synthesis of cyanopyrrolidines 11.

To demonstrate the feasibility of this combination we synthesized cyanopyrrolidines 11a-e

in high yield and selectivity.

Page 150: Biocatalysis & Multicomponent Reactions: The Ideal …...Biocatalysis & Multicomponent Reactions Chapter 2 Monoamine Oxidase N: 41 A Promising Biocatalyst for Asymmetric Synthesis

Chapter 7

150

Figure 2: Cyanopyrrolidines compounds 11a-e obtained using optically pure 12.

The use of acid chlorides derived from amino acids as inputs would give access to amino

acyl pyrrolidine 17 (Scheme 3) and these compounds show even higher inhibitory effects.[23]

We envision starting with commercially available Boc-protected amino acid 13 which is

converted under neutral conditions to the corresponding acyl chloride 14 using Ghosez’s

reagent.[24] These acid chlorides would be added to the in situ generated amino nitrile 10

(Scheme 3) giving Boc-amino acyl pyrrolidine 15.

Scheme 3: Approach towards asymmetric synthesis of amino acyl pyrrolidines.

7.2.2 Amino acid derivatives

It is evident that imine 1 has great potential in the combination with MCRs or complexity

generating reactions. However, amino acid derivatives 16[9] opens up new possibilities for

innovative novel synthetic strategies of pharmaceutically interesting compounds (Scheme 4).

Page 151: Biocatalysis & Multicomponent Reactions: The Ideal …...Biocatalysis & Multicomponent Reactions Chapter 2 Monoamine Oxidase N: 41 A Promising Biocatalyst for Asymmetric Synthesis

Reflections & Outlook

151

Scheme 4: Access to new scaffolds via amino acid derivative 18. PG: protecting group.

Amino acid derivative 16 may be used as a starting point for a Passerini reaction.[25] The

resulting compound (17) may be pharmaceutically very interesting as it resembles several

of the Dolastatin 15 peptides.[26] Dolastatin 15 depsipeptides (e.g. Figure 3) are isolated

from numerous marine cyanobacteria[27] and exhibit interesting anticancer activity.[26,

28] In addition, diastereoselectivity in this particular Passerini reaction is expected to be

considerable as previously been reported when chiral carboxylic acids were involved.[29]

Figure 3: Dolastatin 15 isolated from extracts of the Indian Ocean sea hare Dolabella auricularia.

Furthermore, an intramolecular Ugi reaction (Scheme 4, 18) would also be attractive as this

Ugi-5-centre-4-component reaction (U-5C-4CR)[30] would result in compounds (18) with

potentially interesting biological properties like antiprion[31] (Creutzfeldt–Jakob disease,

Gerstmann–Sträussler–Scheinker syndrome) and antidiabetic activity.[32] Because it is known

Page 152: Biocatalysis & Multicomponent Reactions: The Ideal …...Biocatalysis & Multicomponent Reactions Chapter 2 Monoamine Oxidase N: 41 A Promising Biocatalyst for Asymmetric Synthesis

Chapter 7

152

that α-amino acid derivatives of general formula 19 are capable of significantly controlling

the diastereoselectivity (Scheme 5),[30, 33] we expect this to also be the case when amino acid

derivative 16 is employed in an U-5C-4CR and so, to be most likely highly diastereoselective.

Scheme 5: α-Amino acid derivative 19 employed in U-5C-4CRs resulting in compound 20.

As organocatalysis has become an efficient tool for the synthesis of chiral building blocks,

amino acid 16 due to its versatility, might not only be used as input for a MCR but may

well serve as catalyst mimicking the well-established catalyst; proline. The first example of

proline as a catalyst was developed in the 1970s and is now known as the Hajos–Parrish

reaction.[34] This asymmetric aldol reaction employs achiral triketone 21 and requires just 3%

of L-proline to obtain the optically pure reaction product (22) in high yield and enantiomeric

excess (Scheme 6).

Scheme 6: First example of an asymmetric aldol reaction employing chiral proline.

Later, List and coworkers developed the first proline-catalyzed asymmetric intermolecular

aldol reaction.[35] This reaction between acetone and various aldehydes resulted in moderate

to good yields and enantiomeric excess. A representative example is depicted in scheme 7.

Scheme 7: Proline-catalyzed asymmetric intermolecular aldol condensation between acetone and 4-nitrobenzaldehyde.

Page 153: Biocatalysis & Multicomponent Reactions: The Ideal …...Biocatalysis & Multicomponent Reactions Chapter 2 Monoamine Oxidase N: 41 A Promising Biocatalyst for Asymmetric Synthesis

Reflections & Outlook

153

Therefore, determining if amino acid derivative 24 (Figure 4) due to its steric bulk, would

give better selectivity in an asymmetric intermolecular aldol condensation (compared

to L-proline) would be highly recommended. More interesting would be to establish the

selectivity of amino acid derivative 25 (Figure 4) since List and coworkers have reported

that trans-4-Hydroxy-L-proline shows slightly improved enantioselectivity in relation to

L-proline.[35] It seems that electronic properties or hydrogen bonding influence selectivity

of these kinds of catalysts.

Figure 4: Potential catalysts in asymmetric aldol condensations.

7.2.3 Alcohol Dehydrogenase

This thesis has focused entirely on the use of MAO-N as biocatalyst. Nevertheless, numerous

possibilities exist in combing different biocatalysts with multicomponent chemistry.

Alcohol dehydrogenase (ADH, EC 1.1.1.1) for example, is an attractive target enzyme. This

biocatalyst catalyzes the interconversion between alcohols and aldehydes or ketones and is

mostly NAD(P)H-dependent. Since reduction of a carbonyl group requires stoichiometrical

amounts of expensive cofactor NAD(P)H, regeneration through in situ reduction or whole

cells systems are usually applied. An alcohol dehydrogenase employed as a whole-cell

preparation for the oxidation/reduction with regeneration of the cofactor with acetone and

i-propanol, respectively is ADH from Rhodococcus ruber 44541 (Scheme 8).[36-38] Advantages

of this ADH are its ability to tolerate co-solvents, primarily acetone (25% v/v) and isopropanol

(50% v/v), which, in addition enhance the solubility of lipophilic substrates. Moreover,

these co-solvents can act as co-substrates for ADH activity in the reductive and oxidative

directions, respectively, providing a means for in situ recycling of the requisite nicotinamide

cofactor.

Scheme 8: Oxidation/Reduction of alcohol/ketones utilizing Rhodococcus ruber DSM 44541.

Page 154: Biocatalysis & Multicomponent Reactions: The Ideal …...Biocatalysis & Multicomponent Reactions Chapter 2 Monoamine Oxidase N: 41 A Promising Biocatalyst for Asymmetric Synthesis

Chapter 7

154

Whereas Ugi and Passerini reactions typically suffer from lack of stereochemical

induction, the Petasis (or borono-Mannich) MCR (Scheme 9) generally displays excellent

diastereoselectivities with respect to both the amine and aldehyde input. Thus,

enantiomerically pure aldehydes can also be considered valuable inputs for the Petasis MCR.

Especially α-hydroxy aldehydes have been reported to give excellent induction.[39]

Scheme 9: The Petasis multicomponent reaction employing amines, aldehydes and boronic acids.

The chemical synthesis of enantiomerically pure α-hydroxy aldehydes is not straightforward.

Therefore, synthesis of these aldehydes via enzymatic reduction of substituted keto-

aldehyde 25 using alcohol dehydrogenase would generate α-hydroxy aldehyde 26 in

high enantioselectivity (Scheme 10) which in turn is employed as input for a Petasis MCR

to generate chiral compound 27. Moreover, similar cross-coupling reactions have been

reported using aqueous media[40] and thus, possibilities of performing the biocatalysis and

MCR sequence in a one-pot fashion arise.

Scheme 10: Enzymatic reduction of substituted keto-aldehyde 25 generating α-hydroxy aldehyde 26 which in turn is employed as input for a Petasis MCR to generate chiral compound 27.

The synthesis of optically active compound 27 by means of a Petasis reaction can lead

to attractive pharmaceutical targets like antidiabetic,[41] antifungal[42], anticancer[43] and

immunosuppressant[44] agents.

Page 155: Biocatalysis & Multicomponent Reactions: The Ideal …...Biocatalysis & Multicomponent Reactions Chapter 2 Monoamine Oxidase N: 41 A Promising Biocatalyst for Asymmetric Synthesis

Reflections & Outlook

155

7.3 Conclusions & Outlook

Despite the exciting results obtained with various biocatalysis/MCR sequences presented

in this thesis, the combination of these two types of methodologies has hardly been

described in literature. Still there is a vast potential of feasible combinations between MCRs

with biocatalysis as described above. Emphasis should be placed on full stereochemical

control, one-pot procedures, further optimizing time, resource and energy efficiency.

Ideally, medicinally relevant classes of compounds should be targeted, thereby contributing

significantly to future developments in combinatorial chemistry and drug discovery as well

as process chemistry.

7.4 Experimental Section

General Information

Starting materials and solvents were purchased from Sigma-Aldrich and were used without

further purification. (1R,2S,6R,7S)-4-methyl-4-azatricyclo[5.2.1.02.6]dec-8-ene was prepared

according to literature procedure.[45] Column chromatography was performed on silica gel.1H and 13C NMR spectra were recorded on a Bruker Avance 400 (400.13 MHz for 1H and 100.61

MHz for 13C) or Bruker Avance 500 (500.23 MHz for 1H and 125.78 MHz for 13C) in CDCl3 and

DMSO. Chemical shifts are reported in δ values (ppm) downfield from tetramethylsilane.

Electrospray Ionization (ESI) mass spectrometry was carried out using a Bruker micrOTOF-Q

instrument in positive ion mode (capillary potential of 4500 V). Infrared (IR) spectra were

recorded neat, and wavelengths are reported in cm-1.

General Procedure 1: Preparation of optically active imine 12

Unless stated otherwise: 12 was synthesised according to literature procedure[9] with minor

adjustments. 2.5 g of freeze-dried MAO-N D5 E.Coli were rehydrated for 30 min. in 20 ml

of KPO4 buffer (100 mM, pH = 8.0) at 37 °C. Subsequently 1 mmol amine in 30 ml of KPO4

buffer (100 mM, pH = 8.0) was prepared. The pH of the solution was adjusted to 8,0 by

addition of NaOH and then added to the rehydrated cells. After 16-17 h the reaction was

stopped (conversions were > 95 %) and worked up. For workup the reaction mixture was

centrifuged at 4000 rpm and 4°C until the supernatant had clarified (40 – 60 min.). The pH

of the supernatant was then adjusted to 10-11 by addition of aq. NaOH and the supernatant

was subsequently extracted with t-butyl methyl ether or dichloromethane (4 x 70 mL). The

combined organic phases were dried with Na2SO4 and concentrated at the rotary evaporator.

General procedure 2: Preparation of optically active amino nitriles 11a-e

Unless stated otherwise: Imine (0.50 mmol, 1.0 eq.) was dissolved in 2 ml of DCM followed

by the addition trimethylsilyl cyanide (0.65 mmol, 1.3 eq.). After stirring the reaction mixture

Page 156: Biocatalysis & Multicomponent Reactions: The Ideal …...Biocatalysis & Multicomponent Reactions Chapter 2 Monoamine Oxidase N: 41 A Promising Biocatalyst for Asymmetric Synthesis

Chapter 7

156

for 1 h at RT, the acid chloride (0.65 mmol, 1.3 eq.)was added and the reaction mixture was

stirred for another 16 h at RT. The resulting mixture was concentrated in vacuo and purified

by silica gel flash chromatography (EtOAc (1): cyclohexane (2)).

Note: Two sets of rotameric signals may be observed in the NMR data.

Compound 11a: General procedure 2 was followed using imine 12, (63 mg, 0.47 mmol), trimethylsilyl cyanide (69 mg, 88 µl, 0.70 mmol) and benzoyl chloride (86 mg, 71 µl, 0.61 mmol) giving 11a as a yellow solid, yield 90%.1H NMR (400.13 MHz, CDCl3): δ 7.43–7.35 (m, 5H), 6.18 (bs, 1H), 5.97 (bs, 1H), 4.91 (s, 1H), 3.68–3.63 (m, 1H), 3.17–3.13 (m, 2H), 2.98 (bs, 2H), 2.86 (bs, 1H), 1.53–1.51 (m, 1H), 1.38-1.36 (m, 1H); 13C NMR (100.61 MHz, CDCl3): δ 169.0, 135.4, 135.2 133.7, 130.5, 130.5, 128.3, 127.1, 119,0, 51.4, 51.2, 49.5, 47.7, 46.9, 46.8, 45.5;IR (neat): νmax (cm-1) = 1624 (m),

1393 (m), 718 (m).HRMS (ESI+): calcd for C17H16N2O ([M + H]+) 265.1263, found 265.1341.

Compound 11b: General procedure 2 was followed using imine 12, (60 mg, 0.45 mmol), trimethylsilyl cyanide (61 mg, 78 μl, 0.62 mmol) and p-Toluoyl chloride (90 mg, 77 μl, 0.58 mmol) giving 11b as a white solid, yield 88%.1H NMR (500.23 MHz, CDCl3): δ 7.31–7.30 (m, 2H), 7.18–7.17 (m, 2H), 6.14 (bs, 1H), 5.92 (bs, 1H), 4.90 (s, 1H), 3.66–3.62 (m, 1H), 3.17–3.09 (m, 2H), 2.95 (bs, 2H), 2.82 (bs, 1H), 2.35 (s, 3H), 1.49–1.47 (m, 1H), 1.35–1.33 (m, 1H); 13C NMR (125.78 MHz, CDCl3): δ 168.9, 140.8, 135.1, 133.5, 132.2, 129.3, 129.3, 127.2, 127.2, 119.0, 51.2, 51.2, 49.3,

47.6, 46.8, 46.6, 45.3, 21.3; IR (neat) νmax (cm-1) = 1636 (s), 1397 (m), 1300 (m), 833 (s), 733 (s); HRMS (ESI+): calcd for C18H18N2O ([M + H]+) 279.1419, found 279.1491.

Compound 11c: General procedure 2 was followed using imine 12, (51 mg, 0.38 mmol), trimethylsilyl cyanide (52 mg, 66 μl, 0.53 mmol) and 2-quinoxaloyl chloride (96 mg, 0.50 mmol) giving 11c as a yellow solid, yield 92%.1H NMR (400.13 MHz, CDCl3): δ 9.45–9.35 (m, 1h), 8.20–8.10 (m, 2H), 7.89–7.83 (m, 2H), 6.31–6.21 (m, 1H), 6.13–5.98 (m, 1H), 4.81 (s, 1H), 4.14–3.99 (m, 1H), 3.61–3.54 (m, 1H), 3.36–3.25 (m, 1H), 3.18–2.98 (m, 3H), 1.60–1.56 (m, 1H), 1.47–1.44 (m, 1H); 13C NMR (100.61 MHz, CDCl3): δ 163.9, 146.5, 145.5, 143.0, 140.2, 136.4, 133.9,

132.1, 130.9, 129.8, 129,51, 128.7, 53.2, 51.6, 49.3, 48.1, 47.2, 46.3, 42.0; IR (neat) νmax (cm-1) = 1620 (m), 1427 (m), 1385 (b), 737 (m); HRMS (ESI+): calcd for C19H16N4O ([M + H]+) 317.1324, found 317.1398.

Compound 11d: General procedure 2 was followed using imine 12, (68 mg, 0.51 mmol), trimethylsilyl cyanide (69 mg, 88 µl, 0.70 mmol) and diphenylacetyl chloride (154 mg, 0.67 mmol) giving 11d as a white solid, yield 33%.1H NMR (500.23 MHz, CDCl3): δ 7.35-7.20 (m, 10H), 6.01 (dd, J = 5.7, 3.0 Hz, 1H), 5.51 (dd, J = 5.7, 3.0 Hz, 1H), 4.87 (s, 1H), 4.47 (d, J = 2.3 Hz, 1H), 3.48–3.40 (m, 1H), 3.24–3.21 (m, 1H), 3.17-3.13 (m, 1H), 3.05–2.85 (m, 3H), 1.51–1.43 (m, 1H), 1.36–1.34 (m, 1H); 13C NMR (125.78 MHz, CDCl3): δ 169.8, 138.4, 137.9, 136.0, 134.2, 129.3, 129.3, 129.0, 129.0, 128.8, 128.8, 127.6, 127.6, 127.4, 127.4, 118.9, 56.7, 51.6, 49.9, 49.0,

48.8, 45.6, 43.3, 28.1; IR (neat) νmax (cm-1) = 1647 (m), 1393 (m), 702 (m); HRMS (ESI+): calcd for C24H22N2O ([M + H]+) 355.1732, found 355.1790.

Page 157: Biocatalysis & Multicomponent Reactions: The Ideal …...Biocatalysis & Multicomponent Reactions Chapter 2 Monoamine Oxidase N: 41 A Promising Biocatalyst for Asymmetric Synthesis

Reflections & Outlook

157

Compound 11e: General procedure 2 was followed using imine 12, (66 mg, 0.50 mmol), trimethylsilyl cyanide (67 mg, 86 µl, 0.69 mmol) and crotonoyl chloride (65 mg, 62 µl, 0.62 mmol) giving 11e as a yellow solid, yield 81%.1H-NMR (400.13 MHz, CDCl3): δ 6.92 (dq, J = 21.9, 6.9 Hz, 1H), 6.15–5.99 (m, 2H), 5,90 (dq, J = 15.0, 1.6 Hz, 1H), 4.41 (d, J = 2.3 Hz, 1H), 3.56–3.51 (m, 1H), 3.28–3.25 (m, 1H), 3.11–3.04 (m, 3H), 2.93 (bs, 1H), 1.81–1,80 (d, J = 10 Hz, 3H), 1.51–1.49 (m, 1H), 1.37–

1.35 (m, 1H); 13C NMR (125.78 MHz, CDCl3): δ 167.6, 143.6, 135.6, 134.2, 121.5, 118.8, 52.6, 49.8, 48.3, 48.2, 46.8, 46.5, 45.4, 18.2; IR (neat) νmax (cm-1) = 1659 (m), 1609 (m), 1393 (m), 968 (m), 729 (s); HRMS (ESI+): calcd for C14H16N2O ([M + H]+) 229.1263, found 229.1335.

7.5 References & Notes[1] U. Kusebauch, B. Beck, K. Messer, E. Herdtweck and A. Domling, Org. Lett. 2003, 5, 4021-4024.[2] P. R. Andreana, C. C. Liu and S. L. Schreiber, Org. Lett. 2004, 6, 4231-4233.[3] S. E. Denmark and Y. Fan, J. Am. Chem. Soc. 2003, 125, 7825-7827.[4] S. C. Pan and B. List, Angew. Chem. Int. Ed. 2008, 47, 3622-3625.[5] D. J. Ramon and M. Yus, Angew. Chem. Int. Ed. 2005, 44, 1602-1634.[6] T. Yue, M. X. Wang, D. X. Wang, G. Masson and J. P. Zhu, J. Org. Chem. 2009, 74, 8396-8399.[7] N. J. Turner, Nat. Chem. Biol. 2009, 5, 568-574.[8] K. Faber, Biotransformations in Organic Chemistry, Springer, Berlin, Heidelberg, New York, 1997.[9] V. Kohler, K. R. Bailey, A. Znabet, J. Raftery, M. Helliwell and N. J. Turner, Angew. Chem. Int. Ed. 2010,

49, 2182-2184.[10] A. Znabet, E. Ruijter, F. J. J. de Kanter, V. Kohler, M. Helliwell, N. J. Turner and R. V. A. Orru, Angew.

Chem. Int. Ed. 2010, 49, 5289-5292.[11] A. Znabet, J. Zonneveld, E. Janssen, F. J. J. De Kanter, M. Helliwell, N. J. Turner, E. Ruijter and R. V. A.

Orru, Chem. Commun. 2010, 46, 7706-7708.[12] A. Znabet, M. M. Polak, E. Janssen, F. J. J. de Kanter, N. J. Turner, R. V. A. Orru and E. Ruijter, Chem.

Commun. 2010, 46, 7918-7920.[13] A. E. Weber, J. Med. Chem. 2004, 47, 4135-4141.[14] K. Augustyns, P. Van der Veken, K. Senten and A. Haemers, Curr. Med. Chem. 2005, 12, 971-998.[15] D. Hunziker, M. Hennig and J. U. Peters, Curr. Top. Med. Chem. 2005, 5, 1623-1637.[16] Z. Szeltner and L. Polgar, Curr. Protein Pept. Sci. 2008, 9, 96-107.[17] P. T. Maennisto, J. Vanaelaeinen, A. Jalkanen and J. A. Garcia-Horsman, Drug News Perspect. 2007,

20, 293-305.[18] I. Brandt, S. Scharpe and A.-M. Lambeir, Clin. Chim. Acta 2007, 377, 50-61.[19] B. R. Troen, Drug News Perspect. 2004, 17, 19-28.[20] J. R. Li, E. Wilk and S. Wilk, J. Neurochem. 1996, 66, 2105-2112.[21] E. B. Villhauer, J. A. Brinkman, G. B. Naderi, B. E. Dunning, B. L. Mangold, M. D. Mone, M. E. Russell,

S. C. Weldon and T. E. Hughes, J. Med. Chem. 2002, 45, 2362-2365.[22] R. P. Hanzlik, J. Zygmunt and J. B. Moon, Biochim. Biophys. Acta 1990, 1035, 62-70.[23] J.-U. Peters, Curr. Top. Med. Chem. 2007, 7, 579-595.[24] A. Devos, J. Remion, A. M. Frisquehesbain, A. Colens and L. Ghosez, J. Chem. Soc., Chem. Commun.

1979, 1180-1181.[25] J. Zhu and H. Bienayme, Multicomponent Reactions, Wiley-VCH, Weinheim, 2005.[26] G. R. Pettit, Y. Kamano, C. Dufresne, R. L. Cerny, C. L. Herald and J. M. Schmidt, J. Org. Chem. 1989,

54, 6005-6006.[27] W. H. Gerwick, L. T. Tan and N. Sitachitta, Nitrogen-containing metabolites from marine

cyanobacteria, Academic Press, San Diego, 2001.

Page 158: Biocatalysis & Multicomponent Reactions: The Ideal …...Biocatalysis & Multicomponent Reactions Chapter 2 Monoamine Oxidase N: 41 A Promising Biocatalyst for Asymmetric Synthesis

Chapter 7

158

[28] Z. Cruz-Monserrate, J. T. Mullaney, P. G. Harran, G. R. Pettit and E. Hamel, Eur. J. Biochem. 2003, 270, 3822-3828.

[29] R. Frey, S. G. Galbraith, S. Guelfi, C. Lamberth and M. Zeller, Synlett 2003, 1536-1538.[30] A. Basso, L. Banfi, R. Riva and G. Guanti, Tetrahedron Lett. 2004, 45, 587-590.[31] T. Kimura, J. Hosokawa-Muto, Y. O. Kamatari and K. Kuwata, Bioorg. Med. Chem. Lett. 2011, 21,

1502-1507.[32] A. S. Bell, M. P. Deninno, M. J. Palmer and V. M. S. in Vol. US2004/220186 Pfizer Inc, 2004.[33] A. Basso, L. Banfi, R. Riva and G. Guanti, J. Org. Chem. 2005, 70, 575-579.[34] Z. G. Hajos and D. R. Parrish, J. Org. Chem. 1974, 39, 1615-1621.[35] B. List, R. A. Lerner and C. F. Barbas, J. Am. Chem. Soc. 2000, 122, 2395-2396.[36] W. Stampfer, B. Kosjek, C. Moitzi, W. Kroutil and K. Faber, Angew. Chem. Int. Ed. 2002, 41, 1014-

1017.[37] W. Stampfer, B. Kosjek, W. Kroutil and K. Faber, Biotechnol. Bioeng. 2003, 81, 865-869.[38] G. de Gonzalo, I. Lavandera, K. Faber and W. Kroutil, Org. Lett. 2007, 9, 2163-2166.[39] N. Kumagai, G. Muncipinto and S. L. Schreiber, Angew. Chem. Int. Ed. 2006, 45, 3635-3638.[40] N. Miyaura, T. Ishiyama, H. Sasaki, M. Ishikawa, M. Satoh and A. Suzuki, J. Am. Chem. Soc. 1989, 111,

314-321.[41] A. S. Davis, S. G. Pyne, B. W. Skelton and A. H. White, J. Org. Chem. 2004, 69, 3139-3143.[42] N. A. Petasis and I. Akritopoulou, Tetrahedron Lett. 1993, 34, 583-586.[43] D. R. Li, Y. Q. Tu, G. Q. Lin and W. S. Zhou, Tetrahedron Lett. 2003, 44, 8729-8732.[44] S. Sugiyama, S. Arai and K. Ishii, Tetrahedron: Asymmetry 2004, 15, 3149-3153.[45] S. Michaelis and S. Blechert, Chem.-Eur. J. 2007, 13, 2358-2368.

Page 159: Biocatalysis & Multicomponent Reactions: The Ideal …...Biocatalysis & Multicomponent Reactions Chapter 2 Monoamine Oxidase N: 41 A Promising Biocatalyst for Asymmetric Synthesis

Summary

Page 160: Biocatalysis & Multicomponent Reactions: The Ideal …...Biocatalysis & Multicomponent Reactions Chapter 2 Monoamine Oxidase N: 41 A Promising Biocatalyst for Asymmetric Synthesis
Page 161: Biocatalysis & Multicomponent Reactions: The Ideal …...Biocatalysis & Multicomponent Reactions Chapter 2 Monoamine Oxidase N: 41 A Promising Biocatalyst for Asymmetric Synthesis

Summary

161

Biocatalysis & Multicomponent Reactions: The Ideal Synergy

Asymmetric Synthesis of Substituted Proline Derivatives

The demand for fast, clean, atom and step efficient chemical processes is increasing by the

day. Sustainable synthesis is not only a key feature for the chemical industry but also for

the pharmaceutical world. The key objective of the research in this thesis is to combine

two methodologies that have proven to be efficient and environmentally benign: (i)

multicomponent reactions (MCRs) and (ii) biocatalysis. These methods offer significant

advantages by reducing time, saving money, energy and raw materials, thus resulting in

both economical and environmental benefits. The structurally complex products formed in

MCRs often contain new stereocenters which are difficult to control. For most MCRs, catalytic

asymmetric methods to control the stereochemical outcome of the reaction are so far not

available. The broad repertoire of stereospecific conversions by biocatalysts presents a

unique opportunity to address the stereoselectivity issue of certain MCRs.

We have developed several methods (Scheme 1) making use of the enzymatic

desymmetrization of meso-pyrrolidines by means of a monoamine oxidase N (MAO-N) from

Aspergillus niger optimized by directed evolution and its combination with multicomponent

chemistry.

Scheme 1: Access to different scaffolds and follow-up chemistry using MAO-N and MCRs.

Page 162: Biocatalysis & Multicomponent Reactions: The Ideal …...Biocatalysis & Multicomponent Reactions Chapter 2 Monoamine Oxidase N: 41 A Promising Biocatalyst for Asymmetric Synthesis

Summary

162

Multicomponent reactions (MCRs) combine essentially all the atoms of at least three

components in one pot for atom and process efficiency with reduced waste. MCRs are

significantly more efficient than classical multistep synthesis in terms of time and resources.

Biocatalysis employs enzymes and microorganisms in the chemical or food industry, making

manufacturing processes more environmentally friendly. They can produce high yields of

specific products with low energy use and minimal waste generation. Despite the great

potential of biocatalysis and MCRs, the combination of these two types of methodologies

to generate optically pure complex compounds has hardly been described in literature

(Chapter 1).

The main biocatalyst described in this thesis is monoamine oxidase N (MAO-N) from

Aspergillus niger. This enzyme has proven to be particularly suitable for directed evolution.

Mutant MAO-N tolerates primary, secondary, and tertiary amines as substrates. In addition,

3,4-substituted meso-pyrrolidines were shown to be excellent substrates for biocatalytic

oxidation. Moreover, the stereoselectivity of the biocatalyst is very high (ee > 94%) and no

additional cofactor is required. This makes MAO-N one of the leading biocatalysts for amine

oxidation (Chapter 2).

Since 3,4-substituted meso-pyrrolidines 1 proved excellent substrates for MAO-N,

different optically pure 3,4-disubstituted 1-pyrrolines 2 were generated (Scheme 2). These

3,4-disubstituted 1-pyrrolines offer the possibility to do a subsequent MCR since imines are

intermediates for several of these MCRs. 3,4-Disubstituted 1-pyrrolines were reacted with

carboxylic acids and isocyanides in a highly diastereoselective Ugi-type multicomponent

reaction (Joullié-Ugi 3CR, Scheme 2) to give pharmaceutically relevant substituted prolyl

peptides 3 that would require lengthy reaction sequences using other methods. The resulting

prolyl peptides can also be used as highly active organocatalysts (4) for stereoselective

conjugate addition of aldehydes to nitroolefins (Chapter 3).

Scheme 2: Desymmetrization and diastereoselective Ugi-type 3CR towards optically pure 3,4-substituted prolyl peptides

We envisioned that combining the diastereoselective Joullié-Ugi 3CR approach with

cyclization reactions could further increase the resulting molecular complexity and diversity

considerably. Reacting our 3,4-disubstituted 1-pyrrolines with a-ketocarboxylic acids and

homoveratryl isocyanide afforded the ketoamide intermediate 5 which subsequently

undergoes a Pictet–Spengler cyclization to afford highly complex alkaloid-like polycyclic

compounds (2,5-diketopiperazines; DKPs) 6 with high diversity (Scheme 3). In addition, the

experimental procedure is simple and very efficient. To the best of our knowledge, it also

constitutes the first example of MCR chemistry to synthesize 5-membered ring-fused DKPs

(Chapter 4).

Scheme 3: Diastereoselective Ugi-type 3CR followed by a Pictet-Spengler reaction resulting in DKPs 6.

The efficiency and generality of the approach and the resulting molecular diversity and

complexity makes our methodology highly interesting for medicinal chemistry. Most

notably, a very short and efficient synthesis of important hepatitis C drug Telaprevir

(Incivek™), featuring a biocatalytic desymmetrization and two multicomponent reactions

(Passerini and Joullié-Ugi 3CR) as the key steps was developed. The classical issue of lack of

stereoselectivity in Ugi- and Passerini-type reactions was circumvented. The atom economic

and convergent nature of the synthetic strategy requires only very limited use of protective

groups. The use of protective groups is limited to two intermediate methyl esters and an

acetate. The use of carbamate protective groups is avoided altogether. The synthesis

comprises only eleven steps in total (seven steps in the longest linear sequence) compared

to twenty-four in the originally reported procedure. The total yield (over the longest linear

sequence) is 45% starting from L-cyclohexylglycine methyl ester. Our approach is general

and will be applicable to many other HCV NS3 protease inhibitors that can be derived from

meso-pyrrolidines, such as e.g. Boceprevir (Victrelis™) and Narlaprevir. The combination of

synthetic efficiency and convergence in our approach allows both faster development of

second generation inhibitors and a more economical production of Telaprevir (Chapter 5).

Page 163: Biocatalysis & Multicomponent Reactions: The Ideal …...Biocatalysis & Multicomponent Reactions Chapter 2 Monoamine Oxidase N: 41 A Promising Biocatalyst for Asymmetric Synthesis

Summary

163

We envisioned that combining the diastereoselective Joullié-Ugi 3CR approach with

cyclization reactions could further increase the resulting molecular complexity and diversity

considerably. Reacting our 3,4-disubstituted 1-pyrrolines with a-ketocarboxylic acids and

homoveratryl isocyanide afforded the ketoamide intermediate 5 which subsequently

undergoes a Pictet–Spengler cyclization to afford highly complex alkaloid-like polycyclic

compounds (2,5-diketopiperazines; DKPs) 6 with high diversity (Scheme 3). In addition, the

experimental procedure is simple and very efficient. To the best of our knowledge, it also

constitutes the first example of MCR chemistry to synthesize 5-membered ring-fused DKPs

(Chapter 4).

Scheme 3: Diastereoselective Ugi-type 3CR followed by a Pictet-Spengler reaction resulting in DKPs 6.

The efficiency and generality of the approach and the resulting molecular diversity and

complexity makes our methodology highly interesting for medicinal chemistry. Most

notably, a very short and efficient synthesis of important hepatitis C drug Telaprevir

(Incivek™), featuring a biocatalytic desymmetrization and two multicomponent reactions

(Passerini and Joullié-Ugi 3CR) as the key steps was developed. The classical issue of lack of

stereoselectivity in Ugi- and Passerini-type reactions was circumvented. The atom economic

and convergent nature of the synthetic strategy requires only very limited use of protective

groups. The use of protective groups is limited to two intermediate methyl esters and an

acetate. The use of carbamate protective groups is avoided altogether. The synthesis

comprises only eleven steps in total (seven steps in the longest linear sequence) compared

to twenty-four in the originally reported procedure. The total yield (over the longest linear

sequence) is 45% starting from L-cyclohexylglycine methyl ester. Our approach is general

and will be applicable to many other HCV NS3 protease inhibitors that can be derived from

meso-pyrrolidines, such as e.g. Boceprevir (Victrelis™) and Narlaprevir. The combination of

synthetic efficiency and convergence in our approach allows both faster development of

second generation inhibitors and a more economical production of Telaprevir (Chapter 5).

Page 164: Biocatalysis & Multicomponent Reactions: The Ideal …...Biocatalysis & Multicomponent Reactions Chapter 2 Monoamine Oxidase N: 41 A Promising Biocatalyst for Asymmetric Synthesis

Summary

164

Scheme 4: Synthesis of hepatitis C NS3 protease inhibitor Telaprevir (Incivek™).

We then decided to investigate the possibility of reacting our 3,4-disubstituted 1-pyrrolines

in other MCRs. The Ugi-Smiles reaction, a variation of the Ugi reaction in which the carboxylic

acid component is substituted by an electron-deficient phenol derivative (or a heteroaromatic

analog) is a promising candidate for combination with our biocatalytic oxidation due to its

mechanistic similarity with the Ugi reaction. Reacting our 3,4-disubstituted 1-pyrrolines

with electron-deficient phenols and different isocyanides afforded the N-aryl proline (thio)

amides 7 in high diastereoslectivity (Scheme 4). These 3,4-substituted N-aryl proline (thio)

amides are pharmaceutically relevant since similar compounds have been reported to be

potent VLA-4 antagonists. Therefore, these compounds may be useful in the treatment of

diseases like asthma, atherosclerosis, rheumatoid arthritis, inflammatory bowel disease, and

multiple sclerosis. This method represents the first report of a fully asymmetric Ugi-Smiles

process (Chapter 6).

Scheme 5: Asymmetric synthesis N-aryl proline amides by MAO-N oxidation and Ugi-Smiles 3CR.

In summary, the results described in this thesis demonstrate that biocatalysis and MCRs is a

powerful combination for the production of optically pure complex compounds. We were

able to synthesize two distinct scaffolds; N-aryl proline amides and 3,4-disubstituted prolyl

peptides. The latter showed to be a very versatile scaffold providing access to organocatalysts,

synthetic alkaloids and even to the important Hepatitis C drug Telaprevir (Incivek™).

Page 165: Biocatalysis & Multicomponent Reactions: The Ideal …...Biocatalysis & Multicomponent Reactions Chapter 2 Monoamine Oxidase N: 41 A Promising Biocatalyst for Asymmetric Synthesis

Samenvatting

(Summary in Dutch)

Page 166: Biocatalysis & Multicomponent Reactions: The Ideal …...Biocatalysis & Multicomponent Reactions Chapter 2 Monoamine Oxidase N: 41 A Promising Biocatalyst for Asymmetric Synthesis
Page 167: Biocatalysis & Multicomponent Reactions: The Ideal …...Biocatalysis & Multicomponent Reactions Chapter 2 Monoamine Oxidase N: 41 A Promising Biocatalyst for Asymmetric Synthesis

Samenvatting

167

Biokatalyse & Multicomponentenreacties: De Ideale Synergie

Asymmetrische Synthese van Gesubstitueerde Proline Derivaten

De vraag naar snelle, schone, atoom- en stap- efficiënte chemische processen stijgt met

de dag. Duurzame synthese is niet alleen belangrijke voor de chemische industrie, maar

ook voor de farmaceutische wereld. De belangrijkste doelstelling van het onderzoek in

dit proefschrift is het combineren van twee methodes die hebben bewezen efficiënt en

milieuvriendelijk te zijn: (i) multicomponentreacties (MCRs) and (ii) biokatalyse. Deze

methoden bieden aanzienlijke voordelen door het verkorten van de reactietijd, geld te

besparen, efficiënter gebruik van energie en grondstoffen, hetgeen resulteert in zowel

economische als milieuvoordelen. De structureel complexe producten die in MCRs gevormd

worden bevatten vaak nieuwe stereocentra die moeilijk te controleren zijn. Voor de meeste

MCRs zijn katalytische asymmetrische methoden om de stereochemische uitkomst van

de reactie te sturen tot nu toe niet beschikbaar. Het brede repertoire van stereospecifieke

conversies door biokatalysatoren biedt een unieke gelegenheid om de stereoselectiviteit

van MCRs aan te pakken.

We hebben verschillende methoden (Schema 1) ontwikkeld die gebruik maken van de

enzymatische desymmetrisering van meso-pyrrolidines door middel van een monoamine

oxidase N (MAO-N) uit Aspergillus niger geoptimaliseerd door directed evolution in combinatie

met multicomponentchemie.

Schema 1: Toegang tot verschillende bouwstenen en vervolgchemie gebruikmakend van MAO-N en MCRs.

Page 168: Biocatalysis & Multicomponent Reactions: The Ideal …...Biocatalysis & Multicomponent Reactions Chapter 2 Monoamine Oxidase N: 41 A Promising Biocatalyst for Asymmetric Synthesis

Samenvatting

168

MCRs combineren nagenoeg al de atomen van ten minste drie componenten in één pot

voor atoom- en procesefficiëntie met een lagere afvalproductie. MCRs zijn aanzienlijk

efficiënter dan de klassieke multistapsynthese in termen van tijd en middelen. Biokatalyse

maakt gebruik van enzymen en micro-organismen en wordt vaak gebruikt in de chemische

of voedingsmiddelenindustrie, waardoor productieprocessen milieuvriendelijker efficiënter

en milieuvriendelijker worden gemaakt. Zij kunnen hoge opbrengsten van specifieke

producten genereren met een laag energieverbruik en een minimale afvalproductie.

Ondanks de potentie die biokatalyse en MCRs bezitten, is de combinatie van deze twee

methoden om optisch zuivere complexe verbindingen te maken nauwelijks beschreven in

de literatuur (Hoofdstuk 1).

De belangrijkste biokatalysator beschreven in dit proefschrift is monoamine oxidase N

(MAO-N) uit Aspergillus niger. Dit enzym heeft bewezen bijzonder geschikt te zijn voor directed

evolution. Mutante MAO-N tolereert primaire, secundaire en tertiaire amines als substraten.

Daarnaast bleken 3,4-digesubstitueerde meso-pyrrolidines uitstekend substraten te zijn

voor biokatalytische oxidatie. Bovendien is de stereoselectiviteit van de biokatalysator zeer

hoog (enantiomere overmaat > 94%) en is geen extra co-factor is vereist. Dit maakt MAO-N

een van de belangrijkste biokatalysatoren voor amine-oxidatie (Hoofdstuk 2).

Aangezien 3,4-digesubstitueerde meso-pyrrolidines 1 uitstekende substraten bleken te

zijn voor MAO-N, werden verschillende optisch zuiver 3,4-digesubstitueerde 1-pyrrolines 2

gegenereerd (Schema 2). Deze 3,4-digesubstitueerde 1-pyrrolines bieden de mogelijkheid

om vervolgens in een navolgend MCR te verrichten aangezien iminen intermediairen

zijn voor een groot aantal van deze MCRs. 3,4-digesubstitueerde 1-pyrrolines gingen een

reactie aan met carbonzuren en isocyaniden in een zeer diastereoselectieve Ugi-type

multicomponentreactie (Joullié-Ugi 3CR, Schema 2) hetgeen resulteerde in farmaceutisch

relevante prolylpeptiden 3. Deze verbindingen zouden met behulp van andere methoden

vele reactie stappen vereisen.

De resulterende prolyl peptiden kunnen ook gebruikt worden als zeer actieve

organokatalysatoren (4) voor stereoselectieve geconjugeerde addities van aldehyden aan

nitro-olefines (Hoofdstuk 3).

Page 169: Biocatalysis & Multicomponent Reactions: The Ideal …...Biocatalysis & Multicomponent Reactions Chapter 2 Monoamine Oxidase N: 41 A Promising Biocatalyst for Asymmetric Synthesis

Samenvatting

169

Schema 2: Desymmetrisatie en diastereoselectieve Ugi-type 3CR naar optisch zuiver 3,4-gesubstitueerde prolyl peptiden

Wij voorzagen dat het combineren van de diastereoselectieve Joullié-Ugi 3CR aanpak met

cyclisatiereacties de voortvloeiende moleculaire complexiteit en diversiteit aanzienlijk

zou doen toenemen. Onze 3,4-digesubstitueerde 1-pyrrolines gingen een reactie aan

met a-ketocarbonzuren en homoveratryl isocyanide hetgeen resulteerde in ketoamide

tussenproduct 5 dat vervolgens een Pictet-Spengler cyclisatie onderging en een zeer

complexe alkaloïde-achtige polycyclische verbindingen (2,5-diketopiperazines; DKPs) 6 met

een hoge diversiteit (Schema 3) opleverde. Daarnaast is de experimentele procedure van

deze aanpak eenvoudig en zeer efficiënt. Naar ons weten vormt dit ook het eerste voorbeeld

van MCR chemie in de synthese van vijf ring gefuseerde DKPs (Hoofdstuk 4).

Schema 3: Diastereoselectieve Ugi-type 3CR, gevolgd door een Pictet-Spengler reactie resulterende in DKPs 6.

De efficiëntie en de algemeenheid van de benadering en de daaruit voortvloeiende

moleculaire diversiteit en complexiteit maakt onze methodiek zeer interessant voor

medicinale chemie. Het meest opvallend was de ontwikkeling van een zeer korte en

efficiënte synthese van het belangrijke hepatitis C- medicijn Telaprevir (Incivek™), met

een biokatalytische desymmetrisering en twee multicomponentreacties (Passerini en

Joullié-Ugi 3CR) als de belangrijkste stappen. Het klassieke probleem van een gebrek aan

stereoselectiviteit in de Ugi en Passerini reactie werd omzeild. Het atoom- economische

Page 170: Biocatalysis & Multicomponent Reactions: The Ideal …...Biocatalysis & Multicomponent Reactions Chapter 2 Monoamine Oxidase N: 41 A Promising Biocatalyst for Asymmetric Synthesis

Samenvatting

170

en convergente karakter van de synthetische strategie maakt gebruik van slechts een zeer

beperkt aantal beschermgroepen. Het gebruik van beschermgroepen is beperkt tot twee

methylesters en een acetaat. Het gebruik van carbamaatbeschermgroepen wordt helemaal

vermeden. De synthese omvat slechts elf stappen in totaal (zeven stappen in de langste

lineaire sequentie) in vergelijking met vierentwintig in de oorspronkelijk gerapporteerde

procedure. De totale opbrengst (over de langste lineaire sequentie) is 45% uitgaande van

L-cyclohexylglycinemethylester. Onze aanpak is van algemeen toepassing aangezien vele

andere HCV-NS3 proteaseremmers, zoals Boceprevir (Victrelis™) en Narlaprevir kunnen

worden afgeleid van meso-pyrrolidines. De combinatie van synthetische efficiëntie en

convergentie in onze aanpak maakt het mogelijk om zowel snellere ontwikkeling van tweede

generatie-remmers als een efficiëntere productie van Telaprevir te realiseren (Hoofdstuk 5).

Schema 4: Synthese van hepatitis C NS3 proteaseinhibitor Telaprevir (Incivek™).

We hebben toen besloten om de mogelijkheid te onderzoeken of we onze

3,4-digesubstitueerde 1-pyrrolines zouden kunnen laten reageren in andere MCRs. De

Ugi-Smiles reactie, een variant van de Ugi reactie waarbij de carbonzuurcomponent

wordt vervangen door een elektronenarme fenolderivaat (of een hetero-analoog) is een

veelbelovende kandidaat voor de combinatie met onze biokatalytische oxidatie vanwege

de mechanistische gelijkenis met de Ugi reactie. De reactie van onze 3,4-digesubstitueerde

1-pyrrolines met fenolen en verschillende isocyanides resulteert in de vorming van N-aryl

proline(thio)amides 7 in hoge diastereoslectiviteit (Schema 4). Deze 3,4-gesubstitueerde

N-aryl proline(thio)amides zijn farmaceutisch relevant, omdat soortgelijke verbindingen zijn

gerapporteerd als potente VLA-4-antagonisten. Vandaar dat deze verbindingen misschien

bruikbaar zouden kunnen zijn bij de behandeling van ziekten zoals astma, atherosclerose,

reumatoïde artritis, inflammatoire darmziekten, en multiple sclerose. Deze methode is dan

ook het eerste voorbeeld van een volledig asymmetrisch Ugi-Smiles proces (Hoofdstuk 6).

Page 171: Biocatalysis & Multicomponent Reactions: The Ideal …...Biocatalysis & Multicomponent Reactions Chapter 2 Monoamine Oxidase N: 41 A Promising Biocatalyst for Asymmetric Synthesis

Samenvatting

171

Schema 5: Asymmetrische synthese van N-aryl proline amides via MAO-N oxidatie en Ugi-Smiles 3CR.

Kortom, de resultaten beschreven in dit proefschrift tonen aan dat biokatalyse en MCRs een

krachtige combinatie zijn voor het verkrijgen van optisch zuivere complexe verbindingen.

We waren in staat om twee verschillende kernstructuren te synthetiseren, namelijk N-aryl

proline-amides en 3,4-digesubstitueerde prolylpeptiden. De 3,4-digesubstitueerde

prolylpeptiden bleken veelzijdige bouwstenen te zijn en verschaften toegang tot

organokatalysatoren, synthetische alkaloïden en zelfs de belangrijke Hepatitis C medicijn

Telaprevir (Incivek ™).

Page 172: Biocatalysis & Multicomponent Reactions: The Ideal …...Biocatalysis & Multicomponent Reactions Chapter 2 Monoamine Oxidase N: 41 A Promising Biocatalyst for Asymmetric Synthesis
Page 173: Biocatalysis & Multicomponent Reactions: The Ideal …...Biocatalysis & Multicomponent Reactions Chapter 2 Monoamine Oxidase N: 41 A Promising Biocatalyst for Asymmetric Synthesis

Dankwoord

Page 174: Biocatalysis & Multicomponent Reactions: The Ideal …...Biocatalysis & Multicomponent Reactions Chapter 2 Monoamine Oxidase N: 41 A Promising Biocatalyst for Asymmetric Synthesis
Page 175: Biocatalysis & Multicomponent Reactions: The Ideal …...Biocatalysis & Multicomponent Reactions Chapter 2 Monoamine Oxidase N: 41 A Promising Biocatalyst for Asymmetric Synthesis

Dankwoord

175

Het is niet voor niets dat het dankwoord het meeste gelezen stuk is van een proefschrift.

Promoveren kun je niet alleen en het is daarom zeer evident om waardering uit te spreken

richting de mensen die hun bijdrage hebben geleverd aan het tot stand komen van dit

proefschrift. Ik ben dan ook bang dat op het moment dat dit proefschrift van de drukker

komt, ik erachter kom dat ik iemand vergeten ben te vermelden in mijn dankwoord. Bij deze;

mocht ik iemand vergeten zijn te noemen, alsnog bedankt!

Allereerst wil ik mijn promotor, prof. dr. Romano Orru bedanken. Romano, jij hebt mij de

mogelijkheid geboden om aan dit promotietraject te beginnen. Jouw steun, vertrouwen,

positieve feedback, tijd en hulp hebben ertoe geleid dat mijn project vrij soepel en zeer

aangenaam is verlopen. Verder was je hulp bij mijn mozaïek aanvraag onmisbaar.

Mijn tweede promotor, prof. dr. Marinus Groen, wil ik ook bedanken. Alhoewel we elkaar niet

veel hebben meegemaakt aangezien u met pensioen bent gegaan, ben ik u zeer dankbaar

voor de feedback die u leverde tijdens de werkbesprekingen en voor het nauwkeurig

doorlezen van mijn proefschrift.

Ook dr. Eelco Ruijter, mijn copromotor ben ik zeer erkentelijk. Wat eigenlijk voor Romano

geldt, geldt ook voor jou. Jouw enorme vakkennis en kundigheid maken je tot een groot

voorbeeld.

Prof. dr. Isabel Arends, prof. dr. Pedro Hermkens, prof. dr. Rob Leurs, dr. Tommi Meulemans

en prof. dr. Adri Minnaard wil ik bedanken voor hun bereidheid om hun tijd en expertise ter

beschikking te stellen en zitting te nemen in de lees- en/of promotiecommisie. I am grateful

to prof. dr. Nicholas Turner for reading and correcting this thesis. In addition, I want to thank

you for giving me the opportunity of staying six months at the Manchester Interdisciplinary

Centre. Without you, this thesis would not be what it is now.

Tijdens mijn promotietijd heb ik het genoegen gehad een aantal studenten te mogen

begeleiden. Karl, you were my first student and supervising you has taught me a lot about

how to deal with students. Marloes, jij hebt je Bachelor- en Master- stage bij mij gelopen

en dat heeft ons geen windeieren gelegd. Mede door jou heeft dit geresulteerd in een

publicatie en twee patenten. Job, jij was mijn eerste HLO student. We hebben heel wat

afgelachen op het lab maar natuurlijk werd er ook keihard gewerkt. Je transformatie van een

welpje in een leeuw was compleet. Ook jij hebt er een mooie publicatie aan overgehouden.

Seenakshi, ook jij was een HLO studente. Jij hebt een belangrijke bijdrage geleverd aan

het laatste hoofdstuk van mijn proefschrift. En dan Sara B., op voorspraak van jouw vriend

(lees: Tjøstil) kwam je bij mij stage lopen . Vanaf het eerste moment klikte het tussen ons.

We hebben samen, zowel op het lab als daarbuiten, veel gelachen. En ook jij hebt er een

mooie publicatie aan overgehouden. Xavier, jij was officieel mijn laatste student. In het

Page 176: Biocatalysis & Multicomponent Reactions: The Ideal …...Biocatalysis & Multicomponent Reactions Chapter 2 Monoamine Oxidase N: 41 A Promising Biocatalyst for Asymmetric Synthesis

Dankwoord

176

begin keek je de kat uit de boom maar gaandeweg kwam je steeds losser. Helaas was er

geen tijd om je resultaten in dit proefschrift te verwerken maar we zullen deze zeker nog

gaan publiceren. Alhoewel jij niet aan mijn project werkte beschouw ik jou, Bryan, toch als

een van mijn studenten. Je werd bij mij op het lab geplaatst en zodoende werd ik jouw

praktische begeleider. Ook tussen ons klikte het meteen en onze muzieksmaak (sorry Sara)

bevestigde dat alleen maar.

Verder ben ik veel dank verschuldigd aan dr. Frans de Kanter, dr. Marek Smoluch en dr.

Madeleine Helliwell. Frans, je deur stond altijd open als ik (weer) niet uit een NMR spectrum

kwam. Ook voor jou is de tijd aangebroken om afscheid te nemen van de VU. Geniet van je

pensioen! Marek, I want to thank you for the many HRMS measurements. Last but not least,

Madeleine I am grateful for all X-ray diffraction determinations.

Daarnaast waren er vele collega’s waaronder vaste staf, post-docs, PhD’s en studenten die

de afgelopen vier jaar voor een plezierige werkomgeving hebben gezorgd en die ik graag

wil bedanken.

Bas G., jouw steun in mijn eerste periode als promovendus was onmisbaar. Jij hebt mij

wegwijs gemaakt in het wereldje van het promoveren. Ik ben daarom heel blij dat je

overvliegt uit Engeland om aanwezig te zijn tijdens mijn verdediging. Ik wil jou en Nathalie

dan ook veel geluk en succes wensen in Manchester.

Valentin, I am very grateful for your help and hospitality during my stay in Manchester. The

work we did was not only pleasant and instructive, but eventually also very successful

resulting in two joint publications. Also the (weekly) walk to Brunswick Street for ‘Fish &

Chips’ was very enjoyable. I hope to see you on 26th of January.

Tjøstil, zoals Bas G. mij wegwijs gemaakt heeft in het wereldje dat promoveren heet, heb

ik geprobeerd dat bij jou te doen. Wij hebben vele discussies gevoerd, zowel over “acute”

problemen in de zuurkast, als over de theoretische aspecten van ons onderzoek. Ik ben dan

ook trots dat dit minimaal één gezamenlijke publicatie heeft opgeleverd. Aangezien ik weet

hoe belangrijk traditie voor jou is draag ik hierbij mijn plek aan de koffietafel officieel aan jou

over. Ik wens jou dan ook nog veel succes met je verdere onderzoek.

De rol van een goede secretaresse voor het doel van wetenschappelijk onderzoek is hevig

ondergewaardeerd. Hierbij wil ik dan ook Miep bedanken voor al het regelwerk dat ze voor

mij heeft verricht.

Vanzelfsprekend zijn er nog een aantal collega’s en studenten die ik wil bedanken voor de

afgelopen jaren: Sanne, Paul S., Leon, René, Lisette, Alay, Maurice, Gitte, Rob, Niels E., Niels

T., Rachel, Halil, Helen, Chris, Jeroen, Monica, Loek, prof.dr. Koop Lammertsma, prof.dr. Fritz

Bickelhaupt, Robert, Sebastiaan, Andreas, Bas de J., Federico, Federica, Wannes, Volodymyr,

Page 177: Biocatalysis & Multicomponent Reactions: The Ideal …...Biocatalysis & Multicomponent Reactions Chapter 2 Monoamine Oxidase N: 41 A Promising Biocatalyst for Asymmetric Synthesis

Dankwoord

177

Art, Matthijs, Corien, Tom, Mark R., Elwin en Guido.

Mijn paranimfen, Jalal en Illiass, ik ben zeer vereerd dat jullie mij tijdens de promotie terzijde

willen staan. Alvast bedankt voor alles!

En dan mijn zusjes, Ik wil jullie bedanken voor alle steun die ik de afgelopen jaren van jullie

heb mogen ontvangen. Het is wellicht niet altijd duidelijk geweest waar ik nou eigenlijk mee

bezig was, maar de belangstelling is altijd gebleven.

Daar aangekomen waar dankbaarheid niet meer in woorden uit te drukken is, toch maar

een enkele aanduiding. Mijn ouders hebben me altijd aangemoedigd en gesteund in mijn

schoolloopbaan en studie, en creëerden de randvoorwaarden om mijn mogelijkheden te

ontplooien.

أس

s

Page 178: Biocatalysis & Multicomponent Reactions: The Ideal …...Biocatalysis & Multicomponent Reactions Chapter 2 Monoamine Oxidase N: 41 A Promising Biocatalyst for Asymmetric Synthesis
Page 179: Biocatalysis & Multicomponent Reactions: The Ideal …...Biocatalysis & Multicomponent Reactions Chapter 2 Monoamine Oxidase N: 41 A Promising Biocatalyst for Asymmetric Synthesis

List of Publications/Patents

Page 180: Biocatalysis & Multicomponent Reactions: The Ideal …...Biocatalysis & Multicomponent Reactions Chapter 2 Monoamine Oxidase N: 41 A Promising Biocatalyst for Asymmetric Synthesis
Page 181: Biocatalysis & Multicomponent Reactions: The Ideal …...Biocatalysis & Multicomponent Reactions Chapter 2 Monoamine Oxidase N: 41 A Promising Biocatalyst for Asymmetric Synthesis

Publications/ Patents

181

“Stereoselective Synthesis of N-Aryl Proline Amides by Biotransformation/Ugi-Smiles Sequence”

A. Znabet, S. Blanken, E. Janssen, F. J. J. de Kanter, M. Helliwell, N. J. Turner, R. V. A. Orru, E.

Ruijter, Org. Biomol. Chem. 2011, accepted.

“Palladium-Catalyzed Synthesis of 4-Aminophthalazin-1(2H)-ones by Isocyanide Insertion”

Vlaar T, Ruijter E, Znabet A, Janssen E, De Kanter FJJ, Orru RVA, Org.Lett. 2011, accepted.

“Process for the Preparation of Alpha-Acyloxy Beta-Formamido Amides” E. Ruijter, R. Orru, A.

Znabet, M. Polak, N. Turner, WO2011/103933A1.

“A Process for the Preparation of Substituted Prolyl Peptides and Similar Peptidomimetics” E.

Ruijter, R. Orru, A. Znabet, M. Polak, N. Turner, WO2011/103932A1.

“A Highly Efficient Synthesis of Telaprevir by Strategic use of Biocatalysis and Multicomponent

Reactions” A. Znabet, M. M. Polak, E. Janssen, F. J. J. de Kanter, N. J. Turner, R. V. A. Orru, E.

Ruijter, Chem. Commun. 2010, 46, 7918-7920.

“Asymmetric Synthesis of Synthetic Alkaloids by a Tandem Biocatalysis/Ugi/Pictet-Spengler-type

Cyclization Sequence” A. Znabet, J. Zonneveld, E. Janssen, F. J. J. De Kanter, M. Helliwell, N. J.

Turner, E. Ruijter, R. V. A. Orru, Chem. Commun. 2010, 46, 7706-7708.

“Highly Stereoselective Synthesis of Substituted Prolyl Peptides Using a Combination of

Biocatalytic Desymmetrization and Multicomponent Reactions” A. Znabet, E. Ruijter, F. J. J. de

Kanter, V. Koehler, M. Helliwell, N. J. Turner, R. V. A. Orru, Angew. Chem. 2010, 122, 5417-5420;

Angew. Chem. Int. Ed. 2010, 49, 5289-5292.

“Enantioselective Biocatalytic Oxidative Desymmetrization of Substituted Pyrrolidines” V.

Koehler, K. R. Bailey, A. Znabet, J. Raftery, M. Helliwell, N. J. Turner, Angew. Chem. 2010, 122,

2228-2230; Angew. Chem. Int. Ed. 2010, 49, 2182-2184.

“Scope and Limitations of an Efficient Four-Component Reaction for Dihydropyridin-2-ones” R.

Scheffelaar, M. Paravidino, A. Znabet, R. F. Schmitz, F. J. J. de Kanter, M. Lutz, A. L. Spek, C. F.

Guerra, F. M. Bickelhaupt, M. B. Groen, E. Ruijter, R. V. A. Orru, J. Org. Chem. 2010, 75, 1723-

1732.

“Diastereoselective Multicomponent Synthesis of Dihydropyridones with an Isocyanide

Functionality” M. Paravidino, R. S. Bon, R. Scheffelaar, D. J. Vugts, A. Znabet, R. F. Schmitz, F. J. J.

de Kanter, M. Lutz, A. L. Spek, M. B. Groen, R. V. A. Orru, Org. Lett. 2006, 8, 5369-5372.

Page 182: Biocatalysis & Multicomponent Reactions: The Ideal …...Biocatalysis & Multicomponent Reactions Chapter 2 Monoamine Oxidase N: 41 A Promising Biocatalyst for Asymmetric Synthesis