11.03.2014 Views

Improved Methodology for the Preparation of Chiral Amines

Improved Methodology for the Preparation of Chiral Amines

Improved Methodology for the Preparation of Chiral Amines

SHOW MORE
SHOW LESS

You also want an ePaper? Increase the reach of your titles

YUMPU automatically turns print PDFs into web optimized ePapers that Google loves.

<strong>Improved</strong> <strong>Methodology</strong> <strong>for</strong> <strong>the</strong> <strong>Preparation</strong> <strong>of</strong><br />

<strong>Chiral</strong> <strong>Amines</strong><br />

(Important <strong>Chiral</strong> Building Blocks in Pharmaceutical<br />

Drugs and Natural Products Syn<strong>the</strong>sis)<br />

Mohamed Mahmoud El-Shazly<br />

A <strong>the</strong>sis submitted in partial fulfillment<br />

<strong>of</strong> <strong>the</strong> requirements <strong>for</strong> <strong>the</strong> degree <strong>of</strong><br />

Doctor <strong>of</strong> Philosophy<br />

in<br />

Organic Syn<strong>the</strong>tic Chemistry<br />

Date <strong>of</strong> Defense: August 03, 2009<br />

Approved Thesis Committee<br />

Pr<strong>of</strong>. Dr. Thomas Nugent<br />

Pr<strong>of</strong>essor <strong>of</strong> Organic Chemistry<br />

Jacobs University Bremen<br />

Pr<strong>of</strong>. Dr. Nikolai Kuhnert<br />

Pr<strong>of</strong>essor <strong>of</strong> Organic Chemistry<br />

Jacobs University Bremen<br />

Dr. Pralhad Ganeshpure<br />

Indian Petrochemicals Corporation Limited,<br />

India<br />

School <strong>of</strong> Engineering and Science, Jacobs University, Bremen, Germany.


Declaration<br />

I herewith declare that this <strong>the</strong>sis is my own work and that I have used<br />

only <strong>the</strong> sources listed. No part <strong>of</strong> this <strong>the</strong>sis has been accepted or is<br />

currently being submitted <strong>for</strong> <strong>the</strong> conferral <strong>of</strong> any degree at this<br />

university or elsewhere.<br />

Mohamed El-Shazly<br />

Bremen


This dissertation is dedicated to all those people<br />

who have always given me <strong>the</strong> love, trust, and support<br />

to come to this stage <strong>of</strong> my life<br />

-To My Family-


Abstract<br />

The importance <strong>of</strong> α-chiral amines as building blocks in pharmaceutical drugs, natural<br />

products, fine chemicals and agrochemicals have encouraged scientists to develop different<br />

methodologies <strong>for</strong> <strong>the</strong>ir preparation. Their main goal was to develop a step wise efficient and<br />

low waste production methodology which utilizes inexpensive starting material <strong>for</strong> <strong>the</strong><br />

syn<strong>the</strong>sis <strong>of</strong> α-chiral amines in high yields and enantioselectivity. Different methodologies<br />

have been developed aiming to meet <strong>the</strong>se criteria. These strategies are discussed and <strong>the</strong>ir<br />

importance and limitations are critically analyzed.<br />

Reductive amination is a powerful methodology <strong>for</strong> <strong>the</strong> syn<strong>the</strong>sis <strong>of</strong> chiral amines in high<br />

yields and enantioselectivity. It is a two step strategy beginning from <strong>the</strong> prochiral carbonyl<br />

compound to <strong>the</strong> primary chiral amine. The historical development and <strong>the</strong> latest milestones<br />

in this field are discussed in chapter three. Different drugs and natural products which are<br />

prepared utilizing reductive amination as a key step in <strong>the</strong>ir syn<strong>the</strong>sis are summarized in<br />

chapter four.<br />

Reductive amination utilizing chiral auxiliary/Lewis acid/ heterogeneous catalyst/ molecular<br />

hydrogen has been investigated in our group over <strong>the</strong> last five years. This combination<br />

allowed <strong>the</strong> preparation <strong>of</strong> alkyl-alkyl’ α-chiral amines in mediocre to good yields and<br />

enantioselectivities. This group <strong>of</strong> amines is known historically to be difficult syn<strong>the</strong>tic task.<br />

We developed a new asymmetric reductive amination procedure using Yb(OAc) 3 (50-110<br />

mol %) that allows increased diastereoselectivity (6-15% units) <strong>for</strong> alkyl-alkyl’ α-chiral<br />

amines that previously only provided mediocre to good diastereoselectivity. Different Lewis<br />

acids were tested under different reaction conditions <strong>of</strong> temperature, pressure and solvents<br />

and <strong>the</strong> results <strong>of</strong> <strong>the</strong>se experiments are discussed in chapter five.<br />

1d<br />

O<br />

+<br />

H 2 N<br />

Ph<br />

(S)-α-MBA<br />

Yb(OAc)3 ,MeOH-THF<br />

Raney-Ni, H 2 (120 psi)<br />

(S,S)-2d HN Ph Pd-C<br />

(S)-3d NH 2<br />

H 2 (60 psi)<br />

86% de 85% ee<br />

The use <strong>of</strong> catalytic Lewis acids in reductive amination has never been reported in literatures.<br />

We demonstrated <strong>the</strong> beneficial use <strong>of</strong> 10-15 mol % <strong>of</strong> Yb(OAc) 3 or Ce(OAc) 3 or Y(OAc) 3 in<br />

i


suppressing alcohol <strong>for</strong>mation and promoting reductive amination in good yield but without<br />

enhanced stereoselectivity. Despite <strong>the</strong> fact that <strong>the</strong> use <strong>of</strong> Brønsted acids in reductive<br />

amination is well established no literature reports are available. We have per<strong>for</strong>med and<br />

extensive study on <strong>the</strong> use <strong>of</strong> commercially available Brønsted and mineral acids in reductive<br />

amination. The scope <strong>of</strong> <strong>the</strong> reaction and <strong>the</strong> substrate categories are summarized in chapter<br />

six.<br />

A mechanism <strong>for</strong> <strong>the</strong> reaction has been proposed and <strong>the</strong> basic mechanistic experiments have<br />

been per<strong>for</strong>med. An in situ cis- to trans-ketimine isomerization mechanism, promoted by<br />

Yb(OAc) 3 , has been proposed to account <strong>for</strong> <strong>the</strong> observed increase in diastereoselectivity.<br />

The experiments and <strong>the</strong> proposed mechanism are summarized in chapter seven<br />

ii


Acknowledgement<br />

All <strong>the</strong> work reported in this <strong>the</strong>sis have been carried out at <strong>the</strong> Department <strong>of</strong> Chemistry,<br />

School <strong>of</strong> Engineering and Science, Jacobs University, Bremen, Germany since joining here<br />

on August 2006 till August 2009. I would like to thank Jacobs University <strong>for</strong> <strong>the</strong> financial<br />

support and all <strong>the</strong> laboratory facilities during my stay here. In this regard I would like to<br />

thank Pr<strong>of</strong>. Dr. h. c. Bernhard Kramer <strong>for</strong> approving my PhD scholarship.<br />

I would like to convey my kind regards to my supervisor Pr<strong>of</strong>. Thomas C. Nugent and thank<br />

him <strong>for</strong> all his kind suggestions and deeply appreciate his skillful guidance throughout my<br />

research. It was due to his relentless ef<strong>for</strong>ts that I could master <strong>the</strong> various techniques and<br />

learn to solve <strong>the</strong> different scientific challenges that came by my way. Lastly, I would also<br />

acknowledge his patience and kind understanding.<br />

I would thank Pr<strong>of</strong>. Nikolai Kuhnert <strong>for</strong> his kind consent to become <strong>the</strong> internal examiner <strong>of</strong><br />

this <strong>the</strong>sis.<br />

I would also thank Dr. Pralhad Ganeshpure, Research Centre, Indian Petrochemicals<br />

Corporation Limited, 391 346 Vadodara, India (B-21, Kinnari Duplex Ellora Park, Vadodara,<br />

Gujarat 390023, India) <strong>for</strong> his kind consent to become <strong>the</strong> external examiner <strong>of</strong> this <strong>the</strong>sis.<br />

My sincere appreciation goes to all my lab mates, Dr. Rashmi R. Mohanty, Dr. Vijay N.<br />

Wakchaure, Dr. Abhijit Ghosh, Ahson J. Shaikh, Mohammad Naveed Umar, Mohammad<br />

Shoaib, A. Alvaradomendez, Abdul Sadiq, Dan Hu, Ahtaram Bibi, Satish Wakchaure, Andrei<br />

Dragan, Andrei Iosub and Daniela Negru <strong>for</strong> <strong>the</strong>ir constant help and encouragement in all<br />

respect. I would also thank Mrs. Müller <strong>for</strong> her continuous help.<br />

All my deepest veneration goes to my parents <strong>for</strong> everything that <strong>the</strong>y have given to me. I<br />

would convey my regards to my sister and all my uncles and aunts <strong>for</strong> <strong>the</strong>ir constant support.<br />

I would also thank all pr<strong>of</strong>essors and colleagues in Egypt. Especially I would like to thank<br />

iii


Pr<strong>of</strong>. Mohamed El-Azizi, Pr<strong>of</strong>. Abdel-Nasser Singab and Pr<strong>of</strong>. Nahla Ayoub <strong>for</strong> <strong>the</strong>ir support<br />

and help over <strong>the</strong> past years.<br />

I would like to thank all friends at Jacobs University, Iyad Tumar, Khaled Hassan, Dr. Raed<br />

Mesleh, Mohamed Noor, Hamdy El-Sheshtawy, Salahaldin Juba, Ahmed Moussa, Ahmed El-<br />

Moasry, Hany Elgala and all o<strong>the</strong>r friends in Germany and Egypt <strong>for</strong> <strong>the</strong>ir continuous support<br />

,<br />

,<br />

iv


Abbreviations<br />

Ac<br />

AcOH<br />

aq.<br />

Ar<br />

bs<br />

BINOL<br />

BINAP<br />

BOC<br />

iBu<br />

nBu<br />

conv.<br />

cat.<br />

CDCl 3<br />

COD<br />

d<br />

dd<br />

DCM<br />

de<br />

DIBAL-H<br />

DME<br />

DMF<br />

DMSO<br />

δ<br />

ee<br />

equiv.<br />

ESI<br />

Acetyl<br />

Acetic acid<br />

Aqueous<br />

Aryl<br />

Broad singlet ( 1 H-NMR)<br />

1,1'-Bi-2-naphthol<br />

2,2'-Bis(diphenylphosphino)-1,1'-binaphthyl.<br />

tert-Butyl carbamates<br />

iso-Butyl<br />

n-Butyl<br />

Conversion<br />

Catalyst<br />

Deuterated chloro<strong>for</strong>m<br />

Cycloctadiene<br />

Doublet ( 1 H-NMR)<br />

Doublet <strong>of</strong> doublet ( 1 H-NMR)<br />

Dichloromethane<br />

Diastereomeric excess<br />

Diisobutyl aluminium hydride<br />

1,2-Dimethoxyethane<br />

N,N’-Dimethylfomamide<br />

Dimethylsulfoxide<br />

Chemical shift ( 1 H-NMR)<br />

Enantiomeric excess<br />

Equivalent<br />

Electron spray ionization (Mass<br />

spectroscopy)<br />

v


Et<br />

EtOH<br />

EtOAc<br />

GC<br />

h<br />

HPLC<br />

HRMS<br />

Hz<br />

J<br />

KHMDS<br />

LDA<br />

m<br />

M<br />

MBA<br />

Me<br />

min.<br />

MS<br />

MS<br />

MTBE<br />

MW<br />

m/z<br />

m<br />

NaOtBu<br />

NBD<br />

NMR<br />

o<br />

p<br />

Pd-C<br />

Ph<br />

iPr<br />

nPr<br />

Pt-C<br />

pyr<br />

Ethyl<br />

Ethanol<br />

Ethylacetate<br />

Gas chromatography<br />

Hours<br />

High per<strong>for</strong>mance liquid chromatography<br />

High resolution mass spectrometry<br />

Hertz<br />

Coupling constant ( 1 H-NMR)<br />

Potassium hexamethyldisilazide<br />

Lithium diisopropylamide<br />

Multiplate ( 1 H-NMR)<br />

Molar<br />

Methyl Benzyl Amine<br />

Methyl<br />

Minutes<br />

Molecular sieves<br />

Mass spectroscopy<br />

Methyl-tert-butyl e<strong>the</strong>r<br />

Molecular weight<br />

Mass/charge<br />

Meta<br />

Sodium tert-butoxide<br />

N-Bornadiene<br />

Nuclear Magentic Resonance<br />

Ortho<br />

Para<br />

Palladium on carbon<br />

Phenyl<br />

iso-Propyl<br />

n-Propyl<br />

Platinum on carbon<br />

Pyridine<br />

vi


q<br />

Raney-Ni<br />

Ref.<br />

Rh-C<br />

s<br />

t<br />

t-Bu<br />

tert<br />

temp<br />

TFA<br />

THF<br />

TLC<br />

TMS<br />

Ts<br />

TsOH<br />

tBuLi<br />

Ti(O i Pr) 4<br />

Quartet ( 1 H-NMR)<br />

Raney-Nickel<br />

Reference<br />

Rhodium on carbon<br />

Singlet ( 1 H-NMR)<br />

Triplet ( 1 H-NMR)<br />

tert-Butyl<br />

Tertiary<br />

Temperature<br />

Trifluoroacetic acid<br />

Tetrahydr<strong>of</strong>uran<br />

Thin layer chromatography<br />

Trimethylsilane<br />

Tosyl<br />

p-Toluenesulfonic acid<br />

tert-Butyllithium<br />

Titanium(IV) isopropoxide<br />

vii


Table <strong>of</strong> Contents<br />

Abstract.<br />

Acknowledgment.<br />

List <strong>of</strong> Abbreviations.<br />

i<br />

ii<br />

v<br />

1. Introduction to <strong>Chiral</strong>ity<br />

1.1. <strong>Chiral</strong> Drugs 1<br />

1.2. Isomers and Isomerism 2<br />

1.3. Nature is <strong>Chiral</strong> 3<br />

1.4. <strong>Chiral</strong>ity and Drug-Receptor Interaction 8<br />

1.5. Sources <strong>of</strong> Enantiopure Substances 8<br />

1.5.1. Syn<strong>the</strong>sis <strong>of</strong> Enantiomerically Pure Compounds 9<br />

1.5.2. Resolution 9<br />

1.5.2.1. Preferential Crystallization 9<br />

1.5.2.1. Diastereomer Crystallization 10<br />

1.5.2.2. Kinetic Resolution 11<br />

1.5.3. <strong>Chiral</strong> Pool Approach 12<br />

1.5.4. Stereoselective Conversion <strong>of</strong> Prochiral Substrates to Enantiopure<br />

Compounds (Asymmetric Syn<strong>the</strong>sis) 15<br />

1.5.5. Asymmetric Syn<strong>the</strong>sis vs Kinetic Resolution vs <strong>Chiral</strong> Pool 18<br />

1.6. α-<strong>Chiral</strong> <strong>Amines</strong> Defining Terms 19<br />

1.7. α-<strong>Chiral</strong> <strong>Amines</strong> Importance 20<br />

1.8. α-<strong>Chiral</strong> Amine Syn<strong>the</strong>sis Different Methodologies 22<br />

1.8.1. Imine and Enamide Syn<strong>the</strong>sis 23<br />

1.8.2. Enantioselective Reduction <strong>of</strong> Enamides 23<br />

1.9. Conclusion 28<br />

1.10. References 28<br />

2. Imine Reduction<br />

2.1. Historical View 34<br />

2.2. Asymmetric Reduction <strong>of</strong> N-Phosphinoyl Imines 35<br />

2.2.1. Syn<strong>the</strong>sis <strong>of</strong> N-Phosphinoyl Imines 35<br />

viii


2.2.2. Different Substrates Categories 36<br />

2.2.3. Nguyen Special Substrates 39<br />

2.3. Asymmetric Reduction <strong>of</strong> N-aryl imines 40<br />

2.3.1. Syn<strong>the</strong>sis <strong>of</strong> N-Aryl Imines 40<br />

2.3.2. Different Substrates Categories 43<br />

2.4. Reduction <strong>of</strong> Miscellaneous Imines 49<br />

2.5 Conclusion 49<br />

2.6. References 50<br />

3. Reductive Amination<br />

3.1. Historical View 53<br />

3.1.1. Reductive Amination Utilizing Heterogeneous Catalyst 53<br />

3.1.2. Reductive Amination Utilizing Homogenous Catalysis 55<br />

3.2. Reductive Amination <strong>the</strong> Current State <strong>of</strong> Art 57<br />

3.3. Asymmetric Reductive Amination 60<br />

3.3.1. Asymmetric Reductive Amination Utilizing <strong>Chiral</strong> Catalysts 60<br />

3.3.2. Reductive Amination Utilizing <strong>Chiral</strong> Auxiliary 64<br />

3.3.3. Reductive Amination Utilizing Molecular Hydrogen 65<br />

3.3.4. Asymmetric Reductive Amination Utilizing Transfer<br />

Hydrogenation Conditions 66<br />

3.4.5. Organocatalytic Asymmetric Reductive Amination 67<br />

3.4. Green Chemistry and Reductive Amination 73<br />

3.4.1. Green Chemistry Basic principle 73<br />

3.4.2. Hydrogenation and Green Chemistry 75<br />

3.5. Conclusion 76<br />

3.6. References 76<br />

4. Drugs and Reductive Amination<br />

4.1 Reductive Amination in <strong>the</strong> Syn<strong>the</strong>sis <strong>of</strong> Drugs and Natural Products 81<br />

4.1.1. Syn<strong>the</strong>sis <strong>of</strong> Delavirdine 81<br />

4.1.2. Syn<strong>the</strong>sis <strong>of</strong> Muraglitazar 82<br />

4.1.3. Syn<strong>the</strong>sis <strong>of</strong> Amphetamine 83<br />

4.1.4. Syn<strong>the</strong>sis <strong>of</strong> Sertraline 84<br />

ix


4.1.5. Syn<strong>the</strong>sis <strong>of</strong> Emitine 85<br />

4.1.6. Syn<strong>the</strong>sis <strong>of</strong> Taltobulin 86<br />

4.1.7. Syn<strong>the</strong>sis <strong>of</strong> Perzinfote 87<br />

4.1.8. Syn<strong>the</strong>sis <strong>of</strong> Namindinil 88<br />

4.1.9. Syn<strong>the</strong>sis <strong>of</strong> Ezlopipant 89<br />

4.1.10.Syn<strong>the</strong>sis <strong>of</strong> Monomorine 90<br />

4.1.11. Syn<strong>the</strong>sis <strong>of</strong> Ontazolast 91<br />

4.1.12. Syn<strong>the</strong>sis <strong>of</strong> Pamaquine 92<br />

4.1.13. Syn<strong>the</strong>sis <strong>of</strong> Torcetrapib 92<br />

4.1.14. Syn<strong>the</strong>sis <strong>of</strong> Polyaminocholestanol Derivatives 93<br />

4.1.15. Syn<strong>the</strong>sis <strong>of</strong> piperazinylpropylisoxazoline Analogues 94<br />

4.1.16. Syn<strong>the</strong>sis <strong>of</strong> Ritonavir and Lopinavir 95<br />

4.1.17. Syn<strong>the</strong>sis <strong>of</strong> Tetrahydrocarbazoles 95<br />

4.2. Conclusion 97<br />

4.3. References 97<br />

5. Stoichiometric Use <strong>of</strong> Ytterbium Acetate in Reductive Amination.<br />

5.1. Introduction 98<br />

5.1.1. Ytterbium 100<br />

5.1.1.1. Electronic Overview 100<br />

5.1.1.2. Ytterbium Discovery 101<br />

5.1.1.3. Ytterbium Reactions 101<br />

5.1.1.4. Ytterbium Acetate 102<br />

5.1.2. Commercial Yb(OAc) 3 vs Dried Yb(OAc) 3 110<br />

5.1.3. Optimized Conditions and Useful Substrate Range 112<br />

5.2. Conclusion 115<br />

5.3. References 115<br />

6. Catalytic Lewis Acids in Reductive Amination<br />

6.1. Introduction 117<br />

6.2. Brønsted Acid Promoted Reductive Amination 122<br />

6.3. Conclusion 125<br />

6.4. References 125<br />

x


7. Stereochemical Considerations <strong>of</strong> Proposed Mechanistic Models<br />

7.1. Introduction 127<br />

7.1.1. Mechanism Behind Enhanced Stereoselectivity with Yb(OAc) 3 128<br />

7.1.2. Reasons behind Enhanced Diastereoselectivity <strong>for</strong> Different 133<br />

Substrate Categories<br />

7.1.3. Key Findings <strong>for</strong> Reductive Amination with α-MBA 135<br />

7.2. Conclusion 137<br />

7.3. References 138<br />

8. Appendix<br />

Experimental Section 140<br />

Curriculum Vitae 152<br />

xi


Chapter 1<br />

Introduction<br />

1.1. <strong>Chiral</strong> Drugs<br />

<strong>Chiral</strong> molecules <strong>for</strong>m a large proportion <strong>of</strong> <strong>the</strong>rapeutic agents. Drug chirality is considered a<br />

major <strong>the</strong>me in <strong>the</strong> design, discovery, development, launching and marketing <strong>of</strong> new drugs.<br />

Awareness <strong>of</strong> <strong>the</strong> importance <strong>of</strong> chirality comes from <strong>the</strong> fact that stereoselectivity is an<br />

essential dimension in pharmacology. The recent development <strong>of</strong> bioanalytical tools led to a<br />

better understanding <strong>of</strong> <strong>the</strong> importance <strong>of</strong> stereoselective phramacodyanmics and<br />

pharmacokinetics <strong>of</strong> chiral drugs. In 1984 it was estimated that <strong>the</strong> total proportion <strong>of</strong> drugs<br />

having chiral centre in <strong>the</strong> European market (Swedish survey) was 53%. [1] The percentage<br />

increased up to 57% within less than three years. [2]<br />

It was also estimated that 55% <strong>of</strong> <strong>the</strong> chiral drugs are used as a racemic mixture and <strong>the</strong> rest<br />

is marketed as a single enantiomer. By <strong>the</strong> end <strong>of</strong> <strong>the</strong> last century, <strong>the</strong> market <strong>for</strong> chiral drugs<br />

established major place in <strong>the</strong> overall global drug market. The situation totally changed in<br />

this century. Pharmaceutical companies stopped developing racemic drugs; <strong>the</strong>y only focus<br />

on <strong>the</strong> syn<strong>the</strong>sis <strong>of</strong> single enantiomeric drug entities.<br />

First we have to clarify <strong>the</strong> concept <strong>of</strong> chirality. <strong>Chiral</strong>ity or handedness comes originally<br />

from <strong>the</strong> Greek word cheir which means hand. One <strong>of</strong> <strong>the</strong> simplest definition <strong>of</strong> <strong>the</strong> word<br />

chiral is given by Mislow: An object is chiral if and only if it is not superposable on its mirror<br />

image; o<strong>the</strong>rwise it is chiral. [3] From <strong>the</strong> definition it is clear that <strong>the</strong> term chiral refers to <strong>the</strong><br />

spatial property <strong>of</strong> <strong>the</strong> objects including molecules. It defines that <strong>the</strong> molecule is nonsuperposable<br />

on its mirror image and does not refer to <strong>the</strong> stereochemical composition <strong>of</strong> <strong>the</strong><br />

bulk material. [4]<br />

1


It should be clear that <strong>the</strong> term chiral drug does not indicate that <strong>the</strong> drug is marketed as a<br />

single isomer it may be a racemic or unequal mixture <strong>of</strong> isomers. Through investigating <strong>the</strong><br />

origin <strong>of</strong> chirality it was revealed that <strong>the</strong> concept was introduced long ago. Archimedes<br />

designed Archimedean water screw and studied its chiral structure. Dominique Arge (1811)<br />

discovered <strong>the</strong> rotation <strong>of</strong> plan polarized light in quartz crystal. Later <strong>the</strong> French chemist Jean<br />

Baptiste Biot was <strong>the</strong> first to introduce <strong>the</strong> modern concept <strong>of</strong> chirality when he discovered<br />

rotation <strong>of</strong> light in a sugar solution. [5]<br />

The major breakthrough in understanding <strong>the</strong> concept <strong>of</strong> chirality and its significance in<br />

chemistry was achieved by Louis Pasteur through recrystallization <strong>of</strong> sodium ammonium<br />

tartrate (optically inactive). He noticed that <strong>the</strong> crystals were <strong>of</strong> two types which he<br />

physically separated. The two types <strong>of</strong> crystals were optically active, but rotated <strong>the</strong> plane <strong>of</strong><br />

polarized light in <strong>the</strong> opposite directions. He proposed that <strong>the</strong> molecules came in two <strong>for</strong>ms,<br />

“left handed” and “right handed”. Toge<strong>the</strong>r, <strong>the</strong> mixture <strong>of</strong> <strong>the</strong> two <strong>for</strong>ms is optically<br />

inactive. This finding prompted his famous statement that <strong>the</strong> universe is chiral (l’univers est<br />

dissymme´trique). [6]<br />

Later Van’t H<strong>of</strong>f, a Dutch young scientist proposed that <strong>the</strong> carbon atom is attached to four<br />

different substituents in space having a tetrahedral arrangement. This proposition was faced<br />

by strong opposition from scientists all over <strong>the</strong> world. Later <strong>the</strong>y discovered that his<br />

proposed shape <strong>of</strong> <strong>the</strong> molecule was absolutely right and he was awarded <strong>the</strong> first noble prize<br />

in chemistry <strong>for</strong> his work. [7] <strong>Chiral</strong>ity is manifested by centre <strong>of</strong> dissymmetry, but it can also<br />

be represented in axes or planes <strong>of</strong> dissymmetry. [8]<br />

1.2. Isomers and Isomerism<br />

Isomerism is <strong>the</strong> phenomenon <strong>of</strong> two or more compounds having <strong>the</strong> same number and kind<br />

<strong>of</strong> atoms. [9] Isomers can be subdivided into structural isomers, <strong>the</strong> difference between<br />

isomers is due to a different structural arrangements <strong>of</strong> <strong>the</strong> atoms that <strong>for</strong>m molecules, e.g.<br />

butane and isobutene. The o<strong>the</strong>r division is stereoisomers, <strong>the</strong> isomers have <strong>the</strong> same<br />

structural <strong>for</strong>mula, but differ in <strong>the</strong> spatial arrangement <strong>of</strong> atoms. [9]<br />

There are two types <strong>of</strong> stereoisomers:<br />

2


1. Cis-trans or geometric isomers.<br />

2. Optical isomers<br />

Optical isomers have <strong>the</strong> ability to rotate plane-polarized light. [8] Enantiomers are part <strong>of</strong> <strong>the</strong><br />

optical isomers, toge<strong>the</strong>r with diastereomers. Enantiomers are mirror image optical isomers<br />

having only one chiral centre. Enantiomers posses <strong>the</strong> same physical properties but <strong>the</strong>y<br />

differ in <strong>the</strong>ir biochemical properties. They behave differently only in a chiral medium, such<br />

as when exposed to a polarized light or when participating in a chemical reaction catalyzed<br />

by a chiral catalyst, particularly an enzyme in <strong>the</strong> body. (+)-Glucose (“blood sugar”) is used<br />

<strong>for</strong> metabolic energy whereas (-)-glucose is not. (+)-Lactic acid is produced by reactions<br />

occurring in muscle tissue, and (-)-lactic acid is produced by <strong>the</strong> lactic acid bacteria in <strong>the</strong><br />

souring <strong>of</strong> milk. Diastereomers are non mirror image optical isomers having more than chiral<br />

centre. Diastereomers have different physical properties allowing <strong>the</strong>ir separation.<br />

Enantioselectivity and diastereoselectivity are terms used to express <strong>the</strong> preferential<br />

<strong>for</strong>mation <strong>of</strong> one enantiomer or diastereomer over <strong>the</strong> o<strong>the</strong>r and it is normally expressed as an<br />

enantiomeric excess (ee) or diastereomeric excess (de).<br />

ee (%) =<br />

R(%) - S(%)<br />

R(%) + S(%)<br />

+<br />

100<br />

de (%) =<br />

D 1 (%) - D 2 (%)<br />

D 1 (%) + D 2 (%)<br />

+<br />

100<br />

1.3. Nature is <strong>Chiral</strong><br />

Many naturally occurring substances possess chirality, which is <strong>the</strong> property that a substance<br />

and its mirror image are not superimposable. [10] In every-day life, many examples can be<br />

found as well. Human hands are perhaps <strong>the</strong> most universally-recognized example <strong>of</strong><br />

chirality. The left hand is a non-superimposable mirror image <strong>of</strong> <strong>the</strong> right hand; no matter<br />

how <strong>the</strong> two hands are oriented, it is impossible <strong>for</strong> all <strong>the</strong> major features <strong>of</strong> both hands to<br />

coincide. [11]<br />

In particular life depends on molecular chirality, with many biological functions/processes<br />

inherently based on <strong>the</strong> interaction <strong>of</strong> dissymmetric molecules. Many physiological<br />

phenomena arise from highly preferential molecular interactions in which a chiral host<br />

3


molecule recognizes one <strong>of</strong> two enantiomeric guest molecules. There are numerous examples<br />

<strong>of</strong> enantiomeric effects which are frequently dramatic. Thus, <strong>the</strong> enantiomers <strong>of</strong> limonene,<br />

both are found in nature, smell differently, because our nasal receptors are made <strong>of</strong> chiral<br />

molecules that interact with <strong>the</strong>se enantiomers differently. Similarly one enantiomer <strong>of</strong> <strong>the</strong><br />

amino acid asparagine tastes sweet while <strong>the</strong> o<strong>the</strong>r tastes bitter. Clearly living systems are<br />

very sensitive to chirality and many pharmaceutical drugs consist <strong>of</strong> chiral moieties. <strong>Chiral</strong><br />

drugs are a subgroup <strong>of</strong> drug substances that contain one or more chiral centres. It is well<br />

established that <strong>the</strong> opposite enantiomer <strong>of</strong> a chiral drug <strong>of</strong>ten differs significantly in its<br />

pharmacological, [12] toxicological, [13] pharmacodynamic and pharmacokinetic properties. [14]<br />

A renowned example <strong>of</strong> how chirality affects <strong>the</strong> pharmacological action <strong>of</strong> <strong>the</strong> drugs, a<br />

chiral drug is thalidomide (Thalidomid, Contergan) which was prescribed to pregnant women<br />

in <strong>the</strong> 1960s to alleviate morning sickness. One <strong>of</strong> <strong>the</strong> enantiomeric <strong>for</strong>ms <strong>of</strong> thalidomide<br />

does indeed have sedative and antinausea effects, but <strong>the</strong> o<strong>the</strong>r enantiomer is a potent<br />

teratogen. The racemic drug was approved in Europe <strong>for</strong> <strong>the</strong> treatment <strong>of</strong> pregnant women<br />

suffering from nausea and its use caused severe birth defects. Even <strong>for</strong>mulation <strong>of</strong> <strong>the</strong> pure<br />

nontoxic (R)-enantiomer <strong>of</strong> thalidomide would have been unsafe because racemization takes<br />

place in vivo and <strong>the</strong> teratogenic (S)-enantiomer is rapidly generated in <strong>the</strong> human body.<br />

Since <strong>the</strong> thalidomide tragedy, <strong>the</strong> significance <strong>of</strong> <strong>the</strong> stereochemical integrity <strong>of</strong> biologically<br />

active compounds has received increasing attention and <strong>the</strong> investigation <strong>of</strong> <strong>the</strong><br />

stereodynamic properties <strong>of</strong> chiral molecules has become an integral part <strong>of</strong> modern drug<br />

development. [15]<br />

4


O<br />

O<br />

O<br />

*<br />

NH<br />

Thalidomide<br />

(R)-active agent<br />

(S)-teratogenic<br />

O<br />

*<br />

H<br />

N<br />

OH<br />

Ethambutol<br />

(R,R)-blinding agent<br />

(S,S)-tuberculostatic<br />

2<br />

Limonene<br />

(S)-lemon odor<br />

Limonen<br />

(R)-organge odor<br />

O<br />

Carvone<br />

(S)-caraway<br />

O<br />

Carvone<br />

(R)-spearmint<br />

H 2 N<br />

OH 2 N<br />

H<br />

O<br />

Asparagine<br />

(S)-bitter<br />

OH<br />

HO<br />

O<br />

H<br />

NH 2 O<br />

Asparagine<br />

(R)-sweet<br />

NH 2<br />

HO<br />

H<br />

O<br />

O<br />

H<br />

N<br />

NH<br />

H 2<br />

O<br />

O<br />

O<br />

O<br />

H<br />

O<br />

O<br />

N<br />

H<br />

H 2 N<br />

H<br />

OH<br />

Aspartame<br />

(S,S)-sweet<br />

Aspartame<br />

(R,R)-bitter<br />

Figure 1.1 Absolute Configuration vs Biological Activity.<br />

Ano<strong>the</strong>r example showing <strong>the</strong> importance <strong>of</strong> distinguishing <strong>the</strong> two enantiomers is <strong>the</strong><br />

distinguished effect <strong>of</strong> different isomeric <strong>for</strong>ms <strong>of</strong> <strong>the</strong> nonsteroidal anti inflammatory drugs<br />

(NSAIDs). They include ibupr<strong>of</strong>en (Advil), naproxen (Aleve), ketopr<strong>of</strong>en (Oruvail), and<br />

flurbipr<strong>of</strong>en (Ansaid), which have found widespread use as pain relievers. The antiinflammatory<br />

activity <strong>of</strong> <strong>the</strong>se pr<strong>of</strong>ens resides primarily with <strong>the</strong> (S)-enantiomer. The<br />

enantiomers <strong>of</strong> flurbipr<strong>of</strong>en possess different pharmacokinetic properties and show<br />

substantial racemization under physiological conditions. Although (S)-naproxen is <strong>the</strong> only<br />

pr<strong>of</strong>en that was originally marketed in enantiopure <strong>for</strong>m, in vivo interconversion <strong>of</strong> <strong>the</strong><br />

enantiomers <strong>of</strong> NSAIDs is an important issue in preclinical pharmacological and<br />

toxicological studies. [16]<br />

Also beta-blokers, <strong>the</strong> most widely used pharmaceutical agents <strong>for</strong> angina, hypertension, and<br />

arrhythmias. It is known that <strong>for</strong> most beta-blockers <strong>the</strong> (S)-enantiomer is <strong>the</strong> most active<br />

enantiomer. The S-enantiomer has <strong>the</strong> same three dimensional structure as <strong>the</strong> adrenergic<br />

5


hormone noradrenaline. The (R)-enantiomer <strong>of</strong> <strong>the</strong> betablocker does not give serious sideeffects,<br />

but it does not add to <strong>the</strong> pharmacological effect ei<strong>the</strong>r, so it can be considered as<br />

‘isomeric ballast’. The most sold beta-blockers (propranolol, atenolol, metoprolol) were<br />

developed in <strong>the</strong> 1970s and are still marketed as racemate. If <strong>the</strong>se substances would have<br />

been developed today, it can be expected that <strong>the</strong>y would have been introduced as a single<br />

enantiomer. [17]<br />

There<strong>for</strong>e from <strong>the</strong> points <strong>of</strong> view <strong>of</strong> safety and efficacy, <strong>the</strong> pure enantiomer is preferred<br />

over <strong>the</strong> racemate in many marketed dosage <strong>for</strong>ms. In past decades <strong>the</strong> pharmacopoeia was<br />

dominated by racemates, but since <strong>the</strong> emergence <strong>of</strong> new technologies in <strong>the</strong> 1980s that<br />

allowed <strong>the</strong> preparation <strong>of</strong> pure enantiomers in significant quantities, <strong>the</strong> awareness and<br />

interest in <strong>the</strong> stereochemistry <strong>of</strong> drug action has increased. Although some ‘‘blockbuster’’<br />

drugs, such as fluoxetine hydrochloride (Prozac) is still marketed as racemates. However, <strong>the</strong><br />

recent trend is toward marketing a single-enantiomer drugs. [18]<br />

Previously <strong>the</strong> chiral drug is <strong>of</strong>ten syn<strong>the</strong>sized in <strong>the</strong> racemic <strong>for</strong>m, and it is frequently costly<br />

to resolve <strong>the</strong> racemic mixture into <strong>the</strong> pure enantiomers. Ano<strong>the</strong>r approach by<br />

pharmaceutical companies is what is called racemic switch. This fashionable approach<br />

involves <strong>the</strong> development <strong>of</strong> a pure enantiomer <strong>of</strong> <strong>the</strong> drug that is already marketed as a<br />

racemate. This means if a patent on a drug that is marketed as a racemic mixture is expiring;<br />

it is sometimes possible to obtain a new patent <strong>for</strong> <strong>the</strong> active enantiomer. In this way, <strong>the</strong><br />

pharmaceutical company retains <strong>the</strong> exclusive rights on <strong>the</strong> substance <strong>for</strong> ano<strong>the</strong>r period, but<br />

<strong>the</strong>y will have to change <strong>the</strong>ir manufacturing method as well. [19]<br />

Although only a minority <strong>of</strong> all racemic drugs has proved to be suitable <strong>for</strong> a racemic switch,<br />

this development has boosted <strong>the</strong> development <strong>of</strong> new manufacturing and separation<br />

methods. An example <strong>of</strong> a successful racemic switch is <strong>the</strong> local anaes<strong>the</strong>tic bupivacaine<br />

(AstraZeneca's Marcain). The (S)-isomer is now marketed under <strong>the</strong> trade name Chirocaine.<br />

This isomer was found to be substantially less cardiotoxic than <strong>the</strong> (R)-isomer, and <strong>the</strong>re<strong>for</strong>e<br />

a new patent was granted. Fur<strong>the</strong>rmore, <strong>the</strong> (S)-isomer <strong>of</strong> omeprazole (a proton pump<br />

inhibitor by AstraZeneca, known as Losec/Prilosec) is now marketed as a single enantiomer<br />

under <strong>the</strong> trade name Nexium. [20]<br />

6


One enantiomer may be responsible <strong>for</strong> <strong>the</strong> activity; its paired enantiomer could be inactive,<br />

possess some activity <strong>of</strong> interest, be an antagonist <strong>of</strong> <strong>the</strong> active enantiomer or have a separate<br />

activity that could be desirable or undesirable. To market drug as racemate or as <strong>the</strong><br />

enantiomeric pure <strong>for</strong>m is mainly based on pharmacology, toxicology and economics. From a<br />

pharmaceutical perspective, <strong>the</strong> physical properties <strong>of</strong> both <strong>the</strong> racemate and <strong>the</strong> enantiomer<br />

should be characterized in detail in order to develop a safe, efficacious, and reliable<br />

<strong>for</strong>mulation, no matter whe<strong>the</strong>r <strong>the</strong> racemate or <strong>the</strong> enantiomer pure <strong>for</strong>m is chosen as <strong>the</strong><br />

marketed <strong>for</strong>m. <strong>Chiral</strong>ity <strong>of</strong> a drug can also influence <strong>the</strong> efficiency <strong>of</strong> delivery, which has<br />

not been well investigated in <strong>the</strong> pharmaceutical field. [21]<br />

Density, solubility, dissolution behaviour, stability, and mechanical properties which are <strong>the</strong><br />

many physical properties <strong>of</strong> a crystalline solid, are governed by <strong>the</strong> crystal structure. [22]<br />

Understanding <strong>the</strong> relationship between <strong>the</strong> crystal structure and <strong>the</strong> physical properties, and<br />

<strong>the</strong>ir influence on drug release, may <strong>the</strong>re<strong>for</strong>e provide a clear picture <strong>of</strong> chirality–delivery<br />

relationship.<br />

This discussion supports <strong>the</strong> fact that using single enantiomeric pure <strong>for</strong>m <strong>of</strong> a drug has major<br />

advantages as reducing <strong>the</strong> overall administered dose, improving drug <strong>the</strong>rapeutic window,<br />

reducing any intersubject variability and finally estimating <strong>the</strong> dose response relationship<br />

accurately. [23]<br />

All previously reported reasons have led to an increasing preference <strong>for</strong> production <strong>of</strong> <strong>the</strong><br />

single enantiomers in both industry and regulatory authorities. Regulation regarding control<br />

<strong>of</strong> chiral drugs began in <strong>the</strong> US with a publication in 1992 about <strong>the</strong> <strong>for</strong>mal guidelines on <strong>the</strong><br />

development <strong>of</strong> chiral drugs in a document entitled Policy Statement <strong>for</strong> <strong>the</strong> Development <strong>of</strong><br />

New Stereoisomeric Drugs by FDA and European Union. The major outlines <strong>for</strong> <strong>the</strong><br />

guidelines state that <strong>the</strong> drug applicants must recognize <strong>the</strong> occurrence <strong>of</strong> chirality in <strong>the</strong> new<br />

drugs, attempt to separate <strong>the</strong> stereoisomers, assess <strong>the</strong> contribution <strong>of</strong> <strong>the</strong> various<br />

stereoisomers to <strong>the</strong> activity <strong>of</strong> interest and make a rational selection <strong>of</strong> <strong>the</strong> stereoisomeric<br />

<strong>for</strong>m that is proposed <strong>for</strong> marketing.<br />

Global sales <strong>of</strong> chiral drugs in single-enantiomeric <strong>for</strong>m continue to grow. The annual sales<br />

<strong>of</strong> chiral drugs as a single enantiomeric <strong>for</strong>m increased dramatically, from 27% (US $74.4<br />

7


illion) in 1996, 29% (1997), 30% (1998), 32% (1999), 34% (2000), 38% (2001) to an<br />

estimate <strong>of</strong> 39% (US $151.9 billion) in 2002. [24]<br />

1.4. <strong>Chiral</strong>ity and Drug-Receptor Interaction<br />

As mentioned be<strong>for</strong>e biological systems are based on chirality. For example, enzymes are<br />

considered as chiral biological polymers consisting <strong>of</strong> solely L-amino acids. They are highly<br />

structured compounds: <strong>the</strong>ir secondary and tertiary structure is determined by <strong>the</strong> amino acid<br />

constituents. Enzymes function as molecular receptors by binding selectively to specific<br />

molecules. Due to <strong>the</strong>ir chirality, <strong>the</strong>y commonly interact much stronger with one enantiomer<br />

<strong>of</strong> <strong>the</strong> ‘target’ molecule; or what is known as chiral recognition. ‘lock-and-key’ concept was<br />

introduced in 1894 by Fischer to explain enzyme selectivity. This simple concept sates that<br />

one enantiomer ‘fits’ in <strong>the</strong> enzyme cavity, <strong>the</strong> o<strong>the</strong>r enantiomer does not. This concept was<br />

re<strong>for</strong>mulated later to more complicated model (three point model). [25]<br />

To get a high degree <strong>of</strong> enantioselection, a substrate must be held firmly in three dimensional<br />

space. There must be at least three different points <strong>of</strong> attachment <strong>of</strong> <strong>the</strong> substrate onto <strong>the</strong><br />

active site. Variations and refinements to this rule have been reported. The most important<br />

one is that <strong>the</strong> interactions may be attractive or repulsive. Steric hindrance <strong>of</strong>ten plays an<br />

important role in chiral recognition.<strong>Chiral</strong>ity has also an important role in <strong>the</strong> field <strong>of</strong> fine<br />

chemical industry. Large applications are found in agrochemicals, food and fragrance<br />

industry. There are handful examples <strong>of</strong> enantiomers showing different fragrances because<br />

one is (R) and one is (S).<br />

For a chiral herbicide from <strong>the</strong> class <strong>of</strong> α-aryloxypropionic acids, <strong>the</strong> activity is present<br />

almost solely in <strong>the</strong> (R) enantiomer. Although currently most herbicides are applied as<br />

racemates, attempts towards <strong>the</strong> development <strong>of</strong> convenient large scale methodologies <strong>for</strong> <strong>the</strong><br />

syn<strong>the</strong>sis <strong>of</strong> herbicides were awarded by development <strong>of</strong> metalochlor. This trend will result in<br />

50% reduction <strong>of</strong> dosage, which means 50% less environmental pollution.<br />

1.5. Sources <strong>of</strong> Enantiopure Substances<br />

8


1.5.1. Syn<strong>the</strong>sis <strong>of</strong> Enantiomerically Pure Compounds<br />

The importance <strong>of</strong> chiral compounds and <strong>the</strong> strong need <strong>for</strong> enantiomerically pure<br />

substances has led to develop versatile methodologies to meet this objective. There are three<br />

main approaches <strong>for</strong> <strong>the</strong> preparation <strong>of</strong> chiral compounds as shown in figure 1.2:<br />

1.Resolution <strong>of</strong> racemates;<br />

2.<strong>Chiral</strong> pool approach;<br />

3. Stereoselective conversion <strong>of</strong> prochiral substrates to enantiopure compounds (asymmetric<br />

syn<strong>the</strong>sis via catalytic or stoichiometric process).<br />

Racemates<br />

<strong>Chiral</strong> Pool<br />

Prochiral<br />

Substrates<br />

Preferential<br />

Crystalization<br />

Kinetic<br />

Resolution<br />

Diastereomer<br />

Crystallization<br />

Syn<strong>the</strong>sis<br />

Asymmetric Syn<strong>the</strong>sis<br />

Chemical<br />

Enzymatic<br />

Figure 1.2. Sources <strong>of</strong> Enantiopure Substances.<br />

1.5.2. Resolution<br />

Resolution technique is <strong>the</strong> most classical route to enantiopurity. Although it has many<br />

drawbacks and recently it has been overtaken by asymmetric syn<strong>the</strong>sis, this method is still<br />

persisted with numerous examples on <strong>the</strong> industrial scale till present days. Resolution can be<br />

subdivided into three main techniques.<br />

1.5.2.1. Preferential Crystallization<br />

Preferential crystallization is possible <strong>for</strong> racemates which <strong>for</strong>m conglomerates. The<br />

conglomerates are mechanical mixtures <strong>of</strong> enantiomerically pure crystals <strong>of</strong> one enantiomer<br />

9


and its opposite enantiomer. Molecules in <strong>the</strong> crystal structure have a greater affinity <strong>for</strong> <strong>the</strong><br />

same enantiomer than <strong>for</strong> <strong>the</strong> opposite enantiomer. The melting point <strong>of</strong> <strong>the</strong> racemic<br />

conglomerate is always lower than that <strong>of</strong> <strong>the</strong> pure enantiomer. Addition <strong>of</strong> a small amount <strong>of</strong><br />

one enantiomer to <strong>the</strong> conglomerate increases <strong>the</strong> melting point. Success in this method<br />

depends on <strong>the</strong> fact that <strong>for</strong> a conglomerate <strong>the</strong> racemic mixture is more soluble than ei<strong>the</strong>r <strong>of</strong><br />

<strong>the</strong> enantiomers. Generally only 5-10% <strong>of</strong> racemates <strong>for</strong>m conglomerates. [26]<br />

1.5.2.1. Diastereomer Crystallization<br />

L. Pasteur was <strong>the</strong> one who first to fully develop this methodology in 1854. [26] In this<br />

approach, a racemate interacts with an enantiopure compound to <strong>for</strong>m diastereomeric salt<br />

which can <strong>the</strong>n be separated by crystallization due to unequal solubility in a given solvent.<br />

These enantiopure compounds are called resolving agents and are obtained from <strong>the</strong> chiral<br />

pool, e.g. L-tartaric acid, D-camphor sulfonic acid or some alkaloid bases. In general this<br />

approach is extremely limited to few examples. One example <strong>of</strong> such process is <strong>the</strong><br />

crystallization <strong>of</strong> <strong>the</strong> salt <strong>of</strong> one enantiomer <strong>of</strong> 1,2 diamino cyclohexane obtained from <strong>the</strong><br />

interaction <strong>of</strong> <strong>the</strong> racemic mixture with enantiopure tartaric acid. [27]<br />

O SO 3 H O<br />

HO 3 S<br />

H<br />

X<br />

OH<br />

N<br />

Quinine (X=OMe)<br />

Chinchonidine (X=H)<br />

HO<br />

OH<br />

H 2 O/AcOH<br />

H 3 N<br />

K 2 CO 3<br />

H 2 N<br />

H 2 N<br />

NH 2<br />

HOOC<br />

COOH<br />

90 o C-5 o C<br />

HO<br />

COO<br />

COO<br />

NH 3<br />

H 2 O/EtOH<br />

H 2 N<br />

>98% ee<br />

HO<br />

Scheme 1.1. <strong>Preparation</strong> <strong>of</strong> 1,2 Diamino Cyclohexane<br />

10


1.5.2.2. Kinetic Resolution<br />

Kinetic resolution is based on <strong>the</strong> difference in reactivity rate <strong>of</strong> <strong>the</strong> two enantiomers with a<br />

chiral entity which is used in catalytic amount. [28] The chiral entity can be ei<strong>the</strong>r a biological<br />

catalyst (e.g. enzyme) or a chemical catalyst (e.g. chiral metal complex or organocatalyst).<br />

Rule <strong>of</strong> thumb <strong>for</strong> kinetic resolution to be successful is that one enantiomer must react faster<br />

than <strong>the</strong> o<strong>the</strong>r. In such situation <strong>the</strong>oretically, 50% <strong>of</strong> <strong>the</strong> product from one enantiomer and<br />

50% <strong>of</strong> <strong>the</strong> unreacted enantiomer should be obtained. Scheme 1.2 showing one example<br />

describing this condition in which racemic β-aryl-β-hydroxy esters with different substitution<br />

patterns on <strong>the</strong> aryl moiety provides preferably <strong>the</strong> (R)-enantiomer with 93-98% ee and 32-<br />

41% isolated yield. [29]<br />

Ph<br />

Ph<br />

N<br />

H<br />

Ph<br />

N<br />

HO<br />

BrZnCH 2 CO 2 tBu(8equiv.)<br />

CO 2 tBu<br />

Prolinol ligand (5.0 mol%)<br />

CO 2 tBu<br />

CO 2 tBu<br />

Ar<br />

OH<br />

THF, reflux<br />

Ar<br />

OH<br />

Ar<br />

Scheme 1.2. Kinetic Resolution in Asymmetric Syn<strong>the</strong>sis. .<br />

If <strong>the</strong> unwanted enantiomer is racemized in situ during resolution, a 100% <strong>the</strong>oretical yield <strong>of</strong><br />

<strong>the</strong> enantiopure product can be <strong>the</strong>oretically reached, this is known as a kinetic dynamic<br />

resolution. This approach was successfully applied utilizing enzymes as resolving agents. [30]<br />

11


N<br />

NH 2<br />

CaLB, EtOAc<br />

Toluene<br />

~48h N<br />

NH 2<br />

N<br />

NHAc<br />

>60% yield<br />

N<br />

HN<br />

N<br />

N<br />

O<br />

racemization<br />

Scheme 1.3. Dynamic Kinetic Resolution in Asymmetric Syn<strong>the</strong>sis.<br />

For chemical syn<strong>the</strong>sis, one <strong>of</strong> <strong>the</strong> earliest demonstrations <strong>of</strong> this method is an adaptation <strong>of</strong><br />

<strong>the</strong> Noyori asymmetric hydrogenation. [31]<br />

O<br />

O<br />

OH<br />

R 1 OR 3<br />

R 2<br />

H 2<br />

R 1 OR 3<br />

O<br />

O<br />

O<br />

R 1 OR 3<br />

R 2<br />

(R)-BINAP-Ru<br />

R 2<br />

H<br />

OH O<br />

2<br />

R 1 OR 3<br />

(R)-BINAP-Ru<br />

R 2<br />

a: R 1 =R 2 =CH 3 ;R 3 =C 2 H 5<br />

b: R 1 =R 3 =CH 3 ; R2=NHCOCH 3<br />

c: R 1 =3,4-methylenedioxyphenyl;<br />

R 2 =NHCOCH 3 ;R 3 =CH 3<br />

d: R 1 =3,4-methylenedioxyphenyl;<br />

R 2 =NHCOCH 2 C 6 H 5 ;R 3 =CH 3<br />

e: R 1 =R 3 =CH 3 ;R 2 =CH 2 NHCOC 6 H 5<br />

Scheme 1.4. Chemoselective Dynamic Kinetic Resolution<br />

1.5.3. <strong>Chiral</strong> Pool Approach<br />

Natural sources are <strong>of</strong>ten referred as <strong>the</strong> ‘chiral pool’. The most important classes <strong>of</strong> chiral<br />

pool substances are amino acids, carbohydrates, hydroxy acids, terpenes and alkaloids. [32]<br />

These substances are incorporated into products by chemical processes which involve<br />

retention <strong>of</strong> configuration, inversion or chirality transfer. The chiral starting material is called<br />

chiral synthon which introduces chirality in <strong>the</strong> final compound. This strategy is unlike chiral<br />

auxiliary approach (will be discussed later) in which <strong>the</strong> chirality is installed into <strong>the</strong> achiral<br />

12


compound by <strong>the</strong> auxiliary. The auxiliary is later deattached from <strong>the</strong> final product. Despite<br />

<strong>the</strong> breadth <strong>of</strong> functionality available from nature, limited examples are available in optically<br />

pure <strong>for</strong>m on a large scale. This means that incorporation <strong>of</strong> a “chiral pool” material into a<br />

syn<strong>the</strong>sis can result in a multistep sequence. However, with <strong>the</strong> recent advances in syn<strong>the</strong>tic<br />

methods which added new compounds to <strong>the</strong> chiral pool <strong>the</strong>y are still limited.<br />

Typically chiral pool material should be available on large scale in a reasonable price. One<br />

example is L-aspartic acid, where <strong>the</strong> chiral material can be cheaper than <strong>the</strong> racemate. An<br />

example <strong>of</strong> <strong>the</strong> application <strong>of</strong> chiral pool <strong>for</strong> syn<strong>the</strong>sis <strong>of</strong> pharmaceutical drugs is <strong>the</strong><br />

syn<strong>the</strong>sis <strong>of</strong> (S)-Vigabatrin, a potent GABA-T inhibitor from (R)-methionine by Knaus and<br />

Wei in 96% yield and >98% ee as shown in scheme 1.5. [33]<br />

S<br />

COO<br />

NH 3<br />

1 S COOMe<br />

NHCOOR 1<br />

2 S<br />

NHCOOR 1<br />

COOR 2<br />

a) R 1 =PhCH 2<br />

b) R 1 =Me<br />

3<br />

a) R 1 =PhCH 2 ,R 2 =Et<br />

b) R 1 =PhCH 2 ,R 2 =Me<br />

c) R 1 =Me, R 2 =Et<br />

d) R 1 =Me, R 2 =Me<br />

NH 3<br />

COO<br />

5<br />

O<br />

N<br />

H<br />

4<br />

O<br />

N<br />

H<br />

SMe<br />

1. i)MeOH/SOCl 2 . ii) NaHCO 3 /ClCO 2 R 1 . yield 82%-86%; 2. (R 2 O)P(O)CH 2 CO 2 R 2 /tBuLi/DiBAL-H,62-78% yield;<br />

3. Mg/MeOH. 92-95% yield; 4. i) NaIO 4 . ii)190 °C. 56%yield; 5) KOH/iPrOH/H 2 O. 96% yield. 98% ee.<br />

Scheme 1.5. Syn<strong>the</strong>sis <strong>of</strong> (S)-vigabatrin from (R)-methionine, chiral pool approach.<br />

13


<strong>Chiral</strong> Pool Compounds<br />

Amino acids<br />

Hydroxy acids<br />

NH 2<br />

Valine<br />

COOH<br />

phenylalanine<br />

NH 2<br />

COOH<br />

OH<br />

COOH<br />

lactic acid<br />

OH<br />

Ph COOH<br />

mandelic acid<br />

OH<br />

COOH<br />

HOOC<br />

OH<br />

tartaric acid<br />

Sugars<br />

Terpenes<br />

OH<br />

OH<br />

OH<br />

O<br />

O<br />

HO<br />

HO OH OH HO<br />

HO<br />

glucose<br />

mannose<br />

OH<br />

O<br />

camphor<br />

geraniol<br />

OH<br />

Figure 1.3. Examples <strong>of</strong> <strong>Chiral</strong> Pool Compounds.<br />

Ano<strong>the</strong>r example utilizing this approach is <strong>the</strong> syn<strong>the</strong>sis <strong>of</strong> herbicide (R)-flamprop-isopropyl<br />

starting from L-lactic acid. [34]<br />

OH<br />

COO i Pr<br />

(S)-lactic acid<br />

MeSO 2 Cl<br />

base<br />

OSO 2 Me<br />

COO i Pr<br />

Cl<br />

inversion<br />

F<br />

NH 2<br />

F<br />

Cl<br />

O<br />

N<br />

i PrOOC<br />

PhCOCl<br />

F<br />

Cl<br />

N<br />

H<br />

COO i Pr<br />

(R)-(-)-flamprop-isopropyl<br />

Scheme 1.5.Syn<strong>the</strong>sis <strong>of</strong> Enantiopure Herbicide from L-lactic acid.<br />

14


1.5.4. Stereoselective Conversion <strong>of</strong> Prochiral Substrates to Enantiopure<br />

Compounds (Asymmetric Syn<strong>the</strong>sis)<br />

In asymmetric syn<strong>the</strong>sis a stereogenic centre is created under <strong>the</strong> influence <strong>of</strong> some external<br />

or internal chiral inducing agents. This strategy can be subdivided into three approaches:<br />

substrate-controlled approach; chiral auxiliary approach; and catalyst controlled approach. In<br />

substrate controlled approach, chirality is present internally within <strong>the</strong> molecule directing<br />

remaining groups or faces in stereoselective manner. Limitations <strong>of</strong> this approach come from<br />

<strong>the</strong> fact that enantiopure starting materials are not easily available and <strong>the</strong> reacting sites<br />

should be within close proximity to <strong>the</strong> chiral centre.<br />

Regarding <strong>the</strong> o<strong>the</strong>r two approaches, achiral molecule is converted into chiral entity utilizing<br />

ei<strong>the</strong>r a stoichiometric quantity <strong>of</strong> <strong>the</strong> chiral auxiliary or a catalytic quantity <strong>of</strong> chiral<br />

catalysts. In <strong>the</strong> chiral auxiliary approach, chirality is induced in achiral molecule utilizing<br />

external chiral entity through <strong>for</strong>ming covalent bond with <strong>the</strong> achiral starting material. This<br />

auxiliary is <strong>the</strong>n cleaved from <strong>the</strong> final product in an additional step. Special precautions<br />

should be taken to avoid any racemisation <strong>of</strong> <strong>the</strong> final product in <strong>the</strong> deportation step. One<br />

example <strong>of</strong> this auxiliary approach is shown in scheme 1.6 in which (1S,2S)-(+)-<br />

pseudoephedrine is used as <strong>the</strong> chiral auxiliary to produce diastereomeric alkylated<br />

pseudoephedrine amides which can <strong>for</strong>m enantioenriched carboxylic acids(by hydrolysis),<br />

alcohols and aldehydes (by reduction). [35]<br />

15


OH<br />

N<br />

O<br />

R<br />

1. 2LDA, LiCl<br />

2. R 1 X<br />

THF<br />

OH<br />

N<br />

O<br />

R 1<br />

R<br />

80-99% yield<br />

94-99% de<br />

OH<br />

N<br />

H 2 SO 4 dioxane<br />

O<br />

R<br />

N BH 3 Li<br />

R 1<br />

THF<br />

HO<br />

O<br />

R 1<br />

R<br />

HO<br />

O<br />

R<br />

H<br />

R 1<br />

87-97% yield<br />

95-97% ee<br />

R 1<br />

R<br />

75-92% yield<br />

90-98% ee<br />

80-88% yield<br />

88-99% ee<br />

Scheme 1.6. <strong>Chiral</strong> Auxiliary Approach in Asymmetric Syn<strong>the</strong>sis<br />

The third approach which is <strong>the</strong> catalytic asymmetric trans<strong>for</strong>mation, is promoted by a chiral<br />

entity which is generally used in a catalytic amount enhancing <strong>the</strong> economic value <strong>of</strong> <strong>the</strong><br />

process. The chiral entity can be chiral catalysts (e.g. chiral Lewis acid or base, chiral<br />

organocatalysts, chiral organometallic complexes) or even bio catalysts. One <strong>of</strong> <strong>the</strong> most<br />

fascinated examples was <strong>the</strong> syn<strong>the</strong>sis <strong>of</strong> L-DOPA developed by Knowles. [36]<br />

AcO<br />

H<br />

OMe<br />

COOH [Rh(DiPAMP)]<br />

NHAc H 2<br />

AcO<br />

H<br />

OMe<br />

H<br />

COOH<br />

NHAc<br />

H 3 O +<br />

AcO<br />

H<br />

OMe<br />

H<br />

L-DOPA<br />

97.5% ee<br />

COOH<br />

NH 2<br />

CH 3 O<br />

P<br />

P<br />

OCH 3<br />

Scheme 1.7. Syn<strong>the</strong>sis <strong>of</strong> L-DOPA<br />

Ano<strong>the</strong>r example showing <strong>the</strong> importance <strong>of</strong> this approach, was developed by Royoji<br />

Noyori, [37] In 1980 he developed different derivatives <strong>of</strong> chiral BINAP ligands which were<br />

widely used as chiral ligands <strong>for</strong> Ru and Rh hydrogenation reactions. He was successful in<br />

applying his catalytic system on industrial scale <strong>for</strong> <strong>the</strong> (-)-menthol syn<strong>the</strong>sis from myrcene.<br />

16


It is estimated that 3000 tonnes (after new expansion) <strong>of</strong> menthol are produced (in 94% ee)<br />

by Takasago International Co., using Noyori's method every year. The key step was <strong>the</strong><br />

asymmetric isomerization <strong>of</strong> geranyldiethylamine, promoted by an (S)-BINAP-Rh complex<br />

in THF and <strong>for</strong>ming (R)-citronellal enamine, which upon hydrolysis gives (R)-citronellal in<br />

96-99% ee. This enantiopurity is higher than naturally available product (ee 80%) obtained<br />

from rose oil (scheme 1.8).<br />

Li, (C 2 H 5 ) 2 NH<br />

H R<br />

Hs<br />

myrcene<br />

diethylgeranylamine<br />

N(C 2 H 5 ) 2<br />

[Rh(S)-BINAP] +<br />

CHO<br />

H 3 O +<br />

Hs<br />

H R<br />

(R)-citronellal<br />

ZnBr 2<br />

N(C 2 H 5 ) 2<br />

(R)-citronellal enamine 96-99% ee<br />

OH<br />

OH<br />

H 2 ,Nicat<br />

isopulegol<br />

(-)-menthol<br />

Scheme 1.8 Rhodium-BINAP in Syn<strong>the</strong>sis <strong>of</strong> Menthol.<br />

Noyori BINAP system was applied successfully in <strong>the</strong> syn<strong>the</strong>sis <strong>of</strong> many important<br />

pharmaceutical drugs as <strong>the</strong> anti-inflammatory drug, naproxen, in 97% ee from α-aryl-acrylic<br />

acid. [38] and <strong>the</strong> antibacterial lev<strong>of</strong>loxacin obtained from hydroxyacetone through asymmetric<br />

hydrogenation <strong>of</strong> (R)-1,2-Propanediol. [39,40]<br />

Third major breakthrough in <strong>the</strong> field <strong>of</strong> asymmetric syn<strong>the</strong>sis was introduced by Barry<br />

Sharpless. [41] He developed a highly enantioselective epoxidation <strong>of</strong> allylic alcohols. His<br />

successful result was obtained by <strong>the</strong> use <strong>of</strong> a titanium-tartrate complex as <strong>the</strong> catalyst and<br />

water in a ratio <strong>of</strong> 1:2:1. For <strong>the</strong> process to be catalytically useful only a slight modification<br />

was required. Be<strong>for</strong>e catalyst <strong>for</strong>mation 4 Å molecular sieves had to be added. The molecular<br />

sieves act as a moisture scavenger and, <strong>the</strong>re<strong>for</strong>e, control <strong>the</strong> amount <strong>of</strong> water present in <strong>the</strong><br />

reaction mixture. In addition, <strong>the</strong> <strong>for</strong>mation <strong>of</strong> o<strong>the</strong>r, undesired titanium species which lead to<br />

17


non-enantioselective pathways is diminished (scheme 1.9). Recently, a fur<strong>the</strong>r decrease <strong>of</strong><br />

catalyst loading to 10 mol % has been achieved by replacing water with isopropanol.<br />

R 1 R 2<br />

R 3<br />

OH<br />

tBuOOH, Ti(O i Pr) 4<br />

L-(+)-DET, CH 2 Cl 2 ,-20 o C<br />

R 1 R 2<br />

O OH<br />

R 3<br />

70-90% yield<br />

90-98% ee<br />

HO<br />

L-(+)-DET =<br />

HO<br />

COOEt<br />

COOEt<br />

Scheme 1.9. Enantioselective Epoxidation <strong>of</strong> Allylic Alcohols.<br />

The work <strong>of</strong> those great minds was rewarded with a Noble prize in 2001 by <strong>the</strong> Royal<br />

Swedish academy <strong>of</strong> sciences.<br />

1.5.5. Asymmetric Syn<strong>the</strong>sis vs Kinetic Resolution vs <strong>Chiral</strong> Pool:<br />

From <strong>the</strong> above discussions it can be concluded that each approach <strong>of</strong> <strong>the</strong> three major<br />

approaches has advantages and disadvantages. Resolution suffers from a major drawback<br />

which is <strong>the</strong> low yield <strong>of</strong> <strong>the</strong> desired product; <strong>the</strong> maximum obtainable yield is only 50%<br />

from <strong>the</strong> racemates. In case <strong>of</strong> Kinetic dynamic resolution utilizing enzymes or chiral<br />

catalysts <strong>the</strong> yield can be improved. The yield in <strong>the</strong> kinetic resolutions can be improved by<br />

fast conversion <strong>of</strong> <strong>the</strong> (S)-enantiomer into <strong>the</strong> racemic mixture and <strong>the</strong> (R)-enantiomer reacts<br />

preferably to <strong>for</strong>m <strong>the</strong> desired product in high yield and ee. The ideal dynamic kinetic<br />

resolution reaction which approaches 100% conversion <strong>of</strong> 100% enantiomerically enriched<br />

product is <strong>the</strong> one in which krac>>kR>>kS (figure 1.4). If krac was in fact closer to or even<br />

slower than kR, <strong>the</strong> ee <strong>of</strong> <strong>the</strong> product would be lowered because <strong>the</strong> amount <strong>of</strong> (R) in solution<br />

would not be produced fast enough to make kS negligible.<br />

18


k R<br />

R P 1<br />

k rac<br />

S<br />

k S<br />

P 2<br />

Figure 1.4. Reaction Constants <strong>for</strong> Dynamic Kinetic Resolution.<br />

1.6. α-<strong>Chiral</strong> <strong>Amines</strong> Defining Terms:<br />

Amino compounds with a stereogenic centre at <strong>the</strong> position α-to <strong>the</strong> amino group are known<br />

as α-chiral amines.<br />

Ph<br />

NH 2<br />

Ph<br />

Et<br />

NH 2<br />

tBu<br />

NH 2<br />

NH 2<br />

COOtBu<br />

NH 2<br />

NH2<br />

NH 2<br />

Figure 1.5 Examples <strong>of</strong> α-<strong>Chiral</strong> <strong>Amines</strong>.<br />

They can be addressed as chiral amine <strong>for</strong> simplicity and we will try to stick to this<br />

nomenclature throughout <strong>the</strong> whole <strong>the</strong>sis. <strong>Chiral</strong> amines are useful intermediates <strong>for</strong> alkaloid<br />

natural product syn<strong>the</strong>sis, eg: morphine, codeine and tropane alkaloids. They are also<br />

incorporated in different block buster drugs as <strong>the</strong> billion dollar drugs, e.g. several ACE<br />

inhibitors and Flomax. [42]<br />

To understand <strong>the</strong> importance <strong>of</strong> this moiety in <strong>the</strong> asymmetric syn<strong>the</strong>sis it is estimated that<br />

at least 40% <strong>of</strong> all optically active pharmaceutical drugs contain this moiety. Un<strong>for</strong>tunately,<br />

19


80% <strong>of</strong> <strong>the</strong> syn<strong>the</strong>tic methods still rely on <strong>the</strong> classical resolution methods. [43] Searching<br />

literature in <strong>the</strong> last 30 years revealed that <strong>the</strong>re is a great lack in efficient methodologies <strong>for</strong><br />

syn<strong>the</strong>sis <strong>of</strong> chiral amines. [44] Different syn<strong>the</strong>tic strategies have been developed <strong>for</strong> <strong>the</strong><br />

syn<strong>the</strong>sis <strong>of</strong> chiral amines but as a general conclusion most <strong>of</strong> <strong>the</strong>se strategies suffer from low<br />

yield or stereoselectivity. One <strong>of</strong> <strong>the</strong> main challenges in <strong>the</strong> syn<strong>the</strong>sis <strong>of</strong> chiral amines comes<br />

from <strong>the</strong> lack <strong>of</strong> efficient methodology <strong>for</strong> <strong>the</strong> syn<strong>the</strong>sis <strong>of</strong> alkyl-alkyl amines. [45]<br />

This class <strong>of</strong> chiral amines are accessible in high yield and enantioselectivity through long<br />

tedious procedures. Also starting materials are <strong>of</strong>ten expensive and requires <strong>the</strong> use <strong>of</strong><br />

stoichiometric quantities <strong>of</strong> chirality inducing agents. The overall process is not atom<br />

economical and <strong>the</strong> waste production is high. As a conclusion <strong>the</strong> available processes <strong>for</strong><br />

chiral amine syn<strong>the</strong>sis suffer from many disadvantages resulting in an extreme difficulty <strong>for</strong><br />

<strong>the</strong>ir syn<strong>the</strong>sis on industrial scale utilizing <strong>the</strong> available methodologies. In this study we will<br />

try to identify <strong>the</strong> existing problem, demonstrate all possible available solutions and show our<br />

developed approach to solve this problem.<br />

1.7. α-<strong>Chiral</strong> <strong>Amines</strong> Importance:<br />

Enantiomerically pure amines with an α-stereocenter plays an important role in organic<br />

syn<strong>the</strong>sis. Their applications are innumerable: as chiral resolving agents, [46] chiral<br />

auxiliaries, [47] ligands in various asymmetric trans<strong>for</strong>mations [48] and as advanced building<br />

blocks in pharmaceutical and agrochemical industries. [49] They are also fruitful as chiral<br />

ligands in metal-complex catalysis. [50]<br />

NH 2<br />

(R)-or (S)-α-methylbenzylaine<br />

CO 2 H<br />

N<br />

H<br />

L-proline<br />

Ph<br />

OH<br />

H 2 N<br />

(R)-or (S)-phenylglycinol<br />

NH 2<br />

N<br />

(S)-3-aminoquinuclidine<br />

NH 2<br />

NH 2<br />

(1S,2S)-cyclohexane 1,2-<br />

diamine<br />

Figure 1.6. α-chiral amines Available Commercially.<br />

20


(S)-(α)-Methylbenzylamine and its enantiomer (R), appear to be ideal compounds as chiral<br />

auxiliaries or chiral building blocks <strong>for</strong> pharmaceutical and chemical industry. (S) and (R)<br />

enantiomers are inexpensively available in very high enantiomeric purity, which makes <strong>the</strong>m<br />

attractive as stereodifferentiating agents even <strong>for</strong> industrial scale operations. [51]<br />

It has been used in <strong>the</strong> syn<strong>the</strong>sis <strong>of</strong> biologically active molecules such as Labetalol (βblocker)<br />

and Tamsulosin. [52] Ano<strong>the</strong>r example is α-amino acids, eg; proline. Proline has a<br />

unique value in asymmetric processes; it is used as a ligand in transition metal-catalysis. [53]<br />

Recently proline and its derivatives were applied as highly efficient organocatalysts in<br />

different organic trans<strong>for</strong>mation as asymmetric Aldol, [54] Mannich [55] and Michael<br />

reactions. [56]<br />

Ano<strong>the</strong>r important class <strong>of</strong> chiral α-chiral amines which are used in syn<strong>the</strong>sis <strong>of</strong><br />

pharmaceutical building blocks is <strong>the</strong> quinuclidine family. An example <strong>of</strong> this class includes<br />

enantiopure 3-aminoquinuclidine, an important intermediate in <strong>the</strong> syn<strong>the</strong>sis <strong>of</strong> 5-HT 3<br />

serotonin ligands, [57] such as zacopride. Also diamines as (1S,2S)-Cyclohexane-1,2-diamine is<br />

used as chemo<strong>the</strong>rapeutic agents, [58] chiral auxiliary, transition metal-catalysis and in<br />

organocatalysis. [59] Of course <strong>the</strong>re are more examples <strong>of</strong> <strong>the</strong> available chiral amines which<br />

are used in pharmaceutical and agrochemical industry with great success <strong>for</strong> syn<strong>the</strong>sis <strong>of</strong><br />

natural products and drugs. [60] The following figure shows some examples <strong>of</strong> drugs having<br />

chiral amines(figure 1.7).<br />

HO<br />

N<br />

N<br />

H<br />

(S)-repaglinide<br />

(hypoglycemic agent)<br />

HO<br />

CONH 2<br />

H<br />

N<br />

Labetalol<br />

(β−blocker)<br />

H<br />

Figure 1.7. Examples <strong>of</strong> Drugs with α-<strong>Chiral</strong> <strong>Amines</strong>.<br />

21<br />

Ph<br />

Ph<br />

O<br />

O<br />

OH<br />

MeO<br />

O<br />

NH<br />

O<br />

HO<br />

NH 2<br />

H<br />

N<br />

O<br />

O<br />

amoxicillin<br />

(antibiotic)<br />

SO 2 NH 2<br />

N<br />

H<br />

O<br />

HCl<br />

AcO O OH<br />

HO O<br />

AcO<br />

H<br />

HO OBz<br />

Taxol<br />

(anticancer drug)<br />

H H<br />

S<br />

O<br />

COOH<br />

OEt<br />

(R)-tamsulsin hydrochloride<br />

(benign prostatic hyperplasia)<br />

Cl<br />

H 2 N<br />

H<br />

N<br />

N<br />

O<br />

O<br />

COOMe<br />

H<br />

Ph<br />

rivastigmine<br />

(Alzheimer)<br />

O<br />

N<br />

H<br />

OMe<br />

NMe 2<br />

N<br />

(S)-zacopride<br />

(5-HT 3 agonist)<br />

Ritalin<br />

(treatment <strong>of</strong> hyperdeficit disorder)


1.8. α-<strong>Chiral</strong> Amine Syn<strong>the</strong>sis Different Methodologies<br />

As mentioned previously chiral amines are key components <strong>of</strong> different pharmaceutical and<br />

agrochemical compounds. Over <strong>the</strong> last fifty years different methodologies have been<br />

developed <strong>for</strong> <strong>the</strong>ir syn<strong>the</strong>sis. Some <strong>of</strong> <strong>the</strong> methodologies are industrially viable and o<strong>the</strong>rs<br />

are better suited <strong>for</strong> pilot studies. Of course <strong>for</strong> a methodology to be applicable on industrial;<br />

scale it must fulfil certain features e.g. should be cost effective and waste generation should<br />

be low. Some processes are highly efficient in preparing chiral amines in high yield and<br />

stereoselectivity. Despite <strong>the</strong>ir efficiency <strong>the</strong>y suffer mainly from major drawbacks as lengthy<br />

multistep procedures which hinder <strong>the</strong>ir applications on industrial scale. Among <strong>the</strong> versatile<br />

strategies employed is <strong>the</strong> hydrogenation <strong>of</strong> enamine esters (diastereo and<br />

enantioselective), [61] hydrogenation <strong>of</strong> α- or β-N-acetylenamide esters, [62] 1,4-addition <strong>of</strong><br />

amines to enones, [63] Chemical, [64] and enzymatic [65] reductive amination <strong>of</strong> α- ketoacids,<br />

remote amination via C-H insertion [66] and hydroamination <strong>of</strong> olefins. [67]<br />

Reduction <strong>of</strong> unfunctionalized ketones and aldehydes is one <strong>of</strong> <strong>the</strong> major strategies <strong>for</strong> <strong>the</strong><br />

syn<strong>the</strong>sis <strong>of</strong> chiral amines. This strategy can be subdivided into various subdivisions which<br />

includes <strong>the</strong> following.<br />

1) N-acetylenamide reduction.<br />

2).Transfer hydrogenation or hydrogenation <strong>of</strong> imines<br />

3). Reductive amination <strong>of</strong> ketones<br />

4) Carbanion addition to aldimine and ketimine derivatives.<br />

5) Sequential aminationalkylation <strong>of</strong> aldehydes.<br />

The first three methodologies are closely related as <strong>the</strong>y use hydrogen from different hydride<br />

sources <strong>for</strong> <strong>the</strong> reduction <strong>of</strong> prochiral carbonyl compounds. The asymmetric version <strong>of</strong> <strong>the</strong>se<br />

methodologies has been developed extensively over <strong>the</strong> past few years. N-acetylenamide<br />

reduction and transfer hydrogenation or hydrogenation <strong>of</strong> imines will be discussed in details<br />

trying to shed light on <strong>the</strong>ir advantages and disadvantages and <strong>the</strong>ir applications. Reductive<br />

amination as <strong>the</strong> core <strong>of</strong> my work will be discussed showing its historical development over<br />

<strong>the</strong> last century and <strong>the</strong> major breakthroughs in <strong>the</strong> field during <strong>the</strong> last two decades. The<br />

application <strong>of</strong> reductive amination in pharmaceutical industry will be summarized in <strong>the</strong><br />

22


fourth chapter showing different drug and natural product categories prepared utilizing<br />

reductive amination.<br />

1.8.1. Imine and Enamide Syn<strong>the</strong>sis<br />

A discussion about <strong>the</strong> preparation <strong>of</strong> imine and enamide are necessary as most <strong>of</strong> <strong>the</strong><br />

examples in scientific journals focus mainly on <strong>the</strong> manipulation <strong>of</strong> imines (Nphosphinoylimines)<br />

or N-acyl enamines as starting materials without a clear picture about<br />

<strong>the</strong>ir preparations. The overall yield <strong>of</strong> <strong>the</strong> chiral amine products is very rarely discussed and<br />

<strong>the</strong>re<strong>for</strong>e a perspective in this regard needs to be established.<br />

R<br />

O<br />

R'<br />

NH 2 OH HCl<br />

MeOH<br />

R<br />

NOH<br />

R'<br />

Fe powder<br />

NHAc<br />

Ac 2 O<br />

AcOH. Toluene, 75 o C R R'<br />

Scheme 1.10 Syn<strong>the</strong>sis <strong>of</strong> N-Acyl Enamide from Ketone<br />

The commonly used method <strong>of</strong> enamide [68] syn<strong>the</strong>sis is <strong>the</strong> one in which <strong>the</strong> desired<br />

compound is syn<strong>the</strong>sized from different substituted ketones in two steps as carried out by<br />

Burk (scheme 1.10). [69]<br />

In <strong>the</strong> first step <strong>of</strong> this syn<strong>the</strong>sis <strong>the</strong> ketone is converted to an oxime with hydroxylamine<br />

hydrochloride in MeOH, <strong>the</strong> yield <strong>of</strong> <strong>the</strong> ketoxime is generally >90%. The next step is <strong>the</strong><br />

interaction <strong>of</strong> <strong>the</strong> resultant ketoxime with Iron powder and acetic anhydride with AcOH in<br />

toluene at a temperature <strong>of</strong> 75 °C. The yield <strong>of</strong> <strong>the</strong> enamide is generally between 30-60% in<br />

this step. [70]<br />

In general <strong>the</strong> enamide syn<strong>the</strong>sis methodology is low yielding process. Besides <strong>the</strong> possibility<br />

<strong>of</strong> diacetyl <strong>for</strong>mation which is considered ano<strong>the</strong>r drawback. The E/Z mixtures which are<br />

obtained with R’ -as non-hydrogen atom- are difficult to separate.<br />

1.8.2. Enantioselective Reduction <strong>of</strong> Enamides<br />

23


Enantioselective reduction <strong>of</strong> enamide is very interesting approach <strong>for</strong> <strong>the</strong> syn<strong>the</strong>sis <strong>of</strong> chiral<br />

amines from <strong>the</strong> enantioselectivity prospective. On <strong>the</strong> o<strong>the</strong>r hand, it is a four step procedure<br />

to get <strong>the</strong> final primary amine product. Two step process N-acyl enamide syn<strong>the</strong>sis and a<br />

fur<strong>the</strong>r two steps are involved, reduction <strong>of</strong> <strong>the</strong> enamides and hydrolysis <strong>of</strong> enamide, be<strong>for</strong>e<br />

<strong>the</strong> primary amine is obtained. The overall yield is rarely mentioned in literature. Through<br />

calculating each step yield and estimating <strong>the</strong> overall all yield it is obvious that <strong>the</strong> yield is<br />

low (usually below 50%). Enamides generally are obtained as E and Z mixtures, but this does<br />

not seem to affect enantioselectivity. Specific substrate categories can be only utilized in<br />

enamide reduction protocol, pinacolone and aryl-alkyl ketones. [71]<br />

We will focus on enamide methods allowing alkyl-alkyl and aryl-alkyl substituted α-chiral<br />

amine syn<strong>the</strong>sis. Ding immobilizes Feringa’s MonoPhos/Rh catalyst <strong>for</strong> <strong>the</strong> asymmetric<br />

hydrogenation <strong>of</strong> dehydro-α-amino acid esters and enamides. Treatment <strong>of</strong> <strong>the</strong> ditopic<br />

MonoPhos ligand with [Rh(cod)]BF 4 in DCM/toluene resulted in an immediate precipitation<br />

<strong>of</strong> an amorphous Rh-containing polymer, which were demonstrated to be effective catalysts<br />

<strong>for</strong> <strong>the</strong> asymmetric hydrogenation. Secondary amines were prepared in excellent<br />

enantioselectivity (scheme 1.11). [72]<br />

N<br />

P O<br />

O<br />

linker<br />

O<br />

O P<br />

N<br />

[Rh(cod) 2 ]BF 4<br />

CH 2 Cl 2 /Toluene<br />

O<br />

O<br />

P<br />

N<br />

[Rh]<br />

N<br />

P<br />

O<br />

O<br />

linker<br />

linker a:<br />

b:<br />

n<br />

c: single bond<br />

R<br />

O<br />

OCH 3<br />

NHAc<br />

R=H, CH3, Ph<br />

12a-c, 1 mol%<br />

H 2 , 40 atm, toluene<br />

O<br />

R OCH 3<br />

NHAc<br />

Full conversion<br />

94-96% ee<br />

12a-c, 1 mol%<br />

Ph<br />

NHAc H 2 , 40 atm, toluene Ph NHAc<br />

Scheme 1.11 Heterogeneous Catalysis with Self Supported Rh Catalysts.<br />

24


Burk was successful in reducing aryl-alkyl or alkyl-alkyl enamides with high ee (>95%)<br />

utilizing Rh[Me-DUPHOS] or Rh[Me-BPE] catalysts (Figure 1.7). The substrates are acyclic<br />

and benzocyclic aryl-alkyl ketones, and only two examples <strong>for</strong> alkyl-methyl ketones with<br />

sterically encumbered groups such as t-Bu (pinacolone) or adamantly groups as alkyl<br />

substituents. As described be<strong>for</strong>e <strong>the</strong>re is no in<strong>for</strong>mation about <strong>the</strong> yield from <strong>the</strong> starting<br />

ketone up to <strong>the</strong> final product. Ano<strong>the</strong>r issue regarding this work is <strong>the</strong> limited substrate<br />

breadth. [73]<br />

P<br />

P<br />

P<br />

P<br />

Figure 1.7 Examples <strong>of</strong> <strong>Chiral</strong> Ligands Used by Burk.<br />

Noyori has reported a general and straight<strong>for</strong>ward method <strong>for</strong> syn<strong>the</strong>sizing enantiomerically<br />

pure tetrahydroisoquinoline alkaloids through reduction <strong>of</strong> enamide. Ru-(S)-BINAP and Ru-<br />

(S)-BIPHEMP complexes. These complexes resulted in almost perfect enantioselectivities in<br />

hydrogenation <strong>for</strong> a wide array <strong>of</strong> tetrahydroquinolines. The present reaction provides access<br />

to a wide variety <strong>of</strong> alkaloids as morphinic and syn<strong>the</strong>tic morphinans and benzomorphans<br />

analogues (scheme 1.12). [74]<br />

MeO<br />

MeO<br />

NCHO<br />

1-4 bar H 2<br />

Ru-(S)-BINAP<br />

MeOH<br />

MeO<br />

MeO<br />

NCHO<br />

R=H,<br />

R<br />

OMe<br />

ee>99%<br />

R<br />

OMe<br />

OMe<br />

Scheme 1.12. Noyori Catalyst <strong>for</strong> Enamide.<br />

25


Ano<strong>the</strong>r methodology <strong>for</strong> <strong>the</strong> syn<strong>the</strong>sis <strong>of</strong> N-Boc-(R)-3-amino-2,3,4,5-tetrahydro-1 H-<br />

[1]benzazepin-2-one, which is an important intermediate <strong>for</strong> <strong>the</strong> preparation <strong>of</strong> an angiotensin<br />

converting enzyme inhibitor, based on asymmetric acyclic enamide hydrogenation has been<br />

reported by Merck (scheme 1.13). [75]<br />

N<br />

H<br />

O<br />

NHBoc<br />

3.4 bar H 2<br />

(S)-BINAP RuCl 2<br />

80% yield<br />

N<br />

H O<br />

82% ee<br />

NHBoc<br />

Scheme 1.13. Merck Based <strong>Methodology</strong> <strong>for</strong> Enamide Hydrogenation.<br />

Zhang worked extensively on <strong>the</strong> enantioselective reduction <strong>of</strong> N-acetyl enamides. Various<br />

types <strong>of</strong> chiral ligands were tested with rhodium catalysts showing extremely high<br />

enantioselectivities <strong>for</strong> aryl-alkyl, benzocyclic and ortho substituted aryl ketones using<br />

acceptable catalyst loading (0.1-1 mol %) (figure 1.8). [76]<br />

P<br />

P<br />

P<br />

P<br />

(R,R)-Binaphane<br />

(R,S,R,S)-MePennPhos<br />

OCH 3<br />

H 3 CO Ph<br />

H 3 CO PPh 2<br />

H 3 CO PPh 2<br />

H 3 CO Ph<br />

OCH 3<br />

Ph 2 P<br />

H<br />

H<br />

PPh 2<br />

(R,R)-BICP<br />

PPh 2<br />

O<br />

O<br />

PPh 2<br />

Phosphine-Phosphoramide ligand<br />

P<br />

R<br />

N R= Et or Me<br />

R<br />

(S)-o-Ph-hexMeO-BIPHEP<br />

Figure 1.8. <strong>Chiral</strong> Ligands Used by Zhang in Enamide Reduction.<br />

Rhodium catalyst proved to be <strong>the</strong> catalyst <strong>of</strong> choice <strong>for</strong> N-acetyl enamide hydrogenation <strong>the</strong><br />

following table summarizes <strong>the</strong> latest findings in this field (table 1.1). [77]<br />

26


Table 1.1 Rhodium Catalyzed Reduction <strong>of</strong> Enamide<br />

Substrate Ligand H 2 (bar) ee Configuration<br />

R = H, Ar = Ph Manniphos R=<br />

Me<br />

10 99.5 (R)<br />

R = H, Ar = p-CF 3 Ph Tangphos 1.4 99 (R)<br />

R = H, Ar = p-NO 2 Ph DiSquare P* 2 99 (R)<br />

R = H, Ar = p-ClPh MorPhos 55 99 (R)<br />

R = H, Ar = Ph Aaphos 10 87 (R)<br />

R = H, Ar = Ph (17) 10 93 (R)<br />

R = H, Ar = m-CO 2 MePh t-Bu-BisP* 3 97 (S)<br />

R = H, Ar = Ph (18) 20.6 96.5 (S)<br />

F 3 C CF 3<br />

Ph<br />

O<br />

Fe<br />

O<br />

P CF 3<br />

OR<br />

O<br />

P P<br />

N<br />

O P<br />

O<br />

O<br />

PPh 2 CF 3<br />

Ph O<br />

DiSquare P*<br />

ManniPHOS<br />

17<br />

H O<br />

PR<br />

N P<br />

2 O<br />

O<br />

O P N O PPh 2<br />

PPh 2<br />

Re<br />

CO<br />

Fe<br />

OC CO<br />

AaPHOS (R = Cy)<br />

MorPHOS 18<br />

Figure 1.9. Examples <strong>of</strong> Different Ligands Used in Enamide Reduction.<br />

27


1.9. Conclusion<br />

<strong>Chiral</strong>ity plays an important role in nature and almost all biological reactions are highly<br />

affected by chirality. Pharmaceutical drugs which were sold as racemate proved to have<br />

lower <strong>the</strong>rapeutic activity and more adverse effects compared to <strong>the</strong>ir single isomeric<br />

analogues. Catastrophic incidence <strong>of</strong> misuse <strong>of</strong> drug isomers <strong>for</strong>ced drug regulatory agencies<br />

and pharmaceutical companies to focus on developing new drug entities in a single isomeric<br />

<strong>for</strong>m. Despite <strong>the</strong> fact that agrochemicals and o<strong>the</strong>r fine chemicals are still marketed as<br />

racemate, many alerts suggest that selling <strong>the</strong>se products in a single isomeric <strong>for</strong>m will<br />

dramatically reduce <strong>the</strong> cost and toxicity. The significant importance <strong>of</strong> chiral agents derived<br />

chemists to develop various strategies <strong>for</strong> <strong>the</strong>ir preparation. Asymmetric syn<strong>the</strong>sis is a<br />

powerful convenient way <strong>for</strong> developing new entities <strong>of</strong> chiral agents. Developing new chiral<br />

ligands <strong>for</strong> organometallic catalyst and new organocatalysts <strong>for</strong>ms <strong>the</strong> core <strong>of</strong> organic<br />

chemistry research in <strong>the</strong> last three decades. <strong>Chiral</strong> amines syn<strong>the</strong>sis is one <strong>of</strong> <strong>the</strong> ultimate<br />

goals <strong>for</strong> asymmetric syn<strong>the</strong>sis. Chemoselective and bioselective methodologies have been<br />

developed <strong>for</strong> <strong>the</strong>ir syn<strong>the</strong>sis. N-acetyl enamaide hydrogenation has been extensively<br />

investigated in <strong>the</strong> last two decades <strong>for</strong> α-chiral amine syn<strong>the</strong>sis. It is four steps procedures<br />

<strong>for</strong> <strong>the</strong> final product with an overall yield not exceeding 50%. Several chiral ligands have<br />

been tested with rhodium catalyst <strong>for</strong> <strong>the</strong>ir hydrogenation resulting in 99% enantioselectivity.<br />

1.10. References:<br />

[1] M. Simonyi, Med. Res. Rev. 1984, 4, 359.<br />

[2] E. J. Ariens, E. W. Wuis, Clin. Pharmacol. Ther. 1987, 42, 361.<br />

[3] K. Mislow,Topics Stereochem. 1999, 22, 1.<br />

[4] J. Gal, Enantiomer, 1998, 3, 263.<br />

[5] J. B. Biot, Bull. Soc. Philamoth. Paris 1815, 190.<br />

[6] a) L. Pasteur, C. R. Acad. Sci. 1848, 26, 535; b)Pasteur, L. Ann. Chim. Physique 1848, 24,<br />

442.<br />

[7] a) J. H. van’t H<strong>of</strong>f, Arch. Neerl. Sci. Extracts Nat. 1974, 8, 445; b) E. W. Meijer, Angew.<br />

Chem. Int. Ed. 2001, 40, 3783-3789.<br />

28


[8] E. L. Eliel, S. H. Wilen, Stereochemistry <strong>of</strong> organic compounds, Wiley, New York, 1994,<br />

pp. 12.<br />

[9] M. Nógrádi, Stereoselective Syn<strong>the</strong>sis, Wiley-VCH, Weinheim, Germany, 1995, pp. 1-3.<br />

[10] U. Meierhenrich, Amino Acids and <strong>the</strong> Asymmetry <strong>of</strong> Life, Springer-Verlag, Berlin<br />

Heidelberg, 2008, pp. 17-39.<br />

[11] R. A. Sheldon. Chirotechnology, Industrial syn<strong>the</strong>sis <strong>of</strong> optically active compounds.<br />

Marcel Dekker, New York, 1993.<br />

[12] M. M. Islam, J. G. Mahdi, I. D. Bowen, Drug Safety 1997, 17, 149.<br />

[13] I. Wainer, Drug Stereochemistry: Analytical Methods and Pharmacology, 2nd<br />

Ed.; Marcel Dekker: New York, 1993.<br />

[14] a)D. Drayer, Clin. Pharmacol. Ther. 1986, 40, 125;b) K. M. Midha, G. McKay, M. J.<br />

Rawson, J. W. Hubbard, J. Pharm. Sci. 1998, 87, 797.<br />

[15] P. Knightley, H. Evans, E. Potter, M. Wallace, Suffer The Children: The Story <strong>of</strong><br />

Thalidomide, The Viking Press, New York, 1979.<br />

[16] D. D. Leipold, D. Kantoci, Jr. Murray, E. D., D. D. Quiggle, W. J. Wechter, <strong>Chiral</strong>ity<br />

2004, 16, 379.<br />

[17] C. D. Siebert, A. Hänsicke, T. Nagel, <strong>Chiral</strong>ity 2008, 20, 103.<br />

[18] S. Stinson, Chem. Eng. News 1998, 76, 83.<br />

[19] R. DiCicco, The future <strong>of</strong> S-(þ)-ibupr<strong>of</strong>en and o<strong>the</strong>r racemic switches, In <strong>Chiral</strong>’ 93<br />

USA, Spring Innovations: Stockport, UK, 1993.<br />

[20] a) S. Houlton, Manufacturing Chemist 2002, 73, 28; b) A. M. Rouhi, Chem & Eng News<br />

2003, 5, 56.<br />

[21] E. Francotte , W. Lindner, <strong>Chiral</strong>ity in Drug Research, Wiley-VCH Verlag GmbH &<br />

Co. KGaA, Weinheim, Germany, 2006, pp.3-24.<br />

[22] F. Leusen, G. Engel, Pharm. Pharmacol. 1999, 51, 1.<br />

[23] a) J. Caldwell, Modern Drug Discov. 1999, 2, 51; b) I. Agranat, H. Caner, J. Caldwell,<br />

Nat. Rev. Drug Discov. 2002, 21, 753.<br />

[24] A.M. Rouhi, Chem. Eng. News 2003, 81, 45.<br />

[25] V. A. Davankov, <strong>Chiral</strong>ity 1997, 9, 99. [26] R. A. Sheldon, Chirotechnology: Industrial<br />

Syn<strong>the</strong>sis <strong>of</strong> Optically Active Compounds, 2nd Ed., Marcel Dekker, New York, 1993.<br />

[26] L. Pasteur, C. R. Hebd. Seances Acad. Sci. 1853, 37, 162.<br />

[27] J. F. Larrow, E. N. Jacobsen, Org. Synth. 1998, 75, 1.<br />

29


[28] H. B. Kagan, J. C. Fiaud, Topics in Stereochemistry, E. L. Eliel and S. H. Wilen (Eds.)<br />

Vol. 18, Wiley, New York, 1988, pp. 249.<br />

[29] Y. Kim, E. T. Choi, M. H. Lee, Y. S. Park, Tetrahedron Lett. 2007, 48, 2833.<br />

[30] J. B. Craw<strong>for</strong>d, R. T. Skerlj, G. J. Bridger, J. Org. Chem., 2007, 72, 669.<br />

[31] R. Noyori, T. Ikeda, T. Ohkuma, M. Widhalm, M. Kitamura, H. Takaya, S. Akutagawa,<br />

N. Sayo, T. Saito, J. Am. Chem. Soc. 1989, 111, 9134.<br />

[32] V. Sunjic, F. Kajfez, I. Stromar, N. Blazevic, D. Kolbah, J. Heterocycl. Chem. 1973, 10,<br />

591.<br />

[33] Z. -Y. Wei, E. E. Knaus, Tetrahedron 1994, 50, 5569.<br />

[34] R. M. Scott, G. D. Armitage, E. Haddock, Optically active 4-phenoxyphenoxypropionic<br />

acid derivatives (Shell Internationale Research Maatschappij B. V., Neth.) 1980, 2, 946, 652,<br />

pp. 23.<br />

[35] A. G. Myers, B. H. Yang, H. Chen, J. L. Gleason, J. Am. Chem. Soc. 1994, 116, 9361.<br />

[36] W. S. Knowles, M. J. Sabacky, B. D. Vineyard, D. J. Weinkauf, J. Am. Chem. Soc. 1975,<br />

97, 2567; b) W. S. Knowles, Angew. Chem. Int. Ed. 2002, 41, 1998.<br />

[37] R. Noyori, Adv. Synth. Catal. 2003, 345, 15.<br />

[38] a) T. Ohta, H. Takaya, M. Kitamura, K. Nagai, R. Noyori, J. Org. Chem. 1987, 52, 3174;<br />

b) M. Kitamura, M. Yoshimura, M. Tsukamoto, R. Noyori, Enantiomer 1996, 1, 281.<br />

[39] For o<strong>the</strong>r syn<strong>the</strong>tic applications, see reviews: a) G. M. Ramos Tombo, G. Bellus, Angew.<br />

Chem. Int. Ed. Engl. 1991, 30, 1193; b) A. Börner, J. Holz in Transition Metals <strong>for</strong> Organic<br />

Syn<strong>the</strong>sis, Vol. 2 (Eds.: M. Beller, C. Bolm), Wiley- VCH, Weinheim, 1998, pp.3; c) J. M.<br />

Brown in Comprehensive Asymmetric Catalysis, Vol. 1 (Eds.: E. N. Jacobsen, A. Pfaltz, H.<br />

Yamamoto), Springer, Berlin, 1999, pp. 121; d) T. Ohkuma, M. Kitamura, R. Noyori in<br />

Catalytic Asymmetric Syn<strong>the</strong>sis, 2nd ed. (Ed.: I. Ojima), Wiley-VCH, New York, 2000.<br />

[40] a) M. Kitamura, M. Tokunaga, T. Ohkuma, R. Noyori, Tetrahedron Lett. 1991, 32,<br />

4163; b) M. Kitamura, M. Tokunaga, T. Ohkuma, R. Noyori, Org. Synth. 1993, 71, 1; c) K.<br />

Mashima, K. Kusano, N. Sato, Y. Matsumura, K. Nozaki, H. Kumobayashi, N. Sayo, Y.<br />

Hori, T. Ishizaki, S. Akutagawa, H. Takaya, J. Org. Chem. 1994, 59, 3064.<br />

[41] a)T. Katsuki, K. B. Sharpless, J. Am. Chem. Soc. 1980, 102, 5974; b) R. M. Hanson, K.<br />

B. Sharpless (1986) J. Org. Chem. 1986, 51,1922; c) E. N. Jacobsen, I. Marko , W. S.<br />

Mungall, G. Schröder, K. B. Sharpless, (1988) J. Am. Chem. Soc. 1988,110, 1968; d) G. Li,<br />

H. T. Chang, K. B. Sharpless, Angew. Chem. Int. Ed. 1996, 35, 451; e) R. A. Johnson, K. B.<br />

Sharpless, in Catalytic Asymmetric Syn<strong>the</strong>sis (Ed.: I. Ojima) VCH, Weinheim, 1993.<br />

30


[42] a) J. Blacker, Innovations in Pharmaceutical Technology 2001, 1, 77; b) H. -U. Blaser,<br />

F. Spindler, A. Studer, App. Catal. Gen. 2001, 221, 119; c) H. -U. Blaser, M. Eissen, P. F.<br />

Fauquex. K. Hungerbuhler, E. Schmidt, G. Sedelmeier, M. Studer, Asymmetric Catalysis on<br />

Industrial Scale: Challenges, Approaches, and Solutions; H.-U. Blaser, E. Schmidt Eds.;<br />

Wiley-VCH Verlag GmbH & Co. KGaA: Weinheim, Germany, 2004, pp. 91.<br />

[43] a) J. Blacker, Innovations in Pharmaceutical Technology 2001, 1, 77.<br />

[44] a) P. Kukula, R. Prins, Topics in Catalysis 2003, 25, 29; b) M. Bissen, C. Pinel, Topics<br />

in Catalysis 2003, 25, 43; c) T. Valaivan, W. Bhanthumnavin, Y. S. -Anant, Curr. Org.<br />

Chem. 2005, 9, 1315.<br />

[45] a) T. C. Nugent, A. K. Ghosh, V. N. Wakchaure, R. R. Mohanty, Adv. Synth. Catal.<br />

2006, 348, 1289; b) T. C. Nugent, <strong>Chiral</strong> Amine Syn<strong>the</strong>sis - Strategies, Examples, and<br />

Limitations. In Process Chemistry in <strong>the</strong> Pharmaceutical Industry, Second Edition:<br />

Challenges in an Ever-Changing Climate, T. F. Braish, K. Gadamasetti Eds.; CRC Press-<br />

Taylor and Francis Group: New York, 2008; c) T. C. Nugent, A. K. Ghosh, V. N.<br />

Wakchaure, R. R. Mohanty, WO2006030017, 2006; d) T. C. Nugent, V. N. Wakchaure, A. K.<br />

Ghosh, R. R. Mohanty, Org. lett. 2005, 7, 4967; e) T. C. Nugent, A. K. Ghosh, Eur. J. Org.<br />

Chem. 2007, 3863.<br />

[46] J. Jacques, A. Collet, S. H. Wilen, Enantiomers, Racemates and Resolutions, John Wiley<br />

and Sons: New York, 1981.<br />

[47] J. K. Whitesell, Chem. Rev. 1989, 89, 1581.<br />

[48] E. M. Vogl, H. Groger, M. Shibasaki, Angew. Chem. Int. Ed. 1999, 38, 1570.<br />

[49] a) M. Nogradi, Stereoselective Synthsis, 2 nd Ed. , VCH: Weinheim, Germany, 1995; b) S.<br />

Arrasate, E. Lete, N. Sotomayor, Tetrahedron: Asymmetric 2001, 12, 2077; c) G. M.Coppola,<br />

Asymmetric Syn<strong>the</strong>sis: Construction <strong>of</strong> <strong>Chiral</strong> Molecules using Amino Acids, Wiley, New<br />

York, 1987; d) G. Bringmann, In The Alkaloids, (Ed.: A. Brossi) Acadamic Press, New York,<br />

1986, 141; e) D. Lucet, T. Le Gall, C. Mioskowski, Angew. Chem., Int. Ed. 1998, 37, 2580; f)<br />

J. E. Clifton, I. Collins, P. Hallett, D. Hartley, L. H. C. Lunts, P. D. Wicks J. Med. Chem.<br />

1982, 25, 670; g) R. M. Barmore, S. R. Logan, B. C. Van Wagenen, Tetrahedron Lett. 1998,<br />

39, 3451.<br />

[50] J. W. Canary, C. S. Allen, J. M. Castagnetto, Y. Wang, J. Am. Chem. Soc. 1995, 117,<br />

8484.<br />

31


[51] a) T. Soe, N. Fukui, T. Hino, M. Nakagawa, Heterocycles 1996, 42, 347; b) D. Xu, K.<br />

Prasad, O. Repic, T. J. Blacklock, T. J. Tetrahedron: Asymmetry 1997, 9, 1445; c) C. A.<br />

Parrodi, E. Juaristi, L. Quintero, L., A. Clara-Sosa, Tetrahedron: Asymmetry 1997, 7, 1072.<br />

[52] J. E. Clifton, I. Collins, P. Hallett, D. Hartley, L. H. C. Lunts, P. D. Wicks, J. Med.<br />

Chem. 1982, 25, 670.<br />

[53] a) G. Xu, S. R. Gilbertson, Tetrahedron Lett. 2003, 44, 953; b) Y. B. Kim, M. K. Kim,<br />

S. H. Kang, T. H. Kim, Synlett 2005, 1995.<br />

[54] B. List, R. A. Lerner, C. F. Barbas, J. Am. Chem. Soc. 2000, 122, 2395.<br />

[58] M. Yamaguchi, N. Yokota, T. Minami, J. Chem. Soc., Chem. Commun. 1991, 1088.<br />

[56] B. List, J. Am. Chem. Soc. 2000, 122, 9336.<br />

[57] a) T. C. Nugent, R. Seemeyer, Org. Process. Res. Dev. 2006, 10, 144; b) M. Langlois, C.<br />

Meyer, J. L. Soulier, Synth. Commun. 1992, 22, 1895; c) H. Parnes, E. J. Shelton, J. Labelled<br />

Compd. Radiopharm. 1996, 38, 19.<br />

[58] E. T. Michalson, J. Szmuszkovicz, Prog. Drug. Res. 1989, 33, 135.<br />

[59] D. Lucet, T. Le Gall, C. Mioskowski, Angew. Chem., Int. Ed. 1998, 37, 2580.<br />

[60] G. M. Coppola, Asymmetric Syn<strong>the</strong>sis: Construction <strong>of</strong> <strong>Chiral</strong> Molecules using Amino<br />

Acids, Wiley, New York, 1987.<br />

[61] a) T. Bunlaksananusorn, F. Rampf, Synlett 2005, 17, 2682; b) N. Ikemoto, D. M. Tellers,<br />

S. D. Dreher, J. Liu, A. Huang, N. R. Rivera, E. Njolito, Y. Hsiao, J. C. McWilliams, J. M.<br />

Williams, J. D. Armstrong, Y. Sun, D. Mathre, E. J. J. Grabowski, R. D. Tillyer, J. Am.<br />

Chem. Soc. 2004, 126, 3048; c) Y. Hsiao, N. R. Rivera, T. Rosner, S. W. Krska, E. Njolito, F.<br />

Wang, Y. Sun, J. D. Armstrong, E. J. J. Grabowski, R. D. Tillyer, F. Spindler, C. Malan, J.<br />

Am. Chem. Soc. 2004, 126, 9918.<br />

[62] a) X. –P. Hu, Z. Zheng, Org. Lett. 2005, 7, 419; b) Y. J. Zhang, K. Y. Kim, J. H. Park,<br />

C. E. Song, K. Lee, M. S. Lah, S.-G. Lee, Adv. Synth. Catal. 2005, 347, 563; c) J. You, H.-J.<br />

Drexler, S. Zhang, C. Fischer, D. Heller, Angew. Chem., Int. Ed. 2003, 42, 913.<br />

[63] P. H. Phau, J. G. de Vries, K. K. Hiia, Adv. Synth. Catal. 2005, 347, 1775.<br />

[64] R. Kadyrov, T. H. Reirmeier, U. Dingerdissen, V. I. Tararov, A. Borner, J. Org. Chem.<br />

2003, 68, 4067.<br />

[68] A. Menzel, H. Werner, J. Altenbuchner, H. Groeger, Eng. in Life Sci. 2004, 4, 573.<br />

[66] a) H. Lebel, K. Huard, Org. Lett. 2007, 9, 639; b) M. Kim, J. V. Mulcahy, C. G. Espino,<br />

J. Du Bois, Org. Lett. 2005, 7, 4685; c) C. G. Espino, K. W. Fiori, M. Kim, J. Du Boisn, J.<br />

Am. Chem. Soc. 2004, 126, 15378.<br />

32


[67] a) H. Qin, N. Yamagiwa, S. Matsunaga, M. Shibasaki, J. Am. Chem. Soc. 2006, 128,<br />

1611; b) A. Zulys, M. Dochnahl, D. Hollmann, K. Loehnwitz, J.-S. Herrmann, P. W. Roesky,<br />

S. Blechert, Angew. Chem., Int. Ed. 2005, 44, 7794.<br />

[68] For o<strong>the</strong>r methods <strong>of</strong> enamide preparation, see a) H. B. Kagan, N. Langlois, T. P. Dang,<br />

J. Organomet. Chem. 1975, 90, 353; b) D. Sinou, H. B. Kagan, J. Organomet. Chem. 1976,<br />

114, 325; c) T. Morimoto, M. Chiba, K. Achiwa, Chem. Pharm. Bull. 1992, 40, 2894; d) G.<br />

R. Lenz, Syn<strong>the</strong>sis 1978, 489; e) D. M. Tsachen, L. Abramson, D, Chai, R. Desmond, U. -H.<br />

Dolling, L. Frey, S. Karady, Y. -J. Shi, T. R. Verhoeven, J. Org. Chem. 1995, 60, 4324.<br />

[69] a) M. J. Burk, Y. M. Wang, J. R. Lee, J. Am. Chem. Soc. 1996, 118, 5142; b) M. J. Burk,<br />

G. Casy, N. B. Johnson, J. Org. Chem. 1998, 63, 6084.<br />

[70] a) M. J. Burk, Y. M. Wang, J. R. Lee, J. Am. Chem. Soc. 1996, 118, 5142; b) M. J. Burk,<br />

G. Casy, N. B. Johnson, J. Org. Chem. 1998, 63, 6084.<br />

[71] a) M. J. Burk, G. Casy, N. B. Johnson, J. Org. Chem. 1998, 63, 6084. b) H. Bernsmann,<br />

M. van den Berg, R. Hoen, A. J. Minnaard, G. Mehler, M. T. Reetz, J. G. de Vries, B. L.<br />

Feringa, J. Org. Chem. 2005, 70, 943; c) R. Hoen, M. van den Berg, H. Bernsmann, A. J.<br />

Minnaard, J. G. de Vries, B. L. Feringa, Org. lett. 2004, 6, 1433.<br />

[72] K. Ding, Y. Uozumi, Handbook <strong>of</strong> Asymmetric Heterogeneous Catalysis, Wiley-VCH<br />

Verlag GmbH & Co. KGaA, 2008.<br />

[73] a) M. J. Burk, Y. M. Wang, J. R. Lee, J. Am. Chem. Soc. 1996, 118, 5142; b) M. J. Burk,<br />

G. Casy, N. B. Johnson, J. Org. Chem. 1998, 63, 6084.<br />

[74] M. Kitamura , Y. Hsiao , M. Ohta , M. Tsukamoto , T. Ohta , H. Takaya, R. Noyori, J.<br />

Org. Chem. 1994, 59, 297 .<br />

[75] J. D. Armstrong, K. K. Eng, J. L. Keller, R. M. Purick, F. W. Hartner, Jr. , W.B. Choi,<br />

D. Askin, R. P. Volante, Tetrahedron Lett. 1994, 35, 3239.<br />

[76] a) Z. Guoxin, X. Zhang, J. Org. Chem. 1998, 63, 5871; b) W. Tang, X. Zhang, Angew.<br />

Chem., Int. Ed. 2002, 41, 1612; c) W. Tang, X. Zhang, Angew. Chem., Int. Ed. 2002, 41,<br />

1612.<br />

[77] P.G. Andersson, I. J. Munslow, Modern Reduction Methods, Wiley-VCH, Weinheim,<br />

2008.<br />

33


Chapter 2<br />

Imine Reduction<br />

2.1. Historical View<br />

Reduction <strong>of</strong> imines with chiral catalysts and hydride source to prepare α-chiral amines with<br />

high yield and enantioselectivity represents an important achievement in organic and<br />

pharmaceutical chemistry over <strong>the</strong> last two decades. Different catalytic systems have been<br />

developed, heterogeneous and homogenous systems and recently organocatalysis were<br />

investigated. Few <strong>of</strong> <strong>the</strong>se systems were applicable on industrial scale. [1] Historically<br />

heterogeneous systems were developed first. Many attempts to attach chiral auxiliary to<br />

heterogeneous catalysts (Pt/C, Pd/C and R-Ni) were not successful to attain high<br />

enantioselectivity.<br />

Homogenous systems proved to be <strong>the</strong> systems <strong>of</strong> choice to reduce imines on <strong>the</strong> laboratory<br />

and on <strong>the</strong> industrial scales. In general all <strong>of</strong> <strong>the</strong>se systems require <strong>the</strong> use <strong>of</strong> activated imines<br />

as substrates in which nitrogen atom is attached to a bulk group (phenyl, phosphinoyl, chiral<br />

auxiliary). [2,3] These groups have to be removed in <strong>the</strong> final step which adds to <strong>the</strong> total<br />

number <strong>of</strong> steps from <strong>the</strong> prochiral ketone to <strong>the</strong> α-chiral amine. Removal <strong>of</strong> <strong>the</strong>se groups<br />

requires harsh conditions which is not compatible with many sensitive groups.<br />

In general cyclic imines are easier in reduction with higher enantioselectivity as <strong>the</strong>y do not<br />

have anit/syn con<strong>for</strong>mation. [4] They are considered important intermediates <strong>for</strong> many<br />

pharmaceutical drugs. Acyclic aryl imines were successfully hydrogenated with high<br />

enantioselectivities and yields. Metals as Rh, Ru, Ir and Ti were useful in this process. Ir was<br />

<strong>the</strong> best metal <strong>for</strong> imine reduction with different chiral auxiliaries. Ru catalysts which were<br />

developed by Noyori and proved to be highly effective in <strong>the</strong> reduction <strong>of</strong> ketones showed<br />

limited success. Titanium system which was developed by Buchwald in <strong>the</strong> nineties gave<br />

superior results in terms <strong>of</strong> yield and enantioselectivity but <strong>the</strong>ir industrial application was<br />

not that successful. [1] In this section <strong>of</strong> <strong>the</strong> <strong>the</strong>sis, I will focus on <strong>the</strong> achievments in <strong>the</strong> field<br />

34


<strong>of</strong> imine reduction in <strong>the</strong> past eight years. Of course in <strong>the</strong> nineties great achievements were<br />

accomplished <strong>for</strong> complete picture please refer to <strong>the</strong> following review. [2]<br />

2.2. Asymmetric Reduction <strong>of</strong> N-Phosphinoyl Imines<br />

Of <strong>the</strong> useful imine substrates examined to date, N-phosphinoyl imines hold <strong>the</strong> advantage <strong>of</strong><br />

being reduced with high yields and ees. The steric bulk <strong>of</strong> <strong>the</strong> diphenylphosphine group<br />

affects <strong>the</strong> geometric <strong>for</strong>m <strong>of</strong> imine (only anti isomer is obtained). [5] To access N-phosphinoyl<br />

imines researchers universally begin with a ketone and convert it to an oxime (high yield).<br />

Oximes are readily prepared from ketones and HCl.H 2 NOH, pyridine in ethanol by mixing<br />

1.0 equiv <strong>of</strong> ketone and 1.1 hydroxyl amine hydrochloride and 1.1 equiv <strong>of</strong> pyridine. [6]<br />

Treatment <strong>of</strong> <strong>the</strong> oxime with chlorodiphenylphosphine [Ph 2 P(O)Cl] at –45-78 °C provides <strong>the</strong><br />

N-phosphinoyl imine in mediocre to good yields. For example aryl alkyl N-phosphinoyl<br />

imines provide yield in <strong>the</strong> range <strong>of</strong> 40-70%, while alkyl alkyl N-phosphinoyl imines provide<br />

yields in <strong>the</strong> range <strong>of</strong> 50-70%. The product is purified with column chromatography (scheme<br />

2.1). [7]<br />

2.2.1. Syn<strong>the</strong>sis <strong>of</strong> N-Phosphinoyl Imines.<br />

high<br />

yield<br />

N<br />

OH<br />

R 1 R 2<br />

Ph 2 P(O)Cl<br />

Et 3 N, -45 °C<br />

O<br />

R 1 R 2<br />

+<br />

H 2 N<br />

O<br />

P Ph<br />

Ph<br />

N<br />

R 1 R 2<br />

CH 2 Cl 2<br />

P(O)Ph 2 P(O)Ph2<br />

reduction<br />

high yield<br />

HN<br />

R 1<br />

* R 2<br />

high<br />

NH 2<br />

yield R 1 * R 2<br />

R 1 = aryl, alkyl, heterocylic<br />

R2 = alkyl<br />

1- Ph 2 P(O)Cl, CH 2 Cl 2 ,-78°C<br />

2- hydrolytic work-up<br />

NOH<br />

H 2 NOH<br />

HCl<br />

O<br />

Scheme 2.1. General Strategies <strong>for</strong> Syn<strong>the</strong>sis <strong>of</strong> N-Phosphinoyl Imines.<br />

35


Figure 2.1: General Substrates Categories.<br />

N P(O)Ph 2<br />

N P(O)Ph 2<br />

N P(O)Ph 2<br />

1<br />

2<br />

R<br />

3<br />

Figure 2.2: General Catalysts Categories.<br />

H 3 C H<br />

PR 2<br />

Fe PR 2<br />

O<br />

t-Bu<br />

OMe<br />

josiphos type<br />

[Rh(nbd) 2 ]BF 4<br />

(R)-(S)-R 2 PF-PR 2<br />

R = cycohexyl, t Bu 2<br />

Catalyst 1<br />

O<br />

N N<br />

Co<br />

O O<br />

Catalyst 2<br />

O<br />

O<br />

O<br />

O<br />

P<br />

P<br />

t-Bu<br />

CuCl<br />

t-Bu<br />

2<br />

t-Bu<br />

OMe<br />

2<br />

(R)-(-)-DTBM-SEGPHOS<br />

Ph<br />

Ph<br />

N<br />

Ir<br />

N<br />

H 2<br />

Catalyst 4<br />

Cl<br />

Ph<br />

Ph<br />

N<br />

Rh<br />

N<br />

H 2<br />

Catalyst 5<br />

Cl<br />

NC<br />

Catalyst 3<br />

R<br />

O<br />

N<br />

O Cl<br />

Re<br />

N Cl<br />

O<br />

OPPh 3<br />

R<br />

R= 4- t Bu-ph<br />

Ph<br />

Ph<br />

O<br />

Catalyst 6<br />

Ph<br />

N N<br />

H H<br />

ZnEt 2<br />

Catalyst 7<br />

Ph<br />

O PPh 2<br />

O PPh 2<br />

O<br />

Pd(CF 3 CO 2 ) 2<br />

L-5 (S)-SEGPHOS<br />

Catalyst 8<br />

S<br />

S<br />

n<br />

NH HN<br />

n=2<br />

ZnEt 2<br />

Catalyst 9<br />

S<br />

n<br />

S<br />

2.2.2. Different Substrates Categories.<br />

Phenyl alkyl N-phosphinoyl imines (Structure 1, 2, 3, figure 2.1) have been extensively<br />

investigated over <strong>the</strong> last few years. We will focus our investigation on <strong>the</strong> results <strong>for</strong> <strong>the</strong> last<br />

36


8 years beginning from <strong>the</strong> year 2000.Blaser tested Rh-ferrocenyl-catalyst which he used 1.0<br />

mol % <strong>of</strong> this catalyst (catalyst 1, figure 2.2), 70 bar (1015 psi) <strong>of</strong> H 2 , CH 3 OH at 60 °C over<br />

21 h, <strong>the</strong> ee was 99% with full conversion (structure 1, figure 2.1). [8] He tested also his<br />

system <strong>for</strong> different substituted phenyl alkyl N-phosphinoyl imines. p-OMe phenyl (62% ee),<br />

p-CH 3 phenyl (97% ee), p-CF 3 phenyl (93% ee) were successfully reduced. For p-Cl phenyl<br />

derivative, <strong>the</strong> ee was only 28 % and improved to 67% with ano<strong>the</strong>r chiral ligand (structure 3,<br />

figure 2.1).<br />

Yamada developed <strong>the</strong> use <strong>of</strong> 1.0 mol % <strong>of</strong> cobalt based catalyst (catalyst 2, figure 2.2), 1.5<br />

equiv NaBH 4 in CH 3 Cl, 0 °C, 4 h, providing 97% isolated yield with 90% ee (structure 1,<br />

figure 2.1). [9]<br />

Lipshutz developed <strong>the</strong> use <strong>of</strong> <strong>the</strong> DTBM-SEGPHOS ligand with CuCl (catalyst 3, figure<br />

2.2). [10] He used 6.0 mol % <strong>of</strong> <strong>the</strong> catalyst, 3.0 equiv tetramethyldisiloxane (TMDS), 6.0 mol<br />

% NaOMe, 3.3 equiv t-BuOH, toluene, 25 °C, 17 h, <strong>the</strong> ee <strong>for</strong> (structure 1, figure 2.1) was<br />

96% with 99% isolated yield. Cooling <strong>the</strong> reaction to -25 °C increased <strong>the</strong> ee to 99% with<br />

slightly lower yield (94%) <strong>for</strong> (structure 2, figure 2.1). Different substituted phenyl alkyl N-<br />

phosphinoyl imines were tested. p-Br phenyl (96% ee, 95% yield), p-C 3 F phenyl (97% ee,<br />

94% yield), p-OMe phenyl (94% ee, 98% yield) were reduced successfully (structure 3,<br />

figure 2.1). They were able to reduce sterically hindered imine (phenyl iso-propyl n-<br />

phosphinoyl imine) with 94% ee with 90% yield. The ee was improved to 97% ee with 93%<br />

yield at -25 °C.<br />

Avecia Limited reported <strong>the</strong> use <strong>of</strong> CATHyTM (Catalytic Asymmetric Transfer<br />

Hydrogenation) catalysts (catalyst 4-5, figure 2.1). [11] They utilized 24 equiv <strong>of</strong> Et 3 N/HCO 2 H<br />

(2:5 ratio) <strong>for</strong> reduction <strong>of</strong> phenyl methyl N-phosphinoyl imine (structure 1, figure 2.1) with<br />

86% ee, <strong>for</strong> 1-acetyl naphthalene derivative <strong>the</strong> ee was 99% and <strong>for</strong> 2-octanone derived N-<br />

phosphinoyl imine <strong>the</strong> ee was 95%.<br />

Toste and coworkers developed a highly efficient chiral ligand <strong>for</strong> rhenium metal. [12] The use<br />

<strong>of</strong> this ligand eliminates <strong>the</strong> need <strong>of</strong> restrictive inert condition (open flask technique). Using<br />

3.0 mol % <strong>of</strong> <strong>the</strong> catalyst (catalyst 6, figure 2.2), 2.0 equiv <strong>of</strong> diphenylmethylsilane (DPMS-<br />

H), CH 2 Cl 2 , 25 °C over 72 h product ee was provided in >99% albeit in mediocre yield (51%)<br />

(structure 1, figure 2.1). They tested o<strong>the</strong>r substituted phenyl alkyl N-phosphinoyl imines.<br />

37


Phenyl n-propyl N-suilphinyol imine was reduced with 68% yield and >99% ee. p-OMe<br />

phenyl (98% ee, 61% yield), p-CF 3 phenyl (98% ee, 78% yield), p-I phenyl (99% ee, 71%<br />

yield) methyl N-phosphinoyl imines were reduced. The system was also applicable <strong>for</strong><br />

heterocyclic derivatives.<br />

The use <strong>of</strong> Zn/diamine catalyst was reported by Yun. [13] One <strong>of</strong> <strong>the</strong> problems related to <strong>the</strong><br />

use <strong>of</strong> Zn <strong>for</strong> <strong>the</strong> catalytic enantioselective reduction <strong>of</strong> imines is <strong>the</strong> strong Zn-N bond<br />

<strong>for</strong>med between Zn and amine product. The source <strong>of</strong> hydride should affect this bond without<br />

affecting <strong>the</strong> bond between <strong>the</strong> metal and <strong>the</strong> diamine. They thought that <strong>the</strong> choice <strong>of</strong> <strong>the</strong><br />

substituent attached to <strong>the</strong> imine nitrogen will be crucial, so <strong>the</strong>y selected diphenylphoshinoyl<br />

moiety. Using 5.0 mol % <strong>of</strong> <strong>the</strong> catalyst (catalyst 7, figure 2.2), 3.0 equiv <strong>of</strong><br />

polymethylhydrosiloxane (PMHS), THF/MeOH, 25 °C, 12 h, <strong>the</strong> ee was 97% and 86%<br />

isolated yield <strong>for</strong> phenyl methyl N-phosphinoyl imine (structure 1, figure 2.1). For <strong>the</strong> phenyl<br />

ethyl N-phosphinoyl imine, <strong>the</strong>y achieved 96% ee with 82% yield (structure 2, figure 2.1). p-<br />

Br phenyl (97% ee, 77% yield) and p-OMe phenyl (96% ee, 83% yield) methyl N-<br />

phosphinoyl imines were reduced (structure 3, figure 2.1)<br />

Zhou used Pd(CF 3 CO 2 ) 2 /(S)-SEGPHOS <strong>for</strong> reduction <strong>of</strong> this category <strong>of</strong> imines. [14] Using 2.0<br />

mol % <strong>of</strong> <strong>the</strong> catalyst (catalyst 8, figure 2.2), 69 bar (1015 psi) <strong>of</strong> H 2 , 2,2,2 trifluroethanol, 25<br />

°C, 8-12 h, <strong>the</strong> ee was 96% with 98% yield <strong>for</strong> phenyl methyl N-phosphinoyl imine (structure<br />

1, figure 2.1). His catalyst proved to be highly efficient <strong>for</strong> <strong>the</strong> reduction <strong>of</strong> different<br />

substituted phenyl methyl N-phosphinoyl imines. p-CH 3 phenyl (97% ee, 93% yield), p-F<br />

phenyl (94% ee, 87% yield), p-Cl phenyl (94% ee, 90% yield), p-OMe phenyl (96% ee, 96%<br />

yield), m-OMe phenyl (96% ee, 97% yield), o-OMe phenyl (99% ee, 80% yield) methyl N-<br />

phosphinoyl imines were tested (structure 3, figure 2.1).<br />

Zn-diamino-bis(tert-thiophene) catalyst was tested by Ronchi. [15] Using 5.0 mol % <strong>of</strong> <strong>the</strong><br />

catalyst (catalyst 9, frigure 2.2), 5.0 equiv <strong>of</strong> PMHS, THF/MeOH, 0 °C, 3 h, <strong>the</strong> ee was 97%<br />

and 70% yield <strong>for</strong> phenyl methyl N-Phosphinoyl imine (structure 2, figure 2.1).<br />

38


2.2.3. Nguyen Special Substrates.<br />

Apart from <strong>the</strong> classical substrates (structure 1-3, figure 2.1) investigated, substrates which<br />

were tested by Nguyen were unique. By today’s standards <strong>the</strong> use <strong>of</strong> stoichiometric quantities<br />

<strong>of</strong> a chiral reducing agent are not acceptable, but in this case Nguyen has developed a system<br />

capable <strong>of</strong> accepting a much broader substrate scope and <strong>the</strong>rein lies <strong>the</strong> significance <strong>of</strong> his<br />

research. Using stoichiometric amounts <strong>of</strong> (S)-BINOL/AlMe 3 with isopropanol as a source <strong>of</strong><br />

hydrogen to reduce different N-phosphinoyl imines. Subtle difference between small alkyl<br />

groups could be distinguished. They reported 93% ee with 85% yield with imine derived<br />

from 3-octanone. This class <strong>of</strong> substrates is syn<strong>the</strong>sized utilizing carbanion chemistry because<br />

hydrogen reduction gives low ee (15%). As far as we know this is <strong>the</strong> only example <strong>for</strong><br />

reduction <strong>of</strong> 3-octanonene utilizing hydrogen with such high enantioselectivity. He tested his<br />

system <strong>for</strong> o<strong>the</strong>r N-phosphinoyl imines and reported high yields and ees (table 2.1). [16]<br />

Table 2.1. Different substrate Categories introduced by Nguyen<br />

entry imine yield(product) ee(%)<br />

1<br />

2<br />

3<br />

4<br />

5<br />

6<br />

1<br />

2<br />

3<br />

4<br />

5<br />

6<br />

Aryl<br />

Ph<br />

Ph<br />

Ph<br />

Ph<br />

1-naphthyl<br />

2-naphthyl<br />

Alkyl<br />

Me<br />

Et<br />

n Pr<br />

i Pr<br />

Me<br />

Me<br />

85%<br />

85%<br />

84%<br />

79%<br />

80%<br />

84%<br />

96%<br />

95%<br />

94%<br />

96%<br />

98%<br />

96%<br />

7<br />

Ph<br />

N<br />

P(O)Ph 2<br />

7 84% 94%<br />

Me<br />

8 8 80% 94%<br />

39


9 9 84% 94%<br />

10 10 85% 93%<br />

2.3.Asymmetric Reduction <strong>of</strong> N-aryl Imines<br />

2.3.1. Syn<strong>the</strong>sis <strong>of</strong> N-Aryl Imines.<br />

N-aryl imines are syn<strong>the</strong>sized from <strong>the</strong>ir corresponding ketones and N-aryl amines. They are<br />

mixed in anhydrous toluene in <strong>the</strong> presence <strong>of</strong> NaHCO 3 and 4Å ctivated molecular sieves.<br />

The mixture is heated <strong>for</strong> 12 h at 80 °C. The product is purified by crystallization and<br />

distillation (scheme 2.2). [17]<br />

NH 2<br />

R 3<br />

O<br />

+<br />

R 1 R 2<br />

NaHCO 3 ,4ÅMS<br />

toluene, 80 °C,12h<br />

( mediocre-good yield)<br />

N<br />

R 1 R 2<br />

R 3<br />

reduction<br />

high yield<br />

R 3<br />

HN<br />

cerium ammonium nitrate NH 2<br />

MeOH/H 2 O, 0 °C, 6h<br />

R 1<br />

* R 2<br />

R 1<br />

* R 2<br />

(good-high yield)<br />

Scheme 2.2 General Strategy <strong>for</strong> Syn<strong>the</strong>sis <strong>of</strong> N-Aryl Imines.<br />

Figure 2.3: General Imine Structures:<br />

N<br />

R<br />

N<br />

R<br />

N<br />

R<br />

R 1<br />

1<br />

2<br />

3<br />

40


Figure 2.4 General Catalyst structures:<br />

+<br />

R<br />

PH<br />

N<br />

O<br />

BARF -<br />

P<br />

Fe<br />

P<br />

Ir<br />

[{Ir(cod)Cl} 2 ]<br />

R<br />

Catalyst 2<br />

R= H<br />

Catalyst 1<br />

R<br />

OCH 3<br />

R<br />

R<br />

R<br />

O<br />

P<br />

O<br />

R<br />

O<br />

R<br />

O O<br />

P<br />

O<br />

O<br />

O O<br />

R<br />

R<br />

Ir<br />

Ph 2 P<br />

O PPh 2 O<br />

O<br />

O O<br />

Ph t P<br />

OCH<br />

2 BuSiO<br />

3<br />

O<br />

O<br />

OSi t BuPh O<br />

2<br />

P<br />

H 3 CO<br />

[Ir(COD) 2 ]BF 4<br />

R= t Bu<br />

Catalyst 3<br />

OCH 3<br />

Catalyst 4<br />

S<br />

R 2<br />

R 2<br />

P<br />

Ir<br />

N<br />

O<br />

R 1<br />

BARF<br />

*<br />

P<br />

=<br />

P<br />

R<br />

R<br />

P<br />

P<br />

DuPHOS<br />

R<br />

R<br />

RuCl 2<br />

NH 3<br />

* =<br />

NH 3<br />

H Ph<br />

2 N<br />

H 2 N Ph<br />

DPEN<br />

R 1 = i Pr, R 2 =Ph<br />

Catalyst 5<br />

*<br />

P<br />

P<br />

Cl<br />

Ru<br />

Cl<br />

H 2 N<br />

H 2 N<br />

*<br />

Catalyst 6<br />

Ph 2<br />

P<br />

Ir<br />

O<br />

CF 3 SO -<br />

+<br />

O<br />

BARF -<br />

N<br />

N Ir<br />

P<br />

Ph 2<br />

Ph<br />

N<br />

Ph<br />

P<br />

Ir<br />

N<br />

O<br />

H<br />

R<br />

+<br />

BARF -<br />

(S,R)-15 or (S,S)-15<br />

Catalyst 841<br />

R= i Pr<br />

Catalyst 9<br />

Catalyst 7


t Bu<br />

O<br />

S<br />

R 1 R 2<br />

N PPh 2<br />

[{Ir(cod)Cl 2 }]<br />

PAr 2<br />

O<br />

t Bu<br />

P<br />

O<br />

O<br />

t Bu<br />

Ir + Me<br />

P P BARF -<br />

t Bu<br />

Me<br />

R 1 =Ph,R 2 = i Bu<br />

Catalyst 10<br />

11a, Ar = xyl<br />

11b, Ar = ph, o-OMe-ph<br />

Catalyst 12<br />

Ph 2 P<br />

O<br />

Ir<br />

H<br />

H<br />

PPh 2<br />

O<br />

Catalyst 13<br />

+<br />

PF 6<br />

-<br />

Catalyst 11<br />

O<br />

R<br />

N<br />

P Ir<br />

Ar<br />

Ar<br />

R=Bn,Ar=3,5-DiMe-Ph<br />

+<br />

BARF -<br />

BARF<br />

Ir(COD)<br />

Ph 2 P<br />

OR<br />

N<br />

Fe<br />

R=CH 3<br />

Catalyst 15<br />

Catalyst 14<br />

N<br />

CONH<br />

CHO<br />

Cl 3 SiH<br />

Catalyst 16<br />

Me<br />

N<br />

H<br />

H<br />

N<br />

O<br />

O<br />

Cl 3 SiH<br />

Catalyst 17<br />

O<br />

*<br />

N N<br />

Cl 3 SiH<br />

Catalyst 18<br />

Ph<br />

H 3 C<br />

N<br />

H<br />

H<br />

H<br />

N<br />

O<br />

O Cl 3 SiH<br />

N<br />

Catalyst 19<br />

O<br />

Cl 3 SiH<br />

H<br />

N<br />

O<br />

AcO<br />

Catalyst 21<br />

Me<br />

Me<br />

O<br />

Ph<br />

Ph<br />

H<br />

C 6 F 13<br />

SO 2 (p- t BuPh)<br />

N<br />

N<br />

O<br />

H<br />

N<br />

Ph<br />

O<br />

Cl 3 SiH<br />

Catalyst 22<br />

O<br />

Me<br />

S<br />

N nN<br />

H H<br />

n=5<br />

Cl 3 SiH<br />

N<br />

H<br />

O<br />

S<br />

H<br />

N<br />

O<br />

O Cl 3 SiH<br />

N<br />

H<br />

Cl 3 SiH<br />

Catalyst 23<br />

O<br />

S<br />

Catalyst 20<br />

OH<br />

F<br />

O<br />

Me<br />

Me<br />

n<br />

n=3<br />

H H<br />

N N<br />

N<br />

n<br />

H O<br />

O n=2 O<br />

Cl 3 SiH<br />

O<br />

Catalyst 24<br />

N<br />

H<br />

PH 2<br />

Catalyst 25<br />

42


2.3.2. Different Substrates Categories<br />

Several organometallic and organocatalytic systems were developed <strong>for</strong> <strong>the</strong> reduction <strong>of</strong><br />

phenyl methyl (ethyl) N-aryl imines. The hydrogen source is ei<strong>the</strong>r molecular hydrogen or<br />

hydride reagents. Transition metals having chiral ligands on Ir, Ru, Rh and Ti were used and<br />

resulted in high yield and selectivity.<br />

Pfaltz and Leitner used cationic Ir complexes with chiral phosphinodihydrooxazoles modified<br />

with perfluroalkyl groups. [18] Using 0.09 mol % <strong>of</strong> <strong>the</strong> catalyst (catalyst 1, figure 2.4), 30 bar<br />

(435 psi) H 2 , supercritical carbon dioxide (scCO 2 ) at 40 °C an ee <strong>of</strong> 80% was accomplished<br />

with complete conversion (structure 1, figure 2.2). The choice <strong>of</strong> counter ion dramatically<br />

influenced selectivity with tetrakis-3,5-bis(trifluoromethyl)phenylborate anion (BARF),<br />

resulting in <strong>the</strong> highest selectivity. They later developed <strong>the</strong> use <strong>of</strong> scCO 2 with ionic liquids<br />

and obtained <strong>the</strong> same result. [19]<br />

Zhang and Xiao reported one <strong>of</strong> <strong>the</strong> earliest examples <strong>for</strong> <strong>the</strong> efficient reduction <strong>of</strong> aryl<br />

methyl N-aryl imines utilizing iridium. They introduced <strong>the</strong> use <strong>of</strong> air stable Irbisphospahn<strong>of</strong>errocene<br />

catalyst (catalyst 2, figure 2.4). [20] Using 2.0 mol % <strong>of</strong> <strong>the</strong> catalyst<br />

(catalyst 2, figure 2.4), 70 bar (1015 psi) <strong>of</strong> H 2 , CH 2 Cl 2 , 25 °C over 44 h, 99% ee with 77%<br />

conversion <strong>for</strong> phenyl methyl N-aryl imines (structure 1, figure 2.3). For p-OMe-phenyl<br />

methyl N-aryl imines <strong>the</strong> ee was 98% with 77% conversion and <strong>for</strong> p-CF 3 -phenyl methyl N-<br />

aryl imine <strong>the</strong> ee was 99% with 80% conversion (structure 3, figure 2.3).<br />

Claver and Castillón introduced <strong>the</strong> use <strong>of</strong> sugar derived diphosphite ligands. [21] Using 1.0<br />

mol % <strong>of</strong> <strong>the</strong> iridium catalyst (catalyst 3, figure 2.4), 10 bar (145 psi) H 2 , CH 2 Cl 2 , 25 °C, 18 h<br />

an ee was 57% with 83% conversion <strong>for</strong> phenyl methyl N-aryl imine (structure 1, figure 2.3).<br />

The use <strong>of</strong> 4.0 mol % Bu 4 NI improved conversion (100%) but lowered <strong>the</strong> ee (46%) at 70 bar<br />

(1015 psi) <strong>of</strong> H 2 . Later <strong>the</strong>y reported <strong>the</strong> use <strong>of</strong> o<strong>the</strong>r diphosphinite ligands (catalyst 4, figure<br />

2.4). Using 1.0 mol % <strong>of</strong> <strong>the</strong> catalyst, 70 bar ( 1015 psi) H 2 , CH 2 Cl 2 , 25 °C, 16 h, <strong>the</strong> ee was<br />

70% with complete conversion. [22]<br />

43


Cozzi et al. developed <strong>the</strong> use <strong>of</strong> phosphino oxazolines derived ligands. [23] Using 0.1 mol %<br />

<strong>of</strong> <strong>the</strong> catalyst (catalyst 5, figure 2.4), 50 bar (725 psi) H 2 , CH 2 Cl 2 , 25 °C, 4 h, <strong>the</strong> ee was<br />

86% with complete conversion <strong>for</strong> phenyl methyl N-aryl imine (structure 1, figure 2.3).<br />

Ru<strong>the</strong>nium catalysts earlier developed by Noyori <strong>for</strong> ketone reduction were useful <strong>for</strong> imine<br />

reduction which was tested by Cobley. [24] Using 1.0 mol % <strong>of</strong> RuCl 2 (diphosphine) (diamine)<br />

(catalyst 6, figure 2.4), 15 bar (218 psi) H 2 , 100 mol % <strong>of</strong> t-BuOK in t-BuOH <strong>for</strong> in situ<br />

activation <strong>of</strong> <strong>the</strong> catalyst, 65 °C, 20 h, <strong>the</strong> ee was 91% with complete conversion <strong>for</strong> phenyl<br />

methyl N-aryl imine (structure 1, figure 2.3).<br />

Grützmacher was successful in using mixed phosphane olefin ligand <strong>for</strong> imine reduction. [25]<br />

Using 1.0 mol % <strong>of</strong> <strong>the</strong> iridium catalyst (catalyst 7, figure 2.4), 50 bar (725 psi) H 2 , CHCl 3 ,<br />

50 °C, 2 h an ee <strong>of</strong> 86% with >98% yield <strong>for</strong> phenyl methyl N-aryl imine was reported<br />

(structure 1, figure 2.3).<br />

Niedercorn was able to reduce N-aryl imines with Ir-aminophosphine-oxazoline derived<br />

catalyst. [26] Using 2.0 mol % <strong>of</strong> <strong>the</strong> catalyst (catalyst 8, figure 2.4), 20-50 bar (290-725 psi)<br />

H 2 , CH 2 Cl 2 , 12 h an ee was 90% with full conversion <strong>for</strong> phenyl methyl N-aryl imine was<br />

reported (structure 1, figure 2.3).<br />

Andersson developed a new class <strong>of</strong> chiral phosphine-oxazoline ligands <strong>for</strong> iridium imine<br />

reduction. [27] Using 0.5 mol % <strong>of</strong> <strong>the</strong> catalyst (catalyst 9, figure 2.4), 20 bar (290 psi) <strong>of</strong> H 2 ,<br />

CH 2 Cl 2 , 25 °C over 2 h an ee was 90% with 98% conversion <strong>for</strong> phenyl, methyl N-aryl<br />

imines (structure 1, figure 2.3). He also tested his catalyst <strong>for</strong> reducing p-fluoro phenyl<br />

methyl N-aryl imine and reported 89% ee in 2 h, <strong>for</strong> p-OMe phenyl methyl N-aryl imine <strong>the</strong><br />

ee was 86% within 2-3 h, <strong>for</strong> p-chloro phenyl methyl N-aryl imine <strong>the</strong> ee was 89% within 1.5<br />

h with full conversion. In case <strong>of</strong> o-Me phenyl methyl N-aryl imine <strong>the</strong> ee was lower (83%)<br />

and <strong>the</strong> conversion was much lower (52%) after even 12 h (structure 3, Figure 2.3). Later<br />

<strong>the</strong>y reported 78% ee <strong>for</strong> phenyl ethyl N-aryl imines (structure 2, figure 2.3). 2-naphthyl<br />

methyl N-aryl imine was reduced with 91% ee. [28]<br />

Blom prepared a new class <strong>of</strong> diphenylphosphanyl sulfoximines ligands. [29] Using 1.1 mol %<br />

<strong>of</strong> <strong>the</strong> Ir-Sulxoimine catalyst (catalyst 10, figure 2.4), 2.0 mol % <strong>of</strong> iodine, 20 bar (290 psi) <strong>of</strong><br />

44


H 2 , toluene, 25 °C, 4 h an ee <strong>of</strong> 96% with full conversion was reported <strong>for</strong> phenyl methyl N-<br />

aryl imine and 92% ee <strong>for</strong> phenyl ethyl N-aryl imnes (structure 1,2 , figure 2.3). Reducing <strong>the</strong><br />

catalyst loading to 0.5 mol % resulted in <strong>the</strong> same enantioselectivity. Lower catalyst loading<br />

(0.1 mol %) <strong>the</strong> hydrogen pressure had to increase to 50 bar to achieve <strong>the</strong> same ee with full<br />

conversion. He tested his system <strong>for</strong> substituted phenyl methyl N-aryl imines. The ee was<br />

96% <strong>for</strong> p-Me phenyl methyl N-aryl imine, <strong>for</strong> m-Me phenyl methyl N-aryl imine <strong>the</strong> ee was<br />

93% and <strong>for</strong> o-Me phenyl methyl N-aryl imine <strong>the</strong> ee was 94%. For o-OMe phenyl methyl N-<br />

aryl imine <strong>the</strong> ee was 90%, <strong>for</strong> p-OMe phenyl methyl N-aryl imine <strong>the</strong> ee was 94% and <strong>for</strong><br />

m-OMe phenyl methyl N-aryl imine <strong>the</strong> ee was 96% with full conversion in all cases<br />

(structure 3, figure 2.3).<br />

Pizzano tested Ir-phosphine–phosphites based catalysts <strong>for</strong> reduction <strong>of</strong> N-aryl imines. [30]<br />

Using 1.0 mol % <strong>of</strong> <strong>the</strong> catalyst (catalyst 11a, figure 2.4), 30 bar (436 psi) <strong>of</strong> H 2 , CH 2 Cl 2 , 25<br />

°C, 24 h an ee <strong>of</strong> 82% with complete conversion <strong>for</strong> phenyl methyl N-aryl imine was<br />

achieved (structure 1, figure 2.3). Later he used ano<strong>the</strong>r derivative <strong>of</strong> <strong>the</strong> catalyst (catalyst<br />

11b, figure 2.3) <strong>for</strong> reduction <strong>of</strong> substituted phenyl methyl N-aryl imine. For p-methyl phenyl<br />

methyl N-aryl imine <strong>the</strong> ee was 72%, <strong>for</strong> p-OMe phenyl methyl N-aryl imine <strong>the</strong> ee was 85%,<br />

<strong>for</strong> p-fluoro phenyl methyl N-aryl imine <strong>the</strong> ee was 79% and <strong>for</strong> p-chloro phenyl methyl N-<br />

aryl imine <strong>the</strong> ee was 82%. [31]<br />

Imamoto reduced N-aryl imines utilizing Ir-phosphine catalyst. [32] Using 0.5 mol % <strong>of</strong> <strong>the</strong><br />

catalyst (catalyst 12, figure 2.4), 1 bar (14.5 psi) <strong>of</strong> H 2 , CH 2 Cl 2 , 25 °C, 1.5 h, <strong>the</strong> ee was 99%<br />

with 95% isolated yield <strong>for</strong> phenyl methyl N-aryl imine (structure 1, figure 2.3). They tested<br />

reduction <strong>of</strong> p-OMe phenyl methyl N-aryl imine imines with 83% ee and 98% yield within 2<br />

h. For p-fluoro phenyl methyl N-aryl imine <strong>the</strong> ee was 84% with 92% yield within 1.5 h<br />

(structure 3, figure 2.3).<br />

Dervisi syn<strong>the</strong>sized a new Ir diphosphine catalyst {[(Ir (ddppm)-(COD)]PF 6 } and tested it <strong>for</strong><br />

<strong>the</strong> N-aryl imine reduction. [32] Using 1.0 mol % <strong>of</strong> <strong>the</strong> catalyst (catalyst 13, figure 2.4), 1 bar<br />

(14.5 psi) <strong>of</strong> H 2 , ClCH 2 CH 2 Cl, 25 °C, 24 h, <strong>the</strong> ee was 84% with 99% yield <strong>for</strong> phenyl<br />

methyl N-aryl imine (structure 1, figure 2.3). Operating at atmospheric pressure allowed <strong>the</strong><br />

hydrogenation to be carried using Schlenk technique instead <strong>of</strong> high pressure autoclaves.<br />

They expanded <strong>the</strong>ir investigation to include p-chloro phenyl methyl N-aryl imines which<br />

45


were reduced with 80% ee with 99% yield. For p-OMe phenyl methyl N-aryl imines <strong>the</strong> ee<br />

was 81% with 100% yield (structure 3, figure 2.3).<br />

New chiral phosphine oxazoline ligands was prepared by Zhou <strong>for</strong> <strong>the</strong> Ir reduction <strong>of</strong> imines<br />

(Ir-SIPHOX) (catalyst 14, figure 2.4). [33] Using 1.0 mol % <strong>of</strong> <strong>the</strong> catalyst, 1 bar (14.5) <strong>of</strong> H 2 ,<br />

t-butyl methyl e<strong>the</strong>r (TBME), 4 Å MS, 10 °C over 20 h, <strong>the</strong> ee was 93% with complete<br />

conversion <strong>for</strong> phenyl methyl N-aryl imine (structure 1, figure 2.3). He investigated <strong>the</strong><br />

application <strong>of</strong> his catalyst on <strong>the</strong> reduction <strong>of</strong> different substituted phenyl methyl N-aryl<br />

imine, <strong>for</strong> p-Me <strong>the</strong> (94% ee), p-Cl (90% ee), p-Br (91% ee), m-Cl (93% ee), m-Br (92% ee)<br />

phenyl methyl N-aryl imine derivatives with full conversion in all cases. For 3,4- Di-Me<br />

phenyl methyl N-aryl imine <strong>the</strong> ee was 94% (structure 3, figure 2.3).<br />

New ferrocenyl P,N-ligands was introduced by Knochel and used iridium <strong>for</strong> <strong>the</strong> reduction <strong>of</strong><br />

N-aryl imines. [34] Using 1.0 mol % <strong>of</strong> <strong>the</strong> catalyst (catalyst 15, figure 2.4), 10 bar (145 psi) <strong>of</strong><br />

H 2 , toluene/ MeOH (4:1) at 25 °C over 2 h, <strong>the</strong> ee was 94% with full conversion <strong>for</strong> phenyl<br />

methyl N-aryl imine(structure 1, figure 2.3). He also tested his catalyst <strong>for</strong> <strong>the</strong> reduction <strong>of</strong><br />

substituted phenyl methyl N-aryl imine. p-Ph and p-Cl (92% ee), m-Me (93% ee), o-Me (94%<br />

ee), m-F (93% ee) and p-CF 3 (89% ee) phenyl methyl N-aryl imine were reduced with high<br />

ees (structure 3, figure 2.3). 2-naphthyl methyl N-aryl imine was reduced with 93% ee.<br />

For <strong>the</strong> reduction <strong>of</strong> this category <strong>of</strong> chiral imines organocatalytic methods have proved to be<br />

highly effective. Different organocatalysts have been developed utilizing various silane<br />

derivatives or Hantzsch ester as a source <strong>of</strong> hydride. Although <strong>the</strong>se sources <strong>of</strong> hydrides are<br />

not atom economic, <strong>the</strong>y are commercially available in large quantities at ra<strong>the</strong>r reasonable<br />

prices and <strong>of</strong>fer <strong>the</strong> potential <strong>of</strong> chemoselectivity not possible in <strong>the</strong> presence <strong>of</strong> H 2 .<br />

In 2001 Matsumura and coworkers developed <strong>the</strong> use <strong>of</strong> proline (catalyst 16, figure 2.4)<br />

derivatives <strong>for</strong> <strong>the</strong> hydrosilylation <strong>of</strong> imines. They achieved mediocre enantioselectivity<br />

reporting 66% ee with 52% yield <strong>for</strong> phenyl methyl N-aryl imine (structure 1, figure 2.3). [35]<br />

Inspired by <strong>the</strong> research <strong>of</strong> Matsumura, Kočovský and coworkers developed <strong>the</strong> use <strong>of</strong> valine<br />

derived (<strong>for</strong>mamide/amide based) catalysts <strong>for</strong> <strong>the</strong> hydrosilylation <strong>of</strong> N-aryl imine. [36] Using<br />

10 mol % <strong>of</strong> <strong>the</strong> catalyst (catalyst 17, figure 2.4), 1.5 equiv <strong>of</strong> Cl 3 SiH, CHCl 3 , -20 °C, 16 h,<br />

<strong>the</strong> ee was 92% with 94% isolated yield <strong>for</strong> phenyl methyl N-aryl imine (Structure 1, Figure<br />

46


1). They described <strong>the</strong> role <strong>of</strong> each functional group in <strong>the</strong> catalyst and its importance in<br />

controlling stereoselectivity. They also marked <strong>the</strong> important structural features in <strong>the</strong> imine<br />

which controls <strong>the</strong> total outcome <strong>of</strong> <strong>the</strong> reaction. They tested <strong>the</strong>ir system also <strong>for</strong> <strong>the</strong><br />

reduction <strong>of</strong> substituted phenyl methyl N-aryl imines. p-OMe (85% ee, 86% yield), p-CF 3<br />

(89% ee, 86% yield), o-Me (92% ee, 90% yield) phenyl, methyl N-aryl imines were<br />

successfully reduced (structure 3, figure 2.3).<br />

In 2007 he reported <strong>the</strong> use <strong>of</strong> his catalyst with fluorous tag <strong>for</strong> imine reduction. [37] Fluorous<br />

tags are used to enable recycling <strong>of</strong> <strong>the</strong> catalyst. He was able to reuse <strong>the</strong> catalyst 4-5 times<br />

without significant loss <strong>of</strong> enantioselectivity and yield. Using 10 mol % <strong>of</strong> valine derived<br />

catalyst (catalyst 19, figure 2.4), 2.0 equiv <strong>of</strong> HSiCl 3 , toluene, 18 °C, 16 h, <strong>the</strong> ee was 90%<br />

with 98% yield <strong>for</strong> phenyl methyl N-aryl imine (structure 1, figure 2.3). p-CF 3 (92% ee, 72%<br />

yield, 10 °C) and p-OMe (84% ee, 84% yield) phenyl methyl N-aryl imines were reduced<br />

with high yields and ees (structure 3, figure 2.3). 2-naphthyl methyl N-aryl imine was<br />

reduced with 92% ee and 93% yield.<br />

In 2008 he reported <strong>the</strong> use <strong>of</strong> his catalyst with polymer support, which can be used up to 5<br />

times without significant loss <strong>of</strong> catalyst activity. [38] Using 15 mol % <strong>of</strong> <strong>the</strong> catalyst (catalyst<br />

20, figure 2.4), 2.0 equiv Cl 3 SiH, CH 3 Cl, 25 °C, 16 h, <strong>the</strong> ee was 82% with 84% yield <strong>for</strong><br />

phenyl methyl N-aryl imine (structure 1, figure 2.3). He expanded his study to cover o<strong>the</strong>r<br />

substituted phenyl methyl N-aryl imines. p-OMe (77% ee, 63% yield) and p-CF 3 (81% ee,<br />

67% yield) phenyl methyl N-aryl imines were reduced. Also 2,5 Me-3-furyl phenyl, methyl<br />

N-aryl imine was reduced with 78% ee and 67% yield (structure 3, figure 2.3).<br />

In 2006 he introduced <strong>the</strong> use <strong>of</strong> oxazoline catalyst <strong>for</strong> reduction <strong>of</strong> N-aryl imines. [39] Using<br />

20 mol % <strong>of</strong> <strong>the</strong> catalyst (catalyst 18, figure 2.4), 2.0 equiv <strong>of</strong> HSiCl 3 , CHCl 3 , –20 °C, 24 h,<br />

resulted in 87% ee and 65% yield phenyl methyl N-aryl imine (structure 1, figure 2.3). p-<br />

OMe (87% ee, 51% yield) and p-CF 3 (87% ee, 65% yield) phenyl methyl N-aryl imines were<br />

reduced successfully (structure 3, figure 2.3).<br />

The group <strong>of</strong> Sun also prepared several organocatalysts <strong>for</strong> <strong>the</strong> hydrosilylation <strong>of</strong> imines. In<br />

2006, <strong>the</strong>y tested pipecolinic acid derived <strong>for</strong>mamides catalyst. [40] Using 10 mol % <strong>of</strong> <strong>the</strong><br />

catalyst (catalyst 21, figure 2.4), 2.0 equiv <strong>of</strong> Cl 3 SiH, CH 2 Cl 2 , 0 °C, 16 h, <strong>the</strong> ee was 95%<br />

47


with 97% yield <strong>for</strong> phenyl methyl N-aryl imine (structure 1, figure 2.3). They also tested <strong>the</strong>ir<br />

catalyst <strong>for</strong> different substituted phenyl methyl N-aryl imines. p-OMe (93% ee, 95% yield),<br />

p-Br (95% ee, 98% yield), m-Br (94% ee, 82% yield), p-CF 3 (96% ee, 85% yield), para-NO 2<br />

(95% ee, 96% yield) phenyl methyl N-aryl imines were reduced successfully (structure 3,<br />

figure 2.3). 2-naphthyl methyl N-aryl was reduced with (93% ee, 92% yield) and p-methoxy<br />

substituted naphthyl methyl N-aryl imine (91% ee, 97% yield).<br />

They also reported <strong>the</strong> use <strong>of</strong> <strong>for</strong>mamide derivative <strong>of</strong> piperazine carboxylic acid. [41] Using 10<br />

mol % <strong>of</strong> <strong>the</strong> catalyst (catalyst 22, figure 2.4), 2.0 equiv Cl 3 SiH, CH 2 Cl 2 , -20 °C, 48 h,<br />

resulted in 89% ee with 95% yield <strong>for</strong> phenyl methyl N-aryl imine (structure 1, figure 2.3)<br />

and <strong>for</strong> phenyl ethyl N-aryl imine <strong>the</strong> ee was 94% with 92% yield (structure 2, figure 2.3). p-<br />

NO 2 (90% ee, 99% yield), p-Br (89% ee, 81% yield) and p-Me (85% ee, 71% yield) phenyl<br />

methyl N-aryl were successfully reduced. 2-naphthyl (88% ee, 63% yield) and 6-OMe 2-<br />

naphthyl methyl N-aryl imine (85% ee, 64% yield) were reduced. p-F (95% ee, 87% yield),<br />

p-Cl (94% ee, 83% yield), p-Br (95% ee, 89% yield), p-Me (88% ee, 87% yield) and p-OMe<br />

(90% ee, 83% yield) phenyl ethyl N-aryl imine were reduced with high yields and ees.<br />

Phenyl n-propyl N-aryl imine was reduced with 90% ee and 88% yield and phenyl n-butyl N-<br />

aryl imine was reduced with 89% ee and 84% yield.<br />

They were also successful in using chiral sulfinamides based on Ellman auxiliary [(R)-tertbutansulfinamide.<br />

[42] Using <strong>of</strong> 20 mol % <strong>of</strong> <strong>the</strong> catalyst (catalyst 23, figure 2.4), 2.0 equiv<br />

Cl 3 SiH, CH 2 Cl 2 , -20 °C, 24 h, <strong>the</strong> ee was 92% with 92% yield <strong>for</strong> phenyl methyl N-aryl<br />

imine (structure 1, figure 2.3). p-OMe (93% ee, 98% yield), p-Br (92% ee, 92% yield), p-NO 2<br />

(90% ee, 94% yield) and p-CF 3 (92% ee, 93% yield) phenyl methyl N-aryl imines were<br />

reduced (structure 3, figure 2.3).<br />

In 2007 <strong>the</strong>y reported <strong>the</strong> use <strong>of</strong> proline derived tetramide catalyst <strong>for</strong> imine reduction. [43]<br />

Using 10 mol % <strong>of</strong> <strong>the</strong> catalyst (catalyst 24, figure 2.4), 2.0 equiv Cl 3 SiH, CH 2 Cl 2 , 0 °C, 16<br />

h, <strong>the</strong> ee was 77% with 93% yield <strong>for</strong> phenyl methyl N-aryl imines. For substituted phenyl<br />

methyl imines <strong>the</strong> ees were lower compared with his o<strong>the</strong>r catalytic systems (structure 1,<br />

figure 2.3)<br />

48


More recently, 2008, he described <strong>the</strong> use <strong>of</strong> S-<strong>Chiral</strong> Bissulfinamide catalyst <strong>for</strong> imine<br />

reduction. [44] Using 10 mol % <strong>of</strong> <strong>the</strong> catalyst (catalyst 25, figure 2.4), 2.0 equiv Cl 3 SiH, 0.3<br />

equiv 2,6-lutidine, CH 2 Cl 2 , -20 °C, 24 h, <strong>the</strong> ee was 96% with 91% yield <strong>for</strong> phenyl methyl<br />

N-aryl imine. (structure 1, figure 2.3). p-OMe (95% ee, 83% yield), p-Br (95% ee, 92%<br />

yield), p-NO 2 (93% ee, 90% yield) and p-CF 3 (95% ee, 95% yield) phenyl methyl N-aryl<br />

imines were reduced (structure 3, figure 2.3).<br />

2.4. Reduction <strong>of</strong> Miscellaneous Imines:<br />

O<strong>the</strong>r imine derivatives were reduced as N-benzyl and N-tosyl imines. These substrates were<br />

less investigated compared to <strong>the</strong> previous discussed substrates in <strong>the</strong> last 10 years. For<br />

fur<strong>the</strong>r reading please consult <strong>the</strong> following literatures. [45] Oximes and hydrazones were also<br />

tested but with fewer examples over <strong>the</strong> last 10 years. For fur<strong>the</strong>r reading please consult <strong>the</strong><br />

following literatures. [46] Imines with different chiral auxiliaries were reduced in good to high<br />

enantioselectivity. For fur<strong>the</strong>r reading please consult <strong>the</strong> following references. [47] Cyclic<br />

imines were extensively investigated and extensively reviewed in book chapters and<br />

published reviews. For fur<strong>the</strong>r reading please consult <strong>the</strong> following references. [48]<br />

2.5 Conclusion.<br />

Imine reduction has been extensively investigated by many groups over <strong>the</strong> last <strong>for</strong>ty years.<br />

During <strong>the</strong> last two decades several milestones have been achieved in asymmetric imine<br />

reduction. Imines are reduced with high enantioselectivity and yield. This methodology is<br />

highly efficient <strong>for</strong> obtaining α-chiral amines in 99% enantioselectivity but <strong>the</strong> over all yield<br />

from <strong>the</strong> starting material to <strong>the</strong> final product is usually low and below 50%. Removal <strong>of</strong> <strong>the</strong><br />

auxiliary usually requires harsh acidic conditions which may not compatible with different<br />

acid sensitive groups. Despite <strong>the</strong>se drawbacks <strong>the</strong> high enantioselectivity obtained makes<br />

this method attractive <strong>for</strong> fur<strong>the</strong>r improvements.<br />

49


2.6. References.<br />

[1] a) P. G. Andersson, I. J. Munslow in Modern Reduction Methods, Wiley-VCH,<br />

Weinheim, 2008; b) J. G. De Vries, C. J. Elsevier in The Handbook <strong>of</strong> Homogeneous<br />

Hydrogenation, Wiley-VCH, Weinheim, 2008.<br />

[2] S. Kobayashi, H. Ishitani, Chem. Rev. 1999, 99, 1069.<br />

[3] J. A. Ellman, T. D. Owens, T. P. Tang Acc. Chem. Res. 2002, 35, 984.<br />

[4] E. N. Jacobsen, A. Pfaltz, H. Yamamoto, Comprehensive Asymmetric Catalysis I–III,<br />

Springer-Verlag Berlin Heidelberg, 2000.<br />

[5] F. Spindler, H. -U. Blaser, Adv. Synth. Catal. 2001, 343, 68.<br />

[6] B. Krzyzanowska, W. J. Stec, Syn<strong>the</strong>is 1982, 270.<br />

[7] a) T. Yamada, T. Nagata, K. D. Sugi, K. Yorozu, T. Ikeno, T. Ohtsuka, D. Miyazaki, T.<br />

Mukaiyama, Chem. Eur. J. 2003, 9, 4485.<br />

[8] F. Spindler, H. -U. Blaser, Adv. Synth. Catal. 2001, 343, 68.<br />

[9] Yamada, T. Nagata, K. D. Sugi, K. Yorozu, T. Ikeno, Y. Ohtsuka, D. Miyazaki, T.<br />

Mukaiyama, Chem. Eur. J. 2003, 9, 4485.<br />

[10] H. Lipshutz, H. Shimizu, Angew. Chem. Int. Ed. 2004, 43, 2228.<br />

[11] a) Transfer hydrogenation process. Martin, J.; Campbell; L. A.; Avecia Limited, Great<br />

Britain. Patent publication number WO0112574, 2001; b) Transfer hydrogenation process.<br />

Martin, J.; Campbell, L. A.; Avecia Limited, Great Britain. Patent publication number<br />

US6696608, 2004.<br />

[12] K. A. Nolin, R. W. Ahn, F. D. Toste, J. Am. Chem. Soc. 2005, 127, 12462.<br />

[13] B. M. Park, S. Mun, J. Yun, Adv. Synth. Catal. 2006, 348, 1029.<br />

[14] a)Y. Q. wang, Y. G. Zhou, Synlett 2006, 8, 1189; b) Y. Q. Wang, S. M. Lu, Y. G. Zhou,<br />

J. Org. Chem. 2007, 72, 3729.<br />

[15] M. Bandini, M. M. Melucci, F. Piccinelli, R. Sinisi, S. Tommasi, A. U. Ronchi, Chem.<br />

Commun. 2007, 4519.<br />

[16] C. R. Graves, K. A. Scheidt, S. T. Nguyen, Org. Lett. 2006, 8, 1229.<br />

[17] Z. Wang, X. Ye, S. Wei, P. Wu, A. Zhang, J. Sun, Org. Lett. 2006, 8, 999.<br />

[18] S. Kainz, A. Brinkmann, W. Leitner, A. Pfaltz, J. Am. Chem. Soc. 1999, 121, 6421.<br />

[19] M. Solinas, A. Pfaltz, P. G. Cozzi, W. Leitner, J. Am. Chem. Soc. 2004, 126, 16142.<br />

[20] D. Xiao, X. Zhang, Angew. Chem. Int. Ed. 2001, 40, 3425.<br />

[21] E. Guiu, B. Munoz, S. Castillón, C. Claver, Adv. Synth. Catal. 2003, 345, 169.<br />

50


[22] E. Guiu, M. Aghmiz, Y. Diaz, C. Claver, B. Meseguer, C. militzer, S. Castillón. Eur. J.<br />

Org. Chem. 2006, 627.<br />

[23] P. G. Cozzi, F. Menges, S. Kaiser, Synlett 2003, 6, 833.<br />

[24] C. J. Cobley, J. P. Henschke, Adv. Synth. Catal. 2003, 345, 195.<br />

[25] P. Maire, S. Deblon, F. Breher, J. Geier, C. Böhler, H. Rüegger, H. Schönberg, H.<br />

Grützmacher, Chem. Eur. J. 2004, 10, 4198.<br />

[26] C. Blanc, F. A. Niedercorn, G. Nowogrocki, Tetrahedron: Asymmetry 2004, 15, 2159.<br />

[27] A. Trifonova, J. S. Diesen, C. J. Chapman, P. G. Andersson, Org. Lett. 2004, 6, 3825.<br />

[28] A. Trifonova, J. S. Diesen, P. G. Andresson, Chem. Eur. J. 2006, 12, 2318.<br />

[29] C. Moessner, C. Bolm, Angew. Chem. Int. Ed. 2005, 44, 7564.<br />

[30] S. Vargas, M. Rubio, A. Suárez, A. Pizzano, Tetrahedron Lett. 2005, 46, 2049.<br />

[31] S. Vargas, M. Rubio, A. Suárez, D. del Río, E. Álvarez, A. Pizzano, Organometallics<br />

2006, 25, 961.<br />

[32] T. Imamoto, N. Iwadate, K. Yoshida, Org. lett. 2006, 8, 2289.<br />

[32] A. Dervisi, C. Carcedo, L. L. Ooi, Adv. Synth. Catal. 2006, 348, 175.<br />

[33] S. F. Zhu, J. B. Xie, Y. Z. Zhang, S. Li, Q. L. Zhou, J. Am. Chem. Soc. 2006, 128,<br />

12886.<br />

[34] M. N. Cheemala, P. Knochel, Org. Lett. 2007, 9, 3089.<br />

[35] F. Iwasaki, O. Onomura, K. Mishima, T. Kanematsu, T. Makib, Y. Matsumura,<br />

Tetrahedron Lett. 2001, 42, 2525.<br />

[36] a) A. V. Malkov, A. Mariani, K. N. MacDougall, P. Kočovský, Org. Lett. 2004, 6, 2253;<br />

b) A. V. Malkov, S. Stoncius, K. N. MacDougall, A. Mariani, G. D. McGeoch, P. Kočovský,<br />

Tetrahedron 2006, 62, 264.<br />

[37] A. V. Malkov, M. Figlus, S. Stoncius, P. Kočovský, J. Org. Chem. 2007, 72, 1315.<br />

[38] A. V. Malkov, M. Figlus, P. Kočovský, J. Org. Chem. 2008, 73, 3985.<br />

[39] A. V. Malkov, A. J. P. S. Liddon, P. R. Lopez, L. Bendova, D. Haigh, P. Kočovský,<br />

Angew. Chem. Int. Ed. 2006, 45, 1432.<br />

[40] L. Zhou, Z. Wang, S. Wei, J. Sun, Chem. Commun., 2007, 2977.<br />

[41] Z. Wang, M. Cheng, P. Wu, S. Wei, J. Sun, Org. Lett. 2006, 8, 3045.<br />

[42] D. Pei, Z. Wang, S. Wei, Y. Zhang, J. Sun, Org. Lett. 2006, 8, 5913.<br />

[43] Z. Wang, S. Wei, C. Wang. J. Sun, Tetrahedron: Asymmetry 2007, 18, 705.<br />

[44] D. Pei, Y. Zhang, S. Wei, M. Wang, J. Sun, Adv. Synth. Catal. 2008, 350, 619.<br />

51


[45] a) A. Ros, A. Magriz, H. Dietrich, Mark Ford, R. Fernández, J. M. Lassalettaa, Adv.<br />

Synth. Catal. 2005, 347, 1917; b) J. B. Åberg, J. S. M. Samec, J.E. Bäckvall, Chem.<br />

Commun. 2006, 2771; c) M. T. Reetz, O. Bondarev, Angew. Chem. Int. Ed. 2007, 46, 4523;<br />

d) Q. Yang, G. Shang, W. Gao, J. Deng, and X. Zhang, Angew. Chem. Int. Ed. 2006, 45,<br />

3832.<br />

[46] a) I. Takei, Y. Nishibayashi, Y. Ishii, Y. Mizobe, S. Uemura, M. Hidai, Chem Commun.<br />

2001, 2360; b) E. Fontaine, C. Namane, J. Meneyrol, M. Geslin, L. Serva, E. Roussey, S.<br />

Tissandié, M. Maftouh, P. Roger, Tetrahedron: Asymmetry 2001, 12, 2185; c) M. P.<br />

Krzemiński, M. Zaidlewicz, Tetrahedron: Asymmetry 2003, 14, 1463; d) X. Huang, M. O.-<br />

Marciales, K. Huang, V. Stepanenko, F. G. Merced, A. M. Ayala, W. Correa, M. De Jesús,<br />

Org. Lett. 2007, 9, 1793; e) M. Sugiura, S. Kobayashi, Angew. Chem. Int. Ed. 2005, 44, 5176.<br />

[47] a) G. Chelucci, S. Baldino, R. Solinasa, W. Barattab, Tetrahedron Lett. 2005, 46, 5555;<br />

b) X. Xiao, H. Wang, Z. Huang, J. Yang, X. Bian, Y. Qin, Org. lett. 2006, 8, 139; c) C.<br />

Cimarelli, G. Palmieri, Tetrahedron: Asymmetry 2000, 11 2555; d) B. TöröK, G. K. S.<br />

Prakash, Adv. Synth. Catal. 2003, 345, 165.<br />

[48] a) M. Rueping, A. P. Antonchick, T. Theissmann, Angew. Chem. Int. Ed. 2006, 45,<br />

6751; b) M. Rueping, A. P. Antonchick, T. Theissmann, Angew. Chem. Int. Ed. 2006, 45,<br />

3683; c)M. Rueping, A. P. Antonchick, Angew. Chem. Int. Ed. 2007, 46, 4562; d) P.<br />

Roszkowski, Z. Czarnockia, Mini-Reviews inorganic Chemistry 2007, 4, 190; e) Y.-Q. Wang,<br />

C. –B. Yu, D.-W. Wang, X. –B. Wang, Y. –G. Zhou, Org. Lett. 2008, 10, 2071.f) J. Li, Y.<br />

Zhang, D. Han, Q. Gao, C. Li, Journal <strong>of</strong> Molecular Catalysis A: Chemical 2009 298, 31.<br />

52


Chapter 3<br />

Reductive Amination<br />

3.1. Historical View:<br />

3.1.1. Reductive Amination Utilizing Heterogeneous Catalyst:<br />

Reductive amination can be defined as <strong>the</strong> reaction <strong>of</strong> aldehyde or ketone with ammonia or<br />

with a primary or a secondary amine to give alkylated amines in <strong>the</strong> presence <strong>of</strong> a catalyst<br />

under hydrogenation conditions. Reductive amination was first described in <strong>the</strong> early days <strong>of</strong><br />

twentieth century by Mignonac. [1] Since <strong>the</strong>n it was widely applied <strong>for</strong> <strong>the</strong> preparation <strong>of</strong><br />

different types <strong>of</strong> amines. The process could be described as reductive alkylation <strong>of</strong> ammonia<br />

or reductive amination <strong>of</strong> aldehydes or ketones. [2]<br />

Reductive amination is a powerful methodology <strong>for</strong> <strong>the</strong> syn<strong>the</strong>sis <strong>of</strong> primary, secondary or<br />

tertiary amines. Its main advantage is being a stepwise efficient methodology. It is only two<br />

steps from <strong>the</strong> carbonyl compound to <strong>the</strong> primary amine. Due to <strong>the</strong> significance <strong>of</strong> this<br />

trans<strong>for</strong>mation it was subjected to an intensive investigation by different research groups.<br />

Several obstacles appeared on <strong>the</strong> surface during developing this methodology. Formation <strong>of</strong><br />

alcohol as a main by product resulting in lower yield was one <strong>of</strong> <strong>the</strong> major problem. Over<br />

alkylation <strong>of</strong> <strong>the</strong> amine was an additional problem. Also side reactions, as aldol condensation<br />

which leads to <strong>the</strong> <strong>for</strong>mation <strong>of</strong> side products and lower <strong>the</strong> overall yield was ano<strong>the</strong>r major<br />

concern. Despite <strong>the</strong> fact that most <strong>of</strong> <strong>the</strong>se problems were solved over <strong>the</strong> last three decades,<br />

I will try to describe briefly <strong>the</strong> historical development <strong>of</strong> this methodology and <strong>the</strong> problems<br />

associated with its development.<br />

The introduction <strong>of</strong> catalysis in organic chemistry was <strong>the</strong> most important breackthrough in<br />

<strong>the</strong> field in <strong>the</strong> twentieth century. The use <strong>of</strong> minute quantity <strong>of</strong> a catalyst to accelerate <strong>the</strong><br />

required reaction, accumulate <strong>the</strong> product and inhibit side reactions were <strong>the</strong> main objectives<br />

<strong>of</strong> <strong>the</strong> catalyst use. Historically heterogeneous catalysts were first introduced in <strong>the</strong> main<br />

53


stream <strong>of</strong> organic chemistry. Hydrogenation catalysts as nickel, palladium platinum and o<strong>the</strong>r<br />

transition metal catalysts received <strong>the</strong> maximum attention. These catalysts were initially<br />

introduced <strong>for</strong> alkene hydrogenation and <strong>the</strong>y were later tested <strong>for</strong> o<strong>the</strong>r trans<strong>for</strong>mations as<br />

reductive amination. [3]<br />

Reductive alkylation <strong>of</strong> ammonia was one <strong>of</strong> <strong>the</strong> earliest examples described in literatures. It<br />

generally proceeds under mild conditions using heterogeneous catalyst. The reductive<br />

alkylation <strong>of</strong> ammonia with carbonyl compounds may produce primary, secondary, and<br />

tertiary amines, as well as alcohol as a side product. The origin <strong>of</strong> product selectivity depends<br />

primarily on <strong>the</strong> molar ratio <strong>of</strong> carbonyl compound to ammonia, <strong>the</strong> nature <strong>of</strong> catalyst and<br />

structure <strong>of</strong> <strong>the</strong> carbonyl compound. The reaction <strong>of</strong> benzaldehyde in <strong>the</strong> presence <strong>of</strong> 1.0<br />

equivalent <strong>of</strong> ammonia in ethanol over Raney Ni gave benzylamine in an 89.4% yield while<br />

with 0.5 equivalent <strong>of</strong> ammonia dibenzylamine was obtained in an 80.8% yield. [4]<br />

Reductive amination <strong>of</strong> aliphatic aldehydes having α-hydrogen atoms, especially <strong>of</strong> <strong>the</strong> type<br />

RCH 2 CHO, usually results in lower yields due to <strong>the</strong> <strong>for</strong>mation <strong>of</strong> by products through aldol<br />

or o<strong>the</strong>r condensation reactions. Also lower aliphatic aldehydes usually produce mixture <strong>of</strong><br />

primary, secondary, and tertiary amines. The reaction <strong>of</strong> butyraldehyde with 0.5 equivalent <strong>of</strong><br />

ammonia over Raney Ni also resulted in a mixture <strong>of</strong> 31% <strong>of</strong> butylamine, 17% <strong>of</strong><br />

dibutylamine, and 8% tributylamine. [5] Higher aldehydes usually react selectively with<br />

ammonia producing less by products. [6]<br />

The reductive alkylation <strong>of</strong> ammonia with ketones is per<strong>for</strong>med under conditions similar to<br />

those <strong>for</strong> aldehydes, but appears to proceed with more difficulty. Initially, reductive<br />

amination <strong>of</strong> ketones with ammonia was tested without any additives resulting in lower<br />

yields. [7,8] Primary amines are considered better neocluophihes compared to ammonia.<br />

Despite <strong>the</strong>ir higher nucleophilicity <strong>the</strong>y are more sterically hindered. They were tested in<br />

reductive amination <strong>of</strong> carbonyl compounds utilizing different heterogeneous catalysts as<br />

nickel, platinum oxide, platinum sulphide and nickel sulphide. The following examples are<br />

some early trials <strong>for</strong> <strong>the</strong> preparation <strong>of</strong> secondary amines from primary amines (scheme<br />

3.1). [9-12] NH 2 CH(CH 2 ) 2 CHO<br />

(0.5mol)<br />

Ni-kieselguhr<br />

125 o C,100 bar H 2 ,1h<br />

(0.55mol) 54<br />

(91%)<br />

NHC 4 H 9


.<br />

CH 3 CO(CH 2 ) 5 CH 3 H 2 NCH 2 CH 2 OH<br />

Pt oxide*<br />

C 6 H 13 CH(CH 3 )NHCH 2 CH 2 OH<br />

(1.3mol) (1mol)<br />

100ml EtOH<br />

RT, 1-2 bar H 2 ,7h<br />

(96%)<br />

*Prereduced in 50ml EtOH at 1 bar H 2<br />

O<br />

(1.0mol)<br />

NH 2<br />

(1.90mol)<br />

Ni sulfide*<br />

180 o C,100-120 bar H 2 ,14h<br />

*Supported on montmorillonite (15% Ni)<br />

NH<br />

(95.5%)<br />

PhHN NH 2 MeCOCH 2 CHMe 2<br />

Pt sulfide-C<br />

PhHN NHCH(CH 3 )CH 2 CHMe 2<br />

175-180 o C, 30-40 bar H 2 ,4.5h<br />

(0.86mol) (0.95mol)<br />

(99%)<br />

Scheme 3.1. <strong>Preparation</strong> <strong>of</strong> Secondary <strong>Amines</strong><br />

In <strong>the</strong>ir attempts to overcome <strong>the</strong> problems associated with reductive amination, scientists<br />

tested <strong>the</strong> effect <strong>of</strong> additives on this trans<strong>for</strong>mation. It was found that <strong>the</strong> addition <strong>of</strong> a small<br />

quantity <strong>of</strong> Brønsted acid improved <strong>the</strong> yield dramatically. Dialkyl ketones, especially<br />

sterically hindered ones, tended to produce <strong>the</strong> corresponding alcohols to significant extents<br />

under <strong>the</strong> conditions <strong>of</strong> reductive amination decreasing <strong>the</strong> overall yield <strong>of</strong> <strong>the</strong> amine. The<br />

addition <strong>of</strong> a small amount <strong>of</strong> acetic acid or ammonium acetate is effective in suppressing<br />

alcohol <strong>for</strong>mation. Thus, <strong>the</strong> <strong>for</strong>mation <strong>of</strong> 2-nonanol could be depressed effectively in <strong>the</strong><br />

presence <strong>of</strong> ammonium acetate in <strong>the</strong> reductive amination <strong>of</strong> 2-nonanone (scheme 3.2). [3]<br />

CH 3 OC 7 H 15<br />

10 mL EtOH/0.8 g (0.047mol) NH 3<br />

50 o C, 80 bar H 2 ,Ra-Ni,Brønstedacid<br />

CH 3 CH(NH 2 )C 7 H 15 CH 3 CHOHC 7 H 15<br />

100% 0%<br />

Scheme 3.2. Reductive Amination <strong>of</strong> 2-nonanone<br />

3.1.2. Reductive Amination Utilizing Homogenous Catalysis:<br />

As mentioned be<strong>for</strong>e heterogeneous catalysts were first utilized <strong>for</strong> reductive amination.<br />

After <strong>the</strong> introduction <strong>of</strong> Wilkinson catalyst which opened <strong>the</strong> door <strong>for</strong> <strong>the</strong> use <strong>of</strong><br />

homogenous catalysts in organic syn<strong>the</strong>sis. Interest has been expressed in <strong>the</strong> use <strong>of</strong><br />

homogeneous catalysts <strong>for</strong> reductive amination. Bakos was <strong>the</strong> first to utilize homogenous<br />

55


catalysts <strong>for</strong> reductive amination in 1974. He tested rhodium and cobalt based catalysts <strong>for</strong><br />

<strong>the</strong> reductive alkylation <strong>of</strong> ammonia and aniline derivatives. He found that <strong>the</strong> activity <strong>of</strong><br />

cobalt catalyst is highly influenced by <strong>the</strong> structure <strong>of</strong> phosphine ligand. Also he recognized<br />

that using basic aliphatic amines led to poising <strong>of</strong> <strong>the</strong> cobalt catalyst and no product was<br />

<strong>for</strong>med. On <strong>the</strong> o<strong>the</strong>r hand rhodium was used successfully in <strong>the</strong> reductive amination <strong>of</strong> <strong>the</strong>se<br />

basic amines. [13] Despite <strong>the</strong>se interesting results <strong>the</strong> reaction conditions were harsh (100–300<br />

bar H 2 , 110–200 °C) and no turnover number were reported. In 2000, Börner described more<br />

practical system <strong>for</strong> homogenous reductive amination. [14] Benzaldehyde and piperidine could<br />

be reductively aminated using [Rh(dppb)(COD)]BF 4 or [Rh(1,2-bisdiphenylphosphinitoethane)(COD)]BF<br />

4 under milder conditions (50 bar H 2 , room<br />

temperature) with 500 TON (scheme 3.3).<br />

O<br />

H<br />

H<br />

N<br />

0.2% Rh cat<br />

H 2 (50 bar)<br />

rt,


superior in terms <strong>of</strong> conversion (89-92%) to <strong>the</strong> commercially applied Pt/C catalyst (74%<br />

conversion) (scheme 3.5). [16]<br />

O 10 bar H 2<br />

PhHN NH2 PhHN NH<br />

5h,120°C<br />

[Rh(COD)(PPh 3 ) 2 ]BF 4 ,TON=1060<br />

[Rh(COD)(PPh 3 ) 2 ]BF 4 on MM-K10, TON=1010<br />

Scheme 3.5. Reductive Amination with Rhodium Supported Catalyst.<br />

3.2. Reductive Amination <strong>the</strong> Current State <strong>of</strong> Art:<br />

Reductive amination is a one-pot process in which <strong>the</strong> <strong>for</strong>mation and <strong>the</strong> isolation <strong>of</strong> imines<br />

or enamines are avoided. Over <strong>the</strong> last three decades several research groups studied this<br />

trans<strong>for</strong>mation and factors affecting it. It was proved that pH has important influence on <strong>the</strong><br />

progress <strong>of</strong> <strong>the</strong> reaction. [17] It was proposed that reductive amination passes through reduction<br />

<strong>of</strong> imine or iminium ion. As it is shown in (scheme 3.6), a carbonyl compound combines with<br />

a primary or secondary amine to <strong>for</strong>m a hemiaminal species which <strong>for</strong>ms an iminium ion.<br />

This iminium ion loses hydrogen resulting in <strong>the</strong> <strong>for</strong>mation <strong>of</strong> imine which is <strong>the</strong>n reduced to<br />

<strong>the</strong> amine product. The most critical factor in reductive amination is <strong>the</strong> good choice <strong>of</strong><br />

conditions which favours intermediate reduction over ketone reduction to suppress alcohol<br />

<strong>for</strong>mation. [18]<br />

O<br />

R 1 R 2<br />

H<br />

N R 3<br />

NH 2 R 3<br />

HN R 3<br />

H 2<br />

R 1 R 2<br />

R 1<br />

HO<br />

R 2<br />

+H +<br />

-H 2 O<br />

H<br />

N R 3<br />

N R 4<br />

+H 2 O<br />

-H + R 1 R 2<br />

R 1 R 2<br />

Scheme 3.6. Mechanism <strong>of</strong> Reductive Amination.<br />

The most common strategies in reductive amination representing <strong>the</strong> current state <strong>of</strong> art can<br />

be subdivided into three main strategies. In <strong>the</strong> first strategy reduction is carried out using<br />

molecular hydrogen with heterogeneous catalysts (palladium, platinum or nickel catalysts).<br />

57


This is a straight <strong>for</strong>ward, environmentally friendly with easy procedures, but incompatible<br />

with <strong>the</strong> coexisting functional groups such as nitro, cyano and C-C multiple bonds. [19] The<br />

second strategy is based on <strong>the</strong> transfer hydrogenation conditions utilizing <strong>for</strong>mic acid or one<br />

<strong>of</strong> its derivative (Leuckart-Wallach type). [20] The third strategy uses hydride reductants e.g.<br />

NaBH 3 CN, [21] LiBH 3 CN, [22] [23] [24] [25]<br />

NaBH 3 CN-ZnCl 2, NaBH 3 CNMg(ClO 4 ) 2, NaBH 4 -NiCl 2,<br />

[26]<br />

NaBH 4 -ZnCl 2, borohydride exchange resin, [27] [28]<br />

[29]<br />

ZnBH 4, ZnBH 4 -ZnCl 2, pyridineborane,<br />

[30] picoline-borane, [31] etc. Recently organocatalysts were used in reductive amination<br />

utilizing Hantzsch esters or silanes as hydride sources with organocatalysts as chiral<br />

phosphoric acid and its derivatives. [32]<br />

Reviewing <strong>the</strong> literature <strong>of</strong> <strong>the</strong> last 50 years it is obvious that among <strong>the</strong> different reductive<br />

amination strategies discussed above, hydride reduction with NaBH 3 CN which was<br />

introduced by Borch [33] has been used extensively. This may be due to <strong>the</strong> ease <strong>of</strong> use <strong>of</strong> <strong>the</strong>se<br />

hydride sources. Borohydride salts are fur<strong>the</strong>rmore cheap, available in kg quantities and do<br />

not require special precautions in handling. Despite <strong>the</strong>se advantages borohydride salts suffer<br />

from o<strong>the</strong>r drawbacks.<br />

This reductant is used in excess quantity, toxic and produces toxic byproducts such as HCN<br />

and NaCN upon workup which limits its applications according to <strong>the</strong> new environmental<br />

standards. Abdel-Magid aimed to avoid this toxicity by using NaBH(OAc) 3 , (introduced by<br />

Gribble) as a mild reductant. [34,35] The mild nature <strong>of</strong> this reductant is due to <strong>the</strong> steric and<br />

electronic effects <strong>of</strong> <strong>the</strong> acetoxy groups which stabilize <strong>the</strong> B-H bond. His system was<br />

applicable <strong>for</strong> different types <strong>of</strong> aldehydes and unhindered aliphatic ketones. In spite <strong>of</strong> <strong>the</strong><br />

significant applications <strong>of</strong> <strong>the</strong> above reductants <strong>the</strong>y are not free from limitations regarding<br />

functional group tolerance and side reactions. [36] Also <strong>the</strong> <strong>for</strong>mation <strong>of</strong> tertiary or secondary<br />

amine from primary amine (<strong>the</strong> desired product) due to over alkylation represents ano<strong>the</strong>r<br />

limitation. [37]<br />

Bhattacharyya and coworkers developed a highly efficient mild system <strong>for</strong> reductive<br />

amination utilizing Ti(O i Pr) 4 , and NaBH 4 as <strong>the</strong> hydride donor and an amine source e.g.<br />

ammonia, ammonium chloride or methylamine. He was able to obtain high yields <strong>for</strong><br />

different aldehydes, cyclic ketones and ketone with acid labile groups (scheme 3.7). [38]<br />

58


O<br />

NHR 3 R 4<br />

R 4 R 3 N OTi(O i Pr) 3 NaBH 3 CN R 4 R 3 N<br />

R 1 R 2 Ti(O i Pr) 4<br />

R 1 R R 1<br />

2<br />

Scheme 3.7. Reductive Amination in <strong>the</strong> Presence <strong>of</strong> Ti(O i Pr) 4 .<br />

H<br />

R 2<br />

Ti(O i Pr) 4 is considered as a mild and effective Lewis acid <strong>for</strong> suppressing alcohol <strong>for</strong>mation<br />

in <strong>the</strong> reductive amination <strong>of</strong> ketones and aldehydes. It is compatible with most <strong>of</strong> acidsensitive<br />

functional groups (e.g. acetonides, silyl e<strong>the</strong>rs, esters, amides etc.). [39]<br />

In order to understand <strong>the</strong> role <strong>of</strong> titaiunm isopropoxide in reductive amination, Matson<br />

mixed equimolar ratios <strong>of</strong> amine and ketone with excess amount <strong>of</strong> Ti(O i Pr) 4 . He tried to<br />

isolate and detect <strong>the</strong> intermediates. Nei<strong>the</strong>r imine nor enamine could be detected (by IR<br />

measurements) or isolated. There<strong>for</strong>e, he predicted <strong>the</strong> <strong>for</strong>mation <strong>of</strong> hemiaminal titanate<br />

intermediate (Scheme 3.8) which is an unstable complex and is reduced with NaBH 3 CN to<br />

<strong>for</strong>m <strong>the</strong> amine product. Earlier findings by o<strong>the</strong>r scientists supported this proposal. Reetz has<br />

also predicted a similar titanium intermediate in his reaction between ketone and titanium<br />

amides with diisobutyl aluminium hydride (DIBAL-H) as a reductant. [40]<br />

Also reductive amination was tested under solvent free conditions. The aldehyde or <strong>the</strong><br />

ketone is mixed with <strong>the</strong> amine and <strong>the</strong> mixture is mixed in a mortar with NaBH 4 or α-<br />

picoline borane until TLC showed disappearance <strong>of</strong> starting material (scheme 3.8). [41]<br />

X<br />

O<br />

H<br />

+<br />

PhNH 2<br />

NaBH 4 .H 3 BO 3 (1:1)<br />

grinding<br />

a: x = COMe d: x = CO 2 Me<br />

b: x = CN e: x = NHCOMe<br />

c: x =CO 2 H f: x =NO 2<br />

X<br />

NHPh<br />

H<br />

Scheme 3.8. Solvent Free Reductive Amination.<br />

Baba developed <strong>the</strong> use <strong>of</strong> dibutylchlorotin hydride-HMPA complex as a mild hydride source<br />

<strong>for</strong> <strong>the</strong> reductive amination <strong>of</strong> various ketones and aldehydes. Various aromatic aldehydes<br />

with para or ortho electron withdrawing and electron donating groups were tested producing<br />

59


high yields <strong>of</strong> secondary amines (81-99%). Cyanao, nitro and halogens substituents were<br />

tolerated and <strong>the</strong> amines were prepared in good to high yields (70-99%). Utilization <strong>of</strong><br />

aliphatic amines as n-propyl amine resulted in poor yields (3-56%) compared with aniline<br />

derivatives (70-99%). Aromatic ketones as acetophenone and it substituted derivatives were<br />

also tested showing lower yields (35-69%) compared to aromatic aldehydes. Also aliphatic<br />

ketones showed variable results, benzyl acetone was an excellent substrates af<strong>for</strong>ding 91%<br />

yield. O<strong>the</strong>r aliphatic 2-alakonones showed lower yields (


Scheme 3.9. Syn<strong>the</strong>sis <strong>of</strong> (S)-Metolachlor<br />

One <strong>of</strong> <strong>the</strong> significant examples <strong>for</strong> <strong>the</strong> asymmetric reductive amination was developed by<br />

Zhang. The substituted aromatic and heteroaromatic ketones (acetophenone and substituted<br />

acetophenone) were used as examples and produced ee in <strong>the</strong> range <strong>of</strong> 89-96% and >99%<br />

yield with 1.0 mol % <strong>of</strong> Ir-(S,S)-f- Binaphane catalyst (scheme 3.10). According to <strong>the</strong> report,<br />

Ti(OiPr) 4 did not have any effect on <strong>the</strong> enantioselectivity but facilitates producing imine in<br />

situ from a hemiaminal titanate intermediate (as discussed earlier). He found that <strong>the</strong> presence<br />

<strong>of</strong> iodine is essential <strong>for</strong> <strong>the</strong> reaction to proceed as <strong>the</strong> oxidative addition <strong>of</strong> I 2 to <strong>the</strong> Ir I<br />

precursor generates Ir III complex which is <strong>the</strong>n coordinated with hydrogen <strong>for</strong>ming Ir III -H<br />

complex to which imine is coordinated and starting <strong>the</strong> catalytic cycle. Despite <strong>the</strong>se<br />

fascinating ees and yields <strong>for</strong> <strong>the</strong> aromatic ketones, this system failed to reductively aminate<br />

aliphatic ketones. High catalyst loading and high hydrogen pressure (69 bar) are o<strong>the</strong>r<br />

limitations <strong>of</strong> this methodology. [44]<br />

Ar<br />

O<br />

R<br />

Ir-(S,S)-f-Binaphane (1.0 mol%)<br />

10% I 2 /Ti(OiPr) 4 (1.5 equiv.)<br />

p-anisidine (1.2 equiv.)<br />

H 2 (69 bar), RT<br />

HN<br />

∗<br />

Ar R<br />

OMe<br />

P<br />

Fe<br />

P<br />

Ir-(S,S)-f-Binaphane<br />

Scheme 3.10. Binaphane Iridium Catalyst In Asymmetric Reductive Amination.<br />

61


Pérez reported <strong>the</strong> use <strong>of</strong> BINAP derived palladium catalyst <strong>for</strong> <strong>the</strong> reductive amination <strong>of</strong><br />

alkyl and cycloaliphatic ketones (scheme 3.11). He used o-, m- and p-substituted aniline<br />

derivatives as <strong>the</strong> source <strong>of</strong> nitrogen, molecular hydrogen at 800 psi (55 bar), in CHCl 3 at 70<br />

°C <strong>for</strong> 24 h. The use <strong>of</strong> molecular sieves was crucial <strong>for</strong> obtaining <strong>the</strong> product in high yields.<br />

He observed that <strong>the</strong> presence <strong>of</strong> substituents on <strong>the</strong> aniline improved stereoselectivity with a<br />

little effect on reactivity as <strong>the</strong>y increase <strong>the</strong> steric bulk <strong>of</strong> <strong>the</strong> nitrogen source improving <strong>the</strong><br />

interaction. When isobutyl methyl ketone reacted with aniline <strong>the</strong> ee was 51% but when p-<br />

anisidine was used <strong>the</strong> ee jumped to 90%. One <strong>of</strong> <strong>the</strong> remarkable examples was <strong>the</strong> success in<br />

reductive amination <strong>of</strong> 3-alkanones (2-heptanone). Although <strong>the</strong> ee was mediocre (49-59%)<br />

but it is known that <strong>the</strong>se substrates are only accessible through carbanion chemistry. 2,3-<br />

butanedione underwent chemoselective reductive amination in good yields (83-85%) and low<br />

enantioselectivity (2-20% ee). Aryl ketones were also tested resulting in poor to mediocre ee<br />

(35-43%). One <strong>of</strong> <strong>the</strong> draw backs <strong>of</strong> this methodology is <strong>the</strong> harsh conditions needed <strong>for</strong> <strong>the</strong><br />

removal <strong>of</strong> phenyl ring to obtain <strong>the</strong> primary amine. [45]<br />

(MeCN) 2 PdBr 2 + P P<br />

Benzene<br />

rt, overnight<br />

P<br />

PdBr 2<br />

P<br />

P<br />

=<br />

P<br />

Me<br />

PPh 2<br />

PPh 2<br />

P(p-Tolyl) 2<br />

PPh<br />

PPh 2<br />

P(p-Tolyl) 2<br />

2<br />

Me<br />

O<br />

R 1 R 2<br />

+<br />

R 3<br />

NH 2<br />

Catalyst (2.5 mol%)<br />

5Åms,CHCl 3<br />

H 2 (800 psi),<br />

70 °C, 24 h<br />

HN<br />

R 1 * R 2<br />

R 3<br />

Scheme 3.11. BINAP Derived Palladium Catalyst in Asymmetric Reductive Amination<br />

Xiao extended his work on imine reduction and tested his iridium based catalysts <strong>for</strong><br />

reductive amination (scheme 3.12). He reported high ees and high yields <strong>for</strong> different<br />

acetophenone derivatives with para-anisidine as nitrogen source. The catalyst loading was<br />

1.0 mol% with 5.0 mol% <strong>of</strong> phosphoric acid derivative as Lewis acid. The addition <strong>of</strong><br />

molecular sieves was crucial <strong>for</strong> faster reaction. Using aniline with electron withdrawing<br />

62


groups decreased both <strong>the</strong> enantioselectivities and <strong>the</strong> yields. Ortho substituted acetophenone<br />

derivatives(o-Cl, o-Me,o-OMe, o-F) which are known to be difficult substrates <strong>for</strong> reductive<br />

amination were reduced with high yields and high enantioselectivities using less sterically<br />

hindered catalysts. Aliphatic ketones were also tested and showed high ees and high yields<br />

with para-anisidine and with aniline. [46]<br />

Ph<br />

Ph<br />

O<br />

O NH 2<br />

Ar<br />

S<br />

N<br />

N<br />

H 2<br />

O<br />

Ir<br />

X<br />

OMe<br />

[Ir] (1.0 mol %)<br />

5barH 2<br />

toluene<br />

35 o C, 12h<br />

Ph<br />

Ph<br />

Ar = 2,4,6-(2-C 3 H 7 ) 3 C 6 H 2<br />

a: = Cl; b: X = 7-H<br />

c: Ar = 4-CH 3 C 6 H 4 ,X=7-H<br />

d: Ar = 2,3,4,5,6-(CH 3 ) 5 C 6 ,X=7-H<br />

O<br />

Me<br />

S<br />

N<br />

N<br />

H 2<br />

O<br />

X=7-H<br />

Ir<br />

X<br />

HN OMe OH<br />

Ar<br />

O<br />

O<br />

O<br />

P<br />

OH<br />

Ar<br />

OMe<br />

OMe<br />

OMe<br />

O<br />

HN<br />

O 2 N<br />

HN<br />

HN<br />

O<br />

92 % yield, 95 % ee<br />

OMe<br />

88 % yield, 81 % ee<br />

93 % yield, 95 % ee<br />

HN<br />

Br<br />

OMe<br />

HN<br />

F<br />

HN<br />

92 % yield, 96 % ee<br />

85 % yield, 85 % ee<br />

89 % yield, 95 % ee<br />

Scheme 3.12. Reductive Amination Utilizing Iridium Catalyst and Phosphoric Acid.<br />

63


3.3.2. Reductive Amination Utilizing <strong>Chiral</strong> Auxiliary.<br />

Ellman was <strong>the</strong> first to introduce <strong>the</strong> use <strong>of</strong> t-butylsulfinylamide as a chiral auxiliary <strong>for</strong> <strong>the</strong><br />

asymmetric reductive amination (Scheme 3.13). [47] The imine was generated in situ with<br />

Ti(OEt) 4 at 60-70 °C which was <strong>the</strong>n reduced with NaBH 4 at -48 °C. Ti(OEt) 4 serves as a<br />

desiccant, facilitates imine <strong>for</strong>mation and even helps to improve <strong>the</strong> yield and<br />

diastereoselectivity <strong>of</strong> <strong>the</strong> t-butylsulfinyl protected-amines. He demonstrated that his<br />

methodology is applicable <strong>for</strong> both aryl-alkyl and alkyl-alkyl ketones in 66-86% yield and<br />

80-94% de. Deprotection step is carried out under acidic conditions to obtain <strong>the</strong> primary<br />

amine without any compromise in <strong>the</strong> yield or ee. In his attempt to broaden <strong>the</strong> scope <strong>of</strong> <strong>the</strong><br />

reaction he tested L-Selectride as a hydride donor under similar conditions. The opposite<br />

enantiomer <strong>of</strong> <strong>the</strong> primary amine was obtained. The aliphatic ketone substrates gave similar<br />

yield and de <strong>of</strong> <strong>the</strong> protected amine but benzocyclic ketones showed an improved de with<br />

similar yields. [48]<br />

O<br />

S NH2<br />

O<br />

R 1 R 2<br />

NaBH 4<br />

THF<br />

-48 °C<br />

O R 1<br />

S N R 2<br />

O R 1<br />

S NH R 2<br />

Ti(OEt) 4<br />

THF<br />

L-Selectride<br />

O<br />

R 1<br />

S NH R 2<br />

THF<br />

-48 °C<br />

Scheme 3.13. Reductive Amination <strong>of</strong> Ketones using tert-butylsulfinylamide as Auxiliary.<br />

Pannecoucke used <strong>the</strong> chiral auxiliary developed by Ellman <strong>for</strong> <strong>the</strong> reductive amination <strong>of</strong> α-<br />

fluoro α,β-unsaturated ketones. He reported mediocre to good yields (46-86%) and high des<br />

(96-99%) <strong>for</strong> aryl, aliphatic and sterically aliphatic substrates. The imine was prepared in situ<br />

through combining Ti(OEt) 4 (2.0 equiv), (S)- tert-butylsulfinylamine (2.0 equiv), and 2-<br />

fluoroenone (1.0 equiv) in dry THF under argon and heated to reflux <strong>for</strong> 2 h. The mixture<br />

was allowed to cool to room temperature and <strong>the</strong>n cooled to -78 °C. DIBAL-H (1M in<br />

toluene, 4.0 equiv) was <strong>the</strong>n added dropwise, and <strong>the</strong> mixture was stirred <strong>for</strong> 1 h and <strong>the</strong><br />

64


progress <strong>of</strong> <strong>the</strong> reaction was monitored with NMR. The use <strong>of</strong> DIBAL-H resulted in <strong>the</strong> (S)<br />

con<strong>for</strong>mation and <strong>the</strong> use <strong>of</strong> L-Selectride resulted in (R) con<strong>for</strong>mation (scheme 3.14). [49]<br />

F<br />

R 1<br />

O<br />

H<br />

R 2<br />

H 2 N<br />

O<br />

tBu<br />

1) Ti(OEt) 4 ,THF,reflux<br />

2) Reducing agent, THF<br />

F<br />

HN S O<br />

R 2<br />

tBu<br />

Scheme 3.14. Reductive Amination <strong>of</strong> α-Fluoro α,β-Unsaturated Ketones.<br />

R 1<br />

H<br />

3.3.3. Reductive Amination Utilizing Molecular Hydrogen:<br />

Alexakis, utilized a combination <strong>of</strong> Ti(OiPr) 4 /Pd-C/H 2 to syn<strong>the</strong>size C2 symmetric secondary<br />

amines with 70-82% de and 88-92% yield. His system was only applicable <strong>for</strong> <strong>the</strong> aromatic<br />

substrates. [50] Nugent and Seemayer developed a highly efficient methodology <strong>for</strong> <strong>the</strong><br />

syn<strong>the</strong>sis <strong>of</strong> quinuclidine. [51] Through utilization molecular hydrogen, heterogeneous<br />

hydrogenation catalysts Pt-C or Pd-C and Ti(OiPr) 4 <strong>for</strong> <strong>the</strong> reductive amination <strong>of</strong> a labile α-<br />

chiral quinuclidinone <strong>the</strong>y were successful in incorporation <strong>the</strong> amine without epimerization,<br />

a feat not previously accomplished (scheme 3.15).<br />

N<br />

Ph<br />

O<br />

Ph<br />

Ph NH 2<br />

Ti(OiPr) 4 /Pt-C/H 2 (4.2 bar)<br />

25 o C, 15h, 80% de<br />

N<br />

Ph<br />

NHCH 2 Ph<br />

Ph<br />

N<br />

Ph<br />

NHCH 2 Ph<br />

Ph<br />

Scheme 3.15. Reductive Amination <strong>of</strong> a Quinuclidinone<br />

Nugent group recently developed a two-step methodology relying on <strong>the</strong> asymmetric<br />

reductive amination <strong>of</strong> prochiral ketones with <strong>the</strong> chiral ammonia equivalent (R)- or (S)-αmethylbenzylamine<br />

<strong>for</strong> α-chiral primary amine syn<strong>the</strong>sis. They found that <strong>the</strong> use <strong>of</strong><br />

(Ti(O i Pr) 4 (1.2 equiv)) with ((R)- or (S)-α- methylbenzylamine (α-MBA) (1.1 equiv)) and<br />

heterogeneous catalyst (Ra-Ni, Pd-C, Pt-C) produced <strong>the</strong> secondary amine <strong>of</strong> different 2-<br />

alkanones, cyclic ketones, and aryl alkyl ketones in high yields and diastereoselectivities. The<br />

secondary amine is produced in a single step without <strong>the</strong> need <strong>for</strong> <strong>the</strong> tedious process <strong>of</strong><br />

imine isolation and purification. Simple acid base work is usually enough to purify <strong>the</strong><br />

65


secondary amine product and any impurities like α-MBA (3-5%) can be removed easily<br />

through washing with NH 4 Cl. The primary amine is produced in high ee and yield through<br />

hydrogenolysis using Pd-C in MeOH and <strong>the</strong> ee can be fur<strong>the</strong>r enhanced through simple<br />

crystallization.<br />

Category 1: Raney-Ni<br />

Category 2: Pt-C<br />

HN<br />

Ph<br />

HN<br />

Ph<br />

HN<br />

Ph<br />

HN<br />

Ph<br />

76% yld,87% de<br />

1,2-dichloroethane<br />

94% yld, 74% de<br />

THF<br />

79% yld, 87% de<br />

hexane<br />

82% yld, 92% de<br />

ethanol<br />

Category 3: Pd-C<br />

Ph<br />

HN<br />

Ph<br />

Ph<br />

HN<br />

Ph<br />

NH 2<br />

H 2 N<br />

89% yld, 80% de<br />

methylenechloride<br />

92% yld, 94% de<br />

ethylacetate<br />

92% ee, overall yld 76%<br />

ethylacetate<br />

76% ee, overall yld 64%<br />

methyl-t-butyle<strong>the</strong>r<br />

Figure 3.1. Correlation <strong>of</strong> Heterogeneous Hydrogenation Catalysts with Ketone Structure<br />

and Product example.<br />

Nugent group also investigated o<strong>the</strong>r commercially available Lewis acids and <strong>the</strong>y found that<br />

B(O i Pr) 3 or Al(O i Pr) 3 , can be used to replace Ti(O i Pr) 4 . These Lewis acids are cheaper than<br />

Ti(O i Pr) 4 but must be used in greater quantities. These Lewis acids hold <strong>the</strong> advantage over<br />

Ti(O i Pr) 4 due to <strong>the</strong>ir easier work up procedures. On <strong>the</strong> work-up <strong>of</strong> Ti(O i Pr) 4 reaction a<br />

finely dispersed TiO 2 can be <strong>for</strong>med <strong>for</strong>cing a celite filtration onscale. If no Lewis acid or <strong>the</strong><br />

wrong one is present, large quantities <strong>of</strong> <strong>the</strong> alcohol by product can be expected. [52]<br />

3.3.4. Asymmetric Reductive Amination Utilizing Transfer Hydrogenation Conditions<br />

(<strong>the</strong> Leuchart–Wallach Reaction):<br />

As mentioned be<strong>for</strong>e, <strong>the</strong> source <strong>of</strong> hydrogen can be molecular hydrogen, hydride or through<br />

utilizing transfer hydrogenation conditions. Transfer hydrogenation conditions were applied<br />

successfully <strong>for</strong> <strong>the</strong> reduction <strong>of</strong> ketones to alcohols. The method is highly successful in<br />

terms <strong>of</strong> obtaining high ees and high yields. [53] As it is a highly desirable goal several<br />

66


attempts were directed <strong>for</strong> <strong>the</strong> asymmetric reductive amination <strong>of</strong> ketones under transfer<br />

hydrogenation conditions but with limited success compared to ketone reduction.<br />

Never<strong>the</strong>less, some useful recent breakthroughs have been achieved. [54] A remarkably<br />

effective asymmetric version was reported by Kadyrov, Riermeier and Börner. [55] They<br />

reported <strong>the</strong> use <strong>of</strong> several ru<strong>the</strong>nium and rhodium catalysts <strong>for</strong> <strong>the</strong> conversion <strong>of</strong><br />

acetophenone derivatives to <strong>the</strong> enantiomerically enriched amines (Scheme 3.17 ). According<br />

to <strong>the</strong>ir strategy different aryl-alkyl ketones can be utilized as substrates producing <strong>the</strong><br />

primary amines in high yields (74-92%) and enantioselectivity (89-95%) in a one step<br />

reaction. HCO 2 NH 4 was used as <strong>the</strong> hydride source with NH 3 as <strong>the</strong> nitrogen source and<br />

different BINAP ligands were tested showing <strong>the</strong> best catalyst was [Ru-(R)-(TolBINAP)Cl 2 ].<br />

In <strong>the</strong> reaction along with <strong>the</strong> primary amine a <strong>for</strong>myl derivative (RHNC(O)H) is produced<br />

and in order to improve <strong>the</strong> yield <strong>the</strong> crude product is treated with HCl (EtOH/H 2 O) at reflux<br />

to obtain <strong>the</strong> desired amine in a good to excellent yield. Despite <strong>the</strong>se encouraging results <strong>the</strong><br />

application <strong>of</strong> this method is limited to <strong>the</strong> aromatic substrates in which o<strong>the</strong>r substrates as 1-<br />

indanone (6% yield, no reported ee, chiral Ru catalyst) and aliphatic ketones, e.g. 2-octanone<br />

(44% yield, 24% ee, using chiral Ru catalyst) showed unsatisfactory results (scheme 3.16).<br />

R'<br />

O<br />

R<br />

(i)1 mol% [Ru-(R)-TolBINAP(Cl) 2 ]<br />

NH 4 HCO 2 ,NH 3 /MeOH (15-20%), 85 o C<br />

(ii) 6N HCl, reflux, 1h<br />

R'<br />

H NH 2<br />

R<br />

Aryl group R Yield and ee<br />

Ph Me 92, 95% ee R<br />

Ph Et 89, 95% ee R<br />

3-MeC 6 H 4 Me 74, 89% ee R<br />

4-MeC 6 H 4 Me 93, 93% ee R<br />

4-ClC 6 H 4 Me 93, 92% ee R<br />

4-(NO 2 )C 6 H 4 Me 92, 95% ee R<br />

Scheme 3.16. Transfer Hydrogenation <strong>of</strong> Acetophenone Derivatives.<br />

3.4.5. Organocatalytic Asymmetric Reductive Amination:<br />

The use <strong>of</strong> organocatalysts was slowly introduced to organic chemistry over <strong>the</strong> last two<br />

decades. There were earlier trials <strong>of</strong> using organic compounds to catalyze organic reactions<br />

but <strong>the</strong> low yields and stereoselectivities were discouraging. Recent advances in<br />

spectroscopic and asymmetric techniques have opened <strong>the</strong> door <strong>for</strong> <strong>the</strong> syn<strong>the</strong>sis <strong>of</strong> a big<br />

67


library <strong>of</strong> organocatalysts which were used efficiently <strong>for</strong> different organic trans<strong>for</strong>mations.<br />

Several research groups focused <strong>the</strong>ir ef<strong>for</strong>ts on <strong>the</strong> syn<strong>the</strong>sis <strong>of</strong> novel organocatalysts <strong>for</strong><br />

asymmetric imine reduction and reductive amination. Among <strong>the</strong>se trans<strong>for</strong>mations are imine<br />

reduction and reductive amination. At this time, development in <strong>the</strong> application <strong>of</strong><br />

organocatalysis <strong>for</strong> reductive amination is still in its infantile stage compared to o<strong>the</strong>r<br />

trans<strong>for</strong>mations. [56]<br />

X +2 [H]<br />

H<br />

X<br />

∗<br />

chiral<br />

catalyst R 1<br />

X=CR 2 ,O,NR<br />

R 1 R 2<br />

R 2<br />

Scheme 3.17. Asymmetric Reduction <strong>of</strong> prochiral Compounds.<br />

Inspired by nature and how living organisms reduce imino group through <strong>the</strong> employment <strong>of</strong><br />

organic dihydropyridine c<strong>of</strong>actors such as nicotinamide adenine dinucleotide (NADH) in<br />

combination with enzyme catalysts (figure 3.2). [57]<br />

H<br />

H<br />

O<br />

NH 2<br />

.<br />

- O<br />

O<br />

O<br />

P<br />

O<br />

O<br />

P<br />

O -<br />

O<br />

N<br />

H H<br />

O<br />

H H<br />

OH OH<br />

O H HAdenine<br />

O<br />

H H<br />

OH OH<br />

reduced nicotinamide<br />

adenine dinucleotide<br />

(NADH)<br />

Figure 3.2. Reduced Nicotinamide Adenine Dinucleotide<br />

Scientists started to think <strong>of</strong> NADH analogues and <strong>the</strong>y found that <strong>the</strong> best analogues would<br />

be Hantzsch esters. These hydrogen sources in <strong>the</strong> presence <strong>of</strong> achiral Lewis or Brønsted acid<br />

catalysts proved to be efficient in imine reduction. [58] List investigated <strong>the</strong> catalytic cycle <strong>of</strong><br />

reductive amination utilizing Hantzsch esters. He proposed that reductive amination <strong>of</strong><br />

ketones is initiated by protonation <strong>of</strong> <strong>the</strong> in situ generated ketimine from a chiral Brønsted<br />

acid catalyst (Scheme 3.18). The resulting iminium ion pair, which may be stabilized by<br />

68


hydrogen bonding, is chiral and its reaction with <strong>the</strong> Hantzsch dihydropyridine could give an<br />

enantiomerically enriched amine and pyridine. After screening different phosphoric acid<br />

catalysts, catalyst 9 was found to be <strong>the</strong> best catalyst <strong>for</strong> this reaction and 1.0 mol% <strong>of</strong> <strong>the</strong><br />

catalyst resulted in 93% ee <strong>for</strong> <strong>the</strong> product with an excellent yield <strong>of</strong> 96% (scheme 3.19). [59]<br />

OMe<br />

OMe<br />

O<br />

+PMPNH 2<br />

-H 2 O<br />

N<br />

HN<br />

X*<br />

EtOOC<br />

N<br />

H<br />

COOEt<br />

HN<br />

PMP<br />

H<br />

HX*<br />

H 2 N<br />

H<br />

X*<br />

OMe<br />

EtOOC<br />

N<br />

COOEt<br />

Scheme 3.18. Mechanism <strong>of</strong> <strong>Chiral</strong> Brønsted Acid Catalysed Reductive Amination.<br />

O<br />

9(1mol%),toluene,35 o C,71h, 98%<br />

HN PMP<br />

EtOOC<br />

COOEt<br />

N<br />

H<br />

(1.4 equiv)<br />

93% ee<br />

i-Pr<br />

i-Pr<br />

O<br />

O<br />

i-Pr<br />

P<br />

O i-Pr OH<br />

i-Pr i-Pr<br />

9<br />

Scheme 3.19. Organocatalytic Reductive Amination Developed by List.<br />

List also investigated <strong>the</strong> reductive amination <strong>of</strong> aldehydes. He proposed that under <strong>the</strong><br />

conditions <strong>of</strong> reductive amination an α-branched aldehyde substrate would undergo a fast<br />

racemization in <strong>the</strong> presence <strong>of</strong> <strong>the</strong> amine and acid catalyst via an imine/enamine<br />

tautomerization. The reductive amination <strong>of</strong> one <strong>of</strong> <strong>the</strong> two imine enantiomers would <strong>the</strong>n<br />

69


have to be faster than that <strong>of</strong> <strong>the</strong> o<strong>the</strong>r, resulting in an enantiomerically enriched product<br />

which is ano<strong>the</strong>r successful example <strong>of</strong> a dynamic kinetic resolution process. [60]<br />

He reported high enantioselectivities <strong>for</strong> <strong>the</strong> reductive amination <strong>of</strong> hydratopicaldehyde with<br />

p-anisidine in <strong>the</strong> presence <strong>of</strong> Hantzsch ester and phosphoric acid catalyst 9. [61] List<br />

demonstrated that racemic aldehydes could be successfully converted to branched-chain<br />

secondary amines in an excellent enantiomeric excess. He found that <strong>the</strong> use <strong>of</strong> a highly<br />

hindered phosphate catalyst was essential, and intriguingly <strong>the</strong> best results required <strong>the</strong> very<br />

specific use <strong>of</strong> a particular Hantzsch ester, in this case (Scheme 3.20).<br />

O H 2 NR 3<br />

R 1 R 2<br />

[H]<br />

O<br />

R 1<br />

H H 2 NR 3<br />

R 2 [H]<br />

NHR 3<br />

R 1<br />

∗R 2<br />

α-branched chiral amines<br />

R 1 ∗<br />

β-branched chiral amines<br />

NHR 3<br />

R 2<br />

R 1<br />

R 2<br />

O<br />

H<br />

+H 2 NR 3<br />

-H 2 O<br />

R 1<br />

R 3<br />

N<br />

R 3<br />

HN<br />

H<br />

R 1<br />

H<br />

R 2<br />

R 2<br />

racemization<br />

R 3<br />

N<br />

R 1<br />

R 2<br />

H<br />

H<br />

N<br />

H<br />

HX*<br />

R 3<br />

N<br />

R 1<br />

R 2<br />

H<br />

R 1<br />

NHR 3<br />

R 2<br />

EtOOC<br />

COOEt<br />

EtOOC<br />

COOEt<br />

N<br />

R CHO 9(5mol%),MS5Å,C 6 H 6 ,6 o 1 C, 72 h R 1<br />

NHR 3<br />

R 2 R 2<br />

EtOOC<br />

COOEt<br />

N<br />

H<br />

(1.2 equiv)<br />

F<br />

NHPMP<br />

NHPMP<br />

NHPMP<br />

NHPMP<br />

87%, 96% ee<br />

88%, 98% ee<br />

89%, 94% ee<br />

92%, 98% ee<br />

NHPh<br />

NHPMP<br />

78%, 94% ee 77%, 80% ee<br />

Scheme 3.20. Asymmetric Reductive Amination <strong>of</strong> Aldehydes.<br />

70


Later he utilized this methodology <strong>for</strong> <strong>the</strong> syn<strong>the</strong>sis <strong>of</strong> important pharmaceutical building<br />

blocks. He reported <strong>the</strong> enantioselective syn<strong>the</strong>sis <strong>of</strong> pharmaceutically relevant 3-substituted<br />

cyclohexylamines from 2,6-diketones via an aldolization-dehydrationconjugate reductionreductive<br />

amination cascade that is catalyzed by a chiral Brønsted acid and accelerated by <strong>the</strong><br />

achiral amine substrate, which is ultimately incorporated into <strong>the</strong> product. 2,6-diketone was<br />

treated with 1.0 equiv <strong>of</strong> an achiral amine, 2.0 equiv <strong>of</strong> a Hantzsch ester, and 10 mol % <strong>of</strong> a<br />

chiral Brønsted acid resulted in <strong>the</strong> <strong>for</strong>mation <strong>of</strong> <strong>the</strong> corresponding cyclohexylamines with<br />

mediocre to good yield (35-79%) and with good to high diastereoselectivity (82-96%). Alkyl,<br />

aryl and sterically congested aryl substituted 2,6-diketones were reductively aminated with<br />

high stereoselectivities. [62]<br />

NHR 2<br />

+ NHR 2 + NHR<br />

X - 2<br />

X -<br />

R 1<br />

X<br />

OR 1<br />

R 1<br />

i-Pr<br />

i-Pr<br />

Y<br />

i-Pr<br />

O<br />

O<br />

OOH<br />

P EtO 2 C CO 2 Et<br />

i-Pr<br />

N<br />

O<br />

H<br />

3(2.2equiv)<br />

i-Pr i-Pr<br />

R 2 NH 2 (1.5 equiv)<br />

(R)-TRIP (10mol%)<br />

MS 5Å, cyclohexane, 50<br />

OR o C<br />

1<br />

Y<br />

HN R 2<br />

R 1<br />

Scheme 3.21. Syn<strong>the</strong>sis <strong>of</strong> Pharmaceutical Building Blocks Utilizing Reductive Amination.<br />

Menche have also demonstrated that thiourea acts as an efficient organocatalyst <strong>for</strong> <strong>the</strong><br />

reductive amination <strong>of</strong> aldehydes using aniline derivative and <strong>the</strong> ethyl Hantzsch ester<br />

providing <strong>the</strong> corresponding achiral N-benzylanilines in good to excellent yields (72-93%).<br />

Using 1.1 equiv <strong>of</strong> Hantzsch ester and 1.0 equiv <strong>of</strong> thiourea with molecular sieves in toluene<br />

at 70 °C <strong>for</strong> 24 h substituted benzaldehydes as well as two aliphatic aldehydes were reacted<br />

with p-anisidine <strong>for</strong>ming <strong>the</strong> secondary amines in good isolated yields (scheme 3.22). [63]<br />

71


EtO 2 C<br />

H<br />

H<br />

CO 2 Et<br />

R<br />

O<br />

H<br />

+<br />

H 2 N<br />

OMe<br />

N<br />

H<br />

S<br />

(1.1 equiv)<br />

R<br />

N<br />

H<br />

OMe<br />

H 2 N NH 2<br />

(1.0 equiv)<br />

5 Å MS, toluene, 70 °C<br />

Scheme 3.22. Reductive Amination <strong>of</strong> Aldehydes Utilizing Thiourea.<br />

One <strong>of</strong> <strong>the</strong> milestones in this field was developed by MacMillan, who investigated <strong>the</strong><br />

reductive amination <strong>of</strong> acetophenone with p-anisidine utilizing ethyl Hantzsch ester, and<br />

BINOL-derived phosphoric acids. The catalyst showed high catalytic activity and excellent<br />

enantiocontrol in <strong>the</strong> reductive coupling reaction. Removal <strong>of</strong> water by <strong>the</strong> addition <strong>of</strong> 5Å<br />

molecular sieves proved to be important <strong>for</strong> achieving high catalytic activity and selectivity.<br />

Aryl-alkyl ketones were reduced with good yields (70-87%) and high enantioselectivity (85-<br />

97%) and 2-alkanones were also reduced with mediocre to good yields (49-75%) with good<br />

to high enantioselectivities (81-94%). Different aniline derivatives were tested producing <strong>the</strong><br />

secondary amine with good to high yields (55-92%) and high enantioselectivities (90-95%).<br />

The overall diversity <strong>of</strong> ketone substrates makes his work fascinating from all aspects. [64]<br />

O<br />

15 (10 mol%),PMPNH 2 (1 equiv), MS 5A,C 6 H 6 ,50 o C, 96h, 87%<br />

HN PMP<br />

EtOOC<br />

COOEt<br />

N<br />

H<br />

(1.2 equiv)<br />

87% yield, 94 % ee<br />

SiPh 3<br />

O<br />

O<br />

P<br />

O OH<br />

15<br />

SiPh 3<br />

OMe<br />

OMe<br />

OMe<br />

HN<br />

F<br />

HN<br />

F<br />

HN<br />

79 % yield, 91 % ee<br />

60 % yield, 83 % ee<br />

72


OMe<br />

OMe<br />

O<br />

HN<br />

HN<br />

HN<br />

O 2 N<br />

71 % yield, 95 % ee<br />

75 % yield, 85 % ee<br />

92 % yield, 91 % ee<br />

HN<br />

OMe<br />

HN<br />

OMe<br />

HN<br />

OMe<br />

71 % yield, 83 % ee<br />

2<br />

60 % yield, 90 % ee<br />

73 % yield, 96 % ee<br />

Scheme 3.23. Asymmetric Reductive Amination System Developed by MacMillan<br />

3.4. Green Chemistry and Reductive Amination:<br />

3.4.1. Green Chemistry Basic principles.<br />

Reductive amination is a one pot process <strong>for</strong> chiral amine syn<strong>the</strong>sis. As described previously,<br />

it has many advantages compared to <strong>the</strong> o<strong>the</strong>r available methodologies. One <strong>of</strong> <strong>the</strong> significant<br />

advantages <strong>of</strong> this methodology is that it is considered an environmentally friendly process.<br />

To understand on which bases scientists made such assumption, we have first to know more<br />

about green chemistry and environmentally friendly process.<br />

It is widely acknowledged that <strong>the</strong>re is a growing need <strong>for</strong> more environmentally acceptable<br />

processes in <strong>the</strong> chemical industry and pharmaceutical industry. This trend has led to <strong>the</strong><br />

concept <strong>of</strong> ‘Green Chemistry’. [65]<br />

The new trend differs dramatically from <strong>the</strong> old traditional concepts focused only on process<br />

efficiency and chemical yield. The new trend assigns <strong>the</strong> economic value <strong>of</strong> <strong>the</strong> process<br />

depending on its ability to eliminate waste at source and avoid <strong>the</strong> use <strong>of</strong> toxic and/or<br />

73


hazardous substances. Green chemistry can be defined as <strong>the</strong> new trend in chemistry which<br />

efficiently utilizes (preferably renewable) raw materials, eliminates waste and avoids <strong>the</strong> use<br />

<strong>of</strong> toxic and/or hazardous reagents and solvents in <strong>the</strong> manufacture and application <strong>of</strong><br />

chemical products. [65]<br />

Recently scientists proposed <strong>the</strong> basic principles <strong>of</strong> Green Chemistry which can be<br />

paraphrased as: [65,66]<br />

1. Waste prevention instead <strong>of</strong> disposal.<br />

2. Atom efficiency and atom economy.<br />

3. Less hazardous/toxic chemicals<br />

4. Design safer process and safer product.<br />

5. Safe solvents and auxiliaries<br />

6. Design efficient energy manipulation.<br />

7. Use <strong>of</strong> renewable raw materials.<br />

8. Step wise efficient methodologies.<br />

9. Catalytic ra<strong>the</strong>r than stoichiometric reagents<br />

10. Biodegradable products.<br />

11. Desing analytical techniques <strong>for</strong> pollution control.<br />

12. Inherently safer working environment.<br />

Green chemistry main concern is <strong>the</strong> environmental impact <strong>of</strong> both chemical products and <strong>the</strong><br />

processes by which <strong>the</strong>y are produced. It is well known that prevention is better than cure.<br />

Green chemistry eliminates waste at source, it focuses on primary pollution prevention ra<strong>the</strong>r<br />

than waste remediation (end-<strong>of</strong>-pipe solutions).<br />

To be able to classify any chemical or pharmaceutical process as green process, two<br />

important measures <strong>of</strong> <strong>the</strong> potential environmental acceptability <strong>of</strong> <strong>the</strong> process have been<br />

introduced <strong>the</strong> E factor and <strong>the</strong> Atom Efficiency. [65]<br />

The E factor is defined as <strong>the</strong> mass ratio <strong>of</strong> <strong>the</strong> waste to <strong>the</strong> desired product. A higher E factor<br />

means more waste and, consequently, greater negative environmental impact. The ideal E<br />

factor is zero. It was found that <strong>the</strong> pharmaceutical industry has <strong>the</strong> highest E factor<br />

74


compared to <strong>the</strong> oil refinery industry which has <strong>the</strong> lowest E factor. Atom efficiency is<br />

calculated by dividing <strong>the</strong> molecular weight <strong>of</strong> <strong>the</strong> desired product by <strong>the</strong> sum <strong>of</strong> <strong>the</strong><br />

molecular weights <strong>of</strong> all <strong>the</strong> substances produced in <strong>the</strong> stoichiometric equation.<br />

3.4.2. Hydrogenation and Green Chemistry:<br />

Catalytic hydrogenation–utilizing hydrogen gas and heterogeneous catalysts–can be<br />

considered as <strong>the</strong> most important catalytic method in syn<strong>the</strong>tic organic chemistry on both<br />

laboratory and production scales. Hydrogen is, without doubt, <strong>the</strong> cleanest reducing agent and<br />

<strong>the</strong> heterogeneous robust catalysts have been routinely employed. Catalytic hydrogenation<br />

has distinctive advantages over o<strong>the</strong>r methodologies. Key advantages <strong>of</strong> this technique are: [67]<br />

1. Broad scope, many functional groups can be hydrogenated with high selectivity.<br />

2. High conversions are usually obtained under relatively mild conditions in <strong>the</strong> liquid phase.<br />

3. The large body <strong>of</strong> experience with this technique makes it possible to predict <strong>the</strong> catalyst<br />

<strong>of</strong> choice <strong>for</strong> a particular problem.<br />

4. The process technology is well established and scale-up is <strong>the</strong>re<strong>for</strong>e usually<br />

straight<strong>for</strong>ward.<br />

The field <strong>of</strong> hydrogenation is also <strong>the</strong> area where catalysis was first widely applied in <strong>the</strong> fine<br />

chemical industry. It is a key example <strong>of</strong> green technology, due to <strong>the</strong> low amounts <strong>of</strong><br />

catalysts required, in combination with <strong>the</strong> use <strong>of</strong> hydrogen (100% atom efficient!) as <strong>the</strong><br />

reductant. In general, if chirality is not required, heterogeneous supported catalysts can be<br />

used in combination with hydrogen. Catalytic hydrogenation is considered as <strong>the</strong> green route<br />

<strong>for</strong> <strong>the</strong> syn<strong>the</strong>sis <strong>of</strong> different functional compounds as amines, alcohols, and amino acids.<br />

Once selectivity and chirality is called <strong>for</strong>, homogeneous catalysts and biocatalysts are<br />

applied. The use and <strong>the</strong> application <strong>of</strong> chiral Ru, Rh and Ir catalysts has become a well<br />

developed technology. Homogenous catalytic hydrogenation gives access to a large variety <strong>of</strong><br />

asymmetric trans<strong>for</strong>mations: imines and functionalized ketones and alkenes can be converted<br />

with high selectivity in most cases.<br />

75


3.5. Conclusion:<br />

Different methodologies have been introduced <strong>for</strong> <strong>the</strong> syn<strong>the</strong>sis <strong>of</strong> α- chiral amines. One <strong>of</strong><br />

<strong>the</strong> most important strategy introduced <strong>for</strong> this purpose is reductive amination. Reductive<br />

amination is a stepwise efficient methodology starting from <strong>the</strong> prochiral ketone to <strong>the</strong> α-<br />

chiral amine. Earlier reports described <strong>the</strong> use <strong>of</strong> classical heterogeneous catalysts <strong>for</strong> this<br />

trans<strong>for</strong>mation. Homogeneous catalysts were also introduced in <strong>the</strong> seventies and marked a<br />

significant breakthrough in <strong>the</strong> field. Asymmetric version was introduced utilizing chiral<br />

catalyst, chiral auxiliaries or chiral organocatalysts. The use <strong>of</strong> Brønsted or Lewis acids was<br />

important in most <strong>of</strong> <strong>the</strong> methodologies. Earlier reports suggested <strong>the</strong> role <strong>of</strong> <strong>the</strong> acid as an<br />

efficient desiccant but recent reports suggested more complicated function as <strong>for</strong>cing<br />

hemiaminal <strong>for</strong>mation. Over <strong>the</strong> last two decades several methodologies were evolved<br />

allowing <strong>the</strong> syn<strong>the</strong>sis <strong>of</strong> α- chiral amines from aryl-alkyl ketone and 2-alaknones in good to<br />

high yields and enantioselectivities. [68]<br />

3.6. References:<br />

[1] 1. Mignonac, G. Compt. Rend. 1921, 172, 223.<br />

[2] For review articles, see (a) W. S. Emerson, Org. React. 1948, 4, 174; (b) F. Möller, R.<br />

Schröter, in Methoden der Organischen Chemie (Houben-Weyl), Thieme: Stuttgart; Bd. XI/1,<br />

1957, 602–673.<br />

[3] S. Nishimura, Handbook <strong>of</strong> Heterogeneous Catalytic Hydrogenation <strong>for</strong> Organic<br />

Syn<strong>the</strong>sis, John Wiley &Sons, Inc., New York, 2001.<br />

[4] C. F. Winans, J. Am. Chem. Soc. 1939, 61, 3566.<br />

[5] B. M. Vanderbilt, U.S. Pat. 2,219,879, 1941.<br />

[6] E. H. Pryde, D. E. Anders, J. C. Cowan, J. Am. Oil Chem. Soc. 1969, 46, 67.<br />

[7] A. Skita, F. Keil, Ber. Dtsch. Chem. Ges. 1928, 61, 1682.<br />

[8] R. Cantarel, Compt. Rend. 1940, 210, 403.<br />

[9] Winans, C. F.; Adkins, H. J. Am. Chem. Soc. 1932, 54, 306<br />

[10] A. C. Cope, E. M. Hancock, J. Am. Chem. Soc. 1942, 64, 1503<br />

[11] Dovell, F. S.; Greenfield, H. J. Org. Chem. 1964, 29, 1265<br />

[12] F. S. Dovell, H. Greenfield, J. Am. Chem. Soc. 1965, 87, 2767<br />

[13] L. Marko, J. Bakos, J. Organomet. Chem. 1974, 81, 411<br />

76


[14] V. I. Tararov, R. Kadyrov, T.H. Riermeier, A. Borner, J. Chem. Soc. Chem. Commun.<br />

2000, 1867.<br />

[15] T. Gross, A.M. Seayad, M. Ahmad, M. Beller, Org. Lett. 2002, 4, 2055.<br />

[16] R. Margalef-Catala, C. Claver, P. Salagre, E. Fernandez, Tetrahedron Lett. 2000, 41,<br />

6583.<br />

[17] R. F. Borch, M. D. Bernstein, H. D. Durst, J. Am. Chem. Soc. 1971, 93, 2897. R. F.<br />

Borch, A. I. Hassid, J. Org. Chem. 1972, 37, 1673.<br />

[18] A. F. Abdel-Majid, K. G. Carson, B. D. Harris, C. A. Maryan<strong>of</strong>f, R. D. Shah, J. Org.<br />

Chem.1996, 61, 3849.<br />

[19] a) P. N. Rylander, Catalytic hydrogenation in organic syn<strong>the</strong>sis; Academic:<br />

New York, 1979; p 165; b) M. V. Klyuev, M. L. Khidekel, Russ. Chem. Rev. 1980, 49, 14; c)<br />

T. A. Tarasevich, N. G. Kozlov, Russ. Chem. Rev., 1999, 68, 55.<br />

[20] M. Kitamura, M. Tsukamoto, Y. Bessho, M. Yoshimura, U. Kobs, M. Widhalm, R.<br />

Noyori, J. Am. Chem. Soc. 2002, 124, 6649.<br />

[21] R. F. Borch, M. D. Bernstein, H. D. Durst, J. Am. Chem. Soc. 1971, 93, 2897.<br />

[22] R. F. Borch, H. D. Durst, J. Am. Chem. Soc. 1969, 91, 3996.<br />

[23] S. Kim, C. H. Oh, J. S. Ko, K. H. Ahn, Y. J. Kim, J. Org. Chem.1985, 50, 1927.<br />

[24] J. Brussee, R. A. T. M. van Ben<strong>the</strong>m, C. G. Kruse, A. van der Gen, Tetrahedron:<br />

Asymmetry 1990, 1, 163.<br />

[25] I. Saxena, R. Borah J. C. Sarma, J. Chem. Soc., Perkin Trans. 1, 2000, 503.<br />

[26] S. Bhattacharyya, J. Org. Chem. 1995, 60, 4928.<br />

[27] N. M. Yoon, E. G. Kim, H. S. Son, J. Choi, Synth. Commun. 1993, 23, 1595.<br />

[28] H. Kotsuki, N. Yoshimura, I. Kadota, Y. Ushio, M. Ochi, Syn<strong>the</strong>sis 1990, 401.<br />

[29] S. Bhattacharyya, A. Chatterjee, J. S. Williamson, Synth. Commun. 1997, 27, 4265.<br />

[30] a) A. Pelter, R. M. Rosser, S. Mills, J. Chem. Soc., Perkin Trans. 1, 1984, 717; b) M. D.<br />

Bomann, I. C. Guch, M. DiMare, J. Org. Chem. 1995, 60, 5995.<br />

[31] S. Sato, T. Sakamoto, E. Miyazawa, Y. Kikugawa, Tetrahedron 2004, 60, 7899.<br />

[32] D. Menche, J. Hassfeld, J. Li, G. Menche, A. Ritter, S. Rudolph, Org. lett.2006, 8, 741.<br />

[33] R. F. Borch, M. D. Bernstein, H. D. Durst, J. Am. Chem. Soc. 1971, 93, 2897.<br />

[34] A. F. Abdel-Majid, K. G. Carson, B. D. Harris, C. A. Maryan<strong>of</strong>f, R. D. Shah, J. Org.<br />

Chem.1996, 61, 3849;<br />

[35] a) A. F. Abdel-Majid, S. J. Mehrman, Org. Process Res. Dev. 2006, 10, 971; b) G. W.<br />

Gribble, Org. Process Res. Dev. 2006, 10, 1062.<br />

77


[36] E. W. Baxter, A. B. Reitz, Organic reactions; Wiley: New York, 2002; Vol. 59.<br />

[37] M. B. Smith, J. March, March’s advanced organic chemistry; Wiley: New York, 2001.<br />

[38] a) S. Bhattacharyya, Tetrahedron Lett. 1994, 35, 2401; b) S. Bhattacharyya, J. Org.<br />

Chem. 1995, 60, 4928; c) K. A. Neidigh, M. A. Avery, J. S. Williamson, S. Bhattacharyya, J.<br />

Chem. Soc., Perkin Trans. 1, 1998, 2527.<br />

[39] R. J. Mattson, K. M. Pham, D. J. Leuck, K. A. Cowen, J. Org. Chem. 1990, 55, 2552.<br />

[40] M. T. Reetz, Organotitanium reagents in organic syn<strong>the</strong>sis; Springer-Verlag: Berlin,<br />

1986; p 107.<br />

[41] a) B.T. Cho, S. K. Kang, Tetrahedron 2005, 61, 5725 ; b) S. Sato, T. Sakamoto, E.<br />

Miyazawa, Y. Kikugawa, Tetrahedron 2004, 60, 7899.<br />

[42] a) T. Suwa, E. Sugiyama, I. Shibata, A. Baba, Syn<strong>the</strong>sis 2000, 789; b) T. Suwa, E.<br />

Sugiyama, I. Shibata, A. Baba, Synlett 2000, 556.<br />

[43] H. -U. Blaser, H. -P. Buser, H. -P. Jalett, B. Pugin, F. Spindler, Synlett 1999, 867.<br />

[44] Y. Chi, Y.-G. Zhou, X. Zhang, J. Org. Chem. 2003, 68, 4120.<br />

[45] L. R. Pérez, F. J. P.-Flores, P. Sharma, L. Velasco, A. Cabrera Org. Lett. 2009, 11, 265.<br />

[46] C. Li, B. V.-Marcos, J. Xiao, J. Am. Chem. Soc. 2009, 131, 6967.<br />

[47] G. Borg, D. A. Cogan, J. A. Ellman, Tetrahedron Lett. 1999, 40, 6709.<br />

[48] J. Tanuwidjaja, H. M. Peltier, J. A. Ellman, J. Org. Chem. 2007, 72, 626.<br />

[49] G. Du<strong>the</strong>uil, S. C. -Bonnaire, X. Pannecoucke, Angew. Chem. Int. Ed. 2007, 46, 1290.<br />

[50] A. Alexakis, S. Gille, F. Prian, S. Rosset, K. Ditrich, Tetrahedron Lett. 2004, 45, 1449.<br />

[51] T. C. Nugent, R. Seemeyer, WO2004035575, 2004.<br />

[52] a) T. C. Nugent, A. K. Ghosh, V. N. Wakchaure, R. R. Mohanty, Adv. Synth. Catal.<br />

2006, 348, 1289; b) T. C. Nugent, <strong>Chiral</strong> Amine Syn<strong>the</strong>sis - Strategies, Examples, and<br />

Limitations. In Process Chemistry in <strong>the</strong> Pharmaceutical Industry, Second Edition:<br />

Challenges in an Ever-Changing Climate, T. F. Braish, K. Gadamasetti Eds.; CRC Press-<br />

Taylor and Francis Group: New York, 2008; c) T. C. Nugent, A. K. Ghosh, V. N.<br />

Wakchaure, R. R. Mohanty, WO2006030017, 2006; d) T. C. Nugent, V. N. Wakchaure, A. K.<br />

Ghosh, R. R. Mohanty, Org. lett. 2005, 7, 4967; e) T. C. Nugent, A. K. Ghosh, Eur. J. Org.<br />

Chem. 2007, 3863.<br />

[53] R.Noyori, Adv. Synth. Catal. 2003, 345, 15<br />

[54] M. Kitamura, D. Lee, S. Hayashi, S. Tanaka, M. Yoshimura, J. Org. Chem. 2002, 7,<br />

8685<br />

78


[55] a) R. Kadyrov, T. H. Riermeier, Angew Chem., Int. Ed. 2003, 42, 5472; b) A. Börner, U.<br />

Dingerdissen R. Kadyrov, T. H. Riermeier, V. Tararov, US2004267051, 2004; c) T. H.<br />

Riermeier, K.-J. Haack, U. Dingerdissen, A. Börner, V. Tararov, R. Kadyrov, US6884887,<br />

2005.<br />

[56] a) T. Ohkuma, M. Kitamura, R. Noyori, Asymmetric hydrogenation. In: Ojima I (ed)<br />

Catalytic asymmetric syn<strong>the</strong>sis, 2nd edn. Wiley-VCH, New York, 2000; b)T. Ohkuma, R.<br />

Noyori Hydrogenation <strong>of</strong> imino groups. In: Jacobsen EN, Pfaltz A, Yamamoto H (eds)<br />

Comprehensive asymmetric catalysis, suppl 1. Springer, New York, 2004; c) V. I. Taratov, A.<br />

Börner A, Synlett 2005, 203.<br />

[57] B. D. Sanwal, M. W. Zink, Arch Biochem Biophys, 1961, 94, 430.<br />

[58] T. Itoh, K. Nagata, M. Miyazaki, H. Ishikawa, A. Kurihara, A. Ohsawa Tetrahedron<br />

2004, 60, 6649.<br />

[59] S. H<strong>of</strong>fmann, A. M. Seayad, B. List, Angew. Chem. Int. Ed. 2005, 44, 7424.<br />

[60] a) Noyori, M. Tokunaga, M. Kitamura, Bull Chem Soc Jpn 1995, 68, 36; b) R. S. Ward,<br />

Tetrahedron Asymmetry 1995, 6, 1475; c) H. Stecher, K. Faber, Syn<strong>the</strong>sis 1997, 1; d) H.<br />

Perllissier, Tetrahedron 2003, 59, 8291; e) F.F. Huerta, A. B. E. Minidis, J. E. Bäckvall,<br />

Chem Soc Rev 2001, 30, 321; f) S. Caddick , K. Jenkins, Chem Soc Rev 1996, 25, 447.<br />

[61] S. H<strong>of</strong>fmann, M. Nicoletti, B. List, J. Am. Chem. Soc. 2006, 128, 13074.<br />

[62] J. Zhou, B. List, J. Am. Chem. Soc. 2007, 129, 7498.<br />

[63] D. Menche, F. Arikan, Synlett 2006, 841.<br />

[64] R. I. Storer, D. E. Carrera, Y. Ni, D. W. C. MacMillan, J. Am. Chem. Soc. 2006, 128,<br />

84.<br />

[65] a) P. Anastas, J. C. Warner, Green Chemistry: Theory and Practice, Ox<strong>for</strong>d University<br />

Press, Ox<strong>for</strong>d, 1998; b) P.T. Anastas, M.M. Kirchh<strong>of</strong>f, Acc. Chem. Res. 2002, 35, 686; c)<br />

P.T. Anastas, L.G. Heine, T. C. Williamson, Green Chemical Syn<strong>the</strong>ses and Processes,<br />

American Chemical Society, Washington DC, 2000; c) P.T. Anastas, C. A. Farris, Benign by<br />

Design: Alternative Syn<strong>the</strong>tic Design <strong>for</strong> Pollution Prevention, ACS Symp. Ser. nr. 577,<br />

American Chemical Society, Washington DC, 1994; d) J.H. Clark, D.J. Macquarrie,<br />

Handbook <strong>of</strong> Green Chemistry and Technology, Blackwell, Abingdon, 2002; e) A. S.<br />

Matlack, Introduction to Green Chemistry, Marcel Dekker, New York, 2001; f) M. Lancaster,<br />

Green Chemistry: An Introductory Text, Royal Society <strong>of</strong> Chemistry, Cambridge, 2002; g)<br />

J.H. Clark (Ed.), The Chemistry <strong>of</strong> Waste Minimization, Blackie, London, 1995.<br />

[66] R. A. Sheldon, C. R. Acad. Sci. Paris, IIc, Chimie/Chemistry, 2000, 3, 541.<br />

79


[67] a) R. A. Sheldon, Chemtech 1994, 38; b) R. A. Sheldon, J. Chem. Technol. Biotechnol.<br />

1997, 68, 381; c) R. A. Sheldon, J. Mol. Catal. A: Chemical 1996, 107, 75; d) R. A. Sheldon,<br />

Pure Appl. Chem. 2000, 72, 1233.<br />

[68] a) D. A. Cogan, G. Liu, J. Ellman, Tetrahedron Lett. 1999, 55, 8883; b) D. A. Cogan, J.<br />

A. Ellman, J. Am. Chem. Soc. 1999, 121, 268; c) L. C. Akullian, M. L. Snapper, A. H.<br />

Hoveyda, Angew. Chem., Int. Ed. 2003, 42, 4244; d) L. C. Akullian, J. R. Porter, J. F.<br />

Traverse, M. L. Snapper, A. H. Hoveyda, Adv. Synth. Catal. 2005, 347, 417; e) A. Côté, A.<br />

B. Charette, J. Org. Chem. 2005, 70, 10864.<br />

80


Chapter 4<br />

Drugs and Reductive Amination<br />

4.1 Reductive Amination in <strong>the</strong> Syn<strong>the</strong>sis <strong>of</strong> Drugs and Natural Products:<br />

Different natural products and pharmaceutical drugs contain amino group as an important<br />

part <strong>of</strong> <strong>the</strong>ir structure. Of course some drugs are still sold as racemic compounds but <strong>the</strong> latest<br />

trend over <strong>the</strong> past three decades is to design, develop and market new drug entities as a<br />

single isomeric <strong>for</strong>m. Amino group can be introduced in <strong>the</strong> drug entity through different<br />

strategies one <strong>of</strong> <strong>the</strong>se strategies is reductive amination. Older reports describing <strong>the</strong><br />

syn<strong>the</strong>sis <strong>of</strong> natural products and pharmaceutical drugs utilizing reductive amination did not<br />

involve any source <strong>of</strong> chirality resulting in <strong>the</strong> syn<strong>the</strong>sis <strong>of</strong> racemic product. Recent<br />

literatures focused on utilizing <strong>the</strong> asymmetric versions <strong>of</strong> reductive amination <strong>for</strong> <strong>the</strong><br />

syn<strong>the</strong>sis <strong>of</strong> enantiopure compounds. I will try to give a brief overview on <strong>the</strong> potential<br />

applications <strong>of</strong> this powerful methodology in <strong>the</strong> syn<strong>the</strong>sis <strong>of</strong> natural products and<br />

pharmaceutical drugs. We will try also to show <strong>the</strong> relevance <strong>of</strong> our developed strategy <strong>for</strong><br />

<strong>the</strong> efficient syn<strong>the</strong>sis <strong>of</strong> <strong>the</strong>se entities.<br />

4.1.1. Syn<strong>the</strong>sis <strong>of</strong> Delavirdine:<br />

This compound is a member <strong>of</strong> nonnucleoside HIV-1 reverse transcriptase inhibitors. [1] This<br />

class <strong>of</strong> compounds was discovered by Upjohn scientists from a computer-directed<br />

dissimilarity analysis <strong>of</strong> <strong>the</strong> Pharmacia & Upjohn chemical library to select compounds <strong>for</strong><br />

screening against HIV-1 reverse transcriptase. Syn<strong>the</strong>sis <strong>of</strong> this compound starts with <strong>the</strong><br />

addition <strong>of</strong> piperazine (20) to chloropyridine (21). The nitro group is reduced to <strong>the</strong> amino<br />

group and <strong>the</strong> resultant amine undergoes reductive amination with acetone to provide<br />

pyridylpiperazine (23). Coupling <strong>of</strong> (23) with 6-nitroindole-2-carboxylic acid (24) is<br />

accomplished using ei<strong>the</strong>r 1-ethyl-3-(dimethylamino) propylcarbodiimide (EDC) or 1,10-<br />

81


carbonyldiimidazole (CDI) to give amide (25). The nitro group is reduced under<br />

hydrogenation conditions using Pd-C. The resulting amine is <strong>the</strong>n sulfonylated with<br />

methanesulfonyl chloride to provide delavirdine, which is <strong>the</strong>n trans<strong>for</strong>med to delavirdine<br />

H<br />

N<br />

Cl<br />

N N<br />

H<br />

20 21<br />

mesylate (3).<br />

NO 2 1) CH 2 Cl 2 96%<br />

2) (Boc) 2 O96%<br />

BocN<br />

O 2 N<br />

N<br />

N<br />

22<br />

1) Pd/C, H 2 ,78%<br />

2) acetone<br />

NaCNBH 3 91%<br />

O 2 N<br />

24<br />

N<br />

H<br />

HN<br />

COOH<br />

N<br />

HN<br />

23<br />

N<br />

CDI or EDC<br />

74%<br />

S<br />

H<br />

N<br />

HN<br />

O O CH 3 SO 3 H<br />

N O<br />

H<br />

3<br />

N<br />

N<br />

N<br />

1) H 2 , Pd/C, 66%<br />

2) CH 3 SO 2 Cl,<br />

pyr, CH 2 Cl 2<br />

3) CH 3 SO 3 H<br />

O 2 N<br />

N<br />

H<br />

HN<br />

N<br />

O<br />

25<br />

N<br />

N<br />

Scheme 4.1. Syn<strong>the</strong>sis <strong>of</strong> Delavirdine.<br />

4.1.2. Syn<strong>the</strong>sis <strong>of</strong> Muraglitazar:<br />

Muraglitazar is developed to treat hyperglycemia and dyslipidemia through decrease<br />

triglycerides and increase HDL cholesterol with minimal effects on LDL cholesterol. [2]<br />

Several syn<strong>the</strong>ses have been developed <strong>for</strong> its efficient preparation. Syn<strong>the</strong>sis starts with <strong>the</strong><br />

alkylation <strong>of</strong> 4-hydroxybenzaldehyde with phenyloxazolemesylate (23), which can be easily<br />

syn<strong>the</strong>sized from commercially available alcohol (22), resulting in <strong>the</strong> aldehyde (24)<br />

syn<strong>the</strong>sis. The aldehyde is <strong>the</strong>n treated with glycine methyl ester under reductive amination<br />

conditions to provide secondary amine (25) in an excellent yield. Reaction <strong>of</strong> amine (25) with<br />

4-methoxyphenyl chloro<strong>for</strong>mate followed by hydrolysis <strong>of</strong> <strong>the</strong> methyl ester af<strong>for</strong>ded<br />

Muraglitazar in 94% yield.<br />

82


Scheme 4.2. Syn<strong>the</strong>sis <strong>of</strong> Muraglitazar.<br />

4.1.3. Syn<strong>the</strong>sis <strong>of</strong> Amphetamine:<br />

Amphetamines syn<strong>the</strong>sis is one <strong>of</strong> <strong>the</strong> well known classical examples <strong>of</strong> drugs syn<strong>the</strong>sised<br />

utilising reductive amination. [2] O<strong>the</strong>r methodologies were also developed <strong>for</strong> <strong>the</strong>re syn<strong>the</strong>sis<br />

as <strong>the</strong> direct displacement <strong>of</strong> a leaving group by an amine, nitro alkane addition followed by<br />

reduction <strong>of</strong> <strong>the</strong> nitro group and metal-promoted amination <strong>of</strong> an unsaturated carbon<br />

compound. The reductive amination <strong>of</strong> methyl benzyl ketone was one <strong>of</strong> <strong>the</strong> earliest<br />

strategies used in amphetamine syn<strong>the</strong>sis. Despite being developed in <strong>the</strong> thirties <strong>of</strong> last<br />

century it continues to be one <strong>the</strong> best developed methodologies because <strong>of</strong> its simple<br />

elegance and <strong>the</strong> availability <strong>of</strong> cheap starting materials.<br />

According to <strong>the</strong> developed methodology, oxime is <strong>for</strong>med as a mixture <strong>of</strong> isomers upon<br />

exposure <strong>of</strong> <strong>the</strong> ketone to hydroxylamine hydrochloride under mildly basic conditions.<br />

83


Reduction <strong>of</strong> <strong>the</strong> oxime can be accomplished using a variety <strong>of</strong> reducing agents. The initial<br />

report employed sodium in methanol <strong>for</strong> converting <strong>the</strong> oxime to <strong>the</strong> target amphetamine.<br />

O<strong>the</strong>r modifications <strong>of</strong> this approach have been described in recent literatures.<br />

Scheme 4.3.Syn<strong>the</strong>sis <strong>of</strong> Amphetamine:<br />

The asymmetric syn<strong>the</strong>sis <strong>of</strong> amphetamines was developed in <strong>the</strong> seventies <strong>of</strong> <strong>the</strong> last<br />

century. The syn<strong>the</strong>sis starts with methyl benzyl which is reductively aminated using readily<br />

available chiral α-methyl benzyl amine producing <strong>the</strong> imine intermediate and <strong>the</strong> syn<strong>the</strong>sis is<br />

driven to completion by removing H 2 O with Dean–Stark trap. The resulting imine was<br />

reduced with Raney nickel. The product was isolated and crystallized as HCl salt which is<br />

<strong>the</strong>n hydrogenolyzed with Pd-C producing <strong>the</strong> primary amine in high overall optical purity.<br />

4.1.4. Syn<strong>the</strong>sis <strong>of</strong> Sertraline:<br />

Sertraline is an anti-depressant drug that affects serotonin levels in <strong>the</strong> brain. Initially <strong>the</strong><br />

active isomer was not known when both diastereoisomers were prepared through an<br />

unselective route. [3,4] Syn<strong>the</strong>sis starts with Friedel-Crafts reaction between 1,2-<br />

dichlorobenzene and succinic anhydride <strong>for</strong>ming <strong>the</strong> starting material presented in <strong>the</strong><br />

following scheme.<br />

Scheme 4.4. Syn<strong>the</strong>sis <strong>of</strong> Sertraline:<br />

84


Studies showed that <strong>the</strong> active isomer is <strong>the</strong> syn diastereomer. Reductive amination <strong>of</strong> <strong>the</strong><br />

ketone in <strong>the</strong> final step could be controlled to give 70% syn diastereomer.<br />

4.1.5. Syn<strong>the</strong>sis <strong>of</strong> Emitine:<br />

Emitine is a natural compound which is extracted from ipecacuanha plant (Brazilian root)<br />

and used as <strong>the</strong> primary drug <strong>for</strong> treating amebiasis, leishaniasis, and dysentery. [5,6] It has a<br />

direct amebicidal effect against trophozoites E. histolytica in tissues, and it is not active<br />

against cysts in ei<strong>the</strong>r <strong>the</strong> lumen or intestinal walls, or in o<strong>the</strong>r organs. It blocks protein<br />

syn<strong>the</strong>sis in eukaryotic (but not in prokaryotic) cells. Protein syn<strong>the</strong>sis is inhibited in parasite<br />

and mammalian cells, but not in bacteria. Several syn<strong>the</strong>tic routes have been developed <strong>for</strong> its<br />

syn<strong>the</strong>sis.<br />

Syn<strong>the</strong>sis starts with reductive amination <strong>of</strong> 2-(3,4- imethoxyphenyl)ethylamine and ethyl<br />

ester <strong>of</strong> β–(α-cyano)propylglutaric acid, followed with intramolecular cyclization. The lactam<br />

produced is reacted with phosphorus oxychloride leading to heterocyclization into <strong>the</strong><br />

derivative <strong>of</strong> benzoquinolizine. Subsequent reaction <strong>of</strong> <strong>the</strong> product with homoveratrylamine<br />

produces <strong>the</strong> corresponding amide. Upon reaction with phosphorus oxychloride, this<br />

compound cyclizes to an isoquinoline derivative and <strong>the</strong> pyridine ring is <strong>the</strong>n hydrogenated to<br />

a racemic mixture <strong>of</strong> <strong>the</strong> products, from which <strong>the</strong> desired emetine is isolated.<br />

85


H 3 CO<br />

H 3 CO<br />

H 3 CO<br />

NC C 2 H 5 H2 /PtO<br />

NH 2<br />

O O<br />

C 2 H 5 O OC 2 H 5<br />

H 3 CO<br />

C 2 H 5 OOC<br />

HN<br />

C 2 H 5<br />

COOC 2 H 5<br />

H 3 CO<br />

H 3 CO<br />

N<br />

C 2 H 5<br />

POCl 3<br />

Adams Catalyst<br />

H 3 CO<br />

H 3 CO<br />

O<br />

N<br />

COOC 2 H 5<br />

C 2 H 5<br />

1.<br />

H 3 CO<br />

H 3 CO<br />

NH 2 2.POCl 3<br />

3.H 2 /PtO<br />

COOC 2 H 5<br />

H 3 CO<br />

H 3 CO<br />

H 3 CO<br />

N<br />

C 2 H 5<br />

CH 2<br />

NH<br />

H 3 CO<br />

Scheme 4.5. Syn<strong>the</strong>sis <strong>of</strong> Emitine<br />

4.1.6. Syn<strong>the</strong>sis <strong>of</strong> Taltobulin:<br />

Taltobulin is an anticancer drug which interferes with tubulin function inhibiting <strong>the</strong><br />

<strong>for</strong>mation <strong>of</strong> microtubules that <strong>for</strong>m <strong>the</strong> microskeleton <strong>of</strong> cells. This process has provided<br />

some valuable antitumor activity. [7]<br />

The syn<strong>the</strong>tic strategy depends on <strong>the</strong> separate syn<strong>the</strong>sis <strong>of</strong> two intermediates and combining<br />

<strong>the</strong>m at <strong>the</strong> final step. One arm <strong>of</strong> <strong>the</strong> syn<strong>the</strong>sis begins with <strong>the</strong> construction <strong>of</strong> acrylatecontaining<br />

moiety through condensation <strong>of</strong> <strong>the</strong> t-BOC protected α-aminoaldehyde derived<br />

from valine with <strong>the</strong> arbethoxymethylene phosporane resulting in <strong>the</strong> amino ester. Removal<br />

<strong>of</strong> <strong>the</strong> protecting group is carried out under acidic condition producing <strong>the</strong> free amine. The<br />

o<strong>the</strong>r arm <strong>of</strong> this syn<strong>the</strong>tic strategy starts with condensation <strong>of</strong> that tertiary butyl-substituted<br />

86


aminoacid producing <strong>the</strong> protected amide which can be deprotected under acidic condition.<br />

The second arm <strong>of</strong> this syn<strong>the</strong>tic strategy starts with <strong>the</strong> addition <strong>of</strong> a pair <strong>of</strong> methyl groups to<br />

<strong>the</strong> benzylic position <strong>of</strong> pyruvate through addition <strong>of</strong> methyl iodide to ketoacid in <strong>the</strong><br />

presence <strong>of</strong> hydroxide. The addition <strong>of</strong> methylamine and diborane results in <strong>the</strong> reductive<br />

amination <strong>of</strong> <strong>the</strong> carbonyl group, and thus <strong>for</strong>mation <strong>of</strong> α-aminoacid as a mixture <strong>of</strong> <strong>the</strong> two<br />

isomers. Condensation <strong>of</strong> this moiety with dipeptide <strong>for</strong>med previously under peptide<br />

<strong>for</strong>ming condition resulted in <strong>the</strong> <strong>for</strong>mation <strong>of</strong> amide product which is separated by column<br />

chromatography af<strong>for</strong>ding <strong>the</strong> desired isomer <strong>of</strong> taltobulin.<br />

t-BOC<br />

N<br />

CHO<br />

N<br />

(C 6 H 5 ) 3 P CO 2 C 2 H 5 t-Boc<br />

CO 2 C 2 H 5 H +<br />

HN CO2 C 2 H 5<br />

O<br />

CO 2 H<br />

CH 3 I<br />

NaOH<br />

O<br />

CO 2 H<br />

CO 2 H<br />

NHCH 3<br />

t-BOC<br />

N<br />

H<br />

CO 2 H<br />

DCC<br />

O<br />

N<br />

H<br />

NHCH 3<br />

O<br />

N<br />

CO 2 R<br />

R<br />

N<br />

H<br />

O<br />

N CO 2 C 2 H 5<br />

Scheme 4.6. Syn<strong>the</strong>sis <strong>of</strong> Taltobulin<br />

4.1.7. Syn<strong>the</strong>sis <strong>of</strong> Perzinfote:<br />

Perzinfote is a nonaddictive opiate alternative which is currently used to treat chronic pain. [7]<br />

<strong>Preparation</strong> starts with <strong>the</strong> reductive amination <strong>of</strong> <strong>the</strong> acetaldehyde derivative with<br />

monocarbobenzyloxy ethylenediamine leading to <strong>the</strong> disubstituted ethylenediamine (116).<br />

The amine is reacted with <strong>the</strong> commercially available cyclobutenedione derivative (117)<br />

resulting in <strong>the</strong> replacement <strong>of</strong> one <strong>of</strong> <strong>the</strong> ethoxy groups in (117) by <strong>the</strong> free amino group in<br />

(116) to af<strong>for</strong>d <strong>the</strong> coupled product (118). Transfer hydrogenation <strong>of</strong> <strong>the</strong> (118) with 1,4-<br />

cyclohexadiene/Pd leads to <strong>the</strong> loss <strong>of</strong> <strong>the</strong> carbobenzoxy group and <strong>the</strong> <strong>for</strong>mation <strong>of</strong> <strong>the</strong><br />

transient primary amine (119) which is <strong>the</strong>n cyclised to <strong>for</strong>m eight-membered ring (120).<br />

87


Removal <strong>of</strong> <strong>the</strong> ethyl group on <strong>the</strong> phosphorous is done by treating with trimethylsilyl<br />

bromide resulting in <strong>the</strong> <strong>for</strong>mation <strong>of</strong> free phosphonic acid and thus perzinfotel (121).<br />

NaCNBH 3 HN NHCO2 CH 2 C 5 H 6<br />

H 2 N NHCO 2 CH 2 C 5 H 6<br />

C 2 H 5 O<br />

115<br />

CHO<br />

O C<br />

P<br />

2 H 5 O<br />

O<br />

P<br />

OC 2 H 5<br />

OC 2 H 5<br />

116<br />

O<br />

O<br />

OC 2 H 5<br />

OC 2 H 5<br />

117<br />

O<br />

OC 2 H 5<br />

O<br />

OC 2 H 5<br />

O<br />

N NHCO 2 CH 2 C 5 H 6<br />

Pt<br />

O<br />

N<br />

NH 2<br />

C 2 H 5 O<br />

P O<br />

OC 2 H 5<br />

118<br />

C 2 H 5 O<br />

119<br />

P<br />

O<br />

OC 2 H 5<br />

O<br />

O<br />

HN<br />

N<br />

(CH 3 ) 3 SiBr<br />

O<br />

O<br />

HN<br />

N<br />

HO<br />

P<br />

O<br />

OH<br />

C 2 H 5 O<br />

P<br />

O<br />

OC 2 H 5<br />

121<br />

120<br />

Scheme 4.7.Syn<strong>the</strong>sis <strong>of</strong> Perzinfote:<br />

4.1.8. Syn<strong>the</strong>sis <strong>of</strong> Namindinil:<br />

Namindinil is used to treat hair loss in males. It has a vasodilator action improving blood<br />

circulation in hair follicle capillaries. [7] The convergent syn<strong>the</strong>sis <strong>of</strong> this drug involves<br />

preparation <strong>of</strong> <strong>the</strong> complex alkyl group as a single enantiomer. The process involves <strong>the</strong><br />

preparation <strong>of</strong> imine utilizing p-toluene sulfonic acid. The imine is <strong>the</strong>n reduced with borane-<br />

THF complex yielding <strong>the</strong> secondary amine as a mixture <strong>of</strong> diastereomers. The two<br />

diastereomers were separated by column chromatography. Hydrogenolysis <strong>of</strong> <strong>the</strong> chiral<br />

auxiliary leads to <strong>the</strong> <strong>for</strong>mation <strong>of</strong> <strong>the</strong> primary amine. The primary amine is added to thiourea<br />

derivative <strong>for</strong>ming <strong>the</strong> required compound.<br />

88


O<br />

NH 2<br />

1. p-TsOH/toluene<br />

2. BH 3 -THF/THF<br />

HN<br />

1.Column<br />

Chromatograpgy<br />

2.10 mol% Pd-C/EtOH<br />

NH 2<br />

N<br />

OH<br />

N<br />

N<br />

NH 2<br />

S<br />

C<br />

N<br />

NaCN<br />

NC<br />

S<br />

NH<br />

NH<br />

NC<br />

N<br />

H<br />

N<br />

NH<br />

NH 2<br />

CN<br />

CN<br />

CN<br />

Scheme 4.8. Syn<strong>the</strong>sis <strong>of</strong> Namindinil.<br />

4.1.9. Syn<strong>the</strong>sis <strong>of</strong> Ezlopipant:<br />

Ezlopipant is an antiemetic drug which is prescribed <strong>for</strong> severe nausea and vomiting<br />

associated with chemo<strong>the</strong>rapy. [7] <strong>Preparation</strong> starts with <strong>the</strong> condensation <strong>of</strong> acetonitrile with<br />

ester derivative <strong>of</strong> piperidine. Nitrile group is converted to <strong>the</strong> corresponding acid under acid<br />

hydrolysis. The carboxylic acid undergoes spontaneous decarboxylation <strong>for</strong>ming <strong>the</strong><br />

corresponding ketone which is reacted with bromine yielding bromoketone. This unstable<br />

intermediate undergoes spontaneous internal displacement <strong>for</strong>ming quinuclidine (172) as a<br />

quaternary salt. Debenzylation using palladium leads to <strong>the</strong> <strong>for</strong>mation <strong>of</strong> quinuclidone.<br />

Reductive amination 2-methoxy-4-isopropylbenzylamine (174) af<strong>for</strong>ds ezlopipant (175).<br />

89


CO 2 C 2 H 5<br />

N<br />

CN<br />

base<br />

CN<br />

O H + N<br />

O<br />

N<br />

167 168<br />

169<br />

170<br />

Br 2<br />

N<br />

O<br />

H 2<br />

N<br />

O<br />

O<br />

Br<br />

173<br />

N<br />

172<br />

OCH 3<br />

171<br />

174<br />

NH 2<br />

N<br />

NH<br />

OCH 3<br />

175<br />

Scheme 4.9. Syn<strong>the</strong>sis <strong>of</strong> Ezlopipant.<br />

4.1.10. Syn<strong>the</strong>sis <strong>of</strong> Monomorine:<br />

Monomorine alkaloid is a trail pheromone <strong>of</strong> <strong>the</strong> widespread pharaoh’s ant Monomorium<br />

pharaonis. [8] Syn<strong>the</strong>sis starts by treating amine (125) with 20% titanocene trichloride <strong>for</strong>ming<br />

<strong>the</strong> imidotitanium complex (126). This compound <strong>the</strong>n underwent a [2+2]-cycloaddition with<br />

<strong>the</strong> alkyne to af<strong>for</strong>d intermediate (127). Ring opening and subsequent meta<strong>the</strong>sis. Imine was<br />

reduced with diisobutylaluminum hydride providing amine intermediate followed by<br />

deprotection and intramolecular reductive amination resulting in <strong>the</strong> <strong>for</strong>mation <strong>of</strong> <strong>the</strong> alkaloid<br />

monomorine.<br />

90


O<br />

O<br />

125<br />

NH 2<br />

a<br />

93%<br />

O<br />

O<br />

126<br />

H<br />

Bu<br />

N<br />

Ti<br />

Cp<br />

Cl<br />

H<br />

O<br />

O<br />

Cp Ti<br />

Cl<br />

127<br />

N<br />

Bu<br />

O<br />

O<br />

N<br />

b<br />

95% O O<br />

128 Bu<br />

129<br />

HN<br />

Bu<br />

c,d,e<br />

72% N<br />

(a) Et 3 N, CpTiCl 3 (20% mol); (b) DIBALH; (c) HCl; (d) K 2 CO 3 ; (e) NaCNBH 3<br />

Scheme 4.10. Syn<strong>the</strong>sis <strong>of</strong> Monomorine:<br />

4.1.11. Syn<strong>the</strong>sis <strong>of</strong> Ontazolast:<br />

Ontazolast is an antiasthmatic drug which acts as leukotrienes syn<strong>the</strong>sis inhibitors. [8]<br />

Syn<strong>the</strong>sis starts with <strong>the</strong> condensation <strong>of</strong> pyridine 2-aldehyde with <strong>the</strong> cyclohexylmethyl<br />

magnesium bromide <strong>for</strong>ming carbinol which is oxidized with manganese dioxide to af<strong>for</strong>d<br />

<strong>the</strong> ketone. Reductive amination utilizing ammonium <strong>for</strong>mate/<strong>for</strong>mic acid system converts<br />

<strong>the</strong> carbonyl group into <strong>the</strong> primary amine. The o<strong>the</strong>r half <strong>of</strong> <strong>the</strong> final product is <strong>for</strong>med<br />

through <strong>the</strong> reaction <strong>of</strong> aminotoluol with carbon disulfide in <strong>the</strong> presence <strong>of</strong> a base proceeds<br />

to <strong>the</strong> addition product, benzoxazole. The thiol at <strong>the</strong> 2 position is <strong>the</strong>n replaced by halogen<br />

by reaction with phosphorus oxychloride. Combing this part with <strong>the</strong> primary amine leads to<br />

<strong>the</strong> <strong>for</strong>mation <strong>of</strong> alkylated product <strong>of</strong> ontazolast.<br />

MgBr<br />

OHC<br />

N<br />

1.<br />

2. MnO 2<br />

N<br />

O<br />

HCO 2 NH 4<br />

HCO 2 H<br />

NH 2<br />

N<br />

NH 2<br />

OH<br />

S C S<br />

NaOH<br />

N<br />

O<br />

SH<br />

POCl 3<br />

N<br />

O<br />

Cl<br />

Scheme 4.11. Syn<strong>the</strong>sis <strong>of</strong> Ontazolast<br />

HN<br />

N<br />

N<br />

91<br />

O


4.1.12. Syn<strong>the</strong>sis <strong>of</strong> Pamaquine:<br />

Pamaquine is an antimalarial quinoline derivative. It is very effective against <strong>the</strong> erythrocytic<br />

stages <strong>of</strong> all four human malarias. [8] The syn<strong>the</strong>sis <strong>of</strong> pamaquine is achieved through starting<br />

quinoline (3) which is syn<strong>the</strong>sized through reaction <strong>of</strong> substituted aniline (1) with glycerol in<br />

sulfuric acid and Nitrobenzene. The resulting nitro group is <strong>the</strong>n reduced by molecular<br />

hydrogen giving <strong>the</strong> free amine which is reductively aminated <strong>for</strong>ming pamaquine<br />

O<br />

O<br />

OH<br />

OH<br />

H 2 SO 4<br />

HO<br />

OH<br />

OH<br />

H 3 CO<br />

NO 2<br />

NH 2<br />

NO 2<br />

N<br />

H N<br />

H 2<br />

NO 2<br />

NH 2<br />

H 3 CO<br />

CHO<br />

H 2 SO 4 H 3 CO<br />

H 3 CO<br />

1 2 3 4<br />

Et 2 N<br />

O<br />

N<br />

H 3 CO<br />

HN<br />

N<br />

NEt 2<br />

Scheme 4.12. Syn<strong>the</strong>sis <strong>of</strong> Pamaquine<br />

5<br />

4.1.13. Syn<strong>the</strong>sis <strong>of</strong> Torcetrapib:<br />

Torcetrapib is a cholesterol lowering agent. [8] Dietary cholesterol needs be esterified in order<br />

to be absorbed from <strong>the</strong> gut. The enzyme, cholesteryl ester transfer protein (CETP), <strong>the</strong>n<br />

completes <strong>the</strong> absorption <strong>of</strong> cholesterol. Torcetrapib inhibit this enzyme helping to lower<br />

LDL (Low density lipoproteins) cholesterol and VLDL (Very Low Density Lipoproteins)<br />

cholesterol. It also increases <strong>the</strong> high density–good- lipoprotein cholesterol. <strong>Preparation</strong> <strong>of</strong><br />

torcetrapib starts with <strong>the</strong> reaction <strong>of</strong> <strong>the</strong> trifluoromethylaniline (1) with propanal in <strong>the</strong><br />

presence <strong>of</strong> benzotriazole (2) which af<strong>for</strong>ds aminal (3). The aminal is condensed with vinyl<br />

carbamate <strong>for</strong>ming tetrahydroquinoline ring. The nitrogen group on <strong>the</strong> ring is protected as<br />

92


ethyl carbamate by acylation with ethyl chloro<strong>for</strong>mate. Benzyl carbamate group on <strong>the</strong><br />

nitrogen at 4 position is hydrogenolyzed using ammonium <strong>for</strong>mate over palladium <strong>for</strong>ming<br />

<strong>the</strong> primary amine. The chiral amine is resolved as debenzyl tartarate salt to af<strong>for</strong>d (2R, 4S)<br />

isomer. Bis-trifuoromethyl benzaldehyde and sodiumtriacetoxy borohydride are used <strong>for</strong> <strong>the</strong><br />

reductive amination <strong>of</strong> <strong>the</strong> pure isomeric amine which is <strong>the</strong>n acylated with chloro<strong>for</strong>mate<br />

<strong>for</strong>ming <strong>the</strong> final product.<br />

F 3 C<br />

F 3 C<br />

5<br />

1<br />

HN<br />

N<br />

H<br />

O<br />

NH 2<br />

O<br />

CH 2 C 5 H 5<br />

CHO<br />

ClCO 2 C 2 H 5<br />

CF 3<br />

NH 2 1.<br />

F 3 C<br />

F 3 C CHO<br />

N<br />

2. CH 3 OCOCl<br />

C 2 H 5<br />

O O<br />

7<br />

2<br />

N<br />

N<br />

N<br />

H<br />

F 3 C<br />

O<br />

N<br />

N<br />

N<br />

N<br />

H<br />

CH 2 C 5 H 5<br />

HN O<br />

1. HCO - +<br />

2 NH 4<br />

F 3 C<br />

Pd<br />

N<br />

2.resolve<br />

C 2 H 5<br />

O O<br />

6<br />

O<br />

F 3 C<br />

O<br />

C 2 H 5<br />

O<br />

8<br />

N<br />

N<br />

O<br />

3<br />

CF 3<br />

CF 3<br />

vinyl carbamate<br />

Scheme 4.13.Syn<strong>the</strong>sis <strong>of</strong> Torcetrapib:<br />

4.1.14. Syn<strong>the</strong>sis <strong>of</strong> Polyaminocholestanol derivatives:<br />

Polyaminocholestanol derivatives are used as potent antibiotics <strong>for</strong> highly resistant bacterial<br />

strains (superbugs). [9] The key step in <strong>the</strong>ir syn<strong>the</strong>sis is reductive amination. Different<br />

titanium sources were tested in different solvents. The highest des were obtained when<br />

Ti(O i Pr) 4 was used with MeOH and NaBH as hydride source. Different amine and diamaines<br />

were tested and <strong>the</strong> des were higher than 95% but <strong>the</strong> yields were poor to mediocre (6-77%).<br />

93


AcO<br />

H<br />

O<br />

1) Ti(O i Pr) 4 ,MeOH<br />

20 o C, 5-6 h<br />

2) NaBH 4 ,-78 o C,2 h<br />

RNH 2<br />

3) K 2 CO 3 , MeOH/THF (1:1)<br />

rt, 12 h<br />

AcO<br />

H<br />

NH 2 R<br />

Scheme 4.14. Syn<strong>the</strong>sis <strong>of</strong> Polyaminocholestanol Derivatives.<br />

4.1.15. Syn<strong>the</strong>sis <strong>of</strong> Piperazinylpropylisoxazoline Analogues:<br />

Piperazinylpropylisoxazoline analogues are potent ligands <strong>for</strong> dopamine receptors which<br />

have strong influence on <strong>the</strong> general psychological condition. [10] Syn<strong>the</strong>sis is accomplished<br />

through <strong>the</strong> reductive amination <strong>of</strong> enantiomerically pure (S) or (R) aldehyde (1) with<br />

different piprazine derived amines. The (R) isomer showed higher potent activity <strong>for</strong><br />

dopamine receptors.<br />

O<br />

H<br />

(S)-1<br />

O N<br />

R 1<br />

(S)-1<br />

or<br />

(R)-1<br />

N<br />

N<br />

N<br />

O N<br />

R 1<br />

R 2<br />

N<br />

H<br />

or<br />

R 2<br />

O N<br />

R 1<br />

R 2<br />

N<br />

N<br />

Scheme 4.15. Syn<strong>the</strong>sis <strong>of</strong> piperazinylpropylisoxazoline analogues.<br />

94


4.1.16. Syn<strong>the</strong>sis <strong>of</strong> Ritonavir and Lopinavir:<br />

Ritonavir and lopinavir are HIV-protease inhibitors. Their syn<strong>the</strong>sis depends on <strong>the</strong> syn<strong>the</strong>sis<br />

<strong>of</strong> chiral aminoalcohols. [11] Titanium isopropoxide with polymethylhydrosiloxane (hydride<br />

source) and p-anisidine were utilized <strong>for</strong> <strong>the</strong> syn<strong>the</strong>sis <strong>of</strong> aminoalcohols under reductive<br />

amination conditions. Aliphatic, cyclic, as well as aromatic and heteroaromatic<br />

hydroxyketones were tested showing good to high yields (76-89%) with good<br />

stereoselectivities (de 72-86%).<br />

R 1<br />

OH O<br />

R 2<br />

NR 3 H 2<br />

Ti(O i Pr) 4<br />

PMHS<br />

R 1<br />

O<br />

N<br />

R 2<br />

H + TiLn<br />

OH NR 3 H<br />

R 1<br />

R 3<br />

R 2<br />

S<br />

N<br />

CH 3<br />

N<br />

O<br />

H<br />

N<br />

O<br />

Ph<br />

NH<br />

OH<br />

Ph<br />

H<br />

N<br />

O<br />

O<br />

S<br />

N<br />

HN<br />

O<br />

N<br />

O<br />

Ph<br />

NH<br />

OH<br />

Ph<br />

H<br />

N<br />

O<br />

O<br />

Ritonavir<br />

Lopinavir<br />

Scheme 4.16. Syn<strong>the</strong>sis <strong>of</strong> Aminoalcohols <strong>the</strong> Core <strong>of</strong> Ritonavir and Lopinavir:<br />

4.1.17. Syn<strong>the</strong>sis <strong>of</strong> Tetrahydrocarbazoles:<br />

Tetrahydrocarbazoles is an efficient compound <strong>for</strong> <strong>the</strong> treatment <strong>of</strong> human papillomaviruses<br />

(HPVs). [12] HPV infection is considered <strong>the</strong> most common sexually transmitted disease<br />

throughout <strong>the</strong> world. There are over 5.5 million new cases <strong>of</strong> sexually transmitted HPV in<br />

<strong>the</strong> United States each year, with at least 20 million people currently infected. The chiral<br />

centre in <strong>the</strong> molecule is <strong>the</strong> α-chiral amine in which its syn<strong>the</strong>sis represents <strong>the</strong> key step in<br />

<strong>the</strong> syn<strong>the</strong>sis <strong>of</strong> this tetrahydrocarbazoles. In <strong>the</strong>ir initial attempts <strong>for</strong> building this chiral<br />

moiety <strong>the</strong>y tested chemical resolution. The best results were obtained utilizing dibenzoyl-Dtartaric<br />

acid leading to 86% ee with only 13% yield. This low yield encouraged <strong>the</strong>m to shift<br />

to <strong>the</strong> asymmetric syn<strong>the</strong>sis <strong>for</strong> building <strong>the</strong> chiral centre. Noyori catalyst was tested <strong>for</strong> <strong>the</strong><br />

95


eductive amination <strong>of</strong> <strong>the</strong> prochiral ketone resulting in 80% ee with 60% yield. They also<br />

tested reduction <strong>of</strong> isolated imine resulting in <strong>the</strong> same enantioselectivity and with<br />

unacceptable chemical purity. Higher ees are usually required <strong>for</strong> pharmaceutical<br />

development levels. Later <strong>the</strong>y tested <strong>the</strong> chiral auxiliary approach <strong>for</strong> <strong>the</strong> syn<strong>the</strong>sis <strong>of</strong> chiral<br />

amine moiety. They tested different derivatives <strong>of</strong> MBA and phenylglycinaol. They reported<br />

86% yield with 90% de which was improved by crystallization. The instability <strong>of</strong> substituents<br />

under hydrogenolysis standard conditions was ano<strong>the</strong>r challenge. BCl 3 and BBr 3 gave <strong>the</strong><br />

cleanest N-debenzylation without affecting <strong>the</strong> chloro substituent and tetrahydrocarbazole<br />

moiety.<br />

Cl<br />

HCO 2 NH 4 ,MeOH,60 o C Cl<br />

N<br />

H<br />

O<br />

N NH 2<br />

H<br />

80% ee, yield 60%<br />

NH 3<br />

(7N in MeOH)<br />

TsOH<br />

SO 2<br />

N<br />

Ru<br />

N<br />

H 2<br />

Cl<br />

Cl<br />

N<br />

H<br />

NH<br />

[RuCl 2 (benzene)] 2 Cl<br />

HCO 2 NH 4 ,MeOH,65 o C<br />

N<br />

H<br />

81% ee<br />

NH 2<br />

PPh 2<br />

Cl<br />

N<br />

H<br />

O<br />

p-TsOH, or conc.HCl<br />

toluene, reflux<br />

H 2 N<br />

X<br />

Cl<br />

PPh 2<br />

N<br />

H<br />

N<br />

Y<br />

X<br />

1. NaBH 4<br />

EtOH, -30 o CtoRT<br />

2. HCl<br />

Cl<br />

N<br />

H<br />

(R)<br />

HN<br />

Y<br />

X<br />

Cl<br />

Y<br />

Cl<br />

N<br />

H<br />

HN<br />

OMe<br />

1. BCl 3 ,DCM,0 o C<br />

2.<br />

Scheme 4.17. Syn<strong>the</strong>sis <strong>of</strong> Tetrahydrocarbazoles:<br />

N<br />

COOH,i-PrOH<br />

80-92%<br />

Cl<br />

N<br />

H<br />

N<br />

ee 99.2%<br />

NH 2<br />

COOH<br />

T3P (50% in EtOAc)<br />

i-Pr 2 NEt, DCM, 0 o C<br />

60-87%<br />

Pr<br />

T3P= O<br />

O<br />

P P<br />

O Pr<br />

O O P<br />

Pr<br />

O<br />

Cl<br />

N<br />

H<br />

HN<br />

O<br />

Tetrahydrocarbazoles<br />

ee >99.5%<br />

N<br />

96


4.2. Conclusion<br />

Different important pharmaceutical and natural products are prepared industrially utilizing<br />

reductive amination as a key step in <strong>the</strong>ir preparation. Reductive amination is <strong>the</strong> method <strong>of</strong><br />

choice <strong>for</strong> incorporating amino group in <strong>the</strong> drug entity as it is a single step process which is<br />

highly preferable from <strong>the</strong> industrial point <strong>of</strong> view. Most <strong>of</strong> <strong>the</strong> developed methodologies <strong>for</strong><br />

<strong>the</strong> reductive amination utilized boran as a reducing agent which suffers from many<br />

drawbacks as <strong>the</strong> large toxic waste production. In <strong>the</strong> last ten years scientists focused <strong>the</strong>ir<br />

ef<strong>for</strong>ts on developing an asymmetric version <strong>of</strong> reductive amination utilizing environmentally<br />

friendly hydride source as molecular hydrogen.<br />

4.3. References:<br />

1] D. L. Romero, R. A. Morge, C. Biles, N. Berrios-Pena, P. D. May, J. R. Palmer, P. D.<br />

Johnson, H. W.Smith, M. Busso, C. -K Tan,R. L. Voorman, F. Reusser, I.W. Althaus, K. M.<br />

Downey,A. G. So, L. Resnick, W.G. Tarpley, P. A. Arist<strong>of</strong>f, J. Med. Chem. 1994, 37, 999.<br />

[2] D. S. Johnson, J.J. Li, The Art <strong>of</strong> Drug Syn<strong>the</strong>sis, Wiley & Sons, New Jersey, 2007.<br />

[3] M. Lautens and T. Rovis, J. Org. Chem. 1997, 62, 5246;<br />

[4] E. J. Corey and T. G. Gant, Tetrahedron Lett. 1994, 35, 5373.<br />

[5] S. Takano, M. Sasaki, H. Kanno, K. Shishido, K. Ogasawara, J. Org. Chem.1978, 43,<br />

4169.<br />

[6] T. Fujii, S. Yoshifuji, Tetrahedron 1980, 36, 1539 .<br />

[7] D. Lednicer, The Organic Chemistry <strong>of</strong> Drug Syn<strong>the</strong>sis, Wiley & Sons, New Jersey, 2008.<br />

[8] D. Lednicer, Strategies <strong>for</strong> Organic Drug Syn<strong>the</strong>sis and Design, Wiley & Sons, New<br />

Jersy, 2009.<br />

[9] C. Loncle, C. Salmi, Y. Letourneux, J. M. Brunel, Tetrahedron 2007, 63, 12968.<br />

[10] J. Y. Jung, S. H. Jung a, H. Y. Kohb, European Journal <strong>of</strong> Medicinal Chemistry 2007,<br />

42, 1044.<br />

[11] D. Menche, F. Arikan, J. Li, S. Rudolph, Org. Lett. 2007, 9, 267.<br />

97


[12] S. D. Boggs, J. D. Cobb, K. S. Gudmundsson, L. A. Jones, R. T. Matsuoka, A. Millar, D.<br />

E. Patterson, V. Samano, M. D. Trone, S. Xie, X. –M, Zhou, Org. Process Res. Dev. 2007,<br />

11, 539.<br />

98


Chapter 5<br />

Stoichiometric Use <strong>of</strong> Ytterbium<br />

Acetate in Reductive Amination.<br />

5.1. Introduction.<br />

The general lack <strong>of</strong> literatures describing <strong>the</strong> syn<strong>the</strong>sis <strong>of</strong> alkyl-alkyl chiral amines has<br />

encouraged us to try developing a new methodology <strong>for</strong> <strong>the</strong>ir syn<strong>the</strong>sis. Of <strong>the</strong> commonly<br />

explored strategies <strong>for</strong> <strong>the</strong> α-chiral amine syn<strong>the</strong>sis (described earlier), reductive amination<br />

has <strong>the</strong> advantage <strong>of</strong> being a stepwise efficient methodology with low waste generation. [1]<br />

The use <strong>of</strong> atom economic environmentally friendly hydride source (molecular hydrogen) is<br />

routinely practiced by <strong>the</strong> pharmaceutical industries <strong>for</strong> achiral C-N bond <strong>for</strong>mations but not<br />

<strong>of</strong>ten reported in literatures. (R)- and (S)-α-methylbenzylamine (α-MBA) were chosen as <strong>the</strong><br />

chiral amine auxiliary among. Of course <strong>the</strong>re are o<strong>the</strong>r available auxiliaries such as (R)-and<br />

(S)- phenylglycinol, (R)- and (S)- phenylglycine amide or (R)- or (S)- t-butylsulfinylamide. α-<br />

MBA was chosen <strong>for</strong> three reasons: it is inexpensive, already in use by <strong>the</strong> pharmaceutical<br />

industries and <strong>the</strong> cleavage (hydrogenolysis) <strong>of</strong> this auxiliary is well established high yielding<br />

process. [2]<br />

Table 5.1. Sigma-Aldrich Quote may 2006 <strong>for</strong> Two Common <strong>Chiral</strong> Ammonia<br />

Equivalents. [3]<br />

Chemical Name Quantity(kg) Price (US dollars)<br />

(S)-N-tert-butansulfinamide 1.0 13.125.00<br />

(R)-N-tert-butansulfinamide 1.0 13.125.00<br />

99


(S)-α-methylbenzylamine[(S)-α-MBA] 1.0 798.00<br />

(R)-α-methylbenzylamine[(S)-α-MBA] 1.0 2.220.00<br />

This auxiliary has been previously explored in literature <strong>for</strong> <strong>the</strong> syn<strong>the</strong>sis <strong>of</strong> chiral amines.<br />

The previous methods were based on <strong>the</strong> reduction <strong>of</strong> N-α-methylbenzyl ketimines, not on<br />

reductive amination. [4]<br />

For reductive amination, <strong>the</strong> principal side-reaction is <strong>the</strong> <strong>for</strong>mation <strong>of</strong> an alcohol from <strong>the</strong><br />

competing hydrogenation <strong>of</strong> <strong>the</strong> carbonyl starting material. It could be stated that a desirable<br />

attributes <strong>of</strong> an effective reductive amination method is <strong>the</strong> one capable <strong>of</strong> using atom<br />

efficient reducing reagents and ideally do not allow alcohol <strong>for</strong>mation. In that sense,<br />

molecular hydrogen is an ideal source <strong>of</strong> hydride and <strong>the</strong> reduction <strong>of</strong> imine or iminium ion<br />

intermediates must be fast relative to <strong>the</strong> reduction <strong>of</strong> <strong>the</strong> starting carbonyl compounds. Thus,<br />

in reductive amination <strong>the</strong> correct combination <strong>of</strong> hydride source, catalyst and additives is an<br />

important prerequisite <strong>for</strong> its success. Few reports are available and <strong>the</strong>y require fur<strong>the</strong>r<br />

development <strong>for</strong> optimum reactions. It is a two step process <strong>for</strong> <strong>the</strong> final α-chiral primary<br />

amine inhibiting any possibility <strong>of</strong> imine isolation which is time consuming and low yielding<br />

process. The prochiral ketone is converted to <strong>the</strong> corresponding secondary amine in a good to<br />

high diastereoselectivity and yield in <strong>the</strong> presence <strong>of</strong> Lewis acid/(R)- or (S)-α-MBA/H 2 . The<br />

absolute configuration <strong>of</strong> <strong>the</strong> major and minor diastereomer depends on <strong>the</strong> absolute<br />

configuration <strong>of</strong> <strong>the</strong> chiral amine source. This amine can <strong>the</strong>n be purified by column<br />

chromatography or by crystallization. The chiral primary amine is produced from <strong>the</strong><br />

secondary amine through hydrogenolysis.<br />

As motioned be<strong>for</strong>e, Nugent et al have succeeded to establish a new methodology <strong>for</strong><br />

syn<strong>the</strong>sizing α-chiral amines. Reaction conditions as hydrogen pressure, temperature and<br />

time were milder compared to o<strong>the</strong>r available methodologies. Reaction time is short<br />

compared to o<strong>the</strong>r methodologies adding to <strong>the</strong> advantages <strong>of</strong> this strategy. The use <strong>of</strong> minute<br />

quantities <strong>of</strong> metal catalysts (Pd and Pt) is ano<strong>the</strong>r advantage. Ano<strong>the</strong>r factor which makes<br />

this reaction attractive is <strong>the</strong> low hydrogen pressure needed; almost 80% <strong>of</strong> all ketones are<br />

hydrogenated at 8.0 bar and also at room temperature. Usually o<strong>the</strong>r methodologies require<br />

<strong>the</strong> use <strong>of</strong> higher (up to 100 Bar).<br />

100


The unique feature <strong>of</strong> this methodology is <strong>the</strong> ability to use a wide variety <strong>of</strong> structurally<br />

different substrates. Aliphatic and aromatic ketones with different steric and electronic<br />

environment are successfully reductively aminated in this reaction. The diversity <strong>of</strong> <strong>the</strong> used<br />

ketones will help to broad <strong>the</strong> scope <strong>of</strong> <strong>the</strong> reaction applications. [5]<br />

The success in syn<strong>the</strong>sizing a wide variety <strong>of</strong> amines by this methodology encouraged us to<br />

investigate o<strong>the</strong>r commercially available Lewis acids. I thought that changing <strong>the</strong> Lewis acid<br />

will have an effect on <strong>the</strong> reaction based on our previous results. The use <strong>of</strong> titanium<br />

isopropoxide as a Lewis acid has many advantages and some disadvantages. It is cheap and<br />

available in kg quantities and it is already in use <strong>for</strong> in industrial processes. On <strong>the</strong> o<strong>the</strong>r hand<br />

it is moisture sensitive and it should be used in stoichiometric quantities which complicates<br />

<strong>the</strong> work up process.<br />

Different Lewis acids were tested and some <strong>of</strong> <strong>the</strong> tested Lewis acids showed promising<br />

results. The most pronounced results obtained through using stoichiometric quantity <strong>of</strong><br />

ytterbium acetate. The use <strong>of</strong> this Lewis acid improved <strong>the</strong> de <strong>of</strong> <strong>the</strong> reaction tremendously.<br />

The effect <strong>of</strong> this Lewis acid was significant when aliphatic ketones were used as substrates.<br />

The increase in <strong>the</strong> de reached in certain cases more than 10%. Ytterbium acetate is a solid<br />

Lewis acid and less moisture sensitive compared to titanium isopropoxide which makes it<br />

easier in handling. In <strong>the</strong> following paragraphs I will try to give a brief overview about<br />

ytterbium and its applications in organic syn<strong>the</strong>sis.<br />

5.1.1. Ytterbium.<br />

5.1.1.1. Electronic Overview:<br />

Ytterbium is one <strong>of</strong> <strong>the</strong> Lanthanides rare earth metals. Symbol Yb; atomic number 70; atomic<br />

weight 173.04; valence +2, +3; atomic radius 1.945Å; ionic radius, Yb3+ 0.868Å and 0.98Å<br />

<strong>for</strong> CN 6 and 8; respectively; seven naturally occurring stable isotope: Yb-170 (3.05%), Yb-<br />

171 (14.32%), Yb-172 (21.93%), Yb-173 (16.12%), Yb-174 (31.84%), Yb-176 (12.72%);<br />

twenty-three artificial radioactive isotopes in <strong>the</strong> mass range 151-167, 169, 175, 177-180; <strong>the</strong><br />

101


longest-lived radioisotope Yb-169, t1/2 32.03 days; shortest-lived radioisotope Yb-154, 0.40<br />

second.<br />

5.1.1.2. Ytterbium Discovery:<br />

Ytterbium was discovered in 1878 by J. C. G. de Marignac. The element got its name from<br />

<strong>the</strong> Swedish village Ytterby where this rare earth first was discovered. In 1907, Urbain<br />

separated ytterbia into two components, neoytterbia (oxides <strong>of</strong> ytterbium) and lutecia<br />

(lutecium). The first preparation <strong>of</strong> metallic ytterbium was achieved by Klemm and Bommer<br />

through reduction with potassium metal resulting in an impure ytterbium metal (mixed with<br />

potassium chloride). Daane, Dennison, and Spedding were <strong>the</strong> first to prepare <strong>the</strong> pure metal<br />

in 1953 in gram quantities. Abundance <strong>of</strong> ytterbium in <strong>the</strong> earth’s crust is estimated to be 3.2<br />

mg/kg. Up till now <strong>the</strong> metal had showed very little applications on <strong>the</strong> commercial level. In<br />

elemental <strong>for</strong>m it can be used as a laser source, a portable x-ray source, and as a dopant in<br />

garnets. When added to stainless steel, it improves grain refinement, strength, and o<strong>the</strong>r<br />

properties. Some o<strong>the</strong>r applications, particularly used as oxides mixed with o<strong>the</strong>r rare earths,<br />

include carbon rods <strong>for</strong> industrial lighting, in insulated capacitor and in glass industry. Its<br />

radioactive isoptope is used in detection <strong>of</strong> metal perfection.<br />

5.1.1.3. Ytterbium Reactions:<br />

Ytterbium metal reacts with oxygen above 200°C <strong>for</strong>ming two oxides, <strong>the</strong> monoxide, YbO,<br />

and more stable sesquioxide, Yb 2 O 3 . The metal dissolves in dilute and concentrated mineral<br />

acids. At ordinary temperatures, ytterbium, similar to o<strong>the</strong>r rare earth metals, is corroded<br />

slowly by caustic alkalies, ammonium hydroxide, and sodium nitrate solutions. The metal<br />

dissolves in liquid ammonia <strong>for</strong>ming a deep blue solution. It can react slowly with halogens<br />

at room temperature but progress rapidly above 200°C <strong>for</strong>ming ytterbium trihalides. All <strong>the</strong><br />

trihalides; namely, <strong>the</strong> YbCl 3 , YbBr 3 , and YbI 3 with <strong>the</strong> exception <strong>of</strong> trifluoride, YbF 3 , are<br />

hygroscopic and soluble in water. Ytterbium <strong>for</strong>ms many binary, metalloid, and intermetallic<br />

compounds with a number <strong>of</strong> elements when heated at elevated temperatures. It can <strong>for</strong>m salt<br />

with organic acid triflic or acetic acid. These salts and <strong>the</strong> salts with halogens are used as<br />

Lewis acids in different organic trans<strong>for</strong>mations. [6]<br />

102


Ytterbium triflate is <strong>the</strong> most commonly used <strong>for</strong>m <strong>of</strong> ytterbium in organic syn<strong>the</strong>sis. It is<br />

used in Michael addition <strong>of</strong> β-Ketoester in water, [7] syn<strong>the</strong>sis <strong>of</strong> ethyl arylacetates, [8]<br />

hydrolysis <strong>of</strong> tritylamines and trityloxy compounds to <strong>the</strong> corresponding amines and<br />

alcohols, [9] electrophilic substitution <strong>of</strong> indoles <strong>for</strong> <strong>the</strong> syn<strong>the</strong>sis <strong>of</strong> unnatural tryptophan<br />

derivatives, [10] Friedel–Crafts reaction <strong>of</strong> arylidenecyclopropanes, [11] syn<strong>the</strong>sis <strong>of</strong><br />

polyhydroquinoline derivatives, [12] and syn<strong>the</strong>sis <strong>of</strong> substituted imidazoles. [13]<br />

5.1.1.4. Ytterbium Acetate<br />

Ytterbium acetate is a moderately water soluble crystalline ytterbium source that decomposes<br />

to ytterbium oxide on heating. Acetates are excellent precursors <strong>for</strong> <strong>the</strong> production <strong>of</strong> ultra<br />

high purity compounds and certain catalysts and nanoscale (nanoparticles and nanopowders)<br />

materials. All metallic acetates are inorganic salts <strong>of</strong> a metal cation and <strong>the</strong> acetate anion. The<br />

acetate anion is a univalent (-1 charge) polyatomic ion composed <strong>of</strong> two carbon atoms<br />

ionically bound to three hydrogen and two oxygen atoms (Symbol: CH 3 COO) <strong>for</strong> a total<br />

<strong>for</strong>mula weight <strong>of</strong> 59.05. Ytterbium acetate is applied to fibre amplifier and fibre optic<br />

technologies and in lasing applications. It has a single dominant absorption band at 985 in <strong>the</strong><br />

infra-red useful in silicon photocells to convert radiant energy to electricity.<br />

Two examples were reported in literatures <strong>for</strong> <strong>the</strong> applications <strong>of</strong> ytterbium acetate in organic<br />

chemistry. Fujiwara reported <strong>the</strong> use <strong>of</strong> ytterbium acetate <strong>for</strong> <strong>the</strong> syn<strong>the</strong>sis <strong>of</strong> acetic acid in<br />

water. The method depends on <strong>the</strong> carboxylation <strong>of</strong> methane with carbon monoxide using<br />

ytterbium acetate. Sodium hypochlorite or hydrogen peroxide was used as <strong>the</strong> oxidant in this<br />

reaction. The catalytic activity was improved by <strong>the</strong> addition <strong>of</strong> transition-metal salts such as<br />

manganese acetate. The best result was achieved at a ratio <strong>of</strong> manganese acetate to ytterbium<br />

acetate <strong>of</strong> 1:10. [14]<br />

Oshima reported <strong>the</strong> use <strong>of</strong> ytterbium acetate as additive in oxidation <strong>of</strong> alcohols to aldehydes<br />

and ketones utilizing iodosylbenzene as oxidant. Mixing ytterbium salt with isodosylbenzene<br />

and alcohol in 1,2-dichloroethane and heating <strong>the</strong> mixture at 80 °C <strong>for</strong> 3.5 hours provided <strong>the</strong><br />

ketone products in good to excellent yields. [15]<br />

103


From <strong>the</strong> previous discussion it is obvious that ytterbium acetate was rarely reported in<br />

organic chemistry. In our initial studies I tested different available Lewis acids with benzyl<br />

acetone as <strong>the</strong> ketone substrate. The original reported de was 80% using Ti (O i Pr) 4 / Ra-Ni, α-<br />

MBA, H 2 (120 psi), DCM. Benzyl acetone is an excellent substrate <strong>for</strong> screening as it is<br />

considerably cheap, available in large quantities with high purity from chemical suppliers and<br />

has high molecular weight facilitating its work up. Some <strong>of</strong> <strong>the</strong> tested Lewis acids and results<br />

obtained are summarized in (table 5.2).<br />

Table 5.2. a Different Lewis Acids Tested <strong>for</strong> <strong>the</strong> Reductive Amination <strong>of</strong> Benzyl Actone.<br />

Lewis Acid Ketone Left Alc. Formed Imine Left de%<br />

Dysprosium(III) acetate<br />

hydrate<br />

- - 1.36 77.5<br />

Lanthanum(III) acetate<br />

hydrate<br />

- - 1.3 76.44<br />

Lanthanum(III)<br />

trifluoromethanesulfonate<br />

90 - - NA<br />

Cerium<br />

trifluoromethanesulfonate<br />

90 - - NA<br />

Cerium(III) acetate<br />

hydrate<br />

1.6 - 1.29 76.9<br />

Indium(III) acetate 29.9 - 13.1 79.03<br />

Indium(III)<br />

trifluoromethanesulfonate<br />

90 - - NA<br />

104


Ru<strong>the</strong>nium(III) chloride 31.7 4 4.2 71.29<br />

Cerium(IV) sulfate<br />

tetrahydrate<br />

90 - - NA<br />

Cerium(III) chloride<br />

heptahydrate<br />

90 - - NA<br />

Copper(II)<br />

trifluoromethanesulfonate<br />

90 - - NA<br />

Zinc<br />

trifluoromethanesulfonate<br />

2.8 - - 66.5<br />

Yttrium(III)<br />

trifluoromethanesulfonate<br />

68.17 - - 70.7<br />

Bismuth(III)<br />

trifluoromethanesulfonate<br />

90 - - NA<br />

Dysprosium(III)<br />

trifluoromethanesulfonate<br />

63.5 - - 75.5<br />

Scandium(III) triflate 90 - - NA<br />

Iron(III) bromide 90 - - NA<br />

Aluminum chloride 90 - - NA<br />

105


Bismuth(III) acetate 34.8 - - 68.49<br />

Yttrium(III)<br />

trifluoroacetate hydrate<br />

2.86 - - 76.79<br />

Ytterbium(III) acetate<br />

hydrate<br />

- - - 85<br />

Ytterbium(III) acetate<br />

tetrahydrate<br />

1.5 - - 84<br />

a All reactions per<strong>for</strong>med using 1.0 mmol <strong>of</strong> Benzylacetone, 1.1 mmol <strong>of</strong> (S)-(−)-α-Methylbenzylamine, 1.1<br />

equiv <strong>of</strong> <strong>the</strong> indicated Lewis acid, room temperature, 120 psi (8.3 bar) <strong>of</strong> H 2 , 100 wt % Raney Nickel, and<br />

methanol as a solvent. All components (except <strong>the</strong> Raney Ni and H 2 ) are added toge<strong>the</strong>r and pre-stirred <strong>for</strong> 30<br />

min. The heterogeneous hydrogenation catalyst (Raney Ni) is <strong>the</strong>n added and <strong>the</strong> system pressurized with 8 bar<br />

<strong>of</strong> H 2 . The indicated data is at 12 h <strong>of</strong> reaction from <strong>the</strong> onset <strong>of</strong> hydrogenation.<br />

The data from <strong>the</strong> table indicates that most <strong>of</strong> <strong>the</strong> used Lewis acids showed inferior results<br />

compared to Ti(O i Pr) 4 . The reaction did not even proceed using certain Lewis acids. Only<br />

ytterbium acetate hydrate and ytterbium acetate tetrahydrate showed improvement in <strong>the</strong> de.<br />

These results were encouraging to test ytterbium acetate hydrate or tetrahydrate under<br />

different reaction conditions aiming <strong>for</strong> fur<strong>the</strong>r improvement in <strong>the</strong> de. Different solvents<br />

were tested. The best solvent in terms <strong>of</strong> reaction rate was methanol. This may be due to<br />

partial solubility <strong>of</strong> ytterbium in methanol. The results <strong>of</strong> solvent screening are summarized<br />

in <strong>the</strong> (table 5.3).<br />

Table 5.3. Solvent Screeing with Ytterbium Acetate Hydrate. a<br />

Solvent Ketone Left% Alc. Formed Imine Left de%<br />

Methylene chloride 77 - - 78<br />

Isopropanol 15 - - 76<br />

Toluene 23 - - 80<br />

106


t-Butyl methyl<br />

e<strong>the</strong>r<br />

15 - - 82<br />

Hexane 60 - - 75<br />

THF 20 - - 86<br />

Methanol - - - 81<br />

a<br />

All reactions per<strong>for</strong>med using 1.0 mmol <strong>of</strong> Benzylacetone, 1.1 mmol <strong>of</strong> (S)-(−)-α-Methylbenzylamine, 1.1<br />

equiv <strong>of</strong> Yb(OAc) 3 , room temperature, 120 psi (8.3 bar) <strong>of</strong> H 2 , 100 wt % Raney Nickel, and solvent as<br />

indicated. All components (except <strong>the</strong> Raney Ni and H 2 ) are added toge<strong>the</strong>r and pre-stirred <strong>for</strong> 30 min. The<br />

heterogeneous hydrogenation catalyst (Raney Ni) is <strong>the</strong>n added and <strong>the</strong> system pressurized with 8 bar <strong>of</strong> H 2 . The<br />

indicated data is at 12 h <strong>of</strong> reaction from <strong>the</strong> onset <strong>of</strong> hydrogenation.<br />

The use <strong>of</strong> THF improved <strong>the</strong> de but <strong>the</strong> reaction was slower. O<strong>the</strong>r solvents showed slower<br />

reaction rate with low de. The high de resulting from <strong>the</strong> use <strong>of</strong> THF and <strong>the</strong> fast reaction rate<br />

resulting from <strong>the</strong> use <strong>of</strong> methanol was <strong>the</strong> driving <strong>for</strong>ce to test <strong>the</strong> solvent combination <strong>of</strong><br />

THF-MeOH (1:1). The rate <strong>of</strong> <strong>the</strong> reaction was acceptable with high de (87-89%). O<strong>the</strong>r<br />

solvent combinations also were tested (table 5.3).<br />

Table 5.3. Screening <strong>of</strong> Different Solvent Combinations with Ytterbium Acetate Hydrate. a<br />

Solvent Ketone Left% Alc. Formed Imine Left de%<br />

THF-DMF >90 - - NA<br />

DCM-MeOH 77 - - 78<br />

Isopropanol-<br />

MeOH<br />

15 - - 76<br />

Toluene-MeOH 24 - - 82<br />

EtOAc-MeOH 6 - - 86<br />

THF-MeOH 2 - - 89<br />

DME-MeOH 50 - - 85<br />

DCE-MeOH 51 - - 80<br />

DEE-MeOH 55 - - 84<br />

a<br />

All reactions per<strong>for</strong>med using 1.0 mmol <strong>of</strong> Benzylacetone, 1.1 mmol <strong>of</strong> (S)-(−)-α-Methylbenzylamine, 1.1<br />

equiv <strong>of</strong> Yb(OAc) 3 , room temperature, 120 psi (8.3 bar) <strong>of</strong> H 2 , 100 wt % Raney Nickel, and solvent as<br />

107


indicated. All components (except <strong>the</strong> Raney Ni and H 2 ) are added toge<strong>the</strong>r and pre-stirred <strong>for</strong> 30 min. The<br />

heterogeneous hydrogenation catalyst (Raney Ni) is <strong>the</strong>n added and <strong>the</strong> system pressurized with 8 bar <strong>of</strong> H 2 . The<br />

indicated data is at 12 h <strong>of</strong> reaction from <strong>the</strong> onset <strong>of</strong> hydrogenation.<br />

Ethyl acetate-MeOH and THF-MeOH combinations showed <strong>the</strong> highest possible de with<br />

acceptable reaction time (12 h). The use <strong>of</strong> THF-MeOH resulted in slightly higher de<br />

compared to EtOAc-MeOH encouraging us to proceed using this solvent combination <strong>for</strong> <strong>the</strong><br />

rest <strong>of</strong> <strong>the</strong> study. The addition <strong>of</strong> an anhydrous MgSO 4 or NaSO 4 as desiccants to <strong>the</strong> reaction<br />

mixture be<strong>for</strong>e hydrogenation did not have any effect on <strong>the</strong> reaction pr<strong>of</strong>ile.<br />

After using several bottles <strong>of</strong> Yb(OAc) 3 , which is sold and described as a semihydrated <strong>for</strong>m<br />

(Sigma-Aldrich catalogue no. 544973), I noted that it was sometimes free flowing while o<strong>the</strong>r<br />

bottles from <strong>the</strong> same lot were not. In our attempt to eliminate any variation <strong>of</strong> results and to<br />

get consistent reaction pr<strong>of</strong>ile <strong>for</strong> all substrates, I decided to dry Yb(OAc) 3 powder obtained<br />

from <strong>the</strong> commercial supplier. The powder was high vacuum dried to constant weight at 80<br />

°C (12 h). The dried powder was kept in airtight container and used <strong>for</strong> fur<strong>the</strong>r reactions.<br />

Through this drying procedures all reactions results were reproducible and constant reaction<br />

pr<strong>of</strong>ile was obtained. In <strong>the</strong> rest <strong>of</strong> this study <strong>the</strong> term “dry Yb(OAc) 3 ” means dried as just<br />

stated. The dried Yb(OAc) 3 could be stored in a dry screw cap glass bottle at room<br />

temperature. The container could be repeatedly opened to <strong>the</strong> atmosphere (at least 6 times<br />

without detrimental effect) and <strong>the</strong> desired quantity <strong>of</strong> Yb(OAc) 3 weighed out without <strong>the</strong><br />

need <strong>for</strong> a glovebox. This is one <strong>of</strong> <strong>the</strong> significant advantage <strong>of</strong> Yb(OAc) 3 use in reductive<br />

amination compared to o<strong>the</strong>r air sensitive Lewis acids. Moisture and air stability <strong>of</strong> Yb(OAc) 3<br />

will open <strong>the</strong> door <strong>for</strong> its applications on industrial scale.<br />

After initial optimization <strong>of</strong> <strong>the</strong> reaction using benzylacetone as <strong>the</strong> substrate o<strong>the</strong>r ketone<br />

substrates were also tested. 2-octanone is ano<strong>the</strong>r example <strong>of</strong> 2-alkanones which has been<br />

reductively aminated with Ti(O i Pr) 4 system with a de <strong>of</strong> (72%), was one <strong>of</strong> <strong>the</strong> interesting<br />

substrates to be tested with Yb(OAc) 3 system. Solvent screening was carried also to this<br />

substrate utilizing all findings obtained from benzylacetone results. For example 2-octanone,<br />

in MeOH, was fully consumed within 8 h providing <strong>the</strong> secondary amine in 82% de in <strong>the</strong><br />

presence <strong>of</strong> Raney-Ni (Scheme 5.1). When <strong>the</strong> solvent was changed to THF, <strong>the</strong><br />

stereoselectivity increased to 87-88% de, but 24 h were required to completely consume <strong>the</strong><br />

108


2-octanone starting material. When <strong>the</strong> binary solvent system <strong>of</strong> MeOH-THF (1:1) was<br />

examined, an 86% de was consistently achieved with a fast reaction time <strong>of</strong> 10-12 h. The<br />

same solvent systems which proved to be efficient in <strong>the</strong> reductive amination <strong>of</strong><br />

benzylacetone were also useful in <strong>the</strong> reductive amination <strong>of</strong> 2-octanone. From solvent<br />

screening studies it was obvious that <strong>the</strong> presence <strong>of</strong> MeOH in <strong>the</strong> solvent mixture is essential<br />

<strong>for</strong> <strong>the</strong> fast reaction rate. Replacement <strong>of</strong> THF in THF-MeOH mixture with o<strong>the</strong>r solvents<br />

resulted in lower de or/and prolonged reaction time. The same de was obtained through<br />

replacing THF in <strong>the</strong> THF-MeOH system with toluene, Et 2 O, or 1,3-dioxolane but with<br />

moderately longer reaction times. Replacing MeOH with EtOH in THF-MeOH system<br />

resulted in <strong>the</strong> same de but <strong>the</strong> reaction time was longer (24h). To ensure that this high<br />

diastereoselectivity is maintained and no racemization is occurring, I hydrogenolyzed <strong>the</strong><br />

reductive amination product to ensure that <strong>the</strong> enantiopurity <strong>of</strong> <strong>the</strong> primary amine is<br />

preserved (scheme 5.1.). This level <strong>of</strong> diastereoselectivity represents a 15-16% increase in <strong>the</strong><br />

de over <strong>the</strong> best previously reported <strong>for</strong> 2-octanone and α-MBA.<br />

1d<br />

O<br />

+<br />

H 2 N<br />

Ph<br />

(S)-α-MBA<br />

Yb(OAc)3 ,MeOH-THF<br />

Raney-Ni, H 2 (120 psi)<br />

(S,S)-2d HN Ph Pd-C<br />

(S)-3d NH 2<br />

H 2 (60 psi)<br />

86% de 85% ee<br />

Scheme 5.1. Two-Step Procedure <strong>for</strong> Producing (2S)-Aminooctane in High ee.<br />

The high de obtained in reductive amination <strong>of</strong> benzylacetone and 2-ocatnone utilizing<br />

Yb(OAc) 3 encouraged us to test and evaluate o<strong>the</strong>r ytterbium salts present commercially. As<br />

mentioned be<strong>for</strong>e Yb(OTf) 3 is <strong>the</strong> most common derivative <strong>of</strong> ytterbium used in organic<br />

syn<strong>the</strong>sis. When Yb(OTf) 3 was used in reductive amination <strong>of</strong> 2-ocatnone alcohol was <strong>the</strong><br />

major product. No amine was detected after 10 h or <strong>the</strong> reaction. YbCl 3 provided <strong>the</strong> product<br />

in 66% de, but in only 23 area % (GC) after 24 h (Table 5.4, entries 2 and 3). The use <strong>of</strong><br />

highly expensive salt <strong>of</strong> ytterbium which is Yb(O i Pr) 3 resulted also in alcohol <strong>for</strong>mation and<br />

no secondary amine was detected. From previous findings it is obvious that Yb(OAc) 3 is <strong>the</strong><br />

best <strong>for</strong>m <strong>of</strong> ytterbium to be used in reductive amination. Of course to eliminate any doubts<br />

regarding free acetate in solution modifying <strong>the</strong> heterogeneous metal surface <strong>of</strong> <strong>the</strong> catalyst,<br />

acetate salts were tested alone in <strong>the</strong> reaction. The addition <strong>of</strong> NaOAc was examined (Table<br />

5.4, entry 9). In <strong>the</strong> event, gross quantities <strong>of</strong> <strong>the</strong> alcohol by-product resulted, making this<br />

109


simplified scenario less likely. It is clear that ytterbium and <strong>the</strong> acetate ligand toge<strong>the</strong>r are<br />

needed <strong>for</strong> <strong>the</strong> efficient reductive amination process.<br />

Historically <strong>the</strong> use <strong>of</strong> acetic acid as Brønsted acid in reductive amination is well established<br />

on laboratory and industrial scales. It is cheap, available in kilogram quantities, easy to<br />

handle and usually used in catalytic quantities. Testing acetic acid in our reaction will help to<br />

establish ano<strong>the</strong>r reference point to our work after <strong>the</strong> comparing with Ti(O i Pr) 4 system. The<br />

use <strong>of</strong> (0.2, 0.5 and 1.0 equiv) <strong>of</strong> acetic acid inhibited alcohol <strong>for</strong>mation but did not show any<br />

de improvement. The addition <strong>of</strong> acetic acid to Yb(OAc) 3 reaction resulted in <strong>the</strong> reduction <strong>of</strong><br />

diastereoselectivity <strong>of</strong> <strong>the</strong> secondary amine (72%). All <strong>the</strong>se findings proved that Yb(OAc) 3 is<br />

a unique Lewis acid <strong>for</strong> reductive amination.<br />

As mentioned previously that reductive amination does not require imine separation and<br />

purification which is achieved by <strong>the</strong> addition <strong>of</strong> proper Lewis acid. Some older reports<br />

simplify <strong>the</strong> role <strong>of</strong> Lewis acids in reductive amination to <strong>the</strong> level <strong>of</strong> an efficient desiccant<br />

promoting in situ imine <strong>for</strong>mation in high yield. Extensive studies on <strong>the</strong> mechanism <strong>of</strong><br />

reductive amination and intermediates structures contradicted <strong>the</strong>se simplified speculations<br />

and proved that Lewis acids have greater role than efficient desiccants. To study this effect<br />

more closely, I examined some traditional desiccants. When Yb(OAc) 3 (1.1 equiv) is<br />

replaced by MgSO 4 (5 equiv) or 4Å molecular sieves (4 wt equiv), all vacuum oven dried at<br />

150 ° C <strong>for</strong> 15 h be<strong>for</strong>e use, not only low diastereoselectivities were observed (Table 5.4,<br />

compare entries 1, 6, 7), but gross amounts <strong>of</strong> 2-octanone were reduced to <strong>the</strong> alcohol byproduct.<br />

In relation to <strong>the</strong>se results, alcohol by-product <strong>for</strong>mation could be significantly<br />

suppressed when Ti(O i Pr) 4 (1.25 equiv) was used, but <strong>the</strong> de remained low. These combined<br />

findings clearly establish Yb(OAc) 3 as fulfilling a greater role than that <strong>of</strong> a simple desiccant.<br />

Table 5.4. a Initial Study <strong>of</strong> <strong>the</strong> Role <strong>of</strong> Yb(OAc) 3 in <strong>the</strong> Reductive Amination <strong>of</strong> 2-Octanone. a<br />

Amine 2d<br />

entry additive time (h) (S)-α-MBA (%) b 2-octanol (%) yield (%) de (%)<br />

1 Yb(OAc) 3<br />

c<br />

2 YbCl 3<br />

d<br />

10 0.7 0.3 90.4 86.1<br />

23 77.4 0.0 22.5 66.2<br />

110


3 Yb(OTf) 3<br />

d<br />

4 Ti(O i Pr) 4<br />

e<br />

5 B(O i Pr) 3<br />

e<br />

9 - [f] 96.4 3.5 -<br />

10 2.8 11.4 82.3 67.0<br />

10 14.7 24.4 59.6 71.0<br />

6 MgSO 4 12 19 29.3 52.0 70.5<br />

7 4 Å M.S. 12 18.7 28.8 52.5 69.6<br />

8 none 12 24.2 34.6 41.1 70.8<br />

9 NaOAc 23 26.5 43.3 29.7 70.8<br />

10 HOAc 12 4.3 1.0 94.1 72<br />

a (S)-α-MBA (2.5 mmol, 1.0 equiv), 2-octanone (1.2 equiv), and an additive (entries 1-5, 9, and 10, 1.1 equiv <strong>of</strong><br />

additive, 4Å molecular sieves (4 wt equiv), or MgSO 4 (5.0 equiv)) are stirred in MeOH (1.0 M) <strong>for</strong> 30 min at rt,<br />

<strong>the</strong>n THF (final molarity 0.5 M) and Raney-Ni are added, and <strong>the</strong> reaction pressurized with H 2 (120 psi). All<br />

data is based on GC area % analysis. b Sum <strong>of</strong> (S)-α-MBA and imine remaining at <strong>the</strong> indicated time. c This<br />

result is when Yb(OAc) 3 has not dried, 2-aminooctanone was noted in 6.5 area %. d These Yb salts were only<br />

examined in MeOH. e Under <strong>the</strong> optimal Yb(OAc) 3 conditions used here, <strong>the</strong> Ti(O i Pr) 4 and B(O-i-Pr) 3 Lewis<br />

acids did not provide optimal results. f Not integrated, to emphasize <strong>the</strong> presence <strong>of</strong> <strong>the</strong> dominant alcohol byproduct.<br />

5.1.2. Commercial Yb(OAc) 3 vs Dried Yb(OAc) 3 :<br />

Initial studies were per<strong>for</strong>med using <strong>the</strong> commercially available Yb(OAc) 3 which had a<br />

different physical appearance (powder flowability) in each bottle. Initial results <strong>for</strong> <strong>the</strong><br />

reductive amination <strong>of</strong> 2-octanone with Yb(OAc) 3 in THF-MeOH were promising in terms <strong>of</strong><br />

high de, but <strong>the</strong> isolated (chromatographic) yield <strong>of</strong> secondary amine was 75%. This yield is<br />

considered mediocre <strong>for</strong> a single step process. GC analysis revealed <strong>the</strong> presence <strong>of</strong> unknown<br />

peak with 10-12 area %. Reaction was per<strong>for</strong>med at larger scale (8 mmol) allowing <strong>the</strong><br />

chromatographic separation <strong>of</strong> this compounds. The isolated compound was analyzed using ( 1 H<br />

and 13 C NMR) which revealed that this compound may be as 2-aminooctane. The conditions<br />

under which this primary amine by-product <strong>for</strong>med were when (S)-α-MBA was used as <strong>the</strong><br />

limiting reagent [(ketone 1.2 equiv and undried Yb(OAc) 3 (1.1 equiv)], 2-aminooctanone is<br />

consistently observed at 6-7 area % (GC).<br />

Surprisingly <strong>the</strong> vacuum drying <strong>of</strong> Yb(OAc) 3 , reduced <strong>the</strong> amount <strong>of</strong> 2-aminooctane<br />

dramatically to 1-3 area % (GC). Of course <strong>the</strong> isolated yield <strong>of</strong> <strong>the</strong> secondary amine increased<br />

111


to 86%, and <strong>the</strong> de was consistently reported between 87-88%. Aiming to understand <strong>the</strong><br />

concept behind this effect, H 2 O or AcOH (50 or 100 mol %) was added to <strong>the</strong> reaction mixture<br />

containing dried Yb(OAc) 3 . The amount <strong>of</strong> 2-aminooctane <strong>for</strong>med increased also to ~10 area %<br />

(GC). For this reason I decided to conduct all experiments using only dried Yb(OAc) 3 . Ano<strong>the</strong>r<br />

significant finding from this observation was <strong>the</strong> importance <strong>of</strong> using ketone as <strong>the</strong> limiting<br />

reagent instead <strong>of</strong> α-MBA (1.1 equiv) to obtain <strong>the</strong> optimal de and yield. The use <strong>of</strong> ketone as a<br />

limiting reagent adds to <strong>the</strong> advantages <strong>of</strong> our methodology as it reduces <strong>the</strong> expenses <strong>of</strong><br />

starting materials especially when using expensive ketones on larger scale.<br />

Regarding <strong>the</strong> <strong>for</strong>mation <strong>of</strong> 2-aminooctane, our initial speculation was that <strong>the</strong> reductive<br />

amination products (S,S)- and (R,S)-secondary amine (scheme 5.1.) were slowly being<br />

hydrogenolyzed under <strong>the</strong> reaction conditions, but chiral GC analysis (trifluoroacetamide<br />

derivative) established <strong>the</strong> primary amine side product as a racemate. Based on our prior<br />

experience, regarding <strong>the</strong> stereoinducing capabilities <strong>of</strong> (S)- or (R)-α-MBA with Ti(O i Pr) 4 and<br />

2-octanonone, I considered it unlikely that a racemate would <strong>for</strong>m if a sequential reductive<br />

amination-hydrogenolysis scenario had occurred. This led us to examine <strong>the</strong> possibility that (S)-<br />

α-MBA was being hydrogenolyzed by Raney-Ni, in <strong>the</strong> presence <strong>of</strong> Yb(OAc) 3 , in a small but<br />

significant amount, and <strong>the</strong>reby producing ammonia and ethylbenzene in situ. The subsequent,<br />

but non-productive, reductive amination <strong>of</strong> ammonia with 2-octanone would <strong>the</strong>n account <strong>for</strong><br />

racemic 2-aminooctane <strong>for</strong>mation. In an ef<strong>for</strong>t to support this hypo<strong>the</strong>sis, several reactions were<br />

examined by GC in an ef<strong>for</strong>t to identify ethylbenzene (relative to an au<strong>the</strong>ntic reference<br />

standard), but none was ever observed. To fur<strong>the</strong>r corroborate those findings, I treated α-MBA<br />

(in <strong>the</strong> absence <strong>of</strong> 2-octanone) with Raney-Ni/H 2 . Several repetitions <strong>of</strong> this experiment failed<br />

to allow <strong>the</strong> identification <strong>of</strong> ethylbenzene. The major obstacle <strong>for</strong> identification <strong>of</strong><br />

ethylbenzene was its low boiling point (136 °C). Any low quantities produced could potentially<br />

evaporate on quenching <strong>the</strong> reaction aliquot at room temperature (work-up: drop aliquot into<br />

sat. aq NaHCO 3 /EtOAc). Hoping to reduce ethyl benzene evaporation <strong>the</strong> quenching solution <strong>of</strong><br />

aqueous NaHCO 3 /EtOAc was cooled to 0 °C. Un<strong>for</strong>tunately, this did not allow <strong>the</strong> observation<br />

<strong>of</strong> ethylbenzene by GC.<br />

While [1,3]-proton shift <strong>of</strong> imine is known, it is to our knowledge only accomplished under <strong>the</strong><br />

presence <strong>of</strong> a strong base. [16]<br />

112


If a [1,3]-proton shift <strong>of</strong> <strong>the</strong> initially <strong>for</strong>med imine occurred, followed by hydrolysis, 2-<br />

aminooctane and acetophenone would result. When 2-octanone (1.0 equiv), (S)-α-MBA (1.1<br />

equiv), and Yb(OAc) 3 (1.1 equiv) were added to THF-MeOH (standard reaction conditions),<br />

and stirred in <strong>the</strong> absence <strong>of</strong> Raney-Ni and H 2 , a very small quantity <strong>of</strong> new compound (


4 1d<br />

O<br />

2d<br />

HN<br />

Ph<br />

86 87 15<br />

5 1e<br />

O<br />

2e<br />

HN<br />

Ph<br />

82 85 14<br />

6 1f<br />

O<br />

2f<br />

HN<br />

Ph<br />

80 79 5<br />

a Ketone (2.5 mmol, 1.0 equiv), dried Yb(OAc) 3 (1.1 equiv), (S)-α-methylbenzylamine (1.1 equiv) equiv),<br />

Raney-Ni, H 2 (120 psi), MeOH-THF (1:1) 0.50 M, 22 °C, 12 h. b Isolated yield <strong>of</strong> both diastereomers after<br />

chromatography. c determined by GC analysis <strong>of</strong> crude product 2. d Compared to <strong>the</strong> best previously<br />

reported results. Pinacolone is a Pt-C substrate and requires a T= 50 °C over 22 h. The 6% increase in de<br />

only represents an increase over <strong>the</strong> best reported reductive amination procedure, vs a previously reported<br />

stepwise method <strong>the</strong>re is no change in de.<br />

Through examining <strong>the</strong> results presented in table 5.5 It seems that our methodology is highly<br />

efficient <strong>for</strong> reductive amination <strong>of</strong> ketones having <strong>the</strong> general class R S C(O)CH 3 , linear 2-<br />

alkanones. The subscript serves as a generic reference to <strong>the</strong> steric bulk <strong>of</strong> <strong>the</strong> substituent:<br />

R S = small (any straight chain alkyl substituent, but not a methyl group); R M = medium, e.g. –<br />

CH 2 CH 2 Ph or -CH 2 CH(CH 3 ) 2 ; R L = e.g. -Ar, -i-Pr, -c-hexyl.<br />

For example, <strong>the</strong> longer straight chain 2-alkanones, e.g. 2-octanone (1d) and 2-hexanone<br />

(1e), showed dramatic improvements in de, 15% and 14% respectively, with good isolated<br />

yield (table 5.5, entries 4 and 5). As mentioned be<strong>for</strong>e <strong>the</strong> highest reported de <strong>for</strong> <strong>the</strong><br />

reductive amination <strong>of</strong> 2-octanone (1d) with α-MBA in <strong>the</strong> presence <strong>of</strong> Ti(O i Pr) 4 /Raney-<br />

Ni/H 2 , was 72%. Aiming to prove that <strong>the</strong> addition <strong>of</strong> Yb(OAC) 3 had a dramatic effect on <strong>the</strong><br />

de <strong>of</strong> amine product, I syn<strong>the</strong>sized <strong>the</strong> ketimine <strong>of</strong> 2-octanone. This ketimine was reduced<br />

with Raney-Ni/H 2 in THF-MeOH (1:1) providing 2d in only 64% de.<br />

As <strong>the</strong> chain <strong>of</strong> 2-alkanone gets shorter as <strong>the</strong> steric bulkiness gets smaller reducing <strong>the</strong><br />

enhancement effect <strong>of</strong> Yb(OAc) 3 addition. The short chain 2-butanone (1f) showed a small<br />

but consistent and significant 5% increase in de vs <strong>the</strong> best previously reported result<br />

(Ti(O i Pr) 4 /CH 2 Cl 2 /Raney-Ni: 74% de). Shifting to substrates having medium sized R<br />

substituent residing on 2-alkanone, R M C(O)CH 3 helps to define <strong>the</strong> boundary substrates <strong>for</strong><br />

114


enhanced stereoselectivity. For example when <strong>the</strong> γ-branched benzylacetone (1c) was<br />

examined a 9% increase in de was observed vs <strong>the</strong> previous best reported methods. The use<br />

<strong>of</strong> Yb(OAc) 3 <strong>for</strong> reductive amination <strong>of</strong> β-branched i-butyl methyl ketone (1a), did show any<br />

significant improvement in stereoselectivity (1% increase) (table 5.5, entry 1). Also α-<br />

Branched 2-alkanones, e.g. i-propyl methyl ketone or cyclohexyl methyl ketone, which have<br />

been reductively aminated with a very high diastereoselectivity using Ti(O i Pr) 4 (>98% de),<br />

did not show any improvement when using Yb(OAc) 3 .<br />

Examination <strong>of</strong> alkyl-aryl ketones, e.g. acyclic acetophenone or cyclic benzosuberone, or a<br />

non-2-alkanone, e.g. i-propyl n-propyl ketone, proved problematic. Acetophenone required<br />

higher temperature, 60 °C with 120 psi <strong>of</strong> H 2 , and produced <strong>the</strong> product in 92% de, but with<br />

large amounts <strong>of</strong> <strong>the</strong> corresponding alcohol noted (~20 area %, GC). For benzosuberone and<br />

i-propyl n-propyl ketone repeated attempts to obtain <strong>the</strong> intended product by heating and/or<br />

increasing <strong>the</strong> hydrogen pressure failed.<br />

The previous substrates were successfully reductively amianted using Ti(O i Pr) 4 with good<br />

yield and de. As mentioned previously <strong>the</strong> heterogeneous catalyst used was Raney-Ni. Pt-C<br />

was tested <strong>for</strong> <strong>the</strong> reductive amination <strong>of</strong> i-propyl n-propyl ketone, a 35 area % <strong>of</strong> <strong>the</strong> product<br />

was noted in 76% de (GC). This ketone has been previously reductively aminated with<br />

Ti(O i Pr) 4 /Raney-Ni providing <strong>the</strong> desired product in 76% yield and 87% de.<br />

From previous studies done in our group, it was noted that ketone substrates with an α-<br />

tertiary carbon cannot be reductively aminated using Raney-Ni, even under <strong>for</strong>cing<br />

conditions, instead Pt-C is <strong>the</strong> catalyst <strong>of</strong> choice <strong>for</strong> this class <strong>of</strong> prochiral ketones.<br />

Examination <strong>of</strong> pinacolone (1b) with Yb(OAc) 3 /Pt-C/H 2 , provided a consistent 93% de,<br />

which is 6% greater than <strong>the</strong> previously best reported result reductive amination result (87%<br />

Ti(O i Pr) 4 /Pt-C/H 2 ), but is <strong>the</strong> same as compared to a previously reported stepwise<br />

approach. [17]<br />

5.2. Conclusion:<br />

115


Reductive amination is a step powerful methodology <strong>for</strong> <strong>the</strong> syn<strong>the</strong>sis <strong>of</strong> α-chiral amines. The<br />

correct choice <strong>of</strong> Lewis acid is critical <strong>for</strong> efficient suppression <strong>of</strong> alcohols. The<br />

diastereoselectivity obtained with or without <strong>the</strong> use <strong>of</strong> Lewis acids was <strong>the</strong> same. No<br />

precedent <strong>of</strong> increasing diastereoselectivity in reductive amination with <strong>the</strong> use <strong>of</strong> achiral<br />

Lewis acid was ever reported. The use <strong>of</strong> Yb(OAc) 3 in reductive amination resulted in a<br />

significant increase in diastereoselectivities <strong>for</strong> different 2-alaknones. Diastereoselectivity <strong>of</strong><br />

2-octanone increased 15% compared to <strong>the</strong> highest reported result. Aromatic and cyclic<br />

ketones were not successfully reductively aminated using Yb(OAc) 3 . Ti(O i Pr) 4 proved to be<br />

<strong>the</strong> best Lewis acid in reductive amination <strong>of</strong> <strong>the</strong>se substrates.<br />

5.3. References:<br />

[1] a) H. Lebel, K. Huard, Org. Lett. 2007, 9, 639; b) M. Kim, J. V. Mulcahy, C. G. Espino, J.<br />

Du Bois, Org. Lett. 2005, 7, 4685; c) C. G. Espino, K. W, Fiori, M. Kim, J. Du Boisn, J. Am.<br />

Chem. Soc. 2004, 126, 15378.<br />

[2] a) J. Blacker, Innovations in Pharmaceutical Technology 2001, 1, 77; b) H. -U. Blaser, F.<br />

Spindler, A. Studer, App. Catal. Gen. 2001, 221, 119; c) H. -U. Blaser, M. Eissen, P. F.<br />

Fauquex. K. Hungerbuhler, E. Schmidt, G. Sedelmeier, M. Studer, Asymmetric Catalysis on<br />

Industrial Scale: Challenges, Approaches, and Solutions; H.-U. Blaser, E. Schmidt Eds.;<br />

Wiley-VCH Verlag GmbH & Co. KGaA: Weinheim, Germany, 2004.<br />

[3] T. C. Nugent, <strong>Chiral</strong> Amine Syn<strong>the</strong>sis - Strategies, Examples, and Limitations. In Process<br />

Chemistry in <strong>the</strong> Pharmaceutical Industry, Second Edition: Challenges in an Ever-Changing<br />

Climate, T. F. Braish, K. Gadamasetti Eds.; CRC Press-Taylor and Francis Group: New<br />

York, 2008.<br />

[4] a) L. Storace, L. Anzalone, P. N. Confalone, W. P. Davis, J. M. Fortunak, M.<br />

Giangiordano, J. J. Haley, Jr., K. Kamholz, H.-Y. Li, P. Ma, W. A. Nugent, R. L. Parsons, Jr.,<br />

P. J. Sheeran, C. E. Silverman, R. E. Waltermire, C. C. Wood, Org. Process Res. Dev. 2002,<br />

6, 54; b) E. Juaristi, J. L. León-Romo, A. Reyes, J. Escalante, Tetrahedron: Asymmetry 1999,<br />

10, 2441; c) G. Lauktien, F. -J. Volk, A. W. Frahm, Tetrahedron: Asymmetry 1997, 8, 3457;<br />

d) B. Speckenback, P. Bisel, A. W. Frahm, Syn<strong>the</strong>sis 1997, 1325.<br />

[5] a) T. C. Nugent, A. K. Ghosh, V. N. Wakchaure, R. R. Mohanty, Adv. Synth. Catal. 2006,<br />

348, 1289; b) T. C. Nugent, A. K. Ghosh, V. N. Wakchaure, R. R. Mohanty,<br />

116


WO2006030017, 2006; c) T. C. Nugent, V. N. Wakchaure, A. K. Ghosh, R. R. Mohanty,<br />

Org. lett. 2005, 7, 4967; d) T. C. Nugent, A. K. Ghosh, Eur. J. Org. Chem. 2007, 3863.<br />

[6] P. Patnaik, Handbook <strong>of</strong> Inorganic Chemicals, McGraw-Hill Companies, 2003, PP 973-<br />

975.<br />

[7] E. Keller, B. L. Feringa, Tetrahedron Lett. 1996, 37,1882.<br />

[8] S. Sinha, B. Mandal, S. Chandrasekaran, Tetrahedron Lett. 2000, 41, 9109.<br />

[9] R. J. Lu, D. Liu, R. W. Giese, Tetrahedron Lett. 2000, 41, 2817.<br />

[10] A. Janczuk, W. Zhang, W. Xie, S. Lou, J. P. Chengb, P. G. Wang, Tetrahedron Lett.<br />

2002, 43, 4271.<br />

[11] I. Nakamura, M. Kamada, Y. Yamamoto, Tetrahedron Lett. 2004, 45, 2903.<br />

[12] L.-M. Wang, J. Sheng, L. Zhang, J.-W. Han, Z.-Y. Fan, H. Tiana, C.-T. Qianb,<br />

Tetrahedron 2005,61,1539.<br />

[13] L.-M. Wang, Y.-H.Wang, H. Tian, Y.-F. Yao, J.-H. Shao, B. Liu, J. Fluorine Chem.<br />

2006, 127, 1570.<br />

[14] M. Asadullah, Y.Taniguchi, T. Kitamura, Y. Fujiwara, Appl. Organometal. Chem. 1998,<br />

12, 277.<br />

[15] T. Yokko, K. Matsumoto, K. Oshima, K. Utimoto, Chemistry Letters 1993, 571.<br />

[16] (a) G. Cainelli, D. Giacomini, A. Trerè, P. P. Boyl, J. Org. Chem. 1996, 61, 5139. (b) J.<br />

G. H. Willems, J. G. de Vries, R. J. M. Nolte, B. Zwanenburg, Tetrahedron Lett. 1995, 36,<br />

3917. (c) V. A. Soloshonok, A. G. Kirilenko, S. V. Galushko, V. P. Kukhar, Tetrahedron<br />

Lett. 1994, 35, 5063.<br />

[17] N. Moss, J. Gauthier, J. –M. Ferland, Synlett 1995, 142.<br />

117


Chapter 6<br />

Catalytic Lewis Acids in Reductive<br />

Amination.<br />

6.1. Introduction.<br />

Historically Lewis acids were used in reductive amination in stoichiometric quantities.<br />

Through extensive literature search I did not find any previous example <strong>for</strong> <strong>the</strong> catalytic use<br />

<strong>of</strong> Lewis acids in reductive amination. This was one <strong>of</strong> <strong>the</strong> limitations <strong>for</strong> scaling up <strong>the</strong> use<br />

<strong>of</strong> Lewis acids in reductive amination. The high diastereoselectivity obtained with <strong>the</strong> use <strong>of</strong><br />

Yb(OAc) 3 in reductive amination motivated us to investigate deeply <strong>for</strong> <strong>the</strong> possibility <strong>of</strong><br />

using Lewis acids in catalytic quantities. Table 6.1 shows <strong>the</strong> effect <strong>of</strong> slowly decreasing <strong>the</strong><br />

mol % <strong>of</strong> dried Yb(OAc) 3 from 110 mol % to 10 mol %. The data shows that 80 mol %<br />

Yb(OAc) 3 produces <strong>the</strong> same results as 110 mol %, and interestingly 50 mol % Yb(OAc) 3 is<br />

capable <strong>of</strong> maintaining very similar de (1% less) as compared to 110 mol %. Rapid<br />

deterioration in <strong>the</strong> diastereoselectivity was noticed when <strong>the</strong> Yb(OAc) 3 loading was reduced<br />

below 40%. When <strong>the</strong> Yb(OAC) 3 loading was reduced to 10 mol % no improvement in <strong>the</strong><br />

de was noticed <strong>the</strong> de’s observed were similar to <strong>the</strong> previously reported with non-Yb(OAc) 3<br />

based methods with α-MBA. [1]<br />

Table 6.1. Relationship Between Mol % <strong>of</strong> Yb(OAc) 3 and Diastereoselectivity <strong>of</strong> Amine<br />

Product. a<br />

entry Yb(OAC) 3 mol% de%<br />

1 110 87<br />

2 100 87<br />

3 80 87<br />

118


4 60 86<br />

5 50 86<br />

6 40 84<br />

7 20 79<br />

8 20 80 b<br />

9 10 72<br />

a Ketone (2.5 mmol, 1 equiv), dried Yb(OAc) 3 (1.1 equiv), (S)-α-methylbenzylamine (1.1 equiv), Raney-Ni, H 2 (120 psi),<br />

MeOH-THF (1:1) 0.50 M, 22 °C, 12 h. b 4 Å molecular sieves (4 wt equiv) were also added.<br />

Fur<strong>the</strong>r reduction in <strong>the</strong> Yb(OAc) 3 loading below 10 mol % led to increase <strong>the</strong> <strong>for</strong>mation <strong>of</strong><br />

alcohol byproduct. When 5 mol % <strong>of</strong> Yb(OAc) 3 was used, alcohol by product was detected in<br />

aquantity greater than 2 area % (GC). For this detailed study I reached to <strong>the</strong> conclusion that<br />

<strong>the</strong> least Yb(OAc) 3 loading is 10 mol % which should be used <strong>the</strong> rest <strong>of</strong> <strong>the</strong> study.<br />

Encouraged by <strong>the</strong>se results I investigated a variety <strong>of</strong> transition metal, lanthanide, and<br />

metalloid halide, acetates, alkoxides, and sulfonates, <strong>for</strong> <strong>the</strong>ir ability to allow fast and high<br />

yielding reductive amination reactions to occur. From this extensive study Ce(OAc) 3 (15 mol<br />

%) and Y(OAc) 3 (15 mol %) emerged as useful Lewis acids inhibiting alcohol <strong>for</strong>mation. The<br />

results obtained through using <strong>the</strong>se two Lewis acids were consistently similar to <strong>the</strong> results<br />

<strong>of</strong> Yb(OAc) 3 (10 mol %), regarding reaction times, yield, and diastereoselectivity <strong>for</strong> <strong>the</strong><br />

reductive amination <strong>of</strong> 2-octanone with (S)-α-MBA. Stoichiometric use <strong>of</strong> Ce(OAc) 3 or<br />

Y(OAc) 3 , did not provide enhanced de or any o<strong>the</strong>r added benefit over those reactions<br />

examined at <strong>the</strong> 15 mol % level.<br />

O<strong>the</strong>r transition metal acetate salts showed interesting results in terms <strong>of</strong> promoting reductive<br />

amination in an acceptable time frame. On <strong>the</strong> o<strong>the</strong>r hand, <strong>the</strong> use <strong>of</strong> <strong>the</strong>se Lewis acids<br />

resulted in <strong>for</strong>mation <strong>of</strong> alcohol byproduct in concentration typically 5-15 area % by GC<br />

analysis. These Lewis acids are : In(OAc) 3 , Sc(OAc) 3 , CuOAc, Er(OAc) 3 , Gd(OAc) 3 ,<br />

Dy(OAc) 3 , AgOAc, Zn(OAc) 2 , and Cd(OAc) 2 , <strong>the</strong>y were tested at 15 mol % concentration.<br />

O<strong>the</strong>r commercially available salts <strong>of</strong> <strong>the</strong>se elements were tested but none <strong>of</strong> <strong>the</strong>m proved<br />

useful as <strong>the</strong> alcohol byproduct concentration was always above 25 area % by GC analysis.<br />

From this study it is obvious that all interesting Lewis acids have acetate as counter ion.<br />

Exceptions to this general observation were noted <strong>for</strong> Bi(OTf) 3 , AgCl, ScCl 3 , and scandium<br />

119


hexafluoroacetylacetone which provided 5-15 area % <strong>of</strong> <strong>the</strong> alcohol by-product and/or<br />

observably longer reaction times than Yb(OAc) 3 (10 mol %), Y(OAc) 3 (15 mol %), or<br />

Ce(OAc) 3 (15 mol %).<br />

Thionylchloride which was tested industrially <strong>for</strong> promoting ketimine <strong>for</strong>mation at 5 mol %<br />

concentration was also tested in our study. [2] Contradicting to <strong>the</strong>ir findings, <strong>the</strong> replacement<br />

<strong>of</strong> <strong>the</strong> Lewis acid by thionylchloride, led to <strong>the</strong> amine product in 50% de, which is even lower<br />

than <strong>the</strong> normally achieved 72% de <strong>for</strong> 2-octanone. Phosphorous oxychloride which has not<br />

been previously reported <strong>for</strong> reductive amination was also tested. Its use at a concentration 10<br />

mol % was beneficial in obtaining 72% de <strong>for</strong> 2-octanone but <strong>the</strong> alcohol by-product was<br />

observed in ~4 area % (GC).<br />

The above listed Lewis acids were available commercially in <strong>the</strong>ir semi-hydrated <strong>for</strong>ms, and<br />

used without fur<strong>the</strong>r purification <strong>for</strong> <strong>the</strong> catalytic screening studies. According to <strong>the</strong> findings<br />

from <strong>the</strong> stoichiometric use Yb(OAc) 3 , drying Yb(OAc) 3 was extremely important <strong>for</strong><br />

consistent results. In <strong>the</strong> initial catalytic screening studies, metalloids Bi(OAc) 3 and<br />

Sb(OAc) 3 were included with Yb(OAc) 3 , Y(OAc) 3 , and Ce(OAc) 3 as useful in inhibiting<br />

alcohol <strong>for</strong>mation. During <strong>the</strong> optimization stage <strong>of</strong> <strong>the</strong> catalytic study I recognized that <strong>the</strong><br />

purchased Bi(OAc) 3 and Sb(OAc) 3 smelled strongly <strong>of</strong> AcOH. This raised <strong>the</strong> question <strong>of</strong><br />

what was catalyzing <strong>the</strong> reductive amination, <strong>the</strong> Lewis acid or co-existing acetic acid. The<br />

Lewis acids [Yb(OAc) 3 , Y(OAc) 3 , Ce(OAc) 3 , Bi(OAc) 3 , and Sb(OAc) 3 ] were <strong>the</strong>n dried until<br />

each maintained a constant weight. Reexamination <strong>of</strong> <strong>the</strong>se dried salts showed Bi(OAc) 3 and<br />

Sb(OAc) 3 were no longer efficient catalysts <strong>for</strong> reductive amination, high alcohol by-product<br />

<strong>for</strong>mation (>15 area %, GC) was noted. In stark contrast, Yb(OAc) 3 , Y(OAc) 3 , and<br />

Ce(OAc) 3 , were as effective as be<strong>for</strong>e, although <strong>the</strong>ir solubility in <strong>the</strong> binary reaction solvent<br />

system, THF-MeOH, was visibly reduced. The effect <strong>of</strong> adding H 2 O (1.0 equiv) to <strong>the</strong><br />

reactions with dried Yb(OAc) 3 , Y(OAc) 3 , or Ce(OAc) 3 , resulted in increased alcohol byproduct<br />

<strong>for</strong>mation. While this intentional addition <strong>of</strong> water was clearly not beneficial, <strong>the</strong><br />

indicated dried Lewis acids were routinely weighed without precaution <strong>for</strong> atmospheric<br />

moisture and had no ill effect on <strong>the</strong> reaction pr<strong>of</strong>ile and reaction time, but dry solvents are<br />

always used <strong>for</strong> <strong>the</strong> reactions.<br />

120


The reactions described above all used 100 wt % Raney-Ni, based on <strong>the</strong> limiting ketone<br />

reagent, and this quantity is not untypical <strong>for</strong> <strong>the</strong> use <strong>of</strong> this heterogeneous hydrogenation<br />

catalyst regarding reductive amination and imine reduction. [3]<br />

I tried to study <strong>the</strong> effect <strong>of</strong> reducing heterogeneous catalyst loading on <strong>the</strong> rate <strong>of</strong> <strong>the</strong><br />

reaction and <strong>the</strong> de. I found that using 50 wt % <strong>of</strong> Raney-Ni in combination with dried<br />

Yb(OAc) 3 did not allow complete consumption <strong>of</strong> <strong>the</strong> ketone (8 area % remained unreacted)<br />

within <strong>the</strong> standard room temperature and 12 h reaction time. It is known that increasing <strong>the</strong><br />

temperature increases reaction rate, so I increased <strong>the</strong> temperature (40 o C) and pressure (20<br />

bar) simultaneously (table 6.2, compare entries 2 and 3), <strong>the</strong> reaction could be completed<br />

be<strong>for</strong>e 12 h without compromising <strong>the</strong> diastereoselectivity <strong>of</strong> <strong>the</strong> secondary amine.<br />

When 25 wt % <strong>of</strong> Raney-Ni was used 50% <strong>of</strong> unreacted ketone was detected after 12 h.<br />

Table 6.2. Raney-Ni loading: Reductive Amination <strong>of</strong> 2-Octanone <strong>for</strong> 12 h a<br />

Entry Raney-Ni (wt %) Ketone (area %) [b] de [b]<br />

1 100 0 72<br />

2 50 8 72<br />

3 [c] 50 0 71<br />

4 25 47 71<br />

a<br />

Reaction conditions: 2-octanone (2.5 mmol), Yb(OAc) 3 (15 mol %), (S)-α-methylbenzylamine (1.1<br />

equiv), Raney-Ni, H 2 (8.3 bar), 0.5 M, 12 h, 22 o C. b GC analysis. c H 2 (20 bar), T= 40 o C.<br />

After finalizing <strong>the</strong> different useful Lewis acids categories, I tested <strong>the</strong> use <strong>of</strong> 10 mol % <strong>of</strong><br />

Yb(OAc) 3 <strong>for</strong> reductive amination <strong>of</strong> different ketone substrates. The following table<br />

summarizes <strong>the</strong> results obtained from our study.<br />

Table 6.3. 10 mol % Yb(OAc) 3 Catalyzed Reductive Amination: Substrate Breadth. a<br />

entry ketone 1<br />

secondary amine 2<br />

yield % b<br />

de % c<br />

121


1<br />

1g<br />

O<br />

2g HN Ph<br />

81<br />

98<br />

2<br />

1h<br />

O<br />

2h HN Ph<br />

78<br />

98<br />

3<br />

1i<br />

O<br />

2i HN Ph<br />

63<br />

94<br />

1b<br />

O<br />

2b<br />

HN<br />

Ph<br />

78<br />

92<br />

5<br />

4 d 92<br />

O<br />

1a<br />

HN<br />

Ph<br />

2a<br />

79<br />

6<br />

1c<br />

O<br />

2c<br />

HN<br />

Ph<br />

87<br />

80<br />

7<br />

1f<br />

O<br />

2f<br />

HN<br />

Ph<br />

82<br />

e<br />

79<br />

8<br />

1d<br />

O<br />

2d<br />

HN<br />

Ph<br />

83<br />

72<br />

9<br />

1e<br />

O<br />

2e<br />

HN<br />

Ph<br />

82<br />

71<br />

a Ketone (2.5 mmol), dried Yb(OAc) 3 (10 mol %), (S)-α-methylbenzylamine (1.1 equiv), Raney Ni, 0.50 M<br />

(MeOH-THF, 1:1), 12 h. Entries 1-4: T= 50 o C, H 2 pressure = 290 psi (20 bar); entry 5: T= 50 o C, H 2 pressure =<br />

120 psi (8.3 bar), entries 6-9: T= 22 o C, H 2 pressure = 120 psi (8.3 bar). b Isolated yield <strong>of</strong> both diastereomers<br />

after chromatography. c Determined by GC analysis <strong>of</strong> <strong>the</strong> crude product. d Pt-C was used instead <strong>of</strong> Raney Ni,<br />

T= 50 o C, H 2 pressure = 120 psi. e Non-Yb(OAc) 3 based methods provide 74% de.<br />

Table 6.3 shows <strong>the</strong> breadth <strong>of</strong> prochiral ketones that serve as good substrates <strong>for</strong> our<br />

catalytic method. As might be expected, based on <strong>the</strong> steric considerations, <strong>the</strong><br />

diastereoselectivity <strong>of</strong> <strong>the</strong> reductive amination product increases in a fairly undisturbed and<br />

linear progression (72–98%), on changing <strong>the</strong> R substituent <strong>of</strong> <strong>the</strong> 2-alkanone, RC(O)CH 3 ,<br />

from a straight chain alkyl group to those having –γ, -β, and finally -α branching (table 6.3,<br />

122


e.g. compare entries 1, 5, 6, and 9). This general trend is interrupted only when <strong>the</strong> R<br />

substituent is t-butyl, e.g. pinacolone (table 6.3, entry 4). Pinacolone again fails to react under<br />

<strong>the</strong> standard Raney-Ni catalyst conditions, but <strong>the</strong> desired product is produced when using Pt-<br />

C as <strong>the</strong> hydrogenation catalyst. It is interesting to note that unlike our earlier findings with<br />

Raney-Ni/Ti(O i Pr) 4 , [3] which allow 2-alkanones with α-branching (table 6.3, entries 1-4) to<br />

be reductively aminated at 22 o C and 120 psi H 2 in 12 h, <strong>the</strong> use <strong>of</strong> Raney-Ni/10 mol %<br />

Yb(OAc) 3 requires <strong>the</strong> more <strong>for</strong>cing conditions <strong>of</strong> 50 o C and 290 psi (20 bar) <strong>for</strong> 12 h<br />

reaction times to be accomplished with complete consumption <strong>of</strong> <strong>the</strong> starting ketone.<br />

Regarding aryl-alkyl ketones, acetophenone (table 6.3, entry 3) was sluggish to react even at<br />

50 o C and 432 psi (30 bar) <strong>of</strong> hydrogen, with isolated yields varying between 60-65% and<br />

concomitant alcohol by-product <strong>for</strong>mation always noted. Examination <strong>of</strong> 1-phenylbutanone at<br />

50 o C and 432 psi (30 bar) <strong>of</strong> hydrogen only allowed ~20 area % (GC) <strong>of</strong> <strong>the</strong> expected<br />

product to <strong>for</strong>m after 24 h. Benzosuberone (cyclic aryl-alkyl ketone) and i-propyl n-propyl<br />

ketone, under similar <strong>for</strong>cing conditions (50 o C, 580 psi (40 bar) H 2 , >24 h), showed that<br />

<strong>the</strong>se sterically challenging substrates could not be reductively aminated.<br />

6.2. Stoichiometric and Catalytic Brønsted Acid Promoted Reductive<br />

Amination<br />

Table 6.4. Brønsted Acid Based Reductive Amination <strong>of</strong> 2-Octanone with (S)-α-MBA. a<br />

Entry Brønsted acid ( 20<br />

mol %)<br />

2-octanone<br />

remaining (area %)<br />

alcohol <strong>for</strong>med (area de b<br />

%) b<br />

1 Acetic acid 1 2 72<br />

2 Trifluoroacetic acid - 22.34 71<br />

3 Trichloroacetic acid 1 2.1 72<br />

4 Formic acid - 2.67 71<br />

5 Oxalic acid 1.65 24 73<br />

6 Thiophenol 51.76 3.47 26<br />

7 Phenol - 29 70<br />

123


a 2-Octanone (2.5 mmol), Brønsted acid (20 mol %), (S)-α-methylbenzylamine (2.75 mmol), Raney Ni, 0.50<br />

M (MeOH), 12 h, T= 22 o C, H 2 pressure = 8.3 bar. b Determined by GC analysis at 12 h.<br />

As mentioned previously, <strong>the</strong> main byproduct <strong>of</strong> reductive amination is alcohol. The use <strong>of</strong><br />

optimum Brønsted or Lewis acids inhibit byproduct <strong>for</strong>mation. Brønsted acids were used<br />

historically on industrial scale <strong>for</strong> reductive amination. Despite <strong>the</strong>ir importance <strong>the</strong>re is a<br />

great lack <strong>of</strong> detailed study <strong>for</strong> useful Brønsted acids in primary literatures.<br />

I decided to test <strong>the</strong> effect <strong>of</strong> using different Brønsted and mineral acids under different<br />

loadings in reductive amination <strong>of</strong> 2-ocatnone. Acetic acid, trichloroacetic acid, or <strong>for</strong>mic<br />

acid at 20 mol % successfully catalyzed reductive amination <strong>of</strong> 2-octanone with α-MBA.<br />

Reducing <strong>the</strong> loading <strong>of</strong> AcOH (5 mol %) had detrimental effect <strong>of</strong> allowing significant<br />

alcohol by-product <strong>for</strong>mation (> 5 area %, GC). When <strong>the</strong> loading <strong>of</strong> acetic acid,<br />

trichloroacetic acid, or <strong>for</strong>mic acid was increased to stoichiometric quantities no<br />

improvement <strong>of</strong> de was noticed compared to <strong>the</strong> use <strong>of</strong> stoichiometric quantities <strong>of</strong><br />

Yb(OAC) 3 (de 87%).<br />

Strong mineral and organic acids were also tested showing different reaction pr<strong>of</strong>ile<br />

compared to o<strong>the</strong>r Brønsted and Lewis acids. The use <strong>of</strong> stoichiometric or catalytic (5 or 10<br />

mol %) quantities <strong>of</strong> 12 N HCl or 18 M H 2 SO 4 , which were diluted in MeOH, p-TsOH, or<br />

trifluoroacetic acid, resulted in high alcohol by-product <strong>for</strong>mation (15-30 area %, GC).<br />

Despite <strong>the</strong> lower yields <strong>of</strong> secondary amine <strong>of</strong> 2-octanone, <strong>the</strong> de was always consistent (70-<br />

72%) and no reduction was noted. In all cases <strong>the</strong> use <strong>of</strong> weak and strong Brønsted acids <strong>for</strong><br />

reductive amination <strong>of</strong> 2-octanone shows <strong>the</strong>m be a minimum <strong>of</strong> 15% lower than when using<br />

110 mol % Yb(OAc) 3 .<br />

Solvent screening was also needed <strong>for</strong> choosing <strong>the</strong> best solvent <strong>for</strong> acetic acid promoted<br />

reductive amination. Protic solvents as MeOH and EtOH were optimal solvents allowing<br />

completing <strong>the</strong> reaction within 8 h. The use <strong>of</strong> THF-MeOH (which was optimal <strong>for</strong><br />

stoichiometric Yb(OAc) 3 study) slowed down <strong>the</strong> reaction (12 h). When THF was used as a<br />

sole solvent with 20 mol % AcOH <strong>for</strong> reductive amination <strong>of</strong> 2-octanone <strong>the</strong> reaction rate<br />

was extremely slow showing 30-45% <strong>of</strong> <strong>the</strong> starting ketone after 24 h. Despite this slow rate<br />

reaction, no alcohol was detected after 24 h only starting ketone. The use <strong>of</strong> THF as sole<br />

124


solvent in <strong>the</strong> stoichiometric and catalytic Lewis acid promoted reactions resulted in complete<br />

reaction within 24 h under identical reaction conditions. This different reactivity pr<strong>of</strong>ile <strong>for</strong><br />

different Brønsted and Lewis acids in protic vs aprotic reaction solvent may allow future<br />

substrates with acid labile functional groups or restricted solubility to be reductively<br />

aminated.<br />

Acetic acid (20 mol %) was used as <strong>the</strong> Brønsted acid <strong>of</strong> choice to be tested <strong>for</strong> <strong>the</strong> reductive<br />

amination. The solvent <strong>of</strong> choice was dry MeOH and different ketones were reductively<br />

aminated as 2-octanone (83% isolated yield, 72% de, T= 22 o C, 120 psi (8.0 bar) H 2 ), i-butyl<br />

methyl ketone (80% isolated yield, 92% de, T= 50 o C, 120 psi (8.0 bar) H 2 ), cyclohexyl<br />

methyl ketone (82% isolated yield, 98% de, T= 50 o C, 290 psi (20 bar) H 2 ), and<br />

acetophenone (55% isolated yield, 93% de, T= 50 o C, 435 psi (30 bar) H 2 ). The reaction rate,<br />

isolated yield and de were almost <strong>the</strong> same as <strong>the</strong> optimal catalytic Lewis acids [Yb(OAc) 3<br />

(10 mol %), Y(OAc) 3 (15 mol %), and Ce(OAc) 3 (15 mol %)] results, but never showed <strong>the</strong><br />

enhanced diastereoselectivity possible when using 50-110 mol % Yb(OAc) 3 .<br />

The use <strong>of</strong> 20 mol % <strong>of</strong> AcOH failed to promote reductive amination <strong>of</strong> sterically hindered<br />

ketone substrates benzosuberone, 1-phenylbutanone, or i-propyl n-propyl ketone, Even under<br />

<strong>for</strong>cing conditions <strong>of</strong> high temperatures (22-50 o C) and hydrogen pressure [120-580 psi (8-40<br />

bar)]. 1-phenylbutanone provided <strong>the</strong> desired product only in very low yield (20-25 area %,<br />

GC) after >24 h <strong>of</strong> reaction. These results add to <strong>the</strong> general conclusion that <strong>for</strong> such<br />

sterically congested substrates <strong>the</strong> only effective system <strong>for</strong> <strong>the</strong>ir reductive amination is <strong>the</strong><br />

use <strong>of</strong> stoichiometric quantities <strong>of</strong> Ti(O i Pr) 4 . [3]<br />

Ti(O i Pr) 4 system has ano<strong>the</strong>r major advantage as it allows α-branched and β-branched methyl<br />

ketones to be reductively aminated at mild condition ( 22 o C and 120 psi within 12 h), on <strong>the</strong><br />

o<strong>the</strong>r hand, <strong>the</strong> same ketones under acetic acid promoted reductive amination require harsher<br />

conditions <strong>of</strong> elevated temperature (50 o C and 120 psi) <strong>for</strong> β-branched 2-alkanones or<br />

elevated temperature and pressure (50 o C and 290 psi) <strong>for</strong> α-branched 2-alkanones. Acetic<br />

acid provides a mediocre yield <strong>for</strong> acetophenone (55%) and no improvement even under<br />

harsher conditions.<br />

125


6.3. Conclusion:<br />

I developed <strong>the</strong> use <strong>of</strong> catalytic quantities <strong>of</strong> Lewis acids in reductive amination <strong>for</strong> <strong>the</strong> first<br />

time. The use <strong>of</strong> catalytic quantities <strong>of</strong> Yb(OAc) 3 or Y(OAc) 3 or Ce(OAc) 3 proved to be<br />

efficient in reducing <strong>the</strong> alcohol by product <strong>for</strong>mation and to produce <strong>the</strong> secondary amine in<br />

good yield and normal de. No enhancement <strong>of</strong> <strong>the</strong> de resulted from <strong>the</strong> use <strong>of</strong> catalytic<br />

quantities <strong>of</strong> Lewis acids. O<strong>the</strong>r Lanthanide salts were also successful in suppressing alcohol<br />

<strong>for</strong>mation but with lower efficiency. Brønsted acids were used historically in reductive<br />

amination but without enough reports on <strong>the</strong>ir role in reductive amination. I have conducted<br />

extensive study on <strong>the</strong> use <strong>of</strong> different Brønsted and mineral acids in reductive amination. 20<br />

mol % <strong>of</strong> acetic acid and <strong>for</strong>mic acid proved to be efficient in suppressing alcohol <strong>for</strong>mation<br />

in reductive amination. The use <strong>of</strong> mineral acids resulted in more alcohol <strong>for</strong>mation (20-30<br />

area % GC). Reductive amination <strong>of</strong> sterically congested ketones and also aromatic ketones<br />

are best per<strong>for</strong>med using Ti(O i Pr) 4 not catalytic amount <strong>of</strong> Lewis acids nor Brønsted acids.<br />

6.4. References:<br />

[1] For advances in <strong>the</strong> diastereoselective reduction <strong>of</strong> (R)- or (S)-α-MBA ketimines, see: (a)<br />

Nichols, D. E.; Barfknecht, C. F.; Rusterholz, D. B. J. Med. Chem. 1973, 16, 480. (b) Clifton,<br />

J. E.; Collins, I.; Hallett, P.; Hartley, D.; Lunts, L. H. C.; Wicks, P. D. J. Med. Chem. 1982,<br />

25, 670. (c) Eleveld, M. B.; Hogeveen, H.; Schudde, E. P. J. Org. Chem. 1986, 51, 3635-<br />

3642. (d) Bringmann, G.; Geisler, J.-P. Syn<strong>the</strong>sis 1989, 608. (e) Marx, E.; El Bouz, M.;<br />

Célérier, J. P.; Lhommet, G. Tetrahedron Lett. 1992, 33, 4307. (f) Moss, N.; Gauthier, J.;<br />

Ferland, J.-M. Synlett 1995, 142. (g) Lauktien, G.; Volk, F.-J.; Frahm, A. W. Tetrahedron:<br />

Asymmetry 1997, 8, 3457. (h) Bisel, P.; Breitling, E.; Frahm, A. W. Eur. J. Org. Chem 1998,<br />

729. (i) Gutman, A. L.; Etinger, M.; Nisnevich, G.; Polyak, F. Tetrahedron: Asymmetry 1998,<br />

9, 4369. (j) Cimarelli, C.; Palmieri, G. Tetrahedron: Asymmetry 2000, 11, 2555. (k) Storace,<br />

L.; Anzalone, L.; Confalone, P. N.; Davis, W. P.; Fortunak, J. M.; Giangiordano, M.; Haley,<br />

J. J. Jr.; Kamholz, K.; Li, H.-Y.; Ma, P.; Nugent, W. A.; Parsons, R. L. Jr.; Sheeran, P. J.;<br />

Silverman, C. E.; Waltermire, R. E.; Wood, C. C. Org. Process Res. Dev. 2002, 6, 54.<br />

126


[2] Farina, V.; Grozinger, K.; Müller-Bötticher, H.; Roth, G. P. Ontazolast: The Evolution <strong>of</strong><br />

a Process. In Process Chemistry in <strong>the</strong> Pharmaceutical Industry; K. G. Gadamasetti, Ed.;<br />

Marcel Dekker, Inc.: New York, 1999, pp 107–124.<br />

[3] a) T. C. Nugent, A. K. Ghosh, V. N. Wakchaure, R. R. Mohanty, Adv. Synth. Catal. 2006,<br />

348, 1289; b) T. C. Nugent, <strong>Chiral</strong> Amine Syn<strong>the</strong>sis - Strategies, Examples, and Limitations.<br />

In Process Chemistry in <strong>the</strong> Pharmaceutical Industry, Second Edition: Challenges in an<br />

Ever-Changing Climate, T. F. Braish, K. Gadamasetti Eds.; CRC Press-Taylor and Francis<br />

Group: New York, 2008; c) T. C. Nugent, A. K. Ghosh, V. N. Wakchaure, R. R. Mohanty,<br />

WO2006030017, 2006; d) T. C. Nugent, V. N. Wakchaure, A. K. Ghosh, R. R. Mohanty,<br />

Org. lett. 2005, 7, 4967; e) T. C. Nugent, A. K. Ghosh, Eur. J. Org. Chem. 2007, 3863.<br />

127


Chapter 7<br />

Stereochemical Considerations <strong>of</strong><br />

Proposed Mechanistic Models<br />

7.1. Introduction.<br />

The outstanding diastereoselectivities reported <strong>for</strong> different amines using stoichiometric<br />

quantities <strong>of</strong> Yb(OAc) 3 and <strong>the</strong> application <strong>of</strong> catalytic quantities <strong>of</strong> Lewis acids <strong>for</strong> first time<br />

in reductive amination <strong>of</strong> prochiral ketones were <strong>the</strong> driving <strong>for</strong>ce <strong>for</strong> investigating <strong>the</strong><br />

mechanistic aspects behind this effect. It is known historically that <strong>the</strong> imine con<strong>for</strong>mation<br />

plays <strong>the</strong> major role in determining <strong>the</strong> stereochemical outcome <strong>of</strong> reductive amination.<br />

Analyzing transition state structures <strong>for</strong> trans-and cis imines which controls <strong>the</strong> facial<br />

selectivity during addition <strong>of</strong> hydrogen to <strong>the</strong> imine allows <strong>the</strong> prediction <strong>of</strong> which<br />

diastereomer will be <strong>for</strong>med in excess (figure 7.1). Diastereomeric excess <strong>of</strong> <strong>the</strong> amine<br />

product depends on which enantiomer <strong>of</strong> α-MBA is used, <strong>the</strong> well known concept <strong>of</strong> allylic<br />

1,3-strain [1] and on <strong>the</strong> earlier proposed models. [2]<br />

Although predictions obtained from this model agree with <strong>the</strong> reaction outcome, o<strong>the</strong>r models<br />

were also introduced aiming to describe <strong>the</strong> reason behind <strong>the</strong> improved diastereoselectivity.<br />

It was suggested previously that <strong>the</strong> reductive amination <strong>of</strong> an α-ketoester with α-MBA may<br />

involve a rotamer (about <strong>the</strong> nitrogen-benzylic carbon bond) with <strong>the</strong> phenyl ring <strong>of</strong> α-MBA<br />

coplanar to <strong>the</strong> imine double bond. [3] This proposed idea agrees with <strong>the</strong> known fact <strong>of</strong> π<br />

bonds affinity <strong>for</strong> heterogeneous hydrogenation catalyst surfaces. [4]<br />

Despite this fact, close examination <strong>of</strong> o<strong>the</strong>r two possible trans-ketimine rotamers, having<br />

phenyl group in a coplanar con<strong>for</strong>mation with <strong>the</strong> imine double bond reveals that one rotamer<br />

suffers from high allylic 1,3-strain resulting from steric crowding <strong>of</strong> phenyl group with<br />

methyl group connected to imine carbonyl carbon <strong>the</strong> o<strong>the</strong>r rotamor is less sterically<br />

128


congested which should be favored. Un<strong>for</strong>tunately this rotamer leads to <strong>the</strong> <strong>for</strong>mation <strong>of</strong> <strong>the</strong><br />

wrong diastereomer. It can be stated that phenyl group is not adsorbed on <strong>the</strong> heterogeneous<br />

surface <strong>of</strong> <strong>the</strong> catalyst during hydrogen addition step.<br />

Ph<br />

Ph<br />

Ph<br />

R H<br />

N<br />

CH 3<br />

H<br />

N<br />

H<br />

NH<br />

Re-face addition <strong>of</strong> hydrogen<br />

to <strong>the</strong> cis-(R)-ketimine<br />

H 2<br />

CH 3<br />

R<br />

cis-(R)-ketimine<br />

R<br />

(S,R)-2<br />

Si-face addition <strong>of</strong> hydrogen<br />

to <strong>the</strong> trans-(R)-ketimine<br />

H 2<br />

CH 3<br />

Ph<br />

Ph<br />

CH 3<br />

R<br />

N<br />

H<br />

Ph<br />

N H<br />

R<br />

trans-(R)-ketimine<br />

HN H<br />

R<br />

(R,R)-2<br />

Figure 7.1. Nitrogen-Benzylic Carbon Bond Rotamers Responsible <strong>for</strong> Hydrogen Addition to<br />

cis- and trans-N-α-MBA Ketimines.<br />

7.1.1. Mechanism Behind Enhanced Stereoselectivity with Yb(OAc) 3 :<br />

Through consulting literatures regarding reductive amination, it can be stated clearly that<br />

<strong>the</strong>re is no precedent <strong>for</strong> Lewis acid enhanced diastereoselectivity during reductive amination<br />

has been reported. Also with respect to <strong>the</strong> recent literatures reporting <strong>the</strong> addition <strong>of</strong> Lewis<br />

acid to an N-α-MBA ketimine no enhanced diastereoselectivity was noted.<br />

To be able to understand <strong>the</strong> origin <strong>of</strong> Yb(OAc) 3 effect on diastereoselectivity, different<br />

experiments have been conducted. In <strong>the</strong>se experiments I tried to focus on detecting or<br />

isolating <strong>the</strong> ketimine intermediates. As stated previously <strong>the</strong> ketimine intermediate<br />

con<strong>for</strong>mation is <strong>the</strong> key factor in determining <strong>the</strong> stereochemical outcome <strong>of</strong> <strong>the</strong> reaction.<br />

Also I tried to test <strong>the</strong> reaction be<strong>for</strong>e completion after certain time intervals and compare <strong>the</strong><br />

results with results obtained from Ti(O i Pr) 4 reaction. Reductive amination <strong>of</strong> 2-octanone with<br />

Ti(O i Pr) 4 or Yb(OAc) 3 requires 12 h <strong>for</strong> complete consumption <strong>of</strong> <strong>the</strong> ketone to occur,<br />

examination <strong>of</strong> <strong>the</strong> reaction at 1 h and 3 h showed <strong>the</strong> diastereoselectivity <strong>of</strong> <strong>the</strong> product to be<br />

129


Table 7.1. Closer Examination <strong>of</strong> <strong>the</strong> Reductive Amination <strong>of</strong> 2-Octanone. a Amine 2d<br />

fully consistent with that found at <strong>the</strong> end <strong>of</strong> <strong>the</strong> reaction (table 7.1). The consistent<br />

diastereoselectivity throughout <strong>the</strong> whole reaction suggests that one mechanism is operating<br />

from <strong>the</strong> first minute till <strong>the</strong> end <strong>of</strong> <strong>the</strong> reaction. This conclusion is applicable on Yb(OAc) 3<br />

,Ti(O i Pr) 3 system and even when no Lewis acid was used. Despite <strong>the</strong> importance <strong>of</strong> this<br />

conclusion <strong>the</strong> origin <strong>of</strong> enhanced diastereoselectivity was not clarified.<br />

entry additive time<br />

(h)<br />

2-octanone<br />

(S)-α-<br />

(%) b<br />

MBA (%)<br />

2-octanol (%) yield (%) de (%)<br />

1 Yb(OAc) 3 1 c<br />

3 d 18.1<br />

12.5<br />

6.5<br />

2.5<br />

0.5<br />

1.0<br />

73.9<br />

81.5<br />

87.1<br />

87.2<br />

2 Ti(O i Pr) 4 1<br />

43.9<br />

7.9<br />

0.63<br />

47.6<br />

67.0<br />

3<br />

8.5<br />

4.0<br />

2.1<br />

85.4<br />

67.0<br />

3 none 1<br />

47.1<br />

10.5<br />

16.6<br />

25.8<br />

72.0<br />

3<br />

29.2<br />

1.1<br />

27.6<br />

42.2<br />

72.1<br />

a 2-octanone (2.5 mmol, 1.0 equiv), (S)-α-MBA (1.1 equiv), Yb(OAc) 3 or Ti(O i Pr) 4 (1.1 equiv) are stirred in<br />

MeOH (1.0 M) <strong>for</strong> 90 min at rt, <strong>the</strong>n THF (final molarity 0.5 M) and Ra-Ni (100 wt %) are added, and <strong>the</strong><br />

reaction pressurized with H 2 (8.3 bar/120 psi). All data is based on GC area % analysis. b Sum <strong>of</strong> (S)-α-MBA<br />

and imine remaining. c 1.0 area % <strong>of</strong> (±)-2-aminooctane was noted. d 2.5 area % (±)-2-aminooctane was noted.<br />

In <strong>the</strong> process <strong>of</strong> searching <strong>for</strong> <strong>the</strong> origin <strong>of</strong> enhanced diastereoselectivity I tried to collect<br />

in<strong>for</strong>mation about <strong>the</strong> in situ imine <strong>for</strong>mation. Samples were taken from mixture <strong>of</strong> 2-<br />

octanone and (S)-α-MBA after 30 min with no additive or with Ti(O i Pr) 4 (1.25 equiv) and<br />

analyzed by GC. In both reactions <strong>the</strong> area % <strong>of</strong> <strong>the</strong> imine was almost similar. Comparing<br />

<strong>the</strong>se results with results obtained from mixing 2-octanone and (S)-α-MBA after 30 min with<br />

Yb(OAc) 3 (1.1 equiv) showed that no appreciable amounts <strong>of</strong> imine (< 3 area %) was<br />

detected when Yb(OAc) 3 was used. Extending <strong>the</strong> reaction time upto 12 h aiming to <strong>for</strong>ce <strong>the</strong><br />

imine <strong>for</strong>mation did not show any success (table 7.2).<br />

Table 7.2. In Situ Imine Formation Study: 2-Octanone and (S)-α-MBA. a<br />

130


imine area % (GC analysis)<br />

time (min) no additive Ti(O i Pr) 4 (1.25 equiv) Yb(OAc) 3 (1.1 equiv)<br />

30 16 38


generated trans- and cis-imine mixture. This means that <strong>the</strong> addition <strong>of</strong> Yb(OAc) 3<br />

may results in enrichment <strong>of</strong> trans- over cis- imines be<strong>for</strong>e hydrogenation. This idea<br />

is only applicable if Yb(OAc) 3 was capable <strong>of</strong> isomerizing some <strong>of</strong> <strong>the</strong> cis-imine to<br />

<strong>the</strong> trans-imine.<br />

To test this proposed idea practically I syn<strong>the</strong>sized N-(S)-MBA ketimine <strong>of</strong> 2-<br />

octanone using Dean-Stark trap. To <strong>the</strong> isolated imine, Yb(OAc) 3 (1.1 equiv) in<br />

anhydrous MeOH was added and stirred <strong>for</strong> half an hour at 22 °C be<strong>for</strong>e <strong>the</strong> addition<br />

<strong>of</strong> Raney-Ni slurry in THF and <strong>the</strong> <strong>the</strong>n <strong>the</strong> whole mixture was hydrogenated at 8.1<br />

bar (120 psi) <strong>for</strong> 12h. GC analysis after 12 h showed that all imine was consumed and<br />

<strong>the</strong> product de was 86%. Repeating <strong>the</strong> same reaction but with extended stirring time<br />

<strong>for</strong> <strong>the</strong> imine with Yb(OAc) 3 resulted in similar de (86%). These results correlate with<br />

<strong>the</strong> earlier obtained results <strong>of</strong> enhanced diastereoselectivity utilizing Yb(OAc) 3 in <strong>the</strong><br />

reductive amination <strong>of</strong> 2-octanone. Fur<strong>the</strong>r examination <strong>of</strong> effect <strong>of</strong> changing reaction<br />

conditions on <strong>the</strong> reaction out come as changing <strong>the</strong> solvent during prestirring period<br />

was also tested. THF was used instead <strong>of</strong> MeOH in <strong>the</strong> prestirring period resulting in<br />

lowering <strong>the</strong> de <strong>of</strong> <strong>the</strong> <strong>for</strong>med amine to 77%. This may be a direct result <strong>of</strong> low<br />

solubility <strong>of</strong> Yb(OAc) 3 in THF allowing <strong>the</strong> back round reaction to occur and<br />

showing <strong>the</strong> importance <strong>of</strong> having MeOH as <strong>the</strong> prestirring solvent. All <strong>the</strong>se findings<br />

support <strong>the</strong> proposed hypo<strong>the</strong>sis <strong>of</strong> Yb(OAc) 3 promoted isomerization <strong>of</strong> <strong>the</strong> in situ<br />

<strong>for</strong>med imine.<br />

In our attempt to clarify <strong>the</strong> role <strong>of</strong> Yb(OAc) 3 in reductive amination <strong>of</strong> prochiral<br />

ketones, I proposed a mechanism <strong>for</strong> <strong>the</strong> in situ imine isomerization (scheme 7.1). The<br />

reaction starts with <strong>the</strong> <strong>for</strong>mation <strong>of</strong> a Lewis acid-base pair between <strong>the</strong> imine and<br />

Yb(OAc) 3 in <strong>the</strong> cis con<strong>for</strong>mation. In this adduct, acetate ligand <strong>of</strong> ytterbium attacks<br />

<strong>the</strong> electrophilic carbonyl carbon <strong>of</strong> iminium ion via a six-membered transition state<br />

<strong>for</strong>ming an oxygen-acetylated hemi-aminal in gauche con<strong>for</strong>mation. This<br />

con<strong>for</strong>mation suffers from steric crowding because <strong>of</strong> <strong>the</strong> gauche relationship between<br />

<strong>the</strong> “α-methylbenzyl” substituent on <strong>the</strong> nitrogen atom and <strong>the</strong> “R” substituent <strong>of</strong> <strong>the</strong><br />

<strong>for</strong>mer carbonyl carbon. To relief this steric strain <strong>the</strong> molecule undergoes pyrimidal<br />

inversion at nitrogen atom. This process is a low energy process which readily occurs<br />

at room temperature.<br />

132


Inversion at <strong>the</strong> nitrogen atom <strong>of</strong> <strong>the</strong> gauche con<strong>for</strong>mation results in <strong>for</strong>mation <strong>of</strong> anti<br />

con<strong>for</strong>mation allowing an anti-relationship to exist between <strong>the</strong> “α-methylbenzyl”<br />

substituent <strong>of</strong> nitrogen and <strong>the</strong> “R” substituent <strong>of</strong> <strong>the</strong> <strong>for</strong>mer carbonyl carbon; this<br />

con<strong>for</strong>mation also allows an antiperiplanar arrangement between <strong>the</strong> nitrogen lone<br />

pair and <strong>the</strong> acetate leaving group, allowing facile elimination <strong>of</strong> acetate and transimine<br />

<strong>for</strong>mation. Through this proposed mechanism I can understand <strong>the</strong> role <strong>of</strong><br />

ytterbium acetate imine isomerization.<br />

Ph<br />

R<br />

OAc<br />

OAc<br />

N<br />

Yb O<br />

O<br />

cis-5<br />

R<br />

Yb<br />

N<br />

O<br />

Ph<br />

O<br />

higher energy<br />

cis-ketimine pathway<br />

R CH 3<br />

N<br />

Yb<br />

Ph<br />

O<br />

O<br />

pyrimidal inversion<br />

at nitrogen<br />

R<br />

Yb<br />

O<br />

N<br />

O<br />

CH 3<br />

Ph<br />

gauche-6<br />

anti-6<br />

lower energy<br />

trans-ketimine pathway<br />

Ph<br />

OAc<br />

OAc<br />

R<br />

N<br />

O<br />

Yb<br />

O<br />

Ph<br />

Yb<br />

N<br />

R O<br />

trans-5<br />

O<br />

Scheme 7.1. Proposed Mechanism <strong>for</strong> In Situ Isomerization <strong>of</strong> <strong>the</strong> Ketimine during<br />

Reductive Amination<br />

7.1.2. Reasons Behind Enhanced Diastereoselectivity <strong>for</strong> Different<br />

133


Substrate Categories.<br />

I found that <strong>the</strong> biggest jump <strong>for</strong> de is associated with straight-chain 2-alkanones (e.g.<br />

2-octanone) and γ-branched 2-alkanones (e.g. benzylacetone). These groups <strong>of</strong><br />

substrates are good substrates <strong>for</strong> this methodology in terms <strong>of</strong> high<br />

diastereoselectivity. α- and β-branched 2-alkanones did not show any significant<br />

increase in <strong>the</strong> de. This effect can be rationalized through understanding <strong>the</strong> nature <strong>of</strong><br />

R group attached to <strong>of</strong> <strong>the</strong> two Newman projections in gauche and anti con<strong>for</strong>mations<br />

illustrated in (scheme 7.2).<br />

When α- and β-branched 2-alkanones (when R α or R β = alkyl respectively) are used<br />

with Yb(OAc) 3 , no improvement <strong>of</strong> diastereoselectivity was noticed as Illustrated<br />

from <strong>the</strong> scheme. This finding implies that <strong>the</strong> energy difference between <strong>the</strong> gauche<br />

and <strong>the</strong> anti con<strong>for</strong>mations is not significant. This means that steric crowding <strong>of</strong> <strong>the</strong><br />

“α-methylbenzyl” substituent on <strong>the</strong> nitrogen atom and <strong>the</strong> “R” substituent <strong>of</strong> <strong>the</strong><br />

<strong>for</strong>mer carbonyl carbon in <strong>the</strong> gauche con<strong>for</strong>mation vs <strong>the</strong> steric crowding<br />

experienced when <strong>the</strong> “ytterbium” substituent <strong>of</strong> <strong>the</strong> nitrogen atom is gauche <strong>the</strong> “R”<br />

substituent <strong>of</strong> <strong>the</strong> <strong>for</strong>mer carbonyl carbon <strong>of</strong> <strong>the</strong> anti con<strong>for</strong>mation has almost <strong>the</strong><br />

same energy. On <strong>the</strong> o<strong>the</strong>r hand, examination <strong>of</strong> <strong>the</strong> energy difference <strong>of</strong> <strong>the</strong> gauche<br />

and anti con<strong>for</strong>mations <strong>of</strong> γ-branched 2-alkanones and straight chain 2- alkanones<br />

shows that <strong>the</strong> anti congregation is lower in energy compared to <strong>the</strong> gauche<br />

con<strong>for</strong>mation. The effect is more pronounced in straight chain 2-alkanones compared<br />

to γ-branched 2-alkanones. This difference in energy can be rationalized by noting<br />

that <strong>the</strong> “α-methylbenzyl” substituent <strong>of</strong> nitrogen is sterically very crowded α to <strong>the</strong><br />

nitrogen, while <strong>the</strong> ytterbium-nitrogen bond will be expected to be longer than <strong>the</strong><br />

benzylic carbon-nitrogen bond <strong>of</strong> <strong>the</strong> “α-methylbenzyl” substituent and <strong>the</strong>reby<br />

reduce <strong>the</strong> immediate steric volume next to nitrogen.<br />

The conclusion is that regardless <strong>of</strong> <strong>the</strong> steric bulk <strong>of</strong> <strong>the</strong> “R” substituent, <strong>the</strong>re is<br />

always a high degree <strong>of</strong> steric crowding when <strong>the</strong> “α-methylbenzyl” substituent on<br />

nitrogen is gauche to it (Figure 2, gauche-6). To account <strong>for</strong> <strong>the</strong> observed<br />

enhancement in de, “R” substituents (Scheme 3) having only γ-branching (Figure 3,<br />

134


anti-6) would be expected to have medium steric crowding with a gauche ytterbium<br />

atom, while non-branched “R” substituents would have low steric crowding in<br />

relation to a gauche positioned ytterbium atom. These considerations would thus favor<br />

cis-imine to trans-imine isomerization <strong>for</strong> straight-chain 2-alkanones and γ-branched<br />

2-alkanones, but exclude isomerization <strong>for</strong> α- and β-branched 2-alkanone substrates.<br />

R γ<br />

R α<br />

R α<br />

R<br />

R β<br />

Ph<br />

N<br />

O<br />

CH 3<br />

Yb<br />

O<br />

~=<br />

R<br />

R γ<br />

R β<br />

Yb<br />

O<br />

N<br />

O<br />

CH 3<br />

Ph<br />

gauche-6<br />

R α or R β = alkyl<br />

anti-6<br />

R γ<br />

R γ<br />

R<br />

Ph<br />

N<br />

O<br />

CH 3<br />

Yb<br />

O<br />

><br />

R<br />

Yb<br />

O<br />

N<br />

O<br />

CH 3<br />

Ph<br />

gauche-6<br />

R α and R β = H<br />

anti-6<br />

R<br />

Ph<br />

N<br />

O<br />

CH 3<br />

Yb<br />

O<br />

>><br />

R<br />

Yb<br />

O<br />

N<br />

O<br />

CH 3<br />

Ph<br />

gauche-6<br />

R α , R β , and R γ = H<br />

anti-6<br />

Scheme 7.2. Newman Projections Showing Interactions <strong>of</strong> Ytterbium Acetate with R<br />

Group.<br />

Because no o<strong>the</strong>r lanthanide acetates were identified as capable <strong>of</strong> providing<br />

enhanced de, ytterbium would appear to have unique coordination sphere attributes<br />

(Lewis acidity and high coordination number) allowing <strong>the</strong> proposed in situ<br />

isomerization to occur. Fur<strong>the</strong>rmore a unique and critical role is indicated <strong>for</strong> <strong>the</strong><br />

acetate ligand. The mechanism presented here is more likely than those in which<br />

simple ligation <strong>of</strong> Yb(OAc) 3 to <strong>the</strong> in situ <strong>for</strong>med trans- and cis-imine mixture allows<br />

constructive influence <strong>of</strong> <strong>the</strong> rotamer environment, about <strong>the</strong> benzylic carbon-nitrogen<br />

bond <strong>of</strong> <strong>the</strong> ketimine, and <strong>the</strong>reby improved facial selectivity during reduction. This is<br />

135


supported by <strong>the</strong> fact that o<strong>the</strong>r ytterbium salts, e.g. YbCl 3 and Yb(OTf) 3 , were<br />

ineffective at providing efficient product <strong>for</strong>mation and could not do so with enhanced<br />

diastereoselectivity. Fur<strong>the</strong>rmore none <strong>of</strong> <strong>the</strong> o<strong>the</strong>r Lanthanides or transition metals<br />

examined allowed enhanced diastereoselectivity.<br />

Lastly, Yb(OTf) 3 could in <strong>the</strong>ory undergo a similar imine isomerization process via its<br />

sulfonate oxygen, but as noted earlier only provided <strong>the</strong> alcohol by-product under <strong>the</strong><br />

reductive amination conditions noted here, implying it is too Lewis acidic. To fur<strong>the</strong>r<br />

probe this, I pre<strong>for</strong>med <strong>the</strong> N-α-MBA ketimine <strong>of</strong> 2-octanone and <strong>the</strong>n added<br />

Yb(OTf) 3 (110 mol%) to it. After 30 min Raney Ni (THF slurry) was added followed<br />

by <strong>the</strong> onset <strong>of</strong> hydrogenation, <strong>the</strong> desired product was observed in 50% de.<br />

7.1.3. Key Findings <strong>for</strong> Reductive Amination with α-MBA<br />

These studies, and our earlier ones, complete a body <strong>of</strong> research regarding <strong>the</strong><br />

reductive amination <strong>of</strong> prochiral ketones with (R)- or (S)-α-MBA (1.1 equiv) under<br />

<strong>the</strong> influence <strong>of</strong> an optimal Lewis acid or Brønsted acid. The findings show that<br />

prochiral alkyl-alkyl' and aryl-alkyl ketones can be readily reductively aminated, and<br />

by doing so higher yields and much shorter reaction times are achieved compared to<br />

<strong>the</strong> previously practiced two-step strategy via pre<strong>for</strong>med (R)- or (S)-α-methylbenzyl<br />

ketimines. [5]<br />

Fur<strong>the</strong>rmore, <strong>the</strong> reducing system <strong>of</strong> choice is a heterogeneous hydrogenation catalyst<br />

in <strong>the</strong> presence <strong>of</strong> hydrogen, due to <strong>the</strong> superior diastereoselectivity af<strong>for</strong>ded with<br />

good isolated yield and reaction time vs all o<strong>the</strong>r reducing systems examined to<br />

date. [2]<br />

Use <strong>of</strong> <strong>the</strong> correct acid (Lewis or Brønsted) catalyst is crucial <strong>for</strong> a successful<br />

outcome when reductively aminating a prochiral ketone with α-MBA in <strong>the</strong> presence<br />

<strong>of</strong> Raney-Ni (generally <strong>the</strong> most useful heterogeneous catalyst) and hydrogen (120<br />

psi). Failure to have <strong>the</strong> optimal Lewis acid or Brønsted acid, or no acid at all, results<br />

in gross alcohol by-product <strong>for</strong>mation. Adding catalytic quantities <strong>of</strong> Yb(OAc) 3 ,<br />

Y(OAc) 3 , Ce(OAc) 3 , or catalytic or stoichiometric quantities <strong>of</strong> a weak Brønsted acid,<br />

136


e.g. AcOH, suppresses alcohol by-product <strong>for</strong>mation <strong>for</strong> 2-alkanones below 2%,<br />

providing <strong>the</strong> desired amine product in good yield. Application <strong>of</strong> <strong>the</strong> catalytic Lewis<br />

acid or Brønsted acid systems to aryl-alkyl ketones reveals that only acetophenone<br />

will react and <strong>the</strong>n only under <strong>for</strong>cing conditions (50 ° C, 30 bar) with low yield (63%<br />

and 55 % respectively) <strong>of</strong> <strong>the</strong> desired product.<br />

When stoichiometric quantities <strong>of</strong> Ti(O i Pr) 4 , a Lewis acid, are used <strong>for</strong> reductive<br />

amination <strong>the</strong> reactions are complete within <strong>the</strong> same reaction time and again alcohol<br />

by-product <strong>for</strong>mation is suppressed below 2%; but unlike <strong>the</strong> above mentioned<br />

systems, which require elevated temperature (50 ° C) and/or H 2 pressure (290 psi) <strong>for</strong><br />

α-branched (R L C(O)CH 3 ) and β-branched (R M C(O)CH 3 ) ketones, Ti(O i Pr) 4 only<br />

requires 22 ° C and 120 psi <strong>for</strong> <strong>the</strong>se hindered 2-alkanones. Additionally, aryl-alkyl<br />

ketones and more sterically demanding alkyl-alkyl' ketones, e.g. i-propyl n-propyl<br />

ketone, can be reductively aminated in good yield and de when using Ti(O i Pr) 4 .<br />

When comparing <strong>the</strong> de <strong>of</strong> <strong>the</strong> reductive amination products that are common to<br />

Ti(O i Pr) 4 , Brønsted acids (catalytic or stoichiometric, e.g. AcOH), Yb(OAc) 3 (10 mol<br />

%), Y(OAc) 3 (15 mol %), and Ce(OAc) 3 (15 mol %), <strong>the</strong> de <strong>of</strong> <strong>the</strong> amine product is<br />

<strong>the</strong> same. Fur<strong>the</strong>rmore, if pre<strong>for</strong>med (R)- or (S)-α-MBA ketimines are reductively<br />

aminated <strong>the</strong> same de is observed as when <strong>the</strong> above noted Lewis or Brønsted acids<br />

catalysts are used <strong>for</strong> reductive amination <strong>of</strong> <strong>the</strong> corresponding ketone. In stark<br />

contrast to <strong>the</strong>se stereoselectivity trends, 2-alkanones without branching at <strong>the</strong> α- or<br />

β-carbons, e.g. 2-octanone or benzylacetone, can be reductively aminated with<br />

dramatically increased diastereoselectivity when using as little as 50 mol %<br />

Yb(OAc) 3 , again alcohol by-product <strong>for</strong>mation is suppressed below 2% and good<br />

yields are always realized. These combined findings are summarized in table 7.3.<br />

Table 7.3. Useful Substrate Classes, Optimal Acid Catalysts, and Trends <strong>for</strong> α-<br />

MBA Reductive Amination a<br />

137


ketone class subclass examples de acid catalyst b comment<br />

O R L = i-Pr or c-hexyl 98 Ti(O i Pr) 4 viable alternative<br />

AcOH c<br />

R L CH 3<br />

R L = Ph 95 Ti(O i Pr) 4 o<strong>the</strong>r catatalysts - low<br />

yield<br />

R L<br />

O R L = Ph; R S = n-Pr 90 Ti(O i Pr) 4 o<strong>the</strong>r catalysts - no<br />

product<br />

R S<br />

R L = i-Pr; R S = n-Pr,<br />

n-Bu<br />

87 Ti(O i Pr) 4 o<strong>the</strong>r catalysts - no<br />

product<br />

O R M = i-Bu 93 Ti(O i Pr) 4 viable alternative<br />

AcOH d<br />

R M CH 3<br />

R M = -CH 2 CH 2 Ph 89 Yb(OAc) 3 o<strong>the</strong>r catalysts - low de<br />

O R S = n-hexyl 87 Yb(OAc) 3 o<strong>the</strong>r catalysts - low de<br />

R S CH 3<br />

R S = n-butyl 85 Yb(OAc) 3 o<strong>the</strong>r catalysts - low de<br />

a<br />

Unless o<strong>the</strong>rwise noted, all reactions per<strong>for</strong>med at 22 o C and 120 psi H 2 . The indicated ketone<br />

classes and subclasses provide a starting point <strong>for</strong> assessing near optimal conditions <strong>for</strong> similar<br />

substrates, see reference 1c and this manuscript <strong>for</strong> specific details. b For optimal yield and de,<br />

Ti(O i Pr) 4 is always used in stoichiometric quantities while Yb(OAc) 3 can be used in 50-110 mol %. c<br />

The use <strong>of</strong> 20 mol% AcOH allows very similar results, but only at 50 o C and 290 psi (H 2 ). d The use<br />

<strong>of</strong> 20 mol% AcOH allows very similar results, but only at elevated temperature (50 o C).<br />

7.2. Conclusion:<br />

Strategies <strong>for</strong> α-chiral amine syn<strong>the</strong>sis employing pre<strong>for</strong>med imines or enamines are<br />

stepwise long and can suffer from lower overall yield because <strong>of</strong> mediocre imine or<br />

enamine yield <strong>for</strong>ming steps, as previously commented on. This problems can be<br />

alleviated by using a reductive amination strategy, as outlined here, and avoids <strong>the</strong><br />

normally stepwise excessive procedures <strong>of</strong> chiral auxiliary approaches by<br />

simultaneously incorporating a nitrogen atom (from <strong>the</strong> auxiliary) and a new<br />

stereogenic center at <strong>the</strong> carbonyl carbon during step one (reductive amination). A<br />

second step, hydrogenolysis, allows <strong>the</strong> enantioenriched primary amine to be isolated<br />

in good to high overall yield.<br />

The ytterbium acetate method expounded on here unequivocally demonstrates <strong>the</strong> first<br />

example <strong>of</strong> constructive interference, by any additive, during <strong>the</strong> asymmetric<br />

138


eductive amination <strong>of</strong> a prochiral ketone or an N-α-MBA ketimine. It also represents<br />

<strong>the</strong> first documented use <strong>of</strong> ytterbium <strong>for</strong> reductive amination. The initial mechanistic<br />

investigations elaborated on here suggest an imine isomerization pathway promoted<br />

by Yb(OAc) 3 allowing enhanced diastereoselectivity during reductive amination. A<br />

future study will be required to elaborate on <strong>the</strong>se initial mechanistic proposals, would<br />

likely require computational analysis and include <strong>the</strong> investigation <strong>of</strong> chiral and o<strong>the</strong>r<br />

achiral Lewis acid derivatives. Investigation <strong>of</strong> heterogeneous hydrogenation catalysts<br />

prepared by different methods or supported on different materials (e.g. carbon<br />

nanostructures, alumina, etc.), could provide fur<strong>the</strong>r beneficial insights. Finally, <strong>the</strong><br />

general phenomenon <strong>of</strong> in situ promoted imine isomerization, with Yb(OAc) 3 , would<br />

be expected to have a beneficial impact on <strong>the</strong> study <strong>of</strong> imine/enamine chemistry in<br />

general, e.g. in conjunction with enantioselective organocatalysis.<br />

7.3. References:<br />

[1] (a) R. W. H<strong>of</strong>fmann, Chem Rev. 1989, 89, 1841. (b) K. W. Lee, S. Y. Hwang, C.<br />

R. Kim, D. H. Nam, J. H. Chang, S. C. Choi, B. S. Choi, H. -W. Choi, K. K. Lee, B.<br />

So, S.W. Cho, H. Shin, Org. Process Res. Dev. 2003, 7, 839.<br />

[2] (a) M. B. Eleveld, H. Hogeveen, E.P. Schudde, J. Org. Chem. 1986, 51, 3635; (b)<br />

A. L. Gutman, M. Etinger, G. Nisnevich, F. Polyak, Tetrahedron: Asymmetry 1998, 9,<br />

4369.<br />

[3] G. Siedlaczek, M. Schwickardi, U. Kolb, B. Bogdanovic, D. G. Blackmond, Catal. Lett.<br />

1998, 55, 67.<br />

[4] (a) A. Kraynov, A. Suchopar, L. D'Souza, R. Richards, Physical Chemistry<br />

Chemical Physics, 2006, 8, 1321. (b) T. Bürgi, A. Baiker Acc. Chem. Res. 2004, 37,<br />

909. (c) M. Studer, H. –U. Blaser, C. Exner, Adv. Synth. Catal. 2003, 345, 45.<br />

[5] For advances in <strong>the</strong> diastereoselective reduction <strong>of</strong> (R)- or (S)-α-MBA ketimines,<br />

see: (a) D. E. Nichols,C. F. Barfknecht, D. B. Rusterholz, J. Med. Chem. 1973, 16,<br />

480. (b) J. E.Clifton, I. Collins, P. Hallett, D. Hartley, L. H. C. Lunts, P.D. Wicks, J.<br />

Med. Chem. 1982, 25, 670. (c) M. B. Eleveld, H. Hogeveen, E. P. Schudde, J. Org.<br />

Chem. 1986, 51, 3635-3642. (d) G. Bringmann, J. –P. Geisler, Syn<strong>the</strong>sis 1989, 608.<br />

(e) E. Marx, M. El Bouz,J. P. Célérier, G. Lhommet, Tetrahedron Lett. 1992, 33,<br />

4307. (f) N. Moss, J. Gauthier, J. –M. Ferland, Synlett 1995, 142. (g) G. Lauktien, F. –<br />

139


J. Volk, A. W. Frahm, Tetrahedron: Asymmetry 1997, 8, 3457. (h) P. Bisel, E.<br />

Breitling, A. W. Frahm, Eur. J. Org. Chem 1998, 729. (i) A. L. Gutman, M. Etinger,<br />

G. Nisnevich, F. Polyak, Tetrahedron: Asymmetry 1998, 9, 4369. (j) C. Cimarelli, G.<br />

Palmieri, Tetrahedron: Asymmetry 2000, 11, 2555. (k) L. Storace, L. Anzalone, P. N.<br />

Confalone, W. P. Davis, J. M. Fortunak, M. Giangiordano, J. J. Jr. Haley, K.<br />

Kamholz, H. –Y. Li, P. Ma, W. A. Nugent, R. L. Parsons, Jr.; P.J. Sheeran, C. E.<br />

Silverman, R. E. Waltermire, C. C. Wood, Org. Process Res. Dev. 2002, 6, 54.<br />

140


Appendix<br />

Experimental Section<br />

General Remarks<br />

NMR spectra were recorded on a JEOL ECX 400 spectrometer, operating at 400 MHz ( 1 H)<br />

and 100 MHz ( 13 C) respectively. Chemical shifts (δ) were reported in parts per million (ppm)<br />

downfield from TMS (= 0) or relative to CHCl 3 (7.26 ppm) or D 2 O (4.79 ppm) <strong>for</strong> 1 H NMR.<br />

For 13 C NMR, chemical shifts were reported in <strong>the</strong> scale relative to CHCl 3 (77.0 ppm) or D 2 O<br />

(CH 3 OH internal reference, δ= 49.5 ppm) as an internal reference. Multiplicities are<br />

abbreviated as: s, singlet; d, doublet; t, triplet; q, quartet; m, multiplet; br, broad. The<br />

coupling constants are expressed in Hz. FTIR spectra were obtained on Nicolet Avatar 370<br />

spectrometer. Mass spectra were recorded on a Finnigan MAT 95 (EI) with an ionization<br />

potential <strong>of</strong> 70 eV. Elemental analyses were per<strong>for</strong>med by an external vendor in Lindlar,<br />

Germany on an Elementar Vario EL III instrument. For amine products 2, reaction progress<br />

and diastereomeric excess measurements were obtained using a Shimadzu GC-2010<br />

instrument with a Rtx-5 amine column (Restec, 30 m x 0.25 mm); T inj = 300 °C and T det =<br />

300 °C, and carrier gas He @ 24 psi were always constant. Program A: 50 °C (1 min), <strong>the</strong>n<br />

14 °C/min to 280 °C (hold 2 min); Program B: 50 °C (1 min), <strong>the</strong>n 14 °C/min to 130 °C (hold<br />

9 min), <strong>the</strong>n 20 °C/min to 280 °C (hold 2 min); Program C: 50 °C (1 min); <strong>the</strong>n 14 °C/min to<br />

280 °C (hold 1 min); Program D: 50 °C (1 min); <strong>the</strong>n 14 °C/min to 280 °C (hold 5 min). For<br />

hydrogenolyzed product primary amine 4d <strong>the</strong> enantiomeric excess <strong>of</strong> <strong>the</strong> trifluoroacetamide<br />

derivative was determined by gas chromatography using a Shimadzu GC-2010 instrument on<br />

a <strong>Chiral</strong>dex B-DP column (Astec, 30 m x 0.25mm); T inj = 200 °C, T det = 200 °C, and carrier<br />

gas He @ 24 psi were constant. Program E: 130 °C (20 min), <strong>the</strong>n 20 °C/min to 180 °C (hold<br />

10 min), split ratio 60:1. Column chromatography was per<strong>for</strong>med using silica gel 60 (0.040-<br />

0.063 mm). Thin-layer chromatography (TLC) was per<strong>for</strong>med using precoated plates <strong>of</strong> silica<br />

gel 60 F 254 and visualized under ultraviolet irradiation (254 nm).<br />

All reactions were per<strong>for</strong>med under an inert atmosphere (nitrogen). All reagents were<br />

obtained from Sigma-Aldrich (except cyclohexyl methyl ketone, obtained from ABCR<br />

GmbH &Co) and used without fur<strong>the</strong>r purification. Be<strong>for</strong>e using <strong>the</strong> commercially purchased<br />

141


Yb(OAc) 3 (Aldrich catalog number 544973, 99.999% grade), Y(OAc) 3 (Aldrich catalog<br />

number, 326046, 99.9% grade), and Ce(OAc) 3 (Aldrich catalog number, 529559, 99.999%<br />

grade ), each was dried at 80 °C under high vacuum until a constant weight was achieved (12<br />

h). The dried Lewis acids could be stored in dry screw cap glass bottles at room temperature,<br />

and <strong>the</strong>se containers could be repeatedly opened to <strong>the</strong> atmosphere (at least 6 times without<br />

detrimental effect) without special precaution or need <strong>for</strong> a glovebox. In this way constant<br />

and repeatable results were always observed. The (S)-α-methylbenzylamine (Aldrich catalog<br />

number, 115568) was <strong>of</strong> 98% chemical purity and 98% ee. The Raney-Nickel (in water) was<br />

purchased from Fluka (Catalog number, 83440). Pd(OH) 2 /C [≤ 50% water, 20 wt % loading<br />

(dry basis)] was purchased from Aldrich (catalog number, 212911). Pt/C (1-4% water, 5 wt<br />

% loading) was purchased from Aldrich (Catalog number, 205931).<br />

Experimental Section<br />

Syn<strong>the</strong>sis <strong>of</strong> N-(S)-α-MBA ketimine <strong>of</strong> 2-octanone<br />

p-Toluene sulphonic acid (2 mol %, 80 mg) was added to a double neck 100 mL round<br />

bottom flask, 60 mL <strong>of</strong> toluene was added, 2-octanone (22 mmol, 1.00 equiv, 3.45 mL), and<br />

(S)-α-methylbenzylamine (24.2 mmol, 1.10 equiv, 3.08 mL) were added to <strong>the</strong> flask. The<br />

flask was connected to a Dean-Stark trap which was connected to a refluxing condenser. The<br />

mixture was refluxed <strong>for</strong> 24 h at 120 °C. The mixture was allowed to cool and <strong>the</strong>n <strong>the</strong><br />

toluene was evaporated under vacuum. The residue was dissolved in hexane (~30 mL), <strong>the</strong><br />

solution was passed through filter paper and into a separatory funnel. Aqueous NaHCO 3 (1.0<br />

M, 40 mL) was added and after very brief mixing, <strong>the</strong> hexane layer was separated, washed<br />

with brine, dried over MgSO 4 and filtered. The organic layer was concentrated under rotary<br />

evaporation <strong>the</strong>n under high vacuum at 40 °C (with stirring) <strong>for</strong> 24 h (3.7 g, 57% yield). GC<br />

analysis showed 96 area % <strong>of</strong> <strong>the</strong> imine and 4 area % <strong>of</strong> <strong>the</strong> starting ketone and amine. This<br />

imine was used <strong>for</strong> <strong>the</strong> fur<strong>the</strong>r experiments.<br />

General procedure <strong>for</strong> <strong>the</strong> reduction <strong>of</strong> <strong>the</strong> N-(S)-α-MBA ketimine <strong>of</strong> 2-octanone<br />

The imine (2.0 mmol, 462 mg) was added to a hydrogenation vessel, and <strong>the</strong>n anhydrous<br />

MeOH (2.5 mL) or THF (2.5 mL) was added. Dried Yb(OAc) 3 or Yb(OTf) 3 was <strong>the</strong>n added,<br />

142


and <strong>the</strong> mixture was stirred <strong>for</strong> 30 min. A THF slurry <strong>of</strong> Raney-Ni (100 wt % based on <strong>the</strong><br />

imine, pre-triturated with EtOH (×3) and <strong>the</strong>n with anhydrous THF (×3) be<strong>for</strong>e addition)<br />

(final reaction molarity 0.4 M) was <strong>the</strong>n added and <strong>the</strong> reaction vessel pressurized at 120 psi<br />

(8.3 bar) <strong>of</strong> hydrogen. GC samples were worked-up using NaHCO 3 / EtOAc.<br />

General procedure: Stoichiometric Yb(OAc) 3 (enhanced de)<br />

In a dry reaction vessel Yb(OAc) 3 (0.96 g, 2.75 mmol, 1.1 equiv) was added and<br />

subsequently evacuated under high vacuum <strong>for</strong> 5 min be<strong>for</strong>e flooding with nitrogen,<br />

anhydrous MeOH (2.5 mL, 1.0 M) was <strong>the</strong>n added. To this solution a prochiral ketone 1 (2.5<br />

mmol, 1.0 equiv) and (S)-α-methylbenzylamine (0.35 mL, 2.75 mmol, 1.1 equiv) were added<br />

and subsequently stirred at room temperature <strong>for</strong> 20-30 min. A THF slurry <strong>of</strong> Raney-Ni (100<br />

wt % based on <strong>the</strong> ketone, pre-triturated with EtOH (×3) and <strong>the</strong>n with anhydrous THF (×3)<br />

be<strong>for</strong>e addition) was transferred to <strong>the</strong> reaction mixture using 2.5 mL <strong>of</strong> anhydrous THF<br />

(final molarity <strong>of</strong> reaction solution is 0.5 M) and <strong>the</strong> reaction vessel pressurized at 120 psi<br />

(8.3 bar) <strong>of</strong> hydrogen. After 12 h at 22 o C,


(2S)-4-Methyl-N-((S)-1-phenylethyl)pentan-2-amine (2a)<br />

Reaction details: Yb(OAc) 3 (1.1 equiv), 4-methyl-2-pentanone (0.31 mL, 2.5 mmol, 1.0<br />

equiv), (S)-α-methylbenzylamine (0.35 mL, 2.75 mmol, 1.1 equiv). Reaction time: 12 h; 94%<br />

de. Purification by silica gel flash chromatography (hexanes/EtOAc/NH 4 OH, 83:15:2) gave<br />

<strong>the</strong> mixture <strong>of</strong> diastereomers as a colorless viscous liquid, which was <strong>the</strong>n treated with e<strong>the</strong>ral<br />

HCl to obtain <strong>the</strong> hydrochloride salt (0.467 g, 78% yield) after high vacuum drying. GC<br />

(program A, see: Experimental section (general remarks)): retention time [min]: major (S,S)-<br />

2a isomer, 10.9; minor (R,S)-2a isomer, 10.6. The NMR data <strong>of</strong> (S,S)-2a (free base) matches<br />

that reported in <strong>the</strong> literature. [1]<br />

Major (S,S)-2a: 1 H NMR (400 MHz, CDCl 3 ): δ 7.36-7.20 (m, 5H), 3.89 (q, J = 6.8 Hz, 1H),<br />

2.60-2.52 (m, 1H), 1.63-1.56 (m, 1H), 1.43-1.36 (m 1H), 1.32 (d, J = 6.8 Hz, 3H), 1.13-1.07<br />

(m, 1H), 0.94 (d, J = 6.4 Hz, 3H), 0.89 (d, J = 6.8 Hz, 3H), 0.76 (d, J = 6.8 Hz, 3H). 13 C<br />

NMR (100 MHz, CDCl 3 ): δ 146.5, 128.3, 126.7, 126.5, 55.2, 48.4, 46.6, 25.1, 24.4, 23.6,<br />

22.3, 21.6.<br />

(2S)-4-Phenyl-N-((S)-1-phenylethyl)butan-2-amine (2c)<br />

Reaction details: Yb(OAc) 3 (1.1 equiv), 4-phenyl-2-butanone (0.37 mL, 2.5 mmol, 1.0<br />

equiv), (S)-α-methylbenzylamine (0.35 mL, 2.75 mmol, 1.1 equiv). Reaction time: 12 h; 89%<br />

de. Purification by silica gel flash chromatography (hexanes/EtOAc/NH 4 OH, 78:20:2) gave<br />

<strong>the</strong> mixture <strong>of</strong> diastereomers as a colorless viscous liquid, which was <strong>the</strong>n treated with e<strong>the</strong>ral<br />

HCl to obtain <strong>the</strong> hydrochloride salt (0.625 g, 87% yield) after high vacuum drying. GC<br />

(program D, see Experimental section (general remarks)): retention time [min]: major (S,S)-<br />

2c isomer, 15.5; minor (R,S)-2c isomer, 15.4. The NMR data <strong>of</strong> (S,S)-2c (free base) matches<br />

that reported in <strong>the</strong> literature. [1]<br />

Major (S,S)-2c: 1 H NMR (CDCl3, 400 MHz): δ 7.30-7.14 (m, 10 H), 3.85 (q, J = 6.4 Hz,<br />

1H), 2.69-2.49 (m, 3 H), 1.85-1.83 (m, 1H), 1.63-1.55 (m, 1H), 1.29 (d, J = 6.4 Hz, 3H), 1.02<br />

(d, J = 6.4 Hz, 3H). 13 C NMR (CDCl3, 100 MHz): δ 146.3, 142.5, 128.3, 128.2, 126.6, 126.4,<br />

125.6, 54.9, 49.6, 37.9, 31.9, 24.4, 21.2.<br />

(2S)-N-((S)-1-Phenylethyl)octan-2-amine (2d)<br />

144


Reaction details: Yb(OAc) 3 (1.1 equiv), 2-octanone (0.39 mL, 2.5 mmol, 1.0 equiv), (S)-αmethylbenzylamine<br />

(0.35 mL, 2.75 mmol, 1.1 equiv). Reaction time: 12 h; 87% de.<br />

Purification by silica gel flash chromatography (hexanes/EtOAc/NH 4 OH, 78:20:2) gave <strong>the</strong><br />

mixture <strong>of</strong> diastereomers as a colorless viscous liquid, which was <strong>the</strong>n treated with e<strong>the</strong>ral<br />

HCl to obtain <strong>the</strong> hydrochloride salt (0.580 g, 86% yield) after high vacuum drying. GC<br />

(program A, see Experimental section (general remarks)): retention time [min]: major (S,S)-<br />

2d isomer, 12.9; minor (R,S)-2d isomer, 13.1. The NMR data <strong>of</strong> (S,S)-2d (free base) matches<br />

that reported in <strong>the</strong> literature. [1]<br />

Major (S,S)-2d: 1 H NMR (400 MHz, CDCl 3 ): δ 7.33-7.20 (m, 5H), 3.88 (q, J = 6.4 Hz, 1H),<br />

2.53-2.46 (m, 1H), 1.34-1.20 (m, 14H), 0.94 (d, J = 6.4 Hz, 3H), 0.88 (t, J = 6.4 Hz, 3H). 13 C<br />

NMR (100 MHz, CDCl 3 ): δ 146.5, 128.3, 126.6, 126.5, 55.1, 50.1, 36.4, 31.8, 29.5, 25.7,<br />

24.6, 22.6, 21.3, 14.1.<br />

(2S)-N-((S)-1-Phenylethyl)hexan-2-amine (2e)<br />

Reaction details: Yb(OAc) 3 (1.1 equiv), 2-hexanone (0.31 mL, 2.5 mmol, 1.0 equiv), (S)-αmethylbenzylamine<br />

(0.35 mL, 2.75 mmol, 1.1 equiv). Reaction time: 12 h; 85% de.<br />

Purification by silica gel flash chromatography (hexanes/EtOAc/NH 4 OH, 88:8:2) gave <strong>the</strong><br />

mixture <strong>of</strong> diastereomers as a colorless viscous liquid, which was <strong>the</strong>n treated with e<strong>the</strong>ral<br />

HCl to obtain <strong>the</strong> hydrochloride salt (0.480 g, 80% yield) after high vacuum drying. GC<br />

(program A, see Experimental section (general remarks)): retention time [min]: major (S,S)-<br />

2e isomer, 11.3; minor (R,S)-2e isomer, 11.1. The NMR data <strong>of</strong> (S,S)-2e (free base) matches<br />

that reported in <strong>the</strong> literature. [2]<br />

Major (S,S)-2e: 1 H NMR (CDCl3, 400 MHz): δ 7.33-7.19 (m, 5H), 3.88 (q, J = 6.5 Hz, 1H),<br />

2.51-2.45 (m, 1H),1.52-1.46 (m, 1H), 1.32 (d, J = 6.5 Hz, 3H), 1.28-1.15 (m, 6H), 0.94 (d, J =<br />

6.34 Hz, 3H), 0.88 (t, J = 6.95 Hz, 3H). 13 C NMR (CDCl3, 100 MHz): δ 146.4, 128.3, 126.6,<br />

126.5, 55.1, 50.1, 36.0, 27.9, 24.6, 22.9, 21.3, 14.1<br />

(2S)-N-((S)-1-Phenylethyl)butan-2-amine (2f)<br />

Reaction details: Yb(OAc) 3 (1.1 equiv), 2-butanone (0.22 mL, 2.5 mmol, 1.0 equiv), (S)-αmethylbenzylamine<br />

(0.35 mL, 2.75 mmol, 1.1 equiv). Reaction time: 12 h; 79% de. After<br />

145


stopping <strong>the</strong> hydrogenation, fur<strong>the</strong>r MeOH was added, and this heterogeneous solution was<br />

filtered to remove <strong>the</strong> Raney-Ni, and excess e<strong>the</strong>real HCl was added. This was concentrated<br />

to dryness (rotary evaporation) and <strong>the</strong>n aqueous HCl (2.0 M) and e<strong>the</strong>r were added. The<br />

acidic aqueous layer was removed and <strong>the</strong> Et 2 O was extracted with fur<strong>the</strong>r extracted with<br />

aqueous HCl (2.0 M, 2 × 15 mL). The aqueous acidic layer was basified with NaOH (4.0 M)<br />

to a pH= 12-14 and <strong>the</strong> free amine extracted with CH 2 Cl 2 (4 x 20 mL). The combined organic<br />

extracts were dried over Na 2 SO 4 , filtered and to <strong>the</strong> filtrate e<strong>the</strong>real HCl (2.0 M, 4.0 mL) was<br />

added. This solution was concentrated on rotary evaporator and after high vacuum drying<br />

(≥24 h) af<strong>for</strong>ded <strong>the</strong> HCl salt (0.42 g, 79% yield) after high vacuum drying. GC (program B,<br />

see Experimental section (general remarks)): retention time [min] <strong>for</strong> <strong>the</strong> free base: major<br />

(S,S)-2f isomer, 13.0; minor (R,S)-2f isomer, 12.7. The NMR data <strong>of</strong> (S,S)-2f (free base)<br />

matches that reported in <strong>the</strong> literature. [2]<br />

Major (S,S)-2f: 1 H NMR (CDCl3, 400 MHz,): δ 7.38-7.20 (m, 5H), 3.87 (q, J = 6.4 Hz, 1H),<br />

2.49-2.41 (m, 1H), 1.56-1.49 (m, 1H), 1.34-1.24 (m, 5 H), 0.95 (d, J = 6.4 Hz, 3H), 0.84 (t, J<br />

= 7.2 Hz, 3H). 13 C NMR (CDCl3, 100 MHz): δ 146.4, 128.3, 126.7, 126.5, 55.0, 51.3, 28.6,<br />

24.7, 20.7, 9.8.<br />

(2S)-3,3-Dimethyl-N-((S)-1-phenylethyl)butan-2-amine (2b) (Pt substrate)<br />

In a reaction vessel dry Yb(OAc) 3 (0.96 g, 2.75 mmol, 1.1 equiv) was added and<br />

subsequently evacuated under high vacuum <strong>for</strong> 5 min be<strong>for</strong>e flooding with nitrogen,,<br />

anhydrous MeOH (2.5 mL, 1.0 M) was <strong>the</strong>n added. To this solution 3,3-Dimethyl-2-butanone<br />

(0.31 mL, 2.5mmol, 1.0 equiv), (S)-α-methylbenzylamine (0.35 mL, 2.75 mmol, 1.1 equiv)<br />

were added. The reaction was <strong>the</strong>n stirred at 50 ºC <strong>for</strong> 2 h. Pt-C (98.0 mg, 1.0 mol %) [The 98<br />

mg <strong>of</strong> Pt/C was added in four equal portions, thus 24.5 mg at t= 0 h, t= 2 h, t= 4 h, and finally<br />

at t= 6 h] and THF (2.5 mL, final molarity <strong>of</strong> reaction vessel 0.5 M) was <strong>the</strong>n added and <strong>the</strong><br />

reaction vessel pressurized at 120 psi (8.3 bar) <strong>of</strong> hydrogen. The reaction was <strong>the</strong>n stirred at<br />

50 ºC. After 22 h (


etention time [min]: major (S,S)-2b isomer, 10.9; minor (R,S)-2b isomer, 10.6. The NMR<br />

data <strong>of</strong> (S,S)-2b (free base) matches that reported in <strong>the</strong> literature. [2]<br />

Major (S,S)-2k: 1 H NMR (400 MHz, CDCl 3 ): δ 7.35-7.20 (m, 5H), 3.77 (q, J = 6.4 Hz, 1H),<br />

2.29 (q, J = 6.4 Hz, 1H), 1.27 (d, J = 6.4 Hz, 3H), 0.89-0.84 (m, 12H). 13 C NMR (100 MHz,<br />

CDCl 3 ): δ 147.6, 128.2, 126.7, 126.6, 59.5, 57.0, 34.7, 26.5, 23.7, 16.0.<br />

(S)-2-aminooctane (4d)<br />

The diastereomeric amine mixture (2d) (0.466 g, 2.0 mmol, 86% de) was dissolved in EtOH<br />

(5.0 mL, 0.4 M) and hydrogenolysis was carried out in presence <strong>of</strong> Pd(OH) 2 /C (0.196 g, 7.0<br />

mol %) at 8.3 bar (120 psi) <strong>of</strong> hydrogen pressure at room temperature. After 10 h, <strong>the</strong> catalyst<br />

was filtered through filter paper and which was subsequently washed with EtOH (2 × 10 mL).<br />

2.0 M e<strong>the</strong>ral HCl (4.0 mL) was <strong>the</strong>n added to <strong>the</strong> filtrate, and this solution was evaporated to<br />

dryness to obtain an oil. The oil was triturated with hexane (4 × 10 mL) and <strong>the</strong> residual<br />

hexane evaporated, this was repeated 3-4 times to obtain a white solid. Fur<strong>the</strong>r drying <strong>for</strong> 15<br />

h under high vacuum provided a white solid in qualitative purity (0.25 g, 76% yield). The<br />

trifluoroacetyl derivative <strong>of</strong> 4d had an ee <strong>of</strong> 85% (chiral GC program E, see Experimental<br />

section (general remarks) and Supporting In<strong>for</strong>mation chromatograms). GC retention time<br />

[min]: major (S)-4d trifluoroacetamide isomer, 15.3; minor (R)-4d trifluoroacetamide isomer,<br />

16.5.<br />

4d-HCl salt: 1 H NMR (400 MHz, CDCl 3 ): δ 8.32 (br s, 3H), 3.32-3.29 (m, 1H), 1.82-1.56<br />

(m, 2H), 1.41-1.28 (m, 11H), 0.87 (t, J = 6.4 Hz, 3H). 13 C NMR (100 MHz, CDCl 3 ): δ 46.9,<br />

40.2, 31.8, 29.4, 26.3, 23.9, 22.6, 14.0.<br />

4d-oxalate salt: The reported literature data <strong>for</strong> this compound is that <strong>of</strong> <strong>the</strong> oxalate salt <strong>for</strong><br />

which <strong>the</strong> 1 H NMR is reported. I also <strong>for</strong>med this salt and found <strong>the</strong> 1 H NMR data <strong>for</strong> this<br />

oxalate salt <strong>of</strong> (S)-4d matched that reported: 1 H NMR (400 MHz, D 2 O): δ 3.32-3.27 (m, 1H),<br />

1.64-1.47 (m, 2H), 1.28-1.22 (m, 11H), 0.81 (t, J = 6.4 Hz, 3H). [3] I additionally recorded <strong>the</strong><br />

13 C NMR (100 MHz, D 2 O, CH 3 OH was used as <strong>the</strong> internal reference, δ= 49.5 ppm): δ 164.9,<br />

48.6, 34.7, 31.5, 28.8, 25.2, 22.5, 18.3, 14.0. [4]<br />

General procedure: Catalytic Yb(OAc) 3 , Y(OAc) 3 , or Ce(OAc) 3 (normal de)<br />

147


In a reaction vessel <strong>the</strong> Lewis acid [Yb(OAc) 3 10 mol %, Y(OAc) 3 (15 mol %) or Ce(OAc) 3<br />

(15 mol %)] was added and subsequently evacuated under high vacuum <strong>for</strong> 5 min be<strong>for</strong>e<br />

flooding with nitrogen. To <strong>the</strong> vessel, anhydrous methanol (2.5 mL, 1.0 M), ketone (2.5<br />

mmol, 1.0 equiv), and (S)-α-methylbenzylamine (0.35 mL, 2.75 mmol, 1.1 equiv) were added<br />

and stirred <strong>for</strong> 20-30 min at <strong>the</strong> temperature at which <strong>the</strong> hydrogenation was per<strong>for</strong>med at (22<br />

or 50 o C). A THF slurry <strong>of</strong> Raney-Ni (100 wt % based on <strong>the</strong> ketone, pre-triturated with<br />

EtOH (×3) and <strong>the</strong>n with anhydrous THF (×3) be<strong>for</strong>e addition) was transferred to <strong>the</strong> reaction<br />

mixture using 2.5 mL <strong>of</strong> anhydrous THF (final molarity <strong>of</strong> reaction solution is 0.5 M). The<br />

vessel was <strong>the</strong>n pressurized to <strong>the</strong> indicated pressure 120-290 psi (8-20 bar) <strong>of</strong> hydrogen and<br />

stirred at room temperature or at 50 °C as indicated. At 12 h (< 3 area % <strong>of</strong> ketone and<br />

intermediate imine by GC) <strong>the</strong> reaction mixture was worked-up as delineated in <strong>the</strong> section<br />

entitled: “General procedure: Stoichiometric Yb(OAc) 3 (enhanced de).”<br />

(2S)-4-methyl-N-((S)-1-phenylethyl)pentan-2-amine (2a)<br />

Reaction details: Yb(OAc) 3 (10 mol %), 4-methyl-2-pentanone (0.31 mL, 2.5 mmol, 1.0<br />

equiv), (S)-α-methylbenzylamine (0.35 mL, 2.75 mmol, 1.1 equiv), pre-stirred 30 min at 50<br />

°C, and <strong>the</strong>n hydrogenated at 50 o C and 120 psi (8.3 bar). Reaction time: 12 h; 92% de.<br />

Purification by silica gel flash chromatography (Hexane/EtOAc/NH 4 OH = 87:9:4) gave <strong>the</strong><br />

mixture <strong>of</strong> diastereomers as viscous colorless oil, which <strong>the</strong>n treated with e<strong>the</strong>ral HCl to<br />

obtain <strong>the</strong> hydrochloride salt (0.60 g, 79 yield %). GC (program D, see Experimental section<br />

(general remarks)) retention time [min]: major (S,S)-2a isomer, 11.8; minor (R,S)-2a isomer,<br />

11.6, matched those reported in <strong>the</strong> literature. [1]<br />

(2S)-4-phenyl-N-((S)-1-phenylethyl)butan-2-amine (2c)<br />

Reaction details: Yb(OAc) 3 (10 mol %), 4-phenyl-2-butanone (0.37 mL, 2.5 mmol, 1.0<br />

equiv), (S)-α-methylbenzylamine (0.35 mL, 2.75 mmol, 1.1 equiv), pre-stirred 30 min at<br />

room temperature; hydrogen pressure 8.3 bar (120 psi); hydrogenation was per<strong>for</strong>med at<br />

room temperature; reaction time: 12 h; 80% de. Purification by silica gel flash<br />

chromatography (Hexane/EtOAc/NH 4 OH = 83:15:2) gave <strong>the</strong> mixture <strong>of</strong> diastereomers as<br />

viscous colorless oil, which <strong>the</strong>n treated with e<strong>the</strong>ral HCl to obtain <strong>the</strong> hydrochloride salt<br />

(0.63 g, 87 yield %). GC (program D, see Experimental section (general remarks)) retention<br />

148


time [min]: major (S,S)-2c isomer, 16.5 minor (R,S)-2c isomer, 16.4, matches that reported in<br />

<strong>the</strong> literature. [1]<br />

(2S)-N-((S)-1-phenylethyl)octan-2-amine (2d)<br />

Reaction details: Yb(OAc) 3 (10 mol %), 2-octanone (0.39 mL, 2.5 mmol, 1.0 equiv), (S)-αmethylbenzylamine<br />

(0.35 mL, 2.75 mmol, 1.1 equiv), pre-stirred 30 min at room temperature;<br />

hydrogen pressure 120 psi (8.3 bar); hydrogenation per<strong>for</strong>med at room temperature. Reaction<br />

time: 12 h; 72% de. Purification by silica gel chromatography (Hexane/EtOAc/NH 4 OH =<br />

58:40:2) gave <strong>the</strong> mixture <strong>of</strong> diastereomers as viscous colorless oil, which <strong>the</strong>n treated with<br />

e<strong>the</strong>ral HCl to obtain <strong>the</strong> hydrochloride salt (0.67 g, 83 yield %). GC (program A, see<br />

Experimental section (general remarks) retention time [min]: major (S,S)-2d isomer, 10.9;<br />

minor (R,S)-2d isomer, 10.8 match those reported in <strong>the</strong> literature. [1]<br />

(2S)-N-((S)1-phenylethyl)hexan-2-amine (2e)<br />

Reaction details: Yb(OAc) 3 (10 mol %), 2-hexanone (0.31 mL, 2.5 mmol, 1.0 equiv), (S)-αmethylbenzylamine<br />

(0.35 mL, 2.75 mmol, 1.1 equiv), pre-stirred 30 min at room temperature;<br />

hydrogen pressure 8.3 bar (120 psi); hydrogenation per<strong>for</strong>med at room temperature. Reaction<br />

time: 12 h; 71% de. Purification by silica gel flash chromatography (Hexane/EtOAc/NH 4 OH<br />

= 89.5:5.5:5) gave <strong>the</strong> mixture <strong>of</strong> diastereomers as viscous colorless oil, which <strong>the</strong>n treated<br />

with e<strong>the</strong>ral HCl to obtain <strong>the</strong> hydrochloride salt (0.61 g, 82 yield %). GC (program A, see<br />

Experimental section (general remarks)) retention time [min]: major (S,S)-2e isomer, 9.7;<br />

minor (R,S)-2e isomer, 9.6, match those reported in <strong>the</strong> literature. [2]<br />

(2S)-N-((S)-1-phenylethyl)butan-2-amine (2f)<br />

Reaction details: Yb(OAc) 3 (10 mol %), 2-butanone (0.22 mL, 2.5 mmol, 1.0 equiv), (S)-αmethylbenzylamine<br />

(0.35 mL, 2.75 mmol, 1.1 equiv), pre-stirred 30 min at room temperature;<br />

hydrogen pressure 8.3 bar (120 psi); hydrogenation done at room temperature. Reaction time:<br />

12 h; 79% de. Purification by silica gel chromatography (Hexane/EtOAc/NH 4 OH = 91:4:5)<br />

gave <strong>the</strong> mixture <strong>of</strong> diastereomers as viscous colorless oil, which <strong>the</strong>n treated with e<strong>the</strong>ral<br />

149


HCl to obtain <strong>the</strong> hydrochloride salt (0.54 g, 82 yield %). GC (program A, see Experimental<br />

section (general remarks)) retention time [min]: major (S,S)-2f isomer, 8.5; minor (R,S)-2f<br />

isomer, 8.4, matches that reported in <strong>the</strong> literature. [2]<br />

(1S)-N((S)-1-cyclohexylethyl)-1-phenylethanamine (2g)<br />

Reaction details: Yb(OAc) 3 (10 mol %), cyclohexyl methyl ketone (0.34 mL, 2.5 mmol, 1.0<br />

equiv), (S)-α-MBA (0.35 mL, 2.75 mmol, 1.1 equiv), pre-stirred 30 min at 50 °C; hydrogen<br />

pressure 20 bar (290 psi); hydrogenation per<strong>for</strong>med at 50 °C. Reaction time: 12 h; 98% de.<br />

Purification by silica gel flash chromatography (Hexane/EtOAc/NH 4 OH = 83:15:2) gave <strong>the</strong><br />

mixture <strong>of</strong> diastereomers as a viscous colorless oil, which when treated with e<strong>the</strong>ral HCl<br />

provided <strong>the</strong> hydrochloride salt (0.60 g, 81 yield %). GC (program D, see Experimental<br />

section (general remarks)) retention time [min]: major (S,S)-2g isomer, 14.9; minor (R,S)-2g<br />

isomer, 14.7, matches that reported in <strong>the</strong> literature. [1]<br />

(2S)-3-Methyl-N- ((S)-1-phenylethyl) butan-2-amine (2h)<br />

Reaction details: Yb(OAc) 3 (10 mol %), 3-methyl-2-butanone (0.27 mL, 2.5 mmol, 1.0<br />

equiv), (S)-α-methylbenzylamine (0.35 mL, 2.75 mmol, 1.1 equiv), pre-stirred 30 min at 50<br />

°C; hydrogen pressure 20 bar (290 psi); hydrogenation per<strong>for</strong>med at 50 °C. Reaction time: 12<br />

h; 98% de. Purification by silica gel flash chromatography (Hexane/EtOAc/NH 4 OH =<br />

92.5:3.5:4) gave <strong>the</strong> mixture <strong>of</strong> diastereomers as a viscous colorless oil, which <strong>the</strong>n treated<br />

with e<strong>the</strong>ral HCl provided <strong>the</strong> hydrochloride salt (0.54 g, 78 yield %). GC (program D, see<br />

Experimental section (general remarks)) retention time [min]: major (S,S)-2h isomer, 11.2;<br />

minor (R,S)-2h isomer, 11.1, matches that reported in <strong>the</strong> literature. [1]<br />

Bis((S)-1-phenylethyl)amine (2i)<br />

Reaction details: Yb(OAc) 3 (10 mol %), acetophenone (0.29 mL, 2.5 mmol, 1.0 equiv), (S)-αmethylbenzylamine<br />

(0.35 mL, 2.75 mmol, 1.1 equiv), pre-stirred 30 min at 50 °C; hydrogen<br />

pressure 20 bar (290 psi); hydrogenation per<strong>for</strong>med at 50 °C. Reaction time: 12 h; 94% de.<br />

Purification by silica gel flash chromatography (Hexane/EtOAc/NH 4 OH = 74:25:1) gave <strong>the</strong><br />

mixture <strong>of</strong> diastereomers as a viscous colorless oil, which <strong>the</strong>n treated with e<strong>the</strong>ral HCl to<br />

150


obtain <strong>the</strong> hydrochloride salt (0.41 g, 63 yield %). GC (program D, see Experimental section<br />

(general remarks)) retention time [min]: major (S,S)-2i isomer, 14.4; minor (R,S)-2i isomer,<br />

14.7, matches that reported in <strong>the</strong> literature. [2]<br />

(2S)-3,3-dimethyl-N-((S)-1-phenylethyl)butan-2-amine (2b)<br />

Reaction details: Yb(OAc) 3 (10 mol %), 3,3-dimethyl-2-butanone (0.31 mL, 2.5mmol, 1.0<br />

equiv), (S)-α-methylbenzylamine (0.35 mL, 2.75 mmol, 1.1 equiv), pre-stirred 4h at 50 °C;<br />

<strong>the</strong>n adding Pt/C (instead <strong>of</strong> Raney Ni) in four equal portions at t= 0, 6, 12, 20 h (total added<br />

Pt equals 1.0 mol %), with a total hydrogenation time <strong>of</strong> 30 h at 8.3 bar (120 psi) and at 50<br />

o C. 92% de. Purification by silica gel flash chromatography (Hexane/EtOAc/NH 4 OH =<br />

94.5:1.5:4) gave <strong>the</strong> mixture <strong>of</strong> diastereomers as a viscous colorless oil, which when treated<br />

with e<strong>the</strong>ral HCl to obtain <strong>the</strong> hydrochloride salt (0.58 g, 78 yield %). GC (program C, see<br />

Experimental section (general remark)) retention time [min]: major (S,S)-2b isomer, 11.8;<br />

minor (R,S)-2b isomer, 11.6, matches that reported in <strong>the</strong> literature. [2]<br />

General Procedure: Brønsted acids (normal de)<br />

The reaction vessel was evacuated under high vacuum <strong>for</strong> 5 min be<strong>for</strong>e flooding with<br />

nitrogen. To <strong>the</strong> vessel, anhydrous methanol (2.5 mL, 1.0 M), acetic acid (20 mol %), ketone<br />

(2.5 mmol, 1.0 equiv) (1), and (S)-α-methylbenzylamine (0.35 mL, 2.75 mmol, 1.1 equiv)<br />

were added and stirred <strong>for</strong> 20-30 min at <strong>the</strong> temperature at which <strong>the</strong> hydrogenation was<br />

per<strong>for</strong>med at (22 or 50 o C). The remaining procedural details should be followed as in <strong>the</strong><br />

section entitled: “General procedure: Catalytic Yb(OAc) 3 , Y(OAc) 3 , or Ce(OAc) 3 (normal<br />

de).”<br />

(2S)-4-methyl-N-((S)-1-phenylethyl)pentan-2-amine (2a)<br />

Reaction details: 4-Methyl-2-pentanone (0.31 mL, 2.5 mmol, 1.0 equiv), (S)-αmethylbenzylamine<br />

(0.35 mL, 2.75 mmol, 1.1 equiv), acetic acid (0.028 mL, 20 mol %), prestirred<br />

30 min at 50 °C; hydrogen pressure 8.3 bar (120 psi); hydrogenation per<strong>for</strong>med at 50<br />

°C. Reaction time: 12 h; 92% de. Purification by silica gel flash chromatography<br />

(Hexane/EtOAc/NH 4 OH = 87:9:4) gave <strong>the</strong> mixture <strong>of</strong> diastereomers as a viscous colorless<br />

151


oil, which when treated with e<strong>the</strong>ral HCl provided <strong>the</strong> hydrochloride salt (0.60 g, 80 yield %).<br />

GC (program D, see Experimental section (general remarks)) retention time [min]: major<br />

(S,S)-2a isomer, 11.8; minor (R,S)-2a isomer, 11.6, matches that reported in <strong>the</strong> literature. [1]<br />

(2S)-N-((S)-1-phenylethyl)octan-2-amine (2d)<br />

Reaction details: 2-Octanone (0.39 mL, 2.5 mmol, 1.0 equiv), (S)-α-methylbenzylamine (0.35<br />

mL, 2.75 mmol, 1.1 equiv), acetic acid (0.028 mL, 20 mol %), pre-stirred 30 min at room<br />

temperature; hydrogen pressure 8.3 bar (120 psi); hydrogenation per<strong>for</strong>med at room<br />

temperature. Reaction time: 12 h; 72% de. Purification by silica gel chromatography<br />

(Hexane/EtOAc/NH 4 OH = 58:40:2) gave <strong>the</strong> mixture <strong>of</strong> diastereomers as viscous colorless<br />

oil, which when treated with e<strong>the</strong>ral HCl provided <strong>the</strong> hydrochloride salt (0.68 g, 83 yield %).<br />

GC (program A, see Experimental sectuib (general remarks)) retention time [min]: major<br />

(S,S)-2d isomer, 10.9; minor (R,S)-2d isomer, 10.8, matches that reported in <strong>the</strong> literature. [1]<br />

(1S)-N((S)-1-cyclohexylethyl)-1-phenylethanamine (2g):<br />

Reaction details: Cyclohexyl methyl ketone (0.34 mL, 2.5 mmol, 1.0 equiv), (S)-α-MBA<br />

(0.35 mL, 2.75 mmol, 1.1 equiv), acetic acid (0.028 mL, 20 mol%), pre-stirred 30 min at 50<br />

°C ; hydrogen pressure 20 bar (290 psi); hydrogenation done at 50 °C. Reaction time: 12 h;<br />

98% de. Purification by silica gel flash chromatography (Hexane/EtOAc/NH 4 OH = 83:15:2)<br />

gave <strong>the</strong> mixture <strong>of</strong> diastereomers as a viscous colorless oil, which when treated with e<strong>the</strong>ral<br />

HCl provided <strong>the</strong> hydrochloride salt (0.62 g, 82 yield %). GC (program D, see Experimental<br />

section (general remarks)) retention time [min]: major (S,S)-2g isomer, 14.9; minor (R,S)-2g<br />

isomer, 14.7, matches that reported in <strong>the</strong> literature. [1]<br />

Bis((S)-1-phenylethyl)amine (2i)<br />

Reaction details: Acetophenone (0.29 mL, 2.5 mmol, 1.0 equiv), (S)-α-methylbenzylamine<br />

(0.35 mL, 2.75 mmol, 1.1 equiv), acetic acid (0.028 mL, 20 mol %), pre-stirred 30 min at 50<br />

°C; hydrogen pressure 20 bar (290 psi); hydrogenation per<strong>for</strong>med at 50 °C. Reaction time: 12<br />

h; 93% de. Purification by silica gel flash chromatography (Hexane/EtOAc/NH 4 OH =<br />

74:25:1) gave <strong>the</strong> mixture <strong>of</strong> diastereomers as a viscous colorless oil, which when treated<br />

152


with e<strong>the</strong>ral HCl provided <strong>the</strong> hydrochloride salt (0.37 g, 55 yield %). GC (program D, see<br />

Experimental section (general remarks)) retention time [min]: major (S,S)-2i isomer, 14.4;<br />

minor (R,S)-2i isomer, 14.7, matches that reported in <strong>the</strong> literature. [2]<br />

References and Notes<br />

[1]T. C. Nugent,V. N. Wakchaure, A. K. Ghosh, R. R. Mohanty, Org. Lett. 2005, 7, 4967.<br />

[2] T. C. Nugent, A. K. Ghosh, V. N. Wakchaure, R. R. Mohanty, R. R. Adv. Synth. Catal.<br />

2006, 348, 1289.<br />

[3] B. A. Davis, D. A. Durden, Synth. Commun. 2001, 31, 569.<br />

[4] H. E. Gottlieb,V. Kotlyar, A. Nudelman, A. J. Org. Chem. 1997, 62, 7512.<br />

153


Mohamed Mahmoud El-Shazly<br />

Curriculum Vitae<br />

Date <strong>of</strong> Birth: 09.April.1977<br />

Mailing Address: Research III, Room 132<br />

School <strong>of</strong> Engineering and<br />

Science<br />

Jacobs University,<br />

28759, Bremen, Germany.<br />

Nationality:<br />

Egyptian<br />

E-mail:<br />

m.elshazly@jacobs-university.de, elshazly444@yahoo.com<br />

Phone: +49 160 557 9099<br />

Academic Qualifications:<br />

Degree Month/Year University<br />

B.Sc. Pharmaceutical Science<br />

(Excellent with Honor, GPA 1.0)<br />

Post Graduate Diploma<br />

(Excellent, GPA 1.33)<br />

M.Sc. Nanomolecular science<br />

(Very Good, GPA 1.67)<br />

PhD Organic Chemistry<br />

Sep. 1995-Jun. 2000 Ain-Shams University, Cairo, Egypt.<br />

Sep. 2000-Jul. 2004 Ain-Shams University, Cairo, Egypt.<br />

Aug. 2004-Aug.<br />

2006<br />

Jacobs University, Bremen, Germany.<br />

Sep. 2006-Aug. 2009Jacobs University, Bremen, Germany.<br />

Academic awards/Honors:<br />

• Teaching and Research Assistantship, Department <strong>of</strong> Natural Product Chemistry,<br />

Faculty <strong>of</strong> Pharmacy, Ain Shams University, Cairo, Egypt.<br />

• Best Teaching Award, Ain Shams University, Cairo, Egypt.<br />

• Best Presentation Skills, Ain Shams workshop <strong>for</strong> presentation skills.<br />

• Graduate Student Fellowship, Jacobs University <strong>for</strong> both MSc and PhD, Bremen,<br />

Germany.<br />

Membership:<br />

• American Chemical Association (ACS).<br />

• The Egyptian Federation <strong>of</strong> Red Cross and Red Crescent Societies.<br />

• International Pharmaceutical Federation (FIP).<br />

• Egyptian Pharmacists Syndicate (EPS).<br />

154


Research experience:<br />

Date<br />

Project<br />

Sep. 2000-<br />

Jul. 2004<br />

Phytochemical and Pharmacological Investigation <strong>of</strong> Natural Products Isolated<br />

from Stipagrostis scoparia Family Graminea.*<br />

Project description: Isolation <strong>of</strong> biologically active fractions with antihypertensive<br />

effect (dieresis and vasodilating effect) from Stipagrostis scoparia <strong>for</strong> <strong>the</strong> first time.<br />

Study <strong>the</strong> Application <strong>of</strong> Transfer Hydrogenation in Reductive Amination.<br />

Dec. 2004-<br />

Jun. 2005<br />

Jun. 2005-<br />

Sep. 2005<br />

Sep. 2005-<br />

Jan. 2006<br />

Jan. 2006-<br />

Sep. 2006<br />

Oct. 2006-<br />

Feb. 2007<br />

Feb. 2007-<br />

Aug. 2007<br />

Aug. 2007-<br />

Jun. 2008<br />

Jun. 2008-<br />

Nov. 2008<br />

Project description: Using chiral auxiliary (α-Methyl Benzyl Amine) as chiral<br />

nitrogen source with different hydrogen donors ( isopropanol, <strong>for</strong>mic acid,..etc) and<br />

ru<strong>the</strong>nium or rhodium catalysts <strong>for</strong> reductive amination <strong>of</strong> different ketones.<br />

Conversion < 50% with 40% diastereoselectivity.<br />

Study <strong>the</strong> Effect <strong>of</strong> Additives on Asymmetric Reductive Amination.<br />

Project description: Testing <strong>the</strong> effect <strong>of</strong> different acidic and basic additives on <strong>the</strong><br />

reductive amination <strong>of</strong> ketones. In general acidic compounds improved<br />

diastereoselectivity and vice versa <strong>for</strong> basic compounds.<br />

Syn<strong>the</strong>sis <strong>of</strong> Different Carbenes and Testing <strong>the</strong>ir Application.<br />

Project description: Syn<strong>the</strong>sizing different chiral amines and utilizing <strong>the</strong>m as<br />

building blocks <strong>for</strong> chiral carbenes. Testing carbenes in enantioselective epoxide ring<br />

opening reactions.<br />

Study <strong>the</strong> Effect <strong>of</strong> Different Lewis Acids on Asymmetric Reductive Amination.<br />

Project description: Testing <strong>the</strong> effect <strong>of</strong> different available Lewis acids on<br />

diastereoselectivity <strong>of</strong> reductive amination. The use <strong>of</strong> Yb(OAc) 3 resulted in great<br />

enhancement <strong>of</strong> diastereoselectivity <strong>of</strong> 2-alaknones.<br />

Developing New <strong>Chiral</strong> Modifiers <strong>for</strong> Pt or Pd Metal Surface.<br />

Project description: Developing chiral cinchonidine analogues (chiral urea and<br />

thiourea derivatives) and testing <strong>the</strong>ir applications as modifiers <strong>for</strong> heterogeneous<br />

catalysts in asymmetric ketone reduction reactions.<br />

Syn<strong>the</strong>sis <strong>of</strong> New Thiourea Organocatalysts and Testing <strong>the</strong>ir Application.<br />

Project description: Preparing different chiral thiourea derivatives and testing <strong>the</strong>ir<br />

applications in meso diol desymmetrization through using different acylating agents.<br />

Syn<strong>the</strong>sis <strong>of</strong> New Formamide Organocatalysts and Testing <strong>the</strong>ir Application.<br />

Project description: Syn<strong>the</strong>sizing different chiral <strong>for</strong>mamide and testing <strong>the</strong>ir<br />

applications in <strong>the</strong> asymmetric allylation <strong>of</strong> aldehydes with allyltrichlorosilane.<br />

Syn<strong>the</strong>sis <strong>of</strong> Novel Catalysts <strong>for</strong> Enantioselective Reductive Amination.<br />

Project description: Developing new air stable iridium chiral ligands and testing<br />

<strong>the</strong>ir applications in one pot enantioselective reductive amination without glovebox.<br />

Syn<strong>the</strong>sis <strong>of</strong> Novel <strong>Chiral</strong> <strong>Amines</strong> on Multi-gram Scale.<br />

Nov. 2008-<br />

Feb. 2009 Project description: Preparing different α-chiral primary amines with >98%<br />

chemical purity and >98% enantioselectivity on multi gram scale (50 gram) utilizing<br />

large scale crystallization eliminating <strong>the</strong> need <strong>of</strong> chromatographic purification.<br />

* This project was carried out at Ain-Shams University Cairo Egypt. The rest <strong>of</strong> projects were carried out at<br />

155


Jacobs University, Bremen, Germany.<br />

Conferences and Oral Presentation:<br />

• Nugent, T. C.; El-Shazly, M.; Wakchaure, V. N. ‘Ytterbium acetate promoted<br />

asymmetric reductive amination: Significantly enhanced stereoselectivity’, Abstracts<br />

<strong>of</strong> Papers, 235th ACS National Meeting, New Orleans, LA, United States, April 6-10,<br />

2008.<br />

Internship:<br />

• El Fatooh Pharmaceutical Corporation, Cairo, Egypt (Marketing and Product<br />

Management section, Jun-Sep. 1996).<br />

• Memphis Pharmaceutical Company, Cairo, Egypt (Sterile Products Section Jun.-Sep.<br />

1997).<br />

• Faculty <strong>of</strong> Pharmacy, S<strong>of</strong>ia, Bulgaria (Natural Product Department Jul.-Aug. 1998).<br />

• Drug Analysis and Development Unit (DAAU), Ain-Shams University Cairo, Egypt.<br />

(Herbal Drugs Quality Control Section Jun.-Sep. 1999).<br />

Computer Skills:<br />

• Literature search s<strong>of</strong>tware MDL Cross Fire, SciFinder and ISI Web <strong>of</strong> Knowledge.<br />

• ISIS draw, Chem-sketch, and ChemDraw.<br />

• Windows, Linux, Mac OS operative systems, Micros<strong>of</strong>t Office, Adobe Photoshop,<br />

Adobe Illustrator, LATEx.<br />

Languages:<br />

• English: Excellently written and spoken<br />

• German: Working knowledge written and spoken<br />

• Arabic: Mo<strong>the</strong>r tongue, excellently written and spoken<br />

156

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!