rhodium-catalyzed intermolecular hydroacylation …...ii rhodium-catalyzed intermolecular...

107
Rhodium-Catalyzed Intermolecular Hydroacylation of Unactivated Alkenes and Application to the Total Synthesis of Octaketide Natural Products by Christine Mai Le A thesis submitted in conformity with the requirements for the degree of Masters of Science Department of Chemistry University of Toronto © Copyright by Christine Mai Le 2012

Upload: others

Post on 25-Jan-2020

27 views

Category:

Documents


0 download

TRANSCRIPT

Rhodium-Catalyzed Intermolecular Hydroacylation of Unactivated Alkenes and Application to the Total Synthesis of Octaketide

Natural Products

by

Christine Mai Le

A thesis submitted in conformity with the requirements for the degree of Masters of Science

Department of Chemistry University of Toronto

© Copyright by Christine Mai Le 2012

ii

Rhodium-Catalyzed Intermolecular Hydroacylation of Unactivated

Alkenes and Application to the Total Synthesis of Octaketide

Natural Products

Christine Mai Le

Masters of Science

Department of Chemistry

University of Toronto

2012

Abstract

Transition metal-catalyzed olefin hydroacylation represents an atom-economical approach for the

synthesis of valuable ketone products. To date, the intermolecular variant of this reaction suffers

from several drawbacks, which include limited substrate scope, poor reactivity and/or

regioselectivity for non-activated, non-chelating alkene substrates, and competitive reductive

decarbonylation pathways that lead to catalyst decomposition. Herein, we report the linear-

selective intermolecular hydroacylation of a wide range of electronically diverse olefins with

salicylaldehydes employing catalyst loadings as low as 2 mol%. A unique reactivity profile is

observed for the chiral C2-symmetric phosphoramidite ligand employed in our catalyst system,

and thus, we outline progress made towards the synthesis of new phosphoramidite ligands. We

have applied our methodology in the total synthesis of nine octaketide natural products belonging

to the dothiorelone, cytosporone, and phomopsin families. Due to recent reports demonstrating

the anticancer activity of cytosporone B (Csn-B), we will also discuss progress towards the

synthesis of Csn-B analogues.

iii

Acknowledgments

First and foremost, I would like to thank my supervisor, Prof. Vy Dong for her continual support,

guidance, and encouragement throughout the course of the year. I would also like to thank

Wilmer Alkhas for his hard work around the lab and for always going out of his way to help

students.

I am grateful for having the opportunity to work with such a high-spirited and intellectual group

of people who have made me feel welcome in the lab since the very first day I started. I would

like to send a very appreciative and warm thank you to Dr. Max von Delius, my collaborator on

this project. I am thankful for his patience, encouragement, and willingness to teach.

Lastly, I would like to thank all my family and friends for supporting and believing in me, even

when I doubted myself. I am grateful for having them in my life.

iv

Table of Contents

Acknowledgments .......................................................................................................................... iii

Table of Contents ........................................................................................................................... iv

List of Abbreviations ..................................................................................................................... vi

List of Tables ................................................................................................................................. ix

List of Figures ................................................................................................................................. x

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

Chapter 1 Rhodium-Catalyzed Intermolecular Alkene Hydroacylation ......................................... 1

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

1.1.1 The Origins of Hydroacylation ............................................................................... 1

1.1.2 Recent Advances and Current Limitations in Intermolecular Hydroacylation ....... 2

1.2 Project Objectives ............................................................................................................... 7

1.3 Results and Discussion ....................................................................................................... 8

1.3.1 A Brief Summary of the Method Development and Optimization Studies ............ 8

1.3.2 Investigation of Aldehyde Scope ............................................................................ 9

1.3.3 Investigation of Alkene Scope .............................................................................. 11

1.3.4 Biphasic Hydroacylation ....................................................................................... 14

1.3.5 Mechanistic Considerations .................................................................................. 16

1.4 Importance of Phosphoramidite Ligands in Catalysis ...................................................... 17

1.4.1 Synthesis of Phosphoramidite Ligands: En route to More Active and Selective

Hydroacylation Catalysts ...................................................................................... 19

1.5 Conclusions and Future Work .......................................................................................... 24

1.6 Experimental ..................................................................................................................... 25

1.6.1 General considerations .......................................................................................... 25

1.6.2 General procedure for the linear-selective hydroacylation of monosubstituted

alkenes with salicylaldehydes ............................................................................... 25

v

1.6.3 General procedure for the linear-selective hydroacylation of monosubstituted

alkenes with salicylaldehydes under biphasic conditions ..................................... 26

1.6.4 Characterization of compounds ............................................................................ 27

Chapter 2 Application of Intermolecular Hydroacylation to the Synthesis of Octaketide

Natural Products ....................................................................................................................... 33

2.1 Introduction ....................................................................................................................... 33

2.1.1 Accessing Polyketide Natural Products from Dothiorelone, Cytosporone, and

Phomopsin Families via Hydroacylation .............................................................. 33

2.1.2 Cytosporone B ...................................................................................................... 35

2.2 Research Goals .................................................................................................................. 37

2.3 Results and Discussion ..................................................................................................... 38

2.3.1 Synthesis of Starting Materials for Key Hydroacylation ...................................... 38

2.3.2 Testing Known Hydroacylation Methods ............................................................. 40

2.3.3 Solvent Screen for Hydroacylation of Natural Product Substrates ....................... 40

2.3.4 Synthesis of Octaketide Natural Products ............................................................. 42

2.3.5 Design and Synthesis of Cytosporone B Analogues ............................................. 45

2.4 Conclusions and Future Work .......................................................................................... 47

2.5 Experimental ..................................................................................................................... 47

2.5.1 General Considerations ......................................................................................... 47

2.5.2 General Procedures for the Total Synthesis of Natural Products via Linear-

Selective Hydroacylation ...................................................................................... 48

2.5.3 Characterization of Compounds ........................................................................... 48

Appendix A: NMR Spectra ........................................................................................................... 66

Appendix B: Crystallographic Data .............................................................................................. 96

vi

List of Abbreviations

1H NMR proton nuclear magnetic resonance

13C NMR carbon nuclear magnetic resonance

° C degrees Celsius

SIPHOS-PE 10,11,12,13-Tetrahydrodiindeno[7,1-de:1′, 7′-

fg][1,3,2]dioxaphosphocin-5-bis[(R)-1-phenylethyl]amine N-

Di[(R)-1-phenylethyl]-[(R)-1,1′-spirobiindane-7,7′-diyl]-

phosphoramidite

α alpha

β beta

δ chemical shift

μL microlitre

acac acetylacetonate

aq. aqueous

Ac acetyl

Arg Arginine

atm atmosphere

BINAP 1,1’-binaphthalene-2,2’-diyl-bis-diphenylphosphine

bs broad singlet

COD 1,5-cyclooctadiene

conv. conversion

d doublet; days

dd doublet of doublets

DCE 1,2-dichloroethane

DCM dichloromethane

DMAP 4-dimethylaminopyridine

DMF N,N-dimethylformamide

d.r. diastereomeric ratio

dt doublet of triplets

ee enantiomeric excess

EI electron impact

vii

eq. equation

equiv. equivalents

ESI electron spray ionization

Et ethyl

EtOAc ethyl acetate

g grams

GC-FID gas chromatography-flame ionization detector

GC-MS gas chromatography-mass spectrometry

h hours

HRMS high resolution mass spectrometry

Hz hertz

J coupling constant

LC-MS liquid chromatography-mass spectrometry

m multiplet

M molar

M.P. melting point

M+ parent molecular ion

Me methyl

MeCN acetonitrile

Me-THF 2-methyltetrahydrofuran

mg milligrams

MHz megahertz

min minutes

mL millilitres

mmol millimoles

MsOH methanesulfonic acid

n.r. no reaction

nBu n-butyl

NMR nuclear magnetic resonance

PhH benzene

PDB Protein Data Bank

ppm parts per million

viii

Pro proline

pyr pyridine

q quartet

rt room temperature

s singlet

Ser serine

t triplet

tt triplet of triplets

tBu tert-butyl

THF tetrahydrofuran

Thr threonine

TLC thin layer chromatography

TMS trimethylsilane

Ts tosyl (p-toluenesulfonyl)

ix

List of Tables

Table 1: Substrate scope for salicylaldehyde derivatives .............................................................. 9

Table 2: Solvent screen for the hydroacylation of 1-heptene (4) with 2,4-dihydroxybenzaldehyde

(1f) ................................................................................................................................................. 11

Table 3: Alkene scope .................................................................................................................. 12

Table 4: Testing selected alkene substrates under biphasic conditions ....................................... 15

Table 5: Conditions employed for Friedel-Crafts cyclization ..................................................... 22

Table 6: Solvent screen and reaction monitoring for hydroacylation of 1-heptene (22d) with

aldehyde 21b ................................................................................................................................. 41

Table 7: Synthesis of octaketide natural products ....................................................................... 43

Table 8: Synthesis of cytosporone B analogues ........................................................................... 46

x

List of Figures

Figure 1: Aldehydes bearing a β-chelating heteroatom ................................................................. 4

Figure 2: Common phosphoramidite ligands ............................................................................... 18

Figure 3: Proposed amines to be used in spiro-phosphoramidite synthesis ................................ 20

Figure 4: General structure of octaketide natural products from the dothiorelone, phomopsin,

and cytosporone families .............................................................................................................. 35

Figure 5: (a) Superimposition of Csn-B (yellow) onto the ribbon structure of the Nur77 LBD.

(b) Cytosporone B. (c) Compound 15 ........................................................................................... 36

Figure 6: Reaction progress for hydroacylation of 1-heptene (22d) with aldehyde 21b in Me-

THF and THF ................................................................................................................................ 42

Figure 7: Solid state structure of dothiorelone B (23bb). ............................................................ 44

Figure 8: Ligand binding pocket of Nur77 obtained from the PDB ............................................ 45

xi

List of Schemes

Scheme 1: Transition metal-catalyzed intra- or intermolecular hydroacylation of alkenes or

alkynes leading to ketone products ................................................................................................. 1

Scheme 2: a) Jun’s catalyst system employing 2-amino-picoline co-catalyst; b) Proposed mode

of activation for aldehyde substrates ............................................................................................... 5

Scheme 3: Bisphasic conditions for hydroacylation of 1-octene (2a) with salicylaldehyde (1a) 15

Scheme 4: Proposed catalytic cycle ............................................................................................. 17

Scheme 5: Common synthetic route to access phosphoramidites ................................................ 19

Scheme 6: Synthesis of spirobiindanediol backbone ................................................................... 22

Scheme 7: Synthesis of C2-symmetric chiral amine A1 .............................................................. 24

Scheme 8: Structural revision of phomopsin B and Takahashi’s approach to dothiorelone A .... 34

Scheme 9: Proposed key disconnection for synthesis of dothiorelone A .................................... 34

Scheme 10: Proposed synthesis of octaketides belonging to the dothiorelone, cytosporone, and

phomopsin families ....................................................................................................................... 37

Scheme 11: Synthesis of key aldehydes 21a-c ............................................................................. 39

Scheme 12: Synthesis of key alkenes 22a-c ................................................................................. 39

Scheme 13: Testing Suemune’s and Jun’s conditions for synthesis of phomopsin C ................. 40

1

Chapter 1 Rhodium-Catalyzed Intermolecular Alkene Hydroacylation

1.1 Introduction

1.1.1 The Origins of Hydroacylation

Developing highly efficient and environmentally benign protocols for the synthesis of natural

products and pharmaceutical agents is a key challenge in both academic and industrial settings

today. Towards this end, the direct functionalization of C–H bonds represents an attractive route

for both small and complex molecule synthesis, since pre-activation of formally unreactive C–H

bonds is not required—thereby minimizing chemical waste, while increasing reaction

efficiency.1 Transition-metal catalyzed hydroacylation represents a completely atom-economical

approach for the construction of carbon-carbon bonds, since all the atoms of the starting material

are conserved in the product.2 This particular transformation invokes the activation of an

aldehyde C–H bond by a metal species, followed by the formal addition of the acyl unit and

hydrogen atom across an unsaturated functional group (typically an alkene or alkyne) (Scheme

1).3

Scheme 1: Transition metal-catalyzed intra- or intermolecular hydroacylation of alkenes or

alkynes leading to ketone products.

1 (a) Ritleng, V.; Sirlin, C.; Pfeffer, M. Chem. Rev. 2002, 102, 1731-1769. (b) Labinger, J. A.; Bercaw, J. E. Nature

2002, 417, 507-514. 2 Trost, B. M. Angew. Chem. Int. Ed. Engl. 1995, 34, 259-281.

3 For a general and comprehensive review of transition metal-catalyzed alkene and alkyne hydroacylation, refer to

ref.: (a) Willis, M. C. Chem. Rev. 2010, 110, 725-748. For a microreview of transition metal-catalyzed alkene and

alkyne hydroacylation, refer to ref.: (b) Jun, C.-H.; Jo, E.-A.; Park, J.-W. Eur. J. Org. Chem. 2007, 1869-1881. For a

review of catalytic intermolecular hydroacylation in the absence of chelation-assistance, refer to ref.: (c) Leung, J.

C.; Krische, M. J. Chem. Sci. 2012, 3, 2202-2209.

2

The first example of alkene hydroacylation was reported by Sakai and co-workers in 1972, in

which 4-pentenals could be transformed into their respective cyclopentanones (a) by using a

stoichiometric amount of Wilkinson’s complex (eq. 1).4 While this method is not catalytic, two

significant conclusions can be drawn from this study: (1) trapping the putative acyl-rhodium-

hydride intermediate by an unsaturated functional group is possible, allowing for the rapid

construction of molecular complexity, and (2) the formation of cyclopropane byproducts (b) in

significant quantities indicate that reductive decarbonylation is both a competing and

energetically favourable pathway. This decarbonylation pathway generates a catalytically

inactive Rh-CO species, which becomes problematic when catalytic versions of this

transformation are desired.5 Nonetheless, since Sakai’s seminal report, the development of

catalytic, diastereo- and enantioselective olefin hydroacylations has been achieved with both

relatively low catalyst loadings and expanded substrate scope.3 Although much progress has

been made in the field of intramolecular alkene hydroacylation, this report will focus on its

intermolecular counterpart with an emphasis on rhodium-catalyzed alkene hydroacylation.

1.1.2 Recent Advances and Current Limitations in Intermolecular Hydroacylation

In addition to challenges associated with poor reactivity as well as modest regioselectivity,

finding efficient and general intermolecular hydroacylation protocols that minimize undesirable

decarbonylation pathways has proven to be quite difficult. Towards this end, strategies that

stabilize the key acyl-rhodium intermediate in the catalytic cycle have been successful for

decreasing the propensity for reductive decarbonylation. This approach often involves

heteroatom chelation, whereby the aldehyde or alkene component (or both) can coordinate to the

4 Sakai, K.; Ide, J.; Oda, O.; Nakamura, N. Tetrahedron Lett. 1972, 1287.

5 Fairlie, D. P.; Bosnich. B. Organometallics 1988, 7, 946-954.

3

metal center in a bidentate fashion.3,6

Although this strategy comes at the cost of reduced

substrate scope, an increase in both reactivity as well as regioselectivity has been demonstrated

in such systems.

1.1.2.1 Exploiting β-Chelating Aldehydes

Following Sakai’s report of a stoichiometric system for the synthesis of cyclopentanones (refer

to eq. 1), Miller disclosed that catalysis was possible if ethylene-saturated chloroform was

employed as solvent.7 During this study it was discovered that by exchanging the rhodium

source from Wilkinson’s complex to Rh(acac)(C2H4)2, products a and b, resulting from the

intermolecular hydroacylation of ethylene, could be isolated—where products of type b are a

result of olefin isomerization (eq. 2). The use of enals as substrates proved to be essential for this

type of reactivity, as Miller and coworkers speculated that internal coordination of the alkene

moiety to the metal center assisted in catalyst stabilization. Notably, no products resulting from

reductive decarbonylation were detected and saturated aldehydes did not furnish the desired

hydroacylation adduct.

Initial reports of heteroatom-chelation came from the work of Suggs, where quinoline 8-

carboxylaldehydes (1) were employed for the Rh-catalyzed hydroacylation of 1-octene (Figure

1).8 Seventeen years later, Jun disclosed that benzaldehyde derivatives bearing an ortho-

phosphine substituent (2) could engage this transformation as well.9 In either case, coordination

6 For a review on chelation-assisted hydroacylation, see ref.: Jun, C.-H.; Hong, J.-B.; Lee, D.-Y. Synlett 1999, 1, 1-

12. 7 (a) Lochow, C. F.; Miller, R. G. J. Am. Chem. Soc. 1976, 98, 1281-1283. (b) Vora, K. P.; Lochow, C. F.; Miller, R.

G. J. Organomet. Chem. 1980, 192, 257. 8 Suggs, J. W. J. Am. Chem. Soc. 1978, 100, 640-641.

9 Lee, H.; Jun, C.-H. Bull. Korean Chem. Soc. 1995, 16, 66-68.

4

of the β-heteroatom to the metal center facilitates oxidative addition and results in the formation

of a cyclometallated acyl-Rh intermediate, which has a decreased likelihood of undergoing

decarbonylation, as this would form a strained 4-membered rhodacycle. Willis and Weller have

demonstrated that β-S-aldehydes (3) are effective substrates for the intermolecular

hydroacylation of electron-rich, neutral, and deficient olefins depending on the catalyst system

employed.10

More recently, they have shown that cationic rhodium catalysts containing small-

bite angle diphosphines are extremely efficient at promoting the hydroacylation of β-S-aldehydes

with a wide range of alkenes and alkynes.10c

Catalyst loadings as low as 0.1 mol% and turnover

frequencies of greater than 300 h-1

are reported, which represents the most efficient

intermolecular hydroacylation system to date. While these coordinating heteroatoms serve their

purpose of stabilizing the key acyl-rhodium intermediate, such strategies often necessitate the

incorporation of these motifs in the final product.11

Salicylaldehydes (4), which possess a phenol

moiety, are unique substrates for hydroacylation, since functionalization of such aldehydes can

lead to motifs that are encountered in natural products or to scaffolds that can be easily

derivatized thereafter (vide infra).

Figure 1: Aldehydes bearing a β-chelating heteroatom.

Perhaps one of the most general protocols for the coupling of simple aldehydes with electron-

neutral alkenes comes from the work of Jun, whereby 2-amino-picoline is employed as a co-

catalyst to assist in the oxidative addition and to prevent the decarbonylation of non-chelating

aldehydes, such as benzaldehyde, via reversible imine formation (Scheme 2a).12

In this proposed

mode of activation, the pyridyl nitrogen atom on the co-catalyst is ideally positioned to facilitate

10

(a) Willis, M. C.; McNally, S. K.; Beswick, P. J. Angew. Chem. Int. Ed. 2004, 43, 340-343. (b) Moxham, G. L.;

Randell-Sly, H. E.; Brayshaw, S. K.; Woodward, R. L.; Weller,A. S.; Willis, M. C. Angew. Chem. Int. Ed. 2006, 45,

7618-7622. (c) Chaplin, A. B.; Hooper, J. F.; Weller, A. S.; Willis, M. C. J. Am. Chem. Soc. 2012, 134, 4885-4897. 11

For use of removable β-chelating motifs in hydroacylation, refer to: (a) Willis, M. C.; Randell-Sly, H. E.;

Woodward, R. L.; Currie, G. S. Org. Lett. 2005, 7, 2249-2251. (b) Parsons, S. R.; Hooper, J. F.; Willis, M. C. Org.

Lett. 2011, 13, 998-1000. 12

Jun, C.-H.; Lee, D.-Y.; Lee, H.; Hong, J.-B. Angew. Chem. Int. Ed. 2000, 39, 3070-3072.

5

oxidative addition by coordinating to the rhodium center in the same fashion as typical β-

chelating aldehydes (Scheme 2b). Hydrolysis of the resulting ketimine furnishes the desired

coupling product and permits catalyst turnover. However, the high temperatures required for this

transformation may be considered a disadvantage.

Scheme 2: a) Jun’s catalyst system employing 2-amino-picoline co-catalyst; b) Proposed mode

of activation for aldehyde substrates.

1.1.2.2 The Use of Salicylaldehydes

Over the last decade, salicylaldehydes have emerged as an effective class of substrates for

intermolecular hydroacylation.13,14

In the presence of a catalytic amount of base, the phenolate

of salicylaldehyde can coordinate to the Rh center, allowing for both a facile oxidative addition

and a decreased likelihood of decarbonylation. Miura and co-workers were the first to apply

salicylaldehydes as substrates for intermolecular alkyne hydroacylation, whereby analogous

reactivity with alkenes could not be achieved in the majority of the substrates surveyed.14a,b

The

successful coupling of salicylaldehydes with electron-neutral olefins was realized by Tanaka and

13 Work from our own group: (a) Murphy, S. K.; Petrone, D. A.; Coulter, M. M.; Dong V. M. Org. Lett. 2011, 13,

6216-6219. (b) Murphy, S. K.; Coulter, M. M.; Dong V. M. Chem. Sci. 2012, 3, 355-358. (c) Coulter, M. M.; Kou,

K. G. M.; Galligan, B.; Dong, V. M. J. Am. Chem. Soc. 2010, 132, 16330-16333. (d) Phan, D. H. T.; Kou, K. G. M.;

Dong, V. M. J. Am. Chem. Soc. 2010, 132, 16354-16355. 14

Work from other groups: (a) Kokobu, K.; Matsumasa, K.; Miura, M.; Nomura, M. J. Org. Chem. 1997, 62, 4564-

4565.; (b) Kokobu, K.; Matsumasa, K.; Nishinaka, Y.; Miura, M.; Nomura, M. Bull. Chem. Soc. Jpn. 1999, 72, 303.

(c) Tanaka, M.; Imai, M.; Yamamoto, Y.; Tanaka, K.; Shimowatari, M.; Nagumo, S.; Kawahara, N.; Suemune, H.

Org. Lett. 2003, 5, 1365-1367. (d) Imai, M.; Tanaka, M.; Tanaka, K.; Yamamoto, Y.; Imai-Ogata, N.; Shimowatari,

M.; Nagumo, S.; Kawahara, N.; Suemune, H. J. Org. Chem. 2004, 69, 1144-1150. (e) Imai, M.; Tanaka, M.;

Nagumo, S.; Kawahara, N.; Suemune, H. J. Org. Chem. 2007, 72, 2543-2546.

6

Suemune, in which a double chelation strategy was utilized for the hydroacylation of 1,5-

hexadienes (eq. 3).14c,d

Modest levels of regioselectivity in favour of branched products a were

observed in this particular system. An extension of this methodology to simple non-chelating

olefins was accomplished by the addition of acetonitrile and sodium acetate to the reaction,

however this necessitated significantly higher catalyst loadings (40 mol%) (eq. 4).14e

Our group has demonstrated that certain alkenols can be successfully coupled to salicylaldehydes

when a phosphinite co-catalyst is used in conjunction with a Rh(I) precatalyst (eq. 5).13a,b

Reversible, yet covalent, binding of the allylic or homoallylic alcohol to the phosphinite permits

chelation of this substrate to the metal center, furnishing formal aldol and homoaldol products,

respectively. Branched ketone products resulting from the hydroacylation of homoallylic sulfides

with salicylaldehydes can be accessed with high levels of regio- and enantiocontrol when the

same Rh(I) precatalyst is combined with a sterically-encumbered phosphoramidite ligand, (R)-

SIPHOS-PE (eq. 6).13c

This report represents the first enantioselective intermolecular

hydroacylation with non-activated alkene substrates.

7

1.1.2.3 Current limitations

Although significant achievements have been made in the area of intermolecular olefin

hydroacylation, this transformation remains limited in scope due to a lack of existing catalysts

that can couple a wide range of aldehydes with electronically-diverse olefins. Achieving high

levels of regiocontrol with non-chelating, unactivated olefins, while employing low catalyst

loadings is a remaining challenge to overcome.

1.2 Project Objectives

Based on this literature precedence, Dr. Max von Delius, a post-doctoral fellow in the group,

envisioned that simple, unactivated olefins that do not possess a heteroatom directing group

could undergo regioselective hydroacylation with salicylaldehydes following the judicious

choice of metal and ligand, as well as with the fine tuning of reaction parameters. This ideal

transformation would expand the substrate scope with respect to the olefin component, while

incorporating a biologically relevant arylketone motif in the final coupling product. As such, we

plan to apply intermolecular hydroacylation as a key step in the total synthesis of polyketide

natural products that bear a similar aryl ketone motif to the hydroacylation products obtained

(refer to Chapter 2). In addition to expanding the substrate scope, we aim to find a reactive

hydroacylation system in which catalyst loadings can be lowered to levels below 5 mol%. The

results reported herein are a result of a joint effort between Max and me. The results presented in

the method development represents optimization done solely by Max and were added to inform

the reader of significant background knowledge, as well as for clarity and continuity in the

development of the project.

8

1.3 Results and Discussion

1.3.1 A Brief Summary of the Method Development and Optimization Studies

Based on the high regioselectivity observed for the coupling of homoallylic sulfides with

salicylaldehydes demonstrated by our own group (refer to eq. 6), we wondered whether we could

extend this transformation to traditionally unreactive, electron-neutral olefins that do not possess

directing groups—while maintaining high levels of regioselectivity. Following extensive

optimization efforts, we were pleased to find that the same catalyst system (under slightly

modified conditions) could promote the highly linear-selective hydroacylation of 1-octene (2a)

with salicylaldehyde (1a) (eq. 7) at catalyst loadings as low as 2 mol%. A thorough scan and fine

tuning of different reaction parameters had revealed that elevated temperatures were required for

reactivity and that 1,2-dichloroethane was the ideal solvent for the reaction. One critical

discovery was that the precise 1:1 ratio of heterogenous base to [Rh] was crucial for maintaining

reactivity, and that having a slight excess of base relative to [Rh] was detrimental to the reaction.

A screen of various phosphorous-based ligands led to the following conclusions: (1) only

monodentate phosphorous-based ligands were able to promote hydroacylation15

and (2) of the

surveyed monodentate phosphorous-based ligands, phosphoramidites proved to be the most

effective, with (R)-SIPHOS-PE showing exceptional reactivity and linear-selectivity. While

chiral ligands are typically employed for enantioinduction in metal-catalyzed transformations,

the unique stereoelectronic properties imparted to such ligands can enhance reactivity and

promote regioselectivity in challenging non-asymmetric transformations16

, as demonstrated in

this particular intermolecular hydroacylation. Interestingly, when we subjected salicylaldehyde

15

A screen of several bidentate phosphines proved to be ineffective for this transformation 16

For the use of chiral ligands to enhance reactivity in non-asymmetric transformations, see for example: a) J. P.

Wolfe, S. Wagaw, S. L. Buchwald, J. Am. Chem. Soc. 1996, 118, 7215-7216; b) A. H. Roy, J. F. Hartwig, J. Am.

Chem. Soc. 2003, 125, 8704-8705; c) Q. Shen, S. Shekhar, J. P. Stambuli, J. F. Hartwig, Angew. Chem. 2005, 117,

1395-1399; Angew. Chem. Int. Ed. 2005, 44, 1371-1375.

9

(1a) and 1-octene (2a) to Jun’s conditions (refer to Scheme 2a), only an 8% NMR yield of the

desired hydroacylation product was observed after extended heating. This observation

demonstrates that salicylaldehydes are not compatible substrates for this type of chemistry.

1.3.2 Investigation of Aldehyde Scope

With these optimized conditions in hand, a variety of salicylaldehyde derivatives were tested

(Table 1). Good to excellent yields of the linear products were observed when both sterically-

hindered (entry 1) and electron-deficient (entries 2-4) salicylaldehydes were subjected to the

reaction conditions. A control experiment in which the 2-hydroxy group was protected as the

aryl methyl ether revealed that the free phenolic proton at the 2-position is required for reactivity

(entry 8). Napthylaldehyde 1h and β-S-aldehyde 1j furnished the desired hydroacylation

products in good to modest yields (entries 7 and 9). To the best of our knowledge, entry 9

represents the first report of a neutral Rh-catalyzed hydroacylation with β-S-aldehydes. Electron-

rich salicylaldehyde 1f reacted smoothly under the optimized conditions (entry 5), however

having a free phenol at the 4-position proved to be problematic (entry 6)—presumably due to the

low solubility of this substrate (1g) in DCE. For all cases, linear-to-branched selectivities of

>95:5 were observed.

Table 1: Substrate scope for salicylaldehyde derivatives.

Entry Aldehyde t (h) Product Yielda,b

1 R: 6-Me (1b) 40 3ba 66%

2 R: 5-F (1c) 48 3ca 78%

3 R: 5-Cl (1d) 48c 3da 89%

4 R: 5-I (1e) 60c 3ea 97%

5 R: 4-OMe (1f) 36 3fa 82%

6 R: 4-OH (1g) 24 3ga trace

10

7 (1h) 24 3ha 96%

8 (1i) 48 3ia 0%

9 (1j) 72d 3ja 48%

a Isolated yield.

b Regioisomeric ratio was >95:5 in all cases (determined by integration of crude

1H NMR spectrum).

c 4.0 equiv. of alkene used.

d No base added, 6.0 equiv. of alkene used (neat).

We were interested in finding a set of conditions in which 2,4-dihydroxybenzaldehyde (1g) could

furnish the desired hydroacylation product in reasonable yield and reaction time, since this

particular aldehyde would later serve as a model substrate for our natural product syntheses (vide

infra). By increasing the catalyst loading by two-fold, the hydroacylation of 1g with 1-heptene

(2b)17

went to 67% conversion, but only after prolonged reaction time (i.e. >1 week) (Table 1,

entry 1). Due to the low solubility of 1g in DCE, a screen of various polar solvents was

conducted. At first we reasoned that solvent mixtures consisting of DCE and a more polar

solvent should be ideally suited for solubilizing 1g, while maintaining comparable levels of

efficiency and reactivity. Unfortunately, yields no greater than 26% were obtained when 1:1

solvent mixtures consisting of DCE and THF (entry 2), acetonitrile (entry 3), 1,4-dioxane (entry

4), or acetone (entry 5) were employed. Of these four solvents, THF and 1,4-dioxane were the

most efficient at promoting this transformation. Interestingly, almost identical yields were

obtained upon switching the solvent to THF and 1,4-dioxane entirely (c.f. entries 2 & 8, and

entries 4 & 6). At this point, we wondered whether varying the relative amount of K3PO4 to [Rh]

could have an effect on reactivity. In THF, the amount of base was varied from 20 to 5 mol%

(entries 7 – 9). A marked decrease in yield was observed when excess base relative to [Rh] was

used—a result that is consistent with our previous finding that the precise ratio 1:1 of [Rh] to

base is crucial for reactivity. Surprisingly, when we decreased the loading of base to 5 mol% (i.e.

17

We employed 1-heptene as the alkene coupling partner due to its relevance to the targeted natural products (vide

infra).

11

[Rh]:base = 2:1), a drastic increase in yield was observed (entry 9). Upon switching the ethereal

solvent from THF to 2-Me-THF, a similar trend was observed when we varied the loading of

base from 10 to 5 mol% (entries 10 – 12), whereby a [Rh]:base ratio of 2:1 gave the highest yield

of 78% (entry 12). When the [Rh] loading was decreased to 5 mol%, while maintaining a 2:1

ratio of [Rh]:base, an excellent isolated yield of 95% was obtained for the linear product after

allowing the reaction stir for 120 h (entry 13).

Table 2: Optimization of the hydroacylation of 1-heptene (4) with 2,4-dihydroxybenzaldehyde

(1f).

Entry x:y:z (mol %) Solvent composition NMR yielda (%)

1 5:5:10 DCEb

65

2 5:5:10 DCE:THF (1:1) 26

3 5:5:10 DCE:CH3CN (1:1) <1

4 5:5:10 DCE:1,4-dioxane (1:1) 21

5 5:5:10 DCE:Acetone (1:1) 13

6 5:5:10 1,4-dioxane 20

7 5:5:20 THF trace

8 5:5:10 THF 28

9 5:5:5 THF 66

10 5:5:10 Me-THF 42

11 5:5:7.5 Me-THF 49

12 5:5:5 Me-THF 78

13c

2.5:5:2.5d

Me-THF 95e

a NMR yield determined by referencing

1H spectrum to an internal standard: 1,3,5-trimethoxybenzene.

b Reaction

time >1 week, no observable catalyst decomposition observed. c 1-octene was used instead of 1-heptene. Rxn time =

120 h, product = 3ga d At [Rh] catalyst loadings of <10 mol%, a 1:1 ratio of [Rh]:ligand is necessary for maintaining

reactivity. At [Rh] catalyst loadings of ≥10 mol%, a decrease in ligand loading to 5 mol% has no observable effect

on reactivity. e Isolated yield

1.3.3 Investigation of Alkene Scope

Having identified a general protocol for the coupling of salicylaldehyde derivatives with 1-

octene (2a), we turned our attention to the alkene scope. Overall, a variety of electronically-

12

diverse olefins were tolerated under the reaction conditions (i.e. electron-neutral, rich, or

deficient, as well as activated or unactivated), provided that only monosubstituted alkenes were

employed. Substitution at either the α or β position shut down reactivity completely. Due to the

wide electronic differences in the olefins screened, further optimization studies were carried out,

which entailed subjecting particularly unreactive alkene substrates to a set of five different

reaction conditions. This fine tuning would allow trends in reactivity and selectivity to be

deduced for each class of substrates tested. Yields above 60% and selectivities of >9:1 are

typically observed in the surveyed substrates and the results outlined in Table 3 represent the

highest yielding conditions employed.

Unactivated, electron-neutral alkenes 2c-g (Table 3, entries 1 – 5) displayed excellent reactivity,

as well as linear-to-branched selectivity (>95:5 in all cases) under the optimized reaction

conditions. To the best of our knowledge, entry 5 represents the first example of an acrolein

diacetal (2g) being used as a substrate for intermolecular hydroacylation. Notably, acetal

deprotection of hydroacylation product 3ag would allow for the convenient synthesis of 1,4-

ketoaldehydes. Free alcohol and protected amine functionalities are well tolerated (entries 6 –

10), with alkenes 2h-i (entries 6 – 7) furnishing the desired coupling product with the highest

yields and linear-selectivity. Reduced yields and selectivities are observed for olefins 2j-k in the

series (entries 8 – 9), which can be attributed to potential coordination of the nearby hydroxyl

group to the metal center. A complete reversal in selectivity is observed for 2-hydroxystyrene

(2l)—a case in which we presume branched hydride insertion is favoured to the due formation of

a 5-membered rhodacycle (entry 10). Although diene 2m displayed exceptional reactivity, the

lowered linear-selectivity can be attributed to coordination of the pendant alkene to the metal

center (entry 11). Much to our delight, electron-rich vinyl silane 2n reacted smoothly to give the

hydroacylation product (entry 12); however, modest yield was obtained for electron-rich butyl

vinyl ether (2o) (entry 13). Styrene (2p) and styrene derivatives (2q-r), which are electronically

biased to favour branched product formation, reacted cleanly, furnishing the linear

hydroacylation products with acceptable levels of regioselectivity (entries 14 – 16). When we

subjected α,β-unsaturated ester 2t to the reaction conditions, excellent reactivity was observed,

albeit with modest levels of linear-selectivity (entry 18). Electron-neutral alkene 2u displayed

poor reactivity under the reaction conditions (entry 19), while sterically-hindered alkene 2v did

not react at all (entry 20). Overall, we were pleased to find that our method was effective for a

13

range of electronically-diverse olefins, with electron-neutral alkenes displaying the best

reactivity and linear-to-branched selectivity.

Table 3: Alkene scope.

Entry Alkene t (h) Conditions Product Yield L:Bg

1 (2c) 72 B 3ac 63%a >95:5

2

(2d)

46 A 3ad 94%a >95:5

3

(2e)

24 A 3ae 89%a >95:5

4 (2f) 48 A 3af 86%a >95:5

5

(2g)

70 C 3ag 90%a >95:5

6

(2h)

48 Af

3ah 77%a >95:5

7 (2i)

84 A 3ai 93%a 86:14

8

(2j)

96 A 3aj 60%a 88:12

9

(2k)

46 E 3ak 76%a,d

83:17

10

(2l)

96 A 3al 52%a,d

35:65

11

(2m)

29 A 3am Quant.a,d

63:37

12 (2n) 70 D 3an 93%a >95:5

13 (2o) B 3ao 40%b >95:5

14

(2p)

36 A 3ap 98%a 89:11

15

(2q)

30 A 3aq 90%a 91:9

16

(2r)

24 A 3ar 70%a 88:12

14

17 (2s) 24 A 3as 64%a 58:42

18

(2t)

36 A 3at 91%a 69:31

19

(2u)

28 B 3au 31%a 90:10

20

(2v)

24 A - n.r. -

Conditions A: 2.5 mol% [Rh(COD)Cl]2, 5 mol% (R)-SIPHOS-PE, 5 mol% K3PO4, 2 equiv. alkene, DCE

Conditions B: 2.5 mol% [Rh(COD)Cl]2, 5 mol% (R)-SIPHOS-PE, 2.5 mol% K3PO4, 5 equiv. alkene, neat

Conditions C: 5 mol% [Rh(COD)Cl]2, 5 mol% (R)-SIPHOS-PE, 5 mol% K3PO4, 2 equiv. alkene, DCE

Conditions D: 5 mol% [Rh(COD)Cl]2, 5 mol% (R)-SIPHOS-PE, 5 mol% K3PO4, 4 equiv. alkene, DCE

Conditions E: 2.5 mol% [Rh(COD)Cl]2, 5 mol% (R)-SIPHOS-PE, 5 mol% K3PO4, 4 equiv. alkene, DCE

n.r. = no reaction. a Isolated yield.

b Approximate yield; could not separate from impurities by chromatography.

c

NMR yield. d Linear and branched regioisomers inseparable by chromatography (isolated together).

e Conversion

based on integration of crude 1H NMR relative to starting material reference peak.

f 3.75 mol% K3PO4 used.

g

Linear-to-branched selectivities determined by integration of crude 1H NMR spectrum

1.3.4 Biphasic Hydroacylation

During our optimization studies, we found that biphasic conditions could be employed for the

hydroacylation of 1-octene (2a) with salicylaldehyde (1a). When a 1:1 mixture of DCE:H2O was

used in the reaction, the desired coupling product was furnished in 97% yield after 19 h (Scheme

3a). To the best of our knowledge, this represents the first example of biphasic conditions being

utilized in hydroacylation. Under these conditions, we believe that the catalyst and 1-octene

remain strictly in the organic layer, while the heterogenous base (K3PO4), as well as any

deprotonated aldehyde substrate, remains dissolved in the aqueous layer. In such a scenario, the

reaction would occur at the interface of the two layers. Interestingly, when we conducted the

same experiment with a stoichiometric amount of base, the reaction proceeded with a similar

level of efficiency (Scheme 3b). If we were to use our standard monophasic conditions

employing DCE as solvent, a drastic decrease in yield would be observed, since having an excess

amount of base relative to [Rh] seems to be detrimental to the reaction. We hypothesize that

catalyst decomposition arises from having excess deprotonated salicylaldehyde in solution

(which is directly correlated to the amount of base) and that biphasic conditions help to slow

down decomposition pathways by separating deprotonated salicylaldehyde and rhodium catalyst

into two layers.

15

Scheme 3: Biphasic conditions for hydroacylation of 1-octene (2a) with salicylaldehyde (1a).

We were intrigued by the high efficiency of the biphasic hydroacylation conditions, and thus

decided to apply these conditions to particularly challenging alkene substrates. Much to our

delight, allyl acetate 2w reacted cleanly under the reaction conditions with no observable

byproducts resulting from SN2’-type chemistry (Table 4, entry 1). Although the linear-selectivity

was only moderate for this substrate, we were able to separate and fully characterized the linear

and branched regioisomers, 3aw and 3aw’. We previously observed poor conversion with

acrylamide substrates using our standard hydroacylation conditions. When biphasic conditions

were employed, full conversion of salicylaldehyde was observed and 57% of the desired

hydroacylation adduct was obtained with excellent linear-selectivity. In addition, we isolated a

significant amount of decarbonylation product (i.e. phenol) along with the desired product,

which accounts for the remaining mass balance.

Table 4: Testing selected alkene substrates under biphasic conditions.

Entry

Product

Isolated Yielda

(%)

L:Bb

1

(2w)

3aw/3aw’ 94c

63:37

2

(2x)

3ax 57 >95:5

a Combined isolated yield.

b Linear-to-branched selectivities determined by integration of crude

1H NMR spectrum.

c Linear and branched regioisomers separated by chromatography.

16

1.3.5 Mechanistic Considerations

We have proposed a general catalytic cycle for the linear-selective hydroacylation of electron-

neutral olefins, such as 1-octene, with salicylaldehyde under our optimized reaction conditions

(Scheme 4). This proposal is consistent with literature precedence regarding the generally-

accepted mechanism for the rhodium-catalyzed olefin hydroacylation of related systems.18

In the

first step of the catalytic cycle, deprotonated salicylaldehyde i can coordinate to the Rh center,

which is followed by oxidative addition of the aldehyde C–H bond (step a). From intermediate ii,

one can imagine reductive decarbonylation occurring, which would result in the formation of a

catalytically-inactive Rh-CO species. In the productive pathway (step b), ligand displacement

and olefin coordination gives rise to intermediate iii. Species iv arises from migratory insertion

of the alkene (step c), which upon reductive elimination (step d) furnishes linear hydroacylation

product v and regenerates the Rh(I) catalyst.

18

(a) Campbell, R. E., Jr.; Lochow, C. F.; Vora, K. P.; Miller, R. G. J. Am. Chem. Soc. 1980, 102, 5824-5830. (b)

Campbell, R. E., Jr.; Miller, R. G. J. Organomet. Chem. 1980, 186, C27-C31. (c) Fairlie, D. P.; Bosnich, B.

Organometallics 1988, 7, 946-954. (d) Hyatt, I. F. D.; Anderson, H. K.; Morehead, A. T., Jr.; Sargent, A. L.

Organometallics 2008, 27, 135-147. (e) Moxham, G. L.; Randell-Sly, H. E.; Brayshaw, S. K.; Weller, A. S.;

Willis, M. C. Chem.sEur. J. 2008, 14, 8383-8397.

17

Scheme 4: Proposed catalytic cycle.

1.4 Importance of Phosphoramidite Ligands in Catalysis

Over the past decade, phosphoramidites have emerged as a privileged class of chiral

monodentate ligands and have thus been exploited in a variety of transition metal-catalyzed

transformations.19

These phosphorous-based ligands, when combined with a metal catalyst, have

been shown to promote reactivity and provide high levels of enantioinduction in asymmetric

hydrogenation, conjugate addition, allylic substitution, cycloadditions, cross-coupling, and

hydroformylation. In general, phosphoramidites possess two P–O bonds and one P–N bond,

rendering the phosphorous atom relatively π-acidic, while maintaining good σ-donating capacity

(Figure 2). In such ligand frameworks, several points of chirality may be present in the ligand

(i.e. in either the diol backbone or the amine portion). Representative diols that have been

19

For a comprehensive review of phosphoramidite ligands in asymmetric catalysis, refer to ref.: Teichert, J. F.;

Feringa, B. L. Angew. Chem. Int. Ed. 2010, 49, 2486-2528.

18

previously incorporated into this class of ligands are outlined in Figure 2. Of the diol backbones

presented, BINOL-based phosphoramidites have found the most widespread use in asymmetric

transformations. Although many variations of BINOL-derivatives have been explored, the

number of analogous derivatives containing a spiroiindanediol (SPINOL) backbone is still

limited.20

Figure 2: Common phosphoramidite ligands.

The synthesis and resolution of SPINOL was achieved in 1999 by Birman and co-workers.21

In

2002, Zhou reported the synthesis of a phosphoramidite ligand that incorporated this spirocyclic

framework, as well as its application in the asymmetric hydrogenation of enamides.22

Since this

seminal report, Zhou has developed a series of phosphoramidites based on the SPINOL-

20

For a review on chiral phosphorus ligands based on a spiro scaffold for transition-metal-catalyzed asymmetric

reactions, refer to ref.: Xie, J.-H.; Zhou, Q.-L. Acc. Chem. Res.2008, 41, 581-593.

21 Birman, V.B.; Rheingold, A. L.; Lam, K.-C. Tetrahedron: Asymmetry 1999, 10, 125-131.

22 Hu, A.-G.; Fu, Y.; Xie, J.-H.; Zhou, H.; Wang, L.-X.; Zhou, Q.-L. Angew. Chem. Int. Ed. 2002, 41, 2348-2350.

19

backbone and has applied these ligands to a variety of transformations.20,23

Due to our group’s

growing interest in the use of spirobiindanediol-based phosphoramidites in hydroacylation, we

wondered whether we could expand the library of available SPINOL-based phosphoramidites.

Ultimately, we plan to employ these ligands in rhodium-catalyzed hydroacylation with the hope

of finding a more active and selective catalyst system, particularly for challenging alkene

substrates.

Phosphoramidites can be synthesized in several ways, by which the order of the key P–O or P–N

formation varies in each case.19

The most common synthetic route towards such motifs begins

with the synthesis of the key chlorophosphite intermediate b from diol a, which may or may not

be chiral (Scheme 5). The final step in the synthesis involves a convergent P–N coupling of the

chlorophosphite with either a chiral or an achiral amine (c), furnishing the final phosphoramidite

d in a two-step sequence. The modularity and step-economy of this synthetic approach should

allow rapid entry to a wide variety of phosphoramidite derivatives.

Scheme 5: Common synthetic route to access phosphoramidites.

1.4.1 Synthesis of Phosphoramidite Ligands: En route to More Active and Selective Hydroacylation Catalysts

Due to the large abundance of BINOL-based phosphoramidites in the literature, we believed that

translating these ligand designs to the analogous SPINOL derivatives would be a good starting

point. Taking inspiration from the success of BINOL-based phosphoramidites, for which ligands

23

Xie,J.-H.; Zhu, S.-F.; Fu, Y.; Hu, A.-G.; Zhou, Q.-L. Pure Appl. Chem. 2008, 77, 2121-2132.

20

derived from amines A1-A6 are known in the literature24

(Figure 3), we plan on synthesizing the

corresponding SPINOL derivatives.25

Figure 3: Proposed amines to be used in spiro-phosphoramidite synthesis.

1.4.1.1 Synthesis of Spirobiindanediol Backbone

Following a procedure by Birman and co-workers,21

m-anisaldehyde (4) was condensed with

acetone, furnishing dienone 5 in good yield (Scheme 6). Hydrogenation of 5 using activated

Raney-Ni under 1 atm of hydrogen afforded ketone 6 in quantitative yield. The crude material

was used directly in the next step without purification. Blocking of the 4-position (i.e. para to

methoxy group) was necessary in order to direct bis-cyclization ortho to the methoxy groups in a

later stage of the synthesis. This was accomplished by carrying out an electrophilic aromatic

substitution reaction using bromine as the electrophile, which afforded ketone 7 in 93% isolated

yield. Unfortunately, upon subjecting 7 to Friedel-Crafts type cyclization conditions employing

polyphosphoric acid, only a 15% isolated yield of pure spirobiindane (+/–)-8 was achieved. The

low yielding nature of this cyclization step was not unexpected as persistent emulsions were

24

For A1, A4, A6: Arnold, L. A.; Imbos, R.; Mandoli, A.; de Vries, A. H. M.; Naasz, R.; Feringa, B. L. Tetrahedron

2000, 56, 2865-2878. For A2: Defieber, C.; Ariger, M. A.; Moriel, P.; Carreira, E. M. Angew. Chem. Int. Ed. 2007,

46, 3139-3143. For A3: Choi, Y. H.; Choi, J. Y.; Yang, H. Y.; Kim, Y. H. Tetrahedron: Asymmetry 2002, 13, 801-

804. For A5: Du, H.; Yuan, W.; Zhao, B.; Shi, Y. J. Am. Chem. Soc. 2007, 129, 11688-11689.

25 For A2, A3, A6, SPINOL-based phosphoramidites known, but not commercially available.

A2: Hoffman, T. J.; Carreira, E. M. Angew. Chem. Int. Ed. 2011, 50, 10670-10674. A3: onz lez, A. . Benitez,

D.; Tkatchouk, E.; Goddard, W. A.; Toste, F. D. J. Am. Chem. Soc. 2011, 133, 5500-5507. A6: Xie,J.-H.; Zhu, S.-F.;

Fu, Y.; Hu, A.-G.; Zhou, Q.-L. Pure Appl. Chem. 2008, 77, 2121-2132.

21

formed during the work-up, rendering full mass recovery impossible. Fortunately, there was

enough material to carry forward to the remaining steps in the synthetic route. Removal of the

bromine protecting groups in rac-8 was accomplished via lithium-halogen exchange and

subsequent quenching with methanol, giving rise to spirobiindane 9 in excellent yield.

Deprotection of the aryl methyl ethers using BBr3 and coupling of the free phenols to L-

menthylchloroformate in the presence of NEt3 and catalytic DMAP gave rise to a 1:1

diastereomeric mixture of 11a/11b. Formation of this diastereomeric mixture allows resolution

of pro-(R) and pro-(S)-SPINOL by standard column chromatography. Upon full separation of

diastereomers 11a and 11b, cleavage of the chiral auxiliaries can be carried out to obtain (R) and

(S)-SPINOL in enantiopure form.

22

Scheme 6: Synthesis of spirobiindanediol backbone.

1.4.1.2 Optimization of Friedel-Crafts Cyclization

Additional experiments were carried out in order to probe the origin of the reproducibility issues

for the PPA-catalyzed Friedel-Crafts cyclization. Different grades of PPA were purchased and

used immediately for the test reactions. Entry 1 (Table 5) represents the first attempt at this

cyclization using PPA at a 6.3 mmol scale, which resulted in a low isolated yield of 15% as

mentioned previously (refer to section 1.4.1.1. Synthesis of Spirobiindanediol Backbone). After

23

scaling the reaction down to 0.2 mmol and employing different grades of PPA (entries 2 and 3),

excellent crude yields of 90% were obtained in both cases following standard work-up

procedure. Unfortunately, when the scale of the reaction was increased to 1.3 mmol, a significant

decrease in yield was observed (entry 4). Therefore, we concluded that this particular step could

not be scaled up efficiently, and thus searched for different acids to promote such a

transformation. Indeed, we were pleased to find that upon switching to methanesulfonic acid

(entry 5), we were able to isolate 44% of the pure material after the first recrystallization step.

Notably, this reaction was conducted at a 4.0 mmol scale, which allows access to sufficient

quantities of rac-8 for later stages in the synthesis.

Table 5: Conditions employed for Friedel-Crafts cyclization.

Entry Acid Scale of reaction

(mmol)

Crude yield (%) Isolated yield (%)

1 PPA (115% H3PO4

basis)

6.3 48 15

2 PPA (115% H3PO4

basis)a

0.2 90 -

3 PPA (>83% P2O5 basis)a

0.2 90 -

4 PPA (>83% P2O5 basis)a

1.3 50 -

5 MsOH 4.0 - 44 a Newly purchased reagent

1.4.1.3 Synthesis of C2-Symmetric Amine

The first amine we were interested in synthesizing was A1, due to its stereoelectronic similarity

to the amine portion in (R)-SIPHOS-PE (i.e. absolute configuration at the methine carbons are

the same as those in (R)-SIPHOS-PE), but increase in steric bulk. We hope that this increase in

steric bulk will, at the very least, give rise to higher levels of selectivity.

24

Scheme 7: Synthesis of C2-symmetric chiral amine A1.

Following a literature procedure,26

condensation of chiral amine 12 and 2-napthylketone (13)

was carried out in the presence of montmorillonite under microwave heating, furnishing ketimine

14 (Scheme 7). The reaction had stalled at approximately 75% conversion, even after prolonged

microwave heating (>24h), with the remaining 25% representing unreacted starting material.

Attempts to separate the product from the remaining naphthylketone 14 via column

chromatography were unsuccessful as this compound co-eluted with the 15 in numerous solvent

compositions. Fortunately, after two rounds of recrystallization, only 11% ketone remained in

the final material. Subjecting 14 as a mixture of the ketimine and unreacted ketone to a reduction

using HSiCl3 in the presence of DMF furnished C2-symmetric amine A1 with excellent

diastereoselectivity, albeit in modest yield. The presence of residual ketone did not seem to affect

the diastereoselectivity, but may have had an influence on the yield of the reaction.

1.5 Conclusions and Future Work

We have developed a novel rhodium-catalyzed method for linear-selective hydroacylation that

employs very low catalyst loadings and can be generalized to a range of different salicylaldehyde

and alkene substrates. Progress has been made with respect to phosphoramidite ligand synthesis

and we hope to synthesize and test these ligands in Rh-catalyzed hydroacylation in the near

26

(a) Guizzetti, S.; Benaglia, M.; Rossi, S. Org. Lett. 2009, 11, 2928-2931. (b) Guizzetti, S.; Benaglia, M.;

Biaggi,C.; Celentano, G. Synlett 2010, 1, 134-136.

25

future. In addition, we plan on further investigating and understanding the effects of our biphasic

hydroacylation conditions by testing other alkene substrates and by varying the catalyst system.

1.6 Experimental

1.6.1 General considerations

Commercial reagents were purchased from Sigma Aldrich, Strem or Alfa Aesar and used without

further purification. Reactions were monitored using thin-layer chromatography (TLC) on EMD

Silica Gel 60 F254 plates. Visualization of the developed plates was performed under UV light

(254 nm) or KMnO4 stain. Organic solutions were concentrated under reduced pressure on a

Büchi rotary evaporator. 1H,

31P and

13C NMR spectra were recorded on a Varian Mercury 400,

VRX-S (Unity) 400, or Bruker AV-III 400 spectrometer. 1H NMR spectra were internally

referenced to the residual solvent signal or TMS. 13

C NMR spectra were internally referenced to

the residual solvent signal. Data for 1H NMR are reported as follows: chemical shift (δ ppm),

multiplicity (s = singlet, d = doublet, t = triplet, q = quartet, m = multiplet, b = broad), coupling

constant (Hz), integration. Data for 13

C NMR are reported in terms of chemical shift (δ ppm).

High resolution mass spectra (HRMS) were obtained on a micromass 70S-250 spectrometer (EI)

or an ABI/Sciex QStar Mass Spectrometer (ESI). Infrared (IR) spectra were obtained on a

Perkin-Elmer Spectrum 1000 FT-IR Systems and are reported in terms of frequency of

absorption (cm-1

). Melting point ranges were determined on a Fisher-Johns Melting Point

Apparatus. Column chromatography was performed with Silicycle Silia-P Flash Silica Gel, using

either glass columns or a Biotage SP-1 system. Solvents used in hydroacylations were degassed

by three freeze-pump-thaw cycles. Chiral ligands were purchased from Strem.

1.6.2 General procedure for the linear-selective hydroacylation of monosubstituted alkenes with salicylaldehydes

In a nitrogen-filled glovebox, K3PO4 base (5 mol%), aldehyde (1.0 equiv.), alkene (2.0 equiv.),

[Rh(COD)Cl]2 precatalyst (2.5 mol%) and (R)-SIPHOS-PE ligand (5.0 mol%) were combined in

a one-dram vial and dissolved in solvent (0.4 M). The vial was charged with a stir bar, sealed

with a Teflon-lined cap and the mixture was stirred at 70 °C for the indicated period of time. The

26

crude reaction mixture was passed through a plug of silica and concentrated. A known amount of

standard 1,3,5-trimethoxybenzene was added to determine the NMR yield. The crude reaction

mixtures were purified by silica gel flash chromatography or preparative TLC to determine the

isolated yield.

1.6.3 General procedure for the linear-selective hydroacylation of monosubstituted alkenes with salicylaldehydes under biphasic conditions

In a nitrogen-filled glovebox, K3PO4 base (10 μmol), aldehyde (100 μmol), alkene (200 or 400

μmol), [Rh(COD)Cl]2 precatalyst (5 μmol) and (R)-SIPHOS-PE ligand (5.0 μmol) were

combined in a one-dram vial and dissolved in DCE (125 μL). The vial was charged with a stir

bar, sealed with a Teflon-lined cap, and brought outside the glovebox. Under a funnel of argon,

degassed H2O was added (125 μL) and the mixture was stirred at 70 °C for 48 hours. The crude

reaction mixture was diluted with EtOAc (1.0 M), quenched with 1.0 M HCl (2.0 mL), and

extracted with EtOAc (3 x 2.0 mL). The combined organics were dried over MgSO4 and the

solvent was removed under reduced pressure to afford the crude product, which was then

purified by preparative TLC to determine the isolated yield.

Notes:

• Air sensitivity: For reasons of convenience, all hydroacylation reactions were performed

in a glovebox filled with nitrogen. A control experiment outside the glovebox using standard

Schlenk techniques led to identical reaction efficiency. A second control experiment revealed

that the reaction also proceeds under air, albeit significantly slower (after one week

approximately 60% conversion was observed).

• Variation of reaction parameters: In a number of cases, the standard procedure was

slightly modified to guarantee optimum reaction efficiency. These deviations from the standard

procedure (typically involving changes of solvent, base or Rh equivalents) are specified for each

particular substrate.

27

1.6.4 Characterization of compounds

1-(2-hydroxyphenyl)-4-(trimethylsilyl)butan-1-one (3ac). Prepared according to the general

procedure 1.6.2 (2.5 mol% K3PO4, neat, 5 equiv. alkene, reaction time 72 h). Purification via

preparative TLC (95:5 hexanes/ethyl acetate) afforded the title compound as a colorless oil (33.5

mg, 63%). 1H NMR (400 MHz, CDCl3) δ 12.41 (s, 1H), 7.74 (dd, J = 8.0, 1.6, 1H), 7.47 – 7.42

(m, 1H), 6.97 (dd, J = 8.4, 1.0, 1H), 6.90 – 6.86 (m, 1H), 2.99 (t, J = 7.3, 2H), 1.86 – 1.62 (m,

2H), 0.66 – 0.46 (m, 2H), 0.00 (s, 9H). 13

C {1H} NMR (101 MHz, CDCl3) δ 208.69, 164.31,

137.93, 131.76, 121.17, 120.57, 120.29, 43.74, 21.17, 18.44. HRMS (ESI+): 237.1304 [M+H]

+

(calcd. 237.1311 for C13H21O2Si).

1-(2-hydroxyphenyl)-6-phenylhexan-1-one (3ad). Prepared according to the general procedure

1.6.2 (solvent: DCE; reaction time 46 h). Purification via preparative TLC (95:5 hexanes/ethyl

acetate) afforded the title compound as a colorless oil (50.5 mg, 94%). 1H NMR (400 MHz,

CDCl3) δ 12.41 (s, 1H), 7.76 (d, J = 8.0 Hz, 1H), 7.48 (t, J = 7.8 Hz, 1H), 7.30 (t, J = 7.5 Hz,

2H), 7.20 (d, J = 7.7 Hz, 3H), 7.00 (d, J = 8.4 Hz, 1H), 6.91 (t, J = 7.6 Hz, 1H), 3.00 (t, J = 7.4

Hz, 2H), 2.66 (t, J = 7.7 Hz, 2H), 1.86 – 1.75 (m, 2H), 1.76 – 1.66 (m, 2H), 1.52 – 1.42 (m, 2H).

13C {

1H} NMR (101 MHz, CDCl3) δ 206.88, 162.66, 142.55, 136.33, 130.08, 128.52, 128.42,

125.84, 119.45, 118.95, 118.67, 38.34, 35.86, 31.36, 29.02, 24.43. HRMS (ESI+): 269.1549

[M+H]+ (calc’d. 269.1542 for C18H21O2).

4,4-diethoxy-1-(2-hydroxyphenyl)butan-1-one (3ag). Prepared according to the general

procedure 1.6.2 (10 mol% [Rh], solvent: DCE; reaction time 70 h). Purification via preparative

28

TLC (9:1 hexanes/ethyl acetate) afforded the title compound as a colorless oil (40.4 mg, 80%).

1H NMR (400 MHz, CDCl3) δ 12.30 (s, 1H), 7.80 (dd, J = 8.0, 1.6, 1H), 7.46 (m, 1H), 6.98 (dd,

J = 8.4, 0.9, 1H), 6.90 (m, 1H), 4.60 (t, J = 5.3, 1H), 3.68 (dq, J = 9.4, 7.1, 2H), 3.51 (dq, J =

9.4, 7.0, 2H), 3.11 (t, J = 7.3, 2H), 2.07 (td, J = 7.3, 5.3, 2H), 1.21 (t, J = 7.1, 6H). 13

C {1H}

NMR (101 MHz, CDCl3) δ 206.12, 162.38, 136.24, 130.02, 119.42, 118.89, 118.47, 101.98,

61.83, 33.07, 28.14, 15.31. HRMS (ESI+): 270.1710 [M+NH4]

+ (calc’d. 270.1705 for

C14H24NO4).

6-hydroxy-1-(2-hydroxyphenyl)octan-1-one (3ah). Prepared according to the general

procedure 1.6.2 (5 mol% [Rh], 5 mol% ligand, 3.75 mol% K3PO4; solvent: DCE; 3.0 equiv.

alkene, reaction time 48 h). Purification via preparative TLC (7:3 hexanes/acetone) afforded the

title compound as a yellow oil (36.5 mg, 77%). 1H NMR (400 MHz, CDCl3) δ 12.36 (s, 1H),

7.76 (dd, J = 8.0, 1.5 Hz, 1H), 7.51 – 7.40 (m, 1H), 6.97 (d, J = 8.0 Hz, 1H), 6.93 – 6.83 (m,

1H), 3.55 (bs, 1H), 3.01 (t, J = 7.3 Hz, 2H), 1.86 – 1.70 (m, 2H), 1.63 – 1.36 (m, 6H), 0.94 (t, J

= 7.4 Hz, 3H). 13

C {1H} NMR (101 MHz, CDCl3) δ 206.85, 162.63, 136.38, 130.08, 119.46,

118.99, 118.68, 73.14, 38.35, 36.74, 30.40, 25.50, 24.51, 10.01. HRMS (ESI+): 237.1484

[M+H]+ (calc’d 237.1491 for C14H21O3).

4-hydroxy-1-(2-hydroxyphenyl)pentan-1-one (3aj). Prepared according to the general

procedure 1.6.2 (5.0 mol% K3PO4, DCE, 2 equiv. alkene, reaction time 120 h). Purification via

preparative TLC (8:2 hexanes/ethyl acetate) afforded the title compound as a colorless oil (23.3

mg, 60%). 1H NMR (400 MHz, CDCl3) δ 12.28 (s, 1H), 7.81 (dd, J = 8.0, 1.6 Hz, 1H), 7.47 (m,

1H), 6.98 (dd, J = 8.4, 1.1 Hz, 1H), 6.91 (m, 1H), 3.91 (bs, 1H), 3.36 – 3.02 (m, 2H), 2.00 – 1.92

(m, 1H), 1.89 – 1.78 (m, 1H), 1.27 (d, J = 6.2 Hz, 3H); 13

C {1H} NMR (101 MHz, CDCl3) δ

206.86, 162.60, 136.51, 130.17, 119.48, 119.08, 118.70, 67.51, 34.73, 33.25, 24.07. Note:

missing aliphatic alcohol peak in 1H NMR spectrum

29

1,4-bis(2-hydroxyphenyl)butan-1-one (3ak). Prepared according to the general procedure 1.6.2

(solvent: DCE, 4 equiv. alkene, reaction time 46 h). Purification via preparative TLC (9:1

hexanes/ethyl acetate) afforded a mixture of the linear and branched regioisomers (83:17 lin/br

selectivity) as an orange oil (39.2 mg, 76%27

).1H NMR (400 MHz, CDCl3) δ 12.16 (s, 1H), 7.76

(dd, J = 8.1, 1.5 Hz, 1H), 7.53 – 7.45 (m, 1H), 7.13 (ddd, J = 7.9, 6.9, 2.2 Hz, 2H), 7.01 (dd, J =

8.4, 0.9 Hz, 1H), 6.95 – 6.83 (m, 3H), 6.28 (s, 1H), 3.11 (t, J = 6.5 Hz, 2H), 2.75 – 2.62 (m, 2H),

2.04 (dq, J = 9.7, 6.5 Hz, 2H).13

C {1H} NMR (101 MHz, CDCl3) δ 207.40, 162.62, 154.42,

136.83, 130.27, 130.17, 127.89, 127.20, 120.63, 119.42, 119.20, 118.81, 115.96, 37.07, 29.76,

24.14. HRMS (ESI+): 257.1167 [M+H]

+ (calc’d 257.1178 for C16H17O3).

1-(2-hydroxyphenyl)-6-methylhept-6-en-1-one (3am). Prepared according to the general

procedure 1.6.2 (solvent: DCE; reaction time 29 h). Purification via preparative TLC (97:3

hexanes/ethyl acetate) afforded a mixture of the linear and branched regioisomers (63:37 lin/br

selectivity, 54.1 mg, 99%28

). All spectroscopic data correspond to those reported in the

literature.14c

1H NMR (400 MHz, CDCl3, linear isomer) δ 12.39 (s, 1H), 7.77 (dd, J = 8.0, 1.7,

1H), 7.46 (m, 1H), 6.98 (dd, J = 8.4, 0.9, 1H), 6.90 (m, 1H), 4.72 (s, 1H), 4.69 (s, 1H), 3.01 (t, J

= 7.5, 2H), 2.08 (t, , 2H) 1.79 (m, 2H), 1.72 (s, 3H), 1.25 (m, 2H).

1-(2-hydroxyphenyl)heptane-1,6-dione (3au). Prepared according to the general procedure

1.6.2 (10 mol% [Rh], solvent: DCE; reaction time 72 h). Purification via preparative TLC (8:2

hexanes/ethyl acetate) afforded the title compound as a colourless oil (17.1 mg, 30%).

27

Isolated yield of linear and branched regioisomers 28

Isolated yield of linear and branched regioisomers

30

Spectroscopic data correspond to those reported in the literature.29

1H NMR (400 MHz, CDCl3) δ

12.31 (s, 1H), 7.75 (dd, J = 8.0, 1.3 Hz, 1H), 7.54 – 7.37 (m, 1H), 6.97 (d, J = 8.4 Hz, 1H), 6.89

(dd, J = 11.2, 4.0 Hz, 1H), 3.01 (t, J = 7.0 Hz, 2H), 2.50 (t, J = 7.0 Hz, 2H), 2.15 (s, 3H), 1.86 –

1.56 (m, 4H); 13

C {1H} NMR (101 MHz, CDCl3) δ 208.62, 206.37, 162.63, 136.45, 130.02,

119.43, 119.05, 118.70, 43.52, 38.16, 30.08, 23.86, 23.44.

4-(2-hydroxyphenyl)-4-oxobutyl acetate (3aw). Prepared according to the general procedure

1.6.3 (200 μmol alkene). Purification via preparative TLC (90:10 hexanes/ethyl acetate) afforded

the title compound as a beige solid (15.3 mg, 69%). 1H NMR (400 MHz, CDCl3) δ 12.24 (s, 1H),

7.76 (dd, J = 8.0, 1.6 Hz, 1H), 7.48 (ddd, J = 8.6, 7.3, 1.6 Hz, 1H), 6.99 (dd, J = 8.4, 1.0 Hz, 1H),

6.91 (ddd, J = 8.2, 7.3, 1.1 Hz, 1H), 4.18 (t, J = 6.3 Hz, 2H), 3.10 (t, J = 7.2 Hz, 2H), 2.16 – 2.06

(m, 2H), 2.04 (s, 3H); 13

C {1H} NMR (101 MHz, CDCl3) δ 205.45, 171.16, 162.59, 136.56,

129.90, 119.42, 119.10, 118.76, 63.69, 34.78, 23.30, 21.04. HRMS (ESI+): 223.0976 [M+H]

+

(calc’d 223.0970 for C12H15O4).

3-(2-hydroxyphenyl)-2-methyl-3-oxopropyl acetate (3aw’). Prepared according to the general

procedure 1.6.3 (200 μmol alkene). Purification via preparative TLC (90:10 hexanes/ethyl

acetate) afforded the title compound as a yellow oil (5.5 mg, 25%). 1H NMR (400 MHz, CDCl3)

δ 12.31 (s, 1H), 7.79 (dd, J = 8.1, 1.5 Hz, 1H), 7.50 (ddd, J = 8.6, 7.3, 1.6 Hz, 1H), 7.01 (dd, J =

8.4, 1.0 Hz, 1H), 6.93 (ddd, J = 8.2, 7.3, 1.1 Hz, 1H), 4.43 (dd, J = 10.8, 8.2 Hz, 1H), 4.20 (dd, J

= 10.8, 5.7 Hz, 1H), 3.90 (dq, J = 14.1, 7.1 Hz, 1H), 1.99 (s, 3H), 1.27 (d, J = 7.1 Hz, 3H); 13

C

{1H} NMR (101 MHz, CDCl3) δ 207.45, 171.02, 163.32, 136.89, 129.99, 119.19, 118.99,

118.79, 65.66, 39.65, 20.95, 15.09. HRMS (ESI+): 223.0970 [M+H]

+ (calc’d 223.0973 for

C12H15O4).

29

Butler, J. D.; Conrad, W. E.; Lodewyk, M. W.; Fettinger, J. C.; Tantillo, D. J.; Kurth, M. J. Org. Lett. 2010, 12,

3410-3413.

31

4-(2-hydroxyphenyl)-N,N-dimethyl-4-oxobutanamide (3ax). Prepared according to the

general procedure 1.6.3 (2 equiv. alkene). Purification via preparative TLC (95:5 DCM/MeOH)

afforded the title compound as a beige solid (12.6 mg, 57%). 1H NMR (400 MHz, CDCl3) δ

12.15 (s, 1H), 7.87 (dd, J = 8.0, 1.6 Hz, 1H), 7.45 (ddd, J = 8.6, 7.3, 1.6 Hz, 1H), 6.96 (dd, J =

8.4, 1.0 Hz, 1H), 6.91 (td, J = 7.2, 3.7 Hz, 1H), 3.39 (t, J = 6.5 Hz, 2H), 3.09 (s, 3H), 2.96 (s,

3H), 2.76 (t, J = 6.5 Hz, 2H); 13

C {1H} NMR (101 MHz, CDCl3) δ 205.37, 171.46, 162.36,

136.40, 130.16, 119.58, 119.12, 118.50, 37.25, 35.71, 33.42, 27.06. HRMS (ESI+): 222.1131

[M+H]+ (calc’d 222.1130 for C12H16N1O3).

1-(2,4-dihydroxyphenyl)nonan-1-one (3ga). Prepared according to the general procedure 1.6.2

(2.5 mol% K3PO4, solvent: Me-THF; reaction time 5 d). Purification via preparative TLC (85:15

hexanes/ethyl acetate) afforded the title compound as a beige solid (47.4 mg, 95%). 1H NMR

(400 MHz, CDCl3) δ 12.89 (s, 1H), 7.66 (d, J = 8.6 Hz, 1H), 6.44 – 6.33 (m, 2H), 6.10 (bs, 1H),

2.89 (t, J = 7.5 Hz, 2H), 1.81 – 1.56 (m, 2H), 1.51 – 1.10 (m, 10H), 0.88 (t, J = 6.7 Hz, 3H). 13

C

{1H} NMR (101 MHz, CDCl3) δ 205.71, 165.35, 162.73, 132.60, 113.96, 107.90, 103.74, 38.22,

31.96, 29.52, 29.51, 29.27, 25.17, 22.78, 14.22. LRMS (ESI+) m/z calc’d for C14H20O3 [M+1]+:

237.31; found: 237.31.

1-(2,4-dihydroxyphenyl)octan-1-one (3gb). Prepared according to the general procedure 1.6.2

(10 mol% [Rh], 5 mol% ligand, 5 mol% K3PO4, solvent: Me-THF; reaction time 3 d).

Purification via preparative TLC (85:15 hexanes/ethyl acetate) afforded the title compound as a

beige solid (11.5 mg, 78%). 1H NMR (400 MHz, CDCl3) δ 12.86 (s, 1H), 7.66 (d, J = 9.4 Hz,

32

1H), 6.40 – 6.37 (m, 2H), 5.73 (bs, 1H), 2.89 (t, J = 8 Hz, 2H), 1.80 – 1.63 (m, 2H), 1.48 – 1.17

(m, 8H), 0.88 (t, J = 6.9 Hz, 3H). 13

C {1H} NMR (101 MHz, CDCl3) δ 205.56, 165.40, 162.56,

132.55, 114.04, 107.78, 103.75, 38.21, 31.82, 29.49, 29.23, 25.10, 22.76, 14.21. LRMS (ESI+)

m/z calc’d for C14H20O3 [M+1]+: 237.31; found: 237.31.

33

Chapter 2 Application of Intermolecular Hydroacylation to the Synthesis of

Octaketide Natural Products

2

2.1 Introduction

2.1.1 Accessing Polyketide Natural Products from Dothiorelone, Cytosporone, and Phomopsin Families via Hydroacylation

Polyketides represent a remarkable class of natural products that display diverse structural

complexity and exhibit a wide range of medicinally important activities, such as antibiotic,

anticancer, antifungal, antiparasitic, and immunosuppressive properties.30

Oftentimes, natural

product isolation can be complicated by either low natural abundance or the purification issues

that accompany such a task. As a consequence, the development of efficient routes towards these

complex molecules in a manner that is both environmentally-friendly and feasible on relatively

large scale has been a longstanding goal for synthetic chemists.

Transition metal-catalyzed alkene hydroacylation represents an attractive strategy for C–C bond

formation owing to its atom-economy and amenability to enantioselective catalysis. Unlike the

related process of hydroformylation,31

hydroacylation has not yet found any significant

application in industry or natural product synthesis, which is largely due to the unfavourable

decarbonylation pathways encountered with traditional rhodium catalysts.3c,32

Due to the

prevalence of aryl ketone motifs in polyketide natural products, we wondered whether our

hydroacylation protocol could be applied to the total synthesis of such natural products. More

specifically, we were inspired by a recent report by Takahashi and co-workers, which disclosed

30

Staunton, J.; Weissman, K. J. Nat. Prod. Rep. 2001, 18, 380-416.

31 For selected reviews on hydroformylation, see: (a) Beller, M.; Cornils, B.; Frohning, C. D.; Kohlpaintner, C. W.;

J. Mol. Catal. A: Chem. 1995, 104, 17-85. (b) Breit, B.; Seiche, W. Synthesis 2001, 1, 1-36.

32 To the best of our knowledge, there has only been one report of alkene hydroacylation being utilized for natural

product synthesis. Ref.: Kim, G.; Lee, E.-j. Tetrahedron: Asymmetry 2001, 12, 2073-2076.

34

the total synthesis and structural revision of phomopsin B to a known octaketide, dothiorelone A

(Scheme 8).33

Scheme 8: Structural revision of phomopsin B and Takahashi’s approach to dothiorelone A.

The synthetic route outlined in this report required protecting group chemistry, as well as several

unnecessary changes in oxidation state. Takahashi’s ten step sequence afforded dothiorelone A

in an overall yield of 15%. In order to improve the step and redox economy of this synthesis, we

envisioned that a linear-selective intermolecular hydroacylation between the appropriate

salicylaldehyde derivative and alkene precursor could furnish dothiorelone A in a convergent and

efficient manner (Scheme 9).

Scheme 9: Proposed key disconnection for synthesis of dothiorelone A.

Following our discovery of this initial report, we identified several families of octaketide natural

products that possess a similar framework to dothiorelone A. These particular polyketides belong

to the dothiorelone, cytosporone, and phomopsin families, and have all been isolated within the

33

Izuchi, Y.; Koshino, H.; Hongo, Y.; Kanomata, N.; Takahashi, S. Org. Lett. 2011, 13, 3360–3363.

35

last decade from different genera of endophytic fungi.34

A general feature of these polyketides is

that they possess a tetrasubstituted arene core with a long alkyl chain appended to the aromatic

ketone (Figure 4). Through this hydroacylation disconnection, we can imagine rapid access to

these octaketide natural products—many of which have not been previously synthesized and

display modest to potent levels of biological activity.35

Figure 4: General structure of octaketide natural products from the dothiorelone, phomopsin,

and cytosporone families.

2.1.2 Cytosporone B

Cytosporone B (Csn-B, Figure 5b) is an octaketide natural product that was first isolated in 2000

by Clardy and co-workers from an endophytic fungus (Cytospora sp.) collected in Costa Rica.34a

Following this isolation paper, Zhan et. al. had identified Csn-B as a natural agonist for nuclear

orphan receptor Nur77 (Figure 5a) and demonstrated that this ligand-receptor binding

interaction, which occurs upstream in a biochemical pathway, could inhibit xenograft tumor

growth in live mice.36

Nuclear receptors are expressed in several tissues throughout the human

body and play important roles in T-cell apoptosis, brain development, metabolism, and hormone

regulation, while orphan receptors are those in which no endogenous or exogenous ligands have

34(a) Brady, S. F. Wagenaar, M. M. Singh, M. P. Janso, J. E. Clardy, J. Org. Lett. 2000, 2, 4043-4046. (b) Xu, Q.

Wang, J. Huang, Y. heng, . Song, S. hang, Y. Su, W. Acta Oceanol. Sin. 2004, 23, 541-547. (c) Huang, .

Cai, X. Shao, C. She, . Xia, X. Chen, Y. Yang, J. hou, S. Lin, Y. Phytochemistry 2008, 69, 1604-1608. (d)

Huang, . uo, . Yang, R. Yin, X. Li, X. Luo, W. She, . Lin, Y. Chem. Nat. Prod. 2009, 45, 625-628. (e) Xu

J. et al., Bioorg. Med. Chem. 2009, 17, 7362-7367. (f) han, Y. et al. Nat. Chem. Biol. 2008, 4, 548-556.

35 Previous total syntheses: (a) Huang, H.; Zhang, L.; Zhang, X.; Ji, X.; Ding, X.; Shen, X.; Jiang, H.; Liu, H. Chin.

J. Chem. 2010, 28, 1041-1043 (cytosporone B). (b) Yoshida, H.; Morishita, T.; Ohshita, J. Chem. Lett. 2010, 39,

508-509 (cytosporone B and phomopsin C). (c) Izuchi, Y.; Koshino, H.; Hongo, Y.; Kanomata, N.; Takahashi, S.

Org. Lett. 2011, 13, 3360-3363 (dothiorelone A).

36 Zhan, Y. et. al. Nat. Chem. Biol. 2008, 4, 548.

36

been identified.37

The role of Nur77 in cancer cell apoptosis was independently reported by two

research teams in 1994, and thus, the identification of Csn-B as an agonist for Nur77 has

profound implications for the future of cancer therapy.38

In the account disclosed by Zhan et. al.,

Csn-B was shown to also elevate glucose levels in fasting mice, which may be important in the

development of drugs for the treatment of hypoglycemia.

Figure 5: (a) Superimposition of Csn-B (yellow) onto the ribbon structure of the Nur77 LBD.

(b) Cytosporone B. (c) Compound 15.

A later report by the same research team described the synthesis of simple Csn-B analogues and

identified the pharmacophore necessary for ligand binding as well as for the activation of a

biological response.39

In this study, it was deduced that the ester group was critical for activating

a biological response and that three other elements were required for ligand binding: (1) the

benzene ring, (2) the phenolic hydroxyl group, (3) and the acyl chain of the Csn-B scaffold. Of

the Csn-B derivatives synthesized, compound 15 (which possesses both an elongated alkyl ester

chain and acyl chain) displayed the highest affinity for the LBD of Nur77, and was the most

37

Pols, T. W. H.; Bonta, P. I.; de Vries, C. J. M. Curr. Opin. Lipidol.2007, 18, 515-520.

38 (a) Woronicz, J. D.; Calnan, B.; Ngo, V.; Winoto, A. Nature 1994, 367, 277-281. (b) Liu, Z-G.; Smith, S. W.;

McLaughlin, K. A.; Schwartz, L. M.; Osborne, B. A. Nature 1994, 367, 281-284.

39 Liu, J. et. al. Cancer Res. 2010, 70, 3628-3637.

37

effective for inducing apoptosis in cancer cell lines and decreasing tumor growth in vivo (Figure

5c). The authors synthesized compound 15, as well as similar Csn-B analogues, via a standard

Friedel-Crafts acylation. Notably, this protocol only allowed for the synthesis of simple

derivatives of Csn-B due to low functional group tolerance of the acylation. Since there is high

potential for Csn-B analogues to serve as cancer therapeutics, this highlights a need for simple,

functional-group tolerant methods for the synthesis of more complex and potent activators of this

receptor.

2.2 Research Goals

The overall goal of this project is to apply our methodology to the total synthesis of octaketide

natural products, which would not only demonstrate that our protocol is amenable to more

complex salicylaldehyde derivatives, but would also allow efficient entry to such families of

octaketides and derivatives thereof. We plan to focus our attention on nine different natural

products of this type—namely, dothiorelones A, B, C, cytosporones A, B, M, N, phomopsin C,

and an unnamed polyketide isolated from the mangrove endophytic fungus Phomopsis by Lin

and co-workers in 2009 (Scheme 10).34d

Ultimately, we plan to implement this strategy in the

synthesis of more complex Csn-B analogues, which can then be subjected to biological testing.

Scheme 10: Proposed synthesis of octaketides belonging to the dothiorelone, cytosporone, and

phomopsin families.

38

2.3 Results and Discussion

2.3.1 Synthesis of Starting Materials for Key Hydroacylation

We envisioned accessing the requisite aldehydes for the key hydroacylation through a Vilsmeier-

Haack formylation of the appropriate arene precursors. Such a reaction requires the protection of

acidic functional groups, such as carboxylic acids and phenols. Thus, our synthesis commenced

with the Fischer esterification of phenyl acetic acid 16 (eq. 8) and methylation of the phenols in

phenyl acetate 18 (eq. 9), furnishing arenes 17 and 19, respectively, in excellent yield (Scheme

11). Formylation of substrates bearing free phenol functionalities proved to be unsuccessful

under a variety of conditions. Using AlCl3, we could achieve either bis-deprotection or selective

mono-deprotection of the aryl methyl ethers by carefully monitoring the reaction time. Aryl

ethers ortho to aldehyde functionalities are more susceptible to cleavage due to a chelation

effect.40

Attempts at using BBr3 as the demethylating reagent proved to be unsuccessful, as

products resulting from ester cleavage were observed. Overall, the synthesis of aldehydes 21a-c

was achieved in a total of four steps (or three steps from commercially available starting

materials 17 and 19) with good to excellent yields throughout the synthetic route.

40

(a) Lansinger, J. M.; Ronald, R. C. Synth. Commun. 1979, 9, 341 – 349. (b) Du, Z.-T.; Lu, J.; Yu, H.-R.; Xu, Y.;

Li, A.-P. J. Chem. Res. 2010, 34, 222-227.

39

Scheme 11: Synthesis of key aldehydes 21a-c.

Since these particular octaketides share a common C-7 alkyl chain that differs only in the

position or absence of a single hydroxyl group on this chain, accessing the requisite alkenes for

the key hydroacylation step was quite simple—owing to their commercial availability or ease of

synthesis. Synthesis of alkenes 22a and 22b was achieved by regioselective opening of

enantiopure epoxides using the appropriate alkenyl Grignard reagents (Scheme 12). To access

the final alkene that bears a C-7 hydroxyl substituent (22c), a DIBAL-H reduction of the ethyl

ester was carried out. Finally, requisite alkene 22d (i.e. 1-heptene) is an inexpensive,

commercially available material.

Scheme 12: Synthesis of key alkenes 22a-c.

40

2.3.2 Testing Known Hydroacylation Methods

Initially we wondered whether two known intermolecular hydroacylation methods developed by

Suemune and Jun could potentially furnish our targeted natural product scaffolds. In Suemune’s

report, the hydroacylation of electron-neutral olefins with salicylaldehyde was promoted by the

addition of acetonitrile to the reaction. However, this transformation required 40 mol% of

Wilkinson’s catalyst and six equivalents of alkene to permit good reactivity and full conversion.

We wanted to test the reactivity of our natural product precursors to Suemune’s conditions, but

with a Rh loading comparable to our system (10 mol%). When we subjected aldehyde 21a and 1-

heptene (22d) to these particular conditions, modest conversion was observed, even after

prolonged reaction times (Scheme 13). When we implemented Jun’s conditions, we observed a

62% conversion of the starting material by 1H NMR, but only trace amounts of the desired

product. Evidently, Jun’s method is not compatible with complex aldehyde substrates—

particularly those bearing free phenol substituents. The results from this study demonstrate that

there is a need for more general and functional group compatible strategies for the

hydroacylation of difficult substrates that maintains low catalyst loadings and employs mild

reaction conditions.

Scheme 13: Testing Suemune’s and Jun’s conditions for synthesis of phomopsin C.

2.3.3 Solvent Screen for Hydroacylation of Natural Product Substrates

We were initially concerned about the solubility of our aldehyde substrates in DCE (the optimal

solvent for our hydroacylation method)—particular those bearing two phenolic groups. Since we

41

observed moderate to good reactivity in both THF and Me-THF for 2,4-dihydroxybenzaldehyde

(1g) (refer to section 1.3.2. Investigation of Aldehyde Scope, Table 2), natural product precursors

21b and 22d were subjected to a modified set of reaction conditions41

employing both THF and

Me-THF as solvent. The reaction progress was monitored by 1H NMR spectroscopy, and four

data points were collected at 18, 51, 71, and 91 hours. The results are summarized and plotted in

Table 6 and Figure 6, respectively. Although both Me-THF and THF were excellent at fully

solubilizing substrate 21b, sluggish reactivity was observed in both cases, with the highest

conversion to product 23bd being 36% in Me-THF after 91 hours (Table 6, entry 4). In both

solvents, significant amounts of decarbonylation observed, which is to be expected for highly

substituted, electron-rich aldehydes. No significant conversion is observed in Me-THF beyond

the 51 hour mark, and in THF, the reaction stalled at 13% conversion before the first data point

was even obtained. At this point, we decided to test the reactivity for this hydroacylation in DCE

with the hope that 21b will solubilize slowly over the course of the reaction. Much to our delight,

complete consumption of aldehyde 21b was observed at 51 hours, with 84% representing

hydroacylation product 23bd and the remaining 16% representing decarbonylation product 21b’

(entry 10). These results suggest that the hydroacylation pathway is more favourable than the

decarbonylation pathway in DCE than it is in Me-THF and THF.

Table 6: Solvent screen and reaction monitoring for hydroacylation of 1-heptene (22d) with

aldehyde 21b.

Solvent Entry Time (h) 23bd (%) 21b (%) 21b’ (%)

Me-THF

1 18 24 74 2

2 51 35 45 10

3 71 36 42 22

4 91 36 39 24

THF 5 18 13 60 27

41

We opted to increase the [Rh] loading for the natural product syntheses as these bulkier, electron-rich aldehydes

proved to be quite unreactive when loadings of 5 mol % monomeric Rh were utilized.

42

6 51 13 48 39

7 71 13 44 42

8 91 14 43 43

DCE 9 18 81 6 13

10 51 84 - 16

Figure 6: Reaction progress for hydroacylation of 1-heptene (22d) with aldehyde 21b in Me-

THF and THF.

2.3.4 Synthesis of Octaketide Natural Products

Using the optimized hydroacylation conditions employing DCE as solvent, we were pleased to

find that our method was suitable for the synthesis of eight octaketides, with yields ranging from

59 – 83% and with good to excellent levels of linear-to-branched selectivity (Table 7). In

instances where the linear selectivity was less than 95:5, we were able to separate the

regioisomers by chromatography. It is interesting to note that the union of aldehyde 21b with

alkene 22b only gave a 59% isolated yield of the desired natural product, dothiorelone B, which

represents the lowest yielding reaction of the substrates tested (entry 4). We presume that

coordination of the nearby hydroxyl group on alkene 22b to the metal center decreases the

0

10

20

30

40

50

60

70

80

10 20 30 40 50 60 70 80 90 100

Co

nve

rsio

n (

%)

Reaction time (h)

21b-MeTHF

23bd-MeTHF

21b'-MeTHF

21b-THF

23bd-THF

21b'-THF

43

reactivity for this coupling and promotes catalyst decomposition; however we are unsure of the

exact reasoning for this. Interestingly, when we subjected 21b and alkene 22b to biphasic

conditions, employing a 1:1 mixture of DCE to H2O, an excellent isolated yield of 86% was

achieved. Finally, a ninth natural product, cytosporone A, was synthesized through

saponification of the ethyl ester in cytosporone B (eq. 10).

Table 7: Synthesis of octaketide natural products.

Entry Natural Product Structurea

[α]D lit. [α]D

found

Yieldb

(%)

L:B

ratioc

1 Octaketide 23aa R1

= Et, R2 = Me, R

4 = OH Not

reported

-3.3 (R) 79 90:10

2 Phomopsin C

(23ad)

R1

= Et, R2 = Me - - 78 >95:5

3 Dothiorelone Ad

(23ba)

R1

= Et, R2 = H, R

4 = OH +3.4 (S) +3.5 (R) 79 83:17

4 Dothiorelone Be

(23bb)

R1

= Et, R2 = H, R

3 = OH Not

reported

-4.0 (R) 59, 86f >95:5

5 Dothiorelone C

(23bc)

R1

= Et, R5 = OH - - 83 87:13

6 Cytosporone B

(23bd)

R1

= Et - - 78 93:7

7 Cytosporone M

(23ca)

R1

= Me, R4 = OH -10 -4.0 (R) 73 89:11

8 Cytosporone N

(23cd)

R1

= Me - - 73 94:6

a Non-specified residues are H.

b Combined isolated yield. Reaction times between 18 – 72 h (Refer to Experimental

for details) c Regioisomeric ratio determined by integration of crude

1H NMR spectrum.

d An [α]D.of +3.4 was

reported for synthetic (S)-dothiorelone A.33

e 7.5 mol% (R)-SIPHOS-PE used.

f Biphasic conditions employed, 1:1

mixture of DCE:H2O

44

To the best of our knowledge, this study represents the first total synthesis of octaketides 23aa,

23bb, 23bc, 23ca, and 23cd.35

In general, we observed good agreement between our

spectroscopic data and those reported in the literature.34

However, an exception to this statement

must be made for octaketide 23aa, for which we found significant deviations in the 1H and

13C

NMR (refer to section 2.5 Experimental).34d

Due to the inconsistency between spectroscopic

data, we conclude that the compound isolated by Huang et. al. does not correspond to their

proposed structure. In addition, another discrepancy was noted in the reported 13

C NMR for

dothiorelone B (23bb) (refer to section 2.5 Experimental).34b,d

Fortunately, we were able to

obtain a crystal structure of 23bb, which confirms that we indeed have the correct product and

allows us to assign the absolute configuration as (R) (Figure 7). We have also assigned the

absolute configuration as (R) in polyketides 23aa, 23ba, and 23ca by comparing to known

literature [α]D values (Table 7).

Figure 7: Solid state structure of dothiorelone B (23bb). Single crystals were obtained by slow

diffusion of hexanes into a DCE solution. O: red; C: black; H: white.

Overall, we have synthesized octaketides belonging to the dothiorelone, cytosporone, and

phomopsin families using intermolecular olefin hydroacylation as the final key step. The main

45

advantages of this methodology are the commercial availability and/or ease of access to the

requisite aldehyde and alkene building blocks, and the functional group tolerance of the reaction.

2.3.5 Design and Synthesis of Cytosporone B Analogues

Using our hydroacylation method, we saw the opportunity to synthesize cytosporone B

analogues with the goal of discovering a more potent small-molecular activator for Nur77. For

these particular Csn-B analogues, we wanted to keep the substitution around the aromatic ring

the same, while we varied the functionality on the acyl chain, since our protocol tolerates a wide

range of functional groups on the olefin component. We hope that these derivatives can bind to

the LBD of Nur77 more tightly through additional non-covalent interactions, such as H-bonding

and/or hydrophobic contacts. Taking into account the pharmacophore described by Liu et. al. and

the crystal structure data of the LBD of Nur77 (Figure 8), we synthesized three Csn-B analogues,

which are outlined in Table 8.

Figure 8: Ligand binding pocket of Nur77 obtained from the PDB (code 2QW4).

The importance of C–F bonds in drug design and lead optimization is evidenced by the relatively

high percentage of currently administered drugs that incorporate this functional motif.42

Due to

the metabolic stability of C–F bonds and their ability to form weak hydrogen bonds with H–X

42

(a) Müller, K.; Faeh, C.; Diederich, F. Science 2007, 317, 1881-1886. (b) Purser, S.; Moore, P. R.; Swallowb, S.;

Gouverneur, V. Chem. Soc. Rev. 2008, 37, 320-330.

46

(where X = N, O, or S) and H–Cα groups on amino acid residues, we were inspired to synthesize

a Csn-B analogue that possesses a perfluorinated acyl chain. Coupling of aldehyde 21b to

perfluro 1-octene 22e, which is a commercially available material, afforded 23be in modest

yield, but with excellent linear-to-branched selectivity (Table 8, entry 1). We believe that this

fluoroalkyl chain has the potential to bind more strongly to the LBD of Nur77 through a specific

interaction with Arg-123 in the binding pocket. Crystal structure evidence from the Protein Data

Bank suggests that Arg resides are fluorophilic in nature, whereby the C–F bonds tend to point

directly toward the guanidinium moiety of Arg residues.42a

Our rationale for the synthesis of derivatives 23bf and 23bg, both of which possess a phenyl

group on the acyl chain, stems from the existence and importance of cation-π interactions in

structural biology.43

We hypothesize that a favourable interaction between an aromatic ring and

Arg-123 is feasible, and that this interaction could lead to higher binding affinity. Although the

linear-to-branched selectivity was less than ideal for alkenes 22f and 22g, we were able to fully

separate the regioisomers by chromatography. We are also interested in synthesizing a derivative

that possesses a ketone functionality on the acyl chain (23bu), since this carbonyl group has the

ability to act as a hydrogen-bond acceptor. Due to the presence of amino acid residues that can

act as hydrogen-bond donors (i.e. Pro-46, Ser-47, Thr-48) in the LBD of Nur77, we suggest that

this positive binding interaction is possible.

Table 8: Synthesis of cytosporone B analogues.

Entry Alkene Reaction

time (h) Product

Yielda

(%) L:B

b

1 (22e) 112 23be 57 >95:5

2

(22f)

55 23bf 69 83:17

43

Gallivan, J. P.; Dougherty, D. A. Proc. Natl. Acad. Sci. U. S. A. 1999, 96, 9459-9464.

47

3

(22g)

55 23bg 39 83:17

4

(22u)

- 23bu - -

a Isolated yield of the linear regioisomer.

bRegioisomeric ratio determined by integration of crude

1H NMR spectrum

2.4 Conclusions and Future Work

We have applied our linear-selective intermolecular hydroacylation reaction to the total synthesis

of nine natural products belonging to the dothiorelone, cytosporone, and phomopsin families.

Yields between 73 and 86% and linear-to-branched selectivities between 83:17 and >95:5 were

observed for the key step. We have also applied this methodology to the synthesis of cytosporone

B analogues for the purpose of biological testing. We hope to carry out docking studies in the

near future for the design of more rational Csn-B derivatives.

2.5 Experimental

2.5.1 General Considerations

Commercial reagents were purchased from Sigma Aldrich, Strem or Alfa Aesar and used without

further purification. Reactions were monitored using thin-layer chromatography (TLC) on EMD

Silica Gel 60 F254 plates. Visualization of the developed plates was performed under UV light

(254 nm) or KMnO4 stain. Organic solutions were concentrated under reduced pressure on a

Büchi rotary evaporator. 1H,

31P and

13C NMR spectra were recorded on a Varian Mercury 400,

VRX-S (Unity) 400, or Bruker AV-III 400 spectrometer. 1H NMR spectra were internally

referenced to the residual solvent signal or TMS. 13

C NMR spectra were internally referenced to

the residual solvent signal. Data for 1H NMR are reported as follows: chemical shift (δ ppm),

multiplicity (s = singlet, d = doublet, t = triplet, q = quartet, m = multiplet, b = broad), coupling

constant (Hz), integration. Data for 13

C NMR are reported in terms of chemical shift (δ ppm).

High resolution mass spectra (HRMS) were obtained on a micromass 70S-250 spectrometer (EI)

or an ABI/Sciex QStar Mass Spectrometer (ESI). Infrared (IR) spectra were obtained on a

Perkin-Elmer Spectrum 1000 FT-IR Systems and are reported in terms of frequency of

48

absorption (cm-1

). Melting point ranges were determined on a Fisher-Johns Melting Point

Apparatus. Column chromatography was performed with Silicycle Silia-P Flash Silica Gel, using

either glass columns or a Biotage SP-1 system. Solvents used in hydroacylations were degassed

by three freeze-pump-thaw cycles. Chiral ligands were purchased from Strem.

2.5.2 General Procedures for the Total Synthesis of Natural Products via Linear-Selective Hydroacylation

In a nitrogen-filled glovebox, K3PO4 base (7.5 mol%), aldehyde (1.0 equiv.), alkene (3.0 equiv.),

[Rh(COD)Cl]2 precatalyst (5.0 mol%) and (R)-SIPHOS-PE ligand (5.0 mol%) were combined in

a one-dram vial in the specified solvent (0.4M). The vial was charged with a stir bar, sealed with

a Teflon-lined cap and the mixture was stirred at 70 °C for the indicated period of time. The

crude reaction mixture was passed through a plug of silica, washed with acetone or ethyl acetate

and concentrated. The crude reaction mixtures were purified by silica gel flash chromatography

or preparative TLC to determine the isolated yield.

2.5.3 Characterization of Compounds

ethyl 2-(2-formyl-3,5-dimethoxyphenyl)acetate (20a). To a solution of ethyl 2-(3,5-

dimethoxyphenyl)acetate (1.490 g, 6.5 mmol) in DMF (2.5 mL) was added freshly distilled

POCl3 (0.72 mL, 7.8 mmol). The reaction mixture was stirred at room temperature for 19 h

before diluting with DCM and quenching with H2O (10 mL) at 0 °C. The aqueous layer was

extracted with EtOAc (3 x 20 mL), and the combined organics were washed with brine, dried

with MgSO4, and concentrated under reduced pressure. Purification by flash chromatography

(7:3 hexanes/acetone) afforded the title compound as a white solid (1.381 g, 84%). 1H NMR (400

MHz, CDCl3) δ 10.43 (s, 1H), 6.43 (d, J = 2.3 Hz, 1H), 6.33 (d, J = 2.2 Hz, 1H), 4.18 (q, J = 7.1

49

Hz, 2H), 3.93 (s, 2H), 3.90 (s, 3H), 3.87 (s, 3H), 1.28 (t, J = 7.1 Hz, 3H).13

C {1H} NMR (101

MHz, CDCl3) δ 190.00, 171.12, 165.36, 164.77, 139.36, 117.15, 109.96, 97.14, 60.69, 55.85,

55.53, 40.74, 14.25.

methyl 2-(2-formyl-3,5-dimethoxyphenyl)acetate (20b). To a solution of methyl 2-(3,5-

dimethoxyphenyl)acetate (0.485 g, 2.3 mmol) in DMF (0.89 mL) was added freshly distilled

POCl3 (0.26 mL, 2.8 mmol). The reaction mixture was heated to 55 °C for 48 h before diluting

with ethyl acetate (5.0 mL) and quenching with ice-cold H2O (5.0 mL). The aqueous phase was

extracted with ethyl acetate (3 x 5.0 mL), and the combined organics were washed with brine,

dried with MgSO4, and concentrated under reduced pressure. Purification by flash

chromatography (7:3 hexanes/acetone) afforded the title compound as a white solid (375 mg,

72%). 1H NMR (400 MHz, CDCl3) δ 10.42 (s, 1H), 6.43 (d, J = 2.2 Hz, 1H), 6.33 (d, J = 2.1 Hz,

1H), 3.95 (s, 2H), 3.90 (s, 3H), 3.87 (s, 3H), 3.71 (s, 3H). 13

C {1H} NMR (101 MHz, CDCl3) δ

190.20, 171.69, 165.52, 164.93, 139.34, 117.27, 110.10, 97.34, 55.99, 55.68, 52.04, 40.67.

ethyl 2-(2-formyl-3-hydroxy-5-methoxyphenyl)acetate (21a). To a suspension of anhydrous

AlCl3 (1.34 g, 10.0 mmol) in DCM (6.0 mL) was added a solution of 20a (0.253 g, 1.00 mmol)

in DCM (4.0 mL) at 0 °C. The reaction mixture was heated to 40 °C for 1 hour, diluted with

DCM, and poured into ice-cold water (10 mL). The aqueous phase was extracted with ethyl

acetate (3 x 10 mL), and combined organics were washed with brine, dried using MgSO4, and

concentrated under reduced pressure. Purification by column chromatography (8:2 hexanes/ethyl

acetate) furnished the title compound as a white solid (0.206 g, 87%). M.p. 74-76 °C. 1H NMR

(400 MHz, CDCl3) δ 12.50 (s, 1H), 10.06 (s, 1H), 6.36 (s, 2H), 4.17 (q, J = 7.1 Hz, 2H), 3.84 (s,

3H), 3.82 (s, 2H), 1.25 (t, J = 7.1 Hz, 3H). 13

C {1H} NMR (101 MHz, CDCl3) δ 192.92, 170.33,

50

166.85, 166.62, 139.52, 113.08, 111.67, 100.14, 61.70, 55.81, 38.06, 14.26. HRMS (ESI+):

239.0915 [M+H]+ (calc’d 239.0920 for C12H15O5).

ethyl 2-(2-formyl-3,5-dihydroxyphenyl)acetate (21b). To a suspension of anhydrous AlCl3

(5.280 g, 39.6 mmol) in DCM (15 mL) was added a solution of 20a (0.968 g, 3.84 mmol) in

DCM (10 mL) at 0 °C. The reaction vessel was heated to 40 °C for 45 hours, diluted with DCM,

and poured into ice-cold water (25 mL). The aqueous phase was extracted with ethyl acetate (3 x

25 mL), and combined organics were washed with brine, dried using MgSO4, and concentrated

under reduced pressure. Purification by flash chromatography (7:3 hexanes/ethyl acetate)

afforded the title compound as a white solid (0.751 g, 87%). M.p. 110-112 °C. 1H NMR (400

MHz, CD3CN) δ 12.31 (bs, 1H), 9.96 (s, 1H), 8.02 (bs, 1H), 6.32 – 6.26 (m, 2H), 4.12 (q, J = 7.1

Hz, 2H), 3.88 (s, 2H), 1.21 (t, J = 7.1 Hz, 3H); 13

C {1H} NMR (101 MHz, CD3CN) δ 194.80,

171.56, 167.02, 165.49, 142.21, 113.76, 112.29, 102.73, 62.06, 38.07, 14.39. HRMS (ESI+):

225.0767 [M+H]+ (calc’d 225.0763 for C11H13O5).

methyl 2-(2-formyl-3,5-dihydroxyphenyl)acetate (21c). To a suspension of anhydrous AlCl3

(1.019 g, 7.6 mmol) in DCM (15 mL) was added a solution of 20b (0.182 g, in DCM (8.0 mL) at

0 °C. The reaction vessel was heated to 40 °C for 48 hours, was diluted with DCM, and poured

into ice-cold water (15 mL). The aqueous phase was extracted with ethyl acetate (3 x 15 mL),

and the combined organics were washed with brine, dried with MgSO4, and concentrated under

reduced pressure. Purification by flash chromatography (7:3 hexanes/acetone) afforded the title

compound as a white solid (149 mg, 90%). M.p. 118-120 °C. 1H NMR (400 MHz, Acetone-d6) δ

12.43 (bs, 1H), 10.07 (s, 1H), 9.70 (bs, 1H), 6.41 (d, J = 2.3 Hz, 1H), 6.28 (d, J = 2.3 Hz, 1H),

4.00 (s, 2H), 3.67 (s, 3H). 13

C {1H} NMR (101 MHz, acetone) δ 382.56, 370.98, 348.09, 342.30,

51

318.43, 290.10, 288.84, 279.08, 228.90, 214.12. HRMS (ESI+): 211.0604 [M+H]

+ (calc’d

211.0607 for C10H12O5 ).

(R)-hept-6-en-2-ol (22a). Following a literature procedure,44

a solution of 3-butenylmagnesium

bromide [prepared from 4-bromo-1-butene (4.05 g, 30 mmol) and Mg (1.29 g, 53 mmol)] in THF

(50 mL) was added dropwise to a suspension of CuCN (121 mg, 1.4 mmol) and (R)-propylene

oxide (871 mg, 15 mmol) in THF (20 mL) at -78 °C. The reaction mixture was warmed to room

temperature before quenching with saturated NH4Cl (30 mL). The aqueous layer was extracted

with Et2O (3 x 30 mL). The combined organics were washed with brine, dried with MgSO4, and

concentrated under reduced pressure. Purification by flash chromatography (7:3 pentanes/Et2O)

afforded the title compound as a colorless oil (1.24 g, 72% over two steps). All spectroscopic

data correspond to those reported in the literature.45

1H NMR (400 MHz, CDCl3) δ 5.86 – 5.76

(m, 1H), 5.04 – 4.98 (m, 1H), 4.97 – 4.94 (m, 1H), 3.86 – 3.75 (m, 1H), 2.10- 2.05 (m, 2H), 1.56

– 1.37 (m, 4H), 1.31 (d, J = 3.5 Hz, 1H), 1.19 (d, J = 6.2 Hz, 3H). 98% ee confirmed by chiral

GC-FID (Rt = 6.5 min, 80 °C (isothermal), Cyclodex B column, 30 m x 0.25 mm).

(R)-hept-6-en-3-ol (22b). Following a literature procedure,44

a solution of allylmagnesium

bromide [prepared from allyl bromide (3.00 g, 25 mmol) and Mg (1.09 g, 45 mmol)] in THF (40

mL) was added dropwise to a suspension of CuCN (101 mg, 1.1 mmol) and (R)-1,2-epoxybutane

(901 mg, 12.5 mmol) in THF (15 mL) at -78 °C. The reaction mixture was warmed to room

temperature before quenching with saturated NH4Cl (25 mL). The aqueous layer was extracted

with Et2O (3 x 25 mL). The combined organics were washed with brine, dried with MgSO4, and

concentrated under reduced pressure. Purification by flash chromatography (75:25

pentanes/Et2O) afforded the title compound as a light yellow oil (1.24 g, 70% over two steps). 1H

44

Jastrzebska, I.; Scaglione, J. B.; DeKoster, G. T.; Rath, N. P.; Covey, D. F. J. Org. Chem. 2007, 72, 4837-4843.

45 Takahata , H.; Yotsui, Y.; Momose, T. Tetrahedron 1998, 54, 13505-13516.

52

NMR (400 MHz, CDCl3) δ 5.90 – 5.80 (m, 1H), 5.08 – 5.02 (m, 1H), 5.01 – 4.93 (m, 1H), 3.59 –

3.52 (m, 1H), 2.32 – 2.04 (m, 2H), 1.65 – 1.39 (m, 5H), 1.36 (d, J = 4.7 Hz, 1H), 0.95 (t, J = 7.5

Hz, 3H). (c = 0.032, EtOH) [lit.

(c = 1.7, EtOH) for (S)

enantiomer].46

hept-6-en-1-ol (22c). Following a literature procedure,47

to a solution of ethyl hept-6-enoate

(1.50 g, 9.6 mmol) in Et2O (120 mL) was added DIBAL-H (21 mmol from a 1.0 M solution in

hexanes) at -78 °C. The reaction mixture was stirred at room temperature for 3 hours before

quenching with saturated NH4Cl (30 mL) at -10 °C. The viscous white solution was stirred

vigorously at room temperature for 1 hour, filtered through a pad of celite, and washed with

Et2O. The filtrate was dried over MgSO4, concentrated under reduced pressure, and purified by

flash chromatography (7:3 pentanes/Et2O) to afford the title compound as a colorless oil (944

mg, 86%). All spectroscopic data correspond to those reported in the literature.47

1H NMR (400

MHz, CDCl3) δ 5.86 – 5.76 (m, 1H), 5.03 – 4.97 (m, 1H), 4.97 – 4.91 (m, 1H), 3.67 – 3.63 (m,

2H), 2.09 – 2.04 (m, 2H), 1.64 – 1.51 (m, 2H), 1.49 – 1.32 (m, 4H).

(R)-ethyl 2-(3-hydroxy-2-(7-hydroxyoctanoyl)-5-methoxyphenyl)acetate (23aa). Prepared

according to the general procedure 2.5.2 (150 μmol aldehyde solvent: DCE reaction time 19 h).

Purification via preparative TLC (98:2 chloroform/isopropanol) afforded the title compound as a

colorless solid (41.7 mg, 79%, lin/br 90:10). M.p. 120-122 °C. Spectroscopic data correspond to

those reported in the literature.51b

1H NMR (400 MHz, CDCl3) δ 12.28 (s, 1H), 6.36 (d, J = 2.6

Hz, 1H), 6.32 (d, J = 2.6 Hz, 1H), 4.17 (q, J = 7.1 Hz, 2H), 3.83 (s, 2H), 3.79 (s, 3H), 3.80 – 3.73

46

Bolm, C.; Schlingloff, G.; Harms, K. Chem. Ber. 1992, 125, 1191-1203. 47

Zaed, A.M.; Swift, M. D.; Sutherland, A. Org. Biomol. Chem. 2009, 7, 2678-2680.

53

(m, 1H), 2.84 (t, J = 8.0 Hz, 2H), 1.77 – 1.65 (m, 2H), 1.48 – 1.38 (m, 3H), 1.38 – 1.30 (m, 3H),

1.25 (t, J = 7.1 Hz, 3H), 1.17 (d, J = 6.2 Hz, 3H).; 13

C {1H} NMR (101 MHz, CDCl3) δ 206.54,

171.03, 164.95, 163.70, 136.34, 116.36, 112.54, 100.59, 68.14, 61.49, 55.52, 43.19, 42.02, 39.18,

29.26, 25.61, 24.94, 23.63, 14.28. HRMS (ESI+): 353.1961 [M+H]

+ (calc’d 353.1964 for

C19H29O6). (c = 0.02, MeOH) (no literature reference for [α]D value)

Comparison of NMR Data:

1H NMR (400 MHz,

CDCl3) synthesized

1H NMR (500 MHz,

CDCl3) isolated51b

13C NMR (101 MHz,

CDCl3) synthesized

13C NMR (125 MHz,

CDCl3) isolated51b

12.28 (bs, 1H) 11.39 (s) 206.54 206.8

6.36 (d, J = 2.6, 1H) 6.38 (d, J = 2.0, 1H) 170.03 171.2

6.32 (d, J = 2.6, 1H)

4.17 (q, J = 7.1, 2H)

6.36 (d, J = 2.0, 1H)

4.11 (q, J = 7.2, 2H)

165.95

163.70

161.2

158.7

3.83 (s, 2H) 3.60 (s, 2H) 136.34 134.4

3.79 (s, 3H) 3.80 (s, 3H) 116.36 124.1

3.80 – 3.73 (m, 1H) 3.77 (m, 1H) 112.54 107.7

2.84 (t, J = 8.0, 2H) 2.82 (t, J = 7.2, 2H) 100.59 97.5

1.77 – 1.65 (m, 2H) 1.64 (m, 2H) 68.14 66.8

1.48 – 1.38 (m, 3H) 1.44 (m, 2H) 61.49 60.8

1.38 – 1.30 (m, 3H), 1.42 (m, 2H)

1.33 (m, 2H)

55.52

43.19

55.3

43.9

1.25 (t, J = 7.1, 3H), 1.24 (t, J = 7.2, 3H) 42.02 39.1

1.17 (d, J = 6.2, 3H) 1.16 (d, J = 6.4, 3H) 39.18

29.26

38.8

29.2

25.61 25.5

24.94 23.9

23.63 23.4

14.28 14.1

ethyl 2-(3-hydroxy-5-methoxy-2-octanoylphenyl)acetate, phomopsin C (23ad). Prepared

according to the general procedure 2.5.2 (44 μmol aldehyde; solvent: DCE; reaction time 72 h).

54

Purification via preparative TLC (85:15 hexanes/ethyl acetate) afforded the title compound as a

colorless solid (11.5 mg, 78%, lin/br >95:5). M.p. 62-64 °C. All spectroscopic data corresponded

to those reported in the literature.48

1H NMR (400 MHz, CDCl3) δ 12.59 (s, 1H), 6.39 – 6.32 (m

2H), 4.18 (q, J = 7.1 Hz, 2H), 3.86 (s, 2H), 3.81 (s, 3H), 2.83 (t, J = 8 Hz, 2H), 1.77 – 1.63 (m,

2H), 1.31 – 1.20 (m, 13H), 0.88 (t, J = 6.8 Hz, 3H); 13

C {1H} NMR (101 MHz, CDCl3) δ 206.75,

170.90, 165.50, 163.85, 136.44, 116.04, 112.73, 100.59, 61.50, 55.56, 43.34, 42.19, 31.82, 29.34,

29.22, 25.15, 22.74, 14.30, 14.20. HRMS (ESI+): 337.2017 [M+H]

+ (calc’d 337.2015 for

C19H29O5). Note: Multiplet between 1.31 – 1.20 ppm should integrate to 11H; discrepancy likely

due to a small amount of the branched regioisomer.

Comparison of NMR Data:

1H NMR (400 MHz,

CDCl3) synthesized

1H NMR (500 MHz,

CDCl3) isolated48

13C NMR (101 MHz,

CDCl3) synthesized

13C NMR (125 MHz,

CDCl3) isolated48

12.59 (s, 1H) 12.53 (s, 1H) 206.75 206.6

6.39 – 6.32 (m, 2 H) 6.38 (d, J = 2.0, 1H) 170.90 170.7

4.18 (q, J = 7.1, 2H)

6.33 (d, J = 2.0, 1H)

4.18 (q, J = 7.2, 2H)

165.50

163.85

165.2

163.6

3.86 (s, 2H)

3.81 (s, 3H)

3.86 (s, 2H)

3.81 (s, 3H)

136.44

116.04

136.2

115.9

2.83 (t, J = 8.0, 2H) 2.83 (t, J = 7.2, 2H) 112.73 112.5

1.77 – 1.66 (m, 2H) 1.70 (m, 2H) 100.59 100.4

1.31 – 1.20 (m, 13H) 1.26 – 1.24 (m, 11H) 61.50 61.3

0.88 (t, J = 6.8, 3H) 0.88 (t, J = 7.2, 3H) 55.56

43.34

55.4

43.2

42.19

31.82

42.0

31.6

29.34

29.22

29.0

29.1

48

Huang, Z.; Cai, X.; Shao, C.; She, Z.; Xia, X.; Chen, Y.; Yang, J.; Zhou, S.; Lin, Y. Phytochemistry 2008, 69,

1604-1608.

55

(R)-ethyl 2-(3,5-dihydroxy-2-(7-hydroxyoctanoyl)phenyl)acetate, (-)-dothiorelone A (23ba).

Prepared according to the general procedure 2.5.2 (150 μmol aldehyde; solvent: DCE; reaction

time 18 h). Purification via preparative TLC (95:5 chloroform/isopropanol) afforded the title

compound as a colorless oil (40.2 mg, 79%, lin/br 83:17). Spectroscopic data correspond to those

reported in the literature.49

1H NMR (400 MHz, CDCl3) δ 11.26 (bs, 1H), 7.46 (bs, 1H), 6.23 (s,

2H), 4.18 (q, J = 7.1 Hz, 2H), 3.86 – 3.77 (m, 1H), 3.75 (s, 2H), 2.08 (bs, 1H), 2.84 (t, J = 8.0

Hz, 2H), 1.79 – 1.60 (m, 2H), 1.52 – 1.29 (m, 6H), 1.27 (t, J = 7.1 Hz, 3H), 1.18 (d, J = 6.2 Hz,

3H); 1H NMR (400 MHz, Acetone-d6) δ 6.36 (d, J = 2.3 Hz, 1H), 6.32 (d, J = 2.3 Hz, 1H), 4.09

(q, J = 7.1 Hz, 2H), 3.74 – 3.68 (m, 1H), 3.67 (s, 2H), 2.91 (t, J = 8.0 Hz, 2H), 1.73 – 1.57 (m,

2H), 1.43 – 1.23 (m, 8H), 1.21 (t, J = 7.1 Hz, 3H), 1.11 (d, J = 6.1 Hz, 3H); 13

C {1H} NMR (126

MHz, CDCl3) δ 206.79, 171.59, 163.98, 160.47, 136.74, 116.99, 112.69, 103.29, 68.39, 61.72,

43.28, 41.80, 39.03, 29.15, 25.47, 24.88, 23.61, 14.30. HRMS (ESI+): 339.1820 [M+H]

+ (calc’d

339.1808 for C18H27O6). (c = 0.02, MeOH) [lit. a)

(c = 0.5,

MeOH)50

for (S) enantiomer, synthetic material; b) (c = 0.1, MeOH)

53,

b isolated

material].

Comparison of NMR Data:

1H NMR (400 MHz,

CDCl3) synthesized

1H NMR (500 MHz,

CDCl3) previously

synthesized49

13C NMR (126 MHz,

CDCl3) synthesized

13C NMR (101 MHz,

CDCl3) previously

synthesized49

11.26 (bs, 1H) 12.24 (s, 1H) 206.79 206.5

7.46 (bs, 1H) 5.42 (bs, 1H) 171.59 170.7

6.23 (s, 2H)

6.33 (d, J = 2.3, 1H)

6.28 (d, J = 2.3, 1H)

163.98

160.47

164.7

160.0

49

Izuchi, Y.; Koshino, H.; Hongo, Y.; Kanomata, N.; Takahashi, S. Org. Lett. 2011, 13, 3360-3363.

50 Takahata , H.; Yotsui, Y.; Momose, T. Tetrahedron 1998, 54, 13505-13516.

56

4.18 (q, J = 7.1, 2H) 4.19 (q, J = 7.1, 2H) 136.74 136.9

3.86 – 3.77 (m, 1H) 3.79 (m, 1H) 116.99 116.6

3.75 (s, 2H) 3.84 (s, 2H) 112.69 112.3

2.08 (bs, 1H) - 103.29 103.2

2.84 (t, J = 7.3, 2H) 2.85 (t, J = 7.4, 2H) 68.39 68.1

1.79 – 1.60 (m, 2H) 1.73 (m, 2H) 61.72 61.5

1.52 – 1.29 (m, 6H)

1.27 (t, J = 8.0, 3H)

1.44 (m, 3H)

1.34 (m, 3H)

1.27 (t, J = 7.1, 3H)

43.28

41.80

39.03

43.1

41.8

39.0

1.18 (d, J = 6.2, 3H) 1.18 (d, J = 6.0, 3H) 29.15

25.47

29.1

25.5

24.88 24.8

23.61 23.5

14.30 14.2

(R)-ethyl 2-(3,5-dihydroxy-2-(6-hydroxyoctanoyl)phenyl)acetate, dothiorelone B (23bb).

Prepared according to the general procedure 2.5.2 (94 μmol aldehyde 7.5 mol% ligand; solvent:

DCE; reaction time 60 h). Purification via preparative TLC (2:1 hexanes/acetone) afforded the

title compound as a colorless solid (18.8 mg, 59%, lin/br >95:5). Spectroscopic data correspond

to those reported in the literature.51b

M.p. 88-90 °C. 1H NMR (400 MHz, CDCl3) δ 11.60 (bs,

1H), 6.52 (bs, 1H), 6.26 – 6.24 (m, 2H), 4.19 (q, J = 7.1, 2H), 3.79 (s, 2H), 3.56 (bs, 1H), 2.86 (t,

J = 7.2, 2H), 1.83 – 1.59 (m, 3H), 1.59 – 1.33 (m, 6H), 1.26 (t, J = 7.1, 3H), 0.94 (t, J = 7.4,

3H). 13

C {1H} NMR (101 MHz, CDCl3) δ 206.61, 171.51, 163.70, 160.26, 136.58, 116.97,

112.54, 103.16, 73.28, 61.57, 43.22, 41.61, 36.42, 30.18, 25.18, 24.72, 14.15, 9.89. HRMS

51

a) Xu, Q.; Wang, J.; Huang, Y.; Zheng, Z.; Song, S.; Zhang, Y.; Su, W. Acta Oceanol. Sin., 2004, 23, 541-547; b)

Huang, Z.; Guo, Z.; Yang, R.; Yin, X.; Li, X.; Luo, W.; She, Z.; Lin, Y. Chem. Nat. Prod. 2009, 45, 625-628.

57

(ESI+): 339.1811 [M+H]

+ (calc’d. 339.1808 for C18H27O6).

(c = 0.01, MeOH)

[no lit. data]. Note: In the isolation report, no signal was reported for the CH3 group of the ethyl

ester in the 13

C NMR spectrum (14.15 ppm in our spectrum). Because the remaining

characterization data matches, this is likely due to a mistake by the authors.

Comparison of NMR Data:

1H NMR (400 MHz,

CDCl3) synthesized

1H NMR (500 MHz,

CDCl3) isolated51b

13C NMR (101 MHz,

CDCl3) synthesized

13C NMR (125 MHz,

CDCl3) isolated51b

11.60 (bs, 1H) 11.47 (OH, bs) 206.61 206.6 (C)

6.52 (bs, 1H) 8.05 (OH, s) 171.51 171.2 (C)

6.26 – 6.24 (m, 2H) 6.27 (1H, s) 163.70 163.7 (C)

6.26 (1H, s) 160.26 160.5 (C)

4.19 (q, J = 7.1, 2H) 4.19 (2H, q, J = 7.2) 136.58 136.6 (C)

3.79 (s, 2H) 3.78 (2H, s) 116.97 116.9 (C)

3.56 (m, 1H) 3.55 (1H, m) 112.54 112.7 (CH)

2.86 (t, J = 7.2, 2H) 2.86 (2H, t, J = 7.2) 103.16 103.2 (CH)

1.83 – 1.59 (m, 3H) 1.46 (2H, m) 73.28 73.1 (CH)

1.59 – 1.33 (m, 6H) 1.43 (2H, m) 61.57 61.7 (CH2)

1.26 (2H, m) 43.22 43.3 (CH2)

1.25 (2H, m) 41.61 41.7 (CH2)

1.26 (t, J = 7.1, 3H) 1.24 (3H, t, J = 7.2) 36.42 36.5 (CH2)

0.94 (t, J = 7.4, 3H) 0.94 (3H, t, J = 7.4) 30.18 30.3 (CH2)

29.8 (CH2)

29.7 (CH3)

25.18 25.2 (CH2)

24.72

14.15

9.89 9.9 (CH3)

58

ethyl 2-(3,5-dihydroxy-2-(8-hydroxyoctanoyl)phenyl)acetate, dothiorelone C (23bc).

Prepared according to the general procedure 2.5.2 (200 μmol aldehyde solvent: DCE; reaction

time 48 h). Purification via preparative TLC (95:5 chloroform/isopropanol) afforded the title

compound as a colorless oil (56.2 mg, 83%, lin/br 87:13). Spectroscopic data correspond to those

reported in the literature.51b

1H NMR (400 MHz, CDCl3) δ 6.30 – 6.26 (m, 2H), 4.19 (q, J = 7.1

Hz, 2H), 3.82 (s, 2H), 3.64 (t, J = 6.6 Hz, 2H), 2.83 (t, J = 8.0 Hz, 2H), 1.79 – 1.63 (m, 2H),

1.64 – 1.46 (m, 2H), 1.38 – 1.31 (m, 6H), 1.27 (t, J = 7.1 Hz, 3H); 1H NMR (400 MHz, Acetone-

d6) δ 6.36 (d, J = 2.3 Hz, 1H), 6.31 (d, J = 2.3 Hz, 1H), 4.08 (q, J = 7.1 Hz, 2H), 3.67 (s, 2H),

3.54 (t, J = 6.5 Hz, 2H), 2.90 (t, J = 8.0 Hz, 2H), 1.67 – 1.60 (m, 2H), 1.52 – 1.47 (m, 2H), 1.40

– 1.29 (m, 6H), 1.21 (t, J = 7.1 Hz, 3H); 13

C {1H} NMR (101 MHz, CDCl3) δ 206.84, 171.36,

164.36, 160.38, 136.85, 116.88, 112.63, 103.33, 63.12, 61.69, 43.39, 41.87, 32.69, 29.18, 29.15,

25.60, 24.94, 14.30. HRMS (ESI+): 339.1803 [M+H]

+ (calc’d 339.1808 for C18H27O6). Note:

ROH protons do not appear in the 1H NMR spectrum (rapid exchange).

Comparison of NMR Data:

1H NMR (400 MHz,

CDCl3) synthesized

1H NMR (500 MHz,

CDCl3) isolated51b

13C NMR (101 MHz,

CDCl3) synthesized

13C NMR (125 MHz,

CDCl3) isolated51b

-

-

12.06 (bs, 1H)

7.89 (bs, 1H)

206.84

171.36

206.0

171.3

6.30 – 6.26 (m, 2H)

4.19 (q, J = 7.1, 2H)

3.82 (s, 2H)

3.64 (t, J = 6.6, 2H)

6.32 (d, J = 2.5, 1H)

6.28 (d, J = 2.5, 1H)

4.17 (q, J = 7.2, 2H)

3.82 (s, 2H)

3.65 (t, J = 6.5, 2H)

164.36

160.38

136.85

116.88

112.63

164.6

160.4

136.8

116.6

112.5

2.83 (t, J = 7.4, 2H)

1.79 – 1.63 (m, 2H)

1.64 – 1.46 (m, 3H)

2.82 (t, J = 7.0, 2H)

1.37 (m, 2H)

1.35 (m, 2H)

103.33

63.12

61.69

103.2

63.1

61.4

1.38 – 1.31 (m, 5H)

1.33 (m, 2H)

1.28 (m, 2H)

1.26 (m, 2H)

43.39

41.87

32.69

43.3

41.8

32.7

1.27 (t, J = 7.1, 3H) 1.25 (t, J = 7.2, 3H) 29.18 29.8

59

29.15 29.0

25.60 25.5

24.94 24.9

14.30 14.1

ethyl 2-(3,5-dihydroxy-2-octanoylphenyl)acetate, Cytosporone B (23bd). Prepared according

to the general procedure 2.5.2 (200 μmol aldehyde; solvent: DCE; reaction time 41 h).

Purification via preparative TLC (98:2 DCM/IPA) afforded the title compound as a colorless

solid (22.2 mg, 78%, lin/br 93:7). M.p. 59-61 °C. Spectroscopic data correspond to those

reported in the literature.52 1

H NMR (400 MHz, CDCl3) δ 11.68 (s, 1H), 6.71 (s, 1H), 6.26 – 6.25

(m, 2H), 4.20 (q, J = 7.1 Hz, 2H), 3.80 (s, 2H), 2.82 (t, J = 8.0 Hz, 2H), 1.70 – 1.66 (m, 2H),

1.30 – 1.26 (m, 11H), 0.89 – 0.87 (m, 3H); 13

C {1H} NMR (101 MHz, CDCl3) δ 206.94, 171.91,

164.13, 160.47, 136.62, 116.87, 112.84, 103.42, 61.84, 43.60, 41.85, 31.82, 29.37, 29.22, 25.12,

22.74, 14.27, 14.21. HRMS (ESI+): 323.1849 [M+H]

+ (calc’d 323.1859 for C18H27O5).

Comparison of NMR Data:

1H NMR (400 MHz,

CDCl3) synthesized

1H NMR (400 MHz,

CDCl3) previously

synthesized52

13C NMR (101 MHz,

CDCl3) synthesized

13C NMR (101 MHz,

CDCl3) previously

synthesized52

11.86 (bs, 1H) 12.1 (s, 1H) 206.94 206.7

6.71 (bs, 1H) - 171.91 171.3

6.26 – 6.25 (m, 2 H) 6.29 (d, J = 2.9, 1H)

6.26 (d, J = 2.0, 1H)

164.13

160.47

164.4

160.3

4.20 (q, J = 7.1, 2H) 4.19 (q, J = 6.8, 2H) 136.62 136.6

3.80 (s, 2H) 3.83 (s, 2H) 116.87 116.5

52

Yoshida, H.; Morishita, T.; Ohshita, J. Chem. Lett. 2010, 39, 508-509.

60

2.82 (t, J = 8.0, 2H) 2.82 (t, J = 6.8, 2H) 112.84 112.6

1.70 – 1.66 (m, 2H) 1.50 – 1.69 (m, 2H) 103.42 103.2

1.30 – 1.26 (m, 11H) 1.10 – 1.45 (m, 11H) 61.84 61.6

0.89 – 0.87 (m, 3H) 0.87 (t, J = 7.8, 3H) 43.60

41.85

43.3

41.8

31.82

29.37

31.7

29.2

29.22 29.1

25.12 24.9

22.74 22.6

14.27 14.1

14.21 14.0

(R)-methyl 2-(3,5-dihydroxy-2-(7-hydroxyoctanoyl)phenyl)acetate, cytosporone M (23ca).

Prepared according to the general procedure 2.5.2 (150 μmol aldehyde solvent: DCE reaction

time 48 h). Purification via preparative TLC (95:5 chloroform/isopropanol) afforded the title

compound as a beige solid (35.5 mg, 73%, lin/br 89:11). M.p. 122-124 °C. Spectroscopic data

correspond to those reported in the literature.53

1H NMR (400 MHz, Acetone-d6) δ 9.69 (s, 1H),

8.78 (s, 1H), 6.37 (d, J = 2.3 Hz, 1H), 6.32 (d, J = 2.3 Hz, 1H), 3.74 – 3.69 (m, 1H), 3.68 (s, 2H),

3.62 (s, 3H), 2.90 (t, J = 8.0 Hz, 2H), 1.70 – 1.57 (m, 2H), 1.51 – 1.26 (m, 6H), 1.10 (d, J = 6.2

Hz, 3H).; 13

C {1H}NMR (101 MHz, MeOD) δ 208.94, 173.99, 161.41, 159.86, 136.98, 121.21,

111.80, 102.78, 68.54, 52.32, 45.15, 40.30, 40.08, 30.57, 26.74, 25.57, 23.47. HRMS (ESI+):

325.1655 [M+H]+ (calc’d 325.1651 for C17H25O6).

(c = 0.02, MeOH) [lit.

(c = 0.02, MeOH)

52 for isolated material].

Note: No reference available for 13

C NMR

61

Comparison of NMR Data:

1H NMR (400 MHz,

Acetone-d6)

synthesized

1H NMR (500 MHz,

Acetone-d6)

isolated53

9.69 (bs, 1H) -

8.78 (bs, 1H) -

6.37 (d, J = 2.3, 1H) 6.37 (d, J = 2.3, 1H)

6.32 (d, J = 2.3, 1H) 6.30 (d, J = 2.3, 1H)

3.74 – 3.69 (m, 1H) 3.68 (m, 1H)

3.68 (s, 2H) 3.67 (s, 2H)

3.62 (s, 3H) 3.61 (s, 3H)

2.90 (t, J = 8.0, 2H) 2.92 (t, J = 7.2, 2H)

1.70 – 1.57 (m, 2H) 1.62 (m, 2H)

1.51 – 1.26 (m, 6H) 1.47 – 1.27 (m, 6H)

1.10 (d, J = 6.2, 3H) 1.09 (d, J = 6.3, 3H)

methyl 2-(3,5-dihydroxy-2-octanoylphenyl)acetate, cytosporone N (23cd). Prepared

according to the general procedure 2.5.2 (150 μmol aldehyde; solvent: DCE; reaction time 48 h).

Purification via preparative TLC (95:5 chloroform/isopropanol) afforded the title compound as a

beige solid (34.4 mg, 74%, lin/br 94:6). M.p. 99-101 °C. Spectroscopic data correspond to those

reported in the literature.53,54

1H NMR (400 MHz, Acetone-d6) δ 6.37 (d, J = 2.3 Hz, 1H), 6.31

53

Xu, J.; Kjer, J.; Sendker, J.; Wray, V.; Guan, H.; Edrada, R.; Müller, W. E. G.; Bayer, M.; Lin, W.; Wu, J.;

Proksch, P. Bioorg. Med. Chem. 2009, 17, 7362-7367. [Please note that dothiorelone A is incorrectly labelled

‘dothiorelone B’ in this paper.]

54Shen, Y.; Wu, Q.; Su, W.; Zheng, Z.; Xu, Q.; Du, X.; Huang, Y.; Song, S.; Liu, J.; Hu, Z. CN 1733693, 2006.

62

(d, J = 2.1 Hz, 1H), 3.68 (s, 2H), 3.61 (s, 3H), 2.95 – 2.84 (m, 2H), 1.66 – 1.59 (m, 2H), 1.37 –

1.22 (m, 8H), 0.88 (t, J = 6.9 Hz, 3H).; 13

C {1H} NMR (126 MHz, CDCl3) δ 207.03, 172.41,

163.79, 160.58, 136.43, 116.98, 112.82, 103.46, 52.72, 43.69, 41.48, 31.83, 29.36, 29.21, 25.08,

22.76, 14.21. HRMS (ESI+): 309.1701 [M+H]

+ (calc’d 309.1702 for C17H25O5). Note: Both

phenolic protons do not appear in the 1H NMR spectrum (rapid exchange).

Comparison of NMR Data:

1H NMR (400 MHz,

Acetone-d6)

synthesized

1H NMR (500 MHz,

Acetone-d6)

isolated53

13C NMR (126 MHz,

CDCl3) synthesized

13C NMR (125 MHz,

CDCl3) isolated54

6.37 (d, J = 2.3, 1H) 6.37 (d, J = 1.8, 1H) 207.03 206.7

6.31 (d, J = 2.3, 1H) 6.31 (d, J = 1.8, 1H) 171.41 171.3

3.68 (s, 2H) 3.67 (s, 2H) 163.79 164.6

3.61 (s, 3H) 3.62 (s, 3H) 160.58 160.2

2.95 – 2.84 (m, 2H) 2.89 (t, J = 6.2, 2H) 136.43 136.6

1.66 – 1.59 (m, 2H) 1.62 (m, 2H) 116.98 116.6

1.37 – 1.22 (m, 8H) 1.32 – 1.29 (m, 8H) 112.82 112.4

0.88 (t, J = 6.9, 3H) 0.88 (t, J = 5.8, 3H) 103.46 103.2

52.72 52.4

43.69 43.4

41.48 41.5

31.83 31.7

29.36 29.2

29.21 29.0

25.08 25.0

22.76 22.6

14.21 14.0

63

2-(3,5-dihydroxy-2-octanoylphenyl)acetic acid, cytosporone A (24). Prepared from

cytosporone B by saponification of the ethyl ester. Under N2, cytosporone B (8a, 4.3 mg, 13.3

μmol, 1.0 equiv.) was dissolved in THF (1.0 mL) and a degassed solution of LiOH (5.6 mg, 133

μmol, 10 equiv.) in H2O (0.2 mL) was added. The mixture was stirred for 6 h at room

temperature, the bulk of solvent was evaporated under reduced pressure and 10% NaOH (1.0

mL) and Et2O (1.0 mL) was added. The aqueous layer was acidified with conc. HCl, the mixture

was extracted with ethyl acetate (3 x 2 mL), dried over MgSO4 and the solvent was removed

under reduced pressure. Purification via preparative TLC (6:4 hexanes/ethyl acetate; spiked with

1% AcOH) afforded the title compound as a colorless oil (3.0 mg, 77%). Spectroscopic data

correspond to those reported in the literature.55 1

H NMR (400 MHz, Acetone-d6) δ 6.38 (d, J =

2.1, 1H), 6.34 (d, J = 2.1, 1H), 3.64 (s, 2H), 2.96 (t, J = 7.3, 2H), 1.67 – 1.60 (m, 2H), 1.31 –

1.28 (m, 8H), 0.87 (t, J = 6.8, 3H). 13

C {1H} NMR (126 MHz, Acetone-d6) δ 207.72, 173.54,

161.63, 160.89, 138.36, 120.18, 111.93, 102.63, 44.46, 41.53, 32.61, 30.4*, 30.4*, 25.41, 23.39,

14.44. *: shifts assigned by HSQC 2D NMR (overlap with solvent signal in 1D 13

C spectrum).

13C NMR (101 MHz, CD3OD) δ 209.59, 175.29, 161.35, 159.80, 137.26, 121.35, 111.61, 102.66,

45.15, 40.44, 32.93, 30.50, 30.29, 25.57, 23.70, 14.42. HRMS (ESI+): 295.1554 [M+H]

+ (calc’d

295.1546 for C16H23O5).

Comparison of NMR Data:

1H NMR (400 MHz,

Acetone-d6)

synthesized

1H NMR (500 MHz,

Acetone-d6)

isolated55

13C NMR (126 MHz,

Acetone-d6)

synthesized

13C NMR (125 MHz,

Acetone-d6)

isolated55

6.38 (d, J = 2.1, 1H) 6.36 (s, 1H) 207.7 207.5

6.34 (d, J = 2.1, 1H) 6.34 (s, 1H) 173.5 172.5

55

Brady, S. F.; Wagenaar, M. M.; Singh, M. P.; Janso, J. E.; Clardy, J. Org. Lett. 2000, 2, 4043-4046.

64

3.64 (s, 2H) 3.64 (s, 2H) 161.6 161.0

2.96 (t, J = 7.3, 2H) 2.94 (t, J = 7.5, 2H) 160.9 160.1

1.67 – 1.60 (m, 2H) 1.63 (m, 2H) 138.4 137.6

1.31 – 1.28 (m, 8H) 1.27 – 1.30 (m, 8H) 120.2 120.5

0.87 (t, J = 6.8, 3H) 0.86 (t, J = 7.0, 3H) 111.9 111.7

102.6 102.5

44.5 44.4

41.5 40.6

32.6 32.5

30.4 (HSQC) 30.0

30.4 (HSQC) 29.9

25.4 25.1

23.4 23.3

14.4 14.3

Ethyl 2-(3,5-dihydroxy-2-(4,4,5,5,6,6,7,7,8,8,9,9,9-tridecafluorononanoyl)phenyl)acetate

(23be). Prepared according to the general procedure 2.5.2 (126 μmol aldehyde; solvent: DCE;

reaction time 112 h). Purification via preparative TLC (70:30 hexanes/acetone) afforded the title

compound as a beige solid (40.6 mg, 57%). 1H NMR (400 MHz, Acetone-d6) δ 9.44 (bs, 1H),

8.84 (bs, 1H), 6.43 (d, J = 2.3 Hz, 1H), 6.34 (d, J = 2.3 Hz, 1H), 4.08 (q, J = 7.1 Hz, 2H), 3.70 (s,

2H), 3.34 – 3.24 (m, 2H), 2.69 – 2.47 (m, 2H), 1.21 (t, J = 7.1 Hz, 3H); 13

C {1H} NMR (101

MHz, Acetone-d6) δ 202.18, 171.84, 160.99, 159.14, 137.52, 120.35, 112.03, 102.66, 61.04,

39.79, 35.25, 26.53 (t, J = 21.4 Hz), 14.49. HRMS (ESI+): 571.0796 [M+H]

+ (calc’d 571.0784

for C19H16O5F13). Note: perfluorinated carbon atoms on acyl chain not observable in 13

C NMR.

65

ethyl 2-(3,5-dihydroxy-2-(6-phenylhexanoyl)phenyl)acetate (23bf). Prepared according to the

general procedure 2.5.2 (100 μmol aldehyde solvent: DCE reaction time 55 h). Purification via

preparative TLC (97:3 chloroform/isopropanol) afforded the title compound as a white solid

(25.5 mg, 69%). 1H NMR (400 MHz, CDCl3) δ 11.96 (s, 1H), 7.29 – 7.25 (m, 2H), 7.18 – 7.15

(m, 3H), 6.27 (d, J = 2.5 Hz, 1 H), 6.25 (d, J = 2.5 Hz, 1H), 4.17 (q, J = 7.1 Hz, 2H), 3.81 (s,

2H), 2.87 – 2.73 (m, 2H), 2.69 – 2.55 (m, 2H), 1.73 (dt, J = 15.1, 7.5 Hz, 2H), 1.64 (dt, J = 15.4,

7.6 Hz, 2H), 1.42 – 1.31 (m, 2H), 1.25 (t, J = 7.1 Hz, 3H); 13

C {1H} NMR (101 MHz, CDCl3) δ

206.66, 171.60, 164.42, 160.40, 142.60, 136.76, 128.53, 128.41, 125.82, 116.79, 112.76, 103.41,

61.77, 43.43, 41.90, 35.85, 31.33, 28.96, 24.92, 14.27. HRMS (ESI+): 371.1860 [M+H]

+ (calc’d

371.1859 for C22H27O5).

ethyl 2-(3,5-dihydroxy-2-(4-phenoxybutanoyl)phenyl)acetate (23bg). Prepared according to

the general procedure 2.5.2 (100 μmol aldehyde; solvent: DCE; reaction time 55 h). Purification

via preparative TLC (97:3 chloroform/isopropanol) afforded the title compound as a beige solid

(14.1 mg, 39%). 1H NMR (400 MHz, CDCl3) δ 11.74 (s, 1H), 7.31 – 7.19 (m, 2H), 6.98 – 6.88

(m, 1H), 6.89 – 6.86 (m, 2H), 6.26 – 6.24 (m, 2H), 4.15 (q, J = 7.1 Hz, 2H), 4.02 (t, J = 6.0 Hz,

2H), 3.85 (s, 2H), 3.07 (t, J = 7.1 Hz, 2H), 2.28 – 2.10 (m, 2H), 1.22 (t, J = 7.1 Hz, 3H); 13

C

{1H} NMR (101 MHz, CDCl3) δ 205.70, 171.79, 164.22, 160.50, 158.89, 136.84, 129.60,

120.94, 116.84, 114.64, 112.88, 103.42, 66.92, 61.80, 41.90, 39.96, 24.71, 14.21. HRMS (ESI+):

359.1497 [M+H]+ (calc’d 359.1495 for C20H23O6).

66

Appendix A: NMR Spectra

1H NMR spectrum (400 MHz, CDCl3, 298 K) of 3ga

13C {

1H} NMR spectrum (101 MHz, CDCl3, 298 K) of 3ga

67

1H NMR spectrum (400 MHz, CDCl3, 298 K) of 3ac

13C {

1H} NMR spectrum (101 MHz, CDCl3, 298 K) of 3ac

68

1H NMR spectrum (400 MHz, CDCl3, 298 K) of 3ad

13C {

1H} NMR spectrum (101 MHz, CDCl3, 298 K) of 3ad

69

1H NMR spectrum (400 MHz, CDCl3, 298 K) of 3ag

13C {

1H} NMR spectrum (101 MHz, CDCl3, 298 K) of 3ag

70

1H NMR spectrum (400 MHz, CDCl3, 298 K) of 3ah

13C {

1H} NMR spectrum (101 MHz, CDCl3, 298 K) of 3ah

71

1H NMR spectrum (400 MHz, CDCl3, 298 K) 3aj

13C {

1H} NMR spectrum (101 MHz, CDCl3, 298 K) 3aj

72

1H NMR spectrum (400 MHz, CDCl3, 298 K) of 3ak (mix of linear and branched isomers)

13C {

1H} NMR spectrum (101 MHz, CDCl3, 298 K) 3ak (mix of linear and branched isomers)

73

1H NMR spectrum (400 MHz, CDCl3, 298 K) of 3am (mix of linear and branched isomers)

1H NMR spectrum (400 MHz, CDCl3, 298 K) of 3au

74

13C {

1H} NMR spectrum (101 MHz, CDCl3, 298 K) of 3au

1H NMR spectrum (400 MHz, CDCl3, 298 K) of 3aw (linear isomer)

75

13C {

1H} NMR spectrum (101 MHz, CDCl3, 298 K) of 3aw (linear isomer)

1H NMR spectrum (400 MHz, CDCl3, 298 K) of 3aw’ (branched isomer)

76

13C {

1H} NMR spectrum (101 MHz, CDCl3, 298 K) of 3aw’ (branched isomer)

1H NMR spectrum (400 MHz, CDCl3, 298 K) of 3ax

77

13C {

1H} NMR spectrum (101 MHz, CDCl3, 298 K) of 3ax

1H NMR spectrum (400 MHz, CDCl3, 298 K) of 20a

78

13C {

1H} NMR spectrum (101 MHz, CDCl3, 298 K) of 20a

1H NMR spectrum (400 MHz, CDCl3, 298 K) of 20b

79

13C {

1H} NMR spectrum (101 MHz, CDCl3, 298 K) of 20b

1H NMR spectrum (400 MHz, CDCl3, 298 K) of 21a

80

13C {

1H} NMR spectrum (101 MHz, CDCl3, 298 K) of 21a

1H NMR spectrum (400 MHz, CD3CN, 298 K) of 21b

81

13C {

1H} NMR spectrum (101 MHz, CD3CN, 298 K) of 21b

1H NMR spectrum (400 MHz, Acetone-d6, 298 K) of 21c

82

13C {

1H} NMR spectrum (101 MHz, Acetone-d6, 298 K) of 21c

1H NMR spectrum (400 MHz, CDCl3, 298 K) of (R)-hept-6-en-2-ol 22a

83

1H NMR spectrum (400 MHz, CDCl3, 298 K) of (R)-hept-6-en-3-ol 22b

1H NMR spectrum (400 MHz, CDCl3, 298 K) of hept-6-en-1-ol of 22c

84

1H NMR spectrum (400 MHz, CDCl3, 298 K) of unnamed octaketide 23aa

13C {

1H} NMR spectrum (101 MHz, CDCl3, 298 K) of unnamed octaketide 23aa

85

1H NMR spectrum (400 MHz, CDCl3, 298 K) of phomopsin C (23ad)

13C {

1H} NMR spectrum (101 MHz, CDCl3, 298 K) of phomopsin C (23ad)

86

1H NMR spectrum (400 MHz, Acetone-d6, 298 K) of dothiorelone A (23ba)

13C {

1H} NMR spectrum (126 MHz, CDCl3, 298 K) of dothiorelone A (23ba)

87

1H NMR spectrum (400 MHz, CDCl3, 298 K) of dothiorelone B (23bb)

13C {

1H} NMR spectrum (101 MHz, CDCl3, 298 K) of dothiorelone B (23bb)

88

1H NMR spectrum (400 MHz, Acetone-d6, 298 K) of dothiorelone C (23bc)

13C {

1H} NMR spectrum (101 MHz, CDCl3, 298 K) of dothiorelone C (23bc)

89

1H NMR spectrum (400 MHz, CDCl3, 298 K) of cytosporone B (23bd)

13C {

1H} NMR spectrum (101 MHz, CDCl3, 298 K) of cytosporone B (23bd)

90

1H NMR spectrum (400 MHz, Acetone-d6, 298 K) of cytosporone M (23ca)

13C {

1H} NMR spectrum (101 MHz, MeOD, 298 K) of cytosporone M (23ca)

91

1H NMR spectrum (400 MHz, Acetone-d6, 298 K) of cytosporone N (23cd)

13C {

1H} NMR spectrum (126 MHz, CDCl3, 298 K) of cytosporone N (23cd)

92

1H NMR spectrum (400 MHz, Acetone-d6, 298 K) of cytosporone A (24)

13C {

1H} NMR spectrum (101 MHz, MeOD, 298 K) of cytosporone A (24)

93

1H NMR spectrum (400 MHz, Acetone-d6, 298 K) of 23be

13C {

1H} NMR spectrum (101 MHz, Acetone-d6, 298 K) of 23be

94

1H NMR spectrum (300 MHz, CDCl3, 298 K) of 23bf

13C {

1H} NMR spectrum (101 MHz, CDCl3, 298 K) of 23bf

95

1H NMR spectrum (300 MHz, CDCl3, 298 K) of 23bg

13C {

1H} NMR spectrum (101 MHz, CDCl3, 298 K) of 23bg

96

Appendix B: Crystallographic Data

Crystal data and structure refinement for dothiorelone B (23bb)

Identification code cu_d12103_0m

Empirical formula C18 H26 O6

Formula weight 338.39

Temperature 147(2) K

Wavelength 1.54178 Å

Crystal system Triclinic

Space group P 1

Unit cell dimensions a = 10.0136(5) Å α = 108.517(2)°.

b = 10.1157(5) Å β = 101.863(2)°.

c = 10.9568(5) Å γ = 115.104(2)°.

Volume 874.95(7) Å3

Z 2

Density (calculated) 1.284 Mg/m3

Absorption coefficient 0.791 mm-1

F(000) 364

Crystal size 0.19 x 0.15 x 0.10 mm3

Theta range for data collection 4.64 to 67.39°.

Index ranges -11<=h<=11, -12<=k<=12, -13<=l<=13

Reflections collected 32251

Independent reflections 6063 [R(int) = 0.032]

Completeness to theta = 67.39° 97.9 %

Absorption correction Semi-empirical from equivalents

Max. and min. transmission 0.7528 and 0.6700

Refinement method Full-matrix least-squares on F2

Data / restraints / parameters 6063 / 3 / 443

Goodness-of-fit on F2 1.075

Final R indices [I>2sigma(I)] R1 = 0.0441, wR2 = 0.1238

R indices (all data) R1 = 0.0476, wR2 = 0.1265

Absolute structure parameter 0.1(2)

Largest diff. peak and hole 0.244 and -0.295 e.Å-3