preparation and activity of multidentate beta-amino

152
Preparation and Activity of Multidentate beta-Amino Sulfoxide Ligands and Expedited Catalytic Preparation of Bupropion by Erwin J. Remigio A Thesis presented to The University of Guelph In partial fulfilment of requirements for the degree of Master of Science in Chemistry Guelph, Ontario, Canada © Erwin J. Remigio, January, 2017

Upload: others

Post on 22-Feb-2022

7 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: Preparation and Activity of Multidentate beta-Amino

Preparation and Activity of Multidentate beta-Amino Sulfoxide Ligands and Expedited Catalytic Preparation of Bupropion

by

Erwin J. Remigio

A Thesis presented to

The University of Guelph

In partial fulfilment of requirements for the degree of

Master of Science in

Chemistry

Guelph, Ontario, Canada

© Erwin J. Remigio, January, 2017

Page 2: Preparation and Activity of Multidentate beta-Amino

ii

ABSTRACT

PREPARATION AND ACTIVITY OF MULTIDENTATE BETA-AMINO SULFOXIDE

LIGANDS

Erwin Javier Remigio Advisor: University of Guelph, 2017 Professor A. L. Schwan

Amino sulfoxide (and related derivative) functionalities have seen an increased

utility in the field of asymmetric synthesis. Many chiral ligands have incorporated these

moieties to successfully catalyze carbon-carbon bond forming reaction. In this work, an

N-Boc heteroaryl (2-pyridyl and 8-quinolyl) multidentate amino sulfoxide motif has been

constructed through diastereoselective sulfenate anion-based chemistry and traditional

sulfide oxidation to access the complementary diastereomer. The 2-pyridyl sulfenate

anion provided low diastereoselectivity (d.r. = 1.3:1), while the 8-quinolyl sulfenate anion

provided one lone diastereomer. Unfortunately, the diastereopure sulfoxides generated

via sulfide oxidation were not accessed due to difficulties in recrystallization. These

ligands were further diversified by condensation with an additional heteroaryl aldehyde

(2-picolinaldehyde and salicylaldehyde) to form their imine derivatives, which provided

an additional coordination site. In total, five ligands were prepared and probed for

reactivity in the asymmetric Henry (nitroaldol) reaction. Two ligands were probed as a

diastereomeric mixture, while the remaining three were probed as diastereopure

Page 3: Preparation and Activity of Multidentate beta-Amino

iii

ligands. In some cases high yields were obtained (as high as 99%), however as a

whole, these ligands failed to impart any meaningful enantioselectivity (as high as ee =

32%). The reactivity of the sulfenate anion was also further investigated. Here, lithiated

sulfenate anions were complexed with chiral PyBox ligands in an attempt to generate

enantiopure sulfoxides. Unfortunately, the resultant sulfoxides were accessed in low

yields with poor accompanying enenatioselectivities (49 – 56%; ee = 0 – 1.4%).

Moreover, attempts to generate a [SO]-2 species were unsuccessful and led only to the

expected products arising form reaction with the heteroaryl sulfenate anions.

Page 4: Preparation and Activity of Multidentate beta-Amino

iv

Acknowledgements

First and foremost, I would like to acknowledge Professor Adrian Schwan for

being an excellent supervisor. Thank you for giving me the opportunity to work in your

research group. I have learned a great deal about professionalism, life, and research

under your guidance. The experience I have gained form working under your

supervision is truly invaluable. I would also like to thank the committee members,

Professor William Tam and Professor Marcel Schlaf, for their guidance and feedback.

I owe many thank to Professor France-Isabelle Auzanneau and Jeffrey Davidson

for their guidance and expertise in dealing with HPLC systems and also for allowing me

to use your lab group’s analytical HPLC system.

In addition, I would also like to thank Dr. Kate Stuttaford and Steve Seifried for

their help and willingness to troubleshoot our HPLC system.

I would like to acknowledge Vibrant Pharma Inc. and Dr. Jaipal Nagireddy for

giving us the opportunity to work in collaboration in the Bupropion project. It has been

an invaluable experience to work alongside an industrial client.

In regards to this project, I would like to thank Rebecca Sydor for the work she

put forth in the Bupropion project. Her hard work and diligence helped guide us towards

optimal reaction conditions for Bupropion synthesis.

I would also like to acknowledge my fellow colleagues, past and present, in the

Schwan group. Thank you to Monika Kulak, Dr. Mohanad Shkoor, Marshall Lindner,

Alex Dean, Ashley Chrismas, Marina Lazarakos, Daniel Mok, Matt Sing, Joe Findlay,

Page 5: Preparation and Activity of Multidentate beta-Amino

v

and Michelle Michalski. Thank you for all of the support and feedback you have

provided me.

I would like thank the neighbouring Tam group and 2nd-Floor Island for putting up

with my non-sense, for lending a helping hand, and for providing useful feedback.

I am very grateful for the funding that NSERC has provided, which has made

these projects possible.

I would like to thank all of my friends and family who have provided a wonderful

and supporting environment for me. Thank you for shaping me into the person I am. I

hope to make you all proud.

Page 6: Preparation and Activity of Multidentate beta-Amino

vi

Table of Contents

ABSTRACT ii

Acknowledgements iv

Table of Contents vi

List of Abbreviations ix

List of Figures x

List of Schemes xi

List of Tables xvii

Chapter 1: Introduction 1

1.0 Introduction 2

1.1 Asymmetric catalysis 2

1.2 β-Amino thiol and disulfide ligands 2

1.3 Sulfonamide moiety in chiral ligands 5

1.4 Incorporation of the sulfoxide moiety in chiral ligands 8

1.5 Optically pure sulfoxide ligands 11

1.6 Accessing the sulfoxide moiety through sulfenate anion chemistry 29

Chapter 2: Results and Discussion 32

2.0 Results and Discussion 33

2.1 Chiral ligand design and synthesis 33

2.2 Synthesis of β-amino electrophile 34

Page 7: Preparation and Activity of Multidentate beta-Amino

vii

2.3 Heteroaryl thiols 37

2.4 Traditional oxidative synthesis of β-amino sulfoxide ligands 39

2.5 Sulfenate anion chemistry 45

2.6 Attempts towards enantioselective sulfoxide generation 48

2.7 Attempts towards [SO]-2 extrusion 51

2.8 β-Amino sulfoxide ligands and related derivatives 53

2.9 Probing chiral ligands for catalytic activity in the Henry (nitroaldol) reaction 55

2.10 Proposed model for observed stereoinduction 59

2.11 Future work 61

2.12 Conclusion 64

ABSTRACT lxvi

Chapter 3: Introduction 68

3.0 Introduction 69

3.1 Traditional synthetic pathway towards α-amino carbonyl compounds 69

3.2 Direct C-N oxidative coupling reactions 74

Chapter 4: Results and Discussion 80

4.0 Results and Discussion 81

4.1 Conclusion 96

Chapter 5: Experimental 97

5.0 Experimental 98

5.1 General experimental 98

5.2.0 Amino sulfoxide ligands and related compounds 99

Page 8: Preparation and Activity of Multidentate beta-Amino

viii

5.3.0 Experimental: Synthesis of Bupropion 124

Chapter 5: References 129

Page 9: Preparation and Activity of Multidentate beta-Amino

ix

List of Abbreviations

BINAP (2,2'-bis(diphenylphosphino)-1,1'-binaphthyl)

Bn benzyl

Boc tert-butyloxycarbonyl

DCCA dichloroisocyanuric acid

d.r. diastereomeric ratio

E+ electrophile

ee enantiomeric excess

hd 1,5-hexadiene

HFIP hexafluoroisopropanol

HPLC high pressure liquid chromatography

IPA isopropanol

L* ligand

mCPBA meta-chloroperbenzoic acid

NMR nuclear magnetic resonance

Nu nucleophile

ODNs 2,4-dinitrophenylsulfonate

PG protecting group

PyBox pyridine bis(oxazoline)

SET single electron transfer

Ts para-toluene sulfonate (tosylate)

Page 10: Preparation and Activity of Multidentate beta-Amino

x

List of Figures

Figure 1. Proposed 5-membered chelate transition state to rationalize favoured

production of (S)-product. 15

Figure 2. Probing the effectiveness of the sulfoxide moiety. 19

Figure 3. Enhanced reactivity of the (SS) sulfoxide conformation. 20

Figure 4. Proposed internal complexation transition state to account for observed

sulfenate diastereoselectivity with homochiral amino electrophiles. 31

Figure 5. Chiral amino sulfoxide ligands 54

Figure 6. Copper (II) coordination sites 60

Figure 7. 1H NMR of isolated by-product and corresponding predicted chemical

structure. 82

Figure 8. 1H NMR spectrum of the crude reaction mixture containing Bupropion

and 3’-chloropropiophenone. 84

Page 11: Preparation and Activity of Multidentate beta-Amino

xi

List of Schemes

Scheme 1. Test reactions for the asymmetric reduction of benzaldehyde with

diethyl zinc with amino thiol and amino disulfide ligands. 4

Scheme 2. Ligand screening between amino thiol and corresponding amino

disulfide ligands. 5

Scheme 3. Comparison of amino disulfide and amino thiol ligand. 5

Scheme 4. Selection of lone pair by the activated metal-ligand complex in

aldehydes (top) and ketones (bottom). 6

Scheme 5. Substrate scope tests enantioselective alkylzinc additions to ketones

in presence of Ti(OiPr)4. 7

Scheme 6. Asymmetric alkynation of ketones via isoborneol-10 sulfonamide

lewis acid catalyst. 8

Scheme 7. The first application of the sulfoxide moiety in asymmetric catalysis,

which was employed as a diastereomeric mixture. 9

Scheme 8. Diastereomeric bis-sulfoxide ligand employed in an asymmetric

hydrogenation reaction. 10

Scheme 9. Multidentate sulfoxide ligand used for asymmetric hydrogenation (hd

= 1,5-hexadiene). 11

Page 12: Preparation and Activity of Multidentate beta-Amino

xii

Scheme 10. Optically pure sulfoxide ligands employed in asymmetric aldehyde

alkylation. 12

Scheme 11. Proposed active catalytic species with 1-5 12

Scheme 12. Catalytic cycle for the palladium-catalyzed Tsuji-Trost allylic

substitution. 13

Scheme 13. Test reaction for the palladium-catalyzed Tsuji-Trost allylic

substitution and examination of the reactivity of sulfur moiety in different

oxidation states. 14

Scheme 14. Test reaction for the palladium-catalyzed Tsuji-Trost allylic

substitution. 15

Scheme 15. Test reaction for the palladium-catalyzed Tsuji-Trost allylic

substitution using complementary (RS, S) and (SS, S) sulfoxide diastereomer

ligands. 16

Scheme 16. Proposed 9-membered transition states to rationalize

enenatioselectivity in (RS, S) and (SS, S) sulfoxide diastereomer ligands. 17

Scheme 17. Test reaction for the palladium-catalyzed Tsuji-Trost allylic

substitution using amido phospino sulfoxide ligand. 18

Scheme 18. Enhanced catalytic activity of an amino sulfane over its

diastereomeric amino sulfoxide counterpart. 20

Scheme 19. Endo and exo adducts and respective enantiomers for the

asymmetric Diels-Alder cycloaddition. 21

Page 13: Preparation and Activity of Multidentate beta-Amino

xiii

Scheme 20. Test reaction for asymmetric Diels-Alder cycloaddition and the

proposed active catalytic species. 22

Scheme 21. Proposed 6-membered transition state to rationalize Re-face

approach leading to (R)-endo product. 23

Scheme 22. Proposed 6-membered transition state to rationalize Si-face

approach leading to (S)-endo product. 24

Scheme 23. Addition of methyl group to lock p-tol group in equatorial position. 24

Scheme 24. Proposed transition state to favour diene approach from the Re-face

in either conformation to afford the (S)-product. 25

Scheme 25. Enzyme-mediated de-symmetrisation sulfane with CAL-B lipase and

vinyl acetate to access the de-symmetrized sulfoxide. 26

Scheme 26. Test reaction for the Aza-Henry reaction 27

Scheme 27. Test reaction for Simmons-Ingold cyclopropanation 27

Scheme 28. Imino (Schiff-base) sulfoxide ligands used in the Henry reaction. 28

Scheme 29. Probing the importance of the sulfoxide moiety 29

Scheme 30. O- and S- sulfenate anion alkylation 30

Scheme 31. Retrosynthethic approach towards the construction of heteroaryl

amino sulfoxide ligands. 34

Scheme 32. Reduction of N-Boc protected L-phenylalanine 34

Page 14: Preparation and Activity of Multidentate beta-Amino

xiv

Scheme 33. Synthesis of amino iodide electrophile and formation of cyclic by-

product. 36

a Dissolve iodine in DCM and transfer resultant solution to Ph3P and imid. in DCM.

36

Scheme 34. Attempts towards SNAr substitutions with potassium thioacetate and

potassium thiocyanate. 38

Scheme 35. Reduction of sulfonyl chloride moiety to a thiol vial Ph3P. 39

Scheme 36. Thiolate substitution to access amino sulfane compounds 40

Scheme 37. Removal of N-Boc protecting group via trifluoroacetic acid. 43

Scheme 38. Reported synthesis of N-Ts protected amino sulfoxide ligands 44

Scheme 39. N-Ts protected amino sulfoxides 44

Scheme 40. Conditions for sulfanyl acrylate oxidations 47

Scheme 41. Addition-elimination mechanism for the generation of a lithiated

sulfenate anion. 47

Scheme 42. Heteroaryl sulfenate anion substitution with homochiral amino

electrophile 48

Scheme 43. Lithiated intramolecular sulfenate complexation 49

Scheme 44. Spirocyclic chiral Lewis acid catalyst for enantioselective sulfoxide

formation. 49

Page 15: Preparation and Activity of Multidentate beta-Amino

xv

Scheme 45. Enantioselective sulfoxide formation attempts with chiral PyBox

ligands 51

Scheme 46. Proposed mechanism of SO extrusion. 52

Scheme 47. Attempts towards [SO]-2 extrusion via double nBuLi addition and

release. 53

Scheme 48. Imino sulfoxide ligand synthesis 55

Scheme 49. Probing of other aldehydes as substrates in the Henry reaction under

optimized conditions. 59

Scheme 50. Proposed model for observed stereoinduction and preference for the

(R)-isomer. 61

Scheme 51. Future work towards ligand diversification. 63

Scheme 52. Common synthetic pathway towards α-amino carbonyl compounds

(X = C(O)OR, Y = N or O, Z = lone pair of electrons if Y = O, or Z = C(O)OR if Y = N).

70

Scheme 53. Enantioselective α-amination via tin enolate and silver (I)/BINAP

system. 70

Scheme 54. Asymmetric α-amination via chiral magnesium bis(sulfonamide)

enolate complex. 71

Scheme 55. Photoredox catalytic cycle. 73 Scheme 56. Proline-catalyzed amination via enamine intermediate. 74

Page 16: Preparation and Activity of Multidentate beta-Amino

xvi

Scheme 57. NBS-mediated α-amination of aryl ketones. 75

Scheme 58. Cu(II)-mediated oxidative α-amination of carbonyl compounds. 76

Scheme 59. Proposed I2 radical coupling mechanism 78

Scheme 60. Retrosynthetic approach of Bupropion via direct C-N coupling. 78

Scheme 61. Traditional synthesis of Bupropion. 79

Scheme 62. Failed synthesis of Bupropion via Cu(II)-mediated oxidative coupling.

81

Scheme 63. Test reaction of direct C-N coupling of 3’-chloropropiophenone and

morpholine via Cu(II)-mediated oxidative coupling. 82

Scheme 64. Proposed synthesis of Bupropion. 83

Scheme 65. Proposed ionic mechanism for the synthesis of Bupropion. 91

Scheme 66. Summary of synthetic efforts towards Bupropion. 95

Page 17: Preparation and Activity of Multidentate beta-Amino

xvii

List of Tables

Table 1. Substitution attempts on the amino mesylate 36

Table 2. Oxidation conditions of amino sulfane compounds 41

Table 3. Diastereomeric ratio of recrystallization fractions 43

Table 4. Solvent screening reactions for the Henry Reaction with nitromethane. 57

a Determined by chiral HPLC, Chiralcel OD-H column 0.8 mL/min; 85/15 (v/v)

Hexanes/IPA 57

Table 5. Solvent screening reactions for the Henry Reaction with nitroethane 58

Table 6. Solvent screening reaction trials. (Reaction entries expect #2 and #3

were performed by Rebecca Sydor.) 85

Table 7. Acid and iodine source screening reactions. (Reaction entries in this

table were performed by Rebecca Sydor.) 86

Table 8. Oxidant screening reactions. (Reaction entries in this table were

performed by Rebecca Sydor.) 87

Table 9. Screening reactions with aqueous saturated thiosulfate wash during

aqueous work-up. (Reaction entries in this table were performed by Rebecca

Sydor.) 88

Page 18: Preparation and Activity of Multidentate beta-Amino

xviii

Table 10. Screening reactions performed in screw-top glass pressure vessel and

comparable control reactions. (All reaction entries expect #1 were performed by

Rebecca Sydor.) 89

Table 11. Probing other suitable reagents for α-amination of 3’-

chloropropiophenone. (All reaction entries expect 4 were performed by Rebecca

Sydor.) 90

Table 12. Reaction screening trials with extra equivalents of reagents added.

(Entries marked with asterisk were performed by Rebecca Sydor.) 94

Table 13. Optimized reaction conditions tested with larger scale reactions. 95

Page 19: Preparation and Activity of Multidentate beta-Amino

1

Chapter 1: Introduction

Page 20: Preparation and Activity of Multidentate beta-Amino

2

1.0 Introduction

1.1 Asymmetric catalysis

Asymmetric carbon-carbon bond forming reactions are a central component of

synthetic organic chemistry. They play a large role in modern pharmaceutical chemistry

and natural product synthesis, due to the complex and often multiple stereogenic

centres that are typically sought. Given its importance, the field of asymmetric catalysis

has demanded similar attention regarding the development of highly stereoselective

catalysts. Often, chiral-at-metal complexes are utilized to impart stereoselectivity for

these asymmetric transformations.1-3 Here, the active catalytic species is formed from

the coordination of a chiral ligand and an active metal centre.1 The active species can

then form a chelate involving the substrates involved. The designs of chiral ligands

require the consideration of electronic and steric conformations of the chelate to bring

about stereoselectivity.1

1.2 β-Amino thiol and disulfide ligands

The use sulfur in the design of chiral ligands has garnered an increasing amount

of attention.2-6 Sulfur-carrying starting materials are generally commercially available,

and possess a high stability, allowing for facile handling. Naturally, with the

incorporation of a sulfur moiety in chiral ligands, its various oxidation states for use in

asymmetric catalysis have been probed. Moreover, chiral-at-metal complexes often

Page 21: Preparation and Activity of Multidentate beta-Amino

3

offer an additional accompanying complexing atom to aid in metal coordination to form

an active catalyst. As such, nitrogen atoms are a commonly used, due to the wide array

of natural chirally pure nitrogen containing species, and their commercial availability.5,6

For instance, a series of β-amino thiol ligands, derived from L-valine, were

reported by Anderson et al..7 These ligands were probed for their asymmetric catalytic

activity in addition of organozinc reagents to an aldehyde centre. Previous successful

amino thiol ligands, also used for this reaction, have shown increased catalytic activity

over their related alcohol analogues. This promising feature was thought to arise from

an increased polarizability of the sulfur atom compared to oxygen, and a softer

coordination site, more suited towards transition metal complexation. These ligands

relied on chiral influences by the iso-propyl group and by the differential substitution on

the nitrogen atom, which upon complexation could potentially become a stereogenic

centre, through restricted conformation. Overall, results were modest during their

benchmark reactions using benzaldehyde and diethyl zinc, with enantioselectivities

reaching upwards of 82% for the (R)-stereoisomer. Interestingly, the amino thiol ligands,

which contained mismatched N-substitution, yielded an appreciable improvement in

enantioselectivity over their matched counterparts, thus suggesting that nitrogen itself

could have the potential to be stereogenic (Scheme 1).

In the same report, the authors oxidized the reported amino thiol ligands into their

corresponding disulfides in order to probe them for catalytic activity. Unfortunately, their

activity was generally less efficient with decreased enantioselectivity (ee = 80%).

However the trend of mismatched N-substitution was also observed.7

Page 22: Preparation and Activity of Multidentate beta-Amino

4

Scheme 1. Test reactions for the asymmetric reduction of benzaldehyde with diethyl zinc with amino thiol and amino disulfide ligands.

The literature has provided conflicting results regarding the effectiveness of

amino thiol ligands in comparison to their disulfide counterparts. Kang et al. reported a

set of ligands with impressive results that were slightly more effective in their amino thiol

form as opposed to the disulfide.8 Along with decreased enantioselectivity, reaction

times with the disulfides were often twice as long. However, in most cases the decrease

in effectiveness was only marginal (Scheme 2).

The literature also provides cases where the opposite trend can be found. Hof et

al. have reported a set of amino disulfide ligands more effective than their thiol

counterpart (Scheme 3).9

H

O

ZnEt2+L*

Et

OH

L* =

1R2RN SH

R1 = R2 = Ph: 85%, ee = 74%R1 = R2 = Bn: 91%, ee = 58%R1 = iPr, R2 = Me: 81%, ee = 72%R1 = Ph, R2 = Me: 89%, ee = 82%

L* =

1R2RN S S NR1R2

R1 = R2 = Bn: 67%, ee = 38%R1 = R2 = Ph: 67%, ee = 59%R1 = Ph, R2 = Me: 85%, ee = 80%

Page 23: Preparation and Activity of Multidentate beta-Amino

5

Scheme 2. Ligand screening between amino thiol and corresponding amino disulfide ligands.

Scheme 3. Comparison of amino disulfide and amino thiol ligand.

1.3 Sulfonamide moiety in chiral ligands

A related, but more challenging reaction is the 1,2-addition of organozinc

reagents to ketones. This reaction has proved to be more difficult in imparting

stereoselectivity than its aldehyde counterpart, due to the reduced ability of ketones to

coordinate to a Lewis acid catalyst through steric interactions.5,6 Moreover, aldehydes

have a greater propensity to differentiate which oxygen lone pair is coordinated to the

R H

OZnEt2+ L*

R Et

OH

L* =

R = Ph: 90%, ee = 100%R = pClC6H4: 92%, ee = 100%R = oMeOC6H4: 92%, ee = 99%R = 2-Naph: 98%, ee = 99%

L* =HS N

Ph Ph

S N

Ph Ph

SN

PhPh

R = Ph: 92%, ee = 99%R = pClC6H4: 93%, ee = 99%R = oMeOC6H4: 91%, ee = 99%R = 2-Naph: 93%, ee = 98%

HS

Ph Me

S NHMe

Ph Me

S

L* =L* =

MeHN NHMe

PhMe

H

O

ZnEt2+L*

Et

OH

95%, ee = 80% 75%, ee = 86%

Page 24: Preparation and Activity of Multidentate beta-Amino

6

active catalyst due to lessened steric hindrance of the aldehyde hydrogen compared to

a ketone’s carbon based functionality. In the case of ketones, discrimination between

oxygen lone pairs is greatly reduced since both syn- and anti-coordination tend to have

comparable environments (Scheme 4).5,6

Scheme 4. Selection of lone pair by the activated metal-ligand complex in aldehydes (top) and ketones (bottom).

The reactions of ketones with selected nucleophiles can produce tertiary

alcohols, which are important targets in synthetic chemistry. These products can

typically be accessed via ketone reduction with organolithium or Grignard reagents.

However these protocols require typically harsh conditions and functional group

compatibility suffers. Unlike the organozinc reaction with aldehydes, ketones are not

sufficiently reactive, but difficulties can be overcome through the addition of Ti(OiPr)4

along with the chiral ligand.5,6

The Ti(OiPr)4 is proposed to exchange its own alkoxide for an alkyl group from

the zinc species, allowing the titanium to be the active Lewis acid catalyst, while also

being the carrier of the nucleophilic alkyl reagent (1-1).6 For this reaction the

sulfonamide moiety has been extensively incorporated into the chiral ligands used.

Specifically, the isoborneol-10 sulfonamide motif has been a central component to the

design of these ligands, and has been reported numerous times in various ligand

R H

OMLx

R H

O MLx

R' R"

OMLx

R' R"

O MLx

Page 25: Preparation and Activity of Multidentate beta-Amino

7

designs. An early chiral isoborneol-derived sulfonamide ligand reported by Yus et al.

was quite successful in catalyzing enantioselective alkylzinc additions to ketones in the

presence of Ti(OiPr)4.10 The ligand performed rather well for simple ketone substituents

and alkylzinc reagents. However, the process was troubled by long reaction times and

high catalyst loading (20 mol%) (Scheme 5). In their report, the authors also postulated

a dinuclear titanium complex as the active catalytic species; one titanium centre carried

the ligand and ketone, while the other carried the alkyl nucleophile, and both metal

centres were bridged by two alkoxides.

Scheme 5. Substrate scope tests enantioselective alkylzinc additions to ketones in presence of Ti(OiPr)4.

A report form Lu et al. utilized this same isoborneol-10 sulfonamide motif for the

alkynation of ketones to produce valuable tertiary propargylic alcohols.11 Screening

reactions, which varied substituents at various positions on the isoborneol ring and N-

O

SO2NHTi

O

ArR

H

O

O

iPr

iPr

Ti

Et

OiPrOiPrOiPr

Ar R

H

O L*ZnEt2

Ti(iOPr)4

Active catalytic species:

R1 R2

OZnR32+

Ti(OiPr)4

L* R1 R2

R3 OH

OH

SO2NH

L* =

R1 = Ph, R2 = R3 = Et: 63%, ee = 84%R1 = Ph, R2 = Me, R3 = Et: 85%, ee = 86%R1 = Ph, R2 = nBu, R3 = Et: 78%, ee = 86%R1 = Ph, R2 = Et, R3 = Me: 85%, ee = 86%

1-1

Page 26: Preparation and Activity of Multidentate beta-Amino

8

sulfonamide position of the ligand, led to the use of the same ligand reported by Yus et

al. In this report, the reduced reactivities of ketones and organozinc species was aided

by the addition of the stronger Lewis acid, Cu(OTf)2, in catalytic amounts.11 Overall, this

ligand was able to promote excellent enantioselectivities for the formation tertiary

propargylic alcohols (Scheme 6).

Scheme 6. Asymmetric alkynation of ketones via isoborneol-10 sulfonamide lewis acid catalyst.

1.4 Incorporation of the sulfoxide moiety in chiral ligands

A recent review from Trost and Roa delved into the development of chiral

sulfoxide ligands, and concluded that their use in asymmetric catalysis has much more

left to be explored.12 The sulfoxide functionality offers many distinct characteristics: (I) it

possesses a high optical stability, (II) it efficiently carries chiral information, (III) it can be

accessed in both enantiomeric forms, and (IV) it benefits from a large steric and

electronic differentiation between its surrounding carbon and oxygen environment.12

Ph

OZnMe2

+Cu(OTf)2

L* PhOH

OH

SO2NH

L* =R = H: 92%, ee = 88%R = Br: 65%, ee = 96%R = Cl: 94%, ee = 97%R = F: 91%, ee = 96%

R PhH

Ph

R

Page 27: Preparation and Activity of Multidentate beta-Amino

9

Moreover, the sulfoxide presumably places the location of stereogenecity close to the

reaction centre, which can assist the surrounding chiral carbon backbone. Furthermore,

this highly polarized functionality lends the potential to differentiate between S- and O-

coordination with ‘hard’ and ‘soft’ metals.12,13

James et al. were the first to report a chiral ligand, which included the sulfoxide

functionality in 1976.14 This novel ligand was employed for the hydrogenation of itaconic

acid to form methylsuccinic acid. Although the yields and enantioselectivities were less

than promising, it led them to design two more generations of catalyst for asymmetric

hydrogenation, each with varying design for coordinating groups. The first ligand

disclosed was employed as a diastereomeric mixture and featured an additional chiral

carbon stereocentre (1-2).14 Unfortunately, the results were less than ideal and low

yields with poor accompanying stereoselectivities were reported (Scheme 7).

Scheme 7. The first application of the sulfoxide moiety in asymmetric catalysis, which was employed as a diastereomeric mixture.

The second-generation sulfoxide ligand 1-3 reported from this research group

featured a bis-sulfoxide motif still as a diastereomeric mixture, designed in part to

enhance its chelating capabilities.15 Furthermore, during screening reactions, the

HO

OOH

OHO

OOH

ORuCl3•H2O, L*H2

N,N-dimethylacetamide50 ˚C

50%, 12% ee

L* = SO

d.r. = 1:1

1-2

Page 28: Preparation and Activity of Multidentate beta-Amino

10

substrates were varied in order to test any additional substrate chelating effects.

Itaconic acid underwent the most stereoselective transformation (ee = 25%). However,

atropic and 2-acetamidoacrylic acid underwent hydrogenation with significantly lower

selectivities, which suggested that the presence of an additional carboxylate group aids

the asymmetric transformation (Scheme 8).

This prompted the design of a third generation of sulfoxide-based ligand 1-4,

which incorporated a carboxylate group, once again as a diastereomeric mixture. In

contrast to the previous two reactions this catalyst system utilized a rhodium system and

iPrOH as the hydrogen source.16 Overall, the presence of the carboxylate group

seemingly provided increased enantioselectivities for the hydrogenation (Scheme 9).

Scheme 8. Diastereomeric bis-sulfoxide ligand employed in an asymmetric hydrogenation reaction.

HO

OOH

OHO

OOH

O

RuCl2(DMSO)4, L*H2

N,N-dimethylacetamide50 ˚C

L* =

d.r. = 1:1

HO

O

Ph

HO

O

HN

O

HO

O

Ph

HO

O

HN

O

49%, ee = 25%

17%, ee = 4%

52%, ee = 7%

SO

SO

OO

1-3

Page 29: Preparation and Activity of Multidentate beta-Amino

11

Scheme 9. Multidentate sulfoxide ligand used for asymmetric hydrogenation (hd = 1,5-hexadiene).

1.5 Optically pure sulfoxide ligands

It was not until 1993, that Carreño et al. employed enantiopure β-

hydroxysulfoxide ligands in the alkylation of aromatic aldehydes (Scheme 10).17 A

variety of ligands were reported and varied in carbon skeleton flexibility as well as

crowding around the β-hydroxy centre. Results were modest, but the absolute chirality

of the sulfoxide was given appreciation upon the authors proposing a plausible

mechanism. It was suggested that the sulfinyl oxygen coordinated with the organozinc

species in order to create a more stable six-membered transition state 1-5 (as opposed

to a five-membered ring) (Scheme 11). Interestingly, pre-treatment of one of the ligands

with AlMe3 increased stereoselectivity, however its involvement was not probed further.

R

R'

O

R

R'

OH[RhCl(hd)2], L*

iPrOH, reflux

R = H, R' = Me: 45%, ee = 63%R = H, R' = Et: 21%, ee = 71%R = Me, R' = Me: 31%, ee = 75%

SO

HN

O

OH

OL* =

1-4

Page 30: Preparation and Activity of Multidentate beta-Amino

12

Scheme 10. Optically pure sulfoxide ligands employed in asymmetric aldehyde alkylation.

Scheme 11. Proposed active catalytic species with 1-5

As the sulfoxide moiety was continually incorporated into chiral ligands, it was

paired increasingly with adjacent amine and amine-derived functionalities. Moreover,

the stereogenecity of the sulfur centre itself garnered further attention. This was

highlighted in a report from Williams et al., which further displayed the versatility of the

chiral configuration of the sulfoxide.18 The reported ligands were screened against the

extensively studied Pd-catalyzed Tsuji-Trost allylic substitution, presumably due to its

well-defined catalytic cycle (Scheme 12).6 While many successful sulfur-based ligands

H

O

ZnEt2+L*

Et

OH

OHSO OH

SO

SOHO

SOHO

ee = -9% ee = 22% ee = 11% ee = 45%* when pre-treated

with AlMe3 ee = 55%

L* =

SO

ZnO

RR

HEt

Zn EtO

Et

Ph

H

SO

ZnO

RR

HEt

Zn Et

Et

PhCHO

Proposed active catalytic species:

1-5

Page 31: Preparation and Activity of Multidentate beta-Amino

13

have successfully imparted stereoselectivity in this transformation, phosphorous-based

ligands have still shown to be the most effective.6

Scheme 12. Catalytic cycle for the palladium-catalyzed Tsuji-Trost allylic substitution.

During ligand evaluation, the diastereomeric pairs each provided noticeably

different results (Scheme 13).18 Although the (SS) isomer proved to be more effective

than its complementary pair, the sulfoxide’s presence was found to be nonessential, as

evidenced by the superior performance of the sulfane ligand. This suggested that the

palladium interacts with the ligand through either of the sulfane’s enantiotopic lone-pairs

to form a diastereopure catalyst, and possesses the potential to readily switch to a

preferred configuration upon palladium binding. Moreover, the ineffectiveness of the

sulfone-based ligand, which can only coordinate through its sulfonyl oxygen, further

supports S-Pd coordination.18

X

X

PdL L'

PdL L'

Nu

PdL L'

NuPd

S**S

L L'

XNu

decoordination coordination

oxidativeaddition

nucleophilicaddition

L, L' = ligandS* = solvent/vacantX = leaving groupNu = nucleophile

Page 32: Preparation and Activity of Multidentate beta-Amino

14

Scheme 13. Test reaction for the palladium-catalyzed Tsuji-Trost allylic substitution and examination of the reactivity of sulfur moiety in different oxidation states.

In 1997, Hiroi et al. reported a set of β-amino sulfoxide ligands for use in

asymmetric allylation.19 These ligands utilized the sulfoxide as the lone source of

chirality. In their communication, a variety of ligands were prepared, however only the

(SS)-sulfoxide was probed for catalytic activity. Their reported ligands performed with

moderate efficiency, however it was highlighted that rotationally restricted ligands (1-6)

offered improved stereoselectivities during this coupling (Scheme 14).19

Such an observation from these ligands was rationalized by the proposed

formation of a rigid 5-membered palladacyclic chelate ring, favoured by the planar

phenyl connector. Here, the sulfinyl electron lone pair is the coordination point. A 6-

membered chelate ring is certainly feasible, but the planar influence of the phenyl ring

could act to discourage that. Once the 5-membered chelate 1-7 has been established,

the coordination would provide a chiral environment allowing a nucleophilic attack from

a sulfinyl oxygen approach manner to afford the favoured (S)-product (Figure 1).

Ph Ph

OAc

MeO OMe

O O

Ph Ph

OMeMeO

O O

+[{Pd(n3-allyl)Cl}2], L*

KOAc

CH2Cl2

S O N

O

SO

N

O

S N

O

SO2 N

OL* =

96%, ee = 88% 42%, ee = 55% 69%, ee = 93% 0%, ee = n/a

Page 33: Preparation and Activity of Multidentate beta-Amino

15

Scheme 14. Test reaction for the palladium-catalyzed Tsuji-Trost allylic substitution.

Figure 1. Proposed 5-membered chelate transition state to rationalize favoured production of (S)-product.

Another communication from Hiroi et al. reports a chiral ligand which combined

the amido sulfoxide and phosphine functionalities, which was utilized to couple dimethyl

malonate and rac-(E)-1,3,-diphenylallyl (Scheme 15).20 Interestingly, they also provided

a more in-depth screening on complementary sulfoxide diastereomers. While neither

ligand displayed exceptional effectiveness, the marked difference in their catalytic ability

was rather intriguing. In comparing the complementary sulfinyl diastereomers, the (R)-

diastereomer (1-8) resulted in a low 53% yield, but was accompanied by a moderate

S

N

O

OtBu

O OOAc+ OtBu

O O[PdCl(π−allyl)2]

L* =

Yield = 25% (S)ee = 50%

*

1-6

NSO

Pd

Page 34: Preparation and Activity of Multidentate beta-Amino

16

stereoselectivity with ee = 74% for the (S)-product. The (S)-diastereomer (1-9) resulted

in a similar 58% yield, however it was essentially non-stereoselective (ee = 2%)

(Scheme 15).

Scheme 15. Test reaction for the palladium-catalyzed Tsuji-Trost allylic substitution using complementary (RS, S) and (SS, S) sulfoxide diastereomer ligands.

Such significant differences in stereoselectivity were proposed to arise from

different conformational preferences of the π-allyl group to the 9-membered

palladacyclic complex between either diastereomer, largely due to steric interactions

(Scheme 16).20 In the complex formed from the (R)-diastereomer 1-10, which provided

greater asymmetric control, the chelate resulted in a matched M-type π-allyl

conformation. In this favourable conformation, the phenyl π-allyl phenyl groups are

located in favourable locations, with regards to the sulfoxide and phosphine

substituents. Here, the C3 terminus is the favoured point of attack, since it is placed

trans to the better π-accepting phosphine moiety. In the complementary (S)-

N

O

S Bn

O

N

O

S Bn

O PPh2PPh2

L* =

(RS,S)Yield = 53%ee = 74%

(SS,S)Yield = 58%ee = 2%

Ph Ph

OAc

MeO OMe

O O

Ph Ph

OMeMeO

O O

+[PdCl(π−allyl)2]

1-8 1-9

Page 35: Preparation and Activity of Multidentate beta-Amino

17

diastereomer, an equilibrium exits between the M- and W-type allyl conformation

chelate (1-11, 1-12), to manage the steric interactions between the π-allyl phenyl

groups. Attack at the C3 terminus is still certainly favoured, however the M- and W-π-

allyl equilibrium results in a near racemic mix of (R) and (S)-products (Scheme 16).20

Scheme 16. Proposed 9-membered transition states to rationalize enenatioselectivity in (RS, S) and (SS, S) sulfoxide diastereomer ligands.

Recently, Xiao et al. reported an effective ligand (1-13) for use in this reaction,

which incorporated amido sulfoxide functionalities with a phosphine moiety (Scheme

17).21 In the benchmark test, which coupled dimethyl malonate with rac-(E)-1,3,-

diphenylallyl acetate in the presence of Pd catalyst, K2CO3 and Cs2CO3, the authors

reported a 99% GC-yield along with 99% ee for the (S)-product. While impressive, these

screening reactions were performed at a small scale (0.3 mmol of rac-(E)-1,3,-

diphenylallyl acetate and 0.9 mmol of malonate) and investigation of these parameters

at a larger scale would serve to increase the relevance of these observed results.

Interestingly, they were able to obtain the enantiomeric (R)-product with mono-α-

PdPS

N

Ph

Ph

OR Pd

PS

N

Ph

Ph

RO

O O

Ph Ph Ph Ph1 Ph Ph13

PdPS

N

Ph

Ph

RO

O

1 3

(R)

(R)(S)(S)

3

(S)

Nu Nu Nu1-10 1-11 1-12

M-conformation M-conformation W-conformation

Page 36: Preparation and Activity of Multidentate beta-Amino

18

substituted dimethyl malonate substrates, in similarly high yields and ee’s, although the

authors did not investigate this occurrence further.

Scheme 17. Test reaction for the palladium-catalyzed Tsuji-Trost allylic substitution using amido phospino sulfoxide ligand.

During the development of these ligands, Xiao et al. designed a series of control

experiments to evaluate the sulfoxide moiety’s contribution to the ligands

stereoselective ability. The low catalytic activity of analogous sulfane and sulfone

derivatives of the efficient ligand highlighted the importance of its presence as a chiral

auxiliary (Figure 2). However, in these control and subsequent substrate scope

experiments, only the (RS)-diastereomer was examined for catalytic activity.21 Despite

the limited screening of the chiral sulfoxide influence, its effectiveness certainly offers a

remarkable improvement over amino-derived sulfoxide ligands previously used for this

reaction.

PPh2

NH SO

OBr

L* =

Ph Ph

OAc

MeO OMe

O O

Ph Ph

OMeMeO

O O

+[Pd(C3H5)Cl]2 / L*K2CO3, Cs2CO3

CH2Cl2

Yield = 99% (S) ee = 99%

1-13

Page 37: Preparation and Activity of Multidentate beta-Amino

19

Figure 2. Probing the effectiveness of the sulfoxide moiety.

The inclusion of the sulfoxide moiety in chiral ligands continued to provide varied

results in regards to their efficacy and such ligands were still probed mostly as a curious

afterthought once amino sulfane ligands had been prepared. For instance, a set of

amino sulfoxide ligands 1-14 for the asymmetric hydrogen transfer to ketones was

reported by Petra et al., and was merely included in their optimization tests for their

already successful amino sulfane ligands.22 Unfortunately, such curiosity was met with

both low yields and ee values in the transformation (Scheme 18).

In that same communication from Petra et al., another set of amino sulfoxide

ligands, derived from their successful corresponding amino sulfane, was probed for

catalytic activity. In accessing the amino sulfoxide ligands, both diastereomers were

obtained and tested in the asymmetric hydrogenation of acetophenone. While the

reported stereoselectivities were modest at best, it displayed versatility of the sulfoxide

stereogenicity in imparting enantioselectivity in the resulting product. As such, the (RS)

PPh2

NH SO

OBr

L* =

3 hoursYield = 99% (S) ee = 99%

PPh2

NH SO

60 hoursYield = 5% (R) ee = 54%

PPh2

NH O2SO

60 hoursYield = 4% (R)ee = 38%

OMe

Page 38: Preparation and Activity of Multidentate beta-Amino

20

and (SS) sulfoxide diastereomers (1-15, 1-16) afforded (R) and (S) gem-phenylethanol,

respectively, while the sulfane only obtained the (S) alcohol (Figure 3).22

Scheme 18. Enhanced catalytic activity of an amino sulfane over its diastereomeric amino sulfoxide counterpart.

Figure 3. Enhanced reactivity of the (SS) sulfoxide conformation.

Amino sulfoxide ligands have also seen an increased utility in the Diels-Alder

reaction.6,12 This reaction is yet another significant carbon-carbon bond forming reaction

that has a diverse utility in asymmetric synthesis. This [4+2] cycloaddition can quickly

afford complex stereogenic centres from readily available compounds. In the

cycloaddition shown in Scheme 19, a complex mixture of four chiral compounds is

O OHHCO2H / Et3N[IrCl(COD)]2

L*

S NH2

Me

L* =

S NH2

Me

L* =

O

>99% Yield, 65% ee (Rs) = 32% Yield, 32% ee(Ss) = 9% Yield, 2% ee

1-14

S NH2

MeL* =

S NH2

Me

OO

(Rs) = 56% Yield (R)-gem-phenylethanol, 27% ee

(Ss) = 99% Yield (S)-gem-phenylethanol, 65% ee

1-15 1-16

Page 39: Preparation and Activity of Multidentate beta-Amino

21

readily afforded, which the next few examples utilized as their test reaction. As such,

asymmetric catalysis for this reaction certainly faces challenges in these stereoselective

syntheses.6,12

Scheme 19. Endo and exo adducts and respective enantiomers for the asymmetric Diels-Alder cycloaddition.

Khiar et al. first reported the use of chirally pure sulfoxide-based ligands, which

featured C2-symmetric bis(sulfoxide) motif (Scheme 20).23 The ligand successfully

imparted diastereoselectivity for the endo-adduct (96:4 endo/exo), however

enantioselectivity remained modest (ee = 56%). The ligand was postulated to form a

chelate through the sulfinyl oxygens in order to form an octahedral iron complex 1-17,

which provided enantioselectivity by limiting the diene to an underneath approach from

the less hindered side.23

To this point, amino sulfoxide and related N-derivatives had not yet been

frequently incorporated into ligands for this reaction. The work of Hiroi and co-workers

has helped bring more attention to those functionalities in ligand design.19,20,24-31 Their

work summarized two sets of chiral ligands that both utilize the oxazoline and sulfoxide

O N

OO Catalyst

H

RR

H

HR

(R)

(R)(S)

(S)

endo

exo

O N

OOR =

+

R

H

Page 40: Preparation and Activity of Multidentate beta-Amino

22

moieties, but contrast each other in the amount of rotational freedom. Their results

displayed promise; however they highlighted the challenges regarding cycloadditions of

this nature. In their 2002 communication, the reported ligands with more rotational

freedom consistently delivered great yields (often as high as 93%), with equally

impressive endo-diastereoselectivity (as high as dr = 93%).31 Their reported ligands

were also able to impart enantioselectivity for the endo-diastereomers by variation of the

substituents present on the oxazoline moiety, however the results for either enantiomer

were moderate at best (ee = 57% for (S)-endo, 66% for (R)-endo).31

Scheme 20. Test reaction for asymmetric Diels-Alder cycloaddition and the proposed active catalytic species.

In the (S)-selective ligand, the formation of a six membered copper chelate was

proposed. The chelate utilized the nitrogen and sulfinyl oxygen coordination points and

its most stable conformation 1-18 (Scheme 21) alleviates any steric clashes between

H

O NOO

78%, 96:4 endo/exo, ee = 56%

S SO OL* =

FeI

I

OOOO

SS

N

O

Proposed active catalytic species:

1-17

Page 41: Preparation and Activity of Multidentate beta-Amino

23

the tert-butyl group and the copper chelate coordinated to the dienophile. In this

conformation, the diene is afforded by an approach from the sterically less crowded Re-

face affording the (R)-endo-product, by way of the axial sulfinyl p-tolyl and tert-butyl

placements (Scheme 21).31

Scheme 21. Proposed 6-membered transition state to rationalize Re-face approach leading to (R)-endo product.

The (R)-selective ligand was also proposed to proceed through a 6-membered

chelate. However, the more flexible benzyl of 1-19 group does not create the same

interaction with the copper coordinated dienophile, and the opposite chair conformer is

preferred, with the sulfinyl p-tolyl in its equatorial position. Here, the diene approach

from the Si-face is favoured to afford the (S)-endo-product (Scheme 22).31

This model was further supported by the results of an analogous (R)-selective ligand

with an additional gem-dimethyl functionality (1-20) (Scheme 23). This ligand resulted in

SO N

O

Cu

OS

N O O Cu

SO H

N

X X

Re-face

XX 1-18

Re-face

Cu

O O

N

Cu

O O

NO O

Cu

X X= or

Page 42: Preparation and Activity of Multidentate beta-Amino

24

similar enantioselectivity, rationalized by the presence of the new methyl groups to lock

the sulfinyl p-tolyl in the equatorial position.31

Scheme 22. Proposed 6-membered transition state to rationalize Si-face approach leading to (S)-endo product.

Scheme 23. Addition of methyl group to lock p-tol group in equatorial position.

An earlier communication from Hiroi et al. reported yet another set of imino

sulfoxide ligands.29 Similarly, a high yield and diastereoselectivity for the endo-product

was observed. Moreover, a drastically improved enantioselectivity was reported,

reaching 92% ee for the (S)-product. Its effectiveness was rationalized through the

formation of a 6-membered chelate, where the sulfinyl oxygen and imine were the

coordination points (Scheme 24). Although an equilibrium exists between the two

SO N

O Cu

OS

N O

XX

Si-face

Si-face 1-19

SO N

OCu

OS

N O

XX

1-20

Page 43: Preparation and Activity of Multidentate beta-Amino

25

dienophile-coordination modes, 1-21 and 1-22, the ligands design favours an approach

of the diene from the Re-face in either conformation.29

Scheme 24. Proposed transition state to favour diene approach from the Re-face in either conformation to afford the (S)-product.

These few selected examples of nitrogen-based sulfoxide ligands are just a few

examples of their utility for asymmetric transformation. There have been many effective

ligands reported, and have been applied to a variety of carbon-carbon bond forming

reactions.

The Kiełbasinski group reported a highly versatile tridentate amino sulfoxide

ligand, along with a unique way of accessing its enantiopure sulfoxide centre.32-37 This

ligand motif continues to find utility in increasing amounts of C-C bond forming reaction,

through variation of substituents around the ligand’s amino centre. The construction of

this ligand utilizes traditional oxidation of bis(2-hydroxymethylphenyl) sulfane to afford

the corresponding prochiral sulfoxide. Next it is subjected to enzyme-mediated de-

SO

Mg

N O

OMe

OO

ON

SO

Mg

N O

OMe

OO

O N

Re-face

Re-face

1-211-22

Page 44: Preparation and Activity of Multidentate beta-Amino

26

symmetrization sulfane with CAL-B lipase and vinyl acetate to access the de-

symmetrized sulfoxide 1-23 in excellent yields and stereoselectivities, which served as a

framework to access the desired amino sulfoxide ligands. (Scheme 25).33

This ligand has been utilized for the Aza-Henry reaction with outstanding results

and required no additional metal salts (Scheme 26).34 Recently, this ligand motif has

been reported for use in the asymmetric Simmons-Smith under Sharpless conditions,

and was the first of its kind to utilize a sulfoxide-based ligand (Scheme 27).32

Scheme 25. Enzyme-mediated de-symmetrisation sulfane with CAL-B lipase and vinyl acetate to access the de-symmetrized sulfoxide.

SOHO O

S

HO OH

S*OHO OH

NaIO4

EtOH/H2O 1:1,rt

Olipase,(CAL-B)

CHCl3

O

O

98% Yield, 98% ee

1-23

1. (MeSO2)2O, Et3N, CH2Cl22. R-NH2

SOHO NHR

Amino sulfoxide ligandaccessed via CAL-B

lipase de-symmetrization

Page 45: Preparation and Activity of Multidentate beta-Amino

27

Scheme 26. Test reaction for the Aza-Henry reaction

Scheme 27. Test reaction for Simmons-Ingold cyclopropanation

Recently, Xiao et al. reported an imino sulfoxide ligand for use in this asymmetric

transformation.38 This ligand’s structural motif has served them well, and was in fact the

basis for their highly efficient amido sulfoxide ligands used in the Tsuji-Trost reaction

(Scheme 16 and Figure 1). In the initial screening parameters, these imino sulfoxide

ligands were able to achieve high yields with great stereoselectivity (up to 94%, ee =

85%), and results further improved upon optimization (Scheme 28).38

SOHO

HN

R

H

N Boc

R

HN Boc

NO2L*, Et3NMeNO2, toluene, 35 ˚C

L* = R = H: 97%, ee = 94%R = Me: 97%, ee = 96%R = MeO: 98%, ee = 94%R = NO2: 91%, ee = 86%R = Br: 95%, ee = 94%

SOHO NL* =

R2

R1 OH

R2

R1 OH*

*

L*, Et2AlCl, Et2Zn, CH2I2

DCM, rt

MeO

OH

Cl

OH

OH

Substrate Yield eeAbsolute

Configuration

91%

89%

92%

92%

94%

90%

1S,2S

1S,2S

1S,2S

Page 46: Preparation and Activity of Multidentate beta-Amino

28

Scheme 28. Imino (Schiff-base) sulfoxide ligands used in the Henry reaction.

Despite screening only one sulfoxide diastereomer 1-24, they have reported an

extensive analysis of the influence of the ligand’s moieties in this transformation. The

related sulfonyl ligand performed poorly, indicating that the coordination to the copper

centre likely occurs through the sulfur lone-pairs. Moreover, although the sulfane

derivative was able to deliver high yields, the significant decline of enantioselectivity

suggests that the stereogenic sulfinyl centre allows for proper discrimination of metal to

sulfoxide lone-pair coordination (Scheme 29).38

R H

O Cu(OAc)2•H2O, L*

tBuOH R

OHNO2

L* =

SNO

OH

R = 4-NO2Ph: 94%, ee = 93%R = 2-NO2Ph: 95%, ee = 96%R = 4-CF3Ph: 96%, ee = 94%R = 2,4-Cl2Ph: 98%, ee = 95%R = 2-OMePh: 92%, ee = 94%

1-24

Page 47: Preparation and Activity of Multidentate beta-Amino

29

Scheme 29. Probing the importance of the sulfoxide moiety

1.6 Accessing the sulfoxide moiety through sulfenate anion chemistry

In the selected examples above, the sulfoxide functionality was frequently

accessed through conventional sulfane oxidation. While this is certainly a reliable

protocol to access the sulfoxide, it offers little control over the resultant stereochemistry.

While the presence of other chiral centres may help in directing sulfur stereogenicity, the

resulting stereoinduction varies greatly from substrate to substrate.6,12,13 As a result,

recrystallization protocols have been used to access isomerically pure sulfoxides. Flash

column chromatography has also been able to separate sulfoxide isomers, however its

separation once again varies heavily with the compounds structure.6,12,13

The Schwan group has made significant inroads in the exploration of the

sulfenate anion chemistry.39-43 While this nucleophilic species can undergo “hard” O-

alkylation, the focus has mainly been in its “soft” alkylation, largely due to the ability to

H

O Cu(OAc)2•H2O, L*

tBuOH

OHNO2

L* =

SNO

OH

O2N O2N

O2SN

OH

SN

OH

94%, ee = 85% 59%, ee = 0% 94%, ee = 34%

Page 48: Preparation and Activity of Multidentate beta-Amino

30

form another stereocentre (Scheme 30).39,40,44 This conceptually novel method of

simultaneously introducing a sulfur and oxygen has seen a rapid increase of interest

and presents an alternate pathway to access the sulfoxide group.

Scheme 30. O- and S- sulfenate anion alkylation

The work of Söderman and Schwan in regards to sulfenate anion chemistry has

resulted in a highly diastereoselective synthesis of (RC, SS) or (SC, RS) β-amino

sulfoxides.42,43 The sulfenate anion was accessed via a nucleophilic addition-elimination

mechanism from a 2-sulfinyl acrylate precursor. In short, a homochiral amino iodide

electrophile was utilized in directing stereogenicity at the sulfur centre in the resulting

sulfoxide. The (S)-β-amino iodide electrophiles yielded (RS,SC)-β-amino sulfoxides,

whereas (R)-β-amino iodide electrophiles yielded (SS,RC)-β-amino sulfoxides. A

thorough optimization concluded that sulfenates typically required an aromatic

substituent to be efficiently released from its 2-sulfinyl acrylate form. This method of

generating sulfoxides provided moderate to high yields (up to 91%), and diastereomeric

ratios as high as 94:6.42,43 Such diastereoselectivity was rationalised through internal

complexation of the lithium counterion with the electrophile’s nitrogen lone pair to create

a stable chair form to impart stereoselectivity (Figure 4).43

R S O M R SO

M

R'-XS-Alkylation

R'-XO-Alkylation

R S O R'R SO

R'

Page 49: Preparation and Activity of Multidentate beta-Amino

31

Figure 4. Proposed internal complexation transition state to account for observed sulfenate diastereoselectivity with homochiral amino electrophiles.

Building on these results, we propose to build heteroaryl β-amino sulfoxide

ligand, and probe them for catalytic activity in the Henry reaction. We propose to build

these ligands using traditional oxidative syntheses and sulfenate anion chemistry, with

the aim of probing diastereopure conformations of each of these chiral ligands.

SO Li N

HetI

BnH

HBoc

via:

Page 50: Preparation and Activity of Multidentate beta-Amino

32

Chapter 2: Results and Discussion

Page 51: Preparation and Activity of Multidentate beta-Amino

33

2.0 Results and Discussion

2.1 Chiral ligand design and synthesis

The overall ligand design of the multidentate β-aminosulfoxide ligands can be

deconstructed into two fragments consisting of a nucleophilic source of sulfur, and an

electrophilic amine-containing alkyl group. In an attempt to access diastereopure

compounds, two pathways were envisioned for the construction of these amino-

sulfoxide ligands (Scheme 31). The first pathway begins with the construction of an β-

aminosulfane intermediate, which is to be followed by oxidation to access the β-

aminosulfoxide target (left, Scheme 31). The second pathway involves generation of the

heteroaryl sulfenate anion, which would lead directly to the desired target (right, Scheme

31). Reports from Söderman and Schwan indicate that the sulfenate anion reacted with

an N-Boc amino-acid based electrophile would favour a diastereomeric (SS,RC)

outcome. While the traditional oxidative approach typically lacks diastereoselectivity,

this method may provide a means of accessing the complementary (RS,RC)

diastereomer. Finally, upon accessing the targeted ligand, the N-Boc protecting group

will be removed to yield the desired ligand for use in catalytic studies.

Page 52: Preparation and Activity of Multidentate beta-Amino

34

Scheme 31. Retrosynthethic approach towards the construction of heteroaryl amino sulfoxide ligands.

2.2 Synthesis of β-amino electrophile

The preparation of the electrophilic amine-containing portion began with the

reduction of N-Boc protected L-phenylalanine as the initial source of chirality. It was

envisioned that upon reduction the hydroxyl group would be converted to a mesylate,

thus providing a suitable leaving group (Scheme 32).

Scheme 32. Reduction of N-Boc protected L-phenylalanine

Despite the facile acquisition of this mesylated amino compound 1-25, it proved

to be unsuitable as an electrophile for our purposes. Its reaction with both sulfenate and

thiolate nucleophiles were unsuccessful. Trial reactions with p-tolyl sulfenate anions with

varying amounts of mesylated electrophile resulted in no sulfoxide generation and

recovery of the electrophile (Table 1, entries 1-3). It can be envisioned that the presence

HetAr S*O Bn

NHBocHetAr SH LGBn

NHBoc HetAr S OM LGBn

NHBoc++

Followed by oxidation of corresponding sulfane

Traditional Oxidation Sulfenate Anion

HetAr = 2-pyridyl, 8-quinolyl

HO

ONHBoc

Bn

HO NHBoc

BnO NHBoc

Bn

SMeO O

1. EtOC(O)Cl, Et3N,THF, 0 ˚C to rt

2. NaBH4, H2O,0 ˚C to rt

MsCl, Et3N,0 ˚C to rt

(81%) (94%)1-25

DCM

Page 53: Preparation and Activity of Multidentate beta-Amino

35

of Li+ ions, arising from the required addition of nBuLi, could potentially complex with the

similarly “hard” nucleophilic oxygens form the mesylate moiety, creating steric bulk, and

thereby preventing sulfenate attack. Alternatively, the complexation effect may position

the sulfenate in a manner that precludes backside substitution on the mesylate.

Similarly, thiolate reactions starting 2-mercaptopyridine provided a similar outcome

(Entries 4 & 5, Table 1).

Due to the unsuccessful reactions of sulfur-based nucleophiles with these

mesylated compounds, further investigation with them was not pursued, and efforts to

access an iodinated electrophile, similar to those in the sulfenate reactions reported by

Söderman and Schwan, were made. The initial synthesis of this electrophile proved to

be more difficult, due to their lower yields, difficulty in purification, and the potential

formation of a recalcitrant by-product.

Initial reaction conditions described the addition of an iodine solution in DCM to a

solution of Ph3P and imidazole in DCM (Scheme 33). However, the iodine never fully

dissolved before it was transferred. This led to unreacted Ph3P, which typically co-eluted

with the iodinated product during flash column chromatography, even with low-polarity

mobile phases (95:5 hexanes/EtOAc). Upon formation of the amino iodide it was not

particularly stable. It had a tendency to react intramolecularly when heated or subjected

to aqueous work-up, resulting in the formation of benzylated cyclic carbamate 1-26

through the loss of the N-Boc group (Scheme 33).

Page 54: Preparation and Activity of Multidentate beta-Amino

36

Table 1. Substitution attempts on the amino mesylate

Scheme 33. Synthesis of amino iodide electrophile and formation of cyclic by-product. a Dissolve iodine in DCM and transfer resultant solution to Ph3P and imid. in DCM.

S OLi

Nucleophile

N

SH

Equivalents ElectrophileEntry

1.0

1.5

2.0

1.0

1.0

1

2

3

4

5

Conditions

Sulfenate generation

from 2-sulfinylacrylate a

Hünig's base

Et3N, K2CO3

NucleophileMsO

Bn

NHBoc NuBn

NHBoc

aS*

ptol

O CO2Me nBuLi

(1.0 equiv)+

HO NHBoc

BnI NHBoc

BnPh3P, imid, I2,DCM

0 ˚C to rt

IBn

NH

O OO

NH

O

Bn

ONH

O

Bn

H

BH +

BB

Intramolecular formation of by-product:

(73%)

1-26

Page 55: Preparation and Activity of Multidentate beta-Amino

37

In order to create a more favourable outcome, Ph3P, imidazole, and iodine were

placed in the same reaction vessel, and cooled to -78 ˚C, before solvent was added.

This mixture was stirred until all iodine appeared to be in solution, which would allow

complete reaction with Ph3P. Upon addition of the amino alcohol and a slow

temperature increase to ambient conditions, the crude reaction mixture contained no

Ph3P as confirmed by TLC. To circumvent by-product formation, the crude reaction

mixture was placed directly under reduced pressure, concentrated, and subsequently

loaded into a short silica plug for facile purification (10 g silica/1 g crude material; eluted

with 90:10 hexanes/EtOAc; yield = 67 – 73%). While the yields for this reaction were

only marginally increased, these updated conditions allow for quicker and undemanding

access to this amino iodide electrophile.

2.3 Heteroaryl thiols

The amino iodide electrophile proved to be a suitable electrophile for both

sulfenate and thiolate nucleophiles. Both of the proposed pathways for our ligand

construction began with either 2-mercaptopyridine or 8-mercaptoquinoline (1-27). While

the former is readily accessible, the latter is very costly and multi-gram quantities are

not commercially available. Consequently, more cost-effective precursors were sought.

One protocol began with 8-bromoquinoline, which was treated in separate experiments,

with potassium thioacetate and potassium thiocyanate with the goal of effecting an SNAr

substitution. Traditionally, this reaction requires the presence of strong electron-

Page 56: Preparation and Activity of Multidentate beta-Amino

38

withdrawing groups (e.g. p/o-NO2) on the arene to invoke nucleophilic attack at the ipso-

position.45,46 Here, it was hypothesized that the present quinolyl system could act as the

analogous electron-withdrawing group (Scheme 34). Unfortunately, all attempts towards

an SNAr reaction proved unsuccessful, as starting material failed to react (confirmed via

TLC).

An alternative pathway to access 8-mercaptoquinoline began with the reduction

of its corresponding sulfonyl chloride derivative. A report from Akamanchi et al. detailed

a quick and efficient reduction of this moiety using Ph3P (Scheme 35).47 With their

method, the reduction of 8-quinoline sulfonyl chloride yielded a 72% yield of the

corresponding thiol. No mechanism was included in their report, however one can

envision that the reaction proceed employing Ph3P as the principal reducing agent. The

use of base allows for the access of the corresponding thiolate and acidification affords

the desired thiol 1-27 (Scheme 35).

Scheme 34. Attempts towards SNAr substitutions with potassium thioacetate and potassium thiocyanate.

N

BrK-S-X

N

S X

Key Intermediate

HydrolysisN

SH

K-S-X Conditions

K-S-CN

K-S-C(O)Me

K-S-C(O)Me

DMSO, rt

DMSO, 80 ˚C

DMSO, 80 ˚C

1-27

Page 57: Preparation and Activity of Multidentate beta-Amino

39

Scheme 35. Reduction of sulfonyl chloride moiety to a thiol vial Ph3P.

2.4 Traditional oxidative synthesis of β-amino sulfoxide ligands

In order to access the desired amino-sulfoxide ligands via traditional oxidative

means, the corresponding sulfane 1-28 must first be accessed (Scheme 36). The

optimal conditions for this substitution utilize two equivalents of Hünig’s base and two

equivalents of the amino iodide electrophile. It is imperative that the reaction takes place

under an inert atmosphere, to prevent thiolate oxidation and heteroaryl disulfide

formation. In addition, the single equivalent excess of electrophile aids to drive the

desired substitution reaction forward. Similar reactions, which lacked meticulous effort in

maintaining an inert atmosphere and only one equivalent of electrophile, consequently

formed the disulfide compounds. These by-products have very similar polarities to their

8-quin SO O

Cl PPh3(1st equiv)

8-quin SO O

8-quin SO

O+ PPh3Cl

Ph3P(2nd equiv) Ph3P(O)

8-quin SO

O PPh3

8-quin SO

PPh38-quin S O PPh3Ph3P

(3rd equiv)

8-quinSPh3PPh3P(O)

HO8-quin S

H+

NSH

(76%)

Ph3P(O)

- Cl-

1-27

Page 58: Preparation and Activity of Multidentate beta-Amino

40

corresponding amino-sulfane compounds, which resulted in an arduous purification via

flash-column chromatography and a need for recrystallization afterwards.

Scheme 36. Thiolate substitution to access amino sulfane compounds

Various oxidation protocols were utilized on the amino-sulfane compounds to

probe for any diastereoselective preferences during sulfoxide generation (Table 2).

While diastereomeric ratios of corresponding amino-sulfoxides 1-29 were similar and

modest at best, they provided insight into the effectiveness of each protocol for use in

future sulfane oxidations. Trials of various oxidation procedures led us to prefer

mCPBA-mediated oxidations. These provided the highest yields and typically consumed

nearly all of the starting heteroaryl amino-sulfane materials. Moreover, these reactions

required the least time compared to the other methods (8 - 10 hours). Over-oxidation to

the corresponding sulfone is sometimes a concern when using mCPBA, however this

was circumvented by starting the oxidation at -78 ˚C for the first 3 hours and then raising

it to -35 ˚C until the reaction was complete.

Although oxidations with NaIO4 hold promise due to their undemanding

conditions, its utility in this case was diminished. With this protocol appreciable amounts

of both heteroaryl amino sulfanes were still present after a 48-hour reaction time. The

sluggishness could be attributed to the requisite 1:1 water/MeOH solvent system, which

HetArSH I

NHBoc

Bn Hünig's Base(2 equiv)

DCM, 0 ˚C to rt+

(2 equiv)

SNHBoc

Bn

HetAr

HetAr:2-pyridyl = 93%8-quinolyl = 96%

1-28

Page 59: Preparation and Activity of Multidentate beta-Amino

41

provide less than ideal solubility for starting materials. Similarly, oxidations with aqueous

H2O2 in hexafluoroisopropanol (HFIP) also yielded comparably poor results, in which a

48-hour reaction time left unreacted starting material.

With a preferred method of oxidation in hand, diastereomeric mixtures of the

target heteroaryl amino-sulfoxide ligands were accessed. Although these ligands were

oxidized without any appreciable diastereoselectivity, the literature has provided many

examples where diastereopure sulfoxides were isolated from a diastereomeric mix via

recrystallization protocols. Recent communications from van Leeuwen et al. and Xiao et

al. highlighted this and reported the isolation of at least one diastereopure β-amino-

sulfoxide from a mixture.21,22,38

Table 2. Oxidation conditions of amino sulfane compounds

a d.r. determined via 1H NMR analysis

S NHBoc

Bn

HetAr S* NHBoc

Bn

HetAr

O[O]

Entry HetAr [O] d.r.a

2-pyridyl

2-pyridyl

2-pyridyl

8-quinolyl

8-quinolyl

8-quinolyl

1

2

3

4

5

6

mCPBA

NaIO4

H2O2; HFIP

mCPBA

NaIO4

H2O2; HFIP

Yield (%)

1.4 : 1.0

1.0 : 1.2 95

50

1.0 : 1.3

1.0 : 1.3 63

86

1.0 : 1.0

1.0 : 1.0

45

53

1-29

Page 60: Preparation and Activity of Multidentate beta-Amino

42

Despite having a similar β-amino-sulfoxide motif, the newly acquired ligands

could not be separated by this method. Both amino-sulfoxide ligands were each

recrystallized in a variety of solvent system at varying temperatures which include:

hexanes/EtOAc, pentane/EtOAc, Et2O, pentane/DCM, toluene, toluene/EtOAc,

pentane/EtOH, and Et2O/EtOH. Despite numerous attempts, crystals that formed over

the duration of the recrystallization always contained a diastereomeric mix, with an

expected tendency to favour the diastereomer formed in slight excess during oxidation.

A d.r. = 1.0:2.1 and 1.0:2.3 for the 2-pyridyl and 8-quinolyl-based ligands was obtained

using a pentane/DCM system. However, the reproducibility was very difficult, and

repeated recrystallization on the isolated fractions proved difficult due to minimal

quantities initially isolated. Furthermore, once the mother liquors were re-concentrated

and subjected to further recrystallization, they failed to yield any further separation.

Difficulty in separation arose due the similar rates of crystal formation of each

diastereomer from solution. Therefore a different approach towards recrystallization was

taken (Table 3). Upon initial uniform crystal formation (Fraction 1; typically within the first

five minutes), the hot mother liquor was then decanted, into another heated vessel

(Fraction 2), and this was repeated one final time (Fraction 3). This process was

repeated three times for each heteroaryl amino-sulfoxide ligand. This yielded a better

initial results in the first fraction, however subsequent fractions led to the

recrystallization of both diastereomers as previously observed.

Unfortunately, recrystallization was thus an ineffective method of isolating

diastereopure ligands. Nonetheless, the removal of the N-Boc moiety was carried out, to

Page 61: Preparation and Activity of Multidentate beta-Amino

43

access their free-base forms for use in catalytic trials (Scheme 37). Recrystallization

attempts with the free-base amino-sulfoxide compounds again yielded disappointing

results, and seemingly again caused by similar rates of precipitation of each

diastereomer from solvent.

Table 3. Diastereomeric ratio of recrystallization fractions

Scheme 37. Removal of N-Boc protecting group via trifluoroacetic acid.

A communication from Jiang et al. offered a potential solution to the problem at

hand. The report featured the synthesis and purification of a series of diastereopure β-

amino-sulfoxide ligands, which possessed carbon chirality arising from amino acid

backbones and an N-Ts moiety (Scheme 38). Upon accessing N-Ts β-amino-sulfoxide

N

S*O

NHBoc

Bn

N

S*O

NHBoc

Bn

Fraction initial d.r. = 1.0 : 1.3 initial d.r. = 1.0 : 1.2

1

2

3

1.0 : 2.1 1.0 : 2.3

1.0 : 1.1

1.1: 1.0

1.0 : 1.2

1.0: 1.1

DCM, 0 ˚C to rtS* NH2

Bn

HetAr

HetAr:2-pyridyl (1-L1m) = 78%8-quinolyl (1-L2m) = 81%

OS* NHBoc

Bn

HetAr

O CF3C(O)OH(excess)

Page 62: Preparation and Activity of Multidentate beta-Amino

44

1-30 through non-diastereoselective means, simple flash-column chromatography

proved to be sufficient for isolation of diastereomeric compounds.48

Scheme 38. Reported synthesis of N-Ts protected amino sulfoxide ligands

Aiming to isolate diastereopure compounds, the free-base hetereoaryl amino-

sulfoxide ligands at hand were subjected to N-tosylation (Scheme 39). Disappointingly,

TLC analysis of both heteroaryl compounds in 10:90 MeOH/DCM yielded only one UV-

active spot, indicating failure to separate. In addition, recrystallization of 1-31 resulted in

a rapid and quantitative precipitation of both diastereomers. Although tosylated

compounds are known to crystallize easily, a newly introduced N-Ts moiety on the

heteroaryl amino-sulfoxide compounds could perhaps override each diastereomers

mildly different individual tendency to form crystals or precipitate from solution.

Scheme 39. N-Ts protected amino sulfoxides

tBu S NHTs

Bn

NH2

BnHO

mCPBA

DCM, 0 ˚C tBu S* NHTs

BnO

initial d.r. = 1.2 : 1.0

1-30

DCM, 0 ˚C to rtS* NHTs

Bn

HetAr

HetAr:2-pyridyl = 93%8-quinolyl = 86%

OS* NH2

Bn

HetAr

O TsCl

1-31

Page 63: Preparation and Activity of Multidentate beta-Amino

45

Unfortunately, ligands constructed via traditional oxidative means did not yield

diastereomerically pure compounds. Nonetheless, these diastereomeric mixtures of

ligands can still be probed for catalytic activity and comparison of results from a

diastereopure ligand can offer insight to the effectiveness of each configuration.

2.5 Sulfenate anion chemistry

To construct the proposed heteroaryl amino-sulfoxide ligands via the sulfenate

pathway, the necessary heteroaryl sulfenate anion was generated from a 2-sulfinyl

acrylate precursor, as previously reported by the Schwan group. The synthesis of the

sulfenate synthon began with a Michael addition of the heteroaryl thiols to methyl

propiolate, which resulted in a mixture of cis- and trans-2-sulfanyl acrylates 1-32. These

were then subjected to different oxidation procedures to yield the requisite sulfinyl

acrylates 1-33 (Scheme 40). Oxidation with mCPBA, while maintain a cold temperature,

once again proved to be the most efficient way of oxidation. As seen before, allowing

this reaction to reach room temperature resulted in an undesired mixture of unreacted

sulfanes, sulfoxides, and sulfone. Similarly, oxidations with NaIO4 and H2O2/HFIP

yielded sluggish reaction rates and messy crude reaction mixtures.

The heteroaryl sulfenate anion was generated through an addition/elimination

mechanism using n-butyl lithium as a nucleophile. Upon sulfenate release, two

equivalents of the amino iodide electrophile in solution were added rapidly (Scheme 41).

With this sulfoxidation method, Söderman and Schwan reported excellent

Page 64: Preparation and Activity of Multidentate beta-Amino

46

diastereoselectivities when using similar aryl sulfenates and amino electrophiles and X-

ray crystallography studies confirmed an absolute stereochemistry of (SS,RC) of their

major diastereomers.42,43 Furthermore, they rationalized that such diastereoselectivity

arose from an internal complexation of the lithium counterion with the nitrogen lone-pair

of the amino electrophile, in turn creating a stable chair transition state 1-34.43 Here, a

similar rationale was adopted due to the similarities between their compounds and our

targets (Scheme 41).

However, results obtained from the 2-pyridyl sulfenate anion provided less than

promising results. A diastereomeric ratio of 1.0:0.7 was obtained and such reaction

favoured the same diastereomer that mCPBA oxidation provided. The Schwan group

has synthesized this molecule previously and 1H NMR spectroscopy comparisons

suggest the same major isomer was formed in both cases. In contrast, the 8-quinolyl

sulfenate anion provided one lone diastereomer, which was assumed to possess

(SS,RC) chirality (Scheme 42).

Page 65: Preparation and Activity of Multidentate beta-Amino

47

Oxidant Conditions 2-pyridyl 8-quinolyl

NaIO4 1:1 MeOH/H2O Slow reaction time

H2O2 HFIP Slow reaction

time, messy crude reaction material

Slow reaction time

mCPBA DCM, -78 ˚C to rt Mixture of starting material, sulfoxide and sulfone formed, messy crude reaction material

Complete consumption of starting material, only sulfoxide formed

mCPBA DCM, -78 ˚C 3 hours, 35 ˚ C till completion

Complete consumption of starting material, only sulfoxide formed

Scheme 40. Conditions for sulfanyl acrylate oxidations

Scheme 41. Addition-elimination mechanism for the generation of a lithiated sulfenate anion.

HetAr SH C(O)OMe+1. Et3N, DCM, 0 ˚C to rt

2. H3O+HetAr S C(O)OMe HetAr S* C(O)OMe

O[O]

HetAr:2-pyridyl = 84% 8-quinolyl = 87%

HetAr:2-pyridyl = 81% 8-quinolyl = 77%

1-331-32

HetAr SO

OMe

O

nBuLi

nBu LiHetAr S

OOMe

OnBu Li

HetAr SO

Li HetAr SO Li

HetAr S OLi

-78 ˚CC(O)OMe

nBu

I NHBoc

Bn

-78 ˚C to rtS NHBoc

Bn

HetAr

O

SO Li N

HetArI

BnH

HBoc

1-34

Page 66: Preparation and Activity of Multidentate beta-Amino

48

Scheme 42. Heteroaryl sulfenate anion substitution with homochiral amino electrophile

2.6 Attempts towards enantioselective sulfoxide generation

Given the excellent diastereoselectivity demonstrated by the 8-quinolyl sulfenate

anion, the stereoselective capacity of these heteroaryl sulfenates was further

investigated. These heteroaryl sulfenates feature an additional N-coordination group,

which could form an intramolecular rigid, cyclic complex. Perrio et al. have proposed a

similar intramolecular sulfenate model, 1-35, of complexation in their attempts towards

stereoselective sulfoxide generation (Scheme 43).49

S NHBoc

BnO

N

S NHBoc

BnO

N

82%; d.r. = 1:072%; d.r. = 1.0:0.7

HetAr SO

OMe

O I NHBoc

Bn S NHBoc

Bn

HetAr

O1. nBuLi, THF, -78 ˚C2.

-78 ˚C to rt

Page 67: Preparation and Activity of Multidentate beta-Amino

49

Scheme 43. Lithiated intramolecular sulfenate complexation

In probing this model, it was suggested that the lithiated sulfenate could be

complexed further with an additional chiral ligand to form 1-36. The lithium counterion

should favour an O-Li coordination, allowing the formation of a spirocyclic lithium centre,

and in turn the possibility for sulfur lone pair discrimination leading to stereochemical

induction (Scheme 44).49-52

Scheme 44. Spirocyclic chiral Lewis acid catalyst for enantioselective sulfoxide formation.

Investigation of this concept began with the generation of the sulfenate anion

from the prerequisite 2-sulfinyl acrylates. A solution of chiral PyBox ligand (10 mol%,

catalytic quantity) was then introduced, in the hope of forming the chiral complex. The

SOLi

Me

N(Me)2Me

H NMe Me

S OLi

Proposed model:

1-35

R-X

S ON Li

X

XHetAr S

OOMe

O

nBuLi-78 ˚C HetAr

SOLi

L*S ON Li

Proposed activecatalytic species

Lone-pair discrimination

1-36

Page 68: Preparation and Activity of Multidentate beta-Amino

50

N-coordination sites were chosen for the propensity to coordinate with the similarly

“hard” lithium cation.49,51 The reaction protocol was completed by the addition of the

benzyl bromide in solution and a slow increase to room temperature while stirring for 10

hours (Scheme 45).

Attempts at stereoselective sulfoxide formation were unsuccessful and

accompanied by low yields of the heteroaryl benzyl sulfoxide. The low yields of product

suggest that if complexation was to occur successfully, it does not efficiently facilitate

the alkylation at the stereogenic sulfur centre, let alone with any appreciable

enantioselectivity. Moreover, this lack of stereoselectivity suggests that the complexed

sulfenate anion reacts at a slower rate than one without the PyBox ligands. The

literature suggests that such complexation is viable, and these results here only show a

limited screening of parameters. Another report from Perrio et al. investigated the use of

(–)-spartene as a lithiathed suflenate complexing agent for stereoselective sulfoxide

generation. Similarly, complexed sulfenate anions delivered lower yields and low

enantioselectivities (4 – 59% yields; 0 – 23% ee). Interestingly, they noted that as the

reaction temperature increased, the sulfenate anion delivered higher yields albeit with

lower accompanying %ee values.51

Page 69: Preparation and Activity of Multidentate beta-Amino

51

Scheme 45. Enantioselective sulfoxide formation attempts with chiral PyBox ligands

2.7 Attempts towards [SO]-2 extrusion

One final investigation probing the reactivity of these 2-sulfinyl acrylates

attempted to invoke [SO]-2 extrusion from these compounds. The concept of SO

extrusion has been reported by Hu et al. from 2-pyridyl sulfoxides for use in the Julia-

Kocienski olefination.53 Previously, the authors reported that gem-difluorosulfones can

display increased reactivity for this olefination, and as such, their related gem-

difluorosulfoxide derivatives 1-37 were probed for use in this reaction. Interestingly, the

presence of the gem-fluorides yielded an unexpected major product 1-38, which was

thought to arise from the mechanism in Scheme 46 involving SO extrusion.53

HetArSO

OMe

OHetAr

S*O

NN

OO

N

Q

1-L* PyBox1 =

nBuLi-78 ˚C, THF HetAr

SOLi

SO

N Li

N

N

N-78 ˚C to r.t.,

THF

BnBrL*

PyBox

THF

NN

OO

N

1-L* PyBox2 =

HetAr L* Yield (%) Ee

2-pyridyl 1-L*PyBox1 50 0%

2-pyridyl 1-L*PyBox 2 49 0.9%(S)

8-quinolyl 1-L*PyBox 1 55 1.4% (R)

8-quinolyl 1-L*PyBox 2 56 0%

Page 70: Preparation and Activity of Multidentate beta-Amino

52

Scheme 46. Proposed mechanism of SO extrusion.

To invoke a related [SO]-2 extrusion with the heteroaryl sulfinyl acrylates in hand,

sulfenate anion 1-39 was generated through normal means. However, an additional

equivalent of nBuLi and was added aiming for an ipso-nucleophilic attack to form 1-40

Moreover, the temperature was raised to room temperature before the benzyl bromide

electrophile was added. It was proposed that a similar addition-elimination pathway

would invoke [SO]-2 extrusion (Scheme 47). The resulting sulfur monoxide dianion would

be expected to react with benzyl halide in consecutive reactions.

N

S CHXY

O 1. LiHMDS (2 equiv), DMF -78 ˚C, then HCl (2M)

2. DBU (2 equiv), DMF, rt

If X = Y = For X = H, Y = F

If X = Y = H

Ar

O

Ar+

ArAr

Y X

N

S

O

OYF

ArAr

N

O

F

Ar Ar

NO

SO

F YArAr

N

O

Ar ArS

F Y

N

O

O - F-

ArAr

Y SO

N

O

X S O

ArAr N

O

X S O

ArArN

O

F

Ar Ar

Via:

- SO

1-37

1-38

1-38

Page 71: Preparation and Activity of Multidentate beta-Amino

53

Scheme 47. Attempts towards [SO]-2 extrusion via double nBuLi addition and release.

Unfortunately, 1H NMR analysis of crude reaction mixtures provided no evidence

of dibenzylsulfoxide formation nor the accompanying butylpyridine (Scheme 47).

Specifically, the aromatic region of each spectrum displayed only one set of heteroaryl

aromatic peaks corresponding to the heteroaryl benzyl sulfoxide compound. The first

experiment utilized the 2-pyridyl sulfinyl acrylate and its failure could perhaps be

attributed to the potential insufficient electron delocalization, which prevents the second

nucleophilic ipso-substitution. Thus, the quinolyl sulfinyl acrylate underwent similar

treatment, however similar results were obtained. Given the presence of the heteroaryl

benzyl sulfoxide, the second ipso-substitution could be attributed as the barrier for [SO]-2

extrusion.

2.8 β-Amino sulfoxide ligands and related derivatives

HetArSO

OMe

O

HetArSO

1. nBuLi (2 equiv.)2. BnBr-78 ˚C to r.t.

THF

Not observed:

BnS

Bn

O

SO

OMe

ON N

S OLi

BuLi N

S OLi

Bu

S OLi

Li

Li

nBuLinBuLi nBuLi

N

Bu

Via proposed mechanism shown with 2-pyridylsulfinyl acrylate:

1-39 1-40

HetAr:2-pyridyl = 40%8-quinilyl = 50%

Page 72: Preparation and Activity of Multidentate beta-Amino

54

In summary, two diastereomeric mixtures of ligands (1-L1m, 1-L2m, Figure 5)

were prepared via traditional oxidative means while one diastereopure ligand (1-L2,

Figure 5) was accessed via sulfenate anion chemistry, and subsequently subjected to

N-Boc deprotection (Scheme 48).

Figure 5. Chiral amino sulfoxide ligands

Ligand L2 was further modified through condensation with an additional

heteroaryl aldehyde. A report from Xiao et al. showcased the effectiveness of the β-

imino sulfoxide moieties. Given the similar structures of these ligands, the opportunity to

follow suit was presented. As such, diastereopure 1-L2 underwent condensation with

salicylaldehyde and 2-picolinaldehyde to yield 1-L3 and 1-L4, respectively (Scheme 48).

These newly accessed ligands offer a new imine moiety, as well as an additional

coordination site to a metal centre.

S NH2

BnO

NS* NH2

Bn

N

S* NH2

Bn

N

1-L1m 1-L2m 1-L2

OO

Page 73: Preparation and Activity of Multidentate beta-Amino

55

Scheme 48. Imino sulfoxide ligand synthesis

2.9 Probing chiral ligands for catalytic activity in the Henry (nitroaldol) reaction

The chiral ligands were probed for asymmetric catalytic activity in the Henry

reaction. A review of the literature led to the initial screening conditions.37,38,54-57 Ligand

screening tests within the literature have reported excellent yields and

enantioselectivities with the use of alcoholic solvents and Cu(OAc)2�H2O as the

catalyst.37,38,54,55 Although initial screening reactions with various alcohols led to

moderate to excellent yields, they were accompanied by low enantioselectivities (Table

4). Ligands 1-L1m – 1-L4 were screened within many solvents and did not yield any

promising results. However these screening reactions can offer insight into their

SNHBoc

BnO

N

SNH2

BnO

DCM, 0 ˚C to rt

CF3C(O)OH(excess)

N

1-L2 = 82%

H

O

3:1 DCM/MeOHNa2SO4, reflux

OH

H

O

3:1 DCM/MeOHNa2SO4, reflux

N

S N

BnO

N

1-L3; 91%

OH

S N

BnO

N N

1-L4; 89%

Page 74: Preparation and Activity of Multidentate beta-Amino

56

catalytic activities. These chiral ligands all favoured the formation of the (R)-isomer

(Table 4). In general, the imino sulfoxide ligands (1-L3 and 1-L4) led to higher yields

and ee’s than their corresponding free base form (1-L2) and nitromethane as a solvent

gave to the highest observed enantioselectivities (Entries 16 and 25, Table 4).

Although the complementary diastereomer of 1-L2 was not able to isolated as a

diastereopure ligand, the results from a reaction catalyzed by a diastereomeric mixture

indicate that it does not possess any more catalytic activity (1-L2m: 40% yield, 3.4% ee;

1-L2: 44% yield, 14% ee; Entries 3 and 7, Table 4).

These ligands underwent further screening in the Henry reaction using a

nitroethane (Table 5). Here, the emergence of diastereomeric nitro alcohols allowed the

use of 1H NMR to determine a diastereomeric ratio. Unfortunately, these ligands failed

to impart any appreciable diastereoselectivity, and as such the enantioselectivities from

these reactions were not pursued. Moreover, unlike when nitromethane was used as the

solvent, a nitroethane solvent system failed to improve the diastereomeric ratio (Table 5,

entries 3 and 8).

Page 75: Preparation and Activity of Multidentate beta-Amino

57

Table 4. Solvent screening reactions for the Henry Reaction with nitromethane.

Entry L* Solvent Yield (%) ee (%)a

1 1-L1m MeNO2 37 2.4

2 1-L2m EtOH 55 3.2 3 1-L2m MeNO2 40 3.4 4 1-L2 Toluene 3 24 5 1-L2 MeCN 36 6.7 6 1-L2 CH2Cl2 trace -- 7 1-L2 MeNO2 34 14 8 1-L2 THF 20 10 9 1-L2 DMF 76 11 10 1-L2 EtOH 79 17 11 1-L2 iPrOH 61 6.7 12 1-L2 tBuOH 41 17 13 1-L3 Toluene 35 27 14 1-L3 MeCN 13 21 15 1-L3 CH2Cl2 38 29 16 1-L3 MeNO2 52 29 17 1-L3 THF 72 23 18 1-L3 DMF 99 1.7 19 1-L3 EtOH 50 15 20 1-L3 iPrOH 79 17 21 1-L3 tBuOH 72 15 22 1-L4 Toluene 32 24 23 1-L4 MeCN 59 14 24 1-L4 CH2Cl2 14 25 25 1-L4 MeNO2 73 32 26 1-L4 THF 89 17 27 1-L4 DMF 99 8.8 28 1-L4 EtOH 68 11 29 1-L4 iPrOH 92 13 30 1-L4 tBuOH 91 14

a Determined by chiral HPLC, Chiralcel OD-H column 0.8 mL/min; 85/15 (v/v) Hexanes/IPA

O2N

H

O

O2N

OHNO2

L* (12 mol%) Cu(OAc)2•H2O (10 mol %)

MeNO2 (10 equiv.)Solvent, rt48 hours

Page 76: Preparation and Activity of Multidentate beta-Amino

58

Table 5. Solvent screening reactions for the Henry Reaction with nitroethane

Entry L* Solvent Yield (%) anti:syna 1 1-L1m EtNO2 2 1-L2m EtOH 40 1.21:1.0 3 1-L-2m EtNO2 62 1.29:1.0 4 1-L2 EtOH 64 1.37:1.0 5 1-L2 DMF 46 1.53:1.0 6 1-L3 EtOH 62 1.03:1.0 7 1-L3 DMF 55 1.56:1.0 8 1-L3 EtNO2 71 1.20:1.0

9 1-L4 EtOH 74 1.0:1.0 10 1-L4 DMF 77 1.05:1.0

a determined via 1H NMR analysis

To this point, only 4-nitrobenzaldehyde had been utilized in the Henry reaction.

Although results from the screening reactions, did not yield any exceptional results,

these trials suggested that either ligand 1-L3 or 1-L4 with Cu(OAc)2�H2O in

nitromethane would offer marginally improved enantioselectivities (Table 4, entry 16 and

25). Thus, 4-methylbenzaldehyde, 4-bromobenzaldehyde, and 2-

thiophenecarboxaldehyde were probed as substrates under those conditions (Scheme

50). Unfortunately, none of these substrates were able to produce the nitro alcohol

product. After 48 hours, TLC analysis indicated that no reaction had taken place.

O2N

H

O

O2N

OHL* (12 mol%) Cu(OAc)2•H2O (10 mol%)

EtNO2 (10 equiv.)Solvent, rt48 hours

NO2 O2N

OH

NO2

+

anti syn

Page 77: Preparation and Activity of Multidentate beta-Amino

59

Therefore, 10 mol% of Et3N was added to each reaction with the aim of moving the

reaction forward. However, after an additional 24 hours, TLC and 1H NMR analysis

again indicated that no desired product had been formed.

Disappointingly, the scope of these ligands for this reaction therefore appears to

be very limited. The failure of 4-bromobenzaldehyde to react suggests that this catalyst

system requires a greater electron deficiency in order to yield the desired product.

Moreover, enhanced reactivity of 4-nitrobenzaldehyde suggests that it is not a suitable

screening substrate, since it is not an accurate indicator of catalytic activity.

Scheme 49. Probing of other aldehydes as substrates in the Henry reaction under optimized conditions.

2.10 Proposed model for observed stereoinduction

All ligands probed in the Henry reaction using a nitromethane nucleophile led to a

preference for the R isomer, albeit with low enantioselectivity (Table 4). However, the

diastereopure quinolyl amino sulfoxide ligands (1-L3 and 1-L4) provided the highest

observed enantioselectivities (ee = 29%, 32%; Entry 16 and 25, Table 4). To rationalize

this stereoselectivity, three assumptions were made. (1) Although these ligands have

Ar H

O

N

SO

N

Bn

H

OH

1-L3

1-L3 (12 mol%)Cu(OAc)2•H2O (10 mol%),

MeNO2, rt, 48 hours

(10 mol% Et3Nafter 48 hours)

+ MeNO2(10 equiv)

Ar

OHNO2*

Not observed

Ar:4-Me-C6H44-Br-C6-H42-thienyl

Page 78: Preparation and Activity of Multidentate beta-Amino

60

multiple possible coordination sites, a bidentate coordination mode is assumed for these

hypothesized models; (2) The lone pair of the sulfoxide moiety coordinates with the

copper centre; (3) The imine nitrogen is the other coordination site in this hypothesized

bidentate system.

Upon disruption of an octahedral Cu(II) centre, the impact of the Jahn-Teller

effect is seen.58,59 This results in four possible strong coordination sites arranged in a

square-planar fashion around the copper centre, and an additional two possible weak

coordination sites above and below the plane (1-41, Figure 6).55,58,59 A communication

from Evans et al., which investigated a Cu(II)-bis(oxazoline)-catalyzed Henry reaction,

proposed that the most reactive transition state would place nucleophilic nitromethyl

group in a weak coordination site perpendicular to the plane while the electrophile would

be located in the strong coordination site along the plane.55

Figure 6. Copper (II) coordination sites

Thus, adopting a similar model with these amino sulfoxide ligands would first

place the S and N coordinating groups in a cis-orientation along the plane. Coordination

of the nucleophile and electrophile, in the reactive manner proposed by Evans et al.,

would result in chelate 1-42 (Scheme 50). The orientation of the electrophile in chelate

CuX X

X X

Y

Y

X = Strong coordination siteY = Weak coodination site

1-41

Page 79: Preparation and Activity of Multidentate beta-Amino

61

1-42 would result in the slightly favoured (R)-isomer. However, equilibrium between

chelate 1-42 and 1-43 can be envisioned, where the electrophile conformation is

reversed, since the -OAc group would provide a minimal steric hindrance. Chelate 1-43

would thus result in the (S)-isomer, and this equilibrium also serves to explain the low

enantioselecitivites obtained.

Scheme 50. Proposed model for observed stereoinduction and preference for the (R)-isomer.

2.11 Future work

Although the chemistry towards the access of these chiral sulfoxide ligands has

been established, accessing the pyridyl system in a diastereopure form requires further

investigation. Jiang et al. reported that the N-Ts moiety led to facile isolation of

diastereomers via flash chromatography, while this strategy did not yield separation for

the synthesized ligands, it gives insight as to how additional auxiliary groups could

CuS O

N OAcBn

H

ON

O

H

NO

NO2

CuS O

N OAc

XBn

H

ON

O

H

NONO2

H

NO2

HOO2N

H

NO2

HOO2N

HetAr

(R)

(S)

1-42 1-43

Page 80: Preparation and Activity of Multidentate beta-Amino

62

potentially bring about an efficient separation. For example, condensing the free base of

these ligands with a chiral aldehyde could introduce another chiral centre, which could

affect diastereomeric separation via achiral chromatography or recrystallization

techniques.

The 8-quinolyl sulfenate anion brought about an interesting outcome upon its

reaction with the chiral amino iodide electrophile. It is believed that one lone

diastereomer was produced based on the model that Schwan and Söderman had

reported.42,43 However, the chirality assignments for the sequent amino sulfoxide ligand

are only based on that analogy and an X-ray crystal structure determination would be

needed to determine its true stereochemical configuration.

The synthesized ligands were only screened for the Henry reaction. Although

they did not provide promising results, their effectiveness towards other reactions

remains to be established. Amino sulfoxide ligands have been probed against a variety

of reactions, from hydrogenations to cyclopropanations, and have been met with

success. Some of these examples serve to guide which reactions to probe next.15,32

The lack of stereochemical induction brought about by these ligands could arise from

their lack of rigidity. Substituting the phenylalanine carbon skeleton with that of a proline

would result in a rigid pentacyclic structure within the ligand 1-41 (Scheme 51a).

Moreover, the introduction of steric groups at the alpha position could help form a more

rigid ligand (1-42). A similar structural motif has been reported by Xiao et al. and those

reported ligands have seen success in the Henry reaction and Tsuji-Trost reaction.21,38

This motif can be accessed in a similar way using aziridine substrates. This chemistry

Page 81: Preparation and Activity of Multidentate beta-Amino

63

with a thiolate nucleophile has already been established, however the sulfenate

pathway must still be investigated for its viability (Scheme 51b).38

The free-base form of these amino sulfoxide ligands offers an additional site for

derivation, and potential to include additional components to match the requirements to

a reaction. For example, many ligands utilized in the Tsuji-Trost reaction and

asymmetric Michael additions include a phosphine moiety. This could easily be

introduced via diphenylphosphinobenzaldehyde (1-43) condensation with a ligand in its

free base form to access 1-44 (Scheme 51c).

Scheme 51. Future work towards ligand diversification.

HetAr S OMI N

Boc

*+ HetAr S*

O

NBoc

*

HetAr S OMN

R R'+ HetAr S*

O

R

R

NPG

HetAr S*O

NH2

Bn HetAr S*O

N

Bn

H

P(Ph)2O

H

P(Ph)2

+- H2O

a.

b

c

* * * *

PG

1-41

1-42

1-43 1-44

Page 82: Preparation and Activity of Multidentate beta-Amino

64

2.12 Conclusion

The desired amino sulfoxide targets were accessed through the complementary

traditional oxidative means and sulfenate anion chemistry. Moreover, the ligand design

was diversified via the introduction of an addition heteroaryl aldehyde to access their

imino-derivative. Despite the failure to access diastereopure configurations of each

ligand, the 8-quinolyl sulfenate anion fortunately generated only one diastereomer and

allowed for the access of more diastereopure derivatives.

The sulfenate anion itself provides an opportunity to generate optically enriched

sulfoxides via chiral Lewis acid induction. Although the attempts outlined here did not

yield any promising results, the literature is continually providing more examples of the

chiral sulfenate anion complexes.49-52 Promising chiral inductions have been slow to

develop, however a recent communication from Tan et al. reported enantioselectivities

as high as 94% with the use of chiral halogenated pentanidium salt.50 Despite the poor

results, the literature supports that this is a viable concept, and further screening will

hopefully lead to a more favourable outcome.

The synthesized chiral sulfoxide ligands unfortunately do not appear to have

much ability to influence the stereochemical outcome in the Henry reaction. Although

only one copper salt was screened, the literature suggests that other options are not as

viable. However, a thorough solvent screening with nitromethane and nitroethane

nucleophiles indicated that a slight improvement in enantioselectivity can be achieved

by using nitromethane as a solvent.

Page 83: Preparation and Activity of Multidentate beta-Amino

65

Overall, only a small segment of sulfenate anion reactivity and the capabilities of

the sulfoxide moiety in chiral induction have been probed. However these findings can

help influence and direct future research in these rapidly growing areas of interest.

Page 84: Preparation and Activity of Multidentate beta-Amino

lxvi

ABSTRACT

EXPEDITED CATALYTIC SYNTHESIS OF BUPROPION

Erwin Javier Remigio Advisor: University of Guelph, 2017 Professor A. L. Schwan

α-Amino carbonyl compounds represent a diverse group of synthetically valuable

compounds ranging from synthetically valuable compounds, ranging from synthetic

building blocks to biologically significant natural products. Moreover, the presence of

this motif in pharmaceutical compounds, such as Bupropion, make this a desirable

synthetic target. A common pathway to access this family of compounds involve an

enol-type nucleophile and an electrophilic nitrogen species (e.g. azodicarboxylates and

nitroso compounds). This pathway has been extensively investigated, however it

presents an inherent drawback, which requires additional post-functionalization in order

to access the desired α-amino target. Recently, the literature has provided examples of

direct C(sp3)-N bond forming reaction via catalytic oxidative means. Specifically, the

catalytic Cu(II)-mediated and I2 radical oxidative coupling methodologies, reported by

MacMillan et al. and Guo et al., respectively, have shown promise towards the facile

acquisition of α-amino aryl ketones. Here, these methodologies were probed for their

utility towards Bupropion synthesis, via the coupling of tbutyl amine and 3’-

Page 85: Preparation and Activity of Multidentate beta-Amino

lxvii

chloropropiopheone. The Cu(II)-mediated pathway did not seem to be compatible with

primary amines, and was not able to be generate Bupropion. In contrast, the I2-

mediated pathway was able to yield the desired target. In this system, NH4I in catalytic

amounts and sodium percarbonate in ethyl acetate, led to generation of Bupropion. An

extensive optimization led us to conclude that a reloading the reaction system with

additional sodium percarbonate and amine led to a complete reaction. This reaction was

probed with various iodine and acidic components (e.g. I2/AcOH, and I2/Phenol) in place

of NH4I to successfully yield Bupropion. As such, an alternative ionic mechanism was

proposed in order to account for unfavourable bond dissociation enthalpies required for

the originally proposed radical coupling. Unfortunately, isolation of Bupropion via flash-

column chromatography proved to be problematic. Therefore efforts were put forth

towards reaction optimization to force the complete consumption of the starting ketone

substrate, and its direct subsequent isolation as an HCl salt. Acquisition of HCl salt

proved to be successful, however it was accompanied by the unavoidable formation of

the corresponding tbutyl ammonium salt.

Page 86: Preparation and Activity of Multidentate beta-Amino

68

Chapter 3: Introduction

Page 87: Preparation and Activity of Multidentate beta-Amino

69

3.0 Introduction

3.1 Traditional synthetic pathway towards α-amino carbonyl compounds

Carbonyl compounds possessing an α-amine functionality are a diverse group of

synthetically valuable compounds. They represent significant synthetic building blocks

to biologically significant targets due to their capacity for additional chemistry.60

Moreover, this structural motif itself is present in many natural products and

pharmaceutical compounds.61 Specifically, the α-amino ketone motif is found in many

commonly prescribed drugs, making it an attractive synthetic target.

α-Amino ketones are commonly accessed through the reaction of a nucleophilic

enol, or related derivative with an electrophilic source of nitrogen, which typically

contains electron-withdrawing groups to achieve this reactivity (e.g. azodicarboxylates

and nitroso-compounds) (Scheme 52).60,61 This pathway offers opportunities for

stereoinductive interventions via chiral Lewis acids, whether the chemistry occurs

through a chiral metal enolate, the activation of the nitrogenated electrophile, and in

some instances, both. This reactivity offers much utility due to the need for enantiopure

compounds in drug synthesis.

A communication from Yamamoto et al. reported successful enantioselective α-

amination with tin enolates and a silver (I)/BINAP system complexed to

nitrosobenzene62. However, this electrophile can present chemoselectivity challenges

due to potential lone pair coordination of either the nitrogen or oxygen atoms. However,

it was reported that a 2:1 silver (I)/binap complex led exclusively attack at the nitrogen

Page 88: Preparation and Activity of Multidentate beta-Amino

70

centre, leading to the α-amino adduct. This catalytic system offered effective α-

aminations from cyclic tin enolates (Scheme 53).62

Scheme 52. Common synthetic pathway towards α-amino carbonyl compounds (X = C(O)OR, Y = N or O, Z = lone pair of electrons if Y = O, or Z = C(O)OR if Y = N).

Scheme 53. Enantioselective α-amination via tin enolate and silver (I)/BINAP system.

OHN Y

X

Z

ON Y

XZ+

N NRO(O)C

C(O)ORAr

N O

OHN Y

X

ZO ML*

N YX

Z

ML*

Via Lewis acid activation:

OSnBu3

Ph NO

+O

NOH

Ph

10 mol%(R)-binap•(AgOTf)2

EtOCH2CH2OEt-78 ˚C

O O

N NO

NOH

Ph

OH

Ph Ph

OH

95%; >99% ee 97%; 98% ee 96%; 97% ee

Page 89: Preparation and Activity of Multidentate beta-Amino

71

In another example, Evans and Nelson utilized a chiral magnesium

bis(sulfonamide) enolate complex 2-1 to facilitate asymmetric α-amination with di-tert-

butyl azodicarboxylate.63 Their method required the addition of the carbamate moiety to

their initial carbonyl substrate to act as an additional coordinating group to facilitate the

stereoselective nucleophilic attack.

The authors hypothesized a reactive intermediate, where the spirocyclic

magnesium centre allowed for the sulfonyl p-xylene to block the top-showing face of the

assumed E-enolate, permitting only addition from the rear-showing Si-face. Although the

reaction was sluggish and took upwards of three days in some cases, this catalytic

system afforded excellent yields and enantioselectivities (Scheme 54).63

Scheme 54. Asymmetric α-amination via chiral magnesium bis(sulfonamide) enolate complex.

Mg

N

NSO2

H

HO2S

O

O

ON

ArH

Prevents exposure fromincoming electrophiles

Proposed reactive intermediate:

NAr

O

O

O ML* (10 mol%)pTsNHMe (20 mol%)N

NBoc

Boc+

DCM-75 ˚C - -65 ˚C

N ArO

O

O

BocNNHBoc

N N

Ph Ph

O2S SO2Mg

ML* =

2-1

Ar:Ph = 92 %; ee = 86%4-F-C6H4 = 97%; ee = 90%4-OMe-C6H4 = 93%; ee = 86%

Page 90: Preparation and Activity of Multidentate beta-Amino

72

A report from MacMillan et al. utilized the general reactivity pattern shown in

Scheme 52 towards α-amino aldehyde, however it was achieved via a novel photoredox

mechanism (Scheme 55).64 The authors utilized a chiral electron-rich enamine for

stereoinduction. The radical amine coupling partner was generated from a carbamate

possessing an N-dinitrophenylsulfonate (ODNs) activation handle, chemoselectively

activated by a common household light bulb. The catalytic cycle begins with enamine

formation from the aldehyde and imidazolidinone catalyst. Upon its formation it is able to

couple with the electrophilic nitrogen radical (2-2) to form 2-3. Oxidation of 2-3 via a

photoexcited N-ODNs, N-methyl carbamate (2-4), results in a single electron transfer

(SET) to form the iminium ion (2-5), which is hydrolyzed to regenerate the

imidazolidinone catalyst and enantioenriched α-amino aldehyde. Moreover, this is a key

propagation step, which excites and subsequently delivers another carbamyl radical for

coupling (Scheme 55). Overall, they reported good to moderate yields, with outstanding

enantioselectivities.

Jørgensen et al. were the first to report such aminations using an enamine with a

chiral auxiliary (2-6) (Scheme 56).60 Here, L-proline was utilized to form the chiral

enamine, and its reaction with an azodicarboxylate resulted in impressive results. The

authors rationalized that the nitrogen lone pair and carboxylic acid proton of the proline-

enamine intermediate were able to direct the incoming azodicarboxylate electrophile. A

Zimmerman-Traxler-type transition state has been postulated and it rationalized the

formation of a stable chair conformation 2-7 to explain the observed stereoinductive

effects (Scheme 56).60

Page 91: Preparation and Activity of Multidentate beta-Amino

73

HR1

O

N C(O)OMe

ODNs

NH

NO

•HOTf (30 mol%)

2,6-lutidine(1:1) DMSO/CH3CN

26W CFL

HR1

N C(O)OMe

O

+

Hnhex

N C(O)OMe

O

HCH2-C6H4-4-OMe

N C(O)OMe

O

HiPr

N C(O)OMe

O

71%; ee = 90% 79%; ee = 91% 67%; ee = 94%

H

O

R1

N

NH

OEt

HR1

N

NEtO

HR1

N

NEtO

NR2(C(O)OR3

R2 N

R3

O

HR1

N

NEtO

NR2(CO2R3)

HR1

N

NEtO

NR2(C(O)OR3

HR1

NR2(C(O)OR3O

DNsON

R3

OR2

DNsON

R3

OR2

*DNsO

N

R3

OR2

R2 N

R3

O

SO3NO2

NO2

Imidazolidinone catalyst

hv

+

α−amino aldehyde 2-3

2-5

2-4

2-2

1

Catalytic Cycle:

Scheme 55. Photoredox catalytic cycle.

Page 92: Preparation and Activity of Multidentate beta-Amino

74

Scheme 56. Proline-catalyzed amination via enamine intermediate.

3.2 Direct C-N oxidative coupling reactions

The literature provides many effective approaches for carrying out the general

pattern of reactivity shown in Scheme 52. However this approach towards the α-

amination of carbonyl compounds all share a common disadvantage, which requires the

post-functionalization of the newly acquired amine-moiety. As such, efforts towards a

direct C(sp3)-N coupling have been undertaken, however the inherent nucleophilic

character of both the α-carbonyl centre and the amine present a challenge towards this

goal.65-67

MacMillan et al. provided a versatile one-pot strategy towards the direct coupling

of α-keto carbon centres and amines.66 Their approach utilized the alcohol precursor to

the respective ketone, N-bromosuccinimide (NBS), and open-air conditions to drive the

reaction forward (Scheme 57) A thorough investigation of their reactions using NMR

H Et

O

EtO(O)C NN C(O)OEt

+L-Proline (10 mol%)

DCM, rt H

ON

Me

HN

C(O)OEt

C(O)OEt

77%; ee = 90%

H

N C(O)OH EtO(O)C NN C(O)OEt

N N H

N

O O

EtO(O)CH

EtO(O)C

Proposed transition state:Via:

2-6 2-7

Page 93: Preparation and Activity of Multidentate beta-Amino

75

experiments led them to postulate that the alcohol initially reacts with NBS to form a

reactive hypobromite species 2-8. Its inherent reactivity allows for facile oxidation of this

species to release a bromide ion and form the ketone 2-9. In solution, the bromide is

capable of acquiring a proton to form HBr, which is further oxidized to elemental

bromine by the open-air conditions. The ketone is then subsequently α-brominated, and

the morpholine is able to directly attack the α-keto centre and afford 2-10 (Scheme 57).

Scheme 57. NBS-mediated α-amination of aryl ketones.

Ar

OH

O

HN

+Ar

O

N

O

NBS (1.3 equiv)

1,4-dioxane, rtopen-air

OHNBr

O

O

HN OO

O Br

H

O O

Br

O

N

O

O

HN

HBr

Br2

Br-

HBr +1/2O2 → 1/2 H2O + 1/2Br2Br-

2-8 2-9

2-10

Ar:Ph = 90%4-tBu-C6H4 = 70%4-Cl-C6H4 = 80%3-CF3-C6H4 = 70%2-thienyl = 40%

*

*

*

*

Page 94: Preparation and Activity of Multidentate beta-Amino

76

Another communication from MacMillan et al. further expanded the scope of

direct C(sp3)-amine coupling.65 Similarly, the α-brominated species 2-11 was generated

however it was accessed via a copper-bound enolate 2-12. The subsequent release of

HBr and 2 moles CuBr in the presence of air allowed for the regeneration of the Cu(II)

species. In addition to aryl ketones, this report successfully incorporated the morpholine

moiety onto α-aryl esters under the similar conditions (Scheme 58).

Scheme 58. Cu(II)-mediated oxidative α-amination of carbonyl compounds.

X RO

X RO

Br

X RO

N

O

HN

O

X BrO

R

CuCu

Br

2 CuBr

2 CuBr2

1/2 O2 + 2 HBr2 H2O

HBr

Catalytic cycle:

CuBr2 (10 mol%)

DMSO, 50 ˚Copen-air

X

OR

O

HN

+X

OR

N

O

2-11

2-12

X = 2-furyl, R = Me: 93%X = 4-CF3-C6H4, R = Me: 70%X = MeO, R = Ph: 81%X = MeO, R = 4-NHTs-C6H4: 70%X = MeO, R = 3-MeO-C6H4: 71%

*

**

Page 95: Preparation and Activity of Multidentate beta-Amino

77

Recently, Guo et al. achieved this direct C-N amine coupling via a one-pot

oxidative radical coupling pathway (Scheme 59).67 Their reaction system required the

presence of an iodide source, a weak acid (NH4I acts as both), oxidant, and weak base

(Na2CO2�1.5H2O2 acts as both). In contrast to the generation of an α-halogenated

species, these authors suggested a carbocation as the reactive electrophilic species.

The authors reported that the catalytic cycle begins with the oxidation of iodide anions to

elemental iodine. Iodine was reported to undergo homolytic cleavage, which continued

to abstract an α-hydrogen from the ketone substrate, and a subsequent single electron

transfer (SET) from the radical ketone species provides the reactive carbon electrophile

2-13 for nucleophilic attack. After the initial hydrogen abstraction, the HI by-product is

then transformed into the key elemental iodine species either by treatment with base

followed by oxidation of the anions or through simple oxidation (Scheme 59).

With these new and direct one-pot C-N bond-forming reactions, the synthesis of

small molecule α-aminated ketone compounds can be potentially revisited. Bupropion

((±)1-(3-chlorophenyl)-2-[(1,1-dimethylethyl)amino]-1-propanone hydrochloride), known

on the market as WellbutrinTM or ElontrilTM, is a commonly prescribed anti-depressant

drug. In addition, it has seen utility in smoking-cessation and obesity treatments. It

exerts its effects as a norepinephrine-dopamine reuptake inhibitor (NDRI, which is a

mechanism atypical compared to many antidepressants. Bupropion has a relatively

simple structure containing an α-tbutyl amino group. With these new methods in hand,

an expeditious protocol could potentially be constructed by simply coupling 3’-

chloropropiophenone and tbutyl amine (Scheme 60).

Page 96: Preparation and Activity of Multidentate beta-Amino

78

Scheme 59. Proposed I2 radical coupling mechanism

Scheme 60. Retrosynthetic approach of Bupropion via direct C-N coupling.

2 I

2 HIH2O2

I2

2 I-

H2O2

Base

R R'O

R R'O

R R'O

R R'O

N

O

HN

O

- H+

SET

R R'O R R'

O

N

OO

HN

NH4I (15 mol%)Na2CO3•1.5H2O2 (2 equiv)

MeCN, 50 ˚C+

Catalytic cycle:

[O]2-13

R = Ph, R' = Me: 92%R = 4-CF3-C6H4, R' = Me: 85%R = 4-MeO-C6H4, R' = Me: 80%R = 2-furyl, R' = Et: 66%R = 2-thienyl, R' = Et: 75%

*

*

Cl

O

NH • HCl

Bupropion HClCl

O

NH

Via direct C(sp3)-N coupling

Cl

O

+NH2

**

Page 97: Preparation and Activity of Multidentate beta-Amino

79

The synthesis of Bupropion is carried out as a two-step procedure. 3’-

chloropropiophenone is first subjected to an excess of liquid bromine in DCM to access

the α-brominated intermediate, which is prone to degradation and is also a lachrymator,

as most α-halogenated ketones are. This intermediate is then reacted with tbutyl amine

in DMF at 60 ˚C (Scheme 61) to access the desired product.68

Scheme 61. Traditional synthesis of Bupropion.

The work of MacMillan et al. and Guo et al. offer promising potential to expedite

the acquisition of Bupropion. The use of a one-pot system would reduce the amount of

solvent and reagents used in its synthesis, thereby resulting a more cost-effective,

efficient, and green synthesis. Given Bupropion’s prevalence of use, synthetic

improvements can potentially applied to its industrial preparation. These efforts were of

interest to our industrial client, Vibrant Pharma Inc. and as such this project was

undertaken with their partnership along with fellow laboratory technician, Rebecca

Sydor, who was under my supervision.

Cl

OBr2 (excess)

DCM, 0 ˚C to rt

Cl

O

Br

Must isolate- unstable- lachrymator

DMF,60 ˚C

NH2

Cl

O

NH

**

Page 98: Preparation and Activity of Multidentate beta-Amino

80

Chapter 4: Results and Discussion

Page 99: Preparation and Activity of Multidentate beta-Amino

81

4.0 Results and Discussion

In our hands, the synthetic method provided by MacMillan et al. was unable to afford

Bupropion (Scheme 62). The reaction was performed with 3’-chloropropiophenone and

tbutyl amine using their reported conditions. After a 36-hour reaction time, TLC analysis

displayed plenty of starting material. Although newer UV-active spots had formed, 1H

NMR of the crude reaction mixture only indicated the major presence starting material

along with indiscernible peaks in the aliphatic region.

Scheme 62. Failed synthesis of Bupropion via Cu(II)-mediated oxidative coupling.

The report from MacMillian et al. did not probe 3’-chloropropiophenone as a

substrate, so we examined this compound’s reactivity using with morpholine instead, as

they had originally reported. This substrate required a considerably longer reaction time,

and resulted in low yields of the expected α-amino aryl ketone 2-14 (Scheme 63)

(These sets of reactions were performed by Rebecca Sydor.)

This reaction with morpholine provided an interesting and initially unidentified by-

product 2-15 that the authors did not mention (Figure 6). First, an assumption was

made, that the original carbon skeleton was retained in this by-product. Based on 1H

Cl

O

Cl

O

NH

NH2

+

(3 equiv)

CuBr2 (10 mol%)

DMSO, rtopen-air

*

Page 100: Preparation and Activity of Multidentate beta-Amino

82

NMR analysis, this by-product seemed to contain two morpholine moieties, however

one of them seemed to be in a more polarized environment, due to the two 4H

multiplets, clearly separated by 0.650 ppm. Interestingly, the aromatic region was far

more condensed than usual and the characteristic peak pattern of a meta-disubstituted

phenyl ring was lost. Finally, a lone downfield singlet at 6.39 ppm was reasoned to

belong to an alkenyl proton. Unfortunately, efforts to maximize the yields of either this

by-product or the expected α-morpholine adduct were unsuccessful.

Scheme 63. Test reaction of direct C-N coupling of 3’-chloropropiophenone and morpholine via Cu(II)-mediated oxidative coupling.

Figure 7. 1H NMR of isolated by-product and corresponding predicted chemical structure.

Cl

O

Cl

O

N+

(3 equiv)

CuBr2 (10 mol%)

DMSO, rtopen-airO

HN

OCl

N

Br

H

NO

O

X

Possible by-product

2-14 2-15

*

Page 101: Preparation and Activity of Multidentate beta-Amino

83

The synthetic method towards α-amino aryl ketones provided by Guo et al. led to

more promising results. An initial trial reaction utilized their reported procedure and

provided trace amounts of the desired product formation as indicated by 1H NMR. This

trial was carried through in acetonitrile and formed a 1M solution with respect to the

starting material, however that volume was not sufficient in solvating all of the reactants

(Entry 1, Table 6, Scheme 64).

Scheme 64. Proposed synthesis of Bupropion.

In Figure 7 a sample 1H NMR spectrum of the reaction of Entry 3 (Table 6), which

contains both starting material and Bupropion is shown. The diagnostic signal at 4.33

ppm (q, J = 7.2 Hz, 1H) arises from Bupropion and indicates the α-proton where the

new C-N bond has been created. The signal at 2.99 ppm (q, J = 7.2 Hz, 2H) is indicative

of the starting material, 3’-chloropropiophenone, and indicates the protons from its α-

position. Initial attempts to isolate Bupropion via flash-column chromatography were

unsuccessful. These difficulties could be attributed to the acidic environment provided

by the silica gel, which could allow the newly added amino group to become protonated

allowing for its degradation. Due to the difficulty in isolating Bupropion as a free base,

NH4I (15 mol%)Na2CO3•1.5H2O2 (2 equiv)

MeCN, 50 ˚CCl

O

Cl

O

NH

NH2

+

(3 equiv)

*

Page 102: Preparation and Activity of Multidentate beta-Amino

84

crude reaction materials were characterized via 1H NMR and a ratio of Bupropion to

starting material was used to characterize effectiveness of screening trials. The

integration of the starting material signal at 2.99 ppm was halved to account for relative

contribution of protons compared to Bupropion. Moreover, efforts were placed towards

complete consumption of starting material in order to allow for facile isolation of

Bupropion as an HCl salt.

Figure 8. 1H NMR spectrum of the crude reaction mixture containing Bupropion and 3’-chloropropiophenone.

Increasing the volume of solvent used (0.5 M from 1.0 M wrt ketone) led to

improved product formation (e.g. Entry 3, Table 6). A variety of solvents were tested

during the screening procedures. A non-polar solvent (e.g. toluene) was not a suitable

environment for this reaction to proceed (Entry 10, Table 6). Likewise, polar-protic

alcoholic environments also led to no product formation (Entries 4-9, Table 6). A trial

reaction that utilized polar-aprotic solvents provided identical results (Entries 11-13,

Table 6). Surprisingly, ethyl acetate proved to be an effective solvent, despite its

Page 103: Preparation and Activity of Multidentate beta-Amino

85

potential reactivity as it bears potentially available hydrogens alpha to the carbonyl

moiety (Entries 15-18, Table 6). Furthermore, it provided better solvation even with less

solvent added (1.0 M from 0.5 M wrt ketone, Entry 18, Table 6). However when DMSO

was added as a co-solvent, product formation was inhibited (Entry 14-18, Table 6).

These solvent screening reactions summarized that a reaction temperature of 50 ˚C and

mildly polar aprotic solvent is required for this reaction (e.g. MeCN, EtOAc). A highly

polar aprotic solvent (e.g. DMF, DMSO) could potentially interfere with the reaction via

complexation of the iodine, which must be re-oxidized to drive the reaction forward.

Similarly polar-protic solvents could interfere with the iodine via H-bond interactions

preventing them from re-oxidation, while non-polar solvents do not provide enough

stabilization of these species.

Table 6. Solvent screening reaction trials. (Reaction entries expect #2 and #3 were performed by Rebecca Sydor.)

Entry Solvent Temp. (˚C)

Time (hours)

Notes Prod. Formation

Prod. : Starting Material

1 MeCN rt 48 --- No --- 2 MeCN 50 36 Less solvent Trace --- 3 MeCN 50 36 --- Yes 1.0 : 0.7 4 MeOH rt 48 -- No --- 5 EtOH rt 48 --- No --- 6 EtOH 50 28 --- No --- 7 IPA rt 36 --- No --- 8 tBuOH rt 48 --- No --- 9 tBuOH 50 48 --- No --- 10 Toluene rt 48 --- No --- 11 DMF 80 48 --- No --- 12 DMF 80 48 TBAI was used in

place of NH4I No ---

13 DMSO rt 48 --- No --- 14 EtOAc/D

MSO (4:1)

50 48 --- No ---

Page 104: Preparation and Activity of Multidentate beta-Amino

86

15 EtOAc rt 48 --- Yes 1.0 : 3.2 16 EtOAc 40 48 --- Yes 1.0 : 1.2 17 EtOAc 50 28 --- Yes 1.0 : 0.3 18 EtOAc 50 28 Less solvent

added Yes 1.0 : 0.2

The role of the acidic and iodinated components in this reaction material was also

investigated. Trial reactions, which replaced the NH4I catalyst with TBAI led to no

product formation, which suggested the need for an acidic component in this reaction

(Entry 12, Table 7). Reaction trials, which increased the concentration of iodine and

protons separately both led to significantly increased product formation (Table 7).

Although the presence of a strong acid led to product formation, relatively weaker acids

led to more desirable results (Entries 7 and 8, Table 7).

Table 7. Acid and iodine source screening reactions. (Reaction entries in this table were performed by Rebecca Sydor.)

Entry Solvent Temp.

(˚C) Time (hours)

Notes Prod. Formation

Prod. : Starting Material

1 EtOAc 50 28 Additional 15 mol% of I2 added after 12 hours

Yes 1.0 : 0.2

2 EtOAc 50 28 7.5 mol% NH4I catalyst

Yes 1.0 : 2.1

3 EtOAc 50 28 25 mol% NH4I catalyst

Yes 1.0 : 0.3

4 EtOAc 50 48 25 mol% NH4I catalyst, upscale (__ mmol)

Yes 1.0 : 0.2

5 EtOAc 50 28 TBAI was used in place of NH4I

No ---

6 EtOAc 50 48 4Å M.S. added to reaction mixture

Yes 1.0 : 0.6

7 EtOAc 50 48 Addition of 15 mol% AcOH

Yes 1.0 : 0.08

8 EtOAc 50 48 Addition of 15 mol% TFA

Yes 1.0 : 0.3

Page 105: Preparation and Activity of Multidentate beta-Amino

87

Guo et al. reported that the oxidant, sodium percarbonate, is required to generate

radical iodine species to drive this reaction forward. Different varieties of oxidant were

screened in an attempt to further optimize the reaction for our purposes.

Dichloroisocyanuric acid (DCCA) is an inexpensive oxidizer commonly used as a

swimming pool disinfectant and generates the oxidant hypochlorous acid (HOCl).69

Reaction trials with DCCA led to product formation, however it was not as effective in

consuming the starting material (Entry 1, Table 8). Next, an open-air environment was

probed in an attempt to use atmospheric oxygen as the oxidant to increase atom-

economy and to access a greener procedure however Entries 2-4 from Table 3

suggested that this was not a suitable oxidant.

Table 8. Oxidant screening reactions. (Reaction entries in this table were performed by Rebecca Sydor.)

Entry Solvent Temp.

(˚C) Time (hours)

Notes Prod. Formation

Prod. : Starting Material

1 EtOAc 50 28 DCCA was used in place of percarbonate

Yes 1.0 : 0.9

2 EtOAc 50 48 1 equiv. percarbonate, under air instead of argon, additional wash of sat. aq. Na2S2O3 during work-up

Yes 1.0 : 1.4

3 EtOAc 50 48 Under air instead of Ar(g)

No ---

4 EtOAc 50 48 No percarbonate, under air instead of Ar(g)

No ---

Page 106: Preparation and Activity of Multidentate beta-Amino

88

Crude reaction mixtures after aqueous work-up (aq. sat. NaHCO3, water, brine)

all displayed a dark-brown colour, characteristic of excess iodine. In an attempt to

reduce iodine to its corresponding anions, aqueous saturated thiosulfate washes were

incorporated into the work-up procedure. Although the resulting crude reaction materials

appeared to contain less iodine, the comparative increase of starting material in these

reactions suggested that aqueous thiosulfate washes led to degradation of Bupropion

(Table 9).

Table 9. Screening reactions with aqueous saturated thiosulfate wash during aqueous work-up. (Reaction entries in this table were performed by Rebecca Sydor.)

Entry Solvent Temp. (˚C)

Time (hours)

Notes Prod. Formation

Prod. : Starting Material

1 EtOAc 61 48 4 equiv. amine Yes 1.0 : 0.1 2 EtOAc 50 48 Na2SO4 added to

reaction mixture, additional wash of sat. aq. Na2S2O3 during work-up

Yes 1.0 : 1.0

3 EtOAc 50 48 4 equiv. amine, additional wash of sat. aq. Na2S2O3 during work-up

Yes 1.0 : 3.9

Only a few reaction conditions resulted in near complete consumption of starting

materials after 48 hours. In an attempt to expedite the rate of reaction, screening

reactions were carried out in a screw-top pressure vessel (Entries 1-4, Table 10).

Unfortunately, the standard conditions performed under increased pressure were unable

to yield any Bupropion. Additional catalyst and/or acid were required produce any

desirable results, however the consumption of starting material was only moderate

Page 107: Preparation and Activity of Multidentate beta-Amino

89

(Entries 3 and 4, Table 10). Comparable reactions, which took place at atmospheric

pressure led to increased product formation, which suggest that atmospheric pressures

are more favourable to this reaction.

Table 10. Screening reactions performed in screw-top glass pressure vessel and comparable control reactions. (All reaction entries expect #1 were performed by Rebecca

Sydor.)

Entry Solvent Temp.

(˚C) Time (hours)

Notes Prod. Formation

Prod. : Starting Material

1 EtOAc 50 48 Reaction done in pressure vessel

Yes ---

2 EtOAc 50 48 25 mol% NH4I catalyst, addition of 15 mol% AcOH, pressure vessel

Yes 1.0 : 3.5

3 EtOAc 50 48 25 mol% NaI used in place of NH4I, addition of 15 mol% AcOH, pressure vessel

Yes 1.0 : 3.5

4 EtOAc 50 48 15 mol% NaI used in place of NH4I, addition of 15 mol% AcOH, pressure vessel

Yes 1.0 : 1.3

5 EtOAc 50 48 25 mol% NaI used in place of NH4I, addition of 15 mol% AcOH

Yes 1.0 : 0.6

6 EtOAc 50 48 15 mol% NaI used in place of NH4I, addition of 15 mol% AcOH

Yes 1.0 : 0.3

These screening procedures have summarized that a weak acid, an iodine

source, and the percarbonate oxidant are required to carry this reaction forward. An

attempt to broaden compatible reagents for this reaction resulted in a new set of

conditions that led to promising results. 15 mol% each of elemental iodine and acetic

Page 108: Preparation and Activity of Multidentate beta-Amino

90

acid (Conditions #2) were used in place of NH4I and 1H NMR analysis of the crude

reaction mixture indicated product formation and complete consumption of the starting

material (Entry 1, Table 11). Similarly, 15 mol% each of elemental iodine and phenol

also led to product formation, although in lower yields of Bupropion (Conditions #3,

Entry 4, Table 11).

Table 11. Probing other suitable reagents for α-amination of 3’-chloropropiophenone. (All reaction entries expect 4 were performed by Rebecca Sydor.)

Entry Solvent Temp. (˚C)

Time (hours)

Notes Prod. Formation

Prod. : Starting Material

1 EtOAc 50 48 Conditions #2a Yes 1:0 2 EtOAc 50 48 Conditions #2,

0.10 equiv each of I2 and AcOH

Yes 1.0:0.07

3 EtOAc 50 48 Conditions #2, 0.075 equiv each of I2 and AcOH

Yes 1:0.11

4 EtOAc 50 48 Conditions #3b, Additional equiv. each of amine and percarbonate added after 48 hours (3 mmol scale)

Yes 1.0:1.7

a I2 (15 mol%), AcOH (15 mol%), sodium percarbonate (2 equiv), tbutyl amine (3 equiv), EtOAc b I2 (15 mol%), PhOH (15 mol%), sodium percarbonate (2 equiv), tbutyl amine (3 equiv), EtOAc

Guo et al. proposed a radical catalytic mechanism to account the α-amination of

their substrates. They proposed that I- was oxidized to I2 and underwent homolytic

cleavage to produce radical iodine species, which in turn can abstract a hydrogen atom

from the α-carbonyl position of an aryl ketone. However, bond dissociation enthalpies of

either I2 or HI (151 kJ/mol and 298 kJ/mol, respectively) neither match nor exceed the

Page 109: Preparation and Activity of Multidentate beta-Amino

91

energy required to abstract a hydrogen from the α-position of an aryl ketone

(PhC(O)CH2R = ~388 kJ/mol).70 To account for the discrepancies between the bond

dissociation enthalpies of these species, an ionic mechanism is proposed for this α-

amination reaction and accounts for the compatibility of other reagents for this reaction

summarized in Table 6. The presence of a weak acid helps facilitate formation of the

enol species 2-16, which in turn can form an α-iodinated species 2-17. Upon

nucleophilic attack from tbutyl amine, the desired product, Bupropion is formed. The

catalytic I2 is regenerated from oxidation of 2HI or 2I- (Scheme 65).

Scheme 65. Proposed ionic mechanism for the synthesis of Bupropion.

To this point, only a few reaction conditions had resulted in complete

consumption of starting material after 48 hours to yield Bupropion. In an attempt to

complete the reaction in this time frame extra equivalents of certain reagents were

added after a period of time to help facilitate the reaction. The addition of 3 equivalents

of amine over a span of 30 hours provided a low formation of the desired product,

indicating that at least a three-equivalent excess at the start is required to efficiently

I2

[O]2 I-

2HI

[O]

Base

Cl

O

Cl

OH

Cl

OIH+

+ HI

Cl

OI tBuNH2 + HI

Cl

O

NH

I2 Regenaration of I2

2-16 2-17

2-17

*

**

Page 110: Preparation and Activity of Multidentate beta-Amino

92

drive the reaction forward (Entry 1, Table 12). However, the addition of an additional

equivalent of amine after 24 hours led to the production of Bupropion and complete

consumption of starting material (Entry 4, Table 12). Similarly, an additional equivalent

of percarbonate after 24 hours also led to full consumption of starting material (Entry 6,

Table 12). Additional amounts of catalyst, via I2, AcOH, or NH4I, led to improvements

however were not as effective as the addition of more amine and oxidant (Entry 7 and 9,

Table 12).

With an optimized approach that allowed for complete consumption of starting

material, the scalability of this protocol was probed. Considering the scale of these

reactions previous improvements were employed as well as longer reaction times.

These reactions containing different sources of acid and iodine, and which were

reloaded with amine and percarbonate, resulted in product formation and complete

consumption of starting material (Scheme 66, Table 13).

Due to the difficulty of isolating Bupropion as a free-base via flash-

chromatography and to better serve the industrial client, efforts were placed towards

formation of its corresponding HCl salt. The first attempt combined HCl in IPA (1:1 v/v)

and this mixture was used to dissolve the crude product before removing the excess

under reduced pressure to isolate the salts. A second method dissolved the crude

product in toluene before delivering a saturated solution of HCl in IPA. This method

delivered the desired HCl salt, however this method was less reproducible due to the

difficulty in removing toluene. Both methods led to the desired HCl salt however they

were accompanied with an unwanted by-product. 1H NMR analysis of these HCl salts

Page 111: Preparation and Activity of Multidentate beta-Amino

93

depicted a resonance consistent with an α-tbutyl amino group, and no other protons

attached to a carbon. However, this resonance yielded an integration over 9 protons.

This excess was reasoned to arise from a t-butyl ammonium salt from the

disengagement of the amino group upon protonation. 1H NMR analysis of a separate

tbutyl ammonium salt sample confirmed this impurity. This by-product could have

formed during the acidification of the crude reaction material and as such, efforts to

keep this mixture cold were made. However, this was not sufficient in preventing

formation of this impurity. It should also be noted, that it was ensured that all excess

amine was evaporated from reaction mixtures, before the generation of the

corresponding HCl salt

Page 112: Preparation and Activity of Multidentate beta-Amino

94

Table 12. Reaction screening trials with extra equivalents of reagents added. (Entries marked with asterisk were performed by Rebecca Sydor.)

Entry Solvent Temp.

(˚C) Time (hours)

Notes Prod. Formation

Prod. : Starting Material

1 MeCN 50 30 Amine added in 3 portions over 30 hours

Yes 1.0 : 20

2* EtOAc 50 48 Additional 2 equiv. of amine added after 24 hours

Yes 1.0 : 0.1

3* EtOAc 50 48 Additional 2 equiv. of amine added after 24 hours, upscale (25 mmol)

Yes 1.0 : 0.3

4* EtOAc 50 48 Additional equiv. of amine added after 24 hours

Yes 1 : 0

5 EtOAc 50 72 Additional equiv. of amine added after 24 hours, upscale (36 mmol)

Yes 1.0 : 0.3

6* EtOAc 50 48 Additional equiv. of percarbonate added after 28 hours

Yes 1 : 0

7* EtOAc 50 48 Additional 10 mol% NH4I added after 28 hours

Yes 1.0 : 0.5

8 EtOAc 50 48 Additional 10 mol% AcOH added after 28 hours

Yes 1.0 : 0.4

9 EtOAc 50 48 Additional 10 mol% I2 added after 28 hours

Yes 1.0 : 0.3

10 EtOAc 50 72 Additional equiv. each of amine and percarbonate added after 48

Yes 1 : 0

Page 113: Preparation and Activity of Multidentate beta-Amino

95

hours, upscale (75 mmol)

Scheme 66. Summary of synthetic efforts towards Bupropion.

Table 13. Optimized reaction conditions tested with larger scale reactions.

Entry Solvent Temp. (˚C)

Time (hours)

Notes Prod. Formation

Prod. : Starting Material

1 EtOAc 50 96 Additional equiv. each of amine and percarbonate added after 48 hours, upscale (69 mmol scale)

Yes 1 : 0

2 EtOAc 50 96 Conditions #2, Additional equiv. each of amine and percarbonate added after 48 hours, upscale (68 mmol scale)

Yes 1 : 0

3 EtOAc 50 72 Conditions #3, Additional equiv. each of amine and percarbonate added after 48 hours, upscale (18 mmol scale)

Yes 1 : 0

Conditions #1NH4I (15 mol%)

Na2CO3•1.5H2O2 (2 equiv)

EtOAc, 50 ˚C

Cl

O

Cl

O

NH

NH2

+

(3 equiv)

Conditions #2I2 (15 mol%), AcOH (15 mol%)

Na2CO3•1.5H2O2 (2 equiv)

Conditions #3I2 (15 mol%), PhOH (15 mol%)

Na2CO3•1.5H2O2 (2 equiv)

Page 114: Preparation and Activity of Multidentate beta-Amino

96

4.1 Conclusion

Overall, the synthetic pathway provided by Guo et al. led to a set of conditions

that allowed for the access of Bupropion (Scheme 15). Moreover, modifications led to

the use of less expensive reagents (e.g. I2, AcOH, phenol) and an industrially viable

solvent choice with EtOAc. Despite these improvements, difficulties arose in the

isolation of the Bupropion, both as its free-base and its HCl salt. Nevertheless, the

chemistry was passed onto Vibrant Pharma Inc. for further modification and scale-up.

Page 115: Preparation and Activity of Multidentate beta-Amino

97

Chapter 5: Experimental

Page 116: Preparation and Activity of Multidentate beta-Amino

98

5.0 Experimental

5.1 General experimental

All reactions were carried out in flame-dried glassware under argon unless

specified otherwise. Thin layer chromatography (TLC) was performed on glass-backed

plates coated with Silica Gel 60, which contained a fluorescent indicator. The plates

were visualized under UV light. Flash column chromatography was performed with silica

gel particle size 30 – 60 (mesh 230 – 400) supplied by Silicycle ®.

Infrared (IR) spectra were obtained on a FT-IR spectrometer as a neat film. 1H NMR and

13C NMR Spectra were recorded on a Bruker Avance 300 (300 MHz 1H, 75 MHz 13C), a

Bruker Avance 400 (400 MHz 1H, 100.6 MHz 13C), or a Bruker Avance 600 (600 MHz

1H, 150.9 MHz 13C). Chemical shifts (ppm) and coupling constants (J, Hz) were

obtained from first order analysis of one-dimensional spectra. The proton spectra are

reported as follows δ (multiplicity, coupling constant J, number of protons). 1H NMR data

are reported using standard abbreviations: singlet (s), doublet (d), triplet (t), doublet of

doublet (dd), quartet (q), and multiplet (m). 1H NMR and 13C NMR chemical shifts are

referenced to CHCl3. Analytical thin-layer chromatography (TLC) was performed using

0.25 mm, extra-hard layer layer, 60 Å F254 glass-backed silica gel plates and were

visualized under UV light (254 nm). Pressure vessel reactions were carried out in

heavy-walled cylindrical vessels with an internal thread and a 15 mm Teflon ® bushing

as a pressure seal. HPLC experiments were performed using a Chiralcel OD-H (0.46

cm x 25 cm) column with iPrOH/hexanes as the eluent. Elemental analyses were

Page 117: Preparation and Activity of Multidentate beta-Amino

99

performed by MWH Laboratories of Pheonix, AZ. High-resolution mass spectrometry

was performed by Queen’s University Mass Spectrometry Facility in Kingston, ON.

5.2.0 Amino sulfoxide ligands and related compounds

5.2.1 General procedure for the preparation of aryl thiol from the corresponding

sulfonyl chloride

In a three-neck round bottom flask equipped with a drying tube, argon gas inlet,

and reflux condenser, 8-quinolinesulfonyl chloride (1.0 equiv.) was stirred in anhydrous

toluene (0.1 M, wrt sulfonyl chloride). Ph3P (3.0 equiv.) was added in five portions over

10 minutes. After complete addition of Ph3P, the reaction mixture stirred for 3 hours.

Water was added to the reaction mixture until all solids dissolved, and the mixture

stirred for an additional 30 minutes. The aqueous layer was discarded. The organic

layer was extracted four times with a 10% NaOH aqueous solution. The aqueous layers

were combined and toluene was added and the two-phase mixture was stirred. Water

was added to the mixture to dissolve remaining solids, and the toluene was discarded.

The aqueous layer was then acidified with a 1 M HCl solution until pH = 5 was reached.

The acidified aqueous layer was extracted with CH2Cl2. The organic layers were

combined and then washed with water, and brine. It was then dried over anhydrous

MgSO4, and concentrated under reduced pressure to afford pure product.

Page 118: Preparation and Activity of Multidentate beta-Amino

100

8-Mercaptoquinoline (1-27)47

Obtained as a viscous purple liquid, 76% yield; 1H NMR (400 MHz,

CDCl3), δ: 8.94 (dd, J = 4.4 Hz, 2.0 Hz, 1H), 8.16 (dd, J = 8.0 Hz, 1.6 Hz,

1H), 7.72 (dd, J = 7.2 Hz, 1.2 Hz, 1H), 7.59 (dd, J = 8.2, 1.2 Hz, 1H), 7.46 (dd, J = 8.4

Hz, 4.4 Hz, 1H), 7.42 (app, t, J = 2.8 Hz, 1H), 5.65 (s, 1H); 13C NMR (100 MHz, CDCl3),

δ: 149.50, 143.78, 136.97, 134.80, 128.84, 127.16, 126.63, 124.61, 121.78.

5.2.2 General procedure for the preparation of N-Boc Protected L-Phenylalanine

derived amino alcohol.

N-Boc-L-Phenylalanine was stirred in anhydrous THF (0.1 M wrt amino acid) and

cooled to 0 ˚C. Ethyl chloroformate (1.2 equiv.) was added to the reaction mixture,

followed by dropwise addition of Et3N. The reaction mixture was stirred for 2 hours or

until disappearance of starting material (confirmed via TLC). The solids were filtered off

and the filtrate was added dropwise to NaBH4 (1.5 equiv.) in water (0.25 M wrt hydride)

at 0 ˚C. The mixture was stirred overnight while warming to room temperature. The

reaction was quenched with a 1 M HCl solution until pH = 6 was reached. The acidified

aqueous solution was extracted five times with EtOAc, and the combined organic layers

were washed three times with water, twice with saturated aq. NaHCO3, and once with

brine. It was then dried over anhydrous MgSO4, concentrated under reduced pressure,

and the residue recrystallized with hexanes and EtOAc to afford pure product.

NSH

Page 119: Preparation and Activity of Multidentate beta-Amino

101

N-Boc-(R)-phenylalaninol71

Obtained as a white solid, yield: 81%; mp = 98 - 100 ˚C (lit mp =

96 – 97 ˚C).71 1H NMR (400 MHz, CDCl3), δ: 7.30 – 7.20 (m, 5H),

4.77 (br. d, J = 7.6 Hz, 1H), 3.87 (br. s, 1H), 3.69 (dd, J = 10.8 Hz, 3.6 Hz, 1H), 3.57 (dd,

J = 10.8 Hz, 5.2 Hz, 1H), 2.85 (d, J = 7.2, 2H), 1.32 (s, 9H); 13C NMR (100 MHz, CDCl3),

δ: 156.17, 137.77, 129.28, 128.57, 128.10, 126.55, 79.74, 64.45, 53.72, 37.43, 28.33;

IR (cm-1, neat): 3357, 2979, 2932, 1687, 1528, 1367, 1251, 1169.

5.2.3 General procedure for the preparation of N-Boc Protected L-phenylalanine

derived amino iodide.

Ph3P (1.0 equiv.), imidazole (1.0 equiv.) and I2 (1.1 equiv.) were cooled to -78 ˚C.

Anhydrous CH2Cl2 (0.1 M wrt Ph3P) was added, and after addition the mixture was

stirred and the temperature was raised slowly to -30 ºC. N-Boc protected amino alcohol

(1.0 equiv) in anhydrous CH2Cl2 (0.1 M wrt to amino alcohol) was added dropwise. The

reaction mixture was stirred and while warming to room temperature. Upon

disappearance of starting material (confirmed via TLC), the reaction mixture was filtered

to remove the phosphine oxide and the filtrate was concentrated under reduced

pressure without heat. The crude mixture was run through a short silica column (20 g

silica/1g crude) to afford pure product.

HO NH

Bn

O

O

Page 120: Preparation and Activity of Multidentate beta-Amino

102

(2R)-N-Boc-1-Phenyl-3-iodopropan-2-amine71

Obtained as a white solid, 72% yield; mp = 120 – 123 ˚C (lit mp =

121 – 122 ˚C).71 1H NMR (300 MHz, CDCl3), δ: 7.34 – 7.22 (m,

5H), 4.69 (br. s, 1H), 3.61 (br. s, 1H), 3.38 (dd, J = 10.1 Hz, 4.3 Hz, 1H), 3.18 (dd, J =

10.2 Hz, 3.8 Hz, 1H), 2.92 (dd, J = 13.5 Hz, 5.7 Hz, 1H), 2.78 (dd, J = 13.6 Hz, 8.0 Hz,

1H), 1.43 (s, 9H); 13C NMR (75 MHz, CDCl3), δ: 154.74, 136.96, 129.34, 128.60,

126.76, 79.76, 50.94, 40.54, 28.25, 13.84; IR (cm-1, neat): 3328, 2976, 1666, 1536,

1445, 1436, 1365, 1274, 1171.

5.2.4 General procedure for the preparation of N-Boc protected heteroaryl amino

sulfanes (1-28)

Heteroaryl thiols (1.0 equiv) and N-Boc protected amino iodide (2.0 equiv) were

loaded in a round bottom flask. The flask was evacuated and then purged with inert

argon gas three times, and an argon filled balloon was utilized to keep a constant

pressure of argon. Anhydrous CH2Cl2 (10 mL/mmol thiol) was added and the solution

was cooled to 0 ˚C, which was followed by dropwise addition of N,N-

diisopropylethlyamine (2.0 equiv). The reaction was stirred overnight, while the

temperature slowly rose to room temperature. The reaction mixture was washed three

I NH

Bn

O

O

Page 121: Preparation and Activity of Multidentate beta-Amino

103

times with water, and once with brine. It was dried over anhydrous MgSO4,

concentrated under reduced pressure, and purified by flash column chromatography.

N-Boc-1-Phenyl-3-(2-pyridylsulfanyl)propan-2-amine

Obtained as a white solid, 93% yield; mp = 80 – 81 ˚C. 1H

NMR (400 MHz, CDCl3), δ: 8.39 (d, J = 4.4 Hz, 1H), 7.50

(td, J = 7.6 Hz, 1.6 Hz, 1H), 7.31 – 7.28 (m, 1H), 7.25 – 7.20 (m, 5H), 7.01 – 6.98 (m,

1H), 5.87 (br. d, J = 5.6 Hz, 1H), 4.07 (s, 1H), 3.32 (dd, J = 14.4 Hz, 4.4 Hz, 1H), 3.25

(dd, J = 14.0 Hz, 7.2 Hz, 1H), 3.12 (dd, J = 13.6, 4.8 Hz, 1H), 2.87 (dd, J = 13.6 Hz, 8.4

Hz, 1H), 1.38 (s, 9H); 13C NMR (100 MHz, CDCl3), δ: 158.71, 155.60, 149.10, 137.96,

136.07, 129.46, 128.41, 126.36, 122.47, 119.67, 78.85, 52.84, 40.26, 33.41, 28.32; IR

(cm-1, neat): 3426, 2975, 1686, 1648, 1557, 1162. TOF MS EI calculated for

[C19H24N2O2S]+: 344.1559; found: 344.1556.

N-Boc-1-Phenyl-3-(8-quinolylsulfanyl)propan-2-amine

Obtained as a white solid, 96% yield; mp = 89 – 90 ˚C. 1H

NMR (300 MHz, CDCl3), δ: 8.96 (dd, J = 4.2 Hz, 1.7 Hz,

1H), 8.21 (dd, J = 8.3 Hz, 1.7 Hz, 1H), 7.58 (d, J = 7.9 Hz,

1H), 7.53 (d, J = 6.8 Hz, 1H), 7.43 (m, 1H), 7.41 (m, 1H), 7.28 – 7.20 (m, 5H), 5.38 (d, J

= 7.2 Hz, 1H), 4.22 (d, J = 5.3 Hz, 1H), 3.20 – 3.15 (m, 2H), 3.12 – 3.08 (m, 1H), 3.04

N

SBn

NH

O

O

N

SBn

NH

O

O

Page 122: Preparation and Activity of Multidentate beta-Amino

104

(dd, J = 13.6 Hz, 7.6 Hz, 1H), 1.40 (s, 9H); 13C NMR (75 MHz, CDCl3), δ: 155.16,

149.28, 145.57, 137.53, 137.05, 136.47, 129.34, 128.32, 128.29, 126.57, 126.40,

125.81,124.66, 121.50, 79,16, 50.90, 39.38, 34.82, 28.24; IR (cm-1, neat): 3326, 2988,

1679, 1642, 1532, 1199.

5.2.5 General procedure for the preparation of N-Boc protected heteroaryl amino

sulfoxides (1-29)

5.2.5.1 mCPBA oxidation

N-Boc protected heteroaryl amino sulfanes (1.0 equiv) in anhydrous CH2Cl2 (0.5

M, wrt starting material) were stirred and cooled to -78 ˚C. To this, a solution of mCPBA

(1.0 equiv, 72% active) in anhydrous CH2Cl2 (0.05 M, wrt mCPBA) was added via

canula. The reaction mixture was stirred for an additional 3 hours at -78 ˚C, before

raising the temperature to -35 ˚C and stirring overnight. The reaction mixture was

concentrated under reduced pressure to reduce its volume by half. The reaction mixture

was washed twice with 4:1 solution of saturated aq. Na2S2O3 and NaHCO3, twice with a

saturated aq. NaHCO3, twice with water, and once brine. It was then dried over

anhydrous MgSO4, concentrated under reduced pressure, and purified by flash column

chromatography.

Page 123: Preparation and Activity of Multidentate beta-Amino

105

5.2.5.2 NaIO4 oxidation

N-Boc protected amino sulfanes (1.0 equiv) in MeOH/water solution (1:1 v/v) was

stirred, and CH2Cl2 was added dropwise until starting material had fully dissolved. To

this, NaIO4 (1.0 equiv) was added slowly. The reaction was stirred and its progress was

tracked via TLC. The reaction mixture was filtered through a Büchner funnel and the

filtrate was extracted three times with CH2Cl2. The combined organic layers were

washed three times with water, and once with brine. It was then dried over anhydrous

MgSO4, concentrated under reduced pressure, and purified by flash column

chromatography.

5.2.5.3 H2O2 oxidation

N-Boc protected heteroaryl amino sulfanes (1.0 equiv) in HFIP (1.0 M, wrt

starting material) at room temperature. To this, 30% aqueous H2O2 (2.0 equiv) was

added and the reaction’s progress was tracked via TLC. Excess H2O2 was quenched

with saturated aq. Na2S2O3. The reaction mixture was extracted three times with EtOAc,

and the combined organic layers were dried over anhydrous MgSO4, concentrated

under reduced pressure, and purified by flash column chromatography.

Page 124: Preparation and Activity of Multidentate beta-Amino

106

N-Boc-1-Phenyl-3-(2-pyridylsulfinyl)propan-2-amine71

Obtained as a diastereomeric mixture. White solid. Total

yield: 45 – 86 %; mp = 90 – 94 ˚C. Major isomer: 1H NMR

(400 MHz, CDCl3), δ: 8.61 (d, J = 4.6 Hz, 1H), 7.99 (s, 1H),

7.96 – 7.90 (m, 1H), 7.39 – 7.35 (m, 1H), 7.25 – 7.19 (m, 5H), 4.90 (d, J = 6.1 Hz, 1H),

4.39 (m, 1H), 3.43 (dd, J = 13.5 Hz, 4.6 Hz, 1H), 3.12 (dd, J = 13.7, 7.6, 1H), 3.03 –

2.98 (m, 2H), 1.38 (s, 9H); 13C NMR (100 MHz, CDCl3), δ: 164.53, 155.01, 149.57,

137.99, 136.95, 129.47, 128.45, 126.62, 124.54, 119.98, 79.43, 57.86, 48.07, 40.31,

28.23. Minor isomer: 1H NMR (400 MHz, CDCl3), δ: 8.59 (d, J = 4.7 Hz, 1H), 8.01 (s,

1H), 7.96 – 7.90 (m, 1H), 7.39 – 7.35 (m, 1H), 7.32 – 7.27 (m, 5H), 5.55 (d, J = 6.6 Hz,

1H), 3.32 (m, 1H), 3.19 (m, 1H), 2.99 – 2.94 (m, 2H); 13C NMR (100 MHz, CDCl3), δ:

164.45, 154.74, 149.57, 137.93, 136.95, 129.36, 128.51, 126.62, 124.45, 119.78, 79.43,

58.10, 49.09 40.31, 28.27. Both diastereomers: IR (cm-1, neat): 3466, 2987, 1685, 1632,

1560, 1100, 1042.

N-Boc-1-Phenyl-3-(8-quinolylsulfinyl)propan-2-amine

Obtained as a diastereomeric mixture. Off-white solid. Total

yield: 96%; mp = 120 – 123 ˚C. Major diastereomer: 1H

NMR (600 MHz, CDCl3), δ: 8.70 (s, 1H), 8.15 (m, 1H), 8.09 N

S*O Bn

NH

O

O

N

S*Bn

NH

O

O

O

Page 125: Preparation and Activity of Multidentate beta-Amino

107

(m, 1H), 7.80 (m, 1H), 7.59 (m, 1H), 7.34 (m, 1H), 7.17 – 7.05 (m, 5H), 5.39 (s, 1H),

4.36 (s, 1H), 3.56 (m, 1H), 3.13 (m, 1H), 2.91 (m, 1H), 2.82 (m, 1H), 1.25 (s, 9H); 13C

NMR (150 MHz, CDCl3), δ: 154.52, 149.51, 143.12, 141.11, 137.15, 135.96, 129.96,

129.84, 129.35, 129.18, 127.65, 126.21, 126.19, 121.71, 78.71, 58.41, 50.15, 39.49,

27.21. Minor diastereomer: 1H NMR (600 MHz, CDCl3), δ: 8.75 (s, 1H), 8.27 (d, J = 7.2

Hz, 1H), 8.23 (d, J = 8.4 Hz, 1H), 7.95 (d, J = 8.4 Hz, 1H), 7.76 (app. J = 7.8 Hz, 1H),

7.49 – 7.45 (m, 1H), 7.34 – 7.17 (m, 5H), 4.23 (br. s, 1H), 3.62 (br. s, 1H), 3.25 (br. s,

1H), 3.06 (br. s, 1H), 2.94 (br. s, 1H), 1.37 (s, 9H); 13C NMR (150 MHz, CDCl3), δ:

149.93, 143.96, 136.34, 130.30, 129.70, 129.07, 128.94, 128.32, 128.10, 127.32,

126.74, 126.53, 126.45, 122.08, 78.62, 57.45, 53.79, 41.56, 28.44. Both diastereomers:

IR (cm-1, neat): 3289, 2976, 2929, 1708, 1495, 1169, 1044, 1025.

5.2.6 General procedure for the preparation of heteroaryl amino sulfoxides (N-

Boc de-protection)

To a solution of chiral N-Boc protected heteroaryl amino sulfoxide stirring in

anhydrous CH2Cl2 (0.05 M wrt starting material) at 0˚C, trifluoroacetic acid (5 mL/mmol

starting material) was added dropwise. The resulting solution was stirred, and while the

temperature was slowly raised to room temperature. The excess reagent and solvent

were removed under reduced pressure. Hexane was added and swirled around in the

flask and was removed under reduced pressure, and this process was repeated twice

more. The resulting oil was re-dissolved in CH2Cl2 and washed three times with a

Page 126: Preparation and Activity of Multidentate beta-Amino

108

saturated aq. NaHCO3, twice with water, and once with brine. It was then dried over

anhydrous MgSO4, concentrated under reduced pressure, and purified by flash column

chromatography.

1-Phenyl-3-(2-pyridylsulfinyl)propan-2-amine (Ligand 1-L1m)

Obtained as a diastereomeric mixture. Viscous off-white oil. Total

yield: 81%. Major diastereomer: 1H NMR (600 MHz, CDCl3), δ:

8.61 (d, J = 4.8 Hz, 1H), 8.00 (d, J = 4.0 Hz, 1H), 7.95 (td, J = 7.6 Hz, 1.6 Hz, 1H), 7.40

(m, 1 H), 7.29 - 7.23b (m, 5H), 3.77 (m, 1H), 3.29 (dd, J = 13.6 Hz, 4.6, 1H), 3.02 (dd, J

= 13.2 Hz, 3.0 Hz, 1H), 2.97 (dd, J = 13.3 Hz, 7.9 Hz, 1H), 2.76 (dd, J = 13.4 Hz, 8.4 Hz,

1H), 1.73 (s, 2H); 13C NMR (150 MHz, CDCl3), δ: 165.01, 149.52, 138.07, 137.64,

129.36, 128.59, 126.67, 124.54, 119.61, 61.84, 49.52, 43.97. Minor diastereomer: 1H

NMR (600 MHz, CDCl3), δ: 8.55 (d, J = 4.8 Hz, 1H), 8.02 (d, J = 4.0 Hz, 1H), 7.94 (td, J

= 7.6 Hz, 1.6 Hz, 1H), 7.40 (m, 1 H), 7.34 (m, 3H), 7.11 (app d, J = 8.4 Hz, 2H), 3.61 (m,

1H), 3.15 (dd, J = 13.3 Hz, 9.9 Hz, 1H), 3.02 (dd, J = 13.2 Hz, 3.0 Hz, 1H), 2.83 (dd, J =

13.5 Hz, 5.8 Hz, 1H), 2.72 (dd, J = 13.6 Hz, 8.1 Hz, 1H), 1.73 (s, 2H). 13C NMR (150

MHz, CDCl3), δ: 164.67, 149.59, 137.90, 137.49, 129.19, 128.59, 126.63, 124.38,

119.90, 61.54, 58.24, 44.41. Both diastereomers: IR (cm-1, neat): 3374, 3042, 2903,

1584, 1499, 1028.

N

S*Bn

NH2

O

Page 127: Preparation and Activity of Multidentate beta-Amino

109

1-Phenyl-3-(8-quinolylsulfinyl)propan-2-amine (Ligand 1-L2m)

Obtained as a diastereomeric mixture. Off-white solid. Total yield:

81%; mp = 114 – 118 ˚C. Major diastereomer: 1H NMR (600 MHz,

CDCl3), δ: 8.84 (dd, J = 4.2 Hz, 1.8 Hz, 1H), 8.30 (dd, J = 7.2 Hz,

1.4 Hz, 1H), 8.25 (m, 1H), 7.95 (m, 1H), 7.76 (m, 1H), 7.50 (dd, J = 8.3 Hz, 4.3 Hz, 1H),

7.31 – 7.23 (m, 5H), 3.90 (m, 1H), 3.64 (dd, J = 13.4 Hz, 4.7 Hz, 1H), 3.07 (m, 1H), 2.89

(dd, J = 13.15 Hz, 7.6 Hz, 1H), 2.78 (dd, J = 13.4 Hz, 8.3 Hz, 1H); 13C NMR (100 MHz,

CDCl3), δ: 150.02, 143.71, 141.37, 137.78, 136.26, 130.19, 129.56, 129.04, 128.52,

128.35, 126.59, 126.39, 122.01, 62.07, 49.79, 44.21. Minor diastereomer (Ligand L2

accessed via sulfenate anion pathway, Section 1.4.11.1): 1H NMR (600 MHz, CDCl3), δ:

8.77 (dd, J = 4.3 Hz, 1.7 Hz, 1H), 8.28 (dd, J = 7.2 Hz, 1.4 Hz, 1H), 8.27 (dd, J = 8.4 Hz,

1.7 Hz, 1H), 7.94 (dd, J = 8.2 Hz, 1.3 Hz, 1H), 7.62 (dd, J = 8.0 Hz, 7.3 Hz, 1H), 7.47

(dd, J = 8.3 Hz, 4.3 Hz, 1H), 7.11 – 7.08 (m, 3H), 6.98 (m, 2H), 3.63 – 3.61 (m, 1H),

3.47 (dd, J = 13.2 Hz, 10.0 Hz, 1H), 3.06 (dd, J = 13.2 Hz, 2.2 Hz, 1H), 2.72 – 2.61 (m,

2H), 2.03 (s, 2H); 13C NMR (100 MHz, CDCl3), δ: 150.19, 143,82, 141.41, 137.90,

136.44, 130.38, 129.19, 128.51, 128.28, 126.84, 126.74, 126.55, 122.19, 60.86, 48.81,

44.34. Both diastereomers: IR (cm-1, neat): 3364, 3060, 3028, 2918, 1593, 1493, 1034.

N

S*O Bn

NH2

Page 128: Preparation and Activity of Multidentate beta-Amino

110

5.2.7 General procedure for the preparation of N-Ts protected heteroaryl amino

sulfoxides (1-31)

A mixture of anhydrous CH2Cl2 (0.1 M wrt amino sulfoxide), free base heteroaryl

amino sulfoxide (1.0 equiv), tosyl chloride (1.2 equiv) was stirred at 0 ˚C. Et3N (1.5

equiv) was added dropwise, and the temperature of the resulting solution was raised

slowly to room temperature. The solution was stirred until disappearance of starting

material (verified via TLC). The solvent was removed under reduced pressure and

triturated with hexanes.

N-Ts-1-Phenyl-3-(2-pyridylsulfinyl)propan-2-amine

Obtained as a diastereomeric mixture. White solid. Total

yield: 93%; mp = 168 – 170 ˚C. Major isomer: 1H NMR

(400 MHz, CDCl3), δ: 8.61 (d, J = 4.5 Hz, 1H), 7.95 (m,

1H), 7.89 (m, 1H), 7.62 (d, J = 8.2 Hz, 2H), 7.42 (m, 1H), 7.24 – 7.20 (m, 3H), 7.23 (m,

2H), 7.05 (m, 2H), 5.45 (d, J = 5.6 Hz, 1H), 3.92 (m, 1H), 3.27 (dd, J = 13.6 Hz, 6.3 Hz,

1H), 3.12 (m, 1H), 3.08 (m, 1H), 3.01 (m, 1H), 2.40 (s, 3H); 13C NMR (100 MHz, CDCl3),

δ: 163.92, 149.66, 143.45, 138.24, 136.24,135.71, 129.62, 129.44, 128.72, 127.19,

126.91, 124.84, 119.95, 58.14, 52.00, 40.84, 21.55. Minor isomer: 1H NMR (400 MHz,

CDCl3), δ: 8.53 (d, J – 4.5 Hz, 1H), 7.93 (m, 1H), 7.85 (m, 1H), 7.71 (d, J = 8.2 Hz, 2H),

7.42 (m, 1H), 7.24 – 7.20 (m, 3H), 7.18 (m, 2H), 7.09 (m, 2H), 6.09 (d J = 6.8 Hz, 1H),

N

S*Bn

NH

O O2S

Page 129: Preparation and Activity of Multidentate beta-Amino

111

3.99 (m, 1H), 3.12 (m, 1H), 3.03 (m, 1H), 2.98 (m, 1H), 2.87 (dd, J = 13.8 Hz, 3.8 Hz,

1H), 2.40 (s, 3H); 13C NMR (100 MHz, CDCl3), δ: 164.00, 149.55, 143.35, 138.19,

137.18, 135.98, 129.60, 129.41, 128.72, 127.20, 126.96, 124.70, 119.91, 57.17, 52.10,

41`.15, 21.55. Both diastereomers: IR (cm -1, neat): 3060, 2923, 1453, 1423, 1327,

1157, 1086, 1035; Analysis calculated for C21H22N2O3S2; C, 60.85; H, 5.35; found C,

60.85; H, 5.35.

N-Ts-1-Phenyl-3-(8-quinolylsulfinyl)propan-2-amine

Obtained as a diastereomeric mixture. White solid. Total

yield: 86%; mp = 171 – 172 ˚C. Major isomer: 1H NMR

(400 MHz, CDCl3), δ: 8.88 (dd, J = 4.29 Hz, 1.57 Hz, 1H),

8.29 (dd, J = 8.3 Hz, 1.5 Hz, 1H), 8.26 (m, 1H), 7.99 (m, 1H), 7.83 (d, J = 8.2 Hz, 2H),

7.79 (app t, J = 7.7 Hz), 7.54 (m, 1H), 7.29 (d, J = 8.2 Hz), 2H) 7.21 – 7.08 (m, 5H), 5.99

(d, J = 4.8 Hz, 1H), 3.93 (m, 1H), 3.62 (dd, J = 13.5, 6.3 Hz, 1H), 3.44 (dd, J = 13.7 Hz,

8.9 Hz, 1H), 3.24 (dd, J = 14.0 Hz, 7.4 Hz, 1H), 3.15 (dd, J = 14.0 Hz, 7.4 Hz, 1H), 2.41

(s, 3H); 13C NMR (100 MHz, CDCl3), δ: 150.18, 143.33, 142.77, 140.85, 137.00, 136.78,

136.62, 136.10, 130.51, 129.71, 129.60, 128.53, 128.21, 126.73, 126.63, 126.56,

122.28, 58.66, 53.49, 41.34, 21.57. Minor isomer: 1H NMR (400 MHz, CDCl3), δ: 8.81

(dd, J = 4.3 Hz ,1.6 Hz, 1H), 8.24 (m, 1H), 8.03 (m, 1H), 7.95 (m, 1H), 7.69 (app t, J =

N

S*Bn

NH

O O2S

Page 130: Preparation and Activity of Multidentate beta-Amino

112

7.8 Hz, 1H), 7.58 (d, J = 8.2 Hz, 2H), 7.51 (m, 1H), 7.21 – 7.08 (m, 5H), 7.19 (m, 2H),

3.93 (m, 1H), 3.33 (dd, J = 13.7 Hz, 4.4 Hz, 1H), 3.03 (m, 1H), 3.02 (m, 1H), 2.81 (dd, J

= 13.8 Hz, 2.76 Hz, 1H), 2.39 (s, 3H); 13C NMR (100 MHz, CDCl3), δ: 150.03, 143.31,

140.47, 136.78, 136.62, 136.43, 130.42, 129.64, 129.52, 128.39, 128.14, 127.43,

127.16, 126.70, 126.62, 126.55, 122.25, 58.16, 52.80, 40.24, 21.54 .Both

diastereomers: IR (cm -1, neat): 3061, 3029, 1595, 1493, 1455, 1328, 1202, 1086,

1044; TOF MS EI calculated for [C25H24N2O3S2]+: 464.1228; found: 464.1223.

5.2.8 General procedure for the synthesis of heteroaryl sulfanyl acrylates (1-32)41

Heteroaryl thiols (1.0 equiv) and methyl propiolate (1.3 equiv) were stirred in

anhydrous CH2Cl2 (0.25 M wrt thiol). The mixture was cooled to 0 ˚C and this was

followed by dropwise addition of Et3N (1.1 equiv). The solution was stirred, while the

temperature was slowly raised to room temperature over 60 minutes.. A 10% solution of

HCl(aq) was added when the reaction had reached completion, as confirmed by TLC. The

organic layer was separated. The aqueous layer was extracted by CH2Cl2 (3 x 50 mL)

and organic layers were combined. The organic layer was washed three times with

water, and three times with brine. The organic layer was dried of anhydrous MgSO4,

concentrated under reduced pressure, and purified by flash column chromatography.

Page 131: Preparation and Activity of Multidentate beta-Amino

113

2-Carbomethoxyethenyl 2-pyridyl sulfane

Obtained as mixture of (E) and (Z) isomers. Straw-coloured

solid. Total yield: 81%, mp = 79 – 90 ˚C. Major isomer (Z): 1H

NMR (400 MHz, CDCl3), δ: 8.57 (d, J = 10.3 Hz, 1H), 8.53 (dd, J = 4.9 Hz, J = 1.8 Hz,

1H), 7.62 (td, J = 7.9 Hz, 1.8 Hz, 1H), 7.33 (m, 1H), 7.14 (dd, J = 7.4 Hz, 4.9 Hz, 1H),

6.12 (d, J = 10.3 Hz, 1H), 3.79 (s, 3H); 13C NMR (100 MHz, CDCl3), δ: 167.25, 155.10,

149.63, 142.28, 136.84, 123.29, 121.50, 113.40, 51.54. Minor isomer (E): 1H NMR (400

MHz, CDCl3), δ: 8.60 (d, J = 15.9 Hz, 1H), 8.53 (dd, J = 4.9 Hz, 2.0 Hz, 1H), 7.58 (m,

1H), 7.27 (m, 1H), 7.14 (dd, J = 4.8 Hz, 0.9 Hz, 1H), 6.14 (d, J = 15.9 Hz, 1H), 3.75 (s,

3H); 13C NMR (100 MHz, CDCl3), δ: 165.58, 154.11, 150.12, 142.24, 136.94, 123.16,

121.35, 116.79, 51.54. Both isomers: IR (cm-1, neat): 3042, 2948, 1700, 1574, 1417,

1162, 1120, 1089.

2-Carbomethoxyethenyl 8-quinolyl sulfane

Obtained as mixture of (E) and (Z) isomers. Straw-coloured

solid. Total yield: 81%, mp 80 – 88 ˚C. Major isomer (Z): 1H

NMR (300 MHz, CDCl3), δ: 9.02 (d, J = 4.1 Hz, 1H), 8.20 (d, J =

8.3 Hz, 1H), 7.92 (d, J = 7.3 Hz, 1H), 7.82 (d, J = 8.2 Hz, 1H), 7.56 (m, 1H), 7.49 (d, J =

10.2 Hz, 1H), 7.23 (m, 1H), 6.01 (d, J = 10.2 Hz, 1H), 3.81 (s, 3H); 13C NMR (75 MHz,

N

S CO2Me

N

S CO2Me

Page 132: Preparation and Activity of Multidentate beta-Amino

114

CDCl3), δ: 166.78, 150.48, 148.44, 146.44, 136.55, 135.92, 131.55, 128.92, 128.38,

126.45, 121.84, 113.57, 50.83. Minor isomer (E): 1H NMR (300 MHz, CDCl3), δ: 8.98 (m,

1H), 8.20 (d, J = 8.3 Hz, 1H), 8.04 (d J = 15.1 Hz, 1H), 7. 86 (m, 1H), 7.79 (m, 1H), 7.49

(m, 1H), 7.45 (m, 1H), 5.95 (d, J = 15.1 Hz, 1H), 3.72 (s, 3H); 13C NMR (75 MHz,

CDCl3), δ: 165.66, 150.46, 146.42, 145.21, 136.58, 132.53, 131.10, 128.79, 128.26,

126.70, 122.02, 117.27, 51.51. Both isomers: IR (cm-1, neat): 3028, 2948, 1689, 1564,

1212, 1169. Analysis calculated for C13H11NO2S; C, 63.65; H, 4.52; found C, 63.46; H,

4.71.

5.2.9 General procedure for the preparation of heteroaryl sulfinyl acrylates (1-33)

5.2.9.1 mCPBA oxidation

Heteroaryl sulfanyl acrylates (1.0 equiv.) in anhydrous CH2Cl2 (0.5 M, wrt starting

material) were stirred and cooled to -78 ˚C. To this, a solution of mCPBA (1.0 equiv,

72% mCPBA) in anhydrous CH2Cl2 (0.05 M, wrt mCPBA) was added via canula. The

resulting solution was stirred, and while the temperature was slowly raised to room

temperature. The excess reagent and solvent were removed under reduced pressure.

Hexane was added and swirled around in the flask and was removed under reduced

pressure, and this process was repeated twice more. The resulting oil was re-dissolved

in CH2Cl2 and washed three times with a saturated aq. NaHCO3, three times with water,

Page 133: Preparation and Activity of Multidentate beta-Amino

115

and once with brine. It was then dried over anhydrous MgSO4, concentrated under

reduced pressure, and purified by flash column chromatography.

2-Carbomethoxyethenyl 2-pyridyl sulfoxide71

Obtained as mixture of (E) and (Z) isomers. Total yield: 84%,

mp = 83 – 89 ˚C. Major isomer (Z), straw-coloured oil: 1H NMR

(300 MHz, CDCl3), δ: 8.68; (dd, J = 4.7 Hz, 1.7 Hz, 1H), 8.01

(dt, J = 7.9 Hz, 1.1 Hz, 1H), 7.93 (td, J = 7.5 Hz, 1.7 Hz, 1H), 7.41 (dd, J = 7.4 Hz, 1.2

Hz, 1H), 6.90 (d, J = 10.3 Hz, 1H), 6.41 (d, J = 10.3 Hz, 1H), 3.84 (s, 3H); 13C NMR (100

MHz, CDCl3), δ: 164.26, 163.54, 152.44, 150.33, 137.93, 126.30, 125.15, 120.88, 52.34.

Minor isomer (E), pale-yellow solid: 1H NMR (300 MHz, CDCl3), δ: 8.65 (d, J = 4.7 Hz,

1H), 7.94 (m, 1H), 7.91 (m, 1H), 7.85 (d, J = 15.1 Hz, 1H), 7.41 (m, 1H), 6.72 (d, J =

15.1 Hz), 3.77 (s, 3H); 13C NMR (100 MHz, CDCl3), δ: 164.00, 162.53, 150.74, 149.84,

138.55, 124.86, 124.13, 118.96, 52.16; Both isomers: IR (cm-1, neat): 3049, 2592,

1728,1575, 1435, 1220, 1038.

(E)-2-Carbomethoxyethenyl quinolyl sulfoxide

Obtained as a dark-brown solid. Yield: 87%, mp = 108 -110 ˚C.

1H NMR (300 MHz, CDCl3), δ: 8.97 (dd, J = 4.3 Hz, 1.7 Hz, 1H),

N

S* CO2Me

O

N

S* CO2Me

O

Page 134: Preparation and Activity of Multidentate beta-Amino

116

8.31 (d, J = 15.1 Hz, 1H), 8.28 (dd, J = 8.4 Hz, 1.7 Hz, 1H), 8.18 (dd, J = 7.3 Hz, 1.4 Hz,

1H), 7.98 (dd, J = 8.2 Hz, 1.3 Hz, 1H), 7.68 (dd, J = 8.0 Hz, 7.4 Hz, 1H), 7.57 (dd, J =

8.4 Hz, 4.3 Hz, 1H), 6.78 (d, J = 15.1 Hz, 1H), 3.71 (s, 3H); 13C NMR (75 MHz, CDCl3),

δ: 164.93, 151.06, 150.74, 143.90, 140.01, 136.56, 130.60, 128.44, 127.26, 125.55,

123.38, 122.56, 52.69; IR (cm-1, neat): 3062, 2951, 1725, 1294, 1271, 1071. Analysis

calculated for C13H11NO3S; C, 59.76; H, 4.24; found C, 59.73; H, 4.25.

5.2.10 General procedure for lithiated heteroaryl sulfenate anion generation

Heteroaryl sulfinylacrylates (1.0 equiv.) in anhydrous THF (0.2 M, wrt starting

material) were stirred and cooled to -78 ˚C. The solution was treated with dropwise

addition of nBuLi (1.6 M in hexane, 1.0 equiv.), and stirred for 15 minutes. To this, a

solution of electrophile in anhydrous THF (0.25 M, wrt electrophile) was added quickly.

The solution was stirred for an additional 3 hours at -78 ˚C, and stirred overnight while

slowly warming to room temperature. The reaction mixture was concentrated under

reduced pressure, dissolved in CH2Cl2, and the organic layer was washed three times

with water, and once with brine. It was then dried over anhydrous MgSO4, concentrated

under reduced pressure, and purified by flash column chromatography.

Page 135: Preparation and Activity of Multidentate beta-Amino

117

5.2.10.1 Preparation of N-Boc protected heteroaryl amino sulfoxides (1-29)

(SS,2RC)-N-Boc-1-Phenyl-3-(2-pyridylsulfinyl)propan-2-amine71

A mixture of 2-Carbomethoxyethenyl 2-pyridyl sulfoxide,

nBuLi, and (2R)-N-Boc-1-Phenyl-3-iodopropan-2-amine

afforded a diastereomeric mixture of amino sulfoxides.

Total yield: 72%, d.r. = 1.0:0.7. White solid.

(SS,2RC)-N-Boc-1-Phenyl-3-(8-quinolylsulfinyl)propan-2-amine

A mixture of 2-Carbomethoxyethenyl 8-quinolyl sulfoxide,

nBuLi, and (2R)-N-Boc-1-Phenyl-3-iodopropan-2-amine

afforded one sulfoxide diastereomer. Obtained as an off-

white solid. Yield: 82%, mp = 121 – 122 ˚C. See analysis of minor diastereomer of N-

Boc-1-Phenyl-3-(8-quinolylsulfinyl)propan-2-amine. Removal of N-Boc protecting

group afforded Ligand 1-L2 Analysis calculated for C18H18N2OS (1-L2); C, 69.65; H,

5.85; found C, 70.00; H, 5.83.

N

SBn

NH

O

O

O

N

SO Bn

NH

O

O

Page 136: Preparation and Activity of Multidentate beta-Amino

118

5.2.10.2 Attempts towards enantioselective sulfoxide formation of heteroaryl

benzyl sulfoxide

Upon generation of the lithiated heteroaryl sulfenate anion in cold conditions as

indicated above, a solution of chiral PyBox ligand (10 mol% in 5 mL anhydrous THF)

was quickly added to the reaction mixture. The solution was stirred for an additional 30

minutes at -78 ˚C, before a solution of benzyl bromide in anhydrous THF (0.25 M) was

added quickly. The solution was stirred for an additional 3 hours at -78 ˚C, and stirred

overnight while slowly warming to room temperature. The reaction mixture was

concentrated under reduced pressure, dissolved in CH2Cl2, and the organic layer was

washed three times with water, and once with brine. It was then dried over anhydrous

MgSO4, concentrated under reduced pressure, and purified by flash column

chromatography.

NN

OO

N

L* PyBox1 =

NN

OO

N

L* PyBox2 =

Page 137: Preparation and Activity of Multidentate beta-Amino

119

2-Pyridyl benzyl sulfoxide

A mixture of 2-carbomethoxyethenyl 2-pyridyl sulfoxide, nBuLi,

PyBox ligand and benzyl bromide afforded an enantiomeric

mixture of sulfoxides. Obtained as an off-white solid. Yield: 50%,

mp = 87 – 89 ˚C (lit. mp = 86 – 89 ˚C)72. 1H NMR (300 MHz, CDCl3), δ: 8.67 (d, J = 4.7

Hz, 1H), 7.78 (app t, J = 7.7 Hz, 1H), 7.62 (d, J = 7.8 Hz, 1H), 7.36 (m, 1H), 7.26 – 7.01

(m, 5H), 4.39 (d, J = 13.1 Hz, 1H), 4.08 (d, J = 13.1 Hz, 1H); 13C NMR (100 MHz,

CDCl3), δ: 163.55, 149.24, 137.57, 130.19, 129.20, 128.21, 128.06, 124.53, 120.53,

59.76; IR (cm-1, neat): 3058, 2922, 1602, 1576, 1422, 1051. Determined by chiral HPLC

analysis (Chiralcel OD-H, hexane/isopropanol, 70;30 v.v, 0.5 mL/min, 20 ˚C, UV 220

nm): Retention times: tr = 16.648 for (R)-isomer, tr = 19.526 for (S)-isomer).

8-Quinolyl benzyl sulfoxide

A mixture of 2-Carbomethoxyethenyl 8-quinolyl sulfoxide, nBuLi,

PyBox ligand and benzyl bromide afforded an enantiomeric

mixture of sulfoxides. Obtained as a yellow solid. Yield: 56%, mp

= 90 – 91 ˚C (lit. mp = 94 – 95 ˚C).73 1H NMR (400 MHz, CDCl3), δ: 8.99 (dd, J = 4.2 Hz,

1.7 Hz, 1H), 8.28 (dd, J = 8.3 Hz, 1.7 Hz, 1H), 7.91(dd, J = 8.2 Hz, 1.3 Hz, 1H), 7.83

(dd, J = 7.2 Hz, 1.3 Hz, 1H), 7.21 (m, 1H), 7.16 (m, 1H) 7.26 – 7.13 (m, 3H), 6.91 (m,

2H), 4.60 (d, J = 12.9 Hz, 1H), 4.28 (d, J = 12.9 Hz, 1H); 13C NMR (75 MHz, CDCl3), δ:

N

S*O

S*

N

O

Page 138: Preparation and Activity of Multidentate beta-Amino

120

150.15, 144.11, 140.51, 136.43, 130.24, 130.05, 127.99, 127.91, 127.85, 127.38,

126.50, 121.95, 59.59; IR (cm-1, neat): 3030, 2926, 1561, 1046; Analysis calculated for

C16H13NOS; C, 71.88; H, 4.90; found C, 71.65; H, 5.08. Determined by chiral HPLC

analysis (Chiralcel OD-H, hexane/isopropanol, 70;30 v.v, 0.5 mL/min, 20 ˚C, UV 220

nm): Retention times: tr = 15.288 for (R)-isomer, tr = 18.174 for (S)-isomer.)

5.2.10.3 Attempts towards the release of lithiated [SO]-2

Upon generation of the lithiated heteroaryl sulfenate anion in cold conditions, an

additional equivalent of nBuLi was added to the reaction mixture. The solution was

stirred and slowly raised to room temperature over 3 hours, before a solution of benzyl

bromide in anhydrous THF (0.25 M) was added quickly and allowed to stir overnight.

The reaction mixture was concentrated under reduced pressure, dissolved in CH2Cl2,

and the organic layer was washed three times with water, and once with brine. It was

then dried over anhydrous MgSO4, concentrated under reduced pressure, and purified

by flash column chromatography. Attempts with either 2-carbomethoxyethenyl pyridyl

sulfoxide or 2-carbomethoxyethenyl quinolyl sulfoxide were unsuccessful and the

corresponding benzyl bromides from Section 1.4.11.2 were isolated.

Page 139: Preparation and Activity of Multidentate beta-Amino

121

5.2.11 General procedure for the preparation of quinolyl imino sulfoxides

A mixture of anhydrous CH2Cl2/methanol (3:1) (0.25 M wrt free base amino

sulfoxide), aryl aldehyde (1.0 equiv) and anhydrous Na2SO4 (0.5 g/mmol amino

sulfoxide) was refluxed for 4 h. The solution was filtered and the solvent was evaporated

under reduced pressure. The resulting residue was re-dissolved in CH2Cl2 and washed

three times with water, and once with brine. It was then dried over anhydrous MgSO4,

concentrated under reduced pressure, and triturated with diethyl ether to afford pure

product.

Ligand 1-L3

A mixture of (SS,2RC)-1-phenyl-3-(8-quinolylsulfinyl)propan-2-

amine, and 2-formylphenol afforded Ligand 1-L3. Obtained as

a yellow solid. Yield: 91%; mp = 133 – 134 ˚C. 1H NMR (600

MHz, CDCl3), δ: 8.93 – 8.92 (dd, J = 4.1 Hz, 1.4 Hz, 1H), 8.26 (d, J = 7.1 Hz, 1H), 8.23

(s, 1H), 8.21 (m, 2H), 7.95 (d, J = 8.0 1H), 7.75 (app t, J = 7.6 Hz, 1H), 7.51 (dd, J = 8.3

Hz, 4.2 Hz, 1H), 7.32 (app t, J = 7.0 Hz), 7. 21 – 7.13 (m, 4H), 7.08 (d, J = 7.1 Hz), 7.02

(d, J = 8.3 Hz), 6.86 (app t, J = 7.3 Hz), 4.23 (m, 1H), 4.02 (dd, J = 12.9 Hz, 10.5 Hz,

1H), 3.08 (dd, J = 12.9 Hz, 2.4 Hz, 1H), 2.99 (m, 2H); 13C NMR (150 MHz, CDCl3), δ:

166.78, 150.38, 143.96, 142.16, 137.12, 136.26, 132.59, 131.90, 130.20, 129.57,

128.52, 127.97, 126.62, 126.11, 122.14, 118.82, 117.03, 65.99, 61.55, 42.91; IR (cm-1,

N

SO Bn

N

H

OH

Page 140: Preparation and Activity of Multidentate beta-Amino

122

neat): 3442, 3060, 2918, 1644, 1494, 1106, 1046; [𝛼]!!" 426.4 𝑐 = 0.01 . Analysis

calculated for C25H22N2O2S; C, 72.44; H, 5.35; found C, 72.66; H, 5.57.

Ligand 1-L4

A mixture of (SS,2RC)-1-phenyl-3-(8-quinolylsulfinyl)propan-2-

amine, and 2-formylpyridine afforded Ligand 1-L4. Obtained

as an off-white solid. Yield: 89% yield; mp = 123 – 125 ˚C. 1H

NMR (600 MHz, CDCl3), δ: 8.91 (dd, J = 4.2 Hz, 1.68 Hz, 1H), 8.66 (d J = 4.5 Hz, 1H),

8.29 (s, 1H), 8.28 (d, J = 7.4 Hz), 8.23 (dd, J = 8.3 Hz, 1.7 Hz, 1H), 8.04 (d, J = 7.9 Hz,

1H), 7.94 (d, J = 7.0 Hz, 1H), 7.78 (m, 1H), 7.75 (m, 1H), 7.51 (dd, J = 8.3 Hz, 4.3 Hz,

1H), 7.32 (m, 1H), 7.16 (m, 2H), 7.12 (m, 3H), 4.33 (m, 1H), 4.07 (dd, J = 12.9 Hz, 9.8

Hz, 1H), 3.10 (dd, J = 12.9 Hz, 2.9 Hz, 1H), 3.01 – 2.99 (m, 2H); 13C NMR (150 MHz,

CDCl3), δ: 163.57, 154.21, 150.14, 149.52, 143.85, 142.42, 137.42, 136.41, 136.17,

130.07, 129.55, 128.19, 128.04, 126.64, 126.37, 126.26, 124.77, 122.21, 121.96, 66.59,

60.61, 42.51; IR (cm-1, neat): 3029, 2920, 1628, 1580, 1279, 1046; [𝛼]!!" 59.7 𝑐 = 0.05 .

TOF MS EI calculated for [C24H21N3OS]+: 399.1405; found: 399.1401.

5.2.12 General procedure for the Henry reaction

Chiral amino sulfoxide ligands (12 mol%) and Cu(OAc)2�H2O (10 mol%) were

stirred in solvent (0.1 M wrt aldehyde) in a round-bottom flask. The mixture was stirred

N

SO Bn

N

H N

Page 141: Preparation and Activity of Multidentate beta-Amino

123

at room temperature for 2 hours, then nitroalkyl (10 equiv.) and aldehyde (1 equiv.) were

successively added. The resulting solution was stirred for the specified time. The volatile

components were removed under reduced pressure and purified by flash column

chromatography. Trial reactions with ligands 1-L1m, and 1-L2m were carried out using

20 mol% each of chiral ligand and Cu(OAc)2�H2O.

5.2.12.1 Henry reaction products with nitromethane

2-Nitro-1-(4-nitrophenyl)-ethanol

A mixture of chiral ligand, 4-nitro-benzaldehyde, and

nitromethane afforded an enantiomeric mixture. Obtained as an

off-white solid, mp = 83 – 84 ˚C (lit. mp = 83 – 84 ˚C).38 1H

NMR (300 MHz, CDCl3), δ: 8.30 (d, J = 9.0 Hz, 2H), 8.26 (d = 9.0, 2H), 5.63 – 5.58 (m,

1H), 4.65 – 4.54 (m, 1H), 3.10 (s, 1H); 13C NMR (75 MHz, CDCl3), δ: 148.05, 144.85,

126.53, 124.11, 80.51, 69.86; IR (cm-1, neat): 3612, 2813, 1554, 1519,1349, 1084.

5.2.12.2 Henry reaction products with nitroethane

O2N

OHNO2*

Page 142: Preparation and Activity of Multidentate beta-Amino

124

anti/syn-1-(4-nitropehenyl)-2-nitropropan-1-ol

A mixture of chiral ligand, 4-nitro-

benzaldehyde, and nitroethan afforded an

enantiomeric mixture. Obtained as an

straw-coloured solid, mp = 83 – 85 ˚C (lit mp = 84 – 85 ˚C).74 Anti-diastereomer: 1H

NMR (400 MHz, CDCl3), δ: 8.28 – 8.25 (m, 2H), 7.61 – 7.57 (m, 2H), 5.57 (m, 1H), 4.74

– 4.69 (m, 1H), 2.95 (d, J = 3.6 Hz, 1H), 1.5 (d, J = 6.8 Hz, 3H); 13C NMR (100 MHz,

CDCl3), δ: 147.91, 145.31, 126.96, 123.96, 86.73, 72.80, 11.85. Syn-diastereomer: 1H

NMR (400 MHz, CDCl3), δ: 8.29 – 8.25 (m, 2H), 7.62 – 7.57 (m, 2H), 5.21 – 5.17 (dd, J

= 8.8 Hz, 4.4 Hz, 1H), 4.80 – 4.75 (m, 1H), 2.95 (d, J = 4.4 Hz, 1H), 1.40 (d, J = 6.8 Hz,

3H); 13C NMR (100 MHz, CDCl3), δ: 148.30, 145.13, 127.86, 124.11, 87.70, 75.01,

16.24. Both diastereomers: IR (cm-1, neat): 3526, 3114, 2917, 1550, 1520, 1349.

5.3.0 Experimental: Synthesis of Bupropion

5.3.1 General procedure for the copper-mediated α-amination of 3’-

chloropropiophenone

CuBr2 was (10 mol%) dissolved in DMSO (1.0 M wrt aryl ketone) in a closed

vessel before 3’-chloropropiophenone was added. This solution was stirred for 15

O2N

OH

NO2

antiO2N

OH

NO2

syn

Page 143: Preparation and Activity of Multidentate beta-Amino

125

minutes and alkyl amine (3.0 equiv) was added. The reaction was stirred for the

specified time at room temperature. The crude reaction mixture was diluted with water

and extracted three times with Et2O. The combined organic layer was washed three

times with water and brine. It was then dried over anhydrous MgSO4, concentrated

under reduced pressure, and purified by flash-column chromatography.

5.3.1.1 α-Amination using tBuNH2

A mixture of CuBr2, 3’-chloropropiophenone, tBuNH2 did not yield product.

5.3.1.2 α-Amination using morpholine

2-morpolino-1-(3-chlorophenyl)propan-1-one (2-14)

A mixture of CuBr2, 3’-chloropropiophenone, morpholine afforded the

α-amino aryl ketone. Obtained as a viscous pale-white oil. Yield: 3.6 –

19%. 1H NMR (400 MHz, CDCl3), δ: 8.08 (t, J = 1.2 Hz, 1H), 7.99 (dt, J

= 7.6 Hz, 1.2 Hz, 1H), 7.54 (dd, J = 7.9, 1.9 Hz, 1H), 7.41 (t, J = 8.0 Hz, 1H), 4.04 (q, J =

6.8 Hz,1 H), 3.73 (- 3.64 (m, 4H), 2.64 – 2.53 (m, 4H), 1.29 (d, J = 6.8 Hz, 3 H).

Cl

O

N

O

Page 144: Preparation and Activity of Multidentate beta-Amino

126

Predicted by-product (2-15)

Obtained as a yellow oil. 1H NMR (400 MHz, CDCl3), δ: 7.44

(t, J = 1.6 Hz, 1H), 7.38 (dt, J = 7.6 Hz, 1.6 Hz, 1H), 7.33 (m,

1H), 7.31 (m, 1H), 6.45 (s, 1H), 3.82 (m, 4H), 3.73 – 3.68 (m,

8H), 3.15 (m, 4H).

5.3.2 General procedure for the I2 mediated α-amination of 3’-

chloropropiophenone

NH4I (15 mol%), sodium percarbonate (2.0 equiv), and 3’-chloropropiophenone

(1.0 equiv) were stirred in solvent (0.5 M wrt aryl ketone) at room temperature before

the mixture was heated to 50 ˚C and stirred for the specified time. The crude reaction

mixture was concentrated under reduced pressure, and re-dissolved in CH2Cl2. The

organic layer was washed three times with saturated aq. NaHCO3, twice with water, and

once with brine. It was dried over anhydrous MgSO4, and concentrated under reduced

pressure.

2-tert-butylamino-1-(3-chlorophenyl)propan-1-one (Bupropion)

Conditions #1: A mixture of NH4I (15 mol%), sodium percarbonate (2.0

equiv), and 3’-chloropropiophenone (1.0 equiv) were stirred in EtOAc

Cl

O

NH

Cl

N

Br

H

NO

O

X

Page 145: Preparation and Activity of Multidentate beta-Amino

127

(0.5 M wrt aryl ketone. Conditions #2: A mixture of I2 (15 mol%), AcOH (15 mol%),

sodium percarbonate (2.0 equiv), and 3’-chloropropiophenone (1.0 equiv) were stirred in

EtOAc (0.5 M wrt aryl ketone). Conditions #3: A mixture of I2 (15 mol%), PhOH (15

mol%), sodium percarbonate (2.0 equiv), and 3’-chloropropiophenone (1.0 equiv) were

stirred in EtOAc (0.5 M wrt aryl ketone). Obtained as viscous brown oil. 1H NMR (400

MHz, CDCl3), δ: 7.96 (t, J = 1.7 Hz, 1H), 7.8 (dt, J = 8.9 Hz, 1.2 Hz, 1H), 7.55 (s, 1H),

7.58 (dd, J = 8.0 Hz, 2.1 Hz, 1H), 7.47 (t, J = 8.0 Hz, 1H); 4.33 (q, J = 7.1 Hz, 1H), 1.26

(d, J = 7.2 Hz, 3H), 1.05 (s, 9H).

5.3.3 General procedure for the preparation of Bupropion HCl salt

HCl gas was bubbled into IPA to afford a saturated acidic solution. This was used

to dissolved crude mixtures of Bupropion while stirring at 0 ˚C. This reaction was

concentrated under reduced pressure and this process was repeated twice more. The

resultant viscous liquid was triturated with hexanes and EtOAc to afford the HCl salt.

Bupropion HCl

Obtained as a dark-orange solid. Yield = 67-78%, mp = 233 – 235

˚C. 1H NMR (300 MHz, CDCl3), δ: 8.40 (s, 1H), 8.00 (s, 1H), 7.97

(d, J = 7.9 Hz, 1H), 7.67 (d, J = 6.9 Hz, 1H), 7.54 (d, J = 7.8 Hz,

1H), 5.02 (q, J = 7.2 Hz, 1H), 1.89 (d, J = 7.2 Hz, 3H), 1.47 (s, 9H); 13C NMR (75 MHz,

Cl

O

NHHCl

Page 146: Preparation and Activity of Multidentate beta-Amino

128

CDCl3), δ: 194.38, 136.11, 135.52, 133.07, 130.93, 129.00, 127.02, 59.46, 53.64, 26.58,

18.47.

Page 147: Preparation and Activity of Multidentate beta-Amino

129

Chapter 5: References

Page 148: Preparation and Activity of Multidentate beta-Amino

130

(1) Bauer, E. B. Chem. Soc. Rev. 2012, 41, 3153.

(2) Mellah, M.; Voituriez, A.; Schulz, E. Chem. Rev. 2007, 107, 5133.

(3) Yoon, T. P.; Jacobsen, E. N. Science 2003, 299, 1691.

(4) Fernández, I.; Khiar, N. Chem. Rev. 2003, 103, 3651.

(5) Pellissier, H. Tetrahedron 2007, 63, 1297.

(6) Pellissier, H. Chiral Sulfur Ligands: Asymmetric Catalysis; RSC Publishing:

Cambridge, 2009.

(7) Anderson, J. C.; Cubbon, R.; Harding, M.; James, D. S. Tetrahedron: Asymmetry

1998, 9, 3461.

(8) Kang, J.; Kim, D. S.; Kim, J. I. Synlett 1994, 1994, 842.

(9) Hof, R. P.; Poelert, M. A.; Peper, N. C. M. W.; Kellogg, R. M. Tetrahedron:

Asymmetry 1994, 5, 31.

(10) Ramón, D. J.; Yus, M. Tetrahedron Lett. 1998, 39, 1239.

(11) Lu, G.; Li, X.; Jia, X.; Chan, W. L.; Chan, A. S. C. Angew. Chem., Int. Ed. 2003,

42, 5057.

(12) Trost, B. M.; Rao, M. Angew. Chem., Int. Ed. 2015, 54, 5026.

(13) Sipos, G.; Drinkel, E. E.; Dorta, R. Chem. Soc. Rev. 2015, 44, 3834.

(14) James, B. R.; McMillan, R. S.; Reimer, K. J. J. Mol. Catal. A: Chem. 1976, 1, 439.

(15) James, B. R.; McMillan, R. S. Can. J. Chem. 1977, 55, 3927.

(16) Kvintovics, P.; James, B. R.; Heil, B. Chem. Commun. 1986, 1810.

(17) Carreño, M. C.; García Ruano, J. L.; Maestro, M. C.; Martín Cabrejas, L. M.

Tetrahedron: Asymmetry 1993, 4, 727.

Page 149: Preparation and Activity of Multidentate beta-Amino

131

(18) Allen, J. V.; Bower, J. F.; Williams, J. M. J. Tetrahedron: Asymmetry 1994, 5,

1895.

(19) Hiroi, K.; Suzuki, Y. Heterocycles 1997, 46, 77.

(20) Hiroi, K.; Suzuki, Y.; Abe, I.; Kawagishi, R. Tetrahedron 2000, 56, 4701.

(21) Cheng, H.-G.; Feng, B.; Chen, L.-Y.; Guo, W.; Yu, X.-Y.; Lu, L.-Q.; Chen, J.-R.;

Xiao, W.-J. Chem. Commun. 2014, 50, 2873.

(22) Petra, D. G. I.; Kamer, P. C. J.; Spek, A. L.; Schoemaker, H. E.; Van Leeuwen, P.

W. N. M. J. Org. Chem. 2000, 65, 3010.

(23) Khiar, N.; Fernández, I.; Alcudia, F. Tetrahedron Lett. 1993, 34, 123.

(24) Hiroi, K.; Ishii, M. Tetrahedron Lett. 2000, 41, 7071.

(25) Hiroi, K.; Izawa, I.; Takizawa, T.; Kawai, K.-i. Tetrahedron 2004, 60, 2155.

(26) Hiroi, K.; Suzuki, Y.; Abe, I. Tetrahedron: Asymmetry 1999, 10, 1173.

(27) Hiroi, K.; Suzuki, Y.; Abe, I.; Hasegawa, Y.; Suzuki, K. Tetrahedron: Asymmetry

1998, 9, 3797.

(28) Hiroi, K.; Suzuki, Y.; Kawagishi, R. Tetrahedron Lett. 1999, 40, 715.

(29) Hiroi, K.; Watanabe, K.; Abe, I.; Koseki, M. Tetrahedron Lett. 2001, 42, 7617.

(30) Hiroi, K.; Watanabe, K.; Abe, I.; Koseki, M. Tetrahedron Lett. 2001, 42, 7617.

(31) Watanabe, K.; Hirasawa, T.; Hiroi, K. Chem. Pharm. Bull. 2002, 50, 372.

(32) Rachwalski, M.; Kaczmarczyk, S.; Leśniak, S.; Kiełbasiński, P. ChemCatChem

2014, 6, 873.

(33) Rachwalski, M.; Kwiatkowska, M.; Drabowicz, J.; Kłos, M.; Wieczorek, W. M.;

Szyrej, M.; Sieroń, L.; Kiełbasiński, P. Tetrahedron: Asymmetry 2008, 19, 2096.

Page 150: Preparation and Activity of Multidentate beta-Amino

132

(34) Rachwalski, M.; Lesniak, S.; Kielbasinski, P. Tetrahedron: Asymmetry 2011, 22,

1087.

(35) Rachwalski, M.; Leśniak, S.; Kiełbasiński, P. Tetrahedron: Asymmetry 2010, 21,

1890.

(36) Rachwalski, M.; Leśniak, S.; Kiełbasiński, P. Tetrahedron: Asymmetry 2010, 21,

2687.

(37) Rachwalski, M.; Leśniak, S.; Sznajder, E.; Kiełbasiński, P. Tetrahedron:

Asymmetry 2009, 20, 1547.

(38) Cheng, H.-G.; Lu, L.-Q.; Wang, T.; Chen, J.-R.; Xiao, W.-J. Chem. Commun.

2012, 48, 5596.

(39) O'Donnell, J. S.; Schwan, A. L. J. Sulfur Chem. 2004, 25, 183.

(40) Schwan, A. L.; Söderman, S. C. Phosphorus, Sulfur Silicon Relat. Elem. 2012,

188, 275.

(41) Singh, S. P.; O'Donnell, J. S.; Schwan, A. L. Org. Biomol. Chem. 2010, 8, 1712.

(42) Söderman, S. C.; Schwan, A. L. Org. Lett. 2011, 13, 4192.

(43) Söderman, S. C.; Schwan, A. L. J. Org. Chem. 2013, 78, 1638.

(44) Boldyrev, B. G.; Vid, L. V.; Kolesnikova, S. A.; Kolmakova, L. E.; Lutsenko, L. N.;

Stoyanovskaya, Y. N. Strukt. Mekh. Deistviya Fiziol. Aktiv. Veshchestv 1972, 63.

(45) Bella, M.; Kobbelgaard, S.; Jørgensen, K. A. J. Am. Chem. Soc. 2005, 127,

3670.

(46) Tjosaas, F.; Fiksdahl, A. Molecules 2006, 11, 130.

(47) Bellale, E. V.; Chaudhari, M. K.; Akamanchi, K. G. Synthesis 2009, 2009, 3211.

Page 151: Preparation and Activity of Multidentate beta-Amino

133

(48) Wang, M.; Wei, J.; Fan, Q.; Jiang, X. Tetrahedron 2016, 72, 2671.

(49) Sandrinelli, F.; Perrio, S.; Averbuch-Pouchot, M.-T. Org. Lett. 2002, 4, 3619.

(50) Zong, L.; Ban, X.; Kee, C. W.; Tan, C.-H. Angew. Chem., Int. Ed. 2014, 53,

11849.

(51) Caupène, C.; Boudou, C.; Perrio, S.; Metzner, P. J. Org. Chem. 2005, 70, 2812.

(52) Gelat, F.; Gaumont, A.-C.; Perrio, S. J. Sulfur Chem. 2013, 34, 596.

(53) Gao, B.; Hu, J.; Zhao, Y.; Hu, J. Tetrahedron Lett. 2015, 56, 4180.

(54) Chen, L.-Y.; Chen, J.-R.; Cheng, H.-G.; Lu, L.-Q.; Xiao, W.-J. Eur. J. Org. Chem.

2014, 2014, 4714.

(55) Evans, D. A.; Seidel, D.; Rueping, M.; Lam, H. W.; Shaw, J. T.; Downey, C. W. J.

Am. Chem. Soc. 2003, 125, 12692.

(56) Boruwa, J.; Gogoi, N.; Saikia, P. P.; Barua, N. C. Tetrahedron: Asymmetry 2007,

17, 3315.

(57) Palomo, C.; Oiarbide, M.; Laso, A. Eur. J. Org. Chem. 2007, 2561.

(58) Hathaway, B. J.; Billing, D. E. Coord. Chem. Rev. 1970, 5, 143.

(59) Jahn, H. A.; Teller, E. Proc. R. Soc. A 1937, 161, 220.

(60) Bøgevig, A.; Juhl, K.; Kumaragurubaran, N.; Zhuang, W.; Jørgensen, K. A.

Angew. Chem., Int. Ed. 2002, 114, 1868.

(61) Janey, J. M. Angew. Chem., Int. Ed. 2005, 44, 4292.

(62) Momiyama, N.; Yamamoto, H. J. Am. Chem. Soc. 2004, 126, 5360.

(63) Evans, D. A.; Nelson, S. G. J. Am. Chem. Soc. 1997, 119, 6452.

Page 152: Preparation and Activity of Multidentate beta-Amino

134

(64) Cecere, G.; König, C. M.; Alleva, J. L.; MacMillan, D. W. C. J. Am. Chem. Soc.

2013, 135, 11521.

(65) Evans, R. W.; Zbieg, J. R.; Zhu, S.; Li, W.; MacMillan, D. W. C. J. Am. Chem.

Soc. 2013, 135, 16074.

(66) Guha, S.; Rajeshkumar, V.; Kotha, S. S.; Sekar, G. Org. Lett. 2015, 17, 406.

(67) Jiang, Q.; Xu, B.; Zhao, A.; Jia, J.; Liu, T.; Guo, C. J. Org. Chem. 2014, 79, 8750.

(68) Mehta, N.; Burroughs Wellcome Co,; Meta-Chloro Substituted alpha-

Butylaminopropiophenones. U.S. Patent 3,819,706, 1974.

(69) Pinto, G.; Rohrig, B. J. Chem. Educ. 2003, 80, 41.

(70) Luo, Y.-R. Comprehensive Handbook of Chemical Bond Energies; 3rd ed.; CRC

Press: Boca Raton, 2007.

(71) Söderman, S. C. Ph.D. Dissertation, University of Guelph, Guelph, ON, 2012.

(72) Dean, W. M.; Šiaučiulis, M.; Storr, T. E.; Lewis, W.; Stockman, R. A. Angew.

Chem., Int. Ed. 2016, 55, 10013.

(73) Jia, T.; Zhang, M.; Jiang, H.; Wang, C. Y.; Walsh, P. J. J. Am. Chem. Soc. 2015,

137, 13887.

(74) Shrinidhi, A. Cogent Chem. 2015, 1, 1061412.