213
STUDIES TOWARD THE TOTAL SYNTHESIS OF MULBERRY DIELS- ALDER ADDUCTS MORUSALBANOL A AND SOROCEIN B TEE JIA TI FACULTY OF SCIENCE UNIVERSITY OF MALAYA KUALA LUMPUR 2016

STUDIES TOWARD THE TOTAL SYNTHESIS OF MULBERRY … · 2017-08-14 · Setakat ini, tiada laporan mengenai sintesis sebatian ini sejak pengasingan pertama mereka ... 2.2.3 Chiral-Boron-complex

Embed Size (px)

Citation preview

STUDIES TOWARD THE TOTAL SYNTHESIS OF MULBERRY DIELS-ALDER ADDUCTS

MORUSALBANOL A AND SOROCEIN B

TEE JIA TI

FACULTY OF SCIENCE UNIVERSITY OF MALAYA

KUALA LUMPUR

2016

STUDIES TOWARD THE TOTAL SYNTHESIS OF

MULBERRY DIELS-ALDER ADDUCTS

MORUSALBANOL A AND SOROCEIN B

TEE JIA TI

THESIS SUBMITTED IN FULFILMENT

OF THE REQUIREMENTS FOR

THE DEGREE OF DOCTOR OF PHILOSOPHY

DEPARTMENT OF CHEMISTRY

FACULTY OF SCIENCE

UNIVERSITY OF MALAYA

KUALA LUMPUR

2016

ii

UNIVERSITY OF MALAYA

ORIGINAL LITERARY WORK DECLARATION

Name of Candidate: TEE JIA TI (I.C/Passport No: 851230-01-5116)

Registration/Matric No: SHC 120015

Name of Degree: DOCTOR OF PHILOSOPHY

Title of Project Paper/Research Report/Dissertation/Thesis (“this Work”):

STUDIES TOWARD THE TOTAL SYNTHESIS OF MULBERRY DIELS-

ALDER ADDUCTS MORUSALBANOL A AND SOROCEIN B

Field of Study: TOTAL SYNTHESIS, ORGANIC CHEMISTRY

I do solemnly and sincerely declare that:

(1) I am the sole author/writer of this Work;

(2) This Work is original;

(3) Any use of any work in which copyright exists was done by way of fair dealing

and for permitted purposes and any excerpt or extract from, or reference to or

reproduction of any copyright work has been disclosed expressly and sufficiently

and the title of the Work and its authorship have been acknowledged in this

Work;

(4) I do not have any actual knowledge nor do I ought reasonably to know that the

making of this work constitutes an infringement of any copyright work;

(5) I hereby assign all and every rights in the copyright to this Work to the

University of Malaya (“UM”), who henceforth shall be owner of the copyright in

this Work and that any reproduction or use in any form or by any means

whatsoever is prohibited without the written consent of UM having been first had

and obtained;

(6) I am fully aware that if in the course of making this Work I have infringed any

copyright whether intentionally or otherwise, I may be subject to legal action or

any other action as may be determined by UM.

Candidate’s Signature Date:

Subscribed and solemnly declared before,

Witness’s Signature Date:.

Name:

Designation:

iii

ABSTRACT

Morusalbanol A (42) and sorocein B (62) are biologically active natural products

isolated from moraceous plants. They are postulated to be biosynthetically derived from an

intramolecular cyclization/ketalization of a cis-trans mulberry Diels-Alder adduct. Thus far,

there has been no report on the synthesis of these compounds since their first isolation in

1991. A model studies on construction of the oxabicyclic [3.3.1] core of morusalbanol A

was conducted prior to embarking on the synthesis of morusalbanol A and sorocein B. The

results showed that the required cis-trans Diels-Alder precursors of morusalbanol A was

obtained via the thermal cycloaddition reaction which was proven to be dependent on the

presence of a hydrogen-bonded ortho OH substituent on the chalcones dienophile. Acid

catalyzed intramolecular cyclization of a cis-trans Diels-Alder adduct (endo-181) afforded

the desired oxabicyclic [3.3.1] core of morusalbanol A in a stereocontrolled manner.

Results from the model study have been applied to the synthesis of morusalbanol A (42).

The requisite cis-trans Diels-Alder precursor (endo-204) of morusalbanol A was

successfully prepared via the thermal cycloaddition reaction between the methyl ether

protected chalcones dienophile 88 and the dehydroprenyl diene 191. Selective removal of

the ortho OMe group of the endo-204 by using MgI2 gave (±) morusalbanol A methyl

ethers 206 and 207. A number of key proton and carbon signals in the NMR spectra of 206

and 207 were absent as a result of atroisomerism due to the rotational hindrance of the

Diels-Alder-rings about the C5´´-15´´ and C4´´-C8´´-C9´´ bond. Global demethylation on

the (±) morusalbanol A methyl ethers 206 and 207 by using MgI2, BCl3, TMSI-quinoline

was unsuccessful. An efficient method for preparing 2´2-dimethyl-2H-chromones via

Pd(II)-catalyzed Heck coupling of o-halophenols with 2-methyl-3-buten-2-ol has been

developed during the synthesis of sorocein B. The method is very general and can be useful

to the synthesis of some natural 2´,2-dimethyl-2H-chromones. Similar strategy was used for

construction of the cis-trans Diels-Alder precursor of sorocein B. The thermal

cycloaddition reaction between chalcones dienophile 89 and dehydroprenyl diene 106

afforded the requisite cis-trans Diels-Alder precursor (endo-241) in 35% yield along with

the trans-trans Diels-Alder diastereomer. Subsequently PdCl2 catalyzed cyclization of the

ortho-prenyl group of endo-241 afforded the required 2,2-dimethylchromenyl ring in 230.

However, attempts to remove the methyl ether group of 230 to form sorocein B with BCl3,

MgI2, and TMSI-quinoline were unsuccessful.

iv

ABSTRAK

Morusalbanol A (42) dan sorocein B (62) adalah produk semula jadi biologi aktif

yang diasingkan daripada tumbuhan moraceous. Biosynthetically, mereka berasal dari

intramolecular cyclization / ketalization daripada cis-trans mulberi Diels-Alder adduct.

Setakat ini, tiada laporan mengenai sintesis sebatian ini sejak pengasingan pertama mereka

pada tahun 1991. Sebelum memulakan sintesis morusalbanol A dan sorocein B, kajian

tentang pembinaan model oxabicyclic [3.3.1] teras bagi morusalbanol A telah dijalankan.

Hasil kajian menunjukkan bahawa kehadiran orto OH hidrogen terikat pada dienophile

chalcone itu diperlukan dalam tindak balas cycloaddition haba. Intramolecular cyclization

daripada cis-trans Diels-Alder adduct (endo-181) yang dimangkin oleh asid berjaya

memberi model oxabicyclic [3.3.1] teras bagi morusalbanol A dengan cara yang

stereocontrolled. Hasil daripada kajian model telah digunakan untuk sintesis morusalbanol

A (42). Cis-trans Diels-Alder (endo-204) dari morusalbanol A telah berjaya disediakan

melalui tindak balas cycloaddition haba antara metil eter chalcone dienophile 88 dan

dehydroprenyl diene 191. Penyingkiran dariapada kumpulan orto OMe (endo-204) dengan

menggunakan MgI2 berjaya memberi (±) morusalbanol A metil eter 205 dan 206. Keadaan

atroisomerism menyebabkan ketidakhadiran beberapa proton dan karbon dalam spektrum

NMR bagi 206 dan 207. Keadaan ini berlaku disebabkan terdapat halangan putaran pada

Diels-Alder ring di bond C5´´-15´´ dan C4´´-C8´´-C9´´ itu. Global demethylation ke atas (±)

morusalbanol metil eter A 206 dan 207 dengan menggunakan MgI2, BCl3, TMSI-quinoline

tidak berjaya. Semasa sintesis sorocein B, satu kaedah yang berkesan untuk menyediakan

2'2-dimetil-2H-chromones melalui Pd (II) gandingan Heck -catalyzed o-halophenols

dengan 2-metil-3-buten-2-ol telah dibangunkan. Kaedah ini sangat umum dan boleh

digunakan untuk sintesis beberapa 2,2-dimetil-2H-chromones yang semulajadi. Strategi

yang sama telah digunakan dalam pembinaan cis-trans Diels-Alder sorocein B. Reaksi

cycloaddition haba antara chalcone dienophile 89 dan dehydroprenyl diene 106 memberi

cis-trans Diels-Alder adduct (endo-241) dalam hasil 35% bersama-sama dengan trans-trans

Diels-Alder diastereomer. Selepas itu, cyclization kumpulan orto-prenyl group daripada

endo-241 yang dimangkin oleh PdCl2 berjaya memberikan 2,2-dimethylchromenyl ring

230. Walau bagaimanapun, usaha untuk menghapuskan kumpulan metil eter 230 bagi

membentuk sorocein B dengan BCl3, MgI2 dan TMSI-quinoline tidak berjaya.

v

ACKNOWLEDGEMENTS

This study would not have been possible without the help and assistance of many.

First and foremost, I would like to express my sincere gratitude to my supervisor, Prof Dr

Noorsaadah Abd. Rahman for her invaluable guidance and enthusiasm throughout the research

project and the completion of this thesis. I gratefully acknowledge her for her constructive advices

and suggestions on my research work. It had been a privilege to join the Drug design and

development research group (DDDRG) and work under her supervision.

I appreciate Dr. Chee Chin Fei for his patients, whose encouragement, guidance and

support from the initial to the final level enabled me to complete this project. I am also grateful to

him for reading through the draft of the thesis and advice in the compilation of the thesis.

I am indebted to Dr. Marzieh Yaeghoobi who gave me an extreme support in showing me

the way to think logically in research. To the DDDRG group members with whom I have worked

directly, I would like to collectively thank the group for providing informational and critical

discussions. I am especially proud to be one of them. In addition, I would like to acknowledge to

my friends and lab-mates, especially Assoc. Prof Dr. Michael James Christopher Buck, Dr. Lee

Yean Kee, Dr. Hamid Khaledi for providing valuable comments, stimulating suggestions and

motivation work atmosphere.

Aside from the DDDRG group, I would like to thank all the staff for their generous help.

This work was made possible with SLAI scholarship and research grants from the University of

Malaya (PG020/2014A).

Finally, I would like to thank my family for all their love and encouragement. This thesis

would not have been possible without their continuous support and caring of them.

vi

TABLE OF CONTENTS

Abstract ........................................................................................................................... iii

Abstrak ............................................................................................................................ iv

Acknowledgements .......................................................................................................... v

Table of Contents ........................................................................................................... vi

List of Figures ................................................................................................................. ix

List of Schemes ............................................................................................................... xi

List of Tables ................................................................................................................ xiii

List of Symbols and Abbreviations ............................................................................. xiv

CHAPTER 1: INTRODUCTION AND LITERATURE REVIEW ........................... 1

1.1 Introduction ............................................................................................................. 1

1.2 Mulberry Diels-Alder adducts ................................................................................. 2

1.3 Absolute configurations of mulberry Diels-Alder type adducts .............................. 9

1.4 Biosynthesis of mulberry Diels-Alder adducts ...................................................... 15

1.5 Biological activity of mulberry Diels-Alder type adducts ..................................... 18

1.5.1 Antioxidation ................................................................................................ 19

1.5.2 Anti-inflammation ........................................................................................ 22

1.5.3 Cytotoxicity .................................................................................................. 23

1.5.4 Antimicrobial and antifungus ....................................................................... 25

1.6 Scope and objectives of this thesis ........................................................................ 26

CHAPTER 2: RELATED SYNTHESIS ..................................................................... 28

2.1 Semisythesis of mulberry Diels-Alder adducts ..................................................... 28

vii

2.2 Total synthesis of mulberry Diels-Alder adducts .................................................. 31

2.2.1 Thermal conditions for Diels-Alder reation ................................................. 31

2.2.2 Catalytic Diels-Alder reaction ...................................................................... 41

2.2.3 Chiral-Boron-complex promoted enantioselective synthesis of MDA ........ 51

CHAPTER 3: APPROACHES TOWARD THE SYNTHESIS OF

MORUSALBANOL A ................................................................ 58

3.1 Model study towards biomimetic Diels-Alder reaction Morusalbanol A ............. 59

3.2 DFT calculation for model study ........................................................................... 68

3.3 Approaches towards the synthesis of morusalbanol A .......................................... 73

3.3.1 Retrosynthetic analysis for morusalbanol A ................................................. 74

3.3.2 Synthesis of diene 189 .................................................................................. 75

3.3.3 Synthesis of dienophile 191 .......................................................................... 76

3.3.4 Diels-Alder reaction between 189 and 191 .................................................. 77

3.3.5 Synthesis of diene 190 .................................................................................. 78

3.3.6 Synthesis of dienophile 87 ............................................................................ 78

3.3.7 Diels-Alder reaction between 190 and 87 .................................................... 79

CHAPTER 4: APPROACHES TOWARD THE SYNTHESIS OF

SOROCEIN B .............................................................................. 88

4.1 Retrosynthetic analysis .......................................................................................... 88

4.1.1 Synthesis of chalcone core as diene 105....................................................... 89

4.1.2 Synthesis of dienophile 208 .......................................................................... 90

4.1.3 Method development for installation of 2,2-Dimethyl-2H-chromenes (pyran

moiety) .......................................................................................................... 91

viii

4.1.4 Cycloaddition reaction between 105 and 208 ............................................... 97

4.2 Cycloaddition reaction of diene 232 and dienophile 233 ...................................... 98

4.2.1 Synthesis of diene 232 .................................................................................. 98

4.2.2 Synthesis of dienophile 233 .......................................................................... 99

4.3 Cycloaddition reaction of diene 105 and dienophile 210 .................................... 101

4.4 Cycloaddition reaction of diene 105 and dienophile 88 ...................................... 102

CHAPTER 5: CONCLUSION AND FUTURE WORK ......................................... 107

5.1 Conclusion ........................................................................................................... 107

5.2 Suggestion for Future work ................................................................................. 108

CHAPTER 6: EXPERIMENTALS ........................................................................... 111

References .................................................................................................................... 138

APPENDIX : SPECTRA OF COMPOUNDS .......................................................... 150

LIST OF PUBLICATIONS ........................................................................................ 193

ix

LIST OF FIGURES

Figure ‎1.1: Flavonoids from mulberry tree ....................................................................... 2

Figure ‎1.2: Biogenetic synthesis of MDA and cylised MDA adducts .............................. 3

Figure ‎1.3: Examples of dehydroprenylchalcone adduct .................................................. 6

Figure ‎1.4: Examples dehydroprenylflavonoid adduct ..................................................... 7

Figure 1.5: Examples of dehydroprenylstilbene adducts .................................................. 8

Figure ‎1.6: Examples of prenylarylbenzofuran type adducts ........................................... 9

Figure 1.7:The CD spectra of mulberrofuran C (24) and J (25) and their absolute

configurations……………………………………………………………..10

Figure ‎1.8: The reduction of mulberrofuran C (24) and mulberrofuran J (25) and their

CD spectra of their reduction products........................................................11

Figure ‎1.9: The X-ray analysis of mulberrofuran G pentamethyl ether (30) .................. 12

Figure ‎1.20: Feeding experiments with methyl chalcones 33 and prenyl chalcones 34 to the

Morus alba cell cultures .................................................................................................. 16

Figure ‎1.21: Biosynthesis of sorocenol B (37) and related natural products .................. 17

Figure ‎1.22: Biosynthesis of morusalbanol A (42) ......................................................... 17

Figure ‎1.23: Examples of mulberry Diels-Alder adducts with hypotensive property .... 19

Figure 1.24: Examples of mulberry Diels-Alder adducts with promising inhibition against

malondialdehyde (mda) ................................................................................................... 22

Figure 1.25: The 50% cytotoxicity concentration of sanggenol M (59) and sanggenon C

(45) against human oral squamous cell carcinoma (HSC-2) and human salivary gland

tumuor (HSG) ................................................................................................................. 24

Figure 1.26: Structure of (+)_morusalbanol A (42) and sorocein B (62)… ................... 26

Figure ‎3.1: Morusalbanol A (42) and related mulberry Diels-Alder adducts ................. 58

Figure ‎3.2: Chemical structure for CAS 441772-64-4 (185) .......................................... 63

Figure ‎3.3: Key NOESY correlations leading to relative stereochemistry assignment of 183

......................................................................................................................................... 64

x

Figure ‎3.4: Semisynthesis of wittiorumin F (8) and mulberrofuran F (9) from

chalcomoracin C (6) ........................................................................................................ 65

Figure ‎3.5: Energy profile of transition state structure of (a) 179 (b) 180 (c) 188 (d) 189

where black dashed lines denoted the bond formation that led to cyclisation and blue

dashed lines denoted the intermolecular hydrogen bonding ........................................... 69

Figure ‎3.6: Calculated reaction path for a) 187 and b) 186 ............................................ 71

Figure ‎3.7: Potential energy scans on dihedral angle together the respective energy barrier

and distance for (a) endo-181 and (b) exo-182. The distances between the carbon and

oxygen atom in red circle were calculated in the plot ..................................................... 72

Figure ‎3.8: Morusalbanol A (42) ................................................................................... 74

Figure ‎3.9: Comparison of the 1H NMR spectra for endo-204 at a) day 1 b) day 14 ..... 80

Figure ‎3.10: Comparison of the 13

C NMR spectra for endo-204 at a) day 1 b) day 14 .. 81

Figure ‎3.11: NOESY correlation of moruslabanol A pentamethyl ether 206 ................. 84

Figure ‎4.1: Sorocein B (62) ............................................................................................ 88

xi

LIST OF SCHEMES

Scheme ‎1.1: The general Diels-Alder reaction of mulberry Diels-Alder adducts ............ 2

Scheme ‎1.2: Classification of mulberry Diels-Alder adducts ........................................... 5

Scheme ‎1.3: Establishment of absolute configuration of mulberrofuran C (24) via

mulberrofuran G (29) and aromatized compound 32 ...................................................... 12

Scheme ‎2.1: Semisynthesis of kuwanons G (45) and H (13) ......................................... 29

Scheme ‎2.2: Pyrolysis experiment of chalcomoracin (47).............................................. 30

Scheme ‎2.3: Semisynthesis of mulberrofuran C (24) and chalcomoracin (6) ................ 31

Scheme ‎2.4: Retrosynthesis of kuwanon V pentamethyl ether (73) and dorsterone

pentamethyl ether (74) via a Diels-Alder reaction .......................................................... 32

Scheme ‎2.5: Synthesis of diene 75 .................................................................................. 33

Scheme ‎2.6: Synthesis of dienophile 75 ......................................................................... 34

Scheme ‎2.7: Synthesis of kuwanon V (73) and dorsterone (74) methyl ether ................ 35

Scheme ‎2.8: The coupling constants of the cyclohexene ring of 72 and 73 ................... 35

Scheme ‎2.9: Retrosynthetic analysis of methyl ether mulberrofuran C (85) and methyl

ether chalcomoracin (69)............................................................................ 36

Scheme ‎2.10: Synthesis of diene 87 ................................................................................ 37

Scheme ‎2.11: Synthesis of dienophile 89 ....................................................................... 38

Scheme ‎2.12: Synthesis of mulberrofuran C hexamethyl ether (101), mulberrofuran C

heptamethylether (86), mulberrofuran J hexamethyl ether (102) ............ 39

Scheme ‎2.13: Synthesis of dien Synthesis of chalcomoracin hexamethyl ether (103),

chalcomoracin heptamethyl ether (70), mongolicin F hexamethyl ether

(104)………………………………………………………….................40

Scheme ‎2.14: Retrosynthetic scheme of kuwanon J heptamethyl ether (105) ................ 41

Scheme ‎2.15: Retrosynthetic analysis for sorocenol B (37) ........................................... 42

xii

Scheme ‎2.16: Synthesis of dienophile 110……………………………………………..43

Scheme ‎2.17: Synthesis of diene 111 .............................................................................. 43

Scheme ‎2.18: AgNp’s catalyzed Diels-Alder cycloaddition of 110 and 111 ................. 44

Scheme ‎2.19: Pd(II)-catalyzed oxidation cyclization 107 and 121 ................................. 45

Scheme ‎2.20: Key NOE’s leading to relative stereochemistry assignments 107 and 121...

......................................................................................................................................... 46

Scheme ‎2.21: Pd(II)-catalyzed Oxidation cyclization 109 ............................................. 46

Scheme ‎2.22: Synthesis of sorocenol B (106) ................................................................ 47

Scheme ‎2.23: Biomimetic synthetic design for Brosimone A (121) and B (122) .......... 48

Scheme ‎2.24: Synthesis of Brosimone B (122) .............................................................. 50

xiii

LIST OF TABLES

Table ‎1.1: The optical rotations [α]D of mulberry Diels-Alder adducts isolated from nature

......................................................................................................................................... 14

Table ‎1.2: The cytotoxicity of mulberrofuran F1 (62), mulberrofuran F (9), chalcomoracin

(6) .......................................................................................... …………………………25

Table ‎2.1: Development of the initial methodology employing a model prenylchalcone 49

Table ‎3.1: Development of the initial methodology employing a model and dienophile 62

Table ‎3.2: Crystal data and structure refinement for 184 and 186 .................................. 67

Table ‎3.3: Crystal data and structure refinement for 206 and 207 .................................. 85

Table ‎3.4: 1H and

13C NMR spectra of morusalbanol A pentamethyl ether (206) ......... 86

Table ‎3.5: 1H and

13C NMR spectra of compound 207................................................... 87

Table ‎4.1: Pd-catalysed coupling-condensation of o-iodophenol 212 and 2-methyl-3-buten-

2-ol 82 ............................................................................................................................. 94

Table ‎4.2: Pd-catalysed condensation of o-halophenols 212 and 2-methyl-3-buten-2-ol 82

......................................................................................................................................... 96

Table ‎4.3: Attempt to remove the OMe groups of endo-230 ........................................ 105

xiv

LIST OF SYMBOLS AND ABBREVIATIONS

(o-toly)3-P Tri(o-tolyl)phosphine

∆G Gibbs Free Energy of Reaction

∆G* Activation Energy

A2780 Human Ovaries Carcinoma Cell Line

A549 Human Lung Carcinoma Cell Line

AcCl Acetyl chloride

AgBF4 Silver tetrafluoroborate

AgNP Silver nanoparticle

AgOTf Silver trifluromethnesulfonate

AlBr3 Aluminium bromide

BCl3 Boron trichloride

Bel-7402 Human Liver Carcinoma Cell Line

BF3.Et2O Boron trifluoride Diethyl Etherate

BGC-823 Human Stomach Carcinoma Cell Line

BuNBH4 Tetrabutylammonium

CD Circular Dichroism

CH2Cl2 Dichloromethane

xv

COX Cyclooxygenase

DDQ 2,3-dichloro-5,6-dicyanobenzoquinone

DHDA Dehydrogenative Diels-Alder

DMF N, N-Dimethylformamide

DMSO Dimethyl sulfoxide

EOM Ethoxymethyl

EOM-Cl Ethoxymethyl chloride

Et3N Triethylamine

EtOAc Ethyl acetate

EtOH Ethanol

FeCl3 Iron (III) chloride

Ga(OTf)3 Gallium (III) triflate

H2O Water

HCl Hydrochloric acid

HCT-8 Human Colon Carcinoma Cell Line

HIF-1 Hypoxia-Inducible Factor-1

HMBC Heteronuclear Multiple Bond Correlation

HPLC High Performance Liquid Chromatography

HRMS High Resolution Mass Spectroscopy

xvi

HSC-2 Human Oral Squamous Cell Carcinoma

HSG Human Salivary Gland Tumour

I2 Iodine

ICI Iodine monochloride

K2CO3 Potassium carbonate

KIO3 Potassium iodate

KOAc Potassium acetate

LDA Local Density Approximation

LPS Lipopolysaccharide

LST Linear Synchronous Transit

mda Malondialdehyde

MDA Mulberry Diels-Alder

MeOH Methanol

MgI2 Magnesium iodide

MOM Methoxymethyl

MOMCl Methoxymethyl chloride

MTT (3-(4,5-dimethylthiazol-2-yl)-2,5-diphenyltetrazolium bromide

Na2CO3 Sodium carbonate

NaH Sodium hydride

xvii

NBS N-bromosuccinimide

n-Bu4NCl n-Tetrabutyammonium chloride

n-BuLi n-Butyllithium

NMR Nuclear Magnetic Resonance

NOESY Nuclear Overhauser Effect

Pd(dba)2) Tris[dibenzylideneacetone]dipalladium(0)

Pd(OAc)2 Palladium (II) acetate

Pd(PPh3)4 Tetrakis(triphenylphospine)palladium(0)

Pd(TFA)2 Palladium (II) trifluoroacetate

PdCl2 Palladium (II) chloride

Pt/C Platinum on activated carbon

QST Quadratic Synchronous Transit

SiO2 Silicon dioxide

TCM Traditional Chinese Medicine

THF Tetrahydrofuran

TiCl4 Titanium tetrachloride

TLC Thin Layer Chromatography

TMEDA Tetramethylethylene diamine

TMSI-quinoline 1-Trimethylsilylquinolinium iodide

xviii

TPP Tetraphenylporphyrin

UV Ultraviolet

ZnI2 Zinc iodide

1

CHAPTER 1

INTRODUCTION AND LITERATURE REVIEW

1.1 Introduction

Moraceous plants (family MORACEAE) are a rich source of mulberry Diels-Alder

(MDA) adducts. Moraceae comprises of a large family of sixty genera and nearly 1400

species. Some important genera such as Artocarpus, Ficus and Morus, are found in

temperate and subtropical regions of the world (Venkataraman, 1972). The mulberry tree

has been cultivated for thousands of year and play a key role in agriculture in China,

particularly in providing leaves as indispensable food for silk worms.

The different parts of the mulberry tree, such as root bark, twigs, leaves, and fruits,

have also been commonly used in Traditional Chinese Medicine. Besides Morus alba, the

other species of Morus, such as M. australis and M. cathayana were also used in folk

medicine in China. In the early 1970s, investigation into chemical constituents of the

mulberry tree’s (Morus alba) root bark, which is called “Sang-Bai-Pi” in Traditional

Chinese Medical, was initiated. Many compounds with unique structures were isolated,

such as Diels-Alder type adducts (so-called Mulberry Diels-Alder adducts), benzofurans,

stilbenes and flavonoids (Yang et al. 2014).

Many of mulberry Diels-Alder adducts are biologically active and exhibit a wide

range of pharmacological activities including antibacterial, antioxidant, anti-inflammatory,

anticancer, cyclic AMP inhibition, tyrosinase inhibition, and hypoxia-inducible factor-1

(HIF-1) inhibition (Gan et al. 2010; Nomura, 1988a; Yen, Wu, & Duh, 1996). These

discoveries have stimulated significant research interest into their biogenesis and chemical

synthesis.

2

1.2 Mulberry Diels-Alder adducts

Nomura and co-workers were pioneers in the studies of Mulberry-Diels-Alder

adducts from mulberry tree (Nomura, 1988a). They isolated prenylated flavoids which

displayed an interesting pattern for the isoprenyl group such as, kuwanon A (1), kuwanon B

(2) and kuwanon C (3) (Figure 1.1). Generally, Mulberry Diels-Alder adducts are

trisubstituted methylcyclohexenes derived from intermolecular [4+2] cycloaddition (Diels-

Alder) reaction between a dehydroprenylphenol as the diene and the α,β-unsaturated bond

of a chalcones, as the dienophile (Scheme 1.1).

OHO

OH O

Me

Me

OHO

Me

Me

kuwanon A (1)

OHO

OH O

Me

Me

OHO

kuwanon B (2)

Me

Me

OHO

OH O

Me

Me

OHHO

kuwanon C (3)

MeMe

Figure 1.1. Flavonoids isolated from mulberry tree

OOH

+

Me

Me

O

HO

cis,trans trans-trans

OH

HO

chalcone (dienophile)

dehydroprenylphenol(diene)

+Me

Me

HO

prenylphenol

Me

O

HO

OH

Scheme 1.1. The general Diels-Alder reaction of mulberry Diels-Alder adducts

3

OHOOH

OH Me

OOH

HO

Me

Me OH

OH

+

morachalcone A (4) dehydroprenylmoracin C (5)

Me

OOH

OH

O

HOOH

MeMe

HO

OH

HO

3'' 4''

5''SR

S

[4+2]

intramolecular cyclisation/ketalisation

Me

O

OH

O

O

OHHO

Me Me

OH

OH

HO wittiorumin F (8)

3''4''

5''SR

S+

O O

Me

OH

OH

OH

OH

H

Me

Me

O

HO

SR S

R

3''

4''5''

mulberrofuran F (9)

chalcomoracin (6)(cis-trans, endo)

+

mongolicin F (7) (trans-trans, exo)

Me

OOH

OH

O

HOOH

MeMe

HO

OH

HO

Figure 1.2. Biogenetic synthesis of MDA and cylised MDA adducts

Adducts bearing cis-trans stereochemistry are derived from an endo addition

whereas the trans-trans isomer arises from the exo addition. For example, the reaction of

morachalcone A (4) (as dienophile) and dehydroprenyl moracin C (5) (as diene) in nature

produces a pair of cis-trans and trans-trans adducts, chalcomoracin (6) and mongolicin F

(7), respectively (Figure 1.2). Interestingly, the cis-trans MDA adducts chalcomoracin (6)

underwent acid catalysed regio- and stereo-selective intramolecular cyclization/ketalization

4

in nature to give two different metabolites containing an intriguing oxacyclic core structure

(e.g. wittiorumin F (5) and mulberrofuran F (6) (Figure 1.2)).

According to a report by Nomura and co-workers, the mulberry Diels-Alder adducts

may be classified into four groups on the basis of the phenol nuclei as follows: (a)

dehydroprenylchalcone adducts; (b) dehydroprenylflavonoid adducts; (c)

dehydroprenylstilbene adducts; (d) dehydroprenyl-2-arylbenzofuran adducts (Scheme 1.2)

(Nomura & Hano, 1994).

Amongst the four types of mulberry Diels-Alder adducts listed in Scheme 1.2, the

dehydroprenylchalcone type Diels-Alder type adducts are rare. Kuwanon J (10), kuwanon I

(11) and guangsangon C (12) are some. Examples of these dehydroprenylchalcone type

Diels-Alder adducts as kuwanon J (10), kuwanon I (11) guangsangon C (12) (Figure 1.3).

Kuwanon J (10) and kuwanon I (11) were isolated from the root bark of Morus macroura

and Morus alba (Dai et al. 2004; Nomura & Fukai, 1981), respectively. Guangsangon C

(12), was isolated from the stem bark of Morus macrour (Dai et al. 2004)

5

Scheme 1.2. Classification of mulberry Diesl-Alder adducts

6

OH

OH

Me

Me

O

Me

OH

HO

O OH

OH

HO

HO

3"

4"5"

OH

OH

Me

Me

O

Me

OH

HO

O OH

OH

HO

HO

3"

4"5"

kuwanon J (10) kuwanon I (11)

OH

OH

O

MeHO

OH

OH

HO

HO

3"

4"5"

OH

O

5

3

1

2'

4'

guangsangon C (12)

Figure 1.3. Examples of dehydroprenylchalcone adduct

Dehydroprenylflavonoid Diels-Alder adducts (Scheme 1.2) are abundantly found in

nature. Kuwanon H (13), wittiorumin A (14), wittiorumin B (15), wittiorumin C (16) (Tan

et al., 2009), guangsangon D (17), guangsangon K (18) (Dai et al. 2004) and sanggenon G

(19) (Rollinger et al. 2006) are some of the examples of this type of Diels-Alder adducts

(Figure 1.4).

7

O

OOH

Me

Me

HO OH

Me

O

HO

OH

Me

Me

OH

HO

OHO

OH O

OH

R1 OH

Me

O OH

OH

R2

HO

OH

3''

kuwanon H (13) wittiorumin A (14) R1=R2, 3"Rwittiorumin B (15) R1=OH, R2= prenyl, 3"Rwittiorumin C (16) R1=H, R2= prenyl, 3"S

OHO

O

HOOH

Me

O OH

OH

HO

OH

OHO

O

Me

O OH

OH

HO

OH

OHOH OH

guangsangon D (17) guangsangon K (18)

OH

OH

OMe

Me

HO

OH

OH

O

O

HO

HO OH

sanggenon G (19)

Figure 1.4. Examples of dehydroprenylflavonoid adduct

There are not as many examples of these compounds in the third class of Diels-

Alder adduct, dehydroprenylstilbene. Kuwanon X (20) was isolated from the root bark of

Morus Ihou in Japan and Morus macroura in China. Kuwanon Y (21) was isolated from

8

Morus alba and Morus macroura (Rama Rao et al. 1983). Kuwanon P (22) and

guangsangon B (23) were isolated from Morus Ihou and Morus macroura, respectively

(Dai et al. 2004, (Hano, Tsubura & Nomura, 1986) (Figure 1.5).

O

HO OH

OH

HO

OH

OH

Me

OH

3"

4"

5"

R

OH5

33'

5'

O

HO OH

OH

OH

Me

OH

3"

4"

5"5

3R

HO

OH

kuwanon X (20): 3”α kuwanon P (22): R= OH

kuwanon Y (21): 3”β guangsangon B (23): R= H

Figure 1.5. Examples of dehydroprenylstilbene adducts (Hirakura et al. 1985, Dai et al. 2004)

The fourth class of mulberry Diels-Alder adduct isolated from the genus Morus is

dehydroprenyl-2-arylbenzofuran type. Mulberrofuran C (24), calcomoracin (6),

mulberrofuran J (25), mongolicin F (7) and mulberrofuran U (26) are examples of this type

of Diels-Alder adducts (Basnet et a.l 1993; Nomura et al. 1982; Tan et al. 2009) (Figure

1.6).

9

O

HO OH

OH

O

HO

HO

OH

Me

OH

R

O

HO OH

OH

O

HO

HO

OH

Me

OH

R

O

HO OH

OH

OHO

OH

OH

Me

OH

MeMe

mulberrofuran C (24): R = H mulberrofuran J (25): chalcomoracin (6): R = prenyl mongolicin F (7): R = prenyl

mulberrofuran U (26) Figure 1.6. Examples of prenylarylbenzofuran type adduct (Basnet et al. 1993; Nomura et al. 1982;

Tan et al. 2009)

1.3 Absolute configurations of mulberry Diels-Alder adducts

The absolute configurations of C3”, C4” and C5” in the methylcyclohexene ring of

mulberry Diels-Alder adducts were systematically investigated by the Nomura group. In

general, the stereochemistry and absolute configuration of mulberry Diels-Alder adducts

have been confirmed by two methods (a) circular dichroism (CD) spectroscopic studies and

(b) X-ray analysis (Nomura, 1988b).

The magnitude of ∆ε values in the circular dichroism (CD) spectra of

mulberrofurans C (24) and J (25) is known to be larger than any other mulberry Diels-Alder

adducts (Figure 1.7) (Hano et al. 1988). Mulberrofurans C and J exhibited strong split

Cotton effect in the region of 280-350 nm in the UV for both of the compounds. It has

been suggested that this effect may originated from exciton coupling between 2,4-

dihydroxybenzoyl and 2-arylbenzofuran chromophores (Hano et al. 1988). In order to test

10

this hypothesis, reduction of compounds 24 and 25 were carried out using LiAlH4 to give

dihydromulberrofruan C and J 27 and 28, respectively (Figure 1.9). In the CD spectra of 27

and 28, the magnitude of ∆ε values decreased remarkably (Figure 1.8) which clearly

indicated a strong split Cotton effects due to the exciton coupling between 2,4-

dihydroxybenzoy and 2-arylbenzofuran chromophores (Hano et al. 1988). Additionally, in

the CD spectra of 27 and 28, which are opposite to each other in the region 270-350 nm

(Figure 1.8). which suggested that the stereochemistries of 24 and 25 at the stereogenic

center C3´´ bearing the 2-arylbenzofuran chromophore are antipodal to each other. Since

both 27 and 28 exhibit a positive cotton effect, the absolute configuration of mulberrofuran

C (24) and J (25) were established as 3”S, 4”R, 5”S and 3”R, 4”S, 5”R as shown in Figure

1.8 (Hano et al. 1988).

Me

O

HO OH

OH

OH

OH

OHO

HO

S

R

S3" 5"

mulberrofuran C (24)

Me

O

HO OH

OH

OH

OH

OHO

HO

R

S

R3" 5"

mulberrofuran J (25)

24

25

Figure 1.7. The circular dichroism spectra of mulberrofuran C (24) and J (25) and their absolute configurations (Hano et al. 1988)

11

Me

O

HO OH

OH

OH

OH

OHO

HO

3"

27

28

LiAlH4

Me

HO

HO OH

OH

OH

OH

O

HO

3"

OH H

mulberrofuran C (24): 3”α dihydromulberrofuran C (27): 3”α

mulberrofuran J (25): 3”β dihydromulberrofuran J (28): 3”β

Figure 1.8. Reduction of mulberrofuran C (24) and mulberrofuran J (25) and the CD spectra of

dihydromulberrofuran C and J (Hano et al. 1988)

Another method to determine the absolute configuration of mulberry Diels-Alder

adducts is through X-ray crystallographic analysis. The absolute configuration of

mulberrofuran G pentamethyl ether (30) has been determined by X-ray crystallographic

analysis (Rama Rao et al. 1983) where the relative configuration of the stereogenic center

at C-8” was determined as shown in Figure 1.9. (Rama Rao et al. 1983). Absolute

configuration of mulberrofuran C (24) was confirmed from the following procedure.

Conversion of 24 to mulberrofuran G (29) was achieved under acidic conditions as

described in Scheme 1.3 followed by methylation of 29 to give the mulberrofuran G

pentamethyl ether (30). Crystal structure of the ether 30 was confirmed by X-ray

crystallographic analysis. Treatment of 30 with N-bromosuccinimide (NBS) gave

bromomulberrofuran G pentamentyl ether (31) which was then converted to an aromatic

12

compound 32 through dehydrogenation by 2,3-dichloro-5,6-dicyanobenzoquinone DDQ

(Scheme 1.3). The X-ray crystallographic analysis of 32 showed the absolute configuration

at C-8” to be R. As the correlation between 32 and 24 through 30 was confirmed, the

absolute configuration of mulberrofuran C (24) was determined to be 3”S, 4”R, 5”S (Hano

et al. 1988).

O

HO OH

OH

O

HO

HO

OH

Me

1.5% H2SO4

EtOH

55%

O O

OR1

O

R1O

OR1

Me

OR1

H

H H

OH

3" 5"

4"

3" 20"

17"

13" 11"

OR1R2

8"

6'

mulberrofuran C (24) CH3Imulberrofuran G (29)

30 R1= Me, R2= HNBS

31 R1= Me, R2= Br

O O

OMe

O

MeO

OMe

Me

OMe

3"5" 20"

17"

13" 11"

OMe

8"

6' BrDDQ

32

5"

Scheme 1.3. Establishment of absolute configuration of mulberrofuran C (24) via mulberrofuran G (29)

and aromatized compound 32 (Hano et al. 1988)

13

Figure 1.9. The X-ray analysis of mulberrofuran G pentamethyl ether (30) (Rama Rao et al. 1983)

Optical rotations of some mulberry Diels-Alder adducts isolated from Morus

species are summarized in Table 1.1. The relationships of the absolute configuration,

optical rotation and CD spectrum, provided a more convenient way to classify mulberry

Diels-Alder-type adducts. In case of a negative optical rotation or negative Cotton effects, a

maximum UV absorption will be observed in the CD spectrum, and the relative

configuration of a mulberry Diels-Alder adduct will be trans-trans. Otherwise, if both the

optical rotation and Cotton effect shown in the CD spectrum were positive, then the relative

configuration of a mulberry Diels-Alder adduct will be cis-trans (Hano et al. 1998).

Configuration at the C-3´´ stereogenic center will influence the sign of optical rotation.

Therefore, absolute configurations of the three stereogenic centers (3´´, 4´´, 5´´) in the

methylcyclohexene ring of a cis-trans mulberry Diels-Alder adduct may be specified as 3”S,

4”R, 5”S, and those of the trans-trans adducts as 3´´R, 4´´S, 5´´R.

14

TABLE 1.1. The optical rotations [α]D of some mulberry Diels-Alder adducts isolated from nature

Name

Relative

configuration

Optical

rotation [α]D Name

Relative

configuration

Optical

rotation [α]D australisin C cis-trans +340 (MeOH) mongolicn F trans-trans -283 (MeOH)

australisin C cis-trans +340 (MeOH) mongolicn F trans-trans -283 (MeOH)

cathayanon A cis-trans -194 (MeOH) moracenin D trans-trans -388 (MeOH)

chacomoracin cis-trans +194 (acetone) mulberrofuran J trans-trans -341 (MeOH)

guangsangon E cis-trans +140 (MeOH) mulberrofuran J trans-trans -341 (MeOH)

mulberrofuran C cis-trans +153 (MeOH) sanggenon D trans-trans -145 (MeOH)

mulberrofuran E cis-trans +302 (MeOH) sanggenon E trans-trans -86 (MeOH)

mulberrofuran O cis-trans +196 (MeOH) sanggenon M trans-trans -126 (MeCN)

mulberrofuran T cis-trans +139 (MeOH) sanggenon T trans-trans -194 (EtOH)

sanggenon C cis-trans +304 (MeOH) wittionrumin A trans-trans -415 (MeOH)

sanggenon J cis-trans +98 (MeOH) wittionrumin B trans-trans -443 (MeOH)

sanggenon O cis-trans -64 (MeOH) wittiorumin D trans-trans -575 (MeOH)

wittionrumin C cis-trans +420 (MeOH) wittiorumin E trans-trans -296 (MeOH)

wittiorumin G cis-trans +87 (MeOH) yunanensin B trans-trans -183 (MeOH)

albafuran C trans-trans -302 (MeOH) yunanensin B trans-trans -183 (MeOH)

cathayanon B trans-trans -734 (MeOH) yunanensin C trans-trans +439 (MeOH)

guangsangon A trans-trans -409 (MeOH) albanol B a +118 (CHCl3)

guangsangon D trans-trans -108 (MeOH) australisine A a +523 (MeOH)

guangsangon F trans-trans -112 (MeOH) australisine B a +191 (MeOH)

guangsangon G trans-trans -469 (MeOH) cathayanon C a +11 (MeOH)

guangsangon H trans-trans -128 (MeOH) mongolicn A a +650 (MeOH)

guangsangon I trans-trans -471 (MeOH) mongolicn C a +160 (MeOH)

guangsangon J trans-trans -420 (MeOH) mulberrofuran F a +412 (MeOH)

guangsangon K trans-trans -179 (MeOH) mulberrofuran G a -546 (MeOH)

guangsangon M trans-trans -277 (MeOH) mulberrofuran K a +425 (MeOH)

guangsangon N trans-trans -335 (MeOH) wittiorumin F a +299 (MeOH)

kuwanon G trans-trans -534 (MeOH) yunanensin A a +12 (MeOH)

kuwanon H trans-trans -536 (MeOH) yunanensin D a +161 (MeOH)

kuwanon L trans-trans -227 (MeOH) yunanensin E a +675 (MeOH)

kuwanon O trans-trans -243 (MeOH) mulberrofuran Q a +182 (EtOH)

mongolicn D trans-trans -227 (MeOH) cathayanon C a +11 (MeOH)

a non cis-trans or trans-trans

15

1.4 Biosynthesis of mulberry Diels-Alder adducts

Morus alba callues tissues exhibit a high productivity of mulberry Diels-Alder

adducts. The production of kuwanon J (10) and chalcomoracin (6) as the major secondary

metabolites in Morus alba cell cultures was about 100-1000 times more than that of the

intact plant. Therefore, biosynthesis of mulberry Diels-Alder adducts was studied with the

aid of Morus alba cell cultures (Ueda et al. 1982).

The biosynthesis of the mulberry Diels-Alder adducts was conducted by feeding

experiments of methyl chalcones 33 in Morus alba cell cultures. Chalcone 33 was

administered into cell cultures and this resulted in the isolation of prenyl chalcone 34,

kuwanon J methyl ether 36 and chalcomoracin methyl ether (35) (Figure 1.20). The

formation of 34 from 33 in the cell cultures indicated prenylation occurred first before the

formation of 35 and 36.

16

OOH

HO

OH

OMe

methyl chalcone 33

OOH

HO

OH

OMe

Me

Me

prenyl chalcone 34

M. albacell cultures

M. albacell cultures

OOH

HO

OH

OMe

Me

Me

prenyl chalcone 34

OOH

HO

OH

OMeMe

HO

MeO

O

OH

OH Me

Me

OH

HOMe

HO

MeO

O

OH

OH Me

Me

methyl chalcomoracin 35

methyl kuwanon J 36

O

OH

Figure 1.20. Feeding experiments with chalcone 33 and prenyl chalcone 34 to the Morus alba cell cultures

Biosynthesis pathway of cyclized MDA such as morusalbanol A (42), sorocenol B

(37) (Hano, et al. 1995), mulberrofuran I (39) (Hano et al. 1984), australisin B (40) (Zhang

et al. 2007) and mongolicin C (7) (Kang, Chen, & Yu, 2006) consists of intramolecular

cyclization or oxidative cyclization of their corresponding mulbery Diels-Alder adduct

precursors (Figure 1.21, Figure 1.22).

17

O

OH

OH

OMe

MeMe

OHHO

OOH

OMe

MeMe

OHHO

O

HO

OH

Me

OH

HO

OHO

R

HO

OH

HO

O

OH

[O]

MeO

HO

R

HO

OH

HO

O

OH

-H2O

O

OH

MeO

HO

HO

OH

O

OH

O

OH

sorocenol B (37)

mulberrofuran I (39)R= prenyl, australisin B (40)R= H, mongolicin C (7)

R= prenyl, chalcomoracin (6)R= H, mulberrofuran C (24)

38

[o]

Figure 1.21. Biosynthesis of sorocenol B (37) and related natural products

Me

OHHO

OOH

HO

OH

HO

OHO

OCH3

Me

OH

HO

O

HO

HO

O

HO

OH

O

OCH3

41 42

Figure 1.22. Biosynthesis of morusalbanol A (42)

18

1.5 Biological activities of mulberry Diels-Alder type adducts

Mulberry tree is a highly valuable plant and has been widely used as important

sources in Traditional Chinese Medicine (TCM) (Nomura, 1988b). Studies of phenolic

constituents of Morus root bark were originally undertaken to characterize the components

of the root bark responsible for hypotensive activity. Kuwanon G (43), kuwanon H (13),

kuwanon M (44), mulberrofuran C (24), mulberrofuran F (9), and mulberrofuran G (29)

were shown to have hypotensive properties (Figure 1.23). Compounds 9, 13, 24, 29 and 43

also showed an almost equal transient decrease in arterial blood pressure in doses of 0.1-1

mg/kg in rabbits (Nomura & Fukai, 1980) while compound 44 showed hypotensive action

in hypertensive rats (2 mg/kg). Sanggenon C (45) and sanggenon D (46) have been

characterised as the hypotensive compounds of the crude medicine “Sang-Bai-Pi”.

Sanggenon C (45) showed at 2 mg/kg to have hypotensive effect in rabbits while

sanggenon D (46) required about 0.5-2.0 mg/kg (Figure 1.23) for the same activity in rats.

The promising anti-hypotensive properties of these mulberry Diels-Alder adducts warrant

further investigation for their biological activities.

19

Me

OH

HO

OOH

HO

HO

OH

O

O

HO OH

Me

Me

RO

Me

HO

O

Me

Me

OHHO

O

Me

OH

O

OMe

Me

OHHO

Me

Me

O

O

HO

Me Me

Me

O

OH

HO

OH

HO

3" HOO

R = H, kuwanon G (43)

Hypotensive effect at 1 mg/kg concentration

R = prenyl, kuwanon H (13)

Hypotensive effect at 1 mg/kg concentration

IC50 17.5 μm (anti inhibitory against tyrosinase)

kuwanon M (44)

Hypotensive effect at 2 mg/kg

concentration in rabbits

sanggenon C (45) : 3”β

Hypotensive effect at 2 mg/kg concentration in rabbits

sanggenon D (46) :3”α

Hypotensive effect at 0.5-2.0 mg/kg concentration in

rats Figure 1.23. Examples of mulberry Diels-Alder adducts with hypotensive property

1.5.1 Antioxidation

Diels-Alder adducts originating from flavanoids and other polyphenols have been

reported to show good antioxidant activities. The antioxidant activities determined by the

amount of malondialdehyde (Mda), a compound produced during microsomal lipid per-

oxidation induced by ferrous-cysteine. Mda was detected using the thiobarbituric acid

(TBA) method, and the inhibition rates of Mda were calculated with vitamin E as a positive

control (Dai et al. 2004; Rollinger et al. 2006; Tan et al. 2009). Most Diels-Alder adducts

show more than 50% inhibitory rates of malondialdehyde (mda) formation at a

20

concentration of 10µM. Dehydroprenylflavonoid adducts such as guangsangon M (47),

guangsangon N (48), guangsangon D (17), guangsangon H (49), and guangsangon K (18)

exhibited good antioxidant activity with inhibition of mda ranging from 77-100% at the

concentration of 10 µM. Dehydroprenylarylbenzofuran adducts such is albafuran C (50),

guangsangon E (51), guangsangon J (52) and guangsangon A (53) showed inhibitor activity

ranging 76 - 91% at 10 µM (Dai et al. 2004). Dehydroprenylstilbene addcuts such as

kuwanon X (20), kuwanon Y (21), guangsangon B (23) and cathayanon D (54) displayed

inhibition mda ranging from 71-100% at 10 µM (Dai et al. 2004), (Zhang et al. 2009).

Dehydroprenylchalcone adduct such as guangsangon C (12) displayed inhibition rate of

mda 82% at 10 µM (Figure 1.22) (Dai et al. 2004).

Wittionrumin A (14), wittionrumin B (15), wittionrumin C (16), wittionrumin D

(57) also displayed potent antioxidant activities with inhibitory rates of 73%, 82%, 82%

and 62.5%, respectively, at a concentration of 10 μM. The relatively high inhibitory rates

indicated compounds 14, 15, and 16 possess better antioxidant activities than vitamin E

(Tan et al. 2009).

21

O

OH

Me OH

OHOH

OH

HO

O

O

HO 2

R1

Me

OH

OHO

HO OH

R

OH

O

HO

HO

R2

3"

guangsangon M (47): 2-β, R1, R2 = H albafuran C (50): 3”- α, R = H

98% of inhibition rate of mda 76% of inhibition rate of mda

guangsangon N (48): 2-α, R1, R2 = H guangsangon E (51): 3”-β, R = prenyl

100% of inhibition rate of mda 88% of inhibition rate of mda

guangsangon D (17): 2-α, R1 = OH, R2 = H guangsangon J (52): 3”-α, R = prenyl

77% of inhibition rate of mda 91% of inhibition rate of mda

guangsangon H (49): 2-α, R1 = OH, guangsangon A (53): 3”-α, R = prenyl

R2 = prenyl 85% of inhibition rate of MDA 93% of inhibition rate of mda

Me

OH

OHO

HO OH

ROH

HO OH

3"

Me

OH

OHO

HO OH

OH

HO

O

HO OH

kuwanon X (20): 3”-α, R = OH guangsangon C (12)

81% of inhibition rate of mda 82% of inhibition rate of mda

kuwanon Y (21): 3”-β, R = OH

71% of inhibition rate of mda

guangsangon B (23): 3”-α, R = H

84% of inhibition rate of mda

22

O O

OH

OHOH

OH

HO

Me

HO

HO

O

O OH

R1

OH

Me

O

OH

R2

OH

HO

OH

3"

cathayanon D (54) wittiorumin A (14): 3”-β, R1 = R2 = H

100% of inhibition rate of mda 73% of inhibition rate of mda

wittiorumin B (15): 3”-β, R1 = OH, R2 = prenyl

82% of inhibition rate of mda

wittiorumin C (16): 3”-α, R1 = H, R2 = prenyl

82% of inhibition rate of mda

HO

O

O

HO

OH

Me

O

OH

R

OH

HO

OH

wittiorumin D (55): R = H

62.5% of inhibition rate of mda

Figure 1.24. Examples of mulberry Diels-Alder adducts with promising inhibition against malondialdehyde

(mda)

1.5.2 Anti-inflammation

Cyclooxygenase (COX) is an enzyme, responsible for the formation of prostanoids

which are important biological mediators. The relief from symptoms of pain and

inflammation can be provided from pharmacological inhibition of COX. There are

three known cyclooxygenase isoenzymes, namely COX-1, COX-2, and COX-3.

23

Sanggenon C (45), sanggenon E (56), and sanggenon O (57) have been reported to

show inhibitions against COX-1 and COX-2 with IC50 values ranging from 10-14 and 40-

50 µM, respectively. It is noteworthy that compounds with prenyl group improved the

hydrophobicity, stability and penetrativity of the cell membranes (Rollinger et al. 2005).

Sanggenon B (58) and sanggenon D (46) showed the inhibition of lipopolysaccharide

(LPS)-induced NO production in RAW 264.7 cells, with IC50 values of 18.3 and 59.3 µM,

respectively (Cheon et al. 2000).

1.5.3 Cytotoxicity

Optically active mulberry Diels-Alder adducts have also been examined for

cytotoxicity properties. Shi and co workers reported that sanggenol M (59) and sanggenol C

(45) displayed potent cytotoxicity effect against human oral squamous cell carcinoma

(HSC-2) (CC50 13.0 μM and 18.0 μM) and human salivary gland tumour (HSG) (CC50 13.0

μM and 23.0 μM) (Shi et al. 2001).

Mulberrofuran F1 (60), mulberrofuran F (9), chalcomoracin (6) and kuwanon J (11)

were evaluated for their cytotoxic activities against five human cancer cell lines [A549

(human lung carcinoma cell line), Bel-7402 (human liver carcinoma cell line), BGC-823

(human stomach carcinoma cell line), HCT-8 (human colon carcinoma cell line), and

A2780 (human ovaries carcinoma cell line)] by means of the (3-(4,5-dimethylthiazol-2-yl)-

2,5-diphenyltetrazolium bromide (MTT) cell viability assay (Zhang et al. 2007). It was

found that most of the Diels-Alder type adducts showed moderate cytotoxicities with IC50

values ranging from 1-10 µM. Ethanol extract of leaves of Morus alba yielded four Diels-

Alder type adducts, mulberrofuran F1 (60), mulberrofuran F (9), chalcomoracin (6) and

24

kuwanon J (11) but only compounds (6, 9 and 60) exhibited moderate cytotoxicity effects

as shown in Table 1.2 (Yang, Wang, & Chen 2010).

O

HO OH

HO

OH OMe

Me

OH

OH Me

Me

OHO

HO

O

O

OOH

HO

OH

MeMe

MeOH

HO

OH

HOO

sanggenon M (59) sanggenon C (45) CC50 = 13.0 μM against HSC-2 cell CC50 = 13.0 μM against HSC-2 cell

CC50 = 13.0 μM against HSG cell CC50 = 23.0 μM against HSG cell

Figure 1.25. The 50% cytotoxicity concentration of sanggenol M (59) and sanggenon C (45) against human

oral squamous cell carcinoma (HSC-2) and human salivary gland tumuor (HSG).

25

Table 1.2. The cytotoxicity of mulberrofuran F1 (60), mulberrofuran F (9), chalcomoracin (6)

Compound

IC50/µmolL-1

A549 Bel-7402 BGC-823 HCT-8 A2780

O O

Me

OH

OH

OH

OH

H

O

HO

SR S

R

3''

4''5''

OH

mulberrofuran F1 (60)

8.18 8.29 8.46 >10 8.35

O O

Me

OH

OH

OH

OH

H

Me

Me

O

HO

SR S

R

3''

4''5''

mulberrofuran F (9)

8.48 8.16 8.45 8.43 1.21

Me

OOH

OH

O

HOOH

MeMe

HO

OH

HO

3'' 4''

5''SR

S

chalcomoracin (6)

1.51 1.76 1.51 2.11 2.36

1.5.4 Antimicrobial and antifungus

Sanggenon B (58), sanggenon D (46), mulberrofuran G (29), and albanol B (61)

showed inhibition to microbial growths, including Escherichia coli, Salmonella typhimurim,

26

Staphylococcus epidermidis, and Staphylococcus aureus, with MIC values in the range of

5-50 µg/ml (Sohn et al. 2004). Chalcomoracin (6) completely inhibited the germination of

spore of Fusarium roseum and Bipolaris leersiae at a concentration of 10-100 µM (Yang et

al. 2010).

1.6 Scope and objectives of this thesis

Traditionally, natural products have played an important role in drug discovery and

have been the basis of most early medicines. However, the bioactive ingredients, often the

secondary metabolites isolated from the nature products are usually obtained in minute

quantity. Total synthesis of bioactive complex natural products is a significant challenge in

synthetic organic chemistry. Thus, our obejectives for this study are:

a) To synthesis two mulberry Diels-Alder adduct, morusalbanol A (42) and

sorocein B (62) (Figure 1.26) via a biomimetic Diels-Alder reaction to enable

access to these compounds for future bioactivity testing.

b) To investigate the intramolecular cyclization of their cis-trans and trans-trans

precursors and understand the reason for failure of the trans-trans mulberry

Diels-Alder to cylized in nature.

Figure 1.26. Structure of (+)-morusalbanol A (42) and (+)-sorocein B (62).

HO

OH

O

O

OH

OH

OHHO

(+)-morusalbanol A (42)

OMeO

O O

OH

OH

OOH

HO

OH

O

(+)-sorocein B (62)

HH

H

27

In this thesis, we described the work carried out to achieve the objectives that we set

out for the research. In the first chapter describes the introduction and literature review.

Chapter 2 introduces the other syntheses related to preparation of mulberry adduct. Chapter

3 discussed the approaches towards morusabanol A (42) synthesis while Chapter 4

describes the approaches towards sorocein B (62) synthesis. Chapter 5 is the conclusion of

the work as well as some suggestions for future work and the experimental work and data is

described in Chapter 6.

28

CHAPTER 2

RELATED SYNTHESIS

This chapter presents the semi-synthesis of mulberry Diels-Alder adducts such as

kuwanon G (43), H (13), mulberrofuran C (24), chalcomoracin (7). A brief review of total

syntheses of mulberry Diels-Alder adducts was also described.

2.1 Semisynthesis of mulberry Diels-Alder adducts

Several studies have been reported on the semi-syntheses or partial synthesis of

Mulberry Diels-Alder adduct. Kuwanon G (43) and kuwanon H (13) were the first two

compounds isolated from Morus plant. Kuwanon G (43) and kuwanon H (13) are regarded

as dehydroprenyl flavonoid type Diels-Alder adducts (Nomura et al. 1981). Treatment of

kuwanon G (43) using dimethylsulphate and potassium carbonate in reluxing acetone gave

ether 63 (Scheme 2.1). Pyrolysis of a solution of 63 in toluene in a pressure tube at 280 oC

gave the trans-chalcone tetramethyl ether 65 and the dehydrokuwanon C tetramethyl ether

67. Dehydrokuwanon C tetramethyl ether 67 is a flavonoid which contains a conjugated

diene and a prenyl group. The presence of a diene was supported by signals in the 1H NMR

spectrum at δ 1.91 (br s, Me-5´´), δ 4.88, 4.96 (br s, CH2-4´´), δ 6.80 (d, J = 7.0 Hz, CH2-1´´)

and δ 7.26 (d, J = 16.0 Hz, CH-2´´). Meanwhile the presence of a prenyl group was

supported by signals at δ 1.44 and 1.61 (br s, Me-4´ and 5´), δ 3.09 (br d, J = 7.0 Hz, CH2-

4´) and δ 5.24 (br t, J = 7.0 Hz, CH-2´). Diels-Alder cycloaddition reaction between 65 and

67 at 160 oC in a sealed tube gaved a racemate mixture of cycloaddition products, 63 and

68 in 60% yield. However, detail 1H NMR data was has not been reported for these

products.

29

Similarly, pyrolysis of kuwanon H methyl ethers (62) at 280 oC gave tetramethyl

morachalcone A (66) and dehydrokuwanon C tetramethyl ether (67) (Nomura et al., 1981).

Diels-Alder cycloaddition reaction between 66 and 67 at 160 oC in a sealed tube gave a

racemate mixture of cycloadducts 64 and 69 in 50% yield (Nomura et al. 1981).

O

O

Me

Me

OMeMeO

Me

O

OMeMeO

OMe

R

MeO

63: R = H 64: R = prenyl

toluene

280 oC 65: R = H, 54%66: R = prenyl

O

OMe

OMe

MeO

OMe

R

+

O

O

Me

Me

OMeMeO

Me

1"

3"

8

3 1'

2'

dehydrokuwanon C tetramethyl ether (67)37% from 63

toluene

160 oC

60% (63 and 68)50% (64 and 69)1:1 O

O

Me

Me

OMeMeO

Me

O

OMeMeO

OMe

R

MeO

63: R = H 64: R =prenyl

O

O

Me

Me

OMeMeO

Me

O

OMeMeO

OMe

R

MeO

+

68: R = H 69: R = prenyl

Scheme 2.1. Semisynthesis of kuwanons G (45) and H (13) (Nomura et al. 1981)

The pyrolysis of chalcomoracin (46) has been reported by Takasugi et. al. as shown

in Scheme 2.2. Methylation of chalcomoracin (6) using dimethylsulphate gave the

chalcomoracin methyl ether (69). Subsequently pyrolysis of 70 provided dehydromoracin C

30

trimethyl ether (71) and morachalcone A tetramethyl ether (72) in 24% and 26% yields,

respectively (Takasugi et al. 1980).

Me

HO

HO

O

OH

OH

Me

Me

HO

HO

O

OH

Me2SO4

K2CO3

acetone, reflux

43%

Me

MeO

MeO

O

OMe

OMe

Me

Me

MeO

MeO

O

OMe

toluene

280 oCMeO

Me

OMe

O

OMe

+

O

OMe

OMe

MeO

Me

Me

24% 26%

chacomoracin (6) chalcomoracin methyl ether 70

71 72

Scheme 2.2. Pyrolysis experiment of chalcomoracin (47) (Takasugi et al. 1980)

The semi-syntheses of mulberrofuran G (29) and wittiorumin F (8) are

demonstrated in Scheme 2.3. Treatment of mulberrofuran C (24) with 1.5% sulphuric acid

in ethanol afforded mulberrofuran G (29) in 55% yield via intramolecular ketalization

(Fukai et al. 1985).

Similarly, treatment of chalcomoracin (6) with 5% TFA in methanol gave (+)-

wittiorumin F (8) in 18.3% yield via an intramolecular cyclization (Tan, Yan, Wang, Chen,

& Yu, 2009). The presence of the oxabicylic core in wittiorumin F (8) was proven by the of

1H NMR signals at δ 1.76 (m, H-6’β), δ 2.09 (m, H-6’α), δ 3.40 (m, H-4”), δ 3.64 (d, J = 2.5

Hz, H-3”) and δ 4.43 (d, J = 10.0 Hz, H-4”). Compared to the 1H NMR spectrum of

31

chalcomoracin (6), there are some significant changes for the chemical shifts of the

methylcyclohexene ring. The olefinics proton which signal at δ 5.30 -5.80 ppm as a broad

signlet was disappeared while the chemical shift of the methyl proton was shifted upfield

from δ 1.70 ppm to δ 1.36 ppm as a singlet (Tan et al. 2009).

Me

O

OHHO

OH

OH OH

OH

O

HO

mulberrofuran C (24) mulberrofuran G (29)

1.5% H2SO4

EtOH, 50 oC

24 h

55%

Me

OH

OHO

HO

O O

OH

OH

H

H H

Me

O

OHHO

Me Me

OH

OH OH

OH

O

HO5% TFAMeOH, 50 oC

24 h

HO OH

O OH

OH

Me

Me

Me

O

OHO

HO

18.3%

chalcomoracin (6) wittiorumin F (8)

1"

4"6"5'

Scheme 2.3. Semisynthesis of mulberrofuran C (24) (Fukai et al. 1985) and chalcomoracin (6) (Tan et al. 2009)

2.2 Total syntheses of mulberry Diels-Alder adducts

2.2.1 Thermal Conditions For Diels-Alder Reaction

Thermally promoted Diels-Alder cycloaddition reaction is the most commonly used

strategy for construction of the cyclohexene skeleton. Since Nomura’s pioneering work on

the synthesis of (±)-kuwanon G octamenthyl ether via thermal Diels-Alder reaction was

reported, a number of methods on the total syntheses of mulberry Diels-Alder adducts have

been reported.

32

In 2011, total synthesis of kuwanon V (73) and dorsterone pentamethyl ethers (74)

were reported by Chee and co-workers as outlined in Scheme 2.4 (Chee et al. 2011).

Kuwanon V pentamethyl ether (73) and dorsterone pentamethyl ether (74) were sythesized

via a [4+2] cycloaddition reaction between a chalcone-type diene 75 and a dienophile 76.

The chalcone precursors were prepared from commercially available acetophenone and

aldehyde via Claisen-Schmidt condesation (Scheme 2.4).

MeO

OOMe

Me

OMe

MeO

O

OMe

OH

Me

Me

MeO

OOMe

Me

OMe

MeO

O

OMe

OH

Me

Me

+

kuwanon V pentamethyl ether (73) dorsterone pentamethyl ether (74)

MeO

OOMe

Me

OMe

MeO

O

OMe

OH

Me

Me

75

76

Scheme 2.4. Retrosynthesis of kuwanon V pentamethyl ether (73) and dorsterone pentamethyl ether (74)

via a Diels-Alder reaction (Chee et al. 2011)

The synthesis of diene 75 was accomplished in five steps (Scheme 2.5). First,

acetophenone 77 was iodinated with ICl/CH2Cl2 to give the C-3 iodinated acetophenone 78

in 40% yield. Methylation of aryl iodide 78 followed by Claisen-Schmidt condensation

with 4-methoxybenzaldehyde led to the formation of chalcone 81 in 85% yield. Subjecting

chalcone 81 to Heck coupling [Pd(OAc)2, (o-toly)3-P, Et3N, DMF] (Arkoudis et al., 2009)

33

with 2-methylbut-3-en-2-ol (82) followed by dehydration with acetyl chloride/pyridine

gave the diene 75 (Harrington, Hegedus, & McDaniel, 1987).

O Me

OH

OH

O Me

OH

OH

CH2Cl2, r.t40% I

OMe

O H

KOH, EtOH

acetone, rt

Me2SO4, K2CO3

O Me

OMe

OMe

I

rt, 85%

MeO

I

OMe O

OMe MeO

OMe O

OMe

1) (82)

Pd(OAc)2,(o-tolyl)3P,

Et3N, DMF, 90 oC, 18 h

2) AcCl, pyridine

C6H6, 60 oC

95 %

Me

ICl

77 78 79

81 75

Me

Me

OH

80

Scheme 2.5. Synthesis of diene 75

Dienophile 76 was synthesized from chalcone 83 as shown in Scheme 2.6. Selective

demethylation of the 2-methoxychalcone 83 gave chalcone 84 which was subjected to

prenylation leading to the formation of prenyl ether 85. Then, the prenyl ether 85 then was

treated with Montmorillonite K10 to give dienophile 76 in 45% yield via [1,3]-sigmatropic

rearrangement (Sugamoto et al. 2008).

34

MeO

OMe O

OMe MeO

OH O

OMeBCl3, CH2Cl20 oC, 24 h

70%

MeO

OH O

OMe

Me

Me

Montmorillonite K10,

CH2Cl2, 0 oC, 2 hMeO

O O

OMe

prenyl bromide, K2CO3,acetone, reflux, 8 h

Me

Me

85%

45%

83 84

85 76

Scheme 2.6. Synthesis of dienophile 75

Diels-Alder reaction between diene 75 and dienophile 76 in toluene in a sealed tube

at 160oC afforded a mixture of 73 and 74 in 3:2 ratio in 55% yield (Scheme 2.7). The

coupling constant J3’’,4’’ is 6.8 Hz in 73 (cis) and 10 Hz in the 74 (trans) (Scheme 2.8).

Meanwhile, the coupling constant J4’’,5’’ is 10 Hz (trans) in both 73 and 74. Dorsterone

pentamethyl ether 74 shows evidence of atroisomerism as majority of the signals in the 1H-

NMR spectrum were doubled at room temperature. However, the NMR signals were

resolved by measurement in DMSO-d6 at 80 oC (Chee et al. 2011).

35

MeO

OOMe

Me

OMe

MeO

O

OMe

OH

Me

Me

+

kuwanon V pentamethyl ether (73)

dorsterone pentamethyl ether (74)

75

76

toluene

160 oC, 18 h

55%, 73/74 (3:2)

Me

O

HO OMe

MeMe

OMe

O

MeO

OMe

OMe

+

Me

O

HO OMe

MeMe

OMe

O

MeO

OMe

OMe

Scheme 2.7. Synthesis of kuwanon V (73) and dorsterone (74) methyl ether (Chee et al. 2011)

MeH3"

H4"

O

"H5

(6.8Hz)

(9.6Hz)

MeH3"

H4"

O

"H5

(10Hz)

(10Hz)

R

SR

S

SR

NOESY

Me

O

HO OMe

MeMe

OMe

O

MeO

OMe

OMe

dorsterone pentamethyl ether (74)

Me

O

HO OMe

MeMe

OMe

O

MeO

OMe

OMe

kuwanon V pentamethyl ether (73)

Scheme 2.8 The coupling constants of the cyclohexene ring of 72 and 73 (Chee et al. 2011)

36

In 2010, Rizzacasa and co-workers reported the synthesis of methyl ether

derivatives of mulberrofuran C (86) and chalcomoracin (70) via thermal Diels-alder

reaction. Their retrosynthetic analysis involved a biogenesis-inspired intermolecular [4+2]-

cycloaddition reaction between a dehydroprenylbenzofuran 87 and chalcones dienophile 88

and 89, respectively as shown in Scheme 2.9 (Gunawan & Rizzacasa, 2010).

Me

O

OMe

OMe

OMe

R

MeO

OMe

OMeO

MeO

methyl ether mulberrofuran C (86) : R = H methyl ether chalcomoracin (70) : R = prenyl

Diels-Alder

OMe

OMeO

MeO

Me

OMe

OMe

OMeO

I + BO

O

HO

OH

OH

COOH

OHHO

Me

OH

R

MeO

O

MeO OMe

OH O

Me

HO

O H

OH

OH

+

+

HO

R1

R2O

O

R2O

OR2

Ar

Me

88 R = H89 R = prenyl

87

90 91

77 92

92 93

Scheme 2.9. Retrosynthetic analysis of methyl ether mulberrofuran C (85) and methyl ether chalcomoracin

(69) (Gunawan & Rizzacasa, 2010)

37

Synthesis of the dehydroprenylbenzofuran diene 87 is outlined in Scheme 2.10. A

selective (Fürstner, Heilmann, & Davies, 2007) Sonogashira coupling (Sonogashira, Tohda,

& Hagihara, 1975) between the alkyne 94 and iodide 95 using Cs2CO3 (Rathwell, et al.

2009) afforded the alkyne product 96 in 65% yield. Methanolysis of the acetate 96 provided

phenol 97 which was then cyclized by using TBAF (Hiroya et al. 2000) to produce the

arylbenzofuran halide 90 in good yield. Finally, Suzuki-Miyaura (Miyaura & Suzuki, 1995)

(Coleman, Lu, & Modolo, 2007) coupling was performed on iodide 90 with

pinacolboronate 91 to give diene 87 in 82% yield.

MeO OMe

I

I

OAcMeO

65%

OMe

I

OMe

ORMeO

K2CO3

MeOH

96 R = Ac

97 R = H; 78%

TBAF, THFreflux

82%

MeO OI

OMe

OMe

MeO O

OMe

OMe

Me

BO

O

Me

Pd2(dba)3, AsPh3

K3PO4, DMF, 50 oC

82%

95

94

Pd(PPh3)4, CuICs2CO3, DMF, r.t

90 87

91

Scheme 2.10. Synthesis of diene 87 (Gunawan & Rizzacasa, 2010)

As shown in Scheme 2.11, the synthesis of required dienophile 89 began with an

aldol condensation reaction between 98 and 99 (Ahmed, Wagner, & Razaq, 1978) to afford

the chalcone 88 in 90% yield. Chalcone 88 was utilized for the synthesis of chalcomoracin

38

(6) through O-prenylation, followed by [1,3]-rearrangement using Florisil®

to give 89

(Talamás et al. 1997).

O

OH

OMe

+

O H

OMe

OMe

OH

MeO

O OMe

OMe

KOH, EtOH, r.t.

90%

O

MeO

O OMe

OMe

prenylchalcone,K2CO3, acetone,

84%

Me

Me

OH

MeO

O OMe

OMe

florisil, toulene,

100 oC, [1,3]

Me

Me

98 99

88

100 89

65MeIK2CO3

Scheme 2.11. Synthesis of dienophile 89 (Gunawan & Rizzacasa, 2010)

The Diels-Alder reaction between dehydroprenylbenzofuran 87 and chalcone 88 in

toluene (Scheme 2.12) in a sealed tube at 180 oC gave the endo-adduct 101 and exo-adduct

102, respectively in a 1:1 ratio after separation by preparative HPLC. The endo-adduct 101

corresponded to the hexamethyl ether derivative of mulberrofuran C (24) while the exo-

adduct 102 is the hexamethyl ether derivative of mulberrofuran J (25). Methylation of 101

gave mulberrofuran C heptamethyl ether (86) (Gunawan & Rizzacasa, 2010). The isomers

101 and 102 could be distinguished by 1H NMR spectroscopy where the coupling constant

between H3´´ and H4´´ is 5Hz (cis) in the endo-isomer (101) and 10 Hz in the exo-isomer

(102) (Gunawan & Rizzacasa, 2010). Several attempts toward the deprotection of either

mulberrofuran C heptamethyl ether (86) or mulberrofuran C hexamethyl ether (101) to give

mulberrofuran C (24) only led to incomplete demethylation or decomposition of the

compounds (Gunawan & Rizzacasa, 2010).

39

Me

O OMe

OMe

OMeRO

OMeO

MeO

OMe

+

87

toluene,

180 oC

40%, 1:1

MeIK2CO3

R = H (101)mulberrofuran C hexamethylether

86 R = Me; 98%

endo

exo

+

88

102

O

MeO

OH OMe

OMe

O

Me

OMe

OMeMeO

Me

O OMe

OMe

OMeRO

OMeO

MeO

OMe

Scheme 2.12. Synthesis of mulberrofuran C hexamethyl ether (101), mulberrofuran C heptamethyl ether

(86), mulberrofuran J hexamethyl ether (102) (Gunawan & Rizzacasa, 2010)

The synthesis of methyl ether derivatives of chalcomroacin (6) and mongolicin F (7)

is detailed in Scheme 2.13. Cycloaddition reaction between chalcone 89 and diene 87

proceeded at 180 oC to give endo-adduct 103 and exo-adduct 104, respectively in a 1:1 ratio

after separation by preparative HPLC. The endo-adduct 103 corresponded to hexamethyl

ether derivative of chalcomoracin (6) while the exo-adduct 104 is the hexamethyl ether

derivative of mongolicin F (7).

40

Me

O OMe

OMe

OMeHO

OMeO

MeO

OMe

+

toluene, 180 oC

55%, 1:1

MeIK2CO3

103, R = Hchalcomoracin hexamethyl ether

70, R = Me; 56%

exo

MeMe

mongolicin F hexamethyl ether (104)

O

MeO

OH

Me

Me OMe

OMe

+

OMeOMe

OMe

OMe

Me

O OMe

OMe

OMeRO

OMeO

MeO

OMeendo

MeMe

87

89

Scheme 2.13. Synthesis of chalcomoracin hexamethyl ether (103), chalcomoracin heptamethyl ether (70),

mongolicin F hexamethyl ether (104) (Gunawan & Rizzacasa, 2010)

Rizzacasa and co-workers also investigated on the effect of different protecting

groups on the Diels-Alder reaction (Boonsri et al. 2012). They found that fully methylated

chalcone 65 failed to undergo cycloaddition reaction. They reasoned that the presence of a

H-bonded ortho phenol in the chalcone dienophile is essential to the Diels-Alder reaction.

Their density functional theory calculation suggested that their LUMO-lowering effect of

the OH...O=C hydrogen bond and better co-planarity between the diene and its aryl

substituent in the transition state would lower the energy barrier for the Diels-Alder

41

reaction of the 2-hydroxychalcone compared to that of 2’-methoxychalcone (Boonsri et al.

2012).

Dienophile bearing hydroxy group at the ortho position appeared to be very reactive

while in the absent of an ortho hydroxyl group in dienophile, the diene may undergo

intermolecular dimerisation (Boonsri et al. 2012). They furthered demonstrated that Diels-

Alder reaction of diene 106 and dienophile 89 have kuwnaon J heptamethyl ether 105. The

presence of an ortho hydroxyl group in dienophile 89 was essential for this reaction

(Boonsri et al. 2012).

Me

OOMe

MeOO

OMe

OMe

HO OMe

Me Me

OMe

OMe

[4+2]

Me

Me

OO

H

MeO

OMe

OMe

Me

OOMe

MeO

OMe

OMe

+

kuwanon J heptamethyl ether (105)

89

106

Scheme 2.14. Retrosynthetic scheme of kuwanon J heptamethyl ether (105) (Boonsri et al. 2012)

2.2.2 Catalytic Diels-Alder reaction

In 2012, Porco and co-worker reported the first synthesis of (±)-sorocenol B (37)

(Hano, Yamanaka, Nomura, & Momose, 1995) by employing silver nanoparticle (AgNp)-

catalyzed Diels-Alder reaction (H. Cong & Porco, 2012). The AgNP was prepared from a

3:1 ratio of AgBF4/Bu4NBH4 in CH2Cl2 and then coated on silica gel. The solid product

42

was filtered and then calcinated at 220 oC to give AgNP (Cong et al. 2010). Retrosynthetic

analysis for sorocenol B (37) is shown in Scheme 2.15.

The synthesis of the acetylated chalcone 110 started with Claisen-Schmidt

condensation between ketone 112 and aldehyde 114 with NaH in THF (Nishida &

Kawabata, 2006) to afford chalcone 115 in 96 % yield. Hydrolysis of the MOM protecting

group with 3 M aqueous HCl in refluxing methanol, followed by acetylation provided

chalcone 110 in 79 % yield (two steps) (Scheme 2.16) (Cong et al. 2008).

O

OH OOHHO

MeO

OH

O

OH OOMOMMOMO

MeO

OH

O

OH OOMOMMOMO

Me

OH

HO

O

OH O OAc

OAc+

OMOM

OMOM

Me

HO OH

O

OH O

Me

sorocenol B (37) 107

110111

112 113

3''

Scheme 2.15. Retrosynthetic analysis for sorocenol B (37)

108: 3”-Hβ, endo

109: 3”-Hα, exo

43

OMe

OH O

MeMe

OMe

Me

OH O OMOM

OMOM

OMOM

OMOM

O

H

NaH, THF, r.t48 h, 96%

OMe

Me

OH O OAc

OAc

i) 3 M HCl, MeOH

80 oC, 20 min

ii) Ac2O, py, CH2Cl2, r.t, 8 h, 79% (2 steps)

112115

110

114

Scheme 2.16. Synthesis of dienophile 110

Diene 111 was prepared in four steps from resorcinol 113 (Scheme 2.17). Protection of 113

with MOMCl in DMF gave the MOM ether 116. Regioselective formylation of 116 was

carried out with n-BuLi in the presence of TMEDA followed by quenching the resulting

aryl lithium intermediate with DMF to produce the aldehyde 117 in 83% yield (Eleonora

Ballerini et al, 2009). Aldol reaction of aldehyde 117 with acetone gave 118 in 92% yield

(E. Ballerini, Minuti, & Piermatti, 2010). Wittig olefination of the ketone 118 gave the

desired diene 111 in 79% yield (Cong & Porco, 2012).

NaH, MOMClDMF, r.t, 2h

91%

OH

OH

OMOM

OMOM

OMOM

OMOM

CHO

OMOM

OMOM

Me

OOMOM

OMOM

Me

Ph3PCH3BrNaHMDS

THF, 60 oC, 2 h,

79%

i) n-BuLi, TMEDA,

THF, -10oC, 2 h

ii) DMF, 0 oC, 1 h,

83%

15 mol% NaOHacetone/ water

40 oC, 15 h,

92%

113 116 117

118 111

Scheme 2.17. Synthesis of diene 111

44

The silver nanoparticle-catalyzed Diels-Alder cycloaddition of dienophile 110 and

diene 111 is shown in Shceme 2.18. The reaction using 0.1 mol% catalyst loading of silica-

supported silver nanoparticles (AgNp’s) reacted cleanly in air (50 oC) to afford 90%

combined yield of separable endo/exo diastereomers in a 2:1 ratio (Scheme 2.18).

Meanwhile, the same reaction conducted in thermal condition, without AgNp did not

proceed to completion. As a result, only 53 % combined yield of the endo/exo

diastereomers, 119 and 120 in 2:1 ratio were obtained.

Scheme 2.18. AgNp’s catalyzed Diels-Alder cycloaddition of 110 and 111

Next, deacylation of 119 gave 118 (Scheme 2.19). Oxidative cyclization of 118 by

using Pd(OAc)2, pyridine gave the desired bicylcle [3.3.1] product 107 and its epimer 121

in 50% combined yield. (Trend et al. 2003; Trend, Ramtohul, & Stoltz, 2005) (Scheme

2.19). The relative stereochemistry of 107 and 121 were determined by key NOE signals

(Scheme 2.20) (Cong & Porco, 2012).

119: 3”-Hβ, endo

120: 3”-Hα, exo

119:120 = 2:1

45

O

OH OOMOMMOMO

MeO

OH

MeMe

O

OH OOMOMMOMO

MeO

OH

MeMe6

2622

7

72121

26

6 1+

107 121

50% (two steps)107/121 (2:1 ratio)

Me

OMOMMOMO

AcO OAc

OOH

OMe

Me

3"

Me

OMOMMOMO

HO OH

OOH

OMe

Me

3"

Na2CO3, MeOH/H2O

rt, 12 h

119 108

20 % Pd(OAc)2

40% pyridine

1 atm O2

toluene, 80 oC, 24 h

Scheme 2.19. Pd(II)-catalyzed oxidation cyclization 107 and 121

46

O

OH OOMOMMOMO

MeO

OH

Me

MeO

OH OOMOMMOMO

MeO

OH

Me

Me

4

61

2622

7 721 21

26

61

O

OH

OMe

O

HO MeMe

HH

HH

Ar'

H

HH

3%

2%3%

1%

Me

O

HO

H

HH

Ar'

O

O

Me

Me

HO

HH

2%

10%

3%

NOESY

107 121

Scheme 2.20. Key NOE’s leading to relative stereochemistry assignments 107 and 121 (Cong & Porco, 2012)

However, when the same Stoltz’s conditions for oxidative Wacker cyclization were

applied to 109, it failed to give any cyclize product (Scheme 2.21), possibly due to

stereoelectronic restrictions in the process of oxidative cyclization.

Me

OMOMMOMO

AcO OAc

OOH

OMe

Me

3"

Me

OMOMMOMO

HO OH

OOH

OMe

Me

3"

Na2CO3, MeOH/H2O

rt, 12 h

120 109

20 % Pd(OAc)2

40 % pyridine 1 atm O2

toluene, 80 oC,

24 h

X

Scheme 2.21. Pd(II)-catalyzed oxidative cyclization of 109

47

Finally, treatment of 107 with 3M hydrochloric acid in methanol gave sorocenol B

(106) in 74% without epimerization at C-4 position (Scheme 2.22).

O

OH OOMOMMOMO

MeO

OH

MeMe

3 M HCl, MeOH

80 oC, 10 min

74%

O

OH OOHHO

MeO

OH

MeMe

107 106

Scheme 2.22. Synthesis of sorocenol B (106) (H. Cong & Porco, 2012)

Porco and co-workers also developed another biomimetic and dehydrogenative

Diels-Alder (DHDA) cycloaddition for a total synthesis of the MDA brosimone A (121)

(Messana, Ferrari, & De Araujo, 1988) and brosimone B (122) (Qi et al. 2013). Brosimone

A (121) and B (122) were both isolated from the Brazilian plant, Brosimopsis oblongifolia.

These dimeric structures were derived from protected prenyl chalcones 124 (Scheme 2. 23).

Diene 123 was prepared in situ by dehydrogenation of the prenyl chalcones 124.

48

HO

OH

Me

HO OH

O

Me

OH

HO

O

HO

OH

Me

O OH

OH

Me MeOH

HO

OH

HO

O OH

OH

brosimone A (121) brosimone B (122)

OOH

PgO

Me

OPg

OPg

123

OOH

PgO

Me

OPg

OPg

124Me

Pg = Protecting group

Scheme 2.23. Biomimetic synthetic design for Brosimone A (121) and B (122)(Qi et al. 2013)

Model studies were carried out on a prenyl chalcone 125 with number of catalysts

and oxidants. The results are summarized in Table 2.1. After numerous studies, it was

found that the desired dehydrogenative cycloadducts 126 and 127 can be obtained when

prenyl chalcone 125 was exposed to 10 mol% of platinum on activated carbon (Pt/C)

(Kogan & Herskowitz, 2002; Sebastián et al. 2008) and 0.2 mol% silica-supported AgNPs

in an ambient air atmosphere (Table 1, entry 1). Adding hydrogen scavengers such as

cyclopentene (Lee, Kwon, & Yi, 2012) (entry 3) and norbornene (Wang et al. 1996) (entry

6) in the transfer dehydrogenation reaction gave the cycloadducts 126 and 127 in 61% and

44% yields, respectively. Control experiments showed that reactions with cyclopentene did

49

not proceed without Pt/C (entry 4) and were significantly less efficient in the absence of

AgNPs (entry 5).

Table 2.1. Development of the initial methodology employing a model prenylchalcone (Qi et al. 2013)

Entry Condition details[c]

conv.[a]

endo-126/ exo-127[a]

1 air (1 atm), 10 mol% Pt/C, 88% (52%)[b]

52:48

0.2 mol% AgNP, 90 oC, 24 h

2 O2 (1 atm), 5 mol% Pt/C, 79% (60%)[b]

57:43

0.1 mol% AgNP, 90 oC, 36 h

3 9.0 equiv. cyclopentene, 10 mol% Pt/C, 66% (61%)[b]

42:58

0.1 mol% AgNP, 90 oC, 36 h

4 9.0 equiv. cyclopentene, 0.2 mol% AgNP <10% -

Ar, 90 oC, 36 h

5 9 equiv. cyclopentene, 10 mol% Pt/C, 38% (8%)[b]

72:28

Ar, 90 oC, 36 h

6 10 equiv. norbornene, 10 mol% Pt/C, 76% (44%)[b]

50:50

0.2 mol% AgNP, Ar, 90 oC, 36 h

7 10 mol% Pt/C, 0.2 mol% AgNP, 40%(18%)[b]

55:45

Ar, 90 oC, 36 h

[a] Based on 1H NMR integration.

[b] Yield of isolated product show in parentheses.

With the model reaction established, DHDA cycloaddition of 128 using the

optimized Pt/C–AgNP conditions with cyclopentene as H2 scavenger were carried out to

give the cycloadducts exo-130 and endo-129 (1.2:1) in 64% yield (Qi et al. 2013). Transfer

hydrogenolysis of exo-130 was accomplished by employing 1,4-cyclohexadiene as

hydrogen donor to produce brosimone B. Ammonium formate (HCO2NH4) (Iikubo et al.

2002) was used as additive to further accelerate the hydrogenolysis. Methylation of 122

126: 3”-Hβ, endo

127: 3”-Hα, exo

50

afforded derivative 131. The structure of 131 was determined by X-ray crystal-structure

analysis (Scheme 2.24) (Qi et al., 2013).

OH O

BnO

Me Me

Me

BnO

OH O

O OH

OBn

MeMe

OBn

OBn

OBn

OBn

BnO

BnO

Me

HO

OH O

O OH

OH

MeMe

OH

OH

HO

OH

Brosimone B (122)

i

iii

128

ii

An ORTEP drawing for 131, hydrogen atoms were omitted for clarity

3"

Me

MeO

OMe O

O OMe

OMe

MeMe

OMe

OMe

MeO

MeO

131

Scheme 2.24. Synthesis of Brosimone B (122) (Qi et al., 2013). Reagents and conditions: i)Pt/C (10 mol%),

AgNP (0.2 mol%), cyclopentene, 1,2-dichloroethane (DCE), 110 oC, 48 h, 64%, endo-129:exo-130= 1:1.2 ii)

Pd/C (40 mol%), HCO2NH4 (2 equiv), 1,4-cyclohexadiene, acetone, 40 oC, 24 h, 85% iii) K2CO3, Me2SO4,

acetone, 40 oC, 4 h, 15%.

Brosimine A (121), which biosynthetically originated from brosimone B (102), was

prepared via a second intramolecular dehydrogenative Diels-Alder reaction (Scheme 2.25).

Dehydrogenative Diels-Alder (DHDA) cycloaddition of exo-130 in the presence of DDQ-

AgNps in chlorobenzene at 130 oC gave the exo-exo cycloadduct 132 in 62% yield. Finally,

hydrogenolysis of 132 afforded brosimone A (121) in 91% yields. Methylation of 121 gave

133 as single crystal, thus allow stereochemical assignment for the synthetic brosimone A

(Qi et al. 2013).

129: 3”Hβ endo 130: 3”Hα exo

51

Me

BnO

OH O

O OH

OBn

MeMe

OBn

OBn

BnO

BnO

BnO

BnO

OHO

OBn

Me

BnO

OH

OMe

OBn

BnO

MeO

MeO

OMeO

OMe

Me

MeO

OMe

OMe

OMe

MeO

brosimone A (121) exo-exo 132

i ii

iii

exo-130

HO

HO

OHO

OH

Me

HO

OH

OMe

OH

HO

133

An ORTEP drawing for 133, hydrogen atoms were omitted for clarity

Scheme 2.25 Synthesis of brosimone A (121) (Qi et al., 2013). Reagents and conditions: i) DDQ (1.5

equiv), AgNp (0.3 mol%), PhCl, 130 oC, 72 h, 62%, d.r > 20: 1 ii) Pd/C (60 mol%), HCO2NH4 (2 equiv),

HCO2H, 1,4-cyclohexadiene, acetone, 40 oC, 24 h, 91% iii) K2CO3, Me2I, acetone, 40

oC, 12 h, 13%, Bn

= benzyl.

2.2.3 Chiral-Boron-Complex Promoted enantioselective synthesis

In 2014, Lei and Wulff and their co-workers reported the first enatioselective

total syntheses of (–)-kuwanon I (10), (+)-kuwanon J (9), (–)-brosimone A (121), and (–

)-brosimone B (122) (Han et al., 2014). This successful Diels-Alder cycloaddition

reaction was catalyzed by a chiral boron complex, prepared by coordination of an

axially chiral ligand, such as VANOL or VAPOL (Bao, Wulff, & Rheingold, 1993)

(Heller et al. 2006) to boron. The development of model dienophile 135 and diene 134

using this chiral ligand system was shown in the Scheme 2.26. The enantio- and

diastereoselectivities of the cycloaddition were strongly influenced by the ligand

structure (Han et al., 2014). The reaction in the presence of (R)-VAPOL as ligand gave

a mixture of cycloadducts 136 and 137 in 98% combined yield in a 1:1.4 ratio but

52

moderate ee value (48:72 % ee) (Scheme 2.28, entry 3). When ligand (S)-VANOL was

used, the endo-137 was generated in 80% yield and excellent ee value (97:11 % ee)

(Scheme 2.28, entry 4) (Han et al. 2014).

Me

+

O

OAc

OAc

AcO

OH

134

135

Me

AcO

OAc

O

OH

OAc

entry ligand time (h) yield (%) endo/exo ee (%) endo/exo

1 138 20 92 3.2:1 1:0

2 139 20 97 4.0:1 48:21

3 140 23 98 1.4:1 48:72

4 141 20 99 4.2:1 97:11

O

O

Me

Me OH

OH

(R, R)-TADDOL (138) (S)-BINOL (139) (R)-VAPOL (140) (S)-VANOL (141)

OH

OH Ph

Ph

OH

OH Ph

Ph OH

OH

Scheme 2.26. Chiral-boron-complex-promoted asymmetric Diels-Alder cycloaddition

(Han et al. 2014)

Scheme 2.27 shows the mechanism proposed by Lei and co-workers (Han, Jones,

& Lei, 2015) for the enantioselective Diels-Alder cycloaddition. The reaction was

proposed to proceed through the formation of a chiral boron complex 143, followed by

formation of a tetracoordinate boron complex 144 which then underwent a

enantioselective cycloaddition with diene 134. Lei and co-workers suggested that the

136: 3”Hα: exo

137: 3”Hβ: endo

ii) 135, 5 Å MS, r.t. 1.5 h

iii) 134, r.t.

i) Ligand, BH3.THF

AcOH, THF,

r.t., 25 min

53

enantioselective Diels-Alder cycloaddition may be induced by the following factors

(Han et al. 2015).

a) The coordination bond between boron and dienophile 135 which may lower

energy of the LUMO.

b) Mobility of dienophile 135 being reduced upon complexation.

c) The π-π stacking between chiral ligand and dienophile 135 shielding one face of

chalcones to attack by the diene 134.

Scheme 2.27. Proposed mechanism for the enatioselective Diels-Alder cycloaddition

(Han et al. 2014)

This catalyst was then applied to the synthesis of (–)-kuwanon I (10), (+)-

kuwanon J (9), (–)-brosimone A (121), and (–)-brosimone B (122). For the synthesis of

(–)-kuwanon I (10), (+)-kuwanon J (9), prenylation of 145 followed by sigmatropic

rearrangement gave ortho-prenylated chalcone 148 (37%), along with para-prenylated

chalcone 147 (27%). Chalcone 148 was later converted to chalcone 149 in 35% yield

over two steps. Regisoselctive Schenck ene reaction (Helesbeux et al. 2003) with

[Ru(bpy)3Cl2.6H2O] and MeOH of 149 to form secondary allylic alcohol 150 and

π-π stacking

interaction

136: 3”Hα: exo

137: 3”Hβ: endo

54

tertiary allylic alcohol 151 in 62% combined yield. Dehydration of alcohol 151 in the

presence of SOCl2/DBU proceeded smoothly to provide the desired diene 152 in 75%

yield (Scheme 2.28).

MOMO

MOMO

O OH

OMOM

Br

acetone, K2CO3, reflux, 24 h, 95%

MOMO

MOMO

O O

OMOM

montmorillonite K-10

CH2Cl2, 0 oC, 8 hMOMO

MOMO

O OH

OMOM

MOMO

MOMO

O OH

OMOM+

OAc

AcO

O OH

OAc

i) con. HCl, MeOH, r.t., 20 hii) 4 M NaOH, r.t., 2 hiii) Ac2O, pyridine, CH2Cl2, r.t., 35%

OAc

AcO

O OH

OAcOH

+

OAc

AcO

O OH

OAc

OH

OAc

AcO

O OH

OAcSOCl2, DBU, CH2Cl2

-78 oC to r.t., 12 h, 75%

i) Ru(bpy)3Cl2.6H2O, hv O2, MeOH, r.t., 26 h

ii) Ph3P, CH2Cl2, r.t. 16 h, 62%

145 146

147 27%148 37%

149150

151

152

150/151 = 8:1

Scheme 2.28. Synthesis of dienophile 149 and 152 (Han et al. 2014)

Diene 152 and dienophile 149 were then subjected to asymmetric Diels-Alder

reaction as summarized in Scheme 2.29. Based on the reported result, the chiral ligand

strongly influenced the enantioselectivity of the cycloaddition reaction. The optimal

condition for the formation of kuwanon J precursor, 154 (97% ee, 1.1:1 endo/exo) is

when diene 152 and dienophile 149 were subjected in the present of 2.5 equiv. of (R)-

VANOL, whereas for the formation of kuwanon I precursor exo-153 as major isomer

(84% ee, 1.2:1 exo/endo) when diene 152 and dienophile 149 were subjected in the

55

present of 2.5 equiv. of (S)-8,8’-dimethyl-VANOL. A final global deprotection of endo-

154 and exo-153, under mild basic conditions, efficiently furnished the desired natural

products (-)-kuwanon J (9), (+)-kuwanon I (10) respectively.

Scheme 2.29. Synthesis of (–)-kuwanon I and (+)-kuwanon J (Han et al. 2015)

Similar synthetic route was applied in the synthesis of (–)-brosimone A (121)

and (–)-brosimone B (122). For (–)-brosimone B (122), compound 147 was first

converted to 157 in 33% yield (two steps) (Scheme 2.30). Next, Schenck ene reduction

of 157 in the presence of TPP in MeOH gave 158 in 48% yield. Dehydration of 158

produced the diene 159 in 68% yield. Cycloaddition between para-prenylated

dienophile 157 and diene 159 using (S)-VANOL gave a mixture of 160 and 161 in 71%

3”Hα: kuwanon I (10)

3”Hβ: kuwanon J (9)

153: 3”Hα: exo

154: 3”Hβ: endo

56

yield in 1.2:1 ratio. Remarkably, excellent ee values for both compounds were obtained

(98% ee for endo-161, 93% ee for exo-160). Deprotection of the acetyl groups of exo-

160 gave (-)-brosimone B in 70% yield (Han et al. 2014).

146i

AcO

OH O

OAc

OAc

Me Me

AcO

OH O

OAc

OAc

Me Me

ii

OH

iiiAcO

OH O

OAc

OAc

Me

157 158 159

Me

O

OAc

OAc

OH

AcOO

AcO

OH

Me

OAc

AcOMe

157

+

159

iv v Me

O

OH

OH

OH

HOO

HO

OH

Me

OH

HOMe

Brosimone B (122)

Scheme 2.30. Synthesis of (–)-brosimone B. Reagents and conditions: i) a) 3 M HCl, MeOH, resorcinol,

80 oC, 23 min; b) 4 M NaOH, r.t., 2 h c) Ac2O, pyridine, CH2Cl2 ii) a) hv, O2, TPP, MeOH, r.t., 10 h

b)PPh3, CH2Cl2, r.t. 16 h, 48%; iii) SOCl2, DBU, THF, -78 oC to r.t., 16 h, 48% iv) (S)-VANOL,

BH3.THF, AcOH, 5 Å M. S., THF, r.t., 72 h, 71% (160:161 = 1.2:1), 93% ee for exo-140, 98% ee for

endo-141, recovered (S)-VANOL (90%); v) K2CO3, MeOH/H2O (10:1), r.t., 1 h, 70%

The same diene core 159 was used in the synthesis of brosimone A (121) in a

one-pot inter-/intramolecular Diels-Alder cycloaddition cascade strategy (Roush &

Sciotti, 1998) (Yuan et al. 2013). The (S)-VANOL-boron complex efficiently mediated

the cyloaddition reaction to give endo-endo-162, exo-endo-163 and exo-exo-164

(Scheme 2.30). Global deprotection of the exo-exo-164 gave (-)-brosimone A in 70%

yield (Han et al. 2014).

160: 3”Hα: exo

161: 3”Hβ: endo

57

159

OH

AcO

OMe

OAc

OAc

OAc

OHO

Me

AcO

OAc

OH

AcO

O

Me

OAc

OAc

OAc

OHO

Me

AcO

OAc

+

endo-endo-162 exo-endo-163

OH

AcO

O

Me

OAc

OAc

OAc

OHO

Me

AcO

OAc

+

28%, 98% ee 20%

K2CO3, MeOH-H2O

r.t, 1 h, 70%

OH

HO

O

Me

OH

OH

OH

OHO

Me

HO

OH

Brosimone A (121)exo-exo-16413%, 95% ee

Scheme 2.31. Synthesis of (–)-brosimone A

i) (S)-VANOL, BH3.THF, AcOH, THF, r.t., 25 min

ii) 158, THF, 5 Å MS, r.t., 96 h, 61%

58

CHAPTER 3

APPROACHES TOWARD THE SYNTHESIS OF MORUSALBANOL A

Morusalbanol A (42) is a mulberry Diels-Alder adduct isolated from the bark of

Morus alba which exhibits interesting neuroprotective activity (Chen et al. 2012). It is

characterized to contain an oxabicyclic [3.3.1] skeleton that derived from an

intramolecular cyclization of a cis-trans type mulberry Diels-Alder adduct.

Morusalbanol A shows evidence of atropisomerism due to the rotational hindrance of

the D/E-rings about the C5´´-C15´´and C4´´-C8´´-C9´´ bonds (Figure 3.1) (Chen et al.,

2012). Other examples of natural products in this class are cathayanon E (165) (Q. J.

Zhang, Ni, Wang, Chen, & Yu, 2009), wittiorumin F (8) (Tan et al. 2009), and CAS

441772-64-1 (166) (Figure 3.1) (Abegaz et al. 2002).

O

OH

O

O

OHHO

OH

OH

HO wittiorumin F (8)

3"4"

5"SR

S

S

morusalbanol A (42)

OH

OH

O

O

OH

HO

HO OH

MeO O

cathayanon E (165)

OH

OH

O

O

OH

O

HO OH

O

O

HO

OH

RS

OH

O OH

O

O

HO O

OH

CAS 441772-64-1 (166)

D

E

8"

9"

4"

5"

15"

Figure 3.1 Morusalbanol A (42) and related mulberry Diels-Alder adducts

59

3.1 Model study towards biomimetic Diels-Alder reaction Morusalbanol A

An important issue that needed to be addressed during the planning of the

synthesis of morusalbanol A and related mulberry Diels-Alder adducts is the formation

of a cis-trans (endo) Diels-Alder adduct precursor. Our approach hinged on the

hydrogen bond-assisted regioselective biomimetic Diels-Alder reaction between

chalcone type dienophile 169 and dehydroprenyl diene 170 (Scheme 3.1). This would

result in the formation of a cis-trans and a trans-trans type Diels-Alder adducts in one

step. The feasibility of such a [4+2] cycloaddition reaction has recently been

demonstrated by Porco and co-workers (H. Cong & Porco, 2012; Han et al. 2014),

Rizzacasa and co-workers (Boonsri et al. 2012; Gunawan & Rizzacasa, 2010) and Chee

and co-workers (Chee et al. 2011). It was envisaged that the ortho and para OH groups

on the aryl ring of diene 170 could be selectively protected due to their relative

positions related to the carbonyl group. Selective deprotection of the para OH group in

the cis-trans adducts 168 and subsequent intramolecular cyclization would then produce

morusalbanol A (Scheme 3.1)

Scheme 3.1. Retrosynthetic analysis for morusalbanol A and related mulberry Diels-Alder adducts

To test the feasibility of the biomimetic Diels-Alder reaction, we carried out a

model study on simple diene and dienophile. We examined the [4+2] cycloaddition

reactions with the model diene 172 and dienophile 175 (Table 3.1, entry 1). We

60

anticipated that cis-trans and trans-trans type Diels-Alder adducts 179 and 180 would

undergo intramolecular cyclization to produce the oxabicyclic [3.3.1] skeleton found in

the morusalbanol A. In contrast, we presumed that due to the para methyl ether

protection, the cis-trans and trans-trans type Diels-Alder adducts 181 and 182 would

not undergo intramolecular cyclization. This compound could also be used to further

understand the intramolecular cyclization reaction.

The preparation of diene 172 and 174 is outlined in Scheme 3.2. Regioselective

iodination of commercially available 2,4-dihydroxyacetophenone 76 with I2/KIO3 in

EtOH/H2O utilizing the method of Yu (H. Wang et al., 2014) gave iodobenzene 77 in

high yield. Selective protection of the para OH group in 77 with ethoxymethoxy (EOM)

group, followed by methylation of the remaining ortho OH group with dimethylsulfate

gave the iodide 171. Heck coupling of 171 and subsequent dehydration using

AcCl/pyridine (Harrington et al., 1987) provided the desired diene 172 in 81% yield in

two steps.

For diene 174, iodination of the commercially available 2,4-dihydroxyacetophenone

77 with I2/KIO3 in EtOH/H2O (H. Wang et al., 2014) gave iodobenzene 778 Selective

protection of the para OH group in 78 with dimethylsulfate gave 173. Heck coupling of

173 and subsequent dehydration with AcCl/pyridine (Harrington et al., 1987) provided

the desired diene 174 in 87% combined yield (two steps).

61

HO

OH

Me

O

HO

OH

Me

O

I

OMOE

OMe

Me

O

I

I2, KIO3

EtOH, H2O,

acetone,-10oC,

PdCl2, NaHCO3,

81% combined yield (two steps)

2) AcCl, pyridine, C6H6, 80oC

1) K2CO3

EOMO

OMe

Me

O

Me

2) Me2SO4, K2CO3,

acetone, rt, 12h, 85%77 78

171

172

Bu4NCl, DMF,100oC, 12h

15h, 72%

rt, 24h,92%

MeI, K2CO3,

acetone, rt, 8h, 85%

MeO

OH

Me

O

I

173

PdCl2, NaHCO3,

87% combined yield (two steps)

2) AcCl, pyridine, C6H6, 80oC

MeO

OH

Me

O

Me

174

Bu4NCl, DMF,100oC, 12h

Me

Me

OH1)

Me

Me

OH1)

(82)

(82)

O Cl

Scheme 3.2. Syntheses of model dienes 172 and 174

The thermal Diels-Alder reaction of diene 172 with chalcone-type dienophile

175 in toluene at 110°C for 24h afforded an inseparable mixture of cis-trans (endo) and

trans-trans (exo) adducts 179 and 180 in a 3:2 ratio and 55% yield (Table 3.1, entry 1).

The same reaction between diene 174 and dienophile 176 also gave the desired

cycloadducts in 49% combined yield as a 1:1 mixture of inseparable endo/exo

diastereomers (181 and 182), Table 3.1, entry 2. Following Rizzacasa and co-workers’

report of the substantial rate enhancement for the Diels-Alder reaction for an hydrogen-

bonded ortho OH substituent on a chalcone-type dienophile (Boonsri et al. 2012), we

proceeded to examine the thermal Diels-Alder reaction of diene 172 with dienophiles

177 and 178. Both 177 and 178 lacked an ortho hydroxyl substituent on carbonyl group

of deienophile. As expected, all attempts of thermal Diels-Alder reaction between diene

172 and dienophiles 177 or 178 failed to yield any of the desired products (Table 3.1,

entries 3-4). In each case, the dienophile was recovered while the diene decomposed.

62

This clearly indicated that the presence of an ortho OH substituent on the dienophile is

essential for the Diels-Alder reactivity.

Table 3.1. Development of initial methodology employing a model diene and dienophile

Entry Diene Dienophile Conditionsa Yield

b endo/exo

c

R1 R

2 R

3 R

4

1 172 Me EOM- 175 OH H Thermal 55% 3/2

2 174 H Me 176 OH EOMO- Thermal 49% 1/1

3 172 Me EOM- 177 H H Thermal - -

4 172 Me EOM- 178 OMe H Thermal - -

5 172 Me EOM- 175 OH H AgOTf 38% 0/1

6 172 Me EOM- 175 OH H AgBF4 20% 0/1

7 172 Me EOM- 175 OH H AlBr3 - -

8 172 Me EOM-- 175 OH H FeCl3 - -

9 172 Me EOM- 175 OH H Ga(OTf)3 - -

10 172 Me EOM- 175 OH H TiCl4 - -

11 172 Me EOM- 175 OH H BF3.Et2O - - a Thermal reaction conditions: reaction conducted in toluene (2mL/mmol) at 110°C for 24h in a pressure tube. For

catalytic conditions: reaction conducted in CH2Cl2 (4mL/mmol) at rt for 24h. Molar ratio of diene/dienophile/catalyst

= 1.2/1/0.3 bIsolated yield. cendo/exo ratios are based on 1H NMR integration.

Interestingly, the use of silver catalysts (AgOTf and AgBF4) for the Diels-Alder

reaction between diene 172 and dienophile 175 promoted the selective formation of the

trans-trans adduct, exo-180, albeit in low yield (Table 3.1, entries 5-6). The use of other

Lewis acids such as AlBr3, FeCl3, Ga(OTf)3, TiCl4, and BF3.Et2O did not give any of

63

the desired product (Table 3.1, entries 7-11). In these cases, both the diene and

dienophile decomposed.

Cyclization to form the desired oxabicylic ring was achieved by heating a

solution of endo-179 and exo-180 in 3M aqueous HCl in MeOH for 20 min.

Deprotection of the EOM group and subsequent intramolecular cyclization occurred to

give the desired oxabicyclic [3.3.1] compounds in 90% combined yield as a 3:2 mixture

of separable diastereomers (183 and 184, Scheme 3.3). Diastereomers 183 represents

the core skeleton of the natural product 441772-64-14 (185) (Figure 3.2).

Scheme 3.3. Acid catalyzed intramolecular cyclization of endo-adduct 179 and exo-adduct 180

OH

O OH

O

O

HO O

OH

CAS 441772-64-1

Figure 3.2. Chemical structure for CAS 441772-64-4(185).

64

The stereochemistry of each of the bicyclic diastereomers was confirmed by 1H-

NMR and NOESY experiments. The diastereomer 183, which was derived from the

endo-179 showed a large coupling constant (J = 11.4 Hz) from the splitting between

H4´´ and H5´´ (Figure 3.2). This clearly indicated the trans-diaxial arrangement of H4´´

and H5´´ in diastereomer 183. The small coupling constant (J = 2.8 Hz) between H3´´

and H4´´ implied a cis-orientation with H3´´ being in an equatorial position. The

NOESY correlation between the aromatic protons H14´´ and H4´´ on the cyclohexane

ring and between H3´´ and H4´´ also indicated a cis arrangement (Figure 3.3).

Figure 3.3. Key NOESY correlations leading to relative stereochemistry assignment of 183

The structure and stereochemistry of diastereomer 184 were confirmed by single

crystal X-ray crystallography (Table 3.2). The formation of oxabicyclo[3.3.1] 184 from

its trans-trans precursor exo-180 was rather surprising. Natural products containing

similar oxabicyclic core structures including morusalbanol A (44), wittiorumin F (8),

and cathayanon E (165) were all exclusively derived from their cis-trans Diels-Alder

precursors (Chen et al. 2012; Tan et al. 2009; Q. J. Zhang et al. 2009). Yu and co-

workers hypothesized that the trans-trans Diels-Alder adducts could not meet the

spatial requirement to form an oxabicyclic compound (Tan et al. 2009; Zhang et al.

2009). They demonstrated that the treatment of a cis-trans Diels-Alder adduct,

chalcomoracin (6) in 5% triflic acid/MeOH afforded a mixture of oxabicyclic and

ketalized products (wittiorumin F (8) and mulberrofuran F (9), respectively Figure 3.4).

65

However, when they subjected the recovered trans-trans Diels-Alder adduct,

mulberrofuran J (25) to similar reaction conditions, they did not obtain any product

(Tan et al. 2009).

Figure 3.4. Semisynthesis of wittiorumin F (8) and mulberrofuran F (9) from chalcomoracin C (6)

(Tan et al., 2009)

We examined the intramolecular cyclization of endo-181 and exo-182. When a

solution of endo-181 and exo-182 in 3M aqueous HCl in MeOH was heated for 20 min,

a new oxabicylic compound 186 was obtained in 92% yield (based on endo-181) along

with the recovery of 182. According the observation by Yu and co-workers in the

semisynthesis of wittiorumin F and mulberrofuran F (Figure 3.4), we expected that

endo-181 and exo-182 would undergo intramolecular cyclization to produce a ketalized

compound rather than an oxabicyclic compound due to the para methyl ether protecting

group. Although the 1H-NMR spectrum of compound 186 was observed to be quite

similar to that of 183, its 13

C-NMR spectrum however, did not appear to correspond to

either 183 or ketalized compound 187. Instead, the

13C-NMR spectrum showed signals

for an oxygenated sp3 quaternary carbon (c 76.1 ppm) and two carbonyls (c 199.9 and

66

206.0 ppm). The relative configuration of 186 was determined by HMBC correlations

and NOESY correlations. Single-crystal X-ray structure determination confirmed the

assignment as compound 186 (Scheme 3.4). Interestingly, subjecting the recovered

exo-182 (without the EOM protecting group) to similar reaction conditions did not yield

any product.

The formation of oxabicyclo [3.3.1] 186 from endo-181 is quite surprising.

Natural products containing similar cyclohexenyl core structures including kuwanon I

(11), mulberrofuran J (25) and dorsterone show restricted rotation about the C3”-C3’

bond (as shown in Scheme 3.4) due to the unsymmetrical nature of the aryl ring. Most

of the signals in the 1H-NMR spectrum of these natural products were doubled at room

temperature as a result of atropisomerism due to the high rotation barrier. However, in

the case of 181, the barrier for rotation about the C3”-C3’ bond was negligible, allowing

a 180° rotation followed by intramolecular cyclization to form the respective

oxabicyclic [3.3.1] compound 186.

Scheme 3.4. Formation of diastereomer 186.

67

Table 3.2 Crystal data and structure refinement for 184 and 186

184 186

Empirical formula C29H28O5 C30H29Cl3O6

Formula weight 456.51 591.88

Crystal system, Space group Monoclinic, C2/c Monoclinic, P21/c

a (Å) 39.193(4) 15.3140(9)

b (Å) 7.3592(8) 13.9413(8)

c (Å) 16.0729(15) 13.3869(10)

β () 95.317(7) 106.278(7)

Volume (Å3) 4615.9(8) 2743.5(3)

Z 8 4

Calculated density (Mg/m3) 1.314 1.433

Absorption coefficient 0.089 mm-1

0.378 mm-1

F(000) 1936 1232

Crystal size (mm3) 0.690 x 0.510 x 0.050 0.510 x 0.310 x 0.270

Crystal color and habit colorless plate colorless needle

θ range for data collection 2.087 to 29.615° 2.793 to 30.211°

Reflections collected 15853 37196

Independent reflections 6342 (Rint = 0.0419) 11114 (Rint = 0.1283)

Observed reflections [I > 2σ(I)] 3926 7561

Completeness to θ = 25.242° 99.6 % 99.9 %

Data / restraints / parameters 6342 / 0 / 313 11114 / 13 / 399

Goodness-of-fit on F2 1.039 1.021

Final R indices [I > 2σ(I)] R1 = 0.0530,

wR2 = 0.1099

R1 = 0.0650,

wR2 = 0.1737

R indices (all data) R1 = 0.0963,

wR2 = 0.1285

R1 = 0.0935,

wR2 = 0.1867

Largest diff. peak and hole (e Å-3

) 0.570 and -0.256 0.444 and -0.594

CCDC deposit number for 184 and 186 are 1026385 and 1043889 respectively.

68

3.2 DFT calculation for model study

To understand why the Diels–Alder reactivity is so highly dependent on the

presence of the ortho OH group of the chalcones 175 and 178, we carried out density

functional theory (DFT) calculations on the Diels–Alder reactions of 175 and 178 with

the model diene 172 as shown in Scheme 3.5.

Me

O OMe

Me

OMOE

172

OR O

175: R = H 177: R= Me

Me

O

RO

OMOE

OMe

O Me

Me

O

RO

OMOE

OMe

O Me

+

179 R= H 188 R= Me

180 R= H 189 R= Me

Scheme 3.5. Model Diels-Alder reaction between diene 172 with chalcones 175 and 177 for DFT

calculation

Transition state calculation was performed to gain insights on how the ortho OH

group on the dienophile 175 contributed to the formation of Diels–Alder adducts 179

and 180. The endo and exo transition structures and their activation energies and the

overall reaction energies are shown in Figure 3.4. The transition states were concerted

and synchronous and the cycloaddition reactions were observed to be exothermic.

Hydrogen bonding was observed between the ortho OH group of the dienophile 175

with the adjacent carbonyl oxygen (intramolecular) and the carbonyl oxygen of the

diene 172 (intermolecular). This hydrogen bonding is essential as it leads to the shorter

distance between the carbonyl oxygen of the diene 172 and the ortho OH group of the

dienophile 175 and thus promoting the cyclization.

69

Figure 3.5 showed the energy profile of the transition state structure of 179, 180,

188 and 189. The activation energy (∆G*) for the transition state of endo-179 and exo-

180 were 2.95 kcal/mol and 2.57 kcal/mol, respectively. The Gibbs free energy of

reaction (∆G) for endo-179 and exo-180 were 4.64 kcal/mol and 3.73 kcal/mol,

respectively. The low values of activation energy and the Gibbs free energy explain why

endo-179 and exo-180 can undergo intramolecular cyclization to form oxabicylic [3.3.1]

compound 183 and 184.

TS1 endo-179 TS2 endo-180

∆H* = -0.20 ∆H* = -0.12

∆G* = 2.95 ∆G* = 2.57

∆H = -1.34 ∆H = -1.07

∆G = 4.64 ∆G = 3.73

TS3 endo-188 TS4 endo-189

∆H* = 0.61 ∆H* = -0.27

∆G* = 2.49 ∆G* = 4.85

∆H = -1.11 ∆H = -2.60

∆G = 5.05 ∆G = 8.40

Figure 3.5. Energy profile of transition state structure of (a) 179 (b) 180 (c) 188 (d) 189 where black

dashed lines denoted the bond formation that led to cyclisation and blue dashed lines denoted the

intermolecular hydrogen bonding

70

The calculation for reaction pathway of 186 and 187 is shown in Figure 3.6. The

transition state calculations were carried out using linear synchronous transit (LST) and

quadratic synchronous transit (QST) methods to further understand the reaction

mechanism. The lower energy barrier between the transition state and reactant were

found in the formation of the 186 rather than 187.

In order to have a better insight into the how C3”-C3’ bond rotate endo-181 to

facilitate the formation of the endo-186, an energy scan of the optimized cycloadduct

geometry using local density approximation (LDA) or local functional PWC using

Dmol3 module along the coordinates of the dihedral angle was performed. The torsion

for rotation of 181 and 182 were plotted in Figure 3.6. The torsion angle was scanned

from left to right with the initial optimized structure of each form. Eighteen

conformations with a difference dihedral angle of 20° were obtained. The shortest

bridging distance was 3.61 Å and 2.72 Å for endo-181 and exo-182, respectively. As

highlighted in the potential energy scan graph in Figure 3.7, exo-182 has a barrier to

rotate of approximately 9 times higher than that of endo-181. This observation could

explain why no product was obtained from the acid treatment of exo-182.

71

Figure 3.6. Calculated reaction path for (a) 187 and (b) 186

a) b)

72

(a)

(b)

Figure 3.7. Potential energy scans on dihedral angle together the respective energy barrier and distance for (a)

endo-181 and (b) exo-182. The distances between the carbon and oxygen atom in red circle were calculated

in the plot

0

1

2

3

4

5

6

-1 0 1 2 3 4 5 6 7 8 9

74

.49

94

.49

11

4.4

9

13

4.4

9

15

4.4

9

17

4.4

9

-16

5.5

1

-14

5.5

1

-12

5.5

1

-10

5.5

1

-85

.51

-65

.51

-45

.51

-25

.51

-5.5

1

14

.49

34

.49

54

.49

74

.49

Energy

(kcal/mol)

Distance(Å)

0

1

2

3

4

5

6

0

2

4

6

8

10

-54

.76

-30

.31

-1.8

4

28

.69

51

.99

72

.52

90

.36

10

8.7

9

12

6.0

5

14

3.4

7

16

0.8

0

17

8.8

8

-16

3.2

0

-14

5.1

2

-12

7.6

3

-11

1.2

1

-95

.77

-77

.98

-57

.64

Torsion

Angle(o)

73

In conclusion, the model studies provided a useful approach toward the synthesis of

the oxabicyclic [3.3.1] core system found in morusalbanol A and related mulberry Diels-

Alder adducts. In particular, the required cis-trans (endo) Diels-Alder precursors of

morusalbanol A and 441772-64-1 were obtained via the thermal cycloaddition reaction

which was proven to be dependent on the presence of a hydrogen-bonded ortho-OH

subtituent on the chalcones dienophile. Acid-catalyzed intramolecular cyclization of a cis-

trans (endo) Diels-Alder adduct (endo-179) afforded the desired oxabicyclic [3.3.1] core of

441772-64-1 in a stereoselective manner. Additionally, rotation about the C3”-C3’ bond

for cis-trans (endo) Diels-Alder (endo-181) was observed during acid catalyzed

intramolecular cyclization to form an oxabicyclic [3.3.1] compound. Together with the

results from these studies, it provide important insights into to syntheses of morusalbanol A

and related mulberry Diels-Alder adducts.

3.3 Approaches toward the synthesis of morusalbanol A

Having established the method for construction of the oxabicyclo[3.3.1] core of

morusalbanol A (Section 3.2), we then attempted the synthesis of morusalbanol A (44). The

biosynthesis pathway of morusalbanol A was proposed by Chen and co-workers as

discussed in Section 1.4 (Scheme 1.22) (Chen et al. 2012). In this section, we elaborate

several approaches toward the synthesis of morusalbanol A.

74

OH

HO

O

MeO

O

O

Me

OH

OH

HO

HO

morusalbanol A (42)

Figure 3.8. Morusalbanol A (42) (Chen et al., 2012)

3.3.1 Retrosynthetic analysis for morusalbanol A (42)

A retrosynthetic analysis for morusalbanol A synthesis is detailed in Scheme 3.6.

We envisaged that morusalbanol A (42) could be derived from an intramolecular

cyclization of a cis-trans type mulberry Diels-Alder adduct 43. The required dienes 190 and

191 could be made from a C3-iodinated benzoic ester 193 followed by the Heck coupling

with 2-methyl-3-buten-2-ol (82). The chalcone type dienophiles 192 and 88 could be

readily prepared from commercially available acetophenone 77.

75

HO

HO OH

OMe

O

Me

OH

O

HO

HOHO

HO

OH

OH

MeO

O

O OH

OH

OH

HO

OH

HO

O

MeO

O

O

OH

OH

HO

HO Me

morusalbanol A (42)

43

endo-addition

Me

RO OR

OR O

OMe

190: R = EOM191: R = Me

O OR

RO

OH

OR

192: R = EOM 88: R = Me

HO OH

OH

OH

O

193

HO

OH

Me

O

77

Scheme 3.6. Retrosynthetic strategy of morusalbanol A (42) via Diels-Alder reaction

3.3.2 Synthesis of diene 190

We began our synthesis of the diene 190 from commercially available 2,4,6-

trihydroxybenzoic acid 193 (Scheme 3.7). Esterification followed by mono iodination on

193 with I2/KIO3 in EtOH/H2O (Han et al., 2014) gave iodobenzene 195 in 80% yields.

76

Ethoxymethoxy (EOM) protection of 195 gave 196 in 55% yields. Heck coupling of 196

with 2-methyl-3-buten-2-ol (82) and subsequent dehydration with AcCl/pyridine provided

the diene 190 in two steps.

HO OH

OH

OOH

HO OH

OMe

OOH

i

iii EOMO OMOE

OMe

OEOMO

I

HO OH

OMe

OOH

I

ii

iv, vEOMO OMOE

OMe

OEOMO

Me

193 194 195

196 190

Scheme 3.7. Synthesis of diene 190. Reagents and conditions: (i) CH3I, K2CO3, acetone, rt, 15 h, 60% (ii) I2,

KIO3, ethanol, H2O, rt, 14 h, 80% (iii) EOM-Cl, K2CO3, acetone, 0 oC, 48 h, 55% (iv) 2-methylbut-3-en-2-ol

(82), Pd(OAc)2, K2CO3, DMF, 100oC, 7 h (v) AcCl, pyridine, benzene, 60

oC, 4 h, 80% (2-steps yield)

3.3.3 Synthesis of dienophile 192

The synthesis of dienophile 192 is detailed in Scheme 3.8. Selective protection of

the para OH with EOM group on 77 afforded acetophenone 197. Subsequently Claisen-

Schmidt condensation (Fine & Pulaski, 1973) of 197 with benzaldehyde 198 gave

dienophile 192 in 55% yield.

77

HO

OH

Me

O

i

EOMO

OH

Me

O

ii

EOMO

EOMO

H

O EOMO

OH O

OMOE

OMOE

77 197 192

198

Scheme 3.8. Synthesis of dienophile 192. Reagents and conditions: (i) EOM-Cl, K2CO3, acetone, 0 oC, 16 h,

85% (ii) 50% aqueous KOH, EtOH, 42h, 55%

3.3.4 Diels-Alder reaction between 190 and 192

With the diene 190 in hand, we attempted the [4+2] cycloaddition with chalcone

192. Unfortunately, the Diels-Alder reaction between diene 190 and dienophile 192

resulted in decomposition at 130 oC under refluxing toluene (Scheme 3.9). Presumably,

both diene 190 and dienophile 192 needed to be protected with a more thermally stable

protecting group.

EOMO

EOMO

OMOE

Me OMe

O

+

O

EOMO

OH OMOE

OMOE

190

192

toluene

130 oCX

HO

OMOE

OMOE

MeO

O

O OMOE

OMOE

OMOE

EOMO

Me

exo-200

endo-199

HO

OMOE

OMOE

MeO

O

O OMOE

OMOE

OMOE

EOMO

Me

Scheme 3.9. Cycloaddition reaction of diene 190 and dienophile 192

78

3.3.5 Synthesis of diene 191

We then attempted to synthesise the methyl ether diene 191. The preparation of

diene 191 is outlined in Scheme 3.10. Esterification and methyl ether protection of benzoic

acid 193 gave 201 in 60% yield. Regiospecific iodination of 201 with I2/ KIO3 in

EtOH/H2O gave ester 201 in 80% yields (H. Wang et al., 2014). Methylation of the

remaining ortho OH group with iodomethane gave 203 in 95% yield. Subsequent

installation of the diene moiety by a Heck coupling reaction (Arkoudis et al. 2009) onto

203 gave the required diene 191 in 80% yield.

HO OH

OH

OOH

MeO OMe

OMe

OOH

i iiMeO OMe

OMe

OOH

I

iiiMeO OMe

OMe

OOMe

I

iv,vMeO OMe

OMe

OOMe

Me

193 201 202

203 191

Scheme 3.10. Synthesis of Diene 191: Reagents and conditions: (i) CH3I, K2CO3, acetone, rt, 15 h, 60% (ii) I2,

KIO3, ethanol, H2O, rt, 14 h, 80% (iii) CH3I, K2CO3, acetone, rt, 15 h, 95% (iv) 2-methylbut-3-en-2-ol,

Pd(OAc)2, K2CO3, DMF, 100oC, 6 h (v) AcCl, pyridine, benzene, 60

oC, 4 h, 80%.

3.3.6 Synthesis of dienophile 88

Similarly the synthesis of dienophile 88 methyl ether was attempted from

acetophenone 77. Methylation of the para OH group of 77, followed by Claisen-Schmidt

condensation with 2’,4’-dimethoxybenzaldehyde 99 produced the dienophile 88 in 80%

yield (Scheme 3.11) (Fine & Pulaski, 1973).

79

MeO

OH

Me

O

MeO

OH O OMe

OMe

98 88

99

MeO

OMe

H

O

HO

OH

Me

O

77

i ii

Scheme 3.11. Synthesis of Dienophile 88. Reagents and conditions: (i) CH3I, K2CO3, acetone, rt, 12 h, 90%

(ii) 2,4-dimethoxybenzaldehyde (99), NaH, DMF, 24h, 80%.

3.3.7 Diels-Alder reaction between 191 and 88

With the diene 191 and dienophile 88 in hand, we proceeded to the thermal Diels-

Alder reaction in toluene at 135 °C for 24 h to obtain a separable mixture of endo-204 and

exo-205 in a 3:2 ratio in 55% yield (Scheme 3.12).

OMe

MeO OMe

OMe

O

Me

OMe

O

MeO

OHOMe

+

191

88

HO

OMe

OMe

MeO

O

O OMe

OMe

OMe

MeO

Me

endo-204

exo-205

toluene, 135 oC

24 h, 55%(endo/exo = 3:2)

HO

OMe

OMe

MeO

O

O OMe

OMe

OMe

MeO

Me

+

Scheme 3.12. Cycloaddition reaction of diene 191 and dienophile 88

80

The proton and carbon-13 NMR spectra for endo-204 showed significant changes at

room temperature after 14 days (Figure 3.9 and 3.10) presumably due to atropisomerism. In

the 1H NMR spectrum, chemical shifts at δ 7.25 (d, J = 2.3 Hz), 6.91 (d, J = 9.2 Hz), 6.49

(d, J = 2.3 Hz), 6.46 (d, J = 2.3 Hz) and 6.39 (d, J = 2.3 Hz) were appeared after 14 days.

Additionally, a doublet at δ 6.38 ppm, two singlets at δ 5.78 ppm and δ 5.42 ppm and a

multiplet at the range of δ 3.88-3.87 ppm were also observed. Likewise, new chemical

shifts at the region of δ 170 -100 ppm were observed in 13

C NMR spectrum.

Figure 3.9. Comparison of the 1H NMR spectra for endo-204 at a) day 1 b) day 14

81

Figure 3.10. Comparison of the 13

C NMR spectra for endo-204 at a) day 1 b) day 14

Further studies were undertaken to optimize the yield of the exo-205. Based on the

model study established in Section 3.1, the use of silver catalysts AgOTf and AgBF4 (Table

3.1, entry 5 and 6) in the Diels-Alder reaction promoted the selective formation of exo-180.

Unfortunately, using Lewis acids such as AgOTf, AgBF4 and AgNP (Huan Cong et al.

2010) (H. Cong & Porco, 2012) (Qi et al., 2013) in the Diels-Alder reaction for diene 191

and dienophile 87 only led to decomposition of the reaction mixture.

Next, we examined the intromolecular cyclization of endo-204. Selective removal

of the ortho OMe group of endo-204 with MgI2 smoothly produced ()-morusalbanol A

206 pentamethyl ether in 50% yield (Scheme 3.13). When endo-204 was treated with

82

MgI2 for 6h, a new compound ()-207, was obtained in 36% yield (Scheme 3.14).

Attempts to globally demethylate 206 or 207 with MgI2, BCl3, and 1-

trimethylsilylquinolinium iodide (TMSI-quinoline) to give murasalbanol A were

unsuccessful.

HO

OMe

OMe

MeO

O

O OMe

OMe

OMe

MeO

Me

endo-204

OH

MeO

O

MeO

O

O

Me

OMe

OMe

MeO

MeO

MgI2 (10 equiv)

Et2O: THF (1:1)6 h, 50 %

morusalbanol A pentamethyl ether (206)

Scheme 3.13. Synthesis of morusalbanol A pentamethyl ether (206)

(X-ray )

83

HO

OMe

OMeMeO

O

O OMe

OMe

OMe

MeO

Me

endo-204

OH

MeO

O

O

O

Me

OMe

OMe

MeO

HO

MgI2 (10 equiv)

Et2O: THF (1:1)6 h, 36 %

207

OH

Scheme 3.14. Synthesis of 207.

The relative stereochemistry of 206 was determined by 1H-NMR (Table 3.3) and

NOESY experiments (Figure 3.10). The large coupling constant (11.4 Hz) between H-4”

and H-5” of 206 indicated their trans arrangement. In the NOESY spectrum (Figure 3.11),

correlation between the aromatic protons H15” and H4” on the cyclohexane ring indicated

a cis relationship while correlation between H3” and H4” suggested these were also in a cis

arrangement. The structures of 206 as well as 207 were confirmed by X-ray

crystallography (Table 3.3).

(X-ray )

84

OH

MeO

O

MeO

O

O

Me

OMe

OMe

MeO

MeOO

H

OMe

O

OOMe

O

MeO

OMe

MeO

OMe

H

H

H

H

H

HH

H

HH

H

H

H

H

H

H

H

(NOESY)

1"

3"

4"5"

6"2"

2'

6'

NOEmorusalbanol A pentamethyl ether (206)

9"15"

17"

Figure 3.11. NOESY correlations of morusalbanol A pentamethyl ether (206)

It was noteworthy that a number of proton and carbon signals in the NMR spectra

were absent in compounds 206 and 207. For instance, the absence of proton (H5”) and

carbon (C5” and C20”) resonances in chloroform-d1 at ambient temperature could have

resulted from atropisomerism due to the rotational hindrance of the D/E-rings about the

C5”-C15” and C4”-C8”-C9” bonds. However, when the 1H NMR spectra of 206 was

acquired at 289 K and 353 K in DMSO-d6, resonances of H5 and C20 were observed,

although the resonance for C5 was still not observed. The overall spectral data for ()-206

and ()-207 were in agreement with those reported by Yue et al. for the natural product

(Chen et al. 2012).

85

Table 3.3. Crystal data and structure refinement for 206 and 207

206 207

Empirical formula C69.50H76O20 C31H32O10

Formula weight 1231.30 564.56

Crystal system, Space group Monoclinic, P21/c Triclinic, P-1

a (Å) 13.7493(8) 10.1052(9)

b (Å) 10.2452(5) 11.8237(10)

c (Å) 43.919(2) 12.7062(12)

α () 90 77.712(7)

β () 90.148(5)° 79.804(8)

() 90 70.061(8)

Volume (Å3) 2743.5(3) 1385.4(2)

Z 4 2

Calculated density (Mg/m3) 1.433 1.353

Absorption coefficient (mm-1

) 0.097 0.101

F(000) 2612 596

Crystal size (mm3) 0.320 x 0.120 x 0.090 0.280 x 0.050 x 0.020

Crystal color and habit colorless needle colorless plate

θ range for data collection 2.841 to 27.000° 2.863 to 25.992°

Reflections collected 70030 10730

Independent reflections 13461 (Rint= 0.1344) 5444 (Rint = 0.0566)

Observed reflections [I > 2σ(I)] 7921 2621

Completeness to θ = 25.242° 99.8 % 99.8 %

Data / restraints / parameters 13461 / 0 / 845 5444 / 3 / 378

Goodness-of-fit on F2 1.043 1.023

Final R indices [I > 2σ(I)] R1 = 0.0756, wR2 = 0.1555 R1 = 0.0847, wR2 = 0.1844

R indices (all data) R1 = 0.1362, wR2 = 0.1890 R1 = 0.1783, wR2 = 0.2374

Largest diff. peak and hole (e Å-3

) 0.341 and -0.310 1.460 and -0.319

CCDC deposit number 1402220 (for 206) and 1402153 (for 207).

86

Table 3.4. 1H and

13C NMR spectra of morusalbanol A pentamethyl ether (206)

a signals invisible, bmeasurement at 353 K, ccoupling constant cannot be measured due to broad reasonance.

No. 206 (CDCl3) 206 (toluene) 206 (toluene)b

13C δ(ppm)

1H δ (multiplicity, J in Hz)

13C δ(ppm)

1H δ (multiplicity, J in Hz)

13C δ(ppm)

1H δ (multiplicity, J in Hz)

1” 75.6 - 75.1 - 75.5 -

2” 36.3 1.89 (m)

2.11 (dd, 2.8, 13.3)

36.0 1.60 (bd, 12.8)

1.71 (bd, 11.9)

36.8 1.68 (m)

1.80 (dd,3.2, 13.2)

3” 31.2 3.65 (m) 31.3 3.72 (m) 31.8 3.72 (m)

4” 50.6 4.37 (bs) 51.0 4.21 (m) 51.3 4.26 (d, 11.4)

5” a a a

3.86 (m) a

3.79 (ddd, 4.6, 12.8, 16.9)

6” 45.5 2.19 (m) 45.8 2.30 (bd)c 46.5 2.26 (ddd, 1.8, 4.1, 13.8)

7” 28.6 1.36 (s) 28.4 1.26 (s) 28.7 1.26 (s)

8” 204.1 - 203.9 - 204.2 -

9” 114.3 - 114.5 - 115.3 -

10” 165.4 - 165.4 - 166.0 -

11” 101.1 6.39 (d, 2.3) 100.9 6.35 (d, 2.3) 101.8 6.33 (d, 2.7)

12” 165.3 - 166.2 - 166.4 -

13” 107.1 6.50 (dd, 2.8, 9.1) 107.2 6.45 (d, 8.7) 107.4 6.46 (dd, 2.3, 8.7)

14” 130.5 7.87 (d, 9.1) 130.3 7.79 (d, 9.2) 130.7 7.79 (d, 8.7)

15” 125.6 - 127.6 - 127.6 -

16” 158.5 - 158.3 - 158.4 -

17” 99.1 6.31 (d, 2.3) 99.1 6.16 (d, 1.8) 100.2 6.18 (d, 2.3)

18” 159.1 - 159.3 - 160.0 -

19” 104.1 6.25 (dd, 2.3, 8.2) 104.0 6.06 (dd, 1.8, 8.2) 105.3 6.10 (dd, 2.3, 8.2)

20” a

6.90 (d, 8.2) 125.5 6.83 (d, 8.2) 125.8 6.84 (d, 8.2)

3’ 103.0 - 103.2 - 104.1 -

2’ 154.9 - 155.0 - 155.6 -

1’ 107.1 - 105.5 - 106.6 -

6’ 157.5 - 157.6 - 158.9 -

5’ 86.1 5.88 (s) 86.2 5.60 (s) 87.5 5.68 (s)

4’ 158.3 - 158.4 - 159.1 -

7’ 167.5 - 166.3 - 166.6 -

MeO 54.7 3.38 (s) 54.1 3.18 (s) 54.6 3.20 (s)

MeO 55.3 3.68 (s) 54.2 3.20 (s) 54.9, 3.26 (s)

MeO 55.6 3.74 (s) 54.6 3.21 (s) 55.0 3.26 (s)

MeO 55.5 3.80 (s) 54.6 3.23 (s) 55.3 3.30 (s)

MeO 55.9 3.82 (s) 54.8 3.24 (s) 55.8 3.34 (s)

MeO 52.2 3.88 (s) 51.2 3.72 (s) 51.8 3.72 (s)

OH 12.75 13.36 13.09

87

Table 3.5. 1H and

13C NMR spectra of compound 207

asignals invisible, bmeasurement at 353 K, 13C-NMR was not recorded at 353 K, ccoupling constant cannot be

measured due to broad reasonance. dsignal overlapped.

No. 207 (CDCl3) 207 (DMSO) 207 (DMSO)b

13C δ(ppm)

1H δ (multiplicity, J in Hz)

13C δ(ppm)

1H δ (multiplicity, J in Hz)

1H δ (multiplicity, J in Hz)

1” 80.3 - 77.6 - -

2” 35.9 1.97 (dt, 12.8)

2.22 (dd, 2.7, 13.1)

35.1 1.69 (d, 11.9)

2.38 (d, 11.4)

1.70 (m)

2.39 (d, 11.4)

3” 30.3 3.68 (m)d 30.9 3.45 (m)

d 3.35 (m)

d

4” 49.8 4.37 (bd) c 50.3 4.50 (bd, 11.9) 4.47 (d, 11.2)

5” a a a a a

6” 44.9 2.29 (bd, 12.8) 46.2 2.06 (bd)c 2.09 (bd)

c

7” 28.6 1.56 (s) 28.6 1.35 (s) 1.40 (s)

8” 203.4 - 205.0 - -

9” 114.1 - 114.3 - -

10” 165.5 - 164.9 - -

11” 101.1 6.38 (d, 2.3) 101.4 6.40 (d, 2.3) 6.37 (d, 2.5)

12” 165.0 - 165.8 - -

13” 107.5 6.50 (dd, 1.8, 8.7) 107.5 6.56 (dd, 2.3, 9.2) 6.55 (d, 7.6)

14” 130.3 7.82 (d, 9.1) 132.2 8.19 (d, 8.7) 8.12 (d, 8.7)

15” 122.5 - 124.6 - -

16” 158.2 - 158.1 - -

17” 99.2 6.33 (d, 2.3) 99.0 6.36 (d, 2.3) 6.37 (d, 2.5)

18” 159.3 - 159.1 - -

19” 104.2 6.26 (dd, 2.3, 8.2) 105.0 6.27 (dd, 2.3, 8.7) 6.27 (d, 8.5)

20” a

6.88 (d, 8.7) a

6.98 (d, 8.7) 6.94 (d, 8.7)

3’ 102.0 - 102.6 - -

2’ 155.7 - 157.5 - -

1’ 94.1 - 96.0 - -

6’ 165.5 - 163.5 - -

5’ 92.8 6.02 (s) 91.3 5.91 (s) 5.94 (s)

4’ 162.3 - 161.7 - -

7’ 171.2 - 172.4 - -

MeO 55.2 3.38 (s) 55.4 3.23 (s) 3.28 (s)

MeO 55.3 3.69 (s) 55.4 3.61 (s) 3.62 (s)

MeO 55.5 3.75 (s) 56.0 3.64 (s) 3.67 (s)

MeO 55.6 3.88 (s) 56.2 3.78 (s) 3.80 (s)

OH 12.37 12.13 12.43

OH 12.62 12.54 12.52

COOH 11.38 a a

88

CHAPTER 4

APPROACHES TOWARD THE SYNTHESIS OF SOROCEIN B

Sorocein B (62) was isolated from the root bark of Sorocea bonplandii with an

overall yield of 52 mg per kilogram of dried bark. Thus far, there have been no report on

the synthesis of this bioactive natural product. In this chapter we described several

approaches toward the synthesis of sorocein B.

O O

Me

OH

OH

OOH

HO

OH

OMe

Me

H

H

3"

4"

6"

SR

S

R

Figure 4.1. Sorocein B (62)

4.1 Retrosynthetic analysis

A retrosynthetic analysis towards the synthesis of sorocein B (62) involves a

biogenesis-inspired intermolecular [4+2]-cycloaddition reaction between the chalcone

dienophile 209 and diene 106 as shown in Scheme 4.1. It was envisaged that the

chalcone-type diene 106 could be obtained from a Heck coupling on the chalcone 210 with

2-methyl-3-buten-2-ol (82) (Arkoudis et al., 2009) followed by dehydration (Harrington et

al., 1987). The required dienophile 209 could be readily prepared from commercially

available 2,4-dihydroxyacetophenone 77 and 2´,4´-dihydroxybenzaldehyde 92 via Claisen-

Schmidt condensation (Fine & Pulaski, 1973).

89

HO

OH

Me

O

OH

O

Me

O

O

OMe

HO

OH

OH

OH

Me

HO

OMe

OMeO

MeO

OMe

Me

MeO

O O

OMe

OMe

Me

Diels-Alderreaction

O O

OH

OH

OOH

HO

OH

OMe

Me

H

H

Me

OO

HO HO OH

Me Me

Ar

Me

endo-transition state

Heck coupling

Claisen-Schmidtcondensation

HO

OH

H

O

OMe

MeOO

I

MeO

OMe+HO

OMe

OHO

I

MeO

OMe

Sorocein B (62)208

106 209

2-methylbut-3-en-2-ol

210 211

9277

82

intramolecular ketalization

+

installation of pyran ring

Me

Me

Scheme 4.1. Retrosynthetic analysis of sorocein B (62)

4.1.1 Synthesis of chalcone core as diene 105

The synthesis of diene 106 is outlined in Scheme 4.2. Regiospecific iodination (H.

Wang et al. 2014) of commercially available 2’,4’-dihydroxyacetophenone 77 with I2/KIO3

90

in EtOH/H2O followed by methylation and iodomethane. Subsequent Heck coupling of

210 and dehydration with AcCl/pyridine provided the desired diene 106 in 65% combined

yield in two steps. However, diene 106 was unstable and used immediately after silica gel

purification.

O Me

OH

OH

O Me

OH

OH

I2, KIO3

EtOH, H2Ort, 24 h, 92% I

OMe

OMe

O H

KOH, EtOHacetone, rt15 h, 80%

CH3I, K2CO3

O Me

OMe

OMe

I rt, 70%

MeO

I

OMe O OMe

OMe MeO

OMe O OMe

OMe

1) 2-methylbut-3-en-2-ol (82) PdCl2, K2CO3, DMF, 140oC, 24h

2) AcCl, pyridine

C6H6, 80oC

65 % combined yield

Me

77 78

99

79

210 106

Scheme 4.2. Synthesis of diene 106

4.1.2. Synthesis of dienophile 209

The synthesis of dienophile 209 began with selective iodination of 2´,4´-

dihydroxyacetophenone 77 with I2/KIO3 in EtOH/H2O, followed by methylation to give

acetophenone 173 (Scheme 4.3). Claisen-Schmidt condensation between acetophenone

173 and benzaldehyde 99 furnished the chalcone 211 in 70% yield. Installation of 2,2-

dimethyl chromene ring (pyran moiety) gave chalcone 209 in 65% yield (Scheme 4.3).

91

O Me

OH

OH

O Me

OH

OH

I2, KIO3

EtOH, H2Ort, 24 h, 92% I

OMe

OMe

O H

KOH, EtOHacetone, rt15 h, 80%

CH3I, K2CO3

O Me

OH

OMe

I rt, 70%

MeO

I

OH O OMe

OMeMeO

O O OMe

OMe

Me Me1) 2-methylbut-3-en-2-ol (82)

PdCl2, K2CO3,

DMF, 140oC, 24h

2) SiO2

heat, 130oC

78

99

77 173

211 209

Scheme 4.3. Synthesis of dienophile 209

4.1.3 Method development for installation of 2,2-Dimethyl-2H-chromenes (pyran

moiety)

2,2-Dimethyl-2H-chromenes (referring to the pyran moiety of 209) are commonly

found as parent structure in natural products (Ellis et al. 1977). It is also an important

intermediate for the syntheses of numerous pharmaceutical and biologically active

compounds (Mannhold et al., 1999; Sa e Sant'Anna et al., 2005; Szczepanik et al., 2005).

Therefore, a number of methods have been reported for the construction of 2,2-dimethyl-

2H-chromenes. In particular, condensation of phenols with α,β-unsaturated carbonyls

(aldehydes or acetals) under different reaction conditions, is the most commonly used

method (Adler & Baldwin, 2009; Akanksha & Maiti, 2012; Bröhmer, Volz, & Bräse, 2009;

Chauder et al. 1998; Lamcharfi, Menguy, & Zamarlik, 1993; Lee, Choi, & Yoon, 2005;

North et al. 1995; Prado, Janin, & Bost, 2006; Sartori et al. 1979). Other methods include

condensation of phenols with 2-methyl-3-butyn-2-ol (82) catalyzed by ReCl(CO)5 (Zeng,

92

Ju, & Hua, 2011), BF3.Et2O (Madabhushi et al. 2012) or other Lewis acids (Bigi et al.

1997; Dong, Wang, & Wang, 2008; Gabbutt et al. 2003; Zhao & Carreira, 2003);

condensation of phenols with 3-chloro-3-methylbut-1-yne catalysed by Ph3PAuNTf2

(Lykakis et al. 2011); dehydration of chromanols (Bergmann & Gericke, 1990; Kureshy et

al. 2009); Pd-catalyzed cyclization of o-prenyl phenols (Iyer & Trivedi, 1990; Larock, Wei,

& Hightower, 1998), and photocyclization of 3-aryl-1,1-dimethylprop-2-en-1-ol (Pandey &

Krishna, 1988).

We first attempted to make the chromane (215) via a condensation reaction of a

simple o-iodophenol 212 and 2-methyl-3-buten-2-ol 82 using a variety of Pd catalysts

(PdCl2, Pd(OAc)2, Pd(PPh3)4, Pd(TFA)2, Pd(dba)2), bases (Na2CO3, K2CO3, KOAc),

additive (n-Bu4NCl) and solvents (DMF, DMSO, toluene) and the results are summarised

in Table 4.1. Mixture of 212 (1.0 mmol), 82 (1.5 mmol) and PdCl2 (1 to 10 mol %) in the

present of K2CO3 (3 equiv.) in toluene (1 mL) was heated in a sealed tube under nitrogen at

100C. TLC analysis of the reaction mixture showed 212 to be completely consumed after

18 h. As expected, the Heck coupling product 214 was the major product obtained (35-45%

yield, entries 1-3). The desired product, 2,2-dimethyl-2H-chromene (215), was isolated

only in low yield (5-10%, entries 1-3). Increasing the reaction temperature to 140 C

resulted in a slightly increased yield of 215 (20%, entry 4 vs entry 3). Presumably, a higher

reaction temperature improved the formation of 215, most likely due to increase in the rate

of condensation of 214. Indeed, when the reaction was repeated at 150 C in DMF (entry 5)

and DMSO (entry 6), 215 was obtained in 23% and 25% yields, respectively. Attempts to

increase the yield of 215 through microwave irradiation, however, were unsuccessful

(entries, 7-8). It is noteworthy that identical results are observed under both air and

nitrogen atmosphere (Table 4.1, entries 4 and 5).

93

We next examined the catalytic activity of other Pd catalysts. Notably, the yields of

214 and 215 were slightly increased with catalytic Pd(OAc)2 (Table 4.1, entries 9-10).

Although three other Pd catalysts, Pd(PPh3)4, Pd(TFA)2 and Pd(dba)2, showed activity for

the reaction, they were less effective than Pd(OAc)2 (Table 4.1, entries 11-13). Adding n-

Bu4NCl had no significant effect on formation of 214 and 215 (Table 4.1, entry 10 vs entry

14). The effect of base on the yield of reaction was also evaluated (Table 4.1, entries 15-

17). The yields of 214 and 215 were reduced when Et3N was used (Table 4.1, entry 17). We

also found that both Na2CO3 and KOAc were less efficient for the Heck coupling reactions

(Table 4.1, entries 15-16). Amongst the bases studied, K2CO3 gave the best result (Table

4.1, entries 10 & 14).

94

Table 4.1. Pd-catalysed coupling-condensation of o-iodophenol 212 and 2-methyl-3-buten-2-ol 82

212

+

82: R = H213: R = Ac 214 215

conditions

OH

IMe

Me

OROH

Me

MeOH

+

OMe

Me

aReaction conditions: 212 (1.0 equiv.), 82 (10 equiv.), solvent 1 ml/mmol;

bUnder air condition;

cMicrowave

irradiation; dIsolated yield after chromatography;

eIn present of molecular sieve 4Å (10 equiv.);

f Using 213

(10 equiv.); gSilica gel (SiO2, MW=60, size 230-400 mesh) was added to the reaction mixture after heating for

16 h.

OH

OHMe

Me

SiO2 (20 equiv.)

DMF (1 ml/ mmol)

140 oC, 8 h

96%

OMe

Me

214 215

Scheme 4.4. Controlled experiment

Entry Catalysta

(mol %) Base Additive Solvent Temp

(°C)

Time

(h)

Yield (%)d

(equiv.) (equiv.) (214:215)

1 PdCl2 (1) 81 K2CO3 (1.5) n-Bu4NCl (1.5) toluene 100 24 35:5

2 PdCl2 (5) 81 K2CO3 (2) n-Bu4NCl (1.5) toluene 100 24 40:10

3 PdCl2 (10) 81 K2CO3 (3) n-Bu4NCl (1.5) toluene 100 24 45:10

4 PdCl2 (10) 81 K2CO3 (3) n-Bu4NCl (1.5) DMF 140 24 40:20

5b PdCl2 (10) 81 K2CO3 (3) n-Bu4NCl (1.5) DMF 150 24 43:23

6 PdCl2 (10) 81 K2CO3 (3) n-Bu4NCl (1.5) DMSO 150 24 45:25

7c PdCl2 (10) 81 K2CO3 (3) n-Bu4NCl (1.5) DMF 140 0.5 65:15

8c PdCl2 (10) 81 K2CO3 (3) MS 4Å (10)

e DMF 140 0.5 70:10

9 Pd(OAc)2 (5) 81 K2CO3 (3) n-Bu4NCl (1.5) DMF 140 24 52:15

10 Pd(OAc)2 (10) 81 K2CO3 (3) n-Bu4NCl (1.5) DMF 140 24 55:20

11 Pd(TFA)2 (10) 81 K2CO3 (3) - DMF 140 24 40:12

12 Pd(PPh3)4 (10) 81 K2CO3 (3) - DMF 140 24 35:5

13 Pd(dba)2 (10) 81 K2CO3 (3) - DMF 140 24 38:5

14 Pd(OAc)2 (10) 81 K2CO3 (3) - DMF 140 24 55:15

15 Pd(OAc)2 (10) 81 KOAc (3) - DMF 140 24 40:8

16 Pd(OAc)2 (10) 81 Na2CO3 (3) - DMF 140 24 35:5

17 Pd(OAc)2 (10) 81 Et3N (3) - DMF 140 24 25:3

18f Pd(OAc)2 (10) 212 K2CO3 (3) - DMF 140 24 10: -

19 Pd(OAc)2 (10) 81 K2CO3 (3) SiO2 (10)g DMF 140 24 -:76

20 Pd(OAc)2 (10) 81 K2CO3 (3) SiO2 (20)g DMF 140 24 -:80

95

Subsequently we investigated the condensation reaction of o-iodophenol 212 with

1,1-dimethyl allylic acetate (2-methylbut-3-en-2-yl acetate) 213. Li and co-workers

described a selective Heck reaction of 4-iodoanisole with 213 to give (E)-4-(4-

methoxyphenyl)-2-methylbut-3-en-2-ol in quantitative yield (Liu et al. 2011). Thus we

anticipated that the condensation of 212 with 213 could enhance the formation of 214 and

215. However, condensation of 212 with 213 under the present optimal reaction condition

(Pd(OAc)2 (10 mol%), K2CO3 (3 equiv.), DMF (2M), 140 C) significantly decreased the

yield of 214 and 215 (Table 4.1, entry 18), accompanied with the formation of undesired

side-product.

Since present condensation reaction of 214 and 215 did not proceed well at elevated

temperature or under microwave irradiation, we proceeded to investigated the condensation

under acidic condition, i.e. AcOH (David Cotterill, Iqbal, & Livingstone, 1998), H3PO4

(Goujon, Zammattio, & Kirschleger, 2000; Yus, Foubelo, & Ferrández, 2001).

Considering sensitivity of some functional groups (i.e. ethoxy methoxy group, EOM) in

acidic condition, we decided to use a milder reagent such as molecular sieve or silica gel

(SiO2). However, the condensation reaction of 214 and 215 did not proceed well in present

of molecular sieve (Table 4.1, entry 8). To our delight, the condensation reaction

proceeded extremely well in presence of SiO2 to afford 215 in good yield (76 – 80%, Table

4.1, entries 19-20). A controlled experiment was carried out to examine the role of SiO2 as

a dehydrating agent (Scheme 4.4). Treatment of allylic alcohol 214 in the present of SiO2

afforded chromene 215 in nearly quantitative yield under the standard conditions.

96

Table 4.2. Pd-catalysed condensation of o-halophenols 212 and 2-methyl-3-buten-2-ol 82a

aReaction conditions: Halide (1.0 mmol), 82 (10 mmol), Pd(OAc)2(10 mol %), K2CO3 (3.0 equiv), and DMF (1 mL) at 140 C under air for 8 ~ 14 h. The reaction was monitored by TLC until halide was completely consumed. Silica gel (20 equiv.) then was added to the

reaction mixture, followed by heating at 140°C for another 4 ~ 16h to yield desired product. bIsolated yield.

Entry Halide 1 t

(h)

Product Yield

(%)

1

216

15

215

75

2

217

15

218

80

3

219

12

220

60

4

221

20

222

25

5

223

20

224

70

6

225

20

226

15

7

202

18

227

85

8

228

30

229

17

9

211

24

209

65

97

Consequently, the scope of condensation reaction of o-iodophenols was explored in

order to determine its optimised conditions (Table 4.2). In the presence of Pd(OAc)2 and

K2CO3, a variety of o-iodophenols were reacted with 2-methyl-3-buten-2-ol 82 (Table 4.2,

entries 1-9). The results indicated both electron-rich and electron-deficient aryl iodides to

be suitable for the reaction. Aryl bromide 216 has similar activity as iodide 212. However,

aryl bromides bearing aldehyde and naphthol moieties gave lower product yield (Table 4.2,

entries 4 & 8).

With regards to the synthesis of natural products, isoencecalin (Manners & Jurd,

1976) 220 and 2-demethoxymillepachine 209, it is worthy to note that the present protocol

allows for the installation of 2,2-dimethylchromene ring at the late-stage of the synthesis.

In contrast, the conventional approach, such as in the synthesis of millepachine, the

preparation of 2,2-dimethylchromene ring was performed at an earlier stage of the synthesis

(G. Wang et al., 2012). The late-stage installation of 2,2-dimethylchromene ring was

successfully applied in the synthesis of 209 from chalcones 211 (Scheme 4.3).

4.1.4 Cycloaddition reaction of 106 and 209

With the dienophile 209 in hand, we began to investigate the [4+2] cyloaddition

with diene 106. However, chalcone 209 failed to undergo cycloaddition with diene 106 at

150 oC in refluxing toluene even after several days of reaction (Scheme 4.5). We reasoned

that the ortho OH in chalcone substituent 209 was critical for the success of the [4+2]

cycloaddition as previously described in Chapter 3.

98

O O

OMe

Me Me

OMe

MeO

O OMe

OMe

OMe

MeO

Me

+

toluene

150 oC

endo-230

Me

O

O

Me

Me

OMe

OMe

OMe

OMe

OMe

O

OMe

MeO

x

106

209

exo-231

Me

O

O

Me

Me

OMe

OMe

OMe

OMe

OMe

O

OMe

MeO

+

Scheme 4.5. Cycloaddition reaction of diene 106 with dienophile 209

4.2 Cycloaddition reaction of diene 233 and dienophile 234

The presence of an ortho-OH group in dienophile 209 was found to be essential for

the success of the [4+2] cycloaddition reaction. Thus we directed our attention to a

synthesis of dienophile with an ortho-OH group. Two different protecting groups (EOM

and OMe) were used to protect the other OH groups in dienophile.

4.2.1 Synthesis of diene 233

Synthesis of diene 233 is outlined in Scheme 4.6. Regioselective iodination of

commercially available 2’,4’-dihydroxyacetophenone 77 with I2/KIO3 in EtOH/H2O gave

78. Ethoxymethoxy protection followed by methylation of the phenolic groups of 78 gave

99

171. Claisen-Schmidt condensation between acetophenone 171 and benzaldeyde 198

furnished the chalcones 232 in 50% yield. Heck coupling of 232 with 2-methyl-but-3-en-2-

ol and subsequent dehydration with AcCl/pyridine provided the desired chalcones diene

233 in 55% combined yield in two steps. However, diene 233 was found to be unstable and

used immediately after silica gel purification.

HO

OH

Me

O

HO

OH

Me

O

I

EOMO

OMe

Me

O

I

i ii, iii

EOMO

OMe O

I

EOMO

EOMO

H

O

iv

OMOE

OMOE

v, vi

EOMO

OMe O OMOE

OMOE

Me

77 78 171

198

232 233

Scheme 4.6. Synthesis of diene 233 Reagents and conditions: i) I2, KIO3, EtOH, H2O, rt, 24 h, 92 % ii) EOM-

Cl, K2CO3, acetone, -10 oC, 15 h, 72% iii) Me2SO4, K2CO3, acetone, rt, 12 h, 85 % iv) 50 % aqueous KOH,

40 h, 50% v) 2-methyl-but-3-en-2-ol (82), 10 mol% Pd(OAc)2, K2CO3, DMF, 100 oC, 6 h vi)AcCl, pyridine,

C6H6, 80 oC, 55 % combined yield (two steps)

4.2.2 Synthesis of dienophile 234

Synthesis of EOM protected dienophile 234 is outlined in Scheme 4.7. Iodination of

a commercially available acetophenone 77, followed by EOM protection gave

acetophenone 217. Claisen-Schmidt condensation with the benzaldehye 198 gave the

desired dienophile 234 in 63% yield.

100

O Me

OH

OH

O Me

OH

OH

I2, KIO3

EtOH, H2Ort, 24 h, 92% I

OMOE

OMOE

O H

KOH, EtOH

acetone, -10oC

15 h, 85%

EOM-Cl, K2CO3

O Me

OH

OMOE

I

rt, 50%

EOMO

I

OH O OMOE

OMOE

76 77 216

233

197

Scheme 4.7. Synthesis of dienophile 234

However, subjecting the dienophile 234 to thermal Diels-Alder reaction with

diene 233 resulted in decomposition (Scheme 4.8). Similarly, the Diels-Alder reaction

between diene 233 and dienophile 234 performed in the presence of catalytic amount of

using ZnI2/BuNBH4 and AgBF4/BuNBH4 catalysts also resulted in the decomposition.

O OH

OMOE

OMOE

EOMO

O OMe

OMOE

OMOE

EOMO

Me

+

toluene

150 oC

Me

O

HO

I

OMOE

OMOE

OMOE

OMe

OMOE

OOMOE

EOMO

x

I

Me

O

HO

I

OMOE

OMe

OMOE

OOMOE

EOMO

+

233

234

endo-235

exo-236

OMOE

OMOE

Scheme 4.8. Cycloaddition reaction of diene 233 with dienophile 234

101

4.3 Cycloaddition reaction of diene 106 and dienophile 211

Since the use of EOM protecting group was unstable for the Diels-Alder reaction

between diene 233 and dienophile 234, we decided to change it to a more thermally stable

protecting group with a methyl ether in both diene and dienophile. When the Diels-Alder

reaction between diene 106 and dienophile 211 was performed in toluene at 150 oC in a

pressure tube, a mixture of endo-237 and exo-238 in ratio 3:2 was obtained, respectively

(Scheme 4.9). These adducts were separated by flash chromatography with 30%

EtOAc/hexane as an eluent to give pure endo-237 and exo-238.

O OH

OMe

OMe

MeO

O OMe

OMe

OMe

MeO

Me

+

toluene

150oC

Me

O

HO

I

OMe

OMe

OMe

OMe

OMe

OOMe

MeO

I

Me

O

HO

I

OMe

OMe

OMe

OMe

OMe

OOMe

MeO

+

OOH

MeO

OMe

OMe

O OMe

OMe

OMe

MeOMe

I

endo-addition

OOH

MeO

OMe

OMe

O OMe

OMe

OMe

MeO

Me

I

exo-addition

55%

3:2

106

211

endo-237exo-238

pressure tube

Scheme 4.9. Diels-Alder reaction between diene 106 and dienophile 211

102

Next, we investigated the removal of methyl ether protecting groups in endo-237.

Attempts to use Lewis acid BCl3 at different conditions (-78oC to room temperature) in

CH2Cl2 were unsuccessful. In all cases, endo-237 decomposed (Scheme 4.10).

Me

O

HO

I

OMe

OMe

OMe

OMe

OMe

OOMe

MeOBCl3, CH2Cl2-78 oC - r.t

Me

O

HO

I

OH

OH

OH

OH

OH

OOH

HO

X

237 239

Scheme 4.10. Attempts to remove the OMe groups of endo-237

4.4 Cycloaddition reaction of diene 106 and dienophile 89

Next, we attempted the cycloaddition reaction of diene 106 and dienophile 89. The

dienophile 89 was synthesized from chalcones 88 as shown in Scheme 4.11. Selective

demethylation of the 2’-methoxy group of chalcone 65 afforded chalcone 88. Installation of

prenyl group was applied in chalcones 88 following the method of Romano et al. O-

prenylation of 88 with prenyl chloride in refluxing acetone gave the prenyl ether 100.

Subsequently, the prenyl ether 100 then was subjected to a Montmorillonite K10 promoted

[1,3]-sigmatropic rearangement to give the dienophile 89 in 45% yield.

103

O Me

OH

OH

OMe

OMe

O H

KOH, EtOHacetone, rt15 h, 80%

CH3I, K2CO3

O Me

OMe

OMe

rt, 70%MeO

OMe O OMe

OMe

BCl3, CH2Cl2

0 oC- r.t, 45%MeO

OH O OMe

OMe

MeO

O O OMe

OMe

Me

Me

Cl Me

Me

K2CO3, acetonereflux, 84%

MeO

OH O OMe

OMe

Me

Memontmorillonite K10

CH2Cl2, 0oC,

2 h, 45%

77 240 65

88

100 89

99

Scheme 4.11. Synthesis of dienophile 89

With the dienophile 89 in hand, attention was turned to the Diels-Alder reaction.

Thermal Diels-Alder reaction between 106 and dienophile 89 in toluene at 150 oC in a

pressure tube gave a mixture of the cycloaddition products, endo-241 and exo-242 in 35%

yield in a 1:1 ratio, respectively (Scheme 4.12). The cycloaddition adducts were separated

by flash chromatography with 30% EtOAc/hexane as an eluent to give pure endo-241 and

exo-242.

104

O OH

OMe

OMe

MeO

O OMe

OMe

OMe

MeO

Me

+

toluene

150oC

Me

O

HO OMe

OMe

OMe

OMe

OMe

OOMe

MeOO

HO OMe

OMe

OMe

OMe

OMe

OOMe

MeO

+

O OMe

OMe

OMe

MeOMe

endo-addition

OOH

MeO

OMe

OMe

O OMe

OMe

OMe

MeO

Me

exo-addition

35%

1:1

Me

Me

Me

Me

MeMe MeMe

OHO

MeO

MeO

OMe

Me

Me

106

89

endo-241 exo-242

pressure tube

Scheme 4.12. Diels-Alder reaction between diene 106 and dienophile 89

Installation of the 2,2-dimethylchromene ring (pyran moiety) was performed on

adducts 241 and 242 with 20 mol% PdCl2 in dry ethanol afforded the desired intermediates

230 and 231 in 50% and 45% yield, respectively (Scheme 4.13) after a long period of

reaction.

105

Me

O

HO OMe

OMe

OMe

OMe

OMe

OOMe

MeO

MeMe

20 mol% PdCl2

dry ethanol, r.t 120 h, 50%

dry ethanol, r.t 216 h, 45%

20 mol% PdCl2

Me

O

O OMe

OMe

OMe

OMe

OMe

O

OMe

MeO

Me

Me

242

231

Me

O

O OMe

MeO

OMe

OMe

OMe

O

OMe

MeO

Me

Me

230

Me

O

HO OMe

OMe

OMe

OMe

OMe

OOMe

MeO

MeMe

241

Scheme 4.13. Installation of 2,2-dimethyl chromene in endo-241 and exo-242

Attempt to remove the methyl ether groups of endo-230 and exo-231 with several

reagents such as BCl3, MgI2, TMSI quinoline only resulted in decomposition (Table 4.3).

Table 4.3 Attempt to remove the OMe groups of endo-230

Me

O

O OMe

OMe

OMe

OMe

OMe

OOMe

MeOsee table

Me

O

O OH

OH

OH

OH

OH

OOH

HO

endo-230 243

Entry Reagents Solvent Conditions Result

1 BCl3 CH2Cl2 -78 oC to rt Decomposition

2 MgI2 Et2O:THF 60 oC Decomposition

3 TMSI quinoline net 130 oC Decomposition

106

In summary, a simple and efficient method for preparing 2,2-dimethyl-2H-

chromenes via Pd(II)-catalysed Heck coupling has been developed. Importantly, this

protocol provides a facile synthesis of naturally occurring 2,2-dimethyl-2H-chromenes in

the late stage of synthesis of sorocein B.

Deprotection of endo-237, endo-241 and exo-242 with several reagents such as BCl3,

MgI2, TMSI quinoline were not successful, which only resulted in decomposition. At this

stage, global deprotection remains to be achieved. Effort towards synthesis of sorocien B

was underway.

107

CHAPTER 5

CONCLUSION AND FUTURE WORK

5.1 Conclusion

In conclusion, we have successfully developed a viable method for the synthesis of

morusalbanol A. A model study on construction of the oxabicylic [3.3.1] core system of

morusalbanol A have been developed. The required cis-trans Diels-Alder adduct precursor

of morusalbanol A was obtained via thermal cycloaddition reaction which was proven to be

dependent on the presence of a hydrogen-bonded ortho OH substituent on the chalcone

dienophile. Acid catalyzed intramolecular cyclization of a cis-trans Diels-Alder adduct

afforded the desired oxabicyclic [3.3.1] compound in a stereo-controlled manner.

Additionally, rotation about the C3´´-C3´ bond of a para methyl ether protected, cis-trans

Diels-Alder adduct (endo-186) was observed during acid catalyzed intramolecular

cyclization to form the respective oxabicyclic [3.3.1] core of morusalbanol A.

The strategies that developed from the model study were successfully applied to the

synthesis of ()-morusalbanol A methyl ethers. The key step involved a biomimetic

cycloaddition reaction between a chalcones dienophile and a dehydroprenyl diene.

Intramolecular cyclization of the cis-trans Diels-Alder adduct afforded (±)-morusalbanol A

methyl ethers (endo-204) in a stereo-selective manner. A number of key proton and carbon

signals in the NMR spectra of ()-morusalbanol A methyl ethers were found absent

presumably due to atroisomerism. Global demethylation on (±)-morusalbanol A methyl

ethers (endo-206 or endo-207) with MgI2, BCl3, and 1-trimethylsilylquinolinium iodide

(TMSI-quinoline) were unsuccessful.

108

During the course of synthesis of sorocein B, a method for preparing 2,2-dimethyl-

2H-chromenes via Pd(II)-catalyzed Heck coupling has been developed. The procedure is

simple and efficient. Importantly, the protocol provides a facile synthesis of naturally

occurring 2,2-dimethyl-2H-chromenes, including the precursor of sorocein B.

Similar strategy was used for construction of the cis-trans Diels-Alder precursor of

sorocein B. The thermal cycloaddition reaction between chalcones dienophile 89 and

dehydroprenyl diene 106 afforded the requisite cis-trans Diels-Alder precursor (endo-241)

in 35% yield along with the trans-trans Diels-Alder diastereomer. Subsequently PdCl2

catalyzed cyclization of the ortho-prenyl group of endo-241 afforded the required 2,2-

dimethylchromenyl ring in 230. However, to date attempts to remove the methyl ether

group of 230 to form sorocein B with BCl3, MgI2, and TMSI-quinoline were unsuccessful.

5.2 Suggestion for Future Work

Enantioselective total syntheses of other mulberry Diels-Alder adducts such as

Kuwanon I and J as well as Brosimones A and B utilized a chiral ligand/ boron Lewis acid

was recently reported by Lei’s group (Han, Jones & Lei, 2015). These chiral-boron

complexes promote asymmetric Diels-Alder cycloadition by significantly lower the energy

of the LUMO. Thus, we anticipates that these chiral-boron complexes chiral-boron

complexes would also be useful for the synthesis of morusalbanol A, sorocein B and other

related mulberry Diels-Alder adducts.

For the synthesis of morusalbanol A (42), diene 244 and dienophile 245 were

subjected to asymmetric Diels-Alder reaction as summarized in the Scheme 5.1. The chiral

ligand will strongly influenced the enantioslectivity of the cycloaddition reaction.

109

HO

OH

OH

MeO

O

O OH

OH

OH

HO

OH

HO

O

MeO

O

O

OH

OH

HO

HO

morusalbanol A (42)

Me

PgO OPg

OPg O

OMe

O OPg

PgO

OH

OPg

244

245

+

chiral-boron complex

O OPg

PgO

O

OPg

Me

PgOOPg

PgO O

OMesteric effect

Diels-Alder

B

OO

41

PhPh

Pg = protecting group

Scheme 5.1 Future work for enantioselective total syntheses of morusalbanol A (42)

110

Biosynthesis-inspired asymmetric Diels-Alder cycloaddition promoted by chiral

VANOL@ VAPOL/boron Lewis acid have been recently reported by Lei’s group (Gao,

Han & Lei, 2016). As depicted in Scheme 5.2, the endo-246 could be generated through

asymmetric Diels-Alder cycloadditions from diene 244 and dienophile 245 using (R)-

VANOL as a chiral ligand. With 246 in hand, sulphuric acid was employed to catalyze the

biomimetic intramolecular ketalisation. Following the Lei’s method for ketalisation and

applying our method for a late-stage pyranyl installation of 2,2-dimethylchromene ring on

248 will presumably help to furnish the enantioselective total synthesis of Sorocein B (62).

MOMO

MOMOO

OMOM

OMOM

AcO

OH O

I

OAc

OAc

endo-246

exo-247 (not shown for clarity)

O

HO

I

OAc

AcO

OAc

OMOM

OMOM

OOMOM

MOMO

+

i) (R)-Vanol, AcOH BH3.THF, rt, 25 minii) dienophile, rt, 1.5 h

iii) diene, rt

diene 244

dienophile 245

endo-24610% H2SO4

EtOH, rtO O

OH

OH

OOH

HO

OH

I

OH

HH

H

O O

OH

OH

OOH

HO

OH

O

HH

H

late stage pyran installation

our method

sorocein B (62)

248

Scheme 5.2. Future work for enantioselective total syntheses of sorocein B (62)

111

CHAPTER 6

EXPERIMENTALS

Chemical reagents were purchased from Aldrich, Acros and Merck and were used

as received. Anhydrous solvents were purchased from Merck. HPLC grade tetrahydrofuran,

methylene chloride, diethyl ether, toluene, and benzene were purified and dried by passing

through a PURESOLV® solvent purification system (Innovative Technology, Inc.).

Acetone (HPLC grade) was dried by distillation from activated molecular sieves 4Å. NMR

spectra were obtained using a Jeol ECA 400 (400 MHz) NMR spectrometer with TMS as

the internal standard. All measurements were accomplished in solution in DMSO-d6 or

CDCl3. Chemical shifts are reported in parts per million relative to CDCl3 or TMS. Data

for 1H NMR are reported as follows: chemical shift, integration, multiplicity (br s = broad

singlet, s = singlet, d = doublet, dd = doublet of doublets, ddd= doublet of doublet of

doublets, t = triplet, q = quartet, m = multiplet) and coupling constants. All 13

C NMR

spectra were recorded with complete proton decoupling. Analytical thin layer

chromatography (TLC) was carried out on Merck precoated aluminium silica gel sheets

(Kieselgel 60 F254). Visualization was accomplished under UV light. All products were

purified by silica gel column chromatography using a hexane/EtOAc/chloroform as eluent.

All target compounds were characterized by 1H,

13C, 2D NMR and HRMS (ESI) analyses.

112

Crystallography

Diffraction data for the crystal of 184 was collected on a Bruker SMART Apex II

CCD area-detector diffractometer (graphite-monochromatized Mo-Kα radiation, λ =

0.71073 Å) at 150(2) K. The orientation matrix, unit cell refinement and data reduction

were all handled by the Apex2 software (SAINT integration, SADABS multi-scan

absorption correction). Diffraction data for the crystal of 186, 206, 207 were collected on

an Agilent SuperNova Dual diffractometer with an Atlas detector (graphite-

monochromatized Mo-Kα radiation, λ = 0.71073 Å) at 100(2) K. The data were processed

using CrysAlisPro, Agilent Technologies, Version 1.171.37.34 (release 22-05-2014

CrysAlis171 .NET) and empirical absorption correction using spherical harmonics,

implemented in SCALE3 ABSPACK scaling algorithm (184, 186 and 207) or SADABS

(for 206). The structures were solved using the program SHELXT (Sheldrick, 2012), and

refined by the full matrix least-squares method on F2 with SHELXL-2014/7 (Sheldrick,

2008). All the non-hydrogen atoms were refined anisotropically. All the C-bound hydrogen

atoms were placed at calculated positions and were treated as riding on their parent atoms.

The O-bound hydrogen atoms were found in difference Fourier maps. For all hydrogen

atoms Uiso(H) were set to 1.2-1.5 times Ueq(carrier atom).

The structure of 186 was determined from a two-component twinned crystal with

the twin parameter refined to 0.477(1). The chloroform solvate molecule is disordered over

two positions in a 0.64/0.36 ratio. For the crystal structure of 207, the difference Fourier

map shows one relatively large peak, 1.46 e Å-3

, located at 1.02 Å from H14A and 1.72 Å

from C14. The peak cannot be assigned to any chemical reasonable species in the structure

and could be due to an adventitious incorporation of an impurity during the crystallization

process.

113

Drawing of the molecule was produced with XSEED (Barbour, 2001) (184 and 186),

Mercury (Macrae et al. 2008) (206 and 207). Crystal data and structural refinement

parameters for 184 and 186 are given in Table 3.2, 206 and 207 are given in Table 3.3.

Computational Methodology

Molecular structure for the mechanism study was obtained from trans-chalcone

(CID=637760) with modification. Geometry of the modified structure was then optimized

by using gaussian 09 programme (Frisch et al. 2009). Density Functional Theory (DFT)

method at the B3LYP/6-31G(d,p) level (Becke, 1993; Lee, Yang, and Parr 1988). The

transition state (TS) investigation of the reactants and each of the possible products were

then performed on Material Studio programme 4.3. Geometry optimisation and TS

calculation were carried out using synchronous transit method that embedded in Dmol3

module (Delley, 2000; Delley, 1990) DFT with local density approximation (LDA) of local

functional PWC19

were employed in all the calculations considering the effective core

potential treatment along with the DN or DND basis set. Linear synchronous transit (LST)

accounted for the reaction paths. Meanwhile, optimisation calculation performs a single

interpolation to a maximum energy prior to quadratic synchronous transit (QST) method for

obtaining an energy maximum with constrained minimizations in order to refine the

transition state to a high degree (Perdew & Wang, 1992) Also, each point was treated with

another conjugate gradient minimization. The cycle keeps going until a stationary point is

located or the number of the allowed QST steps has been fully exhausted. The highest

energy points were then optimised to the closest transition state (TS) after the convergence

of the initial paths. Minimum energy path (MEP) between the critical points were

calculated with nudged elastic band (NEB) under TS Optimisation function to ensure the

continuity of the path and projection of the force in order to converge the system to the

114

MEP. On the other hand, potential energy scan was performed on the dihedral angle where

rotations occur under gaussian 09 programme using semi-empirical calculation at PM6

level.

1-(4-(ethoxymethoxy)-3-iodo-2-methoxyphenyl)ethanone (171)

To a solution of 1-(2,4-dihydroxy-3-iodophenyl)ethanone 78 (1.2 g, 4.3 mmol) and K2CO3

(1.5 g, 2.5 equiv) in dry acetone (150 mL) at -10oC was added EOM-Cl (0.44 mL, 4.7

mmol). The reaction mixture was stirred for 15 h. The mixture was diluted with EtOAc,

washed with a sat. NH4Cl aq. Solution, water, and dried over Na2SO4. The crude residue

was purified by flash chromatography eluting with 20% EtOAc/hexane to give 1-(4-

(ethoxymethoxy)-2-hydroxy-3-iodophenyl)ethanone (1.04 g, 72%) as colourless oil. 1H

NMR (400 MHz, CDCl3): δ 13.46 (s, OH, 1H), 7.61 (d, J = 8 Hz, H6, 1H), 6.59 (d, J = 8

Hz, H5, 1H), 5.30 (s, OCH2O, 2H), 3.77 (m, OCH2CH3, 2H), 2.55 (s, Ac, 3H), 1.15 (t, J =

6.9 Hz, OCH2CH3, 3H); 13

C NMR (100 MHz, CDCl3): δ 202.7, 163.4, 162.6, 132.5, 114.9,

105.6, 93.5, 78.1, 65.3, 26.3, 15.1; HRMS calcd for C11H14IO4 [M+H]+

336.9931. Found:

336.9926.

To a solution of 1-(4-(ethoxymethoxy)-2-hydroxy-3-iodophenyl)ethanone (426 mg, 1.27

mmol) and K2CO3 (0.44g, 2.5 equiv) in dry acetone (10 mL) was added dimethyl sulfate

(0.13 ml, 1.1 equiv). The mixture was stirred at r.t. for 12 h. Then, K2CO3 was filtered and

115

the mixture was diluted with EtOAc. The organic phase was extracted with water. The

organic layer was collected, washed with brine, dried on Na2SO4 and evaporated. The

residue was subjected to column chromatography (hexane: EtOAc = 8:2) to afford the

desired compound 171 (378 mg, 85%) as colourless viscous liquid. 1H NMR (400 MHz,

CDCl3): δ 7.61 (d, J = 8.7 Hz, H6, 1H), 6.85 (d, J = 8.7 Hz, H5, 1H), 5.28 (s, OCH2O, 2H),

3.78 (s, OCH3, 3H), 3.71 (m, OCH2CH3, 2H), 2.56 (s, Ac, 3H), 1.15 (t, J = 7.3 Hz,

OCH2CH3, 3H); 13

C NMR (100 MHz, CDCl3): δ 198.1, 160.9, 160.8, 131.9, 127.3, 110.2,

93.6, 86.4, 65.1, 62.6, 24.9, 15.1; HRMS calcd for C12H16IO4 [M+H]+

351.0088. Found:

351.0094.

(E)-1-(4-(ethoxymethoxy)-2-methoxy-3-(3-methylbuta-1, 3-dienyl)phenyl)ethanone

(172)

To a mixture of 2-methyl-3-buten-2-ol (82) (0.52 ml, 5 mmol), Pd(Cl)2 (10 mg, 0.05

mmol) , nBu4NCl (300 mg, 1.08 mmol), NaHCO3 (227 mg, 2.70 mmol), and 171 (378 mg,

1.08 mmol) of in 3 mL of DMF in a pressure tube was heated to 100°C for 12 h. The

mixture was purified by column chromatography eluting with 40% EtOAc/ hexane to yield

a pale yellow viscous liquid (300 mg, 90%). To this liquid (300 mg, 0.97 mmol) in 10 mL

of benzene was added acetyl chloride (87 µL, 1.22 mmol, 1.25 equiv) and pyridine (102 µL,

1.26 mmol, 1.3 equiv). The mixture was then heated at 60°C for 6h. The white precipitate

was filtered, and the organic layers were collected and dried. The residue was purified by

column chromatography (hexane: EtOAc =7:3) to afford 254 mg of diene 1a (90% yield) as

pale yellow viscous liquid. 1H NMR (400 MHz, CDCl3): δ 7.54 (d, J = 9.2 Hz, H6, 1H),

7.25 (d, J = 16.9 Hz, H7, 1H), 6.94 (d, J = 9.2 Hz, H5, 1H), 6.67 (d, J = 16.5 Hz, H8, 1H),

116

5.29 (s, OCH2O, 2H), 5.09 (bs, H11, 2H), 3.73 (m, OCH2CH3, 2H), 3.72 (s, OCH3, 3H),

2.61 (s, Ac, 3H), 1.99 (s, CH3, 3H), 1.21 (t, J = 7.3 Hz, OCH2CH3, 3H); 13

C NMR (100

MHz, CDCl3): δ 199.2, 159.5, 143.1, 137.3, 129.7, 127.1, 120.9, 119.2, 117.7, 110.3, 93.4,

64.9, 61.9, 30.6, 18.3, 15.1; HRMS calcd for C17H23O4 [M+H]+

291.1591. Found: 291.1589.

Endo-179 and exo-180

Diene 172 (160 mg, 0.55 mmol) and dienophile 175 (112 mg, 0.50 mmol) was dissolved in

freshly distilled toluene (1 mL) in a pressure tube. The mixture was heated at 150 oC for 24

h, which afforded a mixture of inseparable endo-179 and exo-180 (ratio 179:180 = 3:2, 140

mg, 54.5% yield). The mixture of endo-179 and exo-180 in 3M aqueous HCl in methanol

was heated for 20 min before addition of water. The residue was extracted with CH2Cl2 (2x).

The organic layers were combined, washed with brine, dried over MgSO4, filtered and

concentrated. The crude residue was purified by flash chromatography eluting with 30%

EtOAc/ hexane to give diastereomers 183 (62 mg) and 184 (50 mg)(90% combined yield)

as viscous liquid.

Diastereomer 183 was obtained as a pale yellow viscous liquid. 1H NMR (400 MHz,

CDCl3): δ 11.94 (s, OH, 1H), 7.89 (d, J = 8.3 Hz, H14”, 1H), 7.48 (d, J = 8.8 Hz, H6’, 1H),

7.40 (t, J = 7.5 Hz, H11”, 1H), 7.13-7.02 (m, H16”-H20”, 5H), 6.91-6.88 (m, H12”, H13”,

2H), 6.62 (d, J = 8.9 Hz, H5’, 1H), 4.11 (dd, J = 2.7, 11.6 Hz, H4”, 1H), 3.64 (d, J = 3.0 Hz,

H3”, 1H), 3.18 (ddd, J= 4.6, 12.5, 16.5 Hz, H5”, 1H), 3.13 (s, OCH3, 3H), 2.44 (s, Ac, 3H),

117

2.27-2.20 (m, H2”, H6”, 2H), 1.98-1.93 (m, H2”, H6”, 2H), 1.41 (s, CH3, 3H); 13

C NMR

(100 MHz, CDCl3): δ 204.4, 199.7, 162.8, 160.6, 159.0, 143.1, 135.8, 130.7, 128.7, 128.5,

128.4, 127.3, 126.6, 124.3, 120.0, 118.7, 118.6, 115.0, 112.0, 75.3, 62.0, 53.1, 48.3, 37.1,

36.1, 32.0, 29.7, 28.5; HRMS calcd for C29H29O5 [M+H]+

457.2010. Found: 457.2018.

Diastereomer 9a was obtained as a pale yellow viscous liquid. Recrystallisation in

methanol gave a pale yellow crystal in very small quantity (not enough for melting point

determination). 1H NMR (400 MHz, CDCl3): δ 12.45 (s, OH), 7.69 (d, J = 8 Hz, H14”, 1H),

7.45 (d, J = 8.8 Hz, H6’, 1H), 7.33 (t, J = 7.6 Hz, H11”, 1H), 7.02 (m, H17”, H19”, 2H),

6.93 (m, H16”, H18”, H20”, 3H), 6.87 (d, J = 8.4 Hz, H12”, 1H), 6.67 (t, J = 7.6 Hz, H13”,

1H), 6.54 (d, J = 8.8 Hz, H5’, 1H), 3.91 (d, J = 8.8 Hz, H4”, 1H), 3.46 (s, OCH3, 3H), 3.30

(br s, H3”, 1H), 3.24 (m, H5”, 1H), 2.50 (s, Ac, 3H), 2.28 (m, H2”, 2H), 2.17 (m, H6”, 2H),

1.55 (s, CH3, 3H); 13

C NMR (100 MHz, CDCl3): δ 209.0, 197.4, 162.0, 157.6, 157.3, 141.5,

135.6, 130.1, 129.4, 127.2, 125.9, 125.4, 122.9, 119.4, 117.7, 117.5, 117.0, 112.3, 74.4,

61.1, 52.4, 39.3, 37.3, 29.8, 28.9, 28.3, 28.2; HRMS calcd for C29H29O5 [M+H]+ 457.2010.

Found: 457.2022.

The procedure using AgOTf and AgBF4 for exo-adduct 180

118

To a solution of diene 172 (60 mg, 0.2 mmol) and dienophile 175 (46 mg, 0.2 mmol) in dry

CH2Cl2 (1mL) was added AgOTf (15 mg, 60 μmol). The reaction mixture was stirred at rt

for 1h. The crude residue was purified by flash chromatography with 30% EtOAc/ hexane

to give exo-180 (39 mg, 38% yield) as a pale yellow viscous liquid. 1H NMR (400 MHz,

CDCl3): δ 11.99 (s, OH, 1H), 7.75 (br d, J = 6.9 Hz, H14”, 1H), 7.37 (d, J = 9.1 Hz, H6’,

1H), 7.31 (t, J = 7.6 Hz, H11”, 1H), 7.02-7.19 (m, H16”, H17”, H19”, H20”, 4H), 6.76-6.73

(m, H12”, H13”, H5’, 3H), 5.49 (br s, H2”, 1H), 4.88 (br s, OCH2O, 2H), 4.55 (br s, H3”,

1H), 4.32 (t, J = 9.6 Hz, H4”, 1H), 3.92 (ddd, J = 6.9, 9.6, 16.0 Hz, H5”, 1H), 3.56 (m,

OCH2CH3, 2H), 3.44 (s, OCH3, 3H), 2.58 (dd, J = 9.6, 17.8, H6”, 1H), 2.44 (s, Ac, 3H),

2.31 (dd, J = 9.1, 17.8 Hz, H6”, 1H), 1.83 (s, CH3, 3H), 1.15 (t, J = 6.9 Hz, OCH2CH3, 3H);

13C NMR (100 MHz, CDCl3): δ 207.3, 200.5, 162.2, 160.1, 145.8, 135.6, 133.3, 130.4,

130.0, 128.5, 127.6, 126.3, 123.1, 122.0, 118.4, 118.0, 109.5, 92.9, 64.4, 63.2, 50.0, 38.9,

34.8, 29.8, 29.7, 23.6, 15.1; HRMS calcd for C32H35O6 [M+H]+ 515.2434, Found: 515.2440.

(E)-1-(2-hydroxy-4-methoxy-3-(3-methylbuta-1,3-dienyl)phenyl)ethanone (174)

To a mixture of 2-methyl-3-buten-2-ol (81) (0.72 ml, 6.85 mmol), Pd(Cl)2 (12 mg, 0.07

mmol), nBu4NCl (190 mg, 0.69 mmol), NaHCO3 (145 mg, 1.73 mmol), and 1-(2-hydroxy-

3-iodo-4-methoxyphenyl)ethanone (200 mg, 0.68 mmol) in 5 mL of DMF was heated at

100°C in a pressure tube for 4 h. The mixture was subjected to column chromatography

119

eluting with 40% EtOAc/ hexane to give a pale yellow liquid (144 mg, 85%). To this

yellow liquid (144 mg, 0.52 mmol) in 10 mL benzene was added acetyl chloride (48 µL,

0.68 mmol, 1.2 equiv) and pyridine (53 µL, 0.68 mmol, 1.2 equiv). The mixture was

heated at 80°C for 2h. Formation of white precipitate was filtered and the organic layer was

collected. The residue was purified by column chromatography (hexane: EtOAc =7:3) to

afford diene 174 (122 mg, 91% yield) as pale yellow viscous liquid. 1H NMR (400 MHz,

CDCl3): δ 13.32 (s, OH, 1H), 7.58 (d, J = 9.2 Hz, H6, 1H), 7.41 (d, J = 16.9 Hz, H7, 1H),

6.80 (d, J = 16.5 Hz, H8, 1H), 6.50 (d, J = 9.2 Hz, H5, 1H), 5.05 (s, H11, 2H), 3.92 (s,

OMe, 3H), 2.55 (s, Ac, 3H), 1.98 (s, Me, H10, 3H); 13

C NMR (100 MHz, CDCl3): δ 203.3,

163.3, 162.5, 143.5, 136.3, 131.0, 118.6, 117.1, 114.3. 113.9, 102.2, 56.0, 26.4, 18.4;

HRMS calcd for C14H17O3 [M+H]+

233.1172, Found: 233.1166.

Endo-181 and exo-182

Diene 174 (128 mg, 0.55 mmol) and dienophile 176 (149 mg, 0.50 mmol) was dissolved in

fresh distilled toluene (1 mL) in a pressure tube. The mixture was heated at 150 oC for 25 h,

which afforded a mixture of inseparable endo-181 and exo-182 (ratio 181:182 = 1:1, 130

mg, 49% yield). The mixture of endo-181 and exo-182 in 5 mL, 3M aqueous HCl in

methanol was heated for 20 min before addition of water. The residue was extracted with

CH2Cl2 (2x). The organic layers were combined, washed with brine, dried over MgSO4,

filtered and concentrated. The crude residue was purified by flash chromatography eluting

with 30% EtOAc/ hexane to give diastereomer 186 (53 mg, 92% based on conversion of

endo-181, which was 65 mg) and the recovery of 182 (40 mg).

120

Diastereomer 186 was obtained as a pale yellow viscous liquid. Recrystallisation in

methanol gave a pale yellow crystal in very small quantity (not enough for melting point

determination). 1H NMR (400 MHz, CDCl3 + MeOD-d4): δ 12.03 (s, OH, 1H), 7.88 (d, J =

7.8 Hz, H14”, 1H), 7.58 (d, J = 8.9 Hz, H6’, 1H), 7.30 (t, J = 7.3 Hz, H11”, 1H), 6.84-6.74

(m, H16”, H17”, H18”, H19”, 4H), 6.51-6.49 (m, H12”, H13”, 2H), 6.20 (d, J = 8.7 Hz,

H5’, 1H), 4.50 (br s, H4”,1H), 3.64 (s, H3”, 1H), 3.30 (br s, H5”, 1H), 3.19 (s, OCH3, 3H),

2.49 (s, Ac, 3H), 2.17-2.14 (m, H2”, H6”, 2H), 1.96 (m, H2”, 1H), 1.78 (d, J =12.8, H6”,

1H), 1.36 (s, CH3, 3H); 13

C NMR (100 MHz, CDCl3): δ 206.0, 199.9, 161.9, 160.8, 157.6,

154.6, 135.7, 130.9, 129.1, 128.8, 127.2, 119.8, 119.7, 119.4, 119.0, 117.9, 115.9, 110.3,

101.4, 76.1, 54.6, 50.8, 44.8, 35.3, 31.9, 31.1, 28.4 (C5 was not observed); HRMS calcd for

C29H29O6 [M+H]+

473.1959. Found: 473.1967.

exo-189 was obtained as a pale yellow viscous liquid. 1H NMR (400MHz, CDCl3): δ 13.0

(s, OH, 1H), 11.95 (s, OH, 1H), 7.33-7.28 (m, H6’, H14”, 2H), 7.17 (d, J = 9.7, H16”, 1H),

7.05 (t, J = 7.8 Hz, H11”, 1H), 6.95-6.88 (m, H19”, 1H), 6.71 (m, H12”, H5’, 2H), 6.56 (d,

J = 8.2 Hz, H17”, 1H), 6.28 (t, J = 7.4 Hz, H13”, 1H), 5.97 (d, J = 8.8 Hz, H18”, 1H), 5.26

121

(s, H2”, 1H), 4.56 (br s, H4”, 1H), 4.27 (br d, H3”, 1H), 3.89 (m, H5”, 1H), 3.53 (s, OCH3,

1H), 2.43 (s, Ac, 3H), 2.17 (m, H6”, 2H), 1.71 (s, CH3, 3H); 13

C NMR (100 MHz, CDCl3):

δ 212.1, 203.2, 163.0, 162.9, 161.6, 153.3, 136.2, 132.3, 131.4, 130.4, 127.6, 123.2, 121.4,

120.8, 118.0, 117.5, 117.2, 117.1, 114.3, 101.8, 55.8, 55.6, 48.5, 38.6, 36.8, 29.8, 26.3, 23.3;

HRMS calcd for C29H29O6 [M+H]+

473.1959. Found: 473.1959.

(E)-methyl-2,4,6-tri(ethoxymethoxy)-3-(3-methylbuta-1,3-dienyl)benzoate (190)

To a mixture of 2-methyl-3-buten-2-ol (82) (0.65 ml, 6.25 mmol), Pd(OAc)2 (29 mg, 0.13

mmol), K2CO3 (345 mg, 2.50 mmol), and 196 (530 mg, 1.25 mmol) of in 1 mL of DMF in

a pressure tube was heated to 100°C for 7 h. The mixture was purified by column

chromatography eluting with 40% EtOAc/ hexane to yield a pale yellow viscous liquid

(387 mg, 70%). To this liquid (387 mg, 0.88 mmol) in 10 mL of benzene was added acetyl

chloride (63 µL, 0.88 mmol, 1.00 equiv) and pyridine (71 µL, 0.88 mmol, 1.00 equiv). The

mixture was then heated at 60°C for 6h. The white precipitate was filtered, and the organic

layers were collected and dried. The residue was purified by column chromatography

(hexane: EtOAc =7:3) to afford 280 mg of diene 190 (75% yield) as pale yellow viscous

liquid. 1H NMR (400 MHz, CDCl3): δ 7.13 (d, J = 16.4 Hz, H7, 1H), 6.83 (s, H5, 1H), 6.59

(d, J = 16.4 Hz, H8, 1H), 5.23 (s, OCH2O, 2H), 5.18 (s, OCH2O, 2H), 5.01 (br s, H11, 2H),

5.00 (s, OCH2O, 2H), 3.87 (s, OCH3, 3H), 3.70 (m, OCH2CH3, 6H), 1.93 (s, CH3, 3H), 1.19

(m, OCH2CH3, 9H); 13

C NMR (100 MHz, CDCl3): δ 166.6, 157.7, 154.3, 153.8, 143.1,

122

135.8, 119.4, 116.8, 115.1, 114.1, 99.1, 98.8, 93.7, 93.6, 65.8, 64.5, 64.4, 52.3, 18.2, 15.1,

15.0, 14.9; HRMS calcd for C22H33O8 [M+H]+

425.2167. Found: 425.2169.

(E)-3-(2,3-bis(ethoxymethoxy)phenyl)-1-1(4-(ethoxymethoxy)-2-hydroxyphenyl)prop-

2-en-1-one (192)

To a solution of acetonphenone 197 (1.1 g, 5 mmol) and benzaldehyde 198 (1.3 g, 5 mmol)

in EtOH (25 mL) was added 50% aqueous KOH (84 mg, 15 mmol). The reaction mixture

was refluxed for 42 h before the addition of water and CH2Cl2. The aqueous phase was

extracted with CH2Cl2 and the combined organic layer were washed with brine, dried over

MgSO4, filtered and concentrated. The crude residue was purified by flash chromatography

eluting with 20% EtOAc/hezane to afford 192 (1.23 g, 55%) as yellow viscous liquid. 1H

NMR (400 MHz, CDCl3): δ 13.48 (s, OH, 1H), 8.17 (d, J = 15.6 Hz, H8, 1H), 7.82 (d, J =

9.2 Hz, H3, 1H), 7.58 (d, J = 16.0 Hz, H9, 1H), 7.53 (d, J = 7.3 Hz, H12, 1H), 6.89 (d, J =

2.8 Hz, H6, 1H), 6.74 (dd, J = 2.8, 9.2 Hz, H5, 1H), 6.63 (d, J = 2.3 Hz, H14, 1H), 6.57 (dd,

J = 2.7, 9.2 Hz, H15, 1H), 5.32 (s, OCH2O, 2H), 5.26 (s, OCH2O, 2H), 5.25 (s, OCH2O,

2H), 3.74 (m, OCH2CH3, 6H), 1.23 (m, OCH2CH3, 9H); 13

C NMR (100 MHz, CDCl3): δ

192.6, 166.2, 163.6, 160.9, 158.2, 140.4, 131.3, 130.1, 118.6, 118.4, 115.1, 109.5, 108.1,

104.0, 103.5, 93.5, 93.1, 92.9, 64.9, 64.8, 64.6, 15.2; HRMS calcd for C24H31O8 [M+H]+

447.2011. Found: 447.2025.

123

Methyl-3-iodo-2,4,6-trimethoxybenzoate (203)

To a solution of 202 (800 mg, 2.37 mmol) and K2CO3 (490 mg, 1.5 equiv) in dry acetone

(50 mL) was added iodomethane (0.16 mL, 1.1 equiv). The mixture was stirred at r.t. for 15

h. Then, K2CO3 was filtred and the organic phase was extracted with EtOAc and water. The

organic layers were collected, washed with brine, dried on MgSO4 and evaporated. The

residue was subjected to column chromatography (hexane: EtOAc= 8:2) to afford 792 mg

compound 203 (95% yield) as colorless viscous liquid. 1H NMR (400 MHz, CDCl3): δ 6.21

(s, H5, 1H), 3.84-3.79 (2s, 4xOMe, 12H); 13

C NMR (100 MHz, CDCl3): δ 165.9, 160.7,

158.9, 158.5, 111.6, 91.5, 72.7, 62.1, 56.5, 56.1, 52.4; HRMS calcd for C11H14IO5 [M+H]+

352.9878. Found: 352.9882.

(E)-methyl-2,3,6-trimethoxy-3-(3-methylbuta-1,3-dienyl)benzoate (191)

To a mixture of 2-methyl-3-buten-2-ol (1.5 ml, 14.2 mmol), Pd(OAc)2 (32 mg, 0.14 mmol),

K2CO3 (400 mg, 2.84 mmol), and 203 (500 mg, 1.42 mmol) of in 1 mL of DMF in a

pressure tube was heated to 100°C for 6 h. The mixture was purified by column

chromatography eluting with 50% EtOAc/hexane to yield pale yellow viscous liquid (242

mg, 55%). To this liquid (242mg, 0.78 mmol) in 10 mL of benzene was added acetyl

124

chloride (69 µL, 0.87 mmol, 1.1 equiv), and pyridine (61 µL, 0.87 mmol, 1.1 equiv). The

mixture was then heated to 60°C for 4h. The white precipitate was filtered, and the organic

layers were collected, dried. The residue was purified by column chromatography (hexane:

EtOAc =7:3) to afford 200 mg of diene 191 (90% yield) as pale yellow viscous liquid. 1H

NMR (400 MHz, CDCl3): δ 7.19 (d, J = 16.5 Hz, H7, 1H), 6.61 (d, J = 16.9 Hz, H8, 1H),

6.26 (s, H5, 1H), 5.03 (d, J = 7.8 Hz, H11, 2H), 3.9-3.7 (4s, 4xOMe, 12H), 1.96 (s, CH3,

3H); 13

C NMR (100 MHz, CDCl3): δ 167.0, 160.3, 157.4, 156.7, 143.4, 135.0, 119.0, 116.7,

112.7, 111.5, 91.6, 62.0, 56.1, 55.9, 52.6, 18.4; HRMS calcd for C16H20O5 [M+H]+

293.1381. Found: 293.1388.

Endo-204 and exo-205

Diene 191 (100 mg, 0.34 mmol) and dienophile 88 (157 mg, 0.34 mmol) were dissolved in

freshly distilled toluene (2 mL) in a pressure tube. The mixture was heated at 135oC for

24h to give a mixture of diastereomers 204 and 205. The mixture of diastereomers 204 and

205 was separated by column chromatography with 80% hexane/EtOAc as eluent to give a

mixture of diastereoisomers 204 (74 mg) and 205 (50 mg) (55% combined yield).

Endo-204 was obtained as a pale yellow viscous liquid. 1H NMR (400 MHz, CDCl3): δ

12.87 (s, OH, 1H), 7.51 (bd, H14”, 1H)*, 6.99 (d, J = 8.8 Hz, H20”, 1H), 6.33 (bd, H11”,

H13”, 2H)*, 6.24 (bd, H17”, H19”, 2H)*, 5.88 (s, H5’, 1H), 5.41 (s, H2”, 1H), 4.23 (bs,

H4”, 1H), 3.84-3.68 (5s, 5xOMe, 15H), 3.40 (s, OMe, 3H), 3.31(s, OMe, 3H), 2.59 (m,

125

H6”, 1H), 2.22(m, H6”, 1H), 1.74 (s, CH3, 3H); 13

C NMR (100 MHz, CDCl3): δ 206.5,

167.5, 165.0, 164.8, 160.9, 159.1, 157.9, 157.2, 132.8, 131.6, 126.8, 122.5, 115.3, 115.2,

110.1, 106.3, 104.2, 100.6, 98.9, 90.7, 62.0, 56.0, 55.5, 55.3, 54.7, 52.4, 47.7, 38.4, 36.0,

23.6; HRMS calcd for C34H38O10Na [M+Na]+ 629.2363. Found: 629.4568.

*coupling constant cannot be measured due to broad resonance

Exo-205 was obtained as a pale yellow viscous liquid. 1H NMR (400 MHz, CDCl3): δ

13.00 (1H, s), 7.39 (bd, H14”, 1H)*, 6.96 (d, J = 8.5 Hz, H20”, 1H), 6.25 (bd, H13”, H11”,

2H)*, 6.21 (bd, H17”, H19”, 2H)*, 6.12 (bd, H2”, 1H)*, 5.82 (s, H5’, 1H), 5.24 (s, H5”,

1H), 4.29 (bd, J = 9.8 Hz, H4”, 1H), 3.92 (s, OMe, 3H), 3.74-3.64 (6s, 6xOMe, 18H), 2.17

(m, H6”, 1H), 1.98 (m, H6”,1H), 1.63 (s, CH3, 3H); 13

C NMR (100 MHz, CDCl3): δ 207.7,

166.2, 164.2, 163.3, 158.5, 17.9, 157.1, 156.1, 131.4, 130.8, 130.5, 123.7, 123.2, 116.6,

114.4, 113.0, 105.8, 105.3, 103.2, 98.9, 97.8, 89.8, 61.0, 54.9, 54.3, 54.1, 51.4, 46.6, 37.3,

30.9, 28.7, 22.2; HRMS calcd for C34H38O10Na [M+Na]+ 629.2363. Found: 629.2402.

Morusalbanol A pentamethyl ether (206) and 207

To a solution of freshly prepared MgI2 (0.23 g, 0.82 mmol) in dry Et2O (10 mL) was added

endo-204 (0.05 g, 0.08 mmol) in THF (10 mL). The mixture was refluxed at 50 oC for 4 h.

Then, the precipitate was filtered, and the organic layers were collected and dried. The

residue was purified by column chromatography (hexane: EtOAc =8:2) to afford 25 mg of

126

Morusalbanol A pentamethyl ether (206) (50% yield) as colorless crystal in very small

quantity (not enough for melting point determination).

Morusalbanol A pentamethyl ether (206)

1H NMR (400 MHz, CDCl3): δ 12.75 (s, OH, 1H), 7.87 (d, J = 9.1 Hz, H14”, 1H), 6.90 (d,

J = 8.2 Hz, H20”, 1H), 6.50 (dd, J = 2.8, 9.1 Hz, H13”, 1H), 6.39 (d, J = 2.3 Hz, H11”, 1H),

6.31 (d, J = 2.3 Hz, H17”, 1H), 6.25 (dd, J = 2.3, 8.2 Hz, H19”, 1H), 5.88 (s, H5’, 1H), 4.37

(bs, H4”, 1H), 3.88-3.68 (5s, 5xOMe, 15H), 3.65 (m, H3”, 1H), 3.38 (s, OMe, 3H), 2.19

(m, H6”, 1H), 2.11(dd, J = 2.8, 13.3 Hz, H2”, 1H), 1.89 (m, H2”, 1H), 1.36 (s, H7”, 3H);

13C NMR (100 MHz, CDCl3): δ 204.1, 167.5, 165.4, 165.3, 159.1, 158.5, 158.3, 157.5,

130.5, 125.6, 114.3, 107.1, 104.1, 103.0, 101.1, 99.1, 86.1, 75.6, 55.9, 55.6, 55.5, 55.3, 54.7,

52.2, 50.6, 45.5, 36.3, 31.2, 28.6; HRMS calcd for C33H37O10 [M+H]+

593.2378. Found:

593.2360.

*coupling constant cannot be measured due to broad resonance

1H NMR (400 MHz, toluene-d8): δ 13.36 (s, OH, 1H), 7.79 (d, J = 9.2 Hz, H14”, 1H), 6.83

(d, J = 8.2 Hz, H20”, 1H), 6.45 (d, J = 8.7Hz, H13”, 1H), 6.35 (d, J = 2.3 Hz, H11”, 1H),

6.16 (d, J = 2.3 Hz, H17”, 1H), 6.06 (dd, J = 1.8, 8.2 Hz, H19”, 1H), 5.60 (s, H5’, 1H), 4.21

(m, H4”, 1H), 3.72-3.21 (5s, 5xOMe, 15H), 3.72 (m, H3”, 1H), 3.18 (s, OMe, 3H), 2.30

(bd, H6”, 1H)*, 1.71 (bd, J = 11.9 Hz, H2”, 1H), 1.60 (bd, J = 12.8 Hz, H2”, 1H), 1.26 (s,

H7”, 3H); 13

C NMR (100 MHz, toluene-d8): δ 203.9, 166.3, 166.2, 165.4, 159.3, 158.4,

127

158.3, 157.6, 130.3, 127.6, 125.5, 114.5, 107.2, 105.5, 104.0, 103.2, 100.9, 99.1, 86.2, 75.1,

54.8, 54.6, 54.2, 54.1, 51.2, 51.0, 45.8, 36.0, 31.3, 28.4.

1H NMR (400 MHz, toluene-d8, 353K): δ 13.09 (s, OH, 1H), 7.79 (d, J = 8.7 Hz, H14”,

1H), 6.84 (d, J = 8.2 Hz, H20”, 1H), 6.46 (dd, J = 2.3, 8.7 Hz, H13”, 1H), 6.33 (d, J = 2.7

Hz, H11”, 1H), 6.18 (d, J = 2.3 Hz, H17”, 1H), 6.10 (dd, J = 2.3, 8.2 Hz, H19”, 1H), 5.68 (s,

H5’, 1H), 4.26 (d, J = 11.4 Hz, H4”, 1H), 3.72-3.26 (5s, 5xOMe, 15H), 3.72 (m, H3”, 1H),

3.20 (s, OMe, 3H), 2.26 (ddd, J = 1.8, 4.1, 13.8 Hz, H6”, 1H), 1.80 (dd, J = 3.2, 13.2 Hz,

H2”, 1H), 1.68 (m, H2”, 1H), 1.26 (s, H7”, 3H); 13

C NMR (100MHz, toluene-d8, 353K): δ

204.2, 166.6, 166.4, 166.0, 165.4, 159.1, 158.9, 158.4, 130.3, 127.6, 125.8, 115.3, 107.4,

106.6, 105.3, 104.1, 101.8, 100.2, 87.5, 75.5, 55.8, 55.3, 55.0, 54.9, 54.6, 51.8, 51.3, 46.5,

36.8, 31.8, 28.7.

Compound 207

To a solution of freshly prepared MgI2 (0.23 g, 0.82 mmol) in dry Et2O (10 mL) was added

to a solution of endo-204 (0.05 g, 0.08 mmol) in THF (10 mL). The mixture was refluxed at

50oC for 6 h. Then, the precipitate was filtered, and the organic layers were collected and

dried. The residue was purified by column chromatography (hexane: EtOAc =8:2) to afford

20 mg of 206 (40% yield) and 18 mg of 207 (36% yield) as colorless crystals in very small

quantity (not enough for melting point determination).

128

1H NMR (400 MHz, CDCl3): δ 12.62 (s, OH, 1H), 12.37 (s, OH, 1H), 11.38 (bs, COOH,

1H), 7.82 (d, J = 9.1 Hz, H14”, 1H), 6.88 (d, J = 8.7 Hz, , H20”, 1H), 6.50 (dd, J = 1.8, 8.7

Hz, H13”, 1H), 6.38 (d, J = 2.3 Hz, H11”, 1H), 6.33 (d, J = 2.3 Hz, H17”, 1H), 6.26 (dd, J

= 2.3, 8.2 Hz, H19”, 1H), 6.02 (s, H5’, 1H), 4.37 (bd, H4”, 1H)*, 3.88-3.69 (3s, 3xOMe,

9H), 3.68 (m, H3”, 1H), 3.38 (s, OMe, 3H), 2.29 (bd, J = 12.8 Hz, H6”, 1H), 2.22 (dd, J =

2.7, 13.1 Hz, H2”, 1H), 1.97 (dt, J = 12.8 Hz, H2”, 1H), 1.56 (s, H7”, 3H); 13

C NMR (100

MHz, CDCl3): δ 203.3, 171.3, 165.7, 165.5, 165.0, 162.3, 159.3, 158.2, 155.7, 130.3, 122.5,

114.1, 107.5, 104.2, 102.0, 101.1, 99.2, 94.1, 92.8, 80.3, 55.6, 55.5, 55.3, 55.2, 49.8, 44.9,

35.9, 30.3, 28.6; HRMS calcd for C31H32O10Na [M+Na]+

587.1893. Found: 587.1890.

1H NMR (400 MHz, DMSO-d6): δ 12.58 (s, OH, 1H), 12.13 (s, OH, 1H), 8.19 (d, J = 8.7

Hz, H14”, 1H), 6.98 (d, J = 8.7 Hz, H20”, 1H), 6.56 (dd, J = 2.3, 9.2 Hz, H13”, 1H), 6.40

(d, J = 2.3 Hz, H11”, 1H), 6.36 (d, J = 2.3 Hz, H17”, 1H), 6.27 (dd, J = 2.3, 8.7 Hz, H19”,

1H), 5.91 (s, H5’, 1H), 4.50 (bd, J = 11.9 Hz, H4”, 1H), 3.78-3.61 (3s, 3xOMe, 9H), 3.45

(m, H3”, 1H), 3.23 (s, OMe, 3H), 2.38 (bd, J = 11.4 Hz, H2”, 1H), 2.06 (bd, H6”, 1H),

1.69 (d, J = 11.9 Hz, H2”, 1H), 1.35 (s, H7”, 3H); 13

C NMR (100 MHz, DMSO-d6): δ

205.0, 172.4, 165.8, 164.9, 163.5, 161.7, 161.7, 159.1, 157.5, 158.1, 124.6, 132.2, 114.3,

107.5, 105.0, 102.6, 101.4, 99.0, 96.0, 77.6, 56.2, 56.0, 55.4, 50.3, 46.2, 35.1, 30.9, 28.6.

129

1H NMR (400MHz, DMSO-d6, 353K): δ 12.52 (s, OH, 1H), 12.43 (s, OH, 1H), 8.12 (d, J =

8.7 Hz, H14”, 1H), 6.94 (d, J = 8.7 Hz, H20”, 1H), 6.55 (d, J = 7.6 Hz, H13”, 1H), 6.37 (d,

J = 2.5 Hz, H11”, 1H), 6.37 (d, J = 2.5 Hz, H17”, 1H), 6.27 (d, J = 8.5 Hz, H19”, 1H), 5.94

(s, H5’, 1H), 4.47 (bd, J = 11.2 Hz, H4”, 1H), 3.80-3.62 (3s, 3xOMe, 9H), 3.35 (m, H3”,

1H), 3.28 (s, OMe, 3H), 2.39 (bd, J = 11.4 Hz, H2”, 1H), 2.09 (bd, H6”, 1H)*, 1.70 (m,

H2”, 1H), 1.40 (s, H7”, 3H). 13

C NMR (100 MHz, DMSO-d6, 353K) not measured.

*coupling constant cannot be measured due to broad resonance

Experiment for compound 224, 226, 227 and 209

A mixture of 2-halophenols (1 eqv.), Pd(OAc)2 (10 mol%), K2CO3 (2 eqv.), and 2-methyl-

3-buten-2-ol (10 eqv.), in 1 mL of DMF in a pressure tube was heated to 100°C for 6 h. The

reaction was monitored by TLC under the starting material completely consumed. 1g of

silica gel then was added to the same mixture, followed by heating at 130°C for another

12h to yield desired product.

8-chloro-2,2-dimethyl-2H-chromene-6-carbaldehyde (224)

1H NMR (400MHz, CDCl3): δ 9.79 (1H, s, CHO), 7.81 (1H, s, phenyl), 7.62 (1H, s,

phenyl), 6.57 (1H, d, J = 9.8 Hz, CH), 5.95 (1H, d, J = 10.1 Hz, CH), 1.46 (6H, s, 2Me);

13C NMR (100MHz, CDCl3): δ 190.5, 153.1, 132.5, 131.3, 130.1, 125.8, 122.5, 120.7,

120.6, 79.5, 28.1; HRMS calcd for C12H12ClO2[M+H]+

223.0518. Found: 223.0512.

130

Methyl 2,2-dimethyl-2H-chromene-8-carboxylate (226)

1H NMR (400MHz, CDCl3): δ 7.47 (dd, J = 1.8, 7.9 Hz, phenyl, 1H), 7.25 (dd, J = 1.8, 7.3

Hz, phenyl, 1H), 6.44 (d, J = 9.8 Hz, CH, 1H), 5.83 (d, J = 10.1 Hz, CH, 1H), 3.77 (s, OMe,

3H), 1.38 (s, 2Me, 6H); 13

C NMR (100MHz, CDCl3): δ 165.7, 152.0, 131.7, 130.0, 122.2,

121.5, 120.1, 119.2, 76.8, 51.7, 27.5; HRMS calcd for C13H15O3[M+H]+

219.1013. Found:

219.1021.

Methyl-5,7-dimethyoxy-2,2-dimethyl-2H-chromene-8-carboxylate (227)

1H NMR (400MHz, CDCl3): δ 6.48 (d, J = 9.8 Hz, CH, 1H), 6.27 (s, phenyl, 1H), 5.86 (d, J

= 10.4 Hz, CH, 1H), 3.84-3.72 (3s, 3OMe, 9H), 1.36 (s, 2Me, 6H); 13

C NMR (100MHz,

CDCl3): δ 165.5, 157.3, 156.2, 150.9, 127.1, 115.8, 105.4, 103.2, 88.6, 76.4, 55.8, 51.7,

27.1; HRMS calcd for C15H19O5[M+H]+

279.1224. Found: 279.1223.

(E)-3-(2,4-dimethyoxyphenyl)-1-(5-methoxy-2,2-dimethyl-2H-chromen-8-yl)prop-2-

en-1-one (209)

131

1H NMR (400MHz, CDCl3): δ 7.72 (d, J = 16.0 Hz, Hβ, 1H), 7.56 (d, J = 9.16 Hz, H5’,1H),

7.52-7.47 (overlap, Hα, H6’,2H), 6.65 (d, J = 9.16 Hz, H6, 1H), 6.60-6.56 (overlap, CH, H3,

H5, 3H), 5.73 (d, J= 10.1 Hz, CH, 1H), 3.83-3.80 (3s, 3OMe, 9H), 1.40 (s, 2Me, 6H); 13

C

NMR (100MHz, CDCl3): δ 189.7, 163.2, 160.3, 158.2, 153.0, 137.1, 131.6, 130.4, 129.9,

125.2, 122.0, 116.6, 106.8, 104.6, 99.1, 77.2, 56.5, 56.3, 56.0, 27.9. HRMS calcd for

C23H25O5[M+H]+

381.1694. Found: 381.1706.

(E)-3-(2,4-bis(ethoxymethoxy)phenyl)-1-(4-(ethoxymethoxy)-3-iodo-2-

methoxyphenyl)prop-2-en-1-one (232)

To a solution of acetophenone 171 (1.51 g, 4.5 mmol) in ethanol (2.5mL/mmol), 50% KOH

(3 eq) was added. After 10 min appropriated benzaldehyde 198 (1.08 g, 4.5 mmol) was

added and the solutions were stirred at room temperature for 40 h. After cooling the

reaction mixtures with ice, the mixtures were neutralized carefully with 1N hydrochloric

acid. The crude mixture was extracted with ethyl acetate, washed with water and brine

afforded chalcones 232 (1.32g, 50% yield). 1H NMR (400 MHz, CDCl3): δ 8.04 (d, J =

16.2 Hz, Hβ, 1H), 7.64 (d, J = 8.6 Hz, H6’, 1H), 7.57 (d, J = 8.8 Hz, H6, 1H), 7.44 (d, J =

15.9 Hz, Hα, 1H), 6.92 (d, J = 8.8 Hz, H5’, 1H), 6.86 (d, J = 2.4 Hz, H3, 1H), 6.70 (dd, J =

2.4 Hz, 8.8 Hz, H5, 1H), 5.34 (s, OCH2O, 2H), 5.27 (s, OCH2O, 2H), 5.22 (s, OCH2O, 2H),

3.76 (s, OMe, 3H), 3.74(m, OCH2CH3, 6H), 1.22 (m, OCH2CH3, 9H); 13

C NMR (100 MHz,

CDCl3): δ 191.3, 160.9, 160.6, 160.4, 158.1, 139.5, 132.3, 129.7, 128.3, 124.1, 118.6, 110.4,

109.6, 103.6, 93.9, 93.6, 93.2, 86.5, 65.2, 64.9, 64.7, 63.0, 15.3; HRMS calcd for C25H32IO8

[M-H]+

585.0994. Found: 585.2892.

132

(E)-3-(2,4-bis(ethoxymethoxy)phenyl)-1-(4-(ethoxymethoxy)-2-methoxy-3-((E)-3-

methylbutan-1,3-dienyl)phenyl)prop-2-en-1-one (233)

To a mixture of 2-methyl-3-buten-2-ol (1.8 ml, 17.1 mmol), Pd(OAc)2 (38 mg, 0.17 mmol),

K2CO3 (472 mg, 3.42 mmol), and 232 (1.0 g, 1.71 mmol) of in 1 mL of DMF in a pressure

tube was heated to 100°C for 6 h. The mixture was purified by column chromatography

eluting with 50% EtOAc/hexane to yield pale yellow viscous liquid (698 mg, 75%). To this

liquid (698 mg, 1.28 mmol) in 10 mL of benzene was added acetyl chloride (0.11 mL, 1.41

mmol, 1.1 equiv), and pyridine (0.10 mL, 1.41 mmol, 1.1 equiv). The mixture was then

heated to 60°C for 4h. The white precipitate was filtered, and the organic layers were

collected, dried. The residue was purified by column chromatography (hexane: EtOAc =7:3)

to afford 495 mg of diene 233 (55% yield, over two steps) as pale yellow viscous liquid. 1H

NMR (400 MHz, CDCl3): 1

H NMR (400 MHz, CDCl3): δ 7.98 (d, J = 15.6 Hz, Hβ, 1H),

7.56 (d, J = 8.3 Hz, H6’, 1H), 7.48 (d, J = 8.8 Hz, H6, 1H), 7.44 (d, J = 16.1 Hz, Hα, 1H),

7.31 (d, J = 16.6 Hz, H7’, 1H), 6.96 (d, J = 9.8 Hz, H5’, 1H), 6.85 (d, J = 2.4 Hz, H3, 1H),

6.74 (d, J = 16.6 Hz, H8’, 1H), 6.70 (dd, J = 2.4 Hz, 8.8 Hz, H5, 1H), 5.30 (s, OCH2O, 2H),

5.26 (s, OCH2O, 2H), 5.21 (s, OCH2O, 2H), 3.72(m, OCH2CH3, 6H), 3.69 (s, OMe, 3H),

1.99 (s, H10’, 3H), 1.20 (m, OCH2CH3, 9H); 13

C NMR (100 MHz, CDCl3): δ 192.5, 160.5,

158.8, 158.7, 157.9, 143.2, 138.8, 129.9, 129.5, 128.1, 124.9, 120.3, 118.6, 116.2, 110.3,

109.5, 103.5, 93.4, 93.3, 93.1, 71.6, 64.8, 64.7, 64.6, 62.2, 29.9, 29.8, 15.2; HRMS calcd

for C30H37O8 [M-H]+

525.2497. Found: 525.2888.

133

(E)-3-(2,4-bis(ethoxymethoxy)phenyl)-1-(4-(ethoxymethoxy)-2-hydroxy-3-

iodophenyl)prop-2-en-1-one (234)

To a solution of acetophenone 217 (0.94 g, 2.8 mmol) in ethanol (2.5mL/mmol), 50% KOH

(3 eq) was added. After 10 min appropriated benzaldehyde 198 (0.71 g, 2.8 mmol) were

added and the solutions were stirred at room temperature for 40 h. After cooling the

reaction mixtures with ice, the mixtures were neutralized carefully with 1N hydrochloric

acid. The crude mixture was extracted with ethyl acetate, washed with water and brine

afforded chalcones 234 in 50% yield. 1H NMR (400 MHz, CDCl3): δ 14.47 (s, OH, 1H),

8.17 (d, J = 15.4 Hz, Hβ, 1H), 7.84 (d, J = 9.0 Hz, H6’, 1H), 7.56 (d, J = 15.4 Hz, Hα, 1H),

7.55 (d, J = 8.8 Hz, H6 , 1H), 6.86 (d, J = 2.2 Hz, H3, 1H), 6.72-6.66 (overlap, H5’, H5,

2H), 5.35 (s, OCH2O, 2H), 5.29 (s, OCH2O, 2H), 5.21 (s, OCH2O, 2H), 3.70(m, OCH2CH3,

6H), 1.21 (m, OCH2CH3, 9H); 13

C NMR (100 MHz, CDCl3): δ 192.3, 164.8, 164.4, 161.2,

158.5, 141.1, 131.3, 130.5, 118.2, 117.9, 115.6, 109.6, 103.5, 93.6, 93.5, 93.1, 78.5, 65.3,

64.9, 64.7, 15.2; HRMS calcd for C24H30IO8 [M+H]+

573.0980. Found: 573.0965.

Endo-237 and exo-238

Diene 105 (79 mg, 0.20 mmol) and dienophile 211 (88 mg, 0.20 mmol) were dissolved in

freshly distilled toluene (2 mL) in a pressure tube. The mixture was heated at 150oC for

24h to give a mixture of diastereomers 237 and 238. The mixture of diastereomers 237 and

238 was separated by column chromatography with 80% hexane/EtOAc as eluent to give a

mixture of diastereoisomers 237 (55 mg) and 238 (37 mg) (55% combined yield, 3:2 ratio).

134

Endo-237 was obtained as a pale yellow viscous liquid. 1H NMR (400 MHz, CDCl3): δ

13.57 (s, OH, 1H), 7.83 (d, J = 15.9 Hz, Hβ, 1H), 7.46 (d, J = 8.6 Hz, H6, 1H), 7.31-7.22

(overlap, Hα, H20”,H5’, 3H), 6.87 (d, J = 8.3 Hz, H14”, 1H), 6.44 (dd, J = 2.1, 8.6 Hz,

H17”, 1H), 6.43-6.40 (overlap, H4’, 1H), 6.37 (d, J = 1.8 Hz, H3, 1H), 6.32 (d, J = 8.6 Hz,

H19”, 1H), 6.24 (bd, *, H5, 1H), 5.41 (s, H2”, 1H), 4.38 (s, H4”, 1H), 3.98 (bd, J = 6.7 Hz,

H3”, 1H), 3.75-3.58 (5s, 5OMe, 15H), 3.52 (bs, H5” , 1H), 3.40 (s, OMe, 3H), 3.18

(overlap, H6”, 1H), 3.17 (s, OMe, 3H), 1.76 (bs, H6”, 1H), 1.73 (s, CH3, 3H); 13

C NMR

(100 MHz, CDCl3): δ 207.6, 192.4, 163.6, 163.5, 163.1, 162.0, 160.3, 159.9, 158.7, 157.5,

157.5, 138.9, 137.7, 131.3, 130.4, 130.3, 130.1, 126.9, 125.9, 124.3, 124.1, 120.1, 117.2,

116.3, 106.5, 105.5, 103.4, 101.4, 98.5, 98.4, 76.4, 62.5, 56.7, 55.6, 55.5, 55.3, 55.2, 54.2,

48.5, 38.2, 31.2, 30.3, 24.2; HRMS calcd for C42H44IO10 [M+H]+

835.1971. Found:

835.1969.

135

Exo-238 was obtained as a pale yellow viscous liquid. 1H NMR (400 MHz, CDCl3): δ 7.89

(d, J = 16.5 Hz, Hβ, 1H), 7.68 (d, J = 8.6 Hz, H20”, 1H), 7.54 (d, J = 8.2 Hz, H6, 1H), 7.41

(d, J = 16.0 Hz, Hα, 1H), 7.26 (overlap, H3, 1H), 6.99 (d, J = 8.2 Hz, H5’. 1H), 6.63-6.41

(overlap, H4’, H14”, H17”, 3H), 6.32 (bd, *, H13”, 1H), 6.23 (bd, *, H19”, 1H), 6.07 (bd, *,

H5, 1H), 5.24 (s, H2”, 1H), 4.81 (m, H4”, 1H), 4.45 (m, H3”, 1H), 3.95-3.65 (5s, 5OMe,

15H), 3.74 (overlap, H5” , 1H), 3.57 (s, OMe, 3H), 3.47 (s, OMe, 3H), 2.24 (m, H6”, 1H),

1.73 (s, CH3, 3H); 13

C NMR (100 MHz, CDCl3): δ 208.7, 192.8, 163.7, 163.2, 162.8, 161.9,

160.3, 160.1, 159.1, 158.2, 139.5, 138.6, 132.6, 132.0, 130.5, 130.3, 127.1, 126.7, 125.0,

124.5, 123.9, 117.1, 116.2, 107.5, 105.7, 104.4, 101.6, 98.9, 98.5, 75.7, 63.8, 63.2, 56.6,

56.4, 56.1, 55.7, 55.6, 55.3, 47.7, 38.8, 29.7, 23.4; HRMS calcd for C42H44IO10 [M+H]+

835.1971. Found: 835.1970.

*coupling constant cannot be measured due to broad resonance

Endo-230

Mixture of endo-241 (39 mg, 0.50 mmol) and PdCl2 (18 mg, 0.01 mmol) in dry

ethanol (20 mL) were stirred in room temperature for 120 h, which smoothly afforded the

desired intermediates 230 in 19 mg, 50% yield.

136

Endo-230 was obtained as a pale yellow viscous liquid. 1H NMR (400 MHz, CDCl3): δ

7.91 (d, J = 16.0 Hz, Hβ, 1H), 7.49-7.42 (overlap, Hα, H6, H20”, 3H), 6.99 (d, J = 7.8 Hz,

H14”, 1H), 6.65 (d, J = 10.0 Hz, CH, 1H), 6.49 (overlap, H4’, 1H), 6.45 (dd, J = 2.3, 8.2

Hz, H17”, 1H), 6.40 (d, J = 2.3 Hz, H3, 1H), 6.30 (bs, H19”, 1H), 6.24 (bd, J = 8.2 Hz, H5,

1H), 5.61 (d, J = 10.0, CH), 5.39 (bs, H2”, 1H), 4.87 (m, H5”, 1H), 4.75 (s, H4”, 1H), 4.26

(bs, H3”, 1H), 3.80 (s, 2OMe, 6H), 3.68-3.63 (4s, 4OMe, 12H), 3.44 (s, OMe, 3H), 2.26

(dd, J = 5.5, 16.0 Hz, H6”, 1H), 2.08 (m, H6”, 1H), 1.72 (s, H7”, 3H), 1.61 (s, CH3, 3H),

1.55 (s, CH3, 3H); 13

C NMR (100 MHz, CDCl3): δ200.6, 193.1, 162.8, 160.4, 158.6, 158.5,

157.7, 153.2, 138.6, 133.6, 131.7, 131.0, 130.5, 128.4, 125.1, 123.8, 122.8, 122.2, 117.7,

117.1, 109.8, 106.6, 105.5, 104.4, 103.0, 99.2, 98.4, 76.9, 63.1, 63.0, 55.8, 55.7, 55.4, 53.6,

46.3, 38.8, 33.8, 29.9, 29.5, 28.4, 28.3, 23.9; HRMS calcd for C47H51O10 [M+H]+

775.3474.

Found: 775.3499.

Exo-231

Mixture of exo-242 (31 mg, 0.40 mmol) and PdCl2 (14 mg, 0.008 mmol) in dry

ethanol (20 mL) were stirred in room temperature for 216 h, which smoothly afforded the

desired intermediates 231 in 14 mg, 45% yield.

137

Exo-231 was obtained as a pale yellow viscous liquid 1

H NMR (400 MHz, CDCl3): δ 7.96

(d, J = 15.6 Hz, Hβ, 1H), 7.73 (d, J = 9.2, H20”, 1H), 7.55-7.46 (overlap, Hα, H6, 2H), 7.40

(d, J = 8.2 Hz, H3, 1H), 7.13 (d, J = 8.7 Hz,H5’, 1H), 6.53-6.30 (overlap, H4’,H13”, H14”,

H17”, H19”, CH, 6H), 6.02 (bd, J = 8.7 Hz, H5, 1H), 5.47 (d, J = 9.6, CH), 5.16 (bs, H2”,

1H), 4.87 (t, J = 10.5 Hz, H4”, 1H), 4.25 (bs, H3”, 1H), 3.95 (bs, H5”, 1H), 3.84 (3s, 3OMe,

9H), 3.70 (2s, 2OMe, 6H), 3.56 (s, OMe, 3H), 3.53 (s, OMe, 3H), 2.27 (bs, H6”, 2H), 1.70

(s, H7”, 3H), 1.54 (s, CH3, 3H), 1.47 (s, CH3, 3H); 13

C NMR (100 MHz, CDCl3): δ 205.4,

191.4, 162.9, 162.6, 160.3, 158.9, 158.7, 158.3, 157.1, 152.4, 137.8, 132.9, 132.8, 131.8,

131.3, 130.6, 130.4, 126.5, 125.6, 124.6, 124.1, 124.0, 117.5, 117.0, 109.9, 107.4, 105.4,

104.3, 102.4, 98.6, 76.3, 63.6, 63.2, 56.9, 55.9, 55.8, 55.6, 55.3, 51.4, 39.2, 38.7, 32.0, 29.8,

28.1, 26.7, 26.1, 23.1; HRMS calcd for C47H51O10 [M+H]+

775.3474. Found: 775.3495.

138

REFERENCES

Abegaz, B. M., Ngadjui, B. T., Dongo, E., Ngameni, B., Nindi, M. N., & Bezabih, M.

(2002). Chalcones and other constituents of Dorstenia prorepens and Dorstenia

zenkeri. Phytochemistry, 59(8), 877-883.

Adler, M. J., & Baldwin, S. W. (2009). Direct, regioselective synthesis of 2,2-dimethyl-2H-

chromenes. Total syntheses of octandrenolone and precocenes I and II. Tetrahedron

Letters, 50(36), 5075-5079.

Ahmed, S., Wagner, H., & Razaq, S. (1978). Synthesis of polyhydroxy-flavone

methylethers with potential cytotoxic activity—III: Synthesis of 5,7-dihydroxy

6,2′.4′,5′-tetramethoxyflavone (tabularin) from chukrasia tabularis a. juss and

5,6,7,2′,4′,5′-hexamethoxyflavone (tabularin dimethylether). Tetrahedron, 34(10),

1593-1594.

Akanksha, & Maiti, D. (2012). Microwave-assisted palladium mediated decarbonylation

reaction: synthesis of eulatachromene. Green Chemistry, 14(8), 2314-2320.

Arkoudis, E., Lykakis, I. N., Gryparis, C., & Stratakis, M. (2009). Biomimetic Synthesis of

Dimeric Metabolite Acremine G via a Highly Regioselective and Stereoselective

Diels−Alder Reaction. Organic Letters, 11(14), 2988-2991.

Ballerini, E., Minuti, L., & Piermatti, O. (2010). High-pressure Diels-Alder cycloadditions

between benzylideneacetones and 1,3-butadienes: application to the synthesis of

(R,R)-(-)- and (S,S)-(+)-Delta8-tetrahydrocannabinol. J Org Chem, 75(12), 4251-

4260. doi: 10.1021/jo100785w

Ballerini, E., Minuti, L., Piermatti, O., & Pizzo, F. (2009). High Pressure Diels−Alder

Approach to Hydroxy-Substituted 6a-Cyano-tetrahydro-6H-benzo[c]chromen-6-

ones: A Route to Δ6-Cis-Cannabidiol. The Journal of Organic Chemistry, 74(11),

4311-4317.

Bao, J., Wulff, W. D., & Rheingold, A. L. (1993). Vaulted biaryls as chiral ligands for

asymmetric catalytic Diels-Alder reactions. Journal of the American Chemical

Society, 115(9), 3814-3815.

Barbour, L. J. (2001). X-Seed -A Software Tool for Supramolecular Crystallography.

Journal of Supramolecular Chemistry, 1, 189-191.

139

Basnet, P. K. S., Terashima, S, Shimizu, M., Namba, T. (1993). Two new 2-arylbenzofuran

derivatives from hypoglycemic activity-bearing fractions of Morus insignis.

Chemical Phamaceutical Bulletin, 41, 1238-1243.

Becke, A. D. (1993a).A new mixing of Hartree–Fock and local density‐functional theories.

Journal Chemical Physics, 98, 1372-1377.

Becke, A. D. (1993b). Density‐functional thermochemistry. III. The role of exact exchange.

Journal Chemical Physics, 98, 5648-5652.

Bergmann, R., & Gericke, R. (1990). Synthesis and antihypertensive activity of 4-(1,2-

dihydro-2-oxo-1-pyridyl)-2H-1-benzopyrans and related compounds, new

potassium channel activators. J Med Chem, 33(2), 492-504.

Bigi, F., Carloni, S., Maggi, R., Muchetti, C., & Sartori, G. (1997). Zeolite-Induced

Heterodomino Reaction. Regioselective Synthesis of 2H-1-Benzopyrans from

Phenols and α-Alkynols. The Journal of Organic Chemistry, 62(20), 7024-7027.

Boonsri, S., Gunawan, C., Krenske, E. H., & Rizzacasa, M. A. (2012). Synthetic studies

towards the mulberry Diels-Alder adducts: H-bond accelerated cycloadditions of

chalcones. Org Biomol Chem, 10(30), 6010-6021.

Bröhmer, M. C., Volz, N., & Bräse, S. (2009). Thieme Chemistry Journal Awardees -

Where Are They Now? Microwave-Assisted Rhodium-Catalyzed Decarbonylation

of Functionalized 3-Formyl-2H-chromenes: A Sequence for Functionalized

Chromenes like Deoxycordiachromene. Synlett, 2009(09), 1383-1386.

Chauder, B. A., Lopes, C. C., Lopes, R. S. C., da Silva, Alcides J. M., & Snieckus, V.

(1998). Phenylboronic Acid-Mediated Synthesis of 2H-Chromenes. Synthesis,

1998(03), 279-282.

Chee, C. F., Lee, Y. K., Buckle, M. J. C., & Abdul Rahman, N. (2011). Synthesis of (±)-

kuwanon V and (±)-dorsterone methyl ethers via Diels–Alder reaction. Tetrahedron

Letters, 52(15), 1797-1799.

Chen, H-D, Ding, Y-Q., Yang, S-P, Li, X-C., Wang, X-J., Zhang, H-Y., Ferreira, D., Yue,

J-M.. (2012). Morusalbanol A, a neuro-protective Diels–Alder adduct with an

unprecedented architecture from Morus alba. Tetrahedron, 68(30), 6054-6058.

140

Coleman, R. S., Lu, X-L., & Modolo, I. (2007). Total Synthesis of 2‘-O-

Methylmyxalamide D and (6E)-2‘-O-Methylmyxalamide D. Journal of the

American Chemical Society, 129(13), 3826-3827.

Cong, H., & Porco, J. A. Jr. (2012). Total synthesis of (+/-)-sorocenol B employing

nanoparticle catalysis. Org Lett, 14(10), 2516-2519.

Cong, H., Becker, C. F., Elliott, S. J., Grinstaff, M. W., & Porco, J. A. Jr. (2010). Silver

Nanoparticle-Catalyzed Diels−Alder Cycloadditions of 2′-Hydroxychalcones.

Journal of the American Chemical Society, 132(21), 7514-7518.

Cong, H., Ledbetter, D., Rowe, G. T., Caradonna, J. P., & Porco, J. A. Jr. (2008). Electron

Transfer-Initiated Diels−Alder Cycloadditions of 2′-Hydroxychalcones. Journal of

the American Chemical Society, 130(29), 9214-9215.

Dai, S. J., Ma, Z. B., Wu, Y., Chen, R. Y., & Yu, D. Q. (2004). Guangsangons F-J, anti-

oxidant and anti-inflammatory Diels-Alder type adducts, from Morus macroura Miq.

Phytochemistry, 65(23), 3135-3141.

Dai, S. J., Mi, Z. M., Ma, Z. B., Li, S., Chen, R. Y., & Yu, D. Q. (2004). Bioactive diels-

alder type adducts from the stem bark of Morus macroura. Planta Med, 70(8), 758-

763.

David, C. W., Iqbal, M, & Livingstone, R. (1998). Preparation of Some 6-Substituted 2,2-

Dimethyl- and 2,2- and 2,4-Diphenyl-naphtho[1,2-b]pyrans. Journal of Chemical

Research, Synopses(1), 2-3.

Delley B. Delley (1990). An All-Electron Numerical Method for Solving the Local Density

Functional for Polyatomic Molecules. Journal of Chemical Physics, 92, 508-517.

Delley B. (2000). From molecules to solids with the DMol3 approach. Journal of Chemical

Physics, 113(18), 7756-7764.

Dong, Y-W, Wang, G-W., & Wang, L. (2008). Solvent-free synthesis of naphthopyrans

under ball-milling conditions. Tetrahedron, 64(44), 10148-10154.

Fine, S. A., & Pulaski, P. D. (1973). Reexamination of the Claisen-Schmidt condensation

of phenylacetone with aromatic aldehydes. The Journal of Organic Chemistry,

38(9), 1747-1749.

Frisch, M. J.; Trucks, G. W.; Schlegel, H. B.; Scuseria, G. E.; Robb, M. A.; Cheeseman, J.

R.; Scalmani, G.; Barone, V.; Mennucci, B.; Petersson, G. A.; Nakatsuji, H.; Caricato,

141

M.; Li, X.; Hratchian, H. P.; Izmaylov, A. F.; Bloino, J.; Zheng, G.; Sonnenberg, J.

L.; Hada, M.; Ehara, M.; Toyota, K.; Fukuda, R.; Hasegawa, J.; Ishida, M.; Nakajima,

T.; Honda, Y.; Kitao, O.; Nakai, H.; Vreven, T.; Montgomery, J. A., Jr.; Peralta, J. E.;

Ogliaro, F.; Bearpark, M.; Heyd, J. J.; Brothers, E.; Kudin, K. N.; Staroverov, V. N.;

Kobayashi, R.; Normand, J.; Raghavachari, K.; Rendell, A.; Burant, J. C.; Iyengar, S.

S.; Tomasi, J.; Cossi, M.; Rega, N.; Millam, J. M.; Klene, M.; Knox, J. E.; Cross, J.

B.; Bakken, V.; Adamo, C.; Jaramillo, J.; Gomperts, R.; Stratmann, R. E.; Yazyev,

O.; Austin, A. J.; Cammi, R.; Pomelli, C.; Ochterski, J. W.; Martin, R. L.; Morokuma,

K.; Zakrzewski, V. G.; Voth, G. A.; Salvador, P.; Dannenberg, J. J.; Dapprich, S.;

Daniels, A. D.; Farkas, Ö.; Foresman, J. B.; Ortiz, J. V.; Cioslowski, J.; Fox, D. J.

(2009) Gaussian 09, Inc. Wallingford CT.

Fukai, T., Hano, Y., Hirakura, K., Nomura, T, Uzawa, J., & Fukushima, K. (1985).

Structures of Two Natural Hypotensive Diels-Alder Type Adducts, Mulberrofurans

F and G, from the Cultivated Mulberry Tree (Morus lhou KOIDZ.). Chemical &

Pharmaceutical Bulletin, 33(8), 3195-3204.

Fürstner, A., Heilmann, E. K., & Davies, P. W. (2007). Total Synthesis of the Antibiotic

Erypoegin H and Cognates by a PtCl2-Catalyzed Cycloisomerization Reaction.

Angewandte Chemie International Edition, 46(25), 4760-4763.

Gabbutt, C. D, Heron, B. Mark, I., Alicia C., Thomas, D. A., Partington, S. M., Hursthouse,

M. B., & Gelbrich, T. (2003). Observations on the Synthesis of Photochromic

Naphthopyrans. European Journal of Organic Chemistry, 2003(7), 1220-1230.

Gan, R. Y., Kuang, L., Xu, X. R., Zhang, Y., Xia, E. Q., Song, F. L., & Li, H. B. (2010).

Screening of natural antioxidants from traditional Chinese medicinal plants

associated with treatment of rheumatic disease. Molecules, 15(9), 5988-5997.

Gao, L., Han, J & Lei, X. (2016). Enantioselective Total Syntheses of Kuwanon X,

Kuwanon Y, and Kuwanol A. Organic Letter, 18 (3), 360–363.

Goujon, J. Y., Zammattio, F., & Kirschleger, B. (2000). Synthesis of various 2H-

benzopyran compounds and their kinetic resolution by asymmetric hydrolysis of

their racemic acetates mediated by lipases. Tetrahedron: Asymmetry, 11(11), 2409-

2420.

Gunawan, C., & Rizzacasa, M. A. (2010). Mulberry Diels−Alder Adducts: Synthesis of

Chalcomoracin and Mulberrofuran C Methyl Ethers. Organic Letters, 12(7), 1388-

1391.

142

Halgren, T. A. and Lipscomb, W. N.(1977). The Synchronous Transit Method for

Determining Reaction Pathways and Locating Transition States. Chemical Physics

Letters, 49, 225-232.

Han, J., Jones, A. X., & Lei, X. (2015). Recent Advances in the Total Synthesis of

Prenylflavonoid and Related Diels–Alder Natural Products. Synthesis, 47(11), 1519-

1533.

Han, J., Li, X., Guan, Y., Zhao, W-J., Wulff, W. D., & Lei, X. (2014). Enantioselective

Biomimetic Total Syntheses of Kuwanons I and J and Brosimones A and B.

Angewandte Chemie, 126(35), 9411-9415.

Hano, Y., Fukai, T., Nomura, T., Uzawa, J., & Fukushima, K. (1984). Structure of

mulberrofuran I, a novel 2-arylbenzofuran derivative from the cultivated mulberry

tree (morus bombycis koidz). Chemical & Pharmaceutical Bulletin , 32(3), 1260-

1263.

Hano, Y., Tsubura, H., Nomura, T. (1986). Constituents of the cultivated mulberry tree. 38.

Structures of kuwanons Y and Z, two new stilbene derivatives from the cultivated

mulberry tree (Morus alba L.) Heterocycles, 24, 2603-2610.

Hano, Y., Suzuki, S., Nomura, T., & Iitaka, Y. (1988). Absolute configuration of natural

Diels-Alder type adducts from the morus root bark. Heterocycles, 27(10), 2315-

2325.

Hano, Y., Yamanaka, J., Nomura, T., & Momose, Y. (1995). Sorocenols A and B, two new

isoprenylated phenols from the root bark of Sorocea bonplandii Baillon.

heterocycles 41(5), 8.

Harrington, P. J., Hegedus, L. S., & McDaniel, K. F. (1987). Palladium-catalyzed reactions

in the synthesis of 3- and 4-substituted indoles. 2. Total synthesis of the N-acetyl

methyl ester of (±)-clavicipitic acids. Journal of the American Chemical Society,

109(14), 4335-4338.

Helesbeux, J-J., Duval, O., Guilet, D., Séraphin, D., Rondeau, D., & Richomme, P. (2003).

Regioselectivity in the ene reaction of singlet oxygen with ortho-prenylphenol

derivatives. Tetrahedron, 59(27), 5091-5104.

Heller, D. P., Goldberg, D. R., Wu, H., & Wulff, W. D. (2006). An examination of

VANOL, VAPOL, and VAPOL derivatives as ligands for asymmetric catalytic

Diels–Alder reactions. Canadian Journal of Chemistry, 84(10),1487-1503.

Hirakura, K., Hano, Y., Fukai, T., Nomura, T., Uzawa, J., Fukushima, K. (1985).

Constituents of the cultivated mulberry tree 21. Structures of three new natural

143

Diels-Alder type adducts, kuwanons P and X, and mulberrofuran J from cultivated

mulberry tree (Morus lhou Koidz.) Chemical Phamaceutical Bulletin, 33, 1088-

1096

Hirakura, K., Hano, Y., Fukai, T., Nomura, T., Uzawa, J., Fukushima, K. (1985). CHiroya,

K., Suzuki, N., Yasuhara, A., Egawa, Y., Kasano, A., & Sakamoto, T. (2000). Total

syntheses of three natural products, vignafuran, 2-(4-hydroxy-2-methoxyphenyl)-6-

methoxybenzofuran-3-carboxylic acid methyl ester, and coumestrol from a common

starting material. Journal of the Chemical Society, Perkin Transactions 1(24), 4339-

4346.

Iikubo, K., Ishikawa, Y., Ando, N., Umezawa, K., & Nishiyama, S. (2002). The first direct

synthesis of α-mangostin, a potent inhibitor of the acidic sphingomyelinase.

Tetrahedron Letters, 43(2), 291-293.

Iyer, M., & Trivedi, G. K. (1990). Synthesis of 2, 2-Dimethyl-2H-chromenes via a

Palladium (II) Catalysed Reaction. Synthetic Communications, 20(9), 1347-1351.

Kang, J., Chen, R. Y., & Yu, D. Q. (2006). Five new diels-alder type adducts from the stem

and root bark of Morus mongolica. Planta Med, 72(1), 52-59.

Kogan, S. B., & Herskowitz, M. (2002). Dehydrogenation of Neohexane to Neohexene on

Platinum Polymetallic Catalysts. Industrial & Engineering Chemistry Research,

41(24), 5949-5951.

Kureshy, R. I., Ahmad, I., Pathak, K., Khan, N. H., Abdi, Sayed H. R., & Jasra, R. V.

(2009). Sulfonic acid functionalized mesoporous SBA-15 as an efficient and

recyclable catalyst for the synthesis of chromenes from chromanols. Catalysis

Communications, 10(5), 572-575.

Lamcharfi, E., Menguy, L., & Zamarlik, H. (1993). One-Pot Synthesis of Analogues of

Precocenes and Isoprenyl-Chromenes. Synthetic Communications, 23(21), 3019-

3028.

Larock, R. C., Wei, L., & Hightower, T. R. (1998). Synthesis of 2H-1-Benzopyrans by Pd-

Catalyzed Cyclization of o-Allylic Phenols. Synlett, 1998(05), 522-524.

Lee, C. T., Yang, W. T., and Parr R. G. (1988). Development of the Colle-Salvetti

correlation-energy formula into a functional of the electron density. Physical

Review B, 37, 785-789.

144

Lee, D-H., Kwon, K-H., & Yi, C. S. (2012). Dehydrative C–H Alkylation and Alkenylation

of Phenols with Alcohols: Expedient Synthesis for Substituted Phenols and

Benzofurans. Journal of the American Chemical Society, 134(17), 7325-7328.

Lee, Y-R., Choi, J-H., & Yoon, S-H. (2005). Efficient and general method for the synthesis

of benzopyrans by ethylenediamine diacetate-catalyzed reactions of resorcinols with

α,β-unsaturated aldehydes. One step synthesis of biologically active (±)-confluentin

and (±)-daurichromenic acid. Tetrahedron Letters, 46(44), 7539-7543.

Liu, Y., Yao, B., Deng, C-L., Tang, R-Y., Zhang, X-G., & Li, J.-H. (2011). Palladium-

Catalyzed Selective Heck-Type Diarylation of Allylic Esters with Aryl Halides

Involving a β-OAc Elimination Process. Organic Letters, 13(5), 1126-1129.

Lykakis, I. N., Efe, C., Gryparis, C., & Stratakis, M. (2011). Ph3PAuNTf2 as a Superior

Catalyst for the Selective Synthesis of 2H-Chromenes: Application to the Concise

Synthesis of Benzopyran Natural Products. European Journal of Organic Chemistry,

2011(12), 2334-2338.

Madabhushi, S., Jillella, R., Godala, Kondal R., Mallu, K. K. R., Beeram, C. R., &

Chinthala, N. (2012). An efficient and simple method for synthesis of 2,2-

disubstituted-2H-chromenes by condensation of a phenol with a 1,1-disubstituted

propargyl alcohol using BF3·Et2O as the catalyst. Tetrahedron Letters, 53(39),

5275-5279.

Manners, G. D., & Jurd, L. (1976). Isomerization of phenolic chromens. Journal of the

Chemical Society, Chemical Communications(12), 448-448.

Mannhold, R., Cruciani, G., Weber, H., Lemoine, H., Derix, A., Weichel, C., & Clementi,

M. (1999). 6-Substituted benzopyrans as potassium channel activators: synthesis,

vasodilator properties, and multivariate analysis. Journal of medicinal chemistry,

42(6), 981-991.

Macrae, C. F.; Bruno, I. J.; Chisholm, J. A.; Edgington, P. R.; McCabe, P.; Pidcock, E.;

Rodriguez-Monge, L.; Taylor, R.; van de Streek, J.; Wood, P. A. (2008). Mercury

CSD 2.0 – new features for the visualization and investigation of crystal structures.

Journal of Applied Crystallography, 41, 466-470

Messana, I., Ferrari, F., & De A., Maria D. C. M. (1988). Constituents of brosimopsis

oblongifolia.31: Structure of a new diels-alder type adduct, brosimone a.

Tetrahedron, 44(21), 6693-6698.

145

Miyaura, N., & Suzuki, A. (1995). Palladium-Catalyzed Cross-Coupling Reactions of

Organoboron Compounds. Chemical Reviews, 95(7), 2457-2483.

Nishida, J., & Kawabata, J. (2006). DPPH radical scavenging reaction of hydroxy- and

methoxychalcones. Biosci Biotechnol Biochem, 70(1), 193-202.

Nomura, t., Fukai, T., Matsumoto, J., Ohmori, T. (1982). Constituents of the cultivated

mulberry tree. VIII. Components of root barks of Morus bombycis. Planta Medica,

46, 28-32.

Nomura, T. (1988a). Phenolic Compounds of the Mulberry Tree and Related Plants. In W.

Herz, H. Grisebach, G. W. Kirby & C. Tamm (Eds.), Fortschritte der Chemie

organischer Naturstoffe / Progress in the Chemistry of Organic Natural Products

(Vol. 53, pp. 87-201): Springer Vienna.

Nomura, T. (1988b). Phenolic compounds of the mulberry tree and related plants. Fortschr

Chem Org Naturst, 53, 87-201.

Nomura, T., & Fukai, T. (1981). Constituents of the cultivated mulberry tree. Planta Med,

42(1), 79-88.

Nomura, T., & Hano, Y. (1994). Isoprenoid-substituted phenolic compounds of moraceous

plants. Natural Product Reports, 11(0), 205-218.

Nomura, T., & Fukai, T. (1980). Kuwanon G, a New Flavone Derivative from the Root

Barks of the Cultivated Mulberry Tree (Morus alba L.). Chemical &

Pharmaceutical Bulletin, 28(8), 2548-2552.

Nomura, T., Fukai, T., Narita, T., Terada, S., Uzawa, J., Iitaka, Y., Takasugi, M., Ishikawa,

S., Nagao, S., Masamune, T. (1981). Confirmation of the structures of kuwanons G

and H (albanins F and G) by partial synthesis. Tetrahedron Letters, 22(23), 2195-

2198.

North, J. T., Kronenthal, D. R., Pullockaran, A. J., Real, S. D., Chen, H. Y. (1995).

Synthesis of 6-Cyano-2,2-dimethyl-2H-1-benzopyran and Other Substituted 2,2-

Dimethyl-2H-1-benzopyrans. The Journal of Organic Chemistry, 60(11), 3397-

3400.

Pandey, G., & Krishna, A. (1988). Synthetic application of photoinduced single electron

transfer reactions: a convenient synthetic approach for the 2,2-dimethyl-2H-

chromene system. The Journal of Organic Chemistry, 53(10), 2364-2365.

146

Perdew J. P. and Wang, Y. (1992). Accurate and simple analytic representation of the

electron-gas correlation energy. Physical Review B, 45(23), 13244-13249.

Prado, S., Janin, Y. L., & Bost, P-E. (2006). A simple two steps ytterbium triflate-catalysed

preparation of 2,2-dimethyl-2h-chromenes from salicylaldehydes and 2-

methylpropene. Journal of Heterocyclic Chemistry, 43(6), 1605-1608. .

Qi, C., Cong, H., Cahill, K. J., Muller, P., Johnson, R. P., & Porco, J. A., Jr. (2013).

Biomimetic dehydrogenative Diels-Alder cycloadditions: total syntheses of

brosimones A and B. Angew Chem Int Ed Engl, 52(32), 8345-8348. .

Rama Rao, A. V., Deshpande, V. H., Shastri, R. K., Tavale, S. S., & Dhaneshwar, N. N.

(1983). Structures of albanols a and b, two novel phenols from Morus alba bark.

Tetrahedron Letters, 24(29), 3013-3016.

Rathwell, D. C., Yang, S. H., Tsang, K. Y., & Brimble, M. A. (2009). An efficient formal

synthesis of the human telomerase inhibitor (+/-)-gamma-rubromycin. Angew Chem

Int Ed Engl, 48(43), 7996-8000.

Rollinger, J. M., Bodensieck, A., Seger, C., Ellmerer, E. P., Bauer, R., Langer, T., &

Stuppner, H. (2005). Discovering COX-inhibiting constituents of Morus root bark:

activity-guided versus computer-aided methods. Planta Med, 71(5), 399-405.

Rollinger, J. M., Spitaler, R., Menz, M., Marschall, K., Zelger, R,. Ellmerer, E. P.,

Schneider, P., Stuppner, H. (2006). Venturia inaequalis-inhibiting Diels-Alder

adducts from Morus root bark. J Agric Food Chem, 54(22), 8432-8436.

Roush, W. R., & Sciotti, R. J. (1998). Enantioselective Total Synthesis of (−)-

Chlorothricolide via the Tandem Inter- and Intramolecular Diels−Alder Reaction of

a Hexaenoate Intermediate. Journal of the American Chemical Society, 120(30),

7411-7419.

Sa e Sant'Anna, S., Evangelista, E. A., Alves, R. B., & Raslan, D. S. (2005). Synthesis of

New (±) Pterocarpans by Heck Oxyarylation. Chemistry of Natural Compounds,

41(4), 385-387.

Sartori, G., Casiraghi, G., Bolzoni, L., & Casnati, G. (1979). General synthesis of 2H-

benzo[b]pyrans (chrom-3-enes) from metal phenoxides and .alpha.,.beta.-

unsaturated carbonyl compounds. The Journal of Organic Chemistry, 44(5), 803-

805.

147

Sebastián, D., Bordejé, E. G., Calvillo, L., Lázaro, M. J., & Moliner, R. (2008). Hydrogen

storage by decalin dehydrogenation/naphthalene hydrogenation pair over platinum

catalysts supported on activated carbon. International Journal of Hydrogen Energy,

33(4), 1329-1334.

Sheldrick, G. M. (2008). A short history of SHELX. Acta Crystallographica Section A.

Foundation and Advances, 64, 112-122.

Sheldrick, G. M. (2012). SHELXT, Universität Göttingen, Germany.

Shi, Y-Q., Fukai, T., Sakagami, H., Chang, W-J., Yang, P-Q., Wang, F-P., & Nomura, T.

(2001). Cytotoxic Flavonoids with Isoprenoid Groups from Morus mongolica1.

Journal of Natural Products, 64(2), 181-188.

Sohn, H. Y., Son, K. H., Kwon, C. S., Kwon, G. S., & Kang, S. S. (2004). Antimicrobial

and cytotoxic activity of 18 prenylated flavonoids isolated from medicinal plants:

Morus alba L., Morus mongolica Schneider, Broussnetia papyrifera (L.) Vent,

Sophora flavescens Ait and Echinosophora koreensis Nakai. Phytomedicine, 11(7-8),

666-672.

Sonogashira, K., Tohda, Y., & Hagihara, N. (1975). A convenient synthesis of acetylenes:

catalytic substitutions of acetylenic hydrogen with bromoalkenes, iodoarenes and

bromopyridines. Tetrahedron Letters, 16(50), 4467-4470.

Sugamoto, K., Kurogi, C., Matsushita, Y., & Matsui, T. (2008). Synthesis of 4-

hydroxyderricin and related derivatives. Tetrahedron Letters, 49(47), 6639-6641.

Szczepanik, M., Obara, R., Szumny, A., Gabrys, B., Halarewicz-Pacan, A., Nawrot, J., &

Wawrzenczyk, C. (2005). Synthesis and insect antifeedant activity of precocene

derivatives with lactone moiety. J Agric Food Chem, 53(15), 5905-5910.

Takasugi, M., Nagao, S., Masamune, T., Shirata, A., & Takahashi, K. (1980).

Chalcomoracin, a natural Diels-Alder adduct from diseased mulberry. Chem Lett, 9,

1573-1576.

Talamás, F. X., Smith, D. B., Cervantes, A., Franco, F., Cutler, S. T., Loughhead, D. G.,

Morgans Jr, D. J., Weikert, R. J. (1997). The Florisil® catalyzed [1,3]-sigmatropic

shift of allyl phenyl ethers — An entryway into novel mycophenolic acid analogues.

Tetrahedron Letters, 38(27), 4725-4728.

Tan, Y. X., Yan, R. Y., Wang, H. Q., Chen, R. Y., & Yu, D. Q. (2009). Wittiorumins A - F,

antioxidant diels-alder-type adducts from Morus wittiorum. Planta Med, 75(3), 249-

255.

148

Trend, R. M., Ramtohul, Y. K., Ferreira, E. M., & Stoltz, B. M. (2003). Palladium-

Catalyzed Oxidative Wacker Cyclizations in Nonpolar Organic Solvents with

Molecular Oxygen: A Stepping Stone to Asymmetric Aerobic Cyclizations.

Angewandte Chemie International Edition, 42(25), 2892-2895.

Trend, R. M., Ramtohul, Y. K., & Stoltz, B. M. (2005). Oxidative Cyclizations in a

Nonpolar Solvent Using Molecular Oxygen and Studies on the Stereochemistry of

Oxypalladation. Journal of the American Chemical Society, 127(50), 17778-17788.

Ueda, S., Nomura, T., Fukai, T., & Matsumoto, J.. (1982). Kuwanon J, a new diesl-alder

adduct and chalcomoracin from callus culture of morus alba L. Chemical &

Pharmaceutical Bulletin, 30(8), 3042-3045. doi: 10.1248/cpb.30.3042

Venkataraman, K. (1972). Wood phenolics in the chemotaxonomy of the moraceae.

Phytochemistry, 11(5), 1571-1586.

Wang, G., Wu, W., Peng, F., Cao, D., Yang, Z., Ma, L., Qiu, N., Ye, H., Han, X., Chen,

J., Qiu, J., Sang, Y., Liang, X., Ran, Y., Peng, A., Wei, Y., Chen, L. (2012). Design,

synthesis, and structure–activity relationship studies of novel millepachine

derivatives as potent antiproliferative agents. European Journal of Medicinal

Chemistry, 54(0), 793-803.

Wang, H., Yan, Z., Lei, Y., Sheng, K., Yao, Q., Lu, K., & Yu, P. (2014). Concise synthesis

of prenylated and geranylated chalcone natural products by regiospecific iodination

and Suzuki coupling reactions. Tetrahedron Letters, 55(4), 897-899.

Wang, K., Goldman, M. E., Emge, T. J., & Goldman, A. S. (1996). Transfer-

dehydrogenation of alkanes catalyzed by rhodium(I) phosphine complexes. Journal

of Organometallic Chemistry, 518(1–2), 55-68.

Yang, Y., Tan, Y. X., Chen, R. Y., & Kang, J. (2014). The latest review on the polyphenols

and their bioactivities of Chinese Morus plants. J Asian Nat Prod Res, 16(6), 690-

702.

Yang, Y., Wang, H. Q., & Chen, R. Y. (2010). [Flavonoids from the leaves of Morus alba

L]. Yao Xue Xue Bao, 45(1), 77-81.

149

Yen, G-C., Wu, S-C., & Duh, P-D. (1996). Extraction and Identification of Antioxidant

Components from the Leaves of Mulberry (Morus alba L.). Journal of Agricultural

and Food Chemistry, 44(7), 1687-1690.

Yuan, C., Du, B., Yang, L., & Liu, B. (2013). Bioinspired Total Synthesis of Bolivianine:

A Diels– Alder/Intramolecular Hetero-Diels–Alder Cascade Approach. Journal of

the American Chemical Society, 135(25), 9291-9294.

Yus, M., Foubelo, F., & Ferrández, José V. (2001). Stereoselective Reductive Opening of

2,3-Benzofuran − A Two-Step Synthesis of 2H-Chromenes Including

Deoxycordiachromene. European Journal of Organic Chemistry, 2001(15), 2809-

2813.

Zeng, H., Ju, J, & Hua, R. (2011). ReCl(CO)5-catalyzed cyclocondensation of phenols with

2-methyl-3-butyn-2-ol to afford 2,2-dimethyl-2H-chromenes. Tetrahedron Letters,

52(30), 3926-3928.

Zhang, Q. J., Ni, G., Wang, Y. H., Chen, R. Y., & Yu, D. Q. (2009). Three new Diels-Alder

type adducts from the stem bark of Morus cathayana. J Asian Nat Prod Res, 11(3),

267-273.

Zhang, Q-J., Tang, Y-B., Chen, R-Y., & Yu, D-Q. (2007). Three New Cytotoxic Diels–

Alder-Type Adducts from Morus australis. Chemistry & Biodiversity, 4(7), 1533-

1540.

Zhao, W., & Carreira, E. M. (2003). Facile one-pot synthesis of photochromic pyrans. Org

Lett, 5(22), 4153-4154.

150

APPENDIX

SPECTRA OF COMPOUNDS

1H NMR and

13C NMR of 1-(4-(ethoxymethoxy)-2-hydroxy-3-iodophenyl)ethanone

151

1H NMR and

13C NMR of compound 171

*impurity

*

152

1H NMR and

13C NMR of 172

153

1H NMR and

13C NMR of 174

154

1H NMR and

13C NMR of exo-adduct 180

*impurity

* *

155

1H NMR and

13C NMR of diastereomer 182 (without the EOM protecting group)

*impurities

**

*

*

*

* *

156

1H NMR and

13C NMR of diastereomer 183

*impurities

* *

*

157

1H NMR and

13C NMR of diastereomer 184

158

1H NMR and

13C NMR of diastereomer 186

*MeOD-d4

159

1H NMR and

13C NMR of compound 190

160

1H NMR and

13C NMR of compound 192

161

1H NMR of compound 202

162

13C NMR of compound 203

163

1H NMR and

13C NMR of compound 191

164

1H NMR and

13C NMR of compound 204

*impurities and solvents peaks

*

165

1H NMR and

13C NMR of compound 205

*impurities/ grease peaks

*

*

* *

166

1H NMR of compound 206

*impurities

*

*

167

13C NMR of compound 206

*

*impurities

*

168

HSQC for morusalbanol A pentamethyl ether 206 (CDCl3)

169

HMBC for morusalbanol A pentamethyl ether 206 (CDCl3)

170

Comparison of 1H NMR of morusalbanol A pentamethyl ether 206 in CDCl3 (top), and toluene-d8 (bottom)

171

Comparison of 13

C NMR of morusalbanol A pentamethyl ether 206 in CDCl3 (top), and toluene-d8 (bottom)

172

DEPT-135 for morusalbanol A pentamethyl ether 206 in toluene-d8

173

Comparison of 1H NMR of morusalbanol A pentamethyl ether 206 in toluene-d8 at 25

oC (top) and 80

oC (bottom)

174

Comparison of 1H NMR of morusalbanol A pentamethyl ether 205 at 25

oC and 80

oC (Expansion in the region 1.5 – 4.5 ppm, toluene-d8)

175

13C NMR for morusalbanol A pentamethyl ether 206 at 80

oC (toluene-d8)

176

1H NMR for 207(400 MHz, CDCl3)

*impurities

*

*

*

177

13C NMR (bottom) and DEPT-135 (top) for 207 (CDCl3)

178

HSQC NMR for 207 (CDCl3)

179

HMBC NMR for 207 (CDCl3)

180

Comparison of 1H NMR of 207 in CDCl3 (top) and DMSO-d6 (bottom)

impurities and solvents peaks

*

*

*

*

*

* *

*

181

Comparison of 13

C NMR of 207 in CDCl3 (top) and DMSO-d6 (bottom)

impurities and solvents peaks

*

*

182

1H NMR and

13C NMR of compound 224

183

1H NMR and

13C NMR of compound 226

184

1H NMR and

13C NMR of compound 227

185

1H NMR and

13C NMR of compound 209

186

1H NMR and

13C NMR of compound 232

187

1H NMR and

13C NMR of compound 233

188

1H NMR and

13C NMR of compound 234

189

1H NMR and

13C NMR of compound 237

190

1H NMR and

13C NMR of compound 238

191

1H NMR and

13C NMR of compound 230

192

1H NMR and

13C NMR of compound 231

193

LIST OF PUBLICATIONS

1) Efficient one-pot synthesis of 2,2-dimethyl-2H-chromenes via Pd(II)catalyzed

coupling and SiO2-promoted condensation of o-halophenols with 2-methy-2-buten-

2-ol. Jia Ti Tee, Marzieh Yaeghoobi, Chin Fei Chee, and Noorsaadah Abd.

Rahman, Accepted in Syn. Commun. 2015

2) Model studies on construction of the oxabicyclic [3.3.1] core of mulberry Diels-

Alder adduct morusalbanol A. Jia Ti Tee, Chin Fei Chee, Michael J. C. Buckle,

Vannajan Sanghiran Lee, Wei Lim Chong, Hamid Khaledi, and Noorsaadah Abd

Rahman, Accepted in Tetrahedron Lett. 2015

3) A short diastereoselective synthesis of ()-morusalbanol A methyl ethers. Jia Ti

Tee, Chin Fei Chee, Hamid Khaledi, and Noorsaadah Abd Rahman, (Accpeted in

Synthesis, 2016)

Presentation at the following conferences:

1) Oral Presentation, 2nd

Junior International Conference on Cutting Edge Organic

Chemistry in Asia, 2nd

ICCEOCA (11-14 December 2012), Nanyang Technology

University (Singapore)

2) Poster Presentation, “Studies towards the synthesis of biological active flavonoid

Diels-Alder”, at 5th

HOPE Meeting (26 February - 2 March, 2013), Grand Prince

Hotel New Takanawa (Tokyo, Japan)

3) Oral Presentation, ‘Approach towards the Total Synthesis of Morusalbanol A and

Sorocein B’ 3rd

Junior International Conference on Cutting Edge Organic Chemistry

in Asia, 3rd

ICCEOCA (22-25 November 2013), Seimei-No-Mori, Nanyang

Technology University (Chiba, Japan)

194

4) Poster Presentation, ‘Total Synthesis of Morusalbanol A and Sorocein B’ 5th

UM-

NUS-Chulalongkorn Trilateral Meeting/Seminar (10-12 February 2014), University

Malaya, Malaysia

5) Poster Presentation, ‘Approach towards the Total Synthesis of Morusalbanol A and

Sorocein B’, 4th

Junior International Conference on Cutting Edge Organic

Chemistry in Asia, JICCEOCA-1 (28-30 November 2014), Chulabhorn Research

Institute (CRI) (Bangkok, Thailand)