176

Organocatalysis and Continuous Flow Reactors as Enabling Tools

  • Upload
    vankiet

  • View
    255

  • Download
    0

Embed Size (px)

Citation preview

Page 1: Organocatalysis and Continuous Flow Reactors as Enabling Tools

9HSTFMG*afagbe+

Page 2: Organocatalysis and Continuous Flow Reactors as Enabling Tools
Page 3: Organocatalysis and Continuous Flow Reactors as Enabling Tools

Aalto University publication series DOCTORAL DISSERTATIONS 43/2013

Organocatalysis and Continuous Flow Reactors as Enabling Tools in Organic Synthesis

Antti Kataja

A doctoral dissertation completed for the degree of Doctor of Science (Technology) (Doctor of Philosophy) to be defended, with the permission of the Aalto University School of Chemical Technology, at a public examination held at the Komppa Auditorium of the school on 6 April 2013 at 12.

Aalto University School of Chemical Technology Department of Chemistry Koskinen Research Group

Page 4: Organocatalysis and Continuous Flow Reactors as Enabling Tools

Supervising professor Professor Ari Koskinen Preliminary examiners Professor Stephen J. Connon, Trinity College Dublin, Ireland Associate Professor Tore Bonge-Hansen, University of Oslo, Norway Opponent Professor David Tanner, DTU, Denmark

Aalto University publication series DOCTORAL DISSERTATIONS 43/2013 © Antti Kataja ISBN 978-952-60-5061-4 (printed) ISBN 978-952-60-5062-1 (pdf) ISSN-L 1799-4934 ISSN 1799-4934 (printed) ISSN 1799-4942 (pdf) http://urn.fi/URN:ISBN:978-952-60-5062-1 Unigrafia Oy Helsinki 2013 Finland

Page 5: Organocatalysis and Continuous Flow Reactors as Enabling Tools

Abstract Aalto University, P.O. Box 11000, FI-00076 Aalto www.aalto.fi

Author Antti Kataja Name of the doctoral dissertation Organocatalysis and Continuous Flow Reactors as Enabling Tools in Organic Synthesis Publisher School of Chemical Technology Unit Department of Chemistry

Series Aalto University publication series DOCTORAL DISSERTATIONS 43/2013

Field of research Organic Chemistry

Manuscript submitted 29 October 2012 Date of the defence 6 April 2013

Permission to publish granted (date) 12 February 2013 Language English

Monograph Article dissertation (summary + original articles)

Abstract Enabling synthetic technologies are a collection of modern tools created to aid synthetic

organic chemists to achieve rapid and efficient synthesis and analysis of complex reaction products in the laboratory environment. These techniques include e.g. microwave reactors, heterogenized catalysts and continuous flow reactor systems. The application of enabling technologies to known methods of organic synthesis such as organocatalysis is paramount for chemists in adapting to a changing world where the ready availability natural resources is by no means a given.

The first part of this thesis describes studies towards elucidating the mechanism of cinchona alkaloid derived bifunctional thiourea organocatalysts. The activity of N-trityl substituted thiourea catalyst in the asymmetric Michael addition of Meldrum’s acid and other soft carbon nucleophiles to nitroalkenes was studied. The catalyst was found to be inactive with malonate diester nucleophiles, and subsequent studies revealed that the catalyst is unable to efficiently activate a non-enolic nucleophile, thus arresting the catalytic cycle.

The second part of this thesis focuses on the use of asymmetric organocatalytic ketone aldol addition in the total synthesis of marine alkaloid (–)-hennoxazole A. Two routes were investigated for the synthesis of the western pyran fragment. A prolineamide derivative was identified as the optimal catalyst. The addition of acetone to a bisoxazole aldehyde was then optimized in batch conditions and studied further in continuous flow conditions. Conditions for the O-methylation of the aldol addition product were also identified.

The final part of this thesis concentrates on the use of continuous flow reactors in industrially important aromatic nitration reactions. The nitration of vanillin was studied as the model reaction. Blockage of flow channels due to product precipitation was identified as a serious problem and the concentration of starting material stock solutions and reaction temperatures had to be adjusted accordingly. Reaction selectivity did not essentially differ from batch conditions in spite of radically different conditions, but product isolation was more difficult. Demethylation of the methoxy substituent of vanillin and subsequent nitration was identified to be the main side reaction, along with a deformylative ipso-nitration. The nitration protocol was studied with other substrates and it was found to function best for relatively electron rich aromatic compounds.

Keywords organocatalysis, continuous flow reactors, bifunctional thiourea catalysts, hydrogen bonding, proline, enamine catalysis, aldol reaction, Michael-addition, aromatic nitration

ISBN (printed) 978-952-60-5061-4 ISBN (pdf) 978-952-60-5062-1

ISSN-L 1799-4934 ISSN (printed) 1799-4934 ISSN (pdf) 1799-4942

Location of publisher Espoo Location of printing Helsinki Year 2013

Pages 171 urn http://urn.fi/URN:ISBN:978-952-60-5062-1

Page 6: Organocatalysis and Continuous Flow Reactors as Enabling Tools
Page 7: Organocatalysis and Continuous Flow Reactors as Enabling Tools

Tiivistelmä Aalto-yliopisto, PL 11000, 00076 Aalto www.aalto.fi

Tekijä Antti Kataja Väitöskirjan nimi Organokatalyysi ja jatkuvatoimiset virtausreaktorit mahdollistavina työkaluina orgaanisessa synteesissä Julkaisija Kemian tekniikan korkeakoulu Yksikkö Kemian laitos

Sarja Aalto University publication series DOCTORAL DISSERTATIONS 43/2013

Tutkimusala Orgaaninen kemia

Käsikirjoituksen pvm 29.10.2012 Väitöspäivä 06.04.2013

Julkaisuluvan myöntämispäivä 12.02.2013 Kieli Englanti

Monografia Yhdistelmäväitöskirja (yhteenveto-osa + erillisartikkelit)

Tiivistelmä Mahdollistavat synteesiteknologiat ovat kokoelma uudenaikaisia työkaluja, joiden

tavoitteena on auttaa orgaanisia kemistejä monimutkaisten reaktiotuotteiden nopeassa ja tehokkaassa synteesissä ja analysoimisessa. Näihin tekniikoihin kuuluvat mm. mikroaaltoreaktorit, immobilisoidut katalyytit sekä jatkuvatoimiset virtausreaktorit. Mahdollistavien synteesiteknologioiden soveltaminen tunnettuihin orgaanisiin synteesimenetelmiin kuten organokatalyysiin on ensiarvoisen tärkeää sopeuduttaessa muuttuvaan maailmaan, missä luonnonvarojen helppo saatavuus ei ole itsestäänselvyys.

Tämän väitöskirjan ensimmäisessä osassa on tutkittu kinkona-alkaloidien tioureajohdannaisia sekä niiden reaktiomekanismeja. N-Trityylisubstituoidun tioureakatalyytin aktiivisuutta tutkittiin Meldrumin hapon ja muiden pehmeiden hiilinukleofiilien asymmetrisessä Michael-additiossa nitroalkeeneihin. Katalyytin huomattiin olevan tehoton malonaattidiesterien suhteen, ja jatkotutkimukset osoittivat, ettei katalyytti pysty tehokkaasti aktivoimaan ei-enolimuotoisia nukleofiilejä, estäen siten katalyyttisen syklin.

Tämän väitöskirjan toisessa osakokonaisuudessa on tutkittu ketonin asymmetrisen organokatalyyttisen aldoliaddition käyttöä meriperäisen alkaloidin (–)-hennoksatsoli A:n kokonaissynteesissä. Kahta eri reittiä tutkittiin läntisen pyraanifragmentin synteesiä varten. Parhaaksi katalyytiksi osoittautui proliinin amidijohdannainen. Asetonin additio bisoksatsolialdehydiin optimoitiin ensin panosolosuhteissa ja sitä tutkittiin sen jälkeen virtausreaktorissa. Aldoliadditiotuotteen O-metyloinnille tutkittiin myös eri olosuhteita.

Tämän väitöskirjan viimeisessä osassa on tutkittu jatkuvatoimisten virtausreaktorien käyttöä teollisesti merkittävässä aromaattisessa nitrausreaktiossa. Vanilliinin nitrausta tutkittiin mallireaktiona. Virtauskanavien tukkeutuminen tuotteen saostumisen vuoksi osoittautui vakavaksi ongelmaksi, ja lähtöaineiden varastoliuosten konsentraatiota sekä reaktiolämpötilaa säädettiin tämän perusteella. Reaktion selektiivisyys ei olennaisesti eronnut panosolosuhteista huolimatta huomattavan erilaisista olosuhteista, mutta tuotteen eristäminen oli haasteellisempaa. Vanilliinin metoksiryhmän O-demetyloituminen ja sitä seurannut nitrautuminen tunnistettiin tärkeimmäksi sivureaktioksi deformyloivan ipso-nitrautumisen ohella. Nitrausmetodia sovellettiin myös muihin lähtöaineisiin ja sen todettiin toimivan parhaiten suhteellisen elektronirikkaiden aromaattisten yhdisteiden kohdalla.

Avainsanat organokatalyysi, jatkuvatoimiset virtausreaktorit, bifunktionaaliset tioureakatalyytit, vetysidokset, proliini, enamiinikatalyysi, aldolireaktio, Michael-additio, aromaattinen nitraus

ISBN (painettu) 978-952-60-5061-4 ISBN (pdf) 978-952-60-5062-1

ISSN-L 1799-4934 ISSN (painettu) 1799-4934 ISSN (pdf) 1799-4942

Julkaisupaikka Espoo Painopaikka Helsinki Vuosi 2013

Sivumäärä 171 urn http://urn.fi/URN:ISBN:978-952-60-5062-1

Page 8: Organocatalysis and Continuous Flow Reactors as Enabling Tools
Page 9: Organocatalysis and Continuous Flow Reactors as Enabling Tools

7

ACKNOWLEDGEMENTS

The research described in this thesis was carried out at the Department of Chemistry, Aalto University (Helsinki University of Technology until 1st January 2010) during 2007-2012. The research on the total synthesis of (–)-Hennoxazole A was carried out at the Department of Chemistry, University of Cambrige, during a research visit to Innovative Technology Centre and the group of Professor Steven V. Ley, from October 2009 to April 2010. The research on the continuous flow nitration of vanillin was carried out at Fermion Corporation, Espoo R&D facility from September 2011 to November 2011.

This work has been funded primarily by the National Technology Agency (TEKES) of Finland, by the project “Enabling Synthesis Technologies” (Pharma program) until February 2012. Funding for finishing the thesis was then received from the Magnus Ehrnrooth Foundation. In addition to the primary funders, Fermion Corporation, the Emil Aaltonen Foundation, Nordforsk network, Graduate School of Organic Chemistry and Chemical Biology, Aalto University and COST Action CM0804 are gratefully acknowledged for supplemental funding and conference travel grants.

I am especially grateful to my supervisor, Professor Ari Koskinen, for taking me in as a member to his research group in 2005, when I started on a research project on synthesizing potential FAAH inhibitors. Over the course of obtaining both M.Sc. and Ph.D. degrees under his supervision, I have come to admire his encyclopedic knowledge on organic chemistry, and his practical approach towards synthesis.

My warmest gratitude goes also to Professor Steven V. Ley, who accepted me as a visiting researcher to the Innovative Technology Centre at the University of Cambridge. It was both a wonderful and educational experience to visit one of the top universities in the world.

My sincere thanks to Mr. Leif Hildén and Mr. Martti Hytönen at Fermion for being able to spend two fascinating months in an industrial research environment, and to Mr. Jarkko Helminen for his insights on process engineering.

Professors Stephen J. Connon (Trinity College, Dublin) and Tore Bonge-Hansen (University of Oslo) are gratefully acknowledged for carefully revising this manuscript and for providing valuable and constructive criticism.

During the preparation of this thesis I have had the chance to consult with several experienced chemists and discuss my ideas. For these helpful and inspiring discussions I would like to thank especially Professor Dieter Seebach, Professor Petri Pihko and Dr. Pekka Joensuu.

The years spent in the laboratory were full of good times, discussions, exchange of ideas, jokes, problems etc. that we faced together as a reseach group. My big thanks to the current

Page 10: Organocatalysis and Continuous Flow Reactors as Enabling Tools

8

and past members of Group Koskinen: Mikko P., Mikko M., Esa, Aki, Aapo, Annika, Essi, Andrejs, Christopher and Oskari. Our visiting post docs have also been of great help during my research, thank you to Damien, Katja and Peter. From the University of Cambridge I would like to thank especially Dr. Ian Baxendale (currently Professor at University of Durham) and Marcus Baumann for their collaboration on the hennoxazole project, and the whole Ley group for a memorable experience in the U.K.

For help in the laboratory, synthesis of starting materials and model studies I would like to thank the following students: Aleksi Tuokko, Laura Kolsi, Sami Tuomi, Päivi Kuosmanen, Simo Halonen and Janne Posti.

For analytical services in recording LC/MS and HRMS spectra and CHN analyses I would like to thank Tiia Juhala, Johanna Mareta (Aalto University) and Ilpo Laitinen (Fermion). For NMR consultation I would like to thank Dr. Jari Koivisto (Aalto) and Dr. Peter Grice (Cambridge). For HPLC services I am grateful to Dr. Richard Turner at Cambridge and especially to Mrs. Tarja Rissanen for ungrudgingly carrying out over 200 HPLC analyses during my brief stay at Fermion. For excellent technical aid and services I would especially like to thank Mr. Melvyn Orriss and Mr. Keith Parmenter at the University of Cambridge.

The years leading to this thesis have (luckily!) also been full of life outside the laboratory. I would like to thank all my dear friends from Dominante choir, Chamber choir Kaamos and Chamber choir Krysostomos, and especially conductors Seppo, Dani and Mikko for wonderful moments in music. Thank you also to all my fellow singers in our various respective vocal groups (especially SherryBeans). True friendship has also endured through the busy years and occasional long distances, thank you especially to Erika, Olli, Janne, Kristina and Tuomo.

My big thank you and love to my parents and close family for making a lot of things possible and enabling me to carve my own path in life, and to my siblings and their families for support and encouragement over the years.

Kun ei enää ulvo kuuta, tippuvat tähdet syliin (Once you stop howling for the moon, the stars may fall into your lap) wrote Kari Hotakainen. Thank you for everything, Iida!

Page 11: Organocatalysis and Continuous Flow Reactors as Enabling Tools

9

TABLE OF CONTENTS

�ABSTRACT ............................................................................................................................... 3

TIIVISTELMÄ (Abstract in Finnish) ........................................................................................ 5

ACKNOWLEDGEMENTS ....................................................................................................... 7�

TABLE OF CONTENTS ........................................................................................................... 9�

ABBREVIATIONS .................................................................................................................. 11�

AUTHOR’S CONTRIBUTION ............................................................................................... 13�

1 INTRODUCTION ................................................................................................................. 15�

2 ASYMMETRIC ORGANOCATALYTIC MICHAEL ADDITIONS ................................. 17�

2.1 The Advent of Asymmetric Organocatalysis .............................................................. 18�

2.2 Hydrogen Bonding in Catalysis .................................................................................. 19�

2.2.1 Non-Covalent Organocatalysis ......................................................................... 19�

2.2.2 Hydrogen Bonds ............................................................................................... 20�

2.2.3 Urea and Thiourea Catalysts ............................................................................. 23�

2.2.4 Cinchona Alkaloid Derived Catalysts ............................................................... 26�

2.2.5 Thiourea Type Organocatalysts in Target-Oriented Synthesis ......................... 32�

2.3 Addition of Meldrum’s Acid to Nitroalkenes ............................................................. 34�

2.4 Results and Discussion ................................................................................................ 38�

2.5 Conclusions ................................................................................................................. 52�

3 STUDIES TOWARDS THE TOTAL SYNTHESIS OF (–)-HENNOXAZOLE A ............. 53�

3.1 Continuous Flow Technologies ................................................................................... 54�

3.1.1 Benefits and Disadvantages of Continuous Flow Conditions ........................... 54�

3.1.2 General Strategies for Continuous Flow Synthesis ........................................... 56�

3.2 Covalent Organocatalysis ............................................................................................ 59�

3.2.1 Enamine Catalysis ............................................................................................. 59�

3.2.2 Asymmetric Organocatalytic Aldol Reactions ................................................. 61�

Page 12: Organocatalysis and Continuous Flow Reactors as Enabling Tools

10

3.2.3 Asymmetric Organocatalysis in Continuous Flow Conditions ......................... 68�

3.3 Total Synthesis of (–)-Hennoxazole A ........................................................................ 73�

3.3.1 Comparison of Previous Syntheses ................................................................... 73�

3.3.2 Previous studies by the Ley group .................................................................... 77�

3.3.3 First Approach to the Synthesis of the Western Fragment ............................... 78�

3.3.4 Second Approach to the Synthesis of the Western Fragment ........................... 80�

3.3.5 Optimization of the Bisoxazole Synthesis ........................................................ 81�

3.3.6 Catalyst Screening ............................................................................................. 83�

3.3.7 Organocatalytic Aldol Step under Continuous Flow Conditions ...................... 87�

3.3.8 Methylation studies ........................................................................................... 89�

3.4 Diastereoselectivity of �-Hydroxy Ketone Reduction with Solid Supported Borohydrides ..................................................................................................................... 90�

3.5 Conclusions ................................................................................................................. 94�

4 NITRATION OF AROMATIC SUBSTRATES UNDER CONTINUOUS FLOW CONDITIONS ......................................................................................................................... 96�

4.1 Common aspects of aromatic nitration ....................................................................... 96�

4.2 Nitration of Vanillin .................................................................................................... 97�

4.2.1 Preliminary studies ............................................................................................ 98�

4.2.2 Pulse experiments ........................................................................................... 102�

4.2.3 Identification of impurities .............................................................................. 104�

4.2.4 Continuous experiments in 5 mL reactor ........................................................ 106�

4.2.5 Continuous experiments in 10 mL reactor ...................................................... 110�

4.2.6 Isolation studies and comparison between batch and flow modes .................. 113�

4.3 Reaction profile and formation of impurities ............................................................ 116�

4.4. Nitration of other substrates ..................................................................................... 119�

4.5 Conclusions ............................................................................................................... 121�

5 EXPERIMENTAL SECTION ............................................................................................ 123�

5.1 General Experimental Considerations ....................................................................... 123�

5.2 Asymmetric Organocatalytic Michael Additions ...................................................... 124�

5.3 Total Synthesis of (–)-Hennoxazole A ...................................................................... 142�

5.4 Continuous Flow Nitration of Aromatic Substrates .................................................. 156�

6 REFERENCES .................................................................................................................... 158�

Page 13: Organocatalysis and Continuous Flow Reactors as Enabling Tools

11

ABBREVIATIONS

Ac acetyl

app apparent

aq. aqueous

BEMP 2-tert-butylimino-2-diethylamino-1,3-dimethylperhydro-1,3,2-diazaphosphorine

Bn benzyl

Boc tert-butoxycarbonyl

bpr back pressure regulator

Bz benzoyl

cat. catalytic amount

Cbz carboxybenzyl

CDI N,N’-carbonyldiimidazole

CPME cyclopentylmethyl ether

Cy cyclohexyl

DAST diethylaminosulfur trifluoride

dba dibenzylideneacetone

DBU 1,8-diazabicyclo[5.4.0]undec-7-ene

DCC dicyclohexylcarbodiimide

DDQ 2,3-dichloro-5,6-dicyano-1,4-benzoquinone

DEPC diethyl cyanophosphonate

(DHQD)2PHAL hydroquinidine 1,4-phthalazinediyl ether

DIAD diisopropyl azodicarboxylate

DIPEA diisopropylethylamine

DME 1,2-dimethoxyethane

DMF dimethyl formamide

DMSO dimethyl sulfoxide

DPPA diphenylphosphoryl azide

EDG electron donating group

ee enantiomeric excess

EWG electron withdrawing group

Page 14: Organocatalysis and Continuous Flow Reactors as Enabling Tools

12

HMTA hexamethylenetetramine

HT-PFA high temperature resistant perfluoroalkoxy polymer

KHMDS potassium hexamethyldisilazane

LDA lithium diisopropylamide

MCA monochloroacetic acid

MTPA �-methoxy-�-(trifluoromethyl)phenylacetic acid

PIFA phenyliodonium bis(trifluoroacetate)

PMP p-methoxyphenyl

PMB p-methoxybenzyl

PPTS pyridinium p-toluenesulfonate

PS polymer supported

PyBroP bromo-tris-pyrrolidino phosphoniumhexafluorophosphate

QP-SA QuadraPureTM sulfonic acid resin

rt room temperature

tR retention time

Rrt relative retention time

TBAF tetra-n-butylammonium fluoride

TBS tert-butyldimethylsilyl

TBDPS tert-butyldiphenylsilyl

TES triethylsilyl

Tf trifluoromethanesulfonyl

TFA trifluoroacetic acid

TFAA trifluoroacetic anhydride

THF tetrahydrofuran

TIPS triisopropylsilyl

TMS trimethylsilyl

Page 15: Organocatalysis and Continuous Flow Reactors as Enabling Tools

13

AUTHOR’S CONTRIBUTION

The author designed and carried out the experiments, recorded the analytical data (with exceptions mentioned in Acknowledgments), and interpreted the results that are presented in this thesis. Publications relevant to this thesis were written by the author.56

Page 16: Organocatalysis and Continuous Flow Reactors as Enabling Tools

14

Page 17: Organocatalysis and Continuous Flow Reactors as Enabling Tools

15

1 INTRODUCTION

Organic synthesis is a fundamental branch of chemistry upon which many givens of modern

life are built. The production and development of existing and new pharmaceuticals, modern

polymer and material sciences, and the production of industrial fine chemicals such as

perfumes and other cosmetics are strongly dependent on organic synthesis and skillful people

who are able to apply the corresponding technologies to use. In the modern world of

dwindling natural resources, this science is ever more important as the industry and academia

need to learn to achieve more with less.

The origin of synthetic technologies traces back to the 19th century and to the great leaps

taken then by experimental chemists in constructing specialized glassware for use in the

laboratory. However, the development of these technologies nearly came to a standstill in the

20th century. Still today undergraduate students are taught to perform organic syntheses in

round bottomed flasks not unlike those e.g. Gustaf Komppa used in the first total synthesis of

a stereochemically elaborate natural product, camphor in 1903. Whereas these basic skills are

important in themselves, the development of new and efficient tools is paramount in facing

the challenges of a changing world. Enabling synthetic technologies comprise a collection of

modern tools to aid the synthetic organic chemist in the laboratory environment.1 To mention

a few, the late 20th and early 21st centuries have seen the rise of solid phase synthesis,

combinatorial chemistry, automation, microwave reactors, continuous flow technologies, new

solvent systems such as ionic liquids and supercritical fluids, and heterogenized chiral

catalysts and biocatalysts as new chemical tools. The purpose of these technologies is to

enable the chemist to reliably synthesize and analyze complex products with high efficiency

and speed, so that the needs of the fast paced industrial and academic research environments

are met.

The importance of asymmetric synthesis and catalysis to modern chemical industry and

academic research cannot be overemphasized, as was shown by awarding the 2001 Nobel

Prize in Chemistry to Ryoji Noyori, William Knowles and K. Barry Sharpless for their

Page 18: Organocatalysis and Continuous Flow Reactors as Enabling Tools

16

development of asymmetric hydrogenation and oxidation reactions. The term Organocatalysis

was recoined2 in the early 2000s as a way to describe the general transformation of substrates

to new compounds catalyzed by small molecules usually consisting only of carbon, hydrogen,

oxygen, nitrogen, sulfur, and/or phosphorus. While in the strictest sense a huge portion of

standard organic chemistry falls into this category (all reactions catalyzed by organic acids

and bases), in terminological sense ‘organocatalysis’ is used especially to describe synthetic

transformations catalyzed by small molecules, with the main focus of modern research being

on asymmetric catalysis. The most common source of chirality in organocatalysis is the chiral

pool provided by Nature. A great variety of organocatalysts is synthesized from amino acids

or other common chiral molecules readily available from natural sources. Depending on the

exact mode of the catalyst, chiral information in organocatalysis may be transferred to the

products either by relayed asymmetric induction (e.g. via covalent intermediates in enamine

and iminium catalysis) or by external asymmetric induction (e.g. via non-covalent

organization of transition states in hydrogen bond catalysis).

This thesis describes my efforts to use asymmetric organocatalysis as an enabling tool in

organic synthesis, from both method development and target oriented points of view. I have

investigated the scope of asymmetric organocatalysis by hydrogen bonding catalysts in

connection with the synthesis of (S)-pregabalin, and proline derivatives in the construction of

stereocenters via enamine aldol reaction in the total synthesis of marine alkaloid (–)-

hennoxazole A. The application of commercial laboratory scale continuous flow reactors to

organocatalytic aldol reaction was also investigated. Finally I will discuss the use of

continuous flow reactors in the nitration of aromatic substrates as a possible tool for industrial

setting.

Page 19: Organocatalysis and Continuous Flow Reactors as Enabling Tools

17

2 ASYMMETRIC ORGANOCATALYTIC MICHAEL

ADDITIONS

The conjugate addition of a stabilized carbanion nucleophile to an �,�-unsaturated carbonyl

compound is called the Michael addition (or Michael reaction), and it is regarded as one of

the fundamental C-C-bond forming reactions in organic chemistry.3 The Michael addition is

an elegant step in the route to synthesizing more complex molecular frameworks. Classical

nucleophiles in the reaction include e.g. malonate diesters, �-ketoesters, �-ketonitriles, and

standard electrophiles are e.g. acrylonitriles, acrylate esters, or vinyl ketones (Scheme 1).

Currently, however, the term is sometimes used to encompass a wide variety of conjugate

addition reactions to C-C double bonds. In modern applications, N-, O-, S- and P-nucleophiles

have also garnered significant attention, especially in asymmetric synthesis.4 The acceptor can

be virtually any double or triple bond in �,�-relation to a resonance stabilizing electron

withdrawing group. The challenging addition into �,�-disubstituted alkenes is likewise of

particular interest for the formation of all carbon quaternary stereocenters.5 Current methods

for carrying out stereoselective Michael additions range from well known metal catalysis to

the blooming field of organocatalysis.6 Due to the focus of this thesis, we shall refrain from

reviewing metal catalyzed methods in any further detail and will concentrate on recent

advances and our own studies in organocatalytic Michael additions.

Page 20: Organocatalysis and Continuous Flow Reactors as Enabling Tools

18

Scheme 1. The Michael manifold

2.1 The Advent of Asymmetric Organocatalysis

Although asymmetric organocatalytic reactions were first documented in early 20th century by

Bredig and Fiske in the form of enantioselective cyanohydrin synthesis,7 with a few seminal

examples appearing in the intervening decades (such as the Hajos-Parrish reaction, discussed

further in Chapter 3), the current avalanche of research was launched in 2000 after the

publications of List and MacMillan.2c,8 By the beginning of the second decade of the 21st

century, several extensive reviews and books have already been published on different aspects

of asymmetric organocatalysis.9-11

Based on the mechanisms utilized by different organocatalysts, a division into two categories

follows quite naturally (Scheme 2). These categories are aptly termed covalent catalysis and

non-covalent catalysis.9b Catalysts belonging to the first category mediate their effect through

formation of new bonds between catalyst and substrate. Examples of this category include the

discoveries of List and MacMillan, who used (S)-proline and imidazolidinones as enamine10

and iminium11 forming catalysts, respectively. Catalysts of the latter category direct the

Page 21: Organocatalysis and Continuous Flow Reactors as Enabling Tools

19

substrates to react with each other through e.g. hydrogen-bond (H-bond) or ion pair formation

and subsequent advantageous spatial arrangement of transition states. Classical examples of

non-covalent catalysts include the Jacobsen and Takemoto thiourea catalysts, and quaternary

ammonium type phase transfer catalysts. Bifunctional catalysts12 may simultaneously activate

both nucleophile and electrophile, resulting in a highly organized transition state assembly,

especially if there are covalent interactions with catalyst and substrate, which usually

increases both reaction rate and stereoselectivity.9c In this thesis, experimental results

concerning both categories will be presented, with Chapter 2 focusing on the use of

bifunctional H-bonding organocatalysis, and Chapter 3 focusing on enamine catalysis,

respectively.

Scheme 2. Division of organocatalytic reactions according to catalysts

2.2 Hydrogen Bonding in Catalysis

2.2.1 Non-Covalent Organocatalysis

Non-covalently functioning organocatalysts can be divided into subcategories according to

their mode of action (Figure 1).9b Several non-covalent organocatalysts mediate their effect

Page 22: Organocatalysis and Continuous Flow Reactors as Enabling Tools

20

via H-bond formation, thus carrying out a formal Lewis acid activation of an electrophile.

Brønsted acids, such as chiral phosphoric acids, are usually considered to belong in this

category, although there usually is momentary complete acid-base dissociation between

catalyst and substrate during the reaction. A catalyst may also function as a Lewis base and

thus trigger a Lewis acidic substrate to react, such as allyl silanes or silyl enol ethers. Another

widely explored category is phase-transfer catalysis, in which an ionic catalyst species

(usually a quaternary ammonium salt) acts as a carrier for reactive ions (e.g. hydroxide ions)

between an aqueous and an organic phase.

Figure 1. Examples of non-covalent organocatalysts: bifunctional H-bond catalyst 1 (Takemoto’s catalyst), chiral Brønsted acid 2, phase transfer catalyst 3.

2.2.2 Hydrogen Bonds

Hydrogen bond is an attractive interaction between an electronegative heteroatom (H-bond

acceptor) and a hydrogen atom (H-bond donor) bound covalently to another electronegative

(hetero)atom. The H-bond donor and acceptor may be parts of the same molecule

(intramolecular H-bond) or belong to separate molecules (intermolecular H-bond). Hydrogen

bonding has a profound impact on the chemical and conformational properties of molecules,

and they have an important role in molecular recognition and supramolecular chemistry.

Hydrogen bonds may encompass a wide variety of bond strengths up to 40 kcal/mol,

depending on the steric and electronic properties of the acceptor atom and the heteroatom of

the donor fragment. Based on this, a division into three subcategories is usually made (Table

1).13 Most commonly occurring H-bonds belong to the lower moderate and weak categories,

bond strengths being typically less than 5-6 kcal/mol. The exact strength and extent of

Page 23: Organocatalysis and Continuous Flow Reactors as Enabling Tools

21

hydrogen bonding between two substrates is also strongly dependent on the surrounding

solvent.

Table 1. Classification of hydrogen bonds according to Jeffrey13

Strong Moderate Weak Type of bonding Mostly covalent Mostly electrostatic Electrostatic Length of H-bond (Å) 1.2–1.5 1.5–2.2 2.2–3.2 Bond angles (°) 175–180 130–180 90–150 Bond energy (kcal/mol) 14–40 4–15 <4

Hydrogen bonding may have several effects that are crucial to successful catalysis. They may

spatially preorganize the substrates prior to reaction, or activate bonds via polarization by

lowering the LUMO of electrophiles through a Lewis acid-base interaction between the H-

bond donor and acceptor moieties, or stabilize charges that are formed in the transition state

of the reaction.14 An early seminal example of H-bond catalysis is given in the works of Hine,

who used biphenylenediol compounds such as 5 to catalyze the epoxide opening of 4 with

diethylamine (Scheme 2).15 Another early example of achiral synthesis was published by

Curran, who used the aromatic urea 8 to catalyze the Claisen rearrangement of allyl vinyl

ether 7.16 The studies of Schreiner established that aromatic thioureas such as 12 are efficient

organic Lewis acid catalysts in e.g. Diels-Alder reactions.17

Page 24: Organocatalysis and Continuous Flow Reactors as Enabling Tools

22

Scheme 3. Examples of achiral H-bond catalysis by Hine, Curran and Schreiner

The excellent activity of urea and thiourea catalysts has been explained by their ability to act

as bidentate H-bond donors, forming two explicit hydrogen bonds with a single substrate. The

crystallization studies of Etter showed that a (thio)urea may indeed form two hydrogen bonds

with e.g. the oxygen atom in THF (Figure 2).18 Through computational studies Schreiner

postulated that in urea/thiourea structures there is also an intramolecular rigidification of the

structure due to S-H-bonding between the thiourea and the aromatic ring, which forces both

N-protons to be in S-trans conformation.17b This effect is pronounced when there are electron

withdrawing groups in meta or para position with respect to the nitrogen. Due to inductive

effects, the Csp2-H bond at the ortho position becomes more polarized, and the acidified

proton can thus function as an H-bond donor to the urea/thiourea oxygen or sulfur. Therefore

the formation of two explicit H-bonds to a carbonyl oxygen is possible, and they may

synergistically activate the substrate towards nucleophilic attack.

Page 25: Organocatalysis and Continuous Flow Reactors as Enabling Tools

23

Figure 2. The bidentate coordination of urea/thiourea with oxygen according to Etter and Schreiner. X-ray structure from Ref. 18c.

The role of hydrogen bonding in synthesis and catalysis has been under intense scrutiny

during the last decade, and the state of the art has been reviewed several times.19 Whereas it is

not practical to repeat the whole history of H-bonding in catalysis here, some seminal

developments are mentioned in the following sections.

2.2.3 Urea and Thiourea Catalysts

In 1998, the Jacobsen group reported an asymmetric Strecker reaction catalyzed by chiral

thiourea 15, derived from tert-leucine, (R,R)-1,2-diaminocyclohexane and a substituted

salicylaldehyde.20 The original aim of their research was the combinatorial synthesis and high

throughput screening of novel tridentate ligands for metal catalysts. They discovered that the

ligands themselves were able to catalyze the Strecker reaction of N-allyl benzaldimine 14 and

TBSCN or HCN in the absence of metals. The N-acylated aminonitriles 16 were formed in

good yields and enantioselectivities in the presence of the optimized catalyst 15 (Scheme 4).

Over the following decade, the Jacobsen group went on to publish a family of different urea

and thiourea catalysts for several reaction types, such as Mannich and acyl-Mannich

reaction,21a,b nitro-Mannich reaction,21c imine hydrophosphonylation,21d acyl-Pictet-Spengler

reaction,21e aza-Morita-Baylis-Hillman reaction,21f and ketone cyanosilylation,21g thus creating

a family of organocatalysts of an almost unprecedented generality (Figure 3).

Page 26: Organocatalysis and Continuous Flow Reactors as Enabling Tools

24

Scheme 4. The first organocatalytic asymmetric Strecker reaction catalyzed by thiourea 15

Figure 3. Jacobsen’s urea/thiourea catalyst family

In 2003, Takemoto reported the asymmetric Michael addition of malonates and other 1,3-

dicarbonyl compounds to nitroalkenes catalyzed by the bifunctional thiourea 1, which

similarly incorporates (R,R)-1,2-diaminocyclohexane, doubling as a chiral scaffold and a

tertiary amine base (Scheme 5).22 The acidity and ease of deprotonation of the pronucleophile

strongly affects the reaction rate, as is shown by the difference in reaction times with diethyl

malonate 23, acetylacetone 24, and malononitrile 25. Takemoto also investigated the thiourea

N-substituent, finding that the electronic properties of the substituent clearly affected the

reaction rate and selectivity. Interestingly, an unsubstituted benzene gave a poorer conversion

Page 27: Organocatalysis and Continuous Flow Reactors as Enabling Tools

25

than both EDG- or EWG-substituents at 3-position of the aromatic ring (Scheme 6), although

the increase in enantioselectivity correlates with decreasing electron density. As a general rule,

electron poor aromatic rings function best as the N-substituent, since they acidify the thiourea

N-protons and hence enhance their H-bond donating ability. It is noteworthy that a catalyst

otherwise identical to 1 but carrying a urea group with less acidic N-protons has a practically

identical activity in this reaction. A usual supposition has been that the lower pKa values of

urea usually correlate with decreasing catalytic activity and selectivity.23 Takemoto’s

observation highlights the importance of the bidentate H-bond formation for catalytic

efficiency. In other instances a urea catalyst has been noted to outperform a corresponding

thiourea in terms of selectivity.24

NH

NH

S

NMe2

CF3

F3C

1 (10 mol %)

toluene, rtPhNO2

26a

PhNO2

Nu+NuH (23-25)

EtO

O

OEt

O

23, 24 h, 86%, 93% ee

O O

24, 1 h, 80%, 89% ee

CNNC

25, 0.25 h, 85%, 25% ee

27a-29a

Scheme 5. Michael additions with Takemoto’s catalyst 1

Scheme 6. Variation of thiourea N-substituent

Page 28: Organocatalysis and Continuous Flow Reactors as Enabling Tools

26

A significant increase in the catalyst activity is observed when the number of H-bond donors

is increased, even though the electron-deficient aromatic substituent is discarded in the

process, as was reported by Wang’s group.25 The thiourea 34 catalyzed addition of cyclic �-

ketoester to nitroalkenes 26a-c proceeded with high yields and selectivities. The aromatic

substituent of the nitroalkene had little effect on the selectivity of the reaction, and aliphatic

substituents were also very well tolerated. The addition of a small H-bond donor may lead to

a tighter transition state assembly, thus explaining the excellent activity and selectivity. A

bulkier sulfonamide group in place of OH showed significantly poorer diastereo- and

enantioselectivity.

Scheme 7. Wang’s catalyst 34 featuring multiple H-bonding sites

2.2.4 Cinchona Alkaloid Derived Catalysts

The highly specialized molecular scaffold of cinchona alkaloids is considered to be a

privileged structure in both organic synthesis and pharmacology.26 In themselves, the

alkaloids have historically been used for their pharmacological properties as antimalarial

(quinine) and antiarrhythmic agents (quinidine). In addition to being held in high regard as

flavoring agents for certain refreshments, the 20th and 21st centuries have seen the widespread

application of cinchona alkaloids to asymmetric synthesis as chiral reagents, ligands for metal

mediated synthesis or as scaffolds for creating new organocatalysts. The good availability and

the relatively easy functionalization of cinchona alkaloids make them attractive and useful

starting materials for methods development. Since the chemistry and scope of cinchona

alkaloid derived catalysts has been recently reviewed,19a,27 we shall consider only a few select

and important examples here.

Page 29: Organocatalysis and Continuous Flow Reactors as Enabling Tools

27

The history of cinchona alkaloids in asymmetric catalysis goes back to the first reported

asymmetric organocatalytic reaction by Bredig and Fiske in 1912. The hydrocyanation of

benzaldehyde 36 catalyzed by quinine 37 (or quinidine) gave a chiral product, with

enantiomeric excess below 10%, however (Scheme 8).7 The next landmark was achieved in

1960, when Pracejus published the asymmetric addition of alcohols to phenyl ketene 39

catalyzed by O-acetylquinine 40.28 The reaction proceeded in good yields and, in context of

the date, in very good enantioselectivities, up to 74% ee (Scheme 9).

Scheme 8. Hydrocyanation of benzaldehyde by Bredig and Fiske

Scheme 9. Addition of methanol to phenyl ketene 39 by Pracejus

Pioneering work in both asymmetric Michael additions and H-bond catalysis was done in the

1970s and 1980s by Wynberg and his group. They reported the stereoselective Michael

additions of �-ketoesters and thiophenols to unsaturated ketones catalyzed by unmodified

cinchona-alkaloids.29 In the single best example of the latter reaction, the addition of 4-tert-

butylthiophenol 42 to 5,5-dimethylcyclohex-2-enone 43 proceeded in good yield (>80%) with

an ee of 75% when catalyzed by cinchonidine 44 (Scheme 10). Wynberg postulated that the

Page 30: Organocatalysis and Continuous Flow Reactors as Enabling Tools

28

alkaloid had a bifunctional role in the reaction. The basic quinuclidine nitrogen functions as a

base and the alkaloid thus forms an unstable ion pair with the deprotonated thiol. The enone is

then directed via H-bond formation between the C9-hydroxyl and carbonyl oxygen into the

optimal position for conjugate addition to take place via a highly structured transition state.

Any modification of the C9-hydroxyl (removal, substitution with chlorine, acetylation,

inversion) led to nearly complete deterioration of stereoselectivity.

Scheme 10. Conjugate addition of phenyl thiols to cyclohexenones by Wynberg

In 2005, the groups of Chen, Soós, Dixon and Connon independently reported the use of

cinchona-derived bifunctional thiourea catalysts 45 and 46 in different asymmetric conjugate

and Michael additions.30 Chen found high activity but poor enantioselectivities in the addition

of phenylthiol to �,�-unsaturated imide, Soós reported a highly enantioselective Michael

addition of nitromethane to chalcones, and the groups of Connon and Dixon reported high

selectivities in the Michael addition of malonate nucleophiles into nitroalkenes (Scheme 11).

All the catalysts incorporate an epi-9-amino-(9-deoxy)cinchona skeleton, and this inversion of

stereochemistry was also proven to be crucial by Soós and Connon. In their studies they

showed that thiourea catalysts having the natural configuration at C9 had little or no activity.

In another interesting coincidence, following Takemoto’s example,22 all four groups also

independently incorporated the highly electron poor 3,5-bis(trifluoromethyl)phenyl moiety as

the other thiourea N-substituent.

Page 31: Organocatalysis and Continuous Flow Reactors as Enabling Tools

29

N

NHN NHS

CF3F3C

N

HNNHN S

F3C CF345 46

OMe

Chen & Dixon Soós & Connon

Ph NH

O

Ph

O

PhSH +45 (10 mol %)

4 Å MS, CH2Cl2, 20 °C, 2 hPh N

H

O

Ph

OPhS

Chen, Ref. 30a

Ph

O

PhMeNO2 + Ph

O

Ph46 (10 mol %)

toluene, 25 °C, 110 h

O2N

99%, ee 17%

94%, ee 96%

Soós, Ref. 30b

47 48

49 50

PhNO2

26a

MeO OMe

OO

51

+46 (2 mol %)

toluene, –20 °C, 30 h PhNO2

(R)-52a

OMe

O

MeO

O

94%, ee 96%

Connon, Ref. 30c

PhNO2

26a

MeO OMe

OO

51

+45 (10 mol %)

CH2Cl2, –20 °C, 30 h PhNO2

(S)-52a

OMe

O

MeO

O

95%, ee 94%

Dixon, Ref. 30d

9 9

Scheme 11. The first cinchona derived thiourea catalysts in conjugate additions

The chemical modification required for the reversal of C9 stereochemistry was originally

reported by Brunner in 1995.31 He developed a one-pot protocol for sequential Mitsunobu

reaction32 and Staudinger reduction33 to avoid isolation of the intermediate azide (Scheme 12).

Practically all subsequent published syntheses of epi-C9-cinchona alkaloids are based on this

protocol, although diphenyl phosphory azide (DPPA) has mostly replaced the use of HN3 as

the azide source.

Page 32: Organocatalysis and Continuous Flow Reactors as Enabling Tools

30

Scheme 12. The inversion of C9 stereocenter of cinchona alkaloids

The groups of Hiemstra and Deng have explored the positioning of the H-bond donor moiety

in cinchona derived catalysts. In 2004, Deng reported the first efficient organocatalytic

Michael addition of malonates into nitroalkenes using 6’-OMe-demethylated quinine 55

(cupreine) as the catalyst.34 In spite of longish reaction times (up to 106 hours), both aromatic

and aliphatic nitroalkenes gave excellent ee:s, and good to excellent yields (Scheme 13). In a

similar fashion, Hiemstra reported the application of catalyst 56 to the asymmetric Henry

reaction.35 In the catalyst structure the thiourea is likewise attached to the 6’-position of the

quinoline ring. No inversion of configuration at C9 was required in this case. Good activity

and selectivity was attained in the addition of nitromethane to aromatic aldehydes, but

aliphatic aldehydes were in general quite unreactive under similar conditions and gave poor

enantioselectivities (Scheme 14). In spite of promising results, catalyst 56 has not found

similar popularity in the literature as 45 or 46, similar catalysts having been used only for

nitroalkene cyanosilylation,36 conjugate addition of thiols to �,�-unsaturated oxazolidin-2-

ones,37 and desymmetrization-fragmentation of cyclic ketoepoxides and ketocyclopropanes.38

Recently, Connon et al also reported studies on C5’-urea-functionalized catalysts and their

use in Henry reaction and sulfa-Michael additions.39

N

OHN

OH

55 (10 mol %)R

NO2MeO OMe

OO

51

+R

NO2

OMe

O

MeO

O

26a: R = Ph26c: R = n-C5H11

52a: R = Ph52c: R = n-C5H11

THF, –20 °C

97%, ee 99%81%, ee 94%

Scheme 13. Conjugate additions by Deng’s C6’-modified cinchona catalyst

Page 33: Organocatalysis and Continuous Flow Reactors as Enabling Tools

31

Scheme 14. Henry reaction by Hiemstra’s catalyst

In 2008, Rawal published a highly active cinchona derived squaramide catalyst 58 for the

Michael-addition of soft nucleophiles to nitroalkenes.40 The addition of acetylacetone 24 into

�-nitrostyrene 26a was carried out with a very low catalyst loading of 0.1 mol % in excellent

yield of 97% and 96% enantioselectivity (Scheme 15). The optimal catalyst loading was

found to be 0.5 mol %, with the ee climbing above 99%. The yields and enantioselectivities

were consistently high for aromatic nitroalkenes, but aliphatic nitroalkenes were for some

reason not tested. For asymmetric acyclic 1,3-diketo- or �-ketoester nucleophiles the

diastereoselectivity was rather poor (at maximum 5:1), but for cyclic substrates the ratio

climbed to 50:1.

NNH

N

NH

OO

58O O

24

PhNO2

26a

+ PhNO2

28a

O O

CH2Cl2, rt

0.5 mol % 58: 8 h, 94%, ee >99%0.1 mol % 58: 20 h, 97%, ee 96%

CF3

F3C

Scheme 15. Rawal’s squaramide catalyst 58 in nitro-Michael addition

The properties of squaramide group in anion recognition have been well studied, and it shows

better H-bond donating abilities than e.g. the corresponding urea compounds.41 This has been

Page 34: Organocatalysis and Continuous Flow Reactors as Enabling Tools

32

explained by the geometry of the squaramide group (wider N-H bite angle) and the better

anion stabilizing characteristics, which ensue from a net gain in aromaticity of the four-

membered ring upon complexation with anions.40,42 Jorgensen calculated the H-bond lengths

between nitroalkene 26a and squaramide N-H:s to be 1.97-2.22 Å,43 but no rigorous

computational and mechanistic studies on the effect of squaramide moiety in organocatalytic

reactions have been published as of yet.

2.2.5 Thiourea Type Organocatalysts in Target-Oriented Synthesis

Since most method development papers concentrate on simple substrates, some examples of

target-oriented examples should be presented.44 An impressive example is the concise and

effective synthesis of (R)-rolipram 62 by Dixon et al. (Scheme 16).45 The key step in the

synthesis of rolipram was the Michael addition of dimethyl malonate 51 into the highly

functionalized aromatic nitroalkene 59, carried out in excellent yield and enantioselectivity,

furnishing the practically enantiopure key intermediate 60 after recrystallization.

O

MeO

NO2

59

N

HNN

45 (10 mol %)

SHN

CF3F3C

CH2Cl2, 48 h at –20 °C, 48 h at 0 °CMeO OMe

OO

51

+O

MeO

NO2

OMeMeO

O O

60 87%, ee >99%afterrecrystallization

2 steps from isovanillin

NiCl2·6H2ONaBH4, EtOH

O

MeO

NH

OEtO

O

61

1) LiOH·H2O, THF

2) toluene, rfxO

MeO

NH

O

(R)-rolipram, 6263% over 6 steps, ee >99%

Scheme 16. Synthesis of (R)-rolipram 62 by Dixon

Page 35: Organocatalysis and Continuous Flow Reactors as Enabling Tools

33

Takemoto used his catalyst 1 in the total synthesis of (–)-epibatidine 66 (Scheme 17).46 The

reaction between �,�-unsaturated �-ketoester 63 and nitroalkene 64 proceeded in a good yield

of 90%, but the ee was only 75%, still being an impressive result considering the challenging

substrates and target-oriented nature of the reaction.

Scheme 17. Michael addition in the synthesis of (–)-epibatidine 66 by Takemoto

Jacobsen has used his thiourea catalysts in the total syntheses of indole alkaloids (+)-

harmicine 69 and (+)-yohimbine 73 via a Pictet-Spengler strategy.47 The synthesis of (+)-

harmicine included an acyl-Pictet-Spengler cyclization of hydroxylactam 67 mediated by

thiourea 68 (Scheme 18), which proceeded in excellent enantioselectivity of 97%.47a

Reduction of the intermediate with LiAlH4 gave (+)-harmicine. The tetrahydro-�-carboline

structure of (+)-yohimbine was obtained from 70 and 71 via an acyl-Pictet-Spengler reaction

in 81% yield over two steps and 94% ee (Scheme 19). 47b

Scheme 18. Synthesis of (+)-harmicine by Jacobsen

Page 36: Organocatalysis and Continuous Flow Reactors as Enabling Tools

34

Scheme 19. Synthesis of (+)-yohimbine 73 by Jacobsen

Takemoto’s catalyst was also used in the synthesis of (+)-esermethole 77 by Barbas III

(Scheme 20).48 Conjugate addition of N-Boc-6-methoxy-4-methyl-3-oxindole 75 to

nitroethene proceeded in moderate yield and goodish ee of 83%, which was improved to 96%

by recrystallization. Further three steps gave the final product 77, which can further be

elaborated to the important parasympathomimetic compound (+)-physostigmine.

Scheme 20. Synthesis of (+)-esermethole 77 by Barbas III

2.3 Addition of Meldrum’s Acid to Nitroalkenes

In 2009, our laboratory published an organocatalytic enantioselective synthesis of (S)-

pregabalin, also known by its trade name Lyrica® (Scheme 21).49 The key step in the

Page 37: Organocatalysis and Continuous Flow Reactors as Enabling Tools

35

synthesis involves the asymmetric addition of Meldrum’s acid 78 to a readily available

aliphatic nitroalkene 26d. After extensive screening of different catalysts, the quinidine-

derived thiourea 79 incorporating a bulky trityl substituent was recognized as the best source

of asymmetric induction, giving an ee of 75%. It is noteworthy that the benchmark catalyst 45

gave a meager enantiomeric excess of 32% and did not give full conversion in 16 hours at a

catalyst loading of 5 mol %.

Scheme 21. Organocatalytic synthesis of (S)-pregabalin 82 from our laboratory

Meldrum’s acid is a synthetically useful building block due to its easy deprotection and

possibilities for further functionalization, and along with its derivatives it has found use e.g. in

the synthesis of natural products.50 It has an intriguingly high acidity, with pKa values of 7.32

in DMSO and 4.83–4.97 in H2O.51 While this is widely regarded as an anomaly, there are

several explanations. The density functional theory calculations based on the reactive hybrid

orbital (RHO) theory developed by Ohwada have shown clear correlation between the energy

levels of unoccupied RHOs of the acidic �-C-H bonds of cyclic 1,3-dicarbonyl compounds,

including Meldrum’s acid, and their corresponding experimental deprotonation energies.52 A

classical explanation invokes the conformational stabilization of esters. An acyclic ester is

stabilized into trans conformation by several interactions between filled and unfilled orbitals:

Page 38: Organocatalysis and Continuous Flow Reactors as Enabling Tools

36

there is donation from free electron pairs of the ‘alcohol’ oxygen to the carbonyl �* and �*

orbitals, and simultaneously donation from carbonyl oxygen free pair to the C-O sigma star

(Figure 4). Meldrum’s acid has a boat like ground state conformation, and the donation effect

of alcohol oxygen lone pair to carbonyl �* is prevented by the cyclic cis arrangement. Hence

the methylene group experiences the heightened reactivity from both carbonyl groups and is

significantly acidified. Thus, Meldrum’s acid is able to form defined ion pairs with practically

any mild base available.

Figure 4. Conformational and stereoelectronic stabilization of esters and Meldrum’s acid

The use of Meldrum’s acid as a nucleophile in asymmetric synthesis has been rather limited,

and more examples exist on e.g. the use of its 5-alkylidene derivatives as electrophiles.53 At

the time of our study, the few published results on the asymmetric addition of 78 to

electrophiles were not very encouraging. Kleczkowska and Sas attempted the asymmetric

Michael-addition of 78 to �-nitrostyrene 26a by using a stoichiometric amount of different

natural alkaloids and their derivatives as catalysts (Table 2).54 In spite of high or good yields,

the enantioselectivities were uniformly low (entries 1-6 and 9-10), and no selectivity was

detected in the presence of hydrocupreidine, �-isocupreidine and (–)-brucine (entries 7-8, 11).

No rationale was given for the low selectivity. Takemoto’s catalyst 1 performed a little better,

giving the Michael adduct 80a in 88% yield and 46% ee (Entry 12).22b The thiourea

derivatives of chiral Tröger’s base were likewise reported to give practically no

stereoselectivity in the same reaction.55

Page 39: Organocatalysis and Continuous Flow Reactors as Enabling Tools

37

Table 2. Asymmetric Michael-additions of Meldrum’s acid prior to our studies.

Entrya Catalyst Yield (%) ee (%) Ref. 1 Cinchonine 88 21 (S) 54 2 Cinchonidine 92 18 (R) 54 3 Quinine 27 22 (R) 54 4 Quinidine 70 25 (S) 54 5 Hydroquinidine 70 19 (S) 54 6 (DHQD)2PHAL 63 5 (S) 54 7 Hydrocupreidine 54 0 54 8 �-Isocupreidine 71 0 54 9 L-proline 63 7 (S) 54

10 (–)-Sparteine 72 15 (R) 54 11 (–)-Brucine 82 0 54 12b 1 88 46 22b

a) The reactions were conducted with a stoichiometric amount of catalyst for 3 hours in CH2Cl2. b) 10 mol % 1, toluene, rt, 24 h

Since the publication of our study (results described below),56 only a few reports on the

asymmetric organocatalytic Michael addition of Meldrum’s acid have surfaced.57 The group

of Ellman reported a highly enantioselective addition of both 78 and cyclohexyl-Meldrum’s

acid 84 to nitroalkenes catalyzed by N-sulfinyl urea catalysts 83 and 85 (Scheme 22).57b They

applied this protocol likewise to the synthesis of (S)-pregabalin, obtaining the intermediate 86

in full conversion and ee of 92% in the key step.

Page 40: Organocatalysis and Continuous Flow Reactors as Enabling Tools

38

Scheme 22. Highly selective additions of Meldrum’s acid 78 and cyclohexyl-Meldrum’s acid 84 to nitroalkenes by Ellman

2.4 Results and Discussion

As follow-up studies to our synthesis of (S)-pregabalin, we set out to further screen the

properties of our catalyst 79. Since it outperformed the widely used 45 in the key step, it was

our aim to study whether this represented a general trend in activity and selectivity. In

addition, we wanted to investigate whether catalyst 79 would outperform the benchmark

catalyst 45 with other substrates as well. Thus, following the general procedure described

previously, a solvent screen was conducted using nitroalkene 26d as a model substrate (Table

3). For the determination of enantioselectivity, the Meldrum’s acid adducts 80 were

transformed into the corresponding anilides 87 by treatment with 1:10 mixture of aniline in

DMF at 100 °C.53c

Page 41: Organocatalysis and Continuous Flow Reactors as Enabling Tools

39

Table 3. The effect of solvent on addition of Meldrum’s acid 78 to nitroalkene 26d

Entry Solvent Time (h) Conversion (%)a ee (%)b 1 CH2Cl2 10 >95 72 2 Et2O 21 >95 69 3 THF 18 >95 72 4 PhCl 16 >95 78 5 MeCN 17 >95 56 6 DMF 18 >95 16

Reaction conditions: 0.8 mmol nitroalkene, 1.6 mmol of Meldrum’s acid, 10 mol % of 79, 0.5 mL solvent, rt. a) Determined by 1H NMR spectroscopy. b) Determined by chiral HPLC from the anilide derivative 87d.

The solvent screen showed little significant difference in enantioselectivity between CH2Cl2,

Et2O or THF, with chlorobenzene giving a maximum ee of 78% (entries 1-4). The

enantioselectivity was slightly suppressed in acetonitrile, and nearly completely eroded in

DMF (entries 5-6). This would be expected for solvents that function as good H-bond

acceptors or donors, as the solvent disrupts the optimal H-bond binding between the catalyst

and substrates. Similar tendencies have been observed with thiourea catalysts by Takemoto

and Soós.22a,30b

Although chlorobenzene gave the best enantioselectivity in the solvent screen, we chose to

complete substrate screening in the original conditions using CH2Cl2 as a solvent, due to the

lower solubility of Meldrum’s acid into chlorobenzene. Results from the substrate studies are

presented in Table 4. There is little difference in activity and selectivity between EDG- and

EWG-substituted aromatic substrates (entries 1-5), a fact noted also by Takemoto and

Wang.22b,25 Heteroaromatic substrates behave in a similar manner (entries 6-7). Surprisingly,

the best enantioselectivities were attained with the original substrate 26d. Aliphatic substrates

in general gave higher selectivities than aromatic nitroalkenes (entries 8-11). This goes

against the common trends reported in connection with bifunctional thiourea catalyzed

Michael additions. In Takemoto’s studies, the enantioselectivities dropped to 81% when

aliphatic nitroalkenes were used as substrates for the Michael-addition of diethyl malonate

23.22b Similar erosion in enantioselectivity was noted by Connon and Dixon in their respective

studies.30c,d

Page 42: Organocatalysis and Continuous Flow Reactors as Enabling Tools

40

Table 4. Addition of Meldrum’s acid to nitroalkenes 26a- l

Entry R Time (h) Conversion (%)a ee (%)b 1 Ph (26a) 8 >95 49 2 4-MeO-C6H4 (26b) 9 >95 53 3 3-MeO-C6H4 (26e) 12 >95 55 4 4-BrC6H4 (26f) 17 >95 57 5 3-BrC6H4 (26g) 8 >95 47 6 2-furyl (26h) 9 >95 47 7 3-furyl (26i) 9 >95 48 8 Cy (26j) 49 91 59 9 n-C4H9 (26k) 19 >95 68

10 i-Bu (26d) 10 >95 72 11 BnOCH2 (26l) 7 >95 62

Reaction conditions: 0.8 mmol nitroalkene, 1.6 mmol of Meldrum’s acid, 10 mol % of 79, 0.5 mL CH2Cl2, rt. a) Determined by 1H NMR. b) Determined by chiral HPLC from the anilide derivative 87a-l.

To compare the performance of our catalyst to the studies of Takemoto, Connon and Dixon,

we turned to other nucleophiles such as dimethyl malonate 51 and acetylacetone 24 (Scheme

23). To our considerable surprise, 79 was unreactive in the simple model reaction between

26a and dimethyl malonate 51, giving a racemic product in 14% yield. Using 24 as the

nucleophile showed slight improvement, giving the product 28a in 37% yield with an ee of

46%.

Scheme 23. The addition of different C-nucleophiles catalyzed by 79

Page 43: Organocatalysis and Continuous Flow Reactors as Enabling Tools

41

These results raised some questions concerning the mechanism of bifunctional thiourea

catalysts. According to the commonly accepted mechanistic hypothesis, the thiourea moiety

of the catalyst associates with the electrophile (e.g. nitroalkene), while the deprotonated

nucleophile forms an ion pair with the tertiary amine. Takemoto determined that the

organocatalytic Michael addition of a malonate nucleophile into nitrostyrene 26a follows

pseudo first order kinetics with respect to all reaction partners. No nonlinear effect was

observed with respect to the enantiomeric purity of the catalyst. Based on the kinetic data and 1H NMR titration experiments of 26a in the presence of the Schreiner thiourea 12, they

proposed a ternary complex to operate in the transition state of the reaction (Figure 5).22b The

groups of Soós and Papai carried out theoretical studies on Takemoto’s catalyst.58 Their

conclusion was that while the coordination scheme proposed by Takemoto soundly explains

the observed selectivity, a reversed situation where the nucleophile (acetylacetone) is

coordinated to the thiourea and the nitroalkene to the quaternary ammonium N-H is in fact

energetically more favorable by 2.7 kcal/mol (Figure 5, gas-phase calculations, 1.9 kcal/mol

with solvation energies accounted for). Based on their calculations, they also confirmed

Takemoto’s mechanistic speculation that the formation of an ion pair between the

deprotonated enolate form of the nucleophile, and protonated catalyst is the first step in the

catalytic cycle. It would also intuitively follow that the nucleophile stays close to its optimal

geometry due to extensive H-bond stabilization from both thiourea and quaternary ammonium

N-Hs. They also found that the same alternative route gives similar results in the cinchona

derived thiourea catalyzed addition of nitromethane into chalcones.58

Figure 5. Proposed transition states of thiourea catalyzed nitro-Michael addition by Takemoto (left) and Pápai (right) (bond lengths exaggerated for clarity)

Comparison of the pKa values of the substrates and catalyst suggests that the quinuclidine

nitrogen alone (pKaH 9.8 in DMSO)59 is by far not basic enough to deprotonate dimethyl

Page 44: Organocatalysis and Continuous Flow Reactors as Enabling Tools

42

malonat

deproto

malonat

pronucl

ternary

the thio

computa

or �-dik

calculat

nucleop

nucleop

enolizat

exists s

extent i

Figure

form.62

bonding

slower t

Figure Compouand 3.56form.

In the c

bonding

te, whereas

nated by n

te51a) in D

eophile, the

complex pr

ourea as sh

ational stud

ketone, is de

tions.22b,58,61

phile activa

philic in com

tion constan

solely in its

n nonpolar

6), and is t

This differ

g catalysts,

than those o

6. 1H NMund 51 show6 ppm are

ase of dime

g of one or

s the unusu

nearly any b

DMSO. Wh

e complete

roposed by

hown on th

dies, the 1,3

epicted in it1 For 1,3-d

ation, since

mparison to

nts of malo

s diketo for

solvents (r

thus much

rence in rea

with react

of acetylace

MR spectra ws no enol 1.90:1 (CH

ethyl malon

both carbon

ually acidic

base. Acety

hile the re

lack of rea

Takemoto

he left in F

-dicarbonyl

ts enol form

dicarbonyl

the enol O

o the uneno

onate dieste

rm (Figure

roughly 4:1

more easily

activity has

tion times

etone.22b,57a,6

of dimethcontent, butvs. CH2), r

ate, it is mo

nyl groups

Meldrum’s

ylacetone ha

eactivities w

activity in t

et al. In bo

Figure 5. In

l nucleophil

m prior to de

compound

O-H is mu

olized form

ers, and acc

6). In cont

enol:diketo

y deproton

been obser

of malonat63

hyl malonatut 24 is 80%roughly cor

ost likely th

to the thiou

s acid (pKa

as a pKa of

with 79 d

the case of

oth cases th

n Takemot

le, whether

eprotonation

s the enol

uch more a

. However,

cording to

trast, acetyl

o according

ated by ter

rved in seve

te diesters

te 51 and % enolized, trresponding

hat the depro

urea moiety

a 7.32 in D

f 13.360 (vs

o reflect t

51 does no

e nitroalken

o’s model

a malonate

n in mechan

form is m

acidic and

little or no

our observ

lacetone is

g to 1H NMR

tiary amine

eral studies

usually bei

acetylacetothe ratio of

to a 3.8:1 r

otonation st

y, thus acidi

DMSO)51a is

s. 15.9 of d

the acidity

ot sit well

ne is coordi

and corres

diester, �-k

nistic propo

more desir

the a-carbo

o data exist

vation, in C

enolized to

MR spectra s

es alone in

on bifunct

ing four to

one 24 in signals at 5ratio of eno

tep is assiste

ifying the �

s readily

dimethyl

y of the

with the

inated to

sponding

ketoester

osals and

able for

on more

ts on the

CD2Cl2 it

o a high

shown in

its enol

ional H-

o tenfold

CD2Cl2. 5.52 ppm ol-diketo

ed by H-

�-protons

Page 45: Organocatalysis and Continuous Flow Reactors as Enabling Tools

and assi

of Schre

and a d

same tim

anilinic

malonat

accordin

in its dik

Figure thioureamol % 1

Based o

N-H-pro

substitu

ability o

which i

Schrein

carbon s

of pKa

substitu

values a

isting the ba

einer thiour

downfield sh

me a maxim

nitrogen.

te pronucleo

ng to 1H NM

keto form.

7. 1H NMRa 12 in CD212. All spec

on these pre

otons of 79

uent preven

of catalyst

is supporte

ner’s measur

substituents

8.5, and t

uted 79 wou

about 19.6 i

asic amine

rea 12 in CD

hift of 0.00

mum shift o

This sugge

ophiles for

MR spectro

R shifts of 2Cl2. Blue: pctra calibrat

emises, two

9 do not co

nts this thro

45 is highl

d by Takem

rements, th

s are alkyl g

the catalyst

uld be closer

in DMSO (F

in the enoli

D2Cl2 show

05 ppm for

of 0.17 ppm

ests that th

the deproto

oscopy in th

dimethyl mpure 51; Reed relative t

different p

ordinate str

ough steric

y assisted b

moto’s, Sch

he thiourea

groups. The

t 88 lies in

r to the Jaco

Figure 8).21e

ization step

w a downfie

the O-Me

m was obser

he bidentat

onation step

he enol cont

malonate 51ed: ~10 mol to CD2Cl2 i

possibilities

rongly enou

interaction

by the elect

hreiner’s an

N-H pKa:s

symmetric

n between

obsen-List te,65

p. 1H NMR

eld shift of 0

protons wa

rved for the

te H-bond

p. However,

tent of 51, w

CH2- and O% 12; Gree

internal sign

now arise i

ugh to the n

ns. In comp

tron withdr

nd Cheng’s

are around

thiourea 12

with pKa

type catalys

titrations of

0.03 ppm fo

as observed

e thiourea p

coordinatio

there is no

which rema

OMe-groupen: ~20 molnal before o

in our case:

nucleophile

parison, the

rawing arom

s studies.17,

d 20 in DM

2 has an inc

value at 1

t 89 in thiou

f 51 in the p

for the CH2-

d (Figure 7)

protons orth

on is impo

o observable

ains practica

p in the prel % 12, Pur

overlaying.

: either the

e or the bul

e H-bond d

matic N-sub,22b,23 Acco

MSO when

credibly high

12.4.64 The

urea acidity

43

presence

-protons,

). At the

ho to the

ortant to

e change

ally fully

sence of rple: ~30

thiourea

lky trityl

donating

bstituent,

rding to

both N-

h acidity

N-trityl

y, having

Page 46: Organocatalysis and Continuous Flow Reactors as Enabling Tools

44

Figure 8. Thiourea N-H acidities of common catalysts according to Schreiner

To test this, we first synthesized catalyst 92 from epi-9-amino-(9-deoxy)quinidine 90 and

cyclohexyl isothiocyanate 91. Compound 90 was synthesized according to literature

procedures via the Mitsunobu-Staudinger protocol.31,49 In the model reaction between 26a and

51 catalyzed by 92, a yield of 73% and an ee of 73% was achieved (Scheme 24). Hence, the

catalyst retained its activity even though the acidity of the thiourea N-H groups was severely

perturbed. This would suggest that the steric interactions between the N-trityl moiety of 79

and dimethyl malonate suppress the required nucleophile enolization step (Figure 9). It is

likewise conceivable that the bulky trityl group forces the thiourea N-H bonds to be in

unfavorable cis/trans conformation with respect to the central thiocarbonyl moiety instead of

the trans/trans conformation usually presented in drawn schemes. At the time of our studies,

less than 20% published crystal structures of thiourea compounds show the trans/trans

conformation.66 An example of cis/trans thiourea conformation within a catalyst was

observed by X-ray crystallography during our studies towards the synthesis of (S)-pregabalin:

a catalyst otherwise analogous to 79 but carrying a N-9-phenylfluorenyl substituent (in

principle a rigidified trityl) adopted said conformation in crystal structure (Figure 10).49,66

Page 47: Organocatalysis and Continuous Flow Reactors as Enabling Tools

45

N H

HN

HN SN

O

92

+O O 92 (10 mol %)

CH2Cl2, rt, 24 hPh

NO2

O O

MeO OMeMeO OMe

73%, ee 73%

51 52a

PhNO2

26a

N H

H2NN

O

+

90

NCS

THF, rt

48%91

Scheme 24. Synthesis of catalyst 92 and its performance in the model reaction

Figure 9. Plausible model of malonate deprotonation by thiourea catalysts and possible steric interactions between N-trityl substituent and malonate nucleophiles with catalyst 79

Page 48: Organocatalysis and Continuous Flow Reactors as Enabling Tools

46

Figure 10. X-ray structure of a 9-phenylfluorenyl substituted thiourea catalyst (from Ref. 49)

A sulfonamide derivative 94 based on the same cinchona scaffold was reported by Lu et al.

during the course of our studies.67 This catalyst was active in the Michael additions of mono-

and bicyclic �-ketoesters into nitroalkenes, indicating that a bidentate H-bond donor is not

always a crucial feature of the catalyst structure in these reactions (Scheme 25). We decided

to investigate further whether a bidentate hydrogen bonding is essential in our case. To this

end we synthesized catalysts 96 and 97, in which one of the thiourea N-H groups is in turn

replaced by an N-Me group (Figure 11).

Scheme 25. Monodentate cinchona derived H-bond catalyst by Lu

Page 49: Organocatalysis and Continuous Flow Reactors as Enabling Tools

47

Figure 11. Catalysts 96 and 97 with a perturbed thiourea moiety

Catalyst 96 was synthesized as shown in Scheme 26. Monomethylation into 99 was carried

out by carbamoylation with EtOCOCl and subsequent LiAlH4-reduction of 98, and the

desired product was isolated after SiO2 chromatography. The standard thiourea formation was

then accomplished by reacting 99 with 3,5-bis(trifluoromethyl)phenyl isothiocyanate 100 to

give 96.

Scheme 26. Synthesis of monomethylated thiourea 96

The counterpart catalyst 97 was achieved by reversing the scaffolds of the isocyanate and the

amine.49 First, 90 was treated with CS2 in the presence of DCC to give the isothiocyanate 101

Page 50: Organocatalysis and Continuous Flow Reactors as Enabling Tools

48

in good yield (Scheme 27). Compound 101 and its other cinchona relatives are versatile

synthons in themselves, and can be used to synthesize a variety of different

organocatalysts.49,68 It was our hope that we could add a methylated aniline into 101 to give

catalyst 97 in a straightforward manner. Monomethylation of 3,5-bis(trifluoromethyl)aniline

102 was attempted by an analogous carbamoylation-LiAlH4-reduction route, but partial

defluorination of carbamate 103 was observed after the reduction step (Scheme 28).69

N

H2NN

OMe

90

CS2, DCC

THF, –15 °C to rt

N

NN

OMe

101

CS

74%

Scheme 27. Synthesis of isothiocyanate 101

Scheme 28. Attempted synthesis of monomethylated aniline 104

Eventually, 104 was synthesized via a three step reaction sequence.70 Aniline 102 was first

acylated with TFAA in CH2Cl2, then treated with K2CO3 and MeI in acetone, and finally

hydrolyzed with K2CO3 in a 5:1 mixture of MeOH:H2O. The crude product was obtained in

90% yield over three steps with acceptable purity, and was used as such (Scheme 29).

Page 51: Organocatalysis and Continuous Flow Reactors as Enabling Tools

49

CF3

F3C NH2

102

TFAACH2Cl2, 0 °C to rt

CF3

F3CHN

105

CF3

OMeI, K2CO3

Acetone, rfx.

CF3

F3C N

106

CF3

O

Me

K2CO3

5:1 MeOH:H2OCF3

F3CHN

104

Me90% over 3 steps, used without purification

Scheme 29. Synthesis of monomethylated aniline 104

Whereas the thiourea formation usually proceeds at ambient temperature without external

catalysts, direct coupling between 101 and 104 to give 97 was in this case unsuccessful, most

likely due to a combination of low nucleophilicity of the highly electron poor aniline and

steric shielding of the isothiocyanate 101. A copper(II)-mediated coupling was attempted but

with no success. Deprotonation of 104 with NaH both with and without the presence of 15-

crown-5 was equally unsuccessful. In both cases, most of the aniline was recovered, but

decomposition of the isothiocyanate was observed. A crossed thiophosgene coupling between

101 and 104 was likewise unsuccessful. The coupling was eventually achieved by

deprotonating 104 with n-BuLi at –78 °C, after which a solution of 101 in THF:DMF was

added to the mixture, which was then gradually warmed to room temperature (Scheme 30).

The reaction gave two products of identical mass which were highly similar according to 1H

NMR spectra. The correct product was assigned by thorough comparison with literature

data.30b

Scheme 30. Synthesis of monomethylated thiourea 97

Page 52: Organocatalysis and Continuous Flow Reactors as Enabling Tools

50

Both catalysts 96 and 97 were nearly completely unreactive in the model reaction, giving little

product with poor enantioselectivity of 13% (Scheme 31). Hence, the bidentate thiourea

moiety is an essential feature of the catalyst structure, and its perturbation leads to loss of

reactivity. The methylation of the anilinic nitrogen also reversed the selectivity, giving the

opposite enantiomer than previously observed.

Scheme 31. Addition of dimethyl malonate 51 into nitroalkene 26a catalyzed by 96 or 97

The inactivity of 96 and 97 supports the proposition that bidentate coordination between 51

and thiourea is necessary for the deprotonation of the nucleophile. To further investigate the

nucleophile deprotonation in the reaction, 1H NMR titration of 51 was performed in the

presence of 7 to 40 mol % of catalyst 88 in CDCl3. In contrast with the Schreiner thiourea 12,

no expected downfield shift was observed for the methylene protons in the presence of 88.

When the 1H NMR spectrum of 51 was recorded in the presence of D2O alone (in MeCN-d3),

there was no significant shift or peak deformation. After addition of Et3N, the CH2 peak was

shifted upfield by 0.22 ppm, indicating a Lewis-basic solvation behavior between Et3N and 51.

No immediate deprotonation was observed, which correlates with the pKa differences between

51 and Et3N. The deprotonation of dimethyl malonate in the presence of the Schreiner

thiourea 12 is very quick. In the presence of 35 mol % of 12 and D2O, the CH2-peak of

dimethyl malonate disappears completely and immediately after the addition of Et3N.

In subsequent experiments we loaded a solution of 51 in MeCN-d3 to an NMR tube in the

presence of 20 mol % of 88. After addition of 0.1 mL of deuterium oxide and quick vigorous

shaking, the first 1H NMR measurement showed the disappearance of the sharp methylene

CH2 peak at 3.38 ppm, and the appearance of a broad singlet at 3.28 ppm. The integrals

matched roughly the intensity relationship 3:1 with the methyl ester peak. The experiment

was repeated in acetone-d6 using MeOD-d4 as the deuterium source. This time we observed a

gradual disappearance of the CH2 signal and the formation of a small amount of

Page 53: Organocatalysis and Continuous Flow Reactors as Enabling Tools

monode

12).

Figure MeOD-50 minpresenta

The exp

observe

40% of

trityl su

moiety

Figure MeOD-addition

euterated 51

12. 1H NM-d4. From bon after addation.

periment wa

ed in 20 min

monodeute

ubstituent ef

of catalyst 7

13. 1H NM-d4. From bn of MeOD-

1, identified

MR spectraottom to top

dition of M

as repeated

nutes. After

erated 51 pr

ffectively b

79, and sub

MR spectraottom to to-d4. Spectra

d by the trip

a of dimethp: 51 and 8

MeOD-d4. S

with 51 in

12 hours, r

resent in the

blocks malo

sequently th

a of dimethop: 51 and 7a have been

plet signal a

hyl malonat88 without MSpectra hav

n the presen

roughly 60%

e mixture (F

onate nucleo

he catalytic

hyl malonat79 without offset manu

t 3.41 ppm

te 51 in theMeOD-d4; 2ve been of

nce of cataly

% of 51 had

Figure 13). H

ophiles from

cycle is als

te 51 in theMeOD-d4; ually for cla

next to the

e presence 2 min; 5 miffset manu

yst 79, but

been depro

Hence we c

m coordinat

so disrupted

e presence 5 min; 20 m

arity of pres

CH2 signal

of catalystin; 10 min;

ually for cl

no deuterat

otonated, wi

conclude tha

ting to the

d.

of catalystmin; 12 hosentation.

51

l (Figure

t 88 and 20 min;

larity of

tion was

ith about

at the N-

thiourea

t 79 and urs after

Page 54: Organocatalysis and Continuous Flow Reactors as Enabling Tools

52

2.5 Conclusions

We have explored the properties of the N-trityl substituted thiourea catalyst 79 and its

applicability to nitro-Michael reaction. The catalyst performs best in the original reaction to

which it was optimized, giving otherwise mediocre enantioselectivities in the Michael

addition of Meldrum’s acid into aromatic and heteroaromatic nitroalkenes. It is noteworthy

that the catalyst gives better enantioselectivities for alkyl substituted than for aryl substituted

nitroalkenes, whereas the reverse trend is commonly observed in method development.

Catalyst 79 also suffers from substrate limitations. It was unable to catalyze the reaction

between dimethyl malonate and aromatic nitroalkenes, possibly due to steric hindrance caused

by the bulky N-trityl substituent. Malonate nucleophiles are ostensibly not able to effectively

coordinate to the thiourea moiety for assisted methylene proton acidification and

deprotonation by the tertiary amine. NMR experiments confirm that the catalyst 79 is unable

to deprotonate malonate like nucleophiles. This was also investigated with thiourea N-

methylated analogues 96 and 97 of Chen-Soós-Connon-Dixon type catalyst. When the

bidentate thiourea moiety was perturbed, the activity and selectivity of both catalysts in the

model reaction was significantly poorer than the original catalyst. The activity of 79 with

Meldrum’s acid would then seem to be mediated through a chiral acid-base ion pair formation.

Whether the thiourea moiety in 79 directs the nitroalkene via two explicit hydrogen bonds

remains unclear, since the bulky trityl substituent may force the N-Hs to adopt a cis/trans-

conformation with respect to central thiocarbonyl group.

Page 55: Organocatalysis and Continuous Flow Reactors as Enabling Tools

53

3 STUDIES TOWARDS THE TOTAL SYNTHESIS OF (–)-

HENNOXAZOLE A

Hennoxazoles (Figure 14) constitute a family of biologically active marine alkaloids isolated

from Polyfibrospongia sp. in the early 1990s.71 Of these, (–)-hennoxazole A (107a) has

shown the most potent antiviral activity against herpes simplex 1 virus with an IC50 of 0.6

�g/mL. The distinctive bisoxazole core combined with a stereochemically elaborate

tetrahydropyran ring and a skipped triene containing a single distant stereocenter at C22 make

107 an interesting target for total synthesis. A similar directly joined bisoxazole moiety has

thus far been identified only in muscorides and diazonamides. To date, four different total

syntheses for (–)-hennoxazole A have been reported, and these are briefly reviewed in section

3.3.1.72

Figure 14. The hennoxazole family of marine natural products. Only the stereochemistry of (–)-hennoxazole A has been confirmed to be that shown here.

Page 56: Organocatalysis and Continuous Flow Reactors as Enabling Tools

54

During the last decade, the Ley group has been developing laboratory scale continuous-flow

reactors and technologies and their application to organic synthesis.73 Their goal was to create

a total synthesis of (–)-hennoxazole A in which all or most of the steps would be carried out

in continuous-flow conditions. An overview of continuous-flow technologies, benefits and

disadvantages is given in below in Section 3.1.

3.1 Continuous Flow Technologies

During the last decade, the application of micro- and mesofluidic continuous flow conditions

in laboratory scale has been developed into a powerful enabling technology for synthetic

chemists.74 Several research groups both in the academic world and in the industry are

currently in the process of exploring the limitations and possibilities of continuous flow

conditions in organic synthesis. Whereas a “do it yourself” mentality has served well in the

early development of continuous flow methodologies, the 2000s have seen the launch of

commercial laboratory scale equipment that can be used in a highly modular fashion.75 These

technologies range from microfluidic chips to fully integrated reactor systems which are

usually based on commercial HPLC equipment, with specific modifications to answer specific

needs. A standard example is the ThalesNano H-Cube® flow hydrogenator, in which a

solution of the substrate is pumped through a cartridge filled with an immobilized

hydrogenation catalyst, e.g. Pd/C.76 Hydrogen is produced on demand in an integrated

electrolytic cell and mixed into the substrate stream before the cartridge, thus circumventing

the problem of storing and handling hazardous gases (although at a nontrivial energy cost).

Below we shall consider the benefits and disadvantages of carrying out organic synthesis in

continuous flow conditions using microreactors.

3.1.1 Benefits and Disadvantages of Continuous Flow Conditions

The mechanism of a reaction obviously does not change when transferred from batch to flow

systems. However, the differences in mass and heat transfer properties of the reactor system

may have a profound effect on the rate and selectivity of the reaction in question. Some recent

reviews shed light on these issues in a highly lucid fashion.74e,f Overall, one of the greatest

advantages of continuous flow conditions is the precise control over reaction parameters due

to the enhanced mass and heat transfer properties.

Page 57: Organocatalysis and Continuous Flow Reactors as Enabling Tools

55

One of the most important issues of carrying out a chemical reaction is the proper and

efficient mixing of reagents. In microreactors (reactor channel diameter 10-500 �m), laminar

flow conditions usually exist (Re <2000), and mixing of reagents takes place via diffusion. It

has been calculated that for a molecule of typical diffusivity, it takes 30 seconds to achieve

complete mixing in a reaction channel of 400 �m i.d. With specifically fabricated

micromixers, this mixing time can be reduced to the order of 10 milliseconds. This is highly

beneficial for extremely fast reactions for which the rates are controlled by mass transfer. For

reactions in which substrates may react further than to the desired product, fast mass transfer

usually results in better selectivities.77

The volume-area relationship of continuous-flow reactors is usually very high, on the order of

5000-50000 m3/m2, which makes heat transfer into and from the reaction system very

efficient, depending on the thermal conductivity of the reactor material. This is highly

beneficial for reactions that release powerful exotherms, such as lithiation reactions. The

group of Yoshida has made significant contributions in this field, carrying out highly

exothermic and fast lithium-halogen exchange reactions with residence times of less than a

second in a cooled glass chip reactor.78 Even though the released energy per reactant mole is

impressive, in a microfluidic environment there are only small amounts of reacting materials

at a given moment, and hence the exotherm may be compensated by the high thermal

conductivity.

The contained environment and relative ease of pressurizing a continuous flow reactor

enables the superheating of reaction solvents above their boiling point in atmospheric pressure

in a safe manner.74g,79 Thus reactions that are accelerated by high pressures (reactions with

negative volume of activation/volume change) and temperatures, such as Diels-Alder

reactions, may be beneficially transferred to continuous flow conditions.80 Unstable,

hazardous and noxious reagents and intermediates are likewise contained in the reactor

system, and thus carrying out certain reactions may be significantly safer.81

The most obvious problem of microfluidic technologies lies in the intrinsic micro aspect.

Reaction channels that are usually less than 0.5 mm in diameter are highly prone to clogging

by salts and products crystallizing out from a reaction mixture.82 At worst this can lead to

powerful pressure buildup in the system and subsequent rupturing of the reactor vessel, coil or

chip, or other damage to the pumps and valves. When dealing with the issue of precipitation

and clogging, the reaction conditions in question usually need to be modified to suit the

microfluidic system. Lowering the concentration and changing the solvent system are obvious

Page 58: Organocatalysis and Continuous Flow Reactors as Enabling Tools

56

methods, which however may lead to suboptimal performance of the reaction and increased

amounts of solvent waste.

When continuous flow conditions are applied to multistep synthesis, reaction monitoring and

coordination of further steps may become problematic (so-called “third stream problem”). In

tubular reactors where the fluid flow is laminar there is always some diversion from an ideal

plug-flow, which leads to axial dispersion and a bell curve shaped reactant concentration with

respect to time and reactor length.83 The injection of subsequent reactant streams has to be

coordinated so that there is minimum loss of substrates. In a continuous process this is an

issue only in the beginning, before reaching an equilibrium state. If a multistep process is

performed in a plug flow manner (e.g. in discovery chemistry for pharmaceuticals), there is a

need for robust in line monitoring and automation so that the injection of further reactant

streams may be triggered when a desirable concentration of product is observed.

3.1.2 General Strategies for Continuous Flow Synthesis

Transferring a reaction from batch to continuous flow conditions often requires a change in

the overall philosophy of chemistry.84 As will also be shown in Chapter 4, homogeneity of the

reaction becomes a major consideration and affects directly reactant concentration in solution.

In line quenching and purification of the reaction mixture is another important consideration.

Solid supported reagents, such as acidic and basic resins have been widely used for both

purposes. However, solid supported reagents suffer from the obvious drawback that they have

a specific loading capacity, and for large scale synthesis, which in flow conditions translates

to long operating time, such reagents should periodically be either replaced with new material

or regenerated with a washing sequence.

Catalysis in continuous flow conditions can be approached with several strategies.

Homogeneous catalysis may be carried out in continuous flow conditions in an analogous

manner to flask chemistry just by injecting the catalyst into the reaction stream.85 Product

isolation and catalyst separation then become the crucial issues. These may be solved by

scavenging reagents, catch and release techniques (of either catalyst or product), or by

incorporating a reactive tag in the catalyst structure which enables further operations.

Immobilizing a catalyst on a solid support is one of the ideal solutions, but it is usually

associated with both benefits and disadvantages.86 Immobilization forces reactions to take

place on the catalyst surface, which usually has a detrimental effect on reaction kinetics.

Another adverse kinetic effect may be caused by the improper porosity of the catalyst, since

Page 59: Organocatalysis and Continuous Flow Reactors as Enabling Tools

57

the diffusion time for substrates to reach a catalytic site within the support matrix is

lengthened. If the porosity is suboptimal, some of the catalyst might be completely occluded

and cannot come into contact with the substrates. In the case of asymmetric catalysis, the

steric effects caused by the structure of the solid support may also deteriorate catalyst

selectivity. Further issues are e.g. the mechanical, thermal and chemical stability of the

support and its swelling properties in different solvents. Most of these issues depend on the

synthesis of the solid support (degree of cross linking in a polymer, polymer beads vs.

monolithic column, sol-gel synthesis of SiO2 with controlled pore size, etc.) and the method

of catalyst immobilization (co-polymerization, grafting, etc.). This is a science in itself and

beyond the scope of this thesis.

The highly automated synthesis of racemic oxomaritidine illustrates several key issues and

strategies in flow chemistry (Scheme 32).87 A solution of bromide 110 in 1:1 CH3CN:THF

was passed through an ion exchange resin to give the azide 111, which was subsequently

immobilized as the corresponding iminophosphorane on a butylphosphine functionalized

resin. Aldehyde 113, obtained by oxidation of the corresponding alcohol by passing a THF

solution through a column packed with immobilized tetra-N-alkylammonium perruthenate,

was pumped with 111 through a phosphine column to yield the aza-Wittig product 114. The

imine was then reduced to the amine 115 with a flow hydrogenator equipped with a 10% Pd/C

catalyst cartridge. After a manual solvent exchange from THF to CH2Cl2, the amine was

trifluoroacetylated in a chip reactor, and then passed through a column loaded with a PIFA-

functionalized resin to effect the oxidative phenolic coupling to give the intermediate 117.

Treatment of 117 with a strongly basic resin in the presence of a 4:1 MeOH:H2O side stream

hydrolyzed the amide, which then underwent a spontaneous conjugate addition to give

racemic oxomaritidine 118 in an overall yield of 40%. Most impurities were scavenged by the

basic resin, and the natural product was obtained with excellent purity.

Page 60: Organocatalysis and Continuous Flow Reactors as Enabling Tools

58

Scheme 32. Continuous flow synthesis of racemic oxomaritidine

The time taken by the synthesis was 8 hours, which is significantly faster than performing the

same synthesis with traditional methods, albeit the scale of the flow synthesis is rather small.

Planning and optimizing each step of a multistep synthesis like oxomaritidine obviously is

quite time consuming. However, providing that a great number of known robust

transformations can be carried out in flow conditions, chemists should be able to perform

multistep flow syntheses routinely in the near future. The synthesis of oxomaritidine also

illustrates one of the greatest benefits of continuous flow systems. With each step telescoped

into the next one, the number of required unit operations is reduced significantly, since there

are no extraction or chromatography steps included, with only two evaporation steps. This

equates directly to reducing waste and saving consumables such as solvents and SiO2, but

even more importantly, time and manpower.

Page 61: Organocatalysis and Continuous Flow Reactors as Enabling Tools

59

3.2 Covalent Organocatalysis

As stated in section 2.1, the other major category in organocatalysis is usually called covalent

catalysis.9b For carbonyl containing substrates this includes e.g. the enamine and iminium

manifolds, which are usually accessed by using chiral primary and secondary amines as

catalysts (Figure 15). Catalysis achieved by structures such as 119-121 is also commonly

called aminocatalysis. Reactions carried out this way include e.g. aldol reactions, Mannich

reactions, conjugate additions, and Diels-Alder reactions among other cyclizations. Other

examples of covalent organocatalysis include the use of tertiary amines or

phosphines/phosphites such as 123 as nucleophilic catalysts, such as in Morita-Baylis-

Hillman reactions. Umpolung type reactions of aldehydes such as Stetter reaction or benzoin

condensation are usually carried out with thiazolium catalysts or other nitrogen containing

heterocycles such as the precatalyst 122 that may form a stable ylide (or carbene).

Figure 15. Representative examples of catalysts in covalent organocatalysis

3.2.1 Enamine Catalysis

Aliphatic aldehydes and ketones are able to form enamines in the presence of primary and

secondary amines under dehydrating conditions. Enamine formation enhances the

nucleophilicity of the parent carbonyl �-carbon, thus making enamines highly useful specific

enol equivalents in organic synthesis.10b Being unstable species and prone to rapid hydrolysis

in aqueous environment, most enamines are difficult to isolate and separate from reaction

mixtures. However, the catalytic formation of enamines in the presence of an amine has

enabled a wide variety of reactions to be carried out under mild organocatalytic conditions.10a

When a chiral amine is used as the catalyst, the resulting enamine carries the chiral

information of the catalyst, and a stereoselective reaction with an electrophile ensues. The

Page 62: Organocatalysis and Continuous Flow Reactors as Enabling Tools

60

topic of enamine catalysis has been widely reviewed recently,9b,c,10 and herein we shall only

consider some crucial theoretical aspects and examples.

The formation of an enamine raises the HOMO energy of the enolate C=C bond and thus

enhances the �-carbon nucleophilicity of a ketone or an aldehyde in comparison to the bare

enol form. The classical enamine catalysis cycle is presented in Scheme 33. The origin of

enantioselectivity in the reactions of e.g. proline has been explained by the Houk-List model

which accounts for intermolecular hydrogen bonding/Brønsted acid-base interactions to take

place between the catalyst-enamine species and the electrophile (Figure 16, I).88 The

minimum energy transition state has the Re-face of the E-enamine attacking the aldehyde

from the Re-face, with coordination assisted by H-bonding, leading to the usually observed

anti selectivity. For catalysts without a directing H-bond donor or Brønsted acid, such as the

Hayashi-Jørgensen type diarylprolinol silyl ethers,89 the bulky substituent blocks the other

side of the enamine, directing the attack to take place from the unhindered side (Figure 16,

II).90

Scheme 33. Enamine catalytic cycle shown for aldol reactions

Page 63: Organocatalysis and Continuous Flow Reactors as Enabling Tools

61

Figure 16. Origin of enantioselectivity in enamine catalysis

3.2.2 Asymmetric Organocatalytic Aldol Reactions

The importance of aldol reactions to organic synthesis cannot be overemphasized, and since

its discovery in the 1800s it has been developed into a versatile tool for C-C bond formation.91

In Nature, the aldol reaction is mediated by class I and II aldolase enzymes. Class I aldolases

contain a lysine residue in the active site, and they mediate their effect via iminium/enamine

formation.92 Class II aldolases incorporate a bivalent metallic Lewis acid cofactor, usually

Zn2+, which assists in the deprotonation of the nucleophile (e.g. dihydroxyacetone

phosphate).93 The studies into catalytically active aldolase antibodies carried out by Barbas

and Lerner in the late 1990s were one of the key factors leading to the discovery of proline-

catalyzed intermolecular direct aldol reactions.94

The first asymmetric organocatalytic aldol reaction was discovered in 1971 by two industrial

research groups independently. The intramolecular aldol reaction of 124a catalyzed by (S)-

proline 125 is known as the Hajos-Parrish reaction (Scheme 34).95 The Hajos-Parrish protocol

proceeds with good yields and enantioselectivities to give the bicyclic aldol addition product

126a, which is then dehydrated to give the versatile synthetic intermediate 127a. Using 124b

as a starting material gives the nonracemic version of the Wieland-Miescher ketone 127b,

albeit in slightly lower enantioselectivity. The concurrently published protocol used by Eder,

Sauer and Wiechert includes a strong acid component, which gives the dehydration product

directly.

Page 64: Organocatalysis and Continuous Flow Reactors as Enabling Tools

62

Scheme 34. Hajos-Parrish reaction

The first proline catalyzed asymmetric intermolecular organocatalytic aldol reaction was

reported by Barbas, List and Lerner in 2000.8 The aldol addition of acetone to 4-

nitrobenzaldehyde 128 in DMSO, catalyzed by 30 mol % (S)-proline gives the product 129 in

68% yield and 76% ee (Scheme 35). Out of simple easily available proline analogues, 4-

trans-hydroxyproline 133 gives a significantly better yield in this model reaction,8 but only

5,5-dimethyl thiazolidine-4-carboxylic acid 134 was able to clearly outperform proline in

enantioselectivity.96 However, the studies by Barbas revealed that proline gives consistently

better yields and is applicable to a wider range of substrates.96

Page 65: Organocatalysis and Continuous Flow Reactors as Enabling Tools

63

Scheme 35. Proline-catalyzed aldol reactions by Barbas & List

One of the greatest challenges of organocatalytic aldol reactions is the control of selectivity in

the intermolecular crossed aldol reaction between two reactive aliphatic aldehyde species.

Unsubstituted aliphatic aldehydes perform rather poorly as acceptors, since they are likewise

apt to form the catalytic enamine species with the catalyst. The situation becomes even more

challenging when ketones are used as the nucleophilic species due to faster enamine

formation between aldehydes and amines. Due to this, most methods development papers

concentrate on using reliable and reactive acceptors such as electron poor aromatic aldehydes.

The first crossed organocatalytic aldol reactions of aliphatic aldehydes were reported by

MacMillan in 2002.97 In their procedure, the donor aldehyde species was slowly added to the

reaction mixture with a syringe pump to suppress self-condensation. As an example,

propionaldehyde 136 and isobutyraldehyde 137 gave the addition product 138 in good yield

and excellent diastereoselectivity and enantioselectivity (Scheme 36).

Scheme 36. Crossed aldol reaction between aliphatic aldehydes by MacMillan

Page 66: Organocatalysis and Continuous Flow Reactors as Enabling Tools

64

The evolution of catalysts for organocatalytic aldol reactions has been rapid ever since the

first publications by List and MacMillan, and the developments have been recently

reviewed.98 Application of these methods to more challenging syntheses and targets has,

however, been modest. Therefore some target oriented examples of organocatalytic aldol

reactions from the literature are highlighted below.

As a classic example, the Hajos-Parrish-Eder-Sauer-Wiechert protocol was applied to the total

synthesis of erythromycin by Woodward et al. (Scheme 37).99 The racemic precursor 139 was

subjected to (R)-proline, which gave the aldol products 140 and 141 in good yield but a

meager ee of 36%. Being an early step in the synthesis, this could be done on large scale, and

the material was recrystallized until optically pure. All stereochemical information of

erythromycin was derived from this step, the rest of the synthesis containing mostly internal

asymmetric induction and diastereoselective methods.

Scheme 37. Proline catalysis in the total synthesis of erythromycin

One of the earliest examples of proline catalyzed crossed aldol reaction in natural product

synthesis is the reaction between propionaldehyde and isobutyraldehyde as reported by Pihko

in 2003.100 The reaction proceeded to give the TBS protected aldol product 142 in 61% yield

over two steps (one pot aldol reaction and TBS protection) in an impressive ee of over 99%

and 40:1 anti:syn diastereoselectivity (Scheme 38). A further Mukaiyama aldol reaction and

lactonization gave the natural product (–)-prelactone B in three operations.

Page 67: Organocatalysis and Continuous Flow Reactors as Enabling Tools

65

Scheme 38. Synthesis of (–)-prelactone B 145

In 2006 Wang et al. published a total synthesis of trichostatin A 147, in which 4-

nitrobenzaldehyde 128 was surprisingly used in a target-oriented setting (Scheme 39).101

Proline was used according to the MacMillan protocol to catalyze the reaction between

propionaldehyde and 128 in good yield over two steps and excellent selectivities (16:1

diastereomeric ratio, >99% ee). The aldol product was immediately protected as the dimethyl

acetal 146 due to the instability of the aldol intermediate, which was prone to decompose on

SiO2. A further 7 steps gave optically pure 147 with a 17% overall yield.

Scheme 39. Organocatalytic aldol reaction in the synthesis of trichostatin A 147

Page 68: Organocatalysis and Continuous Flow Reactors as Enabling Tools

66

An aldol/oxa-Michael reaction cascade was used in a short total synthesis of �-tocopherol by

Woggon and co-workers.102 Diarylprolinol 150 catalyzed the cyclization between phytenal

149 and the highly substituted salicylaldehyde 148 to give the lactol 151 in 97%

diastereomeric excess (Scheme 40). The reaction proceeds via a dienamine pathway, the

phytenal skeleton first adding to 148, and a subsequent ring closure through 1,4-addition of

the phenolic group to the conjugated iminium intermediate. Completion of the synthesis gave

the C2-epimeric �-tocopherol. Synthesis of natural �-tocopherol was achieved by using the R-

enantiomer of the diarylprolinol catalyst. Other chromane and chromanone derivatives have

also been accessed with this method.103

Scheme 40. Aldol/oxa-Michael cascade in the synthesis of �-tocopherol

The MacMillan group used both (S)- and (R)-proline in their total synthesis of Callipeltoside

C (Scheme 41).104 The first step in the tetrahydropyran fragment synthesis was the aldol

reaction between propionaldehyde and Roche aldehyde 152. In spite of poor yield (48%), the

selectivities were once more admirably high (12:1 d.r., 99% ee). A second aldol reaction was

used in the construction of the sugar fragment. TIPS-protected �-hydroxyacetaldehyde 154

was dimerized using (R)-proline in good yield and 99% ee in the synthesis of the mannose

fragment. The NMR spectroscopic data of the finished product did not match with the natural

product. To correct this, the enantiomeric mannose was synthesized by using (S)-proline in

the first step with identical yield and enantioselectivity.

Page 69: Organocatalysis and Continuous Flow Reactors as Enabling Tools

67

PMBO H

O

+

152

H

O

136

PMBO

OH

H

O

153

NH

CO2H

(S)-125 (10 mol %)

DMSO, 4 °CSlow addition (100 h)

48%, ee 99%anti:syn 12:1

H

OOTIPS

154

(R)- or (S)-125H

O

OTIPS

OHOTIPSorH

O

OTIPS

OHOTIPS

155 ent-155

O O NH

CCl3OMeHO

TIPSOor

O O NH

CCl3OMeHO

TIPSO

156 ent-156

DMF, 75%, ee 99%anti:syn not given

Scheme 41. Proline catalyzed aldol reactions in the synthesis of Callipeltoside C

A rare and early example of ketone addition to an �-unsubstituted aldehyde was presented in

the synthesis of (+)-ipsenol 160 by List.105 The addition of acetone to isovaleraldehyde 157

gave enone 158 as the main product via a formal aldol condensation in 42% yield, and the

aldol addition product 159 in 34% yield and 73% ee (Scheme 42). The poor yield and

enantioselectivity may be justified by considering that the reaction is the first step of a

synthesis, and cheap starting materials and catalysts are used. Overall, a substrate screen

carried out by List revealed that the condensation product tends to become the main reaction

product in proline catalyzed ketone aldol additions to unsubstituted aldehydes. Furthermore,

they postulated that the mechanism for the enone formation proceeds through a Mannich-type

addition of the ketone into the iminium species formed from the aldehyde and proline.

Page 70: Organocatalysis and Continuous Flow Reactors as Enabling Tools

68

Scheme 42. Addition of acetone into an unsubstituted aldehyde in the synthesis of (+)-ipsenol

160

In conclusion, the application of organocatalytic aldol reactions into large scale synthesis and

target oriented synthesis is still rather underdeveloped. In most examples, the organocatalytic

step is placed in the early phases of the synthesis, and constricted to simple starting materials.

Taking into consideration the easy availability and low cost of e.g. proline, this approach is

perfectly justified. However, as the current state of the art catalysts require some synthetic

effort and often expensive enantiopure materials, the benefits may quickly be balanced out by

the cost and lost time.

3.2.3 Asymmetric Organocatalysis in Continuous Flow Conditions

Organocatalytic reactions in continuous-flow conditions are still something of a rarity, but

some interesting examples have already been reported.106 More research has been done with

immobilized ligand-metal complexes, e.g. Co(salen)-complexes have been used in dynamic

kinetic resolution107 and Ru-complexes in e.g. cyclopropanation reactions.108 The alkylation

of aryl aldehydes with diethyl zinc has been reported to proceed with good

enantioselectivities when immobilized chiral amino alcohols are used as ligands.109

Enantioselective Michael-addition of malonate nucleophiles to nitroalkenes has been achieved

with an immobilized CaCl2-PyBox-complex.110 Immobilized enzymes have also been used to

carry out e.g. kinetic resolution111 and oxidative dimerization-intramolecular cyclization in a

natural product synthesis setting.112 Asymmetric hydrogenation protocols have also been

investigated recently.113

Page 71: Organocatalysis and Continuous Flow Reactors as Enabling Tools

69

An interesting early example of organocatalysis comes from the group of Lectka.114 They

synthesized �-lactams by gravity assisted flow of substrates through sequential packed

columns (Scheme 43). Acid chloride 161 was used as a ketene precursor, and let flow through

a cooled column packed with PS-BEMP 162 to form the ketene 163 in situ. Imine 164 was

then added in a side stream, and the mixture was let flow through a column packed with

immobilized quinine ester 165 to form the �-lactam via a [2+2] ketene cycloaddition. A third

column was packed with a nucleophilic primary amine resin to scavenge excess reagents and

byproducts. The lactam 167 was obtained in 65% yield and 93% ee after evaporation of

solvents, with a cis/trans ratio of 10:1.

N NMePN NEt2

t-Bu

PS-BEMP 162

PhCl

OPh C

O

–78 °C –40 °C

PS-O-acylquinine 165

EtO2C H

NTs

164

NH2

PS-Benzylamine

NTs O

PhEtO2C

167 65%, 10:1 dree 93%

O

O

O

O

N

N

OMe

161 163

166

Scheme 43. Lectka’s synthesis of �-lactams

The immobilized quinine 165 has also been used in an asymmetric �-chlorination of acid

chlorides by the Lectka group.115 The reaction was a part of a convergent multi-step gravity

driven flow through synthesis of the metalloproteinase inhibitor BMS-275291 (Scheme 44).

Chlorinating agent 168 and acid chloride 169 were fed through a column packed with 165,

and then through an immobilized piperazine-packed column to remove excess acid chloride.

Simultaneously through the other channel was fed a mixture of peptide precursors 172 and

Page 72: Organocatalysis and Continuous Flow Reactors as Enabling Tools

70

173, which were coupled with an immobilized carbodiimide 174 to give the amide 175.

Fmoc-deprotection was then carried out in a column packed with tris-(2-aminoethyl)amine

resin 176. The streams were connected at this point to carry out the amide coupling between

activated acid 171 and deprotected peptide 177. The final step was an SN2 substitution of the

�-chloride with a hydrogen sulfide anion, which was preloaded on a quaternary ammonium

resin. The final product 179 was obtained in 34% overall yield and 83% diastereomeric

excess.

PS-Quinine 165

MeN N

O

O

Cl

O

OCl

ClCl

Cl

ClCl

+

168 169

MeN N

O

O

O

O

Cl ClCl

ClCl

Cl

NNH

PS-Piperazine 170

171

FmocHN

s-BuOH

O

H2N+t-Bu

NHMe

O

172 173

PS-Carbodiimide 174

N C N HN

FmocHN

s-Bu HN

O

175

t-BuNHMe

O PS-Amine 176H2N

s-Bu HN

O

177

t-BuNHMe

O

N

NH2

NH2

171

177

Celite NH

s-Bu HN

O

178

t-BuNHMe

OOCl

N

MeN

OO

NMe3SH

NH

s-Bu HN

O

BMS-275291 179

t-BuNHMe

OOHS

N

MeN

OO

Scheme 44. Solid-phase flow synthesis of BMS-275291 by Lectka

At the time of our studies there was only one published report on asymmetric organocatalytic

aldol reactions in continuous flow conditions by Seeberger’s group.116 They showed the proof

of principle of carrying out an organocatalytic aldol reaction in flow conditions. The addition

of acetone to substituted benzaldehydes could be carried out at elevated temperatures with no

significant deterioration in enantioselectivity when compared to batch conditions (Scheme 45).

However, no in line scavenging or purification procedures were considered, and the reactions

were worked up under standard conditions and purified by SiO2 chromatography after

collection.

Page 73: Organocatalysis and Continuous Flow Reactors as Enabling Tools

71

Scheme 45. Continuous flow aldol reaction with homogeneous catalysis

The current state of asymmetric organocatalytic chemistry in continuous flow is progressing

rapidly, and several research groups are investigating the use of polymer or solid supported

reagents. The Pericàs group has recently made impressive advances in aldol and Mannich

reactions in continuous flow conditions with polymer supported prolines.117 Both

propionaldehyde and isovaleraldehyde could be added to glyoxalate imine 182 with high

selectivities by pumping the starting material stock solutions through a column packed with

triazole linked 4-trans-hydroxy proline 183 on a Merrifield resin (Scheme 46).117a

Scheme 46. Continuous flow Mannich reaction catalyzed by resin bound proline 183

The Pericàs group has also reported the use of a slightly different proline derivative 186 in the

aldol addition of cyclic ketones to aromatic aldehydes. In batch screening they found that

electron poor aromatic aldehydes gave best yields and selectivities. In a continuous-flow

setting, a mixture of 4-nitrobenzaldehyde 128 and cyclohexanone 185 in DMF:H2O (60:40)

was pumped through a column packed with 600 mg of swollen resin containing 186. The flow

rate was 25 �L/min, and residence time 26 minutes. Nearly 20 millimoles of 187 was

Page 74: Organocatalysis and Continuous Flow Reactors as Enabling Tools

72

produced over the course of 30 hours at a constant conversion of above 80% (Scheme 47).117b

As a downside, the product had to be purified by SiO2 chromatography after the reaction to

separate it from possible side products and excess starting material.

Scheme 47. Continuous flow aldol reaction catalyzed by resin bound proline 186

In addition to silica or polymer supported reagents, non-covalent immobilization techniques

have been explored. In 2011 the group of Sels reported the use of chiral primary amino acid

derived diamines immobilized on solid acids as catalysts in the aldol reaction between

aliphatic ketones and aromatic aldehydes.118 The optimal catalyst system was determined in

batch conditions and consisted of phenylalanine derived diamine 190 and sulfonated

fluoropolymer Nafion® NR50. 4-(Trifluoromethyl)benzaldehyde 188 and 2-butanone 189

were pumped through a fixed bed at a flow rate of 1 mL/h (Scheme 48). Over the course of 11

hours, the cumulative yield of the aldol product 191 showed some signs of deterioration, and

the enantiomeric excess peaked at 4 hours at 97% ee, after which it also started to show some

erosion, although staying above 90% during the whole experiment.

Scheme 48. Non-covalent immobilized organocatalyst in continuous-flow aldol reaction

Page 75: Organocatalysis and Continuous Flow Reactors as Enabling Tools

73

3.3 Total Synthesis of (–)-Hennoxazole A

3.3.1 Comparison of Previous Syntheses

Wipf’s synthesis of (+)-Hennoxazole A

The first total synthesis of hennoxazoles was published in 1995 by Wipf.72a The

stereochemistry of the pyranacetal ring and bridging chain was derived from two separate

Sharpless asymmetric epoxidations. The sole stereocenter in the triene fragment was attained

with a Corey-Itsuno reduction of stannane 192 (Scheme 49). A key step in the synthesis was a

Pd-catalyzed coupling between stannane 194 and bromide 195. The final molecular backbone

was constructed via an amide coupling between ester 196 and amine 197 to give the �-

hydroxy amide 198. The bisoxazole structure was then finished via sequential Dess-Martin

oxidation, cyclization, and dehydrohalogenation. Deprotection with TBAF gave the (+)-

enantiomer of hennoxazole A in 23% yield from 198. Overall the synthesis is rather long,

providing key fragments for the final coupling in 21 steps and 3% yield (for 196), and 15

steps and 3% yield (for 197). The stereochemistry of (–)-hennoxazole A was thus confirmed

to be the opposite of the final product attained by Wipf, based on the opposite optical rotation

while other spectral data was identical to (–)-107a.

Page 76: Organocatalysis and Continuous Flow Reactors as Enabling Tools

74

Scheme 49. Key steps in the synthesis of (+)-hennoxazole A by Wipf

Synthesis of (–)-Hennoxazole A by Williams

The second total synthesis of (–)-107a was reported by Williams et al. in 1999.72b The

synthesis is slightly more convergent, with a bisoxazole core 202 synthesized first and then

coupled with pyranacetal precursor 200, and with a late stage Julia-Kocienski olefination to

append the triene tail (Scheme 50). The bisoxazole coupling partner 202 was constructed by a

double sequential mixed anhydride amide coupling, DAST-mediated cyclization and

dehydrogenation, and a final DIBAL-H reduction. The coupling to stannane 200 was

mediated by (R,R)-borane 201 with a 10.5:1 diastereomeric ratio. Stannane 200 was obtained

by transmetalation of the corresponding silane, which in turn was synthesized by copper-

catalyzed Grignard addition of 2-bromo-3-trimethylsilylpropene into epoxide 199. The

stereochemistry at C4 of the pyranacetal 206 was thus derived from (S)-epichlorohydrin which

was used as a starting material, and at C6 from a Terashima reduction of intermediate 206.

The triene precursor 207 for the Julia-Kocienski reaction was synthesized in 11 steps from 3-

butyn-2-one.

Page 77: Organocatalysis and Continuous Flow Reactors as Enabling Tools

75

Scheme 50. William’s synthesis of (–)-hennoxazole A

Synthesis by Shioiri

In their synthesis from 2000, Yokokawa and Shioiri followed Wipf’s strategy of a late stage

amide coupling of amine 211 and acid 212, and subsequent oxidation/cyclization/dehydration

sequence to furnish (–)-hennoxazole A (Scheme 51).72c,d A key feature is a Mukaiyama-aldol

reaction between keteneacetal 209 and aldehyde 208 in the construction of the stereocenter at

C6, giving �-hydroxy ketone 210 as a 88:12 mixture of diastereomers. The remote

stereocenter in the skipped triene is derived from (S)-Roche ester. The first oxazoline

formation was carried out with Deoxo-Fluor mediated cyclization and subsequent

dehydrogenation with BrCCl3/DBU in good yield, but it is very surprising that the same

method was not attempted in the final cyclization, or at least no mention of a failed attempt

was given. The fragments 211 and 212 were synthesized in 18 and 17 steps, respectively, and

a further 6 steps were needed to obtain the final product, rendering the synthesis rather

inefficient. A suprising feature of the synthesis is that the skipped triene moiety is carried

along unchanged for several steps during the endgame, whereas concerns for its stability have

been a directing factor e.g. in the syntheses of Wipf and Williams.

Page 78: Organocatalysis and Continuous Flow Reactors as Enabling Tools

76

Scheme 51. Late stages of Shioiri’s synthesis of (–)-hennoxazole A

Synthesis by Smith

The synthesis published by Smith in 2007 features some of the most elegant solutions to the

construction of (–)-hennoxazole.72e,f The TBS-protected bisoxazole core 217 was synthesized

in 9 steps, beginning with a condensation of serine methyl ester hydrochloride and ethyl

acetimidate to give the first oxazoline 215 (Scheme 52). Sequential dehydrogenation,

saponification and amide coupling with another serine methyl ester gave the amido alcohol

216, which was then cyclized using Deoxo-Fluor. Reduction, acetalization and TBS-

protection of the most labile C-H bond gave 217 in 5 steps. The coupling-ready bromide 222

containing the full triene portion was synthesized in 6 steps from commercial starting

materials with high enantioselectivity (98% ee), using an asymmetric Hoffman pentenylation

and a 2,3-Wittig rearrangement. A drawback in the synthesis is the use of chiral auxiliaries,

which are used to sequentially introduce the stereocenters at C6 and C8 in the pyran ring, and

not without issues with diastereoselectivity (Scheme 53). The key aldol coupling to create the

stereochemistry at C8 proceeded in 6:1 diastereomeric ratio, with a 53% purified yield of 225,

and the second aldol reaction gave 227 in 3:1 dr and 46% purified yield, respectively. The

triene portion was installed by direct alkylation with a strong base, which predicated the

earlier TBS protection of the oxazole. All in all, starting from a rather expensive commercial

carboxylic acid 223 (synthesized in two steps), the longest linear sequence of this synthesis is

14 steps, and overall yield of 4% along the LLS.

Page 79: Organocatalysis and Continuous Flow Reactors as Enabling Tools

77

EtO

NH2Cl SerOMe·HCl, Et3N MeO

ON

O

214 215

1) CuBr2, DBU, HMTA, CH2Cl22) NaOH, MeOH, H2O

3) SOCl2, rfx, then SerOMe·HCl

HN

ON

O

216HO

MeO

O

1) Deoxo-Fluor, CH2Cl2, –20 °C2) DBU, BrCCl3, 0 °C3) DIBAL-H, CH2Cl2, –78 °C4) TMSOMe, TMSOTf, CHCl3, –30 °C5) n-BuLi, TBSOTf, THF, –78 °C

217

N

ON

O

H

OMeMeO TBS

36% from 216

O

H+

218

B O

OCy

Cy

219

OH

220 81%, ee 98%

hexanes, 0 °C

KH, Bu3SnCH2I, THF,then n-BuLi, –78 °C

HO

221

4 steps

222

Br

HO

ON

O

223

Alternative starting material100 EUR/g at Sigma-Aldr ich

CH2Cl2, rfx

Scheme 52. Synthesis of key intermediates by Smith

Scheme 53. Key stereochemical steps by Smith.

3.3.2 Previous studies by the Ley group

The Ley group first generation retrosynthetic analysis of (–)-hennoxazole A is presented in

Scheme 54, consisting of a sp2-sp3 coupling119 to install the triene portion 231 to the

bisoxazole core 230, and a Paterson-type 1,5-anti-aldol reaction120 to couple the pyran

precursor 229 to the nearly complete hennoxazole scaffold in a stereoselective fashion. The

pyran ring would then be finished with OH-directed stereoselective reduction,

Page 80: Organocatalysis and Continuous Flow Reactors as Enabling Tools

78

transacetalization, methylation and deprotection. However, early studies towards the triene

synthesis indicated that the sp2-sp3 coupling is unfeasible,119b and an olefin metathesis

strategy was later adopted.121 As will be shown below, the general strategy to synthesize the

pyranacetal ring was conserved, but the direction of approach was altered.

Scheme 54. Retrosynthesis and first generation approach to pyran ring

3.3.3 First Approach to the Synthesis of the Western Fragment

Our preliminary plan to synthesize and couple the pyranacetal ring was based on an

organocatalytic aldol addition of acetone to the aliphatic aldehyde 237. To synthesize the

substrate, we first attempted a redox neutral selective deprotection of a corresponding diacetal

235, but all attempts with normal acidic deprotection led to dioxolane migration (Scheme 55).

The selective deprotection could be carried out in principle by using TESOTf and 2,6-lutidine,

but this was deemed to be too wasteful due to necessary large scale chromatography.122 The

synthesis of 237 was finally carried out in a straightforward manner over two steps from

ethylacetoacetate 238 by acetal protection and DIBAL-H reduction (Scheme 56).

Page 81: Organocatalysis and Continuous Flow Reactors as Enabling Tools

79

Scheme 55. Attempts at synthesis of 237

O

OEt

O

HO OH

p-TsOH (cat.)CH(OEt)3 (1 equiv)

toluene, rfx OEt

OOO DIBAL-H (1.2 equiv)CH2Cl2, –78 °C

OOOH

238 239 23774% 90%

Scheme 56. Synthesis of aldehyde substrate 237

The organocatalytic aldol addition of ketones to �-unsubstituted aliphatic aldehydes is a

highly challenging process, and in our hands it did not yield promising results. We used

known published catalysts according to literature procedures (Figure 17). Catalysts 125 and

180 were available at hand, and 240 and 241 were synthesized according to published

procedures.123,124 However, none of these gave the desired aldol addition product 242 in any

useful manner (Table 5). Proline was inactive in the reaction (entry 1). The tetrazole analogue

180 was more active (entries 2-3) but upon workup no product was isolated. Catalysts 240

and 241 also performed disappointingly (entries 4-9). Ostensibly the �-unsubstituted aldehyde

competes with acetone in enamine formation with the catalyst, thus possibly leading to self-

condensation products. Previous studies in the group had not yielded very promising results

either for this step. Hence, we decided to revise the synthetic plan.

Page 82: Organocatalysis and Continuous Flow Reactors as Enabling Tools

80

Figure 17. Catalysts screened for the addition of acetone to 237

Table 5. Screening results

Entry Catalyst / Mol % Conditions Result 1 125 / 20 acetone, H2O, rt, 48 h No rxn 2 180 / 10 1:1 acetone:DMF, rt, 48 h Mostly 237, no product 3 180 / 10 4:1 DMSO:acetone, rt, 48 h Full conv., no 242 isolated 4 240 / 10 acetone, rt, 72 h Mostly 237 5 240 / 7 7 mol % TFA, acetone, rt, 72 h Full conv., no 242 isolated 6 240 / 10 acetone, premixing 2 h, rt, 72 h Mostly 237 7 240 / 10 AcOH, acetone, premixing 2 h, rt, 72 h Mostly 237 8 241 / 10 acetone, rt, 72 h Mostly 237 9 241 / 7 7 mol % TFA, acetone, rt, 72 h Almost full conv., no 242

3.3.4 Second Approach to the Synthesis of the Western Fragment

The revised plan is presented in Scheme 57. We decided to use the bisoxazole aldehyde 243

as the aldol acceptor partner in the addition of acetone, as this would better correspond to the

good results in proline-derivative catalyzed aldol reactions. The addition product would then

be alkylated to give 244, which would then be coupled to 237 according to the Paterson 1,5-

anti-aldol protocol.120 The fully functionalized pyran ring would then be achieved in a matter

of 1,3-anti-diastereoselective reduction125 of �-hydroxy ketone 245 and subsequent

transacetalization according to Williams.72b If a protecting group would be needed, an Evans-

Tishchenko reaction could be used to both stereoselectively reduce 245 in 1,3-anti fashion

and install an ester protecting group in one step.126 The presence of a protecting group at C4

hydroxyl might be necessary due to the challenging olefin metathesis that would install the

triene portion. If the pyran ring synthesis could be carried out on the whole bisoxazole-triene-

backbone, no protecting group would be needed.

Page 83: Organocatalysis and Continuous Flow Reactors as Enabling Tools

81

O

O

NH

243

methylation

1,5-anti-aldol

1,3-anti-reduction

OO O

H237

R = H or Pg

transacetalization

N

O acetone aldol

OH

O

N

244

N

O

O OMe

O

N

245

N

O

O

OMe

O

N

246

N

O

OOO OHOMe

O

N

247

N

O

OHOO OR

OMe

O

N

N

OO

OR

HMeOMe

248

44

4

Scheme 57. Revised synthetic plan

3.3.5 Optimization of the Bisoxazole Synthesis

Aldehyde 243 was deemed to be a suitable model substrate, since it could also be used for

screening for olefin metathesis conditions (or the preceding ester 255). The synthesis of 243

was yet unoptimized, and gave the desired product with a 12% yield over 8 steps. By

tweaking the conditions slightly, we were able to raise the overall yield to 36% over the same

route, showing that high throughput would be possible (Scheme 58). The synthesis started

with a CDI mediated coupling of DL-serine methyl ester hydrochloride to 4-pentenoic acid

249 in acetonitrile to give the amide 250. This was then subjected to cyclization conditions

with DAST and K2CO3 in CH2Cl2 at –78 °C and dehydrogenation with BrCCl3 and DBU at

0 °C to give the oxazole 252. The bisoxazole core was obtained by repeating this three

reaction sequence after saponification of 252 with LiOH in THF/H2O. Selective DIBAL-H

reduction of the ester 255 gave the desired aldehyde 243 in good yield after chromatography.

The DAST-mediated cyclizations initially gave poorer yields than those shown below, but

with a fresh reagent the reactions proceeded with excellent yields. The crude products were

also deemed to be pure enough according to 1H NMR analysis for use in the following steps

with no further purification.

Page 84: Organocatalysis and Continuous Flow Reactors as Enabling Tools

82

HO

O

NH

OHO

MeO

O

O

N

MeO

O

O

N

MeO

O

ON

O

NH

O

MeO

O

N

N

O

HMeO

OO

N

N

O

MeO

OO

N

N

O

O

H

DL-SerOMe•HCl,DIPEA, CDIMeCN, rt

89%

DAST, K2CO3

97% crude

BrCCl3, DBUCH2Cl2, 0 °C to rt

84% from 250

1) LiOH2) DL-SerOMe•HClEt3N, CDI, MeCN, rt

79%

crude 95%used as such

78% from 253(42% from 249)

DIBAL-HCH2Cl2, -78 °C

85%, (36% from 249)

HO

249 250 251

CH2Cl2, –78 °C to rt

252253DAST, K2CO3,CH2Cl2, –78 °C

254

BrCCl3, DBUCH2Cl2, 0 °C to rt

255 243

Scheme 58. Optimized synthesis of bisoxazole aldehyde 243

As a test for flow synthesis, we produced 3.2 grams of oxazoline 251 with a continuous flow

DAST-mediated cyclization according to the method developed in the Ley group (Scheme

59).127 The synthesis was carried out by sequential injections of 2-3 mmols of amide 250 and

DAST as stock solutions in anhydrous CH2Cl2 through separate injection loops. The

combined reagent stream was heated to 80 °C in a coil reactor with a residence time of 30

minutes. After the coil the mixture was directed into sequential columns packed with CaCO3

and SiO2. Since colored impurities would eventually leach through the SiO2, we repacked the

SiO2 column after three consecutive runs. The oxazoline 251 was deemed pure enough by 1H

NMR spectroscopy to be carried to the next reaction without further purification.

Scheme 59. Synthesis of 251 with a continuous-flow reactor

Page 85: Organocatalysis and Continuous Flow Reactors as Enabling Tools

83

3.3.6 Catalyst Screening

Our first goal was to find the optimal organocatalyst for our model compound before

transferring the reaction to continuous-flow conditions. Hence, the catalyst screening was

carried out in small scale in batch conditions. The catalysts that were tested are presented in

Figure 18, and the results from the catalyst screening in Table 6. Catalysts 256, 257 and 258

were synthesized according to literature, and others were available from previous experiments

or from commercial sources.123,128

Figure 18. Organocatalysts used in screening for revised route

Page 86: Organocatalysis and Continuous Flow Reactors as Enabling Tools

84

Table 6. Screening results in the addition of acetone to aldehyde 243

Entry Catalyst / Mol % Conditions Time Yield / ee 1 125 / 20 Neat, rt 5 d 11% / 3% 2 180 / 10 Neat, rt 5 d 36% / 9% 3 240 / 10 Neat, rt 5 d 36% / 27% 4 241 / 10 Neat, rt 5 d 73% 243 5 256 / 10 Neat, rt 5 d 33% / 56% 6 256 / 20 Neat, rt 5 d 33% / 47% 7 256 / 100 Neat, rt 2 h 43% / 34% 8 256 / 20 1:1 i-PrOH:acetone, rt 5 d 22% / 24% 9 257 / 10 Neat, rt 6 d 43% / 80%

10 257 / 20 Neat, rt 22 h 48% / 60% 11 258 / 20 Neat, rt 4 d 59% / 46% 12 258 / 20 Neat, rt, 10 mol % AcOH 4.5 h 58% / 57% 13 258 / 20 Neat, rt, 10 mol % MCA 6 h 68% / 83% 14 259 / 20 Neat, rt, 10 mol % AcOH 4 d ee < 10% 15 259 / 20 Neat, rt, 10 mol % AcOH, Na2SO4 4 d ee < 10% 16 260 / 20 Neat, rt 5 d no rxn 17 119 / 20 Neat, rt 2 d no rxn 18 119 / 20 Neat, rt, slow addition 5 d no rxn 19 119 / 20 Neat, rt, 10 mol % AcOH, slow add 5 d no rxn

As a general trend, the heterocyclic aldehyde 243 seemed to be less reactive than most

aromatic aldehydes. Decent reactivity was obtained only with organic acids as co-catalysts.

For entries 1-8, complete conversion was observed only in the cases of 5 and 6. The

prolinamides 257 and 258 reported by Singh128 were shown to be the most active by the initial

experiments without co-catalysts (entries 9-13), the reaction going to completion in 22 hours

when 20 mol % of 257 was used. Entries 14 and 15 gave both condensation and addition

products, obtained in a roughly 1:2 ratio, and the catalyst 259 did not show any asymmetric

induction. Hayashi-Jørgensen-type diphenyl-prolinol derivatives showed no reactivity in these

screening conditions (entries 16-19). Pleasingly, there was little difference whether the

reaction was done in rigorously dried anhydrous acetone or in distilled acetone, which

contains some moisture. The aldol addition product 244 is crystalline, and the enantiomeric

excess could be enhanced by crystallization. As an experiment, crops from several low ee

entries were combined and crystallized to give 244 with an ee of 70%. Generally in the test

conditions small amounts of intensively coloured impurities followed through with the

Page 87: Organocatalysis and Continuous Flow Reactors as Enabling Tools

85

product in column chromatography. However, these could at least partially be removed via

crystallization of the product when larger amounts were handled.

Based on the results from the catalyst screen, the Singh prolinamide 257 was chosen for

further studies. At this time, more of the catalyst was also synthesized (Scheme 60). The Boc-

deprotection was originally attempted with TFA:CH2Cl2, but partial hydrolysis of the amide

was observed.

Scheme 60. Synthesis of catalyst 255

The results from the optimization of the acetone aldol step are presented in Table 7. The

reaction was investigated by varying reaction temperature, co-solvents, catalyst loading, and

the use of a co-catalyst.

Page 88: Organocatalysis and Continuous Flow Reactors as Enabling Tools

86

Table 7. Optimization of organocatalytic conditions for the synthesis of 244

Entry 257 (mol %) Conditions Time Yield / ee 1 10 Neat, rt 6 d 43% / 80% 2 20 Neat, rt 22 h 48% / 60% 3 30 Neat, rt 6 h 50% / 73% 4 20 Neat, rt, 10 mol % TFA 5 d 18% / 90% 5 20 Neat, rt, Na2SO4 22 h 50% / 69% 6 20 1:1 i-PrOH:acetone, rt 6 h 45% / 42% 7 20 1:1 MeCN:acetone, rt 2 d 47% / 76% 8 30 Neat, 0 °C 1 d 53% / 52% 9 20 Neat, rt, 10 mol % AcOH 1 h 50% / 70%

10 20 Neat, 0 °C, 10 mol % AcOH 1.5 h 40% / 81% 11 20 Neat, 0 °C, 10 mol % MCA 3 h 59% / 90% 12 10 Neat, rt, 10 mol % MCA 1 d 68% / 90% 13 20 Neat, rt, 20 mol % HCOOH 1 h 59% / 85%

Catalyst loading had a significant effect on reaction speed (entries 1-3). For entry 4 the

reaction time is inaccurate, since only after work-up and purification it was realized that the

aldol condensation product 263 completely overlaps on TLC with 243, and the reaction in the

presence of TFA was ostensibly a lot faster. The presence of a dehydrating agent had no great

effect (entry 5), and changing the solvent system mostly had a detrimental effect, if any

(entries 6-7). In 0 °C the reaction became sluggish in spite of high catalyst loading, and ee

suffered (entry 8). The presence of acetic acid greatly accelerated the reaction, and yields and

enantioselectivities were further improved by the use of MCA as a co-catalyst (entries 9-12).

Under the best conditions we were able to isolate 244 in 68% yield, with ee slightly above

90%. Entry 13 describes an experiment that was carried out after it was realized that HCOOH

performed better as the co-catalyst under continuous flow conditions.

We attempted to verify the expected absolute configuration 244 by Mosher’s ester analysis,129

but the results were inconclusive (Scheme 61). The first derivatization attempt with (S)-(+)-

MTPA-Cl resulted in complete conversion to the elimination product 263. Another attempt

with DCC-coupling of (R)-(–)-MTPA was successful, and the corresponding crude ester 264

was analyzed by 1H NMR. Downfield shift of 1H NMR signal in the closer oxazole moiety

was observed, along with an upfield shift of the terminal methyl group, giving tentative proof

to the desired R-configuration according to the procedure by Mosher. However, the �-

Page 89: Organocatalysis and Continuous Flow Reactors as Enabling Tools

87

methylene did not completely follow this trend: the group exhibited downfield shift of one

diastereotopic H, and a slight upfield shift of the other.

Scheme 61. Attempt at determination of absolute configuration of 244

3.3.7 Organocatalytic Aldol Step under Continuous Flow Conditions

We began our screening for continuous flow conditions with an approach similar to

Seeberger’s (see Scheme 45).116 A solution of the catalyst 257 and acid co-catalyst in acetone

was injected through one injection loop, and the aldehyde 243 through another according to

the scheme below. The mixture was pumped through the coil reactor at desired temperature

and analyzed with TLC and crude 1H NMR spectroscopy after evaporation. If the conversion

was good enough, the product was isolated after chromatography and analyzed by HPLC to

determine the enantioselectivity. The results of the screening are presented in Table 8.

Page 90: Organocatalysis and Continuous Flow Reactors as Enabling Tools

88

Table 8. Continuous flow studies in organocatalytic aldol addition

Entrya 257 / Cocat (mol %) c (243)b Flow ratec tR Temp 243:244:263d Yield / ee 1 10 / - 0.05 0.5 mL/min 10 min rt 1:0.1 n.d. 2 10 / - 0.05 0.18 mL/min 30 min rt 1:0.1 n.d. 3 10 / 10 MCA 0.05 0.2 mL/min 25 min rt 1:0.3 n.d. 4 20 / 10 MCA 0.05 0.34 mL/min 15 min rt 1:0.75 n.d. 5 10 / 10 MCA 0.15 0.2 mL/min 25 min rt 1:0.3 n.d. 6 10 / 10 MCA 0.05 0.2 mL/min 25 min 50°C 1:0.3:0.05 n.d. 7 10 / 10 MCA 0.05 0.2 mL/min 25 min 70°C 1:0.4:0.1 n.d. 8 20 / 10 MCA 0.10 0.2 mL/min 25 min 60°C 1:1:0.2 n.d. 9 20 / 10 MCA 0.15 0.2 mL/min 25 min rt 1:1 n.d. 10 20 / 20 HCOOH 0.15 0.2 mL/min 25 min 50°C 1:45:trace 60% / 75% 11 20 / 20 HCOOH 0.15 0.1 mL/min 50 min rt 1:6 59% / 80% 12 20 / 20 HCOOH 0.25 0.5 mL/min 40 min 40°C nd 40% / 70% 13 20 / 15 MCA 0.33 0.33 mL/min 60 min rt 1:3.5:0.2 45% / >90%

a) The reactions were carried out by injecting 1 mL of stock solutions via injection loops into the flow reactor. b) Refers to total concentration of 243. c) Total flow rate, individual pumps functioning at 50% of shown rate. d) Determined by 1H NMR of evaporated crude output mixture.

As shown in the table, MCA did not perform well as the co-catalyst under continuous flow

conditions, and it took several attempts to see any significant amount of product (entries 1-9).

Better results were achieved when 20 mol % of formic acid was used (entries 10-12). The

reaction had to be heated to attain good conversion, and quite pleasingly we were able to

determine that the enantioselectivity was nearly as good as in homogeneous reactions, with

some expected erosion occurring. The best yield was achieved when 20 mol % of both 257

and HCOOH were used and the reaction was run at 50 °C with a residence time of 25 minutes

(entry 10). However, a nearly identical yield was achieved when the residence time was

doubled and the reaction was run at room temperature (entry 11). The enantioselectivity was

slightly higher in the latter case. To test whether the poor results with MCA were attributed to

the loading, one more test was run with 15 mol % of co-catalyst, and this time a slightly better

conversion was achieved, and the reaction again gave an ee of 90% (entry 13). The aldol

condensation product 263 could also be detected in the analyses of crude mixtures especially

at higher temperatures.

Page 91: Organocatalysis and Continuous Flow Reactors as Enabling Tools

89

3.3.8 Methylation studies

The methylation of the aldol addition product 244 was briefly investigated in batch and flow

conditions. Several common methylation methods were attempted (Scheme 62). Treatment of

244 with sodium hydride in DMF led to decomposition of the starting material. Reaction with

Meerwein’s salt gave full conversion, but after work-up only the pyridine base was recovered,

indicating either decomposition or methylation of the heterocycle core to form a charged

oxazolium species which would remain in the aqueous phase. Treatment of 244 with dimethyl

sulfate gave no reaction. In the end, methylation with an excess of methyl iodide in the

presence of Ag2O gave a clean conversion in small scale reactions. For optimal reactivity

Ag2O had to be freshly prepared. Due to time restraints, only a proof of principle experiment

was done in continuous-flow conditions. A glass column was filled with a dispersion of Ag2O

in celite, with a small plug of celite in each end. A solution of 244 and an excess of MeI in

CH2Cl2 was pumped through the column, and the product was analyzed by 1H NMR

spectroscopy, showing a 2:1 mixture of unreacted 244 and the desired product 245 (Scheme

63), with no side products. It was also determined by chiral HPLC analysis that the

methylation step with MeI/Ag2O proceeds with no racemization, and the enantiopurity of 245

may likewise be enhanced by recrystallization.

Scheme 62. Methylation attempts of 244

Page 92: Organocatalysis and Continuous Flow Reactors as Enabling Tools

90

Scheme 63. Methylation of 244 in continuous flow

3.4 Diastereoselectivity of �-Hydroxy Ketone Reduction with Solid Supported

Borohydrides

One of the endgame steps in the synthesis would have included a diastereoselective 1,3-anti-

reduction of a �-hydroxy ketone (Scheme 57). While there are established and well-tested

methods for 1,3-anti-reductions, such as the Evans-Carreira-Chapman modification of

Saksena reduction,125 and Evans-Tishchenko reduction,126 at the time of our studies there

were no reports on the diastereoselectivity in the reduction of �-hydroxy ketones with

immobilized borohydrides. Formally, the resin bound borohydride would correspond to a

quaternary ammonium borohydride compound (Scheme 64). Further modification to a

triacetoxyborohydride would then give an immobilized version of the Evans-Saksena reagent

that might be applicable to continuous flow conditions. Hence, we set out to briefly

investigate the matter.

Scheme 64. Reduction of �-hydroxyketone with immobilized borohydrides and the corresponding transition state leading to 1,3-anti-enriched product via complexation and intramolecular hydride attack.

For model compounds we synthesized 265-267 by standard aldol reactions between

corresponding ketones and aldehydes (Figure 19). A commercial polymer supported

borohydride was purchased from SigmaAldrich, but after it showed rather poor activity in the

model reaction, we carried out the immobilization ourselves.130 The borohydride ion was

Page 93: Organocatalysis and Continuous Flow Reactors as Enabling Tools

91

immobilized on Amberlite IRA-400 (Cl– form) by washing the resin first with 1 M HCl and

distilled H2O, and then stirring it in a 1 M aqueous solution of NaBH4 and then drying the

resin cautiously in vacuum.

Figure 19. Model substrates for reduction studies

Reactions with solid supported borohydride were carried out by dissolving 1 mmol of �-

hydroxyketone into an alcoholic solvent, adjusting to desired temperature and then adding 1

gram of borohydride resin (2-3 equivalents based on theoretical loading). The Rychnovsky

cyclic acetonide method was used for determining the diastereoselectivity, and the 1H NMR

signals of CHOH protons at 3.37 (anti) and 3.45 (syn) were used for integration according to

Cohen.131,132 The products were isolated by filtering the reaction mixture and evaporating

solvents to dryness. An example of the reaction is shown in Scheme 65.

Scheme 65. Reduction of 265 with solid supported borohydride

The model reaction gave a syn-selective product, as usually does a standard reduction with

NaBH4 (a control experiment characterized with crude 1H NMR spectroscopy). Pretreating

the resin with glacial AcOH should give the triacetoxyborohydride species.134 The

borohydride resin was stirred in THF at 0°C with 3 equivalents of AcOH for 45 minutes, then

filtered and washed. However, the activity of the treated resin was lower, and still gave a syn-

enriched product with practically same selectivity as the untreated resin (Scheme 66). We also

attempted to perform the reaction in a mixture of AcOH and MeOH, but observed no reaction

Page 94: Organocatalysis and Continuous Flow Reactors as Enabling Tools

92

taking place after strong initial bubbling during which the acid reacted with the borohydride

ions.

Scheme 66. Reduction of 265 with resin pretreated with glacial AcOH

To see how the selectivity could be tuned, we briefly investigated the effect of temperature

and solvents. Scheme 67 highlights the effect of temperature in the model reaction. By

heating the reaction we were able to boost the reaction speed significantly, with no great

effect to diastereoselectivity. This indicates also that the polymer supported borohydride

could theoretically be used in a continuous flow fashion as a packed column reactor, although

resin based borohydrides have been previously reported to be unreliable in flow conditions.76

However, this problem has been solved recently by using a monolithic quaternary ammonium

polymer loaded with borohydride ions.134

Scheme 67. The effect of temperature on reaction speed and selectivity

The effect of solvents is depicted in Table 9. For borohydride reductions, MeOH is usually

the best solvent with respect to reaction speed, since a precomplexation between the solvent

and the borohydride ion apparently results in a more active reagent. When the reaction was

run in i-PrOH, we observed an anti-enriched product although the reaction was very slow and

not as clean as in MeOH (entry 1). A non-alcoholic solvent gave no reaction (entry 2).

Mixtures of MeOH and other solvents gave slower reaction times and varying selectivities

Page 95: Organocatalysis and Continuous Flow Reactors as Enabling Tools

93

(entries 3-6). Mixing i-PrOH with MeOH reversed the selectivity once more towards the syn-

isomer. Heating the reaction in i-PrOH gave a faster conversion (not shown in table), but after

filtration no product was isolated, which may be due to the instability of the resin in higher

temperatures. In entry 6 we carried out a hydrolytic and oxidative work-up with NaOH/H2O2,

which gave a further 50% yield of recovered product after extraction. The product thus

recovered consisted only of the syn-isomer, bringing the total yield to about 80% and

selectivity to 3:1 syn:anti. Keeping in mind the original goal of achieving an anti-selective

diastereoselective reduction in continuous flow conditions, these results were less than

promising, due to the slow kinetics and synthetically unusable selectivities even in the case of

i-PrOH.

Table 9. Effect of solvents on the reduction of 265

Entry Solvent Time Conversion syn:antia Yield 1 i-PrOH 6 d Full 1:~2.5b n.d 2 THF 6 d – – n.d 3 1:1 MeOH:THF 21 h Full 3:1 95% 4 4:1 i-PrOH:MeOH 16 h Full 1-1.5:1 n.d 5 1:1 i-PrOH:MeOH 16 h Full 2.5:1 n.d 6 4:1 s-BuOH:MeOH 5 d Full 1:1.5 30% a) The diastereoselectivity was determined via Rychnovski acetonide method. b) Approximated from crude product 1H NMR spectrum.

Finally, we attempted a simple substrate experiment to determine the effect of substitution.

Whereas compound 267 gave a diastereoselectivity of 2:1 syn:anti, reversing the position of

the ketone and hydroxyl groups gave a completely nonselective reaction (Scheme 68). In the

case of 267 we also treated the remaining resin with NaOH/H2O2 and recovered further 27%

of pure syn-product.

Page 96: Organocatalysis and Continuous Flow Reactors as Enabling Tools

94

Scheme 68. Substrate experiments

3.5 Conclusions

We investigated the application of an organocatalytic aldol reaction into the total synthesis of

(–)-Hennoxazole A, and the possible transfer of the reaction into continuous flow conditions.

Organocatalytic reactions may not be ideally suited to continuous conditions due to the

intrinsic slow kinetics. To increase the rate suitably, the reaction needs to be heated which

leads to increased formation of side products and erosion of enantioselectivity. The best

catalyst for the reaction was 257, previously reported by Singh. The addition of acetone to

bisoxazole aldehyde 243 could be carried out in 60-70% yield and over 90% enantiomeric

excess. The methylation of the aldol product 244 was also investigated and suitable conditions

determined

The reaction was also successfully tested in continuous flow conditions, but further

optimization and development would be needed. It would be beneficial to develop either in

line quenching and scavenging methods for purification of the intermediate product, and to

investigate telescoping the aldol addition to the following methylation step. Another approach

would be the immobilization of the organocatalyst, of which there were no precedents at the

time of our study. The methylation step was also tested in continuous flow conditions, but no

optimization studies were yet carried out.

We briefly investigated the diastereoselectivity of �-hydroxy ketone reductions with polymer

supported borohydride resins. The reduction may be carried out with simple substrates, but

the diastereoselectivities did not show great promise in application to asymmetric synthesis.

In most cases the undesired syn-diastereomer was obtained, and while i-PrOH was the only

solvent that provided the desired anti-diastereomer, the reactions were very slow and of poor

purity. Another issue is the potential formation of six-membered boronate esters between the

Page 97: Organocatalysis and Continuous Flow Reactors as Enabling Tools

95

substrate and the reducing agent, which remains complexed on the polymer beads. Thus,

based on our results we are unable to confirm the validity of using immobilized borohydrides

in a diastereoselective reduction, let alone apply it to the total synthesis of (–)-Hennoxazole A.

Page 98: Organocatalysis and Continuous Flow Reactors as Enabling Tools

96

4 NITRATION OF AROMATIC SUBSTRATES UNDER

CONTINUOUS FLOW CONDITIONS

4.1 Common aspects of aromatic nitration

Nitration of aromatic compounds is a classic reaction, which is still a standard tool of modern

day chemical industry. Nitration reactions are usually carried out using either concentrated

nitric acid or a mixture of concentrated sulfuric acid and nitric acid (generally called the

mixed acid method), depending on the reactivity of the substrate. For electron rich aromatic

substrates even dilute nitric acid may be sufficient, as will be shown below. For unsubstituted

aromatic compounds such as benzene, or electron poor aromatic substrates, forcing conditions

with mixed acid is usually needed. Nitration reactions are frequently carried out in neat

conditions using only the acid mixture as solvent. Acetic acid is a frequently used solvent,

being safe to handle in the presence of concentrated HNO3. Other approaches and reagents

towards nitration include e.g. the in situ formation of acetyl nitrate from acetic anhydride and

concentrated HNO3, ionic complexes of nitronium cation [NO2]+ such as nitronium

tetrafluoroborate, and metal nitrates.

The kinetics of nitration are complex and depend on several factors, e.g. the nitration medium,

solvent, nitronium ion formation and activity of the substrate towards electrophilic aromatic

substitution.135 When electron-rich aromatic subtrates are used, zero order kinetics are usually

observed in respect to the substrate, and first order kinetics with respect to the concentration

of HNO3, or rather the [NO2]+-ion. For unreactive aromatic compounds nitration usually

follows first order kinetics with respect to the substrate. For compounds of intermediate

reactivity, the reaction can be forced towards either zero or first order kinetics based on the

exact conditions used, e.g. acidity of nitration medium and solvent. The nitration of electron

rich aromatic substrates may proceed also via electrophilic attack by nitrosonium ion [NO]+,

and a subsequent oxidation of the nitroso group to nitro group.136

Page 99: Organocatalysis and Continuous Flow Reactors as Enabling Tools

97

4.2 Nitration of Vanillin

Nitration of vanillin 271 is a well-known reaction (Scheme 69).137,138 The highly activated

aromatic ring is reactive enough for the reaction to be carried out by using e.g. a

stoichiometric amount of 60 w-% aqueous nitric acid. In addition to traditional methods,137

the nitration of vanillin has been investigated e.g. with benzoyl nitrate,138a gaseous NO2,138b

SiO2-absorbed acetyl nitrate,138c and different metal nitrate complexes.138d-g The resulting 5-

nitrovanillin 272 has been used in medicinal chemistry as a precursor for e.g. catechol O-

methyltransferase inhibitors,137j,l combretastatin A analogues,137m and HIV-1 integrase

inhibitors.137k

H

OMeO

HO

271

Et2O, 10 °C

HNO3, HNO2H

OMeO

HO

272

NO2

95%, Vogl 1899 (Ref. 137b)

HNO3

AcOH, 0 °C to 20 °C272 83%, Klein 2011 (Ref. 137n)

ZrO(NO3)2·H2Oacetone, rt

272 94%, Venkateswarlu 2006 (Ref. 138e)

272 100%, Francis 1906 (Ref. 138a)BzONO2

CCl4

272 88%, Rodrigues 1999 (Ref. 138c)AcONO2 / SiO2

CHCl3

NO2 (g) 272

84% 272, 10% 273, 6% 274

H

OMeO

HO

273

+

NO2

H

OMeO

HO

274

+NO2

Kaupp 1995 (Ref. 138b)

Scheme 69. Examples of nitration of vanillin 271

In connection with studies for an enhanced process towards an important pharmaceutical

intermediate, it was in our interest to examine whether the reaction could be carried out in

continuous flow conditions, and to see if it would result in improved selectivity, yield, and/or

Page 100: Organocatalysis and Continuous Flow Reactors as Enabling Tools

98

suppression of unwanted side reactions. In general, the reaction is usually carried out as a

semi-batch reaction by slowly adding concentrated pure or aqueous nitric acid into a solution

of vanillin in glacial acetic acid or another solvent. The strongly exothermic reaction is cooled

during the addition of HNO3 to keep the internal temperature as low as possible. The product

precipitates out of the reaction, and is usually isolated by further precipitation from cold water

or via direct filtration of the reaction mixture, depending on the solvent system used. Yields

are generally around 80–85% and the HPLC purity of the isolated material is over 99%.

Higher yields have been reported when the reaction is carried out in diethyl ether.137b

According to literature, the nitration of vanillin is usually highly regioselective, and 272 is

usually isolated in excellent purity. Known side products include 275 and 276, both arising

from a formal ipso-nitration-deformylation cascade (Scheme 70).137f Another reported side

product 277 ostensibly results from a dehydrogenative homocoupling at the 5-position.137c In

our previous studies, an analogous dinitration product to 276 was noted to be the most

common impurity when a slightly different aldehyde was used in large scale experiments. It

was unknown whether this side product was formed via a parallel reaction path, or whether it

was the result of overreaction of the desired product. Determining this was a secondary goal

in our studies.

Scheme 70. Previously known side products 276 and 277 observed by Bentley, and 275 observed by Stostkii in the presence of HNO2.

4.2.1 Preliminary studies

The equipment used for continuous-flow experiments was a commercial Uniqsis FlowSyn

reactor equipped with either a 5 mL HT-PFA reactor coil or a 10 mL stainless steel reactor

Page 101: Organocatalysis and Continuous Flow Reactors as Enabling Tools

99

coil. The internal diameter of both coils is roughly 0.5 millimeters, and it was anticipated that

transferring an essentially heterogeneous reaction to these surroundings might result in

blockages in either the reaction coil or the back pressure regulator which is attached to the end

of the coil. As the nitrating reagent we used 60 w-% aqueous HNO3, the nitrous acid content

of which was not analyzed. The acid was colorless, indicating a low presence of other NOx

species. The reactions were analyzed qualitatively by HPLC, and all subsequent mentions of

HPLC analyses and numerical values in charts and tables refer to area percentages of the

crude chromatograph. The UV-extinction coefficients of 271 and 272 are ostensibly not equal

and hence some error is most likely included in the raw HPLC values.

Our first objectives were to discover reaction conditions that would be as close to

homogeneous as possible, if a completely solution-phase reaction would turn out to be

impossible to attain. These studies were carried out in batch conditions according to published

procedures. The use of co-solvents was also considered. The solubility of 272 into THF,

CH2Cl2, CHCl3, MeCN or PhCl is roughly 5-15 mg/mL at room temperature (on average 0.05

M). The solubility of 272 into glacial acetic acid as a function of temperature was plotted

roughly to get an idea of a workable temperature range (Chart 1). The solubility was

determined by adding measured amounts of 272 into 15 mL of AcOH at determined

temperature until no more dissolved. It was our anticipation that the continuous-flow

conditions for successful nitration would differ drastically from batch conditions, namely that

the reaction concentration would have to be a lot smaller, and conversely temperatures would

have to be higher. After preliminary experiments, we eventually decided to use glacial acetic

acid as the sole solvent to prevent possible (though unlikely) competitive side reactions of

HNO3 with other solvents. Hence all subsequent mentions of stock solutions refer to

concentration of reagents in AcOH unless otherwise stated.

Page 102: Organocatalysis and Continuous Flow Reactors as Enabling Tools

100

� Chart 1. Rough solubility curve of 5-nitrovanillin 272 in glacial acetic acid

Table 10 shows the results from our preliminary experiments. The initiation of the reaction

was screened in a reaction concentration of 0.1–0.5 M. To see if the absence of acetic acid is

important, the reaction was run in THF, with the result that it is a necessary component

(entries 1-2). In several cases the reaction had to be incubated for a long period before

initiation (e.g. entries 3-7). Increasing the amount of HNO3 did not always result in a faster

reaction (entries 4 vs. 5) but eventually led to significant overreaction, and thus no desired

product was isolated in the experiment of entry 5. Cooling the reaction mixture severely

inhibits the initiation in low concentrations (entries 6 and 7). In the case of entry 8, the

reaction did not first proceed at all, but after stirring for 4 hours the characteristic change of

color of the reaction mixture was suddenly observed, and the reaction was over in a matter

moments. Raising the temperature above ambient conditions shows that even in low

concentrations the reaction starts almost immediately after adding HNO3 (entries 9-10). The

reaction could also be carried out by adding a 0.2 M solution of HNO3 in AcOH into a 0.2 M

solution of vanillin in AcOH. However, the reaction only took place after all of HNO3 had

been added (entry 11). As a benchmark, we ran the reaction in typical process concentration

and temperature (entry 12). We observed that the conversion of 271 was incomplete, most

likely due to cocrystallization with the product during the course of the reaction. The product

contained 0.9% of 271, and the filtrate nearly 18% (44% of product, high amounts of

impurities).

0

10

20

30

40

50

60

70

80

20 40 60 80 100

mg/mL

T (°C)

Solubility

Page 103: Organocatalysis and Continuous Flow Reactors as Enabling Tools

101

Table 10. Preliminary experiments into reaction initiation and speed

Entry 271 (g) HNO3 (eq.) Solvent c (M) T (°C) t (init.) Yield (%) HPLC (%)

1 5 1.2 THF 0.55 rt n.d. n.d. n.d.

2 5 1.2 1:1 THF:AcOH 0.55 rt 5 min 77 99.6

3 1 6 1:1 THF:AcOH 0.11 rt n.d. n.d. n.d.

4 1 1.2 1:1 tol:AcOH 0.11 rt o.n. 65 99.9

5 1 6 1:1 tol:AcOH 0.11 rt o.n. n.d. n.d.

6 2 1.1 1:1 tol:AcOH 0.22 0 o.n. 76 99.6

7 2 1.1 1:1 tol:AcOH 0.26 15 o.n. 75 99.8

8 3.3 1.1 1:1 tol:AcOH 0.38 0 4 h 78 99.3

9 2 1.1 1:1 tol:AcOH 0.26 35 1 min 73 99.9

10 2 1.1 AcOH 0.33 50 0 s 75 99.6

11 1.85 1.1 AcOH 0.11 rt 40 min 57 99.9

12 10 1.05 1:2 tol:AcOH 1.5 5 0 s 75 98.8

In a stress experiment, vanillin was dissolved into 1:1 toluene:AcOH, and HNO3 was added to

the cooled reaction mixture (Scheme 71). After the addition of 6 equivalents the reaction had

not yet started. Cooling was stopped and the reaction let slowly warm to ambient temperature.

The reaction took place 15 minutes after the removal of cooling, and it went to completion

immediately. However, only 15% of product was isolated, and according to HPLC it

contained 4.6% of 276 whereas the mother liquor contained 64% of 276. Hence, the use of

excess nitric acid is detrimental to the reaction.

H

O

HO

MeOHNO3 (60 w-%)

1:1 toluene:AcOH, 0 °C to rt

H

O

HO

MeO

NO2

271, 0.2 M, 13 mmol 272

Addition of HNO3:0.2 mL aliquots at 30 s intervals until 2 mL (2 equiv) added0.5 mL aliquots at 60 s intervals until 6 mL (6 equiv) added

15%

Scheme 71. Stress test of reaction initiation at low concentration and temperature

Page 104: Organocatalysis and Continuous Flow Reactors as Enabling Tools

102

In most of the experiments in Table 10, 272 could be seen precipitating in the course of the

reaction. As the concentrations would be lowered for continuous-flow experiments, the yields

could be anticipated to suffer as well. Hence, we briefly experimented on an isolation of a

second crop of precipitated product (Table 11). Carrying out the reaction in room temperature

at a concentration of 0.1 M, simple filtration of the reaction mixture gave a yield of 51–60%

(entries 1-3). Concentration of the mother liquor and subsequent second filtration gave a

further 15%. Both crops were of almost equal purity. In another experiment (entry 4), we first

concentrated the reaction mixture and added some toluene before cooling the mixture to 0 °C.

Filtration then gave a yield of 74%, with a slightly poorer purity of 98.9%, with dinitrated

product 276 being the dominant impurity (0.8%).

Table 11. Isolation of 272 from dilute reaction mixtures

Entry c (271/M) 1st crop HPLC 2nd crop HPLC 1 0.10 51% 99.902 15% 99.663 2 0.10 58% n.d. – – 3 0.10 60% 99.917 15% 99.409 4 0.10 74% 98.876 – –

4.2.2 Pulse experiments

Having observed that the nitration of 271 could be initiated at low concentration and high

temperatures in batch, we began screening for continuous-flow conditions. As a first step we

investigated the range of temperature, concentration, and residence time required for the

initiation and completion of the reaction within the reactor coil. The experiments were carried

out as pulse experiments (or plug flow) by injecting stock solutions of vanillin and HNO3 in

glacial AcOH via injection loops into the reactor coil. A schematic diagram of the setup is

presented in Figure 20. These experiments were carried out using a 5 mL HT-PFA (high-

temperature resistant perfluoroalkoxy polymer) coil. The coil material is transparent, and the

Page 105: Organocatalysis and Continuous Flow Reactors as Enabling Tools

103

starting point of the reaction could thus be observed visually. The results are presented in

Table 12.

271 in AcOH

HNO3 in AcOH

Pump A

Pump B

PV

PV

PV = priming valve

Injector

Injectorreactor coil

BPR

bpr = back pressure regulator

AcOH

Selection valves

collection flask

Figure 20. Schematic representation of the continuous-flow system

Table 12. Pulse studies for the initiation of nitration reaction

Series 1a tR (min) T (°C) Init. Series 2b tR (min) T (°C) Init. 1 5 r.t. - 1 5 30 - 2 10 r.t. - 2 5 40 beaker 3 5 30 - 3 5 55 beaker 4 5 40 - 4 10 55 beaker 5 5 50 - 5 5 70 beaker 6 5 60 - 6 5 85 beaker 7 5 75 beaker 7 2,5 85 bpr

Series 3c tR (min) T (°C) Init. Series 4d tR (min) T (°C) Init. 1 2,5 45 beaker 1 3,3 30 beaker 2 2,5 60 beaker 2 5 40 bpr 3 5 60 beaker 3 10 40 bpr 4 2,5 40 bpr

Series 5e tR (min) T (°C) Init. 5 5 60 bpr 1 5 35 coil 6 10 60 bpr 2 2,5 50 coil 7 10 80 coil 3 2,5 70 coil 8 5 80 coil 4 1,67 70 coil 5 5 70 coil

a) c(271) 0.10 M, c(HNO3) 0.11 M; b) c(271) 0.20 M, c(HNO3) 0.21 M; c) c(271) 0.30 M, c(HNO3) 0.31 M; d) c(271) 0.40 M, c(HNO3) 0.41 M; e) c(271) 0.50 M, c(HNO3) 0.51 M

Page 106: Organocatalysis and Continuous Flow Reactors as Enabling Tools

104

In several cases, the reaction only took place in the collection flask after all of the mixture had

been collected (series 1-3). Going up in concentration and temperature, the reaction could be

seen taking place right after the back pressure regulator, as the exiting mixture had turned

bright yellow (series 4). If the concentration was high enough, the reaction then went to

completion in the beaker and the product could be seen precipitating. When concentration was

increased to 0.4-0.5 M, the reaction could be seen taking place within the coil (series 4-5). If

the temperature was too low (series 5, experiment 1), the product precipitated within the coil,

leading to an immediate blockage at the joint between the back pressure regulator and the

reactor coil. Hence, the pulse experiments led us to conclude that a workable concentration

range would be 0.3-0.5 M and that temperatures should be above 70°C, but care should be

taken to avoid possible blockages.

4.2.3 Identification of impurities

After deciding on a proper range of conditions, we were ready to begin studying the reaction

in earnerst with continuous experiments. First we analyzed and identified the most prominent

impurities that were detected in the reaction mixture and products. A typical HPLC

chromatogram of a test run is presented in Figure 21. The impurity at retention time 11.2 min

was identified to be the ipso-nitration product of vanillin, and this was confirmed with an

authentic sample of 4-nitroguaiacol. We also subjected 4-nitroguaiacol 275 to the standard

batch nitration conditions and thus obtained an authentic sample of 4,6-dinitroguaiacol 276

(Scheme 72). This was then identified as the impurity at retention time 13.2 min, and it was

observed in small amounts e.g. the 1H NMR spectrum of previous experiments. In screening

for other potential impurities, both 271 and 272 were treated with AcCl (Scheme 73) in the

presence of Et3N, but upon HPLC analysis of the crude product, the corresponding acetates

did not give previously identified peaks. Hence, esterification of the phenolic OH does not

take place during the reaction.

Page 107: Organocatalysis and Continuous Flow Reactors as Enabling Tools

Figure and pro

Scheme

Scheme

21. Typicalduct 272 at

HO

MeO

2

e 72. Nitrati

HO

MeO

27

e 73. Synthe

l HPLC chr9.75 min. S

NO2

275

ion of 4-nitr

O

H

R

71, 272

esis of aceta

omatogramSee Experim

1 equiv HNO

AcOH, r

roguaiacol 2

AcCl,

CH2Cl2

ates 278 and

m from a testmental Secti

O3 (60-w%)

rt, 5 min

p

275

, Et3N

2, 0 °C

N

d 279

t run. Startinion for HPL

HO

MeO

precipitated p

2

AcO

MeO

2

22

Neither detec

ng material LC details.

NO2

NO2

product pure

276

278, 279

O

H

R

278: R = H279: R = NO2

cted in HPLC

271 is at tR

on HPLC

H

2

C analyses

105

R 6.6 min

Page 108: Organocatalysis and Continuous Flow Reactors as Enabling Tools

106

HPLC/M

vanillin

mononi

treated w

HPLC a

of 3,4-d

and thu

regioiso

correspo

HO

HO

Scheme

4.2.4 Co

The firs

reactor

The rea

through

system

decided

reaction

MS experim

n, 3,4-dihyd

trated prod

with HNO3

analysis of t

dihydroxy-5

us the corres

omer was n

onding carb

280

O

H

e 74. Nitrati

ontinuous ex

st continuo

in order to

actions wer

h the pump

are cerami

d to use HNO

n stream wa

ments revea

droxybenza

ducts of 280

3, and corre

the reaction

5-nitrobenza

sponding 2

not determ

boxylic acid

1 equiv HN

AcOH,

ion of 3,4-d

experiments

ous experim

be able to

re carried o

s into the

ic and high

O3 as a dilu

as analyzed

aled that th

aldehyde 2

0. To confir

sponding p

n mixture (S

aldehyde 28

- or 6-nitra

ined. HPLC

ds of 281 an

NO3 (60-w%)

rt, 5 min

222

dihydroxybe

in 5 mL rea

ments were

visually mo

out by feed

heated reac

hly resistant

ute solution

d by HPLC

he impurity

280. Impur

rm the resu

eaks of mo

Scheme 74)

81 was analy

ated isomer

C/MS expe

nd 282, but i

)

HO

HO

282: 25 (A)%281: 13% at2,6-dinitratio

enzaldehyde

actor

carried out

onitor both

ding stock

ctor coil. A

t to a wide

in AcOH to

. Temperatu

at 4.14 is

rities at 7

ults, an auth

ononitrated p

. In a simila

yzed and co

282 is the

eriments al

in minor am

NO2281

O

H

% on HPLC art 7.7; isome

on product (p

e 280

t with a tra

reaction in

solutions o

Although th

e range of

o minimize

ure, concen

the demeth

7.7 and 8

hentic samp

products we

ar manner, a

onfirmed as

impurity at

lso indicate

mounts.

HO

HO

2

+

analysis of rxeric 282 (putautative): 17%

ansparent 5

nitiation and

of vanillin a

he pump he

conditions

possible co

ntration of s

hylation pro

8.5 were i

ple of 280 w

ere observe

an authentic

the impurit

at 8.5, but t

ed the pres

282

O

H

NO2 (2- o

xn mixture atative): 12% a% at rt 12.0

5 mL HT-P

d possible b

and HNO3

eads in the

and chemi

orrosive effe

stock soluti

oduct of

isomeric

was also

ed via an

c sample

ty at 7.7,

he exact

sence of

or 6-)

rt 8.5at rt 9.4

PFA coil

blocking.

directly

Uniqsis

cals, we

ects. The

ions and

Page 109: Organocatalysis and Continuous Flow Reactors as Enabling Tools

107

residence time were varied during the experiments. The temperature was varied from 65 °C to

105 °C and residence time from 1 to 5 minutes. The concentration of stock solutions was

varied from 0.30 M to 0.55 M. Some variation in HNO3 stoichiometry was explored, but

mostly a small excess was enough to ensure full conversion. Thus, in a general experiment,

flow rates of both lines were kept equal, and the stoichiometry was determined by the

concentration of the stock solutions.

Chart 2 presents the results according to constant residence time and variable temperature.

The concentration of stock solutions was 0.30 M (271) and 0.32 M (HNO3), and the residence

time 3.33 minutes, equating to a combined flow rate of 1.50 mL/min. From the chart it may

be seen that the conversion of vanillin is enhanced at higher temperature from 98.3% to above

99%, and the amount of 272 is raised correspondingly. The main impurities are the

demethylation product 280, isomeric nitration-demethylation product 282, and 275. Curiously,

the amount of 280 is higher at higher temperatures, and correspondingly the amount of

nitration product 282 decreases.

Chart 2. Nitration of 271 in 5 mL coil at constant residence time of 3.33 min and varying temperature at stock solution concentrations of 0.30 M (271) and 0.32 M (HNO3).

Chart 3 shows the variation of the crude reaction mixture according to residence time at a

constant temperature of 70°C. The stock solution concentrations were 0.35 M 271 and 0.38 M

HNO3. The conversion of vanillin stays nearly constant at over 99.5% under these conditions,

0

1

2

3

4

5

6

7

8

87

88

89

90

91

92

93

70 80 90 100 110

Impurities A-%272 A-%

T(°C)

272

271

280

282

275

276

Page 110: Organocatalysis and Continuous Flow Reactors as Enabling Tools

108

and the longer residence time also seems to be beneficial for the selectivity of 272.

Correspondingly, the main impurity 282 is found in smaller amounts at longer residence times.

The reason for this remains unclear, but it may be due to the change of encounter velocity of

the reactant streams at the T-piece and subsequent diffusion-controlled mixing during the

flow through the reactor. Due to the integrated pressure sensor of the Uniqsis FlowSyn system,

we were not able to test differently shaped mixing pieces, as the pressure data was crucial for

monitoring the condition of the system.

Chart 3. Nitration of 271 at 70 °C and varying residence time at stock solution concentrations of 0.35 M (271) and 0.38 M (HNO3).

Further selected results from continuous experiments in 5 mL coil are presented in Table 13.

The best selectivity from a single experiment is shown in Entry 1, with stock solution

concentrations of 0.40 and 0.43 M. The conversion of 271 was 99.2% and selectivity of 272

94.2%, with 282 being the most significant impurity. From entries 1-4 it can be seen that

increasing the temperature too high may be detrimental to the selectivity. Entries 5-8 highlight

the stoichiometry of HNO3, with the surprising result that going from 1.3 equivalents to 1.7

equivalents at 70 °C has no significant effect on product distribution. The presence of 5% of

H2O in the stock solution of 271 has a slight detrimental effect on conversion at lower

temperatures, but at 90–100 °C the conversions and selectivities are once again at around 99.9%

and 92.2% (entries 9-12). When the reaction is run at 105 °C, shortening the residence time

once again has the effect of lowering selectivity (entries 13-16). Entries 17 and 18 are

performed at the same temperature and with the same residence time, but the reactor coil was

0

1

2

3

4

5

6

7

8

88

89

90

91

92

93

94

0 2 4 6

Impurities A-%272 A-%

Residence time (min)

272

271

280

281

282

275

Page 111: Organocatalysis and Continuous Flow Reactors as Enabling Tools

109

changed from a 5 mL HT-PFA coil to 10 mL stainless steel coil in the latter experiment. The

conversion and selectivity were improved upon changing to the longer coil. The flow rates are

doubled, which might be beneficial for mixing. Likewise, the thermal properties of the reactor

change significantly based on the material, steel being a much better heat conductor than

perfluoroalkoxy polymers. Indeed, the coil and back-pressure regulator were blocked almost

immediately after obtaining the HPLC sample due to rapid cooling of the reaction mixture

and subsequent precipitation of 272. The working concentrations were judged to be nearing

their practical limits within our reactor setup.

Table 13. Selected results from continuous reactions in 5 mL reactor coil

Entry c(271/M) tR (min) T(°C) 271 272 280 281 282 275 276

1 0.40 2 80 0.753 94.22 0.902 0.871 1.517 0.678 0.189

2 0.40 2 90 0.273 92.994 1.297 1.217 1.887 0.845 0.18

3 0.40 1.7 90 0.083 91.901 1.327 1.355 2.207 0.988 0.176

4 0.40 1.7 100 0.085 91.449 1.703 1.503 1.556 1.335 0.137

5 0.35a 2.9 70 20.409 73.853 0.948 n.d. 3.449 0.035 0.043

6 0.35b 2.5 70 3.3 89.438 0.927 n.d. 5.102 0.059 0.065

7 0.35c 2.2 70 0.765 91.792 1.036 n.d. 5.195 0.091 0.084

8 0.35d 2.5 70 0.117 91.953 1.099 n.d. 5.237 0.108 0.119

9 0.35 3.3 70 42.284 53.87 0.79 n.d. 2.198 0.064 0.04

10 0.35 3.3 80 11.955 81.766 0.964 n.d. 3.934 0.072 0.092

11 0.35 3.3 90 0.085 92.189 1.81 1.419 2.331 0.344 0.179

12 0.35 3.3 100 0.114 92.112 2.155 1.515 1.702 0.521 0.147

13 0.30 3.3 105 0.878 91.293 2.303 n.d. 1.946 0.893 0.081

14 0.30 2.5 105 1.119 90.594 2.249 n.d. 1.358 0.847 0.089

15 0.30 1.7 105 0.852 90.179 2.029 n.d. 2.805 0.862 0.119

16 0.30 1 105 0.931 89.818 1.928 n.d. 3.488 0.84 0.092

17 0.55 5 90 9.182 88.949 0.99 1.547 1.656 0.903 0.211

18 0.55 5 90 2.412 91.546 0.802 1.151 1.284 0.677 0.365 a) 0.80 equivalents of HNO3. b) 1.05 equivalents of HNO3. c) 1.30 equivalents of HNO3. d) 1.75 equivalents of HNO3.

Page 112: Organocatalysis and Continuous Flow Reactors as Enabling Tools

110

Experiments performed with stock solutions of 0.45 M 271 and 0.50 M HNO3 are shown in

Table 14. Within this parameter range (temperature 75-95 °C, residence time 1.5-3.5 minutes)

the reaction was functioning quite well in a robust fashion (entries 1-7). The conversion was

excellent in all cases, and selectivity of 272 was nearly constant, with parameter variation

leading mostly to changes in the composition of impurities – even with 2 equivalents of

HNO3 (entry 8). This was the optimal and most consistent series of experiments in the 5 mL

coil. At higher concentrations the conversion of 271 started to suffer, and in many cases full

conversion was not reached.

Table 14. Continuous experiments in 5 mL coil with 0.45 M and 0.50 M stock solutions

Entry tR (min) T(°C) 271 272 280 281 282 275 276

1 3.3 75 0.17 91.87 1.01 0.223 5.475 0.248 0.094

2 2 75 0.284 91.296 1.065 0.149 5.724 0.204 0.085

3 3.3 85 0.126 91.992 1.663 1.071 2.954 0.577 0.151

4 2 85 0.146 91.375 1.477 0.901 3.774 0.587 0.136

5 3.3 95 0.125 92.567 1.839 1.374 1.342 0.764 0.143

6 2 95 0.095 92.142 1.853 1.339 1.878 0.795 0.132

7 1.4 95 0.129 92.124 1.771 1.308 2.247 0.757 0.130

8a 2.2 95 0.156 92.132 0 0.629 0.317 1.89 0.765 a) 2 equivalents of HNO3 were used, and 2.189% of a previously disregarded impurity at Rrt 1.22 observed.

4.2.5 Continuous experiments in 10 mL reactor

Further experiments were carried out in the 10 mL stainless steel reactor coil. The

conversions and selectivities were consistently slightly improved upon this change of setup.

However, the precipitation of the product often became a serious problem at the end of the

coil due to rapid cooling of the long (about 10 cm) tail without heating. This problem was

partly solved by wrapping insulating cloth around the tail of the reactor coil, but as shown

below, the system could be kept running at stable conditions only for a maximum of about 40

minutes before significant pressure fluctuations started to occur due to accumulation of

precipitated material.

Page 113: Organocatalysis and Continuous Flow Reactors as Enabling Tools

111

In a fashion similar to continuous experiments in the 5 mL coil, the reaction was observed at

constant temperature and at constant residence time (Charts 4 & 5). When the reaction was

run at 90 °C at varying residence times, the conversion of 271 was consistently above 99%,

and selectivity of 272 varied from 92.4% to 93.8%, thus giving slightly better overall results.

All impurities remained below 1.5%. A residence time of 4 minutes gave the best selectivity

of 272 and the least amount of impurities.

Chart 4. Nitration of 271 in 10 mL coil at constant temperature at stock solution concentrations of 0.40 M (271) and 0.44 M (HNO3). Lines indicating 280 and 282 overlap each other.

Interestingly, when the residence time was kept constant and the temperature varied from

85 °C to 105 °C, we observed deterioration of conversion and selectivity in higher

temperatures. The optimum temperature was 90 °C. We were unable to determine whether

this effect of high temperature was an artifact, since correlations to both directions could be

seen between separate experiments (See Table 15, Entries 1-2, 3-4 and Chart 5).

0

1

2

3

4

5

90

91

92

93

94

95

2 3 4 5 6 7

Impurities A-%272 A-%

Residence time (min)

272

271

280

281

282

275

276

Page 114: Organocatalysis and Continuous Flow Reactors as Enabling Tools

112

Chart 5. Nitration of 271 in 10 mL coil at constant residence time at stock solution concentrations of 0.45 M (271) and 0.50 M (HNO3).

Table 15 presents selected results from other continuous experiments with 10 mL coil. The

best single run gave a maximum selectivity of 94.4% for 272, which was obtained with

concentrations of 0.50 M and 0.60 M of 271 and HNO3 respectively (entry 1). The residence

time was 3.3 minutes at 95 °C. Small variations of temperature and residence time did not

increase the selectivity (entries 2-4). Again we were close to the practical limits of our reactor

setup, and precipitation and pressure fluctuations were constantly occurring incidents.

Table 15. Selected results from continuous reactions in 10 mL coil

Entry c(271/M) tR (min) T(°C) 271 272 280 281 282 275 276

1 0.50 3.3 95 0.027 94.369 0.553 1.134 1.081 0.812 0.331

2 0.50 3.3 105 0.727 92.871 0.936 1.245 1.04 0.988 0.336

3 0.50 4 95 0.484 93.524 0.616 1.269 1.211 0.889 0.338

4 0.50 4 105 0 94.126 0.756 1.069 0.946 0.912 0.371

By following the reaction for a longer period of time, we determined that the reaction stays at

equilibrium with no significant fluctuations, as long as the pumps are performing smoothly

0

1

2

3

4

5

88

89

90

91

92

93

94

95

80 85 90 95 100 105 110

Impurities A-%272 A-%

T (°C)

272

271

280

281

282

275

276

Page 115: Organocatalysis and Continuous Flow Reactors as Enabling Tools

113

and no precipitation issues occur. A 30 minute run is shown in Chart 6. Stock solutions of

0.43 M (271) and 0.50 M (HNO3) were used with a 10 mL steel coil, at a temperature of

95 °C and residence time of 4 minutes. It can be seen that the conversion of 271, reaction

selectivity, and amount of impurities all stay practically constant during the run.

Chart 6. A continuous nitration equilibrium run

Pressure fluctuations and blockages due to accumulated precipitation into the back pressure

regulator and coil were a constant nuisance during most of the longer experiments performed

in the 10 mL reactor coil. The solution for this, however, lies more in engineering than

chemistry, as the main problem was the rapid dissipation of heat from the unheated part of the

steel coil.

4.2.6 Isolation studies and comparison between batch and flow modes

Due to limited time given to the project we did not carry out rigorous optimization studies for

the isolation of 272. However, some experiments were carried out to see whether the yield

and purity of the product would correspond to results from batch conditions. The isolation

was carried out by collecting the output stream into distilled H2O for a calculated time. The

mixture was then cooled to 0 °C for 1 hour in an ice bath, filtered, washed with cold toluene

and EtOH, dried and analyzed with HPLC. The yields and HPLC purities are presented in

0

1

2

3

4

5

85

86

87

88

89

90

91

92

93

0 5 10 15 20 25 30

Impurities A-%272 A-%

Time (min)

272

280

271

281

282

275

276

Page 116: Organocatalysis and Continuous Flow Reactors as Enabling Tools

114

Table 16. Varying the conditions slightly the yields were from 55% to 76% (entries 1-6). The

purity of the isolated material was overall slightly poorer than in batch conditions (on average

98.5%). The best purity (99.5%) was obtained when the output was collected into toluene

(entry 7), but the yield was significantly lower, only 46%. The fact that the yields are overall

lower than in batch conditions can be directly attributed to the greater amount of solvent used

in continuous flow conditions. When the concentration was raised, precipitation problems

occurred and disrupted the collection (entry 8). The longest collection times were 30 minutes,

the collection starting about 2-3 minutes after reaching equilibrium conditions (entries 9-10).

Although the experiments were ended at 30 minutes, the accumulated solids and remaining

material in the coil were already starting to cause significant precipitation, and pressure

fluctuations were observed by the time clear solvent was flowing through the whole coil.

Table 16. Isolation studies in continuous flow conditions

Entrya c(271/M) tR (min) T(°C) 271 272 281 275 276 Y Coll. t.

1 0.55 5 90 0.22 98.373 0.743 0.499 0.098 60% 13 min

2 0.55 5 95 0.231 97.894 0.954 0.712 0.12 55% 11 min

3 0.40 2.5 90 - 98.824 0.523 0.503 0.127 n.d. 5 min

4 0.45 4 95 0.058 98.706 0.566 0.495 0.15 76% 20 min

5 0.40 4 90 - 98.123 0.773b 0.791 0.135 69% 20 min

6 0.40 4 95 - 98.682 0.541 0.537 0.184 69% 20 min

7 0.40 4 95 - 99.49 0.313 0.098 0.099 46% 20 min

8 0.48 3.3 95 - 98.529 0.623 0.55 0.252 n.d. 13 min

9 0.43 4 95 - 98.384 0.671 0.739 0.065 69% 30 min

10 0.43 3.4 90 0.139 98.454 0.563 0.554 0.102 95% 30 min

a) All reactions were carried out in 10 mL coil except for entries 1-2, which were carried out in 5 mL coil; b) The isolated sample contained also 0.1% of isomeric 282.

Overall, the selectivity of nitration of vanillin was very similar in flow and batch conditions,

even though significantly higher reaction temperatures are used in flow conditions. In fact, the

conversion in flow conditions is usually better due to the fact that some vanillin tends to

cocrystallize with the mononitration product (See Table 10, entry 12). Especially in higher

concentrations this becomes more pronounced. No crystallization occurs in homogeneous

Page 117: Organocatalysis and Continuous Flow Reactors as Enabling Tools

115

flow conditions, and thus all vanillin is able to react with HNO3. For comparison, we carried

out batch reactions at our highest working flow concentrations at different temperatures

(Table 17). Again we osberved high selectivities and the same impurities as in flow reactions

(entries 1-3). The highest selectivity (95.6%) was obtained when a 0.60 M solution of HNO3

was added into a 0.52 M solution of 271 at 80 °C (entry 4). However, at this temperature the

yields remain below 60%, most likely due to complete digestion of some organic material by

HNO3. At higher temperature, the dissociation of HNO3 into nitric oxide was also observed

by the formation of red fumes. When the reaction was run at a usual batch concentration of

1.4 M 271 in AcOH the conversion remained surprisingly low (entry 5). A greater amount of

impurities was also formed at higher temperatures, and new impurities were likewise present

in both filtrate and the product, but were left unidentified.

Table 17. Batch reactions in analogous flow concentrations and higher temperatures. Note: the precipitate was analyzed directly after filtration without washing.

Entry Sample T (°C) 271 272 280 281 282 275 276 Yield (%)

1a Reaction 20 0.128 94,709 1,005 0,118 3,014 - 0,314

Filtrate - 75,584 5,105 0,978 12,081 0,186 1,553

Product 0,148 99,654 - 0,029 0,052 - 0,116 73

2 a Reaction 50 - 94,422 0,967 0,514 2,191 0,193 0,338

Filtrate - 77,762 3,991 2,801 3,971 1,251 1,406

Product 0,013 99,784 - 0,059 0,028 0,008 0,09 69

3 a Reaction 80 0,27 94,572 0,29 0,824 0,775 0,68 0,522

Filtrate 0,545 80,071 0,578 2,358 1,772 2,832 1,674

Product - 99,509 - 0,241 0,032 0,089 0,129 59

4b Reaction 80 0,07 95,576 0,141 0,278 0,192 0,418 0,361

Filtrate - 84,674 - 0,613 0,418 2,016 0,864

Product - 99,768 - 0,089 - 0,089 0,03 58

5 c Reaction 80 15,445 72,806 0,323 0,671 1,043 0,417 2,143

Filtrate 47,333 26,503 0,759 1,528 3,085 1,219 4,15

Product 0,28 97,306 - 0,122 - 0,023 0,892 49 a) c(271) = 0.26 M, HNO3 added neat. b) A 0.60 M solution of HNO3 in AcOH was added into a 0.52 M solution of 271 in AcOH. c) c(271) = 1.4 M, HNO3 added neat.

Page 118: Organocatalysis and Continuous Flow Reactors as Enabling Tools

116

4.3 Reaction profile and formation of impurities

In trying to determine the status of the side reactions (parallel vs. consecutive), we carried out

HPLC analyses over the course of a slow addition of HNO3 in batch conditions. The

conversion profile is presented in Chart 7. As stated before, the main nitration reaction is very

fast and as HNO3 was added in aliquots at regular intervals, the reaction proceeds

immediately, consuming the added HNO3. Thus the graphs are nearly linear, as would be

expected for a reaction of zero order kinetics, even though the experiment was not set up for

rigorous kinetic analysis. The background profile shows that the amount of 3,4-

dihydroxybenzaldehyde 280 stays nearly constant during the addition of HNO3, and the

amounts of mononitrated dihydroxybenzaldehyde isomers 281 and 282 increase.

Chart 7. Reaction conversion profile in batch. Initial concentration of 271 1.0 M in AcOH at 5 °C. 60 w% HNO3 added at a nearly constant rate over 67 minutes in small aliquots. The final vertical shift indicates a 60 minute delay between HPLC analyses after the addition of HNO3 was complete.

The side reaction to 281 can take place via either a nitration-demethylation, or a

demethylation-nitration pathway. To test this, we subjected 5-nitrovanillin to a stress test

under slightly forcing nitration conditions. In the presence of 1 equivalent of HNO3 at 55 °C

no significant demethylation or dinitration was observed during a digestion of 100 minutes

(Chart 8). When another equivalent of HNO3 was added, the amount of UV active material in

0

1

2

3

4

5

6

7

0102030405060708090

100

0 0,2 0,4 0,6 0,8 1 1,2

Impurities A-%271 & 272 A%

HNO3 (equiv)

271

272

280

281

282

275

Rrt 1,22

276

Page 119: Organocatalysis and Continuous Flow Reactors as Enabling Tools

117

the HPLC analysis decreased dramatically, presumably meaning complete oxidation of the

carbon backbone. These results lead us to conclude that the demethylation step takes place as

a parallel reaction pathway with the main nitration. The resulting 3,4-dihydroxybenzaldehyde

ring is more electron rich and nucleophilic than vanillin, and hence is nitrated faster under the

reaction conditions (see Scheme 77). The impurity at relative retention time 1.22 remained

unidentified during the studies, but the most likely candidate would be 5-nitrovanillic acid

based on these results, as nitric acid is able to oxidize aromatic aldehydes to carboxylic acids.

Chart 8. Stress test of 272 with 1 equivalent of HNO3 at 55 °C

The fact that demethylation takes place was rather surprising, considering the relatively mild

reaction conditions. Whereas aromatic ethers are usually dealkylated in harsh conditions or

with strong reagents such as BBr3, HBr, or AlCl3/pyridine, this mild demethylation is

previously unreported in conjunction with the nitration of vanillin. However, a survey of the

literature reveals that the dealkylation of phenolic ethers under nitration conditions has been

reported previously.136,139 The studies by Ingold established that the concomitant dealkylation

is closely associated with the reactive nitration species, be it nitrosonium or nitronium ion, the

former being much more active in the dealkylation process. Ingold et al. were able to isolate

methyl acetate in a nearly quantitative yield corresponding to the amount of demethylation in

the nitration of anisoles. Corresponding phenyl ethyl ethers did not yield more than trace

amounts of ethyl acetate, detectable by olfactory analysis but unisolable from the reaction

0

1

2

3

4

5

88

89

90

91

92

93

94

95

0 20 40 60 80 100 120

Impurities A-%272 A-%

Time (min)

272

281

Rrt 1,22

276

Page 120: Organocatalysis and Continuous Flow Reactors as Enabling Tools

118

mixture. Based on this they proposed that the methyl cation separates as the free cation and is

then attacked by acetic acid (or another available nucleophile in the mixture). They also

proposed that with ethyl ethers, significant oxidation takes place under the same conditions,

oxidizing the ethyl cation eventually to acetic acid, which is then indistinguishable from the

solvent. They ruled out an ordinary acid promoted demethylation process on the basis of a

model experiment in which no significant amounts of dealkylation took place in the presence

of an equimolar amount of H2SO4. Also, 2,6-disubstitution of anisoles effectively inhibited

the demethylation, presumably via steric effects. No kinetic studies have been carried out to

determine whether the demethylation is in fact a unimolecular process, but considering the

stability of primary cations, a bimolecular process would seem more likely (Scheme 75).

Scheme 75. Possible mechanisms for demethylation of vanillin

The ipso-substitution of aromatic compounds under nitration conditions has been reported on

several occasions.140 For example, the substitution of an isopropyl group with a nitro group

has been observed in the nitration of p-cymene.141 When an alkyl substituted carbon

undergoes ipso-attack, the usual outcomes may be a nucleophilic trapping of the cationic

intermediate, or a 1,2-migration of the nitro group, thus giving a formal ortho-nitration

product. The decarbonylative and decarboxylative nitration of electron rich substrates has also

been studied when using dilute nitric acid.142 Bentley reported the formation of 2,4-

dinitroguaiacol 276 from 5-nitrovanillin 272 when subjected to dilute nitration conditions,137c

but in our test experiment we had to use forcing conditions (several equivalents of HNO3,

high temperature) before we saw an increase in the amount of 276. As mentioned above, the

HPLC analysis also showed the disappearance of UV-active material at the same time,

ostensibly via complete oxidation by nitric acid. Hence, based on the actual calculated area,

the increase of 276 in the reaction mixture is rather modest, although clear. Hence we

Page 121: Organocatalysis and Continuous Flow Reactors as Enabling Tools

119

conclude the dominating mechanism for the formation to be first an ipso-substitution of

vanillin to give 275, then via a second nitration to give 276 (Scheme 76).

�Scheme 76. Competing reactions in the nitration of vanillin. Species 275, 276, 280, 281, 282, 287 and 288 were observed experimentally by HPLC or MS.

4.4. Nitration of other substrates

The high-throughput continuous-flow nitration of aromatic compounds has been presented

recently.143 It is no surprise that these publications come from pharmaceutical companies,

since a safe method of nitration using hazardous reagents is of significant interest and

importance. For example, researchers at Novartis performed the nitration of bromo-

quinolinone 289 in good yield and at high production rate (Scheme 77).143a The starting

material was pumped through the reactor as 1.0 M solution in AcOH, and through the other

Page 122: Organocatalysis and Continuous Flow Reactors as Enabling Tools

120

pump was fed fuming HNO3. The mixture was pumped through four consecutive 10 mL

reactor coils which were heated to 90 °C, and then cooled down and quenched by collecting

into H2O. The overall output was 7 mmol/min, corresponding to a production rate of 2.3 kg

per day.

Scheme 77. Continuous-flow nitration of bromoquinolinone 289 by Novartis

We carried out a brief investigation into the applicability of our relatively mild nitration

procedure to other substrates. In general, all procedures were carried out by pumping stock

solutions of starting material and 65 w-% aqueous HNO3 through a heated coil reactor.

However, in all cases the high dilution also hinders the work-up process, and it was difficult

to obtain good yields via either extraction or precipitation. The best results were achieved

with methyl 4-hydroxybenzoate 291. The mononitration product 292 could be isolated by

simple filtration from H2O in 75–85% yield and good purity. The limiting factor was the

solubility of methyl 4-hydroxybenzoate into acetic acid. In a large scale example run, the

reaction was run at steady state for a period of 2.5 hours, producing 18 grams of material

(Scheme 78).

Scheme 78. Continuous-flow nitration of methyl 4-hydroxybenzoate 291

Page 123: Organocatalysis and Continuous Flow Reactors as Enabling Tools

121

With other substrates we were not so successful. With salicylaldehyde we obtained a

maximum yield of only 35% in similar conditions, with an expected mixture of regioisomers.

Anisaldehyde, 4-methoxyacetophenone and 4-methylacetophenone did not react under similar

conditions, and would obviously need more forcing conditions, e.g. fuming nitric acid or

mixed acid as the nitration reagent. While this is possible and precedented, as shown in

Scheme 77, we decided not to explore the subject further as it would stray away from our

original aim of relatively mild nitration in continuous flow conditions.

4.5 Conclusions

We have investigated the nitration of vanillin 271 under continuous flow conditions and

compared the performance of the reactor to standard batch conditions. Overall it was noted

that a highly exothermic nitration could be safely carried out in a continuous flow reactor with

no concerns of a runaway reaction. Moreover, the reaction optimization can be carried out

rapidly, providing that a fast and robust analysis method is at hand. The nitration of 271 is

basically a reactive crystallization, the product 272 precipitating out from the reaction mixture

very early on. This forces severe limitations on the applicability of the reaction into micro-

and mesofluidic conditions. The concentrations need to be lower and temperatures high

enough to ensure that the product does not prematurely precipitate and block the reaction line.

As an upside, a higher conversion can be obtained in the nitration of 271 since it will not

cocrystallize with the product. The isolation of 272 from flow conditions was left unoptimized.

As the main side reaction we identified the O-demethylation of vanillin and subsequent

nitration of the catecholic aldehyde into different regioisomers. Another unpreventable side

reaction arises from the ipso-substitution of the vanillin formyl group, and subsequent

nitration of the still active intermediate 4-nitroguaiacol. The formation of corresponding

carboxylic acids was also detected by LC/MS-experiments, but these were present in

significantly smaller proportions.

Since the usual benefits of continuous processing do not arise from converting a single

nitration step from batch to flow, it should be considered that if vanillin was used as a starting

material in a multistep process, the telescoping of following step(s) into a continuous process

should also be investigated. This could then lead to saving of time, materials and reactor

equipment since several unit operations could be skipped entirely.

Page 124: Organocatalysis and Continuous Flow Reactors as Enabling Tools

122

The continuous flow nitration with a stoichiometric amount of dilute nitric acid is mostly

applicable to highly reactive substrates, as shown by our brief substrate studies. The high

throughput continuous nitration of less reactive aromatic substrates has been illustrated by

several industrial groups, thus confirming that there is pressure towards rethinking and

steering industrial processes towards greener and safer avenues with continuous processing.

Page 125: Organocatalysis and Continuous Flow Reactors as Enabling Tools

123

5 EXPERIMENTAL SECTION

5.1 General Experimental Considerations

Moisture sensitive reactions were carried out under an argon atmosphere, and glassware was

dried with flame in vacuum or in an oven. Dry solvents (CH2Cl2, THF, Et2O, MeCN, toluene)

were obtained with MBraun MB-SPS 800 solvent drying system. Chlorobenzene was distilled

from P2O5 on 4 Å molecular sieves. DMF was distilled from ninhydrin/molecular sieves (4 Å).

Commercial reagents were used without further purification unless otherwise indicated.

Nonaqueous reagents were transferred under argon via syringe or cannula. Analytical TLC

was performed using SiO2 (silica gel 60, F254, 230-400 mesh, precoated aluminum sheets or

glass plates, Merck). TLC analyses were visualized with UV light, or with aqueous KMnO4,

ninhydrin or vanillin solutions. Flash chromatography was carried out on SiO2 (silica gel 60,

F254, 230-400 mesh, Merck). Continuous flow reactions were performed with Vapourtec

R2+/R4 or Uniqsis FlowSyn reactors.

The 1H and 13C NMR spectra were recorded with Bruker Avance DPX-400 spectrometer (1H

399.98 MHz; 13C 100.59 MHz). The chemical shifts are reported in ppm relative to TMS

internal standard (� = 0.00) or residual solvent signal (1H NMR: CDCl3 7.26, DMSO-d6 2.50,

MeCN-d3 1.94, MeOD-d4 3.31, CD2Cl2 5.32; 13C NMR: CDCl3 77.0, DMSO-d6 39.50,

MeOD-d4 49.0). Enantiomeric excesses were determined by HPLC using chiral stationary

phase columns with Waters 501 pump and Waters 2487 UV dual absorbance detector (Aalto

University), Agilent HP 1100 chromatograph (Cambridge), or Waters Alliance 2695 HPLC

system with PDA 996 detector (Fermion). Mass spectra described in Chapter 4 were recorded

with Waters 2960, PDA 996 + EMD 1000 apparatus. IR spectra were recorded with Perkin

Elmer One FTIR spectrometer. Optical rotations were measured with Perkin Elmer 343

polarimeter ( = 589 nm) using a 1 dm cuvette. Elemental analyses were performed with

Perkin Elmer 2400 Series II CHNS/O Analyzer. Melting points (uncorrected) were

determined in open capillaries using either a Stuart SMP3 (Aalto) or SRS OptiMelt apparatus

(Cambridge). HRMS were recorded with Waters Micromass LCT Premier (ESI) spectrometer.

Page 126: Organocatalysis and Continuous Flow Reactors as Enabling Tools

124

For compounds 243 and 250-255 described in section 5.3, CHN analyses had been recorded

previously in the Ley group, and thus were not redetermined, as all other spectral data were in

agreement. The syntheses of known compounds (such as 26a-l, 45, 79, 88, 90, 101, 237, 239,

265-267) were carried out according to published procedures and are not reported here. Due

to low enantiopurity of model reactions and organocatalyst screening, optical rotation was

measured only for stereoisomerically pure compounds such as catalysts and their

intermediates.

5.2 Asymmetric Organocatalytic Michael Additions

5.2.1 General Organocatalytic Screening Procedure

Nitroalkene (0.8 mmol, 100 mol %) was loaded in a small vial equipped with a magnetic

stirrer. Compound 78 (230 mg, 1.6 mmol, 200 mol %) was added, and the mixture was

dissolved in CH2Cl2 (0.5 mL). Catalyst 79 (50 mg, 0.08 mmol, 10 mol %) was added, the vial

was tightly capped and the mixture was stirred at room temperature until completion as shown

by TLC. The solvents were evaporated and the crude mixture analyzed by 1H NMR and used

as such in the derivatization step.

5.2.2 5-(1-(3-Methoxyphenyl)-2-nitroethyl)-2,2-dimethyl-1,3-dioxane-4,6-dione (80e)

Compound 80e was prepared according to the general procedure. Purification of a small

sample for analytical purposes by flash chromatography (10% MeOH in EtOAc) gave 80e as

a yellowish oil. Rf = 0.21 (EtOAc); IR (neat): 2997, 2941, 1734, 1572, 1403, 1375, 1261 cm-1; 1H NMR (MeOD-d4, 400 MHz): (Partial, dicarbonyl H presumably exchanges to D) � 7.12

(app t, J = 7.9 Hz, 1H), 7.03 (s, 1H), 6.98 (d, J = 7.6 Hz, 1H), 6.69 (dd, J = 8.1, 2.0 Hz, 1H),

Page 127: Organocatalysis and Continuous Flow Reactors as Enabling Tools

125

5.21 (dd, J = 12.0, 8.7 Hz, 1H), 5.07 (dd, J = 11.9, 7.5 Hz, 1H), 4.74 (t, J = 8.1 Hz, 1H), 3.74

(s, 3H), 1.54 (s, 6H); 13C NMR (MeOD-d4, 100 MHz): � 169.2, 160.9, 145.3, 129.9, 121.2,

114.4, 112.8, 102.8, 79.2, 76.5, 55.5, 41.3, 25.8; HRMS m/z 346.0906 [C15H17NO7 (M+Na)+

requires 346.0903].

5.2.3 5-(1-(4-Bromophenyl)-2-nitroethyl)-2,2-dimethyl-1,3-dioxane-4,6-dione (80f)

Compound 80f was prepared according to the general procedure. Purification of a small

sample for analytical purposes by flash chromatography (10% MeOH in EtOAc) gave 80f as a

yellow oil. Rf = 0.26 (50% EtOAc in hexane); IR (neat): 2994, 1716, 1579, 1488, 1399, 1374,

1262 cm-1; 1H NMR (MeOD-d4, 400 MHz): (Partial, dicarbonyl H presumably exchanges to

D) � 7.35 (s, 4H), 5.17 (dd, J = 12.1, 8.6 Hz, 1H), 5.09 (dd, J = 12.1, 7.7 Hz, 1H), 4.73 (dd, J

= 8.4, 8.0 Hz, 1H), 1.53 (s, 6H); 13C NMR (MeOD-d4, 100 MHz): � 169.1, 161.2, 143.1,

132.0, 130.9, 120.7, 102.9, 78.8, 76.3, 61.5, 40.9, 25.8; HRMS m/z 415.9742/417.9731

[C14H14BrNO6 (M-H+2Na)+ requires 415.9722/417.9702].

5.2.4 5-(1-(3-Bromophenyl)-2-nitroethyl)-2,2-dimethyl-1,3-dioxane-4,6-dione (80g)

Page 128: Organocatalysis and Continuous Flow Reactors as Enabling Tools

126

Compound 80g was prepared according to the general procedure. Purification of a small

sample for analytical purposes by flash chromatography (10% MeOH in EtOAc) gave 80g as

a yellow oil. Rf = 0.19 (50% EtOAc in hexane); IR (neat): 2994, 2941, 1719, 1578, 1402,

1375, 1262 cm-1; 1H NMR (MeOD-d4, 400 MHz): (Partial, dicarbonyl H presumably

exchanges to D) � 7.61 (t, J = 1.8 Hz, 1H), 7.39 (d, J = 7.7 Hz, 1H), 7.28 (ddd, J = 8.0, 1.9,

1.0 Hz, 1H), 7.14 (t, J = 7.8 Hz, 1H), 5.19 (dd, J = 12.2, 8.6 Hz, 1H), 5.10 (dd, J = 12.1, 7.7

Hz, 1H), 4.74 (t, J = 8.1 Hz, 1H), 1.54 (s, 6H); 13C NMR (MeOD-d4, 100 MHz): � 169.1,

146.4, 131.9, 130.7, 130.2, 127.7, 123.0, 102.9, 78.8, 76.1, 41.1, 25.8; HRMS m/z

415.9731/417.9722 [C14H14BrNO6 (M-H+2Na)+ requires 415.9722/417.9702].

5.2.5 5-(1-(Furan-2-yl)-2-nitroethyl)-2,2-dimethyl-1,3-dioxane-4,6-dione (80h)

Compound 80h was prepared according to the general procedure. Purification of a small

sample for analytical purposes by flash chromatography (10% MeOH in EtOAc) gave 80h as

a yellow oil. Rf = 0.27 (50% EtOAc in hexane); IR (neat): 2995, 2943, 1731, 1578, 1411,

1375, 1262 cm-1; 1H NMR (MeOD-d4, 400 MHz): (Partial, dicarbonyl H presumably

exchanges to D) � 7.31 (m, 1H), 6.26 (dd, J = 3.2, 1.9 Hz, 1H), 6.08 (dt, J = 3.1, 0.9 Hz, 1H),

5.06 (dd, J = 11.9, 8.8 Hz, 1H), 5.02 (dd, J = 11.8, 6.9 Hz, 1H), 4.89-4.84 (m, 1H), 1.58 (s,

6H); 13C NMR (MeOD-d4, 100 MHz): � 169.0, 161.2, 156.3, 141.9, 111.1, 106.5, 102.9, 77.4,

73.8, 35.9, 25.8; HRMS m/z 306.0582 [C12H13NO7 (M+Na)+ requires 306.0590].

Page 129: Organocatalysis and Continuous Flow Reactors as Enabling Tools

127

5.2.6 5-(1-(Furan-3-yl)-2-nitroethyl)-2,2-dimethyl-1,3-dioxane-4,6-dione (80i)

Compound 80i was prepared according to the general procedure. Purification of a small

sample for analytical purposes by flash chromatography (10% MeOH in EtOAc) gave 80i as a

yellow oil. Rf = 0.31 (50% EtOAc in hexane); IR (neat): 2994, 1737, 1578, 1555, 1377, 1261

cm-1; 1H NMR (MeOD-d4, 400 MHz): (Partial, dicarbonyl H presumably exchanges to D) �

7.32 (m, 2H), 6.46 (s, 1H), 5.05 (dd, J = 11.6, 8.7 Hz, 1H), 4.94 (dd, J = 11.6, 7.6 Hz, 1H),

4.66 (t, J = 8.1 Hz, 1H), 1.55 (s, 6H); 13C NMR (MeOD-d4, 100 MHz): � 169.0, 161.2, 143.3,

140.5, 126.9, 111.7, 102.8, 79.1, 75.6, 33.0, 25.8; HRMS m/z 306.0601 [C12H13NO7 (M+Na)+

requires 306.0590].

5.2.7 5-(1-Cyclohexyl-2-nitroethyl)-2,2-dimethyl-1,3-dioxane-4,6-dione (80j)

Compound 80j was prepared according to the general procedure. Purification of a small

sample for analytical purposes by flash chromatography (10% MeOH in EtOAc) gave 80j as a

clear oil. Rf = 0.13 (50% EtOAc in hexane); IR (neat): 2994, 2926, 2852, 1731, 1575, 1406,

1373, 1260 cm-1; 1H NMR (MeOD-d4, 400 MHz): (Partial, dicarbonyl H presumably

exchanges to D) � 4.90 (app. t, J = 10.8 Hz, 1H), 4.62 (dd, J = 10.8, 5.0 Hz, 1H), 3.25-3.17 (m,

1H), 1.84-1.66 (m, 5H), 1.65-1.60 (m, 1H), 1.55 (s, 6H), 1.31-1.10 (m, 3H), 1.10-0.90 (m,

Page 130: Organocatalysis and Continuous Flow Reactors as Enabling Tools

128

2H); 13C NMR (MeOD-d4, 100 MHz): � 169.5, 102.6, 78.2, 74.0, 42.6, 39.9, 32.9, 32.1, 27.6,

27.6, 27.5, 25.9; HRMS m/z 322.1254 [C14H21NO6 (M+Na)+ requires 322.1266].

5.2.8 2,2-Dimethyl-5-(1-nitrohexan-2-yl)-1,3-dioxane-4,6-dione (80k)

O O

O O78

+NO2

26k

79 (cat.) NO2

80k

OO

O O

CH2Cl2, rt

Compound 80k was prepared according to the general procedure. Purification of a small

sample for analytical purposes by flash chromatography (10% MeOH in EtOAc) gave 80k as

a clear oil. Rf = 0.26 (50% EtOAc in hexane); IR (neat): 2956, 2930, 2863, 1782, 1741, 1718,

1553, 1383, 1311, 1206 cm-1; 1H NMR (CDCl3, 400 MHz): � 4.98 (dd, J = 13.3, 10.2 Hz, 1H),

4.57 (dd, J = 13.4, 4.4 Hz, 1H), 3.89 (d, 2.4 Hz, 1H), 3.31-3.22 (m, 1H), 1.81 (s, 3H), 1.79 (s,

3H), 1.64-1.53 (m, 1H), 1.51-1.41 (m, 1H), 1.40-1.27 (m, 4H), 0.94-0.87 (m, 3H); 13C NMR

(CDCl3, 100 MHz): � 164.0, 105.5, 75.8, 46.9, 36.4, 29.5, 29.0, 28.2, 26.9, 22.4, 13.8; HRMS

m/z 296.1106 [C12H19NO6 (M+Na)+ requires 296.1110].

5.2.9 5-(1-(Benzyloxy)-3-nitropropan-2-yl)-2,2-dimethyl-1,3-dioxane-4,6-dione (80l)

Compound 80l was prepared according to the general procedure. Purification of a small

sample for analytical purposes by flash chromatography (10% MeOH in EtOAc) gave 80l as a

yellowish oil. Rf = 0.14 (50% EtOAc in hexane); IR (neat): 2865, 1709, 1579, 1551, 1412,

1376, 1260 cm-1; 1H NMR (MeOD-d4, 400 MHz): (Partial, dicarbonyl H presumably

exchanges to D) � 7.35-7.21 (m, 5H), 4.93 (dd, J = 11.3, 9.2 Hz, 1H), 4.74 (dd, J = 11.3, 5.5

Hz, 1H), 4.50 (q, J = 11.6 Hz, 2H), 3.91-3.76 (m, 2H, overlapping signals), 3.55 (dd, J = 9.1,

Page 131: Organocatalysis and Continuous Flow Reactors as Enabling Tools

129

4.8 Hz, 1H), 1.55 (s, 6H); 13C NMR (MeOD-d4, 100 MHz): � 169.3, 139.9, 129.3, 128.8,

128.5, 102.9, 77.4, 73.7, 72.5, 71.5, 37.1, 25.7; HRMS m/z 360.1064 [C16H19NO7 (M+Na)+

requires 360.1059].

5.2.10 General Procedure for Derivatization

The evaporated crude reaction mixture (100 mg) was loaded in a 5 mL flask and dissolved in

dry DMF (2.5 mL). Distilled aniline (0.25 mL) was added and the mixture was heated to

100 °C for 3 hours. The mixture was partitioned between 10 mL EtOAc and 10 mL 1 M HCl,

and washed further 3 times with 1 M HCl (3 * 10 mL). The organic phase was dried over

MgSO4, filtered and evaporated. The residue was purified by flash chromatography (EtOAc in

hexane). Enantioselectivity was determined by chiral HPLC from the chromatographed

product. For determination of elemental composition, the product was then crystallized from

boiling hexanes by adding a minimum amount of EtOAc for dissolving the product. The yield

was measured from the crystallized product, thus accounting for the low values.

5.2.11 4-Nitro-N,3-diphenylbutanamide (87a)

Compound 87a was prepared according to the general procedure, purified by flash

chromatography (50% EtOAc in hexane) and crystallized from EtOAc:hexane to give 80a as

white crystals (10 mg): Rf = 0.45 (50% EtOAc in hexane); mp. 107-109 °C; IR (neat): 3302,

3138, 3063, 3032, 1661, 1599, 1551, 1498, 1443 cm-1; 1H NMR (CDCl3, 400 MHz): � 7.40-

7.23 (m, 9H), 7.15-7.06 (m, 2H), 4.85 (dd, J = 12.6, 6.4 Hz, 1H), 4.73 (dd, J = 12.6, 7.9 Hz,

1H), 4.07 (app d, J = 7.1 Hz, 1H), 2.83 (dd, J = 15.0, 7.3 Hz, 1H), 2.75 (dd, J = 15.0, 7.1 Hz,

1H); 13C NMR (CDCl3, 100 MHz): � 167.9, 138.4, 137.2, 129.2, 129.0, 128.1, 127.3, 124.7,

120.1, 79.3, 40.7, 40.6; HPLC: Chiralcel OD column (0.46 cm 25 cm), 25% i-PrOH in

Page 132: Organocatalysis and Continuous Flow Reactors as Enabling Tools

130

hexane, 1.0 mL/min, = 254 nm, tR(major) 17.9 min, tR(minor) 23.8 min; Anal. Calcd. for

C16H16N2O3: C, 67.59; H, 5.67; N, 9.85. Found: C, 67.50; H, 5.74; N, 9.85.

5.2.12 3-(4-Methoxyphenyl)-4-nitro-N-phenylbutanamide (87b)

NO2

80b

OO

O O 1:10 aniline:DMF

100 °C NH

ONO2

87bMeO

OMe

Compound 87b was prepared according to the general procedure. Purified by flash

chromatography (50% EtOAc in hexane) and crystallized from EtOAc:hexane to give 87b as

white crystals (12 mg). Rf = 0.38 (50% EtOAc in hexane); mp. 97-99 °C; IR (neat): 3309,

3199, 3138, 3061, 3038, 2969, 2837, 1664, 1600, 1551, 1514, 1499, 1443, 1379, 1303, 1251

cm-1; 1H NMR (CDCl3, 400 MHz): � 7.37 (d, J = 7.6 Hz, 2H), 7.28 (t, J = 7.9 Hz, 2H), 7.19

(br s, 1H), 7.16 (d, J = 8.5 Hz, 2H), 7.10 (t, J = 7.4 Hz, 1H), 6.85 (d, J = 8.7 Hz, 2H), 4.80 (dd,

J = 12.5, 6.5 Hz, 1H), 4.67 (dd, J = 12.5, 7.9 Hz, 1H), 4.06-3.97 (m, 1H), 3.77 (s, 3H), 2.79

(dd, J = 15.0, 7.2 Hz, 1H), 2.71 (dd, J = 15.0, 7.2 Hz, 1H), 13C NMR (CDCl3, 100 MHz): �

168.0, 159.2, 137.2, 130.2, 129.0, 128.4, 124.7, 120.1, 114.5, 79.6, 55.2, 40.8, 39.9; HPLC:

Chiralcel OD column (0.46 cm 25 cm), 25% i-PrOH in hexane, 1.0 mL/min, = 254 nm,

tR(major) 26.1 min, tR(minor) 46.7 min; Anal. Calcd. for C17H18N2O4: C, 64.96; H, 5.77; N,

8.91. Found: C, 64.81; H, 5.63; N, 8.86.

5.2.13 3-(3-Methoxyphenyl)-4-nitro-N-phenylbutanamide (87e)

NO2

80e

OO

O O 1:10 aniline:DMF

100 °C NH

ONO2

87e

MeOOMe

Page 133: Organocatalysis and Continuous Flow Reactors as Enabling Tools

131

Compound 87e was prepared according to the general procedure. Purified by flash

chromatography (50% EtOAc in hexane) and crystallized from EtOAc:hexane to give 87e as

white crystals (16 mg). Rf = 0.39 (50% EtOAc in hexane); mp. 101-103 °C; IR (neat): 3303,

3139, 3059, 2919, 2837, 1661, 1600, 1551, 1498, 1443, 1378, 1260 cm-1; 1H NMR (CDCl3,

400 MHz): � 7.41-7.34 (m, 2H), 7.32-7.23 (m, 3H), 7.14-7.07 (m, 2H), 6.86-6.76 (m, 3H),

4.84 (dd, J = 12.6, 6.4 Hz, 1H), 4.72 (dd, J = 12.6, 7.9 Hz, 1H), 4.09-4.00 (m, 1H), 3.77 (s,

3H), 2.81 (dd, J = 15.1, 7.2 Hz, 1H), 2.75 (dd, J = 15.0, 7.0 Hz); 13C NMR (CDCl3, 100 MHz):

� 167.8, 160.1, 140.0, 137.2, 130.3, 129.0, 124.7, 120.0, 119.4, 113.5, 113.2, 79.2, 55.2, 40.7,

40.6; HPLC: Chiralcel OD column (0.46 cm 25 cm), 25% i-PrOH in hexane, 1.0 mL/min,

= 254 nm, tR(major) 16.6 min, tR(minor) 21.3 min; Anal. Calcd. for C17H18N2O4: C, 64.96; H,

5.77; N, 8.91. Found: C, 64.80; H, 5.50; N, 8.88.

5.2.14 3-(4-Bromophenyl)-4-nitro-N-phenylbutanamide (87f)

Compound 87f was prepared according to the general procedure. Purified by flash

chromatography (40% EtOAc in hexane) and crystallized from EtOH:H2O to give 87f as

white crystals (10 mg). Rf = 0.42 (50% EtOAc in hexane); mp. 127-130 °C; IR (neat): 3302,

3139, 3062, 2918, 1662, 1599, 1551, 1498, 1443, 1377 cm-1; 1H NMR (CDCl3, 400 MHz): �

7.47 (d, J = 8.4 Hz, 2H), 7.39 (d, J = 7.7 Hz, 2H), 7.31 (t, J = 7.9 Hz, 2H), 7.18-7.08 (m, 4H),

4.84 (dd, J = 12.8, 6.3 Hz, 1H), 4.71 (dd, J = 12.8, 8.0 Hz, 1H), 4.11-4.01 (m, 1H), 2.82 (dd, J

= 15.3, 7.3 Hz, 1H), 2.73 (dd, J = 15.3, 6.9 Hz, 1H); 13C NMR (CDCl3, 100 MHz): � 167.5,

137.5, 137.1, 132.3, 129.1, 124.9, 122.1, 120.1, 79.0, 40.3, 39.9; HPLC: Chiralcel OD column

(0.46 cm 25 cm), 25% i-PrOH in hexane, 1.0 mL/min, = 254 nm, tR(major) 14.5 min,

tR(minor) 17.7 min; Anal. Calcd. for C16H15BrN2O3: C, 52.91; H, 4.16; N, 7.71. Found: C,

52.93; H, 3.84; N, 7.55.

Page 134: Organocatalysis and Continuous Flow Reactors as Enabling Tools

132

5.2.15 3-(3-Bromophenyl)-4-nitro-N-phenylbutanamide (87g)

Compound 87g was prepared according to the general procedure. Purified by flash

chromatography (40% EtOAc in hexane) and crystallized from EtOAc:hexane to give 87g as

white crystals (13 mg). Rf = 0.45 (50% EtOAc in hexane); mp. 110-111 °C; IR (neat): 3302,

3199, 3139, 3061, 2918, 1661, 1598, 1551, 1498, 1443, 1376 cm-1; 1H NMR (CDCl3, 400

MHz): � 7.45-7.36 (m, 4H), 7.30 (t, J = 7.9 Hz, 2H), 7.24-7.16 (m, 3H), 7.12 (t, J = 7.3 Hz,

1H), 4.84 (dd, J = 12.8, 6.3 Hz, 1H), 4.71 (dd, J = 12.8, 8.0 Hz, 1H), 4.10-4.01 (m, 1H), 2.81

(dd, J = 15.4, 7.5 Hz, 1H), 2.73 (dd, J = 15.4, 6.9 Hz, 1H); 13C NMR (CDCl3, 100 MHz): �

167.5, 140.8, 137.1, 131.3, 130.7, 130.4, 129.1, 126.2, 124.9, 123.1, 120.1, 78.9, 40.2, 40.0;

HPLC: Chiralcel OD column (0.46 cm 25 cm), 25% i-PrOH in hexane, 1.0 mL/min, =

254 nm, tR(major) 16.3 min, tR(minor) 18.8 min; Anal. Calcd. for C16H15BrN2O3: C, 52.91; H,

4.16; N, 7.71. Found: C, 52.99; H, 3.82; N, 7.71.

5.2.16 3-(Furan-2-yl)-4-nitro-N-phenylbutanamide (87h)

Compound 87h was prepared according to the general procedure. Purified by flash

chromatography (40% EtOAc in hexane) and crystallized from EtOAc:hexane to give 87h as

white crystals (10 mg). Rf = 0.48 (50% EtOAc in hexane); mp. 111-124 °C; IR (neat): 3304,

3141, 2918, 1662, 1600, 1552, 1499, 1444, 1376 cm-1; 1H NMR (CDCl3, 400 MHz): � 7.43 (d,

J = 8.1 Hz, 2H), 7.37 (d, J = 1.1 Hz, 1H), 7.31 (t, J =7.9 Hz, 2H), 7.22 (br. s, 1H), 7.12 (t, J =

Page 135: Organocatalysis and Continuous Flow Reactors as Enabling Tools

133

7.3 Hz, 1H), 6.31 (dd, J = 3.1, 1.8 Hz, 1H), 6.21 (d, J = 3.3 Hz, 1H), 4.84 (dd, J = 12.6, 5.9

Hz, 1H), 4.76 (dd, J = 12.6, 7.1 Hz, 1H), 4.23-4.15 (m, 1H), 2.87 (dd, J = 15.0, 6.6 Hz, 1H),

2.82 (dd, J = 15.1, 7.4 Hz, 1H), 13C NMR (CDCl3, 100 MHz): � 167.6, 151.3, 142.5, 137.2,

129.1, 124.8, 120.1, 110.6, 107.5, 77.1, 38.1, 34.5; HPLC: Chiralcel OD column (0.46 cm

25 cm), 25% i-PrOH in hexane, 1.0 mL/min, = 254 nm, tR(major) 12.0 min, tR(minor) 15.1

min; Anal. Calcd. for C14H14N2O4: C, 61.31; H, 5.14; N, 10.21. Found: C, 61.26; H, 4.98; N,

10.16.

5.2.17 3-(Furan-3-yl)-4-nitro-N-phenylbutanamide (87i)

Compound 87i was prepared according to the general procedure. Purified by flash

chromatography (40% EtOAc in hexane) and crystallized from EtOAc:hexane to give 87i as

white crystals (10 mg). Rf = 0.42 (50% EtOAc in hexane); mp. 101-107 °C; IR (neat): 3302,

3136, 2918, 1664, 1599, 1551, 1499, 1443, 1377 cm-1; 1H NMR (CDCl3, 400 MHz): � 7.44 (d,

J = 7.7 Hz, 2H), 7.40 (t, J = 1.6 Hz, 1H), 7.37 (br s, 1H), 7.32 (app t, J = 7.9 Hz, 2H), 7.19 (br

s, 1H), 7.12 (app t, J = 7.4 Hz, 1H), 6.34 (br s., 1H), 4.80 (dd, J = 12.5, 6.0 Hz, 1H), 4.67 (dd,

J = 12.5, 7.3 Hz, 1H), 4.02 (app d, J = 6.8 Hz, 1H), 2.79 (dd, J = 15.2, 7.3 Hz, 1H), 2.72 (dd,

J = 15.1, 6.7 Hz, 1H); 13C NMR (CDCl3, 100 MHz): � 167.8, 143.9, 139.8, 137.2, 129.1,

124.8, 122.7, 120.0, 108.8, 78.8, 40.1, 31.8; HPLC: Chiralcel OD column (0.46 cm 25 cm),

25% i-PrOH in hexane, 1.0 mL/min, = 254 nm, tR(major) 13.0 min, tR(minor) 16.0 min;

Anal. Calcd. for C14H14N2O4: C, 61.31; H, 5.14; N, 10.21. Found: C, 61.07; H, 5.10; N, 10.18.

Page 136: Organocatalysis and Continuous Flow Reactors as Enabling Tools

134

5.2.18 3-Cyclohexyl-4-nitro-N-phenylbutanamide (87j)

Compound 87j was prepared according to the general procedure. Purified by flash

chromatography (0 to 15% EtOAc in hexane) and recrystallization to give traces of 87j as

white crystals. Rf = 0.61 (50% EtOAc in hexane); IR (neat): 3302, 3139, 3061, 2958, 2931,

2861, 1661, 1600, 1548, 1499, 1443, 1380 cm-1; 1H NMR (CDCl3, 400 MHz): � 7.50 (d, J =

7.7 Hz, 2H), 7.33 (t, J = 7.9 Hz, 2H), 7.28-7.23 (br. s, 1H), 4.62 (dd, J = 12.4, 6.3 Hz, 1H),

4.53 (dd, J = 12.3, 5.1 Hz, 1H), 2.65-2.56 (m, 2H), 2.43 (dd, J = 16.5, 9.5 Hz, 1H), 1.85-1.73

(m, 4H), 1.72-1.65 (m, 1H), 1.55-1.46 (m, 1H), 1.32-1.19 (m, 2H), 1.19-1.10 (m, 1H), 1.09-

0.96 (m, 2H); 13C NMR (CDCl3, 100 MHz): � 169.2, 137.5, 129.1, 124.6, 119.9, 76.8, 40.0,

38.9, 36.5, 30.3, 29.9, 26.2, 26.2; HPLC: Chiralcel OD column (0.46 cm 25 cm), 10% i-

PrOH in hexane, 0.7 mL/min, = 254 nm, tR(major) 22.9 min, tR(minor) 26.8 min; HRMS

m/z 313.1540 [C16H22N2O3 (M+Na)+ requires 313.1528].

5.2.19 3-(Nitromethyl)-N-phenylheptanamide (87k)

Compound 87k was prepared according to the general procedure. Purified by flash

chromatography (0 to 15% EtOAc in hexane) and crystallized from EtOAc:hexane to give

87k as white crystals (9 mg). Rf = 0.59 (50% EtOAc in hexane); mp. 75-76 °C; IR (neat):

3307, 3200, 3139, 3061, 2926, 2854, 1711, 1661, 1600, 1549, 1499, 1443, 1379 cm-1; 1H

NMR (CDCl3, 400 MHz): � 7.51 (d, J = 7.7 Hz, 2H), 7.36-7.28 (m, 3H), 7.13 (t, J = 7.4 Hz,

Page 137: Organocatalysis and Continuous Flow Reactors as Enabling Tools

135

1H), 4.60 (dd, J = 12.1, 5.9 Hz, 1H), 4.53 (dd, J = 12.0, 5.0 Hz, 1H), 2.74-2.63 (m, 1H), 2.52

(dd, J = 15.2, 5.9 Hz, 1H), 2.48 (dd, J = 15.2, 7.7 Hz, 1H), 1.53-1.45 (m, 2H), 1.42-1.29 (m,

4H), 0.91 (t, J = 7.0 Hz, 3H); 13C NMR (CDCl3, 100 MHz): (partial, anilide quaternary C not

seen) � 168.8, 129.1, 124.6, 119.9, 78.4, 39.1, 34.9, 31.1, 28.8, 22.5, 13.9; HPLC: Chiralcel

OD column (0.46 cm 25 cm), 15% i-PrOH in hexane, 1.0 mL/min, = 254 nm, tR(major)

18.6 min, tR(minor) 23.0 min; Anal. Calcd. for C14H20N2O3: C, 63.62; H, 7.63; N, 10.60.

Found: C, 63.59; H, 8.06; N, 10.60.

5.2.20 4-(Benzyloxy)-3-(nitromethyl)-N-phenylbutanamide (87l)

Compound 87l was prepared according to the general procedure. Purified by flash

chromatography (40% EtOAc in hexane) and crystallized from EtOH:H2O to give 87l as

white crystals (17 mg). Rf = 0.51 (50% EtOAc in hexane); mp. 82-85 °C; IR (neat): 3310,

3138, 3062, 3031, 2864, 1663, 1600, 1550, 1498, 1443, 1380 cm-1; 1H NMR (CDCl3, 400

MHz): � 7.43 (d, J = 7.7 Hz, 2H), 7.37-7.27 (m, 8H), 7.12 (t, J = 7.4 Hz, 1H), 4.61 (d, J = 6.4

Hz, 2H), 4.51 (dd, AB system, J = 13.5, 11.9 Hz, 2H), 3.60 (app d, J = 5.3 Hz, 2H), 3.10-3.00

(m, 1H), 2.57 (dd, J = 15.2, 6.8 Hz, 1H), 2.51 (dd, J = 15.2, 6.8 Hz, 1H); 13C NMR (CDCl3,

100 MHz): � 168.3, 137.5, 137.4, 129.0, 128.5, 128.0, 127.8, 124.6, 119.9, 76.3, 73.5, 69.4,

36.4, 35.3; HPLC: Chiralpak AS column (0.46 cm 25 cm), 10% i-PrOH in hexane, 1.0

mL/min, = 254 nm, tR(major) 75.7 min, tR(minor) 86.8 min.; Anal. Calcd. for C18H20N2O4:

C, 65.84; H, 6.14; N, 8.53. Found: C, 65.79; H, 6.35; N, 8.53.

Page 138: Organocatalysis and Continuous Flow Reactors as Enabling Tools

136

5.2.21 1-Cyclohexyl-3-((R)-(6-methoxyquinolin-4-yl)((2R,4S,5R)-5-vinylquinuclidin-2-

yl)methyl)thiourea (92)

To a solution of 90 (180 mg, 0.56 mmol, 100 mol %) in dry THF (2 mL) was added freshly

distilled cyclohexyl isothiocyanate (90 �L, 0.63 mmol, 110 mol %) at room temperature.

Mixture was stirred under Ar overnight. Solvents were evaporated, and purification by flash

chromatography (8% to 10% MeOH in CH2Cl2) gave 92 as white foam (124 mg, 48%). Rf =

0.30 (10% MeOH in CH2Cl2); [�]20D = +218.5 (c = 1.0, CHCl3); IR (neat): 3247 (broad),

3067, 2999, 2933, 2856, 2669, 2212, 1622, 1589, 1542, 1509, 1474, 1452, 1433, 1364, 1337,

1321, 1227 cm-1; 1H NMR (CDCl3, 400 MHz): � 8.75 (d, J = 4.6 Hz, 1H), 8.03 (d, J = 9.1 Hz,

1H), 7.63 (d, J = 1.8 Hz, 1H), 7.50-7.45 (br s, 1H), 7.45 (d, J = 4.6 Hz, 1H), 7.41 (dd, J = 9.2,

2.7 Hz, 1H), 5.89 (ddd, J = 17.2, 10.7, 6.3 Hz), 5.24-5.15 (m, 2H), 3.99 (s, 3H), 3.34-3.20 (m,

1H), 3.20-2.90 (m, 5H), 2.45-2.34 (m, 1H), 1.99-1.87 (m, 1H), 1.83-1.70 (m, 2H), 1.70-1.46

(m, 4H), 1.46-1.37 (m, 1H), 1.37-1.16 (m, 4H), 1.16-0.97 (m, 3H), 0.97-0.79 (m, 1H); 13C

NMR (CDCl3, 100 MHz): � 180.9, 158.1, 147.6, 144.8, 139.3, 131.8, 127.6, 122.4, 120.1,

115.4, 101.3, 61.1, 55.7, 53.3, 48.8, 46.8, 38.5, 32.5, 32.3, 27.1, 25.8, 25.3, 24.8, 24.3; HRMS

m/z 487.2510 [C27H36N4OS (M+H)+ requires 487.2508].

5.2.21 Ethyl (R)-(6-methoxyquinolin-4-yl)((2R,4S,5R)-5-vinylquinuclidin-2-

yl)methylcarbamate (98)

Page 139: Organocatalysis and Continuous Flow Reactors as Enabling Tools

137

To a solution of 90 (715 mg, 2.2 mmol, 100 mol %) in dry THF (10 mL) was added Et3N

(330 �L, 2.4 mmol, 110 mol %), followed by slow addition of ethyl chloroformate (230 �L,

2.4 mmol, 110 mol %). The mixture was stirred at room temperature under Ar for 2 hours.

The solvent was partly evaporated, and the mixture was partitioned between H2O (30 mL) and

Et2O (30 mL). The aqueous phase was washed with Et2O (3 20 mL). The organic phases

were combined and washed with brine, then dried over MgSO4, filtered and evaporated.

Purification by flash chromatography gave 98 as a white foam (463 mg, 53%). Rf = 0.23 (10%

MeOH in CH2Cl2); [�]20D = +176.6 (c = 1.0, CHCl3); IR (neat): 3185, 3075, 2936, 2870, 1709,

1622, 1590, 1508, 1474, 1433, 1365, 1298, 1252, 1227 cm-1; 1H NMR (C2D2Cl4, 400 MHz,

85 °C): � 8.74 (d, J = 4.6 Hz, 1H), 8.06 (d, J = 9.3 Hz, 1H), 7.62 (d, J = 2.7 Hz, 1H), 7.41 (dd,

J = 11.9, 2.7 Hz, 1H), 7.40 (br s, 1H), 6.09 (br s, 1H), 6.01-5.91 (m, 1H), 5.18 (d, J = 1.5 Hz,

1H), 5.11 (dt, J = 6.3, 1.5 Hz, 1H), 5.10 (dd, J = 10.5, 3.0 Hz, 1H), 4.00 (s, 3H), 3.99 (q, J =

7.1 Hz, 2H), 3.09-2.90 (m, 5H), 2.39-2.31 (m, 1H), 1.73-1.68 (m, 1H), 1.65-1.48 (m, 2H),

1.39-1.31 (m, 1H), 1.10 (t, J = 7.1 Hz, 3H), 1.07-.098 (m, 1H); 13C NMR (CDCl3, 100 MHz,

60°C): 157.9, 156.5, 147.6, 145.8, 145.0, 140.5, 131.9, 128.4, 121.7, 119.5, 114.7, 101.7, 60.8,

60.4, 55.5, 49.4, 47.3, 39.3, 27.6, 26.8, 25.5, 14.3; HRMS m/z 396.2293 [C23H30N3O3 (M+H)+

requires 396.2287].

5.2.22 9-N-Methylamino-(9-deoxy)-epi-quinidine (99)

Lithium aluminum hydride (100 mg, 2.6 mmol, 260 mol %) was suspended in dry THF (4 mL)

at 0 °C under Ar. Compound 98 (390 mg, 1.0 mmol, 100 mol %) was added as a solution in

dry THF (8 mL). The reaction was stirred at 0 °C for 6 hours, then allowed to warm to room

temperature and stirred overnight. LiAlH4 added as THF suspension (1 mmol, 100 mol %) to

drive the reaction to completion. The reaction was quenched by adding H2O (1.4 mL), 10 M

NaOH (0.15 mL) and H2O (0.6 mL) sequentially to give a yellow mixture. The organic phase

was dried over MgSO4, filtered and evaporated to give a yellow foam. Purification by flash

Page 140: Organocatalysis and Continuous Flow Reactors as Enabling Tools

138

chromatography (9% MeOH in CH2Cl2, 1% NH4OH) gave 99 as a yellow oil (170 mg, 51%).

Rf = 0.31 (10% MeOH in CH2Cl2); [�]20D = +56.6 (c = 1.0, CHCl3); IR (neat): 3317, 3075,

2937, 2870, 2787, 1666, 1621, 1588, 1508, 1473, 1454, 1432, 1357, 1257, 1240, 1226 cm-1; 1H NMR (CDCl3, 400 MHz, partial spectra due to heavy rotamerism): � 8.77 (br s, 1H), 8.03

(d, J = 9.5 Hz, 1H), 7.77-7.41 (m, 2H), 7.38 (dd, J = 9.2, 2.7 Hz, 1H), 5.92 (ddd, J = 17.5,

10.1, 6.6 Hz, 1H), 5.12-5.05 (m, 2H), 4.36 (br s, 1H), 3.97 (s, 3H), 3.08-2.80 (m, 5H), 2.32-

2.23 (m, 1H), 2.20 (s, 3H), 1.62-1.46 (m, 3H), 1.24-1.15 (m, 1H), 0.89-0.73 (m, 1H); 13C

NMR (CDCl3, 100 MHz, 55°C, partial spectra due to heavy rotamerism): � 157.5, 147.8

(broad, possibly 2 signals), 144.9, 140.9, 131.8, 129.6, 121.3, 114.4, 102.1, 55.5, 49.6, 47.8,

39.7, 34.3, 27.7, 26.8, 25.0; HRMS m/z 338.2227 [C21H28N3O (M+H)+ requires 338.2232].

5.2.23 3-(3,5-Bis(trifluoromethyl)phenyl)-1-((R)-(6-methoxyquinolin-4-yl)((2R,4S,5R)-5-

vinylquinuclidin-2-yl)methyl)-1-methylthiourea (96)

To a solution of 16 (140 mg, 0.42 mmol, 100 mol %) in dry THF (3 mL) was added 3,5-

bis(trifluoromethyl)phenyl isothiocyanate (90 �L, 0.49 mmol, 110 mol %). The mixture was

stirred overnight at room temperature under Ar. The solvents were evaporated and the residue

purified by flash chromatography (5% to 10% MeOH in CH2Cl2) to give 12 as a whitish foam

(181 mg, 71%). Rf = 0.39 (10% MeOH in CH2Cl2); [�]20D = �3.8 (c = 1.0, CHCl3); IR (neat):

3237 (broad), 3070, 2938, 2872, 1622, 1510, 1474, 1370, 1311, 1278, 1242, 1175, 1134 cm-1; 1H NMR (MeOD-d4, 400 MHz): � 8.72 (d, 1H, J = 4.8 Hz), 8.53 (d, 1H, J = 1.8 Hz), 7.98 (s,

2H), 7.93 (d, 1H, J = 9.3 Hz), 7.74 (d, 1H, J = 4.8 Hz), 7.67 (s, 1H), 7.41 (dd, 1H, J = 9.1, 2.7

Hz), 7.36 (d, 1H, J = 11.0 Hz), 5.84 (ddd, 1H, J = 17.2, 10.7, 6.4 Hz), 5.22 (dt, 1H, J = 17.3,

1.6 Hz), 5.03 (dt, 1H, J = 10.6, 1.5 Hz), 4.09-4.00 (m, 1H), 4.00-3.90 (m, 1H), 3.98 (s, 3H),

3.19-3.11 (m, 1H), 3.05 (s, 3H), 3.05-2.91 (m, 2H), 2.36 (q, 1H, J = 7.8 Hz), 1.89-1.79 (m,

Page 141: Organocatalysis and Continuous Flow Reactors as Enabling Tools

139

1H), 1.77-1.66 (m, 2H), 1.59-1.49 (m, 1H), 1.10-1.01 (m, 1H); 13C NMR (MeOD-d4, 100

MHz): � 184.4, 160.2, 147.7, 145.7, 144.4, 143.4, 141.3, 132.2 (q, 2JCF = 33.3 Hz), 131.3,

131.1, 127.2, 124.8 (q, 1JCF = 271 Hz), 124.3, 123.0, 118.7, 115.2, 105.7, 61.0, 57.7, 56.4,

54.8, 50.8, 50.0, 40.3, 32.8, 28.9, 27.6, 27.2. HRMS m/z 609.2133 [C30H31F6N4OS (M+H)+

requires 609.2123].

5.2.24 N-Methyl-3,5-bis(trifluoromethyl)aniline (104)

3,5-Bis(trifluoromethyl)aniline 102 (1.0 mL, 6.4 mmol, 100 mol %) was dissolved in CH2Cl2

(14 mL) and cooled to 0 °C. Trifluoroacetic anhydride (2.7 mL, 19 mmol, 300 mol %) was

added, and the mixture was stirred at 0 °C for 20 minutes. The solvents were evaporated, and

the residue was dissolved in acetone (15 mL). Anhydrous K2CO3 (1.77 g, 12.8 mmol, 200

mol %) and iodomethane (1.2 mL, 19 mmol, 300 mol %) were added, and the mixture was

heated to mild reflux for 2 hours. The mixture was filtered and evaporated, the residue was

dissolved in H2O:MeOH (5 mL : 25 mL), and anhydrous K2CO3 (880 mg, 6.4 mmol, 100

mol %) was added. The mixture was stirred at room temperature for one hour. The solvents

were partly evaporated, and the residue partitioned between H2O (25 mL) and CH2Cl2 (30

mL). The aqueous phase was extracted with CH2Cl2 (30 mL). The combined organic phases

were dried over MgSO4, filtered and evaporated to give 104 as a yellowish oil (1.41 g, 90%).

The product was practically pure with some solvent residues present. No further purification

was done since the compound was somewhat volatile at reduced pressures and could be

partially lost in high vacuum. Rf = 0.23 (5% EtOAc in hexane); IR (neat): 3465, 2927, 1623,

1523, 1485, 1385, 1273, 1168, 1116, 1100, 1072 cm-1; 1H NMR (CDCl3, 400 MHz): � 7.15 (s,

1H), 6.92 (s, 2H), 4.16 (br s, 1H), 2.91 (s, 3H); 13C NMR (CDCl3, 100 MHz): � 150.1, 132.8

(q, 2JCF = 33.0 Hz), 124.0 (q, 1JCF = 272.0 Hz), 111.8, 110.4, 30.7; HRMS m/z 244.0557

[C9H7F6N (M+H)+ requires 244.0561].

Page 142: Organocatalysis and Continuous Flow Reactors as Enabling Tools

140

5.2.25 1-(3,5-Bis(trifluoromethyl)phenyl)-3-((R)-(6-methoxyquinolin-4-yl)((2R,4S,5R)-5-

vinylquinuclidin-2-yl)methyl)-1-methylthiourea (97)

Compound 104 (130 mg, 0.5 mmol, 100 mol %) was dissolved in dry THF (2 mL) under Ar

and cooled to –78 °C. n-BuLi (2.2 M, 230 �L, 0.5 mmol, 100 mol %) was added and the

resulting yellow mixture was stirred for 15 minutes. Isothiocyanate 101 (180 mg, 0.5 mmol,

100 mol %) dissolved in DMF (3 mL) was added, and the mixture was stirred at –78 °C for 40

minutes, then allowed to warm to room temperature and stirred for further one hour. The

reaction was quenched by adding sat. NH4Cl (1.0 mL), and the mixture was partitioned

between H2O (15 mL) and EtOAc (15 mL). The organic layer was washed with H2O (3 x 10

mL), dried over MgSO4, filtered and evaporated. The residue was purified by flash

chromatography (3% to 10% MeOH in CH2Cl2) to give 97 as an amorphous solid (158 mg,

52%). Another product was also isolated (50 mg) which had identical HRMS spectra and

highly similar 1H NMR spectra. The correct product was assigned by thorough comparison

with NMR data reported by Soós (Ref. 29b). Rf = 0.34 (10% MeOH in CH2Cl2); [�]20D =

+220.7 (c = 1.0, CHCl3); IR (neat): 3241 (broad), 3076, 2939, 2876, 1622, 1590, 1508, 1474,

1382, 1278, 1226, 1180, 1138 cm-1; 1H NMR (CDCl3, 400 MHz, 40°C): � 8.62 (d, J = 4.7 Hz,

1H), 8.02 (d, J = 2.2 Hz, 1H), 7.93 (d, J = 9.2 Hz, 1H), 7.87 (br s, 1H), 7.81 (br s, 2H), 7.47 (d,

J = 4.7 Hz, 1H), 7.42 (dd, J = 9.2, 2.7 Hz, 1H), 6.35 (br d, J = 10.7 Hz), 5.95 (ddd, J = 17.3,

10.6, 6.1 Hz, 1H), 5.21 (dt, J = 17.4, 1.6 Hz, 1H), 5.13 (dt, J = 10.6, 1.5 Hz, 1H), 4.01 (s, 3H),

3.56 (s, 3H), 3.36-3.21 (m, 2H, overlapping signals), 2.99 (dd, J = 14.0, 10.6 Hz, 1H), 2.95-

2.85 (m, 2H), 2.40-2.30 (m, 1H), 1.64-1.49 (m, 3H), 1.26-1.17 (m, 1H), 1.04-0.94 (m, 1H); 13C NMR (MeOD-d4, 100 MHz): � 184.5, 159.6, 148.5, 148.2, 148.0, 145.2, 141.7, 134.0 (q, 2JCF = 34.0 Hz), 131.2, 130.2, 129.3, 124.4, (q, 1JCF = 272.7 Hz), 123.7, 121.7, 121.1, 121.1,

115.2, 104.3, 61.2, 56.5, 50.0, 42.6, 40.2, 28.7, 27.4, 26.5; HRMS m/z 609.2137

[C30H31F6N4OS (M+H)+ requires 609.2123].

Page 143: Organocatalysis and Continuous Flow Reactors as Enabling Tools

141

5.2.26 1H NMR Titration Experiment of 51 in the presence of 12

To an NMR tube was loaded a solution of 51 (10 mg, 0.08 mmol) in CD2Cl2 (0.75 mL), and

the 1H NMR spectrum was recorded. A stock solution of 12 (23 mg, 0.05 mmol) in CD2Cl2

(1.5 mL) was added in three increments (0.2 mL, 0.2 mL and 0.3 mL). The tube was shaken

vigorously and the 1H NMR spectrum was recorded after the addition of each increment. The

amount of 12 corresponded roughly to 10, 20 and 30 mol %.

Table 18. 1H NMR shifts of dimethyl malonate 51 and thiourea 12 in titration experiment

Run 51 CH2 51 OMe 12 H2,H6 12 H4 12 N-H

1 3.3698 3.7158 7.9311 7.8031 merged w/ H2,H6

2 3.3790 3.7176 8.1022 7.7462 8.6377 3 3.3871 3.7188 8.0831 7.7515 8.5673 4 3.3953 3.7200 8.0620 7.7577 8.4879

5.2.27 1H NMR Titration Experiment of 51 in the presence of 79 or 88 and MeOD-d4

To an NMR tube was loaded a solution of 51 (7.3 mg, 0.06 mmol, 100 mol %) in acetone-d6

(0.75 mL). Solution of catalyst 79 (3.5 mg, 0.006 mmol, 10 mol %) or 88 (3.3 mg, 0.006

mmol, 10 mol %) in acetone-d6 (0.6 mL for 79, 0.15 mL for 88) was added and the tube was

shaken vigorously. 1H NMR spectrum of the mixture was recorded to show the mixture of 51

Page 144: Organocatalysis and Continuous Flow Reactors as Enabling Tools

142

and catalyst. MeOD-d4 (0.05 mL, 1.2 mmol) was added and the NMR tube was shaken

vigorously for 1 minute. 1H NMR spectra were recorded at time points of 2, 5, 10, 15, 20, 30

and 50 minutes for 88, and at 2, 5, 10, 20 min and 12 h for 79. Formation of 51-d1 and 51-d2

was determined by the appearance of a new signal at 3.41 (t, J = 2.4 Hz, 1H) and the gradual

disappearance of both the signal at 3.43 (s, 2H) and the triplet at 3.41.

5.3 Total Synthesis of (–)-Hennoxazole A

5.3.1 2-(2,2-Dimethoxyethyl)-2-methyl-1,3-dioxolane (235)

3-Oxobutyraldehyde dimethyl acetal (10.0 mL, 72 mmol, 100 mol %) was loaded in a 250 mL

flask and dissolved in toluene (80 mL). Anhydrous ethylene glycol (12.5 mL, 215 mmol, 300

mol %), triethyl orthoformate (12 mL, 72 mmol, 100 mol %) and PPTS (1.5 g, 6 mmol, 8

mol %) were added. The stirred mixture was heated to reflux for 100 minutes, then let cool to

room temperature. Excess solvents were evaporated, the residue dissolved in Et2O (60 mL),

washed with sat. aq. NaHCO3 (50 mL) and brine (50 mL). The organic phase was dried over

MgSO4, filtered and evaporated to give crude 235 as a yellowish oil (9.66 g, 76%). Pure

product was obtained by high vacuum distillation (92 °C, 0.6 mmHg) of the crude product. Rf

= 0.34 (1:3 EtOAc:Hex); IR (neat) 2984, 2884, 1116, 1036; 1H NMR (CDCl3) � 4.97 (t, J =

4.8 Hz, 1H), 4.0-3.8 (m, 8H), 3.82 (s, 2H), 2.04 (d, J = 4.8 Hz, 2H), 1.39 (s, 3H); 13C NMR

(CDCl3) � 108.0, 101.7, 64.7, 64.5, 43.1, 24.4; HRMS m/z 199.0949 [C8H16O4 (M+Na)+

requires 199.0941].

Page 145: Organocatalysis and Continuous Flow Reactors as Enabling Tools

143

5.3.2 4-Hydroxy-5-(2-methyl-1,3-dioxolan-2-yl)pentan-2-one (242)

Racemic: Commercial LDA solution (1.8 M in THF/heptane/ethylbenzene, 0.65 mL, 1.1

mmol, 110 mol %) was added into a flask containing THF cooled to -78 °C under N2.

Acetone (90 �L, 1.2 mmol, 120 mol %) was added and mixture stirred at -78 °C for 30

minutes. Aldehyde 5 (in 1.5 mL THF, 130 mg, 1.0 mmol, 100 mol %) was added dropwise to

this solution and stirring was continued at –78 °C. After 35 minutes TLC showed no aldehyde,

and the reaction was quenched with sat. aq. NH4Cl, then let warm to room temperature. The

mixture was extracted with EtOAc (5 x 7 mL), and the organic phases were dried over

MgSO4, filtered and evaporated to give crude product as a brownish oil. Purification by flash

chromatography (75% EtOAc in hexanes to 100% EtOAc) gave 242 as a thick colourless oil

(110 mg, 59%). Rf = 0.12 (1:1 EtOAc:Hex); IR (neat): 3495, 2983, 2890, 1709, 1376; 1H

NMR (CDCl3) � 4.39-4.31 (m, 1H), 3.98 (s, 4H), 3.57 (br s, OH), 2.65 (dd, J = 16.2, 7.8 Hz,

1H), 2.51 (dd, J = 16.2, 4.6 Hz, 1H), 2.19 (s, 3H), 1.84 (d, J = 5.9 Hz, 2H), 1.37 (s, 3H). 13C

NMR (101 MHz; CDCl3): � 208.0, 109.9, 64.6, 64.3, 50.6, 44.5, 30.9, 24.0; HRMS m/z

211.0953 [C9H15O4 (M+Na)+ requires 211.0946].

Example of organocatalyst screening reaction: Catalyst 240 (11 mg, 0.027 mmol, 10

mol %) was dissolved in distilled acetone (0.5 mL) in a capped vial, and aldehyde 237 (35 mg,

0.27 mmol, 100 mol %) was added to the solution. The reaction was stirred at room

temperature for 72 hours, then quenched with sat. aq. NH4Cl. The mixture was extracted with

EtOAc (5 x 5 mL), and the combined organic phases were dried over MgSO4, filtered and

evaporated to give crude product, which was analyzed by 1H NMR.

Page 146: Organocatalysis and Continuous Flow Reactors as Enabling Tools

144

5.3.3 Methyl 3-hydroxy-2-(pent-4-enamido)propanoate (250)

Pent-4-enoic acid 249 (6.5 mL, 61.7 mmol, 100 mol %) was dissolved in MeCN (100 mL)

and CDI (11.4 g, 70 mmol, 110 mol %) was added. The mixture was stirred at rt for 1 hour,

and then added to a mixture of DL-SerOMe-HCl (11.5 g, 74 mmol, 120 mol %) and DIPEA

(14 mL, 85 mmol, 140 mol %) prepared 1 hour previously. The solution was stirred at rt

overnight, then washed with 1 M HCl (150 mL) and extracted with EtOAc (200 mL + 4*100

mL). Purification by flash chromatography (75% EtOAc in hex) afforded the amide 250 as a

thick clear oil (10.9 g, 88%). Rf = 0.17 (75% EtOAc in hex); IR (neat): 3352, 3309, 3078,

2955, 1739, 1641, 1531, 1437, 1209; 1H-NMR (400 MHz; CDCl3): � 6.47-6.45 (m, 1H), 5.89-

5.79 (m, 1H), 5.09 (dq, J = 17.1, 1.5 Hz, 1H), 5.04-5.01 (m, 1H), 4.68 (dt, J = 7.3, 3.7 Hz,

1H), 3.97 (dd, J = 11.2, 3.9 Hz, 1H), 3.91 (dd, J = 11.2, 3.4 Hz, 1H), 3.79 (s, 3H), 2.43-2.36

(m, 5H); 13C NMR (100 MHz, CDCl3): � 172.8, 171.0, 136.7, 115.7, 63.4, 54.7, 52.7, 35.6,

29.4; HRMS m/z 202.1079 [C9H15NO4 (M+H)+ requires 202.1079].

5.3.4 Methyl 2-(but-3-enyl)-4,5-dihydrooxazole-4-carboxylate (251)

To a stirred solution of 250 (10.5 g, 52 mmol, 100 mol %) in dry CH2Cl2 (110 mL) at –78 °C

was added DAST (8.0 mL, 61 mmol, 120 mol %) dropwise over 15 minutes. The resulting

mixture was stirred at –78 °C for 30 minutes, then cooling was removed and the mixture

stirred for 40 minutes while it was gradually warming towards rt. K2CO3 (10.0 g, 72 mmol,

mol %) was added, and the mixture stirred for a further 1 hour, then cannulated to 110 mL of

vigorously stirred saturated aqueous NaHCO3. Phases were separated, and the aqueous phase

was extracted with CH2Cl2 (150 mL). The combined organic phases were washed with brine,

dried over MgSO4, filtered through SiO2 and evaporated to give the crude oxazoline 251 in

Page 147: Organocatalysis and Continuous Flow Reactors as Enabling Tools

145

excellent purity and yield (9.28 g, 97%). The product was used as such in the next step. Rf =

0.39 (80% EtOAc in hex); IR 2955, 1739, 1659, 1437, 1204, 1176, 982, 914; 1H NMR (400

MHz, CDCl3): � 5.87-5.77 (m, 1H), 5.08-5.03 (m, 1H), 5.01-4.98 (m, 1H), 4.71 (dd, J = 10.6,

7.8 Hz, 1H), 4.47 (dd, J = 8.7, 7.8 Hz, 1H), 4.37 (dd, J = 10.6, 8.7 Hz, 1H), 3.77 (s, 3H), 2.43-

2.37 (m, 4H); 13C NMR (100 MHz, CDCl3): � 184.1, 165.8, 156.4, 143.4, 141.5, 139.3, 135.9,

129.4, 116.2, 30.6, 27.5; HRMS m/z 184.0974 [C9H13NO3 (M+H)+ requires 184.0974].

Preparation of 251 in flow conditions: Vapourtec R2+/R4 was equipped with 5 mL

Rheodyne injection loops and a 10 mL convection heated flow coil reactor. Two OmniFit

glass columns (� 10 mm) were packed with CaCO3 (2.6 g) and SiO2 (2.4 g). The system was

pressurized with a 100 psi back pressure regulator. The combined system was dried by

pumping anhydrous CH2Cl2 through for 30 minutes. Before carrying out the reaction, the

temperature of the reactor coil was adjusted to 80 °C. Stock solutions (0.4-0.6 M) of 250 and

DAST were injected via injection loops through the reactor with a residence time of 30

minutes (flow rates 0.167 mL/min, both channels). The product was collected into a flask

after the CaCO3 and SiO2 columns and the solvents evaporated to give crude oxazoline 251 in

acceptable purity to be used in the following steps. The columns were emptied and repacked

after 3 runs, when colored impurities had almost reached through the SiO2 plug. In 10

sequential runs 3.15 g (17.2 mmols, 66% combined yield) of 251 was obtained.

5.3.5 Methyl 2-(but-3-enyl)-oxazole-4-carboxylate (252)

To a stirred solution of 251 (9.28 g, 50.7 mmol, 100 mol %) in dry CH2Cl2 (110 mL) at 0 °C

was added DBU (11.3 mL, 75.7 mmol, 150 mol %) and BrCCl3 (10 mL, 101 mmol, 200

mol %) dropwise. The mixture was stirred overnight while slowly allowing it to warm to rt.

The mixture was washed with 1 M HCl and the phases were separated. The aqueous phase

was extracted with CH2Cl2 (2 * 100 mL + 90 mL). The combined organic phases were

washed with brine, dried over MgSO4, filtered through SiO2 and evaporated to give the crude

product as a brown oil. Purification by flash chromatography (33% to 50% EtOAc in hex)

Page 148: Organocatalysis and Continuous Flow Reactors as Enabling Tools

146

gave 252 as a yellow oil (7.89 g, 84% from 250). Rf = 0.63 (80% EtOAc in hex); IR 3166,

2953, 1742, 1642, 1586, 1437, 1322, 11??; 1H NMR (400 MHz; CDCl3): � 8.14 (s, 1H), 5.83

(ddt, J = 17.0, 10.3, 6.6 Hz, 1H), 5.07 (dq, J = 17.1, 1.6 Hz, 1H), 5.01 (dq, J = 10.2, 1.4 Hz,

1H), 3.90 (s, 3H), 2.91 (dd, J = 8.2, 7.0 Hz, 2H), 2.57-2.51 (m, 2H); 13C NMR (100 MHz,

CDCl3): � 165.2, 161.7, 143.7, 136.0, 133.1, 116.2, 52.0, 30.7, 27.6; HRMS m/z 182.0819

[C9H11NO3 (M+H)+ requires 182.0817].

5.3.6 Methyl 2-[2-(but-3-enyl)-oxazole-4-carboxamido]-3-hydroxypropanoate (253)

Oxazole 252 (7.89 g, 43.5 mmol, 100 mol %) was dissolved in THF (60 mL) and treated with

2 M LiOH (39 mL, 78 mmol, 180 mol %). The mixture was stirred at rt for 2 hours, then

poured into 1 M HCl (100 mL). Concentrated HCl (2 mL) was added to ensure full

acidification. The phases were separated, and the aqueous phase was extracted with EtOAc (3

* 100 mL). Combined organic phases were washed with brine, dried over Na2SO4, filtered

and evaporated to give the crude acid, which was then dissolved in MeCN (100 mL), treated

with CDI (9.0 g, 55.5 mmol, 130 mol %) and stirred at rt for 1 hour. The mixture was then

added to a previously prepared mixture of DL-SerOMe-HCl (8.34 g, 53.6 mmol, 120 mol %)

and DIPEA (9.5 mL, 57.5 mmol, 130 mol %) in MeCN (70 mL). The resulting solution was

stirred at rt overnight, then washed with 1 M HCl (120 mL) and extracted with EtOAc (200

mL + 2 * 100 mL). The combined organic phases were washed with brine, dried over Na2SO4,

filtered and evaporated. The crude product was purified by flash chromatography (60% to 100%

EtOAc in hex) to give 253 as a yellow oil (8.1 g, 69%). Rf = 0.60 (EtOAc); IR 3400, 2956,

1742, 1655, 1600, 1513, 1214; 1H NMR (400 MHz, CDCl3): � 8.10 (s, 1H), 7.72 (br d), 5.82

(ddt, J = 17.0, 10.3, 6.6 Hz, 1H), 5.07 (dq, J = 17.1, 1.6 Hz, 1H), 5.01 (dq, J = 10.2, 1.4 Hz,

1H), 4.79 (dt, J = 7.8, 3.9 Hz, 1H), 4.08-3.96 (m, 2H), 3.78 (s, 3H), 2.85 (dd, J = 8.2, 7.1 Hz,

2H), 2.53-2.48 (m, 2H); 13C NMR (100 MHz, CDCl3): � 184.1, 165.8, 156.4, 143.4, 141.5,

139.3, 135.9, 129.4, 116.2, 30.6, 27.5; HRMS m/z 269.1133 [C12H16N2O5 (M+H)+ requires

269.1137].

Page 149: Organocatalysis and Continuous Flow Reactors as Enabling Tools

147

5.3.7 Methyl 2-[2-(but-3-enyl)-oxazol-4-yl]-4,5-dihydrooxazole-4-carboxylate (254)

To a solution of 253 (8.0 g, 30 mmol, 100 mol %) in dry CH2Cl2 (100 mL) at –78 °C was

added DAST (4.7 mL, 35 mmol, 120 mol %) dropwise over 40 minutes. The cooling was

stopped and the resulting mixture stirred for 40 minutes while letting it slowly warm up.

K2CO3 was added and the mixture stirred vigorously for further 2 hours, then cannulated into

vigorously stirred saturated aqueous NaHCO3 (100 mL). The mixture was diluted with 100

mL H2O, the phases were separated, and the aqueous phase was extracted with CH2Cl2 (100

mL). The combined organic phases were washed with brine, dried over MgSO4, filtered and

evaporated. The crude product was filtered through SiO2 with EtOAc and the solvents were

evaporated to give a cloudy yellow oil, which was then redried with Na2SO4 to remove

suspected water. Filtration and evaporation gave 254 as a cloudy yellow oil (7.1 g, 95%),

which was of excellent purity by 1H NMR, and the product was used as such in the next step.

Rf = 0.32 (80% EtOAc in hex); IR 2954, 1739, 1676, 1583, 1206, 1104; 1H NMR (400 MHz,

CDCl3): � 8.10 (s, 1H), 5.83 (ddt, J = 17.0, 10.4, 6.6 Hz, 1H), 5.06 (dq, J = 17.1, 1.6 Hz, 1H),

5.01 (q, J = 1.4 Hz, 1H), 4.99 (q, J = 1.4 Hz, ), 4.92 (dd, J = 10.5, 8.0 Hz, 1H), 4.67 (t, J = 8.3

Hz, 1H), 4.57 (dd, J = 10.5, 8.7 Hz, 1H), 3.79 (s, 3H), 2.90 (dd, J = 8.2, 7.1 Hz, 2H), 2.57-

2.51 (m, 2H); 13C NMR (100 MHz, CDCl3): � 171.1, 165.3, 160.1, 141.1, 136.1, 129.8, 116.0,

69.6, 68.4, 52.6, 30.6, 27.5; HRMS m/z 464.2011 [C12H14N2O4 (M+H)+ requires 251.1032].

5.3.8 Methyl 2'-(but-3-enyl)-2,4'-bioxazole-4-carboxylate (255)

To a solution of 254 (7.1 g, 28 mmol, 100 mol %) in dry CH2Cl2 (100 mL) at 0 °C was added

DBU (6.4 mL, 43 mmol, 150 mol %) and BrCCl3 (5.5 mL, 56 mmol, 200 mol %) dropwise.

Page 150: Organocatalysis and Continuous Flow Reactors as Enabling Tools

148

The solution was stirred at 0 °C for 80 minutes, then washed with 1 M HCl (50 mL). The

phases were separated and the aqueous phase was extracted with CH2Cl2 (3 * 50 mL). The

combined organic phases were washed with brine, dried over Na2SO4, filtered and evaporated.

The crude product was purified by flash chromatography (5% to 80% EtOAc in CH2Cl2) to

give 255 as a yellowish solid (5.83 g, 78% from 253). Mp. 118-119 °C; Rf = 0.55 (80%

EtOAc in hex); IR 3148, 2952, 1719, 1637, 1570, 1436, 1317; 1H NMR (400 MHz, CDCl3): �

8.28 (s, 1H), 8.28 (s, 1H), 5.85 (ddt, J = 17.0, 10.3, 6.6 Hz, 1H), 5.09 (dq, J = 17.1, 1.6 Hz,

1H), 5.03 (dq, J = 10.2, 1.3 Hz, 1H), 3.94 (s, 3H), 2.97-2.93 (m, 2H), 2.60-2.55 (m, 2H); 13C

NMR (100 MHz; CDCl3): � 165.6, 161.3, 143.5, 139.2, 135.9, 134.2, 129.5, 116.2, 52.2, 30.6,

27.5; HRMS m/z 271.0688 [C12H12N2O4Na (M+H)+ requires 271.0695].

5.3.9 But-3-enyl-2,4'-bisoxazole-4-carbaldehyde (243)

To a solution of 255 (1.6 g, 6.4 mmol, 100 mol %) in dry CH2Cl2 (35 mL) at –78 °C was

slowly added Dibal-H (1 M in CH2Cl2, 13 mL, 13 mmol, 200 mol %) while keeping the

internal temperature below –65 °C. After addition the mixture was stirred for 1 hour, then

quenched by adding MeOH (3 mL) dropwise. The mixture was stirred for 10 minutes at –

78 °C, then solid Rochelle’s salt (5.6 g) and 2 mL of saturated aqueous Rochelle’s salt were

added. The mixture was stirred vigorously and let warm to rt, then filtered through celite,

washed with H2O, dried over MgSO4, filtered and evaporated. The crude product was purified

by flash chromatography (50% to 66% EtOAc in hex) to give the aldehyde 243 as a yellowish

solid (1.19 g, 85%). Mp. 106-107 °C; Rf = 0.52 (80% EtOAc in hex); IR 3126, 3082, 2830,

1684, 1641, 1553; 1H NMR (400 MHz, CDCl3) 10.00 (s, 1 H), 8.29 (s, 1 H), 8.24 (s, 1 H),

5.87 (ddd, 1 H, J = 17.0, 10.3, 6.6 Hz), 5.10 (dq, 1 H, J = 17.1, 1.6 Hz), 5.04 (dq, 1 H, J =

10.2, 1.4 Hz), 2.97 (dd, 2 H, J = 8.1, 7.1 Hz), 2.62-2.59 (m, 2H); 13C NMR (100 MHz,

CDCl3): � 184.1, 165.8, 156.4, 143.4, 141.5, 139.3, 135.9, 129.4, 116.2, 30.6, 27.5; HRMS

m/z 219.0775 [C11H10N2O3 (M+H)+ requires 219.0770].

Page 151: Organocatalysis and Continuous Flow Reactors as Enabling Tools

149

5.3.10 (S)-2-Amino-4-methyl-1,1-diphenylpentan-1-ol (262)

Commercial PhMgBr (3.0 M in Et2O, 60 mL, 180 mmol, 670 mol %) was loaded in a dry 250

mL flask and diluted with 60 mL dry Et2O to form a roughly 1.8 M solution, which was

cooled to 0 °C. L-Leucine methyl ester hydrochloride (4.9 g, 27 mmol, 100 mol %) was added

portionwise to the flask, and the resulting mixture was stirred overnight while letting it warm

to rt. The reaction was cooled to 0 °C and quenched by slowly adding saturated aqueous

NH4Cl (60 mL). The resulting suspension was partitioned between H2O (200 mL) and EtOAc

(150 mL), and the aqueous phase was extracted with EtOAc (2 *100 mL). The organic phases

were combined and washed with brine, dried over Na2SO4, filtered and evaporated. The crude

product was purified by crystallizing twice from EtOH, and by sequentially precipitating the

remaining product from EtOAc by partially evaporating the solvent and filtering out the

precipitated solid. The crops were combined and washed with cold hexane to give the amino

alcohol as a white solid (5.37 g, 74%). [�]25D = –108 (c = 1.0, CHCl3); Rf = 0.20 (80%

EtOAc:hex); 1H-NMR (400 MHz; CDCl3): � 7.62 (dd, J = 8.4, 1.2 Hz, 2H), 7.48 (dd, J = 8.4,

1.2 Hz, 2H), 7.31 (dd, J = 15.7, 7.4 Hz, 4H), 7.21-7.14 (m, 2H), 3.99 (dd, J = 10.2, 2.1 Hz,

1H), 1.61-1.55 (m, 1H), 1.27 (ddd, J = 14.1, 10.3, 3.8 Hz, 1H), 1.08 (ddd, J = 14.1, 10.7, 2.1

Hz, 1H), 0.90 (d, J = 6.5 Hz, 3H), 0.87 (d, J = 6.6 Hz, 3H). Data in agreement with literature

values.144

5.3.11 (S)-N-[(S)-1-hydroxy-4-methyl-1,1-diphenylpentan-2-yl]pyrrolidine-2-carboxamide

(257)

Page 152: Organocatalysis and Continuous Flow Reactors as Enabling Tools

150

To a solution of L-Boc-proline (3.37 g, 15.7 mmol, mol %) in dry CH2Cl2 at 0 °C was added

Et3N (2.6 mL, 18.8 mmol, mol %) and EtOCOCl (1.7 mL, 17.9 mmol, mol %) dropwise. The

mixture was stirred at 0 °C for 30 minutes followed by addition of 262 (3.95 g, 14.7 mmol,

100 mol %) in one portion. The reaction was let warm to rt and stirred overnight. The mixture

was diluted with CH2Cl2 (90 mL), then filtered to give a white residue and a clear filtrate. The

residue was redissolved in CH2Cl2 and combined with the filtrate. The mixture was washed

with H2O (2 * 150 mL) and brine, dried over Na2SO4, filtered and evaporated to give 293 as a

white solid, which was used without further purification.

Deprotection: HCOOH (20 mL) was slowly added to 293 (2.1 g, 4.5 mmol) at 0 °C, and the

mixture was let warm to rt and stirred overnight. HCOOH was evaporated to give a viscous

oil, which was dissolved in CH2Cl2 and basified to strongly basic with 1 M NaOH. The

mixture was diluted with 50 mL H2O, the phases were separated, and the aqueous phase was

extracted with CH2Cl2 (2 * 50 mL). The combined organic phases were washed with brine,

dried over Na2SO4, filtered and evaporated. Purification by flash chromatography (10%

MeOH in CH2Cl2, 1% NH4OH) gave 257 as a white solid (1.14 g, 69%). NB! Standard

TFA:CH2Cl2 deprotection conditions led to partial hydrolysis of the catalyst, and only 25% of

the desired product was isolated. Mp. 185-190°C; [�]25D = –48 (c = 1.0, CHCl3); Rf = 0.1 (10%

MeOH in CH2Cl2); 1H NMR (400 MHz, CDCl3): � 7.94-7.92 (br d, 1H), 7.57-7.53 (m, 4H),

7.32-7.28 (m, 2H), 7.24-7.22 (m, 4H), 7.20-7.16 (m, 1H), 7.14-7.10 (m, 1H), 4.59 (ddd, J =

11.1, 8.7, 2.2 Hz, 1H), 3.49 (dd, J = 9.4, 4.7 Hz, 1H), 2.83 (m, 1H), 2.58 (m, 1H), 1.91-1.83

(m, 2H), 1.82-1.66 (m, 2H), 1.61-1.52 (m, 1H), 1.49-1.43 (m, 2H), 1.26-1.19 (m, 2H), 0.92 (d,

J = 6.5 Hz, 3H), 0.86 (d, J = 6.7 Hz, 3H); 13C NMR (100 MHz; CDCl3): � 175.8, 146.6, 145.1,

128.2, 127.9, 126.6, 126.4, 125.7, 125.6, 80.9, 60.3, 56.8, 47.0, 37.4, 30.5, 25.7, 25.4, 23.8,

21.5; HRMS m/z 367.2377 [C23H30N2O2 (M+H)+ requires 367.2386].

5.3.12 (R)-4-[2'-(but-3-en-1-yl)-(2,4'-bisoxazol)-4-yl]-4-hydroxybutan-2-one (244)

Representative example of organocatalytic reaction: Catalyst 257 (37 mg, 0.1 mmol, 20

mol %) and MCA (4.7 mg, 0.05 mmol, 10 mol %) were loaded in a small flask and dissolved

Page 153: Organocatalysis and Continuous Flow Reactors as Enabling Tools

151

in acetone (3 mL). The mixture was stirred for 30 minutes at 0 °C. Aldehyde 243 (110 mg, 0.5

mmol, 100 mol %) was added to the flask, and the reaction was stirred overnight while letting

the cooling bath slowly reach room temperature. The reaction was quenched with saturated

aqueous NH4Cl, and the mixture extracted with EtOAc (3 x 5 mL). The combined organic

phases were washed with brine, dried over Na2SO4, filtered and evaporated. The residue was

purified by flash chromatography (60% to 100% EtOAc in hex) to give 244 as a slowly

crystallizing oil. Enantiomeric excess was analyzed by HPLC from a sample dissolved in 1:1

i-PrOH:hexane: ChiralPak AS column (0.46 * 25 cm), 15% i-PrOH in hexane, 1 mL/min, =

254 nm, tR(major) 29.1 min, tR(minor) 36.1 min.

Representative example of flow reaction: The reaction was performed with Uniqsis

FlowSyn reactor. A stock solution of aldehyde 243 in acetone (0.33 M) and catalyst 257 and

MCA in acetone (0.066 M + 0.045 M) were prepared. 1 mL of stock solutions was injected in

sample loops. The streams were mixed in a T-piece and then pumped through a 20 mL

stainless steel reactor coil at room temperature with a residence time of 60 minutes (individual

flow rates 0.16 mL/min). The product stream was collected on saturated aqueous NH4Cl

solution and worked up as in batch conditions.

A racemic sample of 244 was prepared by standard LDA procedure: To a solution of

diisopropylamine (100 μL, 0.7 mmol, mol %) in dry THF (1 mL) at 0 °C was added n-

butyllithium (2.5 M in hexane, 400 μL, 0.8 mmol, mol %). The mixture was stirred at 0 °C for

a few minutes, and then cooled to –78 °C. Anhydrous acetone (60 μL, 0.8 mmol, mol %) was

added, and the mixture stirred for few minutes before the aldehyde 243 (137 mg, 0.63 mmol,

100 mol %) was added dropwise in dry THF (2 mL).

Mp. 59-61 °C (rac) / 69-78 °C (ee 70%); Rf 0.27 (80% EtOAc:hex); IR (neat): 3369, 3109,

2900, 1710, 1632, 1585, 1529, 1407, 1357, 1302, 1104, 1066, 989, 975, 919; 1H NMR (400

MHz; CDCl3): � 8.09 (s, 1H), 7.63 (d, J = 1.2 Hz, 1H), 5.85 (ddt, J = 17.0, 10.3, 6.6 Hz, 1H),

5.18 (ddd, J = 8.7, 3.3, 1.2 Hz, 1H), 5.08 (dq, J = 17.1, 1.6 Hz, 1H), 5.02 (dq, J = 10.2, 1.4 Hz,

1H), 3.15 (dd, J = 17.9, 3.3 Hz, 1H), 2.98 (d, J = 8.7 Hz, 1H), 2.94 (dd, J = 8.2, 7.0 Hz, 3H),

2.60-2.54 (m, 2H), 2.20 (s, 3H). 13C NMR (100 MHz; CDCl3): � 209.0, 165.6, 155.4, 143.6,

138.0, 136.1, 134.6, 130.3, 116.1, 64.1, 48.9, 30.77, 30.70, 27.6; HRMS m/z 277.1175

[C14H16N2O4 (M+H)+ requires 277.1188].

Page 154: Organocatalysis and Continuous Flow Reactors as Enabling Tools

152

5.3.13 Attempted determination of absolute stereochemistry:

Attempt with (S)-(+)-MTPA-Chloride:

To a dry 5 mL flask was sequentially loaded Et3N (150 μL, 110 mg), DMAP (1 mg, 0.01

mmol), (S)-(+)-MTPA-Cl (13 μL, 17 mg, 0.07 mmol), CH2Cl2 (200 μL) and 244 (13 mg, 0.05

mmol). The mixture was stirred at rt overnight, then diluted with Et2O and washed with cold

0.03 M HCl, cold saturated NaHCO3 and brine. The organic phase was dried over Na2SO4,

filtered and evaporated to give a whitish solid, which was directly analyzed by 1H NMR.

Analysis showed the product to consist solely of the corresponding alkene 263. 1H NMR (400

MHz, CDCl3): � 8.17 (s, 1H), 7.84 (s, 1H), 7.36 (d, J = 15.7 Hz, 1H), 7.07 (d, J = 15.7 Hz,

1H), 5.86 (ddt, J = 17.0, 10.3, 6.6 Hz, 1H), 5.09 (dq, J = 17.1, 1.6 Hz, 1H), 5.03 (dq, J = 10.2,

1.4 Hz, 1H), 2.96 (t, J = 7.6 Hz, 2H), 2.61-2.56 (m, 2H), 2.34 (s, 3H).

Attempt with (R)-(–)-MTPA acid:

To a dry 5 mL flask was loaded (R)-(–)-MTPA (34 mg, 0.14 mmol, 140 mol %) which was

then dissolved in dry CH2Cl2 (0.5 mL). To this solution was added 244 (27 mg, 0.1 mmol,

100 mol %) and DMAP (6.7 mg, 0.05 mmol, 50 mol %). The mixture was stirred and cooled

to 0 °C, followed by addition of DCC (47 mg, 0.23 mmol, 230 mol %). The mixture was

stirred at rt overnight, then diluted with Et2O, filtered, and washed with 0.5 M HCl (2 x 5 mL),

saturated NaHCO3 (2 x 5 mL). The organic phase was dried over Na2SO4, filtered and

evaporated to give the crude ester ent-264, along with roughly 10% of the alkene, and some

Page 155: Organocatalysis and Continuous Flow Reactors as Enabling Tools

153

leftover N,N-dicyclohexylurea. 1H NMR (400 MHz, CDCl3): � 8.10 (s, 1H), 7.71 (s, 1H),

7.56-7.30 (m, 6H), 6.47 (dd, J = 8.6, 4.7 Hz, 1H), 5.85 (ddt, J = 17.0, 10.3, 6.6 Hz, 1H), 5.11-

5.00 (m, 2H), 3.45 (s, 3H), 3.35 (dd, J = 17.6, 3.4 Hz, 1H), 3.21-3.15 (m, 1H), 3.11 (dd, J =

17.6, 4.7 Hz, 1H), 2.97-2.93 (m, 2H), 2.59-2.54 (m, 2H), 2.08 (s, 3H).

5.3.14 (R)-4-(2'-(but-3-en-1-yl)-[2,4'-bisoxazol]-4-yl)-4-methoxybutan-2-one (245)

To a solution of 244 (77 mg, 0.28 mmol, 100 mol %) in CH2Cl2 (3 mL) was added MeI (1 mL)

and Ag2O (130 mg, 0.56 mmol, 200 mol %), and the mixture was protected from light and

stirred at rt until completion. The mixture was filtered through celite and the solvents

evaporated. Purification by flash chromatography (60% to 90% EtOAc in hex) gave pure 245

as a slowly crystallizing yellowish oil (29 mg, 36%). HPLC: ChiralPak AS column (0.46 * 25

cm), 5% i-PrOH in hexane, 1 mL/min, = 254 nm, tR(major) 34 min, tR(minor) 55 min; Mp.

62-66 °C; Rf = 0.47 (80% EtOAc in hex); IR (neat): 3133, 2930, 2880, 2823, 1709, 1633,

1585, 1531, 1435, 1362, 1098; 1H NMR (400 MHz; CDCl3): � 8.15 (s, 1H), 7.64 (d, J = 0.6

Hz, 1H), 5.87 (ddt, J = 17.0, 10.3, 6.6 Hz, 1H), 5.11 (dq, J = 17.1, 1.6 Hz, 1H), 5.04 (dq, J =

10.2, 1.4 Hz, 1H), 4.77 (dd, J = 8.3, 4.5 Hz, 1H), 3.35 (s, 3H), 3.09 (dd, J = 16.7, 8.4 Hz, 1H),

2.98-2.90 (m, 3H), 2.62-2.56 (m, 2H), 2.20 (s, 3H); 13C NMR (100 MHz, CDCl3): � 205.9,

165.5, 155.8, 141.2, 138.1, 136.1, 135.5, 130.3, 116.1, 71.9, 57.2, 48.5, 30.8, 27.6; HRMS m/z

313.1166 [C15H18N2O4 (M+Na)+ requires 313.1164].

Proof of principle flow reaction: Through an OmniFit column (� 6.6 mm) packed with a

bed of Ag2O dispersed in celite and capped with celite plugs was pumped a premixed solution

of 244 (27 mg, 0.1 mmol) in 1:1 CH2Cl2:MeI (1 mL) at a flow rate of 0.1 mL/min. The crude

product was collected, evaporated and analyzed by 1H NMR to show a 2:1 mixture of 244:245

with little impurities.

Page 156: Organocatalysis and Continuous Flow Reactors as Enabling Tools

154

5.3.15 Representative Reaction for �-Hydroxy Ketone Reduction With Polymer Supported

Borohydride: (3RS,5RS)-2,6-Dimethylheptane-3,5-diol (268) and (3RS,5SS)-2,6-

Dimethylheptane-3,5-diol (269)

To a 5 mL flask was loaded �-hydroxy ketone 265 (167 mg, 1.05 mmol, 100 mol %) and

dissolved in MeOH (5 mL). Borohydride resin (1.0 g) was added, and the mixture was stirred

for 8 hours until TLC showed no more 265. The mixture was filtered and solvents evaporated

to give the crude product (154 mg, 93%) as a mixture of diastereomers 268 and 269. Spectral

data of compounds was determined to correspond to literature values.125b,132,145

268: 1H NMR (400 MHz; CDCl3): � 3.62 (ddd, J = 10.2, 5.0, 1.3 Hz, 2H), 2.95 (br s, 2H),

1.73-1.62 (m, 2H), 1.59 (dt, J = 14.4, 2.0 Hz, 1H), 1.41 (dt, J = 14.4, 10.3 Hz, 1H), 0.92 (d, J

= 6.8 Hz, 12 H); 13C NMR (100 MHz, CDCl3): � 78.2, 35.8, 34.3, 18.3, 17.4. 269: 1H NMR

(400 MHz; CDCl3): � 3.65 (app q, 2H), 2.21 (br s, 2H), 1.70 (spt, J = 6.7 Hz, 2H), 1.59 (dd, J

= 6.4, 5.1 Hz, 2H), 0.96 (d, J = 6.7 Hz, 6H), 0.90 (d, J = 6.8 Hz, 6H); 13C NMR (100 MHz,

CDCl3): � 74.2, 36.5, 33.7, 18.7, 18.0.

5.3.16 Analysis of Diastereoselectivity

OHOH OO

268 294

p-TsOH, Na2SO4OO

295

+

OMeMeO

Crude 268 (23 mg, 0.14 mmol, 100 mol %) was dissolved in distilled 2,2-dimethoxypropane

(2 mL), and a small amount of Na2SO4 was added, and then a catalytic amount of p-TsOH (8

mg, 0.04 mmol, 30 mol %) was added. The reaction was quenched with NaHCO3 (aq), and

the resulting mixture extracted with Et2O (3 x 10 mL). The organic phases were washed with

brine, dried over Na2SO4, filtered and evaporated to give a crude mixture of 294 and 295,

which was analyzed by 1H and 13C NMR. NMR spectral data of compounds was determined

to correspond to literature values.132,145

Page 157: Organocatalysis and Continuous Flow Reactors as Enabling Tools

294: 1H

1.45 (dt

J = 11.9

CDCl3)

3.36 (ap

0.92 (d,

34.4, 33

5.3.17 R

phenylp

Prepare

literatur

NMR si

the crud

Figure diastere

H NMR (400

t, J = 12.6, 2

9 Hz, 1H),

: � 98.1, 74

pp q, J = 7.

J = 6.6 Hz

3.0, 24.3, 18

Reduction of

pentan-3-on

OH

266

O

267

d according

re values.14

ignals of 27

de reaction m

22. Redeoselectivity

0 MHz; CD

2.4 Hz, 1H)

0.91 (d, J =

4.1, 33.1, 30

.9 Hz, 2H),

, 6H), 0.86

8.9, 17.6.

f 3-hydroxy

e (270 and

Ph

O

6

Ph

OH

7

g to the pro45b,146 Analy

70 at 4.94 p

mixture (Fig

duction of y

Cl3): � 3.44

), 1.38 (d, J

= 6.8 Hz, 6

0.2, 30.2, 19

1.66-1.55

(d, J = 6.8 H

y-4-methyl-1

296)

NMe3B

MeOH

NMe3B

MeOH

ocedure des

ysis of dias

ppm (dd, J =

gures 22 &

f 266 wit

4 (ddd, J =

J = 6.6 Hz, 6

6H), 0.87 (d

9.8, 18.5, 1

(m, 2H), 1

Hz, 6H); 13

1-phenylpen

BH4

BH4

scribed in 5

stereoselect

= 9.0, 3.8 H

23).

th solid

11.5, 6.6, 2

6H), 1.39 (s

d, J = 6.8 H

17.7. 295: 1H

.57 (app t, J

C NMR (10

ntan-3-one a

Ph

OHOH

270

Ph

OH OH

270

5.3.15. NM

tivity was

Hz, 1H) and

supported

.3 Hz, 2H),

, 3H), 1.37

Hz, 6H); 13C

H NMR (40

J = 7.9 Hz,

00 MHz, CD

and 1-hydro

O

+

O

+

MR spectral

determined

d 296 at 5.0

borohydrid

, 1.66-1.56

(s, 3H), 1.0

C NMR (10

00 MHz; C

, 2H), 1.32

DCl3): � 100

oxy-4-methy

Ph

OHOH

296 �

Ph

OHOH

296

data corres

d by integra

07 (app q, 1

de, 1:1 s

155

(m, 2H),

08 (app q,

00 MHz,

DCl3): �

(s, 6H),

0.1, 72.0,

yl-1-

spond to

ating 1H

H) from

syn:anti-

Page 158: Organocatalysis and Continuous Flow Reactors as Enabling Tools

156

Figure diastere

5.4 Con

5.4.1 5-

Represe

mol %)

mL 3-n

monitor

added d

turned b

precipit

kept at

toluene:

272 as a

(d, J =

the mate

23. Redeoselectivity

ntinuous Fl

-Nitrovanill

entative ba

was dissol

neck flask e

ring the int

dropwise wi

brownish im

tating rapidl

that tempe

:AcOH (3:1

a yellow po

1.8 Hz, 1H)

erial to be 9

duction of y

low Nitratio

in (272)

HO

MeO

271

atch proce

lved in a m

quipped wi

ternal tempe

ith an addit

mmediately

ly. After th

erature for

1). The prec

owder (9.7 g

), 3.98 (s, 3

98.8% pure.

f 267 wit

on of Arom

H

O

edure (Tab

mixture of gl

ith a mecha

erature. HN

tion funnel

after the ad

he addition

45 minutes

cipitate was

g, 75%). 1H

3H), corresp

th solid

matic Subst

HNO3

AcOH

ble 10, Ent

lacial AcOH

anical stirre

NO3 (60 w-

over a per

ddition was

was compl

s. The mix

s dried in a

H NMR: � 9

ponding to l

supported

trates

HO

MeO

try 12): Va

H (30 mL)

r. The mixt

-%, 5.5 mL

riod of 3.5 m

HNO3 was

lete, the mi

xture was f

vacuum ov

.88 (s, 1H),

literature da

borohydrid

H

O

NO2

272

anillin (10.0

and toluene

ture was co

L, 69 mmol

minutes. Th

begun, and

xture was c

filtered and

ven overnig

8.11 (d, J =

ata.135n HPL

de, 2:1 s

H

0 g, 66 mm

e (15 mL)

ooled to 5 °

l, 105 mol

he reaction

d the produc

cooled to 0

d washed w

ght at 60 °C

= 1.8 Hz, 1

LC analysis

syn:anti-

mol, 100

in a 250

°C while

%) was

mixture

ct started

0 °C and

with cold

C to give

H), 7.64

showed

Page 159: Organocatalysis and Continuous Flow Reactors as Enabling Tools

157

Representative pulse experiment in flow conditions: Uniqsis FlowSyn reactor was

equipped with a transparent 5 mL HT-PFA reactor coil and a 100 psi back pressure regulator,

and then the system completely flushed with glacial AcOH which was used as the system

solvent. Stock solutions of vanillin and HNO3 of desired concentrations in AcOH were

prepared to separate volumetric flasks. The reactor was primed to the desired temperature and

flow rate, after which 1 mL of both stock solutions was injected to the reactor via injection

ports equipped with proper injection loops. After the indicated residence time, the reaction

mixture was collected to a separate beaker. The initiation of the reaction was observed by a

color change to yellow either in the beaker after collection was ended, right after the reaction

stream exited the bpr, or in the reactor coil.

Representative continuous experiment: Stock solutions of vanillin and HNO3 in AcOH

were prepared in separate volumetric flasks. Uniqsis FlowSyn reactor was equipped with a 10

mL stainless steel reactor coil and a 100 psi back pressure regulator, and then the system was

completely flushed with glacial AcOH which was used as the system solvent. The reactor was

primed to the desired temperature and flow rate, after which the stock solutions were fed

through the pumps into the reactor. The exiting reacted mixture was collected to a separate

beaker partially filled with H2O. The outbound stream was analyzed with HPLC by collecting

the mixture into a 10 mL volumetric flask for a few seconds, and then diluting to indicated

volume with 1:1 H2O:MeCN. The output stream was collected for an indicated time, then

cooled to 0 °C in an ice bath for 1 hour, filtered, and washed with ice cold H2O and EtOH.

The yellow powder was dried in a vacuum oven and analyzed with HPLC.

HPLC analyses: The reactions were characterized by reverse phase HPLC analyses with

Waters Alliance 2695 system uding a Symmetry Shield RP8 column (3.5 �m, 4.6 x 150 mm)

and PDA 996 UV detector ( = 284 nm). Eluent A for gradient elution was H2O:MeCN:TFA

850:150:1, and eluent B was MeCN:TFA 1000:1, and elution speed was 1.5 mL/min. The

gradient was as follows:��

Minutes A(%) B(%) 0-2 100 0 2-20 50 50 20-22 15 85 22-27 15 85 27-29 100 0 29-35 100 0

Page 160: Organocatalysis and Continuous Flow Reactors as Enabling Tools

158

6 REFERENCES

1 a) Kirschning, A.; Solodenko, W.; Mennecke, K. Chem. Eur. J. 2006, 12, 5972-5990. b) O’Brien,

M.; Denton, R.; Ley, S. V. Synthesis 2011, 1157-1192. c) For a collection of articles see Org.

Biomol. Chem. 2010, Web theme issue “Enabling Technologies for Organic Synthesis”

2 a) Originally coined by Ostwald, W. Z. Physik. Chem. 1900, 34, 510. and b) Langenbeck, W.

Chem. Ber. 1927, 60, 930-934. c) Recoined by Ahrendt, K. A.; Borths, C. J.; Macmillan, D. W. C.

J. Am. Chem. Soc. 2000, 122, 4243-4244.

3 a) Michael, A. J. Prakt. Chem. 1887, 35, 349-356. b) Michael, A. Am. Chem. J. 1887, 9, 112-124.

c) Michael, A. J. Prakt. Chem. 1894, 49, 20-29.

4 For representative reviews, see: a) Aza-Michael reaction: Reyes, E.; Fernandez, M.; Uria, U.;

Vicario, J. L. Curr. Org. Chem. 2012, 16, 521-546 and Wang, J.; Li, P.; Choy, P. Y.; Chan, A. S.

C., Kwong, F. Y. ChemCatChem. 2012, 4, 917-924. b) Oxa-Michael reaction: Nising, C. F.;

Bräse, S. Chem. Soc. Rev. 2012, 41, 988-999. c) Sulfa-Michael reaction: Enders, D.; Lüttgen, K.;

Narine, A. A. Synthesis 2007, 959-980. d) Phospha-Michael reaction: Enders, D.; Saint-Dizier, A.;

Lainnou, M.-I.; Lenzen, A. Eur. J. Org. Chem. 2006, 29-49.�

5 For a recent review on metal-catalyzed formation of quaternary stereocenters via conjugate

addition, see: a) Hawner, C.; Alexakis, A. Chem. Commun. 2010, 46, 7295-7306. For other

relevant reviews, see: b) Das, J. P.; Marek, I. Chem. Commun. 2011, 47, 4593-4623. c) Trost, B.

M.; Jiang, C. Synthesis 2006, 369-396. d) Corey, E. J.; Guzman-Perez, A. Angew. Chem. Int. Ed.

1998, 37, 388-401. For a recent example on organocatalytic conjugate addition to �,�-

disubstituted enone, see e) Kwiatkowski, P.; Dudzi�ski, K.; �yzwa, D. Org. Lett. 2011, 13, 3624-

3627.�

6 For recent reviews on asymmetric organocatalytic conjugate and Michael-additions, see: a)

Alma�i, D.; Alonso, D. A.; Nájera, C. Tetrahedron: Asymmetry 2007, 18, 299-356. b) Bella, M.;

Gasperi, T. Synthesis 2009, 1583-1614.

7 Bredig, G.; Fiske, P. S. Biochem. Z. 1912, 46, 7.

8 List, B.; Lerner, R. A.; Barbas, C. F., III J. Am. Chem. Soc. 2000, 122, 2395-2396.

Page 161: Organocatalysis and Continuous Flow Reactors as Enabling Tools

159

9 a) Dalko, P. I.; Moisan, L. Angew. Chem. Int. Ed. 2004, 43, 5138-5175. b) Berkessel, A.; Gröger,

H. Asymmetric Organocatalysis: From Biomimetic Concepts to Application in Asymmetric

Synthesis, Wiley-VCH: Weinheim, 2005. c) Dalko, P. I. Enantioselective Organocatalysis:

Reactions and Experimental Procedures, Wiley-VCH: Weinheim, 2007. d) Pellissier, H.

Tetrahedron 2007, 63, 9267-9331. e) Hernández, J. G.; Juaristi, E. Chem. Commun. 2012, 48,

5396-5409. f) Giacalone, F.; Gruttadauria, M.; Agrigento, P.; Noto, R. Chem. Soc. Rev. 2012, 41,

2406-2447.

10 For reviews on enamine catalysis, see e.g. a) Mukherhee, S.; Yang, J. W.; Hoffman, S.; List, B.

Chem. Rev. 2007, 107, 5471-5569. b) Pihko, P. M.; Majander, I.; Erkkilä, A Top. Curr. Chem.

2010, 291, 29-75.

11 For reviews on iminium catalysis, see e.g. a) Erkkilä, A.; Majander, I.; Pihko, P. M. Chem. Rev.

2007, 107, 5416-5470. b) Brazier, J. B.; Tomkinson, N. C. O. Top. Curr. Chem. 2010, 291, 281-

347.

12 For reviews on metal-based bifunctional catalysis, see e.g.: a) Shibasaki, M.; Kanai, M.;

Matsunaga, S.; Kumagai, N. Acc. Chem. Res. 2009, 42, 1117-1127. b) Grotjahn, D. B. Dalton

Trans. 2008, 6497-6508. c) Yamamoto, H.; Futatsugi, K. Angew. Chem. Int. Ed. 2005, 44, 1924-

1942. d) Ma, J.-A.; Cahard, D. Angew. Chem. Int. Ed. 2004, 43, 4566-4583.

13 Jeffrey, G. A. Introduction to Hydrogen Bonding, Oxford University Press: New York, 2009.

14 Berkessel, A.; Etzenbach-Effers, K. In Hydrogen Bonding in Organic Synthesis, Pihko, P. M., Ed.;

Wiley-VCH: Weinheim, 2009.

15 a) Hine, J.; Linden, S. M.; Kanagasabapathy, V. M. J. Am. Chem. Soc. 1985, 107, 1082-1083. b)

Hine, J.; Linden, S. M.; Kanagasabapathy, V. M. J. Org. Chem. 1985, 50, 5096-5099.

16 a) Curran, D. P.; Kuo, L. H. Tetrahedron Lett. 1995, 36, 6647-6650. b) Studies in allylation

reactions with the same catalyst: Curran, D. P.; Kuo, L. H. J. Org. Chem. 1994, 59, 3259-3261.

17 a) Schreiner, P. R.; Wittkopp, A. Org. Lett. 2002, 4, 217-220. b) Wittkopp, A.; Schreiner, P. R.

Chem. Eur. J. 2003, 9, 407-414.

18 a) Etter, M. C.; Panunto, T. W. J. Am. Chem. Soc. 1988, 110, 5896-5897. b) Etter, M. C. Acc.

Chem. Res. 1990, 23, 120-126. c) Etter, M. C.; Urbañczyk-Lipkowska, Z.; Zia-Ebrahimi, M.;

Panunto, T. W. J. Am. Chem. Soc. 1990, 112, 8415-8426.�

Page 162: Organocatalysis and Continuous Flow Reactors as Enabling Tools

160

19 a) Pihko, P. M. Hydrogen Bonding in Organic Synthesis, Wiley-VCH: Weinheim, 2009. b) Zhang,

Z.; Schreiner, P. R. Chem. Soc. Rev. 2009, 38, 1187-1198. c) Yu, X.; Wang, W. Chem. Asian J.

2008, 3, 516-532. d) Connon, S. J. Chem. Commun. 2008, 2499-2510. e) Akiyama, T. Chem. Rev.

2007, 107, 5744-5758. f) Doyle, A. G.; Jacobsen, E. N. Chem. Rev. 2007, 107, 5713-5743. g)

Connon, S. J. Chem. Eur. J. 2006, 12, 5418-5427. h) Taylor, M. S.; Jacobsen, E. N. Angew. Chem.

Int. Ed. 2006, 45, 1520-1543. i) Takemoto, Y. Org. Biomol. Chem. 2005, 3, 4299-4306. j) Pihko,

P. M. Angew. Chem. Int. Ed. 2004, 43, 2062-2064. k) Schreiner, P. M. Chem. Soc. Rev. 2003, 32,

289-296.

20 For Strecker reaction: a) Sigman, M. S.; Jacobsen, E. N. J. Am. Chem. Soc. 1998, 120, 4901-4902.

b) Vachal, P.; Jacobsen, E. N. Org. Lett. 2000, 2, 867-870. c) Sigman, M. S.; Vachal, P.; Jacobsen,

E. N. Angew. Chem. Int. Ed. 2000, 39, 1279-1281. d) Su, J. T.; Vachal, P.; Jacobsen, E. N. Adv.

Synth. Cat. 2001, 343, 197-200. e) Vachal, P.; Jacobsen, E. N. J. Am. Chem. Soc. 2002, 124,

10012-10014. f) Wenzel, A. G.; Lalonde, M. P.; Jacobsen, E. N. Synlett 2003, 1919-1922.

21 a) Wenzel, A. G.; Jacobsen, E. N. J. Am. Chem. Soc. 2002, 124, 12964-12965. b) Taylor, M. S.;

Tokunaga, N.; Jacobsen, E. N. Angew. Chem. Int. Ed. 2005, 44, 6700-6704. c) Yoon, T. P.;

Jacobsen, E. N. Angew. Chem. Int. Ed. 2005, 44, 466-468. d) Joly, G. D.; Jacobsen, E. N. J. Am.

Chem. Soc. 2004, 126, 4102-4103. e) Taylor, M. S.; Jacobsen, E. N. J. Am. Chem. Soc. 2004, 126,

10558-10559. f) Raheem, I. T.; Jacobsen, E. N. Adv. Synth. Cat. 2005, 347, 1701-1708. g) Fuerst,

D. E.; Jacobsen, E. N. J. Am. Chem. Soc. 2005, 127, 8964-8965.�

22 a) Okino, T.; Hoashi, Y.; Takemoto, Y. J. Am. Chem. Soc. 2003, 125, 12672-12673. b) Okino, T.;

Hoashi, Y.; Furukawa, T.; Xu, X.; Takemoto, Y. J. Am. Chem. Soc. 2005, 127, 119-125.

23 Li, X.; Deng, H.; Zhang, B.; Li, J.; Zhang, L.; Luo, S.; Cheng, J.-P. Chem. Eur. J. 2010, 16, 450-

455.

24 See e.g. a) Allu, S.; Molleti, N.; Panem, R.; Singh, V. K. Tetrahedron Lett. 2011, 52, 4080-4083.

b) Gu, Q.; You, S.-L. Org. Lett. 2011, 13, 5192-5195. c) Kyle, A. F.; Jakubec, P.; Dixon, D. J.;

Cockfield, D. M.; Cleator, E.; Skidmore, J. Chem. Commun. 2011, 10037-10039. d) Tan, B.;

Hérnandez-Torres, G.; Barbas III, C. F. Angew. Chem. Int. Ed. 2012, 51, 5381-5385. e)

Bifunctional urea in cooperative catalysis with CuOTf: Yang, T.; Ferrali, A.; Sladojevic, F.;

Campbell, L.; Dixon, D. J. J. Am. Chem. Soc. 2009, 131, 9140-9141.

25 Zhang, Z.-H.; Dong, X.-J.; Chen, D.; Wang, C.-J. Chem. Eur. J. 2008, 14, 8780-8783.

26 For a discussion on privileged structures, see Yoon, T. P.; Jacobsen, E. N. Science 2003, 299,

1691-1693.

Page 163: Organocatalysis and Continuous Flow Reactors as Enabling Tools

161

27 a) Song, C. E. Cinchona Alkaloids in Synthesis & Catalysis: Ligands, Immobilization and

Organocatalysis, Wiley-VCH: Weinheim, 2009. See also references cited therein. b) Marcelli, T.;

Hiemstra, H. Synthesis 2010, 1229-1279.

28 Pracejus, H. Liebigs Ann. 1960, 634, 9-22.

29 a) Wynberg, H.; Helder, R. Tetrahedron Lett. 1975, 16, 4057-4060. b) Hermann, K.; Wynberg, H.

J. Org. Chem. 1979, 44, 2238-2244. c) Hiemstra, H.; Wynberg, H. J. Am. Chem. Soc. 1981, 103,

417-430.

30 a) Li, B.-J.; Jiang, L.; Liu, M.; Chen, Y.-C.; Ding, L.-S.; Wu, Y. Synlett 2005, 603-606. b)

Vakulya, B.; Varga, S.; Csámpai, A.; Soós, T. Org. Lett. 2005, 7, 1967-1969. c) Ye, J.; Dixon, D.

J.; Hynes, P. S. Chem. Commun. 2005, 4481-4483. d) McCooey, S. H.; Connon, S. J. Angew.

Chem. Int. Ed. 2005, 44, 6367-6370. NB! Each group synthesized several derivatives and

stereoisomers. The scheme illustrates the optimized catalysts which gave best results in model

reactions.

31 Brunner, H.; Bügler, J.; Nuber, B. Tetrahedron: Asymmetry 1995, 6, 1699-1702.

32 a) Mitsunobu, O.; Yamada, M. Bull. Chem. Soc. Jpn. 1967, 40, 2380-2382. b) Mitsunobu, O.;

Yamada, M.; Mukaiyama, T. Bull. Chem. Soc. Jpn. 1967, 40, 935-939.

33 Staudinger, H.; Meyer, J. Helv. Chim. Acta 1919, 2, 635-649.

34 Li, H.; Wang, Y.; Tang, L.; Deng, L. J. Am. Chem. Soc. 2004, 126, 9906-9907.

35 Marcelli, T.; van der Haas, R. N. S.; van Maarseveen, J. H.; Hiemstra, H. Angew. Chem. Int. Ed.

2006, 45, 929-931.

36 Bernal, P.; Fernández, R.; Lassaletta, J. M. Chem. Eur. J. 2010, 16, 7714-7718.

37 Liu, Y.; Sun, B.; Wang, B.; Wakem, M.; Deng, L. J. Am. Chem. Soc. 2009, 131, 418-419.

38 Dickmeiss, G.; De Sio, V.; Udmark, J.; Poulsen, T. B.; Marcos, V.; Jørgensen, K. A. Angew.

Chem. Int. Ed. 2009, 48, 6650-6653.

39 a) Palacio, C.; Connon, S. J. Org. Lett. 2011, 13, 1298-1301. b) Palacio, C.; Connon, S. J. Chem.

Commun. 2012, 2849-2851.

40 Malerich, J. P.; Hagihara, K.; Rawal, V. H. J. Am. Chem. Soc. 2008, 130, 14416-14417.

Page 164: Organocatalysis and Continuous Flow Reactors as Enabling Tools

162

41 a) Amendola, V.; Fabbrizzi, L.; Mosca, L.; Schmidtchen, F.-P. Chem. Eur. J. 2011, 17, 5972-

5981. b) Amendola, V.; Bergamaschi, G.; Boiocchi, M.; Fabbrizzi, L.; Milani, M. Chem. Eur. J.

2010, 16, 4368-4380.

42 Quiñonero, D.; Prohens, R.; Garau, C.; Frontera, A.; Ballester, P.; Costa, A.; Deyà, P. M. Chem.

Phys. Lett. 2002, 351, 115-120.

43 Albrecht, �.; Dickmeiss, G.; Acosta, F. B.; Rodríguez-Escrich, C.; Davis, R. L.; Jørgensen, K. A.

J. Am. Chem. Soc. 2012, 134, 2543-2546.

44 For recent reviews of organocatalytic reactions in the synthesis of natural products, see: a)

Grondal, C.; Jeanty, M.; Enders, D. Nat. Chem. 2010, 2, 167-178. b) Marqués-López, E.; Herrera,

R. P.; Christmann, M. Nat. Prod. Rep. 2010, 27, 1138-1167. c) de Figueiredo, R. M.; Christmann,

M. Eur. J. Org. Chem. 2007, 2575-2600.

45 Hynes, P. S.; Stupple, P. A.; Dixon, D. J. Org. Lett. 2008, 10, 1389-1391.

46 a) Hoashi, Y.; Yabuta, T.; Takemoto, Y. Tetrahedron Lett. 2004, 45, 9185-9188. b) Hoashi, Y.;

Yabuta, T.; Yuan, P.; Miyabe, H.; Takemoto, Y. Tetrahedron 2006, 62, 365-374.

47 a) Raheem, I. T.; Thiara, P. S.; Peterson, E. A.; Jacobsen, E. N. J. Am. Chem. Soc. 2007, 129,

13404-13405. b) Mergott, D. J.; Zuend, S. J.; Jacobsen, E. N. Org. Lett. 2008, 10, 745-748.

48 Bui, T.; Syed, S.; Barbas III, C. F. J. Am. Chem. Soc. 2009, 131, 8758-8759.

49 Bassas, O.; Huuskonen, J.; Rissanen, K.; Koskinen, A. M. P. Eur. J. Org. Chem. 2009, 1340.

50 For reviews on Meldrum’s acid, see: a) McNab, H. Chem. Soc. Rev. 1978, 7 ,345-358. b) Chen,

B.-C. Heterocycles 1991, 32, 529-597. c) Bonifácio, V. D. B. Synlett 1993, 1649-1650. d) Gaber,

A. E.-A. M.; McNab, H. Synthesis, 2001, 2059-2074. d) Ivanov, A. S. Chem. Soc. Rev. 2008, 37,

789-811. e) Dumas, A. M.; Fillion, E. Acc. Chem. Res. 2010, 43, 440-454.

51 a) Arnett, E. M.; Maroldo, S. G.; Schilling, S. L.; Harrelson, J. A. J. Am. Chem. Soc. 1984, 106,

6759-6767. b) Pihlaja, K.; Seilo, M. Acta Chem. Scand. 1968, 22, 3053. c) Pihlaja, K.; Seilo, M.

Acta Chem. Scand. 1969, 23, 3003.

52 a) Hirao, H.; Ohwada, T. J. Phys. Chem. A 2003, 107, 2875-2881. b) Nakamura, S.; Hirao, H.;

Ohwada, T. J. Org. Chem. 2004, 69, 4309-4316.

53 a) Ramachary, D. B.; Chowdary, N. S.; Barbas, C. F. III Angew. Chem. Int. Ed. 2003, 42, 4233-

4237. b) Watanabe, T.; Knöpfel, T. F.; Carreira, E. M. Org. Lett. 2003, 5, 4557-4558. c) Knöpfel,

Page 165: Organocatalysis and Continuous Flow Reactors as Enabling Tools

163

T. F.; Zarotti, P.; Ichikawa, T.; Carreira, E. M. J. Am. Chem. Soc. 2005, 127, 9682. d) Dumas, A.

M.; Fillion, E. Acc. Chem. Res. 2010, 43, 440-454.

54 Kleczkowska, E.; Sas, W. Polish J. Chem. 2007, 81, 1457-1464.

55 Didier, D.; Sergeyev, S. ARKIVOC 2009, xiv, 124-134.

56 Kataja, A. O.; Koskinen, A. M. P. ARKIVOC 2010, ii, 205-223.

57 a) Menguy, L.; Couty, F. Tetrahedron: Asymmetry 2010, 21, 2385-2389. b) Kimmel, K. L.;

Weaver, J. D.; Ellman, J. A. Chem. Sci. 2012, 3, 121-125.

58 Hamza, A.; Schubert, G.; Soós, T.; Pápai, I. J. Am. Chem. Soc. 2006, 128, 13151-13160.

59 Benoit, R. L.; Lefebvre, D.; Fréchette, M. Can. J. Chem. 1987, 65, 996-1001.

60 Olmstead, W. N.; Bordwell, F. G. J. Org. Chem. 1980, 45, 3299-3305.�

61 Zhu, R.; Zhang, D.; Wu, J.; Liu, C. Tetrahedron: Asymmetry 2006, 17, 1611-1616.

62 Reichardt, C. Solvents and Solvent Effects in Organic Chemistry, 3rd Ed.; Wiley-VCH: Weinheim

2003, p. 107.

63 a) Andrés, J. M.; Manzano, R.; Pedrosa, R. Chem. Eur. J. 2008, 14, 5116-5119. b) Jiang, X.;

Zhang, Y.; Liu, X.; Zhang, G.; Lai, L.; Wu, L.; Zhang, J.; Wang, R. J. Org. Chem. 2009, 74,

5562-5567. c) Alma�i, D.; Alonso, D. A.; Gómez-Bengoa, E.; Najéra, C. J. Org. Chem. 2009, 74,

6163-6168. d) Camps, P.; Galdeano, C.; Muñoz-Torrero, D.; Rull, J.; Calvet, T.; Font-Bardia, M.

Tetrahedron: Asymmetry 2011, 22, 745-751. e) Nemoto, T.; Obuchi, K.; Tamura, S.; Fukuyama,

T.; Hamada, Y. Tetrahedron Lett. 2011, 52, 987-991. f) Bae, H. Y.; Some, S.; Oh, J. S.; Lee, Y.

S.; Song, C. E. Chem. Commun. 2011, 47, 9621-9623. g) Ma, Z.-w.; Liu, Y.-x.; Gao, X.; Tao, J.-c.

Tetrahedron: Asymmetry 2012, 23, 443-448.

64 Jakab, G.; Tancon, C.; Zhang, Z.; Lippert, K. M.; Schreiner, P. R. Org. Lett. 2012, 14, 1724-1727.

65 Pan, S. C.; Zhou, J.; List, B. Angew. Chem. Int. Ed. 2007, 46, 612-614.

66 Survey of –HNC(S)NH– moieties (numbering over 700) published in the Cambridge Structural

Database by September 2009. For a representative example of trans/trans, see CCDC-275454

(Ref. 21b), and for cis/trans, see CCDC-685194 (Ref. 47)

Page 166: Organocatalysis and Continuous Flow Reactors as Enabling Tools

164

67 a) Luo, J.; Xu, L.-W.; Hay, R. A.; Lu, Y. Org. Lett. 2009, 11, 437-440. b) Analogous cinchonine

derived sulfonamide catalyst originally reported for in Brunner, H.; Baur, M. A. Eur. J. Org.

Chem. 2003, 2854-2862.

68 a) Shi, X.; He, W.; Hua, L.; Zhang, X.; Zhang. S. Tetrahedron Lett. 2011, 52, 3204-3207. b) Zhao,

M.-X.; Tang, W.-H.; Chen, M.-X.; Wei, D.-K.; Dai, T.-L.; Shi, M. Eur. J. Org. Chem. 2011,

6078-6084. c) Li, W.; Wu, W.; Yu, F.; Huang, H.; Liang, X.; Ye, J. Org. Biomol. Chem. 2011, 9,

2505-2511. d) Li, H.; Zhang, X.; Shi, X.; Ji, N.; He, W.; Zhang, S.; Zhang, B. Adv. Synth. Catal.

2012, 354, 2264-2274.�

69 Carbamoylation of 102: Figner, G. C.; Dickerson, D. R.; Orlopp, D. E.; Ehrmantraut, J. W. J.

Med. Chem. 1964, 7, 572-573.

70 Procedure applied from: Nagashima, N.; Akamatsu, S.; Kawaminami, E.; Kawazoe, S.; Ogami, T.;

Matsumoto, Y.; Okada, M.; Suzuki, K.-I.; Tsukamoto, S.-I. Chem. Pharm. Bull. 2001, 49, 1420-

1432.

71 a) Ichiba, T.; Yoshida, W. Y.; Scheuer, P. J.; Higa, T.; Gravalos, D. G. J. Am. Chem. Soc. 1991,

113, 3173-3174. b) Higa, T.; Tanaka, J.-I.; Kitamura, A.; Koyama, T.; Takahashi, M.; Uchida, T.

Pure. Appl. Chem. 1994, 66, 2227-2230.

72 a) Wipf, P.; Lim, S. J. Am. Chem. Soc. 1995, 117, 558-559. b) Williams, D. R.; Brooks, D. A.;

Berliner, M. A. J. Am. Chem. Soc. 1999, 121, 4924-4925. c) Yokokawa, F.; Asano, T.; Shioiri, T.

Org. Lett. 2000, 2, 4169-4172. d) Yokokawa, F.; Asano, T.; Shioiri, T. Tetrahedron 2001, 57,

6311-6327. e) Smith, T. E.; Kuo, W.-H.; Bock, V. D.; Roizen, J. L.; Balskus, E. P.; Theberge, A.

B. Org Lett. 2007, 9, 1153-1155. f) Smith, T. E.; Kuo, W.-H.; Balskus, E. P.; Bock, V. D.; Roizen,

J. L.; Theberge, A. B.; Carroll, K. A.; Kurihara, T.; Wessler, J. D. J. Org. Chem. 2008, 73, 142-

150.

73 For recent publications, see: a) Koos, P.; Gross, U.; Polyzos, A.; O’Brien, M.; Baxendale, I. R.;

Ley, S. V. Org. Biomol. Chem. 2011, 9, 6903-6908. b) Browne, D. L.; Baumann, M.; Harji, B. H.;

Baxendale, I. R.; Ley, S. V. Org. Lett. 2011, 13, 3312-3315. c) Polyzos, A.; O’Brien, M.;

Petersen, T. P.; Baxendale, I. R.; Ley, S. V. Angew. Chem. Int. Ed. 2011, 50, 1190-1193. d)

Martin, L. J.; Marzinzik, A. L.; Ley, S. V.; Baxendale, I. R. Org. Lett. 2011, 13, 320-323. e)

Baumann, M.; Baxendale, I. R.; Brasholz, M.; Hayward, J. J.; Ley, S. V.; Nikbin, N. Synlett 2011,

1375-1380. f) Qian, Z.; Baxendale, I. R.; Ley, S. V. Chem. Eur. J. 2010, 16, 12342-12348.�

74 For reviews, see: a) Wegner, J.; Ceylan, S.; Kirschning, A. Adv. Synth. Cat. 2012, 354, 17-57. b)

Malet-Sanz, L.; Susanne, F. J. Med. Chem. 2012, 55, 4062-4098. c) Noël, T.; Buchwald, S. L.

Chem. Soc. Rev. 2011, 40, 5010-5029. d) Wegner, J.; Ceylan, S.; Kirschning, A. Chem. Commun.

Page 167: Organocatalysis and Continuous Flow Reactors as Enabling Tools

165

2011, 47, 4583-4592. e) Hartman, R. L.; McMullen, J. P.; Jensen, K. F. Angew. Chem. Int. Ed.

2011, 50, 7502-7519. f) Valera, F. E.; Quaranta, M.; Moran, A.; Blacker, J.; Armstrong, A.;

Cabral, J. T.; Blackmond, D. G. Angew. Chem. Int. Ed. 2010, 49, 2478-2485. g) Razzaq, T.;

Kappe, C. O. Chem. Asian J. 2010, 5, 1274-1289. h) Webb, D.; Jamison, T. F. Chem. Sci. 2010, 1,

675-680. i) Wiles, C.; Watts, P. Eur. J. Org. Chem. 2008, 1655-1671. j) Yoshida, J.-i.; Nagaki, A.;

Yamada, T. Chem. Eur. J. 2008, 14, 7450-7459. k) Kockmann, N.; Gottsponer, M.; Zimmermann,

B.; Roberge, D. M. Chem. Eur. J. 2008, 14, 7470-7477. l) Baxendale, I. R.; Hayward, J. J.;

Lanners, S.; Ley, S. V.; Smith, C. D. in Microreactors in Organic Synthesis and Catalysis, (Ed.:

Wirth, T.), Wiley-VCH, Weinheim, 2008, pp. 84-122. m) Mason, B. P.; Price, K. E.; Steinbacher,

J. L.; Bogdan, A. R.; McQuade, D. T. Chem. Rev. 2007, 107, 2300-2318.

75 For details on commercial equipment, see http://www.vapourtec.co.uk, http://www.uniqsis.com,

and http://www.syrris.com.

76 Saaby, S.; Knudsen, K. R.; Ladlow, M.; Ley, S. V. Chem. Commun. 2005, 2909-2911.

77 a) Selective DIBAL-H reduction of ester to aldehyde: Webb, D.; Jamison, T. F. Org. Lett. 2012,

14, 568-571. b) Friedel-Crafts aminoalkylation: Nagaki, A.; Togai, M.; Suga, S.; Aoki, N.; Mae,

K.; Yoshida, J.-i. J. Am. Chem. Soc. 2005, 127, 11666-11675.�

78 Usutani, H.; Tomida, Y.; Nagaki, A.; Okamoto, H.; Nokami, T.; Yoshida, J.-i. J. Am. Chem. Soc.

2007, 129, 3046-3047.�

79 For a recent example, see: Martin, R. E.; Morawitz, F.; Kuratli, C.; Alker, A. M.; Alanine, A. I.

Eur. J. Org. Chem. 2012, 47-52.

80 Damm, M.; Glasnov, T. N.; Kappe, C. O. Org. Process Res. Devel. 2010, 14, 215-224.

81 For handling of e.g. organic azides, see: a) Baumann, M., Baxendale, I. R.; Ley, S. V.; Nikbin, N.;

Smith, C. D.; Tierney, J. P. Org. Biomol. Chem. 2008, 6, 1577-1586. b) Baumann, M., Baxendale,

I. R.; Ley, S. V.; Nikbin, N.; Smith, C. D. Org. Biomol. Chem. 2008, 6, 1587-1593.

82 Hartman, R. L. Org. Process Res. Devel. 2012, 16, 870-887.

83 a) Nagy, K. D.; Shen, B.; Jamison, T. F.; Jensen, K. F. Org. Process Res. Devel. 2012, 16, 976-

981. b) Hornung, C. H.; Mackley, M. R. Chem. Eng. Sci. 2009, 64, 3889-3902.

84 For a discussion, see: Ley, S. V.; Baxendale, I. R. Chimia 2008, 62, 162-168.

Page 168: Organocatalysis and Continuous Flow Reactors as Enabling Tools

166

85 For recent examples of homogeneous metal catalysis, see e.g. a) Shu, W.; Pellegatti, L.; Oberli,

M. A.; Buchwald, S. L. Angew. Chem. Int. Ed. 2011, 50, 10665-10669. b) Shu, W.; Buchwald, S.

L. Angew. Chem. Int. Ed. 2012, 51, 5355-5358.

86 For selected reviews and discussions, see: a) Ley, S. V.; Baxendale, I. R.; Bream, R. N.; Jackson,

P. S.; Leach, A. G.; Longbottom, D. A.; Nesi, M.; Scott, J. S.; Storer, R. I.; Taylor, S. J. J. Chem.

Soc., Perkin Trans. 1 2000, 3815-4195. b) Clapham, B.; Reger, T. S.; Janda, K. D. Tetrahedron

2001, 57, 4637-4662. c) Leadbeater, N. E.; Marco, M. Chem. Rev. 2002, 102, 3217-3273. d)

Bräse, S.; Lauterwasser, F.; Ziegert, R. E. Adv. Synth. Catal. 2003, 345, 869-929. e) Benaglia, M.;

Puglisi, A.; Cozzi, F. Chem. Rev. 2003, 103, 3401-3429. f) Cozzi, F. Adv. Synth. Catal. 2006, 348,

1367-1390. g) Corma, A.; Garcia, H. Adv. Synth. Catal. 2006, 348, 1391-1412. h) Kristensen, T.

E.; Hansen, T. Eur. J. Org. Chem. 2010, 3179-3204. i) Chem. Rev. 2002, 102, issue 10. �

87 Baxendale, I. R.; Deeley, J.; Griffiths-Jones, C. M.; Ley, S. V.; Saaby, S.; Tranmer, G. K. Chem.

Commun. 2006, 2566-2568.

88 a) Bahmanyar, S.; Houk, K. N. J. Am. Chem. Soc. 2001, 123, 11273. (b) Bahmanyar, S.; Houk, K.

N.; Martin, H. J.; List, B. J. Am. Chem. Soc. 2003, 125, 2475. (c) Hoang, L.; Bahmanyar, S.;

Houk, K. N.; List, B. J. Am. Chem. Soc. 2003, 125, 16. (d) List, B.; Hoang, L.; Martin, H. J. Proc.

Natl. Acad. Sci. U.S.A. 2004, 101, 5839. (e) Clemente, F. R.; Houk, K. N. Angew. Chem. Int. Ed.

2004, 43, 5765. e) First experimental observation of the catalyst-enamine species: Schmid, M. B.;

Zeitler, K.; Gschwind, R. M. Angew. Chem. Int. Ed. 2010, 49, 4997-5003.

89 a) Marigo, M.; Wabnitz, T. C.; Fielenbach, D.; Jørgensen, K. A. Angew. Chem. Int. Ed. 2005, 44,

794-797. b) Hayashi, Y.; Gotoh, H.; Hayashi, T.; Shoji, M. Angew. Chem. Int. Ed. 2005, 44,

4212-4215.�

90 Recent studies on the origin of selectivity with diarylprolinol ethers: a) Burés, J.; Armstrong, A.;

Blackmond, D. G. J. Am. Chem. Soc. 2012, 134, 6741-6750. b) Seebach, D.; Sun, X.; Sparr, C.;

Ebert, M.-O.; Schweizer, W. B.; Beck, A. K. Helv. Chim. Acta 2012, 95, 1064-1078. c) Burés, J.;

Armstrong, A.; Blackmond, D. G. J. Am. Chem. Soc. 2011, 133, 8822-8825. d) Patora-

Komisarska, K.; Benohoud, M.; Ishikawa, H.; Seebach, D.; Hayashi, Y. Helv. Chim. Acta 2011,

94, 719-745. e) Seebach, D.; Grošelj, U.; Badine, D. M.; Schweizer, W. B.; Beck, A. K. Helv.

Chim. Acta 2008, 91, 1999-2034.

91 For a comprehensive overview, see Mahrwald, R. (Ed.) Modern Aldol Reactions, Wiley-VCH:

Weinheim, 2004.

92 a) Rutter, W. J. Fed. Proc. Am. Soc. Exp. Biol. 1964, 23, 1248-1257. b) Lai, C. Y.; Nakai, N.;

Chang, D. Science 1974, 183, 1204-1206.

Page 169: Organocatalysis and Continuous Flow Reactors as Enabling Tools

167

93 Fessner, W.-D.; Schneider, A.; Held, H.; Sunerius, G.; Walter, C.; Hixon, M.; Schloss, J. V.

Angew. Chem. Int. Ed. 1996, 35, 2219-2221.

94 a) Wagner, J.; Lerner, R. A.; Barbas, C. F., III Science 1995, 270, 1797-1800. b) Barbas, C. F., III;

Heine, A.; Zhong, G.; Hoffmann, T.; Gramatikova, S.; Björnestedt, R.; List, B.; Anderson, J.;

Stura, E. A.; Wilson, I. A.; Lerner, R. A. Science 1997, 278, 2085-2092.

95 a) Hajos, Z. G.; Parrish, D. R. German Patent DE 2102623, 1971. b) Eder, U.; Sauer, G. R.;

Wiechert, R. German Patent DE 2014757, 1971. c) Eder, U.; Sauer, G. R.; Wiechert, R. Angew.

Chem. Int. Ed. 1971, 10, 496-497. d) Hajos, Z. G.; Parrish, D. R. J. Org. Chem. 1974, 39, 1615-

1621.

96 Sakthivel, K.; Notz, W.; Bui, T.; Barbas III, C. F. J. Am. Chem. Soc. 2001, 123, 5260-5267.

97 Northrup, A. B.; MacMillan, D. W. C. J. Am. Chem. Soc. 2002, 124, 6798-6799.

98 For a recent review, see: Bisai, V.; Bisai, A.; Singh, V. K. Tetrahedron 2012, 68, 4541-4580 and

references cited therein.

99 Woodward, R. B.; Logusch, E.; Nambiar, K. P.; Sakan, K.; Ward, D. E.; Au-Yeung, B.-W.;

Balaram, P.; Browne, L. J.; Card, P. J.; Chen, C. H.; Chênevert, R. B.; Fliri, A.; Frobel, K.; Gais,

H.-J.; Garratt, D. G.; Hayakawa, K.; Heggie, W.; Hesson, D. P.; Hoppe, D.; Hoppe, I.; Hyatt, J.

A.; Ikeda, D.; Jacobi, P. A.; Kim, K. S.; Kobuke, Y.; Kojima, K.; Krowicki, K.; Lee, V. J.;

Leutert, T.; Malchenko, S.; Martens, J.; Matthews, R. S.; Ong, B. S.; Press, J. B.; Rajan Babu, T.

V.; Rousseau, G.; Sauter, H. M.; Suzuki, M.; Tatsuta, K.; Tolbert, L. M.; Truesdale, E. A.;

Uchida, I.; Ueda, Y.; Uyehara, T., Vasella, A. T.; Vladuchick, W. C.; Wade, P. A.; Williams, R.

M.; Wong, H. N.-C. J. Am. Chem. Soc. 1981, 103, 3210-3213.

100 Pihko, P. M.; Erkkilä, A. Tetrahedron Lett. 2003, 44, 7607-7609.

101 Zhang, S.; Duan, W.; Wang, W. Adv. Synth. Cat. 2006, 348, 1228-1234.

102 Liu, K.; Chougnet, A.; Woggon, W.-D. Angew. Chem. Int. Ed. 2008, 47, 5827-5829.

103 a) Liu, K.; Woggon, W.-D. Eur. J. Org. Chem. 2010, 1033-1036. b) Volz, M.; Brohmer, M. C.;

Nieger, M.; Bräse, S. Synlett 2009, 550-553.�

104 Carpenter, J.; Northrup, A. B.; Chung, d. M.; Wiener, J. J. M.; Kim, S.-G.; MacMillan, D. W. C.

Angew. Chem. Int. Ed. 2008, 47, 3568-3572.

Page 170: Organocatalysis and Continuous Flow Reactors as Enabling Tools

168

105 List, B.; Pojarliev, P.; Castello, C. Org. Lett. 2001, 3, 573-575.

106 For reviews, see: a) Mak, X. Y.; Laurino, P.; Seeberger, P. H. Beilstein J. Org. Chem. 2009, 5, No.

19. b) Yurev, R.; Strompen, S.; Liese, A. Beilstein J. Org. Chem. 2011, 7, 1449-1467 (enzymatic

processes).

107 Solodenko, W.; Jas, G.; Kunz, U.; Kirschning, A. Synthesis 2007, 583-589.

108 Burguete, M. I.; Cornejo, A.; García-Verdugo, E.; Gil, M. J.; Luis, S. V.; Mayoral, J. A.;

Martínez-Merino, V.; Sokolova, M. J. Org. Chem. 2007, 72, 4344-4350.

109 Pericàs, M. A.; Herrerías, C. I.; Solà, L. Adv. Synth. Cat. 2008, 350, 927-932.

110 Tsubogo, T.; Yamashita, S.; Kobayashi, S. Chem. Eur. J. 2012, in press.

http://dx.doi.org/10.1002/chem.201202896

111 Matsuda, T.; Watanabe, K.; Harada, T.; Nakamura, K.; Arita, Y.; Misumi, Y.; Ichikawa, S.;

Ikariya, T. Chem. Commun. 2004, 2286-2287.

112 Baxendale, I. R.; Griffiths-Jones, C. M.; Ley, S. V.; Tranmer, G. K. Synlett 2006, 427-430.

113 a) Newton, S.; Ley, S. V.; Arcé, E. C.; Grainger, D. M. Adv. Synth. Cat. 2012, 354, 1805-1812. b)

Balogh, S.; Farkas, G.; Madarász, J.; Szöll�sy, Á.; Kovács, J.; Darvas, F.; Ürge, L.; Bakos, J.

Green Chem. 2012, 14, 1146-1151. c) Rueping, M.; Bootwicha, T.; Sugiono, E. Beilstein J. Org.

Chem. 2012, 8, 300-307.�

114 Hafez, A. M.; Taggi, A. E.; Dudding, T.; Lectka, T. J. Am. Chem. Soc. 2001, 123, 10853-10859.

115 a) Bernstein, D.; France, S.; Wolfer, J.; Lectka, T. Tetrahedron: Asymmetry 2005, 16, 3481-3483.

b) France, S.; Bernstein, D.; Weatherwax, A.; Lectka, T. Org. Lett. 2005, 7, 3009-3012.

116 Odedra, A.; Seeberger, P. H. Angew. Chem. Int. Ed. 2009, 48, 2699-2702.

117 a) Alza, E.; Rodríguez-Escrich, C.; Sayalero, S.; Bastero, A.; Pericàs, M. A. Chem. Eur. J. 2009,

15, 10167-10172. b) Ayats, C.; Henseler, A. H.; Pericàs, M. A. ChemSusChem 2012, 5, 320-325.

118 Demuynck, A. L. W.; Peng, L.; de Clippel, F.; Vanderleyden, J.; Jacobs, P. A.; Sels, B. F. Adv.

Synth. Cat. 2011, 353, 725-732.

119 a) Review on Suzuki-type sp2-sp3 C-C-couplings: Chemler, S. R.; Trauner, D.; Danishefsky S. J.

Angew. Chem. Int. Ed. 2001, 40, 4544-4568. b) Maddess, M. L.; Ley, S. V., unpublished results.

Page 171: Organocatalysis and Continuous Flow Reactors as Enabling Tools

169

120 a) Paterson, I.; Gibson, K. R.; Oballa, R. M. Tetrahedron Lett. 1996, 37, 8585-8588. b) Evans, D.

A.; Coleman, P. J.; Côté, B. J. Org. Chem. 1997, 62, 788-789. c) Paton, R. S.; Goodman, J. M.

Org. Lett. 2006, 8, 4299-4302.

121 Brasholz, M.; Ley, S. V., unpublished results.

122 Fujioka, H.; Sawama, Y.; Murata, N.; Okitsu, T.; Kubo, O.; Matsuda, S.; Kita, Y. J. Am. Chem.

Soc. 2004, 126, 11800-11801.

123 a) Samanta, S.; Liu, J.; Dodda, R.; Zhao, C.-G. Org. Lett. 2005, 7, 5321-5323. b) Stodulski, M.;

Jazwinzki, J.; Mlynarski, J. Eur. J. Org. Chem. 2008, 5553-5562.�

124 Samanta, S.; Zhao, C.-G. Tetrahedron Lett. 2006, 47, 3383-3386.

125 a) Saksena, A. K.; Mangiaracina, P. Tetrahedron Lett. 1983, 24, 273-276. b) Evans, D. A.;

Chapman, K. T.; Carreira, E. M. J. Am. Chem. Soc. 1988, 110, 3560-3578.

126 Evans, D. A.; Hoveyda, A. H. J. Am. Chem. Soc. 1990, 112, 6447-6449.

127 Baumann, M.; Baxendale, I. R.; Martin, L. J.; Ley, S. V. Tetrahedron 2009, 65, 6611-6625.

128 Raj, M.; Vishnumaya; Ginotra, S. K.; Singh, V. K. Org. Lett. 2006, 8, 4097-4099.

129 a) Sullivan, G. R.; Dale, J. A; Mosher, H. S. J. Am. Chem. Soc. 1973, 95, 512-519. b) Dale, J. A.;

Mosher, H. S. J. Org. Chem. 1973, 38, 2143-2147.

130 Borohydride immobilization on quaternary ammonium resin first reported by Gibson, H. W.;

Bailey, F. C. Chem. Commun. 1977, 815.

131 a) Rychnovsky, S. D.; Skalitzky, D. J. Tetrahedron Lett. 1990, 31, 945-948. b) Rychnovsky, S. D.;

Rogers, B. N.; Richardson, T. I. Acc. Chem. Res. 1998, 31, 9-17.

132 Cohen T.; Jeong, I.-H.; Mudryk, B.; Bhupathy, M.; Awad, M. A. J. Org. Chem. 1990, 55, 1528-

1536.

133 Bhattacharyya, S.; Rana, S.; Gooding, O W.; Labadie, J. Tetrahedron Lett. 2003, 44, 4957-4960.

134 Smith, C. J.; Smith, C. D.; Nikbin, N.; Ley, S. V.; Baxendale, I. R. Org. Biomol. Chem. 2011, 9,

1927-1937.

Page 172: Organocatalysis and Continuous Flow Reactors as Enabling Tools

170

135 For early studies, see: a) Benford, G. A.; Ingold, C. K. J. Chem. Soc. 1938, 929-955. b) Hughes,

E. D.; Ingold, C. K.; Reed, R. I. J. Chem. Soc. 1950, 2400-2440. c) Halberstadt, E. S.; Hughes, E.

D.; Ingold, C. K. J. Chem. Soc. 1950, 2441-2452.

136 Bunton, C. A.; Hughes, E. D.; Ingold, C. K.; Jacobs, D. I. H.; Jones, M. H.; Minkoff, G. J.; Reed,

R. I. J. Chem. Soc. 1950, 2628-2652.

137 Nitration of vanillin with nitric, nitrous or mixed acid: a) Menke, A. E.; Bentley, W. B. J. Am.

Chem. Soc. 1898, 20, 316-317. b) Vogl, W. Monatsh. 1899, 20, 323-400. c) Bentley, W. B. Am.

Chem. J. 1900, 24, 171-181. d) Hayduck, F. Chem. Ber. 1903, 36, 2930-2936. e) Finger, H.;

Schott, W. J. Prakt. Chem. 1927, 115, 281-291. f) Bazanova, G. V.; Stotskii, A. A. Zh. Org. Khim.

1983, 19, 1845-1850 & 2124-2130. g) Benechie, M.; Delpech, B.; Khuong-Huu, Q.; Khuong-Huu,

F. Tetrahedron 1992, 48, 1895-1910. h) Grenier, J. L.; Cotelle, N.; Catteau, J. P.; Cotelle, P. J.

Phys. Org. Chem. 2000, 13, 511-517. i) Malecki, N.; Carato, P.; Rigo, B.; Goossens, J.-F.;

Houssin, R.; Bailly, C.; Henichart, J.-P. Bioorg. Med. Chem. 2004, 12, 641-648. j) Bailey, K.;

Tan, E. W. Bioorg. Med. Chem. 2005, 13, 5740-5749. k) Xia, C.-n.; Liu, F.; Hu, W.-x.; Li, H.-b.

Bioorg. Med. Chem. Lett. 2008, 18, 6553-6557. l) Kiss, L. E.; Ferreira, H.; Torrão, L.; Bonifácio,

M. J.; Palma, P. N.; Soares-da-Silva, P.; Learmonth, D. A. J. Med. Chem. 2010, 53, 3396-3411. m)

Schobert, R.; Biersack, B.; Dietrich, A.; Effenberger, K.; Knauer, S.; Müller, T. J. Med. Chem.

2010, 53, 6595-6602. n) Nitsche, C.; Steuer, C.; Klein, C. D. Bioorg. Med. Chem. 2011, 19, 7318-

7337.

138 Nitration of vanillin with other reagents: a) Francis, F. Chem. Ber. 1906, 39, 3798-3804. b)

Kaupp, G.; Schmeyers, J. J. Org. Chem. 1995, 60, 5494-5503. c) Rodrigues, J. A. R.; Filho, A. P.

de O.; Moran, P. J. S.; Custodio, R. Tetrahedron 1999, 55, 6733-6738. d) Iranpoor, N.;

Firouzabadi, H.; Heydari, R.; Shiri, M. Synth. Commun. 2005, 35, 263-270. e) Selvam, J. J. P.;

Suresh, V.; Rajesh, K.; Reddy, S. R.; Venkateswarlu, Y. Tetrahedron Lett. 2006, 47, 2507-2509. f)

Anuradha, V.; Srinivas, P. V.; Aparna, P.; Rao, J. M. Tetrahedron Lett. 2006, 47, 4933-4935. g)

Lalitha, A.; Sivakumar, K. Synth. Commun. 2008, 38, 1745-1752.

139 Schramm, R. M.; Westheimer, F. H. J. Am. Chem. Soc. 1948, 70, 460-462.

140 Moodie, R. B.; Schofield, K. Acc. Chem. Res. 1976, 9, 287-292.

141 See e.g. Olah, G. A.; Kuhn, S. J. J. Am. Chem. Soc. 1964, 86, 1067-1070 and references cited

therein.

142 Cotelle, P.; Catteau, J. P. Synth. Commun. 1996, 26, 4105-4112.

Page 173: Organocatalysis and Continuous Flow Reactors as Enabling Tools

171

143 a) Brocklehurst, C. E.; Lehmann, H.; La Vecchia, L. Org. Process Res. Devel. 2011, 15, 1447-

1453. b) Gage, J. R.; Guo, X.; Tao, J.; Zheng, C. Org. Process Res. Devel. 2012, 16, 930-933.

144 Da, C.-S.; Che, L.-P.; Guo, Q.-P.; Wu, F.-C.; Ma, X.; Jia, Y.-N. J. Org. Chem. 2009, 74, 2541-

2546.

145 a) Hoffmann, R. W.; Bewersdorf, M.; Krüger, M.; Mikolaiski, W.; Stürmer, R. Chem. Ber. 1991,

124, 1243-1252. b) Schneider, C.; Hansch, M.; Weide, T. Chem. Eur. J. 2005, 11, 3010-3021. c)

Hiraoka, C.; Matsuda, M.; Suzuki, Y.; Fujieda, S.; Tomita, M.; Fuhshuku, K.-i.; Obata, R.;

Nishiyama, S.; Sugai, T. Tetrahedron: Asymmetry 2006, 17, 3358-3367. d) Barbero, A.;

Blakemore, D. C.; Fleming, I.; Wesley, R. N. J. Chem. Soc., Perkin Trans. 1, 1997, 1329-1352. e)

Evans, D. A.; Allison, B. D.; Yang, M. G.; Masse, C. E. J. Am. Chem. Soc. 2001, 123, 10840-

10852. NB! In this reference the spectral data corresponds to anti-structure based on references

143a-d, although syn-structure is depicted.

146 a) Mascarenhas, C. M.; Duffey, M. O.; Liu, S.-Y.; Morken, J. P. Org. Lett. 1999, 1, 1427-1429. b)

Aftab, T.; Carter, C.; Christlieb, M.; Hart, J.; Nelson, A. J. Chem. Soc., Perkin Trans. 1, 2000,

711-722.

Page 174: Organocatalysis and Continuous Flow Reactors as Enabling Tools
Page 175: Organocatalysis and Continuous Flow Reactors as Enabling Tools
Page 176: Organocatalysis and Continuous Flow Reactors as Enabling Tools

9HSTFMG*afagbe+