synthesis of some naturally occurring quinones

190
Synthesis of Some Naturally Occurring Quinones Linda Birgitta Nielsen B.A., B.Sc. (Hons) This thesis is presented for the degree of Doctor of Philosophy of the University of Western Australia Discipline of Chemistry School of Biomedical, Biomolecular and Chemical Sciences 2008

Upload: others

Post on 08-Nov-2021

7 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: Synthesis of Some Naturally Occurring Quinones

Synthesis of Some Naturally Occurring

Quinones

Linda Birgitta Nielsen B.A., B.Sc. (Hons)

This thesis is presented for the degree of Doctor of Philosophy of

the University of Western Australia

Discipline of Chemistry

School of Biomedical, Biomolecular and Chemical Sciences

2008

Page 2: Synthesis of Some Naturally Occurring Quinones

DECLARATION FOR THESES CONTAINING PUBLISHED WORK AND/OR WORK PREPARED FOR PUBLICATION The examination of the thesis is an examination of the work of the student. The work must have been substantially conducted by the student during enrolment in the degree. Where the thesis includes work to which others have contributed, the thesis must include a statement that makes the student’s contribution clear to the examiners. This may be in the form of a description of the precise contribution of the student to the work presented for examination and/or a statement of the percentage of the work that was done by the student. In addition, in the case of co-authored publications included in the thesis, each author must give their signed permission for the work to be included. If signatures from all the authors cannot be obtained, the statement detailing the student’s contribution to the work must be signed by the coordinating supervisor. Please sign one of the statements below. 1. This thesis does not contain work that I have published, nor work under review for publication. Signature......................................................................................................................................................... 2. This thesis contains only sole-authored work, some of which has been published and/or prepared for publication under sole authorship. The bibliographical details of the work and where it appears in the thesis are outlined below. Signature......................................................................................................................................................... 3. This thesis contains published work and/or work prepared for publication, some of which has been co-authored. The bibliographical details of the work and where it appears in the thesis are outlined below.The student must attach to this declaration a statement for each publication that clarifies the contribution of the student to the work. This may be in the form of a description of the precise contributions of the student to the published work and/or a statement of percent contribution by the student. This statement must be signed by all authors. If signatures from all the authors cannot be obtained, the statement detailing the student’s contribution to the published work must be signed by the coordinating supervisor.

Chapter 2 and Chapter 4 include work that has been published. Details of the publication,

together with an estimate of the percentage contribution of each of the authors, appear below:

Linda B. Nielsen (90%) and Dieter Wege (10%), ‘The enantioselective synthesis of elecanacin through an

intramolecular naphthoquinone-vinyl ether photochemical cycloaddition’, Org. Biomol. Chem., 2006, 4,

868.

The corresponding author Assoc. Prof. Dieter Wege has given permission for the results presented in this

publication to be included in this thesis.

Signatures......................................................................................................................................................... Signatures.........................................................................................................................................................

i

Page 3: Synthesis of Some Naturally Occurring Quinones

Summary

Naturally occurring quinones have attracted considerable interest due to their

widespread occurrence, structural diversity and often potent biological activities. The

research outlined in this thesis involves the development of synthetic approaches to two

novel naphthoquinone derivatives, both of which were discovered during investigations

into the bioactive constituents of tropical plant species. Chapter 1 introduces the family

of quinonoid compounds and also considers the important role that natural product

synthesis can play in structural confirmation and in providing an adequate supply of

compounds for further research.

Chapter 2 describes the synthesis of elecanacin 36, an unusual cyclobuta-fused

naphthalene-1,4-dione derivative which has been isolated from the bulbs of the iris

Eleutherine americana Merr. et Heyne (Iridaceae), along with the isomeric and well-

known pyranonaphthoquinones eleutherin 38 and isoeleutherin 39.

O

OO

HMeO

Me H

O

O

O

Me

Me

MeO

O

O

O

Me

Me

MeO

36 38 39

(±)-Elecanacin 36 was prepared, together with its epimer (±)-isoelecanacin 151, by an

intramolecular [2 + 2] photocycloaddition resulting from irradiation of 5-methoxy-2-(2-

vinyloxypropyl)naphthalene-1,4-dione 138.

ii

Page 4: Synthesis of Some Naturally Occurring Quinones

O

O

MeO

O

Me

O

O O

H

MeO

Me

O

O

MeO

O

H

Me

CH2Cl2hv, 350 nm

(±)-elecanacin

15136138(±)-isoelecanacin

(one enantiomer of each product is depicted)

The synthesis of enantiopure elecanacin 36 starting with (R)-propylene oxide was also

achieved. This established the absolute configuration of the natural product and revealed

that the sample isolated from the bulbs possessed an enantiomeric excess of only 14%.

A possible biosynthetic pathway which relates elecanacin 36 to its co-metabolites

eleutherin 38 and isoeleutherin 39 is also discussed.

Chapter 3 focuses on an approach to 3-hydroxymethylfuro[3,2-b]naphtho[2,3-

d]furan-5,10-dione 37, which has been isolated from the wood of the tropical tree

Crescentia cujete L. (Bignoniaceae) and incorporates a rare fully aromatic furofuran

moiety.

O

O

O

O

OH

37

The route to the furofuranonaphthoquinone 37 commenced with the synthesis of

ethyl 3-[3-(2-furyl)-1,4-dimethoxynaphthalen-2-yloxy]propynoate 263, which proved to

be highly unstable and could not be isolated. However, the pendant acetylenic ether

chain within 263 could be trapped intramolecularly by the attached furan moiety. Thus

iii

Page 5: Synthesis of Some Naturally Occurring Quinones

263 was instead converted directly into an advanced intermediate ethyl 5,10-

dimethoxyfuro[3,2-b]naphtho[2,3-d]furan-3-carboxylate 304 by way of a tandem

intramolecular cycloaddition-cycloreversion sequence. The key ester 304 was then

elaborated via a series of standard oxidation and reduction reactions to give the target

quinone 37.

OMe

OMe

O

O

CO2Et O

O

OMe

OMe

CO2Et

NN N

N

py

py

OMe

OMe

O

O

CO2Et

toluene, reflux

299263

304

heat

O

O

O

O

OH

37

1. CAN, MeCN/H2O2. LiAlH4, THF3. ON(SO3K)2

The synthetic routes to elecanacin 36 and furofuranonaphthoquinone 37 arising

from this work represent the first total syntheses of both compounds and have enabled

confirmation of their novel ring systems.

iv

Page 6: Synthesis of Some Naturally Occurring Quinones

Statement of Candidate Contribution

Except where specific acknowledgement of others is made, the work described

in this thesis is original and was carried out by the author in the Discipline of Chemistry

(School of Biomedical, Biomolecular and Chemical Sciences) at the University of

Western Australia, under the supervision of Assoc. Prof. Dieter Wege.

Chapter 2 and Chapter 4 include work that has been published. Details of the

publication, together with an estimate of the percentage contribution of each of the

authors, appear below:

Linda B. Nielsen (90%) and Dieter Wege (10%), ‘The enantioselective synthesis of

elecanacin through an intramolecular naphthoquinone-vinyl ether photochemical

cycloaddition’, Org. Biomol. Chem., 2006, 4, 868.

The corresponding author Assoc. Prof. Dieter Wege has given permission for the results

presented in this publication to be included in this thesis.

Assoc. Prof. Dieter Wege Linda B. Nielsen

v

Page 7: Synthesis of Some Naturally Occurring Quinones

Acknowledgements

First I would like to thank my supervisor Associate Professor Dieter Wege for

all his guidance, support and encouragement. I was lucky enough to be Dieter’s final

PhD student and he very kindly continued to oversee my work after he retired. I have

benefitted greatly from my supervisor’s wisdom, insight and years of experience

solving interesting problems, working at the bench and finding all the best sources in

the library, as well as his super stories. For this I am deeply grateful.

I would also like to give a special thank you to a number of people who helped

me throughout the course of this project. They include Dr Lindsay Byrne who assisted

with the 2D NMR spectra and chiral shift reagent experiments and gave me many

invaluable and enthusiastic NMR lessons, as well as Dr Anthony Reeder who recorded

the mass spectra. I would like to thank Professor Robin Giles (Murdoch University,

Western Australia) who kindly provided the chiral HPLC column, Dr Gavin Flematti

who assisted with HPLC analyses and Professor Y. Imakura (Naruto University,

Takashima) who generously supplied the 13C and 1H NMR spectra of natural

elecanacin. Nicole Hondow’s help with the IR machines was also greatly appreciated. In

addition, I am grateful to the department’s many present and former technical and

administrative staff, including Sarah Davies (store and glass workshop), Greg Cole

(glass workshop), George Sjepcevich (workshop), Kim Foo (first year labs) and Ingrid

Buschmann (office), for all their friendly assistance and expertise.

Thanks must also go to the last members of the Wege Group for many coffees

and much enthusiastic conversation. I would especially like to thank Alan Payne who

made sure that I became well-versed in almost all the lab techniques that I was to face

during my PhD. I would also like to thank Rosenani Haque, along with the members of

vi

Page 8: Synthesis of Some Naturally Occurring Quinones

the Bucat group (especially Janette Head), the Theoretical Chemistry group (including

Daniel Grimwood and Dylan Jayatilaka) and the Berners-Price group (including

Anthony Humphries, Donald Thomas, Joseph Moniodis, Peter Barnard, James Hickey,

Louise Wedlock, Scott McPhee, Junyong Zhang and Mikie Farmer) who all ensured

that I was never lonely and continued to have a great time during the years I spent

working in the Wege lab alone. I express my deepest gratitude to Associate Professor

Emilio Ghisalberti and the full- and part-time members of his group: my office buddy

Gavin Flematti, Laura Clayson and Stuart Ingham (the two nicest honours students

anyone could ever hope to work with), Chuck Heath, Chris Jones, Jon See and Katie

Punch for welcoming me into their lab when we moved into the new chemistry building

and for their continuing friendship over the last couple of years. Thank you also to Dr

Allan McKinley who kindly provided me with a desk and plently of space to write in

his lab during my final year.

Many thanks to my friends from beyond the lab, especially Junming Ho, Alistair

Usher and Nicole Hondow for their long and inspiring letters and my musical friends

Silvia Sze, Birger Dittrich and Helen Roberts who came with me to lots of concerts and

joined me for more than a few coffees afterwards.

Finally I would like to thank my family: Ing-Marie, John and Peter, who have

given me enormous support and encouragement and still continue to take an interest in

my studies after so many years.

vii

Page 9: Synthesis of Some Naturally Occurring Quinones

Table of Contents

Summary ii

Statement of Candidate Contribution v

Acknowledgements vi

Chapter 1: Introduction to Naturally Occurring Quinones 1

1.1 Introduction to Quinones 2

1.2 Naturally Occurring Quinones 4

1.3 Naturally Occurring Quinones as Pharmaceutical Targets 7

1.4 Natural Product Synthesis 13

1.5 Aim of Research 15

Chapter 2: The Synthesis of Elecanacin 17

2.1 Introduction 18

2.1.1 Elecanacin 36 and Other Naphthoquinones from Eleutherine

americana and Eleutherine bulbosa 18

2.1.2 Syntheses of Eleutherin 38 and Isoeleutherin 39 20

2.1.3 Elecanacin 36 as a Synthetic Target 34

2.2 [2 + 2] Photocycloadditions for the Synthesis of Cyclobutane Rings 35

2.2.1 Mechanism of the Enone-Alkene [2 + 2] Photocycloaddition 37

2.2.2 Intermolecular [2 + 2] Photocycloadditions 39

2.2.3 Intramolecular [2 + 2] Photocycloadditions 40

2.2.4 [2 + 2] Photocycloadditions of 1,4-Naphthoquinones 41

2.3 An Approach to Elecanacin 36 44

viii

Page 10: Synthesis of Some Naturally Occurring Quinones

2.4 Preparation of (±)-Elecanacin 36 46

2.4.1 Synthesis of 2-(2-Hydroxypropyl)-5-methoxynaphthalene-1,4-

dione 46 46

2.4.2 Synthesis of 5-Methoxy-2-(2-vinyloxypropyl)naphthalene-1,4-

dione 138 52

2.4.3 Synthesis of (±)-Elecanacin 36 54

2.5 The Enantioselective Synthesis of Elecanacin 36 61

2.5.1 An Approach to Chiral 1-(1-Hydroxy-5-methoxynaphthalen-

2-yl)propan-2-ol 152 using Jacobsen’s Catalyst 61

2.5.2 Synthesis of Epoxides 155, 156 and 168 and Attempted Hydrolytic

Kinetic Resolution 64

2.5.3 A Directed Metallation Approach to (2R)-1-(1-Hydroxy-5-methoxy-

2-naphthalenyl)propan-2-ol 152 70

2.5.4 Final Steps in the Preparation of Enantiopure Elecanacin 36 76

2.6 On the Possible Biosynthesis of Elecanacin 36 77

2.7 Concluding Remarks 82

Chapter 3: The Synthesis of 3-Hydroxymethylfuro[3,2-b]naphtho[2,3-

d]furan-5,10-dione 83

3.1 Introduction to 3-Hydroxymethylfuro[3,2-b]naphtho[2,3-d]furan-5,10-

dione 37 84

3.2 A Synthetic Approach to 3-Hydroxymethylfuro[3,2-b]naphtho[2,3-

d]furan-5,10-dione 37 90

3.3 Previous Work Towards the Synthesis of 3-Hydroxymethylfuro[3,2-

b]naphtho[2,3-d]furan-5,10-dione 37 93

3.4 An Alternative Approach to an Intermediate Acetylenic Ether 100

ix

Page 11: Synthesis of Some Naturally Occurring Quinones

x

3.5 Attempted Synthesis of Acetylenic Ether 263 104

3.6 Synthesis of Key Intermediate Ethyl 5,10-Dimethoxyfuro[3,2-

b]naphtho[2,3-d]furan-3-carboxylate 304 114

3.7 Final Steps in the Synthesis of 3-Hydroxymethylfuro[3,2-b]naphtho[2,3-

d]furan-5,10-dione 37 125

3.8 Concluding Remarks 128

Chapter 4: Experimental 133

4.1 General Details 134

4.1.1 Solvents and Reagents 134

4.1.2 Reactions and Chromatography 135

4.1.3 Characterisation 136

4.2 Experimental for Chapter 2 138

4.3 Experimental for Chapter 3 156

References 167

Page 12: Synthesis of Some Naturally Occurring Quinones

Chapter 1

Introduction to Naturally Occurring Quinones

1

Page 13: Synthesis of Some Naturally Occurring Quinones

1.1 Introduction to Quinones

Quinones are one of the oldest recognized classes of organic compounds and

have fascinated chemists since the early days of modern chemistry. The name

“quinone” (“chinon” in German) originates from the name given in the 1830s to a bright

yellow compound obtained by oxidation of quinic acid in Liebig’s laboratory by

Woskresensky, who referred to the product as “chinoyl” (Scheme 1).1 Quinic acid was

obtained from chinona bark and now is known to have structure 1 and “chinoyl” is 1,4-

benzoquinone 2.

OO

"chinoyl"

2

quinic acid

OHOH

OHOH

HO2C

1

MnO2H2SO4

Scheme 1

Structurally, quinones can be considered to be cross-conjugated systems, which

incorporate alternating double bonds including exocyclic oxygens, as exemplified by

the simplest members of the class 1,4-benzoquinone 2 and 1,2-benzoquinone 3 (Figure

1).

O

O

2

1,4-benzoquinone(p-benzoquinone)

O

O

1,2-benzoquinone(o-benzoquinone)

3

Figure 1

2

Page 14: Synthesis of Some Naturally Occurring Quinones

However, looking beyond these parent compounds, the quinone class encompasses a

remarkable array of molecular structures. Fusion of benzene rings to the parent

structures 2 and 3 generates polycyclic quinones. Some representative compounds are

shown in Figure 2, which includes simple derivatives such as the naphthoquinones 4

and 5, as well as more complex extended compounds such as 7 and 9.

O

O

O

O

O

O

O

O

O O

O O

1,4-naphthoquinone

4 5

1,2-naphthoquinone

6

benzo[a]pyrene-6,12-quinone

8violanthrone 9,10-anthraquinone

9

9,10-anthraquinone

7

Figure 2

One of the important characteristic properties of quinones is the facility with

which they undergo reversible reduction to hydroquinones via a semiquinone

intermediate, as shown in Scheme 2.

3

Page 15: Synthesis of Some Naturally Occurring Quinones

O

O O

OH

OH

OH

2 10 11quinone semiquinone hydroquinone

e- + H+

- ( e- + H+ )

e- + H+

- ( e- + H+ )

Scheme 2

This redox chemistry appears to play a role in the biological function of many naturally

occurring quinones,2 which are introduced in the following section.

1.2 Naturally Occurring Quinones

Quinones occur widely in nature and are most commonly found in plants, fungi

and bacteria. Other naturally occurring quinones are encountered in the animal

kingdom, where they are especially prevalent among arthropods and some marine

organisms, including sea urchins and other echinoderms.3 Several members of the

quinonoid family are of importance due to their involvement in key physiological

processes. In addition, many quinones are of medicinal interest, exhibiting a broad

range of pharmacological activities.

Some quinones have an essential biochemical function, which reflects their

characteristic oxidation/reduction chemistry. These include the widely distributed

ubiquinone 12, also known as coenzyme Q, which plays a vital role in cellular

respiration as an electron carrier (Figure 3).4-6 Coenzyme Q has also been proposed to

function as an anti-oxidant, with the hydroquinone form behaving as an effective free

radical scavenger.5

4

Page 16: Synthesis of Some Naturally Occurring Quinones

O

O

MeO

MeO H

Me

n n = 6-10

12

Me

Figure 3

Other naturally occurring quinones appear to have a defensive role, exhibiting

toxic or unpalatable properties which deter predators or competing species. A well-

known example is the naphthoquinone juglone 15, a plant growth inhibitor which

originates from the walnut tree Juglans nigra. Within the tree, a precursor to juglone

exists in a reduced bound form as a glycoside 13 (Scheme 3).7

OH

OGlcOH

OH

OHOH

O

OOH

oxidationhydrolysis

151413juglone

Scheme 3

Upon release from the leaves, fruit and roots and subsequent exposure to the air, the

hydroquinone 14 undergoes oxidation to the corresponding quinone 15,7, 8 which

accumulates in the soil under walnut trees.9 Juglone 15 has been shown to effectively

inhibit the growth of numerous plant species, even at micromolar concentrations,8, 10

and can reach toxic levels in the soil which eventually causes the death of neighbouring

plant species.9, 11

5

Page 17: Synthesis of Some Naturally Occurring Quinones

Numerous quinones have been isolated which possess a variety of potentially

useful biological properties, including antifungal, antiviral, antibiotic and antitumour

activities. Several bioactive quinone derivatives are shown in Figure 4, which illustrates

the wide-ranging biological properties and structural diversity that exist within the

quinonoid class of compounds.

OH

HO

O

O

O

H OH

HMe

O

OH

Me

HO

Me

OH

Me Me

O

MeO

O

MeMe

N

O

O

Me

MeO Me

O

thysanone, a fungal metabolite from Thysanophora penicilloides)

antiviral properties12

16

18

17

a plant metabolite from Bobgunnia madagascariensis

antifungal properties13

Cryptotanshinone, a plant metabolitefrom Salvia miltiorrhiza

antibacterial,14 and anti-inflammatory15

properties

19

mimosamycin, a marine sponge metabolite from Cribrochalina sp.

antitumour properties16

Figure 4

6

Page 18: Synthesis of Some Naturally Occurring Quinones

The overall importance of quinones in chemistry can be gleaned from a recent

monograph of over 1000 pages which summarizes modern synthetic methods for the

synthesis of quinones and their heteroatom analogues.17

1.3 Naturally Occurring Quinones as Pharmaceutical Targets

Given the often potent biological activity associated with many naturally

occurring quinones, it is not surprising that a considerable effort has been expended in

isolating, evaluating and synthesising novel quinones in the search for compounds with

potentially useful and selective pharmacological properties. In particular, quinones

constitute one of the largest classes of antitumour agents known and several naturally

occurring quinones, as well as a number of synthetic derivatives, have been developed

as important anticancer therapies.2 The quinone moiety and its associated redox

properties appear to be involved in the mechanism of action for cytotoxicity in many

quinones,2, 18 including the powerful antitumour agents mitomycin C 20 and the

anthracycline antibiotics doxorubicin 21 and daunorubicin 22.

The antibiotic mitomycin C 20 was first isolated in 1958 from the soil

microorganism Streptomyces caespitosus and became the first naturally occurring

quinone to be adopted in the clinic as an anticancer therapy (Figure 5).19 Even before

the structure was elucidated in the early 1960s,20-22 mitomycin C 20 received

considerable attention as it became quickly apparent that the compound exhibited

promising broad-spectrum antitumour activity.19, 21, 23

7

Page 19: Synthesis of Some Naturally Occurring Quinones

N NH

O

O

H2N

Me

O

ONH2

OMe

20

Figure 5

Mitomycin C 20 appears to target tumour cells via damage caused by DNA alkylation.

The most widely accepted mode of action for the drug is proposed to operate according

to the pathway in Scheme 4.2, 19, 23 Mitomycin C 20 first requires activation during a

bioreduction step in which 20 is transformed into the corresponding hydroquinone 23,

which is highly unstable and undergoes a spontaneous loss of methanol. Subsequent

protonation and opening of the aziridine ring of 24 generates a highly reactive o-

quinone methide 25, which is susceptible to nucleophilic attack by DNA to give a

covalently bonded drug-DNA complex 26. X-ray crystallography has revealed that

attack on 25 by DNA occurs primarily via the N2 position of a guanine residue.24 This

resulting drug-DNA intermediate 26 is then able to form a cross-linkage to the N2 of

another guanine residue. Such covalent cross-linkages are believed to cause a structural

distortion of the DNA helix, which prevents normal uncoiling and participation in

replication and cell division, hence stopping the growth of tumour cells.25

8

Page 20: Synthesis of Some Naturally Occurring Quinones

N NH

O

O

H2N

Me

O

ONH2

OMe

N

O

OH

H2N

Me

O

ONH2

NH2

H

N

OH

OH

H2N

Me

O

ONH2

NH2

DNA OCONH2

DNAN NH

OH

OH

H2N

Me

O

ONH2

N NH

OH

OH

H2N

Me

O

ONH2

OMe

N

OH

OH

H2N

Me

DNA

NH2

DNA

N

OH

OH

H2N

MeNH2

DNA

MeOH

H

20

28

2726

2425

23

reduction

Scheme 4

9

Page 21: Synthesis of Some Naturally Occurring Quinones

Unfortunately, the high reactivity associated with mitomycin C 20 has also proven to be

destructive towards the constituents of normal cells and produces severe side effects

including vomiting, liver and kidney toxicity, heart damage and bone marrow

depression.19 Although the use of mitomycin C 20 was popular during the 1960s,

especially in Japan where approximately half of all cancer patients received the drug as

part of their treatment, the associated side effects have largely limited mitomycin C’s

utility and the drug has generally fallen out of favour.19 Nevertheless, mitomycin C 20

remains an important component of several combination chemotherapeutic treatments

for breast, lung and prostate cancer.23 A similar bioreductive alkylation mechanism is

believed to play a role in the toxic and in particular antitumour properties of many

quinones, including dehydro-α-lapachone 29 and the alkannin derivative 30 (Figure

6).26

O

O

O Me

Me

O

O OAc

Me

Me

OH

OH

29 30

Figure 6

Doxorubicin (adriamycin) 21 and daunorubicin (daunomycin) 22 are also

naturally occurring quinones that are employed as anticancer drugs (Figure 7). They

were first isolated in the 1960s from Streptomyces peucetius and belong to the large

class of structurally related compounds known as the anthracycline antibiotics.27, 28

10

Page 22: Synthesis of Some Naturally Occurring Quinones

OMeO

O

O

OH

OH

O

OH

O

NH2OH

Me

OH

OMeO

O

O

OH

OH

O

CH3

O

NH2OH

Me

OH

2221

Figure 7

Doxorubicin 21 has proven to be a particularly important drug, possessing potent broad-

spectrum antitumour activity, and it continues to be used widely for the treatment of

acute leukaemias and many solid tumours.2, 27 Daunorubicin 22 has more limited

activity but is an effective treatment in combination with the drug cytarabine for adult

leukaemias.27 Although the exact mode of action for 21 and 22 is not yet understood,

the antitumour activity associated with the anthracycline antibiotics has been proposed

to arise from a combination of several different pathways. First, the planar disk-like

anthracyclines are known to target DNA through intercalation. X-ray crystallography

has revealed that the drugs associate closely with various base pairs, which distorts the

DNA helix.29 This tight binding appears to disrupt normal DNA replication and

transcription, thus halting cell division and the growth of tumours.29, 30 Second, both

doxorubicin 21 and daunorubicin 22 appear to inhibit the action of the enzyme

topoisomerase II, which has an essential function in cleaving and resealing DNA strands

during replication and transcription.31 It has been proposed that the anthracyclines are

able to stabilize the topoisomerase II-cleaved DNA complex. This leads to a large

increase in the number of DNA strand breaks, which disrupts replication and eventually

results in cell death.32 Proliferating cells, including tumour cells, have a particularly

high concentration of topoisomerase II and are therefore very vulnerable to compounds

11

Page 23: Synthesis of Some Naturally Occurring Quinones

which interfere with the enzyme.31, 32 A third likely mode of action for DNA damage

involves a radical based pathway, which is illustrated for doxorubicin 21 in Scheme 5.27

OMeO

O

O

OH

OH

O

OH

O

NH2OH

Me

OH

OMeO

O

O

OH

OH

O

OH

O

NH2OH

Me

OH

OMeO

O

O

OH

OH

O

OH

O

NH2OH

Me

OH

O2O2

O2

O2

one electron reduction

one electron reduction

21

32

31

Scheme 5

This pathway commences with a one electron reduction of the quinonoid drug 21 to

give a semiquinone derivative 31, which is able to react with molecular oxygen to

generate a superoxide radical anion. These radicals are unstable in the aqueous cell

environment and dismutase to hydrogen peroxide, which decomposes on reaction with

metal ions (via Fenton-type chemistry) to give hydroxyl radicals. The hydroxyl radicals

12

Page 24: Synthesis of Some Naturally Occurring Quinones

are highly reactive and can damage DNA and other cell components.2, 18 Unfortunately,

the oxidative damage caused by the generation of superoxide radicals also seems to be

largely responsible for the cardiotoxicity and potentially severe heart damage associated

with the use of anthracycline antibiotics, which greatly limits the dose able to be

administered during chemotherapy.27, 33, 34 Despite this serious side effect, the

anthracycline antibiotics, especially doxorubicin 21, are amongst the most effective and

widely used anticancer pharmaceuticals ever developed.2, 28

1.4 Natural Product Synthesis

Given the wide-spread occurrence and biological activity of naturally occurring

quinones and especially their potential as pharmaceuticals and pharmaceutical leads, it

is important to have a thorough understanding of their properties. This is particularly

true in the area of medicinal chemistry where an understanding of natural product

structure-activity relationships may enable the design of analogues, which exhibit less

toxicity and more selective and potent pharmacological activity.35-37 One of the things

that hampers natural product research is the supply of the compounds of interest as they

are often isolated in only milligram amounts from their source. While these small

amounts are often sufficient for structural elucidation, they do not allow for a thorough

biological evaluation.37 Total synthesis can play an important role in structural

confirmation and in providing an adequate supply of compounds for further research.

This is especially true of natural products that are obtained from hard to access

environments, such as the ocean, or those that are only present in tiny quantities.36, 37

Throughout the 19th Century and for most of the first half of the 20th Century,

total synthesis of a natural product served as the final proof of the proposed structural

13

Page 25: Synthesis of Some Naturally Occurring Quinones

assignment. The latter was formulated on the basis of an often painstaking series of

chemical degradations.38 The period after the Second World War saw rapid advances in

the development of powerful analytical methods including spectroscopy (UV, NMR,

IR), X-ray crystallography and mass spectrometry so that by the late 1960s, the classical

methods of structure determination had been largely replaced by these modern

techniques.38 Today, a combination of these techniques often allows for the correct

structural formulation of quite complex compounds, even before the advent of any

synthetic work.38, 39 However, synthesis continues to play an important role in the area

of structural elucidation. In particular, synthesis maybe necessary for providing

stereochemical information about a natural product, which is often beyond the reach of

X-ray and spectroscopic techniques alone.38, 40 For example, only in occasional cases

when a suitable crystalline sample exists can X-ray analysis be used to establish the

absolute configuration of a natural product.40 More often, at least one or more

stereoisomers will need to be synthesised in order to determine the correct

configuration.38, 40 Moreover, total synthesis remains an important method for verifying

a proposed structure as structural misassignment can occur. As numerous examples

highlighted by Mori41 and by Nicolaou and Snyder38 have demonstrated, it is not

entirely uncommon to complete a synthesis and discover that the original structural

assignment of a natural product is incorrect. One example from the field of quinone

research involves the discovery of the longstanding incorrect structure proposed for the

kinamycin antibiotics. The first members of the family kinamycin A-D were isolated in

the early 1970s from Streptomyces murayamaenis. The compounds were originally

formulated as N-cyanobenzo[b]carbazole derivatives, as exemplified by the postulated

structure for kinamycin C 33, on the basis of spectroscopic, X-ray and chemical

evidence (Figure 8).42 However, the structural assignment was called into question after

14

Page 26: Synthesis of Some Naturally Occurring Quinones

spectroscopic data obtained from several synthetic N-cyano derivatives, including

compound 34, were inconsistent with those of the natural products.43, 44

N

O

O C

AcO

CH3

OAc

OH

OAc

OH

3433

C

O

O N

AcO

CH3

OAc

OH

OAc

OH

35

N

N

O

O CMeO

AcO

CH3

N N

Figure 8

Structural revision in 1994 based on reconsideration of the original X-ray and

spectroscopic evidence, as well as further NMR studies, resulted in the kinamycin

antibiotics being reassigned a diazobenzo[b]fluorene skeleton.44, 45 This revised

structure has recently been confirmed through the first total synthesis of (-)-kinamycin

C 35 by Lei and Porco.46

1.5 Aim of Research

The aim of the research described in this thesis was to develop synthetic

approaches to two novel naturally occurring naphthoquinone derivatives, both of which

were isolated during investigations into the bioactive constituents of tropical plant

species. The first target compound to be considered is a cyclobuta-fused dione 36,

which was isolated from the bulbs of the iris Eleutherine americana Merr. et Heyne

(Iridaceae) and was named elecanacin.47 This compound was deduced as 36 by means

of NMR spectroscopy (Figure 9). However, the absolute stereochemistry was not

determined and the compound was arbitrarily depicted as shown in Figure 9. The

15

Page 27: Synthesis of Some Naturally Occurring Quinones

second compound of interest was isolated from the tropical American tree Crescentia

cujete L. (Bignoniaceae) and was formulated as the furofuranonaphthoquinone 37, again

largely on the basis of NMR spectroscopy (Figure 9).48

O

OO

MeO

HMe

HO

O

O

O

OH

3736

Figure 9

The ring skeletons of both compounds are rather unusual for natural products and have

never previously been reported. Thus, it was felt that confirmation of the structures of

36 and 37 by synthesis was desirable. In addition, it was thought that the development

of an asymmetric approach to elecanacin 36 would possibly allow the absolute

configuration to be established. Work carried out to achieve the synthesis of elecanacin

36 is described in Chapter 2. The naphthoquinone 37 is of particular interest because it

exhibits selective cytotoxic properties and therefore may provide a possible lead

compound in the development of therapeutic anticancer agents. Thus it was thought that

devising a viable route to 37 should not only allow confirmation of the unusual

structure, but may also enable the synthesis of sufficient quantities of the compound for

further investigation into the potentially useful biological activity. Chapter 3 discusses

the development of a synthetic approach to furofuranonaphthoquinone 37.

16

Page 28: Synthesis of Some Naturally Occurring Quinones

Chapter 2

The Synthesis of Elecanacin

17

Page 29: Synthesis of Some Naturally Occurring Quinones

2.1 Introduction

2.1.1 Elecanacin 36 and Other Naphthoquinones from Eleutherine americana and

Eleutherine bulbosa

Extracts from the bulbs of members of the Iridaceae family have long been used

as a traditional medicine by a number of cultures,49-53 and this has led to investigations

into their potential bioactive constituents. In recent years a variety of related

naphthoquinone derivatives, isolated from the bulbs of the Iridaceae species Eleutherine

americana Merr. et Heyne and Eleutherine bulbosa (Miller) Urb., have been a subject

of interest because a number of them exhibit interesting and potentially useful

biological activity.

The white-flowered iris Eleutherine americana is cultivated on the Southern

Chinese island Hainan, and extracts of the bulbs have been used as a traditional

treatment for coronary disorders.49 During a search for bioactive constituents from the

bulbs, Hara and coworkers isolated a novel cyclobuta-fused dione, which they named

elecanacin.47 Elecanacin was assigned the structure 36 on the basis of NMR

spectroscopy and has a tetracyclic ring skeleton, composed of a naphthoquinone ring

fused to a cyclobuta[1,2-b]furan-derived moiety (Figure 10). Along with elecanacin 36,

Hara and coworkers also isolated two known pyranonaphthoquinones, eleutherin 38 and

isoeleutherin 39, which had been isolated previously from Eleutherine americana by

Chen and coworkers49 and are also important constituents of the bulbs of the related iris

Eleutherine bulbosa (Figure 10).

18

Page 30: Synthesis of Some Naturally Occurring Quinones

O

OO

HMeO

Me H

O

O

O

Me

Me

MeO

O

O

O

Me

Me

MeO

36 38 39

Figure 10

Eleutherine bulbosa originates from tropical Central America, South America

and the West Indies,54 where the bulbs have been used traditionally as an antifertility

agent55 and for the promotion of wound healing.51 The bulbs of Eleutherine bulbosa are

also used in Java as a diuretic and purgative, as well as for the treatment of jaundice,56

and have more recently been reported as a South African folk medicine for the treatment

of burns and gastro-intestinal disorders.52

Early work on the chemical constituents of the bulbs of Eleutherine bulbosa was

carried out by Schmid and coworkers who first isolated the pyranonaphthoquinone (+)-

eleutherin 38, along with its diastereomer (-)-isoeleutherin 39.57 The structures of these

compounds were assigned on the basis of chemical degradation and derivatisation

studies,57, 58 and later confirmed by synthesis.59 Eleutherin 38 displays slight

antibacterial activity against Pyococcus aureus, Streptococcus haemolyticus A60 and

Bacillus subtalis61 and also has anticancer properties, inhibiting topoisomerase II

activity.47, 62 Isoeleutherin 39 has been found to have inhibitory activity against HIV,47

and both eleutherin 38 and isoeleutherin 39 possess antifungal activity.53

Further work on the constituents of the bulbs of Eleutherine bulbosa by Alves

and coworkers uncovered another naphthoquinone derivative, (+)-eleutherinone, which

was assigned structure 40 on the basis of spectroscopic evidence (Figure 11).53 The

19

Page 31: Synthesis of Some Naturally Occurring Quinones

absolute stereochemistry was not determined and remains unknown. Biological testing

revealed that (+)-eleutherinone 40 exhibits antifungal activity.

O

O

MeO

O

Me

40

Figure 11

2.1.2 Syntheses of Eleutherin 38 and Isoeleutherin 39

Eleutherin 38 and isoeleutherin 39 have been a particular focus of attention

because they are members of the class of compounds known as the

pyranonaphthoquinone antibiotics, which have the naphtho[2,3-c]pyran-5,10-dione ring

system in common. Like eleutherin 38 and isoeleutherin 39, many

pyranonaphthoquinones exhibit potentially useful biological properties, including

activity against gram-positive bacteria, pathogenic fungi and yeasts, and a number of

them possess antiviral activity.63, 64 It has also been suggested that many

pyranonaphthoquinones, including eleutherin 38 and isoeleutherin 39, may act as

bioreductive alkylating agents with a mechanism of action similar to that of the

antitumour agent mitomycin C 20 discussed in Chapter One (p. 8).26, 63

Eleutherin 38 and isoeleutherin 39 are amongst the simplest members of the

pyranonaphthoquinone family and consequently their syntheses have attracted a

considerable amount of interest.65 A number of different approaches for the construction

of the naphthopyran rings of 38 and 39 have been developed.

20

Page 32: Synthesis of Some Naturally Occurring Quinones

The first reported synthesis of (±)-eleutherin 38 and (±)-isoeleutherin 39 was

devised by Schmid and Eisenhuth,59 and it begins with 5-methoxynaphthalen-1-ol 41

(Scheme 6). Treatment of 41 with allyl bromide gave the allyl ether 42 and a subsequent

Claisen rearrangement produced the naphthol 43, which was oxidised to the

corresponding 1,4-naphthoquinone 44 by treatment with Fremy’s salt. Reductive

cyclisation of 44, followed by oxidative ring-opening then afforded the alcohol

derivative 46. Finally, construction of the pyran ring was effected by condensation with

acetaldehyde under acidic conditions and gave a mixture of (±)-eleutherin 38 and (±)-

isoeleutherin 39.

ON(SO3K)2

O

O

MeO

MeO

OH

O

Me

MeO OH

MeO

O

O

O

O

Me

Me

MeO

Br

K2CO3, Me2CO

CH3CHO

H3PO4

O

O

O

Me

Me

MeO

MeO

O

Me

OH

O

MeO

OH

2. aq. HBr

+

1. SnCl2/ HCl EtOH

FeCl3aq. Me2CO

41

44

42 43

3938

46 45

Scheme 6

21

Page 33: Synthesis of Some Naturally Occurring Quinones

The generation of the pyranonaphthoquinone ring system in the final step of this

synthesis is, at first sight, mechanistically puzzling. Eisenhuth and Schmid have

suggested that the formation of the pyran ring does in fact proceed via the hydroquinone

47, generated by some H donor in the reaction medium, and subsequent intramolecular

electrophilic substitution within the oxacarbenium ion 49, derived from the hemiacetal

48 (Scheme 7). Oxidation of 50 then delivers eleutherin 38 and isoeleutherin 39.

MeO

O

Me

OH

O

O

O

O

Me

Me

MeO

O

O

O

Me

Me

MeO

MeO

OH

Me

OH

OH MeO

OH

Me

O

OH

Me

OH

MeO

OH

Me

O

OH Me

O

OH MeMeO

Me

OH

[H] MeCHO

H+

[O]

+

39

38 50 49

484746

Scheme 7

Support for this sequence comes from the observation that pre-reduction of 46 and later

addition of benzoquinone to the reaction mixture increased the yield of eleutherin 38

and isoeleutherin 39, presumably by facilitating the oxidation step in the generation of

eleutherin 38 and isoeleutherin 39 from 50.

22

Page 34: Synthesis of Some Naturally Occurring Quinones

A biomimetic synthesis of (±)-eleutherin 38 and (±)-isoeleutherin 39 from a

polyketide chain has been achieved by Webb and Harris (Scheme 8).66, 67

Me Me

O O

N

OEt

OEt

O

O

O

OH O

Me

MeOH OH

O

Me

Me

OH

O

Me

Me

MeO MeO OH

O

Me

Me

ON(SO3K)2

N

Me

O O O

Me

O OO

MeO O

O

Me

Me

O

MeO O

O

Me

Me

O

Me

Me

OHOH O

O

1. LDA, THF

2.

H2 , 5% Pd/CEtOH

CH2N2Et2O

cat. CF3CO2HEtOH

39

385857

56 55 54

53

52

51

Scheme 8

23

Page 35: Synthesis of Some Naturally Occurring Quinones

Treatment of diethyl 3-pyrrolidinylglutarate 52 with two equivalents of the

acetylacetone dianion generated the polyketide intermediate 53, which underwent

cyclisation to give the naphthyl diketone 54. Acid-catalysed cyclisation then yielded the

naphthopyran 55. Hydrogenation, followed by monomethylation of 56, by treatment

with diazomethane in the absence of light, gave a 9 : 1 mixture of cis and trans isomers

57 and 58. Finally, oxidation with Fremy’s salt afforded mainly (±)-eleutherin 38, with

(±)-isoeleutherin 39 as the minor product.

A series of three related syntheses of (±)-eleutherin 38 and (±)-isoeleutherin 39

has also been reported. All of them involve construction of the pyran rings of 38 and 39

by way of an intramolecular cyclisation of a key alcohol intermediate 62. The first of

these syntheses was carried out by Kometani and coworkers (Scheme 9).68 Allylation of

2-bromo-8-methoxy-1,4-naphthoquinone 59 with vinylacetic acid, followed by

reductive methylation gave the bromonaphthalene 61. Lithiation, followed by treatment

with acetaldehyde then furnished the key alcohol 62, which participated in an

intramolecular oxymercuration to form the pyran ring, yielding a 1 : 0.9 mixture of cis

and trans isomers 65. The isomers were separated and individually subjected to

oxidative demethylation with ceric ammonium nitrate, giving (±)-eleutherin 38 and (±)-

isoeleutherin 39.

24

Page 36: Synthesis of Some Naturally Occurring Quinones

O

O

MeO

Br CO2H

MeO O

O

Me

Me

O

SnMe3O

O

MeO

Me

OBF3.Et2OCH2Cl2

O

O

MeO

Br

MeO MeO

MeO

Me

O

MeO O

O

Me

Me

O

LiAlH4Et2O

MeO

MeO

O

Me

Me

MeO

MeO MeO

Br

MeO

MeO MeO

MeO

Me

OH

AgNO3, aq. MeCNH2O, (NH4)2S2O8

1. aq. NaHSO3 Et2O

2. KOH, Me2SO4

BuLi, THFMeCHO

1.

2. MeI, K2CO3 Me2CO

1. Hg(OAc)2 THF, H2O2. 3M NaOH NaBH4

CANaq. MeCN

653938

626463

616059

Scheme 9

Uno and coworkers adapted this sequence with an alternative method for

preparing the key alcohol 62 (Scheme 9).69, 70 Nucleophilic addition of

allyltrimethylstannane to the quinone 63, in the presence of the Lewis acid BF3.Et2O,

followed by methylation gave the trimethoxynaphthalene derivative 64. Reduction of 64

then provided the alcohol 62, which was subjected to the operations outlined above to

give (±)-eleutherin 38 and (±)-isoeleutherin 39.

25

Page 37: Synthesis of Some Naturally Occurring Quinones

The final synthesis employing the key alcohol intermediate 62 was carried out

by Giles and coworkers,71, 72 who developed a sequence allowing the selective

preparation of (±)-isoeleutherin 39 (Scheme 10). A stereospecific base-induced

intramolecular cyclisation of the alcohol 62 by brief treatment with potassium tert-

butoxide under anaerobic conditions gave the trans-pyranonaphthalene 66, with only a

trace amount of the cis isomer being produced. Finally, oxidation of 66 gave (±)-

isoeleutherin 39.

MeO MeO

MeO

Me

OH

MeO MeO

MeO

O

Me

Me

MeO O

O

Me

Me

O

39

KOBut

DMFCANaq. MeCN

6662

Scheme 10

Giles and coworkers have accounted for the completely stereoselective nature of

the cyclisation with a mechanistic rationale for a similar system in which a related

alcohol derivative 68 was converted to the trans-pyranonaphthalene 73 (Scheme 11).73

It was noted that a short reaction time (ca 5 min) resulted in the sole formation of 73 in

almost quantitative yield, but that longer reaction times led to the formation of

increasing amounts of the cis isomer 71. After 1.25 h, the trans-pyranonaphthalene 73

and the cis-pyranonaphthalene 71 were formed in 53% and 40% yields respectively.

26

Page 38: Synthesis of Some Naturally Occurring Quinones

OMe

OMe

OH

Me OMe

OMe

O

Me

OMe

OMe

OH

Me

O

Me

Me

OMe

OMe

O

Me

Me

OMe

OMe

O

Me

Me

OMe

OMe

O

Me

Me

OMe

OMe

73

KOBut, DMF

60ºC

KOBut, DMF60ºC

7172

706968

67

Scheme 11

Mechanistically it has been suggested that the reaction commences with deprotonation

of the alcohol 68 and isomerisation of the allyl substituent under the strongly basic

27

Page 39: Synthesis of Some Naturally Occurring Quinones

conditions to give the 2-propenyl intermediate 69. The resultant alkoxide then adds to

the double bond of the alkenyl moiety so that the trans-dimethylpyran anion 72 is

formed. It has been speculated that the formation of the trans compound is favoured

kinetically because the methyl group at C3 adopts the less crowded equatorial position,

avoiding a 1,3 diaxial interaction with the methyl group at C1, which itself adopts a

pseudoaxial configuration thereby avoiding a peri interaction with the methoxy

substituent on the naphthalene ring. Protonation of the anion 72 then delivers the trans-

dimethylpyran derivative 73. Evidence for the initial isomerisation of 68 to 69 comes

from the observation that the alcohol 67 can be substituted for 68 in the above reaction

and also results in the formation of the trans-dimethylpyran derivative 73. It has been

suggested that the cis-dimethylpyranonaphthalene 71 is the thermodynamically

preferred product and that the conversion of the trans-dimethylpyranonaphthalene 73 to

the cis isomer after longer reaction times arises through deprotonation at C4 of the

pyran ring of 73, reforming the anion 72 which undergoes ring opening to give the

alkoxide 69. This can then cyclise to form the cis-pyranonaphthalene anion 70.

Protonation of 70 would then form the cis-pyranonaphthalene 71. The cyclisation of the

alcohol 62 to give the naphthopyran precursor 66 to isoeleutherin 39 presumably occurs

in a similar fashion to that described above and a mechanism for the transformation is

outlined in Scheme 12.

28

Page 40: Synthesis of Some Naturally Occurring Quinones

MeO

MeO

OH

MeMeO

O

Me

Me

MeO

MeO

MeO

MeO

MeO

O

MeMeO

O

Me

Me

MeO

MeO

MeO

O

Me

Me

MeO

MeO

MeO

O

Me

Me

MeO

MeO

MeO

66

KOBut, DMF

60ºC, 15 min

trace

87%

7677

757462

Scheme 12

The addition of alcohol 62 to the unactivated double bond of the alkenyl

substituent to form the pyran ring of 66 is unusual (Scheme 12). Giles and coworkers

have suggested that this novel reaction might be driven by steric effects.74 Studies on

related systems have revealed that such a transformation will only occur if there are two

methoxy substituents at C1 and C4 of the naphthalene moiety, and it has been argued

that the two methoxy groups flanking the hydroxymethyl and alkenyl substituents may

be forcing the two reaction centres together, thereby facilitating the addition to form a

pyran ring.

29

Page 41: Synthesis of Some Naturally Occurring Quinones

A different approach to (±)-eleutherin 38 and (±)-isoeleutherin 39 has been

described by Kobayashi and coworkers,64, 75 who generated the pyran ring of the key

precursor 83 in a single pot via a tandem conjugate addition-cyclisation sequence

(Scheme 13).

O

O

N

Me

iPr

H

Me OHMeO

MeO OH

O

Me

OH

Me

NH

iPr

MeO O

O

Me

Me

O

MeO O

O

Me

Me

O

Et3SiHTFA

N

Me

iPrH

O

O

Me

O

Me

NH

iPr

MeO

MeO O

O

Me

Me

O

O

O

MeO

OH

Me

39

air

NH2iPr

3883

82 81

8078 79

Scheme 13

It has been suggested that the sequence commences with the addition of the enamine 79

to the naphthoquinone 78. Intramolecular cyclisation of 80 and subsequent oxidation,

upon exposure to the air, then gave the naphthoquinone 82. Elimination of the amine,

30

Page 42: Synthesis of Some Naturally Occurring Quinones

during workup and purification on silica gel, provided the key precursor

pyranonaphthoquinone 83. Finally, reduction at room temperature gave a 1 : 5 mixture

of (±)-eleutherin 38 and (±)-isoeleutherin 39. The authors concluded that (±)-eleutherin

38 was the initially formed product, which isomerised to 39 under the reaction

conditions. Repeating the reaction at –20 ºC suppressed the isomerisation resulting in

the selective formation of (±)-eleutherin 38.

Recently, Tewierik and coworkers reported the first enantioselective synthesis of

eleutherin 38 (Scheme 14).76

O

OH O

Me

O

O Me

Me

O

MeO

O

O Me

Me

O

MeONEt3CH2Cl2 O

O Me

Me

O

MeOBr

O

OH Me

Me

MeO

O

OH Me

Me

Br

Br

O

O Me

Me

Br

O

1. MeLi, THF or MeMgBr, ether

2. Et3SiH CF3CO2H CH2Cl2

2 eq. NBSDMF

CANMeCN, H2O

benzene, 80 ºC

150 ºC CH2=CH2

84

89 88 87

8685

(1R, 3S)-38

Scheme 14

31

Page 43: Synthesis of Some Naturally Occurring Quinones

Unlike most of the previous syntheses of (±)-eleutherin 38 and (±)-isoeleutherin 39

described above where the key step involves construction of the pyran rings, this

synthesis takes advantage of the pyran ring already present in the starting lactone (S)-

mellein 84. An important part of the sequence involves formation of the

naphthoquinone moiety of eleutherin 38 via a Diels-Alder cycloaddition. (S)-Mellein 84

is itself a natural product and was prepared following a six step sequence from (S)-

propylene oxide.77 Methylation of 84, followed by stereospecific reduction with

triethylsilane and trifluoroacetic acid yielded the 1,3-dimethylpyran 85. Bromination of

85, followed by oxidaton with ceric ammonium nitrate then generated the

bromonaphthoquinone 87. A Diels-Alder addition of 87 and 1-methoxy-1,3-

cyclohexadiene, followed by treatment with triethylamine and heating to 150 ºC

afforded (1R, 3S)-eleutherin 38 in low yield (13%). This final sequence presumably

occurs by way of dehydrobromination of the Diels-Alder adduct 88 upon treatment with

base, followed by a pyrolysis-induced loss of ethylene to effect aromatisation and

deliver 38.

This recent synthesis has been followed closely by a second enantioselective

route to (1R, 3S)-eleutherin 38, which was devised by Gibson, Andrey and Brimble and

incorporates a key Hauser-Kraus annulation strategy for constructing the naphthopyran

ring of 38 (Scheme 15).78

32

Page 44: Synthesis of Some Naturally Occurring Quinones

O

O

CN

MeO

O

MeO

Me

HO MeMeO

MeO

MeO

MeO

Me

O

OTBDPS

Me

MeO

O

Me

Me

OTBDPS

O

O Me

Me

O

MeO

O

MeO Me

Me

MeO

MeO

OH

OH

Me

O

OTBDPS

Me

MeO

O

O

Me

O

OTBDPS

Me

MeO

(1R, 3S)-38

2. Na2S2O4 tetra-n-butylammonium bromide, H2O, THF; then NaOH, Me2SO4

1. t-BuOK DMSO

90

94 93

9291

95

Et3SiH CF3CO2H CH2Cl2-78 ºC to RT

tetra-n-butylammoniumfluoride, THF

96

CANMeCN/H2O

Scheme 15

The annulation reaction of the known 3-cyanophthalide 90 with a chiral Michael

acceptor 91, followed by a one pot reduction/methylation of the crude reaction product

33

Page 45: Synthesis of Some Naturally Occurring Quinones

gave the naphthalene 94 in 67% yield. Deprotection of the tert-butyldiphenylsilyl ether

with tetra-n-butylammonium fluoride and subsequent intramolecular addition of the

liberated hydroxy group at the ketone moiety then afforded the cyclic hemiacetal 95. A

stereospecific reduction of 95 with triethylsilane and trifluoroacetic acid provided (1R,

3S)-dimethyl benzo[g]isochromene 96, which was converted into the target (1R, 3S)-

eleutherin 38 in excellent yield via an oxidative demethylation reaction with ceric

ammonium nitrate.

The work outlined above illustrates some of the chemistry required to elaborate

the naphthoquinone and dimethyl-substituted pyran moieties of eleutherin 38 and

isoeleutherin 39. Some of this chemistry is relevant to the proposed synthesis of

elecanacin 36, which is one of the themes of this thesis.

2.1.3 Elecanacin 36 as a Synthetic Target

Unlike eleutherin 38 and isoeleutherin 39 which belong to the large family of

pyranonaphthoquinone antibiotics, the structural isomer elecanacin 36 has a unique and

previously unreported tetracyclic ring skeleton. Given this unusual structure, as well as

the continuing interest in naphthoquinones isolated from the Iridaceae Eleutherine

genus, it seemed of interest to develop an approach to the synthesis of elecanacin 36 in

order to confirm the structure of the novel ring skeleton, and to possibly determine its

absolute stereochemistry. The construction of the cyclobutane ring would be a key step

in this synthesis and it was envisaged that this might be possible by way of a [2 + 2]

photocycloaddition.

34

Page 46: Synthesis of Some Naturally Occurring Quinones

2.2 [2 + 2] Photocycloadditions for the Synthesis of Cyclobutane Rings

While there is a variety of methods available for the preparation of cyclobutane

derivatives,79 one of the most common methods for accessing the strained four-

membered ring is by way of a light-induced [2 + 2] cycloaddition between an excited

state enone and a ground state alkene. This reaction has been the subject of many

detailed reviews.80-85

The [2 + 2] photoaddition was discovered by Ciamician and Silber who reported

in 1908 that solutions of carvone 97 left in Italian sunlight for 1 year led to the

formation of an adduct formulated as structure 98 (Scheme 16).86 The resultant

cyclobutane derivative 98 was named carvonecamphor and arises by the intramolecular

cycloaddition of the carbon-carbon double bonds of the cyclohexenone moiety to the

tethered alkenyl side-chain within 97.

MeO

Me

OMe

O

Me

Me

hv (sunlight)1 yr

97 98

Me

Scheme 16

Ciamician and Silber’s observation gained little attention until the result was confirmed

in 1957 by Büchi and coworkers, who repeated the reaction by irradiating a solution of

carvone 97 with Californian sunlight for 6.5 months.87 The structure of 98 was

confirmed using a combination of IR and UV spectroscopy, as well as chemical

degradation and derivatisation techniques.

35

Page 47: Synthesis of Some Naturally Occurring Quinones

Since Büchi’s work in the late 1950s, the reaction has become an important part

of the organic synthetic methodology and has been a key step in the synthesis of a

number of cyclobutane-containing natural products,88 including caryophyllene 10289

and β-panasinene 10590, as well as some theoretically interesting compounds such as

cubane 109 (Scheme 17).91, 92

Br

O

OBr

Br

OBr

OO

H

HMe

MeMe

Me

Me

Me

O

MeMe

OMe

Me

O

Me

Me

Me

OMe

Me

Me

O

O

Br

Br

O Br

Br

O

O

COOH

HOOC

hvpentane

hv, MeOH, HCl

steps

steps

hvbenzene

1. (CH2OH)2 H+

2. aq. HCl

steps

steps50% aq. KOH

99 100 101 102

105104103

106 107 108

109111110

hvpentane

Scheme 17

36

Page 48: Synthesis of Some Naturally Occurring Quinones

The [2 + 2] photoaddition has also become increasingly important for the

preparation of useful synthetic intermediates because the strained cyclobutane ring can

be induced to undergo a number of transformations. These include ring-expansion for

the formation of five- to nine-membered rings, as well as ring-contraction for

cyclopropane rings and ring-opening for acyclic systems.93

2.2.1 Mechanism of the Enone-Alkene [2 + 2] Photocycloaddition

While many details about the mechanism of the [2 + 2] photocycloaddition are

not yet understood,80 the triplet mechanism based on work by Corey94 and de Mayo95, 96

for the addition of an excited state enone to a ground state alkene is now well accepted,

and this can be explained with the help of the figure set out below (Figure 12).

Irradiation results in excitation of the enone 99 to the singlet state S1 (n, π*) 113, which

is short-lived in enone systems. In the case of acyclic enones or large flexible

cycloalkenones (with seven or more members), the energy provided allows them to

undergo cis-trans isomerisation, and the substrate quickly reverts to the ground state

without further reaction. For this reason the [2 + 2] photoaddition is most commonly

carried out on five- or six-membered cycloalkenones and other conformationally

constrained systems because they will not undergo energy-wasting isomerisation around

the carbon-carbon double bond when subjected to irradiation.80-83 However, it should be

noted that this condition is less important for the intramolecular version of the reaction,

where trapping is particularly efficient due to the proximity of the excited state enone

and ground state alkene.80

37

Page 49: Synthesis of Some Naturally Occurring Quinones

O

O O

O O

O

S1 T1

O *

exciplex

intersystemcrossing

hv

and/or

99 112

113 112 112114 115

116 117

118

Figure 12. Mechanism for the enone-alkene [2 + 2] photocycloaddition, illustrated by

the addition of cyclohexenone 99 and ethylene 112 (adapted from a similar scheme by

Horspool)83

In the case of these smaller ring cycloalkenones, the singlet state 113 may

undergo intersystem crossing to the triplet state T1 (n, π* or π, π*) 114 (Figure 12),

which is a fast and efficient process for enones, and this triplet state is sufficiently long

lived for a reaction to occur. 80-84 According to this theory, the enone in its triplet state

forms a complex with the ground state alkene 112, known as an exciplex 115. The

exciplex 115 can decay back to starting materials or proceed to form a triplet 1,4-

diradical intermediate 116 and/or 117. The diradical intermediate then either fragments

38

Page 50: Synthesis of Some Naturally Occurring Quinones

back to starting materials or undergoes spin inversion to a singlet diradical which can

close to form a cyclobutane adduct 118.80, 82, 83 It should be noted that Schuster and

coworkers have presented evidence that the diradical intermediate may be formed

directly without the need for an exciplex precursor.97

2.2.2 Intermolecular [2 + 2] Photocycloadditions

The intermolecular [2 + 2] photoaddition is often of limited use in synthesis

because there are many systems where the regioselectivity of the reaction is

unpredictable.81, 82 When unsymmetrical enones and alkenes undergo photoaddition,

two possible regioisomers may form: the head-to-head and the head-to-tail adduct. The

regiochemical outcome of a reaction tends to be dependent upon the electronic and

steric character of the particular enone and alkene substrates, as well as other factors

such as the temperature and the solvent used in the reaction.80, 82, 84 There are also many

reports of intermolecular photoadditions where the regioselectivity is low and mixtures

of adducts arise.80 One example comes from White and Gupta’s synthesis of the

naturally occurring compound β-bourbonene 123,98, 99 where the key [2 + 2]

photoaddition of cyclopentenone 119 to 2-methyl-5-isopropylcyclopentene 120

proceeded in a non-regioselective fashion, affording a 1 : 1 mixture of two adducts 121

and 122 in 66% yield (Scheme 18).

39

Page 51: Synthesis of Some Naturally Occurring Quinones

O

Me

MeMe

H

H

H

Me

MeMe

O Me

H

H

H

O

MeMe

H

H

H

Me

MeMe

hvpentane

Ph3P=CH2,ether

119 120 121 122

123

Scheme 18

The two regioisomers were separated by chromatography and the sequence was

completed by treatment of the desired head-to-head isomer 121 with

methylenetriphenylphosphorane affording β-bourbonene 123.

2.2.3 Intramolecular [2 + 2] Photocycloadditions

The intramolecular [2 + 2] photoaddition is often more useful than the

intermolecular reaction because the former offers better regiochemical control due to

the constrained nature of such systems.80, 82, 84 These reactions also tend to be more

efficient due to the proximity of the excited state enone subunit and the tethered ground

state alkene.80, 84

40

Page 52: Synthesis of Some Naturally Occurring Quinones

For intramolecular [2 + 2] photoadditions, there is a preference for the formation

of adducts with the cyclobutane ring fused to a five-membered ring, rather than to a

four- or six-membered ring, because five-membered rings form more rapidly.80, 82, 84

This phenomenon has been termed the ‘Rule of Five’ and can be illustrated by Pirrung’s

synthesis of the naturally occurring sesquiterpene (±)-isocumene 126 (Scheme 19),100,

101 where an intramolecular enone-alkene [2 + 2] photoaddition within the

cyclohexenone derivative 124 gave the required cyclopentyl-fused intermediate 125 in

77% yield.

O

Me

Me

MeO

Me

Me

Me

Me

Me

Me

Me

stepshv, 350 nmhexane

124 125 126

Scheme 19

The stereoselectivity of the above photoaddition is also notable and results in the

formation of a single diastereomer. This arises by addition of the terminal alkene to one

of the diastereotopic faces of the cyclohexenone carbon-carbon double bond opposite to

the C4 methyl substituent within 124, thereby avoiding destabilizing steric interactions

(Scheme 19).

2.2.4 [2 + 2] Photocycloadditions of 1,4-Naphthoquinones

1,4-Naphthoquinones are amongst the variety of unsaturated carbonyl

compounds which are known to undergo [2 + 2] photocycloaddition to alkenes, making

them potentially useful substrates for the synthesis of cyclobuta-fused ring systems. The

41

Page 53: Synthesis of Some Naturally Occurring Quinones

intermolecular version of the reaction in particular has attracted some attention and

photoadditions between a variety of alkene and 1,4-naphthoquinone derivatives have

been reported, including the reactions shown in Scheme 20.

O

O

Cl

O

O

Cl

O

O

O

O

O

O

O

OOEt

O

O

OEt

OEt

hv benzene

hv benzene

hv benzene

127 112 128

4 112 130

4 131 132 133

Scheme 20

Naito and coworkers have found that irradiation of 2-chloro-1,4-naphthoquinone 127 in

the presence of ethylene 112 gives rise to the cyclobuta-fused dione 128 in 53%

yield.102 Some further examples come from an investigation carried out by Bryce-Smith

and coworkers who observed that 1,4-naphthoquinone 4 undergoes a similar

photoaddition to ethylene 112 to afford the cyclobutane derivative 130 (28%).103

However, in the case of the photoaddition of 1,4-naphthoquinone 4 to ethyl vinyl ether

131 both a cyclobutane 132 and a spirooxetane adduct 133 were generated in low yield

42

Page 54: Synthesis of Some Naturally Occurring Quinones

(8% total chemical yield). The spirooxetane 133 in this final example arises by way of a

Paterno-Büchi reaction, which involves the light-induced cycloaddition of a carbonyl

carbon-oxygen double bond to an alkene to form a four-membered oxetane ring104 and

is able to compete with cyclobutane formation in some systems.103, 105, 106

Examples of intramolecular [2 + 2] photoadditions of 1,4-naphthoquinones are

rare, but one reaction has been reported by Suzuki and coworkers who found that the

norbornadiene-fused naphthoquinone 134 underwent ready intramolecular cycloaddition

upon irradiation with visible light generating the quadricyclane derivative 135 in 63%

yield (Scheme 21).107 Interestingly, the quadricyclane product absorbs light near the

visible region and 135 was found to undergo a cycloreversion reaction when irradiated

at shorter wavelengths (≤ 370 nm) resulting in a 79 : 21 mixture of the norbornadiene

134 and quadricyclane 135 respectively.

O

O

O

O

hv (> 410 nm) dichloromethane

hv (< 370 nm)

134 135

Scheme 21

43

Page 55: Synthesis of Some Naturally Occurring Quinones

2.3 An Approach to Elecanacin 36

The ability of 1,4-naphthoquinones to undergo light-induced [2 + 2]

cycloaddition and form cyclobutane rings suggested that such a photoaddition may

provide a useful approach to the cyclobuta-fused ring skeleton of elecanacin 36.

Although Suzuki and coworkers’ synthesis of the quadricyclane-fused naphthoquinone

135 appears to be the only reported example of an intramolecular [2 + 2]

photocycloaddition involving a 1,4-naphthoquinone (Scheme 21), intramolecular

photoadditions in cyclohexenone derivatives bearing an alkenyl-substituted side chain

are well known.108 One example is the intramolecular photoaddition reported by Cargill

and coworkers who found that irradiation of 136 in dichloromethane gave the adduct

137 in 83% yield (Scheme 22).109

O O

CH2Cl2

hv

136 137

Scheme 22

On this basis, it was expected that a similar photoaddition could provide access to the

cyclobuta[1,2-b]tetrahydrofuran moiety in elecanacin 36. Leaving aside for the moment

the question of the configuration of the methyl-substituted carbon of the tetrahydrofuran

ring, retrosynthetic cleavage of two cyclobutane bonds in 36 suggested 138 as a

potential key intermediate (Scheme 23).

44

Page 56: Synthesis of Some Naturally Occurring Quinones

O

O O

OMe

Me

O

O

O

MeO

Me

MeO O

O

OH

Me

O

O O

OMe

Me

36

13846

Scheme 23

Considering the synthetic direction, an intramolecular photoaddition of the vinyl ether

moiety to the double bond within the naphthoquinone 138 should in principle give

elecanacin 36, and would allow construction of the cyclobutane and tetrahydrofuran

rings to be achieved in a single reaction. The vinyl ether 138 should in turn be available

from the hydroxy-quinone 46. Although the intermolecular addition of 1,4-

naphthoquinone 4 with ethyl vinyl ether 131 proceeds in low yield (Scheme 20, p. 42),

it was expected that this related intramolecular version should be more efficient given

the proximity of the reaction centres within 138, which should result in a more

favourable entropy of activation.

In order to test the feasibility of the proposed intramolecular [2 + 2]

photoaddition, the synthesis of the racemic vinyl ether 138 was undertaken and

hydroxy-quinone 46 was set as the initial target.

45

Page 57: Synthesis of Some Naturally Occurring Quinones

2.4 Preparation of (±)-Elecanacin 36

2.4.1 Synthesis of 2-(2-Hydroxypropyl)-5-methoxynaphthalene-1,4-dione 46

There have been two previous syntheses of 2-(2-hydroxypropyl)-5-

methoxynaphthalene-1,4-dione, one yielding the racemate 46 and the other giving one

enantiomer 146. However, it was felt that neither synthesis was particularly

straightforward.

As already mentioned, Eisenhuth and Schmid prepared 46 as an intermediate in

their synthesis of eleutherin 38 and isoeleutherin 39 (Scheme 6, p. 21).59 Although they

found that the hydroxy-quinone 46 could be obtained following a route involving an

oxidative ring-opening of the naphthofuran 45 by treatment with ferric chloride in

aqueous acetone (Scheme 6), it was noted that very exact experimental manipulations

were required or the reaction would result in the formation of polymeric material.

A more recent synthesis of the (S)-hydroxy-quinone 146 has been reported by

Tanada and Mori who found that 146 arose unexpectedly during oxidation of

naphthofuran 144 with Fremy’s salt (Scheme 24).110 However, this result was just

mentioned in passing and no experimental detail or yield was given. Although this route

could conceivably be shortened to give the hydroxy-quinone 146 at an earlier stage by

possible oxidation of intermediate 143, the use of tetralone 139 as the starting material

also makes the sequence undesirably expensive (current catalogue price of 139 :

$28/g).111

46

Page 58: Synthesis of Some Naturally Occurring Quinones

O

MeOO

Me

O

Me

OHMeO

O

Me

MeO O

O

ON(SO3K)2 ON(SO3K)2

MeO O

O

OH

Me

O

MeO

Me

OH

OH

MeO

Me

OH

OH

O

MeO

Me

OTroc

O

MeO

Me

OTroc

O

1. LHMDS, toluene

2. 10% Sc(OTf)3

Cl3CCH2OC(O)ClpyridineCH2Cl2

SeO21,4-dioxane

powdered ZnAcOH

DEAD, PPh3THF

139 140 141

142143144

146145

H H

HH

H

Scheme 24

Thus, rather than repeating either of these syntheses, it was envisaged that the

hydroxy-quinone 46 might be prepared more conveniently following a route involving

hydration of the terminal double bond in the allyl-substituted naphthalene 147 via an

oxymercuration-reduction (Scheme 25).

47

Page 59: Synthesis of Some Naturally Occurring Quinones

MeO

OH

Br

MeO O

O

OH

Me

ON(SO3K)2

MeO

O

MeO

OH

OH

Me

K2CO3MeOH

MeO

OAc

MeO

OAc

OH

Me

K2CO3, Me2CO

Ac2O, PhNEt2170 ºC

1. Hg(OAc)2, THF2. NaBH4, 2M NaOH

41 42 147

14814946

Scheme 25

Allylation of 5-methoxynaphthalen-1-ol 41, followed by a Claisen rearrangement and in

situ acetylation of 42 afforded the acetate 147 in 95% yield. While it was appreciated

that the acetate group was unlikely to be robust enough to survive the addition of dilute

base during the oxymercuration-reduction, any hydrolysis of 148 would be of no

consequence because the next step of the sequence would involve removal of the acetyl

group. Unfortunately treatment of 147 with mercuric acetate in aqueous tetrahydrofuran

according to the general method of Brown and Geoghegan,112 followed by sodium

borohydride reduction was unsuccessful and returned a 1 : 1 mixture of starting acetate

147 and hydrolysed starting material. There was no trace of the desired alcohol 148.

The oxymercuration of an alternative allyl-substituted precursor 44 to the key

hydroxy-quinone 46 was also briefly examined (Scheme 26).

48

Page 60: Synthesis of Some Naturally Occurring Quinones

MeO

O

MeO

OH

ON(SO3K)2

MeO O

O

MeO O

O

OH

Me

1. Hg(OAc)2, THF2. NaBH4, 2M NaOH

160-185 ºC

42 43 44

46

Scheme 26

The quinone 44 was prepared according to the procedure of Eisenhuth and Schmid.59 A

Claisen rearrangement of the allyl ether 42 provided the naphthol 43, which underwent

oxidation upon treatment with Fremy’s salt affording 44 in 91% yield. Attempted

oxymeruration of this compound was also unsuccessful and returned a complex mixture

of products which could not be identified.

As preparation of the hydroxy-quinone 46 by way of an oxymercuration reaction

could not be achieved, an alternative approach was investigated and is outlined in

Scheme 27. It begins with the previously prepared acetate 147.

49

Page 61: Synthesis of Some Naturally Occurring Quinones

MeO

OAc

MeO

OAc

H

O

MeO O

O

OH

Me

MeO

OH

OH

Me

ON(SO3K)2

2. (NH2)2CS NaHCO3

1. O3, CH2Cl2/MeOH -78 ºC

147 150

14946

MeMgITHF

Scheme 27

Careful ozonolysis of the acetate 147 followed by thiourea reduction gave the aldehyde

150. The reaction mixture did not turn blue to signify the consumption of the alkene and

it was necessary to monitor the disappearance of the starting material by TLC in order

to avoid over-oxidation of the electron-rich aromatic ring system. The aldehyde 150

proved to be unstable and decomposed on silica during attempted purification. The 1H

NMR spectrum of the crude aldehyde 150 includes a triplet at 9.69 ppm, consistent with

an aldehyde proton and this is coupled to the methylene protons in the formylmethyl

sidechain, which resonate as a doublet at 3.70 ppm. The spectrum also shows two

singlet signals at 4.00 ppm and 2.46 ppm, corresponding to the C5 methoxy group and

C1 acetate methyl group protons respectively.

The problem of the instability of aldehyde 150 was overcome by using the crude

compound directly in the subsequent Grignard reaction (Scheme 27). Immediate

50

Page 62: Synthesis of Some Naturally Occurring Quinones

treatment of 150 with an excess of methylmagnesium iodide gave the required alcohol

149. The molecular formula C14H16O3 of 149 was confirmed by combustion analysis

and mass spectrometry. The 1H NMR spectrum contains signals consistent with the

presence of the hydroxypropyl side chain, including a C3 methyl doublet at 1.27 ppm

and two doublet of doublets arising from the C1 methylene protons at 3.01 ppm and

2.90 ppm (Figure 13). The C2 methine proton appears as a multiplet at 4.36-4.27 ppm.

Figure 13. 1H NMR spectrum of 149 (300 MHz, CDCl3).

Finally, oxidation of 149 with Fremy’s Salt afforded the desired hydroxy-

quinone 46 as a bright yellow crystalline solid in 41% yield over three steps (Scheme

27). In the 1H NMR spectrum of 46, the C3 proton resonates as a triplet at 6.78 ppm due

to long range coupling to the methylene protons in the hydroxypropyl side chain, which

themselves occur as two doublet of doublet of doublets at 2.74 ppm and 2.59 ppm

51

Page 63: Synthesis of Some Naturally Occurring Quinones

(Figure 14). The methylene protons are also coupled to the adjacent methine proton,

which occurs as a multiplet at 4.15-4.03 ppm. A doublet at 1.29 ppm is associated with

the methyl protons in the hydroxypropyl side chain. The 13C NMR spectrum includes

two signals at 186.1 ppm and 184.3 ppm, corresponding to the quinonoid carbonyl

carbons. The electronic spectrum shows absorbances at 247, 268, 354 and 396 nm,

which are consistent with the presence of a naphthoquinone chromophore.113

Figure 14. 1H NMR spectrum of 46 (300 MHz, CDCl3).

2.4.2 Synthesis of 5-Methoxy-2-(2-vinyloxypropyl)naphthalene-1,4-dione 138

With the hydroxy-quinone 46 in hand, the next step involved conversion of 46 to

the target vinyl ether 138. This was achieved in the usual fashion by heating the

52

Page 64: Synthesis of Some Naturally Occurring Quinones

hydroxy-quinone 46, ethyl vinyl ether and a catalytic amount of mercuric acetate under

reflux,114 which delivered 138 as a bright yellow oil (Scheme 28).

O

O

MeO

O

Me

MeO O

O

OH

Me

EtO

cat. Hg(OAc)2

13846

Scheme 28

The NMR spectral properties are in accord with the structure of the vinyl ether 138

(Figure 15). The 1H NMR spectrum shows a signal at 6.75 ppm due to the C3 proton,

which is split into a triplet due to long range coupling to the methylene protons in the

vinyloxypropyl side chain. These methylene protons occur as two doublet of doublets of

doublets at 2.79 ppm and 2.67 ppm. The methylene protons are also coupled to the

adjacent methine, which itself resonates as a multiplet at 4.26-4.15 ppm. In addition, the

spectrum shows a vinylic pattern with components centred at 6.27 ppm, 4.30 ppm and

3.99 ppm. This last signal overlaps with the C5 methoxy group singlet signal at 3.99

ppm.

53

Page 65: Synthesis of Some Naturally Occurring Quinones

Figure 15. 1H NMR spectrum of 138 (300 MHz, CDCl3).

2.4.3 Synthesis of (±)-Elecanacin 36

Having prepared the key vinyl ether 138, efforts were now directed towards

construction of the remaining cyclobutane and tetrahydrofuran rings in elecanacin 36

via the proposed [2 + 2] photocycloaddition. Irradiation at 350 nm of a 0.009 M solution

of vinyl ether 138 in dichloromethane cleanly gave two compounds, with very close Rf

values. An electronic spectrum of the reaction solution measured after irradiation shows

the disappearance of the characteristic long wavelength absorbance attributable to the

naphthoquinone chromophore,113 which occurs at 393 nm in the electronic spectrum of

138 (Figure 16).

54

Page 66: Synthesis of Some Naturally Occurring Quinones

Figure 16. Electronic spectrum of 5-methoxy-2-(2-vinyloxypropyl)naphthalene-1,4-

dione 138 (solid line) and the photoaddition reaction product (broken line) in

dichloromethane.

O

O

MeO

O

Me

O

O O

H

MeO

Me

O

O

MeO

O

H

Me

CH2Cl2hv, 350 nm

(±)-elecanacin

15136138(±)-isoelecanacin

(one enantiomer of each product is depicted)

Scheme 29

55

Page 67: Synthesis of Some Naturally Occurring Quinones

The two compounds were separated by careful radial chromatography. The slightly

more polar compound, isolated in 25% yield, was (±)-elecanacin 36 (Scheme 29). The

1H and 13C NMR spectra were identical to those of the naturally occurring (+)-

enantiomer (provided by Prof. Yasuhiro Imakura of the Naruto University of Education

in Japan)47 and are shown in Figures 18 and 19. This result confirms the unusual

cyclobuta-fused ring skeleton of elecanacin 36.

The more mobile product, which we named isoelecanacin, was obtained in 38%

yield. The 1H and 13C NMR spectra closely resemble those recorded for elecanacin 36

with similar chemical shifts (Figures 20a and 20b), and isoelecanacin was assigned the

isomeric structure 151 on the basis of 2D-NMR (COSY, HMBC, HSQC and NOESY)

evidence. The NMR spectral data for 36 and 151 are collected in Table 1 and key

NOESY correlations within the oxabicyclo[3.2.0]heptyl framework are shown in Figure

17. It should be noted that the aromatic proton signals of both isomers show second

order characteristics, even at 500 MHz, rather than the first order pattern implied47 for

elecanacin 36.

36

elecanacin isoelecanacin

151

O

HnHx

H

H

MeH

Hn

Hx

O

HnHx

H

Me

HH

Hn

Hx

Figure 17. Key NOESY correlations within the 2-oxabicyclo[3.2.0]heptyl framework

of elecanacin 36 (ref. 47) and isoelecanacin 151 (this work).

56

Page 68: Synthesis of Some Naturally Occurring Quinones

Figure 18a. 1H NMR spectrum of synthetic (±)-elecanacin 36 (300 MHz, CDCl3).

Figure 18b. 1H NMR spectrum of natural (+)-elecanacin 36 (300 MHz, CDCl3).

57

Page 69: Synthesis of Some Naturally Occurring Quinones

Figure 19a. 13C NMR spectrum of synthetic (±)-elecanacin 36 (75.5 MHz, CDCl3).

Figure 19b. 13C NMR spectrum of natural (+)-elecanacin 36 (75.5 MHz, CDCl3).

58

Page 70: Synthesis of Some Naturally Occurring Quinones

Figure 20a. 1H NMR spectrum of isoelecanacin 151 (300 MHz, CDCl3).

Figure 20b. 13C NMR spectrum of isoelecanacin 151 (75.5 MHz, CDCl3).

59

Page 71: Synthesis of Some Naturally Occurring Quinones

60

Page 72: Synthesis of Some Naturally Occurring Quinones

The formation of elecanacin 36 and isoelecanacin 151 is due to the addition of

the vinyl ether moiety to the two diastereotopic faces of the carbon-carbon double bond

of the naphthoquinone system of 138 (Scheme 29). Monitoring the photoaddition by

TLC and NMR revealed that both 36 and 151 were formed at the early stages of the

reaction, and control experiments established that both products were photostable and

not interconverted under the irradiation conditions. Interestingly, the vinyl ether 138

also underwent conversion to elecanacin 36 and isoelecanacin 151 when a solution of

vinyl ether 138 in dichloromethane was left exposed to ambient laboratory light for 5h.

2.5 The Enantioselective Synthesis of Elecanacin 36

Having developed a sequence for the synthesis of racemic elecanacin 36, we

decided to extend this approach to the preparation of the enantiopure compound to

establish the absolute configuration of the natural product. In order to do this, a new

asymmetric route to alcohol 149 depicted in Scheme 27 (p. 50) was needed.

2.5.1 An Approach to Chiral 1-(1-Hydroxy-5-methoxynaphthalen-2-yl)propan-2-

ol 152 using Jacobsen’s Catalyst

Epoxides are valuable compounds for the preparation of alcohols because the

strained three-membered ring undergoes ready ring-opening when attacked by a wide

range of nucleophiles, and they can often be conveniently prepared directly from

alkenes or aldehydes.115 Accordingly, it was decided to investigate a route to the chiral

alcohol 152 involving an epoxide precursor, as outlined in Scheme 30.

61

Page 73: Synthesis of Some Naturally Occurring Quinones

MeO

OH

OH

H

Me

MeO

OR

O

MeO

OR

O

153 R = Ac154 R = Bn

152 155 R = Ac156 R = Bn

Scheme 30

It was expected that reductive opening of a chiral epoxide such as 153 or 154, with

lithium aluminium hydride, and removal of the protecting group would provide one

enantiomer of alcohol 152. The enantiopure epoxides 153 and 154 should in turn be

available by resolution of the corresponding racemic epoxide using Jacobsen’s catalyst.

Hydrolytic kinetic resolution with Jacobsen’s catalyst 160 has proven to be a

useful method for accessing chiral terminal epoxides (Figure 21).

N NCo

O OOAc

HH

But

But

But

But

N NCo

O OOAc

HH

But

But

But

But

(R,R) - 160 (S,S) - 160

Figure 21

In this procedure the chiral (salen)-CoIII catalyst 160 promotes opening of the epoxide

ring of one enantiomer over the other in a racemic mixture of a terminal epoxide by

nucleophilic attack of water.116-118 The process is illustrated below for the resolution of

propylene oxide 157 (Scheme 31).

62

Page 74: Synthesis of Some Naturally Occurring Quinones

O

MeMe

OHHOO

Me

0.5 eq H20,(R,R)-160

157 159158

Scheme 31

The reaction results in a mixture of a 1,2-diol 159 and an enantio-enriched starting

epoxide 158. In view of the large boiling point and polarity differences of these

compounds, the desired chiral epoxide can be readily isolated.

Jacobsen’s hydrolytic kinetic resolution has become increasingly popular for

natural product synthesis as the resolved epoxides are often obtained in very high

enantiomeric excess and yield.118-120 A representative example is the recent synthesis of

epothilone A 165 carried out by Liu and coworkers, who used Jacobsen’s catalyst

methodology to introduce the chiral centre at the C3 position of 165 (Scheme 32).121

O

S

N

OHO O

OH

O

OO

OOOOH

HO

OOH

O

O3

3

steps

0.6 eq. H2O2 mol% Jacobsen's catalyst 160

5 mol% Co2(CO)810 mol% 3-hydroxypyridineTHF/MeOH, CO

epothilone A

161 162 163

164165

Scheme 32

63

Page 75: Synthesis of Some Naturally Occurring Quinones

Treatment of the racemic epoxide 161 with Jacobsen’s catalyst and 0.6 equivalents of

water returned the chiral epoxide 163 in 48.3% yield (maximum 50%) and greater than

99% ee. Ring-opening by carbomethoxylation then afforded the alcohol 164 with the

required stereochemistry.

With these considerations in mind, it was envisaged that hydrolytic kinetic

resolution may provide an effective strategy for obtaining the chiral epoxides 153 and

154. Thus, the first step in the enantioselective approach to elecanacin 36 was to prepare

the corresponding racemic epoxides 155 and 156.

2.5.2 Synthesis of Epoxides 155, 156 and 168 and Attempted Hydrolytic Kinetic

Resolution

An initial attempt to prepare 155 was made by epoxidation of the previously

prepared acetate 147 with m-chloroperoxybenzoic acid (Scheme 33a). The reaction

proceeded sluggishly at room temperature and returned mainly starting material, with

only a trace amount of the desired epoxide 155 forming over three days (1H NMR

analysis). An effort to increase the yield of 155 by gentle heating of the reaction

mixture, in the presence of 2,6-di-t-butyl-4-methylphenol, appeared to be accompanied

by degradation of the electron-rich naphthalene system. This resulted in a complex

mixture of products, including the epoxy acetate 155, which could not be isolated in a

pure state. Similarly, the synthesis of the alternative epoxide 156 was attempted. The

naphthol 43 was protected as a benzyl ether, but subsequent treatment of 166 with m-

chloroperoxybenzoic acid again returned starting material along with an unidentifiable

mixture of products (Scheme 33b).

64

Page 76: Synthesis of Some Naturally Occurring Quinones

MeO

OAc

MeO

OH

MeO

OBn

MeO

OBn

O

MeO

OAc

O

PhCH2BrK2CO3, acetone

m-CPBA, CH2Cl2 0 ºC to RT (or 35 ºC, with2,6-di-t-butyl-4-methylphenol)

a)

b)

m-CPBA, CH2Cl2 0 ºC to RT

156

16643

147 155

Scheme 33

Alternative reagents for preparing the epoxides were investigated. Dubois and

coworkers have suggested that peracetic acid accompanied by a ferric phenanthroline

catalyst can be particularly useful for the epoxidation of terminal alkenes.122 However,

no reaction was observed when 147 was subjected to the reaction conditions (Scheme

34).

65

Page 77: Synthesis of Some Naturally Occurring Quinones

MeO

OAc

MeO

OAc

O[((phen)2(H2O)FeIII)2(µ-O)](ClO4)4

CH3CO3HMeCN, 0 ºC

147 155

Scheme 34

More success was encountered using dimethyldioxirane and the epoxy acetate 155 was

synthesised in low yield (23%) by treatment of 147 with a cold solution of

dimethyldioxirane in acetone, prepared according to the general method of Murray and

Singh (Scheme 35).123

MeO

OAc

O O

Me MeMeO

OAc

O

147 155

Me2CO

Scheme 35

As the monosubstituted ethylene moiety of acetate 147 had proven to be quite

resistant to epoxidation with peroxy reagents, an alternative method for preparing the

related epoxide 156 was considered. Aldehydes can be directly converted into epoxides

with dimethylsulfoxonium methylide, following the general method developed by

Corey and Chaykovsky,124 and the required benzyl ether 156 was prepared satisfactorily

using this approach, as shown in Scheme 36.

66

Page 78: Synthesis of Some Naturally Occurring Quinones

MeO

OBn

MeO

OBn

O

MeO

OBn

H

O

1. O3, CH2Cl2-MeOH -78 ºC

trimethylsulfoxoniumiodideNaH, Me2SO

2. (NH2)2CS NaHCO3

166 167

156

Scheme 36

Careful ozonolysis of 166 gave the aldehyde 167. Formation of 167 was rapid but the

reaction mixture did not turn blue at any stage of the reaction and it was necessary to

carefully monitor the disappearance of starting material by TLC in order to avoid over-

oxidation of the electron-rich aromatic ring. Finally, treatment of 167 with

dimethylsulfoxonium methylide gave the desired terminal epoxide 156 in 69 % yield.

With epoxides 155 and 156 in hand, it was now possible to investigate the

proposed hydrolytic kinetic resolution. Unfortunately, treatment with the (salen)-CoIII

catalyst 160 and water (0.5 equivalents) according to standard methods116, 118 did not

lead to ring-opening and both 155 and 156 were recovered essentially unchanged, even

after long reaction times (Scheme 37).

67

Page 79: Synthesis of Some Naturally Occurring Quinones

MeO

OR

O

155 R = Ac156 R = Bn

0.5 eq H2OJacobsen's Catalyst 160THF, 0 ºC to RT no hydrolytic

kinetic resolution

Scheme 37

While the reason for this lack of reactivity is unclear, it was felt that the electron-rich

naphthalene moieties of 155 and 156 could possibly interfere with the catalytic

hydrolytic cycle. This suggested that a more electron-deficient system such as the

quinonoid epoxide 168 would perhaps be more responsive to the resolution process and

the synthesis of 168 was undertaken (Figure 22).

O

O

MeO

O

168

Figure 22

Perez-Sacau and coworkers have demonstrated that naphthoquinone 170 can be

prepared by epoxidation of the alkenyl side chain of 169 by treatment with m-

chloroperoxybenzoic acid (Scheme 38).125

68

Page 80: Synthesis of Some Naturally Occurring Quinones

O

O

Me

Me

OAc

O

O

OAc

O

Me

Me

m-CPBACH2Cl2, 0 ºC

169 170

Scheme 38

Application of the above procedure to the allyl-substituted naphthoquinone 44

proceeded smoothly and the orange crystalline epoxide 168 was obtained in 67% yield

(Scheme 39). However, attempted hydrolytic kinetic resolution with Jacobsen’s catalyst

was unsuccessful and epoxide 168 was also returned unchanged.

O

O

MeO O

O

MeO

O

m-CPBACH2Cl2, 0 ºC to RT

0.5 eq H2OJacobsen's Catalyst 160THF, 0 ºC to RT

no hydrolytickinetic resolution

44 168

Scheme 39

Control reactions using the same batch of catalyst showed that simple epoxides such as

epichlorohydrin and propylene oxide were readily resolved under these conditions, as

described in the literature.116, 118 Although hydrolytic kinetic resolution has been

successfully applied to a broad range of terminal epoxides,118, 119 Jacobsen and

coworkers have noted that careful optimisation of reaction conditions, including the

69

Page 81: Synthesis of Some Naturally Occurring Quinones

type of solvent, the catalyst loading and the nature of the (salen)-CoIII catalyst

counterion, has been required in order to resolve some particularly unreactive

epoxides.118 While resolution of epoxides 155, 156 and 168 may well be achievable

with further experimentation, it was decided to revise the planned route to the chiral

alcohol 152 and another approach was adopted instead.

2.5.3 A Directed Metallation Approach to (2R)-1-(1-Hydroxy-5-methoxy-2-

naphthalenyl)propan-2-ol 152

A possible alternative approach to chiral alcohol 152 involves a directed

metallation sequence. A methoxymethoxy (MOM) substituent is known to be a stronger

ortho director than a methoxy substituent in the lithiation of aromatic rings.126 For

example, methoxymethoxybenzene 172 has been found to undergo ortho lithiation at a

much greater rate than anisole 171 when treated with n-butyllithium in the presence of

TMEDA (Figure 23).127

OMe O OMe

relativerate

1 14

171 172

Figure 23. Relative rates for ortho lithiation of anisole 171 and

methoxymethoxybenzene 172 (n-BuLi/TMEDA, 0 ºC).127

70

Page 82: Synthesis of Some Naturally Occurring Quinones

Kamikawa and Kubo have employed a route to alcohol 175 which takes advantage of

this greater rate (Scheme 40).128 Lithiation of 1-methoxy-5-

methoxymethoxynaphthalene 173 was found to occur preferentially at the C6 position

ortho to the MOM group rather than at the C2 position ortho to the methoxy group.

Subsequent quenching of 174 with n-decanal then afforded alcohol 175 in 77% yield.

MeO

O OMe

Li

MeO

O OMe

MeO

O OMe

CH(CH)8CH3

OH

CH3(CH2)8CHO

n-BuLi, TMEDATHF

173 174

175

Scheme 40

It was envisaged that a similar strategy involving nucleophilic attack of lithiated

naphthalene 174 on chiral propylene oxide 158, followed by deprotection would deliver

alcohol 152 with the required stereochemistry in the side chain (Scheme 42). An

advantage associated with this route is the ready availability of the starting materials.

Thus, 1-methoxy-5-methoxymethoxynaphthalene 173 was prepared in a straightforward

manner by deprotonation of naphthol 41 with sodium hydride, followed by treatment

with methoxymethyl chloride (Scheme 41).128

71

Page 83: Synthesis of Some Naturally Occurring Quinones

MeO

OH

MeO

O OMe

1. NaH, DMF

2. MOMCl

41 173

Scheme 41

(R)-(+)-propylene oxide 158 was also obtained conveniently by hydrolytic kinetic

resolution of racemic propylene oxide with Jacobsen’s catalyst (Scheme 31, p. 63),116

thereby avoiding the cost normally associated with this expensive reagent. Attention

was then turned to the rest of the sequence (Scheme 42).

MeO

O OMe

MeO

O OMe

H

OH H

Me OH

Me

MeO

OHH

OHMe

OMe

H

1. n-BuLi, THF

CBr42-propanol

MeO

O O

173

2. HMPA

158176 177

152

Me

Scheme 42

Lithiation of 173 in the presence of TMEDA, according to the method of Kamikawa

and Kubo,128 followed by addition of (R)-(+)-propylene oxide 158 returned only starting

material. However, lithiation followed by sequential addition of HMPA and 158 was

72

Page 84: Synthesis of Some Naturally Occurring Quinones

more successful and gave a 80 : 20 mixture of two compounds with very close Rf

values. The compounds could not be separated by silica gel filtration but small

analytical samples were isolated by careful radial chromatography. NMR and mass

spectral analysis indicated that both compounds were trisubstituted naphthalene

isomers, derived from nucleophilic attack on propylene oxide 158. The slightly less

polar component, being the major product, was identified as the target alcohol 176

([α]D21 + 8.7). The 1H NMR spectrum shows signals consistent with the presence of the

hydroxypropyl side chain (Figure 24). An ABX pattern is associated with the C1

methylene (3.10-2.94 ppm) and the C2 methine (4.23 ppm) protons. The methine proton

is further coupled to the adjacent hydroxy and methyl group protons, which resonate as

doublets at 2.32 and 1.30 ppm respectively. With the formation of the stereogenic centre

in 176, the methylene protons in the MOM moiety have become diastereotopic. This is

reflected in the 1H NMR spectrum where the methylene protons occur as an AB pattern

with JAB = 5.9 Hz.

Figure 24. 1H NMR spectrum of 176 (300 MHz, CDCl3).

73

Page 85: Synthesis of Some Naturally Occurring Quinones

The minor product was tentatively assigned structure 177 ([α]D21 –26.9), and arises due

to competitive metallation and alkylation ortho to the methoxy substituent. The 1H

NMR spectrum of 177 is similar to that of 176 (Figure 25). However, the MOM

methylene protons occur as a singlet at 5.39 ppm, despite being diastereotopic. An A2X

pattern is due to the C1 methylene and C2 methine protons in the hydroxypropyl side

chain, which resonate at 2.96 and 4.16 ppm respectively. The C2 methine is further

coupled to the C3 methyl protons, which occur as a doublet at 1.28 ppm. A broad

singlet at 2.29 ppm corresponds to the C2 hydroxy group.

Figure 25. 1H NMR spectrum of 177 (300 MHz, CDCl3).

Finally, removal of the MOM protecting group in 176 was achieved by the

action of carbon tetrabromide in refluxing 2-propanol (Scheme 42),129 and the target

alcohol 152 was easily isolated by chromatography at this stage as a white crystalline

solid with [α]D21 –6.7. The enantiomeric excess (ee) within the resulting alcohol 152

74

Page 86: Synthesis of Some Naturally Occurring Quinones

was determined by 1H NMR spectroscopy with the chiral shift reagent europium tris[3-

(heptafluoropropylhydroxymethylene)-(+)-camphorate]. Surprisingly, the chiral shift

reagent failed to induce significant chemical shift changes for protons near the

stereocentre and the only significant induced shift was observed for the H8 aromatic

proton. Treatment of the racemic alcohol 149 with 7.5% of the shift reagent was found

to separate the H8 doublet signal into two doublets in the 1H NMR spectrum (Figure

26). The enantiomeric ratio within the (2R)-alcohol 152 was estimated to be greater than

95 : 5 by examination of the corresponding H8 signal in the 1H NMR spectrum of a

sample of 152 in the presence of the chiral shift reagent (7.5%), indicating that the ee

was greater than 90%. A more precise upper limit for the ee could not be determined at

this stage due to some line-broadening of this signal.

Figure 26. Stacked plot showing the low-field region of the 1H NMR spectra of (2R,S)-

1-(1-hydroxy-5-methoxynaphthalen-2-yl)propan-2-ol 149 and (2R)-1-(1-hydroxy-5-

methoxynaphthalen-2-yl)propan-2-ol 152 in the presence of 7.5% europium tris[3-

(heptafluoropropylhydroxymethylene)-(+)-camphorate] (500 MHz, CDCl3).

75

Page 87: Synthesis of Some Naturally Occurring Quinones

2.5.4 Final Steps in the Preparation of Enantiopure Elecanacin 36

With the key alcohol 152 in hand, it was now possible to complete the synthesis

of chiral elecanacin 36 following the final steps developed for the racemic series

(Scheme 43).

O

O

MeO

H

O

Me

MeO O

O

OH

HMe

EtO

cat. Hg(OAc)2

O

O O

H

MeO

Me

O

O

MeO

O

H

Me

CH2Cl2

MeO

OH

HMe

OH

hv, 350 nm

(-)-elecanacin

15136

152 178 179

(+)-isoelecanacin

ON(SO3K)2

Scheme 43

Oxidation of 152 with Fremy’s salt gave the hydroxyquinone 178 ([α]D –25.3), which

was converted into vinyl ether 179 ([α]D –15.6). Finally, irradiation of 179 followed by

chromatographic separation afforded the target (2R, 3aR, 4aR, 10aR)-elecanacin 36

with [α]D -145.2 (CHCl3) after recrystallisation and an ee of 99.5% (as determined by

HPLC with a chiral stationary phase). The diastereomer (2R, 3aS, 4aS, 10aS)-

isoelecanacin 151 was obtained in 98% ee, with a specific rotation of +110.4.

76

Page 88: Synthesis of Some Naturally Occurring Quinones

The specific rotation reported for natural elecanacin is +20.7 (CHCl3)47 and

therefore the natural major enantiomer is the mirror image of 36, and the configuration

is (2S, 3aS, 4aS, 10aS). From the magnitude of the reported rotation, the enantiomeric

excess within the natural material is only 14%, and thus elecanacin 36 is an example of

a natural product that does not exist in an enantiomerically pure form. Although many

natural products occur as single enantiomers, others are known to exist as enantiomeric

mixtures.130 Another example of a naturally occurring quinone that occurs as an unequal

mixture of enantiomers has been noted by Cotterill and coworkers.131 Dermolactone 180

is an anthraquinone that has been isolated from the fruiting bodies of the Australian

toadstool Dermocybe sanguinea (Figure 27). A significant difference in the specific

rotation values of synthetic (S)-dermolactone 180 ([α]D +169.3) and the natural material

([α]D +45.9), along with chiral shift experiments and chiral HPLC analysis revealed that

natural dermolactone occurs as a 64 : 36 mixture of (S)- and (R)-enantiomers

respectively (28% ee).131

O

OH

MeO

O

O

OH O

H

Me

180

Figure 27

2.6 On the Possible Biosynthesis of Elecanacin 36

The structural similarities between elecanacin 36 and its isomeric co-metabolites

eleutherin 38 and isoeleutherin 38 suggest that these compounds may arise from a

related biosynthetic pathway. Although the biosynthesis of eleutherin 38 and

isoeleutherin 39 does not appear to have been investigated, it presumably involves a

77

Page 89: Synthesis of Some Naturally Occurring Quinones

polyketide pathway, as has been established for other pyranonaphthoquinones.63, 132 For

example, 13C labelling experiments have revealed that the bacterial metabolite

nanaomycin A 182 is assembled from an octaketide by a folding resulting from

orientation 181a, whereas cardinalin 2 183, a metabolite from the toadstool Dermocybe

cardinalis, has been shown to arise by cyclisation of the octaketide with the alternative

orientation 181b (Figure 28).132 It has been suggested that this also may be the pathway

for the synthesis of plant-derived pyranonaphthoquinones, such as eleutherin 38.132

O

HO

Me

O

OMe

OMe

O

Me

O

HO

OH

MeO

H

H

Me

Me

O

OO

OOO

CO2HO

O

OO

Me

OOO

O

CO2H

181a 181b

183

nanaomycin A cardinalin 2182

∗ ∗

O∗

HO O

O

Me

CO2H∗

Figure 28. Incorporation of 1-13C labelled acetate into nanaomycin A 182 and of 1,2-

13C2 labelled acetate into cardinalin 2 183 (identified units shown in bold).

Given that eleutherin 38 and isoeleutherin 39 are likely to arise from a

polyketide pathway, what is the biogenetic origin of elecanacin 36? Formally the vinyl

ether 138 can be derived from a Norrish type II cleavage of 184, the dihydro-derivative

78

Page 90: Synthesis of Some Naturally Occurring Quinones

of eleutherin 38 and isoeleutherin 39, followed by oxidation (Scheme 44). A subsequent

intramolecular [2 + 2] cycloaddition would then deliver elecanacin 36. The

enantiomeric ratio of the product 36 would be determined by the configurational ratio at

C3 inherent within 184. Since both eleutherin 38 and isoeleutherin 39 co-occur in the

plant, the involvement of a biosynthetic precursor such as 184 having both

configurations at the methyl-substituted carbon would explain the low ee observed for

elecanacin 36.

MeO OH

O

O

Me

O

MeO O

O

Me

CH2

HMeO

O

O

Me

O

O

HMe

O

O

MeOH

O

HMe

O

O

MeOH

185184

a

138

36 ent 36

b

Scheme 44

Of the pericyclic reactions possibly involved in biological systems,133 the Diels-Alder

reaction has attracted considerable attention and some evidence has been presented for

the existence of Diels-Alderases.134, 135 Although the suggestion that elecanacin 36

could possibly be generated in vivo by a sequence of pericyclic reactions as outlined in

Scheme 44 must be regarded as highly speculative, some evidence of such reactions in

79

Page 91: Synthesis of Some Naturally Occurring Quinones

other biological systems can be provided in support. For instance, ethylene has been

reported to arise from a Norrish type II fragmentation of enzymatically generated triplet

butanal 188, providing an example of a photobiochemical reaction without light

(Scheme 45).136

O

H

CH2H

CH2

CH2

OH

CH2H

O2H CH2

O

(CH2)2CH3HO C

HCH

O O

(CH2)2CH3

O

OHH

triplet butanal

horseradishperoxidase

112 188

186

189

Norrish type IIcleavage

187

Scheme 45

The cleavage step a in Scheme 44 therefore has some precedence. Furthermore, recent

evidence has been presented that the conversion of isochorismate 190 to salicylate 191

and pyruvate 192, catalysed by the enzyme isochorismate pyruvate lyase, involves a

one-step pericyclic retro-ene process (Scheme 46).137, 138

CO2OH

H

O CO2

CO2

OH

O2CCCH3

OIsochorismatePyruvate Lyase

190 192191

Scheme 46

80

Page 92: Synthesis of Some Naturally Occurring Quinones

Nevertheless, at this stage there appears to be no clear evidence of any enzyme-

mediated analogies for the [2 + 2] cycloaddition, although several such reactions have

been proposed. Hao and coworkers have suggested that the dimeric natural product

sceptrin 194, isolated from the sponge Agelas sceptrum, may arise from a biological [2

+ 2] cycloaddition involving the co-metabolite oroidin 193 (Scheme 47).139 Similarly, it

has been proposed that the plant metabolite Biyouyanagin A 197 from Hypercium

chinese L. var. salicifolium is biosynthesised via a [2 + 2] cycloaddition between the

sesquiterpene 195 and enone 196 (Scheme 47).140, 141 However, so far no enzyme has

been identified as a catalyst for either reaction.

N

NH

NH

NNH2

NH2

NH

O

NH

Br

HN

O

HN

Br

NH

O

HN N

HN

NH2Br

sceptrin

194193

oroidin

enzyme

?

O O

Me

H

MeH

Me

Me

H

H

H

Ph

O

O

Me

biyouyanagin A

197

Me

H

MeH

Me

Me

O O

O

O

Me

Ph

enzyme

?

196195

Scheme 47

The observation that in solution the vinyl ether 138 undergoes ready cyclisation

when exposed to ambient laboratory light also raises the possibility that elecanacin 36 is

81

Page 93: Synthesis of Some Naturally Occurring Quinones

actually an artifact, arising through photochemical cyclisation of 138 during isolation

and workup of the plant extract. However, we consider it unlikely that the vinyl ether

138 is in fact a natural product. The formation of elecanacin 36 under laboratory light is,

as in the preparative reactions, accompanied by isoelecanacin 151. In view of the close

chromatographic Rf values of these products, it seems unlikely that 151 would have

been missed during the isolation of elecanacin 36. Thus elecanacin 36 is more likely to

be a product of a diastereoselective enzyme-mediated reaction rather than an artifact

arising from photochemical transformation of vinyl ether 138. Further work is required

to establish this.

2.7 Concluding Remarks

The current study has established through synthesis the nature of the novel ring

skeleton and absolute configuration of elecanacin 36. The biosynthetic origin of 36

remains to be determined, but may well involve an enzyme-mediated [2 + 2]

cycloaddition.

82

Page 94: Synthesis of Some Naturally Occurring Quinones

Chapter 3

The Synthesis of

3-Hydroxymethylfuro[3,2-b]naphtho[2,3-d]furan-

5,10-dione

83

Page 95: Synthesis of Some Naturally Occurring Quinones

3.1 Introduction to 3-Hydroxymethylfuro[3,2-b]naphtho[2,3-d]furan-5,10-dione 37

Another plant species which has attracted interest as a source of bioactive

naphthoquinones is the common tropical American tree Crescentia cujete L.

(Bignoniaceae). Traditionally, Crescentia cujete has been an important source of folk

medicine across Central America, the Northern half of South America and the

Caribbean,142 with both extracts and pulp of the seeds, fruit, leaves and flowers being

used to treat a variety of ailments. These include colds and other respiratory

illnesses,143-145 hypertension146 and the haemorrhagic effect of venomous snake bites.147

Like many other plant species used in traditional medicine, Crescentia cujete has also

attracted increasing interest as a source of novel and biologically active compounds

which could provide possible lead structures in the development of new

pharmaceuticals.48, 148-150

During an investigation into potentially useful anticancer agents from the wood

of Crescentia cujete, Kingston and coworkers isolated a series of nine related furo[b]-

fused naphthoquinone derivatives,48, 148 which are shown in Figure 29. Although

naphthoquinones 198-201 were known previously, five of the compounds 202-205 and

37 were new. Of particular interest are the two tetracyclic naphthoquinones, which were

isolated as red pigments and assigned structures 3-hydroxymethylfuro[3,2-

b]naphtho[2,3-d]furan-5,10-dione 37 and 9-hydroxy-3-hydroxymethylfuro[3,2-

b]naphtho[2,3-d]furan-5,10-dione 205 on the basis of extensive spectroscopic

analysis.48 Both compounds were found to be cytotoxic and 37 exhibited selective

DNA-damaging activity in an assay involving a DNA repair-deficient strain of yeast,48

suggesting that 37 may be useful in the development of new antitumour drugs. Kingston

and coworkers have also noted that the planar structure of naphthoquinones 37 and 205

84

Page 96: Synthesis of Some Naturally Occurring Quinones

means that intercalation into DNA is likely to contribute to their mode of action for

DNA damage.151

O

O

O

O

OH

205

OH

O

O

O

O

OH

37

O Me

OH

O

O

O Me

OH

O

OOH

O

OH

Me

O

OMeO

MeOH

H

O Me

O

OMeO

MeOH

H

O Me

O

OMeO HH

O Me

O

O HH

O Me

O

OOH HH

198

201

204203202

199 200

Figure 29

As well as having potentially useful biological activity, naphthoquinones 37 and

205 are rather interesting in terms of their structure. The tetracyclic ring system,

composed of a naphthoquinone ring fused to a furo[3,2-b]furan moiety, appears to be

unique to these compounds and represents a new natural product skeleton. In addition,

85

Page 97: Synthesis of Some Naturally Occurring Quinones

the presence of the aromatic furofuran moiety makes the two naphthoquinones highly

unusual for naturally occurring compounds. There are a several examples of natural

products which incorporate a reduced furo[3,2-b]furan ring system (Figure 30),

including panacene 206, a marine metabolite isolated from the sea hare Aplysia

brasiana,152 rutagravine 207, a plant metabolite isolated from a tissue culture of Ruta

graveolens,153 as well as two related xanthones psorofebrin 208 and 5’-

hydroxyisopsorofebrin 209, which have been obtained from the roots of Psorospermin

febrifugum.154 However, naphthoquinones 37 and 205 appear to be the first natural

products containing a fully aromatic furo[3,2-b]furan ring.

N

O

Me

OH

O

OOH

Me

HHO

OEt

H

H

H

CBr

H

H

206 207

O O

OOH

O OMe

HH

OHMe

O O

OMeO

O OH

HH

OH

OH209208

Figure 30

Synthetic compounds incorporating a fully aromatic furo[3,2-b]furan ring are

also very uncommon. The parent heterocycle 210 has not been synthesised to date

(Figure 31), although it has been included in a number of computational studies on

86

Page 98: Synthesis of Some Naturally Occurring Quinones

fused heterocycles.155-158 Similarly, the benzo-fused and naphtho-fused compounds 211

and 212 have not been prepared.

O

O

O

O

O

O

210 212211

Figure 31

In fact, syntheses of any compounds incorporating the aromatic furo[3,2-b]furan ring

system are rare, with only two reports describing the preparation of several benzo-fused

derivatives appearing to date.

Tolmach and coworkers have synthesised benzofuro[3,2-b]benzofuran 216 from

“hydrosalicyloin” 213 according to the route in Scheme 48.159

O

O

O

O

O

OBr

OH

OH

OH OH

hvNBS, benzoyl peroxideCCl4, reflux

silicachromatography

H

H

HBr

H

213 214

215216

DCC35 ºC

Scheme 48

87

Page 99: Synthesis of Some Naturally Occurring Quinones

The fused furofuran framework was constructed in the first step by a simple dehydrative

cyclisation of 213 with dicyclohexylcarbodiimide (DCC). The resulting dihydro

compound 214 was then converted into the bromo derivative 215 via photohalogenation

with N-bromosuccinimide (NBS). Spontaneous dehydrobromination of 215 during this

procedure and subsequent chromatographic purification on silica finally delivered the

fully aromatic compound 216 in 32% yield.

Vaidya and Agasimundin have also achieved the synthesis of three related

benzofuro[3,2-b]furan derivatives, which share the general structure 223 (Scheme

49).160 Compounds 223a, 223b and 223c have been reported to arise from Dieckmann

condensation involving treatment of the appropriate benzofuran 217 with strong base,

followed by acidification. A positive ferric chloride colour test seems to indicate that

these compounds exist at least partially in the enol form. Whilst most β-hydroxyfurans

tend to exist largely as the more stable keto tautomer, there is evidence that the enol

form may be predominant for some compounds.161 For instance, IR and NMR

spectroscopy of 2-acetyl-3-hydroxyfuran 224 provides evidence for only the enol

tautomer (Scheme 49). In this case, it has been suggested that intra- and/or

intermolecular hydrogen bonding plays an important role in stabilizing the β-

hydroxyfuran form.161 Similar opportunities for stabilizing hydrogen bonding would be

expected for the benzofuro[3,2-b]furan derivatives 223a and 223b.

88

Page 100: Synthesis of Some Naturally Occurring Quinones

O

OCH2R

C O

OCHR

C

OC

CO

O

O

O

O

O

O

O

OEt

O

OEt OC

CHO

OEtO

R

O

H

R

O

RR

O

R

OH

Base

where a: R = CO2Et b: R = C(O)Me c: R = CN

217

222 221 220

219218

223

H3O+

O C

Me

O

O H

224

OEt

Scheme 49

To date, there have been no reported syntheses of the two naturally occurring

furofuranonaphthoquinones 37 and 205. The novel structure, along with the potentially

useful selective DNA-damaging activity, make 3-hydroxymethylfuro[3,2-

b]naphtho[2,3-d]furan-5,10-dione 37 a particularly attractive target. Thus, it was

decided to develop an approach to the synthesis of 37, which should allow the structure

of the unusual tetracyclic ring skeleton to be confirmed. The construction of the fused

aromatic furofuran moiety in the natural product would be an important step in this

89

Page 101: Synthesis of Some Naturally Occurring Quinones

synthesis and it was thought that this might be possible by employing a cycloaddition-

cycloreversion strategy.

3.2 A Synthetic Approach to 3-Hydroxymethylfuro[3,2-b]naphtho[2,3-d]furan-

5,10-dione 37

A possible approach to the natural product is shown in Scheme 50.

O

O

OMe

OMe

CO2R

225

O

O

O

O

OH

OMe

OMe

O CO2R

O

OH

OMe

OMe

O

37

O

O

OMe

OMe

CO2R

226227

228

R = alkyl

Scheme 50

90

Page 102: Synthesis of Some Naturally Occurring Quinones

Retrosynthetic analysis suggested that the furofuranonaphthalene 225 would serve as a

suitable precursor to the natural product 37. Considering the synthetic direction, the

target compound should then be available from 225 through a short sequence involving

reduction of the ester group and oxidative demethylation. It was envisaged that the

furofuranonaphthalene 225 would in turn be available from the epoxy-bridged

dihydrobenzene derivative 226 via a cycloreversion strategy. Thus, loss of the etheno

bridge from 226 through a retro-Diels-Alder reaction should in principle provide the

furofuranonaphthalene 225. A possible route to the required epoxy-bridged intermediate

226 could involve an intramolecular Diels-Alder reaction within the substituted

naphthalene 227. Furan may be considered to be an oxygen-bridged diene and many

furan derivatives are able to undergo cycloaddition with various dienophiles, including

suitable acetylenes.162-165 Acetylenic furans of general structure 229 have been used

successfully in intramolecular Diels-Alder reactions to generate epoxy-bridged

polycyclic ring systems (Scheme 51).163

O

O

229 230

Scheme 51

On this basis, it was thought that compound 226 should be accessible through an

intramolecular Diels-Alder addition of the acetylenic ether moiety to the furan diene

within 227 (Scheme 50). Although the use of furan precursors has been somewhat

limited by the tendency of some intramolecular Diels-Alder adducts to undergo

cycloreversion to the starting material,163 the reaction involving acetylenic dienophiles

can be accomplished under milder conditions if the terminal carbon of the acetylene is

91

Page 103: Synthesis of Some Naturally Occurring Quinones

substituted with an activating group.166 Thus, it was reasoned that the presence of the

ester group in the acetylene chain of 227 should facilitate the proposed intramolecular

cycloaddition (Scheme 50). The acetylenic ether 227 may in turn be available from the

naphthol 228.

Buttery, Moursounidis and Wege have used a similar cycloaddition-

cycloreversion approach to prepare a furo[3,4-b]furan derivative 237 in good yield

(Scheme 52).166

O

231

O

O

CO2Me

N

N CO2MeN

O

OCO2Me

O

OCO2Me

O

NN

N

Py

Py

237

233

234

232

NN N

N

Py

Py

Py

Py

O

O

MeO2C

toluenereflux

N2

NN

Py

Py

236

1

2

5

8

9

10

3

235

Scheme 52

92

Page 104: Synthesis of Some Naturally Occurring Quinones

The activated acetylenic furan 231 was converted into 237 via a one pot reaction

involving heating 231 at 110 ºC in the presence of 3,6-di(pyridin-2’-yl)-1,2,4,5-tetrazine

233. This reaction proceeds according to the series of transformations depicted in

Scheme 52. An initial intramolecular Diels-Alder reaction generates adduct 232, which

undergoes cycloaddition at the electron-rich C8-C9 double bond to the electron-

deficient tetrazine 233 (Scheme 52). Spontaneous loss of nitrogen from the resulting

intermediate 234, followed by cycloreversion of 235 then produces the furo[3,4-b]furan

derivative 237. Thus, it was thought that a similar reaction may be applicable to the

synthesis of the natural product 37.

3.3 Previous Work Towards the Synthesis of 3-Hydroxymethylfuro[3,2-

b]naphtho[2,3-d]furan-5,10-dione 37

In order to test the viability of the tandem Diels-Alder-retro-Diels-Alder

approach to the fused furo[3,2-b]furan moiety present in the natural product 37,

previous work within our group by Slamet and Wege has focused on preparing a model

compound furo[3,2-b]benzofuran-3-methanol 238 (Figure 32).167, 168

O

O

238

OH

Figure 32

The first stage in the route to 238 involved preparation of an appropriate

acetylenic ether for the intramolecular Diels-Alder reaction. This was eventually

93

Page 105: Synthesis of Some Naturally Occurring Quinones

accomplished by employing a widely used dehydrohalogenation approach to acetylenic

ethers,169, 170 which has enabled the synthesis of a large variety of acetylenic ether

derivatives.171 The general strategy involves initial preparation of a 1,2-dichlorovinyl

ether 240 by reaction of an alkoxide with trichloroethylene (Scheme 53). Subsequent

dehydrohalogenation of the ether 240 upon treatment with an alkyllithium then

generates an intermediate lithium acetylide 241, which can be functionalized in situ by

trapping with various electrophiles (E-X) to give 242.

ROH

RO

Cl H

Cl

LiRO

ERO

1. Base

2. Cl2C=CHCl

R'Li

E-X

242

241240239

Scheme 53

Applying this approach to 245, the lithium salt of 2-(2’-furyl)phenol 243 was treated

with trichloroethylene, which gave dichloroether 244 in 93% yield (Scheme 54).167

Sequential treatment of 244 with butyllithium and methyl chloroformate then afforded

the desired acetylenic ether 245 in moderate yield (39%).

94

Page 106: Synthesis of Some Naturally Occurring Quinones

OH

O O

O Cl

Cl H

244

O

O

CO2Me

1. LiOMe MeOH

2. Cl2C=CHCl DMF

1. n-BuLi, Et2O2. Cl-CO2Me

243

245

Scheme 54

With the required ether 245 in hand, efforts were directed towards construction

of the furofuran framework via the proposed cycloaddition-cycloreversion sequence.

Encouragingly, treatment of 245 with 3,6-di(pyridin-2’-yl)-1,2,4,5-tetrazine 233 in

refluxing toluene was successful, affording methyl furo[3,2-b]benzofuran-3-carboxylate

247 in 93% yield (Scheme 55). The structure of this compound was confirmed using 2D

NMR techniques. Finally, reduction of ester 247 with lithium aluminium hydride

provided the target compound 238 in 64% yield.

95

Page 107: Synthesis of Some Naturally Occurring Quinones

O

O

CO2Me

NN N

N

py

py

toluene, reflux

O

O

CO2Me

233

246

247

heatO

O

CO2Me

O

O

238

OH

LiAlH4Et2O

245

Scheme 55

The success of the tandem Diels-Alder-retro-Diels-Alder approach to the model

compound 238 showed that this strategy should be applicable to the synthesis of the

natural product 3-hydroxymethylfuro[3,2-b]naphtho[2,3-d]furan-5,10-dione 37.167 In

order to extend this approach to 37, Slamet made a number of attempts to prepare and

isolate the required intermediate acetylenic ether 249 (Scheme 56). Reaction of the

naphthoxide generated from naphthol 228 and lithium methoxide with trichloroethylene

proceeded smoothly and gave the 1,2-dichlorovinyl naphthyl ether 248 in 66% yield.

However, the subsequent dehydrochlorination step proved to be troublesome and failed

to provide 249. It is worth noting that a number of similar alkoxyacetylenic ester

derivatives are reported to be relatively unstable and prone to decomposition.172 In order

to avoid possible degradation of acetylenic ether 249 during purification, an attempt was

also made to prepare the advanced intermediate 251 by subjecting the crude reaction

96

Page 108: Synthesis of Some Naturally Occurring Quinones

product directly to the subsequent cycloaddition-cycloreversion sequence (Scheme

56).167

OMe

OMe

O

O

CO2MeO

O

OMe

OMe

CO2Me

OMe

OMe

O

O

CO2Me

heat

251

250 249

NN N

N

Py

Py

toluene, reflux

233

OMe

OMe

O

OH

228

OMe

OMe

O

O Cl

Cl H

248

1. LiOMe MeOH

2. Cl2C=CHCl DMF

1. n-BuLi, Et2O2. Cl-CO2Me

Scheme 56

Thus treatment of the presumed ether 249 with tetrazine 233 in reluxing toluene

afforded in 2% yield a product, which was formulated as the desired

furofuranonaphthalene 251 on the basis of NMR spectroscopy.167 However, re-

97

Page 109: Synthesis of Some Naturally Occurring Quinones

examination of the 1H NMR spectrum suggests that the correct structure for this

compound is actually 252 (Figure 33). In particular, the chemical shift of the furyl

proton singlet (7.39 ppm) is further upfield than would be expected for the target

furofuranonaphthalene 251 based on the value observed for the analogous proton (8.11

ppm) in the model ester 247167 (Figure 33). Moreover, the furyl proton shift of the

reaction product is more comparable to that observed for the H3 signal in furan 253

(7.57 ppm),173 providing further support for structure 252.

OMe

OMe

O

O

H

252

CO2Me

O

O

CO2Me

247

H

O CO2Me

H

253

δH 7.39

δH 7.57

δH 8.11

Figure 33

This compound is derived from metallation of the furan ring of 248 during the

dehydrohalogenation step (Scheme 57). Reaction with methyl chloroformate at the

furan ring then generates intermediate 255. Finally, an intramolecular cycloaddition

between the acetylene moiety and furan ring within 255, followed by the usual

cycloreversion sequence upon treatment of the resultant adduct 256 with tetrazine 233

delivers the unexpected furofuranonaphthalene 252.

98

Page 110: Synthesis of Some Naturally Occurring Quinones

OMe

OMe

O

O

Li

O

O

OMe

OMe

OMe

OMe

O

O

heat

252

256

254

NN N

N

Py

Py

toluene, reflux

233

OMe

OMe

O

O Cl

Cl H

248

n-BuLi Et2O

Li

OMe

OMe

O

O

H

255

CO2MeCO2Me

CO2Me

1. Cl-CO2Me, Et2O2. H2O

Scheme 57

In view of the problems encountered with the dehydrohalogenation approach to

the required acetylenic ether 249, it was decided to begin the present project by

99

Page 111: Synthesis of Some Naturally Occurring Quinones

investigating an alternative approach to an appropriate acetylenic ether, which could be

used as a key intermediate in the proposed route to the natural product 37 (Scheme 50).

3.4 An Alternative Approach to an Intermediate Acetylenic Ether

A general method that has been employed for preparing acetylenic ethers

involves substitution of a halide with an alkoxide at an acetylenic carbon, which

proceeds by the addition-elimination mechanism shown in Scheme 58.171, 174 For

example, Tanaka and Miller have used this approach to prepare 1-ethoxy-2-

phenylacetylene 261 in 42% yield from sodium ethoxide and 1-chloro-2-

phenylacetylene 260.175

C CX R

RO

C CR

RO

XC CRO R

X

257 258 259

Cl Ph EtO PhNaOEt

DMSO

260 26142%

Scheme 58

Traditionally the synthesis of acetylenic ethers via the reactions of simple

haloacetylenes and alkoxides has been of limited utility because these reactions, like the

example above, tend to be rather low-yielding.171, 176 These low yields are usually

100

Page 112: Synthesis of Some Naturally Occurring Quinones

attributed to a number of competing reactions, including displacement of acetylide

RC≡C- by attack of the alkoxide anion on the halogen atom.176, 177 Nevertheless, it was

envisaged that the target acetylenic ether 263 might be available by a similar strategy

involving the base-catalysed reaction of naphthol 228 and an appropriate haloacetylene

262, as the presence of the ester functional group should activate the acetylene

sufficiently for effective nucleophilic addition (Scheme 59).

OMe

OMe

OH

O

1. Base

2. X CO2Et

OMe

OMe

O

O

CO2Et

263228

262

Scheme 59

Acetylenes substituted with electron-withdrawing groups, including carbonyl, nitrile

and imine groups, are very reactive towards nucleophilic addition because the resulting

carbanionic intermediates can be stabilized through resonance.178, 179 An efficient route

to ethylenic ethers which takes advantage of this reactivity involves conjugate addition

of an alkoxide to an activated acetylene 264, followed by protonation of the vinyl anion

intermediate 265.178, 179 The reaction pathway is outlined for the generic reaction in

Scheme 60, where the activating group in the acetylenic substrate is an ester.

101

Page 113: Synthesis of Some Naturally Occurring Quinones

102

C CH COR

RO

264

O

C C COR

265a

O

C C COR

O

H

RO

H

RO

C CHH

RO

CO2R

[H+ ]

265b

266

Scheme 60

For example, Ciganek has prepared the ethylenic ether 269 in high yield (92%) by a

base-catalysed addition of 2-hydroxy-3-methoxybenzaldehyde 267 to methyl propiolate

268 (Scheme 61).180

CHO

OHO NMe

HC CCO2Me

267 268

CHO

OCO2Me

269

MeCN

Scheme 61

On this basis, it was thought that the target acetylenic ether 263 should be available

according to the reaction pathway outlined in Scheme 62, beginning with a similar base-

catalysed conjugate addition of naphthol 228 to an activated haloacetylene 262. In this

Page 114: Synthesis of Some Naturally Occurring Quinones

case, however, elimination of the halide from intermediate 270 should regenerate the

acetylenic system.

C CX COEt C C COEt

O

X

ArOC CArO CO2Et

270 263

OAr

HB

O

where Ar =

OMe

OMe

O

228 262

Scheme 62

This approach to acetylenic ethers has received only limited attention, which is possibly

due to problems associated with the reactivity of the activated carbon-carbon triple bond

in the product. For instance, Vereshchagin and coworkers have reported that the

acetylenic ether 273, which arose from the base-catalysed reaction of phenol 271 and

the bromoacetylenic ketone 272, was too reactive to be isolated (Scheme 63).181 Instead,

273 underwent further nucleophilic addition with a second equivalent of phenol

affording the ethylenic ether 274 in high yield (94%).

103

Page 115: Synthesis of Some Naturally Occurring Quinones

K2CO3Me2CO

2

Br

OPhOH PhO

O

PhPhO

PhOO

Ph

Ph271

274273272

Scheme 63

Despite these results, it was felt that the reaction outlined in Scheme 62 was worth

examining, particularly since the pendant furan ring within 263 could potentially

intercept the reactive alkyne moiety.

3.5 Attempted Synthesis of Acetylenic Ether 263

The initial step in the synthesis of acetylenic ether 263 required preparation of

the precursor naphthol 228 and a suitable haloacetylene. The naphthol 228 has been

prepared previously within our group by the route shown in Scheme 64.167, 168 A

Hauser-Kraus annulation reaction182, 183 of cyanophthalide 275 and the Michael acceptor

2-(2-furyl)acrolein 277, followed by methylation of the resultant hydroxyquinone 281

gave the aldehyde 282 in 85% yield. Baeyer-Villiger oxidation of 282 with aqueous

hydrogen peroxide and diphenyl selenide and subsequent hydrolysis of the aryl formate

283 then provided the target naphthol 228. However, the Baeyer-Villiger step proved

difficult and the naphthol 228 could be prepared in only moderate yield (49%) at best.167

This suggested that a more efficient approach to 228 would be needed in order to

synthesise a sufficient quantity of starting material for the current work.

104

Page 116: Synthesis of Some Naturally Occurring Quinones

O

O

CNH

LDATHF

O

O

CN

H

O

O

O

O

CNO

H

OO

O

CHO

O

CN

O

O

CHO

O

CHO

OH

OH

O

CHO

OMe

OMe

O

OCHO

OMe

OMe

O

OH

OMe

OMe

O

H3O+

MeI, K2CO3Me2O

275

228

283 282

280 281

278

277276

279

H2O2diphenyl diselenideCH2Cl2

NaOHH2O/MeOH

Scheme 64

105

Page 117: Synthesis of Some Naturally Occurring Quinones

The naphthol 228 was prepared more conveniently following the sequence

outlined in Scheme 65.

CH2OLi

O

O

Br

Br

O

O

OCH2Ph

Br

OBu3Sn

O

O

OCH2Ph

O

OH

OH

O

OCH2Ph

OMe

OMe

O

OCH2Ph

OMe

OMe

O

OH

Pd(Ph3)4CuBrdioxane

Na2S2O4Bu4NBrCH2Cl2/H2O

H2, 10% Pd/CEtOAc

1. NaOH2. (MeO)2SO2

THF 0 ºC

284

287

286285

288289

228290

Scheme 65

The benzyl ether 290 was readily obtained according to the method of Slamet and

Wege,167, 168 beginning with the reaction of 2,3-dibromo-1,4-naphthoquinone 284 and

the lithium salt of benzyl alcohol 285, which gave the naphthoquinone 286. A Stille

106

Page 118: Synthesis of Some Naturally Occurring Quinones

coupling reaction of stannane 287184 and 286, in the presence of

tetrakis(triphenylphosphine)palladium(0) and copper (I) bromide, then provided the red

naphthoquinone 288 (86%), which was converted into the benzyl ether 290 via a one

pot reduction/methylation in 96% yield. The final step involved removal of the benzyl

protecting group. Since furans are susceptible to reduction by catalytic

hydrogenation,185, 186 it was envisaged that a milder method may be required to liberate

the naphthol. Nevertheless, careful hydrogenolysis of 290 with 10% Pd/C catalyst at

room temperature proceeded smoothly and returned the naphthol 228 as a pale yellow

oil in 91% yield (Scheme 64).

Ethyl 3-bromopropiolate 292 was chosen as the acetylenic substrate as it could

be readily obtained by bromination of commercially available ethyl propiolate 291 with

N-bromosucciniminde according to the general method of Leroy187 (Scheme 66).

H CO2Et

NBSAgNO3 Br CO2Et

291 292

Me2CO

Scheme 66

With the naphthol 228 and bromoacetylene 292 in hand, attention was turned to

the synthesis of the acetylenic ether 263 via the proposed base-catalysed addition-

elimination sequence. In an intial experiment, treatment of naphthol 228 in

tetrahydrofuran with butyllithium followed by ethyl 3-bromopropiolate 292 gave only a

complex mixture of products, most of which could not be identified, with no sign of the

target ether 263 (Scheme 67).

107

Page 119: Synthesis of Some Naturally Occurring Quinones

OMe

OMe

O

OH

OMe

O

O

MeO

CO2Et

BuLi THF

Br CO2Et

294

228

30%

OLi

OMe

OMe

O

293

292

OMe

OMe

O

O

CO2Et

Br CO2Et

292

263

Scheme 67

However, the product mixture did include a compound which was assigned structure

294 on the basis of NMR and mass spectral evidence. The mass spectrum showed a

molecular ion at m/z 366 corresponding to a molecular formula of C21H18O6. The

downfield region of the 1H NMR spectrum includes signals integrating for four

aromatic protons and three furyl protons (Figure 34). The upfield region of the spectrum

shows two methoxy signals at 3.75 and 3.48 ppm, along with a methylene quartet and

methyl triplet at 4.22 ppm and 1.28 ppm respectively, corresponding to the ethyl ester

moiety. The target acetylenic ether structure 263 could be ruled out on the basis of the

13C NMR spectrum which includes not only an ester carbonyl signal at 165.7 ppm but

108

Page 120: Synthesis of Some Naturally Occurring Quinones

also a second carbonyl signal at 190.7 ppm attributable to a keto group. The infrared

spectrum shows two carbonyl absorbances at 1713 and 1676 cm-1 and an alkynyl

absorption at 2237 cm-1, confirming the presence of the ketone and acetylenic ester

moieties.

Figure 34. 1H NMR spectrum of 294 (300 MHz, CDCl3).

This compound arises by carbon-alkylation of the naphthoxide 293, rather than

the desired oxygen-alkylation. Although this result was not expected, with hindsight the

isolation of ketone 294 is not surprising as the formation of C-alkylated products in

nucleophilic substitution reactions involving the β-naphthoxide system is well

established,188, 189 and can be explained by considering the ambident nucleophilic nature

of the β-naphthoxide intermediate 293 wherein the negative charge is shared by the

oxygen and α-carbon atoms (Figure 35).

109

Page 121: Synthesis of Some Naturally Occurring Quinones

O O

293b293a

OMe

OMe

O

OMe

OMe

O

Figure 35

Since a β-naphthoxide ion can attack an alkylating agent via either the oxygen or the α-

carbon atom, the formation of both O- and C-alkylation products is possible.188 In

practice, however, the resulting O/C-alkylation product ratio for ambident anions is

largely influenced by solvent effects.189, 190 Although it is unclear why none of the target

acetylenic ether 263 was isolated from the reaction in Scheme 67, the formation of the

C-alkylated product is consistent with the observations of Kornblum and coworkers

who similarly found that a significant amount of C-alkylation occurred when the related

sodium β-naphthoxide salt 295 was treated with benzyl bromide in aprotic non-

dissociating solvents including tetrahydrofuran (Scheme 68).188

ONa PhCH2BrO

CH2Ph OH

CH2Ph

295 297296

60% 36%

THF

Scheme 68

This tendency for C-alkylation of β-naphthoxide species in aprotic non-dissociative

solvents has been interpreted in terms of the strength of the electrostatic attraction

between the naphthoxide ion and counter-cation, which is particularly important in

these solvents due to their poor cation solvating ability.188, 189 In the case of the reaction

in Scheme 68, the hard sodium ion is likely to preferentially associate with the harder

110

Page 122: Synthesis of Some Naturally Occurring Quinones

oxygen centre in the β-naphthoxide ion rather than with the softer α-carbon centre, and

this strong interaction provides a barrier to O-alkylation.189 As the α-carbon is relatively

free to associate with the alkylating agent benzyl bromide, C-alkylation of the

naphthoxide ion is promoted. It would be expected that the stronger the electrostatic

attraction between the cation and naphthoxide oxygen atom, the greater the proportion

of C-alkylation product that should form. Thus smaller cations, such as Li+, which have

higher charge densities should increase the tendency for C-alkylation because they bind

tightly to the hard oxygen atom.188, 189 On reflection therefore, it is not surprising that

the ketone 294 arose from our reaction because the use of lithium β-naphthoxide 293 in

tetrahydrofuran would be expected to promote C-alkylation (Scheme 67). If the desired

O-alkylated product 263 was also formed in this reaction, it may well have been

consumed by side reactions.

On this basis, it was thought that the use of a bulkier cation might be beneficial

in encouraging the formation of the target ether 263 as the lower charge density on the

cation should weaken the naphthoxide oxygen-cation interaction in the intermediate 293

and therefore promote O-alkylation. Accordingly, a number of bases were investigated.

The reaction was attempted by gently heating a solution of naphthol 228 and ethyl 3-

bromopropiolate 292 and N-methyl morpholine in acetonitrile (Scheme 69).

OMe

OMe

OH

O

N OMe , MeCN

(or K2CO3, acetone)

Br CO2Et

228292

OMe

OMe

O CO2Et

OH

298

Scheme 69

111

Page 123: Synthesis of Some Naturally Occurring Quinones

Unfortunately, there was no sign of the desired ether 263, with the reaction returning

only starting material and a complex mixture of products. Although most of these

compounds could not be identified, the mixture did include a small amount of a highly

crystalline product (4%), which was assigned structure 298 on the basis of NMR and

mass spectral evidence. The mass spectrum includes a molecular ion at m/z 366, which

corresponds to a molecular formula of C21H18O6. Importantly, the downfield region of

the 1H NMR spectrum exhibits a singlet at 11.55 ppm, indicating the presence of an

intramolecularly H-bonded phenolic substituent, along with signals for six aromatic

protons, which resonate as three multiplets at 8.36-8.34, 8.3-8.22, 7.54-7.51 ppm, and a

doublet at 7.03 ppm (Figure 36). Further upfield are two singlets at 4.45 and 4.09 ppm,

which are attributable to the two methoxy groups. The ethyl ester methylene and methyl

protons occur as a quartet at 4.57 ppm and a triplet at 1.56 ppm respectively.

Figure 36. 1H NMR spectrum of 298 (500 MHz, CDCl3).

112

Page 124: Synthesis of Some Naturally Occurring Quinones

Similarly, repeating the reaction with potassium carbonate in acetone also afforded the

phenol 298, albeit in higher yield (28%), accompanied by a complex mixture of

products, none of which was the desired ether 263 (Scheme 69). A mechanistic rationale

for the formation of phenol 298 is outlined in Scheme 70, and indicates that both the

target acetylenic ether 263 and the advanced intermediate 299 are being produced as

reactive intermediates.

OMe

OMe

OH

O

OMe

OMe

O

O

CO2Et

O

O

O

OH

CO2Et

OMe

OMe

OMe

OMe

CO2Et

O

OMe

OMe

CO2Et

OH

H

OMe

OMe

O CO2Et

OH

N OMe, MeCN

(or K2CO3, acetone)

Br CO2Et

H+

H+

228

298301

300 299

263

292

Scheme 70

113

Page 125: Synthesis of Some Naturally Occurring Quinones

The initial base-catalysed reaction of naphthol 228 and activated bromoacetylene 292

proceeds as expected to give the ether 263, which is evidently too reactive to be isolated

and is converted into the required Diels-Alder adduct 299 via an intramolecular

cycloaddition between the acetylenic ether moiety and furan ring within 263. Finally,

acid-catalysed opening of the epoxy bridge within adduct 299, possibly during

hydrolytic workup, effects aromatisation, generating the phenol 298.

3.6 Synthesis of Key Intermediate Ethyl 5,10-Dimethoxyfuro[3,2-b]naphtho[2,3-

d]furan-3-carboxylate 304

Despite the generation of the unexpected phenol 298 according to the reaction

pathway outlined in Scheme 70, the formation of the two required intermediates 263

and 299 was rather encouraging as it seemed that the synthesis of the key

furofuranonaphthalene 304 might be achieveable if adduct 299 could be trapped with

3,6-di(pyridin-2’-yl)-1,2,4,5-tetrazine 233 (Scheme 71). Subsequent loss of nitrogen

from the resultant intermediate 302, followed by cycloreversion of 303 should then

generate furofuranonaphthalene 304. Towards this end, a stirred suspension of naphthol

228, ethyl 3-bromopropiolate 292 and potassium carbonate in acetone was heated gently

until the starting material had been consumed (TLC analysis) (Scheme 71). Without

hydrolytic workup, the reaction mixture was filtered, diluted with toluene, and most of

the acetone was removed by distillation. Then, the solution of the presumed adduct 299

was subjected to the action of tetrazine 233 in refluxing toluene. Gratifyingly, this

afforded the target furofuranonaphthalene 304, albeit only in low yield (11%).

114

Page 126: Synthesis of Some Naturally Occurring Quinones

OMe

OMe

OH

O

OMe

OMe

O

O

CO2Et

O

O

OMe

OMe

CO2Et

K2CO3, acetone

Br CO2Et

NN N

N

py

py

toluene, reflux

OMe

OMe

O

O

CO2Et

O

OMe

OMe

CO2Et

O

N Npy

N Npy

O

OMe

OMe

CO2Et

O

N Npypy

N2

228

303

302

233

299

292

263

304

N

N

py

py

236

heat

Scheme 71

The structure of the furofuranonaphthalene 304 follows from its NMR and mass

spectral properties. The mass spectrum exhibits a molecular ion at m/z 340, which

supports the molecular formula C19H16O6. The 1H NMR spectrum is very simple (Figure

115

Page 127: Synthesis of Some Naturally Occurring Quinones

37). The high field region includes two singlets at 4.39 and 4.30 ppm, corresponding to

the two methoxy group protons, as well as a methylene quartet at 4.45 ppm and a

methyl triplet at 1.45 ppm due to the ethyl ester moiety. The aromatic region integrates

for five protons: the four aryl protons, which resonate as three multiplets at 8.33-8.31,

8.28-8.16 and 7.54-7.47 ppm, and the furyl α-hydrogen, which resonates as a singlet at

8.10 ppm. The furyl α-carbon occurs at 150.6 ppm in the 13C NMR spectrum.

Figure 37. 1H NMR spectrum of 304 (500 MHz, CDCl3).

A comparison of the α-furyl chemical shifts of 304, the related model compound 247

prepared previously within our group167, 168 and the natural product 3748 is shown in

Figure 38.

116

Page 128: Synthesis of Some Naturally Occurring Quinones

O

O

O

O

O

O

OH

OMe

OMe

CO2EtO

O

CO2Et

304 247

δC 150.8, δΗ 8.18

δC 150.6, δΗ 8.10 δC 150.8, δΗ 8.11

O δC 142.5, δΗ 7.43

30537

Figure 38

The similarity between these shift values provided support for the presence of the

furofuran moiety in the target furofuranonaphthalene 304. It is interesting to note that

the α-furyl chemical shifts for 304, 247 and 37 are considerably downfield of the

analogous carbon and proton signals in furan191 itself (Figure 38). This is likely to be

due to the electron-withdrawing effect of the ester groups in 304 and 247 and the

carbonyl group in 37, which effectively deshields the α-furyl position (Figure 39).

117

Page 129: Synthesis of Some Naturally Occurring Quinones

O

O

OMe

OMe

C

304a

O

O

O

O

OH

37a

O

OEt

O

O

OMe

OMe

C

O

OEt

304b

O

O

O

O

OH

37b

Figure 39

Although the overall yield of the reaction sequence outlined in Scheme 71 was

disappointing, the successful synthesis of the key intermediate 304 was encouraging as

it validated this approach to the novel furo[3,2-b]naphtho[2,3-d]furan ring system. In an

attempt to improve this yield, a number of different reaction conditions were examined

(Scheme 72). Cesium carbonate has received much attention for its effective use in the

O-alkylation of phenols,192, 193 with the product ethers often returned in yields higher

than those achieved with potassium carbonate.192 In an attempt to improve the

efficiency of the initial ether formation step, and thus the overall yield of the sequence

outlined in Scheme 71, potassium carbonate was replaced with cesium carbonate.

However, this returned the desired furofuranonaphthalene 304 in only 4% yield

(Scheme 72a).

118

Page 130: Synthesis of Some Naturally Occurring Quinones

OMe

OMe

OH

O

OMe

OMe

O

O

CO2Et

O

O

OMe

OMe

CO2Et

i) Cs2CO3, acetone, RTor ii) KOMe, MeOH, 0 ºC to RTBr CO2Et

NN N

N

py

py

toluene, reflux

OMe

OMe

O

O

CO2Et

228

233

299

292

263

304

heat

a)

OMe

OMe

OH

O Br CO2Et

228

292

b)NN N

N

py

py

toluene, reflux

233complex mixture

Scheme 72

Alternatively, performing the initial step of the sequence with potassium methoxide in

methanol was also unsuccessful and afforded only a trace of the target compound

(Scheme 72a). As it seemed possible that the equilibrium concentration of adduct 299 is

very low, it was thought that it may be beneficial to include 3,6-di(pyridin-2’-yl)-

1,2,4,5-tetrazine 233 in the reaction mixture during the intramolecular cycloaddition

step because trapping adduct 299 immediately upon its formation could help to drive the

overall cycloaddition-cycloreversion sequence to completion. In this case, the tetrazine

119

Page 131: Synthesis of Some Naturally Occurring Quinones

233 could also act as as the base for the initial base-catalysed conjugate addition-

elimination step to form the acetylenic ether 263. Unfortunately, when a mixture of

naphthol 228, bromoacetylene 292 and tetrazine 233 in toluene were heated under

reflux, the reaction returned only a complex mixture with no recognizable products

(Scheme 72b).

Although the yield of furofuranonaphthalene 304 could not be increased, it

seems possible that with further experimentation conditions may be found for

improving the efficiency of the sequence. In particular, the use of an aprotic dipolar

solvent, such as N,N-dimethylformamide, which is known to promote O-alkylation of β-

naphthoxide,188, 189 may be beneficial for encouraging the synthesis of the initial ether

intermediate 263. The strong cation-solvating ability of these solvents should result in

effective dissociation of the naphthoxide salt, thereby increasing the nucleophilicity of

the naphthoxide oxygen and promoting attack of the bromoacetylene 292. Similarly, the

use of another effective dissociating agent, such as a crown ether,189 may also be worth

investigating. However, time constraints prevented further experimentation at this stage.

Instead, attention was briefly turned to the route employed by Slamet and

Wege167, 168 for preparing the model furo[3,2-b]benzofuran 247, which involved

synthesis of the precursor acetylenic ether 245 via dehydrohalogenation of a 1,2-

dichloroether, as discussed in Section 3.3 (p. 93). Although an attempt at extending this

approach to one possible furofuranonaphthalene intermediate 251 had been

unsuccessful, it was thought that the sequence might be more successfully applied to the

present target 304, particularly in light of the knowledge that the two intermediates 263

and 299 require careful experimental manipulation due to their high instability.

120

Page 132: Synthesis of Some Naturally Occurring Quinones

The required 1,2-dichlorovinyl l ether 248 was prepared from naphthol 228 by

adapting the method of Slamet and Wege167, 168 described in Section 3.3. Naphthol 228

was treated with n-butyllithium in tetrahydrofuran at 0 ºC and the resulting lithium salt

was treated with trichloroethylene in dimethylformamide to give 248 in 85% (Scheme

73).

OMe

OMe

O

O Cl

Cl H1. n-BuLi, THF, 0 ºC

2. trichloroethylene, DMF, RT

248

OMe

OMe

O

OH

228

Scheme 73

Attention was then turned to the synthesis of the target furofuranonaphthalene 304

(Scheme 74). In an effort to prepare the acetylenic ether 263, a solution of 1,2-

dichlorovinyl ether 248 in tetrahydrofuran at –78 ºC was treated with t-butyllithium.

When TLC analysis indicated that all the starting material had been consumed, ethyl

chloroformate was added to the resulting precipitate and the mixture was allowed to

warm to room temperature over 1 h. Then, without hydrolytic workup, the crude

reaction product was subjected to the action of 3,6-di(pyridin-2’-yl)-1,2,4,5-tetrazine

233 in refluxing toluene.

121

Page 133: Synthesis of Some Naturally Occurring Quinones

OMe

OMe

O

O Cl

Cl H

1. t-BuLi, THF, -78 ºC2. Cl-CO2Et, THF, -78 ºC to RT

248

OMe

OMe

O

O

CO2Et

O

O

OMe

OMe

CO2Et

OMe

OMe

O

O

CO2Et

heat

304299

268

NN N

N

Py

Py

toluene, reflux

233

Scheme 74

However, the sequence was not successful and afforded an unexpected product, which

can be formulated as one of two possible structures 306 and 307 on the basis of NMR

and mass spectral evidence (Scheme 75). Structure 306 arises from lithiation at both the

ethylenic ether and α-position of the furan moiety in 248, followed by reaction with

ethyl chloroformate. The second possible structure 307 is the product of an

intramolecular Diels-Alder addition between the pendant ethylenic ether and the furan

diene within structure 306.

122

Page 134: Synthesis of Some Naturally Occurring Quinones

O

O

OMe

OMe

CO2EtCl

Cl

CO2Et

OMe

OMe

O

O

CO2Et

Cl

Cl

CO2Et

OMe

OMe

O

O Cl

Cl H

1. t-BuLi, THF, -78 ºC2. Cl-CO2Et, THF, -78 ºC to RT

248 306

307

heat

Scheme 75

The high resolution mass spectrum exhibits a molecular ion at m/z 508.0692,

corresponding to the molecular formula C24H22Cl2O8. The 1H NMR spectrum includes

signals for four aromatic protons, which occur as two multiplets at 8.23-8.11 and 7.65-

7.57 ppm (Figure 40). The high field region exhibits two methoxy singlets at 4.00 and

3.89 ppm, as well as resonances consistent with two ethyl ester groups, with two sets of

methylene quartets at 4.39 and 4.31 ppm and two methyl triplets at 1.39 and 1.36 ppm.

The spectrum also includes two doublets, resonating at 7.33 and 6.94 ppm with J = 3.6

Hz, which can be assigned to either the furyl protons in 306 or the vinylic protons in

adduct 307.

123

Page 135: Synthesis of Some Naturally Occurring Quinones

Figure 40. 1H NMR spectrum of compound assigned to either structure 306 or 307 (300

MHz, CDCl3).

However, it seems more likely that the product has structure 306 rather than 307 as the

magnitude of the coupling constant is characteristic of the H3-H4 coupling constants

typically observed for furans.194 For example, this J value compares well with 3.3 Hz

reported for the H3 and H4 protons in a similar system 308195 and is significantly

smaller than the coupling constant (J = 5.2 Hz) reported for the vinylic protons in a

representative 7-oxanorbornene derivative 309196 (Figure 41).

OCO2Et

308

HH

J 3.3 Hz

O

CO2CH3H3CO2C

H H

CH3H

J 5.2 Hz

309

Figure 41

124

Page 136: Synthesis of Some Naturally Occurring Quinones

Signals at 162.1 and 158.7 ppm in the 13C NMR spectrum are consistent with two

carbonyl groups and confirmed the presence of the ethyl ester moieties.

In view of the synthesis of the undesired dichloro compound, any further

attempts to improve the yield of the furofuranonaphthalene intermediate 304 were

deferred at this stage. A small quantity of 304 was available from the reaction sequence

described in Scheme 71. This was used to investigate the remaining steps in the

proposed route of the natural product, which is the subject of the next section.

3.7 Final Steps in the Synthesis of 3-Hydroxymethylfuro[3,2-b]naphtho[2,3-

d]furan-5,10-dione 37

Initially, it was envisaged that the target naphthoquinone 37 would be accessible

by a short sequence involving reduction of the ester group in furofuranonaphthalene

304, followed by oxidative demethylation of the resultant alcohol 310 (Scheme 76).

O

O

OMe

OMe

CO2EtO

O

OMe

OMe

OH

O

O

O

O

OH

LiAlH4THF, 0 ºC

CANMeCN/H2O0 ºC

304 310

37

Scheme 76

125

Page 137: Synthesis of Some Naturally Occurring Quinones

Treatment of 304 with lithium aluminium hydride in tetrahydrofuran, followed by

aqueous workup gave an orange residue, which was kept in the freezer overnight. The

1H NMR spectrum of the crude reaction product provided evidence for the structure of

alcohol 310 with a singlet at 4.83 ppm, attributable to the methylene protons of the

hydroxymethyl moiety, as well the disappearance of the ethyl ester group methylene

and methyl signals, which resonated in the high field region of the 1H NMR spectrum of

the precursor furofuranonaphthalene 304. However, attempted oxidative demethylation

of alcohol 310 with ceric ammonium nitrate in aqueous acetonitrile197 was not

successful and resulted in a complex mixture of products (by 1H NMR), with no trace of

the target naphthoquinone 37. Although the reason for this is unclear, it is possible that

the electron-rich furofuran moiety of 310 may have undergone oxidative degradation.

Fortunately a modified route to naphthoquinone 37, which avoids intermediate

alcohol 310, proved to be successful (Scheme 77).

O

O

O

O

CO2EtO

O

OMe

OMe

CO2Et

O

O

OH

OH

OH

O

O

O

O

OH

CANMeCN/H2O0 ºC

LiAlH4THF, 0 ºC

304 311

31237

ON(SO3K)2

Scheme 77

126

Page 138: Synthesis of Some Naturally Occurring Quinones

The furofuranonaphthalene 304 was first converted into the bright yellow quinone 311

by oxidative demethylation with ceric ammonium nitrate (Scheme 77). Here the

presence of the electron-withdrawing ester group presumably stabilizes the furofuran

system. Subsequent treatment with lithium aluminium hydride, followed by aqueous

workup effected reduction. The intermediate hydroxyquinone 312 was then isolated and

immediately subjected to oxidation with Fremy’s salt; this delivered the target 3-

hydroxymethylfuro[3,2-b]naphtho[2,3-d]furan-5,10-dione 37 as a red solid in 49% yield

over two steps.

Although the natural product is reported to melt at 217-218 ˚C,48 the synthetic

material was observed to sublime and undergo a phase change from ca 204 ˚C onwards

and then melt at 240-241 ˚C. The 1H NMR spectrum of the synthetic material compares

well with that reported for the natural product (Figures 43a and Table 2).48 However, the

furyl proton signal at 8.16 ppm is better resolved and resonates as a triplet due to long

range coupling to the methylene group protons (Figure 44), rather than a broad singlet

as reported for the natural material. The methylene protons occur as a doublet of

doublets at 4.58 ppm and are further coupled to the adjacent hydroxyl proton, which

itself resonates as a triplet at 5.49 ppm. The four aromatic protons occur as two

multiplets with a non-first order spin pattern in the same region as in the reported

spectrum (Figure 44). A computer simulation of these signals is shown in Figure 45 and

allowed the chemical shifts and coupling constants shown in Table 2 to be extracted.

The signals in the 13C NMR spectrum are also consistent with the values reported for

the natural product, except for the position of the methylene carbon signal of the

hydroxymethyl substituent, which resonates at 52.3 ppm compared to 54.5 ppm

reported48 for the natural material (Figure 43b and Table 2). The chemical shift of the

methylene peak remained at 52.3 ppm during a series of dilution experiments, which

127

Page 139: Synthesis of Some Naturally Occurring Quinones

ruled out the possibility that the position of this signal is concentration dependent.

Interestingly, this shift is identical to the value reported by the same authors48 for the

analogous methylene carbon in the related co-metabolite 205 (Figure 42). Since the

phenolic OH group of 205 is unlikely to have an electronic effect on the CH2OH group,

we believe that the shift of 54.5 ppm reported for naturally occurring 37 may be in error.

O

O

O

O

OHδC 52.3

205

HO

Figure 42

Given the similarities between the spectra of the synthetic material and natural

product, we conclude that the structure 37 has been confirmed by total synthesis.

3.8 Concluding Remarks

The current study has confirmed the novel ring system of 3-

hydroxymethylfuro[3,2-b]naphtho[2,3-d]furan-5,10-dione 37 through total synthesis

and represents a new approach to a compound incorporating the unusual fully aromatic

furo[3,2-b]furan ring. However, only a small amount of the target compound 37 was

obtained due to a number of low yielding steps in the preparation of the key

intermediate furofuranonaphthalene 304. It is clear that this route requires further work

if a sufficient quantity of 37 is to be synthesised for any further investigations into the

natural product’s potentially useful biological activity. Preliminary experiments carried

out during this study suggest that the yield of intermediate 304 is at least partly

128

Page 140: Synthesis of Some Naturally Occurring Quinones

dependent upon the reaction conditions employed in the synthesis of the precursor

acetylenic ether 263. Thus, it seems possible that with further experimentation,

especially in regard to variation of the solvent and base, conditions may be found for

improving the efficiency of the sequence. In particular, the use of an effective

dissociating agent, such as an aprotic dipolar solvent, which is known to promote O-

alkylation may be beneficial for encouraging the synthesis of the initial ether

intermediate 263 and would be worth examining (Scheme 78).

OMe

OMe

OH

O

OMe

OMe

O

O

CO2Et

O

O

OMe

OMe

CO2Et

Br CO2Et

NN N

N

py

py

toluene, reflux

OMe

OMe

O

O

CO2Et

228

233

299

292263

304

heat

base,aprotic dipolar solvent(eg DMF)

Scheme 78

129

Page 141: Synthesis of Some Naturally Occurring Quinones

Figure 43a. 1H NMR spectrum of 37 (500 MHz, d6-DMSO).

Figure 43b. 13C NMR spectrum of 37 (125.75 MHz, d6-DMSO).

130

Page 142: Synthesis of Some Naturally Occurring Quinones

Figure 44. Expansions of the low field region in the 1H NMR spectrum of 37 showing

the aromatic H6-H9 and furyl H2 proton signals.

Figure 45. Computer simulation of the 1H NMR signals for the aromatic H6-H9 protons

in 37 obtained using using the programme gNMR version 4.0 (Cherwell Scientific

Publishing Limited).

131

Page 143: Synthesis of Some Naturally Occurring Quinones

132

Page 144: Synthesis of Some Naturally Occurring Quinones

Chapter 4

Experimental

133

Page 145: Synthesis of Some Naturally Occurring Quinones

4.1 General Details

4.1.1 Solvents and Reagents

All solvents were distilled prior to use. Anhydrous solvents were prepared by

refluxing with the reagents shown in Table 3, followed by distillation under an

atmosphere of nitrogen or argon. Ether refers to diethyl ether and light petroleum refers

to the hydrocarbon fraction which distills from 65-70 °C.

Table 3. Drying agents for solvents

Solvent

Drying Agent

acetone

acetonitrile

dichloromethane

dimethylformamide

dioxane

dimethyl sulfoxide

ether

hexane

methanol

2-propanol

tetrahydrofuran

toluene

potassium carbonate

phosphorus pentoxide

calcium hydride

4Å molecular sieves

sodium benzophenone ketyl

calcium hydride

sodium benzophenone ketyl

calcium hydride

3Å molecular sieves

calcium hydride

potassium benzophenone ketyl

calcium hydride

Most liquid reagents were distilled prior to use. A number of liquid reagents,

which were used in reactions carried out under anhydrous conditions, were dried by

134

Page 146: Synthesis of Some Naturally Occurring Quinones

refluxing with the reagents listed in Table 4, followed by distillation under an

atmosphere of nitrogen or argon.

Table 4. Drying agents for liquid reagents

Liquid Reagent Drying Agent

benzyl alcohol

ethyl chloroformate

propylene oxide

trichloroethylene

triethylamine

potassium carbonate, then 4Å molecular

sieves

potassium carbonate

potassium carbonate

potassium carbonate

calcium hydride

Fremy’s salt was prepared according to the method described by Goodgame,198

then stored in a desiccator over calcium oxide and ammonium carbonate.

4.1.2 Reactions and Chromatography

Reaction temperatures refer to bath temperatures. Kugelrohr distillation

temperatures refer to the oven temperature. Unless otherwise stated, all organic extracts

were dried over anhydrous magnesium sulfate. Solvents were evaporated using a rotary

evaporator with minimal heating.

Ozonolyses were performed with a Welsbach model T-408 ozonator. Irradiation

of reaction solutions was carried out through Pyrex in an Oliphant photochemical

chamber reactor equipped with Sylvania F 815/BLB tubes emitting at 350 nm.

135

Page 147: Synthesis of Some Naturally Occurring Quinones

Analytical thin layer chromatography (TLC) was carried out using Merk silica

gel 60 F254 aluminium-backed plates that were visualised under UV light (254 nm) and

by spraying the plates with a 6% ceric sulfate in 2 M sulfuric acid solution, followed by

heating (> 200 ºC). Silica gel filtrations were carried out under water aspirator vacuum

using either Fluka Kieselgel 60 or Merk silica gel 60 as adsorbent on a sintered glass

funnel. Radial chromatography was carried out using a Chromatotron model 7924T

(Harrison Research, Palo Alto, California) using plates coated with Kieselgel 60 PF254

gipshaltig (Merk Art. 7749). For both techniques, increasing proportions of ethyl acetate

in light petroleum were used as eluent and fractions were monitored by TLC.

4.1.3 Characterisation

Melting points were determined on a Kofler hot stage apparatus and are

uncorrected. Microanalyses were performed by MHW Laboratories, Phoenix, Arizona.

Optical rotations were measured with a Perkin Elmer 141 polarimeter in a microcell (1

ml, 10 cm path length).

Nuclear Magnetic Resonance (NMR) spectra were recorded with Varian Gemini

(200 MHz, 1H), Brucker ARX300 (300 MHz, 1H; 75.5 MHz, 13C), Brucker AV500 (500

MHz, 1H, 125.7 MHz, 13C) spectrometers. Chemical shifts are expressed in ppm relative

to CHCl3 (1H, 7.26 ppm), CDCl3 (13C, 77.0 ppm), D3CSOCD2H (1H, 2.49 ppm) or

D3CSOCD3 (13C, 39.5 ppm) as appropriate; J values are given in Hertz (Hz). Routine

assignments of 13C signals were made with the assistance of DEPT 135 and DEPT 90

experiments and full assignments of the 1H and 13 C signals of (±)-isoelecanacin 151

were derived from HMBC, HSQC and NOESY experiments performed using the

Brucker AV500 spectrometer.

136

Page 148: Synthesis of Some Naturally Occurring Quinones

Mass spectra were recorded in the EI mode using a VG Autospec instrument,

except where specified. Only molecular ion peaks and peaks with intensities registering

15% (relative to base peak) and greater are reported. Infrared spectra were measured

using a Perkin Elmer Spectrum One FT-IR spectrometer with absorption recorded in

terms of frequency (vmax) in cm-1. Electronic spectra were recorded using a Milton Roy

Array 3000 Spectrophotometer and are reported in terms of wavelength (λ) in nm.

High Performance Liquid Chromatography (HPLC) was performed using a

Chiracel OD column (Daicel Chemical Industries) fitted to an ICI 1110 pump interfaced

with a Hewlett Packard Series 1050 instrument using UV detection at 254 nm. The

solvent was 2-propanol-hexane 1 : 5 with a flow rate of 0.5 mL min-1.

137

Page 149: Synthesis of Some Naturally Occurring Quinones

4.2 Experimental for Chapter 2

5-Methoxy-1-(2-propenyloxy)naphthalene 42

The procedure was adapted from that described by Eisenhuth and Schmid.59 5-

Methoxy-1-naphthol 41 (3.51 g, 20.2 mmol), allyl bromide (2.8 ml, 3.88 g, 32.1 mmol)

and potassium carbonate (4.28 g, 31.0 mmol) in acetone (90 ml) were refluxed for 3 h

under nitrogen. The reaction mixture was allowed to cool to room temperature, left to

stand overnight and then poured into water (450 ml) and extracted with ether (3 x 60

ml). The combined organic extracts were washed successively with 10% sodium

hydroxide solution (50 ml), water (50 ml) and brine (50 ml), then dried and evaporated

to give 5-methoxy-1-(2-propenyloxy)naphthalene 42 as a yellow crystalline solid (4.17

g, 97%), mp 96-98 °C (lit.,59 98 °C). δ H (200 MHz, CDCl3) 7.91 (1H, d, J 8.5, ArH),

7.85 (1H, d, J 8.5, ArH), 7.43-7.32 (2H, m, 2 x ArH), 6.86 (2H, d, J 7.7, 2 x ArH), 6.19

(1H, ddt, J 17.3, 10.5 and 5.1, CH, vinylic), 5.53 (1H, dtd, J 17.3, 1.5 and 1.5, CH,

vinylic), 5.34 (1H, dtd, J 10.5, 1.5 and 1.5, CH, vinylic), 4.72 (2H, ddd, J 5.1, 1.5 and

1.5, CH2), 4.00 (3H, s, OCH3).

5-Methoxy-2-(2-propenyl)naphthalen-1-yl acetate 147

A stirred solution of 5-methoxy-1-(2-propenyloxy)naphthalene 42 (4.17 g, 19.5 mmol)

in acetic anhydride (31.0 ml, 33.5 g, 328 mmol) and N,N-diethylaniline (100 ml, 93.3 g,

625 mmol) was heated at 165-170 °C (bath) for 6 h under argon, then allowed to cool to

room temperature and left to stir for 24 h. The solution was diluted with water (250 ml)

and extracted with ether (3 x 60 ml). The combined ether extracts were washed with 2

M hydrochloric acid solution (3 x 100 ml), followed by saturated sodium carbonate

solution (80 ml) and brine (80 ml), dried and evaporated to give 5-methoxy-2-(2-

propenyl)naphthalen-1-yl acetate 147 as a yellow-brown oil (4.74 g, 95%), which was

138

Page 150: Synthesis of Some Naturally Occurring Quinones

pure by 1H NMR. δH (200 MHz, CDCl3) 8.11 (1H, d, J 8.8, ArH), 7.46-7.29 (3H, m,

ArH), 6.80 (1H, d, J 7.4, ArH), 6.06-5.82 (1H, m, CH, vinylic), 5.18-5.04 (2H, m, 2 x

CH, vinylic), 3.97 (3H, s, OCH3), 3.43 (2H, ddd, J 6.6, 1.5 and 1.5, CH2), 2.45 (3H, s,

CH3).

5-Methoxy-2-(2-propenyl)naphthalene-1,4-dione 44

The procedure was adapted from that of Eisenhuth and Schmid.59 5-Methoxy-1-

naphthol 41 (4.21 g, 24.2 mmol), allyl bromide (3.4 ml, 4.6 g, 38 mmol) and potassium

carbonate (5.14 g, 37.2 mmol) in acetone (110 ml) were refluxed for 3 h under nitrogen.

The reaction mixture was allowed to cool, then poured into water and extracted with

ether (3 x 60 ml). The extracts were washed with 10% aqueous sodium hydroxide

solution (40 ml) followed by brine, dried and evaporated to give 5-methoxy-1-(2-

propenyloxy)naphthalene 42 as a pale brown solid (4.89 g, 94%), which was pure by 1H

NMR. δ H (200 MHz, CDCl3) 7.91 (1H, d, J 8.5, ArH), 7.85 (1H, d, J 8.5, ArH), 7.43-

7.32 (2H, m, 2 x ArH), 6.86 (2H, d, J 7.7, 2 x ArH), 6.19 (1H, ddt, J 17.3, 10.5, 5.1,

CH, vinylic), 5.53 (1H, dtd, J 17.3, 1.5, 1.5, CH, vinylic), 5.34 (1H, dtd, J 10.5, 1.5, 1.5,

CH, vinylic), 4.72 (2H, ddd, J 5.1, 1.5, 1.5, CH2), 4.00 (3H, s, OCH3). The foregoing

ether 42 (4.89 mg, 22.9 mmol) was heated at 160-185 °C (oil bath) for 2 h 15 min under

argon, affording almost pure 5-methoxy-2-(2-propenyl)naphthalen-1-ol 43 as a pale

brown waxy solid (4.62 g). δ H (200 MHz, CDCl3) 7.80 (1H, d, J 8.5, ArH), 7.73 (1H,

d, J 8.5, ArH), 7.38 (1H, dd, J 8.5, 7.7, ArH, 7.21 (1H, d, J 8.5, ArH), 6.81 (1H, d, J

7.7, ArH), 6.18-5.99 (1H, m, CH, vinylic), 5.50 (1H, s, br, OH), 5.30-5.20 (2H, m, 2 x

CH, vinylic), 3.99 (3H, s, OCH3), 3.58 (2H, ddd, J 6.2, 1.6, 1.6, CH2). A solution of the

phenol 43 (1.49 g, 6.96 mmol) in ether (30 ml) was added to a separating funnel

containing Fremy’s salt (5.23 g, 19.5 mmol) dissolved in an aqueous borax buffer

solution (0.025 M sodium tetraborate, 148 ml; 0.1 M sodium hydroxide, 72 ml). The

139

Page 151: Synthesis of Some Naturally Occurring Quinones

resulting mixture was shaken until TLC indicated that all the starting material had been

consumed (ca 1.5 h). Argon was bubbled through the solution to evaporate most of the

ether, during which time a yellow-brown precipitate separated. The solid was collected

and dried over phosphorus pentoxide to give 5-methoxy-2-(2-propenyl)naphthalene-1,4-

dione 44 (1.44 g, 91%), which was pure by 1H NMR and used in the next reaction

without further purification. A sample recrystallised from dichloromethane-light

petroleum as yellow needles, mp 95-96 °C (lit.,59 96-97 °C) δ H (200 MHz, CDCl3)

7.76-7.57 (2H, m, ArH), 7.26 (1H, d, J 7.7, ArH), 6.68 (1H, s, 3-CH, vinylic), 5.96-5.76

(1H, m, CH, vinylic), 5.22-5.12 (2H, m, 2 x CH, vinylic), 3.98 (3H, s, OCH3), 3.58

(2H, ddd, J 6.8, 1.9 and 1.3, CH2).

5-Methoxy-2-(2-formylmethyl)naphthalen-1-yl acetate 150

Ozone was bubbled through a solution of 5-methoxy-2-(2-propenyl)naphthalen-1-yl

acetate 147 (3.87 g, 15.1 mmol) in dichloromethane/methanol (4:1, 180 ml) at –78 °C

until TLC indicated that all the starting material had been consumed (45 min). The

solution did not turn blue. Oxygen, followed by nitrogen was bubbled through the

solution in order to displace the ozone. The cold solution was added dropwise to a

stirred suspension of thiourea (1.41 g, 18.5 mmol) and sodium bicarbonate (826 mg,

9.84 mmol) in dichloromethane (50 ml) at ice bath temperature and stirred for 1.5 h

under nitrogen. The resulting reaction mixture was diluted with water (350 ml) and the

organic layer was separated. The aqueous layer was extracted with dichloromethane (2

x 40 ml) and the combined organic extracts were washed with brine, dried and

evaporated to give the aldehyde 150 as a yellow oil (4.03 g), which was refrigerated

overnight. δH (200 MHz, CDCl3) 9.69 (1H, t, J 2.4, CHO), 8.20 (1H, d, J 8.8, ArH),

7.48-7.30 (3H, m, ArH), 6.85 (1H, d, J 7.5, ArH), 4.00 (3H, s, OCH3), 3.70 (2H, d, J

140

Page 152: Synthesis of Some Naturally Occurring Quinones

2.4, CH2), 2.46 (3H, s, CH3). Due to its instability, the aldehyde was used directly in the

next reaction without purification or further characterisation.

1-(1-Hydroxy-5-methoxynaphthalen-2-yl)propan-2-ol 149

A solution of methyl iodide (6.2 ml, 14.1 g, 99.7 mmol) in anhydrous tetrahydrofuran

(35 ml) was added dropwise to magnesium turnings (2.28 g, 93.8 mmol) at room

temperature under argon. The mixture was stirred and diluted by the dropwise addition

of anhydrous tetrahydrofuran (70 ml). Upon completion of the reaction, the Grignard

reagent was cooled in an ice bath and a solution of crude 5-methoxy-2-(2-

formylmethyl)naphthalen-1-yl acetate 150 (4.03 g, 15.6 mmol) in anhydrous

tetrahydrofuran (100 ml) was added dropwise. Then the resulting mixture was allowed

to warm to room temperature and stirring was continued for a further 2 h. The reaction

mixture was carefully quenched with water (300 ml), acidified with concentrated

hydrochloric acid, and extracted with ethyl acetate (3 x 70 ml). The combined organic

extracts were washed with water (70 ml) followed by brine (70 ml), dried and

evaporated to give 1-(1-hydroxy-5-methoxynaphthalen-2-yl)propan-2-ol 149 as an

orange oil (3.55 g), which was used directly in the next reaction. A small sample was

subjected to radial chromatography. Elution with 10% ethyl acetate-light petroleum

gave a clear oil, which solidified upon refrigeration and recrystallised from

dichloromethane-light petroleum as white plates, mp 83-84 °C (Found: C, 72.4; H, 7.0.

C14H16O3 requires C, 72.4; H, 6.9%). (Found M+•, 232.1103. C14H16O3 requires

232.1099). Mass Spectrum m/z: 232 (M, 25%), 214 (100), 212 (28), 199 (47), 187 (18),

186 (29), 171 (28), 169 (21), 128 (21), 115 (43). δ H (300 MHz, CDCl3) 7.88 (1H, dt, J

8.5 and 0.8, ArH), 7.74 (1 H, dd, J 8.5 and 0.6, ArH), 7.37 (1 H, dd, J 8.5 and 7.7, ArH),

7.12 (1H, d, J 8.5, ArH), 6.80 (1H, d, J 7.7, ArH), 4.36-4.27 (1H, m, CH), 3.99 (3H, s,

OCH3) 3.01 (1H, dd, J 14.7 and 2.5, CH of methylene), 2.90 (1H, dd, J 14.7 and 7.1,

141

Page 153: Synthesis of Some Naturally Occurring Quinones

CH of methylene), 1.27 (3H, d, J 6.2, CH3). δ C (75.5 MHz, CDCl3) 155.1 (C), 150.9

(C), 129.0 (CH), 126.8 (C), 126.0 (C), 125.0 (CH), 118.9 (C), 114.6 (CH), 113.5 (CH),

103.8 (CH), 70.9 (CH), 55.5 (OCH3), 40.5 (CH2), 23.2 (CH3).

2-(2-Hydroxypropyl)-5-methoxynaphthalene-1,4-dione 46

A solution of 1-(1-hydroxy-5-methoxynaphthalen-2-yl)propan-2-ol 149 (3.48 g, 15.0

mmol) in ethyl acetate (80 ml) was added to a separating funnel containing Fremy’s salt

(8.25 g, 30.8 mmol) dissolved in an aqueous borax buffer solution (0.025 M sodium

tetraborate, 250 ml; 0.1 M sodium hydroxide, 121 ml). The resulting mixture was

shaken until TLC indicated that the starting material had been consumed (ca 40 min).

The mixture was diluted with brine (50 ml) and the organic layer was separated. The

aqueous layer was extracted with ethyl acetate (4 x 60 ml). The combined organic

extracts were washed with brine (80 ml), dried and evaporated to give a yellow oil,

which was subjected to silica gel filtration. Elution with 30% ethyl acetate-light

petroleum gave 2-(2-hydroxypropyl)-5-methoxynaphthalene-1,4-dione 46 as a yellow

oil (1.54 g, 41% over 3 steps) which solidified upon refrigeration. A small sample

recrystallised from ethyl acetate-light petroleum as fine yellow needles, mp 95-96 °C

(lit.,59 96-97 °C). (Found M+•, 246.0894. C14H14O4 requires 246.0892). Mass Spectrum

m/z: 246 (M, 23%), 230 (27), 204 (73), 203 (54), 202 (100), 187 (15), 175 (17), 174

(32), 173 (38), 159 (25), 144 (15), 131 (23), 115 (34). δ H (300 MHz, CDCl3) 7.75 (1H,

dd, J 7.7 and 1.2, ArH), 7.66 (1H, dd, J 8.4 and 7.7, ArH), 7.29 (1 H, dd, J 8.4 and 1.1,

ArH), 6.78 (1H, t, J 1.0, CH, vinylic), 4.15-4.03 (1H, m, CH), 4.00 (3H, s, OCH3) 2.74

(1H, ddd, J 13.8, 4.1 and 1.0, CH of methylene), 2.59 (1H, ddd, J 13.8, 8.0 and 1.0, CH

of methylene), 1.29 (3H, d, J 6.2, CH3). δ C (75.5 MHz, CDCl3) 186.1 (C=O), 184.3

(C=O), 159.4 (C), 145.6 (C), 139.2 (CH), 134.7 (CH), 134.3 (C), 119.8 (C), 119.5 (CH),

142

Page 154: Synthesis of Some Naturally Occurring Quinones

117.8 (CH), 66.7 (CH), 56.4 (OCH3), 39.1 (CH2), 23.7 (CH3). λmax (CH2Cl2) (log ε) 247

(4.14), 268 (3.22), 354 (4.06), 396 (3.54). vmax(CH2Cl2) /cm-1 1658 (C=O).

5-Methoxy-2-(2-vinyloxypropyl)naphthalene-1,4-dione 138

A solution of 2-(2-hydroxypropyl)-5-methoxynaphthalene-1,4-dione 46 (1.43 g, 5.80

mmol) and mercuric acetate (351 mg, 1.10 mmol) in ethyl vinyl ether (35 ml, 26.4 g,

366 mmol) and dichloromethane (10 ml) in a foil-covered flask was refluxed for 6 h

under argon. The solution was kept at room temperature for 2 days, then poured into

water and extracted with dichloromethane (3 x 40 ml). The combined organic extracts

were washed with brine, dried and evaporated to give a yellow residue, which was

subjected to silica gel filtration. Elution with 15% ethyl acetate-light petroleum gave 5-

methoxy-2-(2-vinyloxypropyl)naphthalene-1,4-dione 138 (903 mg, 57%) as a yellow

oil. (Found M+•, 272.1054. C16H16O4 requires 272.1049). Mass Spectrum m/z: 273

(M+1, 17%), 272 (M, 97), 243 (16), 230 (34), 229 (100), 228 (36), 227 (18), 215 (15),

213 (31), 211 (19), 205 (19), 202 (27), 201 (19), 188 (27), 187 (36). δ H (300 MHz,

CDCl3) 7.74 (1H, dd, J 7.7 and 1.2, ArH), 7.66 (1H, dd, J 8.3 and 7.7, ArH), 7.28 (1 H,

dd, J 8.3 and 1.2, ArH), 6.75 (1H, t, J 1.1, CH, vinylic), 6.27 (1H, dd, J 6.7 and 14.2,

CH, vinylic), 4.30 (1H, dd, J 14.2 and 1.7, CH, vinylic), 4.26-4.15 (1H, m, CH), 4.00-

3.99 (1H, dd, J 6.7, 1.7, CH, vinylic), 3.99 (3H, s, OCH3) 2.79 (1H, ddd, J 13.9, 7.4 and

1.1, CH of methylene), 2.67 (1H, ddd, J 13.9, 5.3 and 1.1, CH of methylene), 1.27 (3H,

d, J 6.2, CH3). δ C (75.5 MHz, CDCl3) 185.4 (C=O), 184.3 (C=O), 159.4 (C), 150.3

(CH), 144.7 (C), 139.3 (CH), 134.7 (CH), 134.2 (C), 119.4 (CH), 117.7 (CH), 88.8

(CH2), 77.2 (C), 73.2 (CH), 56.4 (OCH3), 36.1 (CH2), 20.0 (CH3). λmax (CH2Cl2) (log ε)

230 (4.10), 238 (4.15), 262 (4.06), 333 (3.22), 249 (3.24), 393 (3.46).vmax (CH2Cl2)/cm-1

1658 (C=O).

143

Page 155: Synthesis of Some Naturally Occurring Quinones

(±)-Elecanacin 36 and (±)-isoelecanacin 151

A deoxygenated solution of 5-methoxy-2-(2-vinyloxypropyl)naphthalene-1,4-dione 138

(122 mg, 0.45 mmol) in anhydrous dichloromethane (50 ml) was irradiated at 350 nm

through Pyrex for 65 min, when TLC indicated that all the starting material had been

consumed. The solvent was evaporated and the yellow residue was subjected to careful

radial chromatography. Elution with 20% ethyl acetate-light petroleum gave (±)-

isoelecanacin 151 (46 mg, 38%) as a yellow oil, which solidified upon refrigeration and

recrystallised from dichloromethane-light petroleum as white needles, mp 114-115 °C.

(Found M+•, 272.1048. C16H16O4 requires 272.1049). Mass Spectrum m/z: 273 (M+1,

16%), 272 (M, 100), 270 (22), 244 (22), 243 (87), 242 (20), 241 (16), 229 (41), 228

(20), 227 (30), 217 (19), 211 (19), 201 (15), 189 (19), 187 (22), 153 (65), 136 (57), 135

(17), 128 (16), 115 (27), 108 (19), 107 (71), 106 (55), 105 (39), 104 (17), 90 (19), 89

(96). The 13C and 1H NMR spectral data are given in Table 1 (p. 60). vmax (CH2Cl2)/cm-1

1683 (C=O). Analysis on the Chiracel OD column showed two peaks in a ratio of 48 :

52 at retention times 19.2 and 22.6 min. Further elution with 30% ethyl acetate-light

petroleum gave (±)-elecanacin 36 (31 mg, 25%) as a yellow oil, which could not be

induced to crystallise (lit.,47 mp 198 °C for optically active material). (Found M+•,

272.1050. C16H16O4 requires 272.1049). Mass Spectrum m/z: 273 (M+1, 17%), 272 (M,

100), 244 (22), 243 (88), 229 (30), 228 (19), 227 (21), 217 (15), 215 (15), 211 (15), 202

(16), 201 (15), 189 (17), 187 (17), 135 (18), 128 (16), 115 (23). The 13C and 1H NMR

spectral data are shown in Table 1 and are identical with those of natural elecanacin.

vmax (CH2Cl2)/ cm-1 1684 (C=O). Analysis on the Chiracel OD column showed two

peaks in the ratio of 52 : 48 at retention times 26.5 and 39.2 min.

When the irradiation was repeated and the reaction was interrupted at low

conversion of reactant, elecanacin 36 and isoelecanacin 151 were found to be present in

144

Page 156: Synthesis of Some Naturally Occurring Quinones

the same ratio as at complete conversion (TLC and NMR analysis). Irradiation of pure

samples of elecanacin 36 and isoelecanacin 151 in dichloromethane led to no change.

Conversion of the vinyl ether 138 into elecanacin 36 and isoelecanacin 151 also

was observed when a solution of 138 in dichloromethane was kept in ambient

laboratory light.

2-(2,3-Epoxypropyl)-5-methoxynaphthalen-1-yl acetate 155

a) A cold solution of an excess of dimethyldioxirane in acetone (60 ml), prepared

according to the procedure described by Murray and Singh,123 was added to a solution

of 5-methoxy-2-(2-propenyl)naphthalen-1-yl acetate 147 (1.16 g, 453 mmol) in acetone

(10 ml) and left to stir for 16 h at room temperature. The solution was diluted with

water (350 ml) and extracted with ethyl acetate (3 x 70 ml). The extracts were washed

with brine, dried and evaporated to give a brown oil, which was subjected to silica gel

filtration. Elution with 5% ethyl acetate-light petroleum returned unreacted starting

material as a yellow oil (89 mg). Further elution with 10% ethyl acetate-light petroleum

gave a yellow oil, which was subjected to radial chromatography. Elution with 5% ethyl

acetate-light petroleum gave the epoxide 155 as a colourless oil (279 mg, 23%). (Found

M+•, 272.1042. C16H16O4 requires 272.1049). Mass spectrum (FAB) m/z: 273 (M + 1,

74%), 272 (M, 100), 231 (32), 230 (88), 213 (64), 212 (32), 187 (35). δH (300 MHz,

CDCl3) 8.15 (1H, d, J 8.7, ArH), 7.44-7.39 (2H, m, ArH), 7.30 (1H, d, J 8.5, ArH), 6.82

(1H, d, J 7.1, ArH), 3.99 (3H, s, OCH3), 3.22-3.18 (1H, m, CH, X of ABX), 3.01 (1H,

dd, J 14.6 and 5.5, CH of methylene), 2.88-2.79 (2H, m, 2 x CH of methylene), 2.59

(1H, m, CH of methylene), 2.49 (3H, s, CH3). δC (75.5 MHz, CDCl3) 169.3 (C=O),

155.6 (C), 144.3 (C), 128.1 (C), 127.0 (CH), 126.8 (CH), 126.4 (C), 125.9 (C), 120.5

(CH), 113.2 (CH), 104.1 (CH), 55.6 (OCH3), 51.4 (CH), 47.0 (CH2), 33.5 (CH2), 20.7

(CH3).

145

Page 157: Synthesis of Some Naturally Occurring Quinones

b) m-Chloroperoxybenzoic acid (70%, 1.91 g, 7.75 mmol) and 2,6-di-t-butyl-4-

methylphenol (20 mg, 0.09 mmol) were added to a stirred solution of 5-methoxy-2-(2-

propenyl)naphthalen-1-yl acetate 147 (902 mg, 3.52 mmol) in dichloromethane (65 ml)

and the solution was refluxed gently for 1.5 h. The solution was allowed to cool to room

temperature and left to stand overnight, during which time a white precipitate separated.

The mixture was cooled in an ice bath and filtered. The filtrate was washed successively

with 10% sodium bisulfite solution (25 ml), 10% sodium bicarbonate solution (2 x 40

ml) and brine (60 ml), then dried and evaporated to give an orange residue, which was

subjected to radial chromatography. Elution with 10% ethyl acetate-light petroleum

returned unreacted starting material (56 mg). Further elution afforded a fraction, which

was concentrated under reduced pressure to give a yellow oil (114 mg).1H NMR

analysis revealed that the oil contained impure epoxide 155. Further attempts to purify

this fraction by radial chromatography were unsuccessful and the epoxide could not be

isolated.

1-Benzyloxy-5-methoxy-2-(2-propenyl)naphthalene 166

Benzyl bromide (6.13 g, 35.8 mmol) was added to a mechanically stirred suspension of

5-methoxy-2-(2-propenyl)naphthalen-1-ol 43 (6.40 g, 29.9 mmol) and potassium

carbonate (6.52 g, 47.2 mmol) in acetone (200 ml) and the mixture was refluxed for 3 h

under argon. The mixture was diluted with water (400 ml) and extracted with ether (3 x

60 ml). The combined organic extracts were washed with 10% sodium hydroxide

solution (60 ml), followed by brine (70 ml), dried and evaporated to give a yellow oil

(9.70 g). Kugelrohr distillation under a vacuum gave 1-benzyloxy-5-methoxy-2-(2-

propenyl)naphthalene 166 (7.63 g, 84%) as a pale yellow oil, which solidified upon

refrigeration. A sample recrystallised from light petroleum as white plates, mp 46-47

°C. (Found M+•, 304.1455. C21H20O2 requires 304.1463). Mass spectrum m/z: 304 (M,

146

Page 158: Synthesis of Some Naturally Occurring Quinones

60%), 214 (18), 213 (100), 198 (20), 153 (20), 115 (18), 91 (90), 77 (23). δH (300 MHz,

CDCl3) 8.03 (1H, d, J 8.7, ArH), 7.73 (1H, d, J 8.5, ArH), 7.59-7.56 (2H, m, ArH),

7.48-7.32 (5H, m, ArH), 6.82 (1H, d, J 7.3, ArH), 6.10-5.96 (1H, m, CH, vinylic), 5.14-

5.06 (2H, m, 2 x CH, vinylic), 5.02 (2H, s, OCH2), 4.01 (3H, s, OCH3), 3.61 (2H, ddd, J

5.0, 1.4 and 1.4, CH2). δC (75.5 MHz, CDCl3) 155.8 (C), 151.8 (C), 137.5 (C), 137.2

(CH), 129.4 (C), 129.1 (C), 128.6 (CH), 128.0 (CH), 127.7 (CH), 127.5 (CH), 126.0

(CH), 125.9 (C), 118.2 (CH), 115.9 (CH2), 114.4 (CH), 103.6 (CH), 76.2 (CH2), 55.5

(OCH3), 34.0 (CH2).

2-(1-Benzyloxy-5-methoxynaphthalen-2-yl)ethanal 167

Ozone was bubbled through a solution of 1-benzyloxy-5-methoxy-2-(2-

propenyl)naphthalene 166 (2.31 g, 7.60 mmol) in dichloromethane/methanol (4 : 1, 180

ml) at –78 °C, until TLC indicated that the starting material had been consumed (ca 20

min). The solution did not turn blue. Oxygen, followed by argon was bubbled through

the solution in order to displace the ozone. The cold solution was added dropwise to a

stirred suspension of thiourea (687 mg, 9.02 mmol) and sodium bicarbonate (457 mg,

5.44 mmol) in dichloromethane (60 ml) at ice bath temperature and the resulting

mixture was stirred for 1.5 h under argon. The reaction mixture was then diluted with

water (300 ml), the organic layer was separated and the aqueous layer was extracted

with dichloromethane (2 x 60 ml). The combined organic extracts were washed with

brine (80 ml), dried and evaporated to give a yellow oil, which was subjected to silica

gel filtration. Elution with 5% ethyl acetate-light petroleum gave the aldehyde 167 as a

faint yellow oil (1.51 g, 65%). (Found M+•, 306.1251. C20H18O3 requires 306.1256).

Mass spectrum m/z: 306 (M, 24%), 288 (54), 215 (43), 198 (26), 187 (54), 155 (25), 149

(29), 127 (19), 115 (22), 91 (100), 84 (20), 83 (17), 77 (18), 71 (20), 69 (21), 57 (36). δH

(300 MHz, CDCl3) 9.70 (1H, t, J 2.2, CHO), 8.09 (1H, dd, J 8.6 and 0.4, ArH), 7.73

147

Page 159: Synthesis of Some Naturally Occurring Quinones

(1H, d, J 8.5, ArH), 7.50-7.36 (6H, m, ArH), 7.26 (1H, d, J 8.6, ArH), 6.86 (1H, d, J

7.3, ArH), 5.02 (2H, s, OCH2), 4.02 (3H, s, OCH3), 3.78 (2H, d, J 2.2, CH2). δC (75.5

MHz, CDCl3) 199.7 (C=O), 155.9 (C), 152.9 (C), 136.8 (C), 129.2 (C), 128.6 (CH),

128.3 (CH), 128.0 (CH), 127.4 (CH), 126.8 (C), 126.5 (CH), 122.1 (C), 118.9 (CH),

114.3 (CH), 104.2 (CH), 76.2 (CH2), 55.6 (OCH3), 45.3 (CH2).

1-Benzyloxy-2-(2,3-epoxypropyl)-5-methoxynaphthalene 156

Trimethylsulfoxonium iodide (999 mg, 4.54 mmol) was added portionwise over 20 min

to sodium hydride (60% oil dispersion, 186 mg, 4.65 mmol) in anhydrous dimethyl

sulfoxide (4 ml) under argon and the resulting suspension was stirred for 30 min. 2-(1-

Benzyloxy-5-methoxynaphthalen-2-yl)ethanal 167 (526 mg, 1.72 mmol) in anhydrous

dimethyl sulfoxide (4 ml) was added dropwise and the resulting mixture was stirred at

room temperature for 1 h 45 min. The mixture was quenched with ice, diluted with

water (80 ml) and extracted with ether (4 x 40 ml). The combined organic extracts were

washed with brine (50 ml), dried and evaporated to give a yellow oil, which was

subjected to silica gel filtration. Elution with 5% ethyl acetate-light petroleum afforded

the epoxide 156 as a colourless oil (383 mg, 69%). (Found M+•, 320.1422. C21H20O3

requires 320.1412). Mass spectrum m/z: 321 (M+1, 20%), (M, 89), 230 (23), 229 (28),

201 (15), 199 (57), 187 (22), 186 (24), 171 (37), 155 (20), 128 (20), 127 (22), 115 (23),

91 (100), 77 (24), 69 (17), 57 (23). δH (300 MHz, CDCl3) 8.04 (1H, d, J 8.9, ArH), 7.72

(1H, d, J 8.5, ArH), 7.57-7.39 (7H, m, ArH), 6.83 (1H, d, J 7.3, ArH), 5.05 (2H, s,

OCH2), 4.01 (3H, s, OCH3), 3.24-3.18 (1H, m, CH, X of ABX), 3.10 (1H, dd, J 14.3

and 5.3, CH of methylene), 3.00 (1H, dd, J 14.3 and 5.4, CH of methylene), 2.85 (1H,

m, CH of methylene), 2.60 (1H, m, CH of methylene). δC (75.5 MHz, CDCl3) 155.8 (C),

152.3 (C), 137.4 (C), 129.3 (C), 128.6 (CH), 128.1 (CH), 127.7 (CH), 127.4 (CH),

148

Page 160: Synthesis of Some Naturally Occurring Quinones

126.7 (C), 126.3 (C), 126.2 (CH), 118.5 (CH), 114.3 (CH), 103.8 (CH), 76.3 (CH2),

55.6 (OCH3), 52.1 (CH), 47.1 (CH2), 32.9 (CH2).

2-(2,3-Epoxypropyl)-5-methoxynaphthalene-1,4-dione 168

m-Chloroperoxybenzoic acid (70 %, 278 mg, 1.13 mmol) was added to a stirred

solution of 5-methoxy-2-(2-propenyl)naphthalen-1,4-dione 44 (212 mg, 0.93 mmol) in

dichloromethane (10 ml) at ice bath temperature under argon. After 3 h, the mixture was

allowed to warm to room temperature and left to stir for a further 21 h. The mixture was

diluted with dichloromethane (50 ml) and washed with 2.5% sodium bisulfite solution

(20 ml), followed by saturated sodium bicarbonate solution (2 x 25 ml) and brine (30

ml), dried and evaporated to give a yellow-orange solid (257 mg), which was subjected

to radial chromatography. Elution with 40% ethyl acetate-light petroleum returned

unreacted starting material (21 mg). Further elution with 70% ethyl acetate-light

petroleum gave the epoxide 168 as an orange crystalline solid (152 mg, 67%, 74%

based on recovered starting material). A sample recrystallised from ethyl acetate-light

petroleum as orange plates, mp 159-160 °C. (Found: C, 68.9; H, 5.0. C14H12O4 requires

C, 68.85; H, 4.95%). (Found M+•, 244.0731. C14H12O4 requires 244.0736). Mass

spectrum m/z: 246 (M + 2, 35%), 245 (M + 1, 16), 244 (M, 100), 228 (15), 227 (23),

216 (27), 215 (42), 214 (32), 213 (32), 202 (73), 201 (52), 200 (31), 199 (24), 198 (43),

197 (29), 187 (30), 186 (117), 185 (56), 184 (20), 183 (31), 174 (15), 173 (45), 171

(23), 169 (24), 168 (24), 159 (21), 157 (48), 156 (26), 155 (25), 145 (22), 144 (21), 143

(21), 141 (22), 133 (16), 131 (20), 129 (42), 128 (90), 127 (39), 116 (25), 115 (77), 105

(23), 104 (39), 102 (20), 91 (18), 89 (16), 77 (29), 76 (71), 75 (21), 64 (15), 63 (26). δH

(300 MHz, CDCl3) 7.75 (1H, dd, J 7.6 and 1.1, ArH), 7.67 (1H, dd, J 8.3 and 7.6, ArH),

7.30 (1H, dd, J 8.3 and 1.1, ArH), 6.84 (1H, t, J 1.2, 3-CH, vinylic), 4.00 (3H, s, OCH3),

3.22-3.16 (1H, m, CH, X of ABX), 2.91-2.82 (2H, m, CH of methylene and CH of

149

Page 161: Synthesis of Some Naturally Occurring Quinones

oxirane methylene), 2.66-2.57 (2H, m, CH of methylene and CH of oxirane methylene).

δC (75.5 MHz, CDCl3) 185.1 (C=O), 184.1 (C=O), 159.5 (C), 144.4 (C), 138.5 (CH),

134.8 (CH), 134.2 (C), 119.8 (C), 119.4 (CH), 117.9 (CH), 56.5 (OCH3), 50 (CH), 47

(CH2), 32.2 (CH2). vmax (solution, CH2Cl2) 1660 cm-1 (C=O).

Attempted resolution of epoxides 155, 156 and 168

Attempted hydrolytic kinetic resolution of the epoxides 155, 156 and 168 with

Jacobsen’s catalyst under standard conditions116, 118 in each case returned starting

material exhibiting no significant optical rotation.

5-Methoxy-1-methoxymethoxynaphthalene 173

This was prepared as described128 and obtained in 86% yield as colourless needles, mp

74-75 °C (lit.,128 75-76 °C). δH (200 MHz, CDCl3) 7.91 (2H, m, ArH), 7.45-7.35 (2H,

m, ArH), 7.14 (1H, d, J 7.6, ArH), 6.86 (1H, d, J 7.6, ArH), 5.40 (3H, s, CH2), 4.01 (3H,

s, OCH3), 3.56 (3H, s, CH3).

(R)-propylene oxide 158

The procedure described by Jacobsen and coworkers was followed.116, 118 The

precatalyst, (R,R)-N,N’-bis(3,5-di-tert-butylsalicalidene)-1,2-

cyclohexanediaminocobalt(II) (604 mg, 1.00 mmol) and glacial acetic acid (123 mg,

2.05 mmol) in toluene (5 ml) were stirred for 2 h at room temperature while exposed to

the atmosphere. The solvent was evaporated under reduced pressure leaving the crude

(R,R)-(salen)Co(III)(OAc) catalyst 160 as a dark brown residue. Anhydrous propylene

oxide (35 ml, 29.0 g, 499 mmol) was added and the resulting solution was cooled in an

ice-water bath. Water (4.90 ml, 272 mmol) was added slowly to the stirring solution,

ensuring the temperature stayed between 15-20 °C. After the addition was complete (ca

150

Page 162: Synthesis of Some Naturally Occurring Quinones

0.5 h), the solution was allowed to warm to room temperature and left to stir for 18 h.

Distillation gave (R)-propylene oxide 158 bp 35 °C (12.0 g, 41%), [α]D20 + 14.0 (neat)

(lit.,199 [α]D31

+ 13.8 (neat)), followed by 1,2-propanediol bp 46 °C (0.7 mm Hg) (16.9

g, 45%). The remaining residue was diluted with methanol and the red precatalyst (563

mg) was recovered by filtration.

(2R)-1-(5-Methoxy-1-methoxymethoxynaphthalen-2-yl)propan-2-ol 176

n-Butyllithium in hexane (1.6 M, 5.4 ml, 8.6 mmol) was added dropwise to a stirred

solution of 5-methoxy-1-methoxymethoxynaphthalene 173 (1.43 g, 6.56 mmol) in

anhydrous tetrahydrofuran (30 ml) cooled in an ice-water bath under argon. After 2 h

the mixture was treated with hexamethylphosphoramide (3.54 ml, 20.3 mmol), followed

by the immediate dropwise addition of (R)-propylene oxide 158 (491 mg, 8.45 mmol) in

anhydrous tetrahydrofuran (2 ml). The resulting yellow solution was allowed to warm to

room temperature and left to stir for 17 h. The solution was diluted with saturated

ammonium chloride (70 ml) and extracted with ether (3 x 50 ml). The combined organic

extracts were washed with water (40 ml), followed by brine (50 ml), dried and

evaporated to give a yellow oil (2.18 g), which was subjected to silica gel filtration.

Elution with 1% ethyl acetate-light petroleum returned unreacted starting material as a

white crystalline solid (293 mg). Further elution with 10% ethyl acetate-light petroleum

gave (2R)-1-(1-methoxy-5-methoxymethoxynaphthalen-2-yl)propan-2-ol 177 (66 mg)

as a colourless oil, which became a white crystalline solid upon refrigeration and

recrystallised from dichloromethane-light petroleum as white plates, mp 74-75 °C.

[α]D21

- 26.9 (c 0.015 in CH2Cl2). (Found M+•, 276.1366. C16H20O4 requires 276.1362).

Mass spectrum m/z: 276 (M, 41%), 258 (36), 226 (18), 214 (56), 212 (19), 201 (15), 200

(40), 199 (34), 198 (17), 187 (23), 185 (24), 171 (23), 169 (15), 157 (17), 155 (24), 153

(20), 149 (37), 141 (18), 139 (16), 129 (18), 128 (31), 127 (23), 121 (15), 115 (37), 109

151

Page 163: Synthesis of Some Naturally Occurring Quinones

(17), 107 (20), 105 (17), 98 (16), 97 (24), 96 (16), 95 (29), 93 (18), 91 (19), 86 (49), 85

(20), 84 (78), 83 (38), 82 (23), 81 (45), 79 (19), 78 (40), 77 (33), 71 (36), 70 (33), 69

(100), 68 (24), 67 (34), 63 (62), 61 (15), 60 (32), 57 (67), 56 (34). δH (300 MHz,

CDCl3) 8.03 (1H, dd, J 8.6 and 0.4, ArH), 7.73 (1H, dt, J 8.4 and 0.8, ArH), 7.44 (1H

dd, J 8.4 and 7.7, ArH), 7.33 (1H, d, J 8.6, ArH), 7.09 (1H, dd, J 7.7 and 0.8, ArH), 5.39

(2H, s, OCH2O), 4.16 (1H, m, CH), 3.94 (3H, s, OCH3), 3.54 (3H, s, OCH3), 2.96 (2H,

d, J 6.2, CH2), 2.29 (1H, s, br, OH), 1.28 (3H, d, J 6.2, CH3). δC (75.5 MHz, CDCl3)

153.8 (CO), 153.2 (CO), 129.2 (C), 128.2 (CH), 127.5 (C), 126.5 (C), 126.1 (CH),

118.3 (CH), 115.4 (CH), 107.7 (CH), 94.6 (OCH2O), 68.6 (CH), 61.8 (OCH3), 56.2

(OCH3), 39.9 (CH2), 23.2 (CH3). Further elution with 10% ethyl acetate-light petroleum

gave a fraction containing a 15 : 85 mixture of (2R)-1-(1-methoxy-5-

methoxymethoxynaphthalen-2-yl)propan-2-ol 177 and the title alcohol (2R)-1-(5-

methoxy-1-methoxymethoxynaphthalen-2-yl)propan-2-ol 176 as a pale yellow oil (1.13

g), which was used directly in the next reaction. The crude oil obtained from a second

reaction was subjected to careful radial chromatography. Elution with 5% ethyl acetate-

light petroleum allowed the isolation of a small analytical sample of (2R)-1-(5-methoxy-

1-methoxymethoxynaphthalen-2-yl)propan-2-ol 176 as a colourless oil. [α]D21

+ 8.7 (c

0.078 in CH2Cl2). (Found M+•, 276.1359. C16H20O4 requires 276.1362). Mass spectrum

m/z: 276 (M, 21%), 244 (24), 215 (30), 214 (100), 199 (41), 187 (49), 115 (20). δH (300

MHz, CDCl3) 8.03 (1H, d, J 8.6, ArH), 7.59 (1H, d, J 8.6, ArH), 7.41 (1H, dd, J 8.6, 7.6

ArH), 7.33 (1H, d, J 8.6, ArH), 6.81 (1H, d, J 7.6, ArH), 5.18 (1H, d, J 5.9, CH), 5.16

(1H, d, J 5.9, CH), 4.23 (1H, m, CH), 3.99 (3H, s, OCH3), 3.70 (3H, s, OCH3), 3.10-

2.94 (2H, m, CH2), 2.32 (1H, d, J 4.6, OH), 1.30 (3H, d, J 6.2, CH3). δC (75.5 MHz,

CDCl3) 155.7 (C), 151.6 (C), 129.4 (C), 128.3 (C), 127.7 (C), 126.3 (CH), 126.1 (CH),

118.8 (CH), 114.2 (CH), 103.8 (CH), 100.3 (CH2), 68.6 (CH), 57.6 (OCH3), 55.6

(OCH3), 40.1 (CH2), 23.6 (CH3).

152

Page 164: Synthesis of Some Naturally Occurring Quinones

(2R)-1-(1-Hydroxy-5-methoxy-2-naphthalenyl)propan-2-ol 152

A 15 : 85 mixture of (2R)-1-(1-methoxy-5-methoxymethoxynaphthalen-2-yl)propan-2-

ol 177 and (2R)-1-(5-methoxy-1-methoxymethoxynaphthalen-2-yl)propan-2-ol 176

(1.13 g) in anhydrous 2-propanol (20 ml) was treated with carbon tetrabromide (141

mg, 0.425 mmol) and refluxed for 1.5 h under argon. The solution was concentrated

under reduced pressure and the resulting yellow oil was subjected to silica gel filtration.

Elution with 5% ethyl acetate-light petroleum gave (2R)-1-(1-hydroxy-5-methoxy-2-

naphthalenyl)propan-2-ol 152 as a white crystalline solid (696 mg, 73%), which

recrystallised from ethyl acetate-light petroleum as white plates, mp 85-86 °C. [α]D21

- 6.7 (c 0.009 in CH2Cl2). The NMR spectral properties of this material were identical

with those of the (2R,S) compound 149 prepared previously.

Treatment of a sample of the (2R,S)-alcohol 149 with the chiral shift reagent

europium tris[3-heptafluoropropylhydroxymethylene)-(+)-camphorate] (7.5 mol%)

separated the H8 proton doublet signal into two doublets in the 1H NMR spectrum. The

enantiomeric excess within the (2R)-alcohol 152 was estimated to be greater than 90%

by examination of the H8 signal in the 1H NMR spectrum of a sample of 152 which had

been treated with 7.5 mol% of the chiral shift reagent.

(2R)-2-(2-Hydroxypropyl)-5-methoxynaphthalene-1,4-dione 178

A solution of (2R)-1-(1-hydroxy-5-methoxy-2-naphthalenyl)propan-2-ol 152 (696 mg,

3.00 mmol) in ether (15 ml) was added to a separating funnel containing Fremy’s salt

(1.718 g, 6.11 mmol) dissolved in an aqueous borax buffer solution (0.025 M sodium

tetraborate, 84 ml; 0.1 M sodium hydroxide, 41 ml). The resulting mixture was shaken

until TLC indicated that the starting material had been consumed (ca 30 min). The

mixture was extracted with chloroform (4 x 50 ml) and the combined organic extracts

were washed with brine (60 ml), dried and evaporated to give (2R)-2-(2-

153

Page 165: Synthesis of Some Naturally Occurring Quinones

hydroxypropyl)-5-methoxynaphthalene-1,4-dione 178 as a yellow crystalline solid (717

mg, 97%), which was pure by 1H NMR. A sample recrystallised from ethyl acetate-light

petroleum as bright yellow needles, mp 116-117 °C. [α]D20 – 25.3 (c 0.015 in CH2Cl2).

The NMR spectral properties of this material were identical with those of the (2R,S)

compound 46 prepared previously.

(2R)-5-Methoxy-2-(2-vinyloxypropyl)naphthalene-1,4-dione 179

A solution of (2R)-2-(2-hydroxypropyl)-5-methoxynaphthalene-1,4-dione 178 (692 mg,

2.81 mmol) and mercuric acetate (212 mg, 0.66 mmol) in ethyl vinyl ether (17 ml, 12.8

g, 178 mmol) and dichloromethane (4.5 ml) in a foil-covered flask was heated under

reflux for 5.5 h under argon. The solution was left to stand at room temperature

overnight and was then diluted with dichloromethane (40 ml) and washed with water

(80 ml). The organic layer was separated and the aqueous layer was extracted with

dichloromethane (3 x 40 ml). The combined organic extracts were washed with brine

(50 ml), dried and evaporated to give a yellow oil, which was subjected to silica gel

filtration. Elution with 15% ethyl acetate-light petroleum gave the vinyl ether 179 as a

yellow oil (393 mg, 51%, 81% based on recovered starting material). [α]D21 - 15.6 (c

0.010 in CH2Cl2). The NMR spectral properties of this material were identical with

those of the (2R,S) compound 138 prepared previously. Further elution with 40% ethyl

acetate-light petroleum returned unreacted starting material as a yellow crystalline solid

(255 mg).

154

Page 166: Synthesis of Some Naturally Occurring Quinones

(-)-Elecanacin 36 and (+)-isoelecanacin 151 A deoxygenated solution of (2R)-5-methoxy-2-(2-vinyloxypropyl)naphthalene-1,4-

dione 179 (131 mg, 0.482 mmol) in anhydrous dichloromethane (50 ml) was irradiated

at 350 nm through Pyrex for 65 min, when TLC indicated that all the starting material

had been consumed. The solvent was evaporated and the yellow residue was subjected

to careful radial chromatography. Elution with 20% ethyl acetate-light petroleum gave

(+)-isoelecanacin 151 (53 mg, 40%) as a yellow crystalline solid, which recrystallised

from dichloromethane-light petroleum as pale yellow plates, mp 138-139 °C. [α]D21 +

110.4 (c 0.010 in CH2Cl2). Analysis on the Chiracel OD column showed two peaks in

the ratio of 1 : 99 at retention times 20.1 and 23.4 min, giving an ee of 98% for this

material. The NMR spectral properties of this sample were identical with those of the

(2R,S) compound prepared previously. Further elution with 30% ethyl acetate-light

petroleum gave (-)-elecanacin 36 as a pale yellow crystalline solid (31 mg, 24%), which

recrystallised from dichloromethane-light petroleum as faint yellow plates, mp 167-168

°C (lit.,47 198 °C for material of low ee). [α]D21

– 145.2 (c 0.004 in CHCl3) (lit.,47 + 20.7

in CHCl3). Analysis on the Chiracel OD column showed a single peak at retention time

26.1 min. Under these conditions <0.5% of the other enantiomer (expected retention

time 39.2 min) could have been detected. The ee of this sample was thus >99%. The

NMR spectral properties of this material were identical with those of (2R,S) compound

prepared previously.

155

Page 167: Synthesis of Some Naturally Occurring Quinones

4.3 Experimental for Chapter 3

2,3-Dibromonaphthalene-1,4-dione 284

A stirred suspension of 1,4-naphthoquinone (12.8 g, 81.0 mmol), sodium acetate (65.0

g, 793 mmol) and bromine (12.8 ml, 39.8 g, 249 mmol) in acetic acid (320 ml) was

heated at 110 °C (bath) for 2 h under nitrogen, then allowed to cool room temperature

and left to stir for a further 16 h. The reaction mixture was poured into water (1 L) and a

yellow precipitate separated. The solid was collected, washed with water and dried over

phosphorus pentoxide to give 2,3-dibromonaphthalene-1,4-dione 284 (17.0 g), mp 218-

220 °C (lit.,200 216-218 °C), which was used directly in the next reaction without

purification.

2-Benzyloxy-3-bromonaphthalene-1,4-dione 286

The procedure described by Slamet was followed.167 n-Butyllithium in hexane (1.6 M,

35 ml, 56 mmol) was added dropwise to a stirred solution of benzyl alcohol (7.41 g,

68.6 mmol) in anhydrous tetrahydrofuran (180 ml) cooled in an ice bath under argon.

After 20 min the solution was treated with 2,3-dibromonaphthalene-1,4-dione 284 (16.4

g, 52.0 mmol) in anhydrous tetrahydrofuran (50 ml) and the resulting mixture was

stirred for a further 1.5 h. The reaction mixture was quenched with a little ice, then

diluted with water (500 ml) and extracted with ethyl acetate (3 x 100 ml). The combined

organic extracts were washed with water (100 ml), followed by brine (100 ml), dried

and evaporated to give a yellow residue, which was subjected to silica gel filtration.

Elution with 5-30% ethyl acetate-light petroleum gave an oily yellow solid, which

recrystallised from dichloromethane-light petroleum to give 2-benzyloxy-3-

bromonaphthalene-1,4-dione 286 as yellow needles (8.92 g, 50%), mp 103-104 °C

(lit.,167 103-104 °C).

156

Page 168: Synthesis of Some Naturally Occurring Quinones

2-Benzyloxy-3-(2-furyl)naphthalene-1,4-dione 288

The procedure described by Slamet was followed.167 A mixture of 2-benzyloxy-3-

bromonaphthalene-1,4-dione 286 (4.38 g, 12.8 mmol), 2-(tri-n-butylstannyl)furan184

(5.61 g, 15.7 mmol), palladium tetrakis(triphenylphosphine) (546 mg, 0.473 mmol) and

copper (I) bromide (531 mg, 3.70 mmol) in dioxane (60 ml) was heated at reflux for 30

min under argon. Then the reaction mixture was diluted with water (500 ml) and

extracted with ethyl acetate (3 x 60 ml). The combined organic extracts were washed

with water (60 ml), followed by brine (50 ml), dried and evaporated to give a red-brown

residue, which crystallised from dichloromethane-light petroleum to give 2-benzyloxy-

3-(2-furyl)naphthalene-1,4-dione 288 as dark red plates (2.92 g), mp 141 °C (lit.,167

139-140 °C). The mother liquor was concentrated under reduced pressure, adsorbed

onto silica and subjected to silica gel filtration. Elution with 2.5-5% ethyl acetate-light

petroleum gave a further portion of the title naphthoquinone 288 as a bright red solid

(0.825 g, total combined yield 89%). The 1H NMR spectrum was identical with that

described by Slamet.167

2-Benzyloxy-3-(2-furyl)-1,4-dimethoxynaphthalene 290

The procedure described by Slamet was followed.167 A mixture of 2-benzyloxy-3-(2-

furyl)naphthalene-1,4-dione 288 (1.71 g, 5.18 mmol), tetrabutylammonium bromide

(109 mg, 0.295 mmol) and sodium dithionite (2.87 g, 16.5 mmol) in dichloromethane

(45 ml) and water (45 ml) was stirred vigorously under argon. After 30 min the reaction

mixture was treated sequentially with a solution of sodium hydroxide (2.31 g, 57.8

mmol) in water (7 ml) and dimethyl sulfate (3.5 ml, 4.7 g, 37 mmol) and the resulting

solution was left to stir for 17 h. The organic layer was separated and the aqueous layer

was extracted with dichloromethane (2 x 40 ml). The combined organic extracts were

washed with water (40 ml), followed by brine (40 ml), dried and evaporated to give a

157

Page 169: Synthesis of Some Naturally Occurring Quinones

light brown oil, which was subjected to silica gel filtration. Elution with 2.5% ethyl

acetate-light petroleum gave 2-benzyloxy-3-(2-furyl)-1,4-dimethoxynaphthalene 290

(1.80 g, 96%) as a pale yellow oil. The 1H NMR spectrum was identical with that

described by Slamet.167

3-(2-Furyl)-1,4-dimethoxynaphthalen-2-ol 228

A solution of 2-benzyloxy-3-(2-furyl)-1,4-dimethoxynaphthalene 290 (1.86 g, 5.17

mmol) in ethyl acetate (50 ml) was hydrogenated in the presence of 10% palladium on

carbon (280 mg) until TLC indicated that most of the starting material had been

consumed (ca 3.5 h). The reaction mixture was filtered through a plug of Celite and the

filtrate was evaporated to give a yellow oil, which was subjected to silica gel filtration.

Elution with 3% ethyl acetate-light petroleum returned unreacted starting material (157

mg) as a pale yellow oil. Further elution with 7.5% ethyl acetate afforded 3-(2-furyl)-

1,4-dimethoxynaphthalen-2-ol 228 (1.27 g, 91%) as a yellow oil. δH (200 MHz, CDCl3)

8.11 (1H, d, J 8.4, ArH), 7.99 (1H, d, J 8.4, ArH), 7.65 (1H, d, J 1.8, furyl H) 7.56-7.32

(2H, m, ArH), 7.04 (1H, d, J 3.3, furyl H), 6.64 (1H, dd, J 3.3, 1.8, furyl H), 4.00 (3H, s,

OCH3), 3.78 (3H, s, OCH3). The 1H NMR spectrum of this material was identical to that

recorded for the title compound prepared previously by another method.167

Attempted synthesis of ethyl 3-[3-(2-furyl)-1,4-dimethoxynaphthalen-2-yloxy]propynoate 263 a) n-Butyllithium in hexane (1.6 M, 0.65 ml, 1.0 mmol) was added dropwise to a

stirred solution of 3-(2-furyl)-1,4-dimethoxynaphthalen-2-ol 228 (261 mg, 0.967 mmol)

in anhydrous tetrahydrofuran (8 ml) cooled in an ice bath under argon. After 5 min the

solution was treated with ethyl 3-bromopropiolate187 (184 mg, 1.04 mmol) in anhydrous

tetrahydrofuran (1 ml), then the ice bath was removed and the reaction mixture was left

to stir at room temperature for a further 22 h. The mixture was diluted with water (80

158

Page 170: Synthesis of Some Naturally Occurring Quinones

ml) and extracted with dichloromethane (3 x 40 ml). The combined organic extracts

were washed with brine (30 ml), dried and evaporated to give a brown oil, which was

subjected to radial chromatography. Elution with 10% ethyl acetate-light petroleum

returned unreacted starting material (36 mg). Further elution with 20% ethyl acetate-

light petroleum gave ethyl 3-[3-(2-furyl)-1,4-dimethoxy-2-oxo-1,2-dihydronaphthalen-

1-yl]propynoate 294 as a yellow oil (105 mg, 30%, 34% based on recovered starting

material). (Found M+•, 366.1104. C21H18O6 requires 366.1103). Mass spectrum m/z: 366

(M, 51%), 337 (29), 336 (46), 335 (100), 321 (23), 308 (16), 307 (69), 306 (34), 305

(16), 293 (45), 291 (17), 279 (22), 277 (26), 265 (24), 264 (18), 263 (38), 247 (19), 236

(19), 235 (48), 234 (15), 221 (15), 220 (40), 165 (17), 164 (20), 163 (35), 152 (19), 151

(21). δH (300 MHz, CDCl3) 7.90-7.87 (1H, m, ArH), 7.80-7.77 (1H, m, ArH), 7.56-7.45

(3H, m, 2 x ArH, 1 x furyl H), 6.64 (1H, dd, J 3.3, 0.7, furyl H), 6.52 (1H, dd, J 3.3, 1.8,

furyl H), 4.22 (2H, q, J 7.1, CH2), 3.75 (3H, s, OCH3), 3.48 (3H, s, OCH3), 1.28 (3H, t,

J 7.1, CH3). δC (75.5 MHz, CDCl3) 190.7 (C=O), 165.7 (C=O), 152.8 (C), 144.5 (C),

142.6 (CH), 135.4 (C), 131.1 (CH), 129.7 (CH), 128.8 (C), 128.2 (CH), 125.5 (CH),

113.5 (CH), 111.5 (CH), 107.4 (C), 82.3 (C), 79.3 (C), 76.0 (C), 62.3 (CH2), 60.5

(OCH3), 54.5 (OCH3), 13.9 (CH3). vmax (CH2Cl2)/cm-1 1713 (C=O), 1676 (C=O), 2237

(C≡C).

b) A suspension of 3-(2-furyl)-1,4-dimethoxynaphthalen-2-ol 228 (137 mg, 0.507

mmol), ethyl 3-bromopropiolate187 (93 mg, 0.52 mmol) and potassium carbonate (175

mg,1.27 mmol) in acetone (2.5 ml) was stirred at room temperature for 2.5 h under

argon. TLC analysis indicated that starting material still remained so the reaction

mixture was heated at 40 °C (bath) for 3 h, then allowed to cool to room temperature

and left to stir for a further 16 h. The mixture was diluted with water (40 ml) and

extracted with dichloromethane (3 x 20 ml). The combined organic extracts were

159

Page 171: Synthesis of Some Naturally Occurring Quinones

washed with brine (20 ml), dried and evaporated to give a brown oil, which was

subjected to radial chromatography. Elution with 2.5% ethyl acetate-light petroleum

gave ethyl 3-hydroxy-6,11-dimethoxybenzo[b]naphtho[2,3-d]furan-2-carboxylate 298

as a white crystalline solid (52 mg, 28%), which recrystallised from dichloromethane-

light petroleum as pale yellow prisms, mp 154-155 °C. (Found M+•, 366.1099. C21H18O6

requires 366.1103). Mass spectrum m/z: 366 (M, 41%), 351(19), 320 (18), 306 (19), 305

(100), 290 (26). δH (500 MHz, CDCl3) 11.55 (1H, s, OH), 8.36-8.34 (1H, m, ArH),

8.23-8.22 (2H, m, ArH), 7.54-7.51 (2H, m, ArH), 7.03 (1H, d, J 8.5, ArH), 4.57 (2H, q,

J 7.2, OCH2), 4.45 (3H, s, OCH3), 4.09 (3H, s, OCH3), 1.56 (3H, t, J 7.2, CH3). δC

(125.75 MHz, CDCl3) 169.7 (C=O), 162.8 (C), 156.2 (C), 143.9 (C), 142.8 (C), 134.9

(C), 129.3 (CH), 126.1 (C), 125.2 (C), 125.1 (CH), 124.7 (CH), 122.4 (CH), 121.4

(CH), 117.2 (C), 115.2 (C), 113.1 (CH), 99.6 (C), 62.0 (OCH2), 61.6 (OCH3), 60.6

(OCH3), 14.2 (CH3). vmax (CH2Cl2)/cm-1 1670 (C=O).

c) A solution of 3-(2-furyl)-1,4-dimethoxynaphthalen-2-ol 228 (167 mg, 0.618

mmol), ethyl 3-bromopropiolate187 (275 mg, 1.55 mmol) and N-methylmorpholine (30

mg, 0.30 mmol) in anhydrous acetonitrile (2 ml) was stirred at room temperature for 20

h under argon. TLC analysis indicated that starting material was still present so the

reaction mixture was heated at 55-60 ºC (bath) for 4 h, then allowed to cool to room

temperature and left to stir overnight. The mixture was concentrated under reduced

pressure to give a yellow oil, which was subjected to radial chromatography. Elution

with 1% ethyl acetate-light petroleum afforded ethyl 3-hydroxy-6,11-

dimethoxybenzo[b]naphtho[2,3-d]furan-2-carboxylate 298 (8 mg, 4%, 5% based on

recovered starting material) as a white crystalline solid. Further elution with 10% ethyl

acetate-light petroleum returned unreacted starting material (45 mg).

160

Page 172: Synthesis of Some Naturally Occurring Quinones

Ethyl 5,10-dimethoxyfuro[3,2-b]naphtho[2,3-d]furan-3-carboxylate 304

a) A stirred suspension of 3-(2-furyl)-1,4-dimethoxynaphthalen-2-ol 228 (193 mg,

0.715 mmol), ethyl 3-bromopropiolate187 (201 mg, 1.14 mmol) and potassium carbonate

(250 mg, 1.80 mmol) in acetone (5 ml) was heated at 40 °C (bath) under argon until

TLC indicated that the starting material had been consumed (ca 3.5 h). The reaction

mixture was filtered through a plug of Celite and washed through with a little acetone.

The filtrate was diluted with toluene (5 ml) and most of the acetone was removed by

distillation. Then the solution was treated with 3,6-di(pyridin-2’-yl)-1,2,4,5-tetrazine

(169 mg, 0.716 mmol) and heated at reflux for 5.5 h under argon. The resulting dark

brown mixture was adsorbed onto silica and subjected to silica gel filtration. Elution

with 1% ethyl acetate-light petroleum gave a number of fractions containing the impure

title compound, which were combined and further subjected to radial chromatography.

Elution with 2.5-5% ethyl acetate-light petroleum gave ethyl 5,10-dimethoxyfuro[3,2-

b]naphtho[2,3-d]furan-3-carboxylate 304 as a pale pink crystalline solid (27 mg, 11%),

which recrystallised from dichloromethane-light petroleum as faint pink needles, mp

199-200 °C. (Found M+•, 340.0936. C19H16O6 requires 340.0947). Mass spectrum m/z:

341 (M + 1, 21%), 340 (M, 87), 326 (31), 325 (100), 297 (65). δH (500 MHz, CDCl3)

8.33-8.31 (1H, m, ArH), 8.28-8.26 (1H, m, ArH), 8.10 (1H, s, furyl H), 7.54-7.47 (2H,

m, ArH), 4.45 (2H, q, J 7.1, CH2), 4.39 (3H, s, OCH3), 4.30 (3H, s, OCH3), 1.45 (3H, t,

J 7.1, CH3). δC (125.75 MHz, CDCl3) 161.3 (C=O), 150.6 (CH, furyl), 148.7 (C), 146.2

(C), 142.3 (C), 141.5 (C), 134.4 (C), 125.6 (CH), 125.3 (C), 124.5 (CH), 123.1 (C),

122.3 (CH), 121.7 (CH), 110.4 (C), 106.0 (C), 61.6 (OCH3), 61.2 (CH2), 59.9 (OCH3),

14.3 (CH3). vmax (CH2Cl2)/cm-1 1724 (C=O).

161

Page 173: Synthesis of Some Naturally Occurring Quinones

b) A suspension of 3-(2-furyl)-1,4-dimethoxynaphthalen-2-ol 228 (163 mg, 0.604

mmol), ethyl 3-bromopropiolate187 (190 mg, 1.07 mmol) and cesium carbonate (492

mg, 1.51 mmol) in acetone (4 ml) was stirred at room temperature under argon until

TLC indicated that the starting material had been consumed (ca 45 min). The reaction

mixture was filtered through a plug of Celite and washed through with a little acetone.

The filtrate was diluted with toluene (5 ml) and most of the acetone was removed by

distillation. Then the solution was treated with 3,6-di(pyridin-2’-yl)-1,2,4,5-tetrazine

(152 mg, 0.644 mmol) and heated at reflux for 4 h under argon. The reaction mixture

was filtered through a plug of silica and washed through with 20% ethyl acetate-light

petroleum. The filtrate was evaporated and the resulting residue was subjected to radial

chromatography. Elution with 5% ethyl acetate-light petroleum gave ethyl 5,10-

dimethoxyfuro[3,2-b]naphtho[2,3-d]furan-3-carboxylate 304 as a faint pink solid (8 mg,

4%).

c) Potassium methoxide in anhydrous methanol (0.51 M, 1.8 ml, 0.92 mmol) was

added dropwise to a stirred solution of 3-(2-furyl)-1,4-dimethoxynaphthalen-2-ol 228

(232 mg, 0.859 mmol) in anhydrous methanol (1 ml) cooled in an ice bath under argon.

After the addition was completed, the ice bath was removed and the reaction mixture

was stirred for 30 min. The mixture was evaporated, then diluted with acetone (2 ml)

and treated with ethyl 3-bromopropiolate187 (230 mg, 1.30 mmol) and potassium

carbonate (301 mg, 2.18 mmol). The resulting suspension was stirred vigorously for 1 h

at room temperature under argon, then filtered through a plug of Celite and washed

through with a little acetone. Toluene (5 ml) was added to the mixture and most of the

acetone was removed by distillation. The mixture was treated with 3,6-di(pyridin-2’-yl)-

1,2,4,5-tetrazine (201 mg, 0.852 mmol) and heated under reflux for 3 h 20 min. The

resulting brown mixture was adsorbed onto silica and subjected to silica gel filtration.

162

Page 174: Synthesis of Some Naturally Occurring Quinones

Elution with 20% ethyl actetate-light petroleum gave a fraction which was further

subjected to radial chromatography. Elution with 2.5% ethyl actetate-light petroleum

afforded the title compound 304 (3 mg, 1%) as a white solid.

(E)-1,2-Dichloroethenyl 3-(2-furyl)-1,4-dimethoxynaphthalen-2-yl ether 248

The procedure was adapted from that described by Slamet.167 n-Butyllithium in hexane

(1.35 M, 1.2 ml, 1.6 mmol) was added dropwise to a stirred solution of 3-(2-furyl)-1,4-

dimethoxynaphthalen-2-ol 228 (393 mg, 1.46 mmol) in anhydrous tetrahydrofuran (7

ml) cooled in an ice bath under argon. After 10 min the mixture was concentrated under

reduced pressure and the resulting yellow residue was dissolved in dimethylformamide

(10 ml). The reaction mixture was then treated with trichloroethylene (1.74 g, 13.2

mmol) and left to stir overnight under argon. The mixture was diluted with water and

extracted with 50% ether-light petroleum (4 x 50 ml). The combined organic extracts

were washed successively with dilute sodium hydroxide solution and water, dried and

evaporated to give a brown oil, which was subjected to silica gel filtration. Elution with

2% ethyl acetate-light petroleum afforded (E)-1,2-dichloroethenyl 3-(2-furyl)-1,4-

dimethoxynaphthalen-2-yl ether 248 (454 mg, 85%) as a colourless oil. δH (200 MHz,

CDCl3) 8.26-8.09 (2H, m, ArH), 7.63 (1H, dd, J 1.8, 0.8, CH, furyl), 7.61-7.50 (2H, m,

ArH), 6.82 (1H, d, J 0.8, CH, furyl), 6.59 (1H, dd, J 3.4, 1.8, CH, furyl), 5.56 (1H, s,

CH, vinylic), 4.03 (3H, s, CH3), 3.75 (3H, s, CH3). This 1H NMR spectrum is identical

to that described by Slamet.167

163

Page 175: Synthesis of Some Naturally Occurring Quinones

Attempted synthesis of ethyl 5,10-dimethoxyfuro[3,2-b]naphtho[2,3-d]furan-3-

carboxylate 304 from (E)-1,2-dichloroethenyl 3-(2-furyl)-1,4-

dimethoxynaphthalen-2-yl ether 248

tert-Butyllithium in hexane (0.80 M, 3.85 ml, 3.1 mmol) was added dropwise over 10

min to a stirred solution of (E)-1,2-dichloroethenyl 3-(2-furyl)-1,4-

dimethoxynaphthalen-2-yl ether 248 (391 mg, 1.07 mmol) in anhydrous tetrahydrofuran

(35 ml) at –78 °C under argon. After 35 min ethyl chloroformate (0.70 ml, 794 mg, 7.3

mmol) was added, then the cold bath was removed and the resulting mixture was left to

stir for 50 min. The reaction mixture was diluted with anhydrous toluene (15 ml),

treated with 3,6-di(pyridin-2’-yl)-1,2,4,5-tetrazine (219 mg, 0.928 mmol) and most of

the tetrahydrofuran was removed by distillation. The resulting mixture was heated at

reflux for 5 h under argon, then filtered through a plug of Celite and washed through

with dichloromethane. The filtrate was concentrated under reduced pressure and

subjected to radial chromatography. Elution with 20% ethyl acetate-light petroleum

afforded a yellow solid (257 mg, 47%), which recrystallised from dichloromethane-light

petroleum as colourless rods, mp 131-132 °C. This compound was formulated as either

306 or 307. (Found M+•, 508.0692. C24H22Cl2O8 requires 508.0692). Mass spectrum

m/z: 511 (19%), 510 (68), 509 (27), 508 (M, 100), 475 (29), 474 (32), 473 (77), 472

(38), 457 (21), 429 (15), 427 (34), 409 (17), 357 (18), 353 (37), 327 (18), 325 (21), 313

(20), 281 (16). δH (300 MHz, CDCl3) 8.23-8.11 (2H, m, ArH), 7.65-7.57 (2H, m, ArH),

7.33 (1H, d, J 3.6, CH), 6.94 (1H, d, J 3.6, CH), 4.39 (2H, q, J 7.1, CH2), 4.31 (2H, q, J

7.1, CH2), 4.00 (3H, s, CH3), 3.89 (3H, s, CH3), 1.39 (3H, t, J 7.1, CH3), 1.36 (3H, t, J

7.1, CH3). δC (75.5 MHz, CDCl3) 162.1 (C=O), 158.7 (C=O), 151.7 (C), 149.3 (C),

148.6 (C), 144.7 (C), 144.4 (C), 138.9 (C), 128.8 (C), 128.1 (CH), 127.5 (C), 127.0

(CH), 123.3 (CH), 122.3 (CH), 119.1 (CH), 115. 1 (C), 114. 2 (CH), 103.6 (C), 62.6

164

Page 176: Synthesis of Some Naturally Occurring Quinones

(OCH3), 62.5 (OCH3), 62.2 (CH2), 61.0 (CH2), 14.3 (CH3), 14.1 (CH3). vmax

(CH2Cl2)/cm-1 1716 (C=O).

Ethyl 5,10-dioxofuro[3,2-b]naphtho[2,3-d]furan-3-carboxylate 311

A solution of ceric ammonium nitrate (149 mg, 0.272 mmol) in water (2 ml) was added

dropwise to a stirred suspension of ethyl 5,10-dimethoxyfuro[3,2-b]naphtho[2,3-

d]furan-3-carboxylate 304 (31 mg, 0.091 mmol) in acetonitrile (2 ml) cooled in an ice

bath under argon. The mixture was stirred vigorously for 10 min, then diluted with

water (20 ml) and extracted with ethyl acetate (3 x 20 ml). The combined organic

extracts were washed with brine, dried and evaporated to give a bright yellow solid,

which was subjected to radial chromatography. Elution with 20-50% ethyl actetate-light

petroleum gave the ethyl 5,10-dioxofuro[3,2-b]naphtho[2,3-d]furan-3-carboxylate 311

(14 mg, 50%), which recrystallised from dichloromethane-light petroleum as fine

yellow needles, mp 219-220 °C. (Found M+•, 310.0478. C17H10O6 requires 310.0477).

Mass spectrum m/z: 311 (M + 1, 20%), 310 (M, 100), 282 (33), 265 (33), 238 (20). δH

(300 MHz, CDCl3) 8.33 (1H, s, CH, furyl), 8.27-8.20 (2H, m, ArH), 7.84-7.75 (2H, m,

ArH), 4.44 (2H, q, J 7.1, CH2), 1.43 (3H, t, J 7.1, CH3). δC (75.5 MHz, CDCl3) 178.9

(C=O), 173.5 (C=O), 160.0 (C=O), 156.0 (CH, furyl), 154.3 (C), 149.1 (C), 142.6 (C),

134.3 (CH), 133.9 (CH), 132.2 (C), 132.1 (C), 127.1 (CH), 126.0 (CH), 118.2 (C),

110.5 (C), 61.6 (CH2), 14.2 (CH3). vmax (CH2Cl2)/cm-1 1729 (C=O), 1674 (C=O).

3-Hydroxymethylfuro[3,2-b]naphtho[2,3-d]furan-5,10-dione 37

A solution of ethyl 5,10-dioxofuro[3,2-b]naphtho[2,3-d]furan-3-carboxylate 311 (12

mg, 0.039 mmol) in anhydrous tetrahydrofuran (1.5 ml) was added to a suspension of

lithium aluminium hydride (62 mg, 1.6 mmol) in anhydrous tetrahydrofuran (1.5 ml) at

165

Page 177: Synthesis of Some Naturally Occurring Quinones

0 ºC under argon and the resulting mixture was stirred until TLC indicated that the

starting material had been consumed (ca 20 min). The reaction mixture was quenched

with a little ethyl acetate and ice, then diluted with water (5 ml) and extracted with

chloroform (1 x 3 ml), followed by ethyl acetate (4 x 3 ml). The combined organic

extracts were washed with brine and concentrated under reduced pressure. To the

resulting brown residue was added Fremy’s salt (40 mg, 0.15 mmol), ether (3 ml) and

an aqueous borax buffer solution (0.025 M sodium tetraborate, 2.7 ml; 0.1 M sodium

hydroxide, 1.3 ml) and the resulting mixture was shaken vigorously for 20 min. The

organic layer was separated and the aqueous layer was extracted with chloroform (4 x 3

ml). The combined organic extracts were washed with brine, dried and evaporated to

give a red residue, which was subjected to radial chromatography. Elution with 50-70%

ethyl acetate-light petroleum afforded 3-hydroxymethylfuro[3,2-b]naphtho[2,3-d]furan-

5,10-dione 37 as a red solid (5 mg, 49%), mp 204 ºC onwards (sublimed and underwent

crystal phase change), 240-241 °C (melted).(lit.,48 217-218 °C). (Found M+•, 268.0368.

C15H8O5 requires 268.0372). Mass spectrum m/z: 269 (M + 1, 22%), 268 (M, 100), 252

(17). The 1H and 13C NMR spectral data are given in Table 2 (p. 132). vmax (KBr)/cm-1

3485 (OH), 1672 (C=O), 1651 (C=O).

166

Page 178: Synthesis of Some Naturally Occurring Quinones

References 1. (a) L. F. Fieser and M. Fieser, Organic Chemistry, 3rd edn, Reinhold Publishing

Corporation, New York, 1956, p. 709.

(b) A. Woskresensky, Ann. Pharm., 1838, 27, 257.

2. C. Asche, Mini Rev. Med. Chem., 2005, 5, 449.

3. R. H. Thomson, Naturally Occurring Quinones, Academic Press Inc. (London)

Ltd., London, 1971, p. 1.

4. D. V. Banthorpe, in J. Mann, R. S. Davidson, J. B. Hobbs, D. V. Banthorpe and

J. B Harborne, Natural Products: their chemistry and biological significance,

Longman Group UK Ltd., Harlow, London, 1994, p. 295.

5. F. L. Crane, Mitochondrion, 2007, 7S, S2.

6. G. Lenaz, R. Fato, G. Formiggini and M. L. Genova, Mitochondrion, 2007, 7S,

S8.

7. J. B. Harborne, in Comprehensive Natural Products Chemistry, ed. D. Barton

and K. Nakanishi, Elsevier Science Ltd., Oxford, 1999, vol. 8, p. 170.

8. W. J. Rietveld, J. Chem. Ecol., 1983, 9, 295.

9. F. Ponder, Jr. and S. H. Tadros, J. Chem. Ecol., 1985, 11, 937.

10. A. M. Hejl and K. L. Koster, J. Chem. Ecol., 2004, 30, 453.

11. W. J. Rietveld, R. C. Schlesinger and K. J. Kessler, J. Chem. Ecol., 1983, 9,

1119.

12. S. B. Singh, M. G. Cordingley, R. G. Ball, J. L. Smith, A. W. Dombrowski and

M. A. Goetz, Tetrahedron Lett., 1991, 32, 5279.

13. F. Schaller, L. Rahalison, N. Islam, O. Potterat and K. Hostettmann, Helv. Chim.

Acta, 2000, 83, 407.

14. D. -S. Lee, S. -H. Lee, J.-G. Noh and S.-D. Hong, Biosci. Biotechnol. Biochem.,

1999, 63, 2236.

167

Page 179: Synthesis of Some Naturally Occurring Quinones

15. B. Y. Kang, S. W. Chung, S. H. Kim, S. Y. Ryu and T. S. Kim,

Immunopharmacology, 2000, 49, 355.

16. A. Plubrukarn, S. Yuenyongsawad, T. Thammasaroj and A. Jitsue, Pharm. Biol.,

2003, 41, 439.

17. Science of Synthesis, Volume 28: Compounds with Two Carbon-Heteroatom

Bonds: Quinones and Heteroatom Analogues, ed. A. G. Griesbeck. Georg

Thieme Verlag, Stuttgart, 2006. (Review: M. Brimble, Synthesis, 2007, 2924.)

18. P. Kovacic and J. A. Osuna, Jr., Curr. Pharm. Design, 2000, 6, 277.

19. J. Mann, Life Saving Drugs: The Elusive Magic Bullet. The Royal Society of

Chemistry, Cambridge, 2004, pp. 190-191.

20. D. V. Lefemine, M. Dann, F. Barbatschi, W. K. Hausmann, V. Zbinovsky, P.

Monnikendam, J. Adam and N. Bohonos, J. Am. Chem. Soc., 1962, 84, 3184.

21. J. S. Webb, D. B. Cosulich, J. H. Mowat, J. B. Patrick, R. W. Broschard, W. E.

Meyer, R. P. Williams, C. F. Wolf, W. Fulmor, C. Pidacks and J. E Lancaster, J.

Am. Chem. Soc., 1962, 84, 3185.

22. J. S. Webb, D. B. Cosulich, J. H. Mowat, J. B. Patrick, R. W. Broschard, W. E.

Meyer, R. P. Williams, C. F. Wolf, W. Fulmor, C. Pidacks and J. E Lancaster, J.

Am. Chem. Soc., 1962, 84, 3187.

23. J. Cummings, V. J. Spanswick and J. F. Smyth, Eur. J. Cancer, 1995, 31A,

1928.

24. M. Tomasz, R. Lipman, D. Chowdary, J. Pawlak, G. L. Verdine and K.

Nakanishi, Science, 1987, 235, 1204.

25. J. Mann, Life Saving Drugs: The Elusive Magic Bullet. The Royal Society of

Chemistry, Cambridge, 2004, pp. 163-164.

26. H. W. Moore, Science, 1977, 197, 527.

168

Page 180: Synthesis of Some Naturally Occurring Quinones

27. J. Mann, Life Saving Drugs: The Elusive Magic Bullet. The Royal Society of

Chemistry, Cambridge, 2004, pp. 191-193.

28. J. W. Lown, Chem. Soc. Rev., 1993, 165.

29. G. J. Quigley, A. H.-J. Wang, G. Ughetto, G. van der Marel, J. H. van Boom and

A. Rich, Proc. Natl. Acad. Sci. USA, 1980, 77, 7204.

30. J. Mann, Life Saving Drugs: The Elusive Magic Bullet. The Royal Society of

Chemistry, Cambridge, 2004, p. 190.

31. P. D'Arpa and L. F. Liu, Biochim. Biophys. Acta, 1989, 989, 163.

32. A. H. Corbett and N. Osheroff, Chem. Res. Toxicol., 1993, 6, 585.

33. R. D. Olson, R. C. Boerth, J. G. Gerber and A. S. Nies, Life Sci., 1981, 29, 1393.

34. K. Shan, A. M. Lincoff and J. B. Young, Ann. Intern. Med., 1996, 125, 47.

35. K.-H. Lee, J. Biomed. Sci., 1999, 6, 236.

36. J. Mann, Nat. Rev. Cancer, 2002, 2, 143.

37. R. M. Wilson and S. J. Danishefsky, Acc. Chem. Res., 2006, 39, 539.

38. K. C. Nicolaou and S. A. Snyder, Angew. Chem. Int. Ed. Engl., 2005, 44, 1012.

39. G. A. Cordell, Phytochemistry, 1995, 40, 1585.

40. K. Mori, in Comprehensive Natural Products Chemistry, ed. D. Barton and K.

Nakanishi, Elsevier Science Ltd., Oxford, 1999, vol. 8, p. 6.

41. K. Mori, in Comprehensive Natural Products Chemistry, ed. D. Barton and K.

Nakanishi, Elsevier Science Ltd., Oxford, 1999, vol. 8, p. 4-5.

42. S. Omura, A. Nakagawa, H. Yamada, T. Hata, A. Furusaki, and T. Watanabe,

Chem. Pharm. Bull., 1973, 21, 931.

43. G. I. Dmitrienko, K. E. Nielsen, C. Steingart, N. S. Ming, J. M. Willson and G.

Weeratunga, Tetrahedon Lett., 1990, 31, 3681.

44. S. Mithani, G. Weeratunga, N. J. Taylor and G. I. Dmitrienko, J. Am. Chem.

Soc., 1994, 116, 2209.

169

Page 181: Synthesis of Some Naturally Occurring Quinones

45. S. J. Gould, N. Tamayo, C. R. Melville and M. C. Cone, J. Am. Chem. Soc.,

1994, 116, 2207.

46. X. Lei and J. A. Porco, Jr., J. Am. Chem. Soc., 2006, 128, 14790.

47. H. Hara, N. Maruyama, S. Yamashita, Y. Hayashi, K.-H. Lee, K. F. Bastow,

Chairul, R. Marumoto and Y. Imakura, Chem. Pharm. Bull., 1997, 45, 1714.

48. C. E. Heltzel, A. A. L. Gunatilaka, T. E. Glass and D. G. I. Kingston,

Tetrahedron, 1993, 49, 6757.

49. Z. Chen, H. Huang, C. Wang, Y. Li, J. Ding, U. Sankawa, H. Noguchi and Y.

Iitaka, Chem. Pharm. Bull., 1986, 34, 2743.

50. U. Mahmood, V. K. Kaul and L. Jirovetz, Phytochemistry, 2002, 61, 923.

51. L. F. Villegas, I. D. Fernández, H. Maldonado, R. Torres, A. Zavaleta, A. J.

Vaisberg and G. B. Hammond, J. Ethnopharmacol., 1997, 55, 193.

52. J. Lin, T. Puckree and T. P. Mvelase, J. Ethnopharmacol., 2002, 79, 53.

53. T. M. A. Alves, H. Kloos and C. L. Zani, Mem. Inst. Oswaldo Cruz, 2003, 98,

709.

54. T. H. Everett, The New York Botanical Garden Illustrated Encyclopedia of

Horticulture, Garland Publishing, Inc., New York, 1981, vol. 4, p. 1191.

55. B. Weniger, J. Ethnopharmacol., 1982, 6, 67.

56. H. Schmid, T. M. Meijer and A. Ebnöther, Helv. Chim. Acta, 1950, 33, 595.

57. H. Schmid, A. Ebnöther and T. M. Meijer, Helv. Chim. Acta, 1950, 33, 1751.

58. H. Schmid and A. Ebnöther, Helv. Chim. Acta, 1951, 34, 561.

59. W. Eisenhuth and H. Schmid, Helv. Chim. Acta, 1958, 41, 2021.

60. H. Schmid and A. Ebnöther, Helv. Chim. Acta, 1951, 34, 1041.

61. C. Bianchi and G. Ceriotti, J. Pharm. Sci., 1975, 64, 1305.

62. P. Krishnan and K. F. Bastow, Biochem. Pharmacol., 2000, 60, 1367.

63. M. A. Brimble, L. J. Duncalf and M. R. Nairn, Nat. Prod. Rep., 1999, 16, 267.

170

Page 182: Synthesis of Some Naturally Occurring Quinones

64. K. Kobayashi, M. Uchida, T. Uneda, M. Tanmatsu, O. Morikawa and H.

Konishi, Tetrahedron Lett., 1998, 39, 7725.

65. M. A. Brimble, M. R. Nairn and H. Prabaharan, Tetrahedron, 2000, 56, 1937.

66. A. D. Webb and T. M. Harris, Tetrahedron Lett., 1977, 2069.

67. T. M. Harris and C. M. Harris, Tetrahedron, 1977, 33, 2159.

68. T. Kometani, Y. Takeuchi and E. Yoshii, J. Chem. Soc., Perkin Trans. 1, 1981,

1197.

69. Y. Naruta, H. Uno and K. Maruyama, J. Chem. Soc. Chem., Commun., 1981,

1277.

70. H. Uno, J. Org. Chem., 1986, 51, 350.

71. R. G. F. Giles, I. R. Green, V. I. Hugo and P. R. K. Mitchell, J. Chem. Soc.,

Chem. Commun., 1983, 51.

72. R. G. F. Giles, I. R. Green, V. I. Hugo, P. R. K. Mitchell and S. C. Yorke, J.

Chem. Soc., Perkin Trans. 1, 1984, 2383.

73. R. G. F. Giles, I. R. Green, V. I. Hugo, P. R. K. Mitchell and S. C. Yorke, J.

Chem. Soc., Perkin Trans. 1, 1983, 2309.

74. R. G. F. Giles, I. R. Green and J. A. X. Pestana, J. Chem. Soc., Perkin Trans. 1,

1984, 2389.

75. K. Kobayashi, M. Uchida, T. Uneda, K. Yoneda, M. Tanmatsu, O. Morikawa

and H. Konishi, J. Chem. Soc., Perkin Trans. 1, 2001, 2977.

76. L. M. Tewierik, C. Dimitriadis, C. D. Donner, M. Gill and B. Willems, Org.

Biomol. Chem., 2006, 4, 3311.

77. C. Dimitriadis, M. Gill and M. F. Harte, Tetrahedron: Asymmetry, 1997, 8,

2153.

78. J. S. Gibson, O. Andrey and M. A. Brimble, Synthesis, 2007, 2611.

171

Page 183: Synthesis of Some Naturally Occurring Quinones

79. Methods of Organic Chemistry (Houben-Weyl), ed. A. de Meijere, Georg

Thieme Verlag, Stuttgart, 1997, vol. E 17e.

80. M. T. Crimmins and T. L. Reinhold, Organic Reactions, 1993, 44, 297.

81. T. Bach, Synthesis, 1998, 683.

82. J. Mattay, R. Conrads and R. Hoffmann, in Methods of Organic Chemistry

(Houben-Weyl), ed. G. Helmchen, R. W. Hoffmann, J. Mulzer and E.

Schaumann, Georg Thieme Verlag, Stuttgart, 1995, vol. E 21c, p. 3085.

83. W. M. Horspool, in Photochemistry in Organic Synthesis, ed. J. D. Coyle, The

Royal Society of Chemistry, London, 1986, p. 210.

84. P. A. Wender, in Photochemistry in Organic Synthesis, ed. J. D. Coyle, The

Royal Society of Chemistry, London, 1986, p. 163.

85. W. Oppolzer, Acc. Chem. Res., 1982, 15, 135.

86. G. Ciamician and P. Silber, Ber. Dtsch. Chem. Ges., 1908, 41, 1928.

87. G. Büchi and I. M. Goldman, J. Am. Chem. Soc., 1957, 79, 4741.

88. J. Iriondo-Alberdi and M. F. Greaney, Eur. J. Org. Chem., 2007, 4801.

89. E. J. Corey, R. B. Mitra and H. Uda, J. Am. Chem. Soc., 1964, 86, 485.

90. C. R. Johnson and N. A. Meanwell, J. Am. Chem. Soc., 1981, 103, 7667.

91. P. E. Eaton and T. W. Cole, Jr., J. Am. Chem. Soc., 1964, 86, 962.

92. P. E. Eaton and T. W. Cole, Jr., J. Am. Chem. Soc., 1964, 86, 3157.

93. J. C. Namyslo and D. E. Kaufmann, Chem. Rev., 2003, 103, 1485.

94. E. J. Corey, J. Dolf Bass, R. LeMahieu, and R. B. Mitra, J. Am. Chem. Soc.,

1964, 86, 5570.

95. P. de Mayo, Acc. Chem. Res., 1971, 4, 41.

96. R. O. Loutfy and P. de Mayo, J. Am. Chem. Soc., 1977, 99, 3559.

97. D. I. Schuster, G. Lem and N. A. Kaprinidis, Chem. Rev., 1993, 93, 3.

98. J. D. White and D. N. Gupta, J. Am. Chem. Soc., 1966, 88, 5364.

172

Page 184: Synthesis of Some Naturally Occurring Quinones

99. J. D. White and D. N. Gupta, J. Am. Chem. Soc., 1968, 90, 6171.

100. M. C. Pirrung, J. Am. Chem. Soc., 1979, 101, 7130.

101. M. C. Pirrung, J. Am. Chem. Soc., 1981, 103, 82.

102. T. Naito, Y. Makita and C. Kaneko, Chem. Lett., 1984, 921.

103. D. Bryce-Smith, E. H. Evans, A. Gilbert and H. S. Mc Neill, J. Chem. Soc.,

Perkin Trans. 1, 1992, 485.

104. J. A. Porco, Jr. and S. L. Schreiber, in Comprehensive Organic Synthesis, ed. B.

M. Trost, Pergamon Press, Oxford, 1991, vol. 5, p. 151.

105. C. Covell, A. Gilbert and C. Richter, J. Chem. Research (S), 1998, 316.

106. S. Cleridou, C. Covell, A. Gadhia, A. Gilbert and P. Kamonnawin, J. Chem.

Soc., Perkin Trans. 1, 2000, 1149.

107. T. Suzuki, Y. Yamashita, T. Mukai and T. Miyashi, Tetrahedron Lett., 1988, 29,

1405.

108. M. T. Crimmins, Chem. Rev., 1988, 88, 1453.

109. R. L. Cargill, J. R. Dalton, S. O' Connor and D. G. Michels, Tetrahedron Lett.,

1978, 4465.

110. Y. Tanada and K. Mori, Eur. J. Org. Chem., 2001, 4313.

111. Aldrich Handbook of Fine Chemicals, 2007-2008, Sigma-Aldrich Co., 2006.

112. H. C. Brown and P. Geoghegan, Jr., J. Am. Chem. Soc., 1967, 89, 1522.

113. R. H. Thomson, Naturally Occurring Quinones, Academic Press Inc. (London)

Ltd., London, 1971.

114. W. H. Watanabe and L. E. Conlon, J. Am. Chem. Soc., 1957, 79, 2828.

115. A. S. Rao, Tetrahedron, 1983, 39, 2323.

116. M. Tokunaga, J. F. Larrow, F. Kakiuchi and E. N. Jacobsen, Science, 1997, 277,

936.

117. J. M. Keith, J. F. Larrow and E. N. Jacobsen, Adv. Synth. Catal., 2001, 343, 5.

173

Page 185: Synthesis of Some Naturally Occurring Quinones

118. S. E. Schaus, B. D. Brandes, J. F. Larrow, M. Tokunaga, K. B. Hansen, A. E.

Gould, M. E. Furrow and E. N. Jacobsen, J. Am. Chem. Soc., 2002, 124, 1307.

119. D. E. J. E. Robinson and S. D. Bull, Tetrahedon: Asymmetry, 2003, 14, 1407.

120. P. Kumar, V. Naidu and P. Gupta, Tetrahedron, 2007, 63, 2745.

121. Z.-Y. Liu, Z.-C. Chen, C.-Y. Yu, R.-F. Wang, R.-Z. Zhang, C.-S. Huang, Z.

Yan, D.-R. Cao, J.-B. Sun and G. Li, Chem. Eur. J., 2002, 8, 3747.

122. G. Dubois, A. Murphy and T. D. P. Stack, Org. Lett., 2003, 5, 2469.

123. R. W. Murray and M. Singh, Org. Synth., 1997, 74, 91.

124. E. J. Corey and M. Chaykovsky, J. Am. Chem. Soc., 1965, 87, 1353.

125. E. Pérez Sacau, A. Estévez-Bruan, Á. G. Ravelo, E. A. Ferro, H. Tokuda, T.

Mukainaka and H. Nishino, Bioorg. Med. Chem., 2003, 11, 483.

126. M. Schlosser, in Organometallics in Synthesis. A Manual. ed. M. Schlosser,

John Wiley & Sons Ltd., Chichester, 1994, p. 86.

127. S. T. Chadwick, R. A. Rennels, J. L. Rutherford, and D. B. Collum, J. Am.

Chem. Soc., 2000, 122, 8640.

128. T. Kamikawa and I. Kubo, Synthesis, 1986, 431.

129. A. S.-Y. Lee, Y.-J. Hu and S.-F. Chu, Tetrahedron, 2001, 57, 2121.

130. K. Mori, in Comprehensive Natural Products Chemistry, ed. D. Barton and K.

Nakanishi, Elsevier Science Ltd., Oxford, 1999. vol. 8, p. 7.

131. A. S. Cotterill, M. Gill and N. M. Milanovic, J. Chem. Soc., Perkin Trans. 1,

1995, 1215.

132. M. Gill and C. Elsworth, Nat. Prod. Lett., 2000, 14, 411.

133. U. Pindur and G. H. Schneider, Chem. Soc. Rev., 1994, 23, 409.

134. E. M. Stocking and R. M. Williams, Angew. Chem. Int. Ed. Engl., 2003, 42,

3078.

135. H. Oikawa, Bull. Chem. Soc. Jpn., 2005, 78, 537.

174

Page 186: Synthesis of Some Naturally Occurring Quinones

136. F. Da Silva Knudsen, A. Campa, H. A. Stefani and G. Cilento, Proc. Natl. Acad.

Sci. U.S.A, 1994, 91, 410.

137. M. S. DeClue, K. K. Baldridge, D. E. Künzler, P. Kast and D. Hilvert, J. Am.

Chem. Soc., 2005, 127, 15002.

138. M. S. DeClue, K. K. Baldridge, P. Kast and D. Hilvert, J. Am. Chem. Soc., 2006,

128, 2043.

139. E. Hao, J. Fromont, D. Jardine and P. Karuso, Molecules, 2001, 6, 130.

140. N. Tanaka, M. Okasaka, Y. Ishimaru, Y. Takaishi, M. Sato, M. Okamoto, T.

Oshikawa, S. U. Ahmed, L. M. Consentino and K. -H. Lee, Org. Lett., 2005, 7,

2997.

141. K. C. Nicolaou, D. Sarlah and D. M. Shaw, Angew. Chem. Int. Ed., 2007, 46,

4708.

142. J. Bass, J. Lat. Am. Geogr., 2004, 3, 67.

143. L. A. Bentacur-Galvis, J. Saez, H. Granados, A. Salazar and J. E. Ossa, Mem.

Inst. Oswaldo Cruz, 1999, 94, 531.

144. M. V. Berry, Southeast. Geogr., 2005, 45, 239.

145. M. Heinrich, A. Ankli, B. Frei, C. Weimann and O. Sticher, Soc. Sci. Med.,

1998, 47, 1859.

146. C. A. Lans, J. Ethnobiol. Ethnomed., 2006, 2, 45.

147. R. Otero, V. Núñez, J. Barona, R. Fonnegra, S. L. Jiménez, R. G. Osorio, M.

Saldarriaga and A. Díaz, J. Ethnopharmacol., 2000, 73, 233.

148. C. E. Heltzel, A. A. L. Gunatilaka, T. E. Glass and D. G. I. Kingston, J. Nat.

Prod., 1993, 56, 1500.

149. T. Kaneko, K. Ohtani, R. Kasai, K. Yamasaki and N. M. Duc, Phytochemistry,

1997, 46, 907.

175

Page 187: Synthesis of Some Naturally Occurring Quinones

150. T. Kaneko, K. Ohtani, R. Kasai, K. Yamasaki and N. M. Duc, Phytochemistry,

1998, 47, 259.

151. A. A. L. Gunatilaka, D. G. I. Kingston and R. K. Johnson, Pure Appl. Chem.,

1994, 66, 2219.

152. R. Kinnel, A. J. Duggan, T. Eisner and J. Meinwald, Tetrahedon Lett., 1977,

3913.

153. A. Nahrstedt, V. Wray, B. Engel and R. Reinhard, Planta Med., 1985, 44, 517.

154. M. Abou-Shoer, K. Suwanborirux, A. -A. M. Habib, C.-J. Chang and J. M.

Cassady, Phytochemistry, 1993, 34, 1413.

155. J.-i. Aihara, J. Phys. Org. Chem., 2005, 18, 235.

156. G. Subramanian, P. von Ragué Schleyer and H. Jiao, Angew. Chem. Int. Ed.

Engl., 1996, 35, 2638.

157. S. Inagaki and M. Urushibata, Bull. Chem. Soc. Jpn., 1990, 63, 3117.

158. G. Buemi, J. Chim. Phys. Physicochim. Biol., 1987, 84, 1147.

159. E. L. Tolmach, A. B. Kudryavtsev, A. Y. Zheltov and B. I. Stepanov, J. Org.

Chem. USSR, 1989, 25, 1594.

160. V. P. Vaidya and Y. S. Agasimundin, Indian J. Chem., 1983, 22B, 462.

161. J. Elguero, C. Marzin, A. R. Katritzky and P. Linda, The Tautomerism of

Heterocycles, Academic Press, Inc., New York, 1976, p. 224-226.

162. B. H. Lipshutz, Chem. Rev., 1986, 86, 795.

163. E. Ciganek, Org. React., 1984, 32, 1.

164. C. O. Kappe, S. S. Murphree and A. Padwa, Tetrahedron, 1997, 53, 14179.

165. M. V. Sargent and F. M. Dean, in Comprehensive Heterocyclic Chemistry, ed.

C. W. Bird and G. W. H. Cheeseman, Pergamon Press Ltd., Oxford, 1984. vol.

4, pp. 619-634.

166. J. H. Buttery, J. Moursounidis and D. Wege, Aust. J. Chem., 1995, 48, 593.

176

Page 188: Synthesis of Some Naturally Occurring Quinones

167. R. Slamet, PhD Thesis, University of Western Australia, 2001.

168. R. Slamet and D. Wege, unpublished results.

169. A. Moyano, F. Charbonnier and A. E. Greene, J. Org. Chem., 1987, 52, 2919.

170. N. Kann, V. Bernardes and A. E. Greene, Org. Synth., 1997, 74, 13.

171. V. V. Zhdankin, in Comprehensive Functional Group Transformations II, ed. A.

R. Katritzky and R. J. K. Taylor, Elsevier Ltd., Oxford, 2005, vol. 2, pp.1093

1096.

172. F. Camps, J. Coll, A. Llebaria, J. M. Moretó and S. Ricart, Synthesis, 1989, 123.

173. C. J. Pouchert and J. Behnke, The Aldrich Library of 13C and 1H FT NMR

Spectra, Aldrich Chemical Company, Inc., Milwaukee, 1993, vol. 3, p. 28.

174. S. Y. Delavarenne and H. G. Viehe, in Chemistry of Acetylenes, ed. H. G. Viehe,

Marcel Dekker, New York, 1969, p. 651.

175. R. Tanaka and S. I. Miller, Tetrahedron Lett., 1971, 1753.

176. P. J. Stang, in Modern Acetylene Chemistry, ed. P. J. Stang & F. Diederich.

VCH Verlagsgesellschaft mbh, Weinheim, 1995, p. 67.

177. R. Tanaka, M. Rodgers, R. Simonaitis and S. I. Miller, Tetrahedron, 1971, 27,

2651.

178. E. Winterfeldt, Angew. Chem. Int. Ed. Engl., 1967, 6, 423.

179. M. E. Jung, in Comprehensive Organic Synthesis: Selectivity, Strategy, and

Efficiencey in Modern Organic Chemistry, ed. B. M. Trost and I. Fleming,

Pergamon Press, Oxford, 1991, vol. 4, p. 1.

180. E. Ciganek, Synthesis, 1995, 1311.

181. L. I. Vereshchagin, L. D. Gavrilov, R. L. Bol'shedvorskaya, L. P. Kirillova and

L. L. Okhapkina, J. Org. Chem. USSR, 1973, 9, 514.

182. K. Rathwell and M. A. Brimble, Synthesis, 2007, 643.

183. D. Mal and P. Pahari, Chem. Rev., 2007, 107, 1892.

177

Page 189: Synthesis of Some Naturally Occurring Quinones

184. L. S. Liebeskind and J. Wang, J. Org. Chem., 1993, 58, 3550.

185. I. F. Bel'skii and N. I. Shuikin, Russ. Chem. Rev., 1963, 32, 307.

186. M. V. Sargent and F. M. Dean, in Comprehensive Heterocyclic Chemistry: The

Structure, Reactions, Synthesis and Uses of Heterocyclic Compounds, ed. C. W.

Bird and G. W. H. Cheeseman, Pergamon Press, Oxford, 1984, vol. 4, p. 614.

187. Ethyl 3-bromopropiolate 292 was prepared by replacing methyl propiolate with

ethyl propiolate in the procedure described for the synthesis of methyl 3

bromopropiolate in J. Leroy, Org. Synth., 1997, 74, 212.

188. N. Kornblum, R. Seltzer and P. Haberfield, J. Am. Chem. Soc., 1963, 85, 1148.

189. C. Reichardt, Solvent and Solvent Effects in Organic Chemistry, Wiley-VCH

Verlag GmbH & Co, Weinheim, 2003, pp. 269-273.

190. R. W. Alder, R. Baker and J. M. Brown, Mechanism in Organic Chemistry,

John Wiley & Sons Ltd., London, 1971, pp.139-141.

191. C. J. Pouchert and J. Behnke, The Aldrich Library of 13C and 1H FT NMR

Spectra, Aldrich Chemical Company, Inc. Milwaukee, vol. 3, p. 14.

192. R. T. Winters, A. D. Sercel and H. D. H. Showalter, Synthesis, 1988, 712.

193. J. P. Parrish, B. Sudaresan and K. W. Jung, Synth. Commun., 1999, 29, 4423.

194. F. M. Dean and M. V. Sargent, in Comprehensive Heterocyclic Chemistry: The

Structure, Reactions, Synthesis and Uses of Heterocyclic Compounds, ed. C. W.

Bird and G. W. H. Cheeseman, Pergamon Press Ltd., Oxford, 1984, vol. 4, pp.

557-558.

195. J. Guillard, C. Lamazzi, O. Meth-Cohn, C. W. Rees, A. J. P. White and D. J.

Williams, J. Chem. Soc., Perkin Trans. 1, 2001, 1304.

196. A. Bighelli, F. Tomi and J. Casanova, Magn. Reson. Chem., 1992, 30, 1268.

197. P. Jacob, P. S. Callery, A. T. Shulgin and N. Castagnoli, Jr., J. Org. Chem.,

1976, 41, 3627.

178

Page 190: Synthesis of Some Naturally Occurring Quinones

179

198. D. M. L. Goodgame, in Inorganic Experiments, ed. J. D. Woollins, VCH

Verlagsgesellschaft, Weinheim, 1994, p. 41.

199. D. E. White and E. N. Jacobsen, Tetrahedron: Asymmetry, 2003, 14, 3633.

200. S. M. McElvain and E. L. Engelhardt, J. Am. Chem. Soc., 1944, 66, 1077.